Anda di halaman 1dari 11

Progress in Oceanography 120 (2014) 399409

Contents lists available at ScienceDirect

Progress in Oceanography
journal homepage: www.elsevier.com/locate/pocean

Review

The boundary layer at the bottom of a solitary wave and implications for
sediment transport
Giovanna Vittori, Paolo Blondeaux
Department of Civil, Chemical and Environmental Engineering, University of Genoa, Via Montallegro 1, 16145 Genova, Italy

a r t i c l e

i n f o

Article history:
Received 8 November 2012
Received in revised form 28 August 2013
Accepted 3 September 2013
Available online 19 September 2013

a b s t r a c t
The present paper summarizes the theoretical and numerical results of recent studies of the bottom
boundary layer generated by the propagation of a solitary wave which is often used as a model of a tsunami wave. The ow and the bottom shear stress are discussed as function of the parameters of the problem, i.e. (i) the ratio between the height H of the wave and the local water depth h, (ii) the ratio between
the thickness d of the bottom boundary layer and h, (iii) the relative bottom roughness. In particular, the
conditions leading to turbulence appearance, which are obtained by means of a linear stability analysis,
are presented along with those obtained by means of direct numerical simulations of NavierStokes
equations and the integration of the RANS equations. It is shown that turbulence tends to appear during
the decelerating phase of the wave cycle, if the wave height is larger than a critical value which depends
on the ratio between the thickness of the bottom boundary layer and h and the relative bottom roughness. As the height of the wave increases, turbulence appears earlier and becomes more intense, thus
enhancing mixing phenomena and the sediment transport rate.
2013 Elsevier Ltd. All rights reserved.

Contents
1.
2.

3.

4.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Formulation of the problem and solution methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.
Formulation of the problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.
The laminar solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.
The linear stability analysis of the laminar flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4.
The direct numerical simulations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5.
The Reynolds averaged model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
The results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.
The linear stability analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.
The direct numerical simulations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3.
The Reynolds averaged model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.4.
The sediment transport. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1. Introduction
Even though eld measurements show that it is difcult to observe a truly solitary wave (Shore Protection Manual, 1984), the
solitary wave model provides an acceptable description of the
Corresponding author. Tel.: +39 0103532475; fax: +39 0103532546.
E-mail addresses: vittori@dicat.unige.it (G. Vittori), blx@dicat.unige.it (P. Blondeaux).
0079-6611/$ - see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.pocean.2013.09.001

399
401
401
402
402
402
403
403
403
404
406
407
408
409
409

dynamics of sea waves when they approach the coastline and


move into the shallow water region (Munk, 1949). Moreover, landslides and earthquakes can generate long waves which behave
similarly to solitary waves. Hence, a large number of studies have
been devoted to study solitary waves.
In nature, the Reynolds number of the ow eld generated by
the propagation of a solitary wave turns out to be quite large and
it is acceptable to split the uid domain into a core region, where
the uid behaves like an inviscid uid, and boundary layers where

400

G. Vittori, P. Blondeaux / Progress in Oceanography 120 (2014) 399409

viscous effects should be taken into due account. The ow in the


inviscid region can be determined using different approaches and
the paper of Miles (1980) provides a review of the available results.
Even though the thickness of the bottom boundary layer is relatively thin, the description of the ow close to the bottom is quite
important for benthic processes.
In particular, the propagation of a solitary wave can mobilize a
large amount of sediment, thus causing signicant erosion and
deposition processes. To quantify the sediment transport induced
by the propagation of a solitary wave, it is necessary to have a detailed knowledge of the ow induced close to the bottom and to
determine the bottom shear stress. Therefore, in the recent years,
many investigations have been developed to the study of the
boundary layer at the bottom of a solitary wave considering both
the laminar regime and the turbulent regime. Of course, the turbulent regime is the most relevant, to analyze both sea waves in the
near-shore region and tsunami waves.
The boundary layer generated at the bottom of a solitary wave was
considered by Liu and Orla (2004) and Liu et al. (2007) by assuming
the ow regime to be laminar. The solution obtained by Liu and Orla
(2004) and Liu et al. (2007) shows that the horizontal velocity component in the inviscid region always points in the direction of wave
propagation but the uid velocity inside the bottom boundary layer
and the bed shear stress reverse their direction after the passage of
the wave crest, when the external velocity decelerates.
As already pointed out, the theoretical analyzes of Liu and Orla
(2004) and Liu et al. (2007), as well as those of previous investigators (Keulegan, 1948; Mei, 1989), assume that the Reynolds number
is not so large to trigger turbulence appearance. However, under
eld conditions, it is likely that the ow regime becomes turbulent.
To determine the ow eld in the boundary layer at the bottom of a
solitary wave, Vittori and Blondeaux (2008) used direct numerical
simulations of continuity and NavierStokes equations, which allow
turbulence appearance and dynamics to be investigated.
Vittori and Blondeaux (2008) observed that turbulent oscillations start to appear after the passage of the wave crest, if the wave
height is larger than a critical value which depends on the water
depth. After the passage of the wave crest, the ow outside the bottom boundary layer decelerates and the adverse pressure induces
an inectional point in the velocity prole which, for large wave
heights, triggers the instability of the basic laminar ow.
These numerical ndings have been recently conrmed by the
experimental investigations of Sumer et al. (2010) and Tanaka
et al. (2011). Sumer et al. (2010) simulated the ow in the bottom
boundary layer of a solitary wave using an oscillating water tunnel
(U-tube) while Tanaka et al. (2011) employed a new experimental
apparatus which consists of an overow head tank, a conduit water
tunnel and a downstream gate mechanically controlled.
The experimental measurements show that the ow regime
keeps laminar during the whole wave cycle, if the amplitude of
the velocity oscillation induced in the experimental apparatus is
relatively small. If the amplitude is increased, a critical value is
found above which the laminar ow turns out to be unstable. Just
above the critical conditions, the laboratory experiments of Sumer
et al. (2010) show that the instability of the basic laminar ow,
which takes place during the decelerating phase, leads to the
appearance of a regular array of two-dimensional vortices (vortex
tubes) which, later, decay because of viscous effects. If the amplitude of the velocity oscillation is further increased, these twodimensional vortices break-up and give rise to three-dimensional
vortices and a fully turbulent ow. The mechanisms, which give
rise to the coherent vortex tubes and later to turbulence, appear
to be similar to that operating in oscillatory boundary layers
(Costamagna et al., 2003; Carstensen et al., 2010).
For practical applications, it is unfeasible to make direct numerical simulations of the ow in the bottom boundary layer. Hence to

evaluate the ow resistance and the sediment transport induced by


a solitary wave, the effects of the bottom boundary layer on the
nearshore circulation and the coastal morphology, are taken into
account by using empirically-based formulae or simple turbulence
closures. However, even though simple turbulence models can provide a description of the gross features of the boundary layer, they
fail to provide an accurate description of turbulence dynamics.
The turbulent boundary layer at the bottom of a solitary wave
was studied by Liu (2006) by considering the Reynolds averaged
equations. To close the problem, Liu (2006) used the Boussinesq
assumption and assumed the eddy viscosity to be a power function
of the distance from the bottom. The analytical solution of Liu
(2006) allows interesting results to be obtained and in particular
the wave damping rate to be quantied.
However, as pointed out by Blondeaux and Vittori (2012), the algebraic turbulence model employed by Liu (2006) does not allow an
accurate description of turbulence structure. Suntoyo and Tanaka
(2009) obtained a better description of turbulence dynamics by using
the two-equation turbulence model of Menter (1994). This turbulence model gives rise to values of the turbulent kinetic energy and
of the bottom shear stress during the decelerating phase which are
larger than those of the accelerating phase, as observed in the experiments. However, the experimental measurements made by Sumer
et al. (2010) show a well dened second peak of the wall shear stress
which is not reproduced by the model used by Suntoyo and Tanaka
(2009). Moreover, the Reynolds number predicted by Suntoyo and Tanaka (2009) for turbulence appearance is smaller than the value suggested by the laboratory experiments of Sumer et al. (2010).
In the framework of the research project n. 2008YNPNT9-003
Idrodinamica e morfodinamica nella regione dei frangenti, supported by the Ministero dellIstruzione, dellUniversit e della Ricerca, the boundary layer generated by the propagation of a solitary
wave was studied.
(i) By investigating the stability of the laminar ow, using a linear stability analysis and a momentary criterion of instability (Shen, 1961),
(ii) By performing direct numerical simulations of continuity
and NavierStokes equations,
(iii) By using the two-equation turbulence model by Saffman
(1970) and Saffman and Wilcox (1974).
The results of these studies are described in detail in Blondeaux
et al., 2012, Vittori and Blondeaux (2008, 2011), Blondeaux and
Vittori (2012) and are reviewed hereinafter along with the recent
experimental ndings of Sumer et al. (2010) to provide an integrated and exhaustive description of the phenomenon.
Similarities exist between the boundary layer at the bottom of a
surface solitary wave and that generated by long internal waves.
Internal waves have been extensively investigated and a review
of the available results can be found in Garret and Munk (1979),
Ostrovsky and Stepanyants (1989), Grimshaw et al. (1998), Staquet
and Sommeria (2002), Helfrich and Melville (2006).
The inviscid dynamics of propagating long internal waves is
usually determined introducing the weakly nonlinear Kortewegde Vries equation. If a two-layer system is considered and the rigid
lid approximation is introduced, the solution of the Korteweg-de
Vries equation shows the existence of both elevation and depression internal solitary waves. The former take place when thickness
of the upper layer is larger than the thickness of the lower layer
and viceversa.
However, the obtained results are not quantitatively accurate,
since internal waves are often characterized by large amplitudes
and strong nonlinear effects. A useful extension of the Kortewegde Vries model is that which includes cubic nonlinearity (Djordjevic
and Redekopp, 1988). The results obtained by means of the ex-

401

G. Vittori, P. Blondeaux / Progress in Oceanography 120 (2014) 399409

tended Korteweg-de Vries equation show that the wave crest


broadens and tends to becomes at as the amplitude increases.
Then, the Korteweg-de Vries model can be further extended by considering nite amplitude waves in shallow waters or by numerically integrating Euler equations.
Then, the dynamics of inviscid internal waves controls the ow
dynamics in the viscous bottom boundary layer, which is driven by
a sequence of adverse and favorable pressure gradients, the timing
and duration of which depend on the wave phase and on main characteristics of the inviscid wave (e.g. elevation/depression wave).
If the ow regime is laminar, the ow in the bottom boundary
layer can be determined either by using the recent analytical solution of Liu and Orla (2004) or by numerical means. However, for
large wave amplitudes, the laminar solution turns out to be unstable and two-dimensional vortex structures or three-dimensional
turbulence are expected to appear.
The instability characteristics of the ow in the boundary layer
were investigated by many authors (e.g. Bogucki and Redekopp,
1999; Wang and Redekopp, 2001; Stastna and Lamb, 2002; Diamessis and Redekopp, 2006) who focused their attention on the spatially-nonuniform separation region where an inectional velocity
prole is present. In fact, as pointed out by Hammond and Redekopp (1998), the growth rates generated by the inviscid instability,
charateristic of inectional velocity proles, are signicantly larger
than those originated by the viscous mechanism characteristic of
attached boundary layers.
More recently, Lin and Redekopp (2011) made numerical simulations of the boundary layer generated under an internal long solitary wave of depression in a two-layer model of actual density
stratications by considering the turbulent regime and using a
RANS approach and a two-equation turbulence model.
Notwithstanding the similarities between the boundary layers
at the bottom of internal and surface solitary waves, a quantitative
comparison of the results described in the following sections with
those obtained considering internal waves is neither described
hereinafter nor was performed. Indeed, the internal wave case is
usually characterized by a different spatial/temporal forcing term
and quite often a steady current is superimposed to the unsteady
motion generated by the internal wave. Moreover, numerical simulations of the ow generated close to the bottom by internal
waves are sometimes carried out by considering density stratications. The readers can easily make themselves a qualitative
comparison.
The remainder of the paper is organized as follows. In Section 2,
we formulate the problem and describe the different approaches
which are used to obtain the results which are presented and discussed in Section 3. The conclusions are drawn in Section 4.

2.1. Formulation of the problem


A solitary wave of height H is considered (hereinafter a star denotes a dimensional quantity and the same symbol without star its
dimensionless counterpart). The wave is assumed to propagate on
a constant water depth h. As pointed out in the Introduction, the
Reynolds number of the phenomenon is assumed to be large and
the uid domain is split into a core region, where the uid behaves
like an inviscid uid, and boundary layers, where viscous effects
are important. The ow in the core region can be studied introducing the following dimensionless variables:

p
gh
t
;

h

X 1 ; X 2 ; X 3

V 1 ; V 2 ; V 3

V 1 ; V 2 ; V 3
p

H

g h

X 1 ; X 2 ; X 3
;

h

g
H

;
1




where t indicates the time and X 1 ; X 2 ; X 3 denote Cartesian coordinates with the X 1 -axis aligned with the direction of wave propagation and the X 2 -axis in the vertical direction and pointing upward
(X 2 0 describes the bottom). Moreover, in (1), g is the free sur

face elevation with respect to the still water level, V 1 ; V 2 ; V 3 denote the velocity components and the dimensionless parameter

H

h

is a measure of the wave height. Even though the most appropriate horizontal length scale would be a measure L the wavelength
of the solitary wave, the coordinates X 1 and X 3 are scaled by h because the use of h makes the dispersion parameter l = h/L to
disappear from both the problem and its solution (Grimshaw,
1970, 1971). As pointed out in Vittori and Blondeaux (2008), the
reader can easily convert the problem formulated in the following,
as well as its solution, into that obtained using a conventional

length p
scale
L of the solitary wave as horizontal length scale


and L = g h as time scale.
The free surface elevation and the velocity eld in the inviscid
region were obtained by Grimshaw (1970, 1971) by expanding
them in terms of the small parameter H. The leading order solution
reads
2

g s ; V1 s ;

p
V 2  3HX 2 s2 tanh

"r
#
3H
X 1  t
4

In (3), the variable s is dened by

"r
#
3H
X 1  t
s sech
4

The water depth h is not the appropriate length scale to analyze the ow in the bottom boundary layer, which is characterized
q
p

by a thickness of order d 2m h = g  h which turns out to be
much smaller than h. Hence, let us introduce the new spatial
variables,

x1 ; x2 ; x3

X 1 ; X 2 ; X 3
d

and let us denote with (v1, v2, v3) the dimensionless velocity components within the bottom boundary layer.
Then, the hydrodynamic problem turns out to be posed by continuity and NavierStokes equations which, in dimensionless form,
read

@v j
0;
@xj

 
 
@v i
H
@v
H @p 1 @ 2 v i

vj i 

d
d @xi 2 @xn @xn
@t
@xj

i 1; 2; 3

2. Formulation of the problem and solution methods

t

where use is made of Einstein summation convention, the dimensionless dynamic pressure p = p/(H2qgh) is introduced and the
dimensionless parameter d is dened as the ratio between the
dimensional thickness of the boundary layer and the local water
depth

d

h

s
2m
p

h g h

Of course, the velocity far from the bottom should match the irrotational ow and satisfy the no-slip condition at the bottom. The
matching with the irrotation ow far from the bottom, i.e. for x2
much larger than one, suggests the introduction of the modied
~:
pressure p



~
@p
@p
d @V 1

@x1 @x1 H @t X 2 0

402

G. Vittori, P. Blondeaux / Progress in Oceanography 120 (2014) 399409

where @V 1 =@tX2 0 can be computed on the basis of the time derivative of the irrotational velocity (3).

If (13) is plugged into vorticity Eq. (12), the following differential equation is obtained

2.2. The laminar solution

v^ 1 x2 ; t  ctN2 f x2 ; t 

If the ow regime is assumed to be laminar, the leading order


solution of the problem formulated Section 2.1 is described in text
book (e.g. Mei, 1989). The velocity component v2 and v3 vanish and
the streamwise velocity component reads

2
v^ X 1 ; x2 ; t V 1 X 1 ; t  p
p #
"r

Z 1
2
3H x2
2
2
en dn
sech

X

t
1
4 2n2
0

The experimental results of Sumer et al. (2010) show that the


ow described by (9), which is observed for a relatively weak free
stream velocity, turns out to be unstable if the external velocity is
increased. To determine the values of the parameters discriminating between the laminar and turbulent regimes, Blondeaux et al.
(2012) have recently investigated the stability of (9) by means of
a linear stability analysis.
Since Squires theorem holds also in unsteady ows (Conrad
and Criminale, 1965), a two-dimensional perturbation of the ow
described by (9) is considered such that

v 01 ; v 02

10

A linear analysis of the time development of the perturbation can be


performed by assuming that the amplitude  of the perturbation is
much smaller than one. By neglecting terms of order 2 and introducing the streamfunction w such that

v 01 ; v 02

@w
@w
;
@x2
@x1


11

the linearized vorticity equation reads

"
#
@3w
@3w
H
@3w
@3w
@ 2 v^ @w

v^ 3

@t@x21 @t@x22 d
@x1 @x1 @x22 @x22 @x1
!
1 @4w @4w
@4w

2
2 @x41 @x42
@x21 @x22

where the operator N is dened by

N2

@2
 a2
@x22

15

@f x2 ; t
0 f x2 ; t 0 at x2 0
@x2
@f x2 ; t
! 0 f x2 ; t ! 0 for x2 ! 1
@x2

16
17

As pointed out in Blondeaux et al. (2012), the viscous term, which is


of order d/H, is retained in (14) since it is signicant in a viscous
layer close to the wall and within possible critical layers. We notice
that the time variable t and the spatial variable X 1 appear into the
problem posed by (6) and (8) only in the combination

X 1 

p
g h t
X1  t

h

18

which can be considered as a parameter of the problem. Then, Eq.


(14) can be solved for different values of f which correspond at different locations and/or at different phases within the wave cycle. In
the following, X1 is set equal to 0 in such a way that a vanishing value of f corresponds to the passage of the wave crest at X1 = 0. In order to nd a nonvanishing solution of the differential problem just
formulated, it is necessary to force an eigenrelation which provides
the value of c as function of f and the parameters H and d of the
problem. The details of the numerical approach which is employed
to nd the eigenrelation and the eigensolution are described in
Blondeaux et al. (2012).
2.4. The direct numerical simulations

12

The linearity of (12) allows a normal mode analysis to be made


and a generic Fourier component along the x1-axis to be considered. Moreover, since the height H of a solitary wave is usually
much larger than the thickness d of the viscous boundary layer,
the value of the ratio H/d turns out to be much larger than one.
Therefore, the amplitude of the perturbation can be supposed to
growth on a time scale much faster than that which characterizes the time development of the basic ow and a momentary
criterion for instability (Shen, 1961; Blondeaux and Seminara,
1979) can be used. Hence, the function w can be written in
the form



Z
H
csds expiax1 c:c
wx1 ; x2 ; t f x2 ; t exp ia
d

14
2

To close the problem, the following homogeneous boundary


conditions should be forced

2.3. The linear stability analysis of the laminar ow

v 1 ; v 2 v^ ; 0 

@ 2 v^ x2 ; t
1
f x2 ; t
N4 f x2 ; t
2iaH=d
@x22

13

where c.c. stands for the complex conjugate of a complex quantity


and a indicates the streamwise wavenumber of the generic Fourier
component. The real part (cr) of c is related to the wavespeed of the
Fourier component and the imaginary part (ci) controls the growth/
decay of its amplitude.

To determine the velocity eld in the turbulent regime, the governing Eq. (6) are solved numerically (Direct Numerical Simulation) by means of a nite difference approach in a computational
domain characterized by dimensions L1, L2 and L3 in the streamwise, vertical and spanwise directions, respectively.
Appropriate boundary conditions are forced at the boundaries
of the computational domain. At the bottom (x2 = 0) the no-slip
condition is forced. Since, for large values of x2, the ow should
tend to the solution in the core region, at x2 = L2 we force @(v1,
v3)/@x2 = 0 and v2 = 0. Of course L2 should be sufciently large to
be outside the viscous boundary layer. At last, by assuming that
L1 and L3 are much larger than the size of the turbulent eddies
which develop within the boundary layer, turbulence structure is
supposed to be homogeneous in the streamwise and spanwise
directions and periodic boundary conditions are forced along the
x1 and x3 axes. Of course, the use of periodic boundary conditions
is justied if the computational box is large enough to include
the largest eddies generated by turbulence appearance. Checks
on the size of the computational box have been carried out to verify this assumption. The computational mesh is uniform in the
streamwise and spanwise directions while in the vertical direction
a nonuniform mesh is used to cluster the grid points close to the
bottom where velocity gradients are expected to be larger.
The numerical method solves the problem in primitive variables
using standard centered second-order nite difference approximations of the spatial derivatives, while the time-advancement of

G. Vittori, P. Blondeaux / Progress in Oceanography 120 (2014) 399409

403

X d
X p

20

2NavierStokes equations employs the fractional-step method


extensively described by Kim and Moin (1985) and Rai and Moin
(1991). More details on the numerical approach can be found Verzicco and Vittori (1996) and Vittori and Verzicco (1998).
Since the numerical simulations start when the solitary wave is
far from the location where the time development of the boundary
layer is computed, vanishing velocity and pressure elds would be
the most appropriate initial condition.
To trigger transition and turbulence appearance, it is necessary
to introduce a perturbation of the laminar ow. This can be made
either by adding a small random but divergence-free velocity eld
to the initially vanishing velocity eld or by considering wall
imperfections (Blondeaux and Vittori, 1994), i.e. a bottom waviness of quite small amplitude superimposed to the perfectly at
bottom. The former procedure was used by Vittori and Blondeaux
(2008) while the latter procedure was employed by Blondeaux
et al. (2012). The introduction of a bottom waviness has the advantage that a small but known perturbation is present at any wave
phase. On the other hand, the perturbation introduced at the
beginning of the numerical simulation decays during the early
stages of the simulation and attains an uncontrolled amplitude at
the beginning of the decelerating period when the instability of
the laminar ow takes place.
2.5. The Reynolds averaged model
The direct numerical simulations provide a detailed picture of
the ow eld but the power of actual computers does not allow
the use of this approach for practical applications. To investigate
eld cases, the boundary layer generated by the propagation of a
solitary wave can be studied by considering RANS equations and
a two-equation turbulence model. Blondeaux and Vittori (2012)
used the model of Saffman (1970) (see also Saffman and Wilcox
(1974)). This model appears to provide a fair description of unsteady boundary layers as shown by Blondeaux (1987), Foti and Scandura (2004) and Cavallaro et al. (2011) who successfully applied it
to investigate turbulence structure in the boundary layer generated at the bottom of a progressive sea wave. In particular, the results obtained by Blondeaux (1987),Foti and Scandura (2004) and
Cavallaro et al. (2011) show the capability of the model to provide
a fair description of the differences between turbulence structure
during the accelerating and decelerating phases of the wave cycle
and to give an approximate description of the transition process
from the laminar regime to the turbulent regime. The Reynolds
stress tensor is expressed in terms of an eddy viscosity mT and of
the rate of the strain tensor of the mean velocity eld. Then, the
eddy viscosity is assumed to be a function of two turbulence local
properties, namely a pseudo-energy e and a pseudo-vorticity X,
which are assumed to satisfy nonlinear advectiondiffusion
equations.
The introduction of the boundary layer approximation reduces
the problem to the determination of v1(x2, t) which is provided
by the solution of the x1-component of the Reynolds averaged NavierStokes equation

@v 1

@t

@V 1
@t

X 2 0

1 @
2 @x2



mT @ v 1
m @x2


19

Of course in (19), the velocity component v1 is meant to be the Reynolds average of the actual velocity. Moreover, the convective term
is neglected in (19) because the wave height is assumed small enough to use Korteweg and De Vries approach to describe the
dynamics of the solitary wave.
As already pointed out, the pseudo-energy e and the pseudovorticity X obey nonlinear advectiondiffusion equations. By
introducing the dimensionless variables

e
Hg  d

g  h0

the equations of Saffmans model read







@ v 1
@e H
mT @e
 be X 1 @
e ae
1

r
e 
@t
d
2 @x2
@x2
m @x2
"
#




2
@ v 1
@X
H 2
1
@
mT @ X2


X aX
 bX X
1 rX 
d
2 @x2
@t
@x2
m @x2

21
22

where ae, aX, be, bX, ce, cX are assumed to be universal constants.
Moreover, the assumption that turbulence characteristics are fully
determined by the knowledge of e and X and the use of dimensional arguments lead Saffman (1970) to write mT in the form

mT cm

e

X

or

mT cm

mT
e
2cm
m
X

23

The values of the constants were determined by Saffman and


Wilcox (1974)

ae 0:3; be a22 ; re 0:5; bX 0:18;


aX ae



bX 0:32
;

be
ae

cm 1:

24

At the bottom, the velocity satises the no-slip condition. Moreover,


Saffman (1970) postulated that the pseudo-vorticity depends on a
dimensionless roughness parameter zw zr us =m through a univerp
sal function S(zw), zr and us sw =q being the bottom roughness
and friction velocity, respectively and sw is the bottom shear stress

X

sw
Sz
q m ae w

25

Later Saffman and Wilcox (1974) related the function S to the universal logarithmic velocity prole. However, they did not provide an
explicit relationship for S which was obtained by Blondeaux and
Colombini (1985). Finally, the pseudo-energy should vanish for
x2 = 0.
To trigger turbulence appearance, it is necessary to introduce a
perturbation of e either at the beginning of the numerical simulation or at the wall. While the former approach simulates the presence of a disturbance of the initial ow eld, the latter procedure
simulates the effect of wall imperfections which, as already discussed, play a key role in triggering transition to turbulence in unsteady boundary layers. Hence, the second approach is presently
preferred and transition to turbulence depends on the value ew of
e at the wall. The results described in the following are obtained
by xing ew = 1018, but a few runs have been made with different
values of ew to detect the effects that the amplitude of the wall
imperfections has on transition to turbulence. The reader should
consider that, even though the results described in the following
seem to suggest that the model equations can describe the transition process, accurate quantitative predictions of the critical conditions are not expected since the two-equation model of Saffman is
not designed to follow the transition process.
Momentum Eq. (19) along with (21) and (22), which describe
turbulence dynamics, are numerically integrated with a second order nite difference approach to approximate spatial derivatives
and a second order RungeKutta approach to advance in time.
3. The results
3.1. The linear stability analysis
The eigenvalue problem formulated in Section 3.3 can be solved
by xing the values of d and H and by varying the parameter f. The
description of the numerical approach employed to determine the
solution of (14) with boundary conditions (16) and (17) is given in

404

G. Vittori, P. Blondeaux / Progress in Oceanography 120 (2014) 399409

Blondeaux et al. (2012) and it is not repeated herein for the sake of
space.
An example of the results is shown in Fig. 1, where the imaginary part ci of the eigenvalue c is plotted as function of f and a
for xed values of H and d, namely H = 0.12 and d = 5  104. Positive values of ci are found for f larger than fins. Hence, for such values of H and d, the basic laminar regime turns out to be unstable as
soon as f becomes larger than fins, i.e. during the decelerating
phase.
If the fastest growing Fourier component (f.g.F.c) is assumed to
prevail on the other modes, the linear analysis predicts the appearance of a periodic pattern characterized by a wavelength k such
that 2pd/k equals the value amax of a which gives rise to the maximum growth rate. Therefore, the wavelength k of the most unstable Fourier component would seem to depend on the phase within
the wave cycle and the periodic patterns predicted by the stability
analysis would be characterized by a wavelength which depends
on time. However, as pointed out by Blondeaux et al. (2012), once
the perturbation appears, its wavelength cannot change continuously but only through the appearance of defects, the dynamics
of which can be studied only by means of a fully nonlinear
approach.
A reasonable assumption, which can be made, is that the actual
wavelength of the periodic pattern is coincident with the wavelength predicted for f = fins. This assumption is supported by the
experimental visualizations of Sumer et al. (2010), which show
no signicant change of the wavelength of the vortex tubes which
are generated by the growth of the perturbations.
To ascertain the reliability of the analysis, Blondeaux et al.
(2012) compared the predicted wavelengths with the experimental measurements of Sumer et al. (2010). Fig. 2 shows a plane view
of the two-dimensional vortical structures (vortex tubes) visualized by Sumer et al. (2010) for a maximum value
of the free stream
p

velocity U 0m equal to 50.9 cm/s and T  4ph = 3g  H 9:3 s. The
average distance between the axes of adjacent vortex tubes is
about 2.4 cm, a value which is close to the wavelength of the
f.g.F.c. evaluated for f = fins which is about 2.9 cm. Similar results
are obtained when the other experiments made by Sumer et al.
(2010) are considered (see Blondeaux et al., 2012). In particular,
Fig. 3 shows a comparison between the predicted values of k/d
and the range of the observed values for the three movies which
are available at http://journals.cambridge.org and one extra movie
which has been made available by Prof. Sumer who provided also
the range of the observed wavelengths.

The discussion of the results concerning the critical conditions


and the phase of the wave cycle at which the external perturbations start to grow is postponed to Section 3.2, since the theoretical
predictions of the stability analysis can be better understood after
the discussion of the results of the Direct Numerical Simulations.
3.2. The direct numerical simulations
Since the ow in the bottom boundary layer depends on H and
d, Vittori and Blondeaux (2008) made a large number of direct
numerical simulations for different values of H and d. According
to the stability analysis, the numerical results show that the ow
can be either laminar or turbulent depending on the values of
the parameters and on the wave phase.
Fig. 4 shows the time development of the longitudinal velocity
component close to the bottom for a xed value of d (namely
4.75  104) and three different values of H which correspond to
waves of different height propagating in water of xed depth.
When H is relatively small (H = 0.1), the ow regime is laminar
and the velocity computed numerically does not differ from the
velocity predicted by means of (9).
Transition to turbulence is triggered for larger values of H and
turbulence presence is shown by the appearance of random velocity uctuations superimposed to the average velocity eld. For the
intermediate value of H, namely H = 0.2, the velocity uctuations
appear after the velocity in the boundary layer has attained the
minimum negative value. These velocity uctuations have large
amplitudes and an analysis of the vorticity eld shows that these
uctuations are generated by large scale migrating coherent vortices which, after Sumer et al. (2010), are named vortex tubes. At
these phases, even though the vortex structures are generated by
the instability of the basic ow, the regime cannot be dened turbulent since three-dimensional effects are absent.
Only later, the two-dimensional vortices break-up and generate
incoherent eddies and give rise to a fully three-dimensional turbulent ow. If larger values of H are considered, the inviscid ow
close to the bottom is characterized by stronger adverse pressure
gradients. The resulting stronger deceleration induces transition
from the laminar regime to a three-dimensional turbulent ow
without the appearance of coherent transverse vortex tubes.
Fig. 5 shows the ow regimes for all the runs made by Vittori
and Blondeaux (2008). In a few cases, for values of H close to the
critical conditions, the vortex tubes generated by the growth of
Tollmein-Schilchting waves do not break-up but slowly dissipate
because of viscous effects. Hence, the ow cannot be dened tur-

0.6
0.5

0.4
0.3
0.2
0.1

ins

10

12

14

Fig. 1. Growth rate (imaginary part ci of the eigenvalue c) plotted versus f and a for
H = 0.12 and d = 5.0  104. The thin continuous lines correspond to positive values
of ci, the thick continuous lines corresponds to ci = 0, the thin broken lines
correspond to negative values (Dci = 0.005). The thick broken line corresponds to
the fastest growing Fourier component (a = amax).

Fig. 2. Visualization of transverse tubes


which appear during the decelerating
 p
phase for U 0m 50:9 cm/s, T  4ph = 3g  H 9 s (the image has been kindly
provided by Prof. Sumer. The movies of the experimental visualisations can be seen
at journals.cambridge.org/m).

405

40

0.8

35

0.7

30

0.6

25

0.5

*
( /*)predicted

G. Vittori, P. Blondeaux / Progress in Oceanography 120 (2014) 399409

20

0.3

15

laminar regime
transitional regime
turbulent regime

0.2

10

0.1

5
0

0.4

0
0

10

15

20

25

30

35

40

0.0005

0.001

(*/*)observed
Fig. 3. Dimensionless wavelength of the most unstable Fourier mode predicted by
the stability analysis plotted versus the average dimensionless distance between
the adjacent vortex tubes visualized by Sumer et al. (2010).

0.25
0.2
0.15

v1

0.1
0.05
0
-0.05
H=0.5

-0.1
-0.15
-10

-5

H=0.1
H=0.2
5

10

15

0.0015

0.002

0.0025

20

Fig. 4. Time development of the dimensionless streamwise velocity component for


d = 4.75  104 and H = 0.1 (at x2 = 0.34), H = 0.2 (at x2 = 0.34) and H = 0.5 (at
x2 = 0.29). Adapted from Vittori and Blondeaux (2008).

bulent and the term transitional regime is introduced. The


numerical results show that the instability of the basic laminar
ow takes place when the value of H is larger than a critical value
Hc which increases when d is increased. The dependence of Hc on d
was observed also by Sumer et al. (2010) who proposed to introduce a critical
value Rec of the Reynolds number
p
Re 4H3=2 = 3d2 such that for Re larger than Rec, the laminar ow
is unstable.
To discriminate between the laminar regime and the transitional regime, Sumer et al. (2010) proposed a value of Rec equal
to 2  105. Moreover, Sumer et al. (2010) proposed a value Ret
equal to 5  105 to discriminate between the transitional regime
and the fully three-dimensional turbulent regime. The numerical
results, shown in Fig. 5, compared with Sumer et al.s criteria, seem
to suggest larger values of both Rec and Ret.
A possible explanation of the differences between the numerical
results and the experimental measurements is provided by the
external sources of signicant perturbations (e.g. vibrations of
the experimental apparatus) which might play a signicant role
in the transition process, signicantly lowering the critical values
of Re. A similar phenomenon is observed in the oscillatory boundary layer (Stokes layer) generated by a monochromatic sea wave
(Blondeaux and Vittori, 1994).
The existence of a critical value of H which depends on d is
shown also by the linear stability analysis described in Section 2.
However, signicant quantitative differences are found, since the

Fig. 5. Flow regime in the plane (d, H). The continuous line represents the curve
Re = 2  105, which Sumer et al. (2010) suggest to be the limit for the appearance of
the transitional regime, and the broken line represents the curve Re = 5  105 which
is the limit for turbulence appearance. Adapted from Vittori and Blondeaux (2008).

stability analysis predicts values of Hc which are much smaller


than those predicted by the numerical simulations and/or observed
in the experiments. This nding can be explained by assuming that
both in the direct numerical simulations and in the experiments,
the perturbations grow but, for values of H close to Hc, attain
amplitudes too small to be visually detected. This assumption
has been tested running the numerical code of Vittori and Blondeaux (2008).
Fig. 6 shows the dimensionless kinetic energy per unit area K of
the ow perturbations plotted versus the phase f within the wave
cycle for different values of H and d = 1.2  103. For H = 0.025, the
kinetic energy of the perturbation tends to decay as predicted by
the linear stability analysis. If larger values of H are considered
(e.g. H = 0.2), the kinetic energy deviates from that found for the
stable case even though the growth of K is very weak and does
not lead to the appearance of signicant perturbations of the velocity eld. Only when H is equal to 0.5 or larger, the values of K become signicant and the ow deviates from the laminar ow. Of
course, the process is continuous and there is some arbitrariness
in the denition of the value of H which gives rise to a turbulent
ow. Indeed, Vittori and Blondeaux (2008, 2011) introduced an
arbitrary threshold value of K to discriminate between the laminar
and the turbulent regimes.
The results obtained by means of Direct Numerical Simulations
explain also the differences between the values of fins predicted by
the stability analysis and those at which Sumer et al. (2010) ob-

0.1

H=0.6

0.001

K 1e-05
H=0.4

1e-07

H=0.2

1e-09

H=0.025
-5

10

15

20

Fig. 6. Dimensionless kinetic energy K of ow perturbations plotted versus f for


d = 1.2  103 and different values of H. Adapted from Vittori and Blondeaux (2011).

406

G. Vittori, P. Blondeaux / Progress in Oceanography 120 (2014) 399409

served vortex tubes. Indeed, Fig. 6 shows that there is a signicant


time lag between the beginning of the growth of the perturbations
and the phase at which they attain a signicant level such to allow
their visual observation.
From an engineering point of view, the most relevant quantities
are Reynolds averaged quantities and in particular the average
velocity and the average bottom shear stress. Because of the homogeneity of the ow in the two horizontal directions, averaged values are computed by averaging over planes characterized by
constant values of x2.
Sometimes in turbulent ows, the average velocity prole close
to a wall is assumed to follow the logarithmic law. However, the
numerical simulations show that the velocity proles are rather
different from those provided by both the log-law and the Van Driest law, if moderate values of H are considered. Only for large values of H, there are phases for which the velocity prole becomes
close to the logarithmic law. However, the logarithmic law holds
in such a small time interval and for such a restricted range of
the parameters that its general use is questionable (see also Sumer
et al., 2010).
In depth-averaged models, the averaged bottom shear stress s
is usually evaluated by means of the Chezy law and s is assumed
to be proportional to the square of the irrotational velocity evaluated at the bottom (s qcf V 1 jV 1 kX 2 0 , cf being a drag coefcient).
However, the results of Liu et al. (2007) for the laminar regime and
those of Blondeaux and Vittori (2012) for the turbulent regime (see
Fig. 7) show that the bottom shear stress reverses its direction during the late decelerating phase, while V 1 jX 2 0 points always in the
direction of wave propagation. Hence, the relationship
s qcf V 1 kV 1 jX 2 0 certainly fails after the passage of the wave
crest. Moreover, the laminar solution and the turbulent results
show that a phase shift between the maximum of the irrotational
velocity and that of the bed shear stress is also present. In fact, the
time development of the bottom shear stress (Fig. 7) shows a maximum value, which is attained slightly before the passage of the
wave crest. The value of smax as well as its phase shift Dtmax, with
respect to that of the maximum velocity are discussed in Vittori
and Blondeaux (2008) who analyzed also the smallest value smin
of s and its phase.

3.3. The Reynolds averaged model


Both the direct numerical simulations of Vittori and Blondeaux
(2008, 2011) and the experimental measurements of Sumer et al.
(2010) show that the ow regime keeps laminar throughout the
whole wave cycle if small values of H are considered. If the RANS

0.7

model described in Section 2.5 is run for a small value of H, the


pseudo-energy e turns out to be negligible during the whole simulation and the eddy viscosity is practically equal to zero. It follows
that the velocity eld does not differ from the laminar solution. In
this case the ow is not affected by the value of bottom roughness.
If the value of d is kept xed and the value of H is increased, a
critical value Hc is encountered such that for values of H just larger
than Hc, the value of e, which is initially zero, grows explosively
during the decelerating phase, it attains signicant valeus and a
turbulent ow appears. If the value of H is further increased, significant values of the pseudo-energy e appear earlier and e attains
much larger values as shown by Fig. 8 where the value of
R1
Et 0 ex2 ; tdx2 is plotted versus time for a xed value of d
and different values of H.
The ow regimes (laminar/turbulent) determined by means of
the RANS model are shown in Fig. 9. The value of the eddy viscosity, which is used to identify the ow regime, is a continuous function of the parameters. Therefore, to discriminate among laminar,
transitional and turbulent regimes on the basis of a quantitative
criterion, the ow is dened to be laminar if the maximum value
of the ratio mT =m during the wave cycle is smaller than 1. On the
other hand, if the maximum value of the ratio mT =m is larger than
10, the ow regime is dened to be turbulent. Values of the maximum of mT =m within the range (1,10) are supposed to characterize
the transitional regime. The results plotted in Fig. 9 show that transition to turbulence takes place when the parameter H is larger
than a critical value Hc which depends on d. As found by Vittori
and Blondeaux (2011), Hc increases when d is increased. The results
plotted in Fig. 9 have been obtained by xing zr equal to 0.01 d.
Such a value of the bottom roughness leads to the smooth wall regime and both the transition limits and the turbulence characteristics do not depend on zr. Indeed, results equal to those shown in
Fig. 9 are obtained if the value
of zr is slightly changed but zr is kept
p
smaller than 5m =us , us s =q being the shear velocity. To allow
a comparison between the results of the RANS model and the
experimental measurements of Sumer et al. (2010), the empirical
curves proposed by Sumer et al. (2010) are also plotted in Fig. 9.
The use of a RANS model allows the rough bottom case to be
considered, too. Indeed, as soon as zr becomes larger than 5m =us ,
the bottom roughness affects the transition process from the laminar to the turbulent regime as well as the turbulent ow. For
example, Fig. 10 shows the ow regimes as function of H and d
when the dimensionless roughness zr is equal to 1. Comparing
Fig. 10 with Fig. 9, it appears that an increase of zr leads to a significant decrease of the transition limit between the transitional and
turbulent regimes even though the limit between the laminar and

H=0.1
H=0.2
H=0.3
H=0.4
H=0.5

0.6
0.5
0.4

14

H=0.5
10
H=0.4

0.2

0.3

0.1

H=0.3

-0.1
-0.2
-0.3
-20

H=0.6

12

2
-15

-10

-5

10

15

20

Fig. 7. Time development of the dimensionless bottom shear stress s = 2s/


(Hqgd) for d = 0.475  103 and different values of H. Adapted from Vittori and
Blondeaux (2011).

0
-20

-15

-10

-5

H=0.2
H=0.1
5
10
15

20

Fig. 8. Pseudo-energy per unit area of the sea bottom plotted versus f for d = 0.0008
and zr = 0.01. Adapted from Blondeaux and Vittori (2012).

407

G. Vittori, P. Blondeaux / Progress in Oceanography 120 (2014) 399409

0.8

2.5

ew=10

0.7

Re=5 10

0.6

w /max,1

ew=10-36

1.5

0.5

Re=2 105

0.4
0.3

-32

ew=10-40

1
0.5

0.2

laminar regime

0.1

transitional regime

turbulent regime
0

0.0005

0.001

0.0015

0.002

-0.5

0.0025

0.7
Re=5 10

0.6
0.5

-4

-2

Fig. 11. Bottom shear stress plotted versus time for zr = 0.01 and values of d and H
such that Re = 2  106 (H = 0.23, d = 0.00038). Continuous lines = present results,
white points=Sumer et al. (2010)s measurements. Adapted from Blondeaux and
Vittori (2012).

and the measurements of Sumer et al. (2010) is obtained for


ew = 1040. However, since the level of the external disturbances
depends on the experimental equipment, the value of ew to be used
cannot be considered to be universal. Moreover, it is worth pointing out that, even though the value of ew affects the results when H
is close to Hc, the value of ew does not signicantly affect the results, when the amplitude of the solitary wave is much larger than
its critical value (Blondeaux and Vittori, 2012).

0.8

0.4
Re=2 10

3.4. The sediment transport

0.2

laminar regime

0.1

transitional regime

-6

Fig. 9. Flow regime in the plane (d, H) for zr = 0.01. The continuous line represents
the curve Re = 2  105, which Sumer et al. (2010) suggest to be the limit for the
appearance of the transitional regime, and the broken line represents the curve
Re = 5  105 which is the limit for turbulence appearance. Adapted from Blondeaux
and Vittori (2012).

0.3

-8

turbulent regime
0

0.0005

0.001

0.0015

0.002

0.0025

Fig. 10. Flow regime in the plane (d, H) for zr = 1. The continuous line represents the
curve Re = 2  105, which Sumer et al. (2010) suggest to be the limit for the
appearance of the transitional regime, and the broken line represents the curve
Re = 5  105 which is the limit for turbulence appearance. Adapted from Blondeaux
and Vittori (2012).

transitional regimes is not signicantly affected by the changes of


zr .
As already pointed out, for practical applications a detailed
knowledge of the ow within the bottom boundary layer is not
necessary, while it is important to evaluate the value of the bottom
shear stress s. As already discussed in Section 3.2, the relationship
s qcf V 1 kV 1 jX2 0 fails after the passage of the wave crest, when
the bottom shear stress reverses its direction. This qualitative nding is shown also by the results of the RANS model. Indeed, the
explosive growth of turbulence, which takes place during the
decelerating phase of the wave cycle, leads to a signicant difference between the accelerating and decelerating phases of the wave
cycle and to a second positive peak of the bottom shear stress, the
existence of which is also shown by the experimental measurements of Sumer et al. (2010). Fig. 11 shows a comparison between
the present results and some of the laboratory data obtained by Sumer et al. (2010), for Re = 2.0  106. As in Sumer et al. (2010), the
bottom shear stress is scaled with its maximum value smax, which
takes place just before the maximum of the irrotational velocity.
The different curves obtained by means of the RANS model and
plotted in Fig. 11 are characterized by different values ew (we remind to the reader that ew if the value of e forced at the wall).
The best agreement between the prediction of the present model

Once the ow in the bottom boundary layer is determined, the


sediment dynamics can be determined with different approaches.
For example, Vittori (2003) used the results of direct numerical
simulations and computed sediment trajectories integrating Newtons law for each sediment particles. However, this approach cannot be used for practical applications because of the high
computational costs. In the following, a more heuristic approach
is employed and the sediment transport is splitted into bed load
and suspended load (Blondeaux and Vittori, 1999). The former
can be evaluated using an empirical formula. presently, we use
Fredsoe and Deigaards (1992) formula

p
p
Q b
30

Q b q
h  hc h  0:7 hc
p
 3
qs =q  1g  d

26

where the Shields parameter h is dened by

h

s 
q  q g  d

s

27

and hc is the critical value of the Shields parameter for the initiation

of sediment motion. In (27) qs ; d are the density and the size of the
sediment, respectively. To estimate the suspended load, the sediment concentration c can be calculated by integrating the equation
of mass balance, assuming that the sediments are dragged by the
motion of the water, diffuse because of turbulence and fall due to
their weight. Since sediment concentration decays rapidly away
from the bottom and c is signicant only within the boundary layer,
the sediment balance provides the following equation




@c
H @c 1 @
@c
vs

DT
@t
d @x2 2 @x2
@x2

28

where the dimensionless sediment diffusivity DT can be assumed


to
p


be equal to the dimensionless eddy viscosity and v s v s =H g  h
is the dimensionless sediment fall velocity which is assumed to

408

G. Vittori, P. Blondeaux / Progress in Oceanography 120 (2014) 399409

depend on the local value of the sediment concentration and on the


sediment Reynolds number

90
80
70

29

60

Qb,Qs

Rp

q
 3
qs =q  1g  d

100

As already pointed out, c is assumed to vanish for x2 tending to


innity, and c is set equal to cref for x2 = xref, where cref is provided
by an empirical relationship and xref is set equal to twice the size
of the sediment grains. Several formulae can be used to quantify cref.
In this present paper, we use Zysermann and Fredsoes formula
(Zyserman and Fredsoe, 1994).

cref 0:331

cm h  hc 1:75

where the maximum value cm of the concentration is assumed to be


0.32. Then, the ux of sediment in the direction of wave propagation, from the reference level up to the irrotational region, can be
computed by means of the following relationship

Q s
d
q
Q s q
 3

 3



qs =q  1g d
s  1d =h0

30
20
10
0
-10
-10

-8

-6

-4

-2

ucdx2

31

~x2;ref

10

Fig. 13. Bed load (continunous line) and suspended load (broken line) plotted


versus the wave phase f (d = 0.0001,  = 0.5, zr = 3, Rp = 240 and d =h0 0:00015).

Finally, the volume of sand per unit width transported by a solitary wave can be easily estimated starting from the knowledge of
Qb(t) and Qs(t),

where the contribution due to the horizontal diffusion is neglected.


In Figs. 12 and 13 the time development of the sediment transport rate is plotted versus time for different values of the parameters, distinguishing the contributions due to the bed load (Qb) and
the suspended load (Qs). In Fig. 12 the sea bottom is made up of
ne sand (d = 0.2 mm) and the suspended load prevails on the
bed load which is practically negligible. The sediment transport
rate is characterized by the presence of a rst maximum which is
induced by the passage of the wave crest but slightly shifted. Then,
a second relative maximum of the sediment transport rate is observed, which is induced by turbulence which appears explosively
during the decelerating phase and picks up a lot of sediments from
the bottom and put them into suspension. Finally, during the later
stages, the sediment transport reverses its direction and is directed
towards the offshore region because the ow in the proximity of
the bottom reverses its direction. Similar results are obtained for
different values of the parameters even if the variation of sediment
characteristics gives rise to different values of the ratio between
the bed load and the suspended load. Indeed Fig. 13, where the values of Qb and Qs are plotted as function of f for the same values of
the parameters considered in Fig. 12 but for d = 2 mm, shows that
the bed load, which is negligible for very ne sand (see Fig. 12), becomes signicant for coarse sand.

V
h

2

q Z

 3
s  1d =h0

Q b t Q s tdt

32

1

4. Conclusions
The transition from the laminar regime to the turbulent regime
within the smooth boundary layer at the bottom of a solitary wave
is investigated using both a linear stability analysis and direct
numerical simulations of the NavierStokes equations.
The results show that the laminar ow turns out to be unstable
when the wave height is larger than a critical value which depends
on the ratio between the thickness of the bottom boundary layer
and the local water depth. Close to the critical conditions, the
growth of the unstable components of the perturbations leads to
the formation of two-dimensional vortex structures (vortex tubes)
with their axes parallel to the bottom and orthogonal to the direction of wave propagation. The theoretical and numerical results
qualitatively agree with the experimental observations of Sumer
et al. (2010). Moreover, the quantitative differences between the
predictions and the experimental measurements can be explained
on the basis of physical arguments.
From a practical point of view, it can be assumed that:
(1) the laminar regime is unstable for values of the Reynolds
3=2
p 2 larger than 2  105,
number Re 4H
3d

(2) the two-dimensional vortex tubes, which are generated by


the growth the most unstable two-dimensional perturbations, are characterized by a relative distance of order
q
p

30d 30 2m h = g  h ,

1400
1200
1000
800

Qb,Qs

40

30

cm 0:331h  hc 1:75

600
400
200
0
-200
-10

50

-8

-6

-4

-2

10

Fig. 12. Bed load (continunous line) and suspended load (broken line) plotted


versus the wave phase f (d = 0.0001,  = 0.5, zr = 0.3, Rp = 7.5 and d =h0 0:000015).

(3) these two-dimensional vortex structures break and give rise


to a fully turbulent ow, as soon as the value of Re is larger
than 5  105,
(4) close to the critical condition, both vortex tubes and turbulence tend to appear during the decelerating phase, i.e.
behind the wave crest. If the wave height (or equivalently
the Reyolds number) is further increased, turbulence
appears earlier and becomes more intense. Hence, for eld
conditions, it is likely that turbulence dominates a large part
of the wave cycle.
Moreover, the sea bed is usually rough. In order to gain some
information on the transition limits and the characteristics
of the turbulent ow in the rough bottom case, the phenomenon has been investigated using Reynolds averaged

G. Vittori, P. Blondeaux / Progress in Oceanography 120 (2014) 399409

NavierStokes equations and the two-equation turbulence


model of Saffman (1970), which seems to be able to
described the gross features of the transition process in
unsteady boundary layers.
(5) As expected, the bottom roughness destabilizes the boundary layer and turbulence appears for smaller values of the
Reynolds number.
The use of RANS equations and a turbulence model allows a
detailed investigation of the phenomenon in a wide range of
the parameters and the evaluation of the sediment transport
rate. The results obtained show that, for values of the parameters describing eld cases, turbulence picks up a lot of sediment from the bed and carry them into suspesion making
the suspended load much larger than the bed load.
(6) Even though the ow and the bottom shear stress reverse
their direction during the late decelerating phases, the backward ow is weak and the net (time averaged) sediment
transport points always in the direction of wave
propagation.

Acknowledgements
The paper was partially funded by the Ministero dellIstruzione,
dellUniversit e della Ricerca in the framework of the research
Project No. 2008YNPNT9-003 Idrodinamica e morfodinamica nella
regione dei frangenti and by the University of Genoa through the
contract Trasporto solido generato da onde di mare. The authors
are grateful to prof. Sumer (Technical University of Denmark,
DTU Mekanik, Section for Fluid Mechanics, Coastal and Maritime
Engineering) who kindly provided the image of Fig. 2.
References
Blondeaux, P., 1987. Turbulent boundary layer at the bottom of gravity waves.
Journal of Hydraulic Research 25 (4), 447464.
Blondeaux, P., Colombini, M., 1985. Pulsatile turbulent pipe ow. In: V Symposium
on Turbulent Shear Flows, Ithaca, August 79.
Blondeaux, P., Seminara, G., 1979. Transizione incipiente al fondo di unonda di
gravit. Accademia Nazionale dei Lincei 67, 408411, Italian.
Blondeaux, P., Pralits, J., Vittori, G., 2012. Transition to turbulence at the bottom of a
solitary wave. Journal of Fluid Mechanics 709, 396407.
Blondeaux, P., Vittori, G., 1994. Wall imperfections as a triggering mechanism for
Stokes layer transition. Journal of Fluid Mechanics 264, 107135.
Blondeaux, P., Vittori, G., 1999. Boundary layer and sediment dynamics under sea
waves. Advances in Coastal and Ocean Engineering 4, 133190.
Blondeaux, P., Vittori, G., 2012. RANS modelling of the turbulent boundary layer
under a solitary wave. Coastal Engineering 60, 110.
Bogucki, D.J., Redekopp, L.G., 1999. A mechanism for sediment resuspension by
internal solitary waves. Geophysical Research Letters 26 (9), 13171320.
Carstensen, S., Sumer, B.M., Fredsoe, J., 2010. Coherent structures in wave boundary
layers. Part 1. Oscillatory motion. Journal of Fluid Mechanics 646, 169206.
Cavallaro, L., Scandura, P., Foti, E., 2011. Turbulence-induced steady streaming in an
oscillatory boundary layer: on the reliability of turbulence closure models.
Coastal Engineering 58, 290304.
Conrad, P.W., Criminale, W.O., 1965. The stability of time-dependent laminar ows.
Zeitschrift fr Angewandte Mathematik und Physik 16, 233254.
Costamagna, P., Vittori, G., Blondeaux, P., 2003. Coherent structures in oscillatory
boundary layers. Journal of Fluid Mechanics 474, 133.
Diamessis, P.J., Redekopp, L.G., 2006. Numerical investigation of solitary internal
wave-induced global instability in shallow water benthic boundary layers.
Journal of Physical Oceanography 36, 784812.
Djordjevic, V.D., Redekopp, L.G., 1988. Linear stability analysis of nonhomentropic,
inviscid compressible ows. Physics of Fluids 31, 32393245.
Foti, E., Scandura, P., 2004. A low Reynolds number k- model validated for ows
over smooth and rough wall. Coastal Engineering 51, 173184.

409

Fredsoe, J., Deigaard, R., 1992. Mechanics of Coastal Sediment Transport. World
Scientic.
Garret, C., Munk, 1979. Internal waves in the ocean. Annual Review of Fluid
Mechanics 11, 339369.
Grimshaw, R.H.J., 1970. The solitary wave in water of variable depth. Journal of
Fluid Mechanics 42 (3), 639656.
Grimshaw, R.H.J., 1971. The solitary wave in water of variable depth Part 2. Journal
of Fluid Mechanics 46 (3), 611622.
Grimshaw, R.H.J., Ostrovsky, L.A., Shrira, V.I., Stepanyants, Y.A., 1998. Nonlinear
surface and internal gravity waves in a rotating ocean. Surveys in Geophysics
19, 289338.
Hammond, D.A., Redekopp, L.G., 1998. Local and global instability properties of
separation bubbles. European Journal of Mechanics B/Fluids 17 (2), 145164.
Helfrich, K.R., Melville, W.K., 2006. Long nonlinear internal waves. Annual Review of
Fluid Mechanics 38, 395425.
Keulegan, G.H., 1948. Gradual damping of solitary waves. RP1895, 40, U.S.
Department of Commerce, National Bureau of Standards, 487498.
Kim, J., Moin, P., 1985. Application of a fractional-step method to incompressible
NavierStokes equations. Journal of Computational Physics 59, 308323.
Lin, Y., Redekopp, L.G., 2011. The wave-induced boundary layer under long internal
waves. Ocean Dynamics 61, 10451065.
Liu, P.L.F., 2006. Turbulent boundary layer effects on transient wave propagation in
shallow water. Proceedings of the Royal Society of London 462 (2075), 3481
3491.
Liu, P.L.-F., Orla, A., 2004. Viscous effects on transient long-wave propagation.
Journal of Fluid Mechanics 520, 8392.
Liu, P.L.F., Park, Y.S., Cowen, E.A., 2007. Boundary layer ow and bed shear stress
under a solitary wave. Journal of Fluid Mechanics 574, 449463.
Mei, C.C., 1989. The Applied Dynamics of Ocean Surface Waves. Advanced Series on
Ocean Engineering, vol. 1. World Scientic.
Menter, F.R., 1994. Two-equation eddy-viscosity turbulence models for engineering
applications. AIAA Journal 32 (8), 15981605.
Miles, 1980. Solitary waves. Annual Review of Fluid Mechanics 12, 1143.
Munk, W.H., 1949. The Solitary Wave Theory and its Applications to Surf Problems.
New York Academy, Science Annals.
Ostrovsky, L.A., Stepanyants, Y.A., 1989. Do internal solitions exist in the ocean?
Reviews of Geophysics 27, 293310.
Rai, M.M., Moin, P., 1991. Direct simulations of turbulent ow using nitedifference schemes. Journal of Computational Physics 96, 15.
Saffman, P.G., 1970. A model for inhomogeneous turbulent ow. Proceedings of
Royal Society of London. Series A: Mathematical and Physical Sciences 317
(1530), 417433.
Saffman, P.G., Wilcox, D.C., 1974. Turbulence-model predictions for turbulent
boundary layers. AIAA Journal 12, 541546.
Shen, S.F., 1961. Some considerations on the laminar stability of incompressible
time-dependent basic ows. Journal of Aerospace Science 28, 397404, and 417.
Staquet, C., Sommeria, J., 2002. Internal gravity waves: from instabilities to
turbulence. Annual Review of Fluid Mechanics 34, 559593.
Stastna, M., Lamb, K.G., 2002. Vortex shedding and sediment resuspension
associated with the interaction of an internal solitary wave and the bottom
boundary layer. Geophysical Research Letters 29 (11), 1512, 7-1-3.
Sumer, B.M., Jensen, P.M., Soerensen, L.B., Fredsoe, J., Liu, P.L.F., 2010. Coherent
structures in wave boundary layers. Part 2. Solitary motion. Journal of Fluid
Mechanics 646, 207231.
Suntoyo, Tanaka, H., 2009. Numerical modeling of boundary layer ows for a
solitary wave. Journal of Hydro-environment Research 3, 129137.
Tanaka, H., Winarta, B., Suntoyo, Yamaji, H., 2011. Validation of a new generation
system for bottom boundary layer beneath solitary wave. Coastal Engineering
59, 4656.
U.S.A.C.E.; Coastal engineering Res. Center, 1984. Shore Protection Manual.
Vicksburg, Miss., Department of Army, Waterways, Experiment Station, Corps
of Engineers, Coastal Engineering Research Center, Washington DC.
Verzicco, R., Vittori, G., 1996. Direct simulation of transition in Stokes layers. Physics
of Fluids 8 (6), 13411343.
Vittori, G., 2003. Sediment suspension due to waves. Journal of Geophysical
Research 108 (C6), 3173, 4.14.7.
Vittori, G., Blondeaux, P., 2008. Turbulent boundary layer under a solitary wave.
Journal of Fluid Mechanics 615, 433443.
Vittori, G., Blondeaux, P., 2011. Characteristics of the boundary layer at the bottom
of a solitary wave. Coastal Engineering 58 (2), 206213.
Vittori, G., Verzicco, R., 1998. Direct simulation of transition in an oscillatory
boundary layer. Journal of Fluid Mechanics 371, 207232.
Wang, B., Redekopp, L.G., 2001. Long internal waves in shear ows: topographic
resonance and wave-induced global instability. Dynamics of Atmospheres and
Oceans 333, 263302.
Zyserman, J.A., Fredsoe, J., 1994. Data analysis of bed concentration of suspended
sediment. Hydraulic Engineering 120 (9), 10211042.

Anda mungkin juga menyukai