two-particle dispersion in isotropic turbulent flows

30
Two-Particle Dispersion in Isotropic Turbulent Flows Juan P.L.C. Salazar and Lance R. Collins Sibley School of Mechanical and Aerospace Engineering, Cornell University, Ithaca, New York 14853; email: [email protected] Annu. Rev. Fluid Mech. 2009. 41:405–32 First published online as a Review in Advance on September 25, 2008 The Annual Review of Fluid Mechanics is online at fluid.annualreviews.org This article’s doi: 10.1146/annurev.fluid.40.111406.102224 Copyright c 2009 by Annual Reviews. All rights reserved 0066-4189/09/0115-0405$20.00 Key Words two-point statistics, Lagrangian, particle tracking Abstract Two-particle dispersion is of central importance to a wide range of natu- ral and industrial applications. It has been an active area of research since Richardson’s (1926) seminal paper. This review emphasizes recent results from experiments, high-end direct numerical simulations, and modern the- oretical discussions. Our approach is complementary to Sawford’s (2001), whose review focused primarily on stochastic models of pair dispersion. We begin by reviewing the theoretical foundations of relative dispersion, fol- lowed by experimental and numerical findings for the dissipation subrange and inertial subrange. We discuss the findings in the context of the rele- vant theory for each regime. We conclude by providing a critical analysis of our current understanding and by suggesting paths toward further progress that take full advantage of exciting developments in modern experimental methods and peta-scale supercomputing. 405 Annu. Rev. Fluid Mech. 2009.41:405-432. Downloaded from www.annualreviews.org by Lomonosov Moscow State University on 01/27/14. For personal use only.

Upload: lance-r

Post on 21-Dec-2016

220 views

Category:

Documents


4 download

TRANSCRIPT

Page 1: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

Two-Particle Dispersion inIsotropic Turbulent FlowsJuan P.L.C. Salazar and Lance R. CollinsSibley School of Mechanical and Aerospace Engineering, Cornell University, Ithaca,New York 14853; email: [email protected]

Annu. Rev. Fluid Mech. 2009. 41:405–32

First published online as a Review in Advance onSeptember 25, 2008

The Annual Review of Fluid Mechanics is online atfluid.annualreviews.org

This article’s doi:10.1146/annurev.fluid.40.111406.102224

Copyright c© 2009 by Annual Reviews.All rights reserved

0066-4189/09/0115-0405$20.00

Key Words

two-point statistics, Lagrangian, particle tracking

AbstractTwo-particle dispersion is of central importance to a wide range of natu-ral and industrial applications. It has been an active area of research sinceRichardson’s (1926) seminal paper. This review emphasizes recent resultsfrom experiments, high-end direct numerical simulations, and modern the-oretical discussions. Our approach is complementary to Sawford’s (2001),whose review focused primarily on stochastic models of pair dispersion. Webegin by reviewing the theoretical foundations of relative dispersion, fol-lowed by experimental and numerical findings for the dissipation subrangeand inertial subrange. We discuss the findings in the context of the rele-vant theory for each regime. We conclude by providing a critical analysis ofour current understanding and by suggesting paths toward further progressthat take full advantage of exciting developments in modern experimentalmethods and peta-scale supercomputing.

405

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 2: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

There is no royal road to science, and only those who do not dread the fatiguing climb of its steep paths have achance of gaining its luminous summits.

Karl Marx, Das Kapital

1. INTRODUCTION

Arguably the most elemental understanding one can have of any flow field is how particles aremoved by the flow. Conceptually, it is far simpler to consider the trajectory of a particle than it isto fully comprehend the velocity vector field. The motion of particles is also important becauseof its connection with the processes of transport and mixing that impact natural and engineeringflows in such profound ways. Indeed, it is the latter application that drew the attention of some ofthe greatest minds in fluid mechanics to the study of particle motion in turbulence. Taylor’s (1922)original work on single-particle dispersion gave birth to many of the modern statistical tools weuse to study turbulence.

But it is Richardson’s (1926) work that forms the central theme of this review. In his sem-inal study, he examined the relative motion of two particles embedded in isotropic turbulence,establishing the foundations of two-particle dispersion. There is a fundamental link between theformal analysis of pair dispersion and practical problems such as the growth relative to the centerof mass of a cloud of contaminants in the atmosphere, nutrients in the ocean, or chemical speciesin a turbulent reactor. In all these examples that involve fluid flow, a nondimensional parameterthat represents the ratio of inertial and viscous forces is the Reynolds number, here defined interms of the Taylor microscale as Rλ ≡ 〈u2〉1/2λ/ν, where 〈u2〉1/2 is the root mean square of thefluctuating velocity, ν is the fluid kinematic viscosity, λ ≡

√15ν〈u2〉/〈ε〉, and ε is the turbulent en-

ergy dissipation rate. Laboratory and industrial flows are characterized by Rλ O(102–103), whereasgeophysical flows can reach O(104) and higher.

Restricting our attention to particles that are small enough and neutrally buoyant so as tobe considered passive tracers, the kinematic equations governing the separation vector r(t) indifferential and integral form are

drdt

= w(t), r(t) = r0 +∫ t

0w(t′) dt′, (1)

where w(t) is the relative velocity between the two particles, and r0 is the initial separation vector,corresponding to r(0). For the case of a turbulent motion, we are interested mainly in the averagegrowth rate of the separation distance. Defining 〈r2(t)〉 = 〈r(t) ·r(t)〉 as the mean square separationdistance (where 〈·〉 indicates an ensemble average), we can derive an exact differential equation forthis quantity:

12

d 〈r2〉dt

= 〈r(t) · w(t)〉

= 〈r0 · w(t)〉 +∫ t

0〈w(t′) · w(t)〉dt′. (2)

In integral form, this becomes

12〈|r(t) − r0|2〉 =

∫ t

0

∫ t

0〈w(t′) · w(t′′)〉dt′dt′′. (3)

Equations 2 and 3 are nonlinear owing to the implicit dependence of the relative velocity on theseparation vector. Indeed, the principal theoretical challenge is to model this relationship.

Relative dispersion goes both ways. One can also ask, given two particles that are separated bya distance r(t) at time t, on average how far apart were those particles at time t − τ? This is called

406 Salazar · Collins

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 3: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

backward dispersion, a concept originally introduced by Corrsin (1952) for single-particle disper-sion and extended by Durbin (1980) to two-particle dispersion. Backward dispersion constitutesa natural framework for studying turbulent mixing and the associated scalar dissipation (Sawfordet al. 2005, Thomson 2003).

We can divide the process of dispersion into three distinct regimes based on the separation ofthe particles relative to the turbulent scales: (a) The dissipation subrange corresponds to r(t) � η,where η ≡ (ν3/〈ε〉)1/4 is the Kolmogorov length scale; (b) the inertial subrange corresponds to η �r(t) � L, where L is the integral length scale; and (c) the diffusion subrange corresponds to r(t) �L. The analysis and scaling for each subrange are unique, so we discuss each separately below. Thediffusion range is statistically equivalent to the long-time limit of the problem of single-particledispersion initially addressed by Taylor (1922), a topic that is outside the scope of this review.

We limit this review to the forward dispersion of particle pairs in homogeneous isotropic tur-bulence (see Toschi & Bodenschatz 2009 for Lagrangian statistics of more than two particles). Ourgoal is to provide a complementary perspective to Sawford’s (2001) review, which focused primar-ily on modeling approaches. Here we emphasize the empirical evidence obtained from laboratorymeasurements, field measurements, and direct numerical simulations (DNSs), but whenever ap-propriate we introduce theory to help interpret the data or place observations into the propercontext.

In Section 2 we give the analysis for the dissipation subrange, followed by our discussion ofthe inertial subrange in Section 3. We offer a discussion of the results in Section 4 that includes atentative assessment of the theory. The discussion focuses on the more controversial findings inthe inertial subrange. Concluding remarks are given in Section 5, including suggestions for futurelaboratory and field experiments, numerical simulations, and theory.

2. DISSIPATION SUBRANGE

Dispersion in the dissipation subrange is closely related to the mixing of scalars and the rate ofchemical reaction for reacting scalars in the limit of fast chemistry, the prototypical example beingcombustion. Moreover, the deformation of microstructures in the flow such as polymer molecules( Jin & Collins 2007, Massah et al. 1993, Terrapon et al. 2004) or drops (Cristini et al. 2003) canbe linked to the rate at which points on the surface of the microstructure separate.

We begin with an overview of the theoretical framework in Section 2.1, followed by a reviewof more recent theoretical advances in Section 2.2. Section 2.3 discusses the relevant DNS andexperiments.

2.1. Theoretical Framework

The fluid velocity at two neighboring particle locations can be approximated by a Taylor seriesexpansion to first order, an assumption that is sometimes referred to as the locally linear-flowapproximation. The resulting relative velocity is w = r ·Γ, where Γ ≡ ∇u is the velocity gradienttensor in the frame of reference moving with the particle pair. Batchelor (1952b) was the first torecognize that, because the magnitude of the relative velocity is proportional to the separationdistance, the dispersion law is exponential in time: 〈r2〉 ∼ r2

0 exp(ξ t), where ξ is a parameter withunits of inverse time. Batchelor & Townsend (1956) argued that ξ = 2B/τη, where τη ≡ √

ν/〈ε〉 isthe Kolmogorov timescale and B is a dimensionless parameter known as the Batchelor constant.Based on various assumptions regarding the velocity gradient, they estimated 0.35 < B < 0.41.

Saffman (1960) later analyzed the effect of molecular diffusion on pair dispersion. Under theassumption that τc /τη � 1 (where τ c is the average time between molecular collisions), molecular

www.annualreviews.org • Dispersion in Isotropic Turbulence 407

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 4: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

diffusion enhances the dispersion rate in an additive manner to leading order. Secondarily, molec-ular diffusion affects the Lagrangian velocity correlation time. Through an analysis of the scalardiffusion equation, Saffman (1960) arrived at the following relationship for relative dispersionwith molecular diffusion in the short-time limit (three dimensions):

12

d 〈r2〉dt

= B〈r2〉τη

+ 6κ + 2κ(t − t0)2

τ 2η

, (4)

where κ is the molecular diffusivity.

2.2. Modern Theoretical Advances

It is possible to derive the Batchelor constant for the special case of a δ-correlated-in-time ve-locity. The equation for 〈r2(t)〉 is found by substituting the locally linear flow approximation intoEquation 2:

12

d 〈r2〉dt

=∫ ∞

0〈i j (t)ik(t′)r j (t)rk(t′)〉dt′. (5)

We adopt index notation for clarity, where the Einstein summation rule applies to repeated indices.By assuming that i j (t) is δ-correlated in time, Brunk et al. (1997) argued that the motion of theparticles during the time the flow is correlated could be neglected, reducing Equation 5 to

12

d 〈r2〉dt

= 〈r2(t)〉∫ ∞

0〈Si j (t)Si j (t′)〉dt′, (6)

where Si j = (i j + j i )/2 is the rate-of-strain tensor. The Lagrangian correlation for Sij (t), underthe assumption of a short correlation time, takes the form

〈Si j (t)Skm(t + τ )〉 = 〈ε〉20ν

(δikδ jm + δimδ j k − 2

3δi j δkm

)exp

(− τ

τS

), (7)

where τ S is the correlation time for the rate-of-strain tensor. Substituting this into the right-hand side of Equation 6 yields 〈ε〉τS〈r2〉/(15ν) or equivalently B = τS/(15τη). Lundgren (1981)predicted τS = √

5τη, in agreement with DNS (Yeung & Pope 1989) and experiments (Guala et al.2007), implying B = 0.15. Lundgren (1981) also predicted that the probability density function(PDF) for pair separations is log-normally distributed, a result confirmed by DNS (Girimaji &Pope 1990).

As τS/τη ∼ 1, Brunk et al.’s (1997) assumption of a white-noise process is questionable. Indeed,their subsequent work has shown the effect of finite correlation times on the interparticle collisionrate (Brunk et al. 1998a,b). This led Chun et al. (2005) to introduce a new model for the rate-of-strain tensor that relaxes this assumption. They prescribe Sij as a sequence of randomly orienteduniaxial extensional flows of finite (random) duration. The resulting flux of particles at a distancer is given by an integral relationship, which, when equated to the gradient flux approximation,yields B = 0.093.

2.3. Simulations and Experiments

Girimaji & Pope (1990) studied the growth rate of infinitesimal lines in isotropic turbulenceusing Lagrangian velocity gradient data taken from Yeung & Pope’s (1989) DNS. They analyzedd 〈ln r〉/dt and found that it approached an asymptote corresponding to B = 0.13 over the range38 ≤ Rλ ≤ 93. They noted the discrepancy with Batchelor & Townsend’s (1956) original estimateand attributed it to the lack of persistence of the rate-of-strain tensor and the tendency of vorticity

408 Salazar · Collins

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 5: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

to rotate the material line away from the principal extensional axis. A subsequent DNS studyconfirmed this hypothesis (Huang 1996). Kida & Goto (2002) introduced a different statisticalweighting to their analysis and arrived at a slightly higher value of the Batchelor constant, B = 0.17.

Guala et al. (2005) were the first to investigate pair dispersion in the dissipation subrangeexperimentally. Using stereoscopic particle tracking (PT), they were able to measure the velocitygradient tensor in the Lagrangian frame of reference. Computing the source term 〈rir j Si j /r2〉directly, they estimated B = 0.097. The result is very close to the value found by Chun et al.(2005), whose theoretical estimate was confirmed by a novel simulation of a cloud of so-calledsatellite particles surrounding each inertial particle that satisfied

dri

d t= −A/τηri + i j r j , (8)

where A is an adjustable drift parameter. They showed that the resulting PDF of particle pairseparations takes the form P (r) ∼ (η/r)A/B . By fitting the exponent (for known A), it was possibleto backcalculate B. In the limit A → 0, the DNS confirmed their prediction (B = 0.093).

3. INERTIAL SUBRANGE

In this section we present the relevant theory pertaining to relative dispersion in the inertialsubrange (Section 3.1), followed by a summary of the same for two-dimensional turbulence(Section 3.2). In Section 3.3, finite-scale statistics are discussed as an alternative measure of relativedispersion. We then review recent laboratory experiments (Section 3.4), DNSs (Section 3.5), andexperiments in the environment (Section 3.6).

3.1. Theoretical Framework (Three Dimensions)

As the particle pair separates beyond the dissipation scales, the range of motions (or eddy sizes) thatmove them apart varies with the separation distance, in which eddies of scale l ∼ r(t) are most ef-fective in the process of dispersion (Corrsin 1962). Richardson (1926) initially put forth this notionand suggested the following diffusion equation for relative dispersion in isotropic turbulence:

∂q (s , t)∂t

= 1s (d−1)

∂s

[s (d−1) K (s )

∂sq (s , t)

], (9)

where d is the Euclidean space dimension, s is the sample space variable corresponding tothe random separation distance between pairs r, q(s, t) is the PDF of the separation distancebetween particle pairs subject to the normalization [

∫ ∞0 q (s , t)4π s 2ds = 1], and K (s ) [L2 T−1] is

a scale-dependent dispersion coefficient. We note that 〈r p (t)〉 = ∫ ∞0 s p q (s , t)4π s 2ds .

Compiling measurements of the effective eddy diffusion coefficient Keff ≡ 12d

d 〈r2(t)〉dt over a

wide range of scales, Richardson found that Keff ∼ 〈r2(t)〉2/3. This result led him to model K (s ) =CRs 4/3, where CR has units [L2/3 T−1]. This since has been referred to as Richardson’s 4/3 law.

Richardson’s 4/3 law also can be viewed as the outcome of what is now widely referred to asK41 theory (Kolmogorov 1941). K41 states that, provided the Reynolds number is sufficientlylarge and boundary effects are negligible, the energy dissipation rate is determined by the largescales and is independent of the viscosity ν. The inertial subrange is argued to depend only on theaverage dissipation rate 〈ε〉 (second similarity hypothesis), and the dissipation subrange is uniquelydetermined by 〈ε〉 and ν (first similarity hypothesis).

Obukhov (1941), a pupil of Kolmogorov, recognized that the only dimensionally con-sistent form for a dispersion coefficient that depends only on s in the inertial subrange isK (s ) = k0〈ε〉1/3s 4/3, where k0 is a dimensionless constant. With this expression for K(s), Equation 9

www.annualreviews.org • Dispersion in Isotropic Turbulence 409

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 6: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

admits a self-similar solution for dispersion from a point source [i.e., q (s , 0) = δ(s )] with boundaryconditions q (0, t) = q (∞, t) = 0. The result for the three-dimensional case is

qR(s , t) = 42970

√143

2

[1

π〈r2(t)〉]3/2

exp

[−

(1287s 2

8〈r2(t)〉)1/3

], (10)

which implies a mean square displacement,

〈r2(t)〉 = g〈ε〉t3, (11)

where g = 114481 k3

0 . Following convention, we refer to this as the Richardson-Obukhov (R-O) law.Batchelor (1950) applied K41 theory to derive expressions valid for the short-time and

intermediate-time limit of Equation 2. If both the initial r0 and final separation r(t) lie withinthe inertial subrange, then

12

ddt

〈|r(t)|2〉 − 〈r0 · w(t)〉 =

⎧⎪⎨⎪⎩

113

C2 (〈ε〉r0)2/3 t for t � tB

32

g〈ε〉t2 for tB � t � TL,

(12)

where C2 is the Kolmogorov constant for the longitudinal second-order velocity structure function;tB ≡ r2/3

0 〈ε〉−1/3 is the Batchelor time, a timescale after which the dispersion process no longerdepends on r0; and TL is the Lagrangian integral timescale. Upon integration,

〈|r(t) − r0|2〉 =⎧⎨⎩

113

C2(〈ε〉r0)2/3t2 for t � tB

g〈ε〉t3 for tB � t � TL.

(13)

With his work, Batchelor unveiled the existence of a regime not yet considered for t � tB , nowreferred to as the Batchelor regime. We also refer to the second prediction of Equation 13 as theR-O law or R-O regime.

In Equation 13 we have not neglected the 〈r0 · w(t)〉 term, as was done by Batchelor. Ouelletteet al. (2006) demonstrated that this term may not be neglected when the flow is not perfectlyhomogeneous, as is often the case in a realistic experimental scenario (see below). If 〈r0 ·w(t)〉 = 0,then 〈|r(t) − r0|2〉 = 〈r2(t)〉 − r2

0 . The relation 〈r2(t)〉 = g〈ε〉t3 is often referred to as true ortraditional R-O scaling. We believe the r0 term should be retained when analyzing results, partic-ularly when the initial separation lies within the inertial subrange. Richardson (1926) considereddispersion from a point source r0 = 0, hence its absence in his original formulation.

Batchelor (1952a) advocated the use of an explicit time-dependent diffusion coefficient K (t) =k0〈ε〉t2. If we substitute this expression into Equation 9 and solve it under the same initial andboundary conditions used by Richardson, we arrive at the Batchelor PDF:

qB (s , t) =[

32π〈r2(t)〉

]3/2

exp[− 3s 2

2〈r2(t)〉]

, (14)

which also implies Equation 11, with g = 2k0.Klafter et al. (1987) noted that K41 theory does not predict a unique diffusion coefficient

but that any scaling of the form K (s , t) = k0〈ε〉a tb s c satisfying the constraints 2b + 3c = 4 anda = 1 − c /2 is both consistent with K41 and yields 〈r2(t)〉 = g〈ε〉t3, with g = (5/2a)

(3/2a)

( 43 k0a

)1/a ,where (·) is the Gamma function. The Richardson PDF is obtained by setting b = 0 (givinga = 1/3 and c = 4/3), and the Batchelor PDF is obtained by setting c = 0 (giving a = 1 andb = 2). Intermediate combinations of parameters are also admissible by theory (Hentschel &Procaccia 1984, Okubo 1962); however, the resulting distributions, q(s, t), for each combination ofparameters (a, b, c) are quite different. Other forms for q(s, t) have been obtained based on closureapproximations (Kraichnan 1966).

410 Salazar · Collins

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 7: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

Boffetta & Sokolov (2002a) remarked that implicit in the idea that relative dispersion is gov-erned by a diffusion equation is the physical assumption that the velocity field is short correlated intime. In fact, for velocity fields that are δ-correlated in time (Kraichnan ensemble), Equation 9 isexact (Falkovich et al. 2001, Kraichnan 1968). In this context, Sokolov (1999) commented that forsuch a velocity field, Equation 9 with time-dependent and mixed forms of K(s, t) is not admissible,but he also emphasized that it still may apply to other scenarios.

Navier-Stokes turbulence is scale dependent and time correlated (i.e., persistent). Hence it isnot clear if a diffusion-type equation governs the process of relative dispersion. Nonlocal effects areimportant, and it is more likely that the dispersion process is described by an integro-differentialequation (Batchelor & Townsend 1956). Such an avenue has been explored with Levy walks (Klafteret al. 1987; Shlesinger et al. 1987, 1993; Sokolov & Reigada 1999; Sokolov et al. 1999). Levy walksare a generalization of the homonymous random flights, characterized by instantaneous jumps.In Levy walks, a probability density is associated with taking a step of length r and duration t,thus avoiding the infinite mean square displacements of Levy flights because after a finite time t,a finite number of steps have been taken (Shlesinger et al. 1987, 1993).

Other researchers have introduced higher-order time derivatives to account for finite-time correlations. Monin & Yaglom (1975) considered a differential equation of the type∂3q (s ,t)

∂t3 = C q (s , t), where C is a constant and is the Laplacian. This also gives the t3 result.Ogasawara & Toh (2006) proposed a differential equation for q(s, t) based on a telegraph process.The essential features of this model are the consideration of finite-time correlations through theintroduction of a second-order time derivative term to the left-hand side of Equation 9 and the al-lowance of a difference in persistence between expansion and compression of the separation vectorby adding a drift term multiplied by a flow-dependent coefficient to the right-hand side that, theauthors argued, characterizes the average effect of coherent structures in the flow on dispersion.Kanatani et al. (K. Kanatani, T. Ogasawara & S. Toh, submitted manuscript) further explore thismodel. In a similar fashion, T. Faber & J.C. Vassilicos (submitted manuscript) have allowed for con-vergence events within the theory of straining stagnation points given by Goto & Vassilicos (2004).

Falkovich et al. (2001) questioned the possibility of establishing general properties of q(s, t).They arrived at this conclusion by rewriting Equation 2 as

12

ddt

〈|r(t) − r0|2〉 = τwL (t)〈|w|2〉, (15)

whereτw

L (t) =∫ t

0 〈w(t′) · w(t)〉dt′

〈|w|2〉 (16)

is the Lagrangian correlation time of the relative velocities. If the separation distances are in theinertial subrange, then K41 would imply τw

L (t) ∼ t. Under this circumstance, one cannot invoke thecentral limit theorem and cannot regard the dispersion process as a sum of uncorrelated increments.

Although virtually all theories predict 〈r2(t)〉 ∼ t3 at intermediate times, clearly there is noconsensus on the form of q(s, t) once that regime is reached. We consider this question in thediscussions that follow (see Sections 3.4–3.6).

3.2. Theoretical Framework (Two Dimensions)

In statistically stationary forced homogeneous and isotropic two-dimensional turbulence, tworanges of relative dispersion are identified:

12

ddt

〈r2(t)〉 ∼{

〈r2(t)〉 for r(t) � l I

〈r2(t)〉2/3 for l I � r(t) � l0,(17)

www.annualreviews.org • Dispersion in Isotropic Turbulence 411

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 8: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

where lI is the energy-injection length scale, l0 is the largest length scale in the flow, and lη ≡η−1/6ν1/2 is the enstrophy-dissipation length scale, where η = ν〈(∇ω)2〉 is the dissipation rate ofenstrophy 1

2 ω2, where ω is vorticity. The first result of Equation 17 is associated with the directenstrophy cascade and is attributed to Lin (1972), whereas the second result is for the inertialsubrange of the inverse energy cascade, analogous to Richardson’s 4/3 law in three-dimensionalturbulence. Upon integration, we have

〈r2(t)〉 ={

〈r20 〉 exp

(c ′ t

t∗)

for r(t) � l I

〈r20 〉 + g ′〈ε〉t3 for l I � r(t) � l0,

(18)

where c′ and g′ are nondimensional constants, and t∗ is a timescale defined by the enstrophydissipation rate and enstrophy in the local and nonlocal cases, respectively (Babiano et al. 1990).We refer the reader to reviews by Tabeling (2002) and Kraichnan & Montgomery (1980) for amore detailed discussion.

3.3. Finite-Scale Statistics

A complication in analyzing the time evolution of mean particle pair separations is the contam-ination from particle pairs that lie outside the regime of interest. For example, particle pairs inthe dissipation subrange that happen to separate quickly may enter the inertial subrange, therebymodifying the apparent scaling for the dissipation subrange. An alternative approach to studyingparticle dispersion is to consider particle pair statistics at fixed particle separations or intervalsinstead of at fixed times. With so-called finite-scale statistics, it is possible to ensure particle statis-tics are gathered exclusively from the subrange of interest. Below we describe two approaches toobtaining dispersion rates using finite-scale statistics.

3.3.1. Finite-size Lyapunov exponent. Artale et al. (1997) introduced the finite-size Lyapunovexponent (FSLE) as an alternative measure of turbulent relative dispersion by extending conceptsfrom dynamical systems theory. The maximum Lyapunov exponent λmax is a measure of the averagerate of exponential separation δ(t) = δ(0) exp(λmaxt) of infinitesimally close trajectories in a chaoticdynamical system (Aurell et al. 1997). The multiscale nature of turbulence then permits one todefine the FSLE by considering the average response of trajectories as a function of the separationdistance. Finite thresholds δ(n) are defined such that δ(n) = ρnδ(0), where n = 1. . .nmax, δ(0) = r0

is the initial threshold level, nmax defines the maximum threshold, and ρ = δ(n+1)/δ(n) > 1 is thethreshold rate. The time required for a given pair to move from δ(n) to δ(n+1) is denoted by Tρ (δ(n)).The mean exit time (or doubling time) is given by 〈Tρ (δ(n))〉. The FSLE is then defined as

λL (δ) = ln ρ

〈Tρ (δ(n))〉 , (19)

which in the limit of vanishing δ gives the maximum Lyapunov exponent

λmax = limδ→0

λL(δ). (20)

The FSLE has the advantage over finite-time statistics such as 〈|r(t) − r0|2〉 in that the latterhas contributions from pairs that are in different regimes of the dispersion process for any giventime. This is especially true at low to moderate Rλ, at which there is no clear separation of scales.The mean exit time, conversely, has contributions from pairs that predominantly lie between δ(n)

and δ(n+1), effectively removing contamination due to crossover from other scales (Boffetta &Sokolov 2002a). The threshold rate should be chosen judiciously so as to allow for separation ofthe different regimes. We note that 〈r2(t)〉 ∼ tγ corresponds to λ(δ) ∼ δ−2/γ .

412 Salazar · Collins

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 9: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

In the diffusion limit of a delta-correlated velocity field, it is possible to relate the mean exit timeto the Richardson constant from the first-passage problem of the Richardson diffusion equation(Boffetta & Sokolov 2002b). In three dimensions (Boffetta & Sokolov 2002a),

g = 14381

(ρ2/3 − 1)3

ρ2

δ2

〈ε〉〈Tρ (δ)〉3. (21)

3.3.2. Scale-dependent diffusivity. T. Faber & J.C. Vassilicos (submitted manuscript) have pro-posed the use of the mean square diffusivity (see Equation 2) conditioned on the separation scale asa contamination-free measure of relative dispersion. We generalize this approach by conditioningon |r(t) − r0| = ⟨

12

ddt

|r(t) − r0|2∣∣∣∣∣

⟩= 〈(r(t) − r0) · w(t)| 〉 , (22)

which in the inertial subrange takes the form

〈(r(t) − r0) · w(t)| 〉 =

⎧⎪⎪⎨⎪⎪⎩

(113

C2

)1/2

(〈ε〉r0)1/3 for t � tB

32

g〈ε〉1/3 4/3 for tB � t � TL

(23)

for the Batchelor and R-O regimes, respectively. The results above follow from Equations 12 and13. The advantage of this approach over the FSLE is that g and C2 can be determined directly. Iffurther conditioning is made on the sign of [r(t) − r0] · w(t), then one can distinguish between gf

and gb, where the indices f and b refer to forward and backward dispersion, respectively.

3.4. Laboratory Experiments

The laboratory offers a controlled environment wherein experiments can be repeated many timesunder nearly identical conditions. The single greatest disadvantage of laboratory measurementsis the limited range of Reynolds numbers that can be achieved. Recent breakthroughs in three-dimensional stereoscopic particle imaging and high-speed detection have enabled the simultaneousmeasurement of multiple particle trajectories in a turbulent flow field, a technique known as PT(Malik et al. 1993, Virant & Dracos 1997). Below we review modern PT experiments performedin laboratory turbulent flows.

3.4.1. Three-dimensional turbulence. Virant & Dracos (1997) were the first to present resultson the relative dispersion of particle pairs in a laboratory setting. They used essentially the sametracking system described by Maas et al. (1993) in an open channel flow. By tracking a cloud ofquasi-neutrally buoyant particles (d � η) released in the flow centerline, they found that q(s, t)in the homogeneous direction, perpendicular to the mean flow, was in good agreement withqB; Rλ = 101 was relatively low for inertial subrange scaling to hold; the initial cloud size waslarger than the integral scale; and there was strong shear (Ott & Mann 2000). Thus the relativeparticle pair separation PDF is expected to be Gaussian, as the particles are essentially wanderingindependently, and therefore this does not support a Batchelor PDF for relative dispersion in theinertial subrange. The measurement volume was approximately 600η (120LL) in the homogeneousdirection, where LL is a measure of the Lagrangian integral length scale.

Ott & Mann (2000) measured three-dimensional relative dispersion in turbulence generatedby a pair of oscillating grids in a water tank (Rλ ≈ 100). For particles initially separated by ∼10η,〈r2(t)〉 initially exhibited non-R-O-like scaling, which the authors attributed to a t2 dependence,

www.annualreviews.org • Dispersion in Isotropic Turbulence 413

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 10: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

consistent with the Batchelor regime. However, Franzese & Cassiani (2007) have argued that a t2

scaling may arise spuriously from the representation of a nonzero initial separation on a log-logscale. For the appropriate identification of the Batchelor regime, the initial separation must besubtracted [i.e., 〈|r(t)−r0|2〉 should be plotted preferably in coordinates compensated for Batchelorscaling]. To enhance a possible R-O scaling, the authors performed a time shift τ = t − T0 in theirdata, where T0 is given by the root (zero crossing) of a linear fit of 〈r(t)2〉1/3 in the inertial subrange.Rewriting the R-O law as 〈r2(τ )〉 = g〈ε〉τ 3, a τ 3 scaling range is observed for roughly 10τ η. Themeasured Richardson constant for this range is g = 0.5. Particles were followed for up to roughly10tB, and the measurement volume was a cube of length 7L (600η). Computed pair-separationPDFs within the inertial subrange scaling agree well with qR. The procedure of time shifting thedata by the zero crossing T0 has been criticized by Ouellette (2006), who showed that for a givenpower law ∼tm, the procedure can yield a range ∼ tγ for any γ .

Bourgoin et al. (2006) and Ouellette et al. (2006) were able to track hundreds of particles at Rλ

in excess of 800, with very high temporal resolution (up to 27,000 Hz). Turbulence was generatedby coaxial counter-rotating baffled disks in a cylindrical water tank. They achieved the high framerate using state-of-the-art CMOS (complementary metal oxide semiconductor) cameras. Theirresults show unequivocal evidence of the Batchelor regime over two decades of t/tB at Rλ ≈ 815, asseen in Figure 1a, in which the data are compensated to yield a plateau at 1 in the Batchelor range.The excellent collapse is lost if 〈|r(t)|2〉 − 〈|r0|2〉 is plotted, which suggests 〈w(t) · r0〉 �= 0. Thisis an important observation and should be considered in other experiments. The curves deviatefrom the Batchelor regime at approximately t/tB = 0.07. The decrease in the slope beyondt/tB = 0.07 suggests a transition to a regime tγ with γ < 2, in disagreement with a crossover tothe R-O regime. This is attributed to the influence of normal diffusion that occurs when particlepairs separate beyond the integral scale. A bias caused by the finite measurement volume alsois not ruled out. The pair-separation PDFs q ( s = |s − s0|, t) for several initial separations areshown in Figure 1b–g. For the smallest separation distance, there is better agreement with qR, andwith increasing initial separation distance, the distribution approaches qB. The agreement withqB may be a symptom of contamination from the large scales, as the initial separation shown inFigure 1c (387η to 430η) is in the middle of the inertial subrange (60 � r/η � 1000), as identifiedby K41 scaling of the second-order velocity structure function. The ratio of the measurementvolume length to the integral length scale is Lvol/L ≈ 0.7, where L = u′3/〈ε〉 and u′ is thefluctuating root-mean-square velocity. [The value of L = 7.1 cm was taken from an earlier papercharacterizing this facility (Voth et al. 2002).]

Berg et al. (2006) measured relative dispersion in a water tank with turbulence generated byeight rotating propellers located in the corners (Rλ = 172). By using a temporal shift similar to thatof Ott & Mann (2000), the authors found a scaling region that seems consistent with the R-O law.However, they did not claim to have found true R-O scaling. They also used their data to calculatethe pair-separation PDF, finding good agreement with qR (see Figure 2). The initial particle pairseparations are not entirely within the inertial subrange. The measured Richardson constant forforward and backward dispersion is g f = 0.55±0.05 and gb = 1.15±0.05, respectively. It is arguedthat the lack of R-O scaling in the experiments of Bourgoin et al. (2006) and Ouellette et al. (2006)may be related to a bias introduced by their small measurement volume (Lvol/L ≈ 0.7). In thecurrent study Lvol/Lint ≈ 2.5, and only pairs that start within a central subvolume are considered.

3.4.2. Two-dimensional turbulence. Two-dimensional turbulence in the laboratory has beeninvestigated mainly in soap films and in stratified layers of conducting fluid over solid substrates.In the former case, obstacles in the soap film may generate gridlike turbulence, whereas in the

414 Salazar · Collins

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 11: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

a

〈〈∆

r2 〉3 /2 q

(∆s,

t)〈

∆r2 〉3 /

2 q(∆

s, t)

〈∆

r2 〉3 /2 q

(∆s,

t)

0–1 mm 9–10 mm

19–20 mm 29–30 mm

(∆s2/〈∆r 2〉)1/3 (∆s2/〈∆r2〉)1/3

39–40 mm 49–50 mm

2 3 4 5 6 7 8 90.01

2 3 4 5 6 7 8 90.1

2 3

t/tB

1.2

1.0

0.8

0.6

0.4

0.001

102

100

10–2

10–4

10–6

10–8

102

100

10–2

10–4

10–6

10–8

102

100

10–2

10–4

10–6

10–8

102

100

10–2

10–4

10–6

10–8

102

100

10–2

10–4

10–6

10–8

102

100

10–2

10–4

10–6

10–8

0.0 0.5 1.0 1.5 2.0 2.5 0.0 0.5 1.0 1.5 2.0 2.5

0.0 0.5 1.0 1.5 2.0 2.5 0.0 0.5 1.0 1.5 2.0 2.5

0.0 0.5 1.0 1.5 2.0 2.5 0.0 0.5 1.0 1.5 2.0 2.5

〈Δ

r2 (t)〉

11/ 3

C2

(〈ε〉

r 0)2 /

3t2

bb c

d e

f gg

Figure 1(a) Mean square relative dispersion for 50 different bins of initial separations, ranging from 0–1 mm (≈0–43η) to 49–50 mm(≈2107–2150η), compensated by Batchelor scaling (C2 = 2.1). (b–g) Probability density functions (PDFs) of pair separations. The redstraight line is the Richardson PDF, whereas the blue curved line is the Batchelor PDF. The symbols show the experimentalmeasurements. Each plot shows a different initial separation; for each initial separation, PDFs from 20 times ranging from 0–20τη areshown. Rλ = 815, r(t) = |r(t) − r0|, and 1 mm ≈ 43η. Figure adapted from Ouellette et al. 2006.

www.annualreviews.org • Dispersion in Isotropic Turbulence 415

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 12: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

10–6

10–5

10–4

10–3

10–2

10–1

100

101

〈r2 〉

3/2

q(s

, t)

3.02.52.01.51.00.50.0

(s2/〈r 2〉)1/3

a

c

b

a

1.0

0.9

0.8

0.7

0.6

0.5543210

c

1.0

0.9

0.8

0.7

0.6

0.5

〈r(t

)〉2/〈

r2(t

)〉〈

r(t)〉2

/〈r2 (t

)〉

543210t/tB

t/tB

b

Figure 2(a) Pair-separation probability density functions (PDFs) for forward ( gray dots) and backward dispersion( green dots) for pairs with r0 = 12–16η. The red straight line is the Richardson PDF, and the blue curved lineis the Batchelor PDF. (b) 〈r(t)〉2/〈r2(t)〉 versus t/tB for forward dispersion. The different sequencescorrespond to different initial separations: r0 increasing upward from r0 = 4η to r0 = 28η in bins of 4η. Thehorizontal lines are the Richardson prediction at 0.67 and the Gaussian prediction at 0.85. (c) Same as inpanel b, but for the backward case. Figure adapted with permission from J. Berg, B. Luthi, J. Mann, S. Ott.2006. Backwards and forwards relative dispersion in turbulent flow: an experimental investigation. Phys. Rev.E 74:016304. Copyright (2006) by the American Physical Society.

416 Salazar · Collins

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 13: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

latter, electromagnetic forcing is the preferred technique. A recent review is provided by Kellay &Goldburg (2002). Here we describe only experiments that have measured two-particle dispersion.In addition to its likely relevance to intermediate- to large-scale geophysical flows (Charney 1971,McWilliams 1983, Provenzale 1999, Read 2001, Rhines 1979), Boffetta & Celani (2000) claimedthat two-dimensional turbulence is an ideal framework to verify the R-O regime because of theabsence of intermittency of the Eulerian velocity increments in the inverse energy cascade (Boffettaet al. 2000, Paret & Tabeling 1998), even though M.K. Rivera & R.E. Ecke (submitted manuscript)have found anomalous scaling of the Lagrangian structure function exponents by applying theextended self-similarity methodology (Benzi et al. 1993). The lower dimensionality of the flowallows for easier flow characterization and enables a larger separation of scales.

Jullien et al. (1999) generated two-dimensional turbulence in a cell filled with two thin layersof NaCl solution of different concentrations in a stable configuration. Permanent magnets locatedin the solid substrate were oriented such that their magnetization axis was in the vertical direction.The stirring was accomplished by Laplace forces originating from the interaction of the magneticfield with random current impulses driven across the cell. They measured the surface velocity fieldusing particle image velocimetry (PIV) at regular intervals (∼10 Hz). The measured flow fieldwas then used to advect numerical particles in time with a fourth-order Runge-Kutta scheme.They observed a two-dimensional inverse energy cascade that can be characterized by a Reynoldsnumber, Re L ≡ u′/ l I c ≈ 5, where c is a damping coefficient [T−1]. The results (Figure 3) displaya convincing t3 range, with a Richardson constant g ∼ 0.5. The pair-separation PDF has anexponential distribution of the form q (ξ ) ∼ exp[−αξβ ], where ξ = s /〈r2(t)〉1/2, α ∼ 2.6, andβ = 0.5±0.10. The Richardson result is β = 2/3. Another important result from this experimentis the measurement of the Lagrangian correlation time of the relative displacements τ r

L(t) ≈ 0.60t,showing that for any given time the particles are correlated for more than half of the time sincetheir release.

0.001

0.01

0.1

1

10

100

〈〈r2 (t

)〉〉 (c

m2 )

0.12 3 4 5 6 7 8 9

12 3 4 5 6 7 8 9

102 3 4 5 6 7 8 9

100t (s)

~t3

Figure 3Mean square relative dispersion. (Inset) Data compensated by t3. Figure reprinted with permission fromM.C. Jullien, J. Paret, P. Tabeling. 1999. Richardson pair dispersion in two-dimensional turbulence. Phys.Rev. Lett. 82:2872–75. Copyright (1999) by the American Physical Society.

www.annualreviews.org • Dispersion in Isotropic Turbulence 417

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 14: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

Rivera & Ecke (2005) explored a technique similar to Jullien et al.’s (1999), except that theyreplaced the bottom fluid layer with Fluorinert FC-77, which is immiscible and almost two timesheavier than water, allowing for a higher Reynolds number (Re L ≈ 20) without vertical mixingbetween the layers. The velocity field was again obtained via PIV. By advecting numerical particles,they found an approximate t3 range, albeit for initial separations and for mean scales that lie belowthe injection scale (enstrophy range), at which no such t3 range is expected. The pair-separationPDFs show a form similar to qR. By using the mean doubling time to measure relative dispersion,they showed in their data a clear exponential regime up to scales of r/ l I ≈ 1, followed by a power-law range with an exponent of 0.83 instead of the Richardson prediction of 2/3 (corresponding tot2.4 instead of t3).

3.5. Direct Numerical Simulation

DNS refers to the solution of the three-dimensional, constant-property Navier-Stokes equationswith sufficient numerical resolution to ensure that all turbulent scales are resolved adequately. Notunlike laboratory experiments, DNS can achieve only a limited range of Reynolds numbers. Con-versely, it offers the most complete description of the flow available, superior to any experiment.

Yeung (1994) was the first to study relative dispersion with a 1283 DNS of forced homogeneousisotropic turbulence. In spite of the modest Reynolds number Rλ ≈ 90, significant conclusionscould be drawn. Yeung found relative dispersion consistent with the Batchelor regime. No clear t3

range was observed or sought. For particles with r0/η ∼ 1 and for times t/τη � 30, the particle sep-aration and relative velocity were highly intermittent, as measured by the non-Gaussian skewnessand flatness factors. Forcing was accomplished using Eswaran & Pope’s (1988) stochastic scheme.

Boffetta & Sokolov (2002a) found a modest t3 range for relative particle dispersion in their 2563

DNS at Rλ ≈ 200. Energy was held constant in the two lowest wave-number modes, and fluidvelocities at particle positions were obtained via linear interpolation. The initial particle pair sepa-ration was 1/256 of the box size Lbox = 2π , such that r0/η ≈ 1.8 (G. Boffetta, personal communica-tion). The authors found good agreement with qR at short times, which seems to relax to the Gaus-sian form for increasing times. When dispersion was measured via the mean exit time, they found abroader R-O range than that obtained by the traditional measure of the mean-squared separationdistance. Assuming a Richardson form of q(s, t), they found the Richardson constant to be g � 0.55.A simple model led to Lagrangian intermittency corrections for the doubling time that are relatedto the scaling exponents of the Eulerian longitudinal velocity structure functions. The authors per-formed a similar DNS study in the inverse energy cascade of two-dimensional turbulence (Boffetta& Sokolov 2002b), for which the general conclusions above are valid. In this work the authorsderived the long-time asymptotic form for the mean exit time PDF by assuming R-O scaling andshowed that the DNS PDFs, when normalized by the mean exit time, are self-similar and in goodagreement with the theoretical result [∼ exp(−c T/〈T〉), where c is a numerical constant].

Ishihara & Kaneda (2002) used a procedure similar to Ott & Mann’s (2000), in which 〈r2(t)〉1/3

is plotted as a function of t − tB for initial separations ranging from 2.8η to 22.4η, none of whichlies in the inertial subrange. Following this procedure, they estimated a Richardson constant g ≈0.7. For this 10243 DNS, Rλ ≈ 283, and forcing was accomplished through a variable negativeviscosity in the wave-number range k < 2.5 that fixed the kinetic energy.

Building on earlier work (Yeung 1994, 2001), Yeung & Borgas (2004) found evidence of an R-Oregime with g ≈ 0.83 from a 5123 DNS at Rλ ≈ 230 and for r0/η = 16 (this initial separation wasthe closest available to inertial subrange conditions). Of great importance was the recognition ofthe temporal variability in the spatially averaged dissipation rate ε in DNS, which undergoes a low-frequency oscillation in time owing to the forcing. Their recommendation is to use ε averaged over

418 Salazar · Collins

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 15: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

the time of the dispersion process in place of the long-time average when evaluating Lagrangianuniversal scaling laws. It is also recognized that this temporal variability is strongly influenced bythe choice of the forcing scheme (Eswaran & Pope 1988).

Goto & Vassilicos (2004) used two-dimensional DNS to investigate the importance of thedensity of straining stagnation points ns on the process of turbulent relative dispersion, an ideaoriginally presented by Davila & Vassilicos (2003) who used kinematic simulations. The argumentis that pairs tend to separate violently when they encounter straining stagnation points, a viewsupported by the individual tracks from experiments in the environment (LaCasce & Ohlmann2003, Ollitrault et al. 2005) and in the laboratory ( Jullien et al. 1999, Ott & Mann 2000, Virant &Dracos 1997). From three-dimensional kinematic simulations, DNS, and a wind-tunnel experi-ment, Davila & Vassilicos (2003) found a fractal scaling ns ∼ (L/η)Ds for L/η � 1, where Ds = 2is the fractal dimension of the spatial distribution of straining stagnation points. Goto & Vassilicos(2004) further explored the idea and related the exponent γ of the mean square separation of par-ticles ∼ tγ to the fractal dimension Ds (an Eulerian quantity): γ = 2d/Ds , where d is the Euclideanspace dimension, Ds satisfies p + 2Ds /d = 3, p is the power-law exponent of the turbulent kineticenergy spectrum in the inertial subrange [E (k) ∼ k−p], and k is the wave number. They found thatthis relation holds for p = 5/3 and Ds = 4/3 obtained from DNS of two-dimensional turbulence(d = 2).1 They noted that stagnation points are not Galilean invariant and that the application ofthe above theory should be performed in a reference frame that maximizes the persistence of thestreamline topology in a statistical sense. The assumption is made that such a frame of referenceexists and that in isotropic turbulence it is the frame of reference for which the mean fluid velocityis zero. A posteriori DNS supports this argument (Goto et al. 2005).

Biferale et al. (2005) used the mean exit time to obtain an estimate for the Richardson constantof g = 0.50 ± 0.05 for a 10243 DNS at Rλ ∼ 280. This method assumes the validity of theRichardson PDF and is in good agreement with the estimate g = 0.47 obtained by fitting astraight line to 〈r2(t)〉1/3 for an initial separation of r0/η = 2.5 (well below the inertial subrange).The finite-scale statistics do not show dependence on initial pair separation, which for this studylies below r0/η = 20. The separation PDF, measured for r0/η = 1.2 pairs, agrees well with qR inthe range 40–70τη. The authors confirm observations of previous studies at similar Rλ (Boffetta &Sokolov 2002a, Ishihara & Kaneda 2002, Yeung & Borgas 2004), such as strong intermittency inthe dispersion process for particle pairs initially closely separated. Numerical results confirm thatthe correlation times τw

L (t) and τ rL(t) are proportional to the time t (see Figure 4). In this study,

forcing is accomplished by maintaining the total energy constant in the first two wave numbers,and fluctuations in ε were found to be at most 10% of the mean value.

Sawford et al. (2008) presented dispersion results for Reynolds numbers up to Rλ ≈ 650(20483), similar to those measured in experiments by Bourgoin et al. (2006) and Ouellette et al.(2006). The results are compared with earlier DNS (Biferale et al. 2005, Ishihara & Kaneda 2002,Yeung et al. 2006). Following Yeung & Borgas’s (2004) recommendation, the dissipation rate isaveraged over the duration of the dispersion process only:

〈ε1/3〉t = 1t

∫ t

0ε1/3(t′) dt′. (24)

This approach yields an improved collapse of data when compensated by inertial subrangescaling laws that depend on 〈ε1/3〉. The mean square separation distance for the data compiled in

1These relations have been explored experimentally in a two-dimensional electromagnetically driven laminar multiscale flow,in which precise control over forced stagnation points is possible. In the experiment p ≈ 2.5, Ds ≈ 0.5, and d = 2 such thatp + 2Ds /d = 3 (Rossi et al. 2006).

www.annualreviews.org • Dispersion in Isotropic Turbulence 419

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 16: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

1.0

0.8

0.6

0.4

0.2

0.0

D(τ

, t)/

(D(0

, t)

R(τ

, t)/

(R(0

, t)

–1.0 –0.8 –0.6 –0.4 –0.2 0.0τ/t τ/t

–80 –60 –40 –20 0τ/τ

ητ/τ

η

a1.0

0.8

0.6

0.4

0.2

0.0–1.0 –0.8 –0.6 –0.4 –0.2 0.0

1.00.8

0.60.4

0.20

1.00.8

0.60.4

0.20

–80 –60 –40 –20 0

b

Figure 4Normalized correlation functions for initial separation r0/η = 1.2 at different travel times t/τη = {77, 63, 49, 35, 21} (curves from leftto right). (a) D(τ, t)/D(0, t) versus τ/t. D(τ, t) = 〈w||(t)w||(t + τ )〉, where w||(t) = w(t) · r(t), and r(t) is the unit vector in the directionof r(t). (b) R(τ, t)/R(0, t) versus τ/t. R(τ, t) = 〈r(t)r(t + τ )〉. The insets show the same correlation functions plotted versus τ/τη. Figurereprinted with permission from Biferale et al. 2005. Copyright (2005) by the American Institute of Physics.

this study, compensated by R-O scaling, is shown in Figure 5a. Except for Rλ = 38 and r0/η ≥ 16,the data for different Rλ at the same r0/η collapse well for small times, during which a ballisticregime is expected for all r0/η considered. Furthermore, if the time is compensated by tB (seeFigure 5b) for r0/η > 16 and Rλ ≥ 240, the curves collapse at all times, irrespective of r0. Forr0/η ≤ 1 and all Rλ, a minimum in the compensated coordinates occurs before the R-O regime.Sawford (2001) and Boffetta & Sokolov (2002a) attributed this minimum to the contamination ofthe inertial subrange by dissipation subrange effects. For Rλ ≈ 650, the inertial subrange extendsfrom r/η ≈ 16 to r/η ≈ 300. Based on the arguments presented in Section 3.1, R-O scaling isexpected to hold after a time t � tB , and because the initial separation is at the lower end of theinertial subrange, the largest t3 range for the data should occur for r0/η = 16. This is confirmedin Figure 5a. However, contrary to expectation, t3 scaling occurs for t/TL > 1, which suggeststhat the requirement given in Equation 13 for t is too stringent and that the determining factoris the separation scale, not the timescale. We note that tB = (r0/η)2/3τη, such that, even for thecases 64 ≤ r0/η ≤ 256 (which lie in the inertial subrange), R-O scaling is expected to manifestonly after t � 16τη and t � 40τη, respectively. Figure 5b convincingly shows R-O scaling,as demonstrated by the collapse of the curves for different initial separations within the inertialsubrange for t/tB � 1. This data set constitutes the best evidence available of an R-O regime. Themethod of the 〈r2(t)〉1/3 plot described by Ott & Mann (2000) and also used by Ishihara & Kaneda(2002) is not recommended because of the ambiguity in the choice of the appropriate range oftimes over which to apply the fit. Sawford et al. (2008) argued that a more accurate estimate can

be obtained by plotting { d [〈r2(t)/r20 〉1/3]

d [t/tB ] }3. For r0 in the inertial subrange, a plateau independent ofr0 gives the Richardson constant. Figure 5c shows a compilation from several DNS data setsof estimates of g as a function of Rλ from these local slope plots along with a power law andexponential fit, which give g = 0.55 and g = 0.57, respectively, as Rλ → ∞.

3.6. Experiments in the Environment

Environmental flows produce the highest Reynolds numbers on Earth; hence theories such asthe R-O law, which was derived in the limit Rλ → ∞, have the greatest chance of being tested.

420 Salazar · Collins

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 17: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

0.001

0.01

0.1

1

10

100

1000

〈r2 (t

) – r

2 0〉/(〈

ε1 /3 〉 t

t)3

〈|r

(t) –

r0|

2 〉/(〈

ε1 /3 〉 t

t)3

0.1 1 10 100 1000t/τη

t/tB

TL(Rλ = 38) TL(Rλ = 240) TL(Rλ = 650)a

b

a

0.001

0.01

0.1

1

10

100

0.1 1 10 100 1000

b

0.680.66

0.64

0.62

0.60

0.58

0.56

0.54

0.52

g

10008006004002000

c

Figure 5(a) Compensated relative dispersion with time scaled by τη. Nominal initial separations are ( from bottom totop) r0/η = 1

4 , 1, 16, 64, and 256. The dark blue line represents Rλ = 38; the red line, Rλ = 240; the orangeline, Rλ = 650-I; and the light blue line, Rλ = 650-II. The straight lines of slope −2 indicate diffusive limits,and the horizontal line is set at 0.6. (b) The same data as in panel a, but with time scaled by tB.(c) Estimates of the Richardson constant obtained from local slope plots. The solid line is the fit0.57[1 + 0.33 exp(−Rλ/149)], and the dashed line is the fit 0.55(1 + 15.5 R−0.91

λ ). The dotted lines are the95% confidence limits for the fits. Figure reprinted with permission from Sawford et al. 2008. Copyright(2008) by the American Institute of Physics.

However, there are complications in natural flows, such as inhomogeneities of the turbulence,stratification and other buoyancy effects, Coriolis forces, two-dimensional effects due to boundaryconditions, and the vastness of the range of scales of the measurement (millimeters to kilometers)that challenge experimentalists to find measurable conditions in which the theory may apply. De-spite these challenges, historically it has been in the environment that most dispersion experimentshave been performed.

www.annualreviews.org • Dispersion in Isotropic Turbulence 421

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 18: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

Richardson (1926, 1929) reported the first measurements of particle pair dispersion, in theform of Keff. Common traits of these early experiments include both the limitation of the ex-perimental methods employed and the ingenuity of the experimenters. Among the fluid tracersused in Richardson’s first experiments were dandelion seeds, smoke puffs from cigarettes, andballoons of various sizes. In a celebrated experiment, Richardson & Stommel (1948) measuredKeff by throwing carefully cut pieces of parsnips off Blairmore Pier in Loch Long, Scotland, andmeasuring their relative positions with an optical device. Based only on two values of 〈r2〉1/2, theauthors reported a power law Keff ∼ (〈r2〉1/2)1.4.

Many experiments were conducted in the following years in the atmosphere and oceans, mostof which are described in a comprehensive chapter in Monin & Yaglom (1975, pp. 556–67). Toa large extent, the experiments of this period do not offer unequivocal evidence of any regime,although they undoubtedly support the scaling 〈r2(t)〉 ∼ tγ , with γ ∈ [2, 3].

Below we review what we believe to be the most relevant experiments of relative dispersion inthe environment. We purposely omit experiments that are not strictly Lagrangian, based on thepremise that such experiments must rely on indirect measures of two-particle dispersion.

3.6.1. Atmosphere. The EOLE experiment was conducted from October 1971 through January1972. Approximately 480 constant-volume 200-mbar-level balloons were tracked in the mid-latitudes of the southern hemisphere by the French satellite Eole (Morel & Bandeen 1973). Amongthe experiment’s objectives was the measurement of particle pair dispersion, through which apossible form of the energy spectrum in the 100–1000-km range could be inferred. Morel &Larcheveque (1974) found agreement with a k−3 spectrum based on measurements of the observedexponential growth of 1

2d 〈r2(t)〉

dt , with an e-folding time of 2.7 days. This exponential growth wasfollowed by a linear dependence in t. The EOLE data have been reanalyzed by Lacorata et al.(2004), who instead used the FSLE to investigate the scale dependence of relative dispersion.The λL values display a substantially shorter exponential regime (e-folding time of 0.4 days) thanfound originally, and contrary to a linear dependence, the latter interpretation resulted in an R-Oregime.

Er-El & Peskin (1981) analyzed the TWERL (Tropical Wind, Energy Conversion, andReference Level) experiment, conducted between June 1975 and December 1976 ( Julian et al.1977). In the TWERL experiment, 383 tetroons at the 150 mbar level were released in tropicaland mid-latitudes of the southern hemisphere and subsequently tracked via the Nimbus 6 satellite.An exponential range of relative dispersion was found for the mid-latitude releases. Although theauthors claimed to have found a t3 dependence following the exponential range, the scatter in thedata is large, and we believe this claim to be too strong.

3.6.2. Ocean. LaCasce & Ohlmann (2003) used surface-drifter tracks from SCULP (SurfaceCurrent and Lagrangian Drift Program) (Ohlmann & Niiler 2005) to study relative dispersionregimes in the Gulf of Mexico. The drifters were tracked by Doppler ranging of the ARGOSsatellite system five to seven times a day. Exponential growth was found in a range of scales from5 km to 50 km (1 to 10 days) with an e-folding time of approximately 2 days. At larger times, apower-law scaling was observed, tγ , where γ = 2.2 if fixed time statistics were used 〈r2(t)〉, butγ = 3 when the FSLE was adopted. The internal Rossby radius Rint is approximately 45 km.Because drifters are constrained to the ocean surface, the resulting flow is not divergence-free(Falkovich et al. 2001), and therefore predictions for two-dimensional turbulence do not strictlyapply. This, however, does not rule out the possibility of such a scaling (Schumacher & Eckhardt2002). Within the range of scales available, a linear regime was not found.

422 Salazar · Collins

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 19: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

101

102

103

104

105

106

101

102

103

104

105

106

〈〈r2 (t

)〉 (k

m2 )

〈r2 (t

)〉 (k

m2 )

12 3 4 5 6 7 89

102 3 4 5 6 7 89

1002 3 4 5

12 3 4 5 6 7 89

102 3 4 5 6 7 89

1002 3 4 5

t (days) t (days)

4KW(∞)t

αWt3

4KE(∞)t

αEt3a b 30 ≤ r0 < 120 km Minimum criterion (7) 10-day reinitialized (16)

30 ≤ r0 < 120 km Minimum criterion (10) 10-day reinitialized (9)

60-day reinitialized (5)

Figure 6Mean square separation distance versus time obtained from time-shifted float data during the TOPOGULF experiment: (a) westernfloats and (b) eastern floats. r0 < 7.5 km for the 10-day (crosses) and 60-day (open diamonds) reinitialized time series. Asterisks give thecorresponding evolution for portions selected with the minimum separation criterion. This consists of finding the minimum separationdistance over the entire lifetime of the pair and recording the evolution for 120 days thereafter. The remaining record is then searchedfor a new minimum and the process is repeated. Numbers of pairs are indicated within parentheses. αW and αE are numericalcoefficients, and KE (∞) and KW(∞) are the diffusion coefficients in the diffusive limit. Figure adapted from Ollitrault et al. 2005.

Ollitrault et al. (2005) analyzed a small subset of quasi-Lagrangian tracks from theTOPOGULF experiment (Arhan et al. 1989, de Verdiere et al. 1989). During TOPOGULF,26 subsurface floats at the 700-dbar level were deployed in the mid-latitude range of the NorthAtlantic, and their positions were tracked between 1983 and 1989. The subset of float data re-vealed that for initial separations less than Rint = 25 km, there is evidence of an exponential regime(t < 6 days), followed by a t3 regime from scales of 40 km to 300 km (20 to 60 days) after whicha linear t regime was observed (see Figure 6). This is consistent with a two-dimensional directenstrophy cascade at scales smaller than Rint and an inverse energy cascade at scales larger than Rint,although the authors note that there are other possible explanations for these trends, and becausethe plots were constructed under the assumption of an R-O regime (by performing the appro-priate time shift), they cannot prove its existence. Particle-separation PDFs were also computed.For initially close pairs, the PDFs are non-Gaussian, whereas for initially distant pairs, there isstronger resemblance to a Gaussian PDF, in agreement with an earlier surface-drifter experimentoff the northern California coast (Davis 1985a,b).

4. DISCUSSION

Despite the insightful beginning provided by the classic papers, there remain serious questionsabout the dispersion of particle pairs in turbulence. This is, in no small part, due to the manychallenges faced by experimentalists and numerical simulators in making measurements that canaccurately and reproducibly test the theory. Early experiments in the environment have shownthe strongest evidence of the R-O regime; however, there are a number of important caveats. Thedata often have significant scatter, which may be attributable to spatial inhomogeneities in theturbulence parameters and/or measurement errors. Vertical and horizontal mean shear (Bowden1965; Csanady 1969; Kullenberg 1972; Randerson 1972; Saffman 1962a,b; Young et al. 1982)that may enhance dispersion is often not measured. This effect may even dominate turbulentdispersion under the conditions Ri ∼ 1 and ST � 1, where Ri is the Richardson number, S is

www.annualreviews.org • Dispersion in Isotropic Turbulence 423

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 20: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

the mean shear rate, and T is the integral timescale. Moreover, drifters used as surface markers inoceans or lakes are not capable of following the inherently three-dimensional turbulent motion.Their motion on the surface is not equivalent to a two-dimensional, incompressible turbulenceowing to convergence or divergence events that render the surface flow compressible (Falkovichet al. 2001, Okubo 1970). Recent laboratory experiments and numerical simulations have shownthat dispersion on a free surface exhibits a reduced scaling exponent in comparison with the classicalR-O value (Cressman & Goldburg 2003, Cressman et al. 2004). Finally, in studies that report anR-O regime, the values of the Richardson constant can vary by orders of magnitude (Monin &Yaglom 1975).

Nature’s laboratory, while providing an enormous range of scales, exhibits an interplay ofphenomena that must be well understood and accounted for before these results can be used totest theory. Batchelor (1959), reflecting on this very issue at the 1958 International Symposium onAtmospheric Diffusion and Air Pollution at Oxford University, wrote, “needed is a more refinedversion of the Richardson-Stommel parsnip experiment, carried out with careful regard for theconditions under which the similarity theory may be expected to apply, and performed if possiblein the laboratory rather than in the atmosphere.”

Advances in PIV and PT have made this a reality. Overall the turbulence in a laboratoryapparatus is reproducible and better controlled. DNS in two or three dimensions also offers awell-controlled environment in which to study two-particle dispersion.

Results from experimental measurements appear to be converging on the properties of 〈r2(t)〉in the inertial subrange. For example, Jullien et al.’s (1999) two-dimensional experiments thatused PIV to measure the velocity field, and numerically solved for the motion of particles in theflow, yielded convincing R-O scaling. However, the same level of support was not found in Rivera& Ecke’s (2005) more recent experiment of the same nature, in which they found a power-lawrange with a reduced exponent (∼t2.4). One major difference between these experiments is thatJullien et al. performed forcing on the bottom layer, whereas Rivera & Ecke performed forcingon the top layer, on which the particles were tracked. It seems that further investigation of theinfluence of the forcing is warranted. The clever approach of numerically integrating the particlescircumvents some of the problems found with conventional PT (e.g., initialization of particle pairsat controlled separations). Some three-dimensional experiments find t3 scaling (Berg et al. 2006,Ott & Mann 2000), but at Reynolds numbers that are too low to justify attributing this to the R-Oregime. Bodenschatz and coworkers (Bourgoin et al. 2006, Ouellette et al. 2006, Xu et al. 2008)have reached the highest Reynolds numbers in three-dimensional laboratory experiments to date.Although they have found an extensive Batchelor range, they have yet to observe the R-O regime.This can be attributed to a bias caused by the measurement volume or to the Reynolds number,which, although the largest so far, still may not be large enough.

DNS studies of pair dispersion suffer from many of the same limitations as laboratory ex-periments (Boffetta & Sokolov 2002a, Yeung 1994, Yeung & Borgas 2004). Biferale et al. (2005)showed how the use of finite-scale statistics has promise for mitigating some of these problems.Most convincing is Sawford et al.’s (2008) DNS that, at 20483, had a sufficient inertial subrangeto allow the particle pairs to be initialized well within the inertial range. It is encouraging that allthese studies seem to be converging on a Richardson constant in the range 0.5 ≤ g ≤ 0.6, muchimproved relative to measurements in the environment (Monin & Yaglom 1975).

However, the state of affairs for the pair-separation PDF is far less clear. Richardson’s use of adiffusion equation can be rigorously justified only for a velocity field that is δ-correlated in timeand if the process is self-similar at all scales, which does not hold in turbulent flows. A furthercritique, raised by Sawford et al. (2005), is that δ-correlated velocity fields (or a diffusion equationfor relative dispersion) cannot differentiate between forward and backward dispersion. Sokolov

424 Salazar · Collins

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 21: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

(1999) presented heuristic arguments that convincingly show the importance of time correlationsin the analysis of relative turbulent dispersion. If one uses a unique kinematic parameter to specifythe flow over some range of scales, as is the case in K41 theory, then the mean-square separationdistance exhibits the same scaling for a δ-correlated velocity field, for the ballistic regime, andfor two-dimensional isotropic turbulence. Differences among these flows manifest in the form ofthe PDF of pair separations. The question of whether a self-similar form for the PDF even existshas been raised (Falkovich et al. 2001). However, in spite of the apparent flaws, the majority ofrecent experiments and simulations indicate a PDF most similar to the Richardson form, but ourunderstanding of this remains incomplete.

5. CONCLUDING REMARKS

The classical framework created by Richardson (1926), placed in the context of K41 by Obukhov(1941), and extended and clarified by Batchelor (1950) has held up remarkably well over time.Nothing has effectively challenged the existence of the R-O regime. Similarly, there has not beenan experiment that has unequivocally confirmed R-O scaling over a broad-enough range of timeand with sufficient accuracy.

Looking to the future, it appears that tools to solve these puzzles are emerging. We thereforethink it is appropriate to reflect on the lessons learned over the past eight decades of research. Upto now, the primary focus of most experiments has been the search for the R-O regime in the formof t3 scaling. Because nearly all self-consistent theories for the inertial subrange yield a t3 power law(Sokolov 1999), we believe that future experiments and simulations should place greater emphasison Richardson’s constant and the PDF of particle pairs. The latter in particular appears to be abetter discriminator among the various theories. A second challenge is to ensure particle pairs arewell within the inertial range, avoiding contamination from the dissipation subrange that can leadto a spurious t3 regime. To that end, we note that finite-scale statistics offer a clear advantage overanalyzing 〈r2(t)〉. We offer a few additional observations and recommendations specific to each ofthe four approaches to studying pair dispersion.

5.1. Environmental Measurements

Environmental measurements have the complication that the turbulence may not be homoge-neous. Consequently, the mean flow and turbulence should be well characterized over the domainof the experiment and for the duration of the measurement. A second challenge is the vastness ofthe domain to be studied. Experimentalists can exploit developments in remote sensing based onglobal positioning systems and so-called smart dust that contains embedded electronics to transmitits particle location (Butler 2006).

5.2. Laboratory Measurements

The goal of laboratory measurements should be to push the Reynolds number to the largestvalue possible. Supercooled liquid helium, owing to its low molecular viscosity, can achieve muchhigher Reynolds numbers than conventional fluids (Niemela et al. 2000). Likewise, a new windtunnel under construction in Gottingen (at the Max Planck Institute for Dynamics and Self-Organization, directed by Professor Eberhard Bodenschatz) will use compressed SF6 as its workingfluid to achieve Rλ of the order of 10,000. These facilities (and possibly others) can provide awell-conditioned environment to study R-O scaling. The accuracy in the measurement of theRichardson constant may be affected by errors in the measurement of the dissipation rate ε. It

www.annualreviews.org • Dispersion in Isotropic Turbulence 425

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 22: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

0.01

0.1

1

10

0.1 1 10 100

b

10–3

10–2

10–1

100

101

102

103

104

105

106

〈|r

(t) –

r 0|2〉/

(t/τ

η)3

0.1 1 10 100t/τη t/τη

a〈

|r(t

) – r 0

|2〉

0.59L0.78L1.0L1.5L

2.0L

2.0L withperiodicity

Figure 7(a) Mean square separation distance versus t/τη conditioned on volume size for pairs with r0 = 2.4η at Rλ = 284. Cubic volume oflength: 0.59L (orange), 0.78L ( green), 1L (light blue), 1.5L (dark blue), 2L (red ), and 2L with periodicity ( purple). (b) Same as in panel a,but compensated by (t/τη)3 to reveal the Richardson regime. Data from a 10243 direct numerical simulation (Biferale et al. 2005).

is recommended that more than one approach to measuring ε be used to provide an objectiveestimate of any potential bias errors. The finite measurement volume may cause a significant biasin Lagrangian statistics because longer tracks will be associated with slower particles. This hasbeen observed in Mordant et al.’s (2004) single-particle measurements. Assessment of the bias canbe observed by conditioning statistics on track length. Figure 7 illustrates this effect using DNSdata from Biferale et al. (2005). Notice that statistics conditioned on relatively small track lengthsshow no evidence of an inertial subrange t3 scaling, whereas as the conditioning length increases,the more rapidly separating pairs can be observed for a longer time, revealing what appears to beR-O scaling. Concerns over Lagrangian stationarity also have been raised (Ott & Mann 2005).Further investigation of these effects is recommended.

5.3. Direct Numerical Simulations

As with laboratory measurements, the issue of pushing the Reynolds number higher is paramount.There has been steady growth in this parameter over the past three decades, and future peta-scalecomputers (Bement 2007), or even lattice-Boltzmann-based grid computations (Celani 2007),offer the hope of extending DNS beyond the present-day maximum of 40963 (Kaneda et al.2003), corresponding to Rλ ∼ 1000. DNS with 81923 and beyond will be possible on peta-scalecomputers with 105–106 processors. Although DNS may lag behind the highest values of Rλ inexperiments, the fact that the turbulence in most DNS is forced apparently causes the turbulenceto reach the asymptotic limit at a lower Reynolds number than is required in an experiment withdecaying turbulence (Antonia & Burattini 2006).

As relative dispersion inevitably involves scales that approach the energy-containing eddies, therole of forcing must be quantitatively assessed. There are deterministic (Witkowska et al. 1997)and stochastic (Eswaran & Pope 1988) forcing algorithms that can be compared to determine whateffect (if any) they have on two-particle dispersion. A related issue is the low-frequency fluctuationin the dissipation rate that is found in forced turbulence. We recommend the approach used bySawford et al. (2008), who replaced the globally averaged dissipation rate with its value averagedover the time of each dispersion calculation.

426 Salazar · Collins

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 23: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

Homogeneity in DNS is approximated by periodic boundary conditions. Periodicity also allowsfor the analysis of particle pairs with separations that exceed the box size; however, it generates theunphysical situation that two particles separated by integer multiples of the box size in all threedirections will have the same velocity. Yeung (1994) argued that this does not have a significanteffect on Lagrangian statistics because these occurrences are rare and the turbulence is contin-uously evolving, such that the influence on two-time statistics is even weaker. This argument issupported by the results in Figure 7, which compare the mean square separations computed withand without periodicity. The former provide a more convincing R-O regime, most likely due toimproved statistics at large separations. Nevertheless, it seems reasonable that this issue should beanalyzed more carefully in future DNS.

5.4. Theory

The persistence of turbulence challenges the use of a diffusion equation. In the dissipation sub-range, persistence gives rise to nonlocal diffusion. Approaches based on a telegraph scheme showpromise (Chun et al. 2005). In the inertial subrange, the issue may be more serious. In particular,the objection raised by Falkovich et al. (2001) concerning the existence of a self-similar PDFshould be investigated further. They argued that the correlation time for the relative velocitybetween two particles grows indefinitely. Franzese & Cassiani (2007) predicted a similar result.Nevertheless, experimental measurements of PDFs show good agreement with the Richardson orBatchelor predictions (Ouellette et al. 2006), but this does not necessarily confirm R-O scaling.For example, the Gaussian prediction of Batchelor equally would apply to normal diffusion, andtherefore may represent contamination from the large scales. This needs to be further analyzedfrom a theoretical perspective.

Analyses to date have focused on the statistics of particle pairs (e.g., mean separation rate,PDF), with comparatively little attention paid to the underlying dynamics. One exception is Goto& Vassilicos’s (2004) model that connects dispersion with the local strain near stagnation points.The predictions are consistent with R-O scaling, but the proposed mechanism goes further toconnect an Eulerian property (density of straining stagnation points) to the Lagrangian dispersionprocess. Further analysis of the dynamics of particle dispersion along these lines is warranted.Statistical tests should then be devised to evaluate the validity and importance of various proposedmechanisms.

The role of intermittency (Frisch 1995) on pair dispersion remains unclear. Intermittencycorrections have been proposed that lead to a scaling of the form 〈r2(t)〉 ∼ t3(t/TE )κ , with κ ≥ 0and t/TE � 1 (Hentschel & Procaccia 1983, 1984; Shlesinger et al. 1987). The magnitude of thecorrection depends on the model used to account for intermittency. However, the procedure ofincorporating such corrections contains a certain level of ambiguity. Crisanti et al. (1987) haveadvocated a correction for 〈r p (t)〉, where p = 2, with no anomalous scaling for higher exponents,whereas others have proposed corrections that are absent for p = 2, but exist for p > 2 (Boffettaet al. 1999, Novikov 1989). It seems that verification of such corrections will require even larger Rλ.

Moreover, the role of particle inertia on two-particle dispersion should be explored. In partic-ular, the regime of small Stokes numbers can be analyzed by methods currently used with fluidparticles (with minimal further adaptation).

The most important application of two-particle dispersion is in understanding the motion ofpassive scalars in the atmosphere and oceans. Although the Reynolds numbers of natural flowsare large, conditions rarely match the assumptions made in the theory (e.g., isotropic turbulence,absence of two-dimensional effects). On a fundamental level, as discussed by Warhaft (2008), thequestion remains as to whether experiments should be designed to mimic the complexity of natural

www.annualreviews.org • Dispersion in Isotropic Turbulence 427

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 24: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

phenomena or if they should strip away the more complex reality to isolate a particular aspectof the whole. The former approach is referred to as holistic, whereas the latter is referred to asreductionist. Conversely, what is the utility of a theory that can be realized only in laboratoryflows in which all complications have been carefully removed? Ultimately, the theory must begeneralized to account for the nonidealities found in nature. That is why observations in nature,although not as controlled as in the laboratory, remain the ultimate demonstration of the utilityof the theory.

DISCLOSURE STATEMENT

The authors are not aware of any biases that might be perceived as affecting the objectivity of thisreview.

ACKNOWLEDGMENTS

We would like to thank J. Berg, L. Biferale, E. Bodenschatz, M.-C. Jullien, M. Ollitrault, B.L.Sawford, H. Xu, and P.-K. Yeung for kindly sharing their data with us. We are especially ap-preciative of A. Lannotte for reanalyzing the DNS data from Biferale et al. (2005) to produceFigure 7. L.R. Collins would like to acknowledge current and past financial support from the Na-tional Science Foundation (NSF), the National Aeronautics and Space Administration (NASA),and Defense Advanced Research Projects Agency (DARPA). This manuscript was written in partduring a visit to the Kavli Center for Theoretical Physics (KITP) at the University of Californiaat Santa Barbara that is sponsored by the NSF under grant no. PHY05-51164. J.P.L.C. Salazaracknowledges support from the Brazilian Ministry of Education through the CAPES agency.

LITERATURE CITED

Antonia RA, Burattini B. 2006. Approach to the 4/5 law in homogeneous isotropic turbulence. J. Fluid Mech.550:175–84

Arhan M, de Verdiere AC, Mercier H. 1989. Direct observations of the mean circulation at 48◦ N in theAtlantic Ocean. J. Phys. Oceanogr. 19:161–81

Artale V, Boffetta G, Celani A, Cencini M, Vulpiani A. 1997. Dispersion of passive tracers in closed basins:beyond the diffusion coefficient. Phys. Fluids 9:3162–71

Aurell E, Boffetta G, Crisanti A, Paladin G, Vulpiani A. 1997. Predictability in the large: an extension of theconcept of Lyapunov exponent. J. Phys. A 30:1–26

Babiano A, Basdevant C, Le Roy P, Sadourny R. 1990. Relative dispersion in two-dimensional turbulence.J. Fluid Mech. 214:535–57

Batchelor GK. 1950. The application of the similarity theory of turbulence to atmospheric diffusion. Quart.J. R. Meteorol. Soc. 76:133–46

Batchelor GK. 1952a. Diffusion in a field of homogeneous turbulence II. The relative motion of particles.Proc. Camb. Philos. Soc. 48:345–62

Batchelor GK. 1952b. The effect of homogeneous turbulence on material lines and surfaces. Proc. R. Soc. Lond.A 213:349–66

Batchelor GK. 1959. Some reflections on the theoretical problems raised at the symposium. Adv. Geophys.6:449–52

Batchelor GK, Townsend AA. 1956. Turbulent diffusion. In Surveys in Mechanics, ed. GK Batchelor, RMDavies, pp. 352–99. Cambridge: Cambridge Univ. Press

Bement ALJ. 2007. Cyberinfrastructure vision for 21st century discovery. Rep. NSF 07-28, NSF Cyberinfras-tructure Council, http://www.nsf.gov/pubs/2007/nsf0728/index.jsp

428 Salazar · Collins

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 25: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

Benzi R, Ciliberto S, Tripiccione R, Baudet C, Massaioli F, Succi S. 1993. Extended self-similarity in turbulentflows. Phys. Rev. E 48:R29–32

Berg J, Luthi B, Mann J, Ott S. 2006. Backwards and forwards relative dispersion in turbulent flow: anexperimental investigation. Phys. Rev. E 74:016304

Biferale L, Boffetta G, Celani A, Devenish BJ, Lanotte A, Toschi F. 2005. Lagrangian statistics of particlepairs in homogeneous isotropic turbulence. Phys. Fluids 17:115101

Boffetta G, Celani A. 2000. Pair dispersion in turbulence. Phys. A 280:1–9Boffetta G, Celani A, Crisanti A, Vulpiani A. 1999. Pair dispersion in synthetic fully developed turbulence.

Phys. Rev. E 60:6734–41Boffetta G, Celani A, Vergassola M. 2000. Inverse energy cascade in two-dimensional turbulence: deviations

from Gaussian behavior. Phys. Rev. E 61:R29–32Boffetta G, Sokolov IM. 2002a. Relative dispersion in fully developed turbulence: the Richardson’s law and

intermittency corrections. Phys. Rev. Lett. 88:094501Boffetta G, Sokolov IM. 2002b. Statistics of two-particle dispersion in two-dimensional turbulence. Phys.

Fluids 14:3224–32Bourgoin M, Ouelette NT, Xu H, Berg J, Bodenschatz E. 2006. The role of pair dispersion in turbulent flow.

Science 311:835–38Bowden KF. 1965. Horizontal mixing in the sea due to a shearing current. J. Fluid Mech. 21:83–95Brunk BK, Koch DL, Lion LW. 1997. Hydrodynamic pair diffusion in isotropic random velocity fields with

application to turbulent coagulation. Phys. Fluids 9:2670–91Brunk BK, Koch DL, Lion LW. 1998a. Observations of coagulation in isotropic turbulence. J. Fluid Mech.

371:81–107Brunk BK, Koch DL, Lion LW. 1998b. Turbulent coagulation of colloidal particles. J. Fluid Mech. 364:81–113Butler J. 2006. Everything, everywhere. Nature 440:402–5Celani A. 2007. The frontiers of computing in turbulence: challenges and perspectives. J. Turbul. 8:N34Charney JG. 1971. Geostrophic turbulence. J. Atmos. Sci. 28:1087–95Chun J, Koch DL, Rani S, Ahluwalia A, Collins LR. 2005. Clustering of aerosol particles in isotropic turbu-

lence. J. Fluid Mech. 536:219–51Corrsin S. 1952. Heat transfer in isotropic turbulence. J. Appl. Phys. 23:113–18Corrsin S. 1962. Theories of turbulent dispersion. In Int. Colloq. Turbul., pp. 27–52. Marseille: Cent. Natl.

Rech. Sci.Cressman JR, Davoudi J, Goldburg WI, Schumacher J. 2004. Eulerian and Lagrangian studies in surface flow

turbulence. New J. Phys. 6:53Cressman JR, Goldburg WI. 2003. Compressible flow: turbulence at the surface. J. Stat. Phys. 113:875–83Crisanti A, Paladin G, Vulpiani A. 1987. Is multifractal structure relevant for the turbulent diffusion? Phys.

Lett. A 126:120–22Cristini V, Blawzdziewicz J, Loewenberg M, Collins LR. 2003. Breakup in stochastic Stokes flows: sub-

Kolmogorov drops in isotropic turbulence. J. Fluid Mech. 492:231–50Csanady GT. 1969. Diffusion in an Ekman layer. J. Atmos. Sci. 26:414–26Davila J, Vassilicos JC. 2003. Richardson’s pair diffusion and the stagnation point structure of turbulence. Phys.

Rev. Lett. 91:144501Davis RE. 1985a. Drifter observations of coastal surface currents during CODE: the method and descriptive

view. J. Geophys. Res. 90:4741–55Davis RE. 1985b. Drifter observations of coastal surface currents during CODE: the statistical and dynamical

views. J. Geophys. Res. 90:4756–72de Verdiere AC, Mercier H, Arhan M. 1989. Mesoscale variability transition from the western to the eastern

Atlantic along 48◦N. J. Phys. Oceanogr. 19:1149–70Durbin PA. 1980. A stochastic model of two-particle dispersion and concentration fluctuations in homogeneous

turbulence. J. Fluid Mech. 100:279–302Er-El J, Peskin RL. 1981. Relative diffusion of constant-level balloons in the southern hemisphere. J. Atmos.

Sci. 38:2264–74Eswaran V, Pope SB. 1988. An examination of forcing in direct numerical simulations of turbulence. Comput.

Fluids 16:257–78

www.annualreviews.org • Dispersion in Isotropic Turbulence 429

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 26: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

Falkovich G, Gawedzki K, Vergassola M. 2001. Particles and fields in fluid turbulence. Rev. Mod. Phys. 73:913–75

Franzese P, Cassiani M. 2007. A statistical theory of turbulent relative dispersion. J. Fluid Mech. 571:391–417Frisch U. 1995. Turbulence. Cambridge: Cambridge Univ. PressGirimaji SS, Pope SB. 1990. Material-element deformation in isotropic turbulence. J. Fluid Mech. 220:427–58Goto S, Osborne DR, Vassilicos JC, Haigh JD. 2005. Acceleration statistics as measures of statistical persistence

of streamlines in isotropic turbulence. Phys. Rev. E 71:015301Goto S, Vassilicos JC. 2004. Particle pair diffusion and persistent streamline topology in two-dimensional

turbulence. New J. Phys. 6:65Guala M, Liberzon A, Tsinobar A, Kinzelbach W. 2007. An experimental investigation of Lagrangian corre-

lations of small-scale turbulence at low Reynolds number. J. Fluid Mech. 574:405–27Guala M, Luthi B, Liberzon A, Tsinobar A, Kinzelbach W. 2005. On the evolution of material lines and

vorticity in homogeneous turbulence. J. Fluid Mech. 533:339–59Hentschel HGE, Procaccia I. 1983. Fractal nature of turbulence as manifested in turbulent diffusion. Phys.

Rev. A 27:1266–69Hentschel HGE, Procaccia I. 1984. Relative diffusion in turbulent media: the fractal dimension of clouds.

Phys. Rev. A 29:1461–70Huang MJ. 1996. Correlations of vorticity and material line elements with strain in decaying turbulence. Phys.

Fluids 8:2203–14Ishihara T, Kaneda Y. 2002. Relative diffusion of a pair of fluid particles in the inertial subrange of turbulence.

Phys. Fluids 14:L69–72Jin S, Collins LR. 2007. Dynamics of dissolved polymer chains in isotropic turbulence. New J. Phys. 9:360Julian P, Massman W, Levanon N. 1977. The TWERL experiment. Bull. Am. Meteorol. Soc. 58:936–48Jullien MC, Paret J, Tabeling P. 1999. Richardson pair dispersion in two-dimensional turbulence. Phys. Rev.

Lett. 82:2872–75Kaneda Y, Ishihara T, Yokokawa M, Itakura K, Uno A. 2003. Energy dissipation rate and energy spectrum in

high resolution direct numerical simulations of turbulence in a periodic box. Phys. Fluids 15:L21–24Kellay H, Goldburg WI. 2002. Two-dimensional turbulence: a review of some recent experiments. Rep. Prog.

Phys. 65:845–94Kida S, Goto S. 2002. Line statistics: stretching rate of passive lines in turbulence. Phys. Fluids 14:352–61Klafter J, Blumen A, Shlesinger MF. 1987. Stochastic pathway to anomalous diffusion. Phys. Rev. A 35:3081–85Kolmogorov AN. 1941. The local structure of turbulence in an incompressible viscous fluid for very large

Reynolds numbers. Dokl. Akad. Nauk. SSSR 30:299–303Kraichnan RH. 1966. Dispersion of particle pairs in homogeneous turbulence. Phys. Fluids 9:1937–43Kraichnan RH. 1968. Lagrangian-history statistical theory for Burgers’ equation. Phys. Fluids 11:266–77Kraichnan RH, Montgomery D. 1980. Two-dimensional turbulence. Rep. Prog. Phys. 43:547–619Kullenberg G. 1972. Apparent horizontal diffusion in stratified vertical shear flow. Tellus 24:17–28LaCasce JH, Ohlmann C. 2003. Relative dispersion at the surface of the Gulf of Mexico. J. Mar. Res. 61:285–312Lacorata G, Aurell E, Legras B, Vulpiani A. 2004. Evidence for a k−5/3 spectrum from the EOLE Lagrangian

balloons in the low stratosphere. J. Atmos. Sci. 61:2936–42Lin JT. 1972. Relative dispersion in the enstrophy-cascading inertial range of homogeneous two-dimensional

turbulence. J. Atmos. Sci. 29:394–96Lundgren TS. 1981. Turbulent pair dispersion and scalar diffusion. J. Fluid Mech. 111:27–57Maas HG, Gruen A, Papantoniou D. 1993. Particle tracking velocimetry in three-dimensional flows. Part I:

photogrammetric determination of particle coordinates. Exp. Fluids 15:133–46Malik NA, Dracos T, Papantoniou D. 1993. Particle tracking velocimetry in three-dimensional flows. Part II:

particle tracking. Exp. Fluids 15:279–94Massah H, Kontomaris K, Schowalter WR, Hanratty TJ. 1993. The configurations of a FENE bead-spring

chain in transient rheological flows and in a turbulent flow. Phys. Fluids A 5:881–90McWilliams JC. 1983. On the relevance of two-dimensional turbulence to geophysical fluid motions. J. Mec.

Theor. Appl. Suppl. 1983:83–97Monin AS, Yaglom AM. 1975. Statistical Fluid Mechanics, Vol. 2. Cambridge: MIT Press

430 Salazar · Collins

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 27: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

Mordant N, Leveque E, Pinton JF. 2004. Experimental and numerical study of the Lagrangian dynamics ofhigh Reynolds turbulence. New J. Phys. 6:116

Morel P, Bandeen W. 1973. The EOLE experiment: early results and current objectives. Bull. Am. Meteorol.Soc. 54:298–306

Morel P, Larcheveque M. 1974. Relative dispersion of constant-level balloons in the 200-mb general circula-tion. J. Atmos. Sci. 31:2189–96

Niemela JJ, Skrbek L, Sreenivasan KR, Donnelly RJ. 2000. Turbulent convection at very high Rayleighnumbers. Nature 404:837–40

Novikov EA. 1989. Two-particle description of turbulence, Markov property, and intermittency. Phys. FluidsA 326:326–30

Obukhov AM. 1941. On the distribution of energy in the spectrum of turbulent flow. Izv. Akad. Nauk SSSR,Ser. Geogr. Geofiz. 5:453–66

Ogasawara T, Toh S. 2006. Turbulent relative dispersion in two-dimensional free convection turbulence.J. Phy. Soc. Jpn. 75:104402

Ohlmann JC, Niiler PP. 2005. Circulation over the continental shelf in the northern gulf of Mexico. Prog.Oceanogr. 64:45–81

Okubo A. 1962. A review of theoretical models for turbulent diffusion in the sea. J. Oceanogr. Soc. Jpn. 20:286–320

Okubo A. 1970. Horizontal dispersion of floatable particles in the vicinity of velocity singularities such asconvergences. Deep-Sea Res. 17:445–54

Ollitrault M, Gabillet C, de Verdiere AC. 2005. Open ocean regimes of relative dispersion. J. Fluid Mech.533:381–407

Ott S, Mann J. 2000. An experimental investigation of the relative diffusion of particle pairs in three-dimensional turbulent flow. J. Fluid Mech. 422:207–23

Ott S, Mann J. 2005. An experimental test of Corrsin’s conjecture and some related ideas. New J. Phys. 7:142Ouellette NT. 2006. Probing the statistical structure of turbulence with measurements of tracer particle tracks. PhD

thesis. Cornell Univ., IthacaOuellette NT, Xu H, Burgoin M, Bodenschatz E. 2006. An experimental study of turbulent relative dispersion

models. New J. Phys. 8:109Paret J, Tabeling P. 1998. Intermittency in the two-dimensional inverse cascade of energy: experimental

observations. Phys. Fluids 12:3126–36Provenzale A. 1999. Transport by coherent barotropic vortices. Annu. Rev. Fluid Mech. 31:55–93Randerson D. 1972. Temporal changes on horizontal diffusion parameters of a single nuclear debris cloud.

J. Appl. Meteorol. 11:670–73Read PL. 2001. Transition to geostrophic turbulence in the laboratory, and as a paradigm in atmospheres and

oceans. Surv. Geophys. 22:265–317Rhines PB. 1979. Geostrophic turbulence. Annu. Rev. Fluid Mech. 11:401–41Richardson LF. 1926. Atmospheric diffusion shown on a distance-neighbour graph. Proc. R. Soc. Lond. Ser. A

110:709–37Richardson LF. 1929. A search for the law of atmospheric diffusion. Beitr. Phys. Frei. Atmos. 15:24–29Richardson LF, Stommel H. 1948. Note on eddy diffusion in the sea. J. Atmos. Sci. 5:238–40Rivera MK, Ecke RE. 2005. Pair dispersion and doubling time statistics in two-dimensional turbulence. Phys.

Rev. Lett. 95:194503Rossi L, Vassilicos JC, Hardalupas Y. 2006. Electromagnetically controlled multi-scale flows. J. Fluid Mech.

558:207–42Saffman PG. 1960. On the effect of the molecular diffusivity in turbulent diffusion. J. Fluid Mech. 8:273–83Saffman PG. 1962a. Discussion: the effect of wind shear on horizontal spread from an instantaneous ground

source. Quart. J. R. Meteorol. Soc. 89:293–95Saffman PG. 1962b. The effect of wind shear on horizontal spread from an instantaneous ground source.

Quart. J. R. Meteorol. Soc. 88:670–73Sawford BL. 2001. Turbulent relative dispersion. Annu. Rev. Fluid Mech. 33:289–317Sawford BL, Yeung PK, Borgas MS. 2005. Comparison of backwards and forwards relative dispersion in

turbulence. Phys. Fluids 17:095109

www.annualreviews.org • Dispersion in Isotropic Turbulence 431

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 28: Two-Particle Dispersion in Isotropic Turbulent Flows

ANRV365-FL41-19 ARI 14 November 2008 8:57

Sawford BL, Yeung PK, Hackl JF. 2008. Reynolds number dependence of relative dispersion statistics inisotropic turbulence. Phys. Fluids 20:065111

Schumacher J, Eckhardt B. 2002. Clustering dynamics of Lagrangian tracers in free-surface flows. Phys. Rev.E 66:017303

Shlesinger MF, West BJ, Klafter J. 1987. Levy dynamics of enhanced diffusion: application to turbulence.Phys. Rev. Lett. 58:1100–3

Shlesinger MF, Zaslavsky GM, Klafter J. 1993. Strange kinetics. Nature 363:31–37Sokolov IM. 1999. Two-particle dispersion by correlated random velocity fields. Phys. Rev. E 60:5528–32Sokolov IM, Blumen A, Klafter J. 1999. Drude approach to ananomalous diffusion: application to Richardson

dispersion in turbulent flows. Europhys. Lett. 47:152–57Sokolov IM, Reigada R. 1999. Dispersion of passive particles by a quasi-two-dimensional turbulent flow. Phys.

Rev. E 59:5412–16Tabeling P. 2002. Two-dimensional turbulence: a physicist approach. Phys. Rep. 362:1–62Taylor GI. 1922. Diffusion by continuous movements. Proc. Lond. Math. Soc. 20:196–212Terrapon VE, Dubief Y, Moin P, Shaqfeh ESG, Lele SK. 2004. Simulated polymer stretch in a turbulent flow

using Brownian dynamics. J. Fluid Mech. 504:61–71Thomson DJ. 2003. Dispersion of particle pairs and decay of scalar fields in isotropic turbulence. Phys. Fluids

15:801–13Toschi F, Bodenschatz E. 2009. Lagrangian properties of particles in turbulence. Annu. Rev. Fluid Mech.

41:375–404Virant M, Dracos T. 1997. 3D PTV and its application to Lagrangian motion. Meas. Sci. Tech. 8:1539–52Voth GA, La Porta A, Crawford AM, Alexander J, Bodenschatz E. 2002. Measurement of particle accelerations

in fully developed turbulence. J. Fluid Mech. 469:121–60Warhaft Z. 2008. Laboratory studies of droplets in turbulence: towards understanding the formation of clouds.

Fluid Dyn. Res. In pressWitkowska A, Brasseur JG, Juve D. 1997. Numerical study of noise from isotropic turbulence. J. Comput.

Acoust. 5:317–36Xu H, Ouellette NT, Bodenschatz E. 2008. Evolution of geometric structures in intense turbulence. New J.

Phys. 10:013012Yeung PK. 1994. Direct numerical simulation of two-particle relative diffusion in isotropic turbulence. Phys.

Fluids A 6:3416Yeung PK. 2001. Lagrangian characteristics of turbulence and scalar transport in direct numerical simulations.

J. Fluid Mech. 427:241–74Yeung PK, Borgas MS. 2004. Relative dispersion in isotropic turbulence. Part 1. Direct numerical simulations

and Reynolds-number dependence. J. Fluid Mech. 503:93–124Yeung PK, Pope SB. 1989. Lagrangian statistics from direct numerical simulations of isotropic turbulence.

J. Fluid Mech. 207:531–86Yeung PK, Pope SB, Sawford BL. 2006. Reynolds number dependence of Lagrangian statistics in large nu-

merical simulations of isotropic turbulence. J. Turbul. 58:N58Young WR, Rhines PB, Garrett CJR. 1982. Shear-flow dispersion, internal waves and horizontal mixing in

the ocean. J. Phys. Oceanogr. 12:515–27

432 Salazar · Collins

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 29: Two-Particle Dispersion in Isotropic Turbulent Flows

AR365-FM ARI 12 November 2008 17:57

Annual Review ofFluid Mechanics

Volume 41, 2009Contents

Von Karman’s Work: The Later Years (1952 to 1963) and LegacyS.S. Penner, F.A. Williams, P.A. Libby, and S. Nemat-Nasser � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � 1

Optimal Vortex Formation as a Unifying Principlein Biological PropulsionJohn O. Dabiri � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � �17

Uncertainty Quantification and Polynomial Chaos Techniquesin Computational Fluid DynamicsHabib N. Najm � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � �35

Fluid Dynamic Mechanism Responsible for Breaking the Left-RightSymmetry of the Human Body: The Nodal FlowNobutaka Hirokawa, Yasushi Okada, and Yosuke Tanaka � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � �53

The Hydrodynamics of Chemical Cues Among Aquatic OrganismsD.R. Webster and M.J. Weissburg � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � �73

Hemodynamics of Cerebral AneurysmsDaniel M. Sforza, Christopher M. Putman, and Juan Raul Cebral � � � � � � � � � � � � � � � � � � � � � � �91

The 3D Navier-Stokes ProblemCharles R. Doering � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � 109

Boger FluidsDavid F. James � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � 129

Laboratory Modeling of Geophysical VorticesG.J.F. van Heijst and H.J.H. Clercx � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � 143

Study of High–Reynolds Number Isotropic Turbulence by DirectNumerical SimulationTakashi Ishihara, Toshiyuki Gotoh, and Yukio Kaneda � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � 165

Detached-Eddy SimulationPhilippe R. Spalart � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � 181

Morphodynamics of Tidal Inlet SystemsH.E. de Swart and J.T.F. Zimmerman � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � 203

v

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.

Page 30: Two-Particle Dispersion in Isotropic Turbulent Flows

AR365-FM ARI 12 November 2008 17:57

Microelectromechanical Systems–Based Feedback Controlof Turbulence for Skin Friction ReductionNobuhide Kasagi, Yuji Suzuki, and Koji Fukagata � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � 231

Ocean Circulation Kinetic Energy: Reservoirs, Sources, and SinksRaffaele Ferrari and Carl Wunsch � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � 253

Fluid Mechanics in Disks Around Young StarsKarim Shariff � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � 283

Turbulence, Magnetism, and Shear in Stellar InteriorsMark S. Miesch and Juri Toomre � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � 317

Fluid and Solute Transport in Bone: Flow-InducedMechanotransductionSusannah P. Fritton and Sheldon Weinbaum � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � 347

Lagrangian Properties of Particles in TurbulenceFederico Toschi and Eberhard Bodenschatz � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � 375

Two-Particle Dispersion in Isotropic Turbulent FlowsJuan P.L.C. Salazar and Lance R. Collins � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � 405

Rheology of the CytoskeletonMohammad R.K. Mofrad � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � 433

Indexes

Cumulative Index of Contributing Authors, Volumes 1–41 � � � � � � � � � � � � � � � � � � � � � � � � � � � � � 455

Cumulative Index of Chapter Titles, Volumes 1–41 � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � 463

Errata

An online log of corrections to Annual Review of Fluid Mechanics articles may be foundat http://fluid.annualreviews.org/errata.shtml

vi Contents

Ann

u. R

ev. F

luid

Mec

h. 2

009.

41:4

05-4

32. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by L

omon

osov

Mos

cow

Sta

te U

nive

rsity

on

01/2

7/14

. For

per

sona

l use

onl

y.