Next Article in Journal
Study on Shear Velocity Profile Inversion Using an Improved High Frequency Constrained Algorithm
Next Article in Special Issue
Heat Transfer Analysis of Sisko Fluid Flow over a Stretching Sheet in a Conducting Field with Newtonian Heating and Constant Heat Flux
Previous Article in Journal
Condition Assessment of Power Transformers through DGA Measurements Evaluation Using Adaptive Algorithms and Deep Learning
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Heat Transfer of Buoyancy and Radiation on the Free Convection Boundary Layer MHD Flow across a Stretchable Porous Sheet

1
Department of Mathematics and Statistics, University of Victoria, Victoria, BC V8W 3R4, Canada
2
Department of Medical Research, China Medical University Hospital, China Medical University, Taichung 40402, Taiwan
3
Department of Mathematics and Informatics, Azerbaijan University, 71 Jeyhun Hajibeyli Street, AZ1007 Baku, Azerbaijan
4
Center for Converging Humanities, Kyung Hee University, 26 Kyungheedae-ro, Dongdaemun-gu, Seoul 02447, Republic of Korea
5
Department of Mathematics, University of Swabi, Swabi 23561, KPK, Pakistan
6
Department of Mathematics, College of Education, University of Sulaimani, Sulaimani 46001, Iraq
7
Department of Mathematics, Khalifa University, Abu Dhabi P.O. Box 127788, United Arab Emirates
8
Department of Mathematics, Al-Hussein Bin Talal University, Ma’an P.O. Box 33011, Jordan
*
Author to whom correspondence should be addressed.
Energies 2023, 16(1), 58; https://doi.org/10.3390/en16010058
Submission received: 17 November 2022 / Revised: 10 December 2022 / Accepted: 16 December 2022 / Published: 21 December 2022
(This article belongs to the Special Issue Turbulence and Fluid Mechanic)

Abstract

:
Theoretical influence of the buoyancy and thermal radiation effects on the MHD (magnetohydrodynamics) flow across a stretchable porous sheet were analyzed in the present study. The Darcy–Forchheimer model and laminar flow were considered for the flow problem that was investigated. The flow was taken to incorporate a temperature-dependent heat source or sink. The study also incorporated the influences of Brownian motion and thermophoresis. The general form of the buoyancy term in the momentum equation for a free convection boundary layer is derived in this study. A favorable comparison with earlier published studies was achieved. Graphs were used to investigate and explain how different physical parameters affect the velocity, the temperature, and the concentration field. Additionally, tables are included in order to discuss the outcomes of the Sherwood number, the Nusselt number, and skin friction. The fundamental governing partial differential equations (PDEs), which are used in the modeling and analysis of the MHD flow problem, were transformed into a collection of ordinary differential equations (ODEs) by utilizing the similarity transformation. A semi-analytical approach homotopy analysis method (HAM) was applied for approximating the solutions of the modeled equations. The model finds several important applications, such as steel rolling, nuclear explosions, cooling of transmission lines, heating of the room by the use of a radiator, cooling the reactor core in nuclear power plants, design of fins, solar power technology, combustion chambers, astrophysical flow, electric transformers, and rectifiers. Among the various outcomes of the study, it was discovered that skin friction surges for 0.3 F 1 0.6, 0.1 k 1 0.4 and 0.3 M 1.0, snf declines for 1.0 G r 4.0. Moreover, the Nusselt number augments for 0.5 R 1.5, 0.2 N t 0.8 and 0.3 N b 0.9, and declines for 2.5 P r 5.5. The Sherwood number increases for 0.2 N t 0.8 and 0.3 S c 0.9, and decreases for 0.1 N b 0.7.

1. Introduction and Motivation

Magnetohydrodynamics (MHD) is the study of the magnetic impact of an electrically conducted incompressible fluid; it has received attention because of its major applications in different areas, such as X-ray radiation, crystal growth, asthma treatment, tumor therapy, plasma, fusing metal in electric heaters, nuclear reactors, gastric medication, and electrolytic biomedicine. MHD fluid can also be used to manage the cooling rates in certain industrial processes. With homogeneous magnetic fields, Chaudhary and Kumar [1] explored an unsteady 2D flow of an electrically conducted fluid toward a stretchable sheet. Naramgari and Sulochana [2] analyzed the chemical reaction and thermal radiation effects on a two-dimensional MHD steady flow toward a stretchable sheet using injection/suction. By using a chemical reaction, Reddy et al. [3] analyzed the impacts of MHD and buoyancy forces through rotating vertically porous plates. Mishra and Jena [4] explored steady two-dimensional incompressible, laminar, and electrically conducted fluids toward a stretchable sheet using magnetic fields and the viscous dissipation effect. Babu et al. [5] found numerical solutions of a steady two-dimensional MHD boundary layer flow toward a shrinking sheet by using the wall mass suction effect. Using the homotopy analysis method (HAM) technique, Mabood et al. [6] found the impact of radiation on the MHD flow toward an exponential stretchable sheet. Pal and Mondal [7] investigated heat transfer characteristics and the MHD non-Darcy boundary layer flow toward an electrically conducted fluid in the presence of the viscosity impact. The researchers in [8] demonstrated the MHD natural convection-based nanofluid flow toward a porous surface. Using the radiation effect, Mahmoud [9] investigated the heat transport flow of electrically conducted fluids across continuously moving vertical infinite plates with heat flux and suction. With variable heat, radiation effect, and mass diffusion, Kishore et al. [10] described the viscous incompressible MHD flow in porous media. Majeed et al. [11] evaluated the impact of energy activation and chemical reactions on the two-dimensional flow of heat transfer toward an exponential stretchable surface. With radiation influence and the boundary layer flow of viscous fluid, Sharma et al. [12] solved the problem of heat transfer due to an exponentially stretching/shrinking sheet. Poornima and Reddy [13] found the mathematical solutions of the convective flow of radiating boundary layer-combined nanofluids to a non-linear movable sheet using induced magnetic fields. Some studies involving MHD flow and thermal radiation have also been published in [14,15,16,17,18,19,20]. The researchers in [21] discussed the heat generation and Cattaneo–Christov heat flux effect on a hybrid nanofluid flow over two distinct geometries. Yaseen et al. [22] explored the Darcy–Forchheimer model and Cattaneo–Christov heat flux of MoS 2 -SiO 2 /kerosene oil between shrinking and rotating disks. Khan et al. [23] investigated the magnetohydrodynamic thin film flow through a porous stretching sheet by focusing on the impact of thermal radiation and viscous dissipation. Jawad et al. [24] discussed the analysis of the hybrid nanofluid stagnation point flow over a stretching surface by focusing on melting heat transfer.
Heat transfer analysis toward a porous stretchable sheet has received a lot of attention from architects, geologists, engineers, and mathematicians due to its occurrence in many practical and scientific situations, such as electronic cooling, thermal insulation, catalytic reactor, nuclear waste repository, oil production, geothermal system, petroleum industries, and energy storage units (see [25,26,27,28]). In 1856, Henry Darcy worked with homogeneous fluid flow passing through a porous medium. However, the classical Darcy law did not work for higher permeability flows and large velocities. In order to overcome this demerit, the researchers in [29] revised the Darcy velocity formula by including the velocity square into the equation of momentum to calculate the boundary layer flow and inertial effects. Pal and Mondal [30] described the Darcy–Forchheimer model to examine the mixed convection flow of a stretchable sheet using thermal radiation and variable viscosity. Ganesh et al. [31] examined the Darcy–Forchheimer model and hydromagnetic nanofluids on stretchable surfaces. The Darcy–Forchheimer–Casson fluid flow was investigated by Seth and Mandal [32], who discovered the influences of rotational and Casson parameters on the primary velocity. Under the saturated porous medium of the Darcy–Forchheimer flow, Seddeek [33] investigated the combined effects of viscous dissipation and the thermophoresis parameter. Hayat et al. [34] evaluated the 2D Darcy–Forchheimer flow of Maxwell fluids, induced by a stretchable sheet. Rajesh et al. [35] examined the heat transfer rate and the hybrid nanoparticle impact on the MHD flow through a vertical plate. The researchers in [36] studied the heat transfer analysis of nanoparticles (MoS 2 /H 2 O) and hybrid nanoparticles (MoS 2 -SiO 2 /H 2 O-C 2 H 6 O 2 ) between two distinct plates by using heat absorption/generation. Raza et al. [37] explored the influence of Joules heating and the Brownian motion radiative nanofluid toward a stretchable cylinder. Faizan et al. [38] described the behaviors of heat and mass transfer for the Cattaneo–Christov nanofluid toward a Riga sheet.
The injection and suction have remarkable effects on boundary layer flow surfaces. The suction is used to upsurge the skin friction, whereas the injection works in the opposite direction. The injection of fluids through a porous sheet has practical applications in the boundary layer problem, such as wire coating, polymer fiber coating, and film cooling. Furthermore, they are important in the field of engineering, e.g., regarding radial diffusion and the construction of thrust bearings (see [39]). Suction is utilized in chemical progression to remove the reactant, whereas blowing is used to introduce a reactant, cool the surfaces, minimize the drag, and prevent corrosion (see [40,41,42]).
Many phenomena in applied sciences and engineering fields are governed by non-linear BVPs. As a result, BVPs have received attention from engineers, mathematicians, and physicists. Generally speaking, compared to linear ODEs and PDEs, it is significantly more complicated to find the analytical solutions of non-linear ODEs and PDEs, particularly when using analytical methods. Therefore, in 1992, Liao [43] proposed the homotopy analysis method (HAM) [44,45,46,47,48], relying on homotopy in topology. This approach is unaffected by significant large/small physical variations. The HAM approach provides great versatility in terms of changing and choosing the convergence areas and estimating rates. The method has an advantage over typical computational methodologies in that it avoids rounding off errors induced by the discretization procedure.
Judging by the aforementioned and other related literature studies, numerous studies have been presented for buoyancy and radiation influence on the MHD flow toward a stretchable sheet; however, limited studies are available for the Darcy–Forchheimer model with the Brownian motion, heat source, sink, or the thermophoresis effect on the MHD flow across a stretchable porous sheet. Therefore, the present study fills this gap. The novelty of this study is that it enhances the work of Daniel [49] by introducing the species conservation equation and the Darcy–Forchheimer model with the influence of the Brownian motion, heat source, heat sink, or thermophoresis effect on the MHD flow across a stretchable porous sheet. The general form of the buoyancy term in the momentum equation for a free convection boundary layer is derived in this study. The fluid flow is considered laminar, incompressible, and electrically conducted. Mathematica software and homotopy analysis techniques were utilized to solve the modeled equations. The variations of several physical parameters on the flow, as well as thermal and concentration characteristics, are discussed briefly. The present study provides several applications, such as film cooling, wire coating, polymer fiber coating, construction of thrust bearings, radial diffusion, oil production, thermal insulation, X-ray radiation, fusing metal in electric heaters, and gastric medication.
The present study provides the following research questions:
  • What is the general form of the buoyancy term in the momentum equation for a free convection boundary layer? How may it be approximated if the flow is due to temperature variations? What is the name of the approximation?
  • What physical processes are represented by the terms of the momentum Equation (19)? By the energy Equation (20)? By the species conservation Equation (21)?
  • What is the definition of the Prandtl number? How does its value affect the relative growth of the thermal boundary layer for the laminar flow toward a porous stretching sheet?

2. Preliminaries

It is commonly recognized that non-linear PDEs and ODEs for BVPs are significantly harder to solve than linear PDEs and ODEs, particularly when we attempt to use analytical techniques. Traditionally, asymptotic and perturbation approaches are frequently used to obtain analytic approximations of non-linear problems in engineering. Unfortunately, asymptotic and perturbation methods only work for weak nonlinear systems because they generally rely too heavily on small/large physical parameters. Therefore, certain analytical approximation techniques must be developed to be applicable to extremely non-linear problems, independently of any large/small physical parameters. Thus, in 1992, Liao [43] developed one of these semi-analytic approximation techniques.
We begin by presenting the following nonlinear differential equation:
N 1 [ u 1 ( x 1 , t 1 ) ] = 0 ,
where N 1 and u 1 ( x 1 , t 1 ) represent the nonlinear operator and the unknown function, respectively. Liao [43] introduced a two-parameter set of equations in the embedded parameter p [0,1], which is known as the 0 th -order deformation equation given by
( 1 p ) L 1 [ ϕ ( x 1 , t 1 ; p ) u 0 ( x 1 , t 1 ) ] = p N 1 [ ϕ ( x 1 , t 1 ; p ) ] ,
where L 1 , ϕ ( x 1 , t 1 ; p ) and u 0 ( x 1 ) represent an auxiliary linear-operator, unknown function, and initial guess, respectively. We find at p = 0 and p = 1 that
ϕ ( x 1 , t 1 ; 0 ) = u 0 ( x 1 , t 1 ) ,
and
ϕ ( x 1 , t 1 ; 1 ) = u 1 ( x 1 , t 1 ) .
The solution ϕ ( x 1 , t 1 ; p ) of the 0 th -order deformation equation distorts from the initial estimate u 0 ( x 1 , t 1 ) to the actual solution u 1 ( x 1 , t 1 ) of the given equation N 1 [ u 1 ( x 1 , t 1 ) ] = 0 as the embedded parameter p [0,1] increases from 0 to 1. We can extend ϕ ( x 1 , t 1 ; p ) into a Maclaurin expansion with respect to p because it also depends on the embedded parameter p [0,1], as follows:
ϕ ( x 1 , t 1 ; p ) = u 0 ( x 1 , t 1 ) + m = 1 u m ( x 1 , t 1 ) p m ,
where
u m ( x 1 , t 1 ) = 1 m ! m ϕ ( x 1 , t 1 ; p ) p m p = 0 .
Equation (5) is known as the homotopy Maclaurin expansion of ϕ ( x 1 , t 1 ; p ) . Particularly, at p = 1, we have
ϕ ( x 1 , t 1 ; 1 ) = u 0 ( x 1 , t 1 ) + m = 1 u m ( x 1 , t 1 ) .
We obtain the following homotopy series solution if the aforementioned series converges to ϕ ( x 1 , t 1 ; 1 ) , as indicated by Equation (4):
u 1 ( x 1 , t 1 ) = u 0 ( x 1 , t 1 ) + m = 1 u m ( x 1 , t 1 ) ,
which satisfies the given equation N 1 [ u 1 ( x 1 , t 1 ) ] = 0 as proven by Liao [43]. Equation u m ( x 1 , t 1 ) is entirely determined by means of Equation (2). Taking the derivative of Equation (2) m times with respect to the embedded parameter p, if we divide by m ! and finally put p = 0, then we have
L 1 [ u m ( x 1 , t 1 ) χ m u m 1 ( x 1 , t 1 ) ] = D m 1 N 1 [ ϕ ( x 1 , t 1 ; p ) ] ,
where
D m 1 = 1 ( m 1 ) ! m 1 p m 1 p = 0 ,
and
χ m = 0 , if m 1 1 , if m > 1 .

3. Mathematical Formulation

Here, we assume a steady, incompressible, and electrically conducted MHD flow toward a stretching porous sheet that is driven by the buoyancy forces. We also assume that the fluid flow is a laminar, two-dimensional, constant property condition in which the gravity force acts in the negative x-direction. A uniform magnetic field is applied normally to the sheet. The exception is for the influence of the variable density in the buoyancy force since it is this variation that induces fluid motion. The x-axis is taken along the stretching sheet in the direction of the motion while the y-axis is perpendicular to the sheet. Geometric representation of the physical flow model is shown in Figure 1. The transverse applied magnetic field and magnetic Reynolds number are assumed to be very small. Viscous dissipation, induced magnetic field, and the pressure gradient are considered negligible.
Under the above-stated assumption, the continuity equation is written as:
u x + v y = 0 ,
here, the term represents the net outflow (outflow minus inflow) of the mass in the x and y directions, the sum of which must be zero for the steady flow.
The appropriate form of the momentum equation can be represented as:
u u x + v u y = υ 2 u y 2 σ B 0 2 ( x ) ρ u d p ρ d x g 1 u υ k 0 F 0 u 2 ,
where d p ρ d x is the free stream pressure gradient in the region outside the boundary layer. In this region, u = 0, then we have
d p ρ d x = ρ g 1 ,
substituting Equation (14) into Equation (13), we obtain the following expression:
u u x + v u y = υ 2 u y 2 σ B 0 2 ( x ) ρ u + g 1 ( Δ ρ / ρ ) u υ k 0 F 0 u 2 ,
where Δ ρ = ρ ρ .
The third component on the right side of Equation (15) is the buoyancy force; the fluid flow originates because the density ρ is a parameter. If the ρ variation is only due to the temperature variation, then the component may be linked to a fluid property known as the volumetric thermal expansion coefficient.
β 1 = 1 ρ ρ T p .
This thermodynamic property of the fluid provides a measure of the amount by which the density change responds to a change in the temperature at constant pressure. It is expressed in the following approximate form
β 1 Δ ρ ρ Δ T = 1 ρ ρ ρ T T ,
it follows that
( ρ ρ ) ρ β 1 ( T T ) .
This simplification is known as the Boussinesq approximation, and substituting into Equation (15), we obtain the following constituted boundary layers governing the equation (see Daniel [49] and Chamkha [50]):
u u x + v u y = υ 2 u y 2 σ B 0 2 ( x ) ρ u + g 1 β 1 ( T T ) u υ k 0 F 0 u 2 ,
u T x + v T y = α 2 T y 2 + Q 1 ρ c p ( T T ) + β 1 * u ρ c p ( T T ) + 1 c p ρ q r y + τ D B C y T y + D T T T y 2 ,
and
u C x + v C y = 2 C y 2 D B + 2 T y 2 D T T .
The appropriate boundary constraints are given by
u = a 1 x , v = v w , T = T w = T + A 0 x C = C w = B 0 x + C a t y = 0 u 0 , C C , T T a s y .
Here, the velocity components are denoted by u and v along the x-axis and the y-axis, respectively, c p is the specific heat, T is the temperature, B 0 is the applied magnetic induction, T w is the wall temperature, β 1 is the coefficient of thermal expansion, a 1 is the stretching rate, g 1 is the acceleration due to gravity, υ is the fluid kinematic viscosity, k 1 * is the mean absorption coefficient, ρ is the fluid density, v w is the wall injection when v w > 0 and the wall suction when ( v w < 0 ), σ is the electrical conductivity, β 1 * u ( T T ) and Q 1 ( T T ) are the heat generation and absorption, where ( β 1 * and Q 1 are constants) and σ 1 * is the Stefan Boltzmann constant. The first term on the right-hand side of Equations (19)–(21) represents the net force due to the viscous shear stress, the net inflow of thermal energy due to the y-direction conduction, and the net inflow due to the y-direction diffusion, respectively. In terms of the free stream function ψ ( x , y ) , the components of velocity are described as follows:
u = ψ y , v = ψ x .
For the similarity solutions, we use the following acceptable variable:
η = y a 1 υ , θ = T T T w T , ψ ( x , y ) = υ a 1 x f ( η ) , ϕ = C C C w C .
The heat flux q r is calculated as follows:
q r = 4 σ 1 * 3 k 1 * T 4 y .
Here, we have
T 4 4 T 3 T 3 T 4 ,
upon inserting Equations (24)–(26) into Equations (19)–(21), we have
f + f 2 + θ G r + f f M 2 f f 2 F 1 + f k 1 = 0 ,
1 P r ( 1 + 4 R 3 ) θ + ϕ 2 N t + θ f + θ ϕ N b + Δ θ δ 1 θ f = 0 ,
and
ϕ + ϕ f S c + θ N t N b = 0 .
The dimensionless BCs are given
f ( 0 ) = R 1 , θ ( 0 ) = 1 , ϕ ( 0 ) = 1 , f ( 0 ) = 1 , f ( ) = 1 , ϕ ( ) = 0 , θ ( ) = 0 .
The converging parameters involved in the foregoing Equations (27)–(30) are given below:
F 1 = C b k 0 x , N t = τ D T ( T w T ) υ T , k 1 = υ k 0 a 1 , N b = τ D B ( C w C ) υ , S c = υ D B , M = ( σ ρ a 1 ) 1 2 B 0 , δ 1 = β 1 * x ρ c p , R = 4 σ 1 * T 3 k 1 * k 0 , P r = υ α , G r = g 1 β 1 ( T w T ) a 1 2 x , R 1 = v w υ a 1 , Δ = Q 1 ρ c p a 1 .
We explained detail formulation of all governing equations in Appendix A.

Physical Quantities

Some of the most important physical quantities in the engineering processes are S h x , N u x , and C f , which are indicated below:
S h x = x q m D B C w C , N u x = x q w k f T w T , C f = τ w ρ u w 2 ,
where
q m = D B C y y = 0 , q w = k f T y y = 0 , τ w = μ u y y = 0 .
Using Equations (24) and (33) into Equation (32), we have
R e 1 2 S h x = ϕ 0 , R e 1 2 N u x = θ ( 0 ) , R e 1 2 C f = f ( 0 ) .
Here R e = x u w υ denotes the Reynolds number.

4. The HAM Solution

In order to determine an approximate solution for the dimensionless set of Equations (27)–(29) by incorporating the boundary condition as given in Equation (30), we shall apply the well-known homotopy analysis method (HAM) in the Mathematica software. To employ this method, we need initial guesses for the approximate solutions to Equations (27)–(29). These initial guesses are presented as follows:
ϕ 0 ( η ) = e η , θ 0 ( η ) = e η , f 0 ( η ) = R 1 e η + η .
The linear operators L ϕ , L θ , and L f are given below:
L ϕ ( ϕ ) = ϕ , L θ ( θ ) = θ , L f ( f ) = f .
The expanded forms of the linear operators in Equation (36) are given by
L ϕ ( c 1 + c 2 η ) = 0 , L θ ( c 3 + c 4 η ) = 0 , L f ( c 5 + c 6 η + c 7 η 2 ) = 0 .
In this last Equation (37), c i ( i = 1 , 2 , , 7 ) are considered constant. Moreover, on the basis of Equation (35), the non-linear operators N ϕ ^ , N θ ^ , and N f ^ are given as follows:
N ϕ ^ ϕ ^ ( η ; p ) , θ ^ ( η ; p ) , f ^ ( η ; p ) = ϕ ^ η η + ( f ^ η ϕ ^ + ϕ ^ η f ^ ) S c + θ ^ η η N t N b ,
N θ ^ θ ^ ( η ; p ) , ϕ ^ ( η ; p ) , f ^ ( η ; p ) = 1 P r 1 + 4 R 3 θ ^ η η + ϕ ^ η 2 N t + θ ^ η f ^ + θ ^ η ϕ ^ η N b + Δ θ ^ ( 1 + δ 1 ) θ ^ f ^ η ,
and
N f ^ f ^ ( η ; p ) = f ^ η η η + f ^ η 2 + θ ^ G r + f ^ f ^ η η M 2 f ^ η f ^ η 2 F 1 + f ^ η k 1 .
The deformation equations that can be of the 0 th order are listed as follows:
( 1 p ) L ϕ ^ ϕ ^ ( η ; p ) ϕ ^ 0 ( η ) = p ϕ ^ N ϕ ^ ϕ ^ ( η ; p ) , θ ^ ( η ; p ) , f ^ ( η ; p ) ,
( 1 p ) L θ ^ θ ^ ( η ; p ) θ ^ 0 ( η ) = p θ ^ N θ ^ ϕ ^ ( η ; p ) , θ ^ ( η ; p ) , f ^ ( η ; p ) ,
and
( 1 p ) L f ^ f ^ ( η ; p ) f ^ 0 ( η ) = p f ^ N f ^ f ^ ( η ; p ) .
The BCs for Equations (41)–(43) are given by
f ^ ( η ; p ) η = 0 = R 1 , θ ^ ( η ; p ) η = 0 = 1 , ϕ ^ ( η ; p ) η = 0 = 1 , f ^ ( η ; p ) η η = 0 = 1 , f ^ ( η ; p ) η η = = 1 , θ ^ ( η ; p ) η = = 0 , ϕ ^ ( η ; p ) η = = 0 .
Moreover, when p = 0 and p = 1, we have
ϕ ^ ( η ; 0 ) = ϕ 0 ^ ( η ) , θ ^ ( η ; 0 ) = θ 0 ^ ( η ) , f ^ ( η ; 0 ) = f 0 ^ ( η ) ,
and
ϕ ^ ( η ; 1 ) = ϕ ^ ( η ) , θ ^ ( η ; 1 ) = θ ^ ( η ) , f ^ ( η ; 1 ) = f ^ ( η ) .
Now, by using the Maclaurin expansion for ϕ ^ ( η ; p ) , θ ^ ( η ; p ) , and f ^ ( η ; p ) at p = 0, we have
ϕ ^ ( η ; 0 ) = ϕ ^ 0 ( η ) + m = 1 ϕ ^ m ( η ) p m , θ ^ ( η ; 0 ) = θ ^ 0 ( η ) + m = 1 θ ^ m ( η ) p m , f ^ ( η ; 0 ) = f ^ 0 ( η ) + m = 1 f ^ m ( η ) p m .
Here, we have
ϕ ^ m ( η ) = 1 m ! n ϕ ^ ( η ; p ) p n p = 0 , θ ^ m ( η ) = 1 m ! n θ ^ ( η ; p ) p n p = 0 , f ^ m ( η ) = 1 m ! n f ^ ( η ; p ) p n p = 0 .
The deformation equations that can be of the mth-order are defined as follows:
L ϕ ^ ϕ ^ m ( η ) χ m ϕ ^ m 1 ( η ) = ϕ ^ R m ϕ ^ ( η ) ,
L θ ^ θ ^ m ( η ) χ m θ ^ m 1 ( η ) = θ ^ R m θ ^ ( η ) ,
and
L f ^ f ^ m ( η ) χ m f ^ m 1 ( η ) = f ^ R m f ^ ( η ) .
together with the BCs given by
ϕ ^ ( 0 ) = 1 , f ^ ( 0 ) = R 1 , θ ^ ( 0 ) = 1 , f ^ ( 0 ) = 1 , f ^ ( ) = 1 , θ ^ ( ) = 0 , ϕ ^ ( ) = 0 .

5. Results and Discussion

Here, in this section, we shall discuss the convergence of the above-derived solution. The analytical solution includes the convergence control parameter , which can affect the rate of approximation and the converging regions for the HAM solutions. The constant curve is quite rational whenever convergence takes place at p = 1. The convergence of Equations (41)–(43) is completely particularized by the secondary restriction f , θ , and ϕ . There was a choice to control the series solutions and achieve their convergence. The probabilities of are represented by the curve for the fifth-order computed HAM solution. The efficient regions of are −2.0 < f < 1.0, −4.0 < θ < 1.0 and −3.0 < ϕ < 1.0. The convergence of the homotopy analysis method by the curve was employed for f ( 0 ) , θ ( 0 ) and ϕ ( 0 ) , as demonstrated in Figure 2, Figure 3 and Figure 4.

5.1. Velocity Profile

The important characteristics of the flow profile are graphically displayed in this subsection for certain embedded parameters, such as the magnetic parameter, the inertial parameter, the porosity parameter, and the Grashof number. Figure 5, Figure 6, Figure 7 and Figure 8 represent the impact of f ( η ) , which is displayed, respectively, for varying rates of M, F 1 , k 1 , and G r . Figure 5 characterizes the results of f ( η ) for the dissimilar rate of M when G r = 2.0, k 1 = 0.4, and F 1 = 0.5. From this graph, we predict that f ( η ) is dropped with the rising of M. Physically, the magnetic effects create drag forces, which produce resistance and can slow down the fluid motion, thereby reducing f ( η ) . Figure 6 characterizes the result of f ( η ) for the dissimilar rate of F 1 when G r = 2.0, N t = 0.6, k 1 = 0.4, and M = 0.8. It can be observed from Figure 6 that the f ( η ) curves are reduced against the rising of the values of F 1 . The inertial factor is directly proportional to the drag coefficient and porous medium. Thus, for high values of F 1 , both the drag coefficients and porosity of the medium rise. Consequently, a resisting force is boosted and, therefore, lower velocity is obtained, corresponding to bigger values of F 1 . Figure 7 depicts the variations of f ( η ) for various values of k 1 when G r = 2.0, N b = 0.4, F 1 = 0.5, and M = 0.8. This curve shows that f ( η ) decreases for some rising values of k 1 . Porous medium boosted the resistance to the fluid motion. Since the free convection Reynolds number is not an independent parameter and the heat transfer coefficients are small, a new dimensionless grouping of the Grashof number plays an important role, which incorporates the coefficients of the thermal expansion β 1 in the formulation. Figure 8 depicts the behavior of f ( η ) for varying values of G r when k 1 = 0.4, S c = 0.3, F 1 = 0.5, and M = 0.8. This figure shows that f ( η ) increases for larger rates of G r . The Grashof number is a measurement of the ratio of the buoyancy force to the viscous force operating on the fluid.

5.2. Thermal and Concentration Profiles

The important characteristics of the thermal and concentration properties are graphically displayed in this subsection for certain embedded parameters, such as the Brownian motion parameter, the thermophoresis parameter, the radiation parameter, the heat generation/absorption parameter, the Prandtl number, and the Schmidt number. Figure 9, Figure 10, Figure 11, Figure 12, Figure 13, Figure 14, Figure 15, Figure 16 and Figure 17 show the impact of θ ( η ) and ϕ ( η ) , which are, respectively, plotted for certain values of Δ , N b , N t , R, P r , and S c . The behavior of the heat source ( Δ > 0 ) or the heat sink parameter ( Δ < 0 ) on θ ( η ) has been analyzed in Figure 9 and Figure 10. From these figures, it is recognized that the heat source ( Δ > 0 ) enhanced the heat transfer variation. This enhancement in the temperature of the fluid causes the greatly-induced flow toward the plates through the buoyancy impact. Moreover, the figures show that the heat sink parameter ( Δ < 0 ) declines θ ( η ) . This decline in θ ( η ) causes a reduction in the boundary layer flow, which is the result of the buoyancy impact on the thermal problem. Figure 11 and Figure 12 elaborate on the effects of N b on θ ( η ) when N t = 0.7, P r = 2.5, R = 1.3 and ϕ ( η ) when N t = 0.7 and S c = 0.4. These results recognize that θ ( η ) is enhanced for bigger values of N b , whereas the reverse trend is noted for ϕ ( η ) . Figure 13 and Figure 14 indicate the effects of N t on θ ( η ) whenever N b = 0.6, P r = 2.7, R = 1.6, and ϕ ( η ) whenever N b = 0.6 and S c = 0.4. Due to the enhancement in the thermophoretic parameter, the thermophoresis force is enhanced, which helps the fluid to migrate from the heated location to the cooler one. Thus, clearly, θ ( η ) and ϕ ( η ) are enhanced for higher rates of N t . Figure 15 indicates the behavior of R on θ ( η ) . It is necessary to keep in mind that more of an estimation of R strengthens θ ( η ) because the radiation factor produces thermal energy in the flow regions. Therefore, increments have been observed in θ ( η ) . The behavior of P r on θ ( η ) is exhibited in Figure 16. From this graph, we conclude that an increase in P r declines the θ ( η ) curves. The physical interpretations of the Prandtl number follow its definition as a ratio of the momentum diffusivity to the thermal diffusivity. It is common in the problem of heat transfer that P r is utilized to reduce the relative thickness of the momentum and the thermal boundary layers. Figure 17 indicates that there is a reduction in ϕ ( η ) for S c = 0.5, 1.2, 1.8, and 2.4 when N t = 1.2 and N b = 1.4. Physically, the Schmidt number provides the measurement of the relative effectiveness of the mass and momentum transport by diffusion in the concentration and the velocity boundary layer.

5.3. Table Discussions

Our discussions on the numerical outcomes of S h x , C f , and N u x for different convergence factors, such as F 1 , k 1 , M, G r , R, P r , N t , N b , and S c are exhibited in Table 1, Table 2 and Table 3. Table 1 indicates the numerical outcomes of S h x for N t = 0.2, 0.4, 0.6, 0.8, N b = 0.1, 0.3, 0.5, 0.7, and S c = 0.3, 0.5, 0.7, 0.9. The result of S h x is enhanced for a greater estimation of N t and S c while S h x diminishes for the rising credit of N b . It is clearly depicted from Table 2 that C f boots up for higher rates of F 1 , k 1 and M, whereas the opposite behavior is seen for higher values of G r . The variations of P r , N t , N b , and R upon N u x are discussed in Table 3. It is clearly observed from Table 3 that N u x increases for more increments of R, N b , and N t , while it reduces for higher values of P r . Furthermore, Table 4 represents the comparison of θ ( 0 ) for several values of R 1 and δ 1 by using HAM with the results by Daniel [49] and Chamkha [50]. It can also be observed from Table 4 that the obtained results are in close agreement with those published in the earlier investigations.

6. Concluding Remarks

In this study, we proposed a mathematical model in order to investigate how radiation and buoyancy affect the MHD 2D boundary layer’s fluid flow across a stretchable porous sheet. The flow of fluid is considered laminar, incompressible, and electrically conducted. The behavior of thermophoresis, heat source, sink, and the Brownian motion were systematically investigated in this work. The general form of the buoyancy term in the momentum equation for a free convection boundary layer is derived if the flow is due to temperature variations. The results, which we presented in this work, are shown to match with those in the earlier published studies and are found to be in remarkable agreement. Before solving the modeled problem, similarity variables were utilized to convert the fundamental governing PDEs into the corresponding ODEs. A semi-analytical HAM approach was then applied for solving Equations (27)–(29) with the boundary conditions in Equation (30). The analysis was carried out up to the fifth-order HAM approximation solution.
The core findings of this work are summarized as follows:
  • f ( η ) increased for high values of G r and decreased for the increment values of k 1 , F 1 , and M.
  • θ ( η ) enhances with the rising rates of ( Δ > 0 ) , N b , R, and N t , and diminishes with higher P r and negative values of Δ .
  • For more estimations of N t , there is an increment in the ϕ ( η ) curves.
  • ϕ ( η ) gradually diminishes against higher values of S c and N b .
  • Higher estimations of M, F 1 , and k 1 upsurge C f , whereas the reverse is seen for G r .
  • N u x enhances due to increments of N t , R and N b , while it reduces due to higher P r values.
  • S h x reduces due to greater estimation of N b and decreases for higher values of N t and S c .
  • The role of the Grashof number is the same in free convection as that of the Reynolds number in the forced convection.

Author Contributions

Conceptualization, H.M.S., P.O.M., E.A.-S. and R.J.; data curation, M.J.; formal analysis, M.J.; funding acquisition, E.A.-S. and M.J.; investigation, H.M.S., Z.K., P.O.M., E.A.-S. and R.J.; methodology, Z.K., E.A.-S. and M.J.; project administration, H.M.S., P.O.M. and R.J.; resources, Z.K. and R.J.; software, Z.K. and P.O.M.; supervision, H.M.S., P.O.M. and R.J.; validation, M.J.; visualization, R.J.; writing—original draft, H.M.S., Z.K. and P.O.M.; writing—review and editing, E.A.-S. and M.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

B 0 ( x ) applied magnetic induction
a 1 stretching rate
N t thermophoresis parameter
v, uvelocity components along y and x direction (ms 1 )
Tfluid temperature (K)
β 1 thermal expansion
η similarity variable
ψ stream function (m 2 s 1 )
T free stream temperature (K)
ρ fluid density (kgm 3 )
g 1 acceleration due to gravity
σ 1 * Boltzmann constant (Wm 2 K 4 )
c p specific heat (Jm 3 K 1 )
σ electrical conductivity
δ 1 heat generation/absorption
α thermal diffusivity
Mmagnetic parameter
k 1 * mean absorption coefficient
T w      wall temperature (K)
k 1 porosity parameter
Rradiation parameter
υ kinematic viscosity (m 2 s 1 )
convergence control parameter
F 1 inertia parameter
Δ heat source/sink parameter
R e Reynolds number
G r Grashof number
L 1 auxiliary linear-operator
S c Schmidt number
q r heat flux
P r Prandtl number
R 1 suction/injection parameter
N 1 non-linear operator
N b Brownian motion parameter

Appendix A. Derivation of the Flow Problem

The flow problem has the following governing equations and boundary conditions:
Continuity equation:
u x + v y = 0
Momentum equation:
u u x + v u y = υ 2 u y 2 σ B 0 2 ( x ) ρ u + g 1 β 1 ( T T ) u υ k 0 F 0 u 2
Energy equation:
u T x + v T y = α 2 T y 2 + Q 1 ρ c p ( T T ) + β 1 * u ρ c p ( T T ) + 1 c p ρ q r y + τ D B C y T y + D T T T y 2
Concentration equation:
u C x + v C y = 2 C y 2 D B + 2 T y 2 D T T .
Boundary constraints:
u = a 1 x , v = v w , T = T w = T + A 0 x C = C w = B 0 x + C a t y = 0 u 0 , C C , T T a s y .
The dimensionless variables are given as
η = y a 1 υ , θ = T T T w T , ψ ( x , y ) = υ a 1 x f ( η ) , ϕ = C C C w C .

Appendix A.1. Derivation of the Continuity Equation

From
u = ψ y = y ( υ a 1 x f ( η ) ) = a 1 x f ( η )
and
v = ψ x = υ a 1 f ( η ) ,
now
u x = x ( a 1 x f ( η ) ) = a 1 f ( η )
v y = y ( υ a 1 f ( η ) ) = a 1 f ( η ) .
Substitute Equations (A9) and (A10) into Equation (A1), we have
u x + v y = a 1 f ( η ) a 1 f ( η ) = 0 .
Thus, it satisfies the continuity equation.

Appendix A.2. Derivation of the Momentum Equation

u u x = a 1 2 x f 2
v u y = a 1 2 x f f
υ 2 u y 2 = a 1 2 x f
σ B 0 2 ( x ) ρ u = σ B 0 2 ( x ) ρ a 1 x f ( η )
υ k 0 u = υ k 0 a 1 x f ( η )
F 0 u 2 = F 0 a 1 2 x 2 f 2 .
Substituting Equation (A12) to Equation (A17) into Equation (A2), we have
f + f 2 + θ G r + f f M 2 f f 2 F 1 + f k 1 = 0 ,

Appendix A.3. Derivation of the Energy Equation

u T x = 0
v T y = a 1 ( T w T ) f θ
α 2 T y 2 = α ( T w T ) θ a 1 υ
Q 1 ρ c p ( T T ) = Q 1 ρ c p ( T w T ) θ
1 c p ρ 16 σ 1 * T 3 3 k 1 * 2 T y 2 = 1 c p ρ 16 σ 1 * T 3 3 k 1 * ( T w T ) θ a 1 υ
Substituting Equation (A19) to Equation (A23) into Equation (A3), we have
1 P r ( 1 + 4 R 3 ) θ + ϕ 2 N t + θ f + θ ϕ N b + Δ θ δ 1 θ f = 0 ,

Appendix A.4. Derivation of Concentration Equation

u C x = 0
v C y = a 1 ( C w T ) f ϕ
D B 2 C y 2 = D B ( C w T ) ϕ a 1 υ
D T T 2 T y 2 = D T T ( T w T ) θ a 1 υ .
Substituting Equation (A25) to Equation (A28) into Equation (A4), we have
ϕ + ϕ f S c + θ N t N b = 0 .

Appendix A.5. Derivation of Boundary Conditions

The derivation of boundary conditions at η = 0 and η = is as follows:
u = a 1 x
a 1 x f ( η ) = a 1 x
f ( 0 ) = 1
u = 0 a s y
a 1 x f ( ) = 0
f ( ) = 0
v = v w
υ a 1 f ( η ) = R 1 υ a 1
f ( 0 ) = R 1
T = T a s y
T T = 0
( T w T ) θ = 0
θ ( ) = 0
T = T + A 1 x
T T = T w T x x
( T w T ) θ ( 0 ) = T w T
θ ( 0 ) = 1
C = C a s y
ϕ ( ) = 0
C C = B 0 x
C C = ( C w C ) x x
( C w C ) ϕ ( 0 ) = ( C w C )
ϕ ( 0 ) = 1

References

  1. Chaudhary, S.; Kumar, P. Unsteady magnetohydrodynamic boundary layer flow near the stagnation point towards a shrinking surface. J. Appl. Math. Phys. 2015, 3, 921–930. [Google Scholar] [CrossRef] [Green Version]
  2. Naramgari, S.; Sulochana, C. MHD flow over a permeable stretching/shrinking sheet of a nanofluid with suction/injection. Alex. Eng. J. 2016, 55, 819–827. [Google Scholar] [CrossRef] [Green Version]
  3. Reddy, G.V.R.; Reddy, N.B.; Gorla, R.S.R. Radiation and chemical reaction effects on MHD flow along a moving vertical porous plate. Int. J. Appl. Mech. Eng. 2016, 21, 157–168. [Google Scholar] [CrossRef] [Green Version]
  4. Mishra, S.R.; Jena, S. Numerical solution of boundary layer MHD flow with viscous dissipation. Sci. World J. 2014, 2014, 756498. [Google Scholar] [CrossRef]
  5. Babu, P.R.; Rao, J.A.; Sheri, S. Radiation effect on MHD heat and mass transfer flow over a shrinking sheet with mass suction. J. Appl. Fluid Mech. 2014, 7, 641–650. [Google Scholar] [CrossRef]
  6. Mabood, F.; Khan, W.A.; Ismail, A.M. MHD flow over exponential radiating stretching sheet using homotopy analysis method. J. King Saud Univ. Eng. Sci. 2017, 29, 68–74. [Google Scholar] [CrossRef]
  7. Pal, D.; Mondal, H. Effect of variable viscosity on MHD non-Darcy mixed convective heat transfer over a stretching sheet embedded in a porous medium with non-uniform heat source/sink. Commun. Nonlinear. Sci. Numer. Simul. 2010, 15, 1553–1564. [Google Scholar] [CrossRef]
  8. Zeeshan, A.; Ellahi, R.; Hassan, M. Magnetohydrodynamic flow of water/ethylene glycol based nanofluids with natural convection through a porous medium. Eur. Phys. J. Plus 2014, 129, 261. [Google Scholar] [CrossRef]
  9. Mahmoud, M.A.A. Variable viscosity effects on hydromagnetic boundary. Appl. Math. Sci. 2017, 1, 799–814. [Google Scholar]
  10. Kishore, P.M.; Rajesh, V.; Verma, V.S. The effects of thermal radiation and viscous dissipation on MHD heat and mass diffusion flow past an oscillating vertical plate embedded in a porous medium with variable surface conditions. Theor. Appl. Mech. 2012, 39, 99–125. [Google Scholar] [CrossRef]
  11. Majeed, A.; Noori, F.M.; Zeeshan, A.; Mahmood, T.; Rehman, S.U.; Khan, I. Analysis of activation energy in magnetohydrodynamic flow with chemical reaction and second order momentum slip model. Case Stud. Therm. Eng. 2018, 12, 765–773. [Google Scholar] [CrossRef]
  12. Sharma, R.; Ishak, A.; Nazar, R.; Pop, I. Boundary layer flow and heat transfer over a permeable exponentially shrinking sheet in the presence of thermal radiation and partial slip. J. Appl. Fluid Mech. 2014, 7, 125–134. [Google Scholar] [CrossRef]
  13. Poornima, T.; Reddy, N.B. Radiation effects on mhd free convective boundary layer flow of nanofluids over a nonlinear stretching sheet. Adv. Appl. Sci. Res. 2013, 4, 190–202. [Google Scholar]
  14. Singh, P.; Jangid, A.; Tomer, N.S.; Sinha, D. Effects of thermal radiation and magnetic field on unsteady stretching permeable sheet in presence of free stream velocity. Int. J. Phys. Math. Sci. 2010, 4, 396–402. [Google Scholar]
  15. Ahmad, I.; Sajid, M.; Awan, W.; Rafique, M.; Aziz, W.; Ahmed, M.; Abbasi, A.; Taj, M. MHD flow of a viscous fluid over an exponentially stretching sheet in a porous medium. J. Appl. Math. 2014, 2014, 256761. [Google Scholar] [CrossRef] [Green Version]
  16. Shateyi, S.; Motsa, S.S. Thermal radiation effects on heat and mass transfer over an unsteady stretching surface. Math. Probl. Eng. 2009, 2009, 965603. [Google Scholar] [CrossRef] [Green Version]
  17. Waheed, S.E. Flow and heat transfer in a maxwell liquid film over an unsteady stretching sheet in a porous medium with radiation. SpringerPlus 2016, 5, 1061. [Google Scholar] [CrossRef] [Green Version]
  18. Khan, A.S.; Nie, Y.; Shah, Z. Impact of thermal radiation on magnetohydrodynamic unsteady thin film flow of Sisko fluid over a stretching surface. Process 2016, 7, 369. [Google Scholar] [CrossRef] [Green Version]
  19. Metri, P.G.; Tawade, J.; Abel, M.S. Thin film flow and heat transfer over an unsteady stretching sheet with thermal radiation. arXiv 2016, arXiv:1603.03664. [Google Scholar]
  20. Khan, Z.; Srivastava, H.M.; Mohammed, P.O.; Jawad, M.; Jan, R.; Nonlaopon, K. Thermal boundary layer analysis of MHD nanofluids across a thin needle using non-linear thermal radiation. Math. Biosci. Eng. 2022, 19, 14116–14141. [Google Scholar] [CrossRef]
  21. Garia, R.; Rawat, S.K.; Kumar, M.; Yaseen, M. Hybrid nanofluid flow over two different geometries with Cattaneo–Christov heat flux model and heat generation: A model with correlation coefficient and probable error. Chin. J. Phys. 2021, 74, 421–439. [Google Scholar] [CrossRef]
  22. Garia, R.; Rawat, S.K.; Kumar, M.; Yaseen, M. Cattaneo–Christov heat flux model in Darcy–Forchheimer radiative flow of MoS2–SiO2/kerosene oil between two parallel rotating disks. J. Therm. Anal. Calorim. 2022, 147, 10865–10887. [Google Scholar] [CrossRef]
  23. Khan, Z.; Jawad, M.; Bonyah, E.; Khan, N.; Jan, R. Magnetohydrodynamic thin film flow through a porous stretching sheet with the impact of thermal radiation and viscous dissipation. Math. Probl. Eng. 2022, 2022, 1086847. [Google Scholar] [CrossRef]
  24. Jawad, M.; Khan, Z.; Bonyah, E.; Jan, R. Analysis of hybrid nanofluid stagnation point flow over a stretching surface with melting heat transfer. Math. Probl. Eng. 2022, 2022, 9469164. [Google Scholar] [CrossRef]
  25. Dullien, F.A.L. Porous Media: Fluid Transport and Pore Structure, 2nd ed.; Academic Press: New York, NY, USA; London, UK; San Diego, CA, USA, 2012. [Google Scholar]
  26. Nield, D.A.; Bejan, A. Convection in Porous Media, 5th ed.; Springer: Berlin/Heidelberg, Germany; New York, NY, USA, 2006. [Google Scholar] [CrossRef]
  27. Karniadakis, G.E.M.; Beskok, A.; Gad-el-Hak, M. Micro flows: Fundamentals and simulation. Appl. Mech. Rev. 2002, 55, B76. [Google Scholar] [CrossRef]
  28. Hussain, A.; Muneer, Z.; Malik, M.Y.; Ali, S. Formulating the behavior of thermal radiation and magnetic dipole effects on Darcy-Forchheimer grasped ferrofluid flow. Canad. J. Phys. 2019, 97, 938–949. [Google Scholar] [CrossRef]
  29. Forchheimer, P. Wasserbewegung durch boden. Z. Ver. Deutsch. Ing. 1901, 45, 1782–1788. [Google Scholar]
  30. Pal, D.; Mondal, H. Hydromagnetic convective diffusion of species in Darcy—Forchheimer porous medium with non-uniform heat source/sink and variable viscosity. Int. Commun. Heat Mass Transf. 2012, 39, 913–917. [Google Scholar] [CrossRef]
  31. Ganesh, N.V.; Hakeem, A.K.A.; Ganga, B. Darcy–Forchheimer flow of hydromagnetic nanofluid over a stretching/shrinking sheet in a thermally stratified porous medium with second order slip, viscous and Ohmic dissipations effects. Ain Shams Eng. J. 2018, 9, 939–951. [Google Scholar] [CrossRef] [Green Version]
  32. Seth, G.S.; Mandal, P.K. Hydromagnetic rotating flow of Casson fluid in Darcy-Forchheimer porous medium. MATEC Web Conf. 2018, 192, 02059. [Google Scholar] [CrossRef] [Green Version]
  33. Seddeek, M.A. Influence of viscous dissipation and thermophoresis on Darcy–Forchheimer mixed convection in a fluid saturated porous media. J. Colloid Interface Sci. 2006, 293, 137–142. [Google Scholar] [CrossRef]
  34. Hayat, T.; Muhammad, T.; Mezal, S.A.; Liao, S.J. Darcy-forchheimer flow with variable thermal conductivity and Cattaneo–Christov heat flux. Int. J. Numer. Methods Heat Fluid Flow 2006, 26, 2355–2369. [Google Scholar] [CrossRef]
  35. Rajesh, V.; Sheremet, M.A.; Öztop, H.F. Impact of hybrid nanofluids on MHD flow and heat transfer near a vertical plate with ramped wall temperature. Case Stud. Therm. Eng. 2021, 28, 101557. [Google Scholar] [CrossRef]
  36. Yaseen, M.; Rawat, S.K.; Shafiq, A.; Kumar, M.; Nonlaopon, K. Analysis of heat transfer of mono and hybrid nanofluid flow between two parallel plates in a Darcy porous medium with thermal radiation and heat generation/absorption. Symmetry 2022, 14, 1943. [Google Scholar] [CrossRef]
  37. Raza, R.; Naz, R.; Abdelsalam, S.I. Microorganisms swimming through radiative Sutterby nanofluid over stretchable cylinder: Hydrodynamic effect. Numer. Methods Partial. Differ. Equ. 2022, 1–20. [Google Scholar] [CrossRef]
  38. Faizan, M.; Ali, F.; Loganathan, K.; Zaib, A.; Reddy, C.A.; Abdelsalam, S.I. Entropy Analysis of Sutterby Nanofluid Flow over a Riga Sheet with Gyrotactic Microorganisms and Cattaneo–Christov Double Diffusion. Mathematics 2022, 10, 3157. [Google Scholar] [CrossRef]
  39. Bhattacharyya, K. Effects of radiation and heat source/sink on unsteady MHD boundary layer flow and heat transfer over a shrinking sheet with suction/injection. Front. Chem. Sci. Eng. 2011, 5, 376–384. [Google Scholar] [CrossRef]
  40. Mukhopadhyay, S. Slip effects on MHD boundary layer flow over an exponentially stretching sheet with suction/blowing and thermal radiation. Ain Shams Eng. J. 2013, 4, 485–491. [Google Scholar] [CrossRef] [Green Version]
  41. Marin, M.; Öchsner, A. The effect of a dipolar structure on the Hölder stability in Green-Naghdi thermoelasticity. Contin. Mech. Thermodyn. 2017, 29, 1365–1374. [Google Scholar] [CrossRef]
  42. Marin, M. Cesaro means in thermoelasticity of dipolar bodies. Acta Mech. 1997, 122, 155–168. [Google Scholar] [CrossRef]
  43. Liao, S. The Proposed Homotopy Analysis Technique for the Solution of Nonlinear Problems. Ph.D. Thesis, Shanghai Jiao Tong University, Shanghai, China, 1992. [Google Scholar]
  44. Liao, S. A kind of approximate solution technique which does not depend upon small parameters. II. An application in fluid mechanics. Int. J. Nonlinear Mech. 1997, 32, 815–822. [Google Scholar] [CrossRef]
  45. Liao, S. On the analytic solution of magnetohydrodynamic flows of non-Newtonian fluids over a stretching sheet. J. Fluid Mech. 2003, 488, 189–212. [Google Scholar] [CrossRef] [Green Version]
  46. Liao, S. A new branch of solutions of boundary-layer flows over an impermeable stretched plate. Int. J. Heat Mass Transf. 2005, 48, 2529–2539. [Google Scholar] [CrossRef]
  47. Liao, S. An optimal homotopy-analysis approach for strongly nonlinear differential equations. Commun. Nonlinear Sci. Numer. Simul. 2010, 15, 2003–2016. [Google Scholar] [CrossRef]
  48. Liao, S.; Tan, Y. A general approach to obtain series solutions of nonlinear differential equations. Stud. Appl. Math. 2007, 119, 297–354. [Google Scholar] [CrossRef]
  49. Daniel, Y.S.; Daniel, S.K. Effects of buoyancy and thermal radiation on MHD flow over a stretching porous sheet using homotopy analysis method. Alex. Eng. J. 2015, 54, 705–712. [Google Scholar] [CrossRef] [Green Version]
  50. Chamkha, A.J. Thermal radiation and buoyancy effects on hydromagnetic flow over an accelerating permeable surface with heat source or sink. Int. J. Eng. Sci. 2000, 38, 1699–1712. [Google Scholar] [CrossRef]
Figure 1. Geometry of the physical flow problem.
Figure 1. Geometry of the physical flow problem.
Energies 16 00058 g001
Figure 2. Outcomes of f for f ( 0 ) .
Figure 2. Outcomes of f for f ( 0 ) .
Energies 16 00058 g002
Figure 3. Outcomes of θ for θ ( 0 ) .
Figure 3. Outcomes of θ for θ ( 0 ) .
Energies 16 00058 g003
Figure 4. Outcomes of ϕ for ϕ ( 0 ) .
Figure 4. Outcomes of ϕ for ϕ ( 0 ) .
Energies 16 00058 g004
Figure 5. Outcomes of f ( η ) for M where blue, purple, green, and red lines represent for M = 1.1, M = 1.6, M = 2.2, and M = 2.6 respectively.
Figure 5. Outcomes of f ( η ) for M where blue, purple, green, and red lines represent for M = 1.1, M = 1.6, M = 2.2, and M = 2.6 respectively.
Energies 16 00058 g005
Figure 6. Outcomes of f ( η ) for F 1 where blue, purple, green, and red lines represent for F 1 = 1.0, F 1 = 2.0, F 1 = 3.0, and F 1 = 4.0 respectively.
Figure 6. Outcomes of f ( η ) for F 1 where blue, purple, green, and red lines represent for F 1 = 1.0, F 1 = 2.0, F 1 = 3.0, and F 1 = 4.0 respectively.
Energies 16 00058 g006
Figure 7. Outcomes of f ( η ) for k 1 where blue, purple, green, and red lines represent for k 1 = 0.1, k 1 = 0.4, k 1 = 0.7, and k 1 = 1.0 respectively.
Figure 7. Outcomes of f ( η ) for k 1 where blue, purple, green, and red lines represent for k 1 = 0.1, k 1 = 0.4, k 1 = 0.7, and k 1 = 1.0 respectively.
Energies 16 00058 g007
Figure 8. Outcome of f ( η ) for G r where blue, purple, green, and red lines represent for G r = 1.0, G r = 2.0, G r = 3.0, and G r = 4.0 respectively.
Figure 8. Outcome of f ( η ) for G r where blue, purple, green, and red lines represent for G r = 1.0, G r = 2.0, G r = 3.0, and G r = 4.0 respectively.
Energies 16 00058 g008
Figure 9. Outcomes of θ ( η ) for Δ > 0 where yellow, magenta, green, and red lines represent for Δ = 0.5, Δ = 1.0, Δ = 1.5, and Δ = 2.0 respectively.
Figure 9. Outcomes of θ ( η ) for Δ > 0 where yellow, magenta, green, and red lines represent for Δ = 0.5, Δ = 1.0, Δ = 1.5, and Δ = 2.0 respectively.
Energies 16 00058 g009
Figure 10. Outcomes of θ ( η ) for Δ < 0 where yellow, magenta, green, and red lines represent for Δ = −0.5, Δ = −1.0, Δ = −1.5, and Δ = −2.0 respectively.
Figure 10. Outcomes of θ ( η ) for Δ < 0 where yellow, magenta, green, and red lines represent for Δ = −0.5, Δ = −1.0, Δ = −1.5, and Δ = −2.0 respectively.
Energies 16 00058 g010
Figure 11. Outcomes of θ ( η ) for N b where yellow, magenta, green, and red lines represent for N b = 0.1, N b = 0.2, N b = 0.3, and N b = 0.4 respectively.
Figure 11. Outcomes of θ ( η ) for N b where yellow, magenta, green, and red lines represent for N b = 0.1, N b = 0.2, N b = 0.3, and N b = 0.4 respectively.
Energies 16 00058 g011
Figure 12. Outcomes of ϕ ( η ) for N b where yellow, magenta, green, and red lines represent for N b = 0.1, N b = 0.2, N b = 0.3, and N b = 0.4 respectively.
Figure 12. Outcomes of ϕ ( η ) for N b where yellow, magenta, green, and red lines represent for N b = 0.1, N b = 0.2, N b = 0.3, and N b = 0.4 respectively.
Energies 16 00058 g012
Figure 13. Outcomes of θ ( η ) for N t where yellow, magenta, green, and red lines represent for N t = 0.01, N t = 0.09, N t = 0.3, and N t = 0.5 respectively.
Figure 13. Outcomes of θ ( η ) for N t where yellow, magenta, green, and red lines represent for N t = 0.01, N t = 0.09, N t = 0.3, and N t = 0.5 respectively.
Energies 16 00058 g013
Figure 14. Outcomes of ϕ ( η ) for N t where yellow, magenta, green, and red lines represent for N t = 3.0, N t = 3.3, N t = 3.6, and N t = 3.9 respectively.
Figure 14. Outcomes of ϕ ( η ) for N t where yellow, magenta, green, and red lines represent for N t = 3.0, N t = 3.3, N t = 3.6, and N t = 3.9 respectively.
Energies 16 00058 g014
Figure 15. Outcomes of θ ( η ) for R where yellow, magenta, green, and red lines represent for R = 0.1, R = 0.4, R = 0.7, and R = 1.0 respectively.
Figure 15. Outcomes of θ ( η ) for R where yellow, magenta, green, and red lines represent for R = 0.1, R = 0.4, R = 0.7, and R = 1.0 respectively.
Energies 16 00058 g015
Figure 16. Outcomes of θ ( η ) for P r where yellow, magenta, green, and red lines represent for P r = 1.5, P r = 2.5, P r = 3.5, and P r = 4.5 respectively.
Figure 16. Outcomes of θ ( η ) for P r where yellow, magenta, green, and red lines represent for P r = 1.5, P r = 2.5, P r = 3.5, and P r = 4.5 respectively.
Energies 16 00058 g016
Figure 17. Outcomes of ϕ ( η ) for S c where yellow, magenta, green, and red lines represent for S c = 0.5, S c = 1.2, S c = 1.8, and S c = 2.4 respectively.
Figure 17. Outcomes of ϕ ( η ) for S c where yellow, magenta, green, and red lines represent for S c = 0.5, S c = 1.2, S c = 1.8, and S c = 2.4 respectively.
Energies 16 00058 g017
Table 1. Numerical outcomes of S h x for N t , N b and S c .
Table 1. Numerical outcomes of S h x for N t , N b and S c .
N t N b Sc Sh x
0.20.10.32.45038307
0.4 2.36012785
0.6 2.34056085
0.8 2.31059187
0.1 2.26205129
0.3 2.27140738
0.5 2.29025809
0.7 2.29506702
0.32.46230876
0.52.45037182
0.72.44570887
0.92.42071925
Table 2. Numerical outcomes of C f for k 1 , M, G r , and F 1 .
Table 2. Numerical outcomes of C f for k 1 , M, G r , and F 1 .
F 1 k 1 M G r C f
0.30.10.31.01.63205892
0.4 1.64218038
0.5 1.64575398
0.6 1.66094124
0.1 1.23056802
0.2 1.24059735
0.3 1.25192114
0.4 1.25602565
0.3 1.43271729
0.5 1.45952172
0.8 1.47062175
1.0 1.48073426
1.02.03968538
2.01.90250917
3.01.86620846
4.01.80911127
Table 3. Numerical outcomes of N u x for R, P r , N t , and N b .
Table 3. Numerical outcomes of N u x for R, P r , N t , and N b .
R Pr N t N b Nu x
0.52.50.10.22.53802156
1.0 2.52508295
1.2 2.51862063
1.5 2.50752047
2.5 1.54531398
3.5 1.59370589
4.5 1.64950842
5.5 1.66097014
0.2 1.10459009
0.4 1.02054179
0.6 0.85628206
0.8 0.80049531
0.31.27353338
0.51.25904107
0.71.23718056
0.91.21562169
Table 4. Comparison of θ ( 0 ) for R 1 and δ 1 .
Table 4. Comparison of θ ( 0 ) for R 1 and δ 1 .
R 1 δ 1 Daniel [49]Chamkha [50]Present Study
00.50.947650.947690.94661
01.01.078951.079961.07713
0.450.50.823960.823970.82238
0.451.00.961900.961910.96052
−1.50.51.570771.570771.57061
−1.51.01.661821.661841.66047
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Srivastava, H.M.; Khan, Z.; Mohammed, P.O.; Al-Sarairah, E.; Jawad, M.; Jan, R. Heat Transfer of Buoyancy and Radiation on the Free Convection Boundary Layer MHD Flow across a Stretchable Porous Sheet. Energies 2023, 16, 58. https://doi.org/10.3390/en16010058

AMA Style

Srivastava HM, Khan Z, Mohammed PO, Al-Sarairah E, Jawad M, Jan R. Heat Transfer of Buoyancy and Radiation on the Free Convection Boundary Layer MHD Flow across a Stretchable Porous Sheet. Energies. 2023; 16(1):58. https://doi.org/10.3390/en16010058

Chicago/Turabian Style

Srivastava, Hari Mohan, Ziad Khan, Pshtiwan Othman Mohammed, Eman Al-Sarairah, Muhammad Jawad, and Rashid Jan. 2023. "Heat Transfer of Buoyancy and Radiation on the Free Convection Boundary Layer MHD Flow across a Stretchable Porous Sheet" Energies 16, no. 1: 58. https://doi.org/10.3390/en16010058

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop