Top Banner
Resolving the shock-induced combustion by an adaptive mesh redistribution method Li Yuan a, * , Tao Tang b a LSEC, Institute of Computational Mathematics, Academy of Mathematics and Systems Science, Chinese Academy of Sciences, Beijing 100080, PR China b Department of Mathematics, Hong Kong Baptist University, Kowloon Tong, Hong Kong, PR China Received 26 April 2006; received in revised form 5 October 2006; accepted 9 October 2006 Available online 1 December 2006 Abstract An adaptive mesh method is presented for numerical simulation of shock-induced combustion. The method is com- posed of two independent ingredients: a flow solver and a mesh redistribution algorithm. The flow solver is a finite-volume based second-order upwind TVD scheme, together with a lower–upper symmetric Gauss–Seidel relaxation scheme for solv- ing the multispecies Navier–Stokes equations with finite rate chemistry. The adaptive mesh is determined by a grid gen- eration method based on solving Poisson equations, with the monitor function carefully designed to resolve both sharp fronts and the induction zone between them. Numerically it is found that the resolution of the induction zone is particu- larly critical to the combustion problems, and an under-resolved numerical method may cause excessive energy release and spurious runaway reactions. The monitor function proposed in this paper, which is based on the relative rate of change of mass fraction, covered this issue satisfactorily. Numerical simulations of supersonic flows past an axisymmetric projectile in a premixed hydrogen/oxygen (or air) mixture are carried out. The results show that the spurious runaway chemical reac- tions appearing on coarse grids can be eliminated by using adaptive meshes without invoking any ad hoc treatment for reaction rates. The adaptive mesh approach is more effective than the fixed mesh one in obtaining grid-independent results. Finally, discrepancies between the numerical and benchmark experimental results and difficulties in simulating the deto- nation waves are delineated which appeal for further investigation. Ó 2006 Elsevier Inc. All rights reserved. Keywords: Adaptive mesh redistribution; Monitor function; Chemically reacting flow; Shock-induced combustion 1. Introduction Shock-induced combustion is the self-ignited combustion phenomenon of a premixed gas induced by a shock wave propagating in the gas mixture. There have been extensive experimental and theoretical studied on this phenomenon since 1960s, see, e.g. [1,2]. In the past two decades, it has regained widespread interest 0021-9991/$ - see front matter Ó 2006 Elsevier Inc. All rights reserved. doi:10.1016/j.jcp.2006.10.006 * Corresponding author. E-mail addresses: [email protected] (L. Yuan), [email protected] (T. Tang). Journal of Computational Physics 224 (2007) 587–600 www.elsevier.com/locate/jcp
14

Resolving the shock-induced combustion by an adaptive mesh …lsec.cc.ac.cn/~lyuan/jcp07comb.pdf · 2007. 6. 2. · Resolving the shock-induced combustion by an adaptive mesh redistribution

Feb 24, 2021

Download

Documents

dariahiddleston
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Resolving the shock-induced combustion by an adaptive mesh …lsec.cc.ac.cn/~lyuan/jcp07comb.pdf · 2007. 6. 2. · Resolving the shock-induced combustion by an adaptive mesh redistribution

Journal of Computational Physics 224 (2007) 587–600

www.elsevier.com/locate/jcp

Resolving the shock-induced combustion by an adaptivemesh redistribution method

Li Yuan a,*, Tao Tang b

a LSEC, Institute of Computational Mathematics, Academy of Mathematics and Systems Science,

Chinese Academy of Sciences, Beijing 100080, PR Chinab Department of Mathematics, Hong Kong Baptist University, Kowloon Tong, Hong Kong, PR China

Received 26 April 2006; received in revised form 5 October 2006; accepted 9 October 2006Available online 1 December 2006

Abstract

An adaptive mesh method is presented for numerical simulation of shock-induced combustion. The method is com-posed of two independent ingredients: a flow solver and a mesh redistribution algorithm. The flow solver is a finite-volumebased second-order upwind TVD scheme, together with a lower–upper symmetric Gauss–Seidel relaxation scheme for solv-ing the multispecies Navier–Stokes equations with finite rate chemistry. The adaptive mesh is determined by a grid gen-eration method based on solving Poisson equations, with the monitor function carefully designed to resolve both sharpfronts and the induction zone between them. Numerically it is found that the resolution of the induction zone is particu-larly critical to the combustion problems, and an under-resolved numerical method may cause excessive energy release andspurious runaway reactions. The monitor function proposed in this paper, which is based on the relative rate of change ofmass fraction, covered this issue satisfactorily. Numerical simulations of supersonic flows past an axisymmetric projectilein a premixed hydrogen/oxygen (or air) mixture are carried out. The results show that the spurious runaway chemical reac-tions appearing on coarse grids can be eliminated by using adaptive meshes without invoking any ad hoc treatment forreaction rates. The adaptive mesh approach is more effective than the fixed mesh one in obtaining grid-independent results.Finally, discrepancies between the numerical and benchmark experimental results and difficulties in simulating the deto-nation waves are delineated which appeal for further investigation.� 2006 Elsevier Inc. All rights reserved.

Keywords: Adaptive mesh redistribution; Monitor function; Chemically reacting flow; Shock-induced combustion

1. Introduction

Shock-induced combustion is the self-ignited combustion phenomenon of a premixed gas induced by ashock wave propagating in the gas mixture. There have been extensive experimental and theoretical studiedon this phenomenon since 1960s, see, e.g. [1,2]. In the past two decades, it has regained widespread interest

0021-9991/$ - see front matter � 2006 Elsevier Inc. All rights reserved.

doi:10.1016/j.jcp.2006.10.006

* Corresponding author.E-mail addresses: [email protected] (L. Yuan), [email protected] (T. Tang).

Page 2: Resolving the shock-induced combustion by an adaptive mesh …lsec.cc.ac.cn/~lyuan/jcp07comb.pdf · 2007. 6. 2. · Resolving the shock-induced combustion by an adaptive mesh redistribution

588 L. Yuan, T. Tang / Journal of Computational Physics 224 (2007) 587–600

in aerospace science mainly being a promising combustion mechanism for hypersonic propulsion devices suchas pulse detonation engine, oblique detonation wave engine, and ram accelerator [3].

The flowfield of the shock-induced combustion is characterized by the coupling and interactions among theshock wave, combustion front, and combustion instability, which result in various and distinctive flow andwave structures according to the chemical and fluid dynamic conditions. These complicated phenomena posechallenges for numerical simulations. One difficulty is how to capture the coupling and separation betweenshock and combustion fronts. In reality, when the flight velocity of a projectile in a premixed combustiblegas is less than the Chapman–Jouguet (C–J) detonation speed of the gas mixture (referred to as subdetonativespeed), the bow shock and the combustion front are separated by a region called induction zone; when theflight velocity is greater than the C–J detonation speed (referred to as superdetonative speed), the bow shockand the combustion front partially merge to form coupled detonative wave structure [1]. The induction zone isa region of nearly constant pressure, temperature and density, yet many reactions with different reaction ratesare in progress, giving rise to a release of energy, which may trigger a temperature-sensitive explosion. If thenumerical scheme is not accurate enough or the grid size is not sufficiently fine to capture the fastest process,errors in the computed quantities can cause the release of chemical energy much sooner than it should bereleased due to the extreme temperature sensitivity of reaction rates, resulting in so-called spurious runwaychemical reactions [4]. Standard shock-capturing schemes will lead to nonphysical one-grid-cell per time stepspurious detonation velocities when the chemical reactions introduce time scales that are significantly shorterthan the flow time scales [5]. Some empirical approaches such as limiting energy release rate or Damkohlernumber were taken to cure this problem [4,6,7]. However, ad hoc modifications to well defined reaction ratesmight spoil the underlying physics. Another difficulty that plagues numerical simulations of shock-inducedcombustion is false combustion instabilities due to the use of inappropriate reaction mechanisms or numericalschemes.

There are numerous studies utilizing adaptive mesh refinement or mesh redistribution methods for provid-ing required spatial resolution in local regions so as to reduce total computational cost (see e.g. [8–14]), but theissue of resolving the induction zone in shock-induced combustion is pervasive in most calculations except fewones where a wavelet adaptive multilevel method [15], or multi-resolution schemes with adaptive grids [16]were used to resolve the induction zone in one-dimensional detonation problems. Recently, the adaptive meshrefinement method was implemented for high-resolution computation of two-dimensional detonation wavestructures [17]. The moving mesh method [18] was also utilized to solve the reactive Euler equations and itsefficiency was demonstrated for resolving thin structures of the detonation flows. Nevertheless, even if theinduction zone was already known crucial in determining the detonation structures [15,19], it was not inten-tionally resolved in these adaptive mesh simulations.

In this paper, we develop an adaptive mesh redistribution method that is able to resolve shock and combus-tion waves as well as the induction zone. Adaptive mesh method of this type keeps a simple mesh structure andinvariable number of grid points, and has the advantage of separating the mesh redistribution algorithm fromthe flow solver. For the sake of robustness, we choose a standard second-order upwind total variation dimin-ishing (TVD) scheme of Harten–Yee type [20,21] for the flow solution. The chemical source terms are treated bya pointwise implicit approach instead of splitting approaches, which is in line with the lower–upper symmetricGauss–Seidel relaxation scheme as implemented in [22]. Once the flow field is updated at a given time level, thegrid may be redistributed using an iteration procedure [23,24]. The conventional Poisson grid equation ofThompson et al. [25] is used for mesh redistribution, with its control functions determined from equidistribu-tion of the monitor function along each grid direction in a way similar to [26]. We made slight yet critical mod-ification to the control functions given in [26] for robustness. The peculiarity of the present monitor function isthe relative rate of change of mass fractions, which will be shown to be effective for resolving the induction zone.Numerical simulations are compared with Lehr’s experiments [1]. Grid convergence and computational effectsfor two representative hydrogen/oxygen combustion mechanisms are also demonstrated.

This paper is organized as follows. The governing equations and reaction mechanisms are given in Section2. The flow solution method is listed in Section 3, and the adaptive mesh redistribution algorithm is describedin Section 4. Numerical simulations of the steady-state flows corresponding to flow conditions in Lehr’s exper-iments are given in Section 5 and discrepancies between simulations and experiments are delineated. Conclu-sions are presented in Section 6.

Page 3: Resolving the shock-induced combustion by an adaptive mesh …lsec.cc.ac.cn/~lyuan/jcp07comb.pdf · 2007. 6. 2. · Resolving the shock-induced combustion by an adaptive mesh redistribution

L. Yuan, T. Tang / Journal of Computational Physics 224 (2007) 587–600 589

2. Governing equations

The governing equations for the chemically reacting viscous flows are the compressible Navier–Stokesequations with chemical source terms for a mixture composed of N gas species, which are expressed in thefollowing form:

oQ

otþ oE

oxþ oF

oy¼ oEv

oxþ oFv

oyþHþHv þ S; ð1Þ

where

Q ¼ ½q1; . . . ; qN ; qu; qv; qE�T; ð2ÞE ¼ ½q1u; . . . ; qN u; qu2 þ p; quv; uðqE þ pÞ�T; ð3ÞF ¼ ½q1v; . . . ; qN v; quv;qv2 þ p; vðqE þ pÞ�T; ð4Þ

Ev ¼ ½qD1

oc1

ox; . . . ; qDN

ocN

ox; sxx; sxy ; usxx þ vsxy þ qx�

T; ð5Þ

Fv ¼ qD1

oc1

oy; . . . ; qDN

ocN

oy; sxy ; syy ; usxy þ vsyy þ qy

� �T

; ð6Þ

H ¼ �my½q1v; . . . ; qN v; quv; qv2; vðqE þ pÞ�T; ð7Þ

Hv ¼my

qD1

oc1

oy; . . . ; qDN

ocN

oy; sxy ; syy � shh; usxy þ vsyy þ qy

� �T

; ð8Þ

S ¼ ð _x1; . . . ; _xN ; 0; 0; 0ÞT ð9Þ

and

sxx ¼ 2louox� 2

3lr � u ¼ 2

3l 2

ouox� ov

oy� mv

y

� �; ð10Þ

syy ¼ 2lovoy� 2

3lr � u ¼ 2

3l 2

ovoy� ou

ox� mv

y

� �; ð11Þ

sxy ¼ louoyþ ov

ox

� �; ð12Þ

shh ¼ 2lvy� 2

3lr � u; ð13Þ

qx ¼ koToxþ q

XN

k¼1

Dkhkock

ox; ð14Þ

qy ¼ koToyþ q

XN

k¼1

Dkhkock

oy: ð15Þ

The equations describe two-dimensional plane flow if m = 0 and axisymmetric flow if m = 1. u and v are veloc-ity components, p is the pressure, T is the temperature, E is the total energy per unit mass, qk is the density ofspecies k, with total density q ¼

PNk¼1qk, ck = qk/q is the mass fraction, _xk is the mass production rate of spe-

cies k due to chemical reactions, hk is the specific enthalpy, Dk is the mass diffusivity of species k in the mixturedefined as

Dk ¼ ð1� X kÞXj 6¼k

X j

Dkj;

,

where Dkj is the binary diffusivity, Xk is the mole fraction, and l and k are dynamic viscosity and thermal con-ductivity of the gas mixture, respectively. The diffusive transport coefficients, l and k, are determined using

Page 4: Resolving the shock-induced combustion by an adaptive mesh …lsec.cc.ac.cn/~lyuan/jcp07comb.pdf · 2007. 6. 2. · Resolving the shock-induced combustion by an adaptive mesh redistribution

590 L. Yuan, T. Tang / Journal of Computational Physics 224 (2007) 587–600

Wilke’s semi-empirical formula with lk and kk of each species found from NASA thermodynamic data [27].For high speed flow simulations, mass diffusion is greatly simplified by neglecting pressure diffusion and ther-mal diffusion, and assuming the binary diffusivity Dkj to be equal between all components [28]. Its value isobtained by assuming a constant Schmidt number Sc = l/(qDkj) = 0.5, see [27].

It is assumed that all species are thermally perfect, in thermal equilibrium and have the same temperature.The equation of state is that for a mixture of thermally perfect gases

p ¼XN

k¼1

qk

MkRuT ; ð16Þ

where Mk is the molecular weight of species k and Ru is the universal gas constant. The total energy per unitvolume qE is used for implicit evaluation of temperature T by the Newton iteration method through the ther-modynamic relationship qe = qh � p, i.e.

qE � 1

2qu2 ¼

XN

k¼1

qk

Z T

T 0

Cpk

MkdT 0 þ h0

k

� �� RuT

XN

k¼1

qk

Mk; ð17Þ

where Cpkis the specific heat at constant pressure, and h0

k is the heat of formation at reference temperature T0.The specific heats are expressed as functions of temperature in polynomial fit

Cpk

Ru

¼ a1k þ a2kT þ a3kT 2 þ a4kT 3 þ a5kT 4; ð18Þ

where the coefficients aik can be obtained from thermodynamic data file of the chemical kinetics package(CHEMKIN [29]) or from NASA thermochemical polynomial data [30], which are often valid in certain tem-perature ranges. The appropriate speed of sound for the thermally perfect gas mixture model is the frozenspeed of sound

a2 ¼XN

k¼1

ckpqkþ ðc� 1ÞðH � u2Þ

with H ¼ hþ 12u2, pqk

¼ ðc� 1Þð12u2 � hkÞ þ cRuT=Mk, and c = Cp/(Cp � R), where c can be calculated from the

frozen specific heat Cp ¼PN

k¼1ckCpkand mixture gas constant R ¼ Ru

PNk¼1ck=Mk.

In realistic combustion, the general formula for a reaction mechanism with I elementary reactions isexpressed as

XN

k¼1

m0kiX k�

XN

k¼1

m00kiX k; i ¼ 1; . . . ; I : ð19Þ

The mass production rate for each species is

_xk ¼ Mk

XI

i¼1

ðm00ki � m0kiÞ kf i

YNk¼1

qk

Mk

� �m0ki

� kbi

YNk¼1

qk

Mk

� �m00ki

" #; k ¼ 1; . . . ;N : ð20Þ

where the forward and backward reaction rate constants for the ith reaction, kf i and kbi , are given in Arrheniusformula (omitting the subscripts f, b for clarity)

ki ¼ AiT Bi expð�Ci=T Þ: ð21Þ

Two reaction mechanisms for H2–air mixture combustion are used and their simulation effects will be com-pared. One is Evans and Schxnayder’s 12-species 25-step reaction scheme [31], another is Jachimowski’s 13-species 33-step reaction scheme [32]. The first 16 and 19 reactions in Refs. [31,32], respectively, are used forH2/O2 mixture combustion.
Page 5: Resolving the shock-induced combustion by an adaptive mesh …lsec.cc.ac.cn/~lyuan/jcp07comb.pdf · 2007. 6. 2. · Resolving the shock-induced combustion by an adaptive mesh redistribution

L. Yuan, T. Tang / Journal of Computational Physics 224 (2007) 587–600 591

3. Flow solution method

The governing equations for chemically reacting flows are often stiff. There are usually two kinds ofapproach to solve stiff systems: the splitting approach which decouples fluid dynamics from chemical kinetics,and the non-splitting one which solves the fully coupled equations simultaneously. In the present study we usethe latter approach as implemented in [22]. The governing equations (1) are integrated over a quadrilateralmesh cell

Z

X

oQ

otdXþ

IoXðf � nÞ dS ¼

ZXðHþHv þ SÞ dX; ð22Þ

where f = (E � Ev)i + (F � Fv)j. By virtue of divergence theorem, we can obtain the semi-discretized finite vol-ume formulation for a rectangular mesh cell

XoQ

ot

� �I ;J

þ ðEiþ1 � EiÞJ þ ðFjþ1 � FjÞI ¼ ðEv;iþ1 � Ev;iÞJ þ ðFv;jþ1 � Fv;jÞI þ ½XðHþHv þ SÞ�I;J ; ð23Þ

where uppercase I, J denote the center of a mesh volume X, and lowercase i, j denote the cell faces,Ei ¼ ðEiþ FjÞ � ðnSÞi; Ev;i ¼ ðEviþ FvjÞ � ðnSÞi. By applying backward time difference, linearization of all fluxderivative terms and chemical source terms at n + 1 time level with respect to n level, and first-order upwindscheme for resulting terms containing inviscid flux Jacobians, we obtain the fully discretized incremental form:

XDt

I� XPþ Aþ � A� þ Bþ � B� þ Avi þ Av

iþ1 þ Bvj þ Bv

jþ1

� �I ;J

� DQI;J � ðAþI�1 þ Avi;I�1ÞDQI�1;J

þ ðA�Iþ1 � Aviþ1;Iþ1ÞDQIþ1;J � ðBþJ�1 þ Bv

j;J�1ÞDQI ;J�1 þ ðB�Jþ1 � Bvjþ1;Jþ1ÞDQI ;Jþ1 ¼ RHSn; ð24Þ

where

RHSn ¼ �ðEiþ1 � EiÞnJ � ðFjþ1 � FjÞnI þ ðEv;iþ1 � Ev;iÞnJ þ ðFv;jþ1 � Fv;jÞnI þ XðHþHv þ SÞn; ð25Þ

DQ = Qn+1 � Qn, P = oS/oQ is the Jacobian matrix of chemical source vector which is N · N in size as sim-plified in [22], A± and B± are split Jacobian matrices of inviscid fluxes, where we have used the approximateflux Jacobian splitting A� ¼ 1

2ðA� kAIÞ, with A ¼ oE=oQ being the inviscid flux Jacobian and kA its maximum

absolute eigenvalue. The viscous Jacobian matrices are

Avi;I�1 ¼ �oEv;i=oQI�1; Av

i;I ¼ oEv;i=oQI ;

Aviþ1;I ¼ �oEv;iþ1=oQI ; Av

iþ1;Iþ1 ¼ oEv;iþ1=oQIþ1:ð26Þ

Eq. (24) is iteratively solved by the LU-SGS scheme

kD�I ;J � ðAþI�1 þ Av

i;I�1ÞD�I�1;J � ðBþJ�1 þ Bv

j;J�1ÞD�I ;J�1 ¼ RHSn; ð27aÞ

kDQI ;J þ A�Iþ1 � Aviþ1;Iþ1

� �DQIþ1;J þ B�Jþ1 � Bv

jþ1;Jþ1

� �DQI ;Jþ1 ¼ kD�I ;J ; ð27bÞ

where

k ¼ XDt

I� XPþ Aþ � A� þ Bþ � B� þ Avi þ Av

iþ1 þ Bvj þ Bv

jþ1

� �I;J

:

In the evaluation of RHSn, the diffusive terms are approximated by the conventional central difference [33],and the inviscid numerical flux vector are approximated by the second-order upwind TVD scheme due to Har-ten–Yee [20,21]

~Eiþ1 ¼

1

2½Eiþ1ðQIÞ þ Eiþ1ðQIþ1Þ þ Riþ1Uiþ1�; ð28Þ

Page 6: Resolving the shock-induced combustion by an adaptive mesh …lsec.cc.ac.cn/~lyuan/jcp07comb.pdf · 2007. 6. 2. · Resolving the shock-induced combustion by an adaptive mesh redistribution

592 L. Yuan, T. Tang / Journal of Computational Physics 224 (2007) 587–600

where Ri+1 is a matrix whose column vectors are the right eigenvectors of A evaluated at some symmetric aver-age of QI and QI+1, denoted as Qi+1 [34]. The element /l

iþ1 of the dissipation vector Ui+1 for Harten–Yeescheme is

/liþ1 ¼ rðkl

iþ1ÞðglI þ gl

Iþ1Þ � wðkliþ1 þ cl

iþ1Þaliþ1; ð29Þ

where kliþ1 is the eigenvalue of A evaluated at Qi+1, al

iþ1 is the element of the vector aiþ1 ¼ R�1iþ1ðQIþ1 �QIÞ,

and gliþ1 is the limiter function. The function r(z) and variable cl

iþ1 are

rðzÞ ¼ 1

2wðzÞ � Dt

Xz2

� �; ð30Þ

cliþ1 ¼ rðkl

iþ1Þðgl

Iþ1 � glIÞ=al

iþ1; aliþ1 6¼ 0;

0; aliþ1 ¼ 0:

(ð31Þ

The entropy function is given by

wðzÞ ¼jzj; jzjP �;z2þ�2

2�; jzj < �;

(ð32Þ

where � is an empirical coefficient. The limiter function used is

glI ¼

aliðal

iþ1Þ2 þ al

iþ1ðaliÞ

2

ðaliÞ

2 þ ðaliþ1Þ

2: ð33Þ

4. Adaptive mesh redistribution

The present adaptive mesh redistribution method follows from an earlier one by Tang and Tang [23], whichis of grid-location type and has the advantage of decoupling the mesh redistribution from the flow evolution.A continuous conservative interpolation procedure is used for transferring flow variables from previous tonext grid in the iteration process. However, it did not allow grid anisotropy. To achieve grid anisotropy,we use the conventional Poisson grid generation equation due to Thompson et al. [25] to generate and redis-tribute a structured quadrilateral mesh

r2n ¼ g22

gP ;

r2g ¼ g11

gQ;

ð34Þ

where g11 ¼ x2n þ y2

n; g22 ¼ x2g þ y2

g, and P and Q are called control functions which determine the grid pointdistribution. For finite difference solution, the above system is inverted into the computational domain byinterchanging its dependent and independent variables so as to solve for x, y on the computational domain:

g22ðrnn þ P rnÞ þ g11ðrgg þ QrgÞ � 2g12rng ¼ 0; ð35Þ

where r = (x,y), g12 = rn Æ rg. The determination of P and Q depends on whether they are for initial grid gen-eration or for grid redistribution. In the first case when the initial grid is to be generated, the distribution ofgrid points on the boundary is given. From the boundary grid points, the control functions P along n-lineboundaries, and Q along g-line boundaries are determined from projection analysis

P ¼ � rn � rnn

jrnj2;

Q ¼ � rg � rgg

jrgj2;

ð36Þ

and are further modified to take into account the orthogonality on boundaries and the curvature effect [25].The interior P(n,g) and Q(n,g) are subsequently obtained through one-directional interpolation in the

Page 7: Resolving the shock-induced combustion by an adaptive mesh …lsec.cc.ac.cn/~lyuan/jcp07comb.pdf · 2007. 6. 2. · Resolving the shock-induced combustion by an adaptive mesh redistribution

L. Yuan, T. Tang / Journal of Computational Physics 224 (2007) 587–600 593

computational domain. We denote the initial control functions as P0 and Q0, which will play an anchoring rolein subsequent grid adaptation.

In the second case when the grid is to be redistributed, the control functions P and Q are in principleobtained from so-called monitor functions x1 and x2 in n and g curvilinear coordinate direction, respectively.The monitor functions indicate solution errors in some sense [35]. It is very important to choose an appropri-ate monitor function, otherwise the adaptive effect cannot be realized no matter how good a moving meshalgorithm is. Several strategies for designing monitor functions have been proposed e.g. [36,37]. Accordingto Anderson’s analysis [26], the equidistribution of the monitor function xi along a coordinate line ni is equiv-alent to setting control functions as

P ¼ 1

x1

ox1

on;

Q ¼ 1

x2

ox2

og:

ð37Þ

However, we have found that using (37) alone would lead to severely twisted grids near the body surface,resulting in breakdown of a viscous flow computation. To increase robustness, we let (P,Q) be weighted aver-ages between (37) and initial ones

ðP ;QÞT ¼ aðP ;QÞT þ ð1� aÞðP 0;Q0ÞT; ð38Þ

where a = (min(j, jc) � 1)/(jc � 1), and jc is a prescribed index in the direction normal to the wall correspondingroughly to the outer edge of the boundary layer. The grid generation system (35) is solved using Gauss–Seideliteration. After each iteration, the flow variables Q are transferred from the previous grid Gp to the next gridGp+1 by a monotonic conservative interpolation procedure [23]. This enables calculation of the monitor func-tions, hence the control functions for the next iteration. The boundary grid points are also redistributedaccording to the tangential displacement of the neighboring interior grid points when necessary. In orderto maintain a smooth mesh for the viscous flow computation, boundary grid points on the body surfaceare not redistributed.

To achieve grid adaptation near shock and combustion fronts, the following monitor function is often used

xi ¼

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi1þ b1

o�Toni

� �2s

; ð39Þ

with �T being the non-dimensional temperature, and b1 the adaptivity constant, see e.g. [37–39]. However, anefficient monitor function within the induction zone is not obvious due to essentially constant density, pressureand temperature in this region. It is believed that many reactions are occurring, giving rise to variation in com-positions of some chemical species. We have observed large gradients in logarithmic scale of the mass fractionsof some radical species just before the combustion front. The variation of logarithmic mass fraction is equiv-alent to the relative rate of change of mass fraction, namely,

o ln �ck

oni¼ 1

�ck

o�ck

oni; ð40Þ

where �ck ¼ maxðck; 10�9Þ is the tailored mass fraction. Therefore, we suggest a monitor function

xi ¼

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi1þ b1

o�Toni

� �2

þ b2 maxk

1

�ck

o�ck

oni

� �� �2s

; ð41Þ

where b1 and b2 are adaptivity constants. The term 1�ck

o�ckoni

can capture large variations of �ck even if the magnitudeof �ck is small in the induction zone. In the induction zone �ck itself may be very small in magnitude, but itsvariation is critical. Thus, using a relative rate of change makes better sense than using rate of change directly.Indeed, the monitor function (41) is found to give adequate adaption both in the induction zone and near theshock and combustion fronts.

Page 8: Resolving the shock-induced combustion by an adaptive mesh …lsec.cc.ac.cn/~lyuan/jcp07comb.pdf · 2007. 6. 2. · Resolving the shock-induced combustion by an adaptive mesh redistribution

594 L. Yuan, T. Tang / Journal of Computational Physics 224 (2007) 587–600

A summary of the adaptive mesh redistribution algorithm is given below:

(1) Generate the initial grid using the elliptic grid generation system (35), with P0,Q0 obtained from givenboundary grid point distributions and interpolations.

(2) Advance the flow solution M steps by the LU-SGS scheme (27a).(3) In the next time step, after the M steps with the LU-SGS scheme, do grid redistribution:

(a) Calculate the monitor function xi from (41) and do 1–2 times spatial smoothing on xi, then calculatethe control functions P,Q by the weighted averaging (38).

(b) Solve Eq. (35) by one iteration of Gauss–Seidel relaxation to get new grid Gp+1.(c) Interpolate the flow solution Q from old grid Gp onto the new grid Gp+1 by using the conservative

interpolate procedure given in [23].(d) The iteration procedure (a)–(c) on grid redistribution and flow solution interpolation is continued

until the grid do not change significantly from one iteration to the next.

(4) Start new time step (go to step 2).

5. Numerical examples and discussions

Two test cases of shock-induced combustion corresponding to Figs. 1 and 5 in Lehr’s benchmark experi-ments [1] were selected to testify the present adaptive method. A spherical projectile with a diameter of 15 mmis flying at supersonic speeds 1892 m/s (M = 3.55) and 2605 m/s (M = 6.46) in a stoichiometric H2/O2

(2H2 + O2) and H2/air (2H2 + O2+3.76N2) mixture of temperature 293 K at pressure of 186 and 320 mmHg,respectively. Since the detonation speeds of the two mixtures are 2550 and 2055 m/s, respectively, thenM = 3.55 corresponds to the subdetonative speed and M = 6.46 corresponds to the superdetonative speed.We will use M = 3.55 case to demonstrate the effect of mesh adaptation on elimination of spurious runwaychemical reactions, and the effect of the new monitor function, and use M = 6.46 case to compare detonationstructures. The computational domain may include either a hemisphere or a hemisphere plus a cylinder of onediameter in length. Non-slip, adiabatic, and non-catalytic boundary conditions are used on the body surface.All the computations are initialized from the free-stream conditions and are assumed to reach steady statewhen the residual drops to 10�5.

5.1. Role of adaptive mesh and the monitor function

Fig. 1 shows the fixed and adaptive grids and corresponding temperature contours for M = 3.55 case. It canbe seen from Fig. 1b that both the shock wave and the combustion front go out of the computational domainfor the fixed grid computation. The result implies spurious runaway chemical reactions because they are con-trary to experimental stationary shock and combustion fronts. This artifact occurs on low-resolution 652 gridbut not on 1292 and finer grids. However, adaptive mesh calculations using traditional monitor function (39)or present one (41) show that the shock and combustion fronts are captured inside the computational domainon 652 grid, and temperature contours for both monitors are virtually identical as shown in Fig. 1d. Further-more, we can see from Fig. 1e that the adaptive mesh obtained with the new monitor function (41) is clusterednot only near the shock and combustion fronts, but also in the induction zone between them.

We discuss the spurious runaway chemical reactions as shown in Fig. 1b a bit further. They are related toexcessive release of chemical energy in one grid cell in a single time step in numerical calculation of shock-def-lagration problems [4]. If the time step chosen according to the CFL condition is not small enough to resolvechemical reactions, or inexact discretization or insufficient spatial resolution appears near the deflagration,then numerical errors will falsely trigger a temperature-sensitive chemical reaction earlier than it should be.We have found that a low-order numerical scheme is more prone to produce spurious runaway reactions thana high-order one, and that chemical kinetics used also has influence on runaway reactions. For example, cal-culation with Jachimowski’ mechanism produce runaway reactions on 332 but not on 652 grids. In general sit-uations when a reaction mechanism contains very fast exothermic reactions, remedies such as limiting the

Page 9: Resolving the shock-induced combustion by an adaptive mesh …lsec.cc.ac.cn/~lyuan/jcp07comb.pdf · 2007. 6. 2. · Resolving the shock-induced combustion by an adaptive mesh redistribution

-0.5 0 0.5 1

0

0.5

1

1.5

2

11.4

10.6

9.811.0

10.2

10.2

9.8

9.4

8.6

-0.5 0 0.5 1

0

0.5

1

1.5

2

-0.5 0 0.5 1

0

0.5

1

1.5

2

3.4

3.0

2.6

2.2

1.0

8.6

9.4

10.6

10.6

11.4

-0.5 0 0.5 1

0

0.5

1

1.5

2

-0.5 0 0.5 1

0

0.5

1

1.5

2

a b c

ed

Fig. 1. Fixed (a) and adaptive grids (c,e), and corresponding temperature contours for T/T1 (b,d) computed using 652 grid points forM = 3.55 flow in stoichiometric H2/O2 mixture. (c) and (d) are obtained using the conventional monitor function (39), and (e) shows theadaptive grid using the present monitor function (41). Evans’ reaction mechanism [31] is used throughout.

L. Yuan, T. Tang / Journal of Computational Physics 224 (2007) 587–600 595

energy release rate or Damkohler number [4,6,7], or adopting infinitely fast chemistry model, have to beinvoked. However, ad hoc limiting to reaction rates should be avoided. We see the adaptive mesh computationis an alternative approach for eliminating spurious runaway reactions with fewer grid points because it canprovide sufficient spatial resolution near deflagration fronts.

5.2. Grid convergence tests

Fig. 2 shows temperature distributions along the front stagnation line for M = 3.55 case calculated by usingsuccessively fine grid points. One can see that the distributions converge on 1952 adaptive grids, but they donot even on 3852 fixed grids. It seems that the adaptive mesh calculation on 1292 grid at least can give wellconverged shock location. If the results on 1952 adaptive grids are thought to be equivalent to those on3852 fixed grids, it roughly translates into half CPU time saving for the adaptive mesh calculation.

We remark that the converged position of the shock and deflagration fronts are also different for differentreaction mechanisms. The shock and deflagration fronts obtained with Evans’ mechanism are further away

Page 10: Resolving the shock-induced combustion by an adaptive mesh …lsec.cc.ac.cn/~lyuan/jcp07comb.pdf · 2007. 6. 2. · Resolving the shock-induced combustion by an adaptive mesh redistribution

x/R

Tem

per

atu

re (

K)

0.000.050.100.150.200.250.300.350

500

1000

1500

2000

2500

3000

3500

3852 adaptive grid1952 adaptive grid1292 adaptive grid3852 fixed grid1952 fixed grid1292 fixed grid

Fig. 2. Temperature along the stagnation streamline for adaptive and fixed grids using three successively fine grid points. M = 3.55 flow.Evans’ reaction mechanism is used throughout.

596 L. Yuan, T. Tang / Journal of Computational Physics 224 (2007) 587–600

from the body and in better agreement with the experiment. However, Evans’ mechanism is prone to causenumerical unsteady combustion on heavily adapted grids in contrast to the steady state combustion observedin the experiment for M = 6.46 hydrogen/air case, while Jachimowski’s mechanism is not so easy. In thisregard, Jachimowski’s mechanism is more robust than Evans’. We will go back to discuss this issue in the nextsubsection.

Fig. 3 shows comparison between the calculated density contours and the experimental shadowgraph imagefor M = 3.55 case. The computational domain extends one sphere diameter to the cylindrical portion wheresupersonic outflow boundary conditions are used on the exit plane. We see fair agreement between the calcu-lation and the experiment in the upstream, however, the computation does not predict the more inclined com-bustion front and the second shock wave originating from the deflected combustion front in the flank of theafterbody as appeared in the experiment. This discrepancy may be attributed to inaccurate modeling of vis-cous effects, ignorance of the flow transition and base flow region, or other unknown reasons. Further inves-tigation may include base flow region and turbulence modeling, which is not perused in this paper.

5.3. M = 6.46 hydrogen/air case

The M = 6.46 case corresponds to the superdetonative speed, where the shock wave and the deflagrationfront are coupled near the front stagnation line and separated as soon as the velocity component normalto the bow shock wave is equal to the detonation velocity [1]. The exact position of the separation is very sen-sitive to the amount of heat release, hence depends on the reaction mechanism used. Several numerical studieshave tested the appropriateness of different reaction mechanisms using this case (see [40] and referencestherein). However, these earlier calculations used relatively smaller number of grid points that doomed to pre-dict a smaller induction zone than in the experiment, and in particular, veil potential numerical combustioninstability due to excessive numerical diffusions present on coarse grids. Based on the grid convergence testin Section 5.2, we compare the computational effects between Evan’s and Jachimowski’s mechanisms usingmore than 1952 grid points.

Figs. 4a–f compare the density contours for a hemisphere domain calculated with both reaction mecha-nisms on fixed and adaptive grids, respectively. It is seen that the combustion front computed with Evans’mechanism separates from the shock wave earlier and the induction zone is larger than computed with Jachio-mowski’s mechanism, which indicates the former mechanism is better in matching numerical results with theexperimental one. One can also see that the density contours on 1952 adaptive grid (Figs. 4c and d) are com-parable to or better than those on 3852 fixed grid (Figs. 4a and b). The adaptive 3852 grid (Fig. 4e) provideslittle improvement over 1952 adaptive grid because the adaptive extent is limited by numerical combustion

Page 11: Resolving the shock-induced combustion by an adaptive mesh …lsec.cc.ac.cn/~lyuan/jcp07comb.pdf · 2007. 6. 2. · Resolving the shock-induced combustion by an adaptive mesh redistribution

Fig. 3. Comparison of calculated density contours with experimentally obtained shock position (s), combustion front (h) andshadowgraph [1] for M = 3.55 H2/O2 flow. The adaptive mesh has 5132 grid points. Evans’ reaction mechanism is used.

L. Yuan, T. Tang / Journal of Computational Physics 224 (2007) 587–600 597

instabilities occurring on highly adaptive grids for Evans’ mechanism. Nevertheless, it provides evidentimprovement for Jachiomowski’s mechanism as the mesh adaptivity enhances the solution resolution withoutcausing combustion instabilities (Fig. 4f).

Figs. 5a and b show comparisons of calculated density contours in the domain with the cylindrical after-body on 5132 adaptive grids. We can see the shock and combustion fronts for two mechanisms are comparableto the experimental positions. However, the calculated induction zone is smaller in the upstream and wider inthe downstream than in the experiment. Again, the induction zone computed with Evans’ mechanism is largerthan with Jachiomowski’s mechanism, showing the former is in better agreement with the experiment in theupstream region.

At this point, we discuss difficulties encountered in the numerical computations of this particular case. Inobtaining the adaptive grid results, we found that each grid redistribution must be followed by a long, fixed-grid computation to obtain nearly convergent flowfield; otherwise, consecutive adaptive computation wouldlead to unrealistic, extensive combustion instabilities which subsequently result in dramatic grid redistribu-tions to make computation not convergent to steady state. In our computation, only 1–2 times mesh adapta-tions are allowed, further adaptations would again cause unrealistic combustion instabilities.

To find out whether the nonsmoothness of adaptive meshes or the fine mesh size itself is responsible fornumerical combustion instability, calculations were conducted on 7692 and 10252 fixed grid. We found bothmechanisms can converge to steady state on 7692 grid and predicted qualitatively the same result as the exper-imental, but there are unrealistic combustion instabilities on 10252 grid. To our relief, calculations on 10252

grid can converge to steady state for a chemically frozen flow. It is thus believed that the numerical combus-tion instability is caused by very fine mesh size. Thus, the case of M = 6.46 presents a particular difficulty asso-ciated with numerical simulation of detonation waves by using detailed chemistry: when a high resolutionmesh is desirable for eliminating runaway reactions, an excessive fine mesh might lead to false unstable

Page 12: Resolving the shock-induced combustion by an adaptive mesh …lsec.cc.ac.cn/~lyuan/jcp07comb.pdf · 2007. 6. 2. · Resolving the shock-induced combustion by an adaptive mesh redistribution

-0.5 0 0.5 10

0.5

1

1.5

2

(a) 3852 fixed gridmechanism 1

-0.5 0 0.5 10

0.5

1

1.5

2

(b) 3852 fixed gridmechanism 2

-0.5 0 0.5 10

0.5

1

1.5

2

(c) 1952 adaptive gridmechanism 1

-0.5 0 0.5 10

0.5

1

1.5

2

(d) 1952 adaptive gridmechanism 2

-0.5 0 0.5 10

0.5

1

1.5

2

(e) 3852 adaptive gridmechanism 1

-0.5 0 0.5 10

0.5

1

1.5

2

(f) 3852 adaptive gridmechanism 2

Fig. 4. Comparison of calculated density contours and experimental obtained shock position (s) and combustion front (h) for H2/Airand M = 6.46 flow between Evans’ reaction mechanism [31] (left column) and Jachimowski’s mechanism [32] (right column) on 3852 fixed,1952 adaptive, and 3852 adaptive grids, respectively.

598 L. Yuan, T. Tang / Journal of Computational Physics 224 (2007) 587–600

Page 13: Resolving the shock-induced combustion by an adaptive mesh …lsec.cc.ac.cn/~lyuan/jcp07comb.pdf · 2007. 6. 2. · Resolving the shock-induced combustion by an adaptive mesh redistribution

0 1 2 30

0.5

1

1.5

2

2.5

3

(a) 513 2 adaptive gridmechanism 1

0 1 2 30

0.5

1

1.5

2

2.5

3

(b) 513 2 adaptive gridmechanism 2

Fig. 5. Comparisons of calculated density contours obtained with (a) reaction mechanism 1 [31] and (b) mechanism 2 [32] on 5132 adaptivegrid for H2/air M = 6.46 flow. The experimental shock is marked with (s) and the combustion front with (h).

L. Yuan, T. Tang / Journal of Computational Physics 224 (2007) 587–600 599

combustion. We are not sure if it is caused by some subtle interactions of chemical reactions with numericalerrors. The efficiency of an adaptive mesh method will then be limited by such interactions. We present thisdilemma to call for others’ attention.

6. Conclusions

We have developed an adaptive mesh redistribution method for numerical simulation of shock-inducedcombustion. We demonstrate the efficiency of the adaptive method in eliminating spurious runway chemicalreactions and obtaining grid-independent results. One of the main contributors to the high resolution of theadaptive grid method is the monitor function (41) which is based on the relative rate of change of ck’s. Thechoice of the monitor function is found extremely important for the shock-induced combustion problemsdue to large ratios of many variables involved. The proposed monitor function (41) can resolve the inductionzone between the shock wave and combustion front. Our numerical results on moderately fine grids are in fairagreement with classic experiments. Nevertheless, simulations of M = 3.55 case do not reproduce experimen-tally observed deflection of the combustion front and the secondary shock in the flank of the projectile. Fur-thermore, false combustion instabilities occur with very fine mesh or consecutive mesh adaptation forM = 6.46 case. These two unsolved problems appeal for further investigation.

Acknowledgments

We thank the referee for pointing out the equivalence relationship (40) which gives a useful explanation forthe monitor function (41). This work was supported by Natural Science Foundation of China (G10476032,G10531080), state key program for developing basic sciences (2005CB321703), CERG Grants of Hong KongResearch Grant Council and FRG Grants of Hong Kong Baptist University. The computation was conductedon Origin 3800 SMP computer at LSEC of Institute of Computational Mathematics, Chinese Academy ofSciences.

References

[1] H.F. Lehr, Experiments on shock-induced combustion, Astron. Acta 17 (1972) 589–597.[2] J.B. McVey, T.Y. Toong, Mechanism of instabilities of exothermic hypersonic blunt-body flow, Combust. Sci. Technol. 3 (1971) 63–79.[3] J.E. Shepherd, Detonation waves and propulsion, in: J. Buckmaster, T. Jackson, A. Kumar (Eds.), Combustion in High-speed Flows,

Kluwer, Dordrecht, The Nerthlands, 1994, pp. 373–420.[4] E.S. Oran, J.P. Boris, Numerical Simulation of Reactive Flow, Elsevier, New York, 1987 (Chapter 13).

Page 14: Resolving the shock-induced combustion by an adaptive mesh …lsec.cc.ac.cn/~lyuan/jcp07comb.pdf · 2007. 6. 2. · Resolving the shock-induced combustion by an adaptive mesh redistribution

600 L. Yuan, T. Tang / Journal of Computational Physics 224 (2007) 587–600

[5] P. Colella, A. Majda, V. Roytburd, Theoretical and numerical structure for reacting shock waves, SIAM J. Sci. Comput. 7 (1986)1059–1080.

[6] S. Yungster, S. Eberhardt, A. Bruckner, Numerical simulation of hypervelocity projectiles in detonable gases, AIAA J. 29 (1991) 187–199.

[7] V.N. Gamezo, E.S. Oran, Reaction-zone structure of a steady-state detonation wave in a cylindrical charge, Combust. Flame 109(1997) 253–265.

[8] G. Beckett, J.A. Mackenzie, M.L. Robertson, An r-adaptive finite element method for the solution of the two-dimensional phase-fieldequations, Commun. Comput. Phys 1 (2006) 805–826.

[9] M.J. Berger, R.J. LeVeque, Adaptive mesh refinement using wave-propagation algorithms for hyperbolic systems, SIAM J. Numer.Anal. 35 (1998) 2298–2316.

[10] Y. Di, P. Zhang, Moving mesh kinetic simulation for sheared rodlike polymers with high potential intensities, Commun. Comput.Phys. 1 (2006) 859–873.

[11] R. Li, T. Tang, P.-W. Zhang, Moving mesh methods in multiple dimensions based on harmonic maps, J. Comput. Phys. 170 (2001)562–588.

[12] F. Losasso, R. Fedkiw, S. Osher, Spatially adaptive techniques for level set methods and incompressible flow, Comput. Fluids 35(2006) 995–1010.

[13] P.A. Zegeling, On resistive MHD models with adaptive moving meshes, J. Sci. Comput. 24 (2005) 263–284.[14] X. Zeng, J. Lowengrub, A. Anderson, V. Cristini, Adaptive unstructured volume remeshing – II: application to two- and three-

dimensional level-set simulations of multiphase flow, J. Comput. Phys. 208 (2005) 626–650.[15] S. Singh, Y. Rastigejev, S. Paolucci, J. Powers, Viscous detonation in H2–O2–Ar using intrinsic low-dimensional manifolds and

wavelet adaptive multilevel representation, Combust. Theory Modelling 5 (2001) 163–184.[16] B. Bihari, D. Schwendeman, Multiresolution schemes for the reactive Euler equations, J. Comput. Phys. 154 (1999) 197–230.[17] R. Deiterding, Parallel adaptive simulation of multi-dimensional detonation structures, Ph.D. Thesis, Techn. U. Cottbus, September,

2003.[18] B. Azarenok, T. Tang, Second-order Godunov scheme for reactive flow calculations on moving meshes, J. Comput. Phys. 206 (2005)

48–80.[19] A. Matsuo, K. Fujii, Flow features of shock-induced combustion around projectile travelling at hypervelocities, AIAA J. 33 (1995)

1056–1063.[20] A. Harten, High resolution schemes for hyperbolic conservation laws, J. Comput. Phys. 49 (1983) 357–393.[21] H.C. Yee, A class of high-resolution explicit and implicit shock-capturing methods, NASA TM 101088, 1989 (unpublished).[22] J. Shuen, S. Yoon, Numerical study of chemically reacting flows using a lower–upper symmetric successive overrelaxation scheme,

AIAA J. 27 (1989) 1752–1760.[23] H.Z. Tang, T. Tang, Adaptive mesh methods for one- and two-dimensional hyperbolic conservation laws, SIAM J. Numer. Anal. 41

(2003) 487–515.[24] H.Z. Tang, T. Tang, P.-W. Zhang, An adaptive mesh redistribution method for nonlinear Hamilton–Jacobi equations in two- and

three dimensions, J. Comput. Phys. 188 (2003) 543–572.[25] J.F. Thompson, Z.U.A. Warsi, C.W. Mastin, Numerical Grid Generation: Foundations and Applications, North-Holland,

Amsterdam, 1985, Chapter IV. Available from http://www.erc.msstate.edu/publications/gridbook/chap06/index.html .[26] D.A. Anderson, Equidistribution schemes, Poisson generators, and adaptive grids, Appl. Math. Comput. 24 (1987) 211–227.[27] R.N. Gupta, J.M. Yos, R.A. Thompson, A review of reaction rates and thermodynamic and transport properties for the 11-species air

model for chemical and thermal nonequilibrium calculations to 30000K, NASA TM 101528, 1989.[28] T. Poinsot, D. Veynante, Theoretical and Numerical Combustion, R.T. Edwards, Inc., 2001.[29] R.J. Kee, F.M. Rupley, E. Meeks, J.A. Miller, Chemkin-III: a Fortran chemical kinetics package for the analysis of gas-phase

chemical and plasma kinetics, UC405, SAND 96-8216, 1996.[30] W.C. Gardiner Jr., Combustion Chemistry, Springer, New York, 1984.[31] J.S. Evans, C.J. Schexnayder, Influence of chemical kinetics and unmixedness on burning in supersonic hydrogen flames, AIAA J. 18

(1980) 188–193.[32] C.J. Jachimowski, An analytic study of the hydrogen-air reaction mechanism with application to scramjet combustion, NASA TP-

2791, February, 1988.[33] F.A. Greene, R.U. Gupta, Viscous equilibrium computations using LAURA, J. Spacecraft Rockets 29 (5) (1992) 627–632.[34] M. Vinokur, Flux Jacobian matrices and generalized Roe average for an equilibrium real gas, NASA CR-177512, 1988.[35] V.D. Liseikin, Grid Generation Methods, Springer, Berlin, 1999 (Chapter 7).[36] W.M. Cao, W.Z. Huang, R.D. Russell, An error indicator monitor function for an r-adaptive finite-element method, J. Comput.

Phys. 170 (2001) 871–892.[37] H.-Z. Tang, A moving mesh method for the Euler flow calculations using a directional monitor function, Commun. Comput. Phys. 1

(2006) 656–676.[38] W.M. Cao, W.Z. Huang, R.D. Russell, An r-adaptive finite element method based upon moving mesh PDEs, J. Comput. Phys. 149

(1999) 221–244.[39] R. Li, T. Tang, P.W. Zhang, A moving mesh finite element algorithm for singular problems in two and three space dimensions, J.

Comput. Phys. 177 (2002) 365–393.[40] K. Clutter, V. Krishnamurty, W. Shyy, Combustion and turbulent effects in hypersonic projectile flows, AIAA Paper 97-0807, 1997.