You are on page 1of 13

EPJ Plus

EPJ .org
your physics journal

Eur. Phys. J. Plus (2017) 132: 297 DOI 10.1140/epjp/i2017-11572-y

Effects of lubricated surface in the oblique stagnation


point flow of a micro-polar fluid

K. Mahmood, M. Sajid, N. Ali and M.N. Sadiq


Eur. Phys. J. Plus (2017) 132: 297
DOI 10.1140/epjp/i2017-11572-y
THE EUROPEAN
PHYSICAL JOURNAL PLUS
Regular Article

Effects of lubricated surface in the oblique stagnation point flow


of a micro-polar fluid
K. Mahmooda , M. Sajid, N. Ali, and M.N. Sadiq
Department of Mathematics and Statistics, International Islamic University, Islamabad 44000, Pakistan

Received: 3 February 2017 / Revised: 23 May 2017


c Società Italiana di Fisica / Springer-Verlag 2017
Published online: 6 July 2017 – 

Abstract. In this investigation, we have considered steady, two-dimensional, oblique flow of a micro-polar
fluid towards a stagnation point over a lubricated plate. A power-law fluid is utilized for the purpose
of lubrication. To derive the slip condition in the present flow situation, continuity of shear stress and
velocity has been imposed at the fluid lubricant interface. The set of non-linear coupled ordinary differential
equations subject to boundary conditions is solved by a powerful numerical technique called the Keller-
box method. Some important flow features have been analyzed and discussed under the influence of slip
parameter λ, the material parameter K, a free parameter β and ratio of micro-rotation to the skin friction
parameter n. The main purpose of the present article is to analyze the reduction in the shear stress and shift
of stagnation point in the presence of lubrication as compared to the viscous fluid that may be beneficial
during polymeric processing.

1 Introduction
Non-orthogonal stagnation point flow takes place when a stream of fluid strikes the surface at a certain angle. Such a
flow is a combination of orthogonal stagnation point flow and shear flow along the surface. Oblique stagnation point
flow was investigated independently by Stuart [1], Tamada [2] and Dorrepaal [3]. Time-dependent flow towards a
stagnation point was presented by Takemitsu and Matunobu [4] and Wang [5], while Tilley and Weidman [6] discussed
non-orthogonal stagnation point of two fluids. Later, Drazin and Riley [7] and Tooke and Blyth [8] introduced a free
parameter while reviewing the oblique flow problem. Recently, Ghaffari et al. [9, 10] discussed different aspects for the
flows towards an oblique stagnation point.
Wang [11] discussed the effects of the slip parameter on the stagnation point flow of a viscous fluid. Labropulu
et al. [12] examined slip flow due to second-grade fluid impinging orthogonally or obliquely on a surface. Blyth
and Pozrikidis [13] studied stagnation point flow by introducing slip condition at the interface of two viscous fluids.
Axisymmetric stagnation point flow near a lubricated stationary disc has been analyzed by Santra et al. [14]. They used
a power-law fluid as a lubricant. Sajid et al. [15] reconsidered the problem of Santra et al. [14] by applying generalized
slip condition at the fluid-lubricant interface introduced by Thompson and Troian [16]. Recently, Mahmood et al. in
refs. [17] and [18], respectively, investigated flows of second-grade fluids near an oblique stagnation point due to a
lubricated plate and over a lubricated rotating disc.
In recent decades many researchers have been investigating flows of non-Newtonian fluids due to their significant
role in applied sciences, engineering and industrial applications. The governing equations representing these fluids are
highly non-linear and are complicated to solve even by a numerical approach. Among these non-Newtonian fluids,
an important fluid is the micro-polar one. A micro-polar fluid consists of small-sized particles suspended in a viscous
domain with intrinsic rotational micro-motion. Examples of micro-polar fluids are biological fluids like blood, polymeric
additives (suspensions), geomorphological sediments, colloidal, slurries and hematological suspensions, etc. It has been
proved experimentally by Hoyt and Fabula [19] that the fluids containing polymeric additives show a massive reduction
of polymeric concentration and shear stress as explained by Eringen [20]. Eringen [21, 22] also proposed the theory of
micro-polar fluids with which the deformation inside such fluids can be very well explained. Ahmadi [23] investigated
the boundary layer flow of a micro-polar fluid past a semi-infinite surface. Some extensive applications of microploar
fluids has been provided by Ariman et al. [24, 25]. Guram and Smith [26] analyzed strong and weak interaction due to
stagnation point flow of a micro-polar fluid.
a
e-mail: khalidmeh2012@gmail.com (corresponding author)
Page 2 of 12 Eur. Phys. J. Plus (2017) 132: 297

Fig. 1. Schematic diagram for the considered flow problem.

The turbulent flow of a micro-polar fluid has been discussed by Peddieson [27]. Nazar et al. [28] presented the
flow of a micro-polar fluid in the vicinity of a stagnation point over a stretching surface. Heat transfer in the mixed
convection flow of a micro-polar fluid towards a vertical wall with conduction has been investigated by Chang [29].
Lok et al. [30] carried out an analysis on the oblique stagnation point for a micro-polar fluid. A problem investigating
the effects of micro-rotation towards the stagnation point over a vertical surface is discussed by Ishak et al. [31]. They
found the numerical solution by implementing an implicit finite-difference scheme.
Our goal in this communication, is to investigate steady, two-dimensional oblique stagnation point flow of a micro-
polar fluid on a lubricated surface. A power-law fluid is used for the lubrication purpose. The flow problem consists
of a set of coupled non-linear ordinary differential equations along with non-linear coupled boundary conditions.
The Keller-box method [32–34] has been implemented to solve the considered flow problem numerically. Influence of
pertinent parameters on the flow characteristics is discussed through graphs and tables. The validity of the present
study has been checked by comparing results in the limiting case with those exist in the literature.

2 Mathematical formulation
Consider a steady, two-dimensional oblique stagnation point flow of a micro-polar fluid past a semi-infinite lubricated
plate. A power-law fluid is used as lubricant. The lubricated plate is placed along the x-axis, as shown in fig. 1. We
consider a free stream flow containing a combination of orthogonal stagnation point flow with strain rate a and a shear
flow parallel to the plate having strain b.
If U and V are, respectively, the horizontal and vertical velocity components of the power-law fluid, then we have
 H(x)
Q= U (x, y)dy, (1)
0

where Q is the flow rate of the lubricant and H(x) is the variable thickness of the lubrication layer.
Assuming u, v the velocity components in the x- and y-direction and N as the micro-rotation in the z-direction,
we have [30]

∂u ∂v
+ = 0, (2)
∂x ∂y
 
∂u ∂u 1 ∂p µ+k k ∂N
u +v =− + ∇2 u + , (3)
∂x ∂y ρ ∂x ρ ρ ∂y
 
∂v ∂v 1 ∂p µ+k k ∂N
u +v =− + ∇2 v − , (4)
∂x ∂y ρ ∂y ρ ρ ∂x
   
∂N ∂N ∂u ∂v
ρj u +v = γ∇2 N − k 2N + − , (5)
∂x ∂y ∂y ∂x
Eur. Phys. J. Plus (2017) 132: 297 Page 3 of 12

where p is the pressure, ρ is the density, k is the vortex viscosity, γ is the spin gradient viscosity and j is the micro-inertia
density. Eliminating the pressure between eqs. (3) and (4) we obtain

∂2u ∂2u ∂2v ∂2v


 3
∂3u ∂3v ∂3v k ∂2N ∂2N
   
µ+k ∂ u
u +v 2 −u 2 −v − + 3 − − + − = 0. (6)
∂y∂x ∂y ∂x ∂x∂y ρ ∂y∂x2 ∂y ∂x3 ∂x∂y 2 ρ ∂y 2 ∂x2
The wall shear stress is given as   
∂u 
τw = (µ + k) + kN  . (7)
∂y y=0

The boundary conditions at the surface, interface and free stream are as follows.
At the surface, the no-slip boundary conditions imply

U (x, 0) = 0, V (x, 0) = 0. (8)

Assuming that the lubricant film is very thin, we have

V (x, y) = 0, ∀ y ∈ [0, H(x)]. (9)

The interfacial condition between the micro-polar fluid and the lubricant can be obtained by imposing continuity of
velocity and shear stress of both fluids. If µL is the apparent viscosity of the power-law fluid, then continuity of the
shear stress at y = H(x) gives  
∂u ∂U
(µ + k) + kN = µL , (10)
∂y ∂y
where µL is defined as
 m−1
∂U
µL = k1 , (11)
∂y
in which m is power-law index and k1 is the consistency coefficient. Let us assume U (x, y) in the following form:

Ũ (x)y
U (x, y) = . (12)
H(x)

Here Ũ (x) denotes the velocity of both fluids at the interface. Using eq. (1), the thickness H(x) of the lubricant can
be expressed as
2Q
H(x) = . (13)
Ũ (x)
Substituting eqs. (11)–(13), eq. (10) leads to the following slip boundary condition:
 m
∂u 1
(µ + k) + kN = k1 u2m . (14)
∂y 2Q

It is to be mentioned here that Ũ = u according to the continuity of horizontal velocity at y = H(x). Moreover, N at
the surface is  
∂u ∂v
N = −n − , (15)
∂y ∂x
in which the constant n ∈ [0, 1]. The case n = 0 gives rise to strong concentration, implying N = 0 at the surface
indicating that there is no micro-rotation closer to the wall, due to strong concentration [26]. The case n = 1 is utilized
to describe turbulent boundary layer flow [27], and the case n = 0.5 describes weak concentration indicating vanishing
of anti-symmetrical part of stress tensor [23]. The continuity of vertical velocity components of both fluids at y = H(x)
implies
v(x, H(x)) = V (x, H(x)). (16)
Equations(9) and (16) together give
v(x, H(x)) = 0. (17)
Assuming the lubrication layer to be very thin, we can apply boundary conditions (14) and (17) at y = 0. If a and
b represent dimensional constants, then following Tooke and Blythe [8], the boundary conditions for the velocity
components at free stream are
ue = ax + b(y − b), ve = −a(y − a). (18)
Page 4 of 12 Eur. Phys. J. Plus (2017) 132: 297

Furthermore, at free stream, N has the following form:


b
N =− , (19)
a
where b/a represents the shear in the free stream. To express the set of equations into dimensionless form, we introduce


′ ′ a
u = axf (η) + ag (η), v = − aνf (η), N = axq(η) + aT (η), η=y . (20)
ν
In new variables, governing equations subject to boundary conditions become
iv
(1 + K)f + f f ′′′ + f ′ f ′′ − 2f ′ f ′′ + Kq ′′ = 0, (21)
iv ′ ′′ ′′′ ′ ′′ ′ ′′ ′′
(1 + K)g + f g + f g − f g − g f + KT = 0, (22)
′′ ′ ′ ′′
(1 + K/2)q + f q − f q − K(2q + f ) = 0, (23)
′′ ′ ′ ′′
(1 + K/2)T + f T − g q − K(2T + g ) = 0, (24)
′′ ′ 2m ′
f (0) = 0, {1 + (1 − n)K}f (0) = λ(f (0)) , f (∞) = 1, (25)
′′ ′ ′ 2m−1 ′′
g(0) = 0, {1 + (1 − n)K}g (0) = 2mλg (0)(f (0)) , g (∞) = ϕ, (26)
′′
q(0) = −nf (0), q(∞) = 0, (27)
′′ ϕ
T (0) = −ng (0), T (∞) = − , (28)
2
where K = k/µ is the vortex viscosity parameter and ϕ = b/a. The parameter λ, in eqs. (25) and (26), is given by

k1 ν a2m x2m−1
λ= . (29)
µ a3/2 (2Q)m
Integrating eqs. (21) and (22), and using the free stream conditions, we get
2
(1 + K)f ′′′ − f ′ + f f ′′ + 1 + Kq ′ = 0, (30)
′′′ ′′ ′ ′ ′
(1 + K)g + f g − f g + KT + ϕ(α − β) = 0, (31)
where β is a free parameter and α = η ∞ − f (∞).
Introducing g ′ (η) = ϕh(η) and T (η) = ϕS(η), eqs. (24) and (31) reduce to
(1 + K)h′′ + f h′ − f ′ h + KS ′ + α − β = 0, (32)
 
K
1+ S ′′ + f S ′ − qh − K(2S + h′ ) = 0. (33)
2
The corresponding boundary conditions become
′ 2m−1
{1 + (1 − n)K}h (0) = 2mλh(0)(f ′ (0)) , h′ (∞) = 1. (34)
1
S(0) = −nh′ (0), S(∞) = − . (35)
2
From eq. (29) we note that eqs. (23), (30), (32) and (33) possess a similarity solution when m = 1/2. Furthermore, λ
is the ratio between the viscous and lubrication length scales, thus

Lvisc
λ = µ √a = . (36)
k1 2Q
Llub
The case when Llub is small, λ becomes sufficient large and when λ → ∞, the no-slip conditions f ′ (0) = 0, and
h(0) = 0 are restored from eqs. (25) and (34). The case when Llub attains a huge value then λ → 0 and the full
slip boundary conditions f ′′ (0) = 0 and h′ (0) = 0 are obtained. Thus, the parameter λ can be utilized to control the
slip produced by the lubricant and is called slip parameter. Employing (20), the dimensionless skin friction can be
written as
τw = {(1 − n)K + 1}{xf ′′ (0) + g ′′ (0)} = {(1 − n)K + 1}{xf ′′ (0) + ϕh′ (0)}. (37)
At the stagnation point, τw = 0, therefore the location of stagnation point (xs ) on the surface is given by
g ′′ (0) h′ (0)
xs = − = −ϕ ′′ . (38)
f (0)
′′ f (0)
Eur. Phys. J. Plus (2017) 132: 297 Page 5 of 12

Fig. 2. Influence of K on f ′ for both no-slip and partial slip cases when (a) n = 0 (b) n = 0.5.

Fig. 3. Influence of λ on f ′ for two different values of K when (a) n = 0 (b) n = 0.5.

Fig. 4. Influence of K on h′ for both no-slip and partial slip cases when (a) n = 0 (b) n = 0.5.

3 Numerical results and discussions


The values of f ′ , f ′′ , h′ , −q, −q ′ , −S and −S ′ are evaluated for various values of n, λ, K and β by solving
eqs. (23), (30), (32) and (33) subject to boundary conditions (25), (27), (34) and (35) numerically by implement-
ing Keller-box method [32–34].
Figures 2 and 3 have been plotted to show the effects of vortex viscosity parameter K and slip parameter λ on f ′ ,
respectively, while figs. 4–7 display the effects of parameters K, λ and β on h′ . The effects of emerging parameters
on −q are shown in figs. 8 and 9, while that of on −S is presented in figs. 10–13. Streamlines patterns for various
combinations of appearing parameters are shown in figs. 14–16. Influence of pertinent parameters on f ′′ (0) and α for
various values of parameters are shown in fig. 17, while fig. 18 shows the variation in h′ (0) against involved parameters.
Tables 1 and 2 show the comparison of α, f ′′ (0) and h′ (0) in the limiting case of [30].
Page 6 of 12 Eur. Phys. J. Plus (2017) 132: 297

Fig. 5. Influence of λ on h′ for two different values of K when (a) n = 0 (b) n = 0.5.

Fig. 6. Influence of β on h′ for two different values of K and λ = 1 when (a) n = 0 (b) n = 0.5.

Fig. 7. Influence of β on h′ for two different values of λ and K = 1 when (a) n = 0 (b) n = 0.5.

Figure 2 displays the effects of the material parameter K on the horizontal velocity component f ′ . This figure
shows that f ′ decreases with an increase in the magnitude of K. This decrease is more prominent on rough surface
(λ → ∞) as compared to the lubricated surface (e.g., λ → 1) indicating that lubrication enhances the velocity of the
core fluid. The reason is that K is the ratio between vortex and absolute viscosities. If K is large, we can say that
viscous effects are dominant near the wall. That is why f ′ reduces near the surface. Moreover, this decrease is more
significant when there is a strong concentration of micro-particles in the fluid, i.e. when n = 0 (fig. 2(a)) as compared
to the weak concentration of micro-particles, i.e. when n = 0.5 (fig. 2(b)). Variation in f ′ against slip parameter λ is
described in fig. 3. It is evident that f ′ decelerates by enhancing λ. It is also clear, from this figure, that K enhances
the effects of λ. Variation in f ′ is more effective for n = 0 (fig. 3(a)) as compared to when n = 0.5 (fig. 3(b)).
Eur. Phys. J. Plus (2017) 132: 297 Page 7 of 12

Fig. 8. Influence of K on −q for both no-slip and partial slip cases when (a) n = 0 (b) n = 0.5.

Fig. 9. Influence of λ on −q for two different values of K when (a) n = 0 (b) n = 0.5.

Fig. 10. Influence of K on −S for both no-slip and partial slip cases when (a) n = 0 (b) n = 0.5.

Figure 4 was plotted to show the impact of the parameter K on the shear component h′ when β = 0. According to
this figure, h′ decreases by increasing K both for n = 0 and n = 0.5. We observe some oscillations in the magnitude
of h′ over the rough surface (dashed lines) which are more prominent when n = 0 (fig. 4(a)). However, no oscillations
are seen on the lubricated surface. Therefore, it is concluded that introduction of lubrication is very effective to obtain
smoothness in the curves. Influence of λ on h′ in the absence of β is depicted in fig. 5. The analysis shows that h′
enhances by augmenting λ in the absence of micro-rotation (dashed lines) as well as in the presence of micro-rotation
(solid lines). A close inspection of this figure shows that the increase in h′ is more significant on the lubricated surface.
Page 8 of 12 Eur. Phys. J. Plus (2017) 132: 297

Fig. 11. Influence of λ on −S for two different values of K when (a) n = 0 (b) n = 0.5.

Fig. 12. Influence of β on −S for two different values of K and λ = 1 with (a) n = 0 (b) n = 0.5.

Fig. 13. Influence of β on −S for two different values of λ and K = 1 when (a) n = 0 (b) n = 0.5.

Also the parameters n, K and λ increase the effects of each other (fig. 5(b)). Figure 6 is devoted to show the effects
of parameter β on h′ for two different values of K when n = 0 and n = 0.5. It is obvious, through this figure, that
h′ reduces with an increase in the value of β. A reasonable variation is observed in the absence of micro-rotation
(K = 0). This variation is also significant for weak concentration of micro-particles (n = 0.5) even in the presence
of micro-rotation (K = 1) as shown in fig. 6(b). Variation of h′ against the parameter β is displayed in fig. 7. This
figure shows that h′ is a decreasing function of β. This decrease is controlled by introducing a lubricant on the surface.
Eur. Phys. J. Plus (2017) 132: 297 Page 9 of 12

Fig. 14. Streamlines pattern against λ = 0.5 and λ = 10 when ϕ = 4, β = 0, m = 0.5, K = 3 and (a) n = 0 (b) n = 0.5. (a)
From left to right: x1 = −4.88476, x2 = −4.23499. (b) From left to right: x1 = −4.88142, x2 = −4.28137.

Fig. 15. Streamlines pattern against K = 0.5 and K = 5 when ϕ = 4, β = 0, m = 0.5, λ = 1 and (a) n = 0 (b) n = 0.5. (a)
From left to right: x1 = −3.31069, x2 = −5.67413. (b) From left to right: x1 = −3.31209, x2 = −5.67271.

Moreover, it is controlled by introducing a micro-polar fluid with strong concentration (fig. 7(a)). Figure 8 indicates
how the micro-rotation profile −q behaves against K. According to fig. 8(a), −q shows an increasing trend against K,
which is more significant on the rough surface (λ = ∞). The same behaviour is observed in fig. 8(b) except for some
oscillations at the beginning. Figure 9 is presented for the impact of λ on −q. It is clear from this figure that −q is an
increasing function of λ for all values of K. From the above discussion, it is clear that the velocity and micro-rotation
boundary layer thickness augment as K and λ gains the values. Figures 10–13 are displayed to show the effects of
parameters K, λ and β, respectively, on the micro-rotation component −S. These figures show that −S is a decreasing
function of parameters K, λ and β individually for n = 0. The relation S(0) = −nh′ (0) given in eq. (35) indicates
that both h′ and −S have the same pattern near the surface n > 0.
Streamline patterns are presented in figs. 14–16 to see the influence of governing parameters λ, K and β, respectively,
on the stagnation point xs . It is evident from fig. 14 that xs moves towards right when λ is increased from 0.5 to 10
both for n = 0 and n = 0.5. Figure 15 demonstrates that xs moves towards left when K gains the values from 0.5 to
5. According to fig. 16, xs is translated in the right direction when β is increased from −3 to 3. Moreover, a minor
variation in the stagnation point is observed when n varies from 0 to 0.5. It means that concentration of micro-particles
in the fluid has no effect on the movement of the stagnation point.
Page 10 of 12 Eur. Phys. J. Plus (2017) 132: 297

Fig. 16. Streamlines pattern against β = −3 and β = 3 when ϕ = 3, K = 2, m = 0.5, λ = 1 and (a) n = 0 (b) n = 0.5. (a)
From left to right: x1 = −6.24465, x2 = −0.24465. (b) From left to right: x1 = −6.24159, x2 = −0.24159.

Fig. 17. Variation in f ′′ (0) and α against λ and K when n = 0 and n = 0.5. (a) f ′′ (0) against λ and K. (b) α against λ and K.

Fig. 18. Variation in h′ (0) against λ, K and β when n = 0 and n = 0.5. (a) h′ (0) against λ and K. (b) h′ (0) against λ and β.

Figures 17(a) and (b) provide variation in f ′′ (0) and α against λ for different values of K and n. It is observed,
through these figures, that f ′′ (0) and α increase by increasing λ for both n = 0 and n = 0.5. However, f ′′ (0) is a
decreasing and α an increasing function of K. The variation in the magnitude of h′ (0) due to the governing parameters
is displayed in figs. 18(a) and (b). We see that h′ (0) increases by increasing λ for β ≤ 0 and decreases for β > 0.
Eur. Phys. J. Plus (2017) 132: 297 Page 11 of 12

Table 1. Comparison of numerical values of α and f ′′ (0) with [30] for different values of K and n when λ = ∞ and β = 0. The
second entry in each row represents f ′′ (0).

n=0 n = 0.5
Present results Results in [30] Present results Results in [30]
K = 0.0 0.647903 0.647912 0.647903 0.647912
1.232598 1.232627 1.232598 1.232627
K = 0.5 0.786242 0.786249 0.724377 0.724384
0.992640 0.992657 1.102467 1.102486
K = 1.0 0.893914 0.893920 0.793515 0.793521
0.841074 0.841085 1.006409 1.006423
K = 1.5 0.984147 0.984152 0.857094 0.857100
0.736826 0.736835 0.931753 0.931764
K = 2.0 1.063082 1.063087 0.916275 0.916277
0.660473 0.660479 0.871574 0.871584
K = 2.5 1.134033 1.134038 0.971852 0.971858
0.601901 0.601906 0.821728 0.821736
K = 3.0 1.198989 1.198994 1.024422 1.024427
0.555369 0.555374 0.779559 0.779566

Table 2. Comparison of numerical values of h′ (0) with [30] for different values of K and n when λ = ∞ and β = 0.

n=0 n = 0.5
Present results Results in [30] Present results Results in [30]
K = 0.0 1.406556 1.406592 1.406556 1.406592
K = 0.5 1.305958 1.305981 1.406553 1.406579
K = 1.0 1.230820 1.230837 1.406551 1.406572
K = 1.5 1.173988 1.174002 1.406550 1.406569
K = 2.0 1.129847 1.129859 1.406549 1.406566
K = 2.5 1.094635 1.094646 1.406549 1.406563
K = 3.0 1.065878 1.065887 1.406548 1.406561

A rapid change in h′ (0) is seen for n = 0.5. Moreover, h′ (0) reduces by increasing K as well as β for both n = 0 and
n = 0.5. A rapid change in f ′′ (0), h′ (0) and α is observed for the case of strong concentration (n = 0). The relations
q(0) = −nf ′′ (0) and S(0) = −nh′ (0) given in eqs. (27) and (35), respectively, indicate that f ′′ (0) and q(0), and h′ (0)
and −S(0) have the same pattern near the surface for n > 0.
The numerical values of α, f ′′ (0) and h′ (0) for the no-slip case show a good agreement with those of [30] proving
the correctness of the provided numerical solutions (tables 1 and 2).

4 Concluding remarks
In this paper oblique stagnation point flow of a micro-polar fluid over a plate is investigated. A power-law fluid is
introduced on the surface for the lubrication purpose. The flow problem is numerically solved using the Keller-box
method for m = 1/2 to obtain similar solution. The impact of slip parameter λ, material parameter K and free
parameter β on the flow characteristics has been investigated. Some findings are listed below.
i) The velocity of micro-polar fluid f ′ reduces with increasing K and λ. This reduction is more prominent when
K > 0 and λ = ∞ for both n = 0 and n = 0.5.
ii) h′ is an increasing function of λ but a decreasing function of K and β.
iii) −q increases by increasing K and λ. This increase is more significant for K > 0 and λ = ∞.
Page 12 of 12 Eur. Phys. J. Plus (2017) 132: 297

iv) −S decreases by enhancing K and β both for n = 0 and n = 0.5. However, it reduces by augmenting λ for n = 0
and increases for n = 0.5.
v) The velocity and micro-rotation boundary layer thickness augment as K and λ increase.
vi) Stagnation point moves towards right by increasing λ as well as β and shows an opposite behaviour by increasing
K.
vii) f ′′ (0) and α increase by increasing λ for both n = 0 and n = 0.5. However, f ′′ (0) is a decreasing and α is an
increasing function of K.
viii) h′ (0) increases by increasing λ for β ≤ 0 and decreases for β > 0. A rapid change in h′ (0) is seen for n = 0.5.
Moreover h′ (0) reduces by increasing K as well as β for both n = 0 and n = 0.5.

References
1. J.T. Stuart, J. Aerospace Sci. 26, 124 (1959).
2. K.J. Tamada, J. Phys. Soc. Jpn. 46, 310 (1979).
3. J.M. Dorrepaal, J. Fluid Mech. 163, 141 (1986).
4. N. Takemitsu, Y. Matunoba, J. Phys. Soc. Jpn. 47, 1347 (1979).
5. C.Y. Wang, Z. Angew. Math. Phys. 54, 184 (2003).
6. B.S. Tilley, P.D. Weidman, Eur. J. Mech. B/Fluids 17, 205 (1998).
7. P. Drazin, N. Riley, The Navier-Stokes equations: a classification of flows and exact solutions, in London Mathematical
Society Lecture Note Series, Vol. 334 (Cambridge University Press, Cambridge, 2006).
8. R.M. Tooke, M.G. Blyth, Phys. Fluids 20, 033101 (2008).
9. Abuzar Ghaffari, Tariq Javed, Fotini Labropulu, Thermal Sci. (2015) DOI: 10.2298/TSCI150411163G.
10. A. Ghaffari, T. Javed, A. Majeed, Transp. Porous Media 113, 245 (2016).
11. C.Y. Wang, Z. Angew. Math. Phys. 54, 184 (2003).
12. F. Labropulu, D. Li, Int. J. Non-Linear Mech. 43, 941 (2008).
13. M.G. Blyth, C. Pozrikidis, Acta Mech. 180, 203 (2005).
14. B. Santra, B.S. Dandapat, H.I. Andersson, Acta Mech. 194, 1 (2007).
15. M. Sajid, K. Mahmood, Z. Abbas, Chin. Phys. Lett. 29, 024702 (2012).
16. P.A. Thompson, S.M. Troian, Nature 389, 360 (1997).
17. K. Mahmood, M. Sajid, N. Ali, Z. Naturforsch. A 71, 273 (2016).
18. K. Mahmood, M. Sajid, N. Ali, T. Javed, Int. J. Phys. Sci. 11, 96 (2016).
19. J.W. Hoyt, A.G. Fabula, The Effect of Additives on Fluid Friction, U. S., Naval Ordinance Test Station Report, Inyokern,
USA (1964).
20. A.C. Eringen, Theory of Micro Polar Continua, in Proceedings, 9th Midwestern Conference, University of Wisconsin,
Madison, USA (1965).
21. A.C. Eringen, Int. J. Eng. Sci. 2, 205 (1964).
22. A.C. Eringen, J. Math. Mech. 16, 1 (1966).
23. G. Ahmadi, Int. J. Eng. Sci. 14, 639 (1976).
24. T. Ariman et al., Int. J. Eng. Sci. 11, 905 (1973).
25. T. Ariman et al., Int. J. Eng. Sci. 12, 273 (1974).
26. G.S. Guram, A.C. Smith, Comput. Math. Appl. 6, 213 (1980).
27. J. Peddieson, Int. J. Eng. Sci. 10, 23 (1972).
28. R. Nazar et al., Int. J. Non-Linear Mech. 39, 1227 (2004).
29. C.L. Chang, J. Phys. D: Appl. Phys. 39, 1132 (2006).
30. Y.Y. Lok et al., Int. J. Eng. Sci. 45, 173 (2007).
31. A. Ishak et al., Comput. Math. Appl. 56, 3188 (2008).
32. H.B. Keller, T. Cebeci, AIAA J. 10, 1193 (1972).
33. V. Bradshaw, T. Cebeci, I.H. Whitelaw, Engineering Calculation Methods For Turbulent Flows (Academic, London, 1981).
34. H.B. Keller, A New Difference Scheme For Parabolic Problems, in Numerical Solution of Partial Differential Equations,
edited by J. Bramble, Vol. II (Academic, New York, 1970).

You might also like