You are on page 1of 6

ISSN 0015-4628, Fluid Dynamics, 2014, Vol. 49, No. 2, pp. 232–237. © Pleiades Publishing, Ltd., 2014.

Original Russian Text © N.E. Leont’ev, 2014, published in Izvestiya Rossiiskoi Akademii Nauk, Mekhanika Zhidkosti i Gaza, 2014, Vol. 49, No. 2,
pp. 107–112.

Flow past a Cylinder and a Sphere in a Porous Medium


within the Framework of the Brinkman Equation
with the Navier Boundary Condition
N. E. Leont’ev
Moscow State University, Faculty of Mechanics and Mathematics,
Leninskiye Gory 1, Main Building, GSP-1, Moscow, 119991 Russia
e-mail: leontiev n@mail.ru
Received March 2, 2013

Abstract—Exact analytical solutions of the problem of flow past a sphere and a cylinder in a porous
medium are derived within the framework of the Brinkman equation with the Navier boundary condition.
Attention is drawn to the fact that the no-slip condition imposed on the interface between the porous
medium and a solid, used, in particular, in the case of the Brinkman equation, must be in the general
case replaced by a condition that admits nonzero flow velocity at the boundary.
Keywords: porous medium, Brinkman equation, slip boundary condition, no-slip condition.
DOI: 10.1134/S0015462814020112

Slow flows of a Newtonian incompressible fluid in a highly-penetrated porous medium in the absence of
body forces are approximately described by the Brinkman equation [1] and the continuity equation
μ
−∇p − u + μ ′ Δu = 0, div u = 0, (0.1)
k
where p is the pressure, u is the seepage flow, μ is the fluid viscosity, k is the porous medium permeability
for large-scale flows, and μ ′ is a constant with the dimensionality of viscosity which is generally different
from μ [1]. Usually it is assumed that μ and μ ′ are of the same order (different authors give different
estimates for μ ′ [2] but μ ′ → μ as porosity tends to unity) and frequently, starting from the studies of
Brinkman himself [3], it is assumed that μ ′ = μ . This model is applicable for flows in a relatively large
volume of the pore space [2], for example, in fibrous materials and foams, for which, as distinct from the
typical rocks, the permeability k is not necessarily of the order of square of the characteristic dimension of
the intrapore area.
In many studies [4–7], often computational in nature [8], the no-slip condition u = 0 is imposed on the
interface between a porous medium and an impermeable solid, which can apparently be attributed to an
analogy with the formulation of the boundary condition for the Navier–Stokes equations. In a few studies,
in which the slip condition [9, 10] is posed, its use is formally justified by its greater generality.
One of the purposes of this study is to call attention to the fact that the use of the slip condition (if it is not
regarded as an approximation) is generally inadequate. In fact, the seepage velocity [11] used in the contin-
ual description of flows in porous media, is determined as the fluid velocity averaged over a volume much
larger than the dimensions of a microstructure; for this reason the boundary condition at an impermeable
boundary must admit a nonzero flow velocity. This requirement is satisfied by the impermeability condition
u ⋅ n = 0 (n is the normal to the boundary directed toward the porous medium) for the seepage laws which do
not involve the higher derivatives of the flow velocity (Darcy law and its different modifications that account
for the fluid inertia) and the slip condition analogous to the Navier condition [12–14] for a viscous fluid,

232
FLOW PAST A CYLINDER AND A SPHERE IN A POROUS MEDIUM 233

which is applied for the seepage flow laws of the type of the Brinkman equation and can be written in the
form:
b(∇ j ui + ∇i u j )n j (δki − ni nk ) = uk , k = 1, 2, 3. (0.2)
Here, ∇ stands for the covariant derivatives, δki is the Kronecker symbol, and b is a constant with the
length dimension; the summation is over repeating indices (we note that the equality of the normal flow
component to zero automatically follows from Eq. (0.2)).
Below we present analytical solutions of the problems of flow past impermeable cylinders and spheres
embedded in a porous medium within the framework of system (0.1), (0.2). In the case of the no-slip
conditions the solutions of the analogous problems were obtained in [15] and [16], respectively.
1. We will consider transverse flow past an impermeable cylinder, R in radius, embedded in a porous
medium; at infinity the flow velocity U is aligned with the axis x1 of a Cartesian coordinate system (x1 , x2 , x3 ).
Introducing the new dimensionless variables by the formulas

xi u k
xi1 = , u1 = , p1 = p (1.1)
R U μ RU

we arrive at the system (in what follows, the subscript “1” is omitted)

μ′ k
−∇p − u + α 2 Δu = 0, div u = 0, α2 =
μ R2

with the boundary conditions in the Cartesian coordinate system


( ( ) )
∂ ui ∂uj 
β + n j (δik − ni nk ) − uk  = 0,
∂xj ∂ xi r=1

b
lim u = (1; 0; 0), β= ,
r→∞ R
where r is the distance to the cylinder axis.
We will introduce the cylindrical coordinate system (r, θ , z) with θ measured from the x1 axis.
We will seek the r and θ components of the velocity in the form:

ur = F(r) cos θ , uθ = G(r) sin θ , p = H(r) cos θ + const. (1.2)

Substituting expressions (1.2) in the non-dimensionalized system of the porous medium model we obtain
( )
∂p ur 2 ∂ uθ
− − ur + α Δur − 2 − 2
2
= 0,
∂r r r ∂θ
( )
1∂p uθ 2 ∂ ur
− − uθ + α Δuθ − 2 − 2
2
= 0,
r ∂θ r r ∂θ
( )
∂ ∂ uθ 1 ∂ ∂ 1 ∂2
(rur ) + = 0, Δ= r + 2 .
∂r ∂θ r ∂r ∂r r ∂θ2

The boundary conditions take the form:

r→∞: ur → cos θ , uθ → − sin θ ,


( ) (1.3)
∂ uθ
r→1: ur = 0, β − uθ = uθ .
∂r
FLUID DYNAMICS Vol. 49 No. 2 2014
234 LEONT’EV

After some simple algebra we arrive at the following boundary value problem for the functions F, G, and
H dependent on the radius r only

3(r2 + α 2 )F ′ + (r3 − 3α 2 r)F ′′ − 6α 2 r2 F ′′′ − r3 α 2 F IV = 0,

F(+∞) = 1, lim (rF ′ (r)) = 0, F(1) = 0, (1 − β )F ′ (1) − β F ′′ (1) = 0,


r→+∞

G = −F − rF ′ , H = α 2 (4rF ′′ + r2 F ′′′ ) − rF − r2 F ′ .

The solution of this problem is as follows:

C2 C3 1
F(r) = C1 + + K1 (γ r), γ= ,
r 2 r α

α (2β + 1)K0 (γ ) + (β + 4β α 2 + 2α 2 )K1 (γ )


C1 = 1, C2 = − ,
D

2α 2 (2β + 1)
C3 = , D = α (2β + 1)K0 (γ ) + β K1 (γ ),
D

where K0 and K1 are the modified Hankel functions of the zeroth and first order [17].
The streamline patterns in a cylinder cross-section are qualitatively the same for the solution obtained
and the symmetric separationless ideal-fluid flow past the cylinder. As α → ∞ (viscous flow past a cylinder
in the Stokes approximation), the function F(r) vanishes pointwise, which corresponds to the absence of the
problem solution (Stokes paradox). In the α → 0 limit, irrespective of β , the solution obtained pointwise
corresponds to the solution of flow past the cylinder governed by the Darcy law, kinematically analogous
to the circulationless incompressible ideal-fluid flow past the cylinder (in this case, the boundary condition
involving β is not naturally fulfilled for the limiting solution).
The velocity field considerably varies at distances, as minimum of the order of α from the cylinder
surface, the parameter α not being small. In the equatorial plane θ = π /2 the velocity, equal in absolute
value to ∣G(r)∣, increases near the cylinder surface compared with its value at infinity (the velocity profile
has characteristic “ears”), which, for the no-slip case (β = 0) was noted in [15]. At the cylinder surface at
θ = π /2 the velocity is nonzero (at β ∕= 0) and, as α → 0, approaches the double velocity at infinity.
In Fig. 1a for the cases in which β = 0 (no-slip) and β = 1 we have plotted the functions F(r) having
the meaning of the velocity at θ = 0 (curves 1 and 2, respectively) and ∣G(r)∣ (curves 3 and 4).
2. The solution of the problem of axisymmetric flow past an impermeable sphere, R in radius, is obtained
in the same fashion. In the spherical coordinate system (r, θ , χ ) with the azimuthal angle θ measured from
the x1 axis the problem non-dimensionlized using Eq. (1.1) reduces to the solution of the system of equations
( )
∂p 2ur 2 ∂ (uθ sin θ )
− − ur + α Δur − 2 − 2
2
= 0,
∂r r r sin θ ∂θ
( )
1∂p uθ 2 ∂ ur
− − uθ + α Δuθ − 2 2 + 2
2
= 0,
r ∂θ r sin θ r ∂θ

∂ 2 r ∂ (uθ sin θ )
(r ur ) + = 0,
∂r sin θ ∂θ
( ) ( )
1 ∂ 2 ∂ 1 ∂ ∂
Δ= 2 r + 2 sin θ
r ∂r ∂r r sin θ ∂ θ ∂θ

FLUID DYNAMICS Vol. 49 No. 2 2014


FLOW PAST A CYLINDER AND A SPHERE IN A POROUS MEDIUM 235

Fig. 1. Dependence of the flow velocity on the radius r in flows past a cylinder (a) and a sphere (b); α = 1, in the plane (a)
or at the axis (b) of symmetry θ = 0: β = 0 (1), β = 1 (2); in the plane θ = π /2: β = 0 (3), β = 1 (4).

with the boundary conditions (1.3). Substituting the unknown functions in the form (1.2) gives after some
transformations the boundary value problem
( 2 )
4r + 8α 2 F ′ + (r3 − 8α 2 r)F ′′ − 8α 2 r2 F ′′′ − r3 α 2 F IV = 0,

F(+∞) = 1, lim (rF ′ (r)) = 0, F(1) = 0, (1 − 2β )F ′ (1) − β F ′′ (1) = 0,


r→+∞

together with the expressions for the other unknown functions


( )
1 ′ ′ ′′ 1 2 ′′′ 1
G = −F − rF , H = α 2F + 3rF + r F
2
− rF − r2 F ′ .
2 2 2
The solution of the boundary value problem is as follows:
c2 c3
F(r) = c1 + 3 + 3 exp(−γ r)(r + α ),
r r

c1 = 1, c2 = −((6α 3 + 6α 2 + 3α + 1)β + α (3α 2 + 3α + 1))/d,

c3 = 3α 2 (2β + 1) exp(γ )/d, d = (3α + 1)β + α .


As α → ∞, the solution obtained approaches the well-known Basset solution for the creep flow of a
viscous fluid past a sphere with the slip condition imposed on its surface [18]
3(2β + 1) 1 1
F(r) = 1 − + .
2(3β + 1) r 2(3β + 1)r3
It describes, for example, the motion of a spherical particle with a superhydrophobic coating. In other
limiting case, when α → 0, the solution describes flow past a sphere governed by the Darcy law (kinemati-
cally it coincides with the axisymmetric flow of an ideal incompressible fluid past a sphere).
The velocity field behavior is qualitatively analogous to the flow pattern in the case of the cylindrical
body (Fig. 1b). Here, for the cases β = 0 (no-slip) and β = 1 we have plotted the functions F(r) having the
meaning of the velocity at the axis of symmetry of the flow θ = 0 (curves 1 and 2, respectively) and ∣G(r)∣
describing the velocity field in the equatorial plane θ = π /2 (curves 3 and 4).
The qualitative behavior of the streamlines is the same in the slip and no-slip cases and is analogous to
that in the Stokesian flow pasta sphere (Fig. 2).
The presence of the differential term with the seepage velocity in the Brinkman equation manifests itself
as a special feature of the pressure field in the meridional section
( )
c2
p = − r + 2 cos θ + const.
2r
FLUID DYNAMICS Vol. 49 No. 2 2014
236 LEONT’EV

Fig. 2. Streamlines in flow past a sphere; α = 1 (meridional section); β = 1 and 0 (continuous and broken curves).

Fig. 3. Pressure contours in the meridional section in flow past a sphere; α = 1 and β = 1.

It is absent from the Stokesian flow past a sphere and a flow past a sphere governed by the Darcy law.
This feature is characterized by the presence of two saddle points of the function p for any values of α and
β . In Fig. 3, by way of illustration, a fragment of the pressure field is presented for the right half of the flow
x1 > 0 at α = 1 and β = 1. The presence of this singularity leads to a situation in which near the spherical
surface at the axis of symmetry the fluid flows toward a higher rather than a lower pressure, as it would be
in the case of the Darcy law.
An analogous phenomenon, that is, codirection of the pressure gradient and the flow velocity, can also
be observable in creep (Stokesian) steady viscous flows. This is attributable to the fact that the fluid flow is
forced not only by the pressure gradient but also by viscous stresses. A very simple example is the plane
flow with the stream function ψ (x, y) = x(y2 + y3 ) at point x = 0, y = 1. More clearly this is visible in the
case of potential viscous Stokesian flows for which, in the absence of body forces, the pressure is constant
and the viscous forces are the only reason for the motion.
Summary. Attention is drawn to the fact that the condition imposed on the interface between a porous
medium and an impermeable wall must admit fluid flow along the boundary. Because of this, setting the
no-slip condition when using the seepage equations with the higher spatial derivatives (those of Brinkman,
Darcy–Lapwood–Brinkman, and others) is generally inadequate. The use of this condition can lead to a
change in the flow pattern near the boundary.

The study was carried out with the support of the Russian Foundation for Basic Research (projects
Nos. 11-01-00051 and 11-01-00188).

REFERENCES
1. H.C. Brinkman, “A Calculation of the Viscous Force Exerted by a Flowing Fluid on a Dense Swarm of Particles,”
Appl. Sci. Res. 1 (1), 27 (1947).
2. J.-L. Auriault, “On the Domain of Validity of Brinkman’s Equation,” Transp. Porous Med. 79, 215 (2009).
3. H.C. Brinkman, “On the Permeability of Media Consisting of Closely Packed Porous Particles,” Appl. Sci. Res.
1 (1), 81 (1947).

FLUID DYNAMICS Vol. 49 No. 2 2014


FLOW PAST A CYLINDER AND A SPHERE IN A POROUS MEDIUM 237

4. I.B. Stechkina, “Drag of Porous Cylinders in a Viscous Fluid Flow at Low Reynolds Numbers,” Fluid Dynamics
14 (6), 912 (1979).
5. V.A. Babkin, “Investigation of the Relative Motion of a Viscous Fluid and a Porous Medium Using the Brinkman
Equation,” Fluid Dynamics 37 (4), 587 (2002).
6. D.A. Nield, “Spin-up in a Saturated Porous Medium,” Transp. Porous Med. 4, 495 (1989).
7. A.A. Avramenko and A.V. Kuznetsov, “Flow Instability in a Curved Porous Channel Formed by Two Concentric
Cylindrical Surfaces,” Transp. Porous Med. 69, 373 (2007).
8. R. Younsi, A. Harkati, and D. Kalache, “Numerical Simulation of Thermal and Concentration Natural Convection
in a Porous Cavity in the Presence of an Opposing Flow,” Fluid Dynamics 37 (6), 854 (2002).
9. M. Parvazinia, V. Nassehi, R.J. Wakeman, and M.H.R. Ghoreishy, “Finite Element Modelling of Flow Through a
Porous Medium between Two Parallel Plates Using the Brinkman Equation,” Transp. Porous Med. 63, 71 (2006).
10. S.D. Harris, D.B. Ingham, and I. Pop, “Mixed Convection Boundary-Layer Flow near the Stagnation Point on a
Vertical Surface in a Porous Medium: Brinkman Model with Slip,” Transp. Porous Med. 77, 267 (2009).
11. R.E. Collins, Flow of Fluids through Porous Materials, Reinhold Publ. Corp., New York (1961).
12. C.L.M.H. Navier, “Mémoire sur les lois du mouvement des fluides,” Mémoires de l’Académie des sciences de
l’Institut de France 6, 389 (1827).
13. S. Goldstein, “Note on the Condition at the Surface of Contact of a Fluid with a Solid Body,” in: S. Goldstein
(ed.), Modern Development in Fluid Dynamics. Vol. 2, Clarendon Press, Oxford (1938), p. 676.
14. E. Lauga, M.P. Brenner, and H.A. Stone, “Microfluidics: the No-Slip Boundary Condition,” in: C. Tropea et al.
(eds.), Handbook of Experimental Fluid Mechanics, Springer, Berlin (2007), p. 1219.
15. I. Pop and P. Cheng, “Flow past a Circular Cylinder Embedded in a Porous Medium Based on the Brinkman
Model,” Int. J. Engin. Sci. 30, 257 (1992).
16. I. Pop and D.B. Ingham, “Flow past a Sphere Embedded in a Porous Medium Based on the Brinkman Model,”
Int. Comm. Heat Mass Transfer 23, 865 (1996).
17. G.A. Korn and Th.A. Korn, Mathematical Handbook for Scientists and Engineers: Definitions, Theorems, and
Formulas for Reference and Review, McGraw Hill, New York (1968).
18. A.B. Basset, A Treatise on Hydrodynamics, with Numerous Examples. Vol. 2, Deighton, Bell, and Co., Cambridge
(1888).

FLUID DYNAMICS Vol. 49 No. 2 2014

You might also like