Sie sind auf Seite 1von 220

Breakwater stability under tsunami attack

for a site in Nicaragua

ing. K. Cuypers

The background picture of the front page is a well known colour woodcut by
Katsushika Hokusai, a famous late eighteenth century Japanese artist. It is No 20 from
the series Thirty-Six Views on the Mount Fuji. Many textbooks and web sites depict
this wave as a tsunami wave. Indeed it resembles the shape of a tsunami wave when
breaking occurs. But in fact it is a wind generated wave. Nevertheless it has become
an international symbol for tsunamis.
The small pictures in front are:
1. Severe damaged caisson breakwater at the Port of Okushiri after the 1993
Hokkaido tsunami
2. A snapshot of a computer model of the 12 July 1993 tsunami in the East
Japanese Sea
3. Damage due to the 1964 Alaska tsunami
4. The 1946 Aleutian tsunami arrives on the Hawaiian coast as a bore
5. An unbroken tsunami wave arriving at a steep coast

Breakwater stability under tsunami attack


for a site in Nicaragua

Final Report

05-01-2004

ing.K.Cuypers

Colophon
Master Thesis:

Faculty of Civil Engineering, TU Delft

Cooperating company:

HAECON n.v.
Harbour and Engineering Consultants
Deinsesteenweg 110
B-9031 Drongen
Belgium

Supervising committee:
TU Delft
HAECON

Prof. M.J.F. Stive


Prof. G.S. Stelling
ir. H.J. Verhagen
Dr.ir. M. Huygens
ir. N. Gunst

Student:

ing. K. Cuypers
kim_cuypers@hotmail.com
st.nr. 1 055 739

Date:

December 2003

Preface
This report is the final thesis for the masters degree in Civil Engineering at the
Delft University of Technology. The study was conducted in cooperation with
the Belgian Harbor and Engineering Consultants company HAECON n.v.
I would like to thank the following people who attributed to the realization of
this thesis. First of all, I would like to thank my graduation committee at TU
Delft: Prof. M.J.F. Stive, who as chairman of the committee inspired me with
his enthusiasm, Prof. G.S. Stelling, for his help in setting up the mathematical
model and especially ir. H.J. Verhagen, who was always available for giving
advice. Secondly, I would like to thank the people of HAECON for their
continuous support, in particular ir. N. Gunst who was always available to
answer my questions and to provide me with the necessary information
concerning the Nicaragua project. Also, I would like to thank Dr. ir. M.
Huygens and ir. P. de Pooter for their feedback on the conducted research.
Special thanks to Prof. J. De Rouck and Dr.ir. P. Troch, of the University of
Ghent, for their critical reflections on the breakwater stability, and to Dr.ir.
P.H.A.J.M. van Gelder, of TU Delft, for his kind advice regarding some
statistical questions.
Finally, I would like to thank my parents for their support during all these
years.
Kim Cuypers
The Hague, 5 January 2003

CONTENTS

COLOPHON

PREFACE

INTRODUCTION

PART 1: Design Tsunami


1.

GENERAL INTRODUCTION TO TSUNAMIS

2.

TSUNAMIS ON THE PACIFIC SIDE OF CENTRAL AMERICA

BASICS OF TSUNAMI SCIENCE

3.1
3.1.1
3.1.2
3.2
3.2.1
3.2.2
3.2.3
3.2.4
3.2.5
3.3
3.3.1
3.3.2
3.3.3
3.3.4
3.4
3.4.1
3.4.2
4

TSUNAMI GENERATION BY EARTHQUAKES..........................................................................1


Generation mechanism ..............................................................................................................1
Tsunami earthquakes .................................................................................................................3
TSUNAMI WAVES ............................................................................................................................6
Propagation ................................................................................................................................6
Terminology................................................................................................................................9
Different wave forms................................................................................................................10
Numerical calculations............................................................................................................11
Breaking....................................................................................................................................12
TSUNAMI MODELING...................................................................................................................13
Tsunami source modeling ........................................................................................................14
Tsunami propagation ...............................................................................................................15
Calibration................................................................................................................................16
Summary and comments .........................................................................................................17
TSUNAMI SCALES..........................................................................................................................18
Scales ........................................................................................................................................18
Comments from a hydraulic engineering point of view.........................................................23

DETERMINING THE CONCEPT OF THE DESIGN TSUNAMI

4.1
ANALYZING THE TSUNAMI DATA FOR THE PACIFIC SIDE OF CENTRAL AMERICA 1
4.1.1
Homogenizing the dataset..........................................................................................................1
4.1.2
Do the available data for these events contain enough information for deriving a wave
or a wave train? .........................................................................................................................7
4.2
DETERMINING THE LARGEST TSUNAMI EVENT AT THE PACIFIC COAST OF
CENTRAL AMERICA........................................................................................................................9
4.2.1
Occurrence ...............................................................................................................................11
4.2.2
Compared to the entire Pacific Region...................................................................................14
4.3
CONCLUSIONS................................................................................................................................16

DERIVING THE DESIGN TSUNAMI FOR THE SITES OF INTEREST

5.1
5.2
5.2.1
5.2.2
5.2.3
5.3
5.3.1
5.3.2
5.3.3
5.3.4
5.4
5.4.1
5.4.2
5.4.3
5.5
6

INTRODUCTION................................................................................................................................1
MODEL SET-UP .................................................................................................................................2
Model description.......................................................................................................................2
Model input.................................................................................................................................3
Available calibration data .........................................................................................................7
PIE DEL GIGANTE ..........................................................................................................................10
First Calculation ......................................................................................................................10
Possible reasons.......................................................................................................................14
Calibration of the model: the 1992 Nicaraguan tsunami......................................................31
Design wave..............................................................................................................................35
BAHIA DEL SALINAS ....................................................................................................................38
Bathymetry at Bahia Del Salinas ............................................................................................39
Design wave at Bahia Del Salinas..........................................................................................40
Comparison between the design wave at Bahia del Salinas and Pie del Gigante..............41
DESIGN TSUNAMI WAVE .....................................................................................................................42

CONCLUSIONS

PART 2: Breakwater Stability


7

LITERATURE STUDY ON THE CAISSON BREAKWATER STABILITY

7.1
ON THE HYDRAULIC ASPECTS OF TSUNAMI BREAKWATERS IN JAPAN ....................1
Design formulas ..............................................................................................................................1
7.1.1
7.1.2
Comparing to the Goda design formulas for unbroken short period waves ...............................3
7.1.3
Comments ........................................................................................................................................6
7.2
OTHER REFERENCES...................................................................................................................... 7
7.3
CONCLUSION ....................................................................................................................................7
8
8.1
8.2
8.3
8.4
9
9.1
9.2
9.3
9.4
10

DESIGN OF THE CAISSON BREAKWATER


THE CAISSON BREAKWATER FOR PIE DEL GIGANTE.........................................................1
THE DESIGN TSUNAMI WAVE .....................................................................................................1
CAISSON TSUNAMI STABILITY................................................................................................... 2
CONCLUSIONS..................................................................................................................................3
LITEARTURE STUDY ON THE RUBBLE MOUND BREAKWATER STABILITY
LABORATORY STUDY FOR THE DESIGN OF TSUNAMI BARRIER....................................1
DYNAMIC WAVE FORCE OF TSUNAMIS ACTING ON A STRUCTURE .............................7
EXPERIMENTAL STUDIES ON TSUNAMI FLOW AND ARMOR BLOCK STABILITY
FOR THE DESIGN OF A TSUNAMI PROTECTION BREAKWATER IN KAMAISHI BAY .9
CONCLUSIONS................................................................................................................................14
THEORETICAL STABILITY CONCEPT FOR THE RUBBLE MOUND BREAKWATER

10.1
STONE STABILITY UNDER FLOW...............................................................................................2
10.1.1
Izbash ..........................................................................................................................................2
10.1.2
Dutch Formula ...........................................................................................................................3
10.1.3
CERC formula ............................................................................................................................4
10.2
STONE STABILITY GOVERNED BY THE VELOCITY ON TOP OF THE
BREAKWATER CREST ....................................................................................................................5
10.2.1
A comparison between the calculated values...........................................................................7

10.2.2
A comparison between the calculated values and the experimental results...........................8
10.3
STONE STABILITY GOVERNED BY THE VELOCITY ON THE BACKSIDE
SLOPE OF THE BREAKWATER.....................................................................................................8
10.3.1
Velocity at the backside ............................................................................................................9
10.3.2
Comparing the calculated velocities with the experimental results by Kamel (1970)........10
10.3.3
Comparing the required stones with the experimental results by Kamel (1970) ................11
10.4
CONCLUSIONS................................................................................................................................13
11

DESIGN OF THE RUBBLE MOUND BREAKWATER

11.1
11.2
11.3

THE RUBBLE MOUND BREAKWATER FOR PIE DEL GIGANTE..........................................1


THE DESIGN TSUNAMI WAVE .....................................................................................................1
CONCLUSIONS..................................................................................................................................3

12

CONCLUSIONS

13.

SUMMARY AND CONCLUSIONS

PART I DESIGN TSUNAMI.............................................................................................................................1


Occurrence of the 1992 Nicaraguan tsunami ..............................................................................................2
Design tsunami...............................................................................................................................................2
PART II: BREAKWATER STABILITY..........................................................................................................5
Stability Concept for the Caisson and the Rubble Mound Breakwater ...................................................... 5
Design Calculations....................................................................................................................................... 7
14.

RECOMMENDATIONS

APPENDICES
1. Databank of Tsunamis at the Pacific side of the Central American Region
2. A Short Introduction to Seismology
3. Mail correspondence concerning the deriving of tsunami wave characteristics out of tsunami data
or seismic data
4. The 1992 Nicaraguan tsunami
5. Bathymetry
6. Preliminary design of the Caisson Breakwater at Pie del Gigante
7. Comparing the Tanimoto design formulas for unbroken tsunami waves with the Goda design
formulas for unbroken short period waves
8. Calculations of the caisson stability under the design tsunami wave
9. Calculations concerning chapter 10: Theoretical stability concept for the rubble mound breakwater
10. Iteration problems of the Dutch formula
11. Rubble Mound breakwater for Pie del Gigante
BIBLIOGRAPHY

Breakwater stability under tsunami attack


for a site in Nicaragua

Introduction
This thesis is related to the technical feasibility study of the Nicaragua Dry
Canal. The Nicaragua Dry Canal is a container traffic between the Atlantic and
the Pacific coast of Nicaragua. The containers will be transported by train from
a container port at the Pacific coast to a container port at the Atlantic coast. This
thesis deals with the feasibility of the container port at the Pacific side of
Nicaragua. As the region is sensitive to tsunami attack, it is essential to take the
tsunami risk into consideration.

The two proposed harbor locations at the Nicaraguan coast.

This study will focuses on the stability of the breakwater under tsunami attack,
for the two proposed harbor locations, Pie del Gigante and Bahia del Salinas.
The study consists of two parts: the derivation of a design tsunami and the
breakwater stability.
Part 1: Design Tsunami
The aim of part 1 is to formulate a realistic design tsunami for the two harbor
locations at the Pacific Coast of Nicaragua.
At first the basics of tsunami science will be presented. Next out of the
available tsunami data for Nicaragua it will be tried to derive a design tsunami
wave. A statistical approach for deriving the design tsunami is preferred, but it
is uncertain if this will be possible. The design tsunami has to be derived for
deeper water as to minimize local disturbance. And from deeper water the
tsunami has to be derived to the two possible harbor locations. For these wave
calculations a simple set-up is preferred with a 1D hydraulic model.

In chapter 1, a general introduction to tsunamis is given. Chapter 2 gives an


overview of the tsunamis that have occurred in Central America, and of the
available data concerning these events. In chapter 3, the necessary background
knowledge is given to judge the possibility for linking the available tsunami
data to a tsunami wave. Also insight is given into the relevant tsunami science.
Chapter 4 analyses the dataset from chapter 2 and it is decided if a design
tsunami will be derived by a statistical approach or based on a large historical
event. In chapter 5, the design tsunami event is derived for the two locations in
Nicaragua. In chapter 6, the conclusions concerning part 1 are briefly
summarized.
Part 2: Breakwater Stability
The aim of part 2 is to formulate a breakwater design which can withstand the
design tsunami derived in part 1.
Two different breakwater types will be studied: the caisson breakwater and the
rubble mound breakwater. For both structures, first the available literature will
be reviewed. If the available literature turns out to be insufficient for designing
a breakwater, a theoretical stability concept will be set up. Afterwards, the
breakwater design for short period waves will be tested on a tsunami load and
if necessary the design will be adapted.
In chapter 7, the available literature on caisson breakwater stability under a
tsunami attack is presented. In chapter 8, the caisson breakwater is designed to
withstand the design tsunami. The literature study concerning the stability of
rubble mound breakwaters is presented in chapter 9. Because of the limited
specific references concerning the stability of rubble mound breakwaters, a
theoretical stability concept is set up in chapter 10. In chapter 11, the rubble
mound breakwater is designed to withstand the design tsunami load. In chapter
12 the conclusions concerning part 2 are briefly summarized.
Chapter 13 gives a summary of the entire study, together with an overview of
the conclusions. Recommendations for further research are formulated in
chapter 14.

Part I: Design Tsunami

The Waiakea area of Hilo, Hawaii, after the 1960 Chilean tsunami struck the Hawaiian
islands. The force of the debris filled waves bent the parking meters.

1. GENERAL INTRODUCTION TO TSUNAMIS


A tsunami is a series of ocean waves generated by any rapid large-scale
disturbance of the seawater. Most tsunamis are generated by earthquakes
(Satake, 2002). Other occasional causes are, volcanic eruptions, landslides,
undersea landslides, underwater explosions or meteor impacts. All these events
trigger a series of fast moving, long waves of initial low amplitude that radiate
outward in a manner resembling the waves radiating when a pebble is dropped in
a pond.

Figure 1.1
Tsunami waves radiating outward like when a pebble is dropped in a pond.

The term tsunami comes from the Japanese term meaning harbor wave. Because
tsunamis often cause large standing waves in harbors and bays, which can persist
for many hours, tsunamis in ancient Japan became known as great harbor waves.
Other terms for tsunamis are tidal wave or seismic sea wave. The term tidal
wave is an exact translation from the ancient Greek name for a tsunami. It refers
to the initial manifestation of a tsunami at the shore, where it often resembles a
fast ebbing or flooding tide. The term tidal wave is less commonly used, to
avoid the association with tides. It is not only incorrect with regard to its origin as
a tsunami has nothing to do with the tides, but also inappropriate in its descriptive
character. Seismic sea wave is the most descriptive term, as most tsunamis are
generated by large underwater earthquakes.
As stated above far out the most common cause for tsunami generation are
earthquakes. Earthquakes are commonly associated with ground shaking that is a
result of elastic waves traveling through the solid earth. However, near the source
of submarine earthquakes, the seafloor is permanently uplifted and down-dropped,
pushing the entire water column up and down (Figure 1.2).

05-01-2004

Part 1 - Chapter 1 - 1

Figure 1.2
Underwater earthquake resulting in an uplift and down-drop of the ocean surface.

The potential energy that results from pushing water above mean sea level is then
transferred to horizontal propagation of the tsunami wave (kinetic energy) (Figure
1.3). The initial water displacement is split into a tsunami that travels into the
deep ocean (distant tsunami) and another tsunami that travels towards the nearby
coast (local tsunami). In deeper water a tsunami is barely noticeable and will only
cause a small and slow rising and falling of the sea surface as it passes.

Figure 1.3
Wave generation and propagation as a result of the initial surface disturbance

Near their generation point, tsunamis have a great wave length, often exceeding
200 km. The maximum ocean depth lies between 8 to 10 km in the ocean
trenches. This makes a tsunami a typical long wave, as the wave length is much
larger than the water depth. Therefore the propagation speed of tsunamis does
only depends on the water depth ( c = g .h ). As the local tsunami travels over the
continental slope, shoaling begins: the tsunami waves slow down and become
compressed, causing them to grow in height. This results in steepening of the
waves (Figure 1.4).

Part 1 - Chapter 1 - 2

05-01-2004

Figure 1.4
Shoaling process of tsunami wave, clearly showing the steepening of the wave

Contrary to many artistic images of tsunamis, most tsunamis arriving at the coast
do not result in giant breaking waves. Rather, they come in much like very strong
and very fast tides (i.e., a rapid, local rise in sea level) (Bryant, 2001). Much of
the damage inflicted by tsunamis is caused by strong currents and floating debris.
The small number of tsunamis that do break often form vertical walls of turbulent
water called bores.
Because of their great length, arriving tsunamis at the coast travel much farther
inland than normal wind waves. The extend of the area that is affected by the
onshore tsunami is expressed by the run-up height. The run-up height is the height
above a reference level, which the water reaches as the tsunami waves run out on
the coast (Figure 1.5).

Hrun-up

Figure 1.5
Tsunami arriving at a coast as fast incoming tide, causing a large run-up.

Tsunamis rank high on the scale of natural disasters. It has been estimated that
tsunamis cause between 5 and 15% of the earthquake damage worldwide
(Bernard, 2002). Since 1850, tsunamis have been responsible for the loss of over
120,000 lives and billions of U.S. dollars damage to coastal structures and
habitats. Over 99% of these casualties were caused by local tsunamis that occur
about once a year somewhere in the world. The last decade 12 major tsunamis
have struck coastlines around the Pacific Rim, causing more than 4,000 casualties
and an estimated 1 billion U.S. dollars in damage. This makes tsunamis the most
destructive ocean waves of all.
05-01-2004

Part 1 - Chapter 1 - 3

2. TSUNAMIS ON THE PACIFIC SIDE OF CENTRAL


AMERICA
As in the literature no ready-to-use methods are available for a statistical approach
for determining a design tsunami, it will be tried to set up a statistical approach
based on available tsunami data for Nicaragua.
In this chapter the necessary dataset of tsunami events will be derived. On the
basis of the available data it will be decided which extra information is necessary.
Although the sites of interest are located in Nicaragua, due to the exceptional
occurrence one is forced to put the statistical investigation into a broader view of
the region. As a first assumption the Pacific side of Central America (Mexico to
Panama) is regarded as one region.
Tsunamis are classified into three categories, distant (> 750km from the source),
regional (100-750km from the source) and local (< 100km from the source)
(Fernandez, 2000). In principle the Pacific coast of Central America can be
exposed to the three types of tsunamis.
Looking at the entire Pacific Region Ida (1981) sets up a distribution for tsunami
generating earthquakes for different regions in the Pacific (Figure 2.1). As can be
seen Central America is not the most active tsunamigenic region in the Pacific,
but a significant part of all generated tsunamis are generated in Central America:
6.5% during 1900-1980. This makes local and regional tsunamis a significant
hazard for the Central American coast.
Due to geographical conditions, the impact from distant tsunamis is relatively
small for the Central American region. For the total energy radiating from the
circum Pacific zone, the percentage of the received energy was only 6.4% in
Central America during 1900-1992 (Hatori, 1995).
Therefore we can focus in this study on the local and regional tsunamis in Central
America, meaning that the source region can be limited to the Central American
region.

tsumanigenic: tsunami generating (e.g. tsumanigenic earthquakes)

05-01-2004

Part 1 - Chapter 2 - 1

Figure 2.1
Local and regional tsunami activities in the Pacific Ocean, from 1900 to1980
expressed in percent

A dataset for the tsunamis generated at the Pacific side of the Central American
region for the period 1900-2002, was derived out of the Historical Tsunami
Database for the Pacific, 47 B.C. to present. This is the most up to date tsunami
catalogue available. In appendix 1 it is described how the data are retrieved.
The results of this query are presented in Table 2.1. The epicentral locations of the
tsumanigenic earthquakes are plotted on a map of the Central American region
(Figure 2.1).

Part 1 - Chapter 2 - 2

05-01-2004

date

time

generation point

Year

Mo

Da

Hr

Mn

1906

31

15

36

Sec

1907

15

1928

22

17

9.6

1928

17

19

31.8

16.65

earthquake data

tsunami data

generation area

Lat

Long

Dep

Ms

Mw

Mt

Int

Hrmax

-81.5

25

8.6

8.7

8.7

1000

SAM Columbia-Ecuador:TUMACO

16.7

-99.2

25

8.2

7.8

CAM Mexico

16.84

-96.02

15

7.5

7.4

-96.5

7.8

7.9

8.1

-2
1.5

1930

31

40

36.6

33.9

-118.6

5.2

1932

10

36

54.2

19.46

-104.17

25

8.1

7.6

8.2

1932

18

10

12

17.8

19.53

-103.72

71

7.8

7.5

7.8

1932

22

12

59

27.5

19.09

-104.41

6.9

TR

Source Region

CAM MEXICO

CAM

MEXICO: NEAR COAST OF


GUERRER

2.5

-0.5

6.1

CAM S.CALIFORNIA

-0.5

0.7

CAM Central Mexico,Jalisco

0.1

CAM Central Mexico, Jalisco

10

CAM Central Mexico

1934

18

36

27.8

8.01

-82.56

25

7.7

7.6

CAM Costa Rica-Panama

10

1941

12

20

47

3.3

8.61

-83.18

32

7.5

7.3

-3

0.1

CAM Central America

11

1941

12

8.5

-84

33

-3

0.1

CAM Costa Rica-Panama

12

1948

12

22

21.6

-106.7

33

CAM Mexico:MARIA MADRE ISLAND

13

1950

10

16

10.4

-85.2

60

7.7

7.7

7.8

1.5
-2

0.1

CAM Nicaragua,Costa Rica-Panama

14

1950

10

23

16

13

14.3

-91.8

30

7.1

7.5

7.5

-1

0.2

CAM Guatemala-Nicaragua

15

1950

12

14

14

15

16.3

-98.2

50

7.5

7.1

-1

0.3

CAM Acapulco, S.Mexico

16

1951

24

13

-87.5

100

17

1957

28

40

9.8

16.88

-99.29

36

7.9

18

1958

19

14

25

0.99

-79.48

20

7.3

19

1962

12

11

40

15.3

8.09

-82.68

19

6.8

20

1962

11

14

11

56.8

17.14

-99.7

46

7.2

21

1962

19

14

58

13.1

16.99

-99.69

16

7.1

22

1965

23

19

46

1.5

16.17

-95.85

10

7.8

23

1973

30

21

13.5

18.45

-102.96

37

7.5

7.6

24

1979

14

11

15

17.77

-101.23

24

7.6

7.6

25

1979

12

12

59

4.6

1.6

-79.36

24

7.7

8.1

26

1981

10

25

22

15.8

18.12

-102

21

7.3

7.7

10

1.5

2.6

1.5

CAM Potosi, Honduras

68

CAM S.Mexico:GUERRERO

20

SAM Colombia-Ecuador

-3

0.1

CAM Costa Rica-Panama

0.8

CAM S.Mexico

30

CAM S.Mexico

7.5

-1

0.4

CAM MEXICO

-3

0.1

56

CAM S.Mexico:FARIAS,TECOMAN

-1.1

0.4

CAM S.Mexico:GUERRERO

2.5

600

SAM Colombia-Ecuador

7.2

-3

0.1

CAM MEXICO

CAM MEXICO: MICHOACAN: MEXICO CI

27

1985

19

13

17

49.7

18.46

-102.37

21

8.1

1.5

28

1985

21

37

13.5

17.83

-101.62

18

7.5

7.5

1.2

29

1988

30

15

42

32.6

-117.3

CAM MEXIC0: SW COAST: MEXICO CIT

CAM San Diego, S. California

30

1992

16

2.7

11.72

-87.39

45

7.2

7.7

2.8

10.7

170

CAM Nicaragua

31

1994

17

12

30

55.5

34.16

-118.57

14

6.8

6.7

-3

0.1

CAM Southern California

32

1995

14

14

33.9

16.84

-98.61

29

7.2

7.3

-1

0.42

SAM Oaxaca, Mexico

33

1995

10

15

35

54.1

19.05

-104.21

26

7.3

CAM Jalisco, Mexico

16.6

15.93

-98.1

17

6.9

7.1

-3

0.12

CAM Oaxaca, Mexico

17.62

-101.6

2.5

CAM Guerero, Mexico

34

1996

25

35

1999

12

18

7.9

Table 2.1
Databank for the tsunamis at the Pacific side of Central America from 1900 to 2002.
The data are retrieved out of the Historical Tsunami Database for the Pacific,
47 B.C. to present.

Legend
Earthquake data
Dep
Ms
Mw

05-01-2004

:
:
:

depth of the source earthquake


surface-wave magnitude
moment magnitude

Part 1 - Chapter 2 - 3

Tsunami data
Mt
Int
Hrmax
N
D
N
S
M
L
F :
C:
T
L
M
U

: tsunami magnitude
: tsunami intensity
: maximum observed or measured wave height in meters
: total number of available run-up and tide-gauge observations
: damage code
:
non damaging event
:
slight damage
:
moderate damage
:
large (severe) damage
number of reported fatalities due to the event
cause of the tsunami
:
tectonic
:
landslide
:
meteorological
:
unknown

Generation area
TR
CAM
SAM
Source Region

:
:
:
:

tsunamigenic region code


Central America (7 N - 35 N, 125 W - 75 W)
South America (58 S - 7 N, 100 W - 60 W)
source region of the tsunami. Descriptive indication of the tsunami source area

Figure 2.2
The epicentral locations of tsunamigenic earthquakes in Central America for the
period 1900 to 2002

Part 1 - Chapter 2 - 4

05-01-2004

To be able to work with this data, and to make correct interpretations one has to
have the necessary background knowledge concerning tsunamis. In the following
chapter an overview of tsunami science will be given from a hydraulic
engineering point of view.
Looking at the cause of a tsunami it is clear that earthquakes are the dominant
cause : 32 tsunamis on 35 are generated by earthquakes and these events are
responsible for all casualties in Central America. This conforms the general idea
that tsunamis are predominantly generated by earthquakes. Therefore the rest of
the study will be emphasized on earthquake generated tsunamis. In the following
chapter an overview of the most relevant subjects of tsunami science will be
given. In appendix 2 a short introduction to seismology is given, which is a must
when one is not acquainted with seismology.
With the background knowledge out of chapter 3 and appendix 2 it will be tried,
in chapter 4, to decide how a design wave can be coupled to the available dataset;
is a statistical approach possible or does the largest tsunami event has to be taken
as a base for a design tsunami.

05-01-2004

Part 1 - Chapter 2 - 5

3 BASICS OF TSUNAMI SCIENCE


In this chapter an overview is given of the basics of tsunami science. When one
starts to study tsunamis one gets a very fragmented and incomplete notion of the
tsunami subject. Because on one hand a lot of information is easily available in
the so called popular sciences. On the other hand very specific information can be
found in publications. In between, very little information is available. Information
out of popular sciences gives a very superficial insight and they have the tendency
of focusing on exceptional events. Publications are usually focused on one
particular aspect of tsunamis. Therefore compiling a correct and clear
representation about what a tsunami actually is, how it behaves itself, how it can
be described, how it can be modeled is a difficult and time consuming job.
This chapter is a compilation of tsunami science emphasized on the considered
problem in this study: deriving a design tsunami to a coastal location in order to
use it for breakwater stability analysis. The purpose of this chapter is not to give a
total overview of tsunami science, but to give a correct, summarized
representation. It is therefore the key to the understanding of this study.
The chapter is split up into a paragraph about tsunami generation by earthquakes,
a paragraph concerning tsunami waves, a paragraph describing the modeling of a
tsunami and a paragraph giving an overview of the common tsunami scales.

3.1 TSUNAMI GENERATION BY EARTHQUAKES


3.1.1 Generation mechanism
Although landslides, volcanoes and asteroid impacts can all trigger tsunamis, by
far the most common cause are submarine earthquakes. Even if the seafloor itself
does not trigger tsunamis, the shaking may trigger coseismic landslides, which
can cause tsunamis.
Not all earthquakes generate tsunamis. The pattern and extent of vertical ground
deformation from an earthquake uniquely determines whether or not a tsunami is
formed. The ground deformation is determined by the fault geometry. As
explained in appendix 2, earthquake fault geometries basically fall apart in three
fundamental end members: strike-slip, thrust-slip and dip-slip faults (Figure 3.1).
Strike-slip faults involve horizontal motion of the earths crust, while thrust-dip
and dip-slip faults induce vertical motion.

05-01-2004

Part 1 - Chapter 3 - 1

Figure 3.1
Basic fault geometries

Submarine thrust and dip-slip faults produce tsunamis as the seafloor lifts up or
drops down, and either pushes the water up or pulls it down, triggering wave
motion on the ocean surface. On the other hand, strike-slip motions, generally, do
not generate sufficient vertical displacement on the seafloor, yet they may
generate tsunamis through coseismic events. Most faults combine both strike-slip
and thrust motion, but primarily only faults that have predominantly vertical
displacement and create sufficiently large seafloor deformations appear to trigger
tsunamis.

Figure 3.2
Tsunami generation by a subduction zone earthquake. The initial water
displacement equals the bottom displacement

To understand the process of tsunami generation by earthquakes, one is referred to


Figure 3.2 which shows a typical subduction zone earthquake and the initial water
displacement of the oceans surface. The deformed area usually contains an area of
uplift and subsidence, the initial water surface displacement is equal to the bottom
displacement. If the rupture is closer to the shore then the subsidence takes place
primarily onshore and the initial water displacement may only manifest itself as a
solely uplift zone.
An earthquake whose epicenter lies inland will only generate a tsunami if it
produces sufficient vertical displacement offshore on the seafloor; Therefore only
very strong inland thrust earthquakes, as compared to even moderate offshore
Part 1 - Chapter 3 - 2

05-01-2004

earthquakes, are potential tsunami generators (unless of course they trigger a


massive landslide into the sea).
General the larger the moment magnitude of an earthquake, the larger the area that
is deformed and the greater amount of slip, thus producing disproportional larger
tsunamis than smaller events.
In addition to an earthquakes magnitude, the deeper the hypocenter or focus of an
earthquake, the smaller the vertical deformation of the earths surface. A deeper
hypocenter allows the seismic energy to spread over a larger volume. Earthquakes
deeper than about 30km rarely cause sufficient deformation to generate tsunamis.
But truly great megathrust earthquakes that occur deeper than 30km, such as the
1960 Chilean event (Mw=9.5), can occasionally trigger tsunamis.
However often the earthquake magnitude suggests no or only a small tsunami to
be generated, as in fact a considerable tsunami is generated, these earthquakes are
called tsunami earthquakes.

3.1.2 Tsunami earthquakes


Tsunami earthquakes are earthquakes which generate disproportional large
tsunamis for there surface wave magnitude Ms. For most seismic earthquakes in
the Pacific, the size of the tsunami increases as the surface wave magnitude Ms
increases. However it is known that many earthquakes with small and moderate
magnitude Ms can produce large and devastating tsunamis. A well known
example is the Meiji Sanriku earthquake of 1896 (Ms = 7.2) which was barely felt
along the adjacent coastline, yet the tsunami that arrived 30 minutes afterwards
produced run-ups exceeding 30 m at some locations and killing 22,000 people.
The 1992 Nicaraguan earthquake (Ms = 7.2) is also a well know example of a
tsunami earthquake, barely felt by coastal residents, yet producing a large
destructive tsunami.
The generation mechanism of a tsunami earthquake is not well understood, at this
moment two possible causes are put forward; seismic generated submarine
landslides and slow rupturing along fault lines.
3.1.2.1 Seismic generated submarine landslides
Earthquakes can induce submarine landslides, which can amplify the initial water
displacement caused by the seismic bottom movement. This explanation has not
yet been proven conclusively, although recent research of Synolaksis (2003)
suggests that up to one third of the tsunamis is cogenerated by submarine
landslides. The reason that this co mechanism is hard to prove is the difficulty of
determining whether an underwater landslide has occurred.
Figure 3.3 shows the generation mechanism of a tsunami by a landslide.

05-01-2004

Part 1 - Chapter 3 - 3

Figure 3.3
Generation of a tsunami by a landslide

3.1.2.2 Slow rupturing along fault lines


A distinct characteristic for tsunami earthquakes is a slow rupturing process,
causing a deficiency in short period energy. Only broadband seismometers,
sensitive to low frequency waves can detect these slow earthquakes. Therefore
concealing the possibility of tsunami generation for ordinary seismometers.
Figure 3.4 illustrates the difference between a tsunami earthquake and an ordinary
one. The Hokkaido 1993 tsunami was an ordinary event. The earthquake that
generated it lasted for about 80 seconds and consisted of five large and two minor
shock waves. The earthquake was regionally felt along the Northwest coast of
Japan and produced a deadly tsunami on this coast. In contrast, the Nicaraguan
tsunami of 1992 had no distinct peak in seismic wave activity. The earthquake
was hardly felt along the nearby coast, yet it produced a killer tsunami.

Figure 3.4
Comparison of the rate of seismic force between a normal tsunamigenic
earthquake (Hokkaido) and a slow tsunamigenic earthquake (Nicaragua),
a so called tsunami earthquake

The cause of the slow rupturing can be found in the presence of an accretionary
prism developed at the interface of two crustal plates or in a plate interface filled
with soft subducted sediments (Figure 3.5).

Part 1 - Chapter 3 - 4

05-01-2004

The sediment lenses at the accreting margin can amplify the bottom movement
caused by a rupture up to a factor 2 (Synolaksis, 2003). Occasionally also slumps
in this soft layer can be induced by earthquakes, thus amplifying the initial water
displacement.
In subduction zones without large amounts of sediment, the plate interface is filled
with soft sediments. When a shallow earthquake occurs at this interface, the slip
can extend to the surface, breaking through a relatively weak plate interface filled
with sediments. the shallow depth and soft sediments on the interface are then
responsible for efficient tsunami generation and slow rupture propagation
respectively.

Figure 3.5
Accreting and non-accreting margins

3.1.2.3 Discrepancy between Ms and Mw


Regardless of the mechanism, an important diagnostic feature of tsunami
earthquakes is the difference between the surface wave magnitude Ms and the
moment magnitude Mw.
For example the two earthquakes presented in Figure 3.4 are a good illustration of
this feature: the Hokkaido event has an Ms of 6.3 and an Mw of 6.3,whereas the
Nicaragua event has an Ms of 7.2 and an Mw of 7.7. In both cases the Mw clearly
shows the tsunami generation potential of both earthquakes, however this is no
guarantee their will effectively be a generation of a tsunami, other parameters
such as fault geometry and depth of the hypocenter also play an important role.

05-01-2004

Part 1 - Chapter 3 - 5

3.2 TSUNAMI WAVES


3.2.1 Propagation
The earthquake rupture triggers a series of fast-moving, long waves of initial low
amplitude that radiate outward in a manner resembling the waves radiating when a
pebble is dropped in a pond (Figure 3.6). Part of the tsunami travels into the deep
ocean (distant tsunami) and a part travels to the nearby coast (local and regional
tsunami).
Most tsunamis generated by large earthquakes travel as wave trains. These wave
trains contain several long waves with wave lengths that often exceed 200 km in
the deep ocean. Usually one of the waves is more pronounced than the other. In
deep water even the highest wave seldom exceeds 0.5m. Their steepness is so
small that a ship out at sea does not feel a tsunami pass.

Figure 3.6
Tsunami waves radiating outward like when a pebble is dropped in a pond.

The maximum ocean depth lies between 8 to 10 km and tsunamis in the deep
ocean have typical wave lengths of hundreds of kilometers. Thus the wave length
exceeds many times the water depth (L>>d), therefore tsunami waves travel as
shallow water waves. The propagation speed of shallow water waves is solely a
function of the water depth:
c = g .h

c
g
h

(3.1)

: wave celerity (m/s)


: gravitational constant (m/s)
: water depth + water elevation (m)

Part 1 - Chapter 3 - 6

05-01-2004

Due to the propagation speed being solely a function of the water depth, tsunamis
travel at speeds of 150-250 m/s (550-900 km/h) in the deeper ocean (d>2 km),
30-90 m/s (100-300 km/h) across the continental shelf (d<1000 m) and 10-20m/s
(35-70 km/h) in front of the coast (d=50-10m). Equation (3.1) implies that
tsunami waves are amplitude dispersive, meaning that higher waves travel faster
than lower waves. This phenomenon especially becomes significant at smaller
water depths (as h = water depth + wave elevation).
The wave length of a tsunami can be expressed as a simple function of the wave
speed (c) and the period of the wave (T), for a constant depth:
L = c.T

(3.2)

L : wave length (m)


c : wave celerity (m/s)
T : period of the wave (s)
Tsunamis typically have periods of 100-2,000 seconds (1.6-33 min), referred to as
the tsunami window.

Figure 3.7
Shoaling process of a tsunami wave, clearly showing the steepening of the wave

When water depths become shallower, the tsunami waves slows down (equ.(3.1))
and becomes compressed (equ.(3.2)) causing them to grow in height (Figure 3.7).
This is called shoaling of the wave and it is a reversible process as long as the
wave does not break. This makes that the changes in the depth and in the seafloor
cause the tsunami waves to continuously evolve and change in shape as they
propagate.
On approaching land the wave height will have increased dramatically, whereas
the wave length will have significantly decreased. Despite of the steepening of the
tsunami waves, most waves hit the coast as a fast rising or falling tide (Figure
3.8). Occasionally when the wave steepness does cause the tsunami waves to
break, they hit the coast as a bore (Figure 3.9). Very rarely it can occur that
tsunami waves break on the coast.
05-01-2004

Part 1 - Chapter 3 - 7

When the tsunami acts as a rapidly rising tide, the resulting incident current
velocities are relatively low and most initial damage will result from buoyant and
hydrostatic forces and the effects of flooding. Once the tip of the wave has
arrived, the velocities will gradually increase and a lot of damage will occur by
strong currents and floating debris. When the tsunami hits the coast as a bore,
initial damage will not only occur by buoyant and hydrostatic forces but also by
impact forces of the bore front caused by high turbulence, the high water
velocities and by dragged debris in the bore front. Once the bore front has passed
further damage can be attributed to strong currents and floating debris.

Figure 3.8
A tsunami arriving at the coast as a non breaking wave. The first picture is the arrival of
the wave and the last pictures is the retreat of the wave. Notice the absence of the small
truck in front on the last picture. Also notice the darker parts on the seawall, clearly
indicating the maximum water height.

Figure 3.9
A tsunami wave arriving at the coast as a bore

Part 1 - Chapter 3 - 8

05-01-2004

3.2.2 Terminology

L
H0

:
:

h
H
Hr

:
:
:
:
:

wavelength (m)
wave height at deep
water (m)
water depth (m)
wave elevation (m)
water elevation (m)
wave height (m)
wave run-up (m)

Figure 3.10
Various terms to describe tsunami waves

The terminology used to describe tsunami waves is shown schematically in Figure


3.10. Much of this terminology is similar to the terminology used for wind waves.
However, there are some differences. The run-up height is referred to a reference
level, mostly mean sea level and not to the water level at the actual moment the
tsunami occurred. This because the run-up height is mostly measured some days
after the tsunami occurred. So one should be aware that the actual run-up height of
the tsunami is the reported run-up height Hr corrected with the tidal component.
The period of the wave is in principle defined as the time it takes for two
successive peaks to pass a fixed point. But due to the complex wave forms it is
not always convenient and two zero successive down crossings can be taken or
even a more artificial definition has to be used. The tidal recording at Port Orford
of the 1994 Kuril Island tsunami (Figure 3.11) clearly shows the problem
involved in distinguishing the wave period of a tsunami wave in a tidal recording.

Figure 3.11
Tide gauge recording at Port Orford of the 4 October 1994 Kuril Island tsunami

05-01-2004

Part 1 - Chapter 3 - 9

3.2.3 Different wave forms


Tsunamis in the open ocean are approximately sinusoidal in shape (Figure 3.12).
At this stage the waves can be described by the linear wave theory. As they
approach the coast, the waves become more peaked: the wave peak sharpens and
the wave through becomes more flattened. Mathematically the wave in this stage
can be described by the 2nd order Stokes wave theory. As the tsunami waves
approach the coast the wave through disappears and only the wave peak remains.
The wave train evolves to a succession of solitary waves.
Until recently this was the common vision among scientists; most experiments
and calculations assumed that tsunamis at the shore manifested themselves as
solitary waves. However, in the nineties scientists began to doubt this approach.
First of all because almost every big tsunami investigated, was reported by people
having a dipole form: a large through accompanied the tsunami arrival. This kind
of wave form can not be described by a solitary wave, but they can be described
as what mathematicians call N-waves. Nowadays it is believed that most of the
tsunamis manifest themselves on shore as a leading depression N-wave, meaning
that the through of the wave arrives first.

Figure 3.12
Different wave forms of tsunami waves.

One of the reasons it took so long before scientists started to accept the N-wave
form, is that the solitary wave was somewhat easier in describing mathematically
and also easier in generating during experiments. One must realize that before mid
eighties computational modeling and the computational control of experiments
was not so obvious as it is now.
Part 1 - Chapter 3 - 10

05-01-2004

Nevertheless several aspects of solitary waves and N-waves correspond quite


well; the solitary wave and the N-wave both preserve their wave form, and the
propagation is comparable. But N-waves can become steeper than solitary waves
before they break.
This makes that the older work based on solitary waves can still contain a lot of
valuable information.
Some older work can give quantitative insight in the behavior of tsunami waves
and can be used for a preliminary approach to calculate refraction (based on the
linear wave theory), behavior of tsunamis at abrupt depth transitions, wave
reflections, resonance phenomena and many more tsunami wave related
phenomena. A well-documented overview of these theories and their validity is
given by Camfield (1980).

3.2.4 Numerical calculations


Nowadays usually numerical calculations are used to describe tsunami waves. The
standard equations used to model tsunami waves are the shallow water equations
(SW).
The SW equations describe the evolution of the surface elevation and of the
depth-averaged water particle velocity of waves with large wave lengths
compared to the water depth. The equations assume that the pressure distribution
is hydrostatic everywhere, i.e. there is no variation of depth of any of the flow
variables other than the hydrostatic pressure. The SW equations are valid when
L>>d, , as a practical criterion L>20.d is used.
One general form of the SW equations is given below:
The equation of momentum can be written as (Satake, 1995):
VV
V
+ (V .)V = g C f
t
d +

d
Cf
g

:
:
:
:
:

(3.3)

depth averaged horizontal velocity vector (m/s)


water elevation (m)
water depth (m)
nondimensional friction coefficient (-)
gravitational acceleration (m/s)

The equation of continuity can be written as:


(d + )
= {(d + ) V }
t

05-01-2004

(3.4)

Part 1 - Chapter 3 - 11

The equations are coupled and nonlinear, they can be derived directly from the
Navier-Stokes equations, if the viscous effects and vertical accelerations are
neglected. A consequence of these approximations is that these equations are nondispersive, i.e. all waves travel at a speed c = g .h that only depends on the water
level h = d + , thus not of the wave length.
These equations are valid as long as the wave does not reach the point of
breaking; at this point the wave front becomes complex and can no longer be
described accurately by a hydrostatic approach. A better approach is then to
represent the waves by a non-hydrostatic approach, using Boussinesq or
Korteweg-de Vries equations or direct from Reynolds averaged Navier-Stokes
equation as presented in Stelling (2003). However, mostly tsunami waves do not
break, and even when they do the propagation of the wave can still be modeled by
the SW equations, only the representation of the wave front is then incorrect.
When tsunamis travel trough the ocean (distant tsunamis), the sphericity of the
earth must be considered. Therefore the SW equations have to be adapted to a
spherical coordinate system (Camfield, 1980). As Central America is only
exposed to local and regional tsunamis, the use of spherical coordinates is not
necessary.
An overview of the implementation of SW equations in tsunami modeling and
some exact solutions of this theory for simple bathymetries are presented in
Synolaksis (2003).

3.2.5 Breaking
As stated above tsunami waves do rarely break, because of their very long wave
lengths and there modest wave height. A breaking criterion is given by Bryant
(2001):
Br =

2H
( g tan 2 )

Br
H

:
:
:
:

(3.5)
Breaker parameter (Br>1, wave breaks)
wave height (m)
slope of the coast (rad)
radiation frequency of the waves (1/s)

However it is not stated, which wave form is considered. As until recently the
solitary wave was regarded as the tsunami wave form in coastal waters, it is most
likely that the breaker criterion is derived for solitary waves. At this moment the
form of tsunami waves are often regarded as N-waves, which can become steeper
Part 1 - Chapter 3 - 12

05-01-2004

than solitary waves before they break (Synolaksis, 2002). So, conveniently the
presented breaker criterion (equ.(3.5)) can be regarded as a lower limit for
breaking of tsunami waves, as N-waves break later than solitary waves.

3.3 TSUNAMI MODELING

Figure 3.13
Snapshots of a computer model of the 12 July 1993 tsunami in the
East Japan Sea (Byung, 2002). Notice the initial water displacement (00 hours)
and the waves radiating outward and hitting the coasts (01 hours).
After which reflection occurs (02 and 03 hours).

Scientists involved in tsunami research have always been very interested in


models, which can accurately predict the propagation of tsunamis. One of the
most important reasons was and still is providing the necessary data for tsunami
warning systems: an earthquake at A with a certain focal depth and magnitude,
will this generate a tsunami, which causes serious inundation of the coastline at
B? Another area of great interest is to model the extent of the inundation at a
specific coast.
In this study the techniques used for modeling will be used to derive the design
tsunami from deeper water to the two proposed harbor locations.
Most models start from an initial water displacement in deep water and then use a
05-01-2004

Part 1 - Chapter 3 - 13

2D hydraulic model to derive the waves to the coast. The coast is mostly modeled
as a closed boundary, causing reflection at the coast. Sometimes also the coastal
zone is modeled in order to model the extent of the inundation.
Severe tsunami events are modeled in order to efficiently calibrate the models and
to constantly improve the mathematical techniques used in tsunami engineering.
The input for these models is then provided by the derived initial water
displacement for the respective tsunami event. Calibration of the model is done by
using the limited available tsunami data; tide gauge recordings, arrival times on
shore and reported run-up heights. In the paragraphs below, a more detailed
description is given on the modeling of a tsunami event.

3.3.1 Tsunami source modeling


The input for a tsunami propagation model can be provided in two ways; by
solving the inverse tsunami problem or by using an earthquake fault model.
Solving the inverse tsunami problem: readings of tide gauges are transformed by
means of a 2-D hydraulic model to the rupture area and an initial wave profile is
derived for this area.
When the initial wave profile is known, the 2-D propagation/inundation model
can be run.
This method, solely based on tide gauges readings, gives quite some practical
problems in the implementation and obviously does not give very accurate results.
The most accurate input for a tsunami propagation/inundation model is provided
by an earthquake fault model. The tsunamis are treated as gravity waves excited
by the displacement of a large volume of water. The displacement of the oceans
surface follows the vertical displacement of the seabed if the length of the rupture
is at least three to four times the water depth (Figure 3.2). Generally the speed, at
which the displacement of the sea bottom is modeled is taken instantaneous. The
instantaneous assumption is based on the fact that tsunamis propagate at speeds up
to 200 m/s while seismic waves cause rupture to propagate at typical speeds of 2
to 3 km/s, so the seafloor motion is much faster than the speed of the tsunamis
(Synolaksis, 2002). There is little question between tsunami specialists, that the
timing of the seafloor deformation by earthquake generation tsunamis can be
taken instantaneous (Synolaksis, 2002).

Figure 3.14
Schematic view of the bottom deformation due to the earthquake rupture
and the resulting bottom deformation.

Part 1 - Chapter 3 - 14

05-01-2004

Sea floor deformation is provided by seismological source models. These


computer simulations are based on models of elastic deformation on half-spaces.
In a sense, for each earthquake, seismologists estimate what the surface
deformation would be in a material with the same elastic properties as the earths
interior, given an internal displacement of the size inferred from seismic records.
The internal displacement is estimated using the Harvard fault plane solution
(appendix 2).

3.3.2 Tsunami propagation


Tsunamis are usually treated as shallow water waves. The shallow water wave
theory is valid when the wave length is much larger than the water depth, a
practical criterion is given by L>>20.d, which applies for most tsunamis. The
governing equations are given in 3.2.
The non dimensional friction coefficient, Cf,is related to the Chzy coefficient, C,
as C 2f = g / C 2 . Typical values for Cf are given between 5x10-3 to 3x10-2. The
friction coefficient can also be given as function of the depth C f = gn 2 /(d + h)1/ 3 ,
with a typical value for the Manning roughness coefficient, n = 0.03m-1/3.
The input for the bathymetry can be obtained from ocean bathymetry monitoring
programs. As the ETOPO 5 data from NOAA, which are known to be fairly
accurate in deeper water (d>1000m). But inaccuracies tend to occur in shallower
areas (d<1000m), often the data is manually updated using nautical charts.
Especially near the shoreline necessary attention has to be given to an accurate
representation of the shoreline.
Apart from the occasional tide gauge recordings, the run-up height is the only
collected tsunami data available. Therefore one wants to know the run-up height
calculated by the model. Models based on SW equations allow to calculate fairly
accurate the run-up of a wave, provided that accurate input data is used.
Especially estimating the right bottom friction for the onshore locations is a
problem, and sufficiently detailed topography is not always readily available.
Usually the results cant be regarded as very accurate.
Therefore most tsunami models model the coast as a closed boundary and
estimate the run-up height (Hrun-up) out of the reflected wave height (Hrefl) by:
H run up = m.H refl (Imamura, 1993). The multiplification factor, m, can be estimated
taking into account the wave length and the slope of the coast. Unluckily, the
exact formula is not given by Imamura (1993) and the original paper whereto he
refers is in Japanese. For the Nicaraguan coast Imamura (1993) estimates the m to
be 2. Is obvious that a constant value of m for a coastline of a few hundred
kilometers will not provide very accurate results.

NOAA

05-01-2004

National Oceanic and Atmospheric Administration, www.noaa.gov

Part 1 - Chapter 3 - 15

3.3.3 Calibration
The results of the model are compared with the scarce available data of the
tsunami event: tide gauge registrations and reported run-up heights.
Gathering accurate data is always a problem, for recent tsunamis international
survey teams are dispatched a few weeks after the tsunami. One of their main
objectives is to collect tide gauge recordings and run-up data.
The tidal registration is the most accurate data available for calibration; it provides
information about the wave height, the waveform and period. But most tide
gauges are located in or near harbors so the complex layout of the harbor should
be taken into account.
Run-up measurements are not so accurate as small variations in topography can
cause considerable variations in run-up height (Baptista, 1993) and for a strip of
coast of a hundred kilometer usually only at ten or fifteen locations run-up heights
are measured. Therefore the reported run-up height should only be regarded as a
rough indication. This means that comparing the run-up height calculated from the
reflected wave height with the reported wave height can only give a check for the
order of the run-up height, as both the calculated and the reported run-up height
are not very accurate.
In case of inconsequences between the calculated data and the reported data, the
model has to be adapted. Adaptations can be done at the side of the tectonic
source: the length, width and the height of the bottom deformation can be adapted
keeping the seismic moment constant (as the seismic moment is a measure for
the energy of an earthquake). Another possible adaptation is refining the of the
bathymetry near the coast.

The seismic moment of an earthquake is given by :


M o = .u.L.W (Nm)

where is the rigidity around the fault, u is the average slip, L is the fault length and W is the fault width.
The seismic moment of the earthquake is a measure for the energy of the earthquake. Thus by keeping the
seismic moment constant the released energy is kept constant (Appendix 2).

Part 1 - Chapter 3 - 16

05-01-2004

3.3.4 Summary and comments


1e Input

2e Calculation

seismological
source model
Bottom
deformation

earthquake data

input

2D-mathematical model
Wave form
Height
Period
Reflected wave height

3e Feedback

comparing
Tsunami data

feedback for
callibration
bathymetry

Figure 3.15
Schematic overview of the modeling of a tsunami event.

Hydraulic models for modeling the propagation of tsunami waves are developed
by using recent well-documented tsunami events. A schematized overview of the
modeling of tsunami events is given in Figure 3.15 .The modeling of such a
tsunami event starts form the initial bottom deformation as output of a
seismological model. The initial water displacement equals the initial bottom
deformation. The initial water displacement is used as input for a hydraulic model
based on the shallow water equations. The calculated data is then compared with
available data (tide gauge recordings and run-up heights), the comparison is then
used as a feedback for calibration. Adaptations can be done of the bottom
deformations (keeping the seismic moment constant) and in a refining of the
bathymetry.
It is clear that the calculations made with such models are not extremely accurate;
the wave height is not represented with an accuracy of centimeters or a decimeter,
but more in the order of a few decimeters to half a meter.

05-01-2004

Part 1 - Chapter 3 - 17

3.4 TSUNAMI SCALES


At first the most common tsunami scales are presented, together with a short
critical reflection. Secondly some critical comments from a hydraulic engineering
point of view are formulated: do these tsunami scales carry enough information to
describe tsunami waves and to compare different tsunami events.

3.4.1 Scales
A short review will be given of the most common tsunami scales, based on work
of Satake (2002), Hatori (1995), Bryant (2001) and Abe (1983). These scales are
used to quantify tsunami events.
There are two distinct measures for describing tsunamis: intensity and magnitude.
The intensity gives an indication of the strength of the tsunami at a given location
and its magnitude is an indication for the total energy of a tsunami.
Values of historical tsunami events can be looked up in publications concerning a
singular event or in tsunami catalogues. The most complete and up to date
catalogue at this moment is the Historical Tsunami Database for the Pacific, 47
B.C. to Present, which is the joint project of IUGG/TC and ICG/ITSU . This
database can be accessed at http://tsun.sscc.ru/htdbpac.
3.4.1.1 Imamura-Ida scale, m
m = log 2 H r ,max
m
Hr,max

:
:

(3.6)
Imamura-Idas tsunami magnitude scale (-)
the maximum observed run-up height (m)

This is the traditional tsunami magnitude scale, designed by Imamura in the mid
fortys. Because of the obvious strong local influence by using the Hr,max, the
maximum observed run-up height, it is nowadays regarded as an intensity scale.

With a simple example one can illustrate very clearly the distinction between the two concepts:
intensity and magnitude.
Imagine two cars crashing against a monolithic wall with the same energy. One car being a Duck and
one car being a solid Volvo. The Duck will obviously have more damage than the Volvo. The damage
level of the car can be regarded as an intensity. The intensity of the crash with the Duck is greater than
the intensity of the crash with the Volvo, whereas the magnitude (being a measure for the energy) of
the two crashes is the same.
IUGG/TC: the tsunami commission of the International Union of Geodesy and Geophysics
ICG/ITSU: International Coordination Group for the Tsunami Warning System in the Pacific

Part 1 - Chapter 3 - 18

05-01-2004

The maximum run-up height is usually reported along the coast near the tsunami
source.
It is especially convenient for older tsunamis, of which no instrumental records
exist, and, of which the evidence can only be found in eye-witness reports, news
paper articles or old stories.
In order to get an idea of the physical meaning of the Imamura-Ida scale, the
relationship between the Imamura-Ida scale (m) and the maximum run-up (Hr,max)
is given in Table 3.1 and. Figure 3.16.

Table 3.1
Imamura-Ida scale, m

Imamura-Ida scale, m
25
20
15
10
5

-1

-2

0
-3

Mean run-up height


(m)
0.1
0.2
0.35
0.7
1.4
2.8
4.0
5.6
8.0
11.3
16.0

maximum run-up
height (m)

Tsunami
intensity, i
-3.0
-2.0
-1.0
0.0
1.0
2.0
2.5
3.0
3.5
4.0
4.5

m (-)

Figure 3.16
Imamura-Ida scale, m

3.4.1.2 Tsunami intensity scale, i


Soloviev (1970) was the first to point out the inappropriateness of the term
tsunami magnitude as used in the Imamura-Ida scale:
The grades of the Imamura-Ida scale must be designated as the intensity of the
tsunami and not its magnitude. This is because the latter value must characterize
dynamically the processes in the source of the phenomena and the first one must
characterize it at some observational point
Soloviev also pointed out that it is more appropriate to use the mean run-up
(Hr,mean) at one stroke of coast in stead of the maximum run-up (Hr,max). The
differences between mean and maximum run-up height can mostly be attributed to
differences in bathymetry and topography of the regarded strip of beach.
He then defined the tsunami intensity scale i as :
i = log 2

2.H r ,mean

(3.7)

i
: tsunami intensity (-)
Hr,mean : the mean run-up height on the coast nearest to the source (m)
05-01-2004

Part 1 - Chapter 3 - 19

For a tsunami event the maximum intensity (i) on the coast nearest to the source is
used to quantify the tsunami source. Soloviev did not state how long the regarded
strip of beach should be, so the definition of the mean run-up height is still
somewhat ambiguous.
Comparison of Eqs (3.6) and (3.7) suggest that the mean tsunami run-up height is
given as 1/2 times the maximum height (Satake, 2002).
In order to get an idea of the physical meaning of the tsunami intensity scale, the
relationship between the tsunami intensity (i) and the mean run-up (Hr,mean) is
given in Table 3.2 and Figure 3.17.

Table 3.2
Tsunami intensity scale, i

tsunami intensity scale, i


20
15
10
5

-1

-2

0
-3

Mean run-up height


(m)
0.1
0.2
0.35
0.7
1.4
2.8
4.0
5.6
8.0
11.3
16.0

mean run-up height (m)

Tsunami
intensity, i
-3.0
-2.0
-1.0
0.0
1.0
2.0
2.5
3.0
3.5
4.0
4.5

i (-)

Figure 3.17
Tsunami intensity, i

3.4.1.3 Tsunami magnitude, Mt


Both the Imamura-Ida scale, m, and the tsunami intensity scale, i, are based on the
run-up height of the tsunami. Gathering accurate run-up data is not always easy.
Most data were not recorded for scientific analysis. Besides, tsunamis propagating
towards the coast undergo a complicated combination of dynamic processes
strongly related to the local site conditions, which can not be completely
understood without intensive study and modeling. As a result a quantitative
analysis is inevitably accompanied by a large scattering of data.
To overcome this problem an other scale is needed.
Abe (1983) defined a scale based on the maximum tsunami wave amplitude, H, as
recorded on tidal gauges. Its original intended use was to estimate the moment
magnitude, Mw, of historical large earthquakes from available tsunami data.
Therefore Abe used the good correlation between Mw and logH.
Part 1 - Chapter 3 - 20

05-01-2004

The fundamental definition of Mt is given by:


(3.8)

M t = log H + B

Where H is the maximum tsunami wave amplitude measured by tide gauges in


meters and B is a correction term, defined by requiring equalization of the tsunami
magnitude, with the moment magnitude of the seismic event (Mt = Mw) on
average for the calibration set of the specific region.
A distinction is made between regional tsunamis and trans-pacific tsunamis.
Regional tsunamis are tsunamis with a traveled distance between 100 km and
3 500 km, whereas trans-pacific tsunamis have travel distances greater than
3 500 km.
Regional tsunamis (100km < < 3500km)
(formula calibrated for the Japanese region)
(3.9)

M t = log H + log + 5.8

Mt : tsunami magnitude (-)


H : the maximum recorded amplitude on tide gauges (m)
: the distance from the source to the tide gauge (km)
Trans-pacific tsunamis ( > 3500km)
M t = log H + C + 9.1

(3.10)

Mt : tsunami magnitude (-)


H : the maximum recorded amplitude on tide gauge meters (m)
C : a distance factor (-)
The factor C is introduced while calibrating the formula, in order to achieve Mt =
Mw.
It is a small correction factor depending on the combination of the source and the
observation points. The correction factor may vary from 0.6 to 0.5. Table 3.3
lists values of C as determined by Abe (1983).

05-01-2004

Part 1 - Chapter 3 - 21

C
Honolulu

Hilo

California

Japan

Aleutian

+0.2

-0.6

+0.2

0.0

+0.2

Alaska, Aleutian

+0.5

0.0

+0.2

+0.3

Kamchatka, Kurile, Japan

0.0

-0.4

+0.1

-0.2

-0.2

+0.2

-0.3

+0.2

0.0

0.0

Source Region
Peru, Chile

whole region

Table 3.3
Values of C for different source regions and registration regions,
as determined by Abe 1983

At present, few values beyond these given in Table 3.3 have been set, so it is not
possible to give regressed values of Mt for all tsunami events .
The above formula (3.10) was calibrated with the moment magnitude, Mw,
(Bryant, 2001) of earthquakes. Therefore, the average tsunami magnitude, Mt, for
a coastline is in the same order as the Mw-value of the source earthquake. For
most earthquake generated tsunamis the correspondence between Mt and Mw does
not differ more than 0.2, nevertheless Mt and Mw can still differ considerably. The
most important disturbance factors for Mt not corresponding with Mw are:
- the local bathymetry: this can seriously affect the recorded wave height H,
although averaging Mt, as calculated from the different tidal recordings, almost
completely eliminates this effect;
- underwater landslides, triggered by earthquakes can amplify the initial water
displacement, thus giving a higher recorded wave height;
- accredited sediment wedges in the generation area can amplify the initial water
displacement, thus giving a higher recorded wave height.

Part 1 - Chapter 3 - 22

05-01-2004

3.4.2 Comments from a hydraulic engineering point of view


A tsunami wave train consists of a small number of waves, usually one of these
waves is more pronounced. This dominant wave can be used to schematize the
tsunami load. For describing the dominant wave essential parameters are the wave
height, the wave length and the form of the wave (not broken, bore or breaking).
Do the three presented scales carry sufficient information to derive the
dominant tsunami wave?
The Imamura-Ida scale, m, and the tsunami intensity scale, i, are both based on the
recorded run-up. The run-up of a wave is strongly influenced by the local
topography and the bottom friction (influenced by vegetation and buildings). In
catalogues listing tsunami events, this information is not available, it is therefore
considered that m and i do not carry sufficient information to link to the original
tsunami waves.
The tsunami magnitude scale, Mt, is based on the recorded wave height on a tide
gauge and the distance of the source. However it does not give information about
the water depth at the recording side, which is important for the shoaling.
Moreover the Mt value for one tsunami event is calculated by averaging the
calculated Mt values for the different tidal gauges. This averaging filters out the
influence of the local bathymetry and geometry (e.g. amplification by funnel
shaped bays) in order to give good correlation with the earthquake moment
magnitude Mw. But it also makes it impossible for linking the averaged Mt value
with the wave height at one particular depth.
Also deriving information about the wave length and the wave form (not broken,
bore or breaking) is not possible.
Can the three presented scales be used to compare different tsunami events?
The three presented scales were designed to quantify tsunami events, but not all
scales are equally consistent.
The Imamura-Ida scale, m, is based on the maximum run-up. Therefore this scale
is very sensitive for local bathymetry and topography, making it unsuitable to
compare different tsunami events.
The tsunami intensity scale, i, is based on the mean run-up on the nearest coast.
Hereby preventing a strong influence of very local bathymetry and topography

The three parameters wave length, wave height and form of the wave (not broken, bore or breaking), are
not sufficient for an exact representation of the tsunami wave. But as the exact form of a tsunami wave is
hard to derive exactly and because each tsunami event is different, it is believed that these parameters are
sufficient for representing a design load (Kamel 1970). Because of the inaccuracy involved a large
sensitivity analysis should be done.

05-01-2004

Part 1 - Chapter 3 - 23

(e.g. the influence of a funnel shaped bay). Obviously the bathymetry and
topography of the regarded coast does influence the mean run-up, making it
difficult to compare different tsunami events.
However tsunami events occurring in a region with coasts with comparable
bathymetry and topography, can be compared with each other, making it possible
to estimate a reoccurrence period of the tsunami events in that region.
The tsunami magnitude scale, Mt, is based on the recorded wave heights on tide
gauges and the distance to the source. The Mt scale has a strong correlation with
the moment magnitude, Mw, of the generating earthquake.
These factors make the Mt scale the best base for a comparison of different
tsunami events. However an Mt value is not available for all tsunami events.
Are there alternatives for deriving tsunami waves?
Since the early nineties a deepwater tsunami registration program has been set up.
These recordings register the entire tsunami profile. However the registration
program only covers a small part of the Pacific. Consequently these detailed
measurements are not available for all tsunami events.
At this moment the only possible way to get accurate information about the
tsunami wave profile is by modeling the tsunami event right from the initial
bottom deformation, caused by the earthquake.

Part 1 - Chapter 3 - 24

05-01-2004

4 DETERMINING THE CONCEPT OF THE DESIGN


TSUNAMI

In civil engineering the common method of defining a design load is by a


statistical approach. For example defining a design water level for a dike. The
design water level should be higher than the already observed highest water level.
When an extensive dataset of observed water levels is available, a distribution
function is fit to this dataset. And a design water level, for a certain return period,
can then easily be determined. However in reviewed tsunami literature no
examples are found of a statistically defined design tsunami. Mostly for designing
a structure a large historical event is taken and the wave height and the duration
are slightly increased (e.g. Kamel, 1970 - Hitachi, 1994)
At first it will be investigated if a design tsunami can be determined statistically,
with the available tsunami data for Central America from Chapter 2. If this proves
not to be possible, the largest historical event will be taken and slightly amplified.

4.1 ANALYZING THE TSUNAMI DATA FOR THE PACIFIC SIDE


OF CENTRAL AMERICA
In this paragraph it will be examined if the dataset (Table 2.1 and Figure 2.1)
contains enough information for a statistical derivation of a design tsunami.

4.1.1 Homogenizing the dataset


First a homogeneous dataset has to be obtained so that the different tsunami
events can be compared.
Two criteria can be distinguished for the homogenization of this tsunami dataset:
the generation mechanism and the location of the generation.

05-01-2004

Part 1 - Chapter 4 - 1

date

time

generation point

Year

Mo

Da

Hr

Mn

1906

31

15

36

Sec

1907

15

1928

22

17

9.6

1928

17

19

31.8

16.65

earthquake data

tsunami data

generation area

Lat

Long

Dep

Ms

Mw

Mt

Int

Hr,max

-81.5

25

8.6

8.7

8.7

1000

SAM Columbia-Ecuador:TUMACO

TR

Source Region

16.7

-99.2

25

8.2

7.8

CAM Mexico

16.84

-96.02

15

7.5

7.4

-96.5

7.8

7.9

8.1

CAM MEXICO
MEXICO: NEAR COAST OF
GUERRER

2.5

CAM

CAM S.CALIFORNIA

1930

31

40

36.6

33.9

-118.6

5.2

-0.5

6.1

1932

10

36

54.2

19.46

-104.17

25

8.1

7.6

8.2

-0.5

0.7

CAM Central Mexico,Jalisco

1932

18

10

12

17.8

19.53

-103.72

71

7.8

7.5

7.8

-2

0.1

CAM Central Mexico, Jalisco

1932

22

12

59

27.5

19.09

-104.41

6.9

1.5

10

CAM Central Mexico

1934

18

36

27.8

8.01

-82.56

25

7.7

7.6

CAM Costa Rica-Panama

10

1941

12

20

47

3.3

8.61

-83.18

32

7.5

7.3

-3

0.1

CAM Central America

11

1941

12

8.5

-84

33

-3

0.1

CAM Costa Rica-Panama

12

1948

12

22

21.6

-106.7

33

CAM Mexico:MARIA MADRE ISLAND

13

1950

10

16

10.4

-85.2

60

7.7

7.7

7.8

1.5
-2

0.1

CAM Nicaragua,Costa Rica-Panama

14

1950

10

23

16

13

14.3

-91.8

30

7.1

7.5

7.5

-1

0.2

CAM Guatemala-Nicaragua

15

1950

12

14

14

15

16.3

-98.2

50

7.5

7.1

-1

0.3

CAM Acapulco, S.Mexico

16

1951

24

13

-87.5

100

17

1957

28

40

9.8

16.88

-99.29

36

7.9

18

1958

19

14

25

0.99

-79.48

20

7.3

19

1962

12

11

40

15.3

8.09

-82.68

19

6.8

20

1962

11

14

11

56.8

17.14

-99.7

46

7.2

21

1962

19

14

58

13.1

16.99

-99.69

16

7.1

22

1965

23

19

46

1.5

16.17

-95.85

10

7.8

23

1973

30

21

13.5

18.45

-102.96

37

7.5

7.6

24

1979

14

11

15

17.77

-101.23

24

7.6

7.6

25

1979

12

12

59

4.6

1.6

-79.36

24

7.7

8.1

26

1981

10

25

22

15.8

18.12

-102

21

7.3

7.7

10

1.5

2.6

1.5

CAM Potosi, Honduras

68

CAM S.Mexico:GUERRERO

20

SAM Colombia-Ecuador

-3

0.1

CAM Costa Rica-Panama

0.8

CAM S.Mexico

30

CAM S.Mexico

7.5

-1

0.4

CAM MEXICO

-3

0.1

56

CAM S.Mexico:FARIAS,TECOMAN

-1.1

0.4

CAM S.Mexico:GUERRERO

2.5

600

SAM Colombia-Ecuador

7.2

-3

0.1

CAM MEXICO

27

1985

19

13

17

49.7

18.46

-102.37

21

8.1

1.5

28

1985

21

37

13.5

17.83

-101.62

18

7.5

7.5

1.2

29

1988

30

15

42

32.6

-117.3

30

1992

16

2.7

11.72

-87.39

45

7.2

7.7

2.8

10.7

170

CAM Nicaragua

31

1994

17

12

30

55.5

34.16

-118.57

14

6.8

6.7

-3

0.1

CAM Southern California

32

1995

14

14

33.9

16.84

-98.61

29

7.2

7.3

-1

0.42

SAM Oaxaca, Mexico

33

1995

10

15

35

54.1

19.05

-104.21

26

7.3

CAM Jalisco, Mexico

34

1996

25

16.6

15.93

-98.1

17

6.9

7.1

-3

0.12

CAM Oaxaca, Mexico

35

1999

12

18

17.62

-101.6

2.5

CAM Guerero, Mexico

7.9

CAM MEXICO: MICHOACAN: MEXICO CI

CAM MEXIC0: SW COAST: MEXICO CIT

CAM San Diego, S. California

Table 4.1
Databank for the tsunamis at the Pacific side of Central America from 1900 to 2002.
the crossed out events are removed to homogenize the dataset.

Part 1 - Chapter 4 - 2

05-01-2004

Figure 4.1
The epicentral locations of the tsunamigenic earthquakes.
The crossed out events are removed to homogenize the dataset.
Also the tectonic setting of Central America is given.
(MAT: Middle American Trench, PFZ: Panama Fracture Zone)

4.1.1.1 The generation mechanism


In general earthquakes are distinguished as the predominant cause of tsunami
generation. The dataset for Central America confirms this idea; 32 on a total of 35
tsunamis are generated by earthquakes. Therefore the dataset will be limited to
earthquake generated tsunamis.
The following tsunamis events are removed (Table 4.1):
- the 1930 tsunami in Southern California (5), caused by a landslide
- the 1988 tsunami in Southern California (29), caused by meteorological
circumstances
- the 1999 tsunami in Mexico (35), with an unknown cause (this means the cause
was certainly not an earthquake, because even the smallest earthquake is
always detected by sensitive seismographs)

05-01-2004

Part 1 - Chapter 4 - 3

4.1.1.2 The location of the generation


The predominant cause of tsunami generation are earthquakes. Earthquakes are
generated at plate boundaries. The type of the plate boundary affects the
bathymetry and the occurring earthquakes (Appendix 2). Therefore the generation
region will be limited with regard to plate tectonics. A similar approach is used by
Hatori (1995).
Tectonic setting
The tectonic setting of the Pacific side of Central America (Figure 4.1) is given by
the interaction of the Cocos, Caribbean and Nazca plates (fig.1). The Cocos Plate
subducts under the Caribbean Plate along the Middle American trench; from
Mexico to Central Costa Rica. The limit between the Cocos and Nazca plates is
the Panama Fracture Zone (PFZ). The PFZ is composed of north-south trending
faults located in front of the Pacific Coast of Costa Rica and Panama. The
boundary between the Caribbean Plate and the Nazca Plate is still ambiguous,
some authors consider it as a subduction zone and others as a strike-slip fault.
Most earthquakes are related to the Cocos-Caribbean subduction zone, defined by
the Middle American trench. The Middle-American Trench is from 4 to 5 km
deep (Shah et. al., 1975) and extends, approximately 130 km off the Central
American Pacific coast (Montero, 1990), from Mexico to Costa Rica.
Research of Fernandez (2000) reveals that 93% of the total released moment, was
released along the Middle American Trench. The earthquakes can be separated in
two populations, shallow depth (0-30km) and intermediate depth (40-200km)
earthquakes. There are no earthquakes deeper than 200km.
Because shallow and intermediate earthquakes along subduction zones, have the
biggest possibility of causing vertical displacement of the ocean bottom, these
earthquakes are mostly held responsible for tsunami generation. Therefore the
dataset will be limited to earthquakes generated along the Middle American
Trench
The Panama fracture zone is an ocean ridge. Ocean ridges primarily generate
earthquakes without predominant vertical displacement, which is necessary for
tsunami generation.
Therefore it is somewhat ambiguous if the earthquakes of 1934 (9) and 1962 (19)
in Costa Rica-Panama are generated on the middle American trench or on the
Panama fracture zone, these tsunami events are kept in dataset.
The following tsunami generating earthquakes are removed (Table 4.1):
- the 1906 earthquake in Colombia-Ecuador (1)
- the 1930 earthquake in Southern California (5)
- the 1958 earthquake in Colombia-Ecuador (18)
- the 1979 earthquake in Colombia-Ecuador (25)
- the 1988 earthquake in San Diego (29)
- the 1994 earthquake in Southern California (31)
Part 1 - Chapter 4 - 4

05-01-2004

Comments
It is clear that tsunamis generated in Southern California (5, 29 and 31) do not
affect the coast bordered by the Middle American Trench. But the tsunamis
generated at the Columbian subduction zone can obviously affect this coast. This
were the 1906 Columbia-Ecuador tsunami (1) with intensity 3, the 1958
Columbia-Ecuador tsunami (18) with intensity 1.5 and the 1979 ColumbiaEcuador tsunami with intensity 2.5. These tsunamis are so called regional
tsunamis as they are generated at a certain distance of the Pacific coast of Central
America. This means that the large intensities due to the nature of the intensity
scale (defined for the mean run-up on a nearest coast, see 3.4) have no
significance for the Central American coast. They 1906 tsunami was observed
along the entire coast of Central America, but there is no specific reference made
of large damage along this coast due to this specific tsunami (Fernadez, 2000).
Therefore for statistical calculations these three tsunamis are still kept out of the
dataset.

05-01-2004

Part 1 - Chapter 4 - 5

4.1.1.3 Resulting dataset


By restricting the generation cause to earthquakes and the generation area to the
Middle American Trench, a final dataset is obtained with 28 tsunami events for
the period 1900-2002.
date

time
Mn

generation point
Sec

Lat

17

9.6

19

31.8

16.65

earthquake data

Year

Mo

Da

Hr

Long

Dep

Ms

Mw

1907

15

1928

22

16.7

-99.2

25

8.2

7.8

16.84

-96.02

15

7.5

7.4

1928

17

-96.5

7.8

7.9

1932

10

36

54.2

19.46

-104.17

25

8.1

1932

18

10

12

17.8

19.53

-103.72

71

7.8

1932

22

12

59

27.5

19.09

-104.41

6.9

1934

18

36

27.8

8.01

-82.56

25

7.7

7.6

10

1941

12

20

47

3.3

8.61

-83.18

32

7.5

7.3

11

1941

12

8.5

-84

33

tsunami data
Mt

Int

Hr,max

2.5

generation area
F

TR

Source Region

CAM Mexico

CAM MEXICO

CAM

MEXICO: NEAR COAST OF


GUERRER

8.1

7.6

8.2

-0.5

0.7

CAM Central Mexico,Jalisco

7.5

7.8

-2

0.1

CAM Central Mexico, Jalisco

1.5

10

CAM Central Mexico

-3

0.1

-3

CAM Costa Rica-Panama

CAM Central America

0.1

CAM Costa Rica-Panama

12

1948

12

22

21.6

-106.7

33

CAM Mexico:MARIA MADRE ISLAND

13

1950

10

16

10.4

-85.2

60

7.7

7.7

7.8

-2

0.1

CAM Nicaragua,Costa Rica-Panama

14

1950

10

23

16

13

14.3

-91.8

30

7.1

7.5

7.5

-1

0.2

CAM Guatemala-Nicaragua

15

1950

12

14

14

15

16.3

-98.2

50

7.5

7.1

-1

0.3

16

1951

24

13

-87.5

100

10

17

1957

28

40

9.8

16.88

-99.29

36

7.9

1.5

2.6

19

1962

12

11

40

15.3

8.09

-82.68

19

6.8

-3

0.1

0.8

CAM S.Mexico

30

CAM S.Mexico

CAM MEXICO

CAM S.Mexico:FARIAS,TECOMAN

20

1962

11

14

11

56.8

17.14

-99.7

46

7.2

21

1962

19

14

58

13.1

16.99

-99.69

16

7.1

1.5

22

1965

23

19

46

1.5

16.17

-95.85

10

7.8

7.5

23

1973

30

21

13.5

18.45

-102.96

37

7.5

7.6

24

1979

14

11

15

17.77

-101.23

24

7.6

7.6

-1.1

26

1981

10

25

22

15.8

18.12

-102

21

7.3

7.2

-3

27

1985

19

13

17

49.7

18.46

-102.37

21

8.1

1.5

28

1985

21

37

13.5

17.83

-101.62

18

7.5

7.5

30

1992

16

2.7

11.72

-87.39

45

7.2

7.7

32

1995

14

14

33.9

16.84

-98.61

29

7.2

33

1995

10

15

35

54.1

19.05

-104.21

26

34

1996

25

16.6

15.93

-98.1

17

68

-1

0.4

-3

0.1

56

0.4

0.1

N
0

CAM Acapulco, S.Mexico

CAM Potosi, Honduras

CAM S.Mexico:GUERRERO

CAM Costa Rica-Panama

CAM S.Mexico:GUERRERO

CAM MEXICO

CAM MEXICO: MICHOACAN: MEXICO CI

1.2

CAM MEXIC0: SW COAST: MEXICO CIT

2.8

10.7

170

CAM Nicaragua

7.3

-1

0.42

SAM Oaxaca, Mexico

7.3

CAM Jalisco, Mexico

6.9

7.1

-3

0.12

CAM Oaxaca, Mexico

7.9

Table 4.2
Homogenized dataset

Part 1 - Chapter 4 - 6

05-01-2004

Figure 4.2
Homogenized dataset.

4.1.2 Do the available data for these events contain enough


information for deriving a wave or a wave train?
A homogenous dataset was obtained, which enables us to make statistical
calculations. It is now possible to take a parameter, fit a distribution to it and
extrapolate it for a certain return period. But the question is whether these
available parameters can be coupled to wave characteristics? This is examined
below. First it is described how the wave can be schematized. Afterwards it is
described what can be concluded out of the available tsunami data and seismic
parameters.
4.1.2.1 Tsunami wave vs. tsunami wave train
A tsunami is a wave train consisting of a small number of waves, usually less the
10, before reflection occurs. Usually one of the waves is a lot more pronounced
than the others (Murty, 1977).
In literature usually one wave is used to characterize the tsunami load, as well in
theoretical reflections as in experiments (e.g. Kamel, 1970 Ramsden, 1996
Silva, 2000 Tanimoto, 1981).
Therefore for design purpose it is believed to be sufficient to schematize the
05-01-2004

Part 1 - Chapter 4 - 7

tsunami to this one dominant wave, characterized by its wave height, length and
the form of the wave (not broken, bore or breaking) (3.4.2). So it is sufficient to
derive this one dominant wave.
4.1.2.2 Tsunami data
The relevant tsunami data available in the dataset are the maximum run-up
(Hr,max), the tsunami intensity (i) and for some tsunamis the tsunami magnitude
(Mt). in paragraph 3.4 an extensive overview of the scales is given. It is concluded
that nor the maximum run-up (Hr,max), nor the tsunami intensity (i), nor the
tsunami magnitude (Mt) can be used to derive a tsunami wave.
The tsunami intensity (i) can be used to compare different tsunami events as the
coast near to the Middle American Trench is assumed to have a similar
bathymetry and topography along the entire coast. The best base for a comparison
is the tsunami magnitude scale (Mt), but the Mt values are only available for 8
tsunami events.
4.1.2.3 Earthquake scales
The available earthquake data are the depth of the hypocenter (Dep), the surface
wave magnitude (Ms) and the moment magnitude (Mw). This data cant be coupled
directly to wave parameters and they contain insufficient information for deriving
an ocean bottom deformation.
4.1.2.4 Conclusion
The effect of a tsunami wave train can be represented by one dominant wave,
characterized by its wave height, wave length and the form of the wave (not
broken, bore or breaking).
The data available of the tsunami events contains insufficient information to link
to wave characteristics. Therefore a statistical approach for deriving a design
tsunami out of the available data is not possible.
As a statistical approach proves not to be possible, the largest tsunami event in the
Central American region will be identified. The largest event, will then be taken
as a basis for a design tsunami; the wave height and the period will be increased
with 20%.

The conclusions concerning the possibility of linking the tsunami data and the seismic data to tsunami
wave parameters, were the result of an extensive investigation. It was concluded that linking the limited
available data to wave parameters is not possible. However due to the importance of this conclusion for
this study, the expert opinion was asked of some tsunami researchers. They were contacted by mail and
there replies confirm my conclusions. These mails are shown in Appendix 4.

Part 1 - Chapter 4 - 8

05-01-2004

4.2 DETERMINING THE LARGEST TSUNAMI EVENT AT THE


PACIFIC COAST OF CENTRAL AMERICA
The intensity of the tsunamis at the Pacific side of Central America range from 3
to +3.
Looking at the location of tsumanigenic earthquakes it can be seen that all
countries have been hit by tsunamis. Fortunately, most of these tsunamis were
small, causing little damage.
Partially this can be explained by the location of the epicenters of the
tsunamigenic earthquakes. Many of the epicenters are located close to the coast or
even inland. This results in a bottom deformation lying partially on land, so
reducing the generated tsunami (Fernandez, 2000).
Fernandez (2000) also concludes that 45% of the large earthquakes (Ms7.0)
located off-shore or in the vicinity of the Pacific coastline of Central America
generate tsunamis.
Remarks
For the tsunami events of 1928 in Mexico (3) and may 1962 in Mexico (21) no
intensity i value is available. So these two events are removed. Hereby the dataset
is limited to 26 tsunamis.

05-01-2004

Part 1 - Chapter 4 - 9

date

time

generation point
Sec

earthquake data

Year

Mo

Da

Hr

Mn

1907

15

1928

17

19

1932

10

36

54.2

19.46

-104.17

25

8.1

1932

18

10

12

17.8

19.53

-103.72

71

7.8

1932

22

12

59

27.5

19.09

-104.41

6.9

1934

18

36

27.8

8.01

-82.56

25

7.7

7.6

10

1941

12

20

47

3.3

8.61

-83.18

32

7.5

7.3

11

1941

12

8.5

-84

33

31.8

Lat

Long

Dep

Ms

Mw

16.7

-99.2

25

8.2

7.8

16.65

-96.5

7.8

7.9

tsunami data
Mt

generation area

Int

Hr,max

C
T

CAM Mexico

TR

Source Region

8.1

2.5

CAM

7.6

8.2

-0.5

0.7

CAM Central Mexico,Jalisco

7.5

7.8

-2

0.1

CAM Central Mexico, Jalisco

1.5

10

CAM Central Mexico

-3

0.1

-3

MEXICO: NEAR COAST OF


GUERRER

CAM Costa Rica-Panama

CAM Central America

0.1

CAM Costa Rica-Panama

12

1948

12

22

21.6

-106.7

33

CAM Mexico:MARIA MADRE ISLAND

13

1950

10

16

10.4

-85.2

60

7.7

7.7

7.8

1.5
-2

0.1

CAM Nicaragua,Costa Rica-Panama

14

1950

10

23

16

13

14.3

-91.8

30

7.1

7.5

7.5

-1

0.2

CAM Guatemala-Nicaragua

15

1950

12

14

14

15

16.3

-98.2

50

7.5

7.1

-1

0.3

16

1951

24

13

-87.5

100

10

17

1957

28

40

9.8

16.88

-99.29

36

7.9

1.5

2.6

19

1962

12

11

40

15.3

8.09

-82.68

19

6.8

-3

0.1

20

1962

11

14

11

56.8

17.14

-99.7

46

7.2

0.8

22

1965

23

19

46

1.5

16.17

-95.85

10

7.8

7.5

-1

0.4

23

1973

30

21

13.5

18.45

-102.96

37

7.5

7.6

-3

0.1

56

24

1979

14

11

15

17.77

-101.23

24

7.6

7.6

-1.1

0.4

26

1981

10

25

22

15.8

18.12

-102

21

7.3

7.2

-3

0.1

27

1985

19

13

17

49.7

18.46

-102.37

21

8.1

1.5

28

1985

21

37

13.5

17.83

-101.62

18

7.5

7.5

1.2

7.9

68

CAM Acapulco, S.Mexico

CAM Potosi, Honduras

CAM S.Mexico:GUERRERO

CAM Costa Rica-Panama

CAM S.Mexico

CAM MEXICO

CAM S.Mexico:FARIAS,TECOMAN

CAM S.Mexico:GUERRERO

CAM MEXICO

CAM MEXICO: MICHOACAN: MEXICO CI

CAM MEXIC0: SW COAST: MEXICO CIT


CAM Nicaragua

30

1992

16

2.7

11.72

-87.39

45

7.2

7.7

2.8

10.7

170

32

1995

14

14

33.9

16.84

-98.61

29

7.2

7.3

-1

0.42

SAM Oaxaca, Mexico

33

1995

10

15

35

54.1

19.05

-104.21

26

7.3

CAM Jalisco, Mexico

34

1996

25

16.6

15.93

-98.1

17

6.9

7.1

-3

0.12

CAM Oaxaca, Mexico

Table 4.3
Final dataset

Part 1 - Chapter 4 - 10

05-01-2004

Figure 4.3
Final dataset

4.2.1 Occurrence
The occurrence of tsunamis with a certain intensity is shown in Table 4.4. Also
the mean run-up heights (Hr,mean) corresponding with the tsunami intensity (i)
(equ 3.7) are shown.

05-01-2004

Part 1 - Chapter 4 - 11

Hr,mean (m)

4
3,5
3
2,5
2
1,5
1
0,5
0
-0,5
-1
-1,5
-2
-2,5
-3

11,3
8,0
5,7
4,0
2,8
2,0
1,4
1,0
0,7
0,5
0,4
0,3
0,2
0,1
0,1

Number of
tsunamis
0
0
1
0
0
4
3
1
2
1
5
0
3
0
6

Table 4.4
Tsunami intensity versus occurrence.
(the intensities are rounded of to a half)

T su n a m i D ata

num be r

8
6
4
2
0
-3

-2

-1

Figure 4.4
Tsunami intensity versus occurrence

Probably the observation of the small intensity tsunamis (i<0) is an


underestimation of the really occurred events. Because the mean run-up height
(Table 4.4) is so small for these events, they can remain unnoticed. Nevertheless
the big tsunami events will be noticed and reported. Therefore Figure 4.4 may not
be correct with regard for the representation of the entire distribution of
occurrence of a tsunami, but the biggest event can be clearly distinguished. The
Biggest event that occurred at the coast along the Middle American Trench from
1900 to 2002, is the 1992 Nicaraguan tsunami, with an intensity of 2.8. Hatori
(1995) and Fernandez (2000) confirm that the 1992 Nicaraguan tsunami was the
biggest tsunami event in Central America that occurred the past century.
Part 1 - Chapter 4 - 12

05-01-2004

Figure 4.5
Tsunami exceedance probability versus intensity for tsunamis with i 0 for the
Central American region from 1900 to 2002

In Figure 4.5 the tsunamis with an intensity i 0 are plotted against their
exceedance probability. This gives for the 1992 Nicaraguan tsunami (i = 3) a
return period between 85 and 850 years. The best fit distribution function gives an
return period of 120 years for the coast bordering the Middle American Trench.
As only local tsunamis occurred in this region, only a small part of the coast is
affected. In the case of the Nicaraguan tsunami only 250 km of coast was
seriously affected (Bryant, 2001) of the total of 4,000 km along the Middle
American Trench. This is a normal value for tsunamis of such a large intensity.

Figure 4.6
A tsunamigenic earthquake, similar to the 1992 tsunami event, can only affect a
certain location at the coast as the epicenter lies in the zone in front of the location.
As for the 1992 Nicaraguan earthquake the coastal region affected was 250km, the
length of the regarded strip in which the epicenter has to lie is 2x125 km.

05-01-2004

Part 1 - Chapter 4 - 13

To get an impression of the occurrence at one location along the coast, the Middle
American Trench can be divided into strips of 250 km (Figure 4.6). The chance
for a tsunamigenic earthquake, as in Nicaragua1992, can be taken uniformly
distributed over the 16 cells of the Middle American Trench. Thus the chance that
the once in a 120 year tsunami affects one exact location, i.e. occurs in one of the
16 cells, is once in 1920 years. Taking into account the spreading, then the return
period for one location lies between 1360 and 13600 years
Clearly this is only a rough estimation of the return period. The tsumanigenic
earthquakes probably are not uniformly distributed along the Middle American
Trench (Figure 4.3), most tsunamigenic earthquakes seem to occur in Mexico. But
in order to give account for such heterogeneities a dataset for a longer period and
more detailed insight in geological processes is needed.
Nevertheless this simple calculation clearly illustrates the infrequency of large
tsunami events at the Pacific coast of Central America.

4.2.2 Compared to the entire Pacific Region


Compared to the entire Pacific region, one could wonder if this is what should be
expected for Central America .Looking at the entire Pacific region (Table 4.5) for
the period of 1900 to 2002 only 23 tsunamigenic earthquakes with an intensity i
2.8 occurred. So Table 4.5 clearly shows that large tsunami events are infrequent
phenomena. Most of these large tsunamis only affected a few hundred kilometers
of the nearest coast. Figure 4.7 shows that in the North Western part of the Pacific
(Japan, Kamchatka and Alaska) more than half of the large tsunami events were
generated. Therefore this is the most important generation area for large tsunamis.
This has two reasons: first of all this region consists of very active seismic
subduction zones and secondly very local amplification effects can occur. These
very local amplification effects are attributed to local bathymetry such as bays or
straits or large Archipelos (Ida, 1981). Because these two conditions are fulfilled
for the North Western Pacific, most of the large tsunamis occur in this region.
Figure 4.7 is consistent with Figure 2.1, where the percentages of tsunamigenic
earthquakes for all the regions in the Pacific are displayed. Also Figure 2.1 clearly
shows that Central America is a rather modest region for tsunami generation
(6.5%).
The maximum run-up heights of tsunami events comparable to the Nicaraguan
tsunami (i ~ 3), lie mostly between 10 and 15m. Sometimes very local bathymetry
effects can seriously increase the maximum occurred run-up, e.g. the 1964
tsunami in the gulf of Alaska (Hrmax=67m). The enormous maximum run-up
height of the 1958 tsunami in South East Alaska is caused by a massive rockslide
triggered by an earthquake in a confined bay (Bryant, 2001).
Figure 4.7 and Figure 2.1 clearly show that Central America is a rather modest
region for tsunami generation. However large tsunamis, like the 1992 Nicaraguan
tsunami, do occur and they have to be taken account for.

Part 1 - Chapter 4 - 14

05-01-2004

Figure 4.7
For each region in the Pacific the number of generated large tsunamis (i 2.8) are
plotted. (based on Table 4.5)

Year

1900
1906
1918
1923
1923
1923
1927
1933
1939
1946
1946
1952
1957
1958
1960
1960
1963
1964
1968
1969
1975
1977
1992

3
3
3
3
3
3
3
3.5
3
4
3
4
3.5
3.5
3
5
3
4.5
3
3
3
3.5
2.8

Hmax

Source Region

Bismark Sea,New Guinea


5
Tumaco, Colombia-Ecuador
12
E.Urup,S.Kuril Islands
8
SE off Kamchatka
14
E.Kamchatka peninsula.USSR
13
Sagami Bay, Tokaido, Japan
15
Celebes Sea.Indonesia:SULAWES
29.3
Sanriku, Japan
10.5
Solomon Islands Papua New Guinea
35
Unimak Is.,E.Aleutian Islands
30
East of Vancouver Is., Canada
18.6
SE of Kamchatka peninsula
16.2
Central Aleutian Islands
525
Lituya Bay, SE Alaska
0.25
Peru: Arequipa,Chuquibamba,ca
25
Corral, Southern Chile
15
Urup,Kuril Islands.USSR
67.1
Gulf of Alaska, Alaska pen.
10
N.Celebes, Banda Sea.Indonesia
15
Kamchatka,Bering Strait
14.3
Kalapana, Hawaii Is.USA
15
Sunda islands.Indonesia
10.7
Nicaragua
Table 4.5
Earthquake generated tsunami events with an intensity i 2.8, for the entire
Pacific region, for the period 1900-2002 (data compiled out of the Historical
Tsunami Database for the Pacific, 47 B.C. to present).

05-01-2004

Part 1 - Chapter 4 - 15

4.3 CONCLUSIONS
A full statistical approach proves not to be possible, because linking the available
tsunami data and the earthquake data to wave characteristics is not possible. This
conclusion is confirmed by tsunami researchers. Therefore deriving the design
tsunami is based on the biggest tsunami event that occurred along the Pacific
coast of Central America. The biggest tsunami event that occurred in the period
from 1900 to 2002 was the 1992 Nicaraguan tsunami.
Large tsunami events are infrequent phenomena and mostly do only affect a small
part of the coast, making the occurrence frequency of large tsunami events for one
specific location very small. A rough estimation of the return period of the 1992
Nicaraguan event is 1920 years. Large tsunami events are not only infrequent
along the Central American coast, but also in the entire Pacific. Compared to the
Pacific, Central America is a rather modest tsunamigenic region. However large
tsunamis can occur and have to be taken account for.
An overview of the 1992 Nicaraguan tsunami event is given in Appendix 4. In the
following chapter the 1992 Nicaraguan tsunami is used as a base for deriving a
design tsunami. At first a 1D hydraulic model will be set-up which corresponds
with the Nicaraguan event. After this the wave will be modified in order to
become a design tsunami: the wave height and the wave length will be increased
with 20%.

Part 1 - Chapter 4 - 16

05-01-2004

5 DERIVING THE DESIGN TSUNAMI FOR THE SITES


OF INTEREST
5.1 INTRODUCTION

Nicaragua

Figure 5.1
Pacific coast of Nicaragua

The goal of this chapter is deriving the design tsunami wave for the two sites of
interest i.e. Pie del Giagante and Bahia del Salinas. At first the 1992 tsunami wave
to the coast of Pie del Gigante will be derived. Later this tsunami will be adapted
in order to derive the design tsunami for the two sites.
The 1992 tsunami towards Pie del Gigante will be derived by means of a 1D
hydraulic model, starting from the initial bottom deformation of the 1992
Nicaragua tsunami. In order to simulate the 1992 Nicaragua tsunami, the model
will be calibrated using the available data from the 1992 Nicaragua event. These
include tide gauge readings, run-up data and data of available model
investigations. As the first arriving wave was reported as the highest wave, our
modeling efforts will be focused on the first wave.
In order to obtain the design tsunami for Pie del Gigante the bottom deformation
will be adapted in order to obtain a wave with an increased wave height and
length by 20 %. The same bottom deformation is then used to derive the design
tsunami to the site of Bahia Del Salinas. This way the resulting wave at the two
sites can be compared.
05-01-2004

Part 1 - Chapter 5 - 1

5.2 MODEL SET-UP


In 3.3 it is described how tsunami events are modeled. In this chapter a similar
approach is used, but with a 1D mathematical model instead of a 2D model.

5.2.1 Model description


The model used, is based on the 1D nonlinear shallow water equations, an
overview of the 2D nonlinear shallow water equation is given in 3.2.4. An
implementation into a numerical model of the 1D nonlinear shallow water waves
was provided by Prof. Stelling of the department of Fluid Mechanics.
The 1D nonlinear shallow water equations:
equation of momentum:
VV
V
V

+V.
= g.
Cf
t
x
x
d +

d
Cf
g
t
x

:
:
:
:
:
:
:

(5.1)

depth averaged horizontal velocity vector (m/s)


water elevation (m)
water depth (m)
nondimensional friction coefficient (-)
gravitational acceleration (m/s)
time (s)
distance (m)

equation of continuity:
{(d + )V }
(d + )
=
t
x

(5.2)

The coast and the ocean side of the model are modeled as closed boundaries
(V=0).

Part 1 - Chapter 5 - 2

05-01-2004

5.2.2 Model input


5.2.2.1

Bottom deformation

The input of the bottom deformation is based on the modeling of the 1992
Nicaragua tsunami by Satake (1995). In his investigation Satake sets up a 2D
mathematical model, in which he derives an elevation-subsidence profile for the
1992 Nicaragua event. The profile is derived starting from the seismological data
(seismological moment and focal depth) giving input in a geological model in
order to provide the resulting deformation of the ocean bottom. This bottom
deformation is used as input for a 2D hydraulic model and the results are
compared with the available reported data of the tsunami event namely; tide gauge
recordings at Corinto and Puerto Sandino and run-up data collected along the
Nicaraguan coast. This comparison is used as feedback to adapt the bottom
deformation, but keeping the seismic moment constant (Appendix 2).
By doing this, in the end an elevation-subsidence profile is obtained which is
consistent with the earthquake data on one hand and the wave data (tide gauge
records and run-up data) on the other hand.
The elevation-subsidence profile as derived by Satake (1995) is illustrated in
Figure 5.2.

Figure 5.2
The initial bottom deformation as derived by Satake (1995), overlaid on the
bathymetry, the contour interval is 20 cm. A perpendicular on the coast is
constructed to derive the bottom deformation for the 1D model

05-01-2004

Part 1 - Chapter 5 - 3

The bottom deformation is uniform along the long side of the bottom deformation.
Therefore only one bottom deformation as input for the 1D model has to be
derived. The bottom deformation input is derived along a perpendicular to the
coast, as shown in Figure 5.2. The bottom deformation is simplified by
transforming it to rectangular deformations an uplift of 65cm and a subsidence of
27cm (Figure 5.3). The rectangular deformation has the same surface as the
original profile, resulting in equal volumes of displaced water. This
schematization is used for preliminary calculations. Later on it will be evaluated,
whether a good comparison is given by this profile or not.
Bottom displacement
100

displacement (cm)

80
60
schematised bottom
deformation

40
20

original bottom deformation

0
-20

10000

20000

30000

40000

50000

-40
-60
distance (m)

Figure 5.3
Schematized bottom deformation for the 1D model and the original profile

5.2.2.2

Positioning of the bottom deformation

The bottom deformation, for a subduction zone earthquake, is positioned in the


deepest part of the ocean trench, on the edge of the plate boundary (Figure 5.4).

Figure 5.4
Positioning of the bottom deformation, the left side is positioned in
the deepest part of the ocean trench.

Part 1 - Chapter 5 - 4

05-01-2004

5.2.2.3

Bathymetry

The bathymetry is derived out of sea charts of the Nicaraguan region:


- for the shallow areas a chart with a scale of 1/50.000: (DMA, 1981)
- for the deeper area a chart of 1/300.000 : (INETER, 1989)
in Appendix 5; an overview is given of the bathymetry in front of Pie del Gigante,
Bahia del Salinas and Corinto.
The depths are related to the mean sea level. For the bathymetry input for the
model the points were linearly interpolated.
The Nicaraguan coast is a typical leading edge coast, located near an ocean trench
(140 km out of the coast).
The bathymetry falls apart in three parts:
- the foreshore :
the first steep part of the bathymetry, extending from
the still water level to a depth of MSL50m.
- the continental shelf : a milder part of the bathymetry extending from the
foreshore till a sudden increase in depth marks the
beginning of the ocean trench.
- the ocean trench :
the deepest part of the bathymetry, the Middle America
trench, is formed due to the subduction of the Cocos
plate under the Caribbean plate (Appendix 2). The
ocean trench is characterized by two steep slopes
flanking the deepest point, which has a depth of almost
MSL-5000m.

Note that the term foreshore is used for the underwater part of the shore and not for the part of the shore
above the still water line.

05-01-2004

Part 1 - Chapter 5 - 5

As is clearly illustrated by the bathymetry of Pie del Gigante in Figure 5.5 and
Figure 5.6

0
-1000
-2000
-3000
-4000
-5000
-6000

depth (m)

20000

40000

60000

80000

100000

120000

140000

distance (m)

160000

Bathymetry for Pie Del Gigante

continental shelf

ocean trench

Figure 5.5
The bathymetry at Pie Del Gigante, clearly the ocean trench
and the continental shelf can be distinguished.

0
-50
-100
-150
-200
-250
-300
continental shelf

depth (m)

10000

20000

30000

40000

50000

60000

distance (m)

70000

Bathymetry for Pie Del Gigante (shallow)

foreshore

Figure 5.6
The more shallower parts of the bathymetry at Pie Del Giante, the
Continental Shelf and the foreshore can be distinguished.

Part 1 - Chapter 5 - 6

05-01-2004

5.2.2.4

Bottom friction

A depth related bottom friction is used, related to the Mannings roughness


coefficient:
Cf =

g.n 2
(d + h)1/ 3

(5.3)

with :
- Cf : the non dimensional friction coefficient (-)
- n : Mannings roughness coefficient (m-1/3s),
at first n was chosen 0.02 m-1/3s
- d : water depth (m)
- h : water elevation (m)

5.2.3 Available calibration data


5.2.3.1

Tide gauge recordings

There are two tide gauge recordings available, one at Corinto (depth:MSL-10m)
and one at Puerto Sandino (depth:MSL-4 m) (Figure 5.7).

Figure 5.7
Tidal recordings at Corinto and Puerto Sandino.

Due to a mistake at first the value n = 0.02 m-1/3s instead of the suggested value of n = 0.03 m-1/3s
(Bryant, 2001 and Satake, 1995). Later on it will be proven that the resulting differences are minimal.

05-01-2004

Part 1 - Chapter 5 - 7

Figure 5.8
Tidal recording at Corinto after the tidal components
are removed (Satake, 1995).

The tidal recording at Corinto shows an initial water level withdrawal of about 10
cm, followed by a large water elevation showing a peak to through amplitude of
49 cm. The period of the first rising wave is 9.2 minutes. The tidal elevation at
that moment was +80 cm.
The tidal recording at Puerto Sandino shows an initial water level withdrawal of
about 10 cm, followed by a large water elevation causing the tide recorder to go
off scale, the peak to through amplitude is at least 117 cm. The tidal elevation at
this station was MSL+50 cm.
Because the tide gauge at Corinto kept recording during the event, this registration
will be used as a basis for comparison of the computed wave height (Figure 5.8).
5.2.3.2

Run-up heights

Figure 5.9
Maximum reported run-up heights along the
Nicaraguan Coast (Bryant, 2001)

The run-up heights lie mostly between 3 and 8 m MSL, with a maximum of 10.7
m MSL at El Transito.

Part 1 - Chapter 5 - 8

05-01-2004

Tidal correction
The tidal elevation at the arrival time of the tsunami was MSL+80 cm at Corinto
and +50 cm at Puerto Sandino, no detailed tidal calculations for the coast of
Nicaragua are available. Therefore as a first assumption for a tidal correction
MSL+65 cm is proposed.
This gives run-up heights between 2.35 and 7.35 m SL (sea level), referred to the
momentaneous sea level. The significance of the centimeters is negligible, for
easy comparison the values will be round off to 2.5 and 7.5 m SL.
Corrected for the tide the most run-up heights lie between 2.5 and 7.5 m SL.
In Satake (1995) Figure 5.10 is presented which compiles all the measured run-up
data for the 1992 event. The run-up values are corrected for the tidal component.
Notice that the points lying on one vertical represents different run-up heights at
one coastal location, these differences can be attributed to very local differences in
onshore topography.

Figure 5.10
Reported run-up heights for the 1992 Tsunami.
Most run-up heights lie between 2.5 and 7.5m.

Arrival time
The wave arrived onshore between 40-70 minutes after the rupture.
The first arriving wave was reported as the highest wave.
The relationship between run-up heights and reflected wave heights
The run-up heights can be calculated from the reflected wave heights as follows:
H run up = m.H refl

Hrun-up
m
Hrefl

(5.4)

: run-up height (m)


: multiplification factor (-)
: reflected wave height (m)

The multiplification factor m can be estimated taking into account the wave length
and the slope of the coast (Imamura, 1993). Unluckily, the exact formula is not
05-01-2004

Part 1 - Chapter 5 - 9

given by Imamura (1993) and the original paper whereto he refers is in Japanese.
It is stated to be an empirical formula based on relationship of reflecting wave
heights on a vertical wall and the measured run-up of the same waves on a beach.
For the Nicaraguan coast Imamura (1993) estimates the multiplification factor m
to be 2.
Based on the results of his model Satake (1995) suggests that this factor should be
between 1 and 3. Probably this is a more realistic approach, as is clearly indicated
in Figure 5.10, where for one location already large differences in run-up height
can occur. Clearly illustrating the effect of local onshore topography on the runup.

5.3 PIE DEL GIGANTE


5.3.1 First Calculation
A first calculation will be made for Pie del Gigante, in order to compare the
resulting wave with the tide gauge recording of Corinto, also a calculation for
Corinto will be made.
The model is set-up as follows:
Bottom deformation: 5.2.2.1and 5.2.2.2
Grid size : 500m
Time step : 1.25m/s
Friction
: n=0.02 m-1/3s
5.3.1.1

Pie del Gigante

Bathymetry
The bathymetry in front of Pie del Gigante is shown in Figure 5.5 and Figure 5.6.
Model calculations
Two waves are presented:
- at MSL10m (same depth as the tide gauge at Corinto)
- reflected wave height (closed boundary at the coast)

Part 1 - Chapter 5 - 10

05-01-2004

First Run
3,5
3
Reflected
wave

2,5
height (m)

2
1,5

wave at -10m
reflected wave

1
0,5
0
2003
-0,5

2103

2203

-1

2303

2403

2505

2605

2705

2805

2905

period
time (s)

Figure 5.11
First calculations, calculated waves at MSL-10m and the
reflected wave on the closed coast boundary (MSL 0m).

reflected wave :
max height : 2.85m expected run-up height 2.85-8.55 m (5.2.3.2, m=1-3)
at time
: 2852.5s = 47.5 min
arrival time : 2670s = 44.6 min
wave at 10m

peak-to-through amplitude: 2.26m


period: 8.6min

Defining the period is somewhat ambiguous, because of the 2nd wave peak. The
period is defined as is shown in Figure 5.11.
5.3.1.2

Corinto

For comparison with the tide gauge recordings at Corinto, a calculation will be
made with the bathymetry of Corinto.
Bathymetry
The depth profile in front of Corinto is derived using the sea chart of 1/300.000
also used for the depth profile of Pie del Gigante. Only the profile between 12.8
m and 280 m could be derived from this chart. Nearshore no more detailed
soundings are given and the deeper part of the depth profile lies beyond the
borders of the map.
Therefore for the deeper part the same depth profile is used as for Pie del Gigante.
In 5.3.2.3 a sensitivity analysis will be done to investigate the influence of the
deeper parts.
As for the more shallow parts nearshore, the model will only be used to calculate
the wave at 10 m (fictive location of the tide gauge, the point is linearly
05-01-2004

Part 1 - Chapter 5 - 11

interpolated between 12.8m and 0m).


Bathymetry for Corinto

Pie Del Gigante

20000

40000

60000

80000

100000

120000

140000

distance (m)

160000

Corinto

0
-2000
-3000
-4000

depth (m)

-1000

-5000
-6000

Figure 5.12
Bathymetry for Corinto compared with the bathymetry
for Pie Del Gigante.
Bathymetry for Corinto (shallow)
Pie Del Gigante

10000

20000

30000

40000

50000

Corinto
distance (m)

-100
-150

PDG

depth (m)

-50

-200
-250

Figure 5.13
The Shallower parts of the bathymetry for Corinto compared
with the Bathymetry for Pie Del Gigante.

Model calculations
A calculation is made with the bathymetry for Corinto and compared with the
wave calculated for Pie del Gigante.

Part 1 - Chapter 5 - 12

05-01-2004

wave at -10m
2

height (m)

1,5

1
PDG

0,5

Corinto

0
2503

2578

2653

2728

2803

2878

2953

PDG

-0,5

-1

time (s)

Figure 5.14
Comparing the calculated waves at Pie del Gigante (PDG)
and Corinto at MSL-10m.

5.3.1.3

Comparison with the tide gauge readings at Corinto

Comparison between the calculations for Corinto , Pie Del gigante and the tide
gauge recording at Corinto:

first wave
Peak-to-through
period
arrival
highest wave
Max elevation
remark

Model, -10 m
Pie del Gigante
(reference)

Model, -10 m
Corinto

Tidal recording
Corinto (-10 m)

2.26 m

2.18 m

0.49 m

8.6 min
40 min
first wave
+1.73 m

8.0 min
47 min
first wave
+1.63 m

9.2 min
35-50 min
first wave
+0.39 m

significant 2e wave peak significant 2e wave peak

The two calculated waves are quite similar in height (2.18 m 2.25 m) and
period (8.0 min 8.6 min), but the arrival time at Corinto is a bit later (47 min
40 min), probably due to the shallower continental shelf, which slows down
the wave ( c = g .h ).
Therefore comparing the calculated wave at Corinto with the tidal recording at
Corinto gives the same remarks as the comparison between the calculated wave
height at Pie del Gigante and the tide gauge recording: much higher calculated
waves, but the form and the timing is comparable.

05-01-2004

Part 1 - Chapter 5 - 13

Conclusions
- Clearly this doesnt give good correlation with the wave height on the tidal
recording
- The period of the first rising wave is promising: 8.6min and 8.0min9.2min
- The arrival time on shore is acceptable (PDG: 44.6min40-70min)
- Expected run-up heights are also acceptable (PDG: 2.85 - 8.55 m 2.57.5 m, Figure 5.10)

5.3.2 Possible reasons


Possible reasons that can explain the large difference between the tide gauge
recording and the calculations:
- layout of the site of Corinto
- diffraction of the waves
- sensitivity of the model for variations in input variables.
5.3.2.1

Layout of the site of Corinto

Figure 5.15
Layout of the site of Corinto. The location of the tide gauge is marked.

The expected Hrun-up is calculated using equation (5.4) with Hrefl=2.26m and m ranging from 1 to 3

Part 1 - Chapter 5 - 14

05-01-2004

This clearly shows the complex layout at Corinto. The location of the tide gauge
is marked on the map.
Clearly this complex bay will have a serious impact on the wave penetration: the
incoming wave at the bay mount will split up into the different channels. Probably
this will result in lowering of the arriving wave, afterwards complicated
superposition of reflected waves can occur.
As a first approximation it is assumed that the tidal recording at Corinto, gives to
small wave heights for a wave arriving at the coast (due to the complex layout).
But it gives an impression of the form of the first wave: an initial water level
withdrawal, followed by a larger water elevation, with the period of the first rising
wave about 9.2 min.
As this first wave is reported as the highest wave, we will focus on the first wave.
5.3.2.2

Diffraction of the waves

Figure 5.16
Position of Corinto relative to the initial bottom deformation

As can be clearly seen in Figure 5.16 Corinto is located in front of the northwestern end of the initial water disturbance, at a distance of 90 km. So also
diffraction can play an important role in the lowering of the wave height at
Corinto.
5.3.2.3

05-01-2004

Sensitivity of the model for variations of the input variables

Part 1 - Chapter 5 - 15

The sensitivity of the model for variations of the input variables will be tested by
varying:
- the depth of the ocean trench
- the slope of the continental shelf
- the slope of the foreshore
- the length of the bottom deformation
- the amplitude of the bottom deformation
- the bottom friction
- the reflection on the closed coast boundary.
Because the calculated wave at MSL-10m does not differ much for Pie del
Gigante and Corinto and because the total bathymetry profile for Pie del Gigante
is available, the calculations will be carried out with the bathymetry of Pie del
Gigante. The calculations will be compared with the calculations of the first
calculations (5.3.1.1), which will be referred to as the reference model.
I The depth of the ocean trench
Two calculations are made, one with the depth of the ocean trench increased from
MSL-4860 m to 8000 m and another where the depth is decreased to MSL3000
m (Figure 5.17).
Varying the dpth of the ocean trench

shallow trench
reference model

distance (km)

20000

40000

60000

80000

100000

120000

140000

160000

deep trench

-4000
-6000
-8000

depth (m)

-2000

-10000

Figure 5.17
Varying the depth of the ocean trench

The calculations compared with the reference model are displayed in the
following figures:

Part 1 - Chapter 5 - 16

05-01-2004

waves at -10m
2

shallow

1,5

height (m)

1
reference model
0,5

shallow trench
deep trench

0
2503

2578

2653

2728

2803

2878

2953

-0,5

deep

-1

time (s)

Figure 5.18
The calculated waves at -10m compared with the reference model
reflecting wave

deep

3,5
3

height (m)

2,5
reference model

shallow trench

shallow

1,5

deep trench

1
0,5
0
2503

2603

2703

2803

2903

time (s)

Figure 5.19
The calculated reflecting waves on the coast boundary
compared with the reference model

Comparison of the calculations and the referece model:

reflecting wave
Max elevation
arrival time
wave at -10 m
peak-to-through
period

Model,
Dtrench = -8000m

Model,
Dtrench = -3000m

Model (reference)
Dtrench = -4860m

2.91 m
44.5 min

2.76 m
44.7 min

2.85 m
44.6 min

2.31 m

2.19 m

2.26 m

8.6 min

8.6 min

8.6 min

Very large variations in the depth of the ocean trench lead to small variations of
the wave height. Deeper ocean trenches lead to a slightly earlier arrival time of the
05-01-2004

Part 1 - Chapter 5 - 17

wave on shore then shallower ocean trenches. When one looks at the wave of
MSL10m, especially the peak is retarded. Possibly this can be explained by the
bottom elevation lying primary in the now shallower ocean trench. Therefore the
initial water elevation now lies in a more shallower zone. Causing smaller wave
celerity ( c = g .h ), thus leading to a slightly different wave form.
Possibly the variations in the second wave peak at MSL10m, can be partially
explained by the reflected waves on the coast; a slightly different form and later
arrival could induce a higher superimposed wave.
Conclusion
Very large variations in the depth of the ocean trench, lead to small differences in
the resulting wave. Therefore the impact of inaccuracy of the deeper bathymetry
on the calculations is negligible.
II The slope of the continental shelf
Two calculations are made, one with the steepness of the continental shelf
increased from 1/520 to 1/400 and another where the steepness is decreased to
1/750 (Figure 5.20).

Varying the slope of the continental shelf


reference model (1/520)
steeper (1/400)

0
-50
-100
-150
-200
-250

height (m)

10000

20000

30000

40000

50000

60000

distance (m)

70000

milder (1/750)

Figure 5.20
Varying the slope of the continental shelf.

The calculations compared with the reference model are displayed in the
following figures:

Part 1 - Chapter 5 - 18

05-01-2004

reflecting wave

steep

mild

3,5
3

height (m)

2,5
reference model

milder continental shelf


1,5

steeper continental shelf

1
0,5
0
2503

2553

2603

2653

2703

2753

2803

2853

2903

2953

time (s)

Figure 5.21
The calculated waves at -10m compared with the reference model.
reflecting wave

steep

mild

3,5
3

height (m)

2,5
reference model

milder continental shelf


1,5

steeper continental shelf

1
0,5
0
2503

2553

2603

2653

2703

2753

2803

2853

2903

2953

time (s)

Figure 5.22
The calculated reflecting waves on the coast boundary
compared with the reference model.

Comparison of the calculations and the reference model:

reflecting wave
Max elevation
arrival time
wave at -10 m
peak-to-through
period

Model,
1/750

Model,
1/400

Model (reference)
1/520

2.85 m
47 min

2.86 m
42.6 min

2.85 m
44.6 min

2.23 m

2.27 m

2.26 m

8.6 min

8.6 min

8.6 min

Variations in the slope of the continental shelf, practically doesnt affect the wave
heights. Also the form of the waves remain practically unchanged.
But the arrival time of the waves is seriously affected: the milder continental slope
05-01-2004

Part 1 - Chapter 5 - 19

causes the wave to arrive minutes later. This can be explained by the wave
celerity, which is a function of the depth: c = g .h . Milder slopes lead to shallower
depths, which causes the wave to slow down.
Conclusion
Steeper continental slopes cause the wave to arrive earlier ( c = g .h ). But the
slope has practically no influence on the wave height, nor on the wave form.
II The slope of the foreshore
Two calculations are made, one with steepness of the foreshore increased from
1/100 to 1/80 and another where the steepness is decreased from 1/100 to 1/120
(Figure 5.23).
Varyin g th e slop e of th e fore sh o re

0
0

00

00
-1

50

00

20

35

00

00
50

00

65

80

95

d is ta n c e (m )

00

1 /1 2 0
re fe re nc e m o d e l
1 /8 0

-2 0
-4 0
-6 0

h e ig h t

-8 0

Figure 5.23
Varying the slope of the foreshore.

Part 1 - Chapter 5 - 20

05-01-2004

w a ve a t -10m
2,5

1/80

1/120

he ight (m )

1,5

ref erenc e model

1/80
0,5

1/120

0
2503

2553

2603

2653

2703

2753

2803

2853

2903

2953

3003

3053

3103

3153

3203

-0,5

-1

tim e (s)

Figure 5.24
The calculated waves at 10m compared with the reference model.

r e fle ctin g w a ve

3,5

1/120

he igh t (m )

2,5

1/80
ref erenc e model

1/80
1,5

1/120

1
0,5
0
2503

2553

2603

2653

2703

2753

2803

2853

2903

2953

3003

3053

3103

3153

3203

tim e (s )

Figure 5.25
The calculate reflecting waves compared with the reference model.

reflecting wave
Max elevation
arrival time
wave at -10 m
peak-to-through
period

Model,
1/80

Model,
1/120

Model (reference)
1/100

2.88 m
44.3 min

2.76 m
46.7 min

2.85 m
44.6 min

2.45 m

2.47 m

2.26 m

7.8 min

8.1 min

8.6 min

A steeper foreshore leads to an earlier arrival of the wave, partially because the
deeper water depths and partially because of a shorter traveling distance (Figure
5.23). The rest of the variations in the wave, are somewhat more complicated. It is
05-01-2004

Part 1 - Chapter 5 - 21

believed that reflection of the wave partially disturbs the calculations. The
influence of the foreshore is somewhat difficult to distinguish, therefore the
calculations will be redone in 0 with repressing the reflection.
IV The length of the bottom deformation
Two calculations are made, one with length of the bottom deformation increased
with 10% and another where length of the bottom deformation is decreased with
10% . The placement of the bottom deformation with respect to the bathymetry
remains the same (5.2.2.2).
The calculations compared with the reference model are displayed in the
following figures:
waves at -10m
2

L+10%

1,5

height (m)

1
reference model
0,5

L+10%
L-10%

0
2003

2103

2203

2303

2403

2505

2605

2705

2805

2905

-0,5

-1

L-10%

time (s)

Figure 5.26
he calculated waves at -10m compared with the reference model.
reflecting wave

L-10%

2,5

height (m)

2
reference model
1,5

L+10%
L-10%

L+10%

0,5
0
2003

2103

2203

2303

2403

2505

2605

2705

2805

2905

time (s)

Figure 5.27
The calculated reflecting waves on the coast boundary
compared with the reference model.

Part 1 - Chapter 5 - 22

05-01-2004

Comparison of the calculations and the reference model:


Model,
L-10%

Model,
L+10%

Model (reference)
L=58.7 km

2.83 m
45.2 min

2.83 m
44.1 min

2.85 m
44.6 min

peak-to-through

2.26 m

2.12 m

2.26 m

peak-to-through period

127 s

175 s

remarks

lack of 2e wave peak

415 s
high 2e wave peak
2 negative wave peaks

reflecting wave
Max elevation
arrival time
wave at -10 m

Analyzing the reflecting wave at the coast:


The wave heights are almost the same, but there is a clear difference in arrival
time. The smaller the length the later the arrival time, this can be explained by the
longer distance to the coast.
When we look at the waves at 10 m, the situation looks somewhat more complex:
- The different arrival times can be clearly distinguished.
- The calculations with L-10% lacks a second wave peak. The peak-to-through
resembles quite well the reference model as regard for the height but the peakto-through period is much shorter, than the reference model (127s 175s).
Indicating that the wave is significantly shorter.
- The calculations with L+10% shows a higher second wave peak and the
arriving through seems to fall apart in two negative waves. Because the first
wave arrives relatively earlier in than in the reference case, it is believed that
the wave is significantly longer than the reference calculation. Also this implies
a slightly different wave form than the reference calculation, only slightly
different , because no significant differences are noticed in the reflecting wave.
Slightly different incoming waves produce slightly different reflecting waves,
possibly the combination of these two can explain the wave at 10m: a higher
second wave peak and a wave through falling apart in two wave through.
Conclusion
A longer bottom deformation leads to longer waves, but also the form of the wave
is influenced. Due to the shorter travel distances the wave also arrives earlier.
V The amplitude of the bottom deformation
Two calculations are made, one with the amplitude of the bottom deformation
increased with 10% and another where the amplitude of the bottom deformation
is decreased with 10% .
The calculations compared with the reference model are displayed in the
following figures:
05-01-2004

Part 1 - Chapter 5 - 23

w a ve a t -10m

2,5

A+10%

he ight (m )

1,5

ref e renc e mod el

A +1 0%
0,5

A -10%

0
2003

2103

2203

2303

2403

2505

2605

2705

2805

A-10%

2905

-0,5

-1

tim e (s)

Figure 5.30
The calculated waves at -10m compared with the reference model

r e fle ct in g w av e

3,5

A+10%
3

h e ig h t (m )

2,5
ref erenc e model

A +10%
1,5

A -10%

A-10%

0,5
0
2003

2103

2203

2303

2403

2505

2605

2705

2805

2905

t im e ( s )

Figure 5.31
The calculated reflected waves on the coast boundary
compared with the reference model.

Part 1 - Chapter 5 - 24

05-01-2004

Comparison of the calculations and the reference model:

reflecting wave
Max elevation
arrival time
wave at -10 m
peak-to-through
period

Model,
A-10%

Model,
A+10%

Model (reference)
A=+66/-27 cm

2.59 m
44.6 min

3.12 m
44.6 min

2.85 m
44.6 min

2.05 m

2.47 m

2.26 m

8.6 min

8.6 min

8.6 min

A variation of the amplitude of the bottom deformation, clearly has an effect on


the height: a higher amplitude leads to a bigger wave height. The form and the
timing of the waves are practically unaffected
Conclusion
A bigger amplitude of the bottom deformation leads to a higher wave, form and
timing remain unchanged
VI The bottom friction
Different authors suggest mostly the same friction values for the propagation of
tsunami waves. The values presented here are from Bryant (2001) and Satake
(1995).
Calculations will be made with three different values.
For a depth related bottom friction different authors suggest a Manning roughness
coefficient of n=0.03 m-1/3s. Due to a mistake in the reference calculations n was
taken 0.02 m-1/3s, here the difference in results can be compared.
As a typical constant value Cf=0.003 is taken
Also a calculation will be made with Cf=0.01. This is a typical value for tsunami
run-up on a beach and is therefore too large for a wave propagating in water, but it
is interesting for a comparison.
The calculations compared with the reference model are displayed in the
following figures:

05-01-2004

Part 1 - Chapter 5 - 25

wave at -10m
2

Cf=0.01

1,5

height (m)

1
reference model (n=0.02)
n=0.03

0,5

Cf=0.01
Cf=0.003

0
2503

2553

2603

2653

2703

2753

2803

2853

2903

2953

-0,5

-1

time (s)

Figure 5.32
The calculated waves at -10m compared with the reference model.

reflecting wave

Cf=0.01

2,5

height (m)

reference model (n=0.02)


n=0.03

1,5

Cf=0.01
cf=0.003

0,5
0
2503

2553

2603

2653

2703

2753

2803

2853

2903

2953

time (s)

Figure 5.33
The calculated reflecting waves on the coast boundary compared
with the reference model.

Comparison of the calculations and the reference model:

reflecting wave
Max elevation
arrival time
wave at -10 m
peak-to-through
period

Part 1 - Chapter 5 - 26

Model,
n=0.03m-1/3s

Model,
Cf =0.003

Model,
Cf=0.01

Model (reference)
-1/3
n=0.02 m s

2.82 m
44.6 min

2.82 m
44.6 min

2.69 m
44.6 min

2.85 m
44.6 min

2.24 m

2.24 m

2.19 m

2.26 m

8.6 min

8.6 min

8.6 min

8.6 min

05-01-2004

Comparison of the calculations with the values n=0.03m-1/3s and Cf=0.003 reveals
no differences. Comparison of these two calculations with the reference
calculation reveals slightly lower wave heights.
The value for tsunami run-up Cf=0.01shows a small wave height reduction, but
practically no effect in the timing. This suggests that the influence of the friction
only becomes important at shallower areas; the shallow part of the bathymetry is
probably to short to significantly slow down the wave, but is sufficient for a small
lowering of the wave height.
Probably if the bathymetry consists of a longer shallower area or when the
tsunami wave arrives on shore, the bottom friction becomes more important.
Although n=0.02m-1/3s, was used during the previous calculations instead of the
suggested value n=0.03m-1/3s, the differences are supposed to be negligible and
the conclusions are believed to remain valid. Nevertheless it is decided for further
calculations, beyond this sensitivity analysis, to use n=0.03m-1/3s instead of
n=0.02m-1/3s, as this is the suggested value in literature.
Conclusion
With this bathymetry the influence of the bottom friction is limited, when using
the values Cf=0.003, n=0.03m-1/3s or n=0.02m-1/3s, no significant differences are
noticed.
VII The reflection on the closed coast boundary
Comparing all the different runs until now with tide gauge reading at Corinto, it is
striking that all these different model calculations have a significant second wave
peak and the tide gauge reading lacks this wave peak.
It is suspected that this second wave peak for the wave at MSL10 m is caused by
the reflected wave superimposed on the arriving wave.
This is investigated by extending the bathymetry at the coast side with a canal of
a constant depth (-10 m) and a length of 20 km to prevent reflection disturbing
the presented wave.
The calculations compared with the reference model are displayed in the
following figure:

The required minimum length of the canal not to disturb the wave height at the beginning of the canal
(d=-10m):
- T = 8.6min, take as an upper limit 15min
c = g.h 10m / s
this gives L=c.T=9000 m
The minimum required length of the canal is L/2=4.5 km, the canal of 20 km is thus more than sufficient.

05-01-2004

Part 1 - Chapter 5 - 27

wave at -10m
2

height (m)

1,5

1
reference model

0,5

0
2503

canal extension (d=-10m)

2553

2603

2653

2703

2753

2803

2853

2903

2953

canal

-0,5

-1

time (s)

Figure 5.34
The calculated waves at -10m compared with the reference model.

Comparison of the calculations with the reference model:

wave at -10 m
first through
first peak
peak-to-through
period
remark

Model,
canal d=-10 m

Model
(reference)

Tidal recording
Corinto (-10 m)

-0.29 m
+1.7 m
1.99m
9.1 min
no second wave peak

-0.53 m
+1.73 m
2.26 m
8.6 min

-0.10 m
+0.39 m
0.49 m
9.2 min

The arriving through is decreased by almost half, probably this can be explained
by the absence of the negative reflected wave. Also the second wave peak
decreases significantly, which indicates that this second wave peak shown in the
calculations is partially caused by reflection.
However there is still a second wave peak present. It is believed this is a wave
peak superimposed on the first wave peak. This superimposed wave peak travels
slightly slower than the first wave peak, which is somewhat higher. The velocity
difference starts to become more important in shallower water ( c = g .h , h = water
depth + water elevation).
The combination of a superimposed, slightly slower second wave peak and the
reflection on the closed coast boundary explains some of the peculiar differences
between the calculated waves at 10m. The variations in the second wave peak
and the differences in measured periods, can be explained by these two
phenomena. This implies it is not correct measuring the period as is done until
know (Figure 5.13). The elimination of the second wave peak by letting it
propagate in the canal of 10m proves not to be simple. Because the velocity
difference is very small, the friction in the shallow canal becomes more important
Part 1 - Chapter 5 - 28

05-01-2004

and the wave does not retains its form. Therefore it is proposed to keep this
artificial definition of the wave period for comparison of calculated waves.
Comparing the calculations with the recorded profile at Corinto(Figure 5.8), it is
striking that the wave profile more matches the recorded profile, with regard to
the form not for the height. Probably the tide gauge is located in a harbor dock
more or less comparable with our artificial canal. Because of this, the form of the
calculated wave approaches more the tide gauge recording. However comparison
between the calculated period and the period of the recorded wave at Corinto is
difficult because of the 2nd wave peak.
Conclusion
By extending the bathymetry with a canal of -10m and 20 km long, no reflection
disturbs the arriving wave, the arriving through is almost decreased by half and
the second wave peak is decreased significantly. Generally said the wave profile
more matches the recorded profile at Corinto with regard to the form, but not for
the height.
Remark
In 5.3.2.3.II The slope of the foreshore the reflection and the 2nd wave peak
made it impossible to distinguish the influence of the slope of the foreshore. When
redoing these calculations with a model with an extended canal of MSL12.5m
the following waves are become.
w a ve a t -12, 5m (n o re fle ctio n )

1/120

1/80
1,5

he ight (m )

1
ref erenc e model
0,5

1/80
1/120

05

55

05

55

05

55

05

55
29

29

28

28

27

27

26

05

53

03

53

03

53

03

53

03

53

55

26

25

25

24

24

23

23

22

22

21

21

20

20

03

-0,5

-1

ti m e (s)

Figure 5.35
The calculated waves at -12.5 for a varying foreshore slope
compared with the reference model

The only difference between the calculated waves, is the arrival time. The wave
height nor the form are affected by varying the foreshore. The peak-to-through
amplitude is 1.88m for the three calculations. It is concluded that the slope of the
foreshore, does only affect the propagation speed of the wave not the wave itself.
05-01-2004

Part 1 - Chapter 5 - 29

VIII Summary
the depth of the ocean
trench :

Very large variations in the depth of the ocean trench, lead


to small differences in the resulting wave. Therefore the
impact of inaccuracy of the deeper bathymetry on the
calculations is negligible.

the slope of the


continental shelf :

Steeper continental slopes cause the wave to arrive earlier


( c = g .h ). But the slope has practically no influence on the
wave height, nor on the wave form.

the slope of the


foreshore :

The wave remains unchanged. Only the wave propagation


speed is affected.

the length of the bottom A longer bottom deformation leads to longer waves, but
also the form of the wave is influenced. Due to the shorter
deformation :
travel distances the wave arrives earlier.
the amplitude of the
bottom deformation :

A bigger amplitude of the bottom deformation leads to a


higher wave, form and timing remain unchanged

the bottom friction :

With this bathymetry the influence of the bottom friction is


limited, when using the values Cf=0.003, n=0.03m-1/3s or
n=0.02m-1/3s, no significant differences can be found.

the reflection on the


closed coast boundary :

By extending the model with a canal of MSL-10m and 20


km long, no reflection disturbs the arriving wave, the
second wave peak is decreased seriously and the arriving
through is almost decreased by half. Generally said the
wave profile is a better approach of the recorded profile at
Corinto, with regard to the form but not for the height.

Relative large variations in bathymetry (ocean trench continental shelf and


foreshore) do hardly affect the arriving wave. What occurs is a slight difference in
travel time of the superimposed waves (due to variations in depth and bathymetry
length), leading to a slightly different resulting wave. When using the values
Cf=0.003, n=0.03m-1/3s or n=0.02m-1/3s no significant differences can be found.
This makes the model insensitive for these input values, when a canal is included
to suppress reflection.
Only the bottom deformation does affect the arriving wave; the amplitude of the
bottom deformation affects the wave height and the length of the bottom
deformation affects the wave length.
This makes the model a very robust tool for making calculations of earthquake
generated tsunami waves.
Part 1 - Chapter 5 - 30

05-01-2004

5.3.2.4

Conclusions

- It is clear that the small wave height of the tide recording is for the better part
explained by the layout of the Corinto bay and by diffraction. The differences
in height caused by using the bathymetry of Corinto or by varying the different
input parameters is very limited.
Because a 1D model is used the influence of the complex lay-out nor of the
diffraction can be involved, therefore the tidal recording at Corinto can not be
used for calibrating the height of the wave. For the calibration of the wave
height another way has to be found.
- The model is able to approach the form of the wave recorded with the tide
gauge, when no reflection occurs. Even with the simplified rectangular bottom
deformation, therefore only calculations will be done with the simplified
bottom deformation.
- However comparison between the calculated period and the period of the
recorded wave at Corinto is difficult. Therefore it is proposed to take the length
of the bottom deformation as the base for the calibrated wave. Then increasing
the measured period with 20%. And in design calculations, varying the
measured period of the wave to take into account the large uncertainty.

5.3.3 Calibration of the model: the 1992 Nicaraguan tsunami


In Figure 5.36 calculations with the model before calibration are presented, two
waves are shown; the wave at 10m with a suppressed reflection (as described in
5.3.2.3. VII) and the wave reflecting on the coast.
The calculations are carried out in the same way as 5.3.1 First Calculations.
grid size: 500m
time step: 1.25m/s
the same bottom deformation (5.2.2.1).
The only input parameter that is changed is the Mannings roughness coefficient,
n, which is changed from n=0.02 m-1/3s to n=0.03 m-1/3s.

05-01-2004

Part 1 - Chapter 5 - 31

wave at -10m
3
2,5

height (m)

2
1,5
reflecting wave

wave at -10m (no reflection)

0,5
0
1

41

81

121

161

-0,5
-1

time (s)

Figure 5.36
The wave before calibration.

Model,
before calibration
reflecting wave
Max elevation
arrival time
wave at -10 m

5.3.3.1

2.82 m
44.6 min

peak-to-through

1.97 m

period

9.1 min

Time

The arrival time on shore is acceptable: 44.6 min. Reported arrival time range
from 40 to 70min.
5.3.3.2

Form of the wave

Figure 5.37
Tide gauge recording at Corinto.

As can be seen on Figure 5.36 the model approaches the form of the recorded
Part 1 - Chapter 5 - 32

05-01-2004

wave: a small leading through followed by a larger water elevation, the peak-tothrough period is in the order of a few minutes. Comparing the tide gauge
recording with the calculations made for Corinto (Figure 5.38) clearly shows the
same wave form.
As stated earlier comparison between the calculated period and the period of the
recorded wave at Corinto is difficult. Therefore it is proposed to take the length of
the bottom deformation as given by Satake (1995) as the base for the calibrated
wave.

tide gauge
recording

Figure 5.38
The wave at Corinto (d=-10m and the reflection is suppressed) and the tide gauge
recording of the 1992 Tsunami.

5.3.3.3

Height of the wave

As calibration on the recorded wave height at Corinto is not possible, due to the
complex layout of the Corinto bay and to refraction. It will be tried to calibrate the
height of the model by using the results of Satake (1995) concerning the modeling
of the 1992 Nicaraguan tsunami event. Satake (1995) presents a graph which
represents the height of the reflecting wave (on the closed coast boundary) for the
coast of Nicaragua (Figure 5.39).

Figure 5.39
Calculated reflected wave heights of the 1992 Nicaragua tsunami by Satake (1995).

05-01-2004

Part 1 - Chapter 5 - 33

The locations marked are Bahia del Salinas (BDS), Pie del Gigante (PDG), Puerto
Sandino (PS) and Corinto (C).

The computed tsunami height is presented from north to south. Neglecting local
deviations, a gradual lowering of the height occurs at the north side and south
side, probably this can be explained by diffraction. Notice that Pie del Gigante and
Bahia del Salinas are located at the south side, with clearly lower wave heights
than in the middle.
In the middle, a 100 km wide zone can be distinguished with a nearly constant
reflected wave height a bit lower than 2m. It is decided to take this 2 m as the
calibration height of the model.
By adapting the amplitude of the bottom deformation, the reflected wave height is
calibrated to be 2 m for Pie Del Gigante.
Callibrated wave height

2,5

height (m)

1,5
reflecting wave
1

0,5

0
2503

2578

2653

2728

2803

2878

2953

time (s)

Figure 5.40
Calibrating the reflecting wave height at a height of 2m.

In order to become a reflecting wave height of 2 m, the amplitude of the bottom


deformation had to be multiplied with 0.69; the height of the uplifting becomes 45
cm and of the subsidence became 19 cm.
5.3.3.4

The calibrated wave at Pie del Gigante

The resulting calibrated waves are presented in Figure 5.41 ; the waves shown are
the reflecting wave and the wave at 10m (with suppressed reflection).

Part 1 - Chapter 5 - 34

05-01-2004

callib rated w av e
2.5

reflecting wave

he ight (m)

1.5
wave a t -10m (no reflec ting )

reflecting wave

0.5
0
20 02.5 2102.5 2202 .5 23 02.5 2402.5
-0.5

2505

26 05

2705

2805

29 05

tim e (s)

Figure 5.41
The Calibrated wave at Pie del Gigante.

5.3.4 Design wave


The 1992 Nicaraguan tsunami is the biggest tsunami event that occurred in the
Central American region in the period 1900-2002. This event was used to
calibrate the model. Therefore the calibrated tsunami wave is the biggest wave
which could occur during the past century at one location.
In order to design a breakwater we want to increase the load caused by the biggest
tsunami wave. No literature is available for increasing the tsunami load, in a wellfound way. Mostly when a design tsunami is needed, one looks at the biggest
event that occurred in the region. In Japan in the sixties and the early seventies the
highest wave which was known to occur was taken as a design load. Later on one
started to realize that at least the biggest tsunami event should be increased in
order to derive a design tsunami, percentages of 10 or 20% are used to increase
the wave height and the wave length.
Despite of this approach building in a somewhat artificial safety , it is the only
available method for deriving a design tsunami from the biggest tsunami event.
Therefore this method will be used here.
In order to get a design load, an adaptation will be made of the calibrated wave:
- the reflecting wave height will be increased by 20%
- the period of the wave at 10m (without reflection) will be increased by 20%
The adaptation will be done by changing the length and the height of the bottom
deformation.
As stated above calibration of the wave period proves not to be possible, because
comparison between the calculated period and the period of the recorded wave at
Corinto is difficult. Therefore the length of the bottom deformation as the base for
the calibrated wave. In this paragraph the measured period will be increased with
20%. To take this uncertainty with regard to the wave period into account, it is
proposed to vary the measured period of the wave in design calculations.
05-01-2004

Part 1 - Chapter 5 - 35

5.3.4.1

T+20%

At first the length of the bottom deformation is increased in order to get an


increase of the wave period with 20% (9.2min11.2min). The model is used with
an artificial canal at the coast side in order to prevent reflection of the wave.
T+20%
1,4
1,2
1

height (m)

0,8
0,6
0,4

wave at -10m (no reflection)

0,2
0
2003
-0,2

2103

2203

2303

2403

2505

2605

2705

2805

2905

-0,4
-0,6

time (s)

Figure 5.42
The wave at -10m with an increased wave period of 20%.

In order to become a wave period of 11.2min, the length of the bottom


deformation had to be increased with 15%. The length of the uplifting zone
becomes 39.6 km and of the subsidence becomes 27,8 km.
A scan be seen the first arriving wave through falls apart in two negative peaks.
To investigate whether this change in wave form could be caused by reflection
from the ocean boundary, a control run was done with an extended ocean side.
Instead of the ocean side boundary being set at 43 km from the deepest point
(Figure 5.5), the ocean side boundary was set at 100 km from the deepest point. No
difference at all occurred. This proved that the difference in wave form could not
be attributed to reflection on the ocean boundary. Probably the differences can be
attributed to differences in bathymetry when longer bottom deformations are used.
5.3.4.2

Hrefl+20%

The amplitude of the bottom deformation is increased in order to get an increase


of the reflected wave height of 20%.

Part 1 - Chapter 5 - 36

05-01-2004

H+20%

2,5

height (m)

1,5

reflecting wave

0,5

0
2503

2578

2653

2728

2803

2878

2953

time (s)

Figure 5.43
The reflecting wave height on the coast with an increased wave height of 20%.

In order to become a reflected wave height of 2.4m, the amplitude of the bottom
deformation had to be increased with 25%. The height of the uplifting zone
becomes 82.5cm and of the subsidence becomes 33.7cm.
The effect of the increased bottom uplift/subsidence on the period is investigated
in a control calculation.
Effect on the period of the increased wave height
1,6

increased
wave height

1,4
1,2

height (m)

1
0,8
0,6

before the increased wave height

0,4

after the increased wave height

0,2
0
2003
-0,2

2103

2203

2303

2403

2505

2605

2705

2805

2905

-0,4
-0,6

time (s)

Figure 5.44
The wave at -10m (without reflection) before and after the increase
in wave height.

The increase in wave height causes the wave to arrive slightly earlier, but there
seems to be no effect on the period of the wave. The period remains 11.2min.

05-01-2004

Part 1 - Chapter 5 - 37

5.3.4.3

The design tsunami for Pie Del Gigante

The design tsunami as derived for Pie del Gigante:


Design wave for Pie del Gigante
3
2,5

height (m)

2
1,5

wave at -10m (no reflection)


reflecting wave

1
0,5
0
2003

2103

2203

2303

2403

2505

2605

2705

2805

2905

-0,5

time (s)

Figure 5.45
Design wave for Pie del Gigante

Design wave
reflecting wave
Max elevation
arrival time
wave at -10 m (no reflection)

2.4 m
43.9 min

peak-to-through

1.59 m

period

11.2 min

5.4 BAHIA DEL SALINAS


The situation for Bahia Del Salinas is somewhat more complicated. As can be
seen on (Figure 5.46), Bahia Del Salinas lies slightly behind a cape.

Part 1 - Chapter 5 - 38

05-01-2004

Figure 5.46
Location of Bahia del Salinas

Bearing in mind that in deeper water (d>-50m) tsunami waves can be treated by
the linear wave theory, a small qualitative analysis can be made:
- Due to the long wavelength, the tsunami wave will go round the cape
(diffraction). However the direct tsunami wave attack will be reduced by this
obstacle.
- The concave form of the coast behind the cape can very well cause focusing of
the tsunami wave, thus increasing the arriving tsunami wave.
This is only a qualitative reflection for a proper quantitative calculation a 2D
model should be used. Unfortunately time is to limited to conduct such a study.
In order to give a first possible comparison between Pie Del Gigante and Bahia
Del Salinas the available 1D model will be used to derive a design tsunami for
Bahia Del Salinas. As bottom deformation the adapted bottom deformation which
led to the design tsunami for Pie Del Gigante, will be used. And as bathymetry the
profile is given along the path shown in Figure 5.46; starting from the deeper
ocean and going to the coast, slightly turning round the cape.

5.4.1 Bathymetry at Bahia Del Salinas


The bathymetry at Bahia del Salinas is shown below (Figure 5.47 and Figure
5.48).

05-01-2004

Part 1 - Chapter 5 - 39

0
-1000
-2000
-3000
-4000
-5000

depth MSL (m)

20000

40000

60000

80000

100000

120000

140000

d istan ce (m )

160000

180000

Bath ym etrie Bahia Del Salin as

Figure 5.47
Bathymetry at Bahia del Salinas.

0
-100
-200
-300
-400
-500
-600

depth MSL (m)

10000

20000

30000

40000

50000

d istan ce (m )

60000

70000

Bath ym etrie Bah ia Del Salinas

Figure 5.48
The shallower parts of the Bathymetry at Bahia del Salinas

5.4.2 Design wave at Bahia Del Salinas


The following waves were derived:
D e sig n w a ve for B a h ia d e l S a l ina s

reflecting
wave

2,5

he ight (m )

1,5

w av e at - 10m ( no ref lec tion)


ref lec ting w av e

0,5

0
2003

2103

2203

2303

2403

2505

2605

2705

2805

2905

-0,5

ti m e (s)

Figure 5.49
The calculated design waves for Bahia del Salinas.

Part 1 - Chapter 5 - 40

05-01-2004

Design wave
reflecting wave
Max elevation
arrival time
wave at -10 m (no reflection)

2.5 m
45.3 min

peak-to-through

1.72 m

period

9.0 min

5.4.3 Comparison between the design wave at Bahia del Salinas and
Pie del Gigante

Comparison between the rellecting waves


3

Bahia del
Salinas

2,5

height (m)

2
Bahia del Salinas

1,5

Pie del Gigante

1
0,5

0
2003

2103

2203

2303

2403

2505

2605

2705

2805

2905

time (s)

Figure 5.50
Comparison between the reflecting design waves.

C o m p a r iso n b e tw e e n th e w a ve s a t -10 m

1,6

Bahia del
Salinas

1,4
1,2

h e ig h t (m )

1
0,8
0,6

Bah ia de l Salinas

0,4

Pie del Gig ante

0,2
0
200 3
-0 ,2

210 3

220 3

230 3

240 3

250 5

260 5

270 5

280 5

290 5

-0 ,4
-0 ,6

ti m e (s)

Figure 5.51
Comparison between the waves at -10 m

05-01-2004

Part 1 - Chapter 5 - 41

The comparison of the reflecting design waves at Bahia del Salinas and Pie del
Gigante shows a later arrival and a slightly higher wave height at Bahia del
Salinas. Comparing the waves at 10 m, shows a later arrival, but also no second
negative wave peak. The differences between the waves at Pie del Gigante and
Bahia del Salinas probably can be attributed to bathymetry differences which
induce slight differences in wave celerity., leading to individual waves being
superimposed in a different way.
It is believed that the differences between the two sides are so small that for
design calculations it is more appropriate to limit our investigation to one
location. And to vary the height and the period of the wave when looking at the
breakwater stability.
Based on these calculations no large differences in occurring tsunami wave can be
distinguished. This confirms the idea that a 1D bathymetry has only a very limited
influence on the generated waves. It should be kept in mind that for a proper
comparison of the two locations a 2D model is necessary. As differences between
the two locations with regard to reflection is complex, this can only be
investigated with 2D models and is probably very sensitive to the positioning of
the bottom deformation and the length of this bottom deformation.
Nevertheless looking at the calculated reflected wave heights by Satake (1995) in
Figure 5.39, shows that local variations are limited to about half a meter. It is
therefore believed that, for preliminary calculations, the here derived design
waves are sufficient for the preliminary design of a breakwater.

5.5 Design tsunami wave


The design tsunami wave as derived out of the model is shown in Figure 5.52, the
reflecting tsunami wave is shown and the wave at 10m (with repressed
reflection, by adding an artificial canal).
Design wave for Pie del Gigante
3
2,5

reflecting wave

height (m)

2
1,5

wave at -10m (no reflection)


reflecting wave

1
0,5
0
2003

2103

2203

2303

2403

2505

2605

2705

2805

2905

-0,5

time (s)

Figure 5.52
Design tsunami wave, two waves are presented, the reflecting wave at the closed
coast boundary and the wave at -10m with an open coast boundary.

Part 1 - Chapter 5 - 42

05-01-2004

Design wave
reflecting wave
Max elevation

2.4 m

arrival time

43.9 min

wave at -10 m (no reflection)


peak-to-through

1.59 m

period

11.2 min

For a correct representation of the rubble mound breakwater a smaller gridsize


near the location of the breakwater has to be used. The last 5000m to the coast a
gridsize is gradually lowered from dx=500m to dx=6m. The use of a smaller
gridsize leads to a larger influence of the friction in the shallower parts of the
bathymetry resulting in a slightly later arrival of the wave (45min43.9min) and
a significant lowering of the reflecting wave height on the closed coast boundary
(1.69m2.4m). The wave at 10m is less affected a peak-to-through wave height
of 1.43m instead of 1.59m, this clearly illustrates that the very shallow parts of the
bathymetry are responsible for the significant lowering of the wave. The tsunami
wave for a gridsize of 6m near the coast is presented in Figure 5.53.
First calculation with dx=6m
2

reflecting wave
height (m)

1,5
1

wave at -10m (no reflection)


reflecting wave

0,5
0
2002,5

2127,5

2252,5

2377,5

2505

2630

2755

2880

-0,5
time (s)

Figure 5.53
First calculations with a gridsize dx=6m.

05-01-2004

Part 1 - Chapter 5 - 43

Design wave (dx=500m)

First Calculations dx=6m

reflecting wave
Max elevation

2.4 m

1.69 m

arrival time

43.9 min

45 min

wave at -10 m (no reflection)


peak-to-through

1.59 m

1.43 m

period

11.2 min

11.2 min

As the design wave is defined with a reflecting wave height of 2.4m, the resulting
design wave for a small gridsize needs to be recalibrated, by adapting the
amplitude of the bottom deformation. In order to become a reflecting wave height
of 2.4m the amplitude of the bottom deformation has to be divided by 0.69, the
height of the uplifting becomes 0.82cm and of the subsidence becomes 0.33cm.
Note that 0.69 is the inverse multiplification factor used for the initial calibration
for the 1992 Nicaraguan tsunami (5.33). This means for a smaller gridsize the
bottom deformation as derived out of Satakes model would not have to be
adapted in order to become the 1992 Nicaraguan tsunami.
The resulting design wave is shown in Figure 5.54 . The result of the recalibration
is that the wave height at 10m is significantly increased. When the gridsize near
the coast is refined to 3m the reflecting wave height is only slightly lowered
(approximately 3cm). This illustrates that the lowering of the wave height, due to
the refining of the gridsize, is convergent. The difference between a gridsize of
3m and 6m is regarded negligible, therefore the gridsize is kept on dx=6m.
Design wave for Pie del Gigante (dx=6m)
3

reflecting wave

2,5

height (m)

2
1,5

wave at -10m (no reflection)


reflecting wave

1
0,5
0
2002,5
-0,5

2127,5

2252,5

2377,5

2505

2630

2755

2880

time (s)

Figure 5.54
Recallibrated design tsunami wave, near the coast dx=6m. Two waves are
presented, the reflecting wave at the closed coast boundary and the wave at -10m
with an open coast boundary.

Part 1 - Chapter 5 - 44

05-01-2004

Design wave
reflecting wave
Max elevation

2.4 m

arrival time

45 min

wave at -10 m (no reflection)


peak-to-through

2.05 m

period

11.2 min

Can this wave arrive at the coast or will it break before it hits the coast? To
investigate this the breaker criterion out of 3.2.5 is checked for the slope of the
Nicaraguan coast (1/100). The wave height H is plotted against the wave period T
(Figure 5.55), for smaller wave periods the wave breaks. This shows that a wave
with a period of 11.2min will break at a wave height of 11.5m. So the tsunami
wave will arrive at the coast unbroken.

Breaking of Tsunami Waves

not broken

11

9,
5

6,
5

3,
5

broken

800
700
600
500
400
300
200
100
0
0,
5

wave period T (s)

(slope 1/100)

wave height H (m)

Figure 5.55
Breaking criterion of tsunami waves for a slope of the coast of 1/100 .

05-01-2004

Part 1 - Chapter 5 - 45

6 Conclusions
Aim of Part 1: Formulate a realistic design tsunami for two harbor locations
At the Pacific Coast of Nicaragua
An overview of the basic tsunami knowledge is given in chapter 1 and 3. The
other chapters of Part 1 are dedicated to deriving a design tsunami.
A statistical approach for defining a design tsunami proves not to be possible.
Therefore, defining the design tsunami is based on the largest tsunami event that
occurred in the Central American region: the 1992 Nicaraguan tsunami. Also with
regard to the entire Pacific, the 1992 Nicaraguan tsunami was a large event. For
one location along the Middle American Trench, a rough estimation of the return
period is 1 in 1920 years. Taking the spreading into account, the return period for
one location lies between 1360 and 13600 years. This is only a very rough
estimation, but it clearly illustrates that such a large tsunami event is an infrequent
phenomenon for Central America.
The design tsunami is based on the 1992 Nicaraguan tsunami. With a 1D
mathematical model, based on the nonlinear shallow water equations, it was tried
to derive the 1992 Nicaraguan tsunami to Pie del Gigante. The calibration is for
mainly based on the model of Satake (1995), but tide gauge recordings at Corinto,
the reported arrival times and the reported run-up heights were also taken into
account. The first wave that arrived on shore was reported as the highest wave,
therefore the modeling efforts are focused on the first wave.
A sensitivity analysis of the input parameters shows that the wave is almost
insensitive to large variations of the bathymetry (depths of the ocean trench,
steepness of the continental shelf and the steepness of the foreshore (from 50 to
0m)). Also, friction only plays a limited role, probably due to the large water
depths. The resulting wave is mainly determined by the length and the amplitude
of the bottom deformation.
Based on the calibrated model the design wave for Pie del Gigante is derived: the
reflecting wave height is increased by 20% and the wave period at 10m is also
increased by 20%. Calculations for Bahia del Salinas show only minor differences
between the two locations. Therefore design considerations will be limited to one
location: Pie del Gigante.
In order to accurately represent the rubble mound breakwater, the gridsize near the
coast is refined from dx=500m to dx=6m. The smaller gridsize leads to a larger
influence of the bottom friction in the shallower areas of the foreshore, causing a
significant lower wave to arrive at the coast. This lowering proves to be
convergent, as gridsizes smaller than dx=6m give a negligible lowering.
Therefore the model is recalibrated for a gridsize of dx=6m. The resulting design
wave is presented in Figure 6.1.

05-01-2004

Part 1 - Chapter 6 - 1

Design wave for Pie del Gigante (dx=6m)


3

reflecting wave

2,5

height (m)

2
1,5

wave at -10m (no reflection)


reflecting wave

1
0,5
0
2002,5
-0,5

2127,5

2252,5

2377,5

2505

2630

2755

2880

time (s)

Figure 6.1
Design wave for Pie del Gigante

Design wave
reflecting wave
Max elevation

2.4 m

arrival time

45 min

wave at -10 m (no reflection)


peak-to-through

2.05 m

period

11.2 min

Remark: in a further study calculations with a 2D model should be carried out to


investigate if complex reflection and amplification occurs at the two harbor
locations. This way the most favorable of the two locations can be determined.
Nevertheless it is believed that the 1D approach provides a first good estimation
of the occurring wave heights. First of all, because the bathymetry in front of the
two locations is pretty straight. Secondly, because the two models of Satake
(1995) and of Imamura (1993) do not show large variations in reflecting wave
heights on the closed coast boundary.

Part 1 - Chapter 6 - 2

05-01-2004

Part 2: Breakwater Stability

Severe damaged caisson breakwater at the Okushiri Port after the


1993 Hokkaido Tsunami.

7 LITERATURE STUDY ON THE CAISSON


BREAKWATER STABILITY
From the literature study a design method can be presented for a caisson breakwater
under an unbroken tsunami wave. This paper will be discussed in this chapter. Also
a reference is made to some other interesting work concerning caisson breakwater
stability under a tsunami bore.

7.1 ON THE HYDRAULIC ASPECTS OF TSUNAMI


BREAKWATERS IN JAPAN
Tanimoto,1981

One part of the paper is dedicated to the stability of a caisson breakwater against
tsunami forces. Laboratory experiments were carried out in order to investigate the
wave pressure resulting from tsunami waves. It was proven that, when unbroken
tsunami waves hit a vertical wall the pressure distribution becomes hydrostatic.
Tanimoto then uses these findings to derive a pressure distribution on the caisson for
tsunami waves. This pressure distribution is used to calculate the stability against
sliding and overturning. The concept of stability calculations is very similar to
Godas design principles for short waves.
Finally he makes calculations with his formula for a known tsunami induced failure
of the Kawaragi breakwater and good correspondence is found

7.1.1 Design formulas


Tanimoto notes that in order to prevent tsunamis to act as a bore on the structure, the
breakwater should be build in relatively deep water. The greatest difficulty Tanimoto
encounters is the estimation of the incident tsunami wave height H. As design wave
the reflective wave on a closed boundary should be taken, i.e. the positive reflecting
wave height H (in fact this is the positive amplitude).
When the water level behind the breakwater is initially lowered, this should be taken
as the water level behind the breakwater and not the astronomical water level.

hL

Ps

Figure 7.1
Wave pressure distribution due to a non-breaking tsunami wave, the pressure
distribution is assumed to be hydrostatic.

05-01-2004

Part 2 - Chapter 7 - 1

With * (m) the fictive overtopping height of the tsunami wave, d (m) the depth of
the sill, h (m) is the water depth in front of the breakwater structure. The caisson
width is represented as B (m), hc (m) is the distance from the SWL to the top of the
caisson and h (m) is the distance from the SWL to the bottom of the caisson. The
pressure distribution at the seaside heel of the caisson is presented by p and pu
(kN/m). If a water lowering occurs in the harbor, the water lowering is represented
by hL (m) and the resulting pressure distribution is presented by ps (kN/m).
Proposed formulas:

(the forces are calculated per m width)

* = 1.5H
p = pu = 1.1w0 H

(6.1)

ps = w0 hL

w0 is the relative weight of the water.


The total horizontal wave force, P (kN) , the uplift force, U (kN), and the overturning
moments around the heel of the caisson, Mp and Mu (kN/m) are expressed as follows:
h* h*

P = 1 + 1 c c ph '
3H h '
2
1
hc* hc* 1 4 hc* hc*
2
M p = + 1
+
ph '
2
3
'
2
9
'
H
h
H
h

1
pu B
2
1
M u = pu B 2
3

(6.2)

U=

with;

hc* = min {*, hc }

The horizontal force Ps (kN), if a water lowering occurs at the harborside,


calculated as follows:
h

Ps = L + ( h ' hL ) ps
2

is

(6.3)

These are then used to calculate the safety factor against sliding and overturning:
sliding: S .F . =

overturning:

(W0 U )
1.2
P + Ps

S .F . =

Part 2 - Chapter 7 - 2

W0 xW0 M u
M P + Ps xPs

(6.4)

1.2

(6.5)

05-01-2004

With W0 (kN) the weight of the caisson in the water (i.e. the weight of the caisson
minus the buoyancy), Ps the force due to the lowering of the water level behind the
breakwater and xWo and xPs (m) the respective arm lengths around the heel of the
caisson.

7.1.2 Comparing to the Goda design formulas for unbroken short


period waves
As stated above Tanimoto uses the same method as Goda: starting from a pressure
distribution due to the wave load, the resulting forces and moments on the caisson
are calculated and these are used to calculate the safety factors against sliding and
overturning.
The pressure distribution on the caisson assumed by Goda is presented in Figure 7.2 .
The formula for calculating the different pressures are given below. Goda has his own
method for defining the design wave Hmax for details on this method one is referred to
the original publication of Goda (1992).

Figure 7.2
Pressure distribution by short period waves on a caisson breakwater as assumed by
Goda.

The fictive elevation height of the wave crest is denoted as * (m). The resulting
pressure distribution on the caisson is sketched in Figure 7.2. The wave pressure
takes the largest intensity p1 (kN/m) at the SWL and decreases linearly towards the
elevation * and towards the bottom of the caisson and the sea bottom, where the
wave pressure is represented as respectively p3 (kN/m) and as p2 (kN/m). The
maximum vertical pressure at the heel of the caisson is represented as pu (kN/m).
The top of the caisson to the SWL is denoted as hc (m) and the distance from the
SWL to the bottom of the caisson is h (m). The depth of the sill is d (m), the depth of
the entire breakwater structure is h (m).

05-01-2004

Part 2 - Chapter 7 - 3

Goda's design formulas:


fictive crest height
* = 0.75 (1 + cos ) H max

(6.6)

with: the angle of approach of the wave crest


The wave pressures are calculated as follows:
p1 = 0.5 (1 + cos ) (1 + 2 cos2 ) w0 H max
p2 = p1 / cosh kh

(6.7)

p3 = 3 p1
in which:
1 = 0.6 + 0.5 ( 2kh / sinh 2kh )
2 = min

{(( h d ) / 3h ) ( H
b

/ d ) , 2d / H max
2

max

(6.8)

3 = 1 ( h '/ h )(1 1/ cosh kh )


where hb (m) denotes the water depth at the location at a distance 5H1/3 seaward
of the
breakwater.
The maximum uplift pressure pu:
pu = 0.5 (1 + cos ) 1 3 w0 H max

(6.9)

The caisson is subjected to a buoyancy force corresponding to the displaced volume


of water as sketched in Figure 7.2 .
The calculated pressures are then used to calculate the total horizontal wave force, P
(kN) , the uplift force, U (kN), and the overturning moments around the heel of the
caisson, Mp and Mu (kN/m) by simple algebra. The calculated forces and moments
are then used to calculate the safety factors against sliding and overturning:
(W U )
sliding:
S .F . =
1.2
(6.10)
Ps
overturning:

S .F . =

Part 2 - Chapter 7 - 4

(Wt M u )
MP

1.2

(6.11)

05-01-2004

The main difference between the pressure distributions for a short period wave and an
unbroken tsunami wave is clearly illustrated in Figure 7.1 and Figure 7.2. An
unbroken tsunami wave induces a hydrostatic pressure distribution on the caisson.
Where for short period waves a pressure peak is assumed in the vicinity of the SWL,
this pressure peak gradually lowers towards the bottom of the caisson. The vertical
pressure on the seaside heel of the caisson is not equal to the horizontal pressure, this
difference is introduced to improve correspondence with model tests. For tsunami
waves a possible extra load has to be taken into account as a lowering of the water
level at the harborside is expected.
An important parameter in the pressure distribution for short period waves is the
wave length. By adapting the design calculations for the preliminary design of the
caisson breakwater at Pie del Gigante for larger wave lengths eventually the pressures
p1, p2, and p3 become equal, i.e. the pressure distribution becomes almost hydrostatic.
Only the vertical pressure at the heel of the caisson pu is still somewhat lower due to
the artificial adaptation in the Goda formulas, this pressure is exactly hydrostatic. In
Table 7.1 the pressures as calculated for the breakwater at Pie del Gigante are
presented and compared with the same calculation using a wave length of 10,000m
instead of 121m (the calculations are included in appendix 7). Comparing these
pressures with the pressures calculated for a tsunami which causes the same crest
elevation * and no water lowering (Table 7.1) it is obvious that the pressures are
almost the same. Therefore it can be concluded that the design formulas for unbroken
tsunami waves presented by Tanimoto are an end form of the Goda design formulas
for short period waves for waves with very long periods.
Comparing the resulting pressures p1 and p for respectively short period waves and
respectively long tsunami waves, it is obvious that for the same crest elevation, an
unbroken tsunami wave gives a larger pressure than a short period wave. This can be
explained by the fact that Goda assumes unbroken short period waves acting on the
caisson. Thus no dynamic wave pulse acts on the caisson. The reflecting wave
induces a pressure p1, which becomes smaller for shorter wave periods. The smaller
the period, the faster the rising and lowering of the reflecting wave, thus the smaller
the resulting wave pressure. Larger wave periods lead to higher pressure peaks p1,
eventually for very large periods p1 p.
*=15.6m
P1
P2
P3

pu

L=121m
81 kN/m
52 kN/m
56 kN/m
54 kN/m

L=10 000 m
119 kN/m
119 kN/m
119 kN/m
116 kN/m

tsunami
116 kN/m
116 kN/m
116 kN/m
116 kN/m

Table 7.1
Pressure calculations with the Goda formulas (L=121 m and L=10 000 m) and with the
Tanimoto formulas (Tsunamis) for the same crest elevation (*=15.6m).

05-01-2004

Part 2 - Chapter 7 - 5

7.1.3 Comments
The assumption of a hydrostatic load
The assumption of a hydrostatic load due to an unbroken tsunami wave is believed to
be correct as long as the caisson is not overtopped. When overtopping occurs the
pressure distribution on the front of the caisson is no longer hydrostatic, however it is
believed that the hydrostatic load gives an upper limit for the occurring pressure.
The safety factor
Modern calculation standards such as Eurocode make a distinction between acting
and reacting forces. Consequently safety factors are then calculated as follows:
S .F . =

A
R

(6.12)

In the work of Tanimoto (1982) and Goda (1992) this distinction is not made and
safety factors are calculated slightly different. The difference in the resulting safety
factors can be easily illustrated by comparing the safety factor against sliding as
calculated by Tanimoto with the safety factor against sliding as calculated based on
equation 6.12.
S .F . =

Safety factor against sliding (Tanimoto):

(W0 U )
P + Ps

(6.13)

Suppose the safety factor equals 1 (equilibrium of forces), the expression can
easily be rewritten in acting and reacting forces.
1=

(W0 U )
P + Ps

P + Ps + U = W0
1=

W0
=
P + Ps + U

A
R

(6.14)

For equilibrium of forces the safety factor as in equation 6.12 can be easily derived
out of the safety factor as calculated by Tanimoto (equ. 6.13). But when no
equilibrium of forces exists both expressions (equ. 6.13 and equ. 6.14) give a
different safety factor. The difference is usually only in the order of a few percent. In
this study, calculations will be done with the formulas as proposed by Tanimoto
(1982) and Goda (1992). However for the further design it should be considered to
follow the approach as in Eurocode guidelines and to consequently make the
distinction between acting and reacting forces.

Part 2 - Chapter 7 - 6

05-01-2004

7.2 Other references


As for the Nicaraguan coast, tsunamis will almost always hit
unbroken tsunami wave (Figure 5.5). Therefore some interesting
breakwater stability under tsunami bores are not discussed here
Kimura (1996) and Camfield (1993). These are two valuable
studying the breakwater stability under a tsunami bore.

the coast as an
work on caisson
e.g. the work of
references when

7.3 Conclusion
The Tanimoto design formulas are based on the assumption that unbroken tsunami
waves manifest themselves as a hydrostatic load on a caisson. It is a straightforward
design method similar to Godas design principles for short period waves. It is shown
that the Tanimoto formulas are an end form of the Goda design principles for very
long waves.
It is believed that the assumption of a hydrostatic pressure distribution provides an
upper limit for the load of an unbroken tsunami wave. This is believed to be
sufficient for the preliminary design of a caisson breakwater. Nevertheless for further
design also the stability of the sill against a large head difference and the stability of
the soil should be taken into account. For a final design model tests are required.

05-01-2004

Part 2 - Chapter 7 - 7

8 DESIGN OF THE CAISSON BREAKWATER


The design tsunami arrives at the coast as an unbroken tsunami wave. Therefore
the design formulas presented by Tanimoto (1981) will be used to calculate the
stability of the caisson breakwater, which is designed on short period waves. The
preliminary design of the caisson for short period waves is described in Appendix
6.

8.1 THE CAISSON BREAKWATER FOR PIE DEL GIGANTE

Figure 8.1
Caisson breakwater for Pie del Gigante as designed for short waves.

8.2 THE DESIGN TSUNAMI WAVE


High water is normative, then the water depth is -19.8m HHWL. Based on the
design tsunami wave, the reflecting wave at 19.8m HHWL (closed border) was
calculated. The calculated reflecting wave is illustrated below:

05-01-2004

Part 2 - Chapter 8 - 1

D e s ig n w a v e fo r the c ais so n b rea k w a te r


3
2,5
2
he igh t (m )

1,5
1

reflec ting wave on the c ais s on

0,5
0
2002,5
-0,5

2127,5

2252,5

2377,5

2505

2630

2755

2880

-1
-1,5
tim e (s)

Figure 8.2
The reflecting wave at -19.8m.

Positive reflecting wave height: Hposrefl = 2.7m


Negative reflective wave height: Hnegrefl = 0.4m
As a first approach the initial water lowering in the harbor will be taken equal to
the negative reflecting wave height.

8.3 CAISSON TSUNAMI STABILITY


Stability under the design tsunami
The complete calculations are given in appendix 8.
uplift force
horizontal Force
force due to water lowering
relative weight of the caisson
Safety factor
Sliding
overturning

Part 2 - Chapter 8 - 2

U
P
Ps
W0

= 202.6 kN/m
= 564.9 kN/m
= 67.1 kn/m
= 3353.7 kN/m

S.F. = 3.0>1.2
S.F. = 3.6>1.2

05-01-2004

When does one of the safety factors become too low


Suppose an initial water level lowering at the harborside of 1 m
Hposrefl
S.F.sliding
S.F.overturning

= 6.2 m
= 1.20
= 1.32

Suppose an initial water level lowering at the harborside of 3 m


Hposrefl
S.F.sliding
S.F.overturning

= 5.2 m
= 1.20
= 1.39

Waves which can trigger failure of the caisson are waves generated by bottom
deformations approximately twice the initial amplitude of +0..82m and 0.33m.
So the breakwater has a sufficient large safety margin to take into account the
uncertainty concerning the design tsunami wave.

8.4 CONCLUSIONS
The caisson breakwater designed for short waves is sufficiently stable with regard
to the design tsunami. The safety factors become too low for much higher
tsunamis, with bottom deformations approximately for the initial amplitude of
+0..82m and 0.33m.

05-01-2004

Part 2 - Chapter 8 - 3

9 LITERATURE STUDY ON THE RUBBLE MOUND


BREAKWATER STABILITY
In spite of intensive research, only a few specific references concerning the
stability of rubble mound breakwaters under tsunami attack were found.
Although no ready to use design formulas can be formulated, the found
papers are used to give insight in the primary failure modes and the location
of the weak spots of a rubble mound structure under a tsunami wave.

9.1 LABORATORY STUDY FOR THE DESIGN OF TSUNAMI


BARRIER
Kamel, 1970
In the sixties, research on breakwater stability under tsunami attack was
carried out by the American Corps of Engineers*, in order to give
recommendations for the construction of a tsunami barrier at the harbor of
Hilo. Hilo harbor is the second largest harbor of the state of Hawaii. Since
1819 the Hilo area has been hit by seven severe tsunamis, devastating the
Hilo area and 35 others have caused some damage. Therefore in 1960 the
Congress of the United States directed a comprehensive study to be made of
tsunami protection for the city and the harbor of Hilo. 3D Model studies
were used to determine the optimum location, the most appropriate
navigation opening and the design conditions, whereas 2D models were
used to determine the adequacy of the selected design sections. The basic
idea was using the existing rubble mound breakwater and formulate
recommendations for the reinforcement and the extension of this structure,
in order to give sufficient protection against tsunami attack. Here a brief
overview will be given of the findings of the 2D model tests concerning the
stability of the breakwater sections.
Design wave
As design tsunami the tsunami event of 1946 was used, with an increased
wave height and duration of 125%. This is the largest well-documented
tsunami that struck Hilo. 3D model tests were conducted to determine the
overtopping and run-up of different locations along the barrier. The 2D
stability tests were conducted by subjecting a given barrier section to the
attack of bores with increasing percentages of the 1946 tsunami height.

In the original document US units were used, in the presented summary all units are converted to
SI units, however the figures are used in their original form and still contain the US units. The
following conversion is applicable; 1ft = 0.305m , 1lb = 0.453kg , 1US ton = 907,2 kg

05-01-2004

Part 2 - Chapter 9 - 1

Setup
Two different barrier concepts were taken into consideration, overtopping
breakwater sections and nonovertopping breakwater sections.
During the model tests the following wave parameters were measured: the
height of overtopping y and the duration of the overtopping t (for both
sections), the tsunami run-up on the seaside slope of the barrier Y (for
nonovertopping sections), the bore height and the steepness of the bore
front S (for both sections).
Due to the uncertainty involved in predicting the shape of the tsunami with a
reasonable degree of accuracy, the design load was schematized by:
-the height y and the duration t of the overtopping (for overtopping sections)
-the maximum run-up Y (for nonovertopping sections)
Results
Overtopping breakwater sections
Three series of model tests were carried out.
In the first series the tested design sections consisted of the existing
breakwater with an armor layer of either 9 or 18 ton at a 1:1 slope added at
the harborside. These armor stones were specially placed with their
longitudinal axis approximately perpendicular to the harborslide slope. A
grout injection and a concrete slab were also added as illustrated in Figure
9.1
Seaside

Harborside

specially placed
armorstones

Figure 9.1
First tested overtopping section

Part 2 - Chapter 9 - 2

05-01-2004

It was found that the design heights of the overtopping, which the section
could withstand without damage, were about 1 m and 1.6 m for the 9 ton
and the 18 ton armor stones respectively, with the armor stones specially
placed.
Because the design overtopping height for a crest elevation of + 6 m
MLLW, is about 2 m and armor stones of 18 ton being the largest size that
can be obtained in sufficient quantities from the available quarries, it was
decided to design sections with flatter slopes than 1:1.
In the second series the sections consisted of the existing breakwater with a
slope of quarry run material ranging from 9 kg to 2 ton added at the
harborside and an armor layer of 9 or 18 ton covering the whole structure, as
presented in Figure 9.2. The slope at the harborside was varied from 1:2 to
1:7 with 9 ton armor stones and from 1:2 to 1:3 when 18 ton armor stones
were used. The armor stones were not placed with a special orientation.

Figure 9.2
Overtopping section having different harborside slopes

The maximum heights of overtopping, which would cause no damage to the test
section are presented in Table 9.1. From Table 9.1, the harborside slope and the size
of the armor stone could be selected for the design of a satisfactory section to
withstand the overtopping at a certain location along the barrier.

05-01-2004

Part 2 - Chapter 9 - 3

Slope of harborside of barrier

Height of overtopping (m)

9 ton armor stones


1:2.0

0.3

1:2.5

0.8

1:3.0

1.4

1:3.5

1.7

1:4.0

1:4.5

2.3

1:5.0

2.5

1:6.0

2.8

1:7.0

2.9
18 ton armor stones

1:2.0

1:2.5

1.2

1:3.0

1.5

Table 9.1
Maximum heights of overtopping for no damage to test section

The 3D model indicated that the design tsunami at some locations alongside
the barrier gave up to 2.7 m overtopping. A slope of 1:6 with 9 ton armor
stones is considered very expensive, a third series of tests were conducted.
A section was proposed with harborside slopes of 1:1.5 and an armor layer
consisting of stones of 27 ton, this is presented in
Figure 9.3. This is the most expensive section of the barrier, as it consists of
armor stone sizes larger than can be provided by the Hilo quarries .

Part 2 - Chapter 9 - 4

05-01-2004

Figure 9.3
End view of overtopping of the section with the 27 ton armor stones

During testing, both visual and photographic evidence indicated that failure
of the overtopping sections could be attributed to:
1) high porosity of the barrier
2) undermining of the harborside slope of the barrier
3) the method of placement of armor stone on the harborside slope of the
barrier (randomly placed or perpendicular)
4) the steepness of the harborside of the slope
Nonovertopping breakwater sections
Kamel also sets up experiments for nonovertopping breakwater sections and
investigates whether porous flow induced failure of these sections. The
right scaling rules for the modeling of these porous structures were only
recently derived. Therefore the results out of these experiments, conducted
in the seventies, for non-overtopping sections are considered inaccurate.
Therefore these results will not be discussed.
Summary and conclusions
1.

2.

The duration of overtopping and the duration of the maximum run-up


on the seaside of the barrier depend on the shape of the design tsunami
when it strikes the barrier section. Due to the uncertainty involved in
predicting the shape of the tsunami with a reasonable degree of
accuracy, it was believed that the overtopping sections should be
designed to withstand the expected height of overtopping for a
sustained duration.
The stability of a rubble mound tsunami barrier designed as an
overtopping section depends on the height of the overtopping, the slope
of the downstream side of the barrier, and the size of the armor stone on

05-01-2004

Part 2 - Chapter 9 - 5

3.

4.
5.

that side. The larger the height of overtopping, the flatter the slope and
the larger the size of the stone required for the stability of the barrier.
Failure of overtopping sections usually started above the toe of the
barrier due to washing out of some armor stones. This would be
followed by the rolling of armor stones from the downstream side of the
barrier. The longer the duration of overtopping, the greater the damage
to the section.
Efforts to develop an overtopping rubble mound tsunami barrier with a
steep harborside slope were unsuccessful. The required armor stones
were very large and initial failure triggered the failure of the entire
section.
The seaside of the tsunami barrier should be adequately designed to
withstand the effect of design short-period waves.
The use of a rubble mound barrier for the protection of Hilo harbor and
the city of Hilo from future tsunamis proved to be expensive. The great
expense involved in building such a barrier , the uncertainty involved in
predicting the design tsunami, and the catastrophe which usually results
from the failure of protection works made it impractical to employ
barriers as a means of protection of Hilo Harbor from future tsunamis.

Comments
- In the experiments where the slope was varied, no specially placed armor
stones were used, i.e. longitudinal axis approximately perpendicular to
the harborside slope. Although a strong influence can be noticed of the
placement of the stones as can be distinguished by comparison between
the results of the first set of tests (1m and 1.6m for respectively the 10
and 20 ton specially placed armor stones) and the results of the second
set of experiments with no special placement of the armor stones in table
9.1. This clearly shows the positive effect of the special placement on the
allowable overtopping height. An interesting experiment for a further
study is the overtopping on a back slope of accurately placed smooth
concrete armor units. Probably this will strongly attribute to the stability
of the armor layer. However special attention should be paid to the toe of
the backside, because the energy dissipation on the smooth slope is much
less than on a rough slope.
- The purpose of the conducted study is to design a barrier for protection of
the harbor and the city instead of solely designing the breakwater for
surviving a tsunami attack. This could contribute to the assumed high
costs for the construction of this barrier, because extra or stronger design
criteria will have to be taken into account e.g. a minimum crest height
and a harbor entrance reduction, in order to sufficiently reduce the
penetration of the tsunami into the harbor. Maybe the costs for solely
designing the breakwater for a tsunami are lower.
- Scouring behind the breakwater is not taken into account.

Part 2 - Chapter 9 - 6

05-01-2004

9.2 DYNAMIC WAVE FORCE OF TSUNAMIS ACTING ON A


STRUCTURE
Mizutani, 2001
The aim of the study is to carry out hydraulic experiments with the transient
wave-like bore in order to measure the wave force of a bore acting on
seawalls, which function as prevention structures* along the coast.
Setup
One part of the experiments was conducted for a seawall, which could not
perfectly prevent tsunami attack, a flow over the structure and flooding of
the area behind it would be expected. During the experiments the
overflowing wave pressure, which acts on the crest, the back slope and the
back area of an impervious structure was measured
An example of the overflowing wave pressure acting on a structure is
shown in Figure 9.4. For all three experiments the same wave was used, the
back slope was 1:1, but the water depth (h) was varied.

Figure 9.4
Experimental setup and the recorded overflowing wave pressure
as acting on the surface of the different parts of the structure (back slope 1:1)

Numerous prevention works, most of which are seawalls, but also some
breakwaters, were constructed since the 1960 Chilean tsunami attacked the
Japanese coastal area. The function of these structures is preventing the flooding
of the coastal area. The seawalls fulfil this function by retaining the water, the
breakwater do so by reducing the opening to a harbor or a bay.
05-01-2004

Part 2 - Chapter 9 - 7

Results
The maximum overflowing wave pressure occurs when the overflowing
wave collides on the back of the structure. A relationship between the
maximum overflowing pressure and certain parameters is proposed (based
on the law of conservation of momentum):
pom
V H t
=A m w 0
'
gH d
L t

2
sin 2
gH d'

(8.1)

Based on this formula Mizutani suggests that the important parameters for
the maximum overflowing wave pressure, pom, are the maximum velocity on
the crest, Vm, the water depth on the crest, Hw, the angle of the back slope,
2, and the height of the models Hd. The other parameters in the formula
are the non-dimensional overflowing pressure coefficient, A, a value of
0.003 is recommended. The time for the falling water to touch the bottom
after passing the top is represented as t0 and the duration of the maximum
overflow is represented as t, it is the ratio t/ t0 which influences the
maximum overflowing wave pressure (pom). The symbol L is described as
the length acting of the maximum overflowing wave pressure, what is
exactly meant by this parameter isnt clear, probably a poor translation out
of Japanese is to blame.
The relationship and the data are plotted in Figure 9.5.

Figure 9.5
Maximum overflowing wave pressure

Comments
- The high resulting pressures at the backside of the structure are an
indication for high water velocities on the back slope.
- In the experiments there was no water at the backside, with the proposed
breakwater there will be water at the backside. Can the effect on the
Part 2 - Chapter 9 - 8

05-01-2004

resulting overflowing wave pressure be quantified? Probably water at the


backside will lower the resulting wave pressures.
- In spite of some parameters, which are somewhat hard to quantify (A,
t0/t, L), one can clearly distinguish the effect of the following parameters
in relationship to the overtopping wave pressure:
- the slope at the harborside (a steeper slope leads to a higher
pressure)
- the height of the model Hd (a higher crest level leads to a lower
pressure )
- water depth on the model Hw (a higher water level leads to a higher
pressure)
- maximum velocity on the crest Vm (a higher velocity leads to a higher
pressure)

9.3 EXPERIMENTAL STUDIES ON TSUNAMI FLOW AND


ARMOR BLOCK STABILITY FOR THE DESIGN OF A
TSUNAMI PROTECTION BREAKWATER IN KAMAISHI
BAY
Hitachi, 1994
In the nineties the deepest breakwater in the world was constructed at
Kamaishi Bay, Japan with as primary objective protecting the port area from
tsunamis. The aim of this breakwater is to reduce the opening section of the
bay mouth to 5% of the original opening. Consequently strong currents are
generated here when the tsunami attacks.
Detailed information on the tsunami flow around the opening is difficult to
predict, therefore experimental studies were conducted in order to
investigate the tsunami penetration in the bay and the scouring of the rubble
mound foundation in the opening. The investigations confirmed that the
tsunami breakwater is effective in protecting the bay from tsunami attack.
The results concerning the determination of the effective weight and the
coverage area of the armor blocks to protect the rubble mound foundation
against the strong tsunami currents are briefly summarized here.
Design wave
As a design tsunami the two biggest tsunami events that occurred in the
Kamaishi region were used; the Great Sanriku tsunamis of 1893 and 1933.
Since there are no detailed records of these events, detailed wave
characteristics were estimated with a numerical simulation. As an initial
condition the estimated surface wave elevation at the tsunami source area
was used. The tsunami wave profile at the breakwater was estimated by
means of the linear long wave theory in deeper water and the shallow water
wave theory in shallow water.
The results are shown in Table 9.2, as can be seen a strong current of more
05-01-2004

Part 2 - Chapter 9 - 9

than 8 m/s is generated, due to reduction of the opening section to 5.4% of


the original.
Estimated
Design
Tsunami

Great Sanriku Tsunami in 1896


Maximum recorded runup height T.P. +7.9 m, T = 16 min
Great Sanriku Tsunami in 1933
Maximum recorded runup height T.P. +5.4 m, T = 10 min
Maximum tsunami wave height in front of the breakwater
is Hmax = 4.8 m
Maximum flow velocity is V = 8.2 m/s
Table 9.2
Estimated design tsunami

Setup
The layout of the breakwater is illustrated in Figure 9.6, Figure 9.7 and
Figure 9.8. As can be seen the opening section is a caisson-type submerged
breakwater with a crest depth of 19 m, built on a rubble mound sill. To
simulate the design tsunami, periodic flow experiments were carried out, as
well as steady flow experiments. For the unsteady flow experiments
tsunamis were schematized to a wave train of five identical waves with
periods of 10 and 20 min in combination with a peak flow velocity of 8.2
m/s and 9.8 m/s. For the steady flow experiments a flow velocity of 8.2 m/s
was used.

Figure 9.6
Front view of the breakwater

Part 2 - Chapter 9 - 10

05-01-2004

Figure 9.7
Cross Section of the opening section (submerged breakwater)

Figure 9.8
Top view of the test basin

Results
To investigate the stability of the armor stone, a tsunami current
corresponding to a tsunami with a period of T = 16 min was reproduced in
the test basin. When the armor stone S1 (gradation is illustrated in Figure
9.9) was used pronounced scouring took place, the scouring and
accumulation pattern is reported in Figure 9.9. Overturning and subsidence
of the submerged caisson was observed. In successive experiments the
armor stone S1 was replaced by blocks of 10 and 20 ton. The same scouring
and accumulation patterns were observed, but of course with smaller
quantities of moved stones for heavier blocks.
The pronounced block movement takes place at the opening section at
locations where the average flow velocity close to the bed is high and a
strong eddy patterns are formed.

05-01-2004

Part 2 - Chapter 9 - 11

Figure 9.9
Mound scouring and accumulation in the upper right corner the weight
distribution of the armor stones S1 is presented

Theoretical calculations
Calculations were conducted in order to estimate the damage ratio D for the
10 and 20 ton blocks. The following formulas were used:
the required weight of mound armor material W (kg) as determined by
C.E.R.C
W=

rU d6

48 y 6 g 3 ( r / 0 1) ( cos sin )
3

(8.2)

where Ud is the flow velocity acting on the mound material (In the presented
calculation Ud is taken as 65% of Umax at the center of the opening), r
(kg/m) is the density of the armor stones, 0 (kg/m) is the density of the
(sea)water, is the angle of the slope, y a constant expressing the stability
condition of the armor stones which can be estimated using the damage ratio
D.
the damage ratio D :
D=

Number of moved armor blocks/stones in the reference zone


(8.3)
Total number of blocks/stones in the reference zone

Part 2 - Chapter 9 - 12

05-01-2004

The relationship between the damage ratio D and y is given in Figure


9.10.The calculated damage ratio is given in Table 8.3.
The damage ratio obtained from the experiment was of the same order or
less than that calculated.

Umax (m/s)
Weight of
Block (ton)
D

Figure 9.10
Relationship between D and y.

8.2

9.8

10

20

10

20

2%

1%

20%

10%

Table 9.3
Calculated damage ratio D

Summary and Conclusions


1. Block movement is most pronounced around the opening section at
locations where the average flow velocity close to the bed is high and a
strong eddy is formed. In particular under periodic flow conditions, the
block damage ratio inside the harbor around the north breakwater is
high for positive flow.
2. For periodic flow, the block damage ratio for a given peak flow velocity
increases as the period of flow decreases.
3. The damage ratio as assumed in the design calculations is almost equal
to the experimental value. For the design flow velocity of 8.2 m/s, all
observed damage ratios were within 2%.
Comments
- It is assumed that the strong eddy pattern, generated by the two
breakwater heads, is partly responsible for the damage, overtopping of a
normal breakwater generates less eddy patterns, but probably a good
upper limit can be given with the presented formula.
- Probably the tsunami period has a strong influence on the damage ratio,
how can this be quantified in the formula?
- Our proposed breakwater isnt submerged.

05-01-2004

Part 2 - Chapter 9 - 13

9.4 CONCLUSIONS
1.
2.

3.
4.

The seaside slope of the breakwater has to be designed to withstand the


effects of design short period waves, even when the tsunami hits the
breakwater as a bore (Kamel, 1970).
The results of Kamel (1970) suggest that the stability of the stones at
the backside are determined by the velocity of the overtopping water
on the slope. The height of the overtopping water is an indication for
the amount of water overtopping the breakwater, maybe this can be
used to calculate the velocity at the backside.
The results of the water pressure measurements on an overtopped
seawall look like an indication for the high water velocities on the
backside slope.
Hitachi (1994) suggests that ordinary stability formula for flow
conditions are valid, when calculating the stability under a tsunami
wave. The period is that long that the load more resembles a steady
flow than a periodic flow.

Part 2 - Chapter 9 - 14

05-01-2004

10 THEORETICAL STABILITY CONCEPT FOR THE


RUBBLE MOUND BREAKWATER
From the literature study it is concluded that the seaside slope should be designed
for short period waves. Only the armor layer at the backside of the rubble mound
structure should be designed to withstand the tsunami load. The tsunami load on
the backside of the rubble mound structure consists of two components:
- the flow over the breakwater
- the porous flow through the breakwater structure

Figure 10.1
Schematized tsunami load on a rubble mound breakwater

The stability under overtopping flow can be quantified if one knows the governing
flow velocities of the overtopping water, the stability of the armor stones can then
be calculated with classical stability formulas for stones under flow conditions.
The calculations can be compared with the experimental results of Kamel (1970)
on overtopped breakwater structures.
The failure mechanism under porous flow is a lot harder to quantify. Also
calculating the porous flow through the highly porous rubble mound structure
poses some problems: because of the high porosity of the structure the porous
flow is no longer laminar, but becomes turbulent. The basis of calculating
turbulent flow through a porous structure is given in Barends (1981). Calculations
by hand become more complicated than for laminar flow and at the moment no
numerical models are available.
Because the governing stability criterion is taught to be the overtopping water
flow and because the porous flow poses some difficulties in quantifying and in
linking to armor stone stability, the derived theoretical stability concept will
focus on the stability of the backside armor layer under an overtopping tsunami.
Nevertheless the porous flow through the structure can play an important role in
the stability of the breakwater, therefore in the next stage of the project the porous
flow will also have to be taken into account. In this study the preliminary design
of the rubble mound structure under a tsunami loading; will be done with regard
to the stability under overtopping. It is believed that this is sufficient for a first
05-01-2004

Part 2 - Chapter 10 - 1

comparison with the caisson breakwater and to decide which of the two structures
is more favorable under a tsunami load.
In this chapter a theoretical stability concept for a rubble mound breakwater will
be set-up. This stability concept will be based on the flow stability of the armor
stones at the backside of the breakwater. The theoretical concept will be compared
with the experimental results of Kamel (1970) for overtopped breakwaters and if
necessary the model will be adapted.
At first a short review will be given of some common available formulas for flow
stability of stones. Afterwards the governing flow velocities for the stability of the
armor layer will be derived.

10.1 Stone Stability under flow


A short review will be given of some common available formula for flow stability
of stones.

10.1.1

Izbash

(Schiereck, 2001)
Izbash formulated an empirical formula to describe the stone stability under a
current. This formula can be seen as a first approximation for the stone stability
uc = 1.2 2gd

(10.1)

The critical velocity, uc, which gives a limit for stone immobility is dependent on
the relative density, =(s-w)/w and the stone diameter, d.
There is no influence of depth in this formula; in fact Izbash did not define the
place of the velocity, neither it is very clear how the diameter is defined. Further
in the text the nominal diameter dn will be used.
Sloping bed
For stones on a slope a reduction factor can be applied to the strength. In
Schiereck (2001) a reduction factor, K(//), is presented which is based on the
equilibrium of forces. In the case of a slope in the direction of the flow, the
stability factor becomes:
K ( // ) =
with:

sin( )
sin

K(//)

Part 2 - Chapter 10 - 2

(10.2)

slope reduction factor (-)


05-01-2004

:
:

angle of repose (-)


slope angle (-)

When the slope angle is taken into account, the Izbash formula (10.1) becomes;
uc = K ( // )1.2 2gd

10.1.2

(10.3)

Dutch Formula

(Lecture notes HJ Verhagen of the course Breakwaters and Closure Dams)


After completion of the Deltaworks, the Ministry of Public works evaluated all
tests and
prototype measurements done during the design and execution of the closure
works. An Izbash like formula was presented for the stability of stones on the
closure dam during closure:
d n = Au02

(10.4)

with :
A=

K2
C2

C = 18log

(10.5)
6h
dn

(10.6)

in which the nominal diameter, dn, is determined by the relative density, , the
velocity on top of the sill, u0 and a factor A. The value of A is dependent on the
parameters presented in (10.5), in which K is a correction and calibration factor
in the order of 1. To be consistent the Chezy-value, C (m1/2/s), has to be
determined using the above equation, h is the water height on top of the sill. For
the Shields parameter, , a value of 0.04 is used.

05-01-2004

Part 2 - Chapter 10 - 3

Sloping bed
The slope angle can be taken into account in the same way as for the Izbash
formula; by applying the slope reduction factor, K(//) (10.2). The Dutch formula
(10.4) then becomes:
Au02
d n = 2
K ( // )

10.1.3

(10.7)

CERC formula

Hitachi, (1994)
In the paper concerning the design of the breakwater at Kamaishi Bay (Japan)
(9.3) a formula is presented, which was used to calculate the stability of rubble
under a tsunami current. The experiments of Hitachi concern the stability of
rubble on a submerged breakwater at the entrance of the Kamaishi Bay, during a
tsunami induced current entering the bay.
A good similarity between calculations and experiments was found using the
following formula
W=

rU d6

48 y 6 g 3 ( r / 0 1) ( cos sin )
3

(10.8)

where W (kg) is the required weight of the armor material, Ud (m/s) is the flow
velocity acting on the mound material, r (kg/m) is the density of the armor
stones, 0 (kg/m) is the density of the (sea)water, is the angle of the slope, y a
constant expressing the stability condition of the armor stones which can be
estimated using the damage ratio D.
D=

Number of moved armor blocks/stones in the reference zone


Total number of blocks/stones in the reference zone

(10.9)

The relationship between the damage ratio D and y is given in Figure 10.2.

Part 2 - Chapter 10 - 4

05-01-2004

Figure 10.2
Relationship between the damage ratio (D) and the stability factor (y ).

It is referred to as a formula presented by CERC, but the only reference given by


Hitachi is to a paper in Japanese. After some investigation it was found that the
original formula is presented in the Shore Protection Manual (CERC, 1984). But
CERC (1984) presents it as a formula for stone
stability along channel
revetments (side slopes), not for stone stability on a sloping bed (flow in the
direction of the slope). So fundamentally the use of the formula (10.8) by Hitachi
(1994) is wrong, as it is used for the stability of stones on a sloping bed. However
in Japan this is a commonly used formula for stone stability on a sloping bed.
Probably because of the definition of the damage ratio and because of the
presence of a slope angle, the formula (10.8) leads to useful results. However,
scientifically it should be advised to use this formulas as it was originally
intended. Therefore the CERC formula will be presented here only as a
comparison with the results by the Izbash formula (10.3) and the Dutch formula
(10.7).

10.2 STONE STABILITY GOVERNED BY THE VELOCITY ON TOP


OF THE BREAKWATER CREST
With regard to the conclusions of the reviewed papers in chapter 9, it is
considered that the harborside of the breakwater should be designed to withstand a
certain flow velocity, caused by the overtopping tsunami.
In a first approximation the flow velocity on top of the breakwater is considered
as normative. This concept is based on the experience with closure works where
the failure starts on top of the sill.

05-01-2004

Part 2 - Chapter 10 - 5

Figure 10.3
The height of the overtopping water on the breakwater

An upper limit for the velocity on top of the breakwater is calculated using the
formula of free discharge weir:
u = m gh
(10.10)
The flow velocity u (m/s) on top of the breakwater crest is dependent on the
overtopping tsunami height h (m) and the discharge coefficient m (-). The
discharge coefficient m is chosen 0.9, a typical value for a long discharge weir.
The required nominal diameter Dn and required block weight W are calculated
using:
- the Izbash formula (10.3)
- the Dutch formula, a modified Izbash formula (10.7)
- the CERC formula as used in Hitachi (1994) (10.8)
parameters
r
0

=
=
=
=

2600 kg/m (density of the armor stones)


1030 kg/m (density of the seawater)
0.04
(shields parameter)
1% y =1.1
(damage ratio)

An example of the calculations is shown in appendix 9.

free discharge weir: in Dutch reffered to as a volkomen overlaat

Part 2 - Chapter 10 - 6

05-01-2004

9 ton
18 ton

h (m)
0,3
0,8
1,4
1,7
2
2,3
2,5
2,8
2,9

u(m/s)
1,54
2,52
3,34
3,68
3,99
4,28
4,46
4,72
4,80

1
1,2

2,82
3,09

1/ 2
1/ 2,5

1,5

3,45

1/ 3

Izbash
5,05
49,54
181,44
253,69
347,41
465,01
541,43
661,17
667,79

W (kg)
Dutch
3,52
34,72
127,39
178,12
243,94
326,56
379,80
464,35
468,79

CERC
4,35
42,69
156,35
218,60
299,36
400,69
466,55
569,72
575,43

Izbash
0,12
0,27
0,41
0,46
0,51
0,56
0,59
0,63
0,64

dn (m)
Dutch
0,11
0,24
0,37
0,41
0,45
0,50
0,53
0,56
0,56

CERC
0,12
0,25
0,39
0,44
0,49
0,54
0,56
0,60
0,60

0,46
0,38

187,03
167,19

131,19
117,34

161,16
144,07

0,42
0,40

0,37
0,36

0,40
0,38

0,32

223,17

156,65

192,30

0,44

0,39

0,42

slope angle(rad)
1/ 2
0,46
1/ 2,5
0,38
1/ 3
0,32
1/ 3,5
0,28
1/ 4
0,24
1/ 4,5
0,22
1/ 5
0,20
1/ 6
0,17
1/ 7
0,14

Table 10.1
Comparison between experimental results by Kamel (1970) and the
calculated values using the velocity on top of the sill.

10.2.1

A comparison between the calculated values

The calculations with the Izbash and the Dutch formula give nominal diameter,
Dn, in the same order. The Dutch formula systematically gives lower values than
the Izbash formula. The difference can possibly be explained by the influence of
the waterdepth h and the introduction of a stability parameter in the Dutch
formula.
The Dutch formula is in fact based on the Izbash formula, but was calibrated
during intensive experiments conducted for the Deltaworks. Therefore it is
believed that the Dutch formula gives a more accurate estimation of the required
block weight under a certain flow velocity and that the Izbash formula should be
used as a first approximation.
Although the CERC formula is originally intended for stones on channel
revetments, the CERC formula gives nominal diameters Dn in between the Dn
calculated with the Izbash and the Dutch formula. This clearly illustrates why this
formula is used successfully in Japan concerning the stone stability on slopes.

05-01-2004

Part 2 - Chapter 10 - 7

10.2.2
A comparison between the calculated values and the
experimental results
A comparison between the experimental results and the calculated values clearly
reveals a strong dissimilarity. The worst match occurs for very steep slopes and
the best match occurs for more gentle slopes. but still the difference between the
experimental result and the calculated results is more than a factor 13 (an
overtopping height of 2.9 m on a slope of 1/7 requires an armor stone of 9 tons,
where the Izbash formula gives a required block weight of 668 kg).
Possible Reasons
- Are the used formulas wrong?
No, these formulas are commonly used to calculate the stone stability under
flow conditions.
- Is the used flow velocity wrong?
It could be that the maximum velocity on top of the sill is not the normative
failure criterion for the stone stability. Possibly the flow velocity on the
backside of the breakwater is normative.
indications pro
- failure starts above the harborside toe (Kamel, 1970)
- large impact pressure peaks appear at the toe of a seawall, especially
for steep slopes, which is an indication for high water velocities on the
backside slope (Mizutani, 2001)
indications contra
- Experience with closure works indicate that failure starts on top of a
sill, so the flow velocity on top of the sill is normative.
But: The closure dams were submitted to typical tidal differences of 2
to 3 meters. Which leads to a considerable smaller level difference
over the dam than in the case of an overtopping tsunami where,
dependent on the overtopping height, a level difference of 6.5 to 9.5
meters is created (in the experiments of Kamel, 1970).

10.3 STONE STABILITY GOVERNED BY THE VELOCITY ON THE


BACKSIDE SLOPE OF THE BREAKWATER
The velocity on the breakwater crest is too low to explain the large armor stones
required at the backside of the breakwater. Therefore in this paragraph the
velocity at the backside of the breakwater will be calculated. Based on this
calculated velocity it will be tried to give an explanation for the large armor
stones at the backside of the breakwater.
Part 2 - Chapter 10 - 8

05-01-2004

10.3.1

velocity at the backside

Schttrumpf (2001) presents a method for calculating the velocity of a steady flow
on the backside slope of a dike. The steady flow on the slope is governed by the
flow rate and the flow velocity on the crest of the dike. These formulas were
theoretically derived and confirmed by model experiments. The investigations of
Schttrumpf (2001) focus on smooth dike revetments, for rough slopes he advises
the same calculation method, but no model experiments were conducted.

Figure 10.4
Waterflow at the backside of a dike

The velocity on the top vB(0) is calculated using the formula of the free discharge
weir ((10.10)), with as input the thickness of the water layer.
Velocity at the backside
k1hB
kt
.tanh 1
f
2
vB =
f .vB (0)
kt
1+
. tanh 1
hB k1
2
vB (0) +

(10.11)

with
t

k1 =

vB (0)
vB2 (0)
2 sB
+
+
2
2
g sin
g sin g sin

2 fg sin
hB

(10.12)

(10.13)

The velocity vB(sB) (m/s) is the velocity at a distance sB (m) from the crest,
measured along the slope. The thickness of the waterlayer is denoted as hB (m),
is the slope angle and f (-) represents a nondimensional friction coefficient. The
factors t and k1 are defined as is shown in equation (10.12) and (10.13).
05-01-2004

Part 2 - Chapter 10 - 9

10.3.2
comparing the calculated velocities with the experimental
results by Kamel (1970)

Figure 10.5
An overtopped breakwater structure; the maximum velocity lies
in the vicinity of the water level. h is the height of the overtopping
and 1:x is the slope.

The velocity on the given slopes were calculated for two heights under the crown:
-5m and 6m. This is in the vicinity of the water level. The velocity on top of the
crown was calculated using the formula of the free discharge weir, with as input
the given overtopping (h).
For a first comparison of the calculated velocities at the backside of the
breakwater, the critical velocity for the given stone weights on the corresponding
slopes were calculated using the Izbash formula.
The friction coefficient used was f=0.1 (proposed value for rubble, by
Schttrumpf (2001))
An example calculation is included in Appendix 9.
Warmor
(ton)

slope

overtopping
height

crown 5m
u (m/s)

crown 6m
u (m/s)

Izbash uc
(m/s)

1/2

0,3

3,42

3,43

3,61

2/5

0,8

5,14

5,18

4,50

1/3

1,4

6,26

6,35

5,10

2/7

1,7

6,53

6,63

5,54

1/4

6,76

6,86

5,87

2/9

2,3

6,96

7,07

6,13

1/5

2,5

7,03

7,13

6,33

1/6

2,8

7,05

7,15

6,63

1/7

2,9

6,89

6,98

6,85

Part 2 - Chapter 10 - 10

05-01-2004

18

1/2

6,33

6,42

4,05

18

2/5

1,2

6,12

6,2

5,05

18

1/3

1,5

6,44

6,53

5,73

Clearly the match of the critical velocity with the calculated velocities at the
backside is very promising. This is also illustrated in Figure 10.6. The critical
velocity is systematically lower than the calculated velocity, meaning that the
calculated velocity is on the safe side of the calculation.

6,00
4,00

c ritic al veloc ity


for 9 ton

2,00

c alc ulated
veloc ity (5m )

1/

1/

1/

1/

0,00

1/

ve locity (m /s)

8,00

slo p e (-)

Figure 10.6
Calculated velocity 5m under the crown vs. the critical velocity

10.3.3
comparing the required stones with the experimental
results by Kamel (1970)
The results of the velocity calculations are used as an input for stability
calculations in the same way as in 10.2. Calculations are done for velocities (u)
and layer heights (h-5 and h-6) at 5m and 6m under the crest. An example
calculation is included in Appendix 9.

05-01-2004

Part 2 - Chapter 10 - 11

h - 5 ( m)

slo p e

an g le (ra d ) Iz b a sh

W ( kg )
D u tc h

CE R C

Iz b ash

Dn (m )
Dut ch

Dn
C E RC

0,3

3 , 42

0,14

1/ 2

0,46

6 66 8,83

5 13,98

1,3 7

0,5 8

0,8
1,4

5 , 14
6 , 26

0,39
0,75

1/ 2 ,5
1/ 3

0,38
0,32

205 69 ,7 2
313 50 ,8 4

3 06 4,39
6 83 4,20

1,9 9
2,2 9

1,0 6
1,3 8

1,7
2

6 , 53
6 , 76

0,96
1,18

1/ 3 ,5
1/ 4

0,28
0,24

246 30 ,3 1
214 42 ,8 5

6 87 5,66
7 11 7,40

2,1 2
2,0 2

1,3 8
1,4 0

2,3
2,5

6 , 96
7 , 03

1,41
1,59

1/ 4 ,5
1/ 5

0,22
0,20

197 83 ,6 8
172 69 ,3 2

7 46 1,42
7 18 3,62

1,9 7
1,8 8

1,4 2
1,4 0

2,8
2,9

7 , 05
6 , 89

1,88
2,02

1/ 6
1/ 7

0,17
0,14

132 70 ,9 8
9 55 6,55

6 35 1,24
5 03 0,93

1,7 2
1,5 4

1,3 5
1,2 5

1
1,2

6 , 33
6 , 12

0,47
0,6

1/ 2
1/ 2 ,5

0,46
0,38

239 80 ,6 2
101 32 ,6 9

20 663 ,7 4
8 73 1,19

2,1 0
1,5 7

2,0 0
1,5 0

1,5

6 , 44

0,8

1/ 3

0,32

9 40 1,74

8 10 1,34

1,5 3

1,4 6

(tests)

1,91

18 ton

u(m /s)

1,51

9 ton

h (m )

Table 10.2
Stability calculations at 5m under the crest.

h -5 ( m )

slo p e

an g le (ra d ) Iz b a sh

W ( kg )
D u tc h

CE R C

Iz b a sh

Dn ( m )
Du t c h

C ERC

0,3

3 ,43

0,14

1/ 2

0,46

6 78 6,69

5 23, 06

1,3 8

0,5 9

0,8
1,4

5 ,18
6 ,35

0,38
0,74

1/ 2 ,5
1/ 3

0,38
0,32

215 4 9 ,0 5
341 5 4 ,3 1

3 21 0,28
7 44 5,33

2,0 2
2,3 6

1,0 7
1,4 2

1,7
2

6 ,63
6 ,68

0,94
1,16

1/ 3 ,5
1/ 4

0,28
0,24

269 8 1 ,8 6
199 6 4 ,6 3

7 53 2,11
6 62 6,74

2,1 8
1,9 7

1,4 3
1,3 7

2,3
2,5

7 ,07
7 ,13

1,39
1,56

1/ 4 ,5
1/ 5

0,22
0,20

217 3 5 ,4 3
187 9 6 ,6 5

8 19 7,52
7 81 8,96

2,0 3
1,9 3

1,4 7
1,4 4

2,8
2,9

7 ,15
6 ,98

1,85
1,99

1/ 6
1/ 7

0,17
0,14

144 4 1 ,2 4
103 3 0 ,4 3

6 91 1,31
5 43 8,33

1,7 7
1,5 8

1,3 9
1,2 8

1
1,2

6 ,42
6 ,2

0,46
0,59

1/ 2
1/ 2 ,5

0,46
0,38

261 0 0 ,4 7
109 5 3 ,8 4

22 49 0 ,3 8
9 43 8,76

2,1 6
1,6 2

2,0 5
1,5 4

1,5

6 ,53

0,79

1/ 3

0,32

102 1 8 ,1 5

8 80 4,82

1,5 8

1,5 0

Table 10.3
Stability calculations at 5m under the crest.

Stability calculations with the Dutch formula prove not to be possible. Calculation
with these formulas require iteration, but for these input values the iteration
process is not convergent as is shown in Appendix 10. An explanation can be
found in the small water depths (h-5 and h-6) in combination with the large water
velocities, the steep slopes and the large stone diameters. For larger water depths
the formula does lead to solutions. But the situation here exceeds the limits of
validity of the Dutch formula, probably because this formula takes into account
the place of the water velocity.
The other two stability formulas, the Izbash formula (10.3) and the CERC formula
(10.8) do give results. But also the validity of these formulas for the given
circumstances should be questioned. Also the question arises if this flow is
physical possible; these high water velocities in relative small layers on very
rough steep slopes. Can such flows be calculated with the formulas of
Schttrumpf (2001) and what is the effect of the highly porous structure?
Probably the used stability and velocity formulas are on or even beyond the limits
of their application.
Part 2 - Chapter 10 - 12

05-01-2004

Dn
(t es ts)

1,91

18 ton

u (m /s)

1,51

9 ton

h (m )

It is therefore believed that the theoretical stability concept as set up can not be
used to give a first design of the backside armor layer of the breakwater. It is
concluded that for design considerations model tests are inevitable. However,
looking at the experimental results of Kamel (1970), one can conclude that the
armor layer has to be much stronger than required for overtopping of short period
waves, the required blocks are in the order of several tons on moderate slopes. For
making a first comparison possible between the caisson breakwater and the rubble
mound breakwater the following simple failure criterion is put forward: failure of
a rubble mound breakwater as designed for short period waves occurs if
overtopping by a tsunami wave occurs.

10.4 CONCLUSIONS
From the conducted literature review it is concluded that:
- The seaside slope of the rubble mound breakwater should be designed for short
period waves, even when a tsunami hits the breakwater as a bore.
- The armor layer at the harborside should be designed to withstand the load of
an overtopping tsunami wave and the porous flow through the breakwater
structure.
The theoretical stability concept that is set up only takes into account the load of
the overtopping water. For a first base of comparison between the rubble mound
breakwater and the caisson breakwater, this is believed to be sufficient.
It is tried to set up a stability concept that can explain the experimental results by
Kamel (1970). The stability of the armor stones on the backside of the structure
can be calculated using classical stability formulas like: the Izbash formula and
the Dutch formula. Also a third formula is presented ; the CERC formula.
Originally intended for the stones on the banks of a canal, but adapted by the
Japanese for stone stability on a sloping bed. At first impression the formula looks
quite useful, but caution is needed. When applying the right governing velocities
the stability of the armor stones can be calculated using these three formulas. At
first the governing velocities are calculated using the formulas of the free
discharge weir. These velocities proved to be an underestimation. Secondly
velocity calculations were carried out with the method presented by Schttrumpf
(2001). These velocities look very promising, however the question arises if the
used velocity and stability formulas are still valid under these extreme
circumstances: high water velocities in combination with the small water depths,
the steep slopes and the large required armor stones? The Dutch formula is clearly
beyond its scope as the iterative process is no longer convergent. The Izbash
formula and the CERC formula do give results, but one can question the physical
validity of the results. Therefore for the rubble mound breakwater a simplified
stability criterion is put forward: failure occurs if the breakwater structure is
overtopped by the tsunami wave. For a more detailed stability criterion further
study is needed and model tests will be inevitable. Special attention should be
paid to the correct modeling of the porous flow through the breakwater structure.
05-01-2004

Part 2 - Chapter 10 - 13

11 DESIGN OF THE RUBBLE MOUND BREAKWATER


The simplified stability criterion will be tested for the schematized breakwater
profile at Pie del Gigante.

11.1 The rubble mound breakwater for Pie del Gigante

Figure 11.1
Schematized rubble mound breakwater for Pie del Gigante as designed for short period
waves.

The breakwater profile was slightly adapted: the seaside berm was removed. This
was done to make a correct representation of the structure possible in the
mathematical model. In appendix 11 the original breakwater design is presented.

11.2 The design tsunami wave


The run-up of the tsunami wave on the rubble mound breakwater will be modeled
using the 1D mathematical model as set up in chapter 5. As design water level the
HHWL is taken. The height of the breakwater crest is 4.2m above HHWL. The
calculations for the design wave are presented in Figure 11.2. The wave causes a
run-up till 3.1m so the wave does not overtop the breakwater.

05-01-2004

Part 2 - Chapter 11 - 1

Design wave run-up on the rubble mound breakwater


3,5
3
2,5
height (m)

2
1,5
1

design wave

0,5
0
2002,5
-0,5

2127,5

2252,5

2377,5

2505

2630

2755

2880

-1
-1,5
time (s)

Figure 11.2
Design wave for the rubble mound breakwater. The maximum water elevation is 3.1 m.
The breakwater crest lies 4.2m above the SWL, so no overtopping occurs

A sensitivity analysis shows that overtopping occurs when the amplitude of the
bottom deformation becomes larger than 1.3 times the original amplitude
(+0.82m/-0.33m). Variation in the length of the bottom deformation also
influences the run-up as shown in (Figure 11.3). For different amplification
factors the resulting wave run-up is presented. The differences are probably
mostly attributed to differences in superpositioning of individual waves. A
maximum run-up occurs for 0.8 times the length of the initial bottom deformation
(40km/27.5km). But overtopping does not occur.
Run-up on the rubble mound breakwater
0.9L

4
3,5

0.6L

1.2L

3
height (m)

0.8L

1L

1.2L

0.4L

2,5

1L
0.9L

0.8L

1,5

0.6L

0.4L

0,5
0
2002,5

2127,5

2252,5

2377,5

2505

2630

2755

2880

time (s)

Figure 11.3
Run-up of the tsunami wave for different lengths of the bottom deformation
L=(40km/27.5km)

Part 2 - Chapter 11 - 2

05-01-2004

11.3 Conclusions
It can be concluded that the rubble mound breakwater as designed for short period
waves, is not overtopped by the design tsunami wave. With regard to the
simplified stability criterion (overtopping = failing) the breakwater can withstand
the design tsunami wave. The structure will only be overtopped by the design
waves with an amplitude larger than 1.3 times the initial bottom deformation
(+0.82cm/-0.33cm). So a large safety margin exists. But this only takes into
account the simplified failure criterion. Porous flow is not taken into account,
maybe porous flow can trigger failure before overtopping occurs. And does the
harborside armor layer indeed fails when overtopping occurs? Because the armor
stones at the harborside are already very large (30 ton), compared to the model
tests of Kamel (1970). On the other hand the slopes are much steeper than in the
experiments of Kamel (1970).
Obviously more study is needed to be able to give a correct design for the rubble
mound breakwater under a tsunami load and model tests will be inevitable.

05-01-2004

Part 2 - Chapter 11 - 3

12 CONCLUSIONS
Aim of Part 2: Formulate a breakwater design which can withstand the
design tsunami derived in part 1
In the literature study of the caisson breakwater, design formulas for unbroken
tsunami waves are formulated (Tanimoto, 1981). The design formulas are based
on the assumption that unbroken tsunami waves impose a hydrostatic load on
the caisson. The method is similar to Godas design principles for unbroken short
period waves: a pressure distribution on the caisson is used to calculate the
resulting forces on the caisson, based on the resulting forces the safety factors
against sliding and overturning are calculated.
The hydrostatic assumption is believed to be correct as long as no overtopping
occurs. If the tsunami wave overtops the caisson the hydrostatic assumption is
believed to provide an upper limit for the tsunami load. Therefore the design
concept as formulated by Tanimoto (1981) is sufficient for the preliminary design
of the caisson breakwater.
From the literature study concerning rubble mound breakwaters, no ready to use
design formulas can be formulated. However, from the experiments of Kamel
(1970), it can be concluded that the seaside slope of the breakwater should be
designed for short period waves, even when the tsunami hits the breakwater as a
bore. The harborside of the breakwater should be designed on a tsunami load. This
load consists of two components: the flow of the overtopping water on the
harborside and the porous flow through the breakwater structure. Experiments
conducted by Hitachi (1994) show that, due to the long period of the tsunami
wave, the flow under the wave can be characterized by a steady flow and the
stability under a tsunami current can be calculated by stability formulas for stones
under flow conditions. Mizutani (2001) conducts experiments which indicate that
large water velocities occur on the backside of a seawall, when a tsunami wave
overtops the seawall.
A theoretical stability concept is set up which tries to explain the experimental
results of Kamel (1971) concerning overtopping rubble mound breakwaters.
However the question arises if the used velocity and stability formulas are still
valid under these extreme conditions i.e. the high resulting flow velocities in thin
water layers on a steep slope consisting of large armor stones. It is therefore
believed that the theoretical stability concept as set up can not be used to give a
first design of the backside armor layer of the breakwater. It is concluded that for
design considerations model tests are inevitable. However, looking at the
experimental results of Kamel, one can conclude that the armor layer has to be
much stronger than required for overtopping of short period waves. The required
blocks are in the order of several tons on moderate slopes. For making a first
comparison possible between the caisson breakwater and the rubble mound
breakwater the following simple failure criterion is put forward: failure of a
rubble mound breakwater as designed for short period waves occurs if
overtopping by a tsunami wave occurs.
05-01-2004

Part 2 - Chapter 12 - 1

Following from the design calculations for the caisson breakwater it can be
concluded that the design for short period waves is able to withstand the load
imposed by the design tsunami wave (Figure 12.1).
Design wave for the caisson breakwater
3
2,5

height (m)

2
1,5
1

design wave

0,5
0
2002,5
-0,5

2127,5

2252,5

2377,5

2505

2630

2755

2880

-1
-1,5
time (s)

Figure 12.1
Design wave for the caisson breakwater. The positive reflecting wave height is
2.7m.

Failure of the caisson breakwater occurs for a fictive wave with a positive
reflecting wave height Hrefl=6.2m and a water lowering hL=1m or with Hrefl=5.2m
and hL=3m. So clearly the design for short period waves is sufficiently stable for
the design considerations. The waves which can trigger failure of the caisson are
waves generated by bottom deformations approximately 2 times the initial
amplitude of +0.82m and 0.33m. So the breakwater has a sufficient large margin
to take into account the uncertainty concerning the design tsunami wave.
For the design of the rubble mound breakwater the failure criterion was
simplified: failure starts when the tsunami wave overtops the breakwater structure.
The design tsunami wave does not overtop the rubble mound breakwater structure
(Figure 12.2), only for a wave with an initial bottom deformation higher than 1.3
times the original amplitude (+0.82m/-0.33m).

Part 2 - Chapter 12 - 2

05-01-2004

Design wave run-up on the rubble mound breakwater


3,5
3
2,5
height (m)

2
1,5
1

design wave

0,5
0
2002,5
-0,5

2127,5

2252,5

2377,5

2505

2630

2755

2880

-1
-1,5
time (s)

Figure 12.2
Design wave for the rubble mound breakwater. The maximum water elevation is 3.1
m. The breakwater crest lies 4.2m above the SWL, so no overtopping occurs

Both breakwater types fulfill the stability conditions against the design tsunami
wave. However, the caisson breakwater has a larger margin than the rubble
mound breakwater. This is important because of the large uncertainty with regard
to the design tsunami wave.
Of course for the rubble mound breakwater only a very simplified failure criterion
is taken: if overtopping occurs the breakwater fails. Maybe porous flow could
initialize failure without overtopping occurring. And if overtopping occurs, is the
backside layer indeed not strong enough to prevent failure? These two questions
can only be answered with further research, and it is believed that model tests are
a required tool for the further study of a rubble mound breakwater.
Therefore it is concluded that the caisson breakwater is preferred above the
rubble mound breakwater. Not only does the caisson breakwater have a larger
margin of safety against the design criterion used, also the stability concept has
a solid basis. Whereas for a correct design of the rubble mound breakwater further
research is needed. This is needed to gain insight in the failure mechanism and to
set up a correct design criterion for the rubble mound breakwater.

05-01-2004

Part 2 - Chapter 12 - 3

13. Summary and Conclusions


Breakwater stability under tsunami attack
for a site in Nicaragua
Aim: Formulate a breakwater design that can withstand a realistic tsunami
wave for a location at the Pacific Coast of Nicaragua
This study consists of two parts: one part in which a realistic design tsunami is
derived and another part in which the breakwater stability under this design
tsunami is investigated.

Part I Design Tsunami


Aim of Part 1: Formulate a realistic design tsunami for two harbor locations
At the Pacific Coast of Nicaragua

Figure 13.1
The two harbor locations at the Nicaraguan Coast, Pie del Gigante and Bahia del
Salinas.

Before a design tsunami can be derived, the necessary insight in tsunami science
is required: what is a tsunami, how can it be described and how can it be
modeled? An extensive literature study was conducted. In chapter 1 and 3 a
summarized overview is given of tsunami science relevant to this study. These
chapters provide a solid base for further research, including the necessary
references. For a full overview, please refer to chapter 1 and 3.
05-01-2004

Chapter 13 - 1

In the following text the most important findings are presented:


- A tsunami is a series of ocean waves generated by any rapid large-scale
disturbance of the seawater.
- Most tsunamis are generated by bottom deformation caused by earthquakes.
- Most tsunamis arrive at the coast unbroken; they come in like a fast rising or
falling tide. The small number of tsunamis that do break often form vertical
walls of turbulent water called bores.
- Most of the damage inflicted by tsunamis is caused by strong currents and
floating debris. When the tsunami arrives as a bore, also the impact of the bore
is important.
After the overview of tsunami science, the derivation of a design tsunami can
start. A full statistical approach for defining a design tsunami proves not to be
possible. Therefore, defining the design tsunami is based on the biggest tsunami
event that occurred in the Central American region: the 1992 Nicaraguan tsunami.

Occurrence of the 1992 Nicaraguan tsunami


Large tsunami events like the 1992 Nicaraguan tsunami are infrequent
phenomena. On one hand this is the biggest tsunami that occurred in Central
America during the past century, on the other hand only a limited strip of coast is
actually struck by such a large tsunami. The 1992 Nicaraguan tsunami only
affected 250 km coastline. For the 1992 tsunami this gives a return period for one
location along the Middle American Trench, between 1360 and 13600 years, with
a best fit of 1 in 1920 years. This is only a rough estimation, but it clearly
illustrates that large tsunamis are infrequent phenomena at the Pacific coast of
Central America.
Looking at the entire Pacific region, Central America is only a modest region with
regard to tsunami generation. During the period 1900 1980 only 6.5% of all
generated tsunamis in the Pacific were generated along the Central American
Coast (Hatori, 1995). When taking into account all the tsunamis with an equal or
higher intensity than the 1992 Nicaraguan tsunami, it is confirmed that Central
America is only a modest region with regard to the occurrence of tsunamis. But
the 1992 Nicaraguan tsunami is a large tsunami event, even with regard to the
entire Pacific.
Conclusions with regard to the 1992 Nicaraguan tsunami
-With regard to one location this is a rare event (rough estimation of 1 on 1920
years)
-This was a large event, even with regard to the entire Pacific region

Design tsunami
The design tsunami is based on the 1992 Nicaraguan tsunami. With a 1D
mathematical model the 1992 Nicaraguan tsunami is derived at Pie del Gigante.
The modeling of the 1992 tsunami is mainly based on the model of Satake (1995),
Chapter 13 - 2

05-01-2004

but also tide gauge recordings at Corinto, the reported arrival times and the
reported run-up heights were taken into account. The first wave that arrived on
shore was reported as the highest wave, therefore the modeling efforts are focused
on the first wave.
The model was set up starting from the initial bottom deformation as derived by
Satake (1995). The height of the reflecting wave on the closed coast boundary was
calibrated on 2m. Looking at the results of Satake (1995), this can be regarded as
an upper limit. Satakes results also confirm what is already expected: due to the
straight coast, local amplification of the wave height is very limited (Figure 13.2).
The calculated wave form is a so called leading depression N wave, with a small
through followed by a large water elevation (Figure 13.3). This corresponds to the
tide gauge recording for Corinto (Figure 13.4). Defining a wave period is not
simple, due to the superimposed second wave peak.
It is clear that the modeling of a tsunami, in this study as well as in publications, is
not very accurate. The accuracy is not in the order of a few centimeters or a
decimeter, but more in the order of a few decimeters.

Figure 13.2
Calculated reflected wave heights of the 1992 Nicaragua tsunami by Satake (1995).

c allib rated w ave


2.5

reflecting wave

he ight (m)

1.5
wave at -10m (no refl ecting)

reflecting wave

0.5
0
20 02. 5 2102.5 2202 .5 23 02.5 2402.5
-0.5

2505

26 05

2705

2805

29 05

tim e (s)

Figure 13.3
Calibrated wave at Pie del Gigante. Two waves are presented, the reflecting wave at
the closed coast boundary and the wave at -10m with an open coast boundary.

05-01-2004

Chapter 13 - 3

Figure 13.4
Tide gauge recordings of the 1992 Nicaragua tsunami at Corinto after the tidal
components are removed.

A sensitivity analysis of the input parameters shows that the wave is almost
insensitive for large variations of the bathymetry (depths of the ocean trench,
steepness of the continental shelf and the steepness of the foreshore ( 50 to 0m)).
Also the friction only plays a limited role, probably due to the large water depths.
The resulting wave is mainly determined by the length and the amplitude of the
bottom deformation.
Based on the calibrated model the design wave for Pie del Gigante is determined.
The reflecting wave height and the wave period of the calibrated model are
increased by 20%. The resulting bottom deformation is then used as an input in
the 1D model at Bahia del Salinas. Calculations with the 1D model hardly give
any difference between the two locations. For design considerations, the
calculations will therefore be limited to one location: Pie del Gigante.
In order to accurately represent the rubble mound breakwater, the gridsize near the
coast is refined from dx=500m to dx=6m. The smaller gridsize leads to a larger
influence of the bottom friction, causing a significant lower wave to arrive at the
coast. This lowering proves to be convergent, as gridsizes smaller than dx=6m
give a negligible lowering. Therefore the model is recalibrated for a gridsize of
dx=6m. The resulting design wave is presented in Figure 13.5 .
Design wave for Pie del Gigante (dx=6m)
3

reflecting wave

2,5

height (m)

2
1,5

wave at -10m (no reflection)


reflecting wave

1
0,5
0
2002,5
-0,5

2127,5

2252,5

2377,5

2505

2630

2755

2880

time (s)

Figure 13.5
Design wave for Pie del Gigante

Chapter 13 - 4

05-01-2004

Design wave
reflecting wave
Max elevation

2.4 m

arrival time

45 min

wave at -10 m (no reflection)


peak-to-through

2.05 m

period

11.2 min

Remark: in a further study calculations with a 2D model have to be carried out to


investigate if complex reflection and amplification occurs at the two harbor
locations.
Conclusions with regard to the design wave
The design wave is calibrated on a reflecting wave height of 2.4m on the closed
coast boundary, with a period of 11.2min for the wave at a depth of 10m. This
results in a wave height of 2.05m at 10m and an arrival time on shore of 45min.
Because of the large uncertainty with regard to the tsunami wave it is proposed to
conduct a large sensitivity analysis in the design calculations.
An important conclusion, for further stability analysis, is that even large tsunami
waves will not break on the Nicaraguan coast.

Part II: Breakwater Stability


Aim of Part 2: Formulate a breakwater design which can withstand the
design tsunami derived in part 1

Stability Concept for the Caisson and the Rubble Mound Breakwater
In the literature study of the caisson breakwater, design formulas for unbroken
tsunami waves are formulated (Tanimoto, 1981). Based on model experiments
Tanimoto concludes that the tsunami wave load on a caisson manifests itself as a
hydrostatic load. Based on these findings Tanimoto, assumes a pressure
distribution on the caisson. The resulting forces are then used to calculate the
safety factors against sliding and overturning. As an input the reflecting wave
height on a vertical wall is taken. The vertical wall should be positioned at the
breakwater location.
The method is similar to Godas design principles for unbroken short period
waves. It is shown that Tanimotos design formulas for unbroken tsunami waves
05-01-2004

Chapter 13 - 5

are in fact an end form of the Goda design principles for unbroken short period
waves. A tsunami wave imposes a bigger load than a short period wave with the
same crest elevation.
The hydrostatic assumption is believed to be correct as long as no overtopping
occurs. If the tsunami wave overtops the caisson the hydrostatic assumption is
believed to provide an upper limit for the tsunami load. Therefore the design
concept as formulated by Tanimoto (1981) is sufficient for the preliminary design
of the caisson breakwater.
Out of the literature study regarding rubble mound breakwaters, no ready to use
design formulas can be formulated. However from the experiments of Kamel
(1970) it can be concluded that the seaside slope of the breakwater should be
designed for short period waves. Even when the tsunami hits the breakwater as a
bore. The harborside of the breakwater should be designed on a tsunami load. This
load consists of two components: the flow of the overtopping water on the
harborside and the porous flow through the breakwater structure. Experiments
conducted by Hitachi (1994) show that, due to the long period of the tsunami
wave, the flow of the wave can be characterized by a steady flow and the stability
under a tsunami current can be calculated by stability formulas for stones under
flow conditions. Mizutani (2001) conducts experiments which indicate that large
water velocities occur on the backside of a seawall, when a tsunami wave
overtops the seawall.
Based on these findings, a theoretical stability concept is set up. The stability
concept focuses on the stability of the harborside armor layer of a breakwater,
under the flow caused by an overtopping tsunami wave. For a first calculation
porous flow is neglected: quantifying the porous flow through a highly porous
structure is complex and linking to breakwater stability is not easy. Therefore, in
this first assumption the overtopping tsunami flow is considered as the governing
stability factor for the armor layer at the backside of the breakwater structure. In
further research the porous flow should also be taken into account.
A theoretical stability concept is set up that aims to explain the experimental
results of Kamel (1971). Kamel sets up model tests for overtopping tsunami
waves on rubble mound breakwater structures. At first the results are explained in
a similar way as for closure works: the governing flow velocity is calculated by
the formula for the free discharge weir and the stone stability is calculated with
classical stability formulas for stones under flow conditions. This clearly gives an
underestimation of the required stones. Secondly, the velocity on the backside
slope is calculated with the velocity formulas for steady flow on a dike slope as
presented by Schttrumpf (2001). The velocity in the vicinity of the water level is
taken as input for the stability formulas.
However, the question arises if the used velocity and stability formulas are still
valid under these extreme conditions i.e. the high resulting flow velocities in thin
water layers on a steep slope consisting of large armor stones. It is therefore
believed that the theoretical stability concept as set up can not be used for a first
design of the backside armor layer of the breakwater. It is concluded that for
design considerations model tests are required. However, looking at the
experimental results of Kamel, one can conclude that the armor layer has to be
Chapter 13 - 6

05-01-2004

much stronger than required for overtopping of short period waves; the required
blocks are in the order of several tons on moderate slopes. To enable comparison
between the caisson breakwater, and the rubble mound breakwater the following
simple failure criterion is put forward: failure of a rubble mound breakwater as
designed for short period waves occurs if overtopping by a tsunami wave occurs.
Conclusions with regard to the stability concept
- From the literature study design formulas are presented for the caisson
breakwater.
- For the rubble mound breakwater a simplified stability criterion is put forward:
failure occurs if the breakwater structure is overtopped by the tsunami wave. For
a more detailed stability criterion further study is needed and model tests will be
required.

Design Calculations
From the design calculations for the caisson breakwater it can be concluded that
the design for short period waves is able to withstand the load imposed by the
design tsunami wave (Figure 13.6).
Design wave for the caisson breakwater
3
2,5

height (m)

2
1,5
1

design wave

0,5
0
2002,5
-0,5

2127,5

2252,5

2377,5

2505

2630

2755

2880

-1
-1,5
time (s)

Figure 13.6
Design wave for the caisson breakwater. The positive reflecting wave height is
2.7m.

Failure of the caisson breakwater occurs for a fictive wave with a positive
reflecting wave height Hrefl=6.2m and a water lowering hL=1m or with Hrefl=5.2m
and hL=3m. This is significantly higher than the design tsunami characteristics
(Hrefl=2.7m and hL=0.4m). So clearly the design for short period waves is
sufficiently stable under the design considerations. The waves, which can trigger
failure of the caisson, are waves generated by bottom deformations approximately
twice the initial amplitude of +0.82m and 0.33m. The breakwater has a
sufficient large margin to take into account the uncertainty concerning the design
tsunami wave.
05-01-2004

Chapter 13 - 7

For the design of the rubble mound breakwater the failure criterion was
simplified: failure starts when the tsunami wave overtops the breakwater structure.
The design tsunami wave does not overtop the rubble mound breakwater structure
(Figure 13.7). Overtopping only takes place for a wave with an initial bottom
deformation higher than 1.3 times the original amplitude (+0.82m/-0.33m).
Design wave run-up on the rubble mound breakwater
3,5
3
2,5
height (m)

2
1,5
1

design wave

0,5
0
2002,5
-0,5

2127,5

2252,5

2377,5

2505

2630

2755

2880

-1
-1,5
time (s)

Figure 13.7
Design wave for the rubble mound breakwater. The maximum water elevation is
3.1m. The breakwater crest lies 4.2m above the SWL, so no overtopping occurs

Both breakwater types fulfill the stability conditions of the design tsunami wave.
However, the caisson breakwater has a larger safety margin than the rubble
mound breakwater. This is important because of the large uncertainty with regard
to the design tsunami wave.
Of course for the rubble mound breakwater only a very simplified failure criterion
is taken: if overtopping occurs the breakwater fails. Maybe porous flow can
initialize failure without overtopping occurring and if overtopping occurs, is the
backside layer indeed not stable enough to prevent failure? These two question
can only be answered with further study, and it is believed that model tests are a
necessary tool for the further study of a rubble mound breakwater. Therefore it is
concluded that the caisson breakwater is preferred upon the rubble mound
breakwater. The caisson breakwater does not only have a larger margin of safety
against the design criterions used, also the stability concept used has a solid basis.
Whereas for a correct design of the rubble mound breakwater further study is
needed. This further study is needed to gain insight in the failure mechanism and
to set up a correct design criterion for the rubble mound breakwater.

Chapter 13 - 8

05-01-2004

Conclusions with regard to the breakwater stability


It is concluded that the caisson breakwater as designed for short period waves, is
the best alternative of the two breakwater types. Not only does it have a larger
safety margin than the rubble mound breakwater. The stability concept of the
caisson breakwater is also more accurate than the simplified stability criterion for
the rubble mound breakwater. For a correct insight in the rubble mound
breakwater stability under a tsunami wave further research is needed.

05-01-2004

Chapter 13 - 9

14 Recommendations
Concerning the design wave
For a future study it is believed that the 1D approach used in this thesis for deriving
a design tsunami is appropriate. Starting from an initial bottom deformation the
wave is derived to the coast with a mathematical model based on the shallow water
equations. However, local bathymetry can seriously amplify the occurring wave at
one location. Therefore, for further research a 2D model based on the same
principle as the 1D model is required. For the bottom deformation, the same
amplitude and length as in the 1D model should be used, the width should be taken
out of the study of Satake (1995). By setting up the model with a sufficiently fine
grid near the proposed harbor locations (Pie del Gigante and Bahia del Salinas), the
most favorable of the two locations can be determined. The calculations with the
2D model should be carried out for different positions of the bottom deformation
along the Middle American Trench, in order to investigate the effect on the arriving
waves at the two harbor locations. Also a large sensitivity analysis should be
carried out concerning the amplitude and the length of the bottom deformation.
Rubble mound breakwater stability
For further insight in the failure mechanism of the rubble mound breakwater under
a tsunami load, model tests will have to be carried out. Special attention should be
paid to the correct modeling of the porous flow through the breakwater structure.
Two situations have to be investigated:
- the tsunami does not overtop the breakwater structure;
- the tsunami wave overtops the breakwater structure.
For this kind of stability tests an unbroken tsunami load can be modeled as a fast
rising or falling water level.
Effects of a tsunami on harbors
It is hard to predict what will actually happen when a tsunami hits a harbor, but
looking at reports from large tsunami events the following remarks can be given.
Breakwaters will not prevent tsunami waves from entering the harbor, due to the
long wave length. Tsunamis, even small ones, can cause large seiches in harbor
basins which can persist for many hours. These seiches can cause serious damage
to moored boats by rocking them against the quay walls. The risk for the operating
structures of the container terminal can be lowered by building the quay wall at a
sufficient high level above the HHWL. This way, the amount of overtopping will
decrease or no overtopping will occur at all. Less overtopping water leads to
smaller buoyancy forces and smaller velocities, so the damage will be reduced. The
required height of the quay walls should be investigated with a 2D model of the
harbor with the tsunami wave as input. A first estimate, based on the 1D design
wave, is that quay wall with a height of +3.00m HHWL will prevent overtopping.

05-01-2004

Chapter 14 - 1

Appendices:
1. Databank of Tsunamis at the Pacific side of the
Central American Region
2. A Short Introduction to Seismology
3. Mail correspondence concerning the deriving of
tsunami wave characteristics out of tsunami data or
seismic data
4. The 1992 Nicaraguan tsunami
5. Bathymetry
6. Preliminary design of the Caisson Breakwater at Pie
del Gigante
7. Comparing the Tanimoto design formulas for
unbroken tsunami waves with the Goda design
formulas for unbroken short period waves
8. Calculations of the caisson stability under the design
tsunami wave
9. Calculations concerning chapter 10: Theoretical
stability concept for the rubble mound breakwater
10. Iteration problems of the Dutch formula
11. Rubble Mound breakwater for Pie del Gigante

Appendix 1: Databank of Tsunamis at the Pacific side of


the Central American Region
The dataset for the tsunamis at the pacific side of the Central American Region are obtained
out of the Historical Tsunami Database for the Pacific, 47 B.C. to present.

1.

Historical Tsunami Database for the Pacific, 47 B.C. to Present

The Historical Tsunami Database for the Pacific, 47 B.C. to present is the joint project of the
IUGG/TC and ICG/ITSU The project was launched in 1997 and the purpose was to improve
the situation with catalogization of historical tsunamis in the Pacific. By means of organizing
them in the form of the database containing the comprehensive historical tsunami catalog in
the Pacific.
This database summarizes the long-term efforts of many research groups and individuals in
collecting, refining and digitizing the tsunami related data and represents the most updated,
revised and homogenous tsunami data set.
The database is accessible at a site of the Tsunami Laboratory of Siberian Division, Russian
Academy of Sciences : http://tsun.sscc.ru/htdbpac
The specially developed data management software allows for a remote user to make data
retrieval by complex criteria and to obtain the resulting list of tsunamigenic events with their
basic source parameters.

2.

Limiting Criteria

The criteria which were used to retrieve the data are listed below.
Location: the vicinity of Central America
0-35 NB
75-125 WL
In the query a region slightly larger than Central America (Guatemala-Panama) is selected.
This to investigate the influence of the bordering regions.
Period: 1900-2002
Cause: no limitations

IUGG/TC: the tsunami commission of the International Union of Geodesy and Geophysics
ICG/ITSU: International Coordination Group for the Tsunami Warning System in the Pacific

3.

Legend

Blanks in columns indicate that the value of this parameter is unavailable for this particular
event.
year - year [-47, 1999]
mo - month [1,12]
da - day [1,31]
hr - hour (in GMT) [0,23]
mn - minute [0,59]
sec - second [00.0,59.9]
lat - latitude in ('-' for southern latitude) [-65.00,65.00]
long - longitude ('-' for western longitude) [-80.00, 50.00]
dep - source depth in km [0,999]
Ms - surface-wave magnitude [0,9.9]
Mw - moment magnitude [0,9.9]
Mt - Abe's tsunami magnitude [0,9.9]
Int - tsunami intensity on Soloviev scale [-9.9,9.9]
Hmax - maximum observed or measured wave height in meters [0.01,999.99]
N - total number of available run-up and tide-gauge observations [1,999]
D - damage code
N - nondamaging event
S - slight damage
M - moderate damage
L - large (severe) damage
F - number of reported fatalities due to the event [0,99999]
C - cause of the tsunami
T tectonic
V volcanic
L landslide
M meteorological
S seiches
E explosion
I impact
U unknown
TR - tsunamigenic region code
A-A - Alaska and US Pacific coast (35N - 63N, 167E - 112W)
CAM - Central America (7 N - 35 N, 125 W - 75 W)
SAM - South America (58 S - 7 N, 100 W - 60 W)
NZT - New Zealand and Tonga (57 S - 11 S, 160 E - 166 W)
NGS - New Guinea and Solomon Is. (11 S - 5 N, 130 E - 165 E)
IND Indonesia (11 S - 11N, 92 E - 130 E)
PHI Philippines (5 N - 28 N, 104 E - 134 E)
JAP Japan (21 N - 48 N, 114 E - 156 E)
K-K - Kuril-Kamchatka (40 N - 63 N, 131 E - 167 E)
HAW Hawaii (15 N - 35 N, 171 E - 145 W)
Source Region - source region of the tsunami. Descriptive indication of the tsunami source
area
2

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35

time

1906
1907
1928
1928
1930
1932
1932
1932
1934
1941
1941
1948
1950
1950
1950
1951
1957
1958
1962
1962
1962
1965
1973
1979
1979
1981
1985
1985
1988
1992
1994
1995
1995
1996
1999

1
4
3
6
8
6
6
6
7
12
12
12
10
10
12
8
7
1
3
5
5
8
1
3
12
10
9
9
4
9
1
9
10
2
9

31
15
22
17
31
3
18
22
18
5
6
4
5
23
14
3
28
19
12
11
19
23
30
14
12
25
19
21
30
2
17
14
9
25
12

36
8
17
19
40
36
12
59
36
47

22
9
13
15
24
40
7
40
11
58
46
1
7
59
22
17
37
42
16
30
4
35
8

15
6
4
3
0
10
10
12
1
20

0
16
16
14
0
8
14
11
14
14
19
21
11
7
3
13
1
15
0
12
14
15
3
18

2.7
55.5
33.9
54.1
16.6

9.8
25
15.3
56.8
13.1
1.5
13.5
15
4.6
15.8
49.7
13.5

9.6
31.8
36.6
54.2
17.8
27.5
27.8
3.3

Year Mo Da Hr Mn Sec

date

1
16.7
16.84
16.65
33.9
19.46
19.53
19.09
8.01
8.61
8.5
21.6
10.4
14.3
16.3
13
16.88
0.99
8.09
17.14
16.99
16.17
18.45
17.77
1.6
18.12
18.46
17.83
32.6
11.72
34.16
16.84
19.05
15.93
17.62

Lat
-81.5
-99.2
-96.02
-96.5
-118.6
-104.17
-103.72
-104.41
-82.56
-83.18
-84
-106.7
-85.2
-91.8
-98.2
-87.5
-99.29
-79.48
-82.68
-99.7
-99.69
-95.85
-102.96
-101.23
-79.36
-102
-102.37
-101.62
-117.3
-87.39
-118.57
-98.61
-104.21
-98.1
-101.6

Long

generation point

tsunami data

45
14
29
26
17

25
25
15
2
0
25
71
1
25
32
33
33
60
30
50
100
36
20
19
46
16
10
37
24
24
21
21
18
7.2
6.8
7.2
7.3
6.9

8.6
8.2
7.5
7.8
5.2
8.1
7.8
6.9
7.7
7.5
7
7
7.7
7.1
7.5
7
7.9
7.3
6.8
7.2
7.1
7.8
7.5
7.6
7.7
7.3
8.1
7.5
7.7
6.7
7.3
8
7.1

7.5
7.6
7.6
8.1
7.2
8
7.5

7.7

7.7
7.5
7.1
9

7.9

7.8
7.5

8.2
7.8

7.6
7.5
7.6
7.3

8.1

8.7

8.7
7.8
7.4
7.9

2.8
-3
-1
2
-3
2.5

-1
-3
-1.1
2.5
-3
1.5
0

1
-0.5
-0.5
-2
1.5
1
-3
-3
1.5
-2
-1
-1
0
1.5
1.5
-3
0

3
1

0.4
0.1
0.4
5
0.1
3
1.2
5
10.7
0.1
0.42
5
0.12
8

0.1
0.8

0.1
0.2
0.3
10
2.6

2.5
6.1
0.7
0.1
10
2
0.1
0.1

5
2

S
S
N
N
S
N
N
N
M
N
M
S
N
M
N
N
M
N
S

N
N
S
N
N
N

S
M
N
N
M

M
S

Dep Ms Mw Mt Int Hmax D

earthquake data

4, Tsunamis at the Pacific Side of the Central American Region from 1900 to 2002

0
170
0
0
1
0
0

56
5
600

4
30

68
20

1000

T
T
T
T
L
T
T
T
T
T
T
T
T
T
T
T
T
T
T
T
T
T
T
T
T
T
T
T
M
T
T
T
T
T
U

SAM Columbia-Ecuador:TUMACO
CAM Mexico
CAM MEXICO
CAM MEXICO: NEAR COAST OF GUERRER
CAM S.CALIFORNIA
CAM Central Mexico,Jalisco
CAM Central Mexico, Jalisco
CAM Central Mexico
CAM Costa Rica-Panama
CAM Central America
CAM Costa Rica-Panama
CAM Mexico:MARIA MADRE ISLAND
CAM Nicaragua,Costa Rica-Panama
CAM Guatemala-Nicaragua
CAM Acapulco, S.Mexico
CAM Potosi, Honduras
CAM S.Mexico:GUERRERO
SAM Colombia-Ecuador
CAM Costa Rica-Panama
CAM S.Mexico
CAM S.Mexico
CAM MEXICO
CAM S.Mexico:FARIAS,TECOMAN
CAM S.Mexico:GUERRERO
SAM Colombia-Ecuador
CAM MEXICO
CAM MEXICO: MICHOACAN: MEXICO CI
CAM MEXIC0: SW COAST: MEXICO CIT
CAM San Diego, S. California
CAM Nicaragua
CAM Southern California
SAM Oaxaca, Mexico
CAM Jalisco, Mexico
CAM Oaxaca, Mexico
CAM Guerero, Mexico

C TR Source Region

generation area

5.

Map

The tsunami events for the Pacific side of the Central American region from 1900 to 2002
were plotted on the following map (Fig. 1).

Fig. 1 The tsunamis on the Pacific side of the Central Amercan region from 1900 to 2002.

Appendix 2: A Short Introduction to Seismology


As earthquakes are the predominant cause of tsunami generation it is necessary to have some
background knowledge of seismology. In this appendix a brief introduction is given to
seismology.
An earthquake can be defined as the sudden release of elastic energy stored in the rocky
upper mantle of the earths surface.
Most earthquakes are caused by movements of the tectonic plates. These plates form the
earth's crust The tectonic plates are permanently moving. Most of these movements we can't
feel; they are much too small to be noticed by the human senses. When two of these plates
stick together at one location, locally the plates stop moving, inducing a build up of elastic
energy.
This elastic energy is released as these two plates suddenly slip past each other. The elastic
energy, released in this fracture zone, goes into wave energy (the seismic waves responsible
for the shaking) and into heat (due to friction between the two sides of the fracture).
The location in the earths interior where the two plates slip past each other is called
hypocenter, the location directly above the hypocenter at the earths surface is called the
epicenter (Figure 1). The focal depth, which is the depth of the hypocenter, can vary from less
than 65km, for shallow earthquakes, to depths more than 300km, for deep earthquakes.

Figure 1 Terminology of the fault zone of an earthquake.

Where Earthquakes Occur

The Earth is formed of several layers that have very different physical and chemical
properties. The outer layer, which averages about 70 kilometers in thickness, consists of about
a dozen large, irregularly shaped plates (Figure 2), on top of the partly molten inner layer.
These plates slide over, under and past each other. Most earthquakes occur at the boundaries
where the plates meet.

Figure 2 Plate boundaries

There are three types of plate boundaries: spreading zones, transform faults, and subduction
zones (Figure 3).

Figure 3 Types of plate boundaries

At spreading zones, molten rock rises, pushing two plates apart and adding new material at
their edges. Most spreading zones are found in oceans; for example, the North American and
Eurasian plates are spreading apart along the mid-Atlantic ridge. Spreading zones usually
have earthquakes at shallow depths (within 30 kilometers of the surface).
2

Transform faults are found where plates slide past one another. An example of a transformfault plate boundary is the San Andreas fault, along the coast of California and northwestern
Mexico. Earthquakes at transform faults tend to occur at shallow depths and form fairly
straight linear patterns.
Subduction zones are found where one plate overrides, or subducts, another, pushing it
downward into the mantle where it melts. An example of a subduction-zone plate boundary is
found along Pacific coast of Central America (the Middle American Trench). Subduction
zones are characterized by deep-ocean trenches, shallow to deep earthquakes, and mountain
ranges containing active volcanoes. About 90% of all earthquakes occur in subduction zones
(Bolt, 1995).
Earthquakes can also occur within plates, although plate-boundary earthquakes are much
more common. Less than 5% of all earthquakes occur within plate interiors (Bolt, 1995). As
plates continue to move and plate boundaries change over geologic time, weakened boundary
regions become part of the interiors of the plates. These zones of weakness within the
continents can cause earthquakes in response to stresses that originate at the edges of the plate
or in the deeper crust. For Example the New Madrid earthquakes of 1811-1812 and the 1886
Charleston earthquake occurred within the North American plate.
As stated above earthquakes occur in fracture zones mostly located near plate boundaries.
These fraction zones between two blocks of rock are called faults. During an earthquake the
rock on one side of the fault suddenly slips with respect to the other.
Earthquake fault geometry basically falls apart in three fundamental end members: strike-slip,
thrust-dip and dip-slip faults (Figure 4). Strike-slip faults involve horizontal motion of the
earths crust, while thrust-dip and dip-slip faults induce vertical motion. Most faults show a
combination of horizontal and vertical motion, but usually one of the two is more pronounced.
The movement on a fault deep within the earth can break through to the surface, it is then
called surface rupture. This can lead to spectacular fault scarps in the landscape; for example
a sudden rise of a road or a sudden shift of a fence. However not all earthquakes result in
surface rupture. This is mainly determined by the size of the earthquake, the depth of the
hypocenter and the composition of the rocky layers above the hypocenter.

Figure 4 Basic fault geometries

Seismic waves

When two plate boundaries slip over each other, part of the released elastic energy goes into
wave energy.
These seismic waves spread out from the hypocenter through the earths interior and crust.
When they reach the surface, they cause the ground shaking that is felt during an earthquake.
It is this shaking that is responsible for the bigger part of the earthquake damage.
There are several different kinds of seismic waves, and they all move in different ways. The
two main types of waves are body waves and surface waves. Body waves can travel through
the earth's inner layers, but surface waves can only move along the surface of the planet like
ripples on water. Earthquakes radiate seismic energy as both body and surface waves.

2.1

Body Waves

P Waves
The first kind of body wave is the P wave or primary wave. This is the fastest kind of seismic
wave. The P wave can move through solid rock and fluids, like water or the liquid layers of
the earth. It pushes and pulls the rock, and moves through the rock, just like sound waves
push and pull the air (Figure 5). Sometimes animals can hear the P waves of an earthquake.
Usually we only feel the bump and rattle of these waves.

Figure 5 Primary wave

S Waves
The second type of body wave is the S wave or secondary wave, which is the second wave
you feel in an earthquake. An S wave is slower than a P wave and can only move through
solid rock. This wave moves rock up and down, or side-to-side (Fig. 2.5).
The difference in speed between P and S waves can be used to calculate the distance from the
registration point to the epicenter.

Figure 6 Secondary Wave

2.2 Surface Waves


Love Waves
The first kind of surface wave is called a Love wave, named after A.E.H. Love, a British
mathematician who worked out the mathematical model for this kind of wave in 1911. It's the
fastest surface wave and moves the ground from side-to-side (Figure 7).

Figure 7 Love wave

Rayleigh Waves
The other kind of surface wave is the Rayleigh wave, named for John William Strutt, Lord
Rayleigh, who mathematically predicted the existence of this kind of wave in 1885. A
Rayleigh wave rolls along the ground just like a wave rolls across a water surface (Figure 8).
Because it rolls, it moves the ground up and down, and side-to-side in the same direction that
the wave is moving. Most of the shaking felt from an earthquake is due to the Rayleigh wave,
which can be much larger than the other waves.

Figure 8 Raleigh wave

2.3

Registration of the Different Wave Types

An example of a registration of the bottom movements caused by the different seismic waves
is given on the seismographs in Figure 9.
The different seismic waves arrive in a fixed sequence, due to their differences in wave
celerity.
The first waves arriving at one location from the hypocenter are the longitudinal P waves, that
mainly arrive at the surface under a steep angle, producing mainly vertical ground
movements. After a short period the S waves arrive, who produce bottom movements in the
horizontal plane, because they arrive at the surface at a steep angle. Immediately after the S
waves the Love waves arrive, followed by a large train of Raleigh waves. These are the waves
which produce the well known heaving movement of the bottom during earthquakes. The
fading out at the end of an earthquake is a mixture of P-, S-, Love- and Raleigh waves who
propagated at various ways through the complex structure of the interior of the earth.
Figure 9 An example
of the bottom
movement during an
earthquake. Each
earthquake is
reported by three
seismograms, each
corresponding to a
different direction:
East-West (O), NorthSouth (N) and Updown (Z). The
different wave types
can be clearly
distinguished.

Earthquake Scales

The first scientific method introduced for describing the strength of an earthquake was the
magnitude, which is based on the resulting waves recorded on a seismogram. The concept is
that the recorded wave amplitude reflects the earthquake size once the amplitudes are
corrected for the decrease with distance due to geometric spreading and attenuation.
Nowadays seismologists use the moment magnitude as a measure for the size of an
earthquake. The moment magnitude is a scale which is closely connected with the actual
physics of the rupture process.
3.1

Local Magnitude ML

The earliest magnitude scale was introduced by Charles Richter in 1935 for describing
southern Californian earthquakes. It s known as the local magnitude ML, often referred to as
the Richter scale.

He proposed following formula:


M L = log A + 2.76log 2.48 (1)
ML : local magnitude (-)
A : amplitude of the signal (mm)
: distance between the source and the seismometer
(km)
The formula is calibrated for regional earthquakes in Southern California, with a nearly
constant (shallow) depth of the hypocenter. Which are registered with a specific seismograph
with a typical instrument period of 0.8s, known as the Wood-Anderson seismograph.
These factors are incorporated in the constants of the formula.
The nomogram in Figure 10 is a good illustration of how ML is solely dependent on the
maximum recorded amplitude and the distance from the hypocenter.
Richter magnitudes in their original form are no longer used, because earthquakes do not only
solely occur in California, focal depth of earthquakes is very variable and Wood-Anderson
seismographs have a limited registration range. However local (Richter) magnitudes are still
popular by the media. Also because many buildings have resonant frequencies near 1 Hz,
close to the sample frequency of a Wood-Anderson seismograph, ML is often a good
indication of the structural damage an earthquake can cause.

Figure 10 Nomogram for calculating ML by means of a seismogram

3.2

Surface Wave Magnitude Ms

Nowadays seismologists use more refined scales. One of them is the surface wave magnitude
Ms, which can be regarded as a modified Richter scale.
M s = log( A / T ) + 1.66log + 3.3

(2)
Ms : surface wave magnitude (-)
A : amplitude of the signal (mm)
T : dominant wave period (s)
: distance between the source and the seismometer
(km)

The surface wave magnitude is measured using the largest amplitude of the surface waves,
these often are the Raleigh waves.
The Ms -scale is roughly an extended Richter scale for earthquakes at great distances and
gives a good estimate of strong and medium earthquakes. However the Ms -scale cant be
used on deep earthquakes because these dont generate large surface waves. Another
modification for the magnitude scales is needed for this, this goes beyond the scope of this
study and one is referred to specialized literature for instance Bolt, 1995.
As measures of earthquake size, magnitude have two major advantages. First, they are
directly measured from seismograms without sophisticated signal processing. Second, they
yield units of order 1 which are intuitively attractive: magnitude 5 earthquakes are moderate,
magnitude 6 are strong, 7 are major, 8 are great.
However magnitudes have three related limitations. First, they are totally empirical and thus
have no direct connection to the physics of the earthquakes (the 1 and 2 arent even
dimensionally correct). A second difficulty is the variation of the magnitude with the azimuth,
due to the radiation patterns of the waves in the inner of the earth. This can be reduced by
averaging the results.
But most importantly the surface wave magnitude underestimates the true magnitude of low
frequency earthquakes and of very large earthquakes. The Ms -scale only uses the amplitude
of surface waves which have a period about 20 seconds, whereas some earthquakes have
much energy at lower frequencies, so called slow earthquakes. Thus Ms is an underestimation
of the real size of these slow earthquakes. Also as quakes get larger they sent out a greater
portion of their energy in the longer period waves, which tends to saturate the Ms scale. The
saturation begins at around 7.3 and surface wave magnitude never becomes greater than 8.0 to
8.3.
3.3

Moment Magnitude Mw

To cope with the above-mentioned shortcomings a scale was designed which is a true
physical measure for the energy of an earthquake. A magnitude scale was defined based on
the seismic moment Mo of an earthquake. The seismic moment describes the faulting process
itself and considers factors as the area of the fault, the length of the displacement and the
strength of the rock being ruptured.

The seismic moment of an earthquake can be calculated by modeling the earthquake event
using the Harvard fault plane solution, also known as CMT (Centroid-Moment Tensor). The
Harvard solution uses seismologic recordings at different stations around the world and solves
the classic inverse problem in geophysics. This means revealing the location and the rupture
characteristics of the source by means of recordings on distant seismographs. With this
information the seismic moment Mo of an earthquake can be calculated fairly accurate.

Figure 11 the fault plane of the earthquake.

The seismic moment can be calculated by:


M o = .u.L.W

(3)
Mo : seismic moment of an earthquake (Nm)
: rigidity around the fault (N/m)
u : average slip (m)
L : fault length (m)
W : fault width (m)

The seismic moment can also be calculated directly out of data gathered during surveys of the
fault region.
The moment magnitude scale is defined as follows:
M w = log M 0 /1.5 10.73

(4)
Mw : moment magnitud (-)
Mo : seismic moment (Nm)

Defining a magnitude related to the seismic moment Mo has some advantages. First of all it
gives a magnitude directly tied to earthquake source processes and it does not saturate.
Second it preserves the simplicity of the magnitude scale by giving values of order one,
moreover Mw is a convenient extension of Ms, when Ms saturates at values higher than about
7.3, this is clearly illustrated in Figure 12
The disadvantage is that the estimation of Mo requires more analysis of seismograms than for
M s.
Nowadays moment magnitude has become the common measure of the magnitude of large
earthquakes.

Figure 12 Relationship between the moment magnitude and the various magnitude scales.

10

Appendix 3:

Mail correspondence concerning the deriving of tsunami wave


characteristics out of tsunami data or seismic data
This appendix contains an overview of the reactions of some tsunami researchers on my questions concerning the deriving
of a wave from the available tsunami data and the available seismic parameters.

Original mail
"K.F.M. Cuypers" wrote:
> Dear Sir,
>
> I am a last year student coastal engineering at the Technical University of Delft (the Netherlands).
> As part of my master thesis I am doing some research on tsunamis.
>
> I have conducted a small study of literature, there is a lot of material available, but there is one question I can not seem to
get an answer on:
> There is a lot of data available about past tsunami events concerning the seismic parameters and tsunami intensity and/or
magnitude, but for coastal engineering purpose one is particularly interested in the wave parameters offshore (wave period
and wave height). Is it possible to derive these parameters out of the seismic or tsunami data or can they be obtained out of
databases?
>
> I would be very thankful if you could give me a hint about this.
>
> Sincerely Yours,
> Kim Cuypers

Replies
From:
To:
CC:
Date:
Subject:
Attachments:

Rabinovich, Alex
K.F.M.Cuypers@student.TUDelft.NL
CherniawskyJ@pac.dfo-mpo.gc.ca wang@pgc.nrcan.gc.ca kenji.satake@aist.go.jp
02/25/03 07:43 pm
Your question

Dear Kim,
Probably you could not find this information because your question actually is not easy at all!
Certainly there are some empiric relations between parameters of the source and offshore wave heights (Dr. Kenji Satake will
answer to this question better). There is also a simple approximate ratio between the source length (L) and offshore wave period
(T):
T ~ 2L/sqrt(gh),
where h is the mean ocean depth in the source area.
However, you should take into account that near the coast all these parameters will drastically be changed. Topographic resonant
features play the key role in formation wave heights and periods near the coast. You can find more detailed description of this
aspect of the problem in our papers (Rabinovich, JGR, 1997, 102, C6, 12,663-12,676 and Monserrat et al., GRL, 1998, 25, 12,
2197-2200) and in the list of references in these papers.
I can also send you a PDF file of our paper on estimation tsunami wave heights for the coast of Peru (~ 900 kByte) if you are
interested.
Sincerely,
Alexander Rabinovich
Alexander B. Rabinovich
Tsunami Center
Shirshov Institute of Oceanology
Russian Academy of Sciences
Moscow, Russia

From:
To:
CC:

Rabinovich, Alex
K.F.M.Cuypers@student.TUDelft.NL

02/26/03 12:01 am
Date:
Re: Your question
Subject:
Attachments:
Dear Kim,
So if you want to define a design wave, it is in a preliminary fase of costal engineering much safer to use a past tsunami event

then try to make rough estimations?


It is always useful to do both and compare results. There are very few sites where we have enough reliable tsunami statistics.
Imagine also very common situation. There is a tide gauge in a harbour with relatively long historical sea level records and many
recorded tsunamis. So, exactly for this site we can easily estimate possible tsunami wave heights for any return period. However,
the place of a new construction is located in, for example, 10 km aside. Will tsunami wave height be the same?? WE DO NOT
KNOW! We cannot answer to this question without additional investigation (numerical modeling).
In 1993 I participated in a field survey after Okushiri tsunami in the Sea of Japan. In a small valey (about 300 m) tsunami runup
was found to be 32.5 m (!), however in a few kilometers outside it was only 11 m.
I hope this will help you to understand the situation better.
Best regards,
Alexander
Alexander B. Rabinovich
Tsunami Center
Shirshov Institute of Oceanology
Russian Academy of Sciences
Moscow, Russia

From:
To:
CC:
Date:
Subject:
Attachments:

Eric L. Geist
K.F.M.Cuypers@student.TUDelft.NL
hjlee@usgs.gov
02/26/03 07:36 pm
Re: Fwd:

2000JB000139.pdf (536KB)

Dear Kim Cuypers,


It is possible to calculate the tsunami wavefield, if one knows
enough about the seismic parameters that generated the tsunami.
Depending on the level of accuracy needed, one may need to know the
spatial variation of slip and slip direction within the rupture zone
to calculate the tsunami. A good overview of these calculations is
presented by K. Satake (2002) Chapter 28 in International Handbook of
Earthquake and Engineering Seismology (Edited by William H. K. Lee,
Hiroo Kanamori, Paul C. Jennings, and Carl Kisslinger) that I believe
is published by AGU. More information is also available at
http://www.whklee.org/iaspei.html I've also attached a recent paper
that describes the relationship between seismic parameters and the
local tsunami wavefield.
Direct observations of tsunami waveforms are most often made at tide
gauge stations. For example, for the June 23, 2001 Peru tsunami,
data is available at http://www.pmel.noaa.gov/tsunami/peru_pmel.html.
Offshore data by bottom pressure recorders is limited. You can start
exploration of these data at http://www.pmel.noaa.gov/tsunami/Dart/
and http://www.jamstec.go.jp/scdc/top_e.html
I hope you find this information useful. Please let me know if you
have any further questions...Eric
Eric L. Geist Desk: (650)329-5457
U.S. Geological Survey Fax: (650)329-5411
345 Middlefield Rd., MS 999 Email: egeist@usgs.gov
Menlo Park, Ca 94025
Internet: http://walrus.wr.usgs.gov/staff/egeist.html

From:
To:
CC:

grilli
K.F.M.Cuypers@student.TUDelft.NL

02/24/03 10:45 pm
Date:
Re:
Subject:
Attachments: grilli.vcf (391 Bytes)
Kim:
Wave parameters at the continental shelf are usually computed using tsunami generation and propagation models. There is no
simple relationship relating seismic parameters to wave parameters, except through pure empiricism, but this will be quite limited
in scope and range.
Check www.tsunamicommunity.org, for instance, for more info on landslide tsunami models.
Stephan Grilli

Stephan Grilli
http://www.oce.uri.edu/~grilli
Distinguished Engineering Professor and Chairman
Department of Ocean Engineering
University of Rhode Island

From:
To:
CC:
Date:
Subject:
Attachments:

Lori Dengler
K.F.M.Cuypers@student.TUDelft.NL
02/24/03 11:23 pm
Re: your mail

Tsunami research is still in its early stages and most of the modeling
efforts have focussed on trying to match field data from recent tsunamis
to model results. There is an active US modeling effort in the 5 Pacific
States but no archive of all modeling results.
There are a number of tsunami modeling programs including USC: http://www.usc.edu/dept/tsunamis/
Oregon Geological Institute: http://www.ccalmr.ogi.edu/projects/oregonian/
University of Alaska- Fairbanks
http://www.aeic.alaska.edu/tsunami/index.html
NOAA's TIME Center: http://www.pmel.noaa.gov/tsunami/
I suggest contacting either Vasily Titov (titov@pmel.noaa.gov) or Frank
Gonzalez (gonzalez@pmel.noaa.gov) at the TIME center to see if they can
help find what you want.
-----------------------------------------------------------------------------Lori Dengler
Department of Geology
Humboldt State University
Arcata, CA 95521
voice (707) 826-3115 fax (707) 826-5241

Appendix 4: The 1992 Nicaraguan tsunami


The Nicaraguan event was probably the first event documented by extensive seismological and
tsunami data. This was also the first tsunami event for which international survey teams were
dispatched. Extensive data was gathered, giving opportunity for intensive research and modelling as
well of the earthquake as of the tsunami event (Satake, 1995). This makes the 1992 Nicaraguan
tsunami one of the best documented and investigated events, which makes this an ideal event for using
as a base for a design tsunami.
The earthquake causing the tsunami occurred at 7.16 P.M. (00:16 GMT) on 2 September 1992, 70
km offshore from Managua.
The earthquake was barely felt by coastal residents. Yet the movement of the seabed produced a giant
tsunami that reached the coastline 40-70 minutes later. Healthy people who were awake at that time,
were able to outrun the tsunami. The tsunami killed 170 people mostly children who were asleep and
infirm people who could not flee, 490 people were injured and over 13.000 became homeless. Run-up
averaged 4 m along 250 km of the Nicaraguan coastline and reached a maximum value of 10.7 m
MSL near El Transito (Fig. 15.1). Sand was eroded from the beaches and transported, together with
boulders and building rubble, tens of meters inland
The wave also propagated westwards into the Pacific Ocean. A 1 m wave was recorded at the tide
gauges on the Easter- and Galapagos Islands, and a wave a bit higher than 10 cm was measured at Hilo
Hawaii and at Kesennuma, Japan.

Figure 1 Area affected by the 1992 Nicaraguan tsunami. The run-up bars represent the maximum
reported run-up at one location. The run-up bars are scaled relative to each other.

Earthquake Data

The earthquake had a surface wave magnitude, Ms, of 7.2 and a seismic moment, Mo of 3x1020 Nm.
It had a shallow focal depth of 45 km. Aftershocks occurred along the strike, parallel to the coast of
Nicaragua in a band 100 km wide and 200 km long. The rupture occurred as the result of slow
thrusting along the shallow dipping, subduction interface between the Cocos and Caribean Plates.
The earthquakes moment magnitude, Mw, was 7.7, a value at least half an order greater than the
surface wave magnitude, Ms, 7.2.
The reason for this dissimilarity is the slow rupturing process along the fault. The occurrence of slip
on a plate interface filled with soft subducted sediments caused the rupture process to be slower than
in ordinary subduction-zone thrust earthquakes (Kanamori, 1993).

Tsunami Data

The 1992 Nicaragua earthquake was a tsunami earthquake, meaning it generated tsunamis
disproportionately large for its surface wave magnitude, Ms.
The significant disparity between Ms and Mw (7.2-7.7)is clearly a characteristic for tsunami
earthquakes.
The tsunami magnitude, Mt, was estimated at 7.9 +/- 0.4 (Ide, 1993). Because the Mt scale has not yet
been regressed against Mw for the data set in the eastern Pacific, Mt is estimated using the peak-to
trough amplitudes of the tidal measurements in Hawaii and Japan (Chapter 5: tsunami scales). The
tsunami intensity i was estimated at 2.8.
Run-up data along the Nicaraguan coast are presented in Figure 1 and Figure 2, run-up heights are
mostly between 3 and 8 m MSL, with a maximum of 10.7 m MSL at El Transito. It should be noted
that Figure 1 represents the maximum reported run-up heights at one location. During the field surveys
of Baptista (1993) and Abe (not available) at one location more than one cross section was measured
(Figure 2). As is clearly illustrated in Figure 2 at one location serious differences in run-up heights can
occur. It is believed that these differences are caused by on land topographic differences at one
location.

Figure 2 The run-up heights for the 1992 Tsunami reported by the field surveys of Abe and Baptista
(Satake, 1995). The run-up heights are corrected for the tidal elevation during the run-up, tidal elevation
was between 0.5m and 1m MSL. Notice that the points lying on one vertical represents run-up heightys
measured at one location.

The tidal recording at Corinto (Fig. 15.2) shows an initial waterlevel withdrawal of about 10 cm,
followed by a large water elevation showing a peak to through amplitude of 49 cm. The period of the
first rising wave is 9.2 minutes. The tidal recording at Puerto Sandino (Figure 2) shows an initial
waterlevel withdrawal of about 10 cm, followed by a large water elevation causing the tide recorder to
go off scale, the peak to through amplitude is at least 117 cm.
The difference between these two measurements is probably caused by the location of the tide gauges
(waterdepth 10 m at Corinto and 4 m at Puerto Sandino) and the local bathymetry.
Other tidal recordings made: at Baltra (1.12 m) and Libertad (0.26 m) in Ecuador, Socorro Island (0.16
m) and Cabo San Lucas (0.13 m) in Mexico and Valparraiso (0.09 m) in Chile, Kesennuma (0.13 m)
in Japan and Hilo (0.12 m) at Hawaii.

Figure 3 Tidal recordings of the 1992 nicaraguan tsunami at Corinto and Puerto Sandino.

Source Characteristics for the 1992 Nicaraguan Tsunami

This part of the text is based on the work of Imamura (1993) and Satake (1994 and 1995).
Imamura made estimates about the characteristics of the tsunami source based on seismic data (size
and area of the dislocation). A numerical 2D simulation was made with the estimated tsunami source,
the obtained run-up data were compared with the measured run-up data along the Nicaraguan coast.
The measured run-up was 5.6 to 10 times higher than the run-up obtained with the model.
In order to reconcile the discrepancy between seismic data and tsunami data, Satake used various fault
models. He varied the fault length and width, keeping the seismic moment constant by adapting the
average slip and the rigidity.
Comparison of calculated waveforms with tide gauge records of Corinto and Puerto Sandino and the
run-up heights, show best correspondence with a fault width of 40 km and a fault length of 250 km. In
order to correspond with the seismic moment, this tsunami fault model has to have a greater slip than
the seismic fault model as derived by Imamura resulting in bigger bottom deformations as is illustrated
in Figure 3.

The seismic moment of an earthquake is given by :


M o = .u.L.W (Nm)

where is the rigidity around the fault, u is the average slip, L is the fault length and W is the fault width.

Figure 4 shows the original seismological fault area as estimated by Imamura and the fault area
estimated with the tsunami data by Satake.
Figure 5 shows the vertical displacement according to the seismic fault model and Figure 6 shows the
vertical displacement according to the tsunami fault model.

Figure 4 Schematic cross section of the vertical bottom deformation as by


the seismic fault model and by the tsunami fault model. Both
deformations are the result of a rupture with the same seismic moment
Mo.

Figure 5 map of the 1992 Nicaraguan earthquake


source region. Solid circles show the aftershocks
within 1 day of the main shock. The open frame
shows the seismological fault model, whereas the
shaded frame shows the tsunami fault model.

Figure 6 The initial bottom deformation overlaid on


the bathymetry, as estimated from the seismic fault
model by Imamura (1993). The solid lines are for
uplift and the dashed lines are for subsidence. The
contour interval is 2 cm. The star symbol marks the
epicenter of the main shock.

Figure 7 The initial bottom deformation overlaid


on the bathymetry, as estimated from the tsunami
fault model by Satake (1995). The solid lines are
for uplift and the dashed lines are for subsidence.
The contour interval is 20 cm.

Appendix 5: Bathymetry
1. Pie del Gigante
The bathymetry is derived out of nautical charts for the Nicaraguan region:
- for the shallow areas (depth 0m to 66m) a chart with a scale of 1/50 000 (DMA, 1981)
- for the deeper parts (depth >-66m) a chart with a scale of 1/300 000 (INETER, 1989)
The depth profile as derived out of the nautical charts is presented in Table 1 and is plotted in
Figure 1 and Figure 2.

0
-1000
-2000
-3000
-4000
-5000
-6000

depth (m)

20000

40000

60000

80000

100000

120000

140000

continental shelf

ocean trench

Figure 1 The bathymetry at Pie Del Gigante, clearly the ocean trench
and the continental shelf can be distinguished.

0
-50
-100
-150
-200
-250
-300
continental shelf

depth (m)

10000

20000

30000

distance (m)

40000

Bathymetry for Pie Del Gigante (shallow)


50000

-66
-77
-84
-91
-98
-102
-109
-113
-120
-128
-140
-155
-190
-950
-3950
-4860
-4616
-2926
-2926
-2981
-3072

distance (m)

60000

10500
13500
18000
21000
24000
28500
33000
36000
39000
42000
51000
64500
75000
96000
132000
138000
150000
180000
195000
204000
213000

Bathymetry for Pie Del Gigante


160000

diepte
(m)
0
-10
-20
-40
-66

70000

afstand
(m)
0
500
1000
2750
9000

foreshore

Figure 2 The more shallower parts of the bathymetry at Pie Del


Giante, the Continental Shelf and the foreshore can be
distinguished.

Table 1 Bathymetry in front of Pie del Gigante.

For input in the numerical model the bathymetry points were linearly interpolated.

2. Bahia del Salinas


The bathymetry is derived out of nautical charts for the Nicaraguan region:
- for the shallow areas (depth 0m to 66m) a chart with a scale of 1/50 000 (DMA, 1981)
- for the deeper parts (depth >-66m) a chart with a scale of 1/300 000 (INETER, 1989)
The depth profile as derived out of the nautical charts is presented in Table 1 and is plotted in
Figure 1 and Figure 2.

0
-1000
-2000
-3000
-4000
-5000

depth M SL (m)

20000

40000

60000

80000

100000

120000

140000

d is tan c e (m )

160000

B ath ym etrie B ah ia D el S alin as

Figure 3 Bathymetry at Bahia del Salinas.

0
-100
-200
-300
-400
-500
-600

depth M SL (m)

10000

20000

30000

40000

50000

d is ta n ce (m )

60000

B ath ym etrie B ah ia D el S alin as


70000

diepte
(m)
0
-10
-20
-29
-47
-55
-82
-87
-90
-91
-98
-109
-137
-150
-155
-210
-228
-3370
-4440
-3584
-2925
-3072
-3050
-3109
-3107

180000

afstand
(m)
0
500
1125
2625
7500
9000
22500
24000
27000
28500
31500
39000
45000
48000
51000
70500
73500
136500
148500
166500
184500
196500
205500
217500
232500

Figure 4 The shallower parts of the Bathymetry at Bahia del Salinas

Table 2 Bathymetry in front of Pie del Gigante.

For input in the numerical model the bathymetry points were linearly interpolated.

3.

Corinto

The depth profile in front of Corinto is derived using the sea chart of 1/300.000 (DMA, 1981)
also used for the depth profile of Pie del Gigante. Only the profile between 12.8 m and
283m could be derived from this chart. Nearshore no more detailed soundings are given and
the deeper part of the depth profile lies beyond the borders of the map.
Therefore for the deeper part the same depth profile is used as for Pie del Gigante.As for the
more shallow parts nearshore, the point are lineary interpolated between 12.8m and 0m).

distance (m)

41

81

121

161

201

241

281

Bathymetry for Corinto


321

-2000
-3000
-4000

depth (m)

-1000

-5000
-6000

Figure 5 Bathymetry for Corinto.

10000

20000

distance (m)

30000

Bathymetry for Corinto (shallow)


40000

0
-50
-100
-150

depth (m)

diepte
(m)
0
-12,8
-27,5
-35
-44
-58
-66
-69
-73
-84
-91
-102
-106
-100
-117
-120
-208
-283
-950
-3950
-4860
-4616
-2926
-2926
-2981
-3072

50000

afstand
(m)
0
1500
4500
9000
12000
16500
19500
22500
25500
30000
33000
39000
42000
45000
51000
54000
57000
72000
96000
132000
138000
150000
180000
195000
204000
213000

-200
-250

Figure 6 The Shallower parts of the bathymetry for Corinto

Table 3 Bathymetry in front of Corinto,

Appendix 6: Preliminary design of the Caisson


Breakwater at Pie del Gigante
1

Data

The input data is the same as used by HAECON for the preliminary design of the Rubble
Mound breakwater at Pie del Gigante.
The design lifetime of the structure is 50 years.
1.1

Wave Data

The design storm has a return period of 500 years.


Insufficient data is available for the determination of the mean wave period, but it is known
that only 6% of the waves have longer periods than 10s. As a first approximation for the
storm wave a period of 10 s is chosen.
Hs = 6.3 m
Tm = 10 s
= 0 (angle of incidence)
1.2

Site

depth
water level
tidal difference
wind set-up
tan

= -16.0 m C.D.
= +3.0 m MWL
= 1.6 m
= negligible
= 1/100 (slope of the coast)

Due to the steepness of the coast and the lack of shallow areas in front of the breakwater wind
set-up was believed to be negligible. This was confirmed by wind set-up calculations
performed with CRESS.
1.3

Constants
g
w
c
s

= 9.81 m/s
= 1030 kg/m (density of sea water)
= 2400 kg/m (density of reinforced concrete)
= 1800 kg/m (density of ballast sand)

Design

Design calculations were carried out according to Godas principles (Goda, 1992). Godas
design principles are based on an assumed pressure distribution on the caisson face. He then
uses this pressure distribution to calculate the stability against sliding and overturning of the
caisson.
Calculations were made with the proposed formulas embedded in a spreadsheet and compared
with calculations of the Goda formulas carried out in CRESS.
2.1

Design wave at the breakwater location

The design wave was transformed to the breakwater location using:


- the Airy wave theory for shoaling
- Godas principles for determining the design wave as is required in Godas formulas
Hmax
Hs
Tm
L

2.2

= 10.4 m
= 5.8 m
= 10 s
= 121.8 m

Sill

2.2.1
Sea Side
The sill height and length have to be limited in order to prevent waves from breaking on the
sill and thus generating a large dynamic load on the caisson.
Sill height
For an upper limit of the sill height, Miche breaking limit is used:
H/h > 0.78

with:

H = Hmax =10.4 m
depth = depth at lowest low tide = 18.2 m

Height of the sill < 18.2 h


< 4.85 m
The height of the sill is chosen as 4.0 m.
Sill length & slope
As a minimum length for the sill in front of the caisson Goda gives a value of 10 m for heavy
wave attack (Goda, 1985).
A slope of 1:2 is purposed.
Stability
At the toe of the caisson heavy concrete blocks have to be placed. The size of these blocks
ranges from 10 to 40 tons, the required size and extension for this breakwater has to be

determined by model tests. As a first approximation blocks of 20 ton over a length of 5 m are
chosen.
The remaining top of the sill should be covered with an armor layer.
The Dn50 of the armor layer is determined by the formula of Madrigal en Valds:
Hs
h
= (5,8 b 0, 6) N od0,19
hs
Dn50
Hs is the significant wave height. hb/hs is the ratio of the berm depth, hb, and the bottom depth,
hs. Nod is given a value of 2 (acceptable damage).
This leads to a required Dn50 = 0.5 m.
Using rubble of 300-1000 kg in a layer of 1 m, satisfies this condition.
The core of the sill should consist of rubble 80-200 mm, this complies with the geometric
closed filter rules.
2.2.2
Harbor Side
No concrete blocks or armor layer are provided. The core material (80-200 mm) is extended 5
m from the toe of the caisson. The sill is given a slope of 1:2.
2.3
2.3.1

Caisson
Stability against wave action

By trial and error a caisson with the following dimensions was obtained:
H = 21 m
W = 13.5 m
L = 60 m

thickness of the
outer wall = 0.5 m
inner wall = 0.2 m
floor
= 2.5 m
roof
= 3.0 m

This leads to:


the following pressure distribution:
p1 =

81,41 kN/m

p2 =
p3 =

51,63 kN/m
56,14 kN/m

p4 =

59,50 kN/m

pu =

53,84 kN/m

the safety factors then become


sliding:
safety factor
= 1.22 1.2
overturning: safety factor
= 1.20 1.2

the volume of overtopping water


overtopping:
q
= 540 L/m/s
An overtopping of q = 540L/m/s leads to a transmitted wave height Ht = 1 m (dAngremond,
2001). For container terminal operations the limiting wave height perpendicular to a ship is H
= 0.8m (Ligteringen, 2000). However it is accepted that during a short period of time the
terminal operations have to be interrupted, so the high overtopping during design conditions is
expected.
This is just a first approximation, in further designs the port layout and its effect on the wave
penetration and transmission, as well as the actual ship movements hereby caused, should be
taken into account. Also economical considerations should be taken into account.
Because of the sensitivity of caisson breakwater for long period waves a sensitivity analysis is
done. For waves with a longer period than 10 s, the permissible significant wave height on
deep water is determined (Table 1).
Tm
Hs
SF sliding
SF overturning

10 s
6.3 m
1.22
1.20

11 s
5.9 m
1.20
1.20

13 s
5.2 m
1.20
1.21

15 s
4.5 m
1.20
1.23

Table 1 Sensitivity analysis with increasing period Tm

The maximum permissible wave height does decrease, however not dramatically. It is known
that only a small percentage of the waves has a period higher than 10 s. Of which the bigger
part are probably swell waves, with probably a lower Hs than the ordinary wind waves. So
longer waves with smaller heights should not necessarily impose a problem.
When more accurate data on the wave climate is available, this should be investigated
thoroughly, as the wave period, Tm, is regarded as one of the most governing factor in caisson
breakwater design.
2.3.2

Floating Stability

The floating stability of the caisson is important because the caisson is transported floating
from the construction site to the breakwater location.
For insuring the floating stability the metacentre, M, has to lie above the centre of gravity, G,
with a factor of 1.2 (dAngremond, 2001).
This condition is fulfilled:

HM/HG = 6.9/1.4 > 1.2

Fig. 1 Floating stability of the caisson. M: metacentre, G: centre of gravity, B: centre of buoyancy

Note that this is the caisson without a roof, the roof is applied after the caisson is sunk into
place and ballasted.
2.4

Bottom Protection

A bottom protection should be provided. A geotextile under the sill and extended 10m at the
sea side and 5 m at the harbor side is proposed.
2.5

Geotechnical Stability

At this moment the geotechnical data for the site of interest are not yet known. As a first
estimation it is expected that at the Pacific side of Nicaragua the subsoil is not the determining
factor of design. However this should be confirmed by geotechnical investigation.
2.6

Breakwater Dimensions

Fig. 2 Breakwater dimensions

Schematic top view of the caisson:

-total length of the caisson = 60 m


-three inner walls in the length of the caisson
-three inner walls in the width of the caisson
-thickness of each inner wall = 0.20 m

Fig. 3 Schematic top view of the caisson

Appendix 7:

Comparing the Tanimoto design formulas for


unbroken tsunami waves with the Goda design formulas for
unbroken short period waves

In this appendix the calculations leading to table 7.1 will be presented. Table 7.1 is representd
in Table 1
*=15.6m
P1
P2
P3

pu

L=121m
81 kN/m
52 kN/m
56 kN/m
54 kN/m

L=10 000 m
119 kN/m
119 kN/m
119 kN/m
116 kN/m

Tsunami
116 kN/m
116 kN/m
116 kN/m
116 kN/m

Table 1 Pressure calculations with the Goda formulas (L=121 m and L=10 000 m) and with the
Tanimoto formulas (Tsunamis) for the same crest elevation (*=15.6m) (table 7.1).

The second column L=121m are the calculated pressures on the caisson for the design short
periood waves. The pressures were calculated with the Goda design principles (7.1) for a
wave with a crest elevation of *=15.6m and a wave length L=121m. The calculations are
described in appendix 6. The third column are the pressures on the caisson for a wave with the
same crest elevation of *=15.6m but with a very long wave length, L=10 000m . The
calculations were carried out with the Goda design principles. The fourth column are the
pressures on the caisson as calculated with the formulas of Tanimoto for unbroken tsunami
waves (7.1).
The calculations of the third and the second column are presented below:

1. L=10 000m
calculations with Godas design principles (the formulas are presented in 7.1)
input
Hmax = 10.4m
L
= 10 000m

= 0 (angle of incidence of the wave crest, for = 0 the maximum forces occur on
the caisson)
tan = 1/100 (slope of the coast in front of the breakwater)
water = 1030 kg/m
The dimensions of the caisson breakwater are the same a in appendix 6 Figure 2.
calculations

* = 0.75 (1 + cos ) H max


= 0.75 (1 + cos 0 )10.4 = 15.6m

1 = 0.6 + 0.5 ( 2kh / sinh 2kh )

= 0.6 + 0.5 ( 2*6.28*104 *19.8 / sinh 26.28*104 *19.8 ) = 1.10


2

2 = min
= min

{(( h d ) / 3h ) ( H
b

/ d ) , 2d / H max
2

max

{(( 20.25 15.8) / 3*20.25) (10.4 /15.8) , 2*15.8 /10.8} = 0.0317


2

3 = 1 ( h '/ h )(1 1/ cosh kh ) = 1 (16.8 /19.8 ) (1 1/ cosh(6.28*104 *19.8) ) = 1.00


the resulting wave pressures:
p1 = 0.5 (1 + cos ) (1 + 2 cos2 ) w0 H max
= 0.5 (1 + cos 0 ) (1.10 + 0.0317 cos2 0 ) 10.10*10.4 = 118.92kN / m
p2 = 118.92 / cosh(6.28*10 4 *19.8) = 118.91kN / m
p3 = 3 p1 = 1.00*118.92 = 118.92kN / m
pu = 0.5 (1 + cos )1 3 w0 H max

= 0.5 (1 + cos 0 )1.10*1*10.10*10.4 = 115.75

2. Tsunami
calculations with Tanimotos design formulas (the formulas are presented in 7.1)
input
Hrefl = 10.4 m
The dimensions of the caisson breakwater are the same a in appendix 6 Figure 2.
calculations
* = 1.5H = 1.5*10.4 = 15.6m

the resulting wave pressures :


p = p1 = p2 = p3 = pu
= 1.1w0 H = 1.1*10.10* = 115.93kN / m

Appendix 8: Calculations of the caisson stability under


the design tsunami wave
The calculations are carried out with the Tanimoto design formulas for unbroken tsunami
waves (7.1). The caisson as designed for short period waves is shown in Figure 8.1. The
design wave is shown in Figure 8.2. The resulting pressure distribution as assumed by
Tanimoto is shown in Figure 1.

Figure 1 Wave pressure distribution due to a non-breaking tsunami wave as assumed by


Tanimoto.

Input
H
HL
h
Wcaiss
B

water

= Hposrefl = 2.7m (the positive reflecting amplitude of the tsunami)


= 0.4m
= 16.8m
= 5618kN
= 13.5m
= 0.6 (friction factor for the caisson on the rubble)
= 1030kg/m

Calculations
* = 1.5H = 1.5* 2.7 = 4.05m

the calculated wave pressures :


p = pu = 1.1w0 H = 1.1*10.10*2.7 = 30.00kN / m
ps = w0 hL = 10.10*0.4 = 4.04kN / m

the resulting wave forces and moments:


with hc* = min {*, hc } = min {4.05, 4.2} = 4.05m
total horizontal wave force

h* h*
P = 1 + 1 c c ph '
3H h '
4.05 4.05
= 1 + 1
30.00*16.8 = 564.93kN

32.7 16.8
overturning moment around the heel of the caisson

2
1
hc* hc* 1 4 hc* hc*
2
M p = + 1
+
ph '
2 3H h ' 2 9 H h '
2
4.05 4.05 1 4 4.05 4.05
1
2
= + 1
+

30.00*16.8 = 5173.85kNm
2 3*2.7 2.7 2 9 2.7 16.8

uplift force
1
1
U = pu B = 30.00*13.5 = 202.57 kN
2
2
overturning moment around the heel of the caisson
1
1
M u = pu B 2 = 30.00*13.52 = 1823.09kNm
3
3
horizontal wave force due to the water lowering

0.4

Ps = L + ( h ' hL ) ps =
+ (16.8 0.4 ) 4.04 = 67.09kN
2

relative caisson weight

1
( 2h ' hL ) Bcaiss
2
1
= 5618 10.10 ( 2*16.8 0.4 )13.5 = 3353.73kN
2

W0 = Wcaiss w0

Safety factors against sliding and overturning:

(W0 U ) 0.6(3353.73 202.57)


=
= 2.99 1.2
P Ps
564.93 67.09
W0 xW M u 3353.73*8.3 5173.87
overturning: S .F . =
=
= 3.62 1.2
M P + Ps xP
1823.09 + 67.09*8.3
sliding:

S .F . =

Appendix 9: Calculations concerning chapter 10:


Theoretical stability concept for the rubble mound
breakwater

h (m)
0,3
0,8
1,4
1,7
2
2,3
2,5
2,8
2,9

u(m/s)
1,54
2,52
3,34
3,68
3,99
4,28
4,46
4,72
4,80

1
1,2

2,82
3,09

1/ 2
1/ 2,5

1,5

3,45

1/ 3

18 ton

9 ton

1. Example calculations for table 10.1

Izbash
5,05
49,54
181,44
253,69
347,41
465,01
541,43
661,17
667,79

W (kg)
Dutch
3,52
34,72
127,39
178,12
243,94
326,56
379,80
464,35
468,79

CERC
4,35
42,69
156,35
218,60
299,36
400,69
466,55
569,72
575,43

Izbash
0,12
0,27
0,41
0,46
0,51
0,56
0,59
0,63
0,64

dn (m)
Dutch
0,11
0,24
0,37
0,41
0,45
0,50
0,53
0,56
0,56

CERC
0,12
0,25
0,39
0,44
0,49
0,54
0,56
0,60
0,60

0,46
0,38

187,03
167,19

131,19
117,34

161,16
144,07

0,42
0,40

0,37
0,36

0,40
0,38

0,32

223,17

156,65

192,30

0,44

0,39

0,42

slope angle(rad)
1/ 2
0,46
1/ 2,5
0,38
1/ 3
0,32
1/ 3,5
0,28
1/ 4
0,24
1/ 4,5
0,22
1/ 5
0,20
1/ 6
0,17
1/ 7
0,14

The marked line out of the table above will be calculated as an example calculation.
Input
h
slope

stone
0

m
D
K

= 2.9m
= 1/7
= 45
(angle of repose)
= 2600kg/m
= 1030kg/m
= 0.04 (shields parameters)
=0.9
(discharge coefficient)
= 1% (damage ratio)
=1
(calibration factor for the Dutch formula)

Calculations
Velocity with the formula of the free discharge weir:
u = m gh = 0.9 9.81*2.9 = 4.80m / s
Izbash :
K ( // ) =

sin( ) sin(45 8.13)


=
= 0.85
sin
sin 45
2

uc
1
1
4.80
dn50 =
= 0.64m

=
2g K ( // )1.2 2*1.52*9.81 0.85*1.2
Wn50 = r d n350 = 2600*0.643 = 667.79kg

Dutch formula:
K ( // ) =

sin( ) sin(45 8.13)


=
= 0.85
sin
sin 45

after some iteration it is found that Dn50=0.56m

C = 18 log

6h
6*2.9
= 18log
= 28.08m1/ 2 / s
dn
0.56

K2
12
=
= 0.032 s 2 / m
C 2 0.04* 28.082
Au02
0.032* 4.802
dn =
=
= 0.56m
K 2 ( // ) 1.52*0.852
A=

CERC formula:
D = 1% y =1.1 (stability factor out of Figure 10.2)
rU d6
W=
3
3
48 y 6 g 3 ( r / 0 1) ( cos sin )
=

3.14* 2600* 4.80 6


48*1.169.813 ( 2600 /1030 1) ( cos8.13 sin 8.13)
3

1/ 3

W
dn50 = 50
2600

= 575.43kg

1/ 3

575.43
=

2600

= 0.60 m

2. Example calculations for the table on p10-10


The following row is calculated:

Input
vB(0)
tan
f

Warmor
(ton)

slope

overtopping
height

crown 5m
u (m/s)

crown 6m
u (m/s)

Izbash uc
(m/s)

1/7

2,9

6,89

6,98

6,85

= 4.80m
= 1/7
= 0.1
= 45

Calculations
Velocity calculations with the formulas of Schttrumpf (2001)
at a height 5m of the crest
after iteration it is found that
vB (0)
vB2 (0)
2 sB
t
+
+
2
2
g sin
g sin g sin
4.80
4.802
2*35.36

+
+
= 4.47 s
2
2
9.81sin 8.3
9.81 sin 8.3 9.81sin 8.3
k1 =

2 fg sin
2*0.1*9.81sin 8.3
=
= 0.59s 1
hB
2.02

k1hB
0.59* 2.02
kt
0.59*4.47
.tanh 1 4.80 +
. tanh

f
0.1
2
2 =

= 6.89m / s
vB =
0.1*
4.80
0.59*
4.47
f .vB (0)
k
t

1+
.tanh 1
1+
.tanh

2.02*0.59
2
hB k1
2

vB (0) +

at a height 6m of the crest


the calculations are carried out in the same way
The critical velocity calculated with the Izbash formula
K ( // ) =

sin( ) sin(45 8.13)


=
= 0.85
sin
sin 45

W
dn50 = 50
2600

1/ 3

9000
=

2600

1/ 3

= 1.51m

uc = K ( // )1.2 2gdn50 = 0.85*1.2 2*1.52*9.81*1.51 = 6.85m / s

3. Example calculations for table 10.2


The following row is calculated:

18 ton

h - 5 ( m)

slo p e

an g le (ra d ) Iz b a sh

W ( kg )
D u tc h

CE R C

Iz b ash

Dn (m )
Dut ch

Dn
C E RC

0,3

3 , 42

0,14

1/ 2

0,46

6 66 8,83

5 13,98

1,3 7

0,5 8

0,8
1,4

5 , 14
6 , 26

0,39
0,75

1/ 2 ,5
1/ 3

0,38
0,32

205 69 ,7 2
313 50 ,8 4

3 06 4,39
6 83 4,20

1,9 9
2,2 9

1,0 6
1,3 8

1,7
2

6 , 53
6 , 76

0,96
1,18

1/ 3 ,5
1/ 4

0,28
0,24

246 30 ,3 1
214 42 ,8 5

6 87 5,66
7 11 7,40

2,1 2
2,0 2

1,3 8
1,4 0

2,3
2,5

6 , 96
7 , 03

1,41
1,59

1/ 4 ,5
1/ 5

0,22
0,20

197 83 ,6 8
172 69 ,3 2

7 46 1,42
7 18 3,62

1,9 7
1,8 8

1,4 2
1,4 0

2,8
2,9

7 , 05
6 , 89

1,88
2,02

1/ 6
1/ 7

0,17
0,14

132 70 ,9 8
9 55 6,55

6 35 1,24
5 03 0,93

1,7 2
1,5 4

1,3 5
1,2 5

1
1,2

6 , 33
6 , 12

0,47
0,6

1/ 2
1/ 2 ,5

0,46
0,38

239 80 ,6 2
101 32 ,6 9

20 663 ,7 4
8 73 1,19

2,1 0
1,5 7

2,0 0
1,5 0

1,5

6 , 44

0,8

1/ 3

0,32

9 40 1,74

8 10 1,34

1,5 3

1,4 6

(tests)

1,91

Input

stone
0
D

u(m /s)

1,51

9 ton

h (m )

= 45
(angle of repose)
= 2600kg/m
= 1030kg/m
= 1% (damage ratio)

Calculations
The velocity calculations are done in the preceding paragraph
Izbash :
K ( // ) =

sin( ) sin(45 8.13)


=
= 0.85
sin
sin 45
2

uc
1
1
6.89
dn50 =
= 1.54m

=
2g K ( // )1.2 2*1.52*9.81 0.85*1.2
Wn50 = r d n350 = 2600*1.543 = 9556.55kg
CERC formula:
D = 1% y =1.1 (stability factor out of Figure 10.2)
rU d6
W=
3
3
48 y 6 g 3 ( r / 0 1) ( cos sin )
=

3.14* 2600*6.896
48*1.169.813 ( 2600 /1030 1) ( cos8.13 sin 8.13)
3

1/ 3

W
dn50 = 50
2600

= 5030.93kg

1/ 3

5030.93
=

2600

= 1.25m

Appendix 10: Iteration problems of the Dutch formula


For high water velocities in thin waterlayers on steep slopes, the Dutch formula (10.7) does
not lead to solutions, because iteration is no longer possible. This is illustrated in the graphs
below. For the two situations presented in Table 1. Calcualtions were carried out with the
Dutch formula for different flow depths h. A large variation was made in the input nominal
Dn_start. The resulting difference of the calculated nominal diameter and the input nominal
diameter (Dn-Dn_start) was plotted against the calculated diameter Dn.
slope

waterdepth

crown 5m
u (m/s)

1/2
1/7

0,14
1.9

3,42
6,89

1.
2.
Table 1

1. Calculations for row 1

8000
6000
4000
2000
469

4339

6,17

7,24

153

69,4

36,6

20,9

12,4

7,4

4,26

2,15

0
0,42

Dn-Dn_start (m )

iteration Dutch formula

Dn (m )

Figure 1 h=0.14m

5,23

4,41

3,68

3,02

2,44

1,91

1,41

0,93

6
5
4
3
2
1
0
0,28

Dn-Dn_start (m )

iteration Dutch formula

Dn (m )

Figure 2 h=1m

0,4
0,38
0,36
0,34
2,59

2,41

2,23

2,05

1,88

1,71

1,55

1,39

1,22

1,06

0,32
0,9

Dn-Dn_start (m )

iteration Dutch formula

Dn (m )

Figure 3 h=3m
iteration Dutch formula

0,2
0,1
2,01

1,88

1,75

1,63

1,51

1,39

1,26

1,14

-0,2

1,02

-0,1

0,9

0
0,77

Dn-Dn_start (m)

0,3

-0,3
Dn (m )

Figure 4 h=4m

For very small waterdepths, h, the formula can not be iterated (no zero down crossing
occurs).For a waterdepth of h=3m, slowly the graph tends towards zero and for a waterdepth
of h=4 the formula can be iterated, a zero down crossing occurs for Dn=1.44m.

2. Calculations for row 2


Calculations for the second row show the same iteration problem as can be seen in Figure 5
and Figure 6.
Only for large waterdepths does the formula leads iteration of the Dutch formula is possible.

1
0,8
0,6

3,02

2,8

2,59

2,38

2,18

1,98

1,79

1,6

1,41

1,22

0,4
0,2
0
1,03

Dn-Dn_start (m )

iteration Dutch formula

Dn (m )

Figure 5 h=1.9m

iteration Dutch formula

0,15
0,1
0,05
3,18

3,01

2,85

2,7

2,54

2,39

2,25

2,1

1,96

-0,05

1,82

0
1,68

Dn-Dn_start (m )

0,2

Dn (m )

Figure 6 h=4m

Situations like these with high flow velocities in relatively thin water layers on a steep slope,
lead to such large required stone diameters that the situation exceeds the limits of the validity
of the Dutch formula.

Appendix 11: Rubble Mound breakwater for Pie del Gigante


The cross section of the proposed rubble mound breakwater as proposed by HAECON.

Figure 1 Cross section of the rubble


mound breakwater for Pie del Gigante

In Figure 1 the breakwater profile as presented in the feasibility study of the Nicaragua dry
canal. For ordinary rubble the following stone weights were put forward W50=60ton and
W50/2=30ton.
Schematised
For numerical calculations of the tsunami wave hitting the breakwater, the cross-section of the
breakwater will be schematised as shown in Figure 2 The berm at the seaside slope is
removed, because of its small length (8.70m) it would require equally small distance steps to
represent this berm in the model. Also because of its small length compared to the tsunami
wave length (8.7m a few kilometres) it is believed that the effect on the behaviour of the
tsunami wave is negligible.

Figure 2 Schematised breakwater profile

BIBLIOGRAPHY
Angremond, K.d (2001), Breakwaters and Closure Dams, Delft University Publishers,
Delft.
Abe, K. (1983), A New Scale of Tsunami Magnitude, Mt, in Tsunamis-Their Science and
Engineering, Terra Scientific Publishing Company, Tokyo.
Baptista, A.M., Priest, G.R., Murty, T.S. (1993), Field Survey of the 1992 Nicaragua
Tsunami, in Marine Geodesy, Taylor & Francis, Bristol, vol.16, pp.169-203.
Barends, F.B.J., Hannoura, A.A.A. (1981), Non-Darcy Flow, A State of the Art, in Flow
and transport in porous media; Euromech 143, Delft, spet.1981, Balkema, Rotterdam,
pp.37-51.
Bolt, B.A. (1995), Aardbevingen, Segment, Beek.
Bryant, E. (2001), TSUNAMI: The Underrated Hazard, Cambridge University Press,
Cambridge.
Camfield, E. F. (1980), Tsunami Engineering, U.S. Army Coastal Engineering Research
Center, Fort Belvair.
Camfield, F. E. (1993), Dynamic response of structures on tsunami attack, in Tsunamis in
the World, Kluwer Academic Publishers, Dordrecht, pp133-138.
Dudley, C.D. (1998), Tsunami!, University of Hawaii Press, Honolulu.
Fernandez, M, Molina, E., Havskov, J., Atakan, K. (2000), Tsunamis and Tsunami
Hazards in Central America, in Natural Hazards, Kluwer Academic Publishers,
Dordrecht, pp91-116.
Goda, Y (1985), Random Seas and Design of Maritime Structures, World Scientific,
Singapore.
Goda, 1992, The Design of Upright Breakwaters, in Proceedings of the Short Course on
Design and Reliability of Coastal Structures, Instituto di Idraulica Universita di
Bologna, Bologna.
Hatori, T. (1995), Magnitude scale for the Central American tsunamis, in Tsunamis: 19921994; their generation, dynamics, and hazard, Birkhuser, Basel., pp471-479.
Hitachi, S., Kawada, M., Tsurya, H.,(1994) Experimental studies on tsunami flow and
armorblock stability for the design of a tsunami protection breakwater in
KamaischiBay, in Hydro-Port94 Vol.1, Coastal development institute of
technology, pp765-783.
Ida, K. (1981), Some Remarks on the Occurrence of Tsunamigenic Earthquakes around
the Pacific, in Tsunamis their Science and Engineering, Ida, K., Iwasaki, T., Terra
Scientific Publishing Company, Tokyo, pp61-76.
Imaura, F. et.al. (1993), Estimate of the Tsunami Source of the 1992 Nicaraguan
Earthquake from Tsunami Data, in Geophysical Research Letters, American
Geophysical Union, Washington, vol.20, no.14, pp1515-1518.
Kamel, A. M. (1970), Laboratory study for design of tsunami barrier, in Journal of the
Waterways, Harbors and Coastal Engineering Division, ASCE, Reston,
pp766-779.
Kanamori, H., Kikuchi, M. (993), The 1992 Nicaraguan Earthquake a Slow Tsunami
Earthquake Associated with Subducted Sediments, in Nature, Macmillan, London,
vol.361, pp714-716.
Kimura, Y., Kondo, H., Kuwabara, S., Kawamori, A., (1996), Improvement of Composite
Breakwater on Solid Bottom against Sever Tsunamis, in Coastal Engineering 1996,
ASCE, Reston, vol.2 pp1707-1720.
Ligteringen, H. (2000), lecture notes on the course Ports and Terminals (CT4330), faculty
of Civil Engineering Technical University Delft, Delft.

Mizutani, S., Imamura, F., (2001), Dynamic wave force of tsunamis acting on a structure,
in ITS proceedings, publ.?, plaats?, pp941-948.
Murty, T. S. (1977), Seismic Sea Waves: TSUNAMIS, Department of Fisheries and the
Environment Fisheries and Marine service, Ottawa.
Satake, K. (1994), Mechanism of the 1992 Nicaragua Tsunami Earthquake, in
Geophysical Research Letters, American Geophysical Union, Washington, vol.21,
no.23, pp 2519-2522.
Satake, K. (1995), Linear and Nonlinear Computationsof the 1992 Nicaraguan Earthquake
Tsunami, in Pure and Applied Geophysics, Birkhuser Verlag, Basel, vol.144,
pp. 455-469.
Satake, K. (2002), Tsunamis, in International handbook of Earthquake and Engineering
Seismology, International Association of Seismology & Seismology and Physics of
the Earth's Interior, Committee on Education, and International Association for
Earthquake Engineering, Academic Press, Amsterdam, pp436-451.
Schiereck, G.J. (2001), Introduction to Bed Bank and Shore Protection, Delft University
Publishers, Delft.
Shuto, N (1993), Tsunami intensity and disasters, in Tsunamis in the world, Kluwer
Academic Publishers, Dordrecht, pp197-216.
Soloviev, S.L. (1970), Recurrence of tsunamis in the Pacific, in Tsunamis in the Pacific
Ocean, East-West Center Press, Honolulu, pp.149-164.
Stelling, G., Zijlema, M. (2003), An Accurate and Efficient Finite-difference Algorithm
for Non-hydrostatic Free Surface Flow with Application to Wave Propagation, in the
International Journal for Nummerical methods in Fluids, John Wiley & Sons, Bognor
Regis, vol.43, pp1-23.
Synolaksis, C. (2003), Tsunami and Seiche, in The Earthquake Engineering Handbook,
Chen, W.F., Scawthorn, C, CRC Press, Boca Raton, pp.9.1-9.89
Tanimoto, K. (1983), On the hydraulic aspects of tsunami breakwaters in Japan, in
Tsunamis their science and engineering , Terra Scientific Publisching Company,
Tokyo, pp423-435.

Das könnte Ihnen auch gefallen