You are on page 1of 166

STUDY OF RUBBLE MOUND AND CAISSON TYPE BREAKWATERS BY

EXPERIMENTAL AND NUMERICAL MODELLING UNDER EXTREME


WAVES

A THESIS SUBMITTED TO
THE GRADUATE SCHOOL OF NATURAL AND APPLIED SCIENCES
OF
MIDDLE EAST TECHNICAL UNIVERSITY

BY

GÖZDE GÜNEY DOĞAN

IN PARTIAL FULFILLMENT OF THE REQUIREMENTS


FOR
THE DEGREE OF MASTER OF SCIENCE
IN
CIVIL ENGINEERING

FEBRUARY 2016
Approval of the thesis:

STUDY OF RUBBLE MOUND AND CAISSON TYPE BREAKWATERS BY


EXPERIMENTAL AND NUMERICAL MODELLING UNDER EXTREME
WAVES

submitted by GÖZDE GÜNEY DOĞAN in partial fulfillment of the requirements


for the degree of Master of Science in Civil Engineering Department, Middle East
Technical University by,

Prof. Dr. Gülbin Dural Ünver


Dean, Graduate School of Natural and Applied Sciences _____________

Prof. Dr. İsmail Özgür Yaman


Head of Department, Civil Engineering _____________

Assist. Prof. Dr. Gülizar Özyurt Tarakcıoğlu


Supervisor, Civil Engineering Department, METU _____________

Examining Committee Members:

Prof. Dr. Ahmet Cevdet Yalçıner _____________


Civil Engineering Dept., METU

Assist. Prof. Dr. Gülizar Özyurt Tarakcıoğlu _____________


Civil Engineering Dept., METU

Prof. Dr. İsmail Aydın _____________


Civil Engineering Dept., METU

Assist. Prof. Dr. Talia Ekin Tokyay Sinha _____________


Civil Engineering Dept., METU

Assist. Prof. Dr. Aslı Numanoğlu Genç _____________


Civil Engineering Dept., Atılım University

Date: 05.02.2016
I hereby declare that all information in this document has been obtained and
presented in accordance with academic rules and ethical conduct. I also declare
that, as required by these rules and conduct, I have fully cited and referenced all
material and results that are not original to this work.

Name, Last name : Gözde Güney Doğan

Signature :

iv
ABSTRACT

STUDY OF RUBBLE MOUND AND CAISSON TYPE BREAKWATERS BY


EXPERIMENTAL AND NUMERICAL MODELLING UNDER EXTREME
WAVES

Doğan, Gözde Güney


M.S., Department of Civil Engineering
Supervisor: Asst. Prof. Dr. Gülizar Özyurt Tarakcıoğlu

February 2016, 144 pages

Many coastal structures are designed without considering loads of long waves or
extreme waves although they are constructed in areas prone to encounter these waves.
Performance of two different types of coastal structures, rubble mound and caisson
type breakwaters, is investigated under specified extreme wave conditions.
In the first part of the study, laboratory experiments are conducted to observe the
damage and failure phenomenon of the two types of breakwaters simultaneously which
were designed according to wind waves but tested under six different extreme wave
conditions. Pressure measurements along the front surface and bottom surface of the
vertical wall in caisson type breakwater are also carried out within the physical model
experiments for further investigation about the acting wave forces on the concrete
blocks of the caisson breakwater.
In the second part, the caisson type breakwater is modeled using a numerical software,
IH-2VOF, which is one of the RANS models that can be applied to simulate
hydrodynamic conditions around coastal structures. Two cases of the laboratory

v
experiments are simulated by IH-2VOF model and pressure data is obtained from these
simulations.
Finally, the pressure measurements of laboratory experiments are compared with the
results obtained from the numerical model, IH-2VOF and presented for two wave
conditions along the front and bottom surfaces of vertical wall in caisson type
breakwater.

Keywords: rubble mound breakwaters, caisson type breakwaters, physical model


experiments, pressure measurements, numerical modelling, extreme waves

vi
ÖZ

TAŞ DOLGU VE KESON TIPI DALGAKIRANLARIN EKSTREM DALGALAR


ALTINDA DENEYSEL VE SAYISAL MODELLEME ILE ÇALIŞILMASI

Doğan, Gözde Güney


Yüksek Lisans, İnşaat Mühendisliği Bölümü
Tez Yöneticisi: Y. Doç. Dr. Gülizar Özyurt Tarakcıoğlu

Şubat 2016, 144 sayfa

Pek çok kıyı yapısı, uzun dalga ve ya ekstrem dalga koşullarına maruz kalma ihtimali
yüksek olan yerlerde inşa edildiği halde bu dalga koşulları dikkate alınmadan
tasarlanmaktadır. İki farklı temel kıyı yapısı olan taş dolgu ve keson tipi
dalgakıranların belirlenen ekstrem dalga koşulları altındaki performansları
araştırılmıştır.
Çalışmanın ilk kısmında, rüzgar dalgalarına göre tasarlanmış olan taş dolgu ve keson
tipi dalgakıran kesitlerinin ekstrem dalga koşulları altındaki hasar ve yıkılma
durumları gözlemlemek amacı ile fiziksel model deneyleri yapılmış ve bu kesitler altı
farklı ekstrem dalga koşulu için test edilmiştir. Buna ek olarak, keson tipi dalgakıran
kesitindeki beton bloklara etkiyen dalga yüklerini incelemek amacıyla deneylerde
blokların ön yüzeyi boyunca basınç ölçümleri alınmıştır.
İkinci kısımda, keson tipi dalgakıran kesiti ve deney düzeneği IH-2VOF isimli bir
program ile sayısal olarak modellenmiş ve deneylerden ikisinin simulasyonları
yapılmıştır. Keson tipi dalgakıran kesitindeki beton blokların ön yüzüne etkiyen yatay

vii
basınç değerleri ve alt yüzüne etkiyen kaldırma basıncı değerleri sayısal modelleme
yapılarak elde edilmiştir.
Son olarak, sayısal modelleme sonuçlarından elde edilen basınç değerleri ile
laboratuvar ölçümlerinden elde edilen basınç değerleri iki dalga koşulu için
karşılaştırılarak sunulmuştur.

Keywords: taş dolgu dalgakıran, keson tipi dalgakıran, fiziksel model deneyleri,
basınç ölçümü, sayısal modelleme, ekstrem dalgalar

viii
To my beloved family and
To my beloved ones...

ix
ACKNOWLEDGEMENTS

First and foremost, I would like to thank Prof. Dr. Ahmet Cevdet Yalçıner for his
invaluable support at every step of not only my graduate studies but also my life. I am
so grateful to him that he has always listened to my problems patiently, most often
solved my problems heroically and motivated me continuously. He is one of the most
special persons in my life meaning much more than a university teacher. I will never
forget the moments of our conversations and his advices which enlighten my way
through life. I am really fascinated with not only his wide knowledge and experience
in technical issues but also in life, cultures and human relations. He is not only a great
academician but also a great people person. It is a great chance to meet him, to be a
student of him and to have a close relationship with him. I hope with all my heart that
this relationship will continue in the future.

I would also like to thank my supervisor, Assist. Prof. Dr. Gülizar Özyurt Tarakcıoğlu,
and co-supervisor, Dr. Cüneyt Baykal, for their support from the beginning of my
graduate study in METU Coastal and Ocean Engineering Laboratory. Dr. Tarakcıoğlu
encouraged me to work on the projects and told me her ideas about my thesis study
and helped me at each stage of my study. I have always admired her that she has a
great practical perspective. I am also thankfull to her for giving me a chance to
participate in the experimental work which was conducted at Technical University of
Braunschweig, for supporting me there and standing with me there, in my first
experience of living abroad. Dr. Cüneyt Baykal also helped me at each stage of my
experiments and numerical modelling. I was the one knocking on the door of him
during this study and he answered my questions with a great patience. He gave a lot
of importance to this study and always motivated me more than even myself. I hope I
have been a student worthy of his efforts.

x
I would also like to express my sincere thanks to Prof. Dr. Ayşen Ergin who inspired
me in the way of being an academician. Her endless positive energy has always
stimulated me and given me a great motivation. Her unusual, enjoyful and rich content
lessons taught me looking things from different perspectives, most importantly from
the engineering point of view. She is the one I feel gratitude and made me choose
coastal engineering as my area of interest.

I feel also grateful to Dr. Işıkhan Güler for teaching me how to obtain any information
myself, how to search for the solution of a problem in a wide area and how to present
something effectively. Our long discussions about political issues are also so precious
that makes him meaning more than a teacher. I really appreciate his political standing
and participation and contribution to resistance movement. I hope we will always be
in communication.

I have always considered myself lucky to be a member of the family of Coastal and
Ocean Engineering Laboratory. Dr. Ayşen Ergin particularly creates this friendly and
beautiful atmosphere in the lab and makes us feel like at home but I would also like to
express my special thanks to our staff, Nuray Sefa, Yusuf Korkut and Arif Kayışlı for
their endless support in my work and contribution to this beautiful atmosphere.

I am deeply grateful to my lovely friends, Merih Himmetoğlu, Arın Özge Himmetoğlu,


Bircan Işık, Yağmur Öztürk and to my beloved one Burak Uçak for their full support,
encouragement and share of all the important moments with me while walking in this
way. Life would be meaningless without them.

I am so grateful to my lab friends, Ebru Demirci, Betül Aytöre, Gökhan Güler, Deniz
Velioğlu, Rozita Kian, Nilay Doğulu and Duha Metin for all the joyful moments that
we had together. I would also like to express my special thanks to Çağıl Kirezci for
being an excellent work partner. We have always been a ‘’last-minute’’ team with him
but succeded in the end. I will never forget our moments of Matlab coding.

xi
Finally, I would like to express my sincere thanks to my beloved family. My father,
Habib Doğan, has always provided me with unfailing support and continuous
encouragement since my childhood. He has always loved, protected and believed in
me and done his best as a father. I owe him most of the things I have. My mother,
Seval Doğan, also provided me unconditional love and I would never survive without
her support. She is the one that I can share everything in my life and the shoulder that
I can rest forever. I am also grateful to my brother, Ali Doğan that we had so much fun
together, laughed and made jokes during my thesis period. I am also deeply grateful to
my aunts, Nilgün Doğan Sarpkaya and Saniye Özkan for meaning much more than an
aunt in my life. Our long discussions about life and their advices enlighten me about
life. My sincere thanks are for my grandmother, Gülistan Doğan, who is the angel of
my life.

This thesis study is partly supported within the scope of RAPSODI (CONCERT_Dis-
021 and TUBİTAK-113M556) project in the framework of CONCERT-Japan Joint
Call and The Scientific and Technological Research Council of Turkey, and partly
supported within EC funded ASTARTE (Grant no: 603839) project. IH Cantabria is
also acknowledged for providing the numerical model, IH2VOF, which is used in this
thesis.

xii
TABLE OF CONTENTS

ABSTRACT .............................................................................................................. v
ÖZ ............................................................................................................................ vii
ACKNOWLEDGEMENTS ...................................................................................... x
TABLE OF CONTENTS ......................................................................................... xiii
LIST OF TABLES ................................................................................................... xv
LIST OF FIGURES ................................................................................................ .xvi
CHAPTERS

1 INTRODUCTION .......................................................................................... 1

2 LITERATURE REVIEW ............................................................................... 5

2.1 Failure Mechanisms............................................................................ 5


2.2 Forces and Pressures on Structures .................................................. 12
2.3 Numerical Modelling........................................................................ 18

3 PHYSICAL MODEL EXPERIMENTS....................................................... 21

3.1 General Description of Physical Model Experiments ...................... 21


3.2 Wave Flume...................................................................................... 23
3.3 Model Scale ...................................................................................... 25
3.4 Laboratory and Scale Effects ............................................................ 29
3.5 Experimental Setup .......................................................................... 31
3.5.1 Rubble Mound Breakwater Cross-Section ...................................................... 33
3.5.2 Caisson Type Breakwater Cross-Section ........................................................ 35
3.6 Measuring Technique ....................................................................... 38
3.7 Wave Characteristics ........................................................................ 48
3.8 Summary of Physical Model Experiments ....................................... 55

4 NUMERICAL MODELLING ..................................................................... 57

4.1 Introduction to the Numerical Model, IH-2VOF ............................. 57


4.2 Theoretical Background of IH2VOF Model .................................... 59

xiii
4.3 Model Geometry and Mesh Generation ........................................... 66
4.4 Mesh Convergence Checks............................................................... 72
4.5 Hydraulic Conditions for the IH-2VOF Model ................................ 74
4.6 Summary of Numerical Modelling ................................................... 76

5 RESULTS AND DISCUSSION...................................................................79

5.1 Physical Model Experiments ............................................................ 79


5.2 Numerical Modelling Results and Pressure Measurement
Comparison……………………………………………………………………….102

6 CONCLUSION AND FUTURE RECOMMENDATIONS ...................... 115

7 REFERENCES ........................................................................................... 119

8 APPENDICES ............................................................................................ 129

xiv
LIST OF TABLES

Table 2.1: Failure Mode Matrix .............................................................................. 11


Table 3.1: Scaling factors using the similitude law of Froude ................................ 29
Table 3.2: Position of measuring devices in tests with respect to the initial position
of wave generator ............................................................................................. 42
Table 3.3: Wave characteristics before elimination ................................................ 51
Table 3.4: Wave characteristics before elimination, measured values ................... 51
Table 3.5: Characteristics of applied waves in model scale, caisson type breakwater
.......................................................................................................................... 52
Table 3.6: Characteristics of applied waves in model scale, rubble mound
breakwater ........................................................................................................ 52
Table 3.7: Ratio of measured wave heights to the design wave height in prototype53
Table 3.8: Summary of physical model experiments at METU .............................. 55
Table 4.1: Properties of the subzones in the mesh of caisson breakwater model ... 71
Table 5.1: Wave properties in model scale, caisson type breakwater ..................... 80
Table 5.2: Wave properties in model scale, rubble mound breakwater .................. 80
Table 5.3: Overview of detailed breakwater damage under related wave actions .. 97
Table 5.4: Observed properties of applied waves ................................................. 100
Table 8.1: Coordinates of Element Indices of Caisson Breakwater Model Geometry
in IH-2VOF .................................................................................................... 129

xv
LIST OF FIGURES

Figure 2.1: Sketch of stages of sea dike failure a) Overflowing water, b) Creation of
scour-hole at leeward toe, c) Failure of leeward slope, d) Complete failure of
crest, leeward slope and toe (Jayaratne et al., 2013) .......................................... 7
Figure 2.2: Process of seawall overturning (Kato et al., 2012) .................................8
Figure 2.3: Overtopping of a composite breakwater by incoming tsunami wave (note
no armour damage) (Esteban, Miguel, et al., 2013) ........................................... 9
Figure 2.4: Overtopping of a composite breakwater by outgoing tsunami wave (note
the heavy damage) (Esteban, Miguel, et al., 2013) ............................................ 9
Figure 2.5: Comparison of various tsunami-bore velocities as a function of
inundation depth (Nouri et al., 2007) ............................................................... 14
Figure 2.6: Sketch showing some conceptual terms like tsunami height, inundation
depth and run-up height (Japan Weather Association materials,
http://www.japanecho.net/311-data/1205/ ) ..................................................... 15
Figure 2.7: Distribution of design wave pressure (Goda, 1974) ............................. 16
Figure 2.8: Vertical distribution of pressures for simple walls (Allsop, 1996) .......18
Figure 3.1: Flowchart of the model experiments .................................................... 23
Figure 3.2: The smaller basin and the wave flume ................................................. 24
Figure 3.3: a) Piston type wave maker b) Wave absorbers .....................................25
Figure 3.4: General layout of the basin (Top: top view, Bottom: Side view) ......... 25
Figure 3.5: View of horizontal platform in the wave flume ...................................31
Figure 3.6: Caisson Type Breakwater .....................................................................32
Figure 3.7: Rubble Mound Breakwater ...................................................................32
Figure 3.8: Cross-section of the caisson type breakwater with the horizontal
platform (Dimensions are in meters and the figure is not to scale.)................. 33
Figure 3.9: Cross-section of the rubble mound breakwater with the horizontal
platform (Dimensions are in meters and the figure is not to scale.)................. 33

xvi
Figure 3.10: Crown wall element used in rubble mound breakwater from different
perspectives (scale 1:50) .................................................................................. 34
Figure 3.11: Placement of the crown wall elements used in rubble mound
breakwater ........................................................................................................ 34
Figure 3.12: Geometric Details of Rubble Mound Cross-Section (Scale 1:50)...... 35
Figure 3.13: Layout of a concrete block used in caisson breakwater (scale 1:50) . 36
Figure 3.14: Placement of concrete blocks used in caisson type breakwater ......... 36
Figure 3.15: Geometric Details of the Caisson Cross-Section (Scale 1:50) ........... 37
Figure 3.16: Placement of the two cross-sections ................................................... 38
Figure 3.17: Wire type DHI-Wave Meter wave gauge ........................................... 39
Figure 3.18: a) Exemplary arrangement of overtopping wave gauge (WG2) in
caisson breakwater b) Exemplary arrangement of overtopping wave gauge
(WG2) in rubble mound breakwater ................................................................ 39
Figure 3.19: Example graph of calibration for WG3 .............................................. 40
Figure 3.20: Experimental setup including measuring instruments for caisson type
breakwater ........................................................................................................ 41
Figure 3.21: Experimental setup including measuring instruments for rubble mound
breakwater ........................................................................................................ 41
Figure 3.22: Profile measurement lines .................................................................. 43
Figure 3.23: Observation window and video recording system.............................. 43
Figure 3.24: Micro pressure transducer used in pressure measurements ................ 44
Figure 3.25: Pressure measurement points along the front surface of caisson
breakwater (Dimensions are in cm.) ................................................................ 45
Figure 3.26: Placement of two pressure transducers on the front surface of the
vertical wall element at still water level ........................................................... 45
Figure 3.27: Pressure measurement points along the bottom surface of caisson
breakwater (Dimensions are in cm.) ................................................................ 46
Figure 3.28: Results from 6 different static calibration pretests of the transducer 1 46
Figure 3.29: Results of 6 different static calibration pretests of transducer 2 ........ 47
Figure 3.30: Results of the dynamic water tank test of the transducers.................. 47

xvii
Figure 3.31: Frequencies and periods of the vertical motions of the ocean surface
(Holthuijsen, 2007)........................................................................................... 48
Figure 3.32: Approximate regions of validity of analytical wave theories ............. 49
Figure 3.33: Biesel Transfer Function to obtain wave paddle displacement (Biesel,
1951)................................................................................................................. 53
Figure 3.34: Sample surface profile ........................................................................ 54
Figure 4.1: Schematic diagram of main approaches in numerical modelling in
coastal and harbor engineering ......................................................................... 57
Figure 4.2: Sketch of volume averaging process for resolution of the porous flow
(IH-2VOF Course Lecture Notes, 2012) .......................................................... 62
Figure 4.3: Coral Interface, Zones .......................................................................... 67
Figure 4.4: Flowchart of generating a mesh in Coral ............................................. 68
Figure 4.5: Elements indices in defining geometry of caisson type breakwater (IH-
2VOF Course Lecture Notes, 2012) .................................................................70
Figure 4.6: Model geometry for caisson type breakwater.......................................70
Figure 4.7: Generated mesh with the subzones for caisson breakwater ................. 72
Figure 4.8: Quaility check for the generated mesh ................................................. 73
Figure 4.9: IH-2VOF GUI Preprocessing Main Menu ........................................... 74
Figure 4.10: Generation of new wave series ........................................................... 75
Figure 4.11: Setting wave gauge position ............................................................... 76
Figure 5.1: Stages of damage of caisson type breakwater for wave number 9 a)
start of damage b) major damage c) total failure ............................................. 81
Figure 5.2: Stages of damage of rubble mound breakwater for wave number 9
a) start of damage b) major damage c) total failure ......................................... 81
Figure 5.3: Wave Number 9 impact on the vertical wall of caisson type breakwater
.......................................................................................................................... 82
Figure 5.4: Wave Number 9 impact on the crown walls of rubble mound breakwater
.......................................................................................................................... 83
Figure 5.5: Breakwater profiles for caisson type breakwater, Wave number 9 Set 1
(wave approach is from left) ............................................................................ 84

xviii
Figure 5.6: Breakwater profiles for caisson type breakwater, Wave number 9 Set 2
.......................................................................................................................... 85
Figure 5.7: Damage of caisson type breakwater (Top: Set 1, Bottom: Set 2) ........ 86
Figure 5.8: Breakwater profiles for rubble mound breakwater, Wave number 9 Set 1
.......................................................................................................................... 87
Figure 5.9: Breakwater profiles for rubble mound breakwater, Wave number 9 Set 2
.......................................................................................................................... 88
Figure 5.10: Damage of rubble mound breakwater after wave number 9, Set 1 .... 89
Figure 5.11: Overflowing wave acting on the vertical walls of caisson type
breakwater ........................................................................................................ 90
Figure 5.12: Overflowing wave acting on the vertical walls of rubble mound
breakwater ........................................................................................................ 90
Figure 5.13: Breakwater profiles for caisson type breakwater, Wave number 10 Set
2 ........................................................................................................................ 91
Figure 5.14: Damage of caisson type breakwater after wave number 10 Set 2 ...... 92
Figure 5.15: Breakwater profiles for rubble mound breakwater, Wave number 10 Set
2 ........................................................................................................................ 93
Figure 5.16: Damage of rubble mound breakwater after wave number 10, Set 2 .. 93
Figure 5.17: Overturning of vertical walls under wave action ............................... 95
Figure 5.18: Sliding of the crown walls of the rubble mound breakwater under wave
action ................................................................................................................ 96
Figure 5.19: Water surface profiles for wave number 9, WG8............................. 104
Figure 5.20: Water surface profiles for wave number 9, WG2............................. 104
Figure 5.21: Water surface profiles for wave number 10, WG8........................... 105
Figure 5.22: Water surface profiles for wave number 10, WG2........................... 105
Figure 5.23: Wave loading identification, PROVERBS parameter map (Kortenhaus
et al., 1999)..................................................................................................... 106
Figure 5.24: Example graph of measured pressure time series for wave number 9107
Figure 5.25: Definition sketch of an idealized impact wave pressure time history107
Figure 5.26: Example graph of measured pressure time series for wave number 10
........................................................................................................................ 108

xix
Figure 5.27: Local maximum dynamic pressures (point data) and instantaneous
pressure distributions for wave number 9 ...................................................... 109
Figure 5.28: Local maximum quasi-static pressures (point data) and instantaneous
pressure distributions for wave number 9 ...................................................... 110
Figure 5.29: Local maximum uplift pressures (point data) and instantaneous uplift
pressure distributions for wave number 9 ...................................................... 111
Figure 5.30: Local maximum dynamic pressures (point data) and instantaneous
pressure distributions for wave number 10 .................................................... 112
Figure 5.31: Local maximum uplift pressures (point data) and instantaneous uplift
pressure distributions for wave number 10 .................................................... 113
Figure 8.1: Breakwater profile of caisson type breakwater – Line 1 .................... 130
Figure 8.2: Breakwater profile of caisson type breakwater – Line 2 .................... 130
Figure 8.3: Breakwater profile of rubble mound breakwater – Line 1 ................. 131
Figure 8.4: Breakwater profile of rubble mound breakwater – Line 2 ................. 131
Figure 8.5: View of caisson type breakwater after Wave Number 12 .................. 132
Figure 8.6: View of rubble mound breakwater after Wave Number 12 ............... 132
Figure 8.7: Breakwater profile of caisson type breakwater – Line 1 .................... 132
Figure 8.8: Breakwater profile of caisson type breakwater – Line 2 .................... 133
Figure 8.9: Breakwater profile of rubble mound breakwater – Line 1 ................. 133
Figure 8.10: Breakwater profile of rubble mound breakwater – Line 2 ............... 134
Figure 8.11: View of caisson type breakwater after Wave Number 15, Set-2...... 134
Figure 8.12: View of rubble mound breakwater after Wave Number 15, Set-2 ...134
Figure 8.13: Breakwater profile of caisson type breakwater – Line 1 .................. 135
Figure 8.14: Breakwater profile of caisson type breakwater – Line 2 .................. 135
Figure 8.15: Breakwater profile of rubble mound breakwater – Line 1 ............... 136
Figure 8.16: Breakwater profile of rubble mound breakwater – Line 2 ............... 136
Figure 8.17: View of caisson type breakwater after Wave Number 17, Set-1...... 137
Figure 8.18: View of rubble mound breakwater after Wave Number 17, Set-1 ...137
Figure 8.19: Breakwater profile of caisson type breakwater – Line 1 .................. 138
Figure 8.20: Breakwater profile of caisson type breakwater – Line 2 .................. 138
Figure 8.21: Breakwater profile of rubble mound breakwater – Line 1 ............... 138

xx
Figure 8.22: Breakwater profile of rubble mound breakwater – Line 2 ............... 139
Figure 8.23: View of caisson type breakwater after Wave Number 17, Set-2 ..... 139
Figure 8.24: View of rubble mound breakwater after Wave Number 17, Set-2 ... 139
Figure 8.25: Breakwater profile of caisson type breakwater – Line 1 .................. 140
Figure 8.26: Breakwater profile of caisson type breakwater – Line 2 .................. 140
Figure 8.27: Breakwater profile of rubble mound breakwater – Line 1 ............... 141
Figure 8.28: Breakwater profile of rubble mound breakwater – Line 2 ............... 141
Figure 8.29: View of caisson type breakwater after Wave Number 18, Set-1 ..... 142
Figure 8.30: View of rubble mound breakwater after Wave Number 18, Set-1 ... 142
Figure 8.31: Breakwater profile of caisson type breakwater – Line 1 .................. 142
Figure 8.32: Breakwater profile of caisson type breakwater – Line 2 .................. 143
Figure 8.33: Breakwater profile of rubble mound breakwater – Line 1 ............... 143
Figure 8.34: Breakwater profile of rubble mound breakwater – Line 2 ............... 144
Figure 8.35: View of caisson type breakwater after Wave Number 18, Set-2 ..... 144
Figure 8.36: View of rubble mound breakwater after Wave Number 18, Set-2 ... 144

xxi
CHAPTER 1

1 INTRODUCTION

Tsunami is a typical long wave in the ocean, generated by seafloor or water surface
disturbances over a sufficiently large area (Wang, 2009). It has been observed and
recorded since ancient times, especially in Japan and the Mediterranean areas.
Tsunamis may attain large amplitudes in closed basins or shallow regions (Yalciner et
al. 2001; 2002; 2004; and 2005). The interest in tsunami research has increased in
recent years after the terrible consequences of the 2004 Indian Ocean tsunami and of
the 11th March 2011 tsunami in Japan.

2004 Indian Ocean tsunami was one of the deadliest and largest natural hazard all over
the world. ‘Tsunami risk reduction’ issue has come into prominence after this disaster
(Lovholt et al., 2014). Since 2004, six megathrust tsunamis happened in Indonesia,
Samoa, Chile and Japan in 2006, 2007, 2009, 2010 and 2011 (Lovholt et al., 2014).
An undersea earthquake with magnitude 9.0 occurred in Pacific coast of Tohoku, Japan
in 2011. The earthquake is named as "Great East Japan Earthquake" by being the most
powerful earthquake ever recorded in Japan and fifth most powerful earthquake in the
world since 1900 (RAPSODI Project Deliverable, D1). Transmission of the energy
caused giant tsunami waves that reached 40 meters in Miyako. The tsunami caused
nuclear explosions primarily in Fukushima Daiichi Nuclear Power Plant Complex
(RAPSODI Project Deliverable, D1, 2015). Overall world, tsunamis mostly occur in
Pacific Ocean. Despite tsunamis in oceans and open seas, there are great number of
tsunami records in the Mediterranean and Aegean Sea due to undersea earthquakes,
landslides and volcanic eruptions in this region as well. (RAPSODI Project
Deliverable, D1, 2015).
Tsunami history throughout the world indicates that the dense population, coastal
utilization and marine protected areas are vulnerable under tsunami motion. Ports,

1
harbors and marinas are some of the vulnerable places under tsunamis as well as
coastal ecosystems and aquaculture areas. However, all coastal facilities are critical
and must be resilient against natural hazards in order to facilitate recovery operations.
Therefore, it is important to develop a method for the design of coastal structures to
improve resilience against tsunami impacts.

One project focusing on this problem is Risk Assessment and design of Prevention
Structures fOr enhanced tsunami DIsaster resilience (RAPSODI) project. The
RAPSODI consortium consists of three European (NGI, METU, TUBS), and one
Japanese partner (PARI). The main research topics that the project deals with are
vulnerability assessment, the analysis of loads on structures and also numerical
tsunami modelling. The research has been done in three stages as the evaluation of
existing knowledge and comparisons of mitigation strategies (Stage 1), numerical and
experimental studies (Stage 2) and methodology for tsunami vulnerability assessment
and risk management (Stage 3). The first stage includes review and evaluation of tools
available for the numerical modelling, and the assessment of impact loads on
structures, failure modes, and vulnerability. A thorough investigation of the failure
mechanisms of coastal protection structures exposed to tsunamis is carried out in this
thesis within the scope of RAPSODI Project. A catalogue of failure modes of different
types of coastal structures based on observations of the 2011 Tohoku tsunami and
existing knowledge (RAPSODI Project Deliverable, D1, 2015) is prepared and
presented in Chapter 2. On the basis of this failure modes matrix, it is found out that
the existing research gaps and potential research areas are in the field of tsunami-
induced load and damage to coastal structures. Also, most of the missing information
in the literature is about performance of rubble mound breakwaters under tsunami
loads since these types of breakwaters are not common especially in Japan and are not
commonly observed in recent tsunami events. Therefore, laboratory experiments
within the RAPSODI Project were carried out in Technical University of
Braunschweig (TU-BS) hydraulic laboratory in collaboration with project partner
Middle East Technical University (METU) to investigate the damage and failure
mechanisms of rubble mound breakwaters with different configurations under solitary

2
waves and tsunami bores. In addition, it is found out that many coastal structures are
designed without considering loads of tsunami-like waves, long waves or extreme
waves although these structures can be exposed to these waves. Therefore, laboratory
experiments to understand the behavior of different types of coastal structures under
extreme waves are performed as a part of this thesis work to complement the
experimental study of RAPSODI Project. The hydraulic conditions in these
experiments are determined as extreme waves which defines the limitation of the
experimental work of this thesis as mentioned in Chapter 3.

The purpose of this thesis is to understand the damage and failure phenomenon of two
main types of coastal structures; rubble mound and caisson type breakwaters, under
extreme waves. Extreme waves are relatively large surface waves which are rare,
unpredictable, may occur without warning, and can impact with a huge force like
tsunami waves. The generation mechanism of extreme waves or the impact duration
to the structures are different from tsunami waves but large waves can cause flooding
and damage to coastal infrastructure and severe coastal disasters may occur due to
these extreme events. Therefore, it is necessary to consider the extreme wave
conditions in the design of coastal structures as in the case of tsunamis. These waves
are rare events but they occur in seas and oceans and a real danger to ships, platforms
and coastal structures as well as causing accidents resulting in human loss. Thus, 6
different extreme wave conditions are generated in laboratory and rubble mound and
caisson type breakwaters are examined under these wave actions in this thesis work.
Also, pressure measurements are carried out along the surface and bottom of the
concrete blocks of caisson type breakwater to investigate the forces acting on the
structure which cause the damage and failure of the structure.

Physical modelling is one of the classical methods for research in the field of coastal
engineering having some advantages. However, it has also some limitations and
deficiencies so does the other common method in coastal engineering, numerical
modelling. Therefore, coupling one method with the other one is an effective way to
minimize the limitations and deficiencies of each method. A numerical assessment is

3
also included in this study to make a comparison between the pressure measurements
and the numerical results in order to analyse the limitations of simulating the
experiments in the numerical environment. Many experimental tests are performed
with the aim of understanding damage behavior of breakwaters under extreme waves
and a caisson type breakwater is modeled by IH2VOF and the results are compared.
IH-2VOF is a numerical model which solves the two-dimensional (2D) wave flow
using the Reynolds Averaged Navier Stokes (RANS) equations at the fluid region and
Volume Averaged Reynolds Averaged Navier Stokes (VARANS) inside the porous
media (Hsu et al., 2002). The model is mainly used for the flow-structure interaction
especially in case of coastal structures like breakwaters and sea dikes. Detailed
information about the model is provided in Chapter 4.

In Chapter 2, a literature survey is presented including the basic research related to the
thesis topic to provide some background information. In Chapter 3, general description
of the physical model experiments is given including layout and properties of the wave
flume, the experimental setup, breakwater cross-sections, measuring equipment and
wave characteristics. Chapter 4 provides the description of the numerical model,
IH2VOF, theoretical background of the model, the convergence checks, model
geometry and the mesh generation. Chapter 5 includes the results of the experiments,
pressure measurements and numerical model analysis. The comparison experimental
and numerical model results is also discussed. Finally, in Chapter 6, conclusions and
future recommendations are presented.

4
CHAPTER 2

2 LITERATURE REVIEW

Tsunami – structure interaction and long wave – structure interaction research has
gained significant interest in recent years. Effects of these waves on coastal protection
structures, failure mechanisms of these structures or their damage behavior are
important subjects to be well understood in order to have coastal resilience. Many
studies can be reached through literature on these topics.

2.1 Failure Mechanisms

Field surveys carried out after the catastrophic events, 2011 The Great East Japan
Earthquake and Tsunami and 2004 Indian Ocean tsunami, form a great basis for
examining the failure mechanisms of coastal structures. A thorough literature survey
is performed through these survey reports from different authors. A failure modes
matrix is prepared using the reported failure mechanisms of coastal protection
structures in the above summarized reports to highlight the knowledge gaps so that
future studies can be planned.

For this review, overflow is defined as functional failure and it is observed for
seawalls and revetments, sea dikes and breakwaters in several places (Figure 2.1a).
One example is the coastal area in Otsuchi village which was completely destroyed as
the tsunami destroyed the breakwater and propagated inland along the Otsuchi and
Kotsuchi Rivers. Hydraulic control structures and seawalls were completely
overtopped during the inundation. (Mori et al., 2013)

5
Sliding was reported for revetments by Nagasawa and Tanaka (2012) who estimated
that the tsunami external force acted on the revetments seaward and the revetment fell
down seaward and destroyed.

Scour damage was seen on concrete seawalls. The example is two large scour profiles
observed at the leeward slope of the curved 2.5 m high concrete seawall in the east
side of Ishinomaki port extending about 60.0 m in length along the seawall. The
massive concrete platform placed at the toe of the leeward slope disappeared due to
the tsunami, leading to two scour holes being created according to Jayaratne et al.,
(2013). Scour was also reported by Yeh et al. (2012) for solid-concrete seawalls. They
conducted surveys along the Sanriku coast after 2011 event and stated that remarkable
destruction of upright solid-concrete type seawalls was closely related with the
tsunami induced scour. Also, the fast flow velocities with intense turbulence resulted
in severe undermining damage in the rear face of the mound-type seawall in
Kanahama, as well as formation of a large scour hole behind according to Yeh et al.,
(2012). Jayaratne et al. (2013) noticed severe damage to the leeward face and toe of
the sea dike at Soma city. Diverse failure patterns were observed from the north to the
south side, resulting in partial to total failure of the leeward face due to scour. Seaward
flow over the coastal dike may cause scouring at the seaward toe of the dike as also
pointed out in previous studies (e.g., Noguchi et al. 1997). Seaward flow during
tsunami drawdown caused scouring at the seaward toe of the coastal dike and
revetment. Scouring at the landward toe affected the stability of the seaward armor,
resulting in floating away of seaward armor and breaching of the dike. (Kato et al.,
2012) A simple sketch of creation of scour-hole causing complete failure of structure
toe and crest is provided in Figure 2.1b, 2.1c and 2.1d.

6
Figure 2.1: Sketch of stages of sea dike failure a) Overflowing water, b) Creation of
scour-hole at leeward toe, c) Failure of leeward slope, d) Complete failure of crest,
leeward slope and toe (Jayaratne et al., 2013)

Overturning was observed on concrete seawalls in some cases (Figure 2.2). The
concrete seawalls were overturned by return flow rather than by the incoming tsunami
as reported by Earthquake Engineering Research Institute, EERI (2011). Also,
overturning of a seawall which was observed on the Ryoishi Coast, Iwate Prefecture
induced by the Great East Japan Earthquake Tsunami may occur if the overturning
moment induced by the wave force on the seawall during tsunami runup or drawdown
is larger than the resistance moment due to the weight of the seawall reported by Kato
et al. (2012). Many tsunami gates designed to reduce flooding along rivers were also
overturned by the Great East Japan Earthquake and tsunami. (Sagara and Saito, 2013)

7
Figure 2.2: Process of seawall overturning (Kato et al., 2012)

Soil instability is another failure mechanism observed on seawalls and reported by


Yeh et al., (2012). Remarkable destruction of upright solid-concrete type seawalls was
closely related with also soil instability according to them. The rapid decrease in
inundation depth during the return-flow phase caused soil fluidization down to a
substantial depth and this mechanism explains severely undermined foundations
observed in the area along the Sanriku Coast of low flow velocities. They found that
soil instability played a major role in the failures. (Yeh et al., 2012)

In case of breakwaters; scouring, sliding, soil failure, seaward and leeward slope
failures and overturning are the failure mechanisms observed in field surveys.
Scouring, sliding and soil failure due to tsunami overflow was observed in Kamaishi
and Hachinohe Breakwaters. Both breakwaters were modelled in laboratory and
several experiments were performed by Arikawa and Shimosako (2013) to determine
factor of safety regarding sliding and overturning considering the scour at the leeward
(at the foundation level) as well as bearing capacity failure (soil failure). Jayaratne et
al. (2013) states that large concrete armour units had been placed in front of the
seaward slope of the breakwater at Ishinomaki Port. It was found that the primary
armour units on the seaward slope were displaced and scattered in front of the

8
breakwater, and some units were buried under tsunami deposits, though there was no
indication of damage to the front slope. Esteban et al., 2013 states that the exact failure
mechanism for each of the breakwater types is still unclear, and whether armour units
were displaced by the incoming or the outgoing wave could not be easily established
for any of the field failures recorded. Therefore, they carried out some preliminary
laboratory experiments which appear to indicate that although the incoming tsunami
wave can cause some movement to the caisson, the major failure mode of the armour
could occur as a result of the outgoing wave (Figure 2.3 and Figure 2.4). Also, in
Kamaishi City, the tsunami on March 11th overturned the north section (990 m in
length) of the newly completed offshore breakwater and although the south section
(670 m in length) survived mostly intact, it was left inclined as given in Fraser et al.
(2013) but explained in Yagyu (2011).

Figure 2.3: Overtopping of a composite breakwater by incoming tsunami wave (note


no armour damage) (Esteban et al., 2013)

Figure 2.4: Overtopping of a composite breakwater by outgoing tsunami wave (note


the heavy damage) (Esteban et al., 2013)

9
A failure modes matrix is prepared using the reported failure mechanisms of coastal
protection structures in the above summarized reports to highlight the knowledge gaps
so that future studies can be planned. The matrix is given in Table 2.1. In the matrix,
a tick marker represents the match of the structure with the observed failure
mechanism reported in the field surveys. The blanks mean that either a failure type is
not observed for a certain type of structure or there are no reports acquired or it can
not be decided which failure mechanism refer to the specified structure.

10
Table 2.1: Failure Mode Matrix

11
2.2 Forces and Pressures on Structures

As can be seen from Table 2.1 and examples of failures, breakwaters which are one of
the most important elements of coastal protection and exposed elements to tsunami
wave action should be a research topic necessary to observe their damage mechanisms
and collapse phenomenon experimentally, numerically and by field observations. Also
the acting forces on the structure under the wave attack is another point to consider
and investigate since these forces should be classified and identified for new design
considerations and to improve new methods for more resilient structures.

A comprehensive review of tsunami forces is presented by Nistor et al. (2009). He


states that three parameters are essential for defining the magnitude and application of
these forces: (1) inundation depth, (2) flow velocity, and (3) flow direction. The
parameters mainly depend on: (a) tsunami wave height and wave period; (b) coastal
topography; and (c) roughness of the coastal inland. Forces associated with tsunami
bores consist of: (1) hydrostatic force, (2) hydrodynamic (drag) force, (3) buoyant
force, (4) surge force and (5) impact of debris.

The hydrostatic force is generated by still or slow-moving water acting perpendicular


onto planar surfaces. The hydrostatic force per unit width, FHS, can be calculated using
the equation given below.
1 𝑢𝑝 2 2
𝐹𝐻𝑆 = 𝜌𝑔 (𝑑𝑠 + ) [2.1]
2 2𝑔

Where ρ is the seawater density, g is the gravitational acceleration, ds is the inundation


depth and up is the normal component of flow velocity. Equation 2.1 is proposed by
the City and County of Honolulu Building Code (CCH, 2000) and accounts for the
velocity head. The point of application of the resultant hydrostatic force is located at
one third from the base of the triangular hydrostatic pressure distribution.

12
The buoyant force is the vertical force acting through the center of mass of a
submerged body. Its magnitude is equal to the weight of the volume of water displaced
by the submerged body as given in Equation 2.2. Buoyant forces can generate
significant damage to structural elements, and are calculated as follows:
𝐹𝐵 = 𝜌𝑔𝑉 [2.2]
where, V is the volume of water displaced by submerged structure (Nistor et al., 2009).

As the tsunami bore moves inland with moderate to high velocity, structures are
subjected to hydrodynamic forces caused by drag. Currently, there are differences
in estimating the magnitude of the hydrodynamic force. The general expression used
by existing codes is given below (Equation 2.3).
𝜌 𝐶𝐷 𝐴 𝑢2
𝐹𝐷 = [2.3]
2

Where, FD is the drag force acting in the direction of flow and CD is the drag coefficient
that depends on the shape of the surface on which drag forces are applied (Nistor et
al., 2009). Drag coefficient values of 1.0 and 1.2 are recommended for circular piles
by CCH (2000) and FEMA 55 (2003) design codes, respectively. For the case of
rectangular piles, the drag coefficient recommended by the same two codes is 2.0.
Parameter A is the projected area of the body normal to the direction of flow. The flow
is assumed to be uniform, and therefore, the resultant force will act at the centroid of
the projected area A. The term u in Equation 2.3 is the tsunami bore velocity. The
hydrodynamic force is directly proportional to the square of the tsunami bore velocity
as indicated in Equation 2.3. Therefore, the estimation of the bore velocity remains to
be one of the critical elements on which there is significant disagreement in literature
(Nistor et al., 2009). Tsunami-bore velocity and direction can vary significantly during
a major tsunami inundation. The general form of the bore velocity is shown below
(Equation 2.4).
𝑢 = 𝐶 √𝑔𝑑𝑠 [2.4]
Where, C is a constant and ds is the inundation depth. Various formulations were
proposed by FEMA 55 (2003) based on Dames and Moore (1980), CCH (2000),
Kirkoz (1983), Murty (1997), Bryant (2001), and Camfield (1980) for estimating the

13
velocity of a tsunami bore in terms of inundation depth. The velocity inundation depth
relationships proposed are plotted in Figure 2.5.

Figure 2.5: Comparison of various tsunami-bore velocities as a function of


inundation depth (Nouri et al., 2007)

According to Nistor et al. (2009), the surge force is generated by the impact of the
forward-moving water of a tsunami bore on a structure. Due to lack of detailed
experimental research specifically applicable to tsunami bores running up the
shoreline, the calculation of the surge force exerted on a structure has considerable
uncertainty. Based on research conducted by Dames and Moore (1980), CCH (2000)
recommends Equation 2.5 for the surge force FS on walls with heights equal to or
greater than three times the surge height (3h).
𝐹𝑆 = 4.5𝜌𝑔ℎ2 [2.5]

Where, FS is the surge force per unit width of wall, and h is the surge height. The point
of application of the resultant surge force is located at a distance h above the base of
the wall.

14
A tsunami bore traveling towards the land carries debris such as floating automobiles,
floating pieces of buildings, drift wood, boats and ships. The impact of floating debris
can result in significant forces on a structure, leading to structural damage or collapse
(Saatcioglu et al. 2006a, 2006b). Both FEMA 55 (2003) and CCH (2000) codes
account for debris impact forces, using the same approach, and recommend using
Equation 2.6 for estimation of debris impact force.
𝑑𝑢𝑏 𝑢
𝐹𝑖 = 𝑚𝑏 = 𝑚 ∆𝑡𝑖 [2.6]
𝑑𝑡

Where Fi is the impact force, mb is the mass of the body impacting the structure, ub is
the velocity of the impacting body (assumed equal to the flow velocity), ui is approach
velocity of the impacting body (assumed equal to the flow velocity) and Δt is the
impact duration taken equal to the time between the initial contact of the floating body
with the building and the instant of maximum impact force (Nistor et al., 2009).

According to FEMA 55 (2003), the impact force (a single concentrated load) acts
horizontally at the flow surface or at any point below it. Its magnitude is equal to the
force generated by 455 kg (1,000-pound) of debris traveling with the bore and acting
on a 0.092m2 (1 ft2) surface of the structural element. The impact force is to be applied
to the structural element at its most critical location, as determined by the structural
designer.

A sketch which conceptually shows some of the terms used in the force equations is
given in Figure 2.6.

Figure 2.6: Sketch showing some conceptual terms like tsunami height, inundation
depth and run-up height (Japan Weather Association materials,
http://www.japanecho.net/311-data/1205/ )

15
In this study, it is aimed to understand the force-structure interaction through the
pressures acting on the structure since it is difficult to isolate separately these forces
mentioned above and to measure each parameter in the equations. The main focus of
this study is damage and failure mechanisms of the structures. Therefore, the basic
pressure formulas that can be used to obtain the forces on the structures are reviewed.
One of the most commonly used pressure formulas in the literature is given by Goda
(1974). He proposed a pressure formula for composite breakwaters which can be
applied for the whole ranges of wave action from nonbreaking to postbreaking waves.
Goda (1974) gives the distribution of design wave pressure as given in Figure 2.7.

Figure 2.7: Distribution of design wave pressure (Goda, 1974)

The intensities of wave pressures, p1, p2, and p3, are calculated with the following
formula:

2
𝑝1 = 𝑤𝑜 𝐻𝐷 (∝1 +∝2 cos 𝛽) [2.7]

𝑝1
𝑝2 = [2.8]
cosh(2𝜋ℎ⁄𝐿 )

𝑝3 = ∝3 𝑝1 [2.9]

Where

16
1 4ℎ⁄𝐿
∝1 = 0.6 + ( )2 [2.10]
2 sinh4𝜋ℎ⁄𝐿

ℎ −𝑑 𝐻 2 2𝑑
𝑏
∝2 = min { 3ℎ ( 𝑑𝐷 ) , 𝐻 } [2.11]
𝑏 𝐷

ℎ′ 1
∝3 = 1 − [1 − ] [2.12]
ℎ cosh2𝜋ℎ⁄𝐿

HD : design wave height (given in Equation 2.14)


wO : specific weight of sea water
L : wavelength of design wave
min{a,b}: smaller one of a or b
β : angle of wave approach
hb : water depth at which the breaker height is to be evaluated (given in Equation 2.16)

For the calculation of uplift pressure acting on the bottom of upright section, the
triangular distribution is assumed irrespective of wave overtopping (Goda, 1974). The
intensity of toe uplift is given by Equation 2.13.
𝑝𝑢 = ∝1 ∝3 𝑤𝑜 𝐻𝐷 [2.13]

𝐻𝐷 = 𝐻𝑚𝑎𝑥 = min {1.8 𝐻1⁄ , 𝐻𝑏 } [2.14]


3
ℎ 4⁄
𝐻𝑏 = 0.17 𝐿0 {1 − 𝑒𝑥𝑝 [−1.5 𝐿𝑏 (1 + 15𝑡𝑎𝑛 3 𝜃)]} [2.15]
0

ℎ𝑏 = ℎ + 5𝐻1⁄ 𝑡𝑎𝑛𝜃 [2.16]


3

Where
L0 : deepwater wavelength calculated by 𝑔𝑇 2 /2𝜋
tanθ: mean gradient of sea bottom

Goda (1974) calibrated the other existing pressure formula of Hiroi (1920), Sainflou
(as cited in Goda, 1974), and Minikin (1950) with the cases of 21 slidings and 13
nonslidings of the upright sections of prototype breakwaters. He states that the
calibration shows that the new formulas are the most accurate ones. The Shore

17
Protection Manual (SPM) (1984) however states that Sainflou's method may
overestimate wave forces for short non-breaking waves. For long waves of low
steepness, the SPM recommends Sainflou's method (Allsop et al, 1996).

Allsop et al. (1996) also presents wave impacts on simple and composite vertical walls.
They conducted 2D hydraulic model tests to identify the wave pressures on these walls.
They aimed to develop a method to estimate wave forces under wave attack.
They give the vertical distribution of pressures on simple vertical walls as in Figure
2.8. They state that, for pulsating conditions Goda’s trapezoidal distribution of
pressure assumption is reasonably well-supported, but for impact conditions,
agreement is much less good.

Figure 2.8: Vertical distribution of pressures for simple walls (Allsop et al., 1996)

2.3 Numerical Modelling

Numerical modelling of hydraulic processes of wave-structure interaction is also an


essential part of research methods in coastal engineering due to the limitations of
physical modelling such as scale effects, long duration of experiments, laboratory
limitations and costly experiments. Although the numerical models need validation by
experiments or field measurements, they may provide at least successful preliminary

18
results in a shorter time period. There are number of models developed to investigate
the wave-structure interaction by means of solving wave reflection, wave
transmission, wave overtopping, wave diffraction and wave breaking.

IH2VOF is a numerical model that was developed by IH Cantabria. IH2VOF solves


the 2D Reynolds Averaged Navier–Stokes (RANS) equations at the clear fluid region
and the Volume-Averaged Reynolds Averaged Navier–Stokes (VARANS) equations
inside the porous media, based on the decomposition of the instantaneous velocity and
pressure fields into mean and turbulent components, and the κ-ε equations for the
turbulent kinetic energy κ, and its dissipation rate ε (IH2VOF website,
http://ih2vof.ihcantabria.com/). This permits the simulation of any kind of coastal
structure (e.g. rubble mound, vertical or mixed breakwaters). The free surface
movement is tracked by the volume of fluid (VOF) method for one phase only, water
and void (IH2VOF website, http://ih2vof.ihcantabria.com/). For the models based on
RANS equations, the most complete turbulence model is the Reynolds stress closure
model, which attempts to close the Reynolds stresses transport equation directly
(Launder and Spalding, 1974). The k-ε model is the most commonly used alternative
(Launder and Spalding, 1974; Rodi, 1980).

Liu et al. (1999) introduced a RANS model, namely COBRAS (Cornell Breaking
Waves and Structures) to simulate the overtopping phenomena of breaking waves on
a porous structure. The model solves the Reynolds averaged Navier–Stokes equations
to calculate the flow in the fluid region and the corresponding turbulence field is
modelled by an improved k–ɛ model. The flow in porous media was described by the
spatially averaged Navier–Stokes equations (Lara, 2005).

Hsu et al. (2002) introduced the Volume-Averaged/Reynolds Averaged Navier–Stokes


(VARANS) equations to improve the COBRAS model in order to define the surface
wave interaction with coastal structures. The volume-averaged Reynolds stress is
modelled by following the nonlinear eddy viscosity assumption in the VARANS
equations, and the volume averaged turbulent kinetic energy and its dissipation rate

19
are obtained by applying the volume-average of the standard k–ɛ equations. This
model has the advantage of introducing the small-scale turbulence effects as part of
the porous flow (Lara, 2005). Validation of the model is performed using the
experimental set in Liu et al. (1999). However, most of the validation focuses on
regular wave field in this validation study and it is also for a very small domain and
limited simulation time.

Lara (2005) presented an improved version of COBRAS, COBRASUC, developed by


University of Cantabria. The model is used to examine the interaction of random waves
with rubble mound breakwaters focusing on the overtopping phenomena. It is used to
simulate a large numerical wave flume including random waves and long simulation
times. The author states that these may help to reduce some uncertainties in using semi-
empirical formulae and give some statistical information on the overtopping process
(Lara, 2005).

20
CHAPTER 3

3 PHYSICAL MODEL EXPERIMENTS

3.1 General Description of Physical Model Experiments

Physical modelling is a process in which mainly the real-life problems are recreated
by a scale factor in laboratory conditions. Being one of the main techniques in coastal
engineering, physical modelling is used for observing the physical situation of the
hydraulic processes visually. Engineers may develop more creative and innovative
solutions with the help of these observations. It is not always possible to create the
exact environmental conditions of the natural system but an optimization can be made
to obtain reasonable results with reasonable simplifications in model experiments.
Most often, it is a cost and time efficient way compared to field observations and data
gathering in the field. Physical modelling has also an advantage of including the
turbulence effect which always exist in hydraulics.

On the other hand, there are also some disadvantages of laboratory experiments. Scale
effect is one of the major problems in reflection of the stresses and forces in the nature.
The scale effects and the deficiencies resulted from the inaccurately scaled parameters
like fluid density and viscosity may be minimized in large scale experiments but they
may be expensive and time consuming. Additionally, the ability of resources in the
laboratory to generate the necessary conditions could be limited. Thus, not every
condition pertinent to the problem can be observed.

As mentioned in Chapter 1, understanding the damage and failure behavior of coastal


structures is important to increase the coastal resilience. In Chapter 2, it is shown that
most of the missing information in the literature is about the performance of rubble

21
mound breakwaters under tsunami attack since rubble mound structures are not
common in Japan where the recent tsunami events provided much of the available
observations. Furthermore, many coastal structures are designed without considering
loads of tsunami-like waves, long waves or extreme waves although they are
constructed in areas prone to encounter these waves. Therefore, it is very important to
observe the damage behavior and failure mechanisms of rubble mound breakwaters
under different extreme wave conditions. In addition, comparison of the performance
of rubble mound and caisson type structures (which are more commonly observed
under tsunami attack) can provide additional insight to different failure mechanisms.
Observing two types of breakwaters in laboratory environment with accompanying
pressure measurements is the physical modelling part of this thesis. This chapter
presents the wave flume in which the tests are performed, the experimental setup, the
breakwater cross-sections, the measuring technique in the experiments and the
characteristics of the applied waves. The general process followed in this part of the
study is given in Figure 3.1.

22
Identifying the aim of the study

Determining the procedure


- Physical Modelling

Model Setup
- Determining the model scale
- Design of the breakwaters

Wave Generation
- Determining the wave conditions

Data Collection
- Wave Gauge Measurements
- Profile Measurements
- Pressure Measurements
- Photo Collection and Video Recording

Data Analysis

Figure 3.1: Flowchart of the model experiments

3.2 Wave Flume

The experiments were performed in the wave flume constructed in one of the basins
(smaller basin) at METU Coastal and Ocean Engineering Laboratory. The basin has a
length of 26.8 meter and a width of 6.1 meter (Figure 3.2). The dimensions of the wave
flume which is located in the inner channel of this basin are 18m in length, 1.5m in

23
width and 1.0 m in depth. The aim of constructing an inner flume is to reduce reflection
from the border walls of the basin which could distort the wave profile. This wave
flume has a horizontal bottom over the entire length.

Smaller Basin

Inner wave flume

Figure 3.2: The smaller basin and the wave flume

A piston type wave maker is placed at one end of the smaller basin which is capable
of generating regular and irregular waves and was used to generate the intended wave
conditions (Figure 3.3a). On the opposite end of the wave flume, there is a slope of
wave absorbers made of plastic wire scrubbers aiming to prevent reflection of the

24
waves from the end flume walls (Figure 3.3b). The layout of the wave basin is given
in Figure 3.4.

Figure 3.3: a) Piston type wave maker b) Wave absorbers

Figure 3.4: General layout of the basin (Top: top view, Bottom: Side view)

Dimensions are in cm and figure is not to scale.

3.3 Model Scale

Froude similitude law is applied in most of the model studies in the field of coastal
engineering because the effect of surface tension and elastic compression is rather

25
small and therefore can be neglected (Hughes, 1993). Gravity or viscous forces then
become the necessary parameters to define which take modellers to apply Froude
Theorem or Reynolds Theorem. One of these theorems together with the geometric
similarity would provide the appropriate conditions for the hydrodynamic similitude.
Frostick (2011) states that the main forces are gravity, friction and surface tension for
wave models. Therefore, it is suggested that Froude (Fr), Reynolds (Re) and Weber
(We) numbers must be the same in model and prototype for true dynamic similarity.
However, it is not possible to satisfy this condition. If Froude law is used in the model
(since the effects of gravity and inertia are dominant in the wave field) and the model
Reynolds number is in the same range as the prototype, then the Reynolds number
need not be exactly the same. Surface tension is generally negligible in prototype
waves. Therefore, if the model is not too small (wavelengths must be much greater
than 2 cm, wave periods >0.35 s, water depths >2 cm, Le Méhauté, 1976), Weber
similitude can be neglected (Frostick, 2011). In modelling flow through the structure
and forces on structures, if the Reynolds number is large enough and fully turbulent
flow conditions in the armour-/filter layers are maintained (ReD > 30000, e.g. Dai and
Kamel, 1969) the forces can be scaled by Froude. Viscous effects can be limited in the
model if diameters larger than 3–5 mm (in model scale) are used. The following model
boundary values can be derived regarding these conditions so that Froude similitude
law can be applied alone (Le Mehaute, 1976 and Hughes, 1993).
 water depth: >5 cm (water depth=40cm in the experiments)
 wave height >2–3 cm, design wave height: >5 cm (min wave height=9cm,
design wave height=10cm in the experiments)
 wave period realistic wave steepness (steepness range = 0.018 – 0.036 in the
experiments)
 rock diameter >3–5 mm (core layer rock diameter=12-21mm in the
experiments)
 rock armour >25 mm (armor layer rock diameter=28-30mm in the
experiments)

26
For breaking waves, the Weber and Cauchy numbers may be important since they
determine the air intake, water turbulence and water compressibility. However, the
waves applied in the experiments are not breaking waves.

Since all these criteria are met in this model study, Froude similitude law was applied
the physical model experiments. Froude criterion is a parameter which expresses the
relative effect of inertial and gravity forces in a hydraulic flow (Hughes, 1993). It is
commonly expressed in the square of the number and given by Equation 3.1:
𝑢2
𝐹𝑟 2 = [3.1]
𝑔𝑑

Where u is the velocity of water particle, g is gravitational acceleration and d is the


water depth.

According to the similitude law, the Froude numbers in prototype ((Fr)p) and the model
((Fr)m must be equal to each other which implies Equation 3.2.

(𝐹𝑟 ) 𝑝 = (𝐹𝑟 ) 𝑚 [3.2]

The scale factor, λX, is the ratio of any parameter in the model to the parameter in the
prototype. To satisfy the geometric similitude, the length scale factor, λ L, is the ratio
of dimensions (length parameters) of the model (Lm) to the dimensions of the prototype
(Lp) as given in Equation 3.3.

𝐿𝑚
𝜆𝐿 = [3.3]
𝐿𝑝

According to the square root of Equation 3.1., the Froude criterion is given by Equation
3.4.

𝜆𝑢 = √𝜆𝑔 𝜆𝐿 [3.4]

27
Since velocity (u) is the ratio of length to time, the scale ratio for velocity is equal to
Equation 3.5.

𝜆
𝜆𝑢 = 𝜆 𝐿 [3.5]
𝑇

The time scale then can be obtained from equations 3.4 and 3.5 as Equation 3.6 since
the gravitational scale is accepted as unity.
𝜆 𝑇 = √𝜆𝐿 [3.6]
For the weight scaling of the armour units of the breakwaters, the stability numbers in
both the prototype and the model should be equal to each other according to CERC
(1984). From the equalization of the stability numbers, the weight scale (λw) can be
derived as Equation 3.7.

(𝛾𝑟 ) 𝑝 3
⁄(𝛾 ) −1
(𝛾 )
(𝜆𝐿 3 ) (𝛾𝑟 ) 𝑚 [(𝛾𝑟 ) 𝑚
𝑤 𝑝
𝜆𝑊 = ] [3.7]
𝑟 𝑝 ⁄(𝛾 ) −1
𝑤 𝑚

In Equation 3.7, (γr) m and (γr) p are unit weights of armour units that are used in model
and prototype, respectively, which were taken as 2.6 t/m3 and 2.7 t/m3. Also, (γw) m
and (γw) m are unit weight of water that is used in model and unit weight of sea water
in prototype. In the experiments, unit weight of water was taken as 1.0 t/m3 and the
unit weight of sea water (prototype) was taken as 1.025 t/m3.

The pressure scale, 𝜆𝑝 ,is obtained by Equation 3.8 which is used for the scaling of the
pressure values.

(𝜌𝑤 )𝑝
𝜆𝑝 = (𝜌 𝜆𝐿 [3.8]
𝑤 )𝑚

The scale factors to reproduce the basic physical parameters are given in Table 3.1.

28
Table 3.1: Scaling factors using the similitude law of Froude

Scale factor Conversion to model Model


Parameter Unit
λL scale Scale

Length λL 1/λL 1/50 [m]


Volume (λL)3 1/(λL)3 1/125000 [m3]
1 1/7.07
Time √𝜆𝐿 ⁄√𝜆 [s]
𝐿

Weight λw 1 / λw 1/163003.3 [kg]


Pressure λp 1/λp 1/51.25 [kg.m/s2/m2]

The model scale is chosen 1/50 as it was defined according to the laboratory
restrictions and considering the advantages and disadvantages of small scale and large
scale experiments. The water depths that can be obtained in the flume are from 0.3m
to 0.7m. Operation limits of the wave generator which is another restriction of the
laboratory are 0.05 Hz < Frequency < 2.0 Hz and -290 mm < piston amplitude < 290
mm.

3.4 Laboratory and Scale Effects

Hughes (1993) states that both refraction and diffraction in Froude-scaled long-wave
hydrodynamic physical models are correctly recreated for both geometrically
undistorted and distorted models. However, as mentioned in Section 3.1, scale effect
is a problem in laboratory experiments in coastal engineering in reproduction of the
stresses and forces in the nature.

Wave reflection problems are amplified in long-wave models since long waves reflect
more of their energy and it may be more difficult to operate wave absorption systems
in case of longer wavelengths. Accurate reproduction of bathymetry is also required
to obtain reliable results. Also, if the measured hydrodynamic parameters are small,

29
the possibility of measurement error and instrument intrusion are also laboratory
effects that the researchers should be aware of (Hughes, 1993). To minimize these
laboratory effects, a horizontal platform explained in detail in Section 3.5 was
constructed to recreate the bathymetry as much as possible. The surface of the platform
was covered with sand to prevent smoothness of the surface. The measurement
instruments were operated carefully and their calibration was performed by several
trials until obtaining reasonable results.

Scale effects are discussed under the titles of wave reflection, wave transmission,
surface tension and viscosity and friction in Hughes (1993). He stated that the most
important scale effect related with the physical modeling of rubble mound structures
is the viscous forces due to the flow through the underlayers and core of the structure.
It is not a primary issue for the armor layer of the breakwater since the Reynolds
number is sufficiently large based on the dimension of the armor unit to satisfy the
fully turbulent flow. In the underlayers and core of the breakwater on the model, it is
necessary to increase the size of the core and underlayer materials since scaling of
material sizes may cause viscous scale effects which the layers become less permeable
and may cause higher pressures from inside the structure (Hughes, 1993).
To determine the diameter of the material in the model, Dm, methods of Le Méhauté,
(1990) and Keulegan (1973) give a distortion factor, K, that is used in Equation 3.8.

𝐿𝑝 𝐷𝑝
= 𝐾𝐷 or 𝜆𝐿 = 𝐾𝜆𝐷 [3.8]
𝐿𝑚 𝑚

Currently, as also suggested by Hudson et al. (1979), the K values obtained by Le


Méhauté (1990) and Keulegan (1973) methods can be averaged for application of long
wave physical models. Therefore, the diameters of the materials in the filter layer were
corrected by a factor of 1.1 and the diameters of the materials in the core layer were
corrected by a factor of 1.25 in the physical model experiments to minimize the effects
mentioned above.

30
In the physical model experiments, the effect of side walls of the wave flume was
eliminated by taking pressure measurements on the midline of the middle wall element
of the caisson type breakwater which was the most distant position to the side walls.
The boundary effect was also excluded in the damage analysis by observing the cases
which the vertical walls and the crown walls got stuck in the side walls of the flume
and taking into account these conditions in the damage behavior.

3.5 Experimental Setup

The experiments were carried out in the inner wave flume which was further divided
into two parallel sections (0.741 m wide each) by means of a vertical plywood plate to
be able to test the two breakwaters simultaneously. The bathymetry of the model
consisted of a plywood horizontal platform of slope 1:10, on which the breakwater
cross-sections were constructed (Figure 3.5). The platform height was 0.1 m for the
caisson type breakwater side and 0.2 m for the rubble mound breakwater. In order to
prevent the smooth surface of the platform and to reflect the roughness of the seabed
as much as possible, the surface of the platform was covered with aggregates.

Figure 3.5: View of horizontal platform in the wave flume

A caisson type (vertical wall) breakwater with a rubble foundation (Figure 3.6) was
constructed at one section and a rubble mound breakwater with a small berm (Figure

31
3.7) was placed at the other section. The cross-sections were designed according to the
design criteria which has a significant wave height of 5m and a significant wave period
of 9.5s. These cross-sections are generic ones which are widely observed in coastal
areas as vertical wall and rubble mound breakwaters.

Figure 3.6: Caisson Type Breakwater

Figure 3.7: Rubble Mound Breakwater

The caisson type breakwater cross-section is given in Figure 3.8. Both the seaside and
harbor side slopes of the rubble foundation in caisson type breakwater were 1:4. The
total length of the cross-section was 1.55m whereas the horizontal platform had a total
length of 3.63m. The crest elevation of the vertical wall is 0.08m above the still water
level.

The rubble mound breakwater cross-section is given in Figure 3.9. The seaside armor
slope in the rubble mound breakwater was 2:5 whereas the harbor side slope was 1:1.5.
The total length of the cross-section was 1.3m. The platform had a total length of 4.7
m. The crest elevation of the crown wall was 0.04m above the still water level.

32
Figure 3.8: Cross-section of the caisson type breakwater with the horizontal
platform (Dimensions are in meters and the figure is not to scale.)

Figure 3.9: Cross-section of the rubble mound breakwater with the horizontal
platform (Dimensions are in meters and the figure is not to scale.)

3.5.1 Rubble Mound Breakwater Cross-Section


The crown wall, used in the rubble mound breakwater consisted of three similar
concrete elements, which were 0.245 m wide (left element, element in the middle, and
the right element in respect to the direction of wave propagation), 0.09 m long, and
0.09 m high as presented in Figure 3.10 and Figure 3.11.

33
Figure 3.10: Crown wall element used in rubble mound breakwater from different
perspectives (scale 1:50)

Figure 3.11: Placement of the crown wall elements used in rubble mound
breakwater

The stones used for the construction of the rubble-mound breakwater were arranged
manually in three layers, separated by using different colors according to their mass
range:
• core layer (dark grey color): mass of stones between 5 and 25 g (0.4 – 2 t in
prototype),

34
• filter layer (white color): layer thickness of 0.06 m, mass of stones between
15 and 35 g (2 – 4 t in prototype),
• armour layer on the seaside (red color): layer thickness of 0.06 m, mass of
stones between 60 and 75 g (10 - 12 t in prototype),
• armour layer on the side of the port (white color): layer thickness of 0.06 m,
mass of stones between 15 and 35 g (2 - 4 t in prototype).
The arrangement of the layers in the investigated breakwater is shown in Figure 3.12.

Figure 3.12: Geometric Details of Rubble Mound Cross-Section (Scale 1:50)

3.5.2 Caisson Type Breakwater Cross-Section


The concrete blocks used in the vertical wall breakwater consisted of three similar
concrete elements, which were 0.245 m wide (left element, element in the middle, and
the right element in respect to the direction of wave propagation), 0.19 m long, and 0.3
m high as presented in Figure 3.13. The concrete blocks were produced as boxes to be
filled with sand. The blocks were soaked in water for about 24 hours to obtain fully
saturated conditions and then filled with sand by compressing (Figure 3.14).

35
Figure 3.13: Layout of a concrete block used in caisson breakwater (scale 1:50)

Figure 3.14: Placement of concrete blocks used in caisson type breakwater

36
The stones used for the construction of rubble foundation of the vertical wall
breakwater were arranged in two layers,
 core layer (dark grey color): mass of stones between 5 and 25 g (0.4 – 2 t in
prototype),
 armor layer (white color): layer thickness of 0.06 m, mass of stones between 15 and
35 g (2 – 4 t in prototype).

The arrangement of the layers in the investigated breakwaters is shown in Figure


3.15.

Figure 3.15: Geometric Details of the Caisson Cross-Section (Scale 1:50)

The placement of the two breakwater cross-sections in the flume can be seen in Figure
3.16

37
Figure 3.16: Placement of the two cross-sections

3.6 Measuring Technique

Eight wave gauges and two pressure sensors were used in the experiments to collect
data for each breakwater as shown in Figure 3.20 and 3.21 as well as in Table 3.2.

Wave Recording
Wave gauges (WG) were installed along the wave flume and over the breakwater
models to measure the water surface elevation (Figure 3.17). These wire type wave
gauges, (DHI (Danish Hydraulic Institute)-Wave Meter wave gauges) consist of two
parallel stainless steel electrodes which are conductive materials. The electrodes are
placed perpendicular to the wave direction. The bottom part of the wave gauges is the
recompensating electrodes which minimize the effect of salinity and temperature
effects.

38
Figure 3.17: Wire type DHI-Wave Meter wave gauge

Seven of the wave gauges were installed to measure the surface elevation (wave
height) and the wave period (WG1, WG3-WG8). One wave gauge (WG2) was used
for measuring overtopping and it was installed onto the front surface of the crown wall
in rubble mound breakwater and onto the front surface of the concrete block in caisson
type breakwater (Figure 3.18a and 3.18b). The wave gauges except (WG2) were
submerged up to half of the water depth of 0.4 m.

Figure 3.18: a) Exemplary arrangement of overtopping wave gauge (WG2) in


caisson breakwater b) Exemplary arrangement of overtopping wave gauge (WG2) in
rubble mound breakwater

39
For data from the wave gauges recording, an instrument system developed by Danish
Hydraulics Institute (DHI) was used. The wave gauges measure the voltage changes
in water and the system records the signal of these changes and transfers the record to
the computer. This voltage data is collected in the computer as .txt files by using a
software developed by TDG Data Concept Inc. The voltage data is then converted
into water surface elevation as the change of water surface elevation is proportional to
the change in voltage measured by the wave gauges.

The calibration of the recorded data is necessary to obtain this relationship between
recorded raw voltage data and the change of water surface elevation. Thus; the
following procedure was performed before each experiment:
- the still water level was lowered and raised after fixing the wave gauges at the
same level,
- the stillness of the water level was ensured
- the sensor data was recorded at desired water levels in unit volt and the
corresponding water level heights in unit lengths were used to plot calibration
lines.
In the model experiments, calibration was performed at the beginning of each
experiment by changing the water level at +5cm, 0cm and -5cm. The calibration
coefficients were obtained by using the MATLAB code written by Baykal (2009).
A graph showing an example of calibration is given in Figure 3.19.

Calibration - WG3
0,4
0,3
0,2
Voltage (V)

0,1
0
-6 -4 -2 -0,1 0 2 4 6
-0,2
-0,3
-0,4
Water Level (cm)

Figure 3.19: Example graph of calibration for WG3

40
Figure 3.21: Experimental setup including measuring instruments for caisson type breakwater

41
Figure 3.20: Experimental setup including measuring instruments for rubble mound breakwater
In summary, wave gauge WG1 is installed to measure the transmitted wave and WG2
is installed to measure the overtopping height whereas wave gauges WG3 – WG8 are
installed to measure the incident and reflected wave heights. The positions of the wave
gauges WG3 – WG4 and WG6 – WG7 are close to each other (20 cm distance) to
make the reflection analysis and obtain the reflected wave heights. The position of
measuring devices in tests with respect to the initial position of wave generator is given
in Table 3.2.

Table 3.2: Position of measuring devices in tests with respect to the initial position
of wave generator

(Caisson Type Breakwater) (Rubble Mound Breakwater)

Measuring Distance from the wave Measuring Distance from the wave
Device maker Device maker
[m] [m]
WG1 17.669 WG1 18.947
WG2 17.550 WG2 18.842
WG3 16.870 WG3 18.100
WG4 16.670 WG4 17.900
WG5 16.540 WG5 15.600
WG6 7.800 WG6 7.800
WG7 7.600 WG7 7.600
WG8 0.000 WG8 0.000

Profile Measurement
Before and after each test, the profiles of the breakwaters were measured manually at
1cm intervals using a laser meter along two lines that are shown by the white rods in
Figure 3.22.

42
Measurement Axis
Measurement Axis

Figure 3.22: Profile measurement lines

Furthermore, photos of the breakwaters were taken before and after each test to record
and analyse the damage visually. Additionally, there were three video cameras
recording the tests from different angles: one was installed recording the caisson type
breakwater from the side view, one recorded the rubble mound breakwater from the
side view and a GoPro type camera was installed observing the rubble mound
breakwater from the top view.

Figure 3.23: Observation window and video recording system

43
Pressure Measurements
Pressure measurements were taken along the front surface and the bottom of the
caisson type breakwater and along the bottom of the breakwater by two micro pressure
transducers (Figure 3.24). The pressure transducer (model P310) is a disk type pressure
transducer developed by integrated SSK transducer technology and used for hydraulic
experiments as well as for other applications such as soil pressure measurements. The
features of the transducers can be listed as being micro type, low capacity, having high
sensitivity, high response frequency and a waterproof structure.
The transducers are 100 kPa sensors and their natural oscillation frequency is 9.1kHz.
The operating temperature range for the transducers is in between -100C and +550C.

Figure 3.24: Micro pressure transducer used in pressure measurements

Pressure measurements were taken from 2 different points above the still water level,
from 1 point on the still water level and 5 different points below the still water level
along the front surface of the vertical wall element of caisson type breakwater. These
points are shown in Figure 3.25.

44
Figure 3.25: Pressure measurement points along the front surface of caisson
breakwater (Dimensions are in cm.)

Figure 3.26: Placement of two pressure transducers on the front surface of the
vertical wall element at still water level

Pressure measurements were also taken from 6 different points along the bottom
surface of the middle vertical wall element in caisson type breakwater. These points
are shown in Figure 3.27.

45
Figure 3.27: Pressure measurement points along the bottom surface of caisson
breakwater (Dimensions are in cm.)

For the calibration of the data obtained from the transducers and for pretesting of the
transducers, 6 static calibration tests were performed by recording data by placing the
sensors at different water levels. The results obtained from these tests are given in
Figure 3.28 and Figure 3.29 for the two transducers.

100 kPa - P310-1 Ser.12971


10

8
Pressure (kPa)

-2
0 20 40 60 80 100
Sensor height (cm)

Figure 3.28: Results from 6 different static calibration pretests of the transducer 1

46
100 kPa - P310-1 Ser.12972
10

8
Pressure (kPa)

-2
0 20 40 60 80 100
Sensor height (cm)

Figure 3.29: Results of 6 different static calibration pretests of transducer 2

As can be seen from the results, the transducers almost gave consistent results for the
hydrostatic pressure for different levels.

Also, a dynamic water tank test was performed for pretesting of the transducers by
submerging the sensors into a water tank in a periodic manner manually between -5
cm and -95 cm water levels. Results of the test are presented in Figure 3.30.

1,2

1
Water Level (m)

0,8

0,6

0,4

0,2

0
0 10 20 30 40 50 60
Time (s)

P310-1 Ser.12971 P310-1 Ser.12972

Figure 3.30: Results of the dynamic water tank test of the transducers

47
The sampling interval was 5kHz in the experiments to keep the fine data to use it in
case of requirements. The calibration and other processes to analyze the data obtained
from the transducers, a Matlab code was executed. In the code, first the data record as
.txt files was transferred and the parameters like water depth, the sampling interval and
the depths of sensor positions were defined. Since it was difficult to analyze such fine
data, a smoothing process was included in the code. The code sorted the fine data by a
low pass filter process with a cut-off frequency of 20Hz. This process is presented in
the experimental study of Kısacık et al. (2012) on pressure measurements on a vertical
wall. The calibration coefficients were calculated according to the premeasured water
level records for calibration. After that, the mean water level correction was made.
Finally, the offset values resulting from the depths of sensor positions were obtained.

3.7 Wave Characteristics

Holthuijsen (2007) ordered the waves in terms of their period and wavelength as given
in Figure 3.31. Goda (2010) also stated that long-period waves are mainly referred to
the ones with periods of 30 to 300s.

Figure 3.31: Frequencies and periods of the vertical motions of the ocean surface
(Holthuijsen, 2007)

48
Hedges (1995) also suggests that since ocean waves propagate on water together with
currents driven by the wind, the tidal forces or gravity, it is more senseful to use the
wavelength instead of wave period in defining the limits to the validity of the various
analytical theories which exist for describing trains of regular waves. Therefore, he
suggests Ursell number (Equation 3.9) to use as the governing parameter in relatively
shallow water and the wave steepness (H/L) in deep water. The graph suggested
showing approximate regions of validity of analytical wave theories is given in Figure
3.32.
𝐻∗𝐿2
𝑈= [3.9]
𝑑3

Where U is the Ursell number, H is wave height, L is wavelength and d is the water
depth.

Dean and Dalrymple (2000) also defines long wave limit according to wavelength and
the water depth and gives the following limitation:
2𝜋 𝜋
𝑑 < 10 [3.10]
𝐿

Figure 3.32: Approximate regions of validity of analytical wave theories

(Hedges, 1995)

49
The wave characteristics for this study was initially determined by scanning a range of
different wave heights and periods, wave steepnesses, in the range of nonlinear
(cnoidal) waves. Therefore, 18 different waves have been defined at first as given in
Table 3.3. However, some of these waves (wave no: 7, 13 and 14) were eliminated
since they could not be generated due to the limited stroke length of the wave piston.
Waves with wave number 4, 5 and 6 were also eliminated as they were in the linear
region according to Hedges (1995) study. For the other waves, trial experiments were
carried out before the cross-sections were constructed in the flume and the wave
heights and periods were measured at the toe of the structures. According to these
measurements which are given in Table 3.4. wave heights of some of these waves were
very low since they were close to the long wave region (wave no: 1, 2, 3 and 8). These
waves were also not included in the study since the main focus of this study is to
observe the damage and failure mechanisms of the specified breakwaters.

50
Table 3.3: Wave characteristics before elimination

Wave Number 1 2 3 4 5 6 8 9 10 11 12 14 15 16 17 18

H (m) 0.05 0.05 0.05 0.05 0.05 0.05 0.1 0.1 0.1 0.1 0.1 0.15 0.15 0.15 0.15 0.15
H0/L0 0.001 0.002 0.004 0.006 0.01 0.015 0.002 0.004 0.006 0.01 0.015 0.002 0.004 0.006 0.01 0.015
T (s) 5.66 4.00 2.83 2.31 1.79 1.46 5.66 4.00 3.27 2.53 2.07 6.93 4.90 4.00 3.10 2.53
L 11.12 7.79 5.42 4.35 3.25 2.53 11.12 7.79 6.31 4.80 3.84 13.65 9.60 7.79 5.97 4.80
kd 0.23 0.32 0.46 0.58 0.77 0.99 0.23 0.32 0.40 0.52 0.65 0.18 0.26 0.32 0.42 0.52
d/L 0.04 0.05 0.07 0.09 0.12 0.16 0.04 0.05 0.06 0.08 0.10 0.03 0.04 0.05 0.07 0.08
H,prototype 2.5 2.5 2.5 2.5 2.5 2.5 5 5 5 5 5 7.5 7.5 7.5 7.5 7.5
T,prototype 40.0 28.3 20.0 16.3 12.7 10.3 40.0 28.3 23.1 17.9 14.6 49.0 34.7 28.3 21.9 17.9
S0 (m) 0.22 0.16 0.11 0.09 0.07 0.05 0.44 0.31 0.25 0.19 0.15 0.81 0.57 0.47 0.36 0.29

51
Green: Long Wave Limits
Red: Insufficient Stroke

Table 3.4: Wave characteristics before elimination, measured values

Wave Number 1 2 3 4 5 6 8 9 10 11 12 14 15 16 17 18

H,toe - RM (m) 0.004 0.008 0.017 0.036 0.045 0.027 0.01 0.023 0.047 0.099 0.089 0.038 0.059 0.054 0.096 0.151
H,toe -
Vertical (m) 0.021 0.033 0.04 0.05 0.07 0.065 0.039 0.07 0.057 0.08 0.113 0.039 0.086 0.119 0.095 0.118
T (s) 5.2 4 2.8 2.2 1.75 1.5 5.5 4.5 3.2 2.5 2.1 6.8 4.95 4.3 3.1 2.4

Blue: Low Wave Height


Six different wave conditions with wave numbers 9, 10, 12, 15, 17 and 18 were applied
in the experiments at a water depth of 0.4m. The characteristics of these waves are
presented in Table 3.5 and Table 3.6. The characteristics are obtained from the
measurements of WG3 (at breakwater toes) in the flume before the construction of the
breakwater cross-sections. The studied waves are pointed in Figure 3.32 with red
points to illustrate the region where these waves correspond to. The region of the wave
periods of these waves is also illustrated by the red frame in Figure 3.31.

Table 3.5: Characteristics of applied waves in model scale, caisson type breakwater

Wave Number 9 10 12 15 17 18
Period (s) 4 3.2 2 4.95 3.1 2.5
Wave Height (m) 0.22 0.11 0.09 0.17 0.15 0.13
Steepness 0.028 0.018 0.024 0.018 0.026 0.025

Table 3.6: Characteristics of applied waves in model scale, rubble mound breakwater

Wave Number 9 10 12 15 17 18
Period (s) 4 3.2 2 4.95 3.1 2.5
Wave Height (m) 0.2 0.13 0.09 0.19 0.18 0.17
Steepness 0.026 0.021 0.024 0.02 0.03 0.036

As can be seen from Table 3.5 and 3.6, the wave heights are changing in a range of
0.09 – 0.22 m and wave periods are from 2s to 4.95s in model scale. The steepness
range is 0.018 – 0.036. The wave heights are in a range of 4.5m – 11m in prototype
and the wave periods correspond to a range of 14s – 35s in prototype.

Furthermore, the design wave parameters were specified as 5m of significant wave


height (Hs = 5m) and 9.5s of significant wave period (Ts = 9.5s). The cross-sections
were designed according to these wave characteristics but tested under the specified
nonlinear waves which represent the possible extreme wave conditions for such a
design case. The measured wave heights at the structure toe and the ratio of these

52
values to the design wave height (Htoe/Hdesign) in prototype is presented in Table
3.7.

Table 3.7: Ratio of measured wave heights to the design wave height in prototype

Wave No: 9 10 12 15 17 18
Htoe-measured 10.5 6 4.5 9 8.25 7.25
Htoe / Hdesign 2.1 1.2 0.9 1.8 1.65 1.45

To generate the applied waves in the laboratory, the amplitude of paddle displacement
and the frequency values were entered to the Moog device which was connected to the
piston. In order to obtain the paddle displacement (stroke length), Biesel transfer
functions (as cited in Hughes, 1993) which express the relation between wave
amplitude and wave paddle displacement are used. Biesel transfer function for piston
type paddle is given in Equation 3.11 and the graph of the equation is given in Figure
3.33.

𝐻 2 sinh 2 (𝑘ℎ)
= sinh 𝑘ℎ cosh 𝑘ℎ+𝑘ℎ [3.11]
𝑆0

Where H is the wave height, S0 is the paddle displacement and h is the water depth.

Figure 3.33: Biesel Transfer Function to obtain wave paddle displacement (Hughes,
1993)

53
These cnoidal waves were generated as regular waves from the piston. To obtain the
paddle displacement values for the necessary wave parameters, series of trials were
performed by changing the paddle displacement and measuring the recorded data.
Finally, the results of these trials were fitted to a curve which give the wave height
over paddle displacement (H/S0) ratio.

For the analysis of the data obtained from wave gauge measurements, a simple matlab
code was used which execute the calibration and arithmetic correction processes. In
the calibration process, the voltage data was converted into the water level elevation
data. In the arithmetic correction process, mean of all data was subtracted from the
time series recorded to fit the arithmetic mean to zero. Thus, the raw data was
converted into distances from the mean water level. An example of the surface profile
is given in Figure 3.34. The second wave is taken as the representative wave of the
measurements and used for calculations.

Figure 3.34: Sample surface profile

All the experiments were repeated for 2 times to be able to make a comparison between
the repetitions and to obtain as accurate results as possible.

54
3.8 Summary of Physical Model Experiments

A summary of the experimental setup and programme for the tests performed at METU
is presented in Table 3.8.

Table 3.8: Summary of physical model experiments at METU

Experiments at METU
Facility Wave flume: 18×1.5×1.0 m, two parallel sections of width of
0.741 m for testing two breakwaters simultaneously
Model scale 1:50
 Caisson type breakwater with rubble foundation
Breakwaters Rubble mound breakwater with a crown wall unit and a small
berm
 Caisson Type:
- Seaside and harbor slopes of rubble foundation 1:4
- Height: 0.38 m
- Caisson length: 0.19 m
Breakwater- Breakwater basis length: 1.23 m
geometry  Rubble mound:
- Seaside slope 2:5
- Harbor slope 1:1.5
- Height: 0.3 m
- Crown length: 0.09 m
- Breakwater basis length: 1.29 m
 Core layer: 5 - 25 g
Armour  Filter layer: 0.07 m thick, 15 - 35 g
layers  Armour layer on seaside: 0.09 m thick, 60 – 75 g
 Armour layer on harbour side: 0.07 m thick, 15 - 35 g

55
Crown wall  Concrete
units  L-shape, all surfaces perpendicular
 3 units: 0.245/0.245/0.245 m wide, 0.09 m long, 0.09 m high
Concrete  Concrete boxes with full of compressed sand
Block Units  L shape, all surfaces perpendicular
 3 units: 0.245/0.245/0.245 m wide, 0.19 m long, 0.3 m high
 Caisson type breakwater:
Horizontal platform of height of 0.1 m with seaside slope 1:10
Bathymetry and harbour slope 2:5.2
model  Rubble mound breakwater:
Horizontal platform of height of 0.2 m with seaside slope 1:10
and harbour slope 2:4
Measuring  8 wave gauges in each breakwater
devices  2 pressure sensors
 6 different cnoidal waves
Flow regime Wave heights: 0.09 – 0.22m
 Wave periods: 2s – 4.95s
 Water depth: 0.40m

56
CHAPTER 4

4 NUMERICAL MODELLING

4.1 Introduction to the Numerical Model, IH-2VOF

Interest in numerical modelling of coastal structures has increased in recent years since
the computer resources have improved allowing larger storage capabilities, faster
central processing units (CPUs) and cheaper computers. The state-of-the-art numerical
models are also well calibrated and validated in terms of porous media flow,
overtopping phenomenon and pressure fields. Therefore, it can be said that numerical
modelling is now an available tool as complementing the physical modelling and for
design purposes.

As stated in Section 3.1, the classical approach in coastal engineering is physical model
testing. However, it has some handicaps and therefore, new methods to improve the
wave-structure interaction analysis are needed. At this point, integrating numerical
models can provide further insight. Main approaches in numerical modelling can be
schematized as given in Figure 4.1.

Figure 4.1: Schematic diagram of main approaches in numerical modelling in


coastal and harbor engineering

57
Non-linear shallow water equations and Bousinesq Equation models may be cheap
according to the computational cost, may be efficient and allow long simulations with
reliable statistics. On the other hand, wave breaking has to be triggered in the
computations and vertical structures can not be simulated. Also, overtopping can not
be calculated and pressure field is not provided.

SPH models have nonlinear interaction in the fluid and pressure field can be obtained.
It also has an advantage of modelling multi-connected free surfaces. However, these
models are subject to the ongoing research and the porous flow is not solved yet.
Pressure field also needs to be tailored and adapted.

Navier-Stokes models (RANS) solve the non-linear interaction in the fluid and the
pressure field is obtained. They allow modelling of multi-connected free surfaces,
porous media flow and turbulence is considered. The models may have high
computational time (process time) and their efficiency may be disputable. However,
the models are increasingly becoming a methodology of use in coastal engineering due
to several factors listed as the following:
- The number of underlying assumptions is quite low,
- Non-linear flow characteristics are considered,
- Wave dispersion is included,
- Breaking is not triggered artificially,
- Wave induced turbulence can be considered.

The IH-2VOF model solves the two-dimensional Reynolds Averaged Navier-Stokes


(RANS) equations, based on some assumptions. The model characterizes the flow
inside and outside of coastal structures which include permeable layers. It uses Navier-
Stokes model as an engineering tool. It includes mesh design as mesh dimension and
resolution, sea state definition, porous media parameters definition and post-process
and results management like run-up overtopping, loads and moments. Most
importantly for this study, the model allows modelling of both vertical breakwaters

58
and rubble mound breakwaters and has been validated by physical model experiments
and several studies as stated in Chapter 2.
The numerical model has been validated at laboratory scale for wave breaking over
porous and impermeable slopes, for submerged permeable sea dikes, for low crested
permeable sea dikes, for vertical breakwaters and for rubble mound breakwaters (Lara
et al., 2002; Garcia et al., 2004; Guanche et al., 2009). A prototype validation is also
necessary to obtain porous media friction coefficients for laboratory scale and to
overcome scale effects at laboratory scale. The model has been validated for vertical
breakwaters and rubble mound breakwaters in prototype by Guanche et al. (2009).

4.2 Theoretical Background of IH2VOF Model

Fluid domain governing equations: RANS Equations


In a turbulent flow, the instantaneous velocity field, ui, and pressure field, p, can be
split into two parts, the mean velocity and pressure components, 𝑢̅ 𝑖 , and, 𝑝̅ , and the
turbulent velocity and pressure fluctuations, 𝑢𝑖′ and 𝑝′ as given in Equation 4.1.

𝑢𝑖 = 𝑢̅ 𝑖 + 𝑢𝑖′ and 𝑝 = 𝑝̅ + 𝑝′ [4.1]

Where i=1,2 for a bidimensional flow.

Applying the former decomposition to the Navier-Stokes equations, assuming


incompressible fluid and after time averaging, the continuity equation (Equation 4.2)
and the momentum equations (Equation 4.3) imply the following.

̅𝑖
𝜕𝑢
=0 [4.2]
𝜕𝑥 𝑖

̅̅̅̅̅̅̅
′ ′
̅𝑖
𝜕𝑢 ̅
𝜕𝑢 1 𝜕𝑝̅ 1 𝜕𝜏̅ 𝑖𝑗 𝜕𝑢 𝑖 𝑢𝑗
+ 𝑢̅ 𝑗 𝜕𝑥 𝑖 = − 𝜌 𝜕𝑥 + 𝜌 − + 𝑔𝑖 [4.3]
𝜕𝑡 𝑗 𝑖 𝜕𝑥𝑗 𝜕𝑥𝑗

59
𝜏𝑖𝑗 = 2𝜇𝑆𝑖𝑗 where
̅̅̅ [4.4]
̅
1 𝜕𝑢 ̅
𝜕𝑢
𝑆𝑖𝑗 = 2 (𝜕𝑥 𝑖 + 𝜕𝑥 𝑗) [4.5]
𝑗 𝑖

̅̅̅̅̅̅
In the momentum equation (Equation 4.3), the Reynolds stress tensor (𝜌𝑢 ′ ′
𝑖 𝑢𝑗 )

represents the influence of the turbulence fluctuations on the mean flow field and
requires a closure hypothesis. Therefore, in the model, this term is closed with a second
order model. A summary of several closure models can be found in Jaw and Chen
(1998) including the one in the model. In IH-2VOF model, the Reynolds stress tensor
is assumed to be related to the strain rate of mean flow through the algebraic nonlinear
k-ε model as given in Equation 4.6 (Yang and Shih, 1993; Lin and Liu, 1999).

2 𝑘 ̅𝑖
𝜕𝑢 ̅𝑗
𝜕𝑢 2
̅̅̅̅̅̅
𝜌𝑢 ′ ′
𝑖 𝑢𝑗 = 3 𝜌𝑘𝛿𝑖𝑗 − 𝐶𝑑 𝜌 𝜀 (𝜕𝑥 + 𝜕𝑥 ) −
𝑗 𝑖

̅̅̅ 𝜕𝑢
𝜕𝑢 ̅̅̅ ̅̅̅𝑗 𝜕𝑢
𝜕𝑢 ̅̅̅𝑙 ̅̅̅ 𝜕𝑢
2 𝜕𝑢 ̅̅̅̅
𝐶1 (𝜕𝑥𝑖 𝜕𝑥𝑙 + 𝜕𝑥 − 3 𝜕𝑥 𝑙 𝑘
𝛿𝑖𝑗 ) +
𝑙 𝑖 𝑙 𝜕𝑥𝑖 𝑘 𝜕𝑥𝑙
𝑘3
𝜌 𝜀2 [4.6]
𝜕𝑢 ̅̅̅𝑗
̅̅̅𝑖 𝜕𝑢 ̅̅̅𝑙 𝜕𝑢
1 𝜕𝑢 ̅̅̅̅
𝑘 ̅̅̅̅
𝜕𝑢 ̅̅̅̅
𝑘 𝜕𝑢 𝑘 ̅̅̅𝑙 𝜕𝑢
1 𝜕𝑢 ̅̅̅𝑙
+𝐶 ( − 𝛿 ) + 𝐶3 ( 𝜕𝑥 − 3 𝜕𝑥 𝛿 )
[ 2 𝜕𝑥𝑘 𝜕𝑥𝑘 3 𝜕𝑥𝑘 𝜕𝑥𝑙 𝑖𝑗 𝑖 𝜕𝑥𝑗 𝑘 𝜕𝑥𝑘 𝑖𝑗 ]

In which Cd, C1, C2 and C3 are empirical coefficients, 𝛿𝑖𝑗 is the Kronecker delta and k
is the turbulent kinetic energy defined as in Equation 4.7.

1
̅̅̅̅̅̅
𝑘 = 2 (𝑢 𝑖 𝑢𝑖 )
′ ′ [4.7]

The dissipation rate of the turbulent kinetic energy is defined as in Equation 4.8.

̅̅̅̅̅̅̅̅̅̅̅̅̅
𝜕𝑢𝑖′
𝜀 = 𝜗( ⁄𝜕𝑥 ′ ) 2 [4.8]
𝑗

𝜇
Where 𝜗 = ⁄𝜌 is the molecular kinematic viscosity.

60
It is important to note that the condition C1=C2=C3=0 in Equation 4.6 leads to the
conventional linear (isotropic) eddy viscosity model for the Reynolds stresses closure
giving Equation 4.9.
2
̅̅̅̅̅̅
𝑢𝑖′ 𝑢𝑗′ = −2𝜗𝑡 ̅̅
𝜎̅̅
𝑖𝑗 + 3 𝑘𝛿𝑖𝑗 [4.9]

Where 𝜗𝑡 is the eddy viscosity expressed as:


2
𝜗𝑡 = 𝐶𝑑 (𝑘 ⁄𝜀 ) [4.10]

The values for the coefficients C2 and C3 are obtained from experimental results on
turbulent shear flow by Champagne et al. (1970). The C1 coefficient is deduced from
the assumption by Shih et al. (1996): C1 = 2C3. A value for Cd is finally proposed by
Rodi (1980). The values for the coefficients are summarized as the following:

Cd = 0.09, C1 = 0.0054, C2 = -0.0171, C3 = 0.0027 [4.11]

However, considering constant values for these coefficients may lead to inconsistent
physical situations in the momentum equation (Equation 4.3) under some extreme
circumstances such as negative turbulence energy or infinite nonlinear contributions.
Hence, modified expressions for the empirical coefficients have been implemented in
the IH-2VOF model:

2 1 1 1 1
𝐶𝑑 = 3 (7.4+𝑆 ), 𝐶1 = 185.2+𝐷2 , 𝐶2 = 58.5+𝐷2 , 𝐶3 = 370.4+𝐷2 [4.12]
𝑚𝑎𝑥 𝑚𝑎𝑥 𝑚𝑎𝑥 𝑚𝑎𝑥

where;
𝑘 ̅̅̅
𝜕𝑢
𝑆𝑚𝑎𝑥 = 𝜀 𝑚𝑎𝑥 [|𝜕𝑥̅𝑖 |] [4.13]
𝑖

𝑘 ̅̅̅
𝜕𝑢
𝐷𝑚𝑎𝑥 = 𝜀 𝑚𝑎𝑥 [|𝜕𝑥̅̅̅𝑖 |] [4.14]
𝑗

61
It is important to note that all coefficients take their originally proposed values when
Smax and Dmax are zero. The governing equations for 𝑘 and 𝜀 are presented in Rodi
(1980) and Lin and Liu (1999).

Flow in porous media governing equations: Volume-Averaged RANS Equations


(VARANS)
The main assumption of the IH-2VOF model consist in considering that RANS
equations coupled with an appropriate turbulence model (𝑘 − 𝜀 model) can describe
the flow field in the porous media. To make the fluid/porous structure interaction
modelling easier, a volume-averaging process has been applied to the RANS and the
𝑘-𝜀 equations.

The flow in porous media is obtained in the IH-2VOF model through the resolution of
the Volume-Averaged Reynolds Averaged Navier-Stokes equations. The complete
mathematical formulation is presented in Hsu et al. (2002). These equations are
derived by integration of the RANS equations over a control volume. The size of the
averaging volume is chosen much larger than the characteristic pore size (microscopic
scale) but much smaller than the characteristic length scale of the flow, i.e. the scale
of the spatial variation of the physical variables in the fluid domain (macroscopic
scale). Schematic representation of the mathematical approach is given in Figure 4.2.

Figure 4.2: Sketch of volume averaging process for resolution of the porous flow
(IH-2VOF Course Lecture Notes, 2012)

62
The free surface movement is tracked by the volume of fluid (VOF) method, for only
one phase, water and void (http://ih2vof.ihcantabria.com/physics). In order to replicate
solid bodies immersed in the mesh, instead of treating them as sawtooth-shape, the
model uses a cutting cell method first which is given by Clarke et al. (1986). This
method uses orthogonal structured mesh in modelling to save computational time. In
this method, an openness function θ is defined to specify the fraction of volume of free
space in the cell. According to this fraction, if θ=0 it is defined as a ‘’solid cell’’
(entirely occupied by the solid), if θ=1 it is defined as a ‘’fluid cell’’ and if 0<θ<1 it is
defined as a ‘’partial cell’’. The original variables in the cell or on the cell faces are
multiplied by the openness coefficients to be redefined. To numerically carry out this
process, partial cell coefficients are specified at the cell centers and boundaries: θl, θr,
θt, θb which correspond to the openness at the left, right, top and bottom, respectively.
As a result, the redefined governing RANS equations are given as the following:

𝜕(𝜃𝑢𝑖 )
=0 [4.15]
𝜕𝑥𝑖

𝜕(𝜃𝑢𝑖 ) 𝜕(𝜃𝑢𝑖 ) 𝜃 𝜕𝑝 𝜃 𝜕(𝜏𝑖𝑗 )


+ 𝜕𝑢𝑗 = − 𝜌 𝜕𝑥 + 𝜌𝑔𝑖 + 𝜌 + 𝑓𝑏 [4.16]
𝜕𝑡 𝜕𝑥𝑗 𝑖 𝜕𝑥𝑗

The last term in Equation 4.16 is a virtual force. It appears because the IH-2VOF model
considers two different numerical methods to simulate moving bodies within the
computational domain (Raosa, 2012). The first method is ‘’virtual force method’’
(Mittal and Iaccarino, 2005) and the second one is the ‘’direct forcing method’’
(Mohd-Yusof, 1997).

Initial Conditions and Boundary Conditions


The model considers zero velocities and hydrostatic pressure as the initial conditions
for the mean flow in the domain still water.

Regarding turbulence, the initial conditions can not be zero for the mean flow since
the turbulence generation represented by Reynolds stresses is proportional to the

63
turbulent kinetic energy 𝑘, as described in Equation 4.6. If the initial condition for 𝑘
is zero, no turbulence energy would be produced during the whole simulation and
mathematical singularities would occur in the equations. Therefore, the model uses an
initial value for the turbulence energy that produces a numerical disturbance given in
Equation 4.17.

1
𝑘 = 2 𝑢𝑡2 [4.17]

With 𝑢𝑡 = 𝛿𝑐𝑖 where 𝑐𝑖 is the wave celerity in the generation zone and 𝛿 is a constant
equal to 0.0025 (Lin, 1998).

For the turbulent dissipation rate 𝜀, the model considers the expression given in
Equation 4.18.

𝑘2
𝜀 = 𝐶𝑑 𝜗 [4.18]
𝑡

Where 𝜗𝑡 = 𝜉𝜗 and ξ is a constant equal to 0.1 as given in Lin (1998).

Variation of the 𝛿 and ξ values were found to have a negligible effect on the final
results of the computations (Lin, 1998).

In the model, at solid boundaries, two types of conditions for the mean flow can be
considered as no slip and free slip. In the case of turbulent flows, the model considers
a log-law distribution for the mean tangential velocity in the turbulent boundary layer.

The application of an appropriate boundary condition at the mean free surface in


turbulent flows is quite complex as the mean free surface is not clearly defined
(Brocchini and Peregrine, 1996; Liu and Lin, 1997; Lin and Liu, 1999). In the IH-
2VOF model, the mean density fluctuations near the free surface due to mixing and air
intrusion are neglected and similarly to situations of laminar flow, the zero stress and
zero pressure conditions are imposed at the free surface.

64
The open boundary condition in the IH-2VOF model is expressed as the following
equation:

𝜕𝜙 𝜕𝜙
+ 𝑐0 𝜕𝑥 = 0 [4.19]
𝜕𝑡

Where ϕ represents the variable to be evaluated (𝑢̅, 𝑣,


̅ 𝑘, 𝜀, etc.) and 𝑐0 the wave
celerity at the considered position expressed as:

𝑐0 = √𝑔(𝑑 + 𝑎) [4.20]

For long waves, and

𝑔𝜆 2𝜋
𝑐0 = √2𝜋 tanh ( 𝜆 (𝑑 + 𝑎)) [4.21]

For short waves, where a is the wave amplitude, d is the water depth and λ is the wave
length for this depth. More details on the boundary conditions can be found in Rodi
(1980) and Liu and Lin (1997).

Wave Generation
Wave generation is a key factor for all the numerical models devoted to coastal
engineering as waves have to be generated to resemble those in the field or in the
experimental facility. Generated waves will not only cause incorrect results; they may
also cause the solution to be nonphysical.

The IH-2VOF model includes several procedures of wave generation: internal wave
maker, static wave paddle (Dirichlet boundary condition) and dynamic wave paddle
(moving boundary method). A more complete description can be found in Lara et al.
(2006) for internal wave maker, Torres et al. (2010) for static wave paddle method and
Lara et al. (2011) for dynamic wave paddle. In Dirichlet boundary condition, regular

65
waves (first order, second order, fifth order, cnoidal and solitary waves) and irregular
waves (first and second order) are implemented. In internal wave maker, wave
generation occurs in two directions; so there is a need to dissipate one of them.
Therefore, studies are combined with the sponge layer as an absorption method.
Moving boundary method replicates a piston type wave maker. It needs the paddle
position as input. Velocity is then calculated as a first order forward derivative of this
defined position.

Wave absorption boundary conditions are a key factor of IH-2VOF model. They allow
running simulations for a long time, avoiding most of the effects of reflected waves in
the flumes. Active and passive wave absorption methods are applied in the model, the
active wave absorption method can be applied to both static wave paddle and for
dynamic wave paddle. Active wave absorption is applied by following the
methodology developed by Schäffer and Klopman (2000) which is based on shallow
water linear theory. It identifies the waves that reach a boundary and then absorbs them
to prevent their reflection. Passive wave absorption is also provided in the model by a
sponge layer which is applied by following Israeli and Orszag (1981) formulation. The
method consists of a region in which a dissipation model is defined. The aim is to
prevent the reflections on the boundary, as waves generated travel in the two
directions, but it also absorbs any other incident waves from the zone of interest.
Perfect absorption is not realistic; so results up to 10% reflection considered as good.

4.3 Model Geometry and Mesh Generation

The IH-VOF model is operated by a Graphical User Interface (GUI), namely Coral,
developed to assist in generating the mesh. As preprocessing, different geometries can
be created and included within the mesh such as bottom profiles, obstacles
(breakwaters, caissons etc.) and porous media (rubble mounds, cores etc.). Once the
geometrical elements are introduced by the user, Coral creates a grid as output and this
can be read by the numerical model.

66
The Coral main window can be divided into 4 different areas as given in Figure 4.3.
- Zone 1: Graphical representation of the generated mesh and the geometry
- Zone 2: Controls for zooming, dynamic view, adjust view and view options
- Zone 3: Numerical domain and mesh resolution definition
- Zone 4: Obstacle and porous media definition
As can be seen from Figure 4.3, Zone 1 is a black window in which the numerical
domain is displayed in grey colour. Once objects are defined by the user, they appear
in different colours such as obstacles in yellow, porous bodies in red and water bodies
in cyan.

Zone 2 contains four controlling buttons as zoom allowing zoom in and zoom out, a
pan allowing a dynamic view of the mesh and an adjust view button adjusting mesh to
screen size capturing the whole domain. View options panel also displays different
view options for mesh cells depending on their properties.

Zone 3 allows the user to define the geometry on which the IH-2VOF model solves
the governing equations and provide the results. Zone 4 includes the introduction of
water and obstacles within the mesh defined in Zone 3.

Figure 4.3: Coral Interface, Zones

67
Create a new project and save

Define the dimensions of the domain


- the width of the numerical flume
- the height of the numerical flume

Define the geometry


- obstacles
- porous media
- water bodies

Mesh Generation
- define cell sizes in X and Y direction
- use Generate! button

Mesh Quality Check


- ∆x, ∆y and the domain length

Figure 4.4: Flowchart of generating a mesh in Coral

Coral generate structured meshes with both uniform and non-uniform cell size. The
use of uniform (constant cell size) meshes is recommended (IH-2VOF Course Lecture
Notes, 2012) The computing mesh is constructed from a number of submeshes
(subzones) defined at each coordinate direction, X and Y. The origin of the coordinate
system in the mesh is in the upper left hand side of the domain, which implies a positive
Y direction pointing downwards. New subzones could be created in both directions
and then modified by choosing them in the subzone window. General flowchart of
generating a mesh in Coral is given in Figure 4.4.

68
In the numerical modelling of the experiments, firstly, the domain was defined
according to the flume dimensions. Once the domain had been set, the next step was
to define the geometry to be simulated. Obstacles, porous media and water bodies were
defined using closed polygons. Each polygon was specified inserting the vertex
coordinates.

Obstacles are elements with zero porosity. Concrete elements such as caissons or
crown-walls can be defined as obstacles. The bottom bathymetry can also be defined
as impermeable using obstacles polygons.

Porous bodies are created following the same procedure as obstacles. In this case, the
porous media parameters must be introduced. These parameters are porosity, linear
friction coefficient, nonlinear friction coefficient, added mass coefficient and the
material diameter, D50. After introducing all of these parameters, the porous media
characteristics are stored.

Water bodies are introduced using the same method as the previous elements, but in
this case using the button ‘’Water’’.
To perform the model geometry definition for the caisson type breakwater, the
following geometrical features are defined (Figure 4.5).
- Two obstacles (the bathymetry and the caisson)
- Three porous media (two filter layers at the right hand side and the left hand
side, the core layer)
- Water
Figure 4.6 shows the obtained model geometry in the IH-2VOF model for the caisson
breakwater. The coordinates of the element indices which were introduced to the Coral
to obtain this geometry is provided in Appendix A.

69
Figure 4.5: Elements indices in defining geometry of caisson type breakwater (IH-
2VOF Course Lecture Notes, 2012)

Figure 4.6: Model geometry for caisson type breakwater

For the mesh generation for the model, a variable grid is chosen considering the total
number of cells allowing recreation of long domains with good discretization in the
areas of interest. The maximum resolution zone is placed around the breakwater where
a constant value of horizontal, ∆x, and vertical, ∆y, cell size is set. There are some
restrictions about setting the variable grid as the following (IH-2VOF Course Lecture
Notes, 2012):

- The spacing of adjacent cells should not differ by more than 10 – 20 %.

70
- Cell aspect ratios greater than 1 and less than 5 are more desirable. (1 < ∆x/∆y
< 5)

- Changes in the dimension of each cell have to be less than 5%. This criterion
is achieved ∆2x < 0.05 and ∆2y < 0.05.

- The horizontal discretization should be at least 70 - 100 cells per wave length
for nonbreaking waves and should be greater than 100 cells per wave length
for breaking waves.

- 7 – 10 cells per wave height are desirable in the vertical discretization.

- The domain length should be a distance equal to 1.5-2 wave length before an
obstacle or breakwater.

By following these requirements, the mesh properties defined for the mesh are given
in Table 4.1.

Table 4.1: Properties of the subzones in the mesh of caisson breakwater model
X Direction
Subzone 1 Subzone 2 Subzone 3
Center 14.980 20.480 20.520
Division 0.000 15.000 20.500
Number Cells Left 599 274 1
Number Cells Right 1 1 44
Max. Sep. Center 0.020 0.020 0.020
Y Direction
Subzone 1 Subzone 2 Subzone 3
Center 0.545 0.995 1.005
Division 0.000 0.550 1.000
Number Cells Left 49 89 1
Number Cells Right 1 1 29
Max. Sep. Center 0.005 0.005 0.005

71
3 subzones in mesh generation are defined in both X and Y directions regarding the
properties given above. The generated mesh with the subzones shown is given in
Figure 4.7.

Figure 4.7: Generated mesh with the subzones for caisson breakwater

4.4 Mesh Convergence Checks

It is important to check the quality of the generated mesh in the computational domain.
After the mesh is generated, by using the button ‘’Mesh quality’’, a new window
appears in the Coral showing two graphs (Figure 4.8). The upper graph on the window
shows the cell size as black line in X direction along the entire domain and the
derivative of that value as green line representing the variation in the cell size along
the domain. On that graph, X-axis displays the X-coordinate and the cell index
between brackets. Y-axis on the left hand side shows the cell size while Y-axis on the
right hand side shows the value of the derivative. The lower graph shows the same
aspects but for the Y-coordinate.

72
If the condition given in Section 4.3, ∆2x < 0.05 and ∆2y < 0.05, is not satisfied, the
green line on the graphs turns into red meaning that the absolute value of the derivative
curve has reached 0.05. This condition prevents variations in the cell size from being
too sharp. Therefore, if one of these green lines has turned into red, it means that it is
necessary to introduce more cells due to spacing between the adjacent cells.
In modelling of the breakwater, the convergence checks have been applied and the
following graphs (Figure 4.8) are obtained from the quality check.

Figure 4.8: Quaility check for the generated mesh

In order to ensure that the numerical model is able to simulate the regular wave
interaction with the breakwaters, the initial consideration regarding the dimensions of
the domain should also be satisfied. The domain is chosen according to the flume
dimensions given in Section 3.2. The domain length is taken as 22.3 m and the height
is taken as 1.2m. The maximum wave length of the applied waves is 9.7m. The domain
length condition is satisfied considering more than 1.2 wavelength (11.64m) in front
of the breakwaters. The reason behind choosing the domain height as 1.2m (rather than
1m which was the original dimension) is the consideration of possible overtopping
events.

73
4.5 Hydraulic Conditions for the IH-2VOF Model

In the IH-2VOF model, once the mesh generation is completed, the next step is to
establish the wave conditions. This process is carried out by IH-2VOF GUI. The main
menu of GUI is displayed in Figure 4.9.

Figure 4.9: IH-2VOF GUI Preprocessing Main Menu

There are three different possibilities in order to define the wave conditions in IH-
2VOF GUI. One of them is generating a new wave series with specified wave
characteristics such as wave height, wave period, wave series duration and sampling
frequency of the generated signal. Solitary, regular and irregular waves can be
generated by this option. Other option is to import a wave series which may be
generated before and to be used in the simulation in order to create the same conditions
obtained before. The last option is to reconstruct this wave series using a surface
record. A .dat file containing the free surface time series is needed in this option. The
file must include two columns as the time (in seconds) and the free surface elevation

74
(in meters). The wave conditions in the simulation of the two cases in this study were
defined by using the first option, creating new time series by defining the necessary
parameters. The waves were chosen as Cnoidal waves and the sampling frequency was
taken as 20Hz. The wave conditions are applied so that the free surface records at the
toe of the structures are almost the same as in the experiments. An example of defining
wave conditions is given in Figure 4.10.

Figure 4.10: Generation of new wave series

Once the wave series is established, then the wave paddle is generated in Section 3 of
the main preprocessing menu of GUI given in Figure 4.9. Static and dynamic wave
paddle options are provided. In case of dynamic paddle option, the initial position and
the maximum position of the wave paddle must be specified in order to define the
paddle stroke. In the simulations, dynamic paddle option is used since it represents a
piston type wave maker in wave generation in laboratory experiments.

Furthermore, Section 4 in the GUI menu (Figure 4.9) is to be filled in order to introduce
the simulation parameters such as the simulation time and the time step. Left and right
boundary absorption can also be selected in this section. Simulation can be carried out
considering a turbulence model or not. If the turbulence model is to be considered, the

75
parameters; turbulence seed parameter, the eddy viscosity and the parameter for
boundary layer turbulence resolution must be specified.
The properties to obtain and store in the whole domain can also be chosen in Section
4 as shown in Figure 4.10. VOF function, horizontal and vertical velocities, pressure
and turbulence fields can be saved for the entire domain at the specified sampling
frequency.

The GUI also allows locating wave gauges along the domain in order to obtain results
of free surface time history and energy spectra at those specific locations. Different
positions can be defined for these wave gauges by introducing their coordinates. Exact
locations are provided to the model as experiments.

The wave gauge locations are also defined as in the experiments. Sample sketch of
wave gauge places for caisson type breakwater is given in Figure 4.11.

Figure 4.11: Setting wave gauge position

4.6 Summary of Numerical Modelling


In summary, two cases of the physical model experiments (wave number 9 and wave
number 10) were simulated in the model, IH2VOF. The grid size in X direction (∆x)
was chosen 0.02m and in Y direction is 0.005m in the simulations. The initial time
step (dt) was defined as 0.05s in the simulations. The sampling frequency was defined
as 30Hz which corresponds to 0.033s of time step (∆t) but the ∆t values were changed

76
and adjusted around 0.03s by the model for each time step for convergence. The
simulation length was 35s which corresponds to a time duration of ten waves. The
results obtained from the simulations are discussed in Chapter 5.

77
78
CHAPTER 5

5 RESULTS AND DISCUSSION

5.1 Physical Model Experiments

In this section results of the physical model experiments are presented. First, the
analysis of the breakwater damage under the corresponding wave according to the
video recordings, profile measurements and photo documentation is presented. Then,
a summary of the results is provided in a table. Finally, a general discussion based on
these results is provided.

It is useful to give some important details about the laboratory experiments at this
point. There was no constant experiment duration in the physical model experiments
since the tests were finished after the observation of total failure of the breakwaters. In
case of no breakwater failure (wave number 12), the waves were applied as 500 waves,
1000 waves and 1000 waves cumulatively. After the experiments, profile
measurements were carried out via a laser distance meter in two lines on the
breakwaters with 1 cm intervals. The profile measurements obtained from each line on
the cross-sections for the two sets of each test are given in Appendix B. Photo
documentation was performed before and after each test as well as video recording
during the experiments. Surface profile measurements were analyzed by using a
Matlab code and the second wave in the surface profiles was taken as the determinative
wave of the characteristics of the measured waves. The surface profile of each wave
condition can be found in Appendix C. Wave reflection was eliminated by short
duration of the experiments and taking the second wave (it was the wave before the
start of reflection) as the representative wave for the pressure measurements. It may
be also useful to give the properties of the applied waves again which were already

79
given at Section 3.7. It is important to note that the characteristics given in Table 5.1
and Table 5.2 are the measured ones at breakwater toe while the flume was empty.

Table 5.1: Wave properties in model scale, caisson type breakwater

Wave Number 9 10 12 15 17 18
Period (s) 4 3.2 2 4.95 3.1 2.5
Wave Height (m) 0.22 0.11 0.09 0.17 0.15 0.13
Steepness 0.028 0.018 0.024 0.018 0.026 0.025

Table 5.2: Wave properties in model scale, rubble mound breakwater

Wave Number 9 10 12 15 17 18
Period (s) 4 3.2 2 4.95 3.1 2.5
Wave Height (m) 0.2 0.13 0.09 0.19 0.18 0.17
Steepness 0.026 0.021 0.024 0.02 0.03 0.036

The damage behavior of the structures under the applied waves are classified at three
stages as the start of damage, major damage and total failure for both breakwaters. The
descriptions of these stages are summarized as in the following:

Caisson type breakwater (Figure 5.1):


- Start of damage: Start of displacement of the vertical walls
- Major damage: Getting inclined and start of overturning of the vertical walls
- Total failure: Overturning and movement of the vertical walls together with
some damage on the slopes of the rubble foundation

Rubble mound breakwater (Figure 5.2):


- Start of damage: Start of displacement of crown walls and slope instability
- Major damage: Sliding and/or overturning of the crown walls and movement
of the stones together with the crown walls towards onshore

80
- Total failure: Movement of the crown walls towards onshore and total lose of
slope stability

a b

Figure 5.1: Stages of damage of caisson type breakwater for wave number 9 a)
start of damage b) major damage c) total failure

a b

Figure 5.2: Stages of damage of rubble mound breakwater for wave number 9
a) start of damage b) major damage c) total failure

81
Wave Number 9
Wave number 9 had a characteristic of an impact wave which directly hit on the
vertical walls of the caisson breakwater and the crown wall of the rubble mound
breakwater. The visual observation of an impact wave at the time of attack on the
structures is shown in Figure 5.3 and 5.4.

Figure 5.3: Wave Number 9 impact on the vertical wall of caisson type breakwater

82
Figure 5.4: Wave Number 9 impact on the crown walls of rubble mound breakwater

Both of the breakwaters were totally failed in the tests. The caisson type breakwater
started to be damaged at the 1st wave, had major damage by the 2nd wave and totally
failed by the 3rd wave in the tests.
The rubble mound breakwater started to get damage at the 2nd wave, had major damage
at the 5th wave and totally failed by the 7th wave, in the tests.
Sometimes, some difference between the two sets of the same experiment occurred
because the vertical walls and the crown walls got stuck in the side walls of the flume
in some of the tests. This fact was clearly observed from video recordings and was
taken into account in the analysis of the damage behavior.
The breakwater profiles of each set test for caisson type breakwater for Wave number
9 are given in Figure 5.5 and 5.6. The images of damaged breakwater are also given
in Figure 5.7.

83
0,5
Before After
0,45 Line-1
0,4
Y - Vertical Distance (m)

0,35

0,3

0,25

0,2

0,15

0,1

0,05

0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8
X -Horizontal Distance (m)

0,5

0,45 Line-2 Before After

0,4
Y - Vertical Distance (m)

0,35

0,3

0,25

0,2

0,15

0,1

0,05

0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8
X - Horizontal Distance (m)

Figure 5.5: Breakwater profiles for caisson type breakwater, Wave number 9 Set 1
(wave approach is from left)

The difference between the two lines of the profiles which are given in Figure 5.3 is
due to the fact that the three vertical wall elements were overturned and one element
was placed after another after the experiment (Figure 5.5).

84
0,5
Line-1 Before After
0,45

0,4
Y - Vertical Distance (m)

0,35

0,3

0,25

0,2

0,15

0,1

0,05

0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8
X - Horizontal Distance (m)

0,5

0,45 Line-2 Before After

0,4
Y - Vertical Distance (m)

0,35

0,3

0,25

0,2

0,15

0,1

0,05

0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8
X - Horizontal Distance (m)

Figure 5.6: Breakwater profiles for caisson type breakwater, Wave number 9 Set 2

In the second set, the vertical wall elements were overturned and washed away which
is clearly observed from profile measurements given in Figure 5.4. The elements were
placed side-by-side after the experiment which can be seen in Figure 5.5. Therefore,

85
there is almost no difference between the two lines of profile measurements of the
breakwater.

Figure 5.7: Damage of caisson type breakwater (Top: Set 1, Bottom: Set 2)

The breakwater profiles of each set test Wave number 9 for rubble mound breakwater
are given in Figure 5.8 and 5.9. The images of damaged breakwater are also given in
Figure 5.10.

86
0,6
Before After
Line-1
0,5
Y - Vertical Distance (m)

0,4

0,3

0,2

0,1

0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8
X - Horizontal Distance (m)

0,6

Line-2 Before After


0,5
Y - Vertical Distance (m)

0,4

0,3

0,2

0,1

0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6
X - Horizontal Distance (m)

Figure 5.8: Breakwater profiles for rubble mound breakwater, Wave number 9 Set 1

87
0,6

Line-1 Before After


0,5
Y - Vertical Distance (m)

0,4

0,3

0,2

0,1

0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8
X - Horizontal Distance (m)

0,6

Line-2 Before After


0,5
Y- Vertical Distance (m)

0,4

0,3

0,2

0,1

0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6
X - Horizontal Distance (m)

Figure 5.9: Breakwater profiles for rubble mound breakwater, Wave number 9 Set 2

88
Figure 5.10: Damage of rubble mound breakwater after wave number 9, Set 1

Wave Number 10
Wave number 10 had an overflow effect on the vertical walls of the caisson breakwater
and the crown walls of the rubble mound breakwater rather than hitting on these
structures. Therefore, Wave number 10 had a characteristic of overflowing wave
differing from Wave number 9. The visual observation of an overflow wave at the time
of acting on the structures is shown in Figure 5.11 and 5.12.

89
Figure 5.11: Overflowing wave acting on the vertical walls of caisson type
breakwater

Figure 5.12: Overflowing wave acting on the vertical walls of rubble mound
breakwater

Both of the breakwaters were totally failed in the two sets of Wave number 10
experiments. The caisson type breakwater started to be damaged at 7th wave, got major
damage at the 11th wave and totally failed by the 13th wave.

90
The rubble mound breakwater started to get damage at the 3rd wave, got major damage
at the 8th wave and totally failed by the 11th wave. The breakwater profiles for caisson
type breakwater are given in Figure 5.13. The images of damaged breakwater are also
given in Figure 5.14.

0,5
0,45 Before After
Line-1
0,4
Y - Vertical Distance

0,35
0,3
0,25
0,2
0,15
0,1
0,05
0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8
X - Horizontal Distance

0,5

0,45 Line-2 Before After

0,4
Y - Vertical Distance (m)

0,35

0,3

0,25

0,2

0,15

0,1

0,05

0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8
X - Horizontal Distance (m)

Figure 5.13: Breakwater profiles for caisson type breakwater, Wave number 10 Set
2

Although the caisson type breakwater is failed after wave number 10, it is not observed
in the two lines of the profile measurements (Figure 5.13) since the right and left

91
vertical wall elements were stuck between the side walls of the wave flume and the
middle wall element (Figure 5.14).

Figure 5.14: Damage of caisson type breakwater after wave number 10 Set 2

The breakwater profiles of the first set for rubble mound breakwater are given in Figure
5.15. The images of damaged breakwater are also given in Figure 5.16.

0,6

0,5 Before After


Line-1
Y - Vertical Distance (m)

0,4

0,3

0,2

0,1

0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8
X - Horizontal Distance (m)

92
0,6

0,5 Before After


Line-2
Y - Vertical Distance (m)

0,4

0,3

0,2

0,1

0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8
X - Horizontal Distance (m)

Figure 5.15: Breakwater profiles for rubble mound breakwater, Wave number 10 Set
2

Figure 5.16: Damage of rubble mound breakwater after wave number 10, Set 2

Wave Number 12
Wave number 12 was applied in the tests as three sets of 500 waves, 1000 waves and
1000 waves again since the two of the breakwaters did not fail at the end of the sets.
Wave number 12 had a different characteristic behaving like an impact wave but on
the rubble mound breakwater but like an overflowing wave on the caisson type
breakwater. On the rubble mound breakwater, it hit on the stones of the armor layer

93
rather than the crown wall of rubble mound breakwater. Both of the structures had
minor damage after three sets which were applied cumulatively. Breakwater profiles
measured before and after each set and photos are given in Appendix C.

Wave Number 15
The wave number 15 had a characteristic of an impact wave hitting on the vertical
walls of caisson type breakwater and the crown walls of the rubble mound breakwater.
In the tests, both of the breakwaters were totally failed. The caisson type breakwater
started to be damaged at the first wave, had major damage at the 3rd wave and totally
failed by the 4th wave.
The rubble mound breakwater started to get damage at the second wave, had major
damage at the 6th wave and totally failed by the 11th wave. The breakwater profiles and
the images of damaged breakwaters are given in Appendix C.

Wave Number 17
The wave number 17 had a characteristic of an overflowing wave. In the tests, both of
the breakwaters were totally failed. The caisson type breakwater started to be damaged
at the 3rd wave, had major damage at the 10th wave and totally failed by the 11th wave.
The rubble mound breakwater started to get damage at the 4th wave, had major damage
at the 7th wave and totally failed by the 10th wave. The breakwater profiles and the
images of damaged breakwaters are given in Appendix C.

Wave Number 18
The wave number 18 had a characteristic of an impact wave hitting on the vertical
walls of caisson type breakwater but hitting on the stones of the armor layer of rubble
mound breakwater. In the tests, caisson type breakwater failed under the impact of the
wave but rubble mound breakwater did not fail.
Caisson type breakwater started to be damaged at the 4th wave, had major damage at
the 6th wave and totally failed by the 10th wave.

94
The rubble mound breakwater had some damage after many waves but remained
without failure after more than 100 waves. The breakwater profiles and the images of
damaged breakwaters are given in Appendix C.

Key findings on damage analysis


The relevant failure mode of caisson type breakwater in the cases which the breakwater
was failed under wave action was overturning of the vertical walls and movement of
the walls towards onshore. An example of this failure mode at the instant of
overturning of the vertical walls is given in Figure 5.17. The slopes of the rubble base
were stable in most of the experiments having minor damage unless the the vertical
walls move and wash away the stones.

Figure 5.17: Overturning of vertical walls under wave action

The major failure modes in rubble mound breakwater were overturning and sliding of
the crown walls. In case of impact waves, the crown walls were overturned and moved
towards the harbor slope leading the slope instability of the armor layers and total
failure of the structure. In case of overflowing waves, the crown walls slided
subsequently washing away the stones of the armor layer towards onshore. Then, the
slope stability was lost and the structure was totally failed. An example is given in
Figure 5.18.

95
Figure 5.18: Sliding of the crown walls of the rubble mound breakwater under wave
action

The seaside and harbor slopes of the rubble mound breakwater remained almost stable
unless the crown walls fail. However, the scour on the harbor slope which occurred
due to wave overtopping leaded the failure of the crown walls. In case of crown wall
failure, the upper part of the seaside armor layer stones washed away towards the
harbor slope together with the crown walls. Therefore, it can be concluded that the
crown wall stability plays a key role on slope stability and hence on the overall stability
of the breakwater but the armour layer at the harbour side of the breakwater is also
important for crown wall stability. Crown wall failure and slope instability are
interactive processes which one may cause the other subsequently.

A summary of detailed breakwater damage under related wave actions is provided in


Table 5.3.

96
Table 5.3: Overview of detailed breakwater damage under related wave actions

Breakwater Damage
Test No Structure Crown walls–
Seaside Slope Harbor slope Damage
Vertical walls
armor, filter and core
Crown walls slided,
Rubble layer removed over armour layer washed Total
washed away
mound 1/3 of slope length, away failure
towards harbor slope
layers mixed
10-1 Three elements of
Few stones in filter Some stones moved
concrete blocks were
Caisson layer moved towards towards onshore Total
overturned and
Type onshore together with together with failure
moved towards
concrete blocks concrete blocks
onshore
armor, filter and core
Crown walls slided,
Rubble layer removed over armour layer washed Total
washed away
mound 1/3 of slope length, away failure
towards harbor slope
layers mixed
10-2 Middle element first
Few stones in filter Some stones moved
overturned and
Caisson layer moved towards towards onshore Total
prevented other two
Type onshore together with together with failure
concrete blocks from
concrete blocks concrete blocks
overturning
Few stones of armor Few stones in armor
Rubble Minor
layer moved within layer moved within Not moved
mound damage
seaside slope slope downwards
12-1 Few stones in first Few stones in first
Caisson Minor
layer moved within layer moved within Not moved
Type damage
slope slope
Few stones of armor Few stones in armor
Rubble Slightly moved in Minor
layer moved within layer moved within
mound flow direction damage
12-2 seaside slope slope downwards
Few stones in first Few stones in first
Caisson Minor
layer moved within layer moved within Not moved
Type damage
slope slope
Few stones of armor Few stones in armor
Rubble Slightly moved in Minor
layer moved within layer moved within
mound flow direction damage
seaside slope slope downwards
12-3
Few stones in first Few stones in first
Caisson Minor
layer moved within layer moved within Not moved
Type damage
slope slope
Armor layer washed
Armor layer and
away together with
Rubble filter layer stones Crown walls slided Total
stones of seaside
mound moved towarsd and washed away failure
slope and crown
harbor slope
17-1 walls
Three concrete
Few stones in first Few stones in first
Caisson elements overturned Total
layer moved within layer moved within
Type and moved in flow failure
slope slope
direction
Armor layer washed
Rubble Armor layer and Crown walls slided Total
17-2 away together with
mound filter layer stones and washed away failure
stones of seaside

97
Breakwater Damage
Test No Structure Crown walls–
Seaside Slope Harbor slope Damage
Vertical walls
moved towarsd slope and crown
harbor slope walls
Middle and left
concrete elements
Few stones in first Few stones in first
Caisson overturned and Total
layer moved within layer moved within
Type prevented right failure
slope slope
element from
overturning
Some stones moved Many stones in
Middle element
upwards with the armor layer moved
Rubble slided other elements Major
slope and within slope and
mound slightly moved in damage
accumulated in front accumulated at
flow direction
of the crown wall harbor side toe
18-1 Middle and right
concrete blocks
Few stones in first
Caisson Few stones moved overturned and Total
layer moved within
Type within slope moved in flow failure
slope
direction, left
element was stuck
Some stones moved
upwards with the Many stones in
Crown walls slightly
slope and armor layer moved
Rubble moved in flow Major
accumulated in front within slope and
mound direction and slightly Damage
of the crown wall / 1 accumulated at
overturned
stone moved to habor harbor side toe
18-2 slope
Middle and right
concrete blocks
Few stones in first
Caisson Few stones moved overturned and Total
layer moved within
Type within slope moved in flow Failure
slope
direction, left
element was stuck
Armor layer washed
armor, filter and core Crown walls
away together with
Rubble layer removed over overturned and Total
stones of seaside
mound 1/3 of slope length, moved towards Failure
slope and crown
layers mixed harbor slope
9-1 walls
Some stones in first All three elements
Few stones in first
Caisson layer moved towards were overturned and Total
layer moved within
Type onshore by concrete moved towards Failure
slope
blocks onshore
Armor layer washed
armor, filter and core Crown walls
away together with
Rubble layer removed over overturned and Total
stones of seaside
mound 1/3 of slope length, moved towards Failure
slope and crown
9-2 layers mixed harbor slope
walls
Few stones in first Few stones in first
Caisson All three elements Total
layer moved within layer moved towards
Type were overturned Failure
slope onshore

98
Breakwater Damage
Test No Structure Crown walls–
Seaside Slope Harbor slope Damage
Vertical walls
Armor layer washed
armor, filter and core
away together with Crown walls
Rubble layer removed over Total
stones of seaside overturned and
mound 1/3 of slope length, Failure
slope and crown washed away
layers mixed
walls / no slope left
15-1
Many stones in first
Few stones in first
layer moved towards Concrete blocks
Caisson layer moved within Total
onshore together were overturned and
Type slope towards harbor Failure
with moving washed away
slope
concrete blocks
Some stones moved Crown walls
Armor layer washed
upwards with the intended to overturn
away together with
Rubble slope and and wash away but Total
stones of seaside
mound accumulated in front were stuck between Failure
slope and crown
of the crown wall / side walls of the
15-2 walls / no slope left
slope unstable flume
Few stones in first
Few stones in first
Caisson layer moved within All three elements Total
layer moved towards
Type slope towards harbor were overturned Failure
onshore
slope

General Discussion
The damage behavior and failure of the two breakwaters are evaluated based on the
results presented above. Applied waves are classified according to five parameters for
an overall evaluation:
- Wave steepness (H/L)
- Ratio of wave height to water depth (H/d)
- Wave characteristics (impact wave or overflowing wave)
- Wave asymmetry and skewness
- Overtopping height

The properties of the applied waves in the experiments based on these parameters are
given in Table 5.4.

99
Table 5.4: Observed properties of applied waves

Steepness
Wave No Fail (X) (H/L) H/d A S Wave Characteristic Overtop
9 7 0.0258 0.98 6 6 Impact to crown walls 4
10 12 0.0214 0.67 2 5 Overflow 5
Rubble 12 999 0.0238 0.47 3 2 Impact to armor stones 1 Relative
Mound 15 13 0.0197 0.96 1 1 Impact to crown walls 2 Evaluation
17 10 0.0295 0.88 4 4 Overflow 6
18 999 0.0363 0.86 5 3 Impact to armor stones 3

100
9 3 0.0281 0.73 6 6 Impact 14
10 13 0.0178 0.37 2 5 Overflow 4
12 999 0.0244 0.30 3 2 Overflow 1 Measurem
Caisson
15 4 0.0175 0.57 1 1 Impact 15 ent (cm)
17 11 0.0257 0.50 4 4 Overflow 12
18 10 0.0264 0.42 5 3 Impact 13

A: Asymmetry
S: Skewness
Fail (X): Structure failure at the time of Xth coming wave
999: No failure
Based on the parameters given above and observed properties of the applied waves the
following discussions are carried out:

1) Caisson type vs. rubble mound breakwater comparison:

Two breakwaters were similar in cases of wave no 10, 12 and 17 whereas different in
cases of Wave no 9, 15 and 18 in terms of damage behavior (fail or not fail) and failure
times (at which wave the structure failed). In cases of wave 9, 15 and 18 which were
all impact waves, caisson type breakwater failed long before the rubble mound
breakwater. In cases of impact waves, the armor stones of the rubble mound
breakwater absorbed some of the wave energy but the vertical walls in caisson type
breakwater were directly exposed to the impact. Therefore, the caisson type
breakwater was much more unstable in these cases. Cases of wave no 10, 12 and 17
were overflowing waves and the breakwaters failed at similar times in these cases.

2) Rubble mound breakwater damage behavior

When the wave impact was to the armor stones rather than to the crown walls (cases
wave number 12 and 18), the rubble mound breakwater did not fail.
When wave impact was to the crown walls (cases wave number 9 and 15), the
breakwater failed at 7th and 13th waves respectively. This time difference is caused by
steepness and wave shape parameters since H/d and overtopping height are close to
each other in these two cases. It can be deduced that as the wave steepness increases
and wave shape gets non-sinusiodal, the breakwater gets more damage and fails earlier
in case of impact waves.
When wave characteristic was overflow (cases wave number 10 and 17), the
breakwater fails at 12th and 10th waves respectively. The difference arises from H/d
ratio.

101
3) Caisson type breakwater damage behavior

When wave impact occurred (in cases of wave number 9, 15 and 18), the breakwater
failed at 3rd, 4th, and 10th waves respectively. Failure time was similar in cases of wave
number 9 and 15 although the steepness, H/d ratios and wave shape parameters were
different in these two cases (Table 5.4). This is most probably due to the fact that the
wave impacted directly to the blocks and overflow heights were high and similar in
these cases. What makes the case of wave number 18 different from these two cases is
the H/d ratio which was less than the other two cases.
When overflow occurred (in cases of wave number 10, 12 and 17), the breakwater
failed at 13th wave, did not fail and failed at 11th waves respectively. The structure did
not fail in case of wave number 12 because the overtopping height and H/d ratio were
much less than the other two cases whereas the other parameters were similar and
although the steepness of wave number 12 was higher.
It can be concluded that the wave height and hence the overtopping height is a very
important parameter in caisson type breakwaters in both impact and overflowing
waves.

It should be noted that these overall discussios on the physical phenomenon of the
structures are carried out according to the measurements as well as the visual
observations during the experiments and from video recordings.

5.2 Numerical Modelling Results and Pressure Measurement


Comparison

Pressure measurements along the front and bottom surfaces of the vertical wall in
caisson type breakwater were carried out for two wave conditions, wave number 9 and
10. One impact wave and one overflowing wave characteristics were selected from the
applied waves. Pressure measurement experiments were repeated two times for each
case. The measurements correspond to pressure acting on the wall at the time of impact

102
of second wave on the wall. In other words, measured pressures are related to only the
second wave of the wave train.

On the other hand, the caisson type breakwater was modeled in IH2VOF and pressure
results were obtained from the numerical model. The initial time step (dt) was defined
as 0.05s in the simulations. The sampling frequency was defined as 30Hz which
corresponds to 0.033s of time step (∆t) but the ∆t values were changed and adjusted
around 0.03s by the model for each time step for convergence. The simulation length
was 35s.

The two wave conditions were applied in the computational tool, IH-2VOF, by using
the dynamic paddle option which represents a piston type wave maker in wave
generation in laboratory experiments. After the simulations were carried out, the
surface water profiles obtained from IH-2VOF and from the laboratory measurements
were overlapped for offshore (WG8) and overtopping gauges (WG2). The offshore
gauge is the wave gauge which was placed in front of the wave paddle and the
overtopping gauge is the one placed onto the front surface of the vertical walls (see
Figure 3.18 and 3.20). The results are given in Figure 5.19 and 5.20 for wave number
9 and in Figure 5.21 and 5.22 for wave number 10. These wave data are compared to
validate that the proper wave conditions were generated in the IH-2VOF model. After
the validation of the wave data, the obtained pressure data from IH-2VOF and from
the measurements were analysed.

103
Figure 5.19: Water surface profiles for wave number 9, WG8

Figure 5.20: Water surface profiles for wave number 9, WG2

104
Figure 5.21: Water surface profiles for wave number 10, WG8

Figure 5.22: Water surface profiles for wave number 10, WG2

105
Kortenhaus et al. (1999) presented a paremeter map including geometric and wave
parameters in combination to identify the wave impact loading which is given in
Figure 5.23. According to this parameter map, our structure stays on the ‘’composite
breakwater’’ classification and ‘’low-mound breakwater’’ side according to the ‘’hb*’’
definition which is 0.4 in our case. In this classification, our case wave number 9 is in
impact loads category and wave number 10 is in quasi-standing loading. This
classification is consistent with the pressure graphs obtained from analysis of our
measured data which are given in Figure 5.24 and 5.26.

Figure 5.23: Wave loading identification, PROVERBS parameter map (Kortenhaus


et al., 1999)

106
Figure 5.24: Example graph of measured pressure time series for wave number 9

The dynamic and quasi-static parts of pressure time seri of wave number 9 is specified
according to the idealized time history of an impact pressure on vertical walls which
is given in Bullock et al. (2007). The definitional sketch of is given in Figure 5.25.

Figure 5.25: Definition sketch of an idealized impact wave pressure time history

107
Figure 5.26: Example graph of measured pressure time series for wave number 10

In order to obtain the maximum dynamic pressures along the front and bottom surfaces
of the vertical walls of caisson type breakwater, the maximum pressure values of
second incident waves are obtained from the measurements. The same procedure is
also applied to the IH-2VOF pressure data.

Figure 5.27 presents the analysis results of pressure measurement data compared with
IH-2VOF simulation data and two prediction methods of Goda (1974) and Shore
Protection Manual (1984) based on Minikin (1963). In this graph, pointed data
represent the measured local maximum dynamic pressure values for wave number 9
which is classified as impact loading. Goda’s (1974) method is one of the most well-
accepted pressure formulas for vertical and composite walls but the pressure
measurements are slightly different from Goda’s distribution since it does not give
accurate estimates of wave pressures for impact wave loading which is our case in
wave number 9. Minikin’s (1963) method is one of the prediction methods which
account for impact waves however it has been cited erroneously in many publications
(Allsop et al., 1996). The Shore Protection Manual (1984) version of Minikin’s
approach is included in the graph (Figure 5.27) but it gives substantially greater

108
pressures as also quoted so in literature about prediction of impact forces (Allsop et
al., 1996).

Wave No 9
Horizontal Pressure
0,0 1,0 2,0 3,0 4,0 5,0 6,0 7,0 8,0 9,0
0

-5
Y - Vertical Distance (cm)

-10

IH2VOF
-15
Goda (1974)

Measurement
-20
Measurement-Trend

SPM (1984) - Minikin (1963)


-25
X - Maximum Dynamic Pressure (kPa)

Figure 5.27: Local maximum dynamic pressures (point data) and instantaneous
pressure distributions for wave number 9

Figure 5.28 presents the dynamic pressure distributions obtained from IH-2VOF and
Goda’s (1974) prediction method with the quasi-static part of impact wave pressures
obtained from measurements for wave number 9. As can be seen from the graph, IH-
2VOF gives very close results with the measurements for the upper part of still water
level and gives similar results for the lower part. It is obvious that IH-2VOF is far from
being able to give compatible results with measurements in case of impact wave
pressures (Figure 5.28). However, the pressure distribution obtained fron IH-2VOF is
similar with pressure measurement data especially for the upper part of still water level
in case of quasi-static pressure.

109
Wave No 9
Horizontal Pressure
0,0 0,2 0,4 0,6 0,8 1,0 1,2 1,4 1,6 1,8 2,0 2,2 2,4 2,6 2,8 3,0
0

-2,5

-5
Y - Vertical Distance (cm)

-7,5

-10

-12,5

-15 IH2VOF
-17,5 Goda (1974)

-20 Measurement

-22,5 Measurement-
Trend
-25
X - Maximum Dynamic Pressure (kPa)

Figure 5.28: Local maximum quasi-static pressures (point data) and instantaneous
pressure distributions for wave number 9

Figure 5.29 presents the uplift pressures obtained from the measurements, IH-2VOF
simulation and Goda’s (1974) prediction method. The measurement data and Goda’s
formula seem consistent but IH-2VOF simulation results depart from these.

110
X - Horizontal Distance Along Bottom Surface (m)
0 0,05 0,1 0,15 0,2 0,25
0

0,5
Maximum Dynamic Pressure (kPa)

1,5
Measurement
2
Measurement-
Trend
2,5 IH2VOF

Goda (1974)
3

Figure 5.29: Local maximum uplift pressures (point data) and instantaneous uplift
pressure distributions for wave number 9

Figure 5.30 presents the analysis results of pressure measurement data compared with
IH-2VOF simulation data and Goda’s (1974) prediction method. In the graph, full
point data represent the pressure measurements for wave number 10 which are taken
as the maximum values at the second incident wave. Wave number 10 is classified as
quasi-static loading and the results are almost consistent with Goda’s (1974) prediction
method which is more applicable for pressure estimation of quasi-standing waves. In
this case, there are some apocryphal data which are represented by empty points and
the trend line (pressure distribution) of the measured data is drawn without taking into
consideration of these data. These suspicious data on pressure measurements occurred
when the pressure transducers were placed 7cm above the still water level and, 3.5cm,
6.5cm and 9.5cm below the still water level. The further investigation of the suspicios
pressure measurement data shows that there may be some physical phenomenon for
this case which can not be explained with the measurements in this study. The data
show the same inconsistency when the pressure sensors were placed at the same levels
in both of the two repetition of the experiments. Therefore, this situation is most

111
probably related to a physical phenomenon occurring during the wave number 10
attack rather than an error of the transducers. One should note that it may be a further
study and discussion point.

Wave No 10
Horizontal Pressure
0 0,5 1 1,5 2 2,5
0

-5
Y - Vertical Distance (cm)

-10

-15

IH2VOF
-20 Measurement
Measurement-Trend
Goda (1974)
-25
X - Maximum Dynamic Pressure (kPa)

Figure 5.30: Local maximum dynamic pressures (point data) and instantaneous
pressure distributions for wave number 10

Figure 5.31 presents the uplift pressures obtained from the measurements, IH-2VOF
simulation and Goda’s (1974) prediction method for wave number 10. The
measurement data and Goda’s formula seem consistent but IH-2VOF simulation
results again depart from these in terms of uplift pressure distribution in this case.

112
X - Horizontal Distance Along Bottom Surface (m)
0 0,05 0,1 0,15 0,2
0
Maximum Dynamic Pressure (kPa)

0,5

1,5 Measurement

Measurement-Trend

2 IH2VOF

Goda (1974)
2,5

Figure 5.31: Local maximum uplift pressures (point data) and instantaneous uplift
pressure distributions for wave number 10

113
114
CHAPTER 6

6 CONCLUSION AND FUTURE RECOMMENDATIONS

In this study, laboratory experiments on the performance of the rubble mound and
caisson type (vertical wall) breakwaters under extreme waves with the focus on
induced breakwater damage and failure phenomenon are performed. In addition, wave
pressure measurements acting on the vertical wall breakwater were carried out in the
experiments since damage and failure phenomenon are directly related with the forces
and pressures acting on the structures under wave action. Furthermore, the vertical
wall breakwater was modeled by a computational tool, IH-2VOF, to simulate two
cases of the experiments and obtain the pressure data from the software. Finally, the
pressure measurements are compared with the numerical model results and prediction
methods to explain the meaning of the results and to make an overall conclusion.

The conclusions of this study are summarized as the following:

- Physical model experiments are essential for investigation of wave induced


damage on coastal structures since it is not always feasible to observe the
structures in the field. Also, field observations are carried out mostly after the
damage occurs. However, like in the case of rubble mound breakwaters which
are observed in this study, one failure mode triggers another failure mode such
as sliding of the crown walls leads slope instability but also the scour in the
breakwater slope leads the crown wall movement. Therefore, the necessity of
the observation of damage and failure moments arise to investigate failure
mechanisms of the structures in detail.

115
- It is important to observe the wave behavior while acting on the structure
visually since it is not always possible to explain it by formulas and also it
provides a great perspective about the causing factor of the damage. In the
experiments, it is found out that two types of wave behavior which are
classified as impact wave and overflowing waves differ while acting on the
structure and this difference results in different damage behavior and different
times of failure of the structures.

- It is a complex process to evaluate the breakwater damage under different


waves since several wave parameters such as wave height, wave steepness,
overtopping height, wave asymmetry and skewness and wave behavior while
acting on the structure, should be taken into consideration. In general, it can be
concluded that in case of impact waves, the rubble mound breakwater was
more resilient since caisson type breakwater failed long before the rubble
mound breakwater in these cases. Also, wave height and hence the overtopping
height is a very important parameter in caisson type breakwaters in both impact
and overflowing waves.

- Profile measurements which are conducted before and after experiments to


record the breakwater damage need to be taken along more than two lines over
the cross-section in small-scale experiments. Because, as experienced in the
physical model experiments, two-lines profile measurement may not always
represent the overall breakwater damage like in case wave number 10. (See
Figure 5.11 and Figure 5.12). Therefore, as a further study, image processing,
which is a method of 3D modelling of an object in digital form, to obtain
breakwater damage will be applied to the physical model experiments and the
damage will also be analyzed by this method.

- Numerical model, IH-2VOF, give low pressure values for maximum dynamic
pressure on the front surface of the vertical wall breakwater compared to the
measurements and prediction methods for both two different wave conditions.

116
However, the model results depart from the measurements much more in case
of impact pressures. This is probably due to the fact that the model is not able
to resolve the physical phenomenon occurring at the insant of maximum impact
pressure acting on the structures. The mesh grid size is a critical factor on this
issue and therefore it may be a future study to increase the mesh fineness to
obtain more accurate results from the model.

- In order to increase the accuracy of numerical modelling part of this study,


some future studies need to be carried out for the sensitivity analysis of the
numerical model by simulations with different mesh size and time step values
especially in case of simulating impact wave loading conditions.

- For quasi-static wave condition, the vertical distribution of measured dynamic


pressures on the front surface of the caisson type breakwater conformed with
the simple distribution suggested by Goda but differed dramatically in case of
impact loading.

- Impact wave conditions result in very high pressures near to the still water
level, and these pressures can be very severe for vertical wall breakwaters.
Therefore, these conditions should definitely be taken into consideration in the
design of vertical wall breakwaters.

- Two types of breakwaters, rubble mound and caisson type breakwaters, were
designed according to wind waves but tested under extreme waves. Both of the
structures failed under most of the applied waves. Therefore, possible extreme
wave conditions should be considered in the design of breakwaters in order to
avoid any loss which may be caused by the possible extreme events.

117
118
7 REFERENCES

Allsop N.W.H., McKenna J.E., Vicinanza D. and Whittaker T.T.J (1996), "New
Design Methods for Wave Impact Loading on Vertical Breakwaters and Seawalls.",
Coastal Engineering Proceedings 1.25.

Arikawa, T. and Shimosako, K. (2013), “Failure mechanism of breakwaters due to


tsunami; a consideration to the resiliency.”, 6th Civil Engineering Conference in the
Asian Region, Jakarta, Indonesia

Baykal (2009), “Irregular Wave Generation and Analysis”, MATLAB Codes

Brocchini, M., and D. H. Peregrine (1996), "Integral flow properties of the swash zone
and averaging.", Journal of Fluid Mechanics 317, 241-273.

Bryant E.A. (2001), ‘’Tsunami: The Underrated Hazard’’, Cambridge University


Press, London, UK, p. 320.

Bullock G.N., Obhrai C., Peregrine D.H., Bredmose H., (2007), ‘’Violent breaking
wave impacts Part1: results from large-scale regular wave tests on vertical and
sloping walls.’’, CoastalEng.54(8),602–617.

Camfield F. (1980), ‘’Tsunami Engineering’’, Coastal Engineering Research Center,


US Army Corps of Engineers, p. 222, Special Report (SR-6).

Champagne, F. H., V. G. Harris, and Sk Corrsin (1970), "Experiments on nearly


homogeneous turbulent shear flow.", Journal of Fluid Mechanics 41.01, 81-139.

C. C. H. (2000), "City and County of Honolulu Building Code.", Department of


Planning and Permitting of Honolulu Hawaii, Chapter 16, Article 11.
CERC (1984), “Shore Protection Manual”, USACE, Vicksburg, Mississippi, USA

119
Clarke D.K., Hassan H.A., and M.D. Salas. (1986), ‘’Euler calculations for
multielement airfoils using cartesian grids.’’, AIAA Journal, 24, 353–358.

Dames and Moore (1980), ‘’in Design and Construction Standards for Residential
Construction in Tsunami-Prone Areas in Hawaii’’, Prepared for the Federal
Emergency Management Agency.

Dai Y.B. and Kamel A.M. (1969), ‘’Scale effect tests for rubble-mound breakwaters.’’,
U.S. Army Engineer Waterways Experiment Station, Corps of engineers, Vicksburg,
Mississippi.

Dean, R.G. and Dalrymple R.A. (2000), ‘’Coastal Processes with Engineering
Applications.’’, Cambridge University Press.

Department of Planning and Permitting of Honolulu Hawai (2000), Chapter 16, ‘’City
and County of Honolulu Building Code’’, Article 11.

Earthquake Engineering Research Institute (2011), ‘’Learning from Earthquakes the


Tohoku, Japan, Tsunami of March 11, 2011 Effects on Structures’’, EERI Special
Earthquake Report

Esteban M., Jayaratne R., Mikami T., Shibayama T., Mizuno Y., Kinoshita M.,
Matsuba S. (2013), ‘’Analysis of the Stability of Armour Units during the 2004 Indian
Ocean and 2011 Tohoku Tsunami.’’, Journal of Waterway, Port, Coastal, and Ocean
Engineering.

Federal Emergency Management Agency (Jessup, MD, 2003), ‘’Coastal Construction


Manual (3 Vols.)’’, 3rd edn. (FEMA 55).

120
Fraser S., Raby A., Pomonis A., Goda K., Chian S.C., Macabuag J., Offord M., Saito
K., Sammonds P. (2013), ‘’Tsunami damage to coastal defences and buildings in the
March 11th 2011 Mw 9.0 Great East Japan earthquake and tsunami.’’, Bulletin of
Earthquake Engineering 11.1, 205-239.

Frostick, L. E., McLelland S.J., and Mercer T.G. (2011), ‘’Users guide to physical
modelling and experimentation: Experience of the HYDRALAB network.’’, CRC
Press.

Garcia N., Lara J.L. and Losada I.J. (2004), ‘’2-D Numerical analysis of near-field
flow at low-crested permeable breakwaters.’’, Coastal Engineering, ELSEVIER, Vol.
51, no. 10, 991-1020.

Goda, Y. (1974), "New wave pressure formulae for composite breakwaters.", Coastal
Engineering Proceedings, 1.14.

Goda, Y. (2010), ‘’Random seas and design of maritime structures.’’, World


Scientific.

Guanche R., Losada I.J., and Lara J.L. (2009), "Numerical analysis of wave loads for
coastal structure stability.", Coastal Engineering 56.5, 543-558.

Hedges, T. S. (1995), "Regions of validity of analytical wave theories.", Proceedings


of the ICE-Water Maritime and Energy, 112.2, 111-114.

Hiroi, I. (1920), "The force and power of waves." The Engineer 130, 184-185.

Holthuijsen L. H. (2007), ‘’Waves in oceanic and coastal waters.’’, Cambridge


University Press.

121
Hsu, T-J., Tsutomu S., and Philip L-F. (2002), "A numerical model for wave motions
and turbulence flows in front of a composite breakwater.", Coastal Engineering 46.1,
25-50.

Hudson R.Y., Herrmann F.A., Sager R.A. (1979), ‘’Coastal hydraulics models.’’,
Special Report No. 5, Fort Belvoir, USA.

Hughes, S. A. (1993), ‘’Physical models and laboratory techniques in coastal


engineering’’, Vol. 7, World Scientific.

IH-2VOF Course Lecture Notes (2012), IH-Cantabria, Cantabria, Spain.

IH-2VOF (2016), http://ih2vof.ihcantabria.com/physics (last accessed February 25,


2016)

Israeli M. and Orszag S.A. (1981), "Approximation of radiation boundary


conditions.", Journal of Computational Physics, 41.1, 115-135.

Japan Weather Association materials, http://www.japanecho.net/311-data/1205/ (last


accessed February 1, 2016)

Jaw S.Y. and Chen C.J. (1998), "Present status of second-order closure turbulence
models. I: overview.", Journal of Engineering Mechanics 124.5: 485-501.

Jayaratne, R., et al. (2013), ‘’Investigation of coastal structure failure due to the 2011
Great Eastern Japan Earthquake Tsunami.’’, Coasts, Marine Structures and
Breakwaters 2013.

Kato, F., Suwa, Y., Watanabe, K. and Hatogai, S. (2012), ‘’Mechanism of Coastal
Dike Failure Induced by the Great East Japan Earthquake Tsunami.’’, Proc. of 33nd
Int. Conf. on Coastal Engineering Santander, Spain.

122
Keulegan, G. H. (1973), ‘’Wave transmission through rock structures.’’, US Army
Corps of Engineers, WES.

Kirkoz M.S. (1983), ‘’Breaking and run-up of long waves, tsunamis: their science and
engineering.’’, 10th IUGG Int. Tsunami Symposium, Sendai-shi/Miyagi-ken, Japan,
Terra Scientific Publishing, Tokyo, Japan.

Kisacik D., Troch P. and Van Bogaert P. (2012), "Description of loading conditions
due to violent wave impacts on a vertical structure with an overhanging horizontal
cantilever slab.", Coastal Engineering 60: 201-226.

Kortenhaus A., Oumeraci H., Allsop N.W.H., Mcconnell K.J., Van Gelder P.H.A.J.M.,
Hewson P.J., Walkden M., Müller G., Calabrese M., Vicinanza D. (1999), ".1: Wave
Impact Loads-Pressures and Forces.", Final Proceedings, MAST III, PROVERBS-
Project: Vol. IIa: Hydrodynamic Aspects.

Lara J.L., Cowen E. and Sou I. (2002), ‘’A depth-of-field limited particle image
velocimetry technique applied to oscillatory boundary layer flow over a porous bed.’’,
Experiments in Fluids, SPRINGER, 33:47-53.

Lara, J. L. (2005), ‘’A numerical wave flume to study the functionality and stability of
coastal structures.’’, PIANC Magazine, AIPCN n0 121.

Lara, J. L., Garcia N., and Losada I.J. (2006), "RANS modelling applied to random
wave interaction with submerged permeable structures.", Coastal Engineering 53.5,
395-417.

123
Lara, J. L., Ruju A., and Losada I.J. (2011), "Reynolds averaged Navier–Stokes
modelling of long waves induced by a transient wave group on a beach.", Proceedings
of the Royal Society of London A, Mathematical, Physical and Engineering Sciences,
Vol. 467. No. 2129, The Royal Society.

Launder B. E., and D. B. Spalding. (1974), "The numerical computation of turbulent


flows.", Computer methods in applied mechanics and engineering 3.2, 269-289.

Le Méhauté, B. (1976). ‘’Similitude in coastal engineering.’’, Journal of the


Waterways, Harbors and Coastal Engineering Division, American Society of Civil
Engineers, 102(WW3), 317–335.

Le Méhauté, B. (1990), "Similitude.", The sea ocean engineering science 9 , 955-980.

Lin P. (1998), ‘’Numerical modelling of breaking waves’’, Ph.D. Thesis, Cornell


University.

Lin, P., and Liu P.L. (1999), "Internal wave-maker for Navier-Stokes equations
models.", Journal of waterway, port, coastal, and ocean engineering 125.4, 207-215.

Liu, P. L., and Lin P. (1997), ‘’A Numerical Model for Breaking Waves: The Volume
of Fluid Method.’’, No. CACR-97-02. Delaware University Newark Center for
Applied Coastal Research.

Liu, P. L. F., Lin P., Hsu T., Chang K., Losada I.J., Vidal C. and Sakakiyama T.,
(1999), "A Reynolds averaged Navier-Stokes equation model for nonlinear water wave
and structure interactions." Proceedings Coastal Structures. Vol. 99.

Løvholt, F., Setiadi, N.J., Birkmann, J., Harbitz, C.B., Bach, C., Fernando, N., Kaiser,
G., Nadim, F. (2014), ‘’Tsunami Risk Reduction – Are We Better Prepared Today than
in 2004?’’, International Journal of Disaster Risk Reduction 10, 127-142.

124
Minikin, R. R. (1950), ‘’Wind, Waves and Maritime Structures’’, Charles Griffin and
Co., Ltd., p.39.

Minikin, R. C. R. (1963), ‘’Winds, Waves, and Maritime Structures: Studies in


Harbour Making and in the Protection of Coasts’’, Griffin.

Mittal R. and Iaccarino G. (2005), ‘’Immersed boundary methods.’’, Annual Review.


Fluid Mechanics

Mohd-Yusof J. (1997), ‘’Combined immersed-boundary/B-spline methods for


simulations of flow in complex geometries.’’, CTR annual research briefs, 317–327,
Center for Turbulence Research, NASA Ames/Stanford University.

Mori, Nobuhito, et al. (2013), ‘’Overview of the 2011 Tohoku earthquake tsunami
damage and its relation to coastal protection along the Sanriku Coast.’’, Earthquake
Spectra 29. s1, S127-S143.

Murty T.S. (1997), Bull. Fisheries Res. Board of Canada, No. 198, Department of
Fisheries and the Environment, Fisheries and Marine Service, Scientific Information
and Publishing Branch, Ottawa, Canada.

Nagasawa T. and Tanaka H. (2012), ‘’Study of Structural Damages with massive


Geomorphic Change due to Tsunami’’, Journal of Japan Society of Civil Engineers,
Ser. B2 (Coastal Engineering) Vol. 68, No. 2

Nistor I., Palermo D., Nouri Y., Murty T., Saatcioglu M. (2009), "Tsunami-induced
forces on structures.", Handbook of Coastal and Ocean Engineering, Singapore, World
Scientific, 261-286.

Noguchi, K., Sato S., and Tanaka S. (1997), ‘’Large-scale experiments on tsunami
overtopping and bed scour around coastal revetment’’, Proceedings of Coastal
Engineering, JSCE, 44, 296-300 (in Japanese).

125
Nouri Y., Nistor I, Palermo D. and Cornett A. (2007), "Structural Analysis for
Tsunami-Induced Force and Debris Impact.", Coastal Structures 2007, Venice, Italy.

Oumeraci, H. (1984), "Scale effects in coastal hydraulic models.", Symposium on


Scale Effects in Modelling Hydraulic Structures, Technische Akademie Eßlingen.

Pugh, D.T. (1987), ‘’Tides, Surges and Mean Sea-Level: A Handbook for Engineers
and Scientists’’, John Wiley & Sons, Chichester, UK

Raosa, A. N., Zanuttigh B., Lara J.L., Hughes S. (2012), "2DV Rans-Vof modelling of
depths and velocities of overtopping waves at overwashed dikes.", Coastal Engineering
Proceedings 1.33: 62.

RAPSODI Project Deliverable D1, METU (2015), ‘’Existing tools, data, and
literature on tsunami impact, loads on structures, failure modes and vulnerability
assessment.’’, RAPSODI, Norwegian Geotechnical Report 20120768-01-R, 108 pp.
http://www.ngi.no/en/Project-pages/RAPSODI/Reports-and-Publications/ (last
accessed February 3, 2016)

Rodi, W. (1980), "Turbulent models and their application in hydraulics—a state of the
art review.", International Association for Hydraulics Research, Delft.

Saatcioglu M, Nistor I. and Ghobarah A. (2006a), ‘’Earthquake Spectra’’, Earthquake


Engineering Research Institute, 22, pp. 295–320.

Saatcioglu M, Nistor I. and Ghobarah A. (2006b), ‘’Earthquake Spectra’’, Earthquake


Engineering Research Institute, 22, pp. 355–375.

Sagara, J. and Saito K. (2013), ‘’Risk Assessment and Hazard Mapping.’’, World Bank
Knowledge Notes 1-1, Washington, DC.

126
Schäffer H. A., and Klopman G. (2000), "Review of multidirectional active wave
absorption methods.", Journal of waterway, port, coastal, and ocean engineering
126.2, 88-97.

Shih, T‐H., J. Zhu, and John L. Lumley (1996), "Calculation of Wall‐bounded


Complex Flows and Free Shear Flows.", International Journal for Numerical Methods
in Fluids 23.11, 1133-1144.

Shore Protection Manual (1984), U.S. Army Corps of Engineers, Coastal Engineering
Research Center. U.S. Government Printing Office, Washington. D.C.

Torres-Freyermuth A., Lara J.L., and Inigo J. L. (2010), "Numerical modelling of


short-and long-wave transformation on a barred beach.", Coastal Engineering 57.3,
317-330.

Wang, X. (2009), "User manual for COMCOT version 1.7 (first draft).", Cornel
University 65.

Woodroffe, C.D. (2002), ‘’Coasts: form, process and evolution.’’, Cambridge


University Press, Cambridge, UK.

Yagyu, T. (2011), ‘’Sailing ahead’’, PIANC E-Newsletter, 8. Retrieved from


http://www.pianc.org/downloads/sailingahead/SailingAheadApril2011/JapanEarthqu
ake.pdf (last accessed, October 12, 2015)

Yalciner, A. C., Synolakis C.E., Alpar B., Borrero J.C., Altinok Y., Imamura F., Tinti
S., Ersoy S., Kuran U., Pamukcu S., Kanoglu U. (2001), ‘’Field surveys and modeling
1999 Izmit Tsunami’’, Int. Tsunami Symp, ITS 2001, Seattle, Paper 4-6, 557–563.

127
Yalciner, A. C., Alpar, B., Altinok, Y., Ozbay, I., and Imamura, F. (2002), ‘’Tsunamis
in the Sea of Marmara: Historical documents for the past, models for future’’, Marine
Geology, 190(1–2), 445–463.

Yalçiner A.C., Pelinovsky E., Talipova T., Kurkin A., Kozelkov A., Zaitsev A.,
(2004), "Tsunamis in the Black Sea: comparison of the historical, instrumental, and
numerical data.", Journal of Geophysical Research: Oceans (1978–2012) 109.C12.

Yalciner, A. C., Perincek, D., Ersoy, S., Prasetya, G., Rahman, H., and McAdoo, B.
(2005), “Report on January 21–31, 2005 North Sumatra survey on December, 26,
2004 Indian Ocean Tsunami.”, ITST of UNESCO IOC.

Yang Z. and Shih T.H. (1993), ‘’A Galilean and tensorial invariant kE model for near
wall turbulence.’’, Vol. 106263. Lewis Research Center.

Yeh, H., Shinji S., and Yoshimitsu T. (2012), ‘’The 11 March 2011 East Japan
earthquake and tsunami: Tsunami effects on coastal infrastructure and buildings.’’,
Pure and Applied Geophysics, 1-13.

128
8 APPENDICES
A) IH2VOF Model Geometry - Coordinates of Element Indices

Table 8.1: Coordinates of Element Indices of Caisson Breakwater Model Geometry


in IH-2VOF

Element Vertices X Coordinate Y Coordinate


1 17.103 -0.001
2 18.103 0.102
Bathymetry
3 20.473 0.102
4 20.7362 -0.001
1 19.283 0.182
2 19.283 0.482
Caisson – 3 19.313 0.482
Vert. Wall 4 19.313 0.442
5 19.473 0.442
6 19.473 0.182
1 18.763 0.102
2 19.083 0.182
Core Layer
3 19.673 0.182
4 19.993 0.102
1 18.603 0.102
2 19.083 0.222
3 19.283 0.222
Filter Layer-1
4 19.283 0.182
5 19.083 0.182
6 18.763 0.102
1 19.473 0.182
2 19.473 0.222
3 19.673 0.222
Filter Layer-2
4 20.153 0.102
5 19.993 0.102
6 19.673 0.182
1 -0.005 -0.005
2 -0.005 0.402
Water
3 22.305 0.402
4 22.305 -0.005

129
B) Breakwater Profile Measurements and Photos

Wave Number 12
0,5
0,45 Before

0,4
Line 1 After first 500
Y - Vertical Distance (m)

0,35 After first 1000

0,3 After sec 1000


0,25
0,2
0,15
0,1
0,05
0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8
X - Horizontal Distance (m)

Figure 8.1: Breakwater profile of caisson type breakwater – Line 1

0,5
0,45 Before

0,4
Line 2 After first 500
Y = Vertical Distance (m)

0,35 After first 1000


0,3
After sec 1000
0,25
0,2
0,15
0,1
0,05
0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8
X = Horizontal Distance (m)

Figure 8.2: Breakwater profile of caisson type breakwater – Line 2

130
0,6
Before
0,5
Line 1
After first 500
Y - Vertical Distance (m)

After first 1000


0,4
After sec 1000
0,3

0,2

0,1

0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6
X - Horizontal Distance (m)

Figure 8.3: Breakwater profile of rubble mound breakwater – Line 1

0,6
Before
0,5
Line 2 After first 500
Y - Vertical Distance (m)

After first 1000


0,4
After sec 1000

0,3

0,2

0,1

0
0 20 40 60 80 100 120 140 160
X - Horizontal Distance (m)

Figure 8.4: Breakwater profile of rubble mound breakwater – Line 2

131
Figure 8.5: View of caisson type breakwater after Wave Number 12

Figure 8.6: View of rubble mound breakwater after Wave Number 12

Wave Number 15
Set-2
0,5

0,45 Line 1 Before After


0,4
Y - Vertical Distance (m)

0,35

0,3

0,25

0,2

0,15

0,1

0,05

0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8
X - Horizontal Distance (m)

Figure 8.7: Breakwater profile of caisson type breakwater – Line 1

132
0,5
0,45 Line 2 Before After
0,4
Y - Vertical Distance (m)

0,35
0,3
0,25
0,2
0,15
0,1
0,05
0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8
X - Horizontal Distance (m)

Figure 8.8: Breakwater profile of caisson type breakwater – Line 2

0,6
Before After
0,5 Line 1
Y - Vertical Distance (m)

0,4

0,3

0,2

0,1

0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6
X - Horizontal Distance (m)

Figure 8.9: Breakwater profile of rubble mound breakwater – Line 1

133
0,6

0,5
Line 2 Before After
Y - Vertical Distance (m)

0,4

0,3

0,2

0,1

0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6
X - Horizontal Distance (m)

Figure 8.10: Breakwater profile of rubble mound breakwater – Line 2

Figure 8.11: View of caisson type breakwater after Wave Number 15, Set-2

Figure 8.12: View of rubble mound breakwater after Wave Number 15, Set-2

134
Wave Number 17
Set-1
0,5

0,45 Line 1 Before After


0,4
Y - Vertical Distance (m)

0,35

0,3

0,25

0,2

0,15

0,1

0,05

0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8
X - Horizontal Distance (m)

Figure 8.13: Breakwater profile of caisson type breakwater – Line 1

0,5

0,45
Line 2 Before After
0,4
Y - Vertical Distance (m)

0,35

0,3

0,25

0,2

0,15

0,1

0,05

0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8
X - Horizontal Distance (m)

Figure 8.14: Breakwater profile of caisson type breakwater – Line 2

135
0,6

0,5
Line 1 Before After
Y - Vertical DIstance (m)

0,4

0,3

0,2

0,1

0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8
X - Horizontal Distance (m)

Figure 8.15: Breakwater profile of rubble mound breakwater – Line 1

0,6

0,5
Line 2
Before After
Y - Vertical Distance (m)

0,4

0,3

0,2

0,1

0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8
X - Horizontal Distance (m)

Figure 8.16: Breakwater profile of rubble mound breakwater – Line 2

136
Figure 8.17: View of caisson type breakwater after Wave Number 17, Set-1

Figure 8.18: View of rubble mound breakwater after Wave Number 17, Set-1

Set-2
0,5
0,45
Line 1 Before After
0,4
Y - Vertical Distance (m)

0,35
0,3
0,25
0,2
0,15
0,1
0,05
0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8
X - Horizontal Distance (m)

137
Figure 8.19: Breakwater profile of caisson type breakwater – Line 1

0,5

0,45 Line 2 Before After


0,4
Y - Vertical Distance (m)

0,35

0,3

0,25

0,2

0,15

0,1

0,05

0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8
X - Horizontal Distance (m)

Figure 8.20: Breakwater profile of caisson type breakwater – Line 2

0,5
0,45 Line 1 Before After
0,4
Y - Vertical Distance (m)

0,35
0,3
0,25
0,2
0,15
0,1
0,05
0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8
X - Horizontal Distance (m)

Figure 8.21: Breakwater profile of rubble mound breakwater – Line 1

138
0,5
0,45 Line 2 Before After
0,4
Y - Vertical Distance (m)

0,35
0,3
0,25
0,2
0,15
0,1
0,05
0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8
X - Horizontal Distance (m)

Figure 8.22: Breakwater profile of rubble mound breakwater – Line 2

Figure 8.23: View of caisson type breakwater after Wave Number 17, Set-2

Figure 8.24: View of rubble mound breakwater after Wave Number 17, Set-2

139
Wave Number 18
Set-1
0,5
0,45 Line 1 Before After
0,4
Y - Vertical Distance (m)

0,35
0,3
0,25
0,2
0,15
0,1
0,05
0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8
X - Horizontal Distance (m)

Figure 8.25: Breakwater profile of caisson type breakwater – Line 1

0,5
0,45
Line 2 Before After
0,4
Y - Vertical Distance (m)

0,35
0,3
0,25
0,2
0,15
0,1
0,05
0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8
X - Horizontal Distance (m)

Figure 8.26: Breakwater profile of caisson type breakwater – Line 2

140
0,6

0,5
Line 1
Before After
Y - Vertical Distance (m)

0,4

0,3

0,2

0,1

0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6
X - Horizontal Distance (m)

Figure 8.27: Breakwater profile of rubble mound breakwater – Line 1

0,6

0,5
Line 2
Before After
Y - Vertical Distance (m)

0,4

0,3

0,2

0,1

0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6
X - Horizontal Distance (m)

Figure 8.28: Breakwater profile of rubble mound breakwater – Line 2

141
Figure 8.29: View of caisson type breakwater after Wave Number 18, Set-1

Figure 8.30: View of rubble mound breakwater after Wave Number 18, Set-1

Set-2
0,5

0,45
Line 1 Before After
0,4
Y - Vertical Distance (m)

0,35

0,3

0,25

0,2

0,15

0,1

0,05

0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8
X - Horizontal Distance (m)

Figure 8.31: Breakwater profile of caisson type breakwater – Line 1

142
0,5
0,45 Before After
Line 2
0,4
Y - Vertical Distance (m)

0,35
0,3
0,25
0,2
0,15
0,1
0,05
0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6 1,8
X - Horizontal Distance (m)

Figure 8.32: Breakwater profile of caisson type breakwater – Line 2

0,6

Line 1 Before After


0,5
Y - Vertical Distance (m)

0,4

0,3

0,2

0,1

0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6
X - Horizontal Distance (m)

Figure 8.33: Breakwater profile of rubble mound breakwater – Line 1

143
0,6

0,5
Line 2 Before After
Y - Vertical Distance (m)

0,4

0,3

0,2

0,1

0
0 0,2 0,4 0,6 0,8 1 1,2 1,4 1,6
X - Horizontal Distance (m)

Figure 8.34: Breakwater profile of rubble mound breakwater – Line 2

Figure 8.35: View of caisson type breakwater after Wave Number 18, Set-2

Figure 8.36: View of rubble mound breakwater after Wave Number 18, Set-2

144

You might also like