Skip to main content
Scientific Reports logoLink to Scientific Reports
. 2021 Oct 12;11:20225. doi: 10.1038/s41598-021-99163-6

Effect of small heat release and viscosity on thermal-diffusive instability

Keigo Wada 1,
PMCID: PMC8511008  PMID: 34642364

Abstract

The linear stability of a thermal reaction front has been investigated based on the thermal-diffusive model proposed by Zel’dovich and Frank-Kamenetskii, which is called ZFK model. In the framework of ZFK model, heat-conduction and mass-diffusion equations are treated without the effect of hydrodynamic flow. Then, two types of instability appear: cellular and oscillatory instabilities. The cellular instability has only positive real part of growth rate, while the oscillatory instability is accompanied with non-zero imaginary part. In this study, the effect of heat release and viscosity on both instabilities is investigated asymptotically and numerically. This is achieved by coupling mass-conservation and Navier–Stokes equations with the ZFK model for small heat release. Then, the stable range of Lewis number, where the real part of growth rate is negative, is widened by non-zero values of heat release for any wavenumber. The increase of Prandtl number also brings the stabilization effect on the oscillatory instability. However, as for the cellular instability, the viscosity leads to the destabilization effect for small wavenumbers, opposed to its stabilization effect for moderate values of wavenumber. Under the limit of small wavenumber, the property of viscosity becomes clear in view of cut-off wavenumber, which makes the real part of growth rate zero.

Subject terms: Engineering, Mathematics and computing, Physics

Introduction

The analysis of thermal propagation of a flame front has been performed so far based on the thermal-diffusive model, or the ZFK model, proposed by Zel’dovich and Frank-Kamenetskii1,2. In the framework of ZFK model, the effect of hydrodynamic flow is successfully decoupled from the heat and mass-diffusion equations. This is achieved by the assumption of small difference between the values of density, or temperature, evaluated at burned and unburned edges of reaction front3. In this case, we only need to treat the heat-conduction and mass-diffusion equations accompanied with the exothermic reaction term, which is usually expressed by the Arrhenius type46. The thickness of reaction front is characterized by the inverse of Zel’dovich number7. By assuming the large values of Zel’dovich number, the analysis of inner structure of reaction front, where the reaction term is confined, becomes possible8,9. The values of Zel’dovich number larger than 2 has been found to be valid in the context of burning-rate eigenvalue problem1013. The inner structure of reaction front has been investigated based on the method of matched asymptotic expansions14. Then, the boundary conditions across the reaction front are obtained for temperature and mass fraction. We do not pursue the analysis of previous research, and instead, utilize the boundary conditions already derived15. We note that the hydrodynamic flow is also considered in our study. Therefore, the boundary conditions of hydrodynamic quantities across the reaction front are also required and they are presented in the previous work16.

The linear stability analysis of the thermal-diffusive model without hydrodynamic flow has been performed for one reactant17, for two reactants18, for the case of heat loss19 and in the context of ignition temperature20. The nonlinear treatment of this model is also discussed based on Hopf bifurcation21. The thermal-diffusive instability is categorized into two types, the cellular and oscillatory instabilities, depending on values of Lewis number22. The stability boundary of thermal-diffusive instability is usually determined with respect to Lewis number and wavenumber. The cellular instability is observed for some range of Lewis number less than unity. The stability boundary of cellular instability shows monotonic decrease of Lewis number as the wavenumber increases. On the other hand, that of oscillatory instability is characterized by the non-monotonic behavior of Lewis number as a function of wavenumber. These features are obtained as a result of neglecting the effect of thermal expansion, or of no heat release. Our result shows that, for the stability boundary of cellular instability, the existence of heat release changes the monotonic behavior of Lewis number to the non-monotonic one.

This study focuses on the pure thermal-diffusive instability. Therefore, the boundary conditions across the reaction front are different from those used in the hydrodynamic model. Strictly speaking, in the hydrodynamic model, a flame front is treated as a zero-thickness interface and the basic flow, which is assumed to be steady one-dimensional flow, is described by the constant density and velocity, which are discontinuous across the front. Such a discontinuity is the intrinsic property of hydrodynamic instability, or the Darrieus–Landau instability (DLI): the DLI claims that the growth rate of perturbations is an linear function of wavenumber, whose coefficient is positive due to the heat release, and a planar flame front is unstable for any wavenumber. As a consequence of the jump of normal velocity across a flame front, the tangential components of velocity field induced by small perturbations are also discontinuous. This fact is essential for the existence of DLI. However, our study is performed in the scale of thermal thickness of a flame front, or in the framework of thermal-diffusive model. Then, the boundary conditions across the reaction front, which have been derived based on the large activation energy asymptotics, are described by the continuity of all quantities except for the pressure, whose discontinuity is supported by the viscosity. Due to the continuity of normal velocity, the tangential components of velocity becomes continuous for small perturbations. For a basic flow, the reaction front is considered as a density continuous interface, whereas the derivative of density is discontinuous. As a result, the DLI does not appear in our treatment.

In this study, we consider the case of small heat release. Then, we need to deal with not only the thermal-diffusive model but also hydrodynamic one. So far, the coupling of these two models has been achieved by the multi-scale analysis based on the thermal thickness of a flame front, or by introducing the preheat zone inside the front2326. This approach successfully combines the DLI2729 with the Markstein effect which reflects the thermal and mass diffusivities3032. The numerical study to understand the dynamics of a flame front has been performed for cellular and non-cellular fronts33,34 and for cellular front with heat loss3537. Despite of many works on the instability of a flame front, the influence of heat release and viscosity on the stability boundaries of cellular and oscillatory instabilities has not ever been declared clearly. Besides, the effect of higher order terms of wavenumber, such as third and forth order terms, on the growth rate has not been discussed in detail. The existence of third order term of wavenumber has the possibility of leading to two cut-off wavenumbers, opposed to the case of zero heat release, where the cut-off wavenumber is determined uniquely: the cut-off wavenumber means the value of wavenumber at which the growth rate of perturbations changes its sign. In order to reveal these influence, we directly examine the linear stability of reaction front for infinitesimal perturbations with the heat release being a small parameter. Besides, the calculation of numerical solutions is performed based on the TVD-MacCormack scheme3844.

Methods

We employ the following non-dimensional governing equations13,25

ρt+·(ρV)=0, 1
ρVt+(V·)V=-1γMa2PM+PrΔV+13(·V), 2
ρTt+(V·)T=ΔT+qQ, 3
ρYt+(V·)Y=1LeΔY-Q, 4
P=ρT, 5

where the non-dimensional variables, ρ, V=(U,V,W), T, PM and Y are the density of mixture, velocity field, temperature, pressure and mass fraction of a deficient species which is consumed through the exothermic chemical reaction. The parameters γ, Ma, Pr, q and Le are the specific heat ratio, Mach number, Prandtl number, heat release and Lewis number, respectively. The dependent variables are made dimensionless by use of those of fresh mixture at a position far from a flame front, where the flow field is assumed to be steady planar, denoted by ρ~-, S~L, T~-, P~- and Y~-. Especially, S~L is called the laminar flame speed, or the burning velocity, which is the propagation velocity of a planar flame front into a quiescent fuel mixture13. Hereafter, the variables with tilde symbol stand for dimensional ones.

ρ=ρ~ρ~-,T=T~T~-,PM=P~P~-,Y=Y~Y~-,V=V~S~L,l~d=D~thS~L,Le=D~thD~,Pr=μ~c~pλ~,D~th=λ~ρ~-c~p,Ma=S~Lc~s. 6

The parameters l~d, D~th, D~, μ~, c~p and λ~ are the thermal thickness of a flame front, thermal diffusivity, diffusion coefficient of a deficient species, coefficient of viscosity, specific heat at constant pressure and thermal conductivity. We assume that D~, μ~, c~p and λ~ are all constant. The adiabatic sound speed is defined by c~s=(γP~-/ρ~-)1/2. We normalize the coordinates x~ and the time t~, by use of l~d and S~L, as

x=x~/l~d,t=t~S~L/l~d. 7

We introduce the Cartesian coordinate system x=(x,y,z), with an unperturbed steady planar reaction front lying on the xy plane and propagating in the negative direction of z-axis.

According to the previous work25, the reaction term Q is expressed by the following form.

Q=ΛρYε2expTbεT-TbT. 8

The parameter ε is defined by

ε=Tb2N=qβ, 9

where Tb=1+q and N=E~/R~gT~- are the adiabatic temperature, which is evaluated on the burned side of a reaction front, and the non-dimensional activation energy with the dimensional activation energy E~ and universal gas constant R~g. The Zel’dovich number β=E~(T~b-T~-)/(R~gT~b2) will be employed to compare our result with the one obtained in previous works3,32,45. Under the limit of large activation energy (N+), ε becomes a small parameter which characterizes the thickness of reaction front. Although the burning-rate eigenvalue Λ is important to determine the laminar flame speed4,5,1012, its treatment is out of scope of this work. In order to investigate the effect of heat release and viscosity on the thermal-diffusive instability, we only need to utilize the jump conditions across the reaction front derived in previous works by considering the inner structure of the front15,16.

Coordinate system attached to a flame front

We recall the coordinate system attached to the flame front, z=F(x,y,t), by introducing the following variables25.

x=x,y=y,ξ=z-F(x,y,t),t=t. 10

By use of (10), governing Eqs. (1)–(4) are rewritten as

ρt+t·(ρVt)+Mξ=0, 11
ρVtt+ρ(Vt·t)Vt+MVtξ=-tPMγMa2+tFγMa2PMξ+PrΔξVt+13t-tFξt·Vt+Sξ, 12
ρWt+ρ(Vt·t)W+MWξ=-1γMa2PMξ+PrΔξW+13ξt·Vt+Sξ, 13
ρTt+ρ(Vt·t)T+MTξ=ΔξT+qQ, 14
ρYt+ρ(Vt·t)Y+MYξ=1LeΔξY-Q, 15

where S denotes the longitudinal velocity of gases relative to a flame front.

S=W-Ft-(Vt·t)F. 16

The tangential components of velocity field are expressed by Vt=(U,V,0). Based on (16), the mass flux is defined by

M=ρS. 17

The transverse derivative is denoted by t=(/x,/y,0), and we also introduce the notation t2=2/x2+2/y2. Then, the Laplacian is written as

Δξ=(1+|tF|2)2ξ2+t2-(t2F)ξ-2(tF)·tξ. 18

The unit normal on the reaction layer, which is directed toward the burned side (ξ>0), is defined by

n=ξ|ξ|=11+|tF|2-Fx,-Fy,1. 19

Boundary conditions

In the far field from a flame front on the unburned side, where the flow is assumed to be steady planar, all quantities become unity due to the definition (6).

ρ=T=PM=Y=W=1,U=V=0asξ-. 20

The deficient component of combustible mixture is assumed to be depleted during the exothermic chemical reaction inside the reaction front. Then, the mass fraction Y is equivalently zero on the burned side.

Y0(ξ>0). 21

In this study, we employ the zero-Mach-number approximation, Ma1. Under this limit, as readily found from (2), the pressure is assumed to have the following asymptotic form22.

PM=1+γMa2P. 22

In the subsequent discussion, we consider the variation of pressure at O(Ma2).

In order to know the profile of physical quantities outside the reaction front, we need the boundary conditions on it. In the limit of large activation energy, ε(=q/β)1, the following jump conditions across the reaction front have been derived15,16.

[(n·)M]=0, 23
[M]=[V·n]=P-43Pr(n·)(V·n)=0, 24
[n×(V×n)]=0, 25
[T]=[Y]=(n·)T+qLeY=0, 26
(n·)T|ξ=0-=qexpT|ξ=0+-Tb2ε, 27

where the brackets for arbitrary function Φ(x,y,ξ,t) means the difference of its values on the burned and unburned sides, [Φ]=Φ|ξ=0+-Φ|ξ=0-.

In the subsequent calculations, the boundary conditions (20), (21), (23)–(27) are imposed to determine the integral constants of unperturbed and perturbed solutions of (1)–(5) for a weakly corrugated front, whose amplitude is infinitesimally small.

Basic flow

As a basic flow without perturbations, we consider a steady planar flow. The notation of over-bar is used to denote this flow field. For example, the velocity of basic flow is denoted by W¯=W¯(ξ). In this case, the jump conditions (24)–(26) are given by

[M¯]=[W¯]=P¯-43PrdW¯dξ=[T¯]=[Y¯]=ddξT¯+qLeY¯=0. 28

We note that (23) and (27) are automatically satisfied by the solutions obtained by use of (28). For a steady planar flow, the mass conservation Eq. (11) becomes

dM¯dξ=0. 29

From (29), by use of (20) and (28), we readily find that the steady planar mass flux is unity on both sides of reaction front.

M¯=1. 30

Then, the other governing Eqs. (5), (13)–(15) are written down, outside the reaction layer (ξ<0, ξ>0), as

dW¯dξ=-dP¯dξ+43Prd2W¯dξ2, 31
dT¯dξ=d2T¯dξ2, 32
dY¯dξ=1Led2Y¯dξ2, 33
1=ρ¯T¯. 34

We note that the reaction term, included in (14) and (15), is confined inside the reaction front due to the assumption of large activation energy, ε1, with T<Tb in the zero-Mach-number limit25. Subject to (20), (21) and (28), the solutions of unperturbed governing equations (31)–(34) are obtained as

T¯=1+qeξ(ξ<0)1+q(ξ>0),Y¯=1-eLeξ(ξ<0)0(ξ>0),P¯=43Pr-1qeξ(ξ<0)-q(ξ>0),ρ¯=1T¯,W¯=T¯. 35

These solutions are used to calculate small perturbations for a weakly corrugated reaction front below.

Formulation of perturbations

We consider the situation where the reaction front is slightly deviated from its unperturbed state. We assume that the heat release is sufficiently small.

q1. 36

Based on (36), the amplitude of weakly corrugated front is given in the form of harmonic perturbation by

F=q2eikxx+ikyy+ωt, 37

where ω is the growth rate and the wavenumber k is defined by k=(kx2+ky2)1/2. The non-dimensionalization is made as ω=ω~l~d/S~L and k=k~l~d with ω~ and k~ dimensional quantities.

We impose small perturbations on the basic flow as follows.

ρ=ρ¯+ρ,W=W¯+W,Vt=Vt,P=P¯+P,T=T¯+T,Y=Y¯+Y, 38

where the order of each perturbation is assumed to be

ρ=ρp(ξ)qF,W=Wp(ξ)qF,Vt=(Up(ξ),Vp(ξ),0)qF,P=Pp(ξ)qF,T=Tp(ξ)qFO(q3), 39
Y=Yp(ξ)FO(q2). 40

Upon (37)–(40), the perturbed governing equations have the following form with O(q4) terms omitted due to (36).

t+ξρ-Ftdρ¯dξ+t·Vt+Wξ=0, 41
t+ξVt=-tP+(tF)dP¯dξ+Prt2+2ξ2Vt+Pr13tt·Vt+Wξ-13(tF)d2W¯dξ2, 42
t+ξW-FtdW¯dξ=-Pξ+Prt2+2ξ2W+Pr-(t2F)dW¯dξ+13ξt·Vt+Wξ, 43
t+ξT-FtdT¯dξ=t2+2ξ2T-(t2F)dT¯dξ, 44
t+ξY-(1-qeξ)FtdY¯dξ+ρ+WdY¯dξ=1Let2+2ξ2Y-1Le(t2F)dY¯dξ, 45
0=ρ+T. 46

As is readily confirmed from (35), the derivatives of ρ¯, P¯, W¯ and T¯ are O(q) and that of Y¯ is O(1). Unlike the ordinary thermal-diffusive model, the influence of flow field appears in (45) due to the existence of small heat release. In this case, the coupling of hydrodynamic equations and thermal-diffusive ones is possible because the heat equation remains unchanged from the ordinary one and solvable. The detail of calculation is given in the succeeding subsection.

Calculation of perturbations

We calculate perturbations (39) and (40) by solving perturbed governing Eqs. (41)–(46). From (46), we find that once the perturbation of temperature is calculated, that of density is obtained.

ρp=-Tp. 47

Substituting (35), (37) and (39) into the perturbed heat-conduction Eq. (44), the following differential equation is obtained.

d2Tpdξ2-dTpdξ-(ω+k2)Tp=-ω+k2eξ(ξ<0)0(ξ>0). 48

Assuming Re[ω]>0, the requirement that perturbations are bounded in space leads to the solution of (48) as

Tp=θmeλ+ξ+eξ(ξ<0)θpeλ-ξ(ξ>0), 49

where

λ±=1±1+4(ω+k2)2. 50

We note that the signs are in the same order. The integral constants θm and θp are determined in the succeeding subsection.

Next, in order to calculate Pp, we transform the perturbed mass conservation Eq. (41) into

t·Vt+Wξ=t+ξT-qeξFt, 51

where (35) and (46) are used and higher order terms with respect to q are omitted due to (36). Then, taking the transverse derivative t· of (42) and differentiating (43) with respect to ξ, and combining them with the help of (51), we obtain

t2+2ξ2P=43Prt2+2ξ2-t-ξt+ξT-qeξFt+Ft-43Pr(t2F)d2W¯dξ2+(t2F)dP¯dξ. 52

By use of (35), (37) and (39), the combined equation (52) becomes the following differential equation for Pp.

d2Ppdξ2-k2Pp=43Pr-1(λ++ω)2θmeλ+ξ+1-k2eξ(ξ<0)43Pr-1(λ-+ω)2θpeλ-ξ(ξ>0). 53

Solving (53), we obtain

Pp=pmekξ+43Pr-1(λ++ω)θmeλ+ξ+43Pr-1eξ(ξ<0)ppe-kξ+43Pr-1(λ-+ω)θpeλ-ξ(ξ>0), 54

where pm and pp are integral constants.

The perturbations of velocity field are calculated by substituting (35), (37), (39), (49), (51) and (54) into the perturbed Navier–Stokes equations (42) and (43) as follows.

Up=umel+ξ-ikxω+kpmekξ+ikxθmeλ+ξ(ξ<0)upel-ξ-ikxω-kppe-kξ+ikxθpeλ-ξ(ξ>0), 55
Vp=vmel+ξ-ikyω+kpmekξ+ikyθmeλ+ξ(ξ<0)vpel-ξ-ikyω-kppe-kξ+ikyθpeλ-ξ(ξ>0), 56
Wp=wmel+ξ-kω+kpmekξ+λ+θmeλ+ξ+eξ(ξ<0)wpel-ξ+kω-kppe-kξ+λ-θpeλ-ξ(ξ>0), 57

where

l±=1±1+4Pr(ω+Prk2)2Pr, 58

and um, up, vm, vp, wm and wp are integral constants.

The perturbation of mass fraction is calculated by integrating (45) by use of (35), (37), (40), (47), (49) and (57).

Yp=cmeΓ+ξ-LeeLeξ-qLe2ω1-k2+Le(1-ω)e(1+Le)ξ-qLe2(λ+-1)(1+Le)λ++(1-Le)ωθme(λ++Le)ξ-qLe2Pr(1+LePr)l++(1-LePr)ωwme(l++Le)ξ+qLekk2-ω2pme(k+Le)ξ(ξ<0)0(ξ>0), 59

where

Γ±=Le1±1+4(ω+k2/Le)/Le2, 60

and cm is an integral constant. Unlike the classical thermal-diffusive model, the solution of perturbed mass fraction (59) includes the terms reflecting the effect of hydrodynamic flow due to non-zero values of heat release q. Thanks to these terms, it becomes possible to incorporate the influence of heat release and viscosity into the thermal-diffusive instability.

At last, by substituting solutions (49), (55)–(57) into the mass conservation equation (51), we obtain the following relations.

ikxum+ikyvm+l+wm=ikxup+ikyvp+l-wp=0. 61

Dispersion relation

In this subsection, we calculate the dispersion relation of thermal-diffusive instability, which represents the relation between the growth rate and wavenumber for perturbations, by substituting the solutions obtained above into the boundary conditions across the reaction front. For the perturbed quantities, the boundary conditions (23)–(27) are written down as

dρ¯dξ|-(W|--ωF)+ρ¯|-Wξ|-+ρξ|-W¯|-+ρ|-dW¯dξ|-=ρ¯|+Wξ|++ρξ|+W¯|+, 62
U|-=U|+,V|-=V|+, 63
W|-=W|+, 64
P|--43PrWξ|-=P|+-43PrWξ|+, 65
T|-=T|+,Tξ|-=β2T|-, 66
Y|-=0, 67
Tξ|-+1LeYξ|-=Tξ|+. 68

We note that the continuity condition of mass flux in (24) leads to the continuity of density perturbation due to (64), but this gives no information. In fact, this is the same condition as the first one in (66) because of perturbed equation of state (46). From (61)–(67), we determine integral constants of perturbations. After that, (68) is used to calculate the dispersion relation of thermal-diffusive instability including the effect of small heat release accompanied with the viscosity.

At first, the integral constants for thermal-diffusive quantities are easily determined. Indeed, substituting (49) into (66), we obtain

θm+1=θp,λ+θm+1=β2θp. 69

Then, integral constants θm and θp of temperature perturbation are determined by solving (69) as

θm=β-22λ+-β,θp=2λ+-12λ+-β. 70

The expression of cm is obtained from (67), by use of (59), as

cm=Le+qωLe21-k2+Le(1-ω)+qLe2(λ+-1)(1+Le)λ++(1-Le)ωθm+qLe2Pr(1+LePr)l++(1-LePr)ωwm-qLekk2-ω2pm. 71

If q=0, the above solutions (70) and (71) coincides with those in the classical thermal-diffusive model22,32.

Next, the integral constants for hydrodynamic flow field are determined. The substitution of (55) and (56) into (63) gives

um-ikxω+kpm+ikxθm=up-ikxω-kpp+ikxθp,vm-ikyω+kpm+ikyθm=vp-ikyω-kpp+ikyθp. 72

From (61) and (72), um, up, vm and vp are eliminated to give the following relation.

-l+wm+k2ω+kpm-k2θm=-l-wp+k2ω-kpp-k2θp. 73

The continuity of normal velocity (64), with (57) substituted, leads to

wm-kω+kpm+λ+θm+1=wp+kω-kpp+λ-θp. 74

Solving (73) and (74), we find

pm=ω+k2k2(k+l+)wm-(k+l-)wp+(λ++k)kθm-(λ-+k)kθp+k, 75
pp=ω-k2k2(k-l+)wm-(k-l-)wp+(λ+-k)kθm-(λ--k)kθp+k. 76

By use of solutions (54) and (57), the boundary condition of pressure (65) is transformed into

pm+43Pr-1(λ++ω)θm+1-43Prl+wm-k2ω+kpm+λ+2θm+1=pp+43Pr-1(λ-+ω)θp-43Prl-wp-k2ω-kpp+λ-2θp. 77

From (75)–(77), we obtain

wp=l+ω+k2l-ω+k2wm. 78

The continuity condition of derivative of mass flux (62) is reduced, by use of (50), (69) and (73) while being careful about 1/(1+q)1-q+O(q2), to

wm-kω+kpm+3λ+ω+1θm+2ω+4=(2λ-+ω)θp. 79

Substituting (70) and (75) into (79), with the help of (78), we get

wm=l-ω+k2(ω-k)(l--l+)-2(l-ω+k2)22λ+-β×-β2(5λ-+k+2ω)+λ-+(k+2ω)λ++3(ω+k2). 80

Finally, by use of (49), (59), (70), (71) and (80), the dispersion relation is calculated from (68) as follows.

(λ+-λ-)1+q2k+Γ-k-ω+qLeλ-(λ++Γ-)(1+Le)λ++(1-Le)ωβ-22λ+-β+λ+-Γ-+q2k+Γ-k-ωλ+-k-(1+Γ-)qLeω1-k2+Le(1-ω)-qLePr(l++Γ-)(1+LePr)l++(1-LePr)ω+k+Γ-2l--l+l-ω+k2wm=0. 81

The numerical and analytical studies of (81) are given below.

Asymptotic results

The case of q=0

In the pure thermal-diffusive model, or the ZFK model1,2, the effect of hydrodynamic flow was effectively decoupled from the heat and mass-diffusion equations. This situation is easily recovered by taking q=0 in (81). It may be helpful to recall the original dispersion relation for thermal-diffusive instability3,32, which is given by

(λ+-λ-)β-22λ+-β+λ+-Γ-=0. 82

In addition, resorting to the assumption that the Lewis number is nearly equal to unity, which is acceptable for real gases6,13, we introduce O(1) constant le as

Le=1+leβ, 83

under the limit of large Zel’dovich number, β1. By use of (83), the dispersion relation (82) is transformed into the following form.

1-1+4(ω+k2)(1+4(ω+k2))=-le21+2ω-1+4(ω+k2). 84

The numerical calculation of (84) has been shown and summarized by many works3,22,32. The cellular instability is observed for some range of Lewis number less than unity and the oscillatory one appears for large values of Lewis number. The cellular instability is characterized by the positive values of real part of growth rate without imaginary part and the oscillatory one is by those with imaginary part. This tendency remains unchanged for the case of non-zero values of heat release as discussed below.

Computation of dispersion relation

We consider the case of q0. In order to compare our result with the previous research, we apply the assumption (83) to the dispersion relation (81) under β1. Then, we obtain the reduced dispersion relation, with qβ=O(1) and O(q) terms omitted, as follows.

ω+λ-1-2λ+le+2λ-(1-2λ+)-qβ2k+λ-k-ω+λ-ω+k2+2ω(1+λ-)ω+k2-2+qβk+2ω+5λ-2kPr(l+-k)-k-ωk+λ-2(1-2Prl+)+Pr2(l++λ-)(k2+l-ω)l+(1+Pr)-(Pr-1)ω=0. 85

The numerical results of (85) are shown in Figs. 1, 2, 3 and 4 for β=10. The stable range of Lewis number, where the real part of growth rate takes negative values, is greatly altered from that in the previous case of q=0 due to the heat release accompanied with the viscous effect. At first, from Figs. 1 and 3, the inclusion of heat release is found to extend the stable domain for both cellular and oscillatory instabilities. In addition, Fig. 4 shows that the non-zero values of Prandtl number, or the viscous effect, also broadens the stable range of Lewis number for the oscillatory instability. Contrary to such stabilization effect, it should be emphasized that the viscosity has the destabilizing effect on the cellular instability if the wavenumber is small, as shown in Fig. 2. For some moderate values of wavenumber, the cellular instability is stabilized by the viscosity.

Figure 1.

Figure 1

Stability boundary of cellular instability for several values of heat release, q=0 (solid line), 0.05 (dashed), 0.1 (dot-dashed) and 0.15 (dotted), with β=10 and Pr=0.

Figure 2.

Figure 2

The same as Fig. 1 but for two values of Prandtl number, Pr=0 (thin lines) and 3/4 (thick lines).

Figure 3.

Figure 3

Stability boundary of oscillatory instability for several values of heat release, q=0 (solid line), 0.05 (dashed) and 0.1 (dot-dashed), with β=10 and Pr=0.

Figure 4.

Figure 4

The same as Fig. 3 but for two values of Prandtl number, Pr=0 (thin lines) and 3/4 (thick lines).

Cellular instability for small wavenumber and growth rate

As discussed above, we find that the viscosity brings the destabilization effect on the cellular instability for small values of wavenumber, opposed to the stabilizing one for finite values. Therefore, it may be meaningful to investigate the dispersion relation (81) in detail for small wavenumbers to reveal the destabilization effect of viscosity. For this purpose, we consider two cases: ωk21 and ωk41. Especially, in the latter case, our result contains the new effect of third order term with respect to wavenumber in the reduced dispersion relation.

The case of ωk21

At first, the effect of wavenumber is examined up to second order term, which corresponds to the Markstein effect reflecting the curvature effect30,31,45. By assuming that ωk21, the dispersion relation (81) is reduced to

2(1+qβ)ω+2+(Le-1)βLe+2qβ+qβ(1-Pr)(1+Le)(1+LePr)k2=0. 86

For q=0, the above Eq. (86) is consistent with that obtained in the previous work45. In addition, assuming (83) for β1 with qβ=O(1), the reduced dispersion relation (86) gives the following growth rate.

ω=-k22(1+qβ)2+le+qβ25+3Pr1+Pr. 87

The growth rate (87) clearly shows how the variation of le in the range of small wavenumber influences the cellular instability with small heat release. The stability condition of (87), or ω<0, is given by le>-2-qβ(5+3Pr)/(2+2Pr). The stability boundary le=-2, which is obtained in the previous research3,32,45, is extended by the inclusion of heat release q accompanied with the viscous effect represented by Prandtl number Pr. The increase of q leads to the relaxation of lower bound of le. On the other hand, the rise of Pr narrows the stable range of le. This result explains the destabilizing effect of viscosity for small wavenumber as implied by the numerical computation in Fig. 2.

The case of ωk41

Next, we proceed to higher order terms of wavenumber, O(k3) and O(k4) terms. We note that O(k3) term is intrinsic to the effect of heat release and not included in the original dispersion relation of thermal-diffusive model3,32. In the present case, the cut-off wavenumber is no longer determined uniquely and there is the possibility of two cut-off wavenumbers depending on values of Lewis number.

Based on the assumption of ωk41, the dispersion relation (81) is transformed into

2(1+qβ)ω+2+(Le-1)βLe+2qβ+qβ(1-Pr)(1+Le)(1+LePr)k2+qβLe2-Pr-5-2Pr1+LePrk3+2Le-(Le3-1)(β-2)Le3-2qβ-qβPrLe(5-2Pr)(1-LePr)-LePr21+LePrk4=0. 88

Solving the reduced dispersion relation (88), by use of the assumption (83) for β1 with qβ=O(1), we find

ω=12(1+qβ)(ω2k2+ω3k3+ω4k4), 89

with

ω2=-2+le+qβ25+3Pr1+Pr,ω3=qβPr+3-4Pr1+Pr0,ω4=3le-2+qβ2+7Pr(1-Pr)+Pr31+Pr.

We note that the Prandtl number Pr typically takes positive and small values: for example, Pr3/4 in air13. For q=0, the O(k4) term in (89) is consistent with previous work45, which treated the case of le-2, that is ω4k4/2-4k4. For non-zero values of heat release, as confirmed from ω2 and ω3, the rise of Prandtl number increases the growth rate at second order of wavenumber and decreases it at third order. Because the wavenumber is assumed to be small, the viscous effect is not so important at the forth order of wavenumber. By setting ω=0 in (89), the cut-off wavenumber is obtained as

kc=-ω3±ω32-4ω2ω42ω4. 90

In the case of q=0, the cut-off wavenumber is uniquely determined as kc=(le+2)/(3le-2). The set up of cellular instability is given by le=-2 or 2/3. The positive value of growth rate for small wavenumbers, or long wavelengths, is observed for le<-2 and that for large wavenumber, or short wavelength, is observed for le>2/3. For -2<le<2/3, the growth rate is negative for arbitrary wavenumber.

Although (90) implies the existence of instability at some finite value of wavenumber for le>2/3, it should be omitted because the wavenumber is now assumed to be small. In fact, even if the value of le is sufficiently large, kc is not so small: kc1/3+0 as le+.

The new tendency of cellular instability due to the existence of heat release is characterized by ω3 in (89). This term promotes the cellular instability accompanied with two cut-off wavenumbers for some range of le. The stability boundary of le for the cellular instability is obtained from (90) as follows.

le2<le<le3(no cut-off wavenumber), 91
le<le1,le>le4(one cut-off wavenumber), 92
le1<le<le2,le3<le<le4(two cut-off wavenumbers), 93

with

le1=-2-qβ25+3Pr1+Pr, 94
le2=-23-qβ122Pr3-14Pr2+23Pr+191+Pr-43-qβ122Pr3-14Pr2+5Pr-111+Pr2-ω3212, 95
le3=-23-qβ122Pr3-14Pr2+23Pr+191+Pr+43-qβ122Pr3-14Pr2+5Pr-111+Pr2-ω3212, 96
le4=23-qβ32+7Pr(1-Pr)+Pr31+Pr, 97

where le2 and le3 are the solutions of ω32-4ω2ω4=0 and determine the cut-off wavenumber which induces the cellular instability. We note that le1=le2=-2 and le3=le4=2/3 for q=0.

In the stable range of le given by (91), the dispersion relation (89) shows negative values of ω for arbitrary wavenumber and the perturbations disappear in time. On the other hand, in the case of (92), the positive value of growth rate for small wavenumbers is observed for le<le1 and that for large wavenumbers is observed for le>le4. Besides, the new feature of instability due to the existence of heat release is observed in the range of (93). The increase of Pr narrows the range of (93), where two cut-off wavenumbers exist, and decreases the values of kc. The cases of le>le4 and le3<le<le4 are omitted because of the assumption k41. In these cases, the cut-off wavenumbers does not take small values.

The growth rate in the case of (93) is shown in Figs. 5 and 6. The existence of viscosity raises the value of le2, which implies the Lewis number approaches unity. Therefore, the cellular instability tends to be promoted by the viscosity for small wavenumbers. This fact is consistent with the result obtained numerically in Fig. 2. In the case of (92), it is clear that the viscous effect has two aspects on the instability as plotted in Fig. 7. Although the viscosity enhances the growth rate for small wavenumbers, it shrinks the range of wavenumber where the instability exists. For k0, by setting ω=0 in (89), the stability boundary with respect to the values of Lewis number and Prandtl number is given by ω2=0 and shown in Fig. 8. The destabilization due to the viscosity for small wavenumbers is easily comprehended from Fig. 8. Finally, the dependence of cut-off wavenumber on the Prandtl number is shown in Fig. 9. It is apparent that, among two cut-off wavenumbers, the smaller one is reduced due to the viscosity. In contrast, as for the larger cut-off wavenumber, the rise of Pr brings either increase or decrease of it subject to the values of le.

Figure 5.

Figure 5

Growth rate ω v.s. wavenumber k for several values of le, with β=10, q=0.1 and Pr=0. In this case, le2-4.327 and kc0.0538,0.175.

Figure 6.

Figure 6

The same as Fig. 5 but for Pr=3/4. In this case, le2-4.06 and kc0.022,0.0402.

Figure 7.

Figure 7

Growth rate ω v.s. wavenumber k for two values of Prandtl number Pr=0 and 3/4, with β=10, q=0.1 and le=le1(=-4.5).

Figure 8.

Figure 8

Stability boundary with respect to Lewis number and Prandtl number for small wavenumber with β=10. The values of le less than each curve implies the existence of cellular instability for each value of q.

Figure 9.

Figure 9

Cut-off wavenumber kc v.s. Prandtl number Pr for various values of reduced Lewis number le with β=10 and q=0.1. The solid and dashed lines reflect the signs in (90).

Numerical study

Discretization

In this section, the numerical solutions of governing Eqs. (1)–(5) are computed by use of the TVD-MacCormack scheme3844. For simplicity, the two-dimensional flow field in xz plane is considered, where the reaction front propagates in the negative z-direction. In order to employ the MacCormack scheme39,40, (1)–(5) are rewritten into the conservation form:

Ut+Ex+Fz=S, 98

where the flux vectors are

U=ρρUρWρTρY,E=ρUρU2+1γMa2ρT-Pr43Ux-23WzρUW-PrWx+UzρTU-TxρYU-1LeYx, 99
S=000qAρYexp-(1+q)2qβT-AρYexp-(1+q)2qβT,F=ρWρUW-PrWx+UzρW2+1γMa2ρT-Pr43Wz-23UxρTW-TzρYW-1LeYz. 100

Recalling (8) and (9), the coefficient of reaction term is expressed by A=q-2β2Λexp(1+q-1)β. In addition, utilizing the result from burning-rate eigenvalue problem under large-activation-energy asymptotics3,12,13, we may estimate Λq2Tb/(2Le). Therefore, the coefficient A is approximated by

A=1+q2Leβ2exp(1+q-1)β. 101

In this study, the time-splitting method38 is used to calculate numerical solutions of (98). Then, (98) is divided into two one-dimensional problems corresponding to x- and z-directions:

Ut+Ex=S2,Ut+Fz=S2. 102

The one-dimensional explicit MacCormack scheme is applied to solve each equation in (102), respectively. The MacCormack scheme has second-order accuracy in both time and space. The procedure of discretization is expressed by finite-difference operators Lx(Δt) and Lz(Δt)4648. Each operator consists of two steps: the predictor and corrector steps. For each equation in (102), the predictor step is

Ui,jp=Ui,jn-ΔtEi+1,jn-Ei,jnΔx-Si,j2,Ui,jp=Ui,jn-ΔtFi,j+1n-Fi,jnΔz-Si,j2, 103

and the corrector step is

Ui,jc=Ui,jn+Ui,jp2-Δt2Ei,jp-Ei-1,jpΔx-Si,jp2,Ui,jc=Ui,jn+Ui,jp2-Δt2Fi,jp-Fi,j-1pΔz-Si,jp2, 104

where Ep and Fp are calculated by use of Up. The superscript n implies the time step, and the subscripts i and j imply the spatial steps in x- and z-directions, respectively.

Each component of flux vectors is discretized in space as follows.

Ui,j=ρi,j(ρU)i,j(ρW)i,j(ρT)i,j(ρY)i,j,Ei,j=(ρU)i,j(ρU2)i,j+1γMa2Pi,j-Pr43Ui,j-Ui-1,jΔx-23Wi,j+1-Wi,j-12Δz(ρUW)i,j-PrWi,j-Wi-1,jΔx+Ui,j+1-Ui,j-12Δz(ρTU)i,j-Ti,j-Ti-1,jΔx(ρYU)i,j-1LeYi,j-Yi-1,jΔx, 105
Si,j=000qA(ρY)i,jexp-(1+q)2qβTi,j-A(ρY)i,jexp-(1+q)2qβTi,j,Fi,j=(ρW)i,j(ρUW)i,j-PrWi+1,j-Wi-1,j2Δx+Ui,j-Ui,j-1Δz(ρW2)i,j+1γMa2Pi,j-Pr43Wi,j-Wi,j-1Δz-23Ui+1,j-Ui-1,j2Δx(ρTW)i,j-Ti,j-Ti,j-1Δz(ρYW)i,j-1LeYi,j-Yi,j-1Δz. 106

For a flux vector E, the backward and central differences are applied to the partial derivatives with respect to x and z, respectively. On the other hand, for F, the central and backward differences are applied to the partial derivatives with respect to x and z, respectively.

To avoid the oscillation of numerical solutions, we introduce a total variation diminishing (TVD) numerical scheme in the corrector step4144. We employ the Causon’s model44. Instead of calculating the inner product of flux vectors as done in the original method, we simply extend the TVD scheme for a scalar quantity to the case of vector flux in order to reduce the computation time. By denoting each component of a flux vector U as U=(Uk)(k=1,2,3,4,5), each component of TVD vectors TVDx=(TVDxk) and TVDz=(TVDzk) is expressed by

TVDxk,i,j=G(rxk,i,j+)+G(rxk,i+1,j-)ΔUk,i+12,j-G(rxk,i-1,j+)+G(rxk,i,j-)ΔUk,i-12,j, 107
TVDzk,i,j=G(rzk,i,j+)+G(rzk,i,j+1-)ΔUk,i,j+12-G(rzk,i,j-1+)+G(rzk,i,j-)ΔUk,i,j-12, 108

where

ΔUk,i+12,j=Uk,i+1,j-Uk,i,j,ΔUk,i-12,j=Uk,i,j-Uk,i-1,j,rxk,i,j+=ΔUk,i-12,jΔUk,i+12,j,rxk,i,j-=1rxk,i,j+, 109
ΔUk,i,j+12=Uk,i,j+1-Uk,i,j,ΔUk,i,j-12=Uk,i,j-Uk,i,j-1,rzk,i,j+=ΔUk,i,j-12ΔUk,i,j+12,rzk,i,j-=1rzk,i,j+, 110
Grxk,i,j±=CFLx2(1-CFLx)1-ϕrxk,i,j±,Grzk,i,j±=CFLz2(1-CFLz)1-ϕrzk,i,j±, 111
ϕ(r)=min(2r,1)(r>0)0(r0). 112

The Courant–Friedrichs–Lewy factors CFLx and CFLz are defined by use of sound speed and velocity field as follows49.

CFLx=(c~s+maxx,z|U~|)Δt~Δx~=1Ma+maxx,z|U|ΔtΔx,CFLz=(c~s+maxx,z|W~|)Δt~Δz~=1Ma+maxx,z|W|ΔtΔz, 113

where (6) and (7) are used. Accounting for the above formulae, the finite-difference operators of TVD-MacCormack scheme are given by

Lx(Δt)Un:Ui,jp=Ui,jn-ΔtEi+1,jn-Ei,jnΔx-Si,j2Ui,jc=Ui,jn+Ui,jp2-Δt2Ei,jp-Ei-1,jpΔx-Si,jp2+TVDxi,jn, 114
Lz(Δt)Un:Ui,jp=Ui,jn-ΔtFi,j+1n-Fi,jnΔz-Si,j2Ui,jc=Ui,jn+Ui,jp2-Δt2Fi,jp-Fi,j-1pΔz-Si,jp2+TVDzi,jn. 115

Then, the update of flux vector U is made as follows50.

Ui,jn+1=LzΔt2LxΔtLzΔt2Ui,jn. 116

Settings

The initial conditions are given as follows. The development of perturbed reaction front is examined by initially imposing small perturbations on a planar front. We use the sinusoidal displacement periodic in x-direction, which is given by

d=asin(kx), 117

where a is the initial amplitude of displacement and k=2π/λ is a wavenumber with λ the wavelength of perturbations. The initial solutions are given by those of a steady planar flow in (35). Therefore, the initial values of a flux vector U consists of the following components.

ρ=1/(1+qez-9-d)(z9+d)1/(1+q)(z>9+d),ρU=0,ρW=1, 118
ρT=1+γMa243Pr-1qez-9-d(z9+d)1-γMa2q(z>9+d),ρY=(1-eLe(z-9-d))/(1+qez-9-d)(z9+d)0(z>9+d). 119

The perturbations are imposed on the planar reaction front initially located at z=9.

As for the boundary conditions, the spatially periodic conditions are employed in the x-direction. In the z-direction, the far-field condition (20) is used at the upstream edge and zero-gradient conditions are made at both upstream and downstream edges.

We consider the case of k=0.5, correspondingly λ=4π. The parameters are set as γ=1.4, Ma=0.01, β=10 and a=0.1. The Lewis number Le is calculated by use of (83). The computational domain is arranged for one wavelength of a perturbed front, λ, in x-direction. In the propagation direction, we consider 0z10.

We use the Grid Convergence Index (GCI) method for estimation of discretization error5153. A representative grid size h is given by h=(ΔxΔz)1/2 because we consider the case of uniform grid. We select three different sets of grid so as to h1<h2<h3. The grid refinement factor is defined by the ratio of grid size, which should be greater than 1.3. Let r21=h2/h1 and r32=h3/h2. Then, the apparent order p of the numerical method is calculated by

p=1ln(r21)ln(ε32/ε21)+q(p), 120

where

q(p)=lnr21p-sr32p-s,s=signε32/ε21,ε32=ϕ3-ϕ2,ε21=ϕ2-ϕ1. 121

In this study, the key variable ϕ of simulation is set as ρY, which is useful to identify the location of reaction front because Y=0 on the burned side. The percentage of oscillatory convergence is identified by the number of s<0 among entire lattice points. The approximate relative error ea21 and the grid-fine convergence index GCIfine21 are given by

ea21=ϕ1-ϕ2ϕ1,GCIfine21(p)=1.25ea21r21p-1. 122

The numerical uncertainty at each lattice point is estimated by use of GCIfine21(pave) and indicated by error bars in plotting fine-grid solution ϕ1, where pave is an average value of p and measures the global order of accuracy.

Numerical results

The resolution test is shown in Table 1. For 65×513 and 65×1025 grids, the fine-grid solution ϕ1 is plotted in Figs. 10 and 11 with error bars calculated by GCIfine21(pave). In view of GCI, a 65×1025 grid looks like better than others, though the oscillatory convergence occurs at some points and the average apparent order is not near 2, which is a theoretically desirable order of accuracy in the case of MacCormack scheme. By adopting a 65×1025 grid as the spatial resolution, the increments in x- and z-directions are Δx=4π/640.2 and Δz=10/10240.01. The time increment is set as Δt=3×10-5. The temperature distribution is shown in Figs. 12 and 13 for two values of Lewis number, which are chosen so as to show both unstable and stable cases as guessed from Fig. 1. It is possible to estimate the numerical growth rate by considering the front displacement A(t), which is calculated from the difference of maximum and minimum positions of the front. The front is identified by an isoline ρY=0.1. Figures 14 and 15 show the calculated front displacement in time and the estimated exponential curves, -0.005+0.195exp(-0.035t) and -0.01+0.195exp(0.055t), which mean the numerical growth rates are -0.035 for le=-4 and 0.055 for le=-6. There is a concern on the decrease of A(t) at the early stage for le=-6. This might indicate that other grids, which would be finer in z-direction, are demanded. Besides, the case of small wavenumber has not been examined yet, which should be compared with the analytical result.

Table 1.

The resolution test is examined for several grids Nx×Nz, where Nx and Nz are the number of lattice points in x- and z-directions, respectively. All estimation is made at y=2π and t = 0.075 for every lattice points in z-direction.

Nz 513 1025
Nx (17, 33, 65) (33, 65, 129) (17, 33, 65) (33, 65, 129)
pave 1.10185 0.78772 3.61885 7.684307
maxzGCIfine21(pave)(%) 6.8752 8.7668 0.0304172 0.140160
percentage of oscillatory convergence (%) 100 0 1.17302 2.34604

Figure 10.

Figure 10

Semi-log plot of fine-grid solution ϕ1 with error bars calculated by GCIfine21(pave) for a 65×513 grid at y=2π and t = 0.075.

Figure 11.

Figure 11

The same as Fig. 10 but for a 65×1025 grid.

Figure 12.

Figure 12

Temperature distribution for le=-4 and Pr=0.

Figure 13.

Figure 13

The same as Fig. 12 but for le=-6.

Figure 14.

Figure 14

Front displacement for le=-4 with a 65×1025 grid and the numerical growth rate is estimated by an exponential curve.

Figure 15.

Figure 15

The same as Fig. 14 but for le=-6.

Discussion

We remark that our results do not contain the Darrieus-Landau instability (DLI), or the hydrodynamic instability. This is because we have employed the boundary conditions across the reaction front, which have been derived systematically under the large activation energy asymptotics, and focused on the pure thermal-diffusive instability. In our case, the density, or the normal velocity, is continuous across the reaction front even for a basic flow, whereas their gradients are discontinuous. Then, the xy components of velocity field are also continuous across the front, opposed to their discontinuity in the hydrodynamic model. The discontinuity of density is essential for the existence of DLI. The boundary conditions across a flame front, which reflect the density jump, have been studied in detail based on the multi-scale analysis by introducing the preheat zone inside a flame front2326.

We roughly examine why the DLI drops out in the present study by removing the continuity condition of density, or that of normal velocity, which is expected to lead to the appearance of O(k) term in the dispersion relation corresponding to the DLI. In order to confirm this, we consider the following solutions of a basic flow, instead of (35), in accordance with those of hydrodynamic model27,28.

ρ¯=1(ξ<0)1/(1+q)(ξ>0),W¯=1/ρ¯. 123

In this case, boundary conditions (62) and (63) are changed into

Wξ|-+ρξ|-=11+qWξ|++ρξ|+(1+q), 124
U|+-U|-=-qFx,V|+-V|-=-qFy. 125

We employ solutions (47), (49), (54)–(57) and (59), which remain unchanged in order to couple the thermal-diffusive instability with the hydrodynamic one. Substituting these solutions into boundary conditions (64)–(68), (124) and (125), the dispersion relation is calculated, under the assumption of ωk21 and with O(q2) terms omitted, as

2(1+q)ω2+2β-(1+q)1-2Prk-1-LePr1+LePrkLekω+qk3+2β-1-LeLe-q1Le(1+Le)+LePr21+LePrk4=0. 126

The dispersion relation (126) gives the following growth rate for small wavenumbers,

ω=k2-1β(1+q)+12Le+LePr21+LePrk2, 127

which possesses O(k) term, which reflects the influence of hydrodynamic instability. Although this result has been obtained in very rude manner, the intrinsic importance of density discontinuity for the existence of DLI may become apparent. Finally, we note that the present study is restricted to the case of continuous density across the reaction front. Therefore, the DLI does not appear.

Conclusions

We have investigated how the thermal-diffusive instability is affected by the heat release and viscosity. Instead of introducing the effect of thermal thickness of a flame front, or the preheat zone, we directly dealt with the governing equations in the framework of thermal-diffusive model, or ZFK model. This approach has a merit of understanding the viscous effect on the stability boundary of cellular and oscillatory instabilities in a simple manner: the stability boundary is described in the plane with respect to Lewis number and wavenumber. Our main assumption is the small values of heat release with the continuity condition of density across the reaction front, which is valid in the framework of large activation energy asymptotics. Thanks to this assumption, the coupling of hydrodynamic and thermal-diffusive models becomes possible. As a result, the effect of viscosity is incorporated into the dispersion relation for the pure thermal-diffusive instability, though the DLI does not appear.

Due to the existence of heat release and viscosity, the stability boundary of thermal-diffusive instability is largely changed from that obtained in the previous case of no heat release. The heat release has a stabilizing effect on both of cellular and oscillatory instabilities. In other words, for arbitrary wavenumber, the stable range of Lewis number, where the real part of growth rate of perturbations is negative, is widened as the heat release increases. Such a stabilizing effect is also brought by the rise of Prandtl number, which originates from the Navier-Stokes equation, on the oscillatory instability. However, the effect of viscosity has two opposite aspects on the cellular instability. For some moderate values of wavenumber, the stable range of Lewis number for the cellular instability is extended. On the other hand, for small wavenumbers, the stable range is narrowed by the increase of Prandtl number.

The growth rate of cellular instability has been studied in detail by restricting ourselves to small values of growth rate and wavenumber. In this case, we find that the third order term with respect to wavenumber appears in the reduced dispersion relation due to the existence of heat release. This term brings the new band of Lewis number, where two cut-off wavenumbers exist: in other words, there are two wavenumbers at which the growth rate of perturbations changes its sign. The relation between the Lewis number and the cut-off wavenumber is explicitly derived with the viscous effect included. While the value of smaller cut-off wavenumber is reduced as the Prandtl number increases, the behavior of larger one shows both of the increase and decrease depending on values of Lewis number.

The numerical calculation is also made by use of TVD-MacCormack scheme. The numerical growth rate is estimated for different values of Lewis number with small heat release. For moderate value of wavenumber, the numerical result seems to be consistent with the stability boundary obtained in the asymptotic analysis. However, the case of small wavenumbers has not been checked yet. We need to compare this case with the analytical result.

Author contributions

K.W. performed the works for the study.

Data availibility

No datasets were generated or analysed during the current study.

Competing interests

The author declares no competing interests.

Footnotes

The original online version of this Article was revised: The original version of this Article contained errors in Equations 24-26. Full information regarding the corrections made can be found in the correction for this Article.

Publisher's note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Change history

10/21/2021

A Correction to this paper has been published: 10.1038/s41598-021-00632-9

References

  • 1.Zel’dovich YB, Frank-Kamenetskii DA. Theory of thermal flame propagation. Zhu. Fiz. Khim. 1938;12:100–105. [Google Scholar]
  • 2.Zel’dovich YB, Frank-Kamenetskii DA. A theory of thermal propagation of flame. Acta Physicochimica URRS. 1938;9:341–350. [Google Scholar]
  • 3.Joulin G, Vidal P. Hydrodynamics and Nonlinear Instabilities. Cambridge University Press; 1998. [Google Scholar]
  • 4.Spalding DBI. Predicting the laminar flame speed in gases with temperature-explicit reaction rates. Combust. Flame. 1957;1:287–295. doi: 10.1016/0010-2180(57)90015-9. [DOI] [Google Scholar]
  • 5.Spalding DB., II One-dimensional laminar flame theory for temperature-explicit reaction rates. Combust. Flame. 1957;1:296–307. doi: 10.1016/0010-2180(57)90016-0. [DOI] [Google Scholar]
  • 6.Lewis B, von Elbe G. Combustion Flames and Explosions of Gases. 3. Academic Press; 1987. [Google Scholar]
  • 7.Glassman I, Yetter RA. Combustion. 4. Academic Press; 2008. [Google Scholar]
  • 8.Sivashinsky GI. Structure of bunsen flames. J. Chem. Phys. 1975;62:638–643. doi: 10.1063/1.430465. [DOI] [Google Scholar]
  • 9.Sivashinsky GI. On a distorted flame front as a hydrodynamic discontinuity. Acta Astronautica. 1976;3:889–918. doi: 10.1016/0094-5765(76)90001-1. [DOI] [Google Scholar]
  • 10.Friedman R, Burke E. A theoretical model of a gaseous combustion wave governed by a first-order reaction. J. Chem. Phys. 1953;21:710–714. doi: 10.1063/1.1698995. [DOI] [Google Scholar]
  • 11.von Kármán, T. The present status of the theory of laminar flame propagation. In Sixth Symposium on Combustion, Reinhold, New York 1–11 (1957).
  • 12.Bush WB, Fendell FE. Asymptotic analysis of laminar flame propagation for general Lewis numbers. Combust. Sci. Technol. 1970;1:421–428. doi: 10.1080/00102206908952222. [DOI] [Google Scholar]
  • 13.Williams FA. Combustion Theory: The Fundamental Theory of Chemically Reacting Flow Systems. 2. Addison-Wesley; 1985. [Google Scholar]
  • 14.Fife PC. Dynamics of Internal Layers and Diffusive Interfaces. SIAM; 1988. [Google Scholar]
  • 15.Matkowsky BJ, Sivashinsky GI. An asymptotic derivation of two models in flame theory associated with the constant density approximation. SIAM J. Appl. Math. 1979;37:686–699. doi: 10.1137/0137051. [DOI] [Google Scholar]
  • 16.Matalon M, Cui C, Bechtold JK. Hydrodynamic theory of premixed flames: effects of stoichiometry, variable transport coefficients and arbitrary reaction orders. J. Fluid Mech. 2003;487:179–210. doi: 10.1017/S0022112003004683. [DOI] [Google Scholar]
  • 17.Law CK. Combustion Physics. Cambridge University Press; 2006. [Google Scholar]
  • 18.Joulin G, Mitani T. Linear stability analysis of two-reactants flames. Combust. Flame. 1981;40:235–246. doi: 10.1016/0010-2180(81)90127-9. [DOI] [Google Scholar]
  • 19.Joulin G, Clavin P. Linear stability analysis of nonadiabatic flames: Diffusional-thermal model. Combust. Flame. 1979;35:139–153. doi: 10.1016/0010-2180(79)90018-X. [DOI] [Google Scholar]
  • 20.Brailovsky I, Gordon PV, Kagan L, Sivashinsky G. Diffusive-thermal instabilities in premixed flames: Stepwise ignition-temperature kinetics. Combust. Flame. 2015;162:2077–2086. doi: 10.1016/j.combustflame.2015.01.006. [DOI] [Google Scholar]
  • 21.Matkowsky BJ, Olagunju DO. Spinning waves in gaseous combustion. SIAM J. Appl. Math. 1982;42:1138–1156. doi: 10.1137/0142078. [DOI] [Google Scholar]
  • 22.Buckmaster JD. The Mathematics of Combustion. SIAM; 1985. [Google Scholar]
  • 23.Pelce P, Clavin P. Influence of hydrodynamics and diffusion upon the stability limits of laminar premixed flames. J. Fluid Mech. 1982;124:219–237. doi: 10.1017/S002211208200247X. [DOI] [Google Scholar]
  • 24.Clavin P, Williams FA. Effects of molecular diffusion and of thermal expansion on the structure and dynamics of premixed flames in turbulent flows of large scale and low intensity. J. Fluid Mech. 1982;116:251–282. doi: 10.1017/S0022112082000457. [DOI] [Google Scholar]
  • 25.Matalon M, Matkowsky BJ. Flames as gasdynamic discontinuities. J. Fluid Mech. 1982;124:239–259. doi: 10.1017/S0022112082002481. [DOI] [Google Scholar]
  • 26.Clavin P, Garcia P. The influence of the temperature dependence of diffusivities on the dynamics of flame fronts. J. Méc. Théor. Appl. 1983;2:245–263. [Google Scholar]
  • 27.Darrieus, G. Unpublished works presented at La Technique Moderne (1938).
  • 28.Landau L. On the theory of slow combustion. Acta Phys. (USSR) 1944;19:77–85. [Google Scholar]
  • 29.Matalon M. The Darrieus–Landau instability of premixed flames. Fluid Dyn. Res. 2018;50:051412. doi: 10.1088/1873-7005/aab510. [DOI] [Google Scholar]
  • 30.Markstein GH. Experimental and theoretical studies of flame-front stability. J. Aero. Sci. 1951;18:199–209. doi: 10.2514/8.1900. [DOI] [Google Scholar]
  • 31.Eckhaus W. Theory of flame-front stability. J. Fluid Mech. 1961;10:80–100. doi: 10.1017/S0022112061000081. [DOI] [Google Scholar]
  • 32.Clavin P, Searby G. Combustion Waves and Fronts in Flows. Cambridge University Press; 2016. [Google Scholar]
  • 33.Michelson DM, Sivashinsky GI. Nonlinear analysis of hydrodynamic instability in laminar flames—II. Numerical experiments. Acta Astronautica. 1977;4:1207–1221. doi: 10.1016/0094-5765(77)90097-2. [DOI] [Google Scholar]
  • 34.Denet B, Haldenwang P. Numerical study of thermal-diffusive instability of premixed flames. Combust. Sci. Technol. 1992;86:199–221. doi: 10.1080/00102209208947195. [DOI] [Google Scholar]
  • 35.Kadowaki S. The effects of heat loss on the burning velocity of cellular premixed flames generated by hydrodynamic and diffusive-thermal instabilities. Combust. Flame. 2005;143:174–182. doi: 10.1016/j.combustflame.2005.05.012. [DOI] [Google Scholar]
  • 36.Kadowaki S, Takahashi H, Kobayashi H. The effects of radiation on the dynamic behavior of cellular premixed flames generated by intrinsic instability. Proc. Combust. Inst. 2011;33:1153–1162. doi: 10.1016/j.proci.2010.06.069. [DOI] [Google Scholar]
  • 37.Aung TT, Kadowaki S. The diffusive-thermal instability of premixed flames with low unburned-gas temperature under the adiabatic and non-adiabatic conditions. Mech. Eng. J. 2015;2:14–00460. doi: 10.1299/mej.14-00460. [DOI] [Google Scholar]
  • 38.MacCormack, R. W. & Paullay, A. J. Computational efficiency achieved by time splitting of finite difference operators. In 10th Aerospace Sciences Meeting, 72–154 (San Diego, California, 1972).
  • 39.Maccormack, R. W. An efficient numerical method for solving the time-dependent compressible Navier–Stokes equations at high Reynolds number. NASA TM X-73-129, 1–16 (1976).
  • 40.MacCormack RW. A numerical method for solving the equations of compressible viscous flow. AIAA J. 1982;20:1275–1281. doi: 10.2514/3.51188. [DOI] [Google Scholar]
  • 41.Harten A. High resolution schemes for hyperbolic conservation laws. J. Comput. Phys. 1983;49:357–393. doi: 10.1016/0021-9991(83)90136-5. [DOI] [Google Scholar]
  • 42.Davis, S. F. TVD finite difference schemes and artificial viscosity. ICASE Report No. 84-20 (1984).
  • 43.Davis SF. A simplified TVD finite difference scheme via artificial viscosity. SIAM J. Sci. Statist. Comput. 1987;8:1–18. doi: 10.1137/0908002. [DOI] [Google Scholar]
  • 44.Causon, D. M. High resolution finite volume schemes and computational aerodynamics. In Nonlinear Hyperbolic Equations-Theory, Computation Methods, and Applications. Notes on Numerical Fluid Mechanics Vol. 24 (eds Ballmann, J. Jeltsch, R.), 63–74 (Springer, 1989).
  • 45.Sivashinsky GI. Nonlinear analysis of hydrodynamic instability in laminar flames–I. Derivation of basic equations. Acta Astronautica. 1977;4:1177–1206. doi: 10.1016/0094-5765(77)90096-0. [DOI] [Google Scholar]
  • 46.Lax PD, Wendroff B. Difference schemes for hyperbolic equations with high order of accuracy. Comm. Pure Appl. Math. 1964;17:381–398. doi: 10.1002/cpa.3160170311. [DOI] [Google Scholar]
  • 47.Richtmyer RD, Morton KW. Difference Methods for Initial-Value Problems. Interscience; 1967. [Google Scholar]
  • 48.Strang G. On the construction and comparison of difference schemes. SIAM J. Numer. Anal. 1968;5:506–517. doi: 10.1137/0705041. [DOI] [Google Scholar]
  • 49.de Moura CA, Kubrusly CS. The Courant–Friedrichs–Lewy (CFL) Condition. Springer; 2013. [Google Scholar]
  • 50.Ngondiep E. A novel three-level time-split Maccormack scheme for two-dimensional evolutionary linear convection-diffusion-reaction equation with source term. Int. J. Comput. Math. 2021;98:47–74. doi: 10.1080/00207160.2020.1726896. [DOI] [Google Scholar]
  • 51.Celik I, Karatekin O. Numerical experiments on application of Richardson extrapolation with nonuniform grids. J. Fluids Eng. 1997;119:584–590. doi: 10.1115/1.2819284. [DOI] [Google Scholar]
  • 52.Celik I, Ghia U, Roache P, Freitas C. Procedure for estimation and reporting of uncertainty due to discretization in CFD applications. J. Fluids Eng. 2008;130:078001. doi: 10.1115/1.2960953. [DOI] [Google Scholar]
  • 53.Lutsenko NA, Fetsov SS. Numerical model of time-dependent gas flows through bed of granular phase change material. Int. J. Computat. Methods. 2020;17:1950010. doi: 10.1142/S0219876219500105. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

No datasets were generated or analysed during the current study.


Articles from Scientific Reports are provided here courtesy of Nature Publishing Group

RESOURCES