Under consideration for publication in J. Fluid Mech.
1
Segregation of a liquid mixture by a radially
oscillating bubble
By O L I V I E R L O U I S N A R D1 , F R A N C I S C O J. G O M E Z2
AND
R O M A I N G R O S S I E R1
1
Laboratoire de Génie des Procédés des Solides Divisés, Ecole des Mines d’Albi, 81013 ALBI
Cedex 09, FRANCE
2
Laboratorio de Ultrasonidos, Dpto. de Fisica, Universidad de Santiago de Chile, Casilia 302,
Santiago, CHILE
(Received ?? and in revised form ??)
A theoretical formulation is proposed for forced mass transport by pressure gradients in
a liquid binary mixture around a spherical bubble undergoing volume oscillations in a
sound field. Assuming the impermeability of the bubble wall to both species, diffusion
driven by pressure gradients and classical Fick-diffusion must cancel at the bubble wall,
so that an oscillatory concentration gradient arises in the vicinity of the bubble. The
Péclet number Pe is generally high in typical situations and Fick diffusion cannot restore
equilibrium immediately, so that an asymptotic average concentration profile may progressively build up in the liquid over large times. Such a behavior is reminiscent of the
so-called rectified diffusion problem, leading to slow growth of gas bubble oscillating in a
sound field. A rigorous method formerly proposed by Fyrillas & Szeri (1994) to solve the
latter problem is used in this paper to solve the present one. It is based on splitting the
problem into a smooth part and an oscillatory part. The smooth part is solved by a multiple scales method and yields the slowly varying average concentration field everywhere
in the liquid. The oscillatory part is obtained by matched asymptotic expansions in terms
of the small parameter Pe −1/2 : the inner solution is required to satisfy the oscillatory
balance between pressure diffusion and Fick diffusion at the bubble wall, while the outer
solution is required to be zero. Matching both solutions yields a unique splitting of the
problem. The final analytical solution, truncated to leading order, compares successfully
to direct numerical simulation of the full convection–diffusion equation. The analytical
expressions for both smooth and oscillatory parts are calculated for various sets of bubble
parameters: driving pressure, frequency and ambient radius. The smooth problem always
yields an average depletion of the heaviest species at the bubble wall, only noticeable for
large molecules or nano-particles. For driving pressures sufficiently high to yield inertial
oscillations of the bubble, the oscillatory problem predicts a periodic peak excess concentration of the heaviest species at the bubble wall at each collapse, lingering on several
tens of time the characteristic duration of the bubble rebound. The two effects may compete for large molecules and practical implications of this segregation phenomenon are
proposed for various processes involving acoustic cavitation.
1. Introduction
Radially oscillating bubbles forced by a sound field are commonly encountered in acoustic cavitation and sonoluminescence experiments (Crum et al. 1999). Generally, the liquid
surrounding such bubbles is not pure and involves various chemical species, which may
2
O. Louisnard, F. J. Gomez and R. Grossier
either participate in chemical reactions, or undergo phase transitions, like crystallization
processes. The kinetics of such processes depends on the concentrations of the species and
may therefore be influenced by any variation of the spatial homogeneity of the mixture.
Pressure gradients may be a possible source of mixture segregation. Following diffusion
theory (Hirschfelder et al. 1967; Bird et al. 1960), when a mixture of two species is subjected to a pressure gradient, the lightest one is pushed toward low pressure regions. This
forced diffusion process, known as pressure diffusion, generally remains weak unless the
liquid is submitted to high pressure gradients, as in ultracentrifuge applications where it
is used profitably to separate large molecular weight species from a solvent (Archibald
1938). Pressure diffusion is also responsible for gas stratification in a quiescent atmosphere, or solute-solvent segregation in long sedimentation columns (Mullin & Leci 1969;
Larson & Garside 1986). Besides, the effect of pressure diffusion, along with thermal diffusion, on the segregation of a gas mixture inside a radially oscillating bubble has been
investigated by Storey & Szeri (1999) in the context of sonoluminescence.
When a bubble is driven in radial motion by a high amplitude oscillating pressure field,
pressure gradients arise in the liquid, as a result of the bubble wall acceleration. Inertial
cavitation is a situation where the bubble suffers an explosive expansion followed by a
violent collapse. In this case, the pressure gradient reaches a very high value near the end
of the collapse, owing to the strong gas compression in the bubble. The segregation of two
species by pressure diffusion in the neighborhood of a cavitation bubble may therefore
notably influence the liquid homogeneity.
The similar problem of mass transport of a gas dissolved in a liquid around a bubble
undergoing volume oscillations has been studied extensively (Eller & Flynn 1965; Hsieh
& Plesset 1961; Fyrillas & Szeri 1994, 1995): the variations of the gas concentration at the
bubble wall, driven by the bubble oscillations, yield a non-zero average gas flux toward
the bubble, reversing its natural dissolution process, a phenomenon known as rectified
diffusion. In this case, the dissolved gas flux in the liquid arises as a consequence of the
asymmetry in the behavior of the two components at the bubble wall: the gas can cross
the interface, the liquid cannot. In the present problem, assuming a binary mixture of
two non-volatile and non-surface-active fluids, the bubble interface acts as a barrier for
the two species, which would prohibit any relative flux. However, if pressure diffusion
is taken into account, a new asymmetry between the two components arises, owing to
their different densities, and the pressure gradient near the bubble wall separates the
two species. Since the net flux across the bubble wall of any of the two species must
be zero, a non-zero Fick diffusion flux must exactly balance the pressure diffusion flux.
Thus, a concentration gradient should appear near the bubble wall and by continuity in
the whole liquid. Since the pressure gradient reverses as the bubble oscillates, it is clear
that the concentration of each species is an oscillatory quantity, but the question also
arises of a possible average effect, building over several periods, as observed for rectified
diffusion.
In the present paper, an approximate analytical expression of the concentration field in
the liquid is sought in order to be able to draw some conclusions for a given mixture and
given bubble parameters, namely the amplitude of the driving pressure, its frequency and
the ambient radius of the bubble. Since the problem has some common characteristics
with rectified diffusion, the solution method proposed by Fyrillas & Szeri (1994) will
be used. The concentration field is cut in two parts: the oscillating field is required to
fulfill the complicated oscillatory part of the boundary condition at the bubble wall and
is non-zero only in a thin boundary layer near the bubble; the smooth field satisfies the
remaining part of the boundary condition and is uniformly valid everywhere in the liquid.
No specific assumption is made concerning bubble dynamics, apart from its periodic
Segregation of a liquid mixture by a radially oscillating bubble
3
motion, so that the solution obtained is immediately applicable once the bubble radius
is known as a function of time.
The paper is organized as follows: section 2 presents the main convection–diffusion
equation along with boundary conditions and the splitting of the problem in two parts.
In section 3, the oscillatory problem is solved and the splitting is determined unambiguously. The smooth problem is solved in section 4. In section 5, the analytical results are
first validated by comparing them to a full numerical solution of the partial differential
equation. Then, the influence of the bubble parameters on the magnitude of the segregation effect is investigated. In section 6, the model is finally applied to typical mixtures
of water with either small or large molecules and the results are discussed.
2. Formulation
2.1. Bubble motion and liquid fields
We will consider a single bubble oscillating in a liquid mixture of infinite extent, forced
by a oscillating pressure field far from the bubble p∞ (t) = p0 (1−P cos ωt), where ω is the
angular frequency, P the dimensionless forcing pressure and p0 the hydrostatic pressure.
The motion of such a bubble has been widely described in the literature since the early
work of Lord Rayleigh and several refinements of the basic model can be found, including
thermal behavior of the gas, liquid compressibility effects and liquid evaporation at the
bubble wall (see Prosperetti 1999; Brenner et al. 2002, for a recent review).
The model presented here is in itself independent of a specific choice of a bubble
dynamics model, and we defer the choice of the differential equation governing the radial
motion to section 5. However, the mass transport equation used in this work requires
analytical expressions of the velocity and pressure fields in the liquid, and for the sake
of simplicity, we will assume an iso-volume motion of the liquid. Besides, the potential
character of the flow is ensured by the spherical symmetry, and the potential φ̃ and
velocity fields ṽ can be easily obtained from mass conservation:
1 dṼ
,
4πr dt
r dṼ
,
ṽ(r, t) =
4πr3 dt
φ̃(r, t) = −
(2.1)
(2.2)
where Ṽ is the time-dependent bubble volume, and r the distance from the center of the
bubble. Then, using the unsteady Bernoulli law for potential flows between a point in
the liquid of radial coordinate r and a point infinitely far from the bubble, the pressure
field in the liquid is
!2
2
1
dṼ
1 d Ṽ
.
−
(2.3)
p̃(r, t) = p∞ (t) + ρ
4πr dt2
32π 2 r4 dt
The validity of the iso-volume assumption is questionable for strong motion of the bubble,
involving wall velocities near or greater than the speed of sound in the liquid. Accounting for liquid compressibility results in corrections of the order of the Mach number in
both the velocity and pressure field, and therefore also in the bubble dynamics equation
(Prosperetti & Lezzi 1986). The main physical consequence of liquid compressibility is
the formation of shock-waves at the end of the bubble collapse (Hickling & Plesset 1964).
Clearly, since shock-waves are in essence strong pressure gradients, neglecting their effect
on pressure diffusion may appear as a rough approximation. However, taking them into
4
O. Louisnard, F. J. Gomez and R. Grossier
account would require cumbersome expressions of the velocity and pressure fields (see for
example the second-order expressions of Tomita & Shima 1977; Fujikawa & Akamatsu
1980, obtained by the PLK strained-coordinates method). This constitutes a technical
problem especially for the velocity field: as will be seen below, the convective term in the
transport equation can be easily suppressed by a convenient change of variable in the
incompressible case. There is no evidence that a similar change of variable could be found
in the case of a compressible velocity field, which would make the problem untractable.
We therefore chose to sacrifice the compressibility hypothesis in order to draw a general
picture of the pressure diffusion effect. We will however make an exception and keep the
compressibility-induced correction terms in the bubble equation, in order to obtain a
more realistic model for the bubble dynamics. Moreover, since the shock-waves issue is
of practical interest, additional qualitative comments will be proposed in section 6.
Finally, since the liquid considered here is a mixture whose spatial homogeneity is
investigated, the average mixture density may not be constant. It may be reasonably
assumed that the occurrence of such inhomogeneities does neither significantly affect the
liquid fields, nor the bubble motion.
2.2. Mass transport
The mass conservation of a species A in a binary mixture is expressed as
∂ρA
+ ∇ · (ρA ṽ) = −∇ · jA ,
(2.4)
∂t
where ρA is the local density of species A, ṽ is the mass-averaged mixture velocity. The
mass diffusion flux jA is, taking into account pressure diffusion (Hirschfelder et al. 1967;
Bird et al. 1960)
M A MB
V̄A
1
jA = −D ρ∇ωA +
∇p̃ ,
(2.5)
−
ρωA
M RT
MA
ρ
where ωA = ρA /ρ is the mass fraction of species A, MA , MB and M are the respective
molar weight of species A, B and of the mixture, ρ is the mean density of the mixture, V̄A
the specific molal volume of species A, and R the universal gas constant. The first term
in equation (2.5) represents the Fick diffusion flux driven by a concentration gradient,
and the second is the pressure diffusion flux, driven by the local pressure gradient.
Using the mixture mass-conservation equation ∂ρ/∂t + ∇ · (ρṽ) = 0, and inserting the
flux expression (2.5) in equation (2.4), we get:
V̄A
1
MA MB
∂ωA
−
∇p̃ .
(2.6)
+ ṽ · ∇ωA = D∇ · ρ ∇ωA +
ωA
ρ
∂t
M RT
MA
ρ
Although it is tempting to simplify both sides of equation (2.6) by ρ, the latter quantity
is not constant since it depends on the local mass fraction ωA , which is space and time
dependent. The same problem arises with the appearance of the mean molar weight M
and again the density ρ in the pressure diffusion term of equation (2.6), which depends
on the local composition of the mixture. This has the strong consequence that rigorously,
equation (2.6) is non-linear. In view of the method we plan to use for the resolution of the
problem, a linearization of the problem is necessary, paying the price of some additional
assumptions. It is shown in appendix A that in the limit of a dilute mixture (ωA ≪ 1),
the mixture density ρ is approximately constant and equal to MB /V̄B and equation (2.6)
may be simplified as
∂ωA
V̄B
MA
V̄A
−
+ ṽ · ∇ωA = D∇ · ∇ωA +
ωA
∇p̃ ,
(2.7)
∂t
RT
MA
MB
Segregation of a liquid mixture by a radially oscillating bubble
5
or, in spherical coordinates:
∂ωA
∂ p̃
∂ωA
D ∂
∂ωA
2
r
+ ṽ(r, t)
= 2
+ β̃ωA (r, t) ,
∂t
∂r
r ∂r
∂r
∂r
where ṽ(r, t) and p̃(r, t) are given by (2.2),(2.3), and β̃ is:
MA V̄A
V̄B
β̃ =
−
.
RT MA
MB
(2.8)
(2.9)
This conservation equation should be completed with appropriate boundary conditions
at the bubble wall and infinitely far from the bubble. Both components are assumed nonvolatile and therefore cannot cross the bubble wall. Thus, the diffusive flux jA should be
zero at the bubble wall:
∂ p̃
∂ωA
(r = R̃(t), t) + β̃ωA (r = R̃(t), t) (r = R̃(t), t) = 0.
∂r
∂r
(2.10)
We emphasise that the expression of the diffusive flux at the bubble wall must include the
pressure diffusion term, consistently with the transport equation (2.8). It is interesting
to note that a similar boundary condition is used in centrifuge equations (Archibald
1938), to express the impermeability of the sample-tube extremities to any species. The
non-volatility of the species may appear as a drastic limitation. However, relaxing this
hypothesis would have the disadvantage to couple the diffusion problem in the liquid
with the diffusion problem of vapor through the uncondensable gas filling the bubble.
Moreover, the problem would require a liquid-vapor equilibrium condition at the bubble
wall, which may take a complex form in the case of mixtures. Finally, several related
issues such as evaporation–condensation kinetics, or chemical reactions (Storey & Szeri
2000) may further complicate the problem. We therefore leave aside these refinements
for now, and concentrate on the effects of pressure diffusion alone.
Far from the bubble, the concentration field remains undisturbed by the bubble oscillations, so that
ωA (r → ∞, t) = ωA0 ,
(2.11)
ωA (r, t = 0) = ωA0 .
(2.12)
and finally, the liquid mixture is initially assumed homogeneous in space:
2.3. Non-dimensionalization
The equations of the problems are non-dimensionalized as follows: the natural length
scale is the ambient bubble radius R̃0 , the time scale is ω −1 , the inverse of the driving
frequency. The pressure scale is set as 21 ρ0 R̃02 ω 2 , which is the dynamic pressure of the
liquid displaced by the bubble. We therefore set
r = R̃0 ξ,
t = τ /ω,
p̃ =
1 2 2
ρR̃ ω p.
2 0
The bubble instantaneous radius and volume are non-dimensionalized by their ambient
values:
4
R̃ = R̃0 R, Ṽ = π R̃03 V.
3
In the new variables, the dimensionless velocity and pressure field in the liquid are
v(ξ, τ ) =
V̇
,
3ξ
(2.13a)
6
O. Louisnard, F. J. Gomez and R. Grossier
2 V̈
1 V̇ 2
−
,
(2.13b)
3 ξ
9 ξ4
where here, and in what follows, over-dots denote time-derivatives with respect to the
dimensionless time-variable τ . The concentration of species A is non-dimensionalized by
ωA
− 1.
(2.14)
C=
ωA0
p(ξ, τ ) =
The quantity C represents the segregation level: a positive value of C expresses a local
excess of species A above the concentration at rest. The mass transport-equation becomes
!
∂C
1 1 ∂
∂p
V̇
∂C
2 ∂C
ξ
,
(2.15)
+
=
+
β(C
+
1)
∂τ
3ξ 2 ∂ξ
Pe ξ 2 ∂ξ
∂ξ
∂ξ
where Pe = R̃02 ω/D is the Péclet number, and the dimensionless number β is
1
(2.16)
β = β̃ ρR̃02 ω 2 .
2
For later use, we separate the respective contributions of the mixture and the bubble to
the dimensionless parameter β, and write
β = βm R̃02 ω 2 ,
(2.17)
where
1
1 MA V̄A MB
βm = β̃ρ =
−1 ,
2
2 RT MA V̄B
depends only on the mixture considered.
The boundary and initial conditions become, in dimensionless variables
∂p
∂C
(ξ = R(τ ), τ ) + β [C(ξ = R(τ ), τ ) + 1] (ξ = R(τ ), τ ) = 0,
∂ξ
∂ξ
C(ξ → ∞, τ ) = 0,
C(ξ, τ = 0) = 0.
(2.18)
(2.19a)
(2.19b)
(2.19c)
The intrinsic difficulties in the resolution of the above governing equations are similar
to those encountered in the rectified diffusion problem (Hsieh & Plesset 1961; Eller &
Flynn 1965; Fyrillas & Szeri 1994, 1995): on one hand, the boundary condition (2.19a)
at the bubble surface is applied at a moving boundary and is furthermore unsteady. On
the other hand, the velocity field is inhomogeneous and also unsteady. The solution to
overcome the difficulty of the moving boundary and the oscillating velocity field is to
define a Lagrangian radial coordinate, as first suggested by Plesset & Zwick (1952), by
σ = 13 (ξ 3 − V (τ )), which represents physically the dimensionless volume between the
bubble wall and the point of interest in the liquid. A specific liquid particle moves with
a constant σ (under the incompressibility hypothesis) and an observer moving with such
a particle would only see the diffusive transport of species. This may be readily seen by
expressing equation (2.15) in the Lagrangian coordinates (σ, τ ):
∂C
1 ∂
∂C
A(σ, τ )
=
+ βB(σ, τ )(C + 1) ,
(2.20)
∂τ
Pe ∂σ
∂σ
where
A(σ, τ ) = (3σ + V )4/3 ,
(2.21a)
2
V̇ 2
4
B(σ, τ ) = − V̈ +
.
3
9 (3σ + V )
(2.21b)
Segregation of a liquid mixture by a radially oscillating bubble
7
The boundary and initial conditions (2.19) become
B(0, τ )
∂C
(0, τ ) + β
(C(0, τ ) + 1) = 0,
∂σ
A(0, τ )
C(σ → ∞, τ ) = 0,
C(σ, τ = 0) = 0.
(2.22a)
(2.22b)
(2.22c)
It can be readily seen that the convective term of equation (2.15) has indeed disappeared
and that the boundary condition at the bubble wall is now applied at a fixed point,
thanks to the change of variable.
The problem defined by equations (2.20)-(2.22) shares some resemblance with the
rectified diffusion problem and thus, the splitting-method proposed by Fyrillas & Szeri
(1994, 1995, 1996) can be profitably used here. For self-completeness, we will recall here
the main lines of its underlying physical basis. For rectified diffusion, the oscillatory
gas pressure in the bubble drives the gas concentration in the neighbouring liquid in
oscillation, thus producing a periodic inversion of the concentration gradient. This rapidly
oscillating gradient is counteracted by molecular diffusion, but owing to the large value of
the Péclet number, the equilibrium cannot be restored immediately. This delay produces
a long-term average diffusion effect, on a timescale larger than the oscillation period by
a factor of the order of Pe. Therefore the concentration field varies on two time-scales.
The present problem shares this property with rectified diffusion, but here, the source
of the concentration gradient is the segregation of species by pressure diffusion, both in
the liquid bulk and at the bubble wall (see equations (2.20) and (2.22a) respectively).
Moreover, it can be easily seen by looking at equation (2.13b) or (2.21b) that pressure
diffusion is not symmetric over one oscillation period, owing to the V̇ 2 term, representing
the convective acceleration of the fluid in spherical symmetry. Because of this term,
pressure increases when traveling away from the bubble, which may be understood from
the Bernoulli law: because of spherical symmetry, velocity decreases with the distance to
the bubble and this should be balanced by a pressure increase.
For both rectified diffusion and the present problem, the existence of two time scales
justifies a multiple-scales method, but there remains a technical difficulty in the unsteady
character of the boundary conditions at the bubble wall. If the multiple-scales method
were to be applied directly to the set of equations (2.20),(2.22), one would be met with an
impossibility for the solution at leading order to fulfill the oscillating boundary condition
(2.22a). This difficulty is overcome by splitting the concentration field in two parts: an
oscillatory part, which satisfies the oscillating part of the boundary condition, and a
smooth part, to which the remaining part of the boundary condition should be ascribed.
The oscillating part represents physically the perturbation of the concentration field
imposed by the bubble wall forcing term, and is designed to differ from zero only in a
boundary layer of thickness Pe −1/2 . The smooth part extends in the whole liquid and
varies on both the oscillation timescale 1/ω and a slow timescale of order Pe/ω.
2.4. Splitting of the problem
We set the concentration field as C(σ, τ ) = C osc (σ, τ ) + C sm (σ, τ ), where C osc is the
oscillatory part, and C sm the smooth part. We then split the governing equations into
an oscillatory part and a smooth part. The oscillatory problem is defined by
∂C osc
1 ∂
∂C osc
A(σ, τ )
=
+ βB(σ, τ )C osc ,
(2.23a)
∂τ
Pe ∂σ
∂σ
∂C osc
(0, τ ) + βH(τ ) [C osc (0, τ ) + 1] = −G − βH(τ )C sm (0, τ ),
(2.23b)
∂σ
8
O. Louisnard, F. J. Gomez and R. Grossier
and the smooth problem is
∂C sm
1 ∂
∂C sm
sm
=
+ βB(σ, τ ) (C + 1) ,
A(σ, τ )
∂τ
Pe ∂σ
∂σ
∂C sm
(0, τ ) = G.
∂σ
(2.24a)
(2.24b)
The constant G is introduced to add a degree of freedom in the separation process and
will be determined unambiguously by using a splitting condition, to be defined in the
next section. In the boundary condition (2.23b), the function H(τ ) is defined by
H(τ ) =
B(0, τ )
,
A(0, τ )
(2.25)
and represents the dimensionless pressure gradient at the bubble wall.
Finally, both oscillatory and smooth fields are required to fulfill the boundary condition
far from the bubble (2.22b) and the initial condition (2.22c), so that
C sm (σ → ∞, τ ) = C sm (σ, 0) = 0,
C
osc
(σ → ∞, τ ) = C
osc
(2.26)
(σ, 0) = 0.
(2.27)
3. The oscillatory problem
Following Fyrillas & Szeri (1995), we use a matched asymptotic expansion to solve the
oscillatory problem: the inner solution must fulfill the bubble wall boundary condition
(2.23b) while the outer solution is required to be identically zero. To determine the
inner approximation of the oscillatory solution, we define a re-scaled Lagrangian spacecoordinate by s = Pe 1/2 σ. Furthermore, we use the nonlinear time τ̂ first suggested by
Plesset & Zwick (1952), which arises from the spherical symmetry of the problem:
Z τ
R4 (τ ′ ) dτ ′ ,
(3.1)
τ̂ =
0
and for further use, we also define the nonlinear period
Z 2π
T̂ =
R4 (τ ′ ) dτ ′ .
(3.2)
0
Taking τ̂ as the new time-variable, equation (2.23a) becomes
∂C osc
∂C osc
∂
1
′
osc
′
βB (s, τ̂ ; Pe)C
,
A (s, τ̂ ; Pe)
=
+
∂ τ̂
∂s
∂s
Pe 1/2
(3.3)
where
A′ (s, τ̂ ; Pe) = R−4 A(Pe −1/2 s, τ ) =
1+
B ′ (s, τ̂ ; Pe) = R−4 B(Pe −1/2 s, τ ) = −
1
3s
Pe 1/2 V
4/3
2 V̈
4 V̇ 2
+
3 V 4/3
9 V 7/3
,
(3.4a)
1
1
,
(3.4b)
∂C osc
(0, τ̂ ) + βH(τ̂ ) [C osc (0, τ̂ ) + 1] = −G − βH(τ̂ )C sm (0, τ̂ ).
∂s
(3.5)
1+
3s
1/2 V
Pe
and the bubble wall condition reads
Pe 1/2
Segregation of a liquid mixture by a radially oscillating bubble
9
The outer limit of the inner approximation should match the outer approximation which
is identically zero, so that C osc (s, τ̂ ) should satisfy
lim C osc (s, τ̂ ) = 0.
(3.6)
s→∞
We now assume an asymptotic expansion for C osc in the Pe −1/2 parameter:
C osc (s, τ̂ ) = C0osc (s, τ̂ ) +
1
Pe
1/2
C1osc (s, τ̂ ) +
1 osc
C (s, τ̂ ) . . . ,
Pe 2
(3.7)
and we also expand functions A′ and B ′ given by equations (3.4), as well as the separation
constant G appearing in equations (2.23b) and(2.24):
A′ (s, τ̂ ) = 1 + Pe −1/2 A′1 (s, τ̂ ) + Pe −1 A′2 (s, τ̂ ) . . . ,
B ′ (s, τ̂ ) = B0′ (τ̂ ) + Pe −1/2 B1′ (s, τ̂ ) + Pe −1 B2′ (s, τ̂ ) . . . ,
G = G0 + Pe −1/2 G1 + Pe −1 G2 . . . .
A hierarchy of inhomogeneous diffusion problems is obtained, all sharing the same form.
The general solution of such problems is detailed in appendix B, which also yields a
splitting-condition necessary to ensure the matching equation (3.6). It is interesting to
note that in the present case, the oscillatory problem at each order has a Neumann boundary condition, whereas the oscillatory problems in the analysis of surfactants-enhanced
rectified diffusion by Fyrillas & Szeri (1995) involve a Dirichlet boundary condition. The
difference arises from the presence of the Péclet number in the boundary condition in the
latter problem (see equation (2.4) in Fyrillas & Szeri 1995), while the boundary condition
(2.22a) in the present problem is Péclet independent. This is because both pressure and
Fick diffusion terms are proportional to the diffusion coefficient, which thus cancels out
in the null total flux condition (2.10) at the bubble wall.
For further use, it is useful to note that the H function given by (2.25) can also be
expressed in the following forms
2 V̈
4 V̇ 2
R̈
+
= −2 2 .
4/3
3V
9 V 7/3
R
We now turn to solve the oscillatory problems at each order.
H(τ̂ ) = B0′ (τ̂ ) = −
(3.8)
3.1. Zeroth-order
The oscillatory problem at order 0 is
∂C0osc
∂ 2 C0osc
=
,
∂ τ̂
∂s2
∂C0osc
(0, τ̂ ) = 0,
∂s
C0osc (s → ∞) = 0.
(3.9a)
(3.9b)
(3.9c)
The solution is clearly the null one. This can be easily understood on a physical basis as
pressure diffusion does not act to this order, neither in the liquid bulk, nor at the bubble
wall, as can be seen in (3.9a). Therefore the liquid mixture is only submitted to classical
molecular diffusion. Only a non-homogeneous boundary condition could produce a concentration gradient which is not the case, since to this order, the bubble wall condition
only imposes a zero concentration gradient. This is why, contrarily to rectified diffusion
problems (Fyrillas & Szeri 1994, 1995), the zeroth-order oscillatory solution is zero in
the present problem.
10
O. Louisnard, F. J. Gomez and R. Grossier
3.2. First-order
At order 1, using the nullity of
C0osc (0, τ̂ ),
we obtain:
∂ 2 C1osc
∂C1osc
=
,
∂ τ̂
∂s2
(3.10a)
∂C1osc
(0, τ̂ ) + βH(τ̂ ) = −G0 − βH(τ̂ )C0sm (0, τ̂ ),
∂s
(3.10b)
C1osc (s → ∞) = 0.
(3.10c)
G0 = −β hH(τ̂ ) [C0sm (0, τ ) + 1]iτ̂ .
(3.11)
The splitting-condition (B 7) obtained in appendix B yields the separation constant G0 :
The part of the boundary condition ascribed to C1osc is therefore
∂C1osc
(0, τ̂ ) = β [hH(τ̂ )iτ̂ − H(τ̂ )] [C0sm (0) + 1] ,
∂s
where we have used the result, to be demonstrated in section 4, that C0sm is independent
of the fast time-variable τ̂ . The asymptotic solution C̄1osc (s, τ̂ ) of equations (3.10 a–c)
can be obtained from appendix B: expanding H(τ̂ ) as a Fourier series,
H(τ̂ ) = hH(τ̂ )iτ̂ +
m=+∞
X
τ̂
hm exp 2imπ
,
T̂
m=−∞
(3.12)
m6=0
and using equation (B 8), the oscillatory concentration field is
C̄1osc (s, τ̂ )
! 12
T̂
=β
+ 1]
2π
"
1 #
m=+∞
X
|m|π 2
hm
π
τ̂
− (ǫm i + 1)
s , (3.13)
×
exp i 2πm − ǫm
4
|m|1/2
T̂
T̂
m=−∞
[C0sm (0)
m6=0
where ǫm = sgn(m). It is interesting to note that the first order oscillatory solution C̄1osc
depends on the boundary value of the zeroth-order smooth solution C0sm (0), which is to
be determined in the next section.
3.3. Second-order
The second-order oscillatory problem allows the determination of the separation constant
G1 , which, as will be seen below, is enough to solve the smooth problem up to order Pe −1 .
The calculation is detailed in appendix C and yields
G1 = −β hH(τ̂ ) [C1sm (0, τ )]iτ̂ .
(3.14)
C2osc
The second-order oscillatory field
could also be obtained analytically by using appendix B, but is not required in the present analysis.
4. The smooth problem
To treat the smooth problem, time is first rescaled by defining the slow time-variable
λ = τ /Pe, and the smooth field C sm is considered as a function of both fast and slow
Segregation of a liquid mixture by a radially oscillating bubble
11
time-variables, respectively τ and λ. The smooth equation (2.24a) reads, in the new
variables
∂C sm
∂C sm
1 ∂C sm
1 ∂
A(σ, τ )
+
=
+ βB(σ, τ ) (C sm + 1) .
(4.1)
∂τ
Pe ∂λ
Pe ∂σ
∂σ
The smooth field C sm (σ, τ, λ) is next expanded in the small parameter Pe −1/2 :
C sm (σ, τ, λ) = C0sm (σ, τ, λ) +
1
Pe
1/2
C1sm (σ, τ, λ) +
1 sm
C (σ, τ, λ) + . . . ,
Pe 2
(4.2)
which, once introduced in equation (4.1), yields a hierarchy of equations in the small
parameter Pe −1/2 . As in Fyrillas & Szeri (1995), the zeroth- and first-order smooth
equations read simply:
∂C0sm
= 0 ⇒ C0sm (σ, λ)
∂τ
∂C1sm
= 0 ⇒ C1sm (σ, λ)
∂τ
(4.3a)
(4.3b)
which indicates that C0sm and C1sm vary with time only through the slow time-scale λ.
The dependance of these two fields on σ and λ can be obtained by writing the problems
at orders 2 and 3, and using a non-secularity condition. The smooth boundary condition
(2.24b) is written at each order by using the expressions (3.11) and (3.14) of the separation
constants G0 and G1 , and asymptotic solutions for λ → ∞ are sought. The technical
details of the calculation can be found in appendix D (see also Fyrillas & Szeri 1994,
1995). The resulting asymptotic zeroth-order smooth field reads
Z ∞
hB(σ ′ , τ )iτ
′
sm
dσ − 1,
(4.4)
C0,∞ (σ) = exp β
hA(σ ′ , τ )iτ
σ
sm
while the asymptotic first-order smooth field C1,∞
is zero, so that equation (4.4) represents in fact the asymptotic smooth solution up to order 1/Pe.
5. Numerical results
5.1. Bubble dynamics
We will consider hereafter the case of an argon bubble in a mixture of water and some
other species at ambient temperature T = 298 K. The temporal evolution of the bubble radius is calculated by solving the Keller-Miksis equation (Keller & Miksis 1980;
Hilgenfeldt et al. 1996):
d2 R̃
R̃ 2
dt
!
!2
!
1 dR̃
3 dR̃
1 dR̃
1−
+
1−
=
c dt
2 dt
3c dt
(
!
)
1 dR̃
R̃ dp̃g
2σ 4µ dR̃
1
1+
[p̃g − p0 (1 − P cos ωt)] +
, (5.1)
−
−
ρ
c dt
c dt
R̃
R̃ dt
where R̃ is the bubble radius, p̃g the gas pressure in the bubble, assumed homogeneous,
c = 1500 m s−1 , ρ = 998 kg m−3 , µ = 10−3 kg m−1 s−1 are respectively the sound velocity,
density and dynamic viscosity of water, p0 = 101325 Pa the pressure in the liquid at rest,
σ = 0.072 N m−1 the water-gas surface tension, P the dimensionless driving pressure
amplitude and ω the angular driving frequency.
12
O. Louisnard, F. J. Gomez and R. Grossier
Two different models can be used for the bubble interior. The first assumes an isothermal behaviour and a van der Waals equation of state, so that the bubble internal pressure
is
!
2σ
R̃03 − h3
p̃g = p0 +
,
(5.2)
R̃0
R̃3 − h3
where R̃0 is the ambient radius of the bubble and h the van der Waals hard-core radius.
A refined model accounting for water evaporation at the bubble wall and temperature
gradients in the bubble was also used. The details of the model can be found elsewhere
(Toegel et al. 2000; Storey & Szeri 2001). It is known that accounting for such effects
reduces the violence of the collapse and may therefore influence the segregation process
investigated in this paper, as will be seen below.
In the following sections, equation (5.1) will be solved for various sets of parameters
ω, P and R̃0 , over a number of periods sufficiently large to get steady-state oscillations.
The corresponding bubble volume and its time-derivatives on the last period are stored
in tables, from which the time and space dependent coefficients A(σ, τ ) and B(σ, τ ) can
be calculated by equations (2.21 a,b) when needed.
5.2. Comparison with full simulation
In order to check the validity of the approximation obtained from the splitting method,
numerical simulations of the full convection-diffusion problem (2.20)-(2.22) have been
performed, with the help of the FEMLAB software. The present set of equations is
recast without further difficulty in the canonical coefficient form of partial differential
equations allowed in FEMLAB. The interval [0, ∞] was mapped to [0, 1] by using the
variable change x = 1/(σ + 1), the interval [0, 1] was non-uniformly meshed to trap the
boundary layer near the bubble wall, and mesh convergence studies were performed to
ensure good accuracy of the result.
In order to test the analytical approximation obtained in the preceding section, we
first recall that the analytical method yields the concentration field as
1
1
1
sm
osc
sm
C (σ, λ) +
C (σ, τ ) + O
,
C(σ, τ ) = C0 (σ, λ) +
1/2 1
1/2 1
Pe
Pe
Pe
sm
and as
since C0osc = 0. For large times (λ → ∞), C0sm reaches its asymptotic limit C0,∞
sm
osc
shown above, C1 vanishes. The oscillatory field C1 should reach its asymptotic value
(3.13) in a few periods, and therefore one should have
1
1
osc
sm
.
(5.3)
C(σ, τ ) ∼ C0,∞ (σ) +
C̄ (σ, τ ) + O
1/2 1
λ→∞
Pe
Pe
Further averaging on time τ̂ over one period, we get
hC(σ, τ )iτ̂
∼
λ→∞
sm
C0,∞
(σ)
+O
1
Pe
,
(5.4)
since from equation (3.13), C̄1osc has a null τ̂ -average.
Both equations (5.3) and (5.4) were checked against direct numerical simulation for
an argon bubble of ambient radius R0 = 4 µm driven by pressure fields of dimensionless
amplitudes P = 0.3, 0.6 and 0.8 and frequency 26.5 kHz. Since our aim is to check the
analytical model against a numerical result, we take an arbitrary value β = −10−5 rather
than specifying the species A mixed with water. In order to reach the limit λ → ∞
numerically, the final time of the simulation was chosen sufficiently large so that the
system nearly reaches its steady state. The analysis of the smooth problem shows that
Segregation of a liquid mixture by a radially oscillating bubble
13
its steady state should be obtained within a number of periods of the order of Pe. We
therefore chose arbitrarily Pe = 100 and Pe = 500 in order to get reasonable simulation
times. We found in our examples that no noticeable change from one period to the
following one could be observed after about 2Pe periods. The last oscillation period
of the concentration field C num (σ, τ ) obtained numerically was stored, the nonlinear
time τ̂ was calculated, and the nonlinear-average hC num (σ, τ )iτ̂ was calculated over one
sm
(σ) was evaluated by
period at each spatial point σ. The smooth concentration field C0,∞
calculating the integral in equation (4.4) with a Gauss-Jacobi method (see Louisnard &
Gomez 2003, appendix B for details). The asymptotic oscillatory field was calculated from
equation (3.13), after evaluating the Fourier coefficients hm of H(τ̂ ) by a fast Fouriertransform.
Figure 1 shows typical concentration profiles results for a 4 µm argon bubble in water,
driven by an oscillatory pressure of 0.6 bar amplitude and 26.5 kHz frequency, and
Pe = 500. The dashed lines represent the analytical predictions and the solid ones are
the numerical results. The total concentration profile (thin lines) is drawn at four distinct
phases of the acoustic period in order to check equation (5.3). It is seen that the analytical
predictions are in excellent agreement with the numerical result. We also display in
figure 1 the average hC num (σ, τ )iτ̂ (thick solid line) along with the analytical prediction
sm
C0,∞
(σ) (thick dashed line). It can be seen that both quantities are in excellent agreement
(see the magnification in the inset) and we conclude that equation (5.4) is fulfilled.
Besides, it is expected that the analytical approximation would progressively break as
the asymptotic parameter Pe −1/2 increases. Calculations with a smaller Péclet number
(Pe = 100, not presented here), show that this is indeed the case, and yielded a maximum
relative error on the oscillatory field amounting to 11%.
Another validation of the model can be seen in figure 2, which compares the oscillatory
sm
(0) to the analytpart of the numerical solution at the bubble wall C num (0, τ ) − C0,∞
−1/2 osc
C̄1 (0, τ ) over one period of oscillation, for a driving pressure of
ical solution Pe
amplitude P = 0.8, and Pe = 100: here again, the two results are in excellent agreement.
5.3. Parameter-space exploration
The validation of the analytical model being achieved, we now turn to investigate how
the smooth and oscillatory parts vary with the bubble parameters (R0 , P, ω). In order to
get an immediate view of the magnitude of the segregation process, we will focus on the
values of the two fields at the bubble wall.
5.3.1. Smooth part
sm
(0) is obtained by setting σ = 0 in
The smooth concentration at the bubble wall C0,∞
equation (4.4):
sm
C0,∞
(0) = exp(βI) − 1,
(5.5)
where
I=
Z
0
∞
hB(σ ′ , τ )iτ
dσ ′ .
hA(σ ′ , τ )iτ
(5.6)
The value of integral I depends only on the bubble dynamics, and in order to get a
picture independent of the choice of a specific mixture, but containing all the bubble
data, we use the definition (2.17) of the parameter β to obtain
sm
C0,∞
(0) = exp(βm R̃02 ω 2 I) − 1,
14
O. Louisnard, F. J. Gomez and R. Grossier
−7
0.5
x 10
0
τ =0
−0.5
−1
−1.5
τ = 3π/2
−7
C
x 10
−2
0
τ =π
−2.5
−1
−3
−2
τ = π/2
−3.5
−2
10
−4
−2
10
−1
10
0
σ
10
−1
10
0
10
1
10
1
10
2
10
2
10
Figure 1. Comparison between the full numerical solution and the analytical approximation
for a 4 µm argon bubble in water in a 26.5 kHz acoustic field of amplitude P = 0.6. The Péclet
number is 500 and the parameter β is −10−5 . Thin solid lines: concentration profiles C num (σ, τ )
obtained by numerical simulation at different phases of the bubble oscillation. Thin dashed
sm
line: analytical predictions C0,∞
(σ) + Pe −1/2 C̄1osc (σ, τ ). Thick solid line: nonlinear numerical
average of the numerical profile over one period. Thick dashed line: asymptotic zeroth-order
sm
smooth concentration profile C0,∞
(σ). The inset shows a more detailed comparison between the
numerical average and the smooth solution.
where βm , defined by equation (2.18) depends only on the mixture considered. Thus, the
sm
value of R̃02 ω 2 I will be calculated for various bubble parameters and C0,∞
(0) can then
be easily deduced for a specific mixture.
Figure 3 represents R̃02 ω 2 I as a function of the driving pressure, for different ambient
radii and different frequencies. The four bottom curves are calculated for a frequency of
26.5 kHz, for bubble ambient radii ranging from 2 µm to 5 µm. It can be seen that R̃02 ω 2 I
increases with R̃0 in the range considered. The three circles represent the value obtained
from FEMLAB direct simulations, showing again the good agreement with analytical
results. The two top curves are calculated for a 4 µm bubble excited respectively at 50
kHz (dotted line) and 100 kHz (+ signs): it can be seen that the mean segregation process
increases markedly with frequency for small driving pressures, but that all curves merge
for high driving pressures.
It can be noted that in all cases, a marked increase of R̃02 ω 2 I occurs near P = 1
which is approximately the Blake threshold (Akhatov et al. 1997; Hilgenfeldt et al. 1998;
Louisnard & Gomez 2003). Above this driving pressure value, the bubble dynamics becomes inertially driven, yielding large time-variations of V (t) and its time-derivatives,
and therefore large values of the integrand in equation (5.6).
15
Segregation of a liquid mixture by a radially oscillating bubble
−6
1
x 10
C1osc(0, τ )/Pe 1/2
0.5
0
−0.5
−1
−1.5
0
0.2
0.4
τ /T
0.6
0.8
1
Figure 2. Comparison between the analytical oscillatory concentration (dashed line)
sm
Pe −1/2 C̄1osc (0, τ ) at the bubble wall and the numerical solution C num (0, τ ) − C0,∞
(0) (solid
line) for a 4 µm argon bubble driven at P = 0.8 and 26.5 kHz. The Péclet number is 100 and
the parameter β is −10−5 .
5.4. Parameter-space exploration : oscillatory part
Neglecting terms of order O(Pe −1 ), the oscillatory concentration at the bubble wall
C osc (0, τ̂ ) reduces to C̄1osc (0, τ̂ )/Pe 1/2 . Evaluating equation (3.13) at s = 0, we get
C osc (0, τ̂ ) =
β
Pe 1/2
[C0sm (0) + 1] G(τ̂ ),
(5.7)
where
G(τ̂ ) =
T̂
2π
! 12
m=+∞
X
hm
π
τ̂
exp i 2πm − ǫm
.
4
|m|1/2
T̂
m=−∞
(5.8)
m6=0
Using equation (2.17) to express the factor β/Pe 1/2 in terms of the dimensional parameters, equation (5.7) becomes
C̄1osc (0, τ̂ ) = βm D1/2 R̃0 ω 3/2 [C0sm (0) + 1] G(τ̂ ).
(5.9)
In order to identify the contribution of the bubble oscillations independently from the
choice of a specific mixture, the quantity R̃0 ω 3/2 G(τ̂ ) must be calculated for various bubble parameters. For small driving pressure, G(τ̂ ) can be evaluated by summing the series
(5.8) without any specific problem, as was done in section §5.2. However, for driving
pressures high enough to yield inertial cavitation, evaluation of G(τ̂ ) is subject to a technical difficulty linked to the shape of function H(τ̂ ) = −2R̈/R2 , as shown in figure 4(a).
16
O. Louisnard, F. J. Gomez and R. Grossier
2
10
1
R̃02ω 2I( m2 s−2)
10
0
10
−1
10
−2
10
−3
10
−4
10
0.2
0.4
0.6
0.8
P
1
1.2
1.4
1.6
Figure 3. Evolution of R̃02 ω 2 I with driving pressure, from equation (5.6). The four bottom
curves are calculated with f =26.5 kHz, for ambient radii R̃0 =2 µm (thick solid line), 3 µm
(dashed line), 4 µm (thin solid line), 5 µm (dash-dotted line). The two top curves are calculated
for R̃0 =4 µm and respectively with f =50 kHz (dotted line) and f =100 kHz (+ signs). The
three circles represents the results obtained by FEMLAB full simulations for 4 µm bubbles
driven respectively by pressure fields of 0.3, 0.6 and 0.8 driving pressure.
Owing to the huge outward acceleration of the liquid at the end of the bubble collapse,
H(τ̂ ) looks like a series of negative Dirac distributions, the most important being located
at the main collapse, and the other ones at each secondary collapse between the bubble
afterbounces. From the singular shape of function H(τ̂ ), it is expected that its Fourier
spectrum (the coefficients hm in the series (3.12)) spans over a wide frequency range.
Therefore series (5.8) converges very slowly, thus forbidding any numerical estimation.
This is illustrated in figure 4(b), which shows a magnification of the most negative peak
of H(τ̂ ). It is seen that the width of the peak is less than 9 orders of magnitude the
nonlinear period of oscillation, so that one should sum more than 109 terms in the series
to obtain an acceptable result !
We therefore used the following trick: let’s denote τ̂min the time at which H(τ̂ ) reaches
its highest negative peak amplitude Hmin . We fit H by a negative tooth function of
amplitude Hmin and width ∆τ̂ :
τ̂ − τ̂min
Hmin 1 −
, τ̂ ∈ [τ̂min − ∆τ̂ , τ̂min + ∆τ̂ ]
H(τ̂ ) ≃
.
(5.10)
∆τ̂
0,
elsewhere
where ∆τ̂ is determined in such a way that the real and fitted peaks have the same
integral over the interval [τ̂1 , τ̂2 ], where τ̂1 and τ̂2 are the locations of the zeros of H(τ̂ )
17
Segregation of a liquid mixture by a radially oscillating bubble
8
x 10
0
∆τ̂
H
−1
(a)
(b)
−2
−3
0
40
80
τ̂min
−0.5
0
0.5
1
8
10
(τ̂ − τ̂min)/T̂ × 109
τ̂
G/|Gmin|
0
−0.5
(c)
−1
−2
0
2
4
6
9
(τ̂ − τ̂min)/T̂ × 10
Figure 4. (a) Time evolution of function H(τ̂ ) for a 4 µm bubble driven at P = 1.1 and
26.5 kHz. The nonlinear period is 122 in this case. The negative peaks corresponds to the huge
positive values taken by V̈ at the main collapse and subsequent afterbounces. (b) Zoom on the
most negative peak of H(τ̂ ) (solid line) translating the origin of abscissas to the location of this
peak. It is seen that the width of the peak is 9 orders of magnitude smaller than the nonlinear
period; the dashed line is the approximation of H(τ̂ ) defined by equation (5.10) (c) Solid line:
Shape of function G(τ̂ ) given by equation (5.8). Dashed line: shape of function H(τ̂ ).
at each side of τ̂min :
∆τ̂ =
1
Z
τ̂2
H(τ̂ ) dτ̂ .
(5.11)
Hmin τ̂1
∆τ̂ represents physically the characteristic time (in nonlinear form) of the bubble rebound
at the end of the collapse. Figure 4(b) shows the original function H(τ̂ ) (solid line)
compared to the approximation obtained by equation (5.10) (dotted line).
The Fourier coefficients of the tooth function can be easily calculated and introduced
in equation (5.8) to calculate G(τ̂ ). This is done in appendix E and the following approximation of G(τ̂ ) is obtained:
!1/2
m=+∞
X
∆τ̂
1
T̂
T̂
π
τ̂ − τ̂min
2
sin mπ
.
Hmin 2
Gapp (τ̂ ) =
−
cos 2mπ
2π
π ∆τ̂ m=1 m5/2
4
T̂
T̂
(5.12)
For very small ∆τ̂ , which is the case for inertial cavitation, this series is as difficult
to calculate as the original one in equation (5.8). However, a good approximation of
Gapp (τ̂ ) can be found and is detailed in appendix E (see equations (E 3) and (E 12)).
18
O. Louisnard, F. J. Gomez and R. Grossier
Figure 4(c) shows the typical shape of Gapp (τ̂ ): it decreases rapidly down to a minimum
located slightly after the minimum of H and then slowly relaxes to 0. We first restrain
our primary interest to the extremal value attained by C̄1osc (0, τ̂ ) over one period, so
that only the minimum value of Gapp (τ̂ ) is needed. It is shown in appendix E that an
excellent estimate of this minimum is
Gmin
app =
8Γ (1/2)
√
Hmin ∆τ̂ 1/2 .
3 3π
(5.13)
The solid line in figure 5 displays the evolution of R̃0 ω 3/2 Gmin obtained by summing
directly series (5.8) along with a 215 points FFT of H(τ̂ ), while the dashed line represents
R̃0 ω 3/2 Gmin
app calculated from the approximate equation (5.13) for a 4 µm argon-bubble
oscillating at 26.5 kHz in water. It is seen that both results are in agreement up to about
P = 1.05 ( which corresponds approximately to the Blake threshold) and that they
markedly diverge above the threshold, which demonstrates that for inertial motion of
the bubble, H(τ̂ ) becomes too sharp to be correctly represented by a reasonable Fourier
expansion. Therefore, in the inertial regime, the approximate equation (5.13) must be
used to calculate Gmin .
The dash-dotted line in figure 5 also displays the value of R̃0 ω 3/2 Gmin
app calculated
from equation (5.13) but with a refined bubble interior model, taking into account heat
transport and water condensation/evaporation at the bubble wall (Toegel et al. 2000;
Storey & Szeri 2001). At low driving pressures, the results are comparable, but above
the Blake threshold, the refined model predicts values lower by one order of magnitude.
Such a result could be expected since it is known that taking into account heat transport
in the bubble interior yields a less violent collapse than with the isothermal model, and
therefore decreases the amplitude of function H. Similar conclusions have been drawn for
other bubble phenomena directly linked to the violence of the collapse, such as Rayleigh–
Taylor shape instabilities (Lin et al. 2002). Since the refined model is believed to be more
realistic than the isothermal one, it will be used in every result presented hereafter.
Figure 6 displays the influence of frequency on the oscillatory field. It is seen that as
frequency increases, we get a stronger effect at low amplitude but, for high amplitudes,
increasing the frequency reduces the oscillatory segregation effect, despite the ω 3/2 scaling
law. This can be easily explained by the fact that increasing the frequency limits the
expansion phase of the bubble in the inertial regime, which in turn reduces the violence
of the collapse, and therefore the peak value attained by the H function.
Finally, a more practical sense can be given to the time-interval ∆τ̂ appearing in
equation (5.13): since H(τ̂ ) = −2R̈/R2 , and using the definition (3.1) of the nonlinear
time, equation (5.11) can also be expressed as
Z τ2
−2
R2 (τ )R̈(τ ) dτ.
∆τ̂ =
Hmin τ1
Times τ1 and τ2 are located respectively closely before and closely after the time at which
the bubble reaches its minimum radius. Therefore R stays close to Rmin in the interval
2
. Therefore:
[τ1 , τ2 ], so that Hmin ≃ −2R̈max /Rmin
∆τ̂ ≃
4
Rmin
(Ṙ(τ2 ) − Ṙ(τ1 ))
,
R̈max
and since by definition τ1 and τ2 are the zeros of R̈, Ṙ(τ1 ) and Ṙ(τ2 ) are the minimum
and maximum bubble velocities attained before and after the rebound respectively, which
are in fact the minimum and maximum velocities of the bubble over one acoustic period.
19
Segregation of a liquid mixture by a radially oscillating bubble
10
R̃0ω 3/2 Gmin ( m s−3/2)
10
10
10
10
10
10
10
16
14
12
10
8
6
4
2
0.9
1.1
1.3
1.5
P
˛
˛
Figure 5. Solid line: evolution of R̃0 ω 3/2 ˛Gmin ˛ calculated by summing the series in equation
˛
˛
˛
(5.8) from a 215 points FFT of H(τ̂ ). Dashed line: R̃0 ω 3/2 ˛Gmin
app calculated from (5.13). Both
curves are obtained for a 4 µm argon-bubble excited at 26.5 kHz assuming
an˛ isothermal gas
˛
˛
behaviour. The dash-dotted line also represents the evolution of R̃0 ω 3/2 ˛Gmin
app , but calculated
with the refined model of the bubble interior. The values obtained are about one order of
magnitude smaller than with the isothermal model.
Thus:
Ṙmax − Ṙmin
.
(5.14)
R̈max
2
Injecting this value in equation (5.13), and setting Hmin ≃ −2R̈max /Rmin
, we get
h
i
1/2
8Γ (1/2)
√
Gmin
,
(5.15)
R̈max (Ṙmax − Ṙmin )
app ≃ −
3 3π
4
∆τ̂ ≃ Rmin
which can easily be evaluated once the bubble dynamics is known. Figure 7(a) shows
that equation (5.15) gives a reasonable approximation of Gmin
app .
Apart from the minimum value reached by G, it is also of interest to obtain an order of
magnitude of the relaxation time of G (see solid line figure 4c). It is shown in appendix E
that G reaches one tenth of its minimum value after a relaxation time of 42∆τ̂ past τ̂min .
Evaluating ∆τ̂ from equation (5.14), a ready-to-use estimate of the oscillatory segregation
duration can be obtained. The dimensional rebound time ∆t corresponding to ∆τ̂ can
4
be obtained by first converting the latter in linear time by ∆τ ≃ ∆τ̂ /Rmin
and setting
∆t = ∆τ /ω. We obtain:
∆t ≃
1 Ṙmax − Ṙmin
.
ω
R̈max
(5.16)
20
O. Louisnard, F. J. Gomez and R. Grossier
10
10
9
R̃0ω 3/2 Gmin ( m s−3/2)
10
8
10
7
10
6
10
5
10
4
10
3
10
2
10
0.9
1.1
1.3
1.5
P
3/2
˛ min ˛
˛Gapp ˛ calculated from (5.13), for a 4 µm argon-bubble excited at
Figure 6. Evolution of R̃0 ω
26.5 kHz (solid line), 50 kHz (dashed line) and 100 kHz (dash-dotted line). The bubble interior
refined model was used in all cases.
Figure 7(b) displays the dimensional rebound characteristic time ∆t in ns (solid line)
for a 4 µm argon bubble at 26.5 kHz: it rapidly drops from about 300 ns for P = 1 to
10 ps for P = 1.5. For practical applications, the dashed line represents 42∆t during
which the oscillatory segregation stays larger than one tenth of its maximal value. This
is a valuable result, if one wishes to compare the segregation duration to a characteristic
time of some process likely to be enhanced by species segregation.
6. Application and discussion
The above results should now be applied to real binary mixtures to assess the importance of the phenomenon. Rather than selecting specific mixtures, we will try to cover a
wide range of molecule sizes by taking typical values for the other mixture parameters.
We first combine equations (2.14), (5.5) and (5.7) to obtain the segregation ratio at
the bubble wall
β
ωA (0, τ̂ )
G(τ̂ ) .
(6.1)
= exp(βI) 1 +
ωA0
Pe 1/2
Having practical applications in view, we are interested in the average and peak concentrations at the bubble wall, so that in what follows, we will calculate the two quantities:
Ωm = exp(βI),
∆Ωm = exp(βI)
(6.2a)
β
Pe 1/2
Gmin ,
(6.2b)
21
Segregation of a liquid mixture by a radially oscillating bubble
10
10
4
10
(a)
9
3
10
8
2
10
10
7
∆t (ns)
R̃0ω 3/2 Gmin ( m s−3/2)
10
1
10
10
6
0
10
10
5
−1
10
10
4
10
(b)
−2
1
1.1 1.2 1.3 1.4 1.5
10
1
P
1.1 1.2 1.3 1.4 1.5
P
˛
˛
˛
Figure 7. (a) Comparison of R̃0 ω 3/2 ˛Gmin
app evaluated from equation (5.13) (solid line) and from
equation (5.15) (dashed line) for a 4 µm argon-bubble excited at 26.5 kHz. (b) Characteristic
time ∆t of the bubble rebound for the same bubble, calculated from equation (5.16) (solid line).
The dashed line represents 42∆t which is the time necessary for the oscillatory segregation to
reach one tenth of its maximum value.
where Gmin < 0 is calculated from equation (5.13). The two quantities Ωm and ∆Ωm
should be interpreted as follows: the first is the average concentration at the bubble
wall and the second is the maximum algebraic variation of the concentration around the
average, over an oscillation period.
Before specifying the mixture, it is worth recalling that β depends on two physical
properties (see equation (2.18)). On one hand the relative densities of species A and
the host liquid, on the other hand the molar weight of species A. The latter may vary
in a much larger range than the former, so that in what follows, we will study the
predictions of the model for a mixture of water with a heavier species A of apparent
density ρA = MA /V̄A = 2000 kg m−3 , and molecular weights MA ranging from 100 to
107 Da (the symbol Da stands for “Dalton” and 1 Da corresponds to a molar weight of
1 g mol−1 ).
The temperature of the mixture is set to T = 298 K. Consistently with the dilute
mixture hypothesis detailed in appendix A, the density ρ of the mixture is approximated
by the density of pure water. To calculate the oscillatory part under the same conditions,
the additional data of the diffusion coefficient is needed. Since we consider a set of species
which molecular weight varies over a very wide range, the influence of the molecular size
on the diffusion coefficient D should be taken into account. Following the Stokes-Einstein
theory, the diffusion coefficient can be expressed as
D=
kB T
,
6πµRA
(6.3)
22
O. Louisnard, F. J. Gomez and R. Grossier
where kB is the Boltzmann constant and RA the hydrodynamic radius, estimated from
the molecular weight and apparent density of species A by
4 3
Na πRA
ρA = M A ,
3
(6.4)
where Na is the Avogadro number. Under these conditions, the parameter β̃m defined by
(2.18) ranges from −10−5 to −1 s2 m−2 , the hydrodynamic radius from 0.27 to 12.5 nm,
and the diffusion coefficient from 7.9 × 10−10 to 1.7 × 10−11 m2 s−1 .
We consider the case of a 4 µm argon bubble excited at f = 26.5 kHz. The corresponding Péclet number for the above conditions ranges from 3 360 to 15 600, which justifies
a posteriori the asymptotic expansions in terms of Pe −1/2 , and the non-dimensional
parameter β ranges from −4.6 × 10−6 to −4.6 × 10−1 .
The order of magnitude of the average bubble wall concentration of such molecules is
shown in figure 8(a): the Ωm curve for the smallest molecules (MA = 100 Da) remains
indistinguishable from 1 even for high driving pressure so that the mixture is unsegregated
on average. As the weight of the molecules increases, their average depletion at the bubble
wall becomes increasingly high for a given driving pressure. A nearly total depletion of
the heaviest molecules (MA = 5 × 106 and 107 Da) can even be observed for driving
pressures slightly above the Blake threshold.
The amplitude of the oscillatory concentration variation ∆Ωm is shown in figure 8(b),
where it can be seen that the smallest molecule is already over-concentrated by a factor
of 2 at P = 1.5. As MA increases, ∆Ωm first increases, and then decreases again for
very large molecules. This illustrates the opposite effects of the two factors exp(βI) and
βPe −1/2 Gmin in equation (6.2b). For the smallest molecules, the increase of |Gmin | with
P dominates over the decrease of exp(βI), so that ∆Ωm globally increases with driving
pressure, up to nearly 500 for MA = 100000 Da and P = 1.6. For larger molecules, the
opposite occurs, so that the peak value ∆Ωm becomes increasingly masked by the strong
average depletion Ωm and hardly departs from 0 for MA = 107 Da.
Thus, it is seen that both average depletion and peak periodic over-concentration at
the bubble wall compete, depending on the driving level and the molecule sizes. This
suggests that molecules or nano-particles that could undergo some growth or agglomeration process would be periodically concentrated against the bubble wall as long as
they are sufficiently small, but would be held far from the bubble on average, as they
reach some critical size. This may have some strong consequences on polymerization or
nano-particles agglomeration processes for example.
The present results may also help to understand the positive effect of acoustic cavitation on crystal nucleation from a solute (see for example Lyczko et al. 2002, for potassium
sulfate crystallisation). Homogeneous nucleation of crystals in liquids is a first-order phase
transition, which occurs as the solute concentration exceeds the saturation concentration.
There is a fairly general agreement on the so-called classical nucleation theory (Kaschiev
2000), which states that in a metastable solution, the nucleation process occurs through
progressive accumulation of solute molecules, forming multi-mers called “clusters”, up
to a critical radius called nucleus, from which a solid crystal is then free to grow. There
is indeed experimental evidence of the existence of such clusters and their stratification
under gravity has been observed in a sedimentation column (Mullin & Leci 1969; Larson
& Garside 1986). Although stated differently by these authors, the process invoked to
explain cluster sedimentation obeys the pressure diffusion equation considered in this
paper. One may therefore reasonably conjecture that the pressure diffusion effect around
an oscillating bubble would also tend to segregate these clusters to a much larger extent
than gravity, in view of the respective accelerations involved. The present conclusions
23
Segregation of a liquid mixture by a radially oscillating bubble
3
1
10
(a)
102 Da
104 Da
0.8
105 Da
105 Da
(b)
106 Da
2
10
104 Da
1
10
102 Da
0
10
Ωm
106 Da
0.4
∆Ωm
0.6
−1
10
5.106 Da
−2
0.2
10
107 Da
0
5.106 Da
107 Da
−3
0.9
1.1
1.3
P
1.5
10
0.9
1.1
1.3
1.5
P
Figure 8. Segregation ratio in a mixture of water with molecules of apparent density
2000 kg m−3 , of molecular weight MA ranging from 100 to 107 Da around a 4 µm argon bubble
driven at 26.5 kHz: (a) smooth segregation ratio defined by equation (6.2a), (b) oscillatory peak
segregation ratio defined by equation (6.2b).
show that this is indeed the case, and predict that if the nucleating species is heavier
than the liquid, its clusters would be periodically pushed against the bubble wall. There,
since the collision probability varies with the square of the concentration, they could
undergo more frequent attachment events and create bigger clusters. Above a critical
size these clusters would then be held far from the bubble in the liquid, as suggested by
figure 8(a).
It should be added that the conclusions on the smooth effect should be tempered by
considerations on the bubble stability. The smooth effect needs a very large number
of acoustic periods to build up, so that its potential appearance is conditioned by the
bubble stability on such a large timescale. If this stability is well established in singlebubble experiments, there is no definitive conclusion in multi-bubble fields. This issue
has been discussed recently by Louisnard & Gomez (2003). Even in the most optimistic
case, inertial bubbles would rapidly increase their size by rectified diffusion up to the
fragmentation threshold, in a time too small for the smooth effect to build up completely.
Partial build-up remains however possible, and may yield a noticeable smooth effect on
the largest molecules.
Finally, despite the compressibility effects were arbitrarily neglected to reduce the
mathematical complexity of the problem, it is nevertheless an important issue. A spherical shock-wave can build up when the bubble rebounds after the collapse, and the corresponding steepening of the pressure profile may therefore enhance the oscillatory effect.
Such spherical shocks are non-monotonic (see for example figures 7 and 8 of Fujikawa
& Akamatsu 1980) so that just after the collapse, the heaviest species would be very
24
O. Louisnard, F. J. Gomez and R. Grossier
concentrated in a thin layer of fluid surrounding the shock, and would travel with the
shock. An important consequence of this feature is that the excess concentration would
not remain located near the bubble wall, but would be transported toward the bulk liquid. In summary, shock-waves would not only enhance the oscillatory effect, but they
may also extend its influence to a larger spatial region.
7. Conclusion
We have proposed an analytic method to solve the general problem of pressure-gradient
forced diffusion of two non-volatile species around a bubble oscillating radially in the
mixture. The method yields the concentration field in the mixture around the bubble
in two parts: a smooth part, building over a number of acoustic periods of order of the
Péclet number Pe and asymptotically constant in time, and an oscillatory part. Both
expressions are fully analytic and can be easily calculated for a given bubble dynamics.
In the case of inertial cavitation, the oscillatory effect results in a large excess concentration of the heaviest species at the bubble wall at each bubble collapse. This excess
is noticeable even for small molecules, and relaxes with a characteristic time which is
more than one order of magnitude larger than the characteristic duration of the bubble
rebound. Conversely, the smooth effect pushes the heaviest species far from the bubble.
It remains unimportant for small molecules, even for strong driving pressures, but may
almost deplete the bubble wall of large molecules. Both smooth and oscillatory effects
increase with driving pressure. The smooth effect increases with the frequency of the
driving. The oscillatory effect increases with frequency for small driving pressure but
conversely, decreases with frequency in the inertial regime.
For large molecules or nano-particles around an inertial bubble, the two smooth and
oscillatory effects compete: the oscillatory effect dominates for the smallest molecules,
while the smooth one is prominent for the largest ones. This has strong implications
for any physico-chemical process involving molecules or particles undergoing a growing
or agglomeration process, and suggests that species smaller than a given size would be
periodically pushed and concentrated near the bubble wall, while the largest ones are in
average held far from the bubble. Polymerization, agglomeration or cluster formation in
crystal nucleation fall in this specific case and this behaviour may be partly responsible
for the reported enhancement of nucleation by cavitation.
This work is supported by an ECOS-South collaboration program between France and
Chile under grant number C03E05.
Appendix A. Linearization of the convection–diffusion equation
We assume an ideal mixture of two liquids, so that volume is additive. Under these
conditions, the mean density of the mixture is
ρ=
xA MA + xB MB
,
xA V̄A + xB V̄B
(A 1)
where xi , Mi and V̄i are the mole fraction, molecular weight and partial molal volume of
species i, respectively. Using the relation xi = M ωi /Mi between mole and mass fraction,
one readily obtain
1
ρ=
,
(A 2)
ωA v̄A + ωB v̄B
Segregation of a liquid mixture by a radially oscillating bubble
25
where the notation v̄i = V̄i /Mi has been used. The mean molecular weight of the mixture
is defined by
1
ωB
ωA
+
.
=
M
MA
MB
(A 3)
Replacing ωB by 1 − ωA , we can express the density of the solution by
ρ=
1
−1
[1 + α1 ωA ] ,
vB
and the two contributions to diffusion in equation (2.6) become
ρ∇ωA =
MA MB
ωA
M RT
V̄A
1
−
MA
ρ
1
−1
[1 + α1 ωA ] ∇ωA ,
vB
∇p =
MA
(1 − ωA )(1 + α2 ωA )
∇p,
α1 ωA
RT
1 + α1 ωA
(A 4a)
(A 4b)
where parameters α1 and α2 are defined by:
α1 =
v̄A
− 1,
v̄B
(A 5a)
α2 =
MB
− 1.
MA
(A 5b)
2
2
2
),
) and O(α22 ωA
Therefore it is seen that if we neglect terms of order O(ωA
), O(α12 ωA
equation (2.6) becomes
1 ∂ωA
MA
1
+ v · ∇ωA = D ∇ · ∇ωA +
ωA (v̄A − v̄B ) ∇p
(A 6)
v̄B
∂t
v̄B
RT
which is equation (2.7).
Appendix B. Solution of the oscillatory problem and splitting
Each member of the hierarchy of oscillatory problems may be expressed as a nonhomogeneous diffusion partial differential equation
∂Ciosc
∂ 2 Ciosc
osc
−
= F i R(τ̂ ), s, C0osc , . . . , Ci−1
,
2
∂ τ̂
∂s
(B 1)
with a Neumann inhomogeneous boundary condition of the form:
∂Ciosc
osc
(s = 0, τ̂ ) = B i R(τ̂ ), Ci−1
.
∂s
(B 2)
We treat the problem in the manner of Fyrillas & Szeri (1995), with the difference in
that here we have a Neumann condition rather than a Dirichlet one.
The oscillatory solution Ciosc should vanish for s → +∞ since the outer solution of
the boundary layer problem is imposed to be identically 0. The asymptotic oscillatory
solutions C̄ osc at any order have T periodicity in the τ variable, and therefore T̂ = τ̂ (T )
periodicity in the τ̂ variable. Thus the functions F i y B i are also periodic in τ̂ and we
26
O. Louisnard, F. J. Gomez and R. Grossier
expand them in Fourier series, as well as C̄ osc . Setting ωm = 2mπ/T̂ , we get
C̄iosc (s, τ̂ ) =
m=+∞
X
cm (s) exp(iωm τ̂ ),
m=−∞
F i (s, τ̂ ) =
m=+∞
X
fm (s) exp(iωm τ̂ ),
m=−∞
B i (τ̂ ) =
m=+∞
X
bm exp(iωm τ̂ ).
m=−∞
Substituting these series in the problem (B 1)-(B 2), we obtain a set of differential equations relating the coefficients of these series. For any m 6= 0, we obtain
d2 cm (s)
− iωm cm (s) = −fm (s)
ds2
dcm (s)
(s = 0) = bm
ds
The general solution of equation (B 3) vanishing for s → ∞ is
Z ∞
1
fm (s′ ) sinh [km (s − s′ )] ds′ ,
cm (s) = Am exp (−km s) −
km s
(B 3)
(B 4)
(B 5)
with
1
1
2|m|π 2 iǫm π/4
|ωm | 2
=
e
,
km = (1 + ǫm i)
2
T̂
where ǫm = sgn(m). The boundary condition at s = 0 (B 4) yields the following expression for Am :
Z ∞
1
bm
fm (s′ ) cosh(km s′ ) ds′ .
(B 6)
−
Am = −
km
km 0
The zeroth-order harmonics differential equation (m = 0) takes a different form:
d2 c0 (s)
= −f0 (s),
ds2
with the associated boundary condition
dc0 (s)
(s = 0) = b0 .
ds
The solution vanishing for s = ∞ is
Z sZ ∞
c0 (s) =
f0 (s′′ ) ds′′ ds′ .
s′
∞
Applying the Neumann boundary condition at s = 0 yields:
Z ∞
f0 (s′ ) ds′ ,
b0 =
0
and recognizing that b0 = B
reads
i
τ̂
B
and f0 (s) = F i (s) τ̂ , the separation condition finally
i
τ̂
=
Z
0
+∞
F i (s)
τ̂
ds.
(B 7)
Finally, in the special case where F i is identically zero, which is the case in the present
Segregation of a liquid mixture by a radially oscillating bubble
27
paper for i = 1, the separation condition just implies that B i τ̂ should be 0, and the
oscillatory field reads in this case:
! 21 m=+∞
"
1 #
X
T̂
π
|m|π 2
τ̂
bm
osc
C̄i (0, τ̂ ) = −
− (ǫm i + 1)
s .
exp i 2πm − ǫm
2π
4
|m|1/2
T̂
T̂
m=−∞
m6=0
(B 8)
It can be noted that the presence of both m1/2 in the denominator and the π/4 phase
lag recalls the fact that C̄iosc (0, τ̂ ) is the half-order integral of B i (τ̂ ) as could be proved
directly by solving problem (B 1), (B 2) by Laplace transforms.
Appendix C. Solution of the second-order oscillatory problem
The second-order oscillatory problem reads
osc
∂C2osc
∂ 2 C2osc
∂
′ ∂C1
′ osc
A1 s
,
=
+ βB0 C1
+
∂ τ̂
∂s
∂s
∂s2
∂C2osc
(0, τ̂ ) + βH(τ̂ )C1osc (0, τ̂ ) = −G1 − βH(τ̂ )C1sm (0, τ ),
∂s
C2osc (s → ∞) = 0.
The splitting condition (B 7) reads therefore:
(C 1a)
(C 1b)
(C 1c)
h−βH(τ̂ )C1osc (0, τ̂ ) − G1 − βH(τ̂ )C1sm (0, τ )iτ̂
Z ∞
osc
∂
′ osc
′ ∂C1
=
+ βB0 C1
A1 s
ds . (C 2)
∂s
∂s
0
τ̂
Using equations (3.8) and (3.10c), the integral in the right-hand side of equation (C 2)
can also be written
∂C osc
lim A′1 s 1 (s, τ̂ ) − βH(τ̂ )C1osc (0, τ̂ ).
(C 3)
s→∞
∂s
It can be seen from equation (3.13) that the first term of expression (C 3) is zero, so that
(C 2) becomes finally
G1 = −β hH(τ̂ ) [C1sm (0, τ )]iτ̂ .
(C 4)
Appendix D. Solution of the smooth problem
The second and third order smooth equations are
∂C0sm
∂
∂C2sm
∂C0sm
sm
=−
+
+ βB(σ, τ )(C0 + 1) ,
A(σ, τ )
∂τ
∂λ
∂σ
∂σ
∂C1sm
∂C1sm
∂
∂C3sm
sm
A(σ, τ )
=−
+
+ βB(σ, τ )C1
.
∂τ
∂λ
∂σ
∂σ
(D 1a)
(D 1b)
The expansion (4.2) must be uniformly valid and therefore should not contain secular
terms increasing unbounded when τ → ∞. This non-secular behaviour will be satisfied
by C2sm and C3sm only if the right-hand sides of equations (D 1 a,b) have zero τ -averages.
Therefore, C0sm and C1sm should fulfill the respective non-secularity conditions
∂C0sm
∂
∂C0sm
sm
=
+ β hB(σ, τ )iτ (C0 + 1) ,
(D 2a)
hA(σ, τ )iτ
∂λ
∂σ
∂σ
28
O. Louisnard, F. J. Gomez and R. Grossier
∂C1sm
∂C1sm
∂
=
+ β hB(σ, τ )iτ C1sm ,
hA(σ, τ )iτ
∂λ
∂σ
∂σ
(D 2b)
where the independence of C0sm and C1sm on τ has been used.
The associated boundary conditions at the bubble wall, equation (2.24b), can be obtained from expressions (3.11) and (3.14) of the separation constants G0 and G1 . Further
using the independence of C0sm and C1sm on the fast variable τ , these boundary conditions
read
∂C0sm
(0, λ) + β hH(τ̂ )iτ̂ [C0sm (0, λ) + 1] = 0,
(D 3a)
∂σ
∂C1sm
(0, λ) + β hH(τ̂ )iτ̂ C1sm (0, λ) = 0.
(D 3b)
∂σ
Moreover, it can be noticed that, from the definition (2.25) of H, the nonlinear average
hH(τ̂ )iτ̂ also reads
hHiτ̂ =
B(0, τ )
A(0, τ )
τ̂
=
V 4/3 B(0, τ )/A(0, τ )
V 4/3
τ
τ
=
hB(0, τ )iτ
,
hA(0, τ )iτ
since A(0, τ ) = V 4/3 , so that the zeroth- and first-order boundary conditions (D 3 a,b)
at the bubble wall may also be written
hA(0, τ )iτ
∂C0sm
(0, λ) + β hB(0, τ )iτ [C0sm (0, λ) + 1] = 0,
∂σ
(D 4a)
∂C1sm
(0, λ) + β hB(0, τ )iτ C1sm (0, λ) = 0.
(D 4b)
∂σ
sm
for i = 0, 1 of equations (D 2 a,b), by setting
We now seek the asymptotic solutions Ci,∞
sm
∂Ci,∞ /∂λ = 0 for i = 0, 1 in these equations and integrating once with respect to σ.
Making use of boundary conditions (D 4 a,b), this integration yields
hA(0, τ )iτ
hA(σ, τ )iτ
sm
∂C0,∞
sm
+ 1 = 0,
+ β hB(σ, τ )iτ C0,∞
∂σ
(D 5a)
sm
∂C1,∞
sm
+ β hB(σ, τ )iτ C1,∞
= 0.
(D 5b)
∂σ
Now using the condition at infinity (2.26), the zeroth-order equation (D 5a) can be integrated as
Z ∞
hB(σ ′ , τ )iτ
′
sm
dσ − 1.
C0,∞ (σ) = exp β
(D 6)
hA(σ ′ , τ )iτ
σ
hA(σ, τ )iτ
Besides, integration of the first-order equation (D 5b) can only yield the null solution
sm
C1,∞
= 0 in order to fulfill the condition at infinity, equation (2.26). It does not imply
however that C1sm (σ, λ) is zero for finite λ, but just states that its asymptotic limit for
λ → ∞ is zero.
The physical meaning of the asymptotic smooth solution may be understood by at
equations (D 5 a,b). The average pressure diffusion flux (the B term) is exactly balanced
by the average Fick diffusion flux (the A term), and therefore the smooth concentration
field stays constant. The unsteady term in the smooth equations (D 2 a,b) represents the
transitory non-equilibrium between the two average diffusion processes.
Segregation of a liquid mixture by a radially oscillating bubble
29
Appendix E. Numerical estimation of the oscillatory asymptotic
concentration field
We first set x = 2πτ̂ /T̂ , xmin = 2πτ̂min /T̂ and ∆x = 2π∆τ̂ /T̂ . In the new variable x,
the tooth-approximation (5.10) can be written
x − xmin
Hmin 1 −
, x ∈ [xmin − ∆x, xmin + ∆x],
H(x) ≃
(E 1)
∆x
0,
elsewhere,
with Hmin < 0. This function can be Fourier-expanded as
m=+∞
X 1
2
m∆x
Hm ∆x
2
exp [im(x − xmin )] ,
+
Hmin
sin
H(x) =
2π
π∆x
m2
2
m=−∞
(E 2)
m6=0
and function G defined by (5.8) can therefore be approximated as
!1/2
T̂
4
Gapp (τ̂ ) =
F (x),
Hmin
2π
π∆x
(E 3)
with
F (x) =
m=+∞
X
m=1
1
sin2
m5/2
m∆x
2
h
πi
cos m(x − xmin ) −
,
4
(E 4)
which is equation (5.12). In order to get an estimate of the maximum of F (x), we first
reformulate it as:
F (x) = C(x) + S(x),
(E 5)
with
+∞
1 X 1
[2 cos m(x − xmin )
C(x) = √
4 2 m=1 m5/2
− cos m(x − xmin + ∆x) − cos m(x − xmin − ∆x)]
+∞
1 X 1
√
[2 sin m(x − xmin )
S(x) =
4 2 m=1 m5/2
− sin m(x − xmin + ∆x) − sin m(x − xmin − ∆x)]
(E 6a)
(E 6b)
Let’s set:
Z(X) =
+∞
X
m−1/2 (cos mX + sin mX)
(E 7)
m=1
It can be seen that differentiating (E 5), (E 6) twice, F ′′ (x) is the sum of three series of
the form (E 7):
1
F ′′ (x) = √ [Z(x − xmin − ∆x) + Z(x − xmin + ∆x) − 2Z(x − xmin )]
4 2
(E 8)
A theorem by Zygmund (1959) states that
+∞
X
m=1
β−1
m−β cos mX ≃ |X|
X→0
Γ(1 − β) sin π
β
2
(E 9a)
30
O. Louisnard, F. J. Gomez and R. Grossier
+∞
X
m=1
β−1
m−β sin mX ≃ sgn(X) |X|
X→0
Γ(1 − β) cos π
for any β ∈ [0, 1[. Therefore, taking β = 1/2, we get
√
−1/2
2Γ (1/2)
Z(X) ≃ H(X) |X|
X→0
β
2
(E 9b)
(E 10)
where H is the Heaviside function. For small enough ∆x, any x in the neighbourhood of
xmin is also in the neighbourhood of xmin − ∆x and xmin + ∆x, so that, from (E 8), we
can approximate F ′′ (x) as
′′
F ′′ (x) ≃ Fapp
(x)
"
#
Γ (1/2) H(x − xmin − ∆x)
H(x − xmin + ∆x)
H(x − xmin )
=
. (E 11)
+
−2
1/2
1/2
1/2
4
|x − xmin − ∆x|
|x − xmin + ∆x|
|x − xmin |
Integrating twice yields
F (x) ≃ Fapp (x) =
Γ (1/2)
[f (x − xmin − ∆x) + f (x − xmin + ∆x) − 2f (x − xmin )] + Ax + B, (E 12)
3
where f is defined by
3/2
f (X) = H(X) |X|
,
and A, B are two integration constants. Clearly A should be 0 to avoid a spurious
discontinuity of F at x = 2nπ and B must be calculated so that the approximation of F
has a zero average on [0, 2π], as does the original function (E 4). This condition yields
i
Γ (1/2) h
B=
2(2π − xmin )5/2 − (2π − xmin − ∆x)5/2 − (2π − xmin + ∆x)5/2 .
15π
It can be easily checked that Fapp (x) has a maximum at x = xmin + ∆x/3 whose value is
max
Fapp
(x) =
2Γ (1/2)
√ ∆x3/2 + B.
3 3
(E 13)
Owing to the approximation used to obtain (E 11), it is clear that approximation (E 12)
becomes better for smaller ∆x. Figure 9(a,b) shows a comparison of the calculated series
(E 4) (solid line) and its approximation by (E 12) (dashed line) for xmin = π and ∆x =
π/2 (figure 9a) or ∆x = π/10 (figure 9b). It is seen that for ∆x as large as π/2 (in this
case the peak spans over half of the interval), the maximum of F is still predicted with
a relative error as low as 8 %. For ∆x = π/10 is reduced to 1.25 %. Moreover, it can be
noted that the approximation of F is not only good near xmin , where it should be, but
also over the whole interval [0, 2π].
The quality of the approximation of maxx F can be seen in figure 9(c), in which the
max
/F max is displayed as a function of ∆x: since for a
relative error ǫ = F max − Fapp
typical inertial bubble, ∆x amounts to 10−9 , it is clear that the approximation given by
max
equation (E 13) is excellent. We also draw the value of the constant B relative to Fapp
on figure 9(d ), which shows clearly that B can be easily neglected for ∆x smaller than
10−2 .
Finally, it is of interest to know the characteristic relaxation time of function F after it
has reached its maximum. It can be shown after some algebra that F reaches a fraction
of its maximum value αF max after a time approximately equal to 27∆x/64α2 . Applying
31
Segregation of a liquid mixture by a radially oscillating bubble
1
(a)
F
0
0
−0.5
−0.1
−1
0
2
x
4
6
0
−2
4
6
10
(c)
(d)
−3
ǫ
x
0
10
−1
10
10
−4
10
2
B/F max
F
(b)
0.1
0.5
−2
−3
10
−2
10
−1
10
10
−3
10
∆x
−2
10
−1
10
∆x
Figure 9. (a) and (b): comparison of function F calculated numerically from equation (E 4)
(solid line) with function F calculated by approximation (E 12) (dashed line). The dash-dotted
line recalls the shape of the tooth approximation (E 1) of function H. Figure (a) is obtained
with ∆x = π/2 and figure (b) with ∆x = π/10. (c) Relative error on the maximum value of F
calculated from (E 12), as ∆x is varied. (d ) Ratio B/F max as ∆x is varied.
this formula shows that F is still equal to one fifth of its maximum value after 10.5∆x,
and to one tenth after 42∆x.
REFERENCES
Akhatov, I., Gumerov, N., Ohl, C., Parlitz, U. & Lauterborn, W. 1997 The role of
surface tension in stable single bubble sonoluminescence. Phys. Rev. Lett. 78 (2), 227–230.
Archibald, W. J. 1938 The process of diffusion in a centrifugal field of force. Phys. Rev. 53,
746–752.
Bird, R. B., Stewart, W. E. & Lightfoot, E. N. 1960 Transport phenomena. John Wiley
and sons.
Brenner, M. P., Hilgenfeldt, S. & Lohse, D. 2002 Single-bubble sonoluminescence. Rev.
Mod. Phys. 74 (2), 425–483.
Crum, L. A., Mason, T. J., Reisse, J. L. & Suslick, K. S., ed. 1999 Sonochemistry and Sonoluminescence, Dordrecht. Kluwer Academic, proceedings of the NATO Advanced Study Institute on Sonoluminescence and Sonoluminescence, Leavenworth, Washington, USA, 18-29
August 1997.
Eller, A. & Flynn, H. G. 1965 Rectified diffusion during nonlinear pulsations of cavitation
bubbles. J. Acoust. Soc. Am. 37, 493–503.
Fujikawa, S. & Akamatsu, T. 1980 Effects of the nonequilibrium condensation of vapour on
32
O. Louisnard, F. J. Gomez and R. Grossier
the pressure wave produced by the collapse of a bubble in a liquid. J. Fluid Mech. 97,
481–512.
Fyrillas, M. M. & Szeri, A. J. 1994 Dissolution or growth of soluble spherical oscillating
bubbles. J. Fluid Mech. 277, 381–407.
Fyrillas, M. M. & Szeri, A. J. 1995 Dissolution or growth of soluble spherical oscillating
bubbles: the effect of surfactants. J. Fluid Mech. 289, 295–314.
Fyrillas, M. M. & Szeri, A. J. 1996 Surfactant dynamics and rectified diffusion of microbubbles. J. Fluid Mech. 311, 361–378.
Hickling, R. & Plesset, M. S. 1964 Collapse and rebound of a spherical bubble in water.
Phys. Fluids 7, 7–14.
Hilgenfeldt, S., Brenner, M. P., Grossman, S. & Lohse, D. 1998 Analysis of rayleighplesset dynamics for sonoluminescing bubbles. J. Fluid Mech. 365, 171–204.
Hilgenfeldt, S., Lohse, D. & Brenner, M. P. 1996 Phase diagrams for sonoluminescing
bubbles. Phys. Fluids 8 (11), 2808–2826.
Hirschfelder, J. O., Curtiss, C. F. & Bird, R. B. 1967 Molecular theory of gases and
liquids. John Wiley and sons.
Hsieh, D. Y. & Plesset, M. S. 1961 Theory of rectified diffusion of mass into gas bubbles. J.
Acoust. Soc. Am. 33, 206–215.
Kaschiev, D. 2000 Nucleation : Basic theory with applications. Butterworths-Heinemann.
Keller, J. B. & Miksis, M. 1980 Bubble oscillations of large amplitude. J. Acoust. Soc. Am.
68, 628–633.
Larson, M. A. & Garside, J. 1986 Solute clustering in supersaturated solutions. Chem. Eng.
Sci. 41 (5), 1285–1289.
Lin, H., Storey, B. D. & Szeri, A. J. 2002 Rayleigh-taylor instability of violently collapsing
bubbles. Phys. Fluids 14 (8), 2925–2928.
Louisnard, O. & Gomez, F. 2003 Growth by rectified diffusion of strongly acoustically forced
gas bubbles in nearly saturated liquids. Phys. Rev. E 67 (036610), 1–12.
Lyczko, N., Espitalier, F., Louisnard, O. & Schwartzentruber, J. 2002 Effect of ultrasound on the induction time and the metastable zone widths of potassium sulphate. Chem.
Eng. Journ. 86, 233–241.
Mullin, J. W. & Leci, C. L. 1969 Evidence of molecular cluster formation in supersaturated
solutions of citric acid. Phil. Mag. 19 (161), 1075–1077.
Plesset, M. S. & Zwick, S. A. 1952 A nonsteady heat diffusion problem with spherical
symmetry. J. Appl. Phys. 23 (1), 95–98.
Prosperetti, A. 1999 Old-fashioned bubble dynamics. In Sonochemistry and Sonoluminescence
(ed. L. A. Crum, T. J. Mason, J. L. Reisse & K. S. Suslick), pp. 39–62. Dordrecht: Kluwer
Academic, proceedings of the NATO Advanced Study Institute on Sonoluminescence and
Sonoluminescence, Leavenworth, Washington, USA, 18-29 August 1997.
Prosperetti, A. & Lezzi, A. 1986 Bubble dynamics in a compressible liquid. part 1. first-order
theory. J. Fluid Mech. 168, 457–478.
Storey, B. D. & Szeri, A. 2000 Water vapour, sonoluminescence and sonochemistry. Proc.
R. Soc. London, Ser. A 456, 1685–1709.
Storey, B. D. & Szeri, A. 2001 A reduced model of cavitation physics for use in sonochemistry.
Proc. R. Soc. London, Ser. A 457, 1685–1700.
Storey, B. D. & Szeri, A. J. 1999 Mixture segregation within sonoluminescence bubbles. J.
Fluid Mech. 396, 203–221.
Toegel, R., Gompf, B., Pecha, R. & Lohse, D. 2000 Does water vapor prevent upscaling
sonoluminescence ? Phys. Rev. Lett. 85 (15), 3165–3168.
Tomita, Y. & Shima, A. 1977 On the behaviour of a spherical bubble and the impulse pressure
in a viscous compressible liquid. Bulletin of the JSME 20 (149), 1453–1460.
Zygmund, A. 1959 Trigonometric series. Cambridge University Press.