Quadratic convective flow of radiated nano-Jeffrey liquid subject to multiple convective conditions and Cattaneo-Christov double diffusion
Nomenclature |
(U,
V),
| velocity components along x- and y- |
Grx,
| local Grashof number; |
| axes (m·s-1); |
h1,
| heat transfer coefficient (W·m-2·K-1); |
X,
Y,
| coordinates (m); |
h2,
| mass transfer coefficient; |
Bi1, Bi2,
| Biot numbers; |
Le, | Lewis number; |
DB,
| Brownian diffusion coefficient (m2 · |
f, | dimensionless velocity variable; |
| s-1); |
T, | fluid temperature (K); |
DT,
| thermophoretic diffusion coefficient; |
Tf,
| temperature of fluid near wall (K); |
T∞,
| ambient temperature (K); |
N, | buoyancy parameter; |
C,
| volumetric coefficient; |
NB,
| Brownian motion parameter; |
Cw,
| volume fraction of fluid near wall; |
NT,
| thermophoretic parameter; |
C∞,
| concentration far away from surface; |
Nux,
| local Nusselt number; |
cp,
| specific heat capacity (J·kg-1·K-1); |
Shx,
| local Sherwood number; |
k,
| thermal conductivity (W·m-1·K-1); |
qw,
| surface heat flux; |
k*,
| mean absorption coefficient (m-1); |
qm,
| surface mass flux; |
Uw,
| = aX, stretching sheet velocity (m·s-1); |
Rex,
| local Reynolds number; |
a,
| constant (s-1); |
R, | radiation parameter; |
Cf,
| skin friction coefficient; |
Pr,
| Prandtl number. |
g,
| acceleration due to gravity (m·s-1); | |
Greek symbols
|
ρ,
| density of fluid (kg·m-3); |
β3,
| nonlinear volumetric solute expansion coefficient; |
μ,
| dynamic viscosity (kg·m-1·s-1); |
θ, | dimensionless temperature; |
ν,
| kinematic viscosity of fluid (m2 · s-1); |
θw,
| temperature ratio parameter; |
σ*,
| Stefan-Boltzman constant (W·m-2·K-4); |
φ, | nanoparticle volume fraction; |
α1,
a2,
| nonlinear convection parameters; |
λ, | local mixed convection parameter; |
αm,
| = k/(ρcp), thermal diffusivity (m2·s-1); |
λ1,
| ratio of relaxation/retardation time; |
β,
| Deborah number; |
λ2,
| retardation time; |
β0,
| linear volumetric thermal expansion coefficient; |
λE,
| relaxation time of heat flux; |
β1,
| nonlinear volumetric thermal expansion coefficient; |
λC,
| relaxation time of mass flux; |
β2, | linear volumetric solute expansion coef |
τw,
| wall shear stress; |
| ficient; |
τ, | thermophoretic parameter; |
Superscript |
′,
| derivative with respect to η. | |
Subscripts |
f,
| fluid properties at wall; |
∞, | fluid properties at ambient conditions. |
1 Introduction Non-Newtonian fluid models are beneficial in industrial and manufacturing processes, such as hot rolling, paper production, drilling muds, plastic polymers, metal spinning, and cooling of metallic plates in cooling baths. The non-Newtonian Jeffrey liquid is one of the rate type fluid. This liquid model is accomplished by describing the features of retardation and relaxation time. For this reason, the Jeffrey fluid model has gained much attention from the researchers. Kothandapani and Srinivas[1] analyzed the magnetohydrodynamic (MHD) Jeffrey fluid flow under the peristaltic transport in an irregular channel. Shehzad et al.[2] investigated the flow of Jeffrey fluid at the stagnation point with Dufour and Soret effects. An analysis of unsteady flow and heat transport mechanism in the Jeffrey fluid over a stretching sheet was made by Hayat et al.[3]. Ahmad and Ishak[4] discussed the convective Jeffrey liquid flow at a stagnation point over a linearly elongated vertical surface. Ahmed et al.[5] presented exact solutions for the MHD Jeffrey liquid flow through a convectively heated stretched surface. The stagnation-point flow of a Jeffrey liquid due to a convectively heated stretched disk under the effect of viscous dissipation and Joule heating was examined by Hayat et al.[6]. They reported the convergent series solutions to the problem. Khan et al.[7] presented analytical solutions for Jeffrey liquid flow subject to cross-diffusion characteristics due to a stretching of a permeable cylinder in presence of Newtonian flux condition.
There is a huge development in modern nanotechnology due to its broad involvement in the physiological and industrial processes. The mixture of non-metallic or metallic nanoparticles and conventional heat transfer fluids is referred as a nanofluid. The nanoparticles improve the thermal characteristics of the base fluids. Such practical features of nanofluids have special importance in several industrial and technological developments such as vehicle cooling, transformer and electronic device cooling, biomedicine, heat exchanger, and nuclear reactor cooling. Convective heat transfer with nanofluids can be modelled using the single phase or two-phase Buongiorno nanofluid approach. The first assumes that the fluid phase and particles are in thermal equilibrium and move with the same velocity. It only includes the influence of nanoparticle volume fraction aspect. There are several other physical aspects (mainly Brownian motion and thermophoresis) which are the reason for an enhancement in thermal conductivity. The employed model would incorporate these two physical aspects. Also, this approach is simpler and requires less computational time. Choi[8] first introduced the term nanoliquid. The mathematical expressions of nanofluid through thermophoretic and Brownian movement aspects were developed by Buongiorno[9]. The Brownian motion acts against concentration. In addition, thermophoresis moves the particles from hot regions to cold ones. The nanoliquid flow by an elongated surface was initially investigated by Khan and Pop[10]. Hayat et al.[11] addressed the three-dimensional MHD flow of viscous nanofluid over an impermeable nonlinearly stretched surface with the convective condition. The MHD Maxwell nanoliquid flow in a saturating non-Darcy porous medium was investigated by Muhammad et al.[12]. They reported the analytical solutions using the homotopic analysis method. Sheikholeslami et al.[13] studied the MHD flow of nanofluid inside a permeable cavity with hot elliptic obstacle through the lattice Boltzmann method. Later, many researchers have been engaged in recent developments on nanofluids (see Refs. [14]-[24]).
The flowing and mixing of liquid are generated by density variation due to difference in temperature within the liquid, and it is called convection or natural convection. Near the hot surface, the density of the liquid is smaller than that of the cold fluid. The buoyant force appears due to gravity that lifts the heated liquid upward. The problems of convective flow of working fluids have numerous engineering applications, such as solar collectors, food preservation, energy storage, nuclear reactor technology, and cryogenic devices. Moreover, the nonlinear convection produces due to nonlinear density temperature variations in the buoyancy force term, and it has a significant impact on the flow characteristics. Keeping this fact in mind, Vajravelu et al.[25], Kameswaran et al.[26], Shaw et al.[27-28], Ramreddy and Pradeepa[29], Bandaru et al.[30], and Mahanthesh et al.[31] investigated the influence of nonlinear convection in flow and heat transfer under different physical aspects.
The heat transfer process is a hot topic of modern technology, and it arises when there is a difference in temperature between the physical systems. It has plenty of applications in modern industry and technology, for instance, power generation, atomic reactor cooling, and energy production. The theory of heat conduction for heat transfer attributes was estimated by Fourier[32], which is highly useful for characterization of heat transport process in various circumstances. Cattaneo[33] improved the Fourier law[32] by insertion of thermal relaxation time. Christov[34] improved the Cattaneo model[33] by introducing the Oldroyd's upper-convected derivative. The convective flow of Newtonian fluid by considering the Cattaneo-Christov heat flux model was investigated by Straughan[35]. The spinning flow of Maxwell fluid was studied by Mustafa[36] through the Cattaneo-Christov theory of heat transport. Khan et al.[37] analyzed the viscoelastic fluid flow by employing Cattaneo-Christov heat flux aspects. Waqas et al.[38] explored the aspect of Cattaneo-Christov model on the flow of generalized Burgers fluid. The comparison of two classes of viscoelastic fluids through Cattaneo-Christov diffusion expressions was examined by Hayat et al.[39]. Shehzad et al.[40] investigated the three-dimensional hydrodynamic flow of Maxwell liquid through the Cattaneo-Christov heat and mass flux model.
Here, the main goal is to investigate the nonlinear convection flow of Jeffrey liquid across an elongated sheet with heat and mass fluxes via Cattaneo-Christov double diffusion expressions. Further, the influence of thermophoretic and Brownian movement, nonlinear radiation, and convective boundary conditions is accounted. So far, the above-stated model has not been considered yet. The numerical solutions via the Runge-Kutta-Fehlberg (RKF) method along with the shooting system are constructed. For a particular case, the obtained numerical results are compared with the previously reported results of Hayat et al.[3] and good agreement is achieved.
2 Mathematical formulation The nonlinear convective and radiated flow of Jeffrey nanoliquid due to the stretched sheet is considered. Uw (=aX) is the stretching velocity. The coordinate system is assumed that, the sheet is along with the X-axis, while Y measures normally outward to it (see Fig. 1). The sheet is supposed to be heated by a hot fluid with the concentration Cf and the temperature Tf which are related with the mass/heat transfer coefficient h2/h1. T∞ and C∞ are the ambient temperature and the concentration, correspondingly.
The relevant equations are[3, 30, 36]
|
(1) |
|
(2) |
|
(3) |
|
(4) |
with the boundary conditions[11, 23]
|
(5) |
whereas
U and V are velocities, and ν, μ,
ρ, λ1, λ2, g, β0, β1, β2, β3, T, and C symbolize the kinematic viscosity, the dynamic viscosity, the density, the ratio of relaxation to retardation time, the retardation time, the acceleration due to gravity, the linear volumetric thermal expansion coefficient, the nonlinear volumetric thermal expansion coefficient, the linear volumetric solute expansion coefficient, the nonlinear volumetric solute expansion coefficient, the temperature, and the concentration, respectively.
α=k/(ρc)f is the thermal diffusivity, k is the thermal conductivity, (ρc)f is the heat capacity of liquid, (ρc)p is the nanoparticle effective heat capacity,
DB is the Brownian diffusion coefficient, DT is the thermophoretic diffusion coefficient, λE and λC are the relaxation time of heat and mass fluxes, respectively, and σ* and k* are the Stefan-Boltzman constant and the mean absorption coefficient, respectively. Substituting the following similarity variables
|
(6) |
into (1)-(5) yields
|
(7) |
|
(8) |
|
(9) |
with
|
(10) |
where λ is the local mixed convection parameter, Grx is the local Grashof number, Rex is the local Reynolds number,
α1 and α2 are the nonlinear thermal and solutal convection parameters, respectively, N is the buoyancy ratio parameter, β is the Deborah number, Le is the Lewis number,
Pr is the Prandtl number, NB is the Brownian movement parameter, R is the parameter of radiation, NT is the parameter of thermophoretic, Bi1 is the thermal Biot number,
δ1 is the thermal relaxation parameter, δ2 is the concentration relaxation parameter, Bi2 is the concentration Biot number, and θw is the temperature ratio parameter. These parameters are defined as follows:
|
(11) |
The skin friction coefficient, the local Nusselt number, and the local Sherwood number are
|
(12) |
where τw, qw, and qm are
|
(13) |
The dimensionless forms of (12) are
|
(14) |
3 Method of solution and validation The nonlinear ordinary differential equations (7)-(9) subject to the boundary conditions (10) are solved by the shooting technique along with the RKF method. The system of (7)-(9) is reduced to a first-order differential equation system by establishing the following new variables:
Accordingly, we have
with
To determine the unknowns using the shooting method, first, guess the values for m1, m2, m3, and m4. In a while, the subsequent initial value problem is treated numerically by the RKF scheme. We set the convergence criteria and step size as 10-6 and 10-3, correspondingly. The values of f''(0) are compared with the previously published homotopy analysis method (HAM) results of Hayat et al.[3] for the specific case. The comparison results are found in outstanding agreement and presented in Table 1.
Table 1 Comparison of f'' (0) with that of Hayat et al.[3] when λ=0 for isothermal boundary conditions
4 Physical interpretation This section emphasizes the significance of physical parameters on the liquid velocity f'(η), the temperature θ(η), and the concentration φ(η). The behaviors of the skin friction Cf, the local Nusselt number Nux, and the local Sherwood number Shw for various physical constraints are visualized through the numeric data presented in tables. Figure 2 illustrates the impact of the Deborah number on the velocity distribution f'(η). It is visualized that f'(η) increases with an increase in β. In fact, the Deborah number is dependent on the retardation time. Therefore, an increase in β leads to an increase in the retardation time, which reduces the viscous force. Consequently, the liquid velocity field increases.
The impact of λ on f'(η) is illustrated in Fig. 3. It is noted that the velocity distribution shows an increasing behavior corresponding to higher values of the mixed convection parameter. This is because the convection parameter and the buoyancy force are directly proportional.
The influence of λ1 on the velocity is demonstrated in Fig. 4. We visualize that f'(η) decreases for higher values of λ1. Physically, λ1 denotes the ratio of relaxation and retardation time.
f'(η) is higher for large values of α1 for both absence and presence of the Cattaneo-Christov theory of heat and mass diffusion (see Fig. 5). Physically, the nonlinear convection (α1) increases the temperature gradient at the wall and enhances the buoyancy force due to convection which forces the liquid to move away from the sheet. Consequently, f'(η) is increased.
The impact of N on f'(η) is depicted in Fig. 6. Here, as the buoyancy ratio parameter increases, the momentum boundary layer becomes thicker. Figure 7 depicts the variations of Bi1 on θ(η). It is shown that the layer of thermal boundary is superior for larger Biot numbers. The heat transfer coefficient h1 increases at the boundary with an increase in Bi. Consequently, heat energy is restored in liquid. The thermal boundary layer grows thicker.
The variation in θ(η) for thermal relaxation is explored in Fig. 8. The larger thermal relaxation δ1
(=aλE) increases the relaxation time of heat flux λE. As a result, an increment in δ1 shows a reduction in the temperature profile. Figures 9 and 10 depict the influence of α1 and λ on the temperature profile, respectively. Here, θ(η) decays for larger values of both α1 and λ.
Moreover, the temperature field is weaker in the presence of Cattaneo-Christov heat and mass flux compared with their absence. The effect of λ1 on θ(η) is demonstrated in Fig. 11. It is observed that the temperature of fluid is enhanced with higher values of λ1.
The variations of θ(η) due to the effects of NB are shown in Fig. 12. Here, θ(η) enhances for higher NB. Physically, the random motion of nanoparticles increases by increasing NB. As a result, the temperature profile increases. The temperature field is lower for higher values of the buoyancy ratio parameter (see Fig. 13).
Figures 14 and 15 are presented to explore the effects of R and θw on θ(η). In fact, the radiation process generates more heat into the fluid system, due to which θ(η) enhances. It is far clear that an enhancement within the temperature ratio parameter corresponds to a higher wall temperature in comparison with the ambient fluid.
Figures 16 and 17 show the variations of θ(η) and φ(η) for NT, respectively. Here, θ(η) and φ(η) increase significantly for larger NT.
The variation in the concentration profile for the concentration relaxation parameter (δ2) is illustrated in Fig. 18. By increasing δ2 (=aλC), the concentration and its boundary layer thickness decrease. Physically, the relaxation time of mass flux λC raises with an increase in δ2.
Figures 19 and 20 establish the effects of Biot numbers Bi1 and Bi2 on the nanoparticle concentration profile, respectively. The nanoparticle concentration profile increases with the Biot numbers. Thus, the convective conditions can be used as key aspects to control the development of thermal and concentration boundary layers on the surface. φ(η) decreases for large values of α2 (see Fig. 21).
The variation of φ(η) for N is displayed in Fig. 22. Here, φ(η) is reduced with larger values of the buoyancy ratio parameter. Figure 23 depicts the influence of Le on φ(η). As expected, the concentration field diminishes for inflation of the Lewis number parameter. As higher values of the Lewis number Le (=ν/DB) decay the strength of random moment of nanoparticles, φ(η) decreases.
Tables 2-4 present numerical data of the skin friction coefficient (CfRex1/2), the Nusselt number (NuxRex-1/2), and the Sherwood number (ShwRex-1/2) when Bi1=Bi2=λ=λ1=0.5, N=β=0.4, NB=NT=R=0.3, α1=α2=0.3, Pr=0.71, Le=0.5, θw=1.2, and δ1=δ2=0.2. The varying parameter values are specified in the tables. From Table 2, we observe that larger values of Bi1 decay ShwRex-1/2, while they enhance CfRex1/2. CfRex1/2 and NuxRex-1/2 are significantly varied for Bi1 from 0.5 to 20. After that, small variations can be noticed.
Table 2 Numerical results of CfRex1/2, NuxRex-1/2, and ShwRex-1/2 for different values of Bi1 and Bi2
Table 3 Numerical results of CfRex1/2, NuxRex-1/2, and ShwRex-1/2 for different values of λ1, β, λ, N, α1, and α2
Table 4 Numerical results of NuxRex-1/2 and ShwRex-1/2 for different values of NB, NT, R, θw, δ1, δ2, Pr, and Le
From Table 3, it is observed that nonlinear thermal and concentration convections display better friction factors. Also, the heat and mass transfer rate is higher. It is also noticeable that the friction factor is a decreasing function of λ1 whereas it is an increasing function of λ and N. λ and N have propensity to increase the Nusselt number, while an opposite trend is perceived for λ1. This outcome is exactly similar for the Sherwood number (see Table 3). Interestingly, it is also noted that the effect of β is not reliable on the friction factor, Nusselt number, and Sherwood number profiles. All three distributions increase initially by increasing the value of β and then decrease (see Table 3). Finally, from Table 4, it is clear that the rate of heat transfer coefficient enlarges for larger values of R, θw, Pr, and δ1. However, we can see an opposite result with an increase in NB, NT, δ2, and Le.
5 Concluding remarks A nonlinear convection flow of Jeffrey nanofluid is examined through double diffusions. The main features are
(ⅰ) The Deborah number is constructive for the velocity profile.
(ⅱ) The parameter of thermophoresis leads to higher curves of temperature and concentration.
(ⅲ) Consequence of the buoyancy ratio parameter on the velocity and temperature fields is quite opposite.
(ⅵ) The nonlinear convection parameters α1 and α2 give rise to the local Nusselt and Sherwood numbers and also better friction factor.
(ⅴ) The rate of heat transport coefficient increases for R, θw, Pr, and δ1 but decreases for NB, NT, δ2, and Sc.
(ⅳ) The velocity and the heat and mass transport rate are lower in presence of δ1 and δ2 compared with their absence.
Acknowledgements
One of the authors P. B. SAMPATH KUMAR is thankful to University Grant Commission (UGC), New Delhi, for their financial support under National Fellowship for Higher Education (NFHE) of ST students to pursue M. Phil/PhD Degree (F117.1/201516/NFST201517STKAR2228/(SAIII/Website) Dated: 06-April-2016). Also, the author B. MAHANTHESH is thankful to the Management of Christ University, Bengaluru, India, for the support through Major Research Project to accomplish this research work.