Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Eur. Phys. J.

Plus (2019) 134: 176


DOI 10.1140/epjp/i2019-12493-5
THE EUROPEAN
PHYSICAL JOURNAL PLUS
Regular Article

Optimization of entropy generation in thermally stratified


polystyrene-water/kerosene nanofluid flow with convective
boundary condition

Aisha Anjum1 , N.A. Mir1 , M. Farooq1,a , S. Ahmad1 , and Naila Rafiq2


1
Department of Mathematics & Statistics, Riphah International University, Islamabad 44000, Pakistan
2
Department of Mathematics, NUML, Islamabad, Pakistan

Received: 27 November 2018 / Revised: 1 January 2019


Published online: 26 April 2019

c Società Italiana di Fisica / Springer-Verlag GmbH Germany, part of Springer Nature, 2019

Abstract. The refinements of thermodynamic features of systems greatly depend on the generation of
entropy since it produces thermodynamic irreversibility more appropriately. Efficiency of energy trans-
portation in a system can be improved by minimization of rate of entropy generation. Due to such a
crucial role in thermodynamics systems, our main focus is to elaborate the features of entropy generation
in polystyrenerene-water and polystyrene-kerosene nanofluids with the combined phenomena of thermal
stratification and convective boundary condition. Viscous dissipation and stagnation points are imple-
mented to analyze the flow characteristics deformed by the Riga plate. Momentum and energy equations
are modeled in view of the above assumptions. Dimensionless governing equations are acquired by using
similar transformations. Descriptions of numerous physical parameters are scrutinized on the tempera-
ture, entropy generation and velocity distribution. As an outcome, entropy generation can be minimized
by implementing the dominant thermal stratification parameter and the further rate of decay of entropy
generation is more prominent in water-polystyrene in comparison to kerosene-polystyrene nanofluid.

1 Introduction
The second law of thermodynamics reflects more important features in comparison to the first law, because it provides
mathematical tools for the reduction of friction and of the rate of entropy generation. This concept is known as the
exergy of the available energy resources. It helps to improve the efficiency of many engineering processes and electronic
devices because these processes involve heat transfer and further to compute the irreversibility in terms of entropy
generation. To avoid irreversibility losses and to improve the efficiency of various thermodynamical systems, it is
essential to optimize the rate of entropy generation. Bejan [1] for the first time used the ratio of thermal irreversibility
to the loss of total heat due to friction factors of fluids, which is named as the Bejan number. Researchers and
scientists are motivated to execute the features of fluid flow subjected to the thermodynamics second law. Hayat et
al. [2] described the features of entropy generation in Sisko fluid flow deformed by a rotating disk. Khan et al. [3]
discussed the characteristics of slip condition and entropy generation in a micropolar fluid influenced by a rotating
disk. Liu et al. [4] explored the analysis of entropy generation in the fluid flow confined in a curved microchannel with
magnetic and electric fields. Qayyum et al. [5] elaborated the rate of entropy generation of a Williamson fluid confined
between two rotating disks. Hayat et al. [6] demonstrated the characteristics of nanomaterial shapes in peristaltic flow
with entropy generation. Karimzadehkhouei et al. [7] scrutinized the entropy generation in water based nanofluids in
a microtube with constant heat flux.
Many engineering, biomedical and industrial processes are greatly influenced by magnetohydrodynamic fluid flows
which are used to pump, heat levitate and stir liquid metals. Such practical applications of magnetohydrodynamic
flows have captured the curiosity of researchers and scientists. Further these processes involve designing of cooling and
heating systems, instruments of blood flow measurement, nuclear reactors, MHD generators etc. Neutrons diffusion
rate can be regulated in the nuclear reactors by implementing a magnetic field. Deficiency of momentum in the
a
e-mail: m.farooq@riphah.edu.pk (corresponding author)
Page 2 of 13 Eur. Phys. J. Plus (2019) 134: 176

boundary layer of fluids having electrically conducting properties may be regulated by applying magnetic body forces.
Fluids having high electrically conductivity properties can be affected by implementing approximately 1 tesla magnetic
field. Classical MHD flows can be controlled by this technique. On the other hand, for fluids having low electrically
conducting features, an external magnetic field is inadequate for the induction of current. Therefore, the external
electric field is necessary for the generation of the Lorentz force in the direction of the wall. For this purpose, a Riga
plate consists of incorporated electrodes and magnets. Pressure drag and skin friction can be minimized by using a
Riga plate [8]. In this direction researchers have studied the Riga plate via various physical aspects such as Hayat et
al. [9] analyzed a convective heat phenomenon in the squeezed flow deformed by a Riga plate having chemical species,
Kumar et al. [10] presented radiative heat transfer in the nanofluid fluid deformed by a Riga plate having variable
thickness, Anjum et al. [11] demonstrated thermal and velocity slip phenomena in the stagnation point flow caused
by a Riga plate with thermal stratification, Farooq et al. [12] presented the behavior of chemical species in viscous
fluid flow deformed by a Riga plate through melting heat condition, Ayub et al. [13] examined a slip phenomenon in
the flow deformed by a Riga plate using non-Newtonian nanofluid.
Stratification is the arrangement of fluid particles in different layers which arises because of concentration and
temperature variations or mixture of various fluids. The stratification phenomenon plays a key role in the process of
the convective heat phenomenon due to wide spread applications in industrial, natural and engineering processes. Such
application comprises of thermal stratification of oceans and reservoirs, atmosphere having heterogeneous mixtures
etc. Thermal and solutal stratification phenomena in mixed convective flow deformed due to a stretchable surface
saturated with a porous medium is examined by Srinivasacharya and Surender [14]. Features of double stratified
and heat absorption/generation in micropolar fluid are demonstrated by Mishra et al. [15]. RamReddy et al. [16]
presented a thermally stratified phenomenon in nanofluid flow over a stretchable sheet saturated with a non-Darcry
porous medium. Srinivasacharya and Surender [17] examined the thermal stratification phenomenon in nanoliquid flow
deformed by a stretchable. Ahmad et al. [18] analyzed double stratification phenomena in squeezed flow of a Sutterby
fluid with chemical species and radiative heat flux. Ahmad et al. [19] presented solutal and thermal stratification in
squeezed fluid flow with velocity, thermal and solutal slip mechanism.
Human society in the 21st century is greatly concerned with energy resources because of advanced and rapid
technological and biomedical processes and equipment. Thus scientists and researchers show their keen interest to
discover new energy resources. Choi [20] discovered and used firstly the nanoparticles for the enhancement of thermal
conductivity of numerous ordinary fluids and to maximize the storage of useful energy. The homogeneous suspension
of nano-sized particles in the base fluid is categorized as nanofluid. It has applications to enhance the rate of heat
transfer of computers, microchips, microelectronics, fuel cells, biomedicine, transportation, food processing, solid state
lightening and manufacturing. For cooling processes liquids like ethylene, water, glycol oil etc are used in the industries
which have low thermal conductivity. Nanometer sized metallic particles (titanium, gold, copper, iron or their oxides)
are suspended in these liquids in order to achieve higher thermal conductivity efficiency. Nanoparticles have various
shapes for example spherical, rod-like or tubular. Magnetized nanoparticles are used for the treament of cancer cells.
In this treament a medicine containing nanoparticles is injected into the blood vessel near to the tumor and also
a magnet is placed near the tumor. These nanoparticles behave like heat sources by implementing a magnetic field
which destroys cancer tissues at the temperature 42–45 ◦ C [21]. Characteristics of entropy generation in the nanofluid
flow with the emission of nonlinear thermal radiation have been disclosed by Hayat et al. [22]. The description of
Al2 O3 -H2 O and Al2 O3 -C2 H6 O2 in the presence of entropy phenomena is scrutinized by Hayat et al. [23].
A literature survey reflects that a lot of studies have been done by the researchers to elaborate the characteristics
of entropy generation and nanofluid fluid flow using various nanoparticles like Cu, Ag, Al2 O3 , carbon nanotubes etc.
However, no attempt has been made up till now to explore the features of entropy generation in nanofluid using
polystyrene nanoparticles. These particles have applications in plastic surgery treament and soil microorganisms and
enzyme activities. Further these nanoparticles are friendly suitable to soil microbes and in their involved processes.
Hence our main objective is to fill such void. Two types of nanofluids are implemented i.e., water-polystyrene and
kerosene-polystyrene. Stagnation point flow analysis is carried out due to a Riga plate with viscous dissipation. Fur-
ther thermal stratification and convective boundary condition are implemented for heat transportation. Dimensionless
governing equations are solved analytically by the homotopic approach [24–30]. Graphical behavior of temperature,
velocity, skin friction, Nusselt number and entropy generation corresponding to various pertinent parameters is dis-
played and elaborated. We compared our results of skin friction in the limiting case and found excellent agreement
between the present and already published work.

2 Mathematical formulation

We consider the entropy generation in incompressible and two-dimensional stagnation point flow of water-polystyrene
and kerosene-polystyrene nanofluid deformed by a stretchable Riga plate. Combined analysis of convective bound-
ary condition and linear thermal stratification is used to elaborate the features of heat transport (see fig. 1).
Eur. Phys. J. Plus (2019) 134: 176 Page 3 of 13

Fig. 1. Flow goemetry.

Viscous dissipation for the nanofluid flow is also retained. The modeled equations after implementing boundary layer
approximations are as follows:

∂u ∂v
+ = 0, (1)
∂x ∂y
∂u ∂u dUe ∂ 2 u πj0 M  π 
u +v = Ue + νnf 2 + exp − y , (2)
∂x ∂y dx ∂y 8ρnf a1
 2
∂T ∂T ∂2T μnf ∂u
u +v = αnf 2 + , (3)
∂x ∂y ∂y (ρcp )nf ∂y

The implemented boundary conditions are

∂T
u = Uw (x) = ax, v = 0, knf = −hs (Ts − T ) at y = 0,
∂y
u → Ue (x) = bx, T → T∞ = T0 + c1 x as y → ∞. (4)

Here u and v represent the velocity components corresponding to horizontal and vertical directions respectively, wall
stretching and free stream velocities are symbolized as Uw and Ue respectively, μnf , νnf represent the nanofluid
dynamic and kinematic viscosities respectively, j0 represents the applied current density, M (= M0 x) represents the
magnetization, distance between magnets and electrodes is represented by a1 , αnf describes thermal diffusivity of
nanofluid, T and T∞ reflect the fluid temperatures inside and outside the boundary layer, Ts (= T0 + cx) demonstrates
the heated fluid temperature, ρnf represents the fluid’s density, hs depicts the heat transfer coefficient, knf reflects
the nanofluid’s thermal conductivity whereas a, b, c1 are the positive dimensional constants.
The thermophysical properties of nanofluids are mathematically expressed as follows:

μf μnf
μnf = , νnf = , ρnf = (1 − φ)ρf + φρs ,
(1 − φ)2.5 ρnf
knf knf ks + 2kf − 2φ(kf − ks )
αnf = , = , (5)
ρnf (cp )nf kf ks + 2kf + φ(kf − ks )

where μnf , φ, ρs , ρf represent viscosity, nanoparticle volume fraction, density of solid particles and density of base
fluid respectively, while kf , ks and knf demonstrate the thermal conductivities of fluid, solid particles and nanofluid
respectively (see table 1).
Transformations of the form

a √
η=y , u = axf  (η), v = − aνf f (η),
υf
T − T∞
θ(η) = , (6)
Ts − T0
Page 4 of 13 Eur. Phys. J. Plus (2019) 134: 176

Table 1. Thermophysical features of nanoparticles and base fluids.

Physical properties Nanomaterial Base fluids


Polystyrene Water Kerosene Oil
3
ρ (kg/m ) 1053 997 783
cp (J/kgK) 1210 4179 2090
k (W/mK) 0.16 0.613 0.145

Mass conservation law vanishes identically and eqs. (1) to (4) are reduced to


1 Q
 f  + f f  − (f  )2 + S 2 +  exp(−Bη) = 0, (7)
(1−φ) 2.5 1−φ+φ ρρs
f
1 − φ + φ ρρfs
⎛ ⎞
⎝ knf /kf Pr Ec
 ⎠ θ + Pr (f θ − f  θ) − Pr S1f  + f 2 = 0, (8)
1−φ+φ
(ρcp )s (1 − φ) 2.5
(ρcp )f

The boundary conditions take the form


Bi
f (0) = 0, f  (0) = 1, θ (0) = − (1 − S1 − θ(0)),
knf /kf
f  (∞) → S, θ(∞) → 0, (9)
where S represents the ratio of rates, Q represents the modified Hartman number, B is the dimensionless length
parameter, Pr represents the Prandtl number, S1 is the thermal stratification parameter (for S1 = 0 means that the
temperature of the ambient fluid is constant while S1 > 1 represents that the fluid is heated and the plate is cold),
Ec is the Eckert number and Bi represents the Biot number (for Bi = 0 corresponding to an insulated plate while
Bi → ∞ represents the variable temperature at the plate) which are mathematically represented as follows:

μcp b πj0 M0 π υ c1 Uw2
Pr = , S= , Q= , B = , S1 = , Ec = . (10)
k a 8ρf a2 a1 a c Cp (Ts − T0 )
The mathematical expressions for surface drag and heat transfer rate or the Nusselt number for the considered
flow are as follows:
τw xqw
Cf = , N ux = , (11)
2
ρf Uw kf (Ts − T∞ )
   
∂u ∂T
τw = μnf , qw = −κnf . (12)
∂y y=0 ∂y y=0

Dimensionless forms of these quantities are


1 knf θ (0)
Cf Re1/2 = f  (0), N ux Re−1/2 =− , (13)
x
(1 − φ)2.5 x
kf 1 − S1
where Rex = Uw x/ν depicts the Reynolds number.

3 Analysis of entropy generation


Here our main motto is to compute the irreversibility of the system via entropy generation. Entropy generation is
defined as follows:  2  2
knf ∂T μnf ∂u
EG = + . (14)
(T∞ − T0 )
2 ∂y (T∞ − T 0 ) ∂y
The first term is responsible for the system irreversibility due to heat transfer while the second term is responsible for
heat generation via viscous dissipation. The dimensionless relation for entropy generation is expressed as
(T∞ − T0 )2 ( ηy )2
Ns = 2 EG , (15)
knf (Ts − T∞ )
Eur. Phys. J. Plus (2019) 134: 176 Page 5 of 13

by implementing transformations, we have


θ2 Ec Pr kf 2
Ns = + f . (16)
(1 − S1)2 (1 − S1)(1 − φ) knf
2.5

The Bejan number for each point location of the boundary layer is expressed as [1]
Entropy generation due to thermal irreversibility
Be = ,
Total entropy generation
θ 2
(1−S1)2
Be = θ 2 kf
. (17)
(1−S1)2 + Ec Pr
(1−S1)(1−φ)2.5 knf f 2

4 Homotopic solutions
The homotopic technique (developed by Liao [24] in 1992) is implemented for the construction of analytical series
solutions of highly nonlinear equations. This method is preferred over the other methods due to many factors i.e., a)
it is free from any large or small parameters b) guarantees the series solution convergence c) provides great freedom
in in the selection of operators and initial approximation. Thus initial approximations and linear operators are
Bi
f0 (η) = Sη + (1 − S)(1 − exp(−η)), θ0 (η) = knf
(1 − S1) exp(−η), (18)
kf + Bi
d3 f df d2 θ
Lf (f ) = − , Lθ (θ) = − θ, (19)
dη 3 dη dη 2
with
Lf [A∗1 + A∗2 exp(η) + A∗3 exp(−η)] = 0, (20)
Lθ [A∗4 exp(η) + A∗5 exp(−η)] = 0, (21)
where A∗i (i = 1, 2, . . . , 5) represent the arbitrary constants.

4.1 Zeroth-order problem


   
(1 − q)Lf f(η; q) − f0 (η) = qf Nf f(η; q) , (22)
   
 q) − θ0 (η) = qθ Nθ θ(η;
(1 − q)Lθ θ(η;  q), f(η; q) , (23)

f(0; q) = 0, f (0; q) = 1, f (∞; q) = S, (24)


Bi  
θ (0; p) = − knf 1 − S1 − θ(0; q) , 
θ(∞; q) = 0, (25)
kf


  1 ∂ 3 f(η; q) ∂ 2 f(η; q)
Nf f(η, q), θ(η;
 q) =  3
+ f(η; q)
(1−φ)2.5 1−φ+φ ρρs
f
∂η ∂η 2

2
∂ f(η; q) Q
− + S2 +  exp(−Bη), (26)
∂η 1 − φ + φ ρρsf
⎛ ⎞
  knf /kf  q)
∂ 2 θ(η,
Nθ  q), f(η; q) = ⎝ 
θ(η; ⎠
1 − φ + φ (ρcpp)fs
(ρc ) ∂η 2
⎛ 
2 ⎞
 
∂ θ(η; q) ∂ f (η; q)  
∂ f (η; q) Ec 2
∂ f (η; q) ⎠
+ Pr ⎝f(η; q) − θ(η; q) − S1 + ,
∂η ∂η ∂η (1 − φ)2.5 ∂η 2
(27)
where q represents the embedding parameter and has value [0, 1] while non-zero auxiliary parameters are represented
by f and θ .
Page 6 of 13 Eur. Phys. J. Plus (2019) 134: 176

- 0.6

Water-Polystyrene f ''(0)
Kerosene-Polystyrene

- 0.7
φ = 0.1, S=0.1, B=0.3, Q= 0.2, Ec = 0.2, S1 =0.2, Bi = 0.3

- 0.8

- 0.9

ħ
-1.0 -0.8 -0.6 -0.4 -0.2

-1.0

Fig. 2. -curve for f .

4.2 mth-order deformation problems

Lf [fm (η) − χm fm−1 (η)] = f Rfm (η), (28)

Lθ [θm (η) − χm θm−1 (η)] = θ Rθm (η), (29)


 
fm (0) = 0, fm (0) = 0, fm (∞) = 0, (30)

 Bi
θm (0) = θm (0), θm (∞) = 0, (31)
knf /kf


1 
m−1


Rm (η) =
f  fm−1 + fm−1−k fk − fm−1−k

fk
(1−φ)2.5 1−φ+φ ρρs
f k=0

Q
+ S 2 (1 − χm ) +  exp(−Bη), (32)
1 − φ + φ ρρfs
⎛ ⎞
knf /kf 
m−1
Ec

Rθm (η) = ⎝   ⎠ θm−1

+ Pr fm−1−k θk − fm−1−k

θk + f  f 
(ρc )
1 − φ + φ (ρcpp)fs (1 − φ)2.5 m−1−k k
k=0


− Pr S1fm−1 , (33)

⎨0, m ≤ 1,
χm = (34)
⎩1, m > 1.

The general solutions (fm , θm ) of eqs. (28) and (29) are represented by

fm (η) = fm (η) + A∗1 + A∗2 eη + A∗3 e−η , (35)

θm (η) = θm (η) + A∗4 eη + A∗5 e−η , (36)

here fm 
(η) and θm (η) demonstrate the special solutions where A∗i depicts arbitrary constants.

4.3 Convergence analysis

The convergence of the computed analytic solutions is a necessary and important part of the analysis. To ensure the
convergence region, we have demonstrated graphically the -curves in figs. 2 and 3. Auxiliary parameters have ranges
for water-polystyrene nanofluid −0.8 ≤ f ≤ −0.25, −0.3 ≤ θ ≤ −0.1 while for kerosene-polystyrene nanofluid are
−0.6 ≤ f ≤ −0.225, −0.35 ≤ θ ≤ −0.1.
Eur. Phys. J. Plus (2019) 134: 176 Page 7 of 13

1.0

θ '(0) Water-Polystyrene
0.5
Kerosene-Polystyrene

φ = 0.1, S =0.1, B =0.3, Q = 0.2, Ec = 0.2, S1 =0.2, Bi = 0.3


0.0

- 0.5

- 1.0

- 0.8 - 0.6 - 0.4 - 0.2 0.2 0.4 0.6 ħ


- 1.5

Fig. 3. -curve for θ.

f ' (η)
1.2 S= 1.2
φ = 0.3, Q =0.2, B =3.0, S1 = 0.2, Ec = 0.2, Bi =0.1
1.1 S= 1.1

Dashed Lines: Water-Polystyrene


1.0 S=1.0
Solid Lines: Kerosene-Polystyrene
S= 0.9
0.9

S= 0.8
0.8

η
1 2 3 4 5 6
Fig. 4. Description of S on f  (η).

5 Discussion

Here our main focus is to demonstrate the features of numerous pertinent parameters on the temperature and velocity
distributions. Figure 4 reflects the behavior of ratio parameter S on the horizontal velocity for water-polystyrene and
kerosene-polystyrene nanofluids. Dominant behavior of velocity field is noted for higher values of the ratio parameter
for S < 1 and S > 1 i.e., either stretching velocity or free stream velocity is dominant. Thew maximum velocity
exists at the surface of the plate for S < 1, however for S > 1 the maximum velocity is located away from surface.
Further velocity is dominant for water-polystyrene in comparison to kerosene-polystyrene nanofluid. No deformation
or disturbance occurs in the fluid when S = 1 i.e., free stream velocity is equal to stretching velocity in that case
and also no boundary layer exists. The analysis of the modified Hartman number on the velocity field is illustrated
in fig. 5. The velocity field gradually enhances for a higher modified Hartman number. The physical justification is
that the dominant modified Hartman number results in higher intensity of the electric field which consequently grows
the velocity distribution. Water-polystyrenerene nanofluid has dominant velocity compared to kerosene-polystyrene
nanofluid. Description of nanoparticle volume fraction φ on horizontal velocity is illustrated in fig. 6 for two cases i.e.,
i) water-polystyrene and ii) kerosene-polystyrene nanofluid. A higher nanoparticle volume fraction is responsible for
dominant velocity behavior of the fluid particles. It is also concluded that velocity is recessive for kerosene-polystyrene
nanofluid. However, maximum velocity occurs at the surface of the plate. The characteristics of thermal stratification
S1 on the temperature field are displayed in fig. 7. Reduction occurs in the temperature field corresponding to dominant
thermal stratification. A higher thermal stratification classifies the fluid particles with different density regions which
resist the heat transfer. Hence temperature reduces. However, the rate of decay in water-polystyrene is faster than in
kerosene-polystyrene nanofluid. Figure 8 demonstrates the behavior of the Biot number on the temperature field. It is
interesting to note that water-polystyrene and kerosene-polystyrene show dramatically opposite behavior for a higher
Biot number i.e., the temperature field grows for water-polystyrene nanofluid while decays for kerosene-polystyrene
Page 8 of 13 Eur. Phys. J. Plus (2019) 134: 176

f ' (η)
1.0
φ = 0.5, S =0.1, B =3.0, S1 = 0.2, Ec = 0.2, Bi =0.1
0.8
Q = 0.0, 0.7, 1.4, 2.5

Dashed Lines: Water-Polystyrene


0.6 Solid Lines: Kerosene-Polystyrene

0.4

0.2

1 2 3 4 5 6
η
Fig. 5. Description of Q on f  (η).

f ' (η)
1.0
Bi = 0.8, B =0.3, S1 = 0.1, Ec = 0.2, S = 0.1, Q = 0.2

0.8 φ = 0.2, 0.4, 0.5, 0.6

Dashed Lines: Water-Polystyrene


0.6 Solid Lines: Kerosene-Polystyrene

0.4

0.2
η
1 2 3 4 5
Fig. 6. Description of φ on f  (η).

θ (η)
1.0 φ = 0.6, S =0.1, B =0.3, Q = 0.2, Ec = 0.2, Bi =0.7

0.8 S1 = 0.0, 0.2, 0.4, 0.6

Dashed Lines: Water-Polystyrene


0.6
Solid Lines: Kerosene-Polystyrene
0.4

0.2

η
1 2 3 4 5 6
Fig. 7. Description of S1 on θ(η).

nanofluid. Heat transfer coefficient enhances for a higher Biot number. Therefore, the temperature field gradually
increases for water-polystyrene nanofluid. Figure 9 reflects the behavior of the nanoparticle volume fraction on the
temperature field. Increment is observed in temperature for the nanoparticle volume fraction. Figure 10 demonstrates
the analysis of the Eckert number on the temperature field. Here the dominant Eckert number results in enhancement
of the temperature field for both types of nanofluids. The Eckert number is responsible to produce more heat in
the system which alternatively increases the fluid temperature. Figure 11 displays the characteristics of the ratio
Eur. Phys. J. Plus (2019) 134: 176 Page 9 of 13

θ (η)
0.5
φ = 0.5, S =0.1, B =0.3, Q = 0.2, Ec = 0.2, S1 =0.2

0.4 Bi = 0.1, 0.3, 0.5, 0.7


Dashed Lines: Water-Polystyrene
0.3 Solid Lines: Kerosene-Polystyrene

0.2

0.1

1 2 3 4 5 6
η
Fig. 8. Description of Bi on θ(η).

θ (η)
0.14
Bi = 0.2, B =0.3, S1 = 0.2, Ec = 0.2, S =0.2, Q = 0.2
0.12 φ = 0.3, 0.35, 0.4, 0.45

0.10 Dashed Lines: Water-Polystyrene


Solid Lines: Kerosene-Polystyrene
0.08
0.06
0.04
0.02

1 2 3 4 5
η
Fig. 9. Description of φ on θ(η).

parameter and the modified Hartman number on the skin friction coefficient for both types of nanofluids i) water-
polystyrene and ii) kerosene-polystyrene. Skin friction coefficient decays by increasing both modified Hartman number
and ratio parameter. It is interesting to note that polystyrene nanoparticles work more efficiently in water to reduce
the magnitude of skin friction in comparison to kerosene basefluid. Figure 12 is plotted to analyze the behavior of the
Biot number and thermal stratification parameter on the Nusselt number. The Biot number and thermal stratification
parameter reflect their behaviors oppositely on the Nusselt number for both types of nanofluids. However, the rate of
heat transfer is dominant for kersene-polystyrene nanofluid in comparison to water-polystyrene.

6 Production of entropy
The main objective of this section is to concentrate on the behavior of entropy generation corresponding to various
physical parameters. Figure 13 shows the behavior of entropy generation corresponding to the Eckert number. A
higher Eckert number is responsible for more entropy generation. Physically more heat is produced due to a higher
Eckert number and consequently entropy generation enhances. Further it is also noticeable that entropy generation is
minimum for water-polystyrene in comparison to kerosene-polystyrene nanofluid. Figure 14 reflects the characteristics
of the thermal stratification parameter on entropy generation. Decrement occurs in the entropy generation for dominant
values of the thermal stratification parameter. Minimization in entropy is achieved for water-polystryrene nanofluid
in comparison to kerosene-polystryrene nanofluid. Impact of Biot number on entropy generation is illustrated in
fig. 15. Entropy generation enhances for dominant behavior of the Biot number for both types of nanofluids. However,
minimum entropy is generated for the case of water-polystyrene nanofluid. Physically a higher Biot number corresponds
to a higher rate of heat transfer which consequently results in higher entropy generation. Figure 16 demonstrates the
streamlines behavior of flow.
Table 2 reflects that all the results are matched in excellent order.
Page 10 of 13 Eur. Phys. J. Plus (2019) 134: 176

θ (η)
φ = 0.2, S =0.1, B =3.0, Q = 0.3, S1 = 0.2, Bi =0.2
0.20
Ec = 0.4, 0.5, 0.6, 0.7
0.15 Dashed Lines: Water-Polystyrene
Solid Lines: Kerosene-Polystyrene
0.10

0.05

η
1 2 3 4
Fig. 10. Description of Ec on θ(η).

Cf Re1/2
φ = 0.5, B =0.3, Ec = 0.2, S1 =0.2, Bi = 0.8

S=0.4
- 2.0 Dashed Lines: Water-Polystyrene
Solid Lines: Water-Polystyrene
- 2.5

- 3.0

- 3.5 S =0.2
S =0.3 S =0.1

Q
1 2 3 4
Fig. 11. Description of Q and S on Cf Re1/2 .

Fig. 12. Description of S1 and Bi on N ux Re−1/2 .


Eur. Phys. J. Plus (2019) 134: 176 Page 11 of 13

Ns
3.0 φ = 0.2, S =0.1, B =3.0, Q = 0.3, S1 = 0.2, Bi =0.2

2.5 Ec = 0.1, 0.2, 0.3, 0.4

2.0 Dashed Lines: Water-Polystyrene


Solid Lines: Kerosene-Polystyrene
1.5

1.0

0.5

0.5 1.0 1.5 2.0 2.5


η
Fig. 13. Description of Ec on N s.

Ns
φ = 0.1, S =0.1, B =0.3, Q = 0.2, Ec = 0.2, Bi =2.0
0.8
S1 = 0.0, 0.2, 0.4, 0.6
0.6 Dashed Lines: Water-Polystyrene
Solid Lines: Kerosene-Polystyrene
0.4

0.2

0.2 0.4 0.6 0.8 1.0 1.2 1.4 η


Fig. 14. Description of S1 on N s.

Ns
0.8
φ = 0.1, S =0.1, B =0.3, Q = 0.2, Ec = 0.2, S1 =0.1

Bi = 0.1, 1.0, 2.0, 3.0


0.6
Dashed Lines: Water-Polystyrene
Solid Lines: Kerosene-Polystyrene
0.4

0.2

η
0.5 1.0 1.5 2.0 2.5 3.0
Fig. 15. Description of Bi on N s.
Page 12 of 13 Eur. Phys. J. Plus (2019) 134: 176

2.0

φ = 0.2, S =0.1, B =0.3, Q = 0.2, Ec = 0.2, Bi =0.7, S1 = 0.2


1.5 φ = -1.5 φ = 1.5

φ = -1.0 φ = 1.0

φ = -0.5 φ = 0.5
y 1.0
φ = 0.0

Dashed Lines: Water-Polystyrene


0.5 Solid Lines: Kerosene-Polystyrene

0.0
-4 -2 0 2 4
x
Fig. 16. Streamlines for the flow.

Table 2. Analysis of surface drag coefficient (f  (0)) with published works of Mahapatra and Gupta [28], Pop et al. [29] and
Sharma and Singh [30] for various values of S when Q = 0 and φ = 0.

S Mahapatra and Gupta [28] Pop et al. [29] Sharma and Singh [30] Present results
0.1 −0.9694 −0.9694 −0.969386 −0.96939
0.2 −0.9181 −0.9181 −0.9181069 −0.91811
0.5 −0.6673 −0.6673 −0.667263 −0.66726

7 Closing remarks
In the present work we have elaborated the features of combined analysis of thermal stratification and convective
boundary condition in polystyrene-water/kerosene nanofluids in the neighborhood of the stagnation point. The key
points are summarized as follows:
– Velocity distribution enhances for dominant values of the modified Hartman number.
– The temperature field reflects opposite behavior for water-polystryrene and kerosene-polystyrene nanofluids by
increment in the Biot number.
– The drag force decays by increasing the ratio parameter and modified Hartman number.
– Minimization of entropy generation occurs for a higher thermal stratification parameter especially for water-
polystyrene nanofluid.
– Entropy generation can be controlled for a recessive Biot number.
This attempt is hoped to act as stimulus and motivation for nano-medicine treatments. In this method, nanoparti-
cles having medicine are injected into the patient blood especially in cancer treatment. Such nanoparticles are pushed
by an external force or propagated through arteries walls which behave as a boundary layer flow. Further it may
be used for plastic surgery treatments and soil microorganisms and enzyme activities. These nanoparticles might be
friendly suitable to soil microbes and their involved processes.

Publisher’s Note The EPJ Publishers remain neutral with regard to jurisdictional claims in published maps and institutional
affiliations.

References
1. A. Bejan, Entropy Generation Through Heat and Fluid Flow (Wiley, New York, 1982) p. 864.
2. T. Hayat, M.I. Khan, S. Qayyum, M.I. Khan, A. Alsaedi, J. Mol. Liq. 264, 375 (2018).
3. N.A. Khan, F. Naz, F. Sultan, Open Eng. 7, 185 (2017).
4. Y. Liu, Y. Jian, W. Tan, Int. J. Heat Mass Transfer 127, 901 (2018).
5. S. Qayyum, M.I. Khan, T. Hayat, A. Alsaedi, M. Tamoor, Int. J. Heat Mass Transfer 127, 933 (2018).
6. T. Hayat, S. Nawaz, A. Alsaedi, J. Mol. Liq. 248, 447 (2017).
Eur. Phys. J. Plus (2019) 134: 176 Page 13 of 13

7. M. Karimzadehkhouei, M. Shojaeian, A.K. Sadaghiani, K. Şendur, M.P. Mengüç, A. Koşar, Entropy 20, 242 (2018).
8. A. Gailitis, O. Lielausic, Appl. Magnetohydrodyn. Rep. Phys. Inst. Riga 12, 143 (1961).
9. T. Hayat, M. Khan, M. Imtiaz, A. Alsaedi, J. Mol. Liq. 225, 569 (2017).
10. R. Kumar, S. Sood, S.A. Shehzad, M. Sheikholeslami, J. Mol. Liq. 248, 143 (2017).
11. A. Anjum, N.A. Mir, M. Farooq, M.I. Khan, T. Hayat, Results Phys. 9, 1021 (2018).
12. M. Farooq, Aisha Anjum, T. Hayat, A. Alsaedi, J. Mol. Liq. 224, 1341 (2016).
13. M. Ayub, T. Abbas, M.M. Bhatti, Eur. Phys. J. Plus 131, 193 (2016).
14. D. Srinivasacharya, O. Surender, Prog. Eng. 127, 986 (2015).
15. S.R. Mishra, P.K. Pattnaik, G.C. Dash, Alex. Eng. J. 54, 681 (2015).
16. Ch. RamReddy, P.V.S.N. Murthy, A.M. Rashad, A.J. Chamkha, Eur. Phys. J. Plus 129, 25 (2014).
17. D. Srinivasacharya, O. Surender, Appl. Nanosci. 5, 29 (2015).
18. S. Ahmad, M. Farooq, M. Javed, A. Anjum, Results Phys. 8, 1250 (2018).
19. S. Ahmad, M. Farooq, M. Javed, A. Anjum, Results Phys. 9, 527 (2018).
20. S.U.S. Choi, Enhancing thermal conductivity of fluids with nanoparticles, in Developments and Applications of Non-
Newtonian Flows (ASME, 1995) pp. 99–105, FED-231/MD 66.
21. H. Saleh, E. Alali, A. Ebaid, J. Assoc. Arab Univ. Basic Appl. Sci. 24, 206 (2017).
22. T. Hayat, M.I. Khan, S. Qayyum, A. Alsaedi, M.I. Khan, Phys. Lett. A 382, 749 (2018).
23. T. Hayat, Faisal Shah, M.I. Khan, M.I. Khan, A. Alsaedi, J. Mol. Liq. 266, 814 (2018).
24. S.J. Liao, Homotopy Analysis Method in Non-Linear Differential Equations (Springer and Higher Education Press, Heidel-
berg, 2012).
25. N. Muhammad, S. Nadeem, T. Mustafa, Results Phys. 7, 862 (2017).
26. T. Hayat, S. Ali, M.A. Farooq, A. Alsaedi, PLoS ONE 10, e0129613 (2015).
27. T. Hayat, S. Ali, M. Awais, M.S. Alhuthali, Appl. Math. Mech. 36, 61 (2015).
28. T.R. Mahapatra, A. Gupta, Heat Mass Transfer 38, 517 (2002).
29. S. Pop, T. Grosan, I. Pop, Tech. Mech. 25, 100 (2004).
30. P. Sharma, G. Singh, J. Appl. Fluid Mech. 2, 13 (2009).

You might also like