Stochastic suspensions of heavy particles

https://doi.org/10.1016/j.physd.2008.02.022Get rights and content

Abstract

Turbulent suspensions of heavy particles in incompressible flows have gained much attention in recent years. A large amount of work focused on the impact that the inertia and the dissipative dynamics of the particles have on their dynamic and statistical properties. Substantial progress followed from the study of suspensions in model flows which, although much simpler, reproduce most of the important mechanisms observed in real turbulence. This paper presents recent developments made on the relative motion of a pair of particles suspended in time-uncorrelated and spatially self-similar Gaussian flows. This review is complemented by new results. By introducing a time-dependent Stokes number, it is demonstrated that inertial particle relative dispersion recovers asymptotically Richardson’s diffusion associated to simple tracers. A perturbative (homogeneization) technique is used in the small-Stokes-number asymptotics and leads to interpreting first-order corrections to tracer dynamics in terms of an effective drift. This expansion implies that the correlation dimension deficit behaves linearly as a function of the Stokes number. The validity and the accuracy of this prediction is confirmed by numerical simulations.

Introduction

The current understanding of passive turbulent transport profited significantly from studies of the advection by random fields. In particular, flows belonging to the so-called Kraichnan ensemble–i. e. spatially self-similar Gaussian velocity fields with no time correlation–which was first introduced in the late 1960s by Kraichnan [1], led in the mid-1990s to a first analytical description of anomalous scaling in turbulence (see [2] for a review). More recently, much work is devoted to a generalization of this passive advection to heavy particles that, conversely to tracers, do not follow the flow exactly but lag behind it due to their inertia. The particle dynamics is thus dissipative even if the carrier flow is incompressible. This paper provides an overview of several recent results on the dynamics of very heavy particles suspended in random flows belonging to the Kraichnan ensemble.

The recent shift of focus to the transport of heavy particles is motivated by the fact that in many natural and industrial flows finite-size and mass effects of the suspended particles cannot be neglected. Important applications encompass rain formation [3], [4], [5] and suspensions of biological organisms in the ocean [6], [7], [8]. For practical purposes, the formation of particle clusters due to inertia is of central importance as the presence of such inhomogeneities significantly enhances interactions between the suspended particles. However, detailed and reliable predictions on collision or reaction rates, which are crucial to many applications, are still missing.

Two mechanisms compete in the formation of clusters. First, particles much denser than the fluid are ejected from the eddies of the carrier flow and concentrate in the strain-dominated regions [9]. Second, the dissipative dynamics leads the particle trajectories to converge onto a fractal, dynamically evolving attractor [10], [11]. In many studies, a carrier velocity field with no time correlation–and thus no persistent structures–is used to isolate the latter effect. As interactions between three or more particles are usually subdominant, most of the interesting features of monodisperse suspensions can be captured by focusing on the relative motion of two particles separated by R: R̈=1τ[Ṙδu(R,t)], where dots denote time derivatives and τ the particle response time. The fluid velocity difference δu is a Gaussian vector field with correlation δui(r,t)δuj(r,t)=2bij(rr)δ(tt). In order to model turbulent flows, the tensorial structure of the spatial correlation bij(r) is chosen to ensure incompressibility, isotropy and scale invariance, namely bij(r)=D1r2h[(d1+2h)δij2hrirj/r2], where h relates to the Hölder exponent of the fluid velocity field and D1 measures the intensity of its fluctuations. In particular, h=1 corresponds to a spatially differentiable velocity field, mimicking the dissipative range of a turbulent flow, while h<1 models rough flows as in the inertial range of turbulence. In this paper we mostly focus on space dimensions d=1 and d=2; extensions to higher dimensions are just sketched.

The above depicted model flow has the advantage that the particle dynamics is a Markov process. In particular, Gaussianity and δ-correlation in time of the fluid velocity field imply that the probability density p(r,v,t|r0,v0,t0) of finding the particles at separation R(t)=r and with relative velocity Ṙ(t)=v at time t, when R(t0)=r0 and Ṙ(t0)=v0 is a solution of the Fokker–Planck equation tp+i(ri1τvi)(vip)i,jbij(r)τ2vivjp=0, with the initial condition p(r,v,t0)=δ(rr0)δ(vv0). To maintain a statistical steady state, the Fokker–Planck equation (4) as well as the stochastic differential equation (1) should be supplemented by boundary conditions, here chosen to be reflective at a given distance L.

For smooth flows (h=1), the intensity of inertia is generally measured by the Stokes number St, defined as the ratio between the particle response time τ and the fluid characteristic time scale. For St0, particles recover the incompressible dynamics of tracers. In the opposite limit where St is very large, inertia effects dominate and the dynamics approaches that of free particles. In the above depicted model, the Stokes number is defined by nondimensionalizing τ by the typical fluid velocity gradient, i.e. St=D1τ. Note that by rescaling the physical time by τ, it is straightforward to recognize that the dynamics depends solely on St.

Similarly it can be checked that in rough flows (h<1)–with an additional rescaling of the distances by a factor (D1τ)1/(22h)–the dynamics of a particle pair at a distance r only depends on the local Stokes number St(r)=D1τ/r2(1h). This dimensionless quantity, first introduced in [12] and later used in [13], is a generalization of the Stokes number to cases in which the fluid turnover times depend on the observation scale. At large scales, St(r)0 and inertia becomes negligible. Particle dynamics thus approaches that of tracers. At small scales, St(r) and the particle and fluid motions decorrelate, so that the inertial particles move ballistically. In both the large and small Stokes number asymptotics, particles distribute uniformly in space, while inhomogeneities are expected at intermediate values of St(r).

The paper is organized as follows. In Section 2, an approach originally proposed in [14] is used to reduce the dynamics of the particle separation to a system of three stochastic equations with additive noises. This formulation is useful for both numerical and analytical purposes, particularly when studying the statistical properties of particle pairs. In Section 3, we introduce the correlation dimension to quantify clustering as well as the approaching rate which measures collisions. Numerical results for these quantities are reported. In Section 4 we introduce the notion of time-dependent Stokes number which makes particularly transparent the interpretation of the behaviour of the long-time separation between particles. We show how Richardson dispersion, as for tracers, is recovered in the long-time asymptotics. Section 5 briefly summarizes some exact results that can be obtained for the one-dimensional case. Sections 6 Small Stokes number asymptotics, 7 Large Stokes number asymptotics are dedicated to the small and large Stokes number asymptotics, respectively. In particular, the former presents an original perturbative approach which turned out to predict, in agreement with numerical computations, the behaviour of the correlation dimension that characterizes particle clusters. Finally, Section 8 encompasses conclusions, open questions and discusses the relevance of the considered model for real suspensions in turbulent flows.

Section snippets

Reduced dynamics for the two-point motion

In this Section we focus on planar suspensions (d=2). Following the approach proposed in [14] and with the notation R=|R|, the change of variables σ1=(L/R)1+hRṘ/L2,σ2=(L/R)1+h|RṘ|/L2,ρ=(R/L)1h, is introduced to reduce the original system of 2d=4 stochastic equations to the following one of only three equations: σ̇1=σ1/τ[hσ12σ22]/ρ+Cη1,σ̇2=σ2/τ(1+h)σ1σ2/ρ+(1+2h)Cη2,ρ̇=(1h)σ1, where C=2D1/(τL1h)2 and ηi denote two independent white noises. Reflective boundary conditions at R=L in

Correlation dimension and approaching rate

Particle clustering is often quantified by the radial distribution function g(r), which is defined as the ratio between the number of particles inside a thin shell of radius r centred on a given particle and the number which would be in this shell if the particles were uniformly distributed. This quantity enters models for the collision kernel [17]. Following [10], [13], [16], [18], we consider a different, but related way to characterize particle clustering. Instead of the radial distribution

Stretching rate and relative dispersion

This section is devoted to the study of the behaviour of the distance R(t) between two particles at intermediate times t such that R(0)R(t)L. For convenience, we drop the reflective boundary condition at R=L and consider particles evolving in an unbounded domain.

We first consider a differentiable fluid velocity field (h=1). In this case, the time evolution of the distance R(t) is given by (11), so that R(t)=R(0)exp[0tσ1(t)dt] and the particle separation can be measured by the stretching

Exact results in one dimension

A number of analytical results were derived for one-dimensional flows [29], [30], [31]. Although such flows are always compressible, their study helps improving the intuition for the dynamics of inertial particles in higher-dimensional random flows. In particular, several results on caustic formation hold also in two-dimensional (incompressible) flows because the typical velocity fluctuations, which lead to caustic formation, are effectively one-dimensional.

Here, we focus on one-dimensional

Small Stokes number asymptotics

This section reports some asymptotic results related to the limit of small particle inertia. The first part summarizes the approach developed by Mehlig, Wilkinson, and collaborators for differentiable flows (h=1). In analogy to the WKB approximation in quantum mechanics (see, e.g. [32]), the authors construct perturbatively the steady solution to the Fokker–Planck equation associated to the reduced system (8), (9). In the second part of this section original results are reported where the

Large Stokes number asymptotics

Particles with huge inertia (St1) take an infinite time to relax to the velocity of the carrier fluid. They become therefore uncorrelated with the underlying flow and evolve with ballistic dynamics, moving freely and maintaining, almost unchanged, their initial velocities. This limit is particularly appealing for deriving asymptotic theories [16]. In this section, we focus on two aspects, namely the problem of the recovery of homogeneous/uniform distribution for St1 and the problem of the

Remarks and conclusions

Before concluding this paper the results discussed so far are commented in the light of what is known about real turbulent suspensions, which are relevant to most applications. Let us start by recalling the main features of turbulent flows. Turbulence is a multiscale phenomenon [46] which spans length scales ranging from a large (energy injection) scale L to the very small (dissipative) scale η, often called the Kolmogorov scale. This hierarchy of length scales is associated with a hierarchy of

Acknowledgments

We acknowledge useful discussions with S. Musacchio and M. Wilkinson. Part of this work was done while K.T. was visiting Lab. Cassiopée in the framework of the ENS-Landau exchange programme. J.B. was partially supported by ANR “DSPET” BLAN07-1_192604.

References (52)

  • S. Sundby et al.

    Feeding conditions of Arcto-Norwegian cod larvae compared with the Rothschild–Osborn theory on small-scale turbulence and plankton contact rates

    J. Plankton Res.

    (1990)
  • J. Mann et al.

    Predator–prey encounters in turbulent waters

    Phys. Rev. E

    (2002)
  • R.C. Hogan et al.

    Stokes and Reynolds number dependence of preferential particle concentration in simulated three-dimensional turbulence

    Phys. Fluids

    (2001)
  • J. Bec

    Fractal clustering of inertial particles in random flows

    Phys. Fluids

    (2003)
  • G. Falkovich et al.
  • J. Bec et al.

    Clustering of heavy particles in random self-similar flow

    Phys. Rev. E

    (2007)
  • L.I. Piterbarg

    The top Lyapunov exponent for stochastic flow modeling the upper ocean turbulence

    SIAM J. Appl. Math.

    (2002)
  • K. Duncan et al.

    Clustering by mixing flows

    Phys. Rev. Lett.

    (2005)
  • W.C. Reade et al.

    Effect of preferential concentration on turbulent collision rates

    Phys. Fluids

    (2000)
  • J. Bec et al.

    Clustering and collisions of heavy particles in random smooth flows

    Phys. Fluids

    (2005)
  • J.-P. Eckmann et al.

    Ergodic theory of chaos and strange attractors

    Rev. Modern Phys.

    (1985)
  • L. Arnold, Random Dynamical Systems, Springer Monographs in Mathematics, Berlin, New York,...
  • T.D. Sauer et al.

    Are the dimensions of a set and its image equal under typical smooth functions?

    Ergodic Theory Dynam. Syst.

    (1997)
  • L.-P. Wang et al.

    On the collision rate of small particles in isotropic turbulence. Part I. Zero-inertia case

    Phys. Fluids

    (1998)
    Y. Zhou et al.

    On the collision rate of small particles in isotropic turbulence. Part II. Finite-inertia case

    Phys. Fluids

    (1998)
  • B. Mehlig et al.

    Coagulation by random velocity fields as a Kramers problem

    Phys. Rev. Lett.

    (2004)
  • B. Mehlig et al.

    Aggregation of inertial particles in random flows

    Phys. Rev. E

    (2005)
  • Cited by (40)

    View all citing articles on Scopus
    View full text