Download as pdf or txt
Download as pdf or txt
You are on page 1of 23

International Journal of Multiphase Flow 113 (2019) 208–230

Contents lists available at ScienceDirect

International Journal of Multiphase Flow


journal homepage: www.elsevier.com/locate/ijmulflow

A numerical model for multiphase liquid–vapor–gas flows with


interfaces and cavitation
Marica Pelanti a,∗, Keh-Ming Shyue b
a
Institute of Mechanical Sciences and Industrial Applications, UMR 9219 ENSTA ParisTech – EDF – CNRS – CEA, 828, Boulevard des Maréchaux, Palaiseau
Cedex 91762, France
b
Department of Mathematics and Institute of Applied Mathematical Sciences, National Taiwan University, Taipei 10617, Taiwan

a r t i c l e i n f o a b s t r a c t

Article history: We are interested in multiphase flows involving the liquid and vapor phases of one species and a third
Received 2 September 2018 inert gaseous phase. We describe these flows by a hyperbolic single-velocity multiphase flow model com-
Revised 16 January 2019
posed of the phasic mass and total energy equations, the volume fraction equations, and the mixture mo-
Accepted 28 January 2019
mentum equation. The model includes stiff mechanical and thermal relaxation source terms for all the
Available online 29 January 2019
phases, and chemical relaxation terms to describe mass transfer between the liquid and vapor phases of
MSC: the species that may undergo transition. First, we present an analysis of the characteristic wave speeds
65M08 associated to the hierarchy of relaxed multiphase models corresponding to different levels of activation
76T10 of infinitely fast relaxation processes, showing that sub-characteristic conditions hold. We then propose
a mixture-energy-consistent finite volume method for the numerical solution of the multiphase model
Keywords:
Multiphase compressible flows
system. The homogeneous portion of the equations is solved numerically via a second-order wave prop-
Relaxation processes agation scheme based on robust HLLC-type Riemann solvers. Stiff relaxation source terms are treated by
Liquid–vapor phase transition efficient numerical procedures that exploit algebraic equilibrium conditions for the relaxed states. We
Finite volume schemes present numerical results for several three-phase flow problems, including two-dimensional simulations
Riemann solvers of liquid–vapor–gas flows with interfaces and cavitation phenomena.
© 2019 Elsevier Ltd. All rights reserved.

1. Introduction model that we have studied in previous work Pelanti and Shyue
(2014b). This model is composed of the phasic mass and total en-
Liquid–vapor flows are found in a large variety of industrial and ergy equations, the volume fraction equations, and the mixture
technological processes and natural phenomena. Often these flows momentum equation. The model includes thermal relaxation terms
involve one or more additional inert gas phases. For instance, in to account for heat transfer processes between all the phases, and
some processes the dynamics of a liquid–vapor mixture is cou- chemical relaxation terms to describe mass transfer between the
pled to the dynamics of defined regions of a third non-condensable liquid and vapor phases of the species that may undergo tran-
gaseous component. An example is given by underwater explosion sition. Similar hyperbolic multiphase flow models with instanta-
phenomena, where a high pressure bubble of combustion gases neous pressure relaxation have been previously presented for in-
triggers cavitation phenomena in water (Cole, 1948; Kedrinskiy, stance in Petitpas et al. (2009), Le Métayer et al. (2013), and Zein
2005). In other cases liquid–vapor mixtures may contain a diluted et al. (2013). A first contribution of our work is a rigorous deriva-
inert gas component, which may affect the flow features, such as tion of the reduced pressure-relaxed model resulting from the par-
in fuel injectors (Battistoni et al., 2015). We are interested here in ent non-equilibrium multiphase flow model with heat and mass
the simulation of this type of multiphase flows involving the liq- transfer terms in the limit of instantaneous mechanical relaxation.
uid and vapor phases of one species and one or more additional This is done by following the asymptotic analysis technique em-
non-condensable gaseous phases. We describe these multiphase ployed by Murrone and Guillard (2005) for the two-phase case
flows by a hyperbolic single-velocity compressible flow model with with no heat and mass transfer. Moreover, we present an original
infinite-rate mechanical relaxation, which extends the two-phase analysis of the characteristic wave speeds associated to the hierar-
chy of relaxed multiphase models corresponding to different levels
of activation of infinitely fast mechanical and thermo-chemical re-

Corresponding author. laxation processes. Similar to results shown in the literature for the
E-mail addresses: marica.pelanti@ensta-paristech.fr (M. Pelanti), shyue@math. two-phase case (Flåtten and Lund, 2011; Lund, 2012; Linga, 2018),
ntu.edu.tw (K.-M. Shyue).

https://doi.org/10.1016/j.ijmultiphaseflow.2019.01.010
0301-9322/© 2019 Elsevier Ltd. All rights reserved.
M. Pelanti and K.-M. Shyue / International Journal of Multiphase Flow 113 (2019) 208–230 209

we demonstrate that sub-characteristic conditions hold, namely spatial dimension:


the speed of sound of the multiphase mixture is reduced when-

N
ever an additional equilibrium assumption is introduced. Then, we ∂t αk + u · ∇αk = Pk j , k = 1, 3, . . . , N, (1a)
present a finite volume method for the numerical solution of the j=1
multiphase model system based on a classical fractional step pro-
cedure. The homogeneous hyperbolic portion of the equations is
solved numerically via a second-order accurate wave propagation
∂t (α1 ρ1 ) + ∇ · (α1 ρ1 u ) = M, (1b)
scheme, which employs a HLLC-type Riemann solver. In particu-
lar, we present here a new generalized HLLC-solver based on the
idea of the Suliciu relaxation solver of Bouchut (2004), extending
∂t (α2 ρ2 ) + ∇ · (α2 ρ2 u ) = −M, (1c)
the solver that we have recently proposed in De Lorenzo et al.
(2018) for the two-phase case. This HLLC/Suliciu-type solver allows
us to guarantee positivity preservation with a suitable choice of the
∂t (αk ρk ) + ∇ · (αk ρk u ) = 0 , k = 3, . . . , N, (1d)
wave speeds. Stiff relaxation source terms are treated by efficient
numerical procedures that exploit algebraic equilibrium conditions   

N
for the relaxed states, following the ideas of our work (Pelanti ∂t (ρ u ) + ∇ · ρ u  u + αk pk I = 0, (1e)
and Shyue (2014b)). Similar approaches have been previously pre- k=1
sented in the literature for instance in Le Métayer et al. (2013). One
important property of our numerical method is mixture-energy-
consistency in the sense defined in Pelanti and Shyue (2014b)), that ∂t (α1 E1 ) + ∇ · (α1 (E1 + p1 )u ) + ϒ1

is the method guarantees conservation of the mixture total energy N N
|u |2
at the discrete level, and it guarantees consistency by construction =− pI1 j P1 j + Q1 j + gI + M, (1f)
2
of the values of the relaxed states with the mixture pressure law. j=1 j=1

This property is ensured thanks to the total-energy-based formula-


tion of the model system. We present several numerical results for ∂t (α2 E2 ) + ∇ · (α2 (E2 + p2 )u ) + ϒ2
three-phase flow problems, including problems involving liquid– 
N N
|u |2
vapor–gas flows with interfaces and cavitation phenomena, such =− pI2 j P2 j + Q2 j − gI + M, (1g)
as underwater explosion tests. 2
j=1 j=1
This article is organized as follows. In Section 2 we present
the multiphase flow model under study. In Section 3 we derive
∂t (αk Ek ) + ∇ · (αk (Ek + pk )u ) + ϒk
the limit pressure-equilibrium model associated to the considered
multiphase flow model, and we analyze the characteristic speeds 
N 
N
=− pIk j Pk j + Qk j , k = 3, . . . , N. (1h)
of the relaxed models in the hierarchy stemming from the parent
j=1 j=1
relaxation model. In Section 4 we illustrate the numerical method
that we have developed to solve the multiphase flow equations. The non-conservative terms ϒ k appearing in the phasic total en-
Several one-dimensional and two-dimensional numerical experi- ergy Eqs. (1f)–(1h) are given by
ments are finally presented in Section 5.    
→ 
N
ϒk = u · Yk ∇ α j p j − ∇ (αk pk ) , k = 1, . . . , N, (1i)
j=1
α ρ
2. Single-velocity multiphase compressible flow model where Yk = kρ k denotes the mass fraction of phase k. In the sys-
tem above Pk j and Qk j represent the volume transfer and the heat
We consider an inviscid compressible flow composed of N transfer, respectively, between the phases k and j, k, j = 1, . . . , N.
phases that we assume in kinematic equilibrium with velocity u . The term M indicates the mass transfer between the liquid and
In this work we are mainly interested in three-phase flows, N = 3, vapor phases indexed with 1 and 2. The transfer terms are defined
nonetheless we shall present here a general multiphase flow for- as relaxation terms:
mulation. The volume fraction, density, internal energy per unit P k j = μk j ( p k − p j ) , Qk j = ϑk j (T j − Tk ), M = ν ( g2 − g1 ) , (2)
volume, and pressure of each phase will be denoted by α k , ρ k ,
Ek , pk , k = 1, . . . , N, respectively. We will denote the total energy where Tk denotes the phasic temperature, gk the phasic chemical
|u |2 potential, and where we have introduced the mechanical, thermal
for the kth phase with Ek = Ek + ρk 2 . The saturation condition
N N and chemical relaxation parameters μk j = μ jk ≥ 0, ϑk j = ϑ jk ≥ 0,
is k=1 αk = 1. The mixture density is ρ = k=1 αk ρk , the mix- and ν = ν12 = ν21 ≥ 0, respectively. Note that: Pkk = 0, Qkk = 0,
N
ture internal energy is E = k=1 αk Ek , and the mixture total en- Pk j = −P jk and Qk j = −Q jk . The quantities pIk j = pI jk are interface
 |u |2
ergy is E = N k=1 αk Ek = E + ρ 2 . Moreover, we will denote the pressures and gI is an interface chemical potential. We shall as-
kth specific internal energy with εk = Ek /ρk . Mechanical and ther- sume that all mechanical relaxation processes are infinitely fast,
mal transfer processes are considered in general for all the phases. μk j = μ jk ≡ μ → +∞, so that mechanical equilibrium is attained
We assume that one species in the mixture can undergo phase instantaneously between all the phases. Indeed here, following the
transition, so that it can exist as a vapor or a liquid phase, and same idea of Saurel et al. (20 08, 20 09) and Pelanti and Shyue
mass transfer terms are accounted for this species only. We will (2014b)), the parent non-equilibrium multiphase flow model with
use the subscripts 1 and 2 to denote the liquid and vapor phases instantaneous pressure relaxation is used to approximate solutions
of this species. We describe the N-phase flow under consideration to the limiting pressure-equilibrium flow model, which is the phys-
by a compressible flow model that extends the six-equation two- ical flow model of interest. Concerning thermal and chemical re-
phase flow system that we studied in Pelanti and Shyue (2014b). laxation, following the simple approach of Saurel et al. (2008),
The model system is composed of the volume fraction equations we consider in this work that these processes are either inac-
for N − 1 phases, the mass and total energy equations for all the N tive, ϑk j = 0, ν = 0, or they act infinitely fast, ϑk j → +∞, ν → +∞.
phases, and d mixture momentum equations, where d denotes the Heat and mass transfer may be activated at selected locations, for
210 M. Pelanti and K.-M. Shyue / International Journal of Multiphase Flow 113 (2019) 208–230

instance at interfaces for a phase pair (k, j), identified by min (α k , where we have introduced the mixture pressure
α j ) > , where is a tolerance.
The closure of the system (1) is obtained through the specifica- 
N
Yk
pm (ρ , s1 , . . . , sN , Y1 , . . . , YN−1 , α1 , . . . , αN−1 ) = αk pk sk , ρ .
tion of an equation of state (EOS) for each phase pk = pk (Ek , ρk ), αk
k=1
Tk = Tk ( pk , ρk ). For the numerical model here we will adopt the
widely used stiffened gas (SG) equation of state (Menikoff and (8)
Plohr, 1989): From this definition, by noticing that
  
pk (Ek , ρk ) = (γk − 1 )Ek − γk k − (γk − 1 )ηk ρk , (3a) ∂ pk ∂ pk ∂ρk Yk
= = ck2 , (9)
∂ρ sk , Yk , αk
∂ρk sk , Yk , αk
∂ρ sk , Yk , αk
αk

pk + we obtain the expression:


Tk ( pk , ρk ) = k
, (3b)
κvk ρk (γk − 1 ) 
N
cf = Yk ck2 , (10)
where γ k , ϖk , ηk and κ vk are constant material-dependent param-
k=1
eters. In particular, κ vk represents the specific heat at constant vol-
ume. The corresponding expression for the phasic entropy is ∂ pk
where ck = ( ∂ρ )sk is the speed of sound of the phase k, which
γ
k

sk = κvk log(Tk k ( pk + k )−(γk −1) ) + ηk , (3c) can be expressed as ck = k hk + χk , where k = (∂ pk /∂ Ek )ρk
(Grüneisen coefficient), and χk = (∂ pk /∂ ρk )Ek .
where ηk = constant, and gk = hk − Tk sk , with hk denoting the pha-
sic specific enthalpy. The parameters for the SG EOS for the liq- 3. Hierarchy of multiphase relaxed models and speed of sound
uid and vapor phases of the species that may undergo transition
are determined so that the theoretical saturation curves defined By considering different levels of activation of instantaneous re-
by g1 = g2 fit the experimental ones for the considered material laxation processes we can establish from the model (1) a hierarchy
(Le Métayer et al., 2004). The mixture pressure law for the model of hyperbolic multiphase flow models. Here in particular we de-
with instantaneous pressure relaxation is determined by the mix- rive the expression of the speed of sound for the relaxed models
ture energy relation in this hierarchy, similar to Flåtten and Lund (2011) and Flåtten
et al. (2010).

N
E= αk Ek ( p, ρk ), (4)
k=1 3.1. p-Relaxed model

where we have used the mechanical equilibrium conditions pk = p, In the considered limit of instantaneous mechanical relaxation
for all k = 1, . . . , N, in the phasic energy laws Ek ( pk , ρk ). Note that μk j ≡ μ → +∞, the model system (1) reduces to a hyperbolic
for the particular case of the SG EOS, an explicit expression of the single-velocity single-pressure model, which is a generalization of
mixture pressure can be obtained from (4). the five-equation two-phase flow model of Kapila et al. (2001).
Since here we will consider relaxation parameters either = 0 or The reduced pressure equilibrium model, which we shall also call
→ +∞, a specification of the expression for the interface quanti- p-relaxed model, can be derived by means of asymptotic analysis
ties pIkj , gI is not needed. Nevertheless, let us remark that the def- techniques, cf. in particular Murrone and Guillard (2005). We show
inition of these interface quantities must be consistent with the the derivation for the one-dimensional case in Appendix A. Denot-
second law of thermodynamics, which requires a non-negative en- ing with p the equilibrium pressure, we obtain the following re-
tropy production for the mixture. The equation for the mixture to- laxed system, composed of 2N + d equations:

tal entropy S = ρ s , s = N k=1 Yk sk , is found as:
1  N

∂t S + ∇ · (S u ) = HP + HQ + HM , (5a) ∂t α1 + u · ∇α1 = K1 ∇ · u + Q
ρ1 c12 j=2 1 j
where 
ρ c2p  N
j i ρ c2p
− α1 Q − +

N 
N
pk − pIk j 
N 
N
1 ρ1 c12 j,i=1 ji ρ j c2j ρi ci2 ρ1 c12
HP = Pk j , HQ = Q ,
Tk Tk k j i> j
k=1 j=1 k=1 j=1  
gI − g1 gI − g2 
N
αj α
HM = − M. (5b) × (1 (gI − h1 ) + c12 ) + (2 (gI − h2 ) + c22 )
1
M,
T1 T2 j=2
ρ j c2j ρ2 c22
For consistency of the multiphase model (1) with the second (11a)
law of thermodynamics we need HP + HQ + HM ≥ 0. By following
the arguments in Flåtten and Lund (2011), one can infer the follow- k  N

ing sufficient consistency conditions on the interface quantities: ∂t αk + u · ∇αk = Kk ∇ · u + Q


ρk ck j=1 k j
2

pIk j ∈ [min( pk , p j ), max( pk , p j )] j =k



and gI ∈ [min(g1 , g2 ), max(g1 , g2 )]. (6) ρ c2p  N
j i αk
− αk Q ji − + ρ c2p
ρ 2
k ck j,i=1 ρ j c2j ρi ci2 ρk ck2
The model (1) is hyperbolic and the associated speed of sound cf
i> j
(non-equilibrium or frozen sound speed) is defined by 
 2 (gI − h2 ) + c22 1 (gI − h1 ) + c12
∂ pm × − M, k = 3, . . . , N,
cf2 = , (7) ρ2 c22 ρ1 c12
∂ρ sk , Yk , αk , k=1,...,N (11b)
M. Pelanti and K.-M. Shyue / International Journal of Multiphase Flow 113 (2019) 208–230 211

This definition in general does not satisfy the sufficient condition


for entropy consistency (6). Nevertheless, let us note that the nu-
∂t (α1 ρ1 ) + ∇ · (α1 ρ1 u ) = M, (11c)
merical model in Saurel et al. (2008) considers either no mass
transfer or infinite-rate mass transfer, so that the factor multiply-
ing M in (16) has no influence in these specific circumstances.
∂t (α2 ρ2 ) + ∇ · (α2 ρ2 u ) = −M, (11d)
Remark 2. In our previous work (Pelanti and Shyue (2014b)) an
additional term of the form M/ρI was written in the volume frac-
∂t (αk ρk ) + ∇ · (αk ρk u ) = 0, k = 3, . . . , N, (11e) tion equation of the six-equation two-phase flow model corre-
sponding to (1) for N = 2, with ρI representing an interface den-
sity. Similar to Flåtten and Lund (2011), this term is not included in
the present multiphase model (1). The purpose of the term M/ρI
∂t (ρ u ) + ∇ · (ρ u  u + pI ) = 0, (11f)
in Pelanti and Shyue (2014b)) was to indicate the influence of
the mass transfer process on the evolution of the volume fraction.
Nonetheless, the rigorous derivation of the pressure-relaxed model
∂t E + ∇ · ((E + p)u ) = 0, (11g) (11) from the system (1) reveals that indeed mass transfer terms
where affect α k via the pressure relaxation process, as we observe from
  the contribution of M appearing in (11a) and (11b) (and (15) for

N
1 1 ρ c2p the case N = 2). Note that neglecting the term M/ρI in the six-
Kk = ρ c2p αk αj − = αk −1 . (12)
j=1
ρk ck2 ρ j c2j ρk ck2 equation two-phase model of Pelanti and Shyue (2014b) does not
j =k affect the numerical model and the numerical results presented
there, since ν = 0 or ν → +∞, and the numerical procedure for
In the relations above we have introduced the pressure equilibrium
treating instantaneous chemical relaxation consists in imposing di-
speed of sound cp (a generalization of Wood’s sound speed), de-
rectly algebraic thermodynamic equilibrium conditions.
fined by

∂p 3.2. pT-relaxed models
c2p = , (13)
∂ρ s1 ,...,sN ,Y1 ,...,YN
Assuming instantaneous mechanical equilibrium μ jk ≡ μ →
from which we obtain the expression: +∞ for all the phases and thermal equilibrium ϑk j ≡ ϑ → +∞
 − 12 for M phases, 2 ≤ M ≤ N, we obtain a hyperbolic relaxed system
αk 
N
of 2N − M + 1 + d equations characterized by the speed of sound
cp = ρ . (14)
ρ c2
k=1 k k
cpT,M , defined by

As we mentioned above, the pressure equilibrium model (1) is
1 ∂p
= , (18)
indeed the physical flow model of interest. Similar to the two- c pT,M 2 ∂ρ M
k=1 Yk sk ,sM+1 ,...,sN ,Y1 ,...,YN
phase case (Saurel et al., 2009; Zein et al., 2010; Pelanti and Shyue
(2014b)), the non-equilibrium model (11) with instantaneous me- From this definition we obtain the expression:
chanical relaxation is convenient to approximate numerically solu-  2
1 1 ρ T M−1  
M
j k
tions to the p-relaxed model. = 2 + M C pk Cp j − , (19)
c pT,M 2 cp k=1 C pk k=1 j=k+1
ρ j c2j ρk ck2
Remark 1. For the two-phase case N = 2 the p-relaxed model (11)
has a form analogous to the pressure-equilibrium model presented where T denotes the equilibrium temperature, C pk = αk ρk κ pk ,
by Saurel et al. (2008), nonetheless we remark a difference in the κ pk = (∂ hk /∂ Tk ) pk (specific heat at constant pressure), and we re-
expression of mass transfer term appearing in the volume fraction call k = (∂ pk /∂ Ek )ρk . Let us note that in the particular case of
equation. The equation for α 1 obtained from (11) for N = 2 can be thermal equilibrium for all the phases, M = N, the reduced single-
written as: pressure single-temperature pT-relaxed multiphase model has the
 conservative form:
1 2
∂t α1 + u · ∇α1 = K1 ∇ · u + ζ + Q
α1 α2 ∂t (α1 ρ1 ) + ∇ · (α1 ρ1 u ) = M, (20)

1 (gI − h1 ) + c12 2 (gI − h2 ) + c22
+ζ + M, (15)
α1 α2 ∂t (α2 ρ2 ) + ∇ · (α2 ρ2 u ) = −M, (21)
α1 α2
where K1 = ζ (ρ2 c22 − ρ1 c12 ) and ζ = . The equation
α2 ρ1 c12 +α1 ρ2 c22
for the volume fraction α 1 of the relaxed pressure-equilibrium
model reported in Saurel et al. (2008) is:
∂t (αk ρk ) + ∇ · (αk ρk u ) = 0 , k = 3, . . . , N, (22)

  2
1 2 c1 c2
∂t α1 + u · ∇α1 = K1 ∇ · u + ζ + Q+ζ + 2 M. ∂t (ρ u ) + ∇ · (ρ u  u + pI ) = 0, (23)
α1 α2 α1 α2
(16)
We observe that the two formulations are equivalent only with the ∂t E + ∇ · ((E + p)u ) = 0. (24)
following definition of the interface chemical potential gI :
The two-phase (N = 2) version of this model was considered for
α2 1 h1 + α1 2 h2 instance in Lund and Aursand (2012), and more recently in Saurel
gI = . (17)
α2 1 + α1 2 et al. (2016).
212 M. Pelanti and K.-M. Shyue / International Journal of Multiphase Flow 113 (2019) 208–230

3.3. pTg-relaxed models

We now assume instantaneous mechanical equilibrium μ jk ≡


μ → +∞ for all the phases, thermal equilibrium ϑk j ≡ ϑ → +∞
for M phases, 2 ≤ M ≤ N, and, additionally, we consider instanta-
neous chemical relaxation between the liquid and vapor phases 1
and 2, ν → +∞. We consider that at least the liquid–vapor phase
pair is in thermal equilibrium. With these hypotheses we obtain a
hyperbolic relaxed system of 2(N − M + 1 ) + d equations character-
ized by a speed of sound cpTg,M , defined by


2 ∂p
c pTg,M = , (25)
∂ρ M
k=1 Yk sk ,sM+1 ,...,sN ,Y3 ,...,YN

from which we obtain

  2
1 1 ρT 
M
kCpk 1 dT 
M
= + M − C pk ,
c pTg,M 2 c pT,M 2 k=1 C pk k=1
ρk ck2 T dp
sat k=1 Fig. 1. Speed of sound for a three-phase mixture made of liquid water, water vapor
and air versus the total gaseous volume fraction αvg = αv + αg . We use the sub-
(26) scripts l,v,g to indicate the liquid phase, the vapor phase and the non-condensable
gas phase, respectively. cf , c pT = c pT,3 , c pT ( jk ) = c pT,2 , c pTg = c pTg,3 , c pT ( jk )g = c pTg,2 are
the speeds defined in (10), (14), (19), and (26). Here the notation T(jk), j, k = l, v, g,
where we have introduced the derivatives (dT/dp)sat evaluated on
specifies the two phases for which thermal equilibrium is assumed (for instance
the liquid–vapor saturation curve. As expected (cf. e.g. Stewart and cpT(lv) denotes the speed of sound for a mixture characterized by pressure equilib-
Wendroff, 1984), analogously to the two-phase case (Flåtten and rium for all the phases and thermal equilibrium for the liquid and vapor pair only).
Lund, 2011), it is easy to observe from (14), (19), and (26) that
sub-characteristic conditions hold, namely the speed of sound of
the N-phase mixture is reduced whenever an additional equilib-
rium assumption is introduced:

c pTg ≡ c pTg,N ≤ c pTg,M , c pT ≡ c pT,N ≤ c pT,M , and


c pTg < c pT < c p < cf .
⎡ ⎤ ⎡ ⎤
(27) α1 0
⎢ α3 ⎥ ⎢ 0 ⎥
⎢ . ⎥ ⎢ .. ⎥
Let us note that in the particular case of thermal equilibrium for ⎢ . ⎥ ⎢ ⎥
all the phases, M = N, the reduced pTg-relaxed multiphase model ⎢ . ⎥ ⎢ . ⎥
⎢ αN ⎥ ⎢ 0 ⎥
corresponds to the well known Homogeneous Equilibrium Model ⎢ ⎥ ⎢ ⎥
⎢ α1 ρ1 ⎥ ⎢ α1 ρ1 u ⎥
(HEM) (Stewart and Wendroff, 1984), composed of the conserva- ⎢ ⎥ ⎢ ⎥
tion laws for the mixture density ρ , the mixture momentum ρ u , ⎢ α2 ρ2 ⎥ ⎢ α2 ρ2 u ⎥
⎢ . ⎥ ⎢ .. ⎥
and the mixture total energy E. The derivation of the expression ⎢ ⎥ ⎢
q = ⎢ .. ⎥, F (q ) = ⎢ ⎥,
. ⎥
of the speed of sound for the considered hierarchy of multiphase ⎢α ρ ⎥ ⎢ αN ρN u ⎥
⎢ N N⎥ ⎢   ⎥
flow models is detailed in Appendix B. We conclude this section by ⎢ ρ u ⎥ ⎢ρ u  u + Nk=1 αk pk I⎥
showing in Fig. 1 the behavior of the sound speed for different lev- ⎢ ⎥ ⎢ ⎥
⎢ α1 E1 ⎥ ⎢ α1 (E1 + p1 )u ⎥
els of activation of instantaneous mechanical, thermal and chemi- ⎢ ⎥ ⎢ ⎥
⎢ α2 E2 ⎥ ⎢ α2 (E2 + p2 )u ⎥
cal relaxation for a three-phase mixture made of liquid water, wa- ⎢ . ⎥ ⎢ ⎥
ter vapor and air (non-condensable gas). Here we plot the speed of ⎣ .. ⎦ ⎣ .
.. ⎦
sound versus the volume fraction of the total gaseous component αN EN αN (EN + pN )u
αgv = αv + αg for a fixed ratio αg /αv = 0.5, where here α v is the
vapor volume fraction, and α g is the non-condensable gas volume ⎡ ⎤
u · ∇α1
fraction. The reference pressure is p = 105 Pa, and the reference
⎢ u · ∇α3 ⎥
temperature is the corresponding saturation temperature. The pa- ⎢ . ⎥
⎢ . ⎥
rameters used for the equations of state of the phases are the same ⎢ . ⎥
as those of the cavitation tube experiment in Section 5.2 (Experi- ⎢u · ∇αN ⎥
⎢ ⎥
ment 5.2.1). ⎢ 0 ⎥
⎢ ⎥
⎢ 0 ⎥
⎢ . ⎥
ς (q, ∇ q ) = ⎢
⎢ .. ⎥,
⎥ (28b)
⎢ 0 ⎥
4. Numerical method ⎢ ⎥
⎢ 0 ⎥
⎢ ⎥
⎢ ϒ1 ⎥
We focus now on the numerical approximation of the multi- ⎢ ⎥
phase system (1), which we can write in compact vectorial form ⎢ ϒ2 ⎥
⎢ . ⎥
denoting with q ∈ R3N−1+d the vector of the unknowns: ⎣ .. ⎦
ϒN
∂t q + ∇ · F (q ) + ς (q, ∇ q ) = ψμ (q ) + ψϑ (q ) + ψν (q ) , (28a)
M. Pelanti and K.-M. Shyue / International Journal of Multiphase Flow 113 (2019) 208–230 213

⎡ N ⎤ ⎡ ⎤ x, and we denote with Qin the approximate solution of the sys-
P 0
Nj=1 1 j tem at the ith cell and at time tn , i ∈ Z, n ∈ N. The second-order
⎢ j=1 P3 j ⎥ ⎢ 0 ⎥
⎢ ⎥ ⎢ ⎥ wave propagation algorithm has the form
⎢ .. ⎥ ⎢ .. ⎥
⎢ . ⎥ ⎢ . ⎥
⎢ N ⎥ ⎢ ⎥ t + t h
⎢ P ⎥ ⎢ 0 ⎥ Qin+1 = Qin − (A Qi−1/2 + A− Qi+1/2 ) − (F − Fh ) .
⎢ j=1 N j
⎥ ⎢ 0 ⎥ x x i+1/2 i−1/2
⎢ 0 ⎥ ⎢ ⎥
⎢ 0 ⎥ ⎢ 0 ⎥ (29)
⎢ ⎥ ⎢ ⎥
⎢ . ⎥ ⎢ ..⎥ A ∓ Q
ψμ ( q ) = ⎢ .. ⎥, ψθ ( q ) = ⎢ .⎥, Here i+1/2 are the so-called fluctuations arising from Rie-
⎢ ⎥ ⎢ ⎥ mann problems at cell interfaces (i + 1/2 ) between adjacent cells
⎢ 0 ⎥ ⎢ 0 ⎥
⎢ ⎥ ⎢ 0 ⎥ i and (i + 1 ), and Fih+1/2 are correction terms for (formal) second-
⎢  0 ⎥ ⎢ N ⎥
⎢− N p P ⎥ ⎢ j=1 Q1 j ⎥ order accuracy. To define the fluctuations, a Riemann solver
⎢  j=1 I1 j 1 j ⎥ ⎢ N ⎥
⎢− N p P ⎥ ⎢ j=1 Q2 j ⎥ (cf. Godlewski and Raviart, 1996; Toro, 1997; LeVeque, 2002) must
⎢ j=1 I2 j 2 j ⎥ ⎢ ⎥ be provided. The solution structure defined by a given solver for a
⎢ .. ⎥ ⎢ .. ⎥
⎣ .
⎦ ⎣ . ⎦ Riemann problem with left and right data q and qr can be ex-
N N pressed in general by a set of M waves W l and corresponding
− j=1 pIN j PN j j=1 QN j
speeds sl , M  3N. For example, for the HLLC-type solver described
⎡ 0
⎤ below M = 3. The sum of the waves must be equal to the initial
⎢ 0 ⎥ jump in the vector q of the system variables:
⎢ .. ⎥
⎢ ⎥ M

⎢ . ⎥ q ≡ q r − q  = Wl. (30)
⎢ 0 ⎥
⎢ ⎥
⎢ M ⎥ l=1
⎢ −M

⎢ ⎥ Moreover, for any variable of the model system governed by a con-
⎢ .. ⎥ servative equation the initial jump in the associated flux function
⎢ ⎥
ψν ( q ) = ⎢

. ⎥,
⎥ (28c) must be recovered by the sum of waves multiplied by the corre-
⎢ 0 ⎥ sponding speeds. In the considered model the conserved quanti-
⎢ 0 ⎥ ties are α k ρ k , k = 1, . . . , N, and ρ u, therefore in order to guarantee
⎢ ⎥
⎢ gI + |u|2 M ⎥ conservation we need:
⎢ 2 ⎥
⎢ ⎥ M

⎢− gI + |u|2 M⎥  f ( ξ ) ≡ f ( ξ ) ( qr ) − f ( ξ ) ( q ) = sl W l ( ξ )
⎢ 2 ⎥ (31)
⎢ ⎥
⎣ .
. ⎦ l=1
.
for ξ = N, . . . , 2N, where f(ξ ) is the ξ th component of the flux
0
vector f, and W l (ξ ) denotes the ξ th component of the lth wave,
with ϒ k (q, ∇ q) defined in (1i). Above we have put into evi- l = 1, . . . , M . It is clear that conservation of the partial densities

dence the conservative portion of the spatial derivative contribu- ensures conservation of the mixture density ρ = N k=1 αk ρk . In ad-
tions in the system as ∇ · F (q ), and we have indicated the non- dition, we must ensure conservation of the mixture total energy,
conservative term as ς (q, ∇ q). The source terms ψ μ (q), ψ θ (q),
M
 
ψ ν (q) contain mechanical, thermal and chemical relaxation terms, N

respectively, as expressed in (2).


 f E ≡ f E ( qr ) − f E ( q ) = sl W l (2N+k ) , (32)
l=1 k=1
To numerically solve the system (28) we use the same tech-

k=1 αk pk ) is the flux function associated to
niques that we have developed for the two-phase model in where fE = u(E + N
Pelanti and Shyue (2014b). A fractional step method is employed, the mixture total energy E. Once the Riemann solution struc-
where we alternate between the solution of the homogeneous ture {Wil+1/2 , sli+1/2 }l=1,...,M arising at each cell edge xi+1/2 is de-
system ∂t q + ∇ · F (q ) + ς (q, ∇ q ) = 0 and the solution of a se- fined through a Riemann solver, the fluctuations A∓ Qi+1/2 and
quence of systems of ordinary differential equations (ODEs) that
the higher-order (second-order) correction fluxes Fih+1/2 in (29) are
take into account the relaxation source terms ψ μ , ψ ϑ , and ψ ν .
computed as
As in Pelanti and Shyue (2014b), the resulting method is mixture-
energy-consistent, in the sense that (i) it guarantees conservation M

at the discrete level of the mixture total energy; (ii) it guarantees A± Qi+1/2 = (sli+1/2 )± Wil+1/2 , (33)
consistency by construction of the values of the relaxed states with l=1
the mixture pressure law. The method has been implemented by where we have used the notation s+ = max(s, 0 ), s− = min(s, 0 ),
using the libraries of the clawpack software (LeVeque). and

1  l  t  l 
M
4.1. Solution of the homogeneous system h
Fi+1 /2 = si+1/2  1 − si+1/2  Wil+1
h
/2 , (34)
2 x
l=1
To solve the hyperbolic homogeneous portion of (28) we em-
ploy the wave-propagation algorithms of LeVeque (2002, 1997), where Wil+1
h
/2
are a modified version of Wil+1/2 obtained by apply-
which are a class of Godunov-type finite volume methods to ap- ing to Wil+1/2 a limiter function (cf. LeVeque, 2002).
proximate hyperbolic systems of partial differential equations. We One difficulty in the solution of the homogeneous portion of
shall consider here for simplicity the one-dimensional case in the the multiphase system (28) is the presence of the non-conservative
x direction (d = 1), and we refer the reader to LeVeque (2002) for products ϒ k in the phasic energy equations. Although a discus-
a comprehensive presentation of these numerical schemes. Hence sion of the treatment of non-conservative terms is not the main
we consider here the solution of the one dimensional system focus of the present work, it is important to recall the associated
∂t q + ∂x f (q ) + ς (q, ∂x q ) = 0, q ∈ R3N (as obtained by setting u = u issues and challenges. It is well known that a first difficulty of non-
and ∇ = ∂x in (28)). We assume a grid with cells of uniform size conservative hyperbolic systems is the lack of a notion of weak
214 M. Pelanti and K.-M. Shyue / International Journal of Multiphase Flow 113 (2019) 208–230

solution in the distributional framework for problems involving tions (Baer and Nunziato, 1986) was proposed in Tokareva and Toro
shocks. The theory of Dal Maso–LeFloch–Murat (Dal Maso et al., (2010). HLLC-type solvers for two-phase flows were also adopted
1995) has marked an advance by offering a possible definition of for instance in Saurel et al. (2009) and Zein et al. (2010). Still
weak solution, based on the concept of non-conservative products within the class of extended HLL solvers able to represent inter-
as a Borel measure associated to a choice of a family of paths. mediate waves, let us finally mention the HLLEM Riemann solver
Even with this rigorous theoretical framework and the assump- for general conservative and non-conservative hyperbolic systems
tion of a known correct shock wave solution, further difficulties introduced in Dumbser and Balsara (2016). This solver includes
arise in the design of numerical methods able to correctly approxi- the discretization of non-conservative products in the framework
mate non-conservative systems. The path-conservative schemes in- of path-conservative HLL schemes and it was applied in Dumbser
troduced in the seminal article by Parés (2006) are formally consis- and Balsara (2016) to several non-conservative systems, including
tent with the definition of non-conservative products of Dal Maso the Baer–Nunziato equations.
et al. (1995), once a family of paths has been selected. Nonetheless,
this approach has still some known shortcomings as for instance
discussed in Castro et al. (2008) and Abgrall and Karni (2010). 4.1.1. A simple HLLC-type solver
Concerning more specifically the multiphase flow model under We present in this subsection an extension to the multiphase
study with stiff mechanical relaxation, difficulties related to the system (1) of the HLLC-type solver illustrated in Pelanti and Shyue
non-conservative products in the energy equations arise for prob- (2014b) for the two-phase case. This solver is obtained by apply-
lems involving shocks in genuine multiphase mixtures (flow con- ing the standard HLLC method (Toro et al., 1994; Toro, 1997) to
ditions not close to nearly single-phase fluids). The shock jump re- the conservative portion of the multiphase system, neglecting the
lations for two-phase mixtures in kinetic and mechanical equilib- non-conservative terms ϒ k in the phasic energy equations. In the
rium derived by Saurel et al. (2007) are commonly accepted as the next subsection we will present a generalized HLLC-type solver
correct shock conditions for the non-conservative pressure equilib- that takes into account the non-conservative products.
rium model (11) (for N = 2), since they have been validated over a The simple HLLC-type solver consists of three waves W l , l =
large set of experimental data (cf. also e.g. Petel and Jetté, 2010). 1, 2, 3, moving at speeds
These relations allow the construction of an (assumed) exact solu-
tion to the pressure equilibrium model in the presence of shocks s1 = S , s2 = S  , and s3 = Sr , (35)
(Petitpas et al., 2007), and hence a solution to the parent multi-
which separate four constant states q , q , qr and qr . In the fol-
phase model (1) with instantaneous mechanical relaxation. Even
lowing we will indicate with ( · ) and ( · )r quantities correspond-
with the knowledge of shock conditions, the design of efficient
ing to the states q and qr , respectively. Moreover, we will indicate
shock-capturing diffuse-interface numerical methods able to cor-
with ( · ) and ( · )r quantities corresponding to the states q and
rectly compute shocks in multiphase mixtures is still an open chal-
qr adjacent, respectively on the left and on the right, to the mid-
lenge, cf. for instance the methods in Petitpas et al. (2007), Saurel
dle wave propagating at speed S . With this notation, the waves of
et al. (2009), and Abgrall and Kumar (2014).
the HLLC solver are
In the present work for the approximation of the non-
conservative Eq. (28) we propose HLLC-type Riemann solvers that
W 1 = q − q , W 2 = qr − q , and W 3 = qr − qr . (36)
are extensions to the multiphase case of the simple HLLC-type
solver illustrated in Pelanti and Shyue (2014b) and of the Suliciu- The middle states q , qr are defined so as to satisfy the following
type solver developed in De Lorenzo et al. (2018) for the two- Rankine–Hugoniot conditions, based on the conservative portion of
phase case. The simple HLLC-type solver of Pelanti and Shyue the system:
(2014b) omits the discretization of the non-conservative terms ϒ k
  ξ
in the phasic energy equations. The Suliciu-type solver proposed in f (ξ ) q − f (ξ ) (q ) = S q(ξ ) − q( ) , (37a)
De Lorenzo et al. (2018) can be considered as a generalized HLLC-
type method that accounts for the discretization of these non-
conservative products. This solver also includes the simple solver
ξ
of Pelanti and Shyue (2014b) for a special choice of the relax- f (ξ ) (qr ) − f (ξ ) (qr ) = Sr qr( ) − qr (ξ ) , (37b)
ation parameters. For the two-phase case we have numerically in-
vestigated different solvers with different treatments of the non-
   
conservative terms, including the Suliciu-type solver, a Roe-type f (ξ ) (qr ) − f (ξ ) q = S qr (ξ ) − q(ξ ) , (37c)
solver (Pelanti and Shyue, 2014a; Pelanti and Shyue, 2014b; Pelanti,
2017), and several path-conservative solvers (De Lorenzo et al.,
ξ = N, . . . , 3N. The solution structure for the advected volume
2018), following in particular the methods in Dumbser and Balsara
fractions α k simply consists of single jumps αk,r − αk, across the
(2016) and Dumbser and Toro (2011). Typically no relevant differ-
2-wave moving at speed S . Invariance of the equilibrium pressure
ences are observed between results of the various solvers, and re-
p and of the normal velocity u is assumed across the 2-wave, in
sults are found to agree with the exact solution of the pressure
analogy with the exact Riemann solution. Then the speed S is de-
equilibrium model as constructed in Petitpas et al. (2007), except,
termined as Toro (1997)
as expected, for the case of very strong shocks in genuine multi-
phase mixture regions, a type of problem which will not be con- p r − p  + ρ u  ( S  − u  ) − ρr u r ( S r − u r )
S = . (38)
sidered in the present work. We refer the reader in particular to ρ ( S  − u  ) − ρr ( S r − u r )
De Lorenzo et al. (2018) for a discussion on this topic.
To conclude this subsection, let us remark that HLLC-type Rie- A definition for the wave speeds must be provided, see e.g. Toro
mann solvers have gained increased interest in the last decade (1997) and Batten et al. (1997). For the numerical experiments pre-
for applications to multiphase compressible flow models, thanks in sented in this article we have adopted the following classical sim-
particular to the their ability to ensure positivity preservation and ple definition proposed by Davis (1988):
entropy conditions, in addition to the advantage of the inherent
representation of the intermediate contact wave. A first HLLC-type S = min (u − c , ur − cr ) and Sr = max (u + c , ur + cr ).
method for the non-conservative two-phase Baer–Nunziato equa- (39)
M. Pelanti and K.-M. Shyue / International Journal of Multiphase Flow 113 (2019) 208–230 215

The middle states are obtained as


⎛ ⎞  
α1,ι 
N
⎜ α3,ι ⎟ ∂t (ρ u ) + ∂x ρ u + 2
k = 0, (42c)
⎜ ⎟
⎜ .. ⎟ k=1
⎜ . ⎟
⎜ αN,ι ⎟
⎜ ⎟ ∂t (αk ρk Ek ) + ∂x (αk ρk Ek u + k u )
⎜ (α1 ρ1 )ι SSιι −u ι ⎟  
⎜ −S ⎟
⎜ (α2 ρ2 )ι SSιι −u ι ⎟ 
N
⎜ −S ⎟ + u(Yk ∂x  j − ∂x k ) = 0, k = 1, . . . , N, (42d)
⎜ .. ⎟
⎜ ⎟ j=1
ι ⎜ . ⎟
q =⎜ (αN ρN )ι SSιι −u ι ⎟, (40)
⎜ −S ⎟
⎜ ρι Sι −S S
Sι −uι  ⎟ ∂x k + u∂x k + Ck2 /ρ ∂x u = 0, k = 1, . . . , N, (42e)
⎜ ⎟
⎜ ⎟
⎜ (α1 ρ1 )ι SSιι −u ρ1,ι + (S − uι ) S +
E1,ι p1,ι
−S
ι  
ρ1,ι (Sι −uι ) ⎟
⎜ ⎟
⎜ ⎟ ∂xCk + u∂xCk = 0, k = 1, . . . , N. (42f)
⎜ (α2 ρ2 )ι SSιι −u ρ2,ι + (S − uι ) S + ρ2,ι (Sι −uι )
E p
−S
ι 2,ι   2,ι

⎜ ⎟
⎜ .. ⎟ The eigenvalues of this system are:
⎜ . ⎟
⎝ ⎠
(αN ρN )ι SSιι −u 
N
ρN,ι + (S − uι ) S + ρN,ι (Sι −uι )
EN,ι   N,ι p
ι Cm
−S λ˜ 1,5N = u ∓ c˜m , c˜m = , Cm = Ck2 , λ˜ 2 = . . . λ˜ 5N−1 = u.
ρ
ι = , r. Note that in the above formulas pk,ι = pι , for all k = k=1

1, . . . , N, since initial Riemann states satisfy pressure equilibrium (43)


conditions. All the characteristic fields are linearly degenerate, hence we can
easily find the exact solution of the relaxation system through the
4.1.2. A Suliciu-type solver Riemann invariants. The Suliciu Riemann solver uses this exact so-
We present in this section a Suliciu-type Riemann solver for the lution to approximate the Riemann solution of the original system.
multiphase flow model by extending the solver that we have in- The solution structure is analogous to the one of the HLLC solver
troduced in De Lorenzo et al. (2018) for the two-phase case. This and it consists of three waves separating four constant states, the
solver is based on the Suliciu relaxation Riemann solver presented left and right states and two middle states adjacent to a disconti-
in Bouchut (2004) for the Euler equations. Analogously to the case nuity moving with speed u . We will denote quantities correspond-
of the Euler equations, this Suliciu-type solver results to be equiv- 
ing to these middle states with ( · ) adding a subscript (· )Sul if
alent to the classical HLLC solver for the discretization of the con- needed to make a distinction with the previous HLLC-type solver.
servative equations and of the volume fraction equations of the Riemann invariants Across the contact discontinuity associated
multiphase system. We will show indeed that this solver defines a to the eigenvalue u:
class of HLLC-type methods that differ for the definition of some
constant parameters, which affect the discretization of the non- u = const., m = const., (44)
conservative terms. A particular choice of these parameters gives a N
where we have defined m = k=1 k . Across fields associated to
Riemann solver exactly equivalent to the simple HLLC-type method the eigenvalues u ∓ c˜m :
described above that neglects nonconservative terms. The Suliciu
solver (Bouchut, 2004) belongs to the class of relaxation Riemann αk , Yk = const., k = 1, . . . , N, (45a)
solvers (LeVeque and Pelanti, 2001), which are based on the idea
of approximating the solution of the original system by the solu-
1 k
tion of an extended system called relaxation system, which is eas- + = const., k = 1, . . . , N, (45b)
ier to solve. The latter system is assumed to relax to the original ρ Ck2
one, whose variables define the Maxwellian equilibrium. We refer
to Bouchut (2004), Jin and Xin (1995), and LeVeque and Pelanti u ∓ c˜m = const., (45c)
(2001) for details, and we just present the structure of the relax-
ation system associated to (1). Let us introduce N auxiliary relax-
ation variables k , k = 1, . . . , N, which are meant to relax toward Ck2  j − C 2j k = const., k, j = 1, . . . , N, (45d)
the partial pressures, thus at equilibrium k = αk pk , k = 1, . . . , N.
The equations governing the partial pressures,
2k
∂t (αk pk ) + u ∂x (αk pk ) + Yk ck2 ρ ∂x u = 0 , (41) Yk εk − = const., k = 1, . . . , N, (45e)
2Ck2
suggest the form of the equations for new variables k , which are
independent variables of the relaxation system. We introduce the
Ck = const., k = 1, . . . , N. (45f)
constant parameters Ck , k = 1, . . . , N, and we replace in Eq. (41) the
terms Yk ck2 ρ 2 by Ck2 , and (α k pk ) by k , k = 1, . . . , N. In order to be By using (45b) and (45c) we also deduce:
able to specify different constant Ck for the left and right wave
Ck2
structure of the Riemann problem solution, we also introduce ad- k ± u = const., k = 1, . . . , N, (46)
vection equations for Ck . The Suliciu relaxation system associated Cm
to the homogeneous portion of the system (1) in one spatial di- and by using (45d) and (45b):
mension is: 1 m
+ = const. (47)
∂t αk + u∂x αk = 0, k = 1, 3, . . . , N, (42a) ρ 2
Cm
Then, by using (46), we infer:
∂t (αk ρk ) + ∂x (αk ρk u ) = 0, k = 1, . . . , N, (42b) m ± Cm u = const. (48)
216 M. Pelanti and K.-M. Shyue / International Journal of Multiphase Flow 113 (2019) 208–230

Let us note first that (k ),r = (αk pk ),r , and (m ),r = ( pm ),r , Choice of parameters. The parameters Ck need to be chosen so

where pm = N k=1 αk pk . Moreover, since initial Riemann states are that Liu’s subcharacteristic condition (Liu, 1987) holds:
characterized by pressure equilibrium, we can write pm = p and N 2
pmr = pr . The relations (44) and (48) determine the quantities k=1 Ck
u = ur ≡ u and (m ) = (m )r ≡ m : c˜m = ≥ cf , (57)
Sul Sul ρ
ρ c˜m u + ρr c˜mr ur + p − pr where cf is the frozen speed of sound defined in (10). Hence the
u = ,
ρ c˜m + ρr c˜mr (49) simplest less dissipative definition for the parameters of the lo-
ρ c˜ p + ρ ˜
c p − ρ ρ
 r m c˜mr (ur − u )
˜
c cal right and left states would be (Ck2 ),r = (Yk ck2 ρ 2 ),r , which im-
m =  m r r mr 
.
ρ c˜m + ρr c˜mr plies (c˜m ),r = (cf ),r . However, this definition is not suitable when
,r shocks are involved in the solution structure. The idea here is to
The expression (47) determines ρSul :
consider well known robust definitions of the wave speeds used
 −1 for the HLLC solver to define first c˜m and then Ck . To begin with,
1 ρr, c˜mr, (ur − u ) ∓ ( pr − p )
ρSul
,r
= + , (50) let us propose a definition corresponding to the Davis’ wave speeds
ρ,r ρ,r c˜m,r (ρ c˜m + ρr c˜mr ) in (39). We set:
and through (45a) we can determine (ρk ),r
Sul
,r
= (Yk ),r ρSul /(αk ),r . c˜m = max(cf , (cfr + u − ur )), c˜mr = max(cfr , (cf + u − ur )),
Then we can find through (46):
(58)
(Ck )2,r
(k ) ,r
= (k ),r + 2 ( − p,r ), k = 1, . . . , N. (51) and we define Ck as:
(Cm ),r m
(Ck2 ),r = (Yk c˜k2 ρ 2 ),r , k = 1, . . . , N, (59a)
Finally (45e) determines the specific phasic internal energies
(εk ),r
Sul
. Then the intermediate states for the partial phasic ener- where

⎨(ck ),r
⎪ if (cf ),r ≥ (cf )r,
gies per unit volume can be expressed as: 2

+u − ur ,
(αk ρ ε ) ,r
= (αk ρ ) ,r
(εk ),r ( )
c˜k 2,r = (59b)
⎩(ck )r, + 2(2u − ur )(cf )r,
k k Sul k Sul
 ⎪ 2

( )
Ck2 ,r ( ) + ( u − ur ) otherwise.
+ ρSul
,r
(m − p,r )2 + 2k ,r (m − p,r ) , (52)
2(( 2
Cm) )
,r
2 (Cm ),r Another possible definition of the wave speeds is the one pro-
and the corresponding total energies are (αk Ek ),r = (αk ρk εk ),r + posed by Bouchut for the single-phase Suliciu solver (Bouchut,
Sul Sul
2004). We define:
u 2
(αk ρk ),r
Sul
. Let us also note that by using (45d) and (45e) we
2
 (c˜m ),r = (cf ),r + X,r , (60a)
obtain for the mixture specific internal energy ε = N k=1 Yk εk the
invariant: where
"  +
 2
X = β pρr r−p + u − ur
ε− m 

2
= const. . (53) if pr − p ≥ 0 :  p −p
cfr
+ ,
2Cm
Xr = β ρ c˜m + u − ur
r

Having now the intermediate states, the waves of the Suliciu-type "  +
solver are obtained as: Xr = β pρ−p r
+ u − ur
if pr − p ≤ 0 :  pr −p
cf
+ . (60b)
W 1 = q W 2 = qr 
W 3 = qr − qr X = β ρr c˜mr + u − ur
Sul − q , Sul − qSul , and Sul . (54)
and the corresponding speeds are: Then we set:

s1 = u − c˜m , s2 = u , and s3 = ur + c˜mr . (55) ( )


Ck2 ,r = (Yk ),r ((ck2 ),r + X,r
2
+ 2X,r (cf ),r )ρ,r
2
. (61)

We observe that the expressions of the invariants (44), (47), This choice of the wave speeds allows us to rigorously guaran-
(48) and (53) are identical to those of the Suliciu solver for the tee positivity preservation for the partial densities and the mix-
Euler equations with now m and Cm playing the role of the relax- ture internal energy, as long as the constant β ≥ 1 satisfies Bouchut
ation variable associated to the pressure p and the constant C = ρ c (2004):
 # 
of the single-phase case, respectively. Therefore the solution for the
∂ ∂ pm (ρ , sk , sN , αk , Yk )k=1,...,N−1
intermediate states ( · ),r of the mixture quantities of the multi- ρ
phase solver has the same form of the solution for the intermedi- ∂ρ ∂ρ
ate states of the standard single-phase Suliciu solver (see formu- #
las in Bouchut’s book (Bouchut, 2004)). It follows that u = S and ∂ pm (ρ , sk , sN , αk , Yk )k=1,...,N−1
≤β . (62)
the intermediate states for α k and the conserved quantities (par- ∂ρ
tial densities, mixture momentum, mixture total energy) are iden- Assuming a stiffened gas equation of state for each phase, we can
tical to those of the simple HLLC solver presented in the previous maxk γk +1
,r (ξ )
satisfy the condition above by defining β = . Let us recall
= q,r (ξ ) , with q,r given in (40), for the com-
2
subsection, qSul that α k , as well as Yk , are governed by advection equations, hence
ponents ξ = 1, . . . , 2N, as long as positivity is preserved also for these variables. Moreover, since, as
S = u − c˜m and Sr = ur + c˜mr . (56) we have noted above, only the intermediate states of the phasic
energies depend on the individual parameters Ck , if negative phasic
Note that the intermediate states for the conserved quantities de- energies are found for the intermediate states (see (45e)), we can
2 =
N 2
pend merely on the sum Cm k=1 Ck , and only the intermediate always redefine (Ck ),r in order to preserve positivity, still keeping
states for the phasic energies depend on the individual parameters the same values (Cm ),r . Let us finally remark that, given a defini-
Ck . The choice of Ck , k = 1, . . . , N, for a given definition of Cm de- tion of Cm , if we define
fines the partition of the phasic energies within the mixture, based
on the invariant (45d).
(Ck2 ),r = (YkCm
2
),r (63)
M. Pelanti and K.-M. Shyue / International Journal of Multiphase Flow 113 (2019) 208–230 217

then the resulting Suliciu-type solver is entirely equivalent to for k = 1, 2, . . . , N. Imposing the pressure equilibrium conditions
the simple HLLC-type solver described in the previous subsection, pk = p∗ , for all k = 1, . . . , N, at final time the phasic internal ener-
which neglects the discretization of the nonconservative terms in gies are then expressed as Ek∗ = Ek ( p∗ , (αk ρk )0 /αk∗ ). With these re-

lations the system (69) and the constraint N k=1 αk = 1 give N + 1
the phasic energy equations of the system (1). We can also esti-
mate the difference of the wave components for the phasic en- equations for the unknowns αk∗ , k = 1, . . . , N, and p∗ . For the par-
ergies for the case of the new Suliciu/HLLC-type solver based on ticular case of the SG EOS (3) the problem can be reduced to the
a given definition of Ck and the previous simple HLLC-type solver solution of a polynomial equation of degree N for the equilibrium
based on (63): pressure p∗ . In general an iterative solution procedure is needed
to solve this equation. Let us remark that for the most part of the
(αk ρk Ek ),r
Sul
= (αk ρk Ek ),r
HLLC
+ (αk ρk Ek ),r , (64)
three-phase (N = 3) flow numerical tests considered in this work
with we have two gaseous phases governed by a SG EOS with k = 0.
 2 In this particular case the polynomial equation of degree 3 for p∗
ρ,r

(Ck ),r
(αk ρk Ek ),r = (
− p ,r ) 2
− (Yk ),r . (65) reduces to a quadratic equation, whose physically admissible solu-
2(Cm
2)
,r (Cm
2)
,r tion is easily found.

4.2. Relaxation steps 4.2.2. Thermal relaxation


If thermal relaxation terms are also activated, then we con-
After solving the homogeneous system, we solve a sequence of sider the solution of the system (67), with μk j ≡ μ → +∞ for
ordinary differential equations accounting for the relaxation source all phase pairs, and ϑk j ≡ ϑ → +∞ for some desired pairs (k, j).
terms of (1). Here we assume that the characteristic time for me- Let us assume instantaneous thermal equilibrium for M phases,
chanical relaxation is much smaller than the characteristic time 2 ≤ M ≤ N, in addition to mechanical equilibrium for all phases. We
scales for heat and mass transfer (cf. for instance Kapila et al. will denote equilibrium values with the superscript ∗∗ . Then, anal-
(2001)), and we consider that thermal and chemical relaxation oc- ogously to the case of pressure relaxation, we can write (αk ρk )∗∗ =
cur under pressure equilibrium. The steps are the following, using (αk ρk )0 , k = 1, . . . , N, (ρ u )∗∗ = (ρ u )0 , E ∗∗ = E 0 , and E ∗∗ = E 0 .
here the vector notation in (28): Moreover, we write N − M equations of the form (69) with ( · )0
1. Mechanical relaxation. We solve in the limit μk j ≡ μ → +∞ the replaced by ( · )∗ and ( · )∗ replaced by ( · )∗∗ , the mechanical equi-
system of ODEs: librium conditions p∗∗ k
= p∗∗ , for all k = 1, . . . , N, and the thermal
equilibrium conditions Tk∗∗ = T ∗∗ for M phases. All these relations
∂t q = ψμ (q ). (66) give a system of algebraic equations to be solved for the equilib-
rium values αk∗∗ , p∗∗ . As for the mechanical relaxation step, the
This step drives instantaneously the flow to pressure equilib-
solution of this system of algebraic equations can be reduced to
rium, pk = p, for all k.
the solution of a polynomial equation of degree N for the pressure
2. Thermal relaxation. We solve in the limit μ → +∞, ϑk j → +∞:
p∗∗ when the SG EOS is adopted. The problem reduces further to
the solution of a quadratic equation for the case N = 3 with two
∂t q = ψμ (q ) + ψϑ (q ). (67) gaseous phases governed by SG pressure laws with k = 0.
This step drives the chosen phase pairs (k, j) to thermal equi-
librium, while maintaining pressure equilibrium. 4.2.3. Thermo-chemical relaxation
3. Chemical relaxation. We solve in the limit μ → +∞, ϑk j → If thermo-chemical relaxation is activated for the species that
+∞, and ν → +∞: may undergo liquid/vapor transition, then we solve the system of
ODEs (68) with μk j ≡ μ → +∞ for all phase pairs, ϑk j ≡ ϑ → +∞
∂t q = ψμ (q ) + ψϑ (q ) + ψν (q ). (68) for some phase pairs (k, j), and ν → +∞ for the phase pair (1,2).
In the steps 2–3 thermal relaxation can be activated for a sub- Let us assume instantaneous thermal equilibrium for M phases,
set of phase pairs (k, j), however it is always activated for the including at least the phases 1 and 2. We denote the quantities
phases of the species that may undergo phase transition if chem- at thermodynamic equilibrium with the superscript . First, we
ical relaxation (step 3) is also activated. Thermal and chemical re- can write (αk ρk ) = (αk ρk )0 for k = 3, . . . , N, ρ  = ρ 0 , (ρ u
 ) =
laxation processes are typically activated at interfaces only. Simi- (ρ u ) , E = E , and E = E . Moreover, we write N − M equations
0  0  0

lar to Le Métayer et al. (2013) and Pelanti and Shyue (2014b), the of the form (69) with ( · )0 replaced by ( · )∗∗ and ( · )∗ replaced
numerical relaxation procedures to handle infinitely fast transfer by ( · ) , the mechanical equilibrium conditions p k
= p , for all

processes are based on the idea of imposing directly equilibrium k = 1, . . . , N, the thermal equilibrium conditions Tk = T  for M
conditions to obtain a simple system of algebraic equations to be phases, and the chemical equilibrium condition g 1
= g
2
. This set
solved in each relaxation step, as we detail below. of algebraic equations can be solved for the values of the equilib-
rium pressure p , the equilibrium volume fractions αk and the
4.2.1. Mechanical relaxation equilibrium densities ρk . For the case of three-phase flow with
We consider the solution of the system (66) in the limit μ → SG EOS considered here we use a solution procedure similar to
+∞. We denote with superscript 0 the quantities at initial time, the two-phase case Pelanti and Shyue (2014b). First we reduce the
which come from the solution of the homogeneous system, and set of algebraic conditions excluding the chemical equilibrium re-
with superscript ∗ the quantities at final time, which are the quan- lation to the solution of a quadratic equation for the temperature
tities at mechanical equilibrium. First, we easily see that the ex- as a function of the equilibrium pressure, T  = T  ( p ). Then the
act solution of the system of ODEs gives (αk ρk )∗ = (αk ρk )0 , k = expression of T (p ) is introduced into the equilibrium condition
1, . . . , N, and (ρ u
 )∗ = ( ρ u
 )0 , E ∗ = E 0 , hence u
∗ = u
 0 and E ∗ = E 0 . g
1
= g2
. This gives an equation for p , which is solved by New-
We then integrate the equations for the phasic total energies by ton’s iterative method. Let us remark that a physically admissible
approximating the interface pressures pIkj with their values at solution of system (68) might not exist. In such a case we use
equilibrium p∗Ik j = p∗ . We then obtain N equations of the form the same technique that we have proposed in Pelanti and Shyue
(2014b), p. 356): we consider that the species that may undergo
(αk Ek )∗ − (αk Ek )0 = (αk Ek )∗ − (αk Ek )0 = −p∗ (αk∗ − αk0 ) (69) transition consists almost entirely of the phase (liquid or vapor)
218 M. Pelanti and K.-M. Shyue / International Journal of Multiphase Flow 113 (2019) 208–230

that has higher entropy, hence we fix the volume fraction of the


negligible phase to a small tolerance (for instance = 10−8 ), and
this new equation for one volume fraction replaces the equation
g
1
= g
2
in the algebraic system that we have defined with pres-
sure and temperature equilibrium conditions. Among the admissi-
ble solutions of this new system we select the solution that max-
imizes the total entropy. Note that again the problem reduces fur-
ther to the solution of a quadratic equation for the case N = 3 with
two gaseous phases governed by SG pressure laws with k = 0, as
in the experiments with phase transition considered here.

5. Numerical experiments

We present in this section numerical results for test problems


involving three-phase flows (N = 3). All the computations have
Fig. 2. Exact solution for a mechanical equilibrium three-phase shock tube prob-
been performed with the second-order wave propagation algo- lem; the density contour plot in the x–t plane is shown. The horizontal red dashed
rithm with the simple HLLC-type solver described in Section 4.1.1. line plotted in the graph corresponds to the time t = 0.12 s of the snapshots of the
We report results obtained with the generalized HLLC-type Rie- solution shown in Fig. 3. (For interpretation of the references to colour in this figure
mann solver described in Section 4.1.2 based on the Suliciu relax- legend, the reader is referred to the web version of this article.)
ation system only for the first two one-dimensional experiments
Table 1
since for the set of tests considered in this work no relevant dif-
Equation of state parameters employed in Experiment 5.1.2.
ferences are observed between the two solvers.
k (phase) γ ϖ [MPa] η [kJ/kg] η [kJ/(kg K)] κ v [J/(kg K)]
5.1. Test problems with no phase transition 1 (carbon dioxide) 1.03 13.47 0 0 3764
2 (water) 2.85 833.02 0 0 1458
To begin with, we consider test problems without mass transfer 3 (methane) 1.23 10.94 0 0 2382
processes (no chemical relaxation step).
Experiment 5.1.1. Our first numerical example concerns a
three-phase flow problem simulated by Billaud Friess and Kokh agree well with the exact solution. Let us also remark that for this
(2014) by an extended multicomponent formulation of the two- test problems, and more generally for multifluid problems involv-
phase transport five-equation model presented previously in ing defined interfaces between nearly pure inert phases, solutions
Allaire et al. (2002). Our aim here is to verify that our computed to the multiphase pressure-equilibrium model (11) are similar to
solution is an accurate approximation of the exact solution of solutions to the multicomponent extended five-equation model of
the multiphase flow model with instantaneous pressure relaxation. Billaud Friess and Kokh (2014). Hence our results here are sim-
This test involves three fluid phases in mechanical equilibrium in ilar to those in Billaud Friess and Kokh (2014). Let us remark
a one-dimensional domain [0,1] m. The phases are all modeled by however that the five-equation model of Billaud Friess and Kokh
the ideal polytropic gas law with the ratio of the specific heats de- (2014) cannot describe cavitation processes. We refer the reader
fined by γ1 = 1.6, γ2 = 2.4, and γ3 = 1.4. Initially, the velocity is to Murrone and Guillard (2005) for some comparison between the
zero, and the phasic densities are ρ1 = 1 kg/m3 , ρ2 = 0.125 kg/m3 , five-equation transport model (Allaire et al., 2002) and the five-
and ρ3 = 0.1 kg/m3 in the entire domain. There are two initial dis- equation pressure-equilibrium (Kapila et al., 2001) model.
continuities located at x = 0.4 m and x = 0.6 m that separate the Experiment 5.1.2. We now consider a three-phase (CO2 –water–
domain into three parts with the remaining state variables in each methane) flow problem simulated by Morin et al. (2009), where
region set as an extended four-equation model is used for the numerical ap-
⎧ −6 −6
 proximation. Here the three fluid components are assumed to be
⎪ if x ∈ [0, 0.4 ) m,
⎨1 Pa, 1 − 2 · 10 , 10  both in mechanical and thermal equilibrium. Hence this test gives
( p, α1 , α2 ) = 0.1 Pa, 10−6 , 1 − 2 · 10−6 if x ∈ [0.4, 0.6 ) m, us an example to verify our algorithm with the activation of both

⎩ −6 −6
 the mechanical and the thermal relaxation procedure. Initially, the
0.1 Pa, 10 , 10 if x ∈ [0.6, 1] m.
temperature is uniform and equal to T = 310 K throughout the do-
Fig. 2 shows the density contours of the exact solution of main, corresponding to the interval [0,1] m. We set the pressure
the three-phase pressure equilibrium model for this problem in p = 1.5 MPa and the velocity u = 12 m/s in a region where x ∈ [0,
the x − t plane up to time t = 0.14 s. This exact solution has 0.5] m, and p = 0.9 MPa, u = 0 m/s in a region where x ∈ [0.5, 1] m.
been computed based on the algorithm described in Petitpas et al. With these data the phasic density ρ k , k = 1, 2, 3, can be written
(2007). From the graph, it is easy to see that the initial disconti- as a function of p and T in each region, by using (3b) with the
nuity at x = 0.4 m originates a leftward-going rarefaction wave, a material-dependent parameters shown in Table 1. Here a uniform
rightward-going contact discontinuity, and a shock wave. As time composition of the media is used with volume fractions α1 = 0.9,
progresses, the shock wave moves toward the material interface α2 = 0.04, and α3 = 0.06 in each portion of the domain.
at x = 0.6 m, and then collides, yielding a new set of waves from Second-order numerical results obtained by using the HLLC-
the collision. Further collisions then occur. We compute the nu- type solver and the Suliciu-type solver with 10 0 0 grid cells and
merical solution for this test by employing both the simple HLLC- Courant number = 0.5 are shown in Fig. 4 at time t = 1.6 ms.
type solver described in Section 4.1.1, and the Suliciu-type solver The exact solution of this problem is also displayed for compar-
(generalized HLLC-type solver) presented in Section 4.1.2. We use ison. It is easy to observe that our computed solutions are in
10 0 0 grid cells, Courant number = 0.5, and we apply second-order good agreement with the exact solution, composed of a leftward-
corrections with the minmod limiter. Numerical results are dis- going rarefaction wave, a rightward-going contact discontinuity
played in Fig. 3, together with the exact solution for this prob- and a rightward-going shock wave. Again, results obtained with
lem. We observe that results obtained by the HLLC-type solver and the HLLC-type scheme and with the Suliciu-type scheme overlap.
by the Suliciu-type solver are not distinguishable, and they both We note that the exact solution computed here is based on the al-
M. Pelanti and K.-M. Shyue / International Journal of Multiphase Flow 113 (2019) 208–230 219

Fig. 3. Second-order results for the three-phase shock tube problem (Experiment 5.1.1). From left to right and from top to bottom: mixture density, velocity, pressure and
volume fraction α 2 at time t = 0.12 s obtained with the HLLC-type solver (solid blue line) and the Suliciu-type solver (dashed green line). Results computed with the two
solvers overlap. The thin solid red line indicates the exact solution. (For interpretation of the references to colour in this figure legend, the reader is referred to the web
version of this article.)

$ 
ρ01 , ρ02 , ρ03 , p0 , 10−8 , 10
−8 −8
gorithm proposed in Petitpas et al. (2007) for a mechanical equilib- if r < r0 ,
= 
rium flow with a modification to include the thermal equilibrium ρ01 , ρ02 , ρ03 , p0 , 10 , 1 − 2 · 10−8 if r0 ≤ r < r1 ,
condition in the iterative step of the method.
Experiment 5.1.3. As a first test problem in two dimensions, while outside the bubble we have
 
we are interested in a three-phase (air–R22–He) shock-bubble in- (ρ1 , ρ2 , ρ3 , p, α1 , α2 ) = ρ01 , ρ02 , ρ03 , p0 , 1 − 2 · 10−8 , 10−8 .
teraction problem studied numerically by Billaud Friess and Kokh
Behind the shock, we set
(2014). The domain is a closed tube [0, 445] × [0, 89] mm2 , and
solid wall boundary conditions are imposed on all the four sides. (ρ1 , ρ2 , ρ3 , u, v, p, α1 , α2 )

= 1.686 kg/m , ρ02 , ρ03 , −113.5 m/s, 0, 1.59
We set a leftward-going planar Mach 1.22 shock wave in air ini- 3

tially located at x = 275 mm, traveling toward a stationary cylin- 5 −8 −8



drical helium bubble with a R22 shell surrounding it. Here the ra- ×10 Pa, 1 − 2 · 10 , 10 .
dius of the helium bubble is r0 = 15 mm, centered at (x0 , y0 ) = Fig. 5 shows pseudo-color plots of the density at six different times
(225, 0 ) mm, and the thickness of the R22 cylindrical shell is H0 = t = 0, 120, 480, 780, 1020 μs obtained by using the second-order
10 mm. We assume that all the fluid phases, i.e., air, R22 (chlorod- HLLC-type wave propagation method with instantaneous pressure
ifluoromethane), and helium, are modeled by the ideal polytropic relaxation on a 1250 × 250 mesh. Comparing our numerical solu-
gas law and we set (γ , ρ0 )1 = (1.4, 1.225 kg/m ), (γ , ρ0 )2 =
3
tion results with those reported in Billaud Friess and Kokh (2014,
(1.249, 3.863 kg/m3 ), and (γ , ρ0 )3 = (1.6, 0.138 kg/m3 ), respec- cf. Fig. 20 in particular), we observe qualitatively good agreement
tively. The state variables in the region ahead of the shock wave on the geometric structure of the R22-helium bubble at times up
are assumed to correspond to atmospheric condition with pressure to t = 480 μs. At the other times t = 780 and 1020 μs noticeable
p0 = 1.01325 × 105 Pa, and inside the R22–helium bubble with differences can be seen on the outer ring of the helium bubble,
r 2 = (x − x0 )2 + (y − y0 )2 and r1 = r0 + H0 , we set which consists mostly of the phase of R22. We expect that these
visible differences in the results are not related to the differences
in the underlying flow models, as we mentioned above. To ver-
(ρ1 , ρ2 , ρ3 , p, α1 , α2 ) ify our solution, we have performed the same test by employing
220 M. Pelanti and K.-M. Shyue / International Journal of Multiphase Flow 113 (2019) 208–230

Fig. 4. Second-order results for the three-phase CO2 –water–methane Riemann problem (Experiment 5.1.2). From left to right and from top to bottom: mixture density,
velocity, pressure and volume fraction at time t = 0.16 ms obtained with the HLLC-type solver (solid blue line) and the Suliciu-type solver (dashed green line). Results
computed with the two solvers overlap. The thin solid red line indicates the exact solution. (For interpretation of the references to colour in this figure legend, the reader is
referred to the web version of this article.)

a numerical model based on the multicomponent extended five- Table 2


Equation of state parameters employed in Experiment 5.2.1.
equation model (cf. Billaud Friess and Kokh, 2014) in the simula-
tion. Numerical results of this run are presented in Fig. 6 at three k (phase) γ ϖ [Pa] η [kJ/kg] η [kJ/(kg K)] κ v [J/(kg K)]
different times t = 480, 780, and 1020 μs. Agreement of the these 1 (vapor) 1.43 0 2030 −23.4 1040
numerical results with those shown in Fig. 5 can be easily ob- 2 (liquid) 2.35 109 −1167 0 1816
served. 3 (gas) 1.4 0 0 0 718

5.2. Test problems with phase transition


0.5. We perform the simulation with different levels of activation
We now present several numerical experiments involving phase of instantaneous relaxation processes:
transition. For these tests we activate chemical relaxation.
(i) only mechanical relaxation (p-relaxation);
Experiment 5.2.1. We perform a test that is similar to the
(ii) mechanical relaxation for all the three phases and thermal
two-phase cavitation tube experiment presented in Saurel et al.
relaxation for the liquid–vapor pair only (pT(lv)-relaxation);
(2008) and Pelanti and Shyue (2014b). We consider a tube filled
(iii) mechanical and thermal relaxation for all the phases (pT-
initially with liquid water with a uniformly distributed small
relaxation);
amount of water vapor with volume fraction αwv = 10−2 , and a
(iv) mechanical relaxation for all the phases and thermal
small amount of air (non-condensable gas) with volume fraction
and chemical relaxation for the liquid–vapor pair (pT(lv)g-
αg = 10−1 . Air is modeled as an ideal gas, see Table 2 for the
relaxation);
material-dependent parameters of the equations of state employed
(v) mechanical and thermal relaxation for all the phases
in this test. The initial pressure is p0 = 105 Pa, and the initial den-
and chemical relaxation for the liquid–vapor pair (pTg-
sities correspond to the temperature T0 = 354 K. A velocity dis-
relaxation).
continuity is set at initial time at the middle of the tube, with
u0 = −20 m/s on the left and u0 = 20 m/s on the right. We use For the two cases (iv) and (v) of this cavitation tube test
30 0 0 grid cells over the interval [0,1] m, and Courant number = thermo-chemical relaxation is activated if the liquid temperature
M. Pelanti and K.-M. Shyue / International Journal of Multiphase Flow 113 (2019) 208–230 221

Fig. 5. Results for a Mach 1.22 shock wave in air interacting with a helium bubble with a R22 shell surrounding it (Experiment 5.1.3). Pseudo-color plots of the density at
six different times t = 0, 120, 480, 780, and 1020 μs obtained with the numerical multiphase flow model with instantaneous pressure relaxation (1250 × 250 mesh). (For
interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

is greater than the saturation temperature at the local pressure, two evaporation waves develop, causing an increase of the vapor
Tliquid > Tsat (p). Second-order results are displayed at time t = 6 ms mass fraction in the middle region. Note that in these cases the
in Fig. 7 for the pressure, the velocity, the total gaseous vol- pressure decreases in the cavitation zone until the saturation value
ume fraction αwv + αg , and the vapor mass fraction. In the same is reached, whereas the pressure reaches much lower values here if
Fig. 7 we also show the phasic temperatures for the two cases of p- mass transfer is not activated. By inspecting the results we observe
relaxation and pTg-relaxation. In all the cases we observe two rar- that the speed of the leading edges of the two rarefactions de-
efactions propagating in opposite directions that produce a pres- creases for any additional instantaneous thermal equilibrium pro-
sure decrease in the middle region of the tube, and, correspond- cess that we activate in the computation, consistently with the
ingly, an increase of the total gaseous component. For the cases sub-characteristic property demonstrated theoretically for the hi-
with activation of mass transfer, i.e., pT(lv)g- and pTg-relaxation, erarchy of relaxed models in Section 3. Let us note that chemical
222 M. Pelanti and K.-M. Shyue / International Journal of Multiphase Flow 113 (2019) 208–230

Fig. 6. Results obtained using an extended transport five-equation model for the test problem considered in Fig. 5. Pseudo-color plots of the density are shown at three
different times t = 480, 780, and 1020 μs (1250 × 250 mesh). (For interpretation of the references to colour in this figure legend, the reader is referred to the web version
of this article.)

relaxation is not active here around the rarefaction fronts since, as of the wall. In these plots we also show the results of the com-
indicated above, mass transfer in this test is activated only in re- putation done with no activation of mass transfer, in which case
gions where Tliquid > Tsat (p). cavitation is a mechanical process only, and mass fractions do not
Experiment 5.2.2. We now perform a two-dimensional experi- vary. For the two computations, with and without mass transfer,
ment. In this test we simulate a cylindrical underwater explosion we observe some common features: a pressure peak of the same
(UNDEX) close to a rigid surface. Following Xie et al. (2006), we magnitude corresponding to the instant at which the circular shock
consider an initial bubble of highly pressurized gas (combustion hits the wall, the drop of the pressure and consequent growth of
products) surrounded by liquid water and located near an upper a gaseous region in this zone, which then disappears due to the
flat wall. Three fluid components are involved in this problem: subsequent recompression. Later, further weaker processes of cav-
liquid water, water vapor, and combustion gases. We use a grid itation formation and collapse are observed. The behavior of the
of 481 × 280 cells over the domain [−0.6, 0.6] × [−0.7, 0] m2 . The vapor volume fraction is also qualitatively and quantitatively anal-
bubble initially is located at (xb , yb ) = (0, −0.22 ) m, and it has ra- ogous in the two cases, with or without liquid–vapor transition.
dius rb = 0.05 m. Inside the bubble we set initially a pressure p = However, similar to what observed for the one-dimensional cavita-
8290 × 105 Pa, a gas density ρg = 1400 kg/m3 , and volume frac- tion tube experiment (Experiment 5.2.1), the minimum pressure in
tions αwl = αwv = 10−8 for the water phases. Outside the bubble the cavitation region has very different order of magnitude for the
we set p = 105 Pa, T = 303 K, and the volume fractions αwv = 10−4 two computations, as we can observe from the zoom of the pres-
and αg = 10−7 , for water vapor and gas, respectively. The EOS pa- sure history in time in Fig. 9 (upper right plot). In fact, the pressure
rameters for water are those in Table 2. An ideal gas law is used for continues to decrease until very low values if no phase transition
the combustion gases, with γg = 2. In this test we activate thermal is activated, whereas it decreases until the saturation value other-
and chemical relaxation for the liquid–vapor water pair. For com- wise. For instance, at the time t = 0.529 ms corresponding to the
parison, we have also run a simulation with no thermo-chemical maximum value at the wall of the vapor volume fraction, we ob-
relaxation, this allowing us to highlight the effect of mass transfer tain a pressure p = 28 Pa with no mass transfer and p = 4417 Pa
processes. This explosion problem is characterized by a complex with mass transfer. In the literature these type of UNDEX problems
pattern of shocks and rarefaction waves (Xie et al., 2006), and the are typically simulated by simpler single-fluid models (Liu et al.,
likely occurrence of creation and collapse of vapor cavities in the 20 04; Xie et al., 20 06; Zhu et al., 2012; Wang et al., 2014), or by
liquid region close to the wall, due to the strong rarefactions and two-phase flow models (Daramizadeh and Ansari, 2015; Ma et al.,
subsequent recompression. We show in Fig. 8 pseudo-color plots of 2015; Haimovich and Frankel, 2017) that are only able to describe
the pressure at four different times obtained by activating thermo- mechanical cavitation processes, that is growth/collapse of gas cav-
chemical relaxation. At t = 0.075 ms (upper left plot) we can ob- ities due to pressure variations, with no liquid–vapor transition. In
serve the circular shock created by the explosion. At t = 0.2 ms contrast, our three-phase flow model allows a more accurate de-
(upper right plot) this shock has reflected from the wall, at time scription of the thermodynamics of cavitation processes, which in-
t = 0.35 ms (lower left plot) a low pressure cavitation region has volve liquid–vapor phase change. We notice that this is critical for
begun to develop close to the rigid surface, and this region is more an accurate prediction of the pressure field on the wall adjacent to
extended at t = 0.5 ms (lower right plot). The thick solid circle cavitation regions. We have to remark however an important limit
line indicates the water/bubble interface. In Fig. 9 we display the of our current numerical model in relation to the use of the sim-
history in time of the pressure, the water vapor volume fraction ple stiffened gas EOS. This equation of state can be considered a
and the water vapor mass fraction at the point (0,0) at the center linearized version of the Mie–Grüneisen EOS around a reference
M. Pelanti and K.-M. Shyue / International Journal of Multiphase Flow 113 (2019) 208–230 223

Fig. 7. Numerical results for the three-phase cavitation tube test (Experiment 5.2.1). First and second row: results for the pressure, velocity, total gas volume fraction, and
vapor mass fraction for the various relaxation cases. Third row: temperature of the three phases (liquid, vapor, air) for the p-relaxation case (left), and for the pTg-relaxation
case (right).

thermodynamic state (Menikoff and Plohr, 1989). While numeri- instead of an upper rigid solid surface. Many authors in the litera-
cally convenient, this simple EOS is not suited for accurate predic- ture have simulated numerically this type of problem, e.g. Liu et al.
tions of flow conditions with significant variations of the thermo- (2001), Xie et al. (2007), Yeom (2015), Daramizadeh and Ansari
dynamic variables, as we have in the UNDEX problems simulated (2015), and Haimovich and Frankel (2017). However, as for the
here. The current numerical model with the SG EOS is nonethe- previous test, simulations presented in the literature are typically
less able to qualitatively describe the relevant physical phenomena based on simple single-fluid models or two-phase models that do
and it allows us to better understand the effect of the activation of not account for liquid–vapor transition. The setup of this problem
mass transfer processes. has been chosen in order to be able to make qualitative compar-
Experiment 5.2.3. We finally simulate an underwater explosion isons with the laboratory underwater explosion test of Kleine et al.
problem similar to the previous one, but here we set a free surface (2009) (simulated also for instance in Daramizadeh and Ansari,
224 M. Pelanti and K.-M. Shyue / International Journal of Multiphase Flow 113 (2019) 208–230

Fig. 8. Numerical results for the UNDEX experiment near a rigid surface (Experiment 5.2.2). Pressure field at times t = 0.075, 0.2, 0.325, 0.5 ms computed by the HLLC-type
scheme with activation of thermo-chemical relaxation for the liquid–vapor pair. The thick solid line (magenta color) indicates the water/bubble interface. (For interpretation
of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Fig. 9. Numerical results for the UNDEX experiment near a rigid surface (Experiment 5.2.2). History in time at the point (0,0) at the center of the wall (solid line: with mass
transfer, dashed line: no mass transfer). Top plots: Pressure, with a zoom in the cavitation region shown in the right plot. Bottom plots: vapor mass fraction (left) and vapor
volume fraction (right).
M. Pelanti and K.-M. Shyue / International Journal of Multiphase Flow 113 (2019) 208–230 225

Fig. 10. Numerical results for the UNDEX experiment near a free surface (Experiment 5.2.3). Pressure field (left column) and water vapor mass fraction (right column) at
times t = 14, 28, 42 μs computed by the HLLC-type scheme with activation of thermo-chemical relaxation. The thick solid lines (magenta color) indicate the free surface and
the water/bubble interface. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

2015). As before, we consider an initial bubble of highly pressur- Ansari, 2015). We use a grid of 669 × 600 cells over the domain
ized gas surrounded by liquid water. Here the bubble is located [−0.09, 0.09] × [−0.12, 0.042] m2 . The bubble initially is located at
near an air/water interface. Four fluid components are involved (xb , yb ) = (0, −0.034 ) m, and it has radius rb = 0.0035 m. The free
in this problem: liquid water, water vapor, combustion gases (the surface is located at y = 0 m. Inside the bubble we set initially
bubble), and air (region above the free surface). Since here we a pressure p = 8290 × 105 Pa, a gas density ρg = 1400 kg/m3 ,
focus on phase transition phenomena triggered by the explosion, and volume fractions αwl = 10−8 and αwv = 10−7 for the liquid
we make the simplifying assumption that the bubble consists of and vapor phases of water. Outside the bubble, both below and
air at high pressure, so that we need to consider three phases above the free surface, we set p = 105 Pa and T = 298 K. In
only instead of four. Note that other authors choose to model the air region above the free surface we set water volume frac-
the bubble by a high pressure liquid region (Daramizadeh and tions αwl = 10−8 (liquid) and αwv = 10−7 (vapor), while below
226 M. Pelanti and K.-M. Shyue / International Journal of Multiphase Flow 113 (2019) 208–230

Fig. 11. Numerical results for the UNDEX experiment near a free surface (Experiment 5.2.3). Results for the mixture density (left) and the velocity field (right) at t = 42 μs.
The thick solid lines (magenta color) indicate the free surface and the water/bubble interface. (For interpretation of the references to colour in this figure legend, the reader
is referred to the web version of this article.)

the free surface we take αwv = 10−4 and αg = 10−7 , for water to show that the behavior of the wave speed predicted numerically
vapor and air (inert gas), respectively. The SG EOS parameters is consistent with our theoretical findings on the sub-characteristic
for water for this test are (γ , , η, η , κv ) l = (2.62, 9058.29 × interlacing of the characteristic speeds for the hierarchy of relaxed
105 Pa, −1.150975 × 106 kJ/kg, 0, 1606.97 J/(kgK ) ) for the models. The numerical results, finally, show the efficiency of the
liquid phase, and (γ , , η, η , κv ) v = (1.38, 0, 2.060759 × numerical method in modelling complex wave patterns, shocks
106 kJ/kg, −27.225 kJ/(kgK ), 1192.51 J/(kgK ) ) for the vapor phase and interfaces in problems with thermal and mass transfer pro-
(Le Martelot et al. (2014)). An ideal gas law is used for air with cesses where the dynamical appearance of vapor cavities and evap-
γg = 1.4. Heat and mass transfer processes are activated in this oration fronts in a liquid is coupled to the dynamics of a third non-
test. The characteristic features of this explosive event near a free condensable gaseous component governed by its own equation of
surface are similar to those of an event close to a wall: a circular state. An example of application illustrated in the present work is
blast wave is produced by the highly pressurized gas bubble, and the simulation of underwater explosions close to a rigid wall or
it interacts with the air/water interface, leading to a reflected a free surface. In these problems a highly pressurized gas bubble
expansion wave. Due to the consequent pressure decrease, a cav- triggers cavitation processes in a liquid. Another application exam-
itation region is formed just below the free surface. Our numerical ple is the simulation of high speed cavitating underwater systems,
model allows us to describe the liquid–vapor transition processes considered for instance in Petitpas et al. (2009). Some limits of the
occurring in this region. We show in Fig. 10 pseudo-color plots numerical model presented in this work are the choice of a sim-
of the pressure and the vapor mass fraction at three different ple stiffened gas equation of state and the assumption of instan-
times, chosen to have snapshots comparable to the three frames taneous heat and mass transfer. Future work will be dedicated to
of the Schlieren visualization of the experiment of Kleine et al. the extension of the model to more complex and general equations
(2009) (also reported in Daramizadeh and Ansari, 2015). Moreover, of state, such as the IAPWS Industrial Formulation for Water and
in Fig. 11 we also display the mixture density and the velocity field Steam (Wagner et al., 20 0 0), and to finite-rate thermo-chemical re-
at final time. The latter plot in particular allows us to observe the laxation processes.
transmitted shock wave in air, which is too weak to be seen in the
plots of the pressure field. Despite the several simplifications in the
Acknowledgments
model our simulation is able to reproduce qualitatively the main
physical processes observed experimentally (Kleine et al., 2009).
The first author (M. Pelanti) was partially supported by the
French Government Directorate for Armament (Direction Générale
6. Conclusions de l’Armement, DGA) under grant no. 2012.60.0011.00.470.75.01,
and also partially supported by the Research Council of Norway
We have presented a numerical model for multiphase com- grant no. 234126/E30 in the framework of the SIMCOFLOW Project.
pressible flows involving the liquid and vapor phases of one The second author (K.-M. Shyue) was supported in part by the
species and one or more inert gaseous phases, extending the two- Ministry of Science and Technology of Taiwan under grant no. 103-
phase flow model that we have introduced in Pelanti and Shyue 2115-M-002-011-MY2. M. Pelanti is grateful to Tore Flåtten and
(2014b). The model includes mechanical, thermal and chemical re- Gaute Linga for very fruitful discussions.
laxation processes. We have also rigorously derived the associ-
ated pressure-relaxed model by asymptotic techniques, and car-
ried an analysis of the characteristic speeds of the hierarchy of Appendix A. Derivation of the relaxed pressure-equilibrium
relaxed models associated to the parent model. The multiphase model
equations are solved by a mixture-energy-consistent finite volume
wave propagation scheme based on an original HLLC/Suliciu-type In this section we derive the p-relaxed model in (11) from the
Riemann solver, combined with simple and robust procedures for multiphase model in (1). For simplicity, we shall consider the one-
the stiff relaxation terms. Sample one-dimensional tests show the dimensional case d = 1. We follow in particular the technique of
agreement of the computed numerical solution with the exact so- Murrone and Guillard (2005) (see also Chen et al., 1994). First, we
lution for three-phase Riemann problems with mechanical and write the system (1) in one dimension in terms of the vector of
thermal equilibrium. A cavitation tube experiment also allows us primitive variables w ∈ R3N as:
M. Pelanti and K.-M. Shyue / International Journal of Multiphase Flow 113 (2019) 208–230 227

1 Multiplying the above equation by P, by using (72), and by neglect-


∂t w + A(w )∂x w = (w ) + (w ), (70a)
ing terms of order τ , we obtain the reduced model system:
τ
where τ ≡ 1/μ, and
⎡⎤ ⎡ u 0 ... 0 0 0 ... 0 0 0 0 ... 0

α1
⎢ 0 u ... 0 0 0 ... 0 0 0 0 ... 0 ⎥
⎢ α3 ⎥ ⎢ . ⎥
⎢ . ⎥ ⎢ .. .. .. .. .. .. .. .. .. .. .. .. ..

⎢ . ⎥ ⎢ . . . . . . . . . . . . ⎥
⎢ . ⎥ ⎢ 0 ⎥
⎢αN ⎥ ⎢ 0 0 u 0 0 ... 0 0 0 0 ... 0

⎢ ⎥ ⎢ 0 0 ... 0 u 0 ... 0 ρ1 0 0 ... 0 ⎥
⎢ ρ1 ⎥ ⎢ 0 ⎥
⎢ ⎥ ⎢ 0 ... 0 0 u ... 0 ρ2 0 0 ... 0 ⎥
⎢ ρ2 ⎥ ⎢ . ⎥
⎢ . ⎥ ⎢ .. .. .. .. .. .. .. .. .. .. .. .. ..⎥
w=⎢ ⎥
⎢ .. ⎥, A = ⎢
. . . . . . . . . . . .⎥, (70b)
⎢ρ ⎥ ⎢ 0 0 ... 0 0 0 ... u ρN 0 0 ... 0 ⎥
⎢ N⎥ ⎢p − p p3 − p2 pN − p2 α1 α2 αN ⎥
⎢u⎥ ⎢ 1 2
... ... ... ⎥
⎢ ⎥ ⎢ ρ ρ ρ
0 0 0 u
ρ ρ ρ⎥
⎢ p1 ⎥ ⎢ ⎥
⎢ ⎥ ⎢ 0 0 ... 0 0 0 ... 0 ρ 2
1 c1 u 0 ... 0 ⎥
⎢ p2 ⎥ ⎢ ⎥
⎢ . ⎥ ⎢ 0 0 ... 0 0 0 ... 0 ρ 2
2 c2 0 u ... 0 ⎥
⎣ .. ⎦ ⎢ . ⎥
⎣ . .. .. .. .. .. .. .. .. .. .. ..
.
.. ⎦
. . . . . . . . . . . .
pN
0 0 ... 0 0 0 ... 0 ρN cN2 0 0 ... u
⎡ N ⎤
j=1 ( p1 − p j )
⎡ ⎤
 0
⎢ N
j=1 ( p3 − p j )

⎢ ⎥ ⎢ 0 ⎥
⎢ .. ⎥ ⎢ .. ⎥
⎢ . ⎥ ⎢ ⎥
⎢ N ⎥ ⎢ . ⎥
⎢ ( p − p ) ⎥ ⎢ 0 ⎥
⎢ j=1
ρ 1 N
N j ⎥ ⎢ ⎥
⎢ − ( p − p ) ⎥ ⎢ M ⎥
⎢ α1  j=1 1 j ⎥ ⎢ α1 ⎥
⎢ ρ2
− α2 N
j=1 ( p2 − p j )
⎥ ⎢ − M
α2 ⎥
⎢ ⎥ ⎢ ⎥
⎢ . ⎥ ⎢ .. ⎥
=⎢ .. ⎥,  = ⎢ . ⎥. (70c)
⎢  ⎥ ⎢ ⎥
⎢ ρ
− αNN N
j=1 ( pN − p j )
⎥ ⎢ 0 ⎥
⎢ ⎥ ⎢ ⎥
⎢ 0 ⎥ ⎢ N 0 ⎥
⎢ 1 N ⎥ ⎢ 1 M⎥
⎢ −α [  ( E + p ) + χ ρ ] ( p − p ) ⎥ ⎢ α1  j=1 Q1 j + (1 gI + χ1 ) α1 ⎥
⎢ 11 Nj=1 1 1 I1 j 1 1 1 j
⎥ ⎢ 2 N Q2 j − (2 gI + χ2 ) M ⎥
⎢ −α ⎥ ⎢ α2 j=1 α2 ⎥
⎢ 2 j=1 [2 (E2 + pI2 j ) + χ2 ρ2 ]( p2 − p j ) ⎥ ⎢ .. ⎥
⎢ .. ⎥ ⎣ ⎦
⎣ ⎦ .

. N N
− α1N Nj=1 [N (EN + pIN j ) + χN ρN ]( pN − p j ) αN j=1 QN j

We are interested in the behavior of the solutions of (70) in


the limit τ → 0+ (μ = τ1 → +∞). We expect that these solutions
are close to the set U = {w ∈ R3N ; (w ) = 0}. We assume that the
∂t v + PA((v ))dv ∂x v = P((v )). (74)
set of equations (w ) = 0 defines a smooth manifold of dimen-
sion L and that for any w ∈ U we know a parameterization  In the limit of instantaneous pressure relaxation we have pk = p,
(the Maxwellian) from an open subset of RL on a neighbor- ∀k = 1, . . . , N, hence the vector of the variables of the reduced
hood of w in U. For any v ∈ ⊂ RL the Jacobian matrix dv is a pressure-relaxed model is
full rank matrix, moreover, the column vectors of dv form a basis
of ker(  ((v ))) (Murrone and Guillard, 2005). Now let us define
v = [α1 , α3 , . . . , αN , ρ1 , ρ2 , . . . , ρN , u, p]T ∈ R2N+1 . (75)
the matrix C ∈ R3N×3N :
C = [d1v . . . dLv V 1 . . . V 3N−L ] (71)
Note that here L = 2N + 1. The equilibrium state (v) is defined
where d 1 , . . . , d L
v v are the column vectors of dv and by:
{V 1 , . . . , V 3N−L } is a basis of the range of   ((v)). Based on the
observations above, the matrix C is invertible. Let us now denote
 : v → (v ) = [α1 , α3 , . . . , αN , ρ1 , ρ2 , . . . , ρN , u, p, p, . . . , p]T
with P the L × 3N matrix composed of the first L rows of the in-
verse C −1 . We have also the following results (see Murrone and ∈ R 3N . (76)
Guillard, 2005):
P d v = I L and P   ((v )) = 0, (72) The Jacobian dv ∈ R3N×2N+1 of the Maxwellian is:

where IL denotes the L × L identity matrix. Now to obtain a re- ⎡ ⎤


duced pressure equilibrium model we look for solutions in the 0
form w = (v ) + τ z, where z is a small perturbation around the ⎢ I2N .. ⎥
⎢ . ⎥
equilibrium state (v). Using this into the system (70) we obtain ⎢ ⎥
⎢ 0 ⎥
d v = ⎢ ⎥. (77)
⎢ 0 ... 0 1 ⎥
⎢ .. ⎥
∂t ((v )) + A((v ))∂x ((v )) −   ((v )) z = ((v )) + O (τ ). ⎣ ..
. . ⎦
(73) 0 ... 0 1
228 M. Pelanti and K.-M. Shyue / International Journal of Multiphase Flow 113 (2019) 208–230

A basis {V 1 , . . . , V N−1 }, V k ∈ R3N , k = 1, . . . , N − 1, for the range of (11) is obtained from (74) by using the above expression of the
  ((v)) is found as matrix P and by evaluating the matrix A and the source term  in
⎡ ⎤ ⎡ ⎤ the equilibrium state (v) in (76). Let us also note that we use the
(N − 1 ) −1
relations χk = ck2 − k hk in the entries of  in (70c).
⎢ −1 ⎥ ⎢ (N − 1 ) ⎥
⎢ −1 ⎥ ⎢ −1 ⎥
⎢ ⎥ ⎢ ⎥ Appendix B. Speed of sound for the hierarchy of multiphase
⎢ .. ⎥ ⎢ .. ⎥
⎢ . ⎥ ⎢ . ⎥ flow models
⎢ ⎥ ⎢ ⎥
⎢ −1 ⎥ ⎢ −1 ⎥
⎢ − ( N − 1 ) ρ1 ⎥ ⎢ ρ1 ⎥ In this section we show the derivation of the expressions of
⎢ α1 ⎥ ⎢ α1 ⎥
⎢ ρ2 ⎥ ⎢ ρ2 ⎥ the sound speed in (14), (19), and (26), following in particular the
⎢ α2 ⎥ ⎢ α ⎥
⎢ ρ3 ⎥ ⎢ −(N −2 1 ) ρ3 ⎥ work in Flåtten et al. (2010). For the two-phase case, these for-
⎢ α3 ⎥ 2 ⎢ α3 ⎥ mulas have been derived by various authors, see especially Wood
1 ⎢
V =⎢ .. ⎥, V = ⎢ .. ⎥,
⎥ ⎢ ⎥ (1930), Stewart and Wendroff (1984), and Flåtten and Lund (2011).
⎢ . ⎥ ⎢ . ⎥
⎢ ρ N ⎥ ⎢ ρN ⎥ For flow models with N ≥ 2 phases, the expression of the speed of
⎢ αN ⎥ ⎢ αN ⎥
⎢ 0 ⎥ ⎢ 0 ⎥ sound have been studied in Flåtten et al. (2010) for the case of in-
⎢ ⎥ ⎢ ⎥ stantaneous mechanical equilibrium and the case of both mechan-
⎢ −(N − 1 ) αρ11 c12 ⎥ ⎢ ρ1 2
α1 c1 ⎥
⎢ ρ2 2 ⎥ ⎢ ρ2 2 ⎥ ical and thermal equilibrium for all the phases. The novelty here is
⎢ α2 c2 ⎥ ⎢ α2 c2 ⎥
⎢ ρ3 2 ⎥ ⎢ − (N − 1 ) αρ33 c32
⎥ the derivation of (26) for the flow model with chemical potential
⎢ α3 c3 ⎥ ⎢ ⎥ equilibrium and the generalization of (19) to the situation where
⎢ .. ⎥ ⎢ . ⎥
⎣ .
⎦ ⎣ .. ⎦ only 2 ≤ M < N phase are in thermal equilibrium. We begin by re-
ρN 2 ρN 2 calling some definitions and by writing some useful equalities:
αN cN αN cN   
⎡ −1
⎤ ∂ pk ∂ pk ρk ∂ Tk
ck2 = , k = = ,
⎢ −1 ⎥ ∂ρk sk
∂ Ek ρk
Tk ∂ρk sk
⎢ −1 ⎥  
⎢ ⎥ ∂ hk ∂ sk
⎢ .. ⎥ κ pk = = Tk , (80a)
⎢ . ⎥ ∂ Tk ∂ Tk
⎢ ⎥ pk pk
⎢ (N − 1 ) ⎥
⎢ ρ1 ⎥  
⎢ α1 ⎥ ∂ Tk Tk k ∂ρk Tk k ρk
⎢ ρ2 ⎥ = , =− . (80b)
⎢ α2 ⎥ ∂ pk ρk ck2 ∂ sk ck2
⎢ ρ3 ⎥ sk pk
⎢ α3 ⎥
. . . , V N−1 =⎢
⎢ .. ⎥.
⎥ (78) Here a common pressure will be always assumed, pk = p, ∀k =
 N
⎢ . ⎥ 1, . . . , N. Now, by recalling ρ = Nk=1 αk ρk and k=1 αk = 1, we
⎢ −(N − 1 ) αρNN ⎥
⎢ ⎥ can write:
⎢ 0 ⎥
⎢ ρ1 2 ⎥ 
N
d (αk ρk ) 
N
αk N
⎢ α1 c1 ⎥ = d ρk = ρ
1 1
dYk + dρ .
⎢ ρ2 2 ⎥ ρk ρ ρ ρ
(81)
⎢ α2 c2 ⎥ k k
⎢ ρ3 2 ⎥ k=1 k=1 k=1
⎢ α3 c3 ⎥
⎢ .. ⎥ Hence:
⎣ ⎦  
. 1
N
αk N
1
−(N − 1 ) αρNN cN2 dρ = d ρk − ρ dYk . (82)
ρ k=1
ρ k
k=1
ρ k
Note that this structure is associated to the choice of the indepen-
Then we can write, by considering ρk = ρk ( p, sk ):
dent variables in w (where in particular we have chosen α k , k = 2).
Hence we can construct the matrix C ∈ R3N×3N (71), compute the
inverse C −1 , and finally obtain the matrix P ∈ R2N+1×3N by taking
the first 2N + 1 rows of C −1 . We find:
⎡  ⎤
α1
ρ1 c12
ρ c2p Nj=2 ραj cj 2 − ρ αc12 αρ2 c2 ρ c2p − ρ αc12 αρ3 c2 ρ c2p ... − ρ αc21ραNc2 ρ c2p
⎢ j 1 1 2 2

1 1 3 3 1 1 N N

⎢ − ρ αc12 αρ3 c2 ρ c2p − ρ αc22 αρ3 c2 ρ c2p α3
ρ c2
ρ c2p Nj=1 ραcj 2 ... − ρ αc23ραNc2 ρ c2p ⎥
⎢ 1 1 3 3 2 2 3 3 3 3
j=3
j j 3 3 N N

⎢ ⎥
⎢ .. .. .. .. .. ⎥
⎢ . . . . .
N−1 α j

⎢ − ρ αc21ραNc2 ρ c2p − ρ αc22ραNc2 ρ c2p − ρ αc23ραNc2 ρ c2p αN
ρ cp ⎥
⎢ ... ρN cN2
2
j=1 ρ j c2 ⎥
⎢ 1 1 N N 2 2 N N 3 3 N N j ⎥
⎢ I2N N αj α2 α3 αN ⎥
⎢ − c12 ρ c p 2
ρ c2p ρ c2p ... ρ c2p ⎥
P=⎢ j=2 ρ j c2 c12 ρ2 c22
N
c12 ρ3 c32 c12 ρN cN2 ⎥, (79)
⎢ ⎥
1 j
α1 αj α3 αN
⎢ ρ1 c12 c22
ρ c2p − c12 ρ c2p j=1 ρ j c2 c22 ρ3 c32
ρ c2p ... c22 ρN cN2
ρ c2p ⎥
⎢ 2
j=2 j

⎢ .. .. .. .. .. ⎥
⎢ ⎥
⎢ . . . . .
N−1 α j ⎥
⎢ α1
ρ c2p α2
ρ c2p α3
ρ c2p ... − c12 ρ c p 2 ⎥
⎢ ρ1 c12 cN2 ρ2 c22 cN2 ρ3 c32 cN2 j=1 ρ j c2 ⎥
⎢ N j

⎣ 0 0 0 ... 0 ⎦
0 ... 0 ρ c2p ρα1 1c2 ρ c2p ρα2 2c2 ρ c2p ρα3 3c2 ... ρ c2p ραNNc2
1 2 3 N

where cp is the speed of sound of the pressure-equilibrium model


in (14). Finally, the reduced p-relaxed multiphase flow model in
M. Pelanti and K.-M. Shyue / International Journal of Multiphase Flow 113 (2019) 208–230 229

 
∂ρk ∂ρk 1 T ρ The speed of sound CpT,M of the model with pressure equilibrium
d ρk = dp + d sk = d p − k 2k k dsk . (83) for all the phases and temperature equilibrium for M ≤ N phases is
∂p sk
∂ sk p
ck2 ck
defined in (18). Since in this definition we consider ds = 0, dsk =
Hence we obtain: 0 for k = M + 1, . . . , N, and dYk = 0, for k = 1, . . . , N, from (89) we
   obtain the expression of the speed of sound cpT,M in (19), by using

N
αk N
Tk k N
1 also (14).
dρ = ρ dp − αk 2 dsk − ρ dY . (84)
ρ c2
k=1 k k k=1
ck
k=1
ρk k
B3. pTg-Relaxation

B1. p-Relaxation
We now consider pressure equilibrium for all the N phases,
thermal equilibrium at temperature T for M ≤ N phases, Tk = T , for
Let us consider the pressure equilibrium model (11) (p-relaxed
k = 1, . . . , M, and chemical potential equilibrium for the liquid and
model). The speed of sound is defined in (13). Since in this defi-
vapor phase pair (1,2), g1 = g2 (note that this pair is also con-
nition we consider dsk = 0 and dYk = 0, k = 1, . . . , N, from (84) we
sidered in thermal equilibrium). The speed of sound is defined
obtain the expression of the speed of sound cp in (14).
in (25). Since in this definition we consider ds = 0, dsk = 0 for
k = M + 1, . . . , N, and dYk = 0, for k = 3, . . . , N, we can write from
B2. pT-Relaxation (89):
" 
We now consider pressure equilibrium for all the N phases, and, 
αk
N
1
in addition, we assume that M phases, 2 ≤ M ≤ N, are in thermal
dρ = ρ dp + T M
k=1
ρ k ck2 k=1 C pk
equilibrium at a common temperature T, Tk = T for k = 1, . . . , M. 
Recalling the expression of the mixture specific total entropy s =
 2
N 
M−1 M
j k
k=1 Yk sk , we can write
C pk Cp j − dp
k=1 j=k+1
ρ j c2j ρk ck2

M 
N 
N 
Yk dsk = ds − sk dYk − Yk dsk . (85) M
k 1 1
&
−ρ C pk (s2 − s1 )dY1 − ρ − dY1 , (90)
k=1 k=1 k=M+1
ρc
k=1 k k
2 ρ1 ρ2
By considering T = T ( p, sk ), then we can write
  where we have used d Y2 = −d Y1 . Now we use the condition d (g1 −
∂T ∂T T k T g2 ) = 0 to find a suitable expression for dY1 in (90) as a function of
dT = dp + d sk = dp + d sk ,
∂p ∂ sk ρ c
k k
2 κ pk dp. By recalling that gk = εk + ρp − T sk and by noticing that dgk =
sk p k
1
k = 1, . . . , M. (86) ρk d p − sk d T , k = 1, 2, we can write:

Hence we obtain M − 1 equations: 1 1


d ( g2 − g1 ) = − dp − (s2 − s1 )dT . (91)
 ρ2 ρ1
1 1 k+1 k We observe that dT can be expressed as:
d sk − dsk+1 = − , k = 1, . . . , M − 1.
κ pk κ pk+1 ρk+1 ck2+1 ρk ck2 T 1 T
dT = dp + ds . (92)
(87) ρ1 c12 κ p1 1
These M − 1 equations together with Eq. (85) form a system of By using here the expression for ds1 given in (88) (with ds = 0,
M equations for the unknowns dsk , k = 1, . . . , M (the phasic en- dsk = 0 for k = M + 1, . . . , N, dYk = 0, for k = 3, . . . , N, d Y1 = −d Y2 ),
tropy differentials for the M phases in both mechanical and ther- we then obtain:
mal equilibrium). The solution of this system gives, after some ma-  
ρT 1  C pk k
M
nipulations: dT = M (s2 − s1 )dY1 + dp . (93)
   k=1 C pk
ρ ρk ck2 k=1
1 
N 
N
dsk = M C pk ρ ds − s j dY j − Y j ds j Now we use this expression for dT in (91) and we solve d (g2 −
αk ρk j=1 C p j j=1 j=M+1 g1 ) = 0 for the unknown dY1 , obtaining:
 

M
j k  
+ C pk − C dp , (88) 1 1 1 1 
M
1 
M
C pk k
ρ j c2j ρk ck2 p j dY1 = − C pk − dp.
j=1 ρ ρ2 ρ1 T ( s2 − s1 ) 2 s2 − s1
k=1
ρk ck2 k=1
for k = 1, . . . , M. Here we recall that C pk = αk ρk κ pk (heat capaci-
(94)
ties). By introducing (88) in (84) we then obtain
"  By introducing this equation in (90) and after some manipulations
αk
N
1 we obtain:
dρ = ρ dp + T M "   
k=1
ρ k ck2 k=1 C pk

N
αk 1 j
M−1
k 
M 2

dρ = ρ d p + T M −
  2
ρk ck2 C
k=1 pk
ρ j c 2
j
C pk
ρ 2
k ck
Cp j dp
k=1 k=1 j=k+1

M−1 M
j k
C pk Cp j − dp  2 ⎤⎫
ρ j c2j ρk ck2 M
kCpk M ⎬
1 1 1
k=1 j=k+1 + − − C d p⎦ . (95)
  k=1
ρk ck2 ρ2 ρ1 T (s2 − s1 ) k=1 pk

 M
k  N N
−ρ C ds − sk dYk − Yk dsk
ρ c2 pk
k=1 k k k=1 k=M+1
With this result, we finally obtain the expression of the speed of
% sound cpTg,M in (26) by using the Clausius–Clapeyron equation
N
Tk k N
1  −1
− αk 2 dsk − ρ dY . (89) dp
= ( s2 − s1 )
1 1
k=M+1
ck
k=1
ρk k dT ρ2

ρ1
. (96)
sat
230 M. Pelanti and K.-M. Shyue / International Journal of Multiphase Flow 113 (2019) 208–230

References Liu, T.G., Khoo, B.C., Yeo, K.S., 2001. The simulation of compressible multi-medium
flow. II: applications to 2D underwater shock refraction. Comput. Fluids 57,
Abgrall, R., Karni, S., 2010. A comment on the computation of non-conservative 291–314.
products. J. Comput. Phys. 229 (8), 2759–2763. Liu, T.P., 1987. Hyperbolic conservation laws with relaxation. Commun. Math. Phys.
Abgrall, R., Kumar, H., 2014. Numerical approximation of a compressible multiphase 108, 153–175.
system. Commun. Comput. Phys. 15 (5), 1237–1265. Lund, H., 2012. A hierarchy of relaxation models for two-phase flows. SIAM J. Appl.
Allaire, G., Clerc, S., Kokh, S., 2002. A five-equation model for the simulation of in- Math. 72 (6), 1713–1741.
terfaces between compressible fluids. J. Comput. Phys. 181, 577–616. Lund, H., Aursand, P., 2012. Two-phase flow of CO2 with phase transfer. Energy Pro-
Baer, M.R., Nunziato, J.W., 1986. A two-phase mixture theory for the deflagration– cedia 23, 246–255.
to-detonation transition (DDT) in reactive granular materials. Int. J. Multiphase Ma, Z.H., Causon, D.M., Qian, L., Gu, H.B., Mingham, C.G., Ferrer, P.M., 2015. A GPU
Flow 12, 861–889. based compressible multiphase hydrocode for modelling violent hydrodynamic
Batten, P., Clarke, N., Lambert, C., Causon, D., 1997. On the choice of wavespeeds for impact problems. Comput. Fluids 120, 1–23.
the HLLC Riemann solver. SIAM J. Sci. Comput. 18 (6), 1553–1570. Menikoff, R., Plohr, B.J., 1989. The Riemann problem for fluid flow of real materials.
Battistoni, M., Duke, D., Swantek, A.B., Tilocco, F.Z., Powell, C.F., Som, S., 2015. Ef- Rev. Mod. Phys. 61, 75–130.
fects on noncondensable gas on cavitating nozzles. Atomization Spray 25 (6), Morin, A., Aursand, P.K., tten, T.F., Munkejord, S.T., 2009. Numerical resolution of CO2
453–483. transport dynamics. In: SIAM Conference on Mathematics for Industry: Chal-
Billaud Friess, M., Kokh, S., 2014. Simulation of sharp interface multi-material flows lenges and Frontiers (MI09), San Francisco, CA, USA.
involving an arbitrary number of components through an extended five-equa- Murrone, A., Guillard, H., 2005. A five equation reduced model for compressible two
tion model. J. Comput. Phys. 273, 488–519. phase flow problems. J. Comput. Phys. 202, 664–698.
Bouchut, F., 2004. Nonlinear Stability of Finite Volume Methods for Hyperbolic Con- Parés, C., 2006. Numerical methods for nonconservative hyperbolic systems: a the-
servation Laws, and Well-Balanced Schemes for Sources. Birkhäuser-Verlag. oretical framework. SIAM J. Numer. Anal. 44, 300–321.
Castro, M.J., LeFloch, P., Muñoz-Ruiz, M.L., Parés, C., 2008. Why many theories Pelanti, M., 2017. Low Mach number preconditioning techniques for Roe-type and
of shock waves are necessary: convergence error in formally path-consistent HLLC-type methods for a two-phase compressible flow model. Appl. Math.
schemes. J. Comput. Phys. 227, 8107–8129. Comput. 310, 112–133.
Chen, G.Q., Levermore, C.D., Liu, T.P., 1994. Hyperbolic conservation laws with stiff Pelanti, M., Shyue, K.-M., 2014a. A mixture-energy-consistent numerical approxima-
relaxation terms and entropy. Commun. Pure Appl. Math. 47, 787–830. tion of a two-phase flow model for fluids with interfaces and cavitation. In:
Cole, R.H., 1948. Underwater Explosions. Princeton University Press. Hyperbolic Problems: Theory, Numerics, Applications, Proc. 14th Intl. Conf. on
Dal Maso, G., LeFloch, P.G., Murat, F., 1995. Definition and weak stability of noncon- Hyperbolic Problems. AIMS, pp. 839–846.
servative products. J. Math. Pures Appl. 74, 483–548. Pelanti, M., Shyue, K.-M., 2014b. A mixture-energy-consistent six-equation
Daramizadeh, A., Ansari, M., 2015. Numerical simulation of underwater explosion two-phase numerical model for fluids with interfaces, cavitation and evapora-
near air-water free surface using a five-equation reduced model. Ocean Eng. 110, tion waves. J. Comput. Phys. 259, 331–357.
25–35. Petel, O.E., Jetté, F.X., 2010. Comparison of methods for calculating the shock Hugo-
Davis, S.F., 1988. Simplified second-order Godunov-type methods. SIAM J. Sci. Stat. niot of mixtures. Shock Waves 20, 73–83.
Comput. 9, 445–473. Petitpas, F., Franquet, E., Saurel, R., Le Métayer, O., 2007. A relaxation-projection
De Lorenzo, M., Pelanti, M., Lafon, P., 2018. HLLC-type and path-conservative method for compressible flows. part II: artificial heat exchanges for multiphase
schemes for a single-velocity six-equation two-phase flow model: a compara- shocks. J. Comput. Phys. 225 (2), 2214–2248.
tive study. Appl. Math. Comput. 333, 95–117. Petitpas, F., Massoni, J., Saurel, R., Lapebie, E., Munier, L., 2009. Diffuse interface
Dumbser, M., Balsara, D.S., 2016. A new efficient formulation of the HLLEM Rie- models for high speed cavitating underwater systems. Int. J. Multiphase Flow
mann solver for general conservative and non-conservative hyperbolic systems. 35 (8), 747–759.
J. Comput. Phys. 304, 275–319. Saurel, R., Boivin, P., LeMétayer, O., 2016. A general formulation for cavitating, boil-
Dumbser, M., Toro, E.F., 2011. A simple extension of the Osher Riemann solver to ing and evaporating flows. Comput. Fluids 128, 53–64.
non-conservative hyperbolic systems. J. Sci. Comput. 48, 70–88. Saurel, R., Le Métayer, O., Massoni, J., Gavrilyuk, S., 2007. Shock jump relations
Flåtten, T., Lund, H., 2011. Relaxation two-phase models and the subcharacteristic for multiphase mixtures with stiff mechanical relaxation. Shock Waves 16,
condition. Math. Models Methods Appl. Sci. 21, 2379–2407. 209–232.
Flåtten, T., Morin, A., Munkejord, S., 2010. Wave propagation in multicomponent Saurel, R., Petitpas, F., Abgrall, R., 2008. Modelling phase transition in metastable
flow models. SIAM J. Appl. Math. 8, 2861–2882. liquids: application to cavitating and flashing flows. J. Fluid Mech. 607, 313–350.
Godlewski, E., Raviart, P.-A., 1996. Numerical Approximation of Hyperbolic Systems Saurel, R., Petitpas, F., Berry, R.A., 2009. Simple and efficient relaxation methods for
of Conservation Laws. Springer-Verlag, New York. interfaces separating compressible fluids, cavitating flows and shocks in multi-
Haimovich, O., Frankel, H., 2017. Numerical simulations of compressible multicom- phase mixture. J. Comput. Phys. 228, 1678–1712.
ponent and multiphase flow using a high-order targeted ENO (TENO) finite-vol- Stewart, H.B., Wendroff, B., 1984. Two-phase flow: models and methods. J. Comput.
ume method. Comput. Fluids 146, 105–116. Phys. 56, 363–409.
Jin, S., Xin, Z.P., 1995. The relaxation schemes for systems of conservation Tokareva, S.A., Toro, E.F., 2010. HLLC-type Riemann solver for the Baer-Nunziato
laws in arbitrary space dimensions. Commun. Pure Appl. Math. 48, 235– equations of compressible two-phase flow. J. Comput. Phys. 229, 3573–3604.
276. Toro, E.F., 1997. Riemann Solvers and Numerical Methods for Fluid Dynamics.
Kapila, A., Menikoff, R., Bdzil, J.B., Son, S.F., Stewart, D., 2001. Two-phase modeling of Springer-Verlag, Berlin, Heidelberg.
deflagration-to-detonation transition in granular materials: reduced equations. Toro, E.F., Spruce, M., Speares, W., 1994. Restoration of the contact surface in the
Phys. Fluids 13, 3002–3024. HLL Riemann solver. Shock Waves 4, 25–34.
Kedrinskiy, V.K., 2005. Hydrodynamics of Explosion: Experiments and Models. Wagner, W., Cooper, J.R., Dittmann, A., Kijima, J., Kretzschmar, H.-J., Kruse, A.,
Springer. Mareš, R., Oguchi, K., Sato, H., Stöcker, I., Šifner, O., Takaishi, Y., Tanishita, I.,
Kleine, H., Tepper, S., Takehara, K., Etoh, T.G., Hiraki, K., 2009. Cavitation induced by Trübenbach, J., Willkommen, T., 20 0 0. The IAPWS industrial formulation 1997
low speed underwater impact. In: Shock Waves: 26th International Symposium for the thermodynamic properties of water and steam. Trans. ASME 122,
on Shock Waves, vol. 2. Springer, pp. 895–900. 150–182.
Le Martelot, S., Saurel, R., Nkonga, B., 2014. Towards the direct numerical simulation Wang, G., Zhang, S., Yu, M., Li, H., Kong, Y., 2014. Investigation of the shock wave
of nucleate boiling flows. Int. J. Multiphase Flow 66, 62–78. propagation characteristics and cavitation effects of underwater explosion near
Le Métayer, O., Massoni, J., Saurel, R., 2004. Elaborating equations of state of a boundaries. Appl. Ocean Res. 46, 40–53.
liquid and its vapor for two-phase flow models. Int. J. Therm. Sci. 43, 265– Wood, A.B., 1930. A Textbook of Sound. G. Bell and Sons Ltd., London.
276. Xie, W.F., Liu, T.G., Khoo, B.C., 2006. Application of a one-fluid model for large scale
Le Métayer, O., Massoni, J., Saurel, R., 2013. Dynamic relaxation processes in com- homogeneous unsteady cavitation: the modified Schmidt model. Comput. Fluids
pressible multiphase flows. Application to evaporation phenomena. ESAIM Proc. 35, 1177–1192.
40, 103–123. Xie, W.F., Liu, T.G., Khoo, B.C., 2007. The simulation of cavitating flows induced by
LeVeque, R. J., Clawpack. http://www.clawpack.org. underwater shock and free surface interaction. Appl. Numer. Math. 57, 734–745.
LeVeque, R.J., 1997. Wave propagation algorithms for multi-dimensional hyperbolic Yeom, G.-S., 2015. Numerical study of underwater explosion near a free surface
systems. J. Comput. Phys. 131, 327–353. and a structural object on unstructured grid. J. Mech. Sci. Technol. 29 (10),
LeVeque, R.J., 2002. Finite Volume Methods for Hyperbolic Problems. Cambridge 4213–4222.
University Press. Zein, A., Hantke, M., Warnecke, G., 2010. Modeling phase transition for compressible
LeVeque, R.J., Pelanti, M., 2001. A class of approximate Riemann solvers and their two-phase flows applied to metastable liquids. J. Comput. Phys. 229, 2964–2998.
relation to relaxation schemes. J. Comput. Phys. 172, 572–591. Zein, A., Hantke, M., Warnecke, G., 2013. On the modeling and simulation of a
Linga, G., 2018. A hierarchy of non-equilibrium two-phase flow models. Submitted. laser-induced cavitation bubble. Int. J. Numer. Meth. Fluids 73 (2), 172–203.
Liu, T.G., Khoo, B.C., Xie, W.F., 2004. Isentropic one-fluid modelling of unsteady cav- Zhu, J., Liu, T., Qiu, J., Khoo, B.C., 2012. RKDG methods with WENO limiters for un-
itating flow. J. Comput. Phys. 201, 80–108. steady cavitating flow. Comput. Fluids 57, 52–65.

You might also like