You are on page 1of 12

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/257833534

Stability of a disperse-mixture flow in a boundary layer

Article  in  Fluid Dynamics · February 2008


DOI: 10.1134/S0015462808010080

CITATIONS READS
14 75

2 authors:

Sergei A. Boronin Alexander Osiptsov


Skolkovo Institute of Science and Technology Lomonosov Moscow State University
40 PUBLICATIONS   198 CITATIONS    109 PUBLICATIONS   707 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Investigation of vortex ring-like structures in internal combustion engines, taking into account thermal and confinement effects View project

Modelling of aerosols and sprays for medical and automotive applications View project

All content following this page was uploaded by Alexander Osiptsov on 12 October 2014.

The user has requested enhancement of the downloaded file.


ISSN 0015-4628, Fluid Dynamics, 2008, Vol. 43, No. 1, pp. 66–76. © Pleiades Publishing, Ltd., 2008.
Original Russian Text © S.A. Boronin, A.N. Osiptsov, 2008, published in Izvestiya Rossiiskoi Akademii Nauk, Mekhanika Zhidkosti i Gaza, 2008,
Vol. 43, No. 1, pp. 76–87.

Stability of a Disperse-Mixture Flow


in a Boundary Layer
S. A. Boronin and A. N. Osiptsov
Received December 25, 2006

Abstract—The hydrodynamic stability of a dilute disperse mixture flow in a quasi-equilibrium region


of a boundary layer with a significantly nonuniform particle concentration profile is investigated. The
mixture is described by a two-fluid model with an incompressible viscous carrier phase. In addition to
the Stokes drag, the Saffman lifting force is taken into account in the interphase momentum exchange.
On the basis of a numerical solution of the boundary-value problem for a modified Orr–Sommerfeld
equation, neutral stability curves are analyzed and the dependence of the critical Reynolds number on
the governing parameters is studied. It is shown that taking into account the particle concentration
nonuniformity in the main flow and the Saffman lifting force significantly changes the stability limits of
the two-phase laminar boundary layer flow. The effect of these factors on the boundary layer stability is
considered for the first time.
DOI: 10.1134/S0015462808010080
Keywords: stability, boundary layer, disperse mixture, particles, Orr–Sommerfeld equation, Saffman
force.

The problem of the hydrodynamic stability of a plane-parallel flow of a viscous dusty medium was first
considered in [1]. In that paper, a modified Orr–Sommerfeld equation for the main-flow stream function
amplitude was derived on the assumption of a Stokes drag law, a uniform particle distribution, and the
absence of phase velocity slip in the main flow. The equation obtained differs from the classical one by
reason of the presence of a complex analog of the main-flow velocity profile. An asymptotic analysis of the
neutral curves for plane-parallel two-phase flows performed within this formulation in [1, 2] for low-inertia
and high-inertia particles for the case of uniform particle concentration in the main flow demonstrated that
low-inertia particles destabilize (i.e. reduce the critical Reynolds number) and high-inertia particles stabilize
the flow.
There are many publications in which the neutral curves are investigated for different plane-parallel flows
and different values of the particle inertia parameter and mass concentration within a linear formulation
similar to that described above. Thus, the stability of a two-phase flow in a plane channel was studied
numerically in [3–5]. The stability parameters of free two-phase flows were studied in [6, 7] (mixing layer),
[8] (jets and wakes), and [9] (jets). In these studies, the results of the calculations show that the presence
of even a small (below 10%) mass concentration of particles uniformly distributed in the main flow can
significantly (by an order of magnitude) change the stability limits of plane-parallel viscous flows, the effect
of the particles being most significant on flows bounded by rigid walls. The most stabilizing effect on the
Poiseuille flow is produced by particles whose velocity relaxation length is of the order of the channel width.
In [5, 8, 10], the Saffman formulation was modified with account for a particle concentration nonuni-
formity in the main flow. Gaussian model concentration profiles with different locations of the maximum
were considered. The presence of a particle concentration nonuniformity in the main flow results in new
qualitative effects. Thus, Couette flow which is absolutely stable in a pure and uniformly dusty gas becomes
unstable starting from a certain critical value of the particle concentration maximum [10]. The stability of
a two-phase channel flow depends significantly on the position of the concentration maximum relative to

66
STABILITY OF A DISPERSE-MIXTURE FLOW 67

the so-called critical layer (in which, the main-flow velocity coincides with the velocity of the Tollmien–
Schlichting wave). In the papers cited, it was also shown that the most stabilizing effect occurs when the
maximum of the particle concentration coincides with the critical-layer location.Under certain conditions,
the instability region in the “wave number – Reynolds number” plane may be split into two disconnected re-
gions, in the space between which the flow is stable with respect to arbitrary two-dimensional disturbances.
In [11, 12], the stability of model two-phase flows with step-like particle concentration profiles in the main
flow was investigated.
In [13, 14], the stability of a uniformly dusty boundary layer with a Blasius velocity profile was studied
within a formulation similar to [1]. On the basis of a numerical calculation of the neutral curves, it was
shown that the maximum stabilizing effect is produced by particles whose velocity relaxation length is of
the order of the local boundary layer thickness, while low-inertia particles destabilize the flow. Due to a
singularity which arises in the modified Orr–Sommerfeld equation, attributable to the presence of particles,
a discontinuity appears in the dependence of the disturbance growth rate on the wave number for fixed
values of the Reynolds number and the particle inertia parameter. The discontinuities occur only in the
stable region of the spectrum and do not affect the shape of the neutral curve.
In [13, 14], two very important factors were disregarded. First, in the interphase momentum exchange
the lifting force was not considered, although in [15] it has already been shown that taking the Saffman
lifting force into account results in the appearance of instabilities even in a purely shear dispersed-mixture
flow and in [16, 17] it was demonstrated that in boundary layer flows the effect of the Saffman force is
significant even for small particles, several microns in size. Second, in analyzing the boundary layer stability
the initial particle concentration was assumed to be constant, although in [16, 18] it was shown that, in
the nonequilibrium region with respect to the phase velocities, a uniform particle concentration profile is
transformed into an essentially nonuniform profile with a shape that depends on the ratio of the Stokes
and Saffman forces. For “fine” particles, a profile with a monotonous increase in the particle concentration
towards the wall is typically formed, while, for “coarse” particles, in the quasi-equilibrium region the particle
concentration profile has a maximum on the outer edge of the boundary layer. In the present study, we
will extend the formulation of the stability problem for a two-phase boundary layer. Our aim is to take
into account particle concentration nonuniformities in the main flow and the Saffman lifting force in the
interphase momentum exchange and to analyze the effect of these factors on the stability limits of the
laminar flow regime.

1. THE DISPERSED-MIXTURE MODEL AND THE ROLE OF DIFFERENT FORCES


IN THE INTERPHASE MOMENTUM EXCHANGE
For describing the motion of a heterogeneous medium, we will use the two-continuum model [19]. The
mixture consists of a viscous incompressible carrier phase and dispersed particles. The particles are spheres
of the same radius σ and the mass m. The particle volume fraction is small τs  1, and the Einstein
correction to the viscosity is not taken into account. Despite the particle volume fraction being small, the
particle mass concentration is assumed to be finite: α = mNs /ρ ∼ O(1); here, Ns is the characteristic particle
number concentration and ρ is the density of the carrier phase. This condition can be attained when the ratio
of the densities of the carrier phase and the particle material is small. We will consider the range of particle
sizes on which we may neglect the Brownian motion (hence, the stress tensor of the particulate continuum
is zero) and assume a continuum flow regime around the particle. The Reynolds numbers of the flow around
the particles based on the relative flow velocity are assumed to be small: Res = ρ |v − vs |σ /μ  1. Here
and in what follows, the subscript s refers to the dispersed-phase parameters, and μ is the dynamic viscosity
of the carrier phase.
Since we will be considering the problem of the linear stability of the steady-state motion of a dispersed
mixture in the quasi-equilibrium region of the boundary layer, we will analyze the importance of the different
components of the interphase force for the conditions in question. The quasi-equilibrium region of the

FLUID DYNAMICS Vol. 43 No. 1 2008


68 BORONIN, OSIPTSOV

boundary layer is situated at distances L  lτ [16, 18] from the leading edge, where lτ = mU0 /6πσ μ is
the phase velocity relaxation length for the Stokes particle drag and U0 is the characteristic velocity of the
carrier phase (outer-flow velocity).
Under these conditions, the main components of the interphase force are as follows:

fSt = 6πσ μ (v − vs ), (1.1)


2 d 4 dv
fvm = πρσ 3 (v − vs ), fA = πρσ 3 ,
3 dt 3 dt
  
∂u
fM = πρσ [(v − vs ) × ω ],
3
fSa f = j6.46σ (u − us )
2
μρ  ,
∂y
t
2√ d(v − vs ) dt1
fBB = 6ρσ πν √ .
dt t − t1
0

Here, fvm is the virtual mass force, fA is the Archimedes force, fM is the Magnus force [20] (ω is the
particle angular velocity), fSa f is the Saffman force [21] (the unit vector j is perpendicular to the flow
direction), and fBB is the Basset–Boussinesq force [22].
We will estimate the ratio of these forces to the Stokes drag for an unsteady flow deviating slightly from
the steady-state flow in the quasi-equilibrium region of the boundary layer. For this purpose, we introduce
the characteristic scales of the quantities: the boundary layer thickness in the region considered δ , the
distance from the beginning of the boundary layer L, the velocity on the outer edge of the boundary layer U0 ,
the scale of the normal phase velocity in the boundary layer V0 , the scale of the difference in the longitudinal
and transverse phase velocity components in the linear growth stage of deviation of the solution from the
steady-state solution U1 (since the disturbances of the phase velocities may not fall outside the range of
applicability of the boundary layer approximation, we assume that U1 ∼ V0  U0 ), and the time scale L/U0 .
By virtue of the continuity equation, we obtain:

δ V 1 ρ U0 L
∼ 0 ∼√  1, ReL = .
L U0 ReL μ

Assuming also that the angular particle velocity is of the order of the carrier-phase vorticity ω ∼ U0 /δ ,
we obtain the following estimates of the magnitude of the different forces with respect to the Stokes drag:
     
|fA |x 2 σ 2 |fA |y 2 σ 2 |fvm | 1 σ 2
∼ , ∼ ReL , ∼ ,
|fSt |x 9 δ |fSt |y 9 δ |fSt | 9 δ
(1.2)
  |fSa f |y
|fBB | 1σ |fM |y 1 σ 2 6.46 σ 1/4
∼ , ∼ ReL , ∼ Re .
|fSt | πδ |fSt |y 6 δ |fSt |y 6π δ L

The analysis of formulas (1.2) indicates that, when the finiteness of the ratio of the particle size to the
boundary layer thickness is taken into account, the leading contribution is made by the Saffman force, with
the other forces being of lower order. Accordingly, in what follows in the interphase momentum exchange
we will take only the Stokes and the Saffman forces into account.
We will consider a neighborhood of a certain section of the two-phase boundary layer in the quasi-
equilibrium zone and write the system of equations of the two-continuum model [19] in nondimensional
form:

FLUID DYNAMICS Vol. 43 No. 1 2008


STABILITY OF A DISPERSE-MIXTURE FLOW 69

dv 1
div v = 0, = −∇p + Δv − nα fSt − nα fSa f , (1.3)
dt Re
∂n dvs
+ div(nvs ) = 0, = fSt + fSa f ,
∂t dt
 1/2
∂u
fSt = β (v − vs ), fSa f = K j (u − us ),
∂y
ρ U0 δ mN0 ρs 6πσ μδ
Re = , α= = , β= ,
μ ρ ρ mU0

6.46σ 2 (δ μρ )1/2 6.46  ρ
K= = √ β .
mU0 1/2
2π 2 ρs0

Here, ρs0 is the particle material density. In the nondimensionalization, the Cartesian coordinates are
scaled to the local boundary layer thickness δ , the phase velocity components to U0 , the pressure to ρ U02 ,
and the particle concentration to its characteristic value in the outer flow N0 . In addition to the Reynolds
number based on the boundary layer thickness, system (1.3) contains three similarity parameters: the relative
particle mass concentration α , the particle inertia parameter β (inverse Stokes number) equal to the ratio of
the boundary layer thickness to the particle velocity relaxation length, and the parameter K characterizing
the contribution of the Saffman force to the interphase momentum exchange.

2. FORMULATION OF THE STABILITY PROBLEM FOR THE TWO-PHASE


BOUNDARY LAYER
In accordance with [18], we assume that in the quasi-equilibrium region of the boundary layer considered
the main flow corresponds to the Blasius solution with a nonuniform particle distribution across the boundary
layer thickness. The velocity profile takes the form

U0 ρ
V = Vs = {U (ξ , 0)}, U (ξ ) = UB (ξ (1 + α )1/2 ), ξ =y .
μx
Here, UB (·) is the Blasius profile, x and y are the dimensional coordinates.
The flow is represented as a combination of the main flow and a small disturbance:
v = V + v , vs = Vs + vs , n = N + n , p = P + p .
Treating the velocity profile as locally plane-parallel, we can study the stability problem by using the
Fourier method and representing the disturbances in the form of harmonic waves propagating along the
plate. Then, in a certain neighborhood of x0 we assume that V = Vs = {U (y), 0}. We will consider the
system of equations for small disturbances, neglecting the second-order terms. For the carrier phase, we
obtain (in what follows, the asterisks denoting disturbances are omitted):
∂u ∂v
+ = 0, (2.1)
∂x ∂y
∂u ∂u dU ∂p 1
+U + v=− + Δu − α N fx ,
∂t ∂x dy ∂x Re
∂v ∂v ∂p 1
+U =− + Δv − α N fy − α N fsa f ,
∂t ∂x ∂y Re
 1/2
dU
fx = β (u − us ), fy = β (v − vs ), fSa f = K (u − us ),
dy
u(0) = v(0) = u(∞) = v(∞) = 0.

FLUID DYNAMICS Vol. 43 No. 1 2008


70 BORONIN, OSIPTSOV

The boundary conditions are the absence of slip on the wall and the damping of disturbances at infinity.
For the particulate medium, the linearized system takes the form:
∂n ∂N ∂n
+ N div(vs ) + vs +U = 0, (2.2)
∂t ∂y ∂x
∂ us ∂ us dU ∂ vs ∂ vs
+U + vs = fx , +U = fy + fSa f . (2.3)
∂t ∂x dy ∂t ∂x
The continuity equation for the particulate medium has been separated from the equation of motion,
so that Eqs. (2.1) and (2.3) can be solved independently of (2.2). Since in the case under study Squire’s
theorem, which can be proved in the same way as in the case of Stokes particle drag [1], is also valid, we
will consider only two-dimensional disturbances.

3. MODIFIED ORR–SOMMERFELD EQUATION


We will study the stability of the flow using the Fourier method. The disturbances of all the parameters
are taken in the form of individual harmonics

Q(x, y) = q(y) exp[ik(x − ct)].

Here, Q = {u, v, us , vs , p}, q(y) is the disturbance amplitude, c = cr + ici is the disturbance complex
velocity, cr is the wave phase velocity, ci is the wave amplitude growth rate, and k is the wavenumber.
Substituting these expressions for the disturbances in Eqs. (2.3), we obtain the system of two equations
for the phase velocity disturbance amplitudes u, v, us , vs :

− ikcus + ikU us + vsU  = β (u − us ),

− ikcvs + ikU vs = β (v − vs ) + K(U  )


1/2
(u − us ).

We express us and vs in terms of u and v:

β uγ − U  (β v + K(U  )1/2 u)
us (u, v) = , (3.1)
G
γ (β v + K(U  )1/2 u) − β K(U  )1/2U
vs (u, v) = , (3.2)
G
G = γ 2 − K(U  ) ,
3/2
γ = ik(U − c) + β . (3.3)

We introduce the carrier-phase stream function


∂Ψ ∂Ψ
u= , v=− , Ψ(x, y) = ψ (y) exp{ik(x − ct)}.
∂y ∂x
Eliminating the pressure from (2.1) and substituting expressions (3.1) and (3.2), we obtain the Orr–
Sommerfeld equation with additional differential terms describing the interphase momentum exchange

LOS = −αβ ReN(k2 ψ − ikvs (ψ , ψ  ))


  
dN  1/2 d
+ α Re β − N ikK(U ) −β (ψ  − us (ψ , ψ  )), (3.4)
dy dy

LOS = ψ IV − k2 ψ II + k4 ψ − ikRe((ψ II − k2 ψ )(U − c) − U  ψ ).


The boundary conditions take the form:

ψ (0) = ψ  (0) = 0, ψ (∞) = ψ  (∞) = 0. (3.5)

FLUID DYNAMICS Vol. 43 No. 1 2008


STABILITY OF A DISPERSE-MIXTURE FLOW 71

Equation (3.4), together with boundary conditions (3.5), determines the boundary-value problem for
a fourth-order linear ordinary differential equation with variable coefficients. Since both Eq. (3.4) and
boundary conditions (3.5) are uniform, a non-zero solution exists for specific values of the parameters
entering into (3.4). Fixing the parameters Re, α , β , and the imaginary part of the wave velocity ci , we
can determine the values of k and cr , which are called eigenvalues.

4. A NUMERICAL SOLUTION OF THE PROBLEM


For solving Eq. (3.4) we used the orthogonalization method [23]. In the neighborhood of y = ∞, Eq. (3.4)
has an analytical solution

ψ (y) = c1 exp[ky] + c2 exp[−ky] + c3 exp[(k2 + ikRe(U − c))1/2 ] + c4 exp[−(k2 + ikRe(U − c))1/2 ].

To satisfy the boundary condition on the outer edge of the boundary layer, it is necessary to eliminate the
two growing components. After that the basis of the solution of the boundary-value problem far from the
plate takes the form:

{exp[−ky0 ], exp[−k2 + ikRe(U (y0 ) − c))1/2 ]} = {φ1 , φ2 }.

The modified Orr–Sommerfeld equation was integrated from the boundary layer edge towards the plate
on the interval y ∈ [0, y0 ] (y0 = 9) using the fourth-order Runge–Kutta method. In accordance with the
orthogonalization method, the values of the functions ψ1 and ψ2 on the outer edge of the boundary layer
(y = y0 ) are used as the initial conditions for finding the two independent solutions ψ1 and ψ2 of Eq. (3.4).
The general solution is sought in the form ψ (y) = c1 ψ1 (y) + c2 ψ2 (y), where ψi (y0 ) = φi (y0 ), i = 1, 2. On
each integration step, the solutions ψ1 and ψ2 are orthonormalized and scaled with the aim of eliminating
the influence of their fast growing components (the details of the method can be found in [23]). At y = 0,
the boundary conditions take the form:

c1 ψ1 (0) + c2 ψ2 (0) = 0, c1 ψ1 (0) + c2 ψ2 (0) = 0.

A non-zero solution of this system exists under the condition

W (cr , ci , Re, α , β , K) = ψ1 (0)ψ2 (0) − ψ2 (0)ψ1 (0) = 0.

This system of two equations determines the region in parameter space for which a non-zero solution of
the boundary-value problem (3.4), (3.5) exists.
Let us consider the real and imaginary parts of the function G (3.3)

Re{G} = (kci + β )2 − k2 (U (y) − cr )2 − K(U  (y))


3/2
,
(4.1)
Im{G} = 2k(U (y) − cr )(kci + β ).

In the critical layer (y = yc ) where U (yc ) = cr , we have

Re{G} = (kci + β )2 − K(U  (yc ))


3/2
, Im{G} = 0.

Hence, for certain parameters of the problem the denominator of expressions (3.1) and (3.2) may be very
small, which presents significant difficulties for numerical integration. Physically, this signifies the resonant
acceleration of the particles in the critical layer. When the singular point of functions (3.1) and (3.1) is
close to the real axis, we used a procedure for the deformation of the integration contour so chosen that the
denominator in (3.1) and (3.2) is everywhere non-zero and not small. This makes it possible to avoid the
problem with numerical integration. For this purpose, we continued the main-flow velocity profile given

FLUID DYNAMICS Vol. 43 No. 1 2008


72 BORONIN, OSIPTSOV

Fig. 1. Dependence of the critical Reynolds number on the particle inertia parameter β for particle concentration profiles
(5.1) I–III without allowance for the lifting force (curves (1–3), pure gas (α = 0) (4)).

by the modified Blasius formula from the real axis onto the complex plane, replacing the variable y by the
complex variable y = yr + iyi .
Since in a neighborhood of the critical layer at ci = 0 the eigenfunctions of the problem (the amplitudes of
the disturbances of the velocity components of both phases) grow without bound, the linear formulation of
the problem ceases to be valid. The numerical calculations were performed outside a certain neighborhood
of the parameters corresponding to the zero of the function G, where the maximum value of the modulus of
the eigenfunctions was sufficiently small.

5. DISCUSSION OF THE RESULTS


The numerical method with eigenfunction orthogonalization has been used in many earlier studies, in-
cluding studies of dusty-gas flow stability [14]. To test the algorithm, we performed the calculations for
pure- and dusty-gas channel flows. In addition, certain calculations were compared with the eigenvalues
and eigenfunctions given in [14], which showed the agreement to four decimal places.
We will estimate the effect of particles on the stability by finding the location of the neutral curves in
the “Reynolds number – wave number” plane. The minimum Reynolds number Rec corresponding to the
left edge of the neutral curve is called critical. Thus, for Re > Rec inside the region bounded by the neutral
curve the disturbances grow, while, for Re < Rec , all small disturbances are damped.
We will first analyze the effect of particle concentration nonuniformity without taking the lifting force
into account. In the numerical calculations we will use three qualitatively different particle concentration
profiles in the main flow:

I. N(y) = 1 + exp(−y), II. N(y) = 1, III. N(Y ) = 1 − 0.5 exp(−y). (5.1)

As mentioned above, the profiles with increase I and decrease III of the particle concentration towards the
wall qualitatively correspond to the solutions of the problem of steady-state two-phase flow in the boundary
layer [16, 18] for “fine” and “coarse” particles respectively.
We will present the results of the numerical calculations of the stability parameters for the case of Stokes
particle drag. The calculations were performed for the value α = 0.1. The most stable flow is that with
a monotonous increase in the particle concentration towards the wall (Fig. 1). In this case, the maximum
critical Reynolds number, corresponding to the case of particles with a velocity relaxation length of the order
of the boundary layer thickness, is almost an order of magnitude greater than for a pure gas. In the case of
decreasing particle concentration towards the wall, the flow is less stable.

FLUID DYNAMICS Vol. 43 No. 1 2008


STABILITY OF A DISPERSE-MIXTURE FLOW 73

Fig. 2. Dependence of the critical Reynolds number on the particle inertia parameter β for particle concentration profiles
(5.1) I–III (a)–(c). K(β ) (1), K = 0 (2), pure gas (α = 0) (3).

When both the Stokes and Saffman forces are taken simultaneously into account, the particle concen-
tration profiles formed far from the leading edge significantly depend on the parameter κ , which can be
obtained if, in the parameter K introduced above, the boundary layer thickness is replaced by the particle
velocity relaxation length for Stokes drag. Accordingly, the parameters κ and K are independent. By vary-
ing the boundary layer thickness, in the linear problem considered we can vary the value of K at a fixed
value of the parameter κ . Thus, by considering qualitatively different concentration profiles (5.1) and vary-
ing the parameter K in the numerical calculations, we can analyze different ratios between the Stokes and
the Saffman force and study the effect of the Saffman force on the boundary layer stability.
For a fixed ratio of the phase material densities, the parameter K is expressed in terms of β (see (1.3)).
Accordingly, in the numerical calculations, we varied the parameter β over a wide range and the value of K
was found from (1.3) for the fixed value ρ /ρs0 = 10−3 .
For certain values of the governing parameters, in Eq. (3.4) a singularity appears in the critical layer,
which is attributable to the resonant acceleration of the particles in this layer. We can find the value β = β0
for which function (4.1) vanishes in the critical layer on the neutral curve:
  2/3
6.46 ρ
β0 = √ U  (yc ).
2π 2 ρs0
Here, y = yc is the critical layer. The numerical calculations show that the parameter β0 takes the value
0.028 and is almost independent of both the particle concentration in the main flow and the location of the
point on the neutral curve.
As the particle inertia parameter approaches the value β0 from the left, the critical Reynolds number
grows without bound (Fig. 2a–2c). For example, for β = 0.05 and concentration profile (5.1) I:

FLUID DYNAMICS Vol. 43 No. 1 2008


74 BORONIN, OSIPTSOV

Fig. 3. Eigenfunctions of the problem for Re = 15855.03, β = 0.05, α = 0.1, particle concentration profile (5.1) I cor-
responding to the mode k = 0.054281, Cr = 0.156868, Ci = 0. Curves (1–4) correspond to |us (y)|, |vs (y)|, |u(y)|, and
|v(y)|.

Rec = 1.58 × 104 . The eigenfunctions for this case are represented in Fig. 3. Clearly, with approach to
the singularity the particle velocity disturbances increase in the critical layer and hence the singularity of
functions (3.1) and (3.2) is unremovable. In this case, the maximum value of the eigenfunctions cannot be
regarded small in the vicinity of β0 and, for the complete investigation of the problem, it is necessary to take
into account the nonlinear terms omitted in (3.4). In the case when the concentration profile takes the form
(5.1) II or III, with approach to β0 from the right the critical Reynolds number remains finite and differs
only slightly from that obtained when only the Stokes drag is taken into account.
At a fixed Reynolds number, taking the lifting force into account results in flow stabilization. The form
of the dependence of the disturbance growth rate on the wave number is retained (Fig. 4). The presence of
the lifting force may result in the appearance of discontinuities in the spectrum of the problem (this effect is
also observed for purely Stokes drag [14]); however, due to the sharp increase in the eigenfunctions near the
parameters corresponding to zeroes of the function (4.1) these discontinuities are physically meaningless.
We investigated the dependence of the neutral curves on the particle mass concentration. The calcula-
tions showed that an increase in the particle mass concentration results in significant stabilization of the flow
for all finite values of the parameter β . Figure 5 shows the dependence Rec (α ) for Stokes particle drag and
β = 0.08, which approximately corresponds to the peak on the curves in Fig. 1. As the particle mass con-
centration varies from 0 to 0.3, the critical Reynolds number increases nonlinearly and, for the concentration
profile (5.1) I, reaches a value of the order of 2 × 105 , which is three orders of magnitude greater than the
critical Reynolds number for a pure gas.
For all the particle concentration profiles considered, the maximum stabilizing action on the flow is
produced by those particles whose velocity relaxation length is of the order of the boundary layer thickness.
A similar conclusion was reached for a uniform particle concentration profile in the main flow [4, 14]. The
dependence of the critical Reynolds number on the particle inertia parameter corresponds to the asymptotic
solution obtained in [1] for “coarse” particles (β  1) and is valid for a nonuniform particle concentration
profile. It should be noted that, in contrast to the case of uniform particle concentration, for the particle
concentration increasing towards the wall considered, “fine” particles (β  1) noticeably stabilize the flow.
In the case of a uniform particle concentration, the calculated values of the critical Reynolds number agree
with the asymptotic solution for “fine” particles [1].
Taking the lifting force into account results in the appearance of a singularity on the neutral curve at a
certain value of the particle inertia parameter. Since the denominator of expressions (3.1) and (3.2) depends
on the eigenvalues of the problem, the spectrum of the problem is different in the left and right neighbor-
hoods of the value β0 . An analysis of the analytical continuation of the function G (4.1) indicates that, for
β > β0 , the zero of the function G lies above the real axis and, for β < β0 , below the real axis. For all the
particle concentration profiles considered, the critical Reynolds number grows without limit with approach
to β0 from the left. In the case (5.1) I, this effect is observed in the right neighborhood of β0 .

FLUID DYNAMICS Vol. 43 No. 1 2008


STABILITY OF A DISPERSE-MIXTURE FLOW 75

Fig. 4. Dependence of the disturbance growth rate on the wavenumber for Re = 10000, β = 0.021, α = 0.1 and the
concentration profile (5.1) I. K(β ) (1), K = 0 (2), pure gas (α = 0) (3).

Fig. 5. Dependence of the critical Reynolds number on the particle mass concentration α for β = 0.08, K = 0 for the
concentration profiles (5.1) I–III (curves (1–3)).

Thus, as in the case of purely Stokes particle drag, when the lifting force is taken into account the pres-
ence of the particles results in flow stabilization over a narrow range of variation of the particle inertia
parameter, which corresponds to particles whose velocity relaxation length is of the order of the boundary
layer thickness. With approach to β0 , the eigenfunctions of the problem grow without bound in the neigh-
borhood of the critical layer. Accordingly, the linear formulation of the problem becomes invalid and, to
investigate further, it is necessary to consider nonlinear formulations of the stability problem. For all parti-
cle concentration profiles an increase in the particle mass concentration results in a substantially nonlinear
increase in the critical Reynolds number, which agrees with the results of [4] obtained for a uniform particle
concentration profile in the main flow.
Summary. Within the framework of the model of interpenetrating continua, the problem of the linear
stability of a dispersed-mixture flow in a quasi-equilibrium (with respect to the phase velocities) region
of the boundary layer is studied for different nonuniform particle concentration profiles. In the interphase
momentum exchange, both the Stokes drag and the Saffman lifting force are taken into account.
Without the lifting force, the flows with a monotonous increase in the particle concentration towards the
wall are the most stable. For a 10% particle mass concentration on the outer edge of the boundary layer,
a range of the particle inertia parameter exists for which the critical Reynolds number is almost an order
greater than for the pure gas and three times greater than in the case of a uniform particle concentration. The
most stabilizing effect is attained for those particles whose velocity relaxation length is comparable with the
boundary layer thickness.

FLUID DYNAMICS Vol. 43 No. 1 2008


76 BORONIN, OSIPTSOV

Taking the lifting force into account results in significant flow stabilization over a narrow range of vari-
ation of the particle inertia parameter, close to the value at which the maximum stabilization is attained for
purely Stokes particle drag. The maximum stabilization effect is again observed when the particle concen-
tration in the main flow increases monotonously towards the wall.
With increase in the particle mass concentration, the critical Reynolds numbers increase according to an
essentially nonlinear law. Thus, for Stokes drag and a 30% particle mass concentration the critical Reynolds
number is three orders greater than for the pure gas.
The work received financial support from the RFBR (No. 05-01-00502).
REFERENCES
1. P.G. Saffman, “On the Stability of Laminar Flow of a Dusty Gas,” J. Fluid Mech. 13, Pt 1, 120–128 (1962).
2. D.H. Michael, “The Stability of Plane Poiseuille Flow of a Dusty Gas,” J. Fluid Mech. 18, Pt 1, 19–32 (1964).
3. Ch.B. Narmuratov and A.S. Solov’ev, “Effect of Dispersed Particles on the Stability of Plane Poiseuille Flow,”
Fluid Dynamics 21 (1), 38–44 (1986).
4. E.B. Isakov and V.Ya. Rudyak, “Stability of Rarefied Dusty-Gas and Suspension Flows in a Plane Channel,” Fluid
Dynamics 30 (5), 708–712 (1995).
5. E.B. Isakov and V.Ya. Rudyak, “Stability of Poiseuille Flow of a Two-Phase Fluid with a Nonuniform Particle
Distribution,” Prikl. Mekh. Tekh. Fiz. 37 (1), 95–105 (1996).
6. Y. Yang, J.N. Chung, T.R. Troutt, and C.T. Crowe, “The Influence of Particles on the Spatial Stability of Two-
Phase Mixing Layers,” Phys. Fluids A 2 (10), 1839–1845 (1990).
7. X.-L. Tong and L.-P. Wang, “Two-Way Coupled Particle-Laden Mixing Layer. Pt 1: Linear Instability,” Int. J.
Multiphase Flow 25 (4), 575–598 (1999).
8. E.G. Bord, E.B. Isakov, and V.Ya. Rudyak, “The Stability of Laminar Flows of Dilute Dispersed Media,” Fluid
Dynamics 32 (4), 495–499 (1997).
9. J. DeSpirito and L.-P. Wang, “Linear Instability of Two-Way Coupled Particle-Laden Jet,” Intern. J. Multiphase
Flow 27, 1179–1198 (2001).
10. V.Ya. Rudyak, E.B. Isakov, and E.G. Bord, “Instability of Plane Couette Flow of Two-Phase Fluids,” Pis’ma ZhTF
24 (5), 76–80 (1998).
11. O.A. Druzhinin, “The Dynamics of a Concentration Interface in a Dilute Suspension of Solid Particles,” Phys.
Fluids 9 (2), 315–324 (1997).
12. J.A. Hernandez, “Instabilities Induced by Concentration Gradients in Dusty Gases,” J. Fluid Mech. 435, 247–260
(2001).
13. Ch.B. Narmuratov and A.S. Solov’ev, “Stability of a Two-Phase Gas-Solid Particle Flow in a Boundary Layer,”
Fluid Dynamics 23 (2), (1987).
14. E.S. Asmolov and S.V. Manuilovich, “Stability of a Dusty-Gas Laminar Boundary Layer on a Flat Plate,” J. Fluid
Mech. 365, 137–170 (1998).
15. D.A. Drew, “Lift-Generated Instability of the Plane Couette Flow of a Particle-Fluid Mixture,” Phys. Fluids 18
(8), 935–938 (1975).
16. A.N. Osiptsov, “Motion of a Dusty Gas at the Entrance to a Plane Channel and a Circular Pipe,” Fluid Dynamics
24 (6), 867–874 (1988).
17. B.Y. Wang and A.N. Osiptsov, “Near-Wall Boundary Layer Behind a Shock Wave in a Dusty Gas,” Fluid Dynam-
ics 34 (4), 505–515 (1999).
18. A.N. Osiptsov, “Structure of the Laminar Boundary Layer of a Disperse Medium on a Flat Plate,” Fluid Dynamics
16 (4), 512–517 (1980).
19. F.E. Marble, “Dynamics of Dusty Gases,” Annu. Rev. Fluid Mech. 2, 397–446 (1970).
20. S.J. Rubinow and J.B. Keller, “Transverse Force on a Spinning Sphere Moving in a Viscous Fluid,” J. Fluid Mech.
11 Pt 3, 447–458 (1961).
21. P.G. Saffman, “The Lift on a Small Sphere in a Slow Shear Flow,” J. Fluid Mech. 22 Pt 2, 385–400 (1965).
Corrigendum: J. Fluid Mech. 31, 624 (1968).
22. M.R. Maxey and J.J. Riley, “Equation of Motion of a Small Rigid Sphere in a Nonuniform Flow,” Phys. Fluids 26
(4), 883–889 (1983).
23. S.K. Godunov, “Numerical Solution of Boundary-Value Problems for Systems of Linear Differential Equations,”
Usp. Matem. Nauk 16 (3), 171–173.

FLUID DYNAMICS Vol. 43 No. 1 2008

View publication stats

You might also like