Mathematical and Numerical Models For Fluid-Structure Interaction
Mathematical and Numerical Models For Fluid-Structure Interaction
fluid-structure interaction
Fabio Nobile
MOX, Dipartimento di Matematica, Politecnico di Milano
via Bonardi 9, 20133 Milano, fabio.nobile@polimi.it
Introduction
Examples:
I The effect of the wind on civil structures such as bridges, suspended
cables, tall structures like cooling towers or skyscrapers
I The action of the air on aeronautic structures (aeroelasticity, flutter)
I Aero-acoustics
I Vibrations in vessels. Water-hammer effect
I Biomechanics: blood flow in large arteries
I Effect on dams of water movement in a reservoir; sloshing of a fluid
in a tank.
I ...
FSI in aeroelasticity
A Piper PA-30 Twin Commanche,
known as NASA 808, was used at
the NASA Dryden Flight Research
Center as a rugged workhorse in a
variety of research projects associ-
ated with both general aviation and
military projects. The movie shows
a tail flutter test done in 1966.
Df d
Material derivative: (x, t) = f (x(t), t)
Dt dt
Vt
n
f
Z Z Z t
d n
%f u dx = f dx + t dA
dt Vt Vt ∂Vt
where
I f are the external volume forces per unit volume (e.g. gravity∗%f )
I t are the internal surface tractions due to the interaction with the
surrounding fluid particles.
The Cauchy postulate asserts that t = t(x, n, t) depends only on the
unit outward normal vector n to Vt in x and not on the particular
volume Vt considered.
Moreover, the Cauchy theorem asserts that t is necessarily a linear
function of n, i.e. there exists a tensor σ f , (called the Cauchy stress
tensor), such that t = σ f · n.
Finally, σ f has to be symmetric if no concentrated couples exist in
the fluid.
Conservation of momentum
By applying the Reynols transport theorem + the divergence theorem +
arbitrariness of Vt , we get
Momentum conservation equation (first form)
∂%f u
+ div(%f u ⊗ u) = f + div(σ f ), in Ωf , t > 0
∂t
Observe that
∂%f u ∂%f ∂u
+ div(%f u ⊗ u) = u + %f + %f u · ∇u + u div(%f u)
∂t ∂t
∂t
∂%f ∂u
=u + div(%f u) +%f + u · ∇u
∂t ∂t
| {z }
=0(continuity eq.)
Γin
u=0
Γout
I u = g on Γin (inflow velocity profile)
σ.n=0
u=g
I u = 0 on Γwall (perfect adherence to wall)
u=0
Γwall
I σ f · n = 0 on Γout (free stress condition)
Potential flows and Bernoulli law
We consider now an inviscid, incompressible fluid subject to conservative
forces f = ∇V .
If the flow is initially irrotational, it will stay so for all times (Kelvin
theorem)
Assume curl u = 0. Then, for a simply connected domain Ωf , there exists
a potential φ such that u = ∇φ.
I Continuity eq.: div u = 0 =⇒ ∆φ = 0
∂∇φ
Momentum eq.: %f ∂t + u · ∇u + ∇p = ∇V
I
identity u · ∇u = 12 ∇|u|2 :0
using the vector − u ×
curl
u
=⇒ ∇ %f ∂φ 1 2
∂t + %f 2 |u| + p − V = 0
∂φ 1
Bernoulli law %f + %f |u|2 + p − V = const
∂t 2
Potential equations
(
∆φ = 0
∂φ
p = c + V − %f ∂t + 12 |∇φ|2
Effects of viscosity
When considering the flow over a wall, the viscosity forces the fluid
particles to adhere to the wall −→ boundary layers effects. Let
I U: far field velocity; L: characteristic length of the wall
I Re = %f UL/µ: Reynolds number; measures the relative importance
of inertia effects (U ∗ L) versus viscous effects (µ/%f ).
Flow over a flat wall (∂x p = 0)
q
Lx
I boundary layer of thickness z = Re
(increasing along the wall – Blasius
solution) z
y
slope ~
drag forces
x
I shear forces (drag) on the wall
(σ f · n) · ex ≈ µ ∂u
∂z
x
inertial region
E(k)
=c
k ε 2/3 −
k 5/3
production dissipation
of energy tranfer of energy of energy
−1 −1 log(k)
L lk
(integral scale) (Kolmogorov scale)
The decay Ê (k) ∼ 2/3 k −5/3 is one of the main results of Kolmogorov
theory of turbulence, valid for homogeneous isotropic turbulence
Reynolds Averaged Navier Stokes (RANS) models
Since a Direct Numerical Simuation is most of the times impossible, the
idea behind RANS models is to “formally” average over many repetitions
of the experiment/simulation so as to filter out the small fluctuations at
the inertia and dissipation scale.
Split the fluid motion in u = ū + u0 , p = p̄ + p 0 , with
I ū: mean velocity; p̄: mean pressure (mean = ensamble average)
0 0
I u , p : fluctuations around the mean
∂ū
%f + div(%f u ⊗ u) − 2µ div D(ū) + ∇p̄ = f̄
Averaged N-S ∂t
div ū = 0
τ R = %f (ū ⊗ ū − u ⊗ u) = −%f u0 ⊗ u0
Then
RANS equations
% ∂ū + ū · ∇ū − div 2µD(ū) + τ R + ∇p̄ = f̄
f
∂t
div ū = 0
2
Reynolds stresses τ R = 2µT D(ū) − %f kI
3
turbulent viscosity µT = Cµ k 2 /ε
Navier-Stokes equations
∂u
%f + %f u · ∇u − div σ f (u, p) = f f , in Ω, t > 0,
∂t
div u = 0,
in Ω, t > 0,
u = g, on ΓD , t > 0,
σ f · n = d, on ΓN , t > 0,
u = u ,
0 in Ω, t = 0
with σ f (u, p) = 2µD(u) − pI
Weak formulation
Let us define the following functional spaces:
Z Z Z *0 Z
= 2µD(u) : ∇v − pI : ∇v − (σ · n) · v −
f (σ f · n) · v
Ω Ω | {z } Γ
D ΓN | {z }
=p div v =d
Weak formulation
∀(v, q) ∈ V0 × Q
Weak solutions exist for all time ([Leray ’34], [Hopf ’51]). Uniqueness is
an open issue in 3D.
Energy estimate
1 ∂|u|2
Z Z Z
∂u d 1
• %f ·u= %f = %f |u|2
Ω ∂t Ω 2 ∂t dt Ω 2
| {z }
kinetic energy Ek
Z Z
1
Z
:0 Z 1
• %f (u · ∇u) · u = %f u · ∇|u|2 = − %f
(div
u)|u|2 + %f |u|2 u · n
Ω Ω 2 Ω Γ 2
| N {z }
Z Z flux of Ek through ΓN
Vh ⊂ V; Vh0 = Vh ∩ V0 ; Qh ⊂ Q
∀(v, q) ∈ Vh0 × Qh
Algebraic formulation (b)
φj
Nu
I {φi }i=1 : basis of Vh0 ψl
φi
(b) N b
I {φ } u : basis of Vh \ Vh0 (shape functions
j j=1
corresponding to boundary nodes)
Np
I {ψl }l=1 : basis of Qh
Expand the solution (uh (t), ph (t)) on the finite element basis
b
Nu Nu
(b)
X X
uh (x, t) = ui (t)φi (x) + gi (t)φj (x)
i=1 j=1
| {z }
=gh known term
Np
X
ph (x, t) = pi (t)ψl (x)
l=1
M dU + AU + N(U)U + B T P = F (U),
u
dt t>0
BU = Fp ,
with
Z Z
Mij = %f φj φi mass matrix Aij = 2µD(φj ) : ∇φi , stiffness matrix
Ω Ω
Z
N(U)ij = uh · ∇φj · φi , !! depends on U; non linear term
Z Ω
Bli = − ψl div φi , divergence matrix
Ω
Z Z Z „ « Z
∂gh
(Fu (U))i = f · φi + d · φi − %f + uh · ∇gh · φi − 2µD(gh ) : ∇φi
Ω ΓN Ω ∂t Ω
Z
(Fp )l = div gh ψl
Ω
A simpler problem – Stokes equation
Let us consider for the moment the steady state Stokes problem
−2µ div D(u) + ∇p = f, in Ω,
div u = 0, in Ω,
u = g, on ΓD ,
(2µD(u) − pI) · n = d, on ΓN ,
with
Z Z
a(u, v) = 2µD(u) : ∇v, b(v, p) = − p div v,
Z Ω Z Ω
The algebraic system has the typical structure of a saddle point problem
Spurious pressure modes
If these exists a pressure ph∗ such that
this pressure will not be seen by the first equation of (*). It follows that,
if (uh , ph ) is a solution of (*), then also (uh , ph + ph∗ ) is a solution to the
system and we loose uniqueness of the pressure.
Such a pressure ph∗ is called a spurious pressure mode
A necessary and sufficient condition to avoid the presence of spurious
pressure modes is that the finite elements spaces Vh × Qh satisfy the
b(vh , qh )
inf-sup condition: inf sup ≥ βh > 0
qh ∈Qh vh ∈Vh0 kvh kH 1 kqh kL2
P2 / P1 Pbubble
1 / P1 Q2 / Q1
P1 / P1 P1 / P0 Q1 / Q1
Examples of spurious modes
u=1
Couette flow:
σ.n=0
σ.n=0
exact solution: u1 = y ; u2 = p = 0
u=0
Pressure Pressure
IsoValue IsoValue
-9.94024 -3.7491
-8.52021 -3.21351
-7.57352 -2.85645
-6.62683 -2.4994
-5.68015 -2.14234
-4.73346 -1.78528
-3.78677 -1.42822
-2.84008 -1.07116
-1.8934 -0.714103
-0.94671 -0.357045
-2.32106e-05 1.34456e-05
0.946664 0.357072
1.89335 0.71413
2.84004 1.07119
3.78673 1.42825
4.73341 1.78531
5.6801 2.14236
6.62679 2.49942
7.57347 2.85648
9.94019 3.74913
u0 = u0 , un = gn on ΓD − p n n + 2µD(un ) · n = dn on ΓN
In algebraic form this leads to a linear system to solve at every time step,
of the form
1
" 1 #" # " n #
∆t M + A + N(U
n−1
) B T Un Fu + ∆t MUn−1
=
B 0 Pn Fnp
Observe the structure of the matrix, equal to the one for the Stokes
problem.
Temporal discretization – second order scheme
As an example of second order scheme we consider a second order
Backward differentiation (BDF2) with semi-implicit treatment of the
convective term:
Resulting system:
M
" 3M #" # " #
2∆t + A + N(2U
n−1
− Un−2 ) B T Un Fnu + 2∆t (4U
n−1
− Un−2 )
=
B 0 Pn Fnp
Solution of the linear system
After temporal discretization (Implicit Euler, BDF2, ...) we are led to the
linear system
C B T Un
F
= u
B 0 Pn Fp
C = αM + A + N(U∗ ) with α, U∗ depending on the time marching
scheme chosen.
1st eq. C U n + B T Pn = F u Un = C −1 Fu − B T Pn
=⇒
−1 T n −1
2nd eq. BUn = Fp =⇒ |BC {z B } P = BC Fu − Fp
| {z }
Σ χ
with
Z
Mp : mass pressure matrix (Mp )ij = ψj ψi
ZΩ
Kp : stiffness pressure matrix (Kp )ij = ∇ψj · ∇ψi + εψj ψi , ε1
Ω
C B T Un
F
= u
B 0 Pn Fp
∂u
NS eqs. %f + L1 u + L2 u =f
∂t |{z} |{z}
viscous+transport terms incompress. constraint
Split in
ũn − un−1
%f + L1 ũn = f n
Step I: ∆t (transport-diffusion eq.)
un | = 0
∂Ω
un − ũn
n
%f ∆t + ∇p = 0
(L2 -proj. on divergence
Step II: div un = 0
free functions)
n
u |∂Ω · n = 0
Projection methods
I Step II corresponds to the L2 -projection of ũn onto the divergence
free space Hdiv ,0 = {v ∈ L2 (Ω), div v = 0, v · n = 0}. Indeed
Z
ũn − un
Z Z * 0Z
*0
·v = − ∇p n ·v = p n n
∀v ∈ Hdiv ,0 , %f div v− p v · n = 0.
Ω ∆t Ω Ω ∂Ω
Z Z
n n
=⇒ ũ · v = u · v, ∀v ∈ Hdiv ,0
Ω Ω
n n−1
% ũ − u + u ∗,n
· ∇ũn
− div(2µD(ũn )) = f n − ∇p n−1
f
Step I ∆t
n
ũ = 0 on ∂Ω
(
%f
∆δp n = ∆t div ũn in Ω
Step II n
∂n δp = 0 on ∂Ω
∆t
Step III un = ũn − ∇δp n
%f
Step IV p n = p n−1 + δp n
Further improvements
I The incremental Chorin-Temam scheme can be written also for a
BDF2 discretization of the time derivative. In this case, the scheme
reads:
n
3ũ − 4un−1 + un−2
Step I %f + (2un−1 − un−2 ) · ∇ũn
2∆t
− div(2µD(ũn )) = f n − ∇p n−1
3 %f
Step II ∆δp n = div ũn
2 ∆t
I For this scheme it has been proved (see e.g. [Guermond ’99]) that
Kinematics
∂x
velocity : u(ξ, t) = (ξ, t) = ẋ(ξ, t) = η̇(ξ, t)
∂t
∂2x
acceleration : a(ξ, t) = 2 (ξ, t) = ẍ(ξ, t) = η̈(ξ, t)
∂t
Eulerian versus Lagrangian
x=x(ξ,t)
x=x(ξ,t)
dV0 dVt Consider an infinitesimal volume
dx3 dV0 = det(dξ 1 , dξ 2 , dξ 3 )
dξ 3 dξ 2 dx2
dx1
dξ 1
After deformation
dVt
Hence J = det(F) = dV0 measures the change of volumes.
Deformation of surface elements – Nanson’s formula
x=x(ξ,t)
n
Consider an infinitesimal surface element
n0 dx
dA0 with normal n0 and an infinitesimal
dξ dA vector dξ
dA0 infinitesimal volume dV0 = dξ T n0 dA0
After deformation
Conservation of momentum:
Z Z Z
d
%s u dx = f dx + tdA
dt Vt Vt ∂Vt
where we have introduced the nominal stress tensor (also called first
Piola-Kirchhoff stress tensor) σ 0s = Jσ s F−T . Observe that the
divergence divξ is taken with respect to the reference coordinate ξ.
σ 0s = Jσ s F−T
σ 0s n0 dA0 = σ s ndA
I1 (A)= tr A = λ1 + λ2 + λ3
1
(tr A)2 − tr A2
I2 (A)= = λ1 λ2 + λ1 λ3 + λ2 λ3
2
I3 (A)= det A = λ1 λ2 λ3
Incompressible materials (J = 1)
µ
I Neo-Hookean: W (C) = 2 (I1 (C) − 3)
∂W
σ 0s = 2F − pJF−T = µF − p Cof F
∂C
µ1 µ2
I Mooney-Rivlin: W (C) = 2 (I1 (C) − 3) − 2 (I2 (C) − 3)
∂W
σ 0s = 2F − pJF−T = µ1 + µ2 tr(FT F) F − µ2 FFT F − p Cof F
∂C
Compressible materials
I Saint-Venant Kirchhoff: W (C) = λ2 tr(E)2 + µ tr(E2 )
with E = 12 (C − I) and (λ, µ) Lamé constants
∂W ∂W
σ 0s = 2F =F = F (λ(tr E)I + 2µE)
∂C ∂E
Examples of constitutive relations
Quasi incompressible materials
We split the deformation gradient in F = Fiso Fvol where
1
I Fiso = J − 3 F: isochoric deformation, det Fiso = 1
1
I Fvol = J 3 I: volumetric deformation
2 2
Correspondingly: C = FT F = J 3 FT
iso Fiso = J Ciso , with det Ciso = 1.
3
Incompressible materials
(
%0s η̈ − divξ σ 0s (η, p) = f 0 , σ 0s = 2F ∂W
∂C − p Cof F
J=1
∂2η
Z Z Z Z
%0s 2 · φ dξ + σ 0s (η) : ∇ξ φ dξ = 0
f · φ dξ + d · φdA0
Ωs0 ∂t Ωs0 Ωs0 ΓN
Proof: Take φ = η̇
%0 d %0 d
Z Z
%0s η̈ · η̇ dξ = s (η̇)2 dξ = s kη̇k2L2
Ωs0 2 dt Ωs 2 dt
Z Z 0 Z
0 ∂W ∂W
σ s (η) : ∇ξ η̇ dξ = : Ḟ dξ = dξ
Ω s Ωs ∂F Ωs ∂t
0 0 0
Finite element approximation – compressible case
Consider a finite element space Vh ⊂ V (e.g. piecewise continuous
polynomials) defined on a suitable triangulation Th of Ωs0 .
Finite element formulation: find η h (t) ∈ Vh , η h = gh on ΓD , such that
Z 2 Z Z Z
∂ηh
%0s · φh + σ 0s (η h ) : ∇ξ φh = f0s · φh + d · φh dA0
Ωs0 ∂t 2 Ωs0 Ωs0 ΓN
Z Z Z
∂W
σ 0s (η) : ∇ξ φ dξ = 0
f · φ dξ + d · φdA0 , with σ 0s =
Ωs0 Ωs0 ΓN ∂F
and
L(η + δη; φ) − L(η; φ)
L0 (η; δη, φ) = lim
→0
Newton’s method
given η 0 , compute for k = 1, . . . until convergence
∂σ 0s ∂ 2 W
Z Z
L0 (η k ; δη, φ) = ( δF) : ∇ ξ φ = ( ∇ξ (δη)) : ∇ξ φ
Ωs0 ∂F ηk Ωs0 ∂F∂F η k
∂ 2 W ∂δηl ∂φi
XZ
=
Ωs0 ∂Fij ∂Flk η k ∂ξk ∂ξj
ijlk
Tangent problem
∂ 2 W
Z Z Z Z
0
( ∇ ξ (δη)) : ∇ ξ φ = f · φ + d · φ − σ 0s (η k ) : ∇ξ φ
Ωs0 ∂F∂F η k
Ωs0 ΓN Ωs0
(J(η) − 1)q = 0
Ωs0
with
∂ 2 W
Z Z
∂ Cof F
ak (δη, φ) = ( ∇ ξ (δη)) : ∇ ξ φ − p k
∇ ξ (δη) : ∇ξ φ
Ωs0 ∂F∂F ηk Ωs0 ∂F
Z
k
b (φ, δp) = − δp Cof F(η k ) : ∇ξ φ
Ωs0
We see from here that the tangent problem has the usual structure of a
saddle point problem (Stokes) with a constraint that resembles a
“divergence” constraint. At the finite element level, we have to take
compatible finite element spaces (i.e. that satisfy the inf-sup condition).
Time discretization - Newmark scheme
We can use any ODE solver to solve the system (1). A popular scheme is
the following Newmark scheme
ḋns = ḋn−1
s + ∆t(γ d̈ns + (1 − γ)d̈n−1
s ), (Taylor expansion for ḋs )
2
∆t
dns = dn−1
s + ∆t ḋn−1
s + (2β d̈ns + (1 − 2β)d̈n−1
s ), (Taylor expansion for ds )
2
In the Newmark scheme, we can use the expansions for dns and ḋns to
express the acceleration d̈ns in terms of the displacement dns .
1 1 1 − 2β n−1
d̈ns = dn − ζ n , with ζ n = (dn−1 + ∆t ḋn−1 )+ d̈s
β∆t 2 s β∆t 2 s s
2β
In this way, the algebraic system can be written in the displacement only
1
Ms dns + Ks (dns ) = Fns + Ms ζ n
β∆t 2
*0
∂ 2 W ∂ 2 W
∂W
δσ 0s
η=0
= 2δF
+ 2F δC= 4 D(δη)
∂C η=0 ∂C∂C η=0 ∂C∂C η=0
Leads to
! Z
2
∂2W
Z Z Z
0 ∂ δη 0
%s · φ + 4 D(δη) : ∇ ξ φ = δf · φ + δd · φ
Ωs0 ∂t 2 Ωs0 ∂C∂C Ωs0 ΓN
∂ 2 δη
%s − div σ s (δη) = f, in Ωs0 ,
∂t 2
σ s (δη) = λ(div δη)I + 2µD(δη)
[O84] R.W. Ogden. Nonlinear elastic deformations. Halsted Press [John Wiley
& Sons, Inc.], New York, 1984
[BW08] J. Bonet, R.D. Wood. Nonlinear continuum mechanics for finite
element analysis. Second edition. Cambridge University Press, 2008.
[BLM00] T. Belytschko, W.K. Liu, B. Moran. Nonlinear finite elements for
continua and structures. John Wiley & Sons, Ltd., Chichester, 2000.
[ZT00] O.C. Zienkiewicz, R.L. Taylor. The finite element method. Vol. 2.
Solid mechanics. Fifth edition. Butterworth-Heinemann, Oxford, 2000
Coupled fluid-structure problem
Eulerian versus Lagrangian description
We consider now an incompressible fluid interacting with an elastic
structure featuring possibly large deformations.
As we have seen,
I Fluid equations are typically written in Eulerian form
I Structure equations are typically written in Lagrangian form on the
reference domain Ωs0 .
Γt
Γ0
A) Ω ft Ω f0
(u,p)
x2 ζ2
x1 ζ1
Fluid structure coupling conditions
At the common interface Γt (evolving in time), we impose
I Continuity of velocity (kinematic condition)
I Continuity of the normal stress (dynamic condition)
Warning: the continuity of stresses on Γt involves the physical stresses,
i.e. the Cauchy stress tensors.
∂η
(cont. velocity) u(t, x) = (t, ξ), with x = Lt (ξ)
∂t
(cont. normal stress) σ f (u, p)nf = −σ s (η)ns
1
Remember the relation σ s (η) = J(η) σ 0s (η)FT (η) between the Cauchy
stress tensor and the nominal stress tensor.
n nf
Ω st
η
∂u
% f
+ %f div(u ⊗ u) − div σ f (u, p) = f f
∂t in Ωft (η),
div u = 0
∂2η
%0s − divξ σ 0s (η) = f0s , in Ωs0 ,
∂t 2
∂η
u(t, x) = (t, ξ), with x = Lt (ξ) on Γt
∂t
Z Z Z
∂u
%f + div(u ⊗ u) ·v+σ f : ∇v+div uq = f f ·v+ (σ f · nf ) · v
Ωft ∂t Ωft Γt
∂2η
Z Z Z
%0s · φ + σ 0s (η) : ∇ξ φ = f0s · φ + (σ 0s (η) · ns0 ) · φ
Ωs0 ∂t 2 Ωs0 Γ0
Global weak formulation - cont.
V ≡ {(v, q, φ) : v ◦ Lt = φ on Γ0 }
Z
∂u
%f + div(u ⊗ u) · v + σ f (u, p) : ∇v + div uq+
Ωft ∂t
∂2η
Z Z Z
%0s 2 · φ + σ 0s (η) : ∇ξ φ = ff · v + f0s · φ
Ω0s ∂t f
Ωt s
Ω0
∂η
+ coupling condition u ◦ Lt = ∂t on Γ0
Energy inequality
RT R
The term 2µ 0 Ωf D(u) : D(u) dΩ dt represents the energy dissipated
t
by the viscosity of the fluid
Energy Inequality
∂2η
Z Z
%0s 2 · φ + σ 0s (η) : ∇ξ φ = fs0 · φ + < Resf (u, p), Ext(φ) >
Ω0s ∂t s
Ω0
Linear elasticity
∂ 2 ηs
%s − div σ s (η s ) = fs in Ωs
∂t 2
σ s (η s ) = 2µs D(η s ) + λs (div η s )I
Coupling conditions
uf = η̇ s , σ f (uf , pf ) · nf = −σ s (η s ) · ns on Γ
Boundary and initial conditions
s
ΓD
f
ΓD
nf
ns solid
fluid
Ωs
Ωf
Γ
s
ΓN
f
ΓN
Boundary conditions:
uf = ġf on ΓD,f
σ · n = d
f f f on ΓN,f
η s = g s on ΓD,s
σ s · ns = ds on ΓN,s
Initial conditions
(
uf |t=0 = uf ,0 in Ωf
η s |t=0 = η s,0 , η̇ s |t=0 = η̇ s,0 on Ωs
Linear elasticity in mixed form
It is actually more convenient to write a mixed formulation for the
elasticity equations as well, by introducing a pressure
ps = −λs div η s
Linear elasticity
∂ 2 ηs
Z Z Z Z
%s · v + σ s : ∇v = f s · v + ( σ s ns ) · v + ds · v
Ωs ∂t 2 Ωs Γ | {z } ΓN,s
=−σ f nf
Global weak formulation
Similarly, we multiply both the continuity equation of the fluid and the
equation defining the pressure for the solid by the same test function
q ∈ Qfs and sum them together.
global weak formulation
Find (uf (t), pf (t), η s (t), ps (t)) ∈ Vf × Qf × Vs × Qs , with uf = ġf on
ΓD,f , η s = gs on ΓD,s such that
Z
∂uf
% · v + 2µ D(u ) : ∇v − p div v
f f f f
∂t
Ωf Z 2
∂ ηs
· ∇v −
+ % s v + 2µ s D(η s ) : p s div v
ZΩs ∂t 2 Z Z Z
= ff · v + fs · v + df · v + ds · v ∀v ∈ Vfs
Ωf Ωs ΓN,f ΓN,s
Z Z
∂η s 1 ∂ps
∀q ∈ Qfs
div uf q + div q+ q =0
Ωf
Ωs ∂t λs ∂t
uf = ∂ηs
on Γ
∂t
u (0) = u , η (0) = η , ∂ηs (0) = η̇
f f ,0 s s,0 ∂t s,0
Global weak formulation
We define now the following global functions:
(
uf in Ωf
velocity : u = ∂ηs
∂t in Ωs
(R t
0 f
u (s) ds + Ext(η s,0 ) in Ωf
displacement : η(t) =
η s (t) in Ωs
where Ext(η s,0 ) is a suitable extension of η s,0 in the fluid domain
(
pf in Ωf
pressure : p =
ps in Ωs
( (
uf ,0 in Ωf Ext(η s,0 ) in Ωf
initial conditions : u0 = η0 =
η̇ s,0 in Ωs η s,0 in Ωs
( ( (
ff in Ωf df in ΓN,f ġf in ΓD,f
forcing terms : f = d= ġ =
fs in Ωs ds in ΓN,s ġs in ΓD,s
Global weak formulation
... and the following bilinear forms: ∀v, w ∈ Vfs , q ∈ Qfs
Z Z
m(w, v) = %f w · v + %s w · v
Ωf Ωs
Z Z
af (w, v) = 2µf D(w) : ∇v, as (w, v) = 2µs D(w) : ∇v
Ωf Ωs
Z Z Z
b(v, q) = − div vq − div vq = − div vq
Ωf Ωs Ωfs
m(v, v) ≥ Cm,1 kvk2L2 (Ωfs ) , m(w, v) ≤ Cm,2 kwk2L2 (Ωfs ) kvk2L2 (Ωfs )
I aj (w, v), with j = {f , s} are continuous and coercive in Vj,0 , thanks
to the Korn inequality
aj (v, v) ≥ Caj ,1 kvk2H 1 (Ωj ) , aj (w, v) ≤ Caj ,2 kwk2H 1 (Ωj ) kvk2H 1 (Ωj )
I b(v, q) satisfies the usual inf-sup condition between the spaces Vfs
and Qfs .
Global weak formulation
∂u
m( , v) + af (u, v) + as (η, v) + b(v, p)
∂t Z Z
= f ·v+ d·v ∀v ∈ Vfs,0
Ωfs ΓN
Z
1 ∂p
q − b(q, u) = 0 ∀q ∈ Qfs
λs Ωs ∂t
∂η = u
∂t
Finite element approximation (continuous in time)
We define finite element subspaces
Vhfs ⊂ Vfs fs
Vh,0 = Vhfs ∩ Vfs,0 Qhfs ⊂ Qfs
Observe that one should not expect the pressure to be continuous at the
interface. Therefore, it is better to work with a discontinuous pressure
space (or at least a space of pressures discontinuous at the interface).
Algebraic formulation
I let {φj }N fs
j=1 be a basis of Vh
u
N
I p
let {ψk }k=1 be a basis of Qhfs
Expansion of the solution:
X X X
uh (t) = uj (t)φj , η h (t) = ηj (t)φj , ph (t) = pk (t)ψk
j j k
algebraic system
BT
Mfs 0 0 U Af As U Fu
0 d
I 0 Ξ + −I 0 0 Ξ = 0
0 0 %s1λs Ms dt P −B 0 0 P Fp
which now admits a unique solution (the matrix multiplying the time
derivative vectors is now non singular). From this, we infer the existence
and uniqueness of uh , η h and ph |Ωs .
I It remains to show that there exists a unique pressure ph |Ωf also in
the fluid domain. We go back to the original basis {φj } of Vhfs . The
matrix B can be still split according to the split in the basis of Qh in
B T = [BfT |BsT ] where now BfT is not zero anymore.
I We show that the bilinear form b(·, ·) satisfies an inf-sup condition
between the spaces Qhf and Vhfs (rememer that b satisfies already the
inf-sup cond, between Qh and Vhfs ). Indeed,
b(vh , qh ) b(vh , qh )
inf sup ≥ inf sup ≥β>0
qh ∈Qhf v ∈Vfs
h
kvh kH 1 (Ωfs ) kqh kL2 (Ωf ) qh ∈Qh
vh ∈V fs kvh k H 1 (Ω ) kqh kL2 (Ω )
fs fs
h h
I m( ∂uh , uh ) = 1 d % f kuh k2
2 + % s kuh k 2
2
∂t 2 dt L (Ωf ) L (Ωs )
R ∂ph
I b(uh , ph ) = 1 1 d 2
λs Ωs ∂t ph = 2λs dt kph kL2 (Ωs )
from which we deduce that
%f kuh (t)k2L2 (Ωf ) + %s kuh (t)k2L2 (Ωs ) + 2µs kD(η h (t))k2L2 (Ωs )
Z t
1
+ kph (t)k2L2 (Ωs ) + 4µf kD(uh (s))k2L2 (Ωf ) ds = C (u0h , η 0h )
λs 0
2
1 ∂ p ∂uh
R
λs Ωs ∂t 2 qh − b(qh , ∂t ) = 0
h
Estimate for the pressure ph |Ωf . It remains to provide a bound for the
pressure ph |Ωf . We provide actually a bound for ph |Ωfs . From the inf-sup
condition we have
1 b(vh , ph (t))
kph (t)kL2 (Ωfs ) ≤ sup
β vh ∈Vfs kvh kH 1 (Ωfs )
h
1 −m( ∂u
∂t
h
(t), vh ) − af (uh (t), vh ) − as (η h (t), vh )
= sup
β vh ∈Vfs kvh kH 1 (Ωfs )
h
` ´
≤ C k∂t uh (t)kL2 (Ωfs ) + kuh (t)kH 1 (Ωf ) + kη h (t)kH 1 (Ωs )
Well posedness analysis – passage to the limit
We discretize in time with the mid-point algorithm, i.e. all equations are
colocated at t n+1/2 and the time derivatives are discretized with a
centered finite difference.
Mip-point discretization
n+ 1 n+ 1 n+ 1 n+ 1
2
∆t m(uh 2 , vh ) + af (uh 2 , vh ) + ∆t2 as (uh 2 , vh ) + b(vh , ph 2 )
1 1 2
= Ωfs f n+ 2 · vh + ΓN dn+ 2 · vh + ∆t m(unh , vh ) − as (η nh , vh )
R R
n+ 12 n+ 1
2 R
qh − b(qh , uh 2 ) = λs2∆t Ωs phn qh
R
λs ∆t Ωs ph
n+ 21 n+ 12 n+ 12
un+1 = 2uh − unh , p n+1 = 2ph − phn , η n+1
h = ∆tuh + η nh
Algebraic system
2 ∆t
# " n+ 1 # n+ 1
BT
"
∆t Mfs + Af + 2 As U 2 Fu 2 + 2
∆t Mfs U −
n
As Ξn
2 1
=
−B λs %s ∆t Ms Pn+ 2 2
λs %s ∆t Ms P
n
Stability analysis
Consider the homogeneous case (f = d = g = 0) for simplicity. Take
un+1 +un p n+1 −p n
vh = h 2 h and qh = h 2 h . and sum the first two equations.
Then
un+1 +unh phn+1 −phn
I The term b( h , 2 ) cancels out.
2
un+1 −u n
u n+1
+u n
1
m(un+1 n+1
n n
h , uh ) − m(uh , uh )
I m( h h
, h 2 h ) = 2∆t
∆t
η n+1 n n+1
h +η h uh +uh
n
η n+1 +η n η n+1 −η n
I as ( 2 , 2 ) = as ( h 2 h , h ∆t h ) =
1 n+1 n+1
n n
2∆t as (η h , η h ) − as (η h , η h )
R phn+1 −phn phn+1 +phn h i
I 1
λs Ωs ∆t 2 = 2λs1∆t kphn+1 k2L2 (Ωs ) − kphn k2L2 (Ωs )
Therefore
1 1 n+1 2
m(un+1
h , un+1
h ) + as (η n+1
h , η n+1
h ) + kp k L2 (Ωs )
2∆t λs h
un+1
h + unh un+1 + unh
+ af ( , h )
2 2
1 1
= m(unh , unh ) + as (η nh , η nh ) + kphn k2L2 (Ωs )
2∆t λs
Stability analysis
1 n 2
m(unh , unh ) + as (η nh , η nh ) +
kp k 2
λs h L (Ωs )
n
X ui + ui−1 ui + ui−1
+ 2∆t af ( h h
, h h
)
2 2
i=1
= m(uh,0 , uh,0 ) + as (η h,0 , η h,0 ) + λs k div η h,0 k2L2 (Ωs )
I A better estimate for the pressure can be recovered from the inf-sup
condition.
I It can also be proved that the scheme is second order accurate in
time.
Some References
div u = 0
+ b.cs and i.cs
Assume small perturbations around the rest state (ū, p̄) = (0, 0). Then
u · ∇u is negligible
Linearized inviscid fluid model
% ∂u + ∇p = 0,
f
∂t in Ωf , t > 0
div u = 0
u · n = g on ΓD , t > 0, p = d on ΓN , t > 0
f
u|t=0 = u0 , in Ω
Equation for the pressure
Pressure equation
∆p = 0
in Ωf
∂p
∂n = −%f ġ , on ΓD
p = d, on ΓN
Alternative formulation: potential of fluid velocity
I Assume that the fluid velocity is initially irrotational, i.e. curl u0 = 0.
Then the velocity field u(t) will be irrotational at all time.
I For a simply connected domain Ωf , we can introduce (up to a
constant) a potential φ : Ωf × R+ → R such that u(x, t) = ∇φ(x, t)
I Then, the momentum equation can be written as
∂∇φ ∂φ 1
%f + ∇p = 0, =⇒ = − p + c(t)
∂t ∂t %f
that is, the fluid potential is a primitive of the pressure (up to an
arbitrary constant which can be set to zero)
I In what follows, we will use the pressure equation only, although the
potential equation is probably more common.
A toy fluid-structure problem
Piston interacting with an inviscid incompressible fluid
y
η
(−L,H) Γwall u.n=0 (0,H)
Γin Piston
Γp
p=g(t) Fluid .
u.n= η
(−L,0) Γwall u.n=0 (0,0) x
Simplifying assumptions
I Small displacements: the fluid domain is considered fixed.
I Fluid equations linearized around the rest state.
ρf ∂t u + ∇p = 0
div u = 0
Z
Fluid: u · n = 0, on Γwall Piston: m η̈ + c η̇ + kη = pdσ
Γp
p = g (t), on Γin
u · n = η̇, on Γ
p
∆p = 0 in Ω,
∂ p = 0 on Γ
n wall
Pressure equation:
p = g (t) on Γ in
∂n p = −ρf η̈ on Γp
Remarks:
I The effective mass of the piston includes the mass of the fluid
(ma = %f LH) that has to be displaced.
I The presence of the fluid alters (descreases) the natural frequencies
of oscillation of the elastically supported piston
r r
k k
freq. of oscillation: in air: ω = , in the fluid: ω̃ =
m m + ma
More complex structure models
u.n=0
p=0
Σ
p=0
Ωf
Ωs
nf ns
u.n=0
Z Z Z * 0Z
< MA λ, µ >= %f pλ µ = %f pλ ∂n pµ = − %
f pλ ∆pµ +
%f ∇pλ ·∇pµ
Σ Σ
Ωf Ωf
Z
:
0Z
= − %f ∆p pm u +
λ %f ∂n pλ pµ =< λ, MA µ >
Ωf Σ
I MA is positive:
Z
< MA λ, λ >= %f ∇pλ · ∇pλ = %f k∇pλ k2L2 (Ωf ) > 0, ∀pλ 6= const.
Ωf
I Let Vh,s and Vh,f be finite element spaces for the structure
displacement and fluid pressure, respectively (need not be
conforming at the common interface Σ)
I {φj } and {ψl }: basis fuctions for Vh,s and Vh,f , respectively.
Z Z Z
Structure equation: %s η̈ · φ + σ s (η) : ∇φ = − pφ · ns
Ωs Ωs Σ
R
(Ms )ij = R Ωs %s φj φi
=⇒ Ms Ü + Ks U = −G P with (Ks )ij = Ωs σ s (φj ) : ∇φi
R
Gil = Σ ψl φi · ns
Z Z Z
Fluid equation: ∇p · ∇ψ = − %f (η̈ · nf )ψ = %f (η̈ · ns )ψ
Ωf Σ Σ
Z
=⇒ Kf P = %f G T Ü with (Kf )ij = ∇ψj · ∇ψi
Ωf
Algebraic formulation of the fluid structure problem
The coupled fluid-structure problem at algebraic level reads:
Ms 0 Ü Ks G U 0
+ =
−%f G T 0 P̈ 0 Kf P 0
Ms Ü + Ks U = −%f GKf−1 G T Ü
| {z }
MA
Added mass matrix
Ks U = ω 2 (Ms + MA )U
Modal reduction
Assume that the structure displacement can be described as a linear
combination of few modes only
m
X
η(x, y ) = αj (t)η j (x)
j=1
η 1 = e1 η 2 = e2 η 3 = e3
η 4 = y e1 − xe2 η 5 = ze1 − xe3 η 6 = ze2 − y e3
m
X Z m
X Z Z
α̈j (t) %s η j · η i + αj (t) σ s (η j ) : ∇η i = − pη i · ns
Ωs Ω
j=1 | {z } j=1 | s {z } | Σ {z }
(M
fs )ij (K
fs )ij Fei
Pm Pm f
Then p = l=1 α̈l pl and the r.h.s. becomes: Fei = l=1 (M A )il α̈l
p = pa
=⇒ δp(x) = %f g ηz (x)
On the other hand, from the linearized momentum equation
∂δp ∂uz ∂ 2 ηz 1 ∂ 2 δp
(x) = −%f = −%f (x) = − (x)
∂z ∂t ∂t 2 g ∂t 2
Equivalently, this problem can be stated using the potential of the fluid
velocity δu = ∇φ. Linearized momentum equation (frequency domain):
∂δu
%f + ∇δp = 0 =⇒ ∇ iω%f φb + δp
c =0
∂t
Potential equation – frequecy domain
∆φ = 0, in Ωf
b
ω2 b c = −iω%f φb + const
δp
∂z φ = g φ, on Γa
b
+ other b.cs
Fluid-structure with gravity and free surface
Γa p=pa
p=−ρ gz
Σ Ωf state (ū, p̄) = (0, −%f gz)
f
Ωs
f
u.n=0
nf ns I Frequency analysis
Fluid-structure problem
%s ω 2 η + div σ s (η) = 0, in Ωs
σ s (η) · ns = −(p̄ + p)ns ,
on Σ
∆p = 0, in Ωf
∂n p = −%f ω 2 η · ns , on Σ
ω2
∂z p = g p, on Γa
+ other b.cs
Global weak formulation
Z Z Z
2
−ω %s η · φ + σ s (η) : ∇φ + pφ · ns
Ωs Ωs Σ
ω2
Z Z Z Z
∇p · ∇ψ + ω 2 %f ψη · ns − pψ = − p̄φ · ns
Ωf Σ g Γa Σ
Algebraic formulation
Ks G U 2 Ms 0 U G P̄
=ω −
0 Kf P −%f G T S P 0
with Sij = Γa g1 ψj ψi
R
Added mass matrix for free surface flows
We eliminate P from the second equation
P = −ω 2 %f (Kf − ω 2 S)−1 G T U
Remarks
I The added mass depends on the angular frequency ω! This
technique is not very suited for time dependent problems
I The matrix (Kf − ω 2 S) might be singular! This will correspond to
resonance modes of the superficial waves
A better formulation consists in introducing the variable µ = ∂p
I
∂z Γa
(proportional to the vertical displacement of the fluid at the free
surface) and solving the coupled problem in (η, µ).
Some References
[MO95] H. J-P. Morand, R. Ohayon, Fluid structure interaction. Wiley & Sons,
1995.
[N77] J.N. Newman. Marine Hydrodynamics. The MIT Press, 1977.
[M83] C.C. Mei. The Applied Dynamics of Ocean Surface Waves. John Wiley
& Sons Inc, 1983.
[M10] A. Mola. Multi-physics and Multilevel Fidelity Modeling and Analysis of
Olympic Rowing Boat Dynamics. PhD Thesis, o Virginia Polytechnic Institute
and State University, 2010.
[FMMM09] L. Formaggia, E. Miglio, A. Mola, A. Montano. A model for the
dynamics of rowing boats. Int. J. Numer. Methods Fluids 61(2):119–143,
2009.
The fluid equations are defined on a moving domain. How can we treat
them numerically?
Idea Use a moving mesh, that follows the deformation of the domain.
T hs Γt,h
w
f
T h,0 T f
t,h
At
The mesh follows the moving boundary and is fixed on the artificial
(lateral) sections.
Eulerian / Lagrangian / ALE frame of reference
ΓLw
t Γw
t
ΓL
in
Ω ft ΩL t
out
ΓL Γ
in Ω ft ΩA t Γ
t
out
t t t
ΓL t
w
Γw t
(Lagrangian) (ALE)
Lt At
Γ0
w
Γ0 Ω0 Γ0
in f out
u u
Γ0w
At x=x (t, ξ )
Ω0
At : Ωf0 −→ Ωft ,
ξ
Ωt
x(ξ, t) = At (ξ)
∂At
I domain velocity w(x, t) = ◦ A−1
t (x)
∂t
∂u
∂u
I ALE derivative = + w · ∇x u
∂t ξ ∂t x
∂JAt
I Euler expansion = JAt div w, JAt = det(∇At )
∂t ξ
I ALE Transport formula
Z Z " #
d ∂u
u(x, t) = + u div w , ∀Vt ⊂ Ωft
dt Vt Vt ∂t ξ
Navier Stokes equations in ALE form
%f ∂u + %f ((u − w) · ∇)u − div σf (u, p) = f
∂t ξ in Ωt
div u = 0
Z ! Z
∂u
%f + (u − w) · ∇u · v + σ f : ∇v + div uq = f ·v
Ωft ∂t ξ Ωft
f f
T h,0 T t,h
At
Z ! Z
∂uh
%f + (uh − w) · ∇uh · vh + σ f : ∇vh + div uh qh = f · vh
Ωft ∂t ξ Ωft
Z Z Z
d
%f uh · vh + %f div((uh −wh )⊗uh )·vh +σ f : ∇vh +div uh qh = f ·vh
dt Ωft Ωft Ωft
d
(M(t)U(t)) + A(t)U(t) + Nc (t, U − W)U(t) +B T (t)P(t) = F(t)
dt
| {z } | {z } | {z }
transport
stiffness
ALE time der.
R
with Nc (U)ij = Ω
div (uh − wh ) ⊗ φj · φi
I The two formulations (conservative and non conservative) are
equivalent at the time-continuous level. But they lead to different
time discretization schemes.
I Observe that all matrices depend on time!
ALE time discretization
Non Conservative
Un − Un−1
Mn + An Un + N n (Un−1 − Wn )Un + (B n )T Pn = Fn
∆t
Conservative
M n Un − M n−1 Un−1
+ An Un + Ncn (Un−1 − Wn )Un + (B n )T Pn = Fn
∆t
f
T h,0 T f
t,h
At
PNf
In practice, we solve it by the finite element method: xn = n
i=1 xi φi ,
and Xn = [x1n , . . . , xNn f ]T satisfying
Km Xn = Gn
tn − t t − t n−1
At = At n−1 + At n , t ∈ [t n−1 , t n ]
∆t ∆t
I Compute the domain velocity w = ∂A ∂t .
t
A scheme satisfies the (discrete) GCL if the previous relation holds for
the numerical time integration rule used to integrate the ALE transport
n+1
term, here indicated by INTttn .
GCL satisfying time advancing scheme
d s r Possible formula
2 1 1 Mid-point / Trapezoidal
2 2 2 Two-point Gauss
3 1 2 Two-point Gauss
3 2 3 Two-point Gauss
Remark: Only the integral containing the ALE transport term has to be
treated in this special way.
A conservative GCL-satisfying scheme in 2D
Let us consider
I 2D Conservative ALE formulation
n n
1
X X
%f kunh k2n + 2µ∆t kD(uih )k2i− 1 ≤ %f kuh0 k20 + C ∆t kf i− 2 k2i− 1
2 2
i=1 i=1
Z Z
%f %f n− 12
kunh k2n − un−1
h · unh + %f div((un−1
h − wh ) ⊗ unh ) · unh
∆t ∆t Ωfn−1 Ωf 1
n−
2
| {z }
%f A
Z
1
+ 2µkD(un )k2n− 1 = f n− 2 · unh
2
Ωf
n− 1
2
Z
1 1 1 1 n− 21
A≥ kunh k2n − kunh k2n−1 − kun−1 k2n−1 − div(wh )(unh )2
∆t 2∆t 2∆t h 2 Ωf 1
n−
2
1 1
= kunh k2n − kun−1 k2n−1
2∆t 2∆t h
from which the result follows.
The last equality holds thanks to the GCL assumption:
Z tn Z Z
d 1 n− 1
kunh k2n − kunh k2n−1 = (unh )2 = div(wh 2 )(unh )2
t n−1 dt Ωft 2 Ωf
n− 1
2
An illustrative example [Formaggia-N. CMAME ’04]
[TL79] P.D. Thomas, C.K. Lombard, Geometric conservation law and its
application to flow computations on moving grids, AIAA 17:1030–1037, 1979.
Algorithms for fluid-structure interaction
with large displacements
FSI problem - a three field formulation
Coupling conditions
kinematic cond. un ◦ An = η̇ n on Γ0
n n f n s
dynamic cond. σ f (u , p )n = −σ s (η )n on Γt
n n n n
geometric cond. A = ξ + η , Ȧ = η̇ on Γ0
wn = Ȧn ◦ (An )−1 , Ωtn = An (Ωf0 )
I There are other options in between: for instance one can “formally”
write the monolithic problem and then use a block preconditioner to
solve it. In this way one recovers partitioned procedures (see e.g.
[Heil, CMAME ’04])
Partitioned fluid-structure algorithms
A simple loosely coupled partitioned FSI algorithm
Given the solution (un−1 , p n−1 , An−1 , Ȧn−1 , η n−1 , η̇ n−1 ) at time
step t n−1 :
1. Extrapolate the displacement and velocity of the interface: e.g.
˜ n = η̇ n−1 ,
η̇ ˜ n , on Γ0
η̃ n = η n−1 + ∆t η̇
(
Mesh(An , Ȧn ) = 0 in Ωf0
2. Compute the ALE mapping:
˜n
An = ξ + η̃ n , Ȧn = η̇ on Γ0
3. Solve the fluid problem with Dirichlet boundary conditions
(
FlALE (wn ; un , p n ) = 0, in Ωfn = An (Ωf0 )
˜ n ◦ (An )−1 ,
un = η̇ on Γn = An (Γ0 )
ηn−1/2 σ n+1/2
f
.
ηn−1 η n−1 ηn
.
ηn
.
ηn+1 η n+1
Given the solution (un−1/2 , p n−1/2 , An−1/2 , η n , η̇ n ):
1. Extrapolate the interface displacement and velocity at t n+1/2 :
∆t n
η n+1/2 = η n + η̇ , η̇ n+1/2 = η̇ n , on Γ0
2
(
Mesh(An+1/2 ) = 0
2. Compute the ALE mapping at t n+1/2 :
An+1/2 = ξ + η n+1/2
t − t n−1/2 t n+1/2 − t
and A(t) = An+1/2 +An−1/2 , t ∈ [t n−1/2 , t n+1/2 ]
∆t ∆t
3. Solve the fluid problem with BDF2 at t n+1/2
(
GCL
FlALE (w; un+1/2 , p n+1/2 ) = 0, in Ωfn+1/2 = An+1/2 (Ωf0 )
un+1/2 = η̇ n+1/2 ◦ (An+1/2 )−1 , on Γn = An (Γ0 )
η n+1 − η n η̇ n+1 + η̇ n
=
∆t 2
I The ALE mapping at the interface satisfies at t n
η n+1/2 + η n−1/2
n 1 n ∆t n n−1 ∆t n−1
A = = η + η̇ + η + η̇ = ηn
Γ0 2 2 2 2
I Analogously, the mesh velocity in [t n−1/2 , t n+1/2 ] satisfies
An+1/2 − An−1/2 η n+1/2 − η n−1/2
Ȧ(t) = =
Γ0 ∆t Γ0 ∆t
1 ∆t ∆t
= ηn + η̇ n − η n−1 − η̇ n−1 = η̇ n
∆t 2 2
I The ALE mapping is perfectly consistent with the structure
displacement and velocity.
I Moreover, interpreting the fluid velocity as piecewise constant in
time, we have u(t)|Γ0 = un+1/2 = η̇ n = Ȧ(t)|Γ0 in [t n−1/2 , t n+1/2 ]
The fluid velocity matches at the same time the structure and
the mesh velocities
On partitioned procedures
Partitioned procedures may have serious stability problems in certain
applications.
Dangerous situations appear for incompressible fluids when the mass of
the structure is small compared to the mass of the fluid
The instability is due to a non perfect balance of energy transfer at the
interface. For instance, in the algorithm by Farhat & Lesoinne, we have:
Work done in [t n , t n+1 ] by
Z
n+ 1 1 1 1
fluid Wf 2 = (σ f (un+ 2 , p n+ 2 )nf ) · ∆tun+ 2
Γn+ 1
2
η n+1 + η n s η̇ n+1 + η̇ n
Z
n+ 1
structure Ws 2 = (σ s ( )n ) · ∆t
Γn+ 1 2 2
2
n+ 12 n+ 12
Ideally, one would like to have Wf + Ws = 0.
n+ 1 n+ 1
Energy unbalancing: ∆W = Wf 2 + Ws 2
η̇ n+1 − η̇ n
Z
1 1
∆W = −∆t 2 (σ f (un+ 2 , p n+ 2 )nf ) · = O(∆t 2 )
Γn+1/2 2∆t
A fully implicit partitioned FSI algorithm
n n
ζ̃ k = Φ(ζ nk−1 ), rk = ζ̃ k − ζ nk−1
(rk−1 , rk − rk−1 )
ωk = −ωk−1
krk − rk−1 k2
n
ζ nk = ωk ζ̃ k + (1 − ωk )ζ nk−1
In the update of ωk one has to use a suitable inner product (·, ·) and
norm k · k2 on the space of the variable ζ nk . At the discrete level, ζ nk is a
vector and one can use the euclidean inner product and norm.
Load transfer between fluid and structure
At the discrete level, there is a proper way of exchanging information.
Conforming finite element spaces at the interface: let {φbi }N i=1 be the
b
i
111
000
000
111
000
111
000
111
000
111
Γ
000
111
NS
Non-conforming finite element spaces at the interface
Nb
I {φb } f : basis functions (fluid) corresponding to the interface
i i=1
PNb
nodes. uh |Γ = i=1f ui φbi |Γ
N
I {ψ b } bs basis functions (structure) corresponding to the interface
j j=1
PNbs
nodes. η h |Γ = j=1 ηj ψ bj |Γ
Simplifying assumptions
I Inviscid, incompressible fluid.
I Small displacements: the fluid domain is considered fixed.
I Fluid equations linearized around the rest state.
ρf ∂t u + ∇p = 0
Z
Fluid: div u = 0 Piston: m η̈ + c η̇ + kη = pdσ
Γp
u · n = η̇, on Γp
Modified structure equation R R
we write the fluid pressure on the piston as Γp p = Γp pg − MA η̈, where
pg accounts for the pressure on the left boundary and MA is the usual
added mass operator
−∆pλ = 0 in Ω,
−∆p = 0 in Ω,
g Z
∂ p = 0 on Γ ,
n λ wall
∂n pg = 0 on Γwall ∪ Γp MA λ = pλ ,
Γp pλ = 0 on Γin ,
pg = g on Γin .
∂n pλ = %f λ on Γp
Then the equation for the dynamics of the piston modifies as:
Z
(m + MA )η̈ + c η̇ + kη = pg dσ
Γp
Instability of explicit algorithms
Prototype of explicit algorithm: Leap-frog for the piston and Implicit
Euler for the fluid (LF-IE).
Given η n , η n−1 , un−1 , compute
un − un−1
ρ + ∇p n = 0
f
∆p n = 0
∆t
div un = 0
η n − 2η n−1 + η n−2
div
→ n
∂n p = −ρf
n η n − η n−1 ∆t 2
u · n = , on Γp
n
p = g (t n )
∆t
p n = g (t n ), on Γ
in
un+1 − un
ρ f + ∇p n+1 = 0 (
∆t div ∆p n+1 = 0
n+1 →
div u =0 η n+1 −2η n +η n−1
n+1
− ηn ∂n p n+1 = −ρf ∆t 2
un+1 = η
n
∆t
i. ∆pkn+1 = 0 in Ω
n+1
ηk−1 − 2η n + η n−1
∂n pkn+1 = −ρf on Γp
∆t 2
pkn+1 = g (t n+1 ) on Γin
η̃kn+1 n
− 2η + η n−1
η̃kn+1
−η n Z
ii. m +c + k η̃kn+1 = pkn+1 dσ
∆t 2 ∆t Γp
n
St(η nk , η̇ k ) = 0
(R) n n s n n f
αs η̇ k + σ s,k · n = αs uk − σ f ,k · n
with αs 6= αf
I The coefficient αs (resp. αf ) should incorporate (in a single
constant!) as much of the fluid (resp. structure) dynamics as
possible.
I This is not always feasible, but sometimes it works well.
Choice of αs
We have seen that for an inviscid, incompressible fluid, linearized around
the rest state, the effect of the fluid on the structure can be expressed by
the added mass operator.
p Γ ≈ −MA η̈ · nf
wall
Heuristic argument
η̇ n − η̇ n−1 s
(σ ns · ns ) · ns = −(σ nf ·nf )·ns ≈ −p n ≈ MA η̈ n ·nf ≈ −MA ·n
∆t
which resembles a Generalized Robing boundary condition
∆t −1 MA η̇ n + σ ns · ns = · · ·
| {z }
αs
At least the mass term can be included in the Robin boundary conditions:
∆t −1 hs ρs un + σ f (un , p n ) = · · ·
| {z }
αf
In some cases, other terms could be included: e.g. for a linear membrane
type solid one could have a term proportional to the local curvature of
Γwall (see [N.-Vergara, SISC 2008])
(
hs ρs ∆tEhs “ 2 ” E , ν : elastic constants
αf = + 4ρ1 − 2(1 − ν)ρ2 ,
∆t (1 − ν )
2
ρ1 , ρ2 : mean and Gaussian curvatures
Some comments on Robin-Robin partitioned algorithms
Z !
∂u
%f + (u − w) · ∇u · v + σ f (u, p) : ∇v + div uq+
Ωft ∂t ξ
∂2η
Z Z Z
0 f
%s 2 ·φ+σ s (η) : ∇ξ φ = f ·v+ f0s ·φ, ∀(v, q, φ) ∈ VFS
Ωs0 ∂t Ωft Ωs0
complemented with a problem for the ALE mapping, e.g. a Laplace eq.:
find At s.t. At = η(t) on Γ0 and
Z
∇ξ At · ∇ξ ψ = 0, ∀ψ ∈ VA
Ωf0
un − ũn−1
Z
%f + (un − Ȧn ) · ∇un · v + σ f (un , p n ) : ∇v + div un q
An (Ωf0 ) ∆t
η̇ n − η̇ n−1 η n + η n−1
Z
+ %s · φ + σ 0s ( ) : ∇ξ φ
Ωs0 ∆t 2
Z
+ ∇ξ An · ∇ξ ψ
Ωf0
Z Z
f ,n s,n−1/2
= f ·v+ f0 · φ, ∀(v, q, φ) ∈ VFS , ψ ∈ VA
Ωft Ωs0
2 2
η̇ n = (η n − η n−1 ) − η̇ n−1 , Ȧn = (An − An−1 ) − Ȧn−1
∆t ∆t
un ◦ An = η̇ n , An = η n , on Γ0
with
un − ũn−1
Z „ «
Lf (An , un ; p n , v, q) = %f + (un − Ȧn ) · ∇un · v
An (Ωf0 ) ∆t
Z
+ σ f (un , p n ) : ∇v + div un q
An (Ωf0 )
n n−1
η n + η n−1
Z
η̇ − η̇
Ls (η n ; φ) = %s · φ + σ 0s ( ) : ∇ξ φ
Ωs0 ∆t 2
Z
LA (An , ψ) = ∇ξ A n · ∇ ξ ψ
Ωf0