Conference PaperPDF Available

SPH modeling of resonance wave pumping in closed tank: Parametric study

Authors:

Abstract and Figures

The Smoothed Particle Hydrodynamics (SPH) method with the unified semi-analytical wall boundary conditions is used to simulate a novel approach for free-surface wave pumping inspired from Libeau impedance pump: Resonance wave pumping. The results are compared to experimental data. Using the rule of thumb numerical parameters for the weakly in-compressible method (small enough dr and a speed of sound such that c0 ≈ 10 √ ghmax, where hmax is the maximum water detph), good agreement is found for both the quantitative instantaneous flow rate and qualitative surface wave dynamics near the two resonance frequencies of the resonance wave pump considered in this study. It is observed that varying the speed of sound changes significantly the mean flow for the highest resonance frequency considered. This effect is studied using compressible potential flow theory approach. Two effects of the weakly compressible method are identified: shifts of the resonance frequencies and compressible source of pumping. This approach is an interesting way of understanding the origin of the observed effect of c0 on the pumping behaviour and more generally in free-surface dynamics in weakly compressible SPH.
Content may be subject to copyright.
SPH modeling of resonance wave pumping in
closed tank: Parametric study
Remi Carmigniani
Saint-Venant Laboratory for hydraulics
Universit´
e Paris-Est
Chatou, France
rcarmi@me.edu
Antoine Joly
EDF & Saint-Venant Laboratory for hydraulics
Universit´
e Paris-Est
Chatou, France
Agnes Leroy
EDF & Saint-Venant Laboratory for hydraulics
Universit´
e Paris-Est
Chatou, France
Damien Violeau
EDF & Saint-Venant Laboratory for hydraulics
Universit´
e Paris-Est
Chatou, France
Abstract—The Smoothed Particle Hydrodynamics (SPH)
method with the unied semi-analytical wall boundary conditions
is used to simulate a novel approach for free-surface wave
pumping inspired from Libeau impedance pump: Resonance
wave pumping. The results are compared to experimental data.
Using the rule of thumb numerical parameters for the weakly in-
compressible method (small enough dr and a speed of sound such
that c010ghmax , where hmax is the maximum water detph),
good agreement is found for both the quantitative instantaneous
ow rate and qualitative surface wave dynamics near the two
resonance frequencies of the resonance wave pump considered in
this study. It is observed that varying the speed of sound changes
signicantly the mean ow for the highest resonance frequency
considered. This effect is studied using compressible potential
ow theory approach. Two effects of the weakly compressible
method are identied: shifts of the resonance frequencies and
compressible source of pumping. This approach is an interesting
way of understanding the origin of the observed effect of c0
on the pumping behaviour and more generally in free-surface
dynamics in weakly compressible SPH.
I. INTRODUCTION
Smoothed Particle Hydrodynamics (SPH) has been used to
simulate a variety of problems involving the Navier-Stokes
equations. It is a particularly promising method for free-
surface or multi-phase ow applications. In the present study,
a novel approach for free-surface wave pumping is introduced
and the SPH method is used to reproduce the experimental data
available. The system, resonance wave pump, is inspired by
the Liebau impedance pump introduced in 1954 [1]. A Liebau
pump consists of a exible and rigid tube connected together
at each extremities to form a closed loop. The exible tube
is pinched at an off-centred position. For certain frequencies
(resonance frequencies) a unidirectional ow rises in the
tubes. This elastic pump was studied both experimentally
[2]–[5] and numerically [6], [7]. A free-surface version of
such pump was recently investigated and similar behaviour
reported in an experimental study [8]. This near resonance
pumping behaviour results in a complex free-surface dynamic
that seems suitable for the SPH method.
In section II , the SPH method used in this study is briey
described and the main experimental results of the resonance
wave pump are summarised. In section III, the SPH method is
used to simulate the behaviour of the pump with fairly good
agreement with the experiment when using the general strategy
used in the literature to chose the numerical parameters (small
enough resolution, speed of sound of about 10 time the wave
speed and an adiabatic index ζ= 7 in the closure state
equation P V ζ=cte). The different numerical parameters are
then varied. It is outlined that the parameter speed of sound
of the weakly-compressible method has the most inuence on
the results. In the last section, the origin of this dependance
to the speed of sound is investigated using a potential ow
approach. It appears that the compressibility enhances the
pumping mechanism and shifts the resonance frequencies.
II. TH E SPH MET HO D AN D EXP ER IME NTAL RE SU LTS
A. Unied semi-analytical boundary conditions
The SPH method used throughout this paper is presented in
detail in the papers by Ferrand et al. [9] and Mayrhofer et al.
[10], [11]. In the present section a brief summary highlights
the main characteristics of the method.
The main difference from the classical SPH methods is that
the SPH approximation of a function fevaluated at a position
ais given by
[f]a=1
γa
bP
VbfbWab,(1)
where Pis the set of all particles, Vbdenotes the volume
of the particle b,Wab the kernel function as a function of
the distance between the particle band the position awhich,
throughout the paper, is the quintic polynomial by Wendland
[12]. Finally, γis a kernel wall renormalisation factor which
is dened as
γa=Ω
W(rarb)drb,(2)
100
11th international SPHERIC workshop Munich, Germany, June, 14-16 2016
where Ωis the uid domain. Note that this integral is equal to
one inside the uid domain and its value is bounded between
0 and 1 near the vicinity of a boundary. It is possible to
analytically evaluate this integral in 2-D [9] but in the present
study the value is computed throughout a governing equation:
dγa
dt=vR
a·γa,(3)
where vR
ais the relative velocity to the rigid boundary and
γais given by a surface integral:
γa=Ω
W(rarb)ndrb,(4)
with nthe inward oriented unit normal. The gradient and
divergence operators are then given by:
Gγ
a(f) = 1
γabPVb(fa+fb)aWab
1
γasS(fs+fa)γas,(5)
Dγ
a(B) = 1
γabPVb(BaBb)aWab
1
γasS(BsBa)γas.(6)
The additional terms containing the γas come from the
integration by parts that is used to move the differential from
the unknown function to the known kernel. The boundary in
the present method is discretised into discrete elements sS,
refered to segments with no associated mass. The masses are
located at the extremities of these segments where particles are
placed with a mass that depends on the opening angle of the
connected adjacent segments. In 2-D, the boundary segments
are 1-D line segments and:
γas =s
W(rarb)ndrb,(7)
i.e. the integral of the kernel on this boundary segment. Fur-
thermore second-order differential operators are approximated
using:
·(f B)
Lγ
a(f, B) = ρa
γabPmbfa+fb
ρaρb
Bab
r2
ab
rab ·aWab
2
γasSfs(B)·γas,
(8)
where rab =rab.
The method then solves the Navier-Stokes equations for com-
pressible uid:
d
dtρ
u=ρ·u
1
ρ(g+·¯
¯σ),(9)
where g=gezis the gravitational acceleration, and ¯
¯σis the
stress tensor:
¯
¯σ=p¯
¯
I+µu+uT.(10)
In order to close this system of equations, an equation of state
is dened to relate the pressure and the density [13]:
p=ρ0c2
0
ζρ
ρ0ζ
1,(11)
Fig. 1. Sketch of the experimental setup and notations. The PIV windows
shows where the data are recorded in the experiment while the vertical
dashed lines labeled xi{1,2,3}show where the data are sampled for the
SPH simulations. The ow rate is computed by trapezoidal rule integration.
The values of the different parameters are listed in table I. For the simulations
the values are the one without the error bars.
where ρ0is the reference density, c0is the numerical speed
of sound and ζis the adiabatic index equals to 7 for water.
The suitable values for c0are discussed later in this paper but
are generally of the two order of magnitude smaller than the
physical speed of sound in water.
B. Moving wall boundary conditions
At the moving boundary (a wave paddle in the present case,
as explained later), a no-slip boundary condition is imposed
and the velocity of the wall is prescribed. The pressure of a
wall particle has to be calculated from the uid to accurately
approximate the pressure gradient in the uid particle in the
vicinity of the boundary. As suggested by [14], a force balance
at the wall interface gives:
dvw
dt=pw
ρw
+g=aw(12)
where awis the acceleration of the wall. After a bit of algebra
it is possible to show that:
pw=bPpbWwb + (gaw)·bPρbrwb Wwb
bPWwb
.(13)
Even though in the present case it is easy to compute the
acceleration analytically, the wall acceleration is calculated
using a rst order marching scheme, thus only the previous
and current prescribed time steps are required to compute
the acceleration. This enables future integration of generalised
solid-uid interactions in the present method.
C. The resonance wave pump
In Fig. 1 a sketch of the experimental setup is shown. A
tank with a centred submerged plate of length L2Δxand
thickness exed at a water depth of hmin is lled with water
up to hmax. A paddle controlled in heave motion is generating
101
11th international SPHERIC workshop Munich, Germany, June, 14-16 2016
Fig. 2. Experimental results: bulk and rst harmonic ow rate as a function
of the forcing frequency (lled and empty circles and left and right axis
respectively) for a paddle of width 2a = 10.46 cm and a xed stroke amplitude
of S0= 0.69 cm. The points shaded in grey indicate cases where the bulk
ow is larger than the oscillatory rst harmonic ow rate. The red points
outline the frequency considered in this paper.
waves with a controlled frequency f0and S0. The width of the
paddle is 2a. Its mean position is such that the centre of the
paddle is located l1away from the wall and its mean draft is
d. The parameters used in the experiments are shown in table
I. In the experiment, the ow rate, φ, was measured under the
submerged plate using Particle Image Velocimetry (PIV) data
and a second camera was recording the entire tank to visualise
qualitatively the free surface dynamic.
Starting from a quasi at rest tank, the ow rate under the
submerged plate has a transitional stage before reaching an
asymptotic regime. The asymptotic ow rate is analysed using
a Fast Fourier Transform (FFT) and the mean ow (φ) and
rst harmonic (φ1at the frequency f0) are plotted in Fig. 2
as a function of the forcing frequency, f0, on two different
vertical axes (left and right respectively).
In [8], the authors show that the rst harmonic oscillatory
amplitude φ1can be found by solving a potential ow asymp-
totic problem. It was then pointed out by the authors that,
similarly to the exible tube pump, the mean ow is maxi-
mum at forcing frequencies where resonances are predicted.
The interesting behaviour occurs near the forcing frequency
f0= 1.2and 1.6 Hz (or f0L/ghmin = 1.585 and 2.11
respectively). In both cases the oscillatory ow of the paddle
is converted into a pulsating ow. In the rst case, f0= 1.2
Hz, the oscillations are as large as the mean ow resulting in
a pulsating ow while in the second case, f0= 1.6Hz, the
oscillations are much smaller than the mean ow resulting in
a quasi continuous ow. The instantaneous response measure
experimentally are shown in Fig. 3 and Fig. 5 respectively
(gray lines). Such behaviours are particularly interesting for
future applications since the oscillatory motion of the paddle
is converted into a quasi continuous ow. The system behaves
like a rectier: the alternative current (AC) is converted to
direct current (DC) through the system.
TABLE I
EXPERIMENTAL AND NUMERICAL PARAMETERS
Symbol Physical quantity Values
g Gravitational acceleration 9.81 m.s2
L Tank length 77.4 ±0.1 cm
l1Paddle off-centred position 18.7 ±0.1 cm
2 a Paddle length 10.46 ±0.05 cm
d Paddle mean draft 0.07 ±0.1cm
Δx Openings’ length 3 ±0.1cm
hmax Water depth 10.82 ±0.1 cm
W Recirculation section height 6.185 ±0.05 cm
eSubmerged plate thickness 1.2 ±0.05 cm
S0Paddle stroke amplitude 0.69 ±0.05 cm
f0Forcing frequency {1.6,1.2}Hz
xiPosition of the numerical sampling {3.2,38.7,74.2}cm
III. PARAME TR IC ST UDY O F THE RE SO NA NC E WAVE PUMP
The present study focuses on the response of the system
near the fourth resonance peak where the mean ow is larger
than the oscillation resulting in a unidirectional ow under
the submerged plate in the negative direction. The response
near the third peak (f0= 1.2Hz) is also considered at some
point to verify that the SPH method works in other parts of
the resonance wave pump spectrum. In all cases, the response
is near a linear resonance as shown in section IV. Different
numerical parameters are considered in this study. The code
being weakly compressible, one needs to worry about selecting
the proper value for the speed of sound c0. In the literature, it
is usually advised to x the speed of sound at the beginning
of the simulation so that:
c010 max (|u|)t(14)
where max (|u|)tis the maximum speed in modulo of the
uid, obtainable along the whole simulation. In these condi-
tions the Mach number M=u/c0is always less than 0.1 and
it is possible to consider the uid as weakly compressible.
For free surface ow it is usually considered that a proper
value for the maximum value can be taken as gh and then
the Mach number is identied to the Strouhal number. In our
case, we rst select the speed of sound to be c0= 11 m.s1.
Note that this value is about 20ghmin and about 10ghmax
and thus should be reasonable for this kind of simulations.
It is also possible to verify that the maximum speed in the
breaking waves induced in the shallow water is of the order
of magnitude of 1 m.s1. The value of the spatial resolution
is selected to ensure that there are about 70 particles in the
shallow water, thus dr = 0.0005 m (or hmin/dr = 68.7). The
value of dr is also consistent with the error bars of the different
measurement. The value for the state equation adiabatic index
is ζ= 7 for water. The numerical results are compared
to experimental results. In the available experiment data the
starting process of the paddle has not been recorded precisely
so a discrepancy at the beginning of the simulation is expected.
102
11th international SPHERIC workshop Munich, Germany, June, 14-16 2016
SPH with c0=11 m/s, hmin/dr =68.7
Experiment
0 5 10 15 20
25
-40
-30
-20
-10
0
10
20
t(s)
u.dS (cm2.s-1)
Fig. 3. Compared experimental (gray continuous line) and numerical ow rate
using the SPH method (black continuous line) for the numerical parameters
c0= 11 m.s1,ζ= 7,hmin /dr = 68.7,f0= 1.6Hz.
The paddle in the experiment was set-up near its top position
before starting while in the simulation it is started at its
mean position and its displacement is S(t) = S0sin (2πf0t).
In the experiment the authors were mostly interested in the
asymptotic behaviour thus this lack of care on the starting
process. For the sake of simplicity and to ease the comparison
the data of the experiment are shifted such that the rst peak
matches in the ow rate measurement. In the experiment the
ow rate was measured under the submerged plate using PIV.
Since the ow is weakly compressible here to calculate the
mean ow, data are sampled at three different positions under
the submerged plate (x={3.2,38.7,74.2}cm, see vertical
gray dashed lines in Fig. 1), integrated in the entire prole
using a trapezoidal rule and averaged. Figure 3 shows the
results obtained with the parameters described in the previous
paragraph for the 25 rst seconds with the forcing frequency
f0= 1.6Hz. The results are qualitatively and quantitatively
in good agreement with the experiments with these numerical
parameters. The free surface dynamic is also compared to the
experimental data for the ow around 15 seconds after the
start other a wave period. The results are shown in Fig. 4.
The wave dynamic is qualitatively correctly captured by the
SPH method. Finally, the frequency is varied to f0= 1.2Hz
and compared in a similar manner to the experimental results
to validate the method. The ow rate instantaneous response
is given in Fig. 5. There is a good agreement between the
simulation and the experiment. This validates the SPH method
with this set of parameters.
It is now interesting to see what are the effects of separately
varying the resolution and speed of sound. This part of the
study focuses on the frequency f0= 1.6Hz. First, the particles
size is varied. The resolution is reduced with hmin/dr = 34.35
(in blue) and increased hmin/dr = 137.4(in red) and compared
to the previous simulation (in black) and experiment (in gray).
Fig. 6 shows the different instantaneous responses. In all cases,
the pumping occurs. It is visible that the oscillations are
slightly better captured by the most rened simulations. This
results conrms the convergence of the method with these
Fig. 4. Compared numerical (top) and experimental (bottom) surface
dynamics at four different instances during a wave period for the numerical
parameters c0= 11 m.s1,ζ= 7,hmin/dr = 68.7,f0= 1.6Hz. The blue
square in the top images is the paddle position in the simulations.
SPH with c0=11 m/s, hmin/dr =68.7
Experiment
0 5 10 15 20 25
-40
-30
-20
-10
0
10
20
t(s)
u.dS (cm2.s-1)
Fig. 5. Compared experimental (gray continuous line) and numerical ow rate
using the SPH method (black continuous line) for the numerical parameters
c0= 11 m.s1,ζ= 7,hmin /dr = 68.7,f0= 1.2Hz.
parameters (c0= 11 m.s1,ζ= 7).
In the rest, the spatial resolution is now xed to hmin/dr =
68.7and the speed of sound is varied. The speed of sound is
varied between c0= 6 and 44 m.s1. Fig. 7 shows the ow
rate for four different values of the speed of sound for the case
f0= 1.6Hz. In all cases the 5 rst seconds of simulations
are fairly identical. However the behaviour changes drastically
with the different value of c0afterwards. Reducing the speed of
sound leads to a faster asymptotic ow rate while larger speed
of sound and thus more incompressible uid, leads to a slow
down and almost a complete stop of the pumping mechanism.
It is important to outline though than in all cases the mean
ow is negative in most part of the simulation. Also it is
important to remind the reader that up to this point the ow
rate was computed by averaging the horizontal velocity under
the submerged plate at three different locations because the
ow rate we are interested in is the incompressible one and
this way we eliminate the bias of the compressibility.
103
11th international SPHERIC workshop Munich, Germany, June, 14-16 2016
SPH with c0=11 m/s, hmin/dr =34.35
SPH with c0=11 m/s, hmin/dr =68.7
SPH with c0=11 m/s, hmin/dr =137.4
Experiment
0 5 10 15 20 25
-40
-30
-20
-10
0
10
20
t(s)
u.dS (cm2.s-1)
Fig. 6. Compared experimental (gray continuous line) and numerical ow
rate using the SPH method for the numerical parameters c0= 11 m.s1,
ζ= 7,f0= 1.6Hz and different resolutions (in blue, black and red).
Identically, we varied the speed of sound for f0= 1.2Hz.
The results are shown in Fig. 8 for c0= 6 and c0= 22
m.s1. The variation of c0in this range does not have any
signicant impact on the dynamic. It is noticeable though that
the ow is slightly slower for c0= 22 m.s1but the impact
is not as signicant as for the frequency f0= 1.6Hz. In the
next section a potential approach is developed to investigate
the inuence of compressibility on the resonance wave pump
with the forcing frequency f0.
IV. COM PRE SS IBL E POT ENT IA L FLOW HA RMO NI C TH EORY
To investigate the effect of the compressibility on the
simulation results and have a better understanding of the
dependance of the ow rate with c0, identically to what
was done in [8], we analyse the potential ow linearised
response of the system to an innitesimal stroke amplitude.
Since there seems to exist an asymptotic regime, we seek
solution in the form u=U(x) + u1(x, t) + h.o., where
u1=iS0ωϕ(x)eiωt. We further assume that the mean
ow is small compare to the oscillatory part. If this assumption
may seem wrong, it is important to notice that in the Fig. 2
the mean ow is generally largelly smaller than the oscillatory
part. Moreover, the mean ow was reported to vary quasi-
quadratically with the stroke amplitude and thus if the stroke
amplitude is small enough this assumption holds.
The potential scalar ψ=iS0ωϕ (x)eiωtveries the
following set of linearized equations and boundary conditions:
t,tϕc2
02ϕ+gzϕ= 0 ,for xΩ
nϕ= 0 ,for , x Γwall
nϕ=ez·n,for , x Γpaddle
ω2ϕ+gzϕ= 0 ,for , x ΓFS
(15)
where Ωis the uid domain, Γwall the xed wall boundaries,
Γpaddle the moving paddle boundary in contact with water,
ΓFS the free surface (see Appendix A). This problem can be
solved using the Finite Element Method (FEM). In the present
study, we used Mathematicas FEM package to solve this set
of equations. The free surface deformation is reconstructed
using the relation η=S0zϕ|z=0 eiωt. The equivalent
Fig. 7. Compared experimental (gray continuous line) and numerical ow
rate using the SPH method for the numerical parameters ζ= 7,f0= 1.6
Hz, hmin/dr = 68.7and different values of the speed of sound.
problem for incompressible uid is the limit c0 and the
rst equation of the system Eq. 15 simplies to the Laplace
equation for ϕ.
First, we look at the amplitude of oscillations of the ow rate
as a function of frequency dened as the average potential
ow rate at the three position x1,x2and x3:
φ1=S0ω
3W
0
i{1,2,3}
xϕ(xi, z)dz
(16)
The frequency response of the system is shown for different
values of the speed of sound (or Mach numbers M=
104
11th international SPHERIC workshop Munich, Germany, June, 14-16 2016
SPH with c0=6 m/s, hmin/dr =68.7
Experiment
-30
-20
-10
0
10
u.dS (cm2.s-1)
SPH with c0=22 m/s, hmin/dr =68.7
Experiment
0 5 10 15 20
25
-30
-20
-10
0
10
t(s)
u.dS (cm2.s-1)
Fig. 8. Compared experimental (gray continuous line) and numerical ow
rate using the SPH method for the numerical parameters ζ= 7,f0= 1.2
Hz, hmin/dr = 68.7and different values of the speed of sound.
ghmax /c0) near the frequencies f0= 1.2and 1.6 Hz in
Fig. 9. In both cases, a smaller speed of sound (larger Mach
number) shifts the curves to left. The shift is more important
near f0= 1.6Hz resulting in a larger impact of the speed
of sound on the results. We see that for the case f0= 1.2
Hz the difference between the spectrum with values of the
speed of sound c0= 11 m.s1(M= 0.09, black curve)
and the incompressible case is rather small, suggesting that
the solution is fairly converged in the incompressible sense.
However, in the case f0= 1.6Hz, the curve c0= 11
m.s1(M= 0.09) the difference is still large compared to the
incompressible case. This might explain the drastic difference
of response between the different values of the speed of sound
investigated in the previous section. This is outlined in Fig. 10
where the value of the amplitude at the frequencies f0= 1.2
and f0= 1.6Hz are plotted as a function of the Mach number.
Furthermore, one characteristic of the potential theory for
compressible uid is that the amplitude of oscillation of
the ow rate under the submerged plate is a function of
the position x. In particular, the amplitude of oscillation
of the ow rate between the position x1and x3might be
different. A larger amplitude of oscillation at the position
x1compared to x3results in a lower mean pressure at the
position x1that might amplify articially the ow rate. Fig.
11 shows the theoretical amplitude of oscillation of the ow
rate under the submerged plate as a function of the position
nondimensionalised by the incompressible potential solution.
The amplitude of oscillations are larger at the position x1than
x3and thus the effect described before could rise. We see that
M=0.17
M=0.09
M=0.02
M=0
1.2 1.3 1.4 1.5 1.6
0.0
0.1
0.2
0.3
0.4
0.5
0.6
f
(
Hz
)
ϕ1
2a S0ω
Fig. 9. Oscillatory ow rate amplitude as function of frequency in the
compressible potential ow for different Mach numbers. When the Mach
number increases the curves are shifted to the left resulting in a modication
of the resonance frequencies. It appears that this effect is magnied near the
frequency f0= 1.6Hz, while near f0= 1.2Hz the black and red curves
are collapsed with the incompressible one (gray dashed line).
f0=1.2 Hz
f0=1.6 Hz
10 15 20 25 30 35 40
0.0
0.2
0.4
0.6
0.8
1.0
1/M
ϕ1comp/ϕ1incomp
Fig. 10. Ratio of the oscillatory ow rate amplitude for compressible and
incompressible potential ows as function of the Mach numbers. The ratio
goes to 1 when the Mach number goes to zero. For f0= 1.2Hz (in black)
the compressible error is smaller than in the case f0= 1.6Hz (in gray).
this could generate a source of mean ow and that this source
is less important as the speed of sound increases.
To investigate these effects in the SPH method, an FFT
analysis of the ow rate at the three different positions sampled
in the simulations is performed at the end of time series. The
results are shown for the two extreme values of the speed
of sound c0= 6 m.s1(M= 0.17) and c0= 44 m.s1
(M= 0.02) in Fig. 12 and Fig. 13, respectively. The data are
sampled on the last 5 seconds of the simulation for t [20; 25]
s. In the case c0= 44 m.s1(M= 0.02) the FFT mostly
consists of a mean ow and a peak near the forcing frequency
as expected. However the mean ow is smaller than observed
experimentally. For c0= 6 m.s1(M= 0.17), there is a
larger variety of modes. Some of these modes are acoustic
waves in the recirculation section. We do observe that the
105
11th international SPHERIC workshop Munich, Germany, June, 14-16 2016
c0=6 m/s
c0=11 m/s
c0=44 m/s
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
0.0
0.2
0.4
0.6
0.8
1.0
x
-H
-(h+e)ucomp (z).z
-H
-(h+e)uincomp (z).z
Fig. 11. Compressible potential amplitude of oscillation of the rst harmonic
under the submerged plate as a function of the position for different values
of the speed of sound nondimensionalised by the incompressible potential
solution. The gray dashed vertical lines show the position x1and x3. The
amplitude are larger at the position x1than x3in all the cases. The difference
is more important for the lower speed of sound. This might amplify articially
the mean ow.
amplitude of oscillations are larger for case c0= 44 m.s1
(M= 0.02) at the forcing frequency f0than for c0= 6
m.s1(M= 0.17). This is consistent with Fig. 9 where
we see that the shift increases the amplitude of oscillation
at the constant frequency f0= 1.6Hz. We see that the
oscillations are larger at the position x1than x3for the case
with c0= 6 m.s1(M= 0.17). This might help the ow to
pump in the negative direction. Nonetheless, the effect is not
as drastic as the compressible potential theory may suggest.
The oscillations are also smaller at the center for this case. It
is however interesting to point out that the compressibility is
a source of mean ow. The global direction of this secondary
source of mean ow is however uncertain.
V. CONCLUSION
In the present study the SPH method was applied to a
resonance free-surface wave pump near two resonance fre-
quencies. Good agreement is found for a large range of
numerical parameters for both the mean ow generated by
the pump and the free surface dynamic when compared to the
experiment. Nonetheless, it is pointed out that compressibility
might generate pumping that does not occur for larger speed
of sound. A potential approach is developed to demonstrate
this behaviour. It is outlined that the observed inuence to
the mean ow is nonetheless small and this effect should
be negligible. This approach is however an interesting way
of understanding the origin of the observed effect of c0on
the pumping behaviour (and more generally in free-surface
dynamics in weakly compressible SPH). The main difference
appeared to be the shift in the spectrum due to compressibility
which is more likely to explain the difference of behaviour. In
the present study the water depth was not varied in the range
of the experimental error and the particle size was not rened
for larger speed of sound.
Position x1
Position x2
Position x3
02468
0
1
2
3
4
5
6
7
f/f
0
|ϕ|(cm2.s-1)
Fig. 12. FFT of the simulated ow rate at the different positions for
c0= 6 m.s1(or M= 0.17) and a forcing frequency f0= 1.6Hz. The
oscillations are larger at the position x1than x3due to the incompressibility.
This difference in amplitude results in a lower time averaged pressure at the
position x1. Acoustic waves oscillations are visible in the FFT spectrum near
the abscissa 4. The data are sampled on the last 5 seconds of the simulation
for t[20; 25] s.
Position x1
Position x2
Position x3
02468
0
1
2
3
4
5
6
7
f/f0
|ϕ|(cm2.s-1)
Fig. 13. FFT of the simulated ow rate at the different positions for c0= 44
m.s1(or M= 0.02) and a forcing frequency f0= 1.6Hz. There is almost
no difference between the oscillations at the different position and almost no
acoustic waves. The data are sampled on the last 5 seconds of the simulation
for t[20; 25] s.
APPENDIX A
DER IVATIO N OF TH E POT ENT IA L EQUATI ON FO R
CO MP RES SI BL E FLUI D
One can show that the potential scalar ψ=
iS0ωϕ (x)eiωt, veries the generalised Bernoulli
equation for compressible uid:
tψ+1
2(ψ)2+p
pref
dp
ρ(p)+gz =cte (17)
106
11th international SPHERIC workshop Munich, Germany, June, 14-16 2016
where the constant is assumed independent of time. Taking
the time derivative of the Bernoulli equations and noting that
tp
pref
dp
ρ(p)=1
ρtp(18)
it is easy to show that:
t,tψ+1
2tu2+1
ρtp= 0.(19)
Recall the momentum equation for irrotational uid (or
inviscid):
dtui+1
ρxip+gδi3= 0 (20)
where dtis the Lagrangian derivative and xiis the spatial
derivative with respect to the direction i,δij is the kronecker
delta, and using Einstein’s notation. We multiply this equation
by ui. Using the continuity equation and deriving the relation:
1
ρuixip=c2
0xiui1
ρtp, (21)
it comes after some basic algebra:
1
2tu2
i+1
2uixiu2
jc2
0xiui1
ρtp+gu3= 0.(22)
Finally, summing Eq. 19 and Eq. 22 and using the fact that
ui=xiψ, we derive the potential equation for compressible
uid:
t,tψc2
02ψ+gzψ=tu21
2u·u2.(23)
Considering now only the rst order terms, it yields:
t,tψc2
02ψ+gzψ= 0.(24)
Note that in the limit of incompressible uid c0become innite
and we get the Laplace equation for ψ.
ACK NOWL ED GME NT
The authors would like to thank EDF and Caltech for the
help with running the simulations and the assistance in con-
ducting the experiment. We wish to acknowledge the funding
received from EDF to conduct this research and the Gordon
and Betty Moore Foundation for their generous support for
the experiments. Our gratitude also goes to Martin Ferrand
and Alex Ghaitanellis for their support and help to run the
simulations.
REFERENCES
[1] G. Liebau, ¨
Uber ein ventilloses Pumpprinzip,” Naturwissenschaften,
vol. 41, no. 14, pp. 327–327, 1954.
[2] A. I. Hickerson and M. Gharib, “On the resonance of pliant tube as a
mechanism for valveless pumping, Journal of Fluid Mechanics, vol.
555, pp. 141–148, January 2006.
[3] ——, “Flow characterization of a valveless impedance driven pump,”
in American Physical Society, Division of Fluid Dynamics 56th Annual
Meeting, 2003.
[4] D. Rinderknecht, A. I. Hickerson, and M. Gharib, “A valveless micro
impedance pump driven by electromagnetic actuation, Journal of
Micromechanics and Microengineering, vol. 15, no. 4, p. 861, 2005.
[Online]. Available: http://stacks.iop.org/0960-1317/15/i=4/a=026
[5] T. Bringley, S. Childress, N. Vandenberghe, and J. Zhang, An experi-
mental investigation and a simple model of valveless pump,” Physics of
Fluids, vol. 20, no. 033602, 2008.
[6] I. Avrahami and M. Gharib, “Computational studies of resonance wave
pumping in compliant tubes,” Journal of Fluid Mechanics, vol. 608, pp.
139–160, April 2008.
[7] E. Jung, S. Lim, W. Lee, and S. Lee, “Computational models of valveless
pumping using the immersed boundary method,” Comput. Methods Appl.
Mech. Engrg., vol. 197, pp. 2329–2339, 2008.
[8] R. Carmigniani, M. Benoit, D. Violeau, and M. Gharib, “Resonance
wave pumping with surface waves,” Journal of Fluid Mechanics (sub-
mitted), 2016.
[9] M.Ferrand, D.R.Lurence, D.Rogers, D. Violeau, and C. Kassiotis, “Uni-
ed semi-analytical wall boundary conditions for inviscid,laminar or
turbulent ows in the meshless SPH method.” International Journal for
Numerical Methods in Fluids, vol. 71, no. 4, pp. 446–472, 2011.
[10] A. Mayrhofer, B. D. Rogers, D. Violeau, and M. Ferrand, “Study of
differential operators in the context of the semi-analytical wall boundary
conditions,” Proc. 7th SPHERIC International workshop, Prato (Italy),
29 May-1 June 2012, no. 4, 2012.
[11] ——, “Investigation of wall bounded ows using SPH and the unied
semi-analytical wall boundary conditions,” Computer Physics Commu-
nications, vol. 184, no. 11, pp. 2515–2527, 2013.
[12] H. Wendland, “Piecewise polynomial, positive denite and compactly
supported radial functions of minimal degree,” Advances in Computa-
tional Mathematics, vol. 4, no. 1, pp. 389–396, 1995.
[13] Yuan-Hui Li, “Equation of State of Water and Sea Water,” Journal of
Geophysical Research, vol. 72, no. 10, pp. 2665–2678, 1967.
[14] S. Adami, X. Y. Hu, and N. A. Adams, “A generalized wall boundary
condition for smoothed particle hydrodynamics,” Journal of Computa-
tional Physics, vol. 231, no. 21, pp. 7057–7075, 2012.
107
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
Wall boundary conditions in smoothed particle hydrodynamics (SPH) is a key issue to perform accurate simulations. We propose here a new approach based on a renormalising factor for writing all boundary terms. This factor depends on the local shape of a wall and on the position of a particle relative to the wall, which is described by segments (in two‐dimensions), instead of the cumbersome fictitious or ghost particles used in most existing SPH models. By solving a dynamic equation for the renormalising factor, we significantly improve traditional wall treatment in SPH, for pressure forces, wall friction and turbulent conditions. The new model is demonstrated for cases including hydrostatic conditions for still water in a tank of complex geometry and a dam break over triangular bed profile with sharp angle where significant improved behaviour is obtained in comparison with the conventional boundary techniques. The latter case is also compared with a finite volume and volume‐of‐fluid scheme. The performance of the model for a two‐dimensional laminar flow in a channel is demonstrated where the profiles of velocity are in agreement with the theoretical ones, demonstrating that the derived wall shear stress balances the pressure gradient. Finally, the performance of the model is demonstrated for flow in a schematic fish pass where both the velocity field and turbulent viscosity fields are satisfactorily reproduced compared with mesh‐based codes. Copyright © 2012 John Wiley & Sons, Ltd.
Article
Full-text available
Valveless pumping can be achieved through the periodic compression of a pliant tube asymmetrically from its interfaces to different tubing or reservoirs. A mismatch of characteristic impedance between the flow channels is necessary for creating wave reflection sites. Previous experimental studies of the behaviour of such a pump were continued in order to demonstrate the wave mechanics necessary for the build-up of pressure and net flow. Specific measurements of the transient and resonant properties were used to relate the bulk responses to the pump mechanics. Ultrasound imaging through the tube wall allowed visualization of the wall motion concurrently with pressure and flow measurements. For analysis, a one-dimensional wave model was constructed which predicted many of the characteristics exhibited by the experiments.
Article
Full-text available
The P-V-T data on water by Amagat and Kennedy et al. and on sea water by Newton and Kennedy, the compressibility data on water by Diaz Peña and McGlashan, the bulk compression data on water and sea water by Ekman, and the sound velocity data on water and sea water by Del Grosso and Wilson have been analyzed using the Tait and the Tait-Gibson equations. The P-V-T relationship for water can be well represented by the Tait equation, V0(P) = V0(1) - C × log [(B + P)/(B + 1)], and for sea water by the Tait-Gibson equation, V(P) = V(1) - (1 - S × 10−3) × C log [(B* + P)/(B* + 1)] where C = 0.315 × V0(1) and B = 2668.0 + 19.867t - 0.311t² + 1.778 × 10−3t³, for Amagat data in the range of 0 ≤ t ≤ 45°C and 1 ≤ P ≤ 1000 bars and B* = (2670.8 + 6.89656 × S) + (19.39 - 0.0703178 × S)t - 0.223t² for Ekman's sea water data in the range of 0 ≤ t ≤ 20°C, 1 ≤ P ≤ 1000 bars, and 30 ≤ S ≤ 40‰.
Article
Full-text available
The valveless impedance pump is a simple design that allows the producion or amplification of a flow without the requirement for valves or impellers. It is based on fluid-filled flexible tubing, connected to tubing of different impedances. Pumping is achieved by a periodic excitation at an off-centre position relative to the tube ends. This paper presents a comprehensive study of the fluid and structural dynamics in an impedance pump model using numerical simulations. An axisymmetric finite-element model of both the fluid and solid domains is used with direct coupling at the interface. By examining a wide range of parameters, the pump's resonance nature is described and the concept of resonance wave pumping is discussed. The main driving mechanism of the flow in the tube is the reflection of waves at the tube boundary and the wave dynamics in the passive tube. This concept is supported by three different analyses: (i) time-dependent pressure and flow wave dynamics along the tube, (ii) calculations of pressure–flow loop areas along the passive tube for a description of energy conversion, and (iii) an integral description of total work done by the pump on the fluid. It is shown that at some frequencies, the energy given to the system by the excitation is converted by the elastic tube to kinetic energy at the tube outlet, resulting in an efficient pumping mechanism and thus significantly higher flow rate. It is also shown that pumping can be achieved with any impedance mismatch at one boundary and that the outlet configuration does not necessarily need to be a tube.
Article
In this paper, we present a novel extension of impedance (Liebau) wave pumping to a free-surface condition where resonance pumping could be used for hydraulic energy harvesting. Similar pumping behaviours are reported. Surface envelopes of the free surface are shown and outline two different dynamics: U-tube oscillator and wave/resonance pumping. The latter is particularly interesting, since, from an oscillatory motion, a unidirectional flow with small to moderate oscillations is generated. A linear theory is developed to evaluate pseudo-analytically the resonance frequencies of the pump using eigenfunction expansions, and a simplified model is proposed to understand the main pumping mechanism in this type of pump. It is found that the Stokes mass transport is driving the pump. The conversion of energy from paddle oscillation to mean flow is evaluated. Efficiency up to 22 % is reported.
Article
We construct a valveless pump consisting of a section of elastic tube and a section of rigid tube connected in a closed loop and filled with water. By periodically squeezing the elastic tube at an asymmetric location, a persistent flow around the tubes is created. This effect, called the Liebau phenomenon or valveless pumping, has been known for some time but is still not completely understood. We study the flow rates for various squeezing locations, frequencies, and elastic tube rigidities. To understand valveless pumping, we formulate a simple model that can be described by ordinary differential equations. The time series of flow velocities generated by the model are qualitatively and quantitatively similar to those seen in the experiment. The model provides a physical explanation of valveless pumping, and it allows us to identify the essential pumping mechanisms. © 2008 American Institute of Physics.
Article
Mathematical models of valveless pumping can be represented by either a closed loop system or an open tube system. In this paper, we present a three-dimensional model of valveless pumping in a closed loop system. We also present a two-dimensional model using an open elastic cylinder contained in a rigid tank. In both models, we take the periodic compress-and-release action at the asymmetric location of the soft tube and observe the existence of a net flow and the important features of valveless pumping that have been reported in the previous models or experiments. The innovative idea of this work is that we explain the existence of a net flow by introducing the concept of the signed area of the flow-pressure loop over one cycle, which represents the power in the system. The direction and the magnitude of a net flow can also be explained by the sign and the amount of power, which is work done on the fluid by the fluid pressure and the elastic wall over one period, respectively.
Article
We construct a new class of positive definite and compactly supported radial functions which consist of a univariate polynomial within their support. For given smoothness and space dimension it is proved that they are of minimal degree and unique up to a constant factor. Finally, we establish connections between already known functions of this kind.