You are on page 1of 28

Accepted Manuscript

A Riemann problem based method for solving compressible and incompressible flows

Haitian Lu, Jun Zhu, Chunwu Wang, Donghong Wang, Ning Zhao

PII:
DOI:
Reference:

S0021-9991(16)30555-1
http://dx.doi.org/10.1016/j.jcp.2016.10.047
YJCPH 6925

To appear in:

Journal of Computational Physics

Received date:
Revised date:
Accepted date:

27 July 2016
23 September 2016
21 October 2016

Please cite this article in press as: H. Lu et al., A Riemann problem based method for solving compressible and incompressible flows, J.
Comput. Phys. (2016), http://dx.doi.org/10.1016/j.jcp.2016.10.047

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing
this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is
published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.

Highlights

Two-medium flow including compressible and incompressible regions is investigated.


An incompressible discontinuous Galerkin method with nonuniform time step is derived.
A coupled compressible and incompressible Riemann problem is proposed.
Reasonable results are obtained for strong shock wave formed in compressible flow.

A Riemann problem based method for solving compressible and


incompressible flows
Haitian Lu1, Jun Zhu2, Chunwu Wang2, Donghong Wang2, Ning Zhao1,
1 College

of Aerospace Engineering, Nanjing University of Aeronautics and Astronautics, Nanjing, Jiangsu 210016,

P.R. China.
2

College of Science, Nanjing University of Aeronautics and Astronautics, Nanjing, Jiangsu 210016, P.R. China.

Abstract. A Riemann problem based method for solving two-medium flow including compressible and
incompressible regions is presented. The material interface is advanced by front tracking method and the material
interface boundary conditions are defined by modified ghost fluid method. A coupled compressible and
incompressible Riemann problem constructed in the normal direction of the material interface is proposed to
predict the interfacial states. With the ghost fluid states, the compressible and incompressible flows are solved by
discontinuous Galerkin method. An incompressible discontinuous Galerkin method with nonuniform time step is
also deduced. For shock wave formed in compressible flow, the numerical errors for the ghost fluid method in
earlier works are analyzed and discussed in the numerical examples. It shows that the proposed method can
provide reasonable results including shock wave location.
Key words: modified ghost fluid method, compressible and incompressible Riemann problem, discontinuous
Galerkin method, front tracking method.
AMS(MOS) subject classification: 65M60, 76T10, 76T99

1. Introduction
The interaction of liquid droplet with gas medium is difficult in the numerical simulation due
to the large material properties differences. In many earlier works, both the gas and liquid
mediums are treated as compressible and a stiffened gas equation of state (EOS) [34] is used for
the liquid medium. However, a complete compressible treatment is still limited because of the
differences in the speed of sound and the EOS across the material interface. Moreover, only a
small time step is permitted by the restriction of the CFL condition for the liquid medium, which
makes the methods less efficient. When both mediums are modeled as incompressible flow [17, 23,
38], it is unable to simulate the high speed flow problem with shock wave and therefore this model
is more applicable for computing the lower speed flow problem, for example, a bubble rising in
the water or a droplet falling in the gas. For the approaches in which the gas is modeled as
compressible medium and the liquid is modeled as incompressible medium, the material properties
are well disposed and the high speed flow problem with shock wave in the gas medium can be
resolved, providing the material interface is treated precisely.
Generally speaking, the difficulties for simulating two-medium flow are to solve the single
medium flow accurately and treat the moving material interface exactly. In recent years, the
discontinuous Galerkin (DG) method has been a research hotspot in the simulations of single
medium flow. The Runge-Kutta discontinuous Galerkin (RKDG) method proposed by Cockburn

Corresponding author. Email address: lhtgkzy@126.com (H.T. Lu), zhujun@nuaa.edu.cn (J. Zhu),
wangcw@nuaa.edu.cn (C.W. Wang), wangdonghong@nuaa.edu.cn (D.H. Wang), nzhao2000@ hotmail.com (N.
Zhao).

et al. [9-13] performs very well in solving the compressible flow. It employs the total variation
diminishing (TVD) high order Runge-Kutta time discretizations [37] and DG discretization in
space with exact or approximate Riemann solvers as interface fluxes and a total variation bounded
(TVB) limiter [36] to achieve non-oscillatory behavior for strong shock waves. In contrast, only a
few works of the DG method are related to the numerical simulation of the incompressible flow.
Liu et al. [26] introduced a high order DG method for two-dimensional incompressible flow in the
vorticity-function formulation, where the momentum equation is treated explicitly by the DG
method and the stream function is solved by a standard Possion solver with continuous finite
elements. Thereafter, Cockburn et al. [6-8] proposed the local discontinuous Galerkin (LDG)
method for the Stokes, Oseen and incompressible Navier-Stokes (INS) equations. Ferrer et al. [16]
presented a simpler and more efficient DG method for the unsteady INS equations where the
Interior Penalty (IP) spatial discretization is used for the elliptic equations. More related works can
also be found in [4, 22, 33, 35].
As for the treatment of material interface, early algorithms [1, 24, 31] based on classical shock
capturing methods usually yield a smeared contact discontinuity over several spatial nodes due to
excessive numerical diffusions. However, for the front tracking method [19-21], the moving
material interfaces are explicitly tracked and a sharp interface boundary is sustained during the
computation. In this method, a Riemann problem is constructed nearby the material interface for
the sake of handling the sharp jump across the interface. The original ghost fluid method (GFM)
[15] presented an efficient way to treat the material interface. However, when the pressure or the
velocity experiences a large gradient across the interface, the GFM may cause some numerical
inaccuracy. Fedkiw et al. [14], therefore, proposed a new version of GFM (new GFM) which
overcomes the difficulty encountered by the original GFM for the gas-water flow. Caiden et al. [5]
extended the new GFM into coupling compressible and incompressible flows, and this approach
performs well in the numerical simulations. The new GFM defines the interface boundary
conditions where the ghost fluid velocity relies on the liquid medium while the ghost fluid
pressure depends on the gas medium. The new GFM is very simple and the extensions to higher
dimensions are straightforward. Nevertheless, the interface boundary conditions should be
nonlinearly affected by the material properties on both sides of the material interface. This
situation indicates that the definition of the ghost fluid states by simply taking from the real fluid
is not rigorous to confirm such effect. Moreover, the interfacial states should be predicted to
consider the influence of wave interaction in the interfacial evolution. As one of the improved
versions of GFM, for example, the modified ghost fluid method (MGFM) [28] predicts the
interfacial states by solving a Riemann problem approximately near the material interface. The
predicted interfacial states are then used to define the ghost fluid states. The MGFM is effective
and has shown to be much less problem related for the compressible two-medium flow.
If the two-medium flow consists of compressible and incompressible parts, more restrictive
interface boundary conditions should be defined to consider the influences of material properties

and wave interaction near the interface. The motivation of present work is to analyze the
numerical errors for new GFM [5] when shock wave is formed in compressible flow, and
thereafter to propose a Riemann problem based method to predict the interfacial states. The
predicted interfacial states are used to advance the material interface by the front tracking method
and obtain the ghost fluid states directly. With the interface boundary conditions, both
compressible and incompressible flows are resolved separately by the DG method. The
incompressible DG method with nonuniform time step is derived here since the time step is
generally not constant in the numerical simulation of two-medium flow. Extensive numerical
examples are tested and compared with the new GFM [5]. It is found that the proposed method is
robust and can provide convincing results.
The organization of this paper is as follows: in Section 2, we describe the RKDG method for
the compressible flow and derive the incompressible DG method with nonuniform time step. In
Section 3, we illustrate the algorithms of advancing the material interface and construct a coupled
compressible and incompressible Riemann problem at the interface. The causes of numerical
errors for new GFM are analyzed when shock wave is formed in compressible flow. Extensive
numerical tests are presented in Section 4 to show good performances of the proposed method.
Concluding remarks are given in Section 5.
2. Equations and numerical method
2.1. Compressible flow
The Euler equations for two-dimensional compressible flow can be written as follows:
U
+ F ( U ) = 0,
t

(1)

where U = [ , u, v, E]T , F ( U ) = [F1 (U ), F2 (U )], F1 (U) = [ u, u 2 + p, uv,( E + p)u]T , F2 (U) = [

v, uv, v 2 + p, ( E + p )v]T . Here is the density, u and v are the velocities which can be denoted
by the velocity vector V = [u,v ]T , p is the pressure, E is the total energy per unit volume:

E = e + (u 2 + v2 ) / 2,

(2)

where e is the internal energy per unit mass. The EOS for the ideal gas is used:

p = ( 1) e,

(3)

where represents the ratio of the specific heats.


2.2. Incompressible flow
The equations for two-dimensional incompressible flow are the INS equations:
V
= V V p + 2 V ,
t

(4)

V = 0,

(5)

where p = p / , = / and is the viscosity. Both and are assumed to be constant in the

incompressible region. The solutions for the INS equations are denoted by W = [u, v, p]T .
2.3. Adaptive time step
Adaptive time step is used where the overall time step is the minimum of the compressible
and incompressible time steps:

t = min(tc , ti ).

(6)

For compressible flow, the time step defined as:


tc =

Cf
((| u | + a ) / x + (| v | + a ) / y )

(7)

needs to be satisfied at every grid cell, where C f is the CFL coefficient, a is the speed of sound
and x and y are the cell sizes. For the incompressible flow, the time step is [35]:
ti O (

L
),
U ki2

(8)

where L is an integral scale (typically the cell size), U is a characteristic velocity and ki is the
polynomial degree of the DG method for solving the incompressible flow.
2.4. Implementation of DG method
2.4.1. RKDG method for Euler equations
In this section we focus on the description of the two-dimensional RKDG method [9-13] for
solving the compressible flow. Consider the rectangular cell I i , j = [ xi 1/ 2 , xi +1/ 2 ] [ y j 1/ 2 , y j +1/ 2 ]
with the cell center

( xi , y j ) , we denote the cell size by

x = xi +1/ 2 xi 1/ 2

and

y = y j +1/2 y j 1/ 2 . The numerical solutions U h ( x, y, t ) as well as the basis functions are given
by Vhkc = { p : p |Ii , j P kc ( I i , j )}, where P kc ( I i , j ) is the space of polynomials with degrees no
greater than kc on the cell Ii, j. The local orthogonal basis functions over Ii, j are adopted as:

i0, j ( x, y ) = 1,
i1, j ( x, y ) =
i2, j ( x, y ) =

x xi
,
x / 2
y yj
y / 2

i3, j ( x, y ) = i1, j ( x, y )i2, j ( x, y ),


1
3
1
i5, j ( x, y ) = i2, j ( x, y )i2, j ( x, y ) ,
3


i4, j ( x, y ) = i1, j ( x, y )i1, j ( x, y ) ,

The numerical solutions Uh(x, y, t) can be written as:

(9)

Uh ( x, y, t ) = Ui(,l )j (t )il, j ( x, y), for ( x, y) Ii , j ,

(10)

l =0

and Ui(,l )j (t ) are the moments defined by:

Ui(,l )j (t ) =

1
al

Ii , j

Uh ( x, y, t )il, j ( x, y)dxdy, l = 0,1,, L,

(11)

l
2
where al = (i , j ( x, y)) dxdy are the normalization constants. By substituting Eq. (10) into Eq.
Ii , j

(1), and multiplying Eq. (1) with the basis function, integrating over the cell Ii, j by parts, we can
obtain the governing equations for the moments as:

dUi(,l )j (t)
dt

1
( F(Uh (x, y, t)) ne,Ii, j il, j (x, y)d
al eIi, j e

(12)

+ F(U (x, y, t)) grad (x, y)dxdy), l = 0,1, , L ,


h

l
i, j

Ii , j

where ne , I i , j is the outward unit normal to edge e of cell Ii, j. The integrals are approximated by a
Gaussian quadrature formula with sufficient accuracy:

F(U ( x, y, t )) n
h

( x, y)d s F(Uh ( xes , yes , t )) ne, I il, j ( xes , yes ) e ,

l
e, Ii , j i , j

i, j

s =1

(13)

Ii , j

F (U h ( x, y , t )) gradil, j ( x, y ) dxdy r F ( U h ( xIi , j r , y Ii , j r , t )) gradil, j ( xIi , j r , y Ii , j r ) I i , j . (14)


r =1

The flux F ( U h ( xes , yes , t )) ne , Ii , j in this paper is replaced by the simple Lax-Friedrichs flux:

1
F(U h ( xes , yes , t )) ne, Ii , j [(F (U + ( xes , yes , t )) + F (U ( xes , yes , t ))) ne, Ii , j
2
(U + ( xes , yes , t ) U ( xes , yes , t ))],

(15)

where is the biggest absolute value of the eigenvalues of the Jacobian matrix in the ne, Ii , j
direction, and U ( xes , yes , t ) refers to the value of U h in the neighboring cell and in the current
cell Ii, j, respectively.
The semi-discrete equations (Eq. (12)) can be generalized as:
U t = L( U ),

(16)

which is discretized in time by a third-order TVD Runge-Kutta discretization method in [37]:


U (1) = U n + tL(U n ),
3 n 1 (1) 1
U + U + tL(U (1) ),
4
4
4
1 n 2 (2) 2
= U + U + tL(U (2) ).
3
3
3

U (2) =
U n +1

(17)

If there are strong discontinuities in the solutions, spurious oscillations may occur. Thus a local
slope limiter is used to deal with such difficulties. For a scalar equation, we denote:
x

Ui+1/ 2, j = Ui(0)
Ui+1/ 2, j = Ui(0)
, j (t ) + Ui , j ,
, j (t ) Ui , j ,

Ui, j +1/ 2 = Ui(0)


Ui+, j 1/ 2 = Ui(0)
, j (t ) + Ui , j ,
, j (t ) Ui , j ,
where:
L

Ui , j = U i(,l )j (t )il, j ( xi +1/ 2 , y j ), U i , j = U i(,l )j (t )il, j ( xi 1/ 2 , y j ),


x

l =1

y
i, j

l =1

= U (t ) ( xi , y j +1/ 2 ), U i , j = U i(,l )j (t )il, j ( xi , y j 1/ 2 ).


(l )
i, j

l
i, j

l =1

(18)

l =1

x (mod)

(0)
(0)
(0)
= m(U i , j , U i(0)
+1, j (t ) U i , j (t ), U i , j (t ) U i 1, j (t )),

Then we can modify U i , j , U i , j , U i , j , U i , j by:


Ui , j

x (mod)

Ui , j

y (mod)

Ui , j

y (mod)

Ui , j

x
x

(0)
(0)
(0)
= m(U i , j , U i(0)
+1, j (t ) U i , j (t ), U i , j (t ) U i 1, j (t )),
y

(0)
(0)
(0)
= m(U i , j , U i(0)
, j +1 (t ) U i , j (t ), U i , j (t ) U i , j 1 (t )),

(19)

(0)
(0)
(0)
= m(U i , j , U i(0)
, j +1 (t ) U i , j (t ), U i , j (t ) U i , j 1 (t )),

where m is a modified minmod function:

if | a1 | M x 2 ,
a1 ,

m(a1 , , am ) = s min i | ai |, if s = sign(a1 ) =  = sign(am ),


0,
otherwise,

y (mod)

where M is the TVB limiter constant. For U i , j

y (mod)

and U i , j

(20)

, x is switched to y.

For system cases, the limiter is always used with a local characteristic field decomposition
[13].
2.4.2. DG method for INS equations
Let be the incompressible region with boundaries that can be of Dirichlet type ( D )
or Neumann type ( N ). We introduce rectangular tessellation h = {K} of with external
boundaries h and interior boundaries h. The numerical solutions W h ( x, y, t ) and the basis
k
functions are given by Dhki (h ) = { p : p |K P ki ( K )}, where P i ( K ) is the space of polynomials

with degrees no greater than ki on the cell K. For each cell K=Ii,j, the numerical solutions
W h ( x, y, t ) can be written as:
M

Wh ( x, y, t ) = Wi(,mj ) (t )im, j ( x, y), for ( x, y) Ii , j ,

(21)

m =0

where Wi(, mj ) (t ) are the moments and the basis functions are those defined in Eq. (9).
In [16], an incompressible DG method with uniform time step is proposed for solving single
medium flow. However, for the simulation of two-medium flow, the overall time step (see Eq. (6))
should depend on both compressible and incompressible regions, therefore an incompressible DG
method with nonuniform time step is required and derived as follows. The time derivatives in the
INS equations are discretized by a second order backwards-differentitation:

V n +1
) 0 V n +1 0 V n 1V n 1 ,
t

(22)

where 0 = 0 + 1 is required for consistency, the superscript "n-1", "n" and "n+1" denote the
time level tn-1, tn and tn+1, respectively. The parameters 0, 0, 1 are given by:

0 =

2t n + t n 1
,
t n ( t n 1 + t n )

(23)

t n + t n 1
,
t n t n 1

(24)

0 =

1 =

n 1

t n
,
( t n 1 + t n )

(25)

where t n 1 = t n t n 1 and t n = t n +1 t n . Then Eq. (22) can be written as:


(

V n +1
) = 0 ( V n +1 V n ) + 1 ( V n +1 V n 1 )
t
t n +1 V
t n +1 V
dt + 1 n1
dt
= 0 n
t
t
t
t
t n +1

t n+1

(26)

= 0 n fdt + 1 n1 fdt ,

where f is the right hand side of Eq. (4). The nonlinear terms are integrated explicitly while
viscous and pressure terms are treated implicitly [25]. Following this approach, we obtain the
resulting temporal discretization for Eq. (4) based on nonuniform time step:
0 V n + 1 0 V n 1 V n 1 = 0 ( V V ) n 1 ( V V ) n 1 p

n +1

+ 2 V n +1 ,

(27)

where:

0 = 0 t n =

t n + t n 1
,
t n 1

1 = 1 ( t n 1 + t n ) =

t n
.
t n 1

(28)
(29)

It is found that Eq. (27) leads to the temporal discretization in [16] with uniform time step if
tn=tn-1. By introducing the intermediate velocities V and V , Eq. (27) is split into three
distinct steps: an explicit nonlinear advection, an implicit pressure solver and an implicit velocity
correction. Then Eq. (27) can be rewritten as:
0 V 0 V n 1 V n 1 = 0 ( V V ) n 1 ( V V ) n 1 ,
0 ( V V ) = p

n +1

0 ( V n +1 V ) = 2 V n +1 .

(30)
(31)
(32)

Taking the divergence of Eq. (31) and imposing V = 0 , we obtain:


2 p

Solve this pressure Poisson equation for p


velocity Vn+1 by rearranging the Eq. (32) as:

n +1

n +1

= 0 V .

(33)

, update the immediate velocity V to obtain the

2 V n +1 +

0 n +1 0
V = V.

(34)

The symmetric interior penalty Galerkin (SIPG) method is used to spatially discretize elliptic Eq.
(33) and Eq. (34). The details of the derivation and implementation are omitted here but can be
found in [18, 33]. The general formulation is introduced as:
2 q + q = g ,
q = LD , on D ,

(35)

q n = LN , on N ,

where q H 1 () is a scalar solution (but the extension to the vector formulation is obvious),
represents the wave number, n is the outward unit normal on , g L2 () is the forcing term,

LD H 1/ 2 ( D ) and LN L2 ( N ) represent the boundary conditions. Note that =0 is used for


Eq. (33). The weak solution qh of Eq. (35) is obtained by requiring:
a ( q h , ) = l ( ), Dhki (h ),
a(q h , ) =

K h

(q h + q h ) dxdy

({{q }} n [[ ]] + {{}} n [[q


h

e h D

l ( ) =

K h

(36)

( g ) dxdy

( n

e D

]] [[q h ]][[ ]])ds ,

) LD ds +

e N

ds ,

(37)

(38)

across the
where ne is the outward unit normal to the edge e, the jump [[]] and average {{}}
interface between the neighboring cells K1 and K2 are defined as:
[[]] = ( |K1 ) ( |K2 ), {{}}
= (( |K1 ) + (( |K2 )) / 2, e = K1 K 2 ,

(39)

and at boundary edges we define:


[[]] = {{}}
= ( |K1 ), e = K1 h ,

(40)

The penalty parameter is used [35]:


=

( k i + 1)( k i + 2)
max( K ),
K
2

(41)

where K is the ratio of the perimeter and the area of the cell K.
3. The interface treatment
3.1. The interface tracking technique
As indicated in Fig. 1, the compressible flow (medium 1) and incompressible flow (medium 2)
are separated by the fluid interface at time tn. The marker points are intersections of the interface
and the grid lines. Take the marker point E ( xE , yE ) as an example. Point A( xA , yA ) and point

B( xB , yB ) are in different mediums, obtained by the same distance n [21] from the marker point
E:

N 2 N Ey 2
n = Ex +

x y

1
2

x A = xE n N Ex , y A = yE n N Ey ,
xB = xE + n N Ex , yB = yE + n N Ey ,

(42)

(43)

&

where x and y are the cell sizes, and N E = ( N Ex , N Ey ) is the unit normal vector of the marker
point E. The state vectors at point A and point B can be solved directly from the numerical
solutions of the DG method:
L

U A = U(Kl )A (t )Kl A ( xA , y A ),
l =0
M

(44)

WB = W

( m)
KB

m =0

(t ) ( xB , yB ),
m
KB

where KA and KB are the grid cells containing point A and point B. The density, the normal velocity
and the pressure are calculated and denoted by Q A = [ A , uNA , p A ]T and QB = [ B , uNB , p B ]T . A
coupled compressible and incompressible Riemann problem is defined in the normal direction of
the marker point E with the initial conditions:
Q ,
QE = A
Q B .

(45)

This Riemann problem is solved (see Section 3.3) to obtain the predicted interfacial states denoted
by I E = [ LI , RI , u I , p I ]T , where the superscript "I" refers to the interface, and the subscript "L"
and "R" represent the left and right sides of the interface, respectively. Since the incompressible
flow has stiffer property than the compressible flow, the tangential velocity of the marker point
needs to rely on the incompressible flow. In this paper, the tangential velocity of the marker point
E is defined as v I = vTB , where vTB is the tangential velocity of point B. In this way, the marker
point E is advanced by the velocity v E = (u I , v I ) . It is similar for other marker points.

NE

Medium 2

E
D

G
Interface

Medium 1

Fig. 1. Construct the Riemann problem at the marker point.

3.2. The new GFM for shock wave formed in compressible flow
Consider the one-dimensional two-medium flow problem including compressible and
incompressible regions. Physically, the pressure and velocity recover continuity across the
interface instantly in the beginning. However, it may take a few or many computational steps to

reach this state numerically. It is within these steps that the numerical errors are introduced [28].
In the new GFM, the interfacial pressure is determined solely from the compressible flow and
is then used as a Dirichlet boundary condition in solving pressure in the incompressible region.
When a shock wave is formed in the compressible region, the interfacial pressure should be
increased regardless of the initial pressure in the compressible flow. In the first time step, the
interfacial pressure is taken from the initial compressible flow and correct interfacial pressure is
recovered incrementally in the subsequent time steps. If the shock wave formed is weak enough
and the number of time steps needed to reach correct interfacial pressure is not excessive,
acceptable results can still be expected. However, if a strong shock wave is formed in the
compressible region, much more time steps are required for the interfacial pressure to reach
correct magnitude. Therefore, the shock speed is affected in these time steps and a discrepancy of
shock wave location may be observed in the numerical solutions. It should be noted that the
material properties and wave interaction near the interface are very significant in solving
two-medium flow. Consequently, reasonable interfacial states have to be predicted to consider the
wave pattern in the compressible region and incompressible flow properties.
3.3. Compressible and incompressible Riemann problem
In the MGFM [28], the mediums are all compressible and a Riemann problem is constructed
at the interface to predict the interfacial states. However, when the liquid is treated as
incompressible flow where the density is constant, a coupled compressible and incompressible
Riemann problem should be constructed. As is known, the solutions for the compressible flow
tend to the solutions of the corresponding incompressible flow, as the Mach number tends to zero
[2, 3]. Based on this conclusion, we treat the incompressible flow as the compressible flow with an
infinite speed of sound and construct a coupled compressible and incompressible Riemann
problem for the definition of the interface boundary conditions.
When all mediums are treated as compressible, the EOS can be written uniformly as [34]:

p = ( 1) e B,

(46)

where B is the material parameter. For the ideal gas, B is zero and this equation reduces to the Eq.
(3). Consider the Riemann problem with the initial conditions:
Q ,
Q0 = L
Q R ,

(47)

where QL = [ L , uL , pL , L , BL ]T , QR = [ R , uR , pR , R , BR ]T . This Riemann problem is firstly


resolved by an approximate Riemann solver (ARPS) based on a two shock structure [27]:

LI =

L ( pL + BL ) + L ( p I + BL ) + p I pL
,
L ( pL + BL ) + L ( p I + BL ) + pL p I L

(48)

RI =

R ( pR + BR ) + R ( p I + BR ) + p I pR
R ,
R ( pR + BR ) + R ( p I + BR ) + pR p I

(49)

p I p L p I pR
+
+ u R u L = 0,
WL
WR

(50)

where:
WL =

L LI ( p I pL )
R RI ( p I pR )
, WR =
.
I
L L
RI R

(51)

Also consider the solution for uI in [39]:


uI =

1
1
(u L + u R ) + ( f R ( p I , Q R ) f L ( p I , Q L )),
2
2

(52)

where:

fL ( pI , QL ) =

fR ( pI , QR ) =

2( p I pL )

L [ L ( pL + BL ) + L ( p I + BL ) + p I pL ]
2( p I pR )

R [ R ( pR + BR ) + R ( p I + BR ) + p I pR ]

(53)

(54)

Without loss of generality, we assume the left side is compressible flow and the right side is
incompressible flow. As mentioned above, the incompressible flow is treated as compressible flow
and the speed of sound expressed as:
a R = k R / R = R ( p R + BR ) / R ,

(55)

is assumed to be infinite. However, R is a finite value, therefore:


k R = R ( pR + BR ) +.

(56)

Substituting Eq. (56) into Eq. (49) yields:

lim RI = R ,

k R +

(57)

then Eq. (51), (54), (50) and (52) reduce to:

lim WR = +,

(58)

f R ( p I , QR ) = 0,

(59)

p I pL
+ uR u L = 0,
WL

(60)

1
1
(u L + u R ) f L ( p I , Q L ).
2
2

(61)

k R +

uI =

Consider the relation in [39] for the left side:

u I = uL f L ( p I , QL ),

(62)

u I = uR .

(63)

we have:

The iteration is used to solve for pI in Eq. (60) and LI is obtained from Eq. (48). It shows that pI
and LI are nonlinearly affected by the material properties on both sides of the material interface.

Note that RI = R satisfies the condition that the density is constant in the incompressible flow
and u I = uR indicates that the interface velocity only depends on the incompressible flow.
3.4. Modified ghost fluid method
In the MGFM [28], a Riemann problem is defined to predict the interfacial states. However,
since the coupled compressible and incompressible Riemann problem has been constructed when
tracking the material interface in Section 3.1, the predicted interfacial states can be used directly.
As indicated in Fig. 2, the compressible flow (medium 1) and incompressible flow (medium 2) are

))& ))& ))& ))& ))&

separated by the fluid interface . N C , N D , N E , N F , N G are the normal vectors of marker points
))&
and N P is the normal vector of grid cell P adjacent to the interface in the compressible region.
The density at the cell P can be updated by the marker point nearby. The marker point E is
))&
))&
selected if the angle between N E and N P is the minimum compared with other marker points.
We employ the entropy at the marker point E for medium 1 to fix the real density at the cell P to
suppress the possible "overheating". Since in general the derivatives of physical solutions are
discontinuous across the interface, we project the updated density to the basis function space to
obtain the average density in cell P [32]. It is similar to update the density in other real fluid cells
adjacent to the interface in the compressible region.
In order to obtain the ghost fluid states, the advection equation is used:
))&
N = 0,
t

(64)

where is a column vector comprising the density, normal velocity, tangential velocity and

))&

pressure in the ghost fluid region, N is the unit normal vector of the grid cell. The "+" sign is
used if the interface boundary conditions of medium 1 are determined and the "-" sign is used if
the interface boundary conditions of medium 2 are determined. This advection equation is solved
by iterating at each time step. Due to the good compactness of the DG method, we only need to
consider the ghost cells which have the common edge with the real fluid cells and the ghost cells
far from the interface are not used [29, 30]. Once the advection equation is solved, we project the
flow states to the basis function space to obtain the average ghost fluid states, and then the
interface boundary conditions are obtained.
NF

NG

NE
ND

Medium 2

NP

G
Interface

NC

Medium 1

Fig. 2. Update the density adjacent to the interface for the compressible flow.

3.5. Summary of the solution procedures


Here we briefly summarize the solution procedures as follows:
Step 1: Construct the coupled compressible and incompressible Riemann problem to predict
interfacial states.
Step 2: Define ghost fluid states for ghost cells according to the MGFM implementation.
Step 3: Compute the flow field by solving the governing equations with the DG method.
Step 4: Update the interface location and get the final solutions according to the new interface
location.
Step 5: Return to Step 1 and proceed to the next time step.

4. Numerical tests
In this section, several gas-liquid two-medium flow problems are simulated on uniform
Cartesian grids. We fix the polynomial degree as kc=2 and ki=2 in the DG method for solving the
compressible and incompressible flows. The CFL coefficient C f is 0.18 and the TVB limiter
constant M is taken as 0.1 for the RKDG method. Unless otherwise specified, the ratio of the
specific heats =1.4 for compressible gas and the viscosity =0.001137kg/(m s).
4.1. One-dimensional case
All the one-dimensional examples are computed with 200 uniform cells in domain [0m,1m].
The compressible gas is located on the left while the incompressible liquid is on the right. The
results by new GFM are presented to make comparisons. In addition, these examples are the
compressible Riemann problems if both mediums are treated as compressible, and the
corresponding analytical solutions are also shown with the legend "Compressible" in the
numerical results as a reference.

Example 4.1. The gas-liquid interface is located at x=0.5m and the initial conditions are:

L =1kg/m3, uL =1000m/s, pL =1105Pa, for the gas, R =1009kg/m3, uR =0m/s, pR =2107Pa,


for the liquid. Fig. 3 shows the interfacial pressure histories. It is found that the interfacial pressure
by new GFM recovers more slowly and the recovery time to reach the same magnitude as MGFM
is almost at t=8.9310-5s. The pressure and velocity at this recovery time are shown in Fig. 4. For
the compressible analytical solutions, a shock wave is formed in the gas and a rarefaction wave is
propagated in the liquid (but it is not clear to be seen from the velocity result). It is observed that
the shock wave location by MGFM agrees well with the compressible analytical solutions while it
is falling behind by new GFM. It is in this much longer recovery stage of interfacial pressure that
leads to a discrepancy of shock wave location in the gas. As shown in Fig. 4, at the recovery time
the shock wave strengths formed in the gas for MGFM and new GFM are the same, and the only
difference is the shock wave location. Therefore, in order to show the correct shock wave location,
the gas mass errors are calculated:

M e = (M n M 0 M1 ) / ( M 0 + M1 ) ,

(65)

where Mn is the gas mass at the recovery time, M0 is the initial gas mass and M1 is the gas mass
which flows into the left boundary. It has to be emphasized that the mass conservation error is
unavoidable since MGFM and new GFM are all non-conservative methods. However, if the
interfacial pressure cannot recover in few time steps, the shock wave speed is affected and the
mass conservation error will increase quickly. The gas mass errors are calculated as 0.31%
(MGFM) and -8.24% (new GFM), respectively. It indicates that the mass error by MGFM is much
smaller and a more accurate shock wave location is obtained.
Here we replace the initial gas velocity by u L =1500m/s in order to increase the shock wave
strength formed in the gas. The interfacial pressure histories are shown in Fig. 5. The interfacial
pressure by new GFM recovers more slowly and reaches the same magnitude as MGFM almost at
t=2.4510-4s.

The numerical results at this recovery time are shown in Fig. 6. A larger

discrepancy of the shock wave location is observed. The gas mass errors are also calculated as
0.39% (MGFM) and -36.93% (new GFM). It demonstrates that the shock wave location by new
GFM is inaccurate while MGFM can still provide reasonable results with increased shock wave
strength.

Fig. 3. Example 4.1 with uL =1000m/s: interfacial pressure.

Fig. 4. Example 4.1 with uL =1000m/s: pressure (left) and velocity (right).

Fig. 5. Example 4.1 with uL =1500m/s: interfacial pressure.

Fig. 6. Example 4.1 with uL =1500m/s: pressure (left) and velocity (right).

Example 4.2. The gas-liquid interface is located at x=0.4m and the initial conditions are:

L =1kg/m3, uL =800m/s, pL =1105Pa, for the gas, R =1000kg/m3, uR =0m/s, pR =1105Pa,


for the liquid. Fig. 7 shows the interfacial pressure histories. It is found that the interfacial pressure
recovers more slowly by new GFM and reaches the same magnitude as MGFM almost at
t=8.410-5s. The pressure and velocity at this recovery time are shown in Fig. 8. We can find that
the shock wave is formed in both mediums in the compressible analytical solutions (but it is not
clear to be seen from the velocity result). The shock wave location in the gas is almost the same
for the MGFM and the compressible analytical solutions, while it is falling behind by the new
GFM. Since the shock wave strengths for MGFM and new GFM are the same in Fig. 8, the gas
mass errors are calculated to show the accuracy of shock wave location: 0.43% (MGFM), -6.25%
(new GFM). Once again, MGFM provides more accurate shock wave location.
Similarly, we replace the initial gas velocity by u L =1500m/s such that the shock wave
strength in the gas is increased. The interfacial pressure histories are shown in Fig. 9. It is noticed
that the interfacial pressure by new GFM needs more time to recover to the same magnitude by
MGFM. The numerical results at the recovery time t=2.4510-4s are shown in Fig. 10. A larger
discrepancy of shock wave location in the gas is observed. The gas mass errors are calculated as:

0.47% (MGFM), -41.74% (new GFM). It shows that more accurate shock wave location is
obtained by MGFM with increased shock wave strength.

Fig. 7. Example 4.2 with uL =800m/s: interfacial pressure.

Fig. 8. Example 4.2 with uL =800m/s: pressure (left) and velocity (right).

Fig. 9. Example 4.2 with uL =1500m/s: interfacial pressure.

Fig. 10. Example 4.2 with uL =1500m/s: pressure (left) and velocity (right).

Example 4.3. The gas-liquid interface is located at x=0.6m and the initial conditions are:

L =1kg/m3, u L =0m/s, p L =1105Pa, for the gas, R =1000.419kg/m3, u R =-814.8217m/s,


pR =1106Pa, for the liquid. Fig. 11 is the interfacial pressure histories. The interfacial pressure by

new GFM recovers more slowly and the recovery time to the same magnitude as MGFM is almost
at t=1.0510-4s. The pressure and velocity at this recovery time are shown in Fig. 12. For the
compressible analytical solutions, a shock wave is travelling in the gas while no wave should be
formed in the liquid. In general MGFM shows a good agreement with compressible analytical
solutions, while the shock wave location by new GFM is falling behind. The gas mass errors are
calculated as: 0.84% (MGFM), -4.86% (new GFM). A smaller mass error is obtained by MGFM.
The differences of the shock wave locations are not very large because the shock wave formed in
the gas is not very strong.
We further increase the initial liquid pressure to pR =4106Pa while ensure no wave is
formed in the liquid. The initial conditions are: L =1kg/m3, uL =0m/s, pL =1105Pa, for the gas,

R =1001.808kg/m3, uR =-1776.4m/s, pR =4106Pa, for the liquid. The shock wave strength
formed in the gas is increased to 40. The interfacial pressure histories are shown in Fig. 13. It is
obvious that the interfacial pressure by new GFM requires more time to reach the same magnitude
as MGFM. The numerical results at the recovery time t=1.6310-4s are shown in Fig. 14. A larger
discrepancy of shock wave location is observed. The gas mass errors are calculated as: 1.02%
(MGFM), -34.64% (new GFM). It can be seen that the result by MGFM is more reasonable with
much less mass error.

Fig. 11. Example 4.3 with pR =1106Pa: interfacial pressure.

Fig. 12. Example 4.3 with pR =1106Pa: pressure (left) and velocity (right).

Fig. 13. Example 4.3 with pR =4106Pa: interfacial pressure.

Fig. 14. Example 4.3 with pR =4106Pa: pressure (left) and velocity (right).

Example 4.4. The gas-liquid interface is located at x=0.6m. A shock wave which initially
coincides with the gas-liquid interface impacts the liquid from the gas side. The initial conditions
are: L =3.8125kg/m3, uL =814.82m/s, pL =1106Pa, for the post-shocked gas, R =1000kg/m3,
uR =0m/s, pR =1105Pa, for the liquid. Fig. 15 is the interfacial pressure histories. The interfacial

pressure by new GFM recovers to the same magnitude as MGFM almost at t=5.1210-5s. The
pressure and velocity at this recovery time are shown in Fig. 16. The gas mass errors are
calculated as: 0.25% (MGFM), -1.83% (new GFM). The differences of the shock wave locations
are not obvious because the shock wave formed in the gas is not very strong.
In order to increase the shock wave strength formed in the gas, the initial liquid velocity is
replaced by u R =-1000m/s. The interfacial pressure histories are shown in Fig. 17. The interfacial
pressure by new GFM recovers more slowly. The numerical results at the recovery time
t=8.5810-5s are shown in Fig. 18. A larger discrepancy of shock wave location can be observed.
The gas mass errors are calculated as: 0.96% (MGFM), -14.19% (new GFM). It demonstrates
again that MGFM can provide accurate results compared to new GFM when strong shock wave is
formed in the gas.

Fig. 15. Example 4.4 with uR =0m/s: interfacial pressure.

Fig. 16. Example 4.4 with uR =0m/s: pressure (left) and velocity (right).

Fig. 17. Example 4.4 with uR =-1000m/s: interfacial pressure.

Fig. 18. Example 4.4 with uR =-1000m/s: pressure (left) and velocity (right).

4.2. Applications to two-dimensional case


The two-dimensional numerical examples are taken from [5]. The computational domain is
[0m,1m][0m,1m] with 100 mesh cells in each direction and a cylindrical incompressible liquid
droplet surrounded by the compressible gas is centered at (0.5m,0.5m) with the radius of 0.2m.

Example 4.5. The initial conditions are: =1000kg/m3, u=100m/s, v=0m/s, p=1105Pa, for droplet,

=1.226kg/m3, u=0m/s, v=0m/s, p=1105Pa, for compressible gas. The pressure and x-velocity
along y=0.5 at t=510-4s are presented in Fig. 19. Fig. 20 shows the pressure contours at the same
time (note that the dashed line denotes the droplet interface). Since the droplet is moving in the
gas, a shock wave is formed ahead of the droplet and a rarefaction wave is formed behind it. The
interface shape is nearly not changed due to the large stiffness of the droplet. Here we replace the
droplet by a lighter one with an incompressible density of =10kg/m3, and the pressure and
x-velocity along y=0.5 are shown in Fig. 21. The interface locations at t=2.510-3s are presented in
Fig. 22 with the mesh sizes of 0.02m, 0.01m and 0.005m, respectively. The lighter droplet is
deformed more easily and slowed down faster by the compressible gas. Similar to the results in [5],
the Kelvin-Helmholtz instability occurs on the finest mesh. However, the numerical viscosity on
the coarser grids can non-physically weaken this effect. An area loss study of the lighter droplet is
carried out in Table 1. It shows that the area loss is small and decreases with mesh refinement.

Fig. 19. Example 4.5 with incompressible =1000kg/m3: pressure (left) and x-velocity (right).

Fig. 20. Example 4.5: 20 equally spaced pressure contours from 8104Pa to 1.3105Pa.

Fig. 21. Example 4.5 with incompressible =10kg/m3: pressure (left) and x-velocity (right).

(a) 0.02m.

(b) 0.01m.

(c) 0.005m.

Fig. 22. Example 4.5: interfaces with different mesh sizes (the dashed line denotes the initial interface).
Table 1. Area loss study for Example 4.5 with incompressible =10kg/m3.
Mesh size (m)
Area loss

0.02
0.65%

0.01
0.32%

0.005
0.11%

Example 4.6. A shock wave is initially located at x=0.1m with a post-shock state of =2.124kg/m3,
u=89.981m/s, v=0m/s, p=148407.3Pa. The shock travels to the right, impinging on the droplet
with the initial state of =1000kg/m3, u=0m/s, v=0m/s, p=98066.5Pa. The ambient compressible
gas has =1.58317kg/m3, u=0m/s, v=0m/s, p=98066.5Pa. The pressure and x-velocity along y=0.5
at t=1.2510-3s are presented in Fig. 23. Fig. 24 shows the pressure contours at the same time (the
dashed line denotes the droplet interface), where the incident shock wave and reflected shock
wave can be seen clearly, and this relative heavier droplet is nearly not set into motion by the
incident shock wave. Similar with Example 4.5, the same calculations are undertaken for a lighter
droplet with an incompressible density of =10kg/m3. Fig. 25 shows the pressure and the
x-velocity along y=0.5. In Fig. 26, it shows the initial interface location as compared to the
location at t=2.510-3s with the mesh sizes of 0.02m, 0.01m and 0.005m. The lighter droplet is
deformed and set into motion more easily. Note that the Kelvin-Helmholtz instability is appeared
on the finest mesh. The area loss of the lighter droplet is measured and presented in Table 2, which
shows a decreased area loss with mesh refinement again.

Fig. 23. Example 4.6 with incompressible =1000kg/m3: pressure (left) and x-velocity (right).

Fig. 24. Example 4.6: 20 equally spaced pressure contours from 1.0510 5Pa to 1.9105Pa.

Fig. 25. Example 4.6 with incompressible =10kg/m3: pressure (left) and x-velocity (right).

(a) 0.02m.

(b) 0.01m.

(c) 0.005m.

Fig. 26. Example 4.6: Interfaces with different mesh sizes (the dashed line denotes the initial interface).
Table 2. Area loss study for Example 4.6 with incompressible =10kg/m3.
Mesh size (m)
Area loss

0.02
1.75%

0.01
0.41%

0.005
0.30%

5. Concluding remarks
In this paper, a Riemann problem based method is proposed for solving compressible and
incompressible flows. An incompressible DG method with nonuniform time step is derived. The
numerical errors for new GFM are analyzed and discussed when shock wave is formed in the
compressible flow. A coupled compressible and incompressible Riemann problem is constructed
near the interface to predict the interfacial states, which are used to track the interface and define
the interface boundary conditions directly.
Extensive numerical examples are tested to demonstrate the good performances of the new
procedures. For the one-dimensional cases, by changing the initial conditions to increase the
strength of shock wave in the compressible flow, it is found that the new GFM reaches the correct
interfacial pressure more slowly and the discrepancy of shock wave location is observed. However,
MGFM can provide reasonable and accurate results including shock wave location by measuring
the mass conservation error. The two-dimensional numerical results show the applicability of the
extension to higher dimensions and the ability to capture all the important features successfully.

Acknowledgements
The research was supported by the National Basic Research Program of China ("973"
Program) under grant No. 2014CB046200, NSFC grants 11432007, 11372005, 11271188.
Additional support is provided by a project funded by the Priority Academic Program
Development (PAPD) of Jiangsu Higher Education Institutions.

References
[1] Abgrall R. How to prevent pressure oscillations in multicomponent flow calculations: a quasi conservative
approach. J Comput Phys 1996; 125:150-160.
[2] Agemi R. The incompressible limit of compressible fluid motions in a bounded domain. Proc. Japan Acad,
1981; 57: 291-293.
[3] Asano K. On the incompressible limit of the compressible Euler equation. Japan J Appl Math 1987; 4: 455-488.

[4] Bassi F, Crivellini A, Pietro D A D, Rebay S. An artificial compressibility flux for the discontinuous Galerkin
solution of the incompressible Navier-Stokes equations. J Comput Phys 2006; 218: 794-915.
[5] Caiden R, Fedkiw R P, Anderson C. A numerical method for two-phase flow consisting of separate
compressible and incompressible regions. J Comput Phys 2001; 166: 1-27.
[6] Cockburn B, Kanschat G, Schtzau D, Schwab C. Local discontinuous Galerkin methods for the Stokes system.
SIAM J. Numer. Anal. 2002; 40: 319-343.
[7] Cockburn B, Kanschat G, Schtzau D. The local discontinuous Galerkin method for the Oseen equations. Math.
Comp. 2003; 73: 569-593.
[8] Cockburn B, Kanschat G, Schtzau D. A locally conservative LDG method for the incompressible
Navier-Stokes equations. Math. Comp. 2005; 74: 1067-1095.
[9] Cockburn B, Hou S, Shu C-W. The Runge-Kutta local projection discontinuous Galerkin finite element method
for conservation laws IV: the multidimensional case. Math Comput 1990; 54: 545-581.
[10] Cockburn B, Lin S-Y, Shu C-W. TVB Runge-Kutta local projection discontinuous Galerkin finite element
method for conservation laws III: one dimensional systems. J Comput Phys 1989; 84: 90-113.
[11] Cockburn B, Shu C-W. TVB Runge-Kutta local projection discontinuous Galerkin finite element method for
conservation laws II: general framework. Math Comput 1989; 52: 411-435.
[12] Cockburn B, Shu C-W. The Runge-Kutta local projection P1-discontinuous Galerkin finite element method
for scalar conservation laws. Math Model Numer Anal 1991; 25: 337-361.
[13] Cockburn B, Shu C-W. The Runge-Kutta discontinuous Galerkin method for conservation laws V:
multidimensional systems. J Comput Phys 1998; 141: 199-224.
[14] Fedkiw RP. Coupling an Eulerian fluid calculation to a Lagrangian solid calculation with the ghost fluid
method. J Comput Phys 2002; 175: 200-224.
[15] Fedkiw RP, Aslam T, Merriman B, Osher S. A non-oscillatory Eulerian approach to interfaces in multimaterial
flows (the ghost fluid method). J Comput Phys 1999; 152: 457-492.
[16] Ferrer E, Willden R H J. A high order discontinuous Galerkin finite element solver for the incompressible
Navier-Stokes equations. Comput Fluids. 2011; 46: 224-230.
[17] Gibou F, Chen L, Nguyen D, Banerjee S. A level set based sharp interface method for the multiphase
incompressible Navier-Stokes equations with phase change. J Comput Phys 2007; 222: 536-555.
[18] Girault V, Wheeler F. Discontinuous Galerkin methods. Comp Meth Appl Sci. 2008; 16: 3-26.
[19] Glimm J, Grove JW, Li XL, Shyue K-M, Zeng Y, Zhang Q. Three-dimensional front tracking. SIAM J Sci
Comput 1998; 19: 703-727.
[20] Glimm J, Grove JW, Li XL, Zhao N. Simple front tracking. Contemp Math 1999; 238: 133-149.
[21] Glimm J, Grove JW, Zhang Y. Interface tracking for axisymmetric flows. SIAM J Sci Comput 2002; 24:
208-236.
[22] Guzmn J, Sequeira FA, Shu C-W. H(div) conforming and DG methods for incompressible Euler's equations.
IMA J Numer Anal, in press.
[23] Hu XY, Adams NA. An incompressible multi-phase SPH method. J Comput Phys 2007; 227: 264278.
[24] Karni S. Multicomponent flow calculations by a consistent primitive algorithm. J Comput Phys 1994; 112:
31-43.
[25] Karniadakis GE, Israeli M, Orszag SA. High-order splitting methods for the incompressible Navier-Stokes
equations. J Comput Phys 1991; 97: 414-443.
[26] Liu J-G, Shu C-W. A high-order discontinuous Galerkin method for 2D incompressible flows. J Comput Phys
2000; 160: 577-596.
[27] Liu TG, Khoo BC, Yeo KS. The simulation of compressible multi-medium flow. I. A new methodology with
test applications to 1D gas-gas and gas-water cases. Comput Fluids 2001; 30: 291-314.

[28] Liu TG, Khoo BC, Yeo KS. Ghost fluid method for strong shock impacting on material interface. J Comput
Phys 2003; 190: 651-681.
[29] Lu HT, Zhu J, Wang CW, Zhao N. Runge-Kutta discontinuous Galerkin method with front tracking method
for solving the compressible two-medium flow on unstructured meshes. Adv. Appl. Math. Mech 2017; 9: 1-19.
[30] Lu HT, Zhu J, Wang DH, Zhao N. Runge-Kutta discontinuous Galerkin method with front tracking method
for solving the compressible two-medium flow. Comput Fluids 2016; 126: 1-11.
[31] Osher S, Fedkiw RP. Level set methods: An overview and some recent results. J Comput Phys 2001; 169:
463-502.
[32] Qiu J, Liu TG, Khoo BC. Simulations of compressible two-medium flow by Runge-Kutta discontinuous
Galerkin methods with the ghost fluid method. Commun Comput Phys 2008; 3: 479-504.
[33] Riviere B. Discontinuous Galerkin methods for solving elliptic and parabolic equations: theory and
implementation. SIAM Front Appl Math 2008.
[34] Saurel R, Abgrall R. A simple method for compressible multifluid flows. SIAM J Sci Comput 1999; 21:
1115-1145.
[35] Shahbazi K, Fischer P F, Ethier C R. A high-order discontinuous Galerkin method for the unsteady
incompressible Navier-Stokes equations. J Comput Phys 2007; 222: 391-407.
[36] Shu C-W. TVB uniformly high-order schemes for conservation laws. Math Comp 1987; 49: 105-121.
[37] Shu C-W, Osher S. Efficient implementation of essentially non-oscillatory shock capturing schemes. J
Comput Phys 1988; 77: 439-471.
[38] Sussman M, Smith KM, Hussaini MY, Ohta M, Zhi-Wei R. A sharp interface method for incompressible
two-phase flows. J Comput Phys 2007; 221: 469505.
[39] Toro E F. Riemann solvers and numerical methods for fluid dynamics: a practical introduction. Springer,
Berlin, New York, 1997.

You might also like