Вы находитесь на странице: 1из 33

Sandbanks for coastal protection:

implications of sea-level rise

Part 3: wave modelling��

Cui-Ping Kuang and Peter Stansby

February 2006

Tyndall Centre for Climate Change Research Working Paper 88


Sandbanks for coastal protection:
implications of sea-level rise
part 3: wave modelling

Cui-Ping Kuang and Peter Stansby

Manchester Centre for Civil and Construction Engineering, UMIST,


Manchester M60 1QD, UK

Tyndall Centre Working Paper No. 88

1
SUMMARY
An offshore wave energy spectrum is tracked inshore taking account of shoaling and
refraction due to bathymetry and currents, energy losses due to whitecapping and
bottom friction, wind-wave generation, quadruplet wave-wave interactions; and the
nearshore processes of depth-induced breaking and triad wave-wave interactions. The
finite-element code TOMAWAC is used which is efficient and well suited for large-
scale modelling. Diffraction is not modelled and may be significant over shallow
sandbanks of interest here. Comparisons are made with wave fields measured
experimentally over elliptical and circular shoals. Predictions by TOMAWAC are
quite accurate for directional waves with broad spreading, but less so with narrow
spreading. A diffraction code including otherwise similar physics, ARTEMIS,
performs better for narrow spread waves, although not necessarily for broad spread
waves. Field tests for TOMAWAC are also encouraging. Given that TOMAWAC
with broad spread waves appears to avoid, or at least considerably reduce, the need
for diffraction, the nearshore wave fields due to broad and narrow spread waves
offshore were compared for the East Anglian coastal region. Close to the shore, wave
heights were almost identical, suggesting that an efficient ray tracking code like
TOMAWAC with broad spread waves is a valid tool for nearshore wave field
prediction.

1. INTRODUCTION
Wave propagation in coastal zones drives most coastal processes and is thus
fundamental to a regional coastal simulator. There are many complex effects for
linear, random, directional, wave fields while nonlinear, wave-breaking and
boundary-layer effects are significant in shallow water, particularly inshore. There are
broadly two classes of wave model: phase-resolving and phase-averaged. As the name
implies, phase-resolving resolves wave flows within a wave period and a wavelength
in a time-stepping computation with sufficiently fine spatial discretisation. Models are
generally based on the Boussinesq equations which are of similar form to the depth-
averaged shallow-water equations with additional terms to account for frequency
dispersion. There has been a massive research effort in this area in the past decade
following the original suggestion of Peregrine1. Nonlinear wave propagation from
deep water may now be simulated rather accurately, e.g. Madsen et al.2. While

2
shoreline runup may be simulated, e.g. Stansby3, there are rather ad hoc empirical
devices to account for wave breaking processes and, less significantly, bed friction, as
reviewed in Liu and Losada4. Such approaches are highly efficient in 1-D but less so
in 2-D due to the nature of the equations to be solved, involving cross-coupling terms,
and computational domains are typically limited to 1-10km. This approach is however
most realistic for driving nearshore sediment transport processes, and hence
morphology. While progress is also being made with direct solutions to the Navier-
Stokes/continuity/turbulence equations, particularly through the volume-of-fluid
(VOF) approach (e.g. Liu and Losada4), such methods are restricted to several
wavelengths and are far from a practical approach.

Phase-averaged models are presently of most relevance to large-scale coastal


problems and have developed from ray-tracing methods for refraction and shoaling in
the 1960’s, as described in Dean and Dalrymple5 for example. This approach has been
generalised to allow the input of directional wave spectra by solving the wave energy
transport equation. Perhaps the most widely used code of this type is the WAM
(WAve Model) model, described in Hasslemann et al.6, which is intended for ocean
and coastal conditions. The following physical processes are accounted for:
• Wave propagation in time and space
• Wave generation by wind
• Shoaling and refraction due to bathymetry
• Shoaling and refraction due to current
• Energy losses due to whitecapping and bottom friction
• Quadruplet wave-wave interactions

The WAM code is limited in nearshore regions and the SWAN (Simulating WAves
Nearshore) code, described in Booij et al.7 and Ris et al.8, has been developed which
also accounts for:
• Depth-induced breaking
• Triad wave-wave interaction
SWAN solves for the wave action density spectrum rather than energy density,
because in the presence of currents wave action is conserved while energy density is
not. WAM uses spherical coordinates and is explicit in time and SWAN uses a

3
Cartesian mesh and is implicit. Both are widely used and indeed WAM has been used
over a large region to provide boundary conditions for SWAN to give more accurate
inshore wave conditions by Wornom et al.9, who found that depth-induced breaking
had a prominent influence while triad wave-wave interaction did not. Attempts have
been made to include wave diffraction effects through the use of the mild slope
equation to give the propagation velocities but this has not proved numerically robust.
It has however been formulated within a parabolic approximation giving good
numerical stability (Mase10), who tested for various situations with and without
diffraction. While this did not always show exact agreement with analytical results, of
most relevance here, wave spectral propagation over an elliptic shoal was tested
showing quite marked effects of diffraction for directional waves with narrow spread
but very little with broad spread. Indeed for engineering predictions diffraction is
often assumed to be a local effect which does not affect far field wave heights, e.g.
Dean and Dalrymple5.

In this study we use the code TOMAWAC (Telemac-based Operational Model


Addressing wave Action Computation) with an efficient characteristics-based,
mathematical formulation and finite-element spatial discretisation; this is
unstructured, allowing complex domains to be covered and grid refinement in regions
of strong bathymetric variation. This is thus a different formulation from SWAN.
Wave fields are compared with those measured experimentally over circular and
elliptical shoals and with the diffraction code ARTEMIS. Finally TOMAWAC is
applied to the East Anglian coastal region and results for inshore wave fields due to
broad-spread and narrow-spread, offshore, directional waves are compared.

2. MATHEMATICAL FORMULATION AND NUMERICAL METHODS

2.1. TOMAWAC (Telemac-based Operational Model Addressing Wave Action


Computation; Benoit11)

The following form of equation for the propagation of the energy spectrum is used by
TOMAWAC:
∂ ( BF ) • ∂ ( BF ) • ∂ ( BF ) • ∂ ( BF ) • ∂ ( BF )
+x +y +θ + fr = BQ (1)
∂t ∂x ∂y ∂θ ∂f r

4
CC g Cg
where F(x,y,θ,σ,t) is the directional wave energy spectrum, B = = ,
2πσ 2
(2π ) 2 kf r

C = σ / k is the phase velocity, Cg is the group velocity, k is the wave number, σ is

the relative angular frequency, f r is relative frequency and θ is the wave propagation
direction. From the linear wave theory (Komen et al.12), the propagation rates in the
above equation can be expressed as:

x = C g sin θ + U x (2)

y = C g cosθ + U y (3)
r
• 1 ∂σ ~ k ~ r
θ =− Gn (d ) − G n (U ) (4)
k ∂d k
• 1 ∂σ ∂d r r r~ r
fr = [ ( + U∇d ) − C g k Gt (U )] (5)
2π ∂d ∂t
r
where d is the water depth and U is the depth-averaged current velocity vector. The
~ ~
operators Gn and Gt refer to the computation of a function gradient in directions that
r r
are respectively normal ( n ) and tangential ( t ) to the characteristic curve with the
direction θ:
~ rr ∂g ∂g
Gn ( g ) = n ∇g = cosθ − sin θ (6)
∂x ∂y
~ rr ∂g ∂g
Gt ( g ) = t ∇g = sin θ + cosθ (7)
∂x ∂y

Besides, using the dispersion relation σ 2 = g k tanh(kd ) , the following relationship


can be derived:
∂σ σk
= (8)
∂d sinh(2kd )
On the left hand side of the Eq.(1), the first term represents the temporal (or local)
• •
evolution of the spectrum. The next two terms with x and y given by Eqs.(2) and (3)
represent spatial wave propagation with shoaling in finite water depth. The next term

with θ given by Eq.(4) represents refraction from bathymetric variation (first term on
the right hand side of Eq.(4) ) and current gradients (second term on right hand side of

Eq.(4) ). The final term on the left hand side of Eq.(1) with f r given by Eq.(5)

5
represents relative frequency changes resulting from sea level variation, in space and
time, and from spatial current variation.

On the right hand side of the Eq.(1), Q represents the contributions of the following
source terms: i) energy input by wind; ii) energy dissipation by whitecapping; iii)
energy dissipation by bottom friction; iv) energy dissipation due to depth-induced
breaking; and v) non-linear resonating quadruple and triad wave-wave interactions.
Various state-of-the art formulations for each of the terms are available in Benoit et
al.13.

A fractional step method was used to solve the equation: a convection step without
source terms and a time-integration step for the source terms. In the convection step, a
characteristic method (piecewise ray method) is used. If the convection field is
stationary, the characteristics can be traced back only once, at the beginning of the
computation. This step consumes very little computation time since it only consists in
an interpolation operation at each time step. This method is unconditionally stable and
efficient. The time-integration of source terms is semi-implicit following that used in
the WAM-Cycle 4 model (WAMDI Group14, Komen et al.12) which enables the use
of a long time step (about 20-30 min. in an oceanic environment). Further, the finite
element technique allows flexible discretization of the spatial domain.

2.2. ARTEMIS (Agitation and Refraction with TElemac on the MIld Slope
equation; Alebrecht15)

The code ARTEMIS solves the mild slope equation for velocity potential (Berkoff16)
on a finite element mesh. The following elliptic form includes additional dissipative
effects, resulting from bathymetry-induced breaking and bottom friction, in the
complex term (Booij17 , also De Firolamo18):

∇(CC g ∇φ ) + CC g ( k 2 + ikµ )φ = 0 (9)

W
where φ is the reduced velocity potential, µ = is the dissipation coefficient
(CC g )1 / 2

with W a dissipation function. Usual approximations associated with linear theory

6
apply, with irregular waves considered as the linear superposition of regular waves.
The method is suitable for modelling wave resonance and seiching in harbours and
wave fields due to combined refraction/diffraction/reflection in small bays. However
refraction by currents is not included.

To solve the elliptic equation, the real and imaginary parts are separated and the parts
of the velocity potential are solved by a pre-conditioned conjugate gradient method on
a finite-element mesh.

The nearshore processes included in TOMAWAC and ARTEMIS are very similar
apart from the above exceptions.

3. TEST CASES

Comparisons will be made with wave fields measured experimentally over circular
and elliptical shoals. Both TOMAWAC and ARTEMIS input JONSWAP wave
spectra, while the experiments below use TMA spectra. The JONSWAP spectrum can
be written as the following product:

S ( f ,θ ) = J ( f ) G (θ ) (10)

The frequency component J(f) is given by:

⎡ ⎛ f −1 ⎞ ⎤
2
⎢ ⎟ ⎥
exp ⎢ −0.5⎜⎜
fp
⎟ ⎥
⎡ ⎛ f ⎞ ⎤
4
⎢ ⎜ σ ⎟ ⎥
⎝ ⎠ ⎥⎦
J ( f ) = δ H s2 f p4 f −5 exp ⎢ −1.25 ⎜ p ⎟ ⎥ γ ⎣⎢
(11)
⎢⎣ ⎝ f ⎠ ⎥⎦

where Hs is the significant wave height, fp is the peak frequency and σ is a


dimensionless parameter which is determined as follows:
if f ≤ fp, σ = 0.07 (12)
if f > fp, σ = 0.09 (13)

γ is a real number ranging from 1 to 7, which determines the width of the frequency
spectrum and δ is a weight coefficient that depends on γ, such that:

7
0.0624
δ = (14)
0.185
0.230 + 0.0336γ −
1.9 + γ

The directional distribution G(θ) is often chosen in the following form

G = G0 cos 2 s [(θ − θ 0 ) / 2] (15)

where θ is the angle of wave propagation direction, θ0 is the main direction of wave
propagation. s is a positive real exponent such that small s indicates high spread and
vice versa. G0 is a constant determined by normalizing G such that:

θ max

∫ G dθ = 1
ϑmin
(16)

where θmin and θmax indicate the limits of wave propagation direction.

The TMA spectrum (Bouws et al.19) was used in the experiments on waves over
shoals by Vincent et al.20 and Chawla et al.21. The frequency component is similar to
that of the JONSWAP spectrum, but the directional component was obtained using a
wrapped normal directional spreading function (Borgman22), given by:
1 1 J ( jσ m ) 2
D(ϑ ) = + ∑ {exp[− ]cos j (θ − θ 0 )} (17)
2π π j =1 2

where J is the number of terms in the series (chosen as 50) and σm is a parameter
which determines the width of the directional spreading. We can see that the Eq.(17)
is of discrete form. When J is large enough, it tends to a continuous spectrum. A trial-
and-error method was used in transform between σm in Eq.(17) and s in Eq.(15). σm is
computed for different s and the s value corresponding to an experimental value of σm
is obtained by interpolation. Large σm corresponds to small s and vice versa. In this
way the directional spreading functions are made approximately similar.

3.1 Random directional waves over submerged shoals

8
The laboratory experiments of Chawla et al.21 were computed using TOMAWAC and
ARTEMIS. The computational domain is the same as that used in the experiments,
covering an area of 18m × 18m. The centre of the circular shoal was located at (5.0m,
8.98m) with a bottom radius of 2.57 m. The bathymetry at any point in the shoal, in
metres, is given by:

Z = −(h + 8.73) + 82.81 − ( x − 5) 2 − ( y − 8.98) 2 (19)

where h is the water depth away from the shoal. All tests were run in 40 cm water
depth, giving a depth over the centre of the shoal of only 3 cm. This bathymetry and a
mesh of 17344 triangular elements with 8845 nodes are shown in Fig.1 (See
Appendix). The mesh size varied from 0.12 to 0.34 m, with an average value 0.2 m.
Based on the linear dispersion relationship, wavelength is about 1.4m, giving about 7
elements per wavelength.

Four test cases (3-6) in Chawla et al.’s experiments were computed and test
parameters shown in Table 1. A JONSWAP spectrum was input at x = 0 m in the
computations, with parameters as shown in Table 1. A spectral peakedness value of
γ = 10 defines the width of the frequency spectrum. In TOMAWAC, frequency and
direction were discretised as 13 and 18 segments respectively, with a minimum and
maximum frequencies of 1 Hz and 3 Hz respectively. An initial wave height of 0.005
m with the same spectrum was input. The time step was 0.5 s and 240 steps were
computed were computed to reach a steady state. The formulation of Battjes &
Janssen23 was used to compute wave breaking. The formulation for bottom friction
follows Hasselmann et al.6 and Bouws and Komen 24 with a coefficient taken as 0.038
m2/s3 in accordance with the standard value being used in the model WAM-Cycle 4.
As in the experiments, fully absorption condition were used on the boundary at x = 18
m. The two lateral boundaries also used the absorption condition. In ARTEMIS, the
frequency was discretised as 13 segments, with minimum and maximum values of 1
Hz and 1.67 Hz respectively. This reduction appeared necessary for numerical
stability. Other parameters and boundary conditions are the same as for TOMAWAC.
Fig.2 shows an example of a comparison of computed wave fields from TOMAWAC
and ARTEMIS for test 6, with broad directional spreading. The wave fields show

9
some clear differences. The difference in focussing around the centre of the shoal was
more pronounced with narrow spreading. However direct comparison with wave
height measurement is more revealing and Fig.3 shows the measurement locations.
Figs. 4-10 show comparisons of computed and measured wave heights, normalised by
the onset value, for tests 3-6 along transects A-A, B-B, C-C, D-D, E-E, F-F and G-G
respectively. All results show that for broad directional spreading, tests 4 and 6,
TOMAWAC gives good predictions which are generally better than those of
ARTEMIS. For narrow spreading, tests 3 and 5, ARTEMIS generally (but not always)
gives better predictions, in particular showing the increased wave heights over the
shoal due to diffraction.

Table 1 Parameters for Incident Waves in Chawla et al.’s experiments


Case Period Frequency Wave Height γ σm s
(s) (Hz) (m) (deg)
3 0.73 1.37 0.0139 10 5 50
4 0.73 1.37 0.0156 10 20 6
5 0.73 1.37 0.0233 10 5 50
6 0.71 1.41 0.0249 10 20 6

A similar study was made for the experiments of Vincent et al.20 for an elliptic shoal,
in this case with a relatively large minimum depth over the shoal of 15.24cm
compared with a deep water value of 45.72cm. There was a relatively broad
frequency spectrum with γ = 2 . However the results and agreement with experiment
were quite similar and are not reproduced here.

3.2 Wave fields on the East Anglian coast


Wave field predictions using TOMAWAC for a significant height of 8m and a period
of 10s on the Eastern boundary, propagating from east to west, have been presented in
Part 1. Since TOMAWAC gives good predictions with broad spread waves even
when diffraction occurs, examples of results with broad and narrow spread are now
compared for this region. The bathymetry and mesh are shown in Figs. 2 and 3 in Part
1.

10
JONSWAP spectra with γ = 3.3 and a directional spread parameter, s, of 2 and 20
(broad and narrow) were used. Wave frequency and direction were divided into 17
and 36 segments respectively and the frequency was varied from 0.04 Hz to 0.25 Hz.
Use of such limits is standard practice to avoid numerical problems caused by very
low and high frequencies. Free boundary conditions were used on the north and south.
With a time step of 20s, 360 time steps were sufficient to reach a steady state in a
computational time of 2 hours. In the computational domain, the water level was
typical of a spring tide with an average level of 4.0 m above LAT. To start the
computation an initial significant wave height was set to 0.5 m with the same
spectrum as at the incident boundary. Bottom friction dissipation and depth-induced
breaking dissipation were included as in Part 1. Fig. 11 shows the computed wave
fields for broad and narrow directional spreading (s = 2 and 20 respectively). It can be
seen they are almost the same except for the regions near the north and south
boundaries. Figs. 12 and 13 show comparisons for s=2 and 20 of computed wave
height along a line marked by crosses (at depths of about 5 m) and a line marked by
circles (at depths of about 10 m), as shown in Fig. 2 of Part 1. It can be seen there is
negligible difference where the depth is about 5 m, but some difference near the south
and north boundaries where the depths is about 10 m.

3.3 Ocean measurements off France (Benoit et al.25)


The developers of TOMAWAC, Benoit et al.25, used TOMAWAC to simulate the
storm of January 25, 1990 along the French Atlantic Coasts and in the English
Channel, which had been previously modelled with the WAM model (Benoit and
Teisson26). The computed time series of wave heights were compared with buoy
measurements at Ouessant (at the entrance to the Channel with a water depth of 108
m) and Flamanville (inside the Channel with a water depth of 18 m) and the results of
WAM. The two model results compare very well with each other and the agreement
with the buoy data is also satisfactory (Fig.3 in the paper of Benoit et al.25). But there
are great differences between both models in the nearshore zone where wave heights
from the WAM model increase in an unrealistic way for depths less than about 12m.
The inclusion of depth-induced breaking in TOMAWAC keeps this wave height at a
more physical level (Fig. 4 in the paper of Benoit et al.25). Although no measurements
were available at this location during the storm, the ratio of significant wave height to

11
water depth of about 0.4 is consistent with what is expected due to depth-limited
breaking in the surf zone, e.g. Tajima and Madsen27.

4. CONCLUSION
The finite-element code TOMAWAC has been used to compute the propagation of
directional wave spectra inshore. It is efficient and well suited to large-scale
modelling at the cost of omitting diffraction. However comparisons with experiments
on wave fields over circular and elliptic shoals indicate that this is insignificant for
broad-spread waves, but not for narrow-spread waves. For the latter the finite-element
code ARTEMIS, solving the mild-slope equation, generally gives improved results
although there are restrictions associated with numerical stability. Running
TOMAWAC for the East Anglian coastal region with narrow and broad spread waves
showed only small differences at about 10m depth and negligible differences at 5m
depth. This suggests that the use of a code such as TOMAWAC with broad spread
spectra is a useful tool for nearshore wave prediction, even when diffraction over
shoals occurs.

5. ACKNOWLEDGEMENTS
This work (Tyndall project IT 1.37) is part of a larger project ‘towards an integrated
regional coastal simulator for the impact of sea-level rise in East Anglia’ and has been
funded through the Tyndall Centre for Climate Change Research.

REFERENCES

1. Peregrine,D.H. 1967 Long waves on a beach, J. Fluid Mech., 27, 815-827.

2. Madsen,P.A., Bingham,H.B. and Liu, H. 2002 A new Boussinesq method for fully
nonlinear waves from deep to shallow water, J. Fluid Mech., 462, 1-30.

3. Stansby,P.K. 2003 Solitary wave run up and overtopping by a semi-implicit finite-


volume shallow-water Boussinesq model, to appear J. Hydr. Res.

12
4. Liu, P.L.-F. and Losada, I.J. 2002 Wave propagation modelling in coastal
engineering, J. Hydr. Res., 40(3), 229-240.

5. Dean, R.G. and Dalrymple, R.A. 2000 Water wave mechanics for engineers and
scientists, Advanced Series On Ocean Engineering – Vol.2, World Scientific.

6. Hasslemann, S., Hasslemann, K., Buer, E., Janssen, P.A.E.M., Komen,G.J.,


Bertotti,L., Lionello, P., Guillaume, A., Cardone, V.C., Greenwood, J.A., Reistad,M.,
Zambresky, L. and Ewing, J.A. 1988 The WAM model – A third generation ocean
wave prediction model, J. Phys. Oceanog., 18, 1775-1810.

7. Booij, N., Ris, R.C. and Holthuijsen, L.H. 1999 A third generation wave model for
coastal regions, Part I: Model description and validation, J. Geophys. Res., 104, C4,
7649-7666.

8. Ris, R.C., Booij, N. and Holthuijsen, L.H. 1999. A third-generation wave model for
coastal regions, Part II: Verification, J.Geophys. Res. 104, C4, 7667-7681.

9. Wornom, S.F., Welsh, D.J.S. and Bedford, K.W. 2001 On coupling the SWAN and
WAM wave models for accurate nearshore wave predictions, Coastal Engrg. J., 43(3),
161-201.

10. Mase,H. 2001 Multi-directional random wave transformation model based on


energy balance equation, Coastal Engrg. J., 43(4), 317-337.

11. Benoit, M. 2002. TOMAWAC software for finite element sea state modelling –
Release 5.2 theoretical note. HP-75/02/065/A.

12. Komen, G.J., Cavaleri, L.,Donelan, M., Hasselmann, K., Hasselmann S. and
Janssen P.A.E.M. 1994. Dynamics and Modelling of Ocean Waves. Cambridge
University Press, 532.

13
13. Benoit M., Marcos, F. and Becq F. 1996. Development of a third generation
shallow-water wave model with unstructured spatial meshing. Proc. 25th Int. Conf.
Coastal Eng. (ASCE), Orlando, Florida, USA.

14. WAMDI Group, 1988 The WAM-model – A third generation ocean wave
prediction model, J. Phys. Oceanogr., 18, 1775-1810.

15. Aelbrecht, D. 1997. ARTEMIS software – Version 3.0 theoretical note. HE-
42/97/002/B.

16. Berkhoff, J.C.W. 1976. Mathematical models for simple harmonic linear water
waves. Wave diffraction and refraction. Delft Hydraulics Lab. Publication No. 163.

17. Booij, N. 1981. Gravity waves on water with non-uniform depth and current. Ph.D
thesis, Technical University of Delft, The Netherlands.

18. De Girolamo, P., Kostense, J.K. and Dingemans, M.W. 1988. Inclusion of surf-
breaking in a mild-slope model. Comput. Modell. Ocean Engrg., 221-229.

19. Bouws, E., Gunther, H., Rosenthal, W., and Vincent, C. 1985, Simularity of the
wind wave spectrum in finite depth water. J. Geophys. Res. 90, 975-986.

20. Vincent C.L., Briggs M.J. 1989 Refraction-diffraction of irregular waves over a
mound. J. Waterway, Port, and Ocean Engrg., 115, 269-284.

21. Chawla, A., Ozkan-Haller, H.T. and Kirby, J.T. 1998 Spectral model for wave
transformation and breaking over irregular bathymetry, J. Waterway, Port, Coastal
and Ocean Engrg., 124(4), 189-198.

22. Borgman, L.E. (1984) Directional spectrum estimation for the Sxy gages. Tech.
Rep., Coastal Engineering Research Center, Vicksburg, Mississippi.

23. Battjes, J.A. and Janssen, J.P.F.M. 1978. Energy loss and set-up due to breaking
of random waves. Proc. 16th Int. Conf. Coastal Eng., Vol.1, 649-660. Hamburg.

14
24. Bouwse, E. and Komen, G.J. 1983. On the balance between growth and
dissipation in an extreme depth-limited wind-sea in the southern North Sea. J. Phys.
Oceanog., Vol. 13, 1653-1658.

25. Benoit M., Marcos, F. and Becq F. 1997. TOMAWAC: a prediction model for
offshore and nearshore storm waves. Proc. 27th IAHR Congress, San Francisco,
California, USA. or EDF HE-42/97/036/A.

26. Benoit, M. and Teisson, C. 1995. Spectral wave modeling along the French
Atlantic Coasts and in the Channel. Proc. 26th IAHR Congress, Vol.3. 156-161.

27. Tajima,Y. and Madsen, O.S. 2002 Shoaling, breaking and broken wave
characteristics. Proc. 28th Int. Conf. Coastal Engineering, ASCE, Cardiff.

15
APPENDIX

a) b)

Fig.1. Bathymetry (a) and mesh (b) for the circular shoal case of Chawla et al.(1998)

a)

16
b)

Fig.2 Computed normalised wave height field using TOMAWAC (a) and ARTEMIS
(b) for Test 6 (larger wave height with broad directional spread)

17
Fig.3 Scheme of measurement locations for the circular shoal of Chawla et al.(1998)

18
Fig.4 Comparison of computed and measured normalized wave height profiles along
transect A-A (y=8.98 m) for Tests 3-6 in Table 1.

19
Fig.5 Comparison of computed and measured normalized wave height profiles along
transect B-B (x=13.7 m) for Tests 3-6 in Table 1.

20
Fig.6 Comparison of computed and measured normalized wave height profiles along
transect C-C (x=11.0 m) for Tests 3-6 in Table 1.

21
Fig.7 Comparison of computed and measured normalized wave height profiles along
transect D-D (x=8.2 m) for Tests 3-6 in Table 1.

22
Fig.8 Comparison of computed and measured normalized wave height profiles along
transect E-E (x=6.3 m) for Tests 3-6 in Table 1.

23
Fig.9 Comparison of computed and measured normalized wave height profiles along
transect F-F (x=5.0 m) for Tests 3-6 in Table 1.

24
Fig.10 Comparison of computed and measured normalized wave height profiles along
transect G-G (x=3.7 m) for Tests 3-6 in Table 1.

25
a) b)

Fig.11 Computed wave height fields for the East Anglian coastal region using
different directional wave spreading with wave propagation from the East: a) broad
spread, s=2; b) narrow spread, s=20.

26
Fig.12 Comparison of computed wave heights along a line with marked crosses in
Fig.2 (Part 1) for waves from the East with narrow spread (s=20) and broad spread
(s=2).

Fig.13 Comparison of computed wave heights along a line with marked circles in
Fig.2 (Part 1) for waves from the East with narrow spread (s=20) and broad spread
(s=2).

27
The trans-disciplinary Tyndall Centre for Climate Change Research undertakes integrated research into the
long-term consequences of climate change for society and into the development of sustainable responses
that governments, business-leaders and decision-makers can evaluate and implement. Achieving these
objectives brings together UK climate scientists, social scientists, engineers and economists in a unique
collaborative research effort.
Research at the Tyndall Centre is organised into four research themes that collectively contribute to all
aspects of the climate change issue: Integrating Frameworks; Decarbonising Modern Societies; Adapting to
Climate Change; and Sustaining the Coastal Zone. All thematic fields address a clear problem posed to
society by climate change, and will generate results to guide the strategic development of climate change
mitigation and adaptation policies at local, national and global scales.
The Tyndall Centre is named after the 19th century UK scientist John Tyndall, who was the first to prove the
Earth’s natural greenhouse effect and suggested that slight changes in atmospheric composition could bring
about climate variations. In addition, he was committed to improving the quality of science education and
knowledge.
The Tyndall Centre is a partnership of the following institutions:
University of East Anglia
UMIST
Southampton Oceanography Centre
University of Southampton
University of Cambridge
Centre for Ecology and Hydrology
SPRU – Science and Technology Policy Research (University of Sussex)
Institute for Transport Studies (University of Leeds)
Complex Systems Management Centre (Cranfield University)
Energy Research Unit (CLRC Rutherford Appleton Laboratory)
The Centre is core funded by the following organisations:
Natural Environmental Research Council (NERC)
Economic and Social Research Council (ESRC)
Engineering and Physical Sciences Research Council (EPSRC)
UK Government Department of Trade and Industry (DTI)

For more information, visit the Tyndall Centre Web site (www.tyndall.ac.uk) or contact:
Communications Manager
Tyndall Centre for Climate Change Research
University of East Anglia, Norwich NR4 7TJ, UK
Phone: +44 (0) 1603 59 3900; Fax: +44 (0) 1603 59 3901
Email: tyndall@uea.ac.uk
Tyndall Working Papers are available online at
http://www.tyndall.ac.uk/publications/working_papers/working_papers.shtml

Mitchell, T. and Hulme, M. (2000). A Country-by- Köhler, J.H., (2002). Long run technical change
Country Analysis of Past and Future Warming in an energy-environment-economy (E3)
Rates, Tyndall Centre Working Paper 1. model for an IA system: A model of
Kondratiev waves, Tyndall Centre Working Paper
Hulme, M. (2001). Integrated Assessment
15.
Models, Tyndall Centre Working Paper 2.
Adger, W.N., Huq, S., Brown, K., Conway, D. and
Berkhout, F, Hertin, J. and Jordan, A. J. (2001).
Hulme, M. (2002). Adaptation to climate
Socio-economic futures in climate change
change: Setting the Agenda for Development
impact assessment: using scenarios as
Policy and Research, Tyndall Centre Working
'learning machines', Tyndall Centre Working
Paper 16.
Paper 3.
Dutton, G., (2002). Hydrogen Energy
Barker, T. and Ekins, P. (2001). How High are
Technology, Tyndall Centre Working Paper 17.
the Costs of Kyoto for the US Economy?,
Tyndall Centre Working Paper 4. Watson, J. (2002). The development of large
technical systems: implications for hydrogen,
Barnett, J. (2001). The issue of 'Adverse Effects
Tyndall Centre Working Paper 18.
and the Impacts of Response Measures' in the
UNFCCC, Tyndall Centre Working Paper 5. Pridmore, A. and Bristow, A., (2002). The role of
hydrogen in powering road transport, Tyndall
Goodess, C.M., Hulme, M. and Osborn, T. (2001).
Centre Working Paper 19.
The identification and evaluation of suitable
scenario development methods for the Turnpenny, J. (2002). Reviewing organisational
estimation of future probabilities of extreme use of scenarios: Case study - evaluating UK
weather events, Tyndall Centre Working Paper 6. energy policy options, Tyndall Centre Working
Paper 20.
Barnett, J. (2001). Security and Climate
Change, Tyndall Centre Working Paper 7. Watson, W. J. (2002). Renewables and CHP
Deployment in the UK to 2020, Tyndall Centre
Adger, W. N. (2001). Social Capital and Climate
Working Paper 21.
Change, Tyndall Centre Working Paper 8.
Watson, W.J., Hertin, J., Randall, T., Gough, C.
Barnett, J. and Adger, W. N. (2001). Climate
(2002). Renewable Energy and Combined Heat
Dangers and Atoll Countries, Tyndall Centre
and Power Resources in the UK, Tyndall Centre
Working Paper 9.
Working Paper 22.
Gough, C., Taylor, I. and Shackley, S. (2001).
Paavola, J. and Adger, W.N. (2002). Justice and
Burying Carbon under the Sea: An Initial
adaptation to climate change, Tyndall Centre
Exploration of Public Opinions, Tyndall Centre
Working Paper 23.
Working Paper 10.
Xueguang Wu, Jenkins, N. and Strbac, G. (2002).
Barker, T. (2001). Representing the Integrated
Impact of Integrating Renewables and CHP
Assessment of Climate Change, Adaptation
into the UK Transmission Network, Tyndall
and Mitigation, Tyndall Centre Working Paper 11.
Centre Working Paper 24
Dessai, S., (2001). The climate regime from
Xueguang Wu, Mutale, J., Jenkins, N. and Strbac,
The Hague to Marrakech: Saving or sinking
G. (2003). An investigation of Network
the Kyoto Protocol?, Tyndall Centre Working
Splitting for Fault Level Reduction, Tyndall
Paper 12.
Centre Working Paper 25
Dewick, P., Green K., Miozzo, M., (2002).
Brooks, N. and Adger W.N. (2003). Country level
Technological Change, Industry Structure and
risk measures of climate-related natural
the Environment, Tyndall Centre Working Paper
disasters and implications for adaptation to
13.
climate change, Tyndall Centre Working Paper 26
Shackley, S. and Gough, C., (2002). The Use of
Tompkins, E.L. and Adger, W.N. (2003). Building
Integrated Assessment: An Institutional
resilience to climate change through adaptive
Analysis Perspective, Tyndall Centre Working
management of natural resources, Tyndall
Paper 14.
Centre Working Paper 27
Dessai, S., Adger, W.N., Hulme, M., Köhler, J.H., Klein, R.J.T., Lisa Schipper, E. and Dessai, S.
Turnpenny, J. and Warren, R. (2003). Defining (2003), Integrating mitigation and adaptation
and experiencing dangerous climate change, into climate and development policy: three
Tyndall Centre Working Paper 28 research questions, Tyndall Centre Working Paper
40
Brown, K. and Corbera, E. (2003). A Multi-
Criteria Assessment Framework for Carbon-
Watson, J. (2003), UK Electricity Scenarios for
Mitigation Projects: Putting “development” in
2050, Tyndall Centre Working Paper 41
the centre of decision-making, Tyndall Centre
Working Paper 29
Kim, J. A. (2003), Sustainable Development and
Hulme, M. (2003). Abrupt climate change: can the CDM: A South African Case Study, Tyndall
society cope?, Tyndall Centre Working Paper 30 Centre Working Paper 42

Turnpenny, J., Haxeltine A. and O’Riordan, T. Anderson, D. and Winne, S. (2003),


(2003). A scoping study of UK user needs for Innovation and Threshold Effects in
managing climate futures. Part 1 of the pilot- Technology Responses to Climate Change,
phase interactive integrated assessment Tyndall Centre Working Paper 43
process (Aurion Project), Tyndall Centre
Working Paper 31 Shackley, S., McLachlan, C. and Gough, C. (2004)
Xueguang Wu, Jenkins, N. and Strbac, G. (2003). The Public Perceptions of Carbon Capture and
Integrating Renewables and CHP into the UK Storage, Tyndall Centre Working Paper 44
Electricity System: Investigation of the impact
of network faults on the stability of large Purdy, R. and Macrory, R. (2004) Geological
offshore wind farms, Tyndall Centre Working carbon sequestration: critical legal issues,
Paper 32 Tyndall Centre Working Paper 45

Pridmore, A., Bristow, A.L., May, A. D. and Tight, Watson, J., Tetteh, A., Dutton, G., Bristow, A.,
M.R. (2003). Climate Change, Impacts, Future Kelly, C., Page, M. and Pridmore, A., (2004) UK
Scenarios and the Role of Transport, Tyndall Hydrogen Futures to 2050, Tyndall Centre
Centre Working Paper 33 Working Paper 46

Dessai, S., Hulme, M (2003). Does climate policy Berkhout, F., Hertin, J. and Gann, D. M., (2004)
need probabilities?, Tyndall Centre Working Paper Learning to adapt: Organisational adaptation
34 to climate change impacts, Tyndall Centre
Working Paper 47
Tompkins, E. L. and Hurlston, L. (2003). Report to
the Cayman Islands’ Government. Adaptation Pan, H. (2004) The evolution of economic
lessons learned from responding to tropical structure under technological development,
cyclones by the Cayman Islands’ Government, Tyndall Centre Working Paper 48
1988 – 2002, Tyndall Centre Working Paper 35
Awerbuch, S. (2004) Restructuring our
Kröger, K. Fergusson, M. and Skinner, I. (2003). electricity networks to promote
Critical Issues in Decarbonising Transport: The decarbonisation, Tyndall Centre Working Paper 49
Role of Technologies, Tyndall Centre Working
Paper 36 Powell, J.C., Peters, M.D., Ruddell, A. & Halliday, J.
(2004) Fuel Cells for a Sustainable Future?
Ingham, A. and Ulph, A. (2003) Uncertainty, Tyndall Centre Working Paper 50
Irreversibility, Precaution and the Social Cost
of Carbon, Tyndall Centre Working Paper 37 Agnolucci, P., Barker, T. & Ekins, P. (2004)
Hysteresis and energy demand: the
Brooks, N. (2003). Vulnerability, risk and Announcement Effects and the effects of the
adaptation: a conceptual framework, Tyndall UK climate change levy, Tyndall Centre Working
Centre Working Paper 38 Paper 51

Tompkins, E.L. and Adger, W.N. (2003). Agnolucci, P. (2004) Ex post evaluations of CO2
Defining response capacity to enhance climate –Based Taxes: A Survey, Tyndall Centre Working
change policy, Tyndall Centre Working Paper 39 Paper 52
Agnolucci, P. & Ekins, P. (2004) The Adger, W. N., Brown, K. and Tompkins, E. L.
Announcement Effect and environmental (2004) The political economy of cross-scale
taxation, Tyndall Centre Working Paper 53 networks in resource co-management, Tyndall
Centre Working Paper 65
Turnpenny, J., Carney, S., Haxeltine, A., &
O’Riordan, T. (2004) Developing regional and Turnpenny, J., Haxeltine, A., Lorenzoni, I.,
local scenarios for climate change mitigation O’Riordan, T., and Jones, M., (2005) Mapping
and adaptation, Part 1: A framing of the East actors involved in climate change policy
of England, Tyndall Centre Working Paper 54 networks in the UK, Tyndall Centre Working
Paper 66
Mitchell, T.D. Carter, T.R., Jones, .P.D, Hulme, M.
and New, M. (2004) A comprehensive set of Turnpenny, J., Haxeltine, A. and O’Riordan, T.,
high-resolution grids of monthly climate for (2005) Developing regional and local scenarios
Europe and the globe: the observed record for climate change mitigation and adaptation:
(1901-2000) and 16 scenarios (2001-2100), Part 2: Scenario creation, Tyndall Centre
Tyndall Centre Working Paper 55 Working Paper 67

Vincent, K. (2004) Creating an index of social Bleda, M. and Shackley, S. (2005) The formation
vulnerability to climate change for Africa, of belief in climate change in business
Tyndall Centre Working Paper 56 organisations: a dynamic simulation model,
Tyndall Centre Working Paper 68
Shackley, S., Reiche, A. and Mander, S (2004) The
Public Perceptions of Underground Coal Tompkins, E. L. and Hurlston, L. A. (2005) Natural
Gasification (UCG): A Pilot Study, Tyndall Centre hazards and climate change: what knowledge
Working Paper 57 is transferable?, Tyndall Centre Working Paper 69

Bray, D and Shackley, S. (2004) The Social Abu-Sharkh, S., Li, R., Markvart, T., Ross, N.,
Simulation of The Public Perceptions of Wilson, P., Yao, R., Steemers, K., Kohler, J. and
Weather Events and their Effect upon the Arnold, R. (2005) Can Migrogrids Make a Major
Development of Belief in Anthropogenic Contribution to UK Energy Supply?, Tyndall
Climate Change, Tyndall Centre Working Paper 58 Centre Working Paper 70

Anderson, D and Winne, S. (2004) Modelling Boyd, E. Gutierrez, M. and Chang, M. (2005)
Innovation and Threshold Effects Adapting small-scale CDM sinks projects to
In Climate Change Mitigation, Tyndall Centre low-income communities, Tyndall Centre
Working Paper 59 Working Paper 71
Few, R., Brown, K. and Tompkins, E.L. (2004) Lowe, T., Brown, K., Suraje Dessai, S., Doria, M.,
Scaling adaptation: climate change response Haynes, K. and Vincent., K (2005) Does tomorrow
and coastal management in the UK, Tyndall ever come? Disaster narrative and public
Centre Working Paper 60 perceptions of climate change, Tyndall Centre
Working Paper 72
Brooks, N. (2004) Drought in the African Sahel:
Long term perspectives and future prospects, Walkden, M. (2005) Coastal process simulator
Tyndall Centre Working Paper 61 scoping study, Tyndall Centre Working Paper 73
Barker, T. (2004) The transition to Ingham, I., Ma, J., and Ulph, A. M. (2005) How do
sustainability: a comparison of economics the costs of adaptation affect optimal
approaches, Tyndall Centre Working Paper 62 mitigation when there is uncertainty,
irreversibility and learning?, Tyndall Centre
Few, R., Ahern, M., Matthies, F. and Kovats, S. Working Paper 74
(2004) Floods, health and climate change: a
strategic review, Tyndall Centre Working Paper 63 Fu, G., Hall, J. W. and Lawry, J. (2005) Beyond
probability: new methods for representing
Peters, M.D. and Powell, J.C. (2004) Fuel Cells for uncertainty in projections of future climate,
a Sustainable Future II, Tyndall Centre Working Tyndall Centre Working Paper 75
Paper 64
Agnolucci,. P (2005) The role of political morphological modelling Tyndall Centre Working
uncertainty in the Danish renewable energy Paper 87
market, Tyndall Centre Working Paper 76
Peter Stansby and Cui-Ping Kuang, (2006)
Barker, T., Pan, H., Köhler, J., Warren., R and Sandbanks for coastal protection: implications
Winne, S. (2005) Avoiding dangerous climate of sea-level rise – part 3: wave modelling
change by inducing technological progress: Tyndall Centre Working Paper 88
scenarios using a large-scale econometric
model, Tyndall Centre Working Paper 77

Agnolucci,. P (2005) Opportunism and


competition in the non-fossil fuel obligation
market, Tyndall Centre Working Paper 78

Ingham, I., Ma, J., and Ulph, A. M. (2005) Can


adaptation and mitigation be
complements?, Tyndall Centre Working Paper 79

Wittneben, B., Haxeltine, A., Kjellen, B., Köhler, J.,


Turnpenny, J., and Warren, R., (2005) A
framework for assessing the political economy
of post-2012 global climate regime, Tyndall
Centre Working Paper 80

Sorrell, S., (2005) The economics of energy


service contracts, Tyndall Centre Working Paper
81

Bows, A., and Anderson, K. (2005) An analysis of


a post-Kyoto climate policy model, Tyndall
Centre Working Paper 82

Williamson, M. Lenton, T. Shepherd, J. and


Edwards, N. (2006) An efficient numerical
terrestrial scheme (ENTS) for fast earth
system modelling Tyndall Centre Working Paper
83

Kevin Anderson, Alice Bows and Paul Upham (2006)


Growth scenarios for EU & UK aviation:
contradictions with climate policy, Tyndall
Centre Working Paper 84

Michelle Bentham, (2006) An assessment of


carbon sequestration potential in the UK –
Southern North Sea case study Tyndall Centre
Working Paper 85

Peter Stansby, Cui-Ping Kuang, Dominique


Laurence and Brian Launder, (2006) Sandbanks
for coastal protection: implications of sea-
level rise - Part 1: Application to East Anglia
Tyndall Centre Working Paper 86

Peter Stansby and Cui-Ping Kuang, (2006)


Sandbanks for coastal protection: implications
of sea-level rise – Part 2: current and

Вам также может понравиться