Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

A Volume Penalization method coupled with the Lattice Boltzmann ap-

proach for modeling Fluid-Structure Interaction


Erwan Liberge
LaSIE – UMR 7356 CNRS – La Rochelle Université, France, erwan.liberge@univ-lr.fr

Claudine Béghein
LaSIE – UMR 7356 CNRS – La Rochelle Université, France

Abstract. In this paper, two different approaches to calculating fluid forces are compared within the
framework of the Volume Penalty-Lattice Boltzmann (VP-LBM) method. The first method, the mo-
mentum exchange method, uses the variation of distribution functions near the fluid-solid interface,
while the second, the stress integration method, allows the direct integration of fluid forces on this in-
terface. Applied to the VP-LBM, which consists of penalizing the solid in the LBM, these two methods
lead to significantly different results. The tests are carried out to study on the one hand the lift and
drag coefficients of a Naca 0012 profile at different angles of attack, at Reynolds number 1000, and
for a foil on imposed motion (displacement and rotation) in a flow

Keywords: Lattice Boltzmann method; Volume penalization.

NOMENCLATURE

1 INTRODUCTION

Many efficient methods and algorithms for modeling the Fluid-Structure Interaction (FSI) have been
developed in the recent past, always with the joint concern of improving the accuracy and speed of the
calculations. A classical approach consists in coupling a fluid solver for the Navier-Stokes equations
with a structure solver, the fluid solver being obtained by a classical discretization method, such as
the finite element or finite volume method. In this paper, we propose to use the Lattice Boltzmann
Method (LBM) as a fluid solver for numerical simulation of Fluid-Structure Interaction.

The LBM method has been successfully developed for fluid mechanics since the 1990s and has
established itself as an efficient alternative computational method. Based on the Boltzmann equation,
the calculation is performed on a mesoscopic scale to determine macroscopic quantities such as the
velocity field of a flow or the pressure field. One of the interests of the LBM lies in its simplicity of
programming and the low computational time if the algorithm is solved on a graphics card (GPU) (Fan
et al., 2004).

In previous papers (Benamour et al., 2015; Benamour et al., 2017; Benamour et al., 2020), we pro-
posed to couple the Volume Penalization (VP) method and the LBM, under the name VP-LBM. The
Volume Penalization method consists in extending the Navier-Stokes equations to the entire domain
(fluid and solid) and adding a volume penalization term to account for structure. The approach can be
considered as a mixture between the PSBB (Partially Saturated Bounce-Back) (a specific approach
to the LBM method, Bounce Back means that a particle bounces on a boundary) and IB (Immersed
Boundary) methods. However, the volume penalisation method does not require the costly step of
calculating the volume fraction of the solid domain near the interface as in the PSBB methods. The
difference with the IB methods is that the VP method uses a volume force, instead of local forces on
the Lagrangian markers, which reduces the computational time. The capability of penalty methods
for fluid-structure interaction problems has been demonstrated by Destuynder and Liberge (2021). In
previous work, Benamour et al. have shown that VP-LBM performs well for stationary solids. In this
paper, we propose to present the performance of VP-LBM for the interaction between fluid flows and
moving foil.

2 Governing equations

In this section, the numerical models are exposed. The following notations are used : ρ and u are the
macroscopic density and velocity, and bold characters denote vectors.

2.1 VOLUME PENALIZATION

Let us consider a fluid domain Ωf , a solid domain Ωs , Γ the fluid-solid interface, and let us note
Ω = Ωf ∪ Ωs ∪ Γ. The Volume Penalization (VP) method consists in extending the Navier-Stokes
equations to the whole domain Ω, and considering the solid domain as a porous medium with a
very small permeability. The method was introduced by Angot et al. (1999) and already applied to
macroscopic equations for moving bodies (Kadoch et al., 2012). The small permeability of the solid
domain is modelled using a penalization coefficient, hence the desired boundary conditions at the fluid-
solid interface are naturally imposed. With this method, the incompressible Navier-Stokes equations
are written as follows :
∇·u=0
∂u 1 χΩs (1)
+ u · ∇u = − ∇p + ν∇2 u − (u − us )
∂t ρ η

where ß
1 if x ∈ Ωs (t)
χΩs (x, t) = ; η  1 penalization factor (2)
0 otherwise
u denotes the velocity field, p is the pressure field, ρ and ν are the density and the viscosity of the
χΩs
fluid. F = (u − us ) is the penalization term, and us is the velocity field in the solid domain.
η

2.2 Lattice Boltzmann method

Based on the Boltzmann equation (Eq. (3)) proposed in the context of the Kinetic Gaz Theory by
L. Boltzmann in 1870, the Lattice Boltzmann Method has been successfully used to model fluid flow
since the 90’s.
∂f
+ c · ∇x f = Ω (f ) (3)
∂t
This equation models the transport of f (x, t, c), a probability density function of particles with the
velocity c at location x and time t. Ω (f ) is the collision operator. The link between the Boltzmann
equation and the Navier-Stokes equations is well-known since the Chapmann-Enskog expansion pro-
posed in 1915.
The Lattice Boltzmann method considers the discretization of equation (3) according to space and
velocity and leads to the following discretized equations :

fα (x + cα 4t, t + 4t) − fα (x, t) = Ωα (f ) + 4tFα (4)

where fα (x, t) = f (x, cα , t), Fα is a forcing term related to the discrete velocity cα (Guo et al., 2002)
.


 (0, 0)  α=0

  π  π 
cα = cos (α − 1) , sin (α − 1) c α = 1, 2, 3, 4 (5)
   2π   2 π  √
 cos (2α − 9)

, sin (2α − 9) 2c α = 5, 6, 7, 8
4 4
4x
Where c = . Usually 4x = 4y = 4t = 1 are chosen.
4t
4x c6 c2 c5

4y c0
c3 c1

c7 c4 c8

Figure 1. Discrete velocities of the D2Q9 model

The first model proposed by Bhatnagar et al. (1954) is the BGK model which is based on a linear
collision operator with a single relaxation time :
1
Ωα (f ) = − (fα (x, t) − fαeq (x, t)) (6)
τ
where f eq is the equilibrium function,
Ç å
cα · u uu : cα cα − c2s I
fαeq = ωα ρ 1 + + , (7)
c2s 2c4s
c
ωα = {4/9, 1/9, 1/9, 1/9, 1/9, 1/36, 1/36, 1/36, 1/36}, cs = √ and τ is the non dimensional relaxation
3
time which is linked to the fluid viscosity as follows.
Å ã
1
ν = c2s 4t τ − (8)
2

Finally, the macroscopic quantities are computed according to the following expressions :
X X 4t
ρ= fα ρu = cα fα + ρF (9)
α α 2

In the present approach, the volume penalization term is added :


X 4t χΩs
ρu = cα fα − ρ (u − us ) (10)
α 2 η

To avoid instabilities, the term including u in the penalization force is moved to the left hand side of
equation (10). This leads to the modified update step to compute the macroscopic velocity field :
X 4t χΩs
cα fα + ρus
α 2 η
u= (11)
4t χΩs
ρ+ ρ
2 η
In the fluid domain, where χΩs = 0 the classical LBM equation is obtained whereas in the solid domain,
where χΩs = 1, equation (11) forces the velocity field to approach us .

2.3 FLUID FORCES COMPUTATION

Angot et al. Angot et al., 1999 proposed in a context of an integral formulation of the volume penal-
ization problem to compute the fluid forces with the following formula :
Z
Ff = lim u − us dΩ (12)
η−→∞ Ωs

The formula (12) works with finite element or finite volume methods, but fails on our computational
tests. We present in the following the two classical method used in LBM to compute fluid forces.
2.3.1 MOMENTUM EXCHANGE METHOD (MEM)

The fluid forces are computed with the momentum exchange method (MEM) proposed by Wen et al.
(2014). We note xf a boundary node in the fluid domain and xs the image of this boundary node
through the solid interface by a lattice velocity cα , also called incoming velocity( cf. Fig 2). The
intersection point between the fluid-solid interface and the link xf − xs is xΓ , and the outgoing lattice
velocity is denoted cα = −cα .

fluid

• xΓ
xf
• solid


xs

Figure 2. Curved interface on a square lattice : example of a fluid boundary node xf , its image in the
solid domain xs , and the intersection point xΓ located on the interface

The local force at xΓ is computed using the following expression :

F (xΓ ) = (cα − uΓ ) f˜α (xf ) − (cα − uΓ ) f˜α (xs ) (13)

and the total fluid force acting on the solid domain is :


X
Ff = F (xΓ ) (14)

The torque is obtained with X


Tf = (xΓ − xG ) × F (xΓ ) , (15)
with xG the coordinates of the gravity center of the body.

Giovacchini and Ortiz Giovacchini and Ortiz, 2015 showed that the MEM does not depend of the way
the boundary conditions at the solid domain are implemented.

2.3.2 STRESS INTEGRATION METHOD (SIM)

This method is more intuitive in computational fluid dynamics, and consists in integrating the fluid
stress tensor onto the structure structure:
Z Z
Ff = σ · ndS and Tf = r × σ · ndS (16)
∂Ωs ∂Ωs

with Ä ä
σ = −pId + ν ∇u + (∇u)T
ã ÇX å
(17)
Å
2 1 eq
= −ρcs Id − 1 − cα ⊗ cα (fα − fα )
2τ α

and n is the outward normal to the solid interface.


The fα are extrapolated from the closest point in the relevant direction (close to n) in the fluid domain
to the integration points xi located on the surface. Finally, the equation (16) becomes :
X
Ff = Si σ (xi ) · ni (18)
i

Si and ni are the integration surface and the outward normal at integration point xi .

3 APPLICATION

3.1 NACA AIRFOIL

The first application is the study of the NACA 0012 airfoil with different angle attack values at Reynolds
number 1000. This case is well-documented in literature, and the different ME or SI results for VP-LBM
are compared to those obtained by (Ilio et al., 2018; Kurtulus, 2015; Liu et al., 2012).

Liu et al. (2012) use the finite elements method combined with a fine mesh to give accurate numerical
results. Kurtulus (2015) proposes a very complete study, using finite volume method and a lot of
data to compare. Ilio et al. (2018) combine the standard LBM with an unstructured finite volume
formulation in the so-called hybrid lattice Boltzmann method. They have used an overlap between
a standard LBM approach on the whole domain and an unstructured body-fitted grid model where a
finite-volume lattice Boltzmann formulation is applied. This approach has led to very accurate results
close to the body. However, no information has been given on fluid forces calculation. It looks like a
Stress Integration method because the macroscopic values have directly been taken from the body
fitted mesh.

The figure 3 represents the computational domain. Let C be the chord of the NACA 0012. The airfoil

inlet C symmetry outlet


α
U0 H

L1 L2

Figure 3. Shematic of the computational domain around the NACA airfoil

is placed at 4C from the inlet and 9C from the outlet. The height of the computational domain is 7C,
and the NACA is 3.5C from the bottom.

A constant velocity profile has been imposed at the inlet using the classical half-way Bounce-Back
method, and the outflow boundary condition at outlet has been modeled using the convective condition
(Yang, 2013). This condition makes it possible to reduce the distance between the airfoil and the
extreme limit of the computational domain downstream the immersed body. Symmetry boundary
conditions ( u · n = 0) have been imposed at the other boundaries.

The computations have been carried out using the following parameters (in lattice units):

C = 278, U0 = 0.0599, τ = 0.55

Note that τ is close to the stability limit for LBM, but this makes it possible to decrease C and then the
size of the computational problem and also the computational time. Ilio et al. (2018) have used 512
nodes in the chord, and a larger computational domain. However, our interest in VP-LBM solved in
CUDA being the computational time, we try to obtain a good qualitative result without too expensive
computing resources. This is why this set of parameters have been used.

For the Stress Integration method, 849 integrations points have been used. Note that the number of
integration points has been chosen arbitrary. Increasing their number increases the accuracy of the
computation, but not significantly in this case.

(a) Lift coefficient (b) Drag coefficient

Figure 4. Lift and Drag coefficient versus angle. The VP-LBM’s results computed with Stress Integra-
tion and Momentum Exchange methods are compared with literature

The drag and lift coefficients are plotted in figure 4. The VP-LBM gives a good prediction of these
values compared to literature.

For α ≤ 8◦ a steady solution has been obtained, which leads to a small increase in drag. After, up
to 24◦ , a periodic vortex shedding is observed (Figs (5b) and (5c). During this phase, the increase in
drag is constant, and the lift increase remains fairly stable. An irregularity in the lift coefficient appears
at 26◦ . This stall phenomena is well captured with the VP-LBM approach, and the numerical value is
also well computed with ME as well SI.

(a) α = 7◦ (b) α = 10◦

(c) α = 24◦ (d) α = 26◦

Figure 5. Streamlines around NACA 0012 for various angles : α = 7◦ , 10◦ , 24◦ and α = 24◦

Note that the values obtained here are a little higher than those obtained by Kurtulus (2015) and Ilio et
al. (2018), but smaller than those obtained by Liu et al. (2012). Due to the enclosure of our results with
those of Liu et al. (2012) and Ilio et al. (2018), they can be considered as validated. Considering that
the work of Ilio et al. (2018) is the reference one, because a finer mesh is used, the Stress Integration
method gives better results than the Momentum Exchange method. SI allows a better approximation
of the surface, with a direct discretization of the solid boundary, and a true outward normal, while MEM
used a staircase approximation. The limit of SI, which is an extrapolation of the distribution values on
the boundary, seems to have no consequences in these cases.

The lift forces are quite similar, ME or SI does not affect the results. The drag has been slightly over-
estimated with ME, probably because of the approximation of the computational boundary induced
by this method.

3.2 Heaving and piching airfoil

The second application comes from the device proposed by Kinsey and Dumas (2008) of a foil heaving
and pitching to extract energy from an incoming flow. The vertical displacement y (t) and the rotation
θ (t) of the foil are imposed according the following expressions :

y (t) = Y0 sin (γt + φ) and α (t) = α0 sin (γt) (19)

In the following we compare the lift, drag and pitching moment we obtained with the method we
proposed for the following value of the parameters : H0 = C, α0 = 76.33 deg, Re = 1100.
The results are presented in figure 6.The imposition of displacement and a large angle of incidence

(a) Imposed displacement (b) Drag coefficient

(c) Lift coefficient (d) Pitching coefficient

Figure 6. Comparison of drag, lift and pitching moment coefficients over time for a periodic cycle

on the foil during its descent and ascent gives very irregular lift (Fig. 6c), drag (Fig. 6b) and piching
coefficients (Fig 6d). This irregularity is well captured by our proposed method. As in the previous
example, the result given by the stress integration method is of better quality than that obtained by the
MME. However, since the differences are small, this method of calculating stresses is still relevant.
4 CONCLUSIONS

The volume penalization method coupled to the Lattice Boltzmann method (VP-LBM) has been suc-
cessfully applied to two new cases. The available methods for the calculation of fluid loads were
discussed. The VP-LBM method has shown its ability to reproduce the complex physics of an air-
foil at different angles of attack and the stall phenomenon has been well captured. Similarly for the
following example, the complex physics of the flow imposed by the abrupt movements of the foil are
well calculated numerically by the method. For these two applications, both the momentum exchange
(ME) and stress integration (SI) methods perform well, although the SI method seems slightly more
accurate.

REFERENCES

Angot, P., C.-H. Bruneau, and P. Fabrie (1999). “A penalization method to take into account obstacles
in incompressible viscous flows”. In: Numerische Mathematik 81.4, pp. 497–520. DOI: 10 . 1007 /
s002110050401.

Benamour, M., E. Liberge, and C. Béghein (2015). “Lattice Boltzmann method for fluid flow around
bodies using volume penalization”. In: International Journal of Multiphysics 9.3, pp. 299–315. DOI:
10.1260/1750-9548.9.3.299.

— (2017). “A new approach using lattice Boltzmann method to simulate fluid structure interaction”. In:
Energy Procedia 139, pp. 481–486. DOI: 10.1016/j.egypro.2017.11.241.

— (2020). “A volume penalization lattice Boltzmann method for simulating flows in the presence of
obstacles”. In: Journal of Computational Science. ISSN: 1877-7503. DOI: https : / / doi . org / 10 .
1016/j.jocs.2019.101050.

Bhatnagar, P.L., E.P. Gross, and M. Krook (1954). “A model for collision processes in gases. I. Small
amplitude processes in charged and neutral one-component systems”. In: Physical Review 94.3,
pp. 511–525. DOI: 10.1103/PhysRev.94.511.

Destuynder, Philippe and Erwan Liberge (2021). “A few remarks on penalty and penalty-duality meth-
ods in fluid-structure interactions”. In: Applied Numerical Mathematics 167, pp. 1–30. ISSN: 0168-
9274. DOI: https://doi.org/10.1016/j.apnum.2021.04.017.

Fan, Z. et al. (2004). “GPU cluster for high performance computing”. In: IEEE/ACM SC2004 Confer-
ence, Proceedings, pp. 297–308.

Giovacchini, Juan P. and Omar E. Ortiz (Dec. 2015). “Flow force and torque on submerged bodies
in lattice-Boltzmann methods via momentum exchange”. In: Phys. Rev. E 92 (6), p. 063302. DOI:
10.1103/PhysRevE.92.063302.

Guo, Zhaoli, Chuguang Zheng, and Baochang Shi (Apr. 2002). “Discrete lattice effects on the forc-
ing term in the lattice Boltzmann method”. In: Physical Review E 65.4, p. 046308. DOI: 10.1103/
PhysRevE.65.046308.

Ilio, G. Di et al. (2018). “Fluid flow around NACA 0012 airfoil at low-Reynolds numbers with hybrid
lattice Boltzmann method”. In: Computers & Fluids.

Kadoch, B. et al. (2012). “A volume penalization method for incompressible flows and scalar advection-
diffusion with moving obstacles”. In: Journal of Computational Physics 231.12, pp. 4365–4383. DOI:
10.1016/j.jcp.2012.01.036.
Kinsey, T. and G. Dumas (2008). “Parametric Study of an Oscillating Airfoil in a Power-Extraction
Regime”. In: AIAA Journal 46.6, pp. 1318–1330. DOI: 10.2514/1.26253.

Kurtulus, Dilek Funda (2015). “On the Unsteady Behavior of the Flow around NACA 0012 Airfoil with
Steady External Conditions at Re=1000”. In: International Journal of Micro Air Vehicles 7.3, pp. 301–
326. DOI: 10.1260/1756-8293.7.3.301.

Liu, Yan et al. (2012). “Numerical bifurcation analysis of static stall of airfoil and dynamic stall under
unsteady perturbation”. In: Communications in Nonlinear Science and Numerical Simulation 17.8,
pp. 3427–3434. ISSN: 1007-5704. DOI: https://doi.org/10.1016/j.cnsns.2011.12.007.

Wen, Binghai et al. (2014). “Galilean invariant fluid–solid interfacial dynamics in lattice Boltzmann
simulations”. In: Journal of Computational Physics 266, pp. 161–170. DOI: 10.1016/j.jcp.2014.
02.018.

Yang, Z. (2013). “Lattice Boltzmann outflow treatments: Convective conditions and others”. In: Com-
puters and Mathematics with Applications 65.2, pp. 160–171. DOI: 10.1016/j.camwa.2012.11.012.

You might also like