Vous êtes sur la page 1sur 16

Appl. Math. Mech. -Engl. Ed.

, 38(6), 815–830 (2017)


DOI 10.1007/s10483-017-2206-8
Applied Mathematics
c
Shanghai University and Springer-Verlag
and Mechanics
Berlin Heidelberg 2017 (English Edition)

Jeffery-Hamel flow of non-Newtonian fluid with nonlinear viscosity


and wall friction∗

J. NAGLER†
Department of Fluid Mechanics and Heat Transfer, Faculty of Engineering, Tel-Aviv University,
Ramat-Aviv 69978, Israel

Abstract A Jeffery-Hamel (J-H) flow model of the non-Newtonian fluid type inside
a convergent wedge (inclined walls) with a wall friction is derived by a nonlinear or-
dinary differential equation with appropriate boundary conditions based on similarity
relationships. Unlike the usual power law model, this paper develops nonlinear viscosity
based only on a tangential coordinate function due to the radial geometry shape. Two
kinds of solutions are developed, i.e., analytical and semi-analytical (numerical) solutions
with suitable assumptions. As a result of the parametric examination, it has been found
that the Newtonian normalized velocity gradually decreases with the tangential direction
progress. Also, an increase in the friction coefficient leads to a decrease in the normalized
Newtonian velocity profile values. However, an increase in the Reynolds number causes
an increase in the normalized velocity function values. Additionally, for the small values
of wedge semi-angle, the present solutions are in good agreement with the previous results
in the literature.
Key words Jeffery-Hamel (J-H) flow, slip condition, non-Newtonian fluid, friction,
nonlinear viscosity, analytical solution, numerical solution, approximate solution
Chinese Library Classification O373
2010 Mathematics Subject Classification 76A05

1 Introduction
Non-Newtonian fluid mechanics interactions between structures (like surfaces, channels,
tubes, and wedges) and different types of fluid have been investigated during the last four
decades. In 1940, Rosenhead[1] developed an analytical solution in terms of elliptic functions
of the two-dimensional radial viscous fluid flow inside convergent or divergent wedge shape.
About two and a half decades later, three important studies have shed to light. Tanner[2]
found a link between the apparent viscosity and the third rate of the invariant deformation in
the case of inelastic non-Newtonian fluid inside conical walls. Moreover, he suggested a model
at low Reynolds numbers that explains the viscometric experiments of inelastic fluid behavior
in a cone. Simultaneously, Tadmor[3] found analytically the velocity distribution for various
Reynolds numbers of non-Newtonian fluid inside cylindrical annuli based on the power law
model. About two decades later, Moffatt et al.[4] found analytically that velocity discontinuity
was rapidly growing with the increase in the Reynolds number in the case of two-dimensional
∗ Received May 1, 2016 / Revised Dec. 21, 2016
† Corresponding author, E-mail: syankitx@gmail.com
816 J. NAGLER

flow of unmixed fluid between converged walls with the variable viscosity. During the next
two decades, researchers, such as Hull and Pearson[5], Durban[6] , Brewster et al.[7] , and Bird
et al.[8] , used analytical methods to obtain approximate solutions for the converging incom-
pressible slow viscoelastic fluid flow through cones and wedges. Brewster et al.[7] found that
the flow possessed the outer region in the presence of complex boundary layers with transitions
at the walls using explicit solutions. Bird et al.[8] found that, in the case of two-phase flow,
the outlet flow area resistance was beyond the range of other regions where the fraction value
of air bubbles was maximal. They also managed to control the air mass ratio at the end of
the cone channel to reduce the flow resistance in the region. Shortly thereafter, Voropayev et
al.[9] , Rahimi et al.[10] , and Nofal[11] found similarity solutions for the two-dimensional steady
state incompressible viscous flow problem of non-Newtonian fluid inside the convergent and
divergent cones with and without source/sink based on the power-law model. Rahimi et al.[10]
also considered the wall friction and found that the pressure gradient was a key parameter in
the Newtonian and the non-Newtonian fluid flow behaviors, and good agreement between the
lubrication theory and asymptotic solution was found for small cone angles and small friction
parameter values.
Additional two phenomena that may be relevant in the literature context are the magneto-
hydro-dynamics (MHD) and the electroosmotic flow. A quick overview for these issues will be
brought here.
In 1970, Girishwar[12] solved analytically the steady incompressible axisymmetric flow of
electrically conducting-viscous fluid between two concentric rotating cylinders composed of an
insulating material under the influence of radial magnetic field. He found that the velocity field
component values were decreased compared with those appearing in the classical hydrodynamic,
and the tangential velocity became fully developed for the relatively small axial distance due to
the presence of magnetic field. However, it required a greater axial distance for larger Reynolds
numbers. About two decades later, Timol and Timol[13] developed a three-dimensional model
of an incompressible viscous flow over a semi-infinite plate under the influence of the magnetic
field and a pressure gradient with or without suction/injection cases, using the appropriate
similarity transformations. Recently, Makinde and Mhone[14] claimed that a small divergence
of walls might cause the non-stability behavior, while a small convergence had a stabilizing
effect for the case of small disturbances in the Jeffery-Hamel (J-H) flow type including the
MHD effect when magnetic number values were relatively small. Moreover, they found that an
increase in the magnetic field intensity had a strong stabilizing effect on both diverging and
converging channel geometries. Similarly, Sadeghy et al.[15] found analytically that the parame-
ters (the Reynolds number, the Weissenberg number, the half-angle, and the magnetic number)
had crucial effects on the velocity profile. Moreover, the magnetic field played a main role in
delaying flow dependency on fluid’s elasticity when the magnetic force was found to suppress
separation in diverging channels. Two years later, Makinde et al.[16] found analytically using
Bessel and Lommel functions that the axial velocity in the axisymmetric J-H flow through
the porous medium, decreased significantly with an increase in the Hartmann number, and
the tangential velocity was found to decrease due to the increase in the magnetic field. Other
recent studies that were produced by Alam and Khan[17] , Shrama and Singh[18] , Dib et al.[19] ,
Hayat et al.[20] , and Usman et al.[21] concentrated on external MHD field effects on the steady
two-dimensional nonlinear flow through convergent/divergent channels of viscous incompress-
ible electrically conducting fluid. Alam and Khan[17] investigated the instability range limits of
the solution using the power series that was based on Padé-Hermite approximations, and they
found the limit margin for each MHD pattern. In addition, Shrama and Singh[18] concentrated
on the natural convection flow with variable electrical conductivity and heat generation along
an isothermal vertical plate. Their main findings indicated that the fluid velocity was increased
in the presence of heat generation volumetric rate or due to the increase in the electrical conduc-
tivity parameter. Additionally, the fluid temperature was found to increase in the presence of
Jeffery-Hamel flow of non-Newtonian fluid with nonlinear viscosity and wall friction 817

heat generation volumetric rate or due to the increase in the magnetic field intensity whereas it
was decreased due to the increase in the electrical conductivity parameter (not much effective).
Advanced analytical solutions based on the Adomian decomposition method, the numerical
solution, and the homotopy method were obtained by Dib et al.[19] , Hayat et al.[20] , and Usman
et al.[21] , respectively. Hayat et al.[20] found that variable viscosity and thermal conductivity
tended to increase the pressure gradient in the case of a mixed convection peristaltic flow of
electrically conducting fluid in an inclined asymmetric channel when the fluid viscosity and the
thermal conductivity were assumed to vary linearly as a function of temperature. Likewise, the
viscosity parameter, the Grashoff number, and the inclination angle had similar effects on the
velocity while the velocity wasn’t affected by the change in the thermal conductivity parameter.
Additionally, they found that the variable viscosity tended to decrease the fluid temperature
and the viscosity parameter, and the inclination angle had similar effects on the heat transfer
rate near the wall.
Electrostatics has a tremendous effect on many aspects of our life, such as mechanical,
biological, and electrical engineering. In 1999, a research that considered the biological aspect
was performed and presented by Fogolari et al.[22] . The researchers examined electrostatics
phenomena in the context of biomolecular systems, which are linearized forms of Poisson-
Boltzmann equations (PBE) (LPBE). They compared PBE solutions with LPBE solutions and
their application to biomolecular systems. Three years later, Chen and Santiago[23] published
their work on optimal electroosmotic micro-pumps of pressure capacity and flow rate alongside
high thermodynamic efficiency between parallel plates based on the field-induced ion drag.
The researchers found that their model behavior coinciding with experimental results of actual
electroosmotic micro-pump test showed a linear ratio between the pressure and the flow rate.
In addition, setting up high electric field strength values caused the flow rate to increase with
the electrical Joule heating, which yields an upper limit for the operating voltage. Similarly, six
years later, Berli and Olivares[24] found that nonlinear effects induced by the shear-dependent
viscosity of non-Newtonian fluid are limited to the pressure-driven component of the flow.
Moreover, they found a mutual linkage between electroosmosis and streaming current when
considering wall depletion effects of colloids in polymer solutions. In 2008, Zhao et al.[25]
made their own analytical contribution to the electroosmotic flow of power-law fluid through
a slit channel. They found that the dynamic viscosity was decreased monotonically from the
centerline to the wall in the case of pseudoelastic. However, in the case of dilatant fluid, fluid is
in-viscid at the centerline, and then the viscosity increases gradually. Moreover, at the wall, the
dynamic viscosity values are close to the unit one, and the velocity near the center region of the
channel approaches the generalized Smoluchowski velocity regardless of the power law index
values. However, in the case of pseudoplastic fluid, the effect of electrical double layer thickness
parameter plays a main role in the Smoluchowski velocity compared with that in dilatant fluid.
In the same year, Bohinc et al.[26] published their analytical study on the electrolyte solution
inside the cylindrical geometry with charged surfaces. They found that the absolute value of the
electrostatic potential as well as the concentration decreases with the increasing distance from
the charged surface. One year later, Afonso et al.[27] and later Ghosal[28] built an analytical
model for the micro- and nano-fluidic viscoelastic flow between parallel plates, and analytical
solutions were developed under the combined influence of electro-kinetic and pressure forces
with the Debye-Hückel approximation.
In 2010, Vasu and De[29] published their papers on the electroosmotic flow of power-law
fluid in a rectangular microchannel at the high zeta potential. They developed the electrical
double layer potential distribution without assuming the Debye-Hückel linear approximation.
They found that average velocity values of pseudoplastic fluid were larger in the microchan-
nel than that of dilatants for the same given operating conditions. One year later, Zhao and
Yang[30] developed an exact solution for the electroosmotic flow of non-Newtonian fluid inside
the parallel plate microchannel in terms of hypergeometric functions alongside the new exper-
818 J. NAGLER

imental scheme method for determining the rheological properties based on the generalized
Smoluchowski velocity. They also found that either increasing the electro-kinetic parameter or
decreasing the power law index leads to many plug-like velocity profiles where the local velocity
distribution, the average velocity and the generalized Smoluchowski velocity coincide for the
specific case. In 2014, Sherwood et al.[31] gave the analysis on end in the context of the finite
circular cylindrical pore that transverses a membrane with the specific given thickness including
the electroosmosis phenomenon when the membrane outer surfaces are charged. They found
that, if the membrane thickness is relatively larger than the cylindrical pore radius, then the
end effects can be neglected, and both calculation results of infinite and finite cylindrical pores
coincide.
Now, three combined models of both MHD and electroosmotic flow studies are presented.
In 2011, Bhattacharyya and Layek[32] conducted a research study on the MHD boundary layer
flow of an electric conducting incompressible non-Newtonian dilatant fluid flow in a divergent
channel with suction or blowing states when channel walls are designed in the porous shape, and
the electric current was applied along the channel intersection line resulting with the magnetic
field. They found that the boundary layer flow of the dilatant fluid may be generated without
being separated if only the appropriate suction or blowing velocity is being used through its
porous walls. Moreover, they showed that this kind of flow can even exist without the magnetic
field. Moreover, it is found that the boundary layer flow may occur even in the presence of
blowing. About three years later, Escandón et al.[33–34] presented their study on temperature
distributions in a parallel flat plate microchannel with composed steady electroosmotic and
magneto-hydrodynamic driven forces. The analysis[33] was based on the Phan-Thien-Tanner
model, and their main conclusion was that the undesirable Joule heating in the microchannels
can be reduced by a combination of both forces. In addition, the researchers[34] found that
the limits of small Hartmann numbers and low electrical conductivity in the buffer solution
correspond to the range where the electric and magnetic effects can be used to move a charged
solution in the flow control.
The current paper develops the J-H flow of non-Newtonian fluid with nonlinear viscosity
and wall slip conditions based on the previous studies (see Refs. [35]–[37]) inside a convergent
wedge. Unlike the usual power law model, this paper develops nonlinear viscosity which is
dependent only on the tangential coordinate function.

2 Problem formulation model

Consider a steady two-dimensional flow of incompressible viscous fluid inside a convergent


wedge (non-parallel walls) with friction on the walls. The cylindrical polar coordinates (r, θ, z)
are selected to model the flow system as shown in Fig. 1, and the flow is intersected by the z-axis.
It is assumed that the motion is purely radially dependent on r and θ only. Moreover, there
are no changes with respect to the z-axis. The governing equations of motion are elaborated
here as follows. The equation of continuity can be written as[38]

ρ ∂
∇·V =0 → (rur ) = 0, (1)
r ∂r

where V , ρ, and ur are the velocity vector, the constant fluid density, and the radial velocity
component, respectively.
Now, according to the general equation of momentum, it can be written as

 ∂V 
ρ + (V · ∇)V = −∇p + ∇ · τ + FB , (2)
∂t
Jeffery-Hamel flow of non-Newtonian fluid with nonlinear viscosity and wall friction 819

Fig. 1 Power law wedge shape flow model

where FB and τ are the body forces and stress vectors, respectively. Body forces are assumed to
be neglected. The following viscosity terms are dependent on the tangential direction function:
(
µ = µ0 g n−1 , g = g(θ), τ = µ(∇V +(∇V )T ),
∂ur ur µ ∂ur (3)
τrr = −p + 2µ , τθθ = −p + 2µ , τrθ = ,
∂r r r ∂θ
∂(·)
where uθ = uz = 0, and = 0 (no velocities in the perpendicular and tangential axes). The
∂z
general velocity gradient (∇V ) can be represented by the Jacobian matrix as shown in Ref. [39].
Also, µ is the dynamic viscosity function, where µ0 and n represent the material parameter and
the flow consistency index, respectively. As introduced by Zhao et al.[25] , Zhao and Yang[30]
and Rahimi et al.[10] , the apparent (effective) viscosity is defined as a function of the shear rate.
However, in this current study, the apparent viscosity will be assumed to be dependent on the
tangential direction only. Usually, the apparent viscosity is defined as a function of the shear
rate including the radial direction. However, the study will examine the nonlinear viscosity as
a function of the tangential direction regardless the radial direction since the geometry effects
that will be considered are in the tangential direction. Substituting the above relationships (3)
into Eq. (2) results in the following radial and tangential displacement equations:

∂ur ∂P  1 ∂ 1 ∂τrθ τθθ 


rb : ρur =− − (rτrr ) + −
∂r ∂r r ∂r r ∂θ r
∂P ∂µ ∂ur 1 ∂µ ∂ur  ∂2u 1 ∂ 2 ur 1 ∂ur ur 
r
=− +2 + 2 +µ 2
+ 2 2
+ − 2 , (4)
∂r ∂r ∂r r ∂θ ∂θ ∂r r ∂θ r ∂r r
1 ∂P 1 ∂ 1 ∂τ  1 ∂P 1 ∂µ ∂u 2 ∂µ 2µ ∂u
θθ r r
θb : − 2
− 2 (r τrθ )+ = − − ur − 2 = 0. (5)
ρr ∂θ r ∂r r ∂θ ρr ∂θ ρr ∂r ∂θ ρr2 ∂θ ρr ∂θ

The equations of momentum are derived by using the fully cylindrical coordinates flow equations
as developed by Membrado and Pacheco[40]. Hence, Eq. (4) is obtained as

1 ∂P 2 ∂µ ∂ur 1 ∂µ ∂ur µ  ∂ 2 ur 1 ∂ 2 ur 1 ∂ur ur  ∂ur


− + + 2 + 2
+ 2 2
+ − 2 − ur = 0. (6)
ρ ∂r ρ ∂r ∂r ρr ∂θ ∂θ ρ ∂r r ∂θ r ∂r r ∂r

Differentiating Eq. (6) along the tangential direction leads to

1 ∂2P 2 ∂ 2 µ ∂ur 2 ∂µ ∂ 2 ur 1 ∂ 2 µ ∂ur 1 ∂µ ∂ 2 ur


− + + + 2 2 + 2
ρ ∂r∂θ ρ ∂r∂θ ∂r ρ ∂r ∂r∂θ ρr ∂θ ∂θ ρr ∂θ ∂θ2

1 ∂µ ∂ 2 ur 1 ∂ 2 ur 1 ∂ur ur 
+ 2
+ 2 + − 2 + ···
ρ ∂θ ∂r r ∂θ2 r ∂r r
µ  ∂ 3 ur 1 ∂ 3 ur 1 ∂ 2 ur 1 ∂ur  ∂ur ∂ur ∂ 2 ur
+ 2
+ 2 3
+ − 2 − − ur = 0. (7)
ρ ∂r ∂θ r ∂θ r ∂r∂θ r ∂θ ∂θ ∂r ∂r∂θ
820 J. NAGLER

Multiplying Eq. (5) by r together with differentiation performance along the radial direction
yields

1 ∂2P 1 ∂ 2 µ ∂ur 1 ∂µ ∂ 2 ur 2 ∂2µ 2 ∂µ 2 ∂µ ∂ur 2µ ∂ 2 ur


− 2
− − ur + 2 ur − −
ρ ∂r∂θ ρ ∂r ∂θ ρ ∂r ∂r∂θ ρr ∂r∂θ ρr ∂θ ρr ∂θ ∂r ρr ∂r∂θ
2µ ∂ur 2 ∂µ ∂ur
+ 2 − = 0. (8)
ρr ∂θ ρr ∂r ∂r
Therefore, by substituting Eq. (8) into Eq. (7), the following nonlinear differential equation is
obtained:
2 ∂ 2 µ ∂ur 1 ∂µ ∂ 2 ur 1 ∂ 2 µ ∂ur 1 ∂µ ∂ 2 ur 1 ∂µ  ∂ 2 ur 1 ∂ 2 ur 1 ∂ur
+ + 2 2 + 2 + + −
ρ ∂r∂θ ∂r ρ ∂r ∂r∂θ ρr ∂θ ∂θ ρr ∂θ ∂θ2 ρ ∂θ ∂r2 r2 ∂θ2 r ∂r
ur   3
µ ∂ ur 3
1 ∂ ur 2
1 ∂ ur 1 ∂ur  2
2 ∂ µ 2 ∂µ ∂ur ∂ur ∂ur
+ 2 + ···+ + − + − ur − −
r ρ ∂r2 ∂θ r2 ∂θ3 r ∂r∂θ r2 ∂θ ρr ∂r∂θ ρr ∂r ∂r ∂θ ∂r
2 2
∂ ur 1 ∂ µ ∂ur
− ur − =0 (9)
∂r∂θ ρ ∂r2 ∂θ
with the slip boundary conditions as follows:
∂ur
= ∓mu(θ), (10)
∂θ θ=±α
where m > 0 is the friction coefficient factor. These conditions represent the relative slip
friction between the fluid and the wall. A smooth boundary is described by m = 0, and the
perfectly rough wall is obtained by assuming m = ∞. Similar studies on the wall friction
condition (10) are available for a variety of non-Newtonian fluid such as elasto-viscoplastic and
structural fluid[35,41–43] . The third condition is joined to the other two conditions (10), which
provides[35]
Z Z α
Q= ur dA = − ur rdθ, (11)
A −α

where A represents the cross sectional radial surface area of the flow field with the element area
A = rdθ within the deformation zone. α represents the wedge semi-angle where the flow field
domain is confined between rigid rough walls (θ = ±α) that enable the slip condition. Q is the
steady state planar flow rate. In order to simplify the model, the following transformation will
be used using Eq. (1) and the physical description as shown in Fig. 1:

f (θ)
ur = −Q . (12)
r
Thus, substituting Eq. (12) into Eq. (9) yields

4Q ∂ 2 µ Q ∂µ ′ Q ∂2µ ′ Q ∂µ ′′ 1 ∂µ  Q ′′ 4Q  µ  4Q ′ Q ′′′ 
f + f − f − f − f + 3f − f + 3f
ρr2 ∂r∂θ ρr2 ∂r ρr3 ∂θ2 ρr3 ∂θ ρ ∂θ r3 r ρ r3 r
2Q ∂µ Q2 ′ Q ∂ 2 µ ′
− 3 f + 2 3 ff + f = 0. (13)
ρr ∂r r ρr ∂r2
Then, multiplying both sides of Eq. (13) by r3 (r 6= 0) and the density ρ and dividing Eq. (13)
by Q (Q =
6 0) yield

∂2µ  ∂µ ∂ 2 µ  ∂µ ∂µ ∂2µ
4r f+ r − 2 f ′ − (2f ′′ +4f)−µ(4f ′ +f ′′′ )−2 f +2Qρf f ′ +r2 2 f ′ = 0. (14)
∂r∂θ ∂r ∂θ ∂θ ∂r ∂r
Jeffery-Hamel flow of non-Newtonian fluid with nonlinear viscosity and wall friction 821
∂µ
Assuming ∂r = 0 leads to the following nonlinear differential equation:

∂ 2 µ ′ ∂µ
f − (2f ′′ + 4f ) − µ(4f ′ + f ′′′ ) + 2Qρf f ′ = 0. (15)
∂θ2 ∂θ
Substituting Eq. (3) into Eq. (15) yields

(µ0 (n − 1)g ′′ g n−1 + µ0 (n − 1)2 g ′ g n−2 )f ′ − µ0 (n − 1)g ′ g n−1 (2f ′′ + 4f )


− µ0 g n−1 (4f ′ + f ′′′ ) + 2Qρf f ′ = 0. (16)

Now, in order to simplify solutions, two suggestions for g(θ) will be introduced,
 p
g1 (θ) = eκθ when p and κ 6= 0, (17a)
g(θ) =
g2 (θ) = dθs when s and d 6= 0. (17b)

Note that the case of the constant g function was solved and discussed by the author in the
context of nano-fluid (see Ref. [44]). According to Eq. (17), substituting g(θ) into Eq. (16) yields
the following two equilibrium states:

((n − 1)((p − 1)θp−2 + (κp)θ2p−2 )eκn + (n − 1)2 θp−1 eκ(n−1)θ )κpf ′


p

p p
− κp(n − 1)θp−1 eκnθ (2f ′′ + 4f ) − eκ(n−1)θ (4f ′ + f ′′′ ) + 2Ref f ′ = 0, (18)
 s 
(n − 1)(s − 1)θsn−2 + (n − 1)2 θsn−s−1 f ′ − s(n − 1)θsn−1 (2f ′′ + 4f )
d
θs(n−1) Re
− (4f ′ + f ′′′ ) + 2 n f f ′ = 0, (19)
d d

where Re = µ0 , and the appropriate forms of conditions (10) and (11) are

f ′ |θ=±α = ∓mf |θ=±α , (20)


Z α
f dθ = 1. (21)
−α

Due to the flow field symmetry assumption, the condition (20) is simply written as

f ′ (α) = −mf (α), f ′ (0) = 0. (22)

In the next section, an approximate analytical solution to the problem will take place for special
cases.

3 Approximate solutions

In this section, an analysis of Eqs. (18)–(19) will be discussed here for an extreme parameter
value. Moreover, Eqs. (18)–(19) will be solved numerically using MATLAB. The numerical
procedure based on the initial value problem (IVP) will be introduced here. Firstly, Eq. (18)
will be examined. Suppose that we have Newtonian fluid (n = 1) where the inertia effect is
neglected (Re → 0 or µ → ∞), and then the obtained differential equation is

4f ′ + f ′′′ = 0. (23)

The appropriate solution to Eq. (23) is


c1
f |n=1 = a1 cos(2θ) + (24)
4
822 J. NAGLER

with the appropriate constants


1 1 − 0.5c1 α
c1 = sin(2α)
, a1 = . (25)
0.5α + m sin(2α)
8 sin(2α)−0.5m cos(2α)

Now, assume that the half cone angle is small enough such that 0◦ < θ 6 20◦ , and then by
using the Taylor series, the exponential function can be expanded and approximated by the
following relationship:

G2
eG(θ) ≈ 1 + G + , G = κθp . (26)
2!
Note that the numerical error is smaller than 10−2 . Accordingly, the following nonlinear differ-
ential equation is obtained:
  κ2 (n − 1)2 θ2p 
(n − 1)((p − 1)θp−2 + (κp)θ2p−2 )eκn + (n − 1)2 θp−1 1 + κ(n − 1)θp + κpf ′
2
 κ2 n2 θ2p  ′′  κ2 (n − 1)2 θ2p 
− κp(n − 1)θp−1 1 + κnθp + (2f + 4f ) − 1 + κ(n − 1)θp +
2 2
· (4f ′ + f ′′′ ) + 2Ref f ′ = 0. (27)

Equation (27) cannot be solved analytically without far reaching approximations but only by
the complex series solution and will not be discussed here. We will turn to develop semi-
analytical approximate solutions to Eq. (19) based on the series solution. Firstly, suppose that
we have non-Newtonian fluid (n > 0, n =6 1) which fulfills d → ∞,

((s − 1)θsn−2 )f ′ − sθsn−1 (2f ′′ + 4f ) = 0. (28)

Now, Eq. (27) can be separated into two distinct equations by the following conditions:

f ′′ + 2f = 0, when s = 1, (29a)
2s
f′ − θ(f ′′ + 2f ) = 0, when s 6= 1. (29b)
s−1

The solution to Eq. (29a) under conditions (21)–(22) is


√ √
f |n6=1,s=1 = e1 cos( 2θ) + e2 sin( 2θ), (30)

where
1 √ √
e1 = √ √ , e2 = 0, m= 2 tan( 2α). (31)
2 sin( 2α)

Note that the friction coefficient value is also determined and dependent on the wedge semi-
angle.
In the case of Newtonian fluid (n = 1) and d 6= 0, Eq. (19) turns to be

4f ′ + f ′′′ − 2Ref f ′ = 0. (32)

Note that Eq. (32) has no dependency on the constant d. Integrating Eq. (32) along the θ-
direction leads to the following equation:

4f + f ′′ − Ref 2 = c1 , (33)
Jeffery-Hamel flow of non-Newtonian fluid with nonlinear viscosity and wall friction 823

where c1 is a constant. Now, numerically and analytically approximate solutions to Eq. (33)
will be presented. Suppose that f 2 ≈ ±f and 0◦ < θ 6 20◦ , then, the final solutions, including
homogenous and particular parts, are
c1 √
f+ = A1 sin(β1 θ) + A2 cos(β1 θ) + , β1 = 4 − Re, (34a)
β12
c1 √
f− = A1 sin(β2 θ) + A2 cos(β2 θ) + 2 , β2 = 4 + Re. (34b)
β2

Now, the constants A1 , A2 , and c1 will be determined by applying conditions (21)–(22) to


Eqs. (34a) and (34b),

 t2 m 1
c1 = A2 (t sin(tα) − m cos(tα)), A2 = m ,
√m 2α αt sin(tα) + t sin(tα) − m cos(tα) (35)

t± = 4 ∓ Re, A1 = 0.

In the case that the Reynolds number becomes zero, Eqs. (34a) and (34b) turn to be the
solution (24). Another method for solving Eq. (32) will be performed without using the as-
sumption (f 2 ∼ ±f, 0◦ < θ 6 20◦ ). The following method for solving Eq. (32) is based on the
following series:

X
f (θ) = ak θ k , (36)
k=0

where the previous assumption (f 2 ≈ ±f, 0◦ < θ 6 20◦ ) is not considered. Substituting Eq. (36)
into Eq. (32) yields

X ∞
X ∞
X k
X
ak k(k − 1)θk−2 + 4 ak θk − Re θk ak−j aj = c1 . (37)
k=2 k=0 k=0 j=0

Applying subscripts manipulations leads to


∞ 
X k
X 
(k + 2)(k + 1)ak+2 + 4ak − Re ak−j aj θk = c1 . (38)
k=0 j=0

The coefficient development of Eq. (38) for the first eight members including conditions (21)–
(22) applied to Eq. (35) yields
 m

 αa2 + 2α3 a4 + 3α5 a6 + 4a8 α7 = − (a0 + α2 a2 + α4 a4 + α6 a6 + α8 a8 ),

 2
 3 5 7 9
 2a α + 2a α + 2a α + 2a α + 2a α = 1,

0 2 4 6 8
3 5 7 9 (39)

 12a4 + 4a2 − 2Rea0 a2 = 0,

 2

 30a6 + 4a4 − Re(2a0 a4 + a2 ) = 0, 2

56a8 + 4a6 − Re(2a8 a0 + 2a6 a2 + a4 ) = 0

with five unknown constants a0 , a2 , a4 , a6 , and a8 . Note that the constant value c1 is determined
by Eq. (37), but does not participate in the solution process. The first eight coefficients are
sufficient to use in order to obtain accurate solutions, especially for small wedge semi-angles
as shown in the next section. The MATLAB program will be used in order to have a closed
solution by using the FSOLVE function. In the next section, comparison will be performed
between the approximate analytical solutions and the previous results in the literature.
824 J. NAGLER

In the case that n and s 6= 1, the obtained equilibrium is derived by substituting the series
solution substitution (35) into Eq. (29b),

2s X 

X ∞
(k + 2)bk+2 θk − (bk+2 (k + 2)(k + 1) + 2bk )θk = 0. (40)
s−1
k=3 k=0

The coefficient development of Eq. (40) for the first eight members including conditions (21)–
(22) applied to (36) yields
 m


 αb2 + 2α3 b4 + 3α5 b6 + 4b8 α7 = − (b0 + α2 b2 + α4 b4 + α6 b6 + α8 b8 ),

 2


 α3 α5 α7 α9

 2b 0 α + 2b 2 + 2b 4 + 2b 6 + 2b8 = 1,

 3 5 7 9
b2 + b0 = 0,
(41)

 6b 4 + b2 = 0,

 3 s

 b6 − (15b6 + b4 ) = 0,



 2 s − 1

 2b8 − s
(28b8 + b6 ) = 0.
s−1

The suggested parabolic approximations which fulfill boundary conditions (21)–(22) are
brought as follows[35] :
 m 2

 2
1 − 2α+mα 2θ 2α + mα2 − mθ2
 f
 1 (θ) ≈ b 1 (1 − b 2 θ ) = = ,
4α2 + 34 mα3
3
2α − 23 2α+mα

2
(42)

 1 1 m
 b1 = 2α − 2b α3 = 2α − 2 mα3 , b2 = 2α + mα2 ,

2 3 3 2α+mα2

where α 6= 0.

4 Results and discussion

In this section, parametric investigations based on solutions (24), (34), (39), and (41)–(42)
through illustrative results will be presented and discussed. Solutions (24) and (41)–(42) will be
discussed separately. Initially, suppose that the following parameters are considered: α = 2.5◦ ,
and Re = 3. According to the Newtonian case (n = 1) where the variable viscosity behaves
according to Eq. (17b) when d 6= 0, two kinds of flow field solutions have been obtained, i.e., the
analytical solutions (34a)–(34b) and the numerical solution (39), respectively. Seemingly, the
normalized Newtonian velocity (f ) is found to decrease gradually with the tangential direction
progress as shown in Fig. 2. Moreover, the increase in the friction coefficient (m) causes the
normalized Newtonian velocity profile values to decrease as shown in Fig. 2. Here, the case that
the friction coefficient (m) value becomes large enough to represent “infinity” is for m = 106 , and
the condition (22) becomes the no-slip boundary condition f (α) = 0. Moreover, the obtained
solutions (34a), (34b), and (39) coincide for the specific data. The obtained solution has good
agreement with Refs. [17] and [19]. The opposite extreme case that m = 0 or f ′ (α) = 0 is
illustrated in Fig. 2 with the horizontal line parallel to the θ-axis, and good agreement is found
with Ref. [35].
However, the increase in the wedge semi-angle (α) value causes the normalized Newtonian
velocity profile to decrease as shown in Fig. 3, good agreement is found between solutions for
the small wedge semi-angle, and the most accurate solution is the numerical (semi-analytical)
solution (39).
Jeffery-Hamel flow of non-Newtonian fluid with nonlinear viscosity and wall friction 825





 &R B
GPSN
&R C
GPSN
 &R 
GPSN

G

 &R B


GPSN
&R C
GPSN
 &R 
GPSN
 &R B
GPSN
 &R C
GPSN
&R 
GPSN
     
DŽ
( )
Fig. 2 Comparison of normalized Newtonian velocity f (θ) between solutions (34a), (34b), and (39)
for various values of friction coefficient (m) when α = 2.5◦ , and Re = 3

&R B
GPSDŽ
 &R C
GPSDŽ
&R 
GPSDŽ

&R B
GPSDŽ
 &R C
GPSDŽ
&R 
GPSDŽ

&R B
GPSDŽ
 &R C
GPSDŽ
&R 
GPSDŽ
G






          
DŽ
( )
Fig. 3 Comparison of normalized Newtonian velocity f (θ) between solutions (34a), (34b), and (39)
for various values of wedge semi-angle (α) when m = 100, and Re = 3
The initial observation on the Reynolds number (Re) in Fig. 4 shows that the normalized
Newtonian velocity profile values increase as long as Re increases for the chosen parameters
(m = 100, and α = 10◦ ).

Fig. 4 Comparison of normalized Newtonian velocity f (θ) between solutions (34a) and (39) for var-
ious values of Reynolds number (Re) when m = 100, and α = 10◦
826 J. NAGLER

Additionally, for the specific extreme parameters choice (Re = 250, α = 20◦ ), when the
Reynolds number and the wedge semi-angle are relatively large, the normalized Newtonian
velocity (34b) in Fig. 5 does not show physical behavior.








G



 &R C


GPSN
&R 
GPSN
 &R C
GPSN
&R 
GPSN

          
DŽ
( )
Fig. 5 Comparison of normalized Newtonian velocity f (θ) between solutions (34b) and (39) for var-
ious values of friction coefficient (m) when Re = 250, and α = 20◦

Now, we will turn to discuss on another comparison between the parabolic approximation
(42) and the Newtonian flow solution (24). The observation shows that both solutions have no
dependency on the Reynolds number and coincide for each value of the friction coefficient (m)
as shown in Fig. 6. Similarly, both solutions profiles decrease as long as the friction coefficient
increases as shown in Fig. 6. Moreover, like before, it is also obtained that the increase in the
wedge semi-angle value (α) causes the velocity profile to decrease, as shown in Fig. 7.
Finally, the last case that will be discussed here deals with non-Newtonian fluid with non-
linear viscosity that behaves according to (29b) and fulfills d → ∞. It is found that the variable
viscosity power constant (s) has no effect on the solution. In addition, relatively large values of
the wedge semi-angle and friction coefficients have significant effects on the normalized velocity
profile, as shown in Figs. 8 and 9.
In conclusion, it seems that both non-Newtonian and Newtonian flows have similar qualita-
tive behaviors, and the nonlinear viscosity term has influential effects on the flow.







G

&R 
GPSN
 &R 
GPSN
 &R 
GPSN
&R 
GPSN

&R 
GPSN

&R 
GPSN
     
DŽ
( )
Fig. 6 Comparison of normalized Newtonian velocity f (θ) between solutions (24) and (42) for various
values of friction coefficient (m) when α = 2.5◦
Jeffery-Hamel flow of non-Newtonian fluid with nonlinear viscosity and wall friction 827

&R 
GPSDŽ
 &R 
GPSDŽ
 &R 
GPSDŽ
&R 
GPSDŽ

&R 
GPSDŽ
 &R 
GPSDŽ



G






          
DŽ
( )
Fig. 7 Comparison of normalized Newtonian velocity f (θ) between solutions (24) and (42) for various
values of wedge semi-angle (α) when m = 100






G

&R 
GPSN
 &R 
GPSN
 &R 
GPSN
&R 
GPSN

&R 
GPSN
 &R 
GPSN

     


DŽ
( )
Fig. 8 Normalized non-Newtonian velocity f (θ) based on Eq. (41) for various values of friction coef-
ficient (m) when α = 2.5◦

Fig. 9 Normalized non-Newtonian velocity f (θ) based on Eq. (41) for various values of wedge semi-
angle (α) when m = 106

5 Concluding remarks

The J-H flow model of the non-Newtonian flow type inside a convergent wedge (inclined
walls) with the wall friction is derived and expressed with a nonlinear ordinary differential
828 J. NAGLER

equation with appropriate boundary conditions based on similarity relationships. Unlike the
usual power law model, this paper develops the nonlinear viscosity which is based only on the
tangential coordinate function. Two kinds of solutions are developed, i.e., analytical and semi-
analytical (numerical) solutions. However, the analytical solution is developed for Newtonian
and non-Newtonian cases, and the non-Newtonian analytical solution relies on the assumption
that f 2 ∼ ±f and 0 < θ 6 20◦ . On the other hand, the non-Newtonian semi-analytical solution
is not developed by using the latter assumption. In addition, the parabolic approximation is
used to compare between the Newtonian analytical solution and the certain approximation.
Parametric investigations including comparison between various solutions are performed.
It is found that the normalized Newtonian velocity decreases gradually with the tangential
direction progress. Also, the increase in the friction coefficient (m) causes the decrease in the
normalized Newtonian velocity profile values. Additionally, for small values of wedge semi-
angle, the present solutions are in good agreement with the previous results in the literature.
The increase in the wedge semi-angle value causes the decrease in the normalized velocity
profile, although the increase in the Reynolds number (Re) causes the normalized velocity
function values to increase. Additionally, for the specific parameter choice, when the Reynolds
number or the wedge semi-angle is relatively large, the normalized velocity behavior based on
the analytical solution is not found to be suitable for the solution.
Moreover, the parabolic approximation shows excellent agreement with another Newtonian
flow solution, and both solutions have no dependency on the Reynolds number. Similarly, both
solutions profiles decrease as long as the friction coefficient (m) or the wedge semi-angle value
(α) increases.
Finally, the specific case of the non-Newtonian fluid with nonlinear viscosity behavior is
examined. It is found that the relatively large value of wedge semi-angle or friction coefficient
has significant effects on the normalized velocity profile.
To sum it up, both non-Newtonian and Newtonian flows have similar qualitative behaviors,
and the nonlinear viscosity term has influential effects on the flow.

References
[1] Rosenhead, L. The steady two-dimensional radial flow of viscous fluid between two inclined plane
walls. Proceedings of the Royal Society of London Series A, Mathematical and Physical Sciences,
175, 436–467 (1940)
[2] Tanner, R. I. Non-Newtonian fluid parameter estimation using conical flows. Industrial and En-
gineering Chemistry Fundamentals, 5, 55–59 (1966)
[3] Tadmor, Z. Non-Newtonian tangential flow in cylindrical annuli IV. Polymer Engineering and
Science, 6, 203–212 (1966)
[4] Moffatt, H. K., Hooper, A., and Duffy, B. R. Flow of non-uniform viscosity in converging and
diverging channels. Journal of Fluid Mechanics, 117, 283–304 (1982)
[5] Hull, A. M. and Pearson, J. R. A. On the converging flow of viscoelastic fluids through cones and
wedges. Journal of Non-Newtonian Fluid Mechanics, 14, 219–247 (1984)
[6] Durban, D. On generalized radial flow patterns of viscoplastic solids with some applications.
International Journal of Mechanical Sciences, 28, 97–110 (1986)
[7] Brewster, M. E., Chapman, S. J., Fitt, A. D., and Please, C. P. Asymptotics of slow flow of
very small exponent power-law shear-thinning fluids in a wedge. European Journal of Applied
Mathematics, 6, 559–571 (1995)
[8] Bird, C., Breward, C. J. W., Dellar, P., Edwards, C. M., Kaouri, K., Richardson, G., and Wilson,
S. K. Mathematical modeling of pipe-flow and extrusion of composite materials. European Study
Group with Industry, Unilever Research, Lancaster, UK, G1–G16 (2002)
[9] Voropayev, S. I., Smirnov, S. A., and Testik, F. Y. On the case when steady converging/diverging
flow of a non-Newtonian fluid in a round cone permits an exact solution. Mechanics Research
Communications, 31, 477–482 (2004)
Jeffery-Hamel flow of non-Newtonian fluid with nonlinear viscosity and wall friction 829

[10] Rahimi, S., Durban, D., and Khosid, S. Wall friction effects and viscosity reduction of gel propel-
lants in conical extrusion. Journal of Non-Newtonian Fluid Mechanics, 165, 782–792 (2010)
[11] Nofal, T. A. An approximation of the analytical solution of the Jeffery-Hamel flow by homotopy
analysis method. Applied Mathematical Sciences, 5, 2603–2615 (1970)
[12] Girishwar, N. Tangential velocity profile for axial flow through two concentric rotating cylinders
with radial magnetic field. Defence Science Journal, 20, 207–212 (1970)
[13] Timol, M. G. and Timol, A. G. Three-dimensional magnetofluiddynamic flow with pressure gra-
dient and fluid injection. Journal of Engineering Mathematics, 19, 208–216 (1988)
[14] Makinde, O. D. and Mhone, P. Y. Temporal stability of small disturbances in MHD Jeffery-Hamel
flows. Computers and Mathematics with Applications, 53, 128–136 (2007)
[15] Sadeghy, K., Khabazi, N., and Taghavi, S. M. Magnetohydrodynamic (MHD) flows of viscoelastic
fluids in converging/diverging channels. International Journal of Engineering Science, 45, 923–938
(2007)
[16] Makinde, O. D., Bég, O. A., and Takhar, H. S. Magnetohydrodynamic viscous flow in a rotating
porous medium cylindrical annulus with an applied radial magnetic field. International Journal
of Applied Mathematics and Mechanics, 5, 68–81 (2009)
[17] Alam, M. S. and Khan, M. A. H. Critical behviour of the MHD flow in convergent-divergent
channels. Journal of Naval Architecture and Marine Engineering, 7, 83–93 (2010)
[18] Shrama, P. R. and Singh, G. Steady MHD natural convection flow with variable electrical con-
ductivity and heat generation along an isothermal vertical plate. Tamkang Journal of Science and
Engineering, 13, 235–242 (2010)
[19] Dib, A., Haiahem, A., and Bou-Said, B. An analytical solution of the MHD Jeffery-Hamel flow
by the modified adomian decomposition method. Computers and Fluids, 102, 111–115 (2014)
[20] Hayat, T., Abbasi, F. M., Ahmad, B., and Alsaedi, A. MHD mixed convection peristaltic flow
with variable viscosity and thermal conductivity. Sains Malaysiana, 43, 1583–1590 (2014)
[21] Usman, M., Naheed, Z., Nazir, A., and Mohyud-Din, S. T. On MHD flow of an incompressible
viscous fluid. Journal of the Egyptian Mathematical Society, 22, 214–219 (2014)
[22] Fogolari, F., Zuccato, P., Esposito, G., and Viglino, P. Biomolecular electrostatics with the lin-
earized Poisson-Boltzmann equation. Biophysical Journal, 76, 1–16 (1999)
[23] Chen, C. H. and Santiago, J. G. A planar electroosmotic micropump. Journal of Microelectrome-
chanical Systems, 11, 672–683 (2002)
[24] Berli, C. L. A. and Olivares, M. L. Electrokinetic flow of non-Newtonian fluids in microchannels.
Journal of Colloid and Interface Science, 320, 582–589 (2008)
[25] Zhao, C., Zholkovskij, E., Masliyah, J. H., and Yang, C. Analysis of electroosmotic flow of power-
law fluids in a slit microchannel. Journal of Colloid and Interface Science, 326, 503–510 (2008)
[26] Bohinc, K., Iglic̆ A., and Slivnik, T. Linearized Poisson Boltzmann theory in cylindrical geometry.
Elektrotehniski Vestnik, 75, 82–84 (2008)
[27] Afonso, A. M., Alves, M. A., and Pinho, F. T. Analytical solution of mixed electro-
osmotic/pressure driven flows of viscoelastic fluids in microchannels. Journal of Non-Newtonian
Fluid Mechanics, 159, 50–63 (2009)
[28] Ghosal, S. Mathematical modeling of electrokinetic effects in micro and nano fluidics. Microfluidics
and Microfabrication (ed. Chakraborty, S.), Springer Science + Business Media, LLC, New York,
87–112 (2010)
[29] Vasu, N. and De, S. Electroosmotic flow of power-law fluids at high zeta potentials. Colloids and
Surfaces A: Physicochemical and Engineering Aspects, 368, 44–52 (2010)
[30] Zhao, C. and Yang, C. An exact solution for electroosmosis of non-Newtonian fluids in microchan-
nels. Journal of Non-Newtonian Fluid Mechanics, 166, 1076–1079 (2011)
[31] Sherwood, J. D., Mao, M., and Ghosal, S. Electroosmosis in a finite cylindrical pore: simple
models of end effects. Langmuir, 30, 9261–9272 (2014)
[32] Bhattacharyya, K. and Layek, G. C. MHD boundary layer flow of dilatant fluid in a divergent
channel with suction or blowing. Chinese Physics Letters, 28, 084705 (2011)
830 J. NAGLER

[33] Escandón, J. P., Santiago, F., and Bautista, O. E. Temperature distributions in a parallel flat
plate microchannel with electroosmotic and magnetohydrodynamic micropumps. Proceedings of
the 2013 International Conference on Mechanics, Fluids, Heat, Elasticity and Electromagnetic
Fields, 177–181 (2013)
[34] Escandón, J. P., Bautista, O. E., Santiago, F., and Méndez, F. Asymptotic analysis of non-
Newtonian fluid flow in a microchannel under a combination of EO and MHD micropumps. Defect
and Diffusion Forum, 348, 147–152 (2014)
[35] Nagler, J., Durban, D., and Khosid, S. On Planar Radial Flow with Non-Uniform Viscosity and
Wall Friction Without Inertia Effect, M. Sc. dissertation, Technion-Israel Institute of Technology
(2012)
[36] Nagler, J. Laminar boundary layer model for power-law fluids with non-linear viscosity. WSEAS
Transcations on Fluid Mechanics, 9, 19–25 (2014)
[37] Nagler, J. Numerical and approximate solutions of boundary layer development due to a moving
extensible surface. WSEAS Transcations on Fluid Mechanics, 10, 54–68 (2015)
[38] Bird, R. B., Stewart, W. E., and Lightfoot, E. N. Transport Phenomena, John Wiley and Sons,
New York (2001)
[39] Landau, L. D. and Lifshitz, E. M. Fluid Mechanics, 2nd ed., Butterworth Heinemann, Oxford
(1997)
[40] Membrado, M. and Pacheco, A. F. Equations in curvilinear coordinates for fluids with non-
constant viscosity. Latin-American Journal of Physics Education, 5, 702–708 (2011)
[41] Sherwood, J. D. and Durban, D. Squeeze flow of a power law viscoplastic solid. Journal of Non-
Newtonian Fluid Mechanics, 62, 35–54 (1996)
[42] Kaylon, D. M., Yaras, P., Aral, B., and Yilmazer, U. Rheological behavior of a concentrated
suspension: a solid rocket fuel simulant. Journal of Rheology, 37, 35–53 (1993)
[43] Barnes, H. A. A review of the slip (wall depletion) of polymer solution emulsions and parti-
cle suspensions in viscometers: its cause, character, and cure. Journal of Non-Newtonian Fluid
Mechanics, 56, 225–251 (1995)
[44] Nagler, J. Advanced Non-Newtonian and Nano Fluidics Flows, LAP Publishing, Germany, 83–91
(2016)

Vous aimerez peut-être aussi