J. Fluid Mech. (1999), vol. 401, pp. 339–364.
c 1999 Cambridge University Press
Printed in the United Kingdom
339
Steady flow in a dividing pipe
By M. G. B L Y T H
AND
A. J. M E S T E L
Mathematics Department, Imperial College, 180 Queen’s Gate, London SW7 2BZ, UK
(Received 11 November 1998 and in revised form 17 May 1999)
The high Reynolds number flow through a circular pipe divided along a diameter
by a semi-infinite splitter plate is considered. Matched asymptotic expansions are
used to analyse the developing flow, which is decomposed into four regions: a
boundary layer of Blasius type growing along the plate, an inviscid core, a viscous
layer close to the curved wall and a nonlinear corner region. The core solution
is found numerically, initially in the long-distance down-pipe limit and thereafter
the full problem is solved using down-pipe Fourier transforms. The accuracy in the
corners of the semicircular cross-section is improved by subtracting out the singularity
in the velocity perturbation. The linear viscous wall layer is solved analytically in
terms of a displacement function determined from the core. A plausible structure
for the corner region and equations governing the motion therein are presented
although no solution is attempted. The presence of the plate has little effect ahead
of the bifurcation, but wall shear on the curved wall is found to increase from its
undisturbed value downstream.
1. Introduction
We reconsider Smith’s (1977a) problem of the flow of a homogeneous incompressible Newtonian fluid along an infinite straight pipe of circular cross-section
which is divided at some point by a semi-infinite flat plate across its diameter (see
figure 1). Far upstream of the division we assume the motion to be steady and
fully-developed. The main aim of this paper is to examine the entry flow problem
when this uni-directional profile impinges on the divider. A long way downstream the
flow will return to a Poiseuille motion, this time in two semicircular pipes. It is the
intermediate development that is of concern here.
The large number of industrial and biological applications has stimulated much
study of pipe flows. In this paper we will concern ourselves chiefly with the problem’s
physiological relevance. A review of the fluid mechanical aspects of the arterial system
is given in Pedley (1980). In general, haemodynamic flows are rendered difficult to
model by multifarious complicating factors. The problem of a bifurcating artery
seems to have been originally addressed analytically by Zamir & Roach (1973).
The effects of unsteadiness at pipe bifurcations have been considered numerically,
for example in the computer simulations of Friedman, O’Brien & Ehrlich (1975)
and Wille (1984), and experimentally, for example by Rieu & Pelissier (1991) who
measured unsteady flow in a dividing pipe for different angles of bifurcation. The
assumption of steady flow is perhaps the most significant simplification made in the
present work, limiting its physiological applicability. However it may act as a first
step in the model of a bifurcating artery on which future research could be based.
Any reservations about its applicability notwithstanding, it is an interesting problem
worthy of study in its own right. Nevertheless we adopt a biological standpoint
340
M. G. Blyth and A. J. Mestel
y
x
z
U0
Figure 1. Circular pipe with splitter-plate for x > 0.
and in this spirit results are presented (see § § 3, 4) which would be of physiological
interest but which we concede may be far removed from their likely counterparts in
vivo.
Caro, Fitz-Gerald & Schroter (1971) observed that the location on the artery walls
of fatty streaks, which are believed to lead to the onset of atherosclerosis, may be
influenced by certain haemodynamic factors since they are particularly found in regions associated with low wall shear. Preferential sites include the outer walls of
dividing arteries and the inside walls of bends. Here we discuss a simple model
of a bifurcating artery where the angle of bifurcation is taken to be zero. In reality, of course, branching arteries in the body will diverge with daughter tubes
of generally different cross-sectional areas. Many nearly symmetric divisions occur in the airways in the lungs but the only close approximation to a symmetric
branching in the systemic circulation occurs where the aorta splits into the iliac
arteries (Pedley 1980). For simplicity we assume that the branching is symmetric,
although the analysis is easily adaptable to allow for an asymmetric splitter plate.
This is the most straightforward way of allowing for different pressures in the daughter tubes a long way downstream. A more difficult extension would be to admit
different pressures with the geometrical symmetry maintained, in which case the
uni-directionality of the leading order core flow would be lost. The assumption of a
fully-developed upstream flow is also not necessarily valid. Given the many consecutive bifurcations in the arterial system, it may never be allowed the necessary room
to evolve.
Our steady motion occurs in a pipe which is taken to be perfectly straight and
to suffer no bending or torsion. Curvature effects, and those of unsteadiness, have
been considered analytically, notably by Lyne (1970) and Smith (1975), who studied
pulsatile flow in a curved pipe driven by a sinusoidal pressure gradient. In reality
more complicated three-dimensional geometrical effects may also be important and
recently, for example, the effect of torsion has been taken into account by Zabielski
& Mestel (1998a,b), who examined both steady and unsteady flow in a helically
symmetric pipe. Zabielski & Mestel also included a realistic physiological pressure
pulse. Gammack & Hydon (1999) considered pipes with curvature and torsion varying
slowly with the downstream direction.
Arteries are elastic in nature and are to some extent free to move. Some of the
effects of the changing curvature of the coronary arteries during systole and diastole
(the two-part cycle of a heart beat) have been given a preliminary treatment by
Lynch, Waters & Pedley (1996) who investigated unsteady flow in pipes with time-
Steady flow in a dividing pipe
341
dependent curvature. Distensibility is important in the propagation of the pressure
pulse generated by the heart. Here we consider a rigid pipe of infinite length and to
a first approximation believe that distensibility will be of little importance especially
given the many other simplifications adopted. In this respect we follow the example
of other similar studies of pipe flows that have been motivated at least in part by
physiological considerations. This neglect may not be justified for joining veins, where
blood pressure is in general much lower than in arteries, and which are subsequently
more prone to collapse. Clearly in this eventuality elastic properties of the walls may
not be ignored.
The blood rheology is also to be considered. Blood effectively consists of a suspension of red blood cells in a plasma solution. On the lengthscale of a typical arterial
diameter the assumption of homogeneity is justified. Experimental measurements (for
example Whitmore 1968) suggest that blood behaves as a Newtonian fluid although
this might be complicated once unsteadiness is incorporated into the model. The
high Reynolds number limit adopted in the current work may also be supported
by experimental observations. Pedley (1980) reports a peak value of 1000 for the
Reynolds number in a canine femoral artery.
In order to examine the entry flow in a bifurcation, the developing viscous layers and
the three-dimensional dynamical effects of the branching, Smith (1977a) investigated
the steady flow along an infinitely long straight circular pipe with a semi-infinite
splitter plate sliced through the middle. He studied the flow with the down-pipe
distance x scaled with Reynolds number which renders the equations of motion
parabolic and of boundary layer type. Smith argued that the fluid flow may be
divided up into eight distinct regions, with a Blasius-type boundary layer growing
along the plate, a core region above it, and then an inertial and a viscous region
above this on the pipe wall. The remaining regions effect the turning of the secondary
flow in the corners. However, doubt was cast on the scalings Smith used at the
upper edge of the core by his student Bennett (1987) in his PhD thesis. In fact,
Bennett argues (as we do here) that there is a choice of two possible scalings at
the top of the core, the alternative set requiring the core pressure to vanish on
the curved wall (this would correspond to Smith’s P1 = 0). In this case, which we
now believe to be correct, there is no requirement for an intermediate inertial layer
and in fact the whole structure is simplified to just four main zones as detailed
below. Our problem is thus essentially identical to Smith’s although we do not
scale with Reynolds number in the down-pipe direction; instead we solve the flow
in the regime x = O(1) and also consider the behaviour of the core flow when
1 ≪ x ≪ R.
As usual with a boundary layer on a flat plate of infinitesimal thickness, around the
leading edge a tiny region of width O(R −1 ), in this case broadening to O(R −2/3 ) in the
corners, is needed in which the flow is described by the full Navier–Stokes equations.
This zone is unimportant as far as the rest of the flow structure is concerned and a
more detailed discussion is omitted.
The paper is arranged as follows. In § 2 we describe the asymptotic structure of
the flow and lay out the equations governing the separate regions of motion. The
nonlinear equations in the corner region are given but no attempt is made to solve
them. An analytic solution may be found for region III which is dependent on a
function prescribed by the core. The core solution is obtained numerically and is
outlined in § 3, where we consider both x = O(1) and the large x limit. Results and
a discussion of the effect of the splitter plate on the undisturbed flow are presented
in § 4.
342
M. G. Blyth and A. J. Mestel
y
III
II
r
IV
I
h
z
Figure 2. The four distinct flow zones.
2. Flow structure
We assume the steady flow to be Newtonian, incompressible and of constant
viscosity. We therefore use the three-dimensional Navier–Stokes equations
u · ∇u = −∇p + R −1 ∇2 u,
∇ · u = 0,
(2.1a)
(2.1b)
where u = (u, v, w) in the cylindrical polar coordinates (x, θ, r) as in figure 2. Lengths
have been non-dimensionalized with the pipe radius a∗ . For example x∗ = a∗ x, where
an asterisk here denotes a dimensional quantity. Also u∗ = U ∗ u, v ∗ = U ∗ v, w ∗ = U ∗ w,
p∗ = ρ∗ U ∗2 p. U ∗ is a characteristic velocity of the unperturbed flow and ρ∗ is the
density of the fluid, assumed constant. Far upstream the flow is uni-directional and is
given by u = U0 (r), v = w = 0, p = p0 − xR −1 , for constant p0 . The Reynolds number,
R, (assumed R ≫ 1) is based on the pipe radius and is defined by R = U ∗ a∗ /ν ∗ , with
ν ∗ the kinematic viscosity.
We believe that the flow structure can be divided into four main regions (illustrated
in figure 2). A short distance beyond the leading edge a boundary layer I starts to
grow along the plate. Near the centre of the tube this will be of Blasius type, based
on the local mainstream velocity. The boundary layer perturbs the core flow II above
by a finite amount. The displacement thickness of the plate layer is of order R −1/2
and we find that the perturbation to the core is then of order R −1/2 , driven by the
displacement effect of the plate layer. The equations governing the core flow are
inviscid to leading order and consequently a slip velocity is left at the edge of II on
r = 1. Thus another region, the viscous wall layer III, is needed to reduce the velocity
to zero at the curved wall. This layer is linked with the core through the function
A(θ) described below. A singularity arises in III as the corner of the semicircle is
approached and this is smoothed out by a further viscous region IV which otherwise
plays a passive role in the flow structure.
2.1. The core (II )
In the inviscid core the flow is perturbed as follows:
u = U0 (r) + R −1/2 u1 (x, r, θ) + · · · ,
v = R −1/2 v1 (x, r, θ) + · · · ,
w = R −1/2 w1 (x, r, θ) + · · · ,
p = R −1/2 p1 (x, r, θ) + R −5/6 p2 (x, r, θ) + · · · ,
(2.2a)
(2.2b)
(2.2c)
(2.2d)
Steady flow in a dividing pipe
343
where r2 = y 2 + z 2 . The coordinate frame x, y, z is shown in figure 1, with θ = 0
on y = 0, z > 0. The origin of the additional pressure perturbation p2 is discussed
in § 2.3. Far upstream the flow is given by
U0 = 41 (1 − r2 ).
(2.3)
Sufficiently far downstream the core expansions (2.2) will become invalid. From
continuity considerations (discussed in detail in § 2.1.1) we see that u1 ∼ x1/2 as
x → ∞ and so (2.2a) will break down when x ∼ R. This corresponds to the boundary
layer on the plate (I) growing large enough to fill the pipe. Therefore there exists
a down-pipe lengthscale x = RX, where X is of order unity, for which X ≪ 1 will
match to (2.2) as x → ∞, and for which the flow will develop into Poiseuille motion
in a semicircular tube when X ≫ 1. It is the former lengthscale where 1 ≪ x ≪ R
which was considered by Smith (1977a).
Substituting (2.2) into (2.1a) and taking the limit R → ∞, we derive the following
inviscid equations at leading order
(1 − r2 )u1x − 2rw1
(1 − r2 )v1x
(1 − r2 )w1x
u1x + v1θ /r + (rw1 )r /r
= −4p1x ,
= −4p1θ /r,
= −4p1r ,
= 0.
(2.4a)
(2.4b)
(2.4c)
(2.4d)
Following Smith (1976a) we can easily eliminate u1 , v1 , and w1 , and write (2.4) as a
single equation for p1
4r
p1r = 0,
(2.5)
1 − r2
where ∇22 = ∂2 /∂r2 + (1/r)∂/∂r + (1/r2 )∂2 /∂θ2 , the two-dimensional Laplacian for
the cross-section. This equation is elliptic and therefore the possibility of upstream
influence is anticipated. The core flow is driven by the Blasius layer and so this
provides the boundary condition for (2.5) on the base (θ = 0, π). We also need to
consider what happens on the curved wall to determine the boundary condition at
r = 1.
As r → 1 we can approximate (2.5) by
p1xx + ∇22 p1 +
p1nn −
2p1n
= 0,
n
(2.6)
where n = 1 − r. Then, as n → 0, two different scalings are possible for p1 . Either
p1 ∼ 1 or p1 ∼ n3 . Initially for the case x → ∞ Smith (1977a) proposed the former and
solved the core problem with two boundary conditions applied on the base. However
his student Bennett (1987), in his PhD thesis, suggested that in fact the latter choice
is correct and it is these scalings which we adopt here, as argued below. It then seems
reasonable that in the absence of some effect which is dominant for x ∼ 1 but which
is asymptotically small when x is large, p1 = 0 on r = 1 for finite x also. In light of
this we assume
u1 ∝ 1,
v1 ∝ n2 ,
w1 ∝ n,
p1 ∝ n3 ,
(2.7)
as n → 0. This then determines the second boundary condition required to solve (2.5),
p1 = 0 on r = 1.
(2.8)
344
M. G. Blyth and A. J. Mestel
2.1.1. The limit x → ∞
We use a continuity argument to show that the following system represents the
correct limiting problem in region II for large x. We will see in § 2.2 that injection
into the core is proportional to x−1/2 . If we integrate along the upper half of the pipe
from a point upstream of the leading edge of the plate to some point downstream of
it, we argue that as x → ∞ u1 will behave like x1/2 a long way downstream. From
(2.4), either p1 ∼ x1/2 or p1 ∼ x−3/2 . To determine which is correct, we consider the
first, writing u1 = u0 (r, θ)x1/2 , v1 = v0 (r, θ)x−1/2 , w1 = w0 (r, θ)x−1/2 , p1 = p0 (r, θ)x1/2 .
Then as x → ∞, the core equations (2.4) become
− r2 )u0 − 2rw0 = −2p0 ,
p0θ = p0r = 0,
1
u
+
v
+
(1/r)(rw
0θ
0 )r = 0.
2 0
1
(1
2
(2.9a)
(2.9b)
(2.9c)
If we integrate over 0 < θ < π,
1
u
2
+
1
(1 − r2 )u − 2rw
2
[v0 ]π0 + (1/r)(rw)r
= −2p0 π,
= 0,
(2.10a)
(2.10b)
where an overline indicates an averaged
√ variable here and p0 , by (2.9b), is a constant.
From (2.15), [v0 ]π0 = −β(1 − r2 )1/2 / 2. To achieve a balance in (2.10a) as r → 1 (or
n → 0, writing n = 1 − r) we need either u ∼ 1/n or w ∼ 1. Eliminating w from (2.10)
and integrating we find
√
u = −2 2β(1 − r2 )1/2 + const.
(2.11)
and so it must be the case that w ∼ 1 as n → 0 unless p0 = 0. However this conclusion
leads to an imbalance in mass transport inside the viscous wall layer. If w ∼ 1 as
n → 0 then the fluid flux into III is of order unity. The boundary layer thickness is
of order R −1/3 . By (2.11) there will be a mass flux downstream within the boundary
layer of order R −1/3 at most which cannot cater for the larger influx. Therefore we
conclude that neither w ∼ 1 nor u ∼ 1/n as n → 0 and the only option remaining to
us is to set p0 = 0. This corresponds to the constant P1 in Smith (1977a) being zero.
In this case Smith’s complicated asymptotic structure is unnecessary and we take the
boundary condition for (2.5) on r = 1 to be that of vanishing pressure.
2.2. Base boundary layer (I )
The precise nature of the base condition is determined by matching to the plate
layer I. As remarked above, the thickness of this layer is O(R −1/2 ) and so we set
y = R −1/2 Y , where Y is of order unity. A Blasius-type similarity solution is possible
for x > 0. We write
u = (1 − z 2 )f ′ (η)/4 + O(R −1/2 ),
v=R
w=
−1/2
−1/2
2 1/2 −1/2
(1 − z )
x
√
(ηf − f)/(2 2) + · · · ,
R
w1B (x, Y , z) + · · · ,
R −1/2 pB1 (x, Y , z) + · · · ,
′
(2.12a)
(2.12b)
(2.12c)
(2.12d)
√
where η = U0 (z)1/2 Y x−1/2 / 2. Dhanak & Duck (1997) demonstrate that this solution
is non-unique when the outer flow is non-zero on both walls, and that an O(1) crossflow is possible in this layer. Here, however U0 (1) = 0. At leading order z merely
behaves as a parameter and the layer is governed according to (2.12) by the familiar
p=
345
Steady flow in a dividing pipe
Blasius equation
f ′′′ + f f ′′ = 0,
(2.13)
with
f(0) = 0, f ′ (0) = 0, f ′ → 1
as η → ∞.
(2.14)
Then, since as η → ∞, f ∼ η − β + exp, where ‘exp’ denotes exponentially small terms
and β ≈ 1.21678 is the Blasius displacement,
1
1/2
v ∼ R −1/2 √ βU0 x−1/2 ,
2
(2.15)
as η → ∞, x > 0. By (2.4b), the boundary condition on the base of the core becomes
d
1
p1θ = − √ βr(1 − r2 )3/2 (H(x)x−1/2 )
dx
8 2
on θ = 0, π,
(2.16)
valid ∀x. Here H(x) is the Heaviside function. As x → ∞, this condition may be
written as
1
(2.17)
p1θ = √ βr(1 − r2 )3/2 x−3/2 on θ = 0, π.
16 2
2.3. The viscous wall layer (III )
This region is necessarily viscous in order to bring the fluid at the top of the core to
rest and so satisfy the no-slip condition on the curved wall. From (2.1a), the neglect
of viscous terms in the core becomes invalid when inertial terms balance viscous ones.
As r → 1, U0 ∼ 21 (1−r) and so viscous effects become important when (1−r) ∼ R −1/3 .
By (2.2a), the core equations would break down only when (1 − r) ∼ R −1/2 , that is
beyond where viscous effects must be taken into account. Region III is therefore
taken to be of order thickness R −1/3 and we write (1 − r) = R −1/3 ζ, with ζ a new wall
variable of order unity. Note that it is no longer necessary to include an extra layer
here as in Smith (1977a).
We then perturb the mainstream velocities in this viscous wall layer as follows
u = 12 R −1/3 ζ + R −1/2 u1 (x, ζ, θ) + · · · ,
v = R −1/2 v 1 (x, ζ, θ) + · · · ,
w = R −5/6 w 1 (x, ζ, θ) + · · · ,
p = R −5/6 p1 (x, ζ, θ) + · · · ,
(2.18a)
(2.18b)
(2.18c)
(2.18d)
according to the scalings (2.7). The relevant equations of motion are then linear:
1
ζ
2
u1x − 21 w 1 = −p1x + u1ζζ ,
1
ζ
2
v 1x = −p1θ + v 1ζζ ,
p1 = p1 (x, θ),
u1x + v 1θ − w 1ζ = 0,
(2.19a)
(2.19b)
(2.19c)
(2.19d)
with necessary matching conditions to the core
u1 → 21 A(x, θ),
and the no-slip condition
v1 → 0
as ζ → ∞,
u1 = v 1 = w 1 = 0 on ζ = 0.
(2.20)
(2.21)
346
M. G. Blyth and A. J. Mestel
The function A(x, θ) comes from the solution of the core. In § 3 we find it using
numerical methods. By (2.4) and (2.7), as r → 1 we find
1
A
2 xx
= −p1rrr |r=1 .
(2.22)
Note that the azimuthal velocity v is larger than at the top of the core. Here it is of
order R −1/2 , whereas in II, due to (2.7), v ∼ R −7/6 as r → 1, which is considerably
smaller in magnitude. However (2.20) permits a match as we leave the wall layer. The
pressure perturbation here, given by (2.18d ) and required to complete a consistent
structure in the wall layer, is larger than that at the edge of the core and so provokes
the extra perturbation p2 to II seen in (2.2d ).
The equations and boundary conditions laid out above are all valid in the range
−∞ < x < ∞ and it is convenient now to take a generalized Fourier transform in x.
We write
Z ∞
1
∗
u1 = √
u1 (x, ζ, θ)e−ikx dx,
(2.23)
2π −∞
with the other variables similarly defined. Then, transforming (2.19) we obtain the
solutions (cf. Smith 1976a, 1979; Cowley 1983)
Z t
Ai(q) dq − p∗1θθ M(t)/(2ω 5 ),
u∗1 = 23 A∗ (k, θ)
(2.24a)
0
v ∗1 = p∗1θ M(t)/ω 2 ,
Z ζ Z ωζ ′
∗
∗
3
w 1 = 2 ikA
Ai(q) dq dζ ′ ,
0
(2.24b)
(2.24c)
0
where Ai is the Airy function, and with
Z q
Z t
1
Ai(q ′ ) dq ′ dq.
M(t) = Ai(t)
2
Ai
(q)
∞
0
(2.25)
Here t = ωζ, ω = ( 21 ik)1/3 , and we have made the branch cut −3π/2 < arg(ω) 6 π/2
to ensure that Ai(t) → 0 as ζ → ∞. From (2.19), (2.20), and (2.21), p∗1 satisfies the
equation
p∗1θθ − k 2 p∗1 = 3ω 5 Ai′ (0)A∗ .
(2.26)
Assuming the flow has left/right symmetry we have the condition that
p∗1θ = 0 at θ = π/2.
(2.27)
We solve (2.26) in § 4 but first we must determine the second boundary condition.
This comes from a careful consideration of the flow structure in the corner.
2.4. The corner region (IV )
In order to detail the nature of the flow in the corner of the pipe, we need to consider
what is happening in the corner of II. Part of our aim is to determine the second
boundary condition needed to solve for the pressure in III. To this end (2.26) shows
that we will need to consider how A∗ behaves as θ → 0. In the corner of II we are
able to construct a local solution to (2.5) and we use this to examine the behaviour
of A∗ there.
As n → 0 and θ → 0 we may, after a Fourier transform, approximate (2.5) as
∗c
p∗c
1nn + p1θθ =
2p∗c
1n
,
n
(2.28)
Steady flow in a dividing pipe
347
where the neglected −k 2 p∗c
1 term is assumed negligible in comparison with the others
if we move sufficiently close to the corner. Here and throughout the rest of this paper
the superscript or subscript c denotes a corner quantity. We can transform (2.28) to
Laplace’s equation with the substitution
p∗c
1n
.
(2.29)
n
Fixing local polar coordinates (ρ, γ) in the corner, with γ = 0 corresponding to n = 0,
we have the following problem for the local solution:
ψ=
∇2c ψ = 0,
with
(2.30)
(2.31)
ψ = 0 on γ = 0,
ψγ = 3λc k 1/2 ρ1/2 on γ = π/2,
√
1
iπ/4
2
where λc = 16 βe
2, which comes from (2.16) (see also § 3.2.1), and ∇c = ∂2 /∂ρ2
+ (1/ρ)∂/∂ρ + (1/ρ2 )∂2 /∂γ 2 . This has the solution
√
(2.32)
ψ = 6 2λc k 1/2 ρ1/2 sin(γ/2).
Changing back to (n, θ) coordinates we can integrate once to give
p∗c
1 =
12
λ k 1/2 (ξ
5 c
− θ)1/2 (n2 − 31 θ(ξ − θ)),
where ξ 2 = n2 + θ2 . Now, for θ ≪ n ≪ 1
p∗c
1 =
12
λ k 1/2 n5/2
5 c
− 2λc k 1/2 n3/2 θ + O(θ2 ).
(2.33)
(2.34)
This helps to determine the nature of the crossflow in the plate layer as |z| → 1.
According to the scalings (2.12) the momentum equation in the transverse direction
for the crossflow w1B in region I is
B
B
B
+ vs w1Y
= −pB1z + w1Y
us w1x
Y,
(2.35)
where us , vs denote the similarity solutions detailed in (2.12). As Y → ∞, the leading
B
∼ −pB1z . Equation (2.34) implies the behaviour
order balance suggests that us w1x
B
3/2
B
p1z ∼ (1 − z) as z → 1. Hence w1 ∼ (1 − z)1/2 as z → 1. A similarity solution for
w1B can be found, valid as either |z| → 1 or x → ∞ (see Blyth 1999).
Returning to the solution (2.33), we also find that for n ≪ θ ≪ 1
√
2λc k 1/2 θ−1/2 n3 + O(n4 ),
(2.36)
p∗c
1 =
in which case A∗ ∼ θ−1/2 as θ → 0 (see § 3). From (2.26) we must then have p∗1θθ ∼ θ−1/2
as θ → 0, requiring that
p∗1 = α0 + α1 θ + α2 θ3/2 + O(θ2 ),
(2.37)
for small θ. As the corner must effect the turning of the fluid, pressure gradients
should be expected in both the r- and θ-directions. Furthermore, if α1 6= 0 then
by (2.24b) v ∗1 will be of order unity as the corner is approached from within the
viscous wall layer. This is too large, as a consistent match with the base boundary
layer requires the corner region to be at least as thick as the viscous wall layer. We
therefore set α1 = 0, thereby providing the required second boundary condition for
(2.26),
p∗1θ = 0 on θ = 0.
(2.38)
We believe then that IV should form a square region of size O(R −1/3 ), with region I
348
M. G. Blyth and A. J. Mestel
correspondingly thickening at each end. Accordingly we claim the following structure
for the corner when x > 0 (cf. Walton & Smith 1997):
u = R −1/3 uc1 (x, ζ, Θ) + R −1/2 uc2 (x, ζ, Θ) + · · · ,
(2.39a)
w=
(2.39c)
R −2/3 v1c (x, ζ, Θ) + · · · ,
R −2/3 w1c (x, ζ, Θ) + · · · ,
R −5/6 pc1 (x) + R −4/3 pc2 (x, ζ, Θ)
v=
p=
(2.39b)
+ ···,
(2.39d)
with (1 − r) = R −1/3 ζ, θ = R −1/3 Θ, and Θ ∼ 1. The nonlinear governing equations
are
uc1 uc1x + v1c uc1Θ − w1c uc1ζ
c
c
c
− w1c v1ζ
+ v1c v1Θ
uc1 v1x
c
c
c
+ v1c w1Θ
− w1c w1ζ
uc1 w1x
c
c
− w1ζ
uc1x + v1Θ
= uc1ζζ + uc1ΘΘ ,
c
c
+ v1ΘΘ
,
= −pc2Θ + v1ζζ
c
c
c
= p2ζ + w1ζζ + w1ΘΘ ,
= 0,
(2.40a)
(2.40b)
(2.40c)
(2.40d)
with boundary conditions
uc1 = v1c = w1c = 0 on ζ = 0, Θ = 0.
Matching to region I we have, as ζ → ∞ keeping η =
uc1 → 12 ζf ′ (η),
With
rc2
1 1/2
ζ Θx−1/2
2
(2.41)
fixed,
v1c → 12 ζ 1/2 x−1/2 (ηf ′ − f),
(2.42a)
(2.42b)
w1c → x−1/2 ζ 1/2 h(η), pc2 → g(x)ζ 5/2 .
= ζ + Θ we have, matching to region II as rc → ∞ for γ = tan−1 (ζ/Θ) 6= 0,
2
2
uc1 → 21 ζ + κ1 rc−1/2 ,
v1c → κ2 rc1/2 ,
Matching to region III we have, as Θ → ∞
uc1 → 12 ζ + κ5 Θ −1/2 ,
v1c → κ6 Θ 1/2 ,
w1c → κ3 rc1/2 ,
w1c → κ7 Θ −1/2 ,
pc2 → κ4 rc5/2 .
pc2 → κ8 Θ 3/2 .
(2.42c)
(2.42d)
The function h(η) comes from matching to the solution of (2.35) for w1B in the
plate layer, which can be found with little difficulty; κi are functions of x and γ for
i = 1, . . . , 4, and x and ζ for i = 5, . . . , 8, given in the Appendix along with g(x). The
function pc1 is present in (2.39d ) because by (4.3) p∗1 is non-zero on θ = 0, π. It is
purely a function of x since at next order in the x-momentum equation it is found
that
(2.43)
pc1ζ = pc1Θ = 0.
The next-order term in the expansion for u is required to be O(R −1/2 ) in order to
balance −pc1x at second order in the x-momentum equation. In the corners, the tiny
O(R −1 ) Navier–Stokes zone around the leading edge thickens to O(R −2/3 ) due to the
smaller local downstream velocity there. Therefore, as in Rubin (1966), the solution
in the corner would only be valid for 0 < x0 < x where x0 is unknown in advance
but which is presumed to be small.
Although it is nonlinear, and therefore interesting in itself, the corner region is
essentially passive in nature with no controlling influence over the rest of the flow.
It exists to ensure no-slip is satisfied in the corner and to effect the turning of the
secondary motion. Justification of the structure proposed above could be obtained
by a complete, presumably numerical, solution of the given equations which is not
attempted here. Such a numerical solution might proceed along the lines of Rubin &
Steady flow in a dividing pipe
349
Grossman (1971), who studied incompressible viscous flow along a right-angle corner.
Guidance could also be obtained from the simpler, linear corner problem solved by
Cowley (1983).
3. Numerical solution of the core region (II)
Here we briefly describe the numerical procedure used to solve the core equation
(2.5), and present some of our results. Two cases are considered. First we examine
the solution of the core as x → ∞. Thereafter we consider the full x problem. The
numerical procedure is similar in both cases.
3.1. Core solution for x → ∞
We showed in § 2.1.1 that p1 ∼ x−3/2 as x → ∞. We therefore write
u1
v1
w1
p1
= x1/2e
u1 + · · · ,
−1/2
e
=x
v1 + · · · ,
e1 + · · · ,
= x−1/2 w
−3/2
e
=x
p1 + · · · ,
(3.1a)
(3.1b)
(3.1c)
(3.1d)
e
in the large-x limit. Equations similar to (2.4) are then derived. A(θ),
the core
displacement function analogous to A(x, θ), is given by
1e
A
2
= 4e
p1rrr |r=1 .
(3.2)
Once more we can eliminate the other variables to obtain a single equation for e
p1 (r, θ),
∇2 e
p1 +
with boundary conditions
and
4r
e
p1r = 0,
1 − r2
(3.3)
e
p1 = 0 on r = 1,
(3.4)
e θ),
e
p1 = (1 − r2 )φ(r,
(3.6)
1
√ βr(1 − r2 )3/2 on θ = 0, π,
(3.5)
16 2
by (2.17). This is the same as Smith’s (1977a) equation (3.8a) with P1 = 0, although
Smith applied two boundary conditions on the plate. We choose to solve it numerically
using finite differences. It is expedient to make the substitution
e
p1θ = ±
whereupon (3.3) becomes
with boundary conditions
e−4
e ≡ ∇2 φ
Lφ
eθ = ±
φ
1 + r2 e
φ = 0,
(1 − r2 )2
1
√ βr(1 − r2 )1/2
16 2
on θ = 0, π,
(3.7)
(3.8)
e = 0 on r = 1. In this way the equation is simplified somewhat in the
and φ
discretization. Also we anticipate that as r → 1, e
p1 ∼ (1 − r)3 and consequently
2
e
φ ∼ (1 − r) , and the finite difference approximations are more efficient at rendering
this behaviour with the transformed equation. We use central differences and employ
a second-order-accurate derivative approximation for the condition on the base, using
350
M. G. Blyth and A. J. Mestel
‘imaginary’ values one level below the base which are eliminated by application of
the equation on θ = 0, π. Working in polar coordinates, we integrate over a small
e there
half-disc around the origin to obtain a condition for determining the value of φ
(see for example Strikwerda 1989). The discretized equation was then solved using
Gauss–Seidel point iteration with successive over-relaxation. Grids of various mesh
lengths hr , hθ in the r-, θ-directions respectively were tried. Convergence of the grid
was deemed to have been reached when the maximum difference over the grid between
iterations was less than some prescribed tolerance multiplied by the maximum value
over the grid at the present iteration. We used a tolerance of 10−7 . The scheme was
found to be second-order accurate except in the corner of the semicircle. We know the
local solution there exactly, given by (2.33), since the local analysis follows through in
e
a similar manner as x → ∞. From (2.36) we can see that close to the wall, as θ → 0, φ
should behave like θ−1/2 . This square root singularity will clearly not be well resolved
by finite difference approximations. We therefore attempt to subtract it out, writing
where
e−φ
ec,
e=φ
ψ
(3.9)
e c = − 3β (ξ ′ − sin θ)1/2 (σ 2 − 1 sin θ(ξ ′ − sin θ)),
(3.10)
φ
3
40σ
with 2σ = (1 − r2 ), and ξ ′ = σ 2 + sin2 θ. The derivative condition on the base is then
e c has the correct behaviour in the corner and
simplified (see (3.13b) below). Then φ
the singularity should have been extracted to leading order. In addition, in order to
e on r = 1, we need only approximate a second derivative on the wall since,
estimate A
by (3.6),
e rr |r=1 .
e
(3.11)
p1rrr |r=1 = −6φ
The system we now have to solve is as follows:
ec,
Le
ψ = −Lφ
(3.12)
e = 0 on r = 1,
ψ
e θ = 0 on θ = 0, π.
ψ
(3.13a)
(3.13b)
where L is the operator defined in (3.7). The boundary conditions are
Finite differences are used as before and accuracy in the corner is somewhat improved
e c behaves like
although there remains a problem as now near the wall when θ → 0, Lφ
1/2
e θθ . However, we find
nθ and so errors are still likely to arise in the discretization of ψ
that taking hθ small enough results in good agreement in the corner with the known
solution (2.33). Figure 3(a) makes a comparison of e
p1rrr |r=1 used in the calculation of
e according to (3.2). It shows plots of e
A
p1rrr |r=1 for the modified equation (3.7), the
equation with the corner solution subtracted (3.12), and the corner solution (3.10)
itself. We can see that the numerical results approach the local solution as θ → 0,
the best approximation being given by the corner-subtracted solution. However, hθ
has to be taken quite small to obtain this accuracy. Figure 3(b) shows how e
p1 on
θ = 0 compares with n5/2 as n → 0. We believe then that we have solved the core
accurately over the whole semicircle except very far into the corners where, even
having subtracted out the square root singularity in θ, finite differences are unable
to reproduce the local solution for very small θ without taking an unfeasibly large
number of grid points.
351
Steady flow in a dividing pipe
20
(a)
15
10
5
0
0.02
0.04
0.06
0.08
0.10
0.92
0.96
1.00
h
0
(b)
– 0.001
– 0.002
– 0.003
0.80
0.84
0.88
r
Figure 3. Numerical solution of II as x → ∞ in the corner. (a) n ≪ 1, local solution (——),
solving (3.7) (- ⋄ - - ⋄ -), solving (3.12) (· · ·). (b) θ ≪ 1: e
p1 (——) and −(3β/20)n5/2 (· · ·).
e varies with θ. The singular θ−1/2
Figure 4 shows how the displacement function A
behaviour is apparent near θ = 0, π. Figure 5(a) shows contours of pressure e
p1 within
the core. We can see that the pressure is greatest near the coordinate origin. In
figure 5(b) the singular nature of the down-pipe velocity perturbation is reflected
352
M. G. Blyth and A. J. Mestel
3
2
1
2
Ã
1
0
1
2
3
h
e
varying with θ.
Figure 4. 12 A(θ)
(a)
(b)
Figure 5. Both figures refer to the limit x → ∞. Dashed lines indicate negative contours. (a)
p1 |) = 0 on outer contour; Max(|e
p1 |) = 0.019 at the origin. (b)
Pressure contours of e
p1 : Min(|e
Contours of down-pipe velocity perturbation e
u1 : Min(e
u1 ) = 0.003 at the origin; Max(e
u1 ) is at the
corner singularities (see figure 4).
in the way the contours bunch together in the corners. We note that despite the
largest downstream velocities occurring near θ = 0, π, this does not correspond to the
greatest fluid flux since the singularity disappears upon a single integration. Figure 6
displays contours of the secondary velocities in region II. Whilst the bulk of the fluid
is being carried downstream, there is a secondary flux ‘upwards’ into the viscous wall
layer III, in which region this additional fluid is flushed downstream.
3.2. The full x problem
We now solve for the pressure in region II for all values of x. If we take the Fourier
transform of (2.5) and make a similar substitution to that made in § 3.1, namely
p∗1 = (1 − r2 )φ, we have
∇22 φ − 4
1 + r2
φ − k 2 φ = 0,
(1 − r2 )2
(3.14)
Steady flow in a dividing pipe
(a)
353
(b)
Figure 6. Secondary velocities in II as x → ∞ (dashed lines indicate negative contours). (a) Contours
of v1 : Min(|v1 |) = 0 on r = 1, and |v1 | is largest around the origin. (b) Contours of w1 : Min(w1 ) = 0
on r = 1; w1 is largest around the origin. Note that the origin is purely a coordinate singularity.
with
φ = 0 on r = 1,
and boundary condition on the symmetry line,
φθ = ∓ 161 βeiπ/4 r(1 − r2 )1/2 k 1/2
(3.15)
on θ = 0, π,
(3.16)
which comes from taking the transform of (2.16). For negative k the square root is
taken according to the branch cut made in § 2.3. It is convenient to scale the k 1/2 from
(3.16). This leaves us with the task of solving (3.14) according to
φ = 0 on r = 1,
(3.17)
and
(3.18)
φθ = ∓ 161 βeiπ/4 r(1 − r2 )1/2 on θ = 0, π.
Once again a numerical solution is sought using finite differences. The procedure is
similar to the one employed previously, with convergence of the grid being reached
using point iteration. There are similar problems to before in rendering the behaviour
of the solution correctly in the corners. We now discuss the asymptotic behaviour
of the solution for small and large k and show that, except far into the corners, our
numerics approximate the known solutions well.
3.2.1. The case k ≪ 1
When k is small the local analysis has already been performed for p∗1 in § 2.4, the
solution being given by (2.33). Therefore, since p∗1 ∼ 2nφ as r → 1, the local solution
here is given by
(3.19)
φc = 56 λc (ξ − θ)1/2 (n2 − 13 θ(ξ − θ))/n,
with λc = 161 βeiπ/4 21/2 arising from (3.18) as r → 1. A plot similar to that in figure 3(a)
for x → ∞ reveals that if hθ is taken small enough, the numerical solution approaches
the analytic solution for θ → 0.
3.2.2. The case k ≫ 1
From the form of (3.14) we expect φ to decay exponentially as |k| → ∞. Writing
φ ∼ eb(r,θ)k for large positive k we derive the eikonal equation
b2r +
1 2
b = 1,
r2 θ
(3.20)
at O(k 2 ) and so, except near the corners, we anticipate that
φ ∼ e−kr sin θ
as k → ∞.
(3.21)
354
M. G. Blyth and A. J. Mestel
Near the curved wall we consider the approximation to (3.14)
2φ
− k 2 φ = 0.
(3.22)
n2
Far enough into the corner the solution will be similar to (2.33). However when both
n and θ are of order k −1 , the fourth term in (3.22) will become significant. In this case
we write n = k −1 z and θ = k −1 y and expect the solution to be expanded as
φnn + φθθ −
φ = k −3/2 φ0 + · · · ,
(3.23)
because of (3.18). Writing Ψ = φ0z + φ0 /z the problem reduces to
2
∇ Ψ − Ψ = 0,
with
Ψ =0
Ψy = − 32 λc z −1/2
2
on z = 0,
on y = 0,
(3.24)
(3.25a)
(3.25b)
and where ∇ = ∂2 /∂z 2 + ∂2 /∂y 2 . Taking a Fourier transform in z, and requiring that
Ψ be an odd function of z, we identify the exact solution
Z ∞
2
1/2
sin(qz)e−y(q +1)
3λc
dq.
(3.26)
Ψ=√
q 1/2 (q 2 + 1)1/2
2π 0
We are interested in comparing this with our numerical solution and so we use (3.26)
to compute the local approximation to A∗ . Since we know that p∗1 ∼ n3 as n → 0,
p∗1nnn |n=0 ≈ 4kΨz |z=0 ,
(3.27)
where the k 1/2 factor has been scaled back into the right-hand side. The integral for
Ψz |z=0 can be done analytically and thus, recalling (2.22), we find
3
= −k −1 27/4 Γ ( 34 )λc (kθ)−1/4 K1/4 (kθ),
(3.28)
π
where K denotes the modified Bessel function of the second kind. From this we can
see that as θ → ∞, A∗c ∼ θ−3/4 e−kθ , and when θ → 0, A∗c ∼ θ−1/2 as we would expect.
Now A∗c is a purely local solution in the corner for large k and it is possible to
construct a solution which is valid to leading order for large k over the entirety of
region II. Writing Φ = φn + φ/n in (3.22) we approximate the solution to the resulting
equation by solving
(3.29)
∇22 Φ − k 2 Φ = 0
over the whole of the semicircle. Then as k → ∞ the most important terms around
the walls of the semicircle will be encapsulated in (3.29). Its boundary conditions are
Φ = 0 on r = 1, and on the base we take Φθ = g(z) where
1 ∗c
A
2
3
g(z) = − √ λc (1 − z 2 )−1/2 sgn(z),
2
(3.30)
which has been chosen so that Φθ approximates the base condition (3.25b) as z → 1.
It is possible to solve this by the method of images and the use of Green’s theorem
in the plane. Near r = 1
Z
1 1
(K0 (kr1 ) − K0 (kr2 ))g(z) dz,
(3.31)
Φ(r, θ) ≃ −
π −1
355
Steady flow in a dividing pipe
0
(a)
Large-k
corner approxns
–0.025
–0.050
(b)
Numerical solution
–0.075
–0.100
–0.125
0
0.05
0.10
0.15
0.20
h
Figure 7. (a) Comparison in the corner of II at k = 10.0. Large-k corner approximations Re{ 21 A∗c },
Re{ 12 A∗G } (dotted lines) with the real part of the numerical solution (continuous line). (b) Contours
of down-pipe velocity u1 in III at x = 5.0 (shown on a radially exaggerated scale).
where
r12 = z 2 + r2 − 2zr cos θ,
r22 = z 2 + 1/r2 − 2z cos θ/r.
(3.32a)
(3.32b)
Using Laplace’s method it is not hard to show that for large k, away from the corners,
the exponential behaviour of Φ agrees with (3.21). Furthermore we demonstrate that
(3.28) and (3.31) match up in the corner. As θ → 0, from (3.28) we have,
1 ∗c
A
2
∼ − 43 βeiπ/4 θ−1/2 k −3/2 .
(3.33)
∗
Now for the Green’s function solution (3.31) we may approximate A as
1 ∗
A
2 G
= 4k −2 Φr |r=1
(3.34)
∼ − 34 βeiπ/4 θ−1/2 k −3/2 ,
(3.35)
and then as θ → 0 it is possible to show that
1 ∗
A
2 G
and so the two solutions agree. Figure 7(a) shows how Re{ 21 A∗c } and Re{ 21 A∗G }
compare with the numerical approximation of Re{ 12 A∗ } for small values of θ. The
two analytic approximations coincide as predicted and both compare favourably with
the numerics. The discrepancy between analytic and numerical solutions worsens as
θ gets smaller. This is also to be expected since, as discussed above, the numerical
solution becomes increasingly inaccurate in the corners and unlike the case x → ∞
we have made no attempt here to improve accuracy for small θ.
4. Discussion and results
In this section we concentrate on the solution of the viscous wall layer III given
by equations (2.24). We will primarily be interested in obtaining estimates for the
skin frictions on the curved wall as it is these that, despite the gross simplifications
356
M. G. Blyth and A. J. Mestel
(a)
(b)
Figure 8. Radial and azimuthal flow velocities in region III at x = 5.0 (on a radially exaggerated
scale). (a) Contours of w 1 , which is positive everywhere and increases linearly with ζ as the core is
approached. (b) Contours of v 1 . There is no swirl in the layer since v 1 > 0 everywhere, reaching a
maximum near the corner. The fluid moves up along the curved wall and is carried downstream.
that have been made in this model, are of most physiological interest. Now that
the displacement function A∗ has been estimated, it is possible to proceed to obtain
numerical approximations of the transforms of the flow variables and their derivatives
in III. Once this has been done, a fast Fourier transform (FFT) routine is used to
invert them numerically. From (2.24a) and (2.26) we have
τ∗ax = u∗1ζ |ζ=0 = 21 µ0 ωA∗ + µ1 k 2 p∗1 /ω 4 ,
(4.1)
for the axial skin friction, where µ0 = 3Ai(0) + Ai′ (0)/Ai(0), and µ1 = (6Ai(0))−1 .
Also, from (2.24b)
τ∗az = v ∗1ζ |ζ=0 = −2µ1 p∗1θ /ω,
(4.2)
for the azimuthal skin friction. Then (4.1) and (4.2), once inverted, will provide
estimates of the down-pipe and azimuthal shear stresses respectively. In order to
proceed further we need to solve (2.26) with (2.27) and (2.38) for the pressure p∗1 . We
find that
Z
o
n Z π/2
cosh(k(θ − π/2)) π/2 ∗
∗
∗
A (k, t) cosh(kt) dt ,
p1 = C
A (k, t) sinh(k(θ − t)) dt +
sinh(kπ/2)
0
θ
(4.3)
′
5
∗
where C = 3Ai (0)ω /2k, which can then be calculated given A . We compute each
transform for a range of θ values and then transform back with the FFT (see for
example Brigham 1988). Contours of radial and azimuthal velocity perturbations
in the viscous wall layer are shown in figure 8. The flow does not recirculate in
this region since v 1 is always positive. The secondary flow in the core is sketched
in figure 9(b). Down-pipe pressure gradients in the core and viscous wall layer are
plotted in figure 10. Figure 11 shows plots of down-pipe and azimuthal shear stress
perturbations for different θ. It is noticeable that around x = 0 there is a region
where the axial shear stress perturbation has little azimuthal dependence.
We now consider the asymptotic behaviour of the flow variables as x → ±∞. Since
by (3.1d) as x → ∞, p1 ∼ x−3/2 we can deduce that A∗ ∼ k −3/2 as k → 0. Then it
follows that for small k
u∗1ζ |ζ=0 ∼ k −7/6 ,
v ∗1ζ |ζ=0 ∼ k −1/6 ,
p∗1 ∼ k −11/6 .
(4.4a)
(4.4b)
357
Steady flow in a dividing pipe
0.05
0
– 0.05
– 0.10
(b)
p1h
– 0.15
x increasing
– 0.20
– 0.25
– 0.30
(a)
0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
1.6
h
Figure 9. (a) Azimuthal pressure gradients in III for x = 0.25, . . . , 1.5 in steps 0.25. (b) Schematic
diagram of the secondary flow. Fluid is injected into the core from the plate layer I and swept
around the curved wall. The radial velocity is positive throughout the core and the wall layer III,
and mass conservation is maintained by the fluid being flushed downstream.
Whereupon, as x → ∞
u1ζ |ζ=0 ∼ x1/6 ,
v 1ζ |ζ=0 ∼ x−5/6 ,
p1 ∼ x5/6 .
(4.5a)
(4.5b)
How then do the variables behave as x → −∞? In the core if we assume the decay
is algebraic in x, so p1 ∼ x−α , α > 0, we have (2.5) without the x-derivative term
and boundary conditions p1 = 0 on r = 1 and p1θ = 0 on θ = 0, π by (2.16). In the
absence of any eigensolutions to this problem we deduce that p1 = 0 and so it seems
that for large negative x the decay in p1 will be exponential. Since regions II and III
are linked through the displacement function A(x, θ) we would consequently expect
everything there to decay in x at the same rate. We therefore propose in the core that
p1 = eλx f(r, θ) + · · ·
as x → −∞,
(4.6)
and correspondingly in the viscous wall layer
u1 = eλx u−∞ (r, θ) + · · ·
as x → −∞,
(4.7)
with similar expansions for the other variables in III. Then eigenvalues λ correspond
to eigenfunctions f being solutions of
∇22 f +
4r
fr + λ2 f = 0,
1 − r2
(4.8)
with f = 0 on r = 1 and fθ = 0 on θ = 0, π. We Fourier analyse in θ, writing
f=
∞
X
m=0
fm (r) cos mθ,
(4.9)
358
M. G. Blyth and A. J. Mestel
0
(a)
– 0.1
– 0.2
p1x
– 0.3
– 0.4
–2
0
2
4
6
10
8
x
0.0250
(b)
0.0125
0
p1x
– 0.0125
– 0.0250
– 0.0375
–4
–2
0
2
4
6
8
x
Figure 10. Down-pipe variation of pressure gradient perturbations. (a) In the viscous wall layer,
p1x at θ = π/2. (b) In the core, p1x at (r, θ)=(0.5, π/2).
which automatically satisfies the second condition on f. Then
4r3
2 ′′
r fm + r +
fm′ − (m2 − λ2 r2 )fm = 0,
1 − r2
(4.10)
359
Steady flow in a dividing pipe
1.2
1.0
0.8
sax
0.6
h decreasing
0.4
0.2
(a)
0
–1
1
0
2
3
5
4
x
0
– 0.05
h decreasing
– 0.10
saz
– 0.15
– 0.20
– 0.25
– 0.30
(b)
– 0.35
–4
–2
0
2
4
6
8
10
x
Figure 11. Variation of wall stress perturbation. (a) Axial τax . (b) Azimuthal τaz . For both θ
decreases in equal steps in the range π/4 6 θ 6 π/2.
for which we impose that fm (0) be regular and fm (1) = 0. We solve this problem
using the Fortran library Nag routine D02KEF. It is observed that when m = 0
the eigenvalues correspond to the zeros of J1 (r), the first-order Bessel function. This
suggests an exact solution is possible in this case and indeed we find that for m = 0
360
M. G. Blyth and A. J. Mestel
4.0
(a)
3.5
m
3.0
=
0
2.5
fm
2.0
1.5
=
6
1.0
m
0.5
0.1
0
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
r
–4
(b)
–5
–6
–7
–8
–9
–10
–11
–12
–13
–2
– 1.8
– 1.6
– 1.4
– 1.2
– 1.0
– 0.8
– 0.6
– 0.4
– 0.2
0
x
Figure 12. (a) Eigenfunctions fm for m = 0, . . . , 6. (b) Down-pipe comparison of log p1 at
(r, θ) = (0.5, π/2) in the core (——) with a line of gradient 3.83 (· · ·).
the eigenfunction is given by
f0 = J0 (λr) + r2 J2 (λr),
(4.11)
where J1 (λ) = 0. Note that this solution appears in slightly different form in Smith
Steady flow in a dividing pipe
361
(1976c). Despite extensive efforts, we have been unable to find an analytic solution
to (4.10) for general m. The numerical solutions predict that the first non-zero
eigenvalues λm0 for each mode increase more or less linearly with m. Figure 12(a)
shows the eigenfunctions fm for the first few modes. We are interested in the smallest
eigenvalue as this will then determine the slowest rate of decay of p1 . In fact we can
prove that the smallest eigenvalue should occur when m = 0. If we write (4.10) in
self-adjoint form
!′
λ2 r2 − m2
rfm′
fm = 0,
(4.12)
+
2
2
(1 − r )
r(1 − r2 )2
then we can use the Rayleigh–Ritz method (see for example Duff & Naylor 1966)
whereby the Rayleigh quotient
Z 1
(p(wt′ )2 − qwt2 ) dr
0
(4.13)
R=
Z 1
2
swt dr
0
is minimized over a set of l trial functions wt which obey the boundary conditions
on fm to obtain upper-bound estimates Λmj (j = 1, . . . , l) for the first l non-zero
eigenvalues λmj corresponding to mode m with equivalence between the two occurring
when wt is chosen to be the eigenfunction precisely. Here we have
p = r/(1 − r2 )2 ,
q = −m2 /r(1 − r2 )2 ,
s = r/(1 − r2 )2 ,
(4.14a)
(4.14b)
(4.14c)
λ00 < λ10 < · · · < λm0 < · · · ,
(4.15)
and so p, s > 0, q 6 0 for 0 6 r 6 1, ∀m. We wish to estimate λ00 . In general we know
= Λm+1
that λm0 = Λm0 if we choose wt = fm as a single trial function. Similarly λm+1
0
0
m
m
m
if wt = fm+1 . Then if we select wt = fm+1 to estimate λ0 we have λ0 6 Λ0 6 λm+1
.
0
Therefore
and so the first non-zero eigenvalue of the zeroth mode is the smallest. We conclude
then that as x → −∞
0
p1 = eλ0 x f(r, θ) + · · ·
as x → −∞,
(4.16)
with λ00 ≈ 3.8317. How does this compare with our numerical calculations?
Figure 12(b) shows a logarithmic plot of p1 beside a straight line of gradient 3.83. We
can see that even for relatively small x, p1 does seem to be decaying exponentially at
approximately the predicted rate.
5. Concluding remarks
We have extended the work of Smith (1977a) from the large down-pipe limit to
solving over the entire range of the developing flow. The effect of larger constrictions
and blockages on oncoming developed flow is considered by Smith (1977b) and those
in a developing entry flow by Smith (1976b). In both of these cases, for pipe and
channel flows under the correct conditions upstream influence may be significant and
wide-ranging. However, with reference to the comment made in §2.1 that upstream
362
M. G. Blyth and A. J. Mestel
influence may be anticipated, we have found in fact that it is minimal as the pressure
in the core decays exponentially upstream.
From a physiological point of view, bearing in mind the comments made in the
introduction, it is interesting to note that on the curved wall the axial skin friction
perturbation τax increases from its undisturbed Poiseuille value fairly sharply as the
splitter plate is reached and thereafter increases algebraically downstream:
τax =
1
2
+ R −1/6 τax ,
(5.1)
where τax ∼ x1/6 as x → ∞ by (4.5a). To leading order the axial shear stress near
the middle of the splitter plate will approximately be that of the classical Blasius
boundary layer,
1
(5.2)
τBax = R 1/2 √ f ′′ (0)(1 − z 2 )3/2 x−1/2 + O(1)
8 2
with f ′′ (0) ≈ 0.4696. So the axial skin friction is lower on the curved wall than on the
plate, with the two becoming of the same order of magnitude when x ∼ R. Of course
in a dividing pipe with a non-zero angle of bifurcation, separation may occur on the
upper wall which would affect the shear stress there dramatically. In this simplified
scenario however, no such separation is encountered.
There are many possible avenues of extension of the current study. Relaxation
of some of the simplifications mentioned in the introduction to make it of more
biological relevance could constitute much future work. The problem of merging flow
over the trailing edge of a splitter plate might be considered. This has been examined
in the two-dimensional case of a channel by Badr et al. (1985). They solved the flow
over the trailing edge of the plate both numerically and analytically. Working in
the large Reynolds number limit, they studied the flow on a lengthscale x = R 1/7 X,
where X is O(1), surrounding the trailing edge as suggested by Smith (1977b). They
then matched this to the following wake on an O(R) lengthscale, where a three-layer
structure emerges, with a core flow sandwiched between two viscous layers which
interact through a pressure-displacement relation in the core.
Finally we note that the work presented here could fairly easily be adapted to
pipes of different cross-section, for example an ellipse, which might better represent
the shape of an artery. The assumption of up/down symmetry could be relaxed as
the local solution to the core given in §2 can be simply altered for a corner of any
angle. The analysis also extends to other oncoming uni-directional profiles.
Appendix
Here we detail the functions g, κi which are needed for the matching of region IV
to regions I, II and III (see § 2.4). For i = 1, . . . , 4:
√
(A 1a)
κ1 = 3β sec(γ/2)x1/2 H(x)/ 2,
√
3
−1/2
κ2 = 2βcosec(γ) sin (γ/2)x
H(x),
(A 1b)
√
−1/2
H(x)/ 2,
(A 1c)
κ3 = 3β sin(γ/2)x
√
3
−3/2
(A 1d)
H(x)/(5 2),
κ4 = −β sin (γ/2)(2 cos(γ) + 3)x
where H(x) is the Heaviside function. Also,
3
g(x) = κ4 |(γ=π/2) = − 20
βx−3/2 H(x).
(A 2)
Steady flow in a dividing pipe
363
For i = 5, . . . , 8:
κ∗5 = 32 Q
nZ
ωζ
0
o
Ai(q) dq − Ai′ (0)M(ωζ) ,
κ∗6 = 6ω 3 QAi′ (0)M(ωζ),
Z ζ Z ωζ ′
∗
3
Ai(q) dq dζ ′ ,
κ7 = 2 ikQ
0
where
(A 3a)
(A 3b)
(A 3c)
0
√
κ8 = −βAi′ (0)Γ ( 16 )x−7/6 H(x)/(25/3 2π),
√
Q = −12 2λc k −3/2 ,
ω = ( 21 ik)1/3 ,
(A 3d)
(A 4)
and where an asterisk denotes a Fourier transform according to (2.23). These should
be inverted numerically.
REFERENCES
Badr, H., Dennis, S. C. R., Bates, S. & Smith, F. T. 1985 Numerical and asymptotic solutions for
merging flow through a channel with an upstream splitter plate. J. Fluid Mech. 156, 63–81.
Bennett, J. 1987 Theoretical properties of three-dimensional interactive boundary layers. PhD
thesis, University College, London.
Blyth, M. G. 1999 Steady flow in dividing and merging pipes. PhD thesis, Imperial College, London.
Brigham, O. R. 1988 The Fast Fourier Transform and its Applications. Prentice-Hall.
Caro, C. G., Fitz-Gerald, J. M. & Schroter, R. C. 1971 Atheroma and arterial wall shear:
Observation, correlation and proposal of a shear dependent mass transfer mechanism of
atherogenesis. Proc. R. Soc. Lond. B 177, 109–159.
Cowley, S. J. 1983 Steady flow through distorted rectangular tubes of large aspect ratio. Proc. R.
Soc. Lond. A 385, 107–127.
Dhanak, M. R. & Duck, P. W. 1997 The effects of freestream pressure gradient on a corner
boundary layer. Proc. R. Soc. Lond. A 453, 1793–1815.
Duff, G. F. D. & Naylor, D. 1966 Differential Equations of Applied Mathematics. John Wiley &
Sons.
Friedman, M. H., O’Brien, V. & Ehrlich, L. W. 1975 Calculations of pulsatile flow through a
branch. Implications for the haemodynamics of atherogenesis. Circ. Res. 36, 277–285.
Gammack, D. & Hydon, P. E. 1999 Steady flow in pipes of non-uniform curvature and torsion.
J. Fluid Mech. (submitted).
Lynch, D. G., Waters, S. L. & Pedley, T. J. 1996 Flow in a tube with non-uniform, time-dependent
curvature: governing equations and simple examples. J. Fluid Mech. 323, 237–265.
Lyne, W. H. 1970 Unsteady viscous flow in a curved pipe. J. Fluid Mech. 45, 15–31.
Pal, A. & Rubin, S. G. 1971 Asymptotic features of viscous flow along a corner. Q. Appl. Maths
29, 91–108.
Pedley, T. J. 1980 The Fluid Mechanics of Large Blood Vessels. Cambridge University Press.
Rieu, R. & Pelissier, R. 1991 In vitro study of a physiological type flow in a bifurcated vascular
prosthesis. J. Biomechanics 24, No.10, 923–933.
Rubin, S. G. 1966 Incompressible flow along a corner. J. Fluid Mech. 26, 97–110.
Rubin, S. G. & Grossman, B. 1971 Viscous flow along a corner: numerical solution of the corner
layer equations. Q. Appl. Maths 29, 169–186.
Smith, F. T. 1975 Pulsatile flow in curved pipes. J. Fluid Mech. 71, 15–42.
Smith, F. T. 1976a Pipeflows distorted by non-symmetric indentation or branching. Mathematika
23, 62–83.
Smith, F. T. 1976b On entry-flow effects in bifurcating, blocked or constricted tubes. J. Fluid Mech.
78, 709–736.
Smith, F. T. 1976c Flow through constricted or dilated pipes and channels: Part 2. Q. J. Mech. Appl.
Maths 29, 365–376.
364
M. G. Blyth and A. J. Mestel
Smith, F. T. 1977a Steady motion through a branching tube. Proc. R. Soc. Lond. A 355, 167–187.
Smith, F. T. 1977b Upstream interactions in channel flows. J. Fluid Mech. 79, 631–655.
Smith, F. T. 1979 Instability of flow through pipes of general cross-section. Part 1. Mathematika 26,
187–210.
Strikwerda, J. C. 1989 Finite Difference Schemes and Partial Differential Equations. Wadswirth &
Brooks/Cole.
Walton, A. G. & Smith, F. T. 1997 Concerning three-dimensional flow past a tall building on flat
ground. Q. J. Mech. Appl. Maths 50, 97–128.
Whitmore, R. L. 1968 Rheology of the Circulation. Pergamon Press, Oxford.
Wille, S. O. 1984 Numerical simulations of steady flow inside a three dimensional aortic bifurcation
model. J. Biomed. Engng. 6, No. 1, 49–55.
Zabielski, L. & Mestel, A. J. 1998a Steady flow in a helically symmetric pipe. J. Fluid Mech. 370,
297–320.
Zabielski, L. & Mestel, A. J. 1998b Unsteady blood flow in a helically symmetric pipe. J. Fluid
Mech. 370, 321–345.
Zamir, M. & Roach, M. R. 1973 Blood flow downstream of a two-dimensional bifurcation. J. Theor.
Biol. 42, 33–48.