Skip to main content
Proceedings. Mathematical, Physical, and Engineering Sciences logoLink to Proceedings. Mathematical, Physical, and Engineering Sciences
. 2015 Mar 8;471(2175):20140845. doi: 10.1098/rspa.2014.0845

On oscillatory convection with the Cattaneo–Christov hyperbolic heat-flow model

J J Bissell 1,
PMCID: PMC4353040  PMID: 25792960

Abstract

Adoption of the hyperbolic Cattaneo–Christov heat-flow model in place of the more usual parabolic Fourier law is shown to raise the possibility of oscillatory convection in the classic Bénard problem of a Boussinesq fluid heated from below. By comparing the critical Rayleigh numbers for stationary and oscillatory convection, Rc and RS respectively, oscillatory convection is found to represent the preferred form of instability whenever the Cattaneo number C exceeds a threshold value CT≥8/27π2≈0.03. In the case of free boundaries, analytical approaches permit direct treatment of the role played by the Prandtl number P1, which—in contrast to the classical stationary scenario—can impact on oscillatory modes significantly owing to the non-zero frequency of convection. Numerical investigation indicates that the behaviour found analytically for free boundaries applies in a qualitatively similar fashion for fixed boundaries, while the threshold Cattaneo number CT is computed as a function of P1[102,10+2] for both boundary regimes.

Keywords: buoyancy-driven instabilities, thermal convection, hyperbolic heat-flow, oscillatory convection, Rayleigh–Bénard convection

1. Introduction

The notion of thermal transport obeying a hyperbolic rather than diffusive (Fourier-type) law of conduction goes back to Maxwell's early work on kinetic theory [1], and is a topic of considerable ongoing concern. In part, such interest derives from foundational problems in the Fourier theory of heat where—owing to its parabolic nature—thermal disturbances propagate with infinite speed [2,3]; hyperbolic modifications of Fourier's law overcome this problem because thermal signals may then travel as waves (heat waves) [4,5]. However, the significance of thermal waves is deep, and reaches beyond their theoretical implications: for instance, such waves have important applications in the study of energy transport [6] and thermal shocks in solids [7]; heat transport in nanomaterials and nanofluids [811]; biological tissues and surgical procedures [12], including skin burns [13] and radiofrequency heating [14,15]; convection in fluids and porous media [1618]; alongside astrophysical contexts, particularly thermohaline convection [19,20]. In addition, work on hyperbolic theories of thermal transport has inspired related studies in other flux-based problems, for example, those involving advection–diffusion equations [21,22], and discontinuity waves [23,24].

Given such varied applications, and the clear need to improve theoretical understanding of thermal waves, it is the concern of this article to study the effects of hyperbolic heat-flow within a well-established context, namely the classic Rayleigh–Bénard convection problem of a Boussinesq fluid heated from below [25,26]. Indeed, by investigating hyperbolic heat-flow in such a canonical problem, we seek not only to further understanding of thermal waves per se, but also contribute to ongoing research into novel aspects of thermal convection and non-stationary heat conduction more generally [331]. To this end, we focus on the commonly encountered Maxwell–Cattaneo reformulation of Fourier's law in which the heat-flow vector Qi is expressed in terms of both gradients in the local temperature T, and inertial effects owing to the relaxation time τ taken for such gradients to become established [2,32], that is,

τQit+Qi=κTxi, 1.1

where κ is the thermal conductivity, and standard indicial notion applies throughout. The thermal relaxation time τ is typically rather small [3,4], but can take relatively large values (approx. 100 s) in some contexts (e.g. biological tissues [13]). Curiously, while equation (1.1) is perhaps the most frequently encountered model for hyperbolic heat-flow (see, for example, the articles cited above), it is known to yield a paradoxical result for moving bodies whereby the process of heat conduction becomes dependent on the frame of motion [32]. To circumvent such difficulties, Christov and Jordan [32,33] proposed a modification to equation (1.1) by which Galilean invariance is restored after replacing the partial time derivative with a material derivative, i.e.

τ(Qit+vjQixj)+Qi=κTxi. 1.2

As we shall be working with a system in convective motion, we adopt this Cattaneo–Christov formulation throughout. Note that for problems such as ours, inertial effects introduced by the relaxation time are best discussed in terms of the dimensionless Cattaneo number C(τ), which accounts for combined effects, including system length scales (see equation (2.11)).

The problem of thermal convection with a hyperbolic heat-flow model seems to have been first investigated by Straughan & Franchi [16], who examined the instability of a system heated from above; however, it is Straughan's more recent work, in which he considers oscillatory convection with a Cattaneo–Christov model in a fluid layer heated from below [17], that is most relevant to our present discussion. In particular, if we denote the critical Rayleigh numbers for stationary and oscillatory convection by Rc and RS, respectively, then for fluids confined by free boundaries Straughan derives an expression for RS(C) which decreases with Cattaneo number, suggesting a transition from stationary to oscillatory convection as C is increased beyond a threshold value CT for which RS(CT)=Rc; in the case of fixed boundaries, and for Prandtl number P1=6, Straughan then uses numerical methods to bound this value to the interval CT∈(2.2×10−2,2.3×10−2) [17].

Straughan's preliminary investigation yields results which are valuable and compelling; however, Straughan's work raises a number of fundamental problems in need of serious consideration: for example, in the free boundary case it is of pressing importance to determine whether threshold behaviour is actually physical. In fact, by examining the frequency of oscillation γ one may show that oscillatory solutions are forbidden unless the Cattaneo number exceeds some bounding value CB (see §4); for threshold behaviour to obtain generally, therefore, one must confirm the condition CT(P1)>CB(P1) for arbitrary Prandtl number. Indeed, the importance of the Prandtl number itself is something Straughan does not address, possibly because P1 does not enter into calculations of the critical Rayleigh number Rc for stationary convection (when γ=0). However, for oscillatory convection (γ≠0), the role played by the Prandtl number becomes fundamental; in particular, the set of crucial parameters—the threshold Cattaneo number CT, critical wavenumber aS and oscillation frequency γ—all exhibit strong P1 dependence (see §§47).

In this article, therefore, we substantially develop Straughan's ideas [17] to form a comprehensive theory of thermal convection with the Cattaneo–Christov heat-flow model, paying particular attention to the foundational issues described above. Beginning with a short review of the classical theory [26] in §§2 and 3, we proceed to the theory of oscillatory convection in §4, where we generalize and simplify some of Straughan's initial work on the critical Rayleigh number RS, deriving a number of new analytical results and asymptotic expressions in the process. In particular, we present calculations of both the bounding CB and threshold Cattaneo number CT(P1)8/27π20.03 for the first time, and thereby analyse P1-C parameter space generally into regions where either stationary or oscillatory convection is preferred (§5). Such an analysis allows us to show rigorously that the condition CT(P1)>CB(P1) holds for arbitrary P1, and thence—in the case of free boundary conditions—that discontinuous transitions from stationary to oscillatory convection are permitted physically. For fixed boundary conditions, we analyse the system numerically (§6) by calculating the threshold parameters for Prandtl number in the range P1[102,10+2], and—in a qualitative fashion at least—recover much of the behaviour observed analytically for free boundaries. These results, alongside some of their potential consequences for thermal convection more generally, are discussed further in §7.

2. Convection model

The basic fluid model adopted here comprises equations governing the conservation of mass and energy, alongside Christov's Galilean invariant formulation of the Cattaneo heat-flow law [2,26,32]. Thus, denoting the velocity, pressure, temperature and heat-flow as vi, P, T and Qi, respectively, the momentum, energy and heat-flow equations are (cf. Chandrasekar [26] and Straughan [17])

(vit+vjvixj)=1ρ0Pxi+ν2viρρ0gλi, 2.1a
ρ0cV(Tt+vjTxj)=Qjxj 2.1b
andτ(Qit+vjQixj)+Qi=κTxi. 2.1c

Equations (2.1ac) include constant coefficients for the fluid viscosity ν, gravitational acceleration g=|g|, specific heat at constant volume cV, thermal relaxation time τ and thermal conductivity κ. Here, we assume a Cartesian (x,y,z) geometry in which gravity acts in the negative z-direction, i.e. λi is the unit vector λi=(0,0,1), while the Laplacian operator is ∇2≡∂2/∂x2+∂2/∂y2+∂2/∂z2. We also employ the Boussinesq approximation in the buoyancy term, viz

ρ(T)=ρ0[1+α(TαT)], 2.2

where ρ0 is the fluid density when it is at temperature T=Tα, and α is a thermal expansion coefficient. Note that by asserting a material derivative in the Cattaneo heat-flow law, model (2.1) differs from that employed by Straughan, who used an upper convected Oldroyd time derivative [17,32,33]. However, because both models assume an incompressible equation of state, i.e.

vixi=0, 2.3

our focus on a linearized system means that this difference is not important, and the more compact and intuitive form given by equation (2.1c) is preferred.

(a). Steady-state and linearized system

We now suppose that the fluid occupies a semi-infinite region (x,y)R2 confined between two parallel planes, one at z=0, and the other at z=d, where the upper and lower planes are held at constant temperatures Tu and Tl respectively, i.e.

T(0)=TlandT(d)=Tu,withTl>Tu. 2.4

In this way our system corresponds to the classic Bénard problem of a fluid heated from below, with the z-component of the velocity w vanishing on the boundaries, i.e. (cf. Chandrasekhar [26])

vz=w=0atz=0,d. 2.5

In steady-state, solutions to system (2.1) are given by

vi=vi0=0,T=T0=Tlβλjxj,Qi=Qi0=βκλiandP=P0(z), 2.6

where β denotes the temperature gradient, and P0 has a profile such that the buoyancy force is balanced by pressure gradients, that is,

β=|Txi|=TlTudandP0xi=gρ(T0)λi. 2.7

Our main purpose here is to consider the stability of these steady-state solutions once perturbed. However, before proceeding it is expedient to recast our system in a dimensionless form by defining the normalized quantities

x~i=xid,t~=νd2t,v~i=viν/d,P~=Pν2ρ0/d2,T~=αgκd2βν3ρ0cVT,Q~i=dκ(T~T)Qi. 2.8

Consequently, after adding small perturbations {u~i,θ~,q~i,p~} to the steady-state solutions, that is,

v~i=v~i0+u~i=u~i,T~=T~0+θ~,Q~i=Q~i0+q~iandP~=P~0+p~, 2.9

system (2.1) may be linearized to give a set of normalized equations, viz

u~it~=p~x~i+2~u~i+Rθ~λi, 2.10a
P1θ~t~=Rw~q~jx~j 2.10b
and2P1Cq~it~=q~iθ~x~i, 2.10c

where the dimensionless Prandtl number P1, Cattaneo number C and Rayleigh number Ra=R2 are defined

P1=νρ0cVκ,C=τκ2ρ0d2cVandR=Ra=αgd4βρ0cVνκ. 2.11

Hence, denoting the divergence of the heat-flow

Qq=q~ix~i, 2.12

we take the curl of equation (2.10a) twice, the divergence of equation (2.10c) once, and the inner product any vector equations with λi, and thereby eliminate p~ from the linearized system (2.10) to obtain (cf. Chandrasehkar [26])

t2w=R(2θx2+2θy2)+4w, 2.13a
P1θt=RwQq 2.13b
and2CP1Qqt=Qq2θ, 2.13c

where w=uz is the z-component of the velocity field, and the tilde notation has been dropped for brevity. These equations describe the evolution of perturbations to the conductive steady state (2.6) in a form conveniently reduced to three variables; we are thus well placed to investigate the stability of the system by further decomposing the problem into normal modes.

(b). Analysis into normal modes

Let us now write the perturbations in separable form assuming an exponential time dependence such that

w=W(z)f(x,y)eσt,θ=Θ(z)f(x,y)eσtandQq=Q(z)f(x,y)eσt, 2.14

where σ is a constant frequency, W, Θ and Q are some eigenfunctions to be found, and f(x,y) is a plane tiling function satisfying

2f(x,y)=a2f(x,y), 2.15

with a as a characteristic wavenumber or inverse length scale. Then system (2.13) becomes (cf. Chandrasehkar [26] and Straughan [17])

σ(D2a2)W=(D2a2)2Wa2RΘ, 2.16a
σP1Θ=RWQ 2.16b
and2σP1CQ=Q(D2a2)Θ, 2.16c

where with Φ∈{W,Q,Θ} the operators D and D2 are

DΦ=dΦdzandD2Φ=d2Φdz2. 2.17

Equations (2.16) represent the starting point for both the analytical and numerical work on oscillatory convection forming the main basis of this article. However, because we shall be comparing results for oscillatory convection with those of stationary convection, it is appropriate at this stage to briefly review some classical results.

3. Stationary convection

For stationary convection, instability sets in through the marginal state characterized by σ=0, and we can eliminate Θ and Q from equations (2.16) to give [26]

(D2a2)3W=a2R2W=a2RaW. 3.1

For free boundaries, we have that

W=0,D2W=0,Θ=0,atz=0,1, 3.2

so by system (2.16) and equation (3.1), one deduces that all even derivatives of W vanish at the boundaries, and W(z) may be decomposed into odd Fourier modes, thus

W(z)=n=1Wn,withWn=Ansin(nπz) 3.3

as the nth mode weighted by the constant coefficient An. Hence, following substitution of this expansion into equation (3.1), the orthogonality of the Fourier series gives a Rayleigh number Ra for the nth mode [26]

Ra(n)=Λn3a2,whereΛnn2π2+a2. 3.4

Because we seek a minimum value for the onset of convection, differentiating Ra(n) with respect to a thus yields a critical Rayleigh number Rc(n)=Ra(ac) for the nth mode corresponding to critical wavenumber ac(n), where (∂Ra/∂a)ac=0, such that

Rc(n)=27n4π44andac(n)=nπ2. 3.5

Naturally, because the sequence {Ra(n)} increases monotonically with n, the absolute critical Rayleigh number occurs when n=1, i.e. (cf. Chandrasekhar [26]),

Rc=27π44657.511andac=π22.221, 3.6

where we adopt the convention that unless the mode number n is explicitly stated in a parameter's argument, expressions shall be quoted assuming n=1 throughout.

In the case of fixed boundary conditions, that is,

W=0,DW=0,Θ=0,atz=0,1, 3.7

system (2.16) may be solved numerically to obtain (cf. section 6 and Chandrasehkar [26])

Rc1707.762andac3.116. 3.8

Note that because σ=0, in the case of stationary convection we have by system (2.16) that neither the Prandtl number P1 nor Cattaneo number C enter into calculations of critical values, so it is not surprising that we recover the classical results associated with the more usual Fourier law.

4. Oscillatory convection

Proceeding to the focus of the present work, we now consider the onset of instability via some kind of oscillatory mode. Some of the initial results in this section are equivalent to those first obtained by Straughan [17]; however, the theory described here substantially develops Straughan's ideas, and departs from his analysis in key areas. Crucially, for example, we establish rigorously that oscillatory modes are in fact physically permitted solutions, paying particular attention to the role played by Prandtl number P1 (which Straughan did not consider), and thereby derive a number of important analytical results, limiting behaviours, and threshold values (cf. §1).

For oscillatory solutions, we assume that σ may be complex and non-zero, and define γ such that [26]

σ=iγ, 4.1

then eliminating Θ and Q from system (2.16), we have

(D2a2)2[(D2a2)σ]W+σP1(2CP1σ+1)(D2a2)W=(2CP1σ+1)a2RaW. 4.2

Hence, for the free boundary conditions given in equation (3.2), one finds (as we had for stationary convection) that all even derivatives of the eigenfunction W(z) vanish at the boundaries, and our Fourier decomposition (3.3) gives

(Λn32P12CΛn2γ2ΛnP1γ2a2Ra)=(2CP1a2RaΛn2(P1+1)+2CP12Λnγ2)σ. 4.3

From equation (4.3), we see that for an arbitrary γ the Rayleigh number will in general be complex. As observed by Chandrasekhar [26], the physical requirement that Ra be real thus places constraints on the relationship between the real and imaginary components of σ=. We therefore study the onset of convection via a purely oscillatory mode by asserting γ to be real in which case comparison of real and imaginary parts in equation (4.3) yields (cf. Chandrasehkar [26] and Straughan [17])

Λn32P12CΛn2γ2ΛnP1γ2=a2Ra 4.4a

and

Λn2(P1+1)2CP12Λnγ2=2CP1a2Ra. 4.4b

Straughan uses this system to eliminate γ2 and obtain an expression for Ra [17]; however, as we shall discuss further in later sections, examination of the oscillation frequency is integral to establishing the physicality of solutions, and for this reason we instead begin by eliminating Ra to determine γ, viz

γ2(n)=14C2P13[P1(2CΛn1)1]. 4.5

Indeed, because we need γ2>0, for given n, a and P1 this expression places a restriction on the Cattaneo number for physical solutions to obtain; in particular, we require

C>(P1+1)2(n2π2+a2)P1. 4.6

Assuming condition (4.6) to hold, we substitute for γ2 in system (4.4) to give

Ra(n)=14C2P12a2[(P1+1)Λn+2CP12Λn2], 4.7

and then—as we did for stationary convection—solve (∂Ra/∂a)aS=0 to determine the critical Rayleigh number RS(n)=Ra(aS) for the nth mode (cf. Straughan [17])

RS(n)=n2π22C(1+aS2(n)n2π2)2, 4.8

where aS(n) is the corresponding critical wavenumber

aS2(n)=n2π2[(P1+1)2n2π2CP12+1]1/2. 4.9

Here that we have adopted an ‘S’ subscript notation for these critical values, both to emphasize that they represent Straughan's results generalized to the nth mode [17], and to distinguish them from the values for stationary convection outlined in §3. At this stage in the analysis it is appropriate to quote the generalized results, because physical solutions obtain more readily when n is large. Indeed, for the critical modes, that is, a=aS, inequality (4.6) indicates the requirement

C>CB(n)(P1+1)2n2π2(2P1+1),i.e.limnCB(n)=0, 4.10

where CB∈(1/4n2π2,1/2n2π2) is defined here as the ‘bounding’ Cattaneo number.

Note again that the sequence of critical Rayleigh numbers {RS(n)} increases monotonically with n, so that—seeking a minimum value—by equations (4.8) and (4.9) we have an absolute critical Rayleigh number when n=1, i.e. (figure 1)

RS=π22C[1+((P1+1)2π2CP12+1)1/2]2, 4.11

with corresponding wavenumber aS and frequency γ determined by equations (4.9) and (4.5) after taking n=1.

Figure 1.

Figure 1.

Critical Rayleigh numbers Ra (a), wavenumbers a (b), and asymptotic limits (dashed-dotted curves) plotted as a function of C for P1=6 in the case of free boundaries. Note that beyond the threshold value CT≈0.0342 (circles) we have that RS<Rc, in which case oscillatory convection (solid curves) is preferred over stationary convection (dashed curves), and the wavenumber shifts discontinuously (circles). Here, the oscillatory convection curves represent physical solutions only when inequality (4.14) is satisfied with P1=6, i.e. C>CB(6)≈0.0273 (cf. figure 3).

The critical Rayleigh number given by equation (4.11) is equivalent to that derived by Straughan, though here we have expressed RS in a far more compact form that makes its functional dependence on the Cattaneo number more apparent. Indeed, we see immediately that RS is a strictly decreasing function of C, while in the asymptotic limits we have

RS(C){(P1+1)4C2P12,asC0,2π2C0,asC, 4.12a

with

aS2(C){π(P1+1)2CP12,asC0,π2,asC. 4.12b

Similar limits with P1 may also be obtained for given Cattaneo number C. Thus, at fixed P1, we find that for increasing C:0, then the Rayleigh number decreases like RS(C):0.

Naively, therefore, one supposes there to exist a threshold Cattaneo number CT defined such that RS(CT)=Rc, beyond which (C>CT) the preferential form of instability switches from stationary to oscillatory convection (RS<Rc). Furthermore, given our critical wavenumbers

aS2=π2[(P1+1)2π2CP12+1]1/2>ac2=π22, 4.13

one expects this change to be characterized by a discontinuous transition in the value of a (figure 1). Nevertheless, for the n=1 mode to yield physical solutions, inequality (4.14) demands that

C>CB(P1)(P1+1)2π2(2P1+1), 4.14

so for a given Prandtl number P1, these kinds of transitions between forms of convection will only occur in the fashion described provided CT>CB. It is to this particular issue—the problem of determining whether the threshold behaviour is physical—that we now turn.

5. Threshold Cattaneo number

Because the threshold Cattaneo number CT is defined such that RS(CT)=Rc, for a given Prandtl number P1 we have

RS(CT)=π22CT[1+((P1+1)2π2CTP12+1)1/2]2=Rc,i.e.CT3/2[CT1/2(2π2Rc)]=(P1+1)4RcP12, 5.1

an expression which may be solved numerically (figure 2). By equation (5.1), we see that CT is a monotonically decreasing function of P1, obeying the asymptotic behaviour

CT(P1){13π2P13,asP10,2π2Rc=827π2,asP1. 5.2

Assuming all values of C>CT to represent physical solutions, therefore, the curve CT(P1) may be used to divide C-P1 parameter space into regions where either stationary convection (Rc<RS) or oscillatory convection (Rc>RS) with mode n=1 is the preferred mechanism for instability (figure 2). We now show that this assumption is correct.

Figure 2.

Figure 2.

The threshold Cattaneo number CT (solid curve), bounding Cattaneo number CB (dashed curve), and asymptotic limits (dashed-dotted curves) plotted as a function of P1. For C=CT we have RS=Rc, so the curve CT(P1) divides P1-C parameter space into regions where either oscillatory convection (C>CT) or stationary convection (C<CT) by the critical mode n=1 is preferred. Similarly, P1-C space may be divided by the curve CB(P1) into regions where oscillatory convection by the critical n=1 mode is either physically permitted (C>CB), or forbidden (C<CB). Oscillatory convection can occur in regions below the curve CB, but will do so only for mode numbers n>1 or wavenumbers a>aS (see inequality (4.6)). Note that for given C, one may also view the threshold curve in terms of the threshold Prandtl number PT(C) as defined in equation (5.8).

Recall that for the lowest mode (n=1), physical solutions require C>CB, where CB is a lower bound on the Cattaneo number defined in equation (4.14). Then, since RS is a strictly decreasing function of C, for a given Prandtl number P1 we may define

RBRS(CB)=π4(2P1+1)3P12(P1+1) 5.3

as an upper bound on the critical Rayleigh number for oscillatory convection by the n=1 mode, and with limits

RB(P1){π4P12,asP10,8π4=32π44,asP1. 5.4

Furthermore, differentiation of RB with respect to P1 shows that it is a strictly decreasing function of Prandtl number, viz

dRBdP1=π4(P1+2)(2P1+1)2P13(P1+1)2<0,P1R+. 5.5

Hence, by our asymptotic limits (5.4), we have with RS(CB)=RB and RS(CT)=Rc that

RS(CB)>32π44>27π44=RS(CT),P1R+, 5.6

and, because RS(C) is a strictly decreasing function of C, that

CT(P1)>CB(P1),P1R+. 5.7

Thus, for any given value of P1, inequality (5.7) tells us that C>CT also guarantees C>CB, and the values RS(C)<Rc will correspond to physical solutions. Hence, for free boundaries at least, when discussing the transition from stationary to oscillatory convection, it is sufficient—as we have done so far—to restrict our attention to critical values associated with the lowest mode.

Of course, one may express these thresholds equivalently in terms of the Prandtl number P1; indeed, because RS/P1<0, then for fixed Cattaneo number C we can define a threshold Prandtl number PT such that RS(P1)<Rc for all P1>PT (figure 2). By equation (5.1), this value may be written in the closed form

PT=1+1+16RcC3/2(C2π2/Rc)8RcC3/2(C2π2/Rc), 5.8

with asymptotic behaviour corresponding to limits (5.2)

PT(C){,asC2π2Rc,13π2C3,asC, 5.9

where C2π2/Rc is taken from above. As we found in our discussion of CT(P1), therefore, PT(C) is thus a strictly decreasing function of C defined on C(2π2/Rc,).

An important consequence of inequality (5.7) is that the transition from stationary to oscillatory convection occurs with finite frequency. Indeed, for a given Prandtl number P1, one may define a threshold wavenumber aT=aS(CT) and frequency γT=γ(CT) for the critical modes RS(CT)=Rc, where γT>0 and aT>ac. These threshold values, which have asymptotic limits

aT(P1){π(332P1)1/4,asP10,π,asP1 5.10a

and

γT2(P1){9π42P13/2(34)1/4,asP10,135π4256P120,asP1, 5.10b

are discussed further in §7. Because aS decreases with C, and because oscillatory solutions have C>CT, the threshold wavenumber aT represents an upper limit on values taken by aS.

6. Numerical analysis

In the case of fixed boundary conditions, i.e. those given by equation (3.7), system (2.16) may be solved numerically using the Chebyshev tau-QZ method described by Dongerra et al. [34] and outlined in the context of thermal convection by Straughan [17]. Under this approach, one begins by transforming the z-coordinate such that our problem is defined on z∈(−1,1), and then proceeds to eliminate terms with derivatives higher than D2 by introducing an auxiliary variable χ(z) such that χ(z)f(x,y) eσt=∇2w. In this way, system (2.16) may be written in the form

(4D2a2)Wχ=0, 6.1a
(4D2a2)χa2RΘ=σχ, 6.1b
(4D2a2)Θ+Q=2σCP1Q 6.1c
andRW+Q=σP1Θ. 6.1d

Thence, after expanding the quantities Φ∈{W,χ,Θ,Q} as Chebyshev polynomials Tn(z) weighted by constant coefficients ϕn, viz

Φ(z)=n=0N1ϕnTn(z),with N even, 6.2

one may approximate the system to (in principle) arbitrary precision in terms of the eigenvalue problem

Aijxj=σBijxj, 6.3

where xj is a vector of length 4N comprising the ϕn, i.e.

xj=(w0,,wN1,χ0,,χN1,θ0,,θN1,q0,,qN1)T, 6.4

and AijAij(a,R) and BijBij(C,P1) are 4N×4N matrices with constant coefficients inclusive of the expanded boundary conditions (cf. Straughan [17]).

Given R, C and P1, therefore, system (6.3) may be solved for the wavenumber a subject to the condition ℜ{σ}=0; critical values are then obtained by minimizing Ra(a)=R2. In this way, both stationary (γ=0) and oscillatory (γ≠0) solutions may be found such that subsequent variation of C yields an intersection problem for the threshold Cattaneo number CT (figure 3).

Figure 3.

Figure 3.

Critical Rayleigh numbers Ra (a) and wavenumbers a (b) plotted as a function of C for P1=6 in the case of fixed boundaries (cf. figure 1). Beyond the threshold value CT≈0.02223 (circles) we have RS<Rc, in which case oscillatory convection (dashed curves) is preferred over stationary convection (solid curves), and the wavenumber shifts discontinuously (circles). Note that it has not been possible to compute RS for Cattaneo number less than approximately 0.0165; presumably this is because—as we had in the case of free boundaries in figure 1, where CB≈0.0273—there exists some bounding Cattaneo number CB≈0.0165 below which solutions are non-physical. As elsewhere, data in these plots are given to ±0.1% uncertainty, and computed using N=40 polynomials (cf. table 1).

Truncation of the Chebyshev polynomial expansion means that the accuracy of the solutions given by our numerical method are dependent on N. For the range of parameters explored in this article we find—up to six significant figures at least—that changing the number of Chebyshev polynomials from N=40 to N=50 does not impact on computed values, and so all numerical data are quoted assuming N=40. A more important consideration turns out to be the resolution for the scan over wavenumber a; here, we have employed a grid such that the critical wavenumber aS may be determined to within ±0.1%. The threshold Cattaneo number CT then follows by examining RS(C) and finding upper Cu and lower Cl bounds such that RS(Cl)>Rc>RS(Cu); again, by choosing a suitably fine grid for test values of C, this process allows us to resolve the threshold value CT to within ±0.1% (and similarly for γT). Note that investigation of the upper au and lower al wavenumbers associated with the Cu and Cl bounds on CT (the bounding wavenumbers around aT) indicates a corresponding uncertainty on aT to within the ±0.1% for which aS is initially computed, i.e.

12|aualaT|<0.001,whereaT=12(al+au). 6.5

Data corresponding to the intersection problem when P1=6 are displayed graphically in figure 3, in which case we find a threshold Cattaneo number CT=(2.223±0.0022)×10−2, consistent with Straughan's lower precision result of CT∈(2.2×10−2,2.3×10−2) [17]. Note from the logarithmic plot of the Rayleigh number that the oscillatory solution RS appears to approach a power law RS∝1/C as the Cattaneo number becomes large, and therefore—in a qualitative fashion at least—seems to obey similar behaviour to the free boundary solution discussed in §4 (cf. figure 1). Likewise, as C is increased beyond the threshold value CT, the solution switches from stationary convection to oscillatory convection with a discontinuous transition to a larger wavenumber and narrower convection cells, that is, (ac,aS=aT)≈(3.116,4.873) at the threshold. Again, this behaviour corresponds to that seen in the case of free boundaries, with further increases to C resulting in a reduction in the wavenumber (suggesting subsequent broadening of convection cells) as aS tends towards some constant asymptotic limit.

Naturally, one may solve the intersection problem for different values of P1, and thereby investigate the effect of Prandtl number on those values of Cattaneo number CT, wavenumber aT and oscillation frequency γT at which the transition from stationary to oscillatory convection occurs. Indeed, as we did for free boundary conditions in §5, one may then divide our P1-C parameter space into regions where either stationary convection (Rc<RS) or oscillatory convection (Rc>RS) is preferred (figure 4). Here, we compute 401 values for Prandtl number in the range P1[102,10+2] (table 1). In a similar fashion to the free boundary case, note that the threshold Cattaneo number appears to obey some kind of asymptotic behaviour, with CT1/P1 when the Prandtl number is small, tending towards a constant value CT(P1)0.02 when P1 is large (cf. figure 2). This value is comparable to that obtained for free-boundaries, where we found CT≥8/27π2≈0.03 (see §5).

Figure 4.

Figure 4.

The threshold Cattaneo number CT plotted as a function of P1 for fixed boundary conditions. As we found in the free boundary case, for C=CT we have RS=Rc, so the curve CT(P1) divides P1-C parameter space into regions where either oscillatory convection (C>CT) or stationary convection (C<CT) is preferred (cf. figure 2). A subset of the data used to produce this figure is given in table 1.

Table 1.

Numerically computed values for the threshold Cattaneo numbers CT, wavenumbers aT, and oscillation frequencies γT in the case of fixed boundary conditions. This table comprises a subset of the 401 values used to produce figure 5 for Prandtl numbers in the range P1[102,10+2], with data quoted to within ±0.1% uncertainty.

P1 CT aT γT P1 CT aT γT
10−2.0 1.234 5.621 212.5 10+0.0 0.03214 5.175 12.10
10−1.9 0.9894 5.786 199.1 10+0.1 0.02965 5.123 9.928
10−1.8 0.7935 5.954 186.5 10+0.2 0.02768 5.073 8.094
10−1.7 0.6368 6.110 173.5 10+0.3 0.02614 5.027 6.569
10−1.6 0.5116 6.232 159.6 10+0.4 0.02492 4.986 5.307
10−1.5 0.4116 6.295 144.0 10+0.5 0.02396 4.950 4.271
10−1.4 0.3318 6.292 127.3 10+0.6 0.02321 4.918 3.428
10−1.3 0.2681 6.228 110.4 10+0.7 0.02262 4.891 2.746
10−1.2 0.2173 6.127 94.52 10+0.8 0.02215 4.869 2.195
10−1.1 0.1768 6.008 80.24 10+0.9 0.02178 4.850 1.752
10−1.0 0.1445 5.891 67.94 10+1.0 0.02149 4.834 1.397
10−0.9 0.1189 5.783 57.54 10+1.1 0.02126 4.821 1.112
10−0.8 0.09842 5.687 48.81 10+1.2 0.02108 4.811 0.8860
10−0.7 0.08215 5.608 41.44 10+1.3 0.02093 4.802 0.7049
10−0.6 0.06924 5.535 35.18 10+1.4 0.02082 4.796 0.5608
10−0.5 0.05898 5.469 29.80 10+1.5 0.02073 4.790 0.4459
10−0.4 0.05085 5.407 25.16 10+1.6 0.02066 4.785 0.3544
10−0.3 0.04442 5.347 21.14 10+1.7 0.02060 4.781 0.2817
10−0.2 0.03933 5.288 17.66 10+1.8 0.02055 4.778 0.2238
10−0.1 0.03531 5.232 14.67 10+1.9 0.02052 4.776 0.1779
10−0.0 0.03214 5.175 12.10 10+2.0 0.02049 4.774 0.1414

7. Effect of Prandtl number on threshold values

Recall that in the case of free boundary conditions the wavenumber aT2aS2(P1,CT) and frequency γγT(P1,CT) associated with the oscillatory solution at threshold Cattaneo number are defined by

aT2π2[(P1+1)2π2CTP12+1]1/2andγT214CT2P13[P1(2CT(π2+aT2)1)1] 7.1

respectively (cf. equations (4.5) and (4.13)), whereas for fixed boundaries these threshold values are given numerically by the returned values of aS(CT) and γ(CT) at the threshold solution Rc=RS(CT) (table 1). Given their explicit dependence on P1, alongside the P1 dependence of CT itself, it is therefore clear that both aT and γT will be influenced by the Prandtl number rather strongly (figure 5).

Figure 5.

Figure 5.

Threshold wavenumbers aTaS(P1,CT) (a,b) and frequencies γγT(P1,CT) (c,d) as a function of Prandtl number P1 for the case of both free (a,c) and fixed (b,d) boundary conditions. The analytical free boundary solutions (see equations (7.1)) have asymptotic limits with P1 (dashed curves) given by equations (5.10).

Plots of the threshold wavenumbers and frequencies as a function of Prandtl number in the range P1[102,10+2] are given for both free and fixed boundary conditions in figure 5. Note in the free boundary case that aT decreases with P1, with asymptotic behaviour aTP11/4 when P1 is small, converging to the value aT(P1)=π when the Prandtl number is large (see the limiting behaviour expressed in equation (5.10a)). Because aT>π>ac=π/2, this means that the transition from stationary to oscillatory convection is always marked initially by a discontinuous shift to smaller convection cells. Such an effect is to a certain extent replicated in the case of fixed boundary conditions; indeed, again we see what appears to be convergence towards a constant value aT≈4.77 when the Prandtl number is large (table 1), while—on the domain P1[102,10+2] at least—we have aT>ac≈3.116 (see equation (3.8)). In this second case, however, we find that the threshold wavenumber initially increases with P1, before decreasing, reaching a maximum value aT≈6.3 at P10.035.

In the absence of a full nonlinear analysis (which would reveal more detailed information about the geometry of convection cells, but is well beyond the scope of the present article), it is not immediately clear why this discrepancy between the boundary condition regimes should arise at low P1. What is more, our picture is further complicated by the highly coupled nature of the threshold system whereby both Cattaneo number and Prandtl number play a part, with CT dependent on P1. However, some insights may be reached by returning to the definition of P1, viz

P1=νρ0cVκ 7.2

(see equations (2.11)); thus, at fixed conductivity κ, we see that the low Prandtl number regime corresponds to one with vanishing viscosity.1 In the case of fixed boundaries at low P1, therefore, there is a marked contrast between the vanishing ‘stiffness’ of the fluid between the planes of confinement (ν0), and the no-slip condition for the velocity of the fluid at the walls; boundary effects are thus far more significant. Compared with the free boundary case, when no such difference is present, it is then not so surprising that the behaviour at low Prandtl number should be distinct. Of course, why the value of aT should reach a maximum when the boundaries are fixed, rather than simply converging to another constant value as P10, or indeed how P1 affects aT more generally in either regime, are both questions in need of resolution. However, because the wavenumber will correspond to the geometry of the convection cells, it is the opinion of the author that satisfactory answers will require undertaking a more sophisticated nonlinear study than is appropriate to consider here.

On the other hand, a physical basis for the P1 dependence of the threshold oscillation frequency γT is more apparent. Indeed, in both boundary regimes, one expects that high viscosity (i.e. large Prandtl number) will impede the overall motion of the fluid, thereby curtailing oscillatory motion. Such behaviour, whereby γT always decreases with P1, is present for both free and fixed boundary conditions (figure 5), whereas the slightly ‘kinked’ dependence in the latter case seems in some way to reflect the corresponding variation in the wavenumber (cf. the γT dependence on aT in the case of free boundaries indicated by equations (7.1)).

8. Conclusion

We have studied the canonical Rayleigh–Bénard convection problem of a Boussinesq fluid layer heated from below, using the hyperbolic Cattaneo–Christov heat-flow model in place of the more usual Fourier law [2,17,25,26,32,33], and in so doing developed a linear theory for thermal instability by oscillatory modes (which are forbidden classically, see §§14). In the case of free boundary conditions, critical Rayleigh numbers for both stationary and oscillatory convection, denoted Rc and RS respectively, may be determined analytically, and thus allow for a comparison between preferred forms of instability. When considering stationary convection the perturbation time dependence has no component of gyration (σ==0), meaning that both Cattaneo number C and Prandtl number P1 effects vanish (i.e. those terms in σC and σP1). Under these conditions, therefore, solutions are equivalent to the classical results obtained with the more usual Fourier law, and onset of instability occurs with critical Rayleigh number Rc=27π4/4≈657.511, corresponding to a critical wavenumber ac=π/22.221 [26]. For oscillatory convection, however, when γ≠0, the critical Rayleigh number RS acquires both C and P1 dependence, so that by examining the necessary condition for physical solutions γ2>0 overstability is rendered possible provided C is larger than a lower bounding value CB(P1). In this case, the associated critical wavenumber aS also acquires C and P1 dependence, with aS(C,P1)>ac for all C,P1R+, converging on the classical result aSac as C,P1.

Crucially, for given Prandtl number, we have demonstrated that RS(C) is a strictly decreasing function of C, obeying a dependence RS(C)1/C20<Rc when C is large, whereas for relatively small values of the Cattaneo number RS exceeds the classical (Fourier law) value with RS(C)>Rc. Thus, provided C>CB, there exists some threshold Cattaneo number CT such that RS(CT)=Rc beyond which the preferred form of instability switches from stationary to oscillatory convection. Notably, such a transition is marked by a discontinuous shift from the stationary wavenumber ac=π/2 to narrow convection cells in the oscillatory regime, with associated wavenumber aS>ac; for example, at Prandtl number P1=6, when CT≈0.0342, we found a threshold oscillatory wavenumber aT=aS(CT)≈3.347>ac≈2.221 (figure 1).

Indeed, explicit calculations of the threshold Cattaneo number have been presented here for the first time, allowing us to divide P1-C parameter space in a general sense via the curve CT(P1) into regions where either stationary (C<CT) or oscillatory (C>CT) convection is preferred. When expressed at fixed C in terms of the threshold Prandtl number PT, this curve may also be written in closed form (§5). Such an analysis has enabled us to demonstrate rigorously that CT(P1)>CB(P1) for arbitrary Prandtl number, and therefore that threshold behaviour associated with the critical convective mode (n=1) is always permitted physically.

For fixed boundary conditions, both stationary and oscillatory behaviour may be studied numerically using a Chebyshev-tau algorithm [34], after expanding the problem into Chebyshev polynomials and approximating the system in terms of a finite number of linear equations (§6). As expected, in the case of stationary convection one recovers the classical (Fourier law) results Rc≈1707.776 and ac≈3.116 [26], whereas for oscillatory convection one finds that the overstable Rayleigh number RS is less than Rc for C sufficiently large. In this way, we have studied the transition from stationary-to-oscillatory convection for Prandtl numbers in the range P1[102,10+2], paying special attention to the wavenumbers aT and oscillation frequencies γT associated with the threshold solution RS(CT)=Rc. Note that in the fixed boundary case we find that CT0.02, whereas for free boundaries this value is slightly higher, with CT≥8/27π2≈0.03. By using N=40 polynomials, such an approach yields relatively precise values, to within ±0.1% uncertainty. As in the case of free boundaries, we find that the Prandtl number has important consequences for the threshold parameters; indeed, the effect of low P1 on the wavenumber aT is particularly pronounced, possibly owing to the contrast between no-slip conditions at the boundary, and inviscid (or very low ν) motion within the fluid layer (§7).

Overall, our analysis has shown that hyperbolic heat-flow effects have profound consequences for thermal convection, and substantially lower instability thresholds when the Cattaneo number is relatively large, in which case the Rayleigh number for oscillatory modes scales as RS∝1/C2. Why this particular scaling should obtain is not immediately clear; however, the emergence of oscillatory solutions makes sense when we consider the relationship between our problem and overstability in classical systems. Indeed, Cattaneo terms impart a kind of elasticity to the fluid, and allow thermal disturbances to propagate as waves in a semi-analogous fashion to the way in which rotational effects and impressed magnetic fields permit wave propagation and overstability in classical convection [26]. Intuitively, therefore, one expects that increases to the Cattaneo number should make this effect more pronounced, meaning that the onset of oscillatory convection occurs at lower Rayleigh numbers when C is large (with ∂RS/∂C<0), as we have observed.

As discussed in the Introduction, new technologies present heat transfer problems on ever reduced scales, and the phenomena we describe seem likely to be of particular relevance to such systems (where hyperbolic heat-flow effects are known to be especially pronounced), alongside biological contexts for which thermal relaxation times are relatively long [811,1315]. Nevertheless, one of the pressing issues surrounding hyperbolic modifications to the usual Fourier law of heat conduction is to establish which kind of heat-flow model should be properly employed (the model used here is simply one more frequently encountered). Crucially, our discussion emphasizes that a Cattaneo–Christov heat-flow law has implications beyond simply the speed at which thermal signals propagate; rather, that Cattaneo terms lead to pronounced transitions in overall system behaviour: in our case, the emergence of discontinuous transitions from stationary to (otherwise forbidden) oscillatory convection, with marked shifts in wavenumber and gyration frequency. In terms of experimental application, such transitions have the potential to act as very clear signatures of governing dynamics, and thence—in some respects at least—a means of model validation. Indeed, it seems plausible that by studying the transition between forms of convection within an appropriate experimental context (e.g. small-scale system), one may determine various Cattaneo thresholds CT, and thence relaxation times τ. In this respect, a key advantage of our analysis is the inclusion of effects owing to Prandtl number P1, which in some circumstances can be used as a control parameter [35]; for example, we have shown that the Cattaneo threshold CT(P1) can be conceived equivalently as a Prandtl threshold PT(C), so that system bifurcations could potentially be triggered by varying P1.

The theory presented here substantial develops existing work on Cattaneo–Christov heat-flow effects in thermal convection [16,17]; however, a number of theoretical issues precluded by our linear approach remain, not least that of developing an understanding of how the Prandtl number impacts on convection cell geometry, and thus the physical basis for discontinuous transitions between stationary ac and convective aT=aS(CT) wavenumbers at the Cattaneo threshold. Indeed, having investigated Cattaneo–Christov heat-flow effects in the classic Rayleigh–Bénard system, it is important to consider its role in other thermal convection systems, especially those which are known to yield oscillatory or non-stationary instability under the more usual Fourier law, such as magnetized or rotating fluids [26,28]. The author plans to address some of these problems in future publications.

Acknowledgements

The author thanks Professor B. Straughan for many stimulating discussions, and two anonymous referees for insightful comments on the original manuscript which have helped to clarify its argument.

Footnotes

1

Although values of CT are in a sense solely dependent on P1, the Cattaneo number C is itself a function of the thermal conductivity, and so here it is appropriate to avoid ambiguity by asserting that κ is held constant (see equations (2.11)).

Data accessibility

There are no additional datasets associated with this article.

Funding statement

J.J.B. is sponsored by a Leverhulme Trust Grant (Tipping Points Project, University of Durham).

Conflict of interests

The author reports no competing interests.

References

Associated Data

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

Data Availability Statement

There are no additional datasets associated with this article.


Articles from Proceedings. Mathematical, Physical, and Engineering Sciences / The Royal Society are provided here courtesy of The Royal Society

RESOURCES