Next Article in Journal
Development of an Inkjet Setup for Printing and Monitoring Microdroplets
Previous Article in Journal
Dynamic Behaviours of Monodisperse Double Emulsion Formation in a Tri-Axial Capillary Device
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

MHD Dissipative Williamson Nanofluid Flow with Chemical Reaction Due to a Slippery Elastic Sheet Which Was Contained within a Porous Medium

1
Department of Mathematics, Faculty of Science, University of Tabuk, Tabuk 71491, Saudi Arabia
2
Department of Mathematics, Faculty of Science, Benha University, Benha 13518, Egypt
*
Author to whom correspondence should be addressed.
Micromachines 2022, 13(11), 1879; https://doi.org/10.3390/mi13111879
Submission received: 27 September 2022 / Revised: 15 October 2022 / Accepted: 24 October 2022 / Published: 31 October 2022

Abstract

:
The reasons why the model of non-Newtonian nanofluids is more applicable than other models, particularly those that take the porous medium into account, are studied here. Thus, we looked at the heat and mass transfer features of a non-Newtonian Williamson nanofluid flow due to a stretched sheet under the impact of chemical reactions, slip velocity, viscous dissipation, and the magnetic field in this article. The main focus is on a situation in which the properties of Williamson nanofluid, such as viscosity and thermal conductivity, change with temperature. After utilizing the shooting technique, a numerical solution to the suggested problem is provided. As a result, several graphs have been drawn to highlight how various physical characteristics that arise in the problems affect velocity, temperature, and concentration profiles. It was discovered that the heat and mass transmission processes are affected by the viscous dissipation phenomena, the slip velocity assumption, and the magnetic field. Theoretical and numerical results show a high level of qualitative agreement.

1. Introduction

Nanofluids have gained a lot of attention in recent decades due to their claimed unique features and potential applications. When nanoparticles are present, the thermophysical and transport properties of the base fluid are usually altered. The term “nanofluid” refers to a colloidal dispersion of nanoparticles in a base fluid [1]. Rapid technological advancement necessitates the employment of a high-performing coolant in many thermal management activities. Nanofluids have crucial qualities that can not be overlooked when it comes to thermal regulation [2,3]. They also offer a significant potential for improving heat transfer rates in engineering systems, particularly for cooling electronic equipment [4,5,6,7]. The inclusion of nanoparticles in the base fluid, as many authors [8,9,10,11,12,13,14,15,16] have already said, has a considerable impact on the fluid’s physical characteristics and significantly increases the heat transmission mechanism. The Buongiorno nanofluid model [17], Williamson nanofluid model [18], micropolar model [19], and Casson-Williamson nanofluid model [20] have all been proposed to characterize the behaviour of nanofluids. It is also worth noting that MHD nanofluid flow research is crucial because of its essential industrial applications [21,22,23]. On the other hand, many research studies have concentrated on the chemical reaction phenomenon in nanofluid flow in conjunction with magnetic field influence [24], while others have focused on the chemical reaction phenomenon in a porous media [25] and its effect on nanofluid flow in a range of geometries.
The viscous dissipation phenomena is another essential aspect that should not be overlooked when studying nanofluid flow. Internal friction in a viscous flow causes an irreversible conversion of kinetic energy to thermal energy, which is known as the irreversible process of mechanical energy conversion to thermal energy. This important phenomenon only occurs if the fluid has a high viscosity and velocity [26,27,28]. The slip velocity is another important factor that influences the flow behavior of nanofluids. This phenomenon can occur in a variety of physical conditions, including internal cavity polishing, artificial heart valves, and micro electronics [29,30,31]. After reviewing the literature and being inspired by the prominent features of viscous dissipation, variable fluid properties, magnetic field, chemical reaction, and slip velocity phenomena, it was determined that non-Newtonian Williamson nanofluid flow in these significant physical situations has yet to be investigated. Due to the significance of this kind of fluid, the novelty of the current research is based on examining the problem of the flow and heat transmission mechanisms caused by a stretching sheet embedded in a porous medium while taking slip velocity and the impacts of chemical reactions into consideration.

2. Description of the Problem

A two-dimensional non-Newtonian Williamson nanofluid, which is characterized by the Williamson parameter Γ , will flow across an impermeable horizontal heated linearly stretched sheet immersed in a porous liquid in the current theoretical investigation. The nanoparticles in the nanofluid are assumed to circulate due to the Brownian diffusion coefficient D B . In addition, the chemical conversion rate K is supposed to be applied to the nanofluid particles. The viscosity μ and thermal conductivity κ of the Williamson nanofluid are both assumed to be temperature functions in these formulae μ = μ e α θ and κ = κ ( 1 + ε θ ) [32], but its density ρ is assumed to be constant. where μ is a constant viscosity of the nanofluid at the ambient, α is the viscosity parameter, κ and ε are constant thermal conductivity and the thermal conductivity parameter conductivity, respectively. T w and C w are the temperature and concentration near the wall, while T and C are the temperature and concentration far away from the wall, as shown in Figure 1.
In addition, T denotes the temperature of the nanofluid, C the concentration of the nanofluid, and u and v the velocity components of the nanofluid. The problem is modelled using the following governing equations that follow the above-mentioned nanofluid flow conditions [33,34]:
u x + v y = 0 ,
u u x + v u y = 1 ρ y μ u y + μ Γ 2 u y 2 σ B 0 2 ρ u μ ρ k u ,
u T x + v T y = 1 ρ c p y ( κ T y ) + τ D B C y T y + D T T T y 2 + μ ρ c p u y 2 + Γ 2 u y 3 ,
u C x + v C y = D B 2 C y 2 K C C + D T T 2 T y 2 .
As a result, the flow field is governed by the boundary conditions listed below [35]:
u = a x + λ 1 μ μ u y + μ Γ 2 u y 2 , v = 0 , T = T w , C = C w at y = 0 ,
u 0 , T T , C C as y ,
where k is the porous medium’s porosity, σ denotes the electrical conductivity of the nanofluid, B 0 is the magnetic field strength, λ 1 is the factor of the slip velocity and D T is the thermophoretic diffusion coefficient. Now, we begin the dimensionless quantities in the following forms before creating the solution algorithm:
η = a ν 1 2 y , u = a x f ( η ) , v = a ν 1 2 f ( η ) ,
θ ( η ) = T T T w T , ϕ ( η ) = C C C w C .
Equation (1) becomes easily and quickly satisfied after applying these transformations (7) and (8), whereas Equations (2)–(4) with boundary conditions (5) and (6) become
1 + W e f f α θ f 1 + W e 2 f e α θ + f f f 2 M f δ e α θ f = 0 ,
1 P r ( 1 + ε θ ) θ + ε θ 2 + N b θ ϕ + N t θ 2 + f θ + E c f 2 + W e 2 f 3 e α θ = 0 ,
ϕ + S c f ϕ S c G ϕ + N t N b θ = 0 .
In addition, the boundary conditions become:
f = 0 , f = 1 + λ f + W e 2 f 2 e α θ , θ = 1 , ϕ = 1 , a t η = 0 ,
f 0 , θ 0 , ϕ 0 , as η .
The velocity field, as well as the associated condition, is clearly regulated by the local Weissenberg number W e , the magnetic parameter M, the porosity parameter δ and the slip velocity parameter λ , which respectively can be defined as:
W e = Γ x 2 a 3 ν , M = σ B 0 2 a ρ , δ = ν k a , λ = λ 1 a ν .
In addition, governing the temperature field are the Prandtl number P r , the thermophoresis parameter N t , the local Eckert number E c , and the Brownian motion parameter N b , which are defined as follows:
P r = μ c p κ , N t = τ D T ( T w T ) T ν , E c = ( a x ) 2 c p ( T w T ) , N b = τ D B ( C w C ) ν .
Finally, the concentration field is influenced by both basic factors, the chemical reaction parameter G, and the Schmidt number S c , which have the following definitions:
G = K a , S c = ν D B .
As we will see, in this case, the local parameters depending on the length scale x are the Williamson parameter W e and the Eckert number E c . These parameters are a function of x and their changes occur locally throughout the flow action, hence it is important to note that the presented equation is only valid for a locally similar solution [36]. We have provided precise values of W e and E c for the graphical findings, which relate to the flow study at a specific value of x and for all values of y. The following relations may now be achieved utilizing the above dimensionless transformations and well-defined relationships for wall shear stress C f , local Nusselt number N u x , and local Sherwood number S h x :
C f R e x = f ( 0 ) + W e 2 f 2 ( 0 ) e α θ ( 0 ) , N u x R e x = θ ( 0 ) , S h x R e x = ϕ ( 0 ) ,
where R e x = u w x ν is the local Reynolds number.

3. Results and Discussion

To provide a numerical solution for the suggested problem in this study, we used the shooting approach. The boundary value problem is reduced to identifying the initial conditions that produce a root by using the shooting technique. The shooting approach has the advantages of employing the speed and adaptability of approaches for initial value problems. Additionally, compared to other methods, this method significantly improves numerical stability and nonlinearity distribution. We set up a comparison table at the beginning of this section, which is shown in Table 1, to verify our computations, which were made using the shooting method, with previous results. Clearly, the findings are very similar to the previously published research of Khan and Pop [37], implying that the numerical method presented here is very suitable for this type of model.
The major goal of this part is to use graphical figures to illustrate the physics of the numerical non-Newtonian Williamson nanofluid model. The following figures show the impacts of various flow parameters on nanofluid flow behavior and heat mass characteristics, such as the magnetic parameter M, the porosity parameter δ , the slip velocity parameter λ , the thermophoresis parameter N t , the Eckert number E c , the Brownian motion parameter N b , and the chemical reaction parameter G. Figure 2 shows how varying values of the magnetic parameter M affect velocity, temperature, and nanoparticle concentration. It has been demonstrated that boosting the magnetic field parameter M lessens both the momentum thickness and the velocity distribution, whereas the temperature and concentration distributions have the opposite effect. Physically, the magnetic field restricts the boundary layer with a resistive-type force. The only purpose of this force is to reduce the nanofluid’s speed. As a result, the nanofluid consequently absorbs additional heat energy from the same force.
Figure 3 depicts the effect of porosity parameter δ alteration on velocity, temperature, and nanoparticle concentration field. The figure indicates an enhanced tendency in both the temperature and concentration fields as the porosity parameter is increased, whereas the velocity field displays the opposite trend. Physically, increasing the porosity parameter leads to impeding the acceleration of the nanofluid flow, which in turn produces a reduction in the flow’s speed and an increase in the friction between the fluid particles, which raises the temperature field.
Figure 4 shows the velocity of the nanofluid, its temperature, and the concentration of nanoparticles vs. various viscosity parameter α . As seen in this figure, a high value of the viscosity parameter α causes a decrease in the velocity field, whereas the same parameter generates an enhancement in both the temperature and concentration fields, resulting in an increase in the thermal boundary thickness. The pushing force diminishes as the nanofluid becomes more viscous, which results in a reduction in the nanofluid motion throughout the boundary layer, according to the physical explanation for this phenomenon.
The influence of the slip velocity parameter λ on the velocity f ( η ) , temperature θ ( η ) , and nanoparticle concentration ϕ ( η ) profiles is shown in Figure 5. The temperature of the nanofluid θ ( η ) and the concentration profile ϕ ( η ) of nanoparticles both increase significantly as the slip velocity parameter λ is increased, whereas the velocity profile f ( η ) is changed in the opposite direction.
Figure 6 describes the results of the thermal conductivity parameter ε and the Eckert number E c on the temperature profiles θ ( η ) . Clearly, raising both the thermal conductivity parameter ε and the Eckert number E c increases the thickness of the thermal boundary layer as well as the temperature profile of the nanofluid. In terms of physics, the existence of a viscous dissipation phenomena or the dependence of the conductivity of nanofluids on temperature has the result of raising the fluid temperature, which enhances the thickness of the thermal boundary layer.
The influence of the Brownian motion parameter N b on temperature and concentration is depicted in Figure 7. The thermal thickness as well as the temperature distribution θ ( η ) grow rapidly as N b increases, whereas the nanoparticle concentration distribution ϕ ( η ) goes in the other direction with the same parameter. Physically, the Brownian motion parameter depends on the Brownian diffusion coefficient, and an increase in the Brownian motion parameter is associated with an increase in the heat diffusion and the temperature profile.
Table 2 summarizes the findings of the skin friction coefficient C f R e x , wall temperature gradient N u x R e x , and wall concentration gradient S h x R e x as a function of the parameters regulating the boundary layer region studied here. Table 2 shows that the local skin friction coefficient increases as the viscosity parameter, magnetic number, and porosity parameters increase, whereas the local Nusselt number and local Sherwood number for the same parameters show the opposite tendency. As the slip velocity parameter is maintained, the local skin-friction coefficient, surface heat transfer rate, and surface mass transfer rate all reduce. It is due to the fact that, as both the thermal conductivity parameter and the chemical reaction parameter expand, the surface mass transfer rate increases, whilst the local skin-friction coefficient and the local Nusselt number decrease. Additionally, increases in the Eckert number or Brownian motion parameter diminish both the rate of heat transmission and the value of the skin friction coefficient, but the local Sherwood number has the inverse result.

4. Conclusions

In this research, we looked at the heat and mass transfer mechanisms for non-Newtonian Williamson nanofluid flow caused by an elastic sheet with chemical reaction, slip velocity, and viscous dissipation. The sheet is assumed to be stretched linearly, exposed to a magnetic field, and submerged in a saturated porous medium. Using a shooting approach, the governing boundary layer equations are numerically solved for several values of the associated parameters. The following findings can be taken from the research work:
1.
The viscosity parameter, the porous parameter, and the magnetic parameter all show a decrease in velocity.
2.
Temperature rises as the Eckert number, magnetic parameter, porosity parameter, and slip velocity parameter increase.
3.
When the viscosity parameter, magnetic number, and porous parameter are increased, the skin-friction coefficient rises, whereas the chemical reaction parameter, Brownian motion parameter, and Eckert number grow in the opposite direction.
4.
The skin-friction coefficient and the local Nusselt number decline for the mounting values of thermal conductivity parameter and the Eckert number, whereas an enhancement in the local Sherwood number is observed.
5.
When the Brownian motion parameter and the chemical reaction parameter are both increased, the local Sherwood number increases; however, the increasing values of the slip velocity parameter and viscosity parameter have the opposite effect.

Author Contributions

Data curation, M.A.; Formal analysis, H.A.; Funding acquisition, E.A.; Methodology, A.M.M. 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.

Acknowledgments

The authors wish to express their sincere thanks to the honorable referees for their valuable comments and suggestions to improve the quality of the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Choi, S.U.S.; Zhang, Z.G.; Yu, W.; Lockwood, F.E.; Grulke, E.A. Anomalous thermal conductivity enhancement in nanotube suspensions. Appl. Phys. Lett. 2001, 79, 2252–2254. [Google Scholar] [CrossRef]
  2. Ghoneim, N.I.; Megahed, A.M. Hydromagnetic nanofluid film flow over a stretching sheet with prescribed heat flux and viscous dissipation. Fluid Dyn. Mater. Process. 2022, 18, 1373–1388. [Google Scholar] [CrossRef]
  3. Nadeem, S.; Haq, R.U.; Khan, Z.H. Numerical study of MHD boundary layer flow of a Maxwell fluid past a stretching sheet in the presence of nanoparticles. J. Taiwan Inst. Chem. Eng. 2014, 45, 121–126. [Google Scholar] [CrossRef]
  4. Rashad, A.M.; Chamkha, A.J.; Modather, M. Mixed convection boundary-layer flow past a horizontal circular cylinder embedded in a porous medium filled with a nanofluid under convective boundary conditio. Comput. Fluids 2013, 86, 380–388. [Google Scholar] [CrossRef]
  5. Armaghani, T.; Chamkha, A.J.; Maghrebi, J.; Nazari, M. Numerical analysis of a nanofluid forced convection in a porous channel: A new heat flux model in LTNE condition. J. Porous Media 2014, 17, 637–646. [Google Scholar] [CrossRef] [Green Version]
  6. Umavathi, J.C.; Chamkha, A.J.; Mohite, M.B. Convective transport in nanofluid saturated porous layer with crossdiffusion and variation of viscosity and conductivity. Special Topics Rev. Porous Media 2015, 6, 11–27. [Google Scholar] [CrossRef] [Green Version]
  7. Gireesha, B.J.; Mahanthesh, B.; Shivakumara, I.S.; Eshwarappa, K.M. Melting heat transfer in boundary layer stagnation-point flow of nanofluid toward a stretching sheet with induced magnetic field. Eng. Sci. Tech. Int. J. 2016, 19, 313–321. [Google Scholar] [CrossRef] [Green Version]
  8. Gul, H.; Ramzan, M.; Chung, J.D.; Chu, Y.M.; Kadry, S. Multiple slips impact in the MHD hybrid nanofluid flow with Cattaneo-Christov heat flux and autocatalytic chemical reaction. Sci. Rep. 2021, 11, 14625. [Google Scholar] [CrossRef]
  9. Ullah, I. Heat transfer enhancement in Marangoni convection and nonlinear radiative flow of gasoline oil conveying Boehmite alumina and aluminum alloy nanoparticles. Int. Commun. Heat Mass Transf. 2022, 132, 105920. [Google Scholar] [CrossRef]
  10. Muhammad Ramzan, M.; Riasat, S.; Chung, J.D.; Chu, Y.M.; Sheikholeslami, M.; Kadry, S.; Howari, F. Upshot of heterogeneous catalysis in a nanofluid flow over a rotating disk with slip effects and Entropy optimization analysis. Sci. Rep. 2021, 11, 120. [Google Scholar] [CrossRef]
  11. Ullah, I.; Hayat, T.; Alsaedi, A. Optimization of entropy production in flow of hybrid nanomaterials through Darcy-Forchheimer porous space. J. Therm. Anal. Calorim. 2022, 147, 5855–5864. [Google Scholar] [CrossRef]
  12. Ramzan, M.; Shaheen, N.; Chung, J.D.; Kadry, S.; Chu, Y.M.; Howari, F. Impact of Newtonian heating and Fourier and Fick’s laws on a magnetohydrodynamic dusty Casson nanofluid flow with variable heat source/sink over a stretching cylinder. Sci. Rep. 2021, 11, 2357. [Google Scholar] [CrossRef] [PubMed]
  13. Ullah, I.; Alghamdi, M.; Xia, W.-F.; Shah, S.I.; Khan, H. Activation energy effect on the magnetized-nanofluid flow in a rotating system considering the exponential heat source. Int. Commun. Heat Mass Transf. 2021, 128, 105578. [Google Scholar] [CrossRef]
  14. Ullah, I.; Ullah, R.; Alqarni, M.S.; Xia, W.; Muhammad, T. Combined heat source and zero mass flux features on magnetized nanofluid flow by radial disk with the applications of Coriolis force and activation energy. Int. Commun. Heat Mass Transf. 2021, 126, 105416. [Google Scholar] [CrossRef]
  15. Shaheen, N.; Ramzan, M.; Alshehri, A.; Shah, Z.; Kumam, P. Soret-Dufour impact on a three-dimensional Casson nanofluid flow with dust particles and variable characteristics in a permeable media. Sci. Rep. 2021, 11, 14513. [Google Scholar] [CrossRef]
  16. Ramzan, M.; Shahmir, N.; Ghazwani, H.A.S. Hybrid nanofluid flow comprising spherical shaped particles with Hall current and irreversibility analysis: An application of solar radiation. Waves Random Complex Media 2022. [Google Scholar] [CrossRef]
  17. Qasim, M.; Khan, Z.H.; Lopez, R.J.; Khan, W.A. Heat and mass transfer in nanofluid over an unsteady stretching sheet using Buongiorno’s model. Eur. Phys. J. Plus 2016, 131, 1–16. [Google Scholar] [CrossRef]
  18. Hussain, Z.; Hayat, T.; Alsaedi, A.; Ullah, I. On MHD convective flow of Williamson fluid with homogeneous-heterogeneous reactions: A comparative study of sheet and cylinder. Int. Commun. Heat Mass Transf. 2021, 120, 105060. [Google Scholar] [CrossRef]
  19. Patel, H.R.; Mittal, A.S.; Darji, R.R. MHD flow of micropolar nanofluid over a stretching/shrinking sheet considering radiation. Int. Commun. Heat Mass Transf. 2019, 108, 104322. [Google Scholar] [CrossRef]
  20. Yousef, N.S.; Megahed, A.M.; Ghoneim, N.I.; Elsafi, M.; Fares, E. Chemical reaction impact on MHD dissipative Casson-Williamson nanofluid flow over a slippery stretching sheet through porous medium. Alex. Eng. J. 2022, 61, 10161–10170. [Google Scholar] [CrossRef]
  21. Aminossadati, S.M.; Raisi, A.; Ghasemi, B. Effects of magnetic field on nanofluid forced convection in a partially heated microchannel. Int. J. Non-Linear Mech. 2011, 46, 1373–1382. [Google Scholar] [CrossRef]
  22. Ghasemi, B.; Aminossadati, S.M.; Raisi, A. Magnetic field effect on natural convection in a nanofluid-filled square enclosure. Int. J. Therm. Sci. 2011, 50, 1748–1756. [Google Scholar] [CrossRef]
  23. Humane, P.P.; Patil, V.S.; Patil, A.B. Chemical reaction and thermal radiation effects on magnetohydrodynamics flow of Casson-Williamson nanofluid over a porous stretching surface. J. Process. Mech. Eng. 2021, 235, 2008–2018. [Google Scholar] [CrossRef]
  24. Ullah, I.; Waqas, M.; Hayat, T.; Alsaedi, A.; Khan, M. Thermally radiated squeezed flow of magneto-nanofluid between two parallel disks with chemical reaction. J. Therm. Anal. Calorim. 2019, 135, 1021–1030. [Google Scholar] [CrossRef]
  25. Raju, C.S.K.; Babu, M.J.; Sandeep, N. Chemically reacting radiative MHD Jeffrey nanofluid flow over a cone in porous medium. Int. J. Eng. Res. Afr. 2016, 19, 75–90. [Google Scholar] [CrossRef]
  26. Sharma, K.; Gupta, S. Viscous dissipation and thermal radiation effects in MHD flow of Jeffrey nanofluid through impermeable surface with heat generation/absorption. Nonlinear Eng. 2017, 6, 153–166. [Google Scholar] [CrossRef]
  27. Alotaibi, H.; Althubiti, S.; Eid, M.R.; Mahny, K.L. Numerical treatment of MHD flow of Casson nanofluid via convectively heated nonlinear extending surface with viscous dissipation and suction/injection effects. Comput. Mater. Continua. 2021, 66, 229–245. [Google Scholar] [CrossRef]
  28. Mangathai, P.; Reddy, B.R.; Sidhartha, C. Unsteady MHD Williamson and Casson nano fluid flow in the presence of radiation and viscous dissipation. Turk. J. Comput. Math. Educ. 2021, 12, 1036–1051. [Google Scholar]
  29. Alali, E.; Megahed, M.A. MHD dissipative Casson nanofluid liquid film flow due to an unsteady stretching sheet with radiation influence and slip velocity phenomenon. Nanotechnol. Rev. 2022, 11, 463–472. [Google Scholar] [CrossRef]
  30. Sobamowo, G.M.; Jayesimi, L.O. Squeezing flow analysis of nanofluid under the effects of magnetic field and slip boundary conditions using chebychev cpectral collocation method. Fluid Mech. 2017, 3, 54–60. [Google Scholar] [CrossRef] [Green Version]
  31. Noor, N.A.M.; Shafie, S.; Admon, M.A. Slip effects on MHD squeezing flow of Jeffrey nanofluid in horizontal channel with chemical reaction. Mathematics 2021, 9, 1215. [Google Scholar] [CrossRef]
  32. Megahed, A.M. Improvement of heat transfer mechanism through a Maxwell fluid flow over a stretching sheet embedded in a porous medium and convectively heated. Math. Comput. Simul. 2021, 187, 97–109. [Google Scholar] [CrossRef]
  33. Nadeem, S.; Hussain, S.T. Flow and heat transfer analysis of Williamson nanofluid. Appl. Nanosci. 2014, 4, 1005–1012. [Google Scholar] [CrossRef] [Green Version]
  34. Rafique, K.; Alotaibi, H. Numerical simulation of Williamson nanofluid flow over an inclined surface: Keller box analysis. Appl. Sci. 2021, 11, 11523. [Google Scholar] [CrossRef]
  35. Megahed, A.M. Steady flow of MHD Williamson fluid due to a continuously moving surface with viscous dissipation and slip velocity. Int. J. Mod. Phys. C 2020, 31, 1–12. [Google Scholar] [CrossRef]
  36. Megahed, A.M. Flow and heat transfer of powell-eyring fluid due to an exponential stretching sheet with heat flux and variable thermal conductivity. Z. Naturforsch. 2015, 70, 163–169. [Google Scholar] [CrossRef]
  37. Khan, W.A.; Pop, I. Boundary-layer flow of a nanofluid past a stretching sheet. Int. J. Heat Mass Transf. 2010, 53, 2477–2483. [Google Scholar] [CrossRef]
Figure 1. Physical model with coordinates.
Figure 1. Physical model with coordinates.
Micromachines 13 01879 g001
Figure 2. (a) f ( η ) for picked M; (b) θ ( η ) and ϕ ( η ) for picked M.
Figure 2. (a) f ( η ) for picked M; (b) θ ( η ) and ϕ ( η ) for picked M.
Micromachines 13 01879 g002
Figure 3. (a) f ( η ) and ϕ ( η ) for picked δ ; (b) θ ( η ) for picked δ .
Figure 3. (a) f ( η ) and ϕ ( η ) for picked δ ; (b) θ ( η ) for picked δ .
Micromachines 13 01879 g003
Figure 4. (a) f ( η ) and ϕ ( η ) for picked α ; (b) θ ( η ) for picked α .
Figure 4. (a) f ( η ) and ϕ ( η ) for picked α ; (b) θ ( η ) for picked α .
Micromachines 13 01879 g004
Figure 5. (a) f ( η ) and ϕ ( η ) for picked λ ; (b) θ ( η ) for picked λ .
Figure 5. (a) f ( η ) and ϕ ( η ) for picked λ ; (b) θ ( η ) for picked λ .
Micromachines 13 01879 g005
Figure 6. (a) θ ( η ) for picked ε ; (b) θ ( η ) for picked E c .
Figure 6. (a) θ ( η ) for picked ε ; (b) θ ( η ) for picked E c .
Micromachines 13 01879 g006
Figure 7. (a) θ ( η ) for picked G; (b) ϕ ( η ) for picked G.
Figure 7. (a) θ ( η ) for picked G; (b) ϕ ( η ) for picked G.
Micromachines 13 01879 g007
Table 1. Comparison of θ ( 0 ) with the results of Khan and Pop [37] when α = M = δ = ε = E c = λ = W e = 0 and P r = 10 .
Table 1. Comparison of θ ( 0 ) with the results of Khan and Pop [37] when α = M = δ = ε = E c = λ = W e = 0 and P r = 10 .
Nb Nt Khan and Pop [37]Present Work
0.10.10.95240.95239854
0.20.20.36540.36540037
0.30.30.13550.13548752
0.40.40.04950.04949979
0.50.50.01790.01787205
Table 2. Values of C f R e x , N u x R e x and S h x R e x for various values of M , δ , α , λ , ε , N b , E c and G with N t = 0.1 , W e = 0.4 , P r = 3.0 and S c = 0.7 .
Table 2. Values of C f R e x , N u x R e x and S h x R e x for various values of M , δ , α , λ , ε , N b , E c and G with N t = 0.1 , W e = 0.4 , P r = 3.0 and S c = 0.7 .
M δ α λ ε Ec Nb G C f Re x Nu x Re x Sh x Re x
0.10.20.20.20.20.10.50.20.9244950.4423660.503462
0.30.20.20.20.20.10.50.20.9804130.4163180.495631
0.50.20.20.20.20.10.50.21.0304410.3926090.489825
0.30.00.20.20.20.10.50.20.9328340.4394960.502721
0.30.20.20.20.20.10.50.20.9804130.4163180.495631
0.30.40.20.20.20.10.50.21.0232100.3952060.490356
0.30.20.00.20.20.10.50.20.8748100.4289610.501554
0.30.20.20.20.20.10.50.20.9804130.4163180.495631
0.30.20.50.20.20.10.50.21.1477200.3942670.486333
0.30.20.20.00.20.10.50.21.1885410.4348610.514142
0.30.20.20.20.20.10.50.20.9804130.4163180.495631
0.30.20.20.40.20.10.50.20.8304240.3940870.482866
0.30.20.20.20.00.10.50.20.9814870.4501010.490738
0.30.20.20.20.20.10.50.20.9804130.4163180.495631
0.30.20.20.20.50.10.50.20.9791210.3777510.501609
0.30.20.20.20.20.00.50.20.9812670.4825380.483647
0.30.20.20.20.20.20.50.20.9795780.3507670.508082
0.30.20.20.20.20.50.50.20.9770660.1535520.544668
0.30.20.20.20.20.10.30.20.9825580.5240280.441149
0.30.20.20.20.20.10.50.20.9804130.4163180.495631
0.30.20.20.20.20.11.00.20.9757820.2123240.530527
0.30.20.20.20.20.10.50.00.9816270.4870310.278046
0.30.20.20.20.20.10.50.20.9804130.4163180.495631
0.30.20.20.20.20.10.50.50.9796540.3683350.703169
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Alrihieli, H.; Areshi, M.; Alali, E.; Megahed, A.M. MHD Dissipative Williamson Nanofluid Flow with Chemical Reaction Due to a Slippery Elastic Sheet Which Was Contained within a Porous Medium. Micromachines 2022, 13, 1879. https://doi.org/10.3390/mi13111879

AMA Style

Alrihieli H, Areshi M, Alali E, Megahed AM. MHD Dissipative Williamson Nanofluid Flow with Chemical Reaction Due to a Slippery Elastic Sheet Which Was Contained within a Porous Medium. Micromachines. 2022; 13(11):1879. https://doi.org/10.3390/mi13111879

Chicago/Turabian Style

Alrihieli, Haifaa, Mounirah Areshi, Elham Alali, and Ahmed M. Megahed. 2022. "MHD Dissipative Williamson Nanofluid Flow with Chemical Reaction Due to a Slippery Elastic Sheet Which Was Contained within a Porous Medium" Micromachines 13, no. 11: 1879. https://doi.org/10.3390/mi13111879

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