Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

DOI 10.

1007/s10891-022-02613-9
Journal of Engineering Physics and Thermophysics, Vol. 95, No. 6, November, 2022

MATHEMATICAL STUDY OF HEAT TRANSFER


IN A STAGNATION FLOW OF A HYBRID NANOFLUID
OVER A STRETCHING/SHRINKING CYLINDER

T. Poornima,a P. Sreenivasulu,b and B. Souayehc,d UDC 532.529

The steady two-dimensional incompressible boundary-layer, stagnation-point flows of an ordinary nanofluid and a
hybrid CuO + MgO nanofluid over a stretching/shrinking cylinder were investigated using the dimensionless master
Prandtl equations. It was established that the thermal conductivity of the hybrid nanofluid is higher, it transfers a
larger amount of heat from the curved surface of the cylinder, and the skin friction force acting on its flow is larger,
as compared to those of the ordinary nanofluid, and that the skin friction force acting on the flows of these fluids
increases with increase in the volume fraction of nanoparticles in them.

Keywards: hybrid nanofluid, heat source, radiation, stagnation point, stretching/shrinking cylinder.
Introduction. Hybrid or composite nanofluids form a new class of functional fluids which include tiny particles
of volume less than 100 nm and are used in energy transfer processes. The classical base fluids for them are water
(H2O), ethylene glycol (EG), an H2O–EG compound, grease, a lamp oil, and a tuber oil. Among the hybrid nanoclusters
enhancing the energy transfer in fluids are MWCNT–Fe3O4/magnesia, SWCNT–magnesia, and Al2O3–Ag. Initial studies
of different nanofluids are presented [1–2]. The thermal convection in a MHD hybrid nanofluid with a porous medium was
investigated in [3]. The influence of an inclined magnetic field on the Marangoni radiative effect of the convection hybrid
nanofluid flow past a sheet was considered in [4]. The MHD hybrid nanofluid flow past a stretching/shrinking wedge was
investigated in [5].
Since the last decade, the viscous flows near a deformed/curved surface were analysed as applied to many
technological and industrial processes. In [6], the influence of suction and slip on the heat transfer in the flow past a cylinder
stretching or shrinking with time was determined. An unsteady convective mixed viscous flow around a cylinder was
considered in [7]. The effect of thermal radiation on the dissipative nanofluid flow past a cylinder with a deformable porous
surface was investigated in [8]. The nanofluid flow over a stretching/shrinking tubular surface was analysed in [9].
Exact elucidations of the equations for a stagnation-point slip flow past a cylinder are presented in [10]. The suction
at the stagnation point past a shrinking cylinder was examined in [11]. The effect of radiation on the stagnation-point flow
of a nanofluid past a permeable extending sheet with an energy rise/fall was considered in [12]. The inertial point flow of a
hybrid nanofluid near the deformable surface of a cylinder was investigated in [13].
The radiation of convective fluids finds enormous applications in the engineering, industrial, and public health
sectors. In [14], the MHD nanofluid thin film flow past a flat spinning disk dispersing a nonlinear radiation is presented.
The nonlinear radiation of the spherical hybrid nanoflow between two ring plates with a chemical reaction was investigated
in [15]. The authors of [16] focused on the MHD 3D flow of a hybrid nanofluid past an irregular-dimension surface with a
nonlinear radiation.
The release or absorption of energy plays a good role in enhancing the energy transfer from a system and, therefore,
these processes have real world applications. The nonuniform heat release/absorption of a microflow along an inclined
sheet was investigated in [17]. The effects of the Hall current and an exponential heat source in an unsteady dusty nanofluid
were considered in [18]. In [19], a thin film flow of a hydro-magnetic hybrid ferrofluid was investigated with account of the
nonuniform heat rise/drop in it. The employment of cupric oxide (CuO) and magnesia (MgO) in heat exchangers and in the
a
Department of Mathematics, SAS, Vellore Institute of Technology, Vellore, T.N., India; bDepartment of Humanities
and Science, SVEC, Tirupati, A.P., India; cDepartment of Physics, College of Science, King Faisal University, PO Box 400,
Al-Ahsa 31982, Saudi Arabia; d Department of Physics, Laboratory of Fluid Mechanics, Faculty of Sciences of Tunis,
University of Tunis El Manar, 2092, Tunis, Tunisia; email: bsouayeh@kfu.edu.sa. Published in Inzhenerno-Fizicheskii
Zhurnal, Vol. 95, No. 6, pp. 1471–1482, November–December, 2022. Original article submitted June 20, 2022.

0062-0125/22/9506-1443 ©2022 Springer Science+Business Media, LLC 1443


Fig. 1. Physical model of the problem: a) stretching cylinder; b) shrinking cylinder.

production of polymers as well as their antimicrobial and biocide properties motivated the authors of [20–21] to innovate
their applications.
The aim of the present work is numerical investigation of the energy transfer in the flow of a hybrid nanofluid,
comprising CuO + MgO nanoclusters suspended in an H2O–EG mixture, over a stretching/shrinking cylinder.
Mathematical Model. The steady two-dimensional incompressible boundary-layer, stagnation-point flows of an
ordinary nanofluid and a CuO + MgO hybrid nanofluid over a stretching/shrinking cylinder are considered. These fluids
comprise CuO and MgO nanoparticles suspended in an H2O–EG mixture (50–50%). The axial and radial directions of a
fluid flow over the cylinder are coincident with the polar cylindrical coordinates z and r, respectively. The flow pattern is
symmetric around the z axis and about the plane z = 0 with stagnation lines at z = 0 and r = a. The velocity of movement
of the surface of the cylinder is equal to ww(z) = 2bz. It is assumed that the cylinder is static at b = 0, is stretching at
b > 0, and is shrinking at b < 0. The velocity of the surrounding-medium stream is determined by the expression we(z) = 2cz
(c > 0). Diagrams of the fluid flows over the stretching and shrinking cylinders are presented in Fig. 1. The temperature of
the cylinder surface Tw and the ambient temperature T0 are assumed to be constant: Tw > T0.
A nonlinear thermal radiation and an asymmetric heat source/sink are incorporated in the problem on the hybrid
nanofluid flow being considered. It is assumed that the CuO and MgO nanoclusters are spherical in shape, have equal
sizes, do not agglomerate, and their mixture with water and ethylene glycol represents a stable compound. With the above
assumptions the boundary layer of this flow is defined by the equations [11]

 
( rw)  ( ru )  0 , (1)
z r

w w dwe 1   w 
w u  we   hnf r , (2)
z r dz r r  r 

T T 1   T  1 1 
w u   hnf r  ( rqr )  q (3)
z r r r  r  ( C p ) hnf r r
with the boundary conditions
r  a: w  ww , u  0, T  Tw ,
(4)
r  : w  we , T  T .

The thermal radiation released by this flow is defined in the Rosseland (diffusion) approximation [22]:

4  e T 4
qr   . (5)
3 k e r

1444
Since the energy gaps inside the flow are fairly small, Eq. (5) for the temperature T 4 can be expanded linearly with respect
to the temperature T∞. Leaving the higher-order terms in this equation, we obtain the expression

T  T4  4(T  T )T3  ....  4T3T  3T4 . (6)

To determine the effect of the nonlinear thermal radiation released by the fluid flow, we represent Eq. (5) in the form

16  eT3 T
qr   (7)
3 k e r

and rewrite Eq. (3) in view of Eqs. (5) and (7) as

T T 1   16 e 3   T 
w u   hnf 1  T  r   q . (8)
z r r r  3k e k f   r 

An asymmetrical rise or fall in the heat energy of the fluid flow was simulated using the expression [23, 24]

k f ww
q  ( A f ( )(Tw  T )  B  (T  T )) , (9)
z f

where A* > 0 and B* > 0 correspond to the internal heat generated in the fluid, and A* < 0 and B* < 0 correspond to the
internal heat absorption.
The fluid flow over the cylinder is defined by the expressions [5]

 hnf     k hnf
 2 s2   1 s1  (1  1 ) (1  2 ) ,  nf  ,
f f  f  (C p ) hnf

 hnf 1 (C p ) hnf  (C p )s1  (C p )s2


 ,  (1  2 ) (1  1 )  1   2 , (10)
f (1  1 ) (1  2 )5/2
5/2 (C p ) f  (C p ) f  (C p ) f

k hnf k  2k f  22 ( k f  ks2 ) ks1  2k f  21 ( k f  ks1 )


 s2 k nf , k nf  ,   1  2 .
kf ks2  2k f  2 ( k f  ks2 ) ks1  2k f  1 ( k f  ks1 )

Introducing the dimensionless variables [5]


2
r caf ( ) T  T (C p ) f
   , u   , w  2czf ( ) ,   , Pr  ,
a  Tw  T kf
(11)
2  2 
16sT03 Tw b ca 2
k f ba A k f ba B
R  , w  ,   , Re  , A , B 
3k e k f T0 c 2 f 2 f2 2 f2

and using Eqs. (9) and (10), we represent Eqs. (1) and (2) in the form

1   1s1  2s2   
( f )  (1  2 ) (1  1 )       Re  ff   ( f  )  1  0 ,
2
(12)
(1  1 ) 2.5 (1  2 ) 2.5   f  f 

 ks2  2k f  22 ( k f  ks2 ) 


 k k nf  ((1  R(  w  1)) 3 ( ) ) 
 s2  2k f  2 ( k f  ks2 ) 
(13)
  (C p )s1  (C p )s2 
 Pr  (1  2 ) (1  1 )  1   2  Re f   Af   B   0
 (C p ) f  (C p ) f 
 

1445
with the corresponding boundary conditions

f (1)   , f (1)  0 , (1)  1 ; f ( )  1 , ( )  0 . (14)

The important parameters of the hybrid nanofluid flow over the cylinder are its shear stress at the wall, determined
by the skin friction factor of the cylinder surface Cf, and the rate of local heat transfer in the flow, determined by its Nusselt
number Nux. These quantities are determined in the dimensional form by the expressions [25]

2 hnf  w  ak hnf  T 
Cf    , Nu x    r  qr  . (15)
 f we2  r  r  a k f (Tw  T ) r a

Using the similarity transform (12), we represent the indicated quantities in the dimensionless form:

 Re z  1 2k hnf
  Cf  f (1) , Nu x   [1  R(1  (  w  1)(1)) 3 ](1) . (16)
 a  (1  1 ) (1  2 ) 2.5
2.5 kf

Numerical Results and Discussion. The dimensionless coupled nonlinear ordinary differential equations (12)
and (13) with the boundary conditions (14) were solved using the shooting method including the Runge–Kutta integration
technique. The error of the numerical calculations performed with the Mathematica software was 10–5. The calculations
were performed for the flows of the ordinary nanofluid and the hybrid nanofluid over the stretching/shrinking cylinder
being considered for the following values of the parameters of the problem on these flows:  = 1.5 or –1.5, φ1 = 0.1,
φ2 = 0.1, Pr = 10, Re = 1, R = 3, θw = 1.3, A = 0.3, and B = 0.3. The thermal and physical properties of the base fluid and of
the CuO and MgO nanoparticles in it are presented in Table 1. The results of calculations of the parameters of the ordinary
nanofluid flow are compared with the analogous data obtained by other researchers in Table 2. As is seen from this table,
the results of our calculations are in good agreement with the results of analogous calculations performed in [5, 10]. The
results of calculations of the friction coefficient Cf (Re z/a) and the heat transfer coefficient Nu of the ordinary nanofluid
and hybrid nanofluid fluid flow are presented in Tables 3 and 4, and they allow the following conclusions. The skin friction
factors of both fluid flows increase with increase in the volume fraction of nanoparticles in the base fluid. The friction of
the hybrid nanofluid flow near the curved surface of the cylinder is higher than that of the ordinary nanofluid flow. The
Reynolds numbers of these fluid flows significantly influence their properties: an increase in Re of the flows causes their
friction factor to increase. The friction coefficient of the ordinary nanofluid flow is higher than the friction coefficient of the
hybrid nanofluid flow. The shear stresses of the flows at the surface of the cylinder decrease in the process of its stretching or
shrinking. The rate of heat transfer in the hybrid nanofluid flow is larger than that of the ordinary nanofluid flow. An increase
in the volume fraction of nanoclusters in the base fluid causes the thermal energy released from the system to increase. The
rate of heat transfer in the flows increases with increase in their Reynolds number. The heat transfer coefficient of the flows
decreases when the cylinder stretches or shrinks. The nonlinear distribution of the thermal radiation released by the flows is
favourable for increasing the heat transfer in them with increase in their thermal resistance. An increase in the asymmetric
energy rise parameter of the flows leads to a decrease in the rate of heat transfer in them.
The dependences of the velocity and temperature of the ordinary nanofluid and hybrid nanofluid flows over the
shrinking cylinder ( < 0) on the volume fraction of CuO nanoparticles in them are shown in Fig. 2. It is seen from this
figure that the velocities of these flows increase with increase in the volume fraction of CuO nanoparticles in them and that
the ordinary nanofluid flow is superior to the hybrid nanofluid flow. In this case, the thermal energy of a fluid flow decreases
with increase in the sizes of the CuO nanoparticles in it. It is seen from Fig. 3 that, as the sizes of the CuO nanoparticles
in a fluid flow over the stretching cylinder ( > 0) increase, the velocity of the flow decreases because the stretching of the
cylinder acts like a barrier which slows down the fluid motion, while the temperature of the flow increases. An increase
in the sizes of the CuO nanoparticles in a fluid flow leads to a decrease in its thermal resistance, and the energy boundary
layer of the flow expands. The domination of the hybrid nanofluid flow over the ordinary nanofluid flow in this case is seen
clearly, and this is due to the presence, in the hybrid nanofluid, of CuO + MgO clusters whose thermal conductivity is higher
than the thermal conductivity of the nanoparticles in the ordinary nanofluid.
The influence of the volume fraction of MgO nanoparticles in the ordinary nanofluid and hybrid nanofluid flows
over the shrinking cylinder ( < 0) on their momentum and energy is demonstrated in Figs. 4. It is seen that the velocities
of these flows decrease with increase in the volume fraction of MgO nanoparticles in them and that the ordinary nanofluid

1446
TABLE 1. Thermal and Physical Properties of the Base Fluid and the Nanoparticles in It [16, 19]

Base fluid Nanoparticles


Physical parameters
H2O–EG (50–50%) CuO MgO
ρ (kg/m3) 1057 6320 3580
Cp (J/kg∙K) 3287 531.8 960
k (W/m∙K) 0.424 76.5 48.4

TABLE 2. Comparison of the Values of f (1) and 2(1) Obtained in Different Works for a Nanofluid Flow

f (1) 2(1)
Re
[10] [11] [5] Present work [5] Present work
0.2 0.78604 0.786042 0.786042 0.786048 1.508635 1.508632
1.0 1.48418 1.484183 1.484183 1.484182 2.793424 2.793425
10 4.16292 4.162920 4.162920 4.162924 7.701472 7.701473

TABLE 3. Values of the Friction Coefficient Cf (Re z/a) of the Nanofluid Flow (2 = 0) and the Hybrid Nanofluid Flow
(2 = 0.05) at R = 3, Pr = 10, and Different Values of the Other Parameters of These Flows

Hybrid
1 Re  w A B Nanofluid
nanofluid
0.01 0.4 0.5 1.2 0.2 0.2 0.587862 0.64666
0.05 0.670536 0.731909
0.1 0.782239 0.847909
0.05 0.3 0.670536 0.748214
0.5 0.734695 0.819461
0.7 0.792730 0.883934
0.4 1.5 1.44614 1.63995
0 1.20766 1.34852
1.5 –0.78444 –0.87458

flow is superior to the hybrid nanofluid flow. The thermal energies of the flows decrease with increase in the volume fraction
of MgO nanoparticles in them, and the hybrid nanofluid surpasses the ordinary nanofluid in temperature due to its higher
thermal conductivity. Figure 5 shows the effect of the volume fraction of MgO nanoparticles in the ordinary nanofluid and
hybrid nanofluid flows over the stretching cylinder ( > 0) on their velocity and temperature. It is seen that an increase in
the volume fraction of MgO nanoparticles in these flows increases their momentum and thermal energy and that the hybrid
nanofluid flow transports a larger amount of heat compared to the heat transported by the ordinary nanofluid flow, whereas
the opposite effect is observed for the velocities of these flows.
As is seen from Fig. 6, the fluid motion over the shrinking cylinder ( < 0) enhances with increase in its Reynolds
number, which is explained by the fact that an increase in the Reynolds number of a fluid flow over this cylinder increases
the inertial forces rather than the viscous forces and, hence, the velocity of the fluid flow increases. Initially the thermal

1447
TABLE 4. Values of the Heat Transfer Coefficient of the Nanofluid Flow (2 = 0) and the Hybrid Nanofluid Flow (2 = 0.05)
at R = 3, Pr = 10, and Different Values of the Other Parameters of These Flows

1 Re  w A B Nanofluid Hybrid nanofluid


0.01 0.4 0.5 1.2 0.2 0.2 –5.69612 –5.04086
0.05 –5.03013 –4.29203
0.1 –4.11527 –3.26620
0.05 0.3 –5.03013 –4.09735
0.5 –2.99215 –2.18781
0.7 –1.23578 –0.48407
0.4 1.5 –2.03733 –0.93218
0 –4.89258 –3.88407
1.5 –5.04405 –4.22998
0.5 1.1 –5.52236 –4.71545
1.3 –4.45110 –3.40318
1.5 –3.01985 –1.71150
1.2 0.1 0.130737 0.957887
0.2 –5.03013 –4.09735
0.3 –10.16590 –9.12359
0.2 0.1 –0.93896 –0.12549
0.2 –5.03013 –4.09735
0.3 –10. 64980 –9.36685

energy of the fluid flow near the cylinder surface increases, and then the thermal energy of the ambient flow increases. In
this case, the ordinary nanofluid is superior to the hybrid nanofluid. The influence of the Reynolds number of the flows of
these fluids over the stretching cylinder (> 0) on their parameters is demonstrated in Fig. 7. It is seen that an increase in Re
of the flows causes their momentum and temperature to decrease and that the hybrid nanofluid flow surpasses the ordinary
nanofluid flow in momentum, whereas the hybrid nanofluid transports a smaller amount of heat energy compared to that of
the ordinary nanofluid.
Figure 8 shows that the heat energies of the ordinary nanofluid and hybrid nanofluid flows over the shrinking
cylinder ( < 0) decrease as their thermal radiation disperses because of the shrinking of the cylinder surface. Without doubt
the dispersion of the thermal radiation inside the flows in the process of stretching of the cylinder increases their heat energy.
In this case, the temperature of the ordinary nanofluid flow is higher than the temperature of the hybrid nanofluid flow due to
the lower thermal resistance of the hybrid nanofluid. As is seen from Fig. 9, the temperatures of the ordinary nanofluid and
hybrid nanofluid flows over the shrinking cylinder ( < 0) decrease with increase in their nonlinear radiation parameter and
that, in the case of fluid flow over the stretching cylinder ( > 0), the temperature of the ordinary nanofluid flow is higher
than the temperature of the hybrid nanofluid flow. In this case, the temperature of a fluid flow increases with increase in
its nonlinear radiation parameter, which is explained by the fact that the thermal energy of the fluid flow near the surface
of the cylinder increases, and the temperature of the ambient flow decreases as a result of the nonlinear dispersion of the
thermal radiation of the fluid flow. In the case where  > 0, the thermal resistance of the hybrid nanofluid flow is lower and
its thermal conductivity is higher compared to those of the ordinary nanofluid flow.
A release of the heat energy from a fluid flow over the shrinking cylinder ( < 0), defined by the asymmetric
parameter A, has a positive effect on this flow (Fig. 10). With increase in the parameter A, the temperature of the fluid
flow near the shrinking surface of the cylinder increases, and the temperature of the free stream decreases. In this case, the
temperature of the ordinary nanofluid flow is higher than the temperature of the hybrid nanofluid flow. In the case of fluid

1448
Fig. 2. Dependences of the velocity and temperature of the ordinary nanofluid (1) and
hybrid nanofluid (2) flows over the shrinking cylinder on the volume fraction of CuO
nanoparticles in them at  = –1.5.

Fig. 3. Dependences of the velocity and temperature of the ordinary nanofluid (1) and
hybrid nanofluid (2) flows over the stretching cylinder on the volume fraction of CuO
nanoparticles in them at  = 1.5.

Fig. 4. Dependences of the velocity and temperature of the ordinary nanofluid (1) and
hybrid nanofluid (2) flows over the shrinking cylinder on the volume fraction of MgO
nanoparticles in them at  = –1.5.

1449
Fig. 5. Dependences of the velocity and temperature of the ordinary nanofluid (1) and
hybrid nanofluid (2) flows over the stretching cylinder on the volume fraction of MgO
nanoparticles in them at  = 1.5.

Fig. 6. Dependences of the velocity and temperature of the ordinary nanofluid (1) and
hybrid nanofluid (2) flows over the shrinking cylinder on their Reynolds number at
 = –1.5.

Fig. 7. Dependences of the velocity and temperature of the ordinary nanofluid (1) and
hybrid nanofluid (2) flows over the stretching cylinder on their Reynolds number at
 = 1.5.

1450
Fig. 8. Dependences of the temperature of the ordinary nanofluid (1) and hybrid
nanofluid (2) flows over the shrinking (a) and stretching (b) cylinders on their radiation
parameter R: a) = –1.5; b) 1.5.

Fig. 9. Dependences of the temperature of the ordinary nanofluid (1) and hybrid
nanofluid (2) flows over the shrinking (a) and stretching (b) cylinders on their nonlinear
radiation parameter w: a)  = –1.5; b) 1.5.

Fig. 10. Dependences of the temperature of the ordinary nanofluid (1) and hybrid
nanofluid (2) flows over the shrinking (a) and stretching (b) cylinders on their asymptotic
energy release parameter A: a)  = –1.5; b) 1.5.

1451
Fig. 11. Dependences of the temperature of the ordinary nanofluid (1) and hybrid nanofluid
(2) flows over the shrinking (a) and stretching (b) cylinders on their heat source parameter
B: a)  = –1.5; b) 1.5.

Fig. 12. Dependences of the velocity of the ordinary nanofluid (1) and hybrid nanofluid (2)
flows over the shrinking (a) and stretching (b) cylinders on their parameter .

flow over the stretching cylinder ( > 0), the temperature of a fluid flow increases with increase in the nonuniform heat
source parameter. In this case, the hybrid nanofluid flow is superior to the ordinary nanofluid flow, which is explained by the
fact that the thermal conductivity of the clusters of the hybrid nanofluid is higher than that of the particles of the ordinary
nanofluid and, therefore, the heat energy inside the hybrid nanofluid flow increases the temperature of the radiation released
from it.
An increase in the asymmetric energy source parameters (A* > 0, B* > 0) is favourable for increasing the heat
energy inside a fluid flow over the stretching/shrinking cylinder due to the expansion of the thermal boundary layer on its
surface. This increase leads to an increase in the temperature of the fluid flow. Figure 11 shows that, in the case of fluid flow
over the shrinking cylinder ( < 0), the temperature of a fluid flow increases with increase in the heat source parameter B.
The results of the calculations for the base fluid and the nanoclusters selected are almost identical. The heat energy of the
ordinary nanofluid is higher than that of the hybrid nanofluid. In the case of fluid flow over the stretching cylinder ( > 0),
the temperature of a fluid flow increases with increase in the nonuniform heat source parameter, and the hybrid nanofluid
flow is superior to the ordinary nanofluid flow.
As is seen from Fig. 12, the velocity of a fluid flow over the stretching/shrinking cylinder decreases with increase
in the cylinder parameter . Under these conditions, an increase in  enhances the far away stream motion rather than the
shrinking or the stretching of the cylinder, slows down the expansion of the boundary layer, and improves the stress gradient
near the stagnation point. It is interesting that, when  increases, the shrinking of the cylinder activates and occupies its

1452
upper half, whereas the stretching of the cylinder decreases. In this case, the velocity of the hybrid nanofluid fow is higher
than that of the ordinary nanofluid flow. In the case where  > 0, the velocity of the ordinary nanofluid flow decreases with
increase in the rate of deformation of the cylinder surface, and the hybrid nanofluid flow is superior to the ordinary nanofluid
flow.

CONCLUSIONS
1. An increase in the volume fraction of CuO nanoparticles in the base fluid increases the rate of change in the skin
friction of the fluid flow over a shrinking/stretching cylinder and the rate of heat transfer in it.
2. A decrease in the velocity of a fluid flow over the stretching cylinder leads to an increase in its temperature, while
the opposite effect is characteristic for the fluid flow over the shrinking cylinder.
3. An increase in the volume fraction of MgO nanoparticles in the base fluid causes the velocity and temperature
of the fluid flow over the stretching cylinder to increase, and the opposite effect takes place in the case of this flow over the
shrinking cylinder.
4. The rate of energy transfer in the hybrid nanofluid flow over the stretching/shrinking cylinder is larger than that
of the ordinary nanfluid flow at large values of the nonlinear radiation parameter, while the opposite effect takes place at
large values of the nonuniform heat source/sink parameters A and B.
5. The coefficients of friction of the ordinary nanofluid and hybrid nanofluid flows over the stretching/shrinking
cylinder and the rate of heat transfer at the cylinder surface increase with increase in the Reynolds number of these flows.

NOTATION
a and b, constants; A* and B*, space- and time-dependent parameters; c, stretching ratio parameter; Cf, skin friction
factor; Cp, heat capacity, J/K; f, stream function; k, heat conductivity, W/(m∙K); ke, absorption coefficient; Nux, local Nusselt
number; Pr, Prandtl number; Re, local Reynolds number; qr, radiative heat flux, W/m2; r, radial coordinate, m; R, radiation
parameter; T, temperature, K; T0, hot fluid temperature, K; Tw , wall temperature, K; w and u, momentum components, m/s;
ww, stretching or shrinking velocity, m/s; we, free stream velocity, m/s; , thermometric conductivity, m2/s; , independent
variable; , temperature; θw, nonlinear radiation parameter; , stretching/shrinking parameter; , absolute viscosity, Pa∙s
or kg/(m∙s); , momentum diffusivity, m2/s; , density, kg/m3; e, Stefan–Boltzmann constant, W/(m2  K4); 1 and 2,
volume fractions of CuO and MgO nanoparticles. Subscripts: e, environment; f, fluid; nf, nanofluid; hnf, hybrid nanofluid;
r, radiative; s, solid; w, wall; 1 and 2, parameters of CuO and MgO particles, respectively.

REFERENCES
1. P. Sreenivasulu, T. Poornima, Reddy N. Bhaskar, and Reddy M. Gnaneswara, A numerical analysis on UCM Dissipated
nanofluid imbedded carbon nanotubes influenced by inclined Lorentzian force along with nonuniform heat source/sink,
J. Nanofluids, 8, No. 5, 1076–1084 (2019).
2. P. Sreenivasulu, T. Poornima, and R. Bhaskar, Influence of Joule heating and nonlinear radiation on MHD 3D dissipating
flow of Casson nanofluid past a nonlinear stretching sheet, Nonlin. Eng., 8, 661–672 (2019).
3. N. Biswas, N. K. Mann, and A. J. Chamkha, Effects of half-sinusoidal nonuniform heating during MHD thermal
convection in Cu–Al2O3/water hybrid nanofluid saturated with porous media, J. Therm. Anal. Calorim.; 143, No. 10
(2010); https://doi.org/10.1007/s10973-020-10109-y.
4. Qasem M. Al-Mdallal, N. Indumathi, B. Ganga, and A. K. Abdul Hakeem, Marangoni radiative effects of hybrid-
nanofluid flow past a permeable surface with inclined magnetic field, Case Studies Therm. Eng., 17, Article ID 100571
(2020).
5. I. Waini, A. Ishak, and I. Pop, MHD flow and heat transfer of a hybrid nanofluid past a permeable stretching/shrinking
wedge. Appl. Math. Mech. Eng. Ed., 41, 507–520 (2020).
6. Z. Abbas, S. Rasool, and M. M. Rashidi, Heat transfer analysis due to an unsteady stretching/shrinking cylinder with
partial slip condition and suction, Ain. Shams Eng. J., 6, 939–945 (2015).
7. T. Poornima, P. Sreenivasulu, and Reddy N. Bhaskar, Chemical reaction effects on an unsteady MHD mixed convective
and radiative boundary layer flow past a circular cylinder, J. Appl. Fluid Mech., 9, No. 6, 2877–2885 (2016).
8. Alok Kumar Pandey and Manoj Kumar, Natural convection and thermal radiation influence on nanofluid flow over a
stretching cylinder in a porous medium with viscous dissipation, Alex. Eng. J., 56, 55–62 (2017).

1453
9. N. C. Roşca, A. V. Roşca, and I. Pop, Nanofluid flow by a permeable stretching/shrinking cylinder, Heat Mass Transf.,
56, 547–557 (2020).
10. C. Y. Wang, Stagnation flow on a cylinder with partial slip — An exact solution of the Navier–Stokes equations, IMA
J. Appl. Math., 72, 271–277 (2007).
11. Y. Y. Lok and I. Pop, Wang's shrinking cylinder problem with suction near a stagnation point, Phys. Fluids. 23, Article
ID 083102 (2011).
12. Sreenivasulu P. and N. Reddy Bhaskar, Lie group analysis for boundary layer flow of nanofluids near the stagnation point
over a permeable stretching surface embedded in a porous medium in the presence of radiation and heat generation/
absorption, J. Appl. Fluid Mech., 8, No. 3, 549–558 (2015).
13. Najiyah Safwa Khashi' ie, Ezad Hafidz Hafidzuddin, Norihan Md Arifin, and Nadihah Wahi, Stagnation point flow of
hybrid nanofluid over a permeable vertical stretching/shrinking cylinder with thermal stratification effect, CFD Lett.,
12, No. 2, 80–94 (2020).
14. Zahir Shah, Abdullah Dawar, Poom Kumam, Waris Khan, and Saeed Islam, Impact of nonlinear thermal radiation on
MHD nanofluid thin film flow over a horizontally rotating disk, Casson nanofluid flow, Int. J. Appl. Comput. Math., 5,
No. 124 (2019).
15. Naveed Ahmed, Fitnat Saba, Umar Khan, Ilyas Khan, Tawfeeq Abdullah Alkanhal, Imran Faisal, and Syed Tauseef
Mohyud-Din, Spherically shaped (Ag–Fe3O4/H2O) hybrid nanofluid flow squeezed between two plates with nonlinear
thermal radiation and chemical reaction effects, Energies, 12, No. 76 (2019); https://doi.org/10.3390/en12010076.
16. Iskander Tlili, Hossam A. Nabwey, S. P. Samrat, and N. Sandeep, 3D MHD nonlinear radiative flow of CuO–MgO/
methanol hybrid nanofluid beyond an irregular dimension surface with slip effect, Sci. Rep., 10, Article ID 9181 (2020).
17. P. Sreenivasulua, T. Poornima, and P. Bala Anki Reddy, Soret and Dufour effects on MHD non-Darcian radiating
convective flow of micropolar fluid past an inclined surface with nonuniform surface heat source or sink and chemical
reaction, IOP Conf. Ser., Mater. Sci. Eng., 263, Article ID 062014 (2017).
18. B. Mahanthesh, N. S. Shashikuma, B. J. Gireesha, and I. L. Animasaun, Effectiveness of Hall current and exponential
heat source on unsteady heat transport of dusty TiO2–EO nanoliquid with nonlinear radiative heat, J. Comput. Design
Eng., 6, No. 4, 551–561 (2019).
19. Iskander Tlili, M. T. Mustafa, K. Anantha Kumar, and N. Sandeep, Effect of asymmetrical heat rise/fall on the film flow
of magnetohydrodynamic hybrid ferrofluid, Sci. Rep., 10, 66–77 (2020).
20. M. E. Grigore, E. R. Biscu, A. M. Holban, M. Cartelle Gestal, and A. M. Grumezescu, Methods of synthesis, properties
and biomedical applications of CuO nanoclusters, Pharmaceut. 9, No. 75 (2016); doi: 10.3390/ph9040075.
21. P. Askari, A. Faraji, G. Khayatian et al., Effective ultrasound-assisted removal of heavy metal ions As(III), Hg(II), and
Pb(II) from aqueous solution by new MgO/CuO and MgO/MnO2 nanocomposites, J. Iran. Chem. Soc., 14, 613–621
(2017).
22. E. M. Sparrow and R. D. Cess, Radiation Heat Transfer, Chapters 7, 10, and 19, Hemisphere, Washington, DC, USA
(1978).
23. B. J. Gireesha, R. S. R. Gorla, M. R. Krishnamurthy, and B. C. Prasannakumara, Biot number effect on MHD flow and
heat transfer of nanofluid with suspended dust particles in the presence of nonlinear thermal radiation and nonuniform
heat source/sink, Acta Comment. Univ. Tartuensis Math., 22, No. 1, 91–114 (2018).
24. B. J. Gireesha, G. S. Roopa, H. J. Lokesh, and C. S. Bagewadi, MHD flow and heat transfer of a dusty fluid over a
stretching sheet, Int. J. Phys. Math. Sci., 3, No. 1, 171–180 (2012).
25. T. A. Yusuf, F. Mabood, W. A. Khan, and J. A. Gbadeyan, Irreversibility analysis of Cu–TiO2–H2O hybrid-nanofluid
impinging on a 3D stretching sheet in a porous medium with nonlinear radiation: Darcy–Forchhiemer's model,
Alexandria Eng. J., 59, Issue 6, 5247–5261 (2020).

1454

You might also like