Sie sind auf Seite 1von 254

TSUNAMIS IN THE CARIBBEAN SEA

IMPLICATIONS FROM COARSE-CLAST DEPOSITS AND THE IMPORTANCE OF THEIR SHAPE

Von der Fakultät für Bauingenieurwesen der Rheinisch-Westfälischen Technischen


Hochschule Aachen zur Erlangung des akademischen Grades eines Doktors der
Ingenieurwissenschaften genehmigte Dissertation

vorgelegt von

Jan Oetjen

Berichter: Univ.-Prof. Dr.-Ing. Holger Schüttrumpf


Prof. Dr. Helmut Brückner

Tag der mündlichen Prüfung: 09.12.2021

Diese Dissertation ist auf den Internetseiten der Universitätsbibliothek online verfügbar.
„Ich bin so klug, K – L – U – K.“
Homer Jay Simpson
Kurzfassung

Kurzfassung

An vielen Küsten rund um die Erde stellen Tsunamis eine ernsthafte Bedrohung für die Bevölke-
rung dar, welche direkte (z. B. Todesopfer durch Ertrinken) und indirekte Schäden (u. a. Schäden
an Infrastruktur und Beeinträchtigung der Trinkwasserversorgung) nach sich ziehen können.
Tsunamis der jüngeren Vergangenheit, wie 2011 in Japan (Tōhoku-Oki-Tsunami) oder 2004 in
Thailand (Indian-Ocean-Tsunami), haben gezeigt, dass es in vielen Regionen an einem ausrei-
chenden Verständnis zur konkreten lokalen Tsunamigefährdung mangelt.

So waren, obwohl bereits vor 2011 geologische Hinweise auf hochenergetische Tsunamis vorla-
gen, die Tsunami-Managementpläne und Minderungsmaßnahmen in vielen Regionen Japans le-
diglich an den stärksten Ereignissen der vergangenen 100 bis 200 Jahre bemessen und daher
nicht ausreichend, um den Auswirkungen des Tōhoku-Oki-Tsunamis gewachsen zu sein. Da die
vorliegenden geologischen Ablagerungen jedoch bereits auf solch intensive Tsunamis hinwiesen,
wurde im Nachgang zu der Tsunamikatastrophe 2011 in Japan dazu übergegangen, geologische
Hinweise auf nicht dokumentierte historische und prähistorische Tsunamis (Paläo-Tsunamis) ex-
plizit in der Auslegung von Minderungsmaßnahmen und in Managementplänen zu berücksichtigen.
Die Berücksichtigung solcher geologischer Hinweise erfüllt dabei den Nutzen, Informationen zu
Tsunamis bereitzustellen, welche nicht aufgezeichnet oder beobachtet wurden und langjährige
Eintrittsintervalle bzw. sehr niedrige Eintrittswahrscheinlichkeiten aufweisen (Wiederkehrintervalle
von 1.000 Jahren und mehr). Damit erlaubt es die Analyse geologischer Ablagerungen küstenspe-
zifisch abzuschätzen, ob hier Tsunami-Ereignisse auftreten können, welche bisherige Beobach-
tungen oder Abschätzungen übertreffen. So können Risikominderungsstrategien und Manage-
mentpläne entsprechend angepasst werden.

In der vorliegen Dissertation wird die Analyse geologischer Ablagerungen für die Karibische See
und insbesondere die Insel Bonaire (ABC-Inseln, Kleine Antillen) angewendet. Während für die
Karibik eine große Anzahl an Tsunamibeobachtungen aus den vergangenen ca. 520 Jahren exis-
tieren, von denen jedoch nur ein Bruchteil verifiziert ist, liegen für Bonaire und die gesamten ABC-
Inseln keine konkreten Aufzeichnungen vor. Gerade auf Bonaire gibt es jedoch eine große Anzahl
an geologischen Ablagerungen, die auf hochenergetische Wellenereignisse hinweisen. Die Unter-
suchung des möglicherweise tsunami-induzierten Transports einiger dieser Ablagerungen auf
Bonaire in Form experimenteller und numerischer Modelle bilden den Kern der vorliegenden Dis-
sertation, wobei diese von einer ganzheitlichen Analyse der Tsunamigefahr im karibischen Raum
begleitet werden.

In diesem Kontext wurde in einem ersten Schritt das Inventar an Tsunamiablagerungen in der
gesamten Karibik erhoben und der Versuch unternommen, diese Ablagerungen bezüglich des da-
zugehörigen Verlagerungsereignisses (Sturmwellen oder Tsunamis) zu klassifizieren. Dazu wur-
den unter anderem numerische Tsunamisimulationen für die Karibik auf Basis von verifizierten und
angenommenen Erdbebenszenarien als Tsunamitrigger durchgeführt. Im Zuge dieser Simulatio-
nen wurden die Tsunamihöhen entlang aller Küstenlinien in der Karibik berechnet, das besondere
Augenmerk lag jedoch auf den Gebieten Boka Olivia und Spelonk entlang der Nordküste Bonaires.

Im Vergleich der simulierten Tsunamis mit den Ergebnissen aus empirischen Block-Transport-
Gleichungen wurde deutlich, dass die simulierten Tsunamis keine Wellenintensitäten erzeugen,
die ausreichend sind, um die größten Boulder auf der Insel Bonaire umzulagern. Als Ursache für

I
Kurzfassung

diese Diskrepanz konnten u. a. die ungenügende Genauigkeit der empirischen Transport-Glei-


chungen (z. B. infolge zu starker Vereinfachungen) und inadäquate Auswahl oder Reproduktion
der Erdbebenszenarien identifiziert werden.

Unter dem Eindruck dieser Resultate und zur Einschätzung der Genauigkeit von empirischen Boul-
der-Transport-Gleichungen wurde im Anschluss eine umfangreiche Analyse bisheriger physikali-
scher Experimente zum tsunamiinduzierten Block-Transport durchgeführt und deren Schwachstel-
len identifiziert. Als größte Schwachstelle trat dabei die Vernachlässigung von Boulderformen
abseits idealisierter Würfel oder Quader hervor, welche daher als zu untersuchendes Kernelement
der anschließenden physikalischen Experimente ausgewählt wurde. Im Rahmen dieser Recherche
wurde des Weiteren deutlich, dass die heute verfügbaren und vielfach angewendeten empirischen
Gleichungen zur Beschreibung des Block-Transports durch Tsunamis nicht geeignet sind, exakte
Werte zu möglichen Tsunamiintensitäten zu bestimmen. Daher wird empfohlen, stattdessen mit
Wahrscheinlichkeitsbereichen zur Boulder-Mobilisierung zu arbeiten und sich damit in öffentlichen
Aussagen auch auf Bereiche möglicher Tsunamiintensitäten zu beschränken. Als weiteres Ele-
ment der Studie wurde ein Werkzeug entwickelt, welches dem Anwender der empirischen Trans-
port-Gleichungen ermöglicht einzuschätzen, ob auf Basis der lokalen Randbedingungen die An-
wendung der Gleichungen eher eine Unterschätzung oder Überschätzung des tatsächlich
transportursächlichen Tsunamis erwarten lässt.

Für die physikalischen Modellversuche wurde der größte wellentransportierte Block auf der Insel
Bonaire ausgewählt, um insbesondere den Einfluss der Form des Blocks, aber auch der Küsten-
form, des Eintauchgrads des Blocks im Wasser sowie der Block-Ausrichtung auf Transportpro-
zesse zu untersuchen und deren Effekt auf die Anwendung von empirischen Block-Transport-Glei-
chungen zu beurteilen. Dazu wurde dieser Boulder photogrammetrisch erfasst und in einer
Skalierung von 1:50 in physikalischen Modellversuchen im direkten Vergleich mit idealisierten
Block-Modellen untersucht. Aus den Ergebnissen dieser Modellversuche geht klar hervor, dass die
Berücksichtigung der Form in Berechnungen zum Zusammenhang zwischen Boulder-Transport
und notwendiger Tsunamiintensität unerlässlich ist. Gerade stromlinienförmige Boulder werden
erst später durch die angreifende Wellenenergie in Bewegung versetzt und im anschließenden
Transportprozess nur über kürzere Distanzen verlagert.

Abschließend wurde auf Basis eines vorhandenen numerischen Zwei-Phasen-Modells zur Wellen-
erzeugung durch Hangrutschungen die erste Version eines numerischen Block-Transport-Modells
entwickelt, welches zum einen die Signifikanz der Blockform und zum anderen den Einfluss von
Sedimentfracht in der Welle auf den Transport belegen soll. Erste Ergebnisse aus der Anwendung
der aktuellen Modellversion belegen die Wichtigkeit der Form und dessen Ausrichtung zum
Tsunami und deuten darauf hin, dass eine höhere mittlere Dichte der angreifenden Welle einen
transportfördernden Effekt hat.

II
Abstract

Abstract

On many coasts around the world, tsunamis pose a serious threat to the population, causing direct
damage (e.g., drowning deaths) and indirect damage (e.g., damage to infrastructures and impair-
ment of drinking water supplies). Recent tsunamis, such as the 2011 Tōhoku-Oki tsunami in Japan
and the 2004 Indian Ocean tsunami in Thailand, have shown that there is a lack of sufficient un-
derstanding of the specific local tsunami hazard in many regions.

Thus, although geological evidence of high-energy tsunamis existed prior to 2011, tsunami man-
agement plans and mitigation measures in many regions of Japan were based only on the strong-
est events of the past 100-200 years and were, therefore, inadequate to cope with the effects of
the Tōhoku-Oki tsunami. However, because the available geologic deposits already indicated such
intense tsunamis, Japan’s authorities decided to explicitly consider geologic evidence of undocu-
mented historical and prehistoric tsunamis (paleo-tsunamis) in the design of mitigation measures
and in management plans, in the aftermath of the 2011 tsunami disaster in Japan. Consideration
of such geological evidence has the benefit of providing information on tsunamis that have not
been recorded or observed and have long-standing occurrence intervals or very low probabilities
of occurrence (recurrence intervals of 1000 years or more). Thus, the analysis of geological de-
posits allows a coast-specific estimation of whether tsunami events can occur which exceed pre-
vious observations or estimates. Thus, risk mitigation strategies and management plans can be
adapted accordingly.

In this dissertation, the analysis of geological deposits is applied to the Caribbean Sea and, in
particular, the island of Bonaire (ABC Islands, Lesser Antilles). While for the Caribbean Sea a large
number of tsunami observations from the past approximately 520 years exists, of which only a
fraction is verified, no concrete records are available for Bonaire and the entire ABC islands. On
Bonaire in particular, however, there is a large number of geological deposits that indicate high-
energy wave events. The study of the possibly tsunami-induced transport of some of these deposits
on Bonaire by experimental and numerical models forms the key aspect of the present dissertation
and is accompanied by a holistic analysis of the tsunami hazard in the Caribbean region.

In this context, the first step was to evaluate the inventory of tsunami deposits throughout the Car-
ibbean Sea and attempt to assign these deposits to the respective cause of displacement (storm
waves or tsunamis). For this purpose, numerical tsunami simulations were performed for the Car-
ibbean Sea based on verified and assumed earthquake scenarios as tsunami triggers. In the
course of these simulations, tsunami heights were calculated along all coastlines in the Caribbean
Sea, but special attention was paid to the Boka Olivia and Spelonk areas along the northern coast
of Bonaire.

Comparing the simulated tsunamis with results from empirical block transport equations, it became
clear that the simulated tsunamis did not produce wave intensities sufficient to relocate the largest
boulder on the island of Bonaire. The main identified reasons for this discrepancy depict insufficient
accuracy of the empirical transport equations (e.g., as a result of oversimplifications) and inade-
quate selection or reproduction of earthquake scenarios.

Under the impression of these results and in order to assess the accuracy of empirical boulder
transport equations, a comprehensive analysis of published physical experiments on tsunami-in-
duced block transport was carried out and their weaknesses were identified. As a major weakness
the neglect of boulder shapes beyond idealized cubes or parallelepipeds was identified, which was,

III
Abstract

therefore, chosen as the core element to be investigated in the subsequent physical experiments.
During this analysis, it also became clear that the currently available and widely used empirical
equations for the description of block transport by tsunamis are not suitable for determining exact
values of possible tsunami intensities. Instead, it is recommended to work with probability ranges
for boulder mobilization and thus to limit public statements also to ranges of possible tsunami in-
tensities. Another element of the study was the development of a tool that allows the user of the
empirical transport equations to assess whether the application of the equations tends to underes-
timate or overestimate the actual transport-causing tsunami, taking into account local boundary
conditions.

For the physical model experiments, the largest wave-transported boulder on the island of Bonaire
was selected for extensive investigations, especially with respect to the influence of the shape, but
also considering the influence of the coastal shape, the submergence of the boulder as well as the
boulder orientation on transport processes and to assess their effect on the application of empirical
boulder transport equations. For this purpose, this boulder was photogrammetrically recorded and
investigated at a scale of 1:50 in physical experiments and compared to idealized block models.
Based on the experimental results, it could be shown that the consideration of shape in calculations
of the relationship between boulder transport and necessary tsunami intensity is essential. Com-
pared to idealized blocks, streamline-shaped blocks are later mobilized by the impacting wave
energy and are displaced over shorter distances in the subsequent transport process. Further in-
vestigations included experiments with different coastal models, boulder submergence as well as
their orientation to the wave.

Finally, based on an existing numerical two-phase flow model aiming on the wave generation by
landslides, the first version of a numerical boulder transport model was developed, which should
prove the significance of the block shape on the one hand and the influence of sediment load in
the wave on the transport on the other hand. First results from the application of the current model
version prove the importance of the shape and its orientation to the tsunami and indicate that a
higher mean density of the attacking wave has a transport-promoting effect.

IV
Contents

Contents

Kurzfassung .................................................................................................................................... I

Abstract ......................................................................................................................................... III

1 Introduction ......................................................................................................................... 1

1.1 Motivation .............................................................................................................................. 1

1.2 Background ........................................................................................................................... 4

1.2.1 Tsunamis in the Caribbean Sea and at the island of Bonaire ....................................4

1.2.2 Boulders as tsunami proxies ......................................................................................6

1.2.3 Research demands ....................................................................................................8

1.3 Objectives ............................................................................................................................. 9

1.4 Methods .............................................................................................................................. 10


1.5 Declaration of personal contribution ................................................................................... 12

2 Tsunami deposits of the Caribbean – Towards an improved coastal hazard


assessment ........................................................................................................................ 15

2.1 Introduction ......................................................................................................................... 16

2.2 Tsunamis and their onshore deposits ................................................................................. 20

2.2.1 Fine-grained tsunami deposits ................................................................................ 21

2.2.2 Coarse-clast tsunami deposits ................................................................................ 21

2.2.3 The challenges of dating event deposits ................................................................. 22

2.2.4 Deposits of recent extreme-wave events in the Caribbean ..................................... 23

2.3 Physical setting and tsunami triggers of the Caribbean ...................................................... 24


2.3.1 Triggers of tectonic tsunamis ................................................................................... 24

2.3.2 Volcanism, volcanic edifice failure and landslides................................................... 27

2.3.3 Teletsunamis ........................................................................................................... 28


2.4 Alternative processes causing elevated coastal deposits in the Caribbean ....................... 28

2.4.1 Hurricanes ............................................................................................................... 28

2.4.2 Relative sea-level (RSL) changes ...................................................................... 29


2.5 Tsunami deposits of the Caribbean .................................................................................... 29

2.5.1 NW sector ................................................................................................................ 30

2.5.2 N sector ................................................................................................................... 30


2.5.3 E sector .................................................................................................................... 35

2.5.4 S sector .................................................................................................................... 35

2.6 Discussion ........................................................................................................................... 39


2.6.1 Review of extreme-wave deposits ........................................................................... 39

2.6.2 Correlation of tsunami deposits ............................................................................... 42

V
Contents

2.6.3 Quantification of tsunami characteristics based on their deposits ........................... 43

2.6.4 Implications for tsunami hazard assessment ........................................................... 45

2.7 Conclusions ......................................................................................................................... 45

3 Enhanced field observation based physical and numerical modelling of tsunami


induced boulder transport – Phase 1: Physical experiments....................................... 48

3.1 Introduction .......................................................................................................................... 48

3.2 Boulder transport modelling ................................................................................................ 49

3.2.1 Hydrodynamic forces ............................................................................................... 49

3.2.2 Body characteristics ................................................................................................. 50

3.2.3 Environment ............................................................................................................. 50

3.3 Physical boulder transport modelling .................................................................................. 51


3.4 Case study “Island of Bonaire” ............................................................................................ 52

3.5 Physical experiments .......................................................................................................... 54

3.6 Phase 2: Numerical model .................................................................................................. 57

3.7 Summary ............................................................................................................................. 58

4 Experiments on tsunami induced boulder transport – a review .................................. 59

4.1 Introduction .......................................................................................................................... 59

4.1.1 Transport of solid objects by high-energy wave events .......................................... 61

4.1.2 Reviewed publications ............................................................................................. 62

4.2 Classification of experimental studies on tsunami induced transport of bodies.................. 64

4.3 Boulder transport experiments ............................................................................................ 68

4.3.1 General .................................................................................................................... 68

4.3.2 Wave generation mechanism .................................................................................. 68


4.3.3 Experimental focus .................................................................................................. 70

4.4 Experimental investigations on incipient motion ................................................................. 70

4.5 Experimental investigations on tsunami-induced boulder transport ................................... 73


4.6 Model comparability ............................................................................................................ 76

4.6.1 Boulder density and buoyancy................................................................................. 77

4.6.2 Model scale .............................................................................................................. 77

4.6.3 General experimental setup ..................................................................................... 79

4.7 Discussion ........................................................................................................................... 80

4.7.1 Incipient motion ........................................................................................................ 80


4.7.2 Transportation mode ................................................................................................ 80

4.7.3 Reported variabilities ............................................................................................... 81

4.7.4 Parameter influence ................................................................................................. 82

VI
Contents

4.7.5 Assessment of the Flatness Index .......................................................................... 83

4.7.6 Considerations on wave type, impact force and probability .................................... 83

4.7.7 Attempt of a guidance tool for mobilization assessment ......................................... 85

4.7.8 Attempt for a standard case for single-boulder experiments................................... 87

4.8 Summary and outlook ......................................................................................................... 88

5 Experimental models of coarse-clast transport by tsunamis ...................................... 90

5.1 Introduction ......................................................................................................................... 90

5.2 Dimensionless quantities and scaling of experiments ........................................................ 91

5.2.1 Dimensional analysis ............................................................................................... 91

5.2.2 The Froude number and scaling laws ..................................................................... 91

5.2.3 The Reynolds number ............................................................................................. 95

5.3 Measuring approaches in the wave tank ............................................................................ 97

5.4 Types of wave generation ................................................................................................... 99


5.5 Parameters studied in physical experiments .................................................................... 102

5.6 Published wave-tank experiments on tsunami-boulder transport ..................................... 103

5.6.1 Experimental setups .............................................................................................. 103

5.6.2 Key findings ........................................................................................................... 106

5.6.3 Further related studies ........................................................................................... 107

5.7 Link to numerical models .................................................................................................. 109

5.8 Conclusions and recommendations .................................................................................. 110

6 Significance of boulder shape, shoreline configuration and pre-transport setting


for the transport of boulders by tsunamis ................................................................... 113

6.1 Introduction ....................................................................................................................... 113


6.2 Methods ............................................................................................................................ 114

6.2.1 Boulder models used in the experiments .............................................................. 114

6.2.2 Experimental setup ................................................................................................ 116


6.2.3 Video Processing ................................................................................................... 119

6.3 Results .............................................................................................................................. 120

6.3.1 Bore dynamics ....................................................................................................... 120

6.3.2 Boulder transport ................................................................................................... 121

6.4 Discussion ......................................................................................................................... 129

6.4.1 Parameter influence hierarchy............................................................................... 129


6.4.2 Experimental sensitivity ......................................................................................... 130

6.4.3 Influence of shore type .......................................................................................... 130

6.4.4 Initial boulder alignment to the flow ....................................................................... 131

VII
Contents

6.4.5 Boulder shape ........................................................................................................ 132

6.4.6 General observations ............................................................................................. 132

6.5 Future directions and numerical investigations ................................................................. 134

6.6 Conclusions ....................................................................................................................... 135

7 Numerical Boulder Transport Model ............................................................................. 137

7.1 Introduction ........................................................................................................................ 137

7.2 Immersed Boundary Approach ......................................................................................... 140

7.3 General approach.............................................................................................................. 141

7.4 Boundary handling ............................................................................................................ 141

7.4.1 Boundary detection ................................................................................................ 141

7.4.2 Numerical slope ..................................................................................................... 147

7.4.3 Force calculation .................................................................................................... 148

7.5 Calculation of movement ................................................................................................... 151


7.5.1 General .................................................................................................................. 151

7.5.2 Translation ............................................................................................................. 152

7.5.3 Rotation .................................................................................................................. 153

7.5.4 Update of the flow field .......................................................................................... 154

7.6 Simulations ........................................................................................................................ 155

7.6.1 Setup ...................................................................................................................... 155

7.6.2 Investigation on boulder shape (shape a.1 and a.2) ............................................. 156

7.6.3 Investigation on sediment load (shape b.1) ........................................................... 160

7.6.4 Rotation calculations (shapes c.1 and c.2) ............................................................ 165


7.7 Discussion and outlook ..................................................................................................... 169

7.7.1 Results ................................................................................................................... 169

7.7.2 Limitations and outlook .......................................................................................... 169


8 Synthesis and outlook .................................................................................................... 171

8.1 Summary ........................................................................................................................... 171

8.2 Assessment of the research results .................................................................................. 172


8.3 Outlook and future directions in boulder transport research ............................................. 175

References ................................................................................................................................. 177

Appendix .................................................................................................................................... 200


A. Appendix related to Tsunami deposits of the Caribbean – Towards an improved coastal
hazard assessment ........................................................................................................... 200

Supplements .............................................................................................................................. 217

B. Supplements related to Experiments on tsunami induced boulder transport – a review .. 217

VIII
Contents

C. Supplements related to Significance of boulder shape, shoreline configuration and pre-


transport setting for the transport of boulders by tsunamis............................................... 221

Acknowledgments..................................................................................................................... 231

Curriculum Vitae ........................................................................................................................ 232

List of Publications ................................................................................................................... 233

IX
X
List of Figures

List of Figures

Figure 1-1: Simplified sketch of the main transport processes of coarse clasts and fine
sediments during tsunami impact, run-up, backwash and subsequent processes.
.................................................................................................................................... 3
Figure 1-2: Reported tsunami events along the Caribbean Sea coasts (verified and non-
verified reports; NOAA, 2021, modified; basemap: Google Satellite Basemap,
mt1.google.com). Note: Not all events listed in the NOAA Hazards Viewer are
visible. ......................................................................................................................... 4
Figure 1-3: BOL2 on the island of Bonaire (V ≈ 77 m³). The approximated position on
Bonaire is indicated by the star (underlying image by CNES/Airbus/google,
2021). .......................................................................................................................... 6
Figure 1-4: Schematic outline of the thesis from the universal look on tsunamis in the
Caribbean Sea to specific investigations on the hydrodynamics of a single
boulder. The final numerical model and outlook lead back to more universal
investigations. ........................................................................................................... 11
Figure 1-5: a.1 & a.2: Computed water displacement on a plane grid after the trigger model
of Okada (1985). b.1 & b.2 Superelevated illustration of displacement (100:1)
compared to surrounding bathymetry with water level zero (own calculation and
representation). ......................................................................................................... 11
Figure 2-1: Overview map of the Caribbean region .................................................................... 16
Figure 2-2: a) Frequency-size distribution for reported heights of tsunamis vs. number per
year. .......................................................................................................................... 18
Figure 2-3: Overview of the Caribbean basin separated into five sectors (dashed lines). ......... 18
Figure 2-4: The role of palaeotsunami research in a tsunami hazard management process
and schematic depiction and classification of sedimentary archives relevant to
the Caribbean (see also Figure 2-5). ........................................................................ 19
Figure 2-5: Potential sedimentary archives of tsunami- or storm-induced deposits or
landforms in the Caribbean (see categories in Figure 2-4) as exemplified in the
Lac Bai area of southeast Bonaire (Figure 2-1)........................................................ 20
Figure 2-6: Tsunami scenario from a rupture along the western Muertos Thrust Belt (MTB)
and maximum heights along circum-Caribbean coasts. For earthquake
parameters see Table 2-1. ........................................................................................ 26
Figure 2-7: Tsunami scenario from a rupture along the South Caribbean Deformed Belt
(SCDB) ...................................................................................................................... 27
Figure 2-8: Examples of candidate tsunami deposits from Anegada (Sites 23/47, Figure 2-3)
and Bonaire (Site 31/55, Figure 2-3) ........................................................................ 34
Figure 2-9: Examples of storm-induced deposits and landforms from Bonaire (Site 55,
Figure 2-3) (compare with archive categories A–F in Figure 2-4). ....................... 38
Figure 2-10: Compilation of dated candidate tsunami deposits of Holocene age. ................ 40
Figure 3-1: Basic impact forces related to hydrodynamics. FM: Inertia; FD: Drag; FA: Lift;
FG: Gravity; FR: Friction (Pignatelli et al., 2009). ..................................................... 50
Figure 3-2: Boulder transportation modes (a) and initial positions relative to water level (b)
which are commonly considered in numerical models. ............................................ 51
Figure 3-3: Location of Bonaire and the earthquale scenarios (left). Location of the study
areas on Bonaire and the largest boulder BOL2 at Boka Olivia (right). ................... 52
Figure 3-4: Model bathymetry near to Bonaires shoreline. ......................................................... 53
Figure 3-5: Computed wave heights at Boka Olivia. ................................................................... 54

XI
List of Figures

Figure 3-6: Left: Comparison of two test runs for bore generation in the IWW test flume.
Right: Photo of a generated tsunami bore (Peltzer, 2015). ...................................... 55
Figure 3-7: Principal test setup of the flume (not to scale). ......................................................... 55
Figure 3-8: a) Perspective view and mesh model of a boulder deposited on Bonaire. b)
Comparison between the original photograph and the numerical model). c)
Contrasting shapes of idealized and complex-shaped boulders (a, b, c = main
axes of boulders)....................................................................................................... 56
Figure 3-9: Printing process of the bottom part of a boulder replica in 1:50. .............................. 56
Figure 3-10: CAD model of the Sensor Fish Gen 2 device (Deng et al., 2014). ........................... 57
Figure 3-11: Shore replicas used for the physical experiments. ................................................... 57
Figure 4-1: Basic forces acting on a solid body at rest. .............................................................. 62
Figure 4-2: Forces acting on objects by wave impact. ................................................................ 62
Figure 4-3: Definition of the boulder-wave rBW [-] ratio. ............................................................... 64
Figure 4-4: Simplified sketches of commonly applied wave-generation techniques. .................. 70
Figure 4-5: Experimental setup of Harry et al. (2019). ................................................................ 70
Figure 4-6: Size comparison of the considered physical experiments on tsunami induced
boulder transport. ...................................................................................................... 77
Figure 4-7: Size comparison of boulder models in the reviewed studies on tsunami-induced
boulder transport. ...................................................................................................... 77
Figure 4-8: a) Ratios between the boulder impact areas and Froude number of selected
publications. .............................................................................................................. 79
Figure 4-9: Proposal for a supporting tool for transport estimation. As outcome, the diagram
indicates if increased or shortened transport distances can be expected
compared to a standard case ................................................................................... 86
Figure 5-1: Streamlines around a smooth sphere by different Reynolds numbers. .................... 96
Figure 5-2: Boulder tracking algorithm developed in MATLAB. .................................................. 98
Figure 5-3: Dam break and piston mechanism for bore generation. ......................................... 100
Figure 5-4: Pumping-type wave maker ..................................................................................... 101
Figure 5-5: Typical boulder orientation and submergence in physical experiments. ................ 102
Figure 5-6: Comparison of the experimental flumes from Imamura et al. (2008), Nandasena
and Tanaka (2013), Liu et al. (2015) and Bressan et al. (2018). ............................ 104
Figure 5-7: Comparison of the largest applied boulder from Imamura et al. (2008),
Nandasena and Tanaka (2013), Liu et al. (2015) and Bressan et al. (2018). ........ 104
Figure 5-8: Graphical comparison of transport distances and flow velocities in Liu et al.
(2015), Nandasena and Tanaka (2013) and Imamura et al. (2008) for scaled and
unscaled results (Table 5-4). .................................................................................. 107
Figure 5-9: Difference of maximum transport distance and total transport distance for an
initially partially-submerged boulder setup (t1). ...................................................... 111
Figure 6-1: (A) Simplified geological map of the island of Bonaire showing the elevated
‘Lower Terrace’ unit (B, C) that forms a quasi-stepped cliff coast and (D)
represents the location of the complex-shaped boulder BOL 2 .............................. 115
Figure 6-2: The three applied boulder models. From left to right: regular flat boulder, regular
cuboid boulder and complex boulder. ..................................................................... 116
Figure 6-3: Experimental setup ................................................................................................. 118

XII
List of Figures

Figure 6-4: Top: Applied shore types. (a) Type 1 with a uniform inclination; (b) Type 2 with
two different inclinations; (c) Type 3, stepped shore resembling the shore of the
type site of BOL 2 on Bonaire (Figure 6-1). Bottom: initial boulder setups. (d)
Initial alignment to the flow: 90°, long axis perpendicular to the flow; 45°, long
axis at a 45° angle to the flow; 0°, short boulder axis perpendicular to the flow.
(e) Pre-transport positions in relation to water level. .............................................. 119
Figure 6-5: Wave profile 1.5m in front of the shore and at shore tip on shore Type 2 and 17.2
cm bore height. Idealized tsunami profile (natural wave form) after Pedersen and
Gjevik (1983)........................................................................................................... 120
Figure 6-6: First 0.33 s of boulder movement in subaerial conditions (shore Type 1). ............. 124
Figure 6-7: Boulder movement between 0.5 s and 1.9 s after wave impact in subaerial
conditions (shore Type 1). ...................................................................................... 125
Figure 6-8: Transport over time for each tested boulder shape. ............................................... 126
Figure 6-9: (a) Time-distance graph for the subaerial case at shore Type 2. (b) Time-
distance graph for the partially-submerged case at shore Type 2. In both
diagrams, the point of the changing shore inclination is marked by a dotted red
line. ......................................................................................................................... 126
Figure 6-10: Comparison of the total transport distances grouped following the boulder shape
(complex and regular cuboid), initial boulder alignment and submergence on
shore Type 2. .......................................................................................................... 128
Figure 6-11: Comparison of the total transport distances grouped following the boulder shape
(complex and regular cuboid), initial boulder alignment and submergence on
shore Type 3. .......................................................................................................... 128
Figure 6-12: Movement-hampering influence of the stepped shore with partially-submerged
setup (conceptual drawing)..................................................................................... 129
Figure 6-13: Comparison between the simplified horizontal flow fields of (a) Nandasena and
Tanaka (2013), (b) the regular cuboid boulder, (c) the complex boulder. .............. 133
Figure 7-1: Immersed Boundary (IBM) setup. ........................................................................... 140
Figure 7-2 : a) General functionality of the immersed boundary. b) Basic properties
describing the local boundary geometry. ................................................................ 140
Figure 7-3: Exemplary angle base cell assignment. ................................................................. 143
Figure 7-4: Graphical example for the boundary detection algorithm. The numbers follow the
condition 1 to 7. Note: Shown is a simple example in which no exceptions (e.g.,
aligned boundary nodes) occur. ............................................................................. 144
Figure 7-5: Graphical explanations of intercepted exceptions within the boundary detection
algorithm. ................................................................................................................ 147
Figure 7-6: a) The initially plain grid. b) The geometrical layers describing the height and
shape of the IBM. c) Interpolation of the grid elevation. d) Calculation of the
numerical slopes in x and y direction. ..................................................................... 147
Figure 7-7: Exemplary shift of the cut cell for the velocity interpolation on an IBM segment. .. 149
Figure 7-8: Exemplary behaviour of 𝛽, 𝜁 and 𝜃 for the calculation of 𝐹𝐷, 𝑥. ............................. 151
Figure 7-9: Treatment of the flow field around the IBM. ............................................................ 154
Figure 7-10: Standard simulation setup for the numerical investigations. .................................. 156
Figure 7-11: Three-dimensional plots for clarification of the force distribution. .......................... 156
Figure 7-12: First three simulation seconds of a.1 and a.2. ........................................................ 158
Figure 7-13: Further wave development in the a.1 and a.2 setup. The transport distance of
a.2 is significantly above a.1. .................................................................................. 158

XIII
List of Figures

Figure 7-14: Seconds 10 to 13 in the a.1 and a.2 setups. a.1 is remobilized (seconds 11 to
13) due to the reflected wave. ................................................................................. 159
Figure 7-15: Development of the absolute velocities of a.1 and a.2 over the whole simulation
time. ........................................................................................................................ 159
Figure 7-16: Force distribution on a.1 and a.2 within the first 10 simulation seconds. ............... 160
Figure 7-17: Development of the absolute boulder velocity in the b.1 setup. ............................. 161
Figure 7-18: Wave propagation (a) and density distribution (b) in the b.1 setup. ....................... 162
Figure 7-19: Further wave propagtion (a) and density distribution (b) in the b.1 setup. ............. 163
Figure 7-20: Seconds 10 to 13 in the b.1 setup (a: total wave height; b: density distribution). ... 164
Figure 7-21: Force distribution on the boulder within the first 10 seconds of the b.1 setup. ....... 165
Figure 7-22: Exemplary rotation behaviour in the boulder transport model. ............................... 166
Figure 7-23: Force distribution on shapes c.1 and c.2 within the first 4.5 seconds. At second
4.5 an uneven force load occurs at the leading face of c.1 and introduces rotation
(Figure 7-24). .......................................................................................................... 167
Figure 7-24: Seconds 4.5 and 5 of the simulation for shapes c.1 and c.2. For c.1, the induced
rotation from the uneven force load (Figure 7-23) is observable. c.2 is still in
positive rotation. ...................................................................................................... 168
Figure 7-25: Seconds 5.5 to 13 for the simulation of c.2. ............................................................ 168
Figure 7-26: Rough sketch for a physical small-scale validation dam-break setup for the BTM.
The sediment is added immediately before releasing the dam-break. ................... 170
Figure B-1: Sketch of a possible dynamic drag force experiment. a) Initial setup.
Structure/body in origin position. b) Controlled ac-celeration of the structure
during wave impact and measuring of the force development. .............................. 219
Figure B-2: Example of using the proposed decision supporting diagram. ............................... 220
Figure C-1: Repeated wave generation a: comparison of wave velocity. b: comparison of
wave-height. ............................................................................................................ 221
Figure C-2: Wave profile on shore Type 1. Measured 1.5 m in front of the shore tip. ............... 222
Figure C-3: Distances between shore-tip and boulder models depending on initial
submergence and boulder type. ............................................................................. 222
Figure C-4: Distribution graphs for the transport distances on shore-type 1. ............................ 223
Figure C-5: Boxplot for the maximum transport distances on shore Type 1. ............................ 225
Figure C-6: Optical comparison between the transport process of the complex and regular
cuboid boulder on shore Type 2, partially-submerged and 90° initial alignment
(representative experimental run).. ......................................................................... 226
Figure C-7: Optical comparison between the transport process of the complex and regular
cuboid boulder on shore Type 2, subaerial and 90° initial alignment
(representative experimental run).. ......................................................................... 227
Figure C-8: Total transport distances on shore Type 2 grouped for boulder shape based on
131 experiments (black: complex; red: regular cuboid boulder). Only
unidirectional flow and no backwash is considered. ............................................... 228
Figure C-9: Wave profiles for all generated waves on shore Type 3. ........................................ 229
Figure C-10: Initial boulder positions and water-level for the experiments
on the stepped shore. ............................................................................................. 229
Figure C-11: Total transport distances on shore-type 3 grouped for boulder (black:
complex; red: regular cuboid boulder). ................................................................... 230

XIV
List of Tables

List of Tables

Table 1-1: Overview of existing models for tsunami-induced boulder transport (partly taken
from Lodhi et al., 2020). .............................................................................................. 8
Table 1-2: Overview on the published articles and book chapters within this thesis. ................ 10
Table 2-1: Earthquake and fault parameters used for simulating pan-Caribbean tsunamis
originating at the Muertos Thrust Belt (MTB) and the South Caribbean Deformed
Belt (SCBD). ............................................................................................................. 26
Table 3-1: Parameters of the simulated tsunami scenarios. ...................................................... 52
Table 4-1: Terms used across the presented review. Some terms might differ from those in
the original publications. ........................................................................................... 61
Table 4-2: Grouping of the reviewed studies. The parameters initial alignment, scaling and
study scope are not presented to maintain readability. ............................................ 63
Table 4-3: Considered publications on tsunami-induced boulder transport and their main
characteristics (part a). ............................................................................................. 65
Table 4-4: Considered publications on tsunami-induced boulder transport and their main
characteristics (part b). ............................................................................................. 66
Table 4-5: Review publications which are significantly based on empirical or numerical
models. ..................................................................................................................... 67
Table 4-6: Overview on the spreading of transport distance between individual experimental
runs of tsunami-induced boulder transport. .............................................................. 81
Table 4-7: Basic set of parameters which need to be standardized for a common “standard
setup” for studies of single-boulder transport. .......................................................... 87
Table 5-1: Froude scaling laws .................................................................................................. 94
Table 5-2: Examples for parameters of interest in physical experiments and their influence.
A qualitative evaluation is given on the necessary effort to implement parameter
changes in the model setup. ................................................................................... 103
Table 5-3: Comparison of the experimental parameters from Imamura et al. (2008),
Nandasena and Tanaka (2013), Liu et al. (2015) and Bressan et al. (2018). ........ 105
Table 5-4: Scaled comparison (using scaling laws in section 2) between Liu et al. (2015),
Nandasena and Tanaka (2013), Imamura et al. (2008) and Bressan et al. (2018).
No further parameter (boulder dimension, weight, etc.) is considered. If available,
the maximum velocity is taken for comparison. For Bressan et al. (2018), the
calculated average velocity is used. From every calculation, one boulder is taken
for comparison which is in a similar range for FI and density. As leading scaling
parameter, the a-axis is considered. ...................................................................... 106
Table 5-5: Recommendation for standard parameters to be published with every
experimental study. ................................................................................................. 111
Table 6-1: Overview of the experimental setup. Boulder shape depicts the type of applied
boulder (complex, regular cuboid, flat cuboid). ....................................................... 117
Table 6-2: Calculated coefficients of friction (μ [-]) for wet conditions depending on shore
inclination (3.8°, 7.6°, 11.3°) and boulder type (flat cuboid, complex, regular
cuboid). ................................................................................................................... 119
Table 6-3: Time lag between bore impact and the initiation of boulder transport, averaged
over all experimental runs. ...................................................................................... 121
Table 6-4: Statistical results for the experiments on shore Type 1. ......................................... 122
Table 6-5: Transport distances grouped according to boulder shape and submergence
(shore Type 1)......................................................................................................... 122

XV
List of Tables

Table 6-6: Effect of boulder shape and submergence on the total transport distance in
percentages (shore Type 1). ................................................................................... 123
Table 7-1: Used symbols in Chapter 7. .................................................................................... 139
Table 7-2: Applied boulder shapes and in the presented simulations. .................................... 155
Table A-1: Characteristics of sand-dominated tsunami and storm deposits as inferred from
recent and selected historical and palaeo-events................................................... 200
Table A-2: Characteristics of coarse-clast tsunami and storm deposits as inferred from
recent and selected historical and palaeo-events................................................... 206
Table A-3: Sediment-based reconstructions of palaeoenvironments and coastal processes
reviewed for evidence of extreme-wave events...................................................... 208
Table A-4: Coastal coarse-clast deposits reviewed for evidence of extreme-wave events. .... 213
Table B-1: Studies on transportation of artificial bodies by high-energy wave events (part a).
................................................................................................................................ 217
Table B-2: Studies on transportation of artificial bodies by high-energy wave events (part b).
................................................................................................................................ 218
Table B-3: Varied parameter in the considered experimental studies on tsunami induced
boulder transport. .................................................................................................... 219
Table C-1: Overview for the conducted experiments on shore Type 1. ................................... 221
Table C-2: Statistical results for the experiments on shore Type 1. ......................................... 224

XVI
Introduction

1 Introduction

1.1 Motivation

Tsunamis are a serious threat to coastal communities around the world, thus, influencing the plan-
ning of coastal infrastructure and prompting site-specific hazard management strategies. The 2004
Indian Ocean Tsunami (IOT) with about 230,000 fatalities and the 2011 Tōhoku-Oki Tsunami (TOT)
with over 15,000 fatalities recently showed the importance of tsunami awareness and adequate
countermeasures and management plans. Beside direct and immediate damages on population
and (infra-)structure, tsunamis can cause indirect and medium- to long-term impairments to coastal
communities (e.g., destruction of power plants or salt-water intrusions in cultivated delta plains;
Villholth and Neupane 2011; Nakamura et al. 2017). While the tsunami awareness in periodically
tsunami affected areas (e.g., Japan) was yet considerably high before the IOT and TOT, both
events underlined the importance of improved estimations of possible tsunami impacts at coasts
around the world. The destructive impact of potential tsunamis increases further due to the ongoing
population growth in low-elevation coastal areas which is expected to increase between 68 % and
122 % until 2060 from the 2015 baseline (Neumann et al., 2015). Additionally, the mean sea-level
is estimated to rise between 0,17 m and 0,4 m until 2050 (Oppenheimer, 2019). Thus, earthquake
magnitudes which do not cause hazardous tsunami impact today might do so in the future due to
associated lower coastal elevation (Li et al., 2018).

Based on the experiences from the IOT and TOT, researchers from different fields (e.g., geogra-
phers, geologists, engineers, geophysicist) intensified the efforts in assessing the magnitude of
past tsunami events more accurately and identifying areas affected by tsunamis in the past, where
no reliable historical records are available. In Japan, for example, in tsunami mitigation plans, ge-
ological records were only sparsely considered for estimations of potential tsunami magnitudes
before the TOT. Instead, estimations for expectable tsunami magnitudes for Japan’s coastlines
were mainly based on evaluations of seismic activities and potential earthquake magnitudes with-
out systematic investigations of palaeotsunami occurrences and their magnitudes. Nevertheless,
in the aftermath of the TOT, the tsunami mitigation plans were substantially revised and the com-
mittee for Technical Investigation on Countermeasures for Earthquakes and Tsunamis based on
the lessons learned from the “2011 off the Pacific coast of Tohoku Earthquake was established. In
the final report of the committee (CDMC, 2011), the utilization of geological tsunami records (geo-
logical deposits) is now explicitly accounted as additional basis for the estimation of possible ex-
treme tsunami magnitudes (CDMC, 2011; Goto et al., 2014).

As an example, in connection with the TOT, the 869 Jogan tsunami is identified as a possible
predecessor of the TOT. Based on geological records, the inundation area of the Jogan tsunami is
estimated to have reached 3 to 4 km inland from the palaeo-shoreline (Sawai et al., 2007; Goto et
al., 2014), and analysis of the geological record of the Jogan tsunami indicated a necessary earth-
quake magnitude of > 8.3-8.4 Mw (Satake et al., 2008a; Namegaya et al., 2010; Sugawara et al.,
2011; Goto et al., 2014) which is revised to Mw > 8.6 based on enhanced process understanding
from the TOT (Namegaya and Satake, 2014). The conclusions drawn from geological records re-
lated to the Jogan tsunami, were, however, not taken into account in Japan’s tsunami mitigation
plans prior to the TOT. In consequence, the hazard management was not designed for an extreme

1
Introduction

event like the TOT and resulted in the known disastrous impact (e.g., the Fukushima nuclear inci-
dent) since such high tsunami magnitudes were simply not expected (e.g., Suzuki, 2012; Tanaka,
2013; Goff et al., 2014; Engel et al., 2020; Goto et al., 2021).

The above example clarifies that the identification of palaeotsunami impacts and estimations on
their magnitude and recurrence rates, can be a fundamental supporting part of tsunami hazard
management and mitigation plans (e.g., evacuation zones, breakwater design) and, possibly more
important, is able to build public awareness (Engel and Brückner, 2011). Especially in regions
lacking a reliable long-lasting database (like in parts of the Caribbean Sea), palaeotsunami re-
search can help to create awareness where tsunami threats are present. In such regions the threat
due to tsunamis might not be in the minds of governments or population due to the long recurrence
intervals associated to the highest tsunami magnitudes, which can span over multiple lifetimes (up
to over 1000 years; Minoura et al., 2001) and are not preserved by personal experience in the
collective memory. Furthermore, beside estimates on magnitude and timing of past events, analy-
sis of the geological record can also provide insights in the maximum inundated area and impact
direction (Weiss and Bourgeois, 2012; Switzer et al., 2014; Sugawara, 2021; Engel et al., 2020).

In the Caribbean Sea, tsunamis were not recognized as a serious threat for a long time due to the
overwhelming experiences from hazards resulting directly from other threats (earthquakes, river
floods, volcanism, hurricanes), which are more present in the daily life (Bilham, 2010; Fritz et al.,
2013; Engel et al., 2016). Nevertheless, historical reports and observations indicate that the Car-
ibbean Sea has a recognizable history of potential high-magnitude tsunami occurrences (O’Lough-
lin and Lander, 2003) for which no well-founded information on recurrence and magnitude are
available, though (Rowe et al., 2009; Engel et al., 2016). Furthermore, available literature on the
spatial distribution of tsunamis in the Caribbean Sea is mainly based on historical colonial reports,
which are considered to be incomplete (O’Laughlin and Lander, 2003; Engel et al., 2016). Beside
questionable tsunami recognition (due to sparse population density) and correct tsunami identifi-
cation, the record goes back to ca. 1500 AD only, and therefore does not span over typical recur-
rence intervals of the most extreme tsunami events (Jankaew et al., 2008; Brill et al., 2012; Sawai
et al., 2012; Engel et al., 2016). Here, investigations on the geological deposit record can help
improve the knowledge and awareness of expectable high-magnitude tsunami impacts and en-
gage, if necessary, preparation.

The investigation of tsunami deposits can roughly be divided into analysis of fine sediments and
coarse clasts (boulder; Figure 1-1). With regard to fine sediments, their datability, the possibility to
differentiate between individual events through the stratigraphic composition, and the wide
spectrum of proxies have to be considered as advantages compared to coarse clasts. However,
they include also disadvantages like their partially short-term preservation potential and the partly
difficult reachability, which may require drilling (Szczuciński, 2020). Potential coarse clast deposits
relocated by tsunamis, on the other hand, may not require drilling and are generally easier to find,
but may be subject to anthropogenic influences and weathering and might be transported multiple
times by multiple events, for example (Scheffers, 2021). The recognition of coarse clasts as
tsunami proxies, dates back to the nineteenth century (Neale, 1885) while their systematic
application for back-calculations of tsunamis or storm-surges is a comparably young discipline,
originating in the 1980’s (Bourrouilh-Le Jan and Talandier, 1985; Nakata and Kawana, 1995; Nott,
1997; Scheffers and Kelletat, 2003; Etienne et al., 2011; Engel et al., 2016). In boulder-related
palaeotsunami research, the location and physical characteristics of coastal boulder deposits are

2
Introduction

used for recalculating the necessary wave energy and the transport-process for their relocation
from their origin to the present location by using analytical or numerical models. During the back
calculation, several uncertainties need to be considered which potentially lead to over- or
underestimated wave energies. These uncertainties encompass parameters connected to the
boulder itself (e.g., altered weight or shape due to weathering, density; Kelletat et al., 2020), the
not-observed transport process (e.g., remobilization during the backwash and subsequent
underestimations of the maximum transport distance, or storm waves) and
inaccuracies/simplifications in the applied equations (e.g., simplified boulder shapes). While
uncertainties directly connected to the boulder can possibly be tackled by knowledge on general
chemical or physical processes, for the latter exact knowledge on the boulder transport process is
necessary in order to gain accurate mathematical descriptions (i.e., analytical or numerical models)
of the boulder behaviour during tsunami impact. A simplified sketch of the transport processes of
sediments and coarse-clasts during and after tsunamis is shown in Figure 1-1.

Figure 1-1: Simplified sketch of the main transport processes of coarse-clasts and fine sediments during tsunami impact,
run-up, backwash and subsequent processes. Whether a particular process takes place, and the eventual nature of
resulting tsunami deposits heavily depends on the local conditions (e.g., sediment sources, coastal topography, pedo-
genetic processes, climate, population). Note: H denotes the tsunami height immediately before breaking and bore
height during run-up and v the corresponding velocities.

From the first boulder transport model of Nott (1997) to today’s models (e.g., Bressan et al., 2018;
Lodhi et al., 2020), their complexity and accuracy significantly increased due to additionally in-
cluded parameters (e.g., impact force, turbulence) and considered boundary conditions (e.g., in
terms of pre-transport setting or transport mode). However, currently available (in particular ana-
lytical) models tend to overestimate the necessary wave energy for the mobilization and transport

3
Introduction

of coarse clasts (e.g., Bressan et al., 2018; Lodhi et al., 2020) and further research is needed for
improving the knowledge on tsunami-boulder interaction and the main influencing parameters (e.g.,
pre-transport setting, boulder characteristics).

Therefore, the present PhD thesis aims at increasing the knowledge of possible tsunami threats in
the Caribbean Sea and at the island of Bonaire and elsewhere by means of numerical tsunami
simulations and enhancing the understanding of tsunami induced boulder transport processes by
performing experimental and numerical investigations.

1.2 Background

1.2.1 Tsunamis in the Caribbean Sea and at the island of Bonaire

Harbitz et al. (2012) list 85 extreme wave events in the Caribbean Sea from 1498 to 2006 which
might have been tsunamis. However, the authors consider only 35 events among them to be veri-
fied tsunamis and 21 as having “probably“ happened. The remaining 29 events are considered to
be either questionable or very doubtful reports (Harbitz et al., 2012). Figure 1-2 shows (possible)
tsunami events in the Caribbean Sea as reported by the National Oceanic and Atmospheric Ad-
ministration (NOAA) Natural Hazards Viewer. In July 2021, a total of 104 tsunamis (including po-
tential ones) are reported. The oldest record from 1498 refers to a possible tsunami impact at Boca
de la Sierpe (Venezuela) while the first verified tsunami occurred in 1530 at Cumana (Venezuela).
The youngest tsunami occurrence is reported from 2020 off the coast of Cuba with only small run-
ups of approximately 0.1 m. As can be seen from Figure 1-2 several events in the southern Carib-
bean Sea are reported along the coast of Venezuela while no records exist for the ABC Islands
(Aruba, Curacao, Bonaire; Lesser Antilles) over the whole time-range.

Figure 1-2: Reported tsunami events along the Caribbean Sea coasts (verified and non-verified reports; NOAA, 2021,
modified; basemap: Google Satellite Basemap, mt1.google.com). Note: Not all events listed in the NOAA Hazards
Viewer are visible.

Coarse-clast deposits on the island of Bonaire are subject of research for nearly two decades and
have been investigated by several researchers (e.g., Scheffers, 2002; Morton et al., 2008; Spiske
et al., 2008; Engel and May, 2012). There are diverging conclusions and uncertainty regarding the

4
Introduction

exact process of dislocation for the larger deposits on Bonaire. While Spiske et al. (2008) assume
that the deposits have been dislocated by strong wind waves (e.g., during hurricanes), in Scheffers
(2002) and Scheffers et al. (2006) tsunamis are favoured as transport causing event, whereas
Pignatelli et al. (2010) inferred combinations of strong storm waves and several tsunamis. Watt et
al. (2010) consider that the largest boulder on Bonaire (BOL2, see following paragraphs) exhibits
characteristics indicating tsunami as possible cause for the dislocation while they assume storm-
wave driven transport for almost all other deposits.

As last major study focusing on high-energy wave deposits on Bonaire, Engel and May (2012)
applied advanced measurement techniques for calculating boulder characteristics (density, dimen-
sions/volume) and enhanced versions of the analytical models of Pignatelli et al. (2009) and Nan-
dasena et al. (2011a). The findings of Engel and May (2012) support the earlier stated hypothesis
of Scheffers (2002b) who claimed that tsunamis must have been the cause for the dislocation of
the larger deposits with volumes > 10 m³. Furthermore, results of the analytical models revealed
necessary tsunami heights as are reported to have impacted the coast of Venezuela before, sup-
porting the assumption of tsunami induced boulder transport on Bonaire (Engel and May, 2012).
However, since until today no tsunamis were observed on the island of Bonaire, possibly due to
sparse population and assumed tsunami recurrence intervals of 1000s of years (Engel et al., 2013),
uncertainties in analytical models and boulder characterization (e.g., material homogeneity), and
ambiguous tsunami source, the question of tsunami occurrences is not finally solved, yet.

In Engel and May (2012), a set of coarse boulder deposits is identified as potentially relocated by
tsunamis. The boulders encompass volumes between < 1 m³ to approximately 70 m³ with corre-
sponding weights ranging from 1 t to 150 t. The boulder deposits are distributed over two main
locations on Bonaire (Spelonk and Boka Olivia; Chapter 6) while the two largest (denoted as BOL1
with volumes V = 54 m³ and BOL2 with V = 70 m³; later specified to 77 m³ by Oetjen et al., 2020a)
are located at Boka Olivia. For these two boulders, and depending on the applied model, Engel
and May (2012) derived necessary tsunami heights between HT, Pignatelli = 7.3 m and HT, Nan-
dasena = 10.1 m for BOL1 (54 m³) and HT, Pignatelli = 8.9 m to HT, Nandasena = 15.4 m for BOL2 (for the
70 m³ estimate). The corresponding storm wave heights HS of equal transport energy were calcu-
lated to ca. HS, Pignatelli = 29 m to HS, Nandasena = 41 m (BOL1) and ca. HS, Pignatelli = 36 m to HS, Nan-
dasena = 62 m (BOL2). For the experimental study within the present thesis (Chapter 6), BOL2 is
chosen as exemplary boulder as it is the largest and heaviest possible tsunami boulder on Bonaire
(Figure 1-3). While the boulder volume is evaluated with high accuracy by photogrammetry to be
approximately 77 m³, other parameters of the boulder remain unknown and can only be estimated
(e.g., density and density heterogeneity; time of deposition; exact source location).

5
Introduction

Figure 1-3: BOL2 on the island of Bonaire (V ≈ 77 m³). The approximated position on Bonaire is indicated by the star
(underlying image by CNES/Airbus/google, 2021).

1.2.2 Boulders as tsunami proxies

As stated above, the utilization of coarse clasts or fine sediments as proxies for high energy wave
events dates back to the 19th century. However, in recent decades, beginning from the 1980’s, the
utilization becomes more systematic including field studies and corresponding experimental and
numerical investigations (e.g., Atwater, 1987; Dawson et al., 1988; Nott, 1997, 2003; Imamura et
al., 2008; Engel and May, 2012; Yao et al., 2014; Oetjen et al., 2016, Chapter 3; Bressan et al.,
2018; Oetjen et al., 2020a, Chapter 6).

Basically, the use of fine or coarse deposits in tsunami or storm-wave research is related to their
deposition due to wave energy. Even if finer sediments dislocated by high-energy wave events are
comparably prone to alteration after the event (Szczuciński, 2020), they are still useful in the con-
text of palaeo-events since they still might become part of the stratigraphic record in suitable loca-
tions, where a datable archive of extreme-wave events may develop (Figure 1-1). A broad overview
on the topic of utilizing fine sedimentary deposits in palaeotsunami research in the Caribbean Sea
is given in Engel et al. (2016).

On the contrary, research on coarse clasts focuses more on the transport of single boulders up to
fine blocks or accumulations of boulders and the particular force balance of a single boulder. In
this context, and following Blair and McPherson (1999), boulders are defined as particles with in-
termediate axis lengths between 25.6 cm (fine boulders) and 409.6 cm (very coarse boulders)
while “fine block” refers to particles with intermediate axis lengths between 409.6 cm and 820 cm.
Whether a boulder is mobilized and dislocated by wave energy, and in which way, depends on a
set of parameters which encompass the initial boulder setting (submerged, subaerial, partially-
submerged), hydrodynamic wave properties (velocity, height), surrounding bathymetry and topog-
raphy (inclination, bottom roughness), other clasts in the flow (clast-clast interactions), density of
the medium (suspension load), as well as boulder properties (shape, weight; see Chapter 4). For
palaeotsunami boulder deposits, at first only the present-day boulder properties are known while

6
Introduction

other parameters remain largely unknown. Identifying the time of deposition of palaeo-deposits is
a difficult task which might be tackled by investigations of the chemical or stratigraphic composition
and on the boulder (e.g., influence of weathering, lichen cover; Scheffers, 2021). Investigations of
the surrounding area might reveal indications for the boulder’s pretransport location, transport
mode or indicate multiple mobilization (e.g., pits of according dimensions indicating original boulder
location, pedestals indicating longer intermediate deposition, or hitting marks indicating saltating
or rolling transport; Engel and May, 2012; May et al., 2015b; Boesl et al., 2020). Assuming that
only the hydrodynamic wave properties remain unknown, the necessary wave energy can be cal-
culated by comparing the present boulder location with the pretransport location and elaborating
the necessary forces for mobilization and transport.

First analytical equations for evaluating the necessary wave energy go back to Nott (1997) and
subsequent revisions/enhancements (e.g., Nott, 2003). Nott (1997) focussed on distinguishing
whether a boulder was relocated by a tsunami or wind waves generated by tropical cyclones, and
developed case-specific analytical equations. In Nott’s (1997) equations the forces due to drag, lift
and restraining forces are juxtaposed following

𝐹𝐷 + 𝐹𝑙 ≥ 𝐹𝑅 (1-1)

meaning that if the wave induced drag (FD) and lift (Fl) forces exceed the restraining forces (FR,
weight force, friction), the boulder will be set in motion (see also Chapters 2, 4 and 6). These basic
equations by Nott (1997) for determining the necessary wave force initiating boulder transport de-
pict the basis for most analytical and numerical boulder models applied today. However, since
1997 several enhancements of Nott’s equations were developed by Nott himself (2003) and other
researchers (e.g., Imamura et al., 2008; Pignatelli et al., 2009; Benner et al., 2010; Nandasena et
al., 2011a; Bressan et al., 2018; Lodhi et al., 2020). Depending on the complexity of the particular
model (i.e., numerical or analytical; force dimensions), these enhancements encompass develop-
ments regarding variable coefficients of friction, considering additional transportation modes beside
sliding (e.g., rolling and saltation in Imamura et al., 2008), additional pre-transport settings (e.g.,
joint-bounded; Nott, 2003; Engel and May, 2012) or taking into account forces that have been
neglected before (e.g. impact force; Lodhi et al., 2020; Table 1-1).

As can be seen in Table 1-1, available models attributed to tsunami-induced boulder transport
neglect the boulder shape. While the general flatness is often considered by the so-called Flatness
Index (FI, Chapter 4), no further implications on the mobilization or transport due to the shape are
considered in these models, whether numerical or analytical. The neglection of shapes beside
simplified cuboids or prisms in mathematical models has partly to be attributed to the complex
mathematical description arising from shapes without clear rectangular faces or constant radii, in
the case of spheres. Another reason is the unavailability of corresponding physical experiments
and, subsequently, a knowledge gap concerning complex-shaped boulders in highly turbulent
flows. This is especially true regarding the coefficient of drag which is, although known as to be
sensitive to the Froude number and boulder shape/orientation, is commonly assumed to be con-
stant (e.g., Hoerner, 1965; Noji et al., 1993; Nandasena and Tanaka, 2013).

7
Introduction

Table 1-1: Overview of existing models for tsunami-induced boulder transport (partly taken from Lodhi et al., 2020).

Force Boulder shape


Study dimensions Friction standard variation Drag Coefficient
Nott (1997) 1D Constant Cubic Flatness Constant (CD = 1.2 [-])

Luccio et al. (1998)a 1D Constant Cobble/ Diameter/ Constant


Disk height

Nott (2003) 1D Constant Cubic Flatness Constant (CD = 2 [-])

Imamura et al. 2D Variable Cubic Flatness Variable based on Froude


(2008) number

Benner et al. (2010) 1D Constant Cubic Flatness Constant (CD = 1.2 [-])

Nandasena et al. 1D Constant Cubic Flatness Constant (CD = 1.95 [-])


(2011a)

Nandasena et al. 1D Constant Cubic Flatness Constant (CD = 1.95 [-])


(2011b)

Buckley et al. (2012) 1D Range Tabular Shape Varied with respect to flow
factor velocity

Nandasena and 1D Constant Cubic Flatness Constant (CD = 1.95 [-])


Tanaka (2013)

Bressan et al. (2018) 1D Range Cubic Flatness Varied based on literature for
different boulder orientations
to the flow impact (i.e.,
long/short boulder axis)
Lodhi et al. (2020) 1D Constant Cubic Flatness Constant (CD = 1.5 [-])

This PhD thesis 2D Constant Non-self-intersecting 2D poly- Depending on the polygon


gons shape
a
Not addressed to boulder transport; bFor super typhoon Haiyan

Enabling a complete consideration of natural boulder shapes in analytical models, which are in-
tended for comparably easy application, has to be recognized to be more or less impossible since
such models would need to incorporate complex mathematical formulations for accurate descrip-
tions of the boulder shape (e.g., for inertia, centre of mass). On the other hand, the integration of
complex shapes in numerical models (2D or 3D) is a promising step for increasing the accuracy of
such models. Nevertheless, also for analytical models the role of complex shapes should be incor-
porated (e.g., by adjusted drag coefficients), since it is a major parameter influencing the transport
process (e.g., due to the irregular ground contact area or drag coefficient; see also Chapter 6).

1.2.3 Research demands

As mentioned above, possible past tsunami occurrences in the Caribbean Sea, and at the island
of Bonaire in particular, have not yet been clarified completely. However, boulder and sedimentary
deposits throughout the coastlines in the Caribbean Sea indicate that high-energy wave events
have occurred in regions where no corresponding (reliable) reports are available. For several de-
posits the discussion about whether the deposits were dislocated by storm waves, tsunamis or
long-term processes such as sea-level change is still ongoing (e.g., Scheffers, 2002b; Scheffers et
al., 2014; Morton et al., 2008; Spiske et al., 2008; Engel and May, 2012; Rixhon et al., 2018).

8
Introduction

As a first step to test the palaeo-tsunami hypothesis, it is necessary to evaluate the threshold tsu-
nami magnitudes capable of dislocating the boulder deposits and to identify possible tsunami trig-
ger events which might have generated tsunamis of suitable magnitudes. Even if the former was
already tackled by several researchers for the island of Bonaire, the discussion is still ongoing due
to shortcomings in the available empirical boulder transport and mobilization-threshold equations,
and inaccuracies as well as uncertainties in the in-situ measurements (e.g., time of deposition,
boulder density/weight, boulder dimension).

Due to the ambiguities in the tsunami history of the Caribbean Sea and, especially, of the island of
Bonaire, the following research demands are derived:

1) Since possible tsunami trigger events which might have caused the dislocation of BOL2
on Bonaire are not identified yet, it is necessary to identify possible earthquake locations
which might have generated a tsunami of sufficient magnitude.

2) Only few systematic investigations of tsunami induced boulder transport by physical labor-
atory experiments are available as basis for empirical analytical.

3) There is a lack of information on how boulder shapes beside idealized cubes and prisms
influence the mobilization threshold and transport process.

4) It is assumed that sedimentary load in the approaching tsunami wave and bore needs to
be considered as additional parameter, influencing the transport distance (e.g., Kain et al.;
2012).

These research demands depict the basis for the research questions formulated in the subsequent
paragraph.

1.3 Objectives

With respect to the research demands formulated in Chapter 1.2.3, the following research ques-
tions are derived to be addressed in the present PhD thesis and depict the objectives:

1) Which earthquake zones in the Caribbean Sea could be capable of generating tsunamis
impacting Bonaire, and do these have sufficient energy for dislocating the largest boulder
deposits on Bonaire? (Chapters 2 and 3)

2) Which are the main shortcomings of published physical boulder transport experiments,
and which shortcomings are the most important to be solved in order to better reflect real
scenarios and improve the understanding of tsunami-induced boulder transport pro-
cesses? (Chapters 4 and 5)

3) To which extent does the boulder shape influence the tsunami-induced boulder transport
process? (Chapter 6 and 7)

4) Do numerical simulations support the assumption of a transport-supporting effect of sedi-


mentary load in the approaching tsunami? (Chapter 7)

Based on the formulated objectives, the research questions in the present PhD thesis narrow down
from a broad investigation on possible tsunamis throughout the Caribbean Sea and the island of
Bonaire, through identifying the major shortcomings in existing (analytical) models and physical

9
Introduction

experiments to tackling two of the elaborated shortcomings (influence of boulder shape and sedi-
ment load) by physical experiments and a numerical model.

1.4 Methods

The present PhD thesis is a cumulative dissertation composed of three publications in international
journals listed in the Web of Science (Chapters 2, 4, 6), one publication in a journal not listed in the
web of science (Chapter 3), one reviewed book chapter (Chapter 5), and a non-published chapter
which is in preparation for publication (Chapter 7). An overview on the publications is given in Table
1-2. As outlined in the objectives, the present PhD thesis begins with a general investigation of
tsunami occurrences in the Caribbean Sea, continues with detailed analysis on the island of Bon-
aire as well as the shortcomings of published physical boulder transport experiments, and ends
with in-depth investigations on a single boulder deposit (Figure 1-4).
Table 1-2: Overview of the published articles and book chapters within this PhD thesis.

Chapter 2 3 4 5 6 7
Publication Article Article Article Book chapter Article Article in
type preparation
Published in Earth Science Proceedings Earth Science Geological Earth Surface -
Reviews of 35th Con- Reviews Records of Processes
ference on Tsunamis and and Land-
Coastal Engi- Other Ex- forms
neering treme Waves
(Engel et al.,
2020)
Journal im- 12.413 - 12.413 - 4.133 -
pact factor
Listed in
Web of Sci-      -
ence

The rooting research question (1) regarding possible tsunami occurrences on the island of
Bonaire is investigated by reviewing possible tsunami triggers in the Caribbean Sea and
assessing the resulting tsunami heights at Bonaire (Chapter 2). The tsunami scenarios were
estimated by identifying possible triggering fault lines in the Caribbean Sea through literature re-
view. The finally identified tsunami scenarios were than numerically modelled with Deltares Delft3D
and Deltares DelftDashboard, based on the tsunami trigger/earthquake model of Okada (1985), as
exemplary shown in Figure 1-5. The design of the numerical models and their results are presented
in Chapters 3 and 4.

From evaluating research question 1) and as detailed investigations on the island of Bonaire
showed, currently existing analytical models for describing boulder mobilization thresholds are not
sufficiently accurate. To which extent such inaccuracies are connected to the underlying
physical boulder transport experiments is addressed by answering research question 2) by
a literature review on physical boulder transport experiments (Chapters 4 and 5).

10
Introduction

Figure 1-4: Schematic outline of the PhD thesis from the universal look on tsunamis in the Caribbean Sea to specific
investigations on the hydrodynamics of a single boulder. The final numerical model and outlook lead back to more
universal investigations.

Figure 1-5: a.1 & a.2: Computed water displacement on a plane grid after the trigger model of Okada (1985). b.1 & b.2
Superelevated illustration of displacement (100:1) compared to surrounding bathymetry with water level zero (own cal-
culation and representation).

11
Introduction

Evaluating research question 2) revealed that shortcomings in the analytical models result partly
from using oversimplified boulder shapes. Research question 3), the influence of the boulder
shape on boulder transport, is therefore investigated by physical experiments in a flume
utilizing idealized as well as naturally shaped boulders and comparing their behaviour un-
der tsunami bore impact (Chapter 6). Furthermore, alternating shore types were investigated in
these experiments including a stepped shore as resembling of the steep coast at Boka Olivia on
the island of Bonaire. The results of the experimental study are partly shown in Chapter 4 and
completely presented in Chapter 6.

As enhancement to the physical experiments, the first version of a numerical model is de-
veloped based on the Immersed Boundary Method (IBM) for comparing the transport be-
haviour of different idealized boulder shapes as a first step and receive first insights into
the influence of sediment load (research question 4; Chapter 7). In this model, the tsunami
wave can be modelled including sedimentary load for investigating furthermore the influence of
sedimentary load in the wave on boulder transport.

1.5 Declaration of personal contribution

The authors of the different scientific research studies contributed in the following proportions to
conception and design, data collection, data analysis, interpretation and conclusions, and manu-
script preparation of the different chapters:

• Chapter 1: Jan Oetjen defined the objectives of the PhD thesis based on discussions with
Holger Schüttrumpf and Max Engel, designed the structure, created all figures and tables,
wrote and corrected the manuscript. Holger Schüttrumpf and Max Engel proof-read and
commented on the chapter.

• Chapter 2: Max Engel defined the objective of the manuscript, designed the structure and
compiled the database of potential tsunami deposits of the Caribbean region. Jan Oetjen
elaborated the investigated tsunami scenarios, conducted and set-up all numerical simu-
lations and evaluated their results. Helmut Brückner and Matthias May discussed the re-
sults of both the evaluation of regional tsunami deposits and the numerical simulations,
and proof-read the manuscript. The basis for the study were experiences through previous
investigations in the Caribbean Sea in the framework of a project funded by the Deutsche
Forschungsgemeinschaft (DFG; Grant Number BR 877/26-1) while the study itself were
funded by the Max Delbrück Prize for junior researchers (Grant Number DFG ZUK 81/1).

• Chapter 3: Jan Oetjen defined the objectives of the article, designed the structure, created
all figures and tables, wrote and corrected the manuscript. Max Engel, Helmut Brückner,
Shiva P. Pudasaini and Holger Schüttrumpf proof-read the manuscript, and the results
were discussed with Max Engel and Holger Schüttrumpf. The numerical models and phys-
ical experiments were designed, executed, and evaluated by Jan Oetjen and Bache-
lor/Master theses supervised by him (Friesen, 2015; Schechinger, 2015; Klosterhalfen,
2016; Küstermann, 2016; Schönberger, 2017; Maaßen, 2018; Arumugasamy, 2018). The
idea for the experiments and numerical models overall were based on the DFG funded
research project ‘Modelling tsunami-induced coarse-clast transport – combination of phys-
ical experiments, advanced numerical modelling and field observations’ (Grant Numbers

12
Introduction

SCHU 1054/7-1, EN 977/3-1) established by Jan Oetjen, Max Engel, Shiva P. Pudasaini
and Holger Schüttrumpf.

• Chapter 4: Jan Oetjen defined the objectives of the article, designed the structure, created
all figures and tables, wrote and corrected the manuscript. Max Engel and Holger Schüt-
trumpf proof-read the manuscript, and the results were discussed with Max Engel. The
result discussion and idea elaboration of the decision supporting tool were created by Jan
Oetjen and discussed with Max Engel. The idea for the review paper was based on the
DFG funded research project ‘Modelling tsunami-induced coarse-clast transport – combi-
nation of physical experiments, advanced numerical modelling and field observations’
(Grant Numbers SCHU 1054/7-1, EN 977/3-1) established by Jan Oetjen, Max Engel,
Shiva P. Pudasaini and Holger Schüttrumpf.

• Chapter 5: Jan Oetjen defined the objectives of the article, designed the structure, created
all figures and tables, wrote and corrected the manuscript. Max Engel and Holger Schüt-
trumpf proof-read the manuscript, and the results were discussed with both. The idea for
the book chapter was based on the concept of the handbook ‘Geological Records of Tsu-
namis and Other Extreme Waves’ edited by Max Engel, Jessica Pilarczyk, Simon Matthias
May, Dominik Brill and Ed Garret and published by Elsevier.

• Chapter 6: Jan Oetjen defined the objectives of the article, designed the structure, created
all figures (except of Figure 6-1, created by Max Engel) and tables, wrote and corrected
the manuscript. Max Engel, Shiva P. Pudasaini and Holger Schüttrumpf proof-read the
manuscript, and the results were discussed with them. The experimental design, conduc-
tion of the experiments and result evaluation were done by Jan Oetjen and Master/Bach-
elor theses supervised by him (Klosterhalfen, 2016; Küstermann, 2016; Schönberger,
2017; Maaßen, 2018; Arumugasamy, 2018). The results were discussed with Max Engel
and Holger Schüttrumpf. The idea for the article was based on the DFG funded research
project ‘Modelling tsunami-induced coarse-clast transport – combination of physical exper-
iments, advanced numerical modelling and field observations’ (Grant Numbers SCHU
1054/7-1, EN 977/3-1) established by Jan Oetjen, Max Engel, Shiva P. Pudasaini and
Holger Schüttrumpf.

• Chapter 7: Jan Oetjen defined the objectives of the chapter, designed the structure, cre-
ated all figures and tables, wrote and corrected the manuscript. All computational tools in
connection to the Immersed Boundary Method (IBM) and all post-processing tools includ-
ing visualization were developed and programmed by Jan Oetjen. The original two-phase
flow code and the numerical slope in this, were developed and implemented by Shiva P.
Pudasaini. Jan Oetjen transferred the original code to Fortran95, had the idea for and de-
veloped/wrote/implemented the immersed boundary program code. While the basic IBM
approach goes back to Peskin (1972), the basic approach for using forcing nodes in the
approach presented herein is adopted from Liao et al. (2010) and references therein. The
geometric algorithm used to detect the IBM is based on own developments of Jan Oetjen,
although similar approaches exist (e.g., Hylla, 2013). The idea of applying a smoothing
factor to the force goes back to personal discussions with Julia Kowalski. The idea for the
program is based on the DFG funded research project ‘Modelling tsunami-induced coarse-
clast transport – combination of physical experiments, advanced numerical modelling and

13
Introduction

field observations’ (Grant Numbers SCHU 1054/7-1, EN 977/3-1) established by Jan Oet-
jen, Max Engel, Shiva P. Pudasaini and Holger Schüttrumpf.

• Chapter 8: Jan Oetjen defined the conclusions of the PhD thesis, discussed the results,
de-signed the structure of the chapter, wrote and corrected the manuscript. Holger Schüt-
trumpf and Max Engel discussed the results with Jan Oetjen, proof-read and commented
on the manuscript.

14
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

2 Tsunami deposits of the Caribbean – Towards an improved coastal haz-


ard assessment

Max Engel, Jan Oetjen, Simon Matthias May, Helmut Brückner


This is an accepted manuscript of an article published by Elsevier in Earth-Science Reviews in
October 2016. Online available: http://dx.doi.org/10.1016/j.earscirev.2016.10.010

Keywords: Tsunami deposit, Storm deposit, Palaeo-tsunami research, Holocene, Coastal hazard
management, Caribbean basin

Abstract
Coasts worldwide experience considerable population pressure and the demand for reliable hazard
management, such as of tsunamis, increases. Tsunami hazard assessment requires information
on long-term patterns of frequency and magnitude, which are best explained by inverse power-law
functions. In areas with a short historical documentation, long-term patterns must therefore be
based on geological traces. The Caribbean tsunami hazard is exemplified by N80 events triggered
by earthquakes, volcanic activity, or mass wasting within the region or in the far-field during the
last 520 years. Most of these tsunamis had regional or local impacts. Based on two numerical
hydrodynamic models of tsunamis spawning at the Muertos Thrust Belt (MTB) and the South Car-
ibbean Deformed Belt (SCBD), which are two scenarios only marginally considered so far, we show
that pan-Caribbean tsunamis can also be taken into account. We furthermore review almost 60
sites for possible geological evidence of tsunamis in the Caribbean including fine-grained subsur-
face deposits and subaerial coarse clasts, and re-evaluate their implications for tsunami hazard
assessment against state-of-the-art models of onshore sediment deposition by tsunamis and ex-
treme storms. The records span the mid- to late Holocene, with very few exceptions of Pleistocene
age.

Only a limited number of reliable palaeotsunami records with consistent and robust age control
were identified, hampering inter-island or interregional correlation of deposits of the same event.
Distinguishing between storm and tsunami transport of solitary boulders is very difficult in most
cases, whereas those clasts arranged as ridges or incorporated into polymodal ridge complexes,
which line many windward coasts of the Caribbean, can mainly be attributed to long-term formation
during strong storms, overprinting potential tsunami signatures. The quantification of tsunami flood-
ing parameters such as flow depth, inundation distance or flow velocities, by applying inverse and
forward numerical models of sediment transport is still very limited and needs to be extended in
the future. Likewise, sediment-derived hazard implications still await implementation in spatial plan-
ning. As extreme-wave deposits are clearly understudied in the Caribbean, there is great potential
for coastal hazard assessment to be developed and improved. Thus, further studies using common
standards of high-resolution methods of bedform and stratigraphical documentation and robust
chronological models with independent age control, combined with refined inverse and forward
models of sediment transport and deposition are required to reconstruct reliable patterns of mag-
nitude and frequency of palaeotsunamis in the Caribbean and to map hazard-prone areas. To date,
known palaeotsunami deposits from the Caribbean probably represent only a fraction of actually
happened prehistoric tsunamis and, therefore, do not reflect major tsunami inundations of the past
adequately.

15
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

2.1 Introduction

Coasts around the globe experience high population growth, resulting in an increasing number of
humans exposed to hazards associated with the sea and continental margins (Brückner, 2000;
Adger et al., 2005). With > 700 islands and 55,383 km of coastline (UNEP, 2004), and most of its
population, infrastructure, and tourist facilities concentrated in close proximity to the sea (McGregor
and Potter, 1997; von Hillebrandt-Andrade, 2013), the Caribbean region is disproportionally vul-
nerable to coastal hazards (Fitzpatrick, 2012).

The traditional triad of rapid-onset hazards in the Caribbean, as perceived some decades ago and
summarized in Tomblin (1981), comprises earthquakes, volcanism and hurricanes. This was once
more exemplified by the devastating earthquake of Haiti in 2010 with a death toll of > 230,000
(Bilham, 2010; Fritz et al., 2013), the eruption of Mount Pelée on Martinique in 1902, which de-
stroyed the former principal town of the island and killed around 28,000 inhabitants (Tanguy, 1994),
and the Great Hurricane of 1780 with a similar number of fatalities along the Lesser Antilles island
arc (Rappaport and Fernandez-Partagas, 1997).

However, history tells that the Caribbean is also highly susceptible to the hazard of tsunamis, which
is closely linked to the high seismic activity and volcanism, coastal and submarine landslides, and
teletsunamis generated in the open Atlantic Ocean. Even though 127 potential tsunamis were re-
vealed for the Caribbean by O'Loughlin and Lander (2003) based on historical accounts (Figure
2-1), the Caribbean is still lacking well-founded information regarding long-term occurrence pat-
terns of high-magnitude tsunamis (Rowe et al., 2009). However, a regional early warning system
has been installed for the Caribbean and adjacent regions, relying on > 115 seismic stations, 55
sea-level stations, five DART buoys, detailed local evacuation maps, as well as several community
engagement programs (von Hillebrandt-Andrade, 2013).

Figure 2-1: Overview map of the Caribbean region based on the GEBCO One Minute Grid, version 2.0
(http://www.gebco.net) with overlays of the tectonic pattern (Pindell and Kennan, 2009) and vectors of plate movement
(Meschede and Frisch, 1998). Black squares show observations of tsunami run-up since 1498. White rims indicate that
multiple historical tsunamis were recorded (NGDC/WDS, 2016). BVI = British Virgin Islands; B = Bonaire; C = Curaçao;
A = Aruba.

16
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

Similar to other types of natural hazards (Korup and Clague, 2009; Corral et al., 2010), frequency-
magnitude patterns of tsunamis can be explained best by inverse power-law functions, but without
upper truncation (Burroughs and Tebbens, 2005). The tsunami record of the Caribbean (1498–
2014) (NGDC/WDS, 2016) is mainly compiled of a rather low number of non-verifiable eyewitness
observations on tsunami height, which usually do not distinguish between wave height, flow depth
and run-up height. While this record matches the power-law function inadequately, the fit of the
larger, global dataset for the same time period is much better (Figure 2-2a). By disregarding run-
up measurements and unspecific eyewitness records and only considering tide-gauge measure-
ments as well as flow depth and tsunami height derived from post-tsunami measurements from the
global dataset, the confidence level rises to > 0.95 (Figure 2-2b), therefore corroborating the in-
verse power-law relationship.

The spatial distribution of tsunami hazard in the Caribbean based on historical accounts and in-
strumental data is irregular, focusing on the Greater Antilles islands of Jamaica, Hispanola, and
Puerto Rico, followed by the Lesser Antilles island arc (Zahibo and Pelinovsky, 2001; Parsons and
Geist, 2009) (Figure 2-3). Nevertheless, a number of uncertainties are associated with these data:

• Information on historical tsunamis or earthquakes is mainly based on colonial reports.


These reports have to be considered incomplete, since population density was lower and
thus events may have occurred unnoticed. Civil authorities only noted events if they se-
verely affected the economic activities of the colony (O'Loughlin and Lander, 2003).

• The quality and reliability of sources on tsunami occurrence varies greatly, partly due to
the fact that the concept of tsunamis and their trigger mechanisms were not well under-
stood in the past (O'Loughlin and Lander, 2003).

• Historical documentation of Caribbean tsunamis covers the last 517 years, which is ex-
ceeded by recurrence intervals of highest-magnitude tsunamis in other parts of the world
(Jankaew et al., 2008; Brill et al., 2012; Sawai et al., 2012).

17
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

Figure 2-2: a) Frequency-size distribution for reported heights of tsunamis vs. number per year. Each event is only
represented once by its highest value. Triangles = maximum observational and instrumental accounts of wave height,
flow depth or run-up height, documented for each tsunami in the Caribbean 1498–2015 CE (NGDC/WDS, 2016). Circles
= maximum observational and instrumental accounts of wave height, flow depth or run-up height, documented for each
tsunami worldwide 1498–2015 CE (NGDC/WDS, 2016). b) Heights of tsunamis vs. total number per year. Squares =
global dataset of only high-precision tide-gauge measurements and maximum flow depth or water height derived from
post-tsunami measurements (NGDC/WDS, 2016).

Figure 2-3: Overview of the Caribbean basin separated into five sectors (dashed lines). Greyscale quadrants indicate
decadal frequencies of tropical cyclones based on tracks charted by the US Weather Bureau between 1871 and 1987
(Reading, 1990). Circles show the modelled probabilities of tsunami occurrence (run-up N0.5 m) within a period of 30
years (Parsons and Geist, 2009). Numbers refer to reviewed sites in Table A-2 and Table 2-1. Blue frames = fine sedi-
mentary records; red frames = coarse-clast records. The width of the frames refers to the depositional process (storm
or tsunami) inferred by the original author(s) (0 [thinnest frame] = no disturbance in the stratigraphy; 1 = disturbance/dep-
osition attributed to storm; tsunami impact is excluded; 2 = disturbance/deposition documented and storm is favoured
over tsunami, but the latter is not entirely excluded; 3 = disturbance/deposition is ascribed to either hurricane or tsunami;
4 = disturbance/deposition is documented and tsunami is favoured over hurricane, but the latter is not entirely excluded;
5 = disturbance/deposition is documented, ascribed to tsunami, and hurricane is excluded; see also Table A-3 and Table
A-4).

The effects of tsunamis along the Caribbean coasts prior to the era of historical and instrumental
documentation can only be traced by their telltale deposits stored in a variety of sedimentary ar-
chives. These records often cover several millennia and have the potential to significantly add to

18
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

the development of coastal hazard assessment by providing insight into approaching directions of
tsunamis and long-term occurrence patterns (Scheffers and Kelletat, 2006; Weiss and Bourgeois,
2012; Switzer et al., 2014), even though they may be biased by limited sediment preservation
potentials (Szczuciński, 2012; Spiske et al., 2013) or very dynamic coastal environments influenc-
ing patterns of onshore deposition by extreme waves over short time (May et al., 2015a). Tsunami
deposits help to delineate areas prone to tsunami flooding, which is essential to analyse exposure
and vulnerability, calculate potential losses and develop site-specific mitigation concepts (Figure
2-4). Against this background, this paper aims at providing a first comprehensive review of deposits
which are potentially related to tsunami-induced flooding in the Caribbean region. We furthermore
present new hydrodynamic models for major tsunamis potentially related to these deposits and
compile implications for regional coastal hazard assessment.

Figure 2-4: The role of palaeotsunami research in a tsunami hazard management process and schematic depiction and
classification of sedimentary archives relevant to the Caribbean (see also Figure 2-5). Steps 1–4 modified after Dall'Osso
and Dominey-Howes (2010), details of Step 4 partly after Stein and Stein (2013).

19
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

Figure 2-5: Potential sedimentary archives of tsunami- or storm-induced deposits or landforms in the Caribbean (see
categories in Figure 2-4) as exemplified in the Lac Bai area of southeast Bonaire (Figure 2-1) (Satellite image: Digital
Globe, 4 August 2013; oblique aerial insert image courtesy of D. Kelletat).

2.2 Tsunamis and their onshore deposits

Telltale deposits of tsunamis usually comprise a coarser spectrum than their vertically confining
background sediments. They can be found both onshore (Peters and Jaffe, 2010; Brill et al., 2012;
Goff et al., 2012; Spiske et al., 2013) and offshore (Sugawara et al., 2008; Feldens et al., 2012;
Tamura et al., 2015) and are separated into fine-grained (clay, silt, sand, pebble) (Goff et al., 2012)
and coarse-clast (pebble, cobble, boulder, block) (Goto et al., 2010a; Etienne et al., 2011) deposits,
depending on sediment sources, coastal geomorphology, flow patterns, and local preservation po-
tential. Fine-grained tsunami deposits are usually left and preserved in sediment sequences of
coastal lakes (Minoura et al., 1994), lagoons (Minoura and Nakaya, 1991), floodplains (Nanayama
et al., 2000; Jaffe et al., 2003) or swales associated with beach and barrier coasts (Jankaew et al.,
2008; Brill et al., 2012), muddy coasts (Jaffe et al., 2003; Spiske et al., 2013) or deltas (Bourgeois
and Johnson, 2001) and estuaries (Lario et al., 2010). Coarse-clast deposits are found along
(mostly emerging) rocky shores (Scheffers and Kelletat, 2003; Bourgeois and MacInnes, 2010;
Etienne et al., 2011) and adjacent lowlands (Goto et al., 2010a), on coral reef flats (Goto et al.,
2010a; Terry et al., 2013; Lau et al., 2014), but also on beaches, for instance in case of beachrock
clasts (Terry et al., 2013), or inside shallow salt ponds and salt flats backing coral reefs (Atwater
et al., 2012). In the Caribbean, morpho-sedimentary archives can be separated into intertidal to
highly elevated carbonate platforms including reef flats and karst sinkholes, as well as mangrove
swamps, lagoons, coastal lakes and backbarrier mudflats (see categories A–F in Figure 2-4), ex-
emplified by the Lac Bai area on the south Caribbean island of Bonaire (Figure 2-5).

20
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

2.2.1 Fine-grained tsunami deposits

Common features of tsunami deposits include erosional basal contacts, buried plants or soils, ba-
sal load structures, rip-up clasts, landward fining, cross bedding, multimodal grain size distribution,
poor sorting, heavy mineral lamination, heavy metal contamination (limited to recent examples),
macro- and microfossil remains representing a broad range of habitats and states of preservation,
high amounts of reworked benthic marine diatoms, a basal traction carpet, one or several fining-
upward sequences with mud caps coinciding with the number of tsunami waves or even represent-
ing backwash, potentially intercalated with ungraded sections, abrupt changes in pollen concen-
tration and composition, or a marine geochemical signature. Their thickness usually ranges from
millimetres to several decimetres (see details in Table A-1). Many of the features are highly site-
dependent and determined by the hydraulic characteristics of the tsunami, available sediment
sources, local topography and bathymetry, and post-depositional changes (Kortekaas and Daw-
son, 2007; Morton et al., 2007; Sugawara et al., 2008; Mamo et al., 2009; Lario et al., 2010; Peters
and Jaffe, 2010; Engel and Brückner, 2011; Moore et al., 2011; Goff et al., 2012; Szczuciński,
2012).

Preservation of tsunami deposits mostly depends on the mineralogical composition, the onshore
geomorphic setting, rainfall regime, vertical tectonic movement and anthropogenic impact (Spiske
et al., 2013). Tsunami deposits exposed subaerially in tropical carbonate environments, which pre-
vail in the Caribbean, seem to have a limited potential of preservation due to reworking, carbonate
dissolution and bioturbation (Nichol and Kench, 2008; Szczuciński, 2012).

Even though tsunamis differ significantly in hydraulic characteristics from other extreme inundation
processes such as storm surges, both their deposits show many similarities, which often impedes
unequivocal interpretation (Table A-1). It is generally accepted that the presence of distinct sub-
units created by run-up and backwash is associated with tsunamis, and that tsunami deposits lack
foreset bedding (Nanayama et al., 2000), which is very common in subaqueous storm-induced
washover deposits (Sedgwick and Davis, 2003; Brill et al., 2016). Internal mud drapes, often verti-
cally confining graded tsunamigenic sand beds, have not been reported from storm deposits (Ko-
matsubara and Fujiwara, 2007; Morton et al., 2007). Furthermore, the landward extent of tsunami
deposits tends to be larger compared to storms (Morton et al., 2007; Brill et al., 2016), and they
often have a larger source area comprising deeper waters, which might be reflected by a broader
range of microfaunal remains (Mamo et al., 2009; Uchida et al., 2010; Goff et al., 2012). Thus,
identifying the origin of a candidate tsunami deposit requires consideration of a broad variety of
factors, including the “tsunami potential” of a region. It is most valuable to have reference deposits
from either modern or historically well documented tsunami or severe storm events, to ensure the
best deduction as to a deposit's origin (Engel and Brückner, 2011).

2.2.2 Coarse-clast tsunami deposits

The size and pattern of boulder accumulations of tsunamis are mostly determined by flow velocity,
availability of material, surface roughness, the pre-transport setting, coastal topography, rock den-
sity, shape and particle-particle interaction (Bourgeois and MacInnes, 2010; Etienne et al., 2011;
Weiss, 2012; Terry et al., 2013). In carbonate reef settings and steep offshore bathymetries, as
typically found in the Caribbean, tsunami gravel and boulder fields show rather abrupt landward
boundaries instead of exponential landward fining as observed for boulder fields created by tropical

21
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

cyclones (Goto et al., 2010a; Etienne et al., 2011; Watt et al., 2012a; Lau et al., 2014). The for-
mation of polymodal ridge complexes or ramparts, respectively, a common landform built up during
strong storms, has not been directly observed for tsunamis (Etienne and Paris, 2010; Etienne et
al., 2011; Richmond et al., 2011) though it was suggested in some places (Scheffers and Kelletat,
2003) (Table A-2). The main axis of large boulders tends to arrange perpendicular to the flow
direction and can be used to infer the approaching angle and a possible event source (Terry et al.,
2013). Several studies assume that the capacity of a tsunami to transport very large boulders is
significantly higher than in storms (Nott, 2003; Barbano et al., 2010; Benner et al., 2010; Engel and
May, 2012), whereas this view has recently been challenged (Weiss, 2012). In particular, the
coarse-clast record of Typhoon Haiyan (7–9 November 2013) on Eastern Samar (Philippines)
showed that infragravity waves created by the interaction of groups of high and steep storm waves
with fringing reefs reach transport capacities similar to tsunamis (May et al., 2015b). In cliff-top
positions, vertical jets may support the transport of very large boulders during storms (Cox et al.,
2016).

2.2.3 The challenges of dating event deposits

In order to use deposits to establish long-term frequency patterns of tsunamis, they need to be
dated accurately. Age estimates of fine-grained Holocene tsunami deposits are usually derived
from 14C dating. Where quartz or feldspar is dominant and luminescence properties are suitable,
optically stimulated luminescence (OSL) can be used (e.g., Brill et al., 2012), even for deposits
from the last interglacial. In subrecent contexts, 137Cs has been applied successfully (e.g., Barra et
al., 2004). Method-specific uncertainties, such as the reworking of dated organisms and/or varying
reservoir effects in time and space in the case of 14C (e.g., May et al., 2015a; Adomat and Gischler,
2016) or partial bleaching in the case of OSL (e.g., Bishop et al., 2005; Brill et al., 2012), still
complicate the establishment of chronostratigraphies. Embedded into a stratigraphic sequence,
dating of fine-grained tsunami deposits can also be supported by reliably dated volcanic ash layers
(e.g., Sawai et al., 2012) or diagnostic archaeological remains (e.g., Rajendran et al., 2011).

Since subaerial boulder deposits lack stratigraphic contexts in most cases, it is much more difficult
to date the timing of their transport. Marine organisms attached to boulders, such as boring mol-
luscs, vermetids or serpulids, are dated by 14C (e.g., Barbano et al., 2010; Biolchi et al., 2016), but
only provide maximum ages and are prone to uncertainties imposed by diagenetic processes
(Douka et al., 2010). Along reef-dominated coasts, U/Th is preferred over 14C or electron spin res-
onance (ESR) for dating coral material (Schellmann et al., 2011; Scheffers et al., 2014). In temper-
ate regions, lichenometry has been applied for relative age dating on decadal scales (Hall et al.,
2006). Attempts to directly date boulder transport onshore by surface exposure dating ( 36Cl) are
currently being developed (e.g., Rixhon et al., 2016), whereas Sato et al. (2014) revealed cau-
tiously optimistic results through the application of palaeomagnetic dating. Hearty (1997) used
amino acid racemisation (AAR) on carbonate boulders, though this approach permits only relative
age estimations for the age of the substrate. In the near future, OSL surface exposure dating
(Sohbati et al., 2012) might represent an alternative tool for dating boulder transport in the Holo-
cene.

22
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

2.2.4 Deposits of recent extreme-wave events in the Caribbean

In the Caribbean, no sedimentological post-tsunami documentation has ever been carried out. Za-
hibo et al. (2005) and Le Friant et al. (2008) conducted field surveys following the small 21 Novem-
ber 2004 tsunami generated by shallow normal faulting between Basse-Terre (Guadeloupe) and
Dominica. Besides observations on water-level changes, run-up and inundation on Les Saintes
and southern Basse-Terre based on debris and eyewitness interviews, no account is made for
sedimentation or erosion features. Similarly, Fritz et al. (2013) only recorded characteristics and
inundation patterns of the twin tsunamis following the 12 January 2010 Haiti Earthquake. Geo-
morphic and sedimentary impacts of the 11 October 1918 tsunami in the Mona Passage west of
Puerto Rico are briefly touched upon by Reid and Taber (1919, p. 111), who describe the beach
near Point Agujereada as “turned into a sandy waste” by a 5.5–6 m-high wave. Rectangular blocks
(> 1 t) shifted for several tens of metres inland and slightly downslope by a > 4 m wave southwest
of Aguadilla.

Sedimentary effects of recent strong storms representing the most important alternative mecha-
nism of episodic coastal deposition and erosion are much better documented. Atwater et al. (2014)
surveyed effects of Hurricane Earl (30 Aug. 2010, Saffir-Simpson hurricane scale [SSH] category
4) on Anegada, British Virgin Islands. They found limited sediment deposition, mostly in the form
of small washover fans along the south coast and redistribution of microbial detritus in salt ponds,
which stands in stark contrast with extensive deposits of palaeo-wave events on the island. The
hurricane waves induced only limited reworking at pre-existing coral rubble ridges on the shore
(Spiske and Halley, 2014). Based on further monitoring, Spiske (2016) inferred, however, that
these intertidal to supratidal ridges also experience modification during tropical storms and low-
category hurricanes.

All around the coast of Jamaica, pre-existing clasts were shifted during Hurricane Dean (19 August
2007, SSH category 4), including one cliff-top boulder of 80 t, which moved horizontally for a few
metres (Khan et al., 2010). Caron (2012) surveyed effects of Hurricanes Omar (16 Oct. 2008) and
Earl on a beach of southwest St. Bartholomew, where massive erosion of sand- and granule-sized
sediment occurred and previously covered beachrock became exposed.

After Hurricane Lenny (16 November 1999, SSH category 4), Scheffers (2005) found coral-rubble
ridges consisting to 20–30 % of freshly broken material, up to 1 m high and 8 m wide, with small
lobate washover structures, piled up by the storm waves on the leeward coast of Bonaire (Leeward
Antilles). A new coral-rubble spit was created and later modified by Hurricane Ivan (15–16 Sep-
tember 2004, SSH category 5), which created wave heights of up to 12 m at the coast of Bonaire
(Scheffers, 2005; Scheffers and Scheffers, 2006) and other Caribbean islands (Stewart, 2004), as
well as maximum wave heights of up to 28 m measured by buoys on the shelf of the Gulf of Mexico
(Wang et al., 2005). At Playa Funchi, northeast Bonaire, Lenny created a 1.5 m high and 26 m
wide ridge with separate units indicating in-flow and return flow by the orientation of imbricated
clasts (Spiske and Jaffe, 2009). Hurricane Ivan induced erosion of pre-existing polymodal ridges
in the southeast, as well as flattening or seaward steepening of pre-existing ridges in the northwest.
Sands were redistributed to sand sheets and single boulders of up to 25 t were dislodged on top
of the elevated palaeo-reef platform in the northeast, whereas other boulders of up to 6 t were
detached and moved in saltation mode (Scheffers and Scheffers, 2006).

23
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

2.3 Physical setting and tsunami triggers of the Caribbean

The geographical entity of the Caribbean comprises the Caribbean Plate and adjacent coastal
areas (Figure 2-1). It includes the Greater Antilles (Cuba, Jamaica, Hispaniola, Puerto Rico) and
the more eastern and southern Lesser Antilles, the Bahamas and Turks and Caicos Islands, as
well as the coastlines of eastern Central America, northern Colombia and northern Venezuela
(UNEP, 2004). The Caribbean Sea is the Atlantic Ocean's largest marginal sea (1.52 · 106 km2),
has a mostly microtidal regime and an average depth of 4400 m (Gallegos, 1996).

2.3.1 Triggers of tectonic tsunamis

Historical scenarios
The Caribbean basin is characterized by very active geodynamics. Eastward migration of the Car-
ibbean Plate (CP) relative to the North (NAP) and South (SAP) American Plates (Cenozoic offset
c. 1000 km) creates the Lesser Antilles subduction zone (LASZ) in the east and complex patterns
of strike-slip processes in the north and south (Figure 2-1) (Meschede and Frisch, 1998). Neotec-
tonic uplift during the Holocene is generally moderate (Fairbanks, 1989; Milne and Peros, 2013).
At the LASZ, the NAP and SAP are subducted at a rate of c. 2 cm a−1. Seismic activity occurs
where the plates interact during subduction, deeper down along the dipping slab, and within the
CP (Harbitz et al., 2012). The maximum magnitude of an earthquake at the LASZ in historical times
is estimated at 7.5–8, released by the 8 February 1843 megathrust intraplate earthquake between
Antigua and Grand-Terre (Guadeloupe) (Feuillet et al., 2011). In general, the seismic hazard is
considered to be moderate to large as the age and density of the oceanic lithosphere and the rather
slow subduction rate lower the expected maximum magnitudes (Harbitz et al., 2012). However,
Hayes et al. (2014) found that sufficient strain is accumulated in the LASZ offshore of Guadeloupe
for a tsunamigenic earthquake of Mw ~ 8.2 ± 0.4. In the southeast basin, the Motagua/Swan Islands
Fault System has generated earthquakes of M 7–8 and associated tsunamis in the Gulf of Hondu-
ras on 9 August 1956 and c. 900 CE inferred from coseismically uplifted coastal landforms (Cox et
al., 2008).

Further scenarios, which are commonly associated with the Caribbean tsunami hazard, based on
the historical record, and were reconstructed by means of numerical hydrodynamic models in the
past (Zahibo et al., 2003; López-Venegas et al., 2008; ten Brink et al., 2008; Barkan and ten Brink,
2010; Harbitz et al., 2012) include:

• Devastating shallow earthquakes through strike-slip faulting along the northern boundary
of the CP, exemplified by the 12 January 2010 Haiti earthquake (Bilham, 2010; Fritz et al.,
2013).

• Oblique convergence at the tip of a strike-slip fault, in particular near Hispanola and Puerto
Rico (Grindlay et al., 2005; Harbitz et al., 2012).

• Strike-slip faults or parallel normal fault systems along the steep submarine slopes of the
northern CP boundary potentially activating tsunami-triggering slumps, such as during the
11 October 1918 tsunami in the opening Mona Passage (Moya and Mercado, 2006; Horn-
bach et al., 2008b; López-Venegas et al., 2008). Faults are created due to a difference in
migration rates between the Hispanola (slower) and the Puerto Rico-Virgin Islands-Aves
Island (faster) areas (Mann et al., 2002).

24
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

• Complex tectonic deformation patterns between the Virgin Islands and the Lesser Antilles
island arc, resulting in one of the strongest his-tropical tsunamis of the Caribbean in No-
vember 1867 (Zahibo et al., 2003; Barkan and ten Brink, 2010).

• The CP-SAP boundary, which is a N 100 km transpressional boundary running on- and
offshore. Its main source of seismicity is the dextral strike-slip El Pilar fault. Earthquakes
generated along the submarine fault segments reached M s 7.1–7.3 in 1530 and 1853
(Audemard, 2007) and Mw 7.6–7.7 in 1900 (Colón et al., 2015). However, the resulting
regional tsunamis (O'Loughlin and Lander, 2003; Harbitz et al., 2012) may have been en-
hanced through submarine landslides along the offshore fault segment (ten Brink et al.,
2008).

Numerical modelling of scenarios unprecedented in historical times


Only recently, earthquakes along the active Muertos thrust belt (MTB) extending south of eastern
Hispanola and Puerto Rico were identified as a potential, major tsunamigenic source (Figure 2-1).
The low-angle subduction thrust belt, a deformation zone of 250 km width, is a direct result of
oblique CP-NAP convergence inducing a complex pattern of transtension, transpression, and mi-
croplate tectonics (Granja-Bruña et al., 2014), as well as collision with the southeastern Bahamas
platform (Mann et al., 2002). High-magnitude earthquakes and related tsunamis are anticipated by
Granja-Bruña et al. (2014), if the active thrust faults of the MTB rupture along their entire length.

The South Caribbean Deformed Belt (SCDB) represents an underthrust margin of the Colombian
Plain, the largest abyssal zone of the Caribbean (Draper et al., 1994), and separates the Leeward
Antilles terrane from the CP (Figure 2-1). Similar to the MTB in the north, oblique CP-SAP conver-
gence drives the deformation process, focused along the northern margin of the SCDB (Donovan,
1994). Its potential tsunami hazard remains ambiguous and only very few shallow, minor offshore
earthquakes have been recorded between 1976 and 2007. Nevertheless, ten Brink et al. (2008)
suggest a low-probability, “worst-case” scenario of > 500 km tsunamigenic thrust faulting along the
SCDB, a scenario adapted for the Caribbean-wide tsunami warning exercise in 2013 (UNESCO,
2012).

We created hydrodynamic models of both the MTB and SCDB tsunami scenarios using Delft Dash-
board and Delft3D-Flow (4.01.00) software. Both models have the same rectangular numerical grid
of 597,723 cells (1027 × 583 cells, cell size 5 km × 5 km) and bathymetric setting (GEBCO08). The
grid expands over nearly the whole Caribbean Sea from 6°41′42″N, −87°42′54′ E to 24°11′6″N,
−56°54′18″E. Boundary conditions along the four open boundaries are set to Riemann conditions.
The tsunami wave is generated utilizing the Delft Dashboard tsunami toolbox and the Okada model
(Okada, 1985). It is further based on rupture parameters mainly adapted from ten Brink et al. (2009)
for the MTB, and from ten Brink et al. (2008) and UNESCO (2012) for the SCDB (Table 2-1). The
computational time step is set to 0.01 min. Outputs are derived for 1 and 2 min, respectively, at all
138 observation points. The MTB model output indicates a moderate tsunami hazard with coastal
wave heights mostly below 0.5 m and a maximum of 0.8 m on Aruba (Figure 2-6). In contrast, a
tsunami generated at the SCDB poses a hazard both for the central northern and central southern
Caribbean shores. Wave heights at the coast would reach up to 4.6 m on southern Hispanola, up
to 3.5 m at the Venezuelan coast, and up to 3.2 m along the windward ABC Islands (Figure 2-7).

25
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

Table 2-1: Earthquake and fault parameters used for simulating pan-Caribbean tsunamis originating at the Muertos Thrust
Belt (MTB) and the South Caribbean Deformed Belt (SCBD).

Muertos Thrust Belt (MTB) South Caribbean Deformed Belt (SCBD)


Parameter scenario scenario

7.99 (7.8 given in ten 8.8 (8.5 given in ten Brink et al., 2008; UNE-
M
Brink et al., 2009) SCO, 2012)

188 km (170 km given in ten


Fault length 550 km (ten Brink et al., 2008)
Brink et al., 2009)

Strike 92.1–143.3° 74.5° (ten Brink et al., 2008)

61 km (45 km given in ten


Width 100 km (UNESCO, 2012)
Brink et al., 2009)

Depth 24 km (NGDC/WDS, 2016) 10 km (UNESCO, 2012)

10° (Byrne et al., 1985; ten


Dip 17° (ten Brink et al., 2008; UNESCO, 2012)
Brink et al., 2009)

Slip 3.0 m (ten Brink et al., 2009) 7.8 m

Figure 2-6: Tsunami scenario from a rupture along the western Muertos Thrust Belt (MTB) and maximum heights along
circum-Caribbean coasts. For earthquake parameters see Table 2-1.

26
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

Figure 2-7: Tsunami scenario from a rupture along the South Caribbean Deformed Belt (SCDB) and maximum heights
along circum-Caribbean coasts. For earthquake parameters see Table 2-1.

2.3.2 Volcanism, volcanic edifice failure and landslides

Volcanic tsunamis are triggered by volcano-tectonic earthquakes, volcano edifice collapse, pyro-
clastic flows, underwater explosion and shock waves. They are less frequent than seismic tsuna-
mis (Paris, 2015). Twenty-one active volcanoes occur along the Lesser Antilles island arc. Tsuna-
mis of short periods but pronounced local impact may be generated in the region by direct explosive
activity, onshore or underwater, either by associated atmospheric pressure disturbances, or pyro-
clastic flows, lahars, or debris avalanches entering the ocean (Pararas-Carayannis, 2004). The
submarine Kick'em Jenny volcano, which is currently situated c. 150 m below the water surface off
the Grenadine Islands, currently represents the greatest hazard. Eleven eruptions were recorded
since 1939 which triggered several local tsunamis (Smith and Shepherd, 1995; Pararas-Carayan-
nis, 2004). Possible violent eruptions in the future may reach run-ups of up to 8 m on northern
Grenada and the Grenadines (Smith and Shepherd, 1995).

Large submarine structures of volcanic debris mostly resulting from volcanic flank failure of late
Quaternary age frame several active volcanic islands, such as Montserrat, Martinique, St. Lucia,
St. Vincent or Dominica, and possibly have triggered major tsunamis during downslope transport
(Deplus et al., 2001; Le Friant et al., 2009; Lebas et al., 2011; Watt et al., 2012b; Brunet et al.,
2015). Massive amphitheatre-shaped scarps identified at the southern margin of the Puerto Rico
Trench indicate, according to Grindlay et al. (2005), the occurrence of giant submarine slumps and
tsunamis comparable to the early Holocene Storegga event off Norway. Further substantial mass
wasting deposits have been identified along the coasts of Trinidad and Venezuela (Moscardelli et
al., 2010). On Curaçao, a large and partly subaerial mass failure event created the c. 1 km wide
Caracasbaai and apparently induced a large tsunami at the end of Marine Isotope Stage (MIS) 5e

27
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

(Hornbach et al., 2008a, 2010). Recently, Leslie and Mann (2016) pointed to three massive Neo-
gene-Pleistocene mass-transport deposits offshore northern Columbia potentially indicating an ad-
ditional threat of major tsunamis for the southwest and north Caribbean. Alfaro and Holz (2014)
identified mud diapirism and gas hydrates to foster slope destabilisation in this area.

2.3.3 Teletsunamis

The hazard of teletsunamis generated in the open Atlantic Ocean became evident after the 1755
Lisbon Earthquake, which generated run-ups of up to 6.4 m on the northern Lesser Antilles island
arc (ten Brink et al., 2008; Barkan et al., 2009; Harbitz et al., 2012). Scenarios of potential wave
heights reaching > 20 m on the coast of northern South America and up to 10 m along the Lesser
Antilles island arc following a possible future collapse of the Cumbre Vieja volcano on La Palma,
Canary Islands (Ward and Day, 2001), have been challenged by Hunt et al. (2013), who demon-
strate that failures of such volcanic edifices usually occur in several stages and not as one giant
slide. A recent study, however, indicates a potential hazard of megatsunamis for the Atlantic basin
based on observations of large boulders high up on the slopes of Santiago (Cape Verde Islands),
which supposedly were dislocated after a giant late Pleistocene volcanic flank collapse on the ad-
jacent island of Fogo (Ramalho et al., 2015).

2.4 Alternative processes causing elevated coastal deposits in the Caribbean

2.4.1 Hurricanes

The Caribbean basin is located in the tropics of the northern hemisphere, influenced by the trade
winds, and, for most parts, experiences pronounced wet and dry seasons and constantly high tem-
peratures (Blume, 1962). High atmospheric and sea surface temperatures (SST) further east in the
open Atlantic basin during the second half of the year in combination with decreasing wind shear
support the transformation of easterly waves into warm-core tropical cyclones, which track west-
wards. A significant percentage of them cross the Caribbean as high-category hurricanes (Gold-
enberg and Shapiro, 1996; Hobgood, 2005). The highest hurricane frequencies are temporally as-
sociated with La Niña conditions (Caviedes, 1991), and, geographically, with the Bahamas, the
southeast USA coast, Puerto Rico, the Virgin Islands, and the northern part of the Lesser Antilles
island arc (Figure 2-3). Frequencies decrease in a southwestern direction in roughly concentric
half-circles. Nevertheless, as about ten hurricanes enter the Caribbean each year (Reading, 1990;
Walsh and Reading, 1991), high-category hurricanes and associated storm surges, high waves
and coastal flooding have to be taken into account for almost any coastal section of the circum-
Caribbean when interpreting elevated coastal deposits. Information on fluctuations of hurricane
frequencies for both the historical era (Reading, 1990; Caviedes, 1991; Walsh and Reading, 1991;
Chenoweth and Divine, 2008) and the last millennia (e.g. Donnelly and Woodruff, 2007; McCloskey
and Liu, 2013a; see synthesis in Adomat and Gischler, 2016) is controversial, whereas the last
decades seem to have seen an increase in mean intensities and frequencies of high-category
hurricanes in the North Atlantic basin related to elevated SST and associated atmospheric dynam-
ics (Knutson et al., 2010).

28
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

2.4.2 Relative sea-level (RSL) changes

Relative sea level (RSL) is defined as “the level of the sea with respect to land” at a certain location
(Lambeck, 2002, p. 33). Understanding RSL histories is crucial to the interpretation of the palae-
otsunami record (Morton et al., 2006), as marine deposits located in elevated positions today might
have been formed in an intertidal position several millennia ago where RSL falls. In the Caribbean,
RSL was 121 ± 5 m below current mean sea level (b.s.l) during the Last Glacial Maximum 20 ka BP
(Fairbanks, 1989). RSL rose quickly after 15 ka BP until the early Holocene (Fairbanks, 1989;
Toscano and Macintyre, 2003). A slowdown of global ice ablation around 7 ka BP stimulated the
influence of regional (e.g., glacioisostatic adjustment, gravitational effects inducing deformation of
the Earth, upper/lower mantle viscosity) and local (lithosome compaction, vertical neotectonic dis-
placement) factors, and diverging RSL histories evolved (Jackson, 2013; Milne and Peros, 2013).
While RSL rose between 1 mm year−1 in the eastern Caribbean basin and 2.5 mm year−1 in the
southwestern part from 7000 to 4000 years ago, rates dropped below 0.5 mm year−1 in the past
500 years (Jackson, 2013). However, no RSL higher than today has been reported for the Holo-
cene of the Caribbean basin (Fairbanks, 1989; Toscano and Macintyre, 2003; Jackson, 2013; Milne
and Peros, 2013). Therefore, elevated marine deposits of Holocene age can generally be attributed
to extreme-wave conditions.

2.5 Tsunami deposits of the Caribbean

By reviewing almost 59 published coastal sedimentary records for potential fine-grained (Table A-
3) and coarse-clast (Table A-4) tsunami deposits, we divided the Caribbean basin into five different
sectors (Figure 2-3). The sedimentary evidence was classified into six categories following the
interpretation given in the original source:

• 0 = no allochthonous deposits or disturbances present in the stratigraphy;


• 1 = allochthonous deposits or other disturbances are ascribed to storm impact or gradual
coastal changes. Tsunamis are excluded by the authors;
• 2 = allochthonous deposits are ascribed to storm rather than tsunami, but the latter is not
entirely excluded;
• 3 = allochthonous deposits are ascribed to either storm or tsunami;
• 4 = allochthonous deposits are ascribed to tsunami rather than storm, but the latter is not
entirely excluded;
• 5 = allochthonous deposits are ascribed to tsunami and storm is excluded.

Only evidence placed into categories 3 to 5 is presented and discussed here in detail.

The precision and scope in documentation of the traces of extreme-wave events in fine-grained
sedimentary archives varies strongly. Literature sources range from short abstracts to highly tech-
nical reports. Studies focusing on palaeoclimate or palaeoecology, i.e., long-term or periodic sedi-
mentation patterns, often provide only limited information on layers of episodic high-energy sedi-
mentation and assign a lower priority to their interpretation (e.g., Dix et al., 1999; Donnelly and
Woodruff, 2007; Dalman and Park, 2012; Brooks et al., 2015; Caffrey et al., 2015).

Documentation and process-related interpretation of the coarse-clast record, mostly located on


rock platforms from the intertidal level to elevations of up to 20 m, usually focuses on extreme-

29
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

wave events and coastal hazard assessment. Coarse-clast features have regularly been used to
infer palaeotsunamis, even though intense discussions arose in some cases and diverging conclu-
sions on the origin of the deposits exist (e.g., Grand Cayman, ABC Islands) (Jones and Hunter,
1992; Scheffers, 2002b, 2004, 2005; Morton et al., 2008a; Spiske et al., 2008; Rowe et al., 2009;
Khan et al., 2010).

2.5.1 NW sector

A wedge-shaped berm supposedly deposited around 1500 years ago is the only sedimentary fea-
ture from the NW sector linked with a tsunami (Shaw and Benson, 2015, Site 36 on Figure 2-3). It
has a thick base layer of unstructured coarse sand and floating coarse clasts with overlying whitish
and densely packed, reddish brown sand, gravel, and boulders. The berm extends about 50 km
along the east coast of the Yucatán Peninsula. It reaches > 4 m above mean sea level on head-
lands, where it is covered by numerous imbricated boulders with a-axes of up to 1 m length on the
seaward side. At headlands, it reaches up to 150 m in-land, where it gradually becomes finer,
whereas behind low-lying beaches, berm deposits can be found up to 400 m from the shore. Gaps
in the berm are attributed to channelized tsunami backflow (Shaw and Benson, 2015).

2.5.2 N sector

Grand Cayman
On Grand Cayman, Jones and Hunter (1992) describe boulders with a-axes of up to 5.5 m weigh-
ing up to 40 t (Site 37, Figure 2-3). They are located at a distance of up to 100 m from the shoreline
on an elevated carbonate platform. At Blowholes (southeast Grand Cayman), boulders are be-
lieved to have been shifted onshore c. 330 years ago based on 14C ages of coral attached to a
boulder, and were not moved since then. Boulders gather landward of shore-perpendicular inden-
tations on the carbonate platform, where waves seem to become funnelled. Storm waves or tsu-
namis are equally considered as being capable of dislodgement (Jones and Hunter, 1992). Robin-
son et al. (2006) and Rowe et al. (2009) revisited the sites after Hurricane Ivan in 2004 and found
many clasts relocated and partly overturned as indicated by exposed fresh, white surfaces and
downward facing grass patches. Besides solitary boulders and boulder clusters, coastal deposition
on Grand Cayman is dominated by pebble and boulder ramparts and mixed sand-to-boulder ridges,
in particular along the south coast (Rigby and Roberts, 1976; Hernandez-Avila et al., 1977; Jones
and Hunter, 1992).

Jamaica
Very similar patterns were found in Jamaica (Site 39, Figure 2-3). Boulders at Galina (northeast
Jamaica) derive from the cliff edge, weigh up to 120 t, and are clustered in a zone 80–160 m from
the shore (Robinson et al., 2006), fining landward (Rowe et al., 2009). The boulders are backed by
a ridge of polymodal debris, sand and mollusc shells, which is covered by boulders and separates
the boulder field from the densely vegetated area inland (Robinson et al., 2006; Morton et al.,
2008a; Rowe et al., 2009; Miller et al., 2014). While Robinson et al. (2006) describe their internal
architecture as chaotic, Morton et al. (2008a) found particle sorting and distinct stratification in ridge
exposures. At other sites, boulders of up to 200 t were found gathering inland of shoreline inden-
tations (Robinson et al., 2006). Based on eyewitness reports and surveys after the passage of
Hurricane Dean in 2007, where Khan et al. (2010) identified a limestone clast of 80 t having shifted
laterally for 2 m at an elevation of 12 m above current mean sea level (a.s.l.) near Manchioneal,

30
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

east Jamaica, it is anticipated that most boulders have a complex heritage of dislocation potentially
covering the last 4000–5000 years and that hurricane-induced waves may be capable of shifting
even the largest boulders horizontally for a short distance. Displacement by tsunami is not ex-
cluded, in particular in the case of the largest boulders, but no direct evidence exists so far (Rowe
et al., 2009; Khan et al., 2010).

In addition to the coarse clast records, fine-grained deposits of extreme-wave overwash were iden-
tified in sediment cores from the enclosed lagoon of Manatee Bay on the Jamaican south coast
(Site 10, Figure 2-3). Whether the bioclastic carbonate sands deriving from the off-shore coastal
zone represent storms or tsunamis is uncertain (Palmer and Burn, 2012).

Bahamas
Near Glass Window on northern Eleuthera, Bahamas (Site 40, Figure 2-3), Hearty (1997) was the
first to investigate very large oolitic-peloidal boulders weighing up to 2330 t (Hearty, 1997), later
corrected to 958 t by Viret (2008). The boulders rest on aeolianite covered by oolitic entisols at
elevations of up to 20 m a.s.l. High-angle bedding, a lack of possible alternative transport mecha-
nisms, and absence of source areas higher than the boulders are cited as evidence for transport
by extreme waves. Based on stratigraphic evidence, fossil soils and inverse amino acid racemisa-
tion (AAR) ages, transport at the end of MIS 5e was inferred (Hearty, 1997; Viret, 2008; Kindler et
al., 2010). Along with the heavily disputed “chevron ridges” (Hearty et al., 1998, 2002; Kindler and
Strasser, 2000; Kelletat et al., 2004), the giant boulders were used as evidence for “larger and
more frequent cyclonic storms in the North Atlantic than those seen today” (Hearty and Neumann,
2001, p. 1892; Hansen et al., 2016, p. 3784). Others inferred tsunami deposition during MIS 5e
based on numerical modelling of a local submarine slope failure event and the assumption that
storm waves have a limited transport capacity (Samankassou et al., 2008; Hasler et al., 2010).

At the east coast on southern Eleuthera (Site 41, Figure 2-3), low ridges of beachrock and aeo-
lianite slabs with a-axes of up to 3 m have formed low, imbricated ridges at elevations of up to
3 m a.s.l. overgrown by trees. Kelletat et al. (2004) assume that these ridges have remained stable
for centuries. Angular boulders weighing up to 5 t are also found on the leeward side on top of an
elevated carbonate platform. Further north, boulders of up to 30 t and a bimodal deposit of car-
bonate sand, shells and coral branches were documented behind a 15 m-high cliff and inland slop-
ing platform. Soil formation within the sand, mature vegetation cover, and two 14C ages indicate
formation at least several hundred years ago. The Whale Point peninsula in the north also shows
an almost continuous shore-parallel bimodal deposit and boulders of up to N 300 t, with shore-
parallel a-axes, mostly imbricated, arranged into ridges or fields, and separated from the shoreline
by a sediment-free, 80–100 m-wide zone. Very similar findings were made on Long Island (Site
42, Figure 2-3). All deposits are ascribed to two palaeotsunamis generated in the open Atlantic
Ocean around 3000 and 500 years ago (Kelletat et al., 2004).

31
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

Dominican Republic
Coastal lagoons at Bahia de Ocoa, Dominican Republic (Site 15, Figure 2-3), reveal overwash fans
and three coarse sand deposits in multiple cores at varying depths, with the beach and shoreface
as source areas. The overwash deposits, for which a tsunami origin is possible, also contain or-
ganic-rich clasts (Fuentes and Huérfano-Moreno, 2013). At Playa Cosón, northeast coast (Site 17,
Figure 2-3), a light carbonate sand layer extending up to 260 m inland was found vertically confined
by non-carbonate silty clay and a thin top layer of sandy loam. It is dominated by shallow-marine
foraminifera, whereas low numbers of bathyal and abyssal taxa occur as well (Scheucher et al.,
2011). At Puerto Viejo, southwest coast (Site 16, Figure 2-3), carbonate sand with gravel compo-
nents unconformably overlies carbonate-containing loamy clay. It occurs at depths of up to 85 cm
and mostly contains shallow-marine benthic foraminifera (one taxon preferring bathyal/abyssal
depths), as well as an unusually high amount of planktonic taxa (c. 30 %). The deposits were ten-
tatively associated with the 8 August 1946 tsunami (Playa Cosón) and the 18 October 1751 tsu-
nami (Puerto Viejo) (Scheucher et al., 2011).

Puerto Rico
In sediment cores from Aguada Plain at Carrizales, Aguada Swamp at Espinar and Aguadilla
Swamp, northwest Puerto Rico (Site 18, Figure 2-3), Moya and Mercado (2006) found several thin
marine overwash deposits within background sediments of mud to very fine sand, all of which were
linked to tsunamis. Behind the beach on coastal plains and in swamps up to c. 100 m away from
the shore, thin laminated layers containing Halimeda sp. fragments and hematite suggest the
beach as the main source area. A correlation of layers is possible between cores from one site,
but also across sites. Two 14C ages allow maximum age estimates of about 2770–2350 BP and
680–540 BP to be inferred for the corresponding events, whereas the uppermost deposit (> 40 cm
below surface in several cores) is associated with the tsunami of 11 October 1918. Morton et al.
(2006) revisited the area and reported a layer of grey sand at a depth of 127–137 cm with basal
concentrations of heavy minerals within a sequence of sandy mud in a mangrove wetland. Further
to the east, they identified two thin strata of grey sand – the upper one showing shallow marine
gastropods on top – within background muds in a wetland environment behind a breached eolianite
ridge and narrow coastal plain. At Punta Cucharas, south coast (Site 19, Figure 2-3), a grey sand-
and-shell layer with a thickness of only 2 cm was found 1 m below a tidal flat. Depositional pro-
cesses remain uncertain (Morton et al., 2006).

The coastal boulder record of northwest Puerto Rico (Site 44, Figure 2-3) comprises isolated rocks
with a-axes of up to 1 m at a distance of 70 m from the beach inside a small creek at Punta
Borinquen and a large coral (a-axis = 1.2 m) behind a breached, active coastal dune east of Punta
Agujereada. Based on limited storm surge potential in the area, due to a very narrow shelf, and a
strong impact of the 11 October 1918 tsunami in the Mona Passage (Reid and Taber, 1919; López-
Venegas et al., 2008), Moya and Mercado (2006) assume that these boulders were deposited by
tsunami waves. At Isla Mona (Site 45, Figure 2-3), offshore of northwestern Puerto Rico, limited
information is available on several large reefrock clasts (a-axes up to 5 m) dislodged from the Hol-
ocene reef onto late Pleistocene coastal flats along the southwestern coast either during tsunamis
or hurricanes. U/Th dating indicates deposition sometime after 4176 years ago (Taggart et al.,
1993; Gonzalez et al., 1997).

32
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

British Virgin Islands


On Anegada (Site 23, Figure 2-3), a multi-proxy study on disturbed beach ridges and back-barrier
sediments points to an extreme-wave impact between 1600 and 1850 cal CE. Candidate events
include the Antilles tsunami of 1690, the Lisbon teletsunami of 1755, a previously undocumented
tsunami, or a storm surge with a magnitude greater than the most severe hurricane of the past
decades. The event caused overwash and breaching of sandy barriers and deposition of a sand-
and-shell sheet (Figure 2-8-II), now covered by carbonate mud, as well as boulders and cobbles
distributed over a large area. The sand deposit thins and fines inland to a distance of up to 1.5 km
from the breached ridges. The overwash process and associated erosion and sedimentation seem
to have transformed marine inlets into hypersaline, restricted ponds (Atwater et al., 2012). The
taphonomy of shells and encrusting reefal foraminifera (Homotrema rubrum) indicates that source
areas of the sand- and-shell sheet comprise pre-existing storm wrack deposits (sorted beach) and
the reef (Pilarczyk and Reinhardt, 2012; Reinhardt et al., 2012). Two fields of cobbles and boulders
(c. 300 and 800 m away from the coast) (Site 47, Figure 2-3), embedded in the sand-and-shell
deposit (Atwater et al., 2012), derive from adjacent inland outcrops of Pleistocene limestone. The
wide scatter of boulders and absence of ridge-like structures, their distribution far inland, and the
application of sediment transport and coastal inundation modelling indicate that the boulder fields
were likely created by a regional tsunami following a high-magnitude, Puerto Rico Trench outer-
rise earthquake (Buckley et al., 2012; Watt et al., 2012a). Furthermore, 14C ages derived from outer
bands of large coral heads transported inland at several sites on Anegada (Figure 2-8-I) cluster
around 1200–1450 cal CE. Their dislocation is associated with breaches cut into beach ridges and
points to a medieval tsunami according to Weil Accardo et al. (2012) and Atwater et al. (2013b,
2014). Papers in preparation compile further sedimentary evidence and re-evaluate the tsunami
hypothesis for the deposits of 1200–1450 cal CE and 1600–1850 cal CE (B. Atwater, pers. comm.).

33
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

Figure 2-8: Examples of candidate tsunami deposits from Anegada (Sites 23/47, Figure 2-3) and Bonaire (Site 31/55,
Figure 2-3) (compare with archive categories A–F in Figure 2-4). I) Large coral head (Diploria sp.) possibly transported
onto an inland salt flat on Anegada by a tsunami. Periodic flooding of the salt flat during high-category hurricanes is
indicated by a cover of dried microbial mat (photo courtesy of B. Atwater). Calibrated 14C ages measured on this and
other large corals transported inland lie in the range of 1200–1450 CE (Atwater et al., 2014, online supplement). II)
Cerith-dominated mollusc assemblage of the sand-and-shell sheet on Anegada, sieved through 4 mm mesh (photo
courtesy of B. Atwater). The deposit was dated to 1650–1800 CE and could represent either a historical tsunami (e.g.,
1690 or 1755) or an extremely violent, unknown storm (Atwater et al., 2012). III) Tsunami deposit from Boka Bartol on
Bonaire (6.86–6.67 m below surface) 14C-dated to 3300–3000 cal BP (EWE II). It shows several criteria associated with
tsunami deposition and had significant long-term impacts on the local hydroecology by transforming a narrow, mangrove-
fringed open embayment into a hypersaline closed lagoon (Engel et al., 2013). IV) Extensive boulder field near Spelonk
(e.g., Scheffers, 2002b, 2005; Spiske et al., 2008; Engel and May, 2012) on top of an elevated palaeo-reef terrace. Large
singular boulders are possibly dislodged by tsunamis.

34
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

2.5.3 E sector

The Lesser Antilles island arc was subjected to extensive, pioneering surveys of coarse-clast high-
energy wave deposits, most of which were interpreted as tsunamigenic. They comprise:

• Boulder ridges and fields on elevated carbonate platforms (up to 15 m a.s.l.) on Anguilla
and Scrub Island (Site 48, Figure 2-3), in particular along the southeast and east coasts.
The sediment-free zone between the coast and the coarse-clast deposits is used as evi-
dence for hurricanes to predominantly induce erosion. Based on 14C ages, two strong tsu-
namis approaching from the open Atlantic are inferred for around 500 and 1600 years ago
(Scheffers, 2006b; Scheffers and Kelletat, 2006).

• Boulders weighing up to 10 t on St. Martin (Site 49, Figure 2-3) and a “tsunami boulder
spit” at an offshore volcanic island 14C-dated to 500 BP (Scheffers, 2006b; Scheffers and
Kelletat, 2006).

• Boulders up to 5 t scattered at elevations of 2–5 m a.s.l., 30–40 m inland, and a mixed


deposit of sand, shells and coral rubble along with boulders of up to 30 t (up to 10 m a.s.l.)
at the east coast of Grand Terre, Guadeloupe (Site 50, Figure 2-3), indicating tsunami
impact 2400 years ago based on pedogenesis and 14C data (Scheffers et al., 2005).

• Well-rounded basaltic components floating in a volcanic ash matrix at elevations of up to


50 m a.s.l. with exposures of 100 m and a thickness of at least several metres on St. Lucia
(Site 51, Figure 2-3), interpreted as deposits of middle Pleistocene tsunamis (Scheffers et
al., 2005).

• Large singular boulders – one of them weighing up to 170 t is overturned and located
13 m a.s.l. and 30 m away from the cliff edge – boulder ridges, boulder imbrication and
unstable settings with vertical a-axes on Barbados (Site 52, Figure 2-3), attributed to two
tsunamis approaching from the open Atlantic Ocean at around 1500 and 4500 BP
(Schellmann and Radtke, 2004; Scheffers, 2006b; Scheffers and Kelletat, 2006).

• Well-rounded cobbles and boulders in a sand matrix between 0 and 3 m a.s.l., presumably
of Pleistocene age, are covered by tephra at the northwest coast of Grenada (Site 53,
Figure 2-3). Near Halifax Harbour, west coast, a 200 m-long, 2–4 m-high boulder ridge,
consisting of weathered clasts of up to 10 t, is related to a late Holocene local tsunami
supposedly induced by an eruption of the Kick'em Jenny volcano (Scheffers et al., 2005).

2.5.4 S sector

Venezuela
A first multiproxy study from Laguna los Patos, Cumaná (Site 30, Figure 2-3), provides evidence
for marine incursions based on geochemical signatures, grain size and the erosive base of layers
in the uppermost part of the studied sediment core. The signatures were interpreted to have prob-
ably been induced by historical tsunamis (Leal et al., 2014). Oropeza et al. (2015) present silty
sand layers, some of which show landward directed ripples and salt lenses from three lagoons of
the Araya Peninsula, eastern Venezuela (Site 29, Figure 2-3), which are interpreted as possible
traces of tsunami flooding. Another candidate tsunami deposit is reported from the islet of Cayo
Sal off the coast of western Venezuela (Site 31, Figure 2-3), where a massive, ex-situ deposit of

35
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

skeletal muddy sand is associated with an inversion of the sediment core's age model (Weiss,
1979). 14C data indicate an event of the last 600 years.

Schubert (1994) documents coral branches of pebble to boulder size distributed over a terrace of
Mesozoic metamorphic rocks (15–18 m a.s.l.) west of Puerto Colombia, Venezuela (Site 55, Figure
2-3). The coral material has an age of 1300 years based on U/Th and 14C data. Deposition of the
clasts is ascribed to a tsunami, mainly based on comparison with intertidal to supratidal, storm-built
polymodal ridges on La Orchila Island (Schubert and Valastro, 1976). Tsunamis are assumed to
be induced by either strong, local submarine earthquakes, by offshore slumping, or a combination
of both (Schubert, 1994).

ABC Islands (Aruba, Bonaire, Curaçao)


The core record from several onshore coastal depressions (e.g., Figure 2-8-III) along the entire
coast of Bonaire (Sites 32, 56, Figure 2-3), further off the coast of western Venezuela, contains
layers of reworked marine and littoral deposits, and those mixed with terrestrial deposits as indica-
tion of past extreme-wave events. Their sedimentary characteristics comprise one or more nor-
mally graded or non-graded sequences separated by mud caps, rip-up clasts, horizons enriched
with heavy minerals, foraminifera reflecting a dominant shallow-marine sediment source with minor
percentages from deeper waters, low-density particles (wood, plant material, light shells) rafted on
top, and erosive basal contacts (some features visible in Figure 2-8-III) (Engel et al., 2010, 2012,
2013). Source areas of the deposits were identified as shallow areas off the coast, whereas locally,
terrestrial sediment was incorporated as well. Following the lines of evidence in Engel et al. (2012,
2013), the oldest potential tsunami deposit was identified on Klein Bonaire (3.6 ka BP). A well-
preserved one from Boka Bartol with a maximum age of 3.3 ka BP (Figure 2-8-III) has counterparts
on the leeward coast (Klein Bonaire, Saliña Tam, possibly between Saliña Tern and Punt'i Wekua)
and the windward coast (Playa Grandi, possibly Boka Washikemba) and represents the best doc-
umented potential palaeotsunami on Bonaire. A post-2 ka BP tsunami left massive deposits at La-
gun (Engel et al., 2010) and Saliña Tam. Sediments laid down by another younger unspecified
extreme-wave event (1300–500 BP) were found between Saliña Tam and Punt'i Wekua, and at
Boka Washikemba (Engel et al., 2010, 2012).

The ABC Islands also have an intensely studied record of coarse-clast deposits. Coastal “ridges
[…] of loose blocks” (Figure 2-9-I–V) have long been associated with a higher mid-Holocene sea
level (De Buisonjé, 1974) or storms (Zonneveld et al., 1977) without generating too much attention.
Since the late 1990s, the ridges (Figure 2-9-I–IV), ramparts (sensu Focke, 1978; Scheffers, 2005)
or ridge complexes (Figure 2-9-V) (sensu Morton et al., 2008a), respectively, as well as singular
boulders and boulder fields (Figure 2-8-IV) were subjected to systematic investigations (Scheffers,
2002a, b, 2004, 2005; Scheffers et al., 2006, 2009, 2014; Radtke et al., 2003; Morton et al., 2006;
Spiske et al., 2008; Pignatelli et al., 2010; Watt et al., 2010; Engel and May, 2012). Narrow ridges
of unimodal, mostly rounded coral fragments (mostly branches of Acropora cervicornis) based in
intertidal position or close to sea level, with a height of up to 4 m, steep flanks, avalanching features
(Figure 2-9-I– IV,VI) and – where those features are low – distinct washover lobes at their landward
sides, dominate the leeward coasts of all three islands (Scheffers, 2002b) and are deposited and
modified by storm waves (Scheffers, 2002b; Morton et al., 2008a; Scheffers et al., 2014). Along
the windward coasts (eastern and northern exposure), up to 400 m wide polymodal ridge com-
plexes of corals boulders (mostly large branches of Acropora palmata), Pleistocene reefrock,
shells, and sand with a steep seaward slope, tapering out inland at a very low angle prevail on top

36
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

of elevated Pleistocene platforms (Figure 2-9-V,VI). Scheffers (2002b, 2004, 2005, 2006a) associ-
ated their formation with palaeotsunamis generated at the LASZ, the CP-SAP boundary or the
open Atlantic Ocean based on their large inland extent, coarseness of the particles, wide particle
size distributions, the seemingly chaotic internal structure, and large ripple marks identified on aer-
ial imagery. However, Morton et al. (2006, 2008a) showed that textural sorting, imbrications of platy
particles, stratigraphic organization, and even initial palaeosols may be found within the wide ram-
parts (Figure 2-8-V). They concluded, based on comparison with ridge-like features elsewhere,
that the ramparts were the result of multiple extreme-wave events over millennial scales and that
they were not indicative of a certain hydrodynamic process, but only reflect the range of available
source material. Based on age clusters of an initial ESR (Radtke et al., 2003) and 14C dataset
derived from coral branches of the ramparts, Scheffers (2002b, 2004) and Scheffers et al. (2009)
inferred three tsunami impacts at around 500, 1500, and 3500 BP.

Diverging conclusions also existed on the origin of the largest singular boulders on Bonaire (Figure
2-8-IV) (Scheffers, 2004; Spiske et al., 2008), whereas later a general consensus was reached that
the largest boulders carry significant characteristics of tsunami dislodgement (Scheffers, 2002b,
2005; Pignatelli et al., 2010; Watt et al., 2010; Engel and May, 2012). The application of initiation
of motion criteria after Nott (2003) and Nandasena et al. (2011a) indicates that transport by storm-
induced waves is unlikely for the largest boulders of up to 150 t. At the most extensive boulder
fields of Bonaire, boulders and blocks are located on top of an elevated Pleistocene reef platform
(3.5–5 m a.s.l.) at a distance of 40–120 m from the coast (Pignatelli et al., 2010; Engel and May,
2012), whereas recent category 5 hurricanes such as Ivan in 2004 only shifted small clasts close
to the shore (Scheffers and Scheffers, 2006).

On Curaçao and Aruba, the subaerial coarse-clast record is similar to Bonaire with a smaller spatial
extent (Sites 57–58, Figure 2-3). Several tsunamis were inferred by Scheffers (2002b, 2004). A
palaeoenvironmental record from St. Michiel lagoon on Curaçao (Site 33, Figure 2-3) indicates
closure by a barrier of coral rubble at around 3500 BP (or later) based on an abrupt transition from
an open marine to a euryhaline biofacies. Rapid formation of the barrier by a potential tsunami was
taken into account by Klosowska (2003) without further analysis or discussion. An underwater sed-
iment core from Spaanse Waters bay (Site 34, Figure 2-3) contains a “chaotic mixed layer” dated
to 460–310 cal BP (or younger). It contains rock components, terrestrial plant remains, mollusc
shells and older reef material, and was linked to a tsunami by Hornbach et al. (2008a: p. 44). A
short core from Fuik Bay reveals three layers – the lowermost one dated to 500–320 cal BP– con-
taining shell fragments, which were linked to either storm or tsunami events (Hornbach et al.,
2008a).

37
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

Figure 2-9: Examples of storm-induced deposits and landforms from Bonaire (Site 55, Figure 2-3) (compare with
archive categories A–F in Figure 2-4). I) Narrow intertidal ridge of coral rubble formed or at least significantly
modified during Hurricane Ivan, northwestern Bonaire (cf. Scheffers and Scheffers, 2006). II) Sequence of coral
rubble ridges on the low supratidal platform of Klein Bonaire (photograph courtesy of D. Kelletat) formed mostly
by storms over several millennia (Scheffers et al., 2014). III) Supratidal coral rubble ridge near Salina Tern,
northwestern Bonaire, with internal stratification and palaeosol horizons also indicating a multi-phase for-
mation during storms (Engel et al., 2012). IV) Broad rampart/ridge complex of the Washikemba area.
Stratification and textural sorting in a similar ridge complex near Boka Onima (photo insert courtesy of U.S.
Geological Survey, modified) indicates a multi-phase formation (cf. Morton et al., 2008a). V) Cross profiles
show the geomorphic differences between the broad windward ramparts (red) and the more narrow leeward
ridges on Bonaire (Scheffers et al., 2014), which are predominantly a function of the width of the supratidal
platform and approaching direction of major storms.

38
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

2.6 Discussion

2.6.1 Review of extreme-wave deposits

In this chapter, the Caribbean records of potential tsunami deposits are re-evaluated based on
state-of-the-art models of extreme-wave deposition as summarized in Table A-1 and Table A-2.

Fine sediments
At Carrizales, Puerto Rico, successful correlation of the two layers in a 25 m-long trench, the iden-
tification of more well-preserved shells and coral fragments on the trench scale, and the fact that,
according to survivors, the trench site was flooded by the 11 October 1918 tsunami, the tsunami
hypothesis is reasonable. At Espinar, distinct laminated sub-units and the observation of tsunami
flooding in 1918 are the strongest arguments pro tsunami deposition. The sites chosen by Moya
and Mercado (2006) are promising for using tsunami deposits to reconstruct frequency-magnitude
patterns of tsunamis in the wider area of the Mona Passage, but more detailed multidisciplinary
analyses of the event layers and more dating are required to verify their tsunami origin, to infer
hydrodynamic characteristics of the marine incursion and to foster an event chronology.

The record from Anegada seems reliable as well, as the widespread presence of ex-situ articulated
bivalves and angular fragmentation in the sand-and-shell deposit (Reinhardt et al., 2012) and its
sheet-like distribution up to > 1 km inland are congruent with models of tsunami-laid sediments.
Minor depositional effects of strong hurricanes of the past decades (Atwater et al., 2012) and a
relatively high seismic tsunami potential (Parsons and Geist, 2009) further support this evaluation.

The sediments from around Bonaire also meet several criteria associated with the occurrence of
tsunamis as listed above, and age control points to up to four major events (Figure 2-10). However,
due to the lack of recent local tsunami deposits for comparison, storm surges and waves exceeding
those of modern hurricanes in the southern Caribbean cannot entirely be ruled out as a possible
source. Whereas the SCDB tsunami scenario (Figure 2-7) would probably create sufficient inun-
dation to match the spatial distribution of fine-grained candidate tsunami deposits identified so far
on Bonaire, the largest historical scenarios of the southern Caribbean seem to have only moder-
ately affected the island (Oetjen et al., 2015).

The records from the Dominican Republic (Scheucher et al., 2011; Fuentes and Huérfano-Moreno,
2013), Manatee Bay, Jamaica (Palmer and Burn, 2012), and Curaçao (Klosowska, 2003; Hornbach
et al., 2008a) require more details on sediment texture, composition and spatial extent before their
depositional process can be properly evaluated. To date, only conference abstracts are available.
All sites are exposed to potential regional tsunami sources and thus, further in-depth research is
advocated. Similarly, the records from Venezuela (Weiss, 1979; Leal et al., 2014; Oropeza et al.,
2015) only provide first insights. Based on the large number of local historical tsunamis which the
sediments could be linked with, the tsunami hypothesis seems reasonable.

39
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

Figure 2-10: Compilation of dated candidate tsunami deposits of Holocene age. Sites are indicated in Figure 2-3
and details and citations of records are given in Tables A-3 and A-4. Abbreviations for sites on Bonaire: SPE =
Spelonk; KLB = Klein Bonaire; SAT = Saliña Tam; NUK = Nukove. For different areas on Bonaire age-frequency
plots of corrected ESR- and 14C–dated coral rubble from coastal landforms accumulated by extreme events are
displayed. For age-frequency plots bin size is 300 years; the y-axis shows the number of age datings inside each
bin (for details see Scheffers et al., 2014).

Coarse sediments

Solitary boulders and boulder clusters


Boulders are well known from both storm and tsunami deposition (Table A-2) (Etienne et al., 2011;
Richmond et al., 2011). On Grand Cayman (Jones and Hunter, 1992; Robinson et al., 2006) and
Jamaica (Robinson et al., 2006; Khan et al., 2010), they primarily indicate the impact of storm
waves and surge, as the strong storms seem to be capable to shift even some of the largest boul-
ders (Robinson et al., 2006; Rowe et al., 2009; Khan et al., 2010). Even though tsunamis cannot
be excluded, their effects might have been overprinted by the much more frequent high-category
hurricanes as convincingly summarized by Khan et al. (2010).

40
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

Coastal boulders on Eleuthera are much bigger and older (MIS 5e) than anywhere else in the
Caribbean. Considering the entire spectrum of evidence, the boulders may either relate to late
Pleistocene tsunamis (most likely generated in the near-field), major hurricanes, or a combination
of both (Engel et al., 2015). Similarly, the large singular boulders at Whale Point peninsula de-
scribed by Kelletat et al. (2004) show consistencies with both tsunami and storm-wave deposition.
The latter process should not be ruled out according to potentially high storm surges during hurri-
canes of around 4.5 m, and a very active wave climate with waves overtopping the > 10 m-high
platform several times a year (Anglin and MacIntosh, 2005).

The isolated boulders from the beaches nested between elevated coastal terraces near Punta
Agujereada on Puerto Rico might have a multicausal origin as well. However, based on observa-
tions on boulder transport in the area during the 1918 tsunami (Reid and Taber, 1919), Moya and
Mercado (2006) are most certainly correct in suggesting that tsunamis were involved.

The tsunamigenic origin of the boulder scatter and large coral heads transported inland (Figure
2-8-I) on Anegada is also conclusive as it complies with state-of-the-art depositional models (Watt
et al., 2012a) and is backed by inverse sediment transport modelling (Buckley et al., 2012), corre-
lating fine deposits (Reinhardt et al., 2012) and documentary records (Atwater et al., 2012).

Some of the massive singular cliff-top boulders from the E sector – e.g., the one near Bottom Bay
on Barbados (Schellmann and Radtke, 2004; Scheffers and Kelletat, 2006) – are consistent with
patterns of both tsunami and storm deposition. Both processes need to be considered as the haz-
ard imposed by both types of event is high in this region (Reading, 1990; Parsons and Geist, 2009).
The contribution of a tsunami to the formation of the “block-and-ash flows” on St. Lucia (cf. Sig-
urdsson et al., 1980) as suggested by Scheffers et al. (2005) remains equivocal.

The pebble- to boulder-sized corals reported from Venezuela (Schubert, 1994) are difficult to eval-
uate as only limited information is given. Since they are definitely of marine origin, of Holocene
age, and located high up in an area of low probability of hurricane landfall but significant local
tsunami potential, the tsunami hypothesis seems reasonable.

Tsunami deposition of large boulders on Bonaire (Figure 2-8-IV) seems likely and was suggested
based on the large size, scattered distribution (Scheffers, 2002b, 2004, 2005; Watt et al., 2010),
the application of initiation-of-motion criteria (Pignatelli et al., 2010; Engel and May, 2012), and the
minor effects of recent high-category hurricanes passing the island (Scheffers and Scheffers,
2006). However, in view of the clear asymmetry of clast volume and size of individual boulders
between the windward and leeward sides of the ABC islands and recent observations that ex-
tremely large boulders can be deposited in a scattered pattern during tropical cyclones (May et al.,
2015b), a multicausal origin including the impact of strong, landfalling hurricanes cannot be ex-
cluded.

Ridges and ramparts/ridge complexes


Mostly polymodal ridges and ridge complexes (Morton et al., 2006, 2008a) or ramparts (Scheffers,
2002b, 2004, 2005; Scheffers et al., 2014) in varying elevations and distances from the coast seem
to be the most common depositional landform along windward Caribbean coasts. Their height and
width depends on elevation, shore-perpendicular topography, and source material. It is generally
agreed that they mainly form during storms (Morton et al., 2008a), as observed in recent examples
in the Caribbean (Scheffers, 2005; Spiske and Jaffe, 2009) and elsewhere (Table A-2) (Richmond

41
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

et al., 2011; Reyes et al., 2015). Their widespread occurrence on Grand Cayman (Rigby and Rob-
erts, 1976; Hernandez-Avila et al., 1977; Jones and Hunter, 1992), Jamaica (Rowe et al., 2009;
Miller et al., 2014), some of the eastern Bahaman islands (Kelletat et al., 2004), and the entire E
sector (Aguilla, St. Martin, Guadeloupe, Barbados, St. Lucia, Grenada) (Schellmann and Radtke,
2004; Scheffers and Kelletat, 2006; Scheffers et al., 2005) is best explained by the high exposure
to elevated wave action induced by strong, westward-moving hurricanes (Reading, 1990; Walsh
and Reading, 1991), and indicates that any subaerial deposit of tsunamis in these areas is at high
risk to become overprinted by the effects of recurring storm surge and waves. Therefore, after
intense discussions on the polymodal ramparts of Bonaire, Curaçao, and Aruba over more than a
decade (Scheffers, 2002b, 2004, 2005; Morton et al., 2006, 2008a; Watt et al., 2010), and the
analysis of a set of > 400 age estimates from Bonaire, which shows age-distance relationships but
almost no significant overall age clusters, a stepwise evolution of the ramparts by multiple hurri-
canes is agreed upon, without, however, excluding contributions of one or very few major tsunamis
(Scheffers et al., 2014).

For the recently published coastal “berm” from Yucatán, tsunami-induced formation was suggested
based on the moderate geomorphic impact of recent storms such as Hurricane Gilbert in 1988
(Shaw and Benson, 2015). Nevertheless, the feature strongly resembles a polymodal ridge com-
plex inconsistent with tsunami deposition (Morton et al., 2008a; Etienne et al., 2011). The dominat-
ing sand matrix is merely a function of sediment availability. This evaluation is furthermore sup-
ported by a relatively low probability of tsunami occurrence in the W sector (Parsons and Geist,
2009), and by either absence of wave-induced event deposits in adjacent coastal stratigraphic
sequences (e.g., Macintyre et al., 2004; Monacci et al., 2009; Brown et al., 2014) or their reliable
attribution to storm impact (McCloskey and Keller, 2009; McCloskey and Liu, 2013a,b; Adomat and
Gischler, 2015).

2.6.2 Correlation of tsunami deposits

(Supra-)regional correlation of sedimentary records of tsunamis can provide important information


on magnitude and reach. In the Caribbean, such correlation is hampered by (i) a relatively low
density of high-quality palaeotsunami data (see Section 5.1), (ii) the predominant occurrence of
local to sub-regional tsunamis (see Section 3), and (iii) only few cases with consistent age control.
In historical times, only the largest earthquake-induced events potentially affected larger parts of
the Caribbean, as for instance the 1867 Virgin Island tsunami, for which historical accounts and
models of run-up > 0.5 m comprise the entire area between Puerto Rico and the Lesser Antilles
island arc (N and E sectors) (O'Loughlin and Lander, 2003; Zahibo et al., 2003). Historical accounts
of the 1755 Lisbon teletsunami are distributed between Cuba and Barbados (NGDC/WDS, 2016).
However, the majority of tsunamis from submarine landslides or medium to lower-magnitude earth-
quakes within the Caribbean basin probably affected just small areas. Thus, many deposits are
likely to indicate only local and/or regional tsunamis. A sector-wide or even Caribbean-wide corre-
lation of candidate tsunami deposits would, if at all, only work for major events, for which the SCDB
scenario (Figure 2-7) is a possible candidate.

One striking inter-island correlation of palaeotsunami deposits can be made in the N sector, where
both the reliable account from Puerto Rico (Moya and Mercado, 2006) and the large coral heads
transported inland from the N shore of Anegada (Figure 2-8-I) (Weil Accardo et al., 2012; Atwater
et al., 2013b, 2014) are radiocarbon-dated between 1200 and 1450 cal CE. Both records, if con-
nected, point to the Puerto Rico Trench as a major tsunami hazard (Buckley et al., 2012; Atwater

42
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

et al., 2013b). Furthermore, the N sector has several deposits convincingly associated with histor-
ical tsunamis (Moya and Mercado, 2006; Scheucher et al., 2011; Atwater et al., 2012) (Figure
2-10).

Along the Venezuelan coast, possible tsunami deposits show rather poor age constraints. Those
from lagoonal archives might be related to historical events (Weiss, 1979; Leal et al., 2014;
Oropeza et al., 2015), for which the tsunami of 1530 seems to be a likely candidate. Whether other
regional deposits of this age on Curaçao (Hornbach et al., 2008a) or Bonaire (Engel et al., 2010)
can be linked to this regional event, which was directly observed only along the E Venezuelan
coast (NGDC/WDS, 2016), remains speculative (Figure 2-10).

On Bonaire, correlation of candidate tsunami deposits was possible at some sites and led to the
tentative reconstruction of palaeotsunamis at c. 3600 BP, post-3300 BP (extreme-wave event
[EWE] II), post-2000 BP, and 1300–500 BP (Engel et al., 2013). A set of U/Th data of coral rubble
from Spelonk, all dating to shortly before 3300 BP (Scheffers et al., 2014) is probably related to
EWE II, which might have led to the irreversible destruction of fringing coral reefs along central
eastern Bonaire (Scheffers et al., 2006; Engel et al., 2013) (Figure 2-10). Whether the closure of
St. Michiel Lagoon on Curaçao (Klosowska, 2003) is related to either the 3600 BP or post-3300 BP
event identified on Bonaire remains unclear.

Some recent studies suggest significant repercussions of high-magnitude palaeotsunamis to Am-


erindian populations in the southern Caribbean by correlating them with archaeologically docu-
mented population setbacks (Scheffers et al., 2009; Hofman and Hoogland, 2015). As the recent
re-analysis of chronological datasets and the modified interpretation of formation of the ridge com-
plexes on Bonaire as outlined above challenge the inferences of strong palaeotsunamis at
4200 BP, 3100 BP, 1500 BP and 500 BP (Scheffers et al., 2009), such interrelations between tsu-
namis and prehistoric population dynamics remain a matter of debate.

2.6.3 Quantification of tsunami characteristics based on their deposits

The quantification of tsunami inundation parameters using sediment deposits is essential for
coastal hazard assessment, in particular in the case of palaeotsunamis where no other source of
information is available (Weiss and Bourgeois, 2012; Sugawara et al., 2014) (Figure 2-4). Inunda-
tion distance can be inferred from the landward limit of a deposit, even though the relation between
both parameters may vary significantly. Post-tsunami surveys revealed relationships between 50
and 60 % (Sendai Plain, Tohoku-oki Tsunami 2011; Abe et al., 2012) and 90 % (Kuril Islands, Kuril
Island Tsunami 2006; MacInnes et al., 2009) mostly depending on sediment availability and com-
position, onshore topography and local preservation potential.

Inverse numerical models use characteristics of tsunami deposits in order to reconstruct parame-
ters of tsunami inundation, such as flow velocity, wave height and inundation depth. A variety of
models have been developed for both sand and boulder deposits (Sugawara et al., 2014; Jaffe et
al., 2016), whereas their application to Caribbean datasets is clearly underdeveloped.

By applying a simple exponential relationship between flow velocity and the average b-axis of the
five largest clasts adapted from flash flood hydraulics (Costa, 1983), an approach previously used
for tsunami boulders from New South Wales by Young et al. (1996), Hearty (1997) inferred mini-
mum flow speeds of 20 m/s for the giant Eleuthera boulders with b-axes of 5.5–11.5 m. Since this
approach only considers intermediate axis length and does not take into account other factors

43
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

influencing boulder transport such as the remaining axes or boulder shape, density, bottom rough-
ness, inclination, pre-transport setting of the boulder (subaerial/subaqueous or joint-bounded),
boulder source and the MIS 5e geomorphic setting, particle-particle collision or suspension load,
the inferred values are questionable at best.

On Bonaire, initiation-of-motion criteria based on Nott (1997, 2003) were applied by several au-
thors. By using uniform equations of Nott (1997), Scheffers (2002b) infers minimum tsunami wave
heights of up to 46 and 33 m required to move the largest boulders under the assumption of boulder
overturning and rotation perpendicular to the a-axis. Uncertainties still remain, even though three
boulder axes, rock and fluid density, and coefficients of drag and lift are now considered. Spiske et
al. (2008) studied the same boulder fields on Bonaire using the dataset of Scheffers (2002b) and
enhanced initiation-of-motion criteria of Nott (2003). They considered a lower rock density of
1.8 g cm−3 and a submerged pre-transport setting of the boulders based on attached marine or-
ganisms and a relatively low degree of karstification. Reconstructed minimum tsunami heights are
< 3 m which is one reason for these authors to reject the tsunami hypothesis. Pignatelli et al. (2010)
re-evaluated the same field dataset, applied modified equations of Nott (2003) as presented in
Pignatelli et al. (2009), but considered that boulders were detached from the cliff edge and trans-
ported inland by a tsunami (joint-bounded pre-transport setting, JBB). Depending on different pre-
viously published reefrock densities ranging between 1.8 and 2.7 g/cm3 (Scheffers, 2002b; Spiske
et al., 2008), minimum tsunami heights ranging from 6.3–13.4 m were inferred (Pignatelli et al.,
2010). After improvements of the initiation-of-motion criteria were published by Nandasena et al.
(2011a), Engel and May (2012) re-measured the largest boulders' individual densities and volumes
using differential GPS and applied further modifications to the equations in order to account for
realistic boulder volume. The JBB scenario was chosen based on observations of smaller boulders
quarried from the cliff edge and transported inland during a recent storm. Minimum tsunami wave
heights required to shift the largest boulder are 6.7 and 8.9 m, respectively (Engel and May, 2012).
Based on a numerical approach of Pignatelli et al. (2009), the maximum tsunami inundation dis-
tance on top of the elevated carbonate platform of NE Bonaire (Figure 2-8-IV) as indicated by the
largest boulders was calculated to be in the range of 243 m (Engel and May, 2012) and 164–565 m
(Pignatelli et al., 2010), respectively, depending on rock density. No direct relation could be made
between quantification of marine inundation and risk, as the boulder fields are located far from
residential areas and other valuable features such as important infrastructure.

In view of the broad range of reconstructed minimum wave heights on Bonaire resulting from simple
inverse modelling approaches, it becomes clear that diverging approaches of boulder mapping and
measurements of density in heterogeneous reefrock boulders strongly influence model outputs
(e.g., Jaffe et al., 2016). Further uncertainties derive from an ambiguous pre-transport setting (e.g.,
Nott, 2003; Jaffe et al., 2016), oversimplification, such as a stable Froude number, which
significantly varies for tsunamis approaching the coast (Sugawara et al., 2014), the dynamic rela-
tionship between flow velocity and wave height, and non-consideration of the duration of forces,
as well as bottom roughness, slope angle (Weiss and Diplas, 2015) or the boulder's complex shape
(Zainali and Weiss, 2015). However, when using carefully evaluated field data, initiation-of-motion
criteria can provide a useful estimate on wave heights and flow velocities, even though it has to be
considered, in terms of coastal hazard assessment, that they often underestimate real conditions,
as observed in case studies after recent tsunamis (Paris et al., 2010).

44
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

A forward model of sediment transport combines the tsunami source and a hydrodynamic and
sediment-transport model with the local bathymetry and topography on one or several spatial
scales. The model is validated using sedimentary field data (Sugawara et al., 2014). One applica-
tion from the Caribbean exists from Anegada (BVI). Buckley et al. (2012) identified a M 8.0 earth-
quake at the outer rise along the Puerto Rico Trench to have generated a tsunami reaching on-
shore flow velocities of up to 10 m/s, a height of > 5 m at the shoreline and an inundation distance
of > 1000 m by applying a sediment-transport model, which considers fluid drag, inertia, buoyancy,
and lift forces on boulders comprising both sliding and overturning transport modes. In the light of
these promising results, the application of sediment-transport models and their validation with pal-
aeotsunami deposits should be supported and encouraged in the Caribbean in order to serve
coastal hazard assessment.

2.6.4 Implications for tsunami hazard assessment

Even though it is widely accepted that coastal hazard assessment may benefit from palaeotsunami
investigations, a proper transfer with regard to the Caribbean has so far only been made in one
case study from north Jamaica, where, however, the focus is on hurricanes (Rowe et al., 2009;
Miller et al., 2014). Based on the assumption, that boulder ridges from Jamaica accumulate over
time at the landward limit of potential damage induced by major extreme wave events, Miller et al.
(2014) propose to use them as minimum setback distances for coastal development. These rec-
ommendations of up to 130 m-wide no-build zones starkly contrast with current regulations where
setback distances lie between 6 and 30 m, depending on surface elevation and gradient (Miller et
al., 2014).

Considering the spatial distribution of extreme-wave deposits for planning and land use is reason-
able, in particular when characteristics of coastal flooding are reliably quantified through modelling
of the sediment transport. However, at several places in the Caribbean these valuable indicators
of hazard-prone areas are being excessively mined for construction purposes (Scheffers, 2002a,
b, 2005) further leading to an increase of inundation distances (Morton et al., 2006). The dramatic
consequences of carbonate sand and boulder mining in the Caribbean in combination with elevated
wave energy, including coastal erosion and also the inevitable destruction of the archaeological
record, have recently been outlined by Fitzpatrick (2012).

To sum up, a great potential for assessing tsunami hazard in the Caribbean has yet to be unlocked
through interdisciplinary research comprising more detailed studies of extreme-wave deposits, es-
timation of long-term frequency-magnitude patterns based on the deposits' chronology, site-
specific quantification of flooding characteristics in vulnerable areas, and their implementation in –
already existing (von Hillebrandt-Andrade, 2013) – coastal hazard management and planning.

2.7 Conclusions

Many Holocene coastal stratigraphic records have been explored in the Caribbean. Only few of
them report disturbance of background sedimentation and/or allochthonous deposits and relate
these finds to extreme-wave events. A focus on such events and attempts to identify tsunami de-
posits as a contribution to local and regional coastal hazard assessment is rare, in particular when
compared to other regions such as south Japan (Garrett et al., 2016) or the Indian Ocean (e.g.,
Jankaew et al., 2008; Alam et al., 2012; Brill et al., 2012), even though the Caribbean has a

45
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

significant tsunami potential based on the historical record. Several fine-grained deposits con-
sistent with tsunami facies models (Table A-1) were reviewed. More data exist from subaerial
coarse-clast deposits, even though the determination of the exact depositional process (tsunami,
storm) and the timing often remains ambiguous. Therefore, high detail and common standards in
stratigraphic documentation and analysis (e.g., Scheffers and Kelletat, 2003; Mamo et al., 2009)
combined with the application of new innovative approaches (e.g., μCT; May et al., 2016) are
needed in future studies from the Caribbean.

The widespread polymodal ridge complexes, in many places interpreted as tsunamigenic in previ-
ous publications (Table A-4), seem to rather comply to stepwise accumulation mostly during strong
hurricanes (Table A-2) (Morton et al., 2008a; Scheffers et al., 2014) where they mark the terminal
distance of coarse clast transport and long-term limit of transport capacity by storm waves over-
topping the elevated platforms (Khan et al., 2010; Miller et al., 2014). Tsunamis – depending on
availability of source material and coastal topography – rather form laterally unsorted, widely dis-
tributed fields of cobbles and boulders (Watt et al., 2012a). Large singular boulders (up to > 100 t)
found on many elevated coastal platforms are consistent with both, storm and tsunami deposition,
as demonstrated by both the recent Tohoku-oki Tsunami 2011 (Nandasena et al., 2013) and Ty-
phoon Haiyan in the Philippines (May et al., 2015b). However, larger transport distances of tsunami
boulders might still represent a useful criterion to distinguish between both processes (Goto et al.,
2010a; Etienne et al., 2011).

Source areas of fine-grained candidate tsunami deposits from the Caribbean seem to focus on the
shallow subtidal (including the upper coral reef zone), but reveal minor incorporation of sediment
from deeper waters (e.g., Engel et al., 2012) and terrestrial material in the upper part deposited by
the backwash (e.g., Engel et al., 2010). While ridges and ridge complexes are mainly built of reefal
debris, large boulders on top of elevated coastal platforms mainly derive from the cliff edge (e.g.,
Robinson et al., 2006; Khan et al., 2010; Engel and May, 2012). In some places, they preferentially
cluster around terrace indentations (e.g., Jones and Hunter, 1992; Robinson et al., 2006; Watt et
al., 2010).

The spatial correlation of prehistoric tsunami deposits of the same event on a regional scale is so
far only marginally possible due to the low density of high-quality palaeotsunami data, only very
few cases with consistent age constraints, and the predominant occurrence of local to sub-regional
tsunamis on historical scales. If representing the same event, the potential medieval tsunami de-
posits identified on Anegada (Atwater et al., 2013b, 2014) and Puerto Rico (Moya and Mercado,
2006) are an exception as they might indicate a significant tsunami hazard emanating from earth-
quakes at the Puerto Rico Trench (Atwater et al., 2013b). The presented potential numerical sce-
narios of tsunamis generated at the SCBD (Figure 2-7) as well as several giant mass wasting
deposits at the foot of continental slopes (e.g., Deplus et al., 2001; Le Friant et al., 2009; Leslie
and Mann, 2016) reveal the hazard of pan-Caribbean impacts and the possibility of supra-regional
correlation of palaeotsunami deposits and geomorphic traces.

As extreme-wave deposits are unequivocally understudied in the Caribbean, there is an enormous


potential for coastal hazard assessment to be developed. Quantitative information on maximum
flow depth, inundation distance and flow velocities drawn from such deposits by applying improved
numerical models of sediment transport is still very limited. Thus, further palaeotsunami studies
using high-resolution methods of bedform and stratigraphical documentation and generating con-

46
Tsunami deposits of the Caribbean – Towards an improved coastal hazard assessment

sistent chronological models with independent age control, combined with refined inverse and for-
ward models of sediment transport and deposition, are required to reconstruct long-term patterns
of magnitude and frequency of palaeotsunamis in the Caribbean – a prerequisite for reliably map-
ping hazard-prone areas. To date, proposed palaeotsunami deposits from the Caribbean probably
represent only a fraction of actually occurred prehistoric tsunamis and, therefore, reflect major tsu-
nami inundations inadequately.

Acknowledgements

This review was made possible through funds and time budgets associated with a Max Delbrück
Prize for junior researchers, a measure of the future concept of the University of Cologne in the
framework of the Excellence Initiative (DFG ZUK 81/1). It is furthermore based on experience made
through previous research on Caribbean tsunamis funded by the Deutsche Forschungsgemein-
schaft (BR 877/26-1). The manuscript benefitted from thoughtful pre-submission comments by D.
Kelletat and associated discussions. We thank A. Scheffers, D. Kelletat, B. Atwater, B. Richmond,
and G. Gelfenbaum for making photographs available. B. Atwater kindly provided insights into un-
published data. Language editing by K. Jacobson is acknowledged. The manuscript benefitted
from helpful comments by two anonymous reviewers and the handling editor A. Strasser.

47
Enhanced field observation based physical and numerical modelling of tsunami induced boulder transport

3 Enhanced field observation based physical and numerical modelling of


tsunami induced boulder transport – Phase 1: Physical experiments

Jan Oetjen, Max Engel, Helmut Brückner, Shiva P. Pudasaini, Holger Schüttrumpf
This is an accepted manuscript of an article published in Coastal Engineering Proceedings in June
2017. Online available: https://doi.org/10.9753/icce.v35.management.4
Keywords: Tsunami; Boulder transport; Physical experiments; Numerical modelling

Abstract
Coasts around the world are affected by high-energy wave events like storm surges or tsunamis.
By focusing on tsunami impacts, we investigate tsunami-induced transport of boulders by an inter-
disciplinary combination of field observations, laboratory experiments and advanced numerical
modelling. In phase 1 of the project we conduct physical laboratory experiments based on real-
world data. Following the experimental phase we will develop an enhanced numerical boulder
transport model (BTM) based on an existing two-phase model.

3.1 Introduction

Coasts are crucial areas for living, economy, transportation, and all sectors of industry. Many of
them are periodically affected by high-energy wave events. Referring to Neumann et al. (2015) and
depending on the applied scenario for population growth, in 2060 between 1,052 and 1,388 million
people will live within the low-elevation coastal zones. Compared to the 625 million inhabitants
today, this is equivalent with an increase between 68 % and 122 %. Therefore, there is an urgent
need to improve the understanding of the characteristics, maximum magnitudes, recurrence inter-
vals and socio-economic impacts of high-energy wave events (Huntington et al., 2007; Knutson et
al., 2010; Vose et al., 2014). Marine-borne geological deposits in coastal geo-archives provide
essential information on long term, prehistoric frequency-magnitude patterns of high-energy wave
events. This is a basic requirement for long-term coastal hazard management (Engel et al., 2014).
Fine-grained deposits may quickly be altered after such events and may not be preserved in the
sedimentary record (Engel and Brückner, 2011). Coarse clasts with weights of more than 80 t (so
called boulders, a-axis > 0.25 m or blocks, a-axis > 4.1 m), which were dislocated by extreme
waves, represent more reliable records in terms of preservation potential and should Therefore be
used for the investigation of past events. For establishing an accurate chronology and typing of
high-energy wave events and classifying them in terms of risk assessments, it is necessary to
detect whether the boulders were moved by storm or tsunami.

However, differentiating between storm- and tsunami-dislocated clasts is a challenge, and there-
fore the development of reliable approaches and tools to solve this issue, preferably based on field
observations, is essential to accurately assess extreme coastal hazards. Such tools can be numer-
ically coupled tsunami and boulder transport models (BTM) supporting the in situ work of geolo-
gists, geographers, engineers and other researchers. To analyze boulder deposits numerically,
two basic approaches are applied nowadays: (i) inverse modelling uses specific boulder charac-
teristics as input for analytic equations. These equations (e.g. Nott, 1997; Pignatelli et al., 2009;
Nandasena et al., 2011a; Benner et al., 2010; Engel and May, 2012) describe the necessary
threshold wave height or wave velocity to displace a boulder of known dimensions and density by
the use of momentum balances. Those equations only provide rough estimates for minimum wave

48
Enhanced field observation based physical and numerical modelling of tsunami induced boulder transport

energy since they do not consider the complex non-linear interactions between turbulent flow fields
and boulders. (ii) On the other hand, forward models (e.g. Imamura et al., 2008; Zainali and Weiss,
2015) operate numerically and calculate tsunami-induced boulder movements from given wave
characteristics in more detail. Their validation and calibration on the basis of physical measure-
ments or observations is essential. This can be achieved by field observations (e.g. Nandasena et
al., 2011b) and physical experiments (e.g. Imamura et al., 2008). However, current forward models
still suffer from insufficient quantitative understanding of essential processes during boulder
transport by waves. So far, only very few major tsunami events allows the field-observation based
validation of numerical models (e.g. Indian Ocean Tsunami 2004; Goto et al., 2010b; Nandasena
et al., 2011b). Thus, well controlled laboratory experiments are essential for systematic calibration
processes of numerical hydrodynamic models.

3.2 Boulder transport modelling

Wave-induced transport of any three-dimensional solid body is influenced by many dynamic vari-
ables and parameters, which can be classified in three groups:

1. Hydrodynamic forces: Transport-inducing and wave-induced forces accounting for the impact
forces on the body,

2. Body characteristics: Body-inherent characteristics like shape, density or dimensions, etc.

3. Environment: Topography, surface roughness, bathymetry and boulder pre-transport setting,


etc. None of these variables and parameters is simple to determine in the field or easy to
reproduce in the laboratory or numerically, and many uncertainties remain.

3.2.1 Hydrodynamic forces

Existing inverse models commonly consider drag, inertia, lift, friction and (reduced) gravity as im-
pact or restraining forces (Sugawara et al., 2014 and references therein). Figure 3-1 displays these
basic hydrodynamic forces acting on an idealized boulder during wave impact. The lift and reduced
gravity forces are also influenced by the initial boulder location, which will be discussed later in
more detail. By considering particular boulders in the field, in most cases no accurate observations
or even measurements of the transport-inducing event are available. In most cases, the event itself
is not known (e.g., Hisamatsu et al., 2014; Prizomwala et al., 2015). Therefore, none of the hydro-
dynamic variables such as wave length, wave height or amplitude is known; they need to be esti-
mated from today’s obtainable indicators, like boulder characteristics. However, with the existing
numerical models (forward and inverse) it is not possible to determine the necessary wave length,
velocity and height with required and high accuracy which is necessary for distinguishing between
storm and tsunami (e.g. Zainali and Weiss, 2015). Shortcomings include the treatment of the boul-
der as an isolated, single and idealized body or the utilization of rough estimated regionally uniform
coefficients (e.g. for drag coefficient) in available numerical models (Sugawara et al., 2014;
Hisamatsu et al., 2014). Therefore, since wave length and velocity are crucial parameters for the
distinction between storm and tsunami waves, advancements in numerical modelling are essential.

49
Enhanced field observation based physical and numerical modelling of tsunami induced boulder transport

Figure 3-1: Basic impact forces related to hydrodynamics. FM: Inertia; FD: Drag; FA: Lift; FG: Gravity; FR: Friction
(Pignatelli et al., 2009).

3.2.2 Body characteristics

The body characteristics are unique for every single boulder deposit. All known numerical and
laboratory BTM use idealized boulder shapes and homogeneous densities for simulations (e.g.
Imamura et al., 2008; Nandasena and Tanaka, 2013; Liu et al., 2014). However, a simplified shape,
such as a rectangular prism or cube, might be more or less exposed to friction and drag forces
than the natural irregular counterpart. Combined with uncertainties associated with assumptions of
bottom friction, idealized boulder shapes lead to largely inaccurate results for the predominant
transport mode, e.g. rolling, sliding, saltation (Etienne et al., 2011) or a combination of these (Figure
3-2a). Furthermore, the impact of hydrodynamic forces on a complex shape will differ substantially
from the impact on rectangular shapes. Not to mention micro-turbulences on irregular boulder sur-
faces, the impact will noticeably decrease the more the boulder front converges towards a shape
with little drag. Zainali and Weiss (2015) recently pointed out the importance of the boulder aspect
ratio of rectangular bodies for transport in a fluid. Investigating the behavior of more complex bodies
and the comparison to idealized bodies will Therefore lead to new insights of how the shape influ-
ences boulder transport mechanisms.

3.2.3 Environment

Bathymetry and topography strongly influence the flow and wave characteristics. They determine
the distance between the wave breaking point and the initial boulder location. After the wave
breaks, the flow front becomes supercritical with increasing Froude number because the flow be-
comes shallower and faster (e.g. Zainali and Weiss, 2015). This problem can be adequately treated
numerically with high-resolution adaptive computational grids (in time and space). Then, the results
can be validated with accurate field measurements. Furthermore, the onshore transport distance
of a boulder will be significantly influenced by the pretransport setting: whether the boulder is ini-
tially subaerial, submerged, partially submerged and/or joint-bounded (Figure 3-2b) (Zainali and
Weiss, 2015). During the initial wave impact and depending on the submergence of a boulder,
particular forces can be neglected or additional forces need to be considered. A subaerial boulder
will initially be affected by full gravity, while reduced gravity needs to be considered if flow depth
exceeds boulder height. In addition, a boulder connected with the cliff first needs to be detached
and temporarily enhanced restraining forces have to be applied (e.g. Noormets et al., 2004). The

50
Enhanced field observation based physical and numerical modelling of tsunami induced boulder transport

transport mode of the boulder is highly affected by macro- and micro-topography (bottom rough-
ness). A complex shaped boulder with sharp edges, e.g., in karstified reef rock, on a very rough
ground tends to rolling or even saltation, whereas the same boulder on a plain ground will prefer-
ably move by sliding. An accurate numerical boulder transport model needs to account for all pos-
sible initial boulder locations, pre-transport settings, bottom roughness and dominant forces.

Figure 3-2: Boulder transportation modes (a) and initial positions relative to water level (b) which are commonly consid-
ered in numerical models.

3.3 Physical boulder transport modelling

A fundamental requirement of a numerical model is its ability to reproduce nature as accurately as


possible. For verification purposes, several highly developed laboratory experiments on tsunami
generation, propagation and inundation have been established in which researchers follow diver-
gent approaches regarding tsunami characteristics and wave generation (e.g., Penchev, 2008).
However, realistic laboratory experiments focusing on boulder transport mechanism due to tsunami
waves are still rare (Martinez et al., 2011; Nandasena and Tanaka, 2013; Liu et al., 2014; Bressan
et al., 2015); even some of the most basic processes, e.g. influence of boulder shape and sedi-
mentary load on hydrodynamics, are not or not entirely understood. Imamura et al. (2008) con-
ducted laboratory experiments in an open channel ending on a slope, and compared the results
with those of their numerical BTM. This model setup - a plain bed ending on a slope - is to our
knowledge the only today conducted setup type for experimental boulder transport modelling (e.g.
Nandasena and Tanaka, 2013; Bressan et al., 2015). Except for Sangster et al. (unpublished) and
Martinez et al. (2011), who used near to spherical bodies and marbles, respectively, all experi-
ments involve only rectangular boulder replicas. These results in significantly different transport
behavior as compared to realistic shapes due to, for instance, variations of the drag force. Model-
ling the broken tsunami wave as a bore is widely accepted (Sugawara et al., 2008). To our
knowledge, the only laboratory boulder transport experiments in which the velocity and orientation
of boulders are measured directly out of the (idealized) boulder are conducted by Agnieszka
Strusińska-Correia at the Leichtweiß-Institute for Hydraulic Engineering and Water Resources at
the Technical University of Braunschweig. However, so far no results have been published in this
context. Summarized, no laboratory experiments are known which consider for complex boulder
bodies, non-uniformly sloped bathymetries and sedimentary load in a tsunami bore.

51
Enhanced field observation based physical and numerical modelling of tsunami induced boulder transport

3.4 Case study “Island of Bonaire”

The foundation of the presented research project lies in a case study at the Island of Bonaire
(southern Caribbean, Leeward Antilles, Figure 3-3). Large limestone boulders were transported
from the cliff edge of the island landward during high-energy wave events in the past. However,
the transport causing event (paleo-tsunamis or severe hurricanes) is still discussed highly contro-
versial (e.g. Scheffers 2005; Spiske et al., 2008; Watt et al., 2010; Engel and May, 2012).

We applied an iterative approach simulating historical and other potential tsunami events in the
Caribbean Sea in order to identify the most suitable event(s) for creating sediment patterns on
Bonaire as well as the source of possible future threats (Oetjen et al., 2015).

Delft 3D (Deltares, 2014a) and DelftDashboard (Deltares, 2014b) of Deltares systems (NL) was
utilized to compute different tsunami scenarios for Bonaire (Table 3-1, Figure 3-3). The initial, tsu-
nami triggering, earthquake was simulated using DelftDashboard (DD). As result DD generated an
initial input file used for Delft 3D in which the tsunami propagation across the Caribbean Sea and
the impact at Bonaire’s coastline are computed in depth-averaged 2D.

Figure 3-3: Location of Bonaire and the earthquake scenarios (left). Location of the study areas on Bonaire and the
largest boulder BOL2 at Boka Olivia (right).

Table 3-1: Parameters of the simulated tsunami scenarios.

Muertos Southern Carib.


Macuto Virgin Islands Trough Deformed Belt
Witnessed Witnessed Hyphotetic Hyphotetic

Year 1900 1867 [-] [-]


Fault Line Parameter

Magnitude [Mw] 7.79 7 8.79 8.34

Length [km] 144 49 554 302

Width [km] 100 39 130 100

Reference Lander et al. Barkan and ten Bruña et al. IOC (2012)
(2002) Brink (2010) (2014)

52
Enhanced field observation based physical and numerical modelling of tsunami induced boulder transport

We applied a multi-scale modelling approach of three scales where we utilized two coarse models
based on the GEBCO08 bathymetry data set (GEBCO, 2014) and one of higher resolution around
Bonaire applying data provided by Gonzalez-Lopez and Westerink (University of Notre Dame) with
a spatial resolution of 500 – 600 m (Figure 3-4). The topography was generated with SRTM4.1
(Jarvis et al., 2008) for all models.

Figure 3-4: Model bathymetry near to Bonaires shoreline.

None of the simulated scenarios indicated a significant impact at Bonaire’s shoreline. The highest
waves were obtained for the Muertos Trough Scenario at Boka Olivia with a wave height of 3.4 m
(Figure 3-5). According to analytical solutions for boulder transport and results of Engel and May
(2012), such a tsunami does not carry enough hydrodynamic energy to move a boulder like Bon-
aire’s largest of app. 80 m³. Finally the results of our numerical model could not clarify whether the
boulders were transported by storm waves or tsunamis and hence the corresponding discussion
is still alive.

Recognizing these results and considering the scientific state of the art, we founded a research
group aiming at improving the understanding of tsunami induced boulder transport and associated
nearshore tsunami hydrodynamics. Therefore we conduct well controlled downscaled physical ex-
periments utilizing field data from the Island of Bonaire. Beside the new insights gained from the
physical experiments, these data are furthermore used for the validation of a new developed nu-
merical BTM.

53
Enhanced field observation based physical and numerical modelling of tsunami induced boulder transport

Figure 3-5: Computed wave heights at Boka Olivia.

3.5 Physical experiments

Within the physical experiments we want to achieve a better knowledge in terms of the hydrody-
namics of nearshore tsunami behavior. Therefore we investigate three key questions:

1. Influence of realistic / non-uniform boulder shapes,

2. Influence of (non-uniform) bathymetry/onshore topographies,

3. Influence of boulder interactions during transport, on boulder transport path, mode, dis-
tance and deposition pattern.

The experiments are conducted in the tilting flume at Institute of Hydraulic Engineering and Water
Resources Management (IWW, RWTH Aachen University Germany). The 33.5 m x 1 m x 1 m large
flume consists of two centrifugal pumps of 30 kW with 400 l/s conveying capacity. In combination
with the following valves and an automatical control program reproducible experiments are guar-
anteed (Figure 3-6). Bonaire’s largest boulder measures 8.7 m x 4.8 m x 3.8 m. Since the experi-
ments scale is 1:50 the model counterpart of the boulder is approximately 18 cm x 10 cm x 8 cm.
Figure 3-7 shows the general setup of the physical experiments.

54
Enhanced field observation based physical and numerical modelling of tsunami induced boulder transport

Figure 3-6: Left: Comparison of two test runs for bore generation in the IWW test flume. Right: Photo of a generated
tsunami bore (Peltzer, 2015).

Figure 3-7: Principal test setup of the flume (not to scale).

The measurement setup regarding the wave encompasses five ultrasonic sensors, one impeller
flow meter as well as an acoustic doppler velocimeter (ADV). For the measurement of the boulder
behavior we use inertial measurement units (IMU) and image respective video processing with
MATLAB®.

The IMU is used for the measurement of acceleration and inclination of the boulder during the wave
impact. In combination with a drawn 10 cm x 10 cm grid on the shore replica the video processing
shows distance and path of the boulder during the impact.

For the investigation of the influence of shape, boulders at the Island of Bonaire are recorded using
photogrammetry. Therefore, the digital photos (recorded with a Sony Cyber-shot DSC-HX200V)
are computational processed (Figure 3-8). The obtained digital mesh can be processed further
which is necessary for not reachable boulder ranges as the on ground lying part, for example. The
final dataset is printed in three dimensions with a German RepRap X400 PRO ® 3D printer (Figure
3-9). The obtained model from PETG plastic is further used as a template for a negative mask.
This mask is created from silicone rubber (ELASTOSIL ® M, Wacker Chemie AG). This two-part
silicone rubber is highly suitable as the final template for the creation of the concrete boulder model.
The initial fluent rubber fills the gaps of the highly irregular boulder model.

55
Enhanced field observation based physical and numerical modelling of tsunami induced boulder transport

Figure 3-8: a) Perspective view and mesh model of a boulder deposited on Bonaire. b) Comparison between the original
photograph and the numerical model). c) Contrasting shapes of idealized and complex-shaped boulders (a, b, c = main
axes of boulders).

Figure 3-9: Printing process of the bottom part of a boulder replica in 1:50.

The IMU is finally installed inside of the boulder. We apply the Sensor Fish sensor by the Pacific
Northwest National Laboratory of the U.S. Department of Energy (Figure 3-10, Deng et al., 2014).
The Sensor Fish Gen 2 is able to store data for acceleration, rotation velocity and orientation with
a sampling rate of 2048 Hz on a flash memory inside of the device. Therefore, the boulder model
needs the possibility to remove the Sensor fish and download the data. This is realized due to a
closeable thread until the boulders centroid where the sensor is inducted.

Once the non-idealized boulder models are created it is possible to compare the behavior of ide-
alized (cubes, cuboids) to non-idealized boulder shapes in terms of transport path, distance and
mode (Figure 3-8).

Beside the investigation of several boulder shapes we analyze the influence of the shore. Hence
we constructed four divergent shore replicas (Figure 3-11). We use idealized replicas of Bonaire’s
shoreline resembling a low plunging cliff with an elevated platform, emulating the setting the model
coastal boulder field on Bonaire (Engel and May, 2012). The shores are created to change the
setup comparable fast. All shores ends up at the same cuboid. The shores themselves consist of
two parts that can be easily installed by one or two persons.

The experimental plan encompasses several settings regarding the initial boulder position on the
shore and the submergence (full, partial, subaerial). Furthermore, we apply different wave heights.

56
Enhanced field observation based physical and numerical modelling of tsunami induced boulder transport

Finally, utilizing the IMU, video processing and common measurement devices allows us to derive
exactly the necessary tsunami energy to shift a boulder in motion.

Figure 3-10: CAD model of the Sensor Fish Gen 2 device (Deng et al., 2014).

Figure 3-11: Shore replicas used for the physical experiments.

3.6 Phase 2: Numerical model

Parallel to the physical experiments a numerical boulder transport model (BTM) is in development.
The BTM is based on the real two-phase model of Pudasaini (2012). The model of Pudasaini solves
the equations for two-phase flows (fluid and particles) explicit on a Eulerian grid. In order to include
the movement of a rigid boulder with an arbitrary shape the boulder is treated as an immersed
boundary with the direct forcing technique (e.g., Gornak, 2013). In this context the physical behav-
ior of the boulder within the flow will be solved with the pe Rigid Multi-Body Physics Engine of the
Friedrich-Alexander-University Erlangen-Nuremberg (Iglberger and Rüde, 2009).

Hence, the development of the numerical model encompasses two essential steps:

1. Determine the crucial physical forces on the boulder,

2. Well coupling of the pe (FORTRAN 95) and the two-phase model (C++).

57
Enhanced field observation based physical and numerical modelling of tsunami induced boulder transport

The model will be validated using the data from the physical experiments. Especially the data cap-
tured from the IMU is very helpful to determine the accuracy of the BTM highly correct.

3.7 Summary

Regarding to our experience due to the field study on the Island of Bonaire we recognized several
problems in the field of numerical and physical boulder transport modelling. By conducting well
controlled physical experiments utilizing real-world data we want to tackle three crucial gaps (boul-
der and shore shape, boulder interactions) in research. Within our project, we combine the exper-
tise of different research fields like geographers, engineers and (geo-) physicist. By applying real
world data for the boulder and shore models we make a step away from over-simplified model
assumptions both for physical experiments and numerical modelling.

58
Experiments on tsunami induced boulder transport – a review

4 Experiments on tsunami induced boulder transport – a review

Jan Oetjen, Max Engel, Holger Schüttrumpf


This is an accepted manuscript of an article published by Elsevier in Earth-Science Reviews in
September 2021. Online available: https://doi.org/10.1016/j.earscirev.2021.103714
Keywords: Tsunami, Boulder transport, Hydraulic experiment, Physical experiment, Parameter in-
fluence

Abstract
The interest in applying coastal boulders as a proxy for high-energy wave events, i.e., tsunamis
and major storms, increased significantly during the last two decades. Studies include on-site in-
vestigations, numerical modelling or experimental studies. However, inferences on past inundation
processes based on commonly used inverse modelling approaches are associated with significant
uncertainties and the need for validation of relevant parameters has been expressed. Here, we
review experimental studies on the transport of natural boulders (and artificial bodies) by tsunamis,
which provide crucial insights into the physical parameters controlling the transport process. We
summarize and categorize the differences and scopes of existing experimental studies and point
out the importance of the boulders’ wave-facing area and, thus, the drag force, which is highlighted
in most studies. The boulder-transport mode (e.g., rolling or sliding), however, varies strongly
throughout the publications. A considerable number of influencing parameters is not investigated
systematically (e.g., the influence of boulder-boulder interactions) so far, and there is an agreement
that existing hydrodynamic equations for mobilization thresholds tend to overestimate the neces-
sary wave height and velocity due to the existing uncertainties. Therefore, and based on the current
knowledge, working with parameter ranges (e.g., mobilization thresholds, transport-distance
spans) is recommended instead of using exact values. However, we demonstrate that comparisons
of the particular findings of the reviewed studies is not straightforward due to very individual exper-
imental setups. Based on this, we propose a standardized research environment, which might sim-
plify the knowledge transfer between studies and can possibly improve numerical models for tsu-
nami-induced boulder transport.

4.1 Introduction

Scientific interest in the transport of coastal boulders by high-energy waves such as tsunamis and
storm waves has increased in recent years, as they let us infer hydraulic characteristics of such
events in the past. This information is crucial for assessing coastal hazards and to inform commu-
nities and stakeholders in coastal hazard management (Goto et al., 2010a; Etienne et al., 2011;
Nandasena et al., 2011a; Engel et al., 2016; Cox et al., 2020). Tsunamis and storm surges with
superimposed gravity waves represent a coastal hazard in all ocean basins, in tropical and extra-
tropical latitudes (Løvholt et al., 2012; May et al., 2013). These waves may entirely flood low-lying
coastal areas and even overtop high cliffs (e.g., Cox et al., 2020). Coastal boulders are sourced
from these cliff edges, from rocky headlands or the outer part of fringing reefs, and shifted inland
during the same or a subsequent event as a function of the transport capacity of the waves, though
rarely against gravity. These transported boulders are used to reconstruct hydrodynamic charac-
teristics of low-frequency high-magnitude wave events (Bourrouilh-Le Jan and Talandier, 1985;
Etienne and Paris, 2010; Etienne et al., 2011; Switzer and Burston, 2010; Abad et al., 2020) and
depict – in combination with the historical and instrumental record – an essential element of coastal

59
Experiments on tsunami induced boulder transport – a review

hazard assessment (Engel et al., 2016). In recent years, in particular after Supertyphoon Haiyan
in 2013, high-amplitude infragravity waves have also been considered as transport mechanism of
large coastal boulders, with patterns of boulder dislocation very similar to those identified for tsu-
namis. These observations further increase the challenges of hydrodynamic inferences from the
coarse-clast record (May et al., 2015; Kennedy et al., 2017; Soria et al., 2018).

Even though field observations have revealed certain recurring patterns in coastal boulder deposits
for the different hydrodynamic transport mechanisms, it is still often challenging to reconstruct pro-
cesses from old deposits in coastal areas where different types of high-energy waves can be ex-
pected. While boulder fields created by tsunamis often appear unsorted and show rather abrupt
landward boundaries with a generally wider inland distribution, storm waves, which usually occur
at a higher frequency, induce shorter, step-wise transport and result in exponential landward fining
or even build distinct polymodal ridges (Goto et al., 2010a; Switzer and Burston, 2010; Etienne et
al., 2011; Lau et al., 2014; Engel et al., 2016; Boesl et al., 2020; Cox et al., 2020). First attempts
of inversely modelling hydrodynamic characteristics and specific transport processes from the size,
density and shape of boulders by Nott (1997, 2003) and others (e.g. Benner et al., 2010; Lorang,
2011) have been widely applied and modified (e.g. Nandasena et al., 2011a; Engel and May, 2012;
Lau et al., 2015), even though it is now commonly understood that these initiation-of-motion criteria
(or incipient-motion formulas) fail to reproduce the natural system, that comes with dynamic Froude
numbers, complex coastal bathymetries and topographies, dynamic bottom friction, wave-wave
and clast-clast interaction during transport, varying concentrations of suspension load, irregular
shapes and density distribution of boulders, among others (Cox et al., 2020; Nandasena, 2020).
Other unknown factors often include the boulder origin, transport pathway and transport mode.
Some limited quantitative information that may improve estimates on these unknown parameters
in numerical models may be derived from post-tsunami or post-storm field surveys (e.g. Nandasena
et al., 2013; May et al., 2015; Cox et al., 2018a; Soria et al., 2018), in combination with satellite,
aerial images, and local high-resolution orthophotos and digital elevation models (DEMs) using
unmanned aerial vehicles (UAV) that help to track pre- and post-event clast locations (e.g., Boesl
et al., 2020; Hoffmeister, 2020). Nevertheless, the most straightforward way to quantify such un-
known variables are physical experiments of boulder transport by waves (e.g., Imamura et al.,
2008; Oetjen et al., 2020a).

Experimental investigations by physical models help to extend and improve the available database
by investigating crucial parameters for the transport of boulders by high-energy waves in a well-
controllable environment. However, until today only few such experimental studies are published.
Furthermore, most of the studies are based on divergent assumptions and simplifications, and their
comparison is not straightforward. This paper provides an overview on published experimental
studies regarding clast transport by high-energy wave events with a particular focus on tsunami-
related studies. It compares their output with on-site field observations and sums up their benefit
for tsunami research in terms of better understanding the transport process and the parametriza-
tion for numerical modelling. Throughout the existing publications the terminology differs slightly.
Within the present study, terms will be used as in Table 4-1 and might differ from those of the
original references.

60
Experiments on tsunami induced boulder transport – a review

Table 4-1: Terms used across the presented review. Some terms might differ from those in the original publications.

Term Description
Current reflected from a wall or inclined topography located behind
Backwash the boulder and directed towards the sea

Boulder fully covered by water


Submerged

Boulder not fully covered by water (e.g., 50 %)


Partially submerged

Boulder located above water level


Subaerial

Boulder alignment to wave impact (0°: long axis parallel to impact;


Initial alignment 90°: Short axis parallel to impact)

Boulder trajectory between wave impact and final deposition


Transport path

Mode of boulder transport: sliding, rolling, saltating


Transport mode

Maximum extent to the initial boulder location during wave impact


Transport distance (max.)

Total transport distance. Distance between initial boulder location


Transport distance (tot.) and location where the boulder finally comes to rest

Point of time when the boulder is set in motion


Boulder mobilization / incipient motion
Describing boulder flatness by the ratio of the boulder’s long (a),
Flatness Index (FI) intermediate (b) and short (c) axes: FI = (a + b) / 2c (Nandasena
and Tanaka, 2013)

Describing the ratio between boulder height (hB) and a wave or


Boulder-wave ratio rBW = hB/hw [-] bore height (hw) (Figure 4-3). The considered wave height depends
on the values given in the particular publications.

4.1.1 Transport of solid objects by high-energy wave events

The transport of solid objects by waves depends on wave energy (studies on incipient motion, e.g.,
Bressan et al., 2018), boulder properties, pre-transport setting, bathymetry, and topography. An
object is moved by waves if the sum of mobilizing forces exceeds the sum of resisting forces. On
the mobilizing side the main contributing forces are drag, inertia and buoyancy or lift force. The
main resisting forces are friction, gravity (or weight) force and, if present, hydrostatic forces on the
rear side of the object (Figure 4-1). Further forces, such as adhesion between ground and boulder,
may also influence entrainment but are usually considered negligible.

The amount and portion of particular forces initiating boulder transport depend on multiple variable
boundary conditions. In the case of buoyant objects, large lift and buoyancy forces and low bottom
friction enable much quicker entrainment compared to non-buoyant, higher-density material. Fur-
thermore, any sort of inclined shores, in particular for sliding transport, increases resistive gravity
forces, which may be exacerbated by bottom friction. In partially submerged conditions, hydrostatic
forces are applied on the rear of bodies and hamper entrainment. The active portion of drag force
(form drag and skin friction drag) depends on the shape and surface of an object. Body shapes of
low form drag and skin friction coefficient require higher or longer wave impact for entrainment.
Figure 4-1 and Figure 4-2 show the main acting forces for some idealized boulder shapes.

61
Experiments on tsunami induced boulder transport – a review

Figure 4-1: Basic forces acting on a solid body at rest. Fg = force due to gravity and weight; FgH = plane-parallel compo-
nent of Fg; FgN = plane-normal component Fg; Fl= lift force; Ff = friction force.

Figure 4-2: Forces acting on objects by wave impact. Compared to Figure 4-1, additional dynamic forces act on the
body: drag force (FW), dynamic-pressure forces (Fp and Fhs; see Figure 4-1 for explanation). The variability of force
contributions for the different cases is indicated by the arrow widths. For a more comprehensive force description further
forces (skin-drag force, virtual mass, impact force) might need to be considered.

4.1.2 Reviewed publications

Table 4-2, Table 4-3 and Table 4-4 provide an overview of existing studies on boulder transport.
Although this review focusses on the transport of natural boulders by high-energy waves, we further
include slight variations of this setup (e.g., transport of shipping container by tsunamis) in Table
4-2. We do not analyze such studies in high detail since the dynamic behavior of buoyant bodies

62
Experiments on tsunami induced boulder transport – a review

differs significantly from boulders of higher bulk densities and, furthermore, they are usually based
on different bathymetric setups such as harbors instead of natural rocky coastlines. However, due
to the similar scope and often comparable findings, we highlight them where it is suitable and
important. Field sites referred to as quasi-laboratories as in Spiske and Bahlburg (2011), where
the on-site conditions of boulder transport by a real tsunami are relatively well known based on
pre- and post-event observations and measurements, are not included in the analysis.

Several studies (e.g., Luccio et al., 1998; Imamura et al., 2008) conducted physical laboratory
experiments as a basis for hydrodynamic equations and subsequent empirical or numerical mod-
els, e.g., for predicting the necessary wave velocity to mobilize a boulder of specific parameters.
Hydrodynamic equations and numerical models on boulder mobilization are regularly based on
empirical and linear relationships between different contributing parameters (e.g., roughness, drag,
flow velocity), and attempt to derive the necessary flow velocity or height by solving the force or
momentum equilibrium of mobilizing and restraining forces (compare also Figure 4-1 and Figure
4-2). In this review, such theoretical models are not considered in detail; instead, the reader is
referred to the syntheses of Sugawara et al. (2014) and Watanabe et al. (2020). Nevertheless,
Table 4-5 provides an overview on the reviewed publications, which are significantly based on
empirical or numerical models.
Table 4-2: Grouping of the reviewed studies. The parameters initial alignment, scaling and study scope are not presented
to maintain readability.

Wave
Buoyancy generation Study focus Body shape Submergence Publication

cuboids / Bressan et al. (2018)


incipient motion subaerial
prisms Lodhi et al. (2020)

submerged
partially submerged Freund (2014)
subaerial

Imamura et al. (2008)


cuboids /
prisms partially submerged Nandasena and
dam break
Natural boulder

Tanaka (2013)
transport pro-
cess
negative Petroff et al. (2001)
buoyant subaerial
Liu et al. (2014)

Luccio et al. (1998)

spheric subaerial Cytrynbaum (2018)*

Cox et al. (2019)*

incipient motion spheric partially submerged Harry et al. (2019)

modified dam break submerged


transport pro-
other partially submerged Oetjen et al. (2020)
cess
subaerial

Nakamura et al.
(2012)
Artificial bodies

Yao et al. (2014)


various
various
positive (see supple- Goseberg et al.
(see supplementary prisms subaerial
buoyant mentary materi- (2015)
materials)
als)
Stolle et al. (2017a, b;
2019)

Rueben et al. (2015)

*Focusing on storm waves

63
Experiments on tsunami induced boulder transport – a review

4.2 Classification of experimental studies on tsunami induced transport of bodies

Physical boulder-transport experiments can be categorized based on the applied wave (e.g., break-
ing/non-breaking long wave, hydraulic bore), boulder shape(s), the targeted parameters (e.g., mo-
bilization threshold, transport distance) or initial boulder setting. Table 4-2, Table 4-3, and Table
4-4 show how the existing experiments are grouped in this paper. Compared to artificial bodies
(i.e., shipping containers), the most crucial difference is density or buoyancy, respectively, which
in natural boulders usually is much higher (with lower buoyancy) than the density of sea water. The
classification following the boulder density is based on the applied wave type. All studies related to
boulder transport apply broken bores of different generation types (e.g., dam-break, pumps). In
contrary, studies focusing on the transport of shipping containers apply solitary waves breaking on
an inclined surface in most cases. An overview on publications regarding experimental investiga-
tions of artificial bodies can be found in Tables B-1 and B-2 in the supplementary material.

For a rough comparison of the hydrodynamic impact forces on the transported objects, a ratio
rBW [-] (Figure 4-3) following

ℎ𝐵
𝑟𝐵𝑊 =
ℎ𝑊
[−] (4-1)

is introduced, where hB [m] describes the object height and hW [m] the wave height.

Figure 4-3: Definition of the boulder-wave rBW [-] ratio.

64
Experiments on tsunami induced boulder transport – a review

Table 4-3: Considered publications on tsunami-induced boulder transport and their main characteristics (part a).

Boulder
Boulder dimensions Boulder Initial water Investigation of inter-
density Boulder (a x b x c axes or diameter) Flatness In- level at boulder action between multi-
Publication Scope Boulder type Buoyancy [kg/m³] types [cm] dex Wave Wave generation location ple boulders

Luccio et al. 1400 to diameter = 10


transport spherical negative 1 - hydraulic bore dam-break dry 
(1998) 2400 hB = 4 - 6

0.8 x 0.8 x 0.8


Petroff et al. regular cubes
transport negative 2717 14 to 1 to 3.65 hydraulic bore dam-break wet 
(2001) and prisms
3.66 x 1.6 x 0.72

1.6 x 1.6 x 1.6


Imamura et al. regular cubes 1550 to
transport negative 6 to 1 to 2 hydraulic bore dam-break wet 
(2008) and prisms 2710
3.2 x 3.2 x 3.2; 3.2

1.5 x 1.5 x 1.5


Nandasena and regular cubes 1985 to
transport negative 15 to 1 to 3 hydraulic bore dam-break dry and wet 
Tanaka (2013) and prisms 2880
8x4x4

regular cubes 1980 to 8x6x3


Freund (2014) transport negative 6 1 to 2.33 hydraulic Bore dam-break dry and wet 
and prisms 3100 5x5x5

Liu et al. (2015) transport regular prism negative 2400 1 12 x 20 x 12 1.3 hydraulic bore dam-break dry 

3 x 2.8 x 2
Bressan et al. incipient mo- regular cubes 1900 to
negative 4 to 0.96 to 4.1 hydraulic bore dam-break dry 
(2018) tion and prisms 2600
3.14 x 5.92 x 3.08

incipient mo- saturated sedi-


Harry et al. (2019) spherical negative 2600 3 diameter = 1.2; 1.6; 2.5 - hydraulic bore modified dam-break 
tion ment

regular and ir- 13.7 x 15.5 x 3


Oetjen et al. modified dam-
transport regular negative 2200 3 8 x 14 x 6 1.83 to 4.9 hydraulic bore dry and wet 
(2020a) break
shaped ca. 9.6 x 17.5 x 7.6

3x3x3
incipient mo- regular cubes
Lodhi et al. (2020) negative 2670 3 to 1 to 2 hydraulic bore dam-break dry 
tion and prisms
7x5x2

65
Experiments on tsunami induced boulder transport – a review

Table 4-4: Considered publications on tsunami-induced boulder transport and their main characteristics (part b).

Peak wave Peak flow ve-


Publication height locity Ratio Initial boulder loca- Initial boulder Backwash con- Friction or bottom rough-
[cm] [m/s] rBW = hB:hF Shore model Scale tion alignment sidered? ness
kinetic friction coefficient
Luccio et al.
1–9.25 ~1-2 0.67 inclined / horizontal not physically scaled subaerial -  0.1 ± 0.01 and
(1998)
0.4 ± 0.04

Petroff et al. 0°, 90° coarse sand


6.5 1.55 0.12 inclined / horizontal not physically scaled subaerial 
(2001) flat, high (d50 = 0.84)

Imamura et al. kinetic friction coefficient


n.p. 1.5 - inclined not physically scaled partial submerged 0°, 90° 
(2008) 0.5 – 0.7

submerged,
Nandasena and inclined ending on kinetic friction coefficient
9.7-15 ~ 0.7–1.3 0.15 to 0.26 not physically scaled partially submerged, 90° 
Tanaka (2013) horizontal bed ~ 0.5
subaerial
submerged,
kinetic friction coefficient
Freund (2014) 8-14 1.5–1.8 0.43 inclined not physically scaled partially submerged, 0°, 90° 
0.35 – 0.5
subaerial

inclined ending on
Liu et al. (2015) 11.8–13.25 ~ 1.25–1.37 0.92 not physically scaled subaerial 0°, 45°, 90°  n.p.
horizontal bed

Bressan et al. static friction coefficients


0.6–3.75 ~ 0.25 - inclined not physically scaled subaerial 0°, 45°, 90° 
(2018) 0.51 – 0.65

soil with grain diameter


Harry et al. (2019) 2 8.0–8.2 0.6 to 1.25 flat soil 01:40 subaerial - 
0.21 mm

submerged,
Oetjen et al. inclined  kinetic friction coefficient
16 0.7 0.38 not physically scaled partially submerged, 0°, 45°, 90°
(2020a) stepped (partially) 0.21 – 0.6
subaerial

Manning coefficient
Lodhi et al. (2020) n.p. 0.3-0.4 0.19 inclined / horizontal not physically scaled subaerial 90° 
0.42

n.p. = not published

66
Experiments on tsunami induced boulder transport – a review

Table 4-5: Review publications which are significantly based on empirical or numerical models.

Publication Model focus Type


Luccio et al. (1998) Initial cobble velocity after wave impact Analytical / empirical
Complex model for the whole transport pro-
Imamura et al. (2008) Numerical
cess
Complex model for the whole transport pro-
Nandasena and Tanaka (2013) Numerical
cess
Bressan et al. (2018) Velocity threshold for boulder mobilization Analytical / empirical
Harry et al. (2019) Velocity threshold for half buried spheres Analytical / empirical
Lodhi et al. (2020) Velocity threshold for boulder mobilization Analytical / empirical

As can be inferred from Table 4-2, all investigations on artificial bodies consider positive buoyant
objects (e.g., shipping containers), while those on boulder transport consider negative buoyant
clasts. Furthermore, investigations on artificial objects are mostly bounded to specific conditions
resembling man-made prototypes (e.g., harbors, shipping containers), whereas experiments on
natural boulders use simplified models of natural conditions (shoreline setups, transported clasts).
Artificial object-related experiments are usually properly scaled due to the well-known prototype
dimensions. The scaling varies between 1:20 (Rueben et al., 2015) and 1:75 (Nakamura et al.,
2012). Studies on natural boulders often lack specific scaling ratios between wave parameters and
boulder models (e.g., Bressan et al., 2018; Oetjen et al., 2020a). The focus on engineered envi-
ronments in artificial-object studies (harbors, coastal settlements) implies more clearly defined and
regularly shaped shore prototypes. In this context, all artificial-object studies consider a subaerial
pre-transport setting, while those using natural boulders consider also submerged and partially
submerged scenarios, thereby increasing the number of variables and multiplying the number of
experimental runs. Furthermore, natural clasts are not regularly shaped and many transitions be-
tween cuboids/prisms and spheres/ovoids of diverging densities are found in the field (e.g., Costa
et al., 2011; Nandasena et al., 2013).

In principle, physical experiments are designed to resemble natural conditions in a controlled en-
vironment (i.e., the test facility) in order to measure and quantitatively deduce specific physical
parameters. This requires simplifications such as downscaling the nature prototype, leading to re-
sults of more general validity. Most studies on natural boulders apply two main simplifications (be-
side downscaling): a) simplifying the boulder to shapes, which are easily describable, and b) using
shore models without irregularities in the setup (e.g., Imamura et al., 2008; Nandasena and
Tanaka, 2013; Bressan et al., 2018). This approach, however, implies shortcomings regarding
some transport-controlling parameters such as boulder shape. To overcome these shortcomings,
clasts of different regular shapes (e.g., Imamura et al., 2008; Bressan et al., 2018) are applied,
described by the Flatness Index (FI):

(𝑙𝐵 +𝑏𝐵 )
𝐹𝐼 = [−] (4-2)
2ℎ𝐵

where lB [cm] is the boulder length, bB [cm] the boulder width and hB [cm] the boulder height (Nan-
dasena and Tanaka, 2013). Generally, the FI is only valid for cuboid objects, whereas complex,
transitional shapes are most common in nature, as can, for instance, be inferred, for instance, from
the widely scattered distribution of field datasets in the shape diagrams of Costa et al. (2011) and
Abad et al. (2020). Only few experimental studies have considered spherical shapes (Luccio et al.,
1998; Harry et al., 2019), whereas Oetjen et al. (2020a) even applied replicas of real-world boulders

67
Experiments on tsunami induced boulder transport – a review

from a field site in their experiments for comparing the influence of idealized, cuboid and natural
shapes on boulder transport.

So far, all studies on natural boulders apply dam-break setups. However, the exact type of dam
breaks differs significantly regarding the gate-opening mechanism (swinging, sliding). Further de-
velopments encompass dam break-like mechanisms such as pumping (Oetjen et al., 2020a) or
modified setups, where the dam break is realized using a centrifugal setup, in which even the
amount of gravity force is scalable (Harry et al., 2019).

4.3 Boulder transport experiments

4.3.1 General

The majority of boulder-transport experiments use a wave flume with an inclined beach and gen-
erate the bore by a dam-break setup. Several studies (e.g., Nandasena and Tanaka, 2013; Liu et
al., 2015) use a shore model where the inclined section ends on a flat section. An exception is
represented by the setup of Hansom et al. (2008), who implemented a downscaled cliff of 15 m in
nature in order to study processes of cliff-top erosion and transport by storm waves. Different ap-
proaches regarding the pre-transport boulder setting are applied. The most common setup depicts
the subaerial one, fewer ones apply partially submerged setups, and submerged setups are the
least common ones. Further differences emerge from the pre-transport boulder alignment to the
wave impact. Studies considering different wave impact angles commonly apply main-axis align-
ment of 0° and 90°, while some also apply 45° (Liu et al., 2015; Oetjen et al., 2020a). Despite the
general experimental size (not to be mistaken as the model scale) the largest variability in initiation
of motion and the transport patterns arises from the boulder model itself. Density values vary from
1400 to 3100 kg/m3 leading to density ratios of 1.4 to 3.1 between boulder and water. Since the
fluid density is not scaled in the experiments, slightly skewed experimental results are unavoidable.
Nevertheless, compared to other uncertainties arising from scaling issues (cf. Oetjen et al., 2020b),
the inaccuracy due to density variations are of minor importance, even though they should be
considered in data analysis and comparisons. While most studies focus on investigations of solitary
boulders, Nandasena and Tanaka (2013) for tsunami related investigations and several studies on
the interaction of storm waves and boulder extend the experiments to multiple-boulder setups in
order to also consider clast interaction during wave impact and transport (e.g., Cox et al., 2019).

4.3.2 Wave generation mechanism

Dam-break wave generation


Physical experiments applying a dam-break mechanism for wave generation are commonly based
on swing gates (e.g., Nandasena and Tanaka, 2013) or vertical lifting gates (e.g., Bressan et al.,
2018; Figure 4-4), even if other variations exist (e.g., Aleixo et al., 2011). Simple dam-break setups
can be realized in wave flumes by lifting the gate manually by hand. However, the opening time of
the gate is a crucial parameter influencing the wave formation and slow opening times can intro-
duce significant errors, especially in short wave flumes (Lauber and Hager, 1998; von Häfen et al.,
2019; Stolle et al., 2019a). By considering the results of Lauber and Hager (1998), von Häfen et
al. (2019) conducted numerical investigations based on the smoothed-particle hydrodynamics
(SPH) method for identifying the error induced by slowly opened gates. They identified swinging
gates as introducing smaller errors compared to lifted gates. The size of experiments conducted in

68
Experiments on tsunami induced boulder transport – a review

wave flumes with common dam-break mechanism varies greatly. The high-discharge flume of the
Canadian Hydraulics Center (Ottawa, Canada) was modified by a hinged gate in Goseberg et al.
(2015), providing a maximum water volume of approximately 21 m3 as dam-break volume. In con-
trast, the maximum available dam-break volume in Lodhi et al. (2020) is 0.036 m3, dwarfed by a
factor of 580.

Modified dam-break wave generation


Some advanced mechanisms increase the reproducibility of bore parameters. Harry et al. (2019)
used a centrifuge facility in order to minimize the influence of the non-scaled gravity forces. By
installing the experimental setup in a centrifuge chamber, an artificial gravity force can be applied
(Harry et al., 2019; Figure 4-5). Oetjen et al. (2020a) used a dam break-like mechanism operated
by a combination of pumps and valves. Here, the bore is generated over a wet bed by blocking the
pumped water by an artificial barrier and increasing the water level at this point (Figure 4-4). After
reaching the barrier height, the wave passes the barrier resulting in a bore travelling on the wet
bed. The absence of a gate, which has to be opened, generates highly reproducible waves for
each experimental run (Oetjen et al., 2020a).

Piston-type wave generation


Piston-type wave generation is mostly applied in studies on the transport of artificial bodies (e.g.,
Yao et al., 2014; Rueben et al., 2015; Figure 4-4). Nevertheless, Cytrynbaum (2018) and Cox et
al. (2019) investigated boulder transport by storm waves by applying a piston-type wave maker to
reproduce storm wave spectra. Piston-type wave makers are based on accelerating and deceler-
ating the entire water column by horizontally moving plates through the water basin. This principle
of wave generation is commonly applied for generating shallow-water waves like tsunamis. By
installing inclined shore models, which introduce wave shoaling, also wave breaking can be in-
duced to model the near- and onshore behavior of tsunamis. However, in the experiments of Stolle
et al. (2017a, 2017b, 2019b) and Nistor et al. (2017), an advanced wave-generation mechanism
was applied for experiments on shipping containers in the Tsunami Wave Basin at Waseda Uni-
versity (Japan). In such a case, the solitary wave is generated from a water column that is released
by resolving a vacuum. It approaches the wave basin subaquatically, generating a solitary wave.

69
Experiments on tsunami induced boulder transport – a review

Figure 4-4: Simplified sketches of commonly applied wave-generation techniques.

Figure 4-5: Experimental setup of Harry et al. (2019).By conducting experiments within a centrifugal system, the normal
force can be set to a desired value and, therefore, be scaled with respect to the model dimension (redrawn from Harry
et al., 2019).

4.3.3 Experimental focus

The majority of boulder-transport studies focusses on the transport process itself, considering mo-
bilization (e.g., Bressan et al., 2018), transport mode (rolling, saltating, sliding) or transport path
(e.g., Imamura et al., 2008; Nandasena and Tanaka, 2013), whereas some others test advanced
measuring techniques (e.g., Gronz et al., 2016; see categorization in Table 4-2).

Investigating the boulder-transport process is based on a range of parameters, associated with


wave impact, incipient boulder motion, transport and deposition. Experimental foci of the studies
are difficult to categorize as often several parameters are investigated in slightly different ways, as
no standards have yet been developed for experimental studies. Harry et al. (2019), for example,
investigated the incipient motion of boulders half-buried under sediment, but also applied a novel
measuring technique using a proximity sensor, which allows to detect initial boulder motion with
higher accuracy. Furthermore, most studies considering the whole transport process also investi-
gate the incipient boulder motion, even though not as detailed as more specific studies do.

4.4 Experimental investigations on incipient motion

The most important phase of boulder transport is the initial mobilization under wave impact and
subsequent flow. In this phase, where the boulder itself is not accelerated, the highest resisting

70
Experiments on tsunami induced boulder transport – a review

forces of the whole transport process need to be exceeded for initiating the transport. In this con-
text, the initial boulder configuration is crucial especially in terms of its alignment to the flow and
coupling to the environment (e.g., shielded by topography or other clasts). However, studies par-
ticularly addressing incipient motion are rare in the context of tsunami-induced boulder transport.

Bressan et al. (2018) investigated incipient boulder motion in a flume of 11 m length and 0.5 m
width. The dam break used a 2 m long part of the flume as dam-break reservoir. The dam-break
volume was iteratively increased during the experiments to identify the hydrodynamic conditions
initiating boulder motion. An inclined shore model (1:10) was installed at a distance of 3.0 m from
the dam-break gate. The boulders – two regular quasi-cuboids (2.95 × 3.10 × 3.0 cm and
3.0 × 2.8 × 3.0 cm) and two regular prisms (3.1 × 5.45 × 3.0 cm and 3.14 × 5.92 × 3.08 cm) – were
placed in subaerial position on the inclined shore model. Boulder densities varied between
1900 kg/m3 and 2600 kg/m3. The FI ranged between approximately 1 [−] and 1.5 [−]. Flow depths
between 0.003 m and 0.032 m resulted in rBW (boulder-wave ratio) from 1 to 10 [−]. It should be
noted that Bressan et al. (2018) applied a dimensionless parameter for the incipient-motion dia-
grams following

h𝑛 = h𝑊 /hB (4-3)

which is the reciprocal of rBW. Furthermore, a dimensionless key number for the flow velocity fol-
lowing

𝑣2
𝑣𝑛2 = (𝑔∙ℎ (4-4)
𝐵)

with vn [−] as non-dimensional parameter, v [m/s] for flow velocity and g [m/s2] for the acceleration
due to gravity was applied (Bressan et al., 2018). The experimental results were compared to own
hydrodynamic equations describing the velocity threshold for boulder mobilization. As initial align-
ment, Bressan et al. (2018) considered 0°, 45°, and 90° scenarios. Incipient boulder motion oc-
curred regularly for values of rBW < 1 [−], i.e., before the bore height reached the boulder height.
The authors found that their hydrodynamic equation for boulder mobilization generally overesti-
mates the velocity threshold. In particular, it was shown that entrainment occurs well before the
boulder is completely submerged, and that flow velocities were significantly below the threshold
values predicted by the hydrodynamic equation.

Cubic boulders were found to be mobilized when they were covered by the flow by approximately
30 % to 40 %. Velocity thresholds (𝑣𝑛2 ) between 0.02 and 0.1 [−] were identified, whereas the hy-
drodynamic equations predicted much higher values of 𝑣𝑛2 ≤ 0.5 [−]. For rectangular boulders, flow
heights of approximately 40 % of the boulder height were identified. Furthermore, and as to be
expected, the heavier block required higher flow velocities (higher drag force) for mobilization. The
entrainment threshold velocity was found to be highly influenced by the initial boulder alignment.
An alignment with the largest face perpendicular to the wave resulted in a lower minimum flow
velocity compared to a parallel alignment. Generally, lighter boulders providing larger wave impact
areas required lower flow depth and flow velocity. Over the entire series of experiments, Bressan
et al. (2018) recognized a highly sensitive response of boulders to bore impact and a strong diver-
gence of results. In total, three velocity regions were identified: 1) High-velocity regions in which
entrainment is highly likely, 2) an intermediate region covering a wide velocity range in which mo-
bilization might occur, and 3) a low velocity region, where mobilization is impossible.

71
Experiments on tsunami induced boulder transport – a review

Harry et al. (2019) applied spherical bodies on a sediment bed. They conducted experiments on
incipient motion of spherical clasts, which were half-buried in sediment. Three bodies of different
diameter and two setups were used. A proximity sensor beneath the body in one trial allowed to
determine the moment of mobilization with high accuracy. One particular finding of Harry et al.
(2019) is, that inertia forces, which are commonly assumed to highly influence the incipient motion,
are less important, as they are already significantly diminished at the moment of mobilization. In-
stead, the net upward force is crucial for the mobilization of the half-buried spherical bodies. The
drag forces do not contribute significantly to boulder overturning. However, Harry et al. (2019) in-
vestigated spherical bodies instead of the commonly applied cuboids, which highly influences the
drag force. Furthermore, compared to analytical hydrodynamic equations based on force and mo-
ment balances with linear combinations of drag, lift, inertia and body forces (e.g., after Noji et al.,
1993; Petroff et al., 2001; Nott, 2003; Imamura et al., 2008; Nandasena et al., 2011a), they found
a delay for boulder mobilization after the bore impact. This was assigned to two major points: 1)
Since the body was buried in water-saturated soil, the pore-water pressure might have added re-
sisting forces. 2) The applied equations were specified to fully submerged steady flows, which was
not the case here. Due to the half-buried setting and by assuming that the surrounding soil was not
moved by the spherical body, the body could only be shifted in rotational movement due to net-
upward forces (e.g., lift force). Subsequently, the lift force had to exceed the body’s buoyant weight.
Finally, it seems that the approach of linear hydrodynamic equations is not sufficient for determining
the entrainment threshold (Harry et al., 2019). However, the investigations of Harry et al. (2019)
differ significantly from those of the other considered studies on incipient motion (Bressan et al.,
2018; Lodhi et al., 2020) due to the advanced modelling approach and the buried-boulder setup.
Therefore, comparisons are complicated. Nevertheless, the advanced measurement technique
and the attempt to incorporate the scaled normal force (which significantly decrease scaling inac-
curacies) are highly relevant to future research directions.

Lodhi et al. (2020) used 2.91 m of an 8 m long, 0.3 m wide and 0.5 m high flume for dam-break
wave generation. They applied an inclined shore (2:5) ending on a horizontal apron and conducted
the experiments in subaerial conditions. Three dam-break volumes were applied: 0.036 m3,
0.039 m3 and 0.042 m3. Between the dam-break gate and the inclined shore, an initial water level
of 1 cm was kept throughout the experiments. Since Lodhi et al. (2020) focused on incipient motion,
backwash was not considered. The three boulder models had a density of 2670 kg/m 3 and covered
cubic (FI = 1) and prism shapes (R1: FI = 2.75; R2: FI = 3). The boulders were aligned with their
long axis perpendicular to the flow (90°). With Froude numbers between 0.6 and 0.9 [−], the flow
speeds were between 0.3 m/s and 0.4 m/s, depending on the dam-break volume. While sliding
was the predominant transport mode, secondary motions as swinging and rotation were also ob-
served. Repeated runs of identical experimental setups showed differences in the transport dis-
tance and the transport path. While the cubic boulder and R1 were mobilized by flow velocities of
0.3 m/s, boulder R2 was not mobilized below flow velocities of 0.4 m/s. The higher resistance of
R2 was attributed to its higher weight and subsequently larger frictional effects. Furthermore, the
boulders were not mobilized until they were at least partially submerged by the incoming flow (Lodhi
et al., 2020).

Considering artificial clasts, Yao et al. (2014) and Stolle et al. (2017a, 2017b) investigated the
incipient motion of single and multiple positive buoyant objects. Yao et al. (2014) defined the fric-
tional force as the crucial parameter determining entrainment for positive buoyant bodies. If the
drag forces exceed the frictional forces, the body will be shifted in motion. Yao et al. (2014) further

72
Experiments on tsunami induced boulder transport – a review

stated that the body is mobilized faster in the case of an incoming high-energetic steep bore com-
pared to a weak flow.

4.5 Experimental investigations on tsunami-induced boulder transport

Luccio et al. (1998) investigated the motion of cobbles in the swash zone. They conducted exper-
iments in a flume of 2.5 m length, 0.6 m height and 0.3 m width by applying a horizontal as well as
an inclined shore model of two different bottom-friction values. The initial water level in the dam-
break reservoir H0 was kept between 0.05 and 0.3 m, resulting in initial volumes between 0.009 m3
and 0.054 m3. Luccio et al. (1998) used two floor types with different friction values of
Kfric = 0.1 and 0.4 [−]. Hollowed rounded cobbles of polyester resin and fiberglass with diameters
D = 0.1 m and heights hB between 0.04 and 0.06 m were applied, filled with polystyrene or sand to
realized densities between 1500 and 2400 kg/m3. Their initial setting was subaerial at a distance
of 0.5 m from the gate. Bore height was measured at a distance of 30 cm from the gate where the
bore front reached the cobble. The bore height reached values from approximately h0 = 0.01 m
(smooth bottom) to 0.09 m (rough bottom), resulting in rBW = 4 [−] and 0.67 [−]. The authors ob-
served a relatively high initial clast velocity, which, however, stayed significantly below the flow
velocity. The cobble accelerated to a constant velocity approximately 1.25 s after bore impact. On
the rough bottom, the cobble began to decelerate approximately 1.0 s after bore impact. However,
Luccio et al. (1998) found a surprisingly small influence of bottom roughness on transport distance
(< 20 %), attributed to the detachment of cobbles from the ground at high flow velocities.

Petroff et al. (2001) conducted experiments on boulder transport by tsunamis in a 20 m long,


0.45 m high and 0.6 m wide flume. The dam-break mechanism was controlled by a pneumatically
activated gate at a distance of 7 m (dam-break volume of approximately VDB = 1.26 m3, constant
H0 = 0.3 m), which opened by 2 m/s and emptied the tank within 0.2 s. In the initial setup, the dam-
break tank was followed by a wet (h0 = 0.02 m) and horizontal bed ending on an inclined (1:10)
shore section, roughened with coarse sand. Petroff et al. (2001) applied 14 different boulder mod-
els of regular cubic and regular prism shapes with volumes ranging from 0.512 cm3 (5.12 × 10−7 m3)
to 13.824 cm3 (0.13824 × 10−4 m3). The clast height varied from 0.008 m to 0.0366 m. Considering
a bore height (hB) of 0.06 m, this resulted in an average rBW of 0.37 [−]. Seven initial clast orienta-
tions were investigated. The boulders encompassed FIs between 0.3 and 1 [−] (Petroff et al., 2001).

A key finding was that shapes significantly deviating from cuboids show greater sensitivity to the
initial main-axis orientation. Clasts aligned with their short axis parallel to the flow were rotated
after bore impact until the long axis was perpendicular to the flow. This process was expected to
shorten the overall transport distance. Thus, the often-unknown pre-tsunami alignment introduces
additional uncertainties in analytical investigations. Furthermore, significant variations of the
transport distance (up to 20 %) occurred during experiment repetitions. Petroff et al. (2001) con-
cluded that this effect diminishes if transport patterns of multiple bodies are investigated. Therefore,
they recommended considering not only a single-boulder proxy but rather displacements on boul-
der-field scales. Here, an untypical transport behavior of a single boulder can be obscured by the
overall transport pattern of all boulders. However, interactions between the boulders were not in-
vestigated in Petroff et al. (2001).

Imamura et al. (2008) conducted experiments in a 10 m long and 0.3 m wide flume. The dam-break
mechanism was realized by a lifting gate at a distance of 0.3 m resulting in VDB between 0.135 m3

73
Experiments on tsunami induced boulder transport – a review

and 0.27 m3, depending on the initial value of H0. At the other side of the gate, h0 was 0.012 m and
an inclined shore model (1:10) was installed at a distance of 2.5 m from the gate. Imamura et al.
(2008) investigated four different dam-break volumes (H0 = 0.15, 0.2, 0.25, 0.30 m) and six regu-
lar-shaped boulder models with volumes between ca. 4 cm3 (0.016 × 0.016 × 0.016 m) and ca.
32 cm3 (0.032 × 0.032 × 0.032 m). Densities ranged from 1550 kg/m3 to 2710 kg/m3. The pre-
transport setting was partially submerged. The setup consisted of a second gate in the front part
of the flume. After bore run-up and during the subsequent backwash, the second gate was opened
allowing free backflow of the current and preventing reflections. No exact bore height is given, but
the maximum bore velocity was 1.5 m/s (rBW ≅ 0.14 [−]). The boulders were mobilized on the in-
clined shore after bore impact and decelerated with decreasing current velocity. For the cubic
model mainly rolling and saltating transport in upstream direction were registered until the current
velocity decreased and the mode shifted to sliding. Downslope backwash was associated with
rolling and sliding. Regarding prisms, Imamura et al. (2008) differentiated between cases where
the long axis was perpendicular or parallel to the flow. If the long axis was aligned perpendicular
to the flow, transport was comparable to the cubic models, i.e., mainly rolling and saltating during
run-up. When the short axis was perpendicular to the flow, the boulder tended to rotate after bore
impact and the long axis aligned perpendicular to the flow, leading to energy loss and shorter
transport distances. The recurring mobilization of the boulder was found to be strongly depending
on the initially generated run-up bore. In the case of small bores, the backwash was not able to
mobilize the boulder again. Furthermore, differences in the transport distances were attributed to
variations in boulder density.

Nandasena and Tanaka (2013) conducted experiments in an 18 m long, 0.4 m wide and 0.75 m
high flume. The dam-break mechanism was controlled by a hinged gate with
H0 = 0.3, 0.25, 0.2 and 0.15 m. The dam-break reservoir was 3.5 m long resulting in dam-break
volumes from VDB = 0.21 m3 to 0.27 m3. At a distance of 1.17 m from the gate, an inclined shore
model (1:20) was installed. Nandasena and Tanaka (2013) used 15 different boulder models rang-
ing from 0.015 × 0.015 × 0.015 m to 0.08 × 0.04 × 0.04 m, with densities between 1800 and
2880 kg/m3. Shapes encompassed cuboids with a FI of 1 [−] and prisms with FIs from 1.25 to 3 [−].
Bore heights between 0.1 m and 0.15 m were generated leading to an average rBW of 0.32 [−]. The
pre-transport setting was subaerial. As the only study considering natural bodies, they investigated
setups of boulder accumulations, where the boulders were arranged in one line perpendicular to
the flow. Nandasena and Tanaka (2013) found sliding to be the predominant transport mode for
FIs between 2 and 3, while FIs close to 1 were partially transported by rolling. Saltation was very
rare. Independent from the FI, some cases of mixed transport modes (sliding and rolling) were
observed. In these cases, rolling was attributed to the macro-roughness features of the bed. Nan-
dasena and Tanaka (2013) pointed out that the FI is not a suitable parameter for describing rela-
tionships between transport distance and boulder properties, since it neglects properties unrelated
to shape, such as density or weight. More influence as to the FI was directly attributed to the boul-
der weight and the long axis. Their influences were quantified with exponential determination co-
efficients of R2 < 0.48 (long axis) and R2 > 0.43 (boulder weight). Regarding boulder accumula-
tions, they observed a slightly increased maximum transport distance for single boulders in this
setup, which was attributed to momentum transfers between colliding boulders. Furthermore, for
boulder accumulations large ranges of transport distances were documented between the individ-
ual boulders. Nandasena and Tanaka (2013) concluded that boulder-transport processes are

74
Experiments on tsunami induced boulder transport – a review

highly sensitive and mainly governed by axis length, boulder and fluid density, friction and drag
coefficients, boulder alignment to the flow and the possible interaction between multiple boulders.

Liu et al. (2015) conducted experiments in a wave flume of 18 m length and 1 m width. The dam
break was realized by a lifted gate with initial water levels of 0.4, 0.42 and 0.45 m (resulting in initial
waterheads of ΔH = 0.30, 0.32 and 0.35 m, and VDB between 3.2 m3 and VDB = 3.5 m3). Down-
stream of the gate, H0 was 0.1 m. The shore model consisted of an elevated flat part following a
slope of 3.75:100. The boulders were subaerial, placed on the flat part of the shore. Backwash was
not investigated, as the flume had an outlet. Initial alignments of 0°, 45° and 90° were investigated
with a boulder volume of 1.73 cm3 (0.12 × 0.12 × 0.2 m; FI = 0.92 [−]) and a density of 2400 kg/m3.
Bore heights from 0.12 m to 0.13 m result in rBW = 0.96 [−]. For the 90° and 45° setting, Liu et al.
(2015) found a threshold for boulder mobilization regarding the initial waterhead of 0.32 m, since
for the 0.30 m case no mobilization occurred. This finding corresponds to maximum flow heights
of 12.3 cm (ΔH = 0.32 m) and 11.8 cm (ΔH = 0.30 m) accompanied by maximum flow velocities of
1.33 m/s (ΔH = 0.32 m) and 1.37 m/s (ΔH = 0.30 m), measured at a distance of 3 m from the gate
and 3 m in front of the initial boulder setting. Even if the maximum flow velocity of the ΔH = 0.32 m
setting was slightly below the value for ΔH = 0.30 m, two velocity peaks occurred resulting in a
longer period of high velocity in total, while for ΔH = 0.30 m the flow velocity immediately decreased
after the first peak. In the 45° scenario, the main axis oscillated after the bore impact and aligned
to 90°, whereas the transport distance was diverging. For ΔH = 0.35 m, the transport distance for
the 45° scenario was generally below the one for the 90° scenario, whereas the 45° transport
distances exceeded 90° for ΔH = 0.32 m. Liu et al. (2015) attributed this to the larger projected
area for the 45° scenario and energy losses due to oscillations/rotations. No mobilization was ob-
served for the 0° setting, attributed to the small projected impact area. The general transport mode
was sliding, apart from one single experimental run of rolling, where also the transport distance
was shorter due to increased energy dissipation from rolling. Furthermore, a time lag of approxi-
mately 1 s was observed between bore impact and initiation of boulder motion for an initial water-
head of 0.32 m, attributed to a low drag force from the low water level of the bore tip. From the flow
and moving plots in Liu et al. (2015) it can be seen that the flow velocity already significantly de-
creased (from a maximum of ca. 1.2 m/s to ca. 1.0 m/s), while the water level was still increasing
(ca. 0.08 m at the boulder location). However, the bore velocity was measured at a distance of 3 m
from the boulder and the actual value at the boulder is therefore not available. With increasing
water level, the drag force increased, while resistance forces decreased due to buoyancy. Com-
paring the initial waterheads of 0.32 m and 0.35 m, the boulder was transported approximately
twice as far in the 0.35 m case, where the time lag between bore impact and boulder mobilization
was reduced to 0.5 s. After mobilization, the differences between the 0.32 m and 0.35 m cases
only occurred after the first 3 s of the transport process. During transport, Liu et al. (2015) observed
three phases: acceleration, steady movement and deceleration. Differences in the transport dis-
tance are mainly attributed to the last two phases. Furthermore, Liu et al. (2015) observed asyn-
chronous bore and boulder velocity patterns in many of their experimental runs.

Freund (2014) investigated six boulder types of two dimensions (0.08 × 0.06 × 0.03 m,
0.05 × 0.05 × 0.05 m) and three densities (1980 kg/m3, 2400 kg/m3, 3100 kg/m3) in a flume of
19.05 m length. Freund (2014) applied a dam-break setup, an inclined shore of 1:10, boulder align-
ments of 0° and 90°, and submerged, partially submerged and subaerial settings. The prism-
shaped boulder predominantly slid, while the cuboid was mostly transported by rolling. The author

75
Experiments on tsunami induced boulder transport – a review

observed an influence of boulder weight or density, respectively, but no remobilization by back-


wash. Regarding incipient motion, Freund (2014) separated between subaerial and submerged
settings. In submerged settings, there was a clear time lag between bore arrival and mobilization
while the subaerial boulder was mobilized immediately after bore impact.

Oetjen et al. (2020a) conducted experiments in a wave flume by applying three shore types: uni-
form inclination, varying inclination and a stepped shore. The bore was realized by a combination
of pumping and a barrier, which broke the flow (Figure 4-4). Three boulder shapes were applied: a
regular prism (FI = 1.83), a flat (FI = 4.9) and a naturally shaped (FI ≅ 1.58) boulder; the latter was
based on photogrammetric on-site measurements at a field site on Bonaire (Caribbean Sea) (cf.
Engel and May, 2012). As experimental setups, initial boulder alignments of 0°, 45° and 90°, as
well as submerged, partially submerged and subaerial settings were investigated. As in Liu et al.
(2015), three transport phases were identified. One observation was, that the boulder stopped
almost entirely during the transport process, before it accelerated again. Similar to most published
studies (e.g., Liu et al., 2015), Oetjen et al. (2020a) found sliding to be the predominant transport
mode. Regarding the experimental reliability, similar variabilities as in Petroff et al. (2001) occurred
regarding strongly diverging transport distances. The larger the average transport distances (e.g.,
90° initial alignment and subaerial conditions), the higher was the standard deviation (up to 650 %).
The natural boulder was generally transported shorter compared to the regular prism of the same
volume and density, while the flat boulder (FI = 4.9) showed the shortest distances. This was
mainly attributed to the streamline shape of the natural boulder and the small drag coefficient. It is
stated, however, that the effect of the boulder roundness might be superimposed by other param-
eters such as initial submergence or alignment. The investigations on the stepped shore revealed
that not only the backwash needs to be considered in natural settings. Other topographic features
might introduce reflectance of the run-up flow, attenuating energy for boulder transport (Oetjen et
al., 2020a).

4.6 Model comparability

Experimental studies are often based on diverging boundary conditions in terms of their dimen-
sions, scaling, wave type, shore type, as well as boulder size, shape and density. In the following
section, the main differences between experimental approaches are analyzed, and the compara-
bility between the studies regarding specific mobilization and transport parameters is assessed.

Physical experiments on tsunami-induced boulder transport are mostly not physically scaled, since
the focus lies more on the general understanding of the forces involved in the process. Thus, the
applied model setups encompass several combinations of boulder and wave dimensions, which
are not based on nature prototypes. Figure 4-6 and Figure 4-7 show the broad spectrum of dimen-
sions of the studies reviewed here. As the only more or less stable constant in all publications
based on dam-break setups, the dam-break height H0 is around 30 cm (except for Lodhi et al.,
2020, and Bressan et al., 2018, where the dam-break height is not given).

However, boulder dimensions and their dimension-ratio to dam-break volumes, length of the sec-
tion between dam-break and inclination, and inclination angle differ significantly resulting in diverg-
ing hydrodynamic conditions during wave impact (Figure 4-6, Figure 4-7).

76
Experiments on tsunami induced boulder transport – a review

Figure 4-6: Size comparison of the considered physical experiments on tsunami induced boulder transport.

Figure 4-7: Size comparison of boulder models in the reviewed studies on tsunami-induced boulder transport. From
each publication a dimensional representative boulder model is chosen.

4.6.1 Boulder density and buoyancy

In the considered studies on boulder transport, the boulder density varies from 1400 kg/m3 to
3100 kg/m3. As shown by Nandasena and Tanaka (2013), for example, the influence of boulder
density is significant, and deviations must be considered when comparing different studies. Espe-
cially considering experiments on artificial and positive buoyant bodies, e.g., shipping containers,
the body behavior varies significantly compared to studies on negative buoyant boulders. In this
case, the whole transport process from entrainment through backwash to final position is influ-
enced.

4.6.2 Model scale

All experimental studies are based on the Froude scaling laws. In experimental studies following
Froude scaling laws, the Froude number is the same for model and prototype, while other dimen-
sionless quantities (e.g., Reynolds number) cannot be scaled accordingly (Oetjen et al., 2020b).
Holding the Froude number ensures that the similarity of inertia and gravity forces is guaranteed.
Furthermore, independent of the actual model dimensions, physical experiments of the same

77
Experiments on tsunami induced boulder transport – a review

Froude number can be compared in terms of the wave properties, especially regarding the ratio
between flow velocity and length scale. Nevertheless, the Froude number alone does not ensure
comparability between two experimental models of boulder transport, even if they are based on
the same shore and boulder geometries. The ratio between boulder size and wave velocity or
height, boulder and shore material (friction between boulder and shore, respectively) and the range
of validity of Froude scaling (e.g., due to surface tension) influence the transport process and there-
fore the comparability.

Figure 4-8a shows the ratios between impact area (includes boulder height) and Froude number,
where this value is reported (Imamura et al., 2008; Nandasena and Tanaka, 2013; Freund, 2014;
Lodhi et al., 2020; Oetjen et al., 2020a). The ratio between the square-root of the boulder impact
area and the Froude number is used for clarification (for both parameters the maximum given value
is considered here). High ratios indicate that high Froude numbers were applied for boulders of
comparably small impact areas. This ratio varies significantly throughout the studies resulting in
highly diverging hydrodynamic conditions in the experimental setups. Subsequently, especially
during the wave impact on the boulder, the impact force is not straightforward comparable. Studies
with low ratios, for example, state sliding as predominant transport mode, while those of high ratios
(Imamura et al., 2008; Nandasena and Tanaka, 2013) observe a significant percentage of rolling
or saltating transport. Interestingly, the experiments of Freund (2014) are located in the center of
Figure 4-8a with a ratio close to Lodhi et al. (2020) and Oetjen et al. (2020a), even though the
author observed rolling as the major transport mode for the cubic boulder. This may be related to
a high bottom roughness: The comparably rough, gravel-type bottom in Freund (2014) might have
hampered sliding, if the gravel acted as a counterpoint, restricting sliding transport and increasing
the momentum in the vertical dimension and subsequently supporting a lift of the boulder model.
In this regard, rotating transport might also be supported due to the high bottom roughness. How-
ever, the photographs in Freund (2014) indicate that horizontal rotation started before the vertical
lift. From this observation and since the author stated that rotational movements were more com-
mon where only sliding transport occurred, a clear dependence between rotation and bottom
roughness is not evident.

Across all studies, the difference in rBW (Figure 4-8b) exceeds a factor of 10 between Petroff et al.
(2001) and Harry et al. (2019). Nevertheless, excluding the specific experimental setup of Harry et
al. (2019), the largest difference by a factor of approximately 7.5 arises from comparing Petroff et
al. (2001) and Liu et al. (2015). The mean ratio is about rBW = 0.4 [−]. Liu et al. (2015) already
stated that the fundamentally divergent boulder behavior in their experiments compared to Nan-
dasena and Tanaka (2013) might arise from a rBW that is 2.5 times higher.

Even though the data in Figure 4-8 only implicitly considers boulder characteristics such as density
or volume, they still provide an idea on how different proportions between wave and boulder in
physical experiments influence the results, especially in this field where the experiments are com-
monly not physically scaled.

78
Experiments on tsunami induced boulder transport – a review

Figure 4-8: a) Ratios between the boulder impact areas and Froude number of selected publications. The skewness
throughout the publications is clearly visible. b) Comparison of rBW in the considered publications. The applied ratio
between boulder and wave height varies by a factor of >10. Regarding "classic" approaches only, the main difference is
about a factor of 7.5 be-tween Petroff et al. (2001) and Liu et al. (2015).

4.6.3 General experimental setup

Further hindrance to comparison of different studies arises from the general experimental setup.
The experimental design influences the hydrodynamic conditions and, therefore, the transport pro-
cess. Crucial parameters in this context are the shore model design and friction coefficients. The
majority of studies use inclined shore models with slopes between 1:10 and 1:20. Furthermore, as
shown by Stolle et al. (2019b) for positive buoyant objects, the initial body location (on a flat or
inclined bed) has a major influence on the transport pattern especially in combination with the
divergent wave behavior due to the shore geometry. Other uncertainties are introduced due to
different distances between shoreline and boulder.

In experimental studies of tsunami-induced boulder transport, the tsunami is modelled as a hydrau-


lic bore approaching an inclined plane. While in most considered experiments the complete shore
model is designed as an inclined plane, some studies apply horizontal sections after the inclination
(Liu et al., 2015; Lodhi et al., 2020; Oetjen et al., 2020a). Between these types, the friction forces
differ significantly when the boulder is initially located on the horizontal platform or is transported
over this section. Even though in most studies backwash processes are considered, in some the
flume does not permit them. The flume setup in general is crucial for the comparison of experiments
since the overflow type might influence the transport behavior. Basically, three different types exist:

1) In Luccio et al. (1998), for example, the shore model ends on a rigid flume wall, where the flow
is reflected immediately and, thus, generates amplified backflow. This should not be considered
as realistic backwash as long as this effect is not particular addressed (as e.g., in Nakamura et
al., 2012).

2) In Imamura et al. (2008) the run-up does not reach the shore model’s end, and realistic back-
wash can be modelled and must be considered in the evaluation of results.

3) As in Oetjen et al. (2020a), the major amount of the run-up overflows the shore model freely
and only a small amount is reflected as backwash. Its influence on boulder remobilization and
retransportation needs to be treated carefully in such a setup since only a minor portion of the
run-up transforms to backwash, the effect of which is also spatially restricted (approximately to
inclined shore-model sections).

79
Experiments on tsunami induced boulder transport – a review

4.7 Discussion

4.7.1 Incipient motion

The phenomenological description of incipient boulder movement differs throughout the reviewed
publications. However, all authors describing the mobilization process in detail agree that the boul-
ders are not mobilized immediately upon wave impact, but after a short delay, and that the boulder
velocity stays below the flow velocity (e.g., Luccio et al., 1998; Liu et al., 2015; Bressan et al., 2018;
Oetjen et al., 2020a). Bressan et al. (2018) found boulder mobilization before full submergence at
flow heights of 30 % to 40 % of the boulder height. Lodhi et al. (2020) stated that the boulder needs
to be at least partially submerged before mobilization but also observed cases where full submerg-
ence is needed for mobilization. Both studies conclude that the existing hydrodynamic equations
(referring to enhanced versions of equations based on Nandasena et al., 2011a, and Nandasena
and Tanaka, 2013, respectively) for boulder mobilization overestimated the necessary values for
minimum flow height and velocity.

The setup of Harry et al. (2019) using half-buried clasts also revealed that the often-applied hydro-
dynamic equations on boulder mobilization are not sufficiently accurate. In terms of positive buoy-
ant bodies, friction forces are clearly in the focus of the studies. The importance of frictional forces
as resisting forces was particularly highlighted by Stolle et al. (2017a) and Yao et al. (2014), for
example. Interestingly, Stolle et al. (2017b) observed a clear influence on the mobilization behavior
if multiple bodies were exposed to the flow. Therefore, most probably this is also true for natural
boulders, even though such setups have only once been investigated experimentally for natural
boulders in the context of tsunamis, so far, and were focusing on the transport distance (Nan-
dasena and Tanaka, 2013). However, the findings of Nandasena and Tanaka (2013) indicate that
the overall maximum transport distance increases for boulder clusters.

In Oetjen et al. (2020a), the boulder shape was considered to have a significant influence on boul-
der mobilization showing that streamline shapes resist the impact flow for longer, while the friction
coefficients are in the same range. However, the time lag between the idealized prism and naturally
shaped boulder in Oetjen et al. (2020a) was only the range of 20 ms of a transport process encom-
passing 2 s in total, while in other studies it added up to 1 s (Liu et al., 2015).

All reviewed studies agree that an initial boulder alignment of 0° and 45° initiates rotating of the
boulder until it is aligned with the long axis perpendicular (90°) to the flow, even though this move-
ment may not be straightforward and could involve some irregular spinning (e.g., Petroff et al.,
2001; Imamura et al., 2008; Nandasena and Tanaka, 2013; Oetjen et al., 2020a).

4.7.2 Transportation mode

Sliding was observed to be the predominant transport mode for boulders (e.g., Liu et al., 2015;
Lodhi et al., 2020; Oetjen et al., 2020a) and for artificial clasts (Stolle et al., 2017a; Nistor et al.,
2017). In contrary, Nandasena and Tanaka (2013) and Imamura et al. (2008) documented rolling
or saltation as the major transport mode, and the bottom roughness and flatness were identified
as the most significant factor (e.g., Nandasena and Tanaka, 2013). Generally, it is assumed, that
macro-roughness features on the ground provide edges, which initiate overturning of the boulder,
a finding also made for positive buoyant objects (Stolle et al., 2019b). Once tilting is initiated, ad-
ditional uplift forces act on the bottom side of the boulder resulting in rolling or saltating transport.
Considering the flatness, boulders with a FI close to 1 (cubic boulder) are more inclined to rolling

80
Experiments on tsunami induced boulder transport – a review

transport than those with higher FI (prisms, slabs), which was corroborated by Freund (2014), who
applied a very rough bottom in his experiments and compared cubes and prisms. Here, rolling was
the most common transport mode for cubic boulders, especially for higher wave-impact forces,
while prisms were mainly transported by sliding under the same conditions. The bottom was real-
ized from grain material with a d50 of 3.5 mm. Considering boulder heights between 3 cm and 5 cm,
the ratio between the bottom material and boulder would lead to macro-roughness features in the
range of 15 cm (if packing is considered c. 7.5 cm). Therefore, the observation of rolling transport
for cubic boulders in Freund (2014) corroborates the assumption, that macro-roughness elements
support boulder tilting and subsequent rolling. On the other hand, and like in Nandasena and
Tanaka (2013), rolling as transport mode is coupled to cubic boulders in the experiments of Freund
(2014), and will be more likely for real-world boulders with a low FI. Therefore, specific combina-
tions of boulder shape and bottom roughness seem to support transport modes other than sliding.
The evaluation of such combinations could therefore improve the knowledge on both, the influence
of bottom roughness and triggered transport mode. The influence of boulder density on the
transport mode, however, seemed to be negligible. For all investigated boulder densities, the like-
lihood for rolling decreased with decreasing wave velocity and increasing FI (Freund, 2014).

4.7.3 Reported variabilities

It is obvious that tsunami-induced boulder transport is a highly sensitive process, in which the
boulder-transport patterns are influenced by various parameters, which seem to act randomly to
some extent and are not fully controllable in experiments. This is underpinned by spreading
transport-distance data in different studies (Table 4-6).
Table 4-6: Overview on the spreading of transport distance between individual experimental runs of tsunami-induced boul-
der transport.

Nandasena
Petroff et al. Imamura et and Tanaka Yao et al. Oetjen et al.
(2001) al. (2008) (2013) (2014)* (2020a)
Observed
variations
(Variation between min. 120 % 250 % 20 % 300 % 650 %
and max. transport dis-
tance)

*Experiments on positive buoyant objects.

Both, Oetjen et al. (2020a), for boulders, as well as Yao et al. (2014), for artificial bodies, reported
increasing uncertainties with increasing transport distances. Interestingly, Rueben et al. (2015)
reported a highly repeatable behavior for artificial, positive buoyant bodies.

Bressan et al. (2018) also showed that the threshold value for the flow velocity mobilizing a boulder
may vary significantly. Lodhi et al. (2020) observed a cubic boulder to react highly indifferent to a
given pre-transport location and constant flow velocity of vT = 0.3 m/s. Here, the boulder was mo-
bilized only in two third of all cases. Interestingly, a second prism-shaped boulder was mobilized
in all experimental runs of vT = 0.3 m/s, due to the larger impact area, while it was only mobilized
in one of three runs for a flow velocity of vT = 0.4 m/s. Considering the results of Bressan et al.
(2018) and Lodhi et al. (2020), it seems that especially the threshold velocity for mobilizing a boul-
der is very case-specific and hard to determine accurately. It should be investigated as ranges in
the form of categories such as “no mobilization”, “likely mobilization”, “definitive mobilization”, as
already suggested by Bressan et al. (2018).

81
Experiments on tsunami induced boulder transport – a review

4.7.4 Parameter influence

In the reviewed studies on tsunami-induced boulder transport, various sets of influencing parame-
ters were altered or taken as static in the experiments (Table B-3 in the supplementary material).
Therefore, evaluating a hierarchy of the influence of the different parameters is only possible, if
such a parameter is varied in a particular experiment. In most publications, the dam-break height
or wave parameter, respectively, was varied in different setups, and all authors agreed that the
dam-break height is one of, if not the major parameter determining the distance a boulder is trans-
ported. Also in studies on incipient motion, the dam-break height was the leading parameter to be
varied beside boulder properties (Bressan et al., 2018; Lodhi et al., 2020). As expected, in both
studies larger dam-break heights coincided with the mobilization of heavier boulders. However,
this not necessarily the case when focusing on boulders of the same weight but different size and
density. As a further parameter distorting the linear relationship between dam-break height and the
likelihood of boulder mobilization, Bressan et al. (2018) and Lodhi et al. (2020) agreed on the im-
portance of the impact area. Basically, a larger impact area requires lower flow-velocity and height
thresholds for mobilization. The significant influence of the impact area is observed in most of the
studies (Petroff et al., 2001; Imamura et al., 2008; Liu et al., 2015; Oetjen et al., 2020a). An un-
symmetrical impact area (45°) supports rotation of the boulder due to the uneven distribution of
hydrodynamic forces (Bressan et al., 2018). The effect of rotation, however, is not fully understood,
yet. Bressan et al. (2018) stated that boulders were mobilized by lower velocity thresholds if they
rotated during transport. Imamura et al. (2008) also observed rotational boulder movements with
shortened transport distances compared to cases without rotation. Rotation occurred in cases
where the boulders’ short axis was oriented perpendicular to the wave (0°) and the initial impact
area was small. The findings of Imamura et al. (2008) in combination with Bressan et al. (2018)
might indicate that the influence of the initial boulder impact area superimposes the effect of initially
occurring rotations. Comparisons of the experimental findings in Nandasena and Tanaka (2013)
with their numerical model furthermore corroborate that rotations during the transport process
shorten the maximum transport distance.

Thus, all studies agree on the significant influence of the boulder impact area on mobilization and
transport distance. Generally, the transport distance increases with increasing impact area (e.g.,
Petroff et al., 2001; Oetjen et al., 2020a). On the other hand, some cases were reported where this
was not valid for 45° initial orientation because of the unsymmetrical impact area. In Liu et al.
(2015), for example, the boulder was generally mobilized in the two highest of the three applied
dam-break heights. Surprisingly, for the setup with highest flow velocities (dam-break height of
32 cm), the 45° bolder was transported further than the 90° case, which was explained by the
overall larger impact area resulting from the 45° alignment. However, for the 35 cm dam break, the
flow velocity was below the 32 cm dam break and the 90° boulder was transported further than the
45° boulder. Nevertheless, in all studies it was observed that initial boulder alignments of 0° and
45° tended to rotate until their long axis was perpendicular to the flow (e.g., Petroff et al., 2001;
Imamura et al., 2008; Liu et al., 2015; Oetjen et al., 2020a).

As further highly influencing parameters, the boulder weight and density were identified. As ex-
pected, and stated by Luccio et al. (1998), Imamura et al. (2008) and Nandasena and Tanaka
(2013), for example, the transport distance increased with decreasing boulder density. This may
be attributed to increased friction forces with increasing boulder weight (Nandasena and Tanaka,
2013). However, interestingly Luccio et al. (1998), the only considered study with varied bottom

82
Experiments on tsunami induced boulder transport – a review

roughness, did not observe a significant influence of the bottom roughness on boulder transport
during the first transport phase. For the roughest bottom, the transport distance decreased by ap-
proximately 30 %. The actual influence of the bottom roughness or the contact area between boul-
der and ground, respectively, became more unclear considering that Petroff et al. (2001) found that
if the boulder dimension and corresponding side-specific drag coefficient was normalized following

(4-5)
3
𝑋 = √𝑎 ∙ 𝑏 ∙ 𝑐

𝐶𝐷 = 𝑎𝑣𝑔(𝐶𝐷,𝑎 , 𝐶𝐷,𝑏 , 𝐶𝐷,𝑐 ) (4-6)

𝑋
𝑋𝑁 = (4-7)
𝐶𝐷

the transport distance remained almost unchanged from cubes and prisms. This is remarkable
since it held valid for all investigated boulders (encompassing various orientations and boulder
dimensions; Petroff et al., 2001). In Eqs. (4-5), (4-6) and (4-7), a, b and c describe the boulder
dimensions [cm] in x, y, z direction; X [cm] the normalized dimension; CD [−] the averaged and face
specific (CD,a, CD,b, CD,c) coefficient of drag and XN [cm] the corresponding normalized dimen-
sion/drag factor.

4.7.5 Assessment of the Flatness Index

The application of the FI for describing the likelihood of mobilization of a boulder has to be improved
for future studies (Nandasena and Tanaka, 2013). While the FI is a good factor for describing the
ratio between the boulder axes, important parameters like density and the actual impact area are
missing. We propose that the research on a parameter describing the likelihood of boulder mobili-
zation apart from boundary and hydrodynamic conditions should be extended. Ideally, this would
be a parameter describing the ratio between the boulder weight (density and volume), its FI and
the impact area.

4.7.6 Considerations on wave type, impact force and probability

Wave type
The reviewed studies focus on investigations on tsunami-induced boulder transport. For future in-
vestigations, stronger links to findings from other research fields like coastal engineering or, tech-
nical tsunami countermeasures, should be established, as they could help to improve the under-
standing of interactions between tsunamis and coarse clasts.

From studies on bore impact on seawalls it is known that the maximum force of the bore is not
exerted immediately. Wüthrich et al. (2018) and Chock (2016) showed that the maximum momen-
tum was applied on the wall, when the bore or surge flow reached a specific amount (67–90 %) of
its total height. This observation matches the finding that boulders are mobilized only after a short
delay following the impact (e.g., Bressan et al., 2018; Oetjen et al., 2020a, 2020b). However, walls
are fixed and restricted from movement. Therefore, the force exertion is not directly comparable
since loose boulder can be mobilized laterally and vertically, and the momentum transfer between
bore and boulder is more dynamic. To the best knowledge of the authors of this review, no exper-
imental research has been conducted particularly addressing the force exertion of wave impact on
moving bodies, so far. Such a study could provide deeper insights especially in the development
of the drag force on bodies moving in flow direction. A raw sketch on how such experiments might
look like, can be found in the supplementary material (Figure B-1).

83
Experiments on tsunami induced boulder transport – a review

Furthermore, no distinction was made regarding the hydrodynamic differences between bores and
surges. Depending on whether the dam break-induced flow approached the boulder on a wet or
dry bed, the hydrodynamics differed significantly. While for bores the highest momentum transfer
and force exertion occurs shortly after impacting a wall, the momentum transfer increases more
gradually in the case of surges (e.g., Robertson et al., 2008; Wüthrich, 2018). Thus, the short-term
impact force is higher in the case of bores. while it is of longer duration in the case of surges. On
the other hand, bores have a more turbulent tip, which becomes even more turbulent due to flow
reflection on the wall. When the vertically deflected flow falls back on the approaching flow, so-
called rollers are created introducing additional turbulences (e.g., Robertson et al., 2008). Espe-
cially for the investigation of mobilization thresholds, clearer wave-type distinctions are necessary
for improving the accuracy of numerical and analytical models, since in addition to flow depth and
velocity also the impact duration should be considered as crucial parameter affecting the possibility
of mobilization. This is known as the Impulse Concept from particle entrainment research as in
Diplas et al. (2008), Celik et al. (2010) and Valyrakis et al. (2011), for example. A review on this
concept can be found in Dey and Ali (2018).

Turbulences and drag


An attempt to consider turbulences in the context of tsunami-induced boulder transport was made
by Bressan et al. (2018) by calculating the instantaneous turbulent energy using the depth-aver-
aged velocity, the average velocity, the density of water and the temporal velocity variability δvt [−].
Here, the δvt [−] was calculated as the standard deviation of the depth-averaged velocity depending
on the average velocity. In accordance to findings of Lloyd (2016), they concluded that the com-
monly applied static coefficient of drag, as in most inverse models of incipient motion, cannot rep-
resent the actual turbulent flow regime satisfactorily, and recommended the application of velocity
ranges instead of fixed threshold values. However, considering the findings of Lloyd (2016), it might
be interesting to investigate a dynamic drag coefficient depending on estimates of the actual flow
turbulences. Another aspect to consider might be, that the assumption of a depth averaged velocity
profile in the approaching bore is not completely valid for the bore front, since the boundary velocity
layer needs to be considered in this region. Here, the velocity is significantly below the velocity of
the layer above (Wüthrich et al., 2018).

From studies on sediment transport, it is known that turbulences and associated pressure fluctua-
tions in the flow can create vertical gradients which are stronger than the horizontal one (e.g.,
Zanke, 2003; Vollmer and Kleinhans, 2007). As stated above, bore tips are highly turbulent. How-
ever, since the ratio between flow depth and particle size is much higher for finer-sediment
transport compared to the transport of boulders by a tsunami, it is not clear if such pressure fluctu-
ations might also support vertical motion of a boulder. Considering small clasts impacted by a
strong tsunami, such processes might be considered as a mobilization-supporting process, which
could also amplify rolling or saltating transport. The transport of boulders as suspension load of a
tsunami is generally considered “physically unrealistic” based on the calculation of flow depth at
several field sites under a range of realistic Froude numbers (Goff et al., 2010).

Most studies on the drag forces on objects under high-energy wave impact (i.e., tsunamis) focus
on fixed stationary structures. Attempts for investigating a time-dependent dynamic drag coefficient
for bodies moved by waves might support the research on tsunami clasts. The reviewed publica-
tions (i.e., Bressan et al., 2018; Lodhi et al., 2020) underline that especially the hydrodynamic
impact force and turbulences have a significant effect on boulder mobilization and transport and

84
Experiments on tsunami induced boulder transport – a review

need to be considered in order to increase the accuracy of existing analytical models. A first attempt
could be to replace the commonly applied drag coefficient by the resistant coefficient CR [−], which
is already applied for research on the impact of surges on walls. Here, CR [−] is basically estimated
according to the drag coefficient but with respect to the time-dependent variability of surges (Gupta
and Goyal, 1975; Árnason, 2005; Wüthrich, 2018). As shown by Árnason (2005), who investigated
bore impact on solid columns, the Froude number of the bore has a major influence on the impact
force. Depending on the shape of the impacted structure (circular, square) and the bore height, the
resistance coefficient decreases with increasing Froude number.

Probability
Beside the exact calculation of force balances on a single particle, probability-based approaches,
for introducing ranges of probable mobilization, similar to the idea of Bressan et al. (2018), are
useful. In sediment-transport research the threshold functions for particle entrainment are also
based on the concept of balanced forces, even though they only express probabilities of entrain-
ment. Such probability functions can focus on both, the probability for entrainment or non-entrain-
ment (e.g., Einstein, 1950; Cheng and Chiew, 1998; see Dey and Ali, 2018 for an extended review).
However, even if extended probability-based approaches account also for instantaneous flow pa-
rameters (in this case connected to bed shear stress; e.g., Grass, 1970), they are developed for
continuous flows rather than for short-term impacts like tsunami bores and are therefore not easily
transferable to the problem of tsunami-induced boulder transport.

4.7.7 Attempt of a guidance tool for mobilization assessment

Equations derived from experimental investigations and on-site observations have been integrated
in a number of recent studies reconstructing minimum flow velocities or wave heights based on
boulders in the field and for inferring hydrodynamic characteristics (e.g., Engel and May, 2012; Lau
et al., 2015; Abad et al., 2020). However, it has become clear that the uncertainties of such equa-
tions related to the initiation of boulder transport, mostly based on Nott (1997) and Nandasena et
al. (2011a), are not negligible (e.g., Bressan et al., 2018; Cox et al., 2020; Lodhi et al., 2020; Nan-
dasena, 2020).

Figure 4-9 tries to summarize the outcome of this review and represents a qualitative tool to support
interpretation of a single boulder in the field and as a supplement to the application of inverse
modelling approaches to derive the initiation of boulder transport (see review in Nandasena, 2020).
The idea of the diagram is to provide additional information regarding if over- or underestimations
have to be expected by the application of common models (either hydrodynamic equations or nu-
merical models). As a basis for the diagram, a standard case of an idealized cubic boulder in par-
tially submerged conditions is defined, with intermediate size or weight and bottom roughness.
From this case, as origin, the user is able to determine if the site-specific field conditions hamper
or support longer transport distances.

1) Determine the shape of the investigated boulder on the x-axis.

2) Evaluate pre-transport submergence on the vertical axis.

3) Follow the diagonal line representing boulder weight parallel to the dashed line by starting
from the intersection of 1) and 2).

85
Experiments on tsunami induced boulder transport – a review

4) Follow the second diagonal line for bottom roughness parallel to the dot-dashed line. Note
that the influence of bottom roughness is the least understood and quantifiable parameter
in the presented diagram.

As the final outcome, the user ends up either directly on the line of the standard case or to the left
or right of it. By comparing the final location with the color bar above the diagram, a trend is ob-
tained indicating if the conditions tend to increase or shorten transport distances. The transport
mode (rolling, saltating, sliding) is not included since current research indicates that the mode itself
depends on a combination of the considered parameters. A brief example of how to use the dia-
gram is given in Figure B-2 in the supplementary materials. However, if users are able to determine
the present transport mode (e.g., by hitting marks on all sides of the boulder indicating recent
rolling/ saltating transport), an increased transport distance should be considered in the case of
rolling or saltating transport in relation to sliding transport.

Figure 4-9: Proposal for a supporting tool for transport estimation. As outcome, the diagram indicates if increased or
shortened transport distances can be expected compared to a standard case

The diagram should be taken as a concept; general validity is not guaranteed for several reasons:

- Throughout the community no experimental “standard case” is defined so far. Considering


the presented studies, various boulder models are investigated in numerical and physical
experiments without a specific case, which is applied as reference in all studies.
- The boulder behavior is highly sensitive and transport distances can vary significantly for
the same wave, boulder and bathymetric/topographic conditions (Oetjen et al., 2020a).
- The influence of the boulder shape is not fully understood yet. More streamline-shaped
boulders are supposed to be transported for shorter distances than cuboid boulders of the
same volume and weight. However, no studies particularly investigating small variances of
the boulder drag coefficients exist so far and, therefore, shapes other than cuboids are
excluded from the diagram in Figure 4-9.
- The diagram focusses on a single boulder only. Boulder accumulations are not considered,
as there is only very limited quantitative information on clast-clast interaction based on
experimental studies (Nandasena and Tanaka, 2013; Cox et al., 2019).

86
Experiments on tsunami induced boulder transport – a review

- Topography and bathymetry (i.e., bed slope) are not considered, which may influence flow
velocities and wave energy, respectively, and contribute to boulder-transport capacity (Bu-
jan and Cox, 2020).

Figure 4-9 may provide crucial information in the field as it includes site-specific conditions in a
qualitative way. For quantitative inferences, further experimental research is required. In the future,
this concept could help determining a specific coefficient by summarizing several crucial parame-
ters for boulder transport. This coefficient, in turn, could be applied to advanced hydrodynamic
boulder-transport equations for enhancing their accuracy, which at the moment is often insufficient.
However, for achieving this, the relevant parameters need to be investigated further, separately in
standardized experimental setups, in order to find functions describing their particular influence on
tsunami-induced boulder transport.

4.7.8 Attempt for a standard case for single-boulder experiments

As shown in Figure 4-9, we propose to establish a standard case on which future experimental
studies on tsunami-induced boulder transport could be based. For such a standard case, a set of
parameters would have to be standardized (Table 4-7). Beside the parameters listed in Table 4-7,
also the general measurement setup should be standardized as much as possible. This could en-
compass the distance between wave parameter measurements and boulder or shore, for example.
Besides that, a common way for publishing these parameters would significantly support compar-
isons between studies.
Table 4-7: Basic set of parameters which need to be standardized for a common “standard setup” for studies of single-
boulder transport.

Component Parameter

Density
Shape
Boulder Dimension
Initial submergence

Height / velocity
Wave Influence of backwash

Shore type / inclination


General setup Bottom roughness
Wet bed

Defining such a standard case would enable to investigate the influence of particular parameters
in detail and to compare experimental results directly to other studies. Subsequently, enhancing
existing or elaborating novel empirically derived hydrodynamic equations on boulder mobilization
and transport would be simplified due to the broad database. Also, and possibly even more, deriv-
ing corresponding coefficients for specific parameters (e.g., drag, lift) would be supported by a
large set of straightforward comparable experimental findings on variations of the same parameter,
while the remaining parameters are kept constant. Furthermore, since the experiments on tsunami-
induced boulder transport are regularly conducted as Froude-scaled models, the standard case
could straightforward be transferred to flumes of varying sizes, for most parameters at least.

87
Experiments on tsunami induced boulder transport – a review

4.8 Summary and outlook

This paper provides a summary of experiments on the transport of rigid bodies by tsunami-like
waves by focusing on natural boulders. Throughout the reviewed studies, many different ap-
proaches are applied. Due to the significantly diverging approaches and setups it is not straightfor-
ward to derive quantitative rules for transport. The main concern here comprises inconsistent di-
mension ratios between particular studies resulting in different hydrodynamic conditions around
the boulder.

The review shows that the range of physical processes contributing to tsunami-induced boulder
transport is not straightforward to be tackled by physical experiments. However, the research so
far reveals several useful insights into the boulder-transport process and points out important future
research directions. The main conclusions regarding transport-influencing parameters can be sum-
marized as follows:

1) The boulder impact area and, thus, the boulder alignment and drag force significantly in-
fluence boulder mobilization and transport and need to be documented as accurately as
possible in on-site studies.

2) The concept of the Flatness Index (FI) should be revised for boulder-transport studies
since crucial parameters (impact area, density) are not considered.

3) The influence of bottom roughness on transport distance might have been overestimated
in the past. Even though it was only systematically investigated in two studies, both did not
observe a significant influence. Experimental investigations incrementally increasing the
bottom roughness could support the understanding of how and to what extent, the bottom
roughness influences the transport distance (especially regarding sliding transport). Fur-
thermore, it needs to be investigated by incremental changes, if there are thresholds of
combinations of roughness and boulder shape that lead to overturning, or a change in
transport mode to rolling/saltation, respectively. Considering the investigations of Freund
(2014), an influence (depending on boulder shape) on the transport mode is probable.

4) Current hydrodynamic equations (sensu Nott, 1997; Nandasena et al., 2011a) for boulder
mobilization tend to overestimate the necessary wave thresholds. Beside improving the
understanding of particular contributing parameters like boulder shape or boulder sub-
mergence, using findings from adjacent research fields (e.g., sediment transport, technical
tsunami countermeasures) could support the accuracy of such equations. Promising entry
points are the Impulse Concept (e.g., Dey and Ali, 2018) or applying probability-based
approaches, which would be in accordance with the findings of Bressan et al. (2018).

5) Tsunami-induced boulder transport is a highly sensitive process influenced by many prop-


erties of the boulder (e.g., density, impact area), wave (e.g., velocity, depth) and environ-
mental conditions (shore type, macro-roughness). Due to the complex interaction between
this set of parameters, we propose to conduct systematic investigations by varying single
parameters in small steps instead of investigating parameter sets across a large range of
values. This approach could be extended in terms of flow type (e.g., spectra of waves or
continuous flow) or shore setting. Through this, studies on storm waves or artificial bodies
could be included and the various research fields (artificial/natural bodies, storm
waves/tsunamis) could be congregated under one key topic. As can be seen from research

88
Experiments on tsunami induced boulder transport – a review

on boulder deposits shifted by storm waves (e.g., Cox et al., 2019), findings are transfera-
ble to a significant degree.

6) The complex interactions of forces and parameters imply that tsunami-induced boulder
transport can (currently) not be treated as an analytically exact topic and working with
ranges (e.g., as supposed by Bressan et al., 2018) seems to be the most appropriate ap-
proach today.

In this context, we recommend that future experimental studies should be based on consistent
boundary conditions, if possible, and only one specific parameter at a time should be varied in
experimental research designs. This would enable establishing a common database in order to
quantitatively evaluate the influence of each relevant parameter. Since laboratories around the
world own many different types and sizes of test facilities, such a standard case might not be based
on scales but on according dimensions, e.g., regarding the ratio between wave and boulder height,
as a first step. This would allow for comparing results, e.g., on the influence of bottom roughness,
to a defined standard case and furthermore to other experimental research focusing on the same
key parameter.

Furthermore, the experimental measurements should be conducted as similar as possible, in par-


ticular in terms of the wave parameters. Defined distances between the measurement of wave
parameters and the shore or boulder, for example, would significantly improve the comparability of
results of two independent experimental studies. Crucial wave parameters like wave height and
velocity should be clearly published, which is often not the case.

Due to the highly sensitive processes, we propose to alter parameters in small incremental steps
only, for instance applying several intermittent shapes from cuboid to ovoid. Focusing on a single
parameter at first will help to improve our general understanding. Boulder models of the same
weight, volume, density, overall shape and ground contact surface could be designed with diverting
wave facing areas, for example, most probably revealing insights in wave-impact physics, which
are currently hard to determine due to the implicit variation of other parameters like the area of the
wave-facing front (impact area) or bottom roughness.

Acknowledgments

This contribution received funding by the German Research Foundation (Deutsche Forschungsge-
meinschaft, DFG) for the project ‘Modelling tsunami-induced coarse-clast transport –combination
of physical experiments, advanced numerical modelling and field observations’ (SCHU 1054/7-1,
EN 977/3-1). We gratefully thank Mr. Switzer and the anonymous reviewer for their helpful and
constructive comments during the peer review process.

89
Experimental models of coarse-clast transport by tsunamis

5 Experimental models of coarse-clast transport by tsunamis

Jan Oetjen, Max Engel, Holger Schüttrumpf


This is an accepted manuscript of a chapter for the book “Geological Records of Tsunamis and
Other Extreme Waves” published by Elsevier in Earth-Science Reviews in 2020. Online available:
https://doi.org/10.1016/B978-0-12-815686-5.00027-4
Keywords: Boulder, tsunami, experiments, wave tank, scaling, hydrodynamic

Abstract
This chapter summarizes background and setup of wave-flume experiments investigating boulder
transport by tsunamis. It emphasizes both opportunities for better understanding boulder transport
in natural settings and limitations, i.e., the sources of uncertainty, related to these experiments. It
also reviews the most important findings published so far, explains linkages with numerical models,
and draws conclusions on research directions and open questions, which need to be tackled in
future studies.

5.1 Introduction

Physical experiments are, beside in situ observations and theoretical analysis, the foundation of
our understanding of how coastal coarse clasts (boulder- and block-sized) are transported by tsu-
namis and how to differentiate whether they were transported by storm waves or tsunamis. They
provide clues to the development of empirical equations and numerical models describing the
transport processes and fundamental mechanics (see Watanabe et al., 2020 and Nandasena,
2020). Rooted in pioneering tank models on the interaction of waves and clastic sediments, such
as those of Bagnold (1940) simulating pebble beach formation, experiments focusing on the inter-
action of tsunamis and coarse clasts are a comparably young field with only a few well-documented
examples (e.g., Imamura et al., 2008; Nandasena and Tanaka, 2013; Bressan et al., 2018).

Experiments on tsunami-induced boulder transport can be conducted in a well controllable and (in
the optimal case) repeatable environment – in most cases a wave flume or wave basin – and,
therefore, provide quantitative data on the governing parameters of boulder transport. However,
an approximation to natural settings and downscaling of models in the lab is associated with sev-
eral shortcomings such as inappropriate replication of bottom roughness, flow turbulences and
trapped air in the flow (especially during the bore impact), or simplified topographies and boulder
shapes.

This chapter presents typical experimental setups of wave-flume experiments on tsunami-induced


boulder transport. It introduces the most important theoretical background for flume experiments
and commonly applied measurement and wave-generation techniques and methods. Furthermore,
the chapter reviews the most important findings published so far and explains the linkage between
physical experiments and numerical models for boulder transport by tsunamis. It concludes with
future research directions and open questions. While there are similar experimental studies on
coastal boulder transport by storm waves (Hansom et al., 2008; Cytrynbaum, 2018; Cox et al.,
2019), these are not systematically covered and considered beyond the scope of this review due
to their significantly different hydrodynamics.

90
Experimental models of coarse-clast transport by tsunamis

5.2 Dimensionless quantities and scaling of experiments

5.2.1 Dimensional analysis

Physical experiments regarding tsunami-induced boulder transport are typically conducted in a


downscaled environment. The downscaling process is not straightforward, since important physical
laws react sensitively toward changing dimensions, exemplified by the surface tension of water: a
small pebble hitting the surface of a filled coffee cup only induces small uniform, circular waves on
the water surface without strong turbulences or white water formation. In the same scenario with a
25 times larger “coffee cup” and “pebble,” we observe turbulent flows around the impact resulting
in locally high Froude numbers (Fr) with white water and distinct splashes.

To be able to describe fluid mechanics in a scaled environment properly, dimensional analysis


(Kobus, 1974) is applied to find depending variables whose combination leads to a dimensionless
quantity. These dimensionless variables can then be used for planning and optimizing scaled phys-
ical experiments and the proper transmission of their results to nature (Hughes, 1993; Martin and
Pohl, 2000). Furthermore, dimensionless variables allow for reducing the number of variables,
which need to be considered in experimental studies, since they are represented by depending
key numbers (Hughes, 1993; Herwig, 2008). For experimental investigations of gravity-driven
waves (like tsunamis), the Froude number is the most important variable and, thus, is derived in
the following section as an example for dimensional analysis.

5.2.2 The Froude number and scaling laws

The Froude number

v
𝐹𝑟 = [−] (5-1)
√𝑔 ∙ 𝑙

(𝑣 [m/s] = velocity, 𝑔 [m/s²] = gravity and 𝑙 [m] = characteristic length), expresses the ratio between
gravity and inertia forces within the hydrodynamic system (Martin and Pohl, 2000). For the following
example, it is important to recognize that every physical variable, e.g., the length l = 10 m, consists
of a quantity (10), dimension (l) and unit (m). In principle, the scaling ratio for a variable in prototype
and the scaled model is defined as follows:

𝑥𝑃
𝑀𝑥 = (5-2)
𝑥𝑀

where 𝑥 stands for an arbitrary variable in the nature prototype (subscript P) or the scaled model
(subscript M). However, for the dimensional analysis, some basic physical dimensions need to be
defined as a ruling basis, which can rely, e.g., on electrical, thermal or mechanical quantities (Gose-
berg, 2011). For hydrodynamic laboratory experiments with an open-water surface and focusing
on gravity-driven waves, a mechanical similarity between prototype and model is intended. Follow-
ing this, the dimensional analysis is based on the following quantities: length in meter [m], time in
seconds [s], mass in kilogram [kg] (Martin and Pohl, 2000; Herwig, 2008; Goseberg, 2011).

Therefore, density 𝜌 [kg/m³], length 𝑙 [m], volume 𝑉 [m³], time 𝑡 [s] and acceleration 𝑎 [m/s²] are
scaled as follows:

91
Experimental models of coarse-clast transport by tsunamis

𝑙𝑃
Length 𝑀𝑙 = (5-3)
𝑙𝑀
𝜌𝑃
Density 𝑀𝜌 = (5-4)
𝜌𝑀
𝑉𝑃 𝑙𝑃3
Volume 𝑀𝑉 = = 3 = 𝑀𝑙3 (5-5)
𝑉𝑀 𝑙𝑀
𝑡𝑃
Time 𝑀𝑡 = (5-6)
𝑡𝑀
𝑙𝑃
𝑎𝑃 𝑡𝑃2 𝑀𝑙
Acceleration 𝑀𝑎 = = = 2 (5-7)
𝑎𝑀 𝑙𝑃 𝑀𝑡
𝑡𝑃2

Scaling of the inertia or gravity forces:

𝐹𝑇𝑃 𝜌𝑃 ∙ 𝑉𝑃 ∙ 𝑎𝑃
Inertia: = (5-8)
𝐹𝑇𝑀 𝜌𝑀 ∙ 𝑉𝑀 ∙ 𝑎𝑀
𝐹𝐺𝑃 𝜌𝑃 ∙ 𝑉𝑃 ∙ 𝑔𝑃
Gravity: = (5-9)
𝐹𝐺𝑀 𝜌𝑀 ∙ 𝑉𝑀 ∙ 𝑔𝑀

Considering Eqs. (5-3) to (5-7), Eqs. (5-8) and (5-9) can be rewritten as:

𝐹𝑇𝑃 𝜌𝑃 ∙𝑉𝑃 ∙𝑎𝑃 𝑀𝜌 ∙M4


𝑀𝑇 = = = 𝑙
(5-10)
𝐹𝑇𝑀 𝜌𝑀 ∙𝑉𝑀 ∙𝑎𝑀 𝑀𝑡2

𝐹𝐺𝑃 𝜌𝑃 ∙𝑉𝑃 ∙𝑔𝑃


𝑀𝐺 =
𝐹𝐺𝑀
=
𝜌𝑀 ∙𝑉𝑀 ∙𝑔𝑀
= 𝑀𝜌 ∙ 𝑀𝑔 ∙ M𝑙3 (5-11)

Equalizing the scale numbers of two dominant forces leads to dimensionless key numbers and
specified relationships, which allow converting the physical quantities between prototype and
model (Martin and Pohl, 2000). In Froude-scaled models, the dominant forces are inertia and grav-
ity. Therefore:

𝑀𝑇 = 𝑀𝐺 (5-12)

or, considering Eqs. (5-10) and (5-11),

𝑀𝜌 ∙ M𝑙4
= 𝑀𝜌 ∙ 𝑀𝑔 ∙ M𝑙3 (5-13)
𝑀𝑡2

Laboratory experiments are generally conducted by using freshwater. The slightly lower density of
freshwater (approximately 2.7% below salt water in a temperature of 25 °C) is usually neglected in
such experiments. Therefore, density and gravity remain the same in prototype and model, and
lead, by considering Eq. (5-13), to the following scaling rule for the time ratio 𝑀𝑡 (considering the
length ratio or geometric similarity, respectively, as the leading constant) in Froude models:

𝑀𝑡 = √𝑀𝑙 (5-14)
𝑀𝑙
Rewriting, extending with and substituting with 𝑀𝑡 = 𝑀𝑙 (Eq. (5-13)) results in
2
𝑀𝑙

M𝑙 M𝑙 M𝑙 M𝑙 2 M𝑙 2 M𝑣 2
1= = 2 ∙ = 2 = 2 = (5-15)
𝑀𝑡2 2
∙ 𝑀𝑔 𝑀𝑡 ∙ 𝑀𝑔 M𝑙 𝑀𝑡 ∙ 𝑀𝑔 ∙ M𝑙 𝑀𝑡 ∙ 𝑀𝑔 ∙ M𝑡 𝑀𝑔 ∙ M𝑙

which becomes

92
Experimental models of coarse-clast transport by tsunamis

𝑣𝑃 2

1=
𝑣𝑀 2 (5-16)
𝑔 𝑃 𝑙𝑃

𝑔𝑀 𝑙𝑀

or

v𝑃 v𝑀
𝐹𝑟𝑃 = = = 𝐹𝑟𝑀 (5-17)
√𝑔 ∙ 𝑙𝑃 √𝑔 ∙ 𝑙𝑀

ensuring that the Froude number is equal in prototype and model. This dimensional analysis is
focused on mechanically driven Froude models and can be carried out for every depending variable
set (Gibbings, 2011). Further explanations and examples are provided by Buckingham (1914),
Hughes (1993) or Martin and Pohl (2000).

The example for a dimensional analysis regarding the Froude number already shows the scaling
law for time in Eq. (5-14). To account for further scaling effects, mechanical scaling laws are intro-
duced, in hydrodynamics determined by the geometric (shape, 𝝀[−]), kinematic (motion, 𝝉[−]) and
dynamic (forces, 𝜿[−]) similarity between prototype and model. In most cases the geometric simi-
larity is the leading constant that determines the size of the model following:

𝑙𝑃
Geometric similarity 𝜆= (5-18)
𝑙𝑀

𝑡𝑃
Kinematic similarity 𝜏= (5-19)
𝑡𝑀

𝐹𝑃
Dynamic similarity 𝜅= (5-20)
𝐹𝑀

with l [m] for length or geometric dimension, t [s] for time and F [N] for force. To keep the ratio for
the dynamic similarity true for hydrodynamic experiments, the acting forces for inertia, gravity, fric-
tion, surface, and spring (elastic) force need to be considered. The similarity laws for these forces
depend, with varying power, on the geometric ratio between prototype and model. Additionally,
inertia and friction forces are influenced by the time ratio between prototype and model. Therefore,
to keep a complete similarity for the dynamic forces, the following equation (5-21) needs to be
fulfilled:

𝜅𝑖𝑛𝑒𝑟𝑡𝑖𝑎 = 𝜅𝑔𝑟𝑎𝑣𝑖𝑡𝑦 = 𝜅𝑓𝑟𝑖𝑐𝑡𝑖𝑜𝑛 = 𝜅𝑠𝑢𝑟𝑓𝑎𝑐𝑒 𝑓𝑜𝑟𝑐𝑒𝑠 = 𝜅𝑒𝑙𝑎𝑠𝑡𝑖𝑐𝑖𝑡𝑦 (5-21)

𝜌𝑃 𝜆4
𝜅𝑖𝑛𝑒𝑟𝑡𝑖𝑎 =
𝜌𝑀

𝜏2
, (5-22)

𝜌𝑃
𝜅𝑔𝑟𝑎𝑣𝑖𝑡𝑦 = ∙ 𝜆3 , (5-23)
𝜌𝑀

𝜂𝑃 𝜆2
𝜅𝑓𝑟𝑖𝑐𝑡𝑖𝑜𝑛 =
𝜂𝑀

𝜏
, (5-24)

𝜎𝑃
𝜅𝑠𝑢𝑟𝑓𝑎𝑐𝑒 𝑓𝑜𝑟𝑐𝑒𝑠 =
𝜎𝑀
∙ 𝜆, (5-25)

𝐸𝑃
𝜅𝑒𝑙𝑎𝑠𝑡𝑖𝑐𝑖𝑡𝑦 = ∙ 𝜆2 , (5-26)
𝐸𝑀

93
Experimental models of coarse-clast transport by tsunamis

𝜌𝑃 𝜆4 𝜌𝑃 𝜂𝑃 𝜆2 𝜎𝑃 𝐸𝑃
𝜅= ∙ = ∙ 𝜆3 = ∙ = ∙𝜆= ∙ 𝜆2 (5-27)
𝜌𝑀 𝜏2 𝜌𝑀 𝜂𝑀 𝜏 𝜎𝑀 𝐸𝑀

with 𝜌 for density [kg/m³], 𝜆 as the length ratio between prototype and model, 𝜏 as the time ratio
[-], E [N/m²] for elasticity, η [N ∙ t/m²] for dynamic viscosity and 𝜎 [N/m] for surface tension.

Assuming that the investigated medium is the same in prototype and model (ρP = ρM), which is
true for water-related experiments, Eq. (5-27) can be simplified to this:

𝜆4 𝜆2
𝜅= = 𝜆3 = = 𝜆 = 𝜆2 (5-28)
𝜏2 𝜏

In order to keep full similarity between prototype and model Eq. (5-29)

𝜅=𝜏=𝜆 (5-29)

has to be fulfilled, which obviously is not possible with Eq. (5-28), except for a scale of 1:1. Thus,
only the two most important forces are commonly considered in scaled laboratory experiments to
keep the geometric ratio (almost) free of choice. Mainly four scaling laws are available for this
issue, each focusing on the similarity of two specific force types (Martin and Pohl, 2000).

For laboratory experiments on boulder transport by tsunamis, the Froude scaling law is the appro-
priate equation since the impact of a hydraulic bore is mainly gravity-driven (Robertson et al., 2011;
Lloyd, 2016; Li and Chanson, 2017). In experiments scaled according to the Froude scaling law,
the Froude number is the same both for the prototype and for the model:

𝑣𝑃 𝑣𝑀
𝐹𝑟𝑃 =
√𝑔∙𝑙𝑃
=
√𝑔∙𝑙𝑀
= 𝐹𝑟𝑀 (5-30)

It keeps the similarity between inertia and gravity forces (holding a constant Froude number), which
are the most influencing forces in gravity-driven, free-surface flows, i.e., tsunami waves. Table 5-1
shows the most important relationships, which need to be considered while planning an experi-
mental hydrodynamic/hydraulic model following the Froude scaling law.
Table 5-1: Froude scaling laws

Parameter Sign Dimension Scaling law


𝑙𝑃 𝑙𝑃
Length l [m] 𝑙𝑀 = ; 𝜆 = (5-31)
𝜆 𝑙𝑀
𝐴𝑃
Area A [m²] 𝐴𝑀 = (5-32)
𝜆2
𝑉𝑃
Volume V [m³] 𝑉𝑀 = 3 (5-33)
𝜆
𝑡𝑃
Time t [s] 𝑡𝑀 = (5-34)
√𝜆
𝑣𝑃
Velocity v [m/s] 𝑣𝑀 = (5-35)
√𝜆

Acceleration a [m/s²] 𝑎𝑀 = 𝑎𝑃 (5-36)

𝐹𝑃
Force F [N] 𝐹𝑀 = (5-37)
𝜆3

If possible, Froude-scaled experiments should not be carried out in the transition zone between
subcritical and supercritical flows (Fr ≈ 1), as the flow could easily flip between both states caused
by small divergences, e.g., in bottom roughness.

94
Experimental models of coarse-clast transport by tsunamis

Scaling effects regarding turbulence and air inclusions in the bore tip need to be considered even
more carefully in Froude-scaled experiments since these are not properly scalable (She and
Leveque, 1994; Chanson, 2009; Nistor et al., 2017).

Further restrictions regarding downscaling arise from possible transitions from turbulent to laminar
flow conditions (denoted by the dimensionless Reynolds number Re, Chapter 5.2.3). However,
recognizing and considering the Froude scaling law and the influence of the Reynolds number in
laboratory experiments does not allow for scaling as geometrically small as possible (e.g., due to
surface tension of water). A minimum flow depth of 2 cm should be realized in the model to avoid
the influence of surface tension on wave propagation, for example (le Méhauté, 1976).

5.2.3 The Reynolds number

If laboratory experiments are Froude-scaled, the Froude number is conserved between the proto-
type and model, as explained earlier. However, beside the Froude number, the influence of the
Reynolds number

𝜌∙𝑣∙𝑑
𝑅𝑒 = [−] (5-38)
𝜂

with 𝜌 [kg/m³] as the fluid density, 𝑣 [m/s] as the fluid velocity, 𝑑 [m] as the characteristic length
(e.g., flow depth or boulder diameter), and η [kg/(m s)] as dynamic viscosity, needs to be consid-
ered when conducting experiments related to boulder transport by tsunamis. The Reynolds number
describes the ratio between inertia and viscous forces and indicates (but does not prove) the type
of flow in terms of laminar (low Reynolds number) or turbulent (generally high Reynolds number)
flow (Figure 5-1). Froude scaling of natural prototypes with only gentle turbulent flows could result
in laminar conditions in the model due to the lower flow velocity in the downscaled environment
inadequately representing the processes in nature (Martin and Pohl, 2000). Therefore, the Reyn-
olds number of the model should be compared to the transition Reynolds number (Recrit), which
describes the threshold from turbulent to laminar flows (e.g., Recrit ≈ 2320 for flow in a pipe), while
planning physical hydrodynamic experiments (Martin and Pohl, 2000). Since tsunami runup is typ-
ically highly turbulent (Yeh, 1991), laboratory experiments following Froude scaling laws require a
similar turbulent flow regime, even if the Reynolds number cannot be conserved between the pro-
totype and model due to Froude scaling.

95
Experimental models of coarse-clast transport by tsunamis

Figure 5-1: Streamlines around a smooth sphere by different Reynolds numbers. The drag coefficient decreases from
(A) to (D). (A) Low Reynolds number. Inertial forces are much lower than viscous force along the sphere. (B) With
increasing Reynolds number, the stream is detached and stationary eddies occur behind the sphere. (C) Higher Reyn-
olds number. The eddies behind the boulder increase until instabilities initiate their detachment and lead to a broad
turbulent stream behind the boulder. The drag coefficient is fairly constant in this range. (D) Very high Reynolds number.
Comparable to (C) but with narrower turbulent stream behind the sphere.

Moreover, a high Reynolds number ensures that inertial forces are dominating in the experimental
setup compared to the viscous forces, which is an important property regarding the influence of
boundary roughness. Since the active friction is a function of fluid viscosity and boundary rough-
ness, scaling of the roughness might result in violating the threshold from rough to smooth hydraulic
conditions (or vice versa) and cause an unintended change in the influence of boundary roughness
on flow conditions (Martin and Pohl, 2000). Generally, reducing the boundary roughness while
ignoring the geometric similarity can help to avoid unintended changes in the hydraulic friction
conditions (Heller, 2011). However, if the hydraulic conditions in the natural prototype are already
close to the threshold from hydraulic-rough to hydraulic-smooth, applying a fluid of adjusted vis-
cosity or operating a distorted (e.g., superelevated) model should be considered (Martin and Pohl,
2000).

Furthermore, when focusing on tsunami–boulder interaction, the nonlinear influence of the Reyn-
olds number on the shape-specific drag coefficient cd [-] should be considered. The drag coefficient
is an experimentally derived and Re-depending (Spurk and Aksel, 2018), dimensionless number
describing the resistance of an object to the surrounding flow (e.g., the tsunami runup) and influ-
ences the force of drag directly following:
1
𝐹𝑑 = ∙ 𝜌 ∙ 𝑣 2 ∙ 𝑐𝑑 ∙ 𝐴[𝑁]
2
(5-39)

where 𝑣 [m/s] is the flow velocity relative to the fluid, and 𝐴 [m²] is the reference area of the object
(commonly the frontal area of the object). The higher the drag coefficient of a particular body, the
more force is exerted to the exposed object during the flow impact. Therefore, the coefficient of

96
Experimental models of coarse-clast transport by tsunamis

drag and, thus, the Reynolds number, have a strong influence on the transport of boulder by tsu-
namis. To keep the coefficient of drag of the transported body in the range of natural prototypes,
the Reynolds number needs to be in the same range even if it is not possible to conserve the exact
value between prototype and model (Munson et al., 2009; e.g., in the range of Re = 10³ [-] to Re = 2
× 105 [-] for sphere, compare (Figure 5-1). Nevertheless, for setups of high Reynolds numbers and
bodies different from smooth spheres, the drag coefficient becomes essentially independent from
the Reynolds number (Munson et al., 2009).

As shown before, it is not possible to keep the Reynolds number the same between prototype and
model if the model is scaled following Froude scaling laws. Reference values for Reynolds numbers
for some idealized boulder models in laboratory boulder transport experiments can be found in
Bressan et al. (2018) or Nandasena and Tanaka (2013). Further advice concerning scaling effects
and corresponding issues can be found in Martin and Pohl (2000) or Fang (2019). Heller (2011)
provides further rules, hints, and examples to deal with scaling effects for several experimental
hydrodynamic setups.

5.3 Measuring approaches in the wave tank

The reliability of information derived from physical experiments is also closely linked to the record-
ing, usage, and interpretation of data. A broad range of measurement devices is available and
chosen based on study goals, parameters of interest (e.g., mobilization threshold or transport dis-
tance), availability of measurement devices and time for collecting, processing and interpretation
of data. It is recommended to focus on a few crucial parameters only and to consider extending
the study in the case that more time is available or the experiments show unexpected results.

Measuring the wave behavior and parameters is an essential component of physical hydraulic
experiments. For wave measurements, the choice and final placement of devices in the flume as
well as the subsequent data analysis need to be planned carefully.

Ultrasonic wave gauges can be a basis for measuring the wave development during the impact on
the boulder and runup. However, at least in a wave basin, these are point measurements of a
three-dimensional phenomenon. Therefore, if the focus of the experiments is to particularly inves-
tigate the wave behavior, the use of 3D measurement systems might be an option (e.g., Evers,
2018).

Propeller velocimeters are often used to measure current or wave velocity, since they are compa-
rably easy to install and use. However, since measurements are based on revolutions of the pro-
peller per time, inaccuracies occur at the first wave impact (propeller needs time to reach full
speed). Furthermore, they suffer from small particles affecting the rotation mechanics leading to
breakage and measurement errors. Other options for velocity measurements are acoustic doppler
velocimetry, electromagnetic liquid velocity meters, laser measurements, particle-image veloci-
metry, and electromagnetic current meters, as described in more detail by Martin and Pohl (2000)
or Tavoularis (2006).

Investigations of the boulder transport path benefit from the use of an inertial measurement unit
(IMU) or a real-time locating system installed inside the boulder replicate (e.g., Gronz et al., 2016;
Nistor et al., 2017), which records parameters such as acceleration, direction and duration of mo-
tion, or altitude. IMUs, however, can be influenced by electronic waves of other instruments in the
laboratory and alter the center of mass and, thus, the behavior of the boulder. Another alternative

97
Experimental models of coarse-clast transport by tsunamis

is the use of high-speed camera recordings in combination with a grid applied on the model’s sur-
face or to combine both approaches. Video analysis requires automated distance measurements
for each recorded frame, which take surface inclinations of the model into account (e.g., by using
MATLAB scripts; Stolle et al., 2016). Multiple camera positions (e.g., side and top view) are rec-
ommended to compensate for white-water disturbances. Figure 5-2 shows an automatized color
range-based boulder-tracking algorithm developed in MATLAB. It can be seen that the algorithm
is affected by colors similar to the boulder color (white water, white lines). While white water re-
mains problematic, disturbances can be avoided by using clearly contrasting colors. Furthermore,
applying an algorithm based solely on color ranges remains a very simple approach. Applying
algorithms accounting for neighboring pixel or applying more sophisticated filters (e.g., Kalman
Filter in Stolle et al., 2016) could lead to significantly better results.

Figure 5-2: Boulder tracking algorithm developed in MATLAB. Image on top shows the initial state. Here, the boulder is
detected by specifying rgb color ranges encompassing only the boulder color by excluding others. Time steps 2 to 5
show the detection shortcomings especially due to the occurrence of white water, the color of which is too similar.

98
Experimental models of coarse-clast transport by tsunamis

5.4 Types of wave generation

Several approaches exist for wave generation. They range from relatively simple techniques for
tsunami bores (e.g., dam-break scenarios; Imamura et al., 2008; Nandasena and Tanaka, 2013;
Freund, 2014; Liu et al., 2015; Bressan et al., 2018), which are most commonly used in experi-
ments addressing tsunami-induced boulder transport, to more sophisticated techniques with wave
paddles or pumps resembling a desired wave series (e.g., Goseberg et al., 2016; Schimmels et
al., 2016).

In the dam-break scenario, the tsunami is not modelled as a solitary wave but as a nearshore and
already broken tsunami of highly turbulent flow with a high Reynolds number (Figure 5-3). The
tsunami bore is based on the sudden release of a specific amount of water stored behind a gate
(slide- or swivel-type valve), which is either opened manually by hand (Bressan et al., 2018), by a
pulley system (Liu et al., 2017), or by a pneumatic mechanism (Hsu et al., 2012). For long waves,
i.e., the unbroken tsunami, a piston-type wave maker is often applied, where pistons displace the
water column of a connected tank in horizontal direction (Figure 5-3). Piston-type wave makers
can also be installed as multiple devices arranged in a wave basin (e.g., Rhinefrank et al., 2010).

Furthermore, the tsunami or bore generation can be based on pumping systems of different types.
In Oetjen et al. (2017a, 2018), the bore is generated by pumping water through a rectifier on a dry
bed, where the water runs against a barrier of a specific height, which generates the final bore
(Figure 5-4). High-end pump and pneumatic systems such as in Goseberg (2011, 2012), or the HR
Wallingford/UCL Tsunami Generator (Allsop et al., 2014), provide more control on the bore in terms
of wavelength, height etc. (Figure 5-4). Goseberg (2011, 2012) developed an advanced mecha-
nism for the generation of long-waves based on pumps, which allows controlling the discharge
using a predefined wave-generation algorithm. Here, the long wave is generated by adjusting ac-
celeration and deceleration phases (positive and negative pumping phase) of the water column to
generate wave crests and wave troughs. In Allsop et al. (2014), the wave is generated as a pump-
driven dam-break scenario.

However, by generating the wave through a controllable tank in which several water depths and
release velocities can be realized, the Tsunami Generator allows creating more specific wave
shapes of desired heights and velocities (Figure 5-4). A significant advantage of computer-con-
trolled systems as in Goseberg (2011, 2012), Goseberg et al. (2013), Oetjen et al. (2017a, 2018),
or Nistor et al. (2017) is an enhanced reproducibility.

99
Experimental models of coarse-clast transport by tsunamis

Figure 5-3: Dam-break and piston mechanism for bore generation. In dam-break setups the bore is generated by sud-
denly opening a gate and the subsequent water release from an impounded storage (A). In piston-type setups (B) the
wave is generated by accelerating and releasing the water column horizontally using a piston-driven paddle.

100
Experimental models of coarse-clast transport by tsunamis

Figure 5-4: Pumping-type wave maker modified after a) Oetjen et al. (2017a): Bore generation by pumping water out
from a reservoir vertically and subsequent gravity driven flow over an installed barrier; b) HR Wallingford (Allsop et al.,
2014): Wave generation by suctioning and releasing water vertically, and c) Goseberg (2011): Accelerating and decel-
erating the water column by pumping water either in positive or negative direction in a looped channel.

101
Experimental models of coarse-clast transport by tsunamis

5.5 Parameters studied in physical experiments

Typical parameters investigated in experiments on tsunami boulder transport (e.g., Imamura et al.,
2008; Nandasena and Tanaka, 2013; Liu et al., 2015; Oetjen et al., 2017a; Bressan et al., 2018)
are the initial boulder setting (submerged, partially submerged, subaerial), wave parameters (ve-
locity, height), orientation of the main boulder axis to the flow (0 degrees, 45 degrees, 90 degrees)
(Figure 5-5), and boulder shape (cubic, natural, flat). In most existing studies, the shore is kept
constant and designed as a uniformly inclined shore with angles between 5 degrees and 12 de-
grees (e.g., Nandasena and Tanaka, 2013; Liu et al., 2015; Bressan et al., 2018). With the number
of varied parameters, the number of experimental runs hugely increases, following the equation
𝑛

𝑁𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑛𝑒𝑐𝑒𝑠𝑠𝑎𝑟𝑦 𝑟𝑢𝑛𝑠 = ∏ 𝑃𝑖 (5-40)


𝑖

where Pi = the number of variations for the parameter i. Thus, investigating three types of pre-
transport setting (submerged, partially submerged, subaerial), three types of main-axis orientation
(0 degrees, 45 degrees, 90 degrees), and two different shapes, the minimum number of runs is 18.
Since experimental runs are commonly repeated at least three times, the preceeding example
results in 54 runs.

Figure 5-5: Typical boulder orientation and submergence in physical experiments.

The parameters control boulder transport in many different ways. Altering the boulder shape in
terms of flatness or cuboid/spherical shapes changes the available impact area for the incoming
wave. The force transferred from the wave to the boulder decreases with a more streamline boulder
shape. On the other hand, a streamlined boulder shape with a natural, rough surface increases the
energy transfer, which is translated to transport capacity. The orientation of the boulder to the flow
influences the energy transfer in a similar way; the long axis aligned perpendicular to the flow
captures more wave energy. In agreement with field observations (e.g., May et al., 2015b; Boesl
et al., 2019), several experiments showed that boulders tend to rotate around their main axis per-
pendicular to the flow immediately after the wave impact (e.g., Imamura et al., 2008). During this
process, energy is dissipated and no longer available for transport. Changing the bottom rough-
ness during the experiments provides information on the resisting force resulting from bottom fric-
tion. Table 5-2 provides a summary of the most commonly investigated parameters, their influence
on boulder transport, and finally the estimated complexity of the investigation.

102
Experimental models of coarse-clast transport by tsunamis

Table 5-2: Examples for parameters of interest in physical experiments and their influence. A qualitative evaluation is given
on the necessary effort to implement parameter changes in the model setup.

Direct influence on

Available Level of
wave Available Available Impact complexity
Impact Duration impact Active Resisting wave transport force to
Parameter force of impact area friction force energy path over time investigate

Easy to dif-
Wave parameter ficult
(depending
(velocity, height, X X X on wave mak-
shape) ing mecha-
nism)

Boulder shape X X X Medium

Boulder density /
weight
X X Medium

Boulder orienta-
tion
X X X Easy

Bottom rough-
ness
X X Difficult

Submergence X X X Easy

Shore
type/shape
X X X X X Medium

5.6 Published wave-tank experiments on tsunami-boulder transport

5.6.1 Experimental setups

Experiments have been conducted with boulders of several shapes, dimensions and densities. The
dimensions range between main-axis lengths of <2 cm up to 15 cm and densities between
1.5 g/cm³ to 2.71 g/cm³. Generally, the boulder dimensions depend on the focus of the study and
on the applied scale. However, most boulder experiments do not follow a natural prototype setting.
To characterize the shape of boulder models, Nandasena and Tanaka (2013) introduce the “flat-
ness index” (FI) for their experiments, which accounts for the ratio between the boulder axes fol-
lowing the equation

𝑎+𝑏
𝐹𝐼 =
2∙𝑐
(5-41)

with a [cm], b [cm], and c [cm] for the boulders’ long, intermediate, and short (height) axes.

Experimental setups of the most relevant published studies are summarized in Table 5-3 and Fig-
ure 5-6 and Figure 5-7. In a further study not yet published internationally, Freund (2014) investi-
gates six different cuboid boulders of 8 x 6 x 3 cm3 (FI = 2.33) and 5 x 5 x 5 cm3 (FI = 1) with den-
sities of 1.98 g/cm3, 2.4 g/cm3, and 3.1 g/cm3, respectively. Freund (2014) uses a 19.05 m long,
0.3 m wide, and 0.35 m high flume in combination with a dam-break mechanism and a shore incli-
nation of 1:20. The author compares different degrees of pre-transport submergence from
subaerial over partially submerged to fully submerged. In the same laboratory, Strusińska-Correira
et al. (2017) study tsunami-induced boulder transport using two different boulder shapes (cubic

103
Experimental models of coarse-clast transport by tsunamis

and cuboids) made of concrete (density ρ = 2.1 g/cm3) with axis dimension ranging from a = 16-
44 cm, b = 16-22 cm, and c = 8-22 cm. An initial water depth of 0.6 m is used in the flume.

Figure 5-6: Comparison of the experimental flumes from Imamura et al. (2008), Nandasena and Tanaka (2013), Liu et
al. (2015) and Bressan et al. (2018).

Figure 5-7: Comparison of the largest applied boulders (in mm) from Imamura et al. (2008), Nandasena and Tanaka
(2013), Liu et al. (2015) and Bressan et al. (2018).

104
Experimental models of coarse-clast transport by tsunamis

Table 5-3: Comparison of the experimental parameters from Imamura et al. (2008), Nandasena and Tanaka (2013), Liu et
al. (2015) and Bressan et al. (2018).

Nandasena and
Tanaka Imamura et al. Bressan et al. Liu et al.
(2013) (2008) (2018) (2015)
l [m] 18 10 11 18

Dimension h [m] 0.75 n.p. n.p. n.p.

w [m] 0.4 0.3 0.5 1


Uniformly
Uniformly Uniformly Uniformly
Shore [-] inclined
Flume

inclined (1:20) inclined (1:10) inclined (1:10)


(3.75:100)

15 15 40

Initial 20 20 dry 42
water level [cm] 25 25 bed 45

30 30 -
Number of
15 6 4 1
different boulder

Dimension [cm] min 1.5 x 1.5 x 1.5 1.6 x 1.6 x 1.6 3.05 x 3.00 x 2.85
20 x 12 x 12
(follwing the volume) max 8x4x4 3.2 x 3.2 x 3.2 5.92 x 3.14 x 3.08

Flatness index [-] 1-3 1-2 1.03 - 1.47 1.33

min 1723* 1550 1900


Density [kg/m³] 2400
max 2880* 2710 2600
Boulder

0 0 0 0

Orientation [°] 45 - 45 45

90 90 90 90
Partially sub-
merged (seems
also subaerial
Submergence Subaerial Partially submerged Subaerial
and submerged
but not clearly
published)
Wave vel. (app.) [m/s] 0.7 - 1.3 1.5 0.25 1.24 - 1.37
generation
Bore

Max. wave height (app.) [cm] 8.7 - 14.7 n.p. n.p. 11.7 - 13.3

Type Dam break Dam break Dam break Dam break

Boulders with Mainly rolling and Boulder experi- Sliding as


FI = 2 and 3 tend saltating transporta- ments are highly predominant
to sliding. tion for cubic boul- uncertain. transport
der. mode.
Boulders with Shorter transport Identification of Shorter total
FI = 1 tend to roll- distance for non-cu- three velocity transport
ing. bic block. zones regarding distance for
the possibility for rolling.
boulder mobiliza-
Transport mode tion: certain boul- Transport
is influenced by der mobilization, distance is
friction. maybe mobiliza- sensitive re-
Key outcomes

tion, impossible garding the


mobilization. bore height /
velocity.
Impact force Boulder
does not signifi- transport
cantly influence can be
the total transport devided in
distance. accelera-
tion, steady
and deceler-
ation phase.
Pre-Transport
angle influence
the total transport
distance signifi-
cantly.
n.p.: not published
*not clearly published whether measured dry or wet

105
Experimental models of coarse-clast transport by tsunamis

5.6.2 Key findings

Nandasena and Tanaka (2013) find different predominant transport modes depending on boulder
shape. While their rather flat boulders (FI ≈ 2-3) mainly slide, boulders of a lower FI of ≈1 tend to
roll. Imamura et al. (2008), however, do not indicate varying transport modes even though the FI
of their boulder models also ranges between 1 and 2. The correlation between transport mode and
FI found by Nandasena and Tanaka (2013) indicates a highly sensitive boulder behavior even if
dimensionless parameters are applied. Liu et al. (2015) observe rolling transport in only one case,
whereas in all other experimental runs the boulder is transported by sliding, showing that the dif-
ferent scales in which the experiments are conducted are key to understand the diverging transport
modes. Imamura et al. (2008) reach flow velocities of up to 1.5 m/s, which is comparable to the
1.4 m/s in Liu et al. (2015). However, Liu et al. (2015) use a boulder of 20 x 12 x 12 cm3, whereas
the largest one of Imamura et al. (2008) measures 3.2 x 3.2 x 3.2 cm3, i.e., a ratio of 3.75:1 (Table
5-3). By applying the Froude scaling law, the velocity of Imamura et al. (2008) becomes the follow-
ing:

𝑣𝑃
= √𝝀
𝑣𝑀
𝑣𝑃 = √𝝀 ∙ 𝑣𝑀
𝑚 𝑚
2.9 = √3.75 ∙ 1.5
𝑠 𝑠
𝑚 𝑚
𝑣𝐼𝑚𝑎𝑚𝑢𝑟𝑎 = 2.9 ≠ 1.4 = 𝑣𝐿𝑖𝑢
𝑠 𝑠
This example shows the relevance of appropriate scaling when comparing results of different ex-
periments and labs (Table 5-4, Figure 5-8).
Table 5-4: Scaled comparison (using scaling laws in section 2) between Liu et al. (2015), Nandasena and Tanaka (2013),
Imamura et al. (2008) and Bressan et al. (2018). No further parameter (boulder dimension, weight, etc.) is considered. If
available, the maximum velocity is taken for comparison. For Bressan et al. (2018), the calculated average velocity is used.
From every calculation, one boulder is taken for comparison which is in a similar range for FI and density. As leading scaling
parameter, the a-axis is considered.

Scale Max bore Max. transport


Boulder dimension to Liu et velocity distance
a b c V FI al. (2015) voriginal vscaled Δdoriginal Δdscaled
[cm] [cm³] [-] [-] [m/s] [cm]

Liu et al. (2015) 20 12 12 2880 1.42 Reference 1.24 1.24 233 233

Nandasena and Tanaka (2013) 6 4 4 96 1.25 3.33 1.32 2.41 119 396

Imamura et al. (2008) 3.2 1.6 1.6 8.2 1.5 6.25 1.54 3.85 140 875

Bressan et al. (2018) 5.92 3.14 3.08 57.3 1.47 3.38 0.25 0.46 n/a n/a

106
Experimental models of coarse-clast transport by tsunamis

Figure 5-8: Graphical comparison of transport distances and flow velocities in Liu et al. (2015), Nandasena and Tanaka
(2013) and Imamura et al. (2008) for scaled and unscaled results (Table 5-4).

Furthermore, Liu et al. (2015) report a time lag of 1 s between bore impact and boulder mobilization,
which has not been reported by Imamura et al. (2008) and is attributed to the significantly larger
wave height to boulder height ratio (Liu et al., 2015). Regarding incipient motion, Bressan et al.
(2018) find that a boulder may start moving by the approaching bore even before it is completely
submerged and that with decreasing flow depth a higher flow velocity is needed to mobilize the
boulder.

Maximum transport distance is strongly influenced by boulder orientation. Nandasena and Tanaka
(2013) find a 27 % shorter distance in case the main axis is oriented parallel to the flow and a 3 %
shorter distance for a 45-degree orientation. In a two-boulder setting, Nandasena and Tanaka
(2013) find a slightly increased maximum transport distance for one boulder at the expense of the
second one, attributed to momentum transfer during boulder-boulder collisions.

Bressan et al. (2018) identify flow turbulences to significantly influence the boulder behavior result-
ing in a widely scattered transport pattern. Thus, they recommend considering value ranges for
flow depth and velocity instead of using exact values when reconstructing past boulder transport
events. In general, the authors define three stages of transport: an acceleration phase, constant
moving, and deceleration. Before the acceleration phase, a phase of beginning mobilization has
been identified, in which the bore impacts the boulder but resisting forces are not exceeded and
the boulder still rests. This phase can further be divided in velocity zones, where the boulder is
certainly, maybe, or impossibly mobilized (Bressan et al., 2018).

Finally, Freund (2014) finds an increasing influence of the flow velocity on boulder transport with
decreasing density of the boulder. The author observes that submerged boulders are mobilized
after a short time lag when the bore impacts, whereas subaerial ones are immediately set in motion,
mostly through sliding.

5.6.3 Further related studies

Besides the quasi two-dimensional studies on tsunami transport of Imamura et al. (2008), Nan-
dasena and Tanaka (2013), Liu et al. (2015), and Bressan et al. (2018), other experiments focusing
on similar hydrodynamic processes and environments also provide important implications (see also

107
Experimental models of coarse-clast transport by tsunamis

Table 5-5). Stolle et al. (2016) and Nistor et al. (2017) conduct experiments regarding the entrain-
ment of shipping containers at a scale of 1:40, where the focus is on the transport pattern of multiple
arranged bodies due to bore impact.

Hansom et al. (2008) investigate cliff-top erosion and deposition of boulders by storm waves by
simulating prototype coastal settings of the northern and western British Isles with steep cliffs,
concluding that storm waves have a high capacity of moving cliff-top boulders.

Cytrynbaum (2018) and Cox et al. (2019) study boulder transport driven by storm waves at a scale
of 1:100. The bathymetry and the shore are designed following the shape of the coastline of their
prototype site on the Aran Islands off the western coast of Ireland (Cox et al., 2018a). The authors
find that, in agreement with field observations, even very large boulders can be transported by
storm waves, even though specific constraints regarding the wave, boulder, as well as coastal
topography and bathymetry are required.

Becker et al. (2015) investigate the behavior of pebbles in a flume, where channelized flow under
fluvial conditions is simulated. In the same flume and with similar conceptual background, Gronz
et al. (2016) test the application of a Smartstone probe, a new IMU, which is installed inside the
clasts and tracks their movement in running water. However, comparing the movement patterns
derived from Smartstones with high-speed camera tracking still reveals significant deviation in ori-
entation and the clipping of data, which the authors hope to reduce in the future through the appli-
cation of probabilistic modeling (Gronz et al., 2016).

Further experiments investigate the interactions between tsunamis and immovable solid bodies
(e.g., coastal defense structures). Such experiments target the flow behavior of tsunamis at the
shore and around buildings (e.g., Esteban et al., 2017) or evaluate the impact force acting on
tsunami-mitigation structures (e.g., Robertson et al., 2008; Chen et al., 2016). Yao et al. (2014)
study the behavior of solid bodies lighter than water under the impact of solitary waves. Key find-
ings of Yao et al. (2014) include the highly sensitive behavior of the bodies, which increases with
energy of the incoming wave. Additionally, they find that if multiple bodies are arranged in a line,
the maximum inundation distance decreases.

108
Experimental models of coarse-clast transport by tsunamis

Table 5-5: Key information from selected experimental studies on transport of rigid bodies under wave attack not in partic-
ular related to tsunami-induced boulder transport.

Becker et al.
Stolle et al. (2016); Hansom et al. Cytrynbaum (2018); (2015);
Nistor et al. (2017) (2008) Cox et al. (2019) Gronz et al. (2016) Yao et al. (2014)
Motivation Entrainment of ship- Cliff-top erosion and Boulder transport by Pebble Behavior of solid
ping containers deposition of boul- storm waves transport/tracking in bodies lighter than
ders by storm waves steady flow water under the im-
pact of solitary
waves
Experi- 3D wave basin Flume 3D wave basin Flume Flume
mental (74 x 4.6 x 2.2 m3) (l = 2.7, (l = 34, w = 0.55 m)
setup w = 0.265 m)

Solitary wave with Storm wave series JONSWAP (Joint Steady flow (1 m/s) Solitary wave
extended tail and single bore North Sea Wave under fluvial condi-
Project) storm wave tions
spectrum

Trans- Multiple bodies Boulder Pebbles and cobbles Pebbles, cobbles, Cubes of polyeth-
ported (shipping container) (long axis 1.5-3.2 m) (density 1.6 g/cm3 to and cuboids up to ylene
bodies 2.55 g/cm3) 16 cm (10 x 5 x 5 mm3)

Special Real-time locating Downscaled cliff Bathymetry and the IMU sensor;
features system (RTLS) for shore are designed integrating a triaxial
debris tracking following the shape accelerometer, a
of the coastline of a magnetometer, and
prototype site on the a gyroscope, which
Aran Islands (W Ire- is installed inside the
land) clasts and tracks
their movement in
running water

Key Established a func- Storm waves are Even very large Similar influence of Highly sensitive be-
outcomes tion describing the able to mobilize boulders can be the pebble size, havior of the bodies,
maximum inland clifftop boulders transported by storm shape, and pre- which increases with
transport distance waves, even though transport orientation energy of the incom-
and another one de- specific constraints on transport ing wave. If multiple
scribing the spread- regarding the wave, bodies are arranged
ing angle, depending boulder, as well as in a line, the maxi-
on the amount of de- coastal topography mum inundation dis-
bris bodies and bathymetry are tance decreases
required
Random behavior of Advocate caution to Bore velocity, boul-
single debris after associate boulders, der dimension, and
bore Impact which are not depos- wave front slope in
ited in the impact combination with
range of normal wave height are con-
wave activity, strictly sidered main influ-
to tsunamis encing parameters
on boulder transport
Results corroborate
in situ observations
of tsunami-induced
debris transport of
Naito et al. (2014)

5.7 Link to numerical models

Physical laboratory experiments have two main goals: deriving (analytical) descriptions of pro-
cesses and interaction of different parameters, and the development of well elaborated datasets
with quantitative results as basis for numerical models (e.g., Imamura et al., 2008; Nandasena and
Tanaka, 2013). In the latter case, the quantitative results of experiments are used for comparison
with results obtained from a numerical model and for calibration, verification and validation. In this
sense, the parameters of a numerical model are adjusted in order to find the parameter combina-
tion, which generates the numerical results best fitting those of the physical experiments. In nu-
merical models, such parameters may be the (artificial) viscosity of the fluid, the numerical treat-
ment of roughness and turbulences or the computational time step. After the successful calibration,

109
Experimental models of coarse-clast transport by tsunamis

the numerical model can be used for a known event and setting of boulder transport and investi-
gation of parameters which might not be changeable in physical experiments (e.g., roughness). If
the numerical model sufficiently reproduces results for such an event, it is validated, meaning that
the program is able to reproduce events/setups (e.g., experimental results) for which it is not es-
pecially programmed (e.g., real-world bathymetry or initial tsunami/earthquake parameters). An
overview for definitions of calibration, validation and verification can be found in NASA (2008).
Regarding tsunami models, several benchmark tests exist (e.g., Synolakis et al., 2007; NTHMP,
2012), which are used to verify a developed numerical tsunami model for operational use. After-
wards, the model can be applied to unknown or possible future events, by still considering several
restrictions. For boulder transport, no benchmarks similar to those of Synolakis et al. (2007) and
NTHMP (2012) exist so far.

5.8 Conclusions and recommendations

Experimental studies on tsunami-induced boulder transport in particular or focusing on adjacent


research questions (e.g., the transport of shipping containers) give essential insights into tsunami-
induced transport of rigid bodies in nearshore areas and, thus, help to reconstruct past tsunamis
based on the characteristics and spatial distribution of coastal boulders.

The existing experiments also provide a basis for developing and testing numerical models. How-
ever, across the existing publications a certain sensitivity of the boulder transport processes within
high-energy, turbulent flows becomes obvious. Experiments show divergent predominant transport
modes (e.g., rolling and saltation in Imamura et al., 2008 and sliding in Liu et al., 2015) for different
setups and especially the boulder’s FI. The studies of Nandasena and Tanaka (2013) and Imamura
et al. (2008) indicate that boulders with an FI ≈ 1 are more likely transported by rolling compared
to boulders with higher FI, while overall the predominant transport mode seems to be sliding (e.g.,
Freund, 2014; Liu et al., 2015; Nistor et al., 2017). However, the transport mode is highly influenced
by the boulder shape, and since most published experiments use regular-shaped cuboids and
cubes with different FIs, it needs to be investigated further in which amount more irregular shapes
influence the transport mode. Oetjen et al. (2017a, 2018) apply a boulder resembling a natural
prototype and found sliding as the predominant transport mode as well, which is possibly related
to the high FI. The influence on bottom roughness on transport mode, however, is not yet clear and
needs to be addressed in future studies.

In general, authors agree that boulders tend to align the long axis parallel to the flow. Nandasena
and Tanaka (2013) find shorter transport distances for high FI compared to cubic boulders, a hy-
pothesis that is not yet confirmed by others.

For future studies, it would be beneficial if researchers agree on a basic set of parameters regard-
ing their experimental setup, which should be clearly presented in publications (Table 5-6) and
unequivocally defined. Equal terms for the boulder transport distance, for example, should be used
throughout the research community. As in Imamura et al. (2008) the transport distance should at
least be separated regarding the maximum transport distance and total transport distance (Figure
5-9). An additional benefit might be to itemize boulder movements in direction components (e.g.,
in flow direction as well as perpendicular movements) in the result presentation. Regarding the
bore, for example, a standardized point of velocity measurement with a fixed distance from the
shore or boulder should be commonly used.

110
Experimental models of coarse-clast transport by tsunamis

Table 5-6: Recommendation for standard parameters to be published with every experimental study.

Bore Boulder Transport process


Velocity Density Max. transport distance

Height Dimensions Total transport distance


Type of bore generation Shape Mode
(e.g., dam-break or piston system)

(Froude number) Orientation Time-series

Submergence

Flatness index (FI)

Scale (if applicable)

Figure 5-9: Difference of maximum transport distance and total transport distance for an initially partially-submerged
boulder setup (t1). After the maximum flow run-up (t2) the boulder could be transported in backwash direction (t3) before
it finally comes to rest (t4) possibly before the backwash ends on the initial water-level.

The FI requires improvement to a general boulder index that considers shape and density. Due to
the highly random boulder behavior under bore influence, it may even be advantageous to move a
step back and to conduct very basic experiments (idealized shapes of different sizes, uniformly
inclined channel, several wave heights) to better understand the basic processes. Existing experi-
ments show sometimes diverging results leading to uncertain equations. For experiments focusing

111
Experimental models of coarse-clast transport by tsunamis

on boulder behavior, a standard wave bore could be defined facilitating the comparison of results
from different studies with different boulder types. Vice versa, experiments focusing on the influ-
ence of wave parameters could benefit from standardized boulder models.

Focusing on the boulder shape, the step between two shapes should be as small as a divergent
behavior between the shapes can be observed. This would help to find relationships between the
boulder shape and transport behavior and could lead to more reliable analytical equations and a
stronger calibration basis for numerical models. Especially drag, lift, and friction forces are today
considered as relatively widespread coefficients in numerical models and analytical solutions,
which are commonly chosen according to observed and approximated values from literature and
are not based on a fundamental understanding of the ongoing processes.

Even if numerical models provide a good alternative to physical experiments, they still suffer from
physical inaccuracies (Watanabe et al., 2020; Nandasena, 2020). Regarding tsunami-induced
boulder transport, numerical models are at an early stage of development compared to more com-
mon questions in hydrodynamic research, and some crucial and basic mechanisms of nearshore
tsunami hydrodynamics and the exerted forces are not entirely understood. Boulder transport ex-
periments can provide answers to these questions and may help to improve not only specialized
numerical boulder-transport models, but also numerical models focusing on the nearshore tsunami,
its inundation extent, and especially the acting forces on coastal structures.

Acknowledgements

This contribution benefitted from funding by Deutsche Forschungsgemeinschaft (DFG) for the pro-
ject “Modelling tsunami-induced coarse-clast transport – combination of physical experiments, ad-
vanced numerical modelling and field observations” (SCHU 1054/7-1, EN 977/3-1).

112
Significance of boulder shape, shoreline configuration and pre-transport setting for the transport of boulders by tsunamis

6 Significance of boulder shape, shoreline configuration and pre-


transport setting for the transport of boulders by tsunamis

Jan Oetjen, Max Engel, Shiva P. Pudasaini, Holger Schüttrumpf


This is an accepted manuscript of an article published by Wiley in Earth Surface Processes and
Landforms in May 2020. Online available: https://doi.org/10.1002/esp.4870
Keywords: Boulder Transport; Hydraulic Experiment; Statistical Analysis; Boulder Shape; Flatness
Index

Abstract
Research on tsunami-induced coarse-clast transport is a field of rising interest since such deposits
have been identified as useful proxies for extreme-wave events (tsunamis, storm waves) that pro-
vide crucial information for coastal hazard assessment. Physical experiments are, beside in-situ
observations, the foundation of our understanding of how boulders are transported by tsunamis
and provide clues to the development of empirical equations and numerical models describing the
processes and fundamental mechanics. Nevertheless, investigating tsunami-induced boulder
transport is a comparatively young discipline and only a few experimental studies focusing on this
topic have been published so far. To improve the knowledge on nearshore tsunami hydrodynamics,
physical experiments utilizing real-world boulder shapes have been carried out simulating three
different shore types in a wave flume. Crucial insights were gained into boulder transport hydrody-
namics and data resulting from the experiments were analysed in an empirical, statistical, quanti-
tative and qualitative manner. The regular cuboid boulder – one of the specific shapes used in the
experiments – showed the longest transport distances compared to a complex, natural boulder and
a flat cuboid boulder, but also significant fluctuations regarding the total transport distance. The
experiments indicate a strong influence of the shore shape on boulder transport behaviour. Exper-
imental setups of increased mean transport distances also led to a higher spreading of results.
This spreading was further amplified between the idealized-shaped cuboid and the complex-
shaped boulder, which is associated with a lower drag coefficient. Due to the highly sensitive boul-
der reaction to divergent experimental setups, the need to recognize boundary conditions over-
coming commonly considered parameters (e.g., roughness or Flatness Index) in field studies and
numerical models is underlined. Beside the strong influence of initial boulder submergence and
alignment, both the boulder shape and shore type influence the boulder transport pattern, increas-
ing the total transport distance by more than 350 % in some cases.

6.1 Introduction

Subaerial coarse-clast deposits along rocky coasts (mostly boulders to fine blocks sensu Blair and
McPherson, 1999) are valuable indicators for past extreme-wave impacts (storm waves, tsunamis).
While earliest reports on their transport by tsunamis date back to the nineteenth century (Neale,
1885), their systematic use to reconstruct major tsunamis or storms started only in recent decades
(e.g., Bourrouilh-Le Jan and Talandier, 1985; Nakata and Kawana, 1995; Scheffers and Kelletat,
2003; Etienne et al., 2011; Engel et al., 2016). Nott (1997, 2003) was the first to introduce equa-
tions, which describe the initiation of boulder motion and, thus, enable inferences on the physical
characteristics of tsunamis and storm waves. This and other comparable approaches have been
widely used since then (e.g. Noormets et al., 2004; Pignatelli et al., 2009; Paris et al., 2010; Etienne
et al., 2011; Nandasena et al., 2011a; Engel and May, 2012; Sugawara et al., 2014; Watanabe et

113
Significance of boulder shape, shoreline configuration and pre-transport setting for the transport of boulders by tsunamis

al., 2019), even though it is commonly understood that their simplification of the transport process
leads to significant uncertainty (Sugawara et al., 2014; Oetjen et al., 2017b; Bressan et al., 2018;
Nandasena, 2020). These efforts are today complemented by physical experiments on boulder
transport (e.g. Imamura et al., 2008; Bressan et al., 2018) and forward numerical models (e.g.
Zainali and Weiss, 2015) in order to improve our understanding of the transport of boulders by
extreme waves and to use boulders as a proxy for coastal hazards in a more reliable way. However,
the number of studies of boulder transport based on laboratory experiments is surprisingly low and
only a limited spectrum of parameters has been systematically investigated (Oetjen et al., 2020b).
Imamura et al. (2008) studied cubic boulders with a varying Flatness Index (FI = (a + b)/2c;
a = long boulder axis in cm; b = boulder height in cm; c = boulder width in cm) and density in a
10 m-long flume ending in a uniformly inclined shore (1:10). The cubic bodies of low FI tend to roll
or saltate, while the authors also found a significant inverse relationship between FI and transport
distance. Nandasena and Tanaka (2013) conducted experiments in a comparable setup with a
slope of 1:20, boulders with an FI of 1 to 3, different densities, different pre-transport orientations,
and a range of wave heights and velocities. They found that shape and bottom friction significantly
influence the transport mode: while flat boulders and low bottom friction support sliding, cube-like
boulders and higher bottom friction support rolling. After initiation of transport, the main (resp. long)
axis always aligns perpendicular to the flow and, thus, influences the overall transport distance
(Nandasena and Tanaka, 2013). Liu et al. (2015) used an 18 m-long flume with a central inclined
section (3.75:100) and quantitatively investigated the influence of wave velocity and height as well
as the pre-transport setting of the main axis of a cuboid boulder. The authors found sliding to be
the dominant transport mode and a generally shorter distance in the case of rolling transport. Fur-
thermore, transport distance was found to be sensitive to bore height and velocity (Liu et al., 2015).

A range of parameters relevant to boulder transport by waves remains unstudied. In this article,
we investigate the influence of boulder shape on transport mode and transport distance. We use
idealized cuboid boulders, as in previous studies, and compare their performance with subrounded,
naturally shaped clasts. Furthermore, we investigate the pre-transport setting and the influence of
the coastal setup (bathymetry, topography) on transport mode and transport distance by using
three different shore models, a uniformly inclined one, one with altering inclination and a stepped
shore.

6.2 Methods

6.2.1 Boulder models used in the experiments

The complex, naturally shaped boulder model employed in this study is based on a well-studied
cliff-top clast from the island of Bonaire, Leeward Antilles (‘largest boulder’ in Watt et al., 2010;
BOL 2 in Engel and May, 2012) (Figure 6-1). The boulder has main axes of a = 8.7 m, b = 4.8 m
and c = 3.8 m, a density of c. 2.2 g/cm3, a weight of 170 t (previously estimated to be 150 t) and is
located approximately 45 m from the shore at an elevation of c. 6.8 m a.s.l. (above mean sea level).
It was sourced from the cliff edge, as the upper part of the intertidal bio-erosive notch is still visible.
BOL 2 has an irregular, elongated shape, its main axis is parallel to the shoreline and it is located
on a pedestal, indicating that transport occurred hundreds if not thousands of years ago (Engel
and May, 2012). Based on its size, setting and hydraulic parameters inferred from initiation-of-
motion criteria, as well as a comparison with boulder-transport patterns during recent high-category

114
Significance of boulder shape, shoreline configuration and pre-transport setting for the transport of boulders by tsunamis

hurricanes on Bonaire, transport by a tsunami was considered a probable scenario (Watt et al.,
2010; Engel and May, 2012).

For measuring the boulder, structure-from-motion (SfM) was applied (cf. Gienko and Terry, 2014;
Boesl et al., 2019) by using nearly 400 separate overlapping images of a Sony DSC-HX200V cam-
era (constant focal length of 0.5 cm) and Agisoft Photoscan, Meshlab and Autodesk Fusion360
software. The SfM model was completed by creating artificial plain surfaces where images are
either absent, such as at the underside, or where they do not overlap sufficiently, as on the top.
The SfM model has a volume of 77 m3 – compared to 68 m3 derived from a low-resolution DGNSS
(differential global navigation satellite system) point cloud by Engel and May (2012) – resulting in
a weight of 170 t. The SfM model was used to print a positive template of polylactide (PLA) in a
scale of approximately 1:50. The PLA template was then used to create a negative model of sili-
cone rubber (WACKER ELASTOSIL® M 4642) to cast the final boulder model from grouting mortar.

The final 1:50 complex boulder model BOL 2 has (box of outer boundaries; see Figure 6-2) main
axes of a = 17.5 cm, b = 9.6 cm and c = 7.6 cm (FI = 1.58), a volume of 700 cm3 and a weight of
1485 g. Two cuboid boulder models with similar volume and weight were created (Figure 6-2):

• Regular cuboid boulder (approximately 1480 g): a = 14 cm, b = 8 cm, c = 6 cm (FI = 1.83);
this shape is quite common in carbonate reef-top settings.

• Flat cuboid boulder (approximately 1430 g): a = 13.7 cm, b = 15.5 cm, c = 3 cm (FI = 4.9); this
shape is particularly common in beachrock and sandstone-type environments.

All boulder models were created from the same material with a density of approximately 2.2 g/cm3.
Density deviations in a neglectable amount occur due to inhomogeneities during the casting.

Figure 6-1: (A) Simplified geological map of the island of Bonaire showing the elevated ‘Lower Terrace’ unit (B, C) that
forms a quasi-stepped cliff coast and (D) represents the location of the complex-shaped boulder BOL 2 (Engel and
May, 2012), which was considered in this experimental study. Data sources for the map are given in Engel and May
(2012).

115
Significance of boulder shape, shoreline configuration and pre-transport setting for the transport of boulders by tsunamis

Figure 6-2: The three applied boulder models. From left to right: regular flat boulder, regular cuboid boulder and complex
boulder.

6.2.2 Experimental setup

The boulder-transport experiments were conducted in the large tilting flume of the Institute of Hy-
draulic Engineering and Water Resources Management (IWW) of the RWTH University in Aachen
(Germany). The flume has a usable length of 25.5 m and a width of 1 m. The tsunami is modelled
as a broken bore using a combination of specific pumping time, valve position and a low barrier
that initiates wave breaking (Figure 6-3). The pumping time and the valve position are remotely
adjusted for the creation of bores of different heights, lengths and velocities with a high reproduci-
bility (Supporting Information Figure C-1). Two pumps (400 l/s conveying capacity each) extract
the water from an underground reservoir through a perforated metal plate and multiple pipes for
homogenization/abating of the flow. Three meters after the homogenization, the flow passes a
small barrier, which initiates the wave. In front of the barrier, the water level is at between 0.13 m
(standard) and 0.2 m above ground. The wet bed has an extent of approximately 16 m ending at
the shore model, for which we use three different types:

• Type 1: a uniformly inclined slope (approximately 7.5°; Figure 6-4a), a scenario that is used
in all existing experimental studies on boulder transport by tsunamis (Imamura et al., 2008;
Nandasena and Tanaka, 2013; Liu et al., 2015).

• Type 2: a compound slope with two different angles (11° and 4°; Figure 6-4b).

• Type 3: three stepped, horizontal platforms (Figure 6-4c) that are based on the natural setting
of elevated cliff coastlines and abrasive shore platforms. The first two steps are 0.1 m in
height, while the final step is 0.2 m high. The discrepancy in height is due to the aim of keeping
the final height the same as for shore Types 1 and 2 by preserving the possibility for a vertical
shift of the boulder.

All shore types end on a plateau of 1 m length made from polyvinyl chloride (PVC). The experi-
mental programme encompasses multiple setups regarding the submergence (submerged, par-
tially-submerged, subaerial) and alignment (0°, 45°, 90°) of the main boulder axis for each shore
model (Figure 6-4, Table 6-1). The submergence factor is kept constant over the experiments re-
lated to the water column above the boulder (submerged case), the height of the boulder protruding

116
Significance of boulder shape, shoreline configuration and pre-transport setting for the transport of boulders by tsunamis

from the initial water level (partially-submerged) and the distance between water surface and boul-
der (subaerial). Diverting from this, for the stepped shore (Type 3) the water level was varied for
the submerged case (Figure C-10). The actual positions (distance to shore model tip in centime-
tres) differ depending on boulder model and shore type and are shown in detail in the associated
paragraphs.
Table 6-1: Overview of the experimental setup. Boulder shape depicts the type of applied boulder (complex, regular cuboid,
flat cuboid). Alignment describes the initial boulder alignment to the flow: 0°, long axis perpendicular to the flow; 90°, short
axis perpendicular to the flow. Submergence indicates the initial boulder setting (s = submerged: boulder completely under
water; p-s = partially-submerged: part of the boulder is beneath the water surface; s = subaerial: the boulder is completely
above the water surface).

Total
Shore Type 1 setups

Boulder shape Complex Regular cuboid Flat cuboid

Alignment 0° 0° 0° 9

Submergence s p-s sa s p-s sa s p-s sa

Shore Type 2

Boulder Complex Regular cuboid Flat cuboid

Alignment 0° 45° 90° 0° 45° 90° 18

s s s s s s
Submergence p-s p-s p-s p-s p-s p-s
sa sa sa sa sa sa

Shore Type 3

Boulder Complex Regular cuboid Flat cuboid

Alignment 0° 45° 90° 0° 45° 90° 18

s s s s s s
Submergence p-s p-s p-s p-s p-s p-s
sa sa sa sa sa sa

117
Significance of boulder shape, shoreline configuration and pre-transport setting for the transport of boulders by tsunamis

Figure 6-3: Experimental setup: (a) initial state; b) pumping begins. Water is pumped from the reservoir through a per-
forated plate and flow straightener on the flat flume bed. (c) Flow passes the straightener and approaches the installed
barrier, while the hydraulic bore is generated on the wet bed. (d) The bore approaches the shore model and impacts the
boulder.

The wave height was measured by ultrasonic wave gauges at two positions, one directly at the tip
of the shore model and one 3 m further away. Flow velocity was recorded by an impeller velocime-
ter 1.5 m in front of the shore model. The boulder transport was additionally captured by two GoPro
cameras in bird's eye view right above the shore model and at the side of the flume, respectively
(Figure C-2). The friction coefficient was estimated by calculating the necessary force to drag the
boulder across the shore and subsequent division of the necessary drag force by the body-specific
weight force. This leads to varying friction coefficients μ of 0.21 [–] to 0.6 [–] depending on boulder
shape and shore inclination. The friction coefficients for all setups are listed in Table 6-2.

While the original boulder was scaled in a scale of approximately 1:50, all other parameters were
chosen according to the flume dimensions. The exact weight of the original boulder BOL 2 is diffi-
cult to determine due to the unknown heterogeneous density and density distribution of the boulder,
as well as the remaining uncertainties regarding its exact volume. Thus, the boulder replica was
made of a material with a density similar to the assumed bulk density of the original BOL 2
(2.2 g/cm3; Engel and May, 2012). The experiments were conducted using freshwater which con-
sists of a density approximately 3 % below the density of seawater. This difference is neglected in
the experiments since its influence is overprinted by the other scaling effects (e.g. roughness) and
is furthermore constant over all conducted experiments.

118
Significance of boulder shape, shoreline configuration and pre-transport setting for the transport of boulders by tsunamis

Figure 6-4: Top: Applied shore types. (a) Type 1 with a uniform inclination; (b) Type 2 with two different inclinations; (c)
Type 3, stepped shore resembling the shore of the type site of BOL 2 on Bonaire (Figure 6-1). Bottom: initial boulder
setups. (d) Initial alignment to the flow: 90°, long axis perpendicular to the flow; 45°, long axis at a 45° angle to the flow;
0°, short boulder axis perpendicular to the flow. (e) Pre-transport positions in relation to water level.

Table 6-2: Calculated coefficients of friction (μ [-]) for wet conditions depending on shore inclination (3.8°, 7.6°, 11.3°) and
boulder type (flat cuboid, complex, regular cuboid).

Bed Flat cuboid Complex Regular cuboid


0° (horizontal) 0.29 0.21 0.21
3.8° 0.4 0.34 0.38
7.6° 0.45 0.42 0.45
11.3° 0.6 0.54 0.45

6.2.3 Video Processing

Boulder celerity and transport distance after the bore impact were analysed using video recordings.
All shore models contained a 10 cm × 10 cm grid, which allows for a computer-assisted evaluation
of boulder movement. For analysis of the transport, a Matlab® script was established to calculate
position changes between video frames using the underlying grid as reference. The reference dis-
tance was refreshed at characteristic time intervals of the transport: (1) Due to the inclination of
Types 1 and 2 shoreline models and the fish lens of the GoPro camera, the reference distance
varied over the entire transport distance and had to be updated periodically. (2) In case of Type 2,
the reference was updated when the boulder reached the inclination break. (3) When the flow
surrounded or submerged the boulder, the view was disturbed due to the refraction of light. The

119
Significance of boulder shape, shoreline configuration and pre-transport setting for the transport of boulders by tsunamis

boulder movement and bore velocity were eventually calculated by recording the position of a char-
acteristic pixel of the boulder or bore tip over two different frames with respect to the passed time-
interval, which is, in this case, equivalent to the frames-per-second of the video.

6.3 Results

6.3.1 Bore dynamics

The bore hydraulics differed slightly in the three shore setups and were further altered by applying
individual pump and valve preferences. Only one wave was applied over all experiments on shore
Type 1, since the main task was to gain statistically robust results (see wave profile in Figure C-2).
Here, the bore reached a height of c. 0.16 m right in front of the shore and a velocity of v = 0.73 m/s.
From video analysis, the run-up velocity of the bore immediately before the impact on the boulder
is 2 m/s with a height of h = 0.055 cm, resulting in a Froude number of Fr = 2.65 [–].

Two different bores with h = 0.17 cm and h = 0.16 cm were generated in the shore Type 2 setup.
The bore velocity at wave gauge 3 reached 0.87 m/s and increased to 2.0 m/s (Fr = 4.5) during the
run-up. The first wave impact was followed by strong reflections and turbulences in the flume (Fig-
ure 6-5), which resulted from the following pushing wave and which reflect a modified tsunami
profile, if compared with an ideal one as, e.g. in Pedersen and Gjevik (1983). On shore Type 2,
tracking of the boulder using an inertial measurement unit (IMU) was tested. The cavity for the IMU
inside the boulder (radius of 1.5 cm, depth 10 cm) results in a slightly lower volume (approximately
5 % compared to shore Type 1 and 3) and deviating hydrodynamic boulder behaviour.

In the setup of shore Type 3, the bore approached the shore with a velocity of 0.3 m/s, which
increased to 2.2 m/s on the shore at an initial water level of 14 cm. With an initial water level of
0.20 m, the bore approached the shore by approximately 1.93 m/s before the first boulder impact.

Figure 6-5: Wave profile 1.5m in front of the shore and at shore tip on shore Type 2 and 17.2 cm bore height. Idealized
tsunami profile (natural wave form) after Pedersen and Gjevik (1983).

120
Significance of boulder shape, shoreline configuration and pre-transport setting for the transport of boulders by tsunamis

6.3.2 Boulder transport

Experimental designs and numbers of runs (238 runs overall) for the uniformly inclined shore model
(Type 1) are summarized in Supporting Information Table C-1. Ninety-nine runs were conducted
with the complex boulder, 85 runs with the regular cuboid boulder and 54 runs with the flat cuboid
boulder. The experimental runs on the shore Type 2 setup (non-uniformly inclined) encompass 149
runs, while on the stepped shore (Type 3) 230 runs were conducted.

Type 1 – uniformly inclined shore model


In the subaerial pre-transport setting, the boulders were placed on the slope 18 cm from the water
surface and centred in y direction (channel width). For the partially submerged case, the boulders
were placed in a way that half of the object was immersed (Figure C-3). The main axis of the
boulders was aligned perpendicular to the flow.

Boulder transport occurred in two or three steps, respectively. After the first wave impact, a short
time lag occurred before boulders were transported (Table 6-3) at low velocities for a few centime-
tres (Figure 6-6). While the complex and the flat cuboid boulders then stopped for < 0.5 s, the
regular cuboid boulder only slowed down before again experiencing acceleration by the second
wave (Figure 6-7). The regular cuboid boulder with the largest contact surface created a distinct
splash in z direction (vertical).
Table 6-3: Time lag between bore impact and the initiation of boulder transport, averaged over all experimental runs.

Time lag between impact and


Boulder mobilization [ms]
Regular cuboid 70
Complex 90
Flat cuboid 70

Before the deceleration phase of all boulders, the response to the bore impact was similar. How-
ever, this changed after the impact of the second bore, when the flat cuboid boulder stopped im-
mediately after 1.2 to 1.4 s, while the complex boulder reached its final position after about 1.9 s
for the presented run.

The experiments showed that the longer the boulder was transported, the more the individual re-
sults diverged, in particular in the subaerial scenario. In the submerged and partially submerged
scenarios, the results only spread over a small range. For the regular cuboid boulder, more exper-
imental runs were necessary to reach a normalized distribution for transport distance (Figure C-4,
Table C-2). Despite the irregular shape of the distribution groups (Figure C-4), all experimental
setups in Table 6-2 passed the D'Agostino & Pearson omnibus normality test, the Shapiro–Wilk
normality test as well as the KS normality test, ensuring a statistically verified normal distribution
of the results (Table C-2). The 25 % and 75 % percentiles as well as the results for mean, minimum
and maximum transport distances are also presented in Table 6-4.

121
Significance of boulder shape, shoreline configuration and pre-transport setting for the transport of boulders by tsunamis

Table 6-4: Statistical results for the experiments on shore Type 1.

transport distance [cm]


Regular cuboid Complex Flat cuboid
sa ps s sa ps s sa ps s
Minimum 58.44 32.16 8.196 38.19 31.26 3.31 16.47 2.44 0.0

25 % percentile 71.55 49.04 11.68 43.2 34.65 4.63 21.33 4.14 0.0

Median 75.61 54.23 12.03 47.25 36.12 5.39 24.78 5.34 0.0

75 % percentile 83.27 63.18 14.51 50.27 41.61 6.19 26.95 7.92 0.0

Maximum 90.32 76.28 15.22 59.49 49.24 6.64 31.28 9.76 0.0

Mean 76.21 55.67 12.24 47.05 38.12 5.29 24.27 5.9 0.0
Note: sa, subaerial; ps, partially submerged; s, submerged.

The idealized cuboid boulder was accelerated nearly to maximum velocity within the first third of
the transport process (Figure 6-8). After the initial acceleration, it was transported with a fairly con-
stant velocity. The deceleration phase was longer than the acceleration phase. The complex boul-
der had two phases of acceleration and deceleration. While the first acceleration phase was very
short, the second acceleration and deceleration phase described a nearly Gaussian shape until
the boulder reached its final position. The flat cuboid boulder reached its maximum velocity during
the first acceleration phase, after which another lower and slightly broader peak occurred (Figure
6-8).

The experiments showed a strong influence of boulder shape on the total transport distance (Table
6-5, Table 6-6, Figure C-5). Depending on the extent of the initial submergence, the difference in
total transport distance rose up to 900 %, when the regular and flat cuboid boulders were compared
in the partially submerged scenario. The influence of initial submergence was obvious for the flat
cuboid boulder with an increased transport distance of 360 % from partially submerged to
subaerial. Focusing on the shape and the subaerial setup, the transport distance was strongly
reduced between the regular cuboid and complex boulders (−35 %) and also between the complex
and flat cuboid boulder (−47.6 %, Table 6-5). While the gap of the transport distances decreased
between the regular cuboid and the complex boulders (−33.4 %) for the partially-submerged setup,
the gap increased between the complex and flat cuboid boulder (−85.2 %). For the submerged
case, the gap between the regular cuboid and the complex boulder increased again to 55.2 %. The
flat cuboid one showed no significant movement for the submerged case. The only observed
transport mode was sliding.
Table 6-5: Transport distances grouped according to boulder shape and submergence (shore Type 1).

Transport distances
Regular Regular
Flat cuboid Complex cuboid Flat cuboid Complex cuboid
p-s p-s p-s sa sa sa

Average
5.9 38.1 55.7 24.3 47.0 76.2
[cm]
Median [cm] 5.3 36.1 54.2 24.8 47.3 75.6
Note: sa, subaerial; ps, partially submerged; s, submerged.

122
Significance of boulder shape, shoreline configuration and pre-transport setting for the transport of boulders by tsunamis

Table 6-6: Effect of boulder shape and submergence on the total transport distance in percentages (shore Type 1).

Ratio of transport distances


Flat cuboid Complex Regular cuboid Flat cuboid Complex Regular cuboid
p-s p-s p-s sa sa sa
Shape

100 % 677 % 1015 % 100 % 190 % 305 %


flat cuboid complex regular cuboid
100% 433% 660%

Flat cuboid Flat cuboid Complex Complex Regular cuboid Regular cuboid
Submergence

p-s sa p-s sa p-s sa


100% 464% 100% 131% 100% 139%

ps sa
100% 140%
Note: sa, subaerial; ps, partially submerged; s, submerged.

123
Significance of boulder shape, shoreline configuration and pre-transport setting for the transport of boulders by tsunamis

Figure 6-6: First 0.33 s of boulder movement in subaerial conditions (shore Type 1). The boulder velocity vB is averaged
between two presented timesteps and rounded to steps of 0.1 m/s since smaller differences are superimposed by un-
certainties due to light refraction.

124
Significance of boulder shape, shoreline configuration and pre-transport setting for the transport of boulders by tsunamis

Figure 6-7: Boulder movement between 0.5 s and 1.9 s after wave impact in subaerial conditions (shore Type 1). The
boulder velocity vB is averaged between two presented timesteps and rounded to steps of 0.1 m/s, since smaller differ-
ences are superimposed by uncertainties due to light refraction.

125
Significance of boulder shape, shoreline configuration and pre-transport setting for the transport of boulders by tsunamis

Figure 6-8: Transport over time for each tested boulder shape. The qualitative fitting curves are derived from the average
displacements for multiple experimental runs (shore Type 1).

Type 2 – non-uniform inclined shore model


For the subaerial experiments the boulders were placed 110 cm from the shore tip with their long
axis (a-axis) perpendicular to the flow. When the bore reached the boulders, water was reflected
in the y and z directions (0.06–0.16 s in Figure C-6). After the initial impact between 0.0 and 0.06 s,
the regular cuboid boulder started to move slowly with a velocity < 0.1 m/s and slightly rotated
around the a-axis during transport (0.56–0.89 s). Approximately 0.89 s after the bore impact, the
boulder rotated back into its original orientation. Between 0.09 s and 0.29 s, it moved with an av-
erage velocity of 0.57 m/s, compared to 0.52 m/s of the complex boulder. The discrepancy between
the velocity of the regular cuboid (0.93 m/s) and complex (0.42 m/s) boulders increased to approx-
imately 0.5 m/s before the complex boulder came to rest (Figure 6-9a, Figure C-6).

Figure 6-9: (a) Time-distance graph for the subaerial case at shore Type 2. (b) Time-distance graph for the partially-
submerged case at shore Type 2. In both diagrams, the point of the changing shore inclination is marked by a dotted
red line.

In the partially-submerged case, the boulders were placed 60 cm from the shore tip. The water
reflection around the boulders and the initial splash were not as distinct as for the subaerial case
(0.03–0.09 s in Figure C-7). The complex boulder moved immediately after the bore impact, but
only for a short distance. The regular cuboid boulder experienced a time gap between bore impact

126
Significance of boulder shape, shoreline configuration and pre-transport setting for the transport of boulders by tsunamis

and the initiation of transport. After 0.22 s from the impact, the complex boulder was in motion and
had passed the regular cuboid boulder by c. 10 cm (shown as red dotted circle in Figure 6-9a).
After 0.90 s, the complex boulder was overtaken by the regular cuboid boulder, which again rotated
slightly after bore impact (0.36 s). After 0.63 s, the regular cuboid boulder realigned the a-axis
perpendicular to the flow leading to acceleration and a higher velocity. At the later stages, the
distance between both boulders increased constantly, even though the regular cuboid boulder kept
rotating its main axis horizontally for a few degrees to both sides. The complex boulder did not
perform any significant rotation. While the regular cuboid boulder first turned anti-clockwise fol-
lowed by a strong clockwise turn, the complex one shifted its main axis only slightly in both direc-
tions. The complex boulder came to rest after 3.30 s, while the regular cuboid boulder moved for
3.73 s, span for 0.27 s on the spot, and stopped eventually after 4.00 s (Figure 6-9b).

The results on shore Type 2 showed alignment and submergence as dominant parameters influ-
encing transport distance. Furthermore, with longer transport distances, the discrepancy between
the complex and regular cuboid boulders increased until the idealized regular cuboid boulder was
shifted out of the experimental setup in nearly 50 % of the subaerial cases for the 45° and 90°
alignment (marked in Figure 6-10 and Figure 6-11). This occurred only once for the complex boul-
der (subaerial; 90° initial alignment). The increase can be explained by the lower energy transfer
on the complex boulder during the bore impact. This effect is more obvious on longer transport
distances since the available energy is directly dependent on the product of impact force and boul-
der impact area resulting in a non-linear relationship between bore energy, boulder shape (or drag
coefficient) and transport distance. The transport distance generally increased from the 0° to 90°
alignment and from the submerged to subaerial setup. In all experiments except for the 0°,
subaerial and partially-submerged cases, the 95 % percentile of the regular cuboid boulder
transport distance exceeds the distance of the complex one. Figure C-8 shows the total transport
distances on shore Type 2 grouped according to initial boulder alignment, submergence and boul-
der type (complex, regular cuboid).

Figure 6-10 and Figure 6-11 show that pre-transport submergence was a more important control
for transport distance than the impact angle of the wave. Between the submerged and the subaerial
case, the transport distance increased by 2200 % (excluding transportation beyond the step); be-
tween the submerged and partially-submerged scenarios, the distance increased by 1300 %. Fo-
cusing on the impact angles 0° and 90°, the distance increased by 190 %, and by 150 % between
0° and 45° (excluding transportation out of the step). For impact angles of 0° and 45°, the boulders
started to rotate around the z-axis (vertical axis) after the bore impact. After some back and forth
rotation, the main axis aligned perpendicular to the flow.

In general, experimental runs on shore Type 2 showed a longer average transport distance for the
regular cuboid boulder of approximately 90 % (averaged over all submerged and partially-sub-
merged setups) compared to the complex boulder.

Additional experiments showed a correlation between the initial water level in interaction with the
pre-transport setting on the transport distance. For the submerged case, the transport distance
increased when the boulder was placed in shallower water regions. The same was valid for the
subaerial case: the closer the boulder was situated to the water's edge, the longer was the transport
distance.

127
Significance of boulder shape, shoreline configuration and pre-transport setting for the transport of boulders by tsunamis

Figure 6-10: Comparison of the total transport distances grouped following the boulder shape (complex and regular
cuboid), initial boulder alignment and submergence on shore Type 2.

Figure 6-11: Comparison of the total transport distances grouped following the boulder shape (complex and regular
cuboid), initial boulder alignment and submergence on shore Type 3.

Type 3 – stepped shore platforms


The initial water level and wave parameters were altered over the experiments (Figures C-9, C-
10). The variation of the initial water level became necessary since the submergence scenario
could not be realized by simply changing the initial boulder position due to the stepped shore setup.
Therefore, the pre-transport settings were realized using different initial water levels of 14 cm and
20 cm and two different steps of the model (Figure C-11). Corresponding to the divergent initial
water levels, the bore velocity also differed on this shore model.

In accordance with shore Types 1 and 2, the longest transport distances were measured for the
subaerial setups (Figure C-11). However, due to the experimental design, the maximum transport
distance was limited to approximately 125 cm. The maximum transport distance for certain

128
Significance of boulder shape, shoreline configuration and pre-transport setting for the transport of boulders by tsunamis

subaerial setups was therefore not measurable and set to an artificial maximum of 150 cm in Figure
11. From the distribution of transport distances of individual runs it becomes clear that the discrep-
ancies with respect to the boulder shapes and the initial alignments are not as significant as in the
other setups (Figure 6-11). The influence of the impact angle decreased substantially compared to
shore Type 2 (Figure 6-10 and Figure 6-11), while submergence considerably affected the
transport. The influence of the impact angle on the transport distance was almost negligible on
shore Type 3 (shore Type 2: up to > 190 % longer). Again, sliding was the only observed transport
mode within the standardized experimental series and all boulders aligned their long axis perpen-
dicular to the flow during the transport. During the experiments no boulder transport onto the next
step was observed. However, vertical boulder transport would have been possibly caused by bores
of significantly higher velocity and therefore height, which, however, were not part of the standard-
ized experimental series analysed here. Particularly, in the partially-submerged cases, boulder
movement was hampered due to wave reflection at the next vertical step (Figure 6-12).

Figure 6-12: Movement-hampering influence of the stepped shore with partially-submerged setup (conceptual drawing).
t1: approaching bore, no boulder movement. t2: bore reaches and overflows the boulder. No boulder movement. t 3: bore
reaches next step and boulder is shifted along the flow direction. t 4: partial flow and wave reversing due to a hydraulic
jump hamper the boulder movement.

6.4 Discussion

6.4.1 Parameter influence hierarchy

Identifying the parameter from the set of investigated parameters (boulder shape, shore type, sub-
mergence, alignment) with the strongest influence on boulder transport is not straightforward,
mostly due to the synergy of the different parameters and the empirical nature of the study.

However, the experiments underline that the boulder shape (and its ground contact surface) needs
to be considered as a parameter at least as important as the initial submergence for recalculating
the necessary transport energy. The experiments show that the highest transport distances can be
expected for configurations in which the wave impacts a subaerial boulder on its maximum possible
contact surface. In contrast, a submerged boulder of comparable weight and volume but with only
a small contact surface will be transported significantly shorter. The results indicate that a rounded
boulder might be transported approximately a distance 30 % shorter than an idealized cuboid boul-
der of comparable dimensions. However, in nature this effect of the boulder shape is possibly su-
perimposed by other parameters, especially the initial submergence which showed an influence
resulting in an increased transport distance of approximately 40 % between partially-submerged

129
Significance of boulder shape, shoreline configuration and pre-transport setting for the transport of boulders by tsunamis

and subaerial settings (shore Type 1). In particular the parameters submergence and alignment
might amplify or attenuate each other leading to ranges of transport distances fitting to several
combinations of initial submergence and alignment. Nevertheless, the influence of the boulder
shape (beyond the Flatness Index), and also the aspect of roundness, is significant in all investi-
gated setups and should therefore be considered when calculating boulder transport due to tsuna-
mis or storm waves.

In the following paragraphs, first the general experimental sensitivity is evaluated followed by an
analysis of initial alignment, shore type and boulder shape.

6.4.2 Experimental sensitivity

The results of the statistical analysis for shore Type 1 (uniformly inclined) showed clear trends,
even though the number of experimental runs necessary to gain a normal distribution of transport
distances differed significantly depending on boulder shape and initial boulder position (sub-
merged, partially submerged, subaerial). In particular, the scatter of the transport distances, and
therefore the number of necessary runs, increased with the average transport distance, i.e. from
submerged to subaerial and from the regular flat boulder model to the complex shaped and the
regular cuboid boulder (Figure 6-10, Figure 6-11).

For several identical experimental setups of the same initial boulder position/alignment, initial water
level, shore type and bore parameter, the transport distance varied by up to 650 % (shore Type 2,
complex boulder, submerged, 45°). In contrast, the transport distances within the 95 % percentile
of the regular cuboid boulder on shore Type 2 (subaerial, 0°) are significantly shorter than for the
complex boulder, which is not in accordance to the overall trend. These deviations must be ex-
plained by very small, unavoidable differences between individual experimental runs, e.g. due to
the highly turbulent flow, vibrations generated by the pumps and slight deviations (millimeter-scale)
while placing the boulder manually in the model, especially regarding the alignment. However, the
ability to create identical bores for two runs was already a significant improvement compared to
experiments with dam-break type bore generation, especially if the dam break is initiated by man-
ually opening the gate, and leads to sensitivity reduction. In Imamura et al. (2008), for example,
high fluctuations of the transport distances have also been observed for repeated experiments (up
to approximately 250 %), although they occurred mainly for intermediate bore parameters and not
for extrema. Despite the type of wave generation, the most obvious difference between the studies
is the mean boulder size, which is significant smaller in Imamura et al. (2008). Thus, the larger
gradient between the available impact energy (wave) and resisting forces (boulder) seems to su-
perimpose fluctuations of the transport distances, leading to fewer fluctuations in the results.

6.4.3 Influence of shore type

It is remarkable that total transport distances of the three different boulder shapes differed much
more in the inclined shore setups compared to the stepped one. Disregarding the subaerial cases
due to the limitations on the stepped setup, the regular cuboid boulder on average reached 95 %
longer transport distances on shore Type 2 and 90 % longer transport distances on shore Type 1,
compared to the complex boulder. This difference disappears almost entirely in the stepped setup
(shore Type 3). Furthermore, also the difference between the partially-submerged and subaerial
setups as well as the difference between the initial alignments (0°, 45°, 90°) of the boulder is much
smaller in the stepped shoreline scenario. The apparent influence of the shoreline morphology,

130
Significance of boulder shape, shoreline configuration and pre-transport setting for the transport of boulders by tsunamis

however, has not been identified to be linear or easily factorizable since the number of combina-
tions of the other parameters (boulder shape, submergence, alignment, limited step length) has a
major influence on the transport processes (e.g. Goto et al., 2010b; Switzer and Burston, 2010;
Nandasena and Tanaka, 2013). Further investigations with smaller step sizes and increased bot-
tom roughness should be conducted in order to investigate the vertical transport of boulders.

Another observation from the partially-submerged cases on the stepped shore is that enough en-
ergy seems to be available for moving the boulder across the whole shore step but that it is slowed
down due to flow reflection at the following shore step. Thus, the wave velocity is stopped and
partly accelerated in the opposite direction. Since the boulder is not mobilized with the first wave
impact and moves slower than the bore, its movement is reduced by the reflected wave resulting
in a reduced transport distance (Figure 6-12). This hydraulic jump effect seems to superimpose
the influence of boulder shape and impact angle and should be considered in field studies at com-
parable shoreline settings since reflections of the bore might occur even if the main part of the flow
is still moving inland (e.g., Goto et al., 2010b).

6.4.4 Initial boulder alignment to the flow

Liu et al. (2015) stated that a 45° pre-transport alignment leads to higher transport distances than
0° and 90°. In the present study, however, focusing on the complex boulder, transport distances of
the 45° scenario were between those of 0° and 90°, which is in line with observations made by
Nandasena and Tanaka (2013) for idealized shapes. In accordance with their findings, the
transport distance in the 45° scenario exceeded the one in the 90° case for submerged boulders,
but only by a nearly negligible amount in our findings. This indicates that the alignment only has a
minor influence for submerged cases. For partially submerged and subaerial boulders we found
the largest transport distances in the 90° scenario. This might be caused by the higher scale ratio
(lower dimensions) applied in Liu et al. (2015). Nandasena and Tanaka (2013) found a reduced
transport distance of 27 % between an initial alignment of 0° and 90°. The present study confirms
this observation as well as the fact that a boulder which was initially aligned with the long axis
parallel to the flow tends to align with the long axis perpendicular to the flow after impact, as already
indicated by Imamura et al. (2008). However, we found that the influence of the initial boulder
alignment to be highly dependent on the shoreline configuration and observed a higher influence
for an inclined shore and a much lower influence for a stepped shore.

We found a difference of the overall transport distance of approximately 5 % on the horizontal


stepped shore model. This difference rises up to nearly 670 % on shore Type 2 for the partially
submerged case. One reason for the deviation between the observations of Nandasena and
Tanaka (2013) and our experiments could be the significantly larger impact area of the boulders
applied in the present study. The largest applied boulder of Nandasena and Tanaka (2013) had an
impact area of 32 cm2 for the long boulder axis whereas the regular cuboid boulder of the present
experiments has an impact area of 84 cm2. Their boulder had an area of 16 cm2 (48 cm2 in our
case) when the short axis was aligned with the flow. Hence, the ratio between the long and short
axis is two in the case of Nandasena and Tanaka (2013) and 1.75 in the case presented here. The
influence of the ratio between the long and the short impact area is possibly a transport-influencing
parameter not investigated in detail so far and should be considered in future studies in addition to
the Flatness Index (sensu Nandasena and Tanaka, 2013).

131
Significance of boulder shape, shoreline configuration and pre-transport setting for the transport of boulders by tsunamis

6.4.5 Boulder shape

Focusing on the boulder shape, the total transport distance was 47 % longer for the regular cuboid
boulder than for the complex shaped boulder (averaged over all partially-submerged setups). Ex-
periments on the uniformly inclined shore Type 1 showed an additional transport distance of
+177 % for the complex shaped boulder compared with the flat cuboid boulder and a difference of
+231 % between the flat cuboid and the regular cuboid boulder. The results clearly showed the
importance of boulder shape in tsunami-induced boulder transport and that this parameter needs
to be considered beyond the Flatness Index, which is similar for the regular cuboid and the complex
boulder deployed here. The shape determines the area available for wave impact in flow direction
and defined by the a and c axis as well as the roundness of the boulder. It determines the drag
force acting on the boulder, which, in combination with friction, is considered to be the dominant
force controlling boulder transport by a tsunami (Goto et al., 2010b; Nandasena et al., 2011a).
Furthermore, the drag coefficient and, thus, the energy transferred from the bore probably in-
creases first from the complex to the regular cuboid, and further to the flat cuboid boulder. From
the experiments it seems that the effect of the (presumably) higher drag coefficient of the flat cuboid
boulder is superimposed by the higher ground-contact area compared to the other shapes. Based
on our experiments, this is the most reasonable explanation for the shorter transport distance of
the flat cuboid boulder, but further investigations of the dependencies between the coefficients of
drag and friction on transport distance are needed.

However, since we conducted experiments with three boulder models of relatively large variations
in shape, the datasets are not yet sufficient to develop meaningful conclusions and (analytical or
numerical) models considering the shape overcoming idealized (cuboid) shapes. Considering the
sensitive boulder behaviour in the flow, it might be an option to conduct experimental series focus-
ing on several boulder models with smoothly changing shapes in the future, e.g. from egg-like to
cuboid by holding all other parameter constant and repeating the experiments for each boulder
shape until a statistically robust result is reached.

6.4.6 General observations

The comparison of the flow fields around the complex and regular cuboid boulders shows how the
flow responds to boulder shape (Figure 6-13, Figure C-6). The highly turbulent flow field around
the regular cuboid boulder occurring immediately after the impact shows a clear disruption of the
flow in front of and behind the boulder. This provides evidence that a high amount of energy is
transferred from the bore to the boulder. For the complex boulder, turbulences are less distinct and
more water is diverted around the boulder, which indicates lower energy transfer.

132
Significance of boulder shape, shoreline configuration and pre-transport setting for the transport of boulders by tsunamis

Figure 6-13: Comparison between the simplified horizontal flow fields of (a) Nandasena and Tanaka (2013), (b) the
regular cuboid boulder, (c) the complex boulder. For the regular cuboid and complex boulder time step two and three
are shown as photographs, depicting the flow field around the boulder in higher detail. We found a ‘trumpet’ like flow
field behind the boulder where Nandasena and Tanaka (2013) published a more straight-lined field. Time interval be-
tween states is 0.1 s (series a and design from Nandasena and Tanaka, 2013, modified).

Nandasena and Tanaka (2013) analysed the flow field around the boulder within the first 0.3 s after
bore impact. Interestingly, the authors describe the flow after the impact as opening towards the
flow direction, while in the present study the flow field was found to already close after 0.1 s in the
case of the regular cuboid boulder model (Figure 6-13). After 0.2 s, Nandasena and Tanaka (2013)
found the stream closing in flow direction where we identified the beginning of the closing process
a length from the boulder away both for the regular cuboid and the complex shaped boulder. A
reason for the different observations might be the higher maximum Froude number in our experi-
ments resulting in higher flow velocities on the shore creating a larger and longer stable area almost
free of flow perpendicular to the main flow direction. From the video snapshots in Figure 6-13 it
can be seen that the flow field closes significantly faster in the case of the complex boulder com-
pared to the regular cuboid. This observation underlines the lower transfer of energy resulting in
shorter transport distances originating from the streamline shape (lower drag coefficient) and boul-
der parts without ground contact.

We found several agreements with the experimental results in Liu et al. (2015). Liu et al. (2015)
state that the boulder reached a velocity of approximately 50 % of that of the bore in front of the
shore which is in accordance with our results. Furthermore, they found three basic phases in the

133
Significance of boulder shape, shoreline configuration and pre-transport setting for the transport of boulders by tsunamis

boulder transport process: acceleration, steady and deceleration phase. We found an additional
second acceleration phase for the inclined shore setups, but are able to confirm the three-phase
transport process on the horizontally stepped shore setup.

As in Liu et al. (2015), we also found sliding as the general transportation mode for tsunami induced
boulder transport. In the present study, overturning of the boulder occurred only in three of over
600 experimental runs, and even in those cases was the boulder overturned only once and then
transported further by sliding. However, the transport mode depends significantly on roughness-
and bathymetry-related parameters and shape (flatness, sphericity) and it has a strong influence
on the transport distance (Imamura et al., 2008; Nandasena and Tanaka, 2013). Since rolling and
saltating are quite common field observations, in particular in very rough, karstified limestone set-
tings (Etienne et al., 2011; Engel and May, 2012; May et al., 2015), further research is required as
those transport modes are assumed to increase the transport distance due to reduced friction and
centrifugal forces (e.g. Imamura et al., 2008).

Nandasena and Tanaka (2013) also conducted statistical analyses of the dependencies between
several boulder parameters (boulder main axis, Flatness Index, block weight) and the transport
distance. They found the highest determination coefficient for the influence of the weight of the
boulder (R2 = 0.7115), followed by the length of its main axis (approximately R2 = 0.35). Both pa-
rameters were not considered in our study. However, Nandasena and Tanaka (2013) found the
weakest correlation for the FI with R2 = 0.05. Even though we only applied three different boulder
shapes (compared to 15 in Nandasena and Tanaka, 2013), the statistical analyses of the experi-
ments on the uniformly inclined shore (shore Type 1) indicate that the FI needs to be modified in
order to consider shapes other than idealized, cuboid shapes.

6.5 Future directions and numerical investigations

Data and observations from this study will be used for validating a new numerical approach based
on the two-phase mass flow model of Pudasaini (2012). Here, a numerical two-phase mass flow
model is enhanced for tsunami-induced boulder transport by implementing the boulder as a floating
immersed boundary (immersed boundary method, e.g. Peskin, 2002). Once the numerical model
is successfully validated for boulder transport induced by single-phase waves, further physical ex-
periments and numerical investigations will be conducted enhancing the bore as a solid-fluid mix-
ture. The influence of the sedimentary load within the wave is already under discussion but has not
been investigated in detail so far (e.g. Kain et al., 2012). The model of Pudasaini (2012) is already
widely accepted for modelling two-phase mass flows as well as mass flow-induced tsunamis (e.g.
Mergili et al., 2017; Mergili et al., 2018; Drenkhan et al., 2019; Pudasaini and Mergili, 2019), and
therefore provides an excellent opportunity for investigating the influence of a second phase on
tsunami-induced boulder transport.

The validation of other numerical models would benefit from further improvements regarding the
experimental design. Obtaining information of the flow in greater detail could be achieved by em-
ploying boulders of contrasting colour (cf. Cox et al., 2019) and to use e.g. particle-image-veloci-
metry (PIV) techniques for tracking the streamlines. The highly turbulent flow around the boulder
incorporating air in the flow might provide enough information for such approaches. Additional in-
formation could be gained from front or rear-view video recordings which would enable analysis of
the flow behaviour on every side of the boulder. In our experiments, several factors need to be

134
Significance of boulder shape, shoreline configuration and pre-transport setting for the transport of boulders by tsunamis

considered regarding the lack of vertical transport of the boulders in the stepped shoreline model
(Type 3): Firstly, the wave length may have been too short, resulting in an insufficiently long acting
lift force. In contrast, the wave might have been too steep and not able to generate sufficient lift
underneath the boulder before following water hampered a potential lift from the platform. Addi-
tionally, the smooth ground surface of the model supported a sliding motion of the boulders, which
also impeded a possible lift since the energy of the wave is hindered to generate a sufficient lift
force underneath the boulder (e.g., due to tilting).

The experiments showed the strong influence of the boulder shape on the transport process.
Therefore, we recommend to apply a dimensionless factor summarizing not only the ratio of the
three dimensions of an idealized boulder model but furthermore accounting for the ‘energetic'
shape of the boulder which is related to the impact area. This could be done by introducing a
coefficient accounting for both the flatness and the drag coefficient of the boulder, as well as its
roundness, which can be systematically quantified in the field (Cox et al., 2018b). The empirical
derivation of such a coefficient would need a number of experiments based on smooth shape var-
iations from angular-cubic to smooth-spherical by holding the same wave and shore parameter, for
example. Therefore, investigating further boulder shapes with only slightly altered forms could help
to obtain a better understanding of the processes during bore impact and to derive meaningful
numerical and analytical models.

6.6 Conclusions

Boulder transport by high-energy waves is a sensitive process and investigations based on physi-
cal experiments require a well-controlled environment. The presented experiments clearly show
the influence of boulder shape and initial submergence of the boulder on the transport distance.
With decreasing available impact area on the boulder, which is also related to roundness, the
transport distance decreases as well. However, the influence of the shape almost vanishes on the
investigated stepped shore, indicating that the effect of the boulder shape decreases with decreas-
ing shore inclination. Future in-field, numerical and experimental investigations should therefore
also systematically target different shore types and consider their influence on boulder transport.

Investigations of the flow-field around the boulder indicate that the flow disruption is significantly
larger for the regular and idealized cuboid boulder than for the complex one. This observation
underlines that less energy is transferred from the bore to the streamlined, complex boulder result-
ing in shorter transport distances compared to the cuboid boulder. The divergence between the
complex and idealized boulders increases further for longer transport distances due to the non-
linear correlation between boulder drag coefficient, impact force, friction and transport distance.

As a next step, an idealized (cuboid) shape could be incrementally adjusted to a natural shape
(e.g. like the BOL 2 from Bonaire considered here) with several stages in between to elaborate
coefficients exceeding the typical range of applied drag coefficients and flatness numbers.

For validation and calibration issues on numerical boulder transport models, the gained data pro-
vide a useful database even if the deviation of results needs to be handled carefully. For future
investigations regarding the interaction of particle-laden flows and boulders, the extent of experi-
mental setups will have to be reduced (comparable to presented shore Type 1) in order to gain
robust baseline data overcoming ambiguities in statistical significance.

135
Significance of boulder shape, shoreline configuration and pre-transport setting for the transport of boulders by tsunamis

Acknowledgements

This contribution received funding by the German Research Foundation (Deutsche Forschungsge-
meinschaft, DFG) for the project ‘Modelling tsunami-induced coarse-clast transport – combination
of physical experiments, advanced numerical modelling and field observations' (SCHU 1054/7-1,
EN 977/3-1).

Shiva P. Pudasaini gratefully acknowledges the financial support provided by the German Re-
search Foundation (DFG) through the research project, PU 386/5-1: ‘A novel and unified solution
to multi-phase mass flows'.

136
Numerical Boulder Transport Model

7 Numerical Boulder Transport Model

This chapter describes the current state of the developed boulder transport model. All codes in
connection to the immersed boundary/boulder are own developments. All figures representing sim-
ulation results are made with own-developed Python and MATLAB programs.

While the basic Immersed Boundary Method goes back to Peskin (1972), the basic approach for
using forcing nodes in the here presented approach is adopted in a significantly adjusted manner
from Liao et al. (2010), and references therein. The geometric algorithm used to detect the IBM is
based on own developments of Jan Oetjen, although similar approaches exist (e.g., Hylla, 2013).
Shiva P. Pudasaini contributed the original two-phase model, the numerical slope, and discussions
throughout the development. The idea of applying a smoothing factor to the force goes back to
personal discussions with Julia Kowalski.

7.1 Introduction

Coarse and fine coastal tsunami deposits are extensively utilized as proxies for high-energy wave
events like tsunamis (e.g., Bourrouilh-Le Jan and Talandier, 1985; Nakata and Kawana, 1995;
Scheffers and Kelletat, 2003; Etienne et al., 2011; Engel et al., 2016). In recent years, many re-
searchers conducted laboratory (e.g., Imamura et al., 2008; Nandasena and Tanaka, 2013; Bres-
san et al., 2018; Oetjen et al., 2020b) and numerical studies (e.g., Imamura et al., 2008; Nandasena
and Tanaka, 2013; Zainali and Weiss, 2015) for investigating the main mobilization influencing
parameters (e.g., boulder submergence, boulder alignment to impact) and reconstructing boulder
transports of past events in order to get a better understanding of nearshore tsunami hydrodynam-
ics (e.g., Scheffers and Kelletat, 2006, Nandasena et al., 2011a; Engel and May, 2012; Weiss and
Bourgeois, 2012; Switzer et al., 2014).

Numerical boulder and sediment transport models are divided in forward and inverse models. While
inverse models are used to assess transport causing tsunami properties by recalculating them from
the properties of boulder deposits (e.g., boulder dimensions and density), forward models calculate
the tsunami hydrodynamics and the subsequent boulder transport (Watanabe et al., 2020). The
basis for inversed models is the Morison equation (O'Brien and Morison, 1952) which accounts for
the stability of objects regarding hydrodynamic stresses and the equilibrium between drag, inertia,
and friction force (Watanabe et al., 2020). Examples for inverse models can be found in Nott (1997,
2003), Noormets et al. (2004), Pignatelli et al. (2009) or Nandasena et al. (2011a). Inverse models
are broadly applied by researchers in the field (e.g., Kennedy et al., 2007; Spiske et al., 2008; Paris
et al., 2010; Engel and May, 2012). However, as reported in Watanabe et al. (2020), for example,
inverse boulder transport models suffer from several shortcomings: 1) Coefficients for drag, lift,
friction, etc. must be accurately evaluated or estimated for the application. 2) Boulder transport
depends highly on the Froude number which is regularly taken as a constant in inverse models. 3)
The choice of a suitable inverse model (equation) depends on the boulder’s pre-transport location,
especially for joint bound cases (Watanabe et al., 2020). The exact determination of the pre-
transport setup is often a task which cannot be clarified completely.

The advantage of forward boulder transport models originates from their capability of modelling
both, the tsunami hydrodynamic and the thereby induced boulder transport. By fitting the properties
of the computed tsunami to observed boulder transportations, the actual tsunami properties can

137
Numerical Boulder Transport Model

be derived (Watanabe et al., 2020). Forward models are developed by Noji et al. (1993), Imamura
et al. (2008), and Nandasena and Tanaka (2013), for example. Zainali and Weiss (2015) used
Smoothed Particle Hydrodynamics (SPH) for simulating tsunami induced boulder transport in a
fully three-dimensional environment. As inverse models, most forward models suffer form an ap-
propriate estimation of a range of coefficients (e.g., sliding, lift, friction) and the identification of the
pre-transport location of boulders (Nandasena and Tanaka, 2013).

Both model types suffer from uncertainties arising from additional contributing processes (e.g.,
boulder-boulder interaction), not-incorporable features (e.g., micro-topography), and over-simpli-
fied conditions (e.g., topography, boulder shape, density variations in the flow; see also Nandasena
and Tanaka, 2013). Here, we propose to tackle the boulder transport phenomenon by applying a
two-phase model (Pudasaini, 2012) and handling the Boulder as an immersed boundary (e.g.,
Peskin, 1972). From this approach, it becomes possible to simulate the behaviour of boulders of
different shapes, apart from oversimplified cuboids or spheres and to account for, probably,
transport influencing density variations in the tsunami (Watanabe et al., 2020). A particular goal is
investigating the influence of sedimentary load in the tsunami wave as assumed by Kain et al.
(2012). In this context, it is assumed that the increased wave density due to sediments supports
mobilization and transport distance and has, therefore, to be considered in back calculations of
coarse tsunami deposits. After presenting the algorithms and equations on which the presented
model is based on, three simulations are shown for investigating the influence of boulder shape/ori-
entation (Chapter 7.6.2), influence of sedimentary load in the wave (Chapter 7.6.3), and proving
the algorithm for the calculation of rotation (Chapter 7.6.4). The symbols used in this chapter are
listed in Table 7-1.

138
Numerical Boulder Transport Model

Table 7-1: Used symbols in Chapter 7.

Symbol Specification Unit Symbol Specification Unit


𝑟𝑎𝑑
𝛼⃗ Angular acceleration [ ] 𝑛𝑐𝑢𝑡𝑠 Number of IBM cuts [−]
𝑠²
Impact angle of the resulting Translational IBM/boulder 𝑚
𝛼𝐹 force
[𝑟𝑎𝑑] 𝑢𝐵 velocity
[ ]
𝑠
Portion of solid phase to fluid Combined velocity of solid 𝑚
𝛼𝑠 phase
[−] 𝑢𝐹 and fluid flow phases
[ ]
𝑠
Velocity of solid flow 𝑚
𝛼𝑥 , 𝛼𝑦 Numerical slopes [𝑟𝑎𝑑] 𝑢𝑠 [ ]
phase 𝑠
𝑚
𝛽 Primaty force acting factor [−] 𝑢𝑓 Velocity of fluid flow phase [ ]
𝑠
Resulting combined flow 𝑚
𝜑 Angle of velocity [𝑟𝑎𝑑] 𝑢𝑅𝑛,𝑛+1 velocity on a boundary [ ]
𝑠
segment
𝑘𝑔
𝜌𝐵 IBM/boulder density [ ] 𝑥 𝑖,𝑗 Grid cell index [−]
𝑚³
𝑘𝑔 𝑖,𝑗
𝜌𝑓 Density fluid phase [ ] 𝑥𝑥 x position of cell xi,j [𝑚]
𝑚³
𝑘𝑔 𝑖,𝑗
𝜌𝐹 Combined flow density [ ] 𝑥𝑦 x position of cell xi,j [𝑚]
𝑚³
𝑘𝑔 𝑖,𝑗
𝜌𝑠 Density solid phase [ ] 𝑥ℎ Height value of cell xi,j [𝑚]
𝑚³
𝑟𝑎𝑑
𝜔
⃗⃗ Angular velocity [ ]
𝑠

𝜔𝑥𝑛 IBM segment angle [𝑟𝑎𝑑] AnB Area of an IBM segment [𝑚²]

Ω IBM area [−] 𝐴𝐶 Geometrical IBM center [−]

𝜃 Force direction factor [−] 𝐴𝑇 Triangle area [𝑚²]

Geometrical center of
𝜁 Secondary force acting factor [−] 𝐵𝐶 [−]
mass

𝐶𝑓 Coefficient of friction [-]

𝑚
𝑎𝐵 IBM/boulder acceleration [ ] 𝐹𝑑,𝑡𝑟 Translational drag force [𝑁]
𝑠²
𝑛 Langrangian boundary coor-
𝑏𝑥,𝑦 [𝑚] 𝐹𝑑,𝑟𝑜𝑡 Rotational drag force [𝑁]
dinate and index

𝑑𝑥 Grid cell size in x-direction [𝑚] 𝐹𝑔 Reduced weight force [𝑁]

𝑑𝑦 Grid cell size in y-direction [𝑚] 𝐹𝑓 Friction force [𝑁]

𝑑𝑡 Computational timestep [𝑠] 𝐹𝑅 Resulting force [𝑁]

Movement in x-direction per 𝑑𝑥


Δ𝑏𝑥𝑛 timestep [ ] 𝐹𝑟𝑒𝑠 Sum of resisting forces [𝑁]
𝑡
𝑚
𝑔 Acceleration due to gravity [ ] 𝐼𝑧 Moment of intertia [𝑘𝑔 ∙ 𝑚²]
𝑠²

ℎ𝑓 Height of fluid flow phase [𝑚] ⃗⃗⃗⃗⃗⃗


𝑀𝐵 Torque [𝑁 ∙ 𝑚]

ℎ𝑠 Height of solid flow phase [𝑚] 𝑉𝐵 IBM/boulder volume [𝑚³]

𝑚𝐵 IBM/boulder mass [𝑘𝑔]

𝑛 Index of a boundary node [−]

139
Numerical Boulder Transport Model

7.2 Immersed Boundary Approach

The Immersed Boundary Approach (from here denoted as IBM) goes back to Peskin (1972) who
started to use Lagrangian coordinates for describing an object in a flow field calculated on an Eu-
lerian grid. Peskin (1972) initially used IBM for describing blood flow in a beating heart and, thus,
flow in an elastic moving membrane. The scheme of Peskin (1972) was later enhanced for the
simulation of rigid bodies, e.g., by Beyer (1992) and Lai and Peskin (2000). Compared to other
numerical methods based on an Eulerian grid, the application of IBM benefits from comparable low
computational cost and the ability to preserve even complex boundary geometries comparable
easily (Luo, 2013; Dorschner, 2020). In IBM approaches, the immersed boundary is added to a
Cartesian grid by Langrangian coordinates as shown in Figure 7-1 and Figure 7-2. Today the IBM
technique is applied in various research fields, i.e., in medicine or engineering and several different
approaches for handling the interaction between fluids and rigid (non-flexible) objects are devel-
oped. Focusing on rigid bodies, the IBM technique can be divided in techniques based on feed-
back-forcing (FF) and direct-forcing (DF) approaches (Gronskis and Artana, 2016).

Figure 7-1: Immersed Boundary (IBM) setup.

Figure 7-2 : a) General functionality of the immersed boundary. b) Basic properties describing the local boundary geom-
etry.

The IBM is chosen as approach here in order to be independent from the underlying grid regarding
the boulder shape. Since the model is intended to be used in boulder transport research and, in
particular, regarding investigations of the boulder shape, a free choice of the shape is the leading

140
Numerical Boulder Transport Model

stipulation of the program. The IBM is suitable for simulating the movement of rigid bodies, also
with sharp interfaces, as already shown by e.g., Liao et al. (2010), Cai et al. (2017), and Kumar et
al. (2020). The application for highly turbulent flows, like in the here investigated case, is shown in
Iaccarino et al. (2003), Ye et al. (2021) and Constant et al. (2021), for example.

7.3 General approach

The boulder is modelled by applying an IBM. Here, the immersed boundary is described by a set
of point coordinates shaping a closed not interjecting polygon of arbitrary shape. The boulder han-
dling is implemented in the general two-phase flow model of Pudasaini (2012) as a subroutine,
which is called in every time-step of the main code. The general numerical scheme is adapted from
Liao et al. (2010) as follows:

1. Initialize the computational domain.

2. Initialize the IBM.

3. Start main loop of the two-phase flow model.

4. Call IBM Subroutine.

5. Update the velocity field in the two-phase model.

6. Repeat steps 4 and 5 until the final time-step is reached.

The handling of the boulder as immersed boundary requires the following steps in the subroutine:

1. Identify cut cells of the computational grid.

2. Identify the cell nodes of the cut cells inside and outside of the boundary.

3. Identify the forcing cell nodes from which the force calculations are driven.

4. Calculate translational and rotational boundary movement by solving force equations.

5. Update the velocity field around the boulder/IBM.

Step 5 requires interpolation algorithm and time-stepping schemes in common IBM codes. In the
present approach, the velocity field can automatically be updated by using a numerical slope which
describes the boulder geometrically in three-dimensions. Subsequently, the boulder can be han-
dled as a topographic feature and only the effect of boundary displacement has to be addressed
by suitable boundary conditions. Figure 7-2 shows the general application of the IBM on the Car-
tesian grid.

7.4 Boundary handling

7.4.1 Boundary detection

The boundary detection algorithm is mainly based on geometrical processes. As first step, the
number of cuts of a specific grid node is determined. In most cases it is sufficient to determine if
the x coordinate of a cell node is located within a boundary segment ⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗
𝑏 𝑛 𝑏 𝑛+1 in either only positive
or negative direction. However, this basic approach is not valid for a couple of exceptions. Such
an exception depicts the situation that multiple boundary nodes are located in line with the grid

141
Numerical Boulder Transport Model

node, resulting in whether no cut or additional cuts with every boundary node, for example. To
catch that exception, the currently inspected boundary node is tested regarding previous or sub-
sequent boundary nodes on the same 𝑥 or 𝑦 direction. In the following, the detection algorithm and
the interception of exceptions is described for the 𝑦-direction.

At first, the grid cells which have to be investigated are reduced to the rectangular extend of the
IBM by detecting its minimum and maximum 𝑥 and 𝑦 coordinates:

𝑖,𝑗
𝑥𝑚𝑖𝑛 = [𝑚𝑖𝑛(𝑏𝑥 ) − 𝑑𝑥, 𝑚𝑖𝑛(𝑏𝑦 ) − 𝑑𝑦] (7-1)

𝑖,𝑗
𝑥𝑚𝑎𝑥 = [𝑚𝑎𝑥(𝑏𝑥 ) + 𝑑𝑥, 𝑚𝑎𝑥(𝑏𝑦 ) + 𝑑𝑦] (7-2)

From the now gained set of cells, the cells actually cut by the boundary are identified by checking
if a boundary node 𝑏𝑥,𝑦
𝑛
is located inside of the current cell. In this step, boundary nodes can prin-
cipally be assigned to multiple cells (e.g., if boundary node is located exactly on the cell border).
Therefore, a rule is established assigning the cells corresponding to angle between the preceding
and following boundary node (Figure 7-3).

At first, the angles described by the resultant straight line between the preceding and following
boundary nodes is calculated by:

𝑛+1 −𝑏 𝑛−1
𝑏𝑦
(7-3)
𝑦
𝜔𝑥𝑛 = 𝜋 − 𝑎𝑐𝑜𝑠 ( 2 2
)
√[𝑏𝑥𝑛+1 −𝑏𝑥𝑛−1 ] +[𝑏𝑦
𝑛+1 −𝑏 𝑛−1 ]
𝑦

𝑏𝑥𝑛+1 −𝑏𝑥𝑛−1
𝜔𝑦𝑛 = 𝜋 − 𝑎𝑐𝑜𝑠 ( 2 2
) (7-4)
𝑛+1 −𝑏 𝑛−1 ] +[𝑏 𝑛+1 −𝑏 𝑛−1 ]
√[𝑏𝑦 𝑦 𝑥 𝑥

for all nodes. In dependence on the before calculated angles, cells are assigned by

142
Numerical Boulder Transport Model

𝑛 𝜋 𝜋 𝑖,𝑗
𝑏𝑥,𝑦 ∈ 𝑥 𝑖,𝑗 if (𝜔𝑥𝑛 < 𝜋) ∧ (𝜔𝑥𝑛 ≥ ) ∧ (𝜔𝑦𝑛 < ) ∧ (𝜔𝑦𝑛 ≥ 0) ∧ (𝑏𝑥𝑛 ≥ 𝑥𝑥 ) ∧
2 2
𝑖,𝑗 𝑖,𝑗 𝑖,𝑗
(𝑏𝑥𝑛 ≥ 𝑥𝑥+𝑑𝑥 ) ∧ (𝑏𝑦𝑛 > 𝑥𝑦 ) ∧ (𝑏𝑦𝑛 < 𝑥𝑦+𝑑𝑦 )

𝑛 𝜋 𝜋 𝑖,𝑗
𝑏𝑥,𝑦 ∈ 𝑥 𝑖,𝑗 if (𝜔𝑥𝑛 < ) ∧ (𝜔𝑥𝑛 ≥ 0) ∧ (𝜔𝑦𝑛 < ) ∧ (𝜔𝑦𝑛 ≥ 0) ∧ (𝑏𝑥𝑛 ≥ 𝑥𝑥 ) ∧
2 2
𝑖,𝑗 𝑖,𝑗 𝑖,𝑗
(𝑏𝑥𝑛 ≥ 𝑥𝑥+𝑑𝑥 ) ∧ (𝑏𝑦𝑛 > 𝑥𝑦 ) ∧ (𝑏𝑦𝑛 < 𝑥𝑦+𝑑𝑦 )

∨ (7-5)

𝑛 𝜋 𝜋 𝑖,𝑗
𝑏𝑥,𝑦 ∈ 𝑥 𝑖,𝑗 if (𝜔𝑥𝑛 < ) ∧ (𝜔𝑥𝑛 ≥ 0) ∧ (𝜔𝑦𝑛 < ) ∧ (𝜔𝑦𝑛 ≥ 0) ∧ (𝑏𝑥𝑛 ≥ 𝑥𝑥 ) ∧
2 2
𝑖,𝑗 𝑖,𝑗 𝑖,𝑗
(𝑏𝑥𝑛 ≥ 𝑥𝑥+𝑑𝑥 ) ∧ (𝑏𝑦𝑛 > 𝑥𝑦 ) ∧ (𝑏𝑦𝑛 < 𝑥𝑦+𝑑𝑦 )

𝑛 𝜋 𝜋 𝑖,𝑗
𝑏𝑥,𝑦 ∈ 𝑥 𝑖,𝑗 if (𝜔𝑥𝑛 < ) ∧ (𝜔𝑥𝑛 ≥ 0) ∧ (𝜔𝑦𝑛 < ) ∧ (𝜔𝑦𝑛 ≥ 0) ∧ (𝑏𝑥𝑛 ≥ 𝑥𝑥 ) ∧
2 2
𝑖,𝑗 𝑖,𝑗 𝑖,𝑗
(𝑏𝑥𝑛 ≥ 𝑥𝑥+𝑑𝑥 ) ∧ (𝑏𝑦𝑛 > 𝑥𝑦 ) ∧ (𝑏𝑦𝑛 < 𝑥𝑦+𝑑𝑦 )

The above rules are established in order to ensure a boundary shape adjusted cell assignment as
exemplary shown in Figure 7-3.

Figure 7-3: Exemplary angle base cell assignment.

In the following, the rules which has to be fulfilled to consider a grid node as within the boundary
are shown. Graphical explanation of the conditions can be found in Figure 7-4 and Figure 7-5.

143
Numerical Boulder Transport Model

Figure 7-4: Graphical example for the boundary detection algorithm. The numbers follow the condition 1 to 7. Note:
Shown is a simple example in which no exceptions (e.g., aligned boundary nodes) occur.

In several of the geometric operations, it is important to decide in which direction a boundary seg-
ment is defined ⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗
𝑏 𝑛 𝑏 𝑛+1 or ⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗
𝑏 𝑛−1 𝑏 𝑛 . For determining the cuts in y-direction, this is decided by com-
paring the grid node location to the middle axis of the area over which the IBM expands (Figure
7-2a)

𝑖.𝑗
−1 𝑖𝑓 𝑥𝑦 < 𝐴𝐶𝑦
𝛥𝑛 = { 𝑖.𝑗
(7-6)
+1 𝑖𝑓 𝑥𝑦 ≥ 𝐴𝐶𝑦

ACy describes the centre of the rectangular area over which the IBM spans on the main grid as
defined by Eqs. (7-1) and (7-2).

In order to account a boundary segment as cut by the current cell node, the cell node has to be
located in between a segment 𝑏𝑥𝑛 or 𝑏𝑥𝛥𝑛 . This can be achieved by checking the absolute distances
between the cell and boundary node, as shown in Condition 1. Here the distance between the two
boundary nodes needs to be larger than the distance between cell-node and 𝑏𝑥𝑛

144
Numerical Boulder Transport Model

Condition 1:
𝑖,𝑗
|𝑏𝑥𝑛 − 𝑥𝑥 | < |𝑏𝑥𝛥𝑛 − 𝑏𝑥𝑛 |
∨ (7-7)
𝑏𝑥𝑛 = 𝑏𝑥𝛥𝑛

Condition 1 does not contribute to the location of the grid node, whether it is located in positive or
negative 𝑥-direction to the boundary segment. Therefore in

Condition 2:
𝑖,𝑗
|𝑏𝑥𝑛 − 𝑥𝑥 | < |𝑏𝑥𝛥𝑛 − 𝑏𝑥𝑛 |

𝑖,𝑗 𝑖,𝑗
[(𝑏𝑥𝛥𝑛 ≤ 𝑥𝑥 ) ∧ (𝑏𝑥𝑛 > 𝑥𝑥 )]
∨ (7-8)
𝑖,𝑗 𝑖,𝑗
[(𝑏𝑥𝑛 ≤ 𝑥𝑥 ) ∧ (𝑏𝑥𝛥𝑛 > 𝑥𝑥 )]

𝑖,𝑗 𝑖,𝑗
[(𝑏𝑥𝑛 ≤ 𝑥𝑥 ) ∨ (𝑏𝑥𝛥𝑛 ≤ 𝑥𝑥 )]

the valid location is checked, depending on the currently investigated quadrant.

Condition 3 accounts for the case, that the grid node is located below an angular boundary segment
but not above a complete segment. Since cuts are only valid if occurring in positive 𝑥- and 𝑦-
direction, below (or left) lying segments are not allowed to be counted. Therefore, either both or, at
least one boundary segment node has to be located above the current grid node

Condition 3:
𝑖,𝑗 𝑖,𝑗
[(𝑏𝑦𝛥𝑛 < 𝑥𝑦 ) ∧ (𝑏𝑦𝑛 ≥ 𝑥𝑦 )]

𝑖,𝑗
[(𝑏𝑦𝛥𝑛 ≥ 𝑥𝑦 ) ∧ (𝑏𝑦𝑛 ≤ 𝑥𝑦 )]
𝑖,𝑗
(7-9)

𝑖,𝑗 𝑖,𝑗
[(𝑏𝑦𝑛 ≥ 𝑥𝑦 ) ∧ (𝑏𝑦𝛥𝑛 ≥ 𝑥𝑦 )]

Condition 4 and 5 ensure that the segment 𝑏 𝑛 𝑏 𝑛+1 is not describing a perpendicular line on the
⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗
same 𝑥-coordinate on which the cell node is located. Subsequently, only the last segment of such
a straight line is considered as cut. Condition 4 applies only if part of the extended segment de-
scribed by 𝑏 𝑛−1 , 𝑏 𝑛 and 𝑏 𝑛+1 is located directly on the grid line. Then only one intersection can
be counted if either the positive (𝑏 𝑛 𝑏 𝑛+1 ) or negative (𝑏
⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗ 𝑛−1 𝑏 𝑛 ) segment part deviates from the
⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗

grid line.

Condition 4:
𝑖,𝑗
𝑥𝑥 > 𝑏𝑥n−1 ∨ 𝑥𝑥 > 𝑏𝑥𝑛+1
𝑖,𝑗
(7-10)

145
Numerical Boulder Transport Model

Condition 5 works as an additional check in 𝑦-direction since for an outlying grid node which is in
line with a straight boundary segment, also only one intersection would be counted. However, this
error can be avoided by checking also the cuts in 𝑦-direction which can be zero, odd, or even, too
(compare Figure 7-5).

Condition 5:
(𝑏𝑥𝑛 > 𝑏𝑥Δ𝑛 ) ∨ (𝑏𝑥Δ𝑛 > 𝑏𝑥𝑛+1 ) (7-11)

Condition 6 is only false if 𝑏 𝑛 is located directly on a grid node. In this case, the cell node is handled
as outside of the boundary and not counted as interception.

Condition 6:
𝑛 −𝑏 Δ𝑛 𝑖,𝑗 Δ𝑛
𝑏𝑦 𝑥𝑦 −𝑏𝑦
𝑦
Δ𝑛 − ≠0 (7-12)
𝑏𝑥𝑛−𝑏𝑥 𝑖,𝑗
𝑥𝑥 −𝑏𝑥Δ𝑛

Condition 7 accounts for detecting peaks in the boundary only originating from a single boundary
node. Such cases are not counted.

Condition 7:
[(𝑏𝑦𝑛−1 > 𝑏𝑦𝑛 ) ∧ (𝑏𝑦𝑛+1 < 𝑏𝑦𝑛 )] ∨ [(𝑏𝑦𝑛−1 < 𝑏𝑦𝑛 ) ∧ (𝑏𝑦𝑛+1 > 𝑏𝑦𝑛 )] ∨
[(𝑏𝑦𝑛−1 = 𝑏𝑦𝑛 ) ∨ (𝑏𝑦𝑛+1 = 𝑏𝑦𝑛 )]
(7-13)

[(𝑏𝑦𝑛 > 𝑏𝑦𝑛−1 ) ∧ (𝑏𝑦𝑛 > 𝑏𝑦𝑛+1 )] ∨ [(𝑏𝑦𝑛 < 𝑏𝑦𝑛−1 ) ∧ (𝑏𝑦𝑛 < 𝑏𝑦𝑛+1 )]

If conditions 1-7 are fulfilled, the counter for cuts is increased by one. Therefore, finally the number
of cuts is known

𝑛𝑥,𝑐𝑢𝑡𝑠 = number of cuts in x direction (7-14)

At this stage, a pre-selection is conducted in which the location of the cell node is set as inside or
outside following:

∈𝛺 if 𝑛𝑐𝑢𝑡𝑠 is odd
𝑖,𝑗
𝑥𝑥 = { (7-15)
∉𝛺 if 𝑛𝑐𝑢𝑡𝑠 is even

The above-described procedure is conducted both in 𝑥- and 𝑦-direction. Subsequently, the number
of cuts of a grid node is known.

Based on this information, and as additional check regarding condition 5, it is tested if the number
of cuts in in 𝑥- and/or 𝑦-direction is zero or if the sum of the cuts in in 𝑥- and 𝑦-direction is even. If
the sum is odd, this is an indication for an error in count either in 𝑥- and 𝑦-direction. If this is the
𝑖,𝑗
case, the investigated grid node is taken as outside the IBM which is noted as 𝑥𝑥 ∉ 𝛺.

Condition 8:

𝑖,𝑗 ∈𝛺 if 𝑛𝑥,𝑐𝑢𝑡𝑠 ≠ 0 ∧ 𝑛𝑦,𝑐𝑢𝑡𝑠 ≠ 0 ∨ [𝑚𝑜𝑑(𝑛𝑥,𝑐𝑢𝑡𝑠 , 2) + 𝑚𝑜𝑑(𝑛𝑦,𝑐𝑢𝑡𝑠 , 2)] ≠ 0


𝑥𝑥 = { (7-16)
∉𝛺 if 𝑛𝑥,𝑐𝑢𝑡𝑠 = 0 ∨ 𝑛𝑦,𝑐𝑢𝑡𝑠 = 0 ∨ [𝑚𝑜𝑑(𝑛𝑥,𝑐𝑢𝑡𝑠 , 2) + 𝑚𝑜𝑑(𝑛𝑦,𝑐𝑢𝑡𝑠 , 2)] = 0

Now for all cut cells the location of all four grid nodes is known. A simple example is described
graphically in Figure 7-4.

146
Numerical Boulder Transport Model

Figure 7-5: Graphical explanations of intercepted exceptions within the boundary detection algorithm.

7.4.2 Numerical slope

The interaction between the flow field and the IBM is mainly calculated by using a numerical slope.
The numerical slope is derived from the grid elevation which is interpolated from the IBM. In order
to achieve a smooth height elevation and to avoid errors resulting from the interpolation, the IBM
is extended with additional layers describing the height. Figure 7-6 shows this process by a circular
IBM of ten height layers.

Figure 7-6: a) The initially plain grid. b) The geometrical layers describing the height and shape of the IBM. c) Interpola-
tion of the grid elevation. d) Calculation of the numerical slopes in x and y direction.

147
Numerical Boulder Transport Model

The numerical slope is finally calculated on the Cartesian grid by central difference following

𝑖+1,𝑗 𝑖,𝑗 𝑖−1,𝑗


𝛼𝑥 =
𝑥ℎ −2⋅𝑥ℎ +𝑥ℎ
(7-17)
𝑑𝑥 2
𝑖,𝑗+1 𝑖,𝑗 𝑖,𝑗−1
𝑥ℎ −2⋅𝑥ℎ +𝑥ℎ
𝛼𝑦 = (7-18)
𝑑𝑦 2

In order to incorporate the numerical slope in the flow calculations, it is attached to the source
terms of Pudasaini (2012) as described in detail in Kattel (2018) and in Kafle (2019).

7.4.3 Force calculation

The boundary movement is calculated by balancing the forces of drag, friction, and gravity respec-
tively buoyancy, at the boundary. In the following, the calculation of the forces inducing translation
and rotation to the IBM are shown for the x-direction. Indeed, due to the two-dimensional numerical
domain, these equations have to be solved equally in y-direction. The basic equation to be solved
reads

𝐹𝐷 + 𝐹𝑓 + 𝐹𝑔 = 0 (7-19)

where 𝐹𝐷 [𝑁] describes the force of drag, 𝐹𝑓 [𝑁] the force of friction and 𝐹𝑔 [𝑁] the (reduced) gravity
force. The general drag force for a rigid body is calculated following

𝐹𝐷,𝑡𝑟 = 0.5 ⋅ 𝜌𝐹 ⋅ 𝐴 ⋅ |𝑢𝐹 − 𝑢𝐵 | ⋅ (𝑢𝐹 − 𝑢𝐵 ) (7-20)

Additionally, viscous forces and skin drag force are implemented, but currently not activated.

Due the application of an IBM, the forces are not calculated for the boulder/IBM as one object.
Instead, the forces are calculated for every boundary segment (Figure 7-2, Figure 7-7) depending
on the size and alignment of a segment ⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗
𝑏 𝑛 𝑏 𝑛+1 . As advantage, we are able to contribute for the
IBM shape and corresponding properties (e.g., moment of inertia) in detail. In Eq. (7-20), ρF de-
scribes the density of the two-phase flow resulting from the fluid and solid phase following

𝑖,𝑗 ℎ𝑓 ℎ𝑠
𝜌𝐹 = [(
ℎ𝑠 +ℎ𝑓
⋅ 𝜌𝑓 ) + (
ℎ𝑠 +ℎ𝑓
⋅ 𝜌𝑠 )] (7-21)
𝑖,𝑗

The impact area 𝐴𝑛𝐵 is determined by the boundary segment perpendicular to the flow direction and
the wave height of the flow. Subsequently, for the force in in 𝑥-direction the impact area is calcu-
lated by

𝐴𝑛𝐵 = |𝑏𝑦𝑛+1 − 𝑏𝑦𝑛 | ⋅ ℎ𝐹,𝑖,𝑗 (7-22)

where ℎ𝐹,𝑖,𝑗 results from the sum of the solid and fluid flow phases

𝑖,𝑗 𝑖,𝑗 𝑖,𝑗


ℎ𝐹 = ℎ𝑓 + ℎ𝑠 (7-23)

All properties of boundary segments are computed in forward direction and a boundary segment
always refers to a coordinate pair of [𝑏 𝑛 𝑏 𝑛+1 ]. Consequently, and considering calculations of an-
gles, it has to be ensured that the IBM is always sorted in the same direction. In the present algo-
rithm, the boundary has to be initialized in the correct order, therefore. Before the first time-step,
ID-s are assigned to all IBM nodes which are kept throughout the simulation and sorted accordingly
after procedures not conserving the node sequence.

Generally, the forcing velocity on an IBM segment results to

148
Numerical Boulder Transport Model

𝑢𝐹𝑛,𝑛+1 = (1 − 𝛼𝑠
𝑖+𝛥𝐹𝑥,𝑗+𝛥𝐹𝑦
) ⋅ 𝑢𝑓 + (𝛼𝑠
𝑖+𝛥𝐹𝑥,𝑗+𝛥𝐹𝑦
) ⋅ 𝑢𝑠 (7-24)

The flow velocity is taken from cells adjacent to a boundary segment [𝑏 𝑛 𝑏 𝑛+1 ] and depending on
the segment angle (Figure 7-2, Figure 7-7).

Since the grid cells adjacent to the boundary either have velocity vectors of zero (inside the IBM)
or equal to the IBM velocity (outside of the IBM) these cells are not utilized for velocity interpolation.
The corresponding cell is then determined based on the alignment of the IBM segment to the com-
putational grid (Figure 7-7). This procedure can be described as a cell-shift-step in which the cut
cell is shifted in 𝑥 and/or 𝑦 direction, depending on the angle of the IBM segment. The idea of using
forcing nodes adjacent to the IBM and the general computational scheme is taken from Liao et al.
(2010) but was significantly differently implemented. The cell shift is calculated following (given in
degree):

−2
( ) if 𝜓(𝑏𝑖,𝑗 ) ≥ 355∘ ∧ 𝜓(𝑏𝑖,𝑗 ) ≤ 5∘
0
−2
( ) if 𝜓(𝑏𝑖,𝑗 ) > 5∘ ∧ 𝜓(𝑏𝑖,𝑗 ) < 85∘
2
0
( ) if 𝜓(𝑏𝑖,𝑗 ) ≥ 85∘ ∧ 𝜓(𝑏𝑖,𝑗 ) ≤ 95∘
2
2
𝑑𝑥 ( ) if 𝜓(𝑏𝑖,𝑗 ) > 95∘ ∧ 𝜓(𝑏𝑖,𝑗 ) < 175∘
( )= 2 (7-25)
𝑑𝑦 2
( ) if 𝜓(𝑏𝑖,𝑗 ) ≥ 175∘ ∧ 𝜓(𝑏𝑖,𝑗 ) ≤ 185∘
0
2
( ) if 𝜓(𝑏𝑖,𝑗 ) > 185∘ ∧ 𝜓(𝑏𝑖,𝑗 ) < 265∘
−2
0
( ) if 𝜓(𝑏𝑖,𝑗 ) ≥ 265∘ ∧ 𝜓(𝑏𝑖,𝑗 ) ≤ 275∘
−2
−2
{(−2) if 𝜓(𝑏𝑖,𝑗 ) > 275∘ ∧ 𝜓(𝑏𝑖,𝑗 ) < 355∘

Figure 7-7: Exemplary shift of the cut cell for the velocity interpolation on an IBM segment. a) Calculation of the orthog-
onal vector to the IBM segment. b) Shift of the cut cell depending on the alignment of the IBM segment to the computa-
tional grid. Note: The implemented shift has a value of 2dx respectively 2dy.

All flow parameters are taken from these cells (i.e., flow height, flow density, flow velocity).

The resulting velocity 𝑢𝑅,𝑥


𝑛,𝑛+1
contributing to the drag force 𝐹𝐷,𝑥 (𝑏 𝑛,𝑛+1 ) on the boundary segment is
subsequently the flow velocity from solid and fluid phases at a cell, following

149
Numerical Boulder Transport Model

2 2
𝑛,𝑛+1
𝑢𝑅,𝑥 𝑛,𝑛+1
= √(𝛽 ⋅ 𝑢𝐹,𝑥 𝑛,𝑛+1
⋅ 𝑐𝑜𝑠(𝜑𝑥 )) + (𝜁 ⋅ 𝜃 ⋅ 𝑢𝐹,𝑦 ⋅ 𝑠𝑖𝑛(𝜑𝑦 )) ⋅ 𝜃 (7-26)

where 𝜑𝑥 is the angle of ⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗


𝑏 𝑛 𝑏 𝑛+1 to the main grid and 𝛽, 𝜁 and 𝜃, are coefficients for the flow velocity,
regulating their contribution to the resulting flow velocity on ⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗
𝑏 𝑛 𝑏 𝑛+1 .

The angle of a boundary segment in x-direction ⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗


𝑏 𝑛 𝑏 𝑛+1 is simply calculated by (Figure 7-7)

𝑛+1 −𝑏 𝑛
𝑏𝑦
(7-27)
𝑦
𝜑𝑥 = 𝑎𝑐𝑜𝑠 ( 2 2
)
√(𝑏𝑥𝑛+1 −𝑏𝑥𝑛 ) +(𝑏𝑦
𝑛+1 −𝑏 𝑛 )
𝑦

𝛽, 𝜁 and 𝜃 regulate the actual influence of the velocity vectors and their components in 𝑥 and 𝑦
direction (Figure 7-8). If the angle of the boundary to the main grid and the velocity vector do not
correspond regarding their alignment to the basic grid, the velocity vector points away from the
immersed boundary and does, subsequently, not contribute to the drag force:

𝑖,𝑗
1 if 𝑢𝑥 /𝑐𝑜𝑠(𝜑𝐹,𝑥 ) ≥ 0
𝛽= { 𝑖,𝑗
(7-28)
0 if 𝑢𝑥 /𝑐𝑜𝑠(𝜑𝐹,𝑥 ) < 0

Since the immersed boundary is allowed to move freely, the flow velocity vectors in 𝑥 and 𝑦 direc-
tion also act in 𝑦 and 𝑥 direction in the case of diagonal boundary elements, for example (Figure
7-2b). However, as for the main velocity direction, the components do not act if pointing away from
the boundary which is handled by

𝑖,𝑗
1 if 𝑢𝑥 /𝑐𝑜𝑠(𝜑𝐹,𝑦 ) ≥ 0
𝜁= { 𝑖,𝑗
(7-29)
0 if 𝑢𝑥 /𝑐𝑜𝑠(𝜑𝐹,𝑦 ) < 0
𝑖,𝑗
Finally, the direction in which the secondary velocity component (which is 𝑢𝑦 for 𝐹𝐷,𝑥 ) acts has to
be determined. This can be achieved by determining the angle of the boundary section in terms of
the alignment to the main grid

1 if 𝑐𝑜𝑠(φ𝐹,𝑦 ) ≥ 0
θ= { (7-30)
−1 if 𝑐𝑜𝑠(φ𝐹,𝑦 ) < 0

Now, if the secondary velocity component points in boundary direction (𝜁 = 1), θ automatically
regulates in which direction the secondary velocity component acts, depending on the alignment
of the current boundary segment.

150
Numerical Boulder Transport Model

Figure 7-8: Exemplary behaviour of 𝛽, 𝜁 and 𝜃 for the calculation of 𝐹𝐷,𝑥 .

As last parameter the IBM velocity 𝑢𝐵 is determined by the displacement of an IBM node 𝑛 within
one time-step following:

𝑏𝑥𝑛,𝑡 −𝑏𝑥𝑛,𝑡−1
𝑛
𝑢𝐵,𝑥 = (7-31)
Δ𝑡

Finally, the drag force on a single boundary segment 𝑏𝑥𝑛,𝑛+1 in x-direction is calculated by

𝑖+𝐷𝑒𝑙𝑡𝑎𝐹𝑥 ,𝑗+𝐷𝑒𝑙𝑡𝑎𝐹𝑦
𝑛,𝑛+1
𝐹𝐷,𝑡𝑟 = 0.5 ⋅ ρ𝐹 ⋅ 𝐴𝑛,𝑛+1
𝐵
𝑛,𝑛+1
⋅ |𝑢𝑅,𝑥 𝑛
− 𝑢𝐵,𝑥 𝑛,𝑛+1
| ⋅ (𝑢𝑅,𝑥 𝑛
− 𝑢𝐵,𝑥 ) (7-32)

The drag force initiating rotation is calculated accordingly.

The reduced weight force 𝐹𝑔,𝑥 [𝑁] is applied following

𝐹𝑔 = (𝜌𝐵 − 𝜌𝐹,𝑎𝑣𝑔 ) ⋅ 𝑉𝐵 ⋅ 𝑔 (7-33)

with 𝜌𝐹,𝑎𝑣𝑔 taken as the average flow density of the forcing grid cells around the immersed bound-
ary. Friction force is calculated following

𝐹𝑓 = 𝐶𝑓 ⋅ (𝜌𝐵 − 𝜌𝐹,𝑎𝑣𝑔 ) ⋅ 𝑉𝐵 (7-34)

7.5 Calculation of movement

7.5.1 General

The presented algorithm is a variant of so-called trajectory models. Here, the movement is calcu-
lated as a combination of translation (linear unidirectional movement of all boundary nodes) and
rotation. In comparison to the translational movement, rotation needs to be handled more carefully

151
Numerical Boulder Transport Model

since the shift during a time-step differs between the boundary nodes, depending to boulder shape
and distance to the centre of mass of the boulder.

However, both, translation and rotation, are based on the force equilibrium of the boulder. Here,
the resulting force is calculated by integrating the individual forces on boundary segments.

The here applied drag force value of 𝐹𝐷,𝑥 is smoothed over three time-steps following

1 2 1
𝑡
𝐹𝐷,𝑥 = 𝐹𝐷,𝑥
4
𝑡−1
+ 𝐹𝐷,𝑥 𝑡−2
+ 𝐹𝐷,𝑥
4 4
(7-35)

(and accordingly for 𝐹𝐷,𝑦 ) in order to mitigate the effect of artifacts from the flow field around the
boulder resulting from movement.

7.5.2 Translation

The calculation of the boundary movement requires the following steps:

1. Calculation of the forcing nodes of the grid.

2. Calculation of the force acting on every boundary segment ⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗⃗


𝑏 𝑛 𝑏 𝑛+1 .

3. Calculation of the resulting force from the forces acting on the boundary segments.

4. Calculation of the force balance (Eq. (7-19)) and boundary acceleration.

5. Calculation of boundary shift in 𝑥 and 𝑦 direction.

Steps 1 to 3 are described in paragraphs 7.3 and 7.4. The resulting for 𝐹𝐷 is the simply the sum of
all forces linked to the boundary segments

𝐹𝑅,𝑥 = ∑𝑁 𝑛
𝑛=1 𝐹𝐷 (7-36)

The boulder acceleration in 𝑥 and 𝑦 direction is finally calculated by

|𝐹𝑑,𝑡𝑟,𝑥 +𝐹𝑝,𝑥 |−|𝐹𝑓,𝑥 +𝐹𝑔 |


𝑎𝐵,𝑥 = (7-37)
𝑚𝐵

The boundary movement during the current time-step can be calculated following:

𝑑𝑡 2 𝑑𝑡
Δ𝑏𝑥𝑛 = 0.5 ⋅ 𝑎𝐵,𝑥 ⋅ + 𝑢𝐵,𝑥 ⋅ (7-38)
𝑑𝑥 𝑑𝑥

where

𝑡−1
𝑢𝐵,𝑥 = 𝑎𝐵,𝑥 𝑡−2
⋅ 𝑑𝑡 + 𝑢𝐵,𝑥 (7-39)

If the boundary is already in movement and the impacting forces do not exceed the resisting forces
while pointing in opposite direction, the acceleration value is adjusted following

𝐹𝑓,𝑥 +𝐹𝑔 − 𝐹𝑑,𝑡𝑟,𝑥


𝑎𝐵,𝑥,𝑟𝑒𝑔𝑢𝑙𝑎𝑡𝑒𝑑 =
𝑚𝐵
𝑡−1
𝑖𝑓 (𝑢𝐵,𝑥 > 0) ∧ (𝑎𝐵,𝑥 < 0) (7-40)

⇒ 𝑎𝐵,𝑥,𝑟𝑒𝑔𝑢𝑙𝑎𝑡𝑒𝑑 = 0 𝑖𝑓 𝑎𝐵,𝑥,𝑟𝑒𝑔𝑢𝑙𝑎𝑡𝑒𝑑 > 0 (7-40.1)


𝐹𝑓,𝑥 +𝐹𝑔 + 𝐹𝑑,𝑡𝑟,𝑥
𝑎𝐵,𝑥,𝑟𝑒𝑔𝑢𝑙𝑎𝑡𝑒𝑑 =
𝑚𝐵
𝑡−1
𝑖𝑓 (𝑢𝐵,𝑥 < 0) ∧ (𝑎𝐵,𝑥 > 0) (7-41)

⇒ 𝑎𝐵,𝑥,𝑟𝑒𝑔𝑢𝑙𝑎𝑡𝑒𝑑 = 0 𝑖𝑓 𝑎𝐵,𝑥,𝑟𝑒𝑔𝑢𝑙𝑎𝑡𝑒𝑑 < 0 (7-41.2)

152
Numerical Boulder Transport Model

7.5.3 Rotation

For calculating the boundary rotation, additionally the torque and moment of inertia of the boundary
polygon are necessary. The moment of inertia is calculated by deriving all triangles and their areas
described by a boundary segment and the boulder centre

𝐴 𝑇1 = √(𝑏𝑥𝑛+1 − 𝑏𝑥𝑛 )2 + (𝑏𝑦𝑛+1 − 𝑏𝑦𝑛 )


2
(7-42)

𝐴 𝑇2 = √(𝑏𝑥𝑛+1 − 𝐵𝐶𝑥 )2 + (𝑏𝑦𝑛+1 − 𝐵𝐶𝑦 )


2
(7-43)

𝐴 𝑇3 = √(𝑏𝑥𝑛 − 𝐵𝐶𝑥 )2 + (𝑏𝑦𝑛 − 𝐵𝐶𝑦 )


2
(7-44)

Δ𝐴𝑡𝑟𝑖𝑎𝑛𝑔𝑙𝑒 = (𝐴 𝑇1 + 𝐴 𝑇2 + 𝐴 𝑇3 ) ⋅ 0.5 (7-45)

𝐴 𝑇 = √(Δ𝐴𝑡𝑟𝑖𝑎𝑛𝑔𝑙𝑒 ⋅ (Δ𝐴𝑡𝑟𝑖𝑎𝑛𝑔𝑙𝑒 − 𝐴 𝑇1 ) ⋅ (Δ𝐴𝑡𝑟𝑖𝑎𝑛𝑔𝑙𝑒 − 𝐴 𝑇2 ) ⋅ (Δ𝐴𝑡𝑟𝑖𝑎𝑛𝑔𝑙𝑒 − 𝐴 𝑇3 )) (7-46)

𝑖,𝑗 2
𝐼𝑧 = ∑𝑁 𝑛 2 𝑛
𝑛=1 ℎ𝐹 ⋅ ρ𝐵 ⋅ 𝐴 𝑇 ⋅ ((𝑏𝑥 − 𝐵𝐶𝑥 ) + (𝑏𝑦 − 𝐵𝐶𝑦 ) ) (7-47)

Furthermore, for calculating the angular acceleration the torque is needed. For calculating the
torque, the resulting centre of the force on the boundary is derived by

∫ 𝑥𝐹(𝑥)𝑑𝑥
𝑥𝑠,𝐹 =
∫ 𝐹(𝑥)𝑑𝑥
(7-48)

respectively

∫ 𝑦𝐹(𝑦)𝑑𝑦
𝑦𝑠,𝐹 =
∫ 𝐹(𝑦)𝑑𝑦
(7-49)

The torque is than calculated by the force balance following

⃗⃗⃗⃗⃗⃗
𝑀𝐵 = 𝐹𝐷,𝑥,𝑟 ⋅ (𝑦𝑠,𝐹 − 𝐵𝐶𝑦 ) + 𝐹𝐷,𝑦 ⋅ (𝑥𝑠,𝐹 − 𝐵𝐶𝑥 ) (7-50)

where (𝑥𝑠,𝐹 − 𝐵𝐶𝑥 ) and (𝑦𝑠,𝐹 − 𝐵𝐶𝑦 ) are the lever arm of the rotation inducing force. In (7-50), the
actual acting component of the drag force (𝐹𝐷,𝑥,𝑟 ) is adjusted if acting against the current rotational
direction by

𝐹𝐷,𝑥,𝑟 = 𝐹𝐷,𝑥 + 𝐹𝑟𝑒𝑠 𝑖𝑓(𝐹𝐷,𝑥 < 0) (7-51)


⇒ 𝐹𝐷,𝑥,𝑟 =0 𝑖𝑓 𝐹𝐷,𝑥,𝑟 > 0 (7-51.1)
𝐹𝐷,𝑥,𝑟 = 𝐹𝐷,𝑥 − 𝐹𝑟𝑒𝑠 𝑖𝑓(𝐹𝐷,𝑥 > 0) (7-52)
⇒ 𝐹𝐷,𝑥,𝑟 =0 𝑖𝑓 𝐹𝐷,𝑥,𝑟 < 0 (7-52.2)

and equally for 𝐹𝐷,𝑦 .

The angular acceleration of the boundary is than achieved by

⃗⃗⃗⃗⃗⃗⃗⃗
𝑀
α𝐵 =
⃗⃗⃗⃗⃗ 𝐵
(7-53)
𝐼𝑧

and the final movement is achieved by

Δω α𝐵 ⋅ 𝑑𝑡 2 + ω
⃗⃗⃗𝐵 = 0.5 ⋅ ⃗⃗⃗⃗⃗ ⃗⃗⃗ ⋅ 𝑑𝑡 (7-54)

153
Numerical Boulder Transport Model

𝑏𝑥𝑛,𝑡+1 = 𝐵𝐶𝑥 + ((𝑏𝑥𝑛 − 𝐵𝐶𝑥 ) ⋅ 𝑐𝑜𝑠(Δω


⃗⃗⃗𝐵 ) − (𝑏𝑦𝑛 − 𝐵𝐶𝑦 ) ⋅ 𝑠𝑖𝑛(Δω𝑏 )) (7-55)

respectively for the 𝑦 direction

𝑏𝑦𝑛,𝑡+1 = 𝐵𝐶𝑦 + ((𝑏𝑥𝑛 − 𝐵𝐶𝑥 ) ⋅ 𝑠𝑖𝑛(Δω


⃗⃗⃗𝐵 ) − (𝑏𝑦𝑛 − 𝐵𝐶𝑦 ) ⋅ 𝑐𝑜𝑠(Δω𝑏 )) (7-56)

7.5.4 Update of the flow field

The flow field is updated by assigning the boulder velocity as velocity component of the adjacent
cells and equalling the two following cells (Figure 7-9). The adjacent boulder cells have subse-
quently the same velocity as the boulder while the following cells have the same velocity as the
forcing cells (see Chapter 7.4.3). Velocity components inside the boundary area are set to zero.
Furthermore, fluid and solid levels of adjacent cells are equalled to the following. This is the sim-
plest approach and does not ensure continuity and mass conservation and needs, therefore, to be
enhanced in future versions.

Figure 7-9: Treatment of the flow field around the IBM.

The equations for the flow field update are implemented as

𝑑𝑥 𝑑𝑦
𝑖+
𝑢𝑓,𝑥 2
,𝑗+
2
= 𝑢𝑓,𝑥
𝑖+𝑑𝑥,𝑗+𝑑𝑦 (7-57)
𝑑𝑥 𝑑𝑦
𝑖,𝑗
𝑢𝑓,𝑥 = 𝑢𝑓,𝑥 2
𝑖+ ,𝑗+
2 (7-58)
𝑑𝑥 𝑑𝑦
𝑖+
ℎ𝑓,𝑥 2
,𝑗+
2
= ℎ𝑓,𝑥
𝑖+𝑑𝑥,𝑗+𝑑𝑦 (7-59)
𝑑𝑥 𝑑𝑦
𝑖,𝑗
ℎ𝑓,𝑥 = ℎ𝑓,𝑥 2
𝑖+ ,𝑗+
2 (7-60)

exemplary for the fluid phase and in in 𝑥-direction. The equations are solved for both phases in 𝑥-
and 𝑦-direction. 𝑑𝑥 and 𝑑𝑦 refer to the derived interpolation angle from Equation (7-25), in this
context.

154
Numerical Boulder Transport Model

7.6 Simulations

7.6.1 Setup

All shown simulations were executed in the model setup shown in Figure 7-10. The setup resem-
bles a flume-like wave basin with an artificial velocity gradient imposed on the left model boundary.
All basin boundaries are implemented as fully reflective.

The standard setup consists of an initial water-level of 10 m and an initial imposed flow velocity for
both phases of 2.5 m/s which is imposed over a standard period of four seconds. The maximum
boulder/IBM height is 25 m, even if the gravity force of the boulder is calculated for a standard
overall boulder-height of 1 m.

In this chapter, the different boulder behaviour according to its orientation (respectively shape) and
the flow density is shown in Chapter 7.6.2 (shape) and Chapter 7.6.3 (varied flow density). In
Chapter 7.6.4, two additional simulations are shown for proving the ability of modelling boulder
rotation. The used boulder shapes are listed in Table 7-2. The value of the coefficient of friction is
chosen as 𝐶𝑓 = 0.2 [−] which is a comparable low estimate (e.g., to 𝐶𝑓 = 0.42 [-] in Lodhi et al.,
2020) in order to promote movement within the test-simulations.
Table 7-2: Applied boulder shapes and in the presented simulations.

Boulder
a.1/a.2 b.1 c.1 c.2
Shape class triangle circle rectangle club

Area [m²] 129 41 84.5 56.25

Volume [m³]* 129 41 84.5 56.25

Density [kg/m³] 1600 1600 1600 1600

Max. height [m] 25 25 25 25

Mean height [m] 20 20 20 20

Initial distance to 20
10 15 20
left boundary [m] (straight part)

Coefficient of
0.2 0.2 0.2 0.2
friction 𝐶𝑓 [-]
*For the current functionality investigations, the volume is assumed to equal the area

Within this chapter, three-dimensional plots are used to show the force distribution on the boulder.
In these plots, forces acting in positive direction are plotted in red, in negative direction in blue, and
on sections where no force occurs in green. These plots are separated for 𝑥 and 𝑦 direction as
exemplary shown for shape a.1 in Figure 7-11.

155
Numerical Boulder Transport Model

Figure 7-10: Standard simulation setup for the numerical investigations.

Figure 7-11: Three-dimensional plots for clarification of the force distribution. In the example, the flow approaches the
boulder perpendicular to the rear face (right hand side in the figure). The black and the red arrow indicate directions in
which the force values are calculated as acting positively or negatively. On the boulder, red color indicates forces acting
in positive and blue in negative direction. Green markers indicate “neutral” forces in the sense of no impacting force at
the boulder segment.

7.6.2 Investigation on boulder shape (shape a.1 and a.2)

Figure 7-12 to Figure 7-14 show a comparison between the transport distances of differently ar-
ranged triangles within the first thirteen simulation seconds. Triangle a.1 is arranged streamwise
(sharp corner oriented to the flow) while triangle a.2 is arranged in mobilization supporting orienta-
tion (i.e., with the long face perpendicular to the wave impact). The earlier and faster acceleration
of a.2 is clearly observable. On the other hand, triangle a.1’s acceleration is more stable and lasts
longer as can be seen from Figure 7-15. In Figure 7-16, the force distribution on the boulder is
shown for the first 10 seconds of the simulation.

In detail, a.2 reaches a velocity of approximately 1.35 m/s after 0.6 s while a.1 is still in rest. In
contrary to a.2, a.1 accelerates more gradually and reaches its maximum velocity of approximately
1.16 m/s after 3.9 s. However, as can be seen from Figure 7-15, a.1 holds a fairly constant velocity
in comparison to a.2 which acceleration and deceleration phases are unstable over the whole

156
Numerical Boulder Transport Model

movement process. Furthermore, a.1 is remobilized by the reflected wave which now impacts the
long face of the boulder. This is not observed for a.2 where the reflected wave hits the streamline
edge but, as can be seen from Figure 7-15, it also occurs but with a significantly lower magnitude.

It has to be noted that the unsteady behaviour of a.2 is connected to the simple implementation of
the boundary conditions around the boulder. Due to the neglection of proper mass conservation
algorithms the force acting on the boulder is subject to short troughs if the boundary moves across
the border between two cells. This effect is mitigated by Equation (7-35) in which the force is
smoothed over three time-steps but is still observable. A secondary effect of this and the three-
dimensional boulder representation is the possibility of an uneven force distribution on the boulder,
which can result in small rotational movements (Figure 7-14).

157
Numerical Boulder Transport Model

Figure 7-12: First three simulation seconds of a.1 and a.2. The wave approaches from the left hand side to the boulder.
The wave dependent wave propagation is visible from second 1. Shown is the total wave height resulting from solid and
fluid phase.

Figure 7-13: Further wave development in the a.1 and a.2 setup. The transport distance of a.2 is significantly above a.1.

158
Numerical Boulder Transport Model

Figure 7-14: Seconds 10 to 13 in the a.1 and a.2 setups. a.1 is remobilized (seconds 11 to 13) due to the reflected wave.

Figure 7-15: Development of the absolute velocities of a.1 and a.2 over the whole simulation time.

Focussing on the wave impact within the first 10 seconds it can be observed that the symmetric
wave impact results in a positive force acting on the lower diagonal boundary face and a negative
force acting on the upper diagonal boundary face. Subsequently, the resulting force-component in
y-direction stays in equilibrium (i.e., 𝐹𝐷,𝑦,𝑟 = 0 [𝑁]) when the forces on the upper and lower face
are summed and no rotation is induced, proving a correct force interpolation on the boulder (Figure
7-16). At second four, where both shapes reach a peak velocity, the different force distribution on
both boulders can be seen. On a.1 the drag force moves with the wave tip along the diagonal faces
with comparable low values. On a.2 the whole positive force acts on the leading long face where

159
Numerical Boulder Transport Model

the peaks occur on the upper and lower edges. Furthermore, the hampering drag force at the right
boulder face reaches higher values for a.1 due to the straight boulder face.

Figure 7-16: Force distribution on a.1 and a.2 within the first 10 simulation seconds. The differences in the development
of the drag force can clearly be seen at seconds 2 and 4.

7.6.3 Investigation on sediment load (shape b.1)

Figure 7-18 to Figure 7-20 show a comparison between the transport distances caused by different
amounts of available sedimentary material in the first section of the basin. On the left-hand side,
6 % of the total fluid-solid mixture in the first basin section is solid while on the right-hand side only
2 % is solid. As can be seen in Figure 7-18, the whole mixture as well as the solid portion itself
moves faster through the basin in the case of 2 % solids. However, the boulder is faster accelerated
in the 6 % case.

160
Numerical Boulder Transport Model

Focussing on the maximum transport distance, the boulder is significantly further transported in
the 6 % case (Figure 7-20). Since the wave is reflected at the right basin boundary and due to the
faster propagating wave in the 2 % case, the boulder is accelerated again in this setting toward the
left basin boundary (Figure 7-20 and Figure 7-17). However, in the 6 % case a large gradient wave-
height/water-level gradient occurs when the wave passes the boulder (Figure 7-19; seconds 8 and
9.5). This high gradient seems to cause a short period of acceleration in negative in 𝑥-direction
which is rapidly superimposed by the wave-reflections on the left basin boundary resulting from the
initial wave generations (Figure 7-20; seconds 12 and 13). However, this acceleration in negative
direction in the 6 % case needs to be investigated further since the movement is not in accordance
with the absolute point of attack at the boulder.

As in the setup for comparing the boulder shape, small troughs of vanishing force impact are ob-
servable in this setup, too. The boulder is accelerated to its maximum velocity within the first 100
time-steps, respectively under one second (Figure 7-21). After reaching this maximum velocity the
boulder boundary crosses cell-borders and fluctuations in the acceleration are observable.

On the other hand, and due to the faster propagating wave in the 2 % case, the boulder is accel-
erated earlier here with regard to the simulation time but stays below the absolute velocity of the
6 % case until a phase of rapid deceleration in the 6 % case approximately at time-step 750 (sec-
ond 7.5; Figure 7-17). After this phase of rapid deceleration, the phase of acceleration in negative
x-direction can be seen between time-steps 750 and 1000 for the 6 % case, parallel to the phase
of complete deceleration until rest in the 2 % case.

The force distribution on the boulder is presented in Figure 7-21. The difference in the acting drag
becomes even more obvious as in the comparison of boulders a.1 and a.2, especially between
seconds 2 and 6.

Figure 7-17: Development of the absolute boulder velocity in the b.1 setup. As can be seen, both boulders are
remobilized at the end of the simulation time due to the wave reflections and low boulder wheight. The remobilization is
higher in the 2 % scenario due to the significantly higher wave velocity in the 2 % scenario.

161
Numerical Boulder Transport Model

Figure 7-18: Wave propagation (a) and density distribution (b) in the b.1 setup. The setup with the higher sediment
content (6 %, left) shows clearly a faster acceleration of the boulder, compared to the 2 % setup (right). From seconds
1 to 3 it is also observable that the wave propagates significantly faster in the 2 % setup.

162
Numerical Boulder Transport Model

Figure 7-19: Further wave propagtion (a) and density distribution (b) in the b.1 setup. The whole wave propagates faster
in the 2 % setup (right) while the boulder velocity is higher in the 6 % setup (left). The boulder is mobilized in negativ x-
direction in the 6 % case after approximately 8 seconds in the moment when the wave passes the boulder. This could
be introduced by the gradient between the wave height before and after the boulder but needs to investigated further.
However, the short period of backward-movement in the 6 % case is not observable in the 2 % case.

163
Numerical Boulder Transport Model

Figure 7-20: Seconds 10 to 13 in the b.1 setup (a: total wave height; b: density distribution). The boulder is transported
further in the 6 % setup while the boulder is remobilized in the 2 % setup due to the reflected wave at the right basin
boundary.

164
Numerical Boulder Transport Model

Figure 7-21: Force distribution on the boulder within the first 10 seconds of the b.1 setup. Between second 2 and 6 the
significant discripance of acting drag force between the 6 % (left) and 2 % (right) case is observable.

7.6.4 Rotation calculations (shapes c.1 and c.2)

As shown in Figure 7-14, rotation-inducing gradients can occur in the current version due to conti-
nuity/mass conservation errors during translation and rotation over the grid. Therefore, in this par-
agraph, the functionality of rotation is shown by two shapes in order to clarify the correct function-
ality of the rotation algorithm. For this, one shape is designed favouring rotation by an uneven area

165
Numerical Boulder Transport Model

distribution (shape c.2). For representation issues, translation is artificially prevented in the here
presented simulations so that no secondary errors introduced from translation can occur.

Figure 7-22 shows how the boulder shape affects whether rotation or not rotation is induced by the
wave. While the force impact in the symmetrical shape (c.1) remains at the centre point due to the
uniform wave application, the point of application in the asymmetrical shape initially shifts to the
point of first wave contact (Figure 7-22, t = 0.3 s). After the entire boulder c.2 is in the area of the
wave impact, the resulting point of load shifts towards the straight segment, due to the boulder
centre of gravity (Figure 7-22, t = 0.7 s). Subsequently, a lever arm is generated which leads to the
rotation of c.1(Figure 7-22, t = 1 s). In contrary, in the uniform shape the resulting force rests at the
centre of mass, and, since no lever arm is present, no rotation is introduced (seconds 0.7 to 3;
shape c.1).

With progressing simulation time and rotation of c.2, the point of force moves down according to
the altered impact area which aligns more and more to the shape (seconds 1.5 to 3).

Figure 7-22: Exemplary rotation behaviour in the boulder transport model. The black dot indicates the centre of mass
and the white dot the point of load due to wave impact.

As shown in Figure 7-22, the algorithm for rotation works correctly. However, at second 4.5 in
Figure 7-23 it is observable that an uneven force distribution occurs at c.1 (rectangular shape) and
the boulder is forced to rotate, as well (Figure 7-24). This has again to be assigned to the simple
algorithm for flow handling around the boulder. The boulder behaviour itself behaves correctly re-
garding the applied drag force and flow field.

Figure 7-25 shows simulation seconds 5.5 to 13 for c.2 (club shaped). It can be observed that the
boulder behaves correctly within the applied wave field and reacts correctly to shifts in the resulting
wave load. As expected, the imbalance prevents the boulder from aligning itself with its long side
parallel to the flow (Figure 7-25; seconds 6 to 8.5). On the lower side of the boulder, however, a
gradient in the water-level is created as a result of the rotation, which initiates a back rotation,

166
Numerical Boulder Transport Model

which in turn is stopped and reversed by the following current (Figure 7-25; seconds 9 to 12). Since
the gradient in the water level below the boulder has now disappeared, the energy in this time-step
is sufficient for the boulder to turn almost 180° from its initial position (Figure 7-25; seconds 11 and
12.5). Since the boulder is now again impacted on its long side by the following current, a rotation
in the opposite direction now starts again (Figure 7-25; second 13).

Figure 7-23: Force distribution on shapes c.1 and c.2 within the first 4.5 seconds. At second 4.5 an uneven force load
occurs at the leading face of c.1 and introduces rotation (Figure 7-24).

167
Numerical Boulder Transport Model

Figure 7-24: Seconds 4.5 and 5 of the simulation for shapes c.1 and c.2. For c.1, the induced rotation from the uneven
force load (Figure 7-23) is observable. c.2 is still in positive rotation.

Figure 7-25: Seconds 5.5 to 13 for the simulation of c.2. Right before second 5.5, the boulder is aligned nearly parallel
to the flow. Due to wave reflections at the lower domain boundary, the boulder is now accelerated in the opposite rotation
direction (seconds 5.5 to 8). Now the point of force moves again upwards due to the main flow field and rotation is forced
in positive direction, again (seconds 8.5 to 12). Due to the lower water level at the lower side of the boulder, positive
rotation lasts to a longer excursion before the rotation direction is inverted again (seconds 12 to 13). This rotation is now
related to the main flow field and not due to gradients in the flow height.

168
Numerical Boulder Transport Model

7.7 Discussion and outlook

7.7.1 Results

In the previous chapters, the general mechanical functionality of the developed Boulder Transport
Model (BTM) was shown. Comparing streamline and non-streamline orientations (shape a.1 and
a.2), clearly shows the importance of recognizing the shape/orientation in boulder transport calcu-
lations since the transport behaviour differs significantly even at the same flow conditions. The
streamwise oriented boulder is mobilized significantly later and transported more slowly.

Furthermore, simulations of shape b.1 indicate a strong influence of sedimentary load as transport
supporting factor, leading to significant higher acceleration values. However, due to the simple
boundary conditions around the immersed boundary, in future versions of the program it needs to
be checked, how accurate mass conservation will affect the boulder. Depending on the ratio be-
tween sedimentary load and boulder acceleration, a deposition of large portions of the sediment in
front of the boulder could even have a transport hampering effect if the flow/impact force is dis-
tracted from the boulder. Test simulations already indicated that a portion of 10 % sedimentary
load leads to wave velocities too low for mobilizing the boulder. However, this effect needs to be
proved or disproved with proper boundary conditions.

7.7.2 Limitations and outlook

The program has some significant limitations, so far, of which the major ones are the boundary
conditions between boulder and two-phase flow. The currently implemented version based on
equalling the velocity gradients and two-phase flow levels of the cells adjacent to the boulder,
depicts a very rudimentary approach and does not fulfil continuity and mass conservation. Here,
the boundary conditions need to be enhanced to more sophisticated approaches ensuring mass
conservation by pressure correction methods (e.g., MAC-SOLA solver; Kumar et al., 2020).

After the implementation of such correction methods, the model’s boundary conditions should be
adjusted in order to enable probable simulations of dam-break setups which would enable a com-
parison of the results to physical experiments for validation. Since such first validation experiments
should also include inclined shores, the boulder movement equations need to be enhanced for
considering variable friction and weight-load. The validation could then be based on small-scale
experiments by using idealized boulder models of different shapes utilizing a common dam-break
setup enhanced by sedimentary load. A rough sketch of how such experiments might be designed
is shown in Figure 7-26.

Further smaller adjustments might be needed regarding the boundary detection algorithm since
the implemented cell assignment is not mathematically proved, so far, and might fail in some rare
configurations.

Once the general validation is successfully conducted, the model can be enhanced regarding the
wave generation mechanism (especially by implementing the earthquake model of Okada, 1985;
see Chapter 1; Figure 1-5) and enabling importing coastlines (e.g., by coupling the program with
GIS applications; compare Mergili and Pudasaini, 2021).

169
Numerical Boulder Transport Model

Figure 7-26: Rough sketch for a physical small-scale validation dam-break setup for the BTM. The sediment is added
immediately before releasing the dam-break.

The shown first results support the findings from the physical experiments and assumptions re-
garding the influence of sedimentary load. Therefore, further efforts for enhancing the boundary
conditions around the boulder, verificating and validating the program and coupling to GIS appli-
cation are definitely worth to conduct.

170
Synthesis and Outlook

8 Synthesis and outlook

8.1 Summary

In the preceding chapters, two main questions were tackled regarding (i) the possibility of tsunami
occurrences in the Caribbean Sea, and in particular at the island of Bonaire, and (ii) improving the
understanding of the process of tsunami-induced boulder transport process as basis for enhancing
currently existing analytical models (Chapter 1). These research questions were tackled as follows:

▪ At first, geological proxies throughout the Caribbean Sea were reviewed and two witnessed
and two hypothetic tsunami scenarios were numerically modelled for clarifying possible tsu-
nami impact heights, especially at the island of Bonaire (Chapters 2 and 3).

▪ The shortcomings of today’s analytical boulder transport models are connected to their under-
lying physical experiments. Published physical experiments were analyzed, their main short-
comings identified, and possible enhancements highlighted in Chapters 4 and 5.

▪ As revealed from Chapters 4 and 5, a major shortcoming in existing physical experiments on


tsunami-induced boulder transport is the utilization of oversimplified boulder shapes. The ques-
tion of how and to which extend the boulder shape influences the boulder transport is investi-
gated by physical experiments in Chapter 6 and simulated by the numerical boulder transport
model in Chapter 7.

An enormous number of geological tsunami indicators is reported throughout the coasts in the
Caribbean Sea, indicating tsunami events which are not reported so far. However, while tsunamis
cannot be excluded in some regions, strong storms are more likely to be the reason for relocation
(e.g., Grand Cayman, Jamaica; Jones and Hunter, 1992; Robinson et al., 2006; Engel et al., 2016;
Chapter 2). In other regions, combinations of tsunamis and storm waves are likely to have relocated
the boulders (e.g., on Eleuthera; Engel et al., 2015).

On Bonaire, observations during category five hurricanes revealed these hurricanes to be capable
for dislocating small clasts while it seems unlikely that coarse boulders, such as BOL2, were trans-
ported by wind-waves, which is supported by the application of analytical threshold models (Chap-
ter 2). However, due to recent observations during strong and landfalling hurricanes, their disloca-
tion (or remobilization) cannot be excluded, so far.

For proving or disproving the assumption of tsunami-caused transportation of Bonaire’s coarse


clasts, numerical simulations of witnessed and hypothetical tsunami scenarios in the Caribbean
Sea were conducted (Chapters 2 and 3). As turned out from these numerical investigations, either
the assumed tsunami triggers were not able to generate tsunamis as intense as necessary for
generating tsunami magnitudes high enough for dislocating the largest boulders on the island of
Bonaire, or the applied analytical models failed to predict the necessary energy appropriately. In
favor of the latter is that some boulder properties were not assumed properly in literature of which
some are still not discoverable today (e.g., the density distribution inside the boulder). For example,
the DGNSS measurements of BOL2 by Engel and May (2012) underestimated its volume, which
turns out to be approximately 9 m³ higher (68 m³ from DGNSS to 77 m³ from photogrammetry),
based on photogrammetric measurements (Chapters 3 and 6). However, as consequence from
divergent assumptions regarding the deposits on Bonaire (e.g., regarding the boulder density) and
applied models, the calculated minimum tsunami heights for dislocating boulders on Bonaire varies

171
Synthesis and Outlook

between < 3 m (Spiske et al., 2008), 6-13 m (Pignatelli et al., 2009), 7-9 m (Engel and May, 2012)
and 33-46 m (Scheffers, 2002b), showing the still present uncertainties connected to back calcu-
lations of tsunami magnitudes by coarse-clast deposits.

As known from literature and due to the contradictory results for the island of Bonaire, the lack of
accuracy in existing analytical boulder transport models was tackled in the following chapters. As
basis for this, parameters influencing the boulder transport process were investigated in detail for
which a literature review on existing physical boulder transport experiments were conducted (Chap-
ters 4 and 5). The literature review revealed that the oversimplification of the boulder shape in
experiments is one major source of the uncertainties in existing analytical boulder transport mod-
els, and this topic was chosen for further experimental investigations.

In the physical experiments focusing on the boulder shape, the importance of recognizing the
shape during field studies was elaborated (Chapter 6). Especially considering particular shape-
adjusted drag coefficients can lead to improved accuracies of analytical models since the boulder’s
impact and ground contact area have significant influence on mobilization and transportation. The
experiments clearly revealed that less energy is transferred from the wave to the boulder as the
boulder is increasingly streamline-shaped. Additional statistical analyses applied due to the partly
stochastic behavior of the transport process revealed that even smallest divergences in the model
setup, which exceed the limits of reproducibility (e.g., smallest divergences in the boulder orienta-
tion), can lead to significantly diverting results regarding the transport distance. This shows that
incorporating probability functions in boulder-deposit research might support appropriate estima-
tions of necessary tsunami magnitudes, as earlier proposed regarding mobilization thresholds by
Bressan et al. (2018). Based on the findings from the literature review and physical experiments,
a supporting tool for researchers was developed. The tool supports researchers to assess whether
common boulder transport equations tend to over- or underestimate the necessary wave energy,
depending on the local conditions.

An additional unknown factor during boulder transport is the influence of sedimentary load in the
tsunami, which is assumed to increase the maximum and total transport distance. To confirm the
results of the physical experiments (Chapter 6) and getting insights into the influence of sedimen-
tary load on said parameters, a numerical model was developed. First results support the assump-
tion of transport supporting influence of sedimentary load and furthermore confirms the findings of
the influence of boulder shape (Chapter 7).

8.2 Assessment of the research results

This PhD thesis begins with a comprehensive analysis of tsunami occurrences in the Caribbean
Sea, continues with a detailed look on the island of Bonaire and on shortcomings of published
physical boulder transport experiments, and ends with in-depth investigations of the influence of
particular boulder properties on its transportation likelihood. To the author’s knowledge, the present
PhD thesis is the first to overcome oversimplifications in physical boulder transport experiments by
applying a naturally shaped boulder model in physical and numerical experiments and, further-
more, is the first PhD thesis to explore the consequences of sedimentary load in tsunamis on boul-
der mobilization and transport distance.

172
Synthesis and Outlook

By this, the PhD thesis contributes not only to the tsunami hazard assessment in the Caribbean
Sea and on the island of Bonaire in particular, but also to a better understanding of the hydrody-
namic processes contributing to tsunami-induced boulder transport.

With regard to the research questions stated in Chapter 1, the following results are achieved:

1) Which earthquake zones in the Caribbean Sea could be capable of generating tsunamis im-
pacting Bonaire, and do these have sufficient energy for dislocating the largest boulder depos-
its on Bonaire? (Chapters 2 and 3)
➔ Several possible tsunami trigger scenarios were identified (Virgin Islands, Muertos Trough,
Southern Caribbean Deformed Belt, Macuto) and numerically modelled with Delft3D and
DelftDashboard. However, with heights around H = 1.30-3.40 m, the simulated waves were
not high enough for dislocating the largest boulder on the island of Bonaire as assumed by
analytical models (Hnecessary,Tsunami = 9-15 m depending on the analytical approach).

2) Which are the main shortcomings of published physical boulder transport experiments, and
which shortcomings are the most important to be solved in order to better reflect real scenarios
and improve the understanding of tsunami-induced boulder transport processes? (Chapters 4
and 5)
➔ The literature review revealed several shortcomings and oversimplifications in published
experimental studies on tsunami induced boulder transport. The main shortcomings which
need to be solved in order to better reflect real scenarios and increase the understanding
of tsunami-induced boulder transport emerge from oversimplified boulder shapes, neglect-
ing the stochastic nature of boulder mobilization and transport as an important factor, ne-
glecting sedimentary load, misinterpretations of the importance of bottom roughness and
oversimplification of contributing forces and hydrodynamic processes (i.e., impact force,
turbulence).

3) To which extent does the boulder shape influence the tsunami-induced boulder transport pro-
cess? (Chapter 6 and 7)
➔ The experimental study revealed that it is crucial to consider the boulder shape as im-
portant factor for boulder mobilization and transport. Regarding the boulder shape, it is
important to consider not only the boulder’s flatness but also account for its roundness and
streamlined shape. These findings are supported by first investigations with the developed
numerical boulder transport model.

4) Do numerical simulations support the assumption of a transport-supporting effect of sedimen-


tary load in the approaching tsunami? (Chapter 7)
➔ Simulations with the developed numerical boulder transport model support the assumption
that increasing sedimentary load in the wave has a transport- and mobilization-promoting
influence on the boulder transport process.

Numerical simulations of hypothetical and verified tsunami triggers in the Caribbean Sea revealed
no tsunami heights which are able to shift the largest boulder on the island of Bonaire, at least from
the current understanding of the processes (Chapters 2 and 3). However, accurate in-situ meas-
urements of Bonaire’s boulder record shows that there are still significant uncertainties regarding
the boulder’s properties and the process of their relocation which partly can be tackled by the
application of advanced measurement techniques like photogrammetry. This PhD thesis is one of

173
Synthesis and Outlook

the first in which boulder deposits were measured by photogrammetry, an approach which now
spreads throughout the research community (e.g., Rovere et al., 2017; Boesl et al., 2020).

Beside improving measurements in the field, the application of photogrammetry allowed to conduct
experimental investigations by using a naturally-shaped boulder model (Chapter 6): As first exper-
imental study investigating boulder shapes overcoming simplified models like cubes or spheres, it
was possible to contribute fundamentally new insights regarding the behavior of different boulder
shapes under tsunami impact. Even if the effect of the boulder shape can be superimposed by
other parameters (i.e., the pre-transport setting), the importance of recognizing the shape is clearly
shown by differences in the maximum transport distance of ca. 50 % between the naturally shaped
and idealized cubic boulder of the same weight. Transferred to the application of analytical thresh-
old models, which are often applied by researchers without hydrodynamic-engineering back-
ground, this can also result in over- or underestimations of the estimated wave heights, at least in
the same percentage range. Already being aware of the crucial influence of the boulder shape, can
support researchers in the field for more accurate estimations of the threshold tsunami height.

Statistical analysis of the experimental study showed furthermore the importance of recognizing a
kind of stochastic nature of boulder mobilization and transport. Due to the highly turbulent flow,
even the smallest divergences in the initial model setup led to significantly deviating results. Based
on the statistical evaluation, it was possible to show that these deviations increase with an increas-
ing average transport distance of an experimental setup. This is a further factor which needs to be
considered in the field, since this illustrates with how much caution the implications of boulders for
the reconstruction of past extreme-wave events have to be treated with, even if the boundary con-
ditions are perfectly known. As suggested in Bressan et al. (2018) regarding the initiation of motion,
also back calculations based on the transport distance (e.g., for estimations on the inundation area)
should therefore presumably be based on ranges of values and not targeting a fixed result value.
In the experimental study within the present PhD thesis, most previously stated pre-transport set-
ting characteristics (e.g., findings of Liu et al., 2015; Bressan et al., 2018) were confirmed, and it
was shown that submergence and boulder alignment to the flow significantly influence mobilization
and transport distance.

As final part of the PhD thesis, the first version of an advanced numerical model was developed
enabling researchers to investigate not only different boulder shapes and pre-transport settings but
also the influence of sedimentary load in the tsunami (Chapter 7). First results show the importance
of taking these parameters into account, showing that even for idealized shapes (e.g., triangles)
the boulder alignment is crucial. Additional sedimentary load leads to significantly increased accel-
eration and transport distance. However, the current model is not verified and validated so far
which is a crucial prerequisite for its practical application. Before the model can be validated, a
pressure correction boundary condition should be implemented around the boulder in order to en-
sure mass conservation and continuity, which is not fulfilled currently.

As another major outcome, a unique guidance tool for researchers in the field was developed
(Chapter 4) which is intended to support estimations if applied analytical models might tend to over-
or underestimate the actual wave energy depending on the local setting. In this tool, a set of pa-
rameters (boulder flatness and orientation, pre-transport submergence, bottom roughness and
boulder weight) can be evaluated using the given diagram, which reveals if the local conditions
tend to amplify or hamper the boulder transport.

174
Synthesis and Outlook

8.3 Outlook and future directions in boulder transport research

The research on tsunami-induced boulder transport is a comparably young discipline which faces
problems arising from the natural boundary conditions. Instead of dealing with comparably clear
geometrics as, for example, in the case of relocated shipping containers in harbors, the topic of
boulder transport needs to be handled more carefully, recognizing irregular boundary conditions in
terms of boulder shape and shore shape.

Since natural boulders do not consist of simple geometrical shapes such as cuboids, associated
research needs to take into account intermediate shapes while also ensuring practically manage-
able and quantifiable models. Therefore, in future experimental and numerical studies a focus
should be placed on investigating the boulder shape to a larger extent. In the experimental study
of the present PhD thesis, three shapes were compared, and the results clearly indicate that in-
vestigations of intermediate steps are necessary in order to quantify the shape influence accu-
rately. The experimental results revealed that large shifts in the shapes, in the present case from
an idealized cubic over the natural to a flat boulder model, indeed show a significant influence of
the shape on mobilization and transport. However, the coarse increments between the shapes do
not allow to derive a single empirical relationship applicable for all these shapes. For reaching such
empirical relations and incorporating them into analytical models, further investigations using mul-
tiple boulder shapes with a slightly increasing number of faces, and subsequently, shape complex-
ity, are of paramount importance.

An incidental finding of the experimental study was the highly sensitive boulder behavior in the
turbulent flow. Since additional irregular conditions in nature amplify the (kind of) stochastic behav-
ior, applying probability instead of exact functions for the transport distance could support a clearer
understanding of the informative value of the parameters calculated from boulder deposits. Hereby,
the proposed probability-based approach for incipient motion by Bressan et al. (2018) is clearly
supported, and an extension to the transport process overall is recommended. An option might be
to transfer connected findings from fluvial processes which are, however, developed for continuous
streams and small particles, and their transferability to tsunami-induced boulder transport is there-
fore severely limited (Chapters 4 and 6).

Furthermore, existing experimental studies differ not only with regard to the overall experimental
setup but also regarding the type and number of measurements. During the preparation of the
experimental setups of the present PhD thesis and the accompanying literature review, all attempts
for deriving relationships between the different published studies failed since the boundary condi-
tions differ too much. Therefore, and as recommended in the corresponding paper (Chapter 4), it
is proposed to establish a standard approach for boulder transport connected experimental inves-
tigations. This common standard could encompass guidelines regarding the positioning of hydro-
dynamic measurements (e.g., in a certain distance to the shore model) and normalized dimension
ratios to which studies can refer to, in order to enable comparability. In past published studies, for
example, postulations regarding the preferred transport mode (sliding, rolling, saltation) periodically
differ among researchers which is strongly assumed to result from different wave-boulder height
ratios and flow conditions (turbulence). However, such variables are often not clearly communi-
cated and can lead to uncertainties for the reader. As additional benefit, based on such a standard
case, a database could be established incorporating results from different research groups and
supporting comparisons between their results, and giving a set of benchmark tests for numerical
models.

175
Synthesis and Outlook

Regarding numerical models, the future might lead to full three-dimensional simulations for whole
coastal stretches. Such applications of full three-dimensional models for comparable large domains
becomes feasible due to the fast-developing computational speed which is also related to devel-
opments utilizing graphical processing units (GPUs) for solving the numerical schemes. In connec-
tion to boulder transport problems, models, which are verified for simulating the transport of rigid
bodies in highly turbulent flows, would at first allow to reproduce recent transport events on the
basis of accurate boulder and shoreline properties. Once verified based on recent events, such
models will open completely new options for increasing the knowledge on boulder transport phe-
nomena. However, even though the computational resources increase fast, modelling the boulder
transport on complete coastal stretches exceeds the currently available capabilities. Today such
models can hardly encompass areas in the size of small harbor basins (e.g., by Smoothed Particle
Hydrodynamics, SPH; Zainali and Weiss, 2013; Ruffini et al., 2021). However, the numerical results
of Zainali and Weiss (2013) showed partly an excellent agreement with the experimental results of
Imamura et al. (2008), and already indicates the potential of SPH for simulating tsunami- or storm-
wave-induced boulder transport.

The numerical boulder transport model established within the present PhD thesis will be further
developed in the future, but already revealed as functional, and first results support the findings
from the experimental study. Before adding new functionalities to the model and before its appli-
cation to real coastlines, the model equations need to be verified and its functionality needs to be
proofed by physical experiments. After the verification and validation process, the model can be
used for large-scale investigations at shore-lines since the computational needs are comparable
low due to the depth-averaged nature while it is still capable for simulations of non-idealized boul-
der shapes. A promising option in this regard would be its coupling to GIS applications in order to
enable numerical investigations of boulder transport for numerically-untrained researcher.

In conclusion, it should be said that increasing the understanding of boulder-tsunami interactions


and connected processes will significantly support the analysis of palaeotsunamis and, subse-
quently, the identification of threatened coasts and coastal communities. This potential, which can-
not be overestimated, is worth to further intensify the research on coarse tsunami deposits and
invest effort in consolidation of research approaches among different research groups.

176
References

References

Abad, M.; Izquierdo, T.; Cáceres, M.; Bernárdez, E.; Rodriguez‐Vidal J. (2020): Coastal boulder deposit as evidence of an
ocean‐wide prehistoric tsunami originated on the Atacama Desert coast (northern Chile). In: Sedimentology 67 (3), pp.
1505–1528. DOI: 10.1111/sed.12570.
Abe, T.; Goto, K.; Sugawara, D. (2012): Relationship between the maximum extent of tsunami sand and the inundation limit
of the 2011 Tohoku-oki tsunami on the Sendai Plain, Japan. In: Sedimentary Geology 282, pp. 142–150. DOI:
10.1016/j.sedgeo.2012.05.004.
Adger, W. N.; Hughes, T. P.; Folke, C.; Carpenter, S. R.; Rockström, J. (2005): Social-ecological resilience to coastal disasters.
In: Science (New York, N.Y.) 309 (5737), pp. 1036–1039. DOI: 10.1126/science.1112122.
Adomat, F.; Gischler, E. (2015): Sedimentary patterns and evolution of coastal environments during the holocene in Central
Belize, Central America. In: Journal of Coastal Research 31 (4), pp. 802–826. DOI: 10.2112/JCOASTRES-D-14-00093.1.
Adomat, F.; Gischler, E. (2016): Assessing the suitability of Holocene environments along the central Belize coast, Central
America, for the reconstruction of hurricane records. In: International Journal of Earth Sciences (Geologische Rundschau)
106 (1), pp. 283–309. DOI: 10.1007/s00531-016-1319-y.
Alam, E.; Dominey-Howes, D.; Chagué-Goff, C.; Goff, J. (2012): Tsunamis of the northeast Indian Ocean with a particular focus
on the Bay of Bengal region-A synthesis and review. In: Earth-Science Reviews 114 (1-2), pp. 175–193. DOI:
10.1016/j.earscirev.2012.05.002.
Aleixo, R.; Soares-Frazão, S.; Zech, Y. (2011): Velocity-field measurements in a dam-break flow using a PTV Voronoï imaging
technique. In: Experiments In Fluid 50 (6), pp. 1633–1649. DOI: 10.1007/s00348-010-1021-y.
Alfaro, E.; Holz, M. (2014): Seismic geomorphological analysis of deepwater gravity-driven deposits on a slope system of the
southern Colombian Caribbean margin. In: Marine and Petroleum Geology 57, pp. 294–311. DOI: 10.1016/j.mar-
petgeo.2014.06.002.
Allsop, W.; Chandler, I.; Zaccaria, M. (2014): Improvements in the physical modelling of tsunamis and their effects. In: Pro-
ceedings of the 5th International Conference on The Application of Physical Modelling to Port and Coastal Protection,
Coastlab14, Varna, Bulgaria, 29 September – 2 October 2014.
Anglin, D.; MacIntosh, K. (2005): Glass window bridge & causeway project, Bahamas. In: Proceedings of the International
Conference on Coastlines, Structures and Breakwaters 2005, pp. 363–373. DOI: 10.1680/csab2005hsad.34556.0035.
Arnason, H. (2005): Interactions between an incident bore and a free-standing coastal structure. PhD thesis, University of
Washington, Seattle, Washington, USA. https://faculty.washington.edu/cpetroff/Halldor%20dissertation.pdf [Access
date: 22.08.2021].
Arumugasamy, S. (2018): Physikalische Modellversuche zu tsunamiinduziertem Bouldertransport. Master thesis, Institute of
Hydraulic Engineering and Water Resources Management, Faculty of Civil Engineering, RWTH Aachen University. (un-
published)
Atwater, B.F. (1987): Evidence for great holocene earthquakes along the outer coast of washington state. In: Science (New
York, N.Y.) 236 (4804), pp. 942–944. DOI: 10.1126/science.236.4804.942.
Atwater, B. F.; ten Brink, U. S.; Buckley, M.; Halley, R. S.; Jaffe, B. E.; López-Venegas, A. M.; Reinhardt, E. G.; Tuttle, M. P.; Watt,
S.; Wei, Y. (2012): Geomorphic and stratigraphic evidence for an unusual tsunami or storm a few centuries ago at
Anegada, British Virgin Islands. In: Natural Hazards 63 (1), pp. 51–84. DOI: 10.1007/s11069-010-9622-6.
Atwater, B.F.; Cisternas, M.; Yulianto, E.; Prendergast, A.L.; Jankaew, K.; Eipert, A.A.; Starin Fernando, W.I.; Tejakusuma, I.;
Schiappacasse, I.; Sawai, Y. (2013a): The 1960 tsunami on beach-ridge plains near maullín, chile: Landward descent,
renewed breaches, aggraded fans, multiple predecessors [El tsunami de 1960 en una planicie de cordones litorales cerca
de maullín, chile: Descenso tierra adentro, surcos renovados, abanicos agradados, múltiples predecesores]. In: Andean
Geology 40 (3), pp. 393–418. DOI: 10.5027/andgeoV40n3-a01.
Atwater, B. F.; ten Brink, U. S.; Fuentes, Z.; Halley, R. B.; Spiske, M.; Tuttle, M. P.; Wei, Y. (2013b): Further Evidence for Medieval
Faulting along the Puerto Rico Trench. In: AGU Fall Meeting Abstracts, 2013, NH31A-1591.
Atwater, B. F.; Fuentes, Z.; Halley, R. B.; ten Brink, U. S.; Tuttle, M. P. (2014): Effects of 2010 Hurricane Earl amidst geologic
evidence for greater overwash at Anegada, British Virgin Islands. In: Advances in Geosciences 38, pp. 21–30. DOI:
10.5194/adgeo-38-21-2014.
Audemard, F. A. (2007): Revised seismic history of the El Pilar fault, Northeastern Venezuela, from the Cariaco 1997 earth-
quake and recent preliminary paleoseismic results. In: Journal of Seismology 11 (3), pp. 311–326. DOI: 10.1007/s10950-
007-9054-2.
Babu, N.; Babu, D.S.S.; Das, P.N.M. (2007): Impact of tsunami on texture and mineralogy of a major placer deposit in south-
west coast of India. In: Environmental Geology 52 (1), pp. 71–80. DOI: 10.1007/s00254-006-0450-7.
Bagnold, R.A. (1940): Beach formation by waves: some model experiments in a wave tank. (includes photographs). In: Journal
of the Institution of Civil Engineers 15 (1), pp. 27–52. DOI: 10.1680/ijoti.1940.14279.

177
References

Bahlburg, H.; Weiss, R. (2007): Sedimentology of the December 26, 2004, Sumatra tsunami deposits in eastern India (Tamil
Nadu) and Kenya. In: International Journal of Earth Sciences (Geologische Rundschau) 96 (6), pp. 1195–1209. DOI:
10.1007/s00531-006-0148-9.
Bahlburg, H.; Spiske, M. (2012): Sedimentology of tsunami inflow and backflow deposits: Key differences revealed in a mod-
ern example. In: Sedimentology 59 (3), pp. 1063–1086. DOI: 10.1111/j.1365-3091.2011.01295.x.
Barbano, M.S.; Pirrotta, C.; Gerardi, F. (2010): Large boulders along the south-eastern Ionian coast of Sicily: Storm or tsunami
deposits? In: Marine Geology 275 (1-4), pp. 140–154. DOI: 10.1016/j.margeo.2010.05.005.
Barkan, R.; ten Brink, U. S.; Lin, J. (2009): Far field tsunami simulations of the 1755 Lisbon earthquake: Implications for tsunami
hazard to the U.S. East Coast and the Caribbean. In: Marine Geology 264 (1-2), pp. 109–122. DOI: 10.1016/j.mar-
geo.2008.10.010.
Barkan, R.; ten Brink, U.S. (2010): Tsunami simulations of the 1867 Virgin Island earthquake: Constraints on epicenter location
and fault parameters. In: Bulletin of the Seismological Society of America 100 (3), pp. 995–1009. DOI:
10.1785/0120090211.
Barra, R.; Cisternas, M.; Suarez, C.; Araneda, A.; Piñones, O.; Popp, P. (2004): PCBs and HCHs in a salt-marsh sediment record
from South-Central Chile: Use of tsunami signatures and 137Cs fallout as temporal markers. In: Chemosphere 55 (7), pp.
965–972. DOI: 10.1016/j.chemosphere.2003.12.006.
Becker, K.; Gronz, O.; Wirtz, S.; Seeger, M.; Brings, C.; Iserloh, T.; Casper, M. C.; Ries, J. B. (2015): Characterization of complex
pebble movement patterns in channel flow – a laboratory study. In: Cuadernos de Investigación Geográfica 41 (1), p. 63.
DOI: 10.18172/cig.2645.
Benner, R.; Browne, T.; Brückner, H.; Kelletat, D.; Scheffers, A. (2010): Boulder transport by waves: Progress in physical mod-
eling. In: Zeitschrift fur Geomorphologie 54 (SUPPL. 3), pp. 127–146. DOI: 10.1127/0372-8854/2010/0054S3-0000.
Bertran, P.; Bonnissent, D.; Imbert, D.; Lozouet, P.; Serrand, N.; Stouvenot, C. (2004): Caribbean palaeoclimates since 4000 BP:
The Grand-Case Lake record at Saint Martin. In: Comptes Rendus - Geoscience 336 (16), pp. 1501–1510. DOI:
10.1016/j.crte.2004.09.009.
Beyer, R.P.; LeVeque, R.J. (1992): Analysis of a one-dimensional model for the immersed boundary method. In: SIAM Journal
on Numerical Analysis 29 (2), pp. 332–364. DOI: https://doi.org/10.1137/0729022.
Bilham, R. (2010): Lessons from the Haiti earthquake. In: Nature 463 (7283), pp. 878–879. DOI: 10.1038/463878a.
Biolchi, S.; Furlani, S.; Antonioli, F.; Baldassini, N.; Causon Deguara, J.; Devoto, S.; Di Stefano, A.; Evans, J.; Gambin, T.; Gauci,
R.; Mastronuzzi, G.; Monaco, C.; Scicchitano, G. (2016): Boulder accumulations related to extreme wave events on the
eastern coast of Malta. In: Natural hazards and earth system sciences 16 (3), pp. 737–756. DOI: 10.5194/nhess-16-737-
2016.
Bishop, P.; Sanderson, D.; Hansom, J.; Chaimanee, N. (2005): Age-dating of tsunami deposits: Lessons from the 26 December
2004 tsunami in Thailand. In: Geographical Journal 171 (4), pp. 379–384. DOI: 10.1111/j.1475-4959.2005.00175_4.x.
Blair, T.C.; McPherson, J.G. (1999): Grain-size and textural classification of coarse sedimentary particles. In: Journal of Sedi-
mentary Research 69 (1), pp. 6–19. DOI: 10.2110/jsr.69.6.
Blume, H. (1962): Beiträge zur Klimatologie Westindiens. In: Erdkunde 16, pp. 271–289.
Blumenstock, D.I.; Fosberg, F.R.; Johnson, C.G. (1961): The re-survey of typhoon effects on Jaluit Atoll in the Marshall Islands.
In: Nature 189 (4765), pp. 618–620. DOI: 10.1038/189618a0.
Boesl, F.; Engel, M.; Eco, R.C.; Galang, J.B.; Gonzalo, L.A.; Llanes, F.; Quix, E.; Brückner, H. (2020): Digital mapping of coastal
boulders – high-resolution data acquisition to infer past and recent transport dynamics. In: Sedimentology 67 (3), pp.
1393–1410. DOI: 10.1111/sed.12578.
Bourgeois, J.; Johnson, S.Y. (2001): Geologic evidence of earthquakes at the Snohomish delta, Washington, in the past 1200
yr. In: Bulletin of the Geological Society of America 113 (4), pp. 482–494. DOI: 10.1130/0016-7606(2001)113<0482:GE-
OEAT>2.0.CO;2.
Bourgeois, J.; Macinnes, B. (2010): Tsunami boulder transport and other dramatic effects of the 15 november 2006 central
kuril islands Tsunami on the Island of Matua. In: Zeitschrift fur Geomorphologie 54, pp. 175–195. DOI: 10.1127/0372-
8854/2010/0054S3-0024.
Bourrouilh-Le Jan, F.G.; Talandier, J. (1985): Sedimentation et fracturation de haute energie en milieu recifal: Tsunamis, oura-
gans et cyclones et leurs effets sur la sedimentologie et la geomorphologie d’un atoll: Motu et hoa, a rangiroa, Tuamotu,
Pacifique SE. In: Marine Geology 67 (3-4), 263-272,277-323,327-333. DOI: 10.1016/0025-3227(85)90095-7.
Boyajian, G.E.; Thayer, C.W. (1995): Clam calamity: a recent supratidal storm-deposit as an analog for fossil shell beds. In:
Palaios 10 (5), pp. 484–489. DOI: 10.2307/3515050.
Bressan, L.; Antonini, A.; Gaeta, M.G.; Guerrero, M.; Miani, M.; Petruzzelli, V.; Samaras, A. (2015): Boulder transport by tsuna-
mis: A laboratory experiment on incipient motion. In: EGU General Assembly Conference Abstracts 2015, p. 935.
Bressan, L.; Guerrero, M.; Antonini, A.; Petruzzelli, V.; Archetti, R.; Lamberti, A.; Tinti, S. (2018): A laboratory experiment on
the incipient motion of boulders by high-energy coastal flows. In: Earth Surface Processes and Landforms 43 (14), pp.
2935–2947. DOI: 10.1002/esp.4461.

178
References

Brill, D.; Klasen, N.; Jankaew, K.; Brückner, H.; Kelletat, D.; Scheffers, A.; Scheffers, S. (2012): Local inundation distances and
regional tsunami recurrence in the Indian Ocean inferred from luminescence dating of sandy deposits in Thailand. In:
Natural hazards and earth system sciences 12 (7), pp. 2177–2192. DOI: 10.5194/nhess-12-2177-2012.
Brill, D.; May, S.M.; Engel, M.; Reyes, M.; Pint, A.; Opitz, S.; Dierick, M.; Gonzalo, L. A.; Esser, S.; Brückner, H. (2016): Typhoon
Haiyan’s sedimentary record in coastal environments of the Philippines and its palaeotempestological implications. In:
Natural Hazards and Earth System Sciences Discussion, pp. 1–36.
Brooks, G.R.; Devine, B.; Larson, R. A.; Rood, B.P. (2007): Sedimentary development of coral Bay, St. John, USVI: A shift from
natural to anthropogenic influences. In: Caribbean Journal of Science 43 (2), pp. 226–243. DOI: 10.18475/cjos.v43i2.a8.
Brooks, G.R.; Larson, R.A.; Devine, B.; Schwing, P.T. (2015): Annual to millennial record of sediment delivery to US Virgin
Island coastal environments. In: Holocene 25 (6), pp. 1015–1026. DOI: 10.1177/0959683615575357.
Brown, A.L.; Reinhardt, E.G.; van Hengstum, P.J.; Pilarczyk, J. E. (2014): A coastal yucatan sinkhole records intense hurricane
events. In: Journal of Coastal Research 30 (2), pp. 418–428. DOI: 10.2112/JCOASTRES-D-13-00069.1.
Brückner, H. (2000): Küsten - sensible Geo- und Ökosysteme unter zunehmendem Stress. In: Petermanns Geographische
Mitteilungen (143), pp. 6–19.
Brunet, M.; Le Friant, A.; Boudon, G.; Lafuerza, S.; Talling, P.; Hornbach, M.; Ishizuka, O.; Lebas, E.; Guyard, H. (2015): Compo-
sition, geometry, and emplacement dynamics of a large volcanic island landslide offshore Martinique: From volcano
flank-collapse to seafloor sediment failure? In: Geochemistry, Geophysics, Geosystems 17 (3), pp. 699–724. DOI:
10.1002/2015GC006034.
Buckingham, E. (1914): On Physically Similar Systems; Illustrations of the Use of Dimensional Equations. In: Physical Review.
4 (4), pp. 345–376. DOI: 10.1103/PhysRev.4.345.
Buckley, M.L.; Wei, Y.; Jaffe, B.E.; Watt, S.G. (2012): Inverse modeling of velocities and inferred cause of overwash that em-
placed inland fields of boulders at Anegada, British Virgin Islands. In: Natural Hazards 63 (1), pp. 133–149. DOI:
10.1007/s11069-011-9725-8.
Bujan, N.; Cox, R. (2020): Maximal Heights of Nearshore Storm Waves and Resultant Onshore Flow Velocities. In: Frontiers
in Marine Science 7. DOI: 10.3389/fmars.2020.00309.
Burroughs, S.M.; Tebbens, S.F. (2005): Power-law scaling and probabilistic forecasting of tsunami runup heights. In: Pure and
applied geophysics 162 (2), pp. 331–342. DOI: 10.1007/s00024-004-2603-5.
Byrne, D.B.; Suarez, G.; Mccann, W.R. (1985): Muertos trough subduction - Microplate tectonics in the Northern Caribbean?
In: Nature 317 (6036), pp. 420–421. DOI: 10.1038/317420a0.
Caffrey, M.A.; Horn, S.P.; Orvis, K.H.; Haberyan, K.A. (2015): Holocene environmental change at Laguna Saladilla, coastal north
Hispaniola. In: Palaeogeography, Palaeoclimatology, Palaeoecology 436, pp. 9–22. DOI: 10.1016/j.palaeo.2015.06.027.
Cai, S.-G.; Ouahsine, A.; Favier, J.; Hoarau, Y. (2017): Moving immersed boundary method. In: International Journal for Nu-
merical Methods in Fluids 85 (5), pp. 288–323. DOI: 10.1002/fld.4382.
Caron, V. (2011): Contrasted textural and taphonomic properties of high-energy wave deposits cemented in beachrocks (St.
Bartholomew Island, French West Indies). In: Sedimentary Geology 237 (3-4), pp. 189–208. DOI:
10.1016/j.sedgeo.2011.03.002.
Caron, V. (2012): Geomorphic and sedimentologic evidence of extreme wave events recorded by beachrocks: A case study
from the island of St. Bartholomew (Lesser Antilles). In: Journal of Coastal Research 28 (4), pp. 811–828. DOI:
10.2112/JCOASTRES-D-10-00152.1.
Caviedes, C. N. (1991): Five hundred years of hurricanes in the Caribbean: Their relationship with global climatic variabilities.
In: GeoJournal 23 (4), pp. 301–310. DOI: 10.1007/BF00193603.
CDMC (2011): Report of the Committee for Technical Investigation on Countermeasures for Earthquakes and Tsunamis
Based on the Lessons Learned from the “2011 off the Pacific coast of Tohoku Earthquake”. Central Disaster Management
Council, Japan.
Celik, A.O.; Diplas, P.; Dancey, C.L.; Valyrakis, M. (2010): Impulse and particle dislodgement under turbulent flow conditions.
In: Physics of Fluids 22 (4), p. 46601. DOI: 10.1063/1.3385433.
Chanson, H. (2009): Current knowledge in hydraulic jumps and related phenomena. A survey of experimental results. In:
European Journal of Mechanics - B/Fluids 28 (2), pp. 191–210. DOI: 10.1016/j.euromechflu.2008.06.004.
Chen, C.; Melville, B.W.; Nandasena, N.A.K.; Shamseldin, A.Y.; Wotherspoon, L. (2016): Experimental study of uplift loads due
to tsunami bore impact on a wharf model. In: Coastal Engineering 117, pp. 126–137. DOI:
10.1016/j.coastaleng.2016.08.001.
Cheng, N.-S.; Chiew, Y.-M. (1998): Pickup probability for sediment entrainment. In: Journal of Hydraulic Engineering 124 (2),
pp. 232–235. DOI: 10.1061/(ASCE)0733-9429(1998)124:2(232).
Chenoweth, M.; Divine, D. (2008): A document-based 318-year record of tropical cyclones in the Lesser Antilles, 1690-2007.
In: Geochemistry, Geophysics, Geosystems 9 (8). DOI: 10.1029/2008GC002066.
Chock, Gary Y.K. (2016): Design for tsunami loads and effects in the ASCE 7-16 standard. In: Journal of Structural Engineering
142 (11), p. 4016093. DOI: 10.1061/(ASCE)ST.1943-541X.0001565.

179
References

Colón, S.; Audemard, F. A.; Beck, C.; Avila, J.; Padrón, C.; Batist, M. de; Paolini, M.; Leal, A. F.; van Welden, A. (2015): The 1900
Mw 7.6 earthquake offshore north-central Venezuela: Is La Tortuga or San Sebastián the source fault? In: Marine and
Petroleum Geology 67, pp. 498–511. DOI: 10.1016/j.marpetgeo.2015.06.005.
Constant, B.; Péron, S.; Beaugendre, H.; Benoit, C. (2021): An improved immersed boundary method for turbulent flow sim-
ulations on Cartesian grids. In: Journal of Computational Physics 435, p. 110240. DOI: 10.1016/j.jcp.2021.110240.
Corral, A.; Ossó, A.; Llebot, J. E. (2010): Scaling of tropical-cyclone dissipation. In: Nature Physics 6 (9), pp. 693–696. DOI:
10.1038/nphys1725.
Costa, J. E. (1983): Paleohydraulic reconstruction of flash- flood peaks from boulder deposits in the Colorado Front Range.
In: Geological Society of America Bulletin 94 (8), pp. 986–1004. DOI: 10.1130/0016-
7606(1983)94<986:PROFPF>2.0.CO;2.
Costa, P.J.M.; Andrade, C.; Freitas, M.C.; Oliveira, M.A.; Silva, C.M.; Omira, R.; Taborda, R.; Baptista, M.A.; Dawson, A.G. (2011):
Boulder deposition during major tsunami events. In: Earth Surface Processes and Landforms 36 (15), pp. 2054–2068.
DOI: 10.1002/esp.2228.
Costa, P.J.M.; Andrade, C.; Dawson, A. G.; Mahaney, W. C.; Freitas, M. C.; Paris, R.; Taborda, R. (2012): Microtextural charac-
teristics of quartz grains transported and deposited by tsunamis and storms. In: Sedimentary Geology 275-276, pp. 55–
69. DOI: 10.1016/j.sedgeo.2012.07.013.
Cox, R.T.; Lumsden, D.N.; Gough, K.; Lloyd, R.; Talnagi, J. (2008): Investigation of late Quaternary fault block uplift along the
Motagua/Swan Islands fault system: Implications for seismic/tsunami hazard for the Bay of Honduras. In: Tectonophysics
457 (1-2), pp. 30–41. DOI: 10.1016/j.tecto.2008.05.014.
Cox, R.; Zentner, D.B.; Kirchner, B.J.; Cook, M.S. (2012): Boulder ridges on the Aran Islands (Ireland): Recent movements
caused by storm waves, not tsunamis. In: Journal of Geology 120 (3), pp. 249–272. DOI: 10.1086/664787.
Cox, R.; Jahn, K.L.; Watkins, O.G. (2016): Movement of boulders and megagravel by storm waves. In: Geophysical Research
Abstracts 18, EGU2016-EG10535.
Cox, R.; Jahn, K.L.; Watkins, O.G.; Cox, P. (2018a): Extraordinary boulder transport by storm waves (west of Ireland, winter
2013–2014), and criteria for analysing coastal boulder deposits. In: Earth-Science Reviews 177, pp. 623–636. DOI:
10.1016/j.earscirev.2017.12.014.
Cox, R.; Lopes, W.A.; Jahn, K.L. (2018b): Quantitative roundness analysis of coastal boulder deposits. In: Marine Geology 396,
pp. 114–141. DOI: 10.1016/j.margeo.2017.03.003.
Cox, R.; O’boyle, L.; Cytrynbaum, J. (2019): Imbricated coastal boulder deposits are formed by storm waves, and can preserve
a long-term storminess record. In: Scientific Reports 9 (1). DOI: 10.1038/s41598-019-47254-w.
Cox, R.; Ardhuin, F.; Dias, F.; Autret, R.; Beisiegel, N.; Earlie, C.S.; Herterich, J.G.; Kennedy, A.; Paris, R.; Raby, A.; Schmitt, P.;
Weiss, R. (2020): Systematic Review Shows That Work Done by Storm Waves Can Be Misinterpreted as Tsunami-Related
Because Commonly Used Hydrodynamic Equations Are Flawed. In: Frontiers in Marine Science 7. DOI:
10.3389/fmars.2020.00004.
Cuven, S.; Paris, R.; Falvard, S.; Miot-Noirault, E.; Benbakkar, M.; Schneider, J.-L.; Billy, I. (2013): High-resolution analysis of a
tsunami deposit: Case-study from the 1755 lisbon tsunami in southwestern spain. In: Marine Geology 337, pp. 98–111.
DOI: 10.1016/j.margeo.2013.02.002.
Cytrynbaum, J. (2018): Can Storm Waves Move Very Large Boulders? Investigations in a Wave Tank. Thesis, Williams College,
Williamstown, Massachusetts (unpublished).
Dall’Osso, F.; Dominey-Howes, D. (2010): Public assessment of the usefulness of "draft" tsunami evacuation maps from
Sydney, Australia -implications for the establishment of formal evacuation plans. In: Natural Hazards and Earth System
Science 10 (8), pp. 1739–1750. DOI: 10.5194/nhess-10-1739-2010.
Dalman, M.R.; Park, L.E.: Tracking hurricane and climate change records in a Bahamian coastal lake: clear pond, San Salvador
island, Bahamas. In: Proceedings The 15th Symposium on the Geology of the Bahamas and other Carbonate Regions
2012.
Davies, H.L.; Davies, J.M.; Perembo, R.C.B.; Lus, W.Y. (2003): The Aitape 1998 Tsunami: Reconstructing the event from inter-
views and field mapping. In: Pure and applied geophysics 160 (10-11), pp. 1895–1922. DOI: 10.1007/s00024-003-2413-
1.
Dawson, A.G; Long, D.; Smith, D.E (1988): The Storegga Slides: Evidence from eastern Scotland for a possible tsunami. In:
Marine Geology 82 (3-4), pp. 271–276. DOI: 10.1016/0025-3227(88)90146-6.
Dawson, S. (2007): Diatom biostratigraphy of tsunami deposits: Examples from the 1998 Papua New Guinea tsunami. In:
Sedimentary Geology 200 (3-4), pp. 328–335. DOI: 10.1016/j.sedgeo.2007.01.011.
de Buisonjé, P. H. (1974): Neogene and Quaternary geology of Aruba, Curaçao and Bonaire (Netherlands Antilles). In: Neo-
gene and Quaternary Geology of Aruba, Curaçao and Bonaire.
Deltares (2014a): Delft3D-FLOW: Simulation of multi-dimensional hydrodynamic flows and transport phenomena, including
sediments. Deltares, Delft, Netherlands.
Deltares (2014b): DelftDashboard. Deltares, Delft, Netherlands. https://publicwiki.deltares.nl/display/OET/DelftDashboard
[Access date: 01.05.2015].

180
References

Deng, Z.D.; Lu, J.; Myjak, M.J.; Martinez, J.J.; Tian, C.; Morris, S.J.; Carlson, T.J.; Zhou, D.; Hou, H. (2014): Design and implemen-
tation of a new autonomous sensor fish to support advanced hydropower development. In: The Review of scientific
instruments 85 (11), p. 115001. DOI: 10.1063/1.4900543.
Denommee, K.C.; Bentley, S.J.; Droxler, A.W. (2014): Climatic controls on hurricane patterns: A 1200-y near-annual record
from Lighthouse Reef, Belize. In: Scientific reports 4. DOI: 10.1038/srep03876.
Deplus, C.; Le Friant, A.; Boudon, G.; Komorowski, J.-C.; Villemant, B.; Harford, C.; Ségoufin, J.; Cheminée, J.-L. (2001): Subma-
rine evidence for large-scale debris avalanches in the Lesser Antilles Arc. In: Earth and Planetary Science Letters 192 (2),
pp. 145–157. DOI: 10.1016/S0012-821X(01)00444-7.
Dey, S.; Ali, S.Z. (2018): Review Article: Advances in modeling of bed particle entrainment sheared by turbulent flow. In:
Physics of Fluids 30 (6). DOI: 10.1063/1.5030458.
Diplas, P.; Dancey, C.L.; Celik, A.O.; Valyrakis, M.; Greer, K.; Akar, T. (2008): The role of impulse on the initiation of particle
movement under turbulent flow conditions. In: Science (New York, N.Y.) 322 (5902), pp. 717–720. DOI: 10.1126/sci-
ence.1158954.
Dix, G.R.; Patterson, R.T.; Park, L.E. (1999): Marine saline ponds as sedimentary archives of late Holocene climate and sea-
level variation along a carbonate platform margin: Lee Stocking Island, Bahamas. In: Palaeogeography, Palaeoclimatol-
ogy, Palaeoecology 150 (3-4), pp. 223–246. DOI: 10.1016/S0031-0182(98)00184-9.
Donato, S.V.; Reinhardt, E.G.; Boyce, J.I.; Rothaus, R.; Vosmer, T. (2008): Identifying tsunami deposits using bivalve shell ta-
phonomy. In: Geology 36 (3), pp. 199–202. DOI: 10.1130/G24554A.1.
Donnelly, J.P. (2005): Evidence of past intense tropical cyclones from backbarrier salt pond sediments: A case study from Isla
de Culebrita, Puerto Rico, USA. In: Journal of Coastal Research 21 (SPEC. ISS. 42), pp. 201–210.
Donnelly, J.P.; Woodruff, J.D. (2007): Intense hurricane activity over the past 5,000 years controlled by El Niño and the West
African monsoon. In: Nature 447 (7143), pp. 465–468. DOI: 10.1038/nature05834.
Donovan, S.K. (1994): Northern South America. In: Donovan, S.K.; Jackson, T.A. (Eds.): Caribbean Geology: An Introduction,
pp. 229–247. The University of the West Indies Publishers Association, Jamaica. ISBN: 976-41-0033-3.
Dorschner, B.; Colonius, T. (2020): Immersed Boundary Projection Methods. In: Somnath, R.; Ashoke, D.; Elias B. (Eds.): Im-
mersed Boundary Method: Development and Applications. Singapore: Springer Singapore, pp. 3–43. ISBN: 978-981-15-
3940.
Douka, K.; Hedges, R.E.M.; Higham, T.F.G. (2010): Improved AMS 14C dating of shell carbonates using high-precision X-ray
diffraction and a novel density separation protocol (CarDS). In: Radiocarbon 52 (2), pp. 735–751. DOI:
10.1017/S0033822200045756.
Draper, G.; Jackson, T. A.; Donovan, S. K. (1994): Geologic provinces of the Caribbean Region. In: Donovan, S.K.; Jackson, T.A.
(Eds.): Caribbean Geology: An Introduction. The University of the West Indies Publishers Association, Jamaica, pp. 3–12.
ISBN: 976-41-0033-3.
Drenkhan, F.; Huggel, C.; Guardamino, L.; Haeberli, W. (2019): Managing risks and future options from new lakes in the
deglaciating Andes of Peru: The example of the Vilcanota-Urubamba basin. In: Science of the Total Environment 665,
pp. 465–483. DOI: 10.1016/j.scitotenv.2019.02.070.
Dura, T.; Hemphill-Haley, E.; Sawai, Y.; Horton, B.P. (2016): The application of diatoms to reconstruct the history of subduction
zone earthquakes and tsunamis. In: Earth-Science Reviews 152, pp. 181–197. DOI: 10.1016/j.earscirev.2015.11.017.
Einstein, H. A. (1950): The bed-load function for sediment transportation in open channel flows. In: Technical Bulletin (1026).
Engel, M.; Bolten, A.; Brückner, H.; Daut, G.; Kelletat, D.; Schäbitz, F.; Scheffers, A.; Scheffers, S. R.; Vött, A.; Wille, M.; Will-
ershäuser, T. (2009): Reading the chapter of extreme wave events in nearshore geo-bio-archives of Bonaire (Netherlands
antilles)-initial results from Lagun and Boka Bartol. In: Marburger Geographische Schriften 145, pp. 157–178.
Engel, M.; Brückner, H.; Wennrich, V.; Scheffers, A.; Kelletat, D.; Vött, A.; Schäbitz, F.; Daut, G.; Willershäuser, T.; May, S.M.
(2010): Coastal stratigraphies of eastern Bonaire (Netherlands Antilles): New insights into the palaeo-tsunami history of
the southern Caribbean. In: Sedimentary Geology 231 (1-2), pp. 14–30. DOI: 10.1016/j.sedgeo.2010.08.002.
Engel, M.; Brückner, H. (2011): The identification of palaeo-tsunami deposits - A major challenge in coastal sedimentary
research. In: Karius, V., Hadler, H., Deicke, M., von Eynatten, H., Brückner, H., Vött, A. (Eds.), Dynamische Küsten Grund-
lagen, Zusammenhänge und Auswirkungen im Spiegel angewandter Küstenforschung. Proceedings of the 28th Annual
Meeting of the German Working Group on Geography of Oceans and Coasts, 22–25 Apr 2010, Hallig Hooge. Coastline
Reports 17, pp. 65–80.
Engel, M.; May, S.M. (2012): Bonaire’s boulder fields revisited: Evidence for Holocene tsunami impact on the Leeward Antilles.
In: Quaternary Science Reviews 54, pp. 126–141. DOI: 10.1016/j.quascirev.2011.12.011.
Engel, M.; Brückner, H.; Messenzehl, K.; Frenzel, P.; May, S.M.; Scheffers, A.; Scheffers, S.; Wennrich, V.; Kelletat, D. (2012):
Shoreline changes and high-energy wave impacts at the leeward coast of Bonaire (Netherlands Antilles). In: Earth, plan-
ets and space 64 (10), pp. 905–921. DOI: 10.5047/eps.2011.08.011.
Engel, M.; Brückner, H.; Fürstenberg, S.; Frenzel, P.; Konopczak, A.M.; Scheffers, A.; Kelletat, D.; May, S.M.; Schäbitz, F.; Daut,
G. (2013): A prehistoric tsunami induced long-lasting ecosystem changes on a semi-arid tropical island - The case of
Boka Bartol (Bonaire, Leeward Antilles). In: Die Naturwissenschaften 100 (1), pp. 51–67. DOI: 10.1007/s00114-012-0993-
2.

181
References

Engel, M., Brill, D., May, S.M., Reyes, M., Brückner, H. (2014): Philippinen: "Haiyans" Erbe. In: Geographische Rundschau 66
(6), pp. 54–57.
Engel, M., Kindler, P., Godefroid, F. (2015): Interactive comment on “ice melt, sea level rise and superstorms: evidence from
paleoclimate data, climate modeling, and modern observations that 2 °C global warming is highly dangerous” by J.
Hansen et al. In: Atmospheric Chemistry and Physics Discussions (15).
Engel, M.; Oetjen, J.; May, S.M.; Brückner, H. (2016): Tsunami deposits of the Caribbean – Towards an improved coastal
hazard assessment. In: Earth-Science Reviews 163, pp. 260–296. DOI: 10.1016/j.earscirev.2016.10.010.
Engel, M.; May, S.M.; Pilarczyk, J.; Brill, D.; Garrett, E. (2020): Geological records of tsunamis and other extreme waves: con-
cepts, applications and a short history of research. In: Engel, M.; Pilarczyk, J.; May, S.M.; Brill, D.; Garrett, E. (Eds.): Geo-
logical Records of Tsunamis and Other Extreme Waves. Elsevier, pp. 3–20. DOI: https://doi.org/10.1016/B978-0-12-
815686-5.00001-8.
Esteban, M.; Glasbergen, T.; Takabatake, T.; Hofland, B.; Nishizaki, S.; Nishida, Y.; Stolle, J.; Nistor, I.; Bricker, J.; Takagi, H.;
Shibayama, T. (2017): Overtopping of Coastal Structures by Tsunami Waves. In: Geosciences 7 (4), p. 121. DOI:
10.3390/geosciences7040121.
Etienne, S.; Paris, R. (2010): Boulder accumulations related to storms on the south coast of the Reykjanes Peninsula (Iceland).
In: Geomorphology 114 (1-2), pp. 55–70. DOI: 10.1016/j.geomorph.2009.02.008.
Etienne, S.; Buckley, M.; Paris, R.; Nandasena, A.K.; Clark, K.; Strotz, L.; Chagué-Goff, C.; Goff, J.; Richmond, B. (2011): The use
of boulders for characterising past tsunamis: Lessons from the 2004 Indian Ocean and 2009 South Pacific tsunamis. In:
Earth-Science Reviews 107 (1-2), pp. 76–90. DOI: 10.1016/j.earscirev.2010.12.006.
Etienne, S. (2012): Marine inundation hazards in French Polynesia: Geomorphic impacts of tropical cyclone oli in February
2010. In: Geological Society Special Publication 361 (1), pp. 21–39. DOI: 10.1144/SP361.4.
Etienne, S.; Terry, J.P. (2012): Coral boulders, gravel tongues and sand sheets: Features of coastal accretion and sediment
nourishment by Cyclone Tomas (March 2010) on Taveuni Island, Fiji. In: Geomorphology 175-176, pp. 54–65. DOI:
10.1016/j.geomorph.2012.06.018.
Evers, F.M. (2018): Videometric water surface tracking of spatial impulse wave propagation. In: Journal of Visualization 21
(6), pp. 903–907. DOI: 10.1007/s12650-018-0507-1.
Fairbanks, R.G. (1989): A 17,000-year glacio-eustatic sea level record: Influence of glacial melting rates on the Younger Dryas
event and deep-ocean circulation. In: Nature 342 (6250), pp. 637–642. DOI: 10.1038/342637a0.
Fang, C. (2019): An Introduction to Fluid Mechanics. Cham: Springer International Publishing.
Feldens, P.; Schwarzer, K.; Sakuna, D.; Szczuciński, W.; Sompongchaiyakul, P. (2012): Sediment distribution on the inner con-
tinental shelf off Khao Lak (Thailand) after the 2004 Indian Ocean tsunami. In: Earth, planets and space 64 (10), pp. 875–
887. DOI: 10.5047/eps.2011.09.001.
Feuillet, N.; Beauducel, F.; Tapponnier, P. (2011): Tectonic context of moderate to large historical earthquakes in the Lesser
Antilles and mechanical coupling with volcanoes. In: Journal of Geophysical Research: Solid Earth 116 (10). DOI:
10.1029/2011JB008443.
Fitzpatrick, S.M. (2012): On the shoals of giants: Natural catastrophes and the overall destruction of the Caribbean’s archae‐
ological record. In: Journal of Coastal Conservation 16 (2), pp. 173–186. DOI: 10.1007/s11852-010-0109-0.
Focke, J.W. (1978): Limestone cliff morphology on Curaçao (Netherlands Antilles), with special attention to the origin of
notches and vermetid/coralline algal surf benches ("cornices", "trottoirs"). In: Zeitschrift für Geomorphologie 22 (3), pp.
329–349.
Freund, F. (2014): Experimentelle Untersuchung des boreninduzierten Transports von Gesteinsblockmodellen. Bachelor the-
sis. Technische Universität Braunschweig, Braunschweig, Germany. (unpublished)
Friesen (2015): Entwicklung eines kalibrierten numerischen Tsunami Modells für das karibische Meer. Bachelor thesis, Insti-
tute of Hydraulic Engineering and Water Resources Management, Faculty of Civil Engineering, RWTH Aachen University.
(unpublished)
Fritz, H.M.; Hillaire, J.V.; Molière, E.; Wei, Y.; Mohammed, F. (2013): Twin Tsunamis Triggered by the 12 January 2010 Haiti
Earthquake. In: Pure and applied geophysics 170 (9-10), pp. 1463–1474. DOI: 10.1007/s00024-012-0479-3.
Fuentes, Z.; Huérfano-Moreno, V. (2013): Earthquake potential of the Muertos trough: onshore sleuthing for tsunami de-
posits. In: Geological Society of America, Abstracts with Programs 45(2), p. 63.
Gallegos, A. (1996): Descriptive physical oceanography of the Caribbean Sea. In: Small Islands: Marine Science and Sustain-
able Development 51, pp. 36–55.
Garrett, E.; Fujiwara, O.; Garrett, P.; Heyvaert, V.M.A.; Shishikura, M.; Yokoyama, Y.; Hubert-Ferrari, A.; Brückner, H.; Nakamura,
A.; de Batist, M. (2016): A systematic review of geological evidence for Holocene earthquakes and tsunamis along the
Nankai-Suruga Trough, Japan. In: Earth-Science Reviews 159, pp. 337–357. DOI: 10.1016/j.earscirev.2016.06.011.
GEBCO (2014): The GEBCO One Minute Grid. http://www.gebco.net/data_and_products/gridded_bathymetry_data/ [Access
date: 19.07.2021].
Gelfenbaum, G.; Jaffe, B. (2003): Erosion and sedimentation from the 17 July, 1998 Papua New Guinea tsunami. In: Pure and
applied geophysics 160 (10-11), pp. 1969–1999. DOI: 10.1007/s00024-003-2416-y.
Gibbings, J.C. (2011): Dimensional Analysis. London: Springer London. ISBN 978-1-84996-317-6.

182
References

Gienko, G.A.; Terry, J.P. (2014): Three-dimensional modeling of coastal boulders using multi-view image measurements. In:
Earth Surface Processes and Landforms 39 (7), pp. 853–864. DOI: 10.1002/esp.3485.
Gischler, E. (2003): Holocene lagoonal development in the isolated carbonate platforms off Belize. In: Sedimentary Geology
159 (1-2), pp. 113–132. DOI: 10.1016/S0037-0738(03)00098-8.
Gischler, E.; Shinn, E.A.; Oschmann, W.; Fiebig, J.; Buster, N.A. (2008): A 1500-year holocene caribbean climate archive from
the Blue Hole, lighthouse reef, belize. In: Journal of Coastal Research 24 (6), pp. 1495–1505. DOI: 10.2112/07-0891.1.
Goff, J.; Weiss, R.; Courtney, C.; Dominey-Howes, D. (2010): Testing the hypothesis for tsunami boulder deposition from
suspension. In: Marine Geology 277 (1-4), pp. 73–77. DOI: 10.1016/j.margeo.2010.08.003.
Goff, J.; Chagué-Goff, C.; Nichol, S.; Jaffe, B.; Dominey-Howes, D. (2012): Progress in palaeotsunami research. In: Sedimentary
Geology 243-244, pp. 70–88. DOI: 10.1016/j.sedgeo.2011.11.002.
Goff, J.; Terry, J.P.; Chagué-Goff, C.; Goto, K. (2014): What is a mega-tsunami? In: Marine Geology 358, pp. 12–17. DOI:
10.1016/j.margeo.2014.03.013.
Goldenberg, S.B.; Shapiro, L.J. (1996): Physical mechanisms for the association of El Niño and west African rainfall with At-
lantic major hurricane activity. In: Journal of Climate 9 (6), pp. 1169–1187. DOI: 10.1175/1520-
0442(1996)009<1169:PMFTAO>2.0.CO;2.
Gonzalez, L.A., Ruiz, H.M., Taggart, B.E., Budd, A.F., Monell, V. (1997): Geology of Isla de Mona, Puerto Rico. In: In: Vacher,
H.L., Quinn, T. (Eds.): Geology and Hydrogeology of Carbonate Islands. Dev. Sedimentol. (54), pp. 327–358. ISBN: 978-
044-451-644-2.
Gonzalez, L. A.; Ruiz, H. M.; Taggart, B. E.; Budd, A. F.; Monell, V. (2004): Chapter 9 Geology of Isla de Mona, Puerto Rico. In:
Developments in Sedimentology 54 (C), pp. 327–358. DOI: 10.1016/S0070-4571(04)80031-1.
González, C.; Urrego, L. E.; Martínez, J. I.; Polanía, J.; Yokoyama, Y. (2010): Mangrove dynamics in the southwestern Caribbean
since the ’Little Ice Age’: A history of human and natural disturbances. In: Holocene 20 (6), pp. 849–861. DOI:
10.1177/0959683610365941.
Gornak, T. (2013): A goal oriented survey on immersed boundary methods. Fraunhofer-Institut für Techno- und Wirtschafts-
mathematik, ITWM, 2013. http://publica.fraunhofer.de/eprints/urn_nbn_de_0011-n-2664622.pdf [Access date:
22.08.2021].
Goseberg, N. (2011): The run-up of long waves: laboratory-scaled geophysical reproduction and onshore interaction with
macro-roughness elements. PhD dissertation, Gottfried Wilhelm Leibniz Universität, Hannover, Germany. DOI:
https://doi.org/10.15488/7630.
Goseberg, N. (2012): A laboratory perspective of long wave generation. In: Proceedings of the 22nd International Offshore
and Polar Engineering Conference, Rhodes, Greece, 17-22 June 2012.
Goseberg, N.; Wurpts, A.; Schlurmann, T. (2013): Laboratory-scale generation of tsunami and long waves. In: Coastal Engi-
neering 79, pp. 57–74. DOI: 10.1016/j.coastaleng.2013.04.006.
Goseberg, N., Nistor, I., Stolle, J. (2015): Tracking of “Smart” debris location based on the RFID technique. Wallendorf, L.,
Cox, D.T. (Eds.), Coastal Structures and Solutions, American Society of Civil Engineers, Reston, VA. DOI:
https://doi.org/10.1061/9780784480311.
Goseberg, N.; Stolle, J.; Nistor, I.; Shibayama, T. (2016): Experimental analysis of debris motion due the obstruction from fixed
obstacles in tsunami-like flow conditions. In: Coastal Engineering 118, pp. 35–49. DOI: 10.1016/j.coastaleng.2016.08.012.
Goto, K.; Chavanich, S. A.; Imamura, F.; Kunthasap, P.; Matsui, T.; Minoura, K.; Sugawara, D.; Yanagisawa, H. (2007): Distribu-
tion, origin and transport process of boulders deposited by the 2004 Indian Ocean tsunami at Pakarang Cape, Thailand.
In: Sedimentary Geology 202 (4), pp. 821–837. DOI: 10.1016/j.sedgeo.2007.09.004.
Goto, K.; Kawana, T.; Imamura, F. (2010a): Historical and geological evidence of boulders deposited by tsunamis, southern
Ryukyu Islands, Japan. In: Earth-Science Reviews 102 (1-2), pp. 77–99. DOI: 10.1016/j.earscirev.2010.06.005.
Goto, K.; Okada, K.; Imamura, F. (2010b): Numerical analysis of boulder transport by the 2004 Indian Ocean tsunami at
Pakarang Cape, Thailand. In: Marine Geology 268 (1-4), pp. 97–105. DOI: 10.1016/j.margeo.2009.10.023.
Goto, K.; Miyagi, K.; Kawana, T.; Takahashi, J.; Imamura, F. (2011): Emplacement and movement of boulders by known storm
waves - Field evidence from the Okinawa Islands, Japan. In: Marine Geology 283 (1-4), pp. 66–78. DOI: 10.1016/j.mar-
geo.2010.09.007.
Goto, K.; Sugawara, D.; Ikema, S.; Miyagi, T. (2012): Sedimentary processes associated with sand and boulder deposits formed
by the 2011 Tohoku-oki tsunami at Sabusawa Island, Japan. In: Sedimentary Geology 282, pp. 188–198. DOI:
10.1016/j.sedgeo.2012.03.017.
Goto, K.; Fujino, S.; Sugawara, D.; Nishimura, Y. (2014): The current situation of tsunami geology under new policies for
disaster countermeasures in Japan. In: Episodes 37 (4), pp. 258–264. DOI: 10.18814/epiiugs/2014/v37i4/005.
Goto, K.; Ishizawa, T.; Ebina, Y.; Imamura, F.; Sato, S.; Udo, K. (2021): Ten years after the 2011 Tohoku-oki earthquake and
tsunami: Geological and environmental effects and implications for disaster policy changes. In: Earth-Science Reviews
212. DOI: 10.1016/j.earscirev.2020.103417.
Granja Bruña, J.L.; Carbó-Gorosabel, A.; Llanes Estrada, P.; Muñoz-Martín, A.; ten Brink, U.S.; Gómez Ballesteros, M.; Druet,
M.; Pazos, A. (2014): Morphostructure at the junction between the Beata ridge and the Greater Antilles island arc (off-
shore Hispaniola southern slope). In: Tectonophysics 618, pp. 138–163. DOI: 10.1016/j.tecto.2014.02.001.

183
References

Grass, A.J. (1970): Initial Instability of Fine Bed Sand. In: Journal of the Hydraulics Division 96 (3), pp. 619–632. DOI:
10.1061/JYCEAJ.0002369.
Grindlay, N.R.; Hearne, M.; Mann, P. (2005): High risk of tsunami in the Northern Caribbean. In: Eos 86 (12). DOI:
10.1029/2005EO120001.
Gronskis, A.; Artana, G. (2016): A simple and efficient direct forcing immersed boundary method combined with a high order
compact scheme for simulating flows with moving rigid boundaries. In: Computers & Fluids 124, pp. 86–104. DOI:
10.1016/j.compfluid.2015.10.016.
Gronz, O.; Hiller, P. H.; Wirtz, S.; Becker, K.; Iserloh, T.; Seeger, M.; Brings, C.; Aberle, J.; Casper, M. C.; Ries, J. B. (2016): Smart-
stones: A small 9-axis sensor implanted in stones to track their movements. In: CATENA 142, pp. 245–251. DOI:
10.1016/j.catena.2016.03.030.
Gupta, V.P., Goyal, S.C. (1975): Hydrodynamic forces on bridge piers. In: Journal of the Institution of Civil Engineers (56), pp.
12–16.
Hall, A.M.; Hansom, J.D.; Williams, D.M.; Jarvis, J. (2006): Distribution, geomorphology and lithofacies of cliff-top storm de-
posits: Examples from the high-energy coasts of Scotland and Ireland. In: Marine Geology 232 (3-4), pp. 131–155. DOI:
10.1016/j.margeo.2006.06.008.
Hansen, J.; Sato, M.; Hearty, P.; Ruedy, R.; Kelley, M.; Masson-Delmotte, V.; Russell, G.; Tselioudis, G.; Cao, J.; Rignot, E.;
Velicogna, I.; Tormey, B.; Donovan, B.; Kandiano, E.; Schuckmann, K. von; Kharecha, P.; Legrande, A. N.; Bauer, M. (2016):
Ice melt, sea level rise and superstorms: Evidence from paleoclimate data, climate modeling, and modern observations
that 2 °c global warming could be dangerous. In: Atmospheric Chemistry and Physics 16 (6), pp. 3761–3812. DOI:
10.5194/acp-16-3761-2016.
Hansom, J.D.; Barltrop, N.D.P.; Hall, A. M. (2008): Modelling the processes of cliff-top erosion and deposition under extreme
storm waves. In: Marine Geology 253 (1-2), pp. 36–50. DOI: 10.1016/j.margeo.2008.02.015.
Harbitz, C. B.; Glimsdal, S.; Bazin, S.; Zamora, N.; Løvholt, F.; Bungum, H.; Smebye, H.; Gauer, P.; Kjekstad, O. (2012): Tsunami
hazard in the Caribbean: Regional exposure derived from credible worst case scenarios. In: Continental Shelf Research
38, pp. 1–23. DOI: 10.1016/j.csr.2012.02.006.
Harry, S.; Exton, M.; Yeh, H. (2019): Boulder Pickup by Tsunami Surge. In: Journal of Earthquake and Tsunami 13 (05n06).
DOI: 10.1142/S1793431119410069.
Hasler, C.-A.; Simpson, G.; Kindler, P. (2010): Platform margin collapse simulation: The case of the North Eleuthera massive
boulders. In: Abstracts and Program, 15th Symposium on the Geology of the Bahamas and Other Carbonate Regions,
pp. 22–23.
Hawkes, A.D.; Bird, M.; Cowie, S.; Grundy-Warr, C.; Horton, B.P.; Shau Hwai, A.T.; Law, L.; Macgregor, C.; Nott, J.; Ong, J.E.;
Rigg, J.; Robinson, R.; Tan-Mullins, M.; Sa, T.T.; Yasin, Z.; Aik, L.W. (2007): Sediments deposited by the 2004 Indian Ocean
Tsunami along the Malaysia-Thailand Peninsula. In: Marine Geology 242 (1-3), pp. 169–190. DOI: 10.1016/j.mar-
geo.2007.02.017.
Hawkes, A.D.; Horton, B.P. (2012): Sedimentary record of storm deposits from Hurricane Ike, Galveston and San Luis Islands,
Texas. In: Geomorphology 171-172, pp. 180–189. DOI: 10.1016/j.geomorph.2012.05.017.
Hayes, G.P.; McNamara, D.E.; Seidman, L.; Roger, J. (2014): Quantifying potential earthquake and tsunami hazard in the Lesser
Antilles subduction zone of the Caribbean region. In: Geophysical Journal International 196 (1), pp. 510–521. DOI:
10.1093/gji/ggt385.
Hearty, P.J. (1997): Boulder Deposits from Large Waves during the Last Interglaciation on North Eleuthera Island, Bahamas.
In: Quaternary research 48 (3), pp. 326–338. DOI: 10.1006/qres.1997.1926.
Hearty, P.J.; Neumann, A.C.; Kaufman, D.S. (1998): Chevron ridges and runup deposits in the Bahamas from storms late in
oxygen-isotope substage 5e. In: Quaternary research 50 (3), pp. 309–322. DOI: 10.1006/qres.1998.2006.
Hearty, P. J.; Neumann, A. C. (2001): Rapid sea level and climate change at the close of the Last Interglaciation (MIS 5e):
Evidence from the Bahama Islands. In: Quaternary Science Reviews 20 (18), pp. 1881–1895. DOI: 10.1016/S0277-
3791(01)00021-X.
Hearty, P.J.; Tormey, B.R.; Neumann, A.C. (2002): Discussion of "Palaeoclimatic significance of co-occurring wind- and water-
induced sedimentary structures in the last-interglacial coastal deposits from Bermuda and the Bahamas. In: Sedimentary
Geology 147, pp. 429–435.
Heller, V. (2011): Scale effects in physical hydraulic engineering models. In: Journal of Hydraulic Research 49 (3), pp. 293–
306. DOI: 10.1080/00221686.2011.578914.
Hemphill-Haley, E. (1996): Diatoms as an aid in identifying late-holocene tsunami deposits. In: Holocene 6 (4), pp. 439–448.
DOI: 10.1177/095968369600600406.
Hendry, M.D. (1987): Tectonic and eustatic control on late Cenozoic sedimentation within an active plate boundary zone,
west coast margin, Jamaica. In: Bulletin of the Geological Society of America 99 (5), pp. 718–728. DOI: 10.1130/0016-
7606(1987)99<718:TAECOL>2.0.CO;2.
Hernandez-Avila, M.L.; Roberts, H.H.; Rouse, L.J. (1977): Hurricane-generated waves and coastal boulder rampart formation.
In: Proceedings of the 3rd International Coral Reef Symposium 2, pp. 71–78.

184
References

Herwig, H. (2008): Strömungsmechanik - Einführung in die Physik von technischen Strömungen. Wiesbaden, Germany:
Vieweg+Teubner Verlag. ISBN 978-3-8348-9497-7.
Hindson, R.A.; Andrade, C. (1999): Sedimentation and hydrodynamic processes associated with the tsunami generated by
the 1755 Lisbon earthquake. In: Quaternary International 56 (1), pp. 27–38. DOI: 10.1016/S1040-6182(98)00014-7.
Hippensteel, S. P.; Eastin, M. D.; Garcia, W. J. (2013): The geological legacy of Hurricane Irene: Implications for the fidelity of
the paleo-storm record. In: GSA Today 23 (12), pp. 4–10. DOI: 10.1130/GSATG184A.1.
Hisamatsu, A.; Goto, K.; Imamura, F. (2014): Local paleo-tsunami size evaluation using numerical modeling for boulder
transport at Ishigaki Island, Japan. In: Episodes 37 (4), pp. 265–276. DOI: 10.18814/epiiugs/2014/v37i4/006.
Hobgood, J. (2005): Tropical cyclones. In: Encyclopedia of Earth Sciences Series, pp. 750–756. DOI: 10.1007/1-4020-3266-
8_213.
Hoerner, S.F. (1965): Fluid-dynamic drag. Hoerner Fluid Dynamics, Bakersfield, CA.
Hoffmeister, D. (2020): Mapping of subaerial coarse clasts. In: Engel, M.; Pilarczyk, J.; May, S.M.; Brill, D.; Garrett, E. (Eds.):
Geological Records of Tsunamis and Other Extreme Waves. Elsevier, pp. 169–184.
Hofman, C.L.; Hoogland, M.L.P. (2015): Beautiful tropical islands in the Caribbean Sea: Human responses to floods and
droughts and the indigenous archaeological heritage of the Caribbean. In: van Schalk, H.P.J; Willems, W.J.H. (Eds.): Water
and Heritage: Material, Conceptual and Spiritual Connections, ICOMOS Netherlands, UNESCO, pp. 99–120. ISBN: 978-
90-8890-278-9.
Hornbach, M.J.; Mann, P.; Wolf, S.; King, W.; Boon, R. (2008a): Assessing slope stability at Seroe Mansinga and Caracas bay,
Curaçao. In: Final report for APNA, Willemstad, Curaçao. The University of Texas, USA, Austin Jackson School of Geosci-
ences, February, 2008.
Hornbach, M.J.; Mondziel, S.A.; Grindlay, N. R.; Frohlich, C.; Mann, P. (2008b): Did a submarine slide trigger the 1918 Puerto
Rico tsanami? In: Science of Tsunami Hazards 27 (2), pp. 22–31.
Hornbach, M.J.; Mann, P.; Taylor, F.W.; Bowen, S.W. (2010): Estimating the age of near-shore carbonate slides using coral
reefs and erosional markers: a case study from Curaçao, Netherlands Antilles. In: The Sedimentary Record 8 (1), pp. 4–
10. DOI: 10.2110/sedred.2010.1.4
Horton, B.P.; Rossi, V.; Hawkes, A.D. (2009): The sedimentary record of the 2005 hurricane season from the Mississippi and
Alabama coastlines. In: Quaternary International 195 (1-2), pp. 15–30. DOI: 10.1016/j.quaint.2008.03.004.
Horton, B.P.; Sawai, Y.; Hawkes, A.D.; Witter, R.C. (2011): Sedimentology and paleontology of a tsunami deposit accompany-
ing the great Chilean earthquake of February 2010. In: Marine Micropaleontology 79 (3-4), pp. 132–138. DOI:
10.1016/j.marmicro.2011.02.001.
Hsu, H.-C.; Torres-Freyermuth, A.; Hsu, T.-J.; Hwung, H.-H. (2012): Numerical and experimental study of dam-break flood
propagation and its implication to sediment erosion. In: Proceedings of the International Conference on Coastal Engi-
neering 1 (33), p. 7. DOI: 10.9753/icce.v33.sediment.7.
Hughes, S. A. (1993): Physical Models and Laboratory Techniques in Coastal Engineering. In: Advanced Series on Ocean
Engineering 7. DOI: https://doi.org/10.1142/2154
Hunt, J.E.; Wynn, R.B.; Talling, P.J.; Masson, D.G. (2013): Multistage collapse of eight western Canary Island landslides in the
last 1.5 Ma: Sedimentological and geochemical evidence from subunits in submarine flow deposits. In: Geochemistry,
Geophysics, Geosystems 14 (7), pp. 2159–2181. DOI: 10.1002/ggge.20138.
Huntington, K.; Bourgeois, J.; Gelfenbaum, G.; Lynett, P.; Jaffe, B.; Yeh, H.; Weiss, R. (2007): Sandy Signs of a Tsunami's On-
shore Depth and Speed. In: Eos, Transactions, American Geophysical Union Volume 88 (52), p. 577. DOI:
10.1029/2007EO520001.
Hussain, S.M.; Krishnamurthy, R.; Suresh Gandhi, M.; Ilayaraja, K.; Ganesan, P.; Mohan, S.P. (2006): Micropalaeontological
investigations on tsunamigenic sediments of Andaman Islands. In: Current Science 91 (12), pp. 1655–1667.
Hussain, S.M.; Mohan, S.P.; Jonathan, M.P. (2010): Ostracoda as an aid in identifying 2004 tsunami sediments: A report from
SE coast of India. In: Natural Hazards 55 (2), pp. 513–522. DOI: 10.1007/s11069-010-9543-4.
Hylla, E.A. (2013). Eine Immersed Boundary Methode zur Simulation von Strömungen in komplexen und bewegten Geomet-
rien. PhD thesis, Faculty V of Mechanical Engineering and Transport Systems, TU Berlin, Universitätsverlag der TU Berlin
2013. ISBN: 978-3-7983-2530-2. https://depositonce.tu-berlin.de/bitstream/11303/3976/2/hylla_eike_alexander_con-
tent.pdf [Accessed on: 26.08.2021].
Iaccarino, G.; Verzicco, R. (2003): Immersed boundary technique for turbulent flow simulations. In: Applied Mechanics Re-
views 56 (3), pp. 331–347. DOI: 10.1115/1.1563627.
Iglberger, K., Rüde, U. (2009): Massively Parallel Rigid Body Dynamics Simulations. In: Computer Science - Research and
Development 23 (3), pp. 159–167. DOI: 10.1007/s00450-009-0066-8.
Imamura, F.; Goto, K.; Ohkubo, S. (2008): A numerical model for the transport of a boulder by tsunami. In: Journal of Geo-
physical Research: Oceans 113 (1). DOI: 10.1029/2007JC004170.
IOC (2012): Exercise Caribe Wave/Lantex 13. A Caribbean Tsunami Warning Exercise, 20 March 2013. Volume 1: Participant
Handbook. IOC Technical Series No. 101. UNESCO. Paris.

185
References

Ishimura, D.; Miyauchi, T. (2015): Historical and paleo-tsunami deposits during the last 4000 years and their correlations with
historical tsunami events in Koyadori on the Sanriku Coast, northeastern Japan. In: Progress in Earth and Planetary
Science 2 (1). DOI: 10.1186/s40645-015-0047-4.
Jackson, L.P. (2013): Caribbean Sea Level Change: Observational Analysis From Millennial to Decadal Timescales. In: Carib-
bean sea level change: Observational analysis from millennial to decadal timescales. PhD dissertation, University of
Leeds, United Kingdom. ISBN: 978-0-85731-605-9.
Jaffe, B.; Gelfenbaum, G.; Rubin, D.; Peters, R.; Anima, R.; Swensson, M.; Olcese, D.; Bernales, L.; Gomez, J.; Riega, P. (2003):
Tsunami deposits: identification and interpretation of tsunami deposits from the June 23, 2001 Perú tsunami. In: Pro-
ceedings of the International Conference on Coastal Sediments 2003. CD-ROM Published by World Scientific Publishing
Corp and East Meets West Productions.
Jaffe, B.; Buckley, M.L.; Richmond, B.M.; Morton, R.A.; Moya, J.C.; Gelfenbaum, G.; Watt, S.G. (2008): Evidence of tsunami in a
coastal pond in NW Puerto Rico. In: AGU Fall Meeting Abstracts 2008.
Jaffe, B.; Goto, K.; Sugawara, D.; Gelfenbaum, G.; La Selle, S.P. (2016): Uncertainty in tsunami sediment transport modeling.
In: Journal of Disaster Research 11 (4), pp. 647–661. DOI: 10.20965/jdr.2016.p0647.
Jagodziński, R.; Sternal, B.; Szczuciński, W.; Chagué-Goff, C.; Sugawara, D. (2012): Heavy minerals in the 2011 Tohoku-oki
tsunami deposits-insights into sediment sources and hydrodynamics. In: Sedimentary Geology 282, pp. 57–64. DOI:
10.1016/j.sedgeo.2012.07.015.
Jankaew, K.; Atwater, B. F.; Sawai, Y.; Choowong, M.; Charoentitirat, T.; Martin, M. E.; Prendergast, A. (2008): Medieval fore-
warning of the 2004 Indian Ocean tsunami in Thailand. In: Nature 455 (7217), pp. 1228–1231. DOI: 10.1038/nature07373.
Jarvis, A., Reuter, H.I., Nelson, A., Guevara, E. (2008): Hole-filled SRTM for the globe Version 4, available from the CGIAR-CSI
SRTM 90m Database. http://srtm.csi.cgiar.org [Access date: 01.05.2015].
Jessen, C.A.; Pedersen, J.B.T.; Bartholdy, J.; Seidenkrantz, M.-S.; Kuijpers, A. (2008): A late holocene palaeoenvironmental
record from Altona Bay, St. Croix, US Virgin Islands. In: Geografisk Tidsskrift 108 (2), pp. 59–70. DOI:
10.1080/00167223.2008.10649589.
Jones, B.; Hunter, I.G. (1992): Very large boulders on the coast of Grand Cayman: The effects of giant waves on rocky coast-
lines. In: Journal of Coastal Research 8 (4), pp. 763–774.
Kafle, J.; Kattel, P.; Mergili, M.; Fischer, J.T.; Pudasaini, S.P. (2019): Dynamic response of submarine obstacles to two-phase
landslide and tsunami impact on reservoirs. In: Acta Mechanica (230). DOI: 10.1007/s00707-019-02457-0.
Kain, C.L.; Gomez, C.; Moghaddam, A.E. (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]. In: Marine Geology 319-322, pp. 75–76. DOI: 10.1016/j.margeo.2011.08.008.
Kain, C.; Wassmer, P.; Goff, J.; Chagué-Goff, C.; Gomez, C.; Hart, D.; Fierro, D.; Jacobsen, G.; Zawadzki, A. (2016): Determining
flow patterns and emplacement dynamics from tsunami deposits with no visible sedimentary structure. In: Earth Surface
Processes and Landforms 42 (5), pp. 763–780. DOI: 10.1002/esp.4020.
Kattel, P.; Kafle, J.; Fischer, J.-T.; Mergili, M.; Tuladhar, B.M.; Pudasaini, S.P. (2018): Interaction of two-phase debris flow with
obstacles. In: Engineering Geology 242, pp. 197–217. DOI: 10.1016/j.enggeo.2018.05.023.
Kelletat, D.; Scheffers, A.; Scheffers, S. (2004): Holocene tsunami deposits on the Bahaman Islands of Long Island and Eleu-
thera. In: Zeitschrift fur Geomorphologie 48 (4), pp. 519–540. DOI: 10.1127/zfg/48/2004/519.
Kelletat, D.; Scheffers, S.; Scheffers, A. (2007): Field signatures of the SE-Asian mega-tsunami along the West Coast of Thai-
land compared to holocene paleo-tsunami from the Atlantic Region. In: Pure and applied geophysics 164 (2-3), pp. 413–
431. DOI: 10.1007/s00024-006-0171-6.
Kelletat, D.; Engel, M.; May, S.M.; Erdmann, W.; Scheffers, A.; Brückner, H. (2020): Erosive impact of tsunami and storm waves
on rocky coasts and post-depositional weathering of coarse-clast deposits. In: Engel, M.; Pilarczyk, J.; May, S.M.; Brill, D.;
Garrett, E. (Eds.): Geological Records of Tsunamis and Other Extreme Waves. Elsevier, pp. 561–584. DOI: 10.1016/B978-
0-12-815686-5.00026-2.
Kennedy, D.M.; Tannock, K.L.; Crozier, M. J.; Rieser, U. (2007): Boulders of MIS 5 age deposited by a tsunami on the coast of
Otago, New Zealand. In: Sedimentary Geology 200 (3), pp. 222–231. DOI: 10.1016/j.sedgeo.2007.01.005.
Kennedy, A.B.; Mori, N.; Zhang, Y.; Yasuda, T.; Chen, S.-E.; Tajima, Y.; Pecor, W.; Toride, K. (2015): Observations and Modeling
of Coastal Boulder Transport and Loading during Super Typhoon Haiyan. In: Coastal Engineering Journal 58 (1). DOI:
10.1142/S0578563416400040.
Kennedy, A.B.; 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. In: Marine Geol-
ogy 383, pp. 65–77. DOI: 10.1016/j.margeo.2016.11.004.
Khan, S.; Robinson, E.; Rowe, D.-A.; Coutou, R. (2010): Size and mass of shoreline boulders moved and emplaced by recent
hurricanes, jamaica. In: Zeitschrift fur Geomorphologie 54, pp. 281–299. DOI: 10.1127/0372-8854/2010/0054S3-0028.
Kilfeather, A.A.; Blackford, J.J.; van der Meer, J.J.M. (2007): Micromorphological analysis of coastal sediments from Willapa
Bay, Washington, USA: A technique for analysing inferred tsunami deposits. In: Pure and applied geophysics 164 (2-3),
pp. 509–525. DOI: 10.1007/s00024-006-0173-4.

186
References

Kindler, P.; Strasser, A. (2000): Palaeoclimatic significance of co-occurring wind- and water-induced sedimentary structures
in the last-interglacial coastal deposits from Bermuda and the Bahamas. In: Sedimentary Geology 131 (1-2), pp. 1–7.
DOI: 10.1016/S0037-0738(99)00123-2.
Kindler, P.; Strasser, A. (2002): Palaeoclimatic significance of co-occurring wind- and water-induced sedimentary structures
in last-interglacial coastal deposits from Bermuda to the Bahamas: Response to Comment. In: Sedimentary Geology
147, pp. 437–443. DOI: 10.1016/S0037-0738(01)00099-9.
Kindler, P.; Mylroie, J.E.; Curran, H.A.; Carew, J.L.; Gamble, D.W.; Rothfus, T.A.; Savarese, M.; Sealey, N.E. (2010): Geology of
central Eleuthera, Bahamas: a field trip guide. In: Proceedings of 15th Symposium on the Geology of the Bahamas and
Other Carbonate Regions 2012, Gerace Research Center, San Salvador, Bahamas.
Klosowska, B. (2003): Late Holocene Embayment and Salina Record of Curaçao (Dutch Antilles): Criteria to Monitor Environ-
mental Change and Biodiversity. PhD dissertation, Vrije Universiteit Amsterdam, The Netherlands.
Klosterhalfen, J. (2016): Tsunamiinduzierter Boulder- und Sedimenttransport – Konzeptionelle Entwicklung eines experimen-
tellen Versuchsaufbaus. Bachelor thesis, Institute of Hydraulic Engineering and Water Resources Management, Faculty
of Civil Engineering, RWTH Aachen University. (unpublished)
Knutson, T.R.; McBride, J.L.; Chan, J.; Emanuel, K.; Holland, G.; Landsea, C.; Held, I.; Kossin, J. P.; Srivastava, A.K.; Sugi, M. (2010):
Tropical cyclones and climate change. In: Nature Geoscience 3 (3), pp. 157–163. DOI: 10.1038/ngeo779.
Kobus, H. (1974): Anwendung der Dimensionsanalyse in der experimentellen Forschung des Bauingenieurwesens. Bautech-
nik 51 (3), pp. 88-94. DOI: 10.18419/opus-8279.
Komatsubara, J.; Fujiwara, O. (2007): Overview of holocene tsunami deposits along the Nankai, Suruga, and Sagami Troughs,
Southwest Japan. In: Pure and applied geophysics 164 (2-3), pp. 493–507. DOI: 10.1007/s00024-007-0179-y.
Kortekaas, S.; Dawson, A.G. (2007): Distinguishing tsunami and storm deposits: An example from Martinhal, SW Portugal. In:
Sedimentary Geology 200 (3-4), pp. 208–221. DOI: 10.1016/j.sedgeo.2007.01.004.
Korup, O.; Clague, J.J. (2009): Natural hazards, extreme events, and mountain topography. In: Quaternary Science Reviews
28 (11-12), pp. 977–990. DOI: 10.1016/j.quascirev.2009.02.021.
Küstermann, F.M. (2016): Tsunamientstehung durch Erdbeben und Erdrutschungen – Konzeptionelle Versuchsentwicklung.
Bachelor thesis, Institute of Hydraulic Engineering and Water Resources Management, Faculty of Civil Engineering,
RWTH Aachen University. (unpublished)
Kumar, M.; Raj, A.; Roy, S. (2020): Mass Conservation in Sharp Interface Immersed Boundary Method—A GPGPU Accelerated
Implementation. In: Roy, S.; De, A.; Balaras, E. (Eds.): Immersed Boundary Method – Development and Applications.
Springer Singapore (Computational Methods in Engineering & the Sciences), pp. 81–106. DOI: 10.1007/978-981-15-
3940-4_3
Lai, M.-C.; Peskin, C.S. (2000): An Immersed Boundary Method with Formal Second-Order Accuracy and Reduced Numerical
Viscosity. In: Journal of Computational Physics 160 (2), pp. 705–719. DOI: 10.1006/jcph.2000.6483.
Lambeck, K. (2002): Sea-level change from mid-Holocene to recent time: an Australian example with global implications. In:
Ice Sheets, Sea Level and the Dynamic Earth 29, pp. 33–50. DOI: 10.1002/9781118670101.ch3.
Lander, J.F., Whiteside, L.S., Lockridge, P.A. (2002): A brief history of tsunamis in the Caribbean Sea. In: Science of Tsunami
Hazards 20, pp. 57–94.
Lario, J.; Luque, L.; Zazo, C.; Goy, J.L.; Spencer, C.; Cabero, A.; Bardají, T.; Borja, F.; Dabrio, C.J.; Civis, J.; González-Delgado, J.Á.;
Borja, C.; Alonso-Azcárate, J. (2010): Tsunami vs. Storm surge deposits: A review of the sedimentological and geomor-
phological records of extreme wave events (EWE) during the holocene in the gulf of cadiz, Spain. In: Zeitschrift fur
Geomorphologie 54, pp. 301–316. DOI: 10.1127/0372-8854/2010/0054S3-0029.
Lario, J.; Zazo, C.; Goy, J.L. (2016): Tectonic and morphosedimentary features of the 2010 Chile earthquake and tsunami in
the Arauco Gulf and Mataquito River (Central Chile). In: Geomorphology 267, pp. 16–24. DOI: 10.1016/j.geo-
morph.2016.05.019.
Larson, R. A.; Brooks, G. R.; Devine, B.; Schwing, P. T.; Holmes, C. W.; Jilbert, T.; Reichart, G.-J. (2015): Elemental signature of
terrigenous sediment runoff as recorded in coastal salt ponds: US Virgin Islands. In: Applied Geochemistry 63, pp. 573–
585. DOI: 10.1016/j.apgeochem.2015.01.008.
Lau, A.Y.A.; Etienne, S.; Terry, J.P.; Switzer, A.D.; Lee, Y.S. (2014): A preliminary study of the distribution, sizes and orientations
of large reef-top coral boulders deposited by extreme waves at Makemo Atoll, French Polynesia. In: Journal of Coastal
Research 70, pp. 272–277. DOI: 10.2112/SI70-046.1.
Lau, A.Y.A.; Terry, J.P.; Switzer, A.D.; Pile, J. (2015): Advantages of beachrock slabs for interpreting high-energy wave transport:
Evidence from Ludao Island in south-eastern Taiwan. In: Geomorphology 228, pp. 263–274. DOI: 10.1016/j.geo-
morph.2014.09.010.
Lauber, G.; Hager, W.H. (1998): Experiments to dambreak wave: Horizontal channel. In: Journal of Hydraulic Research 36 (3),
pp. 291–307. DOI: 10.1080/00221689809498620.
Le Friant, A.; Heinrich, P.; Boudon, G. (2008): Field survey and numerical simulation of the 21 November 2004 tsunami at Les
Saintes (Lesser Antilles). In: Geophysical Research Letters 35 (12). DOI: 10.1029/2008GL034051.

187
References

Le Friant, A.; Boudon, G.; Arnulf, A.; Robertson, R.E.A. (2009): Debris avalanche deposits offshore St. Vincent (West Indies):
Impact of flank-collapse events on the morphological evolution of the island. In: Journal of Volcanology and Geothermal
Research 179 (1-2), pp. 1–10. DOI: 10.1016/j.jvolgeores.2008.09.022.
le Méhauté, B. (1976): An Introduction to Hydrodynamics and Water Waves. Berlin, Heidelberg: Springer Berlin Heidelberg.
ISBN: 978-3-642-85567-2.
Leal, K.; Scremin, L.; Audemard, F.; Carrillo, E. (2014): Paleotsunamis in the geological record of cumaná, sucre state, Eastern
Venezuela [Paleotsunamis en el registro geológico de cumaná, estado sucre, Venezuela oriental]. In: Boletin de Geologia
36 (2), pp. 45–70.
Lebas, E.; Le Friant, A.; Boudon, G.; Watt, S.F.L.; Talling, P.J.; Feuillet, N.; Deplus, C.; Berndt, C.; Vardy, M.E. (2011): Multiple
widespread landslides during the long-term evolution of a volcanic island: Insights from high-resolution seismic data,
Montserrat, Lesser Antilles. In: Geochemistry, Geophysics, Geosystems 12 (5). DOI: 10.1029/2010GC003451.
Leslie, S.C.; Mann, P. (2016): Giant submarine landslides on the Colombian margin and tsunami risk in the Caribbean Sea. In:
Earth and Planetary Science Letters 449, pp. 382–394. DOI: 10.1016/j.epsl.2016.05.040.
Li, Y., Chanson, H. (2017): Free-surface and velocity characteristics of tidal bore propagation against a slope: experiments on
decelarating bores. In: Proceedings of 37th IAHR World Congress, Kuala Lumpur, Malaysia, 13-18 August 2017., pp.
3413–3422.
Li, L.; Switzer, A.D.; Wang, Y.; Chan, C.-H.; Qiu, Q.; Weiss, R. (2018): A modest 0.5-m rise in sea level will double the tsunami
hazard in Macau. In: Science Advances 4 (8), eaat1180. DOI: 10.1126/sciadv.aat1180.
Liao, C.-C.; Chang, Y.-W.; Lin, C.-A.; McDonough, J.M. (2010): Simulating flows with moving rigid boundary using immersed-
boundary method. In: Computers & Fluids 39 (1), pp. 152–167. DOI: 10.1016/j.compfluid.2009.07.011.
Liu, H.; Sakashita, T.; Sato, S. (2015): An experimental study on the tsunami boulder movement. In: Coastal Engineering
Proceedings 34. DOI: 10.9753/icce.v34.currents.16.
Liu, H.; Liu, H.; Guo, L.; Lu, S. (2017): Experimental study on the dam-break hydrographs at the gate location. In: Journal of
Ocean University of China 16 (4), pp. 697–702. DOI: 10.1007/s11802-017-3470-x.
Lloyd, T.O. (2016): An Experimental Investigation of Tsunami Forces on Coastal Structures. PhD dissertation. University Col-
lege, London, United Kingdom.
Lodhi, H.A.; Hasan, H.; Nandasena, N.A.K. (2020): The role of hydrodynamic impact force in subaerial boulder transport by
tsunami—Experimental evidence and revision of boulder transport equation. In: Sedimentary Geology 408. DOI:
10.1016/j.sedgeo.2020.105745.
López-Venegas, A.M.; ten Brink, U.S.; Geist, E.L. (2008): Submarine landslide as the source for the October 11, 1918 Mona
Passage tsunami: Observations and modeling. In: Marine Geology 254 (1-2), pp. 35–46. DOI: 10.1016/j.mar-
geo.2008.05.001.
Lorang, M.S. (2011): A wave-competence approach to distinguish between boulder and megaclast deposits due to storm
waves versus tsunamis. In: Marine Geology 283 (1-4), pp. 90–97. DOI: 10.1016/j.margeo.2010.10.005.
Løvholt, F.; Glimsdal, S.; Harbitz, C.B.; Zamora, N.; Nadim, F.; Peduzzi, P.l.; Dao, H.; Smebye, H. (2012): Tsunami hazard and
exposure on the global scale. In: Earth-Science Reviews 110 (1-4), pp. 58–73. DOI: 10.1016/j.earscirev.2011.10.002.
Luccio, P.A.; Voropayev, S.I.; Fernando, H.J.S.; Boyer, D.L.; Houston, W.N. (1998): The motion of cobbles in the swash zone on
an impermeable slope. In: Coastal Engineering 33 (1), pp. 41–60. DOI: 10.1016/S0378-3839(98)00003-9.
Luo, H. (2013): Immersed Boundary Method. In: Dongqing L. (Ed.): Encyclopedia of Microfluidics and Nanofluidics. Boston,
MA: Springer US, pp. 1–5. DOI: 10.1007/978-0-387-48998-8.
Maaßen, E. (2018): Physikalische Modellversuche zu tsunamiinduziertem Boulder Transport. Master Thesis. Institute of Hy-
draulic Engineering and Water Resources Management, Faculty of Civil Engineering, RWTH Aachen University. (un-
published)
MacInnes, B.T.; Bourgeois, J.; Pinegina, T.K.; Kravchunovskaya, E.A. (2009): Tsunami geomorphology: Erosion and deposition
from the 15 November 2006 Kuril Island tsunami. In: Geology 37 (11), pp. 995–998. DOI: 10.1130/G30172A.1.
Macintyre, I.G.; Littler, M.M.; Littler, D.S. (1995): Holocene history of Tobacco Range, Belize, Central America. In: Atoll Research
Bulletin 430, pp. 1–18. DOI: 10.5479/si.00775630.430.1.
Macintyre, I.G.; Toscano, M.A.; Lighty, R.G.; Bond, G. B. (2004): Holocene history of the mangrove islands of Twin Cays, Belize,
Central America. In: Atoll Research Bulletin (509-530), pp. 1–16. DOI: 10.5479/si.00775630.510.1.
Macintyre, I.G.; Glynn, P.W.; Steneck, R.S. (2001): A classic Caribbean algal ridge, Holandés Cays, Panamaá: An algal coated
storm deposit. In: Coral Reefs 20 (2), pp. 95–105. DOI: 10.1007/s003380000135.
Malaizé, B.; Bertran, P.; Carbonel, P.; Bonnissent, D.; Charlier, K.; Galop, D.; Imbert, D.; Serrand, N.; Stouvenot, C.; Pujol, C.
(2011): Hurricanes and climate in the caribbean during the past 3700 years BP. In: Holocene 21 (6), pp. 911–924. DOI:
10.1177/0959683611400198.
Mamo, B.; Strotz, L.; Dominey-Howes, D. (2009): Tsunami sediments and their foraminiferal assemblages. In: Earth-Science
Reviews 96 (4), pp. 263–278. DOI: 10.1016/j.earscirev.2009.06.007.
Mann, P.; Calais, E.; Ruegg, J.-C.; DeMets, C.; Jansma, P.E.; Mattioli, G.S. (2002): Oblique collision in the northeastern Caribbean
from GPS measurements and geological observations. In: Tectonics 21 (6), 7-1-7-26. DOI: 10.1029/2001TC001304.

188
References

Maragos, J.E.; Baines, G.B.K.; Beveridge, P.J. (1973): Tropical cyclone Bebe creates a new land formation on Funafuti Atoll. In:
Science (New York, N.Y.) 181 (4105), pp. 1161–1164. DOI: 10.1126/science.181.4105.1161.
Martin Arcos, M.E.; MacInnes, B.T.; Arreaga, P.; Rivera-Hernandez, F.; Weiss, R.; Lynett, P. (2013): An amalgamated meter-
thick sedimentary package enabled by the 2011 Tohoku tsunami in El Garrapatero, Galapagos Islands. In: Quaternary
Research (United States) 80 (1), pp. 9–19. DOI: 10.1016/j.yqres.2013.04.005.
Martin, H., Pohl, R. (2000): Technische Hydromechanik 4 - Hydraulische und numerische Modelle. Beuth publishing, Berlin,
Germany. ISBN: 978-3410241720.
Martinez, C.; Garcia-Martinez, R.; Miralles-Wilhelm, F. (2011): A two-phase debris flow model with boulder transport. In:
International Journal of Safety and Security Engineering 1 (4), pp. 389–402. DOI: 10.2495/SAFE-V1-N4-389-402.
Matsumoto, D.; Naruse, H.; Fujino, S.; Surphawajruksakul, A.; Jarupongsakul, T.; Sakakura, N.; Murayama, M. (2008): Truncated
flame structures within a deposit of the Indian Ocean Tsunami: Evidence of syn-sedimentary deformation. In: Sedimen-
tology 55 (6), pp. 1559–1570. DOI: 10.1111/j.1365-3091.2008.00957.x.
Mattheus, C.R.; Fowler, J.K. (2015): Paleotempestite Distribution across an Isolated Carbonate Platform, San Salvador Island,
Bahamas. In: Journal of Coastal Research 31 (4), pp. 842–858. DOI: 10.2112/JCOASTRES-D-14-00077.1.
May, S.M.; Engel, M.; Brill, D.; Squire, P.; Scheffers, A.; Kelletat, D. (2013): Coastal Hazards from Tropical Cyclones and Extra-
tropical Winter Storms Based on Holocene Storm Chronologies. In: In: Finkl, C.W. (Ed.), Coastal Hazards. Springer Neth-
erlands, pp. 557–585. DOI: 10.1007/978-94-007-5234-4_20.
May, S.M.; Brill, D.; Engel, M.; Scheffers, A.; Pint, A.; Opitz, S.; Wennrich, V.; Squire, P.; Kelletat, D.; Uckner, H.B. (2015a): Traces
of historical tropical cyclones and tsunamis in the Ashburton Delta (North-west Australia). In: Sedimentology 62 (6), pp.
1546–1572. DOI: 10.1111/sed.12192.
May, S.M.; Engel, M.; Brill, D.; Cuadra, C.; Lagmay, A.M.F.; Santiago, J.; Suarez, J. K.; Reyes, M.; Brückner, H. (2015b): Block and
boulder transport in Eastern Samar (Philippines) during Supertyphoon Haiyan. In: Earth Surface Dynamics 3 (4), pp. 543–
558. DOI: 10.5194/esurf-3-543-2015.
May, S.M.; Falvard, S.; Norpoth, M.; Pint, A.; Brill, D.; Engel, M.; Scheffers, A.; Dierick, M.; Paris, R.; Squire, P.; Brückner, H.
(2016): A mid-Holocene candidate tsunami deposit from the NW Cape (Western Australia). In: Sedimentary Geology
332, pp. 40–50. DOI: 10.1016/j.sedgeo.2015.11.010.
McAdoo, B. G.; Ah-Leong, J. S.; Bell, L.; Ifopo, P.; Ward, J.; Lovell, E.; Skelton, P. (2011): Coral reefs as buffers during the 2009
South Pacific tsunami, Upolu Island, Samoa. In: Earth-Science Reviews 107 (1-2), pp. 147–155. DOI: 10.1016/j.earsci-
rev.2010.11.005.
McCloskey, T.A.; Keller, G. (2009): 5000 year sedimentary record of hurricane strikes on the central coast of Belize. In: Qua-
ternary International 195 (1-2), pp. 53–68. DOI: 10.1016/j.quaint.2008.03.003.
McCloskey, T.A.; Liu, K.-B. (2012): A sedimentary-based history of hurricane strikes on the southern Caribbean coast of Nic-
aragua. In: Quaternary Research (United States) 78 (3), pp. 454–464. DOI: 10.1016/j.yqres.2012.07.003.
McCloskey, T.A.; Liu, K.-B. (2013a): A 7000 year record of paleohurricane activity from a coastal wetland in Belize. In: Holocene
23 (2), pp. 278–291. DOI: 10.1177/0959683612460782.
McCloskey, T.A.; Liu, K.-B. (2013b): Sedimentary history of mangrove cays in turneffe Islands, Belize: Evidence for sudden
environmental reversals. In: Journal of Coastal Research 29 (4), pp. 971–983. DOI: 10.2112/JCOASTRES-D-12-00156.1.
McGregor, D.F.M.; Potter, R. B. (1997): Environmental change and sustainability in the Caribbean: Terrestrial perspectives. In:
Sahr, W.D.; Ratter, B.M.W. (Eds.): Land, Sea and Human Effort in the Caribbean, pp. 1–17. ISBN: 978-3-98-031242-4.
Mckee, K.L.; Cahoon, D.R.; Feller, I.C. (2007): Caribbean mangroves adjust to rising sea level through biotic controls on change
in soil elevation. In: Global Ecology and Biogeography 16 (5), pp. 545–556. DOI: 10.1111/j.1466-8238.2007.00317.x.
Mergili, M.; Fischer, J.-T.; Krenn, J.; Pudasaini, S.P. (2017): R.avaflow v1, an advanced open-source computational framework
for the propagation and interaction of two-phase mass flows. In: Geoscientific Model Development 10 (2), pp. 553–569.
DOI: 10.5194/gmd-10-553-2017.
Mergili, M.; Emmer, A.; Juřicová, A.; Cochachin, A.; Fischer, J.-T.; Huggel, C.; Pudasaini, S.P. (2018): How well can we simulate
complex hydro-geomorphic process chains? The 2012 multi-lake outburst flood in the Santa Cruz Valley (Cordillera
Blanca, Perú). In: Earth Surface Processes and Landforms 43 (7), pp. 1373–1389. DOI: 10.1002/esp.4318.
Mergili, M., Pudasaini, S.P. (2021): r.avaflow - The mass flow simulation tool. https://www.avaflow.org [Accessed on:
31.07.2021].
Meschede, M.; Frisch, W. (1998): A plate-tectonic model for the Mesozoic and Early Cenozoic history of the Caribbean plate.
In: Tectonophysics 296 (3-4), pp. 269–291. DOI: 10.1016/S0040-1951(98)00157-7.
Miller, S.; Rowe, D.-A.; Brown, L.; Mandal, A. (2014): Wave-emplaced boulders: implications for development of “prime real
estate” seafront, North Coast Jamaica. In: Bulletin of Engineering Geology and the Environment 73 (1), pp. 109–122. DOI:
10.1007/s10064-013-0517-0.
Milne, G.A.; Peros, M. (2013): Data-model comparison of Holocene sea-level change in the circum-Caribbean region. In:
Global and Planetary Change 107, pp. 119–131. DOI: 10.1016/j.gloplacha.2013.04.014.
Minoura, K.; Nakaya, S. (1991): Traces of tsunami preserved in inter-tidal lacustrine and marsh deposits: some examples from
northeast Japan. In: Journal of Geology 99 (2), pp. 265–287. DOI: 10.1086/629488.

189
References

Minoura, K.; Nakaya, S.; Uchida, M. (1994): Tsunami deposits in a lacustrine sequence of the Sanriku coast, northeast Japan.
In: Sedimentary Geology 89 (1-2), pp. 25–31. DOI: 10.1016/0037-0738(94)90081-7.
Minoura, K.; Imamura, F.; Sugawara, D.; Kono, Y.; Iwashita, T. (2001): The 869 Jogan tsunami deposit and recurrence interval
of large-scaletsunami on the Pacific coast of northeast Japan. In: Journal of Natural Disaster Science 23 (2), pp. 83–88.
Monacci, N.M.; Meier-Grünhagen, U.; Finney, B.P.; Behling, H.; Wooller, M.J. (2009): Mangrove ecosystem changes during
the Holocene at Spanish Lookout Cay, Belize. In: Palaeogeography, Palaeoclimatology, Palaeoecology 280 (1-2), pp. 37–
46. DOI: 10.1016/j.palaeo.2009.05.013.
Moore, A.L.; McAdoo, B.G.; Ruffman, A. (2007): Landward fining from multiple sources in a sand sheet deposited by the 1929
Grand Banks tsunami, Newfoundland. In: Sedimentary Geology 200 (3-4), pp. 336–346. DOI:
10.1016/j.sedgeo.2007.01.012.
Moore, A.; Goff, J.; McAdoo, B.G.; Fritz, H.M.; Gusman, A.; Kalligeris, N.; Kalsum, K.; Susanto, A.; Suteja, D.; Synolakis, C.E.
(2011): Sedimentary Deposits from the 17 July 2006 Western Java Tsunami, Indonesia: Use of Grain Size Analyses to
Assess Tsunami Flow Depth, Speed, and Traction Carpet Characteristics. In: Pure and applied geophysics 168 (11), pp.
1951–1961. DOI: 10.1007/s00024-011-0280-8.
Morton, R.A.; Richmond, B. M.; Jaffe, B. E.; Gelfenbaum, G. (2006): Reconnaissance investigation of Caribbean extreme wave
deposits - preliminary observations, interpretations, and research directions. In: U.S. Geological Survey, Open-File Report
2006-1293. DOI: 10.3133/ofr20061293.
Morton, R.A.; Gelfenbaum, G.; Jaffe, B. E. (2007): Physical criteria for distinguishing sandy tsunami and storm deposits using
modern examples. In: Sedimentary Geology 200 (3-4), pp. 184–207. DOI: 10.1016/j.sedgeo.2007.01.003.
Morton, R.A.; Goff, J.R.; Nichol, S.L. (2008b): Hydrodynamic implications of textural trends in sand deposits of the 2004
tsunami in Sri Lanka. In: Sedimentary Geology 207 (1-4), pp. 56–64. DOI: 10.1016/j.sedgeo.2008.03.008.
Morton, R.A.; Richmond, B. M.; Jaffe, B. E.; Gelfenbaum, G. (2008a): Coarse-clast ridge complexes of the Caribbean: A pre-
liminary basis for distinguishing tsunami and storm-wave origins. In: Journal of Sedimentary Research 78 (9-10), pp.
624–637. DOI: 10.2110/jsr.2008.068.
Morton, R.A. (2010): First-order controls of extreme-storm impacts on the Mississippi-Alabama Barrier-Island Chain. In: Jour-
nal of Coastal Research 26 (4), pp. 635–648. DOI: 10.2112/08-1152.1.
Moscardelli, L.; Hornbach, M.; Wood, L. (2010): Tsunamigenic risks associated with mass transport complexes in offshore
Trinidad and Venezuela. In: Submarine Mass Movements and Their Consequences - 4th International Symposium. DOI:
10.1007/978-90-481-3071-9_59.
Moya, J.C.; Mercado, A. (2006): Geomorphologic and stratigraphic investigations on historic and pre-historic tsunami in
northwestern Puerto Rico: Implications for long term coastal evolution. In: Caribbean Tsunami Hazard, pp. 149–177.
DOI: 10.1142/9789812774613_0007
Munson, B.R., Young, D.F., Okiishi, T.H., Huebsch, W.W. (2009): Fundamentals of Fluid Mechanics. Hoboken, New, Jersey,
U.S.A.: John Wiley & Sons. ISBN: 978-1-11-811613-5.
Mylroie, J.E. (2008): Late Quaternary sea-level position: Evidence from Bahamian carbonate deposition and dissolution cy-
cles. In: Quaternary International 183 (1), pp. 61–75. DOI: 10.1016/j.quaint.2007.06.030.
Naito, C.; Cercone, C.; Riggs, H.R.; Cox, D. (2014): Procedure for Site Assessment of the Potential for Tsunami Debris Impact.
In: J. Waterway, Port, Coastal, Ocean Eng. 140 (2), pp. 223–232. DOI: 10.1061/(ASCE)WW.1943-5460.0000222.
Nakamura, T.; Mizutani, N.; Wakamatsu, Y. (2012): Study on drift behavior of container on apron due to tsunami-induced
incoming and return flow. In: Proceedings of the International Conference on Coastal Engineering 1 (33), p. 16. DOI:
10.9753/icce.v33.currents.16.
Nakamura, M.; Tanaka, K.; Tanaka, F.; Matsuura, Y.; Komi, R.; Niiyama, M.; Kawakami, M.; Koeda, Y.; Sakai, T.i; Onoda, T.; Itoh,
T. (2017): Long-Term Effects of the 2011 Japan Earthquake and Tsunami on Incidence of Fatal and Nonfatal Myocardial
Infarction. In: The American journal of cardiology 120 (3), pp. 352–358. DOI: 10.1016/j.amjcard.2017.05.002.
Nakata, T.; Kawana, T. (1995): Historical and prehistorical large tsunamis in the southern Ryukyus, Japan. In: Tsunami: Pro-
gress in Prediction, Disaster Prevention and Warning, pp. 211–222. DOI: 10.1007/978-94-015-8565-1_15.
Namegaya, Y., Satake, K., Yamamoto, S. (2010): Numerical simulation of the AD 869 Jogan tsunami in Ishinomaki and Sendai
plains and Ukedo river-mouth lowland. In: Annual Report of Active Fault and Paleoearthquake Researches (10), pp. 1–
21.
Namegaya, Y.; Satake, K. (2014): Reexamination of the A.D. 869 Jogan earthquake size from tsunami deposit distribution,
simulated flow depth, and velocity. In: Geophys. Res. Lett. 41 (7), pp. 2297–2303. DOI: 10.1002/2013GL058678.
Nanayama, F.; Shigeno, K.; Satake, K.; Shimokawa, K.; Koitabashi, S.; Miyasaka, S.; Ishii, M. (2000): Sedimentary differences
between the 1993 Hokkaido-nansei-oki tsunami and the 1959 Miyakojima typhoon at Taisei, southwestern Hokkaido,
northern Japan. In: Sedimentary Geology 135 (1-4), pp. 255–264. DOI: 10.1016/S0037-0738(00)00076-2.
Nandasena, N.A.K.; Paris, R.; Tanaka, N. (2011a): Reassessment of hydrodynamic equations: Minimum flow velocity to initiate
boulder transport by high energy events (storms, tsunamis). In: Marine Geology 281 (1-4), pp. 70–84. DOI: 10.1016/j.mar-
geo.2011.02.005.

190
References

Nandasena, N.A.K.; Paris, R.; Tanaka, N. (2011b): Numerical assessment of boulder transport by the 2004 Indian ocean tsu-
nami in Lhok Nga, West Banda Aceh (Sumatra, Indonesia). In: Computers & Geosciences 37 (9), pp. 1391–1399. DOI:
10.1016/j.cageo.2011.02.001.
Nandasena, N.A.K.; Tanaka, N. (2013): Boulder transport by high energy: Numerical model-fitting experimental observations.
In: Ocean Engineering 57, pp. 163–179. DOI: 10.1016/j.oceaneng.2012.09.012.
Nandasena, N.A.K.; Tanaka, N.; Sasaki, Y.; Osada, M. (2013): Boulder transport by the 2011 Great East Japan tsunami: Com-
prehensive field observations and whither model predictions? In: Marine Geology 346, pp. 292–309. DOI: 10.1016/j.mar-
geo.2013.09.015.
Nandasena, N.A.K. (2020): Perspective of incipient motion formulas: boulder transport by high-energy waves. In In: Engel,
M.; Pilarczyk, J.; May, S.M.; Brill, D.; Garrett, E. (Eds.): Geological Records of Tsunamis and Other Extreme Waves. Elsevier,
pp. 641–659. DOI: https://doi.org/10.1016/B978-0-12-815686-5.00029-8.
Naruse, H.; Fujino, S.; Suphawajruksakul, A.; Jarupongsakul, T. (2010): Features and formation processes of multiple deposi-
tion layers from the 2004 Indian Ocean Tsunami at Ban Nam Kem, southern Thailand. In: Island Arc 19 (3), pp. 399–411.
DOI: 10.1111/j.1440-1738.2010.00732.x.
NASA (2008): NPARC Alliance CFD Verification and Validation Web Site: Glossary of Verification and Validation Terms.
https://www.grc.nasa.gov/www/wind/valid/tutorial/glossary.html [Accessed on: 22.08.2021].
Neale, P. (1885): The Krakatoa Eruption. In: The Leisure Hour 34, pp. 348–351.
Neumann, B.; Vafeidis, A.T.; Zimmermann, J.; Nicholls, R.J. (2015): Future coastal population growth and exposure to sea-
level rise and coastal flooding--a global assessment. In: PloS one 10 (3), e0118571. DOI: 10.1371/journal.pone.0118571.
NGDC/WDS (National Geophysical Data Center/World Data Service) (NGDC/WDS) (2016v): Global Historical Tsunami Data-
base. National Geophysical Data Center, NOAA. https://www.ngdc.noaa.gov/hazard/tsu_db.shtml [Accessed on:
22.08.2021].
Nichol, S.L.; Kench, P.S. (2008): Sedimentology and preservation potential of carbonate sand sheets deposited by the De-
cember 2004 Indian Ocean tsunami: South Baa Atoll, Maldives. In: Sedimentology 55 (5), pp. 1173–1187. DOI:
10.1111/j.1365-3091.2007.00941.x.
Nishimura, Y.; Miyaji, N. (1995): Tsunami deposits from the 1993 Southwest Hokkaido earthquake and the 1640 Hokkaido
Komagatake eruption, northern Japan. In: Pure and Applied Geophysics 144 (3-4), pp. 719–733. DOI:
10.1007/BF00874391.
Nistor, I.; Goseberg, N.; Stolle, J.; Mikami, T.; Shibayama, T.; Nakamura, R.; Matsuba, S. (2017): Experimental Investigations of
Debris Dynamics over a Horizontal Plane. In: Journal of Waterway, Port, Coastal, and Ocean Engineering 143 (3). DOI:
10.1061/(ASCE)WW.1943-5460.0000371.
NOAA (2014): Natural Hazard Viewer. http://maps.ngdc.noaa.gov/viewers/hazards/ [Accessed on 22.08.2021].
Noji, M.; Imamura, F.; Shuto, N. (1993): Developing the method of tsunami boulder transport calculation. In: Coastal Engi-
neering, Japan Society of Civil Engineers (40).
Noormets, R.; Crook, K.A.W.; Felton, E. A. (2004): Sedimentology of rocky shorelines: 3. Hydrodynamics of megaclast em-
placement and transport on a shore platform, Oahu, Hawaii. In: Sedimentary Geology 172 (1-2), pp. 41–65. DOI:
10.1016/j.sedgeo.2004.07.006.
Nott, J. (1997): Extremely high-energy wave deposits inside the Great Barrier Reef, Australia: Determining the cause-tsunami
or tropical cyclone. In: Marine Geology 141 (1-4), pp. 193–207. DOI: 10.1016/S0025-3227(97)00063-7.
Nott, J. (2003): Waves, coastal boulder deposits and the importance of the pre-transport setting. In: Earth and Planetary
Science Letters 210 (1-2), pp. 269–276. DOI: 10.1016/S0012-821X(03)00104-3.
NTHMP (National Tsunami Hazard Mitigation Program) (2012): Proceedings and Results of the 2011 NTHMP Model Bench-
marking Workshop. U.S. Department of Commerce/ NOAA/NTHMP (NOAA Special Report). Boulder, U.S.A. https://per-
manent.fdlp.gov/gpo44987/nthmpWorkshopProcMerged.pdf [Accessed on 22.08.2021].
O’Brien, M.P.; Morison, J.R. (1952): The forces exerted by waves on objects. In: Transactions, American Geophysical Union,
Volume 33, Issue 1, p. 32-38 33 (1), pp. 32–38. DOI: 10.1029/TR033i001p00032.
O’Loughlin, K.F.; Lander, J.F. (2003): Caribbean tsunamis: A 500-year history from 1498-1998. In: Caribbean Tsunamis: A 500-
Year History from 1498-1998. Springer Nature, Heidelberg, Germany. ISBN: 978-94-017-0321-5.
Oetjen, J.; Engel, M.; Effkemann, C.; May, S.M.; Pudasaini, S.P.; Wöffler, T.; Aizinger, V.; Schüttrumpf, H.; Brückner, H. (2015):
Numerical modelling of tsunami scenarios for the island of Bonaire (Leeward Antilles). In: Programme and Abstract
Book, 4th International Tsunami Field Symposium, 23–27 March 2015, Kata Beach, Phuket, Thailand, pp. 81–85.
Oetjen, J.; Engel, M.; Pudasaini, S.P.; Schüttrumpf, H.; Brückner, H. (2017a): An advanced three-phase physical, experimental
and numerical method for tsunami induced boulder transport. In: Proceedings of the 19th EGU General Assembly 2017,
Vienna, Austria.
Oetjen, J.; Engel, M.; Brückner, H.; Pudasaini, S.P.; Schüttrumpf, H. (2017b): Enhanced field observation based physical and
numerical modeling of tsunami induced boulder transport: phase 1: physical experiments. In: Proceedings of the Inter-
national Conference on Coastal Engineering 35. DOI: 10.9753/icce.v35.management.4.

191
References

Oetjen, J., Engel, M., Schönberger, J.J., Pudasaini, S.P., Schüttrumpf, H. (2018): Simulation of boulder transport in a flume
comparing cuboid and complex-shaped boulder models. In: EGU General Assembly 2018, Geophysical Research Ab-
stracts 20. DOI: 10.18154/RWTH-2019-03133.
Oetjen, J.; Engel, M.; Pudasaini, S.P.; Schuettrumpf, H. (2020a): Significance of boulder shape, shoreline configuration and
pre‐transport setting for the transport of boulders by tsunamis. In: Earth Surface Processes and Landforms 45 (9), pp.
2118–2133. DOI: 10.1002/esp.4870.
Oetjen, J.; Schüttrumpf, H.; Engel, M. (2020b): Experimental models of coarse-clast transport by tsunamis. In: Engel, M.;
Pilarczyk, J.; May, S.M.; Brill, D.; Garrett, E. (Eds.): Geological Records of Tsunamis and Other Extreme Waves. Elsevier, pp.
585-615. DOI: 10.1016/B978-0-12-815686-5.00027-4.
Okada, Y. (1985): Surface deformation due to shear and tensile faults in a half-space. In: Bulletin of the Seismological Society
of America 75 (4), pp. 1135–1154. DOI: 10.1785/BSSA0750041135.
Oppenheimer, M., B.C. Glavovic, J. Hinkel, R. van de Wal, A.K. Magnan, A. Abd-Elgawad, R. Cai, M. Cifuentes-Jara, R.M. De-
Conto, T. Ghosh, J. Hay, F. Isla, B. Marzeion, B. Meyssignac, Z. Sebesvari (2019): Sea Level Rise and Implications for Low-
Lying Islands, Coasts and Communities. In: Pörtner, H.-O.; Roberts, D.C.; Masson-Delmotte, V.; Zhai, P.; Tignor, M.;
Poloczanska, E.; Mintenbeck, K.; Alegría, A.; Nicolai, M.; Okem, A.; Petzold, J.; Rama, B.; Weyer, N.M. (Eds.). IPCC (IPCC
Special Report on the Ocean and Cryosphere in a Changing Climate). https://www.ipcc.ch/srocc/ [Accessed on
22.08.2021].
Oropeza, J.; Audemard, F. A.; Beck, C.; Vallée, M. (2015): New potential sedimentary evidences of paleotsunamis on coastal
lagoons of Chacopata, State of Sucre, Venezuela. In: 6th International INQUA Meeting on Paleoseismology, Active Tec-
tonics and Archaeoseismology, 19–24 April 2015, Pescina, Fucino Basin, Italy 27, pp. 343–345.
Palmer, S.; Burn, M. (2012): A late-Holocene record of marine washover events from a coastal lagoon in Jamaica, West Indies.
In: Quaternary International 279-280, pp. 365–366. DOI: 10.1016/j.quaint.2012.08.1116.
Pantosti, D.; Barbano, M. S.; Smedile, A.; Martini, P. M. de; Tigano, G. (2008): Geological evidence of paleotsunamis at Torre
degli Inglesi (northeast Sicily). In: Geophysical Research Letters 35 (5). DOI: 10.1029/2007GL032935.
Pararas-Carayannis, G. (2004): Volcanic tsunami generating source mechanisms in the Eastern Caribbean region. In: Science
of Tsunami Hazards 22 (2), pp. 74–114.
Paris, R.; Lavigne, F.; Wassmer, P.; Sartohadi, J. (2007): Coastal sedimentation associated with the December 26, 2004 tsunami
in Lhok Nga, west Banda Aceh (Sumatra, Indonesia). In: Marine Geology 238 (1-4), pp. 93–106. DOI: 10.1016/j.mar-
geo.2006.12.009.
Paris, R.; Wassmer, P.; Sartohadi, J.; Lavigne, F.; Barthomeuf, B.; Desgages, E. et al. (2009): Tsunamis as geomorphic crises:
Lessons from the December 26, 2004 tsunami in Lhok Nga, West Banda Aceh (Sumatra, Indonesia). In: Geomorphology
104 (1-2), pp. 59–72. DOI: 10.1016/j.geomorph.2008.05.040.
Paris, R.; Fournier, J.; Poizot, E.; Etienne, S.; Morin, J.; Lavigne, F.; Wassmer, P. (2010): Boulder and fine sediment transport and
deposition by the 2004 tsunami in Lhok Nga (western Banda Aceh, Sumatra, Indonesia): A coupled offshore-onshore
model. In: Marine Geology 268 (1-4), pp. 43–54. DOI: 10.1016/j.margeo.2009.10.011.
Paris, R.; Wassmer, P.; Lavigne, F.; Belousov, A.; Belousova, M.; Iskandarsyah, Y.; Benbakkar, M.; Ontowirjo, B.; Mazzoni, N.
(2014): Coupling eruption and tsunami records: The Krakatau 1883 case study, Indonesia. In: Bulletin of volcanology 76
(4), pp. 1–23. DOI: 10.1007/s00445-014-0814-x.
Paris, R. (2015): Source mechanisms of volcanic tsunamis. In: Philosophical Transactions of the Royal Society A: Mathematical,
Physical and Engineering Sciences 373 (2053). DOI: 10.1098/rsta.2014.0380.
Park, L.E.; Siewers, F.D.; Metzger, T.; Sipahioglu, S. (2009): After the hurricane hits: Recovery and response to large storm
events in a saline lake, San Salvador Island, Bahamas. In: Quaternary International 195 (1-2), pp. 98–105. DOI:
10.1016/j.quaint.2008.06.010.
Park, L.E. (2012): Comparing two long-term hurricane frequency and intensity records from San Salvador Island, Bahamas.
In: Journal of Coastal Research 28 (4), pp. 891–902. DOI: 10.2112/JCOASTRES-D-11-00065.1.
Parsons, M.L. (1998): Salt marsh sedimentary record of the landfall of hurricane Andrew on the Louisiana coast: Diatoms and
other paleoindicators. In: Journal of Coastal Research 14 (3), pp. 939–950.
Parsons, T.; Geist, E. L. (2009): Tsunami probability in the Caribbean Region. In: Pure and applied geophysics 165 (11-12), pp.
2089–2116. DOI: 10.1007/s00024-008-0416-7.
Pedersen, G.; Gjevik, B. (1983): Run-up of solitary waves. In: Journal of Fluid Mechanics 135, pp. 283–299. DOI:
10.1017/S0022112083003080.
Peltzer, S. (2015): Planung, Entwurf, Konstruktion und Realisierung eines physikalischen Modellversuchs zur Generierung
einer schiffserzeugten Primärwellenbelastung auf Strombauwerke an Wasserstraßen. Master Thesis. Institute of Hydrau-
lic Engineering and Water Resources Management. Faculty of Civil Engineering, RWTH Aachen University. (unpublished)
Penchev, V. (2008): Extreme solitary waves at restricted water depth. In: Book of Abstracts of the Second International Con-
ference on the Application of Physical Modelling to Port and Coastal Protection (CoastLab08), Bari, Italy, July, 2008.
Peros, M.C.; Reinhardt, E.G.; Davis, A.M. (2007a): A 6000-year record of ecological and hydrological changes from Laguna de
la Leche, north coastal Cuba. In: Quaternary research 67 (1), pp. 69–82. DOI: 10.1016/j.yqres.2006.08.004.

192
References

Peros, M.C.; Reinhardt, E. G.; Schwarcz, H. P.; Davis, A. M. (2007b): High-resolution paleosalinity reconstruction from Laguna
de la Leche, north coastal Cuba, using Sr, O, and C isotopes. In: Palaeogeography, Palaeoclimatology, Palaeoecology
245 (3-4), pp. 535–550. DOI: 10.1016/j.palaeo.2006.09.006.
Peskin, C.S. (1972): Flow patterns around heart valves: A numerical method. In: Journal of Computational Physics 10 (2), pp.
252–271. DOI: 10.1016/0021-9991(72)90065-4.
Peskin, C.S. (2002): The immersed boundary method. In: Acta Numerica 11, pp. 479–517. DOI: 10.1017/S0962492902000077.
Peters, R.; Jaffe, B.E. (2010): Identification of tsunami deposits in the geologic record: developing criteria using recent tsunami
deposits. In: US Geological Survey Open-File Report 2010-1239. https://pubs.usgs.gov/of/2010/1239/of2010-1239.pdf
[Accessed on 22.08.2021].
Petroff, C.M., Moore, A.L., Árnason, H. (2001): Particle advection by turbulent bores–orientation effects. In: International
Tsunami Symposium 2001 Proceedings (Session 7, Number 7-23), pp. 897–904.
Phantuwongraj, S.; Choowong, M. (2012): Tsunamis versus storm deposits from Thailand. In: Natural Hazards 63 (1), pp. 31–
50. DOI: 10.1007/s11069-011-9717-8.
Pignatelli, C.; Sansò, P.; Mastronuzzi, G. (2009): Evaluation of tsunami flooding using geomorphologic evidence. In: Marine
Geology 260 (1-4), pp. 6–18. DOI: 10.1016/j.margeo.2009.01.002.
Pignatelli, C.; Scheffers, A.; Scheffers, S.; Mastronuzzi, G. (2010): Assessment of extreme wave flooding from geomorphologic
evidence in Bonaire (Netherlands antilles). In: Zeitschrift fur Geomorphologie 54 (SUPPL. 3), pp. 219–245. DOI:
10.1127/0372-8854/2010/0054S3-0026.
Pilarczyk, J.E.; Reinhardt, E.G. (2012): Homotrema rubrum (Lamarck) taphonomy as an overwash indicator in Marine Ponds
on Anegada, British Virgin Islands. In: Natural Hazards 63 (1), pp. 85–100. DOI: 10.1007/s11069-010-9706-3.
Pilarczyk, J.E.; Horton, B.P.; Soria, J.L.A.; Switzer, A.D.; Siringan, F.; Fritz, H. M. et al. (2016): Micropaleontology of the 2013
Typhoon Haiyan overwash sediments from the Leyte Gulf, Philippines. In: Sedimentary Geology 339, pp. 104–114. DOI:
10.1016/j.sedgeo.2016.04.001.
Pindell, J.L.; Kennan, L. (2009): Tectonic evolution of the Gulf of Mexico, Caribbean and northern South America in the mantle
reference frame: An update. In: Geological Society Special Publication 328, pp. 1–55. DOI: 10.1144/SP328.1.
Prizomwala, S.P.; Gandhi, D.; Ukey, V.M.; Bhatt, N.; Rastogi, B.K. (2015): Coastal boulders as evidences of high-energy marine
events from Diu Island, west coast of India: storm or palaeotsunami? In: Natural Hazards 75 (2), pp. 1187–1203. DOI:
10.1007/s11069-014-1371-5.
Pudasaini, S.P. (2012): A general two-phase debris flow model. In: Journal of Geophysical Research: Earth Surface 117 (3).
DOI: 10.1029/2011JF002186.
Pudasaini, S.P.; Mergili, M. (2019): A Multi-Phase Mass Flow Model. In: Journal of Geophysical Research: Earth Surface 124
(12), pp. 2920–2942. DOI: 10.1029/2019JF005204.
Putra, P.S.; Nishimura, Y.; Yulianto, E. (2013): Sedimentary Features of Tsunami Deposits in Carbonate-Dominated Beach
Environments: A Case Study from the 25 October 2010 Mentawai Tsunami. In: Pure and applied geophysics 170 (9-10),
pp. 1583–1600. DOI: 10.1007/s00024-012-0539-8.
Radtke, U.; Schellmann, G.; Scheffers, A.; Kelletat, D.; Kromer, B.; Kasper, H.U. (2003): Electron spin resonance and radiocarbon
dating of coral deposited by Holocene tsunami events on Curaçao, Bonaire and Aruba (Netherlands Antilles). In: Qua-
ternary Science Reviews 22 (10-13), pp. 1309–1315. DOI: 10.1016/S0277-3791(03)00036-2.
Rahiman, T.I.H.; Pettinga, J.R.; Watts, P. (2007): The source mechanism and numerical modelling of the 1953 Suva tsunami,
Fiji. In: Marine Geology 237 (1-2), pp. 55–70. DOI: 10.1016/j.margeo.2006.10.036.
Rajendran, C.P.; Rajendran, K.; Srinivasalu, S.; Andrade, V.; Aravazhi, P.; Sanwal, J. (2011): Geoarchaeological evidence of a
Chola-period tsunami from an ancient port at Kaveripattinam on the southeastern coast of India. In: Geoarchaeology
26 (6), pp. 867–887. DOI: 10.1002/gea.20376.
Ramalho, R.S.; Winckler, G.; Madeira, J.; Helffrich, G.R.; Hipólito, A.; Quartau, R.; Adena, K.; Schaefer, J.M. (2015): Hazard
potential of volcanic flank collapses raised by new megatsunami evidence. In: Science Advances 1 (9). DOI: 10.1126/sci-
adv.1500456.
Ramcharan, E.K. (2004): Mid-to-late Holocene sea level influence on coastal wetland development in Trinidad. In: Quaternary
International 120 (1), pp. 145–151. DOI: 10.1016/j.quaint.2004.01.013.
Ramcharan, E.K.; McAndrews, J.H. (2006): Holocene development of coastal wetland at Maracas Bay, Trinidad, West Indies.
In: Journal of Coastal Research 22 (3), pp. 581–586. DOI: 10.2112/04A-0001.1.
Rappaport, E.N.; Fernandez-Partagas, J. (1997): The deadliest Atlantic tropical cyclones, 1492-1994. In: The Deadliest Atlantic
Tropical Cyclones, 1492-1994, NOAA Technical Memorandum NWS NHC 47, USA.
Razzhigaeva, N.G.; Ganzei, L.A.; Grebennikova, T.A.; Ivanova, E.D.; Kaistrenko, V.M. (2006): Sedimentation particularities dur-
ing the tsunami of December 26, 2004, in northern Indonesia: Simelue Island and the Medan coast of Sumatra Island.
In: Oceanology 46 (6), pp. 875–890. DOI: 10.1134/S0001437006060130.
Reading, A.J. (1990): Caribbean tropical storm activity over the past four centuries. In: International Journal of Climatology
10 (4), pp. 365–376. DOI: 10.1002/joc.3370100404.
Reid, H.F.; Taber, S. (1919): The Puerto Rico earthquakes of October-November 1918. In: Bulletin of the Seismological Society
of America 9 (4), pp. 95–127.

193
References

Reinhardt, E. G.; Goodman, B. N.; Boyce, J. I.; Lopez, G.; van Hengstum, P.; Rink, W. J.; Mart, Y.; Raban, A. (2006): The tsunami
of 13 December A.D. 115 and the destruction of Herod the Great’s harbor at Caesarea Maritima, Israel. In: Geology 34
(12), pp. 1061–1064. DOI: 10.1130/G22780A.1.
Reinhardt, E. G.; Pilarczyk, J.; Brown, A. (2012): Probable tsunami origin for a Shell and Sand Sheet from marine ponds on
Anegada, British Virgin Islands. In: Natural Hazards 63 (1), pp. 101–117. DOI: 10.1007/s11069-011-9730-y.
Reyes, M.; Engel, M.; May, S.M.; Brill, D.; Brueckner, H. (2015): Life and death after super typhoon Haiyan. In: Coral Reefs 34
(2), p. 419. DOI: 10.1007/s00338-015-1259-1.
Rhinefrank, K.; Schacher, A.; Prudell, Joe; Stillinger, C.; Naviaux, D.; Brekken, T.; von Jouanne, A.; Newborn, D.; Yim, S.; Cox, D.
(2010): High Resolution Wave Tank Testing of Scaled Wave Energy Devices. In: Proceedings of OMAE2010, 29th Inter-
national Conference on Offshore Mechanics and Arctic Engineering, pp. 505–509. DOI: 10.1115/OMAE2010-20602.
Richmond, B.M.; Watt, S.; Buckley, M.; Jaffe, B.E.; Gelfenbaum, G.; Morton, R.A. (2011): Recent storm and tsunami coarse-clast
deposit characteristics, southeast Hawai’i. In: Marine Geology 283 (1-4), pp. 79–89. DOI: 10.1016/j.margeo.2010.08.001.
Richmond, B.; Szczuciński, W.; Chagué-Goff, C.; Goto, K.; Sugawara, D.; Witter, R.; Tappin, D.R.; Jaffe, B.; Fujino, S.; Nishimura,
Y.; Goff, J. (2012): Erosion, deposition and landscape change on the Sendai coastal plain, Japan, resulting from the March
11, 2011 Tohoku-oki tsunami. In: Sedimentary Geology 282, pp. 27–39. DOI: 10.1016/j.sedgeo.2012.08.005.
Rigby, J.K.; Roberts, H.H. (1976): Geology, reefs, and marine communities of Grand Cayman Island, British West Indies. In:
Brigham Young University, Utah, U.S., Geology Studies, Special Publication 4, pp. 1–95.
Rixhon, G.; May, S.M.; Engel, M.; Mechernich, S.; Keulertz, R.; Schröder-Ritzau, A.; Fohlmeister, J.; Frank, N.; Dunai, T.; Brückner,
H. (2016): Multiple dating approach (14 C, U/Th and 36 Cl) of tsunami-transported reef-top megaclasts on Bonaire
(Leeward Antilles) – potential and current limitations. In: Geophysical Research Abstracts 18.
Rixhon, G.; May, S.M.; Engel, M.; Mechernich, S.; Schroeder-Ritzrau, A.; Frank, N.; Fohlmeister, J.; Boulvain, F.; Dunai, T.; Brück-
ner, H. (2018): Multiple dating approach (14C, 230Th/U and 36Cl) of tsunami-transported reef-top boulders on Bonaire
(Leeward Antilles) – Current achievements and challenges. In: Marine Geology 396, pp. 100–113. DOI: 10.1016/j.mar-
geo.2017.03.007.
Robertson, I.N.; Riggs, H.R.; Mohamed, A. (2008): Experimental Results of Tsunami Bore Forces on Structures. In: Proceedings
of the 27th International Conference on Offshore Mechanics and Arctic Engineering, pp. 509–517. DOI:
10.1115/OMAE2008-57525.
Robertson, I.N.; Paczkowski, K.; Riggs, H.R.; Mohamed, A. (2011): Tsunami Bore Forces on Walls. In: Proceedings of the 30th
International Conference on Offshore Mechanics and Arctic Engineering, pp. 395–403. DOI: 10.1115/OMAE2011-49487.
Robinson, E.; Rowe, D.-A.C.; Khan, S.A. (2006): Wave-emplaced boulders on Jamaica’s rocky shores. In: Zeitschrift fur Geo-
morphologie, Supplement 146, pp. 39–57.
Rovere, A.; Casella, E.; Harris, D.L.; Lorscheid, T.; Nandasena, N.A.K.; Dyer, B.; Sandstrom, M.R.; Stocchi, P.; D'Andrea, W.J.;
Raymo, M.E. (2017): Giant boulders and Last Interglacial storm intensity in the North Atlantic. In: Proceedings of the
National Academy of Sciences of the United States of America 114 (46), pp. 12144–12149. DOI:
10.1073/pnas.1712433114.
Rowe, D.-A.; Khan, S.; Robinson, E. (2009): Hurricanes or tsunami? Comparative analysis of extensive boulder arrays along
the southwest and north coasts of Jamaica: lessons for coastal management. In: Global change and Caribbean vulnera-
bility, University of the West Indies Press, Kingston, Jamaica, pp. 49–73. ISBN: 978-9-76-64022-1.
Rueben, M.; Cox, D.; Holman, R.; Shin, S.; Stanley, J. (2015): Optical Measurements of Tsunami Inundation and Debris Move-
ment in a Large-Scale Wave Basin. In: Journal of Waterway, Port, Coastal, and Ocean Engineering 141 (1). DOI:
10.1061/(ASCE)WW.1943-5460.0000267.
Ruffini, G.; Briganti, R.; de Girolamo, P.; Stolle, J.; Ghiassi, B.; Castellino, M. (2021): Numerical Modelling of Flow-Debris Inter-
action during Extreme Hydrodynamic Events with DualSPHysics-CHRONO. In: Applied Sciences 11 (8), p. 3618. DOI:
10.3390/app11083618.
Samankassou, E.; Viret, G.; Kindler, P. (2008): Tsunami evidence during marine oxygen-isotope substage 5e on Eleuthera
Island, Bahamas. In: The 14th Symposium on the Geology of the Bahamas and Other Carbonate Regions, Abstracts and
Program, pp. 21–22.
Satake, K.; Namegaya, Y.; Yamaki, S. (2008): Numerical simulation of the AD 869 Jōgan tsunami in Ishinomaki and Sendai
plains. In: Annual Report of Active Fault and Palaeoearthquake Researches (8), pp. 71–89. (in Japanese)
Sato, T.; Nakamura, N.; Goto, K.; Kumagai, Y.; Nagahama, H.; Minoura, K. (2014): Paleomagnetism reveals the emplacement
age of tsunamigenic coral boulders on Ishigaki Island, Japan. In: Geology 42 (7), pp. 603–606. DOI: 10.1130/G35366.1.
Sawai, Y., Shishikura, M., Okamura, Y., Takada, K., Matsu’ura, T., Aung, T.T., Komatsubara, J., Fujii, Y., Fujiwara, O., Satake, K.,
Kamataki, T., Sato, N. (2007): A study of paleotsunami using handy geoslicer in Sendai Plain (Sendai, Natori, Iwanuma,
Watari, and Yamamoto). In: Miyagi, Japan, Annual Report on Active Fault and Paleoearthquake Researches (7), pp. 47–
80. (in Japanese)
Sawai, Y.; Jankaew, K.; Martin, M. E.; Prendergast, A.; Choowong, M.; Charoentitirat, T. (2009): Diatom assemblages in tsunami
deposits associated with the 2004 Indian Ocean tsunami at Phra Thong Island, Thailand. In: Marine Micropaleontology
73 (1-2), pp. 70–79. DOI: 10.1016/j.marmicro.2009.07.003.

194
References

Sawai, Y.; Namegaya, Y.; Okamura, Y.; Satake, K.; Shishikura, M. (2012): Challenges of anticipating the 2011 Tohoku earth-
quake and tsunami using coastal geology. In: Geophysical Research Letters 39 (21). DOI: 10.1029/2012GL053692.
Schechinger, M. (2015). Numerische Modellierung Tsunami induzierter Sedimentablagerungen auf der Insel Bonaire. Bach-
elor thesis, Institute of Hydraulic Engineering and Water Resources Management, Faculty of Civil Engineering, RWTH
Aachen University. (unpublished)
Scheffers, A. (2002a): Paleotsunami evidences from boulder deposits on Aruba, Curacao and Bonaire. In: Science of Tsunami
Hazards 20 (1), pp. 26–37.
Scheffers, A. (2002b): Paleotsunamis in the Caribbean. Field evidences and datings from Aruba, Curaçao and Bonaire. In:
Essener Geographische Arbeiten 33.
Scheffers, A.; Kelletat, D. (2003): Sedimentologic and geomorphologic tsunami imprints worldwide - A review. In: Earth-
Science Reviews 63 (1-2), pp. 83–92. DOI: 10.1016/S0012-8252(03)00018-7.
Scheffers, A. (2004): Tsunami imprints on the Leeward Netherlands Antilles (Aruba, Curaçao, Bonaire) and their relation to
other coastal problems. In: Quaternary International 120 (1), pp. 163–172. DOI: 10.1016/j.quaint.2004.01.015.
Scheffers, A. (2005): Coastal response to extreme wave events – hurricanes and tsunamis on Bonaire. In: Essener Geogra-
phische Arbeiten (37).
Scheffers, A.; Scheffers, S.; Kelletat, D. (2005): Paleo-tsunami relics on the southern and central Antillean Island Arc. In: Journal
of Coastal Research 21 (2), pp. 263–273. DOI: 10.2112/03-0144.1.
Scheffers, A.; Kelletat, D. (2006): New evidence and datings of Holocene paleo-tsunami events in the Caribbean (Barbados,
St. Martin and Anguilla). In: Caribbean Tsunami Hazard, pp. 178–202.
Scheffers, A.; Scheffers, S. (2006): Documentation of the impact of hurricane ivan on the coastline of bonaire (Netherlands
Antilles). In: Journal of Coastal Research 22 (6), pp. 1437–1450. DOI: 10.2112/05-0535.1.
Scheffers, S.R.; Scheffers, A.; Radtke, U.; Kelletat, D.; Staben, K.; Bak, R. (2006): Tsunamis trigger long-lasting phase-shift in a
coral reef ecosystem. In: Zeitschrift fur Geomorphologie, Supplementband 146, pp. 59–79.
Scheffers, A. (2006a): Ripple marks in coarse tsunami deposits. In: Zeitschrift fur Geomorphologie, Supplementband 146, pp.
221–233.
Scheffers, A. (2006b): Sedimentary impacts of Holocene tsunami events from the intra Americas seas and Southern Europe:
A review. In: Zeitschrift fur Geomorphologie, Supplementband 146, pp. 7–37.
Scheffers, S.R.; Haviser, J.; Browne, T.; Scheffers, A. (2009): Tsunamis, hurricanes, the demise of coral reefs and shifts in pre-
historic human populations in the Caribbean. In: Quaternary International 195 (1-2), pp. 69–87. DOI:
10.1016/j.quaint.2008.07.016.
Scheffers, A.M.; Engel, M.; May, S.M.; Scheffers, S.R.; Joannes-Boyau, R.; Hänssler, E.; Kennedy, K.; Kelletat, D.; Brückner, H.;
Vött, A.; Schellmann, G.; Schäbitz, F.; Radtke, U.; Sommer, B.; Willershäuser, T.; Felis, T. (2014): Potential and limits of
combining studies of coarse-and fine-grained sediments for the coastal event history of a Caribbean carbonate envi-
ronment. In: Geological Society Special Publication 388 (1), pp. 503–531. DOI: 10.1144/SP388.4.
Scheffers A (2021) Tsunami boulder deposits – a strongly debated topic in paleo-tsunami research. In: Shiki, T.; Tsuji, Y.;
Yamazaki, T.; Nanayama, F. (Eds.): Tsunamiites. Elsevier, pp 353–382. DOI: 10.1016/B978-0-12-823939-1.00019-7
Schellmann, G.; Radtke, U. (2004): The marine Quaternary of Barbados. In: Kölner Geographische Arbeiten 81, pp. 1–137.
Schellmann, G.; Radtke, U.; Brückner, H. (2011): Electron spin resonance dating (Esr). In: Encyclopedia of Earth Sciences Series
Part 2, pp. 368–372. DOI: 10.1007/978-90-481-2639-2_75.
Scheucher, L.E.A.; Piller, W. E.; Vortisch, W. (2011): Foraminiferal analysis of tsunami deposits: two examples from the north-
eastern and southwestern coast of the Dominican Republic. In: Sediment 2011 – Sediments: Archives of the Earth Sys-
tem, Leipzig, June 23–26, 2011, Abstracts, pp. 86–87.
Scheucher, L.E.A.; Vortisch, W. (2011): Field survey and hydrodynamics of storm-deposited boulders in the southwestern
Dominican Republic: Playa Azul, Provincia De Barahona. In: Sediment 2011 – Sediments: Archives of the Earth System,
Leipzig, June 23–26, 2011, Abstracts, pp. 88–89.
Schimmels, S.; Sriram, V.; Didenkulova, I. (2016): Tsunami generation in a large scale experimental facility. In: Coastal Engi-
neering 110, pp. 32–41. DOI: 10.1016/j.coastaleng.2015.12.005.
Schönberger, J.J. (2017): Physikalische Modellversuche zu tsunamiinduziertem Bouldertransport. Master thesis, Institute of
Hydraulic Engineering and Water Resources Management, Faculty of Civil Engineering, RWTH Aachen University. (un-
published)
Schubert, C.; Valastro Jr., S. (1976): Quaternary geology of La Orchila Island, central Venezuelan offshore, Caribbean Sea. In:
Bulletin of the Geological Society of America 87 (8), pp. 1131–1142. DOI: 10.1130/0016-
7606(1976)87<1131:QGOLOI>2.0.CO;2.
Schubert, C. (1994): Tsunami in Venezuela. Some observations on their occurrence. In: Coastal Hazards. Perception, Suscep-
tibility and Mitigation 12 (SPEC. ISSUE), pp. 189–195.
Sedgwick, P.E.; Davis Jr., R.A. (2003): Stratigraphy of washover deposits in Florida: Implications for recognition in the strati-
graphic record. In: Marine Geology 200 (1-4), pp. 31–48. DOI: 10.1016/S0025-3227(03)00163-4.

195
References

Shaw, C.E.; Benson, L. (2015): Possible Tsunami Deposits on the Caribbean Coast of the Yucatán Peninsula. In: Journal of
Coastal Research 31 (6), pp. 1306–1316. DOI: 10.2112/JCOASTRES-D-14-00084.1.
Shaw, J.; You, Y.; Mohrig, D.; Kocurek, G. (2015): Tracking hurricane-generated storm surge with washover fan stratigraphy.
In: Geology 43 (2), pp. 127–130. DOI: 10.1130/G36460.1.
She, Z.S.; Leveque, E. (1994): Universal scaling laws in fully developed turbulence. In: Physical review letters 72 (3), pp. 336–
339. DOI: 10.1103/PhysRevLett.72.336.
Sigurdsson, H.; Sparks, R.S.J.; Carey, S.N.; Huang, T.C. (1980): Volcanogenic sedimentation in the Lesser Antilles Arc. In: Jour-
nal of Geology 88 (5), pp. 523–540. DOI: 10.1086/628542.
Smith, M.S.; Shepherd, J.B. (1995): Potential cauchy-poisson waves generated by submarine eruptions of kick ’em Jenny
volcano. In: Natural Hazards 11 (1), pp. 75–94. DOI: 10.1007/BF00613311.
Sohbati, R.; Murray, A.S.; Chapot, M.S.; Jain, M.; Pederson, J. (2012): Optically stimulated luminescence (OSL) as a chronom-
eter for surface exposure dating. In: Journal of Geophysical Research: Solid Earth 117 (9). DOI: 10.1029/2012JB009383.
Soria, J.L.A.; Switzer, A.D.; Pilarczyk, J.E.; Tang, H.; Weiss, R.; Siringan, F.; Manglicmot, M.; Gallentes, A.; Lau, A.; Cheong, A.Y.L.;
Koh, T.W.L. (2018): Surf beat-induced overwash during Typhoon Haiyan deposited two distinct sediment assemblages
on the carbonate coast of Hernani, Samar, central Philippines. In: Marine Geology 396, pp. 215–230. DOI: 10.1016/j.mar-
geo.2017.08.016.
Spiske, M.; Böröcz, Z.; Bahlburg, H. (2008): The role of porosity in discriminating between tsunami and hurricane emplace-
ment of boulders - A case study from the Lesser Antilles, southern Caribbean. In: Earth and Planetary Science Letters
268 (3-4), pp. 384–396. DOI: 10.1016/j.epsl.2008.01.030.
Spiske, M.; Jaffe, B.E. (2009): Sedimentology and hydrodynamic implications of a coarse-grained hurricane sequence in a
carbonate reef setting. In: Geology 37 (9), pp. 839–842. DOI: 10.1130/G30173A.1.
Spiske, M.; Bahlburg, H. (2011): A quasi-experimental setting of coarse clast transport by the 2010 Chile tsunami (Bucalemu,
Central Chile). In: Marine Geology 289 (1-4), pp. 72–85. DOI: 10.1016/j.margeo.2011.09.007.
Spiske, M.; Piepenbreier, J.; Benavente, C.; Bahlburg, H. (2013): Preservation potential of tsunami deposits on arid siliciclastic
coasts. In: Earth-Science Reviews 126, pp. 58–73. DOI: 10.1016/j.earscirev.2013.07.009.
Spiske, M.; Halley, R.B. (2014): A coral-rubble ridge as evidence for hurricane overwash, Anegada (British Virgin Islands). In:
Advances in Geosciences 38, pp. 9–20. DOI: 10.5194/adgeo-38-9-2014.
Spiske, M. (2016): Coral-rubble ridges as dynamic coastal features &ndash; Short-term reworking and weathering processes.
In: Advances in Geosciences 38, pp. 55–61. DOI: 10.5194/adgeo-38-55-2016.
Spurk, J.; Aksel, N. (2018): Strömungslehre - Einführung in die Theorie der Strömungen. Springer, Berlin, Heidelberg, Ger-
many. ISBN: 978-3-662-58764-5.
Srinivasalu, S.; Thangadurai, N.; Jonathan, M.P.; Armstrong-Altrin, J.S.; Ayyamperumal, T.; Ram-Mohan, V. (2008): Evaluation
of trace-metal enrichments from the 26 December 2004 tsunami sediments along the Southeast coast of India. In:
Environmental Geology 53 (8), pp. 1711–1721. DOI: 10.1007/s00254-007-0777-8.
Stein, J.; Stein, S. (2013): Rebuilding Tohoku: A joint geophysical and economic framework for hazard mitigation. In: GSA
Today 22 (10), pp. 42–44. DOI: 10.1130/GSATG154GW.1.
Stewart, S.R. (2004): Tropical cyclone report: Hurricane Ivan. In: Tropical Cyclone Report: Hurricane Ivan, National Hurricane
Center, U.S. NOAA. https://www.nhc.noaa.gov/data/tcr/AL092004_Ivan.pdf [Accessed on 22.08.2021].
Stolle, J.; Nistor, I.; Goseberg, N. (2016): Optical Tracking of Floating Shipping Containers in a High-Velocity Flow. In: Coastal
Engineering Journal 58 (2), 1650005-1-1650005-29. DOI: 10.1142/S0578563416500054.
Stolle, J.; Nistor, I.; Goseberg, N.; Mikami, T.; Shibayama, T.; Nakamura, R.; Matsuba, S. (2017a): Flood-Induced Debris Dy-
namics over a Horizontal Surface. In: Coastal Structures and Solutions to Coastal Disasters Joint Conference 2015. ASCE,
Reston, VA, pp. 54–64. DOI: 10.1061/9780784480311.006.
Stolle, J.; Nistor, I.; Goseberg, N.; Mikami, T.; Shibayama, T. (2017b): Entrainment and Transport Dynamics of Shipping Con-
tainers in Extreme Hydrodynamic Conditions. In: Coastal Engineering Journal 59 (3), 1750011-1-1750011-30. DOI:
10.1142/S0578563417500115.
Stolle, J.; Ghodoosipour, B.; Derschum, C.; Nistor, I.; Petriu, E.; Goseberg, N. (2019a): Swing gate generated dam-break waves.
In: Journal of Hydraulic Research 57 (5), pp. 675–687. DOI: 10.1080/00221686.2018.1489901.
Stolle, J.; Takabatake, T.; Hamano, G.; Ishii, H.; Iimura, K.; Shibayama, T.; Nistor, I.; Goseberg, N.; Petriu, E. (2019b): Debris
transport over a sloped surface in tsunami-like flow conditions. In: Coastal Engineering Journal 61 (2), pp. 241–255. DOI:
10.1080/21664250.2019.1586288.
Strusińska-Correira, A., Hermann, A., Freund, N., Martins, K.A., Oumeraci, H. (2017): Transport characteristics of coarse clasts
under tsunami-bore conditions based on large-scale model experiments. In: In: Costa, P.J.M., Andrade, C., Freitas, M.C.
(Eds.), Abstract Volume of the 5th International Tsunami Field Symposium.
Sugawara, D.; Minoura, K.; Imamura, F. (2008): Tsunamis and Tsunami Sedimentology. In: In: Shiki, T.; Tsuji, Y.; Yamazaki, T.;
Nanayama, F. (Eds.). Tsunamiites: Features and Implications. Elsevier. DOI: 10.1016/B978-0-444-51552-0.00003-5.
Sugawara, D., Imamura, F., Matsumoto, H., Goto, K., Minoura, K. (2011): Reconstruction of the AD869 Jogan earthquake
induced tsunami by using the geological data. In: Journal of Natural Disaster Science 29 (4), pp. 501–516. (in Japanese
with English abstract)

196
References

Sugawara, D.; Goto, K.; Jaffe, B.E. (2014): Numerical models of tsunami sediment transport - Current understanding and
future directions. In: Marine Geology 352, pp. 295–320. DOI: 10.1016/j.margeo.2014.02.007.
Sugawara, D. (2021): Lessons from the 2011 Tohoku-oki tsunami. In: Shiki, T.; Tsuji, Y.; Yamazaki, T.; Nanayama, F. (Eds.).
Tsunamiites: Features and Implications, Second Edition. Elsevier, pp. 155–181. DOI: 10.1016/B978-0-12-823939-
1.00010-0.
Suzuki, Y. (2012): Social Demand Level of Hazard Estimation and the Responsibility. In: tits 17 (8), 8_20-8_24. DOI:
10.5363/tits.17.8_20. (In Japanese)
Switzer, A.D.; Jones, B.G. (2008): Setup, deposition, and sedimentary characteristics of two storm overwash deposits, Abra-
hams Bosom Beach, southeastern Australia. In: Journal of Coastal Research 24 (1 SUPPL. A), pp. 189–200. DOI:
10.2112/05-0487.1.
Switzer, A.D.; Burston, J.M. (2010): Competing mechanisms for boulder deposition on the southeast Australian coast. In:
Geomorphology 114 (1-2), pp. 42–54. DOI: 10.1016/j.geomorph.2009.02.009.
Switzer, A.D.; Srinivasalu, S.; Thangadurai, N.; Ram Mohan, V. (2012): Bedding structures in Indian tsunami deposits that
provide clues to the dynamics of tsunami inundation. In: Geological Society Special Publication 361 (1), pp. 61–77. DOI:
10.1144/SP361.6.
Switzer, A.D.; Yu, F.; Gouramanis, C.; Soria, J.L.A.; Pham, D.T. (2014): Integrating different records to assess coastal hazards at
multi-century timescales. In: Journal of Coastal Research 70, pp. 723–728. DOI: 10.2112/SI70-122.1.
Synolakis, C.E., Bernard, E.N., Titov, V.V., Kânoğlu, U., González, F.I. (2007): Standards, criteria, and procedures for NOAA
evaluation of tsunami numerical models (U.S. NOAA Technical Memorandum OAR PMEL, 135).
https://www.pmel.noaa.gov/pubs/PDF/syno3053/syno3053.pdf [Accessed on: 22.08.2021].
Szczuciński, W. (2012): The post-depositional changes of the onshore 2004 tsunami deposits on the Andaman Sea coast of
Thailand. In: Natural Hazards 60 (1), pp. 115–133. DOI: 10.1007/s11069-011-9956-8.
Szczuciński, W.; Kokociński, M.; Rzeszewski, M.; Chagué-Goff, C.; Cachão, M.; Goto, K.; Sugawara, D. (2012): Sediment sources
and sedimentation processes of 2011 Tohoku-oki tsunami deposits on the Sendai Plain, Japan - Insights from diatoms,
nannoliths and grain size distribution. In: Sedimentary Geology 282, pp. 40–56. DOI: 10.1016/j.sedgeo.2012.07.019.
Szczuciński, W.; Niedzielski, P.; Kozak, L.; Frankowski, M.; Zioła, A.; Lorenc, S. (2007): Effects of rainy season on mobilization
of contaminants from tsunami deposits left in a coastal zone of Thailand by the 26 December 2004 tsunami. In: Envi-
ronmental Geology 53 (2), pp. 253–264. DOI: 10.1007/s00254-007-0639-4.
Szczuciński, W.; Pawłowska, J.; Lejzerowicz, F.; Nishimura, Y.; Kokociński, M.; Majewski, W. et al. (2016): Ancient sedimentary
DNA reveals past tsunami deposits. In: Marine Geology 381, pp. 29–33. DOI: 10.1016/j.margeo.2016.08.006.
Szczuciński, W. (2020): Post-depositional changes to tsunami deposits and their preservation potential. In: Engel, M.; Pilar-
czyk, J.; May, S.M.; Brill, D.; Garrett, E. (Eds.): Geological Records of Tsunamis and Other Extreme Waves. Elsevier, pp.
443–469. DOI: 10.1016/B978-0-12-815686-5.00021-3.
Taggart, B.E.; Lundberg, J.; Carew, J.L.; Mylroie, J.E. (1993): Holocene reef-rock boulders on Isla de Mona, Puerto Rico, trans-
ported by a hurricane or seismic sea wave. In: GSA Annual Meeting, October 25-28, 1993, Boston, Massachusetts, Ab-
stracts with Programs, v. 25, no. 6, p. A61.
Tamura, T.; Sawai, Y.; Ikehara, K.; Nakashima, R.; Hara, J.; Kanai, Y. (2015): Shallow-marine deposits associated with the 2011
Tohoku-oki tsunami in Sendai Bay, Japan. In: Journal of Quaternary Science 30 (4), pp. 293–297. DOI: 10.1002/jqs.2786.
Tanaka, S. (2013): Shift in Disaster Planning Paradigm Following the Great East Japan Earthquake. In: Shakaigaku Hyoron,
JSR 64 (3), pp. 366–385. DOI: 10.4057/jsr.64.366.
Tanguy, J.-C. (1994): The 1902-1905 eruptions of Montagne Pelée, Martinique: anatomy and retrospection. In: Journal of
Volcanology and Geothermal Research 60 (2), pp. 87–107. DOI: 10.1016/0377-0273(94)90064-7.
Tavoularis, S. (2006): Measurement in Fluid Mechanics. Cambridge University Press, Cambridge, United Kingdom. ISBN: 978-
0-52-181518-5.
ten Brink, U.; Twichell, D.; Geist, E.; Chaytor, J.; Locat, J.; Lee, H.; Buczkowski, B.; Barkan, R. (2008): Evaluation of tsunami
sources with the potential to impact the U.S. Atlantic and gulf coasts – an updated report to the nuclear regulatory
commission. In: U.S. Geological Survey Administrative Report. https://www.nrc.gov/docs/ML0829/ML082960196.pdf
[Accessed on: 22.08.2021].
ten Brink, U.S.; Marshak, S.; Granja Bruña, J.-L. (2009): Bivergent thrust wedges surrounding oceanic island arcs: Insight from
observations and sandbox models of the northeastern caribbean plate. In: Bulletin of the Geological Society of America
121 (11-12), pp. 1522–1536. DOI: 10.1130/B26512.1.
Terry, J., Lau, A.Y.A., Etienne, S. (2013): Reef-Platform Coral Boulders – Evidence for High-Energy Marine Inundation Events
on Tropical Coastlines. Springer, New York, U.S.A. ISBN: 978-981-4451-33-8.
Tomblin, J. (1981): Earthquakes, volcanoes and hurricanes: a review of natural hazards and vulnerability in the West Indies.
In: Ambio 10 (6), pp. 340–345.
Toscano, M.A.; Macintyre, I.G. (2003): Corrected western Atlantic sea-level curve for the last 11,000 years based on calibrated
14C dates from Acropora palmata framework and intertidal mangrove peat. In: Coral Reefs 22 (3), pp. 257–270. DOI:
10.1007/s00338-003-0315-4.

197
References

Tuttle, M.P.; Ruffman, A.; Anderson, T.; Jeter, H. (2004): Distinguishing Tsunami from storm deposits in eastern North America:
The 1929 Grand Banks Tsunami versus the 1991 halloween storm. In: Seismological Research Letters 75 (1), pp. 117–
131. DOI: 10.1785/gssrl.75.1.117.
Uchida, J.-I.; Fujiwara, O.; Hasegawa, S.; Kamataki, T. (2010): Sources and depositional processes of tsunami deposits: Analysis
using foraminiferal tests and hydrodynamic verification. In: Island Arc 19 (3), pp. 427–442. DOI: 10.1111/j.1440-
1738.2010.00733.x.
UNEP (2004): Regional Profile – Wider Caribbean Region. URL:. http://www.unep.org/regionalseas/programmes/unpro/car-
ibbean/ [Accessed on: 25.11.2015].
UNESCO (2012): Exercise CARIBE WAVE/LANTEX 13 – a Caribbean tsunami warning exercise, 20 March 2013. In: Volume 1,
Participant Handbook, Intergovernmental Oceanographic Commission Technical Series 101.
Urquhart, G.R. (2009): Paleoecological record of hurricane disturbance and forest regeneration in Nicaragua. In: Quaternary
International 195 (1-2), pp. 88–97. DOI: 10.1016/j.quaint.2008.05.012.
Valyrakis, M.; Diplas, P.; Dancey, C.L. (2011): Entrainment of coarse grains in turbulent flows: An extreme value theory ap-
proach. In: Water Resour. Res. 47 (9). DOI: 10.1029/2010WR010236.
Villholth, K.G.; Neupane B. (2011): Tsunamis as Long-Term Hazards to Coastal Groundwater Resources and Associated Water
Supplies. In: Mokhtari, M. (Ed.). Tsunami – a Growing Disaster. IntechOpen, Limited, London, U.K.. DOI: 10.5772/25125
Viret, G. (2008): Mégablocs au nord d’Eleuthera (Bahamas): preuve de vagues extrêmes au sous-stade isotopique 5e ou
restes érosionnels? In: Mégablocs Au Nord D’Eleuthera (Bahamas): Preuve de Vagues Extrêmes Au Sous-stade
Isotopique 5e Ou Restes érosionnels.
Vollmer, S.; Kleinhans, M.G. (2007): Predicting incipient motion, including the effect of turbulent pressure fluctuations in the
bed. In: Water Resources Research 43 (5). DOI: 10.1029/2006WR004919.
von Häfen, H.; Goseberg, N.; Stolle, J.; Nistor, I. (2019): Gate-Opening Criteria for Generating Dam-Break Waves. In: Journal
of Hydraulic Engineering 145 (3). DOI: 10.1061/(ASCE)HY.1943-7900.0001567.
von Hillebrandt-Andrade, C. (2013): Minimizing caribbean tsunami risk. In: Science 341 (6149), pp. 966–968. DOI:
10.1126/science.1238943.
Vose, R.S.; Applequist, S.; Bourassa, M.A.; Pryor, S.C.; Barthelmie, R.J.; Blanton, B.; Bromirski, P.D.; Brooks, H.E.; DeGaetano,
A.T.; Dole, R.M.; Easterling, D.R.; Jensen, R.E.; Karl, T.R.; Katz, R.W.; Klink, K.; Kruk, M.C.; Kunkel, K.E.; MacCracken, M.C.;
Peterson, T.C.; Shein, K.; Thomas, B.R.; Walsh, J.E.; Wang, X.L.; Wehner, M.F.; Wuebbles, D.J.; Young, R.S. (2014): Monitor-
ing and Understanding Changes in Extremes: Extratropical Storms, Winds, and Waves. In: Bulletin of the American Me-
teorological Society 95 (3), pp. 377–386. DOI: 10.1175/BAMS-D-12-00162.1.
Walsh, R.; Reading, A. (1991): Historical changes in tropical cyclone frequency within the Caribbean since 1500. In: Würz-
burger Geographische Arbeiten. 80, pp. 199–240.
Wang, D.W.; Mitchell, D.A.; Teague, W.J.; Jarosz, E.; Hulbert, M.S. (2005): Ocean science: Extreme waves under Hurricane Ivan.
In: Science (New York, N.Y.) 309 (5736), p. 896. DOI: 10.1126/science.1112509.
Wang, P.; Horwitz, M.H. (2007): Erosional and depositional characteristics of regional overwash deposits caused by multiple
hurricanes. In: Sedimentology 54 (3), pp. 545–564. DOI: 10.1111/j.1365-3091.2006.00848.x.
Ward, S.N.; Day, S. (2001): Cumbre Vieja Volcano-Potential collapse and tsunami at La Palma, Canary Islands. In: Geophysical
Research Letters 28 (17), pp. 3397–3400. DOI: 10.1029/2001GL013110.
Wassmer, P.; Schneider, J.-L.; Fonfrège, A.-V.; Lavigne, F.; Paris, R.; Gomez, C. (2010): Use of anisotropy of magnetic suscep-
tibility (AMS) in the study of tsunami deposits: Application to the 2004 deposits on the eastern coast of Banda Aceh,
North Sumatra, Indonesia. In: Marine Geology 275 (1-4), pp. 255–272. DOI: 10.1016/j.margeo.2010.06.007.
Watanabe, M.; Goto, K.; Imamura, F.; Kennedy, A.; Sugawara, D.; Nakamura, N.; Tonosaki, T. (2019): Modeling boulder
transport by coastal waves on cliff topography: Case study at Hachijo Island, Japan. In: Earth Surface Processes and
Landforms 44 (15), pp. 2939–2956. DOI: 10.1002/esp.4684.
Watanabe, M.; Goto, K.; Imamura, F. (2020): Reconstruction of transport modes and flow parameters from coastal boulders.
In: Engel, M.; Pilarczyk, J.; May, S.M.; Brill, D.; Garrett, E. (Eds.): Geological Records of Tsunamis and Other Extreme Waves.
Elsevier, pp. 617–639. DOI: 10.1016/B978-0-12-815686-5.00028-6.
Watt, S.G.; Jaffe, B.E.; Morton, R.A.; Richmond, B.M.; Gelfenbaum, G. (2010): Description of extreme-wave deposits on the
northern coast of Bonaire, Netherlands Antilles. In: USGS Open-File Report 2010-1180. DOI: 10.3133/ofr2010118.
Watt, S.; Buckley, M.; Jaffe, B. (2012a): Inland fields of dispersed cobbles and boulders as evidence for a tsunami on Anegada,
British Virgin Islands. In: Natural Hazards 63 (1), pp. 119–131. DOI: 10.1007/s11069-011-9848-y.
Watt, S.F.L.; Talling, P. J.; Vardy, M. E.; Heller, V.; Hühnerbach, V.; Urlaub, M. et al. (2012b): Combinations of volcanic-flank
and seafloor-sediment failure offshore Montserrat, and their implications for tsunami generation. In: Earth and Planetary
Science Letters 319-320, pp. 228–240. DOI: 10.1016/j.epsl.2011.11.032.
Weil Accardo, J.; Feuillet, N.; Halley, R. B.; Atwater, B. F.; ten Brink, U. S.; Deschamps, P.; Tuttle, M. P.; Wei, Y.; Fuentes Figueroa,
Z. (2012): Age of overwash and rate of relative sea-level rise inferred from detrital heads and microatolls of medieval
corals at Anegada, British Virgin Islands. In: AGU Fall Meeting 2012, Abstract T41A-2562.
Weiss, M.P. (1979): A saline lagoon on Cayo Sal, western Venezuela. In: Atoll Research Bulletin 232. DOI:
10.5479/si.00775630.232.1.

198
References

Weiss, R. (2012): The mystery of boulders moved by tsunamis and storms. In: Marine Geology 295-298, pp. 28–33. DOI:
10.1016/j.margeo.2011.12.001.
Weiss, R.; Bourgeois, J. (2012): Understanding sediments - Reducing tsunami risk. In: Science (New York, N.Y.) 336 (6085),
pp. 1117–1118. DOI: 10.1126/science.1221452.
Weiss, R.; Diplas, P. (2015): Untangling boulder dislodgement in storms and tsunamis: Is it possible with simple theories? In:
Geochemistry, Geophysics, Geosystems 16 (3), pp. 890–898. DOI: 10.1002/2014GC005682.
Williams, H.F.L. (2009): Stratigraphy, sedimentology, and microfossil content of Hurricane Rita storm surge deposits in south-
west Louisiana. In: Journal of Coastal Research 25 (4), pp. 1041–1051. DOI: 10.2112/08-1038.1.
Williams, H.F.L.; Flanagan, W. M. (2009): Contribution of hurricane rita storm surge deposition to long-term sedimentation
in Louisiana coastal woodlands and marshes. In: Journal of Coastal Research (Spec. Issue 56), pp. 1671–1675.
Williams, H.F.L. (2011): Shell bed tempestites in the Chenier Plain of Louisiana: Late Holocene example and modern analogue.
In: Journal of Quaternary Science 26 (2), pp. 199–206. DOI: 10.1002/jqs.1444.
Woodroffe, C.D. (1981): Mangrove swamp stratigraphy and Holocene transgression, Grand Cayman Island, West Indies. In:
Marine Geology 41 (3-4), pp. 271–294. DOI: 10.1016/0025-3227(81)90085-2.
Woodruff, J.D.; Donnelly, J.P.; Mohrig, D.; Geyer, W.R. (2008): Reconstructing relative flooding intensities responsible for
hurricane-induced deposits from Laguna Playa Grande, Vieques, Puerto Rico. In: Geology 36 (5), pp. 391–394. DOI:
10.1130/G24731A.1.
Wüthrich, D. (2018): Extreme hydrodynamic impact onto buildings. In: PhD dissertation, Ecole polytechnique f´ed´erale de
Lausanne, Civil Engineering Institute, Lausanne, Switzerland.
Wüthrich, Davide; Pfister, Michael; Nistor, Ioan; Schleiss, Anton J. (2018): experimental study of tsunami-like waves generated
with a vertical release technique on dry and wet beds. In: Journal of Waterway, Port, Coastal, and Ocean Engineering
144 (4). DOI: 10.1061/(ASCE)WW.1943-5460.0000447.
Yamada, M.; Fujino, S.; Goto, K. (2014): Deposition of sediments of diverse sizes by the 2011 Tohoku-oki tsunami at Miyako
City, Japan. In: Marine Geology 358, pp. 67–78. DOI: 10.1016/j.margeo.2014.05.019.
Yao, Y.; Huang, Z.; Lo, E.Y.M.; Shen, H.-T. (2014): A preliminary laboratory study of motion of floating debris generated by
solitary waves running up a beach. In: Journal of Earthquake and Tsunami 08 (03). DOI: 10.1142/S1793431114400065.
Yawsangratt, S.; Szczuciński, W.; Chaimanee, N.; Chatprasert, S.; Majewski, W.; Lorenc, S. (2012): Evidence of probable pale-
otsunami deposits on Kho Khao Island, Phang Nga Province, Thailand. In: Natural Hazards 63 (1), pp. 151–163. DOI:
10.1007/s11069-011-9729-4.
Ye, H.; Chen, Y.; Maki, K. (2021): A discrete-forcing immersed boundary method for turbulent-flow simulations. In: Proceed-
ings of the Institution of Mechanical Engineers, Part M: Journal of Engineering for the Maritime Environment 235 (1),
pp. 188–202. DOI: 10.1177/1475090220927245.
Yeh, H.H. (1991): Tsunami bore runup. In: Nat Hazards 4 (2-3), pp. 209–220. DOI: 10.1007/BF00162788.
Young, R.W.; Bryant, E.A.; Price, D.M. (1996): Catastrophic wave (tsunami?) transport of boulders in southern New South
Wales, Australia. In: Zeitschrift fur Geomorphologie 40 (2), pp. 191–207.
Zahibo, N.; Pelinovsky, E.N. (2001): Evaluation of tsunami risk in the Lesser Antilles. In: Natural hazards and earth system
sciences 1 (4), pp. 221–231. DOI: 10.5194/nhess-1-221-2001.
Zahibo, N.; Pelinovsky, E.; Yalciner, A.; Kurkin, A.; Koselkov, A.; Zaitsev, A. (2003): The 1867 Virgin Island tsunami: Observations
and modeling. In: Oceanologica Acta 26 (5-6), pp. 609–621. DOI: 10.1016/S0399-1784(03)00059-8.
Zahibo, N.; Pelinovsky, E.; Okal, E.; Yalçiner, A.; Kharif, C.; Talipova, T.; Kozelkov, A. (2005): The earthquake and tsunami of
November 21, 2004 at Les Saintes, Guadeloupe, Lesser Antilles. In: Science of Tsunami Hazards 23 (1), pp. 25–39.
Zainali, A.; Weiss, R. (2015): Boulder dislodgement and transport by solitary waves: Insights from three-dimensional numer-
ical simulations. In: Geophysical Research Letters 42 (11), pp. 4490–4497. DOI: 10.1002/2015GL063712.
Zanke, U. (2003): On the influence of turbulence on the initiation of sediment motion. In: International Journal of Sediment
Research 18 (1), pp. 17–31.
Zonneveld, J.I.S.; Buisonjé, P.H. de; Herweijer, J.P. (1977): Geomorphology and denudation processes. In: Guide to the Field
Excursions on Curaçao, Bonaire and Aruba, Netherlands Antilles, Gua Papers of Geology 1 (10), pp. 56–68. https://ede-
pot.wur.nl/487902 [Accessed on: 22.08.2021].

199
Appendix

Appendix

A. Appendix related to Tsunami deposits of the Caribbean – Towards an


improved coastal hazard assessment

Table A-1: Characteristics of sand-dominated tsunami and storm deposits as inferred from recent and selected historical and palaeo-events.
Furthermore, their potential in the separation between tsunami and storm deposits is evaluated. IOT = Indian Ocean Tsunami; PNG = Papua
New Guinea; [A] Jagodziński et al. (2012); [B] Richmond et al. (2012); [C] Szczuciński et al. (2012) – all Tohoku-oki Tsunami 2011, Sendai
plain, Japan; [D] Gelfenbaum and Jaffe (2003) – Aitape Tsunami 1998, PNG; [E] Babu et al. (2007) – IOT 2004, southwest India; [F] Uchida
et al. (2010) – specific review; [G] Jaffe et al. (2003) – Peru Tsunami 2001, south Peru; [H] Bahlburg and Weiss (2007) – IOT 2004, southeast
India; [I] Hawkes et al. (2007); [J] Naruse et al. (2010) – both IOT 2004, Andaman coast, Thailand; [K] Nichol and Kench (2008) – IOT 2004,
Maldives; [L] Szczuciński (2012) – IOT 2004, Andaman coast, Thailand; [M] Lario et al. (2016) – Chile Tsunami 2010, central Chile; [N]
Switzer et al. (2012) – IOT 2004, southeast India; [O] Morton et al. (2007) – Aitape Tsunami 1998, PNG; [P] Phantuwongraj and Choowong
(2012) – IOT 2004, Andaman coast, Thailand; [Q] Moore et al. (2011) – West Java Tsunami 2006, Java, Indonesia; [R] Kain et al. (2016) –
tsunamis of 1868 and 1960, Okains Bay, New Zealand; [S] Spiske et al. (2013) – Peru Tsunami 2001, south Peru; [T] Goff et al. (2012) –
specific review; [U] Martin Arcos et al. (2013) – Tohoku-oki Tsunami 2011, Galapagos; [V] Sawai et al. (2009) – IOT 2004, Andaman coast,
Thailand; [W] Brill et al. (2012) – IOT 2004, Andaman coast, Thailand; [X] Cuven et al. (2013) – Lisbon Tsunami 1755, southwest Spain; [Y]
Horton et al. (2011) – Chile Tsunami 2010, south-central Chile; [Z] Paris et al. (2014) – Krakatoa Tsunami 1883, Java and Sumatra, Indonesia;
[AA] Nishimura and Miyaji (1995) – Hokkaido-nansei- oki Tsunami 1993, Oshima Peninsula, Japan; [AB] Pantosti et al. (2008) – historical
tsunamis of 17 CE and 1783 CE(?), northeast Sicily, Italy; [AC] Rajendran et al. (2011) – palaeotsunami, southeast India; [AD] Kilfeather et
al. (2007) – Orphan Tsunami 1700(?), Washington State, USA; [AE] Bahlburg and Weiss (2007) – IOT 2004, Kenya; [AF] Matsumoto et al.
(2008) – IOT 2004, Andaman coast, Thailand; [AG] Minoura et al. (1994); [AH] Ishimura and Miyauchi, 2015 – both historical tsunamis,
Sanriku coast, Japan; [AI] Bahlburg and Spiske (2012) – Chile Tsunami 2010, Isla Mocha, Chile; [AJ] Nanayama et al. (2000) – Hokkaido-
nansei-oki Tsunami 1993, Oshima Peninsula, Japan; [AK] Paris et al. (2010) – IOT 2004, northwest Sumatra, Indonesia; [AL] Donato et al.
(2008) – Makran Tsunami 1945, Sur lagoon, Oman; [AM] Reinhardt et al. (2006) – historical tsunami 115 CE, Caesarea, Israel; [AN] Mamo
et al. (2009) – specific review; [AO] Davies et al. (2003) – Aitape Tsunami 1998, PNG; [AP] Putra et al. (2013) – Mentawai Tsunami 2010,
southern Mentawais, Indonesia; [AQ] Hindson and Andrade (1999) – Lisbon Tsunami 1755, south Portugal; [AR] Hussain et al. (2006) – IOT
2004, Andaman and Nicobar Islands; [AS] Kortekaas and Dawson (2007) – Lisbon Tsunami 1755, southwest Portugal; [AT] Yawsangratt et
al. (2012) – IOT 2004, Andaman coast, Thailand; [AU] Szczuciński et al. (2016) – palaeotsunami, Hokkaido, Japan; [AV] Hussain et al. (2010)
– IOT 2004, southeast India; [AW] Hemphill-Haley (1996) – Orphan Tsunami 1700(?), Washington State, USA; [AX] Dawson (2007) – Aitape
Tsunami 1998, PNG; [AY] Dura et al. (2016) – specific review; [AZ] Tuttle et al. (2004) – Grand Banks Tsunami 1929, Burin Peninsula,
Canada; [BA] Engel et al. (2009) – palaeotsunami, Bonaire, Leeward Antilles; [BB] Szczuciński et al. (2007) – IOT 2004, Andaman coast,
Thailand; [BC] Srinivasalu et al. (2008) – IOT 2004, southeast India; [BD] Costa et al. (2012) – Lisbon Tsunami 1755, south Portugal; [BE]
Costa et al. (2012) – IOT 2004, northwest Sumatra, Indonesia; [BF] Peters and Jaffe (2010) – specific review; [BG] Paris et al. (2007) – IOT
2004, northwest Sumatra, Indonesia; [BH] Abe et al. (2012) – Tohoku- oki Tsunami 2011, Sendai plain, Japan; [BI] acInnes et al. (2009) –
Kuril Island Tsunami 2006, Kuril Islands, Russia; [BJ] Morton et al. (2008b) – IOT 2004, Sri Lanka; [BK] Moore et al. (2007) – Grand Banks
Tsunami 1929, Newfoundland, Canada; [BL] Atwater et al. (2013a) – Chile Tsunami 1960, Maullín, Chile; [BM] Nanayama et al. (2000) –
Miyakojima Typhoon 1959, southwest Hokkaido, Japan; [BN] Switzer and Jones (2008) – storms in 2001, southeast Australia; [BO] Williams
(2009) – Hurricane Rita 2005, southwest Louisiana, USA; [BP] Parsons (1998) – Hurricane Andrew 1992, Louisiana, USA; [BQ] Morton et
al. (2007) – Hurricane Carla 1961, Gulf of Mexico/Hurricane Isabel 2003, western Atlantic Ocean; [BR] Wang and Horwitz (2007) – several
hurricanes 2004/2005, Florida, USA; [BS] Brill et al. (2016) – Typhoon Haiyan 2013, Visayas, Philippines; [BT] Horton et al. (2009) – Hurri-
canes Katrina and Rita 2005, Mississippi, Alabama, USA; [BU] Williams (2011) – Hurricane Ike 2008, southwest Louisiana, USA; [BV] Tuttle
et al. (2004) – Halloween Storm 1991, Massachusetts, USA; [BW] Sedgwick and Davis (2003) – recent storms, Florida; [BX] Phantuwongraj
and Choowong (2012) – recent storms, Gulf of Thailand; [BY] Shaw et al. (2015) – Hurricane Ike 2008, Texas, USA; [BZ] Boyajian and
Thayer (1995) – winter storm 1992, New Jersey, USA; [CA] Hawkes and Horton (2012) – Hurricane Ike 2008, Texas, USA; CB] Pilarczyk et
al. (2016) – Typhoon Haiyan 2013, Visayas, Philippines; [CC] Hippensteel et al. (2013) – Hurricane Irene 2011, North Carolina, USA; [CD]
Kortekaas and Dawson (2007) – post-1755 storm, southwest Portugal; [CE] Williams and Flanagan (2009) – Hurricane Rita 2005, southwest
Louisiana, USA; [CF] Morton (2010) – recent hurricanes, Mississippi, Alabama, USA; [CG] Wassmer et al. (2010) – IOT 2004, northwest
Sumatra, Indonesia; [CH] May et al. (2016) – palaeotsunami, Cape Range peninsula, Western Australia.

Sedimentary Indicative
feature Tsunami deposits Storm deposits potential Remarks

Sediment source • Mostly beach, dunes [A–C] • • Mainly shoreface, beach and dune Medium–high • Tsunami de-
[BM–BO] posits tend to
• A minor terrestrial component, in have a wider
particular in the finer, mud-domi- • Minor part from shallow embay- range of source
nated fraction [A] ments, if present [BP] areas; a bath-
yal or deeper
• Significant part (N60%) may component has
come from offshore [D,E], includ- not been re-
ing deeper shelf and bathyal zone ported from
[F] storm deposits

Lower contact • Generally sharp and erosional as • Generally sharp [BO,BQ–BS] Low • An erosive
a function • of bottom shear lower contact
stress (flow velocity, depth, • bed- • Erosional contact in few cases seems to be
load) [B,G–J] [BR,BT]; more common in the proxi- more common
mal part [BO] or in case of coarse in tsunami de-
• In some cases sharp but non-ero- shell beds [BU] posits
sional [K]

• Sharp contact may vanish within


years after deposition [L]

• In rare cases, sharp contact may


be absent even in fresh deposits
[M]

Lower contact of • Mostly sharp, in some cases • Sharp and in some cases even ero- Not indicative -
subunits even erosional [N] sional, if present [BO]

200
Appendix

Table A-1 (continuation)

Sedimentary Indicative
feature Tsunami deposits Storm deposits potential Remarks

Bedding structures • Entire deposit may be normally • In lower part mostly planar lamina- Medium to • Foreset bedding
graded [D] tion [BQ,BS], in some cases trough high if well pre- in combination
cross bedding observed [BO] served with bottom and
• Multiple normally graded subunits topsets indica-
related to the number of tsunami • Climbing ripples may occur in basal tive of storm
waves [I,O,P] part [BO] overwash

• Lower part of subunits may be in- • Foreset bedding (subaqueous depo- • Low number of
versely graded (traction carpet), sition) very common in the thicker, graded beds
but has lower preservation poten- proximal deposit [BM–BO,BV], but with basal trac-
tial due to immediate • erosion may also occur in the more distal tion carpet indic-
[J,P,Q] part ative of tsunami
deposition
• Bedding is very complex; number • [BW,BX]
of subunits and pattern of grading • Anisotropy of
varies significantly along a shore- • Foreset bedding c. 30°, planar bed- magnetic sus-
perpendicular transect [J,N] ding b4° [BW] ceptibility [CG]
or μCT [BS, CH]
• Deposit may also show no grad- • Foreset bedding is confined to a offer great po-
ing at all[K,R] thickness of N20 cm, thinner depos- tential to study
its show (sub-)planar bedding [BX] such features on
• Basal cross-bedding may occur a very small
close to the shore [H] • Topset may be clearly developed scale, even if
[BY] they are not
• Fining upward may change to macroscopically
coarsening • upward in subaerial • Often (sub-)planar lamination in visible
deposits within years • due to re- case of subaerial deposition, sepa-
moval of the fine fraction in the rated into several normally or un-
uppermost part [K,S] graded laminasets [BN,BQ,BX]

• Inverse grading may also occur [BO]


Preservation of bedding structures
may vary as a function of elevation,
bioturbation, sea-level change and
frequency of storms [BW]

Grain size distri-


bution • Function of highly variable hydro- • Mostly unimodal [BO], but can be bi- Low–me- • Storm deposits
dynamicconditions, sediment modal as function of the sediment dium tend to show
source, local topography and dis- source [BS] unimodal, tsu-
tance to shoreline [T] nami deposits
• Mostly coarser than the background bi- or multi-
• Mostly coarser than the back- sediments [BT] modal distribu-
ground sediments [P,S] tions, but site-
specific devia-
• Mostly bi- to multimodal [S], but tions from this
may also be unimodal in sand [B] pattern need to
be considered
• May be modified within years due
to winnowing of finer fraction
[S,U]

Sorting
• Mostly poor to moderate sorting; • Mostly well sorted [BO,BS,BX], but Low–me- • Storm surge
generallypoorer sorted than verti- may vary within a deposit from poor dium deposits tend to
cally confining deposits [E,P] to very well [BQ] be better sorted
than tsunami
deposits due to
more restricted
source areas;
site-specific de-
viations from
this pattern
need to be con-
sidered

Internal mud la-


minae • Sand deposit has a mud cap • Internal mud laminae not reported Medium • A low number of
where mud is comprised in the [BQ] internal mud
sediment source [D,G] laminae may be
• Mud cap on top may exist [BO] indicative of
• Thin mud drapes may be present tsunami deposi-
within sandy subunits represent- tion
ing suspension load of the wan-
ing stage between waves [G,J]

• Mud layers may contain a range


of different light objects, such as
plant remains [V]

201
Appendix

Table A-1 (continuation)

Sedimentary fea- Indicative po-


ture Tsunami deposits Storm deposits tential Remarks
Intraclasts
• Deposit may contain rip-up clasts • Rip-up clasts usually absent [BQ], Medium • Rip-up clasts of
of underlying cohesive substrate but have been reported in very rare underlying co-
[B,P,W,X] cases [BU,BX] hesive sub-
strate are much
• Intraclasts may comprise pebbles • All sorts of objects may occur as in- more common
[B,Y],pumice [Z], etc. traclasts (asphalt, tree logs, plastic in tsunami de-
litter, wood, pebbles, cobbles, boul- posits
• Intraclasts may comprise recent ders, glass bottles, etc.)
anthropogenic materials [B,AA] or [BM,BR,BV,BX]
ancient artifacts, such as bones,
teeth, glass, pottery etc.; deposit
may even be clast supported
[AB,AC]

• Rip-up clasts may occur on a mi-


croscopic scale [AD]

Heavy minerals
• Heavy mineral laminae often pre- • May be present in foreset and planar Low • Highly site-de-
sent in the lower part of subunits lamination, depending on sediment pendent: can
depending on the sediment source [BQ–BS] be useful in
source [E,AE] combination
• Higher concentration at base of nor- with a thorough
mally graded laminasets [BN] characterization
of source envi-
ronments

Truncated flame
structures/loading • Truncated flame structures in the • Not reported Medium • Even though
structures upper, finer part of a subunit may load structures
occur due to synchronous erosion are only rarely
and deformation induced by an reported from
overlying, coarser subunit [AF] tsunami depos-
its they are
• Basal loading structures may oc- highly indicative
cur in water-saturated sedimen- as they seem to
tary settings [AG,AH] be absent in
storm deposits

Backflow-induced
structures • Not necessarily present; function • Backflow structures mostly absent Low–me- • Backflow cut-
of onshore inclination and topog- [BM,BS] dium and-fill showing
raphy [AI] seaward cross-
• Some small-scale cut-and-fill struc- bedding is
• Strong component of terrestrial tures (channels) may be present in much more
deposits [AJ] upper part of the deposit [BV] common in tsu-
nami deposits
• Seaward cross-bedding or paral- whereas bidi-
lel lamination on top of the de- rectional flow
posit [N,P] can also occur
after storm in-
• Seaward ripple formation may oc- undation as a
cur on top of the deposit [AK] function of on-
shore topogra-
• Seaward imbrication of flat intra- phy
clasts may occur [AJ]

• Backflow structures may have


poor preservation potential [P]

202
Appendix

Table A-1 (continuation)

Sedimentary Indicative po-


feature Tsunami deposits Storm deposits tential Remarks
Mollusc shells
• Deposit may contain mollusc • Source dependent [BQ] Low–medium • Highly site-de-
shells of a wide range of habitats pendent, but a
in low [B] to high [AL] concentra- • Shell laminae more common in the high number
tions, largely depending on sedi- proximal deposit [BO]
ment source • of articulated
• Entire (graded) shell beds occur in shells, angular
• High percentage of articulated bi- widths of b50 m in the backbarrier breaks and
valve shells [AL] if sediment where the concentration of skeletal
source comprises bi valve shell remains is high in the foreshore • stress fractures
habitat [BU,BZ] may rather be
indicative of
• Shell fragments may predomi- • Shells are mostly disarticulated [BU],
nantly show angular breaks and even though a dominance of articu- • a tsunami depo-
stress fractures [AL] lated shells has also been observed sit
[BZ]
• Size of shell fragments may show
overall normal grading [B,X] • Fine shell hash at the top of nor-
mally graded laminasets [BN]
• Larger shells mostly convex-up
and loosely packed [AM]

• In coral reef settings, onshore de-


posits may contain larger coral
fragments and other bioclasts
[K,AK]

Foraminifera
• In general: Deposit may show al- • Presence strongly depending on • storm wave
lochtho nous marine assemblage, sediment source [BN,BS,CB] base may be
mostly dominated by shallow ma- used as indica-
rine/shelf taxa, changes in test • More diverse assemblages than ver- tor for tsunami
numbers, size, or adult/juvenile tically confining deposits [CB] Cal- deposition
ratio compared to background careous forams mostly larger in size
sedimentation, taphonomic than in vertically confining deposits
changes [AN] [CB]

• May include taxa from outer shelf • Test size decreases inland [CB]
to upper bathyal depths [AO,AP] Preservation potential of tests may
be low due to dissolution [CC]; more
• Many tests broken or abraded robust taxa preferentially preserved
[AQ,AR, AS] [BN]

• Dominance of inner-shelf and • Many tests broken [CD]


brackish/freshwater taxa, respec-
tively, may indicate uprush and
backflow component of a deposit
[I]

• Calcareous forams mostly larger


in size than in vertically confining
deposits [I]

• Poor preservation potential: tests


may dis solve quickly within years
after deposition in tropical humid
environments [AT]

• Ancient DNA may help to trace


weathered foraminifers in the de-
posit [AU]

Ostracods
• Onshore deposit may contain al- • Not yet investigated in detail – –
lochtho- • nous marine as-
semblage [AQ,AR]

• Preservation of valves may vary


between rather well [AQ] and
poor [AV]

• Low carapace/valve ratios may


indicate high-energy inflow of tsu-
nami [AV]

203
Appendix

Table A-1 (continuation)

Sedimentary Indicative po-


feature Tsunami deposits Storm deposits tential Remarks
Diatoms
• Deposit may contain diatoms • Deposit may contain diatoms from a Low • Too few infor-
from a range • of different range of different habitats from mation on dia-
habitats from freshwater and freshwater to salt marsh and shallow toms from re-
brackish to fully marine depend- marine, including planktonic taxa cent storm
ing on the sediment source [BP] deposits availa-
[C,V,AW] ble

• Larger beach and shallow sub-


tidal taxa may dominate in the
bedload component of a deposit
(lower part), while lighter marine
and/or planktonic taxa may domi-
nate the suspension load of the
upper part [V]

• Taphonomy of valves may


range from N75% broken [AX] to
mostly well preserved [V,AW]

• Concentration of diatoms in the


deposit varies, depends on sedi-
ment source: mostly low where
beach and dune sands dominate,
higher where material from ter-
restrial soils or marshes is incor-
porated [AY]

Pollen
• Dramatic changes of pollen as- • Not yet investigated in detail – –
semblage in the deposit, e.g.
from a wooded bog to saltmarsh
community [AZ] or through a sud-
den decline of mangroves [BA]

Geochemical sig-
nature • Increased concentrations in dis- • Not yet investigated in detail Low • Peaks in dis-
solved salts • (Na, Ca, Mg, Cl, S) solved salts
compared to vertically confining may indicate
deposits [BB,BC] marine flooding
but provide no
• Salts are quickly removed where evidence for the
the deposit is exposed to high depositional
rates of rainfall [BB] process; how-
ever, increases
• Deposit may be enriched with in terrestrial
trace metals where the sediment ions (e.g. Fe, K)
source is anthropogenically con- may be used to
taminated [BB,BC] trace the back-
wash compo-
nent of a tsu-
nami deposit

Microtexture on
quartz grains • Percussion marks highly abun- • Fresh surfaces dominating, low Low • Very dependent
dant where • sediment concen- number of percussion marks [CE] on local sedi-
tration in tsunami flow is high ment sources
[BD] and sediment
concentration
• Fresh surfaces dominant where during storm or
sediment concentration is lower tsunami inunda-
[BE] tion, too few
empirical stud-
ies

204
Appendix

Table A-1 (continuation)

Sedimentary fea- Indicative po-


ture Tsunami deposits Storm deposits tential Remarks
Thickness • Mostly b30 cm, localized up to 150 • Mostly thick and narrow, N30 cm up
cm, depending on sediment avail- to N1 m, thinning landward [BO,BQ, Low • Highly variable
ability, hydrodynamic conditions BV] and a function
of local topogra-
and onshore topography [BF] • Thickest in topographic lows [BO,CA] phy, sediment
• Usually thickest part up to several availability and
tens of metres landward behind a magnitude of
zone of erosion, then landward the flooding
thinning [BG]
Landward extent • Strongly depending on size of the • Washover fan or sand sheet thins
and fines inland [BO,CE] Medium • Tsunami de-
tsunami,sediment availability, and
posits seem to
local topography [BF] • Landward thinning often abrupt [BO] have a relative
• Sand sheet deposition may occur Storm overwash often builds (coa- greater inland
up to several km inland on a low- lesced) fans behind barrier breaches extent and tend
angle coastal plain [B,BG] [BS,CF] to occupy
higher eleva-
• Relation between landward extent tions than storm
of the deposit and inundation may deposits
vary between

50–60% [BH] and 90% [BI]


• May be locally discontinuous
[Q,BJ]
• In most cases landward fining
trend [C,G,BK]
• May breach barriers and build fans
[BL]

205
Appendix

Table A-2: Characteristics of coarse-clast tsunami and storm deposits as inferred from recent and selected historical and
palaeo-events. Furthermore, their potential in the separation between tsunami and storm deposits is evaluated. IOT = Indian
Ocean Tsunami; BVI = British Virgin Islands; [A] Richmond et al. (2011) – tsunamis of 1868 and 1975, Hawai'i; [B] Yamada
et al. (2014) – Tohoku-oki Tsunami 2011, Sanriku coast, Japan; [C] Rahiman et al. (2007) – Suva Tsunami 1953, Fiji; [D]
Terry et al. (2013) – specific review; [E] Goto et al. (2007) – IOT 2004, Andaman coast, Thailand; [F] Kelletat et al. (2007)
– IOT 2004, Andaman coast, Thailand; [G] Paris et al. (2009) – IOT 2004, northwest Sumatra, Indonesia; [H] McAdoo et al.
(2011) – South Pacific Tsunami 2009, Samoa; [I] Buckley et al. (2012) – historical tsunami, Anegada, BVI; [J] Nandasena
et al. (2013) – Tohoku-oki Tsunami 2011, Sanriku coast, Japan; [K] Goto et al. (2012) – Tohoku-oki Tsunami 2011,
Sabusawa Island, Japan; [L] Bourgeois and MacInnes (2010) – Central Kuril Island Tsunami 2006, Matua Island, Kuril
Islands; [M] Goto et al. (2010) – Meiwa Tsunami 1771, Ishigaki Island, Japan; [N] Paris et al. (2014) – Krakatoa Tsunami
1883, west Java, Indonesia; [O] Etienne et al. (2011) – specific review; [P] Bahlburg and Spiske (2012) – Chile Tsunami
2010, IslaMocha, Chile; [Q] Razzhigaeva et al. (2006) – IOT 2004, Simeulue Island, Indonesia; [R] Scheffers (2002b, 2004)
– palaeotsunamis, Bonaire, Leeward Antilles; [S] Scheffers and Kelletat (2003) – specific review; [T] Scheffers et al. (2006)
– Hurricane Ivan 2004, Bonaire, Leeward Antilles; [U] Khan et al. (2010) – Hurricane Dean 2007, Jamaica; [V] Goto et al.
(2011) – recent typhoons,Okinawa Islands, Japan;[W] Cox et al. (2012) – recentwinter storms, Aran Islands, Ireland; [X]May
et al. (2015b) – TyphoonHaiyan 2013, Eastern Samar, Philippines; [Y] Maragos et al. (1973) – Tropical cyclone Bebe 1972,
Funafuti Atoll, Tuvalu; [Z] z-Avila et al. (1977) – (sub-)recent hurricanes, Grand Cayman; [AA] Spiske and Jaffe (2009) –
Hurricane Lenny 1999, Bonaire, Leeward Antilles; [AB] Etienne and Terry (2012) – Tropical cyclone Tomas 2010, Fiji; [AC]
Etienne (2012) – Tropical cyclone Oli, French Polynesia; [AD] Spiske andHalley (2014) – historical hurricanes, Anegada,
BVI; [AE] Kennedy et al. (2015) – TyphoonHaiyan 2013, Eastern Samar, Philippines; [AF] Brill et al. (2016) – TyphoonHai-
yan 2013, central Visayas, Philippines; [AG] Miller et al. (2014) – historical hurricanes, Jamaica; [AH] Cox et al. (2016) –
winter storms 2013–2014, Aran Islands, Ireland; [AI] Etienne and Paris (2010) – subrecent storms, Reykjanes Peninsula,
Iceland; [AJ]Morton et al. (2008a) – past hurricanes, Caribbean; [AK] Blumenstock et al. (1961) – Typhoon Ophelia 1958,
Marshall Islands; [AL] Richmond et al. (2011) – historical storms, Hawai'i; [AM] Scheffers et al. (2014) – past hurricanes,
Bonaire, Leeward Antilles; [AN] Scheffers (2002b) – past hurricanes, Bonaire, Leeward Antilles; [AO] Spiske (2016) – Hur-
ricane Earl 2010, Anegada, BVI.

Sedimentary/ge-
omorphic fea- Indicative
ture Tsunami deposits Storm deposits potential Remarks
Sediment source • Subaerial cliff edges [A,B] • Subaerial cliff edges or cliff top [T–X] Low • Boulder sources very similar
• Reef crest or slope [C,D] • Reef crest or slope [D,V] for transport during tsunamis
• Living or dead coral colonies or • Living or dead coral colonies or large and storms
large coral slabs in shallow water coral slabs in shallow water [T,Y–AA]
[E–G] • Exhumed beachrock outcrops [AB]
• Exhumed beachrock outcrops • Artificial structures such as rip rap boul-
[G,H] ders [AC]
• Onshore rock outcrops [I,J] • Most coarse clasts rest in subtidal
• Intertidal or onshore artificial struc- [T,AA,AD–AF], intertidal [V] or supratidal
tures: seawalls, rip-rap boulders, [X,AE] position prior to transport as step-
roads, concrete structures, etc. wise transport inland is very common
[B,G,K] [AG]

• Most boulders rest in subtidal [L] to


supratidal [B] position prior to
transport
Singular mega- • Very large isolated boulders or • Very large isolated boulders or fine Low-medium • Boulder size or weight alone
boulders fine blocks up to several hundred blocks up to several hundred tons in in- is not suitable to distinguish
tons [B,J,M,N] tertidal [V,W] or cliff-top [X,AH] position between tsunami and storm
• Distributed in intertidal or suprati- • Only short transport distance for largest • The transport distance of very
dal position, up to 750 m inland boulders N100 t recorded [U,W,AH] large tsunami boulders
where onshore topogra- • phy • Presence is a function of availability, seems generally greater; they
is low [B,M] storm magnitude, and other processes are distributed further inland
• Presence is a function of availa- associated with the storm, such as infra-
bility and tsunami magnitude [O] gravity waves [W] or vertical jets at cliffs
[AH]

206
Appendix

Table A-2 (continuation)


Sedimen-
tary/geomor- Indicative
phic feature Tsunami deposits Storm deposits potential Remarks
Boulder fields • Fields of scattered boulders very • Fields of boulders occur, for instance, on Medium • Landward fining in a boulder
common [A,E,G] intertidal reef flats [V], but their inland distribution seems to be indic-
• In most cases no shore-perpen- extent is rather small [AE] ative of repeated storm im-
dicular fining trend [A,E,G,P] • Exponential landward fining is common pact in case non-linear wave
• Abrupt landward fining only at [V,AE,AI] phenomena such as infra-
the landward limit of boulder dis- gravity waves do not play a
tribution [B] role
• General landward fining may oc-
cur where return flow plays a
major role [K]
• Boulders may show imbrication
[G]
Linear construc- • Not observed in tsunamis • Highly depending on availability of sedi- High • Large shore-parallel con-
tional landforms [A,D,K,O] ments, clast size, and wave run-up [AJ] structional landforms are
(ramparts, ridges) Upper intertidal ridges or ramparts
• Coarse-clast ridges may form highly indicative of long-term
right after the tsunami as an im- (broader ridge-complexes) of coral rub- formation dominated by re-
mediate response to the erosive ble close to the coast or along the edge curring storms
event [G] of a shallow reef flat
• Small rampart observed after the [T,Y,Z,AB,AF,AK,AL]
IOT 2004 was attributed to low- • Small, narrow ridges, cliff-top ridges
amplitude sea-level variations at [W,AL] and ramparts [AM] in elevations
the final stage of the tsunami [Q] up to 40 m [W]
• In some cases, ramparts have • Ridges more narrow, steeper, better
been associated with palaeotsu- sorted, unimodal, from gravel to boulder
namis [R,S] size, low sand content, up to several
metres high [T,AM], may be segmented
or cuspate in higher position [AL]
• Ramparts (ridge complexes) very wide
(up to 200 m), steep seaward side, low
inclination inland (wedge-shaped),
poorly sorted (sand to boulder), multi-
modal, layered [Y,AJ,AM,AN]
• Smaller, narrow ridges close to sea
level may be built during one single
storm [T,AF], but are prone to rework-
ing by subsequent events [AK,AO]
• Coarse-clast ridges may form washover
lobes at their landward side [AA,AK]
• Clast imbrication may indicate flow di-
rection [AA,AJ]
• Ridges/ramparts may extend for sev-
eral kilometres parallel to the shore
[X,Z,AM]

207
Appendix

Table A-3: Sediment-based reconstructions of palaeoenvironments and coastal processes reviewed for evidence of extreme-wave events. All sites are indicated in Figure 2-3. For all sites, the 30-year
probability of tsunami run-up N0.5m(Parsons and Geist, 2009), the calculated amplitude of the Virgin Island Tsunami of 1867 (Zahibo et al., 2003), tsunami height in the SCDB scenario (Figure 2-7), and
the decadal number of hurricanes (Reading, 1990) is shown. The depositional process is coded as follows: 0= no disturbance in the stratigraphy/no coarse-clast record evident; 1=disturbance/deposition
attributed to storm; tsunami impact is excluded; 2=disturbance/deposition documented and storm is favoured over tsunami, but the latter is not entirely excluded; 3=disturbance/deposition is ascribed to
either hurricane or tsunami; 4=disturbance/deposition is documented and tsunami is favoured over hurricane, but the latter is not entirely excluded; 5=disturbance/deposition is documented, ascribed to
tsunami, and hurricane is excluded.

Interpretation
Time covered Tsunami in the original
by sediment ar- Age of the can- 30-year prob. Of height Decadal source (see or-
chive (years Type of sedi- didate tsunami tsunami run-up Amplitude 1867 SCDB sce- number of der of literature
No. Author(s) Goal of the study Location BP) ment archive Evidence for extreme wave events deposit > 0.5 m tsunami (m) nario (m) hurricanes sources)
1 McCloskey and Reconstruction of Gales Point, 5000–0; C/D: Swamp, Hurricanes inferred, since absence - 0 (Belize City n/d n/d 5-9 1
Keller (2009); palaeotempests Mullins River, 7000–0 lagoon of historical accounts on tsunamis
McCloskey and Commerce Bight and low tsunami potential of the area
Liu (2013a) (Belize)

2 Adomat and Reconstructing Holo- Manatee La- c. 8000–0 C/D: Coastal Multiple hurricane overwash and - 0 (Belize City) n/d n/d 5-9 1; 2
Gischler (2015, cene coastal envi- goon, Colson lagoons, sediment deposition is inferred at all
2016) ronments and pal- Point Lagoon, mangrove sites, tsunamis seem unlikely due to
aeo-hurricanes Commerce Bight swamp the low tsunami potential of the area
Lagoon, Sapo-
dilla Lagoon,
Mullins River
Beach (Central
Belize)

3 Macintyre et al. Reconstructing rates Tobacco Range; c. 7000–0 C: Mangrove No indication for extreme-wave - 0 (Belize City) n/d n/d 5–9 0
(1995, 2004); of peat formation Twin Cays (Be- island events; sand lenses interpreted as
McKee et al. and mangrove com- lize) shoreline migration
(2007) plexes

4 Gischler (2003); Reconstructing la- Turneffe Islands, 8500–0 A/C: Man- Shell beds interpreted as reworking - 0 (Belize City) n/d n/d 5-9 1:1:1:1/2
Gischler et al. goonal development Glovers Reef, grove island, by storms and/or bioturbation, e.g.
(2008); McClos- and climate varia- Lighthouse Reef shallow reef, approx. 4500 BP ago; pattern of
key and Liu tions (Belize) subaquatic mixed layer of peat clasts in a car-
(2013b); De- sinkhole bonate matrix pure carbonate depos-
nommee et al. its bracketed by mangrove peat are
(2014) ascribed to seismic activity or hurri-
canes

5 Monacci et al. Reconstructing man- Spanish Lookout 8000–0 C: Mangrove Continuous peat record, no disturb- - 0 (Belize City) n/d n/d 5-9 0
(2009) grove dynamics Cay (Belize) island ance detected

6 Brown et al. Evaluating the po- Laguna Chum- c. 7100–6700 A: 80 m deep Moderate to strong hurricanes are - 5-10 n/d 0.21 10-14 0
(2014) tential of karst sink- kopó (Yucatán, and the last karst sinkhole represented by thin graded se-
holes for archiving Mexico) 50 years quences of resuspended material; no
historical hurricanes tsunamis discussed

7 Woodroffe Spatio-temporal re- North Sound >2100–0 C: Mangrove No extreme-wave impact inferred - 10.79 n/d 0.25 10-14 0
(1981) construction ofmaine (Grand Canyan) swamp
transgression

8 Peros et al. Reconstructing pal- Laguna de la c. 8000–0 E: Coastal Storm impact inferred based on - c. 0 n/d n/d 10-14 0
(2007a, b) aeoenvironmental Leche (Cuba) lake changes in C and O isotope ratios
change

208
Appendix

Table A-3 (continuation)

Tsu- Interpreta-
nami tion in the
Time cov- height original
ered by Age of the 30-year SCDB Decadal source (see
sediment Type of candidate prob. Of tsu- Amplitude sce- number order of lit-
archive sediment Evidence for extreme wave tsunami nami run-up 1867 tsu- nario of hurri- erature
No. Author(s) Goal of the study Location (years BP) archive events deposit > 0.5 m nami (m) (m) canes sources)

Hendry (1987) Determining tectonic The Great Mo- Holocene C/D/F: La- No extreme-wave impact inferred - c. 5 n/d 0.33 5-9 0
and eustatic influ- rass (northwest goon, beach,
9
ence on sedimenta- Jamaica) wetland
tion

10 Palmer and Identifying marine Manatee Bay “Millennial D: Enclosed Several overwash strata, overwash - c. 22 n/d 1.02 5-9 3
Burn (2012) washover events (Jamaica) scale” lagoon fan in the east

11 Hearty et al. Landward-pointing Northwestern MIS 5e A: Shallow Ridges with thick, tabular beds with End of MIS c. 0–5 0.18 10–14 1;0
(1998, 2002); v-shaped chevron Bahamas shelf fenestrae at high elevations indicate 5e
Kindler and ridges, partly extreme storm-wave action (Hearty
Strasser (2000, et al., 1998); Kindler and Strasser
subaerial
2002) (2000) suggest an aeolian origin, fe-
nestrae are attributed to rain

12 Dix et al. (1999) Reconstructing pal- Lee Stocking Is- 1500–0 E: Shallow Possible storm surge overwash in- - c. 0 n/d n/d 10-14 1
aeoenvironmental land (Bahamas) ponds ferred from ex-situ marine deposits
change

13 Park (2012); Evaluating ecologi- San Salvador Is- up to N3500– E: Shallow Salt pond: tsunami overwash possi- - c. 3–5 n/d n/d 10-14 0;2;1;1
Park et al. cal response to hur- land (Bahamas) 0 pond ble, but considered to be unlikely,
(2009); Dalman ricane impacts due to lack of modern tsunami depo-
and Park sition and presence of correlating
(2012); Mat- freshening (rainwater input); 19
theus and events marked by peaks in sand
Fowler (2015) (Park et al., 2009); Clear Pond and
others: sandy overwash deposits
identified in the sediment cores are
interpreted as hurricane-induced
(Dalman and Park, 2012; Mattheus
and Fowler, 2015)

14 Caffrey et al. Reconstructing envi- Laguna Saladilla c. 8000–0 E: Freshwa- Strom surge incursions inferred from - c. 0–5 n/d 0.15 5-9 1
(2015) ronmental change ter lake (c. 5 peak occurrences in marine diatoms
km from the
coast)

15 Fuentes and Searching for tsu- Bahia de Ocoa ? D: Shallow Several overwash deposits of coarse - 5–7 0.2–0.5 2.29 5-9 3
Huérfano- nami deposits (Dominican Re- lagoons sand detected in cores (Barahona, (Penin-
Moreno (2013) public) southwestern sula de
Dominican Peder-
Republic) nales)

16 Scheucher et Detecting tsunami Puerto Viejo 300–0? Beach (?) Sand layer is ascribed to a historical 18 October 5–7 0.2–0.5 2.29 5-9 5
al. (2011) deposits and forami- (southwestern tsunami 1751 tsu- (Barahona, (Penin-
niferal characteriza- Dominican Re- nami southwestern sula de
tion public) Dominican Peder-
Republic) nales)

209
Appendix

Table A-3 (continuation)


Interpreta-
tion in the
➢ Time cov- Tsunami original
ered by Age of the 30-year height Decadal source (see
Type of
sediment candidate prob. Of tsu- Amplitude SCDB number order of lit-
archive sediment Evidence for extreme wave tsunami nami run-up 1867 tsu- scenario of hurri- erature
No. Author(s) Goal of the study Location (years BP) archive events deposit > 0.5 m nami (m) (m) canes sources)

17 Scheucher et Detecting tsunami Playa Cosón ➢ ? Beach (?) Sand layer is ascribed to the 8 Au- 08 August c. 10 n/d 0.30 10-14 5
al. (2011) deposits and forami- (northeastern gust 1946 tsunami 1946 tsu-
niferal characteriza- Dominican Re- nami
tion public)

18 Morton et al. Reconstructing pre- Aguadilla– ➢ 2500–0 C/F: swamps 1–2 thin, possibly tsunamigenic sand (Post-) c. 10–20 Run-up 11 1.04 10-14 5
(2006); Moya historical tsunamis Rincón (north- layers in various depths at different 2770–2350 October
and Mercado western Puerto locations, no age estimates; over- BP, (post-) 1918 tsu-
(2006); Jaffe et Rico) wash deposits correlated with histori- 680–540 BP, nami: 3–6 m
al. (2008) cal and prehistorical tsunamis 11 Oct. 1918

19 Morton et al. Evaluating coastal Punta Cucha- ? C: Wetlands A grey sand-and-shell layer with a ? c. 15 3.3–5.1 2.03 10-14 3
(2006) hazards using sedi- ras, south thickness of only 2 cm was found 1 (Guayama,
mentary records Puerto Rico m below a tidal flat. Depositional pro- Puerto Rico)
cesses remain uncertain

20 Donnelly and Reconstructing fluc- Southwestern 5000–0 D/E: Coastal Periodic hurricane overwash in- - c. 15 3.0–3.5 m 0.62 10-14 1
Woodruff tuations in hurricane Vieques (Puerto lagoon ferred; tsunamis excluded tsunami (eastern
(2007); Wood- frequency Rico) wave ampli- Puerto
ruff et al. (2008) tude (south- Rico)
east coast of
Vieques), 3
main waves
within 30
min.

21 Donnelly (2005) Reconstructing se- Isla de Culebrita 2200–0 D/E: Back- Massive sand layer interpreted as c. 15 6.1 m at ad- 0.62 10–14 2
vere tropical cy- (Puerto Rico) barrier salt 1867 hurricane-born; 1867 tsunami jacent Cule- (eastern
clones pond also possible bra Puerto
Rico)

22 Brooks et al. Identifying human in- St. Thomas, St.➢ 5000-0 Different Multiple marine overwash inferred - 13.85% (Road 3.2–7.5 0.45 10-14 1;1;2
(2007, 2015); fluence on sedimen- Croix, St. John coastal envi- from N100 sediment cores Town, BVI) (Cruz Bay, (Anegada)
Larson et al. tation (USVI) ronments, St. John)
(2015) marginal ma-
rine embay-
ment

23 Atwater et al. Identifying marine Central and c. 400-0 D/E: Coastal Tsunami or storm overwash inferred; - c. 10-15 2.3–5.2 0.45 10-14 3;3;4
(2012); Pi- Overwash western negada ponds taphonomic characteristics and spa- (Road Town,
larczyk and (BVI) tial extent of a shell-rich sand sheet Tortola)
Reinhardt indicate tsunami deposition
(2012); Rein-
hardt et al.
(2012)

24 Jessen et al. Reconstructing pal- Altona Bay (St. c. 4500-0 D: Lagoon One event inferred, post-1960 CE 1650–1800 c. 10-15 3.0–3.9 n/d 10-14 1
(2008) aeoenvironments Croix, USVI) Hurricane CE

25 Bertran et al. Determining variabil- Grand-Case (St. c. 4000-0 D/E: Coastal Up to 21 event layers; sand layers - c. 3-5 Run-up 1867 0.52 (St. 10-14 2
(2004); Malaizé ity of hurricanes Martin) pond ascribed to hurricanes, but authors tsunami: c. Kitts and
et al. (2011) “cannot exclude that some sand lay- 1.5 m Nevis)
ers correspond to tsunamis”

210
Appendix

Table A-3 (continuation)

Interpreta-
tion in the
Time cov- Tsunami original
ered by Age of the 30-year height Decadal source (see
sediment Type of candidate prob. Of tsu- Amplitude SCDB number order of lit-
archive sediment Evidence for extreme wave tsunami nami run-up 1867 tsu- scenario of hurri- erature
No. Author(s) Goal of the study Location (years BP) archive events deposit > 0.5 m nami (m) (m) canes sources)

26 Caron (2011, Evaluating the St. Bartholomew ? Beachrock Considered as unlikely to be tsu- - < 10 n/d 0.52 (St. 10-14 2
2012) preservation poten- namigenic Kitts and
tial of extreme wave Nevis)
deposits by
beachrock

27 Morton et al. Evaluating coastal Northwestern ? C: Coastal No marine overwash identified - c. 10-12 0.6–1.6 0.88 10-14 0
(2006) hazards using sedi- Basse-Terre plains and (Basse-
mentary records (Guadeloupe) wetlands Terre)

28 Ramcharan Reconstructing Hol- Maracas Bay 7000-0 C: Swamps No marine sedimentary impact - 0 (Port-of- 0.5–1.2 0.45 5-9 0
(2004); Ram- ocene wetland de- (Trinidad) Spain) (Port-of-
charan and velopment Spain)
McAndrews
(2006)

29 Oropeza et al. Identification of pal- Bocaripo, Cha- ? D: Lagoon Preliminary findings: 30 cm-thick al- After 960 ± 6.27 (Cu- 0.3–1.0 0.28-0.49 5-9 5
(2015) aeotsunami deposits copata, and ternating sequence of reddish sand 30 BCE? maná)
Morro de Cha- and beige sandy silt (Bocaripo); mol-
copata lagoons lusc-rich layer (36 cm-thick) might
(Sucre, north- correspond to tsunami impact (Cha-
east Venezuela) copata); discontinuous, 3 cm-thick
sand layers with landward-directed
ripples may indicate tsunami incur-
sion (Morro de Chacopata)

30 Leal et al. Identification of pal- Los Patos La- 5000-? (age D/E: Coastal Tsunami deposits identified in the Historical? 6.27 (Cu- 0.3–1.0 0.30 5-9 5
(2014) aeotsunami deposits goon near Cu- inversions) lagoon upper part of the sediment core, age maná)
maná (northeast remains unknown
Venezuela)

31 Weiss (1979) Sedimentological Cayo Sal (Vene- 3800-0 D: Enclosed Hurricane or tsunami is suggested to 770–500 BP c. 3–5 0.6–0.8 0.94 2-4 3
and geomorphologi- zuela) lagoon have disturbed sedimentation of (Puerto Ca-
cal characterization Cayo Sal bello)

32 Engel et al. Evaluating the im- Bonaire 8000-0 C/D/F: En- Potential tsunamis inferred 3600 or c. 7 (Willem- 0.1–0.8 Up to 1.30 2-4 4
(2010, 2012, pact of extreme closed la- later, 3300 stad, Curaçao)
2013) wave events on Hol- goons, flood- or later,
ocene coastal plains 2000 BP or
later, plus
ecosystems
one younger
event

33 Klosowska Reconstructing Lagoon St. 5000-0 D: Saline La- Rapid closure of lagoon at around At around c. 5–10 n/d 0.93 2-4 3
(2003) coastal palaeoenvi- Michiel (Cura- goon 3500 BP might be related to tsunami 3500 BP
ronments çao) impact

34 Hornbach et al. Reconstructing Spaanse Wa- c. 1000-0 A: Sheltered Surface Halimeda sand attributed to 460–310 cal 7.04 (Willem- Negligible 1.57 2-4 4
(2008a) storm and tsunami ters, Lagoen coastal em- recent storms; “chaotic mixed” layer BP stad, Curaçao)
events Jan. Thiel, Fuik bayments interpreted as tsunamigenic
Bay (Curaçao)

211
Appendix

Table A-3 (continuation)

Interpreta-
tion in the
Time cov- Tsunami original
ered by Age of the 30-year height Decadal source (see
sediment Type of candidate prob. Of tsu- Amplitude SCDB number order of lit-
archive sediment Evidence for extreme wave tsunami nami run-up 1867 tsu- scenario of hurri- erature
No. Author(s) Goal of the study Location (years BP) archive events deposit > 0.5 m nami (m) (m) canes sources)

35 González et al. Reconstructing man- Bahia Honda c. 3000-0 (hi- C: Mangrove Major erosive disturbance event: - c. 3 Negligible n/d 2-4 1
(2010) grove dynamics (San Andres, atus from c. embayment 1605 hurricane?
Colombia) 2500-
500/400)

36 Urquhart Evaluating man- Bluefields 8000-0; 5000- E: Lagoon, Single major tsunami or hurricane (Post-)3340 0 0–0.2 0.47 2-4 2;1
(2009); McClos- grove response to (southeast Nica- 0 17 km from (Urquhart, 2009); clastic layers are ± (Puerto Cab-
key and Liu hurricane impacts ragua) the coast; interpreted as hurricane-induced due ezas)
40 BP
(2012) coastal to low tsunami potential of the region
marsh (McCloskey and Liu, 2012) (Urquhart,
2009)

212
Appendix

Table A-4: Coastal coarse-clast deposits reviewed for evidence of extreme-wave events. All sites are indicated in Figure 2-3. For all sites, the 30-year probability of tsunami run-up N0.5m(Parsons and
Geist, 2009), the calculated amplitude of the Virgin Island tsunami of 1867 (Zahibo et al., 2003), tsunami height in the SCDB scenario (Figure 2-7), and the decadal number of hurricanes (Reading, 1990)
is shown. The depositional process is coded as follows: 0=no disturbance in the stratigraphy/no coarse-clast record evident; 1=disturbance/deposition attributed to storm; tsunami impact is excluded;
2=disturbance/deposition documented and storm is favoured over tsunami, but the latter is not entirely excluded; 3=disturbance/deposition is ascribed to either hurricane or tsunami; 4 = disturbance/dep-
osition is documented and tsunami is favoured over hurricane, but the latter is not entirely excluded; 5 = disturbance/deposition is documented, ascribed to tsunami, and hurricane is excluded.

Interpretation
Time covered Tsunami in the original
by sediment Age of the can- 30-year prob. height Decadal source (see or-
archive (years Type of sedi- didate tsunami Of tsunami run- Amplitude 1867 SCDB sce- number of der of literature
No. Author(s) Goal of the study Location BP) ment archive Evidence for extreme wave events deposit up > 0.5 m tsunami (m) nario (m) hurricanes sources)
37 Shaw and Ben- Investigation of an Tankah to c. 1500–0 A: Coastal Elevated coastal berm out of reach c. 1500 BP 5-10 0.21 10-14 4
son (2015) anomalous sand and Puerto del Car- berm (sand for hurricane waves, lack of distinct
coarse-clast berm men (Yucatán, and rubble) internal bedding, covered by boul-
Mexico) ders, landward extent of up to 400 m

38 Jones and Study effect of ex- Great Pedro Mid- to late B: Elevated Large boulders of up to 40 t at- Two boul- 10.79 n/d 0.25 10-14 3;2
Hunter (1992); treme-wave condi- Point/Blowholes Holocene; carbonate tributed to deposition during hurri- ders moved
Robinson et al. tion on rocky coasts (Grand Cayman) historical platform (+ canes or tsunamis; Robinson et al. around
(2006); Rowe et 7–12 m.s.l.) (2006) favour storms 1625–1688
al. (2009) CE or later

39 Rigby and Rob- Calculate storm- Spots Bay, Mid- to late A/B: Elevated Ramparts of coral pebbles and boul- - 10.79 n/d 0.25 10-14 1
erts (1976); wave force required Breakers, Old Holocene carbonate ders were accumulated mainly dur-
Hernandez- to induce breakup of Isaacs (Grand platform up to ing hurricanes of 1931 and 1932
Avila et al. corals as source ma- Cayman) 4.5 m a.s.l.
(1977); Jones terial of ramparts;
and Hunter study effect of ex-
(1992) treme-wave condi-
tion on rocky coasts

40 Robinson et al. Identification of geo- Several sites on Mid- to late B: Elevated Large boulders mostly shifted by hur- 20th cen- Up to c. 22 n/d Up to 5-9 3
(2006); Morton logical evidence of Jamaica Holocene carbonate ricanes; for some boulders tsunamis tury? 2.01
et al. (2008a); tsunamis in service platform are not excluded; boulders of up to
Rowe et al. of coastal hazard as- 80 t were moved during Hurricane
(2009); Khan et sessment Dean in 2007; two particular boul-
al. (2010); Miller ders moved during the 20th century
et al. (2014)

41 Hearty (1997); Reconstruction of Glass Window, MIS 5e B: Elevated Large boulders and blocks of up to End of MIS <2 n/d 0.18 10-14 3;0;0;4;4;4;3;1
Hearty et al. transport mechanism northern Eleu- carbonate 2000 t attributed to tsunamis or 5e? (see details in
(2002); Mylroie and time of deposi- thera (Bahamas) platform (10– storms exceeding the magnitude of column “evi-
(2008); Saman- tion of large coastal 20 m a.s.l.); the largest storms known in history dence for ex-
kassou et al. boulders intertidal plat- (Hearty, 1997; Hearty et al., 2002); treme wave
(2008); Viret form; evidence for wave emplacement was events”)
(2008); Hasler emerged fos- challenged by Mylroie (2008) associ-
et al. (2010); sil beach (3 ating the boulders with tower karst
Kindler et al. m a.s.l.); features; Samankassou et al. (2008),
(2010); Hansen washover ba- Hasler et al. (2010) and Kindler et al.
et al. (2016) sin (6–8 m (2010) interpret them as tsunami-em-
a.s.l.) placed boulders based on assump-
tions of lower transport capacity of
storm waves and speculate on a
landslide tsunami either on the Ca-
nary Islands or the adjacent platform
margin; Hansen et al. (2016) link the
boulders with superstorms unprece-
dented in the historical era

213
Appendix

Table A-4 (continuation)


Tsu- Interpreta-
nami tion in the
Time cov- height original
ered by Age of the 30-year SCDB Decadal source (see
sediment Type of candidate prob. Of tsu- Amplitude sce- number order of lit-
archive sediment Evidence for extreme wave tsunami nami run-up 1867 tsu- nario of hurri- erature
No. Author(s) Goal of the study Location (years BP) archive events deposit > 0.5 m nami (m) (m) canes sources)

42 Kelletat et al. Mapping of coarse- Eleuthera (Ba- Mid- to late B: Elevated Boulders of up to 200 t, partly imbri- 3000 and <2 n/d 0.18 10-14 5
(2004) clast tsunami depos- hamas) Holocene carbonate cated, and bimodal deposits indicate 500 BP
its platform (5– tsunami wave heights of up to 20 m (based on
15 m a.s.l.) uncal. 14C
data)

43 Kelletat et al. Mapping of coarse- Long Island (Ba- Mid- to late A/B: (Ele- Large boulders of up to 50 t, boulder 3000, <2 n/d 0.27 10-14 5
(2004) clast tsunami depos- hamas) Holocene vated) car- and pebble fields and boulder ridges 1600(?), and
its bonate plat- (bimodal) 500 BP
form, upper (based on
supratidal up uncal. 14C
to 15 m a.s.l. data

44 Scheucher and Reconstructing ex- Playa Azul (Do- Mid- to late A: Elevated Boulders up to 7 t form a ridge paral- - 5-7 0.2–0.5 2.29 5-9 1
Vortisch (2011) treme-wave events minican Repub- Holocene carbonate lel to the shoreline, 1–1.5 m a.s.l., in- (Barahona, (Penin-
from coastal depos- lic) platform terpreted as the result of storms southwestern sula de
its Dominican Peder-
Republic) nales)

45 Moya and Mer- Reconstructing pre- Aguadilla– Mid- to late A/B/F: Cliffs, Onshore transport of coral colonies 11 Oct. 1918 c. 15-20 Local run-up 0.95 10-14 5
cado (2006) historical tsunamis Rincón (north- Holocene beach and (up to 1.5 m) and large boulders is tsunami 11 October
west Puerto back of the ascribed to prehistoric tsunami im- 1918 tsu-
Rico) dunes pact nami: 3–6

46 Taggart et al. Reconstructing Isla de Mona Mid- to late A/B: Intertidal Coarse-clast ridge complex (sand, After 4176 c. 5 <0.9 0.95 10-14 3;3;2
(1993); Gonza- transport processes (Puerto Rico) Holocene to elevated cobbles and fine boulders) erosion years ago (north-
lez et al. (1997); of coastal boulders; carbonate and deposition is rather associated west
Morton et al. evaluating coastal platform (up with hurricane-induced waves by Puerto
(2008a) hazards using sedi- to several Morton et al. (2008a) Rico)
mentary records metres a.s.l.)

47 Buckley et al. Identifying and da- Southeast of Mid- to late F: Coastal Cobbles and fine boulders probably 1200–1450 c. 10-15 2.3–5.2 0.45 10-14 4
(2012); Watt et ting marine over- Windlass Bight, Holocene flats behind deposited by tsunami overwash CE; 1650 (Road Town,
al. (2012a); wash; documenta- north shore of beach ridges based on spatial distribution, main CE–1800 CE Tortola)
Weil Accardo et tion of a cobble and Anegada (BVI) axes orientation, numerical simula-
al. (2012); At- boulder field and re- tion and correlating finer deposits At-
water et al. construction of dep- water et al. (2012)
(2013b, 2014) ositional process

48 Spiske and Hal- Reconstruct the dep- Soldier Wash, A few tens of A: Limestone Clast-supported structure, imbrica- - c. 10-15 2.3–5.2 0.45 10-14 1
ley (2014) ositional process of north shore of years (mini- platform, tions and the exclusively well- (Road Town,
a ridge of coral rub- Anegada (BVI) mum age) 0.4–0.5 m rounded material of the boulder ridge Tortola)
ble high indicate build up during hurricane
overwash

49 Scheffers Documentation of Anguilla and Mid- to late B: Elevated boulders of up to 100 t, bimodal tsu- N1000 BP; < c. <2 n/d 0.52 (st. 5-9 5
(2006b); coarse-clast tsunami Scrub Island Holocene limestone nami deposits at the east coast, tsu- 610 BP Kitts and
Scheffers and deposits platforms nami boulder ridges, elongated bars (based on Nevis)
Kelletat (2006) of coral debris; boulder ridge at the uncal. 14C
southern coast data)

214
Appendix

Table A-4 (continuation)

Tsu- Interpreta-
nami tion in the
Time cov- height original
ered by Age of the 30-year SCDB Decadal source (see
sediment Type of candidate prob. Of tsu- Amplitude sce- number order of lit-
archive sediment Evidence for extreme wave tsunami nami run-up 1867 tsu- nario of hurri- erature
No. Author(s) Goal of the study Location (years BP) archive events deposit > 0.5 m nami (m) (m) canes sources)

50 Scheffers Documentation of St. Martin Mid- to late A/B: Elevated Very few boulders up to 10 t, “tsu- 500 BP c. <5 n/d 0.52 (St. 5-9 5
(2006b); coarse-clast tsunami Holocene limestone nami boulder spit” on an offshore (based on Kitts and
Scheffers and deposits platforms volcanic island; a coral rubble ridge uncal. 14C Nevis)
Kelletat (2006) described as tsunamigenic at the data)
east coast was dated to 500 BP

51 Scheffers et al. Evaluating coastal Guadeloupe Mid- to late A/B: Intertidal Boulders of up to 30 t along the east 2700–2500 11.79 (Basse- 0.6–1.6 0.88 10-14 5;2
(2005); Morton hazards using sedi- Holocene to cliff-top po- coast and bimodal deposits indicate BP or slightly Terre) (Basse-
et al. (2006, mentary records sition (up to tsunami impact (Scheffers et al., later (based Terre)
2008a) several me- 2005); since these deposits are on uncal.
tres a.s.l.) within the range of hurricane-induced 14C data)
inundation and their morphology,
Morton et al. (2006, 2008a) infer
storm-dominated deposition

52 Scheffers et al. Documentation of St. Lucia Mid-Pleisto- Volcano flank Mixture of fines to boulder incorpo- Mid-Pleisto- 5.52 n/d 0.95 5-9 5
(2005) coarse-clast tsunami cene exposure of rated into volcanic tephra represents cene
deposits pyroclastic mid-Pleistocene tsunami deposit
deposits, 15–
50 m a.s.l.

53 Schellmann and Summarizing the in- Barbados Mid- to late B: Elevated Some polymodal ramparts/ridge 4500 and n/d n/d 0.58 5-9 2;5
Radtke (2004); ventory of marine Holocene limestone complexes at the northeast coast as- 1400 BP
Scheffers and Quaternary deposits; platforms sociated with deposition during hurri- (based on
Kelletat (2006) documentation of canes uncal. 14C
coarse-clast tsunami data)
(Schellmann and Radtke, 2004); Iso-
deposits lated boulders up to 170 t up to 15 m
a.s.l. and other boulder ridge fea-
tures indicate at least two strong tsu-
namis (Scheffers and Kelletat, 2006)

54 Scheffers et al. Documentation of Grenada Pleistocene to A/B: Suprati- Cobbles and fine boulders in sand Pleistocene 2.48 1.1–2.8 0.37 5-9 5
(2005) coarse-clast tsunami Holocene dal of rocky matrix, 0–3 m a.s.l., and semi-circu- and late Hol-
deposits shore lar boulder ridge indicate tsunami ocene
deposition

55 Schubert (1994) Presenting geologi- Puerto Colombia Mid- to late B: Erosional Coral pebbles to boulders with an c. 1300(?) <2 0.6–0.8 0.52 2-4 4
cal evidence for a (Venezuela) Holocene terrace average age of 1300 years distrib- years ago (Puerto Ca-
prehistoric tsunami (clifftop) uted at an elevation of 10–20 m a.s.l. bello)

215
Appendix

Table A-4 (continuation)

Tsu- Interpreta-
nami tion in the
Time cov- height original
ered by Age of the 30-year SCDB Decadal source (see
sediment Type of candidate prob. Of tsu- Amplitude sce- number order of lit-
archive sediment Evidence for extreme wave tsunami nami run-up 1867 tsu- nario of hurri- erature
No. Author(s) Goal of the study Location (years BP) archive events deposit > 0.5 m nami (m) (m) canes sources)

56 Scheffers Evaluating the depo- Bonaire Mid- to late A/B: Intertidal Largest blocks considered to be 4300, 3900, c. 7 (Willem- 0.1–0.8 Up to 5-9 See details in
(2002a, b, sitional process for Holocene to elevated “consistent with tsunami deposits” 3300, 1200, stad, Curaçao) 1.30 column “evi-
2004, 2005); different types of carbonate (Watt et al., 2010, p. 60) by all au- and 500 BP dence for ex-
Scheffers et al. coarse-clast depos- platform thors, apart from Spiske et al. (based on treme wave
(2006, 2014); its (2008); Scheffers (2002a, b, 2004, ESR and un- events”
Radtke et al. 2005) and Scheffers et al. (2006, cal. 14C
(2003); Morton 2009) ascribe the vast bimodal ram- data;
et al. (2006, parts along the windward coast to Scheffers,
2008a); Spiske tsunami deposition which has been 2005)
et al. (2008); criticized by Morton et al. (2006,
Pignatelli et al. 2008a); a scattered clast field north
(2010); Watt et of Boka Olivia is associated with tsu-
al. (2010); En- nami deposition by Scheffers
gel and May (2002a, b, 2004, 2005), Morton et al.
(2012) (2006) and Watt et al. (2010)

57 Scheffers Evaluating the depo- Curaçao Mid- to late A/B: Intertidal Boulder fields and ridge complexes 3500, 1500, c. 7 (Willem- 0.1–0.8 Up to 2-4 5
(2002a, b, sitional process for Holocene to elevated are attributed to tsunami impact and 500 BP stad, Curaçao) (Bonaire) 2.90
2004); Radtke different types of carbonate (based on
et al. (2003) coarse-clast depos- platform cal. 14C
its data;
Scheffers,
2002b)

58 Scheffers Evaluating the depo- Aruba Mid- to late A/B: Intertidal Boulder fields and ridge complexes 3500, 1500, c. 7 (Willem- 0.1–0.8 Up to 2-4 5
(2002a, b, sitional process for Holocene to elevated are attributed to tsunami impact and 500 BP stad, Curaçao) (Bonaire) 3.26
2004); Radtke different types of carbonate (based on
et al. (2003) coarse-clast depos- platform cal. 14C
its data;
Scheffers,
2002b)

59 Macintyre et al. Reconstructing for- Holandés Cays, Late Holo- A: Coral reef A cemented ridge on the reef flat - 17.56 (Colón) 0–0.2 (Co- 0.66 0-1 1
(2001) mation of an inter- Panama cene (intertidal) with coralline algae cover was inter- lón)
tidal ridge preted as storm-built around 2000–
2800 ago

216
Supplements

Supplements

B. Supplements related to Experiments on tsunami induced boulder


transport – a review

Table B-1: Studies on transportation of artificial bodies by high-energy wave events (part a).

Debris di-
mensions Peak
(l x w x h or Debris wave
Buoy- Debris diameter) Flatness Wave height
Publication Scope Body type ancy types [cm] Index Wave generation [cm]
Transporta-
Non-break-
Nakamura Artificial tion / effect of Shipping con- 3.2 x 16.3 x 3.22 to
Positive 2 2.8 ing solitary Piston
et al. (2012) objects seawall reflec- tainer 3.5 5.28
wave
tions

Regular-
Yao et al. Artificial Transport dis- shaped 1 x 0.5 x Solitary
Positive 1 1.5 Piston 4
(2014) objects tance, path house mod- 0.5 wave
els

100 (soliton
Rueben et Artificial Body acceler- Shipping con- 60 x 60 x offshore) /
Positive 1 1.5 Broken bore Piston
al. (2015) objects ation tainer 40 30 (broken
bore)

~ 15 (off-
Stolle et al. Nearshore Waseda
Artificial Transporta- Shipping con- shore) /
(2015; Positive 1 15 x 6 x 6 1.75 solitary wave gener-
objects tion tainer ~6 (on-
2017) Wave ation
shore bore)
~ 15 (off-
Nearshore Waseda shore) /
Nistor et al. Artificial Transporta- Shipping con- ~6 (on-
Positive 1 15 x 6 x 6 1.75 Solitary wave gener-
(2017) objects tion tainer shore bore)
Wave ation

~ 14 (off-
Friction influ- Waseda shore) /
Stolle et al. Artificial Shipping con- 12 x 4.5 x ~8 (on-
ence on Positive 3 1.83 Bore wave gener-
(2019b) objects tainer 4.5 shore bore)
movement ation

217
Supplements

Table B-2: Studies on transportation of artificial bodies by high-energy wave events (part b).

Peak wave Ratio Initial boul- Additional Tests with


velocity rBW = hB:hF Initial boul- der align- fixed ob- multiple
Publication [m/s] (based on peak values) Shore model Scale der location ment Bottom roughness stacle debris
Nakamura et al. seawall with horizon- 0°, 30°, 45°,
n.p. 0.67 1:75 subaerial n.p.  
(2012) tal apron 60°, 90°

90° (not
not physically
Yao et al. (2014) n.p. 0.125 inclined / horizontal subaerial clearly pub- n.p.  
scaled
lished)

Rueben et al. (2015) 2 (onshore) 0.4 / 1.33 inclined / horizontal app. 1:20 subaerial 90° (Fr. Coeff 0.2 – 0.3)  (partially) 

Stolle et al. (2017a,


~ 0.7 0.4 / 1 elevated apron 1:40 subaerial 0°, 90° 0.4 (static)  
2017b)

Nistor et al. (2017) ~ 0.7 0.4 / 1 elevated apron 1:40 subaerial 0°, 90° 0.4 (static)  

inclined and elevated Frict. Coeff: 0.41 -


Stolle et al. (2019b) 2.4 0.3 / 0.6 1:50 subaerial 90°  
apron 0.85

n.p. = not provided; rBW = boulder-wave ratio

218
Supplements

Table B-3: Varied parameter in the considered experimental studies on tsunami induced boulder transport.

Oetjen et al.
Petroff et al.
Luccio et al.

and Tanaka
Nandasena
Imamura et

Lodhi et al.
Bressan et
al. (2008)

al. (2018)
Liu et al.

Freund.
(1998)

(2001)

(2013)

(2014)

(2014)

(2020)

(2020)
Dam-break height         *

Boulder orientation         

Boulder volume         

Boulder shape         

Boulder density         

Submergence         

Fluid density         

Shore inclination / shape         

Bottom roughness         

*Not investigated regarding influence

Figure B-1: Sketch of a possible dynamic drag force experiment. a) Initial setup. Structure/body in origin position. b)
Controlled acceleration of the structure during wave impact and measuring of the force development.

219
Supplements

Figure B-2: Example of using the proposed decision supporting diagram. The exemplary case is compound from a
boulder shape of extended width and length to a cubic while the width exceeds the length. The initial submergence is
assumed to be near to the boulders height and its weight is slightly larger than in the standard case. Finally, the bottom
roughness is set slightly rougher than in the assumed standard case.

220
Supplements

C. Supplements related to Significance of boulder shape, shoreline config-


uration and pre-transport setting for the transport of boulders by tsuna-
mis

Figure C-1: Repeated wave generation a: comparison of wave velocity. b: comparison of wave-height.

Table C-1: Overview for the conducted experiments on shore type 1.

submergence quantity

subaerial 56

partially submerged1) 15
complex

partially submerged2) 16

submerged1) 3

submerged2) 9

subaerial 39

partially submerged1) -
idealized

partially submerged2) 37

submerged1) -

submerged2)
9
subaerial
36
partially submerged1)
flat cuboid

9
partially submerged2)
-
submerged1)
-
submerged 2)
9
2)
All boulder types are placed on the same coordinate in x and y direction. Different water
covering.
1)
The boulder is placed on different x, y coordinates in order to realize the same water
height over the boulder.

221
Supplements

Figure C-2: Wave profile on shore Type 1. Measured 1.5 m in front of the shore tip.

Figure C-3: Distances between shore-tip and boulder models depending on initial submergence and boulder type.

222
Supplements

Figure C-4: Distribution graphs for the transport distances on shore Type 1.

223
Supplements

Table C-2: Statistical results for the experiments on shore type 1.

Regular cuboid Complex Flat cuboid


sa ps s1) sa ps1) s1) sa ps1) s1)
Minimum 58.44 32.16 8.196 38.19 31.26 3.31 16.47 2.44 0.0

25% percentile 71.55 49.04 11.68 43.2 34.65 4.63 21.33 4.14 0.0

Median 75.61 54.23 12.03 47.25 36.12 5.39 24.78 5.34 0.0

75 % percentile 83.27 63.18 14.51 50.27 41.61 6.19 26.95 7.92 0.0
transport distance [cm]

Maximum 90.32 76.28 15.22 59.49 49.24 6.64 31.28 9.76 0.0

Mean 76.21 55.67 12.24 47.05 38.12 5.29 24.27 5.9 0.0

Std. deviation 7.85 9.971 2.28 4.66 5.24 1.03 3.62 2.35 0.0

Std. error of mean 1.257 1.639 0.86 0.623 1.31 0.34 0.6 0.78 0.0

Lower 95 % CI of mean 73.66 52.34 10.14 45.8 35.33 4.50 23.05 4.09 0.0

Upper 95 % CI of mean 78.75 58.99 14.35 48.3 40.91 6.08 25.5 7.7 0.0

D'Agostino & Pearson omnibus normality test


N too N too
K2 0.8351 0.2841 0.7559 1.985 1.181 0.6327 0.2334
small small
P value 0.6587 0.8676 0.6853 0.3706 0.5540 0.7288 0.8898

Passed? Yes Yes Yes Yes Yes Yes Yes

Shapiro-Wilk normality test


normality test

N too
W 0.9716 0.9821 0.9157 0.9807 0.9298 0.9639 0.9762 0.9734
small
P value 0.4185 0.8032 0.4366 0.507 0.2424 0.8384 0.6176 0.922

Passed? Yes Yes Yes Yes Yes Yes Yes Yes

KS normality test
KS distance 0.1084 0.1352 0.2589 0.07712 0.1972 0.1179 0.1118 0.1494
> >
P value > 0.1000 0.0853 > 0.1000 0.0969 > 0.1000 > 0.1000
0.1000 0.1000
Passed? Yes Yes Yes Yes Yes Yes Yes Yes

224
Supplements

Figure C-5: Boxplot for the maximum transport distances on shore Type 1. Grouped following submergence and boul-
der shape. The initial orientation was kept as 90° (long axis perpendicular to the flow) over all experiments on shore-
Type 1.

225
Supplements

Figure C-6: Optical comparison between the transport process of the complex and regular cuboid boulder on shore
Type 2, partially-submerged and 90° initial alignment (representative experimental run).

226
Supplements

Figure C-7: Optical comparison between the transport process of the complex and regular cuboid boulder on shore
Type 2, subaerial and 90° initial alignment (representative experimental run).

227
Supplements

Figure C-8: Total transport distances on shore Type 2 grouped for boulder shape based on 131 experiments (black:
complex; red: regular cuboid boulder). Only unidirectional flow and no backwash is considered.

228
Supplements

Figure C-9: Wave profiles for all generated waves on shore Type 3.

Figure C-10: Initial boulder positions and water-level for the experiments on the stepped shore.

229
Supplements

Figure C-11: Total transport distances on shore-type 3 grouped for boulder (black: complex; red: regular cuboid
bouder). The red line marks the transportation limit (extended movements are set to step size in the figure).

230
Acknowledgments

Acknowledgments

First of all, I would like to express my special thanks to Univ.-Prof. Dr.-Ing. Holger Schüttrumpf who
made it possible for me to write my dissertation at the Institute of Hydraulic Engineering and Water
Resources Management (IWW) at the RWTH Aachen University and who gave me every freedom
to research off the usual paths. I also want to gratefully thank Prof. Dr. Helmut Brückner for taking
the second review and providing excellent suggestions on the thesis. Additional thanks go to the
German Research Foundation (DFG), which enabled the funding of my PhD position and neces-
sary material.

Furthermore, I thank Max Engel and Shiva P. Pudasaini. The contribution of Max to this thesis is
beyond words. I am very glad to have had him as support. The same goes for Shiva. He has been
very patient with me and has always inspired me to try unusual and progressive approaches to
improving research.

It is particularly important to me to thank all my colleagues who supported me during my PhD: the
family Banhold, Christiane Eichmanns, Christian Grimm, Sebastian Hudjetz, Elena Klopries, Vere-
na Krebs, Moritz Kreyenschulte, Simone Lechthaler, Elena Pummer, Babette Scheres, Tobias
Schruff, Harish Selvam, and Christian Vogelgesang. You all supported me in your personal and
different ways. I am very grateful for all of you!

I also want to thank all my friends outside of the University for being there, every time needed: The
BET gang, the Entsorgungsingenieurwesen gang (and all associated, including the economy
agent), and Mark Kliewer.

A very special thanks goes to my family: Antje, Hans, Hedi, Yara, Yona, Dirk and my grandma. I
can always count them.

Finally, I am very thankful for my girlfriend: Nora Paul. She endured all my “difficult moods”, espe-
cially during the last six months of my PhD with admirable equanimity. Thank you for being there
and here’s to many tours with Dullie.

231
Curriculum Vitae

Curriculum Vitae

Personal Data
Name Jan Oetjen
Date of birth 8th March 1984
Place of birth Jülich

Academic Education
08/1996 – 07/2003 Allgemeine Hochschulreife
Willy-Brandt Gesamtschule, Übach-Palenberg

10/2004 – 10/2012 Bachelor of Science – Waste Management Engineering


Waste Management Engineering
Faculty of Georesources and Materials Engineering
RWTH Aachen University

Bachelor thesis
Effects of climate change on the water balance in North Rhine-
Westphalia with regard to the resource groundwater: state of re-
search and assessment
Chair of Engineering Geology and Hydrogeology
RWTH Aachen University

10/2012 – 05/2014 Master of Science – Waste Management Engineering


Waste Management Engineering
Faculty of Georesources and Materials Engineering
RWTH Aachen University

Master thesis
Development of a concept for the establishment of a real-time
groundwater monitoring network
Institute of Hydraulic Engineering and Water Resources
Management
RWTH Aachen University

09/2014 – 12/2021 Doctorate


Institute of Hydraulic Engineering and Water Resources
Management
RWTH Aachen University

232
List of Publications

List of Publications

Peer Reviewed – Scientific Journals


Engel, M.; Oetjen, J.; May, S.M.; Brückner, H. (2016): Tsunami deposits of the Caribbean – To-
wards an improved coastal hazard assessment. In: Earth-Science Reviews 163, pp. 260–296.
DOI: 10.1016/j.earscirev.2016.10.010.
Oetjen, J.; Engel, M.; Pudasaini, S.P.; Schüttrumpf, H. (2020): Significance of boulder shape,
shoreline configuration and pre‐transport setting for the transport of boulders by tsunamis. In:
Earth Surface Processes and Landforms 45 (9), pp. 2118–2133. DOI: 10.1002/esp.4870.
Oetjen, J.; Engel, M.; Schüttrumpf, H. (2021): Experiments on tsunami induced boulder transport
– A review. In: Earth-Science Reviews 220. DOI: 10.1016/j.earscirev.2021.103714.
Oetjen, J.; Sundar, V.; Venkatachalam, S.; Reicherter, K.; Engel, M.; Schüttrumpf, H.; Sannasiraj,
S.A. (in review): A Comprehensive Review on Structural Tsunami Countermeasures. In: Natural
Hazards. Manuscript Number: NHAZ-D-20-01613.

Book Contributions
Oetjen, J.; Schüttrumpf, H.; Engel, M. (2020b): Experimental models of coarse-clast transport by
tsunamis. In: Engel, M.; Pilarczyk, J.; May, S.M.; Brill, D.; Garrett, E. (Eds.): Geological Records
of Tsunamis and Other Extreme Waves. Elsevier, pp. 585-615. DOI: 10.1016/B978-0-12-
815686-5.00027-4.

Conference Contributions
Oetjen, J.; Engel, M.; Effkemann, C.; May, S.M.; Pudasaini, S.P.; Wöffler, T.; Aizinger, V.; Schüt-
trumpf, H.; Brückner, H. (2015): Numerical modelling of tsunami scenarios for the island of Bon-
aire (Leeward Antilles). In: Programme and Abstract Book, 4th International Tsunami Field Sym-
posium, 23–27 March 2015, Kata Beach, Phuket, Thailand, pp. 81–85.
Oetjen, J.; Engel, M.; May, S.M.; Schüttrumpf, H.; Brückner, H.; Pudasaini, S.P. (2016): Tsunami-
induced boulder transport – combining physical experiments and numerical modelling. In: Eu-
ropean Geoscience Union (EGU) General Assembly, Vienna. Geophysical Research Abstracts,
Vol. 18, EGU2016-5810.
Engel., M.; Oetjen, J.; May, S.M.; Brückner, H. (2016): Extreme-wave deposits in the Caribbean –
towards an improved tsunami hazard assessment. In: European Geoscience Union (EGU) Gen-
eral Assembly, Vienna. Geophysical Research Abstracts, Vol. 18, EGU2016-13628.
Oetjen, J.; Engel, M.; Pudasaini, S.P.; Schüttrumpf, H.; Brückner, H. (2017): An advanced three-
phase physical, experimental and numerical method for tsunami induced boulder transport. In:
European Geoscience Union (EGU) General Assembly, Vienna. Geophysical Research Ab-
stracts, Vol. 19. EGU2017-15455
Oetjen, J.; Engel, M.; Brückner, H.; Pudasaini, S.P.; Schüttrumpf, H. (2017): Enhanced field obser-
vation based physical and numerical modeling of tsunami induced boulder transport: phase 1:
physical experiments. In: Proceedings of the International Conference on Coastal Engineering
35. DOI: 10.9753/icce.v35.management.4.
Oetjen, J.; Engel, M.; Pudasaini, S.P.; Schüttrumpf, H.; Brückner, H. (2017): Physical modelling of
tsunami-induced boulder transport. In: 03–07 Sep 2017: 5th International Tsunami Field Sym-
posium, Lisbon, Portugal.
Oetjen, J.; Engel, M.; Schönberger, J.J.; Pudasaini, S.P.; Schüttrumpf, H. (2018): Simulation of
boulder transport in a flume comparing cuboid and complex-shaped boulder models. In: EGU
General Assembly 2018, Geophysical Research Abstracts 20. DOI: 10.18154/RWTH-2019-
03133.
Oetjen, J.; Engel, M.; Schüttrumpf, H.; Pudasaini, S.P. (2019): Tsunami boulder transport simulated
in a numerical two-phase mass flow model and flume experiments – a combined approach. In:
25–31 Jul 2019: 20th INQUA Congress, Dublin, Ireland.

233
List of Publications

Oetjen, J.; Engel, M.; Schüttrumpf, H.; Pudasaini, S.P. (2019): Advanced physical numerical mod-
elling of tsunami induced boulder transport. In: European Geoscience Union (EGU) General
Assembly, Vienna. Geophysical Research Abstracts, Vol. 21, EGU2019-7728-1.
Hess, K.; Engel, M.; Oetjen, J.; Patel, T.; Schön, I.; Dawson, S.; Heyvaert, V.M.A. (2021): Historical
records of storm frequency on the Shetland Islands (UK) - Preliminary insights from lake sedi-
ment cores and coastal wave modelling. European Geoscience Union (EGU) General Assem-
bly, Vienna. EGU General Assembly 2021, EGU21-8583. DOI: 10.5194/egusphere-egu21-
8773.
Oetjen, J.; Engel, M.; Schüttrumpf, H. (2021): Significance of boulder shape for the transport of
boulders by tsunamis. European Geoscience Union (EGU) General Assembly, Vienna. In: EGU
General Assembly 2021, EGU21-8583. DOI: 10.5194/egusphere-egu21-8583.

234

Das könnte Ihnen auch gefallen