1 - Physica A

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Physica A 470 (2017) 330–344

Contents lists available at ScienceDirect

Physica A
journal homepage: www.elsevier.com/locate/physa

Stagnation point flow and heat transfer on a thin porous


sheet: Applications to flow dynamics of the circulatory
system
J.C. Misra a,∗ , A. Sinha b , B. Mallick a
a
Centre for Healthcare Science and Technology, Indian Institute of Engineering Science and Technology, Shibpur, Howrah - 711103,
India
b
Department of Mathematics, Yogoda Satsanga Palpara Mahavidyalaya, Purba Medinipur-721457, India

highlights
• Stagnation point flow of biofluids on a thin porous sheet is investigated.
• Effect of magnetic induction and heat transfer are accounted.
• Problem is solved numerically by developing finite difference scheme.
• Magnetic field plays an important role in controlling flow and heat transfer.

article info abstract


Article history: The paper is concerned with the modeling and analysis of stagnation point flow and
Received 22 July 2016 heat transfer on a thin porous sheet under the action of an induced magnetic field. The
Received in revised form 15 September fluid is considered to be incompressible viscous and electrically conducting. The study
2016
is motivated towards exploring some interesting phenomena in the micro-circulatory
Available online 24 October 2016
system. Heat transfer is considered to be governed by the heat equation. In order to take
care of the induced magnetism that affects the flow process, the flow equations are coupled
Keywords:
MHD stagnation point flow
with magnetic field variables. The analysis has been performed under the purview of the
Induced magnetic field boundary layer theory, together with the use of similarity transformation. The transformed
Stretching porous sheet equations are solved by developing an appropriate numerical method. Numerical results
Heat transfer have been computed for a typical situation of the fluid in motion. The results are displayed
Finite difference method graphically/in tabular form, which depict the distribution of velocity and temperature
under the action of the induced magnetic field and permeability of the porous sheet. The
study shows that the flow of the fluid reduces, as the strength of the induced magnetic
field increases. However, the reduction in velocity is accompanied by an enhancement of
the temperature field.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction

A stagnation point, as we know, refers to a point in a given flow field, at which the local velocity vanishes. It was
mentioned by Clancy [1] that the existence of stagnation points is very apparent on the surface of objects in the flow field,

∗ Corresponding author.
E-mail address: misrajc@gmail.com (J.C. Misra).

http://dx.doi.org/10.1016/j.physa.2016.10.051
0378-4371/© 2016 Elsevier B.V. All rights reserved.
J.C. Misra et al. / Physica A 470 (2017) 330–344 331

Nomenclature
V Velocity vector
H Induced magnetic field vector
(x, y) Cartesian coordinates of a representative point
(u, υ ) Velocity components along x-, y-axis respectively
(H1 , H2 ) Magnetic components along x, y-axis respectively
H0 Uniform magnetic field at infinity
He (x) x-component of the magnetic field at the edge
T0 Reference temperature
Tw Wall temperature
T∞ Ambient temperature
T Fluid temperature
ue x-component of the velocity at the edge
l Characteristic length
Pr Prandtl number
υ0 Suction/Injection velocity
s Suction/Injection parameter
Gr x Local Grashof number
Rex Local Reynolds number
P Magnetohydrodynamic pressure
p Fluid pressure
k⋆ Permeability of capillary wall
k1 Thermal conductivity
g⋆ Acceleration due to gravity
cp Specific heat at constant pressure
Q Volumetric rate of heat generation/absorption
uw Wall velocity
k Permeability parameter
Cf Skin friction coefficient
τw Wall shear stress
Nu Nusselt number
qw Heat flux at the wall
a Positive constant proportional to free stream velocity
c Proportionality constant of the velocity of the sheet

Greek Symbol
µe Magnetic permeability
ρ Fluid density
ν Kinematic viscosity
βT Volumetric coefficient of heat transfer
ψ Stream function
η Similarity variable
η0 Magnetic diffusivity
β Magnetic parameter
λ1 Buoyancy parameter
λ Reciprocal magnetic Prandtl number
µ Dynamic viscosity
α Heat source parameter
κ Thermal conductivity

Subscript
w Condition at the surface
∞ Condition at infinity

Superscript

Differentiation with respect to η
332 J.C. Misra et al. / Physica A 470 (2017) 330–344

in which the object is responsible for reducing the fluid velocity to zero. Static pressure of the fluid (which is maximum at
the stagnation point) is often called the stagnation pressure and the temperature at a stagnation point is referred to as the
stagnation temperature (cf. Fox and Mc Donald [2]). It may be mentioned further that at a stagnation point, the whole kinetic
energy is converted to internal energy, since the fluid velocity is zero at the stagnation point. Van Wylen and Sonntag [3]
made an observation that at all points on the streamline leading to the stagnation point, the stagnation temperature equals
the total temperature.
Owing to their importance in engineering and industry as well as in the study of physiological fluid dynamics, several
researchers dwelt upon the dynamics of stagnation flow/heat transfer under various situations. The two-dimensional flow
near a stagnation point was discussed by Beard and Walters [4] in the context of elastico-viscous boundary-layer flows. The
laminar two-dimensional flow of a non-Newtonian fluid (second grade fluid) in the vicinity of a stagnation point was studied
by Ariel [5]. Stagnation point flows/heat transfer in plates were analyzed by Wang [6,7], by considering velocity-slip. On the
basis of these studies, the author found that the stagnation flow and heat transfer are significantly affected by velocity-slip
and mentioned that such studies on stagnation flow have applications in situations, where a thin film of light oil is attached
to the plate, or when the plate has coatings, like a thick monolayer of hydrophobic octadecyltrichlorosilane (OTS).
Apart from various applications of studies on boundary layer behavior in different physiological processes, they have a
variety of applications in many industrial processes. They are of particular importance in the production of paper and glass-
fibers, spinning of laments, hot rolling, casting and cooling of metallic sheets, electronic chips, polymer melts and solutions,
crystal growth processes. The effects of variation in certain fluid properties on hydromagnetic flow and heat transfer over
a stretching sheet were studied by Prasad et al. [8]. Two dimensional mixed convection equations were employed by Hayat
et al. [9] in their study of stagnation point flow in the presence of thermal radiation. The study was performed for a situation,
when a vertical stretching surface is embedded in a fluid saturated porous matrix. The problem was solved using homotopy
analysis method. Two dimensional steady stagnation point flow of a non-Newtonian fluid (second grade) impinging on a
rigid wall was studied by Labropulu and Li [10], by considering Navier-slip. The effect of fluid viscosity on the stagnation
point flow was emphasized in particular. Malvandi et al. [11] discussed slip effects on unsteady stagnation point flow of a
nano fluid on a stretching sheet, by considering heat transfer. They made an observation that velocity-slip has an enhancing
effect on the heat transfer rate. Steady stagnation point flow of an incompressible viscous fluid over a stretching sheet was
studied by Sinha and Misra [12]. The associated problem of heat transfer was also paid due consideration. They made an
important observation that the temperature gradient diminishes on the surface of the sheet with an increase in Prandtl
number. References of some previous publications on stagnation point flows are also available in Ref. [12].
Importance of studies on stagnation point flows in physiological fluid dynamics has been discussed recently by Humphrey
and Rourke [13] and Ambrosi et al. [14]. While performing the simulation of physiological flows, Roldan-Alzate [15] studied
stagnation flows also. On the basis of his study, he presented the streamlines indicating the stagnation region. In the context
of several studies on the dynamics of blood flow, stagnation point flows have been investigated by several authors. It is
known that platelets constitute an important component of whole blood. For examining platelet decomposition, stagnation
point flow of blood was studied experimentally by Affeld et al. [16], by considering flow dynamics of platelet-rich-plasma
(PRP), whose flow characteristics were studied by the use of computational methods. Platelet flow along the wall was
estimated by using boundary layer theory. The experimental results reported in this paper reveal that platelets are deposited
at a particular shear rate. By accounting for convection, surface reaction and diffusion, platelet adhesion in stagnation point
flow was studied theoretically by David et al. [17]. The study shows that so long as the reaction rate remains constant, the
maximum platelet flux occurs at the stagnation point streamline.
Studies on flows through porous media/permeable membranes bear the prospect of huge applications in biomechanical
and biochemical engineering, cardiovascular fluid dynamics and in the design and construction of artificial organs. Organs
in the human body that are porous, include skin, bones, lungs, gall bladder, bile duct, arterioles and capillaries. Porosity in
these organs increases in the pathological state. It has been mentioned by Levick [18] that the permeability of capillaries
to small lipophobic solutes like potassium and sodium ions is found to have wide variations, so far as different organs of
the human body are concerned. We may assert that although ‘lipophobicity’ (also called lipophobia) literally means ‘fear of
fat’, it refers to a property of chemical compounds, which means ‘fat rejection’. Lipophobic materials are those which have
no affinity for fats/organic solvents. Levick [18] further mentioned that the permeability of capillaries to water and small
lipid-insoluble molecules is much more than that of the membranes of most cells. However, from his argument it appears
that the physiological variation in permeability from one organ to another may not be due to variation in the dimension of
different pores, but it is likely to be due to differences in the pore density (i.e. ratio between the area occupied by all the pores
and the total area); it may also be due to variation in the length of the pores. In a recent book, Vadasz [19] has emphasized
upon studies dealing with porous media approach and their potential applications in various physiological processes of the
human body, including flow of liquids in biological tissues, e.g. the human brain and cardiovascular flow of blood in the
human circulatory system.
The application potential of studies on MHD flow of fluids in hemodynamics/physiological fluid dynamics has been
discussed by Misra and his collaborators [20–23]. Importance of consideration of slip velocity in model studies pertaining
to the flow of blood in the micro-circulatory system has been emphasized in several publications of different researchers,
including Misra and Adhikary [24]. By developing a mathematical model, the authors [24] showed that blood viscoelasticity
lowers flow velocity significantly and that heat transfer rate reduces with increasing Prandtl number. Recently, a couple
J.C. Misra et al. / Physica A 470 (2017) 330–344 333

Fig. 1. Physical sketch of the problem.

of theoretical studies of flow and heat transfer through porous media, having applications to flow in micro-vessels were
reported by Shit et al. [25,26].
Being motivated by all the discussions made above, the present study aims at exploring some information regarding
the MHD stagnation point flow and heat transfer during blood flow in the micro-circulatory system. For this purpose, a
theoretical model has been formulated and analyzed mathematically by developing a computational method that is suitable
for finding the numerical estimates of some physical variables associated with the present investigation. The physical model
consists of MHD stagnation point flow and heat transfer of an electrically conducting fluid over a sheet that is in stretching
motion. The effects of induced magnetic field and injection/suction have been duly taken into account. This model is fairly
applicable to the study of stagnation point flow and heat transfer in capillaries, even in the normal physiological state, under
the physical conditions stated earlier, because the capillaries of the micro-circulatory system are in a state of stretching
motion and can be approximated as stretching sheets. The study shows that the stagnation point flow dynamics and heat
transfer are affected significantly by the induced magnetic field and the permeability of the capillary wall. It also indicates
that with an increase in the Prandtl number, thermal boundary layer reduces, but it increases as the value of the heat source
parameter increases.

2. Mathematical Formulation of the model

With reference to Fig. 1, let us take up the study of flow and heat transfer on a sheet, which is porous and is in a state
of stretching motion. As shown in the figure, two-dimensional Cartesian coordinates (x, y) have been used, with the axis of
x along the horizontal direction and y-axis along the transverse direction; the origin O being taken at the stagnation point
of the flow medium. The components of fluid velocity along and perpendicular to the axis of the sheet will be denoted by u
and υ respectively. A physical sketch of the problem is shown in Fig. 1.
Formulation of the model is performed on the basis of assumptions listed below.

(a) Velocity of the external fluid and the temperature at the surface are proportional to the distance from the stagnation
point O (origin of the coordinate system).
(b) A uniform induced magnetic field of strength H0 is applied in the transverse direction, such that at the wall the parallel
component equals H0 , while the normal component is zero.

In view of the assumption (a), we can write

ue (x) = ax
x
and Tw (x) = T∞ + T0 (1)
l

a being a positive constant, with dimension (time)−1 , proportional to free stream velocity far away from the stretching
sheet, T∞ the ambient temperature, l the characteristic length while T0 (x), Tw (x) are the reference temperature and wall
temperature respectively at a distance x.
334 J.C. Misra et al. / Physica A 470 (2017) 330–344

The basic equations for the flow of a viscous, electrically conducting, incompressible fluid can be written in the following
form:
∇ · V = 0, ∇ ·H=0 (2)
µe 1  1 
(V · ∇)V − (H · ∇)H = − ∇ P + ν ∇ 2 V − ⋆ V + g ⋆ βT (T − T∞ ) (3)
4π ρ ρ k
∇ × (V × H) = η0 ∇ 2 H (4)
k1 Q
(V · ∇)T = ∇ 2T + (T − T∞ ) (5)
ρ cp ρ cp
where V is the fluid velocity vector, H is the induced magnetic field vector, P = (p+µe |H|2 /8π ) is the magnetohydrodynamic
pressure, p is the fluid pressure, µe denotes the magnetic permeability, ν is the kinematic coefficient of viscosity, T is the
temperature, T∞ is the temperature of the ambient fluid, k⋆ is the permeability of the capillary wall, η0 is the magnetic
diffusivity, g ⋆ being the gravitational acceleration, βT the volumetric coefficient of expansion of heat transfer, ρ is the density,
k1 is the thermal conductivity, Q is the volumetric rate of heat generation or absorption and cp is the specific heat at constant
pressure.
Under the assumptions stated above, using Boussinesq approximation the flow and heat transfer are governed by the
following equations:
∂ u ∂υ
+ =0 (6)
∂x ∂y
∂ H1 ∂ H2
+ =0 (7)
∂x ∂y
∂u ∂υ µe ∂ H1 ∂ H1 µe ∂ 2u νu
 
due dHe
u +υ − H1 + H2 = ue − He + ν 2 − ⋆ + g ⋆ βT (T − T∞ ) (8)
∂x ∂y 4π ρ ∂x ∂y dx 4π ρ dx ∂y k

∂ H1 ∂ H1 ∂u ∂u ∂ 2 H1
u +υ − H1 − H2 = η0 2 (9)
∂x ∂y ∂x ∂y ∂y
∂T ∂T k1 ∂ 2 T Q
u +υ = + (T − T∞ ). (10)
∂x ∂y ρ cp ∂ y2 ρ cp
In the above equations H1 and H2 are the induced magnetic field components along the x and y directions respectively (cf.
Fig. 1). ue and He are respectively the velocity and magnetic field components at the edge of the boundary layer.
The boundary conditions of Eqs. (6)–(10) are as follows:
∂ H1 x
u = uw = cx, υ = −υ0 , = H2 = 0, T = Tw = T∞ + T0 at y = 0
∂y l
x
u = ue (x) = ax, H1 = He (x) = H0 , T = T∞ at y → ∞ (11)
l
where c is proportionality constant of the velocity of the stretching sheet, having dimension (time)−1 and υ0 is a constant
velocity. Here, H0 is the uniform magnetic field at infinity and l is the characteristic length.

3. Relevance of the model to hemodynamical studies

One of the serious cardiovascular diseases is the formation of thrombus (accumulation of platelets of blood) in the lumen
of arteries. Due to formation of thrombus the arterial lumen is often occluded and blood flow is disturbed. While discussing
the role of the dynamics of blood flow in the deposition of platelets, Bluestein et al. [27] made an observation that formation
of thrombus can be one of the consequences of stagnation flow. For examining the growth of a thrombus, stagnation flow
geometry has been found to be useful (cf. Petschek et al. [28]). In the case of stagnation point flow, the location of thrombus
formation has been found to be strongly dependent on flow conditions [29]. Stagnation point flow experiments for blood
were carried out by Wurzinger et al. [30] with a motivation of bringing about reduction in platelet aggregation before
adhesion takes place. Affeld et al. [31] performed an experiment on platelet rich plasma, on the basis of which they made a
conjecture that maximum number of platelets are deposited at a point where the shear rate has a critical value. Hazel and
Pedley [32] carried out an investigation on time-averaged mean wall shear stress in the vicinity of an oscillatory stagnation
point, on the basis of which they made an observation that time-averaged wall shear stress is considerably changed, if the
motion of the stagnation is accounted for.
All these experimental/theoretical observations reported by previous researchers emphasize the importance of studies
on stagnation point flows in the area of hemodynamics. Results of the present investigation bear the promise of important
J.C. Misra et al. / Physica A 470 (2017) 330–344 335

applications in clinical sciences and medical technology, particularly when heat transfer is very prominent during stagnation
point flow of blood in normal physiological/pathological states in the presence of thrombus, which might have been formed
due to accumulation of platelets of blood. Since heat transfer during stagnation point flow of blood has hardly been taken into
account in previous researches, this communication can be regarded as a pioneering study in physiological fluid dynamics.

4. Analysis

∂ψ ∂ψ
The stream function ψ is defined by u = ∂ y and υ = − ∂ x by which the continuity equation is automatically satisfied.
To facilitate numerical solutions, we introduce the following similarity by transformations:

1 T − T∞  c  12 H0 x ′  ν  12
ψ = (c ν) 2 xf (η), θ= , η= y, H1 = g (η), H2 = − H0 g (η). (12)
Tw − T∞ ν l cl2
Using (12), from Eqs. (8)–(10) we have the following set of non-linear differential equations:

a2 f′
f ′′′ + ff ′′ − (f ′ )2 + − + β[(g ′ )2 − gg ′′ − 1] + λ1 θ = 0 (13)
c2 k
λg ′′′ + fg ′′ − f ′′ g = 0 (14)
1
(θ ′′ + α) + f θ ′ − θ f ′ = 0. (15)
Pr
Also by using (8) the boundary conditions defined in (11) assume the following forms:
υ0
f ′ (0) = 1, f (0) = 1
= s, g (0) = 0, g ′′ (0) = 0, θ (0) = 1
(c ν) 2

a
f ′ (∞) = , g ′ (∞) = 1, θ (∞) = 0. (16)
c
In (12)–(16), primes denote derivatives with respect to η, k is the permeability parameter, β the magnetic parameter, λ1
the constant buoyancy parameter. Gr x and Rex are local Grashof number and local Reynolds number respectively, λ the
reciprocal magnetic Prandtl number, Pr the Prandtl number and α the heat source parameter. These are given by
 2
ck⋆ µe H 0 g ⋆ βT (Tw − T ∞)x3 /ν 2 Gr x
k= , β= , λ1 = = ,
ν 4π ρ lc x4 c 2 /ν 2
Re2x
η0 µcp Qν
λ= , Pr = , α= . (17)
ν k1 ck1
Some more physical quantities of interest in the present study are the skin friction coefficient Cf and the local Nusselt number
Nu defined by:
τw xqw
Cf = , Nu = (18)
ρ u2w k(Tw − T∞ )
in which, τw stands for the wall shear stress in the y direction and qw for the heat flux at the wall, given by
 ∂u   ∂T 
τw = µ , qw = −κ (19)
∂y y=0 ∂y y=0

where µ and κ are respectively the dynamic viscosity and the thermal conductivity of the fluid.
Using non-dimensional parameters, we obtain
1
− 12
Cf Rex2 = f ′′ (0), Nu Rex = −θ ′ (0). (20)
cx2
Rex = ν = uwν x being the local Reynolds number.

5. Numerical procedure

Our numerical procedure is based on a finite difference scheme with central differences along with an iterative procedure
(cf. Kafoussias and Williams [33]).
Writing f ′ = F and g ′ = G in Eqs. (13) and (14), we have

a2 F
F ′′ + fF ′ − F 2 + − + β[G2 − gG′ − 1] + λ1 θ = 0 (21)
c2 k
λG′′ + fG′ − F ′ g = 0. (22)
336 J.C. Misra et al. / Physica A 470 (2017) 330–344

Using the same notations, the boundary conditions (16) for Eqs. (21)–(22) read
υ0
F (0) = 1, f (0) = 1
= s, g (0) = 0, G′ (0) = 0
(c ν) 2
a
F (∞) = , G(∞) = 1. (23)
c
We write the central difference scheme for derivatives with respect to η as follows:
Vi+1 − Vi−1
(V ′ )i = + O((δη)2 )
2δη
Vi+1 − 2Vi + Vi−1
(V ′′ )i = + O((δη)2 ). (24)
(δη)2
In (24), V stands for F , G and θ ; i for the grid index in η-direction, where ηi = i ∗ δη; i = 0, 1, . . . , m and δη stands for the
increment along the η-axis. To linearize the discretized equations, we apply Newton’s Linearization method:
When the values of the dependent variables at the nth iteration are known, the corresponding values of these variables
at the next iteration can be obtained by using equation

Vin+1 = Vin + (1Vi )n (25)


in which (1Vi ) represents the error at the nth iteration, i = 0, 1, 2, . . . , n. On the boundary, (1Vi ) = 0, since the values
n n

at the boundary are known.


Applying (24) in (15) and (21)–(22) and making use of (25) in (21) and finally dropping the quadratic terms in (1Vi )n ,
we obtain a system of block tri-diagonal equations as follows:

Ai 1Fin−+11 + Bi 1Fin+1 + Ci 1Fin++11 = Dni ,


Qi Gin−+11 + Ri Gin+1 + Si Gni++11 = Tin , (26)
θ
Li in−+11 + Mi in+1 θ + θ
Ni in++11 = Oni ,
where,
4.0
Ai = 4 − 2fin δη, Bi = −8 − 8Fin δη2 − δη2 , Ci = 4 + 2fin δη,
k
  a
Di = β 4δη2 + 4gin (gin+1 − 2gin + gin−1 ) − (gin+1 − gin−1 )2 − 4λ1 θin δη2 − 4 δη2
c
−Ai Fin−1 − (Bi + 4Fin δη2 )Fin − Ci Fin+1 , (27)
 
Qi = 2λ − fi δη, Ri = −4λ,
n
Si = 2λ + fi δη, Ti = fi+1 − 2fi + fi−1 ,
n
2gin n n n

1  4 1  2αδη2
Li = 2 − Prfin δη , Mi = − − (fi−
n
1 − fi+1 )δη,
n
Ni = 2 + Prfin δη , Oi = − .
Pr Pr Pr Pr
Corresponding boundary conditions are:
f0n = s, F0n = 1, g0n = 0, Gn0 = Gn1 , θ0n = 1
a a
fmn = fmn−1 + δη, Fmn = , n
gm = gmn −1 + δη, Gnm = 1, θmn = 0. (28)
c c
These equations can be solved by using ‘‘Thomas algorithm’’. The error committed during the determination of the
distribution of the function f (η) is the difference between the calculated values of f (η) at two consecutive operations, say
(n + 1)th and nth. In the present study, the error ϵ = |f n+1 (η) − f n (η)| and is estimated to be less than 10−6 .

6. Discussion

In this section, the computational estimates of different dimensionless parameters involved in the present study,
have been presented in graphical/tabular forms. Using the numerical procedure described in the preceding section,
we solved the system of Eqs. (13)–(15) subjected to the boundary conditions (16). Here we have made use of the
numerical values of different dimensionless parameters that are available in scientific literatures published by previous
experimental/theoretical researchers. The data used for the computation in the present study are given in Table 1.
Numerical computation of the present study was executed by taking δη = 0.0125 with 441 grid points. The
computational results have been displayed graphically through Figs. 2–16. Distributions of axial velocity (f ′ ) for different
values of the magnetic parameter β , the ratio a/c, suction parameter s, permeability parameter k and local buoyancy
parameter λ1 are presented in Figs. 2, 5, 8, 10 and 11 respectively, while Figs. 3, 6 and 9 respectively indicate the changes that
J.C. Misra et al. / Physica A 470 (2017) 330–344 337

Table 1
Parametric values used in the present analysis.
Parameter Symbol Value Reference number

Permeability parameter k 0.05, 0.1, 0.15, 0.2 (cf. Ref. [35])


Suction parameter s −1.0, −0.5, 0.5, 1.0 (cf. Ref. [36])
Magnetic parameter β 0.0, 2.0, 4.0, 6.0 (cf. Ref. [34])
Buoyancy parameter λ1 −7.0, −1.0, 5.0, 10.0 (cf. Ref. [37])
Reciprocal magnetic Prandtl number λ 1.0, 5.0, 10.0, 15.0 (cf. Ref. [12])
Prandtl number Pr 20.0, 21.0, 22.0, 25 (cf. Refs. [38,39])
Heat source parameter α 0.0, 1.0, 3.0, 4.0 (cf. Ref. [40])
a/c 2.0, 2.5, 3.0, 4.0, 5.0 (cf. Ref. [34])

Fig. 2. Velocity distribution for different values of β when s = 0.5, λ1 = 5.0, Pr = 21.0, λ = 1.0, a/c = 2.5, k = 0.1, α = 1.0.

Fig. 3. Variation of induced magnetic field in x-direction for different values of β when s = 0.5, λ1 = 5.0, Pr = 21.0, λ = 1.0, a/c = 2.5, k = 0.1,
α = 1.0.

take place in the induced magnetic field distribution when the values of the parameters β, a/c and s alter. The temperature
distributions for varied values of β, a/c , Pr and α are presented through Figs. 4, 7, 12 and 13 respectively. Fig. 14 depicts the
effects of β and λ on local skin friction and Fig. 15 illustrates the variation of local Nusselt number due to change in α and Pr.
In order to establish the validity of the present study, we made a comparison between the results of the present study and
those of Ali et al. [34]. The comparative study reported in Fig. 16 shows that there is an excellent agreement of our results
with those presented in Ref. [34].
Fig. 2 shows that the fluid velocity decreases with an increase in the value of the magnetic parameter β . One finds an
exactly similar change in the induced magnetic field for the change in β (Fig. 3). However, Fig. 4 reveals that increase in the
value of β brings about an enhancement of the temperature field. A look into the physics of the problem suggests that with
increase in β , the Lorentz force associated with the magnetic field increases and this may be interpreted as being responsible
for making the boundary layer thinner and for enhancing the temperature field.
338 J.C. Misra et al. / Physica A 470 (2017) 330–344

Fig. 4. Temperature distribution for different values of β when s = 0.5, λ1 = 5.0, Pr = 21.0, λ = 1.0, a/c = 2.5, k = 0.1, α = 1.0.

Fig. 5. Change in velocity distribution for different values of a/c when s = 0.5, λ1 = 5.0, Pr = 21.0, λ = 1.0, β = 2.0, k = 0.1, α = 1.0.

Fig. 6. Variation of induced magnetic field in x-direction for different values of a/c when s = 0.5, λ1 = 5.0, Pr = 21.0, λ = 1.0, β = 2.0, k = 0.1,
α = 1.0.

From Fig. 5, one may observe that the boundary layer thickness increases, as the value of the ratio a/c increases. Here we
can make an important observation that although the boundary layer thickness increases with a/c increasing, the boundary
layer structure is altered, when the value of a/c exceeds 3. This may be attributed to the fact that for a fixed value of ‘c’
corresponding to the stretching of the sheet, the increase in ‘a’ implies an enhancement in the stretching motion in the
J.C. Misra et al. / Physica A 470 (2017) 330–344 339

Fig. 7. Change in temperature distribution for different values of a/c when s = 0.5, λ1 = 5.0, Pr = 21.0, λ = 1.0, β = 2.0, k = 0.1, α = 1.0.

Fig. 8. Velocity distribution for different values of s when a/c = 2.5, λ1 = 5.0, Pr = 21.0, λ = 1.0, β = 2.0, k = 0.1, α = 1.0.

Fig. 9. Variation of induced magnetic field in x-direction for different values of s when a/c = 2.5, λ1 = 5.0, Pr = 21.0, λ = 1.0, β = 2.0, k = 0.1,
α = 1.0.

vicinity of the stagnation region and as a result, the acceleration of the external stream increases. Fig. 6 shows that with an
increase in the value of the ratio a/c, the induced magnetic field strength increases. The initial formation of the respective
profile is different for the two cases, when a/c < 3 and a/c > 3. In the former case, one can observe concave initial
340 J.C. Misra et al. / Physica A 470 (2017) 330–344

Fig. 10. Velocity distribution for different values of k when a/c = 2.5, λ1 = 5.0, Pr = 21.0, λ = 1.0, β = 2.0, s = 0.5, α = 1.0.

Fig. 11. Velocity distribution for different values of λ1 when a/c = 2.5, λ = 1.0, Pr = 21.0, k = 0.1, β = 2.0, s = 0.5, α = 1.0.

Fig. 12. Change in temperature distribution for different values of Pr when s = 0.5, λ1 = 5.0, a/c = 2.5, λ = 1.0, β = 2.0, k = 0.1, α = 1.0.

formation, while a convex formation is seen for the latter one. From Fig. 7 that gives the temperature distribution in the
fluid for different values of a/c, it is worthwhile to notice that the thermal boundary layer thickness reduces with increasing
a/c.
J.C. Misra et al. / Physica A 470 (2017) 330–344 341

Fig. 13. Change in temperature distribution for different values of α when s = 0.5, λ1 = 5.0, a/c = 2.5, λ = 1.0, β = 2.0, k = 0.1, Pr = 21.0.

Fig. 14. Variation in skin friction β for different values of λ, when a/c = 2.5, s = 0.5, λ1 = 5.0, k = 0.1, α = 1.0, Pr = 21.0.

Fig. 15. Variation in local Nusselt number with α for different values of Pr, when s = 0.5, λ1 = 5.0, a/c = 2.5, λ = 1.0, β = 2.0, k = 0.1.

Fig. 8 elucidates the influence of the suction parameter s on fluid velocity distribution. Both the cases, (i) s < 0 and
(ii) s > 0, which correspond to injection and suction respectively, have been investigated here. This figure reveals that as
the value of the suction parameter s increases, the fluid velocity reduces up to a certain point, beyond which an opposite
342 J.C. Misra et al. / Physica A 470 (2017) 330–344

Fig. 16. Comparison of results (velocity) between the present study and the Ref. [34].

Table 2
1
− 12
Values of Cf Rex2 and Nu Rex for variation of β, k, s, λ1 , α, Pr when λ = 1.0 and a/c = 2.5.
1
− 21
β k s λ1 α Pr Cf Rex2 Nu Rex

0.0 0.1 0.5 5.0 1.0 21.0 −1.1584 11.6713


2.0 0.1 0.5 5.0 1.0 21.0 −1.6659 11.5275
4.0 0.1 0.5 5.0 1.0 21.0 −2.1710 11.3363
2.0 0.05 0.5 5.0 1.0 21.0 −3.3241 11.1099
2.0 0.1 0.5 5.0 1.0 21.0 −1.6659 11.5275
2.0 0.15 0.5 5.0 1.0 21.0 −0.8005 11.7356
2.0 0.1 −0.5 5.0 1.0 21.0 −0.8565 1.8084
2.0 0.1 0.5 5.0 1.0 21.0 −1.6659 11.5275
2.0 0.1 1.0 5.0 1.0 21.0 −1.9118 18.7785
2.0 0.1 0.5 −2.0 1.0 21.0 −2.1581 11.4501
2.0 0.1 0.5 2.0 5.0 21.0 −1.8747 11.4956
2.0 0.1 0.5 5.0 5.0 21.0 −1.6659 11.5275
2.0 0.1 0.5 5.0 0.0 21.0 −1.7563 12.3075
2.0 0.1 0.5 5.0 1.0 21.0 −1.6659 11.5275
2.0 0.1 0.5 5.0 3.0 21.0 −1.5006 10.080
2.0 0.1 0.5 5.0 1.0 19.0 −1.6383 10.5900
2.0 0.1 0.5 5.0 1.0 21.0 −1.6659 11.5275
2.0 0.1 0.5 5.0 1.0 22.0 −1.6783 11.9892

trend is noticed. However, the reduction takes place at a faster rate in the case of injection than in the case of suction. Fig. 9
shows that a similar change occurs for the case of injection (s < 0) for induced magnetic field, but one may observe that
there is always a rise in magnetic induction for increasing suction (s > 0).
The influence of the permeability of the porous sheet on which the fluid flow and heat transfer take place is illustrated
in Fig. 10. It is revealed that permeability affects the fluid velocity quite significantly. The fluid velocity is greatly enhanced
when there is an increase in the permeability parameter. However, it may be noted that when the permeability is small
(k = 0.05), the fluid velocity reduces along the axis at a rate faster than when the permeability is higher. Fig. 11 gives several
plots depicting the variation in the velocity distribution for different values of the mixed convection parameter λ1 , when
the values of others parameters are kept fixed. Although the nature of velocity distribution is found not to be significantly
affected by the values of λ1 (in the range studied here), one can notice that there is a sharp rise in the velocity in the vicinity
of the wall.
Fig. 12 gives a picture of the change in temperature distribution in the thermal boundary layer due to change in the
Prandtl number (Pr). This figure shows that a rise in Prandtl number brings about a reduction in the thermal boundary
layer thickness. One can further observe from this figure that the temperature gradient at the wall rises with an increase
in the Prandtl number. This implies that when the flow attains a higher value of the Prandtl number, the capacity of the
fluid to transfer the energy is drastically reduced. Fig. 13 explores how the temperature distribution changes with rise in the
value of the heat source parameter α . The results presented here show that thermal radiation plays an important role in the
formation of the temperature profiles. From this figure, one finds that with a rise in the value of the heat source parameter
in the fluid mass that is in motion, there is a rise in the thickness of the boundary layer. This indicates that there occurs a
rise in the energy transfer capacity of the fluid, when the heat source parameter increases.
J.C. Misra et al. / Physica A 470 (2017) 330–344 343

In order to make the study more exhaustive, it is important to examine the variation of skin-friction and local Nusselt
number defined by Eqs. (18)–(20) for different values of the parameters. Variation in the values of skin-friction and local
Nusselt number can be observed from Table 2, for fixed values of the reciprocal magnetic Prandtl number λ and the ratio
a/c. This enables us to have a complete picture of the problem under consideration. This table shows that the magnitude
of the skin friction increases, while that of the local Nusselt number reduces, when the strength of the magnetic field
increases. However, with an increase in permeability, the magnitude of the skin friction reduces, but that of the Nusselt
number increases. One can similarly have an idea of the effects of changing the values of buoyancy parameter (λ1 ), suction
parameter (s) and heat source parameter (α ) on the skin friction and the local Nusselt number by having a glance through
Table 2. Moreover, Fig. 14 gives the variation of local skin-friction coefficient for different values of the parameters β and λ.
It is observed from Fig. 14 that the skin-friction coefficient has increasing effect on λ while the magnetic parameter β has
reducing effect. Fig. 15 represents the variation of local Nusselt number with α for different values of Pr. It is seen that local
Nusselt number decreases as α increases, but it increases with increasing Pr.
The validity of the present study has been examined by comparing our results with those of a simplified study performed
earlier by Ali et al. [34]. The results obtained from our study along with those reported in Ref. [34] have been presented
in Fig. 16. For the sake of a proper comparison, the results presented in Fig. 16 for our study were computed by taking
β = 5.0, s = 0.0, Pr = 0.72, a/c = 3.0, λ = 1.0, λ1 = 0.0 and α = 5.0, considered in Ref. [34]. Fig. 16 shows that our
results conform very well to those of Ref. [34].

7. Summary and conclusion

The present investigation has been carried out for the mixed convection stagnation point flow and heat transfer of a
magnetohydrodynamic incompressible fluid on a thin porous stretching sheet in the presence of a heat source. However, the
study has been motivated with the purpose of exploring some relevant phenomena that occur during the physiological flow
of blood in the microcirculatory system. The experimental studies on microcirculatory system reveal that the microvessels,
e.g. capillaries, behave like a stretching sheet. An attempt has been made to formulate the problem accordingly. Numerical
estimates of different physical variables involved in the flow and heat transfer problem have been obtained by employing
a suitable numerical procedure that provides better efficiency, more accuracy and better stability. The results have been
presented and adequately discussed in Section 6. The observations of the study lead to the following conclusions:
• The induced magnetic field reduces the fluid velocity, but it gives rise to an enhancement of the fluid temperature.
• With a rise in the suction/injection velocity, the fluid velocity reduces up to a certain point, beyond which an opposite
trend is observed.
• The temperature of the moving mass of the fluid in the boundary layer reduces as the Prandtl number increases.
The study will find important applications in validating the results of relevant experimental studies, as well as the results
of studies for more complex flows that may be carried out in the future by applying numerical methods. The study will also be
helpful in making a proper design of fluid flow in micro-fluidic devices. The results presented here will also find applications
in drug delivery and species separation. It is worthwhile to make a mention here the relevant experimental work of Feusner
et al. [41] on platelet counts in capillary blood, in which samples of both capillary and venous blood were used in an attempt
to assess platelet counts in the capillary bed. On the basis of their experimental observation, they arrived at the conclusion
that venous blood gives a more accurate estimate of platelet counts than that in capillary blood. Stagnation point flow in a
capillary plug (uniform thin walled capillaries) was studied by Gadgil in his experimental researches on anisotropic porous
chemical vapor deposition (APCVD) reactor design [42] and also optimization of stagnation flow reactor design [43]. Thus
the present study reported here bears the promise of far reaching applications also in exploring important issues involved
in chemical/biochemical industries.

Acknowledgments

The authors wish to express their deep sense of gratitude to Science and Engineering Research Board, Department of
Science and Technology, Government of India, New Delhi for the financial support of this investigation through Grant No.
SB/S4/MS: 864/14. The authors are thankful to the esteemed reviewers for their valuable comments based upon which the
manuscript has been revised.

References

[1] L.J. Clancy, Aerodynamics, Pitman, 1975.


[2] R.W. Fox, A.T. McDonald, Introduction to Fluid Mechanics, fourth ed., Wiley, 2003.
[3] G.J. Van Wylen, R.E. Sonntag, Fundamentals of Classical Thermodynamics, John Wiley and Sons, Inc., New York, 1965.
[4] D.W. Beard, K. Walters, Elastico-viscous boundary-layer flows, J. two-dimensional flow near a stagnation point, Proc. Cambridge Philos. Soc. 60 (1964)
667–676.
[5] P.D. Ariel, A numerical algorithm for computing the stagnation point flow of a second grade fluid with/without suction, J. Comput. Appl. Math. 59 (1)
(1995) 9–24.
[6] C.Y. Wang, Stagnation flows with slip: Exact solution of the Navier–Stokes equations, Z. Angew. Math. Phys. 54 (2003) 184–189.
[7] C.Y. Wang, Stagnation slip flow and heat transfer on a moving plate, Chem. Eng. Sci. 61 (23) (2006) 7668–7672.
344 J.C. Misra et al. / Physica A 470 (2017) 330–344

[8] K.V. Prasad, K. Vajravelu, P.S. Datti, The effects of variable fluid properties on the hydro-magnetic flow and heat transfer over a non-linearly stretching
sheet, Int. J. Therm. Sci. 49 (2010) 603–610.
[9] T. Hayat, Z. Abbas, I. Pop, S. Asghar, Effects of radiation and magnetic field on the mixed convection stagnation-point flow over a vertical stretching
sheet in a porous medium, Int. J. Heat Mass Transfer 53 (2010) 466–474.
[10] F. Labropulu, D. Li, Stagnation point flow of a second grade fluid with slip, Int. J. Nonlinear Mech. 43 (9) (2008) 941–947.
[11] A. Malvandi, F. Hedayati, D.D. Ganji, Slip effects on unsteady stagnation point flow of a nanofluid over a stretching sheet, Powder Technol. 253 (2014)
377–384.
[12] A. Sinha, J.C. Misra, Effect of induced magnetic field on magnetohydrodynamic stagnation point flow and heat transfer on a stretching sheet, ASME. J.
Heat Transfer 136 (11) (2014).
[13] J.D. Humphrey, O. Rourke, L. Sherry, An Introduction to Biomechanics: Solids and Fluids, in: Analysis and Design, Springer, New York, 2015.
[14] D. Ambrosi, A. Quarteroni, G. Rozza (Eds.), Modeling of Physiological Flows, Springer, 2012.
[15] A. Roldán-Alzate, Simulation of physiological flows (Ph. D. Dissertation), Univ. Wisconsin-Madison, 2008.
[16] K. Affeld, A.J. Reininger, J. Gadischke, K. Grunert, S. Schmidt, F. Thiele, Fluid mechanics of the stagnation point flow chamber and its platelet
decomposition, Artif. Organs 19 (1995) 597–602.
[17] T. David, S. Thomas, P.G. Walker, Platelet deposition in stagnation point flow: an analytical and computational simulation, Med. Eng. Phys. 23 (5)
(2001) 299–312.
[18] J.R. Levick, An Introduction to Cardiovascular Physiology, Butterworth-Heinemann, London, 2013.
[19] P. Vadasz, Fluid Flow and Heat Transfer in Rotating Porous Media, Springer, 2016.
[20] J.C. Misra, G.C. Shit, H.J. Rath, Flow and heat transfer of a MHD viscoelastic fluid in a channel with stretching walls: Some Applications to
Haemodynamics, Comput. & Fluids 37 (1) (2008) 1–11.
[21] J.C. Misra, G.C. Shit, Flow of a biomagnetic visco-elastic fluid in a channel with stretching walls, J. Appl. Mech. (ASME) 76 (6) (2009) 061006.
[22] J.C. Misra, B. Pal, A.S. Gupta, Hydromagnetic flow of a second-grade fluid in a channel: Some applications to physiological systems, Math. Models
Methods Appl. Sci. 8 (08) (1998) 1323–1342.
[23] J.C. Misra, A. Sinha, Effect of thermal radiation on MHD flow of blood and heat transfer in a permeable capillary in stretching motion, Heat Mass Transf.
49 (5) (2013) 617–628.
[24] J.C. Misra, S.D. Adhikary, MHD oscillatory channel flow, heat and mass transfer in a physiological fluid in presence of chemical reaction, Alexandria
Eng. J. 55 (1) (2016) 287–297.
[25] G.C. Shit, A. Mondal, A. Sinha, P.K. Kundu, Electro-osmotically driven MHD flow and heat transfer in micro-channel, Physica A 449 (2016) 437–454.
[26] G.C. Shit, A. Mondal, A. Sinha, P.K. Kundu, Electro-osmotic flow of power-law fluid and heat transfer in a micro-channel with effects of Joule heating
and thermal radiation, Physica A 462 (2016) 1040–1057.
[27] D. Bluestein, L. Niu, R.T. Schoephoerster, M.K. Dewanjee, Steady flow in an aneurysm model: correlation between fluid dynamics and blood platelet,
J. Biomech. Eng. 118 (3) (1996) 280–286.
[28] H. Petschek, D. Adamis, A.R. Kantrowitz, Stagnation flow thrombus formation, Trans. Am. Soc. Artif. Intern. Organs 14 (1968) 256–260.
[29] L.J. Wurzinger, P. Blasberg, M. Van de Loecht, W. Suweklack, H. Schmid-Schonbein, Model experiments on platelet adhesion in stagnation point flow,
Biorheology 21 (4) (1984) 649–659.
[30] L.J. Wurzinger, P. Blasberg, F. Horii, H. Schmid-Schobein, A stagnation point flow technique to measure platelet adhesion onto polymer films from
native blood - A technical report, Thromb. Res. 44 (3) (1986) 401–406.
[31] K. Affeld, A.J. Reininger, J. Gadischke, K. Grunet, S. Schmidt, F. Thiele, Fluid mechanics of the stagnation point flow chamber and its platelet deposition,
Artif. Organs 19 (7) (1995) 597–602.
[32] A.L. Hazel, T.J. Pedley, Alteration of mean wall shear stress near an oscillating stagnation stagnation point, J. Biomech. Eng. 120 (2) (1998) 227–237.
[33] N.G. Kafoussias, E.M. Williams, Improved approximation technique to obtain numerical solution of a class of two-point boundary value similarity
problems in fluid mechanics, Internat. J. Numer. Methods Fluids 17 (2) (1993) 145–162.
[34] F.M. Ali, R. Nazar, N.M. Arifin, I. Pop, MHD stagnation-point flow and heat transfer towards stretching sheet with induced magnetic field, Appl. Math.
Mech. 32 (4) (2011) 409–418.
[35] H. Rosali, Anuar Ishak, Ioan Pop, Stagnation point flow and heat transfer over a stretching/shrinking sheet in a porous medium, Int. Commun. Heat
Mass Transfer 38 (2011) 1029–1032.
[36] K. Bhattacharyya, G.C. Layek, Effects of suction/blowing on steady boundary layer stagnation-point flow and heat transfer towards a shrinking sheet
with termal radiation, Int. J. Heat Mass Transfer 54 (2011) 302–307.
[37] A. Ishak, R. Nazar, I. Pop, Mixed convection on the stagnation point flow toward a vertical, continuously stretching sheet, ASME J. Heat Transfer 129
(8) (2006) 1087–1090.
[38] V. D’Ambrosio, F. Dughiero, M. Forzan, Numerical models of RF-thermal ablation treatments, Int. J. Appl. Electromagn. Mech. 25 (2007) 429–433.
[39] J.C. Misra, A. Sinha, Effect of thermal radiation on MHD flow of blood and heat transfer in a permeable capillary in stretching motion, Heat Mass
Transfer 49 (2013) 617–628.
[40] N. Freidoonimehr, M.M. Rashidi, B. Jalilpour, MHD stagnation-point flow past a stretching/shrinking sheet in the presence of heat
generation/absorption and chemical reaction effects, J. Braz. Soc. Mech. Sci. Eng. (2015) 1–10. http://dx.doi.org/10.1007/s40430-015-0456-8.
[41] J.H. Feusner, J.A. Behrens, J.C. Detter, T.C. Cullen, Platelet counts in capillary blood, Am. J. Clin. Pathol. 72 (3) (1979) 410–414.
[42] P.N. Gadgil, Capillary plug flow distributor for stagnation point flow in APCVD reactor, Mater. Lett. 20 (5) (1994) 351–354.
[43] P.N. Gadgil, Optimization of a stagnation point flow reactor design for metalorganic chemical vapor deposition by flow visualization, J. Cryst. Growth
134 (3) (1993) 302–312.

You might also like