Home Marangoni convection in layers of water-based nanofluids under the effect of rotation
Article Open Access

Marangoni convection in layers of water-based nanofluids under the effect of rotation

  • Abeer H. Bakhsh and Abdullah A. Abdullah EMAIL logo
Published/Copyright: September 8, 2021

Abstract

A linear stability analysis is performed for the onset of Marangoni convection in a horizontal layer of a nanofluid heated from below and affected by rotation. The top boundary of the layer is assumed to be impenetrable to nanoparticles with their distribution being determined from a conservation condition while the bottom boundary is assumed to be a rigid surface with fixed temperature. The motion of the nanoparticles is characterized by the effects of thermophoresis and Brownian diffusion. A modification model is used in which the effects of Brownian diffusion and thermophoresis are taken into consideration by new expressions in the nanoparticle mass flux. Also, material properties of the nanofluid are modelled by non-constant constitutive expressions depending on nanoparticle volume fraction. The steady-state solution is shown to be well approximated by an exponential distribution of the nanoparticle volume fraction. The Chebyshev-Tau method is used to obtain the critical thermal and nanoparticle Marangoni numbers. Different stability boundaries are obtained using the modified model and the rotation.

MSC 2010: 76E06; 76E09; 76E25

1 Introduction

The last decade witnessed great interest in nanofluids due to their wide applications in science and engineering because of their thermal and mechanical properties. This type of fluid appears to increase the heat transfer performance significantly in comparison to ordinary fluids. Nanofluids are not obtained naturally but are synthesized in laboratories and consist of base fluids, such as water and organic solvents, in addition to metallic or metallic oxide nanoparticles with a maximum diameter of 100 nm.

Buongiorno [1] proposed a model for convective transport in nanofluids, which combines the effect of Brownian motion and thermophoresis. His model is used by many researchers to study the onset of thermal instability of a nanofluid layer. For example, Tzou [2,3] showed that the presence of nanoparticles in a base fluid could reduce the threshold of the Rayleigh-Benard instability, and Nield and Kuzentsov [4,5] studied the onset of convection in a horizontal nanofluid layer under several effects including the effect of porous medium. The onset of Rayleigh-Benard convection in nanofluids under the effect of rotation is studied by Yadav et al. [6]. Gupta et al. [7] studied the onset of Rayleigh-Benard convection in nanofluids under the effect of magnetic field. Shivakumara and Dhananjaya [8] studied the penetrative Brinkman convection in an anisotropic porous layer saturated by a nanofluid and Abdullah and Lindsay [9,10] investigated the onset of Marangoni convection in a layer of nanofluid.

The onset of Marangoni convection in a base fluid is induced by the dependence of the surface tension on temperature. Pearson [11] studied the stability of Marangoni convection of a layer of base fluid heated from below and showed that instability is caused by surface tension and not by the buoyancy forces. Nield [12] demonstrated that the effect of buoyancy forces can be neglected if the fluid layer depth is less than 1 mm. The results of Pearson [11] and Nield [12] have been expanded and refined by many researchers for deformable and non-deformable surfaces (e.g., Takashima [13,14], Benguria and Depassier [15], Wilson [16], Shivakumara et al. [17], Hashim and Arifin [18], and Shivakumara et al. and [19]).

Previous studies in the thermal stability of nanofluid layers assume that material properties such as effective thermal conductivity and effective viscosity behave as constant functions across the nanofluid layer. Abdullah et al. [20] studied the stability of Marangoni convection in a layer of nanofluid using a modification model of the nanoparticle mass flux in which the effects of Brownian diffusion and thermophoresis are taken into consideration. Also, they assumed that material properties of the nanofluid layer such as effective thermal conductivity and effective viscosity are modelled by non-constant constitutive expressions which depend on temperature and nanoparticle volume fraction.

The literature enriched also by the study of Brownian motions of a particle in a fluid (see e.g. [21,22,23]) revealed that no research has been conducted regarding the stability of Marangoni convection in a nanofluid layer under the effect of rotation. Therefore, the object of the present study is to investigate this problem for layers of distilled water (DW)/alumina and DW/cupric oxide nanofluids using a modified model with material properties which are modelled by non-constant constitutive expressions that depend on nanoparticle volume fraction.

The numerical calculations will adopt the models of Khanafar and Vafai [24] for the effective dynamic viscosity and effective thermal conductivity of these nanofluids. Other models will be adopted for comparison such as Hamilton and Crosser’s model [25] for the effective thermal conductivity and Brinkman’s model [26] for the effective dynamic viscosity.

The work is set out as follows. Section 2 formulates the problem, introduces the field equations that describe the model, constructs the steady state solution and produces the non-dimensional linearized equations and boundary conditions. Section 3 establishes the eigenvalue problem to be solved. Section 4 formulates the normal mode analysis. Section 5 describes the numerical procedure used to treat the eigenvalue problem. Section 6 presents results and compares stability boundaries based on the classical model of convection under the effect of rotation and the general model of convection under the effect of rotation in which constitutive functions are assumed to be non-constants. Section 7 concludes.

2 Field equations

Following the work of Tzou [2,3] and Nield and Kuznetsov [4,5], for the motion of a nanofluid in the absence of chemical reactions leads to the development of the following field equations:

(1) v j x j = 0 ,

(2) ρ eff ( φ ) v j t + v k v j x k = P x j + x k μ eff ( φ ) v j x k + v k x j + 2 ( v × Ω ) j ,

(3) φ t + v k φ x k = 1 ρ p J k x k ,

(4) ( ρ c ) eff ( φ ) T t + v k T x k = q k x k + h p ( T ) J k x k ,

where P = p + 1 2 ( Ω × v ) 2 is the sum of hydrostatic pressure and centrifugal forces, v j is the velocity, J k is the mass flux of nanoparticles within the nanofluid, q k is the heat flux, μ eff is the effective dynamic viscosity and the quantities c eff and ρ eff are the effective specific heats and density of nanofluid. The density and specific heat of the nanoparticle material, ( ρ p , c p ) , and the base fluid, ( ρ f , c f ) , are related to c eff and ρ eff , by the expressions

(5) ρ eff = φ ρ p + ( 1 φ ) ρ f ,

(6) ( ρ c ) eff = φ ρ p c p + ( 1 φ ) ρ f c f .

Equations (1)–(4) are affected by Buongiorno’s assumption, which resides in the specification of the nanoparticle mass flux with constitutive formula

(7) J k = ρ p C B T φ x k + φ C T T T x k ,

where the first component describes Brownian diffusion of nanoparticles and the second component describes the influence of thermophoresis. C B and C T are two parameters that have the constitutive specifications

(8) C B = k 3 π μ f d p , C T = μ f ρ f 0.26 κ f 2 κ f + κ p ,

where k is the Boltzman constant, d p is the nanoparticle diameter, μ f is the dynamic viscosity of the base fluid and κ f and κ p are, respectively, the thermal conductivities of the base fluid and the nanoparticle material. The heat flux vector q = q k e k in equation (4) has the constitutive expression:

(9) q k = κ eff ( φ ) T x k + h p ( T ) J k ,

where κ eff ( φ ) is the effective thermal conductivity of nanofluid and h p ( T ) is the specific enthalpy of nanoparticle material. Substituting this expression into the energy equation, (4), we get

(10) ( ρ c ) eff T t + v k T x k = x k ( κ eff ( φ ) T x k ) c p T x k J k ,

where c p = h p T .

2.1 Boundary conditions

According to the problem formulation, the nanofluid is assumed to rest on a rigid boundary, which is maintained at a constant temperature T 0 at x 3 = 0 , whereas on x 3 = d , the surface is non-deformable and loses heat by convection to an environment at temperature T ext . These conditions are embodied in the equations:

(11) T ( t ; x β , 0 ) = T 0 ,

(12) q 3 ( t ; x β , d ) = R ( T ( t ; x β , d ) T ext ) ,

where R is a constant that represents a heat transfer coefficient and β takes the values 1 and 2. Mechanical conditions require the fluid velocity on the boundary x 3 = 0 to be zero and taking into account the influence of surface tension; we assume the boundary x 3 = d to be flat and free of shear stress. These conditions are brought together into the conditions

(13) v k ( t ; x β , 0 ) = 0 ; k = ( 1 , 2 , 3 ) ,

(14) v 3 ( t ; x β , d ) = 0 ,

(15) σ 3 λ ( t ; x β , d ) = γ x β ,

where σ 31 and σ 32 are the shear components of nanofluid stress in the plane x 3 = d . At the boundary x 3 = d , the condition (14) can simplify the surface tension boundary condition (15) to

(16) μ eff ( φ ) v β ( t ; x β , d ) x 3 = γ T T ( t ; x β , d ) x β + γ φ φ ( t ; x β , d ) x β ,

where φ and T are evaluated on x 3 = d . The no slip condition on a rigid boundary implies that the horizontal components of the fluid velocity and all of x 1 , x 2 partial derivatives of each component of the fluid velocity vanish, i.e., v 1 = v 2 = 0 and from equation (1) it can be obtained that

(17) v 3 x 3 = 0 .

If we introduce ξ i = ε i j k v k , j to be the fluid vorticity, then

(18) ξ 3 ( t ; x β , 0 ) = 0 , ξ 3 ( t ; x β , d ) x 3 = 0 .

Finally, the confining boundaries x 3 = 0 and x 3 = d are assumed to be impervious to the flow of nanoparticles and so the appropriate boundary conditions require that fluxes of nanoparticles crossing both boundaries are zero. Consequently,

(19) J 3 ( t ; x β , 0 ) = J 3 ( t ; x β , d ) = 0 ,

subject to the constraint that

(20) lim A 1 A A 1 d 0 d φ ( t ; x β , x 3 ) d x 3 d x β = φ ¯ ,

where φ ¯ is the average volume fraction of nanoparticles across the layer and A is an arbitrary region of the ( x 1 , x 2 ) plane enclosing area A .

2.2 Models to be investigated

This problem will be investigated using two models:

Model I. It is the usual classical model in which the effective thermal conductivity and effective dynamic viscosity behave as parameters and the nanoparticle flux has the form:

(21) J k = ρ p C B T 0 φ x k + φ ¯ C T T 0 T x k .

Model II. In which the effective dynamic viscosity and effective thermal conductivity are non-constant functions of nanoparticle volume fraction, φ , and the nanoparticle mass flux has the form:

(22) J k = ρ p C B T φ x k + φ C T T T x k .

2.3 Steady state

Equations (1)–(3) and (10) have a steady state solution in which the nanofluid is at rest, the diffusion mass flux is zero, the temperature is T ( x 3 ) and the volume fraction of nanoparticles is Φ ( x 3 ) .

Model I. In this case, we can show that

(23) T ( x 3 ) = T 0 + Nu ( T ext T 0 ) Nu + ( κ eff / κ f ) x 3 d , Φ ( x 3 ) = Φ 0 + η φ ¯ T 0 Nu ( T 0 T ext ) Nu + ( κ eff / κ f ) x 3 d ,

where η = C T / ( C B T 0 ) , Φ 0 is the volume fraction of nanoparticle on the boundary x 3 = 0 with value determined from conservation of nanoparticles and the Nusselt number is Nu = R d / κ f . In this model, the temperature gradient and the gradient of the volume fraction of nanoparticles are constants and lead to the distributions described in equations (23). These equations show that both temperature and nanoparticle volume fraction are well described by linear functions of depth. Hence, in this model the steady-state solution is well approximated by a linear distribution of temperature and nanoparticle volume fraction.

Model II. The solution of temperature and the volume fraction of nanoparticles in the steady state assume that κ eff is a function of Φ . The steady state volume fraction Φ ( x 3 ) and temperature T ( x 3 ) satisfy the equations

(24) J 3 = ρ p C B T d Φ d x 3 + Φ C T T d T d x 3 = 0 , κ eff ( Φ ) d T d x 3 + R ( T d T ext ) = 0

with the boundary condition T ( 0 ) = T 0 , T d = T ( d ) along with the constraint condition

(25) 1 d 0 d Φ d x 3 = φ ¯ .

Using the non-dimensional spatial variable, x i = x i / d , we can show that

(26) κ eff ( Φ ) κ f d T d x 3 + Nu ( T d T ext ) = 0 ,

(27) Φ = Φ 0 exp η ( T 0 T ) T .

Equation (27) shows that the nanoparticle volume fraction is well approximated by an exponential distribution. Abdullah et al. [20] solved equations (26) and (27) numerically for T ( x 3 ) and Φ ( x 3 ) and proved that the linear temperature gradient across the layer is a robust assumption, but this is not the case for the nanoparticle volume fraction in this model.

2.4 Linearized equations

Equations (1)–(3) and (10) and their associated boundary conditions are linearized by introducing the perturbations

(28) v j = ε v ˆ j , P = P ( x 3 ) + ε P ˆ , T = T ( x 3 ) + ε T ˆ , φ = Φ ( x 3 ) + ε φ ˆ , J j = ε J ˆ j and ξ i = ε ξ ˆ i .

Thus, the linearized equations have the form:

(29) v ˆ j x j = 0 ,

(30) ρ eff ( Φ ) v ˆ j t = P ˆ x j + x k μ eff ( Φ ) v ˆ j x k + v ˆ k x j + 2 ( v ˆ × Ω ) j ,

(31) φ ˆ t + v ˆ 3 d Φ d x 3 = 1 ρ p J ˆ k x k ,

(32) ( ρ c ) eff ( Φ ) T ˆ t + v ˆ 3 d T d x 3 = x k κ eff ( Φ ) T ˆ x k + x 3 φ ˆ d κ eff ( Φ ) d Φ d T d x 3 c p d T d x 3 J ˆ 3 .

The associated linearized boundary conditions on x 3 = 0 are

(33) v ˆ 3 ( t ; x β , 0 ) = v ˆ 3 ( t ; x β , 0 ) x 3 = ξ ˆ 3 ( t ; x β , 0 ) = T ˆ ( t ; x β , 0 ) = J ˆ 3 ( t ; x β , 0 ) = 0 ,

and the associated linearized boundary conditions on x 3 = d are

(34) v ˆ 3 ( t ; x β , d ) = ξ ˆ 3 ( t ; x β , d ) x 3 = J ˆ 3 ( t ; x β , d ) = 0 ,

(35) v ˆ β ( t ; x β , d ) x 3 = 1 μ eff ( Φ ) γ T T ˆ ( t ; x β , d ) x β + γ Φ φ ˆ ( t ; x β , d ) x β ,

(36) κ eff ( Φ ) T ˆ ( t ; x β , d ) x 3 + d T d x 3 φ ˆ d κ eff ( Φ ) d Φ + R T ˆ ( t ; x β , d ) = 0 .

The linearized constrained condition requires that

(37) lim A 1 A A 1 d 0 d φ ˆ ( t ; x β , x 3 ) d x 3 d x β = 0 .

In Model I, the functions ρ eff ( Φ ) , μ eff ( Φ ) , ( ρ c ) eff ( Φ ) and κ eff ( Φ ) behave as parameters and the following expression in equations (32) and (36) vanishes, i.e.,

(38) φ ˆ d κ eff ( Φ ) d Φ d T d x 3 = 0 ,

whereas in Model II these functions are functions of x 3 .

2.5 Non-dimensional equations

Following standard procedures, equations (29)–(32) are non-dimensionalized using the scale length d and time t = d 2 α t and replacing v ˆ k , P ˆ , T ˆ , φ ˆ and J ˆ k by the following expressions:

(39) v ˆ k = α d v k , P ˆ = α μ f d 2 p , T ˆ = T d T 0 θ , φ ˆ = Φ d Φ 0 ψ , J ˆ k = 1 d ρ p C B T 0 Φ d Φ 0 J k ,

where α = κ f / ( ρ c ) f is the diffusivity. The final form of the non-dimensional linearized equations becomes:

(40) v j x j = 0 ,

(41) f 1 Pr v j t = p x j + x k f 2 v j x k + v k x j + T a ε j k i v k δ i 3 ,

(42) Le ψ t + v 3 Φ d Φ 0 d Φ d x 3 = J k x k ,

(43) f 3 θ t + v 3 T d T 0 d T d x 3 = x k f 4 θ x k + x 3 1 T d T 0 d T d x 3 f 5 θ N B Le 1 T d T 0 d T d x 3 J 3 ,

where f 1 , , f 5 are functions with expressions

(44) f 1 ( x 3 ) = ρ eff ( Φ ) ρ f , f 2 ( x 3 ) = μ eff ( Φ ) μ f , f 3 ( x 3 ) = ( ρ c ) eff ( Φ ) ( ρ c ) f , f 4 ( x 3 ) = κ eff ( Φ ) κ f , f 5 ( x 3 ) = Φ d Φ 0 Φ κ eff ( Φ ) κ f ,

where Pr = μ f c f κ f is the Prandtl number, T a = 4 Ω 2 d 4 μ f 2 is the Taylor number, Le = α C B T 0 is the Lewis number and N B = ( ρ c ) p ( ρ c ) f Φ d Φ 0 is the modified particle density increment. In Model II, the functions f 1 , , f 5 depend on x 3 whereas, in Model I these functions behave as parameters. The associated non-dimensional forms of the lower boundary conditions at x 3 = 0 are

(45) v 3 ( t ; x β , 0 ) = v 3 ( t ; x β , 0 ) x 3 = ξ 3 ( t ; x β , 0 ) = θ ( t ; x β , 0 ) = J 3 ( t ; x β , 0 ) = 0 ,

and on the upper boundary, at x 3 = 1 , the non-dimensional boundary conditions are

(46) v 3 ( t ; x β , 1 ) = ξ 3 ( t ; x β , 1 ) x 3 = J 3 ( t ; x β , 1 ) = 0 , f 2 ( 1 ) 2 v 3 ( t ; x β , 1 ) x 3 2 M T Δ 2 θ ( t ; x β , 1 ) M Φ Δ 2 ψ ( t ; x β , 1 ) = 0 , f 4 ( 1 ) θ ( t ; x β , 1 ) x 3 + Nu θ ( t ; x β , 1 ) + 1 T d T 0 d T d x 3 f 5 ( 1 ) ψ ( t ; x β , 1 ) = 0 ,

where Δ 2 is the Laplacian in two dimensions and M T and M Φ are, respectively, the temperature and nanoparticle Marangoni numbers which have the following definitions:

(47) M T = T d T 0 d α μ f γ T , M Φ = Φ d Φ 0 d α μ f γ Φ .

The non-dimensional forms of the nanoparticles mass flux J ˆ k with respect to Model I and Model II are as follows:

(48) Model I. J k = ψ x k + N A θ x k , Model II. J k = T T 0 ψ x k + N A T 0 T Φ φ ¯ θ x k + 2 N A η d d x 3 ( Φ φ ¯ ) θ + η T d T d x 3 ψ δ k 3 ,

where N A = η T d T 0 Φ d Φ 0 φ ¯ T 0 is the modified diffusivity ratio.

3 Eigenvalue problem

The eigenvalue problem is constructed by taking the curl operator of equation (41). Thus, the third component of the resulting equation has the form:

(49) f 1 Pr ξ t = f 2 Δ 3 ξ + T a d w d x 3 , in Model I.

(50) f 1 Pr ξ t = d f 2 d x 3 d ξ d x 3 + f 2 Δ 3 ξ + T a d w d x 3 , in Model II.

Taking the curl once again, the third component of the resulting equation has the form:

(51) f 1 Pr Δ 3 w t = f 2 Δ 3 2 ξ T a ξ x 3 , in Model I.

(52) 1 Pr f 1 Δ 3 w t + d f 1 d x 3 x 3 w t = f 2 Δ 3 2 w + 2 d f 2 d x 3 ( Δ 3 w ) x 3 + d 2 f 2 d x 3 2 2 w x 3 2 Δ 2 w T a ξ x 3 , in Model II.

in which w , ξ represent the third components of the velocity field and vorticity, respectively, and Δ 3 is the Laplacian in three dimensions. The third and fourth equations that complete the eigenvalue problem are derived from equations (42) and (43) where the expression of J k is determined by the choice of the model. Thus, in Model I, the third components of equations (42) and (43) have the form:

(53) Le ψ t + H v 3 = Δ 3 ψ + N A Δ 3 θ ,

(54) f 3 θ t H v 3 = f 4 Δ 3 θ H N B Le ψ x 3 + N A θ x 3 .

For Model II, the third components of equations (42) and (43) have the form:

(55) Le ψ t + H Φ φ ¯ v 3 = T T 0 Δ 3 ψ H T d T 0 T 0 1 + η T 0 T ψ x 3 η T d T 0 2 T 2 ψ + N A Φ φ ¯ T 0 T Δ 3 θ + N A H Φ φ ¯ T d T 0 T 0 2 + T 0 2 T 2 + η T 0 T θ x 3 + 2 η N A Φ φ ¯ T d T 0 2 T 0 2 θ ,

(56) f 3 θ t + v 3 T d T 0 d T d x 3 = x 3 f 4 θ x 3 H x 3 ψ d f 5 d Φ H N B Le T T 0 ψ x 3 + N A Φ φ ¯ T 0 T θ x 3 + 2 N A H T d T 0 T 0 Φ φ ¯ θ + η T d T d x 3 ψ .

4 Normal mode analysis

A normal mode analysis is used to investigate the stability of the boundary value problem of each model by writing the perturbed quantities in the form:

(57) χ ( x , t ) = χ ( x 3 ) exp ( σ t + i ( n x 1 + m x 2 ) ) ,

where χ = ( w , θ , ψ , ξ ) , n , m are wave numbers along e 1 and e 2 directions, respectively, and σ is the growth rate. Thus, the final forms of the eigenvalue problem in each model are written as follows:

Model I.

(58) σ Pr f 1 ξ = f 2 ( D 2 a 2 ) ξ + T a D w ,

(59) σ Pr f 1 ( D 2 a 2 ) w = f 2 ( D 2 a 2 ) 2 w T a D ξ ,

(60) Le σ ψ = Le H w + ( D 2 a 2 ) ψ + N A ( D 2 a 2 ) θ ,

(61) f 3 σ θ = f 3 H w + f 4 ( D 2 a 2 ) θ H N B Le ( D ψ + N A D θ ) .

The corresponding boundary conditions on x 3 = 0 have the form:

(62) w ( 0 ) = D w ( 0 ) = ξ ( 0 ) = θ ( 0 ) = D ψ ( 0 ) + N A D θ ( 0 ) = 0 ,

and on x 3 = 1

(63) θ ( 1 ) = w ( 1 ) = D ξ ( 1 ) = D ψ ( 1 ) + N A D θ ( 1 ) = 0 , f 2 ( 1 ) D 2 w ( 1 ) + a 2 M T θ ( 1 ) + a 2 M Φ ψ ( 1 ) = 0 , f 4 ( 1 ) D θ ( 1 ) + Nu θ ( 1 ) = 0 .

Model II.

(64) σ Pr f 1 ξ = D f 2 D ξ + f 2 ( D 2 a 2 ) ξ + T a D w ,

(65) σ Pr ( f 1 ( D 2 a 2 ) w + D f 1 D w ) = f 2 ( D 2 a 2 ) 2 w + 2 D f 2 ( D 2 a 2 ) D w + D 2 f 2 ( D 2 + a 2 ) w T a D ξ ,

(66) Le σ ψ = Le H Φ φ ¯ w + T T 0 ( D 2 a 2 ) ψ H T d T 0 T 0 1 + η T 0 T D ψ η T d T 0 2 T 2 ψ + N A Φ φ ¯ T 0 T ( D 2 a 2 ) θ + N A H Φ φ ¯ T d T 0 T 0 2 + T 0 2 T 2 + η T 0 T D θ + 2 N A η Φ φ ¯ T d T 0 2 T 0 2 θ ,

(67) f 3 σ θ = f 3 w T d T 0 d T d x 3 + f 4 ( D 2 a 2 ) θ + D f 4 D θ H D ψ d f 5 d Φ H N B Le T T 0 D ψ + N A T 0 T Φ φ ¯ D θ + 2 N A H T d T 0 T 0 Φ φ ¯ θ + η T d T d x 3 ψ .

The associated boundary conditions at x 3 = 0 are

(68) w ( 0 ) = D w ( 0 ) = ξ ( 0 ) = θ ( 0 ) = 0 , D ψ ( 0 ) + N A Φ 0 φ ¯ D θ ( 0 ) + η T 0 d T d x 3 ψ ( 0 ) = 0 .

While on x 3 = 1 , the boundary conditions have the form:

(69) w ( 1 ) = D ξ ( 1 ) = 0 , f 2 ( 1 ) D 2 w ( 1 ) + a 2 M T θ ( 1 ) + a 2 M Φ ψ ( 1 ) = 0 , f 4 ( 1 ) D θ ( 1 ) + Nu θ ( 1 ) H f 5 ( 1 ) ψ ( 1 ) = 0 , D ψ ( 1 ) + N A T 0 T d 2 Φ d φ ¯ D θ ( 1 ) + η T 0 T d 2 d T ( 1 ) d x 3 ψ ( 1 ) + 2 N A η T 0 T d d d x 3 Φ φ ¯ θ ( 1 ) = 0 ,

where D = d d x 3 , and a = n 2 + m 2 is the wave number.

5 Numerical method

Equations (58)–(61) of Model I together with the corresponding conditions (62) and (63) and equations (64)–(67) of Model II with the corresponding conditions (68) and (69) are solved numerically using the method of expansion of Chebyshev polynomials when the nanofluid layer is heated from below, i.e., H = 1 . This method is known by its high accuracy and it allows stationary and over-stable modes to be treated simultaneously [27]. Let us define the variables y 1 ( z ) , , y 10 ( z ) by the expressions

(70) y 1 = w , y 2 = D w , y 3 = ( D 2 a 2 ) w , y 4 = ( D 2 a 2 ) D w , y 5 = ψ , y 6 = D ψ , y 7 = θ , y 8 = D θ , y 9 = ξ , y 10 = D ξ ,

where the functions y 1 ( z ) , , y 8 ( z ) are assigned the Chebyshev polynomial expansions of the form:

(71) y r = k = 0 N a k r T k ( z ) , 1 r 10 ,

where N is a user-specified even integer and T k ( z ) are the Chebyshev polynomials of the first kind. The system of equations of Model I and Model II leads to the eigenvalue problem A Y = σ B Y , where B is a singular matrix. The eigenvalues, σ , are calculated using a special routine. In Model I, the numerical solutions are obtained quite easily; however, in Model II this is not the case. In fact, the complexity in this case stems from the inhomogeneity of the eigenvalue problem, namely the multiplication of the non-constant functions f 1 ( z ) , , f 5 ( z ) with the variables of the problem.

6 Results and discussion

Results are obtained for two types of nanofluids, DW/alumina and DW/cupric oxide for nanoparticles of size 30, 70 and 100 nm. Calculations will take advantage of the work of Khanafer and Vafai [24] who have proposed the following constitutive model for the effective thermal conductivity

(72) κ eff κ f = 1.0 + φ 1.0112 + 114.5625 d p + 0.040457 κ p ,

for these nanofluids at working temperatures close to room temperature ( 298 ° K ). In expression (72), the volume fraction satisfies ϕ [ 0.01 , 0.10 ] and the nanoparticle diameter (expressed in nanometers) satisfies d p [ 11 , 150 ] . Also, the following effective dynamic viscosity expressions proposed by Khanafer and Vafai [24] will be used in the calculations

(73) Distilled-water Alumina μ eff = a 0 + a 1 φ + a 2 φ 2 + a 3 φ 3 + a 4 T + a 5 φ 2 T 2 + a 6 φ T 3 + a 7 φ 2 d p 2 + a 8 φ 3 d p 2 ,

(74) Distilled-water Cupric oxide μ eff = b 0 + b 1 φ + b 2 φ 2 + b 3 φ 3 + b 4 T + b 5 T 2 + b 6 φ T + b 7 φ 2 T 2 + b 8 φ T 3 .

In these expressions φ [ 0.01 , 0.09 ] , Celsius temperature T [ 20 , 70 ] and in the case of the DW/alumina nanofluid, the nanoparticle diameter (expressed in nanometers) satisfies d p [ 13 , 133 ] . For comparison with the results obtained using the models proposed by Khanafer and Vafai [24], for the effective thermal conductivity and effective dynamic viscosity, another model proposed by Hamilton and Crosser [25] is used for the effective thermal conductivity of nanofluid, namely

(75) κ ( φ ) κ f = ( n 1 ) ( 1 φ ) κ f + ( n φ + 1 φ ) κ p κ p ( 1 φ ) + ( n 1 + φ ) κ f ,

and Brinkman’s model [26] for the effective dynamic viscosity is used, namely

(76) μ ( φ ¯ ) = μ f ( 1 φ ¯ ) 5 / 2 .

The investigation of the stability boundaries is obtained based on the classical model of convection of a layer of nanofluid under the effect of rotation and compared with the general model of convection of a layer of nanofluid under the effect of rotation where the constitutive functions are assumed to be non-constants.

The main condition that must be satisfied is that all eigenvalues of the linear problem have negative real parts. Hence, the location of the stability boundary is determined by comparing the values of the true temperature Marangoni number, M T , with its value on the stability boundary when the true value of the nanoparticle Marangoni number, M Φ , is known. If the true value of the temperature Marangoni number exceeds the value on the stability boundary, then the layer is unstable, otherwise it is stable. The true values of the temperature Marangoni number and the nanoparticle Marangoni number are calculated using equations (47).

The results are obtained for a layer of 1 mm thickness, nanoparticles of size d p = 30 nm and Nu = 0.2 . The average nanoparticle volume fractions are ranging from 0.01 to 0.03 for DW/alumina nanofluid and from 0.01 to 0.06 for DW/cupric oxide nanofluid. The rotational effects are considered for T a = 0 , 100, 1000. The lower boundary of the layer has a constant temperature of 30 0 ° K and the atmospheric temperature at the upper boundary is 29 8 ° K .

The true and critical temperature Marangoni numbers for the layer of DW/alumina nanofluid, are evaluated for both Model I, Figure 1a, and Model II, Figure 1b, with different values of T a using the thermal conductivity and dynamic viscosity constitutive functions proposed by Brinkman-Hamilton-Crosser. The graphs of Model I in Figure 1a indicate that the layer of nanofluid is unstable for all values of φ ¯ [ 0.01 , 0.03 ] and for all values of T a , whereas the graphs of Model II in Figure 1b indicate that the nanofluid layer is unstable for all values of φ ¯ when T a = 0 , 100 . However when T a = 1000 , the nanofluid layer is stable when φ ¯ 0.005 . Moreover, the graphs in Figure 1 indicate that the critical Marangoni number increases as the Taylor number increases showing that the rotation has a stabilizing effect. Also the graphs show that the critical Marangoni number in Model I are lower than the critical Marangoni number in Model II, which indicates that Model II is more stable than Model I. This result clarifies that the assumption of non-constant thermal conductivity and dynamic viscosity functions across the layer stabilizes the DW/alumina nanofluid layer.

Figure 1 
               The true and critical temperature Marangoni numbers for DW/alumina nanofluid with 
                     
                        
                        
                           
                              
                                 d
                              
                              
                                 p
                              
                           
                           =
                           30
                           
                           nm
                        
                        {d}_{p}=30\hspace{0.33em}{\rm{nm}}
                     
                  , 
                     
                        
                        
                           Δ
                           T
                           =
                           
                              
                                 2
                              
                              
                                 
                                    ∘
                                 
                              
                           
                           K
                        
                        \Delta T={2}^{\circ }{\rm{K}}
                     
                   for average volume fractions ranging from 0.01 to 0.03. Stability boundary curves for 
                     
                        
                        
                           
                              
                                 M
                              
                              
                                 T
                              
                           
                        
                        {M}_{T}
                     
                   are calculated based on (a) Model I and (b) Model II, using the thermal conductivity and dynamic viscosity constitutive functions proposed by Brinkman-Hamilton-Crosser.
Figure 1

The true and critical temperature Marangoni numbers for DW/alumina nanofluid with d p = 30 nm , Δ T = 2 K for average volume fractions ranging from 0.01 to 0.03. Stability boundary curves for M T are calculated based on (a) Model I and (b) Model II, using the thermal conductivity and dynamic viscosity constitutive functions proposed by Brinkman-Hamilton-Crosser.

Figure 2 illustrates the results of the true and critical temperature Marangoni numbers for DW/alumina nanofluid using the thermal conductivity and dynamic viscosity constitutive functions calculated by Khanafer and Vafai. Stability boundaries based on Model I in Figure 2a show that the layer is unstable for all values of φ ¯ [ 0.01 , 0.03 ] . In contrast, the layer in Model II is stable for all values of the Taylor number, T a , when φ ¯ exceeds specific values. In fact, when T a = 0 , 100 , and 1000, the specific values of φ ¯ at which stability ensue are 0.02 , 0.007 and 0.0025, respectively, as shown in Figure 2b. The aforementioned results of Model II clarify that the assumption of non-constant thermal conductivity and dynamic viscosity functions across the layer stabilizes the DW/alumina nanofluid layer. This stability increases as the Taylor number increases showing that rotation has a stabilizing effect.

Figure 2 
               The true and critical temperature Marangoni numbers for DW/alumina nanofluid with 
                     
                        
                        
                           
                              
                                 d
                              
                              
                                 p
                              
                           
                           =
                           30
                           
                           nm
                        
                        {d}_{p}=30\hspace{0.33em}{\rm{nm}}
                     
                  , 
                     
                        
                        
                           Δ
                           T
                           =
                           
                              
                                 2
                              
                              
                                 
                                    ∘
                                 
                              
                           
                           K
                        
                        \Delta T={2}^{\circ }{\rm{K}}
                     
                   for average volume fractions ranging from 0.01 to 0.03. Stability boundary curves for the temperature Marangoni number are calculated based on (a) Model I and (b) Model II, using the thermal conductivity and dynamic viscosity constitutive functions proposed by Khanafer and Vafai.
Figure 2

The true and critical temperature Marangoni numbers for DW/alumina nanofluid with d p = 30 nm , Δ T = 2 K for average volume fractions ranging from 0.01 to 0.03. Stability boundary curves for the temperature Marangoni number are calculated based on (a) Model I and (b) Model II, using the thermal conductivity and dynamic viscosity constitutive functions proposed by Khanafer and Vafai.

The evaluations of the true and critical temperature Marangoni numbers for the layer of DW/cupric oxide are plotted in Figure 3 using the thermal conductivity and dynamic viscosity constitutive functions proposed by Brinkman-Hamilton-Crosser and in Figure 4 using the thermal conductivity and dynamic viscosity constitutive functions proposed by Khanafer and Vafai. The graphs of Model I and Model II in these figures show that the nanofluid layer is stable for all values of T a when φ ¯ exceeds specific values. A significant result is seen in Figure 3b when T a = 1000 in which the nanofluid layer is stable for all values of φ ¯ [ 0.01 , 0.06 ] . As in Figure 1a and b, rotation and the assumption of non-constant thermal conductivity and dynamic viscosity functions across the layer stabilize the nanofluid layer.

Figure 3 
               The true and critical temperature Marangoni numbers for DW/cupric oxide nanofluid with 
                     
                        
                        
                           
                              
                                 d
                              
                              
                                 p
                              
                           
                           =
                           30
                           
                           nm
                        
                        {d}_{p}=30\hspace{0.33em}{\rm{nm}}
                     
                  , 
                     
                        
                        
                           Δ
                           T
                           =
                           
                              
                                 2
                              
                              
                                 
                                    
                                       °
                                    
                                 
                              
                           
                           K
                        
                        \Delta T={2}^{<mml:mpadded xmlns:ali="http://www.niso.org/schemas/ali/1.0/"
                                        xmlns:xsi="http://www.w3.org/2001/XMLSchema-instance">^\circ </mml:mpadded>}{\rm{K}}
                     
                   for average volume fractions ranging from 0.01 to 0.06. Stability boundary curves for the temperature Marangoni number are calculated based on (a) Model I and (b) Model II, using thermal conductivity and dynamic viscosity constitutive functions proposed by Brinkman-Hamilton-Crosser.
Figure 3

The true and critical temperature Marangoni numbers for DW/cupric oxide nanofluid with d p = 30 nm , Δ T = 2 ° K for average volume fractions ranging from 0.01 to 0.06. Stability boundary curves for the temperature Marangoni number are calculated based on (a) Model I and (b) Model II, using thermal conductivity and dynamic viscosity constitutive functions proposed by Brinkman-Hamilton-Crosser.

Figure 4 
               The true and critical temperature Marangoni numbers for DW/cupric oxide nanofluid with 
                     
                        
                        
                           
                              
                                 d
                              
                              
                                 p
                              
                           
                           =
                           30
                           
                           nm
                        
                        {d}_{p}=30\hspace{0.33em}{\rm{nm}}
                     
                  , 
                     
                        
                        
                           Δ
                           T
                           =
                           
                              
                                 2
                              
                              
                                 
                                    
                                       °
                                    
                                 
                              
                           
                           K
                        
                        \Delta T={2}^{<mml:mpadded xmlns:ali="http://www.niso.org/schemas/ali/1.0/"
                                        xmlns:xsi="http://www.w3.org/2001/XMLSchema-instance">^\circ </mml:mpadded>}{\rm{K}}
                     
                   for average volume fractions ranging from 0.01 to 0.06. Stability boundary curves for the temperature Marangoni number are calculated based on (a) Model I and (b) Model II, using the thermal conductivity and dynamic viscosity constitutive functions proposed by Khanafer and Vafai.
Figure 4

The true and critical temperature Marangoni numbers for DW/cupric oxide nanofluid with d p = 30 nm , Δ T = 2 ° K for average volume fractions ranging from 0.01 to 0.06. Stability boundary curves for the temperature Marangoni number are calculated based on (a) Model I and (b) Model II, using the thermal conductivity and dynamic viscosity constitutive functions proposed by Khanafer and Vafai.

Figure 5 represents a general comparison between the stability boundaries calculated based on Model I (Figure 5a) and those based on Model II (Figure 5b). For Models I and II, the figure shows that for the DW/alumina nanofluid the model of Khanafer and Vafai makes the nanofluid more stable compared to the model of Brinkman-Hamilton-Crosser. Moreover, for the DW/cupric oxide nanofluid the model of Brinkman-Hamilton-Crosser makes the nanofluid more stable compared to the model of Khanafer and Vafai if φ ¯ within a specific range.

Figure 5 
               The stability boundaries of critical temperature Marangoni numbers for DW/alumina and DW/cupric oxide nanofluids for 
                     
                        
                        
                           
                              
                                 T
                              
                              
                                 a
                              
                           
                           =
                           
                              
                              1000
                              
                           
                        
                        {T}_{a}=\hspace{0.1em}\text{1000}\hspace{0.1em}
                     
                  , using the thermal conductivity and dynamic viscosity constitutive functions proposed by Brinkman-Hamilton-Crosser and Khanafer and Vafai based on (a) Model I and (b) Model II.
Figure 5

The stability boundaries of critical temperature Marangoni numbers for DW/alumina and DW/cupric oxide nanofluids for T a = 1000 , using the thermal conductivity and dynamic viscosity constitutive functions proposed by Brinkman-Hamilton-Crosser and Khanafer and Vafai based on (a) Model I and (b) Model II.

Finally, Figure 6 shows a comparison between Model I and Model II by calculating the critical temperature Marangoni number, M T c r i t . , for a range of values of the average volume function of nanoparticles, φ ¯ , when T a = 1000. Clearly, Model II is more stable than Model I for both types of nanofluids (DW/cupric oxide and DW/alumina) and for both models of Brinkman-Hamilton-Crosser (Figure 6a) and Khanafer and Vafai (Figure 6b).

Figure 6 
               The stability boundaries of critical temperature Marangoni numbers for DW/alumina and DW/cupric oxide nanofluids based on Model I and Model II with 
                     
                        
                        
                           
                              
                                 T
                              
                              
                                 a
                              
                           
                           =
                           
                              
                              1,000,
                              
                           
                        
                        {T}_{a}=\hspace{0.1em}\text{1,000,}\hspace{0.1em}
                     
                   using the models proposed by (a) Brinkman-Hamilton-Crosser and (b) Khanafer and Vafai.
Figure 6

The stability boundaries of critical temperature Marangoni numbers for DW/alumina and DW/cupric oxide nanofluids based on Model I and Model II with T a = 1,000, using the models proposed by (a) Brinkman-Hamilton-Crosser and (b) Khanafer and Vafai.

7 Conclusions

This work investigated the linear stability of the onset of Marangoni convection for horizontal layers of DW/alumina and DW/cupric oxide nanofluids rotated about the x 3 -axis, heated from below and losing heat from their upper surfaces by convection to a region at constant temperature. The lower boundaries of the layers are assumed to be rigid with fixed temperature and impermeable to nanoparticles while the top boundaries are assumed to be non-deformable free surfaces at which surface tension variations due to nanoparticle volume fraction and temperature are allowed.

The stability of these nanofluids is investigated using two models. These models are distinguished by the condition that Model I assumes that constitutive properties are uniform across the layer of nanofluid, whereas in Model II the same properties are allowed to depend on local volume fraction of nanoparticles and satisfy constitutive expressions developed by Hamilton-Crosser-Brinkman and Khanafer and Vafai based on experimental evidence. Moreover, in Model I the classical expression of the nanoparticle mass flux is used in which the temperature and nanoparticle volume fraction are constants, whereas in Model II the temperature and nanoparticle volume fraction are assumed to be non-constants. Therefore, these changes contributed extra terms to the linear stability analysis for Model II that are not present in the same analysis based on Model I. The presence of these extra terms largely increases the stability boundary in Model II for both types of nanofluids and for both types of constitutive expressions developed by Hamilton-Crosser-Brinkman and Khanafer and Vafai. Findings also indicate that the effect of rotation plays a significant role in increasing the stability boundary in Model I and Model II for both types of nanofluids and for both models developed by Hamilton-Crosser-Brinkman and Khanafer and Vafai.

The numerical results were obtained using spectral eigenfunction expansions in terms of Chebyshev polynomials. Different stability boundaries were obtained and several comparisons were made for Model I and Model II to clarify which of the two nanofluids under consideration are more stable depending on the values of the nanoparticle volume fraction and on the type of the constitutive expressions used.

  1. Conflict of interest: Authors state no conflict of interest.

References

[1] J. Buongiorno , Convective transport in nanofluids, J. Heat Transf. 128 (2006), no. 3, 240–250, https://doi.org/10.1115/1.2150834 . 10.1115/1.2150834Search in Google Scholar

[2] D. Y. Tzou , Instability of nanofluids in natural convection, J. Heat Transf. 130 (2008), no. 7, https://doi.org/10.1115/1.2908427 . 10.1115/1.2908427Search in Google Scholar

[3] D. Y. Tzou , Thermal instability of nanofluids in natural convection, Int. J. Heat Mass Transf. 51 (2008), no. 11–12, 2967–2979, https://doi.org/10.1016/j.ijheatmasstransfer.2007.09.014. Search in Google Scholar

[4] D. A. Nield and A. V. Kuznetsov , Thermal instability in a porous medium layer saturated by a nanofluid, Int. J. Heat Mass Transf. 52 (2009), no. 25–26, 5796–5801, https://doi.org/10.1016/j.ijheatmasstransfer.2009.07.023. Search in Google Scholar

[5] D. A. Nield and A. V. Kuznetsov , The onset of convection in a horizontal nanofluid layer of finite depth, Eur. J. Mech. B/Fluid 29 (2010), no. 3, 217–223, https://doi.org/10.1016/j.euromechflu.2010.02.003. Search in Google Scholar

[6] D. Yadav , G. Agrawal , and R. Bhargava , Thermal instability of rotating nanofluid layer, Int. J. Eng. Sci. 49 (2011), no. 11, 1171–1184, https://doi.org/10.1016/j.ijengsci.2011.07.002. Search in Google Scholar

[7] U. Gupta , J. Ahuja , and R. Wanchoo , Magneto convection in a nanofluid layer, Int. J. Heat Mass Transf. 64 (2013), 1163–1171, https://doi.org/10.1016/j.ijheatmasstransfer.2013.05.035. Search in Google Scholar

[8] I. Shivakumara and M. Dhananjaya , Penetrative Brinkman convection in an anisotropic porous layer saturated by a nanofluid, Ain Shams Eng. J. 6 (2015), no. 2, 703–713, https://doi.org/10.1016/j.asej.2014.12.005. Search in Google Scholar

[9] A. A. Abdullah and K. Lindsay , Marangoni convection in a thin layer of nanofluid: Application to combinations of water or ethanol with nanoparticles of alumina or multi-walled carbon nanotubules, Int. J. Heat Mass Transf. 104 (2017), 693–702, https://doi.org/10.1016/j.ijheatmasstransfer.2016.08.099. Search in Google Scholar

[10] A. A. Abdullah , S. Althobaiti , and K. Lindsay , Marangoni convection in water-alumina nanofluids: Dependence on the nanoparticle size, Eur. J. Mech. B/Fluid 67 (2018), 259–268, https://doi.org/10.1016/j.euromechflu.2017.09.015. Search in Google Scholar

[11] J. Pearson , On convection cells induced by surface tension, J. Fluid Mech. 4 (1958), no. 5, 489–500, https://doi.org/10.1017/S0022112058000616 . 10.1017/S0022112058000616Search in Google Scholar

[12] D. Nield , Surface tension and buoyancy effects in cellular convection, J. Fluid Mech. 19 (1964), no. 3, 341–352, https://doi.org/10.1017/S0022112064000763. Search in Google Scholar

[13] M. Takashima , Surface tension driven instability in a horizontal liquid layer with a deformable free surface. I. Stationary convection, J. Phys. Soc. Jpn. 50 (1981), no. 8, 2745–2750, https://doi.org/10.1143/JPSJ.50.2745. Search in Google Scholar

[14] M. Takashima , Surface tension driven instability in a horizontal liquid layer with a deformable free surface. II. Overstability, J. Phys. Soc. Jpn. 50 (1981), no. 8, 2751–2756, https://doi.org/10.1143/JPSJ.50.2751. Search in Google Scholar

[15] R. Benguria and M. Depassier , On the linear stability theory of Bénard-Marangoni convection, Phys. Fluid Fluid Dynam. 1 (1989), no. 7, 1123–1127, https://doi.org/10.1063/1.857336. Search in Google Scholar

[16] S. Wilson , The effect of a uniform magnetic field on the onset of steady Bénard-Marangoni convection in a layer of conducting fluid, J. Eng. Math. 27 (1993), no. 2, 161–188, https://doi.org/10.1063/1.868417. Search in Google Scholar

[17] I. Shivakumara , M. Venkatachalappa , and S. Suma , Exact analysis of Marangoni convection with throughflow, Acta Mech. 136 (1999), no. 1–2, 109–117, https://doi.org/10.1007/BF01292301. Search in Google Scholar

[18] I. Hashim and N. Arifin , Oscillatory Marangoni convection in a conducting fluid layer with a deformable free surface in the presence of a vertical magnetic field, Acta Mech. 164 (2003), no. 3–4, 199–215, https://doi.org/10.1007/s00707-003-0008-7. Search in Google Scholar

[19] I. Shivakumara , C. Nanjundappa , and K. Chavaraddi , Darcy-Benard-Marangoni convection in porous media, Acta Mech. 52 (2009), no. 11–12, 2815–2823, https://doi.org/10.1016/j.ijheatmasstransfer.2008.09.038. Search in Google Scholar

[20] A. A. Abdullah , N. Alraiqib , and K. Lindsay , Modelling the stability of Marangoni convection in a layer of nanofluid, Int. J. Therm. Sci. 151 (2020), 106228, https://doi.org/10.1016/j.ijthermalsci.2019.106228. Search in Google Scholar

[21] H. M. Ahmed , M. M. El-Borai , A. S. O. El Bab , and M. E. Ramadan , Approximate controllability of noninstantaneous impulsive Hilfer fractional integrodifferential equations with fractional Brownian motion, Bound. Value Probl. 2020 (2020), 120, https://doi.org/10.1186/s13661-020-01418-0. Search in Google Scholar

[22] P. Duan , Existence and exponential stability of almost pseudo automorphic solution for neutral stochastic evolution equations driven by G-Brownian motion, Filomat 34 (2020), no. 4, 1075–1092, https://doi.org/10.2298/FIL2004075D . 10.2298/FIL2004075DSearch in Google Scholar

[23] S. Polidoro and M. A. Ragusa , Sobolev Morrey spaces related to an ultraparabolic equation, Manuscripta Math. 96 (1998), no. 3, 371–392, https://doi.org/10.1007/s002290050072. Search in Google Scholar

[24] K. Khanafer and K. Vafai , A critical synthesis of thermophysical characteristics of nanofluids, Int. J. Heat Mass Transf. 54 (2011), no. 19–20, 4410–4428, https://doi.org/10.1016/j.ijheatmasstransfer.2011.04.048. Search in Google Scholar

[25] R. Hamilton and O. K. Crosser , Thermal conductivity of heterogeneous two-component systems, Ind. Eng. Chem. Fundam. 1 (1962), no. 3, 187–191, https://doi.org/10.1021/i160003a005. Search in Google Scholar

[26] H. G. Brinkman , The viscosity of concentrated suspensions and solutions, J. Chem. Phys. 20 (1952), no. 4, 571–571, https://doi.org/10.1063/1.1700493. Search in Google Scholar

[27] C. Canuto , M. Hussaini , A. Quarteroni , and A. Thomas, Jr. , Spectral Methods in Fluid Dynamics, Springer Science & Business Media, Berlin Heidelberg, 2012. Search in Google Scholar

Received: 2021-04-22
Revised: 2021-06-13
Accepted: 2021-07-07
Published Online: 2021-09-08

© 2021 Abeer H. Bakhsh and Abdullah A. Abdullah, published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Regular Articles
  2. Sharp conditions for the convergence of greedy expansions with prescribed coefficients
  3. Range-kernel weak orthogonality of some elementary operators
  4. Stability analysis for Selkov-Schnakenberg reaction-diffusion system
  5. On non-normal cyclic subgroups of prime order or order 4 of finite groups
  6. Some results on semigroups of transformations with restricted range
  7. Quasi-ideal Ehresmann transversals: The spined product structure
  8. On the regulator problem for linear systems over rings and algebras
  9. Solvability of the abstract evolution equations in Ls-spaces with critical temporal weights
  10. Resolving resolution dimensions in triangulated categories
  11. Entire functions that share two pairs of small functions
  12. On stochastic inverse problem of construction of stable program motion
  13. Pentagonal quasigroups, their translatability and parastrophes
  14. Counting certain quadratic partitions of zero modulo a prime number
  15. Global attractors for a class of semilinear degenerate parabolic equations
  16. A new implicit symmetric method of sixth algebraic order with vanished phase-lag and its first derivative for solving Schrödinger's equation
  17. On sub-class sizes of mutually permutable products
  18. Asymptotic solution of the Cauchy problem for the singularly perturbed partial integro-differential equation with rapidly oscillating coefficients and with rapidly oscillating heterogeneity
  19. Existence and asymptotical behavior of solutions for a quasilinear Choquard equation with singularity
  20. On kernels by rainbow paths in arc-coloured digraphs
  21. Fully degenerate Bell polynomials associated with degenerate Poisson random variables
  22. Multiple solutions and ground state solutions for a class of generalized Kadomtsev-Petviashvili equation
  23. A note on maximal operators related to Laplace-Bessel differential operators on variable exponent Lebesgue spaces
  24. Weak and strong estimates for linear and multilinear fractional Hausdorff operators on the Heisenberg group
  25. Partial sums and inclusion relations for analytic functions involving (p, q)-differential operator
  26. Hodge-Deligne polynomials of character varieties of free abelian groups
  27. Diophantine approximation with one prime, two squares of primes and one kth power of a prime
  28. The equivalent parameter conditions for constructing multiple integral half-discrete Hilbert-type inequalities with a class of nonhomogeneous kernels and their applications
  29. Boundedness of vector-valued sublinear operators on weighted Herz-Morrey spaces with variable exponents
  30. On some new quantum midpoint-type inequalities for twice quantum differentiable convex functions
  31. Quantum Ostrowski-type inequalities for twice quantum differentiable functions in quantum calculus
  32. Asymptotic measure-expansiveness for generic diffeomorphisms
  33. Infinitesimals via Cauchy sequences: Refining the classical equivalence
  34. The (1, 2)-step competition graph of a hypertournament
  35. Properties of multiplication operators on the space of functions of bounded φ-variation
  36. Disproving a conjecture of Thornton on Bohemian matrices
  37. Some estimates for the commutators of multilinear maximal function on Morrey-type space
  38. Inviscid, zero Froude number limit of the viscous shallow water system
  39. Inequalities between height and deviation of polynomials
  40. New criteria-based ℋ-tensors for identifying the positive definiteness of multivariate homogeneous forms
  41. Determinantal inequalities of Hua-Marcus-Zhang type for quaternion matrices
  42. On a new generalization of some Hilbert-type inequalities
  43. On split quaternion equivalents for Quaternaccis, shortly Split Quaternaccis
  44. On split regular BiHom-Poisson color algebras
  45. Asymptotic stability of the time-changed stochastic delay differential equations with Markovian switching
  46. The mixed metric dimension of flower snarks and wheels
  47. Oscillatory bifurcation problems for ODEs with logarithmic nonlinearity
  48. The B-topology on S-doubly quasicontinuous posets
  49. Hyers-Ulam stability of isometries on bounded domains
  50. Inhomogeneous conformable abstract Cauchy problem
  51. Path homology theory of edge-colored graphs
  52. Refinements of quantum Hermite-Hadamard-type inequalities
  53. Symmetric graphs of valency seven and their basic normal quotient graphs
  54. Mean oscillation and boundedness of multilinear operator related to multiplier operator
  55. Numerical methods for time-fractional convection-diffusion problems with high-order accuracy
  56. Several explicit formulas for (degenerate) Narumi and Cauchy polynomials and numbers
  57. Finite groups whose intersection power graphs are toroidal and projective-planar
  58. On primitive solutions of the Diophantine equation x2 + y2 = M
  59. A note on polyexponential and unipoly Bernoulli polynomials of the second kind
  60. On the type 2 poly-Bernoulli polynomials associated with umbral calculus
  61. Some estimates for commutators of Littlewood-Paley g-functions
  62. Construction of a family of non-stationary combined ternary subdivision schemes reproducing exponential polynomials
  63. On the evolutionary bifurcation curves for the one-dimensional prescribed mean curvature equation with logistic type
  64. On intersections of two non-incident subgroups of finite p-groups
  65. Global existence and boundedness in a two-species chemotaxis system with nonlinear diffusion
  66. Finite groups with 4p2q elements of maximal order
  67. Positive solutions of a discrete nonlinear third-order three-point eigenvalue problem with sign-changing Green's function
  68. Power moments of automorphic L-functions related to Maass forms for SL3(ℤ)
  69. Entire solutions for several general quadratic trinomial differential difference equations
  70. Strong consistency of regression function estimator with martingale difference errors
  71. Fractional Hermite-Hadamard-type inequalities for interval-valued co-ordinated convex functions
  72. Montgomery identity and Ostrowski-type inequalities via quantum calculus
  73. Universal inequalities of the poly-drifting Laplacian on smooth metric measure spaces
  74. On reducible non-Weierstrass semigroups
  75. so-metrizable spaces and images of metric spaces
  76. Some new parameterized inequalities for co-ordinated convex functions involving generalized fractional integrals
  77. The concept of cone b-Banach space and fixed point theorems
  78. Complete consistency for the estimator of nonparametric regression model based on m-END errors
  79. A posteriori error estimates based on superconvergence of FEM for fractional evolution equations
  80. Solution of integral equations via coupled fixed point theorems in 𝔉-complete metric spaces
  81. Symmetric pairs and pseudosymmetry of Θ-Yetter-Drinfeld categories for Hom-Hopf algebras
  82. A new characterization of the automorphism groups of Mathieu groups
  83. The role of w-tilting modules in relative Gorenstein (co)homology
  84. Primitive and decomposable elements in homology of ΩΣℂP
  85. The G-sequence shadowing property and G-equicontinuity of the inverse limit spaces under group action
  86. Classification of f-biharmonic submanifolds in Lorentz space forms
  87. Some new results on the weaving of K-g-frames in Hilbert spaces
  88. Matrix representation of a cross product and related curl-based differential operators in all space dimensions
  89. Global optimization and applications to a variational inequality problem
  90. Functional equations related to higher derivations in semiprime rings
  91. A partial order on transformation semigroups with restricted range that preserve double direction equivalence
  92. On multi-step methods for singular fractional q-integro-differential equations
  93. Compact perturbations of operators with property (t)
  94. Entire solutions for several complex partial differential-difference equations of Fermat type in ℂ2
  95. Random attractors for stochastic plate equations with memory in unbounded domains
  96. On the convergence of two-step modulus-based matrix splitting iteration method
  97. On the separation method in stochastic reconstruction problem
  98. Robust estimation for partial functional linear regression models based on FPCA and weighted composite quantile regression
  99. Structure of coincidence isometry groups
  100. Sharp function estimates and boundedness for Toeplitz-type operators associated with general fractional integral operators
  101. Oscillatory hyper-Hilbert transform on Wiener amalgam spaces
  102. Euler-type sums involving multiple harmonic sums and binomial coefficients
  103. Poly-falling factorial sequences and poly-rising factorial sequences
  104. Geometric approximations to transition densities of Jump-type Markov processes
  105. Multiple solutions for a quasilinear Choquard equation with critical nonlinearity
  106. Bifurcations and exact traveling wave solutions for the regularized Schamel equation
  107. Almost factorizable weakly type B semigroups
  108. The finite spectrum of Sturm-Liouville problems with n transmission conditions and quadratic eigenparameter-dependent boundary conditions
  109. Ground state sign-changing solutions for a class of quasilinear Schrödinger equations
  110. Epi-quasi normality
  111. Derivative and higher-order Cauchy integral formula of matrix functions
  112. Commutators of multilinear strongly singular integrals on nonhomogeneous metric measure spaces
  113. Solutions to a multi-phase model of sea ice growth
  114. Existence and simulation of positive solutions for m-point fractional differential equations with derivative terms
  115. Bernstein-Walsh type inequalities for derivatives of algebraic polynomials in quasidisks
  116. Review Article
  117. Semiprimeness of semigroup algebras
  118. Special Issue on Problems, Methods and Applications of Nonlinear Analysis (Part II)
  119. Third-order differential equations with three-point boundary conditions
  120. Fractional calculus, zeta functions and Shannon entropy
  121. Uniqueness of positive solutions for boundary value problems associated with indefinite ϕ-Laplacian-type equations
  122. Synchronization of Caputo fractional neural networks with bounded time variable delays
  123. On quasilinear elliptic problems with finite or infinite potential wells
  124. Deterministic and random approximation by the combination of algebraic polynomials and trigonometric polynomials
  125. On a fractional Schrödinger-Poisson system with strong singularity
  126. Parabolic inequalities in Orlicz spaces with data in L1
  127. Special Issue on Evolution Equations, Theory and Applications (Part II)
  128. Impulsive Caputo-Fabrizio fractional differential equations in b-metric spaces
  129. Existence of a solution of Hilfer fractional hybrid problems via new Krasnoselskii-type fixed point theorems
  130. On a nonlinear system of Riemann-Liouville fractional differential equations with semi-coupled integro-multipoint boundary conditions
  131. Blow-up results of the positive solution for a class of degenerate parabolic equations
  132. Long time decay for 3D Navier-Stokes equations in Fourier-Lei-Lin spaces
  133. On the extinction problem for a p-Laplacian equation with a nonlinear gradient source
  134. General decay rate for a viscoelastic wave equation with distributed delay and Balakrishnan-Taylor damping
  135. On hyponormality on a weighted annulus
  136. Exponential stability of Timoshenko system in thermoelasticity of second sound with a memory and distributed delay term
  137. Convergence results on Picard-Krasnoselskii hybrid iterative process in CAT(0) spaces
  138. Special Issue on Boundary Value Problems and their Applications on Biosciences and Engineering (Part I)
  139. Marangoni convection in layers of water-based nanofluids under the effect of rotation
  140. A transient analysis to the M(τ)/M(τ)/k queue with time-dependent parameters
  141. Existence of random attractors and the upper semicontinuity for small random perturbations of 2D Navier-Stokes equations with linear damping
  142. Degenerate binomial and Poisson random variables associated with degenerate Lah-Bell polynomials
  143. Special Issue on Fractional Problems with Variable-Order or Variable Exponents (Part I)
  144. On the mixed fractional quantum and Hadamard derivatives for impulsive boundary value problems
  145. The Lp dual Minkowski problem about 0 < p < 1 and q > 0
Downloaded on 9.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/math-2021-0073/html
Scroll to top button