ArticlePDF Available

Particle separation by Stokes number for small neutrally buoyant spheres in a fluid

Authors:

Abstract and Figures

It is a commonly observed phenomenon that spherical particles with inertia in an incompressible fluid do not behave as ideal tracers. Due to the inertia of the particle, the planar dynamics are described in a four-dimensional phase space and thus can differ considerably from the ideal tracer dynamics. Using finite-time Lyapunov exponents, we compute the sensitivity of the final position of a particle with respect to its initial velocity, relative to the fluid, and thus partition the relative velocity subspace at each point in configuration space. The computations are done at every point in the relative velocity subspace, thus giving a sensitivity field. The Stokes number, being a measure of the independence of the particle from the underlying fluid flow, acts as a parameter in determining the variation in these partitions. We demonstrate how this partition framework can be used to segregate particles by Stokes number in a fluid. The fluid model used for demonstration is a two-dimensional cellular flow.
Content may be subject to copyright.
Particle segregation by Stokes number for small neutrally buoyant spheres in a fluid
Phanindra Tallapragada*and Shane D. Ross
Department of Engineering Science and Mechanics, Virginia Polytechnic Institute and State University, (VPISU),
Blacksburg, VA 24061, USA
Received 12 March 2008; revised manuscript received 14 June 2008; published 10 September 2008
It is a commonly observed phenomenon that spherical particles with inertia in an incompressible fluid do not
behave as ideal tracers. Due to the inertia of the particle, the planar dynamics are described in a four-
dimensional phase space and thus can differ considerably from the ideal tracer dynamics. Using finite-time
Lyapunov exponents, we compute the sensitivity of the final position of a particle with respect to its initial
velocity, relative to the fluid, and thus partition the relative velocity subspace at each point in configuration
space. The computations are done at every point in the relative velocity subspace, thus giving a sensitivity
field. The Stokes number, being a measure of the independence of the particle from the underlying fluid flow,
acts as a parameter in determining the variation in these partitions. We demonstrate how this partition frame-
work can be used to segregate particles by Stokes number in a fluid. The fluid model used for demonstration
is a two-dimensional cellular flow.
DOI: 10.1103/PhysRevE.78.036308 PACS numbers: 47.52.j, 47.51.a
I. INTRODUCTION
It has long been observed that particles with a finite size
and mass have different dynamics from the ambient fluid.
Because of their inertia the particles do not evolve as point-
like tracers in a fluid. This leads to preferential concentra-
tion, clustering, and separation of particles as observed in
numerous studies 13. The inertial dynamics of solid par-
ticles can have important implications in natural phenomena,
e.g., the transport of pollutants and pathogenic spores in the
atmosphere 4,5, formation of rain clouds 6by coales-
cence around dust particles, and formation of plankton colo-
nies in oceans 7. Similarly, the inertial dynamics of reactant
particles is important in the reaction kinetics and distribution
of reactants in solution for coalescence-type reactions 8.
Mixing-sensitive reactions in the wake of bubbles have been
shown to be driven by buoyancy effects of reactants 9.
Recently, a principle of asymmetric bifurcation of laminar
flows was applied to the separation of particles by size and
demonstrated the separation of flexible biological particles
and the fractional distillation of blood 10,11. Innovative
channel geometries have been empirically designed to focus
randomly ordered inertial particles in microchannels 12.
These phenomena and related applications rely on the non-
trivial dynamics of inertial particles in a fluid.
Many theoretical and numerical studies have been done
on the dynamics of inertial particles in some model flows,
using the Maxey-Riley equation 13. Maxey 14studied the
settling properties and retention zones of nonbuoyant inertial
particles in vertical cellular flows, under the influence of
gravity. The sensitive dependence on the initial conditions of
inertial particle trajectories and clustering was further studied
by 15using a two-dimensional cellular flow for neutrally
buoyant particles.
Several studies have been concerned with characterizing
the surface in physical space to which inertial particles clus-
ter, using, for instance, fractal dimension and rates of con-
vergence to this surface. Among these, 1618studied clus-
tering of inertial particles in two-dimensional flow past a
cylinder. Similar studies 1922considered inertial particle
clustering in turbulent flows. In the turbulent flow studies,
the spectrum of long-timeLyapunov exponents was com-
puted to calculate the fractal dimension of the surfaces of
particle clustering. While we also use Lyapunov exponents,
we use them in a different way, as described below.
A method to segregate inertial particles from an initial
mixture by different sizes i.e., Stokes numberswas numeri-
cally demonstrated in 23, but no physical description was
given of how segregation arose. We provide a description by
partitioning phase space into zones of initial phase space
locations where inertial particles will evolve to different final
locations, according to Stokes number. To achieve this, we
calculate the phase space distribution of short-time Lyapunov
exponents, and use topological features of this distribution to
find the partition boundaries. This partition is parametrized
by the Stokes number. We demonstrate this method using a
simple test model of two-dimensional flow, namely, cellular
flow, demonstrating that this partitioning scheme can explain
the sensitive dependence of trajectories on initial conditions
and the consequent clustering effects. We argue that this
methodology can be used as a systematic tool to achieve
segregation of inertial particles in a fluid.
We employ a simplified form of the Maxey-Riley equa-
tion 13as the governing equation for the motion of inertial
particles in a fluid. The dynamics of a single particle occur in
a four-dimensional phase space. The sensitive dependence of
the particle motion on initial conditions is quantified using
the finite-time Lyapunov exponents FTLEs. It has been
shown previously 24,25, that the ridges in the FTLE field
act as separatrices. These are in general time dependent and
go by the name of Lagrangian coherent structures LCSs.
We chose to do a simplified sensitivity analysis by perturbing
the initial conditions in only two dimensions, in the initial
relative velocity subspace. We obtain a sensitivity field akin
to a FTLE field but restricted to the relative velocity sub-
space, and demonstrate numerically that the ridges in this
*tssap@vt.edu
sdross@vt.edu
PHYSICAL REVIEW E 78, 036308 2008
1539-3755/2008/783/0363089©2008 The American Physical Society036308-1
field act as separatrices. The partitions in the relative velocity
subspace created by these separatrices determine the even-
tual spatial distribution of particles in the fluid. Using this
partitioning scheme we show how the Stokes number acts as
a parameter in the separation of particles of different inertia
or size.
The paper is organized as follows. In Sec. II we review
the equation governing the inertial particle dynamics in a
fluid and its simplified form. In Sec. III we briefly review the
background theory of phase space distributions of finite-time
Lyapunov exponents, which we use to quantify the sensitiv-
ity of the physical location of inertial particles with respect
to perturbations in the initial relative velocity. We also de-
scribe our computational scheme to obtain the sensitivity
field in the relative velocity subspace. In Sec. IV we present
results for the sensitivity field of the inertial particles in a
cellular flow. In Sec. V we demonstrate our procedure for the
segregation of particles by their Stokes number using the
results from Sec. IV In Sec. VI we give numerical justifica-
tion for the robustness of the sensitivity field to perturbations
in the velocity field of the fluid. In Sec. VII we discuss the
results and give conclusions.
II. GOVERNING EQUATIONS
Our starting point is Maxey and Riley’s equation of mo-
tion of a rigid spherical particle in a fluid 13:
p
dv
dt =
f
Du
Dt +
p
fg9
␯␳
f
2a2
vua2
6
2u
f
2
dv
dt D
Dt
ua2
10
2u
9
f
2a
0
t1
t
d
d
vua2
6
2u
d
,1
where vis the velocity of the solid spherical particle, uthe
velocity field of the fluid,
pthe density of the particle,
fthe
density of the fluid,
the kinematic of the viscosity of the
fluid, athe radius of the particle, and gthe acceleration due
to gravity. The terms on the right-hand side are the force
exerted by the undisturbed flow on the particle, the force of
buoyancy, the Stokes drag, the added mass correction, and
the Basset-Boussinesq history force, respectively. Equation
1is valid under the following restrictions:
avu/
1,
a/L1,
a2
冊冉
U
L
1, 2
where Land U/Lare the length scale and velocity gradient
scale for the undisturbed fluid flow. The derivative
Du
Dt =
u
t+u·u3
is the acceleration of a fluid particle along the fluid trajec-
tory, whereas the derivative
dv
dt =
v
t+v·v4
is the acceleration of a solid particle along the solid particle
trajectory.
Equation 1can be simplified by neglecting the Faxen
correction and the Basset-Boussinesq terms 15. We restrict
our study to the case of neutrally buoyant particles, i.e.,
p
=
f. Writing W=vu, the relative velocity of the particle
and the surrounding fluid, the evolution of Wbecomes
dW
dt =−J+
IW,5
and the change in the particle position is given by
dr
dt =W+u,6
where Jis the gradient of the undisturbed velocity field of
the fluid, r=x,yis the position of the solid particle, and
=2
3St−1 is a constant for a particle with a given Stokes
number St. Equations 5and 6can be rewritten as the
vector field
d
dt =F
,7
with
=r,W=x,y,Wx,WyR4. Equation 7defines a
dissipative system with constant divergence −4
3
. It has been
shown by Haller 17that an exponentially attracting slow
manifold exists for general unsteady inertial particle motion
as long as the particle Stokes number is small enough. For
neutrally buoyant particles this attractor is W=0the xy
plane. Despite the global attractiveness of the slow mani-
fold, domains of instability exist in which particle trajecto-
ries diverge 15,16,18.
III. SENSITIVITY ANALYSIS
The Lyapunov characteristic exponent is widely used to
quantify the sensitivity to initial conditions. A positive
Lyapunov exponent is a good indicator of chaotic behavior.
We have used the finite-time version of the Lyapunov expo-
nents, the FTLEs, as a measure of the maximum stretching
for a pair of phase points.
We review some important background regarding the
FTLEs below, following 2426. The solution to Eq. 7can
be given by a flow map
t0
t, which maps an initial point
t0
at time t0to
tat time t,
PHANINDRA TALLAPRAGADA AND SHANE D. ROSS PHYSICAL REVIEW E 78, 036308 2008
036308-2
t=
t0
t
t0.8
The evolution over a time Tof the displacement between two
initially close phase points
t0and
t0+
t0is given by
t0+T=
d
t0
t0+T
d
t0+O共储
2.9
Neglecting the higher-order terms, the magnitude of the per-
turbation is
t0+T兲储 =
t0,
d
t0
t0+T
*
d
d
t0
t0+T
d
t0
.
10
The matrix
C=
d
t0
t0+T
*
d
d
t0
t0+T
d
11
is the right Cauchy-Green deformation tensor. Maximum
stretching occurs when the perturbation
is along the ei-
genvector nmax corresponding to the maximum eigenvalue
max of C. The growth ratio is given by
t0+T兲储/
t0兲储 =e
1
t0T,12
where
1
t0=1
Tln maxC兲共13
is the maximal finite-time Lyapunov exponent. One can as-
sociate an entire spectrum of finite-time Lyapunov exponents
with
t0, ordering them as
1
t0
2
t0
3
t0
4
t0.14
The entire spectrum of the Lyapunov exponents can be com-
puted from the state transition t,t0=d
t0
t
/d
matrix
using singular value decomposition, where t=t0+T,
t,t0=Bt,t0t,t01/2Rt,t0.15
The diagonal matrix gives all the Lyapunov exponents
while
tf,t0=lntf,t01/2T,16
where T=tft0and tf,t0=diag
1,...,
4. An arbitrary
perturbation in the fixed basis can be transformed using a
time-dependent transformation 26
t=At,t0,tf
t,17
such that in the new basis the primed frame, the variational
equations become
˙t=tf,t0
t.18
Since tf,t0is a constant diagonal matrix, we have
t=ett0tf,t0
t0.19
The first coordinate in the new frame grows as
1
t
=ett0
1
1
t0. The time-dependent transformation Atis
given by 26
At,t0,tf=ett0tf,t0Rtf,t0*Rt,t0t,t0−1/2Bt,t0.
20
A. Sensitivity to initial relative velocity
Since the dynamics of the inertial particle is in a four-
dimensional phase space, the separatrices, that is, LCSs de-
fined by ridges in the field of the maximal FTLEs, are three-
dimensional surfaces see 25兴兲. However, because the
system is dissipative and the global attractor is the xy sub-
space, we can obtain meaningful information by restricting
the computations to a lower-dimensional subdomain of the
phase space. This we do by considering an initial perturba-
tion only in the relative velocity subspace and study how this
perturbation grows in the xy plane, the configuration space,
i.e.,
t0=0,0, Wx,Wy*,21
where Wxand Wyare the perturbations in the relative
velocity subspace. It is to be noted that when perturbations
are applied to the initial velocity of the solid particle, an
initial drag is experienced by the particle, but this effect is
negligible 14.
Using the time-dependent transformation At,t0,tfthe
evolution of the perturbation is given by
t=A−1tett0tf,t0At0
t0.22
The growth of perturbations in the xy plane is given by the
first two components of the above vector. The last two com-
ponents are the evolution of the perturbations in the relative
velocity subspace. Since the xy plane is a global attracting
set, these tend to zero. One can choose a finite time Tsuch
that the evolution of the initial perturbation comes arbitrarily
close to the xy plane. In this way the sensitivity of the final
spatial location of the particles with respect to initial relative
velocity can be computed.
B. Numerical computation of the sensitivity field
The evolution of a perturbation is along the four basis
vectors. For an arbitrarily oriented initial perturbation the
growth may not be dominated in the direction of greatest
expansion for short integration times. This can be overcome
by sampling multiple perturbations in the different direc-
tions. A reference point and its neighbors are identified and
after a finite time their positions in configuration space are
PARTICLE SEGREGATION BY STOKES NUMBER FORPHYSICAL REVIEW E 78, 036308 2008
036308-3
computed. The state transition matrix can then be computed
at each point in the xy plane, by using a central finite-
difference method. For initial perturbations restricted to
WxWysubspace, this gives
,rW =
xi,j,k+1,lt0+Txi,j,k−1,lt0+T
Wxt0
xi,j,k,l+1t0+Txi,j,k,l−1t0+T
Wyt0
yi,j,k+1,lt0+Tyi,j,k−1,lt0+T
Wxt0
yi,j,k,l+1t0+Tyi,j,k,l−1t0+T
Wyt0
.23
In the four-dimensional finite-difference grid indexed by
i,j,k,l, each reference point has eight neighboring points,
one along each of the positive and negative directions in
each phase space direction. Since we are looking at the initial
perturbations in relative velocity only, we fix the initial spa-
tial position of the particle i,jbut vary its relative velocity
k,l. In the relative velocity, points are separated by con-
stant amounts Wx,Wyin the x,yrelative velocity direc-
tions, respectively.
The relative velocity sensitivity field
Wx,Wyis given
by
Wx,Wy=1
Tln max,rW
*,rW.24
Ridges on this sensitivity surface are one dimensional struc-
tures similar to LCSs. The ridges in the maximal sensitivity
field
Wx,Wypartition the relative velocity subspace. We
applied the above procedure to a cellular flow.
We make a note on the terminology used here. The field
measuring the sensitivity of the final location of particles in
configuration space with respect to perturbations in initial
relative velocity is analogous to the FTLE field, but not iden-
tical. To obtain the true FTLE field, one would have to com-
pute the 44 state transition matrix . Using the notation of
Eq. 23,
=
,rr ,rW
,Wr ,WW
.25
The FTLE field is then given by
x,y,Wx,Wy
=1/T兩兲lnmax*. Ridges in this field are three-
dimensional structures and represent the true LCSs. Ridges
in the relative velocity sensitivity field
Wx,Wyare one-
dimensional structures which can be considered “slices” of
the full three-dimensional structure, where the slices are pa-
rametrized by the two-dimensional location of the initial spa-
tial point x,y.
IV. EXAMPLE FLUID MODEL: CELLULAR FLOW
We demonstrate the computation of the relative sensitivity
field using a simple test model of two-dimensional flow. We
choose cellular flow, as it has been used in previous studies
14,15, and is a simple example of a fluid with separatrices.
This flow is described by the stream function
x,y,t=acos xcos y.26
The velocity field is given by
u=−acos xsin y,27
v=asin xcos y28
There are heteroclinic connections from the stable and
unstable manifolds of the fixed points 2n+1兲共
/2, shown
by the arrows in Fig. 1, which are also the boundaries of the
cells. These coincide with LCSf, the LCSs for fluid particles,
which have no relative velocity Wx=Wy=0and evolve ac-
cording to the fluid velocity field, Eqs. 27and 28. The
LCSfis to be distinguished from the LCS of the inertial
particle in the full four-dimensional phase space. By choos-
ing initial perturbations of the form given by Eq. 21at
different points along a streamline, we follow how these per-
turbations grow in the xy plane by integrating the particle
FIG. 1. Streamlines of
=acos xcos yform an array of cells.
The arrows indicate the heteroclinic fluid trajectories connecting the
fixed points of the velocity field formed by
. For this velocity
field, the heteroclinic trajectories coincide with the LCSf, i.e., the
separatrices or transport barriers, for fluid particles.
PHANINDRA TALLAPRAGADA AND SHANE D. ROSS PHYSICAL REVIEW E 78, 036308 2008
036308-4
trajectories numerically from which the sensitivity field is
computed. Figure 2shows the sensitivity field computed for
initial perturbations in the relative velocity subspace, at dif-
ferent points on the streamline
x,y,t=0. The ridges in
this field have high values of sensitivity. It can be seen that
there is a continuous variation in the ridges of the sensitivity
(
a)
(
b)
(
c
)
(
d)
(
e)
(
f)
(g
)
(
h)
(
i)
(
j)
(
k
)
(
l
)
FIG. 2. Ridges in the sensitivity field for
=acos xcos y. Initial spatial position varies from a兲共x0,y0=
/2,
/2to k兲共0,
/2,at
the points shown in l, along
=0, at intervals of 0.05
. The plots show a smooth variation in the structure of the ridges in the sensitivity
field. Parameters: a=100, St=0.2, T=0.24.
PARTICLE SEGREGATION BY STOKES NUMBER FORPHYSICAL REVIEW E 78, 036308 2008
036308-5
field with respect to the initial x,ycoordinates. In each case
the sensitivity field at a given point depends on the underly-
ing LCSfof the fluid flow.
The ridges in the sensitivity field have meaningful infor-
mation about the dynamics of inertial particles even when
computed at points far from the saddle points of the fluid
flow. This is shown in Fig. 3awhich is the sensitivity field
computed at x,y=3
/8,3
/8. The ridges in the sensitiv-
ity field partition the relative velocity subspace according to
the final location of particles. In Fig. 3bthe ridges in the
sensitivity field are used to identify regions in the relative
velocity subspace, which produce qualitatively different tra-
jectories. Particles that start at the same physical location,
but are in different regions of the relative velocity subspace,
are neatly separated from particles that started in other re-
gions, as shown in Fig. 3c. Thus the ridges in the sensitivity
field have the property of a separatrix.
V. SEGREGATION OF PARTICLES BY STOKES NUMBER
Equation 5can be diagonalized as
dWd
dt =
0
0
Wd,29
where are the eigenvalues of the Jacobian of the fluid
velocity field. If
=2
3St−1 is very large, then both the com-
ponents of Adwill decay. For low values of
, one compo-
nent of Adwill grow. Therefore the dynamics of an inertial
particle depend on the value of
, that is, on the Stokes
number. It is reasonable to expect that the computations of
the sensitivity of the particle location to the initial relative
velocity also would depend on the Stokes number. That this
is indeed the case is shown by the computations of the sen-
sitivity field for a particle with Stokes number 0.1 for the
time-independent flow, as shown in Fig. 4a. The thick lines
are the ridges in the sensitivity field for particles with St
=0.1 and the hatched lines are those of St=0.2. It can be seen
that, though the structure of the sensitivity field is similar, the
ridges are present at different locations in the relative veloc-
ity subspace. This fact can be exploited to design a process to
separate particles by their Stokes number. In this section we
illustrate a simple procedure for doing this.
The ridges of the sensitivity fields computed for the two
different particles of Stokes number 0.1 and 0.2, respectively,
are superimposed on the same plot, as shown in Fig. 4. The
subdomain of the relative velocity subspace sandwiched be-
tween the ridges of the sensitivity fields of the two types of
particles forms a zone of segregation. One such sample zone
is shown in gray in Fig. 4a. Two particles, with St=0.1 and
0.2, respectively, with common initial coordinates x,y
=3
/8,3
/8and initial relative velocities belonging to the
gray region, have trajectories that separate in physical space.
To illustrate this, the trajectories of 500 particles of each
Stokes number, starting at the same initial physical point
x,y=3
/8,3
/8and with initial relative velocities val-
ues belonging to the gray region were computed. Figures
4b4jshow snapshots of the particle positions as a func-
tion of time. The particles are completely segregated into two
(
a)
(b)
(
c)
FIG. 3. aRidges in the sensitivity field partition the velocity
subspace into regions of distinct qualitative dynamics. Three such
partitioned regions are shown. bParticles starting with relative
initial velocities belonging to distinct partitions in the relative ve-
locity subspace are segregated into different cells in the xy plane.
Same parameters as in Fig. 2:a=100, St=0.2, T=0.24. The initial
position of all particles is x0,y0=3
/8,3
/8, shown by the
marker in c.
PHANINDRA TALLAPRAGADA AND SHANE D. ROSS PHYSICAL REVIEW E 78, 036308 2008
036308-6
(a) (b) (c)
(d)
e
(f)
(g) (h) (i)
(j)
FIG. 4. aRidges in the sensitivity field for particles with St= 0.2 hatchedand St = 0.1 thick. The initial position of all particles is
x0,y0=3
/8,3
/8. The gray patch is a sample region sandwiched between the ridges of the two Stokes numbers. bjA mixture of
St=0.1 and St= 0.2 particles starting at x0,y0=3
/8,3
/8, with initial relative velocity in the gray patch in a, are separated into
different cells in the xy plane after a short time. t=b0.005, c0.030, d0.060, e0.085, f0.110, g0.135, h0.160, i0.185, and
j0.210.
PARTICLE SEGREGATION BY STOKES NUMBER FORPHYSICAL REVIEW E 78, 036308 2008
036308-7
different cells after a short time t0.060. Notice that the
particles separate as they approach a portion of the LCSf, the
boundary of the two cells. The above procedure can be ap-
plied to any other region sandwiched between the two types
of ridges and can be extended to more than two particle
sizes.
VI. ROBUSTNESS OF THE SENSITIVITY FIELD TO
PERTURBATIONS IN THE STREAM FUNCTION
The time-independent flow given by the stream function
in Eq. 26is perturbed by making it weakly time dependent.
The modified fluid flow is given by the stream function 15
x,y,t=acosx+bsin
tcos y.30
The velocity field is given by
u=−acosx+bsin
tsin y,31
v=asinx+bsin
tcos y.32
For time-varying vector fields the location of the LCSf
and LCSs depends on the choice of initial time. For the com-
putation of the sensitivity field, the locations of ridges in the
relative velocity subspace depend on the initial spatial coor-
dinates of the particle as well as the initial time. However,
our computations show that the dependence of the ridge
structure on the initial time is weak. Figure 5shows the
ridges in the sensitivity field. As the initial time is increased,
it is seen that there is a “squeezing” of the sensitivity field in
some regions of the relative velocity subspace. A comparison
of Fig. 4with Fig. 3shows that the ridge locations in the
sensitivity field remain qualitatively the same for the three
cases in Fig. 5where the initial time is small. This offers
numerical evidence that the sensitivity field is robust to small
perturbations in the fluid velocity.
VII. CONCLUSION
The dynamics of inertial particles in a fluid flow can ex-
hibit sensitivity to initial conditions. We demonstrated that
ridges in the relative velocity sensitivity field at each spatial
point effectively partition phase space into zones of different
particle fates, i.e., inertial particles initially located on either
side of a ridge will evolve to different spatial locations after
a short time. The phase space location of these ridges de-
pends on the Stokes number, and by implication the size of
the inertial particles of interest. This dependence can be ex-
ploited to make particles of different sizes cluster in different
regions of the fluid and thus separate and segregate them.
We used this method to achieve segregation using a
simple test model of two-dimensional flow: cellular flow. By
“injecting” a mixture of inertial particles of different sizes
into the fluid at a common relative velocity range that is
sandwiched between the ridges of different Stokes number,
the particles are segregated by size in a short time. Though
we have based our results on only cellular flow, the method-
ology presented only requires that the underlying flow 1
has a spatial partition, i.e., separatrices in the fluid itself and
2is of low Reynolds number. These requirements ensure
(
a)
(
b)
(
c)
FIG. 5. aRidges in the sensitivity field for the time-dependent
stream function
x,y,t=acosx+bsin
tcos yfor x0,y0
=3
/8,3
/8. The hatched and thick lines are the ridges corre-
sponding to St=0.2 and 0.1, respectively. Parameters: a= 100, b
=0.25,
=1, T= 0.24. Initial times t0=a0; b0.25; c0.5.
PHANINDRA TALLAPRAGADA AND SHANE D. ROSS PHYSICAL REVIEW E 78, 036308 2008
036308-8
that segregated particles do not remix. The method does not
rely on any other flow characteristic or specific stream func-
tion. Extending this to turbulent flows may present difficul-
ties, since the ridges in the sensitivity field may not persist
for long enough times to achieve a clean separation of the
inertial particles. This aspect requires further investigation.
In future work, the approach employed here can be adapted
to segregate non-neutrally-buoyant particles, and to segre-
gate particles by other characteristics, e.g., density and
shape, with a goal of designing flows that can fractionally
separate particles for a range of inertial parameters.
1G. Segre and A. Silberberg, Nature London189, 209 1961.
2M. Tirumkudulu, A. Tripathi, and A. Acrivos, Phys. Fluids 11,
507 1999.
3T. Shinbrot, M. M. Alvarez, J. M. Zalc, and F. J. Muzzio, Phys.
Rev. Lett. 86, 1207 2001.
4S. A. Isard, S. H. Gage, P. Comtois, and J. H. Russo, Bio-
Science 55, 851 2005.
5D. G. Schmale, A. S. Shah, and G. C. Bergstro, Phytopathol-
ogy 95, 472 2006.
6M. C. Facchini, M. Mircea, S. Fuzzi, and R. J. Charlson, Na-
ture London401, 257 1999.
7E. R. Abraham, Nature London391, 577 1998.
8T. Nishikawa, Z. Toroczkai, and C. Grebogi, Phys. Rev. Lett.
87, 038301 2001.
9A. Athanas and G. K. Johannes, Chem. Eng. Technol. 29,13
2006.
10L. R. Huang, C. E. Cox, R. H. Austin, and J. C. Sturm, Science
304, 987 2004.
11J. A. Davis, D. W. Inglis, K. J. Morton, D. A. Lawrence, L. R.
Huang, S. Y. Chou, J. C. Sturm, and R. H. Austin, Proc. Natl.
Acad. Sci. U.S.A. 103, 14779 2006.
12D. DiCarlo, D. Irimia, R. G. Tompkins, and M. Toner, Proc.
Natl. Acad. Sci. U.S.A. 104,482007.
13M. R. Maxey and J. J. Riley, Phys. Fluids 26, 883 1983.
14M. R. Maxey, Phys. Fluids 30, 1915 1987.
15A. Babiano, J. H. E. Cartwright, O. Piro, and A. Provenzale,
Phys. Rev. Lett. 84, 5764 2000.
16I. J. Benczik, Z. Toroczkai, and T. Tel, Phys. Rev. Lett. 89,
164501 2002.
17G. Haller and T. Sapsis, Physica D 237, 573 2008.
18G. Haller and T. Sapsis, Phys. Fluids 20, 017102 2008.
19J. Bec, Phys. Fluids 15, L81 2003.
20E. Balkovsky, G. Falkovich, and A. Fouxon, Phys. Rev. Lett.
86, 2790 2001.
21K. Duncan, B. Mehlig, S. Ostlund, and M. Wilkinson, Phys.
Rev. Lett. 95, 240602 2005.
22J. Bec, Phys. Fluids 18, 091702 2006.
23I. J. Benczik, Z. Toroczkai, and T. Tel, Phys. Rev. E 67,
036303 2003.
24S. C. Shadden, F. Lekien, and J. Marsden, Physica D 212, 271
2005.
25F. Lekien, S. C. Shadden, and J. E. Marsden, J. Math. Phys.
48, 065404 2007.
26K. Ide, D. Small, and S. Wiggins, Nonlinear Processes Geo-
phys. 9, 237 2002.
PARTICLE SEGREGATION BY STOKES NUMBER FORPHYSICAL REVIEW E 78, 036308 2008
036308-9
... with ϕ as the velocity potential and ψ as the real-valued flow streamfunction; the asterisk denotes complex conjugation. Subsequently, the inhaled particulate motion was analytically derived using a simplified version of the Maxey-Riley equation [35], in the two-dimensional vector form [36,37]: ...
Preprint
Full-text available
Could the microdroplets formed by viscoelastic stretching and break-up of mucosal liquids in the upper respiratory tract (URT), when inhaled further downwind, explain the brisk pace at which deep lung infections emerge following onset of initial infection at the URT? While it is well-established that particulates inhaled from outside can possibly penetrate to the lower airway only if they are < 5 microns, the fate of particulates (many > 5-microns in diameter) sheared away from the intra-URT mucosa during inhalation remains an open question. These particulates predominantly originate at the nasopharynx, oropharynx, and laryngeal chamber with the vocal folds. To resolve the posed question, this study considers a CT-based 3D anatomical airway reconstruction and isolates the tract from the laryngeal vocal fold region, mapping the entire tracheal cavity and concluding at generation 2 of the tracheobronchial tree. Through the delineated geometry, airflow simulation is conducted using the LES scheme to replicate relaxed inhalation at 15 L/min. Against the ambient air flux, numerical experiments have been performed to monitor the transport of liquid particulates with diameters 1-30 microns, bearing physical properties akin to aerosolized mucus with embedded virions. The full-scale numerical transmission trends to the lower airway were found consistent with the findings from a reduced-order mathematical model that conceptualized the impact of intra-airway vortex instabilities on local particle transport through point vortex idealization in an anatomy-guided 2D potential flow domain. The results collectively demonstrate markedly elevated trends of deep lung penetration by the URT-derived particulates, even if they are as large as 10- and 15 microns. The high viral load carried by such droplets to the bronchial spaces could mechanistically explain the accelerated seeding of infection in the lungs.
... Taking a dynamical systems point of view, several papers [12][13][14][15][16] , have predicted the clustering and size based segregation of particles in some canonical two and three dimensional fluid flows. The calculations in these papers are based on a simplification of the Maxey-Riley equation 17 , that governs the motion of a small spherical neutrally buoyant particle, ...
Article
Full-text available
Size based separation and identification of particles in microfluidics through purely hydrodynamic means has gained significant interest due to a number of possible biomedical applications. Curved micro-channels, particularly spiral micro-channels with rectangular cross-section and the dynamics of particles in such channels have been extensively researched to achieve size based separation of particles. In this paper we present evidence that sheds new light on the dynamics of particles in such curved channels; that a unique particle slip velocity is associated with the focusing positions in the cross sections, which leads to a balance of forces. Our experiments therefore imply that the forces acting on the particle lead to convergence to an attractor in both the physical space (the cross section of the channel) and the slip velocity space.
... We construct two examples of dynamical systems with invariant manifolds that despite being unstable at every point in the normal direction are nevertheless globally stable. The examples are inspired by some recent findings, [4], on the dynamics of inertial particles modeled using a simplified Maxey-Riley equation, [5][6][7]. The planar motion of inertial particles in a fluid generates a four dimensional dynamical system with the fluid streamlines forming the invariant manifold for a time independent fluid flow. ...
Article
Full-text available
We construct two examples of invariant manifolds that despite being locally unstable at every point in the transverse direction are globally stable. Using numerical simulations we show that these invariant manifolds temporarily repel nearby trajectories but act as global attractors. We formulate an explanation for such global stability in terms of the `rate of rotation' of the stable and unstable eigenvectors spanning the normal subspace associated with each point of the invariant manifold. We discuss the role of this rate of rotation on the transitions between the stable and unstable regimes.
... In this paper, we choose t 0 = 0 to compute the FTLE field for the 4body system, denoted as σ T . In order to compute the FTLE, following Tallapragada and Ross (2008); Gawlik et al. (2009);Ross et al. (2010), we set a regularly spaced rectilinear grid of tracers in a n-dimensional phase space to advect the grid of tracers forward in time by the fixed time T employing the Runge-Kutta-Fehlberg integrator, as in Press et al. (1992). To compute the FTLE numelicaly, we need to discretize ...
Article
Full-text available
In this paper, we show a low energy Earth–Moon transfer in the context of the Sun–Earth–Moon–spacecraft 4-body system. We consider the 4-body system as the coupled system of the Sun–Earth–spacecraft 3-body system perturbed by the Moon (which we call the Moon-perturbed system) and the Earth–Moon–spacecraft 3-body system perturbed by the Sun (which we call the Sun-perturbed system). In both perturbed systems, analogs of the stable and unstable manifolds are computed numerically by using the notion of Lagrangian coherent structures, wherein the stable and unstable manifolds play the role of separating orbits into transit and non-transit orbits. We obtain a family of non-transit orbits departing from a low Earth orbit in the Moon-perturbed system, and a family of transit orbits arriving into a low lunar orbit in the Sun-perturbed system. Finally, we show that we can construct a low energy transfer from the Earth to the Moon by choosing appropriate trajectories from both families and patching these trajectories with a maneuver.
... Ideal tracers in a flow should be subject to the same dynamics as the ambient fluid, yet because of their inertia, the density spheres used as particles do not behave as point-like tracers (Tallapragada & Ross, 2008). In this experiment, the particle tracers needed to be as close as possible to small, neutrally buoyant spheres, and thus had to fulfill three requirements: be of a small, spherical shape to mitigate inertia, yet big enough to be captured on camera for particle tracking; have a density between 1.26 g/cm 3 and 1.29 g/cm 3 to float at the interface between the layers; and be of a color such that they would be greatly contrasted on the recorded images. ...
Thesis
Describing transport in fluid flows has been a long-standing challenge in dynamical systems theory, with applications to industrial and natural flows. The detection of Lagrangian structures that stay coherent over time helps gain insight into the evolution of a system's dynamics and the fate of transport. Whereas most techniques to detect coherent structures rely on a dense velocity field, techniques based on sparse datasets are increasingly being developed. The braid theory approach to detect Lagrangian coherent structures from sparse sets of trajectories is tested through a periodic, two-dimensional Stokes flow, the rotor-oscillator flow. Combined theoretical and numerical studies have shown that this flow can offer chaotic regimes with islands of coherence. The flow was recreated experimentally in a laboratory based on the findings of these theoretical studies. The braid theory approach was found to successfully detect coherent groups from sparse trajectories, although it is very sensitive to the quality of that data available.
... Kuznetsov et al. (2002) have examined the role played by Lagrangian structures obtained in the Gulf of Mexico. In particular, they have analyzed data from the Colorado Recent studies (Tallapragada and Ross, 2008;Beron-Vera et al., 2015) have considered deviations in the evolution of drifters from that of purely advected particles by taking into account inertial effects produced by the buoyancy of the object and their finite size. Their approach considers the Maxey-Riley equation (Maxey and Riley, 1983) which holds for small rigid spheres. ...
Article
Full-text available
The disappearance of Malaysia Airlines flight MH370 on the morning of the 8 March 2014 is one of the great mysteries of our time. Perhaps the most relevant aspect of this mystery is that not a single piece of debris from the aircraft has been found. Difficulties in the search efforts, due to the uncertainty in the plane's final impact point and the time that has passed since the accident, bring the question on how the debris has scattered in an always moving ocean, for which there are multiple data sets that do not uniquely determine its state. Our approach to this problem is based on the use of Lagrangian Descriptors (LD), a novel mathematical tool coming from dynamical systems theory that identifies dynamic barriers and coherent structures governing transport. By combining publicly available information supplied by different ocean data sources with these mathematical techniques, we are able to assess the spatio-temporal state of the ocean in the priority search area at the time of impact and the following weeks. Using this information we propose a revised search strategy by showing why one might not have expected to find debris in some large search areas targeted by the Australian Maritime Safety Authority (AMSA), and determining regions where one might have expected impact debris to be located and that have not been subjected to any exploration.
... We consider the example of cell flow, [22,28] given by the stream function ...
Article
Full-text available
Finite sized particles exhibit complex dynamics that differ from that of the underlying fluid flow. These dynamics such as chaotic motion, size dependent clustering and separation can have important consequences in many natural and engineered settings. Though fluid streamlines are global attractors for the inertial particles, regions of local instability can exist where the inertial particle can move away from the fluid streamlines. Identifying and manipulating the location of the so called stable and unstable regions in the fluid flow can find important applications in microfluidics. Research in the last two decades has identified analytical criteria that can partition the fluid domain into locally stable and unstable regions. In this paper, we identify two new mechanisms by which neutrally buoyant inertial particles could exhibit globally stable dynamics in the regions of the fluid flow that are thought to be locally unstable and demonstrate this with examples. The examples we use are restricted to the simpler case of time independent fluid flows.
Preprint
Full-text available
Saharan dust events, having great ecological and environmental impacts, are the largest producers of the world's dust by far. Identifying the mechanisms by which the dust is transported across the Atlantic is crucial for obtaining a complete understanding of these important events. Of these events, the so-called ``Godzilla'' dust intrusion of June 2020 was the largest and most impactful in the last two decades and underwent a particularly interesting transport pattern. By uncovering dominant, organizing structures derived from the wind velocity fields, known as Lagrangian coherent structures, we demonstrate the ability to describe and qualitatively predict certain aspects related to the evolution of the dust plume as it traverses the atmosphere over the Atlantic. In addition, we identify regions of high hyperbolicity, leading to drastic changes in the shape of the plume and its eventual splitting. While these tools have been quite readily adopted by the oceanographic community, they have still yet to fully take hold in the atmospheric sciences and we aim to highlight some of the advantages over traditional atmospheric transport methods.
Article
Saharan dust events, having great ecological and environmental impacts, are the largest producers of the world’s dust by far. Identifying the mechanisms by which the dust is transported across the Atlantic is crucial for obtaining a complete understanding of these important events. Of these events, the so-called "Godzilla" dust intrusion of June 2020 was the largest and most impactful in the last two decades and underwent a particularly interesting transport pattern. By uncovering dominant, organizing structures derived from the wind velocity fields, known as Lagrangian coherent structures, we demonstrate the ability to describe and qualitatively predict certain aspects related to the evolution of the dust plume as it traverses the atmosphere over the Atlantic. In addition, we identify regions of high hyperbolicity, leading to drastic changes in the shape of the plume and its eventual splitting. While these tools have been quite readily adopted by the oceanographic community, they have still yet to fully take hold in the atmospheric sciences and we aim to highlight some of the advantages over traditional atmospheric transport methods.
Preprint
Full-text available
The problem of phase space transport, which is of interest from both the theoretical and practical point of view, has been investigated extensively using geometric and probabilistic methods. Two important tools to study this problem that have emerged in recent years are finite-time Lyapunov exponents (FTLE) and the Perron-Frobenius operator. The FTLE measures the averaged local stretching around reference trajectories. Regions with high stretching are used to identify phase space transport barriers. One probabilistic method is to consider the spectrum of the Perron-Frobenius operator of the flow to identify almost-invariant densities. These almost-invariant densities are used to identify almost invariant sets. In this paper, a set-oriented definition of the FTLE is proposed which is applicable to phase space sets of finite size and reduces to the usual definition of FTLE in the limit of infinitesimal phase space elements. This definition offers a straightforward connection between the evolution of probability densities and finite-time stretching experienced by phase space curves. This definition also addresses some concerns with the standard computation of the FTLE. For the case of autonomous and periodic vector fields we provide a simplified method to calculate the set-oriented FTLE using the Perron-Frobenius operator. Based on the new definition of the FTLE we propose a simple definition of finite-time coherent sets applicable to vector fields of general time-dependence, which are the analogues of almost-invariant sets in autonomous and time-periodic vector fields. The coherent sets we identify will necessarily be separated from one another by ridges of high FTLE, providing a link between the framework of coherent sets and that of codimension one Lagrangian coherent structures. Our identification of coherent sets is applied to three examples.
Article
Full-text available
Understanding the properties of clouds in the Earth's atmosphere is currently limited by difficulties at the fundamental level of adequately describing the processes of cloud droplet nucleation and growth. Small changes in droplet population may significantly influence cloud albedo as well as formation of precipitation. Models of cloud formation based on laboratory studies with idealized composition of nuclei suggest that organic solutes significantly lower surface tension-one of the parameters determining droplet population but the lack of data on composition and properties of the organic material in the atmosphere precludes realistic laboratory or model studies. Here, we report measurements on vacuum-evaporated samples of cloud water from the Po Valley, Italy, that show a large decrease in surface tension, by up to about one-third relative to pure water, for realistic concentrations of organic solutes expected to exist in growing droplets. Such large surface-tension changes, if they occur in cloud droplets near the critical size for nucleation, lead to an increase in droplet population and hence in cloud albedo. The error produced in ignoring this effect is estimated to be comparable to other calculated direct and indirect influences of aerosols on scattering and absorption of solar radiation.
Article
Full-text available
Aerial transport alone is seldom responsible for the introduction of nonindigenous species into distant regions; however, the capacity to use the atmospheric pathway for rapid spread in large part determines the invasive potential of organisms once they are introduced. Because physical and biological features of Earth's surface influence the routes and timing of organisms that use the atmospheric pathway, long-distance movement of aerobiota is largely regular and thus predictable. Soybean rust (Phakopsora pachyrhizi), potentially the most destructive foliar disease of soybean, recently invaded North America. The concepts presented in this article form the basis of the soybean rust aerobiology prediction system (SRAPS) that was developed to assess potential pathogen movement from South America to the United States. Output from SRAPS guided the scouting operations after the initial discovery of soybean rust in Louisiana. Subsequent observations of P. pachyrhizi in the southeastern United States provide validation of the modeling effort.
Article
Full-text available
Diffusive processes are often used to represent the formation of spatial patterns in biological systems. Here I show how patchiness may be generated in planktonic ecosystems through non-diffusive advection. Plankton distributions in oceanic surface waters can be characterized by the spectra of concentrations obtained along ship transects. Such spectra are inevitably found to have a power-law form over horizontal scales ranging from 1 to 100km (ref. 2). Phytoplankton have distributions similar to those of physical quantities such as sea surface temperature, with much less variability at shorter length scales. In contrast, zooplankton density may be almost as variable at short scales as long ones. Distributions of this form are generated in a model of the turbulent stirring of coupled phytoplankton and zooplankton populations. The characteristic spatial patterns of the phytoplankton and zooplankton are a consequence of the timescales of their response to changes in their environment caused by turbulent advection.
Article
This paper develops the theory and computation of Lagrangian Coherent Structures (LCS), which are defined as ridges of Finite-Time Lyapunov Exponent (FTLE) fields. These ridges can be seen as finite-time mixing templates. Such a framework is common in dynamical systems theory for autonomous and time-periodic systems, in which examples of LCS are stable and unstable manifolds of fixed points and periodic orbits. The concepts defined in this paper remain applicable to flows with arbitrary time dependence and, in particular, to flows that are only defined (computed or measured) over a finite interval of time. Previous work has demonstrated the usefulness of FTLE fields and the associated LCSs for revealing the Lagrangian behavior of systems with general time dependence. However, ridges of the FTLE field need not be exactly advected with the flow. The main result of this paper is an estimate for the flux across an LCS, which shows that the flux is small, and in most cases negligible, for well-defined LCSs or those that rotate at a speed comparable to the local Eulerian velocity field, and are computed from FTLE fields with a sufficiently long integration time. Under these hypotheses, the structures represent nearly invariant manifolds even in systems with arbitrary time dependence. The results are illustrated on three examples. The first is a simplified dynamical model of a double-gyre flow. The second is surface current data collected by high-frequency radar stations along the coast of Florida and the third is unsteady separation over an airfoil. In all cases, the existence of LCSs governs the transport and it is verified numerically that the flux of particles through these distinguished lines is indeed negligible.
Article
In an earlier paper, Maxey and Corrsin [J. Atmos. Sci. 43, 1112 (1986)] studied the motion of small aerosol particles settling under gravity through an infinite, periodic, cellular flow field subject to the effects of a Stokes drag force and inertia of the particles. Particle inertia was shown to have an important influence on the motion: No permanent suspension in the flow occurred, particles generally settled more rapidly than in still fluid, and the particle paths merged into isolated asymptotic trajectories. This study is continued for particles that are not necessarily much denser than the surrounding fluid but vary in density. Two basic responses are identified: an aerosol response for particles denser than the fluid, similar to that mentioned, and a bubble response for particles less dense. For both, particle accumulation is still a recurring feature. Results of numerical simulations are discussed, together with the stability of equilibrium points and the role of particle or fluid inertia.
Article
We derive analytic criteria for the existence of hyperbolic (attracting or repelling), elliptic, and parabolic material lines in two-dimensional turbulence. The criteria use a frame-independent Eulerian partition of the physical space that is based on the sign definiteness of the strain acceleration tensor over directions of zero strain. For Navier-Stokes flows, our hyperbolicity criterion can be reformulated in terms of strain, vorticity, pressure, viscous and body forces. The special material lines we identify allow us to locate different kinds of material structures that enhance or suppress finite-time turbulent mixing: stretching and folding lines, Lagrangian vortex cores, and shear jets. We illustrate the use of our criteria on simulations of two-dimensional barotropic turbulence.
Article
We identify a new instability when a horizontal Couette device is partially filled with a monodisperse suspension of neutrally buoyant spherical particles in a Newtonian liquid in which the particle concentration equals or exceeds 15%. When the inner cylinder of the horizontal Couette device is rotated and the outer cylinder is held fixed, the suspension separates itself into alternating bands of high and low particle concentration along the length of the device. The influence of the rotation rates, the suspension concentration, the fill levels and the gap width on the onset and the degree of segregation will be discussed.
Article
The forces on a small rigid sphere in a nonuniform flow are considered from first principles in order to resolve the errors in Tchen's equation and the subsequent modified versions that have since appeared. Forces from the undisturbed flow and the disturbance flow created by the presence of the sphere are treated separately. Proper account is taken of the effect of spatial variations of the undisturbed flow on both forces. In particular the appropriate Faxen correction for unsteady Stokes flow is derived and included as part of the consistent approximiation for the equation of motion.
Article
Detailed, high-resolution numerical simulations have been performed of the buoyancy-driven motion of deformable, chemically reacting bubble and droplet swarms. Mass transfer rates and chemical reaction selectivities were determined and a comparison is presented between the results for bubble/droplet swarms and single bubbles. The mixing in the wake of bubbles was characterized as well. It was shown that for mixing-sensitive reaction networks, the hydrodynamics of the bubble swarm may significantly impact the reaction selectivity. Four special cases are highlighted, i.e., highly exothermic reactions, heterogeneously catalyzed reactions, animal cell cultures and experiments. The most important outcome of this work is that bubble swarms and the impact of the swarm hydrodynamics on reacting systems can be studied in detail. This is the first study of this kind reported in the literature. Its importance is paramount as knowledge of the unique flow dynamics inside bubble swarms is crucial to understanding the mechanisms controlling mass transport and chemical reactions and is a prerequisite for effective process intensification.