Vous êtes sur la page 1sur 25

Evaluation of TVD high resolution schemes for unsteady

viscous shocked ows


Virginie Daru
a
, Christian Tenaud
b,
*
a
ENSAM, Lab. Sinumef, 151 Bd de l'Hopital, 75013 Paris, France
b
LIMSI-UPR CNRS 3251, BP 133, Campus d'Orsay, F-91403 Orsay Cedex, France
Received 14 August 1998; received in revised form 25 November 1999; accepted 1 February 2000
Abstract
The goal of this study is to evaluate the accuracy of several high resolution total variation diminishing
schemes in solving complex unsteady viscous shocked ows. Two types of discretization, namely a
combined time and space discretization, and an independent time and space discretization are
considered. Both methods are associated with several limiters, among which a more accurate new family
of limiters depending on the local wave velocity. The accuracy properties of each scheme are rst
reviewed on inviscid 1D and 2D test cases, in order to establish a ranking with respect to their
dissipative error. We then study the ow produced by the interaction of a reected shock wave with the
incident boundary layer in a shock tube. The calculations are performed for two values of the Reynolds
number. At Re = 200, convergence is attained and it is shown that the combined time and space
discretization method converges faster. Good classical limiters do almost the same job as the new family
of limiters. When the Reynolds number is increased to the value of 1000, the ow becomes much more
complex. Although convergence is hard to reach, the close examination of the results leads us to
conclude that the combined time and space discretization method associated with the new limiter gives
from far the best results. 7 2000 Elsevier Science Ltd. All rights reserved.
Keywords: Shock boundary-layer interaction; Shock tube; Reected shock; Computational uid dynamics
Computers & Fluids 30 (2001) 89113
0045-7930/01/$ - see front matter 7 2000 Elsevier Science Ltd. All rights reserved.
PII: S0045- 7930( 00) 00006- 2
www.elsevier.com/locate/compfluid
* Corresponding author. Tel.: +33-1-6985-8130; fax: +33-1-6985-8088.
E-mail address: tenaud@limsi.fr (C. Tenaud).
1. Introduction
Total variation diminishing (TVD) high resolution schemes have proven to be very eective
for computing inviscid ows. However, for high Reynolds number viscous ows, the numerical
diusion which is introduced by the limiter functions can have undesirable eects, generally
leading to misrepresentation of the viscous eects. There exists now a large number of limiters,
among them the Minmod limiter which is known to be very diusive, the Superbee limiter of
Roe, well adapted to the computation of discontinuities but which tends to ``square'' smooth
proles, or the Van Leer limiter which gives intermediate results and can be taken as a
reference. All of these classical limiters do not depend on the Courant number, and retain only
a simplied condition to satisfy the TVD constraints. We must underline that these limiters are
associated with schemes which can eventually depend on the Courant number.
More recently, in the inviscid case, several authors [13] have used a family of limiters with
some success which, if not really new (the rst work in this way was done by Roe and Baines
[4]), did not receive much attention for ow computations until now. The main interest of this
family of limiters lies in the fact that, in addition to prevent numerical oscillations, it can
increase the order of accuracy of the scheme from 2 to 3 in smooth regions, at least in the
linear scalar case. As these limiters depend on the local Courant number, they are well adapted
only for unsteady ows.
For the computation of unsteady ows, not only the limiter, but also the scheme which is
used is of importance. Our aim here is to evaluate the accuracy of a number of TVD high
resolution schemes to solve complex unsteady viscous ows. We will focus our study on the
dicult problem of the viscous interaction between the boundary layer generated behind a
shock wave travelling in a shock tube and the reected shock produced after its reection at
the end wall. This ow conguration can be encountered in high enthalpy ow facilities or in
industrial equipments for the distribution of pressurized gas for instance, where a good
knowledge of the phenomena involved is of particular interest in terms of safety purposes.
Within an incident shock Mach number range, the interaction results in a lambda-shape like
bifurcated shock wave pattern, rst described in 1958 by H. Mark in the laminar case. This
phenomenon can be explained by the fact that the stagnation pressure of the boundary-layer
becomes lower than the pressure behind the reected shock. Therefore, the uid cannot pass
under the reected wave and a separated ow region appears. This separated region forms a
``bubble'' which is dragged upstream with the reected shock (Fig. 1). The analysis of recent
experimental data [5] suggests that the ow is turbulent in the bubble carried along under the
reected shock.
To our knowledge, very few numerical investigations of this problem have been reported in
the literature (see [6] and references therein), until the recent work of Weber et al. [7], who
have carried out detailed studies on this problem, using the FCT algorithm. They have
reviewed the results of sensitivity to the Reynolds number, the Mach number and the wall
temperature. But, as it is mentioned in their paper, it is not clear whether or not convergence
with respect to the grid renement was attained. In this paper, we highlight the diculty to
obtain a converged solution when the Reynolds number value exceeds a few hundred. In the
meantime the above new family of limiters, associated with a Mac-Cormack type scheme, is
clearly demonstrated to be the more accurate for solving this kind of problem. Finally, we
V. Daru, C. Tenaud / Computers & Fluids 30 (2001) 89113 90
think that the reected shock-boundary layer interaction problem could constitute an
interesting benchmark test case to evaluate the accuracy of numerical schemes for the
computation of unsteady viscous ows.
The paper is organized as follows: after a brief recall of the governing equations (Section 1),
we present in Section 2 the so-called ``O3'' and ``O3Sup'' limiters, and the dierent numerical
schemes we have used (a predictorcorrector Mac-Cormack scheme, a Harten-Yee second-
order nite dierence upwind scheme, and a MUSCL scheme). In Sections 3 and 4, we report
the results obtained for inviscid 1D and 2D test cases. The rest of the paper is devoted to the
evaluation and the comparison of the dierent numerical approaches for solving the reected
shock-boundary layer interaction problem. The Reynolds number is naturally a key parameter
in this study, and we limit ourselves to relatively low values (below 1000) in order to limit the
grid size requirements.
2. Governing equations
Our study is limited to the two-dimensional case (it could represent the ow in the middle
section of a cubic shock tube). The ow is treated as laminar, so no turbulence model is used.
We solve the NavierStokes equations written in cartesian coordinates, expressed using non-
dimensional variables as:
w
t

_
f(w) f
v
(
w, w
x
, w
y)
_
x

_
g(w) g
v
(
w, w
x
, w
y)
_
y
= 0 (1)
where w is the vector of conservative variables (r, ru, rv, rE), r being the density, u and v the
uid velocity components, and E the total energy. f and g are the inviscid uxes, f
v
and g
v
are
the viscous part of the uxes. We have:
f(w) =
_
_
_
_
ru
ru
2
p
ruv
(rE p)u
_
_
_
_
g(w) =
_
_
_
_
rv
ruv
rv
2
p
(rE p)v
_
_
_
_
Fig. 1. Bifurcated shock pattern.
V. Daru, C. Tenaud / Computers & Fluids 30 (2001) 89113 91
f
v
(w) =
1
Re
_
_
_
_
_
_
0
t
xx
= l
(
u
x
v
y)
2mu
x
t
xy
= m
(
u
y
v
x)
ut
xx
vt
xy

gm
Pr
e
x
_
_
_
_
_
_
g
v
(w) =
1
Re
_
_
_
_
_
_
0
t
yx
= t
xy
t
yy
= l
(
u
x
v
y)
2mv
y
ut
yx
vt
yy

gm
Pr
e
y
_
_
_
_
_
_
As usual, p denotes the pressure, given by the perfect gas law p = (g 1)re, where e is the
specic internal energy, and Re is the Reynolds number. For simplicity we restrict our
attention to an ideal gas with constant specic heat ratio g = 1X4, constant viscosity coecients
l and m and Prandtl number P
r
= 0X73X
3. Numerical schemes
We now present the numerical schemes we have used in the computations.
3.1. Combined time and space discretization: the MCO3 scheme
In such an approach, we use the Mac-Cormack scheme, to which is added a limiting
correction term in order to render the scheme TVD. Beside classical limiters, we have used
recent specic limiters which particularity is that they hybridize a third order scheme and a rst
order scheme in the scalar linear case. When implemented in a correction term added to a
second order scheme, they can act either by increasing the order of the scheme to three in
smooth regions, either by diminishing it to one to satisfy TVD constraints. Let us describe this
limiter by considering the 1D linear model equation:
du
dt
a
du
dx
= 0
where a is a positive constant, and u = u(x, t)X If we denote by f
n
j
the discrete quantity f
estimated at a grid point x
j
= jd
x
and at a time t = n dt, dt and dx being the time and space
steps, a limited Lax-Wendro scheme can be written:
u
n1
j
= u
n
j

dt
dx
_
F
j1a2
F
j1a2
_
where n = adtadx is the Courant number, du
n
j1a2
= u
n
j1
u
n
j
, and F
j1a2
= au
n
j

1
2
a(1
n)du
n
j1a2
F
j1a2
with the usual denition of r
j
: r
j
=
du
n
j1a2
du
n
j1a2
X Now, if the limiting function F is
equal to 1
1n
3
(1 r
j
), we get the third order upwind scheme. This leads to the denition of a
new family of limiters F(r, n), giving the third order scheme in smooth regions, and restricted
elsewhere such as to satisfy the TVD constraints established by Harten. Such a function was
proposed by the authors at the ICFD conference at Oxford (1995, unpublished) and in [3]. It is
written as:
V. Daru, C. Tenaud / Computers & Fluids 30 (2001) 89113 92
F
O3
(r, n) =
_

_
0 r0
1
(
1 r
)(
1 (n 2)r) 0 ` r1a3
1
(
1 r
)
1 n
3
1a3 ` r3
1
(
1 r
)
1 n
r
r b 3
Arora et al. [2] and Jeng et al. [1] proposed to take the upper bound of the TVD constraints. It
is given by:
F
O3Sup
(r, n) = max
_
0, min
_
2r
n
, 1
1 n
3
(
1 r
)
,
2
1 n
__
The above limiter functions have proven to be very well behaved, as well in capturing
discontinuities as for computing smooth proles. In particular, the latter are not ``squared o''
as it is the case when using the SuperBee limiter. Arora et al. [2] have found that F
O3Sup
is too
compressive for non-linear systems. Our experience using F
O3
did not show such a behaviour.
We must underline that this limiter family has been built in a combined time and space
approach, using a ux limiter formulation. It cannot be easily transposed to a slope limiter
approach. This is the reason why the F
O3
and F
O3Sup
limiters will only be associated to a
combined time and space discretization. To solve the Euler equations, we make use of the Roe-
averaged matrices. In the multi-dimensional case, the correction is applied separately in each
space direction. To solve the system of conservation laws:
dw
dt

df
dx

dg
dy
= 0
the scheme reads, denoting by f
n
i, j
the discrete quantity f estimated at a grid point x
i
= i dx,
y
j
= j dy and at time t = ndt:
w
n1
i, j
= w
MC
i, j
C
x
i1a2, j
C
x
i1a2, j
C
y
i, j1a2
C
y
i, j1a2
where w
MC
i, j
stands for the solution given by the Mac-Cormack scheme. The correction term
C
x
i1a2, j
is given by:
C
x
i1a2, j
=

l
_
[n
l
[
2
_
1 [n
l
[
__
1 F
_
r
sl
, [n
l
[
__
da
l
d
l
_
i1a2, j
where, as usual, a
l
and d
l
are the eigenvalues and eigenvectors of the Roe-averaged jacobian
matrix A =
df
dw
, da
l
is the contribution of the l-wave to the variation dw, s = sign(a
l
) and n
l
=
dt
dx
a
l
X The ratios r
l
and r
l
are given by:
r
l
=
da
l
i1a2, j
da
l
i1a2, j
, r
l
=
da
l
i3a2, j
da
l
i1a2, j
The term C
y
is analogous.
V. Daru, C. Tenaud / Computers & Fluids 30 (2001) 89113 93
In the system case, the underlying high order scheme is no more third order but it should be
emphasized that the correction term, beside being a TVD correction of a second order scheme,
should also reduce the truncation error in smooth regions.
The extension to the NavierStokes equation is made by using central second order accurate
dierences to approximate the viscous uxes. Following Abarbanel et al. [8], since the Mac-
Cormack dierencing method cannot be applied straightforwardly to the case of mixed
derivatives present in the viscous uxes, we rewrite the NavierStokes equations in the way:
w
t

_
f(w) f
vP
(
w, w
x)
f
vM
(
w, w
y)
_
x

_
g(w) g
vP
(
w, w
x)
g
vM
(
w, w
x)
_
y
= 0
where the notations P and M stand for the parabolic or mixed derivatives parts of the viscous
uxes, ie:
f
vP
(w) =
1
Re
_
_
_
_
_
_
_
0
t
P
xx
= lu
x
2mu
x
t
P
xy
= mv
x
ut
P
xx
vt
P
xy

gm
Pr
e
x
_
_
_
_
_
_
_
f
vM
(w) =
1
Re
_
_
_
_
_
_
0
t
M
xx
= lv
y
t
M
xy
= mu
y
ut
M
xx
vt
M
xy
_
_
_
_
_
_
and similarly in the y direction. The scheme is then implemented as follows:
w
+
i, j
= w
n
i, j

dt
dx
_
_
f f
vP
_
n
i1, j

_
f f
vP
_
n
i, j
_

dt
dy
_
_
g g
vP
_
n
i, j1

_
g g
vP
_
n
i, j
_

dt
2Xdx
_
f
vM
n
i1, j
f
vM
n
i1, j
_

dt
2Xdy
_
g
vM
n
i, j1
g
vM
n
i, j1
_
where the derivatives terms appearing in the parabolic uxes are expressed by backward
dierencing, while the derivative terms appearing in the mixed derivative uxes are expressed
by central dierencing.
w
++
i, j
= w
+
i, j

dt
dx
_
_
f f
vP
_
+
i, j

_
f f
vP
_
+
i1, j
_

dt
dy
_
_
g g
vP
_
+
i, j

_
g g
vP
_
+
i, j1
_

dt
2Xdx
_
f
vM
+
i1, j
f
vM
+
i1, j
_

dt
2Xdy
_
g
vM
+
i, j1
g
vM
+
i, j1
_
here derivative terms appearing in the parabolic uxes are expressed by forward dierencing,
while the derivative terms appearing in the mixed derivative uxes are still expressed by central
dierencing.
w
n1
i, j
=
1
2
_
w
n
i, j
w
++
i, j
_
C
x
i1a2, j
C
x
i1a2, j
C
y
i, j1a2
C
y
i, j1a2
As the scheme is explicit, the stability condition on the time step dt is expressed by:
dt = CFL min
_
dt
E
, dt
V
_
, CFL0X5
V. Daru, C. Tenaud / Computers & Fluids 30 (2001) 89113 94
where
dt
E
=
min(dx, dy)
(
u
2
v
2
)
1a2
c
, dt
V
=
1
2
(min(dx, dy))
2
Pr
g
r
m
Re
c being the sound velocity. In our case, there is no need to implicit the viscous terms, as the
Euler time step dt
E
is generally smaller than the viscous time step dt
V
, at least after reexion of
the shock wave on the end wall, due to large values of the sound velocity.
3.2. Independent time and space discretization methods: the HY and MUSCL schemes
The resolution of the governing Eq. (1) has also been performed by means of a nite volume
method. The discrete equations read as follows:
_
dw
dt
_
i, j
=
1
dx
_
f
n
i1a2, j
f
n
i1a2, j
f
v
n
i1a2, j
f
v
n
i1a2, j
_

1
dy
_
g
n
i, j1a2
g
n
i, j1a2
g
v
n
i, j1a2
g
v
n
i, j1a2
_
(2)
3.2.1. Temporal integration
The time integration of the previous Eq. (2) represented as
dw
dt
= v(w) (3)
is performed by means of a third order Runge-Kutta method [9]
w
0
= w
n
w
1
= w
0
dtv
(
w
0)
w
2
=
3
4
w
0

1
4
w
1

1
4
dtv
(
w
1)
w
(n1)
=
1
3
w
0

2
3
w
2

2
3
dtv
(
w
2)
(4)
where dt is the time step and w
n
stands for the conservative variable vector evaluated at the
time n dtX This method has been choosen since it does not increase the total variation of the
R.H.S. of Eq (2). The stability limits of this temporal scheme are the ones of an explicit
scheme, that means, the Courant number (CFL) and the diusion number (D) must be less
than, 1. and 0.5. respectively. In the present calculations, the time step (dt) is prescribed to
ensure that the CFL and D numbers satisfy the stability limits. Let us mention that the local
CFL number is calculated using the spectral radius of the Jacobian matrices of the Euler
uxes.
V. Daru, C. Tenaud / Computers & Fluids 30 (2001) 89113 95
3.2.2. Space discretization
The integration of the convective terms is based on a Roe's approximate Riemann solver
[10]. Two shock capturing schemes have been implemented:
. A second-order upwind total variation diminishing (TVD) scheme, developed by Harten [11]
and Yee et al. [12]. The numerical ux f
i1a2, j
is expressed as
f
i1a2, j
=
1
2
_
f
i1, j
f
i, j
R
x
i1a2, j
r
x
i1a2, j
_
r
x
i1a2, j
is a 4 4 matrix; to ensure that the scheme is a second order upwind TVD scheme,
the r
l
x
i1a2, j
elements must be written [11,12]:
r
l
x
i1a2, j
=
1
2

a
l
x
i1a2, j

_
F
l
x
i1, j
F
l
x
i, j
_

a
l
x
i1a2, j
z
l
x
i1a2, j

da
l
x
i1a2, j
(5)
with da
x
i1a2Y j
= R
1
x
i1a2Y j
(w
i1Y j
w
iY j
) is the forward dierence of the local characteristic
variables in the x direction, a
l
x
i1a2, j
represent the eigenvalues of the Jacobian of the Euler
ux (dfadw), R
x
i1a2Y j
is the matrix whose columns are composed of the eigenvectors of
dfadwX The ``additional eigenvalues'' are expressed as:
z
x
i1a2, j
=
1
2

a
l
x
i1a2, j

_
_
F
l
x
i1, j
F
l
x
i, j
_
da
x
i1a2, j
if da
x
,=0X
0X if da
x
= 0X
We must mention that, contrary to the original scheme, no entropic parameter is used as we
look for unsteady solutions of the equations. Two limiter functions, which give a second-
order TVD scheme, have been implemented:
*
The rst one is the classic Super-Bee function [13]
F
l
x
i, j
= S max
_
0, min
_
2

da
l
x
i1a2, j

, S da
l
x
i1a2, j
_
, min
_

da
l
x
i1a2, j

, 2 S da
l
x
i1a2, j
__
(6)
where S is the sign of da
l
x
i1a2, j
X
*
The second limiter is the Van-Leer Harmonic function
F
l
x
i, j
= S max
_
da
l
x
i1a2, j
da
l
x
i1a2, j

da
l
x
i1a2, j
da
l
x
i1a2, j

_
_
da
l
x
i1a2, j
da
l
x
i1a2, j
_ (7)
To estimate the values at the mid-point (i 1a2, j) and to be sure that the numerical ux
(f) is consistent with f, we have used the classic mass-weighted average introduced by Roe
[10] between states w
i1, j
and w
i, j
X A similar formulation is used to express the numerical
uxes (g) in the y direction. While the integration is third order accurate in time, the
numerical scheme is only second order in both time and space due to the space
discretization.
V. Daru, C. Tenaud / Computers & Fluids 30 (2001) 89113 96
. A MUSCL-TVD scheme using the local characteristic approach [14]. the convective
numerical ux functions (f
i1a2,j
and g
i, j1a2
) are written as follows:
f
i1a2, j
=
1
2
_
f
_
w
r
i1a2, j
_
f
_
w
l
i1a2, j
_
R
x
i1a2, j
r
x
i1a2, j
_
(8)
Spurious oscillations are eliminated using ``slope'' limiter applied on the conservative
variables (w):
w
r
i1a2, j
= w
i1, j

1
4
_
(1 Z)d
i3a2
(1 Z)d

i1a2
_
(9)
w
l
i1a2, j
= w
i, j

1
4
_
(1 Z)d

i1a2
(1 Z)d
i1a2
_
(10)
In the following calculations, a minmod ``slope'' limiter has been used with Z = 1a3, which
gives a third order upwind-biased scheme:
d
i1a2
= minmod
_
d
i1a2
, bd
i1a2
_
(11)
d

i1a2
= minmod
_
d
i1a2
, bd
i3a2
_
(12)
where d
i1a2
=w
i1Y j
w
iY j
X Several values of the b coecient have been checked. The value
b = 4 has been retained since it gives the less diusive results. The minmod function is
expressed following:
minmod
(
p, q
)
= sign
(
p
)
max
{
0, min[[p[, q sign ( p)]
The elements of the matrix r
x
i1a2, j
are written
r
l
x
i1a2, j
=

a
l
x
i1a2, j

R
1
x
i1a2, j
_
w
r
i1a2, j
w
l
i1a2, j
_
(13)
where a
l
x
i1a2, j
represents the eigenvalues and R
x
i1a2, j
are the right eigenvector matrix of
dfadw evaluated using the Roe average between the states w
r
i1a2, j
and w
l
i1a2, j
X
Concerning the diusive uxes, a central dierencing scheme is applied, giving a second
order accuracy in space.
4. A 1D inviscid test case
The O3 limiter is compared with the Van Leer limiter on the 1D test case proposed in [15]
concerning a moving Mach = 3 shock interacting with sine waves in density. The 1D Euler
equations are solved on the spatial domain x c [0, 10]X Absorbing boundary conditions are
used. The solution is initially prescribed as:
r = 3X857; u = 2X629; p = 10X333 when x ` 1
V. Daru, C. Tenaud / Computers & Fluids 30 (2001) 89113 97
r = 1 0X2 sin(5x); u = 0; p = 1 when xr1 (14)
A Lax-Wendro-type discretization is used as described in Section 3.1. Figs. 2 and 3 show the
density distribution obtained at a dimensionless time t = 1X8, using the two limiters, for 400
and 800 grid points and a CFL number equal to 0.8 and 0.1, respectively. The thin solid lines
are numerical solutions with 3200 grid points using the same algorithm, they can be regarded
as the exact solution. It is clear that the O3 limiter improves the solution whatever the CFL
number is, compared to the Van Leer limiter which is more diusive. The comparison with the
results obtained using the O3Sup limiter (see [2]) shows that O3 is slightly more diusive than
O3Sup. However, these authors mentioned that O3Sup is too compressive for non linear
systems, a behaviour that was not encountered using O3. Let us notice that the O3 and O3Sup
limiters using 800 points are comparable to ENO schemes using 400 points (see results in [15]),
but the latter is more expensive in CPU cost and storage.
Now, it is worth to compare the above results with the results obtained using the HYVL
scheme. The latter are represented in Fig. 4, for a CFL number equal to 0.8 (quite similar
results are obtained for CFL = 0.1). One can see that the use of a Mac-Cormack time
discretization give much better results for this CFL number. This can be explained by the fact
that in the separate timespace discretization approach, the limiting procedure is applied in
space at each step of the Runge-Kutta algorithm. By doing so, the underlying scheme which is
used around extrema is rst order in space, but still third order in time. This imply that the
leading term of the truncation error becomes
1
2
dxau
xx
around extrema in the scalar linear case
(so it is independent of the CFL number). In the case where a combined time-space
discretization is used, we rather have
1
2
dxa(1 n)u
xx
as we get the upwind (rst order in time
and space) scheme. These errors are equivalent if one uses a small CFL number, but the
diusive error generated by the Runge-Kutta procedure is much higher if the CFL number is
close to 1. This is attenuated in the system case, as the local CFL number is close to 1 only for
the higher velocity wave. Anyway, it is to be expected that the combined time-space
discretization will give better results for unsteady ows, whatever the limiter is.
Fig. 2. Density distribution for the Shu-Osher test case at t = 1X8: comparison of the O3 and Van Leer limiters, for
400 (left) and 800 (right) grid points, CFL = 0.8.
V. Daru, C. Tenaud / Computers & Fluids 30 (2001) 89113 98
5. A 2D inviscid test case: advection of a vortex
We compare here the results obtained on the inviscid test case of the advection of a vortex
at low Mach number. This test case is treated in [16]. The vortex is described initially by an
analytical form of Sculley. The angular velocity of the initial vortex is given as:
v
y
V
max
=
_

_
r
a
0
if ra
0
exp
_

(
r a
0)
2
O
_
if r b a
0
where V
max
= 0X3, a
0
= 0X06 and O = 0X065X The pressure and density are obtained from the
Fig. 3. Density distribution for the Shu-Osher test case at t = 1.8: comparison of the O3 and Van Leer limiters, for
400 (left) and 800 (right) grid points, CFL = 0.1.
Fig. 4. Density distribution for the Shu-Osher test case at t = 1X8, HYVL scheme, for 400 (left) and 800 (right) grid
points, CFL = 0.8.
V. Daru, C. Tenaud / Computers & Fluids 30 (2001) 89113 99
relations:
dp
dr
=
rv
2
y
r
gp
(g 1)r

v
2
y
2
=
gp
o
(g 1)r
o
The vortex is initially superimposed on a uniform ow with Mach number M
o
= 0X1X The
computational domain considered is [1, 4] [1, 1]X
We compare the results obtained using the three following numerical schemes described
above:
1. the Mac-Cormack scheme, using the O3 and Van Leer limiters hereafter noted MCO3 and
MCVL;
2. the Harten-Yee scheme equipped with the Van Leer limiter, called HYVL;
3. and the MUSCL scheme with a Minmod(b = 4) limiter, called MUMI4.
The mesh is cartesian uniform, composed of 251 101 cells. In Fig. 5 is represented the
minimum pressure (core pressure ), reached at the center of the vortex, as a function of the
length of vortex travel Xcore/a
0
, where Xcore is the abscissa of the center of the vortex. The
ow being inviscid and low-speed, the true solution is almost a pure advection of the initial
vortex, i.e. the change of the core pressure is mainly due to the numerical diusion of the
numerical scheme and provides then a good evaluation of the dissipative properties of the
schemes. The following comments can be done:
Fig. 5. Evolution of pressure at the center of the vortex (O3 and VanLeer limiters) vs. the length of the vortex
travel.
V. Daru, C. Tenaud / Computers & Fluids 30 (2001) 89113 100
1. the best results, which are nearly equivalent, are obtained using the MCO3 or MUMI4
schemes;
2. the Van Leer limiter, which is used in the two other results, is more diusive. We notice that
the Harten-Yee scheme is more diusive than the Mac Cormack scheme when using the
same limiter (Van Leer). This again should be attributed to the fact that the diusion error
of the RK3 algorithm around extrema is higher than for the Mac Cormack algorithm. The
dierence is weak in this case because the Mach number is low, implying that the local CFL
number for the material wave which transports the vortex is small (around 0.05).
The next gure, Fig. 6, shows the longitudinal distribution of the static pressure at a
dimensionless time t = 30 for the four schemes we used. Let us mention that the O3 and
O3Sup limiters give almost the same results for this test case, so only the O3 results are shown.
We should also mention that the same computation has been conducted with the SuperBee
limiter, using both the combined and independent timespace discretization methods. It is well
known that this limiter tend to ``square o'' smooth proles. In the test case considered here,
this behavior leads to a very surprising result (see Fig. 7): the core pressure in the vortex
decreases as time goes on, which is very unphysical. Although we expect that this behavior will
be counteracted in the viscous case by the viscous eects, numerical problems can arise as the
core pressure can become too low or negative in the center of vortices.
Simulations have also been performed using the MUSCL approach with other values of the
b coecient, showing that the smaller the value of b is, the more diusive the results are.
As a conclusion following the previous results, we will use three schemes to compute the
shock-boundary layer interaction:
. the TVD Harten-Yee scheme with the Van-Leer Harmonic limiter (HYVL);
. the MUSCL-TVD scheme using the Minmod limiter with b = 4 (MUMI4);
Fig. 6. Longitudinal distribution of the static pressure at y = 0 and for a dimensionless time t = 30X
V. Daru, C. Tenaud / Computers & Fluids 30 (2001) 89113 101
. the Mack-Cormack scheme using the O3 or O3Sup limiter functions (MCO3 and
MCO3Sup). The latter did not show the compressive behavior mentioned in [2] in the
inviscid test cases, so we will use it as well in the viscous case.
The other schemes we have tested will not be used in the following, as either they give
equivalent results, or they are not suciently ``safe''.
6. Reected shock-boundary layer interaction in a shock tube
6.1. Description of the ow
We consider a unit side length square shock tube with insulated walls (Fig. 8). The
diaphragm is initially located in the middle of the tube (x = 0X5). The initial state, in terms of
dimensionless quantities, is on the left of the diaphragm: r
L
= 120, p
L
= r
L
ag, u
L
= v
L
= 0, and
on the right: r
R
= 1X2, p
R
= r
R
ag, u
R
= v
R
= 0X At the initial time, the diaphragm is broken.
The inviscid solution is represented in Fig. 9, showing the evolution of the density in the x t
plane. A shock wave, followed by a contact discontinuity, moves to the right (the shock Mach
number is equal to 2.37). The incident shock wave is weak, and reects at the right end wall
approximately at time t = 0X2X After this reection, it interacts with the contact discontinuity.
Then, complex interactions occur. The contact discontinuity stays stationary, close to the right
end wall. Afterwards, the reected shock wave begins to interact with the rarefaction wave at
time t = 0X4X On the other side, the rarefaction wave reects on the left wall, at a dimensionless
time t = 0X5, modifying the propagation of the incident rarefaction wave.
In the viscous case, the incident contact discontinuity and shock wave, during their
propagation, interact with the horizontal wall, creating a thin boundary layer. After its
Fig. 7. Evolution of pressure at the center of the vortex (SuperBee limiter) vs. the length of the vortex travel.
V. Daru, C. Tenaud / Computers & Fluids 30 (2001) 89113 102
reection on the right wall, the shock wave interacts with this boundary layer, modifying the
ow pattern near the horizontal wall. This is illustrated in Fig. 10, showing the density
contours at t = 0X6X The shock-boundary layer interaction results in a lambda-shape like shock
pattern. As the stagnation pressure in the boundary layer is lower than that of the outow
region, a separated boundary layer ``bubble'' takes place under this shock pattern. The bubble
is delimited by a supersonic shear layer in which a lot of instabilities occur. The triple point
emerging from the lambda-shape like shock pattern generates a slip line which rolls up around
Fig. 8. Representation of the shock tube.
Fig. 9. xt diagram of the density.
V. Daru, C. Tenaud / Computers & Fluids 30 (2001) 89113 103
the vortex downstream of the boundary layer bubble. The bubble is dragged upstream with the
reected shock and separations and vortices are generated near the wall inside the bubble.
The solution being supposed to be symmetric in the vertical direction (i.e. parallel to the
diaphragm), the computations have been performed on half of the shock tube, using the three
algorithms presented above. Two Reynolds numbers have been considered (Re = 200,
Re = 1000). The results will be discussed on the density eld for the dimensionless time t = 1X
In all the computations, we used equally spacing grid such as that dx = dyX
6.2. Wall boundary conditions
On the solid walls, we impose a no-slip condition, i.e. the velocity is set to zero. We treat the
case of an adiabatic wall, so the normal derivative of the temperature is set to zero using a
second order forward dierence formula, giving the temperature at the wall. The last value
needed at the wall can then be obtained from two dierent procedures:
. the rst one allows the density to be computed from the continuity equation, where the
tangential derivatives are set to zero. A second order forward dierence formula is used for
the space discretization. The time discretization uses either a Crank-Nicolson or a Runge-
Kutta method. The pressure is then deduced from the equation of state for the perfect gas;
. the second procedure is more classical, using the equation of moment normal to the wall
where the velocity is set to zero at the wall, leading to an equation giving the normal
derivative of the pressure at the wall, which is discretized using second order dierence
formulae.
The rst procedure is simple to implement, as there are no viscous term in the equation for the
density. Nevertheless, the time derivative term, not present in the second procedure, does
introduce additional discretization errors. We have studied the inuence of these boundary
conditions. At low Reynolds numbers, these two procedures lead to very similar results. The
dierences increase with the Reynolds number. No evident conclusion emerges to choose which
Fig. 10. Re = 200, t = 0X6, density contours and streamlines.
V. Daru, C. Tenaud / Computers & Fluids 30 (2001) 89113 104
procedure is the more accurate. In the following paragraphs, all the presented results have been
obtained using boundary conditions based on the solution of the continuity equation.
6.3. Re = 200
For this low Reynolds number value, the computations have been performed using the three
algorithms MUMI4, HYVL and MCO3 on three meshes (dx = 4 10
3
, dx = 2 10
3
and
dx = 10
3
). The inuence of the grid renement is presented on the distribution of the density
along the horizontal wall (Fig. 11) at a dimensionless time t = 1 and for the HYVL and
MCO3 schemes. It can be stated that the coarsest grid is insucient for this ow. The middle
grid is not yet sucient for convergence, but all the structures that are seen on the nest grid
are correctly represented. We can also observe, when comparing the results for the two shemes,
that there still exists some dierences even in the nest grid, notably on the peak levels. Fig. 12
shows a comparison of the density eld obtained on the ``middle grid'' (501 251) with the
three schemes at a dimensionless time t = 1X We see that large vortices are generated within the
boundary layer bubble, and are ejected downstream the lambda-shape like pattern. The
position of these large scale patterns are in agreement with the peaks and valleys recorded in
the longitudinal evolution of the density along the wall. Comparing the results obtained with
the studied schemes, some dierences can be pointed out. In the MUSCL result, we can notice
small oscillations behind the reected shock, which do not appear for the other schemes. This
is probably a numerical artifact due to the compressive behavior of the Minmod (b = 4)
limiter. There are also some dierences in the orientation of the large vortex and in the pattern
of the contact discontinuity which rolls up in the lower right corner of the tube. The MCO3
and HYVL schemes give very similar results, but the HYVL results are more diusive. This
can be seen by examining more closely some details of the ow close to the bottom wall inside
the boundary layer bubble, for example around x = 0X6 or x = 0X76X
Fig. 11. Distribution of the density along the bottom wall at a time t = 1X Left: HYVL scheme, right: MCO3
scheme.
V. Daru, C. Tenaud / Computers & Fluids 30 (2001) 89113 105
When rening the mesh, setting dx = 10
3
, the HYVL and MCO3 results show few
dierences with the precedent mesh (see Fig. 13). The main dierence lies in the large vortex,
which is slightly more rotated in the MCO3 result. The MCO3 scheme again shows a less
diusive behavior than the HYVL scheme. This can also be seen on the longitudinal
distribution of the density in the ``inviscid'' region (y = 0X3), at time t = 1 (Fig. 14), in the
contact discontinuity region close to the right-end wall.
Let us also say that we have run the MCVL and MCO3Sup schemes for this case and found
that the results are very close to the MCO3 results in the middle mesh. In the ne mesh, the
Fig. 12. Re = 200, t = 1, density contours. Top left: MUMI4 scheme. top right: HYVL scheme, bottom MCO3
scheme. Mesh size dx = 2 10
3
.
Fig. 13. Re = 200, t = 1, density contours. Left: HYVL scheme, right: MCO3 scheme, Mesh size dx = 10
3
.
V. Daru, C. Tenaud / Computers & Fluids 30 (2001) 89113 106
results are almost indistinguishable. This leads us to state that we have attained convergence in
the ne mesh when using the Mac-Cormack time discretization and that for this low Reynolds
number value the O3 limiters are almost equivalent to the Van Leer limiter. In comparison, the
HYVL algorithm seems to converge rather slowly, as there still exists some signicant
dierences with the MCO3 results in the ne mesh. This is probably explained by the
arguments we developed in the 1D case and the fact that this is a high Mach number ow.
Even in the boundary layer bubble, we nd high values of the Mach number, as can be seen in
Fig. 15 where is represented the Mach number distribution along y = 0X01X
Fig. 14. Comparison of the longitudenal distribution of the density at y = 0X3 given by the HYVL and MCO3
schemes on the ne grid (1000 500) at time t = 1 (Re = 200).
Fig. 15. Mach distribution along y = 0X01 at a time t = 1X MCO3 scheme, dx = 10
3
X
V. Daru, C. Tenaud / Computers & Fluids 30 (2001) 89113 107
6.4. Re = 1000
The rst computation for this Reynolds number has been performed on the middle mesh we
previously used (dx = 2 10
3
). We compare the density elds obtained at t = 1 (Fig. 16)
using the MUMI4, HYVL and MCO3Sup schemes (we have used the O3Sup limiter rather
than the O3 limiter in order to have better chances to reach grid convergence, as the O3Sup
limiter is slightly less diusive; let us say, however, that the results obtained using the MCO3
or MCO3Sup schemes are very similar). At this Reynolds number, the discrepancies between
the results are very important. This is attributed to a lack of resolution in the boundary layer,
leading to a misrepresentation of the separation points at the wall. The eect of the numerical
error specic to each scheme is here very marked. The MUMI4 scheme reveals, in this case, a
strange behavior since the lambda-shape like shock pattern has almost completely disappeared
(Fig. 16). The boundary-layer bubble and especially the large vortex at the bottom of the
bubble are attened. This result seems to be completely unrealistic. It is probably due to the
compressive nature of the limiter (Minmod with b = 4). A conclusion which can be drawn
from this result is that compressive limiters, although suitable in some cases as they counteract
the numerical diusion, should not be used to compute viscous ows. This is the reason why,
in the following, the MUMI4 scheme will not be used.
The results obtained using the MCO3 and HYVL schemes are notably dierent concerning
Fig. 16. Re = 1000, t = 1, density contours. Mesh size dx = 2 10
3
X Top left: MUMI4 scheme, top right: HYVL
scheme, bottom: MCO3Sup scheme.
V. Daru, C. Tenaud / Computers & Fluids 30 (2001) 89113 108
the vortical structures inside the boundary layer bubble (Fig. 16). The triple point is higher in
the MCO3Sup case. One can also remark that the ``inviscid'' zone, above the slip line initiated
at the triple point, is now perturbed by acoustic waves generated by the growth of the vortices
within the boundary layer.
The rst mesh we used is in fact very insucient to give a correct representation of the
solution, whatever the scheme is. We then have made a series of computations on successively
rened meshes, corresponding to dx = 10
3
, dx = 6X7 10
4
and dx = 5 10
4
X The results
are represented in Fig. 17. We can see that small structures are generated just downstream of
the boundary layer separation. These structures produce distortions of the unstable line
delimiting the separation under the lambda-shape like shock pattern. These warpings induce
oblique shock waves moving along this slip line. We can also notice that shocklets are
Fig. 17. Re = 1000, t = 1, density contours. Left: HYVL scheme, right: MCO3Sup scheme. Mesh size dx = 10
3
(top), dx = 6X7 10
4
(middle), dx = 5 10
4
(bottom).
V. Daru, C. Tenaud / Computers & Fluids 30 (2001) 89113 109
generated inside the large vortices at many locations. All these ow characteristics are very
sensitive to small perturbations. As a consequence, it is extremely dicult to ascertain the
quality of the schemes, because there does not probably exist a uniquely dened convergence
process due to the instability of the ow. However, as a general comment, we can observe that
the HYVL results are smoother and the variation of the density eld when the mesh is rened
seems weak. On the contrary, the MCO3Sup results are very sensitive to the mesh size. Also,
they are more chaotic, especially inside the boundary layer bubble where there is no regular
structures as in the HYVL results. Nevertheless, the last two MCO3Sup results seem
suciently close one to the other to think that convergence is not far from the last mesh.
In the case of the HYVL scheme, the ner result shows that the regularity of the solution
Fig. 18. Re = 1000, t = 1, velocity vector eld and instantaneous streamlines. Left: HYVL scheme, right: MCO3Sup
scheme. Mesh size dx = 10
3
(top), dx = 6X7 10
4
(middle), dx = 5 10
4
(bottom).
V. Daru, C. Tenaud / Computers & Fluids 30 (2001) 89113 110
begins to be broken. More complex structures appear, like for example the structure on the left
side of the big vortex, which is very similar to the one in the MCO3Sup result. The
convergence of this scheme is slower, but we can expect that if we rene the mesh again we
will obtain a solution close to the one given by the MCO3Sup scheme.
It is interesting, in order to give some more justications to our previous conclusions about
the convergence of the schemes, to get a closer insight in the right bottom corner. We have
represented in Fig. 18, the velocity vector eld obtained using the two schemes in the three
ner meshes. Here, we clearly see that the MCO3Sup scheme converges much more faster than
the HYVL scheme: the results on the two ner meshes are almost the same, indicating that the
solution is almost converged in this area on the intermediate mesh. Moreover, the four vortices
are already captured in the coarsest mesh, though the small one in the corner is barely
outlined. In contrast, we can see that three vortices are hardly captured using the HYVL
scheme on the intermediate mesh. The four vortices appear only on the ner mesh. The
coarsest mesh solution reproduces only one vortex. The examination of this gure also shows
that the convergence process of the two schemes are quite dierent. This is probably due to the
fact that the truncation error of the HYVL scheme is less dependent on the local CFL number.
Finally, in order to illustrate the contribution of the limiter, we show in Fig. 19, the results
obtained using the MCVL scheme on the mesh dened by dx = 10
3
X The density eld is seen
to be more regular than the MCO3Sup result (it looks more like the HYVL result). If now we
inspect the corner ow, we can again say that the combined timespace discretization is
without doubt the more accurate, but it is also seen that the O3Sup limiter gives better results,
as the small vortex in the corner is not present in the MCVL solution, and the vortex at the
left is not as well developed as in the MCO3Sup result. This shows that the combination of a
Mac-Cormack type scheme with a O3-type limiter is the best choice to do.
Fig. 19. Re = 1000, t = 1, density contours. MCVL scheme. Left: density contours, right: velocity vectors and
instantaneous streamlines. Mesh size dx = 10
3
X
V. Daru, C. Tenaud / Computers & Fluids 30 (2001) 89113 111
6.5. Comments
When we observe the dependence of the lambda-shape like pattern on the Reynolds number,
our results show that the slope of the leading oblique shock wave stays constant in the
Reynolds number range when we studied. This is in agreement with Mark's theoretical
predictions based on an inviscid assumption. However, a tendency of the triple point to raise
as Re increases, is noticeable on our results. This is in contradiction with the conclusion drawn
by Weber et al. [7], who observed that the asymptotic bifurcation height is larger for lower
Reynolds numbers. This could be related to the grid resolution, which have been shown in our
results to have a very important inuence on the ow pattern. We have also noticed that the
results are sensitive to the cell mesh ratios dxadyX As there exists a lot of small eddy structures
in the viscous ``bubble'', it seems more appropriate to use a uniform mesh inside this area.
7. Conclusions
Our goal in this study was to evaluate the capability of some high resolution TVD schemes
to solve complex unsteady viscous shocked ows. We have considered two types of
discretization, namely a combined time and space (CTS) discretization and an independent
time and space (ITS) discretization. Both methods are associated with several limiters. In the
case of the combined time and space discretization, a new family of limiters have been
considered.
The accuracy properties of each scheme have rst been reviewed on inviscid 1D and 2D
computations. When the same limiter is used, the best results are obtained using the CTS
method. This is explained by showing that around extrema, the recovered rst order scheme is
less diusive for this method than for the ITS method. The so-called O3 family of limiters is
shown to be more accurate than the classical ones.
We have then studied the application of the dierent schemes to the ow produced by the
interaction of a reected shock wave with the incident boundary layer in a shock tube. The
calculations have been performed for two values of the Reynolds number. At Re = 200, we
have shown that the CTS method converges much more faster than the ITS method. The O3
limiter does not improve the results signicantly compared to the Van Leer limiter.
When the Reynolds number is increased to the value of 1000, the ow becomes much more
complex. The boundary layer separates in several places, giving rise to the development of a lot
of vortices with shocklets and large compressibility eects. For this ow, we can only exhibit
an ``almost converged'' numerical solution due to the excessive grid requirements. However,
the close examination of the results leads us to conclude that the CTS method associated with
a O3 limiter is from far the best scheme in terms of accuracy and convergence properties. A
sensible improvement is notably due to the limiter himself, when compared to good quality
classical limiters.
To conclude, we think that this viscous shock tube problem could constitute an interesting
test case to compare high resolution numerical schemes in the case of compressible viscous
unsteady ows, at least for low values of the Reynolds number. We have considered in this
work only TVD schemes. There exists other approaches, particularly the ENO and WENO
V. Daru, C. Tenaud / Computers & Fluids 30 (2001) 89113 112
schemes which are gaining popularity as they allow a higher order of accuracy, generally at the
expense of an increased cost. It should be interesting, in a future work, to test these schemes
on this ow case.
Acknowledgements
Part of the computations have been carried out on the Cray C98 of the I.D.R.I.S./C.N.R.S.
The authors greatly acknowledge the support of these institutions.
References
[1] Jeng NY, Payne UJ. An adaptive TVD limiter. Journal of Computational Physics 1995;118:22941.
[2] Arora M, Roe PL. A well-behaved TVD limiter for high-resolution calculations of unsteady ows. Journal of
Computational Physics 1997;132:311.
[3] Daru V, Fernandez G, Tenaud C. On CFD to investigate bifurcated shock wave pattern. In: Houwing AFP,
editor. Proceedings of the 21st International Symposium on Shock Waves, Australia. Fyshwick: Panther
Publishing & Printing, 1997. p. 10917.
[4] Roe PL, Baines MJ. Algorithms for advection and shock problems. In: Viviand H, editor. Proceedings of the
Fourth GAMM Conference on Numerical Methods in Fluid Mechanics. Braunschweig: Vieweg, 1981. p. 281
90 Notes on Numerical Fluid Mechanics, 5.
[5] Kleine H, Lyakhov LG, Gvozdeva LG, Gronig H. Bifurcation of a reected shock wave in a shock tube. In:
Proceedings of the 18th International Symposium on Shock Waves, Sendai. Berlin: Springer-Verlag, 1991. p.
2616.
[6] Wilson GJ, Sharma SP, Gillespie WD. Time-dependent simulation of reected-shock/boundary layer
interaction in shock tubes. In: Proceedings of the 19th International Symposium on Shock Waves, Marseille.
Berlin: Springer-Verlag, 1993. p. 43944.
[7] Weber YS, Oran ES, Boris JP, Anderson JD. The numerical simulation of shock bifurcation near the end wall
of a shock tube. Physics of Fluids 1995;7:247588.
[8] Abarbanel S, Gottlieb D. Optimal time splitting for two and three-dimensional NavierStokes equations with
mixed derivatives. Journal of Computational Physics 1981;41:133.
[9] Shu CW, Zang TA, Erlebacher G, Whitaker D, Osher S. High-order ENO schemes applied to two- and three-
dimensional compressible ow. Applied Numerical Mathematics: Transactions of IMACS 1992;9:4571.
[10] Roe PL. Approximate Riemann solvers, parameter vectors and dierence schemes. Journal of Computational
Physics 1981;43:35772.
[11] Harten A. High resolution schemes for hyperbolic conservation laws. Journal of Computational Physics
1983;49:35793.
[12] Yee HC, Harten A. Implicit TVD schemes for hyperbolic conservation laws in curvilinear coordinates. AIAA
Journal 1987;25:26674.
[13] Yee, H.C., Numerical experiments with a symmetric high resolution shock-capturing scheme. NASA TM-
88325, (1986).
[14] Yee, H.C., On the implementation of a class of upwind schemes for system of hyperbolic conservation laws.
NASA TM-86839, (1985).
[15] Shu CW, Osher S. Ecient implementation of essentially non-oscillatory shock-capturing schemes. II, Journal
of Computational Physics 1989;83:3279.
[16] Lin SY, Chin YS. Comparison of higher resolution Euler schemes for aeroacoustic computation. AIAA Journal
1995;33:23745.
V. Daru, C. Tenaud / Computers & Fluids 30 (2001) 89113 113

Vous aimerez peut-être aussi