Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Application of Boundary Element

Anatoliy Khait1
Method for Determination of the
Mem. ASME
School of Mechanical Engineering,
Faculty of Engineering,
Wavemaker Driving Signal
Tel Aviv University, A method for the generation of steep nonlinear broad-banded wave trains having an arbi-
Tel Aviv 6997801, Israel trary prescribed shape is developed. It is shown that the second-order contributions to the
e-mail: haitanatoliy@gmail.com velocity field are negligible in deep water, while the second-order bound components of the
surface elevation are significant. This fact allows improvement of an iterative method of
Lev Shemer the wavemaker driving signal adjustment that increases the accuracy of excitation of
Mem. ASME wave train with the prescribed free waves’ spectrum. The decomposition of the complex
School of Mechanical Engineering, amplitude spectrum of the surface elevation into free and bound components is based on
Faculty of Engineering, the approach adopted in the derivation of the Zakharov model. The iterative adjustment
Tel Aviv University, of the driving signal is carried out using the numerical wave tank based on the boundary
Tel Aviv 6997801, Israel element method. It is demonstrated that accurate wave train excitation is attained for dif-
e-mail: shemer@eng.tau.ac.il ferent values of the wave steepness. The method allows decreasing the number of iterations
needed for the driving signal adjustment. The surface elevation values measured in the lab-
oratory wave tank agree closely with those obtained in the numerical simulations. The mea-
sured and the simulated frequency spectra are in agreement as well.
[DOI: 10.1115/1.4042942]

Introduction second-order in the wave steepness theory may be attributed to the


fact that it neglects higher-order contributions. A corrected transfer
Mechanically driven wavemakers are often applied to excite
function based on experimental results was offered in Ref. [5].
either deterministic or random wave sequences in laboratory
Houtani et al. [6,7] adopted the higher-order spectral method
conditions. The linear transfer function relating the oscillatory
(HOSM) to upgrade the results of Ref. [2] for directional wave exci-
wavemaker motion at a constant frequency to the resulting mono-
tation. The HOSM-WG (here WG stands for the wave generation)
chromatic surface elevation in the tank is given in Ref. [1]. The the-
method developed in Ref. [6,7] was shown to be superior to the con-
oretical analysis predicts that in addition to the designed sinusoidal
ventional wave generation when wave steepness increases, the fre-
wave, undesirable (the so-called evanescent) modes are excited in
quency spectrum broadens or the directional spreading becomes
the tank. These modes that result from inconsistency of the vertical
large. Particular attention was paid to the generation of the freak waves.
distribution of the horizontal component of the wavemaker motion
An alternative approach proposed in Refs. [3–5] is restricted to an
with that of the orbital wave velocity are indeed observed in exper-
accurate excitation of single-frequency waves. The approach is
iments; these waves decay fast with the distance from the wave-
based on the utilization of the force-feedback controller system
maker. Since the evanescent modes do not affect the wave field at
that employs the impedance matching procedure. In contrast to
large distances, they are often disregarded. Experiments, however,
Ref. [2], which applies theoretically calculated second-order
demonstrate that additional propagating modes, in particular at
driving signal correction, in Refs. [3–5] forces acting on the wave
the second harmonic of the dominant frequency, are observed.
board surface by force transducers are measured and serve as feed-
These undesirable propagating waves contaminate the designed
back to the computer-aided wavemaker system. An additional
wave field over much longer distances.
second-order correction was also implemented. It was found out that
Analytical second-order wavemaker theory for excitation of
wavemakers driven by force-controlled procedures, as opposed to
multiple harmonics in deep and intermediate-depth water was pro-
the displacement control, significantly reduce the content of the
posed by Schäffer [2]. He also suggested a second-order correction
second-order spurious waves. The method presented in Refs.
to the wavemaker motion in order to suppress the spurious propagat-
[3,4] also has an additional significant advantage due to its ability
ing waves. This theory was generally confirmed experimentally for
to simultaneously generate and absorb any unwanted waves, includ-
regular waves, wave groups, and irregular wave trains. Nevertheless,
ing the reflected ones, by a single mechanical device. On the other
certain deviations of the measured surface elevation from the theoret-
hand, it is limited to simple nearly monochromatic signals and
ical predictions based on the Schäffer theory [2] were observed,
requires high-precision force transducers to measure the wave
indicating that spurious waves are not completely removed. Spinne-
board torque, thus significantly complicating the experimental setup.
ken and Swan [3,4] and Aknin and Spinneken [5] re-examined
The need to generate accurately broad-banded wave trains
the theory of Schäffer for a wide range of excitation condi-
prompted the development of empirical methods for wavemaker
tions: shallow-to-intermediate, intermediate, and deepwater-wave
motion adjustment. To this end, an experimental technique to gen-
regimes, with wave steepness in the range of ε = 0.05–0.25 and for
erate the wave packet focusing at the prescribed location in the
either flap or piston wavemakers. Differences between the theoreti-
wave tank was applied in Ref. [8]. Adjustment of phases at the
cal predictions and experiments were found to exist for wave ampli-
wave board yielded satisfactory convergence of the wave energy
tudes as well as for phases of the spurious waves. These differences
at the specified focus point. In contrast, a numerical wave tank
persist for all experimental conditions. The limited accuracy of the
(NWT) based on the boundary element method (BEM) was used
in Ref. [9] to iteratively adjust the phases of the wave components
of the wave packet yielding the required wavemaker motion. The
1
Corresponding author. phase shifting was checked by a numerical wave probe and
Contributed by the Ocean, Offshore, and Arctic Engineering Division of ASME for
publication in the JOURNAL OF OFFSHORE MECHANICS AND ARCTIC ENGINEERING. Manu-
further verified experimentally.
script received August 31, 2018; final manuscript received February 3, 2019; published Shemer et al. [10,11] applied an experimentally determined
online March 20, 2019. Assoc. Editor: Felice Arena. wavemaker transfer function in order to determine amplitudes and

Journal of Offshore Mechanics and Arctic Engineering DECEMBER 2019, Vol. 141 / 061102-1
Copyright © 2019 by ASME
phases of the spectral components of the wavemaker driving signal In laboratory facilities, waves are usually generated by the hori-
for generation of a propagating wave train with a prescribed shape. zontal motion of the wave board that prescribes the vertical distribu-
The spatial version of the Zakharov equation was used as a theoret- tion of the horizontal water velocity component. In linear
ical model for computation of spectral changes during the wave approximation, this condition allows one to determine the transfer
train propagation along the experimental facility. Schmittner et al. function c (1) [1]:
[12] utilized the phase–amplitude iteration scheme to experimen-
tally synthesize extreme deterministic wave sequences. They indi- ∂ψ (1) ∂S(1)
u(1) (x = 0, z = 0, t) ≡ − (x = 0, z = 0, t) = c(1) (z = 0, t) (4)
cate that this scheme is applicable even without a priorily ∂x ∂t
knowing the transfer functions of the wave generator. However,
Here, S(z,t) is the horizontal wavemaker coordinate, u is the hor-
their study demonstrates a noticeable discrepancy between the the-
izontal velocity of the fluid, and x = 0 corresponds to the mean
oretical and the measured surface elevations. Buldakov et al. [13]
wavemaker location. For weakly nonlinear waves, the horizontal
improved the iterative wave generation method by application of
velocity can be expanded around the small parameter ε resulting
the harmonics separation technique for linearization of the ampli-
in the second-order boundary condition at the wavemaker. For
tude spectrum. They showed that experimental adjustment of the
piston-type wavemaker, it reads as
linear part of the spectrum only considerably improves the genera-
  2 (1) (1)
tion accuracy. The method requires four experimental runs (with ∂ψ ∂ψ (1)  ∂ψ (2)  ∂ ψ ∂ψ
different shifts in phases of the driving signal harmonics) at every = + − dt (5)
∂x ∂x x=0 ∂x x=0 ∂x2 ∂x
iteration to complete the separation of harmonics, thus resulting  
 
 

x=0
in a significant extension of the duration of experiments, especially u(1) u(2) u(2)
d
in large experimental facilities.
Existing theoretical models of nonlinear waves based on Schrö- Although Eq. (5) is valid for any vertical coordinate z, we further
dinger, Dysthe, Zakharov equations, etc. [14–18] describe evolution consider only the mean surface level z = 0. In Eq. (5), the linear
of free water gravity waves only, i.e., waves that satisfy the deep or velocity u (1) estimated according to Eq. (4) is supplemented with
intermediate-depth dispersion relation between the wave numbers k two additional second-order terms. The term u(2) d is caused by
and the angular frequencies ω: finite displacement of the wave board, while u (2) is associated
with the second-order bound waves η (2). In contrast to propagating
ω2 = gk tanh (kh) (1) waves η (1), bound waves are determined by free waves and do not
Due to nonlinearity, the free waves induce the corresponding satisfy the dispersion relation (1). Substitution of ψ (1) and ψ (2) into
higher-order bound waves that are totally defined by the carrier Eq. (5) yields the following:
free waves and cannot be controlled separately. ka
This paper is aimed at the improvement of the accuracy of the |û(1) |(ka, kh) = √
steep broad-banded wave trains generation in deepwater conditions. tanh (kh)
√
We develop a fully nonlinear method to adjust the driving signal of 3 cosh (2kh) tanh (kh)
the wavemaker to generate wave trains with the prescribed broad |û(2) |(ka, kh) = (ka)2 (6)
4 sinh4 (kh)
free waves’ spectrum. An additional benefit of this method is the
improvement of the convergence rate of the iteration process (ka)2
since it distinguishes harmonics according to their order of nonlin- d |(ka, kh) = 
|û(2)
2 tanh3 (kh)
earity. To improve the accuracy of the iteration procedure, measure-
ments of the surface elevation variation should be carried out in the The relative second-order contributions to the horizontal velocity
vicinity of the wavemaker. amplitudes at the free surface are plotted in Fig. 1 as a function
of the steepness ka and water depth kh. Deepwater conditions
Wave Generation Method exist when kh > π (tanh (kh) → 1), whereas the range π/10 < kh < π
(0.3 < tanh (kh) < 1) corresponds to the intermediate depth [1]. For
Nonlinear Wave Generation. Assumption of weak nonlinear- steep waves, the surface elevation amplitudes of the bound wave
ity within potential wave theory [1] allows perturbation expansion |η̂(2) | may become essential compared with |η̂(1) | even in deep
of the surface elevation η and potential ψ. Taking the wave steep- water. However, the amplitude of the second-order horizontal
ness ε = ka as a small parameter, k and a being the wave number
velocity |û(2) | vanishes with an increase of kh and in the limiting
and the surface elevation amplitude, the second-order expansion
case of deep water it is lower by few orders than the velocity ampli-
can be written as follows:
tude of the propagating wave |û(1) |. Therefore, it can be assumed

η(x, t) = η(1) (x, t) + η(2) (x, t) that the second-order velocity field can be neglected in the
(2) process of generation of deepwater waves, while the surface eleva-
ψ(x, z, t) = ψ (1) (x, z, t) + ψ (2) (x, z, t)
tion variations due to bound waves should be accounted for.
where x defines the propagation direction of unidirectional waves, z Figure 2 exhibits the velocity fields at different orders for the
is directed upward, and t is the time. The dimensionless amplitudes
 regular wave with ka = 0.3 and kh = π. It is seen that the
are defined as η̂ = ηA k and ψ̂ = ψ A k 3 /g, where ηA and ψA are the second-order velocities remain well below those of the first-order
corresponding dimensional values. For regular waves, the absolute even in the vicinity of the steep crest, while the shape of the
values of those amplitudes are given as follows [1]: surface elevation is noticeably asymmetric with respect to z = 0.
The difference in the behavior of η (2) and u (2) in Fig. 1 can be
|η̂(1) |(ka, kh) = ka related to the vertical velocity w of the surface elevation at a
fixed location x:
(ka)2 cosh (kh)[2 + cosh (2kh)]   
|η̂(2) |(ka, kh) = ∂η ∂ψ (1)  ∂ψ (2) 
4 sinh3 (kh) = − −
(3) ∂t x=0 ∂z x=0,z=0 ∂z x=0,z=0
ka
|ψ̂ (1) |(ka, kh)|z=0 = √  

 

tanh (kh) w(1) w(2)


√ 2 (1)
(7)
3 cosh (2kh) tanh (kh) ∂η ∂ψ(1) (1)
(1) ∂ ψ
|ψ̂ (2) |(ka, kh)|z=0 = (ka)2 + −η
8 sinh4 (kh) ∂x ∂x ∂z2 x=0,z=0
 

where kh is the dimensionless water depth. w(2)


st

061102-2 / Vol. 141, DECEMBER 2019 Transactions of the ASME


Fig. 2 Normalized velocity fields of (a) the propagating and
(b) second-order components of a regular wave at kh = π and
ka = 0.3

Fig. 1 Relative contribution of (a) the second-order bound


waves to the surface elevation and (b) the second-order horizon-
tal velocity component, as a function of wave steepness and
water depth

Here, w (2) represents the second-order contribution to the verti-


cal velocity due to the instantaneous potential variation with z,
whereas w(2)st is caused by the motion of the free surface in the
Eulerian coordinate system. Substitution of (2) into (7) yields the
following:

|ŵ(1) |(ka, kh) = ka tanh (kh)
√
3 cosh (kh) tanh (kh)
|ŵ(2) |(ka, kh) = (ka)2
2 sinh3 (kh) (8)
(ka)2
st |(ka, kh) = √
|ŵ(2)
tanh (kh)
Figure 3 compares different contributions to the vertical velocity
of the free surface (8) for the regular wave with steepness ka = 0.3.
Both w (2) and wst(2) contribute significantly to the surface elevation
variation in an intermediate depth causing the appearance of the Fig. 3 Comparison of the contributions (8) to the vertical motion
bound waves η (2). In deep water, w (2) is negligibly small and of the free surface of regular wave with ka = 0.3
bound waves η (2) are stipulated by wst(2). This explains the differ-
ences in behavior of |η̂(2) | and |û(2) | in Fig. 1. Fig. 4. Within this approach, the surface elevation at xt is repre-
The unwanted spurious propagating waves [2] appear when sented in the spectral form containing free as well as bound compo-
both second-order terms u (2) and u(2) d in Eq. (5) are not taken into nents η(xt, t) = η (1) + η (2):
account in the wavemaker motion. Since u (2) is negligibly small  
in deep water, u(2) 
d is the only essential source of inaccuracies at
η(x, t) = Re aj e i[kj (x−xt )−ωj t]
(9)
the second-order in this case. As seen from Eq. (5), the expression
j
d incorporates only the first-order terms. The wavemaker
for u(2)
motion can therefore be adjusted iteratively so that it satisfies the where kj and ωj are the wave number and the angular frequency of
nonlinear boundary condition (5) in deep water. the jth component, respectively.
The wave generation method consists of three steps. First, the
free complex amplitude spectrum of the target wave train is deter-
Wave Generation in Deep Water. The suggested iterative mined and the surface elevation at the wavemaker location x = 0
generation method for a broad-banded wave train is most suitable is calculated. The linear transfer function c (1) (4) is then applied
for deepwater conditions. The iterative adjustment of the wave- to convert the temporal variation of the surface elevation into the
maker motion was performed with respect to the surface elevation required motion of the wave board. Next, the wave train is
variation with time recorded at the virtual probe location xt, see excited either experimentally or numerically employing a fully

Journal of Offshore Mechanics and Arctic Engineering DECEMBER 2019, Vol. 141 / 061102-3
Fig. 4 Schematic of the NWT

nonlinear numerical wave tank using the given wavemaker motion, spectral components are identical and equal to zero, thus at the pre-
and the full frequency spectrum of the surface elevation is deter- scribed focusing location, the extreme wave crest appears with the
mined at xt. height equal to the sum of amplitudes of all free harmonics.
As shown earlier, for wave generation in deep water, the Since deepwater and intermediate-depth waves are dispersive,
second-order velocity field u (2) is vanishingly small and may be phase velocities cj = ωj/kj differ for different harmonics, resulting
neglected. Decomposition into free and bound waves according to in wave train spreading away from the focusing location with the
the order of nonlinearity based on the Zakharov theory [15–18] is corresponding reduction of the wave steepness [10,11].
utilized here. The leading order free components that satisfy the dis- The accuracy of the wave generation should be controlled as
persion relation (1) are separated from the contribution of higher- close to the wavemaker as possible in order to avoid spectral
order bound waves by an iterative procedure [10,11] based on the changes due to nonlinear effects. However, the surface elevation
analytic relations between the first-order free and the second-order at short distances from the wavemaker (usually not exceeding
bound components of the surface elevation spectra (see the Appen- about 3 mean water depths h) is affected by evanescent modes
dix). The spectrum of the propagating first-order waves is recalcu- [1]. Therefore, the wave generation accuracy is monitored here at
lated to the location of the wavemaker x = 0 by shifting the phases the target location xt = 2 m = 3.3 h (see Fig. 4).
of each harmonic according to (9).
At the next step, a modification of an iterative scheme given in
Ref. [12] is utilized for adjustment of the wavemaker motion Experimental Facility. The experiments were carried out in a
aimed at improvement of agreement of the linear part of the 18 m long, 1.2 m wide wave tank filled to mean water depth of
surface elevation spectrum at x = 0 with the target one. For each h = 0.6 m. The tank is equipped with a paddle-type wavemaker
free harmonic ωj, the wavemaker amplitudes Swm(ωj) and phases hinged at the bottom. The wave-energy absorbing sloping beach
ϕwm(ωj) at iteration l + 1 are adjusted by correcting their values at that is 3 m long is located at the opposite end of the tank. Although
the previous iteration l by the difference between the target and the reflection from the far end of the tank is significantly reduced by
the measured values of the amplitude and the phase in the the beach, the experiments are planned so that all measurements are
complex surface elevation spectra: completed prior to the arrival of the wave train to the beach.
The experiments, including the wave generation and data record-
 free ing, are controlled by a single LABVIEW program ensuring precise
j,l+1 = S j,l + c {a j − a j,l }
Swm wm (1) tar
(10) synchronization. The actual instantaneous paddle inclinations are
free
j,l+1 = ϕ j,l + {ϕj − ϕ j,l }
ϕwm wm tar recorded as well. Resistance-type wave gauge can be moved
along the tank; in the present experiments, it is usually located at
where c (1) is the linear transfer function and atar and ϕtar are ampli- xt = 2 m (see Fig. 4). The wave gauge was statically calibrated
tudes and phases at the target free wave spectrum, respectively. The prior to every experimental run using a fully automatic procedure.
iterations continue until convergence is attained. The sampling interval was set at T0/128, where T0 is the carrier
wave period.
Parameters of the wave train (11) are as follows: T0 = 0.7 s,
m = 1.2, and carrier wave steepness ε0 = k0η0 = 0.25.
Formulation of the Problem
We investigate here the excitation of a unidirectional surface
gravity wave train with a prescribed spectrum. A Gaussian-shaped Numerical Wave Tank. The iterative adjustment of the wave-
free-wave envelope of the surface elevation at the prescribed loca- maker motion is performed in the fully nonlinear two-dimensional
tion x = xf within the tank is considered as a generic shape of a wave NWT based on transient BEM. This allows reduction of the exper-
train with no restrictions on the spectral width: imental effort for the determination of the wavemaker driving
signal.
    The numerical domain Ω is bounded by the bottom Γb and lateral
t 2 ΓL boundaries, the wavemaker Γwm, and the free surface Γfs, as
η(t) = η0 exp − cos (ω0 t) (11)
mT0 shown in Fig. 4. Every boundary is defined by several nodes. The
nodes on Γb and ΓL are fixed and nodes on Γwm follow the
Here, η0 is the envelope amplitude, T0 and ω0 = 2π/T0 are the motion of the wavemaker, while the nodes on Γfs move freely deter-
carrier wave period and angular frequency, respectively, and the mining the instantaneous shape of the free surface. Every boundary
parameter m can be adjusted to generate wave train with any contains end nodes that coincide geometrically with the end nodes
desired spectral width. The linear amplitude spectrum aj = a(ωj) of the neighbor boundaries. These corner nodes are marked in Fig. 4
of the wave train also has a Gaussian shape. The phases of all and treated in a special manner [19,20].

061102-4 / Vol. 141, DECEMBER 2019 Transactions of the ASME


Potential and irrotational fluid flow is governed by the Laplace
equation with appropriate boundary conditions: impenetrability
conditions for Γb and ΓL, the known linear time-dependent veloc-
ity profile for Γwm, and the kinematic and the dynamic boundary
conditions at the free surface Γfs [19–21]. Green’s identity is
applied to transform the Laplace equation into a boundary integral
equation:
 
∂ψ ∂ξ
α(rs )ψ(rs ) = (r)ξ(r, rs ) − ψ(r) (r, rs ) dΓ (12)
∂n ∂n
Γ Fig. 5 Assessment of effectiveness of the numerical wave
where r(x,z) is the radius vector in the coordinate system x–z absorption techniques
(Fig. 4), ξ(r, rs) = −1/(2π) ln(|r − rs|) is the fundamental solution
that represents the potential flow at point r due to a source
located at rs, α is a geometric coefficient, α = π for regular nodes where μd, pd, and bε are parameters; and xorig and xend are coordi-
and α ≈ π/2 for corner nodes, and Γ = Γb ∪ ΓL ∪ Γwm ∪ Γ fs . For a nates of both ends of the numerical absorbing beach (see Fig. 4).
source located at the node p, the equation determines the value of Two approaches were considered: pressure action opposing the
the potential ψp for a known normal gradient (∂ψ/∂n)p and vice free surface motion (term 1 in Eq. (14)) [22] and an artificial viscos-
versa. Setting the position of the source at different grid nodes ity (term 2) [23]. Beach function bf(x) defines the rate of increase of
leads to a system of linear equations. both terms along the beach; bf (x ≤ xorig) = 0:
Equation (12) is discretized by sliding cubic 4-node boundary   
1 x − xorig
elements as shown schematically in Fig. 4. The central part of bf (x) = 1 + tanh bε (15)
every single boundary element is used for approximation of the 2 xend − xorig
contour Γ, whereas the outer parts define the curvature of the It was found that the absorption efficiency is mostly affected by
element. The full set of the boundary elements constitutes a contin- the dimensionless parameters μd and pd. Based on numerical simu-
uous closed curve Γ. Local curvilinear coordinate system s–n lations performed to determine the optimal values for these param-
(Fig. 4) was introduced to integrate Eq. (12) numerically by the eters, the value of bε = 6 was selected. Numerical simulation results
Gaussian quadrature method. A more detailed description of the for the 16 m long wave tank with no absorbing beach were taken as
BEM can be found in Refs. [19–21]. a basis for estimation of the beach performance; the root-mean-
The solution depends on the spatial resolution of the grid. The square (rms) deviation of the surface elevation profile from the
convergence study disclosed that grid density at the free surface basis served as a measure for the waves damping effectiveness.
of at least 40 nodes/m is required to obtain a grid-independent solu- Figure 5 compares the beach effectiveness for different model
tion. Taking into account the width of the Gaussian spectrum and parameters. The most effective absorption occurs at pd = 2; this
the Nyquist condition for the highest frequencies, the grid density value was adopted for subsequent simulations. The introduction
at the free surface was taken to be 60 nodes/m. The number of of artificial viscosity appeared to be less effective and thus was
grid nodes for different boundaries was thus set as follows: Nfs = not implemented. Note that the values of μd exceeding 2000 Pa · s
360, Nb = 210, NL = 37, and Nwm = 37. cause numerical instability.
The radius vector of free surface nodes and their potentials were
expanded into a truncated second-order Taylor series; the time inte-
gration of the free surface boundary conditions was performed as Results and Discussion
suggested in Ref. [21]. Quadratic five-node sliding element was
adopted to estimate the expansion terms. The value of the integra- Excitation of Steep Waves (εc = 0.21). Generation of initially
tion time step was taken as T0/1024 exceeding the Courant steep highly nonlinear waves is now investigated. It is expected
number requirement for explicit schemes CFL ≤ 0.1. The Courant that strong nonlinear effects at the wavemaker may complicate
number is defined as follows [20]: the generation process. High initial steepness is attained by linear
focusing of the Gaussian-shaped wave train (11) at xf = 4 m. The
 Δt surface elevation at x = 0 was calculated using the linear theory.
CFL = gh (13)
(Δr)min The wave steepness at x = 0 is defined as εc ≡ k0Ax=0 = 0.21, Ax=0
being the maximum of the free wave envelope at the wavemaker
where g is the acceleration due to gravity, Δt is the time step, and location. The corresponding linear wavemaker motion was then
(Δr)min is the minimal distance between adjacent nodes. applied as the boundary condition for NWT.
Free surface elevation is recorded at xt applying cubic interpola- The surface elevation spectrum estimated from the data recorded
tion between neighboring nodes. The configuration of the domain by the virtual probe of NWT was decomposed into free and
Ω is similar to the physical wave tank (h = 0.6 m) except for its second-order bound components as shown in Fig. 6(a). Although
length since there is no need to simulate full length of the tank bound components have peak amplitudes that are lower by an
for determination of the surface elevation at xt = 2 m. The order of magnitude than the free ones, their contribution to the
domain length was thus reduced to LNWT = 6 m. The numerical surface elevation is essential and cannot be ignored, see Fig. 6(b).
absorbing beach is developed to decrease wave field contamination This figure also demonstrates that free modes exist in the frequency
by the reflected waves. The length of the beach was taken to be range of the second-order bound waves, whereas wave amplitudes
Lb = 3 m, see Fig. 4. in the target spectrum at those frequencies are negligibly small.
These waves are thus classified as spurious and should be elimi-
Numerical Absorbing Beach. The absorbing beach is built by nated during an iterative adjustment of the driving signal.
modification of the dynamic free surface boundary condition: The wavemaker motion adjustment procedure relies on Eq. (10).
The convergence of the free surface elevation to the target one
∂ψ 1 p μ ∂2 ψ is attained in 3–4 iterations, as seen from Table 1. A very rapid
= gz + |∇ψ|2 + bf (x) + 2 d 2 bf (x)
∂t 2 ρ ρ ∂s drop in the discrepancy during the first iteration results from fast
 
 
adjustment of the phases. Further improvement of the convergence
term 1 term 2 (14)
is mainly due to the modification of amplitudes. Subsequent itera-
 ∂ψ tions do not affect the error noticeably. The remaining error may
p = −pd ρ gh
∂n be attributed to the limited accuracy of the NWT due to its finite

Journal of Offshore Mechanics and Arctic Engineering DECEMBER 2019, Vol. 141 / 061102-5
Fig. 6 Separation of the spectrum into free and bound Fig. 7 Amplitude and phase adjustment for free components of
components the surface elevation

Table 1 Convergence of the adjustment procedure for steep


waves

 
2 0.5  
ηBEM − ηtarg ηtarg 
max
Iteration no.

1 26.3%
2 1.43%
3 1.05%
4 1.01%

spatial and temporal resolutions and the disregarded third- and


higher-order bound waves.
In Fig. 7(a), the simulated free amplitude spectra of surface ele-
adj
vation at the first and the final iterations (ainit
BEM and aBEM , respec-
tively) are plotted. The target spectrum at is also given for
comparison. The corresponding deviations of the initial and final Fig. 8 Adjustment of the numerically simulated and the actual
phases from the target are depicted in Fig. 7(b). The peak and the spectra of the wavemaker motion
low-frequency parts of the initial spectrum have lower amplitudes
than those of the target, whereas high-frequency modes are overval-
ued. Such an asymmetry signifies the importance of nonlinear The required correction of the driving signal is demonstrated in
effects accompanying waves’ excitation. The phase shift between Fig. 8. Here, Sinit
BEM represents the amplitude spectrum of the actual
the surface elevation and the wavemaker motion within the range paddle motion resulting from the initially computed driving
of meaningful modes is close to π/2 but not constant. A slight signal, whereas Sadj adj
BEM and ϕBEM correspond to the amplitudes and
increase is observed for higher frequencies, in agreement with phases of the paddle motion, respectively, at NWT after the final
Ref. [5]. adjustment of the required driving signal. The amplitudes and
For most frequencies, the adjusted amplitude spectrum agrees phases of the driving signal that causes the required paddle motion
drv and ϕdrv . This correc-
corr
well with the target one, Fig. 7(a). A slight deviation can be identi- in experiments are depicted in Fig. 8 as Scorr
fied at high frequencies where bound waves dominate. The phases tion of the driving signal results in the root-mean-square deviation of
of the frequency harmonics of the adjusted wave train for all signif- the actual motion from the designed one of less than 1%.
icant modes are also in agreement with the target, Fig. 7(b), except It is seen from Fig. 8(a) that the corrected amplitude spectrum
for relatively high frequencies. The observed disagreement with the of the driving signal Scorr
drv is shifted to higher frequencies as com-
target values of both amplitudes and phases can be attributed to the pared with the spectrum of the wavemaker motion. Thus, the trans-
finite spatial and temporal resolution of the NWT that imposes lim- fer function correction that relates the driving signal to the actual
itations on the accuracy of simulation of short (high-frequency) paddle motion is frequency-dependent. The phase difference
adj
waves. It may also be caused by the disregarded third- and higher- drv − ϕBEM is, however, almost constant within a wide frequency
ϕcorr
order bound waves. range remaining close to −π/2. It should be stressed that the devia-
For experimental verification of the simulated results, the wave- tion of the driving signal from the actual paddle motion is a feature
maker motion computed by NWT served as a driving signal. In of the particular wavemaker and, therefore, the importance of the
experiments, however, a significant deviation of the actual motion suggested driving signal correction may be wavemaker-dependent.
of the wavemaker paddle from the driving signal was observed. Upon completion of the driving signal correction procedure, the
The complex spectrum of the driving signal was therefore corrected surface elevation at the position of xt was measured, and the appro-
to improve agreement of the actual wavemaker motion with the pre- priate spectrum was calculated. The free waves were separated from
dicted one by the NWT. the bound components as described previously, and the results are

061102-6 / Vol. 141, DECEMBER 2019 Transactions of the ASME


Fig. 9 The measured spectrum of free waves after correction of
the wavemaker motion

Fig. 11 Separation of the surface elevation spectrum into free


plotted in Fig. 9. Although a small discrepancy for the high- and bound waves
frequency part remains, very good agreement is obtained at lower
frequencies. These results confirm the applicability of the suggested
method and NWT based on BEM for derivation of the driving Table 2 Convergence of the adjustment procedure for
signal for generation of unidirectional broad-banded gravity waves. medium-steep waves
Comparison of the simulated and the measured waveforms with 
 2 0.5  
the target is carried out in Fig. 10. The plotted experimental and ηtarg 
ηBEM − ηtarg
numerical data contain free as well as bound components. The Iteration no.
max

target linear waveform containing only free waves was therefore


augmented by corresponding bound components computed using 1 40.5%
expressions given in Refs. [16–18]. The experimental and numeri- 2 1.12%
cal waveforms agree well with the resulting modified target surface 3 0.67%
elevation variation with time. 4 0.60%
5 0.59%

Excitation of Waves With Lower Steepness (εc = 0.12). The


effect of the initial wave steepness on the accuracy of the generation
is now assessed. To this end, the focusing location of the Gaussian in Fig. 11(a) are notably lower than those in Fig. 6(a), as expected
wave train (11) was moved farther away from the wavemaker to for less steep waves. There is no overlapping of free and bound
xf = 8.5 m, while retaining all other governing parameters including waves in the spectral domain, thus enabling improved separation
the envelope amplitude η0. The linear dispersion causes spreading of harmonics.
of the wave train at the wavemaker with the corresponding reduc- The convergence of the iteration procedure required for the deter-
tion of the height of the steepest crest. The location of the wave mination of the driving signal in the NWT is illustrated in Table 2.
gauge and of the virtual probe remained unchanged. The accuracy is now nearly twice better than that in Table 1, but
As in the case of steep waves excitation for xf = 4 m, the surface four iterations are still needed to arrive at the optimal adjusted
elevation at the wavemaker was transformed into the wavemaker wavemaker motion. The accuracy of the waves’ generation thus
motion by application of the linear transfer function. The obtained strongly depends on the initial wave steepness, whereas the conver-
signal served at the initial iteration for the wavemaker boundary gence rate of the iterative method remains almost independent of ε.
condition in the NWT. The amplitude spectrum of the surface ele- Figure 12 demonstrates the variation of the free wave spectrum
vation derived from the virtual probe output and decomposed into during the iterative adjustment of the paddle motion. While the
free and bound modes is plotted in Fig. 11(a), while the wave peak and the low-frequency parts of the initial amplitude spectrum
forms are given in Fig. 11(b). The amplitudes of the bound waves are slightly underestimated, the high-frequency region is very close

Fig. 10 Comparison of experimental, numerical, and target wave forms

Journal of Offshore Mechanics and Arctic Engineering DECEMBER 2019, Vol. 141 / 061102-7
Fig. 15 Adjustment of (a) amplitudes and (b) phases in the
Fig. 12 Amplitude and phase adjustment for free components
spectra of the wavemaker motion needed for excitation of wave
utilizing NWT
trains with different steepnesses

attributed to the reduced contribution of spurious waves, especially


at high frequencies.
The correction of the driving signal is still required for this
lower steepness case. The procedure applied is identical to that
described in relation to Fig. 8. The measured spectrum of free com-
ponents and the corresponding temporal variation of the surface
elevation comprising all free and bound modes are compared with
the target in Figs. 13 and 14, respectively. The convergence of
the measured records with the target improved noticeably as com-
pared with the case of steeper waves at the wavemaker (cf., Figs.
9 and 10).
Figure 15 compares the correction of the wavemaker motion
obtained as a result of the iterative adjustment in the NWT for
wave trains of different steepness. The corrections to both ampli-
tudes and phases of the wavemaker motion are almost independent
of εc in the vicinity of the peak frequency fp, cf., Figs. 15(a) and
15(b). However, a more significant correction is required for a
steeper wave train (εc = 0.21) at frequencies closer to the free
Fig. 13 Experimental spectrum of free waves after correction of wave spectrum band denoted by vertical dashed lines in Fig. 15.
the wavemaker motion This observation agrees with the theory of Schäffer [2], where the
impact of steepness of the dominant free waves on the amplitudes
of the spurious propagating waves is reported.
to the target right from the beginning. This is the main difference
between excitation by the wavemaker of waves with different steep-
ness (cf., Figs. 7 and 12).
The initial phase difference in Fig. 12(b) is close to π/2 but varies Conclusions
slightly, similar to Fig. 7(b). The phases of the adjusted wave train Theoretical analysis of regular steep waves demonstrated that
are almost identical to those of the target within the whole range of the second-order contributions to the instantaneous velocity field
significant modes contrary to Fig. 7(b). This improvement can be are negligibly small in deep water. This simplifies the iterative

Fig. 14 Comparison of experimental, numerical, and target wave forms

061102-8 / Vol. 141, DECEMBER 2019 Transactions of the ASME


adjustment of the wavemaker driving signal for generation of unidi- η, ψ = surface elevation and potential
rectional broad-banded surface gravity waves with the prescribed ω, k = angular frequency and wave number
nonlinear spectrum of surface elevation. Expressions obtained in η̂, ψ̂, û, ŵ = dimensionless amplitudes of surface elevation,
the course of derivation of the Zakharov model [15–18] are potential, horizontal, and vertical velocities
applied for decomposition of the surface elevation spectrum into
components according to their order of nonlinearity [10,11]. The
developed method allows considerable improvement of the numeri-
cal effectiveness of the computational procedure, including a higher Appendix: Analytic Relation Between First-Order Free
convergence rate of the iteration process and the wave generation and Second-Order Bound Components of the Surface
accuracy. The importance of verification of the wave generation Elevation Spectra According to Zakharov Theory
accuracy in the vicinity of the wavemaker should be stressed. Surface elevation due to each spectral component of the first-
The fully nonlinear numerical wave tank based on the boundary order free waves may be presented as follows:
element method served for the iterative adjustment of the wave-  
maker motion. The results of the numerical simulations agree (1) (1)
η(1) = Re a(1) ei(k x−ω t) (A1)
well with the experimental data. The duration of the simulations
was reduced by shortening the numerical wave tank and by imple- where a (1) is its complex amplitude. The Zakharov theory [15–18]
mentation of the numerical beach that effectively prevents wave
n , n = 1 … N,
relates the complex spectrum of the free harmonics a(1)
reflection from the far end. The artificial pressure incorporated to the complex amplitudes of the corresponding second-order bound
into the dynamic free surface boundary condition was found to be components. For every pair of first-order harmonics ωm(1) and ω(1) n
significantly more effective for damping the reflected waves in with complex amplitudes am(1) and a(1)n , three second-order bound
the numerical procedure as compared with the introduction of a waves appear with the following frequencies ωj , wave numbers
(2)
viscous dissipative term. It should be stressed that the application k(2) (2)
j , and complex amplitudes aj :
of the numerical wave tank for adjustment of the driving signal is
particularly important in large-scale experiments. 1 = ωm + ωn ; k1 = km + kn
• ω(2) (1) (1) (2) (1) (1)

The accuracy of the generation process was found to be strongly 


dependent on the wave steepness at the wavemaker. Considerable 2gω p,1 V1 (ω p,1 , ω(1)
m , ωn , k1 , km , kn ) (1) (1)
(1) (2) (1) (1)
deviation of the free wave spectra from the target was observed a1 = −π
(2)
am an
ω(1)
m ωn
(1)
ω p,1 − ω(1)
m − ωn
(1)
for steep waves with the characteristic steepness exceeding εc ≈
0.2. Reduction of wave steepness at the wavemaker to εc ≈ 0.12 (A2)
substantially reduces the contribution of unwanted waves. The iter-
ative adjustment of the driving signal applied in the present study
demonstrated its effectiveness in the generation of the waves that • ω(2)
2 = −ωm + ωn ; k2 = −km + kn
(1) (1) (2) (1) (1)

are steep in the vicinity of the wavemaker. 


2gω p,2 V2 (ω p,2 , ω(1)
m , ωn , k2 , km , kn ) (1) * (1)
(1) (2) (1) (1)
a2 = −π
(2)
(am ) an
Funding Data ωm ωn
(1) (1)
ω p,2 + ω(1)
m − ωn
(1)

• Israel Ministry of Science, Technology and Space (Grant No. (A3)


3-12473).
• ω(2)
3 = −ωm − ωn ; k3 = −km − kn
(1) (1) (2) (1) (1)

Nomenclature 
h = mean water depth 2gω p,3 V3 (ω p,3 , ω(1)
m , ωn , k3 , km , kn ) (1) * (1) *
(1) (2) (1) (1)
a3 = −π
(2)
(am ) (an )
xf = focusing location of the wave train relative to the ωm ωn
(1) (1)
ω p,3 + ω(1)
m + ωn
(1)

wavemaker
(A4)
xt = location of the wavegauge in the experimental and
numerical wave tanks
a, ϕ = amplitudes and phases of the surface elevation
Here * denotes complex conjugate. In (A2)–(A4), the angular fre-

Sdrv, ϕdrv = amplitudes and phases of the driving signal
Swm, ϕwm = amplitudes and phases of the wavemaker motion quency ω p,j = gkj(2) tanh (kj(2) h) corresponds to that of a free wave
εc = wave steepness calculated as a product of the with the wave number k(2) j . The radian frequencies ωp,j do not
amplitude of the wave train envelope and the carrier appear in the bound waves’ spectrum and are used to calculate
wave number the coefficients:

     


1 g 1 ω0 ω1 ω2 2 ω2 ω0 ω1 2
V1 (ω0 , ω1 , ω2 , k0 , k1 , k2 ) = + k1 k2 − − k0 k1 (A5)
4π 2 2 ω1 ω2 g ω0 ω1 g

      


1 g ω2 ω0 ω1 2 ω1 ω0 ω2 2 ω0 ω1 ω2 2
V2 (ω0 , ω1 , ω2 , k0 , k1 , k2 ) = + k0 k1 − − k0 k2 − − k1 k2 (A6)
4π 2 ω0 ω1 g ω0 ω2 g ω1 ω2 g

    
1 g ω2 ω0 ω1 2 1 ω0 ω1 ω2 2
V3 (ω0 , ω1 , ω2 , k0 , k1 , k2 ) = + k0 k1 + + k1 k2 (A7)
4π 2 ω0 ω1 g 2 ω1 ω2 g

Journal of Offshore Mechanics and Arctic Engineering DECEMBER 2019, Vol. 141 / 061102-9
References [12] Schmittner, C., Kosleck, S., and Hennig, J., 2009, “A Phase-Amplitude Iteration
Scheme for the Optimization of Deterministic Wave Sequences,” ASME Paper
[1] Dean, R. G., and Dalrymple, R. A., 1991, Water Wave Mechanics for Engineers No. OMAE2009-80131.
and Scientists, World Scientific, Singapore. [13] Buldakov, E., Stagonas, D., and Simons, R., 2017, “Extreme Wave Groups in a
[2] Schäffer, H. A., 1996, “Second-Order Wavemaker Theory for Irregular Waves,” Wave Flume: Controlled Generation and Breaking Onset,” Coastal Eng., 128,
Ocean Eng., 23, pp. 47–88. pp. 75–83.
[3] Spinneken, J., and Swan, C., 2009, “Second-Order Wave Maker Theory Using [14] Mei, C. C., 1989, The Applied Dynamics of Ocean Surface Waves, World
Force-Feedback Control. Part I: A New Theory for Regular Wave Generation,” Scientific, Singapore.
Ocean Eng., 36(8), pp. 539–548. [15] Zakharov, V. E., 1968, “Stability of Periodic Waves of Finite Amplitude on the
[4] Spinneken, J., and Swan, C., 2009, “Second-Order Wave Maker Theory Using Surface of a Deep Fluid,” J. Appl. Mech. Tech. Phys., 9, pp. 190–194.
Force-Feedback Control. Part II: An Experimental Verification of Regular [16] Stiassnie, M., and Shemer, L., 1984, “On Modification of Zakharov Equation for
Wave Generation,” Ocean Eng., 36(8), pp. 549–555. Surface Gravity Waves,” J. Fluid Mech., 143, pp. 47–67.
[5] Aknin, D., and Spinneken, J., 2017, “A Laboratory Investigation Concerning the [17] Stiassnie, M., and Shemer, L., 1987, “Energy Computations for Evolution of
Superharmonic Free Wave Suppression in Shallow and Intermediate Water Class I and II Instabilities of Stokes Waves,” J. Fluid Mech., 174, pp. 299–312.
Conditions,” Coastal Eng., 120, pp. 112–132. [18] Krasitskii, V. P., 1994, “On the Reduced Equations in the Hamiltonian Theory of
[6] Houtani, H., Waseda, T., Fujimoto, W., Kiyomatsu, K., and Tanizawa, K., 2015, Weakly Nonlinear Surface Waves,” J. Fluid Mech., 272, pp. 1–20.
“Freak Wave Generation in a Wave Basin With HOSM-WG Method,” ASME [19] Grilli, S. T., and Svendsen, I. A., 1990, “Corner Problems and Global Accuracy in
Paper No. OMAE2015-42284. the Boundary Element Solution of Nonlinear Wave Flows,” Eng. Anal. Bound.
[7] Houtani, H., Waseda, T., Fujimoto, W., Kiyomatsu, K., and Tanizawa, K., 2018, Elem., 7(4), pp. 178–195.
“Generation of a Spatially Periodic Directional Wave Field in a Rectangular Wave [20] Grilli, S. T., and Subramanya, R., 1996, “Numerical Modeling of Wave
Basin Based on Higher-Order Spectral Simulation,” Ocean Eng., 169, pp. 428–441. Breaking Induced by Fixed or Moving Boundaries,” Comput. Mech., 17,
[8] Chaplin, J. R., 1996, “On Frequency-Focusing Unidirectional Waves,” pp. 374–391.
Int. J. Offshore Polar Eng., 6(2), pp. 131–137. [21] Grilli, S. T., Skourup, J., and Svendsen, I. A., 1989, “An Efficient Boundary
[9] Lugni, C., 2000, “An Investigation on the Interaction Between Free-Surface Element Method for Nonlinear Water Waves,” Eng. Anal. Bound. Elem., 6(2),
Waves and Floating Structure,” Ph.D. thesis, University of Rome, Italy (in Italian). pp. 97–107.
[10] Shemer, L., Goulitski, K., and Kit, E., 2006, “Steep Waves in Tanks: Experiments [22] Grilli, S. T., and Horrillo, J., 1997, “Numerical Generation and Absorption of
and Simulations,” ASME Paper No. OMAE2006–92547. Fully Nonlinear Periodic Waves,” J. Eng. Mech., pp. 1060–1069.
[11] Shemer, L., Goulitski, K., and Kit, E., 2007, “Evolution of Wide-Spectrum [23] Tian, Z., Perlin, M., and Choi, W., 2010, “Energy Dissipation in Two-
Unidirectional Wave Groups in a Tank: An Experimental and Numerical Dimensional Unsteady Plunging Breakers and an Eddy Viscosity Model,”
Study,” Eur. J. Mech. B Fluids, 26(2), pp. 193–219. J. Fluid Mech., 655, pp. 217–257.

061102-10 / Vol. 141, DECEMBER 2019 Transactions of the ASME

You might also like