You are on page 1of 10

A non-Fourier approach towards the analysis of heat transfer enhancement with

water based nanofluids through a channel


Taimoor Dil, and M. Sabeel Khan

Citation: AIP Advances 8, 055311 (2018); doi: 10.1063/1.5005870


View online: https://doi.org/10.1063/1.5005870
View Table of Contents: http://aip.scitation.org/toc/adv/8/5
Published by the American Institute of Physics

Articles you may be interested in


Graphene-insulator-semiconductor capacitors as superior test structures for photoelectric determination of
semiconductor devices band diagrams
AIP Advances 8, 055203 (2018); 10.1063/1.4976624

A new method for spatial structure detection of complex inner cavities based on 3D γ-photon imaging
AIP Advances 8, 055205 (2018); 10.1063/1.4984027

The effect of surface wettability on the performance of a piezoelectric membrane pump


AIP Advances 8, 045010 (2018); 10.1063/1.5017993

Texture-enhanced Al-Cu electrodes on ultrathin Ti buffer layers for high-power durable 2.6 GHz SAW filters
AIP Advances 8, 045212 (2018); 10.1063/1.5017091

Implementing Bayesian networks with embedded stochastic MRAM


AIP Advances 8, 045101 (2018); 10.1063/1.5021332

Magnetic field manipulation of spin current in a single-molecule magnet tunnel junction with two-electron
Coulomb interaction
AIP Advances 8, 045309 (2018); 10.1063/1.5019651
AIP ADVANCES 8, 055311 (2018)

A non-Fourier approach towards the analysis of heat


transfer enhancement with water based nanofluids
through a channel
Taimoor Dil1,a and M. Sabeel Khan2,b
1 Department of Mechanical Engineering, Institute of Space Technology,
Islamabad 44000, Pakistan
2 Department of Applied Mathematics and Statistic, Institute of Space Technology,

Islamabad 44000, Pakistan


(Received 20 September 2017; accepted 3 December 2017; published online 8 May 2018)

In this article, a non-Fourier approach to model the heat transfer phenomenon in


nanofluids having application to automotive industry is studied. In this respect, a
recently proposed hyperbolic heat flux equation is embedded into the heat energy
equation and thereby incorporating the effect of thermal relaxation time. Nanofluids
are formed by considering copper oxide (CuO), Titanium dioxide (TiO2 ) and Alu-
minum oxide (Al2 O3 ) nano-solid particles in the base fluid. The flow governing system
of PDEs along with boundary conditions is transformed into its respective coupled
system of nonlinear ODEs using suitable similarity functions. Runge-Kutta-Fehlberg
(RK-5) numerical scheme embedded with shooting method is implemented and used
to solve the obtained boundary value problem. Numerical simulations are performed
and tabulated to analyze the influence of solid volume fraction on local coefficient of
skin-friction and Nusselt number. A comparison is made between the results by Fourier
and present heat flux model. We conclude that the presented new approach is more
general and thus allows predicting the influence of thermal relaxation time on the heat
transfer characteristics. Moreover, consideration of present model over the Fourier
model helps to predict the actual temporal behavior of solution. © 2018 Author(s).
All article content, except where otherwise noted, is licensed under a Creative
Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/).
https://doi.org/10.1063/1.5005870

I. INTRODUCTION
The phenomenon of heat transfer using nanofluids is of great industrial interest in many physical
applications.1–3 Since the operational speed of many electromechanical devices depends on the cool-
ing rate, therefore for this reason enhancement of heat transfer is always in demand of the associated
industry. The numerical study to analyze the heat transfer is therefore of great significance. Nanoflu-
ids have shown a good tendency in enhancing the heat transfer rate. Some of the experimental
observations in this respect are listed here. Pak and Cho4 illustrated that increasing volume frac-
tion and Reynolds number results in increasing Nusselt number for Al2 O3 –water and TiO2 –water
nanofluids. For nanofluids in the laminar flow regime a similar conclusion was reported by Yang
et al.5 Among other research reporting on the enhancement of heat transfer using nanofluids is the
work.6–11,23,37 An experimental study was also conducted by Xuan and Li6 to investigate flow features
and convective heat transfer coefficient of nanofluids. According to their results as flow velocity and
volume fraction of nanoparticles increases the heat transfer coefficient of nanofluid also increases.
Nanna et al.11,12 experimentally observed enhancement in convective heat transfer using nanoflu-
ids. Das et al.13 demonstrated experimentally that increase in temperature results an increase

a
e-mail: taimoor dil@yahoo.com
b
e-mail: drmsabeel@gmail.com

2158-3226/2018/8(5)/055311/9 8, 055311-1 © Author(s) 2018


055311-2 T. Dil and M. S. Khan AIP Advances 8, 055311 (2018)

in the nanofluid thermal conductivity. Moreover, a study on the nanoparticles volume fraction,
aspect ratio of the cylinder as well as the particle density was experimentally performed in
an enclosed horizontal cylinder by Putra et al.14 Ding et al.15 in their investigation found that
an increase in nanoparticle concentration systematically decreases natural convective heat trans-
fer coefficient. Natural convection is investigated with aqueous suspension of Al2 O3 micro-
particle in thin enclosures by Chang et al.16 Their results show that for vertical enclosure
the particles have an insignificant effect on the values for Nusselt number. Moreover nat-
ural convection heat transfer of nanofluid with various volume fractions of nanoparticles
(Al2 O3 ) and Rayleigh numbers was experimentally discussed by Ho et al.17 in a vertical
enclosure.
In the numerical analysis of heat transfer using nanofluid a numerous studies have been presented
in literature. For instance, Yirga and Shankar 18 investigated the convective heat transfer in nanofluid
through porous medium due to stretching surface subjected to viscous dissipation and chemical
reaction effects. Raza et al.19 studied the incompressible Cu-Water nanofluid and analyzed the heat
transfer in a porous channel with stretching walls. Li et al.20 presented an investigation of radiation heat
transfer of nanofluid in thin film with the effect of heat generation and thermophoresis. Pourmehran
et al.21 discussed the effect of external magnetic field and heat transfer in the analysis of nanofluid
flow derived by stretching sheet. Heat transfer of Newtonian nanofluid over a stretching surface
with varying temperature and velocity is discussed by Elyasi et al.22 Khanafer et al.23 determine
natural convection heat transfer of nanofluids by conducting a numerical study in an enclosure with
different physical conditions. Their results showed that an increase in the particle volume fraction
the average Nusselt number also increases for different Grashof numbers. Other related numerical
investigations on heat transfer using nanofluids include the studies (Refs. 24–33, and references
therein).
In all these previous numerical studies (Refs. 18–23, and references therein) so far parabolic
heat conduction law34 has been used in order to investigate heat transfer using nanofluids. But,
due to paradox of heat conduction35 the utilization of Fourier law of heat conduction in energy
balance is not much appreciated in literature.34 One of the major disadvantages of doing so is that
the energy balance equation becomes parabolic; moreover, any initial disturbance has an infinite
speed of propagation. Which in principle is not true, therefore Cattaneo35 suggested an alternate
to the Fourier law of heat conduction34 as a replacement of Fourier law of heat conduction. It was
Christov36 who made it possible to explicitly obtain the temperature field equation. The advantage of
this new heat conduction equation, which afterwards named Cattaneo-Christov heat flux equation, is
that it is hyperbolic in nature. As a result any initial disturbance in the temperature field will have a
finite speed of propagation in the medium.
Here, we apply for the first time in literature the Cattaneo-Christov heat flux equation in
analyzing heat transfer characteristics of water based nanofluids with application to automo-
bile radiator. Some interesting effects of thermal relaxation time on heat transfer rate are ana-
lyzed. Moreover, a comparison is made between the results obtained by Fourier heat flux equa-
tion and Cattaneo-Christov heat flux equation. The rest of this paper is structured as follows:
Section II describes the physical problem and formulates and describes the differential equa-
tions governing the dynamics of flow. Section III presents the discussion on the results from
the numerical experiments. Finally, Section IV states the conclusions drawn based on numerical
experiments.

II. PROBLEM FORMULATION


Let us consider a steady, incompressible and laminar flow of water-based nanofluid through the
channel of an automobile radiator. Copper-oxide, Alumina, and Titanium nano-solid particles are used
to form the nanofluid. As a representative problem consider the flow past a vertical two-dimensional
surface which is heated convectively by fluid temperature T f with having a heat transfer coefficient
hfluid . Assume that there is no-slip between the nano-solid particles interface and the base fluid. Since
the nanofluid behaves as a continuous medium therefore the following continuity, momentum and
energy balance equations37 govern the dynamics of the flow
055311-3 T. Dil and M. S. Khan AIP Advances 8, 055311 (2018)

∂u ∂v
+ = 0, (1)
∂x ∂y

∂u ∂u µnanofluid ∂ 2 u
u+v = + βnanofluid g(T − T∞ ), (2)
∂x ∂y ρnanofluid ∂y2
!2
∂T ∂T ∂u
  !
ρcp u +v = −∇ · q + µnanofluid , (3)
nanofluid ∂x ∂y ∂y
with the following boundary conditions
∂T  
u(0) = U0 , v(0) = 0, −knanofluid (0) = hfluid Tf − T (0) , u(∞) = 0, T (∞) = T∞ . (4)
∂y
In above equations u and v represent the fluid particle’s velocities, µnanofluid abbreviate the kine-
matic viscosity of nanofluid, β nanofluid represent the coefficient of thermal expansion, g denotes the
gravitational acceleration ρ represents density, k is thermal conductivity, and cp is the specific heat
constant. The heat flux q in eq. (3) is now determined by a recently proposed law of heat conduction
by Cattaneo-Christov.36 They incorporated the thermal relaxation time in the Fourier law of heat
conduction in the following way:
∂q   !
knanofluid ∇T + q + β − q · ∇V + V · ∇q + ∇ · V q = 0, (5)
∂t
!2
∂T ∂T ∂2T ∂u
  !
ρcp u +v = knanofluid 2 + µnanofluid
nanofluid ∂x ∂y ∂y ∂y
∂u ∂u ∂T ∂v ∂v ∂T ∂2T ∂2T ∂2T
( ! ! )
−β u +v + u +v + u2 2 + v 2 2 + 2uv . (6)
∂x ∂y ∂x ∂x ∂y ∂y ∂x ∂y ∂x∂y
Now, by assuming a uniform distribution of nanoparticles in the base fluid we can calculate thermo-
physical properties of resulting nanofluid from following equations
ρnanofluid = (1 − φ) ρfluid + φ ρnanoparticle , (7)

βnanofluid = (1 − φ) βfluid + φ βnanoparticle , (8)

µfluid
µnanofluid = , (9)
(1 − φ)2.5
     
ρcp = (1 − φ) ρcp + φ ρcp , (10)
nanofluid fluid nanoparticle

3 (r − 1) φ
" #
σnanofluid = σfluid 1 + , (11)
(r + 2) − (r − 1) φ
σnanoparticle
r= , (12)
σfluid
   
knanofluid knanoparticle + 2kfluid − 2φ kfluid − knanoparticle
=     . (13)
kfluid knanoparticle + 2kfluid + φ kfluid − knanoparticle
Wherein eqs. (7)–(13), φ is the nanoparticles volume fraction where φ = 0 corresponds to regular base
fluid. The constants β nanofluid , ρ, k, cp , are the coefficient of thermal expansion, density, coefficient
of thermal conductivity, constant of specific heat respectively. Now by introducing the following
similarity transformations
√ T − T∞
q .
η = a υfluid y, ψ = aυfluid xf (η) , θ (η) = , (14)
Tw − T∞
055311-4 T. Dil and M. S. Khan AIP Advances 8, 055311 (2018)

into eqs.(1),(2) and eq. (6) we obtain the coupled nonlinear differential equations as follows

ρnanoparticle  00
! 
1−φ+φ ff − f 02

 

ρ


 



 fluid 

f + (1 − φ) 
000 2.5   = 0,
 
(15)
β ρ
! ! 
nanoparticle nanoparticle
 
φ φ φ φ θ
 
+Gr 1 − + 1 − +

 

β ρ

 

 fluid fluid 

 
 ρc 
1 + kfluid p *.1 − φ + φ p nanoparticle +/ βf 2  θ 00 + kfluid Ec Pr
f 00 2

r
knanofluid (1 − φ)
 2.5
 knanofluid ρcp 

 , fluid -
 
kfluid
 ρCp 
nanoparticle  
pr 1 − φ + φ   1 − βf 0 f θ 0 = 0,

+   (16)
knanofluid  ρCp 
 nanofluid 

and eq. (4) reads


 
f |η=0 = 0, f 0 |η=0 = 1, θ 0 |η=0 = −Bi 1 − θ| η=0 , and f 0 |η=∞ = 0, θ 0 |η=∞ = 0. (17)

Where in above equations prime denote derivative with respect to η, and Gr, Pr, Ec, and Bi represents
the Grashof, Prandtl, Eckert and Biot number respectively and are stated as
 
βfluid g Tfluid − T∞ υfluid U0 2 hfluid υfluid x
r
Gr = , Pr = , Ec =  , Bi = . (18)
αfluid

U0 a Cpfluid Tfluid − T∞ kfluid U0

As for as engineering applications are concerned the quantities of physical interest are Nu the Nusselt
number and local coefficient C f of skin-friction, and are given by
τw xqw
Cf = , Nu = . (19)
ρfluid U0 2

kfluid Tfluid − T∞

Wherein eq. (19), τ w denotes the skin friction of the surface and qw represents the heat flux at the
surface and are calculated as
∂u ∂T
! !
τw = µnanofluid , qw = −knanofluid . (20)
∂y y=0 ∂y y=0

The local-skin-friction C fx and local-Nusselt number Nux can be obtained by computing τ w and qw
using eq. (20) and putting respective values in eq. (19). The resulting values are

Cfx = Rex Cf (1 − φ)2.5 = f 00 (0) ,


p
(21)
knanofluid 0
Nux Rex = Nu = − θ (0) ,
p
(22)
kf

where Rex is the local-Reynolds number and is defined by Rex = Uv0xf .

III. NUMERICAL SOLUTION METHODOLOGY


The governing system of coupled nonlinear differential equations i.e. eq. (15) and eq. (16) are
first converted to first order differential equations by the substitutions f = d 1 and θ = d 4 . In this way,
we convert the boundary value problem in Eqs. (15)–(17) into the following initial value problem
(IVP)
055311-5 T. Dil and M. S. Khan AIP Advances 8, 055311 (2018)

f 0 = d10 = d2 ,
f 00 = d20 = d3 ,
ρnanoparticle  00
 !  
1−φ+φ ff − f 02


 


ρ

 

 fluid 
f = d3 = −(1 − φ) 
000 0 2.5   = 0, ,
 
β ρ
! ! 
nanoparticle nanoparticle
 
 +Gr 1 − φ + φ β 1−φ+φ θ


 

ρfluid
 

 fluid 
θ 0 = d40 = d5 ,
(23)
1
θ 00 = d50 = ( ! )
kfluid (ρcp )nanoparticle
1+ knanofluid pr 1−φ+φ βf 2
(ρcp )fluid
 
 k ρCp 
fluid nanoparticle +
  0


p *.1 − φ + φ  βf 0
− 1 f θ


r

  / 

knanofluid ρC

 

p
 

× , nanofluid -  = 0,


 
k E



 fluid c Pr 00 2 



 − f 

φ) 2.5
 
k (1 −
 
 nanofluid 
subjected to the initial conditions
d1 (0) = 0, d2 (0) = 1, d3 (0) = q, d4 (0) = r, d5 (0) = −Bi (1 − r) . (24)
The above IVP is solved using RK-5 method with some initial choices for q and r. In order to refine
the guesses for q and r Newton method is used until the following stopping criterion is achieved
max {|d2 (12) − 0| , |d5 (12) − 0|} < ε
The tolerance ε for all the numerical results presented in this paper is taken to be ε = 10 8 .

IV. NUMERICAL RESULTS AND DISCUSSION


In this section we present and discuss computational results obtained by solving the boundary
value problem in Eqs. (15)–(17). Numerical experiments are performed for different values of thermal
relaxation time parameter. A finite domain [0, 12] is taken into account for the analysis instead of

TABLE I. Thermo-physical properties of water and nano-solid particles.

Materials ρ [kg/m3 ] β × 10 5 [K 1 ] σ [S/m] cp [J/kgK] k [W /mK]

Water (H2 O) 997.5 21.4 5.5 × 10 6 4178 0.628


Titania (TiO2 ) 4250 0.9 2.38 × 106 686.2 8.9538
Copper oxide (CuO) 6310 0.85 5.96 × 107 550.5 32.9
Alumina (Al2 O3 ) 3900 0.84 3.50 × 107 779 40

TABLE II. Nusselt number and Local skin-friction coefficient for CuO-H2 O nanofluid with varying nanoparticle-
concentration.

Rex −1/ 2 Nu Rex 1/ 2 Cf


φ k nanofluid FHM34 CCHM FHM34 CCHM

0 0.628 0.084138154 0.077851951 0.993989767 0.979710458


0.05 0.72141 0.084156656 0.080885778 1.212989968 1.21600044
0.1 0.82458 0.085508861 0.081543703 1.381935016 1.420309744
0.15 0.9391 0.085951918 0.083138796 1.58355532 1.639025602
0.2 1.06696 0.086160334 0.085475222 1.791982095 1.828859258
055311-6 T. Dil and M. S. Khan AIP Advances 8, 055311 (2018)

TABLE III. Nusselt number and Local skin-friction coefficient for Al2 O3 -H2 O nanofluid with varying nanoparticle-
concentration.

Rex −1/ 2 Nu Rex 1/ 2 Cf


φ k nanofluid FHM34 CCHM FHM34 CCHM

0 0.628 0.082889587 0.082889587 1.019001134 1.019001134


0.05 0.72239 0.086342294 0.086342294 1.159290297 1.159290297
0.1 0.82676 0.086390776 0.089318155 1.3092473 1.315008197
0.15 0.94275 0.087228597 0.081350589 1.436783851 1.479135835
0.2 1.07241 0.087525896 0.074965368 1.606078134 1.671223622

TABLE IV. Nusselt number and Local skin-friction coefficient for TiO2 -H2 O nanofluid with varying nanoparticle-
concentration.

Rex −1/ 2 Nu Rex 1/ 2 Cf


φ k nanofluid FHM34 CCHM FHM34 CCHM

0 0.628 0.074010991 0.074010991 1.017373496 1.017373496


0.05 0.70808 0.084510885 0.085397692 1.166120926 1.167191847
0.1 0.79527 0.08331349 0.082295046 1.325979797 1.324743045
0.15 0.89057 0.068886254 0.0707198 1.48915902 1.492035456
0.2 0.99515 0.041313107 0.043963353 1.662519924 1.66653377

an infinite domain. In all these numerical experiments the thermo-physical properties are taken from
Table I and the values for different physical parameters are computed from Eqs. (7)–(13). The Nusselt
number and local skin-friction coefficient for varying values of nano-particle concentration are shown
in Tables II–IV. A comparison is made between the numerical values of Nusselt number and skin-
friction coefficient obtained by our presented model with the values obtained by using Fourier heat
flux equation. The CuO-water nanofluid velocity profile for varying values of thermal relaxation
time is shown in Fig.1. The plotted velocity profiles are obtained by both the Fourier heat flux
model and Cattaneo-Christov heat flux model. It is observed that the fluid velocity in case of Fourier
heat flux model (FHM) is higher than the velocity obtained by Cattaneo-Christov heat flux model
(CCHM).

FIG. 1. Velocity profile of FHM and CCHM with varying values of β.


055311-7 T. Dil and M. S. Khan AIP Advances 8, 055311 (2018)

FIG. 2. Temperature profile with varying values of β.

In Fig.2 the temperature profiles obtained by FHM and CCHM are shown. It is observed that the
temperature values are always lower in case of CCHM than in case of FHM. Moreover a decrease
in the temperature boundary layer thickness is seen with an increase in the thermal relaxation time.
In Fig.3 the temperature profile of Al2 O3 -water nanofluid are plotted with both FHM and CCHM
for varying values of thermal relaxation time parameters. It can be seen that the temperature bound-
ary layer thickness decreases with an increase in the thermal relaxation time parameter. Moreover
the temperature values in case of CCHM are always lower than in case of FHM. Figure 4 shows
temperature profile of TiO2 -water nanofluid by FHM and CCHM. It is depicted from the plots that
temperature values obtained by CCHM are lower than FHM. Furthermore a decrease in thermal
boundary layer thickness is observed by increasing thermal relaxation time parameter. In Fig. 5 a
comparison is made between the temperature profiles with FHM and CCHM for different nanofluids.
It is seen that the temperature values in case of CCHM are always lower than the temperature values
obtained by FHM for every nanofluid case presented here.

FIG. 3. Temperature profile of Al2 O3 -water by FHM and CCHM with varying values of β.
055311-8 T. Dil and M. S. Khan AIP Advances 8, 055311 (2018)

FIG. 4. Temperature profile of TiO2 by FHM and CCHM with varying values of β.

FIG. 5. Temperature profile of different nanofluids by FHM and CCHM.

V. CONCLUSIONS
In this paper, a non-Fourier approach is used to study the heat transfer characteristics of water
based nanofluids and the results are compared using parabolic (FHM) and hyperbolic (CCHM) heat
flux models. In this study three different types of nanoparticles viz. CuO, Al2 O3 and TiO2 nanoparticle
are taken into account. Partial differential equations (PDEs) describing the flow dynamics together
with the boundary conditions are stated. To convert the PDEs into system of coupled nonlinear
differential equations similarity transformations are used. As a result boundary value problem in
ordinary differentia equations is obtained which afterwards is solved by using shooting technique
and Newton-Raphson method embedded with RK-5. The main findings of this investigation are stated
below:
1. Temperature and thermal boundary thickness have inverse proportionality with the thermal
relaxation-time. This result agrees with the observations of Han et al.38
055311-9 T. Dil and M. S. Khan AIP Advances 8, 055311 (2018)

2. Velocity is inversely proportional to the thermal relaxation time. As thermal relaxation time
increases velocity decreases and as a result fluid takes more time to move on plate and thus
absorbs more heat from the plate of the radiator.
3. It is also observed that the temperature values with present heat conduction model are smaller
than those obtained from the Fourier’s model. This is highly due to the damped hyperbolic equa-
tion used in present model which enables to resolve paradox of heat conduction in the Fourier
model. Thus allowing resolving the shortcoming of Fourier model which is completely due to
the constitutive relationship between heat flux and temperature field. Our present consideration
render the heat equation to a damped hyperbolic equation and as a consequence any heat distur-
bance hence propagates with a finite speed in the numerical simulation, thus resulting in lower
temperature field values in comparison to the Fourier heat flux model.
4. The consideration of present model over the Fourier model helps to predict the actual temporal
behavior of solution since the energy equation becomes hyperbolic instead of parabolic.
5. Moreover, the influence of solid volume fraction on local coefficient of skin-friction and Nusselt
number for the case of three different nanofluids is also tabulated in this investigation.
1 K. M. Shirvan, M. Mamourian, S. Mirzakhanlari, and R. Ellahi, Powder Technol. 313, 99–111 (2017).
2 M. Mamourian, M. S. Kamel, and S. Mirzakhanlari, J. Mol. Liq. 220, 888–901 (2016).
3 S. U. Rahman, R. Ellahi, S. Nadeem, and Q. M. Zaigham Zia, J. Mol. Liq. 218, 484–493 (2016).
4 B. C. Pak and Y. I. Cho, Exp. Heat Transfer. 11, 151–170 (1999).
5 Y. Yang, Z. Zhang, E. Grulke, W. Anderson, and G. Wu, Int. J. Heat Mass Tran. 48, 1107–1116 (2005).
6 Y. M. Xuan and Q. Li, ASME J. Heat Transfer 125, 151–155 (2003).
7 D. S. Wen and Y. L. Ding, J. Nanopart. Res. 7, 265–274 (2005).
8 B. X. Wang, L. P. Zhou, and X. F. Peng, Int. J. Heat Mass Tran. 46, 2665–2672 (2003).
9 S. K. Das, N. Putra, and W. Roetzel, Int. J. Heat Mass Tran. 46, 851–862 (2003).
10 S. K. Das, N. Putra, and W. Roetzel, Int. J. Multiphas. Flow. 29, 1237–1247 (2003).
11 A. G. A. Nanna, T. Fistrovich, K. Malinski, and S. U. S. Choi, In Proceedings of ASME International Mechanical Engineering

Congress and RD&D Exposition, Anaheim, California, USA. (2004).


12 A. G. A. Nnanna and M. Routhu, In Proceedings of ASME Summer Heat Transfer Conference, San Francisco, California,

USA. (2005).
13 S. K. Das, N. Putra, P. Thiesen, and W. Roetzel, J. Heat Transf. 125, 567–574 (2003).
14 N. Putra, W. Roetzel, and S. K. Das, Heat Mass Transfer. 39, 775–784 (2003).
15 Y. Ding et al., J. Particle Powder. 25, 23–36 (2007).
16 B. H. Chang, A. F. Mills, and E. Hernandez, Int. J. Heat Mass Tran. 51, 1332–1341 (2008).
17 C. J. Ho, W. K. Liu, Y. S. Chang, and C. C. Lin, Int. J. Therm. Sci. 49, 1345–1353 (2010).
18 Y. Yirga and B. Shankar, Int. J. Comp. Meth. Eng. Sci. and Mech. 16(5), 275–284 (2015).
19 J. Raza, A. M. Rohni, and Z. Omar, Int. J. Heat Mass Tran. 103, 336–340 (2016).
20 Li et al., J. Taiwan Inst. Chem. E 67, 226–234 (2016).
21 O. Pourmehran, M. Rahimi-Gorji, and D. D. Ganji, J. Taiwan Inst. Chem. E 65, 162–171 (2016).
22 P. Elyasi and A. Shateri, Transport Phenomena in Nano and Micro Scales 4(2), 28–40 (2016).
23 K. Khanafer, K. Vafai, and M. Lightstone, Int. J. Heat Mass Tran. 46, 3639–3653 (2003).
24 K. M. Shirvan, R. Ellahi, M. Mamourian, and M. Moghiman, Int. J. Heat Mass Tran. 107, 1110–1118 (2017).
25 R. Ellahi, M. H. Tariq, M. Hassan, and K. Vafa, J. Mol. Liq. 229, 339–345 (2017).
26 J. A. Esfahani, M. Akbarzadeh, S. Rashidi, M. A. Rosen, and R. Ellahi, Int. J. Heat Mass Tran. 109, 1162–1171 (2017).
27 E. S. Rashidi, A. J. Esfahani, and R. Ellahi, Applied Sciences 7, 431 (2017).
28 M. M. Hassan, A. Zeeshan, A. Majeed, and R. Ellahi, J. Magn. Magn. Mater. 443, 36–44 (2017).
29 R. Elahi, M. Hassan, and A. Zeeshan, Asia-Pac. J. Chem. Eng. 11(2), 179–186 (2016).
30 M. Akbarzadeh, S. Rashidi, M. Bovand, and R. Ellahi, J. Mol. Liq. 220, 1–13 (2016).
31 N. Shehzad, A. Zeeshan, R. Ellahi, and K. Vafai, J. Mol. Liq. 222, 446–455 (2016).
32 R. Ellahi, A. Zeeshan, and M. Hassan, Int. J. Numerical Methods for Heat and Fluid Flow 26(7), 2160–2174 (2016).
33 M. Sheikholeslami, Q. M. Zaigham Zia, and R. Ellahi, Applied Science 6, 1–11 (2016).
34 J. B. J. Fourier, (Paris, 1822).
35 C. Cattaneo, “Sulla conduzionedelcalore,” AttiSemin. Mat. Fis. Univ. Modena Reggio Emilia 3, 83–101 (1948).
36 C. I. Christov, Mech. Res. Commun. 36, 481–486 (2009).
37 M. S. Khan and T. Dil, AIP Advances 7, 045018 (2017).
38 S. Han, L. Zheng, C. Li, and X. Zhang, Appl. Math. Lett. 38, 87–93 (2014).

You might also like