You are on page 1of 12

Engineering Applications of Computational Fluid Mechanics Vol. 1, No.4, pp.

325336 (2007)

A METHOD BASED ON RIEMANN PROBLEM IN TRACKING


MULTI-MATERIAL INTERFACE ON UNSTRUCTURED MOVING GRIDS
Bing Wang* and Houqian Xu
College of Power Engineering, Nanjing University of Science & Technology,
Nanjing, P.R. China 210094
* E-Mail: evancfd@163.com (Corresponding Author)
ABSTRACT: The material interface is tracked by solving the arbitrary Lagrangian-Eulerian (ALE) formulation during
the simulation of the compressible multi-material flow. The material interface is looked upon as a Lagrangian interface
which can move freely and is composed of a number of edges of the unstructured grids and the state vectors of the points
on the interface have two different definitions corresponding to the two different fluids. Then, Riemann problem is
solved by the two-shock approximation method for general form of equation of state on both sides of the interface to
track the interface accurately and the grids are moving automatically with the motion of the interface. The 1D
spherically symmetric underwater explosion model is computed by using ALE method and the numerical results agree
well with the experimental data, which indicates that the interface tracking method is reasonable. Furthermore,
interaction between shock and water surface is also simulated to show that this method is suitable for solving microdeforming interface problem.
Keywords: multi-material interface, unstructured moving grids, Riemann problem, ALE method, HLLC scheme,
underwater explosion

to solve the multi-material flows accurately with a


micro-deforming interface.
In this paper, an ALE method based on moving
grids (Hirt, Amsden and Cook, 1974) is used during
the simulation of the compressible multi-material
flow. The material interface is looked upon as a
Lagrangian interface which can move freely and is
composed of a number of edges of the unstructured
grids and the state vectors of the points on the
interface
have
two
different
definitions
corresponding to the two different fluids. Then, the
Riemann problem is solved on both sides of the
interface to track the interface accurately and the
grids are moving automatically with the motion of
the interface. The 1D spherically symmetric
underwater explosion model is computed by using
ALE method and the numerical results agree well
with the experimental data, which indicates that the
interface tracking method is reasonable.
Furthermore, interaction between shock and water
surface is also simulated to show that this method is
suitable for solving micro-deforming interface
problems.

1. INTRODUCTION
Multi-material flow, where a freely moving
interface exists between two immiscible fluids, can
be found in many engineering problems. It is still
difficult to track the material interface accurately
and efficiently in computational fluid dynamics.
There are a number of numerical methods
developed to solve the interface problems, such as
VOF method (Hankin, 2001; Wang et al., 2004;
Tang, Wrobel and Fan, 2004), Level-set method
(Xiao and Ebisuzaki, 1998; Xu, 1997), Gamma
based model method (Abgrall, 2003), etc. However,
some common problems are encountered in using
these methods. First, the interfaces location is
obtained with all of these methods by solving some
special formulations, which cannot show directly
the motion of the interface. Second, these methods
are all based on the stationary grids, where the state
vectors near the interface are obtained through
interpolation, and even a small error may lead to
collapse or meaningless computation (Fedkiw,
Marquina and Merriman, 1999; Blom, 2000) unless
some special method is introduced at the interface,
such as Ghost Fluid Method (Fedkiw et al.,
1999). Lastly, it is difficult for all of these methods
Received: 16 Apr. 2007; Revised: 7 Jun. 2007; Accepted: 10 Jul. 2007
325

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 4 (2007)

2. ALE FORMULATIONS

3. NUMERICAL METHODS

The inviscid compressible Euler formulations on a


free deformable finite control volume (t) are

The numerical form of ALE formulations can be


written as

()
t

Ud +

()
t

(F x& U ) nd =

Sd

(t )

(1)
U n +1

and the flow variable vector U and flux vector F are


defined by

U = v
E

v
F = v v + p
Ev + pv

(2)

the HLLC flux vector Fi,HLLC


on their common
j
boundary i, j can be obtained by

FiHLLC
,j

(3)

j
( S j un j ) j

1
*

( S j un j )( v) j + ( p p j )n
U j = ( v) j =

( E ) S j S M

( S j un j )( E ) j p j un j + p S M

(9)

When the control volume is moving or deforming


arbitrarily, the conservation of ALE formulation (1)
will probably collapse unless the following
geometric conservation law (Nkonga, 2000;
Thomas and Lombard, 1978) (GCL) is satisfied.

()
t

(7)

(8)

(4)

x& nd = 0

SM 0 S j
Sj < 0

i
( S i u ni ) i

1
*

U i = ( v ) i =
( Si uni )( v)i + ( p pi )n
( E ) Si S M ( S u )( E ) p u + p S
i
ni
i
i ni
M

Si > 0
Si 0 < S M

F(Ui )
F ( U )

i
=

F (U j )
F (U j )

where

where e is the specific internal energy and related to


the specific total energy by
v
E =e+
2

(6)

where U n is the cell-averaged variable vector on


the control volume at the time t n , and N is the
number of the volumes boundaries. F HLLC is the
numerical flux vector computed in the HLLC
method (Toro, Spruce and Speares, 1994), which is
a numerical method based on the approximate
Riemann problems.
When j is one of the neighboring volumes of i ,

S is the symmetry source vector in the symmetric


flow and otherwise S = 0 . , v, p and E are density,
vector of velocity, pressure and specific total energy
of the fluid, respectively. n denotes the unit outward
normal vector to the moving boundary of the
volume (t), whose moving velocity is x& . If x& = 0 ,
Eqs. (1) and (2) correspond to the Eulerian
description of the flow problem. If x& = v , the
equations correspond to the Lagrangian description
of the flow problem, where the control volume
moves instantaneously with the fluid.
This set of formulations is completed by the
addition of an equation of state (EOS) which
establishes the relationship between, at most, three
thermodynamic variables. Here, the general simple
EOS form is
p = p( , e )

n n N HLLC
i t + n U n
S Fi
i =1

=
n +1

(5)

In this paper, GCL is satisfied by computing n +1


from Eq. (5) during the iteration in solving Eq. (1).

S M i

Fi F(U i ) =
S M ( v )i + p n

S ( E ) + ( S + x& n) p
i
M
M

(10)

S M j

F j F(U j ) =
S M ( v) j + p n

S M ( E ) j + ( S M + x& n) p

(11)

)(

p = i u n Si u n S M + pi
i

= j u n S j u n S M + p j
j
j

326

(12)

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 4 (2007)

uni = ( vi x& ) n
SM =

un j = ( v j x& ) n

j un ( S j un ) iun ( Si un ) + pi p j
j

j ( S j un ) i ( Si u n )
j

(13)

Interface
A

(14)

( )

(15)

( )

(16)

S i = max u n j + c j , v x& n + c

(A*, v*, p*)


Fig. 1

with v and c being the Roes average variables for


velocity and sound speed.

One dimension

As it is known that the material interface is actually


a kind of contact discontinuity, motion of the
material interface can be tracked by solving a
Riemann problem (Cocchi and Saurel, 1997; Koren
et al., 2002; Colella, 1985) at the interface. Fig. 1
shows the Riemann problem at the material
interface between fluid A and B in one dimension.
The interface node has two different definitions
corresponding to different fluids. The two variable
vectors of fluid A and B at the interface node are
U A and U B , which define the initial condition of
the Riemann problem at the interface.
vA
vB

pA )

pB )

(17)

(B*, v*, p*)

Riemann problem at the material interface.

Method I: A pure Lagrangian formulation at


the material interface

As it is mentioned above, the material interface is


also the moving boundary of the grids near the
interface and its velocity v* is already obtained by
solving the Riemann problem defined at the
interface. So x& in the ALE formulation (1) at the
material interface is equal to v* and the numerical
flux at the material interface F MI can be written as
follows:

The two state vectors of (A*, v*, p*)T and (B*, v*,
p*)T can be obtained by solving the Riemann
problem (see Appendix: Approximate Riemann
solution for general EOS), where v* is the velocity
of the contact discontinuity, that is to say, the
velocity of the material interface and p* is the
pressure at the material interface.

F MI

As the velocity of the material interface node is


known, the moving grid system is employed to
avoid the negative grid volumes near the material
interface. In this paper, the spring analogy (Blom,
2000) is applied to obtain other grids velocities so
that the whole fluids grid nodes are moving
together with the material interface. Then the
motion of the material interface is tracked at any
time. For some multi-fluid flow with large-scale or
complex deformations of material interface, the
interface rebuilding technology and the local
remeshing method are more effective, which is the
authors current work.
To solve the integral Eq. (1) on the control volume
adjacent to the material interface, the numerical flux
at the material interface must be solved. After the
Riemann problem is resolved, two different
methods to compute the numerical flux at the
material interface can be considered here.

4. RIEMANN PROBLEM AT THE


MATERIAL INTERFACE

U = ( A
U0 = A
U B = ( B

UB

S i = min u ni ci , v x& n c

4.1

UA

* v*
0
* *
* * *
*
= (F x& U ) n = v v + p n x& v n = p * n
*E
* E * v* + p* v*
p * v *

327

(18)

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 4 (2007)

analytically (Chang and Liou, 2007). Note that both


shock tube problems are computed on grids with
202 points and CFL = 0.8.

From this equation, it is known that the numerical


flux at the material interface is decided only by the
moving velocity of the interface v* and the pressure
at the interface p*, which also shows the
conservation of mass, momentum and energy near
the material interface. It is reasonable that there is
no mass exchange through the material interface
and the exchange of momentum and energy is
brought up by pressure and velocity at the interface.
y

1) Gas/gas shock tube problem


This problem is a simple common problem with no
large pressure gradient. The shock tube contains
two different kinds of gas labeled as A and B. The
initial conditions in this computation are the
following:

Method II: A virtual fluid defined at the other


side of the material interface

The interface can be considered as a special interior


boundary where a virtual fluid is defined at the
other side of each material (Fig. 2). Such as
material A, whose state vector is (A, vA, pA)T, there
is a virtual fluid A* on the other side of it. Here, the
fluid state vector of A* at the interface is equal to
(A*, v*, p*)T. Then the numerical flux at the
material interface FMIA between A and A* can be
computed by HLLC scheme (7), that is, FMIA =
F HLLC(UA, UA*). Similarly for material A, FMIB can
be obtained by B and B*. Obviously, F MIA F MIB.
However, they are approximately the same in the
numerical point of view as in the Ghost Fluid
Method (GFM) (Fedkiw et al., 1999).

A
UA

Fig. 2

A*
(A*, v*, p*)

B*
(B*, v*, p*)

uA = 0

p A = 1.0

A = 1.667

PA = 0

0 x < 0.5

B = 0.125

uB = 0
PB = 0

0.5 x 1

B = 1.200

pB = 0.1

Fig. 3(a) and (b) show the analytic and numerical


results at t = 0.2 computed by Method I and II. Both
numerical results agree well with the analytic
solution and the gas interface (x = 0.667) is sharply
captured.
2) Gas/water shock tube problem
A two-phase gas-liquid shock tube problem is
considered here, with a large pressure gradient near
the gas-liquid interface. The air in the left of the
tube is labeled as A and the water in the right is
labeled as B. The initial conditions in this
computation are the following:

B
UB

The boundary condition on the material


interface.

A = 1.271

uA = 0

p A = 9.11925 10 9

A = 1.4

PA = 0

0 x < 0.5

B = 1.0

uB = 0

p B = 1.01325 10 6

B = 7.0

P B = 3.03975 10 9

0.5 x 1

Fig. 4(a) and (b) show the computing results of


Methods I and II at t = 1.55910-4. As a whole, the
two numerical results agree well with the analytic
solution. However, in the density distribution
computed by Method I, there is obviously a small
oscillation near the gas/water interface (x = 0.543)
and Method II performs better. This phenomenon
shows that the Lagrangian method results in some
numerical oscillations at the interface and cannot
deal with material problems with a large pressure
gradient near the interface.
Although Method I seems more reasonable than
Method II and performs well enough in dealing
with the common material interface, it dose not
work well for the problem with large pressure
gradient near the interface. In this paper, Method II

Comparison between Method I and Method II

Both methods seem to be feasible and robust in the


computing of the flux at the material interface.
Here, two simple shock tube problems with free
material interface are introduced to examine the two
methods. Stiffened Gas EOS is used in both
problems:

p = ( 1)( e P ) P

A = 1.000

(19)

where is the ratio of specific heats and P is a


prescribed pressure-like constant about the material.
For all ideal gases, P = 0; for water,
P = 3.03975109 d/cm2. Analytic solutions to
shock tube problems based on Stiffened Gas EOS
can be obtained by solving the Riemann problem
328

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 4 (2007)

is applied to solve the underwater explosion


problem which has a large pressure gradient near

the interface between the explosive gas product


bubble and the ambient water.

(a) Density

Fig. 3

(b) Pressure

Computed results for gas/gas shock tube problem at t = 0.2 by Method I and II.

(a) Density

Fig. 4

(b) Pressure

Computed results for air/water shock tube problem at t = 1.55910-4 by Method I and II.

vBn

v*

vBt

v An

v At
Interface

Fig. 5

Velocitys definition in Riemann problem


at the 2D material interface.

329

Fig. 6

t +t

The motion of the 2D material


interface.

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 4 (2007)

4.2

= 1.00037984 g/cm3
p = 1.00107 d/cm2

Two dimension

The definition of the Riemann problem in 2D is


similar to that in 1D. The difference between them
is the definition of the velocitys direction. In 2D,
the Riemann problem is determined in the normal
direction of the material interface, as shown in
Fig. 5. The initial condition of the Riemann
problem in 2D at the material interface can be
written as

U = ( A
U0 = A
U B = ( B

v An
v Bn

pA )

The Tait and JWL equations of state (Fedkiw et al.,


1999) (EOS) are used to describe the water and
bubble (gaseous detonation products), respectively:
Trait:

(20)

x i
t

(22)

B = 3.31 109 d/cm


0 = 1.0 g/cm3

Then the Riemann solution based on this initial


condition is (A*, v* + vAt, p*) and (B*, v* + vBt, p*)
respectively for fluids A and B. vAt and vBt are the
distributions of different material velocities in the
tangential direction of the material interface edge.
As shown in Fig. 6, v* is the velocity of the moving
material interface edge in the normal direction and
the moving velocity of material interface point can
be obtained by

vi =


p = B 1 + A

where

pB )

u = 0 cm/s
3.5287 r 5000 cm

JWL:

A = 1.0 10 6 d/cm
= 7.15

R

p = A1
exp 1 0

R1 0

R

+ B 1
exp 2 0 + e

R2 0

(23)

where
2

(21)

where vi and xi are the moving velocity and


displacement of the material interface point i. The
numerical flux at the 2D material interface is
computed by Method II, same as the 1D multimaterial problem. At the same time, the spring
analogy is employed to obtain the velocity of other
grid points to ensure the quality of the whole grid
system.

A = 3.712 1012 d/cm


R1 = 4.15

B = 3.23 1010 d/cm


R2 = 0.95
= 0.30

0 = 1.63 g/cm3

e0 = 4.29 1010 d cm/g

The computation is performed on 352 grids in the


bubble domain and 4964 grids in the water domain.
Fig. 7(a)(c) show the evolution of the computed
solution through different phases of spherically
symmetric underwater explosion model. The initial
phase, illustrated in Fig. 7(a), starts with a primary
shock wave traveling to the right into the water and
an expansion wave moving to the left toward the
bubble origin. The expansion wave reflects from the
origin, resulting in a region of very low pressure
near the origin. The outward inertia of the
expanding gas is eventually overcome by the
centripetal pressure gradient and the gas reverses
direction, forming an inward moving shock wave
which in turn reflects as a shock wave from the
origin. This reflected secondary shock wave then
propagates outward toward the gas/water interface.
Fig. 7(b) shows the shock and free-surface
interaction phase, where the secondary shock wave
propagates to the gas/water interface and interacts
with it. This interaction generates a reflected shock
wave moving back into the bubble toward the
origin, and a transmitted shock wave traveling
outward from the bubble into the water, as showed
in Fig. 7(c). This process repeats numerous times
and each time at a reduced shock strength. At the

5. NUMERICAL EXAMPLES
5.1 1D spherically symmetric underwater
explosion model
Simulation of underwater explosion is an essential
component of platform vulnerability and weapon
lethality assessments. The computational conditions
of this model are: detonation of a 300 g of TNT
sphere (radius of 3.5287 cm) at a depth of 91.4 m
(ambient pressure of 1.0107 d/cm2, where
1 d/cm2 = 0.1 Pa). Then the initial conditions can be
given by
= 1.63 g/cm3 u = 0 cm/s p = 8.385631010 d/cm2
0 r 3.5287 cm

330

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 4 (2007)

the bubble are in reasonable agreement with the


experiment data from Andrew, Wardlaw and Hans
(1998), as showed in Table 1. It is indicated that the
ALE method illustrated here is feasible and
rigorous.

same time, the primary shock travels in water at a


reduced level (Fig. 7(d) and (e)) and the impulsive
pressure gradient at the interface results in the
impulsion of the bubble. Fig. 7(f) shows that the
radius of the bubble impulses with time. The
computed maximum radius and impulsing period of

Fig. 7

(a)

(b)

(c)

(d)

(e)

(f)

(a)(e): Pressure distributions during 1D spherically symmetric underwater explosion;


(f): Bubbles impulsing with time.

331

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 4 (2007)

Axis

Water tank

Table 1 Comparison of results of computation and


experiment on bubble pulse.

3.5287 cm

23 cm
TNT
8 cm

15 cm

Fig. 8

Experiment
Computation
Error

Period (ms)

Maximum radius (cm)

29.8
29.92
0.4027%

48.1
48.9662
1.80%

Initial physical conditions for 2D cylindrically


symmetric underwater explosion model.

(a) t = 0.86403E-5 s

(b) t = 0.25200E-4 s

(c) t = 0.45615E-4 s

(d) t = 0.58542E-4 s

(e) t = 0.86561E-4 s

(f) t = 0.96446E-4 s

Fig. 9

Pressure contours at different times in 2D cylindrical symmetric underwater explosion.


332

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 4 (2007)

reduced strength. The primary shock wave reflects


at the top and wall of the tank and then generates
two reflected shock waves moving toward the
bubble (Fig. 9(d)). Then the reflected shock waves
interacts with the bubble interface (Fig. 9(e)) and
results in a complex set of contours (Fig. 9(f)). The
interactions between shock waves and the bubble
interface make the bubble unsymmetrical. In this
computation, there are no spurious oscillations at
the bubble interface, which clearly indicates that the
ALE method has a satisfactory performance in the
simulation of 2D cylindrically symmetric
underwater explosion model.

5.2 2D cylindrical symmetric underwater


explosion model
We examine the ALE method in a 2D cylindrically
symmetric underwater explosion problem. The
initial physical conditions of this problem are
described by Fig. 8. A 300 g of TNT sphere denotes
in water tank which is at a depth of 91.4 m in the
water (ambient pressure is 1.0107 d/cm2). The
same JWL and Tait equations of state with the 1D
spherically symmetric underwater explosion model
are used for the gaseous detonation products and the
water. Fig. 9(a)(f) are the pressure contours at
different times in the water tank with 4748 grids in
the bubble domain and 73892 grids in water
domain. The black bold curve stands for the bubble
interface. Fig. 10 shows the grids of the whole
computational domain at the moment of 0.964465 s.

5.3 Interaction between aerofoil and water


surface
This model is designed to examine the performance
of the ALE method in the micro-deforming
interface problem. We simplify the aerofoil as an
ellipse of diameters 3 by 5 m. The aerofoil travels
with a Mach number of 1.47 at an altitude of 5 m
above a flat body of water. The air is regarded as an
ideal gas at 15 C. In this model, the perfect gas
equation of state is used for air, that is,

p = ( 1) e

(24)

where is the ratio of specific heats. And the Tait


equations of state are used for the water. The
computational domain contains 80484 grids and the
pressure contours are displayed in Fig. 11. Good
resolution of shock waves and reflected shock
waves is clearly showed. Fig. 12 shows the shape of
air/water interface, where the micro-deformation is
well resolved. It is clearly difficult for other
capturing method to capture such small
deformation. So ALE method is more advantageous
in solving micro-deforming interface problems.

Fig. 10 The grids of the water tank domain at


t = 0.96446 s.

Fig. 9(a) shows that a primary shock wave travels


into the water and an expansion wave moves in the
bubble and toward the bubble origin. The primary
shock wave then hits the bottom of the tank, which
generates a reflected shock wave moving toward
the gas/water interface (Fig. 9(b)). After the
interaction with the interface, this reflected shock
wave is diffracted into two waves: a reflected
rarefaction wave moving into the water and a
transmitted shock wave in the bubble (Fig. 9(c)).
The rarefaction wave then hits the bottom of the
tank and generates a second-reflected rarefaction
wave moving toward the bubble interface with a

Fig. 11 Pressure contours near the airfoil.

333

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 4 (2007)

3. Nkonga B (2000). On the conservative and


accurate CFD approximations for moving
meshes and moving boundaries. Computer
Methods
in
Applied
Mechanics
And
Engineering 190(13):18011825.
4. Chang CH and Liou MS (2007). A robust and
accurate approach to computing compressible
multiphase flow: stratified flow model and
AUSM+-up scheme. Journal of Computational
Physics 225(1):840873.
5. Cocchi JP and Saurel R (1997). A Riemann
problem based method for the resolution of
compressible multimaterial flows. Journal of
Computational Physics 137(2):265298.
6. Toro EF, Spruce M and Speares W (1994).
Restoration of the contact surface in the HLLRiemann solver. Shock Waves 4(1):2534.
7. Fedkiw RP, Aslam T, Merriman B and Osher S
(1999). A non-oscillatory Eulerian approach to
interfaces in multimaterial flows (the ghost
fluid method). Journal of Computational
Physics 152(2):457492.
8. Fedkiw RP, Marquina A and Merriman B
(1999). An isobaric fix for the overheating
problem in multimaterial compressible flows.
Journal of Computational Physics 148(2):545
578.
9. Blom FJ (2000). Considerations on the spring
analogy. International Journal for Numerical
Methods in Fluids 32(6):647668.
10. Hankin RKS (2001). The Euler equations for
multiphase compressible flow in conservation
form: simulation of shock-bubble interactions.
Journal of Computational Physics 172(2):808
826.
11. Hirt WC, Amsden AA and Cook LJ (1974). An
arbitrary
LagrangianEulerian
computing
method for all flow speeds. Journal of
Computational Physics 14(3):227253.
12. Koren B, Lewis MR, van Brummelen EH and
van Leer B (2002). Riemann-problem and
level-set approaches for homentropic two-fluid
flow computations. Journal of Computational
Physics 181(2):654674.
13. Liu TG, Khoo BC and Yeo KS (2001a). The
simulation of compressible multi-medium flow.
Part I: A new methodology with test
applications to 1D gasgas and gaswater
cases. Computers & Fluids 30(3):291314.

Fig. 12 The micro-deformation of the water surface.

6. CONCLUSIONS
The material interface is tracked by solving ALE
formulations on the compressible multi-material
flow. The material interface is considered as a
Lagrangian interface which is composed by a
number of edges of the unstructured grids and state
vectors of the points on the interface have two
definitions according to the two different fluids.
Then, Riemann problem is solved on both side of
the interface to track the interfaces movement
accurately and the grids are moving automatically
with the motion of the interface.
1D spherically symmetric underwater explosion
module is computed using ALE method and the
numerical results agree well with the experimental
data, which indicates that the interface tracking
method is reasonable. Furthermore, interaction
between shock and water surface is also simulated
to show that this method is suitable for solving
micro-deforming interface problems.
ACKNOWLEDGEMENTS
The work described in this paper is supported by
the National Natural Science Foundation of China
(Project No. 10476011).
REFERENCES
1. Abgrall R (2003). Efficient numerical
approximation of compressible multi-material
flow for unstructured meshes. Computers &
Fluids 32(4):571605.
2. Andrew B, Wardlaw Jr and Hans UM (1998).
Spherical solutions of an underwater explosion
bubble. Shock and Vibration 5(2):89102.
334

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 4 (2007)

product gas in the underwater explosion model


where the JWL EOS must be used. And in many
situations Tait EOS for water is much more
vigorous than stiffened EOS. At the same time, no
exact Riemann solutions for either JWL EOS or
Tait EOS exists in CFD domains because their
forms are non-linear and complex.
General EOS is used to describe any material in this
paper, similar in form to Mie-Grneisen EOS
(Shyue, 2001). The general EOS is written by

14. Liu TG, Khoo BC and Yeo KS (2001b). The


simulation of compressible multi-medium flow.
Part II: Applications to 2D underwater shock
refraction. Computers & Fluids 30(3):315337.
15. Nourgaliev RR, Dinh TN and Theofanous TG
(2006).
Adaptive
characteristics-based
matching for compressible multifluid dynamics.
Journal of Computational Physics 213(2):500
529.
16. Thomas PD and Lombard CK (1978).
Geometric conservation law and its application
to flow computations on moving grids. AIAA
Journal 17(10):10301037.
17. Colella P (1985). A direct Eulerian MUSCL
scheme for gas dynamics. SIAM Journal on
Scientific and Statistical Computing 6(1):104
117.
18. Wang SP, Anderson MH, Oakley JG, Corradini
ML
and
Bonazza
R
(2004).
A
thermodynamically consistent and fully
conservative
treatment
of
contact
discontinuities
for
compressible
multicomponent
flows.
Journal
of
Computational Physics 195(2):528559.
19. Shyue K (2001). A fluid-mixture type algorithm
for compressible multicomponent flow with
Mie-Grneisen equation of state. Journal of
Computational Physics 171(2):678707.
20. Tang H, Wrobel LC and Fan Z (2004).
Tracking of immiscible interfaces in multiplematerial mixing processes. Computational
Materials Science 29(1):103118.
21. Xiao F and Ebisuzaki T (1998). An efficient
numerical model for multi-phase fluid
dynamics. Advances in Engineering Software
29(36):345352.
22. Xu K (1997). BGK-based scheme for
multicomponent flow calculations. Journal of
Computational Physics 134(1):122133.

e = f () p + g ()

(A1)

where f() and g() are general functions of density


. The sound speed c can be obtained by the
following formulation:
pf + g p(1 + f ) + g
+
f
f

c2 =

and f =

(A2)

f
g
, g =
.

For this general EOS, the two-shock approximation


method is applied to solve the Riemann problem,
whose main idea is to consider rarefaction as a
simple shock jump during iteration schemes. This
method is efficient for general EOS with only a
small loss of accuracy. Consider the following
initial conditions in the Riemann problem:

(
(

uL
uR

pL
pR

)
)

x<0

x>0

(A3)

The Riemann resolution to be resolved is defined by


(L*, R*, u*, p*)T. Then two sets of energy state
formulations can be obtained from the Hugoniot
relations:
W L2 =

APPENDIX

WR2 =

APPROXIMATE RIEMANN SOLUTION FOR


GENERAL EOS

p*2 p L2
2 e*L e L

, W L2 = *L L

p* p L
*L L

p*2 p R2
p* p R
, WR2 = *R R *
R R
2 e*R eR

(A4)

(A5)

where W is the Lagrangian wave speed and e is the


specific internal energy, defined by Eq. (A1). Then
Eqs. (A4) and (A5) can be simplified to two sets of
equations:

The stiffened EOS is applied in quite a lot of multifluid computations as the exact Riemann solution
exists (Chang and Liou, 2007). It is certainly
efficient in tracking the material interface.
However, stiffened EOS cannot describe all kinds
of material accurately enough, such as the explosion

( )

f1 x, y = e x, p* e x K , p K

) y 12 p

*2

p K2 = 0

(A6)
335

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 4 (2007)

1 1
f 2 x, y =
y p* p K = 0
xK x

( )

where

(A7)

and
p* =

WL p R + WR p L + WLWR (u L u R )
WL + W R

(A8)

where
y = WK2

x = *K

(A9)

K = L, R

Newton iteration scheme is used to solve Eqs. (A6)


and (A7), to obtain the steady results of x and y.
That is,

n +1

1
=x n
J

f1 x n , y n

f 2 x n , y n

f1 x n , y n

1
n +1
n
x

y
=y n
J f x n , y n

x
f1 x n , y n

x
n
J =
f 2 x n , y n

f1 x n , y n

y
f 2 x n , y n

y =

yn < q

yn + 1 yn
yn + 1 yn
yn

xn q

yn q

W u + WR u R (p R p L )
u* = L L
WL + WR

f1 x n , y n

(A11)

f 2 x n , y n

f1 x n , y n

y
f 2 x n , y n

(A12)

where

f1 (x, y )
e
= y
x
x

f1(x, y)
= e x, p* e xK , p K (A13)
y

f 2 (x, y ) y
= 2
x
x

f 2 (x, y ) 1 1
=

xK x
y

(A14)

The error condition must be satisfied before the


iteration is finished. The error condition can be
written as
max x , y <

xn < q

(A16)

is the submitting error and q is a control constant


to choose for relative error or absolute error. In the
authors computer codes, = 10-5 and q = 10-3.
When the iteration is finished, most of the
approximate Riemann resolutions, that is, L*, R*
and p* are obtained. Then the velocity of the contact
u* can be known by

(A10)

xn + 1 xn

x = xn + 1 x n

xn

(A15)

336

(A17)

You might also like