Open Access

Global Stability For Double-Diffusive Convection In A Couple-Stress Fluid Saturating A Porous Medium


Cite

Introduction

Conventional hydrodynamic stability theory is mainly concerned with the determination of critical values of Rayleigh number, demarcating a region of stability from that of instability. The potentials of linear theory of stability and the energy method are complementary to each other in the sense that the linear theory gives conditions under which the hydrodynamic system is definitely unstable. It cannot with certainty conclude stability. On the other hand, the energy theory gives conditions under which the hydrodynamic system is definitely stable. It cannot with certainty conclude instability. Suffering from its basic assumptions, the validity of the linearized stability theory becomes questionable.

Hence, the non-linear approach becomes inevitable to investigate the effect of finite disturbances. The formulation and derivation of the basic equation of a layer of fluid heated and soluted from below in the porous medium using Boussinesq approximation has been given in a treatise by Joseph [1]. When a fluid flows in an isotropic and homogenous porous medium, the gross effect is represented by the Darcy’s law. The study of a layer of fluid heated and soluted from below in the porous media is motivated both theoretically and by its practical application in engineering. Among the application in engineering disciplines, one can find the food process industry, chemical process industry and solidification and centrifugal casting of metals. The oldest method of non-linear stability analysis that can deal with finite disturbances is the energy method, originated by Orr [2], and its recent revival has been inspired by the work of Serrin [3] and Joseph [1, 4, 5]. Rapid improvements of the classical energy theory have been made in recent years [6]. The approach adopted in the present article is by the application of the energy method, pioneered and developed in its modern use way by Straughan [7, 8]. Straughan [9] developed a sharp non-linear energy stability analysis for the saturated porous medium, and the results obtained are the best possible showing that subcritical instabilities are not possible. By selecting the optimal, it has been possible to sharpen the stability bound in many physical problems (Straughan [8]). A nonlinear stability analysis of fluids by using generalized energy stability theory has been considered by many authors[10, 11, 12, 13, 14, 15]).

There are a lot of analyses of performance and experiment in the couple-stress lubricant. Stokes [16] proposed a simplest theory called the Stokes microcontinum theory, which could be used for the simulation of couple-stress fluid. One of the applications of couple-stress fluid is its use in the study of the mechanism of lubrication of synovial joints, which has become the object of scientific research. A human joint is a dynamically loaded bearing that has auricular cartilage as the bearing and synovial fluid as the lubricant. When a fluid film is generated, squeeze-film action is capable of providing considerable protection to the cartilage surface. Ramanaian [17] applied the couple-stress fluid model to analyze the long slider bearing. Sharma and Thakur [18] and Sharma et al. [19] have studied the problems of couple-stress fluid heated and soluted from below in the hydromagnetic porous medium and rotation separately. Recently, Sunil and Mahajan [20, 21, 22, 23] studied the non-linear stability analysis for thermal convection in a magnetized ferrofluid heated from below saturating a porous medium. Sunil et al. [24, 25] studied the non-linear stability analysis for thermal convection in a couple-stress fluid heated from below saturating a porous medium. Hsu et al. [26] studied the combined effects of couple stress and surface roughness using journal bearings lubricated with the non-Newtonian fluid. It was found that the combined effects of couple stress and surface roughness can improve the load carrying capacity and decrease the attitude angle and friction parameter. Recently, Lahmar [27] also found that the lubricants with couple-stress fluid would increase the load carrying capacity and stability and decrease the friction factor and the attitude angle.

The purpose of the present article was to study the non-linear stability analysis of couple-stress fluid heated and soluted from below, saturating a porous medium of high permeability [28]. The really interesting situation from a mathematical viewpoint arises when the layer is simultaneously heated and salted from below. In the standard Bènard problem, the instability is driven by a density difference caused by a temperature difference between the upper and lower surfaces bounding the fluid. If, additionally, the fluid layer has salt dissolved in it, then there are potentially two destabilizing sources for the density difference, the temperature field and the salt field. When there are two effects such as this, the phenomenon of convection that arises is called double-diffusive convection. The driving force for many studies in double-diffusive or multicomponent convection is largely physical applications. The double-diffusive convection problems have been studied by many authors [13-14, 29, 30, 31, 32, 33, 34, 35, 36, 37, 38, 39,]. In porous media, an alternative to Darcy’s equation is what is known as Brinkman’s equation [28]. It is believed that for the flow of a high-porosity porous medium, the Brinkman equation removes some of the deficiencies and gives preferable result. In the work of Qin and Kaloni [40], it was remarked that for high porosity materials and when boundary layer effects need to be taken into account, the Brinkman model is superior to Darcy’s model. Here, we establish the optimal result, that is the linear instability and non-linear stability of Rayleigh numbers are the same, i.e. R𝓁fRef We also find that the critical value of thermal Rayleigh number for the couple-stress fluid is higher than the critical value of thermal Rayleigh number for the ordinary fluid; hence, the couple-stress fluid is thermally more stable than the ordinary fluid. This problem, to the best of my knowledge, has not been investigated yet.

Figure 1

Geometrical configuration of the problem.

Mathematical Formulation Of The Problem

Here, we consider an infinite, horizontal layer of thickness ‘ d ’ of incompressible thin couple-stress fluid with constant viscosity that is heated and soluted from below, saturating an isotropic homogeneous porous medium of porosity ε and medium permeability K1 .

The fluid is assumed to occupy the layer z(d2,d2)$z\in \left( -\frac{d}{2},\,\frac{d}{2} \right)$with gravity field g = (0,0,-g ) pervading the system in the negative z-direction.

The equations governing the flow of an incompressible couple-stress fluid (utilizing the Boussinesq approximation) are given as follows [13, 24, 41]:

Mass balance:

q=0.$$\nabla \cdot \mathbf{q}=0.$$

Momentum balance:

ρ0ε[qt]=p+ρ0[1α(TTa)+α(CCa)]g1K1(μμ2)q+μ˜2q.$$\frac{{{\rho }_{0}}}{\varepsilon }\left[ \frac{\partial \mathbf{q}}{\partial t} \right]=-\nabla p+{{\rho }_{0}}\left[ 1-\alpha \left( T-{{T}_{a}} \right)+{\alpha }'\left( C-{{C}_{a}} \right) \right]g-\frac{1}{{{K}_{1}}}\left( \mu -{\mu }'{{\nabla }^{2}} \right)\mathbf{q}+\tilde{\mu }{{\nabla }^{2}}\mathbf{q}.$$

Temperature equation:

(ρ0C0)mTt+(ρ0C0)fqT=.(kT).$${{\left( {{\rho }_{0}}{{C}_{0}} \right)}_{m}}\frac{\partial T}{\partial t}+{{\left( {{\rho }_{0}}{{C}_{0}} \right)}_{f}}\mathbf{q}\cdot \nabla T=\nabla .\left( k\nabla T \right).$$

Solute equation:

(ρ0C0)mCt+(ρ0C0)fqC=.(kC).$${{\left( {{\rho }_{0}}{{C}_{0}} \right)}_{m}}\frac{\partial C}{\partial t}+{{\left( {{\rho }_{0}}{{C}_{0}} \right)}_{f}}\mathbf{q}\cdot \nabla C=\nabla .\left( {k}'\nabla C \right).$$

Here ρ, ρ 0 , q , g , t , p , μ , μ' , μ˜$\tilde{\mu }$, ĸ , ĸ] ' , K1 , ε ,α , , α' and C0 are the fluid density, reference density, filter velocity, acceleration due to gravity, time, pressure, coefficient of viscosity, coefficient of visco-elasticity, effective viscosity, thermal diffusivity, solute diffusivity, permeability of porous medium, porosity, thermal expansion coefficient, concentration expansion coefficient analogous to the thermal expansion coefficient and specific heat at constant pressure, respectively. The subscripts ‘m’ and ‘f’ refer to the fluid– solid mixture and the fluid, respectively. Ta and Ca are the average temperature and solute concentration given by Ta=(TL+TU)2${{T}_{a}}=\frac{\left( {{T}_{L}}+{{T}_{U}} \right)}{2}$and Ca=(CL+CU)2${{C}_{a}}=\frac{\left( {{C}_{L}}+{{C}_{U}} \right)}{2}$respectively, where TL , TU and CLCU are the constant average temperatures and solute concentrations of the lower and upper surfaces of the layer, respectively, and β(=|dTdz|)andβ(=|dCdz|)$\beta \left( =\left| \frac{dT}{dz} \right| \right)\text{and}\,{\beta }'\left( =\left| \frac{dC}{dz} \right| \right)$are uniform temperature and solute gradients, respectively.

The basic state is assumed to be the quiescent state and is given by

q=qb=(0,0,0),p=pb(z),ρ=ρb(z)=ρ0(1+αβzαβz),T=Tb(z)=βz+Ta,C=Cb(z)=βz+Ca,β=TLTUd,β=CLCUd,$$\begin{align}& \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\mathbf{q}={{\mathbf{q}}_{b}}=\left( 0,0,0 \right),\,\,\,\,\,p={{p}_{b}}\left( z \right), \\ & \rho ={{\rho }_{b}}\left( z \right)={{\rho }_{0}}\left( 1+\alpha \beta z-{\alpha }'{\beta }'z \right),\,T={{T}_{b}}\left( z \right)=-\beta z+{{T}_{a}}, \\ & C={{C}_{b}}\left( z \right)=-{\beta }'z+{{C}_{a}},\,\beta =\frac{{{T}_{L}}-{{T}_{U}}}{d},\,{\beta }'=\frac{{{C}_{L}}-{{C}_{U}}}{d}, \\ \end{align}$$

where the subscript 'b' denotes the basic state.

We shall analyze the stability of the basic state by introducing the following perturbations:

q=qb+q,ρ=ρb+ρ,p=pb(z)+p,T=Tb(z)+θandC=Cb(z)+γ.$$\begin{align}& \mathbf{q}={{\mathbf{q}}_{b}}+\mathbf{{q}'},\,\,\,\,\rho ={{\rho }_{b}}+{\rho }',\,\,\,p={{p}_{b}}\left( z \right)+{p}',\,\,T={{T}_{b}}\left( z \right)+\theta \,\,\text{and} \\ & \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,C={{C}_{b}}\left( z \right)+\gamma . \\ \end{align}$$

The non-linear equations for the perturbations q' = (u,v,w), p', ρ' ', θ and γ , which represent perturbations in velocity q , pressure p, density ρ, , temperature T and concentration C , respectively, are given by

ρ0εqt=p+ρ0g(αθαγ)k^μK1q+μK1μ2q+μ˜2q'$$\frac{{{\rho }_{0}}}{\varepsilon }\frac{\partial \mathbf{{q}'}}{\partial t}=-\nabla {p}'+{{\rho }_{0}}g\left( \alpha \theta -{\alpha }'\gamma \right)\mathbf{\hat{k}}-\frac{\mu }{{{K}_{1}}}\mathbf{{q}'}+\frac{\mu }{{{K}_{1}}}{\mu }'{{\nabla }^{2}}\mathbf{{q}'}+\tilde{\mu }{{\nabla }^{2}}\mathbf{q}'$$q=0,$$\nabla \cdot \mathbf{{q}'}=0,$$Aθt+qθ=κ2θ+βw,$$A\frac{\partial \theta }{\partial t}+\mathbf{{q}'}\cdot \nabla \theta =\kappa {{\nabla }^{2}}\theta +\beta w,$$Aγt+qγ=κ2γ+βw,$$A\frac{\partial \gamma }{\partial t}+\mathbf{{q}'}\cdot \nabla \gamma ={\kappa }'{{\nabla }^{2}}\gamma +{\beta }'w,$$

where A=(ρ0C0)m(ρ0C0)f,κ=k(ρ0C0)fandκ=k(ρ0C0)f.$A=\frac{{{\left( {{\rho }_{0}}{{C}_{0}} \right)}_{m}}}{{{\left( {{\rho }_{0}}{{C}_{0}} \right)}_{f}}},\,\,{\kappa }'=\frac{{{k}'}}{{{\left( {{\rho }_{0}}{{C}_{0}} \right)}_{f}}}\,\,\,\,\,\,\,\,\text{and}\,\,\,\,\kappa =\frac{k}{{{\left( {{\rho }_{0}}{{C}_{0}} \right)}_{f}}}.$

The boundary conditions are

q=0,θ=0,γ=0at  z=±d2,$$\mathbf{{q}'}=\mathbf{0},\,\,\theta =0,\,\gamma =0\,\,\,\,\,\,\text{at}\,\,\,\,\,\,\,\,\,\text{ z}=\pm \frac{d}{2},$$

with q',θ andγ satisfying the plane tiling periodicity.

Non-Linear Stability Analysis

To investigate the non-linear stability analysis, the governing equations (7), (8), (9), (10) in the non-dimensional form (dropping asterisk) can be written as

1Vaqt=p+R1/2θk^q+(F+D˜a)2qS1/2Leγk^,$$\frac{1}{{{V}_{a}}}\frac{\partial \mathbf{q}}{\partial t}=-\nabla p+{{R}^{{1}/{2}\;}}\theta \mathbf{\hat{k}}-\mathbf{q}+\left( F+{{{\tilde{D}}}_{a}} \right){{\nabla }^{2}}\mathbf{q}-\frac{{{S}^{{1}/{2}\;}}}{Le}\gamma \mathbf{\hat{k}},$$q=0,$$\nabla \cdot \mathbf{q}=0,$$Aγt+qγ=1Le2γ+S1/2w,$$A\frac{\partial \gamma }{\partial t}+\mathbf{q}\cdot \nabla \gamma =\frac{1}{{{L}_{e}}}{{\nabla }^{2}}\gamma +{{S}^{{1}/{2}\;}}w,$$Aγt+qγ=1Le2γ+S1/2w,$$A\frac{\partial \gamma }{\partial t}+\mathbf{q}\cdot \nabla \gamma =\frac{1}{{{L}_{e}}}{{\nabla }^{2}}\gamma +{{S}^{{1}/{2}\;}}w,$$

where the following non-dimensional quantities and parameters are introduced:

t*=κd2t,q*=dκq,θ=R1/2βdθ,p*=K1μκp,z*=1dz,Vr=μ˜μ,Va=εvd2κK1,F=μvρ0d2,Da=K1d2,Le=κκ,Pr=vk,S=gαβρ0d2K1μκ,γ*=S1/2βdγ,R=gαβρ0K1d2μκ,D˜a=μ˜K1μd2=VrDa.$$\begin{align}& \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,{{t}^{*}}=\frac{\kappa }{{{d}^{2}}}{t}',\,{{\mathbf{q}}^{*}}=\frac{d}{\kappa }{q}',\,{\theta }'=\frac{{{R}^{{1}/{2}\;}}}{\beta d}{\theta }',\,\,{{p}^{*}}=\frac{{{K}_{1}}}{\mu \kappa }{p}', \\ & \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,{{z}^{*}}=\frac{1}{d}z,\,\,{{V}_{r}}=\frac{{\tilde{\mu }}}{\mu },\,{{V}_{a}}=\frac{\varepsilon v{{d}^{2}}}{\kappa {{K}_{1}}}, \\ & F=\frac{{{\mu }'}}{v{{\rho }_{0}}{{d}^{2}}},\,{{D}_{a}}=\frac{{{K}_{1}}}{{{d}^{2}}},\,\,{{L}_{e}}=\frac{\kappa }{{{\kappa }'}},\,\,{{P}_{r}}=\frac{v}{k},\,\,S=\frac{g{\alpha }'{\beta }'{{\rho }_{0}}{{d}^{2}}{{K}_{1}}}{\mu {\kappa }'},\,\,\,{{\gamma }^{*}}=\frac{{{S}^{{1}/{2}\;}}}{{\beta }'d}\gamma , \\ & \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,R=\frac{g\alpha \beta {{\rho }_{0}}{{K}_{1}}{{d}^{2}}}{\mu \kappa },\,\,{{{\tilde{D}}}_{a}}=\frac{\tilde{\mu }{{K}_{1}}}{\mu {{d}^{2}}}={{V}_{r}}{{D}_{a}}. \\ \end{align}$$

Here, R is the Rayleigh–Darcy number, which is the product of Darcy number and the usual Rayleigh number for a clear viscous fluid; a is the Darcy–Brinkman number; Da is the Darcy number; Va is the Vadasz number (as named by Straughan [9]), F is the Couple-stress parameter, S is the solute Rayleigh–Darcy number and Le is the Lewis number.

On multiplying (12) by q, 14) byθ , (15) by γ and integrating over V, we get (after using equation (11), the boundary conditions and the divergence theorem):

12Vad||q||2dt=||q||2(F+D˜a)||q||2+R1/2wθS1/2Lewγ,$$\frac{1}{2{{V}_{a}}}\frac{d{{\left| \left| \mathbf{q} \right| \right|}^{2}}}{dt}=-{{\left| \left| \mathbf{q} \right| \right|}^{2}}-\left( F+{{{\tilde{D}}}_{a}} \right){{\left| \left| \nabla \mathbf{q} \right| \right|}^{2}}+{{R}^{{1}/{2}\;}}\left\langle w\theta \right\rangle -\frac{{{S}^{{1}/{2}\;}}}{{{L}_{e}}}\left\langle w\gamma \right\rangle ,$$A2d||θ||2dt=||θ||2+R1/2wθ,$$\frac{A}{2}\frac{d{{\left| \left| \theta \right| \right|}^{2}}}{dt}=-{{\left| \left| \nabla \theta \right| \right|}^{2}}+{{R}^{{1}/{2}\;}}\left\langle w\,\theta \right\rangle ,$$A2d||γ||2dt=1Le||γ||2+S1/2wγ,$$\frac{A}{2}\frac{d{{\left| \left| \gamma \right| \right|}^{2}}}{dt}=-\frac{1}{{{L}_{e}}}{{\left| \left| \nabla \gamma \right| \right|}^{2}}+{{S}^{{1}/{2}\;}}\left\langle w\,\gamma \right\rangle ,$$

where $\left\langle \cdot \right\rangle $denotes the integration over V, ||||$\left| \left| \cdot \right| \right|$denotes the L2 (V ) norm and V denotes a typical periodicity cell.

To study the non-linear stability of the basic state (5), an L2 energy, E(t), is constructed using equations (17)(19), and the evolution of E (t) is given by

dEdt=I0D0,$$\frac{dE}{dt}={{I}_{0}}-{{D}_{0}},$$

where

E=A2||θ||2+λ12Va||q||2λ2A2||γ||2,$$E=\frac{A}{2}{{\left| \left| \theta \right| \right|}^{2}}+\frac{{{\lambda }_{1}}}{2{{V}_{a}}}{{\left| \left| \mathbf{q} \right| \right|}^{2}}-\frac{{{\lambda }_{2}}A}{2}{{\left| \left| \gamma \right| \right|}^{2}},$$I0=(λ1+1)R1/2wθ(λ2+λ1Le)S1/2wγ,$${{I}_{0}}=\left( {{\lambda }_{1}}+1 \right){{R}^{{1}/{2}\;}}\left\langle w\,\theta \right\rangle -\left( {{\lambda }_{2}}+\frac{{{\lambda }_{1}}}{{{L}_{e}}} \right){{S}^{{1}/{2}\;}}\left\langle w\,\gamma \right\rangle ,$$D0=||θ||2+λ1||q||2+λ1(F+D˜a)||q||2λ2Le||γ||2,$${{D}_{0}}={{\left| \left| \nabla \theta \right| \right|}^{2}}+{{\lambda }_{1}}{{\left| \left| \mathbf{q} \right| \right|}^{2}}+{{\lambda }_{1}}\left( F+{{{\tilde{D}}}_{a}} \right){{\left| \left| \nabla \mathbf{q} \right| \right|}^{2}}-\frac{{{\lambda }_{2}}}{{{L}_{e}}}{{\left| \left| \nabla \gamma \right| \right|}^{2}},$$

with λ1 and λ2 as two positive coupling parameters.

Here, the negative sign with the λ2A2||γ||2$\frac{{{\lambda }_{2}}A}{2}{{\left| \left| \gamma \right| \right|}^{2}}$term in the energy equation (21) shows that energy of the system is consumed due to solute concentration as the system is soluted from below. Now, we take the assumption that the energy consumed due to solute concentration is less than the energy produced due to velocity and temperature. We also assume that the energy dissipated by the solute concentration is less than the energy dissipated by the velocity and temperature. These assumptions will ensure that all the terms on the right-hand side of (21) and (23) are always less than those on the left-hand side of these equations.

We now define,

m=maxHI0D0,$$m=\underset{H}{\mathop{\max }}\,\frac{{{I}_{0}}}{{{D}_{0}}},$$

where H is the space of admissible solutions.

Then, we require m<1 so that

dEdta0D0$$\frac{dE}{dt}\le -{{a}_{0}}{{D}_{0}}$$

where a0 =1-m(> 0) .

From the Poincaré inequality, we have

D0π2||θ||2+λ1[1+π2(F+D˜a)]||q||2λ2Le||γ||2k*E,$${{D}_{0}}\ge {{\pi }^{2}}{{\left| \left| \theta \right| \right|}^{2}}+{{\lambda }_{1}}\left[ 1+{{\pi }^{2}}\left( F+{{{\tilde{D}}}_{a}} \right) \right]{{\left| \left| \mathbf{q} \right| \right|}^{2}}-\frac{{{\lambda }_{2}}}{{{L}_{e}}}{{\left| \left| \gamma \right| \right|}^{2}}\ge {{k}^{*}}E,$$

where k* = 2 π 2 min (A,Va1).$\left( A,V_{a}^{-1} \right).$

This gives

dEdta0k*E$$\frac{dE}{dt}\le -{{a}_{0}}{{k}^{*}}E$$

implying

E(t)ea0k*tE(0).$$E\left( t \right)\le {{e}^{-{{a}_{0}}{{k}^{*}}t}}E\left( 0 \right).$$

Thus, E decays at least exponentially fast, and non-linear stability is assured for all values of E (0). It is important to note that the result holds for all initial data.

Variational Problem

We now return to equation (24) and use calculus of variation to find the maximum problem at the critical argument m =1. The associated Euler–Lagrange equations after taking transformations q^=λ1qandγ^=λ2γ$\mathbf{\hat{q}}=\sqrt{{{\lambda }_{1}}\mathbf{q}}\,\text{and}\,\,\hat{\gamma }=\sqrt{{{\lambda }_{2}}}\gamma $(dropping caps) are

2(F+D˜a)2q2q+R1/2(1+λ1)1λ11/2θk^S1/2(λ2+λ1Le)1λ11/2λ1/22γk^=2p,$$2\left( F+{{{\tilde{D}}}_{a}} \right){{\nabla }^{2}}\mathbf{q}-2\mathbf{q}+{{R}^{{1}/{2}\;}}\left( 1+{{\lambda }_{1}} \right)\frac{1}{\lambda _{1}^{{1}/{2}\;}}\theta \mathbf{\hat{k}}-{{S}^{{1}/{2}\;}}\left( {{\lambda }_{2}}+\frac{{{\lambda }_{1}}}{{{L}_{e}}} \right)\frac{1}{\lambda _{1}^{{1}/{2}\;}{{\lambda }^{{1}/{2}\;}}_{2}}\gamma \mathbf{\hat{k}}=2\nabla p,$$22θ+R1/2(1+λ1)1λ11/2w=0,$$2{{\nabla }^{2}}\theta +{{R}^{{1}/{2}\;}}\left( 1+{{\lambda }_{1}} \right)\frac{1}{\lambda _{1}^{{1}/{2}\;}}w=0,$$2Le2γ+S1/2(λ2+λ1Le)1λ11/2λ1/22w=0,$$\frac{2}{{{L}_{e}}}{{\nabla }^{2}}\gamma +{{S}^{{1}/{2}\;}}\left( {{\lambda }_{2}}+\frac{{{\lambda }_{1}}}{{{L}_{e}}} \right)\frac{1}{{{\lambda }_{1}}^{{1}/{2}\;}{{\lambda }^{{1}/{2}\;}}_{2}}w=0,$$

where p is a Lagrange’s multiplier introduced, since q is solenoidal.

On taking curl curl of equation (29) and then taking the third component of the resulting equation, we find

2(F+D˜a)2w22w+R1/2(1+λ1)1λ11/212θS1/2(λ2+λ1Le)1λ112λ1/2212γ=0.$$2\left( F+{{{\tilde{D}}}_{a}} \right){{\nabla }^{2}}w-2{{\nabla }^{2}}w+{{R}^{{1}/{2}\;}}\left( 1+{{\lambda }_{1}} \right)\frac{1}{{{\lambda }_{1}}^{{1}/{2}\;}}\nabla _{1}^{2}\theta -{{S}^{{1}/{2}\;}}\left( {{\lambda }_{2}}+\frac{{{\lambda }_{1}}}{{{L}_{e}}} \right)\frac{1}{{{\lambda }_{1}}^{\underset{2}{\mathop{1}}\,}{{\lambda }^{{1}/{2}\;}}_{2}}\nabla _{1}^{2}\gamma =0.$$

Now, we assume a plane tiling form

(w,θ,γ)=[W(z),Θ(z),Γ(z)]g(x,y),$$\left( w,\theta ,\gamma \right)=\left[ W\left( z \right),\Theta \left( z \right),\Gamma \left( z \right) \right]g\left( x,y \right),$$

where 12g+a2g=0,$\nabla _{1}^{2}g+{{a}^{2}}g=0,$a being the wave number [9, 42].

The wave number is found a posteriori to be non-zero; thus, from equations (29), (30), (31), we see that W and Θ satisfy

2(F+D˜a)(D2a2)2W2(D2a2)WR1/2a2(1+λ1)1λ11/2Θ+S1/2a2(λ2+λ1Le)1λ11/2λ1/22γ=0,$$\begin{align}& 2\left( F+{{{\tilde{D}}}_{a}} \right){{\left( {{D}^{2}}-{{a}^{2}} \right)}^{2}}W-2\left( {{D}^{2}}-{{a}^{2}} \right)W-{{R}^{{1}/{2}\;}}{{a}^{2}}\left( 1+{{\lambda }_{1}} \right)\frac{1}{{{\lambda }_{1}}^{{1}/{2}\;}}\Theta \\ & \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,+{{S}^{{1}/{2}\;}}{{a}^{2}}\left( {{\lambda }_{2}}+\frac{{{\lambda }_{1}}}{{{L}_{e}}} \right)\frac{1}{{{\lambda }_{1}}^{{1}/{2}\;}{{\lambda }^{{1}/{2}\;}}_{2}}\gamma =0, \\ \end{align}$$2Le(D2a2)Γ+S1/2(λ2+λ1Le)1λ11/2W=0,$$\frac{2}{{{L}_{e}}}\left( {{D}^{2}}-{{a}^{2}} \right)\Gamma +{{S}^{{1}/{2}\;}}\left( {{\lambda }_{2}}+\frac{{{\lambda }_{1}}}{{{L}_{e}}} \right)\frac{1}{{{\lambda }_{1}}^{{1}/{2}\;}}W=0,$$2Le(D2a2)Γ+S1/2(λ2+λ1Le)1λ11/2λ21/2W=0.$$\frac{2}{{{L}_{e}}}\left( {{D}^{2}}-{{a}^{2}} \right)\Gamma +{{S}^{{1}/{2}\;}}\left( {{\lambda }_{2}}+\frac{{{\lambda }_{1}}}{{{L}_{e}}} \right)\frac{1}{{{\lambda }_{1}}^{{1}/{2}\;}{{\lambda }_{2}}^{{1}/{2}\;}}W=0.$$

Thus, the exact solution to the equations (34), (35), (36) subject to boundary conditions

W=0,D2W=0,Θ=0,Γ=0atz=±12$$W=0,\,\,{{D}^{2}}W=0,\,\,\,\Theta =0,\,\,\Gamma =0\,\,\,\text{at}\,\,\text{z}=\pm \frac{1}{2}$$

is written in the form

W=A0cosπz,Θ=B0cosπz,Γ=C0cosπz$$W={{A}_{0}}\cos \pi z,\,\,\Theta ={{B}_{0}}\,\cos \,\pi z,\,\Gamma ={{C}_{0}}\cos \pi z$$

where A0 , B0 and C0 are constants. Substituting solution (37) in equations (34), (35), (36), we get the equations involving coefficients of A0 , B0 and C0 . For the existence of non-trivial solutions, the determinant of the coefficients of A0 , B0 and C0 must vanish. This determinant on simplification yields

Re=4Le(1+x)2{1+(F1+D^a)(1+x)}+xS1λ1λ2(λ2+λ1Le)2x1λ1Le(1+λ1)2,$${{R}_{e}}=\frac{\frac{4}{{{L}_{e}}}{{\left( 1+x \right)}^{2}}\left\{ 1+\left( {{F}_{1}}+{{{\hat{D}}}_{a}} \right)\left( 1+x \right) \right\}+\frac{x{{S}_{1}}}{{{\lambda }_{1}}{{\lambda }_{2}}}{{\left( {{\lambda }_{2}}+\frac{{{\lambda }_{1}}}{{{L}_{e}}} \right)}^{2}}}{x\frac{1}{{{\lambda }_{1}}{{L}_{e}}}{{\left( 1+{{\lambda }_{1}} \right)}^{2}}},$$

where

Re=Rπ2,D^a=π2D˜a,x=a2π2,F1=Fπ2andS1=Sπ2.$${{R}_{e}}=\frac{R}{{{\pi }^{2}}},\,{{\hat{D}}_{a}}={{\pi }^{2}}{{\tilde{D}}_{a}},\,x=\frac{{{a}^{2}}}{{{\pi }^{2}}},\,{{F}_{1}}=F{{\pi }^{2}}\,\text{and}\,{{\text{S}}_{1}}=\frac{S}{{{\pi }^{2}}}.$$

The maximum value of λ1 and λ2 is determined by the conditions SdRedλ1=0and dRedλ2=0$\text{S}\frac{d\,{{R}_{e}}}{d{{\lambda }_{1}}}=0\,\,\text{and }\frac{d\,{{R}_{e}}}{d{{\lambda }_{2}}}=0$and is found to be

λ1=1,λ2=1Le.$${{\lambda }_{1}}=1,\,\,{{\lambda }_{2}}=\frac{1}{{{L}_{e}}}.$$

Using (40) in equation (39), we have

Re=(1+x)2{1+(F1+D^a)(1+x)}x+S1.$${{R}_{e}}=\frac{{{\left( 1+x \right)}^{2}}\left\{ 1+\left( {{F}_{1}}+{{{\hat{D}}}_{a}} \right)\left( 1+x \right) \right\}}{x}+{{S}_{1}}.$$

We obtain the fluid-based thermal Rayleigh number as

Ref=ReDa=(1+x)2{(1+x)Vr+1Da{1+F1(1+x)}}x+Sf.$${{R}_{ef}}=\frac{{{R}_{e}}}{{{D}_{a}}}=\frac{{{\left( 1+x \right)}^{2}}\left\{ \left( 1+x \right){{V}_{r}}+\frac{1}{{{D}_{a}}}\left\{ 1+F{{\,}_{1}}\left( 1+x \right) \right\} \right\}}{x}+{{S}_{f}}.$$

As a function of x , R ef given by equation (42) attains its minimum when

P3x3+P2x2+P0=0,$${{P}_{3}}{{x}^{3}}+{{P}_{2}}{{x}^{2}}+{{P}_{0}}=0,$$

where

P3=2(Vr+F1Da),P2=[1Da+(Vr+F1Da)]andP0=[Vr+(1+F1Da)].$${{P}_{3}}=2\left( {{V}_{r}}+\frac{{{F}_{1}}}{{{D}_{a}}} \right),\,\,{{P}_{2}}=\left[ \frac{1}{{{D}_{a}}}+\left( {{V}_{r}}+\frac{{{F}_{1}}}{{{D}_{a}}} \right) \right]\text{and}\,\,{{P}_{0}}=-\left[ {{V}_{r}}+\left( \frac{1+{{F}_{1}}}{{{D}_{a}}} \right) \right].$$

The thermal Rayleigh number R e f is minimized with respect to x, and we use the Newton–Raphson iterative scheme to obtain the value of critical wave number and the corresponding critical thermal Rayleigh number R ce f (see Tables 13).

The variation in the fluid-based critical thermal Rayleigh number Rcef with the couple-stress parameter F1 for Da=0.02, Vr =1 and Sf = 100.

F1xceR cef
010.9636307.90
20.6154661.40
30.56821001.9
40.54831340.8
50.53741679.0
60.53052017.0
70.525823537
80.52232696.0
90.51973030.3
100.51763367.9
0.51593705.6

For analyzing the linear instability results, we return to the perturbed equations (7), (8), (9), (10), neglecting the nonlinear terms. We again perform the standard stationary normal mode analysis and look for the solution of these equations in the form (38). The boundary conditions in the present case are same, i.e. (37) (here, the thermal Rayleigh number).

Rlf=(1+x)2{(1+x)Vr+1Da{1+F1(1+x)}}x+Sf=Ref$${{R}_{\ell f}}=\frac{{{\left( 1+x \right)}^{2}}\left\{ \left( 1+x \right){{V}_{r}}+\frac{1}{{{D}_{a}}}\left\{ 1+{{F}_{1}}\left( 1+x \right) \right\} \right\}}{x}+{{S}_{f}}={{R}_{ef}}$$

In the absence of solute ( S f = 0 ), this further simplifies to

Rlf=(1+x)2{(1+x)Vr+1Da{1+F1(1+x)}}x=Ref$${{R}_{\ell f}}=\frac{{{\left( 1+x \right)}^{2}}\left\{ \left( 1+x \right){{V}_{r}}+\frac{1}{{{D}_{a}}}\left\{ 1+{{F}_{1}}\left( 1+x \right) \right\} \right\}}{x}={{R}_{ef}}$$

i.e., in both the case, the linear instability boundary ≡ the non-linear stability boundary, and so no subcritical instabilities are possible for the case of couple-stress fluid. This result is equivalent to the result given by Joseph [4, 5] for the standard Bénard problem.

Discussion Of Results And Conclusion

The critical wave numbers xc and xce and critical thermal Rayleigh number R ef =R ce f depends on the parameters Vr , F1, S f and Da . The variation in x c and R cef with the variation in F1 is given in Table 1, that with the variation in Da is given in Table 2 and that with the variation in S f is given in Table 3, and the

The variation in the fluid-based critical thermal Rayleigh number R cef with the couple-stress parameter (Da ) for F1=2, Vr=1 and Sf = 100.

DaxceRcef
0.010.56841897.1
0.020.56821001.9
0.030.56797017
0.040.56765532
0.050.56734636
0.060.5671405.2
0.070.5668365.0
0.080.5665330.6
0.090.5663305.7
0.100.5660285.8
0.110.5667269.5

The variation in the fluid-based critical thermal Rayleigh number R cef with the solute gradient (Sf) for Da, Vr =1.

SRcef(F1=1xce=0.6164)${{R}_{cef}}\left( \begin{align}& {{F}_{1}}=1 \\ & {{x}_{ce}}=0.6164 \\ \end{align} \right)$Rcef(F1=2xce=0.5682)${{R}_{cef}}\left( \begin{align}& {{F}_{1}}=2 \\ & {{x}_{ce}}=0.5682 \\ \end{align} \right)$Rcef(F1=3xce=0.5483)${{R}_{cef}}\left( \begin{align}& {{F}_{1}}=3 \\ & {{x}_{ce}}=0.5483 \\ \end{align} \right)$
100661.41001.91340.8
200761.41101.91440.8
300861.41201.91540.8
400961.41301.91640.8
5001061.41401.91740.8
6001161.41501.91840.8
700271.41601.91940.8
8001361.41701.92040.8
9001461.41801.92140.8

results are further illustrated in Figs. 24, which represent the plot of critical thermal Rayleigh number R cef versus the parameter F1, Da and S f , respectively. Figure 2 indicates that the parameter F1 has the stabilizing effect on convection because as F1 increases, the value of R cef increases. We also note that the value of critical thermal Rayleigh number remains the same for both the theories (linear theory and non-linear theory) and no subcritical instabilities are possible.

Figure 2

The variation in the critical thermal Rayleigh number Rcef with the couple-stress parameter F1.

This conclusion is further strengthened in Fig. 3, which shows the variation in Rcef with Darcy number Da at F = 2, Vr =1. Here, an increase in Da leads to a decrease in Rcef , rendering the system prone to instability. Figure 4 indicates that the solute gradient S f has a stabilizing effect because with the increase in S f , the values of Rcef also increase. Here, two diffusing components heat and salt are present that produce the density differences required to derive the motion. The components make opposing contributions to the vertical density gradient as motion is encouraged due to heating and solute acts to prevent motion through convection overturning. Thus, these two physical effects are competing against each other. We also note that the value of critical thermal Rayleigh number remains the same for both the theories (linear theory and non-linear theory) and no subcritical instabilities are possible. In other words, medium permeability destabilizes the flow. To investigate our result, we must review the results and its physical explanation. When the fluid layer is assumed to be flowing through an isotropic and homogenous porous medium, then the medium permeability has a destabilizing effect. This is because, as medium permeability increases, the void space increases, and as a result of this, the flow quantities perpendicular to the planes will clearly be increased. Thus, an increase in heat transfer is responsible for early onset of convection. Hence, an increase in Da leads to a decrease in Rcef

Figure 3

The variation in the critical thermal Rayleigh number Rcef with the couple -stress parameter Da

Figure 4

The variation in the critical thermal Rayleigh number Rcef with the solute gradient Sf

The principal conclusion from the above analysis is as follows:

The result we establish is that boundaries of nonlinear stability and linear instability analyses coincide with each other. So, no subcritical instabilities are possible.

The couple stress has the tendency to slow down the motion of the fluid in the boundary layer, thus reducing the heat transfer from bottom to top. The a decrease in heat transfer is responsible for delaying the onset of convection. Thus, the couple-stress parameter F1 promotes stabilization.

The medium permeability is found to have destabilizing effect on the system.

It is observed that solute gradient delays the onset of convection and thus has a stabilizing effect on the system.

eISSN:
2083-831X
Language:
English
Publication timeframe:
4 times per year
Journal Subjects:
Geosciences, other, Materials Sciences, Composites, Porous Materials, Physics, Mechanics and Fluid Dynamics