Advanced Numerical Models For Simulating Tsunami Waves and Runup
Advanced Numerical Models For Simulating Tsunami Waves and Runup
Editors
Harry Yeh
Oregon State University, USR
Costas Synolakis
University of Southern California, USR
r pWorld Scientific
N E W JERSEY LONDON SINGAPORE - BElJlNG SHANGHAI - HONG KONG - TAIPEI - CHENNAI
A-PDF Merger DEMO : Purchase from www.A-PDF.com to remove the watermark
Published by
World Scientific Publishing Co. Pte. Ltd.
5 Toh Tuck Link, Singapore 596224
USA office: 27 Warren Street, Suite 401-402, Hackensack, NJ 07601
UK office: 57 Shelton Street, Covent Garden, London WC2H 9HE
For photocopying of material in this volume, please pay a copying fee through the Copyright
Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to
photocopy is not required from the publisher.
ISBN-13 978-981-270-012-4
ISBN-10 981-270-012-9
Printed in Singapore.
Preface
v
August 26, 2008 16:56 WSPC/Trim Size: 9in x 6in for Review Volume Preface
vi Preface
States. The format of the second workshop was designed to focus more on
discussions than on formal presentations. Four benchmark problems were
selected before the workshop so that numerical models can be compared,
evaluated and discussed among the participants. During the workshop,
seven discussion themes were organized as follows: laboratory, analytical,
finite-difference, finite-element, vertical-plane models, boundary-integral-
element models, and marker-and-cell models. The presentations and dis-
cussions were edited and published in a book entitled “Long-Wave Runup
Models” (Yeh, et al. 1996).
There was no doubt that the benchmark-problem exercises used in the
second workshop proved extremely useful in identifying absolute and com-
parative modeling capabilities. Overall, in terms of tsunami runup mod-
eling, significant advances had been made between two workshops, due
mainly to the advancement of computational capabilities, and because of
the generation of a large 2-D and 3-D laboratory data set and the fortuitous
field measurements in 1992–1995, all of which have contributed to model
calibrations. The tsunami modeling efforts had become more directed to-
wards their implementation for real tsunami predictions and hind-castings
than ever before. At the time of the 1990 Catalina workshop, large differ-
ences between computed runup results and field measurements might have
been attributed to both errors in the seismic estimates of the source mo-
tion and to the hydrodynamic calculations. Much advances had been made
by the 1995 Friday Harbor workshop; researchers became much more con-
fident in the hydrodynamic calculations, at least for non-breaking waves.
It was equally clear that reduction and even elimination of numerical dis-
persion and numerical dissipation effects would — if not already — soon
become reality. At the same time, the workshop participants recognized
additional and important problems arising from modeling improvements,
such as determination of highly accurate initial wave conditions, modeling
the three-dimensional flow effects, and turbulence. Actual tsunami runup
motions are turbulent; the runup flow patterns, impacts, scouring effects,
and sediment transport are all affected by turbulence in the runup motions.
From 1995 to the summer of 2004, there have been at least six additional
large tsunamis resulting in catastrophic loss of life and property. They are
the Irian Jaya tsunami (Indonesia) in 1996, the Peru tsunamis in 1996 and in
2001, the Papua New Guinea tsunami in 1998, the Turkey tsunami in 2000,
and the Stromboli (Italy) tsunami in 2002. Among these six tsunamis, the
Turkey tsunami was definitely caused by land subsidence and slides associ-
ated with earthquakes, and the Stromboli tsunami was caused by landslides
August 26, 2008 16:56 WSPC/Trim Size: 9in x 6in for Review Volume Preface
Preface vii
caused by volcanic eruption. On the other hand, the source of the Papua
New Guinea tsunami, which killed more than 2000 people and destroyed
completely three villages, remains controversial and has been postulated as
due either to co-seismic seafloor dislocation or sediment slump. Because of
the occurrence of these tsunamis, research interest and efforts on the mod-
eling of landslide generated tsunamis have been intensified in recent years.
The landslide-generated tsunamis have very different characteristics from
the tsunamis generated by earthquakes. The traditional depth-integrated
shallow water equations are not always adequate for modeling the landslide
generated tsunamis. Other noteworthy advances in recent years are the de-
velopment of several computational fluid dynamics models calculating the
nearshore waves and their interactions with structures with the considera-
tion of frequency dispersion and turbulence.
The primary objectives of the third workshop were to provide a
platform for discussing both old and new numerical models and their ap-
plications to various critical issues concerning tsunami runup and wave-
structure interactions. To accomplish this goal, four benchmark problems
were selected and posted on the workshop website before the workshop.
http://www.cee.cornell.edu/longwave/. The workshop participants were
given the solutions, in the form of laboratory data or analytical solutions,
a week before the workshop. The workshop participants discussed their nu-
merical model and the comparisons between their numerical results and
solutions for the benchmark problems. Their presentations are also posted
on the workshop website. The four benchmark problems are as follows:
This review volume is divided into two parts. The first part includes
five review papers on various numerical models. Pedersen provided a brief
but thorough review on the theoretical background for depth-integrated
wave equations, which are employed to simulate tsunami runup. LeVeque
and George describe high-resolution finite volume methods for solving the
nonlinear shallow water equations. They have focused their discussion on
the applications of these methods to tsunami runup. In recent years, sev-
eral advanced 3D numerical models have been introduced to the field of
Coastal Engineering to calculate breaking waves and wave-structure inter-
actions. These models are still being developed and are at different stage of
August 26, 2008 16:56 WSPC/Trim Size: 9in x 6in for Review Volume Preface
viii Preface
References
Liu, P. L.-F., Synolakis, C. and Yeh, H., A report on the international
workshop on long wave runup, J. Fluid Mech. 229 (1991), 678–88.
Yeh, H., Liu, P. L.-F. and Synolakis, C. (ed.), Long-Wave Runup Models
(World Scientific, 1996), 403p.
January 10, 2008 19:59 WSPC/Trim Size: 9in x 6in for Review Volume Cliff
ix
January 10, 2008 19:59 WSPC/Trim Size: 9in x 6in for Review Volume Cliff
funding until Dr. Astill took over the program. Dr. Astill’s enthusiasm and
steady support in tsunami research over the last 20 years is the catalyst
for all the achievement made to date. He has indeed been a mentor for the
US tsunami research community until his retirement from NSF in January
2004.
Dr. Astill had a vision that tsunami research ought to be interdisci-
plinary. On one hand, he supported the fundamental research in develop-
ing mathematical models and performing basic laboratory experiments and
numerical simulations for various aspects of physical processes related to
tsunami generation, propagation and runup. On the other hand, he rec-
ognized the importance of the hazard mitigation practice by supporting
research in the areas of social sciences, policy-making and planning, and
data management. Fifteen years ago, he initiated such multi-disciplinary
activities by strongly encouraging researchers to submit cooperative group
proposals, which were unheard of in those days. Dr. Astill was always proac-
tive in promoting promising young researchers. In many occasions, using
himself as a source of information, he introduced young researchers to senior
researchers and encouraged them to form a possible collaborative team.
Dr. Astill also firmly believed that an effective means to develop multi-
discipline and cross-disciplinary research programs is through informal
small group meetings and formal workshops. Over the years he has sup-
ported a numerous such meetings and workshops (see the list below).
Through his encouragement, the tsunami community has expanded its
membership to include not only hydrodynamists and seismologists but
also geologists, social scientists, oceanographers, mathematicians, and data
managers. When we look back, Dr. Astill led us to the research mode of “col-
laboratory” more than 15 years ago, which is the fundamental philosophy
of NSF NEES project (Network for Earthquake Engineering Simulations:
http://www.nees.org/index.php). Dr. Astill was not afraid of offering his
opinions and making suggestions to workshop organizers. For example, the
procedure of “Blind Tests” in validating numerical models was first recom-
mended by Dr. Astill and implemented in the 1995 Second International
Workshop on Long-Wave Runup Models at Friday Harbor. Prior to the
workshop — April 1994 — Dr. Astill wrote:
Dr. Astill was a strong supporter for the reconnaissance field surveys for
tsunamis. Since tsunamis are rare and high impact events, lessons learned
from the field surveys are invaluable. Dr. Astill believed that the strong
January 10, 2008 19:59 WSPC/Trim Size: 9in x 6in for Review Volume Cliff
links between the real-world observations and the laboratory and/or theo-
retical understandings are crucial for natural hazards research. Based on the
field observations, many topics of fundamental research were identified and
several theoretical hypotheses were validated. Dr. Astill’s effort to support
reconnaissance tsunami survey began after the 1992 Nicaragua Tsunami,
which was the first major tsunami since 1983 Nihonkai Chubu Tsunami.
He developed several options, which enabled researchers to quickly go to
the filed without being burden by paper works. He clearly understood the
importance of the survey and its critical timing for capturing perishable
data. The tsunami community made a numerous tsunami surveys during
his tenure (see the list below), which resulted in substantial advances in
tsunami research.
Dr. Astill also proactively encouraged international collaboration by
inviting international experts to serve on the advisory committees for sev-
eral group research projects. He also suggested that we should participate
more in international conferences and meetings: for example, the Interna-
tional Travel Grants to attend the General Assembly of European Geophys-
ical Society in 2001 and 2002 were funded by Dr. Astill.
Dr. Astill’s career at NSF can be best described by his own words on
May 31, 2004.
Tsunami Sources:
o Workshop on Seafloor Deformation Models, Santa Monica, CA,
May 1997
o Workshop on the Prediction of Underwater Landslide and Slump
Occurrence and Tsunami Hazards, Los Angeles, CA, March 2000
Development of NEES:
o Workshop for Tsunami Research Facilities, Baltimore, May 1998
o Ad-Hoc Planning Working Group for the NSF Major Research
Equipment Program: National Earthquake Engineering Simula-
tions, 1997–1999
o Workshop on Research with NEES Tsunami Facility, Corvallis,
OR, March 2001
Contents
Preface v
Chapter 1
Modeling Runup with Depth Integrated Equation Models . . . . . . . . . 3
G. Pedersen
Chapter 2
High-Resolution Finite Volume Methods for the Shallow Water
Equations with Bathymetry and Dry States . . . . . . . . . . . . . . . . . . . . . . . 43
R. J. LeVeque and D. L. George
Chapter 3
SPH Modeling of Tsunami Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
B. D. Rogers and R. A. Dalrymple
Chapter 4
A Large Eddy Simulation Model for Tsunami and Runup
Generated by Landslides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
T.-R. Wu and P. L.-F. Liu
Chapter 5
Free-Surface Lattice Boltzmann Modeling in Single Phase Flows . . . 163
J. B. Frandsen
Chapter 6
Benchmark Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
P. L.-F. Liu, H. Yeh and C. E. Synolakis
xvii
August 26, 2008 16:56 WSPC/Trim Size: 9in x 6in for Review Volume Contents
xviii Contents
Chapter 7
Tsunami Runup onto a Plane Beach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231
Z. Kowalik, J. Horrillo and E. Kornkven
Chapter 8
Nonlinear Evolution of Long Waves over a Sloping Beach . . . . . . . . . . 237
U. Kânoǧlu
Chapter 9
Amplitude Evolution and Runup of Long Waves;
Comparison of Experimental and Numerical Data
on a 3D Complex Topography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
A. C. Yalciner, F. Imamura and C. E. Synolakis
Chapter 10
Numerical Simulations of Tsunami Runup onto a
Three-Dimensional Beach with Shallow Water Equations . . . . . . . . . . 249
X. Wang, P. L.-F. Liu and A. Orfila
Chapter 11
3D Numerical Simulation of Tsunami Runup onto a
Complex Beach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255
T. Kakinuma
Chapter 12
Evaluating Wave Propagation and Inundation Characteristics
of the Most Tsunami Model over a Complex 3D Beach . . . . . . . . . . . . 261
A. Chawla, J. Borrero and V. Titov
Chapter 13
Tsunami Generation and Runup Due to a 2D Landslide . . . . . . . . . . . 269
Z. Kowalik, J. Horrillo and E. Kornkven
Chapter 14
Boussinesq Modeling of Landslide-Generated Waves and Tsunami
Runup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 273
O. Nwogu
August 26, 2008 16:56 WSPC/Trim Size: 9in x 6in for Review Volume Contents
Contents xix
Chapter 15
Numerical Simulation of Tsunami Runup onto a Complex Beach
with a Boundary-Fitting Cell System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279
H. Yasuda
Chapter 16
A 1-D Lattice Boltzmann Model Applied to Tsunami Runup onto a
Plane Beach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283
J. B. Frandsen
Chapter 17
A Lagrangian Model Applied to Runup Problems . . . . . . . . . . . . . . . . . 311
G. Pedersen
Appendix
Phase-Averaged Towed PIV Measurements for Regular Head Waves
in a Model Ship Towing Tank . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319
J. Longo, J. Shao, M. Irvine, L. Gui and F. Stern
April 3, 2008 3:12 WSPC/Trim Size: 9in x 6in for Review Volume divider
Part 1
Review Papers
This page intentionally left blank
August 8, 2008 11:12 WSPC/Trim Size: 9in x 6in for Review Volume 01˙Pedersen
CHAPTER 1
G. Pedersen
Mechanics Division, Department of Mathematics, University of Oslo
E-mail: [email protected]
This survey starts with a sketch of long wave theory, reflecting some of
its diversity. Next, general aspects of runup on sloping beaches are briefly
discussed in the scope of long waves, including important analytical solu-
tions, breaking criteria, significance of non-hydrostatic effects and exper-
iments. Then the literature on numerical modeling of runup is reviewed.
The models are loosely organized into classes and then described chrono-
logically within each class. Primarily, we refer models that are published
in international journals, while matter from the vast number of pro-
ceedings and internal reports are mentioned occasionally. Moreover, we
highlight the verification of the numerical methods by comparison to an-
alytical solutions or experiments. Applications to real tsunami or storm
surge events are, with a few exceptions, not included.
1. Introduction
Comprehension of wave runup on beaches is essential for prediction of beach
erosion and coastal impact of tsunamis and storm surges. This is one reason
for the lasting attention that runup topics have received in the literature
of hydrodynamics and coastal engineering. Moreover, this complicated, yet
commonplace phenomenon, that can be observed at every beach with in-
cident swells or wind waves, pose mathematical and conceptual challenges
that appeal to many scientists. Wave runup also inherits similarities to other
important phenomena in hydrodynamics, such as “green water” (wave over-
topping on vessels), slamming and interaction between the ocean pycnocline
and bathymetry.
As for many other wave phenomena, long wave theory has been crucial
for our understanding of wave runup on beaches and derivation of closed
form solutions supporting this understanding. Also numerical models for
3
August 8, 2008 11:12 WSPC/Trim Size: 9in x 6in for Review Volume 01˙Pedersen
4 G. Pedersen
large scale ocean waves in general, such as tides, storm surges and tsunamis,
have traditionally been based on shallow water theory, due to the simplic-
ity and efficiency of this formulation. Certainly, more general models are in
progress, also for runup on beaches23 – 25,43,50,51,70,109 , but do still involve
heavy computations. For instance, most fair sized three-dimensional wave
problems are still dependent on approximate methods, such as long wave
theory. In fact, even an idealized, two-dimensional simulation modeling the
propagation of a tsunami wave or swell from deep water, through shoal-
ing/surf and to runup on a shore will often border on the limits of models
based on the Navier-Stokes equations or full potential theory. Therefore,
for many applications we must rely on long wave models also in the years
to come.
In the context of depth integrated equations, the moving shoreline on a
beach is in some respects similar to a free water surface that is deformed
by waves; the flow field is confined by a time dependent boundary with a
position that is unknown a priori. Furthermore, in both cases the surface
excursions correspond to motion of material particles with gravity as restor-
ing force. However, there are important differences. In runup the medium is
“thinning” toward the shoreline, resulting in a singularity in the governing
equations, though not in the physically relevant solutions. In addition, wave
breaking and topographical effects are often important for the near-shore
flow. If we look beyond the long wave (depth-integrated) description, other
aspects appear, such as the dynamics of a contact point between a free sur-
face and a rigid no-slip boundary, boundary layers, and effects of bottom
porosity. These are presumably most important on laboratory scale and are
anyway outside the scope of the present article.
An early review on long wave runup is found in Meyer and Taylor
(1972)67 . The predecessors to the “The Third International Workshop On
Long-Wave Runup Models”, leading to this volume, also initiated articles
summarizing aspects of runup research53,113. Among the contributions are
overviews of finite difference and element methods for tsunamis, that also
address long wave runup32,105 .
This review is meant to provide a brief and updated status on long wave
runup models, with an inclination toward applicability of the different long
wave approximations. We do not include related themes such as runup on
vertical walls, edge waves, rip currents, and impact problems. Moreover,
we focus on the principal aspects of physics and modeling rather than case
studies of real events of tsunamis and storm surges. Wave breaking is the
subject of Chapter 2 (by LeVeque and George) in this volume48 . Still, for
August 8, 2008 11:12 WSPC/Trim Size: 9in x 6in for Review Volume 01˙Pedersen
6 G. Pedersen
ηt = −∇ · (h + η)(v + µ2 M) + O(µ4 ),
(2)
2 2 1 2
vt + ∇(v ) = −∇η − µ z ∇∇ · vt + zα ∇∇ · (hvt )
2 2 α
+ µ2 ∇(D1 + D2 + 2 D3 ) + O(µ4 ) + N + E, (3)
with
1 2 1 2 2 2 1
M = z − (h − hη + η ) ∇∇ · v + zα + (h − η) ∇∇ · (hv),
2 α 6 2
1 2 1
D1 = η∇ · (hvt ) − zα v · ∇∇v − zα v · ∇∇ · (hv) − (∇ · (hv))2 ,
2 2
1
D2 = η 2 ∇ · vt + ηv∇∇ · (hv) − η∇ · (hv)∇ · v,
2
1
= η 2 v · ∇∇ · v − (∇ · v)2 ,
D3
2
and where the index t denotes temporal differentiation, h is the equilibrium
depth, η is the surface elevation, v is the horizontal velocity evaluated at
z = zα (x, y) and ∇ is the horizontal component of the gradient operator.
The heuristic terms N and E in the momentum equation (3) represent
bottom drag and artificial diffusion and will be discussed below. This par-
ticular form of the leading dispersion (0 µ2 ) term was discussed and tested
by Nwogu (1993)71 , while additional nonlinearity was added by Wei et al.
August 8, 2008 11:12 WSPC/Trim Size: 9in x 6in for Review Volume 01˙Pedersen
h2
h
vt + ∇(v2 ) = −∇η + µ2 ∇∇ · (hvt ) − ∇∇ · vt
2 2 6
+ O(µ2 , µ4 ) + N + E. (5)
This set has been used much in numerical simulations and for the first time
by Peregrine (1967)78 for the shoaling of a solitary wave. We will subse-
quently refer to formulations like (4,5) as “standard” Boussinesq equations,
whereas those with more accurate dispersion properties, such as (2,3) with
zα = −0.531h, will be denoted as “improved”.
Further information on, and extensions of, higher order long wave equa-
tions are found in a number of papers in the literature16,58 – 61 . However,
progress has primarily been made with respect to higher order nonlinearity
8 G. Pedersen
Eq. 1 q
∗
√c zα = ( 13 − 1)h
gd
zα = −0.531h
KdV
LSW
λ/d
Fig. 1. The phase speed as function of wavelength for long wave equations compared
to that of the fully inviscid set. The curve for the Korteweg-deVries (KdV) equation is
included for comparison.
where subscripts indicate partial derivation, H = h+η, and u and v are the
velocity components in the x- and y-directions, respectively. Naturally, a
corresponding equation applies to the y-direction. We observe that the right
hand side is a momentum source due to a sloping bottom. Often the depth
integrated fluxes are used as primary unknowns instead of the velocities.
This may cause difficulties during runup, since velocities are needed for the
propagation of the shoreline and the relation between fluxes and velocities
may be poorly defined when H → 0.
A popular transformation of the NLSW equations is obtained through
introduction of characteristics. For plane waves the characteristic equations
become
p
βt± + (u ± c)βx± = hx , where β ± = u ± 2c, c = h + η. (7)
Characteristics played an important role in the early analytic work on runup
of breaking waves41,87 and are used in some numerical runup models as
well72,101,102 . In these references the technique is extended to two horizontal
directions by operator splitting102 or a partial characteristic description72 .
At the shoreline c becomes zero and both the characteristic equations and
variables in (7) coalesce. Alternative equations or extrapolation must then
be employed in a numeric model. Moreover, as long as ∂(h + η)/∂x is
nonzero at the shoreline, the spatial derivative of c is non-finite, which is a
challenge from a computational point of view. Still, as demonstrated in the
references runup can be treated accurately and efficiently by (7).
If the two-dimensional NLSW equations are linearized, we may eliminate
the velocity to obtain a very simple equation
ηtt − (hηx )x = 0. (8)
It might seem counter-intuitive for such an apparently nonlinear phe-
nomenon, but linear equations work surprisingly well on a number of runup
problems.
10 G. Pedersen
where A, δ are constants and J0 is the Bessel function of zero order that
may be expanded in even powers of its argument. For the singular solution
of (8) we have η ∼ ln(x) for small x, while the horizontal velocity has a pole
of order 1. As a consequence there is a volume source at the singularity.
Hence, runup models that do not conserve volume at the shoreline should be
checked extra carefully for accuracy and spurious behavior. The mathemat-
ical structure with regular and singular solutions carries over to nonlinear
equations as well as full potential theory33 and is further discussed in a long
wave context by Meyer65,66 and others.
For nonlinear equations the shoreline is signified by H = h + η =
0. In addition the position of the shoreline must be traced. We confine
the discussion to two dimensions and assume that the shoreline always
August 8, 2008 11:12 WSPC/Trim Size: 9in x 6in for Review Volume 01˙Pedersen
12 G. Pedersen
Such terms have often been included to stabilize numerical models, under
cover of being a physical eddy viscosity effect. With a constant ν (13)
is quite inefficient as a tool for controlling wave breaking. Zelt (1991)114
made ν dependent on the velocity gradient in a similar term. ν was zero
until the gradient reached a threshold value and then increased with the
gradient beyond this. The critical value of the velocity gradient was related
to the highest stable solitary wave, which has an amplitude 0.78 times the
depthb . Kennedy et al.42 expressed ν by ηt and pursued the idea further by
including an explicit time relaxation for ν after activation of the diffusion.
This method has been demonstrated to reproduce bores from experiments
very well42,57 . However, as pointed out in a reference56 , this procedure may
not recognize a nearly stationary bore on a current, like the one that often
forms during drawdown at beaches.
14 G. Pedersen
v u, v u, v
y y
u η u ∆y η ∆y
v u, v u, v
x x
Fig. 2. Left panel: The C-grid that defines a basic configuration for mass conserving
FD methods. It is used also for many Navier-Stokes models with η replaced by pressure.
Right panel: The less used B-grid. This is well suited for representation of deformation,
as needed in Lagrangian models. u and v are the velocity compenents in the x- and
y-directions, respectively.
z∗ λ
A x∗
θ ` d
φλλ − (σφσ )σ = 0.
16 G. Pedersen
patching conditions. This idea has later been exploited by others, among
which the work of Synolakis (1987)94 has made most impact. Before going
into details we will remark on some general properties of this approach.
Since the transformed equations, as well as far-field and patching are lin-
ear, the solutions in σ and λ equal linear solutions in the physical plane
with σ 2 /16 and λ/2 substituted for x and t, respectively. Nonlinearity en-
ters first through the transformation (14) back to the physical variables.
When u is zero, as at the point of maximum excursion of the shoreline,
the nonlinearity also disappears from this transformation. Hence, the max-
imum runup height for the combined linear/non-linear theory will be as for
a purely linear one. Moreover, ` only gives a time shift when the constant
depth region is governed by the LSW equations. For the wave tank problem
in Fig. 3 the maximum runup height then becomes
R = AF (κ), (15)
3◦
5◦
7.18◦
R/A
Eq. (17)
10.54◦ Boussinesq
LSW
A/d
Fig. 4. Shallow water and Boussinesq predictions for maximum runup heights of solitary
waves for different beach inclinations as indicated on the curves. The thick and thin
columns represent breaking limits, (20), during runup and drawdown, respectively.
the closest node half a grid increment from the shore to implement the
condition of zero flux, makes this a trivial task. The results are subsequently
shown in Fig. 4.
The effect of dispersion on solitary wave runup is frequently overlooked
when NLSW models are compared to experiments or dispersive models are
compared with analytic NLSW solutions. This makes the model assessments
imprecise, with a negative bias concerning model performance. We may
employ optics to obtain a simple estimate on the variation of dispersion
with depth. Assuming that the period is unchanged during shoaling, we
1
find that the wavelength is reduced as h− 2 and the relative size of the
leading dispersion term in the wave celerity vanishes proportional to h near
shore. However, this result will not be valid in the proximity of the shore and
nonlinear effects will lead to steepening of the wave front, which in turn
enhance dispersion. Therefore, we resort to employment of a Lagrangian
Boussinesq type model34,76 (Sec. 4.1), that was readily at hand for the
author. Converged results of both the LSW and Boussinesq models are
compared to (17) in Fig. 4, that also contains the breaking criteria as given
in Eq. (20). Shallow water theory over-predicts runup slightly for the larger
A/d, but is surprisingly close to the Boussinesq theory even beyond the
hydrostatic breaking limit. For large θ and small A/d, meaning small κ,
the numeric solutions coincide, while the asymptotic formula (17) deviates.
Hence, we clearly observe the asymptotic nature of the closed form solution
August 8, 2008 11:12 WSPC/Trim Size: 9in x 6in for Review Volume 01˙Pedersen
18 G. Pedersen
3◦
R/A
Eq. (17)
Boussinesq
10.54◦ 0.5
0.9
NLSW
A/d
Fig. 5. Maximum runup heights for solitary waves. Results from combined Boussi-
nesq/NLSW theory are marked by the transition depth h(ac ) at equilibrium.
η∗
d
x∗ /d
1 Du∗
g Dt∗
full full
hydrostat. hydrostat.
x∗ /d x∗ /d
Fig. 6. Runup of a solitary wave with A/d = 0.15 and θ = 7.18◦ . Upper panel: Surfaces
for the incident wave, an early stage in runup, an intermediate stage and maximum
runup. Lower panels: Comparison of the full acceleration to the part that stems from
the hydrostatic pressure gradient. Left: Vicinity of beach for the second time (thin solid
line in upper panel). Right: The incident wave (rightmost wave profile in upper panel).
the NLSW equation will often do for the near-shore region and runup. An
equally important lesson is that dispersion should not readily be ignored for
the constant depth propagation in simulations of wave tank experiments.
There are other analytic runup solutions that deserve mentioning. Some
are related to those above, but involve different initial conditions, slides or
modified geometries3,6,39,49,74,77,92,95,97,103 .
Another class of solutions contains the eigenoscillations in basins of
parabolic bottom shapes that are summarized by Thacker100. There exist
closed form solutions of the NLSW equations for the two lowest eigenmodes.
In the first mode the whole fluid body moves back and forth in a uniform
horizontal motion, driven by a uniform pressure gradient associated with a
linear surface. Setting d equal to the maximum depth, identifying the half-
width of the of equilibrium surface with L and choosing d as the amplitude
of the lowest harmonic in η ∗ , we obtain simple formulas for the 2D case
√ √ 1 √
η = cos( 2t)x − cos2 ( 2t), u = − √ sin( 2t). (18)
4 2
August 8, 2008 11:12 WSPC/Trim Size: 9in x 6in for Review Volume 01˙Pedersen
20 G. Pedersen
c0
H∗
d
2c0
x∗ /d
Fig. 7. The dambreak solution according to (19). Surfaces are depicted for t ∗ = 0,
∆t∗ .., where ∆t∗ = 2d/c0 .
22 G. Pedersen
Plunging breakers are outside the long wave regime, but fully developed
bores may be crudely presented as discontinuities in shallow water solutions.
This theme is treated in Chapter 2 (by LeVeque and George) in this volume
and we will merely attach a few comments on the runup aspects. The most
remarkable feature for bore runup28,41,67,87 is that the bore collapses rapidly
at the shoreline, giving birth to a long thin runup tongue with a height
that is proportional to the square of the distance from the instantaneous
shoreline. The swash is dominated by gravity and the runup trajectory
becomes a parabola in the x, t plane. This transition from a steep incident
wave to a swash is closely related to dam-break34,67,81 (see Fig. 7). The
runup of steep non-breaking waves and finite width bores are quite similar
to that of the idealized bore solution; a steep front is transformed into a
thin swash zone, where gravity often dominates over pressure effects34,114 .
Moreover, there are large accelerations and velocities in the early phases of
runup, when the steep front vanishes. Small timing errors in experiments or
models may lead to large deviations temporarily, while quantities like runup
height and overall pattern is less influenced. This may cause an under-rating
of model performance. The occurrence of a very thin runup layer that even
may vanish as the square of the distance from the tip, seemingly argue
against the application of a linear onshore extrapolation of the surface that
is employed in some models. Even more so, perhaps, since the rundown
phase display even thinner layers. On the other hand, bottom roughness,
finite width bores and turbulence will modify the swash zone dynamics.
Since, in addition, the feedback of a thin swash to the main body of fluid
may be small, it is quite likely that surface extrapolation may be acceptable
also in such cases.
3.3. Experiments
There is only a limited number of published experimental investigations
that are commonly used for assessment of theoretic models. Hall and Watts
(1953)26 measured runup heights of solitary waves for different ampli-
tudes and beach inclinations. Synolakis (1987)94 published runup heights
of breaking and non-breaking solitary waves on a 1.0 in 19.85 slope and
amplitudes in the range 0.004 < A/d < 0.6. In addition, data from a set
of wave gauges, on and off the beach, were combined to rather complete
surface elevations at different times. These surfaces have been valuable for
validation of numerical models, in particular for the breaking cases. One
such data-set is displayed in Fig. 10.
August 8, 2008 11:12 WSPC/Trim Size: 9in x 6in for Review Volume 01˙Pedersen
island base
4.64
7.2
13
equilibrium shoreline
27.43
Wavemaker
Fig. 8. Definition sketch of conical island experiment, with scales in meter. The outer
dashed circle is the base of the island, the fully drawn circle is the equilibrium shoreline
when d = 0.32m.
c TheBoussinesq model34 employed in Sec. 3.1 indicates breaking during retreat for
A/d = 0.2, but not for the smaller amplitudes.
August 8, 2008 11:12 WSPC/Trim Size: 9in x 6in for Review Volume 01˙Pedersen
24 G. Pedersen
90 1 a) 90 1 b) 90 1 c)
120 60 120 60 120 60
Fig. 9. Measured2 (stars) runup on conical island with recent simulations by Lynett
et al.57 (lines). (a): A/d = 0.05, (b): A/d = 0.1, (c): A/d = 0.2. Reproduced from Lynett
et al. by courtesy of the Coastal Engineering.
4. Runup Models
We may recognize at least three key problems in runup modeling:
26 G. Pedersen
28 G. Pedersen
a series of NLSW methods for tidal flooding, mainly based on the C-grid, on
a small amplitude second eigenmode in a parabolic basin100 . To a greater
or lesser extent all techniques gave near-shore fluctuations. However, the
degree of nonlinearity was too small to allow any inference on the general
runup performance of the models.
Apart from simulations of flooding based on linear theory86 , the first
proper runup model was reported by Sielecki and Wurtele (1970)90 . They
employed three different finite difference techniques, among them the Lax-
Wendroff method, for the NLSW equations with the Coriolis term. A sim-
ple extrapolation combined with an asymmetric representation was used to
follow the shoreline. Among the test cases were CG-i5 , that was well repro-
duced except for a small discrepancy at shoreline, and a three-dimensional
eigenoscillation in a parabolic basin100 that was simulated perfectly. In 1975
Flather and Heaps13 published a 3D FD technique for the NLSW equations,
based on the C-grid, with a scheme to update wet and dry regions. However,
the model was applied to tidal oscillations in a bay without any verification
of its runup performance. Hibberd and Peregrine (1979)27 employed the
Lax-Wendroff technique to the NLSW equations in conservative form and
obtained a model for non-breaking waves as well as bores. According to the
authors the simple shoreline method of Sielicki and Wurtele90 did not suf-
fice for these problems. After attempting various methods they settled for a
predictor-corrector sequence of linear extrapolations and applications of the
physical equations in the vicinity of the shoreline. Even though some near-
shore deviations were identified, both CG-i5 and CG-s5 were reproduced
well. Very close agreement were reported with analytical runup heights67 ,
while the theoretical values overshot measurements69 . This was explained
by viscous effects that must influence the thin swash zone significantly on
laboratory scale. Later Kobayashi (1987)46 adopted the method and applied
it to runup on rough slopes.
Kowalik and Murty (1993)47 employed linear onshore extrapolation of
velocity and surface elevation with a C-grid discretization of the NLSW
equations. Their method gave reasonably good results for CG-i5 and the
first mode in a parabolic basin, but with significant noise at the shoreline.
Then they used the method to compute the inundation in a real tsunami
event.
With the experiments on solitary wave runup on a conical island Briggs
et al. (1995)2 (Sec. 3.3) provided what still is the only 3D dataset well
suited for verification of models. In a companion paper Liu et al.52 pre-
sented NLSW simulations with a C-grid model with upwind representation
August 8, 2008 11:12 WSPC/Trim Size: 9in x 6in for Review Volume 01˙Pedersen
30 G. Pedersen
0.6
0.4
0.2
0 a)
−0.2
−20 −15 −10 −5 0 5 10
0.6
0.4
0.2
0 b)
−0.2
−20 −15 −10 −5 0 5 10
0.6
0.4
z/h
0.2
0 c)
−0.2
−20 −15 −10 −5 0 5 10
0.6
0.4
0.2
0 d)
−0.2
−20 −15 −10 −5 0 5 10
Fig. 10. Breaking solitary wave runup. Experiments from Synolakis94 are indicated
by markers, while the fully drawn lines are solutions from Lynett et al.57 . Panel (c)
includes also results from Zelt114 (dots), Titov and Synolakis101 (dash-dot) and the N-S
computations of Lin et al.51 (dashes). Reproduced from Lynett et al. by courtesy of the
Coastal Engineering.
was accurately reproduced. Also the conical island test was made with
good results. For runup of a breaking wave (A/d = 0.3) impressive agree-
ment was obtained with both experiments94 and a turbulent Navier-Stokes
simulation51 (Fig. 10). Even though this is unlikely to be the reason for the
good outcome, it must be noted that the choice of the coefficient in the
August 8, 2008 11:12 WSPC/Trim Size: 9in x 6in for Review Volume 01˙Pedersen
32 G. Pedersen
bottom-drag was optimized. As pointed out by the authors, there are still
some unresolved shortcomings in the model for features like over-topping
and reentry of a runup tongue into quiescent water56 .
Another line of runup modeling is based on the NLSW equations in con-
servative form, finite volume techniques and approximate Riemann solvers
to include bores as sharp jumps in the solution. This branch will be covered
in depth in Chapter 2 in this volume by LeVeque and George48 . Hence, we
refer briefly only a selection of contributions, with emphasis on runup tests.
The shoreline may be treated by the special Riemann problem related to
penetration of gas into vacuum or dam-break. However, this leads to overes-
timation of the pressure gradient and acceleration at the shoreline. Hence,
modelers often resort to complete or partial exclusion of very shallow cells
from the general computation scheme. The flux and source terms in the
momentum equation (see below Eq. (6)) are generally separated by opera-
tor splitting, even though Watson et al. (1992)106 suggested the inclusion
of the source term in the Riemann problem by a local transformation to an
accelerating frame of reference. Sleigh et al. (1998)91 invoked a discretiza-
tion into triangular volumes which enables flexible local grid refinement.
The method was tested on various dam-break problems and bore-runup.
Good agreement was observed with the analytical solution for the standard
dam-break problem, save from small deviations at the toe. Two-dimensional
simulations were presented by Hu et al. (2000)30 including runup of non-
breaking solitary waves, that agreed closely with an analytic solution 94 ,
and the standard dam-break problem. The latter case demonstrated that
exemption of shallow cells provide a better solution near the fluid tip than
application of a modified Riemann solution. In Brocchini et al. (2001)3 oper-
ator splitting was employed on a rectangular grid and a Riemann technique
was used for the shoreline. Results compare well with the analytic solution
CG-i5 and the standard dam-break problem. However, for the latter there
was noticeable spurious behavior close to the water line. The particular
strength of Hubbard and Dodd (2002)31 is a hierarchical mesh refinement,
which means that finer rectangular grids may be superimposed on coarser
ones in regions that require high resolution, as the shore and swash zone.
Close agreement, save for small irregularities at the shore, was obtained
with CG-s5 . Surprisingly, the reproduction of the first nonlinear eigenmode
in a parabolic basin was less good.
A common objective for this kind of models is to preserve sharp bore
fronts. It would have been interesting to see a detailed comparison with
real, finite width bores, as breaking solitary waves94 .
August 8, 2008 11:12 WSPC/Trim Size: 9in x 6in for Review Volume 01˙Pedersen
The finite element method has a much weaker merit in Eulerian runup
modeling than the finite difference counterpart. Several methods for in-
cluding a moving shoreline in a static FE grid have been proposed105 , but
few have been validated by comparison to reference solutions or system-
atic grid refinement. An interesting attempt on solving a two-dimensional
Boussinesq equation with the differentiable Hermitian cubic shape func-
tions was made by Gopalakrishnan and Tung (1983)18 . The pressure gra-
dient was used directly to compute the shoreline acceleration which then
was integrated to yield the position. The shore element was then modified
at each time step, and split/deleted when appropriate. However, the only
application was solitary wave runup, without any comparison to other so-
lutions nor application of grid refinement tests. Moreover, the presented
solutions clearly displayed noise and spurious features. Later the method
was modified17 to the three-dimensional NLSW equations and applied to
tidal flooding, again with inadequate testing of the runup performance.
Umetsu (1995)104 employed partially dry triangular elements, where the
velocities on dry near-shore nodes are imposed as averages of the values
from the adjacent wet ones. The technique performed well for broken dam
flow, but produced outspoken deviation from the CG-i5 , in particular close
to the shore. In Takagi (1996)98 the technique was applied to the conical
island problem with reasonably good, but not convincing, result.
5. Discussion
The crucial point in runup modeling is the representation of the shoreline
motion, for which the presence of a singularity (Sec. 2.4) calls for caution.
The literature refer many rather crude as well as sophisticated methods for
the moving boundary. Even though noise in the vicinity of the shoreline is
often observed there are few severe problems that have been attributed to
the singularity. It is tempting to assume that this problem is not that grave
after all, or that a logarithmic singularity is too weak to be fatal. On the
other hand, too many authors have treated numerical convergence rather
loosely. More rigid grid refinement studies might reveal that there indeed
are irregularities at the shore, even though they in most applications will
be overshadowed by other error sources. Anyway, the bottom line is that
there is a diversity of working runup models about.
The recent progress in long wave runup modeling may roughly be di-
vided into two directions: one where the runup facility is integrated in gen-
eral purpose wave propagation models with high order inherent dispersion
August 8, 2008 11:12 WSPC/Trim Size: 9in x 6in for Review Volume 01˙Pedersen
34 G. Pedersen
Acknowledgments
The author wishes to thank Professor P. L.-F. Liu and Professor P. Lynett
for their helpful assistance.
References
1. A. Balzano. Evaluation of methods for numerical simulation of wetting and
drying in shallow water flow models. Coast. Eng. 34, 83–107 (1998).
2. M. J. Briggs, C. E. Synolakis, G. S. Harkins and D. R. Green. Laboratory
experiments of tsunami runup on circular island. Pure and Applied Geo-
physics 144(3/4), 569–593 (1995).
3. M. Brocchini, R. Bernetti, A. Mancinelli and G. Albertini. An efficient solver
for nearshore flows based on the WAF method. Coast. Eng. 43, 105–129
(2001).
4. G. F. Carrier. Gravity waves on water of variable depth. J. Fluid Mech. 24,
641–659 (1966).
5. G. F. Carrier and H. P. Greenspan. Water waves of finite amplitude on a
sloping beach. J. Fluid Mech. 4, 97–109 (1958).
6. G. F. Carrier, T. T. Wu and H. Yeh. Tsunami run-up and draw-down on a
plane beach. J. Fluid Mech. 475, 79–99 (2003).
7. Y. Chen and P. L.-F. Liu. Modified Boussinesq equations and associated
parabolic models for water wave propagation. J. Fluid Mech. 288, 351–381
(1995).
8. Q. Chen, R. A. Dalrymple, J. T. Kirby, A. B. Kennedy and M. C. Haller.
Boussinesq modeling of a rip current system. J. Geophys. Res. 104, 20617–
20637 (1999).
9. Q. Chen, J. T. Kirby, R. A. Dalrymple, A. B. Kennedy and A. Chawla.
Boussinesq modeling of wave transformation, breaking, and run-up. Part
II: 2D. J. Waterw., Port, Coast., Ocean Engrg. 126(1), 48–56 (2000).
August 8, 2008 11:12 WSPC/Trim Size: 9in x 6in for Review Volume 01˙Pedersen
36 G. Pedersen
10. Y.-S. Cho, K.-Y. Park and T.-H. Lin. Run-up heights of nearshore tsunamis
based on quadtree grid system. Ocean engineering 31, 1093–1109 (2004).
11. E. D. Cokelet. Breaking waves. Nature 267, 769–774 (1977).
12. B. Elfrink and T. Baldock. Hydrodynamics and sediment transport in the
swash zone: a review and perspectives. Coastal Engineering 45, 149–167
(2002).
13. R. A. Flather and N. S. Heaps. Tidal computations for Morecambe Bay.
Geopshys. J.R. Astr. Soc. 42, 489–517 (1975).
14. K. Fujima. Application of linear theory to the computation of runup of
solitary waves on a conical island. In Long-wave runup models, Eds. H. Yeh,
C. E. Synolakis and P. L.-F. Liu (World Scientific Publishing Co., 1996),
p. 221–230.
15. B. Gjevik and G. Pedersen. Run-up of long waves on an inclined plane.
Preprint series in applied mathematics, Dept. of Mathematics, University
of Oslo (1981).
16. M. F. Gobbi, J. T. Kirby and G. Wei. A fully nonlinear Boussinesq model
for surface waves. Part 2. Extension to O(kh)4 . J. Fluid Mech. 405, 181–210
(2000).
17. T. C. Gopalakrishnan. A moving boundary circulation model for regions
with large tidal flats. Int. J. Num. Meth. Eng. 28, 245–260 (1989).
18. T. C. Gopalakrishnan and C. C. Tung. Numerical analysis of a moving
boundary problem in coastal hydrodynamics. Int. J. Num. Meth. Fluids 3,
179–200 (1983).
19. D. G. Goring. Tsunamis: The propagation of long waves onto a shelf.
Ph.D. thesis Rep. KH-R-38, Calif. Inst. Technol., Pasadena, California
(2003).
20. C. Goto. Nonlinear equations of long waves in the Lagrangian description.
Coast. Eng. Japan 22, 1–9 (1979).
21. C. Goto and N. Shuto. Run-up of tsunamis by linear and nonlinear theories.
In Coastal Engineering. Proc. of the seventeenth Coastal Engineering Conf.,
vol. 1 (Sydney, Australia, 1980), p. 695–70.
22. C. Goto and N. Shuto, Numerical simulation of tsunami propagations and
run-up. In Tsunamis–Their Science and Engineering, Eds. K. Lida and
T. Iwasaki (TERRAPUB, 1983), p. 439–451.
23. S. Grilli and I. A. Svendsen. Computation of nonlinear wave kinematics dur-
ing propagation and run-up on a slope. In Water wave kinematics (Kluwer
Academic Publishers, 1990), p. 378–412.
24. S. T. Grilli, I. A. Svendsen and R. Subramanya. Breaking criterion and
characteristics for solitary waves on slopes. J. Waterw., Port, Coast., Ocean
Engrg. (1997).
25. S. Guignard, R. Marcer, V. Rey, K. Kharif and Fraunié. Solitary wave break-
ing on sloping beaches; 2-D two phase flow numerical simulation by SL-VOF
method. Eur. J. Mech. B-Fluids 20, 57–74 (2001).
26. J. V. Hall and J. W. Watts. Laboratory investigation of the vertical rise
of solitary waves on impermeable slopes. Tech. Memo. 33, Beach Erosion
Board, U.S. Army Corps of Engrs. (1953).
August 8, 2008 11:12 WSPC/Trim Size: 9in x 6in for Review Volume 01˙Pedersen
38 G. Pedersen
September 1998.
URL: http://chinacat.coastal.udel.edu.
46. N. Kobayashi, A. K. Otta and I. Roy. Wave reflection and run-up on rough
slopes. J. Waterw., Port, Coast., Ocean Engrg. Div. ASCE. 113, 282–298
(1987).
47. Z. Kowalik and T. S. Murty. Numerical simulation of two-dimensional
tsunami run-up. Marine Geodesy 16, 87–100 (1993).
48. R. J. LeVeque and D. L. George. High-resolution finite volume methods for
the shallow water equations with bathymetry and dry states. In Advanced
Numerical Models for Simulating Tsunami Waves and Runup, eds. P. L.-
F. Liu, H. Yeh and C. Synolakis (World Scientific Publishing Co., 2008),
p. 43–73.
49. Y. Li and F. Raichlen. Solitary wave runup on plane slopes. J. Wa-
terw., Port, Coast., Ocean Engrg. 127(1), 33–44 (2001).
50. Y. Li and F. Raichlen. Non-breaking and breaking solitary wave run-up.
J. Fluid Mech. 456, 295–318 (2002).
51. Pengzhi Lin, Kuang-An Chang and Philip L.-F. Liu. Runup and rundown
of solitary waves on sloping beaches. J. Waterw., Port, Coast., Ocean En-
grg. 125(5), 247–255 (1999).
52. P. L.-F. Liu, Y.-S. Cho, M. J. Briggs, U. Kânoǧlu and C. E. Synolakis.
Runup of solitary waves on a circular island. J. Fluid Mech. 302, 259–285
(1995).
53. P. L.-F. Liu, C. E. Synolakis and H. Yeh. Report on the international work-
shop on long-wave run-up. J. Fluid Mech. 229, 675–688 (1991).
54. S. Longo, M. Petti and I. J. Losada. Turbulence in the swash and surf zones:
a review. Coastal Engineering 45, 129–147 (2002).
55. D. R. Lynch and W. G. Gray. Finite element simulation of flow in deforming
regions. J. Comp. Phys. 36, 135–153 (1980).
56. P. J. Lynett and P. L.-F. Liu. Coulwave model page (2004).
URL: http://ceprofs.tamu.edu/plynett/COULWAVE/default.htm.
57. P. J. Lynett, T.-R. Wu and P. L.-F. Liu. Modeling wave runup with depth-
integrated equations. Coast. Eng. 46, 89–107 (2002).
58. P. A. Madsen, H. B. Bingham and H. Liu. A new Boussinesq method for
fully nonlinear waves from shallow to deep water. J. Fluid Mech. 462, 1–30
(2002).
59. P. A. Madsen, R. Murray and O. R. Sörensen. A new form of the Boussinesq
equations with improved linear dispersion characteristics. Coast. Eng. 15,
371–388 (1991).
60. P. A. Madsen and H. A. Schäffer. Higher-order Boussinesq-type equations
for surface gravity waves: derivation and analysis. Phil. Trans. R. Soc. Lond.
A 356, 3123–3184 (1998).
61. P. A. Madsen and H. A. Schäffer. A review of Boussinesq-type equations for
surface gravity waves. Advances in Coastal and Ocean Engineering, vol. 5
(World Scientific Publishing Co., 1999), p. 1–95.
August 8, 2008 11:12 WSPC/Trim Size: 9in x 6in for Review Volume 01˙Pedersen
40 G. Pedersen
CHAPTER 2
1. Introduction
We will present a brief introduction to a class of high-resolution finite vol-
ume methods for hyperbolic problems and discuss the application of these
methods to long-wave runup problems using the shallow water equations. To
solve the benchmark problems for this workshop we have used such meth-
ods in one and two space dimensions that work robustly with bathymetry
(bottom topography) and dry states, and that automatically handle the
moving interface between water and land. Some results on the benchmark
problems are presented in Secs. 6 and 8, and more results, along with some
animations, may be found at the website15 .
We use a mathematical framework known as the wave-propagation al-
gorithm that has been implemented in the software package clawpack
(Conservation Laws Package) in 1, 2, and 3 space dimensions, and which
43
July 24, 2008 10:23 WSPC/Trim Size: 9in x 6in for Review Volume 02˙leveque
and hence
qt + f (q)x = 0, (3)
which is the PDE form of the conservation law (with subscripts denoting
partial derivatives). This equation is called hyperbolic if the Jacobian matrix
f 0 (q0 ) ∈ lRm×m is diagonalizable and has real eigenvalues for any physically
relevant state q0 . Hyperbolic problems typically model wave propagation
and the eigenvalues correspond to the propagation velocities if we linearize
about the state q0 . For the shallow water equations given by (2),
0 0 1
f (q) = , (4)
−u2 + gh 2u
with eigenvalues
p p
λ1 = u − gh, λ2 = u + gh, (5)
and corresponding eigenvectors
1 1 1 1
r1 = √ = , r2 = √ = . (6)
u − gh λ1 u+ gh λ2
A finite volume method in conservation form updates the cell average
Qni of the solution over the grid cell using an expression
∆t n
Qn+1 = Qni − [F n
− Fi−1/2 ], (7)
i
∆x i+1/2
where
Z xi+1/2
1
Qni ≈ q(x, tn ) dx,
∆x xi−1/2
Z tn+1
(8)
n 1
Fi−1/2 ≈ f (q(xi−1/2 , t)) dt
∆t tn
are the numerical approximations to the cell average and interface flux,
respectively. The update (7) comes directly from integrating (1) in time
from tn to tn+1 and dividing by ∆x. Equation (7) can also be viewed as
n
a direct discretization of the PDE (3), but viewing the value Fi−1/2 as
an approximation to the interface flux is key in developing high-resolution
methods for nonlinear problems.
July 24, 2008 10:23 WSPC/Trim Size: 9in x 6in for Review Volume 02˙leveque
A = RΛR−1 ,
We use W p to denote the pth wave in this Riemann solution. The vector of
coefficients αpi−1/2 is given by αi−1/2 = R−1 ∆Qi−1/2 . Godunov’s method
in the linear case is then defined by setting
| |
Fi−1/2 = f (Q∨ ∨
i−1/2 ) = AQi−1/2 ,
|
where Q∨ i−1/2 denotes the value at the interface xi−1/2 in the Riemann
solution,
| X p
Q∨i−1/2 = Qi−1 + Wi−1/2 .
p:λp <0
July 24, 2008 10:23 WSPC/Trim Size: 9in x 6in for Review Volume 02˙leveque
Multiplying by A gives
m
X
Fi−1/2 = AQi−1 + αpi−1/2 (λp )− rp
p=1 (10)
= AQi−1 + A− ∆Qi−1/2 ,
where
and obtain
m
X
Fi−1/2 = AQi + αpi−1/2 (λp )+ rp
p=1 (12)
= AQi + A+ ∆Qi−1/2 ,
where
For a linear problem the resulting method is simply the first-order upwind
method, extended from the scalar advection equation to a general system
by diagonalizing the system and applying the upwind method to each char-
acteristic component in the appropriate direction based on the propagation
velocity.
For nonlinear problems, such as the shallow water equations, the exact
solution to the Riemann problem is harder to calculate but can still be
worked out (see, e.g. LeVeque14 and Toro19 ) and the resulting interface flux
used for Fi−1/2 . In practice, however, it is usually more efficient to use some
approximate Riemann solver to obtain Fi−1/2 . One popular choice is to use
a “Roe solver” following the work of Roe17 for gas dynamics, in which the
data Qi−1 , Qi is used to define a “Roe-averaged” Jacobian matrix Ai−1/2
by a suitable combination of f 0 (Qi−1 ) and f 0 (Qi ). The numerical flux is
then determined by solving the Riemann problem for the linear problem
qt + Ai−1/2 qx = 0. The Roe average is chosen to have the property that
This leads to nice properties in the approximate solution. The Roe matrix
for the shallow water equations is easily computed (see, e.g. LeVeque14 ),
and is simply the Jacobian matrix (4) evaluated at the Roe-averaged state
p √
hi + hi−1 ui−1 ghi−1 + ui ghi
ĥi−1/2 = , ûi−1/2 = p √ . (15)
2 ghi−1 + ghi
dimension of the system, but could be different. The Riemann solver must
also return the fluctuations A− ∆Qi−1/2 and A+ ∆Qi−1/2 , two vectors that
are used to update the solution according to
∆t +
Qn+1
i = Qi − (A ∆Qi−1/2 + A− ∆Qi+1/2 ). (18)
∆x
These fluctuations should have the property that
Then (19) is satisfied (note the shift in index) and (18) reduces to the
flux-differencing update formula (7). The form (18) is used in clawpack
and the general formulation of the wave-propagation algorithms because it
is more flexible and allows the extension of these methods to hyperbolic
problems that are not in conservation form, in which case there is no “flux
function” (see LeVeque14 ).
July 24, 2008 10:23 WSPC/Trim Size: 9in x 6in for Review Volume 02˙leveque
For a linear problem or a nonlinear problem when the Roe solver is used,
this agrees with the definition (20).
The first-order method (18) only uses the fluctuations returned from
the Riemann solver, and does not make explicit use of the waves W p or
speeds sp . These quantities are used in the high-resolution correction terms
discussed in the next section.
oscillations can be avoided by using the form (23) and applying appropriate
limiters to the correction terms. To this end we rewrite (24) as
m
1X ∆t p fp
F̃i−1/2 = I− |λ | |λp |W i−1/2 , (26)
2 p=1 ∆x
p
where Wi−1/2 = αpi−1/2 rp are the waves obtained from the Riemann solu-
tion, as in (9), and Wfp
i−1/2 represents a “limited” version of the wave. Each
wave is limited by comparing it to the wave in the same family arising from
the Riemann problem at the neighboring interface in the upwind direction,
i.e. we set
fp
W p p
i−1/2 = limiter(Wi−1/2 , WI−1/2 ), (27)
where
i−1 if λp > 0
I=
i+1 if λp < 0.
p p
If WI−1/2 and Wi−1/2 are “comparable” in some sense then this compo-
nent of the solution is presumed to be smoothly varying. In this case the
corresponding term in (26) can be expected to give a useful correction that
will help improve accuracy and the limiter should return W fp p
i−1/2 ≈ Wi−1/2 .
p p
However, if WI−1/2 and Wi−1/2 quite different then this component is not
smoothly varying and attempting to add an additional term from the Tay-
lor series may make things worse rather than better. In this case the limiter
should modify the wave, typically be reducing its magnitude. There is an
extensive theory of limiters that will not be further discussed here.
The method (23) with correction fluxes (26) is easily extended to non-
linear problems. Recall that the (approximate) Riemann solver returns fluc-
p
tuations A± ∆Qi−1/2 , waves Wi−1/2 , and speeds spi−1/2 . The only change in
the formulas required in order to apply (23) to a general nonlinear problem
is to replace λp in (26) by the local speed spi−1/2 , obtaining
Mw
1X ∆t p
F̃i−1/2 = I− |s | |spi−1/2 |W
fp
i−1/2 . (28)
2 p=1 ∆x i−1/2
p
to handle variable bathymetry. Recall that the waves Wi−1/2 correspond to
a splitting of the jump in Q at the interface xi−1/2 ,
Mw
X p
Qi − Qi−1 = Wi−1/2 .
p=1
If the matrix Ai−1/2 satisfies Roe’s condition (14), then we simply have
p
Zi−1/2 p
= spi−1/2 Wi−1/2 . For other approximate Riemann solvers it is nec-
essary to determine an appropriate splitting of f (Q) based on the splitting
of Q
The vectors Z p are called f-waves because they carry jumps in f rather
p
than jumps in q. Since Zi−1/2 = sgn(spi−1/2 )|spi−1/2 |Wi−1/2
p
for linearized
Riemann solvers, the natural generalization of the correction flux (28) for
the f-wave formulation is
Mw
1X ∆t
F̃i−1/2 = I− sgn(spi−1/2 ) Zei−1/2
p
, (29)
2 p=1 ∆x
p p
where Zei−1/2 is a limited version of Zi−1/2 calculated using the same limiter
p
as previously applied to W .
One advantage of the f-wave approach is that any linearly independent
p
set of vectors ri−1 can be used to define the splitting of ∆f into waves
p
Zi−1/2 and the method remains conservative. Of course a reasonable choice
is required in order to maintain consistency with the differential equation,
but for example the eigenvectors of any reasonable approximate Jacobian
matrix based on Qi−1 and Qi could be used without needing to impose
the Roe condition (14). For the shallow water equations this suggests using
vectors
1 1
r1 = 1 , r2 = 2 , (30)
s s
where s1 and s2 are some approximations to the wave speeds of the two
waves in the Riemann solution. Taking s1 and s2 to be the Roe speeds
July 24, 2008 10:23 WSPC/Trim Size: 9in x 6in for Review Volume 02˙leveque
(16) recovers the Roe solver, but in some cases this is not a good choice.
In particular the Roe solver can fail when dry states are expected in the
solution. In Sec. 4 we present a different choice of speeds that can be used
much more robustly.
4. Dry States
It is well known that if the Roe solver is used to solve a Riemann problem
in which ui−1 < ui with sufficient difference in velocities, then the approxi-
mate Riemann solution will have a negative depth in the intermediate state
(see Fig. 15.3 in LeVeque14 for an illustration of this). This nonphysical be-
havior often causes the computation to crash. In reality a dry state should
July 24, 2008 10:23 WSPC/Trim Size: 9in x 6in for Review Volume 02˙leveque
form if the velocity difference is sufficiently large, although the Roe solver
can fail even for smaller velocity differences (see Einfeldt et al.6 for some
discussion of related issues for the Euler equations). Similar problems arise
when solving a Riemann problem with one state initially dry, hi−1 = 0 or
hi = 0, as happens at some cell interfaces for any problem involving wave
motion at the shore.
This difficulty can be avoided by using the f-wave approach discussed
in Secs. 2 and 3 using eigenvectors r 1 and r2 from (30) with a better choice
of speeds than the Roe speeds. In most cases we use the “Einfeldt speeds”
p p
s1E = min(ui−1 − ghi−1 , ŝ1 ), s2E = max(ui + ghi , ŝ2 ), (37)
obtained by comparing the Roe speeds with the characteristic speeds λ1i−1
and λ1i (the eigenvalues of the Jacobian matrices in states Qi−1 and Qi ).
This choice is adapted from the suggestion of Einfeldt5 that speeds cor-
responding to these be used in the HLL method for gas dynamics. The
HLL approximate Riemann solver (after Harten et al.9 ) simply uses two
waves to approximate the Riemann solution (regardless of the dimension
m of the system) with speeds s1 and s2 approximating the minimum and
maximum wave speeds arising in the system. The HLLE method, using the
Einfeldt speeds, chooses these speeds by comparing the Roe speed, a rea-
sonable choice if the wave is a shock, and the extreme characteristic speed,
which may be faster if the wave is instead a spreading rarefaction wave.
Since m = 2 in the one-dimensional shallow water equations, the method
we use is closely related to the HLLE method, though not the same and
will be modified further to handle source terms and dry states below. (See
LeVeque and Pelanti16 for some more discussion of the relation between the
HLL and Roe solvers and the f-wave approach.)
Using the f-wave approach with the choice of speeds (37) and corre-
sponding eigenvectors (30) nearly always maintains non-negative depth. In
fact, it can be shown that the total mass in the intermediate Riemann
solution is always positive given these speeds and eigenvectors. However,
as explained below, it is sometimes necessary to further modify the wave
speeds to maintain non-negativity, at least in the subcritical case when
s1E < 0 < s2E . In the supercritical case when both speeds have the same
sign, no modification is needed.
We first consider the case of preserving non-negativity in a cell that has
positive depth initially, hi−1 > 0 or hi > 0, and later consider the case of
preserving non-negativity in a cell that is already initially dry. Even in the
case where both cells have a positive depth initially, hi−1 > 0 and hi > 0,
July 24, 2008 10:23 WSPC/Trim Size: 9in x 6in for Review Volume 02˙leveque
a negative depth can be generated if s1E < 0 < s2E when the choice (37)
is used. Although the total mass in the intermediate Riemann solution is
positive, it may happen that the approximate Riemann solution leads to
the mass going negative on one side of the interface. In this case it can
be shown that the mass must be increasing on the other side by at least
the same amount, and so negativity can be avoided by a transfer of mass
that can be accomplished by increasing the speed on the side losing mass.
Working out the formula to increase this speed just to the point where the
negative state reaches h = 0, we find that in general the following speeds
can always be used:
p
s1 = min s1E , s2E /2 − (s2E /2)2 + max(0, s2E ∆1 − ∆2 )/hi−1 , (38)
p
s2 = max s2E , s1E /2 + (|s1E |/2)2 + max(0, ∆2 − s1E ∆1 )/hi . (39)
That is, given hi−1 > 0 initially, (38) will maintain that hi−1 ≥ 0, and
given hi > 0 initially (39) will maintain that hi ≥ 0. Here ∆1 and ∆2 are
the components of the modified flux difference
which takes into account the bathymetry. Each speed (38) and (39) always
corresponds to the Einfeldt speed unless a negative state would arise on
that side, so for a given Riemann problem, at most one of s1 and s2 is
different from the Einfeldt speed, and only when necessary to maintain
non-negativity.
Maintaining the non-negativity in a cell with a depth that is initially 0
is handled somewhat differently. If Bi−1 = Bi , and only one of hi−1 or hi
is positive, then the true solution to this Riemann problem consists only of
√
a rarefaction wave with the leading
p edge propagating at speed ui − 2 ghi
if hi−1 = 0 or speed ui−1 + 2 ghi−1 if hi = 0. We use these speeds as s1
or s2 if hi−1 = 0 or hi = 0 respectively. As stated above, using the speed
(38) will preserve hi−1 ≥ 0 if it is initially positive, and (39) will preserve
hi ≥ 0 if it is initially positive. Using the speed of the leading edge of a
rarefaction wave for the complimentary speed however, will not necessarily
prevent the dry state from becoming negative. It can be shown however
that negativity is only
p possible if the true rarefaction wave is transonic
√
(s1 < 0 < ui−1 + 2 ghi−1 if hi = 0 or ui − 2 ghi < 0 < s2 if hi−1 = 0).
Transonic rarefactions are a more general problem, and are discussed in the
following section.
July 24, 2008 10:23 WSPC/Trim Size: 9in x 6in for Review Volume 02˙leveque
5. Entropy Conditions
Integral conservation laws can be satisfied by discontinuous weak solutions
subject to the Rankine-Hugoniot jump conditions. This results in possible
non-uniqueness of weak solutions — an initial value problem might be sat-
isfied by both a smooth solution and a discontinuous one. Determining the
physically relevant solution requires additional admissibility conditions, of-
ten taking the form of an “entropy” function that is conserved except across
a discontinuity. The name arises from the Euler equations of gas dynam-
ics, where the entropy must increase when gas passes through a shock. For
the shallow water equations, the “entropy” function is actually mechanical
energy, which must decrease when passing through a discontinuity.
Since the integral conservation laws alone do not guarantee a unique
solution to an initial value problem, a numerical method based on the
conservation laws alone might converge to an entropy violating weak so-
lution. An “entropy fix” is therefore often needed. Determining such a fix
for Godunov-type methods requires a careful look at the particular Rie-
mann solver being used. If an approximate solver is used, true solutions to
Riemann problems which consist of m waves — any combination of rar-
efactions and shock waves, are replaced by m jump discontinuities locally
at each grid cell. These discontinuities might approximate a physically rel-
evant shock wave, or perhaps a smooth rarefaction. In the latter case, the
jump discontinuity still approximates the conservation law, however it more
closely resembles the entropy violating shock than the physically relevant
rarefaction.
With the wave propagation methods, the waves arising from a particu-
lar Riemann problem at a grid cell interface are averaged onto the adjacent
cells. Therefore, the local discrepancy between an entropy violating dis-
continuity and a smooth rarefaction may have no effect on the numerical
solution, if both produce the same average within a grid cell. This will be the
case if the wave structure of the rarefaction remains entirely within a grid
cell. The exceptional case is a transonic rarefaction — a rarefaction where
one of the eigenvalues passes through zero. This type of rarefaction has a
wave structure that should overlap a cell interface, yet it is approximated
by a jump discontinuity moving either to the left or the right. This does
affect the numerical solution, and can cause a method to converge globally
to an entropy violating weak solution. The fix is to determine when the
correct entropy solution to a Riemann problem corresponds to a transonic
July 24, 2008 10:23 WSPC/Trim Size: 9in x 6in for Review Volume 02˙leveque
rarefaction, and then split the entropy violating single wave, apportioning
it to the adjacent grid cells appropriately.
An alternative and numerically more useful formulation of the entropy
condition for the shallow water equations is that a physically correct shock
in the pth (p = 1, 2) characteristic family must have the p-characteristics
impinging on it. That is,
λp (ql ) > s > λp (qr ), (41)
p th
where s is the shock speed and λ is the p eigenvalue evaluated at ql
and qr — the states directly to the left and right of the shock respectively.
Therefore, if the entropy solution to a Riemann problem contains a shock
connecting the left state Qi−1 to the middle state, denoted Qm , then
λ1 (Qi−1 ) > λ1 (Qm ). (42)
Similarly if a shock connects the right state Qi to the middle state Qm ,
then
λ2 (Qm ) > λ2 (Qi ). (43)
If (42) or (43) is violated, then in fact a rarefaction connects the corre-
sponding states in the entropy solution. As explained above, an entropy
violating Riemann solution will not affect the numerical solution, except in
the case of a transonic rarefaction. Therefore the only cases in which an
entropy fix is required are when
λ1 (Qi−1 ) < 0 < λ1 (Qm ) (44)
or
λ2 (Qm ) < 0 < λ2 (Qi ), (45)
which indicate a transonic rarefaction in the first or second families
respectively.
It is easy to check λ1 (Qi−1 ) and λ2 (Qi ). However, with the f-wave ap-
proach, Qm is never explicitly computed so we cannot simply evaluate
λ1 (Qm ) and λ2 (Qm ) directly. In fact the f-wave approach is not equiv-
alent to using a single value for Qm , but two middle values, one to the
left and one to the right of the cell interface. The approach we’ve taken
is to instead compare the Roe speeds, ŝ1i− 1 and ŝ2i− 1 , with the right and
2 2
left speeds, λ1 (Qi−1 ) and λ2 (Qi ). The motivation for this approach is that
the Roe speeds can serve as an estimate for the shock speeds, allowing
us to estimate whether (41) is satisfied. For instance, to detect the pres-
√
ence of a transonic rarefaction in the second family, ui + ghi is compared
July 24, 2008 10:23 WSPC/Trim Size: 9in x 6in for Review Volume 02˙leveque
√
to ŝ2i− 1 . If ui + ghi > ŝ2i− 1 then most likely the true Riemann solu-
2 2
tion has a rarefaction in the second family. Furthermore ŝ2i− 1 serves as
2
an estimate for the characteristic speed at the center of the rarefaction
fan, ŝ2i−1/2 ≈ 21 (λ2 (Qm ) + λ2 (Qi )) and hence λ2 (Qm ) ≈ 2s2i−1/2 − λ2 (Qi ).
Therefore if
p p
2(ŝ2i− 1 ) − (ui + ghi ) < 0 < ui + ghi , (46)
2
the sign of its first component. The gross mass-flux out of the ith grid cell,
due to these correction fluxes, is therefore
h i
1 1
Mi = max(0, F̃i+ 1 ) − min(0, F̃
i− 1 ) . (49)
2 2
If ∆tMi is larger than the mass present after the Godunov update,
∆x(QG 1
i ) , the correction fluxes could potentially create a negative depth
in this cell. This is prevented by re-limiting the correction fluxes based on
which cell they take mass away from
F̃i− 12 → ϕi− 12 F̃i− 21 , (50)
where
(
min(1, ∆x(QG 1
i ) /∆tMi ) if 1
F̃i− 1 < 0
ϕi− 12 = 2
(51)
min(1, ∆x(QG 1
i−1 ) /∆tMi−1 ) if
1
F̃i− 1 > 0.
2
20
0
m
−20
−40
PSfrag replacements
−60
−80
200 300 400 500 600 700 800
m
u(x, t) at time t = 160 seconds
15 computed
exact
10
5
u (m/s)
−5
−10
PSfrag replacements
−15
m/s
200 300 400 500 600 700 800
m
Fig. 1. Top: Water surface elevation at t = 160s, shown near the beach. Bottom: Veloc-
ity field at t = 160s in the same region. Both computations were done on a 1000-point
grid.
1000 points so that the numerical results are still distinguishable from the
analytical solution. The surface elevation and the velocity field are shown
at t = 160s in Fig. 1. Note that the figures show only a small portion of
the domain near the beach. Figure 2 shows the motion of the shoreline
for the same computation. For additional results at other times and grid
resolutions, see the website15 .
July 24, 2008 10:23 WSPC/Trim Size: 9in x 6in for Review Volume 02˙leveque
100 0
50
meters
m/s
−5
0
−50 −10
−100
−15
−150
PSfrag replacements
η
Fig. 4. The computed solution during the primary runup of the wave. The primary
runup occurred in the first 30 seconds. Contour lines show the topography that was
initially above the water surface, including an island. Gray scale indicates elevation
above the original water surface.
July 24, 2008 10:23 WSPC/Trim Size: 9in x 6in for Review Volume 02˙leveque
Fig. 5. Comparison of the numerical solution (right) with snapshots from the overhead
movie of the laboratory experiment (left). Frames 11, 26, 41, 56 from the movie are shown
and computed results are shown at corresponding 30 second intervals. (The movie shows
30 frames per second.)
July 24, 2008 10:23 WSPC/Trim Size: 9in x 6in for Review Volume 02˙leveque
0.03 0.04
0.02 0.03
meters
meters
0.01 0.02
0 0.01
−0.01 0
−0.02 −0.01
0 10 20 30 40 50 0 10 20 30 40 50
sec. sec.
0.03 0.025
0.02
meters
meters
0.02 0.015
0.01
0.01
0.005
0
0
−0.005
−0.01 −0.01
0 10 20 30 40 50 0 50 100 150 200
sec. sec.
Fig. 6. Comparison of the surface elevation, with the laboratory measurements at the
three wave gages (Channels 5, 7, 9). The numerical solution is comparable during the
primary runup in the first 50 seconds. The bottom right figure shows the comparison for
the duration of the laboratory measurements, at one of the gages (Channel 5). Sustained
oscillations in the laboratory measurements are not seen in the numerical solution, as
discussed in the text. The other gages showed similar patterns at later times.
and was followed by other waves in the wave tank for which no data was
provided.
with load balancing for adaptive refinement on parallel computers, and sup-
port for implicit algorithms that may be needed if dissipative or dispersive
terms are added to the equations.
A rectangular grid is refined by covering a portion of the domain by one
or more rectangular patches of grids that are finer by a factor of k, some
integer. This process can be repeated recursively, with each grid level fur-
ther refined by grid patches at a higher level or refinement. The maximum
number of levels is specified along with the refinement factor at each level.
The grid patches are chosen by flagging grid cells at each level that “need
refining” (see below) and then clustering these cells into a set of rectangular
patches that cover these cells and a limited number of other cells. This is
done by solving an optimization problem (using the algorithm of Berger
and Rigoutsos3 ) that takes into account the trade-off between refining too
many cells unnecessarily and creating too many grid patches, since there is
some overhead associated with applying the algorithm on each patch. The
boundary data (ghost cell values) on a patch must be generated by space-
time interpolation from data on the coarser grid. A time step of length
∆t is first taken on the coarse grid, boundary data is then generated from
the fine grid patches, and k 0 time steps of length ∆t/k 0 are then taken on
the fine grids to reach the same time. For most AMR applications on hy-
perbolic problems we take k 0 = k, refining in time by the same factor as
in space in order to maintain a comparable Courant number at all levels.
However, for tsunami propagation and runup applications where we only
have the finest grids near the shore, it may be desirable to refine in time
√
by a smaller factor k 0 < k. Recall that the wave speeds are roughly gh,
which differs by more than an order of magnitude between the deep ocean
and the coastal regions. The smaller wave speed near shore allows a larger
time step.
After updating to the same time on the fine grids, the coarse grid so-
lution on any grid cell covered by the finer grid is then replaced by the
average of the fine grid values in this cell in order to transfer the more ac-
curate solution to the coarse grid. Additional modifications to the values are
needed near the patch edge to maintain global conservation as described in
Berger and LeVeque2 . This is done recursively at each level. While the inner
workings are somewhat complicated, the general formulation in amrclaw
allows extension of most clawpack computations directly to adaptively
refined grids. The computational time required for the overhead associated
with multiple grids is often negligible compared to the savings achieved by
concentrating fine grids only where needed.
July 24, 2008 10:23 WSPC/Trim Size: 9in x 6in for Review Volume 02˙leveque
We can flag the cells that “need refinement” however we wish. For the
calculation presented below, we have allowed some refinement everywhere
based on a measure of the variation in the solution, so that the propagating
wave is well resolved. In addition, we allow additional levels of refinement
near the shore in particular regions of interest, though only once the wave
is approaching. Regridding is performed every few time steps to modify the
region of refinement as the wave propagates.
In the course of this work, we ran into several difficulties at the interfaces
between grids that have not been observed in other applications of the
AMR codes. These arise from the representation of the bathymetry and
shore on a Cartesian grid. One problem, for example, is that a coarse grid
cell that is dry (if the bathymetry value in this cell is above sea level) may
be refined into some cells that are above sea level and others below. Even
though h = 0 on the coarse cell we cannot initialize h to zero on all the
fine cells without generating nonphysical wave motion. The fine cells below
sea level must be initialized with a positive h to maintain the constant
sea level. This problem and related difficulties could only be solved by
some substantial reworking of the amrclaw code. The result is a special-
purpose program that incorporates these algorithmic modifications and can
now be applied to many tsunami scenarios. It is currently being tested by
comparing predictions with measurements made at various places around
the Indian Ocean in the wake of the 26 December 2004 Sumatra earthquake.
Some preliminary results are shown in Fig. 7, and additional results and
movies are linked from the webpage15 and will be reported more fully in
future publications.
The top two frames of Fig. 7 show the Bay of Bengal at two early
times. A coarse grid is used where nothing is yet happening and the grid
cells are shown on this “Level 1 grid”, which has a mesh width of one
degree (approximately 111 km). The rectangular region where no grid lines
are shown is a Level 2 grid with mesh width 8 times smaller, about 14
km. Red represents water elevation above sea level, dark blue is below the
undisturbed surface. Figure 7(c) shows a zoomed view of the southern tip of
India and Sri Lanka at a later time. The original Level 1 grid is still visible
along the left-most edge, but the rest of the region shown has been refined
by a Level 2 grid. Due north from Sri Lanka, along the coast of India, there
is a region near Chennai where two additional levels of refinement have
been allowed at this stage in the computation. The grid lines on Level 3
are not shown; the mesh width on this level is about 1.7 km, a factor of 8
finer than Level 2. A Level 4 grid is also visible in the center of this region
and appears as a small black rectangle.
July 24, 2008 10:23 WSPC/Trim Size: 9in x 6in for Review Volume 02˙leveque
(a) Time 01:07:10 (550 seconds) (b) Time 01:34:40 (2200 seconds)
(c) Time 03:13:12.5 (8112.5 seconds) (d) Time 03:15:30 (8250 seconds)
(e) Time 03:24:40 (8800 seconds) (f) Time 03:35:30 (9450 seconds)
Fig. 7. Propagation of the 26 December 2004 tsunami across the Indian Ocean, using
adaptive mesh refinement with refinement by a factor of 4096 from the coarsest grid
shown in (a)–(b) to the finest grid shown in (d)–(f). The latter figures show zoomed
views of the region near the harbors of Chennai, India. Times are GMT (and seconds
since initial rupture at 0:58 GMT). See the text for more details.
July 24, 2008 10:23 WSPC/Trim Size: 9in x 6in for Review Volume 02˙leveque
Figure 7(d) shows a further zoomed view of the coast near Chennai. In
this figure the grid lines show the Level 3 grid. The Level 4 grid is refined
by a factor of 64 relative to the Level 3 grid, so the mesh width is about
27 meters. Grid lines on this finest level are not shown. The rectangle in
this figure shows a region that is expanded in Figs. 7(e) and 7(f) to show
the two harbors of Chennai. The fine-scale bathymetry used in this com-
putation was obtained by digitizing navigational charts. In this simulation
the commercial harbor to the south is inundated, with the tsunami over-
topping the surrounding sea wall, while the fishing harbor to the north is
largely untouched. This appears to agree with what was actually observed,
although we are still investigating this. Moreover, we do not yet have suffi-
ciently accurate data on the height of the sea walls enclosing each harbor.
One data point known accurately from tide gage data is the arrival time
for the initial wave in Chennai, which was at 9:05 local time (3:35 GMT).
This is well matched by our simulation: this is essentially the time shown
in Fig. 7(f).
A more careful study of this region will be performed in the future and
presented elsewhere, including comparisons with runup data collected by
Harry Yeh as part of the Tsunami Survey Team that visited this area in
February, 2005. See the webpage15 for more recent results and movies of
the simulations. In the future we plan to also compare with data collected
by other teams at various other points around the Indian Ocean. We will
also make our computer code available to the community for others to use,
in a form that should allow the local bathymetry from other regions to be
easily incorporated. See the webpage for more details.
Adaptive mesh refinement is crucial for this simulation. The calculation
shown here was run on a single-processor 3.2 GHz PC under linux. Figure
7(c) was obtained after about 15 minutes of running time, Fig. 7(d) about
2.5 hours later, indicating that most of the grid cells are concentrated on
the Level 4 grid, which is introduced only when the wave approaches the
shore in this one region. Were it possible to use the finest grid over the
entire domain, the result of Fig. 7(c) would require more than 4000 years
of computing on the same machine.
Acknowledgments
This work was supported in part by NSF grants DMS-0106511 and CMS-
0245206, and by DOE grant DE-FC02-01ER25474. The authors would like
to thank Marsha Berger and Donna Calhoun for assistance with the adap-
tive refinement aspect of this work.
References
1. D. Bale, R. J. LeVeque, S. Mitran, and J. A. Rossmanith, A wave-propagation
method for conservation laws and balance laws with spatially varying flux
functions, SIAM J. Sci. Comput., 24 (2002), pp. 955–978.
2. M. J. Berger and R. J. LeVeque, Adaptive mesh refinement using wave-
propagation algorithms for hyperbolic systems, SIAM J. Numer. Anal., 35
(1998), pp. 2298–2316.
3. M. J. Berger and I. Rigoutsos, An algorithm for point clustering and grid
generation, IEEE Trans. Sys. Man & Cyber., 21 (1991), pp. 1278–1286.
4. P. Colella et al., chombo software, http://seesar.lbl.gov/anag/chombo/,
2005.
5. B. Einfeldt, On Godunov-type methods for gas dynamics, SIAM J. Num.
Anal., 25 (1988), pp. 294–318.
6. B. Einfeldt, C. D. Munz, P. L. Roe, and B. Sjogreen, On Godunov type
methods near low densities, J. Comput. Phys., 92 (1991), pp. 273–295.
7. L. Gosse, A well-balanced flux-vector splitting scheme designed for hyperbolic
systems of conservation laws with source terms, Comput. Math. Appl., 39
(2000), pp. 135–159.
8. J. M. Greenberg, A. Y. LeRoux, R. Baraille, and A. Noussair, Analysis
and approximation of conservation laws with source terms, SIAM J. Numer.
Anal., 34 (1997), pp. 1980–2007.
July 24, 2008 10:23 WSPC/Trim Size: 9in x 6in for Review Volume 02˙leveque
1. Introduction
While the mechanisms that generate the tsunami flows are generally
understood, the ability to predict the flows they produce at coastlines still
represents a formidable challenge due to the complexities of coastline
formations and the presence of numerous coastal structures that interact
and alter the flow. Indeed, overland flows kill more people alone than
any other aspect of tsunamis but are still not well understood or
accurately predictable. Numerous experimental investigations have
focused on different instances of tsunami runup such as circular islands
(Briggs et al. 1994). However, the accurate prediction of runup onto
real coastlines with structures and human habitation still remains
incomplete.
75
76 B. D. Rogers and R. A. Dalrymple
2. Numerical Discretization
where Vj is the volume of the jth particle located at xj with scalar quantity
( )
fj, and W x − x j is the weighting function referred to as the smoothing
kernel. Prior to simulation, the kernel is specified by an analytical
expression making its evaluation in Eq. (2.1) straightforward. In all
results presented, a quadratic kernel has been used for computational
SPH Modeling of Tsunami Waves 79
speed and to avoid the problems associated with inflexion points in the
derivatives of spline kernels (Liu et al. 2003):
C
W (x − x j ) = N
( 1
4 )
q2 − q +1 0≤q≤2
(2.2)
0 otherwise ,
where q = r h , and r = x − x j . In 1-D, C N = C 1 = 3 4 h , in 2-D
C N = C 2 = 3 (2 π h 2 ) and in 3-D C N = C 3 = 15 (16 πh ) . The kernel has a
3
d ui P Pj
= −∑ m j i2 + 2 + Π ij ∇ iWij + g , (2.6b)
ρ
dt j i ρj
where mj is the mass of the jth particle, ∇iWij = ∇iW (x − x j ) , the first
subscript i referring to the derivative of W with respect to the coordinates
of particle i, and Πij is an artificial viscosity term. The artificial viscosity
term (see Monaghan 1992) is used to account for the viscous terms in the
governing equations. For free-surface hydrodynamics, the artificial
viscosity term also prevents the simulation from becoming unstable and
is given by (Monaghan 1992)
αµ c ij ij
Π ij = − , (2.7)
ρ ij
where α is an empirical coefficient (usually taken as 0.01 – 0.1),
( ) ( )
cij = 12 ci + c j , ρ ij = 12 ρ i + ρ j and
hu ij ⋅ x ij
µ ij = . (2.8)
rij2 + η 2
Here, u ij = u i − u j , rij = x i − x j and η 2 = 0.01h2 to prevent
singularities.
Within our SPH scheme we discretize the SPS stresses according to the
symmetric formulation given by Lo and Shao (2002)
1 τ i τj
∇ ⋅ τ = ∑ m j 2 + 2 ⋅ ∇ iWij .
(2.15)
ρ ρ ρ j
i j i
SPH Modeling of Tsunami Waves 83
the local water depth and velocity of the water particle normal to the
boundary
ε ( z, u ⊥ ) = ε ( z ) + ε (u ⊥ ) , (2.20)
where
0.02 z≥0
ε ( z ) = z ho + 0.02 0 > z ≥ − ho (2.21)
1 z ho > 1 ,
and
0 u⊥ > 0
ε (u ⊥ ) = 20u ⊥ co 20u ⊥ < co (2.22)
1 20u ⊥ > co .
In Eqs. (2.20)–(2.22), z is the elevation above the local still-water level
( )
ho, u ⊥ = u WaterParti cle − u BoundaryParticle ⋅ n and c o = Bγ ρ o .
This technique uses conventional SPH particles that remain part of the
prescribed boundary. While the position and velocity of these particles is
known a priori, the density of these particles is predicted by Eqs. (2.6a)
or (2.10a). This density is then inserted into the equation of state (2.5)
to give the pressure that a boundary particle exerts on a fluid particle
when evaluating the momentum equations (2.6a) or (2.10a). This type of
boundary condition has proved robust for three-dimensional SPH
simulations (Gómez-Gesteira and Dalrymple 2004). However, when
this boundary condition was used for three-dimensional simulations in
Sec. 3.2, it was found to have too much wall friction which became
critical near the shoreline.
our present model. This highlights one of the current limitations of SPH
in that its application to huge domains remains inappropriate and beyond
practical use. In all cases, the governing equations were advanced using
a second-order predictor-corrector time integration scheme.
−2
−4
y (m)
−6
−8
−10
−12
0 20 40 60 80 100 120
x (m)
Fig. 1. Test 3 — Tsunami generation by a two-dimensional landslide.
SPH Modeling of Tsunami Waves 87
tND = 0.5
2
SPH
1.5 Analytical
free-surface (m)
0.5
0
0 20 40 60 80 100 120
-0.5
-1
x (m)
tND = 1.0
2
SPH
1.5 Analytical
free-surface (m)
0.5
0
0 20 40 60 80 100 120
-0.5
-1
x (m)
tND = 2.5
2
SPH
1.5 Analytical
free-surface (m)
0.5
0
0 20 40 60 80 100 120
-0.5
-1
x (m)
tND = 4.5
2
SPH
1.5 Analytical
free-surface (m)
0.5
0
0 20 40 60 80 100 120
-0.5
-1
x (m)
water has been drawn down in generating the offshore propagating wave.
This is confirmed in Fig. 4 which displays a comparison of the free-
surface data at a wave gage located 1.83m offshore from the still-water
line. The 2-D SPH-LES scheme vastly overpredicts both the initial wave
and the following drawdown of the free-surface. This simulation was
performed using 8600 particles where ∆x = 0.09m taking 3 hours of cpu
time. A convergence study showed no change in the predicted numerical
behavior allowing us to conclude that the cause of such a large
discrepancy is that important three-dimensional effects of the drawdown
and violent splash collisions are neglected.
1
frame= 65 t= 2.6s
0
−1
y (m)
−2
−3
−4
−5
−1 0 1 2 3 4 5 6 7 8 9 x (m)
Fig. 3. Test 4 — Tsunami generation by a three-dimensional landslide, run 30 (initially
partially submerged): 2-D SPH-SPS simulation snapshot.
0.2
0.1
0
free-surface (m)
-0.2
-0.5
tim e (s)
When the wedge is placed just under the water surface so that it is only
just completely submerged, the simulation exhibits the same behavior as
for run 30 above. Figure 5 shows the comparison of the water surface at
the same wave gage. The same large initial wave and drawdown are
visible. A similar convergence study again confirmed the lack of three
dimensionality.
0.2
0.1
0
free-surface (m)
-0.2
-0.3
-0.4
experimental data
2-D SPH
-0.5
time (s)
93
simulation, snapshot 1.
94
B. D. Rogers and R. A. Dalrymple
Fig. 7. Test 4 — Tsunami generation by a three-dimensional subaerial landslide, run 30: Monaghan boundary condition, 3-D SPH
simulation, snapshot 2.
SPH Modeling of Tsunami Waves
Fig. 8. Test 4 — Tsunami generation by a three-dimensional subaerial landslide, run 30: Monaghan boundary condition, 3-D SPH
simulation, snapshot 3.
95
96 B. D. Rogers and R. A. Dalrymple
0.2
0.1
free-surface (m)
0
0 0.5 1 1.5 2 2.5 3
-0.1
-0.2
experimental data
-0.3
SPH 2-D
-0.4 SPH 3-D
-0.5
time (s)
Fig. 9. Test 4 — Tsunami generation by a three-dimensional sub-aerial landslide, run 32:
Comparison of 3-D and 2-D SPH-SPS with wave gage 1 data.
Overall however, the results from this case clearly demonstrate one
of the essential drawbacks of SPH. The resolution needed to perform an
accurate simulation of this test case must be on the order of O (∆x, ∆y,
∆z) ≈ 0.005m. For the dimensions of the current case, this would mean ~
32×107 particles, which can only be accomplished using high-
performance parallel computing, this is unavailable to the authors at the
present time.
far-field flow which are then linked to SPH which models the
complicated 3-D fluid-structure interaction. Hence, of particular
engineering merit will be the patching together of different numerical
models with SPH whereby flow information is passed between each
model. At JHU, work is already underway investigating how SPH can
be linked to shallow water and Boussinesq-type models (FUNWAVE
from the University of Delaware) that will predict the far-flow field
utilizing SPH to focus on localised flow features. Boussinesq-type
models are now a well established technique for predicting the
propagation of dispersive and nonlinear waves in shallow water. Recent
advances (e.g. Madsen et al. 2003) have greatly extended the accuracy of
the linear dispersion characteristics and nonlinear properties of
Boussinesq-type models. In the depth-integrated approach, the modeling
of wave breaking itself has been difficult requiring approximate
submodels such as modifications of Svendsen’s (1984) decay of a
turbulent bore. SPH will clearly be able to handle the turbulent flow
field with the propagation of the breaking wave.
Following tsunami propagation into shallow water, SPH will also be
ideally suited to examining the three-dimensional flow fields of tsunami-
interaction with coastal structures and inundation of dry land. This will
allow modelers to address the important issues of tsunami-structure
interaction and modeling debris flow following the passage of the
tsunami wave.
A natural consequence of being able to model tsunami-structure
interaction is that we also envisage using SPH to model the rupture of
safety critical buildings such as oil storage tanks. In contrast to other
formulations of fluids flow and intense wave breaking, SPH facilitates
easily the modeling of multi-fluid flow enabling us to examine the
transport and dispersal of contaminants and pollutants following such
catastrophic failures.
Acknowledgment
This work was supported by the Office of Naval Research (ONR).
98 B. D. Rogers and R. A. Dalrymple
References
1. Batchelor, Sir. G. K., An Introduction to Fluid Dynamics (Cambridge University
Press, 1967) pp. 635.
2. Blin, L., A. Hadjadj and L. Vervisch, Large eddy simulation of turbulent flows in
reversing systems, Selected Proceedings of the 1st French Seminar on Turbulence
and Space Launchers, CNES-Paris, 13–14 June, 2002, eds. P. Vuillermoz, P. Comte
and M. Lesieur, 2002.
3. Briggs, M. J., C. E. Synolakis and G. S. Harkins, Tsunami runup on a conical
island, Proc. Waves — Physical and Numerical Modeling, University of British
Columbia, Vancouver, Canada, 446– 455, 1994.
4. Christensen, E. D. and R. Deigaard, Large eddy simulation of breaking waves,
Coastal Eng. 42, 53–86, 2001.
5. Dalrymple, R. A. and O. Knio, SPH modeling of water waves, Proc. Coastal
Dynamics 2001, ASCE, 779–787, Lund, Sweden, 2001.
6. Dalrymple, R. A., O. Knio, D. T. Cox, M. Gesteira, and S. Zou, Using a Lagrangian
particle method for deck overtopping, Proc. Ocean Wave Measurement and
Analysis, ASCE, 1082–1091, 2001.
7. Gingold, R. A. and J. J. Monaghan, Smoothed particle hydrodynamics: theory and
application to non-spherical stars, Mon. Not. R. Astron. Soc., 181, 375–389, 1977.
8. Gómez-Gesteira, M. and R. A. Dalrymple, Using a 3D SPH method for wave
impact on a tall structure, J. Waterways, Port, Coastal, and Ocean Engineering,
ASCE, 130(2), 63–69, 2004.
9. Gotoh, H., T. Sakai and M. Hayashi, Lagrangian model of drift-timbers induced
flood by using moving particle semi-implicit method, J. Hydroscience and Hyd.
Eng. 20(1), 95–102, 2002.
10. Gotoh, H., T. Shibihara and T. Sakai, Sub-particle-scale model for the MPS method
— Lagrangian flow model for hydraulic engineering, Computational Fluid
Dynamics J., 9(4), 339–347, 2001.
11. Gotoh, H., S. Shao and T. Memita, SPH-LES model for numerical investigation of
wave interaction with partially immersed breakwater, Coastal Engineering J., 46(1),
39–63, 2004.
12. Grilli, S. T., S. Vogelmann and P. Watts, Development of a 3D Numerical Wave
Tank for modeling tsunami generation by underwater landslides, Eng. Analysis
Boundary Elemt. 26(4), 301–313, 2002.
13. Johnson, G. R. and S. R. Beissel, Normalized smoothing functions for SPH impact
computations, Int. J. Num. Meth. Eng. 39, 2725–2741, 1996.
14. Koshizuka, S., A. Nobe and Y. Oka, Numerical analysis of breaking waves using
the moving particle semi-implicit method, Int. J. Numer. Meth. Fluids, 26, 751–
769, 1998.
15. Li, S. and W. K. Liu, Meshfree and particle methods and their applications, Appl.
Mech. Rev. 55(1), 1–34, 2002.
SPH Modeling of Tsunami Waves 99
16. Li, Y. and F. Raichlen, Energy balance model for breaking solitary wave runup, J.
Waterways, Port, Coastal and Ocean Engineering, 47–59, 2003.
17. Lin, P., K.-A. Chang and P. L.-F. Liu, Runup and Rundown of Solitary Waves on
Sloping Beaches, Journal of Waterways, Port, Coastal and Ocean Engineering,
125(5), 247–255, 1999.
18. Liu, P. L.-F., P. Lynett and K. Synolakis, Analytical solutions for forced long waves
on a sloping beach, Journal of Fluid Mechanics, 478, 101–109, 2003.
19. Lo, E. Y. M. and S. Shao, Simulation of near-shore solitary wave mechanics by an
incompressible SPH method, Applied Ocean Research, 24, 275–286, 2002.
20. Lucy, L. B., A numerical approach to the testing of fusion process, Astronomical J.,
88, 1013–1024, 1977.
21. Lynett, P., T.-R. Wu and P. L.-F. Liu, Modeling Wave Runup with Depth-
Integrated Equations, Coastal Engineering, 46(2), 89–107, 2002.
22. Madsen, P. A., H. B. Bingham and H. A. Schaffer, Boussinesq-type formulations
for fully nonlinear and extremely dispersive water waves: derivation and analysis,
Proc. Roy. Soc. London, Series A, 459(2033), 1075–1104, 2003.
23. McGuire, B., Apocalypse: a Natural History of Global Disasters, Cassell, London,
1999.
24. Monaghan, J. J., Smoothed particle hydrodynamics, Annual Review of Astronomy
and Astrophysics, 30, 543–574m, 1992.
25. Monaghan, J. J., Simulating free surface flows with SPH, J. Computational Physics,
110, 399–406, 1994.
26. Meneveau, C. and J. Katz., Scale-invariance and turbulence models for large-eddy
simulation, Annual Rev. Fluid Mechanics, 32, 1–32, 2000.
27. Monaghan, J. J., SPH without a tensile instability, J. Computational Physics, 159,
290–311, 2000.
28. Monaghan, J. J. and A. Kos, Solitary waves on a Cretan beach, J. Waterway, Port,
Coastal, and Ocean Eng, ASCE, 125, 3, 1999.
29. Morris, J., P. Fox and Y. Zhu, Modeling low Reynolds number incompressible
flows using SPH, Journal of Computational Physics, 136, 214–226, 1997.
30. Panizzo, A. and R. A. Dalrymple, SPH Modelling of Underwater Landslide
Generated Waves, Proc. 29th Intl. Conference on Coastal Engineering, Lisbon,
World Scientific Publishing Co., 1147–1159, 2004.
31. Raichlen, F. and C. E. Synolakis, Run-up from three dimensional sliding mass,
Proceedings of the Long Wave Symposium 2003, (Briggs, M, Coutitas, Ch.) XXX
IAHR Congress Proceedings, 247–256, 2003.
32. Randles, P. W. and L. D. Libersky, Smoothed Particle Hydrodynamics: Some
recent improvements and applications, Comput. Methods Appl. Mech. Eng., 139,
375–408, 1996.
33. Shankar, N. J. and M. P. R. Jayaratne, Wave run-up and overtopping on smooth and
rough slopes of coastal structures, Ocean Engineering, 30(2), 153–295, 2003.
100 B. D. Rogers and R. A. Dalrymple
34. Shao S. D., and H. Gotoh, Simulating coupled motion of progressive wave and
floating curtain wall by SPH-LES model, Coastal Engineering Journal, 46(2), 171–
202, 2004.
35. Smagorinsky, J., General circulation experiments with the primitive equations. I.
The basic experiment, Mon. Weather Rev., 91, 99–164, 1963.
36. Svendsen, I., Wave heights and set-up in a surf zone, Coastal Engineering, 8, 303–
329, 1984.
37. Toro, E. F., M. Olim and K. Takayama, Unusual increase in tsunami wave
amplitude at the Okushiri island: Mach reflection of shallow water waves, Proc.
22nd Int. Symp. on Shock Waves, Imperial College, London, UK, July 18–23 1999.
38. Tulin, M. P. and M. Landrini, Breaking waves in the ocean and around Ships, Proc.
23rd ONR Symposium on Naval Hydrodynamics, Val de Reuil, France, 2000.
39. Yeh, H. and M. Petroff, Bore in a box experiment, http://engr.smu.waves.edu/solid.html,
2003.
40. Yoshizawa, A., Statistical theory for compressible turbulent shear flows, with the
application to subgrid modeling, Phys. Fluids A, 29, 2152–2164, 1986.
CHAPTER 4
1. Introduction
Tsunamis could be generated by many different geophysical phenomena,
such as earthquake, landslide, and volcano eruption. Aerial and
submarine landslides have been documented as a source of several
destructive tsunamis. For example, in 1958 a rockslide, triggered by an
8.3 magnitude earthquake, occurred in Lituya Bay, Alaska. An estimated
volume of 30.6 million m3 of amphibole and biotite slide down into the
Gilbert Inlet at the head of Lituya Bay, generating large tsunamis. The
runup height on the opposite side of the slide in the Gilbert Inlet was
estimated as 524 m. On November 29, 1975, a landslide was triggered by
101
102 T.-R. Wu and P. L.-F. Liu
for the simulation of various breaking wave related problems [e.g., Lin
and Liu (1998a, 1998b)]. In their studies, the interface between air and
water was reconstructed by either a horizontal or a vertical line. This
piecewise constant interface reconstruction scheme is relatively crude.
Rider and Kothe (1998) provided a second-order reconstruction scheme,
in which the interface is approximated as piecewise linear. In this paper,
we will use the VOF method and adopt the piecewise linear interface
reconstruction scheme in a LES model.
In addition to the treatment of free surface, modeling the landslide-
generated waves faces another difficulty: the treatment of moving solid
boundaries (i.e., landslides). One possible method is to use curvilinear
boundary-fitted coordinates. However, for a problem with moving or
deforming boundaries the boundary-fitted grids must be either
regenerated or deformed as the boundary geometry changes, adding
considerable complexity to the computations. Another approach is the
immersed boundary method (IBM). This method introduces a body-force
field so that a desired velocity distribution can be assigned over a solid
boundary (e.g., Mohd-Yusof, 1997; Goldstein et al., 1993). This method
allows the imposition of the boundary conditions on a given surface
which does not coincide with the computational grids. The IBM has been
successfully applied to the internal combustion (IC) piston simulations
(Fadlun et al., 2000; Verzicco et al., 2000). The IBM has the advantage
of providing highly accurate results for moving boundary problems.
However, since this method depends on the interpolated velocity on the
boundary to evaluate the immersed body-force, it requires fine resolution
on the solid boundary.
Heinrich (1992) developed a partial cell method, in which a source
function is added to the continuity equation to represent the solid
boundary movement. He applied this method to simulate waves
generated by a two-dimensional landslide and obtained very good
numerical results for water surface profiles as compared to the laboratory
data. This moving solid algorithm does not require any interpolation
scheme and is independent from the grid system. Thus, it is extended to
three-dimensional problems in the current study.
The present numerical examples concern the solid landslide motion
as well as the landslide generated breaking wave runup and rundown on
106 T.-R. Wu and P. L.-F. Liu
∂u 1 1
+ ∇ ⋅ (uu ) = − ∇p + g + ∇ ⋅τɶ, (2.2)
∂t ρw ρw
where u represents the velocity vector, ρw water density, τɶ the stress
tensor, g the gravity force vector, t time, and p pressure. The viscous
stress tensor is a function of the molecular viscosity µw and the strain
rate ∇u :
τɶ = µ w (∇u + ∇T u ). (2.3)
∂u i ∂uiu j 1 ∂p 1 ∂τij
+ =− + gi + , (2.5)
∂t ∂x j ρw ∂xi ρw ∂x j
∂u ∂u
τij = µw i + .
j
(2.6)
∂x j ∂xi
ui = usi . (2.7)
Γ1 Γ2
0
Γ1
u si
Ω
usi0
Fig. 2.1. A sketch of the flow domain and boundaries. The gray parts indicate the solid
material.
∂u n
− p + 2µw = S n
∂n
∂uT , on Γ 2 (2.8)
∂un
µ w k
+
= STk
∂n ∂Tk
where the subscripts n and Tk denote the outward normal direction and
two tangential directions ( k =1, 2), respectively, on the free surface. Sn
and STk are the normal and tangential stress components specified by the
air flow on the free surface. The kinematic boundary condition requires
that the free surface be a sharp interface separating the two fluids so that
no flow is allowed to go through it. The mathematical expression of the
kinematic boundary can be derived from the equation that describes the
free surface. If the free surface is expressed as J (x, t ) = 0 , the total
derivative of J ( x, t ) must be zero on the surface:
DJ ( x, t )
=0 (2.9)
Dt on J (x ,t )=0 or on Γ 2
or
∂J
+ u ⋅∇J = 0 on J (x, t ) = 0, or on Γ2 . (2.10)
∂t
For the initial conditions, the whole flow field has to be prescribed:
∂ui0
= 0, in Ω
∂xi
∂ui0 ∂ui u j 1 ∂τij
0 0 0
1 ∂p 0
+ =− + gi + , in Ω
∂t ∂x j ρw ∂xi ρw ∂x j
ui0 = usi0 , on Γ1
∂un0 (2.13)
− p 0 + 2µ w = S n0 , on Γ2
∂n
∂uT0 ∂u 0
µw k + n = ST0k , on Γ2
∂n ∂Tk
∂J 0 ∂J 0
+ ui0 = 0, on Γ2 .
∂t ∂xi
2.2.1. Filtering
The filtered N-S equations can be derived by applying the filter function
to the N-S equations. The filtered continuity and momentum equations
are:
∂u ∂u
i = i = 0, (2.18)
∂x ∂x
i i
∂ (ui ) ∂ (uiu j ) 1 ∂p
+ = − + gi
∂t ∂x j ρw ∂xi
(2.19)
1 ∂ ∂ui ∂u j
+ µw + ,
ρw ∂x j ∂x j ∂xi
Eddy Simulation Model for Tsunami and Runup Generated by Landslides 111
ui u j ≠ ui u j . (2.20)
τijR ≡ ui u j − ui u j . (2.21)
In LES, τijR is also called the SGS Reynolds stress. The residual
kinetic energy is
1
kr ≡ τiiR , (2.22)
2
and the anisotropic residual-stress tensor is defined by
2
τijr ≡ τijR − kr δij . (2.23)
3
The isotropic residual stress is included in the modified filtered pressure
2
p ≡ p + kr . (2.24)
3
Substituting (2.23) and (2.24) into (2.19), the filtered momentum
equation can be rewritten as:
D (ui ) 1 ∂p
=− + gi
Dt ρw ∂xi
(2.25)
1 ∂ ∂ui ∂u j 1 ∂τij
r
+ µw + − ,
ρw ∂x j ∂x j ∂xi ρw ∂x j
where the substantial derivative based on the filtered velocity is:
112 T.-R. Wu and P. L.-F. Liu
D ∂
≡ + u ⋅∇. (2.26)
Dt ∂t
In (2.25), the SGS Reynolds stress contains the local average of the small
scale field, therefore the SGS model should be based on the local
velocity field. The most commonly used SGS model is the Smagorinsky
model (Smagorinsky, 1963). It is essentially a linear eddy viscosity
model:
∂u ∂u
τijr = −ν t i + = −2ν t Sij .
j
(2.27)
∂x j ∂xi
The Smagorinsky model relates the residual stress to the filtered
strain rate. The coefficient ν t ( x, t ) is the subgrid-scale eddy viscosity of
the residual motions. Based on the dimensional analysis, the subgrid-
scale eddy viscosity is then modeled as:
ν t = ℓ 2S S
(2.28)
= (CS ∆) S,
2
S ≡ (2 Sij Sij ) .
12
(2.29)
where dx1 , dx2 , and dx3 are the three components of the grid lengths.
Under the isotropic turbulence condition, the Smagorinsky coefficient
CS ≈ 0.2 . However, in general, CS is not a constant; its value varies
Eddy Simulation Model for Tsunami and Runup Generated by Landslides 113
∂ui ∂ui u j 1 ∂p
+ =− + gi
∂t ∂x j ρw ∂xi
(2.31)
1 ∂ ∂u ∂u j
+ µw,eff i + ,
ρ w ∂x j ∂x j ∂xi
where µw,eff = µw + µt is the sum of molecular and eddy viscosities.
In the near-wall viscous sub-layer region, the largest local turbulent eddy
sizes are limited by the viscous scales. The well resolved LES requires
grids nearly as fine as those used in the DNS. This restriction should be
applied not only in the wall-normal direction but also in the streamwise
direction. The near-wall resolution requirement clearly limits the
application of LES in simulating high Reynolds number flows.
Therefore, a modeling strategy is required to solve the practical
applications. Instead of modeling every detail in the near-wall region,
this study uses a wall function approach to reduce the number of
computational cells. Cabot and Moin (2000) derived a set of near-wall
damping functions and used them to approximate the eddy viscosity in
114 T.-R. Wu and P. L.-F. Liu
the first cell adjacent to the wall. The eddy viscosity ν t is obtained from
a mixing-length eddy viscosity model with near-wall damping:
νt
( )
+ 2
= κ yw+ 1− e− yw / A , (2.32)
ν
where yw+ = ywuτ / ν is the distance to the wall in wall units, κ = 0.41
and A = 19 (Cabot and Moin, 2000). The filtered velocity field is
assumed to satisfy the no-slip condition on the wall.
3. Numerical Implementation
The filtered continuity and momentum equations are solved by using the
finite-volume two-step projection method (Bussmann et al., 2002). The
forward time difference method is used to discretize the time derivatives.
In order to track the free-surface, the VOF method is used. The VOF
method was originally developed by Hirt and Nichols (1981) and
improved by Kothe et al. (1999) to second-order accuracy on the free
surface reconstruction. The current numerical model is modified from
Truchas 1.8.4 developed by Doug Kothe, Jim Sicilian and their Telluride
team members at Los Alamos National Laboratory. The original program
solves N-S equations by using the projection method with the finite
volume discretization. A few new algorithms have been added in order to
simulate turbulent free surface flows generated by a moving landslide.
The VOF method as well as the moving boundary algorithm will be
described in Section 3.2.
In order to simulate the free-surface motion in a fixed grid system,
which include both air and water, a volume of fluid function, f ,
representing the volume fraction of water within a computational cell is
introduced. The f value equals to one if the cell is full, zero if empty,
and 0 < f < 1 if the cell is partially filled with water (Fig. 3.1). Since f
is conserved and governed by the transport equation, it can be described
by:
∂f ∂ui f
+ =0. (3.1)
∂t ∂xi
Eddy Simulation Model for Tsunami and Runup Generated by Landslides 115
The f value represents the spatial position of the fluid body. It also
provides the kinematic information on the fluid.
f =
Fig. 3.1. A sketch of the volume fraction f , where f = 1 indicates that the cell is filled with
water, f = 1 filled with air, and 0 < f < 1 contains an air-water interface.
∂ρui ∂ρui u j ∂p
+ =− + ρ gi
∂t ∂x j ∂xi
(3.2)
∂ ∂u ∂u
+ µ i + j ,
eff ∂x
∂x j j ∂xi
ρ = f ρw (3.3)
∂ρu
+ ∇⋅ (ρ u u ) = −∇p + ρg + ∇⋅ µeff (∇u + ∇T u) ,
n
(3.4)
∂t
where µeff is the cell viscosity.
∂φ
+ ∇⋅ (uφ ) = S (φ ) , (3.5)
∂t
where S is a source term which is generally dependent upon φ . The
finite volume method starts from the volume integral form of the partial
differential equation, e.g., (3.5), we have,
∂φ
∫ + ∇⋅ (uφ ) = S (φ ) dV . (3.6)
∂t
By adopting the Gauss divergence theorem, (3.6) can be expressed as
∂φ
∫ ∂t dA ⋅(uφ) = ∫ S (φ) dV ,
dV + ∫ (3.7)
n1
n2 dA1
dA2
dV
n4
dA4
dA3
n3
Fig. 3.2. A sketch of the control volume and face normal vectors. Each face has a unique
normal vector n and face area dA.
integral is over the bounding surface of the cell. Equation (3.7) can then
be approximated by a discrete numerical scheme:
φin+1 − φin 1
dt
+
Vi
∑ A ⋅ u
F
n [φ ]n = Sin ,
F F (3.9)
where superscript n indicates the nth time step, dt is the time step size,
Vi is the i th cell volume, and subscript F denotes the cell face.
On the left-hand side of (3.9), the surface integral has been
approximated as a sum over discrete cell faces F with an area vector
A F ; Vi is the control volume, and φι is a local volume-averaged value,
∫ φdV = 1 φdV ,
φi = ∫ (3.10)
∫ dV V i
φi = φ (xc ) , (3.11)
dVwn, F
f Fn = . (3.14)
dVFn
Once the solution to (3.13) is found, the fluid volumes are marched
forward in time.
Eddy Simulation Model for Tsunami and Runup Generated by Landslides 119
uF dt uF dt
[ f 2 ]F
n
dV2,F dV2,F f 2,nF
uF uF
[ f1 ]F
n
dV1,F dV1,F f1,nF
Fig. 3.3. The advected mass through a cell face. The left one tends to have larger error
than the right one.
To solve the volume fraction f n+1 at the new time-step from (3.13), we
need the information on f Fn as well as the free surface reconstruction. In
this paper, a multidimensional PLIC (piecewise linear interface
calculation) (Rider and Kothe, 1998) is utilized to construct the plane in
each free surface cell. The plane is then used to estimate f Fn .
The PLIC algorithm consists of two steps: the planar reconstruction
and the calculation of the volume of water across a cell face. As for the
first step, the free surface within a cell is assumed to be a plane. Using
the information on f n , the orientation of the free surface can be
estimated as the gradient of f n . As for the second step, the volume of
water across a cell face can be evaluated by multiplying the volume flux
by dt . This volume is then used to update the volume fraction in the cell
and to transport other quantities in the momentum equations. We shall
discuss the detailed algorithm in the following sections.
∇f
nˆ = . (3.16)
∇f
The cell center and cell face gradient calculations are carried out via the
least square algorithms.
By assuming that the free surface geometry is piecewise linear
(planar), the free surface in each free surface cell can be described as:
nˆ ⋅ x − C p = 0 , (3.17)
where x is the position vector of a point on the plane and C p is the plane
constant to be determined.
After determining the normal vector, n̂ , on the free surface, the free
surface locations can be calculated iteratively. For each free surface cell,
the free surface plane divides space into regions inside and outside of
water body, depending upon the direction chosen by n̂ . Based on the
definition expressed in (3.16), n̂ points to the interior of the water body.
Hence, nˆ ⋅ x − C p will be positive for any point inside the water body;
zero for any point x on the free surface plane; and negative for any point
x outside the water body. The volume Vtr is the truncation volume lying
within the water body and separated by (3.17). By varying C p , Vtr will
also be changed. Our goal is to find the constant C p iteratively (see
Fig. 3.4) so that a nonlinear function defined as:
E (C p ) = Vtr (C p ) − f ∗ V (3.18)
Xp
Fig. 3.4. Locating the interface. The interface can be moved up and down by changing C
(Eq. (3.17)) and followed the direction of n̂ . This is constrained by the volume
conservation.
The projection method has been widely used to solve the NS equations.
This method was first proposed by Chorin (1968) and then used by Bell
and his coworkers to solve the constant-density (Issa, 1986) or variable-
density (Rider and Kothe, 1998) incompressible N-S equations. In this
study, the projection method will be used to solve the filtered N-S
equations.
In the projection method, momentum equations, (3.2), can be
described by two fractional steps:
ρ n+1u* − ρ n u n
dt
n n
(
= −∇⋅ (ρ n u u) + ∇⋅ µeff (∇u + ∇T u) ,
n
) (3.19)
ρ n+1u n+1 − ρ n u n
= −∇⋅ (ρ u u ) −∇p n+1 + ρ n+1g
n
dt (3.21)
+∇⋅ µ ( n
eff (∇u n
+ ∇ u) .
T n
)
No additional approximation results from this decomposition.
Equation (3.20) relates u n+1 to u* . By adopting the solenoidal
condition, (2.18), we have:
∇p n+1 u*
∇⋅ = ∇⋅ + g . (3.22)
ρ n+1 dt
The above equation is also called the Poisson Pressure Equation (PPE).
The cell-centered pressure p n+1 at the new time step can be obtained by
solving (3.22), and cell-face pressure gradient ∇pFn+1 is calculated from
the same stencil as that of the density interpolation to faces. Equation
(3.20) is then used to calculate the solenoidal face velocity field:
∇pFn+1
u n+1
= u − dt n+1 − g .
*
(3.23)
F F
ρ F
) dV ≈ ∑
(ρuudV
∫ ∇⋅uu
n
dt dt∇⋅ ≈ M FnM
u nF u
F
(3.28)
= dt ∑ ρw ( A F ⋅ u nF ) f Fn u nF ,
F
reduces to a sum of the dot product of the face normal vector with the
local velocity gradient multiplied by the face area:
(
∇⋅ µeff
n
(∇u + ∇T u)
n
)
(3.29)
, F AF n F ⋅ (∇u F + ∇ u F ) .
= ∑ µeff ˆ
n
n T
F
2
µeff , F = , (3.31)
1 µnb +1 µ p
where µnb is the neighbor cell viscosity and µ p is the cell center
viscosity.
3.3.3. Projection
where the cell face pressure gradient ∇pFn+1 is calculated from the same
stencil as that of the density interpolation to faces:
n+1
n +1
pnb − p pn+1
∇p F = . (3.33)
dx
Here pnb n +1
is the neighbor cell pressure, p np+1 the cell pressure, and δ x
the centroid-to-centroid distance.
Equation (3.32) is first solved for the cell-centered pressure p n+1 .
Equation (3.23) is then used to calculate the solenoidal face velocity field
u nF+1 . The last step is to interpolate ∇pFn+1 / (ρ Fn+1 ) − g to cell centers,
which, along with (3.20) and u* , is used to obtain cell center velocity
field u n+1 .
Equation (3.32), the PPE, is a set of linear algebraic equations for the
pressure field that can be solved by iterative methods. In this study, the
conjugate gradient (CG) algorithm with symmetric successive over-
relation (SSOR) (Varga, 1962) pre-conditioner (Golub and Van Loan,
1989) will be used to solve the linear algebraic equations.
volume flux information obtained directly from the previous step, and
the viscous force is obtained via (3.29).
Eddy Simulation Model for Tsunami and Runup Generated by Landslides 127
dVobst(t)
∫ u ⋅ ndA =
A dt
. (3.34)
The new continuity equation for the volume V can be expressed as:
1 dVobst(t )
∇⋅ u = = φ ( x, y , z , t ) , (3.35)
V dt
where φ = internal source function. Substituting the new continuity
equation, (3.35), into the original governing equations, the momentum
equation becomes:
∂ (ρui ) ∂ (ρ ui u j ) ∂p
+ = − + ρ gi
∂t ∂x j ∂xi (3.36)
∂ ∂ui ∂u j
+ µeff + + ρu iφ.
∂x j ∂x j ∂xi
128 T.-R. Wu and P. L.-F. Liu
The main procedure of the moving solid algorithm can be expressed as:
1. Update the new time step solid body VOF (Fig. 3.6).
2. Add a positive source function to the region where water will be
pushed out, and a sink term to the wake zone where water will be
sucked in (Fig. 3.7).
3. Solve the new time step velocities.
t1
t2
Fig. 3.6. Update the solid body VOF from old time step (t1) to new time step (t2).
One of the advantages of using the moving solid algorithm is that the
grid system does not need to fit the solid boundary. Therefore, a part of
the internal cells will be occupied by the solid material. In order to deal
with a cell occupied partially by solid materials and water, a simple
partial cell treatment will be applied.
t1
t3 (=t2 + dt)
Fig. 3.7. Add a positive source term in front of the solid body, and a negative source term
(sink term) in the wake zone. Where dt = t2 – t1, t3 = t2 + dt.
t1 t1
t2
t2
t3 (=t2 + dt)
Fig. 3.8. The comparison between Heinrich’s method and present moving solid
algorithm.
130 T.-R. Wu and P. L.-F. Liu
The purpose of the partial cell treatment is to adopt the moving solid
algorithm and to deal with unstructured interior obstacles. Since the
obstacle inside the computational domain occupies a part of the interior
fluid volume, the effective cell volume will be a fraction of the original
cell volume:
Veff = (1− f solid )V
(3.37)
= θV ,
where Veff is the effective cell volume, f solid is the volume fraction
occupied by solid material in each cell, V is the original cell volume,
and θ is the effective volume fraction (Fig. 3.9).
Veff = V Veff = θV θV uF = 0
uF = 0
θV θV uF = 0
uF = 0
θV uF = 0
(Solid material)
uF = 0
Fig. 3.9. A sketch of the partial cell treatment. If a cell contains partial solid material, the
effective cell volume will be a fraction of the original cell volume. The face velocity will
be set to zero if at least one side of the cell is fully filled with solid material.
If a cell contains partially water and solid material, the flow solver
has to deal with it. Cell faces are defined either to be entirely closed or
opened. Cell faces are “closed” only if at least one of the two
immediately neighboring cells is entirely occupied by solid material. If
Eddy Simulation Model for Tsunami and Runup Generated by Landslides 131
the cell faces are “closed”, the face velocity of the cell is set to zero, and
the face pressure is no longer calculated in the pressure solution. On the
other hand, if any face between two cells, containing at least a partial cell
volume of fluid, is “open”, the code solves the velocities and pressure
gradients.
Compared to the conventional way to treat the irregular solid
material that creates the “saw-tooth” boundary, the partial cell treatment
introduced here is a better choice to represent the smooth real geometry
of the boundary.
z
η
−δ x
Fig. 4.1. A definition sketch for the physical variables employed in this study.
pe
Slo
1:2
Fig. 4.2. The numerical setup and gauge positions. Unit is in meter.
The numerical simulations are carried out in a numerical wave tank and
the landslide is modeled by the moving solid algorithm. Because this
study focuses on the runup and water waves generated in the near field,
the length of the computational domain is shorter than the laboratory
Eddy Simulation Model for Tsunami and Runup Generated by Landslides 133
setup (6.6 m). Furthermore, since the vortex shading behind the moving
wedge is small, we assume that the flow field is symmetric with respect
to the centerline and thus, only half of the physical domain is simulated
(Fig. 4.2). The number of grid point is 60 in the streamwise (the direction
of landslide movement) direction, 50 in the span-wise direction, and 60
in the vertical direction. A non-uniform grid is used in the streamwise
and vertical directions. The finest cell is located at the onshore corner to
have a finer resolution for the runup simulations. Coarser resolution is
applied to the offshore deep water corner. The side walls as well as the
end-wall are treated as impermeable free-slip walls. Therefore, the
landslide generated waves will be reflected by the end-wall, which does
not exist in the laboratory experiments. However, since the maximum
and minimum runup will occur before the reflected waves reach the
domain of interest, the reflecting end-wall boundary condition is still a
good choice to simplify the numerical setup. Other types of boundary
conditions such as numerical sponge layer and advective open boundary
condition (Liu et al., 2005) can also be used to avoid the reflection
from the end-wall. The boundary condition on the ceiling of the
computational domain is the zero-pressure open boundary condition. The
1:2 slope is designed to extend from upper-right corner to lower-left
corner (Fig. 4.2), so the solid material will occupy 50% of the cell
volume on the slope. This design ensures that the effective cell volume
will not be smaller than 50% of the total cell volume, therefore the time
marching will not be limited by certain cells with extremely small
effective cell volumes due to the Courant number restriction. This design
also helps us to clearly identify the shoreline location without being
interfered with the irregular effective cell volume. The time step is
determined dynamically based on the Courant number criteria, which is
0.45 in this study.
Similar to the laboratory setup, three runup gauges are installed close
to the moving landslide; four wave gauges numbered from #4 to #7 are
installed in the offshore direction; and one wave gauge array with six
wave gauges numbered from #8 to #11 is installed on the lateral side of
the moving wedge (Fig. 4.2). However, the gauge locations might not
coincide with grid centroid. Thus, the nearest neighbor interpolation
134 T.-R. Wu and P. L.-F. Liu
4
S (m)
0
0 0.5 1 1.5 2 2.5 3
time (sec)
3
Moving speed (m/sec)
0
0 0.5 1 1.5 2 2.5 3
time (sec)
Fig. 4.3. The displacement (S) and the speed of the moving slide. In this subaerial case,
the initial wedge position ( δ ) is 0.454 m above the still water surface. The specific
weight ( γ ) of the wedge is 2.64.
In the laboratory setup, two runup gauges were installed on the slope to
provide the time history of the runup information. The first runup gauge
(Gauge #2) is located at the edge of the sliding wedge, the other one is
located at one wedge width away from the centerline. Figure 4.4 shows
the comparison between the numerical results and laboratory data. The
solid lines are the numerical results and the broken lines are the
experimental measurements. The numerical runup heights are determined
by the contour line where the water occupies 50% of the effective cell
volume. A very good agreement is shown at Gauge #2. However, at
Gauge #3 the numerical solutions slightly over-predict the maximum
runup heights and the disagreement is about 10%. Overall, the numerical
simulation is able to capture the maximum runup heights in the near field
region, and most importantly the model is able to accurately predict the
arrival time of the maximum runup height.
0.15
0.1 Gauge #2
0.05
R (m)
−0.05
−0.1
0.15
0.1 Gauge #3
0.05
R (m)
−0.05
−0.1
Fig. 4.4. The comparison between numerical results (solid lines) and experimental data
(broken lines) for the time history runup height at Gauge #2 and Gauge #3. δ = 0.454 m.
γ = 2.64.
136 T.-R. Wu and P. L.-F. Liu
There are four fixed wave gauges installed in front of the landslide and
they are labeled as gauge #4 ~ #7 (see Fig. 4.2). Figure 4.5 shows the
comparison of numerical results and wave gauge data. Again, the solid
lines are the numerical solutions and the broken lines are the
measurements. Gauge #4 and #6 are the ones closer to the shoreline. The
comparison shows that the numerical simulation successfully predicts
the leading wave height as well as the phase speed. Gauge #5 and #7
show that the numerical solutions under-predict the leading wave
heights. However, the numerical model is able to predict the phase of the
propagating waves.
0.15 0.15
Gauge #4 Gauge #5
0.1 0.1
0.05 0.05
η (m)
η (m)
0 0
−0.05 −0.05
−0.1 −0.1
0 1 2 3 4 0 1 2 3 4
Time (Sec) Time (Sec)
0.15 0.15
Gauge #6 Gauge #7
0.1 0.1
0.05 0.05
η (m)
η (m)
0 0
−0.05 −0.05
−0.1 −0.1
0 1 2 3 4 0 1 2 3 4
Time (Sec) Time (Sec)
Fig. 4.5. The comparison between numerical results (solid lines) and experimental data
(broken lines) for the time histories of free surface fluctuations at wave gauge #4 ~ #7;
δ = 0.454 m, γ = 2.64. The coordinates for gauges are: Gauge #4: (x, y) = (1.83, 0);
Gauge #5: (x, y) = (2.74, 0); Gauge #6: (x, y) = (1.83, 0.61); Gauge #7: (x, y) = (2.74,
0.61). The unit is in meter.
Eddy Simulation Model for Tsunami and Runup Generated by Landslides 137
Wave gauges #8 ~ #13 are installed on the side of the moving slide to
record the lateral movement of landslide-generated waves. Figure 4.6
shows both the numerical results and laboratory data. The comparisons
show that the numerical solutions successfully predict the wave heights
of the leading and secondary waves. The numerical solutions also
accurately capture the phase of the waves. However, from gauge #8, #9,
and #11, around time = 1.75 sec, the comparisons show that the
numerical solutions cannot predict some small scale motions. This might
be cause by the lack of fine resolution in numerical simulations. It could
also be the experimental errors because there are significant wave
breaking and air bubble entrainment. Fortunately, these errors do not
seem to affect the predictions of the secondary waves.
η (m)
η (m)
0 0 0
0 1 2 3 4 0 1 2 3 4 0 1 2 3 4
Time (Sec) Time (Sec) Time (Sec)
η (m)
η (m)
0 0 0
0 1 2 3 4 0 1 2 3 4 0 1 2 3 4
Time (Sec) Time (Sec) Time (Sec)
Fig. 4.6. The comparison between numerical results (solid lines) and experimental data
(broken lines) for the time histories of free surface fluctuations at wave gauge #8 ~ #13;
δ = 0.454 m, γ = 2.64. The coordinates for gauges are: Gauge #8: (x, y) = (0.4826,
1.092); Gauge #9: (x, y) = (0.8636, 1.092); Gauge #10: (x, y) = (1.2446, 1.092); Gauge
#11: (x, y) = (0.635, 0.4826); Gauge #12: (x, y) = (0.635, 0.8636); Gauge #13: (x, y) =
(0.635, 1.2446). The unit is in meter.
138 T.-R. Wu and P. L.-F. Liu
Figure 4.7 shows the snapshots of the free surface profiles. The initial
shoreline is at the intersection of x = 0 and z = 0 . The landslide first
pushes the water in front and generates the outgoing waves (time = 0.0 ~
0.8 sec). Then the landslide submerges into water and the free surface
caves in, generating strong lateral converging flows (time = 0.8 ~ 1.2
sec). While the landslide keeps submerging, the converging lateral flows
in the wake zone generate strong reflecting (rooster-tail) waves (time =
1.2 ~ 3.0 sec). These rooster-tail waves are the key source to the
maximum runup. After time = 3.0 sec, the solution is gradually
contaminated by the waves reflected from the side walls and will be
excluded from this discussion.
4.3.5. Velocities
Figure 4.8 presents the detailed velocity vectors at the centerline cross-
section. From time = 0.0 ~ 0.8 sec, the velocities are generated by the
entry of the landslide. After time = 0.8 sec, the elevation of the top of the
slide is lower than the still water surface and water starts to flood into the
wake area above the slide. At time = 1.2 sec, the slide is fully submerged.
A very complex three-dimensional flow pattern can be observed right
after the slide has fully submerged into water (time = 1.2 ~ 1.5 sec). The
depth-integrated equation models are not suitable for modeling this type
of flows. From time = 2.1 to 2.7 sec, the velocity distributions show that
a strong flow current has been generated by the slide motion, which has a
thickness about 1.5 ~ 2.0 times the front face height ( H slide ) of the slide.
Above this current, there exists a returning current and generates a
negative wave. Between these two currents, there exists an eddy that can
be clearly seen from time = 2.1 sec to 2.7 sec. The size of the eddy
depends on the moving speed of the slide. In this case, the eddy size is
about one front face height ( H slide ) of the slide. After t = 2.7 sec, the
slide stops. However, the current behind the slide still keeps on moving
with the decreasing strength.
Eddy Simulation Model for Tsunami and Runup Generated by Landslides 139
Fig. 4.7. Snapshots of free-surface profile for the sliding wedge with δ = 0.454 m, and
γ = 2.64. The unit is in meter.
140 T.-R. Wu and P. L.-F. Liu
Fig. 4.8. Snapshots of velocity vectors on the centerline vertical plane for the sliding
wedge with δ = 0.454 m, and γ = 2.64. The unit is in meter.
142 T.-R. Wu and P. L.-F. Liu
Figure 4.9 shows the snapshots of the shoreline movement. The snapshot
at time = 1.2 sec shows that the slide has fully submerged into water
and an air bubble is trapped there. After time = 1.5 sec, the shoreline
starts to move up, spreads out, and then reaches the maximum runup
height at time = 2.7 sec. From the sequence of the snapshots we
observe that at time = 2.7 sec, both the centerline and the near field
region (−1 m ≤ y ≤ 1 m) reaches the maximum runup. Therefore, the
intersection of the shoreline and z = 0 m at time = 2.7 sec can be treated
as the maximum inundation area.
4.3.7. x-Cross-Section
Fig. 4.9. Snapshots of shoreline movement for the sliding wedge with δ = 0.454 m, and
γ = 2.64. The unit is in meter.
144 T.-R. Wu and P. L.-F. Liu
Fig. 4.10. Velocity vectors on the vertical x-planes at time = 1.5 sec for the sliding wedge
with δ = 0.454 m, and γ = 2.64. The unit is in meter. The magnitude of the reference
vector indicates the speed of the wedge.
Eddy Simulation Model for Tsunami and Runup Generated by Landslides 145
4.3.8. y-Cross-Section
4.3.9. z-Cross-Section
5.1. Setup
Fig. 4.11. Velocity vectors on the vertical y-planes at time = 1.5 sec for the sliding wedge
with δ = 0.454 m, and γ = 2.64. The unit is in meter. The magnitude of the reference
vector indicates the speed of the wedge.
Eddy Simulation Model for Tsunami and Runup Generated by Landslides 147
Fig. 4.12. Velocity vectors on the horizontal z-planes at time = 1.5 sec for the sliding
wedge with δ = 0.454 m, and γ = 3.24. The unit is in meter. The magnitude of the
reference vector indicates the speed of the wedge.
148 T.-R. Wu and P. L.-F. Liu
4
S (m)
0
0 0.5 1 1.5 2 2.5 3
time (sec)
3
Moving speed (m/sec)
0
0 0.5 1 1.5 2 2.5 3
time (sec)
Fig. 5.1. The displacement (S) and the speed of the moving slide. In this subaerial case,
the initial wedge position ( δ ) is −0.1 m above the still water surface. The specific weight
( γ ) of the wedge is 2.64.
0.15
0.1 Gauge #2
0.05
R (m)
−0.05
−0.1
0.15
0.1 Gauge #3
0.05
R (m)
−0.05
−0.1
Fig. 5.2. The comparison between numerical results (solid lines) and experimental data
(broken lines) for the time history runup height at Gauge #2 and Gauge #3. δ = −0.1 m.
γ = 2.64.
Figure 5.4 shows the comparison of Gauge #8 ~ Gauge #13. These six
wave gauges mainly monitor the energy transfer in the lateral direction.
150 T.-R. Wu and P. L.-F. Liu
0.15 0.15
Gauge #4 Gauge #5
0.1 0.1
0.05 0.05
η (m)
η (m)
0 0
−0.05 −0.05
−0.1 −0.1
0 1 2 3 4 0 1 2 3 4
Time (Sec) Time (Sec)
0.15 0.15
Gauge #6 Gauge #7
0.1 0.1
0.05 0.05
η (m)
η (m)
0 0
−0.05 −0.05
−0.1 −0.1
0 1 2 3 4 0 1 2 3 4
Time (Sec) Time (Sec)
Fig. 5.3. The comparison between numerical results (solid lines) and experimental data
(broken lines) for the time histories of free surface fluctuations at wave gauge #4 ~ #7;
δ = −0.1 m, γ = 2.64. The coordinates for gauges are: Gauge #4: (x, y) = (1.83, 0);
Gauge #5: (x, y) = (2.74, 0); Gauge #6: (x, y) = (1.83, 0.61); Gauge #7: (x, y) = (2.74,
0.61). The unit is in meter.
0.15 0.15 0.15
Gauge #8 Gauge #9 Gauge #10
0.1 0.1 0.1
η (m)
η (m)
0 0 0
0 1 2 3 4 0 1 2 3 4 0 1 2 3 4
Time (Sec) Time (Sec) Time (Sec)
η (m)
η (m)
0 0 0
0 1 2 3 4 0 1 2 3 4 0 1 2 3 4
Time (Sec) Time (Sec) Time (Sec)
Fig. 5.4. The comparison between numerical results (solid lines) and experimental data
(broken lines) for the time histories of free surface fluctuations at wave gauge #8 ~ #13;
δ = −0.1 m, γ = 2.64. The coordinates for gauges are: Gauge #8: (x, y) = (0.4826,
1.092); Gauge #9: (x, y) = (0.8636, 1.092); Gauge #10: (x, y) = (1.2446, 1.092); Gauge
#11: (x, y) = (0.635, 0.4826); Gauge #12: (x, y) = (0.635, 0.8636); Gauge #13: (x, y) =
(0.635, 1.2446). The unit is in meter.
Eddy Simulation Model for Tsunami and Runup Generated by Landslides 151
The comparisons show that before time = 3 sec, the numerical solution
has a good agreement with the laboratory data. After time = 3 sec, the
solution and data are contaminated by the reflected wave from the side
walls, and will be excluded from current discussion.
5.2.4. Snapshots
5.2.5. Cross-Sections
In this section, the detailed velocity fields (time = 1.5 sec) at different
cross-sections will be presented from Figs. 5.8–5.10. Because the major
observations are close to that of the subaerial case, only the differences
and important phenomena will be addressed.
152 T.-R. Wu and P. L.-F. Liu
Fig. 5.5. Snapshots of free-surface profile for the sliding wedge with δ = −0.1 m, and
γ = 2.64. The unit is in meter.
Eddy Simulation Model for Tsunami and Runup Generated by Landslides 153
Fig. 5.6. Snapshots of velocity vectors on the centerline vertical plane for the sliding
wedge with δ = −0.1 m, and γ = 2.64. The unit is in meter.
Eddy Simulation Model for Tsunami and Runup Generated by Landslides 155
Fig. 5.7. Snapshots of shoreline movement for the sliding wedge with δ = −0.1 m, and
γ = 2.64. The unit is in meter.
156 T.-R. Wu and P. L.-F. Liu
Fig. 5.8. Velocity vectors on the vertical x-planes at time = 1.5 sec for the sliding wedge
with δ = −0.1 m, and γ = 2.64. The unit is in meter. The magnitude of the reference
vector indicates the speed of the wedge.
Eddy Simulation Model for Tsunami and Runup Generated by Landslides 157
Fig. 5.9. Velocity vectors on the vertical y-planes at time = 1.5 sec for the sliding wedge
with δ = −0.1 m, and γ = 2.64. The unit is in meter. The magnitude of the reference
vector indicates the speed of the wedge.
158 T.-R. Wu and P. L.-F. Liu
Fig. 5.10. Velocity vectors on the horizontal z-planes at time = 1.5 sec for the sliding
wedge with δ = −0.1 m, and γ = 2.64. The unit is in meter. The magnitude of the
reference vector indicates the speed of the wedge.
Eddy Simulation Model for Tsunami and Runup Generated by Landslides 159
6. Concluding Remarks
In this paper we have presented a LES model for 3D free surface flows.
Although the applications shown in the paper are for landslide generated
waves and shoreline movements, the model can be used as a tool for
studying other practical coastal problems. The advantage of the
numerical simulation approach is its ability in providing a large amount
of information in the entire flow domain of interest, which is not possible
for most of laboratory observations. This is especially true for a large
scale experiment.
The success in developing a moving solid algorithm greatly extends
the model capability to simulate wave-structure interaction problems.
However, further improvement on the turbulent closure modeling for
flow separation and wave breaking is essential.
Acknowledgment
This work was supported by research grants from National Science
Foundation to Cornell University.
References
1. Bardet, J.-P., Synolakis, C. E., Davies, H. L., Imamura, F., and Okal, E. A. (2003)
Landslide Tsunamis: Recent Findings and Research Directions. Pure Appl. Geohys.
160, 1793–1809.
2. Barth, T. J. (1995) Aspects of unstructured grids and finite-volume solvers for Euler
and Navier Stokes equations, VKI/NASA/AGARD Special Course on Unstructured
Grid Methods for Advection Dominated Flows, AGARD Publication R-787.
160 T.-R. Wu and P. L.-F. Liu
3. Bussmann, M. Kothe, D. B., and Sicilian J. M. (2002) Modeling high density ratio
incompressible interfacial flows. Proceedings of FRDSM’02, 2002 ASME Fluids
Engineering Division Summer Meeting. FEDSM2002-31123, 1–7.
4. Cabot, W. and Moin, P. (2000) Approximate wall boundary conditions in the large-
eddy simulation of high Reynolds number flow. Flow Turb. Combust. 63, 269–291.
5. Chorin, A. J. (1968) Numerical solution of the Navier-Stokes equations. Math.
Comp., 22, 745–762.
6. Christensen, E. D., and Deigaard R. (2001) Large eddy simulation of breaking
waves, Coast. Eng. 42, 53–86.
7. Cox, D. C. and Morgan, J. (1977) Local tsunamis and possible local tsunamis in
Hawaii, HIG 77–14, Hawaii Institute of Geophysics, University of Hawaii, pp. 118.
8. Deardorff, J. W. (1970) A numerical study of three-dimensional turbulent channel
flow at large Reynolds numbers. J. Fluid Mech., 41(2), 452–480.
9. Dommermuth, D. G., Rottman, J. W., Innis, G. E., and Novikov, E. A. (2002)
Numerical simulation of the wake of a towed sphere in a weakly stratified fluid, J.
Fluid Mech. 473, 83–101.
10. Fadlun, E. A., Verzicco, R., Orlandi, P., and Mohd-Yusof, J. (2000) Combined
immersed-boundary finite-difference methods for three-dimensional complex flow
simulations. J. Comp. Phys. 161, 35–60.
11. Goldstein, D., Handler, R., and Sirovich, L. (1993) Modeling a no-slip flow
boundary with an external force field, J. Com. Phys. 105, 354.
12. Golub, G. H. and Van Loan, C. F. (1989) Matrix Computations. Johns Hopkins
University Press.
13. Grilli, S. T., Vogelmann, S., and Watts, P. (2002) Development of a 3D numerical
wave tank for modeling tsunami generation by underwater landslides, Eng Analysis
Boundary Elemt. 26(4), 301–313.
14. Harlow, F. H. and Welch, J. E. (1965) Numerical calculation of time-dependent
viscous incompressible flow. Phys. Fluids, 8, 2182–2189.
15. Heinrich, P. (1992) Nonlinear water waves generated by submarine and aerial
landslides. J. Waterw., Port, Coast., Ocean Eng. ASCE, 118(3), 249–266.
16. Hendrikson, K., Shen, L., Yue, D. K. P., Dommermuth, D. G., and Adams, P.
(2003) Simulation of steep breaking waves and spray sheets around a ship: The last
frontier in computational ship hydrodynamics. The Resource Newsletter, U.S. Army
Engineer Research and Development Center Information Technology Laboratory,
Fall, 2–7.
17. Hirt, C. W. and Nichols, B. D. (1981) Volume of fluid (VOF) method for the
dynamics of free boundaries. J. Comp. Phys., 39, 201–225.
18. Issa, R. I. (1986) Solution of implicitly discretized fluid flow equations by operator-
splitting. J. Comput. Phys., 62, 40–65.
19. Kawata, Y., Benson, B. C., Borrero, J. C., Borrero, J. L., Davies, H. L., DE Lange,
W. P., Imamura, F., Letz, H., Nott, J., and Synolakis, C. E. (1999) Tsunami in
Eddy Simulation Model for Tsunami and Runup Generated by Landslides 161
Papua New Guinea was as Intense as First Thought, EOS. Trans. Am. Geophys.
Union 80, 101, 104–105.
20. Jiang, L., and LeBlond, P. H. (1994) Three-dimensional modeling of tsunami
generation due to a submarine landslide, J. Phys. Oceanog., 24, 559–572.
21. Kim, J., Moin P., and Moser R. (1987) Turbulence statistics in fully developed
channel flow at low Reynolds number. J. Fluid Mech., 177, 133–166.
22. Kothe, D. B., Williams, M. W., Lam K. L., Korzewa, D. R., Tubesing, P. K., and
Puckett E. G. (1999) A Second-order accurate, linearity-preserving volume tracking
algorithm, for free surface flows on 3-D unstructured meshes, Proceedings of the
3rd ASME/JSME Joint Fluids Engineering Conference. FEDSM99-7109, July 18–
22.
23. Leonard, A. (1974) Energy cascade in large eddy simulation of turbulent fluid flow.
Adv. Feopgys. 18A, 237–248.
24. Li, C.-W. and Lin, P. (2001) A numerical study of three-dimensional wave
interaction with a square cylinder. Ocean Eng., 28, 1545–1555.
25. Lilly, D. K. (1967) The representation of small-scale turbulence in numerical
simulation experiments. In H. H. Goldstine (Ed.), Proc. IBM Scientific Computing
Symposium on Environmental Sciences, IBM Form No. 320-1951, Yorktown
Heights, New York, pp. 195–210.
26. Lin, P. and Li, C. W. (2002) A σ -coordinate three-dimensional numerical model for
surface wave propagation. Int. J. Numer. Meth. Fluids, 38: 1048–1068.
27. Lin, P. and Li, C. W. (2003) Wave-current interaction with a vertical square
cylinder. Ocean. Eng., 30, 855–876.
28. Lin, P. and Liu, P. L.-F. (1998a) A numerical study of breaking waves in the surf
zone. J. Fluid Mech., 359, 239–264.
29. Lin, P. and Liu, P. L.-F. (1998b) Turbulence transport, vortices dynamics, and
solute mixing under plunging breaking waves in surf zone. J. Geophys. Res.,
103(C8), 15677–15694.
30. Liu, P. L.-F., Lin, P. Z., and Chang, K. A. (2001) Numerical modeling of wave
interaction with porous structures – Closure, J. Waterw., Port, Coast., and Ocean
Eng., ASCE 127(2): 124–124.
31. Liu, P. L.-F., Lin P. Z. and Chang, K. A. (1999) Numerical modeling of wave
interaction with porous structures, J. Waterw., Port, Coast., and Ocean Eng., ASCE
125(6): 322–330.
32. Liu, P. L.-F., Wu, T.-R., Raichlen, F., Synolakis, C. E., and Borrero, J. C. (2005)
Runup and rundown generated by three-dimensional sliding masses, J. Fluid
Mechanics 536: 107–144.
33. Lynett, P., Wu, T.-R. and Lim P. L.-F. (2002) Modeling Wave Runup with Depth-
Integrated Equations, Coastal Engineering, 46(2), 89–107.
34. Mohd-Yusof, J. (1997) Combined immersed boundaries/B-splines methods for
simulations of flows in complex geometries, CTR Annual Research Briefs, NASA
Ames/Stanford University.
162 T.-R. Wu and P. L.-F. Liu
CHAPTER 5
J. B. Frandsen
School of Civil Engineering
The University of Sydney, NSW 2006, Australia
Email: [email protected]
163
August 26, 2008 17:6 World Scientific Review Volume - 9in x 6in 05˙main
164 J. B. Frandsen
There are several kinetic or mesoscopic methods, e.g. the lattice gas cellu-
lar, the lattice Boltzmann equation, the gas kinetic schemes, the smoothed
particle hydrodynamics, and the dissipative particle dynamics method. The
Lattice Boltzmann (LB) method with roots in the Boltzmann equation
stems from the ideas of Ludwig Boltzmann (1844–1906) who made advances
in electromagnetism and thermodynamics. We seek to utilize his gas dy-
namics theory (Boltzmann, 1964). Lattice gas models and LB models have
been used in fluid mechanics since the 1990s. About 1000 articles have been
published over the last fourteen years. The majority of these publications
are related to physics and computer sciences (e.g. soft condensed matter,
porous media, microfluidics, electrokinetic flows, magnetohydrodynamics,
etc.). The applications are broad. Sukop and Thorne (2006) have pro-
vided an overview of a number of papers published (1992–2004), divided
into various areas of science, in the introductory chapter of their recent
book.
The purpose of this chapter is to introduce the LB Equations (LBE),
the simplified Boltzmann Equation, with application to free-surface water
waves. Applying the LB approach will allow us to solve the fluid flow at
the mesoscopic scale level, i.e. length scales ranging between the atomistic
or micro and the continuum mechanics level (assuming the grid refinement
is adequate). The latter is hereafter referred to as the macro scale level.
Somewhere between the atomistic and macro scale level (Fig. 1(a)), the
continuum model approach breaks down where part of the physical sys-
tems cannot be assumed to be continuous (e.g. breaking waves). This is
one of the main motivations behind exploring the LB approach. Herein, we
shall, however, focus on single phase LB models only, as this is our first ex-
perience with LB modeling. We have highlighted some questions which we
asked ourselves before we chose to explore the LB approach. Some of these
modeling features and issues are shown in Fig. 1(b). We have further indi-
cated our progress on what we have/have not tested. The first important
question to ask was whether or not we could predict nonlinearities at the
free-surface. We have got some promising indication that the LB method
may work, as shown in the test case studies in §6. Our investigations have
so far been limited to shallow water depths. We do not yet know how accu-
rate the method would be regarding wave breaking predictions. One feature
which motivated us to chose the LB approach is that we can model bub-
ble break-up. Regarding Fluid-Structure interaction (FSI) we have so far
August 26, 2008 17:6 World Scientific Review Volume - 9in x 6in 05˙main
Newton Lattice
Gas
Liouville Lattice
Liouville
Boltzmann Lattice
Boltzmann
found that the LB approach can provide a relatively easy means of simulat-
ing flow around single and multiple bluff bodies. The standard LB method
is based on uniform lattices. This presents the usual difficulties as to grid
refinement requirement, and thus yields a host of questions on accuracy
and efficiency. This is on-going research, as described in the review part of
this chapter.
Further, we should point out that our intention with the LB modeling
approach is to use it in local areas where physical details are needed such
as the nearshore regions and in the breaking wave process itself. Therefore
a realistic goal and potential improvement to the current tsunami model
literature would be to propose an LB model coupled with a macro-scale
level model for large domain tsunami model predictions.
The key issues to bear in mind when reading through this chapter
are that the standard LBE scheme/approach has the following model
properties:
• the fluid mass is collected in discrete particles, located on corner
points of the lattice site;
• the velocity in a lattice node is discretized (in addition to time and
space);
• the particles move according to a finite, discrete set of velocities;
• the viscosity (ν) is introduced via the single time relaxation param-
eter (τ ) and is a discretized equation with uniform lattice spacing
2
∆x and time step ∆t, ν = c3 (τ − ∆t 2 ) where c = ∆x/∆t. In the
incompressible flow simulation herein, we must ensure that ν is
constant;
August 26, 2008 17:6 World Scientific Review Volume - 9in x 6in 05˙main
166 J. B. Frandsen
From our viewpoint, the LB model is mainly proposed due to the molecular
properties of the formulation and its ease of handling large data sets in rela-
tion to understanding underlying mechanisms of breaking ocean waves. As
outlined above, there are of course several other important reasons an LB
modeling approach appears to be a promising simulation method in fluid dy-
namics. Some of the numerical model advantages of the LB method are: (1)
no Poisson equation, and (2) a reduction from second-order to first-order
partial differential equations, and therefore a simplification of nonlinear
free-surface numerical modeling. However, as mentioned, the major motiva-
tion of the present work is driven by the mathematical framework allowing
molecular level modeling of fluid flow which differs from any conventional
free-surface models. The locality of the formulation is another advantage,
essential for CPU efficiency and thus practicality. As for most fluid dynam-
ics problems, high CPU requirement is a barrier in advancing free-surface
models which typically involve high CPU time near the moving boundary,
especially when discontinuities occur. The LB method has been demon-
strated to show good parallel computing performance (e.g. Aidun et al.
(1998); Thürey et al. (2006a); Körner et al. (2006)). Furthermore, com-
pared to traditional continuum mechanics solvers, the LB solver has also
shown promise when simulating fluid interaction with multiple bluff-bodies
in terms of low CPU time and an easy means of rearranging bluff-bodies to
study optimum solutions for flow-induced vibration problems with/without
August 26, 2008 17:6 World Scientific Review Volume - 9in x 6in 05˙main
168 J. B. Frandsen
This brief review is meant to provide the reader with insight into the work
and progress on development and applications of LBE, in particular, with
flow around bluff-bodies and free-surface flows in mind. We emphasize that
this approach has not been applied and validated in the area of free-surface
ocean wave predictions. Therefore, a review in a more general format is
presented. We outline the development of various branches of LB modeling.
In doing so we highlight ideas of expansion and development of LB-based
approaches relevant to the ocean and coastal engineering communities.
The LB formulation offers an alternative treatment for description of
fluid particle motion. It is based on statistical mechanics concepts and
recovers the Navier-Stokes equations in the nearly incompressible limit, as
mentioned by Chen and Doolen (1998). The approach has been used exten-
sively in molecular dynamics and physics but little attention has been given
to the modeling of free-surface water waves. The LB method was devel-
oped as an improvement of the method of Lattice Gas Automata (LGA),
as described by Rothman and Zaleski (1997). It was first introduced by
McNamara and Zanetti (1988). Comprehensive reviews of the LB method
have been given (Benzi et al. (1992); Chen and Doolen (1998); Nourgaliev
et al. (2003)). The LB method belongs to a class of the pseudocom-
pressible solvers of the Navier-Stokes equations and can be classified as a
Lagrangian, local equilibrium, finite-hyperbolicity approximation. The nu-
merical formulation typically (but not necessarily) separates the equations
into the dissipative and non-dissipative parts. The most popular form of LB
equations is the Lattice-BGK (LBGK) incorporating a single-time approxi-
mation of the Boltzmann equation, as described by Bhatnagar et al. (1954),
hereafter the LBGK model. In this model the mean free path between par-
ticles is assumed to be the same. It should also be noted that the governing
LB formulation is second-order accurate space and first-order accurate in
time, as shown by Junk et al. (2005). The disadvantages inherent in the
LBGK model can be reduced by using a Multi-Relaxation-Time (MRT)
approach. It separates the relaxation times for different kinetic modes and
improves numerical stability and accuracy. For more detailed discussions
August 26, 2008 17:6 World Scientific Review Volume - 9in x 6in 05˙main
The majority of LBE model predictions are based on uniform regular spaced
lattices. Compared to traditional CFD, few initiatives have been done on
unstructured LB lattices. Although an increase of contributions is evident,
a literature review still reveals only relatively few publications on this topic,
e.g. He et al. (1996); Tölke et al. (1998); Peng et al. (1999); Kandhai et al.
(2000); Van der Sman and Ernst (2000). Mesh refinement techniques are
also discussed by Crouse et al. (2002); Ubertini et al. (2003); Crouse et al.
(2003). Recent progress on grid refinement method using a nested adap-
tive grid approach are described by Tölke et al. (2006), in which numerical
examples with rising bubble simulations are presented. Geller et al. (2006)
simulates flow around multiple cylinders based on unstructured grids at
Re=200. Good agreement was found with traditional CFD solvers. Several
investigators have made contributions regarding the treatment of curved
boundaries on uniform grids in 2-D, e.g. Filippova and Hänel (1997); Mei
et al. (1999). Expansion into 3-D flows have also been developed, as de-
scribed by Mei et al. (2000). We should also mention that “stair-step”
effects on the boundary can be overcome by using mapping techniques,
like in traditional finite difference models. For example, He and Doolen
(1997) developed a curvilinear formulation to simulate flow past a circular
cylinder.
Today, the research group directed by Professor Krafczyk at the Insti-
tute for Computer Applications in Civil Engineering, Technical University
of Braunschweig, Germany, is at the forefront on LB-grid method devel-
opment. We should also mention that they are in the forefront in many
other LB developments as well, including contributions to civil/mechanical
engineering applications. The reader is invited to visit their web site which
has many publications available for downloading including a nice article
search engine (www.cab.bau.tu-bs.de).
With respect to LB-model development with application to steep and
breaking free-surface water waves, it would be necessary to have models
August 26, 2008 17:6 World Scientific Review Volume - 9in x 6in 05˙main
170 J. B. Frandsen
with non-uniform grids, just like it is when using traditional CFD finite
volume and finite element solvers. In contrast to the macro mechanics
models, it should be noted that non-uniform grid implementation in the
LB context is non-trivial.
Studying papers in these volumes alone would bring the reader quickly up
to date in the area of Lattice Boltzmann modeling (we have referenced sev-
eral articles herein). We should also mention that fundamentals on LBE
August 26, 2008 17:6 World Scientific Review Volume - 9in x 6in 05˙main
172 J. B. Frandsen
similarity between the equations for shallow water theory and the funda-
mental differential equations of gas dynamics (see the references listed in
§4.7). We shall later introduce kinetic theory which forms the background
of LB models (§4.1).
In the context of kinetic modeling of water waves, the approximation of
the collision integral in the Boltzmann equation becomes very important.
We should say that the LBGK approximation is the crudest approximation
of the possibilities the LB model series would offer. Since, the LB free-
surface water wave research is not an established area, the LBGK approach,
the simplest form, would however be a natural starting point for model
development. As far as we know, there exists no attempt yet which includes
a LB Navier-Stokes formulation for non-overturning, as well as breaking
waves, in shallow water depth which has undergone validation (the same can
be said for LB deep water predictions which also has not yet been explored).
We believe that this has a simple explanation, that is, LB investigators come
with very different backgrounds and interests, not to mention the numerical
approach is a relative new research area.
The first attempt on LB non-overturning free-surface flows was under-
taken by Salmon (1999a). The single phase LBGK model of Salmon obeys
the shallow water equations in rotating flows. Salmon’s interest was to use
the LB model for 3-D ocean circulation models [Salmon (1999b)]. There-
fore, a relatively large flow domain in the order of 4000×4000 km2 was
prescribed. It is also notable that the numerical results were based on rel-
atively coarse models of 100×100 lattice points. Dellar (2002) undertakes
a critical review of the stability of the shallow water LBE. Zhong et al.
(2005) is also interested in a large scale rotational flow field with applica-
tion to ocean circulation. They expanded the model of Salmon (1999a) to
include a fully explicit second-order accurate model. Ghidaoui et al. (2001)
and Deng et al. (2001) developed a finite volume BGK model for open chan-
nel flows and contaminant flows, respectively. Xu (2002) investigated the
performance of a BGK scheme, especially the treatment of shocks. Sim-
ilar to the formulation of Salmon (1999a), Zhou (2002) explored LBGK
solutions including an eddy viscosity model which represented shallow wa-
ter equations with rotating flows. Zhou (2004) accumulated many of his
investigations in book format with a focus on flow effects in channels on
rough beds with application to river mechanics problems.
Recent progress in development and applications are by Que and
Xu (2006); Ghidaoui et al. (2006); Frandsen (2006a); Thömmes et al.
(2007). Thömmes et al. (2007) applied their LBGK solver to complex
August 26, 2008 17:6 World Scientific Review Volume - 9in x 6in 05˙main
174 J. B. Frandsen
176 J. B. Frandsen
LB model approach and our attraction to this approach is that one can
take the collision approach much further to include soft sphere modeling,
as highlighted in Fig. 1(b). For example, Professor Yeoman and co-workers
(Oxford, Theoretical Physics) have undertaken work on this aspect of LB
modeling in the general area of binary fluids, wetting/spreading, and the ap-
plication of droplet impact predictions (funded by the inkjet/laserjet printer
industry). The group has also made advances made in liquid crystal dy-
namics, e.g. Denniston et al. (2004). Abraham and co-workers (Purdue
University, Mechanical Engineering) have conducted a variety of bubble in-
teraction dynamics experiments, e.g. McCracken (2004); Premnath (2004);
Premnath and Abraham (2005, 2007). These experiences are useful start-
ing points for the free-surface breaking-wave application. Furthermore,
Professor Aidun and co-workers (Georgia Tech, Mechanical Engineering)
have simulated stokes flow with deformable particles amongst a long list of
contributions in the area of suspended particles hydrodynamics, e.g. Ding
and Aidun (2003). Some of the fundamental research have been applied
to biomedical problems, e.g. Ding and Aidun (2006). Recently, Aidun
et al. have introduced an LB-fluid model with deformable finite-elements
particles to simulate mesoscale blood flow. The work, in preparation for
publication, was presented at the 2006 meeting of the American Physical
Society, Division of Fluid Dynamics.
Finally we note that it is still an unresolved LB problem as to how to
tackle the high density ratio difference at the free-surface interface in two
phase flow in ocean and costal engineering applications. Some work on
this has been initiated in relation to bubble and droplets, as described by,
e.g. Inamuro et al. (2004); Lee and Lin (2005).
These LB experiences mentioned in this section, although not obvious at
first glance, are all very useful starting points for the free-surface breaking-
wave application in ocean and coastal engineering.
(2006a). The LBGK solver did not reveal any sign of numerical dissipa-
tion but an excessive resolution was required to simulate the weak jump at
the surface. We have done some further stability assessment through the
wave runup studies herein; see benchmark test case 1 reported in Chapter
16 in this volume. These studies suggest that high-order finite difference
LB schemes improve the instability problems typically found in the evo-
lution of the velocity components. Having said this, we cannot do direct
comparisons between the bore studies and wave runup problems, as the
latter test case is governed and influenced by the treatment of the wet-dry
interface.
Our observations based on the current progress do suggest advantages
when using high-order finite-difference LB schemes. Some of our LB ex-
periences indicate that this kind of LB model should be a prerequisite for
breaking LB wave models. In our free-surface review, it was notable that
some investigators have applied the standard LBE with turbulence models
and a breaking-wave algorithm (VOF). Some investigators have used the
means of turbulence models to suppress instabilities. Some LB solvers have
also gone to the extent of tuning the viscosity, i.e. so that the fluid is not
a water representation. From a numerical water wave mechanics develop-
ment view point, none of these are appropriate steps to take. Although
the above-mentioned research forms some of the background necessary for
the development of the LB free-surface framework, it is still uncertain if
an LB approach can simulate water wave behavior accurately. Little re-
search on water waves using LB modeling have been undertaken. The LB
research is very much in the beginning of entering the group of numerical
solver options available to ocean and coastal ocean engineering applications.
On a final note, we conclude that it has not yet been demonstrated that
non-overturning water waves based on standard LBE formulations or other
LB models are an adequate means of describing free-surface water wave
behavior.
178 J. B. Frandsen
The LB Equation (LBE) originates from the kinetic theory of gases and is
a minimal form of the Boltzmann kinetic equation in which all the details
of molecular motion are removed, except those that are strictly needed
to recover hydrodynamic behavior at the macroscopic level such as mass
and momentum conservation on sufficient resolution of the lattices (Succi,
2001). The LB method discretizes kinetic theory.
Kinetic theory assumes that the fluid is described by a large number
of molecular constituents whose motions obey Newtonian mechanics. The
objective is not to know the motion of every individual molecule, but the
collective behavior for which we need a statistical description of the system,
as illustrated in Fig. 3.
i 0 1 2 3 4 5 6 7 8
ci,x 0 +c 0 -c 0 +c -c -c +c
ci,y 0 0 +c 0 -c +c +c -c -c
(lattice) particles. The particles are by default located at the corner nodes
of a regular lattice. The particles move according to a finite, discrete set
of velocities. The standard LB model notation follows a Dn Qm reference
where n is dimension and m denotes number of particles. A simple example
is the 1-D diffusion model or following the notation a D1 Q2 lattice model.
A 2-D convection-diffusion model would be referred to as a D2 Q9 lattice
model and the 3-D counter part as a D3 Q19 lattice model. The velocity
components of the D2 Q9 model are illustrated in Fig. 4.
17
14 6
13
2 10 9
6 5 2
18
0
4 3
16
3 0 1 k 1
7 8
11 12
γ 5
y z β j
7
4 8 y α 15
x (a) x i (b)
where the lattice velocity c = ∆x/∆t, ∆x is the lattice size (uniform) and
∆t is the duration of the time step. The rest particle corresponds to i=0
(ci=0 = 0) and others represent lattice vectors in the direction of the nearest
neighbors. The complete velocity set of a D2 Q9 model is listed in Table 1.
August 26, 2008 17:6 World Scientific Review Volume - 9in x 6in 05˙main
180 J. B. Frandsen
The Boltzmann equation relates the time evolution and spatial variation of
a collection of molecules to a collision operator that describes the interaction
of the molecules,
∂f 1
+ ci · ∇f = J(f, f ) , (2)
∂t
π 2π
1
Z Z Z
J(f, f ) = dφ dΦ dV B(φ, V ) (f10 f20 − f1 f2 ) , (3)
2m 0 0
2 r0 f1
2
+c
f2
1 f1 −c
+c
φ V
1
r0 Φ
−c
f2
(a) (b)
182 J. B. Frandsen
184 J. B. Frandsen
i=0
8
ci fieq = ρ u ;
X
(11)
i=0
8
1 2
ci cj fieq =
X
g ρ δij + ρ u u ,
i=0
2
where g is the acceleration of gravity. The density in (11) becomes the total
water depth h in the shallow water applications, as shown later. Satisfying
the equilibrium distribution functions is essential in obtaining valid physics.
Note that the form of the functions would be dependent on the application
of interest. For example, the deep water formulation would have different
functions compared to the shallow water equation counter-part. These
functions would typically have to be derived through Taylor expansions
and Chapman-Enskog analysis from which it can be shown that the Navier-
Stokes equation can be derived from the Boltzmann equation. This kind of
analysis will be described in the following §4.4.
August 26, 2008 17:6 World Scientific Review Volume - 9in x 6in 05˙main
There are mainly two ways to truncate an infinite series of transport equa-
tions. As mentioned, one technique is based on the Chapman-Enskog
method and another is introduced by Grad (1949). Herein we highlight
the former method. Chapman (1888–1970) in England and Enskog (1884–
1947) in Sweden described independently the general solution to the Boltz-
mann equation in a series of papers published 1911–21, e.g. Chapman (1916,
1918); Enskog (1921). The primary use is the calculation of transport
equations. As we know these quantities are unknown in the macroscopic
theories. The Chapman-Enskog procedure considers an expansion of the
equations for the moments of fi and assumes that the time dependence of
fi occurs through mass density, velocity and temperature. The Chapman-
Enskog expansion corresponds to a multi-scale expansion to first-order in
space and second-order in time and assumes that the diffusion time scale
is much slower than the convection time scale.
Below we shall briefly demonstrate that the Boltzmann equation (2) sat-
isfies the Navier-Stokes equation using the second-order Chapman-Enskog
expansion technique. We assume the moments of the equilibrium distri-
bution is equal to the Maxwell-Boltzmann distribution. The analysis is
performed for the Lattice BGK scheme. The Chapman-Enskog expansion
parameter is the Knudsen number (). Our definition of the coordinate
system and the subscript indices follow the velocity set of the LB cube, as
shown in Fig. 4(b).
To perform the Chapman-Enskog expansion, a Taylor expansion of the
discretized Boltzmann equation (6) is first undertaken,
1 1
fi (~x + ∆~x, t + ∆t) ≈ 1 + D + D2 + D3 + ..... fi (~x, t) . (12)
2 6
186 J. B. Frandsen
188 J. B. Frandsen
∆x
∆y 6 2 5
∆y /2 8 4 2
7
Wall
6 5
Ghost cells
∆x /2 ∆x
8 2 5
∆y 7
6 6
3 1
∆y /2 1 7
7 Wall
5 2 6 8
4 8
Ghost cells
(a) (b)
Fig. 6. Free slip boundary conditions at wall and corner (a) before and (b) after
streaming.
August 26, 2008 17:6 World Scientific Review Volume - 9in x 6in 05˙main
190 J. B. Frandsen
Junk et al. (2005) undertook a detailed analysis of the LBE. They related
the LBE to the finite discrete-velocity model of the Boltzmann equation
with diffusive scaling and obtained the incompressible Navier-Stokes equa-
tions as opposed to the compressible Navier-Stokes equations obtained by
the Chapman-Enskog analysis with convective scaling. Junk et al. proved
that the LBE is second-order accurate in space. For the incompressible
Navier-Stokes, they also proved that the LBE is first-order accurate in
time asymptotically. Gou et al. (in press) also recently contributed to this
concern. Through unsteady flow predictions, Guo et al. observed that LBE
is second-order accurate in time.
Further, we note that the standard bounce-back condition, the most
applied boundary condition, is first-order in numerical accuracy at the
boundaries. This, of course, degrades the LB method. As mentioned in
the review, several investigators have contributed to second-order accurate
boundary conditions. New contributions to the LBM literature are for this
reason (and others) still on-going research.
In this chapter, we present results based on the standard LBE. The wave
runup test case, presented in Chapter 16 in this volume, has undergone fur-
ther tests including high-order finite-discretization schemes, thus improving
the standard LBE. We should mention that it did not originally occur to in-
vestigators that the LBE could be expanded this way (like traditional CFD
methods). It is still not common to consider a particular discretization for
the discrete Boltzmann equation. However, some progress on high-order
LB discretization schemes have been reported. Cao et al. (1997) were the
first investigators to view the LBM as a special finite difference (FD) dis-
cretization. Further elaborations on this topic have been undertaken since,
e.g. Kandhai et al. (2000); Junk (2001); Seta and Takahashi (2002); Sofonea
and Sekerka (2003); Van der Sman (2006). Some Finite Volume (FV) LB
models have also been developed, e.g. Ghidaoui et al. (2006); Que and Xu
(2006). As far was we are aware of, there has been no comparison under-
taken for FDLB and FVLB model, so the difference in solutions accuracy
is not clear to us. Furthermore, we should also mention that various dis-
cretization schemes for gas dynamics computations are described in details
in the book by Laney (1998).
Fig. 7. Kinetic modeling approach to water wave predictions with roots in the gas
dynamics theory of Boltzmann. Comments and equations derivations on this idea can
be found in the listed summary of references. (The photograph is a wave runup snapshot
from the Northshore of Oahu, Hawai’i.)
192 J. B. Frandsen
We shall in the remaining part of the chapter focus on applying and testing
a Lattice Boltzmann modeling approach to free-surface water waves. We
assume that the collisions of particles are represented through the single
time relaxation scheme (LBGK approximation). It should be noted that
there is no deeper reason for starting the Boltzmann implementation this
way other than this is the first logical modeling development step. How-
ever, this is not necessarily the most accurate and stable approach due
to the BGK collision approximation. Our LB model contains equilibrium
distribution functions which are based on particle motions in shallow wa-
ter and therefore our test cases are limited to shallow water depths. The
discretized equations are based on viscous free-surface waves. Therefore,
flows with horizontal velocity dominating free-surface waves is applicable
and is our focus only. Furthermore, the present LBGK model explores the
free-surface behavior of non-overturning water waves on uniform grids. For
validation purposes, we first investigate the dynamics of viscous liquid and
water waves in tanks with flat and slope beds; in 1-D and 2-D, respectively.
Then, we test the tsunami runup on sloped beaches (workshop benchmark
test 1). The latter test case is presented in Chapter 16 in this volume.
We consider the flow of water with a free-surface under gravity in a 3-D
domain where x − y denotes a horizontal plane while z defines the vertical
direction. The free-surface elevation coincides with the z-axis. Assuming
that the vertical component is negligible in the field of interest, the fluid
motion can be described by the shallow water equations with rotating flows,
∂h ∂(h ux ) ∂(h uy )
+ + = 0;
∂t ∂x ∂y
∂ h2 ∂ ∂ux ∂ ∂ux
= −g + hν + hν + Fx ; (27)
∂x 2 ∂x ∂x ∂y ∂y
∂ h2 ∂ ∂uy ∂ ∂uy
= −g + hν + hν + Fy ,
∂y 2 ∂x ∂x ∂y ∂y
August 26, 2008 17:6 World Scientific Review Volume - 9in x 6in 05˙main
where h = h0 + ζ, h0 is the still water depth (or initial water depth) and ζ
denotes the free-surface elevation measured vertically above still water level,
t is time, g = 9.81 m/s2 denotes the acceleration of gravity, ρ = 1000 kg/m3
is the water density, ν = 1 × 10−6 m2 /s is the kinematic viscosity, ux , uy
are the horizontal depth-averaged velocity components, and F denotes force
terms.
Equivalent to (27), the proposed single phase LBGK formulation for
shallow liquid rotating flows can be written as
2
∂fi ∂fi (fi − fieq ) 1 1 X
+ ci =− + c i Fi , where Ni = ci ci ,
∂t ∂x τ Ni c 2 c2 i=0
(28)
∆t ∆t
fi (~x + ∆~x, t + ∆t) − fi (~x, t) = − fi (~x, t) − fieq (~x, t) + c i Fi .
τ Ni c 2
(29)
194 J. B. Frandsen
196 J. B. Frandsen
6. Case Studies
180
180
170
170
160 160
150
y
150
y
140 140
130 130
120 120
180 200 220 240 260 180 200 220 240 260
x x
Fig. 10. LBGK flow past circular cylinder. Velocity snapshots (Re = 200).
Frandsen (2006b). This particular article also showed some initial at-
tempts on predictions of free-surface bluff-body interactions (e.g. Fig. 2).
These studies included single and multiple bluff-bodies studies at Re ≤ 250
(Fig. 9). In general, we found that the LB model provided an easy means of
handling multiple bluff-bodies with sharp edges. We are currently modeling
flow around circular cylinders (Fig. 10). The curved boundary is modeled
through the second-order accurate boundary condition proposed by Filip-
pova and Hänel (1997). Good agreement with other investigators have been
achieved for Re ≤ 250. We have also implemented a curvilinear formula-
tion, as developed by He and Doolen (1997). This approach seems to work
equally well for flow around circular cylinder.
The free-surface test cases, as illustrated in Fig. 11, have so far in-
cluded (1) weak bore simulations; (2) simulations of wave interactions with
surface-piercing bodies; (3) dam-break predictions; (4) force term assess-
ment (e.g. beach slope); and (5) wave runup. We have done extensive
studies of bore predictions (without bluff-bodies) and wave runup.
We shall in the following, present some of the bore studies and dam-
break predictions. The wave runup studies follow the description of the
benchmark test 1 set-up problem, which is reported in Chapter 16 in this
volume.
198 J. B. Frandsen
model in the horizontal plane (x-y). The bounce-back scheme for slip walls
have been prescribed. We should emphasize that the solutions in this first
test series are based on highly viscous liquid. This is not only motivated
by our comparison source (Wu et al (2001)) but also because we were at
first not certain if we could simulate free-surface behavior at all, including
both highly viscous liquid and water. Therefore, we began our studies with
highly viscous liquid.
The numerical predictions are compared with the linearized Navier-
Stokes solution derived by Wu et al. (2001) who showed that the wave
elevation at an arbitrary depth is given by
ζ0 g k2 tanh(k2 h0 ) tanh(α2 h0 ) λ t
Z
ζ = ζ0 + e dλ , (33)
2π ν i Γ C2
p
where k2 = 2π/b is the wave number for n2 and αn=2 = kn2 + λ/ν,
where λ is a complex number and b denotes the width of the tank. The
denominator in (33) is complex and is presented in detail in the article of
Wu et al. (an analytical solution exists for λ = 0 only). The deep water
limit of (33) is explored by Wu et al. (2001) for Re = 2, 20, 200, with
√
Re = h0 g h0 /ν, in a rectangular tank with an aspect ratio h0 /b = 0.5.
In this paper, it is the shallow liquid limit of (33) which is used in the
August 26, 2008 17:6 World Scientific Review Volume - 9in x 6in 05˙main
ζ0 g k22 h0 ν eλ t
Z
ζ = ζ0 + × eλ t dλ . (35)
2π ν i Γ −λ3 − 16π 2 νλ2 + g 4π 2 h0 λ
In this form, the denominator has three poles in the complex plane which
allows the integral to be calculated analytically. The integration path Γ is
the path for calculating inverse Laplace transforms. Letting ωn2 = g kn2 h0
where ωn represents the sloshing frequency in shallow liquid, (35) reduces
to
ζ0 ωn2 −ν kn2 t
ζ = ζ0 + e f (ν kn2 t) , (36)
kn2 ν
where f (ν kn2 t) is the inverse Laplace transform which solution can be found
in Poularikas (1996). The derivation leads to following solution for the
viscous time evolution of the free-surface in shallow liquid,
ζ0 h
ζ= p 1 − kn2 ν
2 4 − ωn2 /(kn2 ν)2
√
2 /(k2 ν)2 ν k2 t
p −2− 4−ωn (37)
− k 2 ν 2 + 4 − ω 2 /(k 2 ν)2 e
n n n
n n
√
2 /(k2 ν)2 ν k2 t
i
p −2+ 4−ωn
+ kn2 ν 2 − 4 − ωn2 /(kn2 ν)2 e n n
.
Initially the velocity in the flow domain is zero, and the free-surface
ζ0 (x, n) = a cos(kn x), where a is the amplitude of the initial wave profile
and x is the horizontal distance from the left wall. Nonlinear free-surface
motions are investigated by varying the wave steepness, defined herein as
= a ωn2 /g and is a measure of non-linearity at the free-surface in the
fixed-tank studies. The results presented are for a tank of aspect ratio h0 /b
= 0.05 and for Re = 20, 200 with corresponding ν = 1.75×10−3 m2 s−1 ,
1.75×10−4 m2 s−1 . The Re = 20 corresponds to glycerine type of liquid
whereas Re = 200 is oil-based liquid (SAE 20W-20). The time histories of
the free-sloshing motions are presented in non-dimensional form using the
sloshing frequency ωn and the initial amplitude a.
August 26, 2008 17:6 World Scientific Review Volume - 9in x 6in 05˙main
200 J. B. Frandsen
1 1
ε = 0.1972 ε = 0.1972
ε = 0.0296 ε = 0.0296
ε = 0.002 ε = 0.002
Wu et al. Wu et al.
ζ/a
ζ/a
0 0
−1 −1
0 20 40 0 25 50
txω t x ω2
2
Figure 12 shows the time histories of the free-surface elevation for the
second sloshing mode (n = 2) for the liquids with Re = 20 and 200, re-
spectively. Each of these Re solutions are based on a grid resolution of
100 and 500 nodes, respectively. Solutions are presented for time steps ∆t
= 0.01 s, 0.002 s and ∆x = 0.01 m, 0.002 m yielding τ = 1.025 for low-
est grid density/Re case and τ = 0.7625 for highest grid density/Re case.
The increase of nodes and associated reduced time step were necessary in
order to maintain accuracy of the explicit solver for higher Re. The free-
surface time histories at the left wall are shown for small to moderate waves
( = 0.002, 0.197). The Re = 20 result shows one oscillation cycle only.
Hereafter, due to the high viscosity, the amplitude decays to zero. This is
also true for all of the other wave steepness cases considered shown. The
small initial wave = 0.002 case agrees fairly well with the linear solution
of Wu et al. The increasing wave steepness introduces a phase-shift with
larger peaks and lower troughs. The higher Re = 200 test case generates
larger wave steepness (k2 × ζmax = 0.43) with associated amplitude skew-
ness and an increase in oscillation cycles, which eventually decays toward
zero, representing the physics well. We should point out that the skewness
of the amplitudes is the characteristic of a bore as opposed to non-skewed
amplitudes of sloshing motion. Bores are known to form at shallow liquid
depths for a relatively large forcing frequency. The next validation study
is on water bores.
Sway
b/2
y
3 x Surge
b /4
2
1m
b /4
1m
1
l /2 l /2
Fig. 13. Shake table with shallow water sloshing in a square tank with base dimensions
of 1×1 m2 . Definition of excitation modes and plan view of tank with wave gauge
locations (1–3). (The photograph is from the laboratory of Frandsen located at Louisiana
State University, USA.)
202 J. B. Frandsen
z
ζ U0
h0
uy
h1
α
h2
y
b
capture off- and near-resonance free-surface water waves while the forc-
ing amplitude are ah /b = 0.006 or ah /b = 0.02. Initially, the free-surface
elevation and the velocity field are prescribed to be zero. The non-
dimensional horizontal forcing parameter κh = ah ωh2 0 /g and the Froude
√
number F r = U0 / g h1 (where U0 is the propagating speed and h1 is the
developed water depth) are used as a measure of non-linearity and bore
strength, respectively.
The numerical results are compared with experimental data, as de-
scribed elsewhere (Frandsen (2006a)). We also compare the LB results
with a low-order Riemann solution. Herein we use a first-order upwind
method with simple linear interpolations, as described by Toro (1999). We
realize this is a shortcoming in our comparison study as it is well known that
schemes which are first-order accurate in space and time can produce arti-
ficial viscosity. Indeed, this was experienced. Recognized shock-capturing
schemes typically contain high-order interpolations with second-order accu-
racy methods including limiters. Extensive theories of these schemes exist
in terms of flux functions (e.g. Roe, Godunov, Van Leer, Steger-Warming,
etc.), reconstruction methods and limiters, however, this will not be fur-
ther discussed here. Instead, the reader is referred to the comprehensive
descriptions of the host of Riemann solvers in the books by Toro (2001);
LeVeque (2002); and Chapter 2 in this volume by LeVeque.
All results are presented in dimensionless form. The free-surface (ζ) is
non-dimensionalized by the still water depth (h0 ) and the non-dimensional
time T = ω1 t and the non-dimensional time step ∆T = ω1 ∆t where the
first initial
p linear natural sloshing frequency in shallow water is defined as
ω1 = g (π/b)2 h0 . The first numerical tests carried out were designed to
check the sensitivity of the LBGK scheme to the time step and the grid res-
olution (Frandsen, 2006a). The 1-D simulations are based on D1 Q3 lattices.
The bore formation was found to occur at Fr1 =0.93, showing two distinct
developed water depths h1 /h2 = 1.34. All grid resolutions used yielded
August 26, 2008 17:6 World Scientific Review Volume - 9in x 6in 05˙main
0.4 0.2
T =128.80 T =129.43 T
6 7 8
T =130.06
8
T =130.69
9
0
ζ/h
ζ/h0
−0.4 −0.15
−0.5 0 0.5 −0.086 −0.085
y/b y/b
Fig. 15. LB wave profiles (Fr1 = 0.93; h1 /h2 = 1.34) near resonance (ωh /ω1 = 1; κh =
0.003). Grid resolution, nodes: −−, 20,000 (τ =0.547); − · −, 30,000 (τ =0.578); − + −,
40,000 (τ =0.591); − ◦ −, 50,000 (τ =0.643); −, 60,000 (τ =0.553).
similar wave profiles and thus represent grid converged solutions except at
the bore tip, as shown in Fig. 15. Higher grid resolution is required for
the near resonance cases compared to outside resonance, as expected. It
was found that the LB solutions for κh = 0.003 required a minimum of
about 10,000 nodes outside resonance and 20,000 nodes near resonance,
respectively, to capture the physics experienced experimentally. However,
in order to represent the bore tip adequately, this would not yield accurate
solutions. The LBGK and Riemann solutions agree fairly well in terms of
the free-surface elevation and velocity predictions. However, a phase lag
exists which appears to be more pronounced in the wave profiles than in
the time evolutions at various locations in the tank. The phase lag is caused
by numerical viscosity related to the first-order upwind scheme of the Rie-
mann solution. The low-order Riemann solution produced a slower bore
propagation compared to the LBGK bores. We found that, to completely
suppress the bore tip oscillations, a 60,000 node resolution for the standard
LBGK model would be required (Fig. 15). The first-order Riemann solution
would require about a four times smaller time step to achieve equivalent
accuracy. Further, it should be noted that the reason for the requirement of
the high resolution is due to the vertical jump in the surface. Lesser resolu-
tion would yield a sloped jump. Moreover, the Riemann solution with 2,000
nodes would not be able to capture a bore whereas the LB solution would.
In general, the low-order Riemann solution would have a tendency to pre-
dict sloped jumps whereas the LB solution would have vertical jumps for
the same grid resolution (analogous to conservative versus non-conservative
methods). However, the LB bore front showed numerical oscillations at the
tip (Fig. 15) which could be suppressed with an increase in resolution, or
August 26, 2008 17:6 World Scientific Review Volume - 9in x 6in 05˙main
204 J. B. Frandsen
0.6
a h /b = 0.02
ζmax/h0
a h /b = 0.006
0
0.8 1.0 1.2 1.4
ωh/ω1
Fig. 16. Maximum free-surface elevation at the wave gauge 2 (x, y) = (0, -b/4) location.
−, LB solution; −−, Experimental solution. •, time evolution comparison.
0.3 0.3
ζ / h0
ζ / h0
0 0
−0.3 −0.3
0 15 30 0 15 30
t × ω1 t × ω1
0.4 0.4
ζ / h0
ζ / h0
0 0
−0.4 −0.4
190 200 210 190 200 210
t × ω1 t × ω1
equivalent (ah /b = 0.02). It was found that the experimental tank required
greater initial momentum than the numerical tank, simply occurring in
the physical process of activating the movement of the tank. Therefore,
it is not possible to compare test series for the exact same forcing am-
plitudes. Although the dominating horizontal flow maintained its overall
one-dimensionality inside the tank, away from the bore tongue and walls,
discrepancies were found at the bore location and at the wall due to vertical
components present in the experimental data. This could also be a reason
the bore is developed for smaller forcing amplitudes numerically. Further,
it should be noted that the skewness of the numerical free-surface is not
typical. It is caused by the bore which occurred periodically in the tank.
The numerical simulations predicted a bore (Fr1 = 0.93, h1 /h2 = 1.34)
whereas the physical experiments captured traveling waves (Fr0 = 1.0,
h1 /h2 ≈ 1). Although the numerical solutions agree overall fairly well
with the experimental data, as shown in the close-up look (e.g. Fig. 18), it
should be emphasized that the vertical velocity components occurring at
the walls and at the location of the bore are not captured by the governing
August 26, 2008 17:6 World Scientific Review Volume - 9in x 6in 05˙main
206 J. B. Frandsen
0.8 0.8
0.4 0.4
ζ / h0
ζ / h0
0 0
−0.4 −0.4
−0.8 −0.8
190 200 210 190 200 210
t×ω t×ω
1 1
equations, and are thus a source of error inherently built into the numerical
formulation. Further details and discussions on the 1-D solution can be
found in Frandsen (2006a). Two-dimensional simulations with sway base
excitations have also been undertaken. We found that the 1-D and 2-D
weak bore solution compares fairly well.
Fig. 19. Snapshots of wave profile comparisons for sway and surge base excitation.
Our third numerical test to be presented involves sway and surge base
excitation for ahx = ahy =0.006 (expanding the above pure sway case). The
fluid flow is characterized by D2 Q9 lattices. These preliminary test series
include low resolution grids 500×500 (∆t × ω1 =0.002). Figure 19 shows
August 26, 2008 17:6 World Scientific Review Volume - 9in x 6in 05˙main
Cl
y
lw1
x
Flow 2
6 5
lb
Cl
l
3 0 1
h1 h2
lw2
y
7 8
4
b (a) x (b)
Fig. 20. (a) Definition sketch of the dam-break test case. (b) D2 Q9 velocity set.
August 26, 2008 17:6 World Scientific Review Volume - 9in x 6in 05˙main
208 J. B. Frandsen
Fig. 21. Snapshots of dam-break surface flow through asymmetric wall configuration
at 1.6 s, 4.9 s, 7.2 s.
(a) (b)
10
10
] ][m]
1s 3s 5s 7.2s
5200
5
0
0 y
y/b [m] 200
(c)
Fig. 22. Snapshots of the free-surface elevation contours of dam-break flow through the
asymmetric wall configuration (in meters). (a) LBGK, (b) BGK, after Ghidaoui et al.
(2001), (c) LBGK free-surface profiles at selected instances (x = lw1 + lb /2).
(a)
(b)
(c)
Fig. 23. (a) Dam-break problem including (b) wall movement and (c) bed slope (α = 4);
5 s between snapshots.
August 26, 2008 17:6 World Scientific Review Volume - 9in x 6in 05˙main
210 J. B. Frandsen
the physics of the flow processes. Second, instead of the regular standard
lattice scheme one could test the LB performance using adaptive irregular
lattices.
Based on our current knowledge, we highly recommend that that a
second-order FD LB scheme form the basis for free-surface water-wave de-
velopments. We are currently developing an LB model applicable at any
water depth. This involves an entirely new LB model formulation which
satisfies the Navier-Stokes equations. Our focus is on breaking wave pre-
dictions and we are investigating free-surface algorithms which will allow
for steep and overturning waves. Closely related, we are also undertaking
detailed bubble break-up studies based on soft-sphere collision models. An
LB sediment-transport model is also part of the LB development. At some
point in time, we should be able to simulate the wave transformation from
deep to shallow water to runup.
On a final note, it is our hope that this book chapter represents a useful
starting point for those interested in learning about and/or applying LB
models. In our opinion, LB simulations are a promising route in calculation
methods and may be a helpful method in discovering new physics.
Acknowledgments
I wish to thank Professor P. L.-F. Liu, Cornell University (USA), for giving
me the opportunity to write about the first time experiences with the Lat-
tice Boltzmann generated water waves. Thanks are also due to Professor
G. Pedersen, University of Oslo (Norway), who provided the NLSW and
Boussinesq data allowing us to assess the Lattice Boltzmann solutions in
detail. I would also like to mention that I met Professor H. Yeh, Oregon
State University, in Spring 2004, who actually introduced me to the wave
runup problem. In many ways, this was a turning point for the Lattice
Boltzmann model development and other new research directions. I am
also thankful to Professor C. Dalton, University of Houston (USA), for his
many valuable comments and corrections to the manuscript.
Further thanks to the many students, including under-represented mi-
nority and early career researchers, who got partly involved in this project
of LB model learning. Especially, I would like to thank visiting scholar, Dr.
J. Zhang, University of Hawai’i Manoā (UHM), who assisted producing the
results for the benchmark test case 1 at the final phase of the chapter devel-
opment. Finally, many thanks to the organizers of the faculty ambassadors
program of UHM, for giving me the opportunity to share with middle and
August 26, 2008 17:6 World Scientific Review Volume - 9in x 6in 05˙main
212 J. B. Frandsen
References
Aidun, C. K., Lu, Y. and Ding, E.-J. (1998) Direct analysis of particulate sus-
pensions with inertia using discrete Boltzmann equation, Journal of Fluid
Mechanics 373, pp. 287–311.
Benzi, R., Succi, S. and Vergassola, M. (1992) The Lattice Boltzmann equation:
theory and applications, Physics Reports (review section of Physics Letters)
222(3), pp. 145–197.
Bhatnagar, P., Gross, E. and Krook, M. A. (1954) A model for collision processes
in gases i: small amplitude process in charged and neutral one component
systems, Physics Review A 94, pp. 511–526.
Boltzmann, L. (1964) Lectures on Gas Theory (Dover Publications).
Buick, J. M. and Greated, C. A. (2003) Lattice Boltzmann modelling of interfacial
gravity waves, Physics of Fluids 10(6), pp. 1490–1511.
Burgers, J. M. (1969) Flow Equations for Composite Gases (Academic Press).
Cao, N., Chen, S., Jin, S. and Martinez, D. (1997) Physical symmetry and lattice
symmetry in the Lattice Boltzmann method, Physical Review E 55(1),
pp. R21–R24.
Carrier, G. F., Wu, T. T. and Yeh, H. (2003) Tsunami run-up and draw-down on
a plane beach, Journal of Fluid Mechanics 475, pp. 79–99.
Cercignani, C. (1988) The Boltzmann Equation and its Applications (Springer-
Verlag).
Chapman, S. (1916) On the law of distribution of molecular velocities, and
on the theory of viscosity and thermal conduction, in a non-uniform
simple monatomic gas, Philosophical Transactions Royal Society A 216,
pp. 279–348.
August 26, 2008 17:6 World Scientific Review Volume - 9in x 6in 05˙main
Chapman, S. (1918) On the kinetic theory of gas. part ii: a composite monatomic
gas: diffusion, viscosity, and thermal conduction, Philosophical Transac-
tions Royal Society A 217, pp. 115–197.
Chapman, S. and Cowling, T. G. (1958) The Mathematical Theory of Non-
Uniform Gases (Cambridge University Press).
Chen, H., Chen, S. and Matthaeus, W. H. (1992) Recovery of the Navier-Stokes
equations using a lattice-gas Boltzmann method, Physical Review A 45(8),
pp. R5339–R5342.
Chen, H., I. Saroselsky, I., Orszag, S. and Succi, S. (2004) Expanded analogy
between boltzmann kinetic theory of fluids and turbulence, Journal of Fluid
Mechanics 519, pp. 301–314.
Chen, S. and Doolen, G. D. (1998) Lattice Boltzmann method for fluid flows,
Annual Review of Fluid Mechanics 30(4), pp. 329–364.
Crouse, B., Rank, E., Krafczyk, M. and Tölke, J. (2003) A LB-based approach
for adaptive flow simulations, International Journal of Modern Physics B
17(1 & 2), pp. 109–112.
Crouse, M., B. Krafczyk, Kühner, S., Rank, E. and Van Treeck, C. (2002) In-
door air flow analysis based on Lattice Boltzmann methods, Energy and
Buildings 34, pp. 941–949.
Dalrymple, R. A. and Rogers, B. (2006) A note on wave celerities on compressible
fluid, in Proceedings, 30th International Conference on Coastal Engineering
(San Diego, USA).
Dellar, P. J. (2001) Bulk and shear viscosities in Lattice Boltzmann equations,
Physics Review E 64, 031203.
Dellar, P. J. (2002) Non-hydrodynamic modes and a priori construction of shallow
water lattice boltzmann equations, Physics Review E 65, 036309.
Deng, J. Q., Ghidaoui, M. S., Gray, W. G. and Xu, K. (2001) A Boltzmann-based
mesoscopic model for contaminant transport in flow systems, Advances in
Water Resources 24, pp. 531–550.
Denniston, C., Marenduzzo, D., Orlandini, E. and Yeomans, J. M. (2004) Lattice-
Boltzmann algorithm for three-dimensional liquid crystal hydrodynamics,
Philosophical Transactions, Royal Society of London Series A Mathematical
Physical and Engineering Sciences 362(1821), pp. 1745–1754.
Derksen, J. J. (2005) Simulations of confined turbulent vortex flow, Computers
and Fluids 34(3), pp. 301–318.
D’Humières, D., Ginzburg, I., Krafczyk, M., Lallemand, P. and Luo, L.-S. (2002)
Multiple-relaxation-time for the Lattice Boltzmann models in three di-
mensions, Philosophical Transactions, Royal Society London A. 360(1),
pp. 437–451.
Ding, E.-J. and Aidun, C. K. (2003) Extension of the Lattice Boltzmann method
for direct simulation of suspended particles near contact, Journal of Statis-
tical Physics 112(314), pp. 685–708.
Ding, E.-J. and Aidun, C. K. (2006) Cluster size distribution and scaling for
spherical particles and red blood cells in pressure-driven flows at small
reynolds number, Physical Review Letters 96, 204502.
Enskog, D. (1921) The numerical calculation of phenomena in fairly dense gases,
Arkiv Mat. Astr. Fys. 16(1).
August 26, 2008 17:6 World Scientific Review Volume - 9in x 6in 05˙main
214 J. B. Frandsen
Hou, S., Sterling, J., Chen, S. and Doolen, G. D. (1996) A Lattice Boltzmann
subgrid model for high Reynolds number flows, Fields Institute Communi-
cations 6, pp. 151–165.
Inamuro, T., Ogata, S., Tajima, S. and Konishi, N. (2004) A Lattice Boltzmann
method for incomressible two-phase with large density differences, Journal
of Computational Physics 198, pp. 628–644.
Junk, M. (2001) A finite difference interpretation of the Lattice Boltz-
mann method, Numerical Methods Partial Differential Equations 17(4),
pp. 383–402.
Junk, M., Klar, A. and Luo, L.-S. (2005) Asymptotic analysis of the Lattice
Boltzmann equation, Journal of Computational Physics 210, pp. 676–704.
Kalikmanov, V. I. (2001) Statistical Physics of Fluids. Basic Concepts and
Applications (Springer).
Kandhai, D., Soll, W., Chen, S., Hoekstra, A. and Sloot, P. (2000) Finite-
difference Lattice-BGK methods on nested grids, Computer Physics Com-
munications 129, pp. 100–109.
Karniadakis, G. E. and Beskok, A. (2002) Micro Flows. Fundamentals and Sim-
ulations (Springer-Verlag).
Keller, H. B., Levine, D. A. and Whitham, G. B. (1960) Motion of a bore over a
sloping beach, Journal of Fluid Mechanics 7, pp. 302–316.
Koelman, J. M. V. A. (1991) A simple Lattice Boltzmann scheme for Navier-
Stokes fluid flow, Europhysics Letters 15(6), pp. 603–607.
Körner, C., Pohl, T., Rüde, U., Thürey, N. and Zeiser, T. (2006). Parallel Lat-
tice Boltzmann Methods for CFD Applications (Numerical Solution of Par-
tial Differential Equations on Parallel Computers. Eds. A. M. Bruaset, A.
Tveito. Lecture Notes in Computational Science and Engineering, Vol. 51,
Springer).
Körner, C., Thies, H., M., T. Thürey, N. and Rüde, U. (2005) Lattice Boltzmann
model for free surface flow for modeling foaming, Journal of Statistical
Physics 121(1-2), pp. 179–196.
Krafczyk, M., Tölke, J. and Luo, L.-S. (2003) Large-eddy simulations with a mul-
tiple relaxation time LBE model, International Journal of Modern Physics
B 17(1), pp. 33–39.
Krafczyk, M. and Tölke, J. (2004) Lattice-Boltzmann methods — basics and
recent progress, in NAFEMS CFD Workshop on Simulation of Complex
Flows (Germany).
Lallemand, P. and Luo, L.-S. (2000) Theory of the Lattice Boltzmann method:
dispersion, dissipation, isotropy, galilean invariance, and stability, Physical
Review E 61, pp. 6546–6562.
Laney, C. B. (1998) Computational Gasdynamics (Cambridge University Press).
Lee, K., Yu, D. and Girimaji, S. S. (2006) Lattice Boltzmann DNS of decay-
ing compressible isotropic turbulence with temperature fluctuations, Inter-
national Journal of Computational Fluid Dynamics 20, pp. 401–413, doi:
doi:10.1080/10618560601001122, URL http://www.ingentaconnect.com/
content/tandf/gcfd/2006/0000002%0/00000006/art00006.
August 26, 2008 17:6 World Scientific Review Volume - 9in x 6in 05˙main
216 J. B. Frandsen
Lee, T. and Lin, C.-L. (2005) A stable discretization of the Lattice Boltzmann
equation for simulation of incomressible two-phase flows at high density
ratio, Journal of Computational Physics 206, pp. 16–47.
Lesieur, M., Métais, O. and Comte, P. (2005) Large-eddy Simulations of Turbu-
lence (Cambridge University Press).
LeVeque, R. J. (2002) Finite Volume Methods for Hyperbolic Problems (Wiley).
LeVeque, R. J. and George, D. L. (2008) “High-resolution finite volume meth-
ods for the shallow water equations with bathymetry and dry states”, in
Advanced Numerical Models for Simulating Tsunami Waves and Runup,
Advances in Coastal and Ocean Engineering, Vol. 10, eds. Liu, P. L.-F.,
Yeh, H. and Synolakis, C. (World Scientific Publishing Co.).
Li, Y., Shock, R., Zhang, R. and Chen, H. (2004) Numerical study of flow past
an impulsively started cylinder by the Lattice-Boltzmann method, Journal
of Fluid Mechanics 519, pp. 273–300.
Liang, J. H., Ghidaoui, J. Q., M. S. Deng and Gray, W. G. (2007) A Boltzmann-
based finite volume algorithm for surface water flows on cells of arbitrary
shapes, Journal of Hydraulic Research 45(2).
Lockard, D. P., Luo, L.-S., Milder, S. D. and Singer, B. A. (2002) Evaluation of
PowerFLOW for aerodynamic applications, Journal of Statistical Physics
107(1/2), pp. 423–478.
Luo, L.-S. (2000) The lattice-gas and Lattice Boltzmann methods: Past, present,
and future, in Proceedings of the International Conference on Applied Com-
putational Fluid Dynamics (Beijing, China), pp. 52–83.
Luo, L.-S., Qi, D. and Wang, L. P. (2002) Applications of the Lattice Boltzmann
method to complex and turbulent flows (Eds. M. Breuer, F. Durst, and C.
Zenger. Lecture Notes in Computational Science and Engineering, Vol. 21,
pp. 123-130.).
Lynett, P. J., Wu, T.-R. and Liu, P. L.-F. (2002) Modeling wave runup with
depth-integrated equations, Coastal Engineering 46, pp. 89–107.
McCracken, M. E. (2004) Development and Evaluation of the Lattice Boltzmann
Models for Investigations of Liquid Break-up, PhD Thesis, Purdue Univer-
sity, USA.
McNamara, G. R. and Zanetti, G. (1988) Use of the Boltzmann equa-
tion to simulate lattice gas automata, Physics Review Letter 61(20),
pp. 2332–2335.
Mei, C. C. (1992) The Applied Dynamics of Ocean Surface Waves (World
Scientific Publishing Co.).
Mei, R., Luo, L.-S. and Shyy, W. (1999) An accurate curved boundary treatment
in the Lattice Bolzmann method, Journal of Computational Physics 155,
pp. 307–330.
Mei, R., Shyy, W., Yu, D. and Luo, L.-S. (2000) Lattice Bolzmann method for
3-D flows with curved boundaries, Journal of Computational Physics 161,
pp. 680–699.
Noble, R., Chen, S., Georgiadis, J. and Buckius, R. (1995) A consistent hydrody-
namic boundary condition for the Lattice Boltzmann method, Physics of
Fluids 7(1), pp. 203–209.
August 26, 2008 17:6 World Scientific Review Volume - 9in x 6in 05˙main
218 J. B. Frandsen
Part 2
Extended Abstracts
This page intentionally left blank
CHAPTER 6
BENCHMARK PROBLEMS
223
224 P. L.-F. Liu, H. Yeh and C. E. Synolakis
cm
(a) Whole model area
m
(b) Detailed topography near the maximum runup
The problem statement, data file, input file, bathymetry file and the
experimental data file can be found in the following websites,
respectively:
http://www.cee.cornell.edu/longwave/data/problem02.doc
http://www.cee.cornell.edu/longwave/data/Benchmark_2.txt
http://www.cee.cornell.edu/longwave/data/Benchmark_2_input.txt
http://www.cee.cornell.edu/longwave/data/Benchmark_2_Bathymetry.txt
http:// www.cee.cornell.edu/longwave/data/benchmark2/output_ch5-7
9.xls
where
H(x) = x tan β ,
2
2
x µ g
ho = δ exp − 2 − µt ,
δ tan β δ
g
CASE A: µ t = 0.1, 0.5, 1.0, 1.5
δ
g
CASE B: µ t = 0.5, 1.0, 2.5, 4.5
δ
The data files for this benchmark problem are named as readme1.txt,
bench3A_t05.txt, bench3B_t10.txt, bench3A_t15.txt, bench3B_t05.txt,
bench3B_t10.txt, bench3B_t25.txt, and bench3B_t45.txt; they can be
found in the following websites:
http://www.cee.cornell.edu/longwave/data/benchmark3/readme1.txt
http://www.cee.cornell.edu/longwave/data/benchmark3/bench3A_t05.txt
http://www.cee.cornell.edu/longwave/data/benchmark3/bench3A_t10.txt
http://www.cee.cornell.edu/longwave/data/benchmark3/bench3A_t15.txt
http://www.cee.cornell.edu/longwave/data/benchmark3/bench3B_t05.txt
http://www.cee.cornell.edu/longwave/data/benchmark3/bench3B_t10.txt
http://www.cee.cornell.edu/longwave/data/benchmark3/bench3B_t25.txt
http://www.cee.cornell.edu/longwave/data/benchmark3/bench3B_t45.txt
Included in each data file are four columns of data: x (m); H (m); ho (m);
ζ (m). The free surface elevations, ζ , are those predicted by an analytical
solution. Agreement with CASE B using any model other than the linear
shallow water wave equations may be poor. For more information, please
refer to Liu, Lynett and Synolakis (J. Fluid Mech., 478, 101–109, 2003).
http://www.cee.cornell.edu/longwave/data/Benchmark_4_readme.txt
http://www.cee.cornell.edu/longwave/data/Benchmark_4_landslide_setu
p.pdf
http://www.cee.cornell.edu/longwave/data/Benchmark_4_run30.txt
http://www.cee.cornell.edu/longwave/data/Benchmark_4_run32.txt
The wave gage data and runup data can be located in the following
websites:
http://www.cee.cornell.edu/longwave/data/benchmark4/gage_data_read
me.txt
http://www.cee.cornell.edu/longwave/data/benchmark4/run30_wave_gag
e_1.txt
http://www.cee.cornell.edu/longwave/data/benchmark4/run30_wave_gag
e_2.txt
http://www.cee.cornell.edu/longwave/data/benchmark4/run30_runup_ga
ge_2.txt
http://www.cee.cornell.edu/longwave/data/benchmark4/run30_runup_ga
ge_3.txt
http://www.cee.cornell.edu/longwave/data/benchmark4/run32_wave_gag
e_1.txt
http://www.cee.cornell.edu/longwave/data/benchmark4/run32_wave_gag
e_2.txt
http://www.cee.cornell.edu/longwave/data/benchmark4/run32_runup_ga
ge_2.txt
http://www.cee.cornell.edu/longwave/data/benchmark4/run32_runup_ga
ge_3.txt
For more information, please refer to Liu, Wu, Raichlen, Synolakis, and
Borrero (J. Fluid Mech., 536, 107–144, 2005).
August 8, 2008 11:58 WSPC/Trim Size: 9in x 6in for Review Volume 07˙kowa
CHAPTER 7
231
August 8, 2008 11:58 WSPC/Trim Size: 9in x 6in for Review Volume 07˙kowa
T
ζjm+1 = ζjm − (f luxm+1 m+1
j+1 − f luxj ),
h
(H +H )
where f luxj = um+1 p
m
ζj−1 + um+1
n ζjm + um+1
j
j
2
j−1
, up = 0.5(uj + |uj |)
and un = 0.5(uj − |uj |). T is the time step, h is the space step. Indices m
and j = 1, 2, 3, ...n−1 stand for the time stepping and horizontal coordinate
points, respectively. For the runup condition, the following steps are taken
when the dry point (jwet − 1) is located to the left of the wet point jwet ,
thus: if (ζ m (jwet ) > −H(jwet − 1)), then um m
jwet = ujwet +1 , see Kowalik and
8
Murty.
~um+1 − ~ũ 1
= − m ∇pm+1 . (7)
T ρ
Equation (7) and the continuity equation, ∇ · ~um+1 = 0, can be combined
into a single equation (Poisson equation) for the solution of the pressure as
∇ · ~ũ
1 m+1
∇ · m ∇p = . (8)
ρ T
The fluid free surface is described by the discrete VOF method, see Nichols
and Hirt5 and Nichols et al.6
Analytical
0.03
First Order
Leap−Frog
FNS−VOF
0.02
0.01 t=3.0
815
(220
s)
η (dimensionless)
0 s)
(175 s)
512 (160
t= 2.4 411
t=2.2
−0.01
Slop
e=(
1:10
−0.02
)
−0.03
−0.04
Analytical
First Order
Leap−Frog
0.1 t=2
.24
11
(16
0s
ec
)
0.05
u (dimensionless)
MWL
0
t=3.081
5 (220
se c)
−0.05
c )
se
75
(1
−0.1 Slop
12
e=(
1:10
45
)
2.
t=
−0.15
−0.2
−0.04 −0.02 0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
x (dimensionless)
3.5
Analytical
First Order
Leap−Frog
FNS−VOF
3.0815 220
3
t (dimensionless)
t(seconds)
2.5
2.4512 175
2.2411 160
1.5
−0.06 −0.04 −0.02 0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
x (dimensionless)
match very well with the analytical solution. The extrapolation of the veloc-
ity from the immediate wet cell to the dry cell facilitates runup, improving
the timing. The leap frog method does well in predicting the analytical
solution of the shoreline evolution, sea level and velocity profile as well.
However, due to the small difference in timing, some discrepancy in the
velocity profile can be seen, i.e. at time 220s. The FNS-VOF method gives
a frame of reference to validate the NLSW solutions. Some differences in
wave profile, shore line evolution and timing are quite plausible, since FNS
approximation allows vertical fluid velocity/acceleration while the NLSW
theory does not. From Figs. 1 and 3, it is clear that dispersion effects are
important. NLSW and analytical solutions underestimate the runup and
overestimate the rundown. Timing of maximum runup and rundown occur
slightly earlier in the NLSW solutions.
References
1. G. Carrier, T. T. Wu and H. Yeh, “Tsunami runup and draw-down on a plane
beach”, Journal of Fluid Mechanics, 475:79–99, 2003.
2. A. J. Chorin, Numerical solution of the Navier-Stokes equations, Math. Comp.,
22:745–762, 1968.
August 8, 2008 11:58 WSPC/Trim Size: 9in x 6in for Review Volume 07˙kowa
Utku Kânoğlu
Department of Engineering Sciences, Middle East Technical University
Ankara, 06531, Turkey
[email protected]
The solution method presented by Kânoğlu (2004) for the initial value
problem solution of the nonlinear shallow water-wave equations over a
sloping beach is applied to the benchmark problem. Here only the
description of the method and how it is applied to the benchmark
problem will be explained. Details of the method summarized here can
be found in Kânoğlu (2004).
The dimensionless nonlinear shallow water-wave equations that
describe a propagation problem over the undisturbed water of variable
depth h( x) = x are
[u (η + h)]x + ηt = 0, ut + u u x + η x = 0.
Here u(x, t), η(x, t), and β are the horizontal depth-averaged velocity, the
free-surface elevation, and the beach angle from the horizontal
respectively. The initial shoreline is chosen at x = 0 with x increasing in
the seaward-direction. The dimensionless variables are defined using
~ ~ ~
l , ( l tan β ), and l /( g~ tan β ) as the characteristic length, height, and time
scales; g~ being the gravitational acceleration. The major analytical
237
238 U. Kânoğlu
ϕ ϕ u2 u 2 ϕλ σ 2 λ
u= σ , η= λ − , x= − + , t =u− ,
σ 4 2 2 4 16 2
and reduced the nonlinear shallow water-wave equations into the second-
order linear equation: σϕ λλ − (σϕσ )σ = 0. The instantaneous shoreline is
defined at σ = 0 in the transform space with this transformation. Given
initial condition in terms of an initial wave profile with zero initial
velocity, Carrier and Greenspan (1958) provided the following solution:
∞ ∞
ϕ (σ , λ ) = −
∫ ∫
0 0
(1/ ω ) ζ 2 J 0 (ωσ ) sin(ωλ ) J1 (ωζ ) φ (ζ ) dω dζ ,
0
η(m)
−5
−10
0 10 20 30 40 50
x(km)
Fig. 1. Comparison of the initial wave profiles resulted from the present nonlinear
solution and given in the benchmark problem. The circles represent the initial wave
profile given in the benchmark problem.
240 U. Kânoğlu
0 9
6
u(m/s)
−10
η(m)
t = 160s 3 t = 160s
−20
0
−30 −3
0 200 400 600 800 0 200 400 600 800
0 0
u(m/s)
−10 −5
η(m)
t = 175s t = 175s
−20 −10
−30 −15
0 200 400 600 800 0 200 400 600 800
20 2
u(m/s)
0
η(m)
10 t = 220s t = 220s
−2
0 −4
−200 0 200 400 600 800 −200 0 200 400 600 800
x(m) x(m)
Fig. 2. Comparison of the present nonlinear solution and benchmark solution for the
spatial variation of the surface elevations and velocities at three specific times. The
circles represent the benchmark solution.
Nonlinear Evolution of Long Waves Over a Sloping Beach 241
300
(a)
200
100
x(m)
−100
−200
0 50 100 150 200 250 300 350
10
(b)
0
u(m/s)
−10
−20
0 50 100 150 200 250 300 350
t(s)
Fig. 3. Comparison of the present nonlinear solution and benchmark solution for the
shoreline (a) position and (b) velocity. The circles represent the benchmark solution.
References
1. G. F. Carrier and H. P. Greenspan, J. Fluid Mech. 4, 97 (1958).
2. G. F. Carrier, T. T. Wu and H. Yeh, J. Fluid Mech. 475, 79 (2003).
3. U. Kânoğlu, J. Fluid Mech. 513, 363 (2004).
This page intentionally left blank
CHAPTER 9
1. Introduction
Tsunami runup onto a complex 3D dimensional beach is one of the
Benchmark Problems of tsunami runup. The extreme runup height of
32 m that was measured near the village of Monai in Okushiri Island by
1993 Okushiri tsunami is one of the examples of this kind of benchmark
problems. There also exists experimental data in a 1/400 scale laboratory
243
244 A. C. Yalciner, F. Imamura and C. E. Synolakis
2. Modeling
The numerical model, TUNAMI-N2, used for the simulation of the
propagation of long waves is authored and developed by Prof. Imamura
in Disaster Control Research Center in Tohoku University, Japan,
through the Tsunami Inundation Modeling Exchange (TIME) program
(Shuto et al. (1990)). TUNAMI-N2 is one of the key tools for the studies
for propagation and coastal amplification of tsunamis in relation to
different initial conditions. It solves the nonlinear form of long wave
equations with bottom friction by finite difference technique, and
computes water surface fluctuations and depth averaged velocities at all
locations even at shallow and land regions within the limitations of grid
size on the complex bathymetry and topography. Consequently the
program simulates the propagation and coastal amplification of long
waves. Shuto et al. (1990), Nagano et al. (1991), Imamura (1996), Goto
et al. (1997), Imamura et al. (1999), Yalciner et al. (2001a, 2001b, 2002,
2003, 2004), are some of the studies used in TUNAMI-N2.
The computation domain is selected as 5.488 m by 3.402 m (393
nodes × 244 nodes with the grid size of 0.014 m). The bathymetry data is
the same as the one used in laboratory experiments where the minimum
depth is −0.125 m (land) and maximum depth is 0.13535 m. A typical
value of 0.025 is used as Manning’s coefficient (Munson et al. (1998)).
The time step was chosen as 0.05 sec to satisfy the CFD condition. The
propagation of the wave in the basin with respect to the inputted wave
form (from laboratory data) is simulated.
Amplitude Evolution and Runup of Long Waves 245
Fig. 1. The computed snapshots (at t = 12.5 and 18.4 sec. in the model) of the wave
propagation (left two figures) and maximum positive amplitudes during 22.5 sec
simulation.
246 A. C. Yalciner, F. Imamura and C. E. Synolakis
Fig. 2. The measured and computed values of temporal water surface variations
measured at different gauge locations.
results for three gauge locations are compared. These results show that
the simulation using TUNAMI-N2 provides fairly consistent results with
experiments.
Acknowledgments
The tsunami propagation model TUNAMI-N2 is a registered copyright
of Imamura, Yalciner and Synolakis.
References
Goto, C., Ogawa, Y., Shuto, N. and Imamura, F. (1997) Numerical method of tsunami
simulation with the leap-frog scheme, IOC Manual 35, U.N. Educ., Sci. and Cult. Org.,
Paris.
Imamura, F., Koshimura, S. and Yalciner, A. C. (1999) Field survey and numerical
modeling of tsunami generated by Turkish earthquake of August 17, 1999 (in Japanese),
Proc. Coastal Engineering in Japan, 47, 331–333.
Lin, P., Chang, K. A. and Liu, P. L.-F. (1999) Runup and Rundown of Solitary Waves on
Sloping Beaches, Journal of Waterway, Port, Coastal, and Ocean Engineering,
September/October 1999.
Nagano, O., Imamura, F. and Shuto, N. (1991) A numerical model for a far field
tsunamis and its application to predict damages done to aquaculture, Nat. Hazards, 4,
235–255.
Shuto, N., Goto, C. and Imamur, F. (1990) Numerical simulation as a means of warning
for near field tsunamis, Coastal Engineering in Japan, 33(2), 173–193.
Yalciner, A. C., Pelinovsky, E., Okal, E. and Synolakis, C. E. (Eds.) (2001a) NATO
ARW, Underwater Ground Failures on Tsunami Generation, Modeling, Risk and
Mitigation, North Atlantic Treaty Org., Istanbul, Turkey.
Yalciner, A. C., Synolakis, A. C., Alpar, B., Borrero, J., Altinok, Y., Imamura, F.,
Tinti, S., Ersoy, Ş., Kuran, U., Pamukcu, S. and Kanoglu, U. (2001b) Field Surveys and
Modeling 1999 Izmit Tsunami, International Tsunami Symposium ITS 2001, Session 4,
Paper 4–6, Seattle, August 7–9, 2001, pp. 557–563.
Yalciner, A. C., Alpar, B., Altinok, Y., Ozbay, I. and Imamura, F. (2002) Tsunamis in
the Sea of Marmara: Historical documents for the past, models for future, Mar. Geol.
190(1–2), 445–463.
Yalçiner, A. C., Pelinovsky, E., Talipova, T., Kurkin, A., Kozelkov, A. and Zaitsev A.,
(2004) Tsunamis in the Black Sea: Comparison of the historical, instrumental, and
numerical data, J. Geophys. Res., 109, C12023, doi:10.1029/2003JC002113.
Yeh, H., Liu, P. L.-F. and Synolakis, C. (Eds.) (1996) Long-Wave Runup Models (World
Scientific Publishing Co.).
This page intentionally left blank
August 8, 2008 12:8 WSPC/Trim Size: 9in x 6in for Review Volume 10˙wang
CHAPTER 10
Alejandro Orfila
IMEDEA(CSIC-UIB), Miquel Marques, 21.07190 Esporles, Spain
[email protected]
1. Introduction
In this paper we present the numerical results for the benchmark problem
#2 proposed. The model implemented is the Cornell Multigrid Coupled
Tsunami Model, COMCOT (Liu et al., 1998). Although a multi-grid sys-
tem, dynamically coupled up to three regions (either spherical or cartesian)
can be implemented, in the present simulations only one layer has been used.
The governing equations for the model are based on the shallow water wave
equations. The continuity and momentum equations are expressed as,
∂η ∂P ∂Q
+ + = 0, (1)
∂t ∂x ∂y
P2
∂P ∂ ∂ PQ ∂η
+ + + gH + τx = 0 , (2)
∂t ∂x H ∂y H ∂x
Q2
∂Q ∂ PQ ∂ ∂η
+ + − gH + τy = 0 , (3)
∂t ∂x H ∂y H ∂y
where η is the free-surface displacement, h is the still water depth, H = h+η
is the total water depth and u, v are the depth averaged velocities in the
249
August 8, 2008 12:8 WSPC/Trim Size: 9in x 6in for Review Volume 10˙wang
gn2 1/2
τy = 7/3
Q P 2 + Q2 , (5)
H
where n is the roughness coefficient. The model solves (1)–(3) with an
explicit modified Leap-Frog finite difference scheme (Cho, 1995).
2. Numerical Results
The initial surface profile provide by the benchmark problem is set as the
initial wave input in the left boundary and the wave propagation is sim-
ulated for 150 s. Solid reflective walls are implemented in the other three
boundaries. Numerical results are compared with data provided at three
gage locations A = (4.521,1.196); B = (4.521,1.696) and C = (4.521,2.196).
0.04
Coarser grid
a)
Amplitude (m)
Finer grid
0.02
-0.02
0 50 100 150
time (s)
0.04
No friction
b)
Amplitude (m)
With friction
0.02
-0.02
0 50 100 150
time (s)
0.06
Linear SW
Amplitude (m)
Non-linear SW
0.04
c)
0.02
-0.02
0 50 100 150
time (s)
Fig. 1. Comparison between (a) grid sizes; (b) bottom friction effects and (c) linear vs.
non-linear simulations.
water equations are compared with the linearized version of the equations.
Figure 1(c) shows the numerical results for the linear SW equations (black)
and for the nonlinear SW equations (grey) where the main differences are
obtained at the small scales providing overall similar results.
0.03 0
0.01
0
0.02
0.04
0 05
Fig. 2. (a) Free-surface elevation at gage A; (b) B and (c) C provided by the numeri-
cal model black and the experiments (grey). (d) Maximum runup superimposed to the
bathymetric contours.
4. Conclusions
The numerical results show that the problem can be well simulated as a first
attempt with the linear system without including bottom friction. Results
indicate that the grid size reduction does not lead to a much accurate
results. COMCOT results match the experimental data very good for both
the arrival time and amplitude of leading waves. In the near shore region,
the waves becomes very nonlinear and will break. In this zone, the present
model is no longer capable to deal with it and the wave amplitude may be
exaggerated.
August 8, 2008 12:8 WSPC/Trim Size: 9in x 6in for Review Volume 10˙wang
References
1. P. L.-F. Liu, S.-B. Woo and Y. S. Cho, Computer Program for Tsunami Prop-
agation and Inundation, Cornell University (1998).
2. Y.-S. Cho, Numerical Simulations of Tsunami Propagation and Runup, Ph.D.
thesis, Cornell University (1995).
This page intentionally left blank
CHAPTER 11
Taro Kakinuma
Tsunami Research Center, Port and Airport Research Institute
3-1-1 Nagase, Yokosuka, Kanagawa 239-0826, Japan
E-mail: [email protected]
1. Introduction
From offshore to coastal zones, phenomena having various scales on
space and time should be solved economically. For this purpose we have
developed a hybrid model, STOC, 1 which consists of 3D models,
multilevel models and connection models. As shown in Fig. 1, the
multilevel models are used for wide-area calculation, where we treat
hydrostatic pressure, while the connection models are applied to overlap
regions for smooth connection between the multilevel models and the 3D
models, where we solve pressure without assumption of hydrostatic
pressure. By local application of the 3D models to narrower areas
255
256 T. Kakinuma
Fig. 3. Highest runup at Point A when t = 16.5 s. This water surface profile was calculated
by STOC-VF under the slip condition.
Fig. 4. Velocity vectors and constant-pressure curves on the x–z plane including Point A
in Fig. 3 when t = 16.5 s. These were calculated by STOC-VF under the slip condition.
3. Conclusions
Two numerical, hydrodynamic models, STOC-NS and STOC-VF, have
been applied to Benchmark No. 2, where water surface displacement is
determined by different ways. The calculation results by both models
under the slip-bed condition generally show correspondence with the
260 T. Kakinuma
References
1. T. Kakinuma and T. Tomita, Proc. 29th Int. Conf. on Coastal Eng., 1552 (2004).
2. T. Kakinuma and T. Tomita, OCEANS/TECHNO-OCEAN Conf. Proc., 146 (2004).
3. M. Matsuyama and H. Tanaka, Proc. Int. Tsunami Symposium 2001, 879 (2001).
August 26, 2008 17:11 WSPC/Trim Size: 9in x 6in for Review Volume 12˙titov
CHAPTER 12
1. Introduction
Runup measurements from the 1993 tsunami off the coast of Okushuri is-
land in Japan showed that the highest wave was measured at the end of a
narrow gully in a cove near the village of Monai. The measured runup mark
at 32 m was significantly higher than other recorded runup values around
the island, and was the largest recorded measurement in Japan over the last
century.2 To determine how the local bathymetric and topographic features
could act as a focusing mechanism for wave energy, a laboratory scale ex-
periment of this region was carried out at the Central Research Institute
for Electric Power Industry (CRIEPI) in Akibo, Japan. This experimental
data set provides a valuable benchmark for evaluating long-wave tsunami
runup models and was part of a recent international workshop on long-wave
runup models.
In this paper the laboratory data set is used to evaluate a 2D long-wave
model (MOST), which is the primary tsunami inundation model used by the
NOAA Center for Tsunami Research at the Pacific Marine and Engineering
Laboratory in Seattle. It is currently being implemented as a forecast model
at the two U.S. Tsunami Warning Centers.4 The model has been developed
by Titov and Synolakis3 and has been validated with analytical solutions,
the conical island experimental studies of Briggs et al.1 and several historic
tsunami events. The aim here is to characterize how accurately the model
can simulate wave propagation and inundation processes in a more complex
but still controlled environment.
261
August 26, 2008 17:11 WSPC/Trim Size: 9in x 6in for Review Volume 12˙titov
2. Laboratory Experiments
Simulations of wave runup around the village of Monai in Okushuri island
were carried out with a 1/400 scale laboratory experiment using a large scale
tank (205 m long, 6 m deep, 3.4 m wide) at the Central Research Institute
for Electric Power Industry (CRIEPI) in Akibo, Japan. This benchmark
problem was part of the “3rd International Workshop on Long-Wave Runup
Models” held on June 17–18, 2004, at the Wrigley Marine Science Center
in Catalina Island California.
Figure 1 shows a side view of the experimental setup. Waves are shoaled
from the deep water using a series of linearly sloping beaches. The area
marked as “the model” in the figure is where the complex bathymetric and
topographic features from the field have been reproduced and marks the
limit of the numerical modeling exercise. The bathymetry for this model
area is shown in Fig. 2. The wave propagates from left to right, with solid
walls along the sides and the right end. The incoming wave is determined
from a gage located near the offshore boundary and is given in Fig. 3.
3. Model
MOST is a 2D long-wave model that uses the method of fractional steps5 to
reduce the 2D shallow water wave equations into a series of 1D equations
along each spatial dimension. Each set of 1D equations is re-written in
characteristic form and solved using an explicit finite difference scheme
(Method of undetermined coefficients). The scheme is second-order accurate
in space and first-order accurate in time. Solving a series of 1D equations
is computationally more efficient than solving the 2D equations. Detailed
3 −0.12
−0.08
2.5
Gage 5
−0.01
2
0.12 0 0 7 −0.08
y(m)
0.08 Gage
0.06 0.04 −0.02
1.5 0.02 0.01 −0.04
Gage 9
0.5
0
0 1 2 3 4 5
x(m)
Fig. 2. Domain bathymetry in the model area. The smallest depth is −0.125 m and
the deepest offshore depth is 0.135 m. Both the cove and the offshore island from the
field have been reproduced in this physical model. The three gage locations used for
model–data comparisons are given by the solid black circles.
0.02
0.015
0.01
0.005
η (m)
−0.005
−0.01
−0.015
0 5 10 15 20 25
time (sec)
4. Results
The simulations were carried out using two grids of different resolutions,
with the coarser grid covering the full domain and the finer grid nested
inside. The coarser grid had a resolution of ∆x = 0.028 m and a time step
∆t = 0.02 s. The surface time series in Fig. 3 was provided as the boundary
condition along the left boundary (x = 0). The corresponding horizontal
velocity was determined from shallow water linear wave theory. Reflecting
boundary conditions were used along the remaining three boundaries. The
internal grid had a resolution of ∆x = 0.014 m and a time step ∆t = 0.01 s.
Its boundary conditions were determined from the solutions of the outer
grid. Imposing boundary conditions from the coarser grid on the incoming
waves and allowing the outgoing waves to radiate out are trivial with the
MOST model since it solves a set of characteristic equations along each
dimension. Wetting and drying was switched on for both the grids, with
one-way coupling between the coarse and fine grid. The simulation was
run on an a single processor of a 4 processor Intel Xeon 3.6 GHz machine
running Red Hat Linux and took approximately 2.7 minutes to complete.
Time series comparison of model and data for the three wave gages in
shallow water (Fig. 2) are shown in Fig. 4. The gap in the model time series
at gage 9 corresponds to drying. There was a phase lag of 0.6 s between the
data and the model at all the three gages. This could either be because there
is an offset in the start time of the model runs as compared to the data, or
that the offshore gage is not exactly at x = 0 as in the model. Accounting
August 26, 2008 17:11 WSPC/Trim Size: 9in x 6in for Review Volume 12˙titov
gage 5
0.04
0.02
η (m)
−0.02
0 5 10 15 20
gage 7
0.04
η (m)
0.02
0 5 10 15 20
gage 9
0.04
η (m)
0.02
0 5 10 15 20
time (sec)
Fig. 4. Model (solid line) and data (dashed line) comparisons at the three gage loca-
tions. See Fig. 2 for gage locations.
for that phase lag, we see from the figure that the model compares very
well with the data. The high frequency waves at the wave crest are not
reproduced in the model simulations. Considering the shallow water depth
at these gages (approximately 1 cm), the high frequency waves are probably
caused by higher order details of wave breaking which are not simulated
in the long wave model. Maximum alongshore (y-direction) wave runup
from the simulations is shown in Fig. 5. The highest value of 10.047 cm is
observed at the cove.
August 26, 2008 17:11 WSPC/Trim Size: 9in x 6in for Review Volume 12˙titov
2.8
2.6
2.4
Gage 5
2.2
y (m)
1.8 Gage 7
−0 −
1.6 0 .02 0.0 −0 −−0.1
4 .060.08
1.4
Gage 9
1.2 0.0
2
1
4 4.2 4.4 4.6 4.8 5 5.2 5.4 2 4 6 8 10 12
x (m) runup (cm)
5. Conclusions
Earlier laboratory validation studies of the MOST tsunami model3 have
shown that the model compares well in simple domains. The model is cur-
rently being used to develop a short-term tsunami forecasting capability4
and has been considerably revised over the last year. Some of the changes
include a generalized methodology for nesting grids and the capability to
have wetting and drying in all the grids. These comparisons with a labora-
tory study in a complex domain show that the model performs reasonably
well. This increases our confidence in the model.
Acknowledgments
This publication is [partially] funded by the Joint Institute for the Study of
the Atmosphere and Ocean (JISAO) under NOAA Cooperative Agreement
No. NA17RJ1232, Contribution number 1315.
References
1. M. J. Briggs, C. E. Synolakis, G. S. Harkins and D. R. Green, Laboratory
experiments of tsunami runup on circular island, Pure and Applied Geophys.
144 (1995) 569–593.
2. N. Shuto and H. Matsutomi, Field survey of the 1993 Hokkaido Nansei-Oki
tsunami, Pure and Applied Geophys. 144 (1995) 649–664.
August 26, 2008 17:11 WSPC/Trim Size: 9in x 6in for Review Volume 12˙titov
CHAPTER 13
269
August 26, 2008 19:8 WSPC/Trim Size: 9in x 6in for Review Volume 13˙Kornkven
1 dVo ∂~u 1
∇ · ~u = = φ and + ∇ · (~u~u) − φ~u = − ∇p + ν∇2 ~u + ~g , (1)
V dt ∂t ρ
& ! " '(" *)+, - #./0- '- . 1 243 &65 $
7 *81" *)+, - #./0- '9-, . 1 3 &65 $
: - ;*" *) <=79" >#. ;*7" #!" ? . !A@ !-B;*7 * *) .C9D E9 F
: 7* <=)
3C 9 I !- J " K . 8A!@" 7 '*" '
G
C & " '17 L 9 - . . 8 ;*M# #
C NO 7 ;*8A*" !P
E
C 3C 9
P GQ
. 7 E$ Q PC
*R CS3
.C RC
K K
. R C#T- . K Q G$ P .
K
KC9U V Q W
. C
G !'" )
'*C H $
.C 9 '.C H 9 * ;#$
. C
2
Solid lines: First Order (NLSW) Case B:
Thin dash lines: First Order (LSW)
L=10m. Horz. length of
Thick dash lines: FNS−VOF landslide
Circles: Analytical solution
1.5 δ = 1m. Slide max.
t=4.5 vert. thickness
β=5.7 Beach slope
µ=δ/L=0.1
tan(β)/µ=1
1 x’=Lx
t=2.5
η’=η
η (Dimensionless)
t’=sqrt(g/δ) µ t
g=9.81 m/s2
t=1.0
0.5
t=0.5
0
Models:
a) First Order
dx= 1e−3 ( 0.01m)
dt=3e−5 (0.0001sec)
−0.5
b) FNS−VOF
dx= 0.025 (0.250m)
dy=0.125 (0.125m)
dt=3e−5 (0.0001sec)
−1
0 2 4 6 8 10 12 14
x (Dimensionless)
t=0.5 t=1.0
0.6 Analytical
0.6
0.5 First Order LSW
First Order NLSW
0.4 0.4
FNS−VOF
0.3
0.2
y
0.2
0.1 0
0 −0.2
−0.1
−0.2 −0.4
0 0.5 1 1.5 0 0.5 1 1.5 2
0.2
−0.4
0
−0.6
−0.2
−0.4 −0.8
0 0.5 1 1.5 2 2.5 3 0 1 2 3 4
x x
References
1. A. J. Chorin, Numerical solution of the Navier-Stokes equations, Math. Comp.,
22:745–762, 1968.
2. F. H. Harlow and J. E. Welch, Numerical calculation of time-dependent viscous
incompressible flow of fluid with a free surface, The Physics of Fluids, 8:2182–
2189, 1965.
3. P. L.-F. Liu, P. Lynett and C. E. Synolakis, Analytical solution for forced long
waves on a sloping beach, Journal of Fluid Mechanics, 478, 101–109, 2003.
4. Z. Kowalik and T. S. Murty, Numerical Modeling of Ocean Dynamics, World
Scientific Publishing Co., 1993.
CHAPTER 14
Okey Nwogu
Dept. of Naval Architecture and Marine Engineering, University of Michigan
Ann Arbor, MI 48109, U.S.A.
E-mail: [email protected]
1. Introduction
The massive destruction and loss of life associated with recent tsunamis
(Indonesia, 2004; Papua New Guinea, 1998) has underscored the need to
develop and implement tsunami hazard mitigation measures such as
early warning systems in tsunami-prone areas. Given either the seabed
deformation or ocean measurements of water surface elevation by buoys,
hydrodynamic models can be used to forecast in real-time the
propagation and transformation of tsunamis across the world’s ocean
basins and the subsequent inundation of low-lying coastal areas.
Numerical models based on the non-dispersive shallow water
equations are often used to simulate tsunami propagation and runup (e.g.
Titov and Synolakis [6]). The shallow water equations assume that the
vertical fluid motions are much smaller than the horizontal fluid motions
and the resulting fluid pressure is hydrostatic. The hydrostatic pressure
assumption is reasonable for seismic tsunamis since most ocean basins
have depths of O(1km) while earthquake-generated tsunamis typically
have wavelengths of O(100km).
Landslide-generated tsunamis are characterized by shorter
wavelengths [O(10km)], depending on the relative slide speed or depth-
based Froude number. The propagation of tsunamis with water depth to
273
274 O. Nwogu
−
wavelength ratios of O(10 1) can be more accurately modeled by
dispersive water wave evolution equations such as Boussinesq-type
equations (e.g. Nwogu [3]). For dispersive waves, the horizontal
velocities are no longer uniform over the depth and the pressure field is
no longer hydrostatic. The velocity and pressure field is derived by
asymptotically expanding the velocity potential in terms of a frequency
dispersion parameter. The flow field is then substituted into the mass
and momentum equations to yield Boussinesq-type equations.
Boussinesq equations have also been used to simulate landslide-
generated waves by modifying the boundary condition at the seabed (e.g.
Lynett and Liu [2]). This paper documents the extension of a Boussinesq
model to landslide tsunami problems and the validation of the model
with benchmark problems as part of the Third International Workshop
on Long-Wave Runup Models.
φ ( x, y, z, t ) = φα ( x, y, t ) + µ 2 ( zα − z ) [ht + ∇φα ⋅ ∇h ]
1 2
+ µ ( zα + h )2 − ( z + h )2 ∇ 2φα , (1)
2
h +η
ηt + ∇ ⋅ ( h + η ) u = − ht − ∇ ⋅ (h + η ) ( zα + h ) − ∇ht ,
2
( )
uα ,t + g∇η + uη ⋅ ∇ uη + wη ∇wη
− ( uα ,t ⋅ ∇h ) + (h + η )∇ ⋅ uα ,t ∇η
+ ( zα − η ) ∇ ( uα ,t ⋅ ∇h ) + (∇ ⋅ uα ,t )∇h
(2)
1
( zα + h ) − (h + η )2 ∇ (∇ ⋅ uα ,t )
2
+
2
1 1
+ ∇ [ν (h + η )∇ ⋅ uα ] + fw ub ub = − ( zα − η ) ∇htt ,
h +η h +η
3. Benchmark Problems
3.1. Tsunami Runup onto a Complex Three-dimensional Beach
Ch. 9
2 -0.01
y(m)
-0.1
-0.12
-0.08
0.02
-0.04 Ch. 7 0
-0.
-0.01
06
Ch. 5
1
0
0 1 2 3 4 5
x(m)
Fig. 1. 2-D map of bathymetry for laboratory model study of Okushiri tsunami.
Num. Model
0.02
0
0 10 20 30 40 50
Time (s)
Fig. 2. Measured and predicted wave elevation time histories at Gauge #9 for Okushiri
tsunami.
( ) ,
2
h( x, t ) = x tan β + δ exp − 2 xhµ2 / δ tan β − hµ t g / δ (3)
0
0 10 20 30 40 50
x(m)
Elevation (m)
1.6
tnd = 4.5 NLSW
0.8 Boussinesq
Analytical
0
-0.8
0 50 100 150 200
x(m)
Fig. 3. Numerical and analytical profiles of waves generated by sliding Gaussian mass.
References
1. P.L.-F. Liu, P. Lynett, and C. Synolakis, J. Fluid. Mech., 478, 101 (2003).
2. P. Lynett and P.L.-F. Liu, Proc. R. Soc. Lond., 458, 2885 (2002).
3. O. Nwogu, J. Waterw. Port Coastal Ocean Eng., 119, 618 (1993).
4. O. Nwogu and Z. Demirbilek, Tech. Rep. ERDC/CHL 01-25, U.S. Army Engineer R
& D Center, Vicksburg, MS (2001).
5. O. Nwogu and Z. Demirbilek, Coastal Engineering, accepted (2005).
6. V. Titov and C.E. Synolakis, J. Waterw. Port Coastal Ocean Eng., 121, 308 (1995).
This page intentionally left blank
August 26, 2008 17:30 WSPC/Trim Size: 9in x 6in for Review Volume 15˙yasuda
CHAPTER 15
Hiroyasu Yasuda
River Engineering Division, Civil Engineering Research Institute of Hokkaido
Hiragishi 1-3, Toyohira, Sapporo, 062-8602, Japan
E-mail: [email protected]
Tsunami runup onto a complex beach was modeled and analyzed us-
ing a method that allows accurate reproduction of complex topographic
features. The analysis produced accurate results and confirmed that to-
pography greatly affects the flow regime.
1. Introduction
The tsunami caused by the 1993 Hokkaido-Nansei-Oki Earthquake off
southwestern Hokkaido left traces of a great runup at Monai Beach on
Okushiri Island. This runup is regarded as one of the highest of the 20th cen-
tury, and its severity is attributed in part to the very complex 3-dimensional
trough-shaped topography of that beach. This study reproduced the runup
using 2-dimensional analysis that incorporates a “boundary-fitting cell”
(BFC) system. The system allows calculations to flexibly incorporate com-
plex topographic shapes.
279
August 26, 2008 17:30 WSPC/Trim Size: 9in x 6in for Review Volume 15˙yasuda
280 H. Yasuda
1X
k
∂η
− Qi = 0 , (2)
∂t A i=1
where t is the time coordinate, Q is flow rate at the sides of the cell, s is
the plane coordinate, η is the water level, D is the water depth, g is the
gravitational acceleration, n is Manning’s roughness coefficient, A is the
area of the cell where the water level is to be obtained, and k is the number
of sides of the cell.
The numerical scheme used in Eqs. (1) and (2) is based on the Staggered
Leap-Frog scheme, which is an explicit differential scheme.
Experiments, Calculation
Water level (cm)
reflective boundary and tsunami was able to run up onto land. The entire
surface of the left edge was considered an incidence boundary and a water
level was assigned. For the BFC domain, this was the boundary condition:
the water levels obtained from the Cartesian grid calculation were assigned
to cells at the domain boundary.
For both domains, the initial conditions were these: at 0 cm height,
static water surface was assigned; and at points over 0 cm height, the water
depth was set at 0 cm.
3. Analysis Results
3.1. Cartesian Grid Domain
The accuracy in calculating sea level variations is an indication of the valid-
ity of boundary conditions in the BFC domain. Figure 1 shows time-lapse
waveforms of water levels obtained from the hydraulic experiment and nu-
merical analysis. The three points in Fig. 1 show that the time-lapse water
levels for both methods are in close agreement.
282 H. Yasuda
t =17.70 sec, .
M aximum height: t = 17.90(s)
2.50
Maximum runup at Monai 7.10cm (by BFC)
7.
50
0.0
(by Rectangle Grid) 5.
00
2.
50
2.5
0
BFC Area 2.50
7.5
0
5. Max. heig
ht:
00 7.30cm
z=
5.
00
y 7.
5
2. 0cm
50
0.
cm
0
x=3.00
0c
y=1.00 7.50
m
x
Water Level (cm) D epth of W ater (cm )
-2.25 -1.50 -0.75 0.00 0.75 1.50 2.25 0.0 0.5 1.0 1.5 2.0 2.5 3.0
(a) Water level distribution at Monai Beach (b) Flow regime of flood from the
at the time of maximum flood runup BFC method
Fig. 2. Analysis results for the tsunami runup at Monai Beach. (a) shows the results of
analysis using the Cartesian grid: the maximum runup height at Monai Beach is 7.1 cm.
In BFC analysis, the water level on these results are obtained.
4. Conclusion
This numerical analysis calculated tsunami runup onto a complex beach,
using the BFC system, which accurately reproduces complex topographic
features. The calculations obtained maximum runup heights that were
similar to results, without 3-dimensional calculation. This shows the im-
portance of using discretization methods that are appropriate for the flow
characteristics and topography.
References
1. H. Yasuda, M. Shirato, C. Goto and T. Yamada, Journal of Hydraulics, Coastal
and Environment Engineering No.740/II–64 (2003) (in Japanese).
August 26, 2008 17:43 WSPC/Trim Size: 9in x 6in for Review Volume 16˙frandsen
CHAPTER 16
J. B. Frandsen
School of Civil Engineering, The University of Sydney
NSW 2006, Australia
Email: [email protected]
283
August 26, 2008 17:43 WSPC/Trim Size: 9in x 6in for Review Volume 16˙frandsen
284 J. B. Frandsen
x x + ∆x
f1 f2
t _c
+c
(1 − α )f2 (1 − α ) f1
t
α f1 α f2
f f
t + ∆t 2 1
∆x
j (x + 2 ) = c ( f2 f1 )
We introduce the single phase 1-D LBGK formulation for shallow water
flows on sloped beds as
2
∂fi ∂fi (fi − fieq ) ci ∆t ∂hb 1
+ ci =− − gh , where Ni = ci ci ,
∂t ∂x τ N i c2 ∂x c2 i=0
(3)
with Ni as a constant which depends on the lattice geometry and is defined
Ni = 2 for the D1 Q3 velocity set. The first term represents the effect of
the local change of the fluid motion in time and the second term describes
the convection. Note the convection operator of the Lattice Boltzmann
Method (LBM) is linear. This property originates from kinetic theory. The
first term of (3) on the right-hand-side (RHS) is the non-equilibrium distri-
bution function which describes the effect of the collisions (Succi (2001)).
The second term on the RHS represents the force term F . The time-scale
parameter τ = 3 ν/c2 + ∆t/2 where ν is the kinematic viscosity of wa-
ter and c = ∆x/∆t (the time step is ∆t and the grid spacing is ∆x). It
describes the collisional relaxation to the local equilibrium and 1/τ is the
collision frequency. In the present formulation τ is limited to a single value
(BGK approximation). To ensure numerical stability τ > ∆t/2 when using
the standard LBGK scheme. Further, we note that assuming a constant
value of τ is considered to be a crude approximation. The equilibrium dis-
tribution functions fieq represents the invariant function under collision (no
gradients are involved) and is dependent upon the microscopic velocity ci .
It is essential that the equilibrium distribution functions satisfy the Navier-
Stokes equation for any water wave model. In our case, the depth-averaged
shallow-water equations with rotational flow. The fieq functions can be de-
rived from the Boltzmann equation using a Chapman-Enskog expansion.
Details of the Chapman-Enskog method for the classical Boltzmann equa-
tion can be found in Gombosi (1994). The LBGK based fieq expressions
are velocity dependent, both at the micro and macroscale levels. A new set
of equilibrium distribution functions following the 1-D flow approximation
has been derived and are given as,
g h2 h u2 gh2 hu h u2
f0eq = h − − ; eq
fi=1,2 = + c i + , (4)
2c2 c2 4c2 2c 2c2
where the discrete velocity set is ci=1 = 1 and ci=2 = −1, respectively.
August 26, 2008 17:43 WSPC/Trim Size: 9in x 6in for Review Volume 16˙frandsen
286 J. B. Frandsen
For the present application, it is noted that the pressure field exhibit near
hydrostatic behavior and therefore the dynamic pressure term (hu2 ) is neg-
ligible. The macroscopic variables of the free surface (ζ) and the depth
averaged velocities (u) are calculated as the first and second moments of
the distribution function,
2
2
1
ζ= f i − h0 and u = ci f i . (6)
i
(h0 + ζ) i
of low Mach numbers where c2s = ∂p/∂ρ = c2 /3 is the speed of sound and
ρ is the fluid density.
In the following we use the Chapman-Enskog technique to show that
the present LB formulation recovers the NLSW equations in rotational flow.
The Chapman-Enskog expansion assumes that the convection and diffusion
processes operate at different time scales. The diffusion time scale being
much slower than the convective counter part. First, we perform a Taylor
expansion of (7) and retain the terms up to second-order,
∂ ∂ ∆t2 ∂ ∂ fi − fieq
∆t( + ci )fi + ( + ci )2 fi = − + Fi ∆t . (8)
∂t ∂x 2 ∂t ∂x τ
Then we expand the particle distribution function as
(1) (2)
fi = fieq + εfi + ε2 fi + ··· . (9)
Substituting (9) into (8), the equations with the first-order and second-order
of the small parameter ε are obtained as
(1)
∂fieq ∂f eq f
ε1 : + ci i = − i + Fi , (10)
∂t ∂x τ
eq (1) (1)
2 ∂fi ∂fieq 1 ∂fi ∂fi
ε : + ci +ε 1− · + ci
∂t ∂x 2τ ∂t ∂x
(1) (2)
+ εfi fi
=− + Fi . (11)
τ
The conservation of mass and momentum requires
(n) (n)
fi = 0, fi ci = 0, n = 1, 2, · · · . (12)
i i
From (6) and (12), we obtain the macro continuity equation (1). Next, we
multiply (11) by ci and sum over i giving
∂f eq (1) (1)
∂fieq 1 ∂fi ∂fi
ci i
+ ci ci +ε 1− · ci + ci
i
∂t i
∂x 2τ i
∂t ∂x
(1)
fi + εfi
(2)
=− ci + ci Fi . (14)
i
τ i
August 26, 2008 17:43 WSPC/Trim Size: 9in x 6in for Review Volume 16˙frandsen
288 J. B. Frandsen
Then we use (6), (12) and ci ci fieq = 12 gh2 + h u2 leading to the macro
i
momentum equation of the NLSW equation (1),
∂(hu) ∂(hu2 ) ∂ h2 ∂Fν
+ = −g ( ) + + Fi , (15)
∂t ∂x ∂x 2 ∂x
∂(h2 /2) 2
)
where Fν = ν · ∂(hu)
∂x + ∂t + 3 ∂(hu
∂t and the kinematic viscosity is
c2
ν= 3 (τ − 12 ).
290 J. B. Frandsen
6. Solution Procedure
Initially we prescribe the waveform of Fig. 3(b) together with the non-
physical values of the velocity u = 0 and the distribution functions fi = fieq .
We solve (7) using a non-splitting operator scheme for both the standard
LBGK and the FD scheme. The solution procedure is outlined in Fig. 2.
The links relevant to the LB schemes herein are labeled: standard LBGK
and FD LB , respectively. The hydrodynamic moments of the equilibrium
distribution functions of (5) are conserved at every time step. It is assumed
that the Mach number Ma = u/cs 1, where cs is the speed of sound. The
August 26, 2008
17:43
WSPC/Trim Size: 9in x 6in for Review Volume
A 1-D Lattice Boltzmann Model
standard LBGK
1.17
FD scheme
1.7
1.20
16˙frandsen
291
Fig. 2. Solution procedure. Non-splitting operator scheme applies to the standard LBGK and FD LB solvers.
August 26, 2008 17:43 WSPC/Trim Size: 9in x 6in for Review Volume 16˙frandsen
292 J. B. Frandsen
Peclet number Pe = u ∆x/ν < 2 and the Courant number Cr = u ∆t/∆x <
1 are also satisfied.
4
3
z
ζ [m]
ζ
x
11111111111111111
00000000000000000
00000000000000000
11111111111111111 z0
00000000000000000
11111111111111111
h0
00000000000000000
11111111111111111
00000000000000000
11111111111111111
hb
00000000000000000
11111111111111111
00000000000000000
11111111111111111
−8.81
−10
00000000000000000
11111111111111111 0 8.25 20.55 35 50
L (a) x [km] (b)
Fig. 3. (a) Definition sketch of wave runup study. (b) Initial wave profile.
Initially the velocity in the flow domain is zero and the free-surface de-
scribed by the form of a leading depression N-wave shape, typically caused
by an offshore submarine landslide,
2 2
ζ = a1 e(−k1 (x−x1 ) ) − a2 e(−k2 (x−x2 ) )
and u = 0 at t=0 , (22)
where a1 = 13 a2 = 0.006, k1 = 19 k2 = 0.4444, x1 = 4.1209 and x2 = 1.6384,
after Carrier et al. (2003). The initial wave profile is also shown in Fig. 3(b).
8. Results
We shall, in the following, present 1-D wave runup results based on our stan-
dard LBE and second-order accurate FD LBM solvers. First we undertook
August 26, 2008 17:43 WSPC/Trim Size: 9in x 6in for Review Volume 16˙frandsen
294 J. B. Frandsen
200
x [m]
−200
10
0
us [m/s]
−10
−20
100 160 175 220 280
t [s] (b)
230
200
x [m]
−166
−200
100 170 215 280
t [s] (a)
10
7.4
0
us [m/s]
−15.2
−20
100 161 181 215 280
t [s] (b)
Fig. 5. FD LB solutions. Grids: − −, 1000 nodes; −, 5000 nodes; − · −, 6000 nodes.
August 26, 2008 17:43 WSPC/Trim Size: 9in x 6in for Review Volume 16˙frandsen
296 J. B. Frandsen
15 15
10 10
5
5
0
0
ζ [m]
ζ [m]
−5
−5
−10
−10
−15
−15 −20
−20 −25
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
t [s] t [s]
(a) (b)
15 18
16
10
14
5
ζ [m]
ζ [m]
12
0
10
−5
8
−10 6
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
t [s] t [s]
(c) (d)
10 20
8
15
6
10
4
5
2
u [m/s]
u [m/s]
0 0
−2
−5
−4
−10
−6
−15
−8
−10 −20
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
t [s] t [s]
(e) (f)
25 30
20
20
15
10 10
5
u [m/s]
u [m/s]
0
0
−5 −10
−10
−20
−15
−20
−30
−25
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
t [s] t [s]
(g) (h)
Fig. 6. FD LB free-surface (a–d) elevation and (e–h) velocity solutions of wave form
transformation x ∈ [−166, 500] m where x = −166 m and x = +230 m represent maxi-
mum runup and run-down locations. (a) Offshore, (b–c) Approaching shoreline, (d) The
shoreline motion behavior towards maximum runup. (a,e) −, 500 m; − −, 400 m; − · −,
300 m. (b,f) −, 230 m; − −, 200 m; − · −, 150 m. (c,g) −, 100 m; − −, 50 m; − · −, 0 m.
(d,h) −, −50 m; − −, −100 m; − · −, −166 m.
August 26, 2008 17:43 WSPC/Trim Size: 9in x 6in for Review Volume 16˙frandsen
10 20
8
15
6
10
4
5
2
u [m/s]
u [m/s]
0 0
−2
−5
−4
−10
−6
−15
−8
−10 −20
−20 −10 0 10 20 −30 −20 −10 0 10 20 30
ζ [m] ζ [m]
(a) (b)
20 20
15 15
10 10
5 5
u [m/s]
u [m/s]
0 0
−5 −5
−10 −10
−15 −15
−20 −20
−30 −20 −10 0 10 20 30 −30 −20 −10 0 10 20 30
ζ [m] ζ [m]
(c) (d)
20 20
10 10
u [m/s]
u [m/s]
0 0
−10 −10
−20 −20
30 30
15 15
300 300
0 0
200 200
−15 100 −15 100
−30 0 −30 0
ζ [m] t [s] ζ [m] t [s]
(e) (f)
20 20
10 10
u [m/s]
u [m/s]
0 0
−10 −10
−20 −20
30 30
15 15
300 300
0 0
200 200
−15 100 −15 100
−30 0 −30 0
ζ [m] t [s] ζ [m] t [s]
(g) (h)
Fig. 7. FD LB velocity and free-surface trajectories of wave form transformation x ∈
[−166, 500] m where x = −166 m and x = +230 m represent maximum runup and
run-down locations. (a) Offshore, (b–c) Approaching shoreline, (d) The shoreline motion
behavior towards maximum runup. (a,e) −, 500 m; − −, 400 m; − · −, 300 m. (b,f) −,
230 m; − −, 200 m; − · −, 150 m. (c,g) −, 100 m; − −, 50 m; − · −, 0 m. (d,h) −, −50
m; − −, −100 m; − · −, −166 m.
August 26, 2008 17:43 WSPC/Trim Size: 9in x 6in for Review Volume 16˙frandsen
298 J. B. Frandsen
200
x [m]
0
−200
0
us [m/s]
−10
−20
100 160 175 220 280
t [s] (b)
20
t=220s
10
ζ [m]
0
t=175s
−10
t=160s
−20
−200 0 1000
x [m] (c)
Runup [m] −147 (t = 226 s) −166 (t = 215 s) −164 (t = 216 s) −164 (t = 217 s)
Run-down [m] +226 (t = 170 s) +230 (t = 170 s) +239 (t = 172 s) +242 (t = 173 s)
Shoreward velocity [m/s] −11.29 (x = 130 m) −15.22 (x, t) = (181 m,180 s) −17 (x = 176 m) −15.76 (x, t) = (174 m,177 s)
Offshore velocity [m/s] +7.09 (x = 179 m) +7.44 (x, t) = (161 m,162 s) +7.28 (x = 164 m) +7.28 (x, t) = (164 m,162 s)
No. of nodes 100,000 5,000 3,171
∆x [m] 0.5 10 15.78
∆t [s] 0.002 0.00015 0.09
τ [−] 0.732 0.5 —
CPU time [s] 6702 1975 ≈30
16˙frandsen
299
August 26, 2008 17:43 WSPC/Trim Size: 9in x 6in for Review Volume 16˙frandsen
300 J. B. Frandsen
70 40
35
60
30
50
25
f [m]
f0 [m]
40 20
0
15
30
10
20
5
10 0
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
t [s] (a) t [s] (b)
25 9
8
20 7
6
15
5
f0 [m]
f0 [m]
4
10
3
5 2
0 0
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
t [s] (c) t [s] (d)
−3 −4
x 10 x 10
2
5
1.5
1
0
0.5
f1 [m]
f1 [m]
−0.5 −5
−1
−1.5 −10
−2
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
t [s] (e) t [s] (f)
−4 −4
x 10 x 10
4 1
2 0.5
0 0
f [m]
f [m]
−2 −0.5
1
−4 −1
−6 −1.5
−8 −2
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
t [s] (g) t [s] (h)
Fig. 9. The temporal variations of the distribution function (a–d) f0 and (e–h) f1 for
x ∈ [−166, 500] m where x = −166 m and x = +230 m represent maximum runup and
run-down locations. (a,e) Offshore, (b–c, f–g) Approaching shoreline, (d,h) The shoreline
motion behavior towards maximum runup. (a,e) −, 500 m; − −, 400 m; − · −, 300 m.
(b,f) −, 230 m; − −, 200 m; − · −, 150 m. (c,g) −, 100 m; − −, 50 m; − · −, 0 m. (d,h)
−, −50 m; − −, −100 m; − · −, −166 m.
August 26, 2008 17:43 WSPC/Trim Size: 9in x 6in for Review Volume 16˙frandsen
0
(a) (b)
0
(c) (d)
−4 −4
x 10
5
0
[m]
feq
1
[m]
−5
−10
2 1
0.5
0
0
feq [m]
feq [m]
−2
1
−0.5
−4
−1
−6 −1.5
−8 −2
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
t [s] (g) t [s] (h)
Fig. 10. The temporal variations of the equilibrium distribution function (a–d) and f0eq
(e–h) f1eq . (a,e) Offshore, (b–c, f–g) Approaching shoreline, (d,h) The shoreline motion
behavior towards maximum runup. (a,e) −, 500 m; − −, 400 m; − · −, 300 m. (b,f) −,
230 m; − −, 200 m; − · −, 150 m. (c,g) −, 100 m; − −, 50 m; − · −, 0 m. (d,h) −, −50 m;
− −, −100 m; − · −, −166 m.
August 26, 2008 17:43 WSPC/Trim Size: 9in x 6in for Review Volume 16˙frandsen
302 J. B. Frandsen
−4 −4
x 10 x 10
1.5
4
1
2
0.5
0
f −feq [m]
f −feq [m]
0
0
0
−2
−0.5
0
0
−1 −4
−1.5 −6
−2 −8
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
t [s] (a) t [s] (b)
−4 −4
x 10 x 10
1
1
0.5
0 0
f −feq [m]
[m]
−0.5
f0−feq
0
−1
0
−1
−2
−1.5
−2
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
t [s] (c) t [s] (d)
−5 −4
x 10 x 10
10
8 2
6
1
4
f −feq [m]
[m]
2
f1−feq
1
0
1
−2
−1
−4
−6
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
t [s] (e) t [s] (f)
−4 −5
x 10 x 10
2 10
8
1.5
6
1
4
f −feq [m]
f −feq [m]
0.5 2
1
1
1
0
0
−2
−0.5
−4
−1 −6
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
t [s] (g) t [s] (h)
Fig. 11. The temporal variations of the collision term (a–d) and (e–h) f1 − f1eq . f0 − f0eq
(a,e) Offshore, (b–c, f–g) Approaching shoreline, (d,h) The shoreline motion behavior
towards maximum runup. (a,e) −, 500 m; − −, 400 m; − · −, 300 m. (b,f) −, 230 m;
− −, 200 m; − · −, 150 m. (c,g) −, 100 m; − −, 50 m; − · −, 0 m. (d,h) −, −50 m; − −,
−100 m; − · −, −166 m.
August 26, 2008 17:43 WSPC/Trim Size: 9in x 6in for Review Volume 16˙frandsen
−4 −4
x 10 x 10
4 4
3 3
2 2
1 1
f −feq [m]
[m]
0 0
f1−feq
0
1
0
−1 −1
−2 −2
−3 −3
−4 −4
−200 0 200 400 600 800 −200 0 200 400 600 800
x [m] (a) x [m] (b)
Fig. 12. Snapshots of collision term variations for the discrete particle velocities (a) c0
and (b) c1 . −, t = 160 s; − · −, t = 175 s; − −, t = 220 s.
(a) (b)
Fig. 13. Distribution functions at the maximum runup location (t = 175 s), for (a) c0
and (b) c1 . −, fi ; − −, fieq ; − · −, (fi − fieq ).
August 26, 2008 17:43 WSPC/Trim Size: 9in x 6in for Review Volume 16˙frandsen
304 J. B. Frandsen
collision term fi − fieq does not follow any macroscopic patterns. Due to
the coarse grid, we can unfortunately not make any further interpretation in
terms of bridging the behavior of the mesoscopic and macroscopic variables.
Figure 12 shows the variations of the collision term (fi −fieq ) for the discrete
microscopic velocities ci at t = 160 s, 175 s, and 220 s. We can observe that
the deviations from equilibrium are largest in the run-down region near
the occurrence at maximum velocity. This is further confirmed in curves
of Fig. 13 and thus agrees with the observations of the macroscopic vari-
ables. Finally, we should note that the LB maximum velocity, maximum
runup and run-down occurred at slightly different instances than the semi-
analytical solution of Carrier et al. and thus the LB extreme values would
deviate from the solutions shown.
400
1/20
1/10
200
1/5
x [m]
−200
−400
0 100 200 300
t [s]
for all three slopes of the FD LB model. For example, one would expect
that the long wave models would be less accurate for relatively steep slopes,
however, the steepest slope of 1/5 we tested showed good agreement be-
tween the LB and the NLSW solutions. The largest discrepancy occurred
at maximum run-down. The grid convergence study of Pedersen (Chapter
17) showed no runup, i.e. breaking around maximum withdrawal where pre-
dicted with minor differences between the NLSW and Boussinesq solutions
when ∂hb /∂x = 1/20 whereas the LB model did predict shoreline motion
profile similar to the 1/10 case (just enlarged in shape). It is unclear why
10 20
10
us [m/s]
ζ [m]
0
−10
−4 −20
−0.05
0 5 7 100 152 210 280
x [m] 4
x 10 (a) t [s] (b)
4 20
10
0
us [m/s]
ζ [m]
−10
−10 −20
−0.05
0 5 7 100 166 182 280
x [m] x 10
4
(c) t [s] (d)
10 20
10
us [m/s]
ζ [m]
0
−10
−4 −20
−0.05
0 5 7 100 142 207 280
x [m] x 10
4
(e) t [s] (f)
Fig. 15. Shoreline velocity time history (b,d,f) based on initial wave forms (a,c,e). −,
FD LB solutions; −−, Carrier et al. (2003).
August 26, 2008 17:43 WSPC/Trim Size: 9in x 6in for Review Volume 16˙frandsen
306 J. B. Frandsen
20 20
10 10
us [m/s]
us [m/s]
0 0
−10 −10
−20 −20
−400 −200 0 200 400 −400 −200 0 200 400
x [m] (a) x [m] (b)
20 20
10 10
u [m/s]
us [m/s]
0 0
s
−10 −10
−20 −20
−400 −200 0 200 400 −400 −200 0 200 400
x [m] (c) x [m] (d)
Fig. 16. Shoreline velocity versus shoreline location. (a) Initial Gaussian wave form,
case a, (b) negative Gaussian shape, case b, (c) leading depression N-wave, case c,
(d) waveform caused by submarine landslide, case d (bench mark test 1). −, FD LB
solutions; −−, Carrier et al. (2003).
Pedersen’s model solutions did not predict runup for the slope of 1/20. The
other beach slope cases showed similar trajectory effects on the shoreline
motion when comparing the LB with the NLSW and Boussinesq solutions
(Pedersen (Chapter 17)).
Finally, we tested the LB solver for other initial wave forms, after Car-
rier et al. (2003). We found fairly good agreement for the 5000 LB node
solutions, as shown in Figs. 15 and 16.
308 J. B. Frandsen
shore and at the open boundary. The near shore physics could be repre-
sented better, as the bed friction and sediment transport effects are not
included in our results. Nor is the effect of wave breaking considered. To
delay wave reflections at the open boundary, we adopted a simple extrapo-
lation equation at the open boundary and extended the flow domain with
500 m to mimic some form of sponge layer. Other algorithms could be
explored to improve accuracy and efficiency.
In a final remark, we should emphasize that our intention with the
LB modeling approach is to use it in local areas where physical details
are needed such as the nearshore regions. Therefore a realistic goal and
potential improvement to the current tsunami model literature would be
to propose a LB-model coupled with a macro-scale level model for large
domain tsunami model predictions.
We attempted to write-up the LB solutions of benchmark test case 1
as self-contained as possible. We realize, however, the shortcomings of ex-
planations, and would like to refer the reader to further discussions, details
and references about the LB method which are given in Chapter 5 in this
volume.
References
Balzano, A. (1998) Evaluation of methods for numerical simulation of wetting
and drying in shallow water flow, Coastal Engineering 34, pp. 83–107.
Bhatnagar, P., Gross, E. and Krook, M. A. (1954) A model for collision processes
in gases i: small amplitude process in charged and neutral one component
systems, Physics Review A 94, pp. 511–526.
Carrier, G. F., Wu, T. T. and Yeh, H. (2003) Tsunami run-up and draw-down on
a plane beach, Journal of Fluid Mechanics 475, pp. 79–99.
Frandsen, J. B. (2008) “Free-surface Lattice Boltzmann modeling in single phase
flows”, in Advanced Numerical Models for Simulating Tsunami Waves and
Runup, Advances in Coastal and Ocean Engineering, Vol. 10, eds. Liu, P.
L.-F., Yeh, H. and Synolakis, C. (World Scientific Publishing Co.).
Gombosi, T. I. (1994) Gaskinetic Theory (Cambridge University Press).
Guo, Z. and Zhao, T. S. (2003) Explicit finite-difference lattice Boltzmann method
for curvilinear coordinates, Physical Review E 67, 066709.
Junk, M., Klar, A. and Luo, L.-S. (2005) Asymptotic analysis of the Lattice
Boltzmann equation, Journal of Computational Physics 210, pp. 676–704.
Kennedy, A. B., Chen, Q., Kirby, J. T. and Dalrymple, R. A. (2000) Boussinesq
modeling of wave transformation. breaking and runup. I:1D, Journal of
Waterway, Port, Coastal, and Ocean Engineering 126(1), pp. 39–47.
Lynett, P. J., Wu, T.-R. and Liu, P. L.-F. (2002) Modeling wave runup with
depth-integrated equations, Coastal Engineering 46, pp. 89–107.
August 26, 2008 17:43 WSPC/Trim Size: 9in x 6in for Review Volume 16˙frandsen
CHAPTER 17
G. Pedersen
Mechanics Division, Department of Mathematics, University of Oslo
E-mail: [email protected]
1. The Models
The main runup model of this section is based on Lagrangian Boussinesq
equations that are fully nonlinear and possess standard dispersion proper-
ties. It is run in both hydrostatic (NLSW) and dispersive mode. For com-
parison we employ also an Eulerian FDM code for the standard Boussinesq
equations and a Boundary Integral Method (BIM) for full potential the-
ory. The first is used for linear computations and is without any particular
runup feature; runup heights are found from the vertical elevation at the
shore. The BIM is based on Cauchy’s formula, Lagrangian surface nodes
and spline interpolation. All models allow for a variable spatial resolution,
but this feature is employed only for the BIM and the Eulerian models.
Systematic grid refinement is employed for all applications reported.
Expressed in terms of the depth averaged velocity, u, and the total
depth, H the Lagrangian long wave equations read
∂x ∂u H ∂H ∂h
H = H0 (a), (1−r1 )= −g +g −r2 +S1 +S2 (b),
∂a ∂t H0 ∂a ∂x
3
∂ h ∂ h ∂u ∂ 2 h ∂H
2 2
∂3h
1 ∂ h 1
S1 = H 2 , S2 = − 2u + 2 − +Hu ,
2 ∂t ∂x g ∂t∂x ∂t ∂t ∂t∂x ∂t ∂t∂x2
(1)
where a is the Lagrangian coordinate, h is the equilibrium depth and
H0 = H(a, 0). The position x relates to the velocity according to ∂x/∂t = u.
Only the shallow water (hydrostatic) terms are spelled out in (1b). A more
complete description of the equations in absence of a slide, including the
311
August 26, 2008 17:46 WSPC/Trim Size: 9in x 6in for Review Volume 17˙Geir
312 G. Pedersen
dispersion terms (r1 and r2 ), is found in Jensen et al. (2003)2 and refer-
ences therein. The principal forcing due to a time dependent bottom is
the modified source of horizontal momentum from the second term on the
right hand side of (1b). Higher order source terms, S1 and S2 , have been
included particularly for benchmark 3. Replacing H by h in S1 we obtain
the linear part that is employed in the Eulerian Boussinesq equation, while
S2 is exclusively nonlinear.
A moving shoreline is associated with a fixed a, where the condition
H = 0 is employed. The set (1) is solved by finite differences on a grid
that is staggered in space and time. The hydrostatic (NLSW) version is
explicit, while dispersion requires implicitness. No smoothing or filtering is
employed with the long wave models.
Bouss(NL) ref.
ref. 220s NLSW
NLSW
Bouss(L)
SW(L)
BIM(NL)
175s
160s
t/sec x/km
Fig. 1. Left: withdrawal (m). Right: surface (m) at selected times.
2. Benchmark Problem 1
The reference solution1 exists in a semi-infinite domain, but have been
made available until 50 km from the coast. We add a deep water region
(x > 50 km) to avoid that reflections from the seaward boundary reach the
shore region for t < 280 s. The alternatives are open boundary conditions
or sponge layers that may introduce additional errors and uncertainties.
The inundation lengths shown in Fig. 1 are all close to convergence
(increment ∆a = 16 m for the nonlinear long wave models). The NLSW
solution agrees very well with the reference solution (ref.). Dispersion has
some effect and reduces the draw-down and increases the runup. The maxi-
mum runup and drawdown are nearly equal for linear and nonlinear models.
August 26, 2008 17:46 WSPC/Trim Size: 9in x 6in for Review Volume 17˙Geir
Since the initial condition inherits very small elevation/depth ratios this is
to be expected in the hydrostatic approximation4 . That also the linear
(L) and nonlinear (NL) Boussinesq models yield nearly identical extrema
indicates that dispersion and nonlinearity are important in the deep and
shallow regions, respectively, with no or little overlap.
The Boussinesq and full potential (BIM) models yield nearly identi-
cal solutions near shore, while there are small discrepancies in the waves
propagating into deeper water (not shown).
From a modeling point of view the NLSW solution is the most inter-
esting (challenging) one. At maximum drawdown there is a (near) cusp on
the inundation curve corresponding to huge accelerations, a solution that
is close to being “weak” and slow convergence of the numerical model. In
Fig. 2 we observe that the accuracy for the velocity is much poorer than
for the inundation length. The convergence of the numerical solution is
particularly slow close to t = 173 s where the reference solution is slightly
multi-valued. These features have not been pursued further. Moreover, for
the maximum withdrawal the NLSW model displays a linear convergence
except for extremely fine resolutions. For the smooth Boussinesq solution
we obtain quadratic convergence (dashes in right panel).
Bouss(NL)
15.8 NLSW
7.9 ref.
3.9
2.5
252
63 ref.
16
8
ref.
t/sec ∆a (m)
Fig. 2. Left and mid panels: shoreline velocities (m/s). NLSW computations, marked by
∆a (m), are compared to the reference solution. Mid panel: blow-up of region (172, 176)×
(−18, 5). Right: Convergence of maximum withdrawal (m).
3. Benchmark Problem 3
In this problem3 a slide with time dependent shape and acceleration pen-
etrates the water. All results for this benchmark is given in meters and
August 26, 2008 17:46 WSPC/Trim Size: 9in x 6in for Review Volume 17˙Geir
314 G. Pedersen
p
dimensionless time units, δ/gµ2 , as defined in the reference. There is a
singularity at the rear of the slide body (x = 0), while it is undefined for
x < 0. Hence, if the fluid reaches x = 0 before the height of the slide body
becomes negligible at this point, conceptual and practical problems arise in
the numerical solution. For case A this occur before the first runup max-
imum. This is not so for case B for which the NLSW and Boussinesq(∗ )
equations yield:
The solutions for the different models are shown in Fig. 3. For case A and
t = 1.5 the reference solution agree closely with the nonlinear solutions ex-
cept for a small discrepancy very close to the shore. The differences between
the linear and nonlinear solutions are significant for case B at t = 2.5, while
effects of dispersion are visible only at the front of the offshore propagating
wave. The Eulerian model SW(L) is close to the reference solution for both
cases (difference up to 0.0025 m with shoreline resolution ∆x = 15 m for
case A and t = 1.5). For case B inclusion of the dispersive source term, S1 ,
ref. ref.
Bouss(NL) Bouss(NL)
NLSW NLSW
SW(L) SW(L)
x/m x/m
Fig. 3. η (m); comparison of models.
reduces the maximum height of the outgoing wave from 1.42 m to 1.33 m at
t = 4.5. S2 then increases the height by 0.01 m. For case A and t = 1.5 the
term S1 reduces the wave height from 0.698 m to 0.692 m, while the effect
of S2 is negligible (less than 10−5 m).
August 26, 2008 17:46 WSPC/Trim Size: 9in x 6in for Review Volume 17˙Geir
4. Remarks
Lagrangian (and ALE) models are undoubtedly useful for idealized stud-
ies, as benchmarks 1 and 3, where they produce very accurate results.
However, such models have poorer prospects for complex tsunami studies.
Further discussion and references on related models are given in Chapter 1
in this volume (Pedersen)4 .
References
1. G. F. Carrier, T. T. Wu and H. Yeh, J. Fluid Mech. 475, 79 (2003).
2. A. Jensen, G. Pedersen and D. J. Wood, J. Fluid. Mech. 486, 161 (2003).
3. P. L.-F. Liu, P. Lynett and C. E. Synolakis, J. Fluid Mech. 478, 101 (2003).
4. G. Pedersen, Modeling runup with depth integrated equation models, in
Advanced Numerical Models for Simulating Tsunami Waves and Runup, eds.
P. L.-F. Liu, H. Yeh and C. Synolakis (World Scientific Publishing Co., 2008),
p. 3–41.
This page intentionally left blank
April 3, 2008 3:12 WSPC/Trim Size: 9in x 6in for Review Volume divider
Appendix
The following article was omitted from Volume 9 of the review series,
entitled “PIV and Water Waves” (eds. J. Grue, P. L.-F. Liu and G. K.
Pedersen).
This page intentionally left blank
APPENDIX
Joe Longo, Jun Shao, Marty Irvine, Lichuan Gui and Fred Stern
IIHR Hydroscience and Engineering
300 S. Riverside Drive
The University of Iowa, Iowa City, IA, USA 52242
E-mail: [email protected]
1. Introduction
Focus of experimental ship hydrodynamics research is moving into
unsteady viscous flows in support of unsteady Reynolds-averaged Navier-
Stokes (RANS) code development for simulation-based design. The
forward-speed diffraction problem1, i.e. restrained body advancing in
regular headwaves, is a building block problem towards ultimate goal of
physical understanding and simulation of viscous nonlinear seakeeping
and 6DOF maneuvering. The present study is precursory for measurement
of the unsteady nominal wake of a naval combatant and provides
documentation of the data-acquisition and reduction procedures along
with detailed data of the incident headwave and comparisons with 2D
progressive-wave theory.
319
320 J. Longo et al.
2. Test Design
The tests are conducted in the 100 × 3 × 3 m IIHR tank, which is equipped
with a drive carriage and trailer, wavemaker, and moveable wave
dampeners.
(a) (b)
Fig. 1. Experimental setup for towed, unsteady PIV measurements.
z
-0.020 -0.020 -0.020 -0.020
Fig. 2. Vector field for four time instances in one encounter period.
4. Future Work
Future work consists of steady and unsteady PIV measurements and UA of
the nominal wake plane of model 55126 for the same conditions presented
herein. The data and UA will be archived at www.iihr.uiowa.edu/~towtank.
322 J. Longo et al.
Acknowledgments
This research was sponsored by the Office of Naval Research under Grant
N00014-01-1-0073 under the administration of Dr. Pat Purtell.
References
1. H. Rhee and F. Stern, Int. J. Num. Meth. Fluids, 37 (2001) 445.
2. L. Gui, J. Longo and F. Stern, Exp. Fluids, 31 (2001) 336.
3. L. Gui, J. Longo, B. Metcalf, J. Shao and F. Stern, Exp. Fluids, 31 (2001) 674.
4. L. Gui, J. Longo, B. Metcalf, J. Shao and F. Stern, Exp. Fluids, 32 (2002) 27.
5. Test uncertainty: Instruments and apparatus, ASME PTC (Performance Test Code)
19.1–1998 (1998), 112.
6. J. Longo, J. Shao, M. Irvine and F. Stern, 24th Symposium on Naval Hydrodynamics
(2002).