29.01.2015 Views

A New Crank-Nicolson-Like ALE Finite Volume Scheme Verifying ...

A New Crank-Nicolson-Like ALE Finite Volume Scheme Verifying ...

A New Crank-Nicolson-Like ALE Finite Volume Scheme Verifying ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Conference on Modelling Fluid Flow (CMFF’03)<br />

The 12 th International Conference on Fluid Flow Technologies<br />

Budapest, Hungary, September 3 - 6, 2003<br />

A NEW CRANCK-NICOLSON-LIKE <strong>ALE</strong> FINITE VOLUME SCHEME VERIFYING THE<br />

DISCRETE GEOMETRIC CONSERVATION LAW FOR 3D FLOW COMPUTATIONS ON<br />

MOVING UNSTRUCTURED MESHES<br />

Ingrid LEPOT 1 , Didier VIGNERON 2 and Jean-André ESSERS 3<br />

University of Liège, ASMA Department<br />

Institut de Mécanique et Génie Civil (Bât B52/3)<br />

1, Chemin des Chevreuils – B-4000 Liège, Belgium<br />

1<br />

FNRS Research Fellow<br />

Aerodynamic Group<br />

2<br />

FRIA Research Fellow<br />

Aerodynamic Group and<br />

Turbomachinery Group<br />

3<br />

Professor<br />

Aerodynamic Group<br />

ABSTRACT<br />

This paper describes a finite volume solver for<br />

the computation of unsteady Euler flows on moving<br />

unstructured grids. The Arbitrary Lagrangian<br />

Eulerian formulation takes the grid velocity into<br />

account. A Cranck-<strong>Nicolson</strong> time integration<br />

scheme and a quadratric spatial reconstruction<br />

technique are applied and yield a second order time<br />

and space truncation errors. A new way of satisfying<br />

the Discrete Geometric Conservation Law adapted<br />

to this scheme is proposed. This new method is<br />

computational efficient as it leaves the number of<br />

fluxes evaluations equal to their fixed grid<br />

counterpart. Results highlighting the robustness and<br />

accuracy of the proposed approach are presented.<br />

Key Words: discrete geometric conservation<br />

law, high order finite volume scheme, moving<br />

unstructured grids.<br />

NOMENCLATURE<br />

h [ m ] characteristic grid size<br />

J<br />

2<br />

[ m ] jacobian<br />

M ∞<br />

[-] Mach number<br />

n [-] unitary normal vector<br />

t [s] time<br />

u [ ] conservative variables vector<br />

v G<br />

[m/s] grid velocity vector<br />

x [m] spatial coordinates<br />

F [ ] advective fluxes vector<br />

R [ ] spatial discretized operator<br />

T [ ] grid vibration period<br />

α [-] time parameter<br />

ε [-] relative error<br />

ξ<br />

[-] surface integration parameter<br />

η<br />

[-] surface integration parameter<br />

μ<br />

2<br />

[ m ] non unitary normal<br />

ω [1/s] grid vibration pulsation<br />

∆ t [s] numerical time step<br />

θ [-] time integration scheme parameter<br />

2<br />

Σ [ m ]<br />

3<br />

Ω [ m ]<br />

cell surface<br />

cell volume<br />

1. INTRODUCTION<br />

In many unsteady fluid mechanics applications,<br />

the fluid movement is induced by the boundary<br />

movement and the flow problem is necessarily<br />

solved on a deforming grid. In those cases, the<br />

Euler equations must be written in Arbitrary<br />

Lagrangian Eulerian (<strong>ALE</strong>) formulation taking the<br />

grid velocity into account. The semi-discrete finite<br />

volume form of these equations is as follows:<br />

d<br />

d<br />

dt<br />

∫ u Ω<br />

Ω ( t)<br />

∫<br />

( ) ( )<br />

+ ⎡ ⋅ − ⋅ ⎤ dΣ<br />

=<br />

Σ ( t)<br />

⎣F u n u vG<br />

n ⎦ 0<br />

(1)<br />

where u collects the conservative variables at the


different mesh nodes, Ω(t) are the cell volumes,<br />

F(u).n the normal advective fluxes and v G<br />

.n the<br />

normal grid velocity. The time integration of these<br />

semi-discrete equations raises the issue of how to<br />

evaluate the normal fluxes integral on moving<br />

surfaces.<br />

An answer to this question issued in the early<br />

days of Computational Fluid Dynamics (CFD) [1]<br />

by considering the Geometric Conservation Law<br />

(GCL) stated below<br />

d<br />

d<br />

d<br />

dt<br />

∫ Ω −<br />

Ω ( t) ∫ v ⋅ Σ =<br />

Σ ( t)<br />

G<br />

n 0<br />

(2)<br />

An exact time integration turns this relation to<br />

the continuous formulation of GCL<br />

n 1<br />

t<br />

n 1 n<br />

Ω + +<br />

− Ω = ∫ dΣ<br />

n<br />

t ∫ v ⋅<br />

Σ ( t)<br />

G<br />

n<br />

(3)<br />

This formula expresses that the change in<br />

volume between t n and t n+1 is equal to the volume<br />

swept by its boundary during the time step Δt = t n+1 -<br />

t n . During their deformation, the control volumes<br />

always respect this relation and, hence, the Euler<br />

equations in <strong>ALE</strong> formulation has the natural<br />

property of accepting a solution u≠0 uniform in<br />

space and constant in time.<br />

After space and time discretization, this<br />

particular property is not necessarily preserved. A<br />

solution for obtaining a new numerical scheme for<br />

moving grid that respects this property from a wellknown<br />

classical scheme designed for fixed grid<br />

computations is to evaluate the geometric<br />

parameters, such as cell volume and grid velocity, in<br />

order to preserve a constant uniform solution at<br />

each time step. Therefore an adequate discrete<br />

version of the GCL (DGCL), depending not only on<br />

the semi-discretization but also on the time<br />

integration procedure, has to be derived for each<br />

numerical scheme. This DGCL constitutes a<br />

guideline to determine the geometrical quantities in<br />

the normal fluxes integrals.<br />

Violating the DGCL introduces spurious<br />

oscillations affecting the accuracy of the numerical<br />

scheme but the real necessity of the DGCL has long<br />

remained hazy. Guillard and Farhat [2] have shown<br />

that the DGCL enforcement is close to a time<br />

integration consistency condition. More recently,<br />

the relation of the DGCL enforcement to linear [3]<br />

and nonlinear [4] stability has also been<br />

investigated, assessing its practical importance.<br />

Several ways of imposing the DGCL condition<br />

for each numerical time step are possible. The GCL<br />

can be regarded as an additional conservation law<br />

[5] that has to be solved numerically using the same<br />

dt<br />

scheme that is used to integrate the fluid equations<br />

to provide a self-consistent solution for the local<br />

cell volumes. However, a more computationally<br />

efficient way is to avoid solving an additional<br />

equation and to adapt the numerical scheme with<br />

specific geometrical quantities to ensure the DGCL<br />

is not violated by integrating the physical quantities<br />

conservation equations.<br />

In this paper, the vertices movement considered<br />

between two mesh configurations is assumed to be<br />

linear in time, which yields a constant mesh<br />

velocity per time interval. The spatial discretization<br />

of the advective terms is first outlined. The implicit<br />

geometrically conservative time integration<br />

procedure is then described in detail. Finally,<br />

preliminary results assessing the accuracy and<br />

robustness of the proposed approach are presented.<br />

2. SPATIAL DISCRETIZATION<br />

This cell-centered finite volume solver is<br />

designed for unstructured meshes whose volume<br />

faces may be either triangles or (possibly non<br />

planar) quadrilaterals, allowing among others for<br />

tetrahedral, hexahedra, wedges or pyramids as<br />

control volumes. The design of high order schemes<br />

requires two elements : an accurate interpolation (or<br />

reconstruction) of the advective fluxes from the cell<br />

centers to the cell faces, and an accurate surface<br />

integration of the normal fluxes. Once the<br />

reconstruction has been performed, Roe’s flux<br />

difference splitting is used to compute the upwinded<br />

advective normal fluxes.<br />

The constant reconstruction scheme (often<br />

called first-order) only uses the left and right<br />

neighbours values on each face and can be proved<br />

to be inconsistent on irregular meshes. The<br />

consistent scheme needs a linear interpolation of the<br />

fluxes or variables to the face’s Gauss points. A<br />

third order method, which leads to a second order<br />

truncation error for the advective derivatives, is<br />

obtained with a quadratic reconstruction. It requires<br />

the calculation of centered second-order accurate<br />

first derivatives and first-order accurate second<br />

derivatives.<br />

The key component of high order schemes is<br />

the computation of the variables derivatives, used to<br />

perform the reconstruction, at the cell centers. A<br />

fixed stencil is selected for each node, which is the<br />

set of neighbouring nodes whose variables values<br />

are linearly combined to provide the<br />

aforementioned centered derivatives. This<br />

combination is defined in such a way that the<br />

difference between the reconstructed values at the<br />

location of the stencil nodes and their actual values


is minimized in a least square sense [6,7]. For the<br />

deforming meshes considered in this work, the<br />

derivative stencils remain constant throughout the<br />

computation, while the weights are recomputed for<br />

each new mesh configuration, i.e. at each time step.<br />

If each surface point x is represented by the two<br />

parameters ξ and η, the unitary normal n and the<br />

jacobian matrix determinant J can be expressed as<br />

x′ ξ<br />

∧ x′<br />

η<br />

n =<br />

x′ ′<br />

ξ<br />

∧ xη<br />

(4)<br />

J = x′ ′<br />

ξ<br />

∧ xη<br />

(5)<br />

and the advective fluxes surface integration<br />

becomes<br />

∫ ⎡ ( ) ( ) d d<br />

Σ ( t)<br />

⎣F u − u v ⎤⎦<br />

⋅ ′ ′<br />

G<br />

xξ<br />

∧ xη<br />

ξ η<br />

(6)<br />

For convenience, we define below the non unitary<br />

normal as<br />

μ = x′ ′<br />

ξ<br />

∧ xη = n dΣ<br />

(7)<br />

Those integrals are performed by a Gauss<br />

quadrature. Table 1 summarizes the minimum<br />

number of Gauss points per face necessary not to<br />

spoil the accuracy of the reconstruction procedure<br />

on static grids.<br />

Table 1. Number of face Gauss points on static grid<br />

Triangles Quadrangles<br />

Constant reconstruction 1 1<br />

Linear reconstruction 1 4<br />

Quadratic reconstruction 3 4<br />

On dynamic grids, the additional <strong>ALE</strong> term,<br />

composed of the reconstructed variables multiplied<br />

by the normal grid velocities which are at most biquadratic<br />

functions of (ξ,η), causes the necessary<br />

number of Gauss points to be increased in order to<br />

preserve the accuracy of the spatial discretization<br />

and to respect the DGCL condition.<br />

Table 2. Number of face Gauss points on dynamic<br />

grid<br />

Triangles Quadrangles<br />

Constant reconstruction 1 4<br />

Linear reconstruction 3 4<br />

Quadratic reconstruction 4 9<br />

3. TIME INTEGRATION<br />

Defining R as the Euler equations spatial terms<br />

( , ,<br />

G<br />

⋅ n)<br />

=<br />

⎡F ( u) n u ( v n)<br />

R u n v<br />

∫<br />

− ⋅ − ⋅ ⎤ dΣ<br />

Σ ( t)<br />

⎣<br />

G ⎦<br />

(8)<br />

the semi-discrete form of Eq. (1) can be written as<br />

d<br />

( u Ω ) = R ( u, n, vG<br />

⋅n,<br />

dΣ<br />

)<br />

dt<br />

(9)<br />

The time integration of this equation is<br />

performed by a one-step semi-implicit scheme<br />

leading to the following nonlinear equations system<br />

for the unknown variables u n+1<br />

n+ 1 n+<br />

1 n n<br />

u Ω − u Ω<br />

n+<br />

1<br />

= θ R + ( 1−θ<br />

)<br />

n<br />

∆t<br />

(10)<br />

Those nonlinear equations are solved by means<br />

of a <strong>New</strong>ton iterative process at each time step and<br />

the resulting linear system is solved by a matrixfree<br />

GMRES algorithm. The preconditioning of the<br />

linear system is critical to achieve a good<br />

preformance of the Krylov subspace solvers. Hence,<br />

a right-preconditioning of the GMRES, based on an<br />

cheap approximation of the jacobian matrix<br />

obtained by assuming a constant reconstruction<br />

scheme and an Incomplete LU decomposition<br />

solver (ILU), has been implemented.<br />

Such a scheme yields second order accuracy in<br />

time for θ = 0.5 ; this one-step trapezoidal rule (or<br />

Cranck-<strong>Nicolson</strong> scheme) appears as a natural<br />

optimum method since it is the most accurate A-<br />

stable linear multi-step method. Hence, thanks to<br />

the implicit treatment of the equations, the time step<br />

size is not restricted by a stability condition and can<br />

be as large as the smallest characteristic physical<br />

time of the fluid flow. It has been demonstrated in<br />

[4] that an implicit scheme violating the DGCL can<br />

exhibit the same nonlinear stability behaviour either<br />

on moving or fixed grid, if the time step and the<br />

mesh distortion are small enough. In our case, the<br />

DGCL is a crucial stability condition since the time<br />

steps could be quite large.<br />

Deriving the corresponding DGCL for this<br />

particular scheme is done by imposing a constant<br />

uniform solution u≠0 in Eq. (9). In this case, the<br />

integration of the normal fluxes F(u).n over surface<br />

cell volume is equal to zero and Eq. (9) yields<br />

n+<br />

1 n<br />

Ω − Ω<br />

n<br />

∆t<br />

= ∑ ∑<br />

⎡θ<br />

G<br />

⋅ + −<br />

G<br />

⋅<br />

⎣<br />

faces Gauss<br />

points<br />

R<br />

n+<br />

1<br />

( vμ ) ( 1 θv) ( μ )<br />

(11)<br />

Using the relation of Eq. (3), the latter equation<br />

turns to<br />

n<br />

n<br />

⎤<br />


n+<br />

1<br />

t<br />

∫ n<br />

t ∫ v<br />

Σ ( t)<br />

G<br />

∑ ∑<br />

faces Gauss<br />

points<br />

⋅n<br />

dΣ<br />

dt<br />

n+<br />

1<br />

( vμ ) ( 1 θv) ( μ )<br />

= ⎡θ<br />

G<br />

⋅ + −<br />

G<br />

⋅<br />

⎣<br />

(12)<br />

The right hand side (RHS) of Eq. (12) clearly<br />

appears as a discretized approximation of the left<br />

hand side (LHS). However that equation could be<br />

exact depending on the way the product v G<br />

.n<br />

depends on time. The latter requirement should be<br />

met for the DGCL to be satisfied. Since the vertices<br />

coordinates are considered to vary linearly in time<br />

during a time step, the grid velocity is constant and<br />

the non unitary normal is quadratic for 3D meshes<br />

and therefore the product v G<br />

.n is also quadratic. As<br />

the θ-scheme does not provide an exact integration<br />

of a quadratic function even for θ = 0.5 , the non<br />

modify Cranck-<strong>Nicolson</strong> scheme described in Eq.<br />

(10) does not meet the DGCL for 3D moving<br />

meshes.<br />

A straightforward possibility to enforce the<br />

DGCL is to evaluate both R n and R n+1 on each of the<br />

intermediate configurations and take their weighted<br />

average. Such a procedure however considerably<br />

raises the computational burden by doubling the<br />

number of normal fluxes computations.<br />

A more computationally efficient procedure [4]<br />

takes advantage of the fact that the normal fluxes<br />

become linear with respect of the normal for a<br />

uniform flow. Thanks to this property, one can<br />

choose the geometric parameters for each R n and<br />

R n+1 as to respect the DGCL. The distinction<br />

between the mean normal procedure and the distinct<br />

normals procedure is made below.<br />

Mean normal procedure<br />

An identical n for both normal fluxes, equal to<br />

the arithmetical mean between the normals at the<br />

intermediate configurations can be used as it has<br />

been proposed in [4]. Those intermediate<br />

configurations are taken at time t * and t **<br />

corresponding to the Gauss points positions on the<br />

interval time step Δt :<br />

* n 1 3<br />

t t ⎛ ⎞<br />

= + 1− ∆t<br />

2 ⎜<br />

3 ⎟<br />

⎝ ⎠ (13)<br />

** n 1 3<br />

t t ⎛ ⎞<br />

= + 1+ ∆t<br />

2 ⎜<br />

3 ⎟<br />

⎝ ⎠ (14)<br />

This mean normal procedure leads to the<br />

following time discretization scheme<br />

n<br />

⎤<br />

⎦<br />

u<br />

Ω − u<br />

n<br />

∆t<br />

Ω<br />

n+ 1 n+<br />

1 n n<br />

n+<br />

θ R ( u 1 , n, vG<br />

n,<br />

dΣ<br />

)<br />

n<br />

( 1 θ ) R ( u , n, vG<br />

n,<br />

dΣ<br />

)<br />

= ⋅<br />

+ − ⋅<br />

(15)<br />

where the mean geometric parameters are computed<br />

as<br />

1 * **<br />

n = 2 ⎡ ⎣n + n ⎤ ⎦<br />

(16)<br />

1<br />

* **<br />

v ⋅ = ( ) ( )<br />

2 ⎡ ⋅ + ⋅<br />

⎣<br />

⎤<br />

G<br />

n vG n vG<br />

n<br />

⎦<br />

(17)<br />

1 * **<br />

dΣ = ⎡( dΣ ) + ( dΣ<br />

) ⎤<br />

2 ⎣<br />

⎦<br />

(18)<br />

The DGCL is then satisfied independently of<br />

the θ parameter. Such a normal no longer<br />

corresponds to a physical intermediate<br />

configuration as the face normals defined in Eq. (4)<br />

are nonlinear functions of the vertices coordinates.<br />

Distinct normals procedure<br />

A new way of enforcing the DGCL is now<br />

proposed. Integrating the product of two functions<br />

ψ(t) and δ(t) can be done by the following time<br />

integration method<br />

n+<br />

1<br />

1 t<br />

ψ<br />

n<br />

( t ) δ ( t ) dt<br />

∆t<br />

∫t<br />

* n<br />

** n+<br />

1<br />

( 1 θ ) ψ ( t ) δ ( t ) θψ ( t ) δ ( t )<br />

= − +<br />

( t )<br />

+ O ∆<br />

with the intermediate times t * and t ** chosen as<br />

* n *<br />

t = t + α ∆t<br />

** n **<br />

t = t + α ∆t<br />

(19)<br />

(20)<br />

(21)<br />

This scheme is generally first order accurate in<br />

time. However, if θ=0.5 and α * +α ** =1, the second<br />

order of accuracy of the trapezoidal rule is<br />

recovered. In the following, we require formula in<br />

Eq. (19) to be exact if the product ψ(t) δ(t) is a<br />

quadratic function of time. This requirement is met<br />

by choosing the α * and α ** as follows<br />

⎡<br />

⎤<br />

*<br />

α =<br />

1<br />

2<br />

⎢1<br />

−<br />

⎣<br />

3<br />

θ<br />

( 1−θ<br />

) ⎥ ⎦<br />

(22)<br />

⎡ ⎤<br />

** 1 1−θ<br />

α = ⎢1<br />

+ ⎥<br />

2 ⎣ 3θ<br />

⎦<br />

(23)<br />

To ensure that α * and α ** , which now depend on<br />

the θ parameter, belong to the interval [0,1], θ must


e chosen between 0.25 and 0.75. For the Cranck-<br />

<strong>Nicolson</strong> scheme (θ=0.5), we naturally recover the<br />

Gauss points positions expressed in “Eqs. (13) and<br />

(14)”.<br />

Since the <strong>ALE</strong> term of Eq. (8) appears as a<br />

product of two functions, the ψ(t) function is<br />

assumed to contain the geometrical part such as the<br />

grid velocity (considered constant during each time<br />

step) multiplied by the non unitary normal μ while<br />

the δ(t) function contains the conservative variables<br />

part u. For 3D flow computations, the non unitary<br />

normal is a quadratic function of time while u is<br />

considered constant while checking the DGCL<br />

condition. Hence, the resulting product ψ(t) δ(t) is a<br />

quadratic function of time and the formula in Eq.<br />

(19) is therefore exact. As a result the scheme below<br />

satisfies the DGCL.<br />

n+ 1 n+<br />

1 n n<br />

u Ω − u Ω<br />

n<br />

∆t<br />

n+<br />

**<br />

( ( )<br />

)<br />

1 , ** , , dΣ<br />

**<br />

G<br />

= θ R u n v ⋅n<br />

*<br />

( )<br />

* dΣ<br />

*<br />

G<br />

n<br />

( θ ) R u n ( v n)<br />

+ 1 − , , ⋅ ,<br />

(24)<br />

This procedure, like the mean normal<br />

procedure, leaves the number of fluxes evaluations<br />

equal to their fixed grid counterpart. However, the<br />

distinct normals used for the normal fluxes<br />

evaluations now correspond to a physical<br />

intermediate configuration on the considered time<br />

interval. This allows to take the positions of the face<br />

Gauss points at these configurations and, hence, this<br />

procedure could also be applied when the spatial<br />

coordinates explicitly appear in the fluxes<br />

expressions (e.g. Euler equations in cylindrical<br />

coordinates), which can’t be handled by mean<br />

normal procedure.<br />

4. RESULTS<br />

In this section, numerical results are presented.<br />

Starting with a uniform mesh with grid size h, we<br />

require that grid to vibrate randomly for each test<br />

case. This leads to a completely distorded mesh a<br />

each time step. Each vertex vibrates around its<br />

initial position with an amplitude Δx 0<br />

, randomly<br />

chosen in amplitude and direction, and its position<br />

at time t n can be calculated as<br />

n 0 0 n<br />

x = x + ∆x<br />

sin ( ω t )<br />

(25)<br />

where the pulsation ω can be chosen.<br />

To make sure that the proposed schemes really<br />

meet the DGCL, a first test case has been computed<br />

solving the Euler equations for a uniform flow on a<br />

3D mesh.<br />

The time accuracy is then analysed on a simple<br />

linear scalar hyperbolic equation with a known<br />

analytical solution for different mesh movements.<br />

For these two test cases, comparison is made<br />

between the normal mean procedure, the distinct<br />

normals procedure and the non modified Cranck-<br />

<strong>Nicolson</strong> scheme which takes all the geometric<br />

parameters in t n and t n+1 .<br />

Finally, some convergence properties of the<br />

distinct normal approach are outlined on a fluid<br />

flow in a channel with a bump.<br />

4.1. Uniform flow<br />

This first test case consists in a uniform flow<br />

(M ∞<br />

=0.2) through a rectangular duct. The goal is to<br />

check if the DGCL is satisfied and to measure the<br />

time evolution of the error for non-DGCL schemes.<br />

The grid is required to vibrate with a fixed pulsation<br />

ω=500π and several distorsion amplitudes. The<br />

numerical time step Δt is constant and is such that<br />

the mesh movement period is discretized on 16<br />

points.<br />

Realtive error<br />

2,51E-04<br />

2,01E-04<br />

1,51E-04<br />

1,01E-04<br />

5,10E-05<br />

dx0 = 0.05 h<br />

dx0 = 0.1 h<br />

dx0 = 0.15 h<br />

1,00E-06<br />

2,50E-04 2,00E-03 3,75E-03 5,50E-03 7,25E-03<br />

Time t (s)<br />

Figure 1. Time evolution of the relative error for<br />

the non respective DGCL Cranck-<strong>Nicolson</strong> scheme<br />

for different mesh distosion amplitudes.


Relative error<br />

1,82E-12<br />

1,62E-12<br />

1,42E-12<br />

1,22E-12<br />

1,02E-12<br />

8,20E-13<br />

6,20E-13<br />

4,20E-13<br />

2,20E-13<br />

Distinct normals procedure<br />

Normal mean procedure<br />

2,00E-14<br />

2,50E-04 1,75E-03 3,25E-03 4,75E-03 6,25E-03<br />

Time t (s)<br />

Figure 2. Time evolution of the relative error for<br />

both the distinct normals and the normal mean<br />

approach for a mesh distorsion amplitude of dx0 =<br />

1.5 h.<br />

The relative error ε occurring at each time step<br />

is computed with the following formula<br />

ε =<br />

∑<br />

i<br />

( ui<br />

− u )<br />

∑ ( ui<br />

)<br />

i<br />

2<br />

i<br />

2<br />

(26)<br />

with u u<br />

i and<br />

i respectively denoting the numerical<br />

solution and the analytical solution.<br />

Figure 1 shows the evolution of error with time<br />

for the non modified Cranck-<strong>Nicolson</strong> scheme<br />

while Figure 2 shows it for both normal mean and<br />

distinct normals approaches.<br />

These examples clearly show that, as opposed<br />

to the non-DGCL Cranck-<strong>Nicolson</strong> scheme, either<br />

the normal mean and the distinct normals<br />

techniques lead to a DGCL scheme.<br />

4.2. Advection scalar test<br />

The accuracy analysis is here made by solving<br />

the following simple linear scalar hyperbolic<br />

conservation equation on a cubic domain.<br />

∂u<br />

∂u<br />

+ = 0<br />

∂t<br />

∂x<br />

(27)<br />

The initial condition consists in the continuous<br />

spherical field<br />

2⎛<br />

1 ⎞<br />

u = 1+<br />

cos ⎜ π r ⎟<br />

⎝ 2 ⎠ if r < 1<br />

(28)<br />

u = 1<br />

if r ≥ 1<br />

(29)<br />

with the radius r defined as<br />

r =<br />

2<br />

2<br />

( x − 0.3) + ( y − 0.5) + ( z − 0.5)<br />

with R=0.25.<br />

Error<br />

7.50E-03<br />

6.50E-03<br />

5.50E-03<br />

4.50E-03<br />

3.50E-03<br />

2.50E-03<br />

1.50E-03<br />

R<br />

Violating DGCL scheme<br />

Distinct normal scheme<br />

Normal mean scheme<br />

Fixed grid error limit<br />

2<br />

(30)<br />

5.00E-04<br />

2 4 8 16 32 64<br />

T/(2dt)<br />

Figure 3. Time integration error on a vibrating grid<br />

for the respecting DGCL schemes (distinct normal<br />

and normal mean procedures) and for the non<br />

modified Cranck-<strong>Nicolson</strong> scheme violating the<br />

DGCL.<br />

The analytical solution is simply a translation of<br />

that initial field in the x direction with a unitary<br />

speed. For accuracy analysis, the numerical solution<br />

u obtained at t=0.08 with a time step Δt=2.5 is<br />

compared with the analytical solution u on each<br />

node i for the different procedures. The relative<br />

error ε is computed by Eq. (26).<br />

Here, the distorsion amplitude Δx 0<br />

is maintained<br />

constant for each test while the vibration period T of<br />

the grid movement goes from small to large number<br />

of time steps. Figure 3 shows the different errors<br />

obtained for each procedure.<br />

This example highlights that the non modified<br />

Cranck-<strong>Nicolson</strong> scheme, which does not meet the<br />

DGCL, exhibits larger errors for the high frequency<br />

mesh movements and that its behaviour depends of<br />

how the mesh movement is time discretized. On the<br />

other hand the two DGCL schemes are less<br />

dependent of the grid velocity. Finally we can also<br />

observe that the obtained error for very slow mesh<br />

motions tends to be equal to the fixed grid error<br />

whether the scheme is DGCL or not. This remark<br />

means that satisfying the DGCL is really essential<br />

when the characteristic time of the grid movement<br />

becomes comparable or smaller that the numerical<br />

time step.


4.3. Steady plane flow in a channel with a<br />

bump using a 3D vibrating mesh<br />

A steady subsonic flow (M ∞<br />

=0.3) in a channel<br />

with a bump is computed on a tetrahedral mesh<br />

(19378 cells) and serves as initial field before the<br />

mesh starts to vibrate.<br />

Number of Krylov vectors<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

NM-dx0=h<br />

DN-dx0=h<br />

NM-dx0=0.5 h<br />

DN-dx0=0.5 h<br />

NM-dx0=0.25 h<br />

DN-dx0=0.25 h<br />

1 2 3 4 5 6<br />

Nonlinear iterations<br />

Figure 4. Convergence curves : number of Krylov<br />

vectors per nonlinear <strong>New</strong>ton iterations for several<br />

mesh distorsion amplitudes (Δx 0<br />

= 0.25h, Δx 0<br />

= 0.5h<br />

and Δx 0<br />

= h) using the normal mean procedure (MN)<br />

and the distinct normals procedure (DN).<br />

The only unsteadyness is here brought by the<br />

mesh movement. Figure 4 shows the necessary<br />

number of Krylov vectors within each nonlinear<br />

<strong>New</strong>ton iteration until the convergence is obtained<br />

from the numerical solution at time t n to the new<br />

solution at time t n+1 .<br />

For a random Δx 0<br />

= 0.25h distorsion, both mean<br />

normal and distinct normals approaches exhibit the<br />

same behaviour. However, for a Δx 0<br />

=0.5h<br />

distorsion, the distinct normals approach shows an<br />

improved convergence rate. This becomes more<br />

critical for a Δx 0<br />

=h distortion. For the latter, it<br />

should be noted that even some mesh crossovers<br />

appear (about 4.8% of the volumes) ; this however<br />

doesn’t alter the conservativity of the finite volume<br />

discretization and solutions could be obtained by<br />

both methods on the cases presented here with such<br />

extreme conditions.<br />

showed an improved convergence rate with respect<br />

to a normal mean approach, with the same accuracy.<br />

This new technique could also be applied when the<br />

spatial coordinates explicitly appear in the fluxes<br />

expressions (e.g. Euler equations in cylindrical<br />

coordinates), which can’t be handled by mean<br />

normal procedure.<br />

Further investigations need of course to be<br />

carried out for unsteady flows with moving<br />

boundaries.<br />

REFERENCES<br />

[1] Thomas P.D., and Lombard C.K.,<br />

“Geometrical Conservation Law and its<br />

Applications to Flow Computations on Moving<br />

Grids”, AIAA Journal, 17:1030, 1979.<br />

[2] Guillard H., and Farhat C., “On the<br />

Significance of The Geometric Conservation Law<br />

for Flow Computations on Moving Meshes”,<br />

Comput. Meth. Appl. Mech. Eng., 190:1467, 2000.<br />

[3] Fromaggia L., and Nobile F., “A Stability<br />

Analysis for the Arbitrary Lagrangian Eulerian<br />

Formulation with <strong>Finite</strong> Elements”, East-West J.<br />

Numeric Math., 7:107, 1999.<br />

[4] Farhat C., Geuzaine P., and Grandmont C.,<br />

“The Discret Geometric Conservation Law and the<br />

Nonlinear Stability of <strong>ALE</strong> <strong>Scheme</strong>s for the<br />

Solution of Flow Problems on Moving Grids”,<br />

Journal of Computational Physics, 174:669, 2001.<br />

[5] Batina J.T., “Unsteady Euler Airfoil<br />

Solutions Using Unstructured Dynamic Meshes”,<br />

Jan 1989. AIAA Paper 89-0115.<br />

[6] Delanaye M., and YenLiu, “Quadratic<br />

Reconstruction <strong>Finite</strong> <strong>Volume</strong> <strong>Scheme</strong>s on 3D<br />

Arbitrary Unstructured Polyhedral Grids”, June<br />

1999. AIAA Paper 99-3259.<br />

[7] Lepot I., Meers F., and Essers J.-A.,<br />

“Multilevel Parallel High Order <strong>Scheme</strong>s for<br />

Inviscid Flow Computations on 3D Unstructured<br />

Meshes”, In Proceedings of the ECCOMAS<br />

Computational Fluid Dynamics Conference, Sep<br />

2001.<br />

5. CONCLUSION<br />

A geometrically conservative high order scheme<br />

has been presented. The importance of verifying the<br />

Discrete Geometric Conservation Law for<br />

computations on highly moving grids has been<br />

outlined on simple test cases.<br />

The new proposed DGCL enforcement has

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!