ArticlePDF Available

Particle-Turbulence Interaction in a Homogeneous, Isotropic Turbulent Suspension

Authors:

Abstract and Figures

A review is given of numerical, analytical, and experimental research regarding the two-way coupling effect between particles and fluid turbulence in a homogeneous, iso- tropic turbulent suspension. The emphasis of this review is on the effect of the suspended particles on the spectrum of the carrier fluid, in order to explain the physical mechanisms that are involved. An important result of numerical simulations and analytical models (neglecting the effect of gravity) is that, for a homogeneous and isotropic suspension with particles with a response time much larger than the Kolmogorov time scale, the main effect of the particles is suppression of the energy of eddies of all sizes. However for a suspension with particles with a response time comparable to or smaller than the Kol- mogorov time, the Kolmogorov length scale will decrease and the turbulence energy of (nearly) all eddy sizes increases. For a suspension with particles with a response time in between the two limiting cases mentioned above the energy of the larger eddies is sup- pressed, whereas the energy of the smaller ones is enhanced. Attention is paid to several physical mechanisms that were suggested in the literature to explain this influence of the particles on the turbulence. In some of the experimental studies, certain results from simulations and models have, indeed, been confirmed. However, in other experiments these results were not found. This is attributed to the role of gravity, which leads to turbulence production by the particles. Additional research effort is needed to fully un- derstand the physical mechanisms causing the two-way coupling effect in a homoge- neous, isotropic, and turbulently flowing suspension. This review contains 47 references. DOI: 10.1115/1.2130361
Content may be subject to copyright.
Christian Poelma
Gijs Ooms1
e-mail: g.ooms@wbmt.tudelft.nl
Laboratory for Aero- and Hydrodynamics,
J.M. Bergerscentrum,
Delft University of Technology,
Mekelweg 2,
2628 CD Delft, The Netherlands
Particle-Turbulence Interaction
in a Homogeneous, Isotropic
Turbulent Suspension
A review is given of numerical, analytical, and experimental research regarding the
two-way coupling effect between particles and fluid turbulence in a homogeneous, iso-
tropic turbulent suspension. The emphasis of this review is on the effect of the suspended
particles on the spectrum of the carrier fluid, in order to explain the physical mechanisms
that are involved. An important result of numerical simulations and analytical models
(neglecting the effect of gravity) is that, for a homogeneous and isotropic suspension with
particles with a response time much larger than the Kolmogorov time scale, the main
effect of the particles is suppression of the energy of eddies of all sizes. However for a
suspension with particles with a response time comparable to or smaller than the Kol-
mogorov time, the Kolmogorov length scale will decrease and the turbulence energy of
(nearly) all eddy sizes increases. For a suspension with particles with a response time in
between the two limiting cases mentioned above the energy of the larger eddies is sup-
pressed, whereas the energy of the smaller ones is enhanced. Attention is paid to several
physical mechanisms that were suggested in the literature to explain this influence of the
particles on the turbulence. In some of the experimental studies, certain results from
simulations and models have, indeed, been confirmed. However, in other experiments
these results were not found. This is attributed to the role of gravity, which leads to
turbulence production by the particles. Additional research effort is needed to fully un-
derstand the physical mechanisms causing the two-way coupling effect in a homoge-
neous, isotropic, and turbulently flowing suspension. This review contains 47
references. DOI: 10.1115/1.2130361
1 Introduction
The occurrence of dispersed two-phase flows in nature and in-
dustrial applications is abundant. Despite a significant research
interest, there are still several open questions in this topic, limiting
a good quantitative prediction of these flows. In this publication, a
review is given of recent progress relating to one of the more
fundamental problems: the so-called two-way coupling between
the dispersed phase and turbulence. Two-way coupling refers to
the effects of the fluid on the particles and vice versa. The review
is limited to rigid particles in homogeneous and isotropic turbu-
lence, in order to avoid further complications. For the same rea-
son, no work on shear flows is incorporated. Even though inclu-
sion of these publications would greatly increase the amount of
available data, it would also significantly complicate the analysis,
since, among other phenomena, shear-induced turbulence produc-
tion and particle migration might obscure other effects.
Because of the growth of computer power, detailed direct nu-
merical simulations DNSof the behavior of particles in the tur-
bulent fluid velocity field and of the two-way coupling effect be-
came possible. Also, theoretical models were developed that try to
capture the main physical mechanisms occurring in turbulent sus-
pensions. Moreover, because of the progress in experimental tech-
niques, new and accurate measurement results became available.
We will give a review of publications about the two-way cou-
pling effect in turbulent suspensions that appeared in the literature
during the last decade. It is, of course, impossible to discuss all
important publications. However, we hope that the review will
give a clear picture of the progress in understanding of turbulent
suspensions. Attention will be paid to numerical simulations, the-
oretical models, and experiments. Emphasis will be placed on the
influence of the particles on the turbulent kinetic energy spectrum
of the carrier fluid. We know that turbulence spectra are not suf-
ficient to characterize turbulence completely. However, in the lit-
erature particular attention was given to the influence of particles
on the turbulence spectrum in order to derive a physical under-
standing of the two-way coupling effect.
In the first part of this review we will integrate the knowledge
from numerical simulations with respect to the dependence of the
turbulence spectrum of the carrier fluid on the presence of the
particles in the fluid. In the second part such an integration will be
given with respect to theoretical explanations given in the litera-
ture on the two-way coupling effect, in general, and the depen-
dence of the turbulence spectrum on the particles, in particular.
The third part is devoted to a comparison between experimental
data and results from numerical simulations and theoretical
models.
Several good reviews about particle-laden turbulent flows were
published during the past decade, see for instance Hetsroni 1,
Elghobashi 2, Crowe et al. 3, and Mashayek and Pandya 4.
Our review is different, in our opinion, in the sense that we pay
particular attention to the physical mechanisms that play an im-
portant role in the interaction between particles and fluid turbu-
lence in a turbulently flowing suspension. In recent years some
interesting publications have been written in which detailed infor-
mation is given about these mechanisms.
2 Relevant Parameters
We define a number of parameters that are of importance for
turbulently flowing suspensions and that will be used in this re-
view. The volume fraction of the dispersed phase the particlesis
defined as
1To whom correspondence should be addressed.
Transmitted by Assoc. Editor A. Tsinober.
78/Vol. 59, MARCH 2006 Copyright © 2006 by ASME Transactions of the ASME
=
Vd
V1
where
Vdis the volume of the dispersed phase in the volume
V
of the suspension assumed large enough to ensure averaging.
The ratio of the average particle spacing dspacingand the particle
diameter dpis related to the volume fraction by
dspacing
dp
=
6
1/3
2
The particle mass loading is given by
=
p
f
3
in which
pand
fare the densities of the particles and the carrier
fluid, respectively. The particle response time to changes in the
surrounding fluid is expressed by the Stokes time
p=2
pdp
2
9
f
4
where
is fluid kinematic viscosity. For the fluid phase the Kol-
mogorov time is defined by
k=
2/
5
with
the Kolmogorov length scale. The integral time scale is
given by
=/u6
in which is the integral length scale of turbulence and uthe
root-mean-square value of the turbulent velocity fluctuations. The
Stokes number is defined in this review as the ratio of the particle
time scale and the Kolmogorov time scale
St =
p/
k7
For St0, the particle behaves as a fluid tracer. For St, the
particle is unresponsive to the fluctuations of the flow. Obviously,
the physically most interesting situation occurs when it ap-
proaches unity: particles follow the large scale fluid motions.
Finally, the flow regime around the particles can be character-
ized by means of a Reynolds number
Rep=uTVdp
8
In this equation, uTV represents the terminal velocity the ve-
locity a particle attains falling in a quiescent medium. Alterna-
tively, it can be defined using, e.g., the root-mean-square value of
the free-stream turbulence. In has been postulated by Hetsroni 1
that Repcan be used to determine whether particles attenuate
Rep100or enhance Rep400the fluid turbulence level.
3 Direct Numerical Simulations
3.1 Effect of Particles on Turbulence Spectrum. An early
indication about the influence of the particles on the turbulence
spectrum was given by Squires and Eaton 5,6. They used direct
numerical simulation to study statistically stationary, homoge-
neous, isotropic turbulence. They considered particle motion in
the Stokes regime. Gravitational settling was neglected. Compu-
tations were performed using both 323and 643grid points. The
particles were treated as point particles. To achieve the stationary
flow, a steady nonuniform body force was added to the governing
equations. Particle sample sizes up to 106were used in the simu-
lations. Mass loadings of 0.1, 0.5, and 1.0 were considered. They
used the following values for the ratio of the particle response
time
pand the Kolmogorov time scale of turbulence
k:
p/
k=0.75, 1.4, 1.5, 5.2, 7.5, and 15.0. Squires and Eaton
calculated the effect of the mass loading on the spatial turbulent
kinetic energy spectrum of the carrier fluid. The dimensionless
spatial spectrum for
p/
k=1.5 as function of the dimensionless
wave number for different mass loadings is shown in Fig. 1. E
is the turbulent energy spectrum as function of wave number. As
mentioned earlier
is the Kolmogorov length and
the mass
loading. The spectra are normalized using q2, the total turbulent
kinetic energy for each particular case. It can be seen that with
increasing mass loading, the energy at large wave numbers in-
creases relative to the energy at small wave numbers. As Squires
and Eaton found that the total turbulent energy decreases with
increasing mass loading, it can be concluded that at small wave
numbers where most of the energy is locatedthe energy de-
creases with increasing mass loading compared to the particle-free
case. Whether the energy at large wave numbers decreases or
increases with respect to the particle-free case cannot directly be
concluded from the results of Squires and Eaton. For that it would
be necessary to multiply the spectra of Fig. 1 with the total energy
q2. The relative increase of the distribution of energy at large
wave numbers with respect to the energy at small wave numbers
was found for all particle response times used in the simulations.
More details about the influence of particles on the spatial tur-
bulence spectrum of the fluid became available via the work of
Elghobashi and Truesdell 7. In contrast to Squires and Eaton,
they examined turbulence modulation by particles in decaying iso-
tropic turbulence. A point-particle approximation was again made.
They used the particle equation of motion derived by Maxey and
Riley 8. Like Squires and Eaton they found that the coupling
between the particles and fluid results in an increase in the turbu-
lent energy at the large wave numbers relative to the energy at
small wave numbers. Moreover, they concluded from their calcu-
lations that the large wave number energy for a flow with particles
with a sufficient small response timeis even larger than the large
wave number energy for the particle-free case at the same time in
the decay process.
Like Squires and Eaton, also Boivin et al. 9made a detailed
DNS study of the modulation of statistically stationary, homoge-
neous, and isotropic turbulence by particles. Gravitational settling
was neglected and the particle motion was assumed to be gov-
erned by drag. The ratio of the particle response time to the Kol-
mogorov time scale had the following values: 1.26, 4.49, and
11.38, and the particle mass loading was equal to: 0.2, 0.5, and
Fig. 1 Effect of mass loading on spatial energy spectra for
p/
k=1.5. With increasing mass loading the energy at large
wave numbers increases relative to the energy at small wave
numbers. From Ref. 5.
Applied Mechanics Reviews MARCH 2006, Vol. 59 /79
1.0. The velocity field was made statistically stationary by forcing
the small wave numbers of the flow. Again the effect of particles
on the turbulence was included by using a point-force approxima-
tion. Fluid turbulence spatial energy spectra, derived by Boivin et
al. 9for different mass loadings are shown in Fig. 2 for
p/
k
=1.26 and in Fig. 3 for
p/
k=11.38. Note that the figures as they
are represented here are replotted with double-logarithmic axes, to
facilitate comparison to the other figures. In their simulations the
dimensionless maximum value of kis about equal to the number
of grid points in each direction of their cubic simulation domain.
So the region around k=1 represents the energy-containing eddies
and we can say that the wave number along the horizontal axis is
made dimensionless by means of the integral length scale. In gen-
eral, there is a similar damping effect on the smaller wave num-
bers of the fluid turbulence by both particles with a large response
time and particles with a small response time. At the larger wave
numbers the turbulence kinetic energy is attenuated by particles
with a large response time, but increased by particles with a small
response time. It should be noted that in Figs. 1–3 the highest
wave numbers appear to be contaminated with noise, since they
show a distinct “plateau” Fig. 1or even increase Figs. 2 and 3
at these wave numbers. Nevertheless, even if this part of the graph
is ignored, the trends mentioned above remain.
A next step in achieving information about the dependence of
the turbulence spectrum on the presence of particles in a turbulent
suspension was provided by Sundaram and Collins 10. Like El-
ghobashi and Truesdell they performed DNS simulations of
particle-laden isotropic decaying turbulence. The particle response
time was in the range 1.6
p/
k6.4. The ratio of the particle
density and fluid density was of the order 103. The drag force on
the particles was described by Stokes law, and the influence of the
gravitational force was neglected. The DNS results showed again
that the turbulent energy spectrum of the fluid is reduced at small
wave numbers and increased at large wave numbers compared to
the particle-free caseby the two-way coupling effect. They also
concluded that the location of the cross-over point the wave num-
ber where the influence of the particles changes from a
turbulence-damping effect to a turbulence-enhancing oneis
shifted toward larger wave numbers for larger values of the par-
ticle response time
p.
In their DNS calculations Ferrante and Elghobashi 11fixed
both the volume fraction
=10−3and mass fraction
=1for
four different types of particles, classified by their ratio of the
particle response time and the Kolmogorov time scale of turbu-
lence. The ratio
p/
khad the values 0.1, 0.25, 1.0, and 5.0. The
ratio of the particle density
pand the fluid density
fis 103.
In Fig. 4 the spatial turbulent energy spectrum Et,
for the
carrier fluid in the suspension is given for the case without grav-
ityat a certain moment during the decay process.
is the wave
number made dimensionless by means of the the integral length
scale L. In the figure the result indicated by case A is for the
particle-free flow, the results indicated by cases B, C, D, and E are
for the carrier fluid in the suspension with particles of increasing
response time
p/
k=0.1, 0.25, 1.0, and 5.0, respectively. Micro-
particles case Bincrease Et,
relative to the particle-free flow
case Aat wave numbers
12 and reduce Et,
relative to
case A for
12, such that Et=Et,
d
in case B is larger
than in case A the particle-free case. Also for the cases C, D, and
E, the particles dampen the turbulence at small wave numbers
compared to the particle-free flow and enhance the turbulence at
Fig. 2 Turbulence kinetic energy spectrum for
p/
k=1.26:
=0 solid line,0.2dotted,0.5dashed, and 1.0 dashed-
dotted. With increasing mass loading the energy at large wave
numbers is enhanced relative to the energy at small wave num-
bers. Data taken from Ref. 9.
Fig. 3 Turbulence kinetic energy spectrum for
p/
k=11.38:
=0 solid line,0.2dotted,0.5dashed, and 1.0 dashed-
dotted. With increasing mass loading the energy at all wave
numbers decreases. Data taken Ref. 9.
Fig. 4 Kinetic energy spectrum of the carrier fluid at t=5.0.
The cross-over wave number increases with increasing particle
response time. From Ref. 11.
80/Vol. 59, MARCH 2006 Transactions of the ASME
large wave numbers. However the cross-over wave number in-
creases with increasing particle response time. As can be seen
from Fig. 4 large particles case Econtribute to a faster decay of
the turbulent kinetic energy by reducing the energy content at
almost all wave numbers, except for
87, where a slight in-
crease of Et,
occurs.
3.2 Physical Mechanisms. In the publication of Squires and
Eaton 6, a first proposal is made to explain the physical mecha-
nism responsible for the nonuniform distortion of the turbulence
energy spectrum by particles. In their opinion this nonuniform
distortion is due to the preferential concentration of particles in
the turbulent flow field. They showed that particles with a small
response time
p/
1exhibit significant effects of preferential
concentration in regions of low vorticity and high strain rate. The
effect of high concentrations of particles in these regions leads to
an increase in small-scale turbulent velocity fluctuations. This pro-
duction of small-scale fluctuations subsequently causes the vis-
cous dissipation rate in the carrier fluid to be increased for par-
ticles with a small response time. Squires and Eaton also showed
that preferential concentration causes a significant disruption of
the balance between production and destruction of dissipation,
again leading to a selective modification of the turbulence spec-
trum. More research is, in our opinion, needed to fully understand
the contribution of preferential concentration to the two-way cou-
pling effect.
Boivin et al. 9found that in a turbulently flowing suspension
the cascade process energy transport from the large to the small
eddiesis influenced by the particles. They calculated the spec-
trum of the fluid-particle energy exchange rate. In the small wave
number part of this spectrum the turbulent fluid motion transfers
energy to the particles, i.e., the particles act as a sink of kinetic
energy. At larger wave numbers of the spectrum the energy ex-
change rate is positive, indicating that particles are capable of
adding kinetic energy to the turbulence. This energy, “released”
by the particles, is not immediately dissipated by viscous effects
but is, in fact, responsible for the relative increase of small-scale
energy compared to the particle-free case observed in the energy
spectra for particles with a small response time.
A different type of physical mechanism for the two-way cou-
pling effect was proposed in the publication by Ferrante and Elg-
hobashi 11. They provided explanations for the behavior of mi-
croparticles and large particles. Because of their fast response to
the turbulent velocity fluctuations of the carrier fluid, the micro-
particles are not ejected from the vortical structures of their initial
surrounding fluid. The inertia of the microparticles causes their
velocity autocorrelation to be larger than that of the surrounding
fluid. Since the microparticles’ trajectories are almost aligned with
fluid points’ trajectories, and their kinetic energy is larger than that
of the surrounding fluid, the particles will transfer part of their
own energy to the fluid. The idea is that the microparticles accel-
erate quickly in an eddy to the velocity of the surrounding fluid.
After that acceleration period the small but finite inertia of the
particles causes, on average, a larger kinetic energy for the par-
ticles than the surrounding fluid, as the particles tend to retain
their velocity for a longer time. Much more quantitative details,
based on their DNS calculations, are given in the publication by
Ferrante and Elghobashi.On the other hand the microparticles
increase the viscous dissipation rate relative to that of the particle-
free flow. The reason is that the microparticles remain in their
initially surrounding vortices, causing these vortical structures to
retain their initial vorticity and strain rates longer than for the
particle-free flow. Again, it is the small but finite inertia of the
particles that causes this effect.The net effect is positive for the
turbulent kinetic energy of the carrier fluid, as the gain in energy
due to the transfer of energy from the particles is larger than the
increase in viscous dissipation. The DNS calculations also show
that the microparticles directly interact with the small scales of
motion, augmenting their energy content. The triadic interaction
of wave numbers then alters the energy content of the other scales
of motion, such that after few integral time scales the energy
spectrum is modified at all the wave numbers as compared to the
particle-free case.
For large particles the explanation is different: because of their
significant response time, large particles do not respond to the
velocity fluctuations of the surrounding fluid as quickly as micro-
particles do, but rather escape from their initial surrounding fluid
crossing the trajectories of fluid points. Large particles retain
their kinetic energy longer than the surrounding fluid. However,
because of the “crossing trajectories” effect, the fluid velocity au-
tocorrelation is larger than the correlation between the particle
velocity and the fluid velocity, causing a transfer of energy from
the fluid to the particles. On the other hand, large particles reduce
the lifetime of eddies, causing a viscous dissipation rate which is
smaller than for the particle-free flow.It is shown that large
particles interacting with a clockwise vortex create a counter-
clockwise torque on the fluid, which, in turn, reduces the vortic-
ity.The net result of the two opposing effects is a reduction of
turbulent kinetic energy for a suspension with large particles at
nearly all wave numbers relative to the kinetic energy for the
particle-free turbulent flow. Again for more quantitative details the
publication of Ferrante and Elghobashi should be consulted. We
think that their analysis is a very interesting and important step in
the direction of a physical understanding of the two-way coupling
effect.
3.3 Effect of Finite Particle Size. Numerical work in which
the particles are fully resolved is slowly becoming available in the
literature. An example of this type is the work by ten Cate 12,13,
who carried out numerical simulations of a homogeneous and
isotropicturbulent suspension taking into account the finite size
of the particles by satisfying the no-slip boundary condition at the
particle-fluid interface. Nevertheless, the short-range interactions
between the particles had to be added explicitly, since they could
not be resolved on the grid. For the generation of sustained tur-
bulent conditions a spectral forcing scheme was implemented us-
ing the lattice-Boltzmann technique. In these simulations the par-
ticle volume concentration is varied between 2% and 10% which
is probably no longer in the two-way coupling regimeand the
particle to fluid density ratio was between 1.15 and 1.73. The
Taylor-scale Reynolds number was 61. Results were presented
concerning the influence of the particle phase on the turbulent
energy spectrum. Fluid motion was generated at length scales in
the range of the particle size, which resulted in a strong increase
in the rate of energy dissipation at the small length scales. With
respect to the turbulent energy spectrum, little difference was
found between the spectra with and without particles at the lowest
wave numbers. At the intermediate wave numbers the particles
reduced the fluid kinetic energy see Fig. 5. At the larger wave
numbers the spectra crossed at a clear cross-over or pivot point,
and the particles increased the kinetic energy with respect to the
particle-free case.
In contrast to suspensions studies, an alternative is the very
fundamental approach of studying a single particle. Examples of
this type of work, in which a fully resolved particle in a turbulent
flow is studied, can be found in, e.g., publications by Mittal 14
and Bagchi and Balachandar 15. In the former work, the produc-
tion of turbulence due to vortex shedding is studied. It is found
that this can contribute to turbulence production when Rep300,
yet only when the free-stream turbulence level is sufficiently low.
In the work by Bagchi and Balachandar, various forces inertial,
viscous, historyare studied, in detail, in homogeneous and strain-
ing flow. This type of work may also clarify to what level the
point-force approximation, that is used in most numerical and
theoretical work, is valid.
3.4 Discussion of DNS Results. Some general conclusions
can be drawn from the literature on numerical simulations of tur-
bulently flowing suspensions. In the numerical simulations, the
Applied Mechanics Reviews MARCH 2006, Vol. 59 /81
effect of the turbulence generation by the particle wakes and by
the vortices shed by the particles was not taken into account. It
would also have been difficult to include this effect of turbulence
generation, as the particles in these publications were treated as
point-particles apart from the work by ten Cate. This is a sig-
nificant simplification, and the results should thus be interpreted
with care. From the simulations it can be concluded that for a
suspension with particles with a response time much larger than
the Kolmogorov time scale, the main effect of the particles is
suppression of the energy of eddies of all sizes at the same energy
input into the suspension as for the particle-free case. So for such
a suspension the total turbulent energy of the carrier fluid will be
smaller than the total turbulent energy of the fluid for the particle-
free case at the same energy input. However, for a suspension
with particles with a response time comparable to or smaller than
the Kolmogorov time, the Kolmogorov length scale will decrease
and the turbulence energy of nearlyall sizes increases. In that
case the total turbulent energy of the carrier fluid can be larger
than the total turbulent energy of the fluid for the particle-free
case. For a suspension with particles with a response time in be-
tween the two limiting cases mentioned above, the energy of the
larger eddies is suppressed, whereas the energy of the smaller
ones is enhanced. Several physical mechanisms have been pro-
posed to explain these effects. It is not possible, at the moment, to
decide which mechanism is the most important one, or whether
they all contribute to explain the influence of particles on the
turbulence of the carrier fluid in a suspension. More research is
needed on this subject. It is also desirable to extend the DNS work
with finite-size particles to investigate, in detail, the important
influence of turbulence generation by particle wakes and vortices.
4 Theoretical Models
4.1 Introduction. The starting point for analytical models de-
scribed in the literature is often the Navier-Stokes NSequation
for the velocity of the carrier fluid ut,rwith external forces
f
t+u·
2
u+p=fp+f9
pt,ris the pressure and
fis the fluid density. The random
vector field ft,rrepresents the stirring force responsible for the
maintenance of the turbulent flow. The equation includes also the
force fpt,rcaused by the friction of the fluid with the particles,
which is often approximated by
fpt,r=
f
p
vt,rut,r兲兴 共10
Here vt,ris the velocity field of the particles, considered as a
continuous medium with density mp/l3=
f
, where mpis the
mass of a particle, l3suspension volume per particle, and
the
mass loading parameter
=mp/
f311
The validity to represent fpt,rin the form of Eqs. 10and 11
is based on the assumption of space homogeneity of the particle
distribution. If the particles are not homogeneously distributed,
one can add additional equations, e.g., the number density. It is
assumed that the particles are small enough for the Stokes drag
law to be valid. For the simple case of monosize particles moving
under the influence of the Stokes drag, the equation of motion for
the particles considered as a continuous phasehas the following
form see, for instance, 16兴兲
mp
l3
t+v·
v=−fp12
4.2 Physical Mechanisms.
4.2.1 Two-Fluid Models. The treatment mentioned above of
the particle phase as a continuum, i.e., the assumption of two
interpenetrating fluids, is the basis for the so-called two-fluid
models. Many publications have been written about the two-way
coupling effect using a version of the two-fluid model, and good
reviews about this type of research are available see, for instance,
4兴兲. As mentioned before, we will concentrate on publications in
which special attention is given to the influence of the particles on
the turbulence in a homogeneous, isotropic suspension and in
which an attempt is made to understand the underlying physical
mechanism.
The equations given above were used by Baw and Peskin 17
to derive a set of “energy-balance” equations for the following
functions:
Effk—energy spectrum of the fluid turbulence Ekin the
nomenclature used in other publications
Eff,pk—energy spectrum of the fluid turbulence along a
particle trajectory
Efpk—fluid-particle covariance spectrum
Eppk—particle energy spectrum
In the balance equations, the following energy transfer functions
occur:
Tff,fk—energy transfer in fluid turbulence
Tfp,fk—transfer of fluid-particle correlated motion by the
fluid turbulence along the particle path
Tfp,pk—transfer of fluid-particle correlated motion by the
particles
Tppk—transfer of particle-particle correlated motion by the
particle motion
q,fk—fluid-particle energy exchange rate.
Baw and Peskin made a set of simplifying assumptions in order to
be able to analyze the balance equations. First, they assumed that
the particles do not respond to the fluid velocity fluctuations due
to their very largeinertia. Therefore
Eff,pk=Effk兲共13
and
Fig. 5 Scaled energy spectra E=Ek/
2/3
5/3…… of the single-
phase simulations dottedand two-phase simulations con-
tinuous. The wavenumber is made dimensionless with the par-
ticle wave number: kp=2
/dp.
=0.005,
p/
f=1.414, StK
=0.207. Data taken from Ref. 13.
82/Vol. 59, MARCH 2006 Transactions of the ASME
Tfp,fk=Tfp,pk=Tpp,pk=0 14
This assumption is, of course, not realistic for particles satisfying
the Stokes’ approximation. Their next assumption
q,f=
EfpkEff,pk兲兴/
p15
may be understood as a statement that the fluid-particle exchange
rate is statistically the same for all scales characterized by a
k-independent frequency
p=
/
p. This is reasonable for par-
ticles with very large inertia, but then Stokes law is not valid. For
particles satisfying Stokes law, assumption Eq. 15has to be
replaced with a more realistic, k-dependent frequency
pk.
A serious difficulty in the derivation of the energy-balance
equations is how to find a closure expression for third-order ve-
locity correlation functions, responsible for the various energy
transfer functions. Baw and Peskin assumed that Tff,fkcan be
expressed similarly as in the case of a pure single phaseflow
Tff,fk=− d
dk
f
1/3k5/3Effk
16
where
fis the viscous dissipation in the pure fluid without par-
ticlesand
is the so-called Kolmogorov constant. This assump-
tion is questionable. According to the spirit of the Richardson-
Kolmogorov cascade picture of turbulence, one may express
inertial range objects, like Tff,fkin terms of again inertial range
quantities, like k,Effk, and
k兲共the energy flux in kspace.In
a single-phase flow, indeed
k=
f. However, this is not the case
for a turbulent suspension due to the fluid-particle energy ex-
change, given by Eq. 15. With this simplified model, Baw and
Peskin predicted the following influences on the energy spectrum
of the fluid turbulence due to the particles:
a decrease of the energy in the energy-containing range of
the spectrum
an increase in the inertial range of the spectrum
a decrease in the viscous dissipation range.
Boivin et al. 9used the same model as Baw and Peskin 17.
They also applied assumptions similar to Eqs. 15and 16. For-
tunately, they took into account the response of the particles to the
turbulent velocity fluctuations by relaxing assumptions of Eqs.
13and 14and also accounted for the very important physical
effect of the energy dissipation due to the drag around the par-
ticles. For that reason they approximated Tff,fkand Tfp,fkas
follows:
Tff,fk=− d
dk
fq,fk兲兴1/3k5/3Effk兲共17
and
Tfp,fk=− d
dk
fq,fk兲兴1/3k5/3Efpk兲共18
Note that this closure has the same weakness as Eq. 16, involv-
ing the dissipation range value
f. With the above-described
changes with respect to the model as developed by Baw and Pe-
skin, Boivin et al. found an increase in the viscous dissipation
range of the fluid turbulence spectrum for small values of the
particle response time
p.
4.2.2 Single-Fluid Models. Some analytical models were de-
veloped, in which the turbulent suspension was treated as a single
fluid with effective frequency- and wave-number-dependent
physical properties. Felderhof and Ooms 18兴共see also the
follow-up publications Ooms and Jansen 19and Ooms et al.
20兴兲 developed an analytical model for the dynamics of a suspen-
sion of solid spherical particles in an incompressible fluid based
on the linearized version of the Navier-Stokes equation. In par-
ticular, the effect of the particles-fluid interaction on the effective
transport coefficients and on the turbulent energy spectrum of the
suspension was studied. Also the hydrodynamic interaction be-
tween the particles and the influence of the finite size of the par-
ticles were incorporated. However, it is needless to say that the
nonlinearity of the Navier-Stokes equation is of crucial impor-
tance in the problem of turbulence. The authors were well aware
of this problem, but as mentionedwanted to study, in particular,
the influences of the particle-particle hydrodynamic interaction
and of the finite particle size at a relatively high particle volume
concentration. In order to improve the turbulence modeling, they
included in one of their publications a wave-number-dependent
turbulence viscosity. However, as the turbulence cascade process
was not properly accounted for, they never found an increase of
the turbulent kinetic energy at large wave numbers for particles
with a small response time as found by numerical simulations
and some experiments. The importance of this work lies in the
description of the particle-particle hydrodynamic interaction at
high particle volume fractions and of the finite particle size.
L’vov et al. 21develop a one-fluid theoretical model for a
stochastically stationary, homogeneous, isotropic turbulently
flowing suspension. It is based on a modified Navier-Stokes equa-
tion with a wave-number-dependent effective density of suspen-
sion and an additional damping term representing the fluid-
particle friction described by Stokes’ law. It can be considered as
an improvement of the work of Felderhof and Ooms because in
their model, L’vov et al. incorporated a description of the cascade
process of turbulence. Ooms and Poelma 22extended the theo-
retical model in such a way that it can be applied to a decaying,
homogeneous, and isotropic turbulent suspension and can be com-
pared to the DNS data of Ferrante and Elghobashi 11. They
calculated, for instance, the energy spectra E
for the carrier
fluid in the suspension for the five cases A, B, C, D, and E
discussed by Ferrante and Elghobashi and compared the predic-
tions with their DNS results. Here only the results are shown for
the particle-free flow case A, the microparticles case B, and the
large particles case E. The results for the other particles cases C
and Dare in between those for cases B and E. For a certain
moment in the decay process, the results are shown in Fig. 6.
There is a difference between model predictions Fig. 6and the
DNS results Fig. 4at small values of
1. This is due to a
difference in the boundary condition at the small wave-number
end between the two methods. It is clear from Fig. 6 that the
particles dampen the turbulence for small values of
large ed-
diesand enhance the turbulence for large values of
small
Fig. 6 Turbulent kinetic energy spectrum of the carrier fluid
É1.5. The cross-over wave number increases with increas-
ing particle response time in accordance with numerical simu-
lations. From Ref. 22.
Applied Mechanics Reviews MARCH 2006, Vol. 59 /83
eddies. However, there is a difference. The microparticles case
Benhance the turbulence over a much larger range of
values
than the large particles case E. For microparticles the enhance-
ment is so strong that the total energy over all eddies is larger than
for the particle-free flow. That is not the case for the large par-
ticles. The cross-over wave number the wave number where the
influence of the particles changes from a turbulence-damping ef-
fect to a turbulence-enhancing oneincreases with increasing par-
ticle response time. This result is the same as found in the DNS
calculations.
The following physical mechanism is proposed by L’vov et al.
21to explain the obtained results. According to them an impor-
tant effect of the particles is, that they increase the effective den-
sity of the suspension. As the dynamic viscosity is not much in-
fluenced at low values of the particle volume fraction, the
kinematic viscosity of the suspension will decrease compared to
the kinematic viscosity for the particle-free case. This will de-
crease the Kolmogorov length scale and hence elongate the iner-
tial subrange of the energy spectrum. There is a second effect that
is, in particular, important in the inertial subrange. There are two
competing effects in that subrange: an energy suppression due to
the fluid-particle friction and an energy enhancement during the
cascade process due to the decrease of the effective density of the
suspension with decreasing eddy size. Particles become less in-
volved in the eddy motion with decreasing eddy size. A more
detailed investigation of this effect has been made by L’vov et al.,
and it is shown that this effect can lead to a significant enhance-
ment of the turbulence in the inertial subrange dependent on con-
ditions, such as the ratio of the particle response time the integral
time scale. It is the combination of the two effects mentioned
above, that explains the phenomena observed in Fig. 6 in terms of
the model.
Druzhinin 23兴共see also Druzhinin and Elghobashi 24兴兲 stud-
ied the two-way coupling effect on the decay rate of isotropic
turbulence laden with solid spherical microparticles whose re-
sponse time is much smaller than the Kolmogorov time scale. An
asymptotic analytical solution was obtained for the kinetic energy
spectrum of the carrier fluid, which indicated that the two-way
coupling increases the fluid inertia in the fluid momentum equa-
tion by the factor 1+
.
is the particle mass fraction.The net
result is a reduction of the decay rate of turbulence energy as
compared to that of the particle-free turbulence. The analytical
solution was also extended up to the first order in
p/
kand is
applicable for particles with small but finite inertia.
4.2.3 Phenomenological Models. Yuan and Michaelides 25
presented a model for the turbulence modification in particle laden
flows based on the interaction between particles and turbulence
eddies. Two predominant mechanisms for the suppression and
production of turbulence were identified: ithe acceleration of
particles in eddies is the mechanism for the turbulence reduction
and iithe flow velocity disturbance due to the wake of the par-
ticles or the vortices shed by the particles is taken to be the re-
sponsible mechanism for turbulence enhancement. The effect of
the two mechanisms were combined to yield the overall turbu-
lence intensity modulation. The model exemplifies the effect of
several variables, such as particle size, relative velocity, Reynolds
number, ratio of densities, etc. A comparison to available experi-
mental data confirmed that the model predicts rather well the ob-
served changes in turbulence intensity. In the model no attention
is devoted to the influence of the particles on the turbulent energy
spectrum of the carrier fluid. Although the model shows fair
agreement with experimental data for the turbulence intensity, it
involves ad hoc modeling procedures leaving many questions re-
garding turbulence modulation unanswered. It has to be empha-
sized, however, that the publication by Yuan and Michaelides is
one of the few that pays attention to the generation of turbulence
in the wake behind the particles for large values of the particle
Reynolds number. Similar models that also considers the turbu-
lence generation have been published by Yarin and Hetsroni 26
and Crowe 27.
4.3 Discussion of Results From Theoretical Models. As a
general conclusion from the theoretical work, it can be stated that
the more recent analytical models for a homogeneous, isotropic,
turbulent suspension predict the same effect of the particles on the
turbulent energy spectrum of the carrier fluid as the effect pre-
dicted by direct numerical simulations. For a suspension with par-
ticles with a response time much larger than the Kolmogorov time
scale, the main effect of the particles is suppression of the energy
of eddies of all sizes at the same energy input into the suspension
as for the particle-free case. Thus for such a suspension, the total
turbulent energy will be smaller than the total turbulent energy for
the particle-free case at the same energy input. However, for a
suspension with particles with a response time comparable to or
smaller than the Kolmogorov time, the Kolmogorov length scale
will decrease and the turbulence energy of nearlyall sizes in-
creases. In that case the total turbulent energy can be larger than
the total turbulent energy for the particle-free case. For a suspen-
sion with particles with a response time in between the two lim-
iting cases mentioned above, the energy of the larger eddies is
suppressed, whereas the energy of the smaller ones is enhanced.
An interesting point is that it seems possible to give different
physical explanations for the influence of the particles on a de-
cayinghomogeneous, isotropic turbulent suspension. One expla-
nation given by Ferrante and Elghobashiis based on a micro-
scopic picture about the interaction between individual particles
and their local fluid flow environment. The other one given by
21兴兲 uses a macroscopic picture with eddy-size-dependent sus-
pension properties, such as effective density. Both pictures give a
satisfactory explanation, not only in words but also mathemati-
cally for details, see the relevant publications.
It is important to point out that only in a few analytical models
the turbulence generation due to the particle wakes or vortices
shed by the particles is taken into account. In this respect it is
interesting to mention briefly the work of Parthasarathy and Faeth
28. It will also be discussed in the section on experimental work.
They investigated theoretically and experimentally the continuous
phase properties for the case of nearly monodisperse glass par-
ticles falling in a stagnant water bath. This yielded a stationary,
homogeneous flow in which all turbulence properties were due to
the effects of turbulence modulation by the particles. Their theo-
retical model is based on a linear superposition of undistorted
particle wakes in a nonturbulent environment. The model predicts
many properties of the flow reasonably well. However, it yields
poor estimates of the integral length scale and streamwise spatial
correlations. This model seems a good starting point for further
research on the turbulence generation by particle wakes.
As a final remark we stress the point that more attention may
also be given to the theoretical and numericalinvestigation of
the influence of preferential concentration clusteringof particles
in the turbulent flow field on the two-way coupling effect. During
the last ten years several publications concerning theoretical and
numerical studies of particles clustering in a turbulent flow field
have appeared in the literature. Some of them are summarized
below. We first emphasize, however, that in these publications the
effect of turbulence on particle clustering is studied. The influence
of this clustering on the turbulence of the fluid velocity field the
two-way coupling effect that is the subject of this reviewis not
considered. Elperin et al. 29proposed a theory in which particle
clusters are caused by the combined influence of particle inertia
leading to a compressibility of the particle velocity fieldand a
finite velocity correlation time of the fluid flow field. Particles
inside turbulent eddies are carried to the boundary regions be-
tween them by inertial forces. This clustering mechanism acts on
all scales of turbulence and increases toward small scales. The
turbulent diffusion of particles decreases toward smaller scales.
Therefore, the clustering instability dominates at the Kolmogorov
84/Vol. 59, MARCH 2006 Transactions of the ASME
scale. An exponential growth of the number of particles in the
clusters is inhibited by collisions between the particles. The end
result can be a strong clustering whereby a finite fraction of par-
ticles is accumulated in the clusters, or a weak clustering when a
finite fraction of collisions occurs in the clusters. A crucial param-
eter for clustering is the particle radius, which has to be larger
than a certain critical value. Also Balkovsky et al. 30considered
clustering of inertial particles suspended in a turbulent flow and
developed a statistical theory of this phenomenon based on a La-
grangian description of turbulence. The initial growth of concen-
tration fluctuations from a uniform state is studied, as is its satu-
ration due to finite-size effects, imposed either by the Brownian
motion or by a finite distance between the particles. The statistics
of these fluctuations is independent of the details of the velocity
statistics, which allows the authors to predict that the particles
cluster at the Kolmogorov scale of turbulence. Also the probabil-
ity distribution of the concentration fluctuations is calculated.
They discuss the possible role of the particle clustering in the
physics of atmospheric aerosols, in particular, cloud formation.
5 Experimental Work
5.1 Introduction. In this section, an overview is given of the
experimental studies of dispersed turbulent flows found in the
literature. Experimental work on the interaction of particles and
turbulence in homogeneous, isotropic flows is mostly done in
grid-generated turbulence. This is the closest approximation to
true homogeneous, isotropic turbulence, while still being experi-
mentally feasible. Most experiments use a static grid through
which a fluid moves with a constant mean velocity. Alternatively,
an oscillating grid in a enclosed tank can be used 31–33. The
advantage of these is that the turbulence level that can be attained
is relatively high. Additionally, the flow does not have a mean
flow component. This can be beneficial if Lagrangian measure-
ments are required, such as tracking particles over a longer time to
study, e.g., dispersion properties. The biggest drawback of these
experiments is the fact that there often is a strong gradient in the
particle concentration. To overcome this drawback, experiments
are being done in microgravity 34. The focus of this review will
be on the classic approach: by placing a grid in a wind tunnel or
water channel.
5.2 Grid-Generated Turbulence. The main concept of grid-
generated turbulence is simple: a fluid passes a grid with a certain
solidity. Strong gradients in the axial direction are generated i.e.,
“jets” emerging from the openings in the grid, which break up to
form nearly isotropic turbulence. The macroscopic length scale of
the turbulence is determined by the mesh spacing M. This length
scale, as any other length scale of the flow, grows proportional to
the square root of the downstream position: Lx1/2. Usually, it is
found that at a downstream distance of 20 mesh spacings the flow
is reasonably isotropic 10% difference between axial and trans-
versal components. Because of the mixing behavior of a turbulent
flow, no influence of the separate mesh bars can be observed from
this distance on; the flow is homogeneous in the plane perpendicu-
lar to the mean flow direction. Typically, the turbulence level,
defined as the ratio of the root-mean-square of the fluid fluctua-
tions uto the mean flow velocity U, is of the order of a few
percent. It is mainly determined by the mean fluid velocity and the
solidity of the grid. The turbulence obviously decays, therefore,
strictly speaking, the flow is not homogeneous in the axial direc-
tion. With respect to a typical particle time scale e.g., the Stokes
time, however, this decay is slow. Therefore, the flow can be
considered to be homogeneous in the context of particle-
turbulence interaction. The decay of the turbulent kinetic energy is
usually assumed to decay proportional to the inverse of the dis-
tance u2x−1. The conventional way of representing the results
is therefore by plotting the reciprocal value of the turbulent kinetic
energy versus the distance to the grid. For an extensive treatment
of grid-generated flow, one is referred to the classic experiments
by, e.g., Comte-Bellot and Corrsin 35or Batchelor 36.
5.3 Overview of Experiments. The first thing that becomes
clear when reviewing the available literature is the fact that each
of the experiments has a relatively narrow scope. Even though
most papers claim to study the two-way interactions between the
phases, in general, the authors are often limited to the study of
only one or two related phenomena. Two-phase flows are notori-
ously difficult to study experimentally 37, so each of the experi-
ments has to be more or less tailored to study a certain aspect of
the interaction. A broad classification of the experiments can be
made:
changes in carrier-phase properties
particle-induced turbulence
clustering, preferential concentrations
particle-phase properties
The focus of the numerical and theoretical sections was mainly
on the first class: the influence of particles on the carrier-phase
spectrum. Only a few experiments have reported this because of
the difficulties in obtaining good measurements in such flows.
Additionally, the decay of the total turbulent kinetic energy
i.e., the integral of the spectrumis discussed. A full discussion of
all related phenomena is clearly unfeasible. For example, the work
on the effective settling velocity of particles in a turbulent flow is
an important parameter in many engineering models and the topic
of numerous publications. The settling velocity of a particle is a
direct result of particle-fluid interaction. But since settling obvi-
ously indicates a strong contribution from gravitational forces, it
is therefore inherently anisotropic. Since most theories and nu-
merical works exclude gravity, this topic is not considered here.
5.4 Influence of the Particles on the Carrier Phase
Spectrum. Only three experiments have reported measurements
of the influence of particles on the carrier phase turbulence in
homogeneous isotropic turbulence: Schreck and Kleis 38, Hus-
sainov et al. 39and Geiss et al. 40. The first one used solid
particles in water, yielding a density ratio of the order of unity.
The latter two used solid particles in a vertical wind tunnel den-
sity ratio of order 103. Work in progress on the topic has recently
been reported by Nishino et al. 41and Poelma et al. 42, which
will hopefully contribute to the data available in the near future.
Where needed, their preliminary results are discussed.
Schreck and Kleis 38used two types of particles: glass rela-
tive density 2.5and neutrally buoyant plastic particles. Decaying
grid turbulence offers the chance to study the dynamics of turbu-
lent suspensions. In Fig. 7, the reciprocal value of theturbulent
kinetic energy is plotted for the single-phase and particle-laden
case. These data were obtained using laser Doppler anemometry
LDA. As can be seen, the turbulence level is lower for all mass
loadings compared to the single-phase flow. This was also the case
for the neutrally buoyant particles. The slopes in Fig. 7 corre-
sponding to the particle-laden cases are somewhat steeper than the
single-phase case, indicating an increased dissipation rate.
In the longitudinal spectrum a very small decrease in the energy
at large scales could be observed, as well as an increase in energy
at small scales Fig. 8. This was most evident for the neutrally
buoyant particles. The Stokes number of these particles was 1.9. It
should be noted that the Stokes number decreases as the turbu-
lence decays because of the growth of the Kolmogorov scales. For
the transversal spectra the reverse effect was seen: the small scales
have less energy for the particle-laden flows. The overall effect of
these phenomena resulted in almost identical shapes of the one-
dimensional spectrum for the particle-laden and particle-free
flows. Therefore, these results seem in contradiction to the out-
come of the numerical work, which predicts changes in the one-
dimensional spectrum at this Stokes number and mass loading
Applied Mechanics Reviews MARCH 2006, Vol. 59 /85
=3.8%. Obviously, there is a significant difference in volume
load e.g., Ferrante and Elghobashi:
=0.1%, Schreck and Kleiss:
=1.5%, yet one would expect more of an effect with higher
loads.
Hussainov et al. 39used particles that were similar to those
used by Schreck and Kleis 38, yet instead of a water channel
they used a wind tunnel. This led to a mass fraction that was
significantly larger
=10%, but more importantly also provided
very large Stokes numbers i.e., of the order 103. Measurements
are again done using LDA. Surprisingly, the effects that Hus-
sainov and co-workers measure are less than those obtained by
Schreck and Kleiss despite the higher mass load: the decay rate is
somewhat larger, but very similar to the single-phase decay rate
see Fig. 9. In the far-downstream region they find that the equi-
librium turbulence level i.e., fully developed pipe/channel flow
is lower than for the particle-free case.
In the reported transversal spectra see Fig. 10an increase of
energy is observed at higher frequencies viz. above 1 kHz. Even
though this agrees with Schreck and Kleis, it can be debated that
this is well within their measurement uncertainty and there is no
difference between the single-phase and particle-laden fluid spec-
tra. Similar results, yet less pronounced, were found in experi-
ments with a grid with mesh spacing twice as large. A possible
explanation for the absence of a clear cross-over in the spectrum
might be the large Stokes number of the particles. Since they are
unresponsive to most fluid fluctuations, true two-way coupling
effects cannot be expected; there is only influence of the particle
on the fluid and not vice versa.
In a recent publication, Geiss et al. 40reported very similar
experiments as Hussainov’s. The main difference is the used par-
ticle size 120, 240, and 480
m, which is significantly smaller
than the experiments mentioned earlier. Still, the Stokes times of
the glass particles are very much larger than unity the Kolmog-
orov scales are not reported, only Stokes numbers based on the
integral time scale are reported. Obviously these are smaller than
the Kolmogorov scale-based values. For their measurements,
they used a phase Doppler anemometry. This enabled them to
measure both the fluid and the particle velocity simultaneously. At
a mass fraction of up to 0.077, they do not find any influence on
the normalizedcarrier phase spectrum within the accuracy of the
measurements. On the other hand, there was an influence on the
total kinetic energy of the carrier phase and also on the decay rate.
One of their findings was the fact that there seems to be a thresh-
Fig. 7 Influence of mass load on decay of turbulent kinetic
energy; 0.65 mm glass particles in water. Reproduced from Ref.
38.
Fig. 8 Longitudinal spectrum for unladen and laden flow;
0.65 mm neutrally buoyant particles in water. Reproduced from
Ref. 38.
Fig. 9 Decay of turbulent kinetic energy for particle-free and
particle-laden flow. Data taken from Ref. 39.
Fig. 10 Smoothened energy spectra for single-phase
=0
and particle-laden
=0.005flows. Data taken from Ref. 39.
86/Vol. 59, MARCH 2006 Transactions of the ASME
old in the mass load above which turbulence dampening occurs.
However, this effect seems rather weak considering their results
given in Figs. 11 and 12, which show the decay of the carrier
phase turbulent kinetic energy for various mass loads. A remark-
able result of their measurement is the fact that the flow becomes
anisotropic while it decays. This was not observed by e.g.,
Schreck and Kleis, but has been confirmed by recent work of
Poelma et al. 42and Poelma 43. An example of their decaying
axial and transversal kinetic energy components is shown in Fig.
13. Poelma et al. added ceramic particles
=0.1%, dp= 280
m,
p/
f=3.8, Rep=18to their decaying grid-generated turbulence
in a water channel. It is clear that in the initial stages after the
grid, there is less turbulence compared to the single-phase flow at
the same centerline velocity. The axial component appears to de-
cay slower than the transversal component, a result very similar to
that reported in Figs. 11 and 12. Note that in the work of Poelma,
the mass fraction is significantly lower because the experiments
were done in a water channel. The reason for the slower decay is
probably particle-induced turbulence; the overal decay rate ap-
pears lower because the particles generate turbulence, predomi-
nantly in the axial gravitationaldirection. This behavior is in
contradiction to the often-cited rule of thumb that particles start
producing turbulence when Rep400.
The detailed mechanisms of the generation of turbulence by
particles has been studied by an number of authors. Parthasarathy
and Faeth 28performed studies using LDA of turbulence gen-
eration due to falling glass particles in stagnant water. An impor-
tant conclusion was the fact that the turbulence generated by a
falling particle is anisotropic; the mean streamwise i.e., in the
direction of gravitycomponent is roughly a factor 2 higher than
the cross-stream component. Furthermore, it was found that the
spectrum of the fluid phase showed energy at a large range of
frequencies, even though the particle Reynolds numbers were
small. This indicates that energy is generated at or transported to
scales larger than the particle size. Chen et al. 44and Lee et al.
45did similar experiments in a counterflow wind tunnel. They
found that even at low volume loads 0.003%, the particles can
generate a turbulence level of up to 5%.
In a recent study in grid-generated turbulence, Nishino et al.
41also found strong anisotropic particle-induced turbulence.
Again, grid-generated turbulence in a water channel was studied.
The particles they used were 1.0 and 1.25 mm glass particles,
their size being in between the Kolmogorov and integral length
scales Rep=138, 208. Instead of having a constant mass load
they studied a decaying mass load using a innovative idea: the
upwardmean flow was chosen equal to the mean settling veloc-
ity of the particles. The particles were trapped in the test section
with only a small hole at the top. This led to a slowly decreasing
mass/volume load, which was measured using image processing
techniques. The turbulence statistics of the fluid, as measured by a
particle image velocimetry PIVsystem, could thus be measured
as a function of the volume load. Figure 14 shows the influence of
the mass load on the turbulence level. The most important feature
Fig. 11 Decay of the axial streamwiseand transversal kinetic energy for single-phase and
particle laden case: left:
=0.37%, right:
=2.67%. Data replotted from Ref. 40.
Fig. 12 Decay of the axial streamwiseand transversal kinetic
energy for single-phase and particle laden case.
=4.99%.
Data replotted from Ref. 40.
Fig. 13 Decay of axial and transversal kinetic energy for
single-phase and particle-laden case
=0.1%,
=0.38%, dp
=280
m,
p/
f=3.8. Data taken from Ref. 42.
Applied Mechanics Reviews MARCH 2006, Vol. 59 /87
of this graph is the fact that while the axial component increases
significantly and more or less linearlywith volume load, the
transversal component increases only slowly. In total, the turbu-
lence level is roughly three times larger than for the particle-free
flow. These results are in contradiction with the experiments of
Schreck and Kleiss 38, who found a slight damping with some-
what smaller particles of the same density. The explanation given
by Nishino et al. 41is the fact that the potential energy of the
particles is transferred to kinetic energy. This is the same effect as
was studied by Parthasarathy and Faeth 28, apart from the pres-
ence of ambient turbulence in Nishino’s work.
In addition to the particle-induced anisotropy, they observed
large fluctuations in the concentration of particles. More specific,
they observed what they called “columnar particle accumulation,”
i.e., vertical bands of high concentration. This could also be ob-
served in the transversal autocorrelation function, which changed
drastically. Similar phenomena have also been observed by
Poelma et al. 42, also using PIV. These large density fluctuations
may also be the cause of the increased turbulence level. Obvi-
ously, these effects could only be identified due to the whole-field
character of the PIV measurements, in contrast to the earlier
single-point LDA work.
The inhomogeneities in particle distribution observed by
Nishino et al. 41and others have been investigated by a number
of authors, and the effect is often referred to as “clustering” or
“preferential concentration.” Eaton and Fessler 46collected a
significant amount of experiments and simulations done in this
field. The main conclusion that can be drawn from this overview
is that these effects occur when the Stokes number of the particles
is close to unity. Work by Fallon and Rogers 34, based on im-
aging of particles in microgravity, indicated that the presence of
gravity or any other body forcereduces preferential concentra-
tion phenomena. Preferential concentration can have significant
effects. For instance, Aliseda et al. 47studied the effect of pref-
erential concentration on the settling velocity of particles. The
system was studied using imaging-based techniques. They ob-
served a quasi-linear relationship between the effective settling
velocity and the local concentration. The settling velocity of par-
ticles is an important parameter in, e.g., many engineering models,
but a discussion of this parameter is beyond the scope of this
review.
5.5 Discussion of Experimental Work. No significant
changes in the shape of the carrier-phase turbulence spectrum
normalized by the total turbulent kinetic energyare observed in
the experiments. However, the total turbulent kinetic energy of the
fluid phase is lower for most experiments, which indicates that
there is some way of coupling between the phases. The fact that
the Stokes number of the particles used in most experiments is an
order of magnitude larger than unity, might explain the absence of
bigger changes in the spectrum. It seems to us that a damping
effect due to the particles with large Stokes numbertakes place
at the largest scales of turbulence, and that because of the cascade
process of turbulence all other smaller eddies also receive less
energy than in the particle-free case. If energy would be added or
taken at any other position in the spectrum, this would become
evident in the shape of the spectrum. On the other hand, particles
are able to generate turbulence. The particle-induced turbulence is
anisotropic, and energy is generated at a large range of scales in
the spectrum. The two effects, idamping of the fluid motions at
large scale and cascade transport to smaller scales and ii
particle-induced turbulence production, are competing processes
in dispersed two-phase flows. A more quantitative description us-
ing the presently available experimental data is impossible. The
governing parameters such as density ratio, ratio of particle size
and fluid length scale, and volume load, are too different to be
able to compare the present results directly. More research is
therefore needed, preferably varying one single parameter as was
done in, e.g., Nishino’s work with the volume fraction.
6 Conclusion
The numerical work from different researchers for the two-way
coupling effect in a homogeneous, isotropic, turbulently flowing
suspension agrees reasonably well with respect to the effect of
particles on the turbulence spectrum of the carrier phase: low
wave numbers are suppressed, while energy is gained at higher
wave numbers dependent on the Stokes number. The cross-over
point—the wave number above which the energy is larger com-
pared to the single-phase case—shifts to larger wave numbers for
larger Stokes numbers. The overall effect can be either a damping
of the turbulence level or an increase, depending on the particle
Stokes number and the volume load. Several physical explana-
tions for this phenomenon have been given in the literature.
Analytical theories for the two-way coupling have been devel-
oped based on the physical mechanisms that are thought to govern
the system. For instance, the pivoting of the carrier-phase spec-
trum is reasonably well predicted, even though the results are still
somewhat qualitative. At the moment, no comprehensive theory
that integrates all phenomena exists, however. In particular, pref-
erential concentration effects are not included. This effect can play
an important role.
The available experimental data for a homogeneous, isotropic
turbulent suspension are scarce. The pivoting of the spectrum was
not observed in the obtained spectra, neither in solid/liquid nor
solid/gas flows, even though the mass loads were comparable to
those used in the numerical simulations. The fact that the Stokes
number of the used particles was rather high in most experiments
may partly account for the discrepancy. According to the physical
explanations mentioned above, a large Stokes number implies that
the cross-over pivotingpoint would move to very high wave
numbers; thus, effectively damping on all scales occurs. This is in
agreement with what is observed in most experiments.
Another important conclusion from the experiments is the ef-
fect of gravity, which generates a strongly anisotropic system.
This effect is not included in most numerical simulations and
analytical theories; therefore, direct comparison is not trivial. It is
important to extend the simulations and theories so that the effect
of gravity is incorporated.
Acknowledgment
Parts of this work was supported by the Technology Foundation
STW, Applied Science Division of NWO, and the technology pro-
gramme of the Ministry of Economic Affairs. The authors would
like to thank Dr. Dreizler and Dipl.-Ing. Chrigui for supplying the
original data used in Figs. 11 and 12.
Fig. 14 Turbulence level modification by 1.0 mm glass par-
ticles in water as function of volume concentration, Rep=140.
Data taken from Ref. 41.
88/Vol. 59, MARCH 2006 Transactions of the ASME
References
1Hetsroni, G., 1989, “Particles-Turbulence Interaction,” Int. J. Multiphase
Flow, 155, pp. 735–746.
2Elghobashi, S., 1994, “On Predicting Particle-Laden Turbulent Flows,” Appl.
Sci. Res., 52, pp. 309–329.
3Crowe, C., Troutt, T., and Chung, J., 1996, “Numerical Models for Two-Phase
Flows,” Annu. Rev. Fluid Dyn., 28, pp. 11–43.
4Mashayek, F., and Pandya, R., 2003, “Analytical Description of Particle/
Droplet-Laden Turbulent Flows,” Prog. Energy Combust. Sci., 294, pp. 329–
378.
5Squires, K., and Eaton, J., 1990, “Particle Response and Turbulence Modifi-
cation in Isotropic Turbulence,” Phys. Fluids A, 27, pp. 1191–1203.
6Squires, K., and Eaton, J., 1994, “Effect of Selective Modification of Turbu-
lence on Two-Equation Models for Particle-Laden Turbulent Flows,” ASME J.
Fluids Eng., 116, pp. 778–784.
7Elghobashi, S., and Truesdell, G., 1993, “On the Two-Way Interaction Be-
tween Homogeneous Turbulence and Dispersed Solid Particles, I: Turbulence
Modification,” Phys. Fluids A, 57, pp. 1790–1801.
8Maxey, M., and Riley, J., 1983, “Equation of Motion for a Small Rigid Sphere
in Nonuniform Flow,” Phys. Fluids, 264, pp. 883–889.
9Boivin, M., Simonin, O., and Squires, K., 1998, “Direct Numerical Simulation
of Turbulence Modulation by Particles in Isotropic Turbulence,” J. Fluid
Mech., 375, pp. 235–263.
10Sundaram, S., and Collins, L., 1999, “A Numerical Study of the Modulation of
Isotropic Turbulence by Suspended Particles,” J. Fluid Mech., 379, pp. 105–
143.
11Ferrante, A., and Elghobashi, S., 2003, “On the Physical Mechanisms of Two-
Way Coupling in Particle-Laden Isotropic Turbulence,” Phys. Fluids, 152,
pp. 315–329.
12ten Cate, A., Derksen, J., Kramer, H., Rosmalen, G., and Van den Akker, H.,
2001, “The Microscopic Modeling of Hydrodynamics in Industrial Crystallis-
ers,” Chem. Eng. Sci., 56, pp. 2495–2509.
13ten Cate, A., 2002, “Turbulence and Particle Dynamics in Dense Crystal Slur-
ries,” Ph.D. thesis, Delft University Press, The Netherlands.
14Mittal, R., 2000, “Response of the Sphere Wake to Freestream Fluctuations,”
Theor. Comput. Fluid Dyn., 13, pp. 397–419.
15Bagchi, P., and Balachandar, S., 2003, “Inertial and Viscous Forces on a Rigid
Sphere in Straining Flows at Moderate Reynolds Numbers,” J. Fluid Mech.,
481, pp. 105–148.
16Zhang, D., and Prosperetti, A., 2003, “Momentum and Energy Equations for
Disperse Two-Phase Flows and Their Closure for Dilute Suspensions,” Int. J.
Multiphase Flow, 154, pp. 868–880.
17Baw, P., and Peskin, R., 1971, “Some Aspects of Gas-Solid Suspension Tur-
bulence,” ASME J. Basic Eng., 93, pp. 631–635.
18Felderhof, B., and Ooms, G., 1990, “Effect of Inertia, Friction and Hydrody-
namic Interactions on Turbulent Diffusion,” Eur. J. Mech. B/Fluids, 9, pp.
349–368.
19Ooms, G., and Jansen, G., 2000, “Particles-Turbulence Interaction in Station-
ary, Homogeneous, Isotropic Turbulence,” Int. J. Multiphase Flow, 26, pp.
1831–1850.
20Ooms, G., Poelma, C., and Westerweel, J., 2002, Advances in Turbulence IX,
proc. of ETC-9 Southampton (UK), CIMNE, Barcelona, pp. 369–372.
21L’vov, V., Ooms, G., and Pomyalov, A., 2003, “Effect of Particle Inertia on
Turbulence in a Suspension,” Phys. Rev. E, 674, pp. 1–21.
22Ooms, G., and Poelma, C., 2004, “Comparison Between Theoretical Predic-
tions and DNS-Results for a Decaying Turbulent Suspension,” Phys. Rev. E,
695, p. 056311.
23Druzhinin, O., 2001, “The Influence of Particle Inertia on the Two-Way Cou-
pling and Modification of Isotropic Turbulence by Microparticles,” Phys. Flu-
ids, 1312, pp. 3378–3755.
24Druzhinin, O., and Elghobashi, S., 1999, “On the Decay Rate of Isotropic
Turbulence Laden With Microparticles,” Phys. Fluids, 11, pp. 602–610.
25Yuan, Z., and Michaelides, E., 1992, “Turbulence Modulation in Particulate
Flows: A Theoretical Approach,” Int. J. Multiphase Flow, 18, pp. 779–785.
26Yarin, L., and Hetsroni, G., 1994, “Turbulence Intensity in Dilute Two-Phase
Flow. 3: The Particles-Turbulence Interaction in Dilute Two-Phase Flow,” Int.
J. Multiphase Flow, 20, pp. 27–44.
27Crowe, C., 2000, “On Models for Turbulence Modulation in Fluid-Particle
Flows,” Int. J. Multiphase Flow, 26, pp. 719–727.
28Parthasarathy, R., and Faeth, G., 1990, “Turbulence Modulation in Homoge-
neous Dilute Particle-Laden Flow,” J. Fluid Mech., 220, pp. 485–514.
29Elperin, T., Kleeorin, N., L’vov, V., Rogachevskii, I., and Sokoloff, D., 2002,
“Clustering Instability of the Spatial Distribution of Intertial Particles in Tur-
bulent Flows,” Phys. Rev. E, 663, pp. 036302.
30Balkovsky, E., Falkovich, G., and Fouxon, A., 2001, “Intermittent Distribution
of Inertial Particles in Turbulent Flows,” Phys. Rev. Lett., 8613, pp. 2790–
2793.
31Bennet, S., and Best, J., 1995, “Particle Size and Velocity Discrimination in a
Sediment-Laden Turbulent Flow Using Phase Doppler Anemometry,” ASME
J. Fluids Eng., 117, pp. 505–510.
32Huppert, H., Turner, J., and Hallworth, M., 1995, “Sedimentation and Entrain-
ment in Dense Layers of Suspended Particles Stirred by an Oscillating Grid,”
J. Fluid Mech., 289, pp. 269–293.
33Orlins, J., and Gulliver, J., 2003, “Turbulence Quantification and Sediment
Resuspension in an Oscillating Grid Chamber,” Exp. Fluids, 346, pp. 662–
677.
34Fallon, T., and Rogers, C., 2002, “Turbulence-Induced Preferential Concentra-
tion of Solid Particles in Microgravity Conditions,” Exp. Fluids, 332, pp.
233–241.
35Comte-Bellot, G., and Corrsin, S., 1971, “Simple Eulerian Time Correlations
of Full-and Narrow-Band Velocity Signals in Grid-Generated Isotropic Turbu-
lence,” J. Fluid Mech., 48, pp. 273–337.
36Batchelor, G., 1953, The Theory of Homogeneous Turbulence, Cambridge Uni-
versity Press, Cambridge, England.
37Boyer, C., Duquenne, A.-M., and Wild, G., 2002, “Measuring Techniques in
Gas-Liquid and Gas-Liquid-Solid Reactors,” Chem. Eng. Sci., 57, pp. 3185–
3215.
38Schreck, S., and Kleis, S., 1993, “Modification of Grid-Generated Turbulence
by Solid Particles,” J. Fluid Mech., 249, pp. 665–688.
39Hussainov, M., Karthushinsky, A., Rudi, Ü., Shcheglov, I., Kohnen, G., and
Sommerfeld, M., 2000, “Experimental Investigation of Turbulence Modulation
by Solid Particles in a Grid-Generated Vertical Flow,” Int. J. Heat Fluid Flow,
21, pp. 365–373.
40Geiss, S., Dreizler, A., Stojanovic, Z., Chrigui, M., Sadiki, A., and Janicka, J.,
2004, “Investigation of Turbulence Modification in a Non-Reactive Two-Phase
Flow,” Exp. Fluids, 36, pp. 344–354.
41Nishino, K., Matsushita, H., and Torii, K., 2003, “Piv Measurements of Tur-
bulence Modification by Solid Particles in Upward Grid Turbulence of Water,”
Proc. of 5th Int. Symp. on PIV, Busan (Korea),p.3118.
42Poelma, C., Westerweel, J., and Ooms, G., 2003, “Turbulence Modification by
Particles, Experiments Using PIV/PTV,” European Fluid Mechanics Confer-
ence, Toulouse (France).
43Poelma, C., 2004, “Experiments in Particle-Laden Turbulence,” Ph.D. thesis,
Delft University of Technology, The Netherlands.
44Chen, J., Wu, J., and Faeth, G., 2000, “Turbulence Generation in a Homoge-
neous Particle-Laden Flow,” AIAA J., 384, pp. 636– 642.
45Lee, K., Faeth, G., and Chen, J.-H., 2003, “Properties of Particle-Generated
Turbulence in the Final-Decay Period,” AIAA J., 417, pp. 1332–1340.
46Eaton, J., and Fessler, J., 1994, “Preferential Concentration of Particles by
Turbulence,” Int. J. Multiphase Flow, 201, pp. 169–209.
47Aliseda, A., Cartellier, A., Hainaux, F., and Lasheras, J., 2002, “Effect of
Preferential Concentration on the Settling Velocity of Heavy Particles in Ho-
mogeneous Isotropic Turbulence,” J. Fluid Mech., 468, pp. 77–105.
Christian Poelma received his M.Sc. degree in Chemical Engineering in 1999 from the Delft University of
Technology (TUD), The Netherlands. At the TUD, he started his Ph.D. research in the Laboratory for Aero
and Hydrodynamics. His main research topics are experimental fluid mechanics, turbulence, and dispersed
two-phase flows. In 2004, he obtained his Ph.D. on experimental work on turbulence modification by
particles. This review paper was prepared as part of his thesis. In the summer of 2004, he started as a
postdoc (funded by a NWO Talent Fellowship) at the California Institute of Technology in Pasadena to study
the unsteady aerodynamics of insect flight.
Applied Mechanics Reviews MARCH 2006, Vol. 59 /89
Gijs Ooms studied applied physics at the Delft University of Technology (TUD). He received his doctoral
degree in 1971 on a topic from the field of fluid mechanics. Thereafter he worked for Shell (in Amsterdam,
Houston, and Rijswijk) and the TUD. At Shell he was involved in research and technology development.
Although he gradually developed into a manager of research, he always continued carrying out his own
research projects. At the TUD he is professor of fluid mechanics at the Laboratory for Aero- and Hydro-
dynamics. He is also scientific director of the J.M. Burgers Centre (the national research school for fluid
mechanics). His scientific interest is on two-phase flow, turbulence, and the influence of high-frequency
acoustic waves on the flow through a porous material.
90/Vol. 59, MARCH 2006 Transactions of the ASME
... Later, Geiss et al. (2004) observed that the continuous phase turbulence intensity only changed with a minimum particle mass loading (0.058-0.077) which depends on the particle diameter distribution. However, Poelma and Ooms (2006) analysed the experimental data of Geiss et al. (2004), and they found the augmentation of turbulence intensity for various mass loading ratios (∼0.37 %, 2.67 %, and 4.99 %), which seems to contradict with the result previously reported by Geiss et al. (2004). Furthermore, Geiss et al. (2004) reported that the original isotropic turbulence produced by the grid inside the wind tunnel was modulated to anisotropic turbulence in the presence of different sizes of particles. ...
... This observation agrees well with the experimental spectra reported by Schreck and Kleis (1993); however, a clear cross-over was apparent in the latter study. The presence of observed cross-over is attributed to the large particle Stokes number based on the hypothesis that large Stokes number particles are not affected by the fluid flow hence a two-way coupling is not applicable (Poelma and Ooms 2006). Geiss et al. (2004) did not observe any significant changes in the energy spectrum slope; however, Poelma et al. (2007) reported a less steep slope in the inertial region of the spectrum compared to Kolmogorov's −5/3 slope due to the presence of particles. ...
Article
Full-text available
In multiphase particulate systems, the turbulence of the continuous phase (gas or liquid) is modulated due to interactions between the continuous phase and the suspended particles. Such phenomena are non-trivial in the essence that addition of a dispersed phase to a turbulent flow complicates the existing flow patterns depending on the physical properties of the particles leading to either augmentation or attenuation of continuous phase turbulence. In the present study, this aspect has been comprehensively analysed based on the available experimental data obtained from the well-studied turbulent flow systems such as channel and pipes, free jets and grids. Relevant non-dimensional parameters such as particle diameter to integral length scale ratio, Stokes number, particle volume fraction, particle momentum number, and particle Reynolds number have been utilised to characterise the reported turbulence modulation behavior. Some limitations of these commonly used dimensionless parameters to characterise turbulence modulation are discussed, and possible improvements are suggested.
... An increase in the energy of small scales happens due to modification in the energy transfer rate that reduces at large scales and increases at small scales for the particle-laden flows, which is also observed in different flow configurations. [9][10][11][12][13][14][15][16][17][18] Sundaram and Collins 11 performed DNS for homogeneous isotropic turbulence and found the magnitude of particle-induced dissipation and viscous dissipation scale with particle number density at low volume fraction. The authors also observed a decrease in energy at low wavenumbers and an increase at high wavenumbers for fluid energy. ...
Article
The particle phase attenuates the fluid fluctuations with an increase in volume fraction, and a sudden collapse in the turbulence is observed at a particular particle volume fraction, called critical particle volume loading (CPVL) [P. Muramulla et al. J. Fluid Mech. 889, A28 (2020)]. The present study reports the capability of two different classes of large eddy simulation (LES), viz. anisotropic and eddy viscosity-based, models to capture the turbulence modulation and the sudden disruption of the fluid fluctuations in the particle-laden vertical channel flows. The simulations are performed at two bulk Reynolds numbers of 3300 and 5600 based on the channel width and the bulk averaged fluid velocity. Our study on different LES models shows that approximate deconvolution (ADM) and scale similarity (SS) models accurately predict the critical loading for the Reynolds number of 3300. However, these models predict the critical loading qualitatively only for the Reynolds number of 5600 in the sense that they fail to predict the discontinuity as shown by the direct numerical simulation (DNS) study. The coherent structure model (CSM) predicts the critical loading with an 80% accuracy at both Reynolds numbers. The energy spectral density, production, and particle-induced dissipation spectra are plotted to analyze the distribution across wavenumbers. For all the LES models, a decrease in more than one order of magnitude is observed in the energy spectrum density at the critical loading compared to the unladen flow. The energy density decreases more in the channel center than in the near-wall region for the same particle volume loading. The mean component of particle-induced dissipation is almost two orders of magnitude larger than the particle dissipation spectra of fluctuating energy. The magnitude of streamwise and spanwise dissipation spectra of fluctuating components is higher in the near-wall region than the channel center. However, the magnitude of wall-normal dissipation spectra is higher in the channel center than near the wall region.
... In particle-laden turbulent flows, the dispersed and continuous phases can have a significant effect on each other's dynamics. This situation, usually termed two-way coupling, is common and has been the focus of much of the recent research in the field (Poelma & Ooms 2006;Balachandar & Eaton 2010;Kuerten 2016). Indeed, the portion of the parameter space where the dynamics can be considered one-way coupled is limited: the particle size d p will be smaller than the Kolmogorov scale η, and the solid volume fraction φ V and mass fraction φ m will be below threshold levels for highly dilute suspensions (Brandt & Coletti 2022). ...
Article
Full-text available
When inertial particles are dispersed in a turbulent flow at sufficiently high concentrations, the continuous and dispersed phases are two-way coupled. Here, we show via laboratory measurements how, as the suspended particles modify the turbulence, their behaviour is also profoundly changed. In particular, we investigate the spatial distribution and motion of sub-Kolmogorov particles falling in homogeneous air turbulence. We focus on the regime considered in Hassaini & Coletti ( J. Fluid Mech. , vol. 949, 2022, A30), where the turbulent kinetic energy and dissipation rate were found to increase as the particle volume fraction increases from $10^{-6}$ to $5\times 10^{-5}$ . This leads to strong intensification of the clustering, encompassing a larger fraction of the particles and over a wider range of scales. The settling rate is approximately doubled over the considered range of concentrations, with particles in large clusters falling even faster. The settling enhancement is due in comparable measure to the predominantly downward fluid velocity at the particle location (attributed to the collective drag effect) and to the larger slip velocity between the particles and the fluid. With increasing loading, the particles become less able to respond to the fluid fluctuations, and the random uncorrelated component of their motion grows. Taken together, the results indicate that the concentrated particles possess an effectively higher Stokes number, which is a consequence of the amplified dissipation induced by two-way coupling. The larger relative velocities and accelerations due to the increased fall speed may have far-reaching consequences for the inter-particle collision probability.
... All numerical experiments are conducted using a single flow realization (to ensure that the initial condition for the continuous phase is identical; see Section 2.2). However, it is well-known that the presence (as well as the deformation/breakup and coalescence) of dispersed drops, modulate the turbulence [45][46][47][48][49][50][51][52][53][54][55][56][57]. Thus, if the drops have different turbulence modulating effects, they do not necessarily experience the same turbulent stress, despite starting from the same flow realization. ...
Article
Full-text available
Turbulent emulsification is an important unit operation in chemical engineering. Due to its high energy cost, there is substantial interest in increasing the fundamental understanding of drop breakup in these devices, e.g., for optimization. In this study, numerical breakup experiments are used to study turbulent fragmentation of viscous drops, under conditions similar to emulsification devices such as high-pressure homogenizers and rotor-stator mixers. The drop diameter was kept larger than the Kolmogorov length scale (i.e., turbulent inertial breakup). When varying the Weber number (We) and the disperse-to-continuous phase viscosity ratio in a range applicable to emulsification, three distinct breakup morphologies are identified: sheet breakup (large We and/or low viscosity ratio), thread breakup (intermediary We and viscosity ratio >5), and bulb breakup (low We). The number and size of resulting fragments differ between these three morphologies. Moreover, results also confirm previous findings showing drops with different We differing in how they attenuate the surrounding turbulent flow. This can create ‘exclaves’ in the phase space, i.e., narrow We-intervals, where drops with lower We break and drops with higher We do not (due to the latter attenuating the surrounding turbulence stresses more).
... In addition, for St = 2.5, the cascades reveal the damping of energy and enstrophy at the large scale which is consistent with the physical discussion of Ψ β . This enhancement/reduction of turbulence at the small/large scales is also referred to as pivoting effect in the literature 21,37 . It has to be further emphasized that the critical wavenumbers determined in the present study, where the multiphase energy spectra deviate from their single-phase counterparts (figures 12 c,d), agree well with the point-particle Euler-Lagrange simulations of Letournel et al. 22 . ...
Article
Full-text available
Understanding the evolution of turbulence in multiphase flows remains a challenge due to the complex inter-phase interactions at different scales. This paper attempts to enlighten the multiphase turbulence phenomenon from a new perspective by exploiting the classical concept of vorticity and its role in the evolution of the turbulent energy cascade. We start with the vorticity transport equations for two different multiphase flow formulations, which are one-fluid and two-fluid models. By extending the decaying homogeneous isotropic turbulence (HIT) problem to the multiphase flow context, we performed two highly-resolved simulations of HIT in the presence of (i) a thin interface layer, and (ii) homogeneously distributed solid particle. These two configurations allow for the investigation of interfacial turbulence and particulate turbulence, respectively. Besides the analysis of the global flow characteristic in both cases, we evaluate the spectral contribution of each production/dissipation mechanism in the vorticity transport equation to the distribution of vortical energy (enstrophy) across the scales. We base our discussion on the role of the main inter-phase interaction mechanisms in vorticity transport (i.e. the surface tension for interfacial turbulence and drag force for particulate turbulence), and unveil a similar contribution from these mechanisms to the multiphase turbulence cascade. The results also explain the deviation of kinetic energy and enstrophy spectra of multiphase HIT problems from their single-phase similitudes, confirming the validity of this approach for establishing a universal description of multiphase turbulence.
Article
In canonical wall-bounded flows, point particle-laden turbulence exhibits a substantial interaction between scales with a variety of regimes, and the dynamics of the point particle-laden fluid are primarily identified by the Reynolds number. Such interactions are even more augmented in curved channels with variable curvature, and fixed Reynolds numbers demonstrate distinct flow behavior, as shown by Brethouwer [J. Fluid Mech. 931, A21 (2022)]. In this work, we demonstrate the characteristics of wall-bounded point particle-laden turbulent flows in sharply bent channels by evaluating the time-averaged velocity profiles at the straight section, at the bend, and in the inclined sections. The mean (time-averaged) normalized velocity profiles retain their well-known logarithmic features, with the von Kármán and additive constants taking different values depending on the acute inclination of the bend. Near-wall fluctuations at the bend are found to be intensified due to the bend that leads to increased turbulent activity. On examining the friction Reynolds number along the bent channel walls in the streamwise direction, a modulated behavior with an abrupt change at the bend is observed. Budgets of turbulent kinetic energy (TKE) are delineated for various inclinations of the bend at different sections of the channel and are compared with the unladen sharply bent turbulent channel flows, which illustrate that TKE is modulated at the bend and there is an overall attenuation of TKE on loading the channel with point particles.
Article
Full-text available
Separate and joint droplets, clusters, and voids characteristics of sprays injected in a turbulent co-flow are investigated experimentally. Simultaneous Mie scattering and interferometric laser imaging for droplet sizing along with separate hotwire anemometry are performed. The turbulent co-flow characteristics are adjusted using zero, one or two perforated plates. The Taylor-length-scale-based Reynolds number varies from 10 to 38, and the Stokes number estimated based on the Kolmogorov time scale varies from 3 to 25. The results show that the mean length scale of the clusters normalized by the Kolmogorov length scale varies linearly with the Stokes number. However, the mean void length scale is of the order of the integral length scale. It is shown that the number density of the droplets inside the clusters is approximately 7 times larger than that in the voids. The ratios of the droplets number densities in the clusters and voids to the total number density are independent of the test conditions and equal 5.5 and 0.8, respectively. The joint probability density function of the droplets diameter and clusters area shows that the droplets with the most probable diameter are found in the majority of the clusters. It is argued that intensifying the turbulence broadens the range of turbulent eddy size in the co-flow which allows for accommodating droplets with a broad range of diameters in the clusters. The results are of significance for engineering applications that aim to modify the clustering characteristics of large-Stokes-number droplets sprayed into turbulent co-flows.
Article
Full-text available
Utilizing a fan-stirred chamber and two-dimensional particle image velocimetry, we analyse the modification of homogeneous and isotropic turbulence ( $50 \leq Re_\lambda \leq 140$ , with supplementary data out to $Re_\lambda = 310$ , where $Re_\lambda$ is the longitudinal Taylor Reynolds number) induced by both a non-volatile (water) and a volatile (ethanol) isolated and anchored droplet in the range $(0.3 \leq d/\eta \leq 5.1)$ , where $d/\eta$ is the ratio of droplet diameter to the Kolmogorov length scale. The dissipation rate, $\varepsilon$ , is calculated via the corrected spatial gradient method, and the resultant fields of both turbulent kinetic energy, $k$ , and $\varepsilon$ are presented as spatial heat maps and as shell averages, ${\overline {k_{\Delta r}}}$ and ${\overline {\varepsilon _{\Delta r}}}$ , vs the radial coordinate normalized by the droplet radius, $r/R$ . The dissipation rate near the water droplet surface may exceed the corresponding unladen flow value by a factor of twenty or more. The normalized radius of recovery, $r^*$ , which designates the radial location where ${\overline {k_{\Delta r}}}$ or ${\overline {\varepsilon _{\Delta r}}}$ has returned to within 10 % of the unladen value, is reasonably expressed as $r^* \propto (d/\lambda )^{-C_2}$ in either case, where $\lambda$ is the longitudinal Taylor microscale and $C_2$ is a positive empirical fitting parameter. Recovery of ${\overline {k_{\Delta r}}}$ and ${\overline {\varepsilon _{\Delta r}}}$ may take up to 14 normalized radii when $d/\lambda$ is small. Trend line extrapolation suggests that the attenuation region becomes negligible as $d/\lambda \to 1$ . Ethanol, which evaporates up to five times faster than water, induces a much smaller dissipation spike near the surface. The mass ejection phenomenon appears to reduce the strong near-surface damping of the radial root-mean-square component. However, the radius of recovery trend for fields surrounding a volatile ethanol droplet falls directly in line with the non-volatile water droplet data for both $k$ and $\varepsilon$ , indicating that droplet vaporization has little effect on the far-field return to isotropy.
Article
Full-text available
Despite decades of investigations, there is still no consensus on whether inertial particles augment or dampen turbulence. Here, we perform the first experimental study in which the particle concentration is varied systematically across a broad range of volume fractions $\varPhi _{v}$ , from nominally one-way coupled to heavily two-way coupled regimes, keeping all other parameters constant. We utilize a zero-mean flow chamber where steady, homogeneous and approximately isotropic air turbulence is realized, with a Taylor-microscale Reynolds number $Re_{\lambda } = 150\unicode{x2013}300$ . We consider spherical solid particles of two sizes, both much smaller than the Kolmogorov length, and yielding Stokes numbers $St_{\eta } = 0.3$ and 2.6 based on the Kolmogorov time scale. By adjusting the turbulent intensity, the settling velocity parameter is kept constant for both cases, $Sv_{\eta } = V_{t}/u_{\eta }\approx 3$ (where $V_{t}$ is the still-air terminal velocity, and $u_{\eta }$ is the Kolmogorov velocity scale). Unlike previous studies focused on massively inertial particles, we find that the turbulent kinetic energy increases with particle loading, being more than doubled at $\varPhi _{v} =5\times 10^{-5}$ . This is attributed to the energy input associated with gravitational settling: the particles release their potential energy into the fluid and increase its dissipation rate, while the time scale associated with the inter-scale energy transfer is not strongly changed. Two-point statistics indicate that the energy-containing eddies become vertically elongated in the presence of falling particles, and that the latter redistribute the energy more homogeneously across the scales compared to unladen turbulence. This is rooted in an enhanced cascade, as shown by the nonlinear inter-scale energy transfer rate.
Article
This review is motivated by the fast progress in our understanding of the physics of particle-laden turbulence in the last decade, partly due to the tremendous advances of measurement and simulation capabilities. The focus is on spherical particles in homogeneous and canonical wall-bounded flows. The analysis of recent data indicates that conclusions drawn in zero gravity should not be extrapolated outside of this condition, and that the particle response time alone cannot completely define the dynamics of finite-size particles. Several breakthroughs have been reported, mostly separately, on the dynamics and turbulence modifications of small inertial particles in dilute conditions and of large weakly buoyant spheres. Measurements at higher concentrations, simulations fully resolving smaller particles, and theoretical tools accounting for both phases are needed to bridge this gap and allow for the exploration of the fluid dynamics of suspensions, from laminar rheology and granular media to particulate turbulence. Expected final online publication date for the Annual Review of Fluid Mechanics, Volume 54 is January 2022. Please see http://www.annualreviews.org/page/journal/pubdates for revised estimates.
Article
Full-text available
Many flows are kept turbulent by the action of the bottom stress, and this turbulence is also responsible for maintaining sedimenting particles in suspension and in some cases entraining more particles from the bed. A convenient 1-D analogue of these processes is provided by laboratory experiments conducted in a mixing box, where a characterizable turbulence is generated by the vertical oscillation of a horizontal grid. This paper reports the results of a series of experiments with a grid located close to the bottom boundary to simulate the action of stresses acting at a rough boundary, and compare the results with those obtained using the more extensively studied geometry in which a similar grid is located in the interior of a stirred fluid layer. -from Authors
Article
Space-time correlation measurements in the roughly isotropic turbulence behind a regular grid spanning a uniform airstream give the simplest Eulerian time correlation if we choose for the upstream probe signal a time delay which just ‘cancels’ the mean flow displacement. The correlation coefficient of turbulent velocities passed through matched narrow-band niters shows a strong dependence on nominal filter frequency ([similar] wave-number at these small turbulence levels). With plausible scaling of the time separations, a scaling dependent on both wave-number and time, it is possible to effect a good collapse of the correlation functions corresponding to wave-numbers from 0·5 cm−1, the location of the peak in the three-dimensional spectrum, to 10 cm−1, about half the Kolmogorov wave-number. The spectrally local time-scaling factor is a ‘parallel’ combination of the times characterizing (i) gross strain distortion by larger eddies, (ii) wrinkling distortion by smaller eddies, (iii) convection by larger eddies and (iv) gross rotation by larger eddies.
Article
An analysis is given for the apparent particle energy spectrum function in addition to the analysis of the effect of particles on the fluid energy spectrum in a turbulent gas-solid suspension flow. The analysis assumes that the problem of the motion of a continuous medium containing solid particles can be treated as two interacting continuous media, namely, the gas—and the solid—phase. The results obtained show that upon introduction of the particles the energy density of the fluid decreases more rapidly than for the case of pure fluid as wave number increases.
Article
We study the turbulent diffusion of particles suspended in an incompressible fluid on the basis of an Eulerian continuum picture. Account is taken of particle inertia and friction, as well as hydrodynamic interactions between the particles. We derive a time- dependent effective diffusion coefficient which describes the average dispersion of particles.
Article
The objective of this paper is to examine the modulation of isotropic decaying turbulence by particles whose diameter is smaller than the Kolmogorov length scale, and their response time, Τp, is smaller than the Kolmogorov time scale, Τk (hence microparticles), and the influence of increasing the particle inertia on the two-way coupling. The particle volume fraction is considered small enough, so that particle–particle interactions are neglected. On the other hand, the particle material density &rgr;p≫&rgr;f, the fluid density, and the mass loading of the particles is large enough to modify the carrier flow. The particle Reynolds number is smaller than unity, and the gravitational settling of the particles is neglected. We obtain an asymptotic analytical solution describing the spectrum of the instantaneous two-way coupling source term, &PSgr;p(k,t), in the equation for the fluid turbulence kinetic energy (TKE) spectrum, E(k,t), as a series in powers of the ratio (Τp/Τk). Recent results of Druzhinin and Elghobashi [Phys. Fluids 11, 602 (1999)] for particles whose Τp≪Τk show that to the zeroth order in (Τp/Τk),&PSgr;p(k,t) is proportional to the fluid spectral dissipation function, &egr;(k,t). In the present paper, the asymptotic solution is extended up to the first order in (Τp/Τk) and is applicable for particles with small but finite inertia. We also perform direct numerical simulation (DNS) of particle-laden isotropic turbulence using the Eulerian–Lagrangian approach. The results obtained for particles whose Τp⩽0.4Τk show that both the TKE and its dissipation rate, &egr;(t), as well as the spectral transfer of the fluid kinetic energy, are increased by the two-way coupling as compared to the particle-free case, and the increase is more pronounced for smaller Τp. The asymptotic solution for the two-way coupling source term spectrum, &PSgr;p(k,t), is found in good qualitative and quantitative agreement with the numerical results. Both the asymptotic solution and the DNS results for the instantaneous source term spectrum, &PSgr;p(k,t), show that as the particle response time is increased, the magnitude of the maximum of &PSgr;p(k,t) is reduced and its location is shifted toward higher wave numbers, as compared to the limiting case Τp≪Τk. The DNS results also show that for particles with sufficiently high inertia (whose Τp⩾0.5Τk), a negative peak of &PSgr;p(k,t) is created at low wave numbers, whereas the fluid spectral energy transfer is reduced, as compared to the one-way coupling case. The development of the negative peak of &PSgr;p(k,t) is accompanied by a well-pronounced preferential accumulation of particles. The net two-way coupling effect is the reduction of the TKE by particles with sufficiently high inertia (whose Τp=0.8Τk in our DNS), as compared to the particle-free flow. In this case, our results are in qualitative agreement with the DNS results of Boivin &etal; [J. Fluid Mech. 375, 235 (1998)], who considered particles whose Τp⩾1.26Τk. Therefore, our results show that there occurs a qualitative transition in the two-way coupling effect of particles on isotropic turbulence as the particle response time is increased from Τp≪Τk, in the limit of microparticles, to Τp≃Τk, for particles with finite inertia. In the case of microparticles (whose Τp≪Τk), the instantaneous spectrum of the two-way coupling source term, &PSgr;p(k,t), is positive at all wave numbers so that the particles add the energy to the fluid motion and increase the turbulence kinetic energy, as compared to the one-way coupling case. On the other hand, in the case of particles with higher inertia (whose Τp≃Τk), the positive contribution of the source term, &PSgr;p(k,t), is reduced at high wave numbers whereas a negative peak of &PSgr;p(k,t) is created at low wave numbers. In this case, the net two-way coupling effect is the reduction of the TKE by the particles, as compared to the one-way coupling case.