You are on page 1of 9

Applied Mathematical Modelling 46 (2017) 572–580

Contents lists available at ScienceDirect

Applied Mathematical Modelling


journal homepage: www.elsevier.com/locate/apm

Elasto-viscoplastic fluid flow in tubes of arbitrary


cross-section
Mario F. Letelier a, Dennis A. Siginer a,b,∗, Amaru González a
a
Centro de Investigación en Creatividad y Educación Superior and Departamento de Ingeniería Mechánica, Universidad de Santiago de
Chile, Avenida Libertador Bernardo O’Higgins No. 3363 Santiago, Chile
b
Department of Mathematics and Statistical Sciences and Department of Mechanical, Energy and Industrial Engineering, Bostwana
International University of Science and Technology, Private Bag 16, Palapye, Botswana

a r t i c l e i n f o a b s t r a c t

Article history: Elasto-viscoplastic fluid flow in tubes of arbitrary cross-section is explored. A constitutive
Received 13 September 2016 equation superposing elasticity on the viscoplastic behavior through a linear combination
Revised 9 January 2017
of the non-linear simplified Phan–Thien–Tanner constitutive model of viscoelasticity and
Accepted 17 January 2017
the non-linear Bingham constitutive equation of viscoplasticity is developed. The equa-
Available online 25 January 2017
tion of motion is solved analytically for the longitudinal field for steady flow in tubes of
Keywords: non-circular cross-section. The evolution of the plug and stagnant zones, hallmarks of vis-
Elasto-viscoplastic fluid flow coplastic behavior, is studied when elasticity is present, and the rate of flow is determined
Arbitrary cross-sectional tube in terms of the Weissenberg and Bingham numbers. Elasticity tends to enhance the rate
Weissenberg and Bingham numbers of flow for given viscoplastic conditions, and stagnant and plug zone configurations are
Plug and stagnant zones substantially altered.
Flow enhancement © 2017 Elsevier Inc. All rights reserved.

1. Introduction

Viscoplastic fluids flowing in straight tubes of arbitrary cross-sectional geometry are likely to develop plug zones and
stagnant zones attached to the corners. These flow structures lead to extra resistance which requires additional energy input
to drive the flow. This resistance can be determined once the rate of flow is computed and related to the pressure gradient.
Flow blockage may occur when the yield stress of the fluid reaches or exceeds a critical value for a given pressure gradient.
Many industrial and natural materials such as concrete, slurries, mud, pastes, paint and others behave as viscoplastic fluids
which, in many cases, have large yield stress and, therefore, require significant amounts of energy for mass transport. The
constitutive characteristics of viscoplastic fluids spawn complex structures of the velocity and shear fields in tube flows
with non-circular cross-sections even in the case of the Bingham model perhaps the simplest mathematical description of
this type of fluid. Plasticity implies existence of yield stress, which may induce both plug zones and stagnant zones in the
tube cross-section depending on the tube geometry. Such solid regions require applying greater pressure gradients to keep
a given rate of flow which, in turn, leads to greater energy dissipation and heating in the tube.
Many industrial fluids exhibit yield stress, and are found in areas such as mining (slurries), food (pastes), construction
(concrete and mud) and cosmetics among others. Flow description in tubes and other configurations already considerably


Corresponding author at: Centro de Investigación en Creatividad y Educación Superior and Departamento de Ingeniería Mechánica Universidad de
Santiago de Chile Santiago, Chile and concurrently at Department of Mathematics and Statistical Sciences and Department of Mechanical, Energy and
Industrial Engineering Bostwana International University of Science and Technology, Palapye, Bostwana.
E-mail addresses: dennis.siginer@usach.cl, siginerd@biust.ac.bw (D.A. Siginer).

http://dx.doi.org/10.1016/j.apm.2017.01.058
0307-904X/© 2017 Elsevier Inc. All rights reserved.
M.F. Letelier et al. / Applied Mathematical Modelling 46 (2017) 572–580 573

difficult due to the non-linear constitutive structure of the fluid is compounded by the geometry except for very simple
shapes. Some recent research in this field include flow around a cylinder, Tokpavi et al. [1], particle sedimentation, Yu
and Wachs [2], flow in an eccentric annular tube, Wachs [3] and Walton and Bittleston, [4], a general analysis for flow
in non-circular ducts, Letelier and Siginer [5], and numerical studies on the plug zone geometry, Saramito and Roquet [6].
Viscoelastic fluids are encountered in diverse areas an unusual example of which is ophthalmology, Wilkie and Willis [7]. A
very relevant example of elasto-viscoplastic material behavior is found in tailing disposal, Nguyen and Boger [8]. Both costs
and environmental impact can be reduced substantially through the correct modeling of the elasto-viscoplastic behavior in
tailing disposal the study of which becomes thus particularly relevant.
This paper presents an analytical method to describe the elasto-viscoplastic flow through tubes of non-circular cross-
section. To characterize the behavior of the fluid we adopt a linear combination of the simplified Phan–Thien Tanner (SPTT)
model, Siginer [9,10], and the Bingham model to account for the effects of viscoelasticity and viscoplasticity, respectively.
The geometry of the non-circular cross-section is described by the shape factor method a one-to-one mapping, which maps
a circular base contour onto families of contours characterized by two mapping parameters, α and ε , Siginer and Letelier
[11,12]. The field variables are expanded in asymptotic series in the Weissenberg number We leading to a set of hierarchical
equations to be solved successively. In this paper we only develop the primary longitudinal field to study the effect of
inserting elasticity in the structure of an essentially visco-plastic fluid to determine the change in the flow rate as well
as in the flow configuration. We show for the triangular and approximately square cross-sections the plug zones (PZ) and
compute the rate of flow for different combinations of viscoelastic and viscoplastic parameters. The primary longitudinal
field is the solution of the first and second orders in the Weissenberg number We. At the first two orders the secondary
field is null. A non-zero secondary field develops for the first time at the third order O(We3 ). The solution for the secondary
field, which presents a different set of mathematical challenges, will be developed in a separate paper.

2. Elasto-viscoplastic model and analysis

To describe the behavior of the elasto-viscoplastic fluid a superposition of the Bingham and SPTT models is formulated.
The Bingham constitutive equation and the linear momentum balance in dimensionless variables and in polar coordinates
read as:
 2  2  12
∂w 1 ∂w
I= + , (1)
∂r r ∂θ
 
N ∂w
 
N 1 ∂w
τrz = − 1 + , τθ z = − 1 + , τrz ≥ N, τθ z ≥ N, (2)
I ∂r I r ∂θ
∂ τrz τrz 1 ∂ τθ z ∂P
+ + =− . (3)
∂r r r ∂θ ∂z
Here (r, θ , z) are cylindrical coordinates, w is the axial velocity, τ rz and τ θ z are the coordinate-oriented axial shear-
stresses, N is the dimensionless yield stress, and P is the piezometric pressure. The radius a of the base contour (circu-
lar tube), the average velocity w0 , ηw0 /a (η is the viscosity of the fluid), and a2 /ηw0 are introduced as scale factors to
non-dimensionalize the radial coordinate r, the velocity v, the shear stress τ and the pressure gradient, respectively. The
dimensionless yield stress N, the Bingham number, is defined in terms of its dimensional counterpart τ y ,
a
N= τy (4)
η w0
The SPPT model which describes the effect of viscoelasticity in dimensionless variables is given by,
gτ + W e(v · ∇τ − Lτ − τ LT ) = −2D, (5)

g = 1 + We 0 tr (τ ). (6)
We represents the Weissenberg number, a dimensionless measure of the elasticity of the fluid, 0 is a material parameter
representing elongational properties of the fluid and D, τ and L = ∇vT stand for the deformation rate, extra-stress and the
transpose of the velocity gradient tensors. Now consider a steady elasto-viscoplastic fluid flow with constant fluid properties
in a straight tube with constant cross-section. For the description of this flow behavior a frame indifferent combination
of the two previous models obtained by substituting (1 + NI )2D for the deformation rate tensor D in (5) resulting in the
formulation (7) is suggested:
 N

gτ + W e (v · ∇τ − Lτ − τ LT ) = − 1 + 2D. (7)
I
The constitutive and momentum equations herein proposed to describe the behavior of the elasto-viscoplastic fluid are
(1), (3), (4), (6) and (7); (2) and (5) is replaced by (7). For We = 0 (7) reduces to (2) and for N = 0 (5) is recovered. The tube
contour is described by the one-to-one mapping G, also called shape factor, defined as:
G = 1 − r 2 + ε r α sin αθ , (8)
574 M.F. Letelier et al. / Applied Mathematical Modelling 46 (2017) 572–580

where ε < 1 is a mapping parameter that will also serve as the perturbation parameter in the hierarchical algorithm to
be described below. The parameter α is an integer number for closed cross-sectional contours. Combinations of ε and α
yield a large spectrum of contour shapes for G = 0. This shape factor can always be made more general by adding more
perturbation terms to generate almost any mapped tube contour desired, Siginer and Letelier [11,12]. Nonetheless, (8) is
suitable for modeling general symmetric cross-sections and is enough for the purpose of this paper.
It should be noted that the mapping (also perturbation) parameter ε has maximum allowable values ε c for a given value
of α , Siginer and Letelier [11,12],
2 α − 2 (α−2)/2
εc = ( ) . (9)
α α
The mathematical structure of the longitudinal velocity field w implies that the tube boundary can be specified in an
infinite number of different forms. The range of the spectrum of mapped closed tube contours may be substantially extended
if a finite number of terms of the form ε i rα i sin α i θ , ε i rα i cos α i θ are incorporated in the shape factor G defined in (8). That
makes it possible to describe very complex shapes depending on the number of terms included. Shape factor G describes
exact ellipses for α = 2 and any ε , an exact equilateral triangle for the limiting case of ε = ε c = 0.3849, and regular figures of
number of sides equal to α in which sides are slightly curved when ε = ε c as in (9) if only one term is considered in (8). In
the present analysis two families of shapes are considered; shapes that evolve from a circle to an equilateral triangle (α = 3,
0 ≤ ε ≤ ε c = 0.3849) and shapes that evolve from a circle to a square (α = 4, 0 ≤ ε ≤ ε c = 0.25) with slightly curved sides
when ε = ε c . These tube contours were selected because they are the contours with the largest departure from the circle,
the base contour, and thus are more likely to bring out the dynamic characteristics of elasto-viscoplastic fluid flows.
The velocity is expressed as,
 
w = G f0 + ε f1 + O ε2 . (9)
Functions fi are unknowns to be determined. The structure of (9) satisfies the no-slip boundary condition on the contour
of the tube. We also define the following asymptotic expansion in terms of the Weissenberg number We,

w = w0 + We w1 + W e2 w2 + O W e3 , (10)
where the zero superscript indicates the purely plastic solution:
0 0 0
w 0 = w 0 + ε w 1 + ε 2 w 2 + O ( ε 3 ). (11)

2.1. Velocity field w0 of the purely plastic Bingham fluid flow up to and including O(ε )

0
The leading term w0 in (11) is the well-known velocity field of the Bingham fluid in round tubes. The structure of (11)
ensures that asymptotic conditions for ε = 0 and for N = 0 are met. Substituting (11) into (2) and using (1) yields,
⎡ ⎤
0
0 0  0 2
dw0 ∂ w1 ⎢ ∂w N ∂ w1 ⎥ 2 
τrz = N − − ε−⎣ 2 +  2 ⎦ε + O ε 3 , (12)
dr ∂r ∂r 0 ∂θ
2r 2 dw0 /dr
⎡ ⎤
 0  ∂ w 1
0  0  0  0
1 dw0 ∂θ ⎢ N ∂ w1 ∂ w1 2 ∂ w2 ⎥ 2
τθ z = ε−⎣    ε + O ( ε 3 ).
∂θ ⎦
N− 2 − (13)
r dr dw0
0
0 ∂θ ∂r 0
dw0 /dr
dr r dw0 /dr

Substituting (12) and (13) into the momentum balance (3) yields at zeroth order the well-known Bingham formula for
flow in round tubes, i.e.:
0
w0 = 1 − r 2 − N (1 − r ) (14)
Defining the velocity w0 in the same spirit as (9):
 
0
w 0 = G f0 + ε f10 + O( ε 2 ) . (15)

0 0
Equating (15) and (11) and matching powers of ε the functions f0 and f1 are completely determined after imposing
0 0
the condition that f0 and f1 are continuous at r = 1; thus we find:

0 N
f0 =1− (16)
1+r
The dimensionless pressure gradient is set as (−4). Substitution of (12), (13), (14) and (16) in the momentum balance (3)
0
yields the following equation for w1 in the asymptotic expansion (11),
 N
  ∂ 2 w 0 1 ∂ w 0  ∂ 2 w 0
1 1 1
r r− + + =0 (17)
2 ∂ r2 r ∂r ∂θ2
M.F. Letelier et al. / Applied Mathematical Modelling 46 (2017) 572–580 575

(17) is solved by setting,


0
w1 = Rα (r ) sin α θ,
where α is the mapping parameter in (8). We will develop the solution of (17) for α = 3 and α = 4 corresponding to the
equilateral triangle and the square, respectively. A change of variable r̄ = 2Nr transforms (17) into (18), and leads to the
hypergeometric differential equation,
d2 Rα (r̄ ) dRα (r̄ )
r̄ (1 − r̄ ) + (1 − r̄ ) + α 2 Rα (r̄ ) = 0. (18)
dr̄ 2 dr̄
The physical solution to (18), well behaved at the origin, is the hypergeometric function 2 F1 (a1 , a2 ; b1 ; r̄ ), also known
as Gauss’s hypergeometric function. It comes out of the generalized hypergeometric function p Fq (a1 , . . . ., a p ; b1 , . . . ., bq ; r̄ )
by setting p = 2 and q = 1. Thus, the solution of (18) reads,
2−N
 2

Rα (r̄ ) = F −1 a, b; c; 2 F1 (a, b; c; r̄ )
2 2 1 N
a = −α , b = α , c = 1.
In the expression for Rα (r̄ ), the overall multiplicative constant has been fixed by comparing (15) and (11) and imposing
0
the condition that f1 be continuous at r = 1. 2 F1 (−α , α ; 1; r̄ ) is reduced to polynomials for positive, integer values of α .
Evaluating the solution for α = 3 and α = 4 we obtain R3 and R4 , corresponding to the equilateral triangle and the square,
respectively,
N3 9 2 9
R3 ( r ) = − + N r− Nr2 + r3 , (19)
80 40 10

R4 (r ) = 0.00179 N 4 − 0.05714 N 3 r + 0.42857N 2 r 2 − 1.14286N r 3 + r 4 (20)


0
Comparison of (15) and (11) yields f1 ,

0 (2 − N )(1 + r )Rα (r ) − 2Rα (1 )(1 + r − N )r3


f1 =  sin 3θ . (21)
2Rα ( 1 ) ( 1 + r ) 1 − r 2
Then (15) together with (8), (16) and (21) determine the velocity field of the purely plastic fluid in an arbitrary cross-
sectional straight tube which symmetries, the number of sides in this case, are fixed by the mapping parameter α ,
 
N (2 − N )(1 + r )Rα (r ) − 2Rα (1 )(1 + r − N )r3 
w 0 = G 1 − +ε  sin 3θ + O ε 2
. (22)
1+r 2Rα ( 1 ) ( 1 + r ) 1 − r 2

For instance α = 3 would correspond to an equilateral triangle, and the velocity field w0 for the equilateral triangular
cross-section up to and including O(ε ) reads,
 
 N (2 − N )(1 + r )R3 (r ) − 2R3 (1 )(1 + r − N )r3 
w0 = 1 − r 2 + ε r 3 sin αθ 1− +ε  sin 3θ + O ε 2
,
1+r 2R3 ( 1 ) ( 1 + r ) 1 − r 2

where R3 (r) is given by (19). The non-vanishing components of the stress-tensor for the purely plastic fluid are then com-
puted using (12) and (13),
 
0 9N 2 9N
τzr = 2r − ε − r + 3r 2 sin 3θ + O( ε 2 ), (23)
40 5
 
6ε N3 9N 2 9N 2
τzθ0 = − + r− r + r3 cos 3θ + O( ε 2 ). (24)
N − 2r 80 40 10
0
The second order velocity w2 can be computed from the partial differential equation of second order derived in a
similar way to (17),
 N
  ∂ 2 w 0 1 ∂ w 0  ∂ 2 w 0
r r− 2
+ 2
+ 2
= K (r, θ ) (25)
2 ∂ r2 r ∂r ∂θ2
0 0 0
where K(r, θ ) is a complex function that involves terms in w0 and w1 . Eq. (25) can be solved analytically and hence f2
0 0 0
can be found in a similar fashion to f0 and f1 . However computing f2 does not change results qualitatively as compared
0 0 0
to the case where only f0 and f1 are considered. Analytical expressions for fi , i = 3, . . . , M can be determined in a
similar way but make only increasingly small quantitative contributions to the values of the field variables, and do not
change at all the qualitative outcomes as is implied by the asymptotic nature of the expansions.
576 M.F. Letelier et al. / Applied Mathematical Modelling 46 (2017) 572–580

3. Elasto-viscoplastic fluid velocity field w up to and including the first significant order O(We2 ,ε)

To determine the velocity field w for the elasto-viscoplastic flow, the momentum Eq. (3) is solved next at different orders
in We and ε by writing the components of the shear-stress tensor τ and the pressure P as asymptotic expansions in the
Weissenberg number We as:

= ( 0 ) + W e ( 1 ) + W e2 ( 2 ) + O ( W e3 ), (26)
where = {P, I, τ , w}, leading to a set of hierarchical equations to be solved. Likewise, (10) is used to introduce the depen-
dence of τ and I on the velocity w through the constitutive relations. In the following P(0) is given a value of (−4), while
P(i) = 0 ∀ i > 0.

3.1. Solution at O(We, ε )

As tr(τ (0) ) = 0, it follows from (6) that g = 1 up to O(We). The inverse of the invariant I (1) is expanded up to this order
as:
 
∂ w10
1
I−1 = − 1 − W e ∂ r0 + O(W e2 ) (27)
∂ w00 ∂ w0
∂r ∂r
If α = 3, (27) takes the following form where (11)–(13) and (23) and (24) have been used,
   
−1 9N 2 9N ∂ w 1 
I −1
= (N − 2r ) − ε sin (3θ ) − r + 3r 2 − We + O W e2 . (28)
( N − 2r ) 2 40 5 ∂r
Using the proposed elasto-viscoplastic constitutive Eqs. (7) and (28) the stress-tensor components for the elasto-
viscoplastic fluid flowing in the equilateral triangular cross-sectional tube at O(We, ε ) are determined as,
 2 
1 9N 9N
τzz = 4r ( N − 2r ) + 2ε ( 4r − N ) − r + 3r sin 3θ
2
40 5
∂ w 1 ∂ w 1  2

τzr1 = − , τzθ1 = ,
∂r ∂θ N − 2r
τθθ1 = τrr1 = τrθ1 = 0. (29)
The momentum Eq. (3) at O(We) takes the form,
1
1 ∂τzθ τzr1 ∂τzr1
+ + = 0. (30)
r ∂θ r ∂r
Substituting the expressions for the stress components into (30), the following equation for the velocity field is obtained
at O(We):

∂ 2 w 1 1 ∂ w 1 1  2  ∂ 2 w 1


+ + = 0. (31)
∂ r2 r ∂r r 2r − N ∂θ2
Since no source-like term is present in (31) and w1 must satisfy the no-slip boundary condition and consequently
vanishes at the tube boundary, the solution of (31) is w1 = 0. There is no contribution linear in the Weissenberg number
We to the velocity field of the elasto-viscoplastic fluid.

3.2. Solution at O(We2 , ε )

At this order for α = 3 using (6) we find,


 2 
9N 9N
g = 0 W e 4r (N − 2r ) + 2ε sin (3θ )(4r − N )
2
− 2
r + 3r , (32)
40 5

and the invariant I − 1 is expanded up to order O(We2 ),


   
9N 2 9N ∂ w2
I−1 = −(N − 2r )−2 (N − 2r ) − ε sin (3θ ) − r + 3r 2 − W e2 (33)
40 5 ∂r
The extra-stress tensor components at this order in terms of the velocity are then computed using the proposed consti-
tutive Eq. (7), the asymptotic expansions (10) and (26) as well as (32) and (33),

∂ w 2
τzr2 = − − 80 r 2 (N − 2r ),
∂r
M.F. Letelier et al. / Applied Mathematical Modelling 46 (2017) 572–580 577

∂ w 2  2 
τθz2 = 2 2 2 2
, τzz = τθ θ = τrr = τθ r = 0. (34)
∂θ N − 2r
Substituting in the momentum Eq. (3), the following non-homogeneous equation is obtained at order O(We2 , ε ):
1
 2
 ∂ 2 w 2 1 ∂ w 2 ∂ 2 w 2
+ + = 80 r [N − 4r]. (35)
r 2r − N ∂θ 2 r ∂r ∂ r2
The partial differential Eq. (35) is structurally similar to (17) except for the non-homogeneous source term independent
2 2
of θ . Then, the velocity is obtained by seeking a solution of the form w2 = w0 + ε w1 or equivalently:
2
w2 = w0 + ε KR(r ) sin (3θ ) + C0 , (36)
2 2
9 0 N r − 20
8 3
where R(r) satisfies (18) and w0 = r4 . Rewriting (36) in terms of the shape factor (8) we introduce f0 and
2
f1 to be computed for each cross-section,
 
2 2
w 2 = G f 0 + ε f1 . (37)

2 2
For the triangular cross-section (α = 3) the functions f0 and f1 are computed as,
 
2 − 89 0 N 1 + r + r + 20 (r + 1 ) r 2 + 1
2
f0 = ,
(1 + r )
4 0 (3−N ) R3 (r )
2 3 R3 ( 1 )
− r 3 g0
f1 = sin (3θ )
1 − r2
2 2
and for the square cross-section (α = 4) f0 and f1 are found to be,
 
< 2>
− 89 0 N 1 + r + r + 20 (r + 1 ) r 2 + 12
f0 = ,
(1 + r )
    
sin (4θ ) 4 (3 − N )R4 (r ) − 89 N 1 + r + r 2 + 2(r + 1 ) r 2 + 1
f1<2> = − r4
1 − r2 3R4 ( 1 ) 1+r

The functions R3 (r) and R4 (r) are given by (19) and (20), respectively. In the previous solutions both constants of integra-
tion C0 and K are fixed by requiring continuity at r = 1, similar to constructing w0 .
In this fashion, the axial velocity field for the elasto-viscoplastic flow is computed up to O(We2 , ε ), which in the form
of (10) is given as:
w = w 0 + W e 2 w 2 ,
where the perturbations around the circular solutions have been carried out in terms of the mapping parameter ε as w<i> =
i i i i
w0 + ε w1 + O( ε 2 ) = G( f0 + ε f1 ), i = 0, 2. This combined scheme shows that the interplay of all three ingredients,
namely elasticity, plasticity and geometry, appears at the order O(We2 ), while w0 only encapsulates the last two.

4. Results and discussion

Isovels (equal velocity lines) for different combinations of the viscoplastic and viscoelastic properties of the fluid are
shown for the triangular (α = 3) and square (α = 4) cross-sections. Yield surfaces are identified using the standard criterion
introduced with the Bingham model given by Eq. (2). Regions where the shear stresses fall below the yield stress, i.e. τ rz ≤
N and τ θ z ≤ N, are identified as plug zones where material particles move all together with a uniform velocity as a solid.
The curve of maximum axial velocity, the uniform velocity of the plug, determines the yield surface, the surface of the plug.
Since the fluid abides by the elasto-viscoplastic constitutive model (7), and the flow problem is solved using a perturbation
approach in the elastic parameter We, higher order terms in We could in principle distort the discrimination of yielded from
unyielded regions using the criteria τ rz ≤ N and τ θ z ≤ N, but given that the elastic contributions are successively smaller,
possible variations on the computed plug velocity (maximum isovel) are expected to be no greater than We3 ∼ 0.1 based
on Wemax ∼ 0.5. In the following figures the PZ are represented as shaded regions.
In Figs. 1 and 3 the dimensionless yield stress takes the constant value N = 0.5, while Figs. 2 and 4 correspond to N = 1.
Figs. 1–4 show that by incorporating viscoelasticity flow velocity is increased with respect to the purely viscoplastic case,
thus enhancing the rate of flow Q, independently of the plasticity and cross-sectional shape of the tube. Figs. 1–4 also show
that the PZ is modified in shape and area by the viscoelasticity embedded in the constitutive structure. For fixed 0 and
We the PZ area increases with plasticity in both geometries triangular and square, in agreement with the standard behavior
of a plastic fluid. The precise effect of the viscoelasticity on the PZ area depends on the value of the yield stress N. For a
578 M.F. Letelier et al. / Applied Mathematical Modelling 46 (2017) 572–580

a b c

Fig. 1. Isovels for α = 3, ε = 0.3849, N = 0.5. Velocity values on the isovels from border to center in each case: (a) 0, 0.015, 0.3, 0.45, 0.562; (b) 0, 0.015, 0.3,
0.45, 0.587; (c) 0, 0.25, 0.50, 0.757. Shaded regions represent plug zones.

a b c

Fig. 2. Isovels for α = 3, ε = 0.3849, N = 1.0. Velocity values on the isovels from border to center in each case are: (a) 0, 0.075, 0.15, 0.225, 0.249; (b) 0,
0.01, 0.2, 0.266; (c) 0, 0.13, 0.26, 0.386. Shaded regions represent plug zones.

a b c

Fig. 3. Isovels for α = 4, ε = 0.204, N = 0.5. Velocity values on the isovels from border to center in each case are: (a) 0, 0.15, 0.3, 0.45, 0.562; (b) 0, 0.15,
0.3, 0.45, 0.587; (c) 0, 0.25, 0.50, 0.75. Shaded regions represent plug zones.

purely plastic flow with N = 1 in the triangular cross-section shown in Fig. 2(a) the PZ area is 0.931, whereas in the situation
depicted in Fig. 2(b) the area is 0.870 for 0 = We = 0.5, accounting for a 6.6% drop. The converse situation is observed for
a low plasticity such as N = 0.5, where the area of the PZ increases up to 3.2%. Remarkably elasticity seems to remove the
stagnant zones hallmarks of plasticity in the corners in the range of the parameters studied in particular in the triangular
cross-section.
The rate of flow is computed as a function of the elasticity measure the dimensionless Weissenberg number We for
different values of the dimensionless yield stress N for the triangular (α = 3) and square (α = 4) cross-sections in Fig. 5(a)
and (b), respectively. The rate of flow Q clearly increases with increasing elasticity for a fixed value of the yield stress. For
instance, in the triangular case shown in Fig. 1, in the case of purely plastic fluid (We = 0 = 0) with N = 0.5 the rate of flow
is Q = 1.17, whereas in the elasto-viscoplastic case this value increases up to Q = 1.69 for 0 = We = 0.5, a 44% enhancement
with respect to the previous case.
M.F. Letelier et al. / Applied Mathematical Modelling 46 (2017) 572–580 579

a b c

Fig. 4. Isovels for α = 4, ε = 0.204, N = 1.0 . Velocity values on the isovels from border to center in each case are: (a) 0, 0.075, 0.15, 0.225, 0.249; (b) 0,
0.075, 0.15, 0.225, 0.267; (c) 0, 0.086, 0.173, 0.260. Shaded regions represent plug zones.

a b

Fig. 5. Rate of flow for N = 1.65, 1.5, 1.0 and 0.5 upwards as a function of We = 0 . (a) and (b) correspond to the triangular (α = 3, ε = 0.3849) and square
(α = 4, ε = 0.204) cross-sections, respectively.

The phenomenon of flow rate enhancement due to the addition of elastic components to a base flow follows the general
behavior of shear thinning fluids. In viscoelastic fluids, shearing tends to stretch out the polymer chains, which straightens
out some segments and aligns them in the direction of shear, an orientation that minimizes resistance to shearing. The
stretching and alignment increase with shear rate, thus a polymer solution is always shear thinning. Analytical studies on the
tube flow of Phan–Thien–Tanner and FENE-P fluids with Newtonian solvents show that the flow rate increases significantly
when the concentration of viscoelastic polymer increases [13]. This effect is similar to drag reduction, also known as Toms
effect, the addition of tiny amounts of long polymer chains to organic solvents, which can lead to significant turbulent drag
reduction [14]. Drag reduction due to the presence of polymeric additives finds many practical applications, e.g. in pipeline
material transport and oil well operations. Elastic additives to a viscoplastic base fluid result in a flow rate enhancement
mechanism similar in its effect on the flow behavior to that of the drag reducing fluids with polymeric additives to a
Newtonian base fluid. The flow rate enhancement of elasto-viscoplastic fluids equivalent to reduced frictional energy losses
is relevant to many industrial applications pipeline transport and tailing disposal among them.
The present authors investigated the flow of non-affine magnetorheological visco-elastoplastic fluids with yield stress in
round tubes characterized by a similar constitutive equation [15]. Their findings also predict the enhancement of the flow
rate and the deformation of unyielded regions by means of an exact, analytical approach. The behavior reported in this work
is consistent with the results in the literature, where it is shown for example that the inclusion of a yield stress in the fluid
dampens the mobility of the material, decreasing elastic phenomena such as vortex activity [16].

5. Conclusions

The effects of superimposing elasticity on a viscoplastic fluid are analyzed for steady flow in tubes of non-circular cross-
sections. The cross-sectional geometry is determined by the shape factor method, which maps the circular base contour onto
a family of shapes characterized by two mapping parameters, α and ε . To predict the behavior of the elasto-viscoplastic fluid
a linear combination of the simplified Phan–Thien–Tanner model and the Bingham model is constructed. The effect of the
elasticity is analyzed, and the evolution of the plug zones and rate of flow is determined in terms of the Weissenberg
580 M.F. Letelier et al. / Applied Mathematical Modelling 46 (2017) 572–580

number We, a material parameter of the SPTT model 0 , and the Bingham number N (the dimensionless yield stress). We
find that elasticity tends to enhance the rate of flow for given viscoplastic conditions.
The inclusion of even a small amount of viscoelastic component to the plastic fluid increases the velocity and modifies
the size of the plug and stagnant zones, and consequently increases the flow rate, supporting previous findings suggesting
that even a small viscoelastic contribution to the constitutive structure may enhance the flow velocity [13,15]. In particular
in the range of the parameters studied in this paper elasticity has the effect of removing the stagnant zones. Then, clearly
energy savings in transporting the fluid can be achieved by incorporating some degree of elasticity into the structure of the
viscoplastic fluid. This effect may be relevant in several flow situations, such as the transport of tailing, concrete, slurries,
mud, pastes, paint and other industrial materials.
We note that secondary flows of the elasto-viscoplastic fluid can be studied using the approach introduced in this paper
through the investigation of the third order problem O(We3 ) in the hierarchical perturbative scheme. The velocity field
herein computed can also be used to study the temperature distribution in this type of flow with different thermal boundary
conditions. Both of these problems spawn their own mathematical challenges and will be dealt with in separate publications.

Acknowledgments

The authors gratefully acknowledge the financial support of DICYT at USACH and of FONDECYT, through Grant 1130346.
The authors are also grateful to Ercio Báez and Nicolás Díaz for their help with the computations.

References

[1] D.L. Tokpavi, A. Magnin, P. Jay, Very slow flow of Bingham viscoplastic fluid around a circular cylinder, J. Non Newton. Fluid Mech. 154 (2008) 65–76.
[2] Z. Yu, A. Wachs, A fictitious domain method for dynamic simulation of particle sedimentation in Bingham fluids, J. Non Newton. Fluid Mech. 145
(2007) 78–91.
[3] A. Wachs, Numerical simulation of steady Bingham flow through an eccentric annular cross-section by distributed Lagrange multiplier fictitious domain
and augmented Lagrangian methods, J. Non Newton. Fluid Mech. 142 (2007) 183–198.
[4] I.C. Walton, S.H. Bittleston, The axial flow a Bingham plastic in a narrow eccentric annulus, J. Fluid Mech. 222 (1991) 39–60.
[5] M.F. Letelier, D.A. Siginer, On the flow of a class of viscoinelastic-viscoplastic fluids in tubes of non-circular contour, Int. J. Eng. Sci. 45 (2007) 873–881.
[6] P. Saramito, N. Roquet, An adaptive finite element method for viscoplastic fluid flows in pipes, Comput. Methods Appl. Mech. Eng. 190 (2001)
5391–5412.
[7] D.A. Wilkie, A.M. Willis, Viscoelastic materials in veterinary ophthalmology, Vet. Ophthalmol. 2 (1999) 147–153.
[8] Q.D. Nguyen, D.V. Boger, Application of rheology to solving tailing disposal problems, Int. J. Miner. Process. 54 (1998) 217–233.
[9] D.A. Siginer, Developments in Tube Flow of Complex Fluids, Springer Inc., New York, 2015.
[10] D.A. Siginer, Isothermal tube flow of non-linear viscoelastic fluids, part ii: transversal field, Int. J. Eng. Sci. 49 (6) (2011) 443–465.
[11] D.A. Siginer, M.F. Letelier, Laminar flow of non-linear viscoelastic fluids in straight tubes of arbitrary contour, Int. J. Heat Mass Transf. 54 (2011)
2188–2202.
[12] D.A. Siginer, M.F. Letelier, Heat transfer asymptote in laminar flow of non-linear viscoelastic fluids in straight non-circular tubes, Int. J. Eng. Sci. 48
(2010) 1544–1562.
[13] D.O.A. Cruz, F.T. Pinho, P.J. Oliveira, Analytical solutions for fully developed laminar flow of some viscoelastic liquids with a Newtonian solvent contri-
bution, J. Non Newton. Fluid Mech. 132 (2005) 28–35.
[14] A. Toms B., Some observations on the flow of linear polymer solutions through straight tubes at large Reynolds numbers, in: Proceedings of the First
International Congress of Rheology, 2, 1949, pp. 135–141.
[15] M.F. Letelier, D.A. Siginer, On the steady flow of magnetorheological fluids in pipes, Int. J. Appl. Mech. Eng. 10 (3) (2005) 463–474.
[16] J.E. López-Aguilar, M.F. Webster, H.R. Tamaddon-Jahromi, et al., Numerical modelling of thixotropic and viscoelastoplastic materials in complex flow,
Rheol. Acta 54 (2010) 307–325.

You might also like