0% found this document useful (0 votes)
31 views18 pages

Manuscript

COMPARISON OF OVERSET MESH WITH MORPHING MESH: FLOW OVER A FORCED OSCILLATING AND FREELY OSCILLATING 2D CYLINDER

Uploaded by

Sahil Jawa
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
31 views18 pages

Manuscript

COMPARISON OF OVERSET MESH WITH MORPHING MESH: FLOW OVER A FORCED OSCILLATING AND FREELY OSCILLATING 2D CYLINDER

Uploaded by

Sahil Jawa
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 18

Volume 2 [Full Papers], Pages 13–30

ISSN: 2753-8168

COMPARISON OF OVERSET MESH WITH MORPHING MESH: FLOW OVER A


FORCED OSCILLATING AND FREELY OSCILLATING 2D CYLINDER

MICHAEL ALLETTO

Private person
Email address: [email protected]

DOI: 10.51560/ofj.v2.47
Version(s): OpenFOAM® v2006
Repo: -

Abstract. Fluid structural interactions involve temporal change of the boundaries of the computational
domain as a result of the fluid forces acting on the solid structure. In order to take into account this
change, OpenFOAM offers two different methods: the overset method and the morphing mesh method.
The morphing mesh method takes into account the boundary movement by changing the topology of
the mesh but keeping the connectivity between cells unchanged. The overset methods consists in non
overlapping meshes which are moving relative to a background mesh. Usually the meshes do not change
the topology. The coupling between meshes is achieved by interpolation of the variables between the
meshes. Both methods are compared with each other and reference experiments and simulation by
taking a two-dimensional laminar flow around a cylinder. The results are found to agree reasonable well
with the references. Both mesh types give similar results for similar number of cells.

1. Introduction
Fluid structure interaction is associated with a multitude of physical phenomena: the stability and
response of aircraft wings, the flow through arteries, the vibration of heat exchanger, the vibration of
compressor and turbine blades and many more [1]. OpenFOAM offers two type of solvers two deal with
the deformation of the structure (actually the deformation of a bounding surface): morphing mesh solvers
and overset solvers. The class of morphing mesh solvers update the position of the vertices forming the
cell in order to follow the motion of a boundary surface, i.e., the topology of the mesh is changed. The
connectivity between the cells is kept unchanged. See [2] for an overview. Regarding the overset method,
there is a static (not moving) background mesh and one or more meshes attached to the moving bodies.
Usually, both meshes do not change their topology. In order to couple the solutions of both meshes, the
variables are interpolated between the cells closed to the boundary of the moving mesh and the static
background mesh. For a description of the method see e.g. [3, 4].
Fluid structure problems are characterized by a mutual influence of the fluid phase and the deforming
solid: The fluid exerts a force on the solid via the pressure and surface stress distribution on the solid
boundary. This forces lead to a deformation or position change of the surface. The change of the surface
shape or position influences the flow motion. In order to capture the phenomena associated with moving
boundaries accurately, it is important to carefully validate both fluid and structure solvers first separately
and after that both together. In this spirit, we will first analyze the solvers used in OpenFOAM to solve
fluid-structure interaction separately and then together. The chosen configuration consists in a circular
cylinder in a two dimensional laminar flow. Despite its simplicity and the reduced computational costs
(we need only a 2D mesh) this configuration is associated with a variety of interesting phenomena. For
this reason it is an ideal test case to have a first impression about the accuracy and performance of the
two method implemented in OpenFOAM to deal with displacing boundaries.
The first step in this validation is to look at the structure part of OpenFOAM solely. OpenFOAM
uses a six degree of freedom solver to advance the location and orientation of a rigid body subject to fluid
forces and restrains like springs and damper in time. The solver solves a set of ordinary differential which
are derived from Newtons second law. For a case without fluid forces analytical solution exist which can
be compared with the numerical results.

Received: 16 July 2021, Accepted: 12 January 2022, Published: 24 January 2022

© 2022 The Authors. This work is an open access article under the CC BY-SA 4.0 license

13
14 MICHAEL ALLETTO

As second step we will examine the inline oscillatory motion of a cylinder in a uniform flow. Inline
means that the cylinder is oscillating parallel to the incoming flow by a prescribed motion. This kind of
configuration is often used to validate solvers aiming to simulate large body motions (see e.g. [5, 6, 7]).
In the works of Su et al. [5] and Liau et al. [6] the inline oscillation cylinder was one of the test cases
to validate their immersed boundary method code. Chandar et al. [7] used this configuration (among
others) to validate the OpenFOAM based overset method. The reason is that that despite its simplicity
this configuration is associated with a variety of complex physical phenomena. One of this phenomena is
frequency lock in: When the frequency of the oscillating cylinder is close to twice the shedding frequency
of the stationary cylinder fst , the frequency of the vortices shed became equal to the one of the cylinder.
The amplitude of the cylinder motion which is required to maintain the lock in regime is increasing with
departure from twice the shedding frequency of a fixed cylinder [8]. Another interesting phenomena is
the increase of the lift coefficient Cl when the cylinder is oscillating at 2fst [9]. For frequencies which
are either lower or higher than 2fst , Cl is much lower [9]. Leontini et al. [10] made simulations with an
in line oscillating cylinder at a Reynolds number of Re = 175. For this Re number the flow around a
fixed cylinder is two dimensional. For an oscillation frequency fc of 2fst and an amplitude of 65% of the
cylinder diameter the number of peaks in the frequency spectrum of the lift force increases considerably.
For lower amplitudes of the oscillation only a few peaks are present in the frequency spectrum. Regarding
OpenFOAM simulation of this particular configuration, only one work is known to the author of the
present article. It is the work of Chandar et al. [7]. They used OpenFOAM together with an in-house
developed overset library OPERA. They focused the attention on the numerical oscillations of the lift
and drag generated by the overset interpolation rather than performing a comparison with reference
experiments or simulations. Unfortunately no work was found using this test case with the morphing
mesh method implemented in OpenFOAM.
In the third part of our validation study will examine the capability of OpenFOAM to simulate the
vortex induced vibration on a circular cylinder in a 2D laminar configuration. The cylinder is free to
vibrate in stream wise and in transverse direction. This kind of flow configuration is often used to validate
codes which aim to describe the interaction of a fluid with a solid structure (see e.g. [11, 12]). Besides
that, this configuration is used to answer fundamental questions of the basic physical mechanisms playing
a role in the interaction between the flow and the solid body (see e.g. [13]). Recent studies concentrate
on the phenomena which lead to large amplitude oscillation of the cylinder and the synchronization
of the vortex shedding frequency with the natural frequency of the spring-mass-damper system (see
[14, 15]). Regarding OpenFOAM specific simulation of this test case, the author could only find the PhD
thesis of Conceição [16]. He performed a systematic validation of the one and two degrees of freedom
vortex induced simulation of a circular cylinder using the solver pimpleFoam together with a morphing
mesh approach. Unfortunately no work was found using this test case with the overset mesh method
implemented in OpenFOAM.
Summing up, the purpose of this work is to compare the overset method and the morphing mesh
method implemented in OpenFOAM with reference studies available in the literature. By using a two
dimensional laminar configuration, the sources of errors are restricted to the numeric and the choice of the
artificial boundaries. Assuming that the errors associated with imposing artificial boundaries conditions
are not too big, one can directly make conclusions about the accuracy of methods used for the grid
motion. Especially for the overset method an accurate estimation of the numerical error is required since
the overset interpolation is known to not be mass conservative (see e.g. [3, 17]).
The paper is organized as follows: Section 2 describes the governing equations, Section 3 describes
the test cases used for this validation study, Section 4 describes the meshes used, Section 5 gives a brief
description of the numerical methods used, the results are exposed in Section 6 and the conclusion are
drawn in Section 7.

2. Governing equations
2.1. Fluid. The equation of motion used in OpenFOAM for moving meshes are written in the Arbitrary-
Euler-Lagrange (ALE) formulation. This formulation is one of the most popular if morphing meshes are
used to describe the solid body deformation or displacement [18]. If the overset method is used, this
formulation avoids to formulate the equation of motion in multiple frames of references [19]. For the
derivation of the equation see [20, 21]. The continuity and momentum equations are in this form given
by
Z I

ρdV + ρn · (U − Ug )dS = 0, (1)
∂t
COMPARISON OF OVERSET MESH WITH MORPHING MESH 15

Z I I I

ρUdV + ρn · (U − Ug )UdS − ρνn · ∇UdS + pndS = 0. (2)
∂t
In the above equations U is the fluid velocity, p the pressure, ν the molecular viscosity and ρ the
density. The only difference to the equation of motion in a fixed mesh is the introduction of a new
variable, i.e. the grid velocity Ug . In order to determine the grid velocity the space conservation law
(SCL) has to be solved as (see also [2, 18])
Z I

dV − n · Ug dS = 0. (3)
∂t
In order to not introduce additional mass conservation errors, the temporal discretisation of the SCL
3 should be the same as used in the other conservation equation (see [20, 2, 18]).

2.2. Solid. The equation of motion for the displacement xs of the rigid bodies involved in the FSI
problem, are derived from Newtons second law, as
I I
mx¨s = dx˙s + kxs + npdS + n · τvis dS. (4)

The above equation represent the usual spring mass damper system with additional two term which
account for the influence of the fluid forces on the movement of the solid. τvis is the viscose stress tensor.
k represents the spring constant, m the mass of the solid and d is the damping constant.

2.3. Mesh motion. If the morping mesh approach is adopted, an additional equation is solved for the
point displacement pd . Note that for the overset grid the topology of the mesh remains unchanged. The
point displacement is added to the position of the points at the previous time step p0 to get the new
points pn , as

pn = p0 + pd . (5)
In order to get a smooth transition from the displacement of the moving boundaries towards a usually
zero displacement of the outer boundaries, a Laplacian equation is solved. In the version of OpenFOAM
used in this work, an equation of the cell center displacement cd is solved rather than one for the point
displacement, given by

·
∇ (γ∇cd ) = 0. (6)
The diffusivity γ is proportional to the inverse distance from the moving boundary. Once cd is obtained,
it is interpolated to the points of the cells by using an inverse distance interpolation to get pd .

3. Test cases
3.1. One degree of freedom spring mass damper system. Figure 1 shows the setup of the simu-
lation of the one degree of freedom spring mass damper system: We have a closed square with a fluid
initially at rest. Actually the fluid part does not play a role in this test case since we are interested
only in the movement of the one degree of freedom spring-mass-damper system. In order to exclude any
influence of the fluid force on the movement of the circular cylinder two possible approaches are possible:
set the density ρ and the viscosity ν to zero (actually to a very small value to avoid possible divisions by
zero) or set the test keyword in the dynamicMeshDict to yes. Setting the test keyword to yes deactivates
the fluid forces. Both options deliver the same results for the present configuration. In this way the
pressure force which acts normal to the solid body surface and the shear stress which acts parallel to it
do not influence the motion of the cylinder.
The mass m is set to m = 0.03575 kg, the spring constant k of the linear spring is set to k = 69.48
N/m. Two different damping cases are considered. The first one is an undamped system. In the second
case the damping constant d is set to d = 0.0039 Ns/m which corresponds to a weakly damped system.
For undamped and weakly damped systems the oscillation induced by the fluid on the structure can have
potentially disastrous effects. The spring has an initial extension of z0 = 0.0016 m which corresponds to
the diameter of the cylinder.
For this configuration we have an analytical solution which we can use to verify our solver, given by
16 MICHAEL ALLETTO

p
z(t)/z0 = e−0.5(d/m)t cos(t k/m). (7)

Figure 1. Setup of the simulation of the one degree of freedom spring mass damper
system

3.2. Inline oscillating cylinder. Figure 2 shows the setup of the simulation for the inline oscillating
cylinder: At the inflow a uniform velocity of Uinf = 1m/s in x-direction is applied. The pressure is set
to zero gradient at this boundary. At the top and bottom walls a slip boundary condition is applied.
At the outlet the pressure is to be set equal to zero and for the velocity a zero gradient boundary
condition is applied. The kinematic viscosity ν = 0.01m2 /s is set in order that the Reynolds number
Re = Uinf D/ν based on the cylinder diameter D = 1m is equal to Re = 100. For this Reynolds number
the flow over a stationary cylinder is two dimensional. For this reason the flow is assumed to be two
dimensional. This is imposed by using only one cell in span wise direction and setting both boundary
conditions in this direction to the type empty. At the cylinder wall a no slip boundary condition is
applied for the velocity and zero gradient for the pressure. The cylinder oscillates in x-direction with a
prescribed sinosoidal motion: xc = Asin(2πfc t) = Asin(ωt). Four different frequencies are computed:
fc = 0, 0.16, 0.33, 0.66 1s . The amplitude is set to A = 0.14m. The main purpose of this test case
was to check whether the increase of the lift at fc = 2fst can be captured and if both techniques to
simulate moving bodies deliver comparable results. The static frequency fst is computed by performing
a simulation around a fixed cylinder. The force signal in flow-normal direction is analyzed by fast Fourier
transform (fft) and the frequency with the highest amplitude is used to compute fst . The domain extends
20 D in upstream direction, 40 D in downstream direction and 20 D in both lateral direction for most
simulation performed. This chosen domain size is in agreement with the settings used by other authors
for inline oscillating cylinders: Leontini et al. [10] placed the outflow boundary at 50 D from the center
of the cylinder. The location of the other boundaries are not mentioned. Liao et al. [6] used a domain
which extends 50 D in stream wise direction and 70 D in lateral direction. Unfortunately, the coordinate
of the cylinder center was not given by the author. The outer boundaries in the simulations performed
by Do et al. [22] where located 10 D upstream, 40 D downstream and 12 D in lateral direction from the
center of the cylinder. However, in order to confirm that the location of the outer boundaries have little
influence on the solution, for the coarse overset mesh (see Section 4 for the description) and a frequency
of fc = 3fst four additional simulations are made: one where the upstream boundary was located at 40
D from the center of the cylinder and all other boundary location are kept unchanged (abbreviated as
40DUS), one where the downstream boundary was located at 80 D from the center of the cylinder and all
other boundary location are kept unchanged (abbreviated as 80DDS), one where the lateral boundaries
are located at 40 D from the center of the cylinder and all other boundary location are kept unchanged
(abbreviated as 40DL) and a last one where the upstream boundary was located at 40 D, the downstream
boundary at 80 D and the lateral boundaries at 40 D from the cylinder center (abbreviated as DD). Only
the overset mesh is used for the domain size study since the influence of the domain boundaries should
be similar for both mesh types investigated. Furthermore it is very easy for the overset mesh to ensure
that the same grid is used in the overlapping regions of the five different meshes used for the domain size
investigation.
3.3. Vortex induced vibration of a cylinder. Figure 3 shows the setup of the simulation were we
analyze the vortex induced vibration of a circular cylinder: At the inflow a uniform velocity of Uinf =
0.0656m/s in x-direction is applied. The pressure is set to zero gradient at this boundary. At the top
COMPARISON OF OVERSET MESH WITH MORPHING MESH 17

Figure 2. Setup of the simulation of the inline oscillating cylinder

and bottom walls a slip boundary condition is applied. At the outlet the pressure is to be set equal
to zero and for the velocity a zero gradient boundary condition is applied. The kinematic viscosity
ν = 1.05 × 10−6 m2 /s is set in order that the Reynolds number Re = Uinf D/ν based on the cylinder
diameter D = 0.0016m is equal to Re = 100. For this Reynolds number the flow over a stationary
cylinder is two dimensional. For this reason the flow is assumed to be two dimensional. This is imposed
by using only one cell in span wise direction and setting both boundary conditions in this direction to
the type empty. At the cylinder wall a no slip boundary condition is applied for the velocity and zero
gradient for the pressure. The cylinder is free to oscillate in x- and y-direction. In this direction a linear
spring restrain type is applied. The linear damping is set to zero in order to be comparable with the
reference simulation of Singh and Mittal [23]. The domain extends 20 D in upstream direction, 40 D
in downstream direction and 20 D in both lateral direction for most simulations. For this size of the
computational domain, Prasanth and Mittal [24] found that the results are almost independent of the
location of the boundaries. In order check the influence of the location of the domain boundaries, four
additional simulation for the coarse overset mesh (see Section 4 for the description) and a U ∗ = fUn∞D =5.5
are performed. The same domain sizes as described in Section 3.2 are used.

Figure 3. Setup of the vortex induced vibration simulation

Four different non dimensional velocities are computed: U ∗ = fUn∞D =4, 4.8, 5 and 8.5. U∞ is the

ω k/m
incoming flow velocity and fn = 2π = 2π is the natural frequency of the spring-mass system. The
lowest and highest non-dimensional velocity correspond to cases in the work by Singh and Mattal [23],
where the shedding frequency is not synchronized with the spring-mass system and therefore low am-
plitude oscillation of the mass occur. A U ∗ =4.8 correspond to the onset of synchronization with large
amplitude oscillation (y/D ≈ 0.6) in the reference experiment. A U ∗ =5.5 correspond to a region in the
middle of the synchronization regime. Also for this case the reference experiment reported large ampli-
tude oscillations. Having defined the diameter D of the cylinder, the incoming flow velocity U∞ and the
mass m = 0.03575 kg, after that one can determine the four different linear spring constants k =148.2,
102.97, 78.42 and 32.83 N/m for a non-dimensional velocity of U ∗ =4, 4.8, 5.5 and 8.5, respectively.
18 MICHAEL ALLETTO

(a) Overview (b) Zoomed view

Figure 4. Mesh used for the overset method

(a) Overview (b) Zoomed view

Figure 5. Mesh used for the morphing mesh method

The last non-dimensional parameter to match is the non-dimensional mass m∗ = m/mf = 4m/(ρD2 H).
H = 0.12m is the height of the cylinder and mf is the mass of the fluid occupied by the cylinder. In
order to match this non-dimensional parameter the density of the cylinder is set to 14817 kg/m3 .

4. Mesh generation
In this section the meshes used for the simulations are described. Two types of meshes are used: one
for the morphing mesh simulation and one for the overset simulations. Since the topology of the mesh
is identical for all simulations and only the number of cells changes, the meshes employed are described
only once.
The mesh used for the overset mesh simulation is shown in Figure 4. We can see that it is formed
by the mesh surrounding the cylinder and the background mesh. The mesh surrounding the cylinder is
generated by simply stretching the circular form of the cylinder outward. The background mesh consist
of a rectangular block mesh with different refinement zones. In the outer part of the rectangle where
the gradients of the flow field are smaller, a coarser mesh is used. Closer to the region surrounding the
cylinder the mesh is finer. Note that when generating an overset mesh, it has to be paid attention that
the cells at the intersection of the moving and the background mesh should have roughly the same size
to minimize the interpolation errors.
Figure 5 shows the morphing mesh. It is generated using the utility blockMesh. Note that the
dictionary for the generation of the mesh was taken online* and slightly modified in order to adopt it to
the present test case. It is evident that the constitution of the mesh is much more complex compared to
the overset mesh.

5. Numerical methods
Both solvers used for the simulations (pimpleFoam for the morphing mesh method and overPim-
pleDyMFoam for the overset method) use the PIMPLE algorithm. It offers the possibility to use either
the SIMPLE or the PISO algorithm to achieve an iterative pressure velocity coupling. The user can
specify the number of outer iteration used to account for the non-linearity of the governing equations and
the number of inner iterations which adjust the pressure in order to fulfill the mass conservation. For the
details on the SIMPLE and the PISO algorithm the interested reader is referred to the book of Ferziger
et al. [20]. For the details of the implementation in OpenFOAM the interested reader may have a look at
the book of Moukalled et al. [25]. In the present study three outer and one inner iteration was used for
*https://curiosityfluids.com/2016/07/19/oscillating-cylinder-in-crossflow-pimpledymfoam/
COMPARISON OF OVERSET MESH WITH MORPHING MESH 19

(a) Undamped (b) Damped

Figure 6. Response of the one degree of freedom mass-spring-damper system to an


initial displacement for the overset mesh

all test cases. The divergence terms in the momentum equation are dicretized using a fourth-order cubic
scheme. This scheme was found to be less diffusive compared to a linear scheme leading to higher lift and
drag coefficients for a stationary cylinder. A second-order backward scheme was used for the temporal
discretization. In order to study the influence of the disretization scheme for the divergence and the
transient term on a moving cylinder, simulations with the first-order Euler scheme for the transient term
and a second-order linear scheme for the divergence term are performance. A more detailed description
is provided in the result section. For the overset method an inverse distance interpolation was used to in-
terpolate the variables between the grids. Note that a fairly good description of the method implemented
in OF1906 can be found in the work by Tisovska [26]. In order to advance the position of the solid body
in time, two different methods are testes: The Symplectic methods developed by Dullweber et al. [27]
and the Newmark method [28].

6. Results
6.1. One degree of freedom spring mass damper system. Figure 6 shows the results obtained by
the overset mesh. Two different time step sizes of dt = 0.0005 s and dt = 0.001 s are compared with each
other. Furthermore two different method to integrate the ordinary differential equation are investigate:
the Newmark method and the symplectic method. It has to be mentioned that the Newmark method
for the overset method is explicit since the spring and damper force are not updated during the outer
iterations of the pimple loop. In the Newmark method used for the morphing mesh approach the spring
and damper forces are updated during the outer iterations of the pimple loop. For this reason it can be
regarded as implicit. The symplectic method is explicit for both approaches compared here (spring and
damper forces are kept constant within a time step). It is obvious that the Newmark method is unstable
for both undamped and weakly damped system. The error decreases with decreasing time step. The
symplectic method reproduces the analytical result exactly.
The results shown in Fig. 7 are obtained for the morphing mesh. Four different simulation are
performed for the undamped (left) and damped case (right). The time step size was dt = 0.001 s. For
one simulation the symplectic method was used to compute the evolution of the velocity and position of
the solid. For the other three the Newmark method was used. For one simulation the acceleration of
the solid body was not updated during the outer iterations (labeled as Newmark no outer correction).
For the remaining simulations the acceleration of the solid body was updated during the two (labeled
as Newmark 2 outer) and four (Newmark 4 outer) outer iterations. It is evident that for the cases
where the Newmark method without updating the acceleration during the outer iteration is used, are not
stable. The symplectic method is stable. The observations that the Newmark method is unstable if the
acceleration is not updated during the outer iteration of the pimple loop agrees well with the findings
achieved using the overset solver.
6.2. Inline oscillating cylinder. Before starting with the comparison between the overset mesh and
the morphing mesh, the analysis of the influence of the domain size is summarized in Table 1. The
coarse mesh is used for this study. A frequency of ffstc = 3 is taken for this study since it is believed
that the domain size should have the biggest influence on the results for this frequency since the drag
20 MICHAEL ALLETTO

(a) Undamped (b) Damped

Figure 7. Response of the one degree of freedom mass-spring-damper system to an


initial displacement for the morphing mesh

Table 1. Comparison of the Strouhal number, the mean drag coefficient Cd,mean and
the maximum lift coefficient Cl,max for different domain sizes

oc oc 40DUS oc 80DDS oc 40DL oc DD


St 0.157 0.157 0.157 0.157 0.157
Cd,mean 1.39 1.39 1.39 1.38 1.36
Cl,max 0.42 0.43 0.42 0.42 0.42

force is highest. It is evident that there are only small differences between the five different domain sizes
analyzed when looking at integral quantities like the Strouhal number, the mean drag coefficient Cd,mean
and the maximum lift coefficient Cl,max . For this reason the smallest domain is sufficient to guarantee
results which are independent of the domain size.
The comparison between overset mesh and the morphing mesh starts with a study on the influence
of the discretization of the convective term and the time derivative in the momentum equation. For the
purpose also the reduced frequency of ffstc = 3 is taken. The coarse mesh is used. Three simulation for
each mesh type are made: One with a fourth-order discretization for the convective term and a second-
order discretization for the time derivative (abbreviated as t2d4), one with a fourth-order discretization
for the convective term and a first-order discretization for the time derivative (abbreviated as t1d4)
and one with a second-order discretization for the convective term and a second-order discretization
for the time derivative (abbreviated as t2d2). The abbreviation oc stands for overset coarse and the
abbreviation mc stands for morphing coarse. The comparison of the Strouhal number, the mean drag
coefficient Cd,mean and the maximum lift coefficient Cl,max for different discretizations and mesh types
is shown in Table 2. It is evident that the Strouhal number St is not affected by the chosen schemes.
The mean drag coefficient Cd,mean decreased for the overset mesh if a first-order Euler scheme is used
instead of second-order backward. For the morphing mesh it remains unchanged. Changing the order
of the discretisation of the convective scheme from fourth to second leads to a decrease in Cd,mean for
the overset mesh and increase for the morphing. The maximum lift coefficient Cl,max increases for the
overset mesh and remains unchanged for the morphing mesh changing the temporal discretisation from
second to first order. Changing the convective scheme from fourth to second order leads to an increase
of the maximum lift coefficients Cl,max for the overset mesh and a slight reduction of the morphing
mesh. The root mean square lift coefficient Cl,rms is very little affected by the choice of the discretisation
schemes. The reason why the morphing mesh and the overset mesh show different trends regarding the
discretisation is not very clear. For the vortex-induced vibration case (see Table 6) the trend with respect
to the discretisation schemes is however the same for the overset and the morphing mesh.
As advisable for every new simulation technique, we start with a small mesh convergence study. A
mesh convergence study allows to guess the numerical errors (at least for laminar flows where no modeling
errors are introduced) and to find a trade off between accuracy and time to solution. The results are
presented in Table 3. The mesh study was made for two setups: the stationary cylinder and a reduced
frequency of ffstc = 3. The purpose was to check if the influence of the mesh resolution is different for
COMPARISON OF OVERSET MESH WITH MORPHING MESH 21

Table 2. Comparison of the Strouhal number, the mean drag coefficient Cd,mean , the
maximum lift coefficient Cl,max and the root mean square lift coefficient Cl,rms for dif-
ferent discretizations and mesh types

oc t2d4 mc t2d4 oc t1d4 mc t1d4 oc t2d2 mc t2d2


St 0.157 0.157 0.157 0.157 0.157 0.157
Cd,mean 1.39 1.32 1.31 1.32 1.34 1.34
Cl,max 0.42 0.46 0.46 0.47 0.45 0.44
Cl,rms 0.20 0.18 0.22 0.18 0.19 0.18

different oscillation frequencies. For the purpose of the mesh study, we analyzed three overset meshes: one
with 17 300 (abbreviated as of), one with 13 300 (abbreviated as om) and one with 5 800 cells (abbreviated
as oc). Tree morphing meshes one with 23 000 (abbreviated as mf), one with 12 700 (abbreviated as
mm) and one with 7 400 cells (abbreviated as mc) are compared as well.
Table 3 shows the results of the mesh study. Reference simulations and experiments of a stationary
cylinder are also included. The variables which are compared are the Strouhal number St, the mean drag
coefficient Cd,mean and the maximum lift coefficient Cl,max . The Strouhal number St is computed by
analyzing the signal of the lift (force in z-direction). The time signal generated by the function object
is interpolated to a equidistant time step size array of 10 000 points. An equidistant time spacing is
required by the fft Python packages used. The last 3 000 points are used to compute the statistics in
order to exclude the initial transient stage of the simulation. The Strouhal number is computed from
the frequency with the highest amplitude. The simulations were run for 200 s. The signals of the last
60 s of the simulation are used to compute these three quantities. Since the purpose of this work is
to evaluate the capability of OpenFOAM to simulate an oscillating cylinder, the first necessary step is
to check the quality of the simulations for a fixed cylinder. It is evident that the agreement with the
references is reasonable. The results of the overset mesh and the mesh generated with blockMesh are
very close. This means that the overset interpolation does not deteriorate the quality of the results for
this case and the errors introduced by the non-mass conservative interpolation is small. Regarding the
analysis of the violation of the mass conservation see e.g. [3, 17] or Fig. 12. Interestingly the St number
does not show any mesh influence. Regarding the overset mesh the paramters analyzed are close for all
three meshes used. We see a slight increase of the maximum lift coefficient Cl,max and the mean drag
coefficient Cd,mean remains almost unchanged. Regarding the morphing mesh, the lift coefficient Cl,max
slightly increases when the resolution is increased. The results shown are computed with a Python script
which analyzes the forces computed around the cylinder. Of course the Python script is also provided
within the downloadable files. Regarding the mesh study for ffstc = 3, we can conclude that the solution
still changes slightly. Both Cd,mean and Cl,max increase when changing from the morphing medium mesh
to the morphing fine mesh. For the overset mesh Cd,mean decreases from the medium to the fine mesh and
Cl,max remains unchanged. It is therefore recommended to use even finer meshes for both approaches in
order to reach a mesh independent solution. To access the grid size necessary to have mesh independent
results is not topic of the paper. The main purpose of this paper is to compare the results of the morphing
mesh approach with the overset approach with similar number of cells.
Table 4 shows the comparison of the mean drag coefficient Cd,mean and the maximum lift coefficient
Cl,max with other simulation (sim) for different oscillation frequencies fc . For the present simulations
also the Strouhal number computed with the frequency of the lift signal with the maximum amplitude is
shown. It can see seen that both mean drag coefficient Cd,mean and the maximum lift coefficient Cl,max
have a maximum at a frequency twice the shedding frequency of a fixed cylinder fc = 2fst . The difference
between the morphing mesh and the overset mesh are marginal and the agreement with the reference
simulations are reasonable. All simulation (the one of this work and also the reference simulations) report
an almost tripling of the lift coefficient with respect to the stationary cylinder (see Table 3). The Strouhal
number for fc = 2fst is the same as for the stationary cylinder (see Table 3). That means that also the
frequency lock in can be reproduced with both the overset and the morphing mesh. As shown for the
previous study, the integral quantities (except the Strouhal number St) slightly vary between the coarse
and the medium meshes used.
Figure 8 shows the comparison of the temporal evolution of drag coefficient in the last 60 s (it is
the time span where the statistics shown in the tables are computed) of the simulation for the medium
mesh. It is evident that for increasing frequency fc the amplitude of the drag coefficient increases since
the velocity at which the cylinder is displacing through the fluid increases. The results of the morphing
mesh and the overset mesh are very close. The oscillatory part of the drag coefficient seems to by given
22 MICHAEL ALLETTO

Table 3. Comparison of the mean drag coefficient Cd,mean and the maximum lift coef-
ficient Cl,max with other simulation (sim) and experiments (exp) for different oscillation
frequencies fc different meshes

overset morphing
St Cd,mean Cl,max St Cd,mean Cl,max
coarse ffstc = 0 0.164 1.30 0.31 0.164 1.33 0.27
medium ffstc = 0 0.164 1.33 0.33 0.164 1.33 0.30
fine ffstc = 0 0.164 1.32 0.32 0.164 1.30 0.31
coarse ffstc = 3 0.157 1.39 0.44 0.157 1.39 0.47
medium ffstc = 3 0.157 1.43 0.50 0.157 1.36 0.50
fine ffstc = 3 0.157 1.37 0.50 0.157 1.44 0.52
[6] (sim) ffstc = 0 1.36 0.34
[5] (sim) ffstc = 0 1.42 0.32
[29] (exp) ffstc = 0 0.166

Table 4. Comparison of the mean drag coefficient Cd,mean and the maximum lift coef-
ficient Cl,max with other simulation (sim) for different oscillation frequencies fc

overset morphing
St Cd,mean Cl,max St Cd,mean Cl,max
coarse ffstc = 1 0.157 1.36 0.45 0.157 1.36 0.40
coarse ffstc = 2 0.164 1.52 0.93 0.164 1.62 0.90
coarse ffstc = 3 0.157 1.39 0.44 0.157 1.39 0.47
medium ffstc = 1 0.157 1.37 0.46 0.164 1.36 0.42
medium ffstc = 2 0.164 1.57 0.92 0.164 1.58 0.91
medium ffstc = 3 0.157 1.43 0.50 0.157 1.36 0.50
[6] (sim) ffstc = 2 1.71 0.95
[5] (sim) ffstc = 2 1.70 0.97

by the movement of the cylinder. This is clearly visible by the functional evolution of Cd . The lower
the frequency of the moving cylinder, the lower is also the frequency of the drag coefficient. There is no
phase shift visible between the overset mesh and the morphing mesh. For both approaches the functional
evolution of the position of the cylinder is the same. The drag force can be divided into friction and
pressure contribution. Actually the pressure contribution can be divided into pressure difference between
front and rear part of the cylinder as a result of the wake and a contribution resulting from the acceleration
of the fluid induced by the acceleration of the cylinder (it’s called added mass effect). Both contribution
are influenced by the resolution. The grid close to the cylinder is very similar for the overset mesh and the
morphing mesh. This can partially explain the not visible differences. The marginal difference between
morphing and overset mesh can be seen also by the mean drag coefficient displayed in Table 4.
Figure 9 shows the comparison of the temporal evolution of lift coefficient during the initial transient
stage. The temporal evolution of the lift coefficient is similar for both simulation approaches except for a
frequency of fc = 3fst . For this frequency the initial transient stage of the overset mesh is much shorter
compared to the morphing mesh. The reason is not clear. A cautious guess is that the numerical errors
of the overset mesh lead to a faster destabilization of the initial steady solution. Figure 10 shows the
comparison of the temporal evolution of lift coefficient during the fully developed stage of the simulation.
It is obvious that for a frequency of fc = 2fst the amplitude of the lift coefficient is highest. Again the
results of the morphing mesh and the overset mesh are very similar. From the signals it is also evident,
that the higher frequencies content increases with fc . Figure 11 shows the amplitude spectrum of the
lift coefficient for the four oscillation frequencies fc analyzed using both mesh types. The results of both
mesh types are again very similar. The amount of peaks increase with increasing fc . Note that the
amplitudes of the lift coefficient Cl calculated by the fft of Python are very close to the one recorded
(compare Fig. 11 with Fig. 10) demonstrating the usefulness of the provided Python scripts to retrieve
the frequencies with the highest amplitudes.
COMPARISON OF OVERSET MESH WITH MORPHING MESH 23

(a) Morphing mesh (b) Overset mesh

Figure 8. Comparison of the drag coefficient of the morphing mesh and overset mesh

(a) Morphing mesh (b) Overset mesh

Figure 9. Comparison of the lift coefficient of the morphing mesh and overset mesh at
the beginning of the simulation

(a) Morphing mesh (b) Overset mesh

Figure 10. Comparison of the lift coefficient of the morphing mesh and overset mesh
for the developed phase

A final remark should be made regarding the computational time. For the fixed cylinder and the case
with an oscillation frequency fc = 0.16, the simulations with the morphing mesh are much faster. For
increasing frequency however, the maximum Courant number Comax had to be decreased from 0.75 for
fc = 0.16 to Comax = 0.5 for fc = 0.33 and Comax = 0.25 for fc = 0.66 for the morphing mesh. The
24 MICHAEL ALLETTO

(a) Morphing mesh (b) Overset mesh

Figure 11. Comparison of the amplitude spectrum of lift for the morphing mesh and
overset mesh

Table 5. Comparison of the St number, the r.m.s drag coefficient Cd,rms , the r.m.s
lift coefficient Cl,rms , the r.m.s of the x-displacement xrms /D and the maximum z-
displacement zmax /D for different domain sizes

oc oc 40DUS oc 80DDS oc 40DL oc DD


St 0.175 0.175 0.175 0.175 0.175
Cd,rms 0.36 0.35 0.35 0.35 0.34
Cl,rms 0.46 0.48 0.46 0.49 0.51
zmax /D 0.57 0.57 0.57 0.56 0.57
xrms /D 0.01 0.009 0.008 0.009 0.008

purpose was to avoid nonphysical pressure peaks which occurred for higher Comax . For this reason the
simulation time for the cases with the highest frequency was comparable between both mesh types. For
the overset mesh the maximum Courant number Comax = 0.75 was the same for all frequencies and the
coarse and the medium mesh. For the fine mesh however, a Courant number of Comax = 0.5 had to be
used to avoid nonphysical high amplitude and high frequency oscillation in the lift and drag signal. The
reason for this is not really clear. According to Benjamin et al. [30] the mesh Courant number for moving
meshes should be of order unity to guarantee high quality results. For the highest frequency fc = 0.66
we get a maximum velocity of the cylinder of 0.58 m/s. The smallest cell size in the region where the
background mesh and the mesh around the cylinder overlap is ∆x = 0.125 m. With these quantities
we get the time needed for the mesh to cover a distance equal to the smallest cell size of 0.21 s. The
typical time step size used in the simulation is of the order of 0.01 s and hence much smaller than the
time required for the moving mesh to cross a cell of the background mesh.
6.3. Vortex induced vibration of a cylinder. Before starting with the comparison between the
overset mesh and the morphing mesh, the analysis of the influence of the domain size is summarized in
Table 5. It is evident that the St number is independent of the size of the domains analyzed. We observe
a small reduction of the r.m.s. drag coefficient Cd,rms and small increase of the r.m.s lift coefficient Cl,rms
if the domain size is doubled. The maximum z-displacement zmax /D remains unchanged between the
difference position of the domain boundaries. The maximum z-displacement zmax /D is reduced slightly
if the domain size is doubled. Since the changes of these integral quantities are very small, the smallest
domain size is used in order to save computational time.
The comparison between the overset mesh and the morphing mesh starts as for the forced oscillation
case with a study on the influence of the discretization of the convective term and the time derivative
in the momentum equation. For the purpose also the reduced velocity of U ∗ = 5.5 is taken. The coarse
mesh is used. Three simulations for each mesh type are made: One with a fourth-order discretization
for the convective term and a second-order discretization for the time derivative (abbreviated as t2d4),
one with a fourth-order discretization for the convective term and a first-order discretization for the time
derivative (abbreviated as t1d4) and one with a second-order discretization for the convective term and a
second-order discretization for the time derivative (abbreviated as t2d2). Table 6 shows the comparison
COMPARISON OF OVERSET MESH WITH MORPHING MESH 25

Table 6. Comparison of the Strouhal number, the root mean square drag coefficient
Cd,rms and the root mean square lift coefficient Cl,rms , the root mean square of the x-
displacement xrms /D and the maximum z-displacement zmax /D for different discretiza-
tions and mesh types

oc t2d4 mc t2d4 oc t1d4 mc t1d4 oc t2d2 mc t2d2


St 0.175 0.177 0.175 0.177 0.175 0.177
Cd,rms 0.36 0.34 0.36 0.34 0.37 0.37
Cl,rms 0.46 0.32 0.53 0.38 0.45 0.32
zmax /D 0.57 0.55 0.56 0.56 0.57 0.56
xrms /D 0.01 0.008 0.01 0.009 0.01 0.01

Table 7. Mesh comparison study for U ∗ = 5.5: Comparison of the St number, the r.m.s
drag coefficient Cd,rms , the r.m.s lift coefficient Cl,rms , the r.m.s of the x-displacement
xrms /D and the maximum z-displacement zmax /D

of om oc mf m. m. mm [23]
St 0.180 0.174 0.173 0.179 0.181 0.184 0.179
Cd,rms 0.35 0.35 0.35 0.34 0.34 0.34 0.33
Cl,rms 0.47 0.45 0.44 0.31 0.29 0.29 0.24
zmax 0.57 0.57 0.57 0.55 0.55 0.55 0.53
xrms 0.006 0.006 0.006 0.0057 0.0057 0.0059 0.0056

of the Strouhal number, the root mean square drag coefficient Cd,rms and the root mean square lift
coefficient Cl,rms , the root mean square of the x-displacement xrms /D and the maximum z-displacement
zmax /D for different discretizations and mesh types. It is evident that the Strouhal number St is not
affected by the discretisation schemes. The r.m.s drag coefficient Cd,rms increases when switching from a
fourth-order to a second-order convective scheme. The temporal discretisation has little influence on this
quantity. The trend is the same for the overset and the morphing mesh. The r.m.s. lift coefficient Cl,rms
is almost not affected by the chosen convective scheme. Using the first-order Euler scheme instead of the
second-order backward scheme increases the Cl,rms . Again the trend is the same for the overset and the
morphing mesh. The root mean square of the x-displacement xrms /D and the maximum z-displacement
zmax /D are not affected by the schemes chosen.
As advisable for every new simulation technique, we start with a small mesh convergence study. For
this purpose we analyzed three overset meshes: one with 17 300 (abbreviated as of), one with 13 300
(abbreviated as om) and one with 5 800 cells (abbreviated as oc). Three morphing meshes one with 23
000 (abbreviated as mf), one with 12 700 (abbreviated as mm) and one with 7 400 cells (abbreviated as
mc) are compared as well.
Table 7 shows the mesh comparison study for a U ∗ = 5.5. The results of the fine, medium and the
coarse mesh are compared to reference simulations. The table shows the St number St = f D/U , the
r.m.s drag coefficient Cd,rms , the r.m.s lift coefficient Cl,rms , the r.m.s of the x-displacement xrms /D and
the maximum z-displacement zmax /D. The shedding frequency f is calculated using the time signal of
the lift force. The time signal generated by the function object is interpolated to a equidistant time step
size array of 3 000 points. An equidistant time spacing is required by the fft Python packages used. The
last 1 000 points are used to compute all statistics shown in Table 7 and Table 8 in order to exclude
the initial transient stage of the simulation. The results are compared with the simulations of Singh and
Mittal [23]. From these integral coefficients it is evident that no big differences are visible between the
coarse, medium and fine overset mesh and the coarse, medium and fine morphing mesh. For this reason
all further computations are made with the coarse version of both meshes in order to save computational
resources. The agreement with the reference simulation is reasonable.
Since it is known that the overset methodology is not mass conservative, Figure 12 shows the normalized
difference ṁn = (ṁin − ṁou )/ṁin ∗ 100 of the mass flow rate at the inlet ṁin and the mass flow rate
at the outlet ṁou . Note that the normalized mass flow difference is shown in percent. It is evident that
there exists a small imbalance between the flow rate at the inlet and the outlet. It decreases for a finer
mesh.
Table 8 compares the St number, the r.m.s drag coefficient Cd,rms , the r.m.s lift coefficient Cl,rms , the
r.m.s of the x-displacement xrms /D and the maximum z-displacement zmax /D for different U ∗ with the
simulations of Singh and Mittal [23]. It is evident that the best agreement with the reference simulation
26 MICHAEL ALLETTO

Figure 12. Normalized flow rate difference in percent of the three overset meshes used.

Table 8. Comparison of the St number, the r.m.s drag coefficient Cd,rms , the r.m.s
lift coefficient Cl,rms , the r.m.s of the x-displacement xrms /D and the maximum z-
displacement zmax /D for different U ∗

St Cd,rms Cl,rms zmax /D xrms /D


overset U ∗ = 4 0.156 0.006 0.20 0.012 0.002
morping U ∗ = 4 0.167 0.0059 0.25 0.016 0.001
[23] U ∗ = 4 0.172 0.0 0.30 0.028 0.00019
overset U ∗ = 4.8 0.156 0.0047 0.24 0.031 0.001
morping U ∗ = 4.8 0.186 0.100 0.65 0.268 0.003
[23] U ∗ = 4.8 0.198 0.47 0.63 0.56 0.0063
overset U ∗ = 5.5 0.175 0.36 0.46 0.57 0.01
morping U ∗ = 5.5 0.177 0.34 0.32 0.55 0.008
[23] U ∗ = 5.5 0.179 0.33 0.24 0.53 0.0056
overset U ∗ = 8.5 0.146 0.023 0.10 0.11 0.002
morping U ∗ = 8.5 0.156 0.013 0.15 0.065 0.001
[23] U ∗ = 8.5 0.161 0.0 0.18 0.037 0.00023

is obtained for the morphing mesh. For U ∗ = 4.8 where Singh and Mittal [23] obtained already a locked
in state, the overset mesh simulation reports only small oscillation and the morphing mesh simulation
seems to be in a transitional regime right before the onset of lock in (see Fig. 14 displaying the evolution
of the z-displacement). For the highest non-dimensional velocity U ∗ = 8.5 the overset mesh predicts
higher oscillations compared with Singh and Mittal [23] and the morphing mesh. It has be be said that
this U ∗ lays very close to the end of the lock in region reported by Singh and Mittal [23]. It seems that
the boarders of the lock in region obtained with the overset mesh are shifted towards higher U ∗ compared
with the morphing mesh and the reference simulation. This is also confirmed by looking at the r.m.s lift
coefficient Cl,rms at a U ∗ = 5.5 for the overset mesh. The values are much higher compared with the
morphing mesh and the results of Singh and Mittal [23]. Singh and Mittal [23] reported a drastic increase
of Cl,rms at U ∗ = 4.7 (when lock in occurs). The maximum of Cl,rms is also at U ∗ = 4.7. After that
Cl,rms monotonically decreases with U ∗ . The displacements also decrease monotonically after U ∗ = 4.7
but not so fast as the lift. This means that if for the overset mesh the onset of lock in is predicted at
higher U ∗ respect to the reference, it is in line with this observation that in the lock in regime Cl,rms is
higher for the overset mesh compared with the reference.
Figure 13 compares the work done from the start of the simulation (t = 0) until the time t by the fluid
on the cylinder for the two mesh types and the four non-dimensional velocities analyzed. It is calculated
as follows:
Z t
W (t) = Fz (τ )vz (τ )dτ (8)
0
Fz is the force in the z-direction exerted by the fluid on the cylinder and vz is the z-velocity. According
to Mittal et al. [14] and Morse and Williamson [31], the main mechanisms which lead to large amplitude
oscillation is a positive energy transfer (or work done) over an oscillation cycle from the fluid to the
COMPARISON OF OVERSET MESH WITH MORPHING MESH 27

(a) Overset mesh (b) Morphing mesh

Figure 13. Work done by the fluid on the cylinder

(a) Overset mesh (b) Morphing mesh

Figure 14. Displacement in z-direction for different U ∗ and the two mesh types

oscillator. In order to check if this mechanisms can be found also in the present study, the work done by
the fluid is plotted in figure 13. It is evident that for the cases with large oscillations in the z-direction
the work done experiences a sharp increase and after that it remains constant. The large work done
by the fluid is necessary to increase the amplitude of the displacement. After that a mean steady state
is reached where the work done by the fluid over an oscillation cycle has to be zero for the present
simulations without structural damping [14]. For this reason the total work done by the fluid on the
oscillator converges to a constant value.
Figure 14 shows the temporal evolution of the z-displacement of the cylinder for the overset mesh
and the morphing mesh. It is evident that the maximum displacement is observed for U ∗ =5.5 for both
meshes. For U ∗ = 4 both mesh types show small oscillations. A remarkable difference between the two
mesh types is observed for U ∗ = 4.8: While for the overset mesh the amplitude of the oscillations is still
small as in the pre-lock in regime, for the morphing mesh we can observed an intermittent behavior as
observed by Prasanth and Mittal [24] in the small transition region before lock in occurs. For U ∗ = 8.5
the amplitude of the overset mesh is higher compared to the morphing mesh.
Figure 15 shows the temporal evolution of the x-coordinate of the cylinder for the overset mesh and the
morphing mesh. We can see that the two approaches deliver slightly different equilibrium position of the
x-coordinate. The oscillations around the equilibrium position are very small for both approaches. The
maximum fluctuation of the x-coordinate are observed for both mesh types for U ∗ = 5.5. The magnitude
is less than 5% of the cylinder for this case. For the reduced velocities U ∗ = 5.5 and 8.5 a slightly shorter
initial transient phase can be observed for the morphing mesh approach compared to the overset mesh.
For the other two reduced velocities analyzed no visible difference in the duration of the initial transient
phase can be seen.
28 MICHAEL ALLETTO

(a) Overset mesh (b) Morphing mesh

Figure 15. Displacement in x-coordinate for different U ∗ and the two mesh types

Figure 16 shows snapshots of z-vorticity for the overset and morphing mesh for U ∗ = 4, U ∗ = 5.5 and

U = 8.5. For the first two non-dimensional velocities the shedding modes are different. For the lower
non-dimensional velocity the shedding mode is similar to a stationary cylinder: Two single vortecies are
shed per cycle from the cylinder. This shedding mode is called 2P mode which is also observed by Singh
and Mittal [23] and Prasant and Mittal [24] in the low amplitude case. For the high amplitude oscillations
case these two vortecies merge giving rise to the C(2P) mode also observed by Singh and Mittal [23] and
Prasant and Mittal [24]. The flow pictures obtained with the overset mesh and the morphing mesh are
very similar for these two U ∗ . For U ∗ = 8.5, where the fluid flow is outside the lock in regime, the flow
picture is again similar to a stationary cylinder resulting in a 2P mode. The same is observed also by
Singh and Mittal [23] and Prasant and Mittal [24].

7. Conclusion
In this work we analyzed the behavior of the overset method and the morphing mesh method with
the help of a two-dimensional cylinder configuration. First the solution obtained with the six-degree of
freedom solver was compared with an analytical solution of spring-mass-damper system. As next step
the complexity of the case analyzed was increased and the inline oscillation of a cylinder was simulated.
Finally the vortex induced oscillation was tackled. The purpose of increasing the complexity of the
test cases step by step is to isolate potential hurdles. Furthermore it is easier to understand a complex
configuration if beforehand easier problems are analyzed and understood.
Regarding the one degree of freedom spring mass damper system we can draw following conclusions:
• It is very useful to verify numerical methods with analytical solutions. From my experience as
model developer and programmer one can get rid of the majority of bugs and errors in the settings
if one compares the results of the simulation with an analytical solution where one exactly knows
the outcome.
• If the overset mesh is used, it is recommended to use the symplectic solver. The Newmark is
unstable for a undamped and weakly damped system.
• If the morphing mesh is used, it is recommended to update the acceleration of the solid body
during the outer iterations if the Newmark method is used.
Regarding the inline oscillation cylinder we can draw following conclusions:
• The results of the overset mesh are very similar to the one obtained with the morphing mesh.
• The maximum of the lift occurs at twice the shedding frequency of the stationary cylinder. This
is in line with the findings of other simulations and experiments.
• Frequency lock in at fc = 2fst can be reproduced with both overset and morphing mesh.
• A mesh study has revealed that for an oscillating cylinder more grid points are required to reach
a mesh independent result compared to a fixed cylinder.
Regarding the vortex induced oscillation we can draw following conclusions:
• Both overset mesh and morphing mesh can predict the large amplitude oscillations of the cylinder
in the lock in regime. Outside this regime (for lower and higher values of non-dimensional velocity
U ∗ ) the reported amplitudes of the structural oscillations are low. Inside the lock in regime the
COMPARISON OF OVERSET MESH WITH MORPHING MESH 29

(a) Overset, U ∗ = 4 (b) Morphing, U ∗ = 4

(c) Overset, U ∗ = 5.5 (d) Morphing, U ∗ = 5.5

(e) Overset, U ∗ = 8.5 (f) Morphing, U ∗ = 8.5

Figure 16. Snapshots of z-vorticity for the overset and morphing mesh for U ∗ = 4,
U ∗ = 5.5 and U ∗ = 8.5

two single vortices shed per cycle merge resulting in the so called C(2P) mode. Outside the lock
in regime the two single vortices are maintained also further downstream. The same observations
are made also by Singh and Mittal [23] and Prasant and Mittal [24].
• It seems that the boarders of the lock in region obtainable with the overset mesh with the
settings used in this work are shifted towards higher U ∗ compared with the morphing mesh and
the reference simulation.
• For the current settings the morphing mesh is up to two times faster compared with the overset
mesh for similar number of cells. The reason is probably that the pressure solver requires more
iterations for the overset mesh compared to the morphing mesh. For this reason it is better to
use the former if the movements of the structure are not too large to deteriorate the mesh quality.
Note also that all cases are run in serial. The performance analysis should be done also for a
parallel run. This is not in the scope of the current work.
• The work done by the fluid on the cylinder is highest for the cases with the highest amplitude
oscillations.
Author Contributions: Conceptualisation, M.A.; methodology, M.A.; software, M.A; validation, M.A.; formal
analysis, M.A.; investigation, M.A.; resources, M.A.; data curation, M.A.; writing—original draft preparation,
M.A.; writing—review and editing, M.A.; visualisation, M.A.; supervision, M.A.; All authors have read and agreed
to the published version of the manuscript.

References
[1] E. H. Dowell and K. C. Hall, “Modeling of fluid-structure interaction,” Annual review of fluid mechanics, vol. 33, no. 1,
pp. 445–490, 2001.
[2] H. Jasak, “Dynamic mesh handling in openfoam,” in 47th AIAA aerospace sciences meeting including the new horizons
forum and aerospace exposition, 2009, p. 341.
[3] S. Völkner, J. Brunswig, and T. Rung, “Analysis of non-conservative interpolation techniques in overset grid finite-
volume methods,” Computers & Fluids, vol. 148, pp. 39–55, 2017.
30 MICHAEL ALLETTO

[4] H. Gopalan, R. Jaiman, and D. D. Chandar, “Flow past tandem circular cylinders at high reynolds numbers using
overset grids in openfoam,” in 53rd AIAA Aerospace Sciences Meeting, 2015, p. 0315.
[5] S.-W. Su, M.-C. Lai, and C.-A. Lin, “An immersed boundary technique for simulating complex flows with rigid
boundary,” Computers & fluids, vol. 36, no. 2, pp. 313–324, 2007.
[6] C.-C. Liao, Y.-W. Chang, C.-A. Lin, and J. McDonough, “Simulating flows with moving rigid boundary using immersed-
boundary method,” Computers & Fluids, vol. 39, no. 1, pp. 152–167, 2010.
[7] D. D. Chandar, B. Boppana, and V. Kumar, “A comparative study of different overset grid solvers between openfoam,
starccm+ and ansys-fluent,” in 2018 AIAA Aerospace Sciences Meeting, 2018, p. 1564.
[8] O. M. Griffin and M. Hall, “Vortex shedding lock-on and flow control in bluff body wakes,” 1991.
[9] Y. Tanida, A. Okajima, and Y. Watanabe, “Stability of a circular cylinder oscillating in uniform flow or in a wake,”
Journal of Fluid Mechanics, vol. 61, no. 4, pp. 769–784, 1973.
[10] J. S. Leontini, D. L. Jacono, and M. C. Thompson, “A numerical study of an inline oscillating cylinder in a free
stream,” Journal of Fluid Mechanics, vol. 688, pp. 551–568, 2011.
[11] L. Shen, E.-S. Chan, and P. Lin, “Calculation of hydrodynamic forces acting on a submerged moving object using
immersed boundary method,” Computers & Fluids, vol. 38, no. 3, pp. 691–702, 2009.
[12] M. Pomarede, E. Longatte, and J.-F. o. Sigrist, “Benchmark of numerical codes for coupled csd/cfd computations on
an elementary vortex induced vibration problem,” in ASME Pressure Vessels and Piping Conference, vol. 43673, 2009,
pp. 537–546.
[13] C. Williamson and R. Govardhan, “Vortex-induced vibrations,” Annu. Rev. Fluid Mech., vol. 36, pp. 413–455, 2004.
[14] S. Mittal et al., “Lock-in in vortex-induced vibration,” Journal of Fluid Mechanics, vol. 794, pp. 565–594, 2016.
[15] D. Sabino, D. Fabre, J. Leontini, and D. L. Jacono, “Vortex-induced vibration prediction via an impedance criterion,”
Journal of Fluid Mechanics, vol. 890, 2020.
[16] P. D. d. S. Conceição, “Numerical simulation of two-degree-of-freedom vortex induced vibration in a circular cylinder
with openfoam,” Ph.D. dissertation, 2016.
[17] S. Lemaire, G. Vaz, M. Deij-van Rijswijk, and S. R. Turnock, “On the accuracy, robustness and performance of
high order interpolation schemes for the overset method on unstructured grids,” International Journal for Numerical
Methods in Fluids, 2021.
[18] M. Breuer, G. De Nayer, M. Münsch, T. Gallinger, and R. Wüchner, “Fluid–structure interaction using a partitioned
semi-implicit predictor–corrector coupling scheme for the application of large-eddy simulation,” Journal of Fluids and
Structures, vol. 29, pp. 107–130, 2012.
[19] W. J. Horne and K. Mahesh, “A massively-parallel, unstructured overset method to simulate moving bodies in turbulent
flows,” Journal of Computational Physics, vol. 397, p. 108790, 2019.
[20] J. H. Ferziger, M. Perić, and R. L. Street, Computational methods for fluid dynamics. Springer, 2002, vol. 3.
[21] M. Buchmayr, Development of Fully Implicit Block Coupled Solvers for Incompressible Transient Flows. TU-Graz,
2014.
[22] T. Do, L. Chen, and J. Tu, “Numerical simulations of flows over a forced oscillating cylinder,” 2007.
[23] S. Singh and S. Mittal, “Vortex-induced oscillations at low reynolds numbers: hysteresis and vortex-shedding modes,”
Journal of Fluids and Structures, vol. 20, no. 8, pp. 1085–1104, 2005.
[24] T. Prasanth and S. Mittal, “Vortex-induced vibrations of a circular cylinder at low reynolds numbers,” Journal of
Fluid Mechanics, vol. 594, p. 463, 2008.
[25] F. Moukalled, L. Mangani, M. Darwish et al., The finite volume method in computational fluid dynamics. Springer,
2016, vol. 113.
[26] P. Tisovska, “Description of the overset mesh approach in esi version of openfoam,” Proceedings of the CFD with
OpenSource Software; Nilsson, H., Ed, 2019.
[27] A. Dullweber, B. Leimkuhler, and R. McLachlan, “Symplectic splitting methods for rigid body molecular dynamics,”
The Journal of chemical physics, vol. 107, no. 15, pp. 5840–5851, 1997.
[28] N. M. Newmark, “A method of computation for structural dynamics,” Journal of the engineering mechanics division,
vol. 85, no. 3, pp. 67–94, 1959.
[29] C. H. Williamson, “Vortex dynamics in the cylinder wake,” Annual review of fluid mechanics, vol. 28, no. 1, pp.
477–539, 1996.
[30] T. K. J. Benjamin, H. Hesse, and P. C. Wang, “Wing-tail interaction under forced harmonic pitch,” in AIAA AVIATION
2021 FORUM, 2021, p. 2517.
[31] T. Morse and C. Williamson, “Prediction of vortex-induced vibration response by employing controlled motion,”
Journal of Fluid Mechanics, vol. 634, p. 5, 2009.

You might also like