Download as pdf or txt
Download as pdf or txt
You are on page 1of 29

The current issue and full text archive of this journal is available on Emerald Insight at:

https://www.emerald.com/insight/0961-5539.htm

Buoy
A numerical and experimental interacting
study of a buoy interacting with waves

with waves
Jonathan Núñez Aedo and Marcela A. Cruchaga
Departmento de Ingeniería Mecanica,
Facultade de Ingeniería, Universidad de Santiago, Santiago, Chile, and Received 25 January 2023
Revised 24 June 2023
13 September 2023
Mario A. Storti Accepted 11 October 2023
Centro de Investigacion en Metodos Computacionales CIMEC,
Universidad Nacional del Litoral UNL, Consejo Nacional de Investigaciones
Científicas y Tecnicas CONICET, Santa Fe, Argentina

Abstract
Purpose – This paper aims to report the study of a fluid buoy system that includes wave effects, with
particular emphasis on validating the numerical results with experimental data.
Design/methodology/approach – A fluid–solid coupled algorithm is proposed to describe the motion of
a rigid buoy under the effects of waves. The Navier–Stokes equations are solved with the open-source finite
volume package Code Saturne, in which a free-surface capture technique and equations of motion for the solid
are implemented. An ad hoc experiment on a laboratory scale is built. A buoy is placed into a tank partially
filled with water; the tank is mounted into a shake table and subjected to controlled motion that promotes
waves. The experiment allows for recording the evolution of the free surface at the control points using the
ultrasonic sensors and the movement of the buoy by tracking the markers by postprocessing the recorded
videos. The numerical results are validated by comparison with the experimental data.
Findings – The implemented free-surface technique, developed within the framework of the finite-volume
method, is validated. The best-obtained agreement is for small amplitudes compatible with the waves evolving
under deep-water conditions. Second, the algorithm proposed to describe rigid-body motion, including wave
analysis, is validated. The numerical body motion and wave pattern satisfactorily matched the experimental data.
The complete 3D proposed model can realistically describe buoy motions under the effects of stationary waves.
Originality/value – The novel aspects of this study encompass the implementation of a fluid–structure
interaction strategy to describe rigid-body motion, including wave effects in a finite-volume context, and the reported
free-surface and buoy position measurements from experiments. To the best of the authors’ knowledge, the numerical
strategy, the validation of the computed results and the experimental data are all original contributions of this work.
Keywords Fluid–structure interaction, Numerical simulation, Experimental analysis,
Numerical verification
Paper type Research paper

The authors would like to thank the support given by the institutions listed in the funding list. The
first exploratory experimental test was conducted in collaboration with the Ing. Maitane Uribe; the
authors thank her for her strong dedication in the early stages of the study.
Funding: The study was supported by grants ANID-Fondecyt 1210228, Departamento de
Investigaciones Científicas y Tecnologicas, Universidad de Santiago de Chile DICYT 052016C, and
Universidad Nacional del Litoral CAI+D 50620190100110LI, ANPCyT PICT 2018-2920 and 2018-
International Journal of Numerical
1607, PICT-E 2018-0271. Methods for Heat & Fluid Flow
Ethics statement: This study did not include studies on human subjects, data, human tissues, © Emerald Publishing Limited
0961-5539
animals, or any other objects under legal protection. DOI 10.1108/HFF-01-2023-0040
HFF Nomenclature
h(t) = Free-surface evolution in time;
hmax = Maximum value of free surface;
hmin = Minimum value of free surface;
Atank = Amplitude imposed on the tank;
ftank = Frequency imposed on the tank;
xbuoy, ybuoy, zbuoy = Buoy position in x-, y- and z-axes;
r = Density of the fluid;
m = Dynamic viscosity of the fluid;
v = Velocity vector;
r = Stress tensor;
b = Body force;
g = Gravity;
ar = Acceleration of the frame of reference;
p = Pressure;
I = Identity tensor;
s = Viscous stress tensor;
$ = Spatial gradient operator;
X = Domain of analysis;
Y = Time of analysis;
c = Convective velocity;
vm = Velocity of the mesh (ALE technique);
w(x, t) = Level-set distance function;
Subindexes: w, a, FS, m = Refers to water, air, free surface (air–water) and mesh (ALE); respectively;
Xm = Mesh coordinates (ALE);
Mbuoy = Buoy mass;
€z buoy = Buoy acceleration;
Ffluid = Hydrodynamic forces on the buoy;
Cb = Buoy boundary;
b, g = Parameters of Newmark method;
_
ðÞ ; ðÞ = Dot/bar over a variable denotes time derivative/known initial value;
respectively;
Dx, Dy, Dz = Mesh sizes in x-, y- and z-axes;
Dt = Time step;
S0, S1, S2, S3 = Ultrasonic sensor location and signals;
hR = Wave amplitude at right to the buoy (at S1 or S0); and
hL = Wave amplitude at left to the buoy (at S2 or S3).

1. Introduction
Fluid–solid interaction (FSI) problems are commonly encountered in many engineering
applications, such as energy generation (Bazilevs et al., 2011; Ducassou et al., 2017;
Gonzalez et al., 2017; Tampier and Grueter, 2017; Pozzi et al., 2017), the food industry
(Orona et al., 2018), naval engineering (Calderer et al., 2014) and biomechanics (Hron and
Turek, 2006).
To mathematically model fluid-structure problems, the fluid and solid governing
equations must be coupled. Staggered schemes for coupling fluid and solid fields (Hu et al.,
2001; Storti et al., 2012; Zamora et al., 2019) are an alternative methodology to monolithic or
unified schemes (Hron and Turek, 2006; Idelsohn et al., 2008; Robinson-Mosher et al., 2011),
which solve both fields simultaneously but present a higher computational cost. In a
staggered context, the fluid and solid equations are solved independently, and coupling is Buoy
achieved by exchanging variables at the end of each time-step analysis or at each iteration interacting
within the loop used to solve the nonlinear equations inside each time step. This strategy
reduces the computational cost but requires the evaluation of its stability and precision to
with waves
guarantee minimal errors in the numerical predictions.
Two distinct formulations can be applied to describe the interface between the solid and
fluid fields, i.e. the solid contour. Body-fitted approaches exactly represent the body, and the
mesh adjusts it for every time step of the analysis. These techniques require mesh motion
algorithms and remeshing techniques to adapt the discretization of the fluid surrounding the
solid. Arbitrary Lagrangian–Eulerian (ALE) techniques (Sarrate et al., 2001; Hu et al., 2001;
Baiges et al., 2017; Cai et al., 2021) have been extensively used to include mesh motions in the
governing equations. Sliding meshes have also been proposed for tracking rigid solids
(Devolder et al., 2018).
As an alternative, a solid can be described using embedded schemes on fixed-fluid
meshes. Representative techniques within this group of formulations include the fictitious
domain (Glowinski et al., 2001; Ducassou et al., 2017), immersed boundary (Mittal and
Iaccarino, 2005; Lo et al., 2014) and embedded boundary (Löhner et al., 2004; Yang et al.,
2008) approaches. They can easily represent complex body geometries at relatively low
computational costs.
In problems where free-surface analysis must be included, such as in wave energy
converters, fluid structure formulations also require inclusion of this topic (Idelsohn
et al., 2008; Cruchaga et al., 2012; Koh et al., 2013; Calderer et al., 2014; Ducassou et al.,
2017). The free surface can also be represented by fitting mesh techniques, computing
only the fluid as a unique phase or considering both air and fluid regions. The last two-
phase approach is commonly used in fixed-mesh techniques and requires an additional
equation to describe the free surface. In this context, Eulerian descriptions, such as
level-set techniques (Zhang et al., 2014; Bai et al., 2015; Zhao and Chen, 2015; Schillaci
et al., 2021) and volume-of-fluid methods (Celebi and Akyildiz, 2002; Kleefsman et al.,
2005; Connell et al., 2018; Zou and Abdelkhalik, 2021; Wang et al., 2021) have been
extensively used. The moving Lagrangian approach (Cruchaga et al., 2001) has been
proposed as an alternative for describing free surfaces (Cruchaga et al., 2006, 2010).
These approaches have also been used to describe fluid–solid interfaces (Gonzalez et al.,
2017, 2019).
In particular, wave–structure interaction analyses have been used to study devices
that convert wave energy to mechanical energy. A significant effort has been made more
recently to numerically model these devices and validate the numerical responses with
experimental data (Connell et al., 2018; Jafari et al., 2018; Tran et al., 2019; Ransley et al.,
2021; Amaechi et al., 2021; Schillaci et al., 2021; Wang et al., 2021). A list of plausible
models and codes applied to buoys of different geometries can be found in the literature
(Amaechi et al., 2021; Ransley et al., 2021). Many proposed numerical models were built
using well-established commercial software (Tampier and Grueter, 2017; Connell et al.,
2018; Jafari et al., 2018; Wiegard et al., 2019; Zou and Abdelkhalik, 2021; Wang et al., 2021)
and open-source (Devolder et al., 2018) or partially open-source codes (Jafari et al., 2018;
Tran et al., 2019). Regarding experimental validation of the simulations, in Devolder et al.
(2018), the buoy motion was described with one degree of freedom, and the model was
able to solve more than one buoy. The numerical results were compared with
experimental data reported by the same group of researchers in Stratigaki et al. (2014).
The experimental data obtained by Wang et al. (2017) were used to validate the numerical
results reported in Wang et al. (2021). The work presented in Connell et al. (2018)
HFF encompassed a numerical study of a wave energy converter experimentally tested at a
laboratory scale of 1:50. This list of works provides examples of simulations within
different computational frameworks and the current interest in performing experimental
validations using reduced scales. In Núñez et al. (2022), a study at different laboratory
scales evaluated the possibility of comparing waves between those generated in a
hydraulic channel and during sloshing. That work opened the possibility of
experimentally studying buoys subjected to sloshing as a preliminary stage for the
design of systems such as those proposed in this work, where the effect on a buoy of fluid
dynamics considering waves is specifically studied. In this experimental study, we focus
on analyzing the behavior of a buoy using a 1:8 geometric scale model of a buoy adopted
by Tampier and Grueter (2017).
In this study, we propose the development of a strategy for solving fluid–rigid solid
problems under free-surface effects. The fluid domain was described using the finite
volume Code Saturne (CS) (EDF R&D, 2017). The rigid body equations are implemented
in this code. A staggered scheme is used to solve the coupling, and a body-fitting mesh
technique is used to adjust the fluid domain to the current solid position. In addition,
free-surface evolution is captured using a level-set-type technique also implemented in
CS. The formulation numerically solves the problem within a two-phase flow scenario,
i.e. the free surface is strictly speaking a gas–liquid (air–water in this case) interface.
Note that the current version of the code considers the solution of the Navier–Stokes
equations for finite volumes and the fluid–structure interaction of flow around an
immersed body. In this study, a technique for capturing a free surface and its interaction
with a floating body is developed. Moreover, a laboratory-scale experiment design to
study the behavior of a buoy subjected to sloshing is reported. The computed numerical
predictions of the waves and buoy motion over time are compared with experimental
results. Moreover, to verify the dynamics of the system and the advantages of the
present formulation, a model based on the potential theory is developed. The main
aspects of this study are as follows:
 An FSI formulation for floating rigid bodies is developed into an open-source code.
 A fluid–rigid solid experiment including wave evolution is originally reported.
 The numerical results are validated using experimental data.
 A linear potential-flow model is used as a first approach to confirm the natural
frequencies and modes.

The remainder of this paper is organized as follows. Section 2 summarizes the proposed
coupled formulation. the fluid equations, the moving boundary techniques used, the solid
model, and the coupled strategy are presented in Subsections 2.1, 2.2, 2.3 and 2.4,
respectively. The experimental task is described in Section 3. The laboratory-scale
experimental layout is presented in Subsection 3.1, and the experimental data are
presented in Subsection 3.2. The validation of the implemented model is presented in
Section 4, where the finite-volume results are compared with results computed using a
model built within a finite element method (FEM) previously reported (Cruchaga et al.,
2001, 2006, 2013; Núñez Aedo et al., 2020), and with the experimental data for sloshing
(Subsection 4.1). The FSI analysis is reported in Subsection 4.2. In addition, a linear
potential solution to the FSI problem is presented in Subsection 4.3 because it is the
easiest approximate way to determine natural frequencies of the system for validating.
These frequencies can serve as a reference to evaluate the experimental data and the
results obtained in the most complex models. The FEM is also used to solve the problem
within the assumption of linear potential-flow theory (Sonzogni et al., 2002; Storti et al., Buoy
2000; D’Elía et al., 2002). Finally, concluding remarks are presented in Section 5. interacting
with waves
2. Numerical method for the fluid–solid interaction model
The study of the proposed buoy problem requires a description of the fluid-structure
interaction with free-surface evolution. The algorithms used to solve the problem were
implemented in an open-source code with high performance in computational fluid
dynamics (EDF R&D, 2017). The code was modified as follows:
 capture the free-surface evolution;
 describe a rigid-body motion; and
 include a fluid-structure interaction scheme.

2.1 Fluid dynamic model


The fluid dynamic is described using the incompressible Navier–Stokes system of
equations:
 
@v
r ð Þ
þ v  $ v  $  r ¼ b in X  Y (2.1)
@t

$  v ¼ 0 in X  Y

where r is the fluid density, v is the velocity, r is the stress tensor and b is the body force.
One contribution to the body force is the fluid weight (rg). Sloshing cases are solved in
a non-inertial frame of reference, where the observer moves with the sloshing tank, and
then a second contribution to the body force is the non-inertial force produced by the
acceleration of the frame of reference (r ar) as reported in Section 4.1. In addition,
X is the whole domain of analysis occupied by water and air (X ¼ Xw | Xa) and Y is the
time interval of interest. The Navier–Stokes system, equation (2.1), was solved using the
SIMPLEC strategy (Van Doormaal and Raithby, 1984; Aguerre et al., 2020) and Euler
implicit time integration scheme. Note that in the numerical context of ALE
formulations, the convective term is modified using a velocity c as (c · $)v, with c ¼ v 
vm, and vm is the velocity of the mesh. The mesh velocity is null for fixed meshes, i.e.
sloshing without the buoy, and this velocity is not zero when considering the buoy
because the mesh moves to fit the buoy at any time, as we explain later. Moreover, an
additional equation is used to determine the current free-surface position. The parts of
the domain occupied by water and air are identified according to these free-surface
positions (Subsection 2.2.1).

2.2 Moving boundaries


In problems with free-surface flows and fluid–structure interactions, there are two interfaces
of relevance: gas–liquid and solid–liquid. Of course, there is also a solid–gas interface, but in
most cases, the forces coming from the gas are much smaller, and the solution of the gas
phase is irrelevant. These interfaces must be represented either by a moving (deforming)
mesh or capturing techniques (using a level set or similar). In the moving mesh approach,
the mesh deforms in time following the interfaces. The mesh movement is considered by
adding the corresponding ALE terms. The advantage of this strategy is that there is a sharp
definition of the interfaces, and local mesh refinements close to those interfaces move with
HFF them, thus ensuring that there is a good representation of the physical phenomena (e.g.
boundary layers). The problem with this type of methodology is that for large displacements
of the interfaces, the mesh may collapse, requiring fully automatic remeshing of the region
and projection of the current state of the fluid from the old mesh to the new mesh. Capturing
techniques use a fixed mesh and have level-set functions that define the gas–liquid and
solid–liquid interfaces. In general, this methodology is less accurate but does not suffer from
the problem of mesh collapse.
In this study, a mixed strategy was adopted; a two-phase flow level-set strategy was
used to define and follow the air–water interface, whereas the mesh was deformed using an
ALE strategy to track the movement of the solid. The reason for choosing this mixed
strategy is that the body is rigid, and its movement is not very large; therefore, deforming
the mesh to track the body is relatively simple. Note that only the fluid domain (water and
air) is used in the computations. However, the movement of the free surface is more complex;
therefore, the level-set strategy is more robust.

2.2.1 Free-surface capturing: the level-set strategy. To solve the free-surface problem, i.e.
to follow the air–water interface, we adopted a capturing technique in which the interface
was described by the solution of an additional equation. In this two-phase flow scenario, a
level-set algorithm (Grotle et al., 2016; Gibou et al., 2018; Ge et al., 2018) is implemented to
capture the air–water interface over time.
As it is well-known, the scalar function w(x, t) of a leve set represents a continuous field
defined over the entire domain describing the water and air regions, and the surface between
them is defined with a given value of that function. We assume that w(x, t) > 0 if x [ Xw,
w(x, t) < 0 if x [ Xa, and w(x, t) ¼ 0 if x [ CFS. The level-set function w(x, t) is updated using
the following equation:
@w
þ ðv  $Þw ¼ 0 in X  Y (2.2)
@t
To avoid interface spreading over time, a renormalization technique previously
developed using the FEM (Battaglia et al., 2010), was implemented in the CS
context.
2.2.2 Moving mesh fitting the buoy contour. When the fluid interacts with the buoy, the
fluid domain varies its boundary according to the buoy movement. The buoy boundary Cb
is updated according to the solid velocity computed with the solid dynamic model described
in Subsection 2.3. To update the fluid mesh, an ad hoc technique is proposed for the present
discretization. The mesh is updated with a 1 D movement in the Z direction. The nodes of the
mesh are divided (according to Z) into three sections AB, BC and CD (Figure 1). Zone BC
includes the buoy. Therefore, when the buoy moves up, the entire BC zone moves rigidly
together with the body, zone AB contracts linearly, and zone CD expands linearly. The
reciprocal occurs when the buoy goes down. Note that this ad hoc technique only applies to
displacements that are not so large.
For larger mesh movements, other techniques need to be used. Numerical approaches
such as those proposed using a pseudoelastic formulation (Behr and Abraham, 2002; Charlot
et al., 2015; Castorrini et al., 2019), solving a Laplace problem (Löhner and Yang, 1996; Yong
and Baozeng, 2017) or minimizing the elemental distortion of the mesh (Lopez et al., 2007;
Battaglia et al., 2022) can be adapted either to gas–liquid or solid–liquid interface moving
mesh analyses.
Buoy
interacting
with waves

Figure 1.
Mesh motion
technique

2.3 Solid dynamic model


The buoy motion is computed according to the equilibrium equation:
Mbuoy€z buoy ¼ Ffluid  Mbuoy g (2.3)

where Mbuoy is the buoy mass, €z buoy is the buoy acceleration and Ffluid is the hydrodynamic
force acting instantaneously on the buoy. Equation (2.3) is presented only in the Z direction
because the motion of the buoy is restricted to this direction only, as will be described and
justified in Section 3.2. Ffluid is computed by integrating the pressure and viscous stress
tensor onð the buoy boundary Cb considering both fluids, air and water, that is
Ffluid ¼ ^ dS. As the mesh fits the buoy at any time, the hydrodynamic stresses are
sn
Cb
computed directly from the velocity and pressure fields.
To solve equation (2.3), the bNewmark second-order integration scheme is used:
  
1
ztþDt ¼ zt þ Dtz_ t þ Dt 2
 b €z t þ b€z tþDt (2.4)
2

z_ tþDt ¼ z_ t þ Dt½ð1  gÞ€z t þ g€z tþDt 

with b ¼ 0.25 and g ¼ 0.5.

2.4 Numerical coupled strategy


In summary, the implemented interaction strategy considers the following steps for
marching from time t to time t þ Dt in a partitioned way (Table 1).
Notice that at the beginning of each step, a predicted position of the buoy (Step 6) is used
to determine the current boundary of the domain where the fluid dynamics is computed.
HFF 1.
   
 Init structure (buoy)
z0buoy ; z_ 0buoy z buoy ; z_ buoy
 
2. v0 ; p0 ; w0  v; p; w  Init fluid and free surface

3. Xm 0
CMD z0buoy  Compute initial mesh displacement
according to Subsection 2.2.2
4. f or n ¼ 0; nstep  1 do
5. tn ¼ nDt; tnþ1 ¼ t þ Dt  Solve FSI in the interval [tn, tn þ 1]
6. nþ1;p
zbuoy znbuoy þ Dt z_ nbuoy  Predicted buoy position
7. nþ1   Update mesh according to Subsection 2.2.2
Xm CMD znbuoy

8. v ; p ; w
nþ1 nþ1 nþ1  Solve fluid dynamics including free

surfaces according to Subsections 2.1 and
CFD v n ; pn ; wn ; X nm ; X nþ1 m 2.2.1 [system of equations (2.1) and (2.2)]

9. nþ1
Ffluid Forces vnþ1 ; pnþ1  Compute hydrodynamic forces on the buoy
    according to Subsection 2.3
10. nþ1 _ nþ1
zbuoy ; z buoy CSD znbuoy ; z_ nbuoy ; Ffluid
nþ1  Solve solid dynamic according to
Subsection 2.3
11. end do

Notes: CMD = computational mesh dynamics operator; CFD = computational fluid dynamics code
including free surface computation; Forces() integrates hydrodynamic stress tensor on the buoy boundary;
Table 1. CSD = computation of the rigid solid dynamics
Numerical strategy Source: Table by authors

The predicted position from Step 6 increases the time precision of the algorithm (Storti et al.,
2008). After that, in Step 8, the fluid dynamics problem is solved using an ALE strategy (Hu
et al., 2001; Hron and Turek, 2006; Storti et al., 2008, 2012) with a given buoy position. This
nþ1
technique requires the mesh position X nm at time t n and nþ1
 X m at t  to evaluate the mesh
velocity (vm) used in the ALE strategy, e.g, vm ¼ X nþ1 n
m  X m =Dt. Finally, the solid
dynamic is solved (Step 10) as a correction step of the predicted buoy position and velocity.
The coupling between fluid and solid is explicit.

3. Experimental method
This section presents an experimental study on a buoy subjected to waves. This experiment
was conducted to collect data from wave evolution and buoy motion to validate the models
proposed in Section 2.

3.1 Problem layout


The laboratory layout is illustrated in Figure 2. The laboratory-scale experiment helps
obtain feasible and reproducible data under different imposed conditions and focuses the
study on the fluid–solid dynamics of the system. The main limitation when using reduced
scales is the lack of adjusting solid inertia. Nevertheless, the experiment helps the
understanding of the mechanism of motion under wave hydrodynamics effects. The tanks
and buoy geometries are shown in Figure 2(b) and 2(c), respectively. A rectangular tank
with an aspect ratio of 2:1 and height of 500 mm was filled with water up to a level of
300 mm. This filling level reproduces deep-water conditions, as confirmed by Núñez Aedo
et al. (2020). The buoy has a cylindrical shape with a height H ¼ 40 mm and radius R ¼
40 mm and ends with a hemisphere of radius R. The tank was mounted on a shaking table to
promote controlled motion. The imposed motion on the tank was set as xtank (t) ¼ Atank sin
(2pftank t) with an imposed amplitude Atank of 5 and 10 mm, and an imposed frequency ftank
Buoy
interacting
with waves

Figure 2.
Experimental setup

between [0.6  1.8] Hz. The movement of the shaking table starts with an initial ramp that
weights the imposed amplitude for the first three periods until the established periodic
movement. The shaking table motion is imposed during hundreds of periods to guarantee
the measurements at time-periodic regime. Regarding the waves and buoy motions, their
initial transient regime evolves during the first 20 or 30 periods of the imposed shaking
motion, and they could be slightly affected by the initial ramp of the shaking table.
Nevertheless, the initial ramp does not affect the results during the time-periodic regime,
these are the reported results. Moreover, filter was not used in the reported experimental
data. Consistent with previous work on wave evolution patterns without buoys
HFF (Cruchaga et al., 2013; Battaglia et al., 2018; Núñez Aedo et al., 2020), these imposed
parameters promote free surface evolution only in the XZ plane [Figure 2(b)] when the
frequencies are lower than the first natural frequency f1 ¼ 1.38 Hz. As the frequency of the
waves increases, the 3D free surface evolves, leading to larger effects on the system with
buoy. The buoy is attached with a string passed through a hole drilled along the vertical
axis of the body, and the string is pulled taut and tied to the top of the tank with a clip. This
fastening allows the movement in the vertical axis of the buoy and its rotation around that
axis. From experimental observations, rotations evolve together with 3D free-surface effects,
and they do not appear for free-surface planar modes. The motion induced by the flexibility
of the string was experimentally evaluated, as will be presented later, to validate the
hypothesis of vertical buoy motion used in the numerical model presented in Section 2.
The free-surface evolution was registered using ultrasonic sensors at a sampling rate of
400 Hz. The most significant advantage of ultrasonic sensors is that they can record large
amounts of data over time. However, they require some care for their use since they can fail
due to the reflection of the ultrasonic beam in the wall of the container or the distance
between the point of emission and the free surface. All of these factors are considered when
installing the sensors. Furthermore, the sensor records may not be reliable under certain
experimental conditions for which the free surface can be strongly distorted, such as sharp
waves or high 3D effects. Nevertheless, in the experiments we will report in this work, the
free surface does not have high distortions. In addition, the measurements obtained with the
sensors are confirmed with data provided by the camera using as post process a motion-
capturing technique for the interface. Information on the sensors and the verification
procedures applied to their measurements can be found in previously reported work (Castillo,
2019; Núñez Aedo et al., 2020; Núñez et al., 2022; Battaglia, 2022). Considering these facts, the
sensors were positioned 25 mm away from the tank walls, as shown in Figure 2(d). This
configuration was adopted to avoid wall interference with ultrasonic signals. These sensors
have a precision of 60.5 mm according to the data sheet from the supplier (Banner, 2022).
The buoy motion was registered using a camera at 120 fps with an image resolution of
1696  1710 pixels. Markers fixed to the buoy were used to track motion. Moreover, a
marker fixed to the tank helps confirm the motion imposed on the tank and synchronizes the
measurements. The videos were recorded during the periodic time regime of the system. The
error in measurements is bounded by 6 1 pix ¼ 60.2 mm, which is consistent with previous
works (Cruchaga et al., 2013; Castillo et al., 2019; Núñez Aedo et al., 2020).
Surface tension and meniscus effects could be an error source in the experiments due to
the relatively small scale. To quantify that error, the simple Jurin’s law applied to water at a
temperature of 25°C (surface tension 72.8  103 N/m and 75° for the contact angle with
acrylic) results in a meniscus of 0.14 mm for the distance where the ultrasonic sensors are,
being this value under the declared error for the ultrasonic sensors and for the data
capturing with the camera, i.e. surface tension and meniscus effects are included in the
ranges of error for the measurements.

3.2 Experimental data and analysis


To assess the influence of waves on the buoy, different tank motion conditions were used to
study the fluid–structure interaction effects, including the free-surface evolution. A
frequency sweep study was conducted to promote different wave patterns; that is, for a
given amplitude Atank the imposed frequency ftank was modified.
Figures 3, 4 and 5 show zooms from frames taken at certain instants and under different
shaking table conditions where the buoy motion is observed. Note that only the buoy motion
is evaluated from videos, and the free surface is slightly observed on them because the
Buoy
interacting
with waves

Figure 3.
Snapshots capturing
buoy positions for
Atank ¼ 5 mm and
ftank ¼1.0 Hz at
different instants
during the time-
periodic regime

experimental conditions are adjusted to have good marker definition. The free surface is
simultaneously registered by the ultrasonic sensors (outside the zoom shown in the figures),
synchronized with the camera and the shake table motion. The frames shown in Figure 3
were obtained at ftank ¼ 1 Hz for Atank ¼ 5 mm, and Figure 4 reports frames from the
experiment conducted at the same frequency but using Atank ¼ 10 mm. These figures
illustrate the cases in which no 3D effects evolve. Markers added to the buoy and tank are
used to track the objects when postprocessing images. From that tracking, the vertical and
horizontal positions of the buoy and the tank are obtained, as shown in Figure 6. This figure
shows that the horizontal motion of the buoy coincided with tank motion, and no rotation
effects were detected. In addition, no motion along Y has been detected, i.e. ybuoy % 0 is null
within the experimental error. Therefore, the buoy can be studied as a system with one
degree of freedom that can move vertically in such cases (as was assumed in Section 2). In
contrast, and for the completeness of the experimental observations, Figure 5 reports an
experiment for ftank ¼ 1.7 Hz and Atank ¼ 5 mm and shows a strong 3 D behavior not studied
in the present numerical work.
The experimental data for the fluid–buoy interaction are summarized in Figure 7, where
the maximum and minimum wave heights and buoy displacements are reported during the
periodic regime of the sloshing responses for each imposed tank motion (also listed in
Tables 2 and 3). Notice that at the current laboratory configuration, the wave amplitudes are
limited by the height of the tank. Due to this fact, the evolution of stationary waves is not
registered near or at natural frequency, because the waves exceed the height of the tank (a
HFF

Figure 4.
Snapshots capturing
buoy positions for
Atank ¼ 10 mm and
ftank ¼ 1.0 Hz at
different instants
during the time-
periodic regime

detailed study was made in Núñez et al., 2022). For this reason, there are no experimental
results in those frequencies.
The linearity of the wave behavior on the right (sensors S0 and S1) and left (sensors S2
and S3) sides of the buoy, and buoy dynamics can be evaluated using the ratio hmax/jhminj,
as shown in Tables 2 and 3. Note that all registered conditions denote nonlinear behavior;
that is, hmax/(jhminj) = 1. Moreover, greater differences between the ratios observed for
waves on the right and left sides of the buoy and the buoy itself appeared when 3D effects
were clearly observed in the buoy motion.
The experimental data are also reported in Section 4, along with numerical results. It is
worth mentioning that the pattern of waves without a buoy was also determined
experimentally in this work to verify the effect of the buoy on the waves and to validate the
proposed numerical model for the free surface without fluid–structure interaction effects as
reported in Subsection 4.1. Therefore, Figure 9 of that section shows the evolution of the
experimental free surface with its respective error associated with the experimental
precision during the periodic regime. The periodicity of the measurement is considered
2 min after the start of the imposed motion when the maximum and minimum amplitudes of
the sensors were constant (or with a dispersion lower than the experimental precision).
Furthermore, no 3D behavior was observed on the free surface; the pairs of sensors S0-S1
and S2-S3 recorded the same wave amplitude over time. Table 5 presents the experimental
data for the maximum and minimum wave amplitudes for different imposed frequencies,
compared with the numerical models based on the finite volume method (FVM) and FEM
Buoy
interacting
with waves

Figure 5.
Snapshots capturing
buoy positions for
Atank ¼ 10 mm and
ftank ¼ 1.7 Hz at
different instants
during the time-
periodic regime

Figure 6.
Buoy displacements
obtained from motion
capturing technique
(error bounded by
60.2 mm), sloshing at
Atank ¼ 5 mm and
ftank ¼ 1.0 Hz
HFF (details are provided in Section 4). In Subsection 4.2, the numerical simulation of the fluid–
buoy interaction is performed, and the wave heights and buoy motions are compared with
the experimental data. Subsection 4.3 presents the results computed with a model based on
linear potential theory, and they will serve as a second reference to evaluate the
experimental data, and the results obtained in the most complex models.

4. Numerical simulation and experimental validation


This section describes the validation of our proposed model. To this end, the free-surface
evolution during sloshing (without buoy) is first analyzed to validate the model
implemented by comparison with other numerical techniques (Cruchaga et al., 2013;
Battaglia et al., 2018; Núñez Aedo et al., 2020) and the experimental data reported here. In the
second step, the FSI model is applied to simulate the buoy experiment. Finally, a potential
flow model is included to evaluate the scope and limitations of the different approaches.

4.1 Sloshing analysis using the finite volume method


The implemented formulation used to describe the free surface was applied to solve the
sloshing problem in the current experimental configuration (Figure 2, without a buoy). The
geometry of the computational domain is shown in Figure 8, with dimensions of
400  650 mm in the x-and z-directions for the 2 D cases and 200 mm in the y-direction for
the 3D extended model.
The discretization is built by zones shown in Figure 8(c), with zP1 ¼ 280 mm and
zP2 ¼ 320 mm. For 2D cases, Mesh 1 has 93,600 cells with uniform size along X of Dx ¼
4 mm, and varying linearly along Z where in Zone 1 (from Dz ¼ 12 mm at the bottom to the
mesh size of Zone 2) and Zone 3 (from the mesh size of Zone 2 to Dz ¼ 6 mm at the top), uniform
sizes of Dz ¼ 0.3 mm are used in Zone 2 to capture the free-surface adequately. A second mesh
(Mesh 2) with half-sizes in all directions results in 374,000 cells, and the finer Dz ¼ 0.15 mm for
Zone 2. A 3D mesh is built based on Mesh 1 and taken a uniform cell distribution along Y of
Dy ¼ 22.2 mm. The time step used in the simulations is Dt ¼ 0.0002 s.
Slip boundary conditions were considered along the vertical walls, indicating that the
fluid slides perfectly during the time-periodic forced motion, as observed during the
experiments; that is, convection effects were dominant. Bottom wall conditions were studied
to assess the influence of such a wall, a weak influence on fluid motion was obtained when
slip or nonslip boundaries were used for the analyzed water level (Núñez Aedo et al., 2020).

Figure 7.
Frequency sweep
analyses at imposed
tank amplitudes of (a)
Atank ¼ 5 mm and (b)
Atank ¼ 10 mm (error
bounded by 60.5 mm
for sensors and
60.2 mm for zbuoy)
Frequency S0 S1 S2 S3 Buoy
Buoy
[Hz] [mm] [mm] [mm] [mm] [mm] interacting
with waves
hmax 60.5 mm zmax 60.2 mm
0.6 1.3 1.5 1.6 1.7 1.0
0.7 2.4 2.4 2.5 2.5 2.0
0.8 2.3 2.3 2.4 2.4 2.0
0.9 4.9 5.0 4.8 4.9 G*
1.0 7.0 7.1 7.3 7.2 5.3
1.1 11.7 11.7 12.3 12.3 8.8
1.2 22.8 22.8 23.1 23.1 18.2
1.6 34.1 34.1 34.1 34.1 18.4
1.7 23.4 23.4 21.5 21.6 2.3
hmin 60.5 mm zmin 60.2 mm
0.6 1.9 1.7 2.0 2.0 1.0
0.7 2.6 2.6 2.5 2.6 1.2
0.8 4.3 4.4 4.6 4.5 1.2
0.9 4.9 5.0 5.5 5.5 G*
1.0 7.3 7.6 8.5 8.3 6.4
1.1 11.4 11.4 12.2 11.3 8.9
1.2 20.9 20.9 21.7 21.7 17.2
1.6 26.1 26.0 28.1 28.1 22.6
1.7 19.6 17.8 18.2  18.2 19.2 z 
max
hmax zmax =jzmin j6 d
hmax =j hmin j 6d zmin
hmin
0.6 0.7 6 0.5 0.9 6 0.5 0.8 6 0.5 0.9 6 0.5 1.0 6 0.4
0.7 0.9 6 0.4 0.9 6 0.4 1.0 6 0.4 1.0 6 0.4 1.7 6 0.4
0.8 0.5 6 0.2 0.5 6 0.2 0.5 6 0.2 0.5 6 0.2 1.7 6 0.4
0.9 1.0 6 0.2 1.0 6 0.2 0.9 6 0.2 0.9 6 0.2 G*
1.0 1.0 6 0.1 0.9 6 0.1 0.9 6 0.1 0.9 6 0.1 0.82 6 0.06
1.1 1.0 6 0.09 1.0 6 0.09 1.0 6 0.08 1.1 6 0.09 1.0 6 0.05
1.2 1.1 6 0.05 1.1 6 0.05 1.1 6 0.05 1.1 6 0.05 1.0 6 0.02
1.6 1.3 6 0.05 1.3 6 0.05 1.2 6 0.04 1.2 6 0.04 0.8 6 0.02
1.7 1.2 6 0.06 1.3 6 0.06 1.2 6 0.06 ! 1.2 6 0.06 0.1 6 0.01
  Table 2.
hmax 1 hmax
Notes: Ratios with uncertainty d ¼ þ 0:5 mm and Wave amplitudes
hmin j hmin j j hmin j2
  ! hmax, hmin and ratios
zmax 1 zmax hmax/jhminj obtained
d ¼ þ 0:2 mm
zmin jzmin j jzmin j2 for Atank ¼ 5 mm
(G* means that the
Source: Table by authors buoy rotates)

The motion is driven by ar ðtÞ ¼ €x tank ðtÞ^i and gravity az ¼ g ¼ 9:81 m=s ^k, i.e.
Equation (2.1) is solved using b ¼ r(g  ar). The fluid properties are r1 ¼ 998.2 kg/m3, m1 ¼
103 Pa · s, r2 ¼ 1.225 kg/m3 and m2 ¼ 1.8 ·105 Pa · s for the density and viscosity of water
(1) and air (2), respectively. Due to the jump in properties, the shear stresses near the
interface could promote instability despite the lack of evidence of turbulence observed in the
experiments where slight disorder at the free surface physically evolved only in some cases
(Figure 5). To describe these situations, a simple mixing-length
 turbulencemodel was used.
pffiffiffiffiffiffiffiffiffiffiffiffi
In such cases, the viscosity adopted is mt ¼ min m þ lmix r 2« : «; mmax , where mmax is
2

the cutoff value and lmix is a characteristic length that measures a representative vortex
dimension and can be defined locally (i.e. mesh size depending) or globally (Tezduyar, 2003;
HFF Frequency S0 S1 S2 S3 Buoy
[Hz] [mm] [mm] [mm] [mm] [mm]

hmax 60.5 mm zmax 60.2 mm


0.75 5.9 5.9 6.1 6.3 0.5
0.85 7.8 7.9 7.9 8.2 4.3
1.0 14.9 15.0 14.6 14.6 9.6
1.1 23.2 23.8 28.1 28.1 17.8
1.6 48.4 48.0 48.4 47.4 G*
1.7 37.4 36.9 37.3 36.5 11.3
1.8 26.5 33.0 29.9 28.7 11.8
hmin 60.5 mm zmin 60.2 mm
0.75 5.0 5.2 5.6 5.1 0.5
0.85 7.5 7.7 7.5 7.6 3.4
1.0 13.7 13.9 14.4 14.5 8.7
1.1 19.0 19.3 21.0 21.0 9.6
1.6 35.0 35.6 34.8 34.9 G*
1.7 27.0 27.3 28.1 28.1 11.3
1.8 23.0 17.4 20.4  18.9 11.8  
hmax zmax
hmax =j hmin j 6d zmax =jzmin j6 d
hmin zmin
0.75 1.2 6 0.2 1.2 6 0.2 1.1 6 0.2 1.2 6 0.2 1.0 6 0.8
0.85 1.0 6 0.1 1.0 6 0.1 1.1 6 0.1 1.1 6 0.1 1.3 6 0.1
1.0 1.09 6 0.08 1.08 6 0.07 1.01 6 0.07 1.0 6 0.07 1.11 6 0.05
1.1 1.22 6 0.06 1.23 6 0.06 1.34 6 0.06 1.34 6 0.06 1.85 6 0.06
1.6 1.38 6 0.03 1.35 6 0.03 1.39 6 0.03 1.36 6 0.03 G*
1.7 1.38 6 0.04 1.36 6 0.04 1.33 6 0.04 1.30 6 0.04 3.19 6 0.07
1.8 1.15 6 0.05 1.89 6 0.08 1.46 6 0.06 ! 1.51 6 0.07 1.02 6 0.03
Table 3.  
hmax 1 hmax
Wave amplitudes Notes: Ratios with uncertainty d ¼ þ 0:5 mm and
hmin j hmin j j hmin j2
hmax, hmin and ratios   !
hmax/jhminj obtained zmax 1 zmax
d ¼ þ 0:2 mm
for atank ¼ 10 mm zmin jzmin j jzmin j2
(G* means that the
buoy rotates) Source: Table by authors

Cruchaga et al., 2009 and 2013; Battaglia et al., 2018, 2022). The parameters used in the
present analysis are mmax ¼ 0.1 Pa · s and lmix ¼ 0.15 m, according to their good numerical
behavior in the previous sloshing analyses reported.
The results obtained using Atank ¼ 5 mm and ftank ¼ 0.8 Hz are plotted in Figure 9
together with experimental data and those computed using a well-reported finite element
code (Cruchaga et al., 2001, 2006) successfully applied in Cruchaga et al. (2013) to the sloshing
analysis for a filling level of h ¼ 300 mm and in Núñez Aedo et al. (2020) for the study of deep
water conditions in stationary waves. The FEM study was performed in a 2D domain (no 3D
effects were registered in the experiments for this case). The boundary conditions and mesh
distribution are sketched in Figure 8(c), with zP1 ¼ 230 mm, zP2 ¼ 400 mm and
zP3 ¼ 500 mm. The element sizes are Dx ¼ 4 mm in the horizontal direction and Dz varies
as Dz ¼ 3 mm in Zones 1 and 3 and Dz ¼ 0.57 mm in Zone 2, resulting in 34,500 four-node
finite elements. The time step used in the FEM simulations is Dt ¼ 0.001 s.
Figure 9 shows that the FVM and FEM numerical results are in good agreement with the
experimental data (considering the experimental error bound); thus, both can be considered
valid for this level of precision.
Buoy
interacting
with waves

Figure 8.
Meshes (schematic
representation, only
illustrative) and
boundary conditions
used for FEM and
FVM models

This sloshing case for Atank ¼ 5 mm and ftank ¼ 0.8 Hz is used to perform a short mesh
convergence analysis when using the previously described 2D Meshes 1 and 2 and the 3D
mesh. Table 4 reports the computed results. Note that all predictions are within the
experimental range; due to this fact mesh 1 is used in the analyses. We assume that mesh
convergence is reached when the difference in wave amplitude between two successive
meshes are within the experimental error.
3D mesh case helps to illustrate some details on the computing time requirements.
Figure 10 reports the speed up and the CPU time per time step for three different runs
(using different numbers of cores) performed in the Pirayu cluster (CIMEC, 2023) with
processors Intel(R) Xeon(R) CPU E5-2650 with 20 cores. Note that the analysis
HFF exhibits good weak scalability with a computing rate of 33000 cells/(cores s) per time
step.
In Table 5, the maximum and minimum wave amplitudes (related to the free-surface
position at rest) are reported using other imposed motions of the tank. In such a table, the
experimental error is the one that corresponds to the ultrasonic sensors, and the numerical
error is on the order of the smallest size of the spatial discretization. These results also
showed a satisfactory fit between the numerical models and experiments.

Figure 9.
Free surface
evolution during the
time-periodic regime
for the sloshing
problem using Atank
¼ 5 mm and ftank ¼
0.8 Hz, comparison
between numerical
models and
experimental data
obtained from
ultrasonic sensors
(error bounded by
60.5 mm)

Table 4.
Comparison of hmax
and hmin between Model hmax [mm] hmin [mm]
different meshes for
sloshing obtained 2D Mesh 1 3.2 6 0.3 3.1 6 0.3
during the time- 2D Mesh 2 2.98 6 0.15 2.85 6 0.15
periodic regime at 3D mesh 3.2 6 0.3 3.1 6 0.3
Experiment 3.2 6 0.5 3.1 6 0.5
ftank ¼ 0.8 Hz for
atank ¼ 5 mm Source: Table by authors

80 cores
40 cores
20 cores
80 cores
Figure 10. 40 cores
20 cores
CPU time
requirements for the 0 1 2 3 0.00 0.30 0.60 0.90 1.20 1.50
3D sloshing Speed Up CPU/time step
tested case
Source: Figure by authors
With these sloshing results, the verification of the free surface model implemented in the Buoy
Saturne Code under deep-water conditions is considered satisfactory for this level of interacting
precision. The FSI strategy is validated in the following subsection.
with waves
4.2 Fluid–solid interaction analysis
In this subsection, a simulation of the experiment presented in Figure 2 is presented. The
domain was discretized, as shown in Figure 11. The mesh fits the buoy. A no-slip boundary
condition is adopted for the buoy wall. As previously mentioned, the buoy motion was
constrained to move only in the z-direction according to the experimental conditions. The
numerical study was performed using Atank ¼ 5 mm at different imposed frequencies, ftank.
As in the previous analyses, the discretization is built by zones shown in Figure 8(c). Cell
distribution along Z varies in Zone 1 from Dz ¼ 14 mm at the bottom to the mesh size of
Zone 2, and for Zone 3 from the mesh size of zone 2 to Dz ¼ 8.4 mm at the top; uniform sizes
of Dz ¼ 0.5 mm are used in Zone 2. The mesh has 1.2 millions of cells.
Other meshes were tested to perform a short analysis of the CPU time for FSI cases,
summarized in Figure 12. The figure shows the dependence of speed up and the average
CPU time per time step on the mesh size and number of cores. The runs have been made
using the Pirayu cluster (CIMEC, 2023) with processors Intel(R) Xeon(R) CPU E5-2650 with
20 cores. Note that the analysis exhibits good weak scalability with a computing rate of
57000 cells/(cores s) per time step.

Table 5.
Experiment [mm] FVM [mm] FEM [mm] Comparison of hmax
ftank [Hz] hmax hmin hmax hmin hmax hmin and hmin between
experimental data
60.05 60.5 60.5 60.3 60.3 60.6 60.6 and numerical
0.6 1.6 1.6 1.5 1.5 2.0 2.0 models for sloshing
0.7 2.5 2.4 2.2 2.1 2.3 2.4 obtained during the
0.8 3.2 3.2 3.2 3.1 3.1 3.4 time-periodic regime
0.9 4.7 4.6 4.2 4.5 4.7 5.0 at different imposed
1.0 5.5 8.0 5.8 5.4 7.7 6.5
frequencies ftank for
Source: Table by authors atank ¼ 5 mm

Figure 11.
Finite volume grid
used in the FSI
analysis
HFF Figure 13 shows the maximum and minimum wave heights and buoy displacements during
the time-periodic regime of sloshing responses at different imposed frequencies. In this
figure, the average values between sensors S0 and S1 (subindex R) and between sensors S2
and S3 (subindex L) are plotted. Moreover, wave amplitudes computed without the buoy are
also reported. Note that wave amplitudes increase when the buoy acts, i.e. higher waves
evolve. Table 6 summarizes the numerical and experimental data. The numerical
formulation only includes buoy vertical motion, while the experiments show a 3 D behavior
for ftank ¼ 0.9 Hz. This is the reason for the discrepancies observed at that ftank. The 3 D
behavior of the buoy is reported as G* in the tables, i.e. buoy rotates during the experiments.
Figure 14 shows the numerical results together with the experimental measurements for
the buoy position and free surface at control points S0, S1, S2 and S3 during the time-
periodic regime for sloshing at ftank ¼ 0.8 Hz. Note that the evolution of the free surface in S1
and S0 coincides, as well as in S2 and S3, showing that in this case, the wave evolves in a 2D
plane, that is, the XZ plane. Small differences were found between experiments and

20 cores 40 cores 80 cores 20 cores 40 cores 80 cores


4 7
CPU Time per step [s]
6
3 5
Speed Up

4
2
3
1 2
1
Figure 12. 0 0
CPU time 0.3 1.2 2 6.6 8.2 0.3 1.2 2 6.6 8.2
requirements for the Number of cells [in millons] Number of cells [in millons]
3D FSI cases
Source: Figure by authors

Figure 13.
Frequency sweep
analysis performed
using FSI
implemented in Code
Saturne (filled
symbols)
numerical predictions, being approximately 0.5 mm near the maximum and minimum Buoy
amplitudes of motion. interacting
with waves
4.3 Potential flow analysis
The problem was also studied using linear potential-flow analysis, as this type of
formulation is extensively used to solve wave-propagation problems (Tampier and Grueter,
2017). The computations are made in the frequency domain, i.e. for a given frequency, a
linear system is solved; hence, the steady harmonic solutions for wave amplitudes and buoy
positions are obtained. The main objective for running this model was to evaluate the results
reported in the previous sections.
Following Clauss et al. (1992), the fluid–structure interaction problem with rigid bodies
can be solved by assuming that the velocity gradients induced by the body in the
surrounding fluid are negligible. Clauss et al. (1992) determined that this condition arises
when 2 R/l < 0.2, as in this case. Therefore, we present the potential model results computed
using the numerical package PETSc-FEM (Sonzogni et al., 2002; Storti et al., 2000; D’Elía
et al., 2002) to add information for comparison. The 3D domain was discretized using a mesh
of 85000 hexahedral finite elements, as shown in Figure 15. The solutions are obtained for
imposed frequencies in the range [0.5  2.1] Hz and unitary amplitudes for Atank and Abuoy
because the superposition principle can be applied, i.e. the solution is computed considering
the solution of the sloshing problem for a fixed position of the buoy and, the solution due to
the buoy’s motion at a given fixed tank position.
Figure 16 summarizes the computed maximum wave amplitudes and buoy positions
(normalized with respect to Atank). The subindex R indicates the positions of sensors S0
and S1, and L indicates the positions of sensors S2 and S3 [Figure 2(d)]. Note that the
potential solution is linear, i.e. in the evolution of the wave heights and buoy motion
related to their equilibrium position, the maximum and minimum values coincide (i.e.
hmax ¼ jhminj).
Therefore, Figure 16 reports unique values for the variable (wave height or buoy motion)
at each frequency analyzed. Moreover, positive values indicate that the variable (wave
evolution or buoy position) is in phase with the tank motion (i.e. the maximum values are

ftank hLmax hLmin hRmax hRmin Zbuoymax Zbuoymin


60.05 60.5 60.5 60.5 60.5 60.5 60.5
[Hz] Analysis [mm] [mm] [mm] [mm] [mm] [mm]

0.8 FSI 2.8 3.1 3.1 3.0 2.9 2.0


Exp. 2.3 4.3 2.4 4.6 2.5 1.5
0.9 FSI 3.8 4.2 4.1 3.9 3.7 3.0 Table 6.
Exp. 4.9 5.0 4.9 5.5 G* G* Comparison of hmax
1.0 FSI 5.4 5.3 5.8 5.4 5.3 4.3 and hmin between the
Exp. 7.1 7.4 7.3 8.4 5.3 6.4 experimental data
1.1 FSI 9.0 7.8 8.9 8.0 8.6 6.2 and FSI numerical
Exp.       model obtained
1.2 FSI 12.1 9.8 12.4 10.0 12.0 9.9 during the time-
Exp.      
1.3 FSI 14.1 12.7 15.8 10.6 15.4 14.2
periodic regime at
Exp.       different imposed
frequencies ftank for
Source: Table by authors Atank ¼ 5 mm
HFF

Figure 14.
Comparison between
experimental and FSI
numerical (FVM
implemented into
Code Saturne CS)
results for
Atank ¼ 5 mm and
ftank ¼ 0.8 Hz

Figure 15.
Finite element mesh
used to compute the
coupled potential
flow model using the
(a) XZ plane, (b) YZ
plane and (c) half
view of the 3D
domain

reached when the maximum tank displacement occurs), and negative values are in
counterphase (i.e. minimum values are reached when maximum tank displacements occur).
In addition, the potential model is 3D but only reproduces symmetric waves with respect to
the plane Y ¼ 100 mm [Figure 2(b)].
The predictions obtained using the linear potential model are close to the average Buoy
experimental values, i.e. h ¼ (hmaxþ jhminj)/2, for the imposed amplitude of Atank ¼ 5 mm, interacting
as listed in Table 7, despite the experiments reporting nonlinear wave and buoy motion
behaviors (see Figure 7(a) and Table 2). Although the potential solutions did not match the
with waves
average experimental values when the tank motion increased to Atank ¼ 10 mm, (as shown
in Table 8), this is because the potential solutions are based on a linear approach, while the
experiments show strong nonlinear behavior [as shown in Figure 7(b) and Table 3].
These results justify the application of more complex models to describe problems where the
waves are not linear. It also confirms that under conditions of low nonlinearity, the
responses can be compared if average values are considered.
As is known, in any induced oscillatory motion, the amplitudes of the system are
amplified when the imposed frequency approaches the natural frequency. Taking this into
account, in Figure 16, three natural frequencies can be observed. The first and third, f1 ¼
1.324 Hz and f3 ¼ 1.925 Hz, were closer to the free-surface
 natural
 frequencies related to the
first and second planar modes, that is, vi ¼ gp L tanh p L h for i ¼ 1,2. The experimental
2 i i

(Figure 7) and numerical (Figure 16) results show an asymptotic trend at the first natural
frequency, confirming its occurrence in the observations. The second frequency that appears

Figure 16.
Linear potential-flow
model, numerical
responses of
maximum wave
amplitudes at right
(S0 and S1) and left
(S2 and S3) to the
buoy and the
maximum vertical
buoy’s position

Frequency [Hz] Experimental [mm] Potential flow [mm] Table 7.


ftank 6 0.05 h R 60:5 h L 60:5 Z buoy 60:2 hR hL Zbuoy Comparison between
numerical results of
0.6 1.8 1.6 1.0 1.8 1.6 1.2 the linear potential
0.7 2.5 2.5 1.6 2.7 2.3 1.9 flow model and
0.8 3.6 3.3 1.6 3.8 3.4 2.9
0.9 5.2 4.9 G* 5.6 5.0 4.6
average experimental
1.0 7.8 7.3 5.8 8.3 7.5 7.7 data for Atank ¼
5 mm (G* means that
Source: Table by authors the buoy rotates)
HFF in Figure 16, f2 ¼ 1.64 Hz, can be attributed to the natural frequency of the buoy. This result
also justifies the behavior described in Figure 7(b) near 1.6 Hz. For frequencies near 1.6 and
higher arose from the development of 3D effects confirmable by the rotation of the buoy at
that frequency.

5. Conclusions
This work presents a numerical formulation implemented in an open source code and the
validation of its predictions for the simulation of a fluid–rigid interaction problem on a
laboratory scale that includes wave effects. The numerical formulation was developed
within the framework of the finite volume method. Specifically, a free-surface level set
capturing method and a coupled strategy between the fluid and rigid solid equations were
implemented and tested. The experiment consisted of a tank partially filled with water, with
a floating buoy that could only move vertically and rotate. The tank is subjected to
controlled motions using a shake table. The free-surface evolution is registered using
ultrasonic sensors, and a motion capturing technique is used to track the buoy.
The computed results were verified using the experimental data. First, the free-surface
evolution is validated. Good agreement was found between the finite volume predictions and
those computed using a very different finite element numerical strategy. The computed free-
surface evolution satisfactorily matched the experimental data. Subsequently, a full-fluid
buoy numerical model was built, and a good match between the numerical results and
experiments was found when no strong 3D effects appeared in the buoy motion.
The numerical predictions follow the experimental trends. Clearly, the free surface
evolves in a nonlinear way in all the analyzed cases, although when its behavior is in the
plane of movement, the maximum and minimum heights of the waves do not have the same
value. The 3D effects (i.e. when the waves in the planes in front and back of the tank differ)
evolve for stronger imposed conditions where the buoy tends to have rotational
displacements.
Potential flow analysis is a standard tool that helps in understanding the physics of
systems such as the one proposed here. Therefore, we include an analysis in this context.
Their results are comparable to average numerical and experimental data only for weakly
nonlinear waves. The frequencies obtained qualitatively agree with those derived from the
finite volume simulations and experimental observations.
Although we found a numerical limitation in the description of the buoy rotation modes,
the present results support the future extension of the numerical methodology to study more
complex behaviors, an aspect that needs to be explored.

Frequency [Hz] Experimental [mm] Potential flow [mm]


ftank 6 0.05 h R 60:5 h L 60:5 Z buoy 60:2 hR hL Zbuoy
Table 8.
Comparison between 0.75 5.6 5.8 0.5 6.4 5.6 4.7
numerical results of 0.85 7.7 7.8 3.8 9.2 8.2 7.4
the linear potential 1.0 14.3 14.5 9.1 16.5 15.1 15.4
flow model and 1.1 21.3 24.6 13.7 26.2 24.5 28.3
1.6 41.8 41.4 G* 74.3 31.7 322.6
average experimental 1.7 32.2 32.5 23.7 93.3 69.3 163.7
data for Atank ¼ 1.8 25.0 24.5 11.9 40.3 38.3 44.3
10 mm (G* means
that the buoy rotates) Source: Table by authors
References Buoy
Aguerre, H.J., Venier, C.M., Pairetti, C.I., Santiago Marquez, D. and Nigro, N.M. (2020), “A simple-based interacting
algorithm with enhanced velocity corrections: the complex method”, Computers and Fluids,
Vol. 198, p. 104396.
with waves
Amaechi, C., Wang, F. and Ye, J. (2021), “Mathematical modelling of bonded marine hoses for single
point mooring (spm) systems, with catenary anchor leg mooring (calm) buoy pplication—a
review”, Journal of Marine Science and Engineering, Vol. 9 No. 11, p. 1179.
Bai, W., Liu, X. and Koh, C.G. (2015), “Numerical study of violent lng sloshing induced by realistic ship
motions using level set method”, Ocean Engineering, Vol. 97, pp. 100-113.
Baiges, J., Codina, R., Pont, A. and Castillo, E. (2017), “An adaptive fixed-mesh ale method for free
surface flows”, Computer Methods in Applied Mechanics and Engineering, Vol. 313,
pp. 159-188.
Banner (2022), “Datasheet sensor”, available at: https://info.bannerengineering.com/cs/groups/public/
documents/literature/110738.pdf (accessed 13 October 2022).
Battaglia, L., Storti, M.A. and D’Elía, J. (2010), “Bounded renormalization with continuous penalization
for level set interface-capturing methods”, International Journal for Numerical Methods in
Engineering, Vol. 84 No. 7, pp. 830-848.
Battaglia, L., Lopez, E.J., Cruchaga, M.A., Storti, M.A. and D’Elía, J. (2022), “Mesh-moving arbitrary
Lagrangian–Eulerian three-dimensional technique applied to sloshing problems”, Ocean
Engineering, Vol. 256, p. 111463.
Battaglia, L., Cruchaga, M., Storti, M., D’Elía, J., Núñez Aedo, J. and Reinoso, R. (2018), “Numerical
modelling of 3d sloshing experiments in rectangular tanks”, Applied Mathematical Modelling,
Vol. 59, pp. 357-378.
Bazilevs, Y., Hsu, M.-C., Kiendl, J., Wüchner, R. and Bletzinger, K.-U. (2011), “3d simulation of wind
turbine rotors at full scale. Part ii: Fluid–structure interaction modeling with composite blades”,
International Journal for Numerical Methods in Fluids, Vol. 65 Nos 1/3, pp. 236-253.
Behr, M. and Abraham, F. (2002), “Free surface flow simulations in the presence of inclined walls”,
Computer Methods in Applied Mechanics and Engineering, Vol. 191 Nos 47/48,
pp. 5467-5483.
Cai, Z., Topa, A., Djukic, L.P., Herath, M.T. and Pearce, G.M.K. (2021), “Evaluation of rigid body force in
liquid sloshing problems of a partially filled tank: Traditional cfd/sph/ale comparative study”,
Ocean Engineering, Vol. 236, p. 109556.
Calderer, A., Kang, S. and Sotiropoulos, F. (2014), “Level set immersed boundary method for coupled
simulation of air/water interaction with complex floating structures”, Journal of Computational
Physics, Vol. 277, pp. 201-227.
Castillo, E., Cruchaga, M.A., Baiges, J. and Flores, J. (2019), “An oil sloshing study: adaptive fixed-mesh
ale analysis and comparison with experiments”, Computational Mechanics, Vol. 63 No. 5,
pp. 985-998.
Castorrini, A., Corsini, A., Rispoli, F., Takizawa, K. and Tezduyar, T.E. (2019), “A stabilized ALE
method for computational fluid–structure interaction analysis of passive morphing in
turbomachinery”, Mathematical Models and Methods in Applied Sciences, Vol. 29 No. 5,
pp. 967-994.
Celebi, M. and Akyildiz, H. (2002), “Nonlinear modeling of liquid sloshing in a moving rectangular
tank”, Ocean Engineering, Vol. 29 No. 12, pp. 1527-1553.
Charlot, L., Etienne, S., Hay, A., Pelletier, D. and Garon, A. (2015), “High-order time integrators for front-
tracking finite-element analysis of viscous free-surface flows”, International Journal for
Numerical Methods in Fluids, Vol. 77 No. 11, pp. 668-693.
CIMEC (2023), available at: https://cimec.org.ar/c3/pirayu/Pirayucluster (accessed June 2023).
HFF Clauss, G., Lehmann, E. and Östergaard, C. (1992), Offshore Structures: volume I: conceptual Design and
Hydromechanics, Springer, New York, NY.
Connell, K.O., Thiebaut, F., Kelly, G. and Cashman, A. (2018), “Development of a free heaving owc
model with non-linear pto interaction”, Renewable Energy, Vol. 117, pp. 108-115.
Cruchaga, M., Celentano, D. and Tezduyar, T. (2001), “A moving Lagrangian interface technique for
flow computations over fixed meshes”, Computer Methods in Applied Mechanics and
Engineering, Vol. 191 Nos 6/7, pp. 525-543.
Cruchaga, M.A., Celentano, D.J. and Tezduyar, T.E. (2009), “Computational modeling of the collapse of
a liquid column over an obstacle and experimental validation”, Journal of Applied Mechanics,
Vol. 76 No. 2.
Cruchaga, M., Löhner, R. and Celentano, D. (2012), “Experimental and numerical analysis of a sphere
falling into a viscous fluid”, International Journal for Numerical Methods in Fluids, Vol. 69 No. 9,
pp. 1496-1521.
Cruchaga, M., Celentano, D., Breitkopf, P., Villon, P. and Rassineux, A. (2006), “A front remeshing
technique for a Lagrangian description of moving interfaces in two-fluid flows”, International
Journal for Numerical Methods in Engineering, Vol. 66 No. 13, pp. 2035-2063.
Cruchaga, M., Celentano, D., Breitkopf, P., Villon, P. and Rassineux, A. (2010), “A surface remeshing
technique for a Lagrangian description of 3d two-fluid flow problems”, International Journal for
Numerical Methods in Fluids, Vol. 63, pp. 415-430.
Cruchaga, M.A., Reinoso, R.S., Storti, M.A., Celentano, D.J. and Tezduyar, T.E. (2013), “Finite element
computation and experimental validation of sloshing in rectangular tanks”, Computational
Mechanics, Vol. 52 No. 6, pp. 1301-1312.
D’Elía, J., Storti, M.A. and Idelsohn, S.R. (2002), “Applied hydrodynamic wave-resistance computation
by Fourier transform”, Ocean Engineering, Vol. 29 No. 3, pp. 261-278.
Devolder, B., Stratigaki, V., Troch, P. and Rauwoens, P. (2018), “Cfd simulations of floating point
absorber wave energy converter arrays subjected to regular waves”, Energies, Vol. 11 No. 3,
p. 641.
Ducassou, B., Núñez, J., Cruchaga, M. and Abadie, S. (2017), “A fictitious domain approach based on a
viscosity penalty method to simulate wave/structure interaction”, Journal of Hydraulic Research,
Vol. 55 No. 6, pp. 847-862.
EDF R&D (2017), “Code Saturne version 5.0.0 practical user’s guide”.
Ge, Z., Loiseau, J., Ch, Tammisola, O. and Brandt, L. (2018), “An efficient mass-preserving interface-
correction level set/ghost fluid method for droplet suspensions under depletion forces”, Journal
of Computational Physics, Vol. 353, pp. 435-459.
Gibou, F., Fedkiw, R. and Osher, S. (2018), “A review of level-set methods and some recent
applications”, Journal of Computational Physics, Vol. 353, pp. 82-109.
Glowinski, R., Pan, T.-W., Hesla, T.I., Joseph, D.D. and Periaux, J. (2001), “A fictitious domain approach
to the direct numerical simulation of incompressible viscous flow past moving rigid bodies:
application to particulate flow”, Journal of Computational Physics, Vol. 169 No. 2, pp. 363-426.
Gonzalez, F.A., Cruchaga, M.A. and Celentano, D.J. (2017), “Analysis of flow past oscillatory cylinders
using a finite element fixed mesh formulation”, Journal of Fluids Engineering, Vol. 139 No. 8,
pp. 081202.
Gonzalez, F.A., Bustamante, J.A., Cruchaga, M.A. and Celentano, D.J. (2019), “Numerical study of flow
past oscillatory square cylinders at low reynolds number”, European Journal of Mechanics-B/
Fluids, Vol. 75, pp. 286-299.
Grotle, E.L., Bihs, H., Pedersen, E. and Æsøy, V. (2016), “Cfd simulations of non-linear sloshing in a
rotating rectangular tank using the level set method”, International Conference Offshore and
Arctic Engineering. American Society of Mechanical Engineers. V002T08A019.
Hron, J. and Turek, S. (2006), “A monolithic fem/multigrid solver for an ale formulation of fluid- Buoy
structure interaction with applications in biomechanics”, Fluid-Structure Interaction, Springer,
New York, NY, Vol. 53, pp. 146-170.
interacting
Hu, H.H., Patankar, N.A. and Zhu, M.Y. (2001), “Direct numerical simulations of fluid–solid systems
with waves
using the arbitrary Lagrangian–Eulerian technique”, Journal of Computational Physics, Vol. 169
No. 2, pp. 427-462.
Idelsohn, S.R., Marti, J., Limache, A. and Oñate, E. (2008), “Unified Lagrangian formulation for elastic
solids and incompressible fluids: application to fluid– structure interaction problems via the
pfem”, Computer Methods in Applied Mechanics and Engineering, Vol. 197 Nos 19/20,
pp. 1762-1776.
Jafari, M., Babajani, A., Hafezisefat, P., Mirhosseini, M., Rezania, A. and Rosendahl, L. (2018),
“Numerical simulation of a novel ocean wave energy converter”, Energy Procedia, Vol. 147,
pp. 474-481.
Kleefsman, K.M.T., Fekken, G., Veldman, A.E.P., Iwanowski, B. and Buchner, B. (2005), “A volume-of-
fluid based simulation method for wave impact problems”, Journal of Computational Physics,
Vol. 206 No. 1, pp. 363-393.
Koh, C.G., Luo, M., Gao, M. and Bai, W. (2013), “Modelling of liquid sloshing with constrained floating
baffle”, Computers and Structures, Vol. 122, pp. 270-279.
Lo, D.C., Hsieh, C.-M. and Young, D.L. (2014), “An embedding finite element method for viscous
incompressible flows with complex immersed boundaries on cartesian grids”, Engineering
Computations, Vol. 31 No. 4, pp. 656-680.
Löhner, R. and Yang, C. (1996), “Improved ALE mesh velocities for moving bodies”, Communications in
Numerical Methods in Engineering, Vol. 12 No. 10, pp. 599-608.
Löhner, R., Baum, J.D., Mestreau, E., Sharov, D., Charman, Ch. and Pelessone, D. (2004), “Adaptive
embedded unstructured grid methods”, International Journal for Numerical Methods in
Engineering, Vol. 60 No. 3, pp. 641-660.
Lopez, E.J., Nigro, N.M., Storti, M.A. and Toth, J.A. (2007), “A minimal element distortion strategy for
computational mesh dynamics”, International Journal for Numerical Methods in Engineering,
Vol. 69 No. 9, pp. 1898-1929.
Mittal, R. and Iaccarino, G. (2005), “Immersed boundary methods”, Annual Review of Fluid Mechanics,
Vol. 37 No. 1, pp. 239-261.
Núñez Aedo, J., Cruchaga, M. and Castillo del Barrio, E. (2020), “Study on the dependence with the
filling level of the sloshing wave pattern in a rectangular tank”, Physics of Fluids, Vol. 32 No. 1.
Núñez, J., Cruchaga, M. and Tampier, G. (2022), “Wave analysis based on genetic algorithms using data
collected from laboratories at different scales”, European Journal of Mechanics – B/Fluids,
Vol. 95, pp. 231-239.
Orona, J.D., Zorrilla, S.E. and Peralta, J.M. (2018), “Sensitivity analysis using a model based on
computational fluid dynamics, discrete element method and discrete phase model to study a
food hydrofluidization system”, Journal of Food Engineering, Vol. 237, pp. 183-193.
Pozzi, N., Bracco, G., Passione, B., Sirigu Sergej, A., Vissio, G., Mattiazzo, G. and Sannino, G. (2017),
“Wave tank testing of a pendulum wave energy converter 1: 12 scale model”, International
Journal of Applied Mechanics, Vol. 9, p. 1750024.
Ransley, E.J., Brown, S.A., Hann, M., Greaves, D.M., Windt, C., Ringwood, J., Davidson, J., Schmitt, P.,
Yan, Sh, Wang. and J.X., Others. (2021), “Focused wave interactions with floating structures: a
blind comparative study”, Proceedings of the Institution of Civil Engineers-Engineering and
Computational Mechanics, Vol. 174 No. 1, pp. 46-61.
Robinson-Mosher, A., Schroeder, C. and Fedkiw, R. (2011), “A symmetric positive definite formulation
for monolithic fluid structure interaction”, Journal of Computational Physics, Vol. 230 No. 4,
pp. 1547-1566.
HFF Sarrate, J., Huerta, A. and Donea, J. (2001), “Arbitrary Lagrangian–Eulerian formulation for fluid–rigid
body interaction”, Computer Methods in Applied Mechanics and Engineering, Vol. 190 Nos 24/25,
pp. 3171-3188.
Schillaci, E., Favre, F., Troch, P. and Oliva, A. (2021), “Numerical simulation of fluid structure
interaction in free-surface flows: the wec case”, Journal of Physics: Conference Series, v2116, IOP
Publishing.
Sonzogni, V.E., Yommi, A.M., Nigro, N.M. and Storti, M.A. (2002), “A parallel finite element program on
a beowulf cluster”, Advances in Engineering Software, Vol. 33 Nos 7/10, pp. 427-443.
Storti, M.A., Garelli, L. and Paz, R.R. (2012), “A finite element formulation satisfying the discrete
geometric conservation law based on averaged Jacobians”, International Journal for Numerical
Methods in Fluids, Vol. 69 No. 12, pp. 1872-1890.
Storti, M.A., Nigro, N.M., Paz, R.R. and Dalcín, L. (2008), “Strong coupling strategy for fluid structure
interaction problems in supersonic regime via fixed point iteration”, Journal of Sound and
Vibration, Vol. 320 No. 4-5, pp. 859-877.
Storti, M.A., D’Elía, J., Bonet Chaple, R.P., Nigro, N.M. and Idelsohn, S.R. (2000), “The dnl absorbing
boundary condition: applications to wave problems”, Computer Methods in Applied Mechanics
and Engineering, Vol. 182 Nos 3/4, pp. 483-498.
Stratigaki, V., Troch, P., Stallard, T., Forehand, D., Kofoed, J.P., Folley, M., Benoit, M., Babarit, A. and
Kirkegaard, J. (2014), “Wave basin experiments with large wave energy converter arrays to
study interactions between the converters and effects on other users in the sea and the coastal
area”, Energies, Vol. 7 No. 2, pp. 701-734.
Tampier, G. and Grueter, L. (2017), “Hydrodynamic analysis of a heaving wave energy converter”,
International Journal of Marine Energy, Vol. 19, pp. 304-318.
Tezduyar, T.E. (2003), “Computation of moving boundaries and interfaces and stabilization
parameters”, International Journal for Numerical Methods in Fluids, Vol. 43 No. 5,
pp. 555-575.
Tran, T.T., Krueger, A.M., Gunawan, B. and Alam, M.-R. (2019), “Predicting the dynamic
characteristics of a fully submerged wave energy converter subjected to a power take off failure
using a high-fidelity computational fluid dynamics model”, Tech. Rep. Sandia National Lab.
(SNL-NM), Albuquerque, NM.
Van Doormaal, J.P. and Raithby, G.D. (1984), “Enhancements of the simple method for predicting
incompressible fluid flows”, Numerical Heat Transfer, Vol. 7 No. 2, pp. 147-163.
Wang, W.-H., Ran, X.-M., Zhao, Z.-H. and Huang, Y. (2021), “Comparative analysis on various
components of heave damping for sandglass-type floating body”, Ocean Engineering, Vol. 221,
p. 108555.
Wang, W.-H., Du, Y.-Z., Wang, L.-L., Yao, Y.-X., Gao, H. and Huang, Y. (2017), “Experimental analysis
on behaviour in waves for sandglass-type floating body”, Ships and Offshore Structures, Vol. 12
No. 3, pp. 433-441.
Wiegard, B., Radtke, L., König, M., Abdel-Maksoud, M. and Düster, A. (2019), “Simulation of the fluid-
structure interaction of a floating wind turbine”, Ships and Offshore Structures, Vol. 14 No. sup1,
pp. 207-218.
Yang, J., Preidikman, S. and Balaras, E. (2008), “A strongly coupled, embedded-boundary method for
fluid–structure interactions of elastically mounted rigid bodies”, Journal of Fluids and
Structures, Vol. 24 No. 2, pp. 167-182.
Yong, T. and Baozeng, Y. (2017), “Simulation of large-amplitude three-dimensional liquid sloshing in
spherical tanks”, AIAA Journal, Vol. 55 No. 6, pp. 2052-2059.
Zamora, E., Battaglia, L., Storti, M., Cruchaga, M. and Ortega, R. (2019), “Numerical and experimental
study of the motion of a sphere in a communicating vessel system subject to sloshing”, Physics of
Fluids, Vol. 31 No. 8.
Zhang, Y.-X., Wan, D., Ch. and Takanori, H. (2014), “Comparative study of mps method and level-set Buoy
method for sloshing flows”, Journal of Hydrodynamics, Vol. 26 No. 4, pp. 577-585.
interacting
Zhao, Y. and Chen, H.-C. (2015), “Numerical simulation of 3d sloshing flow in partially filled lng tank
using a coupled level-set and volume-of-fluid method”, Ocean Engineering, Vol. 104, pp. 10-30. with waves
Zou, S. and Abdelkhalik, O. (2021), “A numerical simulation of a variable-shape buoy wave energy
converter”, Journal of Marine Science and Engineering, Vol. 9 No. 6, p. 625.

Corresponding author
Marcela A. Cruchaga can be contacted at: marcela.cruchaga@usach.cl

For instructions on how to order reprints of this article, please visit our website:
www.emeraldgrouppublishing.com/licensing/reprints.htm
Or contact us for further details: permissions@emeraldinsight.com

You might also like