26.11.2012 Views

Non-Newtonian turbulence: viscoelastic fluids and binary mixtures.

Non-Newtonian turbulence: viscoelastic fluids and binary mixtures.

Non-Newtonian turbulence: viscoelastic fluids and binary mixtures.

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

UNIVERSITÀ DEGLI STUDI DI TORINO<br />

DIPARTIMENTO DI FISICA GENERALE “A. AVOGADRO”<br />

UNIVERSITÉ DE NICE-SOPHIA ANTIPOLIS<br />

UFR SCIENCES<br />

École Doctorale “Sciences fondamentales et appliquées”<br />

DOTTORATO DI RICERCA IN FISICA<br />

XIX CICLO<br />

<strong>Non</strong>-<strong>Newtonian</strong> <strong>turbulence</strong>:<br />

<strong>viscoelastic</strong> <strong>fluids</strong> <strong>and</strong> <strong>binary</strong> <strong>mixtures</strong>.<br />

Tesi presentata da: Tutors:<br />

Dr. Stefano Berti Prof. Guido Boffetta<br />

Dr. Antonio Celani<br />

Coordinatore del ciclo:<br />

Prof. Stefano Sciuto<br />

ANNI ACCADEMICI 2003 / ’04 - 2004 / ’05 - 2005 / ’06<br />

SETTORE SCIENTIFICO-DISCIPLINARE: FIS01


UNIVERSITÀ DEGLI STUDI DI TORINO<br />

DIPARTIMENTO DI FISICA GENERALE “A. AVOGADRO”<br />

UNIVERSITÉ DE NICE-SOPHIA ANTIPOLIS<br />

UFR SCIENCES<br />

École Doctorale “Sciences fondamentales et appliquées”<br />

THÈSE<br />

Présentée pour obtenir le titre de<br />

Docteur en SCIENCE<br />

Spécialité: Physique<br />

par<br />

Stefano Berti<br />

<strong>Non</strong>-<strong>Newtonian</strong> <strong>turbulence</strong>:<br />

<strong>viscoelastic</strong> <strong>fluids</strong> <strong>and</strong> <strong>binary</strong> <strong>mixtures</strong>.<br />

Soutenue le 1 décembre 2006 devant le jury composé de:<br />

Directeur Guido Boffetta<br />

Co-directeur Antonio Celani<br />

Examinateur Mario Ferraro<br />

Examinateur Andrea Mazzino<br />

Examinateur Miguel Onorato<br />

Examinateur Alain Pumir<br />

Rapporteur Hamid Kellay<br />

Università di Torino, Torino, Italia


Riassunto<br />

Questa tesi presenta uno studio teorico e numerico del problema della turbolenza<br />

in fluidi non <strong>Newtonian</strong>i. La dinamica di questi sistemi può essere modellizzata<br />

nel contesto del trasporto di campi attivi e costituisce un tema di interesse generale<br />

per la fisica dei fluidi complessi. Le loro proprietà reologiche peculiari li rendono,<br />

inoltre, interessanti per applicazioni ingegneristiche.<br />

La maggior parte del lavoro riguarda il problema della turbolenza in soluzioni<br />

diluite di polimeri, ovvero fluidi <strong>viscoelastic</strong>i. Vengono considerate due questioni:<br />

la statistica delle piccole scale, a valori di elasticità moderati in regime<br />

di turbolenza pienamente sviluppata; la destabilizzazione di un flusso laminare<br />

attraverso nonlinearità puramente elastiche.<br />

L’effetto dell’aggiunta di polimeri sulle piccole scale della turbolenza viene studiato<br />

attraverso un modello semplificato di fluido <strong>viscoelastic</strong>o, in una configurazione<br />

omogenea isotropa. Sono state considerate le modifiche indotte sulla<br />

cascata turbolenta e sono state esaminate le loro conseguenze per la statistica a<br />

piccola scala, in particolare per l’accelerazione.<br />

Nel limite opposto di nonlinearità inerziali trascurabili, un flusso può essere destabilizzato<br />

dai gradi di libertà polimerici, purché l’elastictà della soluzione sia sufficientemente<br />

elevata. Al crescere dell’elasticità si osserva una transizione verso<br />

stati caotici ed, alla fine, turbolenti. La fenomenologia che emerge dagli esperimenti<br />

è stata riprodotta numericamente per un flusso bidimensionale, e sono state<br />

caratterizzate le proprietà statistiche.<br />

Un altro tema preso in considerazione è quello delle miscele binarie. È stata esaminata<br />

la separazione di fase tra due fluidi, studi<strong>and</strong>o il processo di ordinamento<br />

in presenza di un campo di velocità forzato dall’esterno. È stata analizzata la competizione<br />

tra forze di segregazione termodinamiche e sforzi di taglio locali, sia in<br />

miscele attive che passive. Infine, si è sottolineato il ruolo marginale del caos<br />

Lagrangiano nel fenomeno dell’arresto della crescita dei domini di fase.<br />

i<br />

i


ii<br />

Resumé<br />

Cette thèse présente une étude théorique et numérique du problème de la <strong>turbulence</strong><br />

dans des fluides non-Newtoniens. La dynamique de ces systèmes peut<br />

être modelisée dans le contexte du transport de champs actifs et constitue un sujet<br />

d’intérêt général pour la physique des fluides complexes. Leurs propriétés<br />

rhéologiques particulières les rendent, en outre, intéressants pour des applications<br />

en ingénierie.<br />

La plus gr<strong>and</strong>e partie du travail regarde le problème de la <strong>turbulence</strong> dans des<br />

solutions diluées de polymères, voire des fluides viscoélastiques. Deux questions<br />

sont considérées: la statistique aux petites échelles, pour des valeurs d’élasticité<br />

modérées dans un régime de <strong>turbulence</strong> pleinement développée; la déstabilisation<br />

d’un écoulement laminaire par nonlinéarités purement élastiques.<br />

L’effet de l’addition de polymères sur les petites échelles de la <strong>turbulence</strong> est<br />

étudié à travers un modèle simplifié de fluide viscoélastique, dans une configuration<br />

homogène isotropique. Les modifications produites sur la cascade turbulente<br />

ont été considérées, et leurs conséquences pour la statistique à petite échelle, notamment<br />

l’accélération, ont été examinées.<br />

Dans la limite opposée de nonlinéarités inertielles négligeables, un écoulement<br />

peut être déstabilisé par les degrés de liberté polymériques, pourvu que l’élasticité<br />

de la solution soit suffisamment élevée. En augmentant l’élasticité on observe une<br />

transition vers des états chaotiques et, finalement, turbulents. La phénoménologie<br />

qui émerge des expériences a été reproduite numériquement pour un écoulement<br />

bidimensionnel, et les propriétés statistiques ont été caractérisées.<br />

Un autre sujet pris en considération est celui des mélanges binaires. On a examiné<br />

la séparation de phase entre deux fluides, en étudiant le procès d’organisation en<br />

présence d’un champ de vitesse forcé. On a analysé la compétition entre les forces<br />

de ségrégation thermodynamiques et les cisaillements locaux, aussi bien dans les<br />

mélanges actifs que passifs. On a enfin souligné le rôle marginal du chaos Lagrangien<br />

pour le phénomène de l’arrêt de la croissance des domaines de phase.<br />

ii


Abstract<br />

This thesis presents a theoretical <strong>and</strong> numerical study of the problem of <strong>turbulence</strong><br />

in non-<strong>Newtonian</strong> <strong>fluids</strong>. The dynamics of these systems can be modeled<br />

in terms of transported active fields <strong>and</strong> constitutes a subject of general interest<br />

in the physics of complex <strong>fluids</strong>. Their peculiar rheological properties make them<br />

attractive also for engineering applications.<br />

The major part of the work concerns the problem of <strong>turbulence</strong> in dilute polymer<br />

solutions, i. e. <strong>viscoelastic</strong> <strong>fluids</strong>. Two issues are considered: the small-scale<br />

statistics at moderate values of elasticity in a regime of fully developed <strong>turbulence</strong>;<br />

the destabilization of a laminar flow by means of purely elastic nonlinearities.<br />

The effect of polymer addition on small-scale <strong>turbulence</strong> has been investigated<br />

within a simplified <strong>viscoelastic</strong> fluid model, in a homogeneous isotropic configuration.<br />

The modifications induced on the turbulent cascade have been addressed,<br />

<strong>and</strong> their consequences on small-scale statistics, such as acceleration, have been<br />

examined.<br />

In the opposite limit of negligible inertial nonlinearities, polymeric degrees of<br />

freedom can destabilize a flow, provided the elasticity of the solution is sufficiently<br />

large. At growing elasticity, a transition to chaotic, <strong>and</strong> finally turbulent,<br />

states is observed. The basic experimental phenomenology has been numerically<br />

reproduced for a two-dimensional flow, <strong>and</strong> the statistical properties have been<br />

characterized.<br />

Another item considered is that of <strong>binary</strong> <strong>mixtures</strong>. Phase separation between two<br />

<strong>fluids</strong> has been investigated, studying the phase ordering process in the presence<br />

of an externally forced velocity field. The competition between thermodynamic<br />

segregation forces <strong>and</strong> local shears has been examined in both active <strong>and</strong> passive<br />

<strong>mixtures</strong>. Finally, the marginal role of Lagrangian chaos in the phenomenon of<br />

coarsening arrest has been highlighted.<br />

iii<br />

iii


Contents<br />

Riassunto . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . i<br />

Resumé . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ii<br />

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii<br />

Introduction 1<br />

Part I 5<br />

1 <strong>Newtonian</strong> <strong>turbulence</strong> 7<br />

1.1 Navier-Stokes equation . . . . . . . . . . . . . . . . . . . . . . . 8<br />

1.1.1 Reynolds number . . . . . . . . . . . . . . . . . . . . . . 9<br />

1.1.2 Energy balance . . . . . . . . . . . . . . . . . . . . . . . 10<br />

1.1.3 Energy transfer . . . . . . . . . . . . . . . . . . . . . . . 11<br />

1.2 Phenomenology of <strong>turbulence</strong> . . . . . . . . . . . . . . . . . . . 15<br />

1.2.1 Turbulent cascade . . . . . . . . . . . . . . . . . . . . . . 15<br />

1.2.2 Kolmogorov K41 theory . . . . . . . . . . . . . . . . . . 16<br />

1.2.3 Intermittency . . . . . . . . . . . . . . . . . . . . . . . . 19<br />

1.2.4 Acceleration statistics . . . . . . . . . . . . . . . . . . . 21<br />

1.3 Two-dimensional <strong>turbulence</strong> . . . . . . . . . . . . . . . . . . . . 25<br />

1.3.1 Vorticity equation . . . . . . . . . . . . . . . . . . . . . . 25<br />

1.3.2 Inviscid invariants . . . . . . . . . . . . . . . . . . . . . 26<br />

1.3.3 Direct <strong>and</strong> inverse cascades . . . . . . . . . . . . . . . . 27<br />

2 Polymer solutions 31<br />

2.1 Polymer dynamics in <strong>fluids</strong> . . . . . . . . . . . . . . . . . . . . . 32<br />

2.1.1 Freely jointed chain . . . . . . . . . . . . . . . . . . . . . 33<br />

2.1.2 Weissenberg number . . . . . . . . . . . . . . . . . . . . 36<br />

2.1.3 Dumbbell model . . . . . . . . . . . . . . . . . . . . . . 37<br />

2.2 Hydrodynamic models . . . . . . . . . . . . . . . . . . . . . . . 41<br />

2.2.1 Oldroyd-B model . . . . . . . . . . . . . . . . . . . . . . 41<br />

2.2.2 FENE-P model . . . . . . . . . . . . . . . . . . . . . . . 44<br />

v


vi CONTENTS<br />

2.3 Drag reduction . . . . . . . . . . . . . . . . . . . . . . . . . . . 46<br />

2.4 Elastic <strong>turbulence</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . 48<br />

3 Phase separation in <strong>binary</strong> <strong>fluids</strong> 51<br />

3.1 Thermodynamics of <strong>binary</strong> <strong>mixtures</strong> . . . . . . . . . . . . . . . . 52<br />

3.1.1 Spinodal decomposition . . . . . . . . . . . . . . . . . . 52<br />

3.1.2 Surface tension . . . . . . . . . . . . . . . . . . . . . . . 53<br />

3.2 Coarsening in a fluid at rest . . . . . . . . . . . . . . . . . . . . . 54<br />

3.2.1 Domain growth . . . . . . . . . . . . . . . . . . . . . . . 55<br />

3.3 Hydrodynamics <strong>and</strong> coarsening . . . . . . . . . . . . . . . . . . . 56<br />

3.3.1 Growth laws . . . . . . . . . . . . . . . . . . . . . . . . 57<br />

Part II 59<br />

4 Small-scale statistics of <strong>viscoelastic</strong> <strong>turbulence</strong> 61<br />

4.1 Uniaxial model . . . . . . . . . . . . . . . . . . . . . . . . . . . 62<br />

4.1.1 Numerical settings . . . . . . . . . . . . . . . . . . . . . 64<br />

4.2 Coil-stretch transition . . . . . . . . . . . . . . . . . . . . . . . . 65<br />

4.3 Energy cascade . . . . . . . . . . . . . . . . . . . . . . . . . . . 66<br />

4.4 Acceleration statistics . . . . . . . . . . . . . . . . . . . . . . . . 68<br />

4.4.1 Fluctuations of <strong>viscoelastic</strong> accelerations . . . . . . . . . 68<br />

4.4.2 Role of the back-reaction . . . . . . . . . . . . . . . . . . 69<br />

4.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71<br />

5 Elastic <strong>turbulence</strong> in two-dimensional flows 73<br />

5.1 Viscoelastic model . . . . . . . . . . . . . . . . . . . . . . . . . 74<br />

5.1.1 Elastic instabilities . . . . . . . . . . . . . . . . . . . . . 74<br />

5.1.2 Numerical setup . . . . . . . . . . . . . . . . . . . . . . 75<br />

5.2 Momentum budget . . . . . . . . . . . . . . . . . . . . . . . . . 76<br />

5.3 Transition to <strong>turbulence</strong> . . . . . . . . . . . . . . . . . . . . . . . 77<br />

5.3.1 Eulerian Lyapunov exponent . . . . . . . . . . . . . . . . 78<br />

5.3.2 Intermediate regime: elastic waves . . . . . . . . . . . . . 79<br />

5.3.3 Turbulent energy spectrum . . . . . . . . . . . . . . . . . 80<br />

5.4 Mixing: Lagrangian Lyapunov exponent . . . . . . . . . . . . . . 81<br />

5.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82<br />

6 Turbulence <strong>and</strong> coarsening in <strong>binary</strong> <strong>mixtures</strong> 85<br />

6.1 Phase separation under stirring . . . . . . . . . . . . . . . . . . . 86<br />

6.1.1 Evolution equations . . . . . . . . . . . . . . . . . . . . 86<br />

6.1.2 Numerical analysis . . . . . . . . . . . . . . . . . . . . . 87<br />

vi


CONTENTS vii<br />

6.2 Active <strong>mixtures</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . 88<br />

6.2.1 Unstirred case . . . . . . . . . . . . . . . . . . . . . . . . 88<br />

6.2.2 Stirred case . . . . . . . . . . . . . . . . . . . . . . . . . 88<br />

6.3 Passive <strong>mixtures</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . 91<br />

6.3.1 Role of Lagrangian chaos . . . . . . . . . . . . . . . . . 92<br />

6.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94<br />

Conclusions 95<br />

Bibliography 97<br />

Acknowledgements 103<br />

vii


Introduction<br />

This thesis presents a study on <strong>turbulence</strong> in <strong>viscoelastic</strong> <strong>fluids</strong> <strong>and</strong> <strong>binary</strong> <strong>mixtures</strong>.<br />

In both these systems, belonging to the general class of non-<strong>Newtonian</strong><br />

<strong>fluids</strong>, due to the presence of additives or the coexistence of different phases, the<br />

usual relation between the stress <strong>and</strong> velocity gradient tensors is no longer valid,<br />

giving rise to deep rheological consequences. Though physically different, these<br />

systems share common properties that can be described in terms of transported<br />

active fields.<br />

The major part of the work deals with the problem of <strong>turbulence</strong> in dilute<br />

polymer solutions (i. e. <strong>viscoelastic</strong> <strong>fluids</strong>). In such flows stability is controlled<br />

not only by inertial instabilities but by elastic ones as well. Two problems, corresponding<br />

to opposite limits in the space of stability parameters are considered:<br />

the study of small-scale statistics at moderate values of elasticity in a regime of<br />

fully developed <strong>turbulence</strong>, <strong>and</strong> the destabilization of a laminar flow by means of<br />

elastic nonlinearities.<br />

A key question in the study of <strong>turbulence</strong> of <strong>viscoelastic</strong> solutions is whether the<br />

turbulent cascade survives or not to the addition of polymers <strong>and</strong>, in case, how it is<br />

modified. This question is addressed through the analysis of small-scale statistics<br />

in a three-dimensional homogeneous isotropic configuration. By means of Direct<br />

Numerical Simulations (DNS) it is shown that indeed the turbulent cascade survives<br />

to polymer injection <strong>and</strong> the flow actually possesses non trivial small-scale<br />

features. Specifically, acceleration statistics is considered <strong>and</strong> it is shown how<br />

fluctuations of turbulent acceleration are affected by the presence of polymers.<br />

In the opposite limit of negligible inertial nonlinearities, polymeric elastic degrees<br />

of freedom can destabilize an originally laminar flow, provided elasticity is<br />

large enough. At growing elasticity, a transition to chaotic, <strong>and</strong> eventually turbulent,<br />

states is observed. This phenomenon is investigated in a two-dimensional<br />

flow with nonzero mean velocity. The phenomenology observed in experiments<br />

is reproduced numerically <strong>and</strong> the transition to chaos characterized by means of<br />

Eulerian <strong>and</strong> Lagrangian Lyapunov exponents, thus allowing the identification of<br />

turbulent-like features.<br />

The other subject considered in the thesis is the interplay between coarsening<br />

1


2 Introduction<br />

<strong>and</strong> <strong>turbulence</strong> in <strong>binary</strong> <strong>mixtures</strong>. Phase separation between two <strong>fluids</strong> in twodimensions<br />

is investigated, studying the phase ordering process in the presence<br />

of an external stirring acting on the velocity field. For both active <strong>and</strong> passive<br />

<strong>mixtures</strong> it is found that, for sufficiently strong stirring, coarsening is arrested in<br />

a stationary dynamical state characterized by a continuous rupture <strong>and</strong> formation<br />

of finite domains. Coarsening arrest is shown to be independent of the chaotic<br />

or regular nature of the flow, as it is a consequence of the competition between<br />

thermodynamic forces <strong>and</strong> stretching induced by local shears.<br />

The thesis is organized in six chapters, divided in two parts. Part I constitutes<br />

an introduction to the dynamics of <strong>Newtonian</strong> <strong>turbulence</strong>, polymer solutions <strong>and</strong><br />

phase separating <strong>binary</strong> <strong>mixtures</strong>; part II contains theoretical <strong>and</strong> numerical results<br />

on <strong>turbulence</strong> in <strong>viscoelastic</strong> <strong>and</strong> <strong>binary</strong> <strong>fluids</strong>.<br />

Part I<br />

The first chapter is intended as an introduction to fully developed <strong>turbulence</strong> in<br />

<strong>Newtonian</strong> <strong>fluids</strong>. The basic phenomenology is presented <strong>and</strong> the differences between<br />

the three <strong>and</strong> two-dimensional cases are stressed.<br />

In the second chapter <strong>viscoelastic</strong> <strong>fluids</strong> are introduced. After a brief description<br />

of polymers <strong>and</strong> their structural properties, their behaviour in a flow is considered.<br />

Starting from a "microscopic" point of view, corresponding to the dumbbell<br />

model, fully hydrodynamic models are constructed.<br />

The third chapter contains an introduction to the physics of <strong>binary</strong> <strong>mixtures</strong>.<br />

Coarsening in a fluid at rest is shortly reviewed, before presenting the analysis<br />

of the effect of an underlying velocity field on the process of phase separation.<br />

Part II<br />

The fourth chapter contains theoretical <strong>and</strong> numerical analysis of small-scale statistics<br />

in <strong>viscoelastic</strong> turbulent flows.<br />

The fifth chapter deals with elastic <strong>turbulence</strong>. The basic physical mechanisms<br />

of inertial <strong>and</strong> elastic instabilities are briefly described, <strong>and</strong> the numerical results<br />

reported.<br />

The sixth chapter reports a numerical analysis of the competition between stirring<br />

<strong>and</strong> coarsening in both active <strong>and</strong> passive symmetric <strong>binary</strong> <strong>mixtures</strong>, in two<br />

dimensions.<br />

2


The results presented in the thesis have been documented in the following<br />

publications:<br />

S. Berti, A. Bistagnino, G. Boffetta, A. Celani, S. Musacchio, Smallscale<br />

statistics of <strong>viscoelastic</strong> <strong>turbulence</strong>, Europhysics Letters 76, 63<br />

(2006). [chapter 4]<br />

S. Berti, A. Bistagnino, G. Boffetta, A. Celani, S. Musacchio, Elastic<br />

<strong>turbulence</strong> in 2D <strong>viscoelastic</strong> flows (submitted to the 11 th European<br />

Turbulence Conference). [chapter 5]<br />

S. Berti, G. Boffetta, M. Cencini, A. Vulpiani, Turbulence <strong>and</strong> coarsening<br />

in active <strong>and</strong> passive <strong>binary</strong> <strong>mixtures</strong>, Physical Review Letters<br />

95, 224501 (2005). [chapter 6]<br />

3<br />

3


Part I<br />

5


Chapter 1<br />

<strong>Newtonian</strong> <strong>turbulence</strong><br />

Fluid <strong>turbulence</strong> is a relevant problem for many different areas of scientific research<br />

such as meteorology, oceanography, astrophysics, engineering, chemistry.<br />

Most flows occurring in nature <strong>and</strong> engineering applications are turbulent: jets<br />

in atmosphere <strong>and</strong> oceanic currents, the photosphere of stars <strong>and</strong> interstellar gas<br />

clouds, boundary layers growing on aircraft wings, flows involved in combustion<br />

processes. Turbulent motion is indeed closer to being a typical behaviour in fluid<br />

dynamics, rather than an exception.<br />

From a fundamental point of view, a complete self-consistent theory of <strong>turbulence</strong><br />

is still lacking, though the equations governing it have been known since<br />

more than 180 years. The first scientific studies date back to the scripts <strong>and</strong> drawings<br />

of Leonardo da Vinci. In the nineteenth century the works of L. M. H. Navier,<br />

G. G. Stokes, O. Reynolds have laid the basis of a research field which is still open.<br />

This chapter presents an introduction to the basic concepts <strong>and</strong> phenomenology<br />

of <strong>turbulence</strong> in <strong>Newtonian</strong> <strong>fluids</strong>, both in the three <strong>and</strong> two-dimensional case,<br />

in a homogeneous isotropic configuration, except when explicitly said. The aim<br />

is not to provide a review of the subject, which can be found in [1, 2, 3, 4, 5], but<br />

just to introduce concepts <strong>and</strong> terminology that will be used in the thesis.<br />

7


8 1. <strong>Newtonian</strong> <strong>turbulence</strong><br />

1.1 Navier-Stokes equation<br />

The dynamics of an incompressible viscous fluid is described by the Navier-<br />

Stokes (1823) equation for the velocity field u(x, t), supplemented by the incompressibility<br />

condition:<br />

∂tu + u · ∇u = − 1<br />

∇P + ν∆u + f (1.1)<br />

ρ<br />

∇ · u = 0 (1.2)<br />

where P is the pressure, ρ is the density of the fluid, ν = µ/ρ its kinematic<br />

viscosity <strong>and</strong> f the resultant per unit mass of the external forces sustaining the<br />

motion.<br />

Let us briefly inspect the different terms appearing in Navier-Stokes equation:<br />

• u · ∇u is the inertial, nonlinear, term responsible for the transfer of kinetic<br />

energy in the turbulent cascade.<br />

• −∇P are the pressure gradients, ensuring incompressibility of the flow. In<br />

ρ<br />

absence of external forces, they are determined by the Poisson equation<br />

∆P = −ρ∂i∂juiuj<br />

obtained taking the divergence of eq. (1.1).<br />

(1.3)<br />

• ν∆u is the viscous dissipative term originated by the Reynolds stresses.<br />

This is the dominant term in the laminar regime.<br />

It is easy to underst<strong>and</strong> the physical meaning of eqs. (1.1)-(1.2); these are nothing<br />

else than conservation of momentum <strong>and</strong> mass per unit volume, respectively:<br />

dui<br />

dt<br />

1 ∂Tij<br />

=<br />

ρ ∂xj<br />

+ fi<br />

(1.4)<br />

∂ρ<br />

+ ∇ · (ρu) = 0 (1.5)<br />

∂t<br />

where T is the stress tensor. For a <strong>Newtonian</strong> fluid this is linearly dependent on<br />

the deformation tensor eij = 1<br />

2 (∂jui + ∂iuj) <strong>and</strong> is given by [6]<br />

Tij = −Pδij + µ(∂jui + ∂iuj − 2<br />

3 δij∂kuk) + ζδij∂kuk<br />

where the viscosity coefficients µ <strong>and</strong> ζ are positive functions of pressure <strong>and</strong><br />

temperature that will be assumed to be constant in the following. Let us observe<br />

8


1.1. Navier-Stokes equation 9<br />

that the expression of T is obtained from that of the ideal case, just adding to the<br />

pressure term a contribution accounting for the irreversible transfer of momentum<br />

due to processes of internal friction, whose explicit expression can be derived by<br />

simmetry considerations.<br />

Finally, for an incompressible flow, the stress tensor becomes:<br />

Tij = −Pδij + µ(∂jui + ∂iuj). (1.6)<br />

The incompressibility assumption is consistent as long as typical velocities u<br />

are smaller than the speed of sound cs in the fluid. Since eventual density fluctuations<br />

are swept away as sound waves, for small values of the Mach number<br />

M ≡ u/cs, the density can be considered constant in space <strong>and</strong> time ρ(x, t) = ρ<br />

<strong>and</strong> mass conservation (1.5) leads to the divergence-less condition (1.2) for the<br />

velocity field. It is common to assume the constant density equal to unity or,<br />

equivalently, to consider dynamical quantities per unit mass of fluid. For instance,<br />

we will often refer to the square modulus of the velocity as kinetic energy.<br />

1.1.1 Reynolds number<br />

Because of its nonlinearity, Navier-Stokes equation typically displays a sensitive<br />

dependence on initial conditions, i.e. chaotic behaviour. The degree of nonlinearity<br />

can be quantified through the Reynolds number, defined as<br />

Re ≡ UL<br />

ν<br />

(1.7)<br />

where U <strong>and</strong> L are, respectively, characteristic velocity <strong>and</strong> lenght scales; e.g.,<br />

in a pipe flow, U is the mean velocity <strong>and</strong> L is the diameter of the pipe. It was<br />

introduced by Osborne Reynolds, who showed that a transition from laminar to<br />

turbulent flow occurs when this number exceeds a critical value, depending on the<br />

geometry. For a fixed geometrical configuration, the critical value is universal,<br />

meaning that Re is the only adimensional parameter controlling the stability of a<br />

viscous flow. By a simple dimensional analysis, it can be seen that the Reynolds<br />

number quantifies the relative weight of nonlinear to linear term in eq. (1.1):<br />

u · ∇u<br />

ν∆u<br />

UL<br />

∼ . (1.8)<br />

ν<br />

Fully developed <strong>turbulence</strong> is achieved in the limit Re → ∞, corresponding by<br />

definition of Re to the zero-viscosity limit ν → 0. In such a case there is clearly no<br />

point in looking for a solution of Navier-Stokes equation but instead a statistical<br />

approach will be required for the investigation of <strong>turbulence</strong>.<br />

9


10 1. <strong>Newtonian</strong> <strong>turbulence</strong><br />

1.1.2 Energy balance<br />

The energy balance in absence of external forcing for Navier-Stokes equation follows<br />

from eqs. (1.1)-(1.2). The total kinetic energy of the fluid is<br />

E = 1<br />

�<br />

d<br />

2<br />

3 x|u| 2<br />

(1.9)<br />

<strong>and</strong> its time derivative is<br />

dE<br />

dt =<br />

�<br />

d 3 �<br />

x[ui∂tui + uiuj∂jui] =<br />

d 3 x[−ui∂iP + νui∂j∂jui]. (1.10)<br />

Assuming periodic boundary conditions on a cubic volume of size L<br />

u(x + nL, y + mL, z + qL) = u(x, y, z) ∀x, y, z ∈ R ∀n, m, q ∈ Z (1.11)<br />

or null boundary conditions on a volume V<br />

u|∂V = 0 (1.12)<br />

the nonlinear <strong>and</strong> pressure gradient contributions vanish; using the identity<br />

|∇ × u| 2 = (ɛijk∂juk)(ɛilm∂lum) = ∂j(uk∂juk) − ∂juk∂kuj − uk∂j∂juk (1.13)<br />

one finally gets<br />

�<br />

dE<br />

= ν<br />

dt<br />

d 3 �<br />

xu · ∆u = −ν<br />

d 3 x|∇ × u| 2 = −ν<br />

�<br />

d 3 x|ω| 2<br />

(1.14)<br />

where we have introduced the vorticity of the fluid ω = ∇ ×u. Defining the total<br />

enstrophy as<br />

Z = 1<br />

�<br />

d<br />

2<br />

3 x|∇ × u| 2<br />

(1.15)<br />

the energy balance reads:<br />

dE<br />

= −2νZ<br />

dt<br />

(1.16)<br />

<strong>and</strong> shows that in absence of external forcing <strong>and</strong> for ν = 0 energy is conserved,<br />

i.e. it is an inviscid invariant. On the other h<strong>and</strong>, in the limit ν → 0 the energy<br />

dissipation rate does not vanish, but approaches a constant value:<br />

lim 2νZ = ɛ. (1.17)<br />

ν→0<br />

This phenomenon is known as dissipative anomaly <strong>and</strong> represents one of the fundamental<br />

laws of <strong>turbulence</strong>. It implies that for ν → 0 the enstrophy must grow as<br />

Z ∼ ν −1 to compensate the decreasing viscosity. The unbounded growth of enstrophy<br />

in three dimensions is the physical origin of the dissipative anomaly, <strong>and</strong><br />

it is caused by the vortex stretching which produces diverging velocity gradients<br />

in the limit Re → ∞.<br />

10


1.1. Navier-Stokes equation 11<br />

1.1.3 Energy transfer<br />

In the energy balance eq. (1.16) the nonlinear term of Navier-Stokes equation<br />

is not involved, meaning it cannot change the global kinetic energy of the fluid.<br />

Nevertheless, this term has a major role in <strong>turbulence</strong>, since it is responsible for<br />

the energy transfer between different modes at the origin of the turbulent cascade.<br />

To describe its action, let us consider the energy balance in Fourier space. For the<br />

sake of simplicity we will consider the infinite volume limit, in which the fluid is<br />

assumed to fill the entire space. The Fourier transform reads<br />

<strong>and</strong> its inverse is<br />

ûα(k) = 1<br />

(2π) 3<br />

�<br />

uα(x) =<br />

�<br />

d 3 xe −ik·x uα(x) (1.18)<br />

d 3 ke ik·x ûα(k) (1.19)<br />

The reality condition on the velocity field u ∗ α(x) = uα(x) implies û ∗ α(k) =<br />

ûα(−k) in Fourier space, <strong>and</strong> the derivatives become multiplicative operators<br />

(∇ → ik), so that incompressibility is now expressed by k · u = 0. In Fourier<br />

space Navier-Stokes equation has the form:<br />

�<br />

∂ûα(k)<br />

= −i d<br />

∂t<br />

3 p(kβ − pβ)ûβ(p)ûα(k − p) +<br />

+i kα<br />

k2 �<br />

d 3 ppγ(kβ − pβ)ûβ(p)ûγ(k − p) +<br />

−νk 2 ûα(k) (1.20)<br />

where k = |k| <strong>and</strong> it is still possible to recognize the inertial term, the pressure<br />

term <strong>and</strong> the dissipative term, whereas the forcing has been omitted.<br />

Using incompressibility <strong>and</strong> the simmetry of the integrals for (p, k − p) →<br />

(p − k, p) it is possible to rewrite eq. (1.20) as<br />

� �<br />

∂<br />

+ νk2 ûα(k) = −i<br />

∂t<br />

Introducing the tensors<br />

�<br />

kβδαγ − kαkβkγ<br />

k 2<br />

Navier-Stokes equation becomes<br />

� �<br />

∂<br />

+ νk2 ûα(k) = −<br />

∂t i<br />

2 Pαβγ(k)<br />

�<br />

� �<br />

d 3 pûβ(p)ûγ(k − p) (1.21)<br />

Pαβ(k) ≡ δαβ − kαkβ<br />

k2 (1.22)<br />

Pαβγ(k) ≡ kβPαγ(k) + kγPαβ(k) (1.23)<br />

11<br />

d 3 pûβ(p)ûγ(k − p) (1.24)


12 1. <strong>Newtonian</strong> <strong>turbulence</strong><br />

Let us now introduce some notations. The two-point correlation function is defined<br />

as<br />

Qαβ(r) ≡ 〈ûα(x)ûβ(x + r)〉 (1.25)<br />

where 〈...〉 st<strong>and</strong>s for the average over the volume V 〈f(x)〉 = 1<br />

�<br />

V V d3xf(x). Its<br />

Fourier transform is<br />

Sαβ(k) = 1<br />

(2π3 �<br />

)<br />

<strong>and</strong> the correlation function in Fourier space reads<br />

Under isotropy the tensor Sαβ takes the form<br />

d 3 re −ik·r Qαβ(r) (1.26)<br />

〈ûα(k)ûβ(k ′ )〉 = δ(k + k ′ )Sαβ(k) (1.27)<br />

Sαβ(k) = A(k)kαkβ + B(k)δαβ<br />

(1.28)<br />

where A <strong>and</strong> B are functions of the wavevector modulus k. Multiplying eq. (1.27)<br />

by kβ <strong>and</strong> using incompressibility one gets B(k) = −k 2 A(k); substituting this in<br />

eq. (1.28) brings<br />

Sαβ(k) = Pαβ(k)B(k) (1.29)<br />

The energy spectrum is defined as the integral of the square modulus of velocity<br />

over a shell with fixed radius k:<br />

E(k) = 1<br />

�<br />

k<br />

2<br />

2 |û(k)| 2 dΩk<br />

(1.30)<br />

<strong>and</strong> the total kinetic energy is its integral E = � ∞<br />

0 dkE(k). By definition Sαα(k) =<br />

〈|û(k)| 2 〉; from eq. (1.29) we have Sαα(k) = 2B(k), so that the following relation<br />

holds:<br />

B(k) = 1<br />

(1.31)<br />

4πk 2E(k)<br />

which allows to derive the relation between the energy spectrum <strong>and</strong> the Fourier<br />

transform of the two-point correlation function:<br />

Sαβ(k) = E(k)<br />

4πk 2 Pαβ(k) (1.32)<br />

The evolution equation for the transorm of the correlation function is obtained<br />

from Navier-Stokes equation <strong>and</strong> reads<br />

� �<br />

∂<br />

+ νk2 Sαβ(k) = −<br />

∂t i<br />

2 Pαµν(k)<br />

�<br />

d 3 pTβµν(−k, p)<br />

− i<br />

2 Pβµν(−k)<br />

�<br />

d 3 pTαµν(k, p) (1.33)<br />

12


1.1. Navier-Stokes equation 13<br />

where the three point correlation function<br />

〈ûα(k)ûβ(k ′ )ûγ(k ′′ )〉 = δ(k + k ′ + k ′′ )Tαβγ(k, k ′ ) (1.34)<br />

has been introduced.<br />

The energy balance in Fourier space is obtained from eq. (1.33) remembering<br />

the relation (1.32) between the energy spectrum <strong>and</strong> the two-point correlation<br />

function. Using the antisimmetry Pαβγ(−k) = −Pαβγ(k) <strong>and</strong> the reality condition<br />

Tαβγ(−k, p) = T ∗ αβγ (k, −p), one finally gets<br />

� �<br />

∂<br />

+ 2νk2 E(k) = T(k) (1.35)<br />

∂t<br />

where the energy transfer T(k) has the expression:<br />

T(k) = −4πk 2 ��<br />

kµIm d 3 �<br />

pTαµα(k, p)<br />

Defining the enstrophy spectrum as:<br />

Z(k) = 1<br />

�<br />

2<br />

(1.36)<br />

k 2 |ˆω(k)| 2 dΩk = k 2 E(k) (1.37)<br />

<strong>and</strong> restoring the external force f in Navier-Stokes equation, the energy balance<br />

can be rewritten as<br />

∂tE(k) = −2νZ(k) + T(K) + F(k) (1.38)<br />

where F(k) is the injection energy spectrum:<br />

�<br />

F(k) = k 2 û(k) · ˆ f(k)dΩk<br />

(1.39)<br />

In the case of large scale forcing with correlation length L, the injection spectrum<br />

dominates the energy balance at small wavenumbers k ∼ kf ∼ 1/L, whereas<br />

large wavenumbers are characterized by strong contributions from viscous dissipation<br />

2νZ(k) = 2νk 2 E(k). In the intermediate wavenumber range, where both<br />

energy input <strong>and</strong> output are negligible, the dominant term in eq. (1.38) is the energy<br />

transfer T(k). In this inertial range energy is conserved <strong>and</strong> transferred by<br />

triadic interactions between modes with wavenumbers such that k + k ′ + k ′′ � 0.<br />

Finally, let us observe that equation (1.24) for the velocity field involves the<br />

two-point correlation function, <strong>and</strong> equation (1.33) for the two-point correlation<br />

function requires the three-point one. Clearly, the presence of a quadratic term in<br />

Navier-Stokes equation reproduces this problem at every order, i.e. the evolution<br />

13


14 1. <strong>Newtonian</strong> <strong>turbulence</strong><br />

of the correlation function of order n involves that of order n + 1; it will be then<br />

necessary to make assumptions on the statistics of the velocity field in order to<br />

close the equations of motion for correlation functions. Despite the importance of<br />

the problem, instead of concentrating on the study of closure approximations, in<br />

the following we will focus on scaling properties which can be inferred by means<br />

of phenomenological arguments.<br />

Let us end this section introducing an important length scale that can be formed<br />

from the correlation function. If uL is the longitudinal (parallel to the separation<br />

vector r) velocity component, then the longitudinal correlation coefficient is defined<br />

as:<br />

f(r) ≡ 〈uL(x)uL(x + r)〉<br />

u2 (1.40)<br />

rms<br />

where < u2 i >= u2rms ∀i, due to isotropy. Suppose that we exp<strong>and</strong> f(r) about<br />

r = 0. Then, recalling that f must be a symmetric function of r, we can write<br />

where we defined the Taylor scale λ this way:<br />

f(r) = 1 − r2<br />

2λ 2 + O(r4 ) (1.41)<br />

1<br />

λ 2 ≡ −f ′′ (0) (1.42)<br />

that is, by fitting a parabola to the longitudinal correlation function for small values<br />

of r.<br />

It is possible to show that the Taylor scale can be equivalently defined as<br />

1<br />

λ2 ≡ 〈(∂xux) 2 〉<br />

〈u2 x〉<br />

(1.43)<br />

Under isotropy we have: u 2 rms = 2E/3 <strong>and</strong> ɛ = 15ν < (∂xux) 2 >, so that the<br />

definition (1.43) of λ implies the relation<br />

λ 2 = 15ν u2 rms<br />

ɛ<br />

= 5E<br />

Z<br />

(1.44)<br />

between the Taylor scale <strong>and</strong> global quantities such as the energy dissipation rate<br />

ɛ, the kinetic energy E <strong>and</strong> the enstrophy Z.<br />

With the new length λ, the Taylor-scale Reynolds number can be constructed:<br />

Rλ ≡ urmsλ<br />

ν = √ 15Re (1.45)<br />

which is commonly used for experimental data, because it is easier to measure<br />

than the integral scale Reynolds number defined in section 1.1.1.<br />

14


1.2. Phenomenology of <strong>turbulence</strong> 15<br />

1.2 Phenomenology of <strong>turbulence</strong><br />

In the previous section the role of the nonlinear term in Navier-Stokes equation<br />

was shown to be responsible for the production of small scales <strong>and</strong> to correspond<br />

to an energy tranfer in Fourier space between modes in the inertial range.<br />

A different, much simpler, approach bringing the same results is the phenomenological<br />

one. This substantially consits in dimensional arguments, able to catch<br />

some essential features of fully developed <strong>turbulence</strong>.<br />

1.2.1 Turbulent cascade<br />

The basic phenomenology of <strong>turbulence</strong> can be recovered by using the picture of<br />

the turbulent cascade, proposed by Richardson [7].<br />

Kinetic energy is supposed to be injected by an external forcing feeding the<br />

motion of large scale eddies. Fluid dynamics then deforms these large scale structures,<br />

until they eventually break into smaller eddies, <strong>and</strong> the process is repeated<br />

such that energy is tranferred to smaller <strong>and</strong> smaller structures. Finally, the smallest<br />

length scale is reached, where energy is dissipated in the form of heat by the<br />

action of viscosity. The whole process (sketched in fig. 1.1) of energy transfer<br />

from larger to smaller scale eddies is known as turbulent cascade. It is important<br />

to remember that eddies do not correspond to real vortices, but they are just<br />

a visual description of the triadic interactions between modes which have been<br />

formally derived in the previous section.<br />

Figure 1.1: Turbulent cascade à la Richardson.<br />

Let uℓ be the rms velocity at scale ℓ <strong>and</strong> τℓ ∼ ℓ/uℓ the corresponding eddy<br />

turnover time, i. e. the typical time for a structure of size ℓ to be significantly<br />

15


16 1. <strong>Newtonian</strong> <strong>turbulence</strong><br />

deformed due to relative motion. Thus τℓ is also the typical time to transfer energy<br />

from scales of order ℓ to smaller ones.<br />

Dimensional analysis of the different terms in Navier-Stokes equation allows<br />

to identify three different ranges of scales:<br />

• injective range corresponding to large scales of order L where the forcing<br />

pumps energy into the system;<br />

• inertial range where both injection <strong>and</strong> viscous dissipation are negligible<br />

so that energy simply cascades to small scales;<br />

• dissipative range corresponding to small scales of order η where viscous<br />

dissipation dominates <strong>and</strong> the cascade terminates.<br />

The existence of a statistically steady state for the turbulent flow requires a<br />

constant energy flux Π(ℓ) in the inertial range, i. e. a constant rate of energy transfer<br />

that must be equal to the energy dissipation rate ɛ:<br />

Π(ℓ) ∼ E(ℓ)<br />

τℓ<br />

∼ u3 ℓ<br />

ℓ<br />

= ɛ (1.46)<br />

The above relation determines the Kolmogorov scaling of characteristic velocities<br />

<strong>and</strong> times:<br />

uℓ ∼ ɛ 1/3 ℓ 1/3<br />

τℓ ∼ ɛ −1/3 ℓ 2/3<br />

(1.47)<br />

(1.48)<br />

In this inertial range, the eddy turnover time at scale ℓ is much smaller than<br />

the viscous time at the same scale τ (diss)<br />

ℓ ∼ ℓ2 /ν; hence the energy is conserved<br />

<strong>and</strong> transported to smaller scales.<br />

The border between the inertial <strong>and</strong> the dissipative range is identified by the<br />

Kolmogorov dissipative scale η, where τℓ ∼ τ (diss)<br />

ℓ :<br />

� � 3 1/4<br />

ν<br />

η ∼<br />

(1.49)<br />

ɛ<br />

Below the Kolmogorov scale dissipation dominates <strong>and</strong> the velocity field is<br />

smooth <strong>and</strong> differentiable.<br />

1.2.2 Kolmogorov K41 theory<br />

There is presently no fully deductive theory of <strong>turbulence</strong> explaining the basic experimental<br />

observations of finite energy dissipation rate <strong>and</strong> power law spectrum<br />

16


1.2. Phenomenology of <strong>turbulence</strong> 17<br />

of exponent −5/3, in the limit of very large Reynolds number. Nevertheless, it<br />

is possible to make assumptions compatible with these laws <strong>and</strong> leading to further<br />

predictions. In Kolmogorov 1941 (K41) theory, hypotheses are formulated<br />

as to reproduce the experimental behaviour of the second order structure function<br />

(i. e. of the spectrum).<br />

If δuℓ(x) ≡ [u(x + ℓ) − u(x)] · ℓ/|ℓ| is the longitudinal velocity increment,<br />

then the p-th moment of its distribution:<br />

is called longitudinal structure function of order p.<br />

Let us reformulate Kolmogorov’s hypotheses:<br />

Sp(ℓ) ≡ 〈(δuℓ) p 〉 (1.50)<br />

H1 in the limit Re → ∞, all the simmetries of Navier-Stokes equation, broken<br />

by the mechanisms producing the turbulent flow, are restored in a statistical<br />

sense at small scales (ℓ ≪ L) <strong>and</strong> far from the boundaries;<br />

H2 in the same limit, the turbulent flow is self-similar at small scales. A unique<br />

scaling exponent h ∈ R therefore exists, such that:<br />

δuλℓ(x) = λ h δuℓ(x) ∀λ ∈ R+, ∀x : |x| ≪ L (1.51)<br />

H3 again, in the same limit, the energy dissipation rate has a finite nonvanishing<br />

limit 0 < ɛ < ∞.<br />

Starting from the Karman-Howarth-Monin relation [2], Kolmogorov derived<br />

the following exact result:<br />

four-fifths law. In the limit of infinite Reynolds number the third order (longitudinal)<br />

structure function of homogeneous isotropic <strong>turbulence</strong>, evaluated for<br />

increments ℓ small compared to the integral scale, is given in terms of the mean<br />

energy dissipation rate ɛ by<br />

S3(ℓ) ≡ 〈(δuℓ) 3 〉 = − 4<br />

ɛℓ (1.52)<br />

5<br />

The four-fifths law allows to determine the value of the scaling exponent h. Under<br />

rescaling of the increment ℓ by a factor λ, thanks to hypothesis H2, 〈(δuℓ) 3 〉 in<br />

(1.52) is multiplied by a factor λ3h while −4 5<br />

ɛℓ is multiplied by a factor λ, which<br />

implies h = 1/3.<br />

If we now consider the structure function of arbitrary order p > 0, this scales<br />

as Sp(λℓ) = λ p/3 Sp(ℓ), so that Sp(ℓ) ∼ ℓ p/3 . Since the physical dimensions of<br />

(ɛℓ) p/3 are those of Sp, we have<br />

Sp(ℓ) = Cpɛ p/3 ℓ p/3<br />

17<br />

(1.53)


18 1. <strong>Newtonian</strong> <strong>turbulence</strong><br />

The dimensionless constants Cp are not known, except for the universal value<br />

C3 = −4/5.<br />

An interesting case corresponds to p = 2; the K41 prediction for the second<br />

order structure function then is<br />

S2(ℓ) ∼ ɛ 2/3 ℓ 2/3<br />

The above scaling implies the power law energy spectrum:<br />

E(k) = C2ɛ 2/3 k −5/3<br />

where C2 is called Kolmogorov constant.<br />

(1.54)<br />

(1.55)<br />

Figure 1.2: Turbulent spectra in the time domain for data from the S1 wind tunnel<br />

ONERA [9] (Rλ = 2720, left) <strong>and</strong> a low temperature helium gas flow between<br />

counter-rotating cylinders [10] (Rλ = 1200, right).<br />

Longitudinal structure functions are convenient also for an experimental approach.<br />

Indeed, longitudinal velocity increments can be obtained by means of<br />

st<strong>and</strong>ard techniques such as, e. g., hot wire anemometry. Let us suppose to have<br />

a velocity field u which can be decomposed in a mean flow U = (U, 0, 0) <strong>and</strong><br />

a turbulent fluctuating part u ′ = u − U whose intensity is assumed to be small<br />

compared to the mean flow 〈|u ′ | 2 〉 1/2 ≪ U. Due to the cooling produced by the<br />

flow, the resistance of a hot wire perpendicular to the mean flow, say parallel to<br />

the z direction, is reduced. From the measure of this reduction, it is possible to<br />

obtain the time series of the velocity integrated in the direction of the wire:<br />

uN = [(u ′ x + U) 2 + u ′2<br />

y ] 1/2 = U[1 + u′ x<br />

U + O(u′2 x<br />

U 2)]<br />

(1.56)<br />

where it has been supposed that amplitudes of fluctuations in the two directions<br />

perpendicular to the wire are of the same order u ′ x ∼ u′ y . Within Taylor hypothesis,<br />

18


1.2. Phenomenology of <strong>turbulence</strong> 19<br />

i. e. assuming that for a short time delay τ the turbulent velocity field is almost<br />

frozen <strong>and</strong> simply transported by the fast mean flow U, the time series u ′ x (t) can<br />

be reinterpreted as a space series<br />

u ′ x (x, t + τ) = u′ x (x − Uτ, t) (1.57)<br />

<strong>and</strong> the structure function can be obtained.<br />

Spectra such as (1.55) have been observed in a variety of physical situations,<br />

from experiments in tidal channels [8], to more recent measurements in wind<br />

tunnels [9] <strong>and</strong> in low temperature helium gas flows between counter-rotating<br />

cylinders [10] (see fig. 1.2).<br />

1.2.3 Intermittency<br />

The central hypothesis of Kolmogorov K41 theory is the self-similarity of the<br />

r<strong>and</strong>om velocity field at inertial range scales. Nevertheless, experimental turbulent<br />

signals, once high-pass filtered, typically reveal intermittent features: their activity<br />

is restricted to only a fraction of the time, which decreases with the time scale<br />

under consideration. In other words, the velocity field can be thought as a r<strong>and</strong>om<br />

alternation of quiet periods <strong>and</strong> chaotic bursts of intense fluctuations.<br />

These intermittent features are reflected in the shape of the probability distribution<br />

function (pdf) of velocity increments (see fig. 1.3). Experimental data<br />

[11, 12, 13] show that the pdf is essentially gaussian at large scales (ℓ0) but, as<br />

the scale gets smaller, it develops higher <strong>and</strong> higher tails accounting for larger <strong>and</strong><br />

larger probabilities of strong fluctuations. Near the Kolmogorov dissipative scale<br />

η, the pdf takes the shape of a stretched exponential.<br />

A convenient measure of intermittency is given by the flatness:<br />

F(ℓ) ≡ S4(ℓ)<br />

(S2(ℓ)) 2 = 〈(δuℓ) 4 〉<br />

〈(δuℓ) 2 〉 2<br />

(1.58)<br />

Since the fourth order moment receives higher contributions from the tails of the<br />

probability distribution function (pdf), the flatness gives a measure of the frequency<br />

of events larger than the st<strong>and</strong>ard deviation. While for an intermittent<br />

function the flatness grows with decreasing scale, either in the gaussian (F = 3)<br />

or in the self-similar case this is not true, being F independent of the scale ℓ.<br />

Measurements of high order structure functions [9], allowed to the detect the<br />

intermittent nature of the turbulent velocity field, both in the dissipative <strong>and</strong> the<br />

inertial range. The results suggest that structure functions follow power laws in<br />

the inertial range:<br />

Sp(ℓ) = 〈(δuℓ) p 〉 ∼ ℓ ζp (1.59)<br />

but the exponents ζp differ from the K41 prediction ζ K41<br />

p<br />

19<br />

= p/3.


20 1. <strong>Newtonian</strong> <strong>turbulence</strong><br />

Figure 1.3: Probability distributions, normalized with their st<strong>and</strong>ard deviations<br />

σr =< δu 2 r >1/2 , of transverse velocity increments in a turbulent air jet at Rλ =<br />

695 for different separations: r ∼ η (dotted line); r ∼ ℓ0 shifted of three decades<br />

(long dashed line); intermediate scales (solid line) [13].<br />

A phenomenological model of intermittency is the multifractal one, introduced<br />

by Parisi <strong>and</strong> Frisch [14]. In the multifractal approach [14, 15], instead of global<br />

scale invariance, local scale invariance is assumed; more specifically, hypothesis<br />

H2 is modified into:<br />

HMF in the limit Re → ∞, the turbulent flow possesses a range of scaling<br />

exponents I = [hmin, hmax]. For all h ∈ I, there is a set Sh ⊂ R3 of<br />

dimension D(h), such that for x ∈ I <strong>and</strong> ℓ → 0:<br />

� �h δuℓ(x) ℓ<br />

∼<br />

(1.60)<br />

u0<br />

where, of course, u0 is the large scale velocity.<br />

The structure function of order p is then expressed as a superposition, weighted<br />

with a measure dµ(h), of different contributions originated by different scaling<br />

exponents:<br />

Sp(ℓ)<br />

u p<br />

0<br />

�<br />

∼<br />

I<br />

ℓ0<br />

� �ph+3−D(h) ℓ<br />

dµ(h)<br />

ℓ0<br />

(1.61)<br />

where the factor (ℓ/ℓ0) 3−D(h) is the probability of being within a distance ℓ of the<br />

set Sh. The exponent ζp can be obtained in the limit ℓ → 0 by a saddle-point<br />

20


1.2. Phenomenology of <strong>turbulence</strong> 21<br />

argument, which gives the relation:<br />

ζp = inf<br />

h [ph − D(h) + 3] (1.62)<br />

in which one recognizes the Legendre transform of the fractal dimension D(h).<br />

1.2.4 Acceleration statistics<br />

The motion of test particles in a turbulent flow as they are pushed by fluctuating<br />

accelerations is a subject of interest for a wide variety of problems, ranging from<br />

cloud formation [16] to stirred chemical reactions <strong>and</strong> combustion processes [17].<br />

From a more fundamental point of view, the study of acceleration statistics is<br />

essential for transport <strong>and</strong> mixing in <strong>turbulence</strong>.<br />

Figure 1.4: Particle trajectory in a turbulent water flow, Rλ = 970. A sphere<br />

marks the measured position of the particle in each of 300 frames taken every<br />

0.014ms (≈ τη/20). The shading indicates the acceleration magnitude, with the<br />

maximum value of 12000ms −2 corresponding to approximately 30 st<strong>and</strong>ard deviations<br />

[18].<br />

A typical turbulent trajectory is shown in fig. 1.4. This is a three-dimensional<br />

measure, with extremely high resolution in time, of the trajectory of a tracer in<br />

a water flow [18], obtained using silicon strip detectors originally designed for<br />

high-energy physics experiments. This trajectory shows that tracers in a turbulent<br />

flow experience a highly irregular motion, characterized by violently fluctuating<br />

accelerations; the particle enters the detection volume from the upper right, is<br />

21


22 1. <strong>Newtonian</strong> <strong>turbulence</strong><br />

pushed to the left by a burst of acceleration <strong>and</strong> comes nearly to a stop before<br />

being rapidly accelerated upward by a fluctuation roughly equal to 30 times the<br />

root mean square value.<br />

The acceleration a of a fluid particle in a turbulent flow is given by the Navier-<br />

Stokes equation:<br />

a ≡ du<br />

= −∇p + ν∆u + f (1.63)<br />

dt<br />

Provided f is a large scale forcing <strong>and</strong> Re is large enough, the statistics of a<br />

is essentially determined by that of pressure gradients. Experimental data [18]<br />

indicate that the acceleration is an extremely intermittent variable <strong>and</strong> the shape<br />

of its pdf is a stretched exponential (fig. 1.5).<br />

Figure 1.5: Probability distribution functions of a normalized acceleration component<br />

at three Reynolds numbers. The solid line is a stretched exponential parameterization<br />

of the highest Rλ data, the inner dotted line is a gaussian for reference.<br />

Inset: flatness F = 〈a 4 〉/〈a 2 〉 2 as a function of Rλ [18].<br />

Let us now derive the shape of the acceleration pdf by means of simple phenomenological<br />

arguments. By definition<br />

δuτ<br />

a ≡ lim<br />

τ→0 τ<br />

� δuτη<br />

τη<br />

(1.64)<br />

where τη is the eddy turnover time associated with the Kolmogorov dissipative<br />

scale η. The velocity fluctuations along a particle trajectory may be considered as<br />

the superposition of different contributions from eddies of all sizes. In a time lag<br />

τ the contributions from eddies smaller than a given scale ℓ are uncorrelated <strong>and</strong><br />

one may then write δuτ ∼ δuℓ. We assume that ℓ <strong>and</strong> τ are linked by the typical<br />

22


1.2. Phenomenology of <strong>turbulence</strong> 23<br />

eddy turnover time at the given spatial scale, τℓ ∼ ℓ/δuℓ. Therefore, we can write<br />

2 (δuη)<br />

a �<br />

η<br />

(1.65)<br />

Imposing the scaling δuℓ = u0(ℓ/ℓ0) h , we have: η/ℓ0 = Re −1/(1+h) , <strong>and</strong> hence<br />

(Rλ ∼ Re 1/2 ) we get the estimate:<br />

a = u2 0<br />

ℓ0<br />

R −22h−1<br />

1+h<br />

λ<br />

(1.66)<br />

In the framework of K41 theory h = 1/3 <strong>and</strong> the only fluctuating quantity is<br />

the large scale velocity u0. The pdf of acceleration p(a) is then simply obtained<br />

by the change of variable p(a)da = p(u0)du0 <strong>and</strong> the assumption of gaussianity<br />

for the statistics of the large scale velocity. This gives the stretched exponential:<br />

p(a) ∼ a −5/9 e −a8/9<br />

(1.67)<br />

where the st<strong>and</strong>ard deviation of u0 was implicitly assumed to be unity.<br />

From (1.66) the acceleration variance is expressed in terms of the energy dissipation<br />

rate ɛ by:<br />

〈a 2 〉 = u20 Rλ ∼ ν<br />

ℓ0<br />

−1/2 ɛ 3/2<br />

(1.68)<br />

known in the literature as Heisenberg-Yaglom relation. This is experimentally verified<br />

in the limit of very large Reynolds numbers; at moderate values of Rλ there is<br />

experimental <strong>and</strong> numerical evidence of a weak dependence of the proportionality<br />

coefficient on ɛ.<br />

The estimate (1.67) only roughly reproduces the large tails of experimental<br />

acceleration probability distributions. This is not surprising, since this expression<br />

is derived in the context of K41 theory <strong>and</strong> thus cannot take into account intermittency.<br />

Recently, several attempts have been made to improve the agreement<br />

between theory <strong>and</strong> experiments for Lagrangian statistics, such as acceleration.<br />

Among these, a considerable number focuses on the construction of models based<br />

on generalized equilibrium statistics (see, e. g., [19, 20, 21, 22], critically reviewed<br />

in [23]).<br />

A possible alternative approach, rooted in the phenomenology of <strong>turbulence</strong>,<br />

is based on the multifractal model [24]. Let us briefly present its main issues. The<br />

starting point is expression (1.66) of the acceleration, that can be rewritten as<br />

a(h, u0) ∼ ν 2(h−1)/(1+h) u 3/(1+h)<br />

0 ℓ −3h/(1+h)<br />

0<br />

(1.69)<br />

In the multifractal framework, the scaling exponent h is allowed to fluctuate within<br />

the interval I = [hmin, hmax]. The pdf of a can be obtained integrating (1.69) over<br />

23


24 1. <strong>Newtonian</strong> <strong>turbulence</strong><br />

all h <strong>and</strong> u0, weighted with their respective probabilities, [τη(h, u0)/τℓ0(u0)] [3−D(h)]<br />

(1−h)<br />

<strong>and</strong> p(u0). The latter, as before, is reasonably approximated by a gaussian of st<strong>and</strong>ard<br />

deviation σ2 u = 〈u20 〉. Integration over u0 gives<br />

p(a) ∼<br />

h∈I<br />

�<br />

h∈I<br />

�<br />

× exp<br />

dha [h−5+D(h)]/3 ν [7−2h−2D(h)]/3 ℓ D(h)+h−3<br />

0 σ −1<br />

u<br />

− a2(1+h)/3ν2(1−2h)/3ℓ2h 0<br />

2σ2 u<br />

�<br />

(1.70)<br />

From (1.70) the dependence of the acceleration moments on Rλ can be derived.<br />

For example, in the limit of large Rλ the second order moment is given by 〈a2 〉 ∼<br />

R χ<br />

λ , where χ = suph{2[D(h) −4h −1]/(1+h)}. The precise value of χ depends<br />

on the choice of D(h), but it is typically close to the K41 value χK41 = 1. In terms<br />

of the dimensionless acceleration ã = a/σa, (1.70) becomes<br />

�<br />

p(ã) ∼ dhã [h−5+D(h)]/3 R y(h)<br />

�<br />

λ exp − 1<br />

2 ã2(1+h)/3R z(h)<br />

�<br />

λ (1.71)<br />

where y(h) = χ[h−5+D(h)]/6+2[2D(h)+2h−7]/3 <strong>and</strong> z(h) = χ(1+h)/3+<br />

4(2h − 1)/3. Let us observe that the K41 prediction (1.67) can be recovered from<br />

(1.71) with h = 1/3, D(h) = 3 <strong>and</strong> χ K41 = 1. Comparison with data from high<br />

resolution DNS, as reported in [24], shows that the multifractal prediction (1.71)<br />

captures the shape of the acceleration pdf much better than its K41 counterpart<br />

(see fig. 1.6).<br />

Figure 1.6: Log-linear plot of the acceleration pdf. The crosses are DNS data, the<br />

solid line is the multifractal prediction, <strong>and</strong> the dashed line is the K41 prediction.<br />

The DNS statistics was calculated along the trajectories of 2 × 10 6 Lagrangian<br />

particles amounting to 1.06 × 10 10 events in total. Inset: ã 4 p(ã) for DNS data<br />

(crosses) <strong>and</strong> the multifractal prediction [24].<br />

24


1.3. Two-dimensional <strong>turbulence</strong> 25<br />

1.3 Two-dimensional <strong>turbulence</strong><br />

Two-dimensional <strong>turbulence</strong> describes the behaviour of high-Reynolds-number<br />

solutions of Navier-Stokes equation which depend only on two cartesian coordinates<br />

(x, y). In this case, the third component of the velocity uz satisfies an<br />

advection-diffusion equation without back-reaction on the horizontal (x, y) flow.<br />

Hence, without loss of generality, one may assume that the velocity has only two<br />

components.<br />

There exist numerous situations, in natural flows <strong>and</strong> laboratory experiments,<br />

which are constrained to quasi-two-dimensional motion. The most important examples<br />

probably arise in geophysics <strong>and</strong> plasma physics. Indeed, the intermediatescale<br />

dynamics of the oceans <strong>and</strong> the atmosphere, due to the combined effect of<br />

their stratification <strong>and</strong> earth’s rotation, can be roughly described as being twodimensional.<br />

Similarly, a strong magnetic field can confine the turbulent motions<br />

of a plasma in the plane perpendicular to its axis <strong>and</strong> the dynamics can be described<br />

by two-dimensional magneto-hydrodynamics (2D MHD) [25].<br />

The classical theory of 2D <strong>turbulence</strong> originates from the works of of Batchelor,<br />

Kraichnan <strong>and</strong> Leith [26, 27, 28], where it was shown that the conservation<br />

of vorticity along streamlines, occurring in two dimensions, produces radical<br />

changes in the behaviour of <strong>turbulence</strong>.<br />

Finally, Navier-Stokes equation in two dimensions is less dem<strong>and</strong>ing on a<br />

computational level than the three dimensional case, thus allowing to reach relatively<br />

high Reynolds numbers in direct numerical simulations.<br />

1.3.1 Vorticity equation<br />

In two dimensions the incompressible velocity field u can be expressed in terms<br />

of the stream function ψ as:<br />

u = (∂yψ, −∂xψ) (1.72)<br />

The vorticity field, defined as the curl of velocity ω = ∇ × u, in two dimensions<br />

has only one nonzero component, normal to the plane of velocity, which is related<br />

to the stream function by<br />

ω = −∆ψ (1.73)<br />

Instead of describing the flow in terms of the two velocity components, which<br />

are not independent because of the incompressibility condition, it is convenient to<br />

work with the evolution equation for the scalar field ω, obtained taking the curl of<br />

the 2D Navier-Stokes equation. This reads:<br />

∂ω<br />

∂t<br />

+ u · ∇ω = ν∆ω − αω + fω<br />

(1.74)<br />

25


26 1. <strong>Newtonian</strong> <strong>turbulence</strong><br />

The linear dissipative term on the right-h<strong>and</strong> side of (1.74) accounts for friction<br />

between the thin layer of fluid which is considered <strong>and</strong> the rest of the three dimensional<br />

environment. The term fω represents an external forcing that counteracts<br />

dissipation by viscosity ν <strong>and</strong> friction α <strong>and</strong> allows to obtain a statistically steady<br />

state.<br />

To solve eq. (1.74) it is necessary to specify boundary conditions, which are<br />

required to solve the Poisson equation (1.73) for the stream function. In most<br />

studies on two-dimensional <strong>turbulence</strong> these are assumed to be of periodic type<br />

in both the directions. The presence of realistic no-slip boundaries gives rise to a<br />

source of vorticity fluctuations.<br />

1.3.2 Inviscid invariants<br />

The main difference between the two <strong>and</strong> three-dimensional problem is the conservation<br />

of vorticity along fluid trajectories in 2D, when viscosity, friction <strong>and</strong><br />

forcing are ignored.<br />

The origin of this phenomenon is due to the absence of the vortex-stretching<br />

term (ω ·∇)u in the two-dimensional vorticity equation. In three dimensions this<br />

term is responsible for the unbounded growth of enstrophy in the limit of infinite<br />

Reynolds number.<br />

In the inviscid limit ν = 0, with no forcing <strong>and</strong> friction, eq. (1.74) simply<br />

states that the material derivative of ω vanishes:<br />

Dω<br />

Dt<br />

= ∂ω<br />

∂t<br />

+ u · ∇ω = 0 (1.75)<br />

thus implying conservation of vorticity of fluid particles. Moreover, all the integrals<br />

of the form � d 2 rf(ω) are inviscid invariants of the flow. In particular, this<br />

yields conservation of the of the enstrophy<br />

Z = 1<br />

�<br />

2<br />

d 2 r|ω| 2<br />

If f = α = 0 in eq. (1.74), the enstrophy balance equation<br />

(1.76)<br />

dZ<br />

dt = −ν〈|∇ω|2 〉 = −2νP (1.77)<br />

states that Z is bounded since, by definition, the palinstrophy P is a positivedefinite<br />

quantity:<br />

P = 1<br />

2 〈|∇ω|2 〉 = 1<br />

�<br />

d<br />

2<br />

2 r|∇ω 2 | (1.78)<br />

26


1.3. Two-dimensional <strong>turbulence</strong> 27<br />

Therefore, at variance with the three-dimensional case, in two-dimensional <strong>turbulence</strong><br />

the viscous dissipation of energy vanishes in the limit ν → 0:<br />

dE<br />

lim<br />

ν→0 dt<br />

= lim(−2νZ)<br />

= 0 (1.79)<br />

ν→0<br />

but there still is dissipative anomaly for enstrophy (when friction is not considered).<br />

The complete balance equations are:<br />

dZ<br />

= −2νP − 2αZ + ζ (1.80)<br />

dt<br />

dE<br />

= −2νZ − 2αE + ɛ (1.81)<br />

dt<br />

where ɛ <strong>and</strong> ζ are, respectively, energy <strong>and</strong> enstrophy input terms. From the above<br />

equations is clearly seen that, in the limit of vanishing viscosity, the friction term<br />

(−2αE) is responsible for the existence of a stationary state; indeed was it α = 0<br />

in eq. (1.81), energy would grow indefinitely. Moreover, in the same limit ν → 0,<br />

it is known [29] that the presence of friction causes the enstropy dissipation to<br />

vanish as well.<br />

1.3.3 Direct <strong>and</strong> inverse cascades<br />

Since viscous energy dissipation vanishes in the limit Re → ∞, in fully developed<br />

two-dimensional <strong>turbulence</strong> it is not possible to have a cascade of energy with<br />

constant flux towards small scales as in three dimensions. Indeed, the presence<br />

of two quadratic inviscid invariants, energy <strong>and</strong> enstrophy, modifies the picture of<br />

the turbulent cascade.<br />

In order for both energy <strong>and</strong> enstrophy to be conserved, the net transfer by<br />

each triad interaction must be out of the middle wavenumber into both smaller<br />

<strong>and</strong> larger wavenumbers. Starting from the hint that interactions should act towards<br />

producing equilibrium, a state which is never reached because of viscous<br />

dissipation, Kraichnan showed that in two-dimensional <strong>turbulence</strong> enstrophy is<br />

mainly transferred to small scales, where it is dissipated by viscosity, giving rise<br />

to the direct enstrophy cascade; on the contrary, energy is transported to lower<br />

wavenumbers, in the inverse energy cascade (see fig. 1.7).<br />

The scaling laws in both cascades can be obtained from dimensional analysis<br />

of Navier-Stokes equation as well as in the three dimensional case.<br />

Inverse energy cascade<br />

For the inverse energy cascade, the assumption of constant energy flux Π(ℓ) = −ɛ<br />

towards large scales reproduces 3D-like scaling laws for velocities <strong>and</strong> character-<br />

27


28 1. <strong>Newtonian</strong> <strong>turbulence</strong><br />

Figure 1.7: Schematic double cascading spectrum of forced (wavenumber kF )<br />

two-dimensional <strong>turbulence</strong>.<br />

istic times:<br />

uℓ ∼ ɛ 1/3 ℓ 1/3<br />

τℓ ∼ ɛ −1/3 ℓ 2/3<br />

(1.82)<br />

(1.83)<br />

This means that the velocity field in the inverse cascade is rough, with scaling<br />

exponent h = 1/3, exactly as in the three dimensional case. The prediction for<br />

the energy spectrum is:<br />

E(k) = Cɛ 2/3 k −5/3<br />

(1.84)<br />

The hypothesis of locality of interactions in the inverse cascade is consistent<br />

with the k −5/3 spectrum. The transfer is associated with the distortion of the<br />

velocity field by its own shear. The effective shear at a given wavenumber k<br />

is expected to be negligibly affected by wavenumbers ≪ k because the integral<br />

� ∞<br />

0 k2 E(k)dk, which measures the mean square shear, converges at k = 0. Also<br />

the contribution from high wavenumbers ≫ k is negligible because vorticity associated<br />

with those wavenumbers fluctuates rapidly in space <strong>and</strong> time <strong>and</strong> gives a<br />

small effective shear across distances of order k −1 .<br />

In the absence of a large scale sink of energy, the inverse cascade can only be<br />

quasy-steady because the peak kE of the energy spectrum keeps moving down to<br />

ever lower wavenumbers as<br />

kE ∼ ɛ −1/2 t −3/2<br />

(1.85)<br />

while the total energy grows linearly in time as E(t) = ɛt. If the energy input is<br />

turned on for a sufficiently long time, the cascade can eventually reach the integral<br />

28


1.3. Two-dimensional <strong>turbulence</strong> 29<br />

scale <strong>and</strong> begins to accumulate at the lowest wavenumber. This pile-up of energy<br />

can produce a large scale spectrum steeper than k −3 , which violates the hypothesis<br />

of locality of interactions.<br />

The presence of friction stops the energy cascade at the wavenumber<br />

kE ∼ ɛ −1/2 α 3/2<br />

(1.86)<br />

where the energy dissipation balances the energy transfer.<br />

In two-dimensional <strong>turbulence</strong> it is possible to demonstrate the analogous of<br />

Kolmogorov’s four-fifths law. In the limit of infinite Reynolds number the third<br />

order (longitudinal) structure function of two-dimensional homogeneous isotropic<br />

<strong>turbulence</strong>, evaluated for increments ℓ small compared to the integral scale <strong>and</strong><br />

larger than the forcing correlation length, is given in terms of the mean energy<br />

flux ɛ by<br />

S3(ℓ) ≡ 〈(δuℓ) 3 〉 = 3<br />

ɛℓ (1.87)<br />

Together with the scaling hypothesis for the structure functions Sp(ℓ) ∼ ℓ ζp , the<br />

three-halves law allows to obtain the equivalent of K41 theory for the inverse<br />

energy cascade.<br />

At variance with the three-dimensional case, where intermittency modifies the<br />

dimensionally predicted scaling exponents, the statistics of velocity fluctuations<br />

in the inverse-cascade range of scales is found to be roughly self-similar [30], with<br />

small deviations from gaussianity.<br />

Direct enstrophy cascade<br />

At scales smaller than the forcing correlation length, the hypothesis of constant<br />

enstrophy flux Πω(ℓ) = ζ leads to a different scaling. The enstrophy contained in<br />

the eddies of size ℓ can be estimated as Z(ℓ) ∼ E(ℓ)/ℓ 2 ∼ u 2 ℓ /ℓ2 <strong>and</strong> its flux is<br />

so that velocities scale as<br />

Πω(ℓ) ∼ Z(ℓ)<br />

τℓ<br />

2<br />

∼ u3 ℓ<br />

ℓ 3<br />

(1.88)<br />

uℓ ∼ ζ 1/3 ℓ (1.89)<br />

Therefore the velocity field is smooth in the direct cascade. The dimensional<br />

prediction for characteristic times gives a single time scale τ ∼ ζ −1/3 , which<br />

provides an estimate of the inverse of the Lyapunov exponent of the flow. The<br />

prediction for the energy spectrum is<br />

E(k) = C ′ ζ 2/3 k −3<br />

(1.90)<br />

A spectrum k −3 means that the integral � ∞<br />

0 k2 E(k)dk, a measure of the mean<br />

square shear, has a logarithmic divergence in the infrared cutoff. Thus the hypothesis<br />

of locality of interactions can be violated in the direct enstrophy cascade.<br />

29


Chapter 2<br />

Polymer solutions<br />

It is known since the 1940’s that the addition of small amounts of polymers to<br />

a fluid can produce dramatic changes of its properties. Due to their molecular<br />

structure, polymers have elastic degrees of freedom which must be taken into account<br />

in the description of the mechanical response of the fluid to which they are<br />

added. Indeed, while for a <strong>Newtonian</strong> fluid the stress is proportional to the deformation<br />

rate, the coefficient being the viscosity, in an elastic material the stress<br />

is proportional, via the Hook modulus, to the deformation itself. A solution of a<br />

<strong>Newtonian</strong> fluid <strong>and</strong> polymers can be thought of as a mixture of these idealized<br />

situations, because it presents features of both viscous <strong>and</strong> elastic materials, <strong>and</strong><br />

it is thus called a <strong>viscoelastic</strong> fluid.<br />

The above behaviour manifests itself in a variety of spectacular phenomena.<br />

Among these, to have an idea of the different response of <strong>Newtonian</strong> <strong>and</strong> viscoealstic<br />

<strong>fluids</strong>, let us briefly consider the Weissenberg effect [31]. When a <strong>Newtonian</strong><br />

fluid is put in rotation, it is pushed away from the center by the centrifugal force<br />

<strong>and</strong> a dip appears on the free surface, which takes the shape of a paraboloid. On<br />

the contrary, when a rotating rod is inserted in a tank filled with a <strong>viscoelastic</strong><br />

fluid, the fluid tends to climb up the rod, as in fig. 2.1.<br />

The effects produced by polymer additives in <strong>fluids</strong> range from the change<br />

of flow stability <strong>and</strong> transition to <strong>turbulence</strong>, to the modification of mixing properties<br />

<strong>and</strong> of transport of heat, mass <strong>and</strong> momentum. The vast bibliography by<br />

Nadolink <strong>and</strong> Haigh [32] gives an idea of the efforts devoted to the underst<strong>and</strong>ing<br />

of <strong>viscoelastic</strong> <strong>fluids</strong>’ behaviour, which is clearly of interest both from a purely<br />

theoretical <strong>and</strong> a more application-oriented point of view.<br />

In this chapter the basic dynamics of polymers in <strong>fluids</strong> is introduced. Starting<br />

from a "microscopic" description in terms of single polymer dynamics, full hydrodynamic<br />

models that are typically used to describe <strong>viscoelastic</strong> solutions will<br />

be constructed. Finally, the phenomena of drag reduction <strong>and</strong> elastic <strong>turbulence</strong><br />

will be briefly introduced, as they will be discussed in further detail, or referred<br />

31


32 2. Polymer solutions<br />

Figure 2.1: Rod-climbing or Weissenberg effect (picture from web.mit.edu).<br />

to, in the following chapters.<br />

2.1 Polymer dynamics in <strong>fluids</strong><br />

Polymers are molecules of high molecular mass, composed by a large number of<br />

repeating units, the monomers, joined by chemical bonds to form a long chain.<br />

There exist both natural <strong>and</strong> synthetic polymers. Examples of the first type are<br />

DNA, proteins, starches, cellulose, latex. Synthetic polymers are commercially<br />

produced for a huge variety of uses; among these, we find all the materials commonly<br />

called plastics.<br />

In what follows, simple models describing polymer dynamics will be developed.<br />

The essential characteristics of polymers we focus on are:<br />

• long polymers: the degree of polymerization M, i. e. the number of monomers,<br />

is very large, typically M ∼ 10 6 .<br />

• flexible polymers: the angle formed by a bond connecting two monomers<br />

is fixed. This means that at the scale of the monomer the polymer is rigid.<br />

However, if one looks at the polymer at a scale ℓp, called persistence scale,<br />

one sees a flexible coil. When ℓp is much smaller than the total length of the<br />

polymer the molecule is said to be highly flexible.<br />

• homopolymers: only a single type of monomer is present.<br />

• no branching: we will consider only single chain molecules, though there<br />

exist polymers formed by several branches.<br />

32


2.1. Polymer dynamics in <strong>fluids</strong> 33<br />

Let us now briefly state the essential characteristics that models are required<br />

to reproduce. Because of thermal agitation, at equilibrium the molecule assumes<br />

the aspect of a statistically spherical coil, as in fig. 2.2(a), whose average radius is<br />

typically of the order of 0.1µm. Once elongated, polymers asymptotically relax<br />

to the equilibrium configuration on a time scale τ. For relatively small extensions<br />

polymers counteract elongation with a recalling force proportional to the extension<br />

itself.<br />

(a) (b)<br />

Figure 2.2: (a) Sketch of the coiled equilibrium configuration of a polymer. (b) A<br />

freely jointed chain.<br />

2.1.1 Freely jointed chain<br />

The simplest description of a polymer is that of the freely jointed chain [fig. 2.2(b)],<br />

in which the molecule is approximated by a chain of M segments of length b with<br />

r<strong>and</strong>om indipendent relative orientations. The different nodes of the chain are<br />

labelled by a set of vectors Pn, with n = 1, ..., M, <strong>and</strong> the bond vectors are:<br />

The characteristic size of the chain is :<br />

rn = Pn − Pn−1<br />

〈|R| 2 〉 1/2 = 〈|<br />

M�<br />

n=1<br />

rn| 2 〉 1/2<br />

(2.1)<br />

(2.2)<br />

which, thanks to the r<strong>and</strong>om independent relative orientations of the bonds, gives:<br />

〈|R| 2 〉 1/2 = M 1/2 b (2.3)<br />

33


34 2. Polymer solutions<br />

Introducing the maximum extension (contour length) Rmax = Mb of the chain,<br />

we have:<br />

〈|R| 2 〉 = Rmaxb (2.4)<br />

A quantity often used to characterize polymer elongation at rest is the radius<br />

of gyration R0:<br />

R0 ≡ 〈 1<br />

2M2 M� M�<br />

|Pn − Pm| 2 〉 (2.5)<br />

n=1 m=1<br />

which is the square mean distance between two r<strong>and</strong>omly selected nodes. For<br />

flexible polymers, it is possible to show that the following scaling holds:<br />

R 2 0 = 〈|R|2 〉<br />

6<br />

= Mb2<br />

6<br />

(2.6)<br />

This model is equivalent to a diffusion process: a particle moving on the segments<br />

of the chain performs a r<strong>and</strong>om walk. The equilibrium pdf of the elongation<br />

R is then obtained by the central limit theorem:<br />

Peq(R) =<br />

�<br />

3<br />

2πMb2 �3/2 3R2<br />

−<br />

e 2Mb2 (2.7)<br />

The above pdf allows to determine the entropic force restoring the equilibrium<br />

coiled configuration, once polymers have been stretched. Being Z ∝ Peq the<br />

partition function, the free energy F reads:<br />

F = −kBT ln Z =<br />

3kBTR 2<br />

2Rmaxb<br />

+ const (2.8)<br />

where, of course, kB is the Boltzmann constant <strong>and</strong> T the temperature. The free<br />

energy variation due to a variation of elongation is<br />

f = ∂F<br />

∂R<br />

= 3kBTR<br />

2Rmaxb<br />

(2.9)<br />

meaning that the polymer as a whole behaves like a spring of elastic constant<br />

H = 3kBT<br />

2Rmaxb .<br />

Direct proportionality between recalling force <strong>and</strong> extension implies a linear<br />

relaxation ˙ R = −R/τ. A convenient estimate of the relaxation time τ has been<br />

introduced by Zimm [33], within a different chain model:<br />

τ � µR3 0<br />

kBT<br />

34<br />

(2.10)


2.1. Polymer dynamics in <strong>fluids</strong> 35<br />

Figure 2.3: Single DNA molecule (40µm) relaxing to the coiled state. In this<br />

experiment a latex bead (1µm) is tethered to an end of the molecule. The DNA,<br />

coloured with a fluorescent dye, is stretched by a uniform flow <strong>and</strong> successively<br />

let free to relax. Images are taken at 5s time intervals, from left to right [34].<br />

where µ is the dynamic viscosity of the solvent. Experiments with DNA molecules<br />

[34] confirm that relaxation can be safely considered linear, provided the elongation<br />

is small compared to the maximum extension R ≪ Rmax (see fig. 2.3).<br />

Actually, the relaxation process can be much more complex than the simple<br />

description of Zimm model. Several microscopic models of polymeric dynamics<br />

have been developed to characterize this process. An introduction to this subject<br />

can be found in the book by Doi <strong>and</strong> Edwards [35]. Nevertheless, the simple linear<br />

relaxation is able to catch, at least qualitatively, the basic features of polymer<br />

dynamics in <strong>fluids</strong>.<br />

Finally, let us mention that the description of a polymer as a freely jointed<br />

chain does not consider excluded-volume effects, which are experimentally known<br />

to alter the scaling R0 ∼ M 1/2 of the gyration radius with the degree of polymerization.<br />

The scaling exponent is, indeed, found to be ν = (0.55 ÷0.6). The origin<br />

of these effects is the possibility that segments of the chain have, in the model, to<br />

intersect or superpose to each other, which is not realistic. A better approximation<br />

of the chain dynamics is based on a self-avoiding r<strong>and</strong>om walk. While the theory<br />

of the st<strong>and</strong>ard r<strong>and</strong>om walk is quite simple, the treatment of the self-avoiding<br />

problem is more difficult. A theoretical estimation of ν has been given by Flory<br />

[36] for spatial dimensionality d < 4:<br />

ν = 3<br />

d + 2<br />

(2.11)<br />

which, for d = 3, gives the value ν = 3/5 = 0.6, not far from the experimental<br />

measures; in two dimensions, Flory’s result is exact. When d ≥ 4 the value<br />

ν = 1/2 of the ideal chain is recovered, meaning that the segments have "enough<br />

room" to avoid each other.<br />

35


36 2. Polymer solutions<br />

2.1.2 Weissenberg number<br />

The behaviour of polymers in a flow critically depends on the gradients of the<br />

velocity field. In a non-homogeneous flow, the polymer molecule can be deformed<br />

into an elongated shape <strong>and</strong> its extension R can be significantly larger than the<br />

equilibrium value R0. The deformation is determined by the competition between<br />

the stretching exerted by velocity gradients <strong>and</strong> the relaxation due to Brownian<br />

collisions with molecules of the fluid.<br />

The relative strength of stretching with respect to relaxation is quantified by<br />

the adimensional Weissenberg number:<br />

Wi ≡ γτ (2.12)<br />

defined as the product of the characteristic velocity gradient γ <strong>and</strong> the relaxation<br />

time τ. When Wi ≪ 1 relaxation is fast <strong>and</strong> polymers stay in their coiled configuration.<br />

For Wi ≫ 1, on the contrary, polymers are substantially elongated. The<br />

transition point is called coil-stretch transition <strong>and</strong> corresponds to Wi = O(1).<br />

The coil-stretch transition has been demonstrated to occur under general conditions<br />

in unsteady flow [37, 38]; in the case of steady flow, it is present in purely<br />

elongational flows, while it can be suppressed by rotation, becasue polymers do<br />

not always point in the stretching direction [39].<br />

In a turbulent velocity field polymers are stretched by a chaotic smooth flow,<br />

because their linear size is typically smaller than the Kolmogorov dissipative scale<br />

η. In such a case, the characteristic velocity gradient is measured by the Lagrangian<br />

Lyapunov exponent λ, that is the average logarithmic divergence rate of<br />

initially very close trajectories, <strong>and</strong> the Weissenberg number is:<br />

Wi = λτ (2.13)<br />

The Lyapunov exponent λ of a turbulent flow is, by dimensional analysis, the<br />

inverse of the shortest eddy turnover time. Within K41 scaling, this is τη ∼<br />

ɛ −1/2 ν 1/2 , so that λ ∼ Re 1/2 , as well as Wi ∼ Re 1/2 ; hence at a critical Reynolds<br />

number, the coil-stretch transition occurs.<br />

The growth of the molecule size R is limited by the back reaction of polymers<br />

themselves on the flow. When these are significantly elongated, elastic stresses<br />

become comparable to the visocus ones <strong>and</strong>, consequently, the flow can be modified.<br />

Specifically, this mechanism can reduce the Lyapunov exponent [39, 40],<br />

i. e. stretching, giving rise to a dynamical steady state characterized by a constant<br />

average elongation. For the sake of clearness, in the following, the Weissenberg<br />

number will always be defined in terms of the Lyapunov exponent of the <strong>Newtonian</strong><br />

fluid flow: Wi ≡ λNτ.<br />

36


2.1. Polymer dynamics in <strong>fluids</strong> 37<br />

2.1.3 Dumbbell model<br />

A model which is very often used for its simplicity is the dumbbell model [41],<br />

where the complex structure of the polymer molecule is replaced by a couple of<br />

beads of negligible mass, connected by a spring. Such a spring has the same<br />

properties of an entire freely jointed chain (see sec. 2.1.1) of Hook modulus<br />

H = 3kBT<br />

2Rmaxb . The evolution of the dumbbell end-to-end vector R = x2 − x1<br />

is determined by different contributions: the hydrodynamic drag force acting on<br />

the polymer, thermal noise, <strong>and</strong> the elastic recalling force of the spring. In a homogeneous<br />

flow, the equation of motion for R then is:<br />

˙R = − H<br />

R + Bξ (2.14)<br />

β<br />

where β is the friction coefficient <strong>and</strong> ξ is a zero mean Brownian process with<br />

correlation 〈ξi(t)ξj(t ′ )〉 = δijδ(t − t ′ ). The relaxation time is clearly introduced<br />

as:<br />

τ ≡ β<br />

H<br />

(2.15)<br />

The dependence of the constant B on the gyration radius R0 <strong>and</strong> the relaxation<br />

time τ can be derived as follows. The formal solution of eq. (2.14) allows to<br />

estimate the long time behaviour of the square mean elongation as 〈R2 〉 ∼ τB2<br />

. At 2<br />

is comparable<br />

equilibrium, the elastic energy of the spring U ∼ H〈R2 〉eq ∼ HR2 0<br />

to the thermal energy kBT , which implies B = (2R2 0 /τ)1/2 .<br />

U ( x1)<br />

R(t)<br />

U (x2 )<br />

Figure 2.4: Sketch of the dumbbell in a velocity field u(x).<br />

In a non-homogeneous flow, polymers can also get stretched, because of different<br />

velocities of the two beads (see fig. 2.4). Therefore, a term ˙ R = u(x2, t) −<br />

u(x1, t) must be added to the evolution equation (2.14). Since the flow is smooth<br />

at the scale of the polymer, we can approximate this stretching term with the ve-<br />

37


38 2. Polymer solutions<br />

locity gradient, <strong>and</strong> finally write the complete equation of motion:<br />

˙R = (R · ∇)u − 1<br />

R +<br />

τ<br />

�<br />

2 2R0 ξ (2.16)<br />

τ<br />

The linear description of the dumbbell model is adequate for relatively small<br />

elongations R ≪ Rmax. When this condition is not anymore fulfilled, it is reasonable<br />

to expect that different structures will have different elasticities, namely for<br />

very large elongations the friction coefficient β depends on the elongation itself,<br />

<strong>and</strong> so does the relaxation time. A possible generalization taking into account<br />

these effects is the Finitely Extensible <strong>Non</strong>linear Elastic model (FENE) [31], in<br />

which τ → τ R2 max −R2<br />

R 2 max−R 2 0<br />

. This simple modelling works particularly well for syn-<br />

thetic polymers (such as PolyEthileneOxide <strong>and</strong> PolyAcrylaMide).<br />

Estimating the friction coefficient as β = 6πµR0 <strong>and</strong> plugging this expression<br />

in (2.15), the Zimm relaxation time (2.10) is found. This unique characteristic<br />

time has to be intended as the slowest relaxation time of the polymer dynamics.<br />

In other words, the dumbbell model retains only the fundamental elastic mode<br />

of the polymer chain. Although higher oscillatory modes, with faster relaxation<br />

times, have been experimentally observed in DNA [42], they can be only weakly<br />

excited by the velocity gradients of a turbulent flow. Consequently, in a simplified<br />

rheological model it is enough to keep only the slowest relaxation mode.<br />

Coil-stretch transition<br />

The dumbbell model is able to reproduce the coil-stretch transition, happening<br />

when the stretching operated by the flow overcomes the relaxation of polymers. In<br />

what follows the main points of the theory developed in ref. [38] will be presented.<br />

Since our interest is now focused on the competition between stretching by<br />

velocity gradients <strong>and</strong> relaxation, thermal noise is irrelevant <strong>and</strong> the evolution<br />

equation (2.16) for the end-to-end separation vector R of the dumbbell, for the<br />

component Rα, reads:<br />

d<br />

(2.17)<br />

dt Rα = Rβ∂βuα − Rα<br />

τ<br />

To get rid of inessential degrees of freedom responsible for the orientation of<br />

the molecule, we can write R = Rn, passing to the modulus R of the vector R.<br />

Then we obtain from eq. (2.17):<br />

dρ<br />

dt<br />

1<br />

= ζ −<br />

τ<br />

(2.18)<br />

dnα<br />

dt = nβ∂βuα − ζnα (2.19)<br />

R = R0e ρ , ζ = nαnβ∂βuα (2.20)<br />

38


2.1. Polymer dynamics in <strong>fluids</strong> 39<br />

For turbulent flows, where the velocity field u is a r<strong>and</strong>om function of time,<br />

a statistical approach is needed. Specifically the problem can be attacked with a<br />

large deviations approach. Integration of eqs. (2.18)-(2.19) brings:<br />

ρ(t) = ρ0 + z − t<br />

, z =<br />

τ<br />

� t<br />

dt<br />

0<br />

′ ζ(t ′ ) (2.21)<br />

where ρ0 is the value of ρ at t = 0. The integral z in eq. (2.21) has universal properties,<br />

for times much larger than the correlation time τζ of the r<strong>and</strong>om process ζ,<br />

which in a turbulent flow is estimated as the characteristic time of the Lagrangian<br />

motion on the vicous scale. For t ≫ τζ, the variable z is the sum of a large number<br />

of independent variables <strong>and</strong> the statistics of fluctuations of z near its mean value<br />

can be deduced by the central limit theorem. In order to study large deviations, a<br />

more general formulation is needed, corresponding to the following pdf of z:<br />

G(t, z) ≈<br />

�<br />

λ = 〈ζ〉, ∆ =<br />

1<br />

z−λt<br />

√ e<br />

−tS( t ) (2.22)<br />

2π∆t<br />

dt ′ � 〈ζ(t)ζ(t ′ )〉 − λ 2�<br />

(2.23)<br />

where the "entropy density" S is a positive convex function <strong>and</strong> λ is the principal<br />

Lyapunov exponent of the flow.<br />

In order to derive the probability distribution of molecular elongations, let us<br />

consider typical fluctuations determining the largest contribution to the pdf itself.<br />

If the polymer initially has a nearly equilibrium shape (ρ0 ∼ 1) <strong>and</strong> it is successively<br />

stretched by velocity gradients up to ρ ≫ 1, then the function G(t, ρ+t/τ)<br />

provides an estimate of the contribution due to fluctuations of stretching period t.<br />

This has a sharp maximum at t∗ = ρ/β, determined from dG(t, ρ + t/τ)/dt = 0,<br />

where β is such that<br />

βS ′ (β + 1<br />

τ<br />

− λ) − S(β + 1<br />

τ<br />

− λ) = 0 (2.24)<br />

The probability distribution is thus dominated by fluctuations of duration t∗, which<br />

means:<br />

(2.25)<br />

or, equivalently, in terms of R:<br />

1<br />

−t∗S(β+<br />

G ∼ G(t∗, ρ + t∗/τ) ∼ e τ −λ)<br />

P(R) ∼ R α 0R −1−α<br />

α = S ′ (β + 1<br />

τ<br />

(2.26)<br />

− λ) (2.27)<br />

Let us observe that the characteristic value of the velocity gradient for relevant<br />

fluctuations is given by ζ ∼ ρ/t∗ + 1/τ <strong>and</strong> is of order 1/τ.<br />

39


40 2. Polymer solutions<br />

The convexity <strong>and</strong> positivity of the function S ensure that α is positive if λ <<br />

1/τ. As it is clear from eq. (2.27), α decreases with increasing λ, tending to zero<br />

when λ approaches 1/τ. In the latter case, the quadratic approximation S(x) ≈<br />

x 2 /(2∆) holds, leading to:<br />

β = 1<br />

− λ (2.28)<br />

τ<br />

α = 2<br />

� �<br />

1<br />

− λ<br />

(2.29)<br />

∆ τ<br />

The above discussion can be interpreted as the criterion for the coil-stretch transition.<br />

When the value of the Weissenberg number Wi = λτ is smaller than<br />

unity, i. e. for α > 0, the normalization integral � dRP(R) is determined by small<br />

R <strong>and</strong> the majority of polymers have nearly equilibrium size. Above unity, instead,<br />

α < 0 <strong>and</strong> the normalization integral diverges at large R; then most of the<br />

molecules are strongly stretched.<br />

40


2.2. Hydrodynamic models 41<br />

2.2 Hydrodynamic models<br />

The previous microscopic models describe the behaviour of a single polymer in<br />

a fluid, but they do not include the back-reaction that polymers have on the flow.<br />

In order to take into account this feedback mechanism it is necessary to move<br />

to a macroscopic hydrodynamic description. Indeed, the elastic properties of the<br />

solution require to include an extra stress term in the law of momentum conservation.<br />

This elastic stress term provides the macroscopic description of polymer<br />

back-reaction on the flow.<br />

2.2.1 Oldroyd-B model<br />

A simple linear <strong>viscoelastic</strong> model is the widely used Oldroyd-B model [31],<br />

based on the dumbbell model.<br />

The passage from the microscopic to the macroscopic point of view requires to get<br />

rid of microscopic degrees of freedom, such as thermal noise. Polymer behaviour<br />

is conveniently described by the conformation tensor:<br />

σij ≡ 〈RiRj〉<br />

R 2 0<br />

(2.30)<br />

where the average is taken over thermal noise or, equivalently, over a small volume<br />

containing a large number of molecules. By construction the tensor σ is<br />

symmetric, positive definite <strong>and</strong> its trace trσ is a measure of the square polymer<br />

elongation. The evolution equation for the conformation tensor follows from the<br />

linear equation (2.16) <strong>and</strong> reads:<br />

∂tσ + (u · ∇)σ = (∇u) T · σ + σ · (∇u) − 2<br />

(σ − 1) (2.31)<br />

τ<br />

where τ is the polymer relaxation time defined in eq.(2.10) <strong>and</strong> the matrix of velocity<br />

gradients is defined as (∇u)ij = ∂iuj. The conformation tensor is normalized<br />

with the equilibrium size R0 <strong>and</strong> therefore in absence of fluid flow it relaxes<br />

to the unit tensor 1.<br />

Equation (2.31) must be supplemented by the evolution equation for the velocity<br />

field, which is derived from the momentum conservation law:<br />

Dui<br />

Dt = fi + 1 ∂Tij<br />

ρ ∂xj<br />

(2.32)<br />

where f is the sum of body forces per unit mass <strong>and</strong> T is the stress tensor of the<br />

fluid.<br />

41


42 2. Polymer solutions<br />

The stress tensor of a <strong>Newtonian</strong> fluid is linear in the deformation rate tensor<br />

eij = 1/2(∂jui + ∂iuj) <strong>and</strong>, in the incompressible case, is expressed by [6]:<br />

T N ij = −Pδij + µ(∂jui + ∂iuj) (2.33)<br />

where µ is the dynamic viscosity <strong>and</strong> P pressure.<br />

In the case of a <strong>viscoelastic</strong> solution, the stress tensor is given by the sum of<br />

the <strong>Newtonian</strong> stress tensor T N <strong>and</strong> the elastic stress tensor T P , which takes into<br />

account elastic forces due to the presence of polymers. While for a <strong>Newtonian</strong><br />

fluid the stress tensor is proportional to the deformation rate tensor via the viscosity,<br />

in the Hookean approximation for the elasticity of the single polymer, the<br />

elastic stress tensor is proportional via the Hook modulus to the deformation tensor<br />

Tij = nHRiRj. The elastic stress tensor per unit volume of fluid is obtained<br />

summing the average contribution given by each polymer:<br />

T P ij = nH〈RiRj〉 = nHR 2 0 σij<br />

(2.34)<br />

where n is the concentration of polymer molecules.<br />

For an incompressible fluid of constant density ρ, plugging the expression of<br />

the stress T = T N + T P in the momentum conservation law, one obtains:<br />

∂tu + (u · ∇)u = −∇P + ν∆u + 2νη<br />

∇ · σ + f (2.35)<br />

τ<br />

The solvent kinematic viscosity is denoted by ν = µ/ρ <strong>and</strong> here η is the zero-shear<br />

contribution of polymers to the total viscosity of the solution νT = ν(1 + η):<br />

η = nHR2 0 τ<br />

2µ<br />

(2.36)<br />

Equation (2.35) is a generalization of Navier-Stokes equation for the viscoelatic<br />

solution <strong>and</strong> together with eq. (2.31) completely determines the dynamics<br />

of Oldroyd-B model.<br />

Viscosity renormalization<br />

In the limit τ → 0, the elastic force originated from thermal motion keeps the<br />

molecules near their equilibrium configuration, <strong>and</strong> the polymer solution is expected<br />

to behave like a <strong>Newtonian</strong> fluid. Indeed, a perturbative expansion in τ for<br />

the conformation tensor<br />

σij = σ (0)<br />

ij<br />

substituted in eq. (2.31) gives at zero <strong>and</strong> first order:<br />

σ 0 ij = δij; σ (1)<br />

ij<br />

+ τσ(1)<br />

ij + O(τ2 ) (2.37)<br />

= 1<br />

2 (∂jui + ∂iuj) = eij<br />

42<br />

(2.38)


2.2. Hydrodynamic models 43<br />

Therefore, at first order in τ the elastic stress is proportional to the deformation<br />

tensor<br />

T P ij = nHR 2 0[δij + τeij + O(τ 2 )] (2.39)<br />

<strong>and</strong> the fluid is <strong>Newtonian</strong> up to higher order corrections O(τ 2 ). Nevertheless, the<br />

presence of polymers affects fluid properties also in this <strong>Newtonian</strong> limit, because<br />

the fluid is partially trapped in the coiled polymers, producing a change of the<br />

total viscosity of the solution µT , which is renormalized as:<br />

Energy balance<br />

µT = µ + 1<br />

2 nHR2 0τ = µ(1 + η) (2.40)<br />

The free energy of the <strong>viscoelastic</strong> fluid is the sum of kinetic <strong>and</strong> elastic contributions:<br />

�<br />

F = d 3 �<br />

1<br />

x<br />

2 ρu2 + ηνρ<br />

�<br />

[trσ − log det σ]<br />

τ<br />

(2.41)<br />

where the last term represents the entropy of polymer molecules. In order to<br />

compute the free energy time derivative, let us separately consider the rate of<br />

change of different contributions. Through eqs. (2.31)-(2.35) one obtains:<br />

u 2<br />

2 = f · u − ν(∂iuj) 2 − ην<br />

τ [σij∂jui + σij∂iuj] (2.42)<br />

d<br />

dt<br />

d<br />

dt trσ = [σij∂jui + σij∂iuj] − 2<br />

tr[σ − 1] (2.43)<br />

τ<br />

d<br />

2<br />

log det σ =<br />

dt τ tr[σ−1 − 1] (2.44)<br />

The forcing pumps kinetic energy into the system, that is partially dissipated by<br />

viscosity <strong>and</strong> polymer relaxation. The term σij∂jui+σij∂iuj, representing the exchange<br />

between kinetic <strong>and</strong> elastic energy, does not have a definite sign. Summing<br />

together the different contributions, the free energy rate of change is:<br />

�<br />

∂F<br />

= ρ d<br />

∂t 3 �<br />

x f · u − ν(∂iuj) 2 − 2ην<br />

τ2 tr[σ − 21 + σ−1 �<br />

] (2.45)<br />

Since the conformation tensor is positive definite <strong>and</strong> symmetric, it can be<br />

decomposed as the product of two symmetric matrices S:<br />

σij = SikSkj<br />

(2.46)<br />

Using this decomposition, the trace term in eq. (2.45), representing the energy<br />

dissipation rate due to polymers, can be rewritten as<br />

tr[σ − 21 + σ −1 ] = tr[(S − S −1 ) 2 ] (2.47)<br />

43


44 2. Polymer solutions<br />

which shows that it has a definite sign.<br />

In a statistically steady state the average free energy is constant <strong>and</strong> the energy<br />

balance reads:<br />

ɛI = ν〈(∂iuj) 2 〉 + 2ην<br />

τ2 � � −1<br />

〈trσ〉 − 2tr1 + 〈trσ 〉 (2.48)<br />

where ɛI is the average energy input per unit mass. Assuming kinetic energy <strong>and</strong><br />

polymer elongation are statistically constant, it follows that the average entropy<br />

production rate vanishes:<br />

〈trσ −1 〉 − tr1 = 0 (2.49)<br />

<strong>and</strong> the entropy of polymer molecules is conserved.<br />

2.2.2 FENE-P model<br />

The linear Oldroyd-B model is based on the assumption that polymers can be<br />

modeled as Hookean springs <strong>and</strong>, consequently, it allows infinite extension of<br />

polymer molecules. This is clearly unphysical because the polymer end-to-end<br />

separation R is bounded by the maximum length Rmax. When R approaches<br />

Rmax, elastic nonlinearities become important <strong>and</strong> the linear approximation fails.<br />

A more refined, but still conceptually simple, description is that of FENE-P model<br />

[31], where the elastic constant H is replaced by the function:<br />

H(R) = H R2 max − R2 0<br />

R2 (2.50)<br />

max − R2<br />

diverging for R → Rmax, which means that the recoil force becomes infinite when<br />

the polymer is extremely stretched.<br />

However, the introduction of nonlinearities produces a non-closed evolution<br />

equation for the conformation tensor 〈RiRj〉. A commonly accepted closure is<br />

the Peterlin approximation [43], consisting in a pre-averaging in the nonlinear<br />

function:<br />

H(R 2 ) → H(〈R 2 〉) = H R2 max − R2 0<br />

R2 max − 〈R2 〉<br />

(2.51)<br />

The coupled equations for the conformation tensor <strong>and</strong> the velocity field, in<br />

the FENE-P model are:<br />

∂tu + (u · ∇)u = −∇P + ν∆u + 2ην<br />

H(trσ)∇ · σ + f (2.52)<br />

τ<br />

∂tσ + (u · ∇)σ = (∇u) T · σ + σ · (∇u) − 2<br />

(H(trσ)σ − 1) (2.53)<br />

τ<br />

where the nonlinear factor has been rewritten as:<br />

H(trσ) = trmax − tr1<br />

(2.54)<br />

trmax − trσ<br />

44


2.2. Hydrodynamic models 45<br />

with trmax = R 2 max /R2 0 .<br />

The FENE-P model is able to reproduce features of polymer solutions not captured<br />

by Oldroyd-B model, like the shear thinning, i. e. the decrease of viscosity<br />

at increasing shear rates. Moreover, the finite molecular extensibility reduces the<br />

onset of numerical instabilities due to strong gradients of the conformation tensor<br />

field. For these reasons this model is generally adopted in numerical simulations<br />

of <strong>viscoelastic</strong> channel flows [44].<br />

45


46 2. Polymer solutions<br />

2.3 Drag reduction<br />

Since the work of the chemist Toms [45], it is known that the addition of small<br />

amounts of polymers to a turbulent flow drastically reduces the friction drag.<br />

Specifically, Toms observed that the addition of a few parts per million in weight<br />

of long chain polymers to a pipe flow has the consequence of reducing the drag of<br />

about 80%.<br />

Friction drag in a pipe flow is usually quantified by the adimensional Fanning<br />

friction factor f:<br />

f = ∆P<br />

ρU2 R<br />

L<br />

(2.55)<br />

where R is the radius of the pipe, ρ is the fluid density, ∆P is the pressure drop on<br />

a distance L <strong>and</strong> U is the mean velocity across the section. Physically, the friction<br />

factor is the ratio of the external pressure difference sustaining the motion, <strong>and</strong> the<br />

kinetic energy of the fluid. A significant reduction of this factor thus allows to let<br />

a fluid flow in a pipe at a much lower cost in terms of energy, which clearly has a<br />

high relevance for practical applications.<br />

Figure 2.5: A schematic illustrating the onset of drag reduction <strong>and</strong> the MDR<br />

asymptote in Pr<strong>and</strong>tl-Karman coordinates. The Pr<strong>and</strong>tl-Karman law corresponds<br />

to the turbulent behaviour of <strong>Newtonian</strong> <strong>fluids</strong>. The dotted line qualitatively represents<br />

the friction reduction in the <strong>viscoelastic</strong> case. The dashed line is for a larger<br />

polymer concentration.<br />

In <strong>Newtonian</strong> flows the drag coefficient f is a function of the Reynolds number,<br />

which is Re = 2RU/ν for a pipe flow. The dependence is conventionally<br />

shown in Pr<strong>and</strong>tl-Karman (P-K) coordinates: 1/ √ f versus log(Re √ f) (see<br />

46


2.3. Drag reduction 47<br />

fig. 2.5). In the laminar regime the drag decreases as Re −1 until a critical Reynolds<br />

number is reached. Transition to <strong>turbulence</strong> causes a sudden increase of the coefficient<br />

f which, in fully developed <strong>turbulence</strong>, almost reaches a constant value<br />

with only a weak logarithmic dependence on Re, corresponding to a straight line<br />

in P-K coordinates.<br />

Dilute polymer solutions deviate from the Pr<strong>and</strong>tl-Karman law: below a critical<br />

value of Re their behaviour is similar to that of a <strong>Newtonian</strong> fluid, but at larger<br />

Reynolds numbers the drag is drastically reduced with respect to the <strong>Newtonian</strong><br />

case <strong>and</strong> it finally reaches a universal asymptote, independent of the specific type<br />

of polymer <strong>and</strong> of the concentration, known in the literature as Maximum Drag<br />

Reduction asymptote (MDR).<br />

Many theories have been proposed to explain what is the basic physical mechanism<br />

of the drag reduction phenomenon. Lumley [46] proposed a quantitative<br />

approach based on the phenomenology of the channel flow. Another approach is<br />

due to de Gennes [47], who suggested an elasticity-based criterion. Much effort<br />

to the underst<strong>and</strong>ing of the problem has been dedicated also by Piva, Casciola <strong>and</strong><br />

coworkers [44], by Procaccia <strong>and</strong> coworkers [48], by Benzi <strong>and</strong> coworkers [49].<br />

Finally, recent experimental results are reported in [50, 51, 52]. Despite the considerable<br />

number of studies, a quantitative fully satisfactory explanation of drag<br />

reduction is still lacking, <strong>and</strong> the phenomenological theories are not universally<br />

accepted [51]. The experimental <strong>and</strong> theoretical studies on drag reduction have<br />

been performed in wall bounded flows, due to technological <strong>and</strong> industrial applications.<br />

It is worth remarking that this effect can take place also in the absence of<br />

material boundaries [53].<br />

47


48 2. Polymer solutions<br />

2.4 Elastic <strong>turbulence</strong><br />

A remarkable phenomenon of recent experimental discovery [54, 55, 56, 57] is the<br />

onset of elastic <strong>turbulence</strong>. This effect appears in a somehow opposite limit in the<br />

space of stability parameters of the flow, with respect to drag reduction. Whereas<br />

drag reduction is a high Reynolds number phenomenon, elastic <strong>turbulence</strong> occurs<br />

in the limit of vanishing Re, provided the elasticity El ≡ Wi/Re is high enough.<br />

The presence of polymers modifies the stability of a laminar flow. Indeed,<br />

when stretched by a primary shear flow, polymers can trigger elastic instabilities.<br />

These instabilities give rise to a secondary flow that further stretches the<br />

molecules until a stationary state is reached. Such a state displays features typical<br />

of turbulent flows, such as spatial <strong>and</strong> temporal irregular behaviour, with a broad<br />

range of active scales (see fig. 2.6). The possibility to achieve turbulent states for<br />

arbitrarily small Reynolds numbers is clearly very important for mixing related<br />

applications in microchannel flows [56, 58], where mixing efficiency is limited by<br />

the practical difficulty of reaching large Re.<br />

Figure 2.6: Representative snapshots of elastic <strong>turbulence</strong> in a flow generated by<br />

two coaxial rotating disks. [(a), (b)] Polymer solution at Wi = 6.5, Re = 0.35;<br />

[(c), (d), (e)] polymer solution at Wi = 13, Re = 0.7; (f) pure solvent at Re = 1<br />

[57].<br />

A measure of elastic <strong>turbulence</strong>, as suggested in [55], is ratio between the<br />

average shear stress <strong>and</strong> its corresponding value for a laminar flow. The results<br />

48


2.4. Elastic <strong>turbulence</strong> 49<br />

reported refer to a flow between two coaxial rotating disks. When the relative<br />

angular velocity of the plates was increased the rescaled shear stress significantly<br />

growed, exhibiting a sharp transition. The maximum stress value was found to<br />

correspond to that of a <strong>Newtonian</strong> flow at Re ∼ 10 4 , whereas the measures were<br />

taken at Re < 1, showing the elastic origin of this effect.<br />

Numerical analysis of elastic <strong>turbulence</strong> in a two-dimensional flow will be the<br />

subject of chapter 5.<br />

49


Chapter 3<br />

Phase separation in <strong>binary</strong> <strong>fluids</strong><br />

The quenching of a system from a disordered phase into an ordered one produces<br />

a time dependent growth process of ordered regions. The temporal evolution of<br />

these regions is the subject of phase ordering dynamics, in which the problem is<br />

approached by means of a statistical description based on rather general forms of<br />

equations of motion for an order parameter [59, 60]. The generality of the approach<br />

<strong>and</strong> the universality properties of the coarsening process make the subject<br />

of interest for a varied number of physical systems ranging from solid alloys, to<br />

polymer blends, multiphase <strong>fluids</strong> <strong>and</strong> nematic liquid crystals [60, 61, 62, 63].<br />

The first to address the problem were, in the late 1950’s, Cahn <strong>and</strong> Hilliard [64]<br />

who investigated the behaviour of metallurgical systems, specifically the spinodal<br />

decomposition of <strong>binary</strong> alloys. Similar phenomena occur in the phase separation<br />

of <strong>binary</strong> <strong>fluids</strong>, i. e. <strong>fluids</strong> composed by either two phases of the same chemical<br />

species, or phases of different composition. In this case, however, the phenomenology<br />

is complicated by the interplay with the fluid dynamics. For instance,<br />

since Siggia’s seminal work [61] it is well known that hydrodynamics may accelerate<br />

the domain growth (see references [65] <strong>and</strong> [66] for recent developments in<br />

three- <strong>and</strong> two-dimensional <strong>fluids</strong>, respectively). Finally, by assuming more complex<br />

algebraic structures for the order parameter, such as vector <strong>and</strong> tensor fields<br />

rather than the scalars usually considered, it is possible to describe phase ordering<br />

in nematic liquid crystals [60].<br />

This chapter is intended as an introduction to the physics of phase separation<br />

in <strong>binary</strong> <strong>fluids</strong>, in view of a study on the interaction between <strong>turbulence</strong> <strong>and</strong><br />

domain growth, presented in chapter 6. The main thermodynamical aspects are<br />

introduced <strong>and</strong> growth laws are discussed, both in the case of a fluid at rest <strong>and</strong> in<br />

presence of a hydrodynamic velocity field.<br />

51


52 3. Phase separation in <strong>binary</strong> <strong>fluids</strong><br />

3.1 Thermodynamics of <strong>binary</strong> <strong>mixtures</strong><br />

Let us consider the cooling of a fluid consisting of two species A <strong>and</strong> B, starting<br />

from a completely mixed state. Below a critical temperature Tc , the originally<br />

homogeneous phase separates into distinct A-rich <strong>and</strong> B-rich components <strong>and</strong> the<br />

system spontaneously evolves into two phases separated by an interface.<br />

The thermodynamic behaviour of the system can be understood by considering<br />

the free energy F. Assuming the volume is fixed, this is a functional of a single<br />

composition variable, accounting for the local fraction of the two <strong>fluids</strong>:<br />

θ(x, t) ≡ nA − nB<br />

nA + nB<br />

(3.1)<br />

where nA,B(x, t) are the number densities of fluid A <strong>and</strong> B, respectively, having<br />

assumed unit mass for A <strong>and</strong> B particles.<br />

In what follows, we will focus the attention on phase separation taking place<br />

through spinodal decomposition [60, 61, 62, 65].<br />

3.1.1 Spinodal decomposition<br />

When a <strong>binary</strong> fluid is quenched into the coexistence region of its phase diagram<br />

(see fig. 3.1) it can phase separate in two different ways. The first mechanism<br />

occurs when the point to which the fluid is quenched is near the binodal. In this<br />

case phase separation generally proceeds via nucleation of droplets of one phase<br />

forming in the other phase. For example, if the fluid is quenched to a state point<br />

inside the binodal on the liquid side, then bubbles of the gas phase can nucleate in<br />

the bulk of the metastable liquid.<br />

However, if the fluid is quenched to a state point well inside the binodal, then<br />

a different separation mechanism, called spinodal decomposition, prevails. This<br />

type of decomposition is characterized by the formation of a bicontinuous structure<br />

of the two phases. Indeed, below the spinodal line the free energy changes<br />

curvature <strong>and</strong> the system becomes locally unstable. The free energy can then be<br />

lowered, in any local neighbourhood, by creating two domains whose composition<br />

differs only infinitesimally from the initial one. Accordingly, infinitesimal<br />

fluctuations will grow by diffusion until there is local coexistence of domains at<br />

compositions approaching the equilibrium values of the single phases.<br />

In order to describe this process, it is convenient to set up a continuous formulation<br />

in terms of the coarse-grained order parameter field θ(x, t). The L<strong>and</strong>au-<br />

Ginzburg free energy of the mixture is:<br />

� � 2 ξ<br />

F[θ]= dx<br />

2 |∇θ|2 �<br />

+ V (θ)<br />

(3.2)<br />

52


3.1. Thermodynamics of <strong>binary</strong> <strong>mixtures</strong> 53<br />

Figure 3.1: Phase diagram for spinodal decomposition. Below the critical temperature<br />

Tc the system starts to separate, whereas for higher temperatures it remains<br />

completely mixed.<br />

where V (θ) is a st<strong>and</strong>ard double well potential (fig. 3.2):<br />

V (θ) = 1<br />

2 r(T)θ2 + 1<br />

4 uθ4<br />

(3.3)<br />

The coefficient r(T) ∝ (T − Tc) in (3.3) becomes negative at low temperatures<br />

T < Tc, <strong>and</strong> the resulting two minima of V correspond to the equilibrium values<br />

θ = ±1 in the bulk phases. Since the interest is here on quenches into the ordered<br />

phase, it is reasonable to assume T = 0 <strong>and</strong> |r| = |u| = 1; moreover, we adopt<br />

the convention V (0) = 0, so that V (±1) = −1/4. The gradient term penalizes<br />

the formation of sharp gradients in composition <strong>and</strong> ξ gives the length scale of the<br />

interface width.<br />

3.1.2 Surface tension<br />

The gradient term in the free energy functional (3.2) provides a nonzero surface<br />

tension. This can be calculated as follows [60, 65]. Stationarity of F requires<br />

ξ 2 ∆θ = V ′ (θ) (3.4)<br />

Let us consider a flat interface; introducing a coordinate g normal to it <strong>and</strong> setting<br />

θ(0) = 0, integration of eq. (3.4) across the domain wall gives<br />

ξ 2<br />

2<br />

� �2 dθ<br />

= V (θ) − V (±1) (3.5)<br />

dg<br />

53


54 3. Phase separation in <strong>binary</strong> <strong>fluids</strong><br />

0.3<br />

0.2<br />

0.1<br />

2 �1 1 2 Θ<br />

�0.1<br />

�0.2<br />

V�Θ�<br />

Figure 3.2: Model potential for phase separation.<br />

Surface tension is the energy per unit area of the wall:<br />

�<br />

σ ≡ dg ξ 2 |∇θ| 2<br />

interface<br />

Using (3.5) <strong>and</strong> the fact that ∇θ = 0 in the bulk, the above expression becomes<br />

σ =<br />

� +1<br />

dθ ξ 2<br />

−1<br />

1/2 [V (θ) − V (±1)] 1/2<br />

(3.6)<br />

(3.7)<br />

Given a form of the potential V (θ) the value of surface tension can thus be calculated.<br />

Let us observe that the order parameter varies smoothly in the interface<br />

region <strong>and</strong> saturates to θ = ±1 in the bulk of the domains. Then, it follows that σ<br />

is localized in the domain wall, <strong>and</strong> that the driving force for the domain growth is<br />

the wall curvature, since the energy of the system can decrease only by reducing<br />

the total wall area.<br />

3.2 Coarsening in a fluid at rest<br />

In phase separation the order parameter is conserved on average, due to the local<br />

exchange of particles of species A <strong>and</strong> B, which leads to a diffusive evolution<br />

equation, known as Cahn-Hilliard equation. In absence of any external flow, this<br />

reads:<br />

∂θ<br />

∂t<br />

= Γ∆<br />

� �<br />

δF<br />

= Γ∆µ (3.8)<br />

δθ<br />

where Γ is a mobility parameter that, here <strong>and</strong> in the following, will be assumed<br />

constant, <strong>and</strong> µ = −θ+θ 3 +ξ 2 ∆θ is the chemical potential governing the exchange<br />

54


3.2. Coarsening in a fluid at rest 55<br />

of particles. It is worth remarking that Cahn-Hilliard equation can be written in<br />

the form of a continuity equation:<br />

∂tθ = −∇ · j (3.9)<br />

j = −∇µ (3.10)<br />

where adimensional units have been introduced by rescaling of lengths with ξ <strong>and</strong><br />

of times with the diffusive time tm = ξ 2 /Γ.<br />

3.2.1 Domain growth<br />

After the temperature quench phase separation starts, forming, in a short time of<br />

the order of tm, well defined domains of the single phases, which then grow until<br />

complete separation is achieved.<br />

Figure 3.3: DNS of Cahn-Hilliard equation at resolution 512 2 . Snapshots of the<br />

composition field θ at times growing from left to right. Black/white codes θ ± 1.<br />

Inspection of a time sequence, as the one shown in fig. 3.3, suggests that domain<br />

growth is a scaling phenomenon, provided that the typical domain size L(t)<br />

is larger than the interface width <strong>and</strong> smaller than the whole system size. In this<br />

range, the growth law can be derived according to the following argument [60].<br />

In the bulk of a single phase the order parameter is constant, so that ∂tθ = 0<br />

<strong>and</strong> the chemical potential obeys the Laplace equation<br />

∆µ = 0 (3.11)<br />

From the behaviour at the boundaries, it can be inferred that<br />

µ ∼ σ<br />

L<br />

(3.12)<br />

where σ is surface tension <strong>and</strong> 1/L is the curvature of the domain wall. The<br />

interface velocity is v ∼ j ∼ ∇µ. Estimating the gradient as 1/L, the equation of<br />

motion for L then is:<br />

dL<br />

dt<br />

= σ<br />

L 2<br />

55<br />

(3.13)


56 3. Phase separation in <strong>binary</strong> <strong>fluids</strong><br />

Integration of eq. (3.13) gives the scaling L(t) ∼ (σt) 1/3 ; reinserting the physical<br />

dimensions, we finally get:<br />

�<br />

t<br />

L(t) ∼ ξ<br />

tm<br />

� 1/3<br />

3.3 Hydrodynamics <strong>and</strong> coarsening<br />

(3.14)<br />

A detailed description of phase separation in <strong>fluids</strong> requires to consider hydrodynamic<br />

effects [60, 62, 65]. To do this Cahn-Hilliard equation must be modified<br />

to include an advective term, accounting for transport of the order parameter operated<br />

by a fluid flow. Moreover, coupling with Navier-Stokes equation for the<br />

velocity field must be considered.<br />

The interaction between coarsening <strong>and</strong> fluid dynamics arises as a consequence<br />

of inhomogeneities in composition. These are controlled by the chemical<br />

potential µ = δF/δθ, which describes the variation of free energy resulting from<br />

a small local change in composition. If the latter is not uniform, then a thermodynamic<br />

force density −θ∇µ acts on the fluid. The two species are pulled in<br />

opposite dircetions by the chemical potential gradient <strong>and</strong> the net force vanishes<br />

only if there is a single fluid or when the two <strong>fluids</strong> are perfectly mixed, i. e. when<br />

θ = 0. The evolution equations then become:<br />

∂tθ + v · ∇θ = Γ∆µ = Γ(−∆θ + ∆θ 3 − ξ 2 ∆ 2 θ) (3.15)<br />

ρ[∂tv + (v · ∇)v] = −∇p + µv∆v − θ∇µ + f (3.16)<br />

where ρ = nA + nB is the mean fluid density, v is the (incompressible) velocity<br />

field, p the pressure, µv dynamic viscosity <strong>and</strong> f an external stirring source.<br />

From eq. (3.16) it is seen that the order parameter is now an active field, since<br />

it is not simply transported by the flow, but it is able to react on it. Let us now<br />

rewrite the feedback term in a different manner:<br />

−θ∇µ = −ξ 2 ∆θ∇θ + ∇<br />

�<br />

1<br />

2 θ2 − 3<br />

4 θ4 + ξ 2 �<br />

θ∆θ<br />

(3.17)<br />

Now, the last term on the right-h<strong>and</strong> side of eq. (3.17) can be reabsorbed in the<br />

pressure gradient, <strong>and</strong> the stress tensor results modified by this chemical pressure<br />

term, with respect to the <strong>Newtonian</strong> case.<br />

Finally, the equations of motion are:<br />

∂tθ + v · ∇θ = Γ∆µ = Γ(−∆θ + ∆θ 3 − ξ 2 ∆ 2 θ) (3.18)<br />

∂tv + (v · ∇)v = −∇p + ν∆v − ξ 2 ∆θ∇θ + f (3.19)<br />

56


3.3. Hydrodynamics <strong>and</strong> coarsening 57<br />

where ν = µv/ρ is kinematic viscosity <strong>and</strong> it has been assumed ρ = 1.<br />

Observe that eq. (3.19) formally is the MHD equation for the velocity field,<br />

provided one identifies the composition variable with the magnetic vector potential.<br />

A discussion on the similiraties in the phenomenology of MHD <strong>and</strong> that of<br />

phase separation can be found in ref. [62].<br />

3.3.1 Growth laws<br />

When hydrodynamics is taken into account, the coupling between Cahn-Hilliard<br />

<strong>and</strong> Navier-Stokes equations strongly limits the possibility to derive growth laws<br />

for the typical domain size. Nevertheless, some dimensional arguments can help<br />

finding out how fast domains grow in different regimes individuated by the relative<br />

strength of interfacial contributions with respect to the hydrodynamic ones.<br />

Let us now reconsider eqs. (3.18)-(3.19) in absence of external forcing, i. e. for<br />

f = 0. Then, the only driving force for the velocity field is the feedback term<br />

−ξ 2 ∆θ∇θ. In order to compare different contributions, we rescale time, length<br />

<strong>and</strong> velocity by means of natural units that can be defined in terms of the coarsening<br />

parameters ξ <strong>and</strong> Γ, namely:<br />

t → t<br />

tm<br />

; tm = ξ2<br />

Γ<br />

(3.20)<br />

x → x<br />

; lm = ξ (3.21)<br />

lm<br />

v → v<br />

vm<br />

; vm = Γ<br />

ξ<br />

After rescaling, the evolution equations become:<br />

(3.22)<br />

∂tθ + v · ∇θ = −∆θ + ∆θ 3 − ξ 2 ∆ 2 θ (3.23)<br />

∂tv + (v · ∇)v = ν λξ2<br />

∆v − ∆θ∇θ (3.24)<br />

Γ Γ2 where λ = 1 is put only for dimensional reasons, its dimensions being those of a<br />

squared velocity. Let us discuss different regimes, assuming the initial condition<br />

v = 0 for eq. (3.24).<br />

• If Γ is extremely high, such that λξ 2 /Γ 2 ≪ 1, the velocity remains very<br />

small <strong>and</strong> eq. (3.24) is almost not effective, allowing to disregard the advective<br />

term in eq. (3.23). The system then evolves according to Cahn-Hilliard<br />

equation <strong>and</strong> the domain size grows as L(t) ∼ t 1/3 .<br />

• If ν/Γ ≫ λξ 2 /Γ 2 , then viscous damping overcomes the forcing produced<br />

by composition inhomogeneities <strong>and</strong>, therefore, we fall again in the previous<br />

case, <strong>and</strong> hence L(t) ∼ t 1/3 .<br />

57


58 3. Phase separation in <strong>binary</strong> <strong>fluids</strong><br />

• When, instead, ν/Γ ≪ λξ 2 /Γ 2 , the viscous term is negligible <strong>and</strong> the force<br />

from the interface will be balanced by the hydrodynamic inertial terms. Estimating<br />

−θ∇µ ∼ σ/L 2 , ∇ ∼ 1/L, v ∼ dL/dt, the resulting growth law<br />

[65, 67] is:<br />

L(t) ∼ t 2/3<br />

(3.25)<br />

However, it should be noted that it can be not completely correct to consider<br />

this regime as universally described by the growth law t 2/3 [66, 68]. Indeed,<br />

the physics of the domains depends on the value of the fluidity, defined as<br />

the ratio of interfacial to viscous terms (λξ 2 /Γ 2 )/(ν/Γ) = λξ 2 /(Γν) [68].<br />

When this number is very large, inertial terms become so important that<br />

phase separation might be partially inhibited <strong>and</strong> the order parameter fails<br />

in reaching its equilibrium values θ = ±1. Different phenomena, such as<br />

double phase separation [68], may then arise <strong>and</strong> different growth laws can<br />

be simultaneously present [66]. Finally, the morphology of the domains,<br />

displaying nested structures, can be significantly different from the bicontinuous<br />

structure characteristic of symmetric <strong>mixtures</strong>, suggesting essential<br />

deviations from the growth mechanism of spinodal decomposition.<br />

• In 3D balancing of the viscous term with the feedback due to coarsening<br />

produces a linear growth law L(t) ∼ t [61, 65]. In 2D this regime has not<br />

been observed.<br />

In conclusion, the interplay between coarsening <strong>and</strong> fluid dynamics is still<br />

far from being completely understood <strong>and</strong> the scaling laws are still debated. In<br />

chapter 6 we will focus on the case of symmetric <strong>binary</strong> <strong>mixtures</strong> in which well<br />

defined domains form <strong>and</strong> grow in time with the power law L(t) ∼ t 2/3 , studying<br />

the problem in presence of an external stirring acting on the velocity field.<br />

58


Part II<br />

59


Chapter 4<br />

Small-scale statistics of<br />

<strong>viscoelastic</strong> <strong>turbulence</strong><br />

Polymer additives are able to deeply modify the rheology of the fluid they are<br />

dissolved into, even when they are present in very low concentrations. In high<br />

Reynolds number flows, the most renowned effect is probably the reduction of<br />

friction drag (see chapter 2 <strong>and</strong> references therein).<br />

Most studies relative to this fully developed <strong>turbulence</strong> regime focused on<br />

dilute polymer solutions in channel or pipe geometry, where boundary effects are<br />

important [69]. Nevertheless, recent experimental [52, 70] <strong>and</strong> numerical [40, 53,<br />

71] works have shown that polymers affect the turbulent flow even far from (or<br />

in absence of) boundaries. In particular, ref. [71] studied the modification of the<br />

turbulent cascade induced by polymers in numerical simulations of homogeneous<br />

isotropic <strong>turbulence</strong> of the FENE-P model.<br />

This chapter deals with the effects of polymer addition on the small-scale<br />

statistics in fully developed homogeneous isotropic <strong>turbulence</strong>. The behaviour<br />

of a simplified <strong>viscoelastic</strong> model, namely the uniaxial model, is investigated by<br />

means of direct numerical simulations, in three dimensions [72].<br />

We show that, by increasing the elasticity of polymers, the energy flux in the<br />

turbulent cascade is partially suppressed <strong>and</strong> transferred to the elastic degrees of<br />

freedom. This suppression remains partial even for large values of elasticity: as<br />

a consequence the energy flux to small scales remains finite <strong>and</strong> the small-scale<br />

statistics, such as acceleration probability density function, retain some characteristics<br />

of <strong>Newtonian</strong> flows.<br />

61


62 4. Small-scale statistics of <strong>viscoelastic</strong> <strong>turbulence</strong><br />

4.1 Uniaxial model<br />

The mechanism by which dilute polymer solutions can influence turbulent flows is<br />

the extreme extensibility of polymers, as discussed in chapter 2. The description<br />

of polymeric <strong>fluids</strong> in terms of <strong>viscoelastic</strong> models gets greatly simplified within<br />

the uniaxial model (first introduced in [39]), that will now be constructed. In order<br />

to do this, it is worth to preliminarily summarize the main issues of the interaction<br />

between polymeric <strong>and</strong> fluid dynamics.<br />

Polymers, typically composed by a large number of monomers, at equilibrium<br />

are coiled in a ball of radius R0. In presence of a nonhomogeneous flow, the<br />

molecule is deformed in an elongated structure characterized by its end-to-end<br />

distance R which can be much larger than R0. The deformation of molecules<br />

is the result of the competition between the stretching induced by differences of<br />

velocities <strong>and</strong> the entropic relaxation of polymers to their equilibrium configuration.<br />

This relaxation is linear, provided the elongation is small compared with the<br />

maximum extension R ≪ Rmax, <strong>and</strong> can be characterized by a typical relaxation<br />

time τ.<br />

These ingredients lead to the simplest model which describes the behaviour of<br />

a polymer in a flow, the dumbbell model (sec. 2.1.3):<br />

dR<br />

dt = (∇u)TR − 1<br />

�<br />

2 2R0 R + ξ (4.1)<br />

τ τ<br />

where ξ is a Brownian process with correlation 〈ξi(t)ξj(t ′ )〉 = δijδ(t − t ′ ). The<br />

relative importance between polymer relaxation <strong>and</strong> stretching is measured by<br />

the Weissenberg number Wi: when Wi ≪ 1 polymers are in the coiled state,<br />

whereas for Wi ≫ 1 they are substantially elongated; Wi = O(1) corresponds to<br />

the coil-stretch transition.<br />

The development of a fully hydrodynamic model requires to include the backreaction<br />

of polymers on the flow. In the case of dilute solutions, for which the<br />

polymer concentration n satisfies nR 3 0 ≪ 1, the influence of polymers in the<br />

coiled state on the fluid is negligible. Above the coil-stretch transition, polymers<br />

start to affect the flow. This regime is characterized by large elongations R ≫<br />

R0, which allow to disregard the thermal noise in (4.1). Polymer solutions at<br />

macroscopic scales, i.e. at scales much larger than typical interpolymer distances,<br />

can be described by a local elongation field R(x, t) which evolves according to<br />

∂R<br />

∂t + u · ∇R = (∇u)T · R − R<br />

τ<br />

(4.2)<br />

Taking the divergence of (4.2) one easily sees that ∇ · R decays in time <strong>and</strong> thus<br />

R can be taken solenoidal.<br />

62


4.1. Uniaxial model 63<br />

The effect of polymers on the fluid is in the modification of the stress tensor<br />

through an additional elastic component T P which takes into account the elastic<br />

forces of polymers T P<br />

ij = (2νη/τ)(RiRj/R 2 0) where ν is the solvent viscosity <strong>and</strong><br />

η (proportional to polymer concentration) represents the zero-shear contribution<br />

of polymers to the total solution viscosity ν(1 + η). The Navier-Stokes equation<br />

for the incompressible velocity field u(x, t) thus becomes<br />

∂u<br />

∂t<br />

+ u · ∇u = −∇p + ν∆u + 2νη<br />

τ<br />

R · ∇R<br />

R 2 0<br />

(4.3)<br />

Equations (4.2) <strong>and</strong> (4.3) are the so-called uniaxial model for <strong>viscoelastic</strong> flows.<br />

Observe that by introducing the rescaled variable B = � 2νη/τ(R/R0),<br />

equations (4.2)-(4.3) formally become the MHD equations for a plasma at zero<br />

resistivity with a linear damping −B/τ [25, 73]. In this representation the coefficient<br />

η disappears <strong>and</strong> thus the dynamics of the uniaxial model is independent of<br />

the concentration (which is physically consistent with the assumption of linearity<br />

<strong>and</strong> strong elongation). In the following we will thus take η = 1 <strong>and</strong> absorb R0 in<br />

the definition of R: R/R0 → R.<br />

The total energy (kinetic plus elastic) ET = (〈u 2 〉 + 〈B 2 〉)/2 of the flow is<br />

dissipated at a rate<br />

dET<br />

dt = −ν〈|∇ × u|2 〉 − 1<br />

τ 〈B2 〉 = −ɛν − ɛτ<br />

(4.4)<br />

where ɛν is the viscous dissipation while the second term ɛτ represents the additional<br />

dissipation due to the relaxation of polymers to their equilibrium configuration.<br />

Oldroyd-B model: limit of large elongations<br />

The uniaxial model can also be derived from Oldroyd-B model, by taking the limit<br />

of large elongations [39, 74]. Following ref. [74], in such a limit the Oldroyd-B<br />

evolution equation for the conformation tensor can be written in a Lagrangian<br />

reference frame as:<br />

∂tσ = (∇u) T · σ + σ · (∇u) − 2<br />

σ (4.5)<br />

τ<br />

where the velocity gradients are evaluated along the Lagrangian trajectories<br />

It is possible to write the solution of eq. (4.5) as<br />

∂tx = u(t, x) ; x(t0, r) = r (4.6)<br />

σ(t, x) = W(t, t0)σ(t0, r)W T (t, t0)e −(t−t 0 )<br />

τ (4.7)<br />

63


64 4. Small-scale statistics of <strong>viscoelastic</strong> <strong>turbulence</strong><br />

The Lagrangian mapping matrix W , defined by:<br />

∂tW(t, t0) = (∇u) T (t)W(t, t0); W(t0, t0) = 1 (4.8)<br />

describes the deformation of an infinitesimal fluid element along a Lagrangian<br />

trajectory. The meaning of eq. (4.7) is clear: polymers are advected along the<br />

Lagrangian trajectories being stretched by the velocity gradient <strong>and</strong> relaxing due<br />

to their elasticity.<br />

The matrix W can be decomposed as<br />

W = MΛN (4.9)<br />

where M <strong>and</strong> N are orthogonal matrices <strong>and</strong> Λ is diagonal. At times much<br />

larger than the velocity gradients correlation time the main eigenvalue of Λ is<br />

much larger than all the others <strong>and</strong> the molecule tends to allineate to the stretching<br />

direction. Then the matrix σ has to be uniaxial:<br />

σij = SiSj<br />

<strong>and</strong> the <strong>viscoelastic</strong> fluid evolves according to eqs. (4.2)-(4.3), with S = R<br />

R0 .<br />

4.1.1 Numerical settings<br />

(4.10)<br />

In the following we will consider the statistics of stationary turbulent solutions of<br />

the <strong>viscoelastic</strong> model (4.2)-(4.3) integrated in a periodic box of size L = 2π at<br />

resolution 128 3 by means of a st<strong>and</strong>ard pseudo-spectral code for different values<br />

of the relaxation time τ.<br />

For each τ, a statistically stationary state is obtained by adding to (4.3) an<br />

external forcing term which acts on the largest scales by keeping their energy<br />

constant [75]. In stationary conditions the forcing injects energy with a mean rate<br />

ɛI which balances the dissipation (4.4), ɛI = ɛν + ɛτ.<br />

The turbulent regime of a <strong>viscoelastic</strong> flow is controlled by two dimensionless<br />

parameters, the Reynolds number Rλ = urmsλ/ν (where λ = urms/〈(∂xux) 2 〉 1/2<br />

is the Taylor microscale) <strong>and</strong> the Weissenberg number which is defined here as<br />

Wi = τ/τK, where τK = (ν/ɛν) 1/2 is the Kolmogorov time.<br />

As a reference run, we integrated the st<strong>and</strong>ard Navier-Stokes equation (4.3)<br />

with η = 0, for which we have Rλ � 87. In this <strong>Newtonian</strong> limit, we have also<br />

computed the Lagrangian Lyapunov exponent λL which is a measure of the mean<br />

stretching rate. The dimensionless number λLτη � 0.13 is consistent with known<br />

simulations [76]. The <strong>viscoelastic</strong> runs are performed for different values of the<br />

relaxation times corresponding to a Weissenberg number in the range 4.8 ≤ Wi ≤<br />

24 (i.e. 0.63 ≤ λLτ ≤ 3.14).<br />

64


4.2. Coil-stretch transition 65<br />

4.2 Coil-stretch transition<br />

Although the uniaxial model is derived in the limit of strong elongation, it displays<br />

a clear coil-stretch transition as a function of Wi. The behaviour of the fluid, at<br />

progressively increasing values of the Weissenberg number, varies as follows.<br />

p(R)<br />

10 -1<br />

10 -2<br />

10 -3<br />

10 -4<br />

10 -5<br />

10 -6<br />

<br />

40<br />

20<br />

0<br />

0 5 10 15 20 25 30<br />

R<br />

0 5 10 15 20 25<br />

Figure 4.1: Probability density functions of polymer elongations above the coilstretch<br />

transition at Wi = 9.5 (squares), Wi = 12 (circles) <strong>and</strong> Wi = 24 (triangles).<br />

For Wi < Wi ∗ = 6.5 the distribution is p(R) = δ(R). In the inset we show<br />

the growth of 〈R 2 〉 vs Wi.<br />

When Wi → 0, i.e. τ → 0, eq. (4.2) implies R = 0, so that the feedback term<br />

(2νη/τ)R · ∇R is negligible in (4.3) <strong>and</strong> the usual Navier-Stokes equation for a<br />

<strong>Newtonian</strong> fluid is recovered. Polymers are coiled <strong>and</strong> their pdf of elongation is<br />

p(R) = δ(R). This state persists until the mean stretching rate is comparable with<br />

the inverse relaxation time, λLτ ∼ 1 [39].<br />

For larger values of Wi, the relaxation time of polymers is larger than the<br />

smallest characteristic time of the turbulent flow <strong>and</strong> molecules start to be elongated.<br />

Figure 4.1 shows the pdf of elongations p(R) at different values of Wi<br />

together with the mean square elongation, from which the transition at Wi ∗ =<br />

6.5±0.5 (corresponding to λLτ = 0.86) is evident. We remark that because in the<br />

linear model (4.2) there is no maximum polymer extension [77], the stationarity<br />

65<br />

Wi


66 4. Small-scale statistics of <strong>viscoelastic</strong> <strong>turbulence</strong><br />

of the distribution p(R) (<strong>and</strong> its exponential tail) is entirely due to the feedback<br />

on the velocity field in eq. (4.3).<br />

4.3 Energy cascade<br />

A natural <strong>and</strong> relevant question arising in the study of <strong>viscoelastic</strong> <strong>turbulence</strong><br />

is how the energy cascade is modified by polymeric contributions, in the strong<br />

back-reaction regime defined by the condition Wi > Wi ∗ .<br />

According to the Lumley criterion [37, 46], as Wi increases above Wi ∗ , polymers<br />

start to affect the dynamics of the turbulent cascade at the scale ℓL at which<br />

the eddy turnover time is of the same order of the relaxation time. Assuming the<br />

K41 scaling δuℓ ∼ ɛ 1/3 ℓ 1/3 for the velocity increments, the Lumley scale ℓL can<br />

be estimated:<br />

τℓ ∼ ɛ −1/3 ℓ 2/3 ∼ τ ⇒ ℓL ∼ (ɛτ 3 ) 1/2<br />

(4.11)<br />

For scales ℓ > ℓL we expect that the turbulent cascade is unaffected by polymers,<br />

i.e. with constant energy flux equal to the energy input ɛI, while for ℓ < ℓL we<br />

expect a reduced energy flux ɛ = ɛν < ɛI because below the Lumley scale part of<br />

the flux is removed by elastic dissipation.<br />

By increasing Wi above Wi ∗ two different scenarios are possible:<br />

• the first is that elastic dissipation in eq. (4.4) increases with Wi <strong>and</strong> energy<br />

flux at scales ℓ < ℓL vanishes, meaning that the Lumley scale ℓL becomes<br />

the new dissipative scale.<br />

• A second alternative is that elastic dissipation removes only a fixed fraction<br />

of the flux. In this case ɛν becomes independent of Wi <strong>and</strong> thus the energy<br />

cascade proceeds below ℓL, though with a reduced flux. This latter scenario<br />

has been observed in shell models of <strong>viscoelastic</strong> <strong>fluids</strong> [78].<br />

Our numerical simulations at increasing values of Wi indicate that the second<br />

scenario occurs. The inset of fig. 4.2 shows that the ratio ɛν/ɛI, which is by<br />

definition unity for Wi ≤ Wi ∗ , decreases for Wi > Wi ∗ but already at Wi �<br />

15 � 2Wi ∗ saturates to a new value ∼ 0.85. These results are in agreement with<br />

those reported in [70], where the reduction of vorticity in a <strong>viscoelastic</strong> solution<br />

was experimentally measured, <strong>and</strong> the Taylor microscale λ was observed to be<br />

practically unaffected by the presence of polymers.<br />

The above picture is supported by the comparison of the kinetic energy spectrum<br />

of a <strong>viscoelastic</strong> flow above the coil-stretch transition <strong>and</strong> the spectrum of<br />

the reference <strong>Newtonian</strong> flow. Figure 4.2 shows that, although the energy content<br />

at small scales is reduced by polymers, a power-law spectrum, characteristic of<br />

a turbulent cascade à la Kolmogorov, is present in the <strong>viscoelastic</strong> case as well.<br />

66


4.3. Energy cascade 67<br />

E K (k)/ε I 2/3<br />

1<br />

0.1<br />

0.01<br />

0.001<br />

ε ν /ε I<br />

1<br />

0.9<br />

0.8<br />

0 5 10 15 20 25<br />

Wi<br />

D<br />

1 10<br />

k<br />

0.25<br />

0.2<br />

0 5 10 15 20 25<br />

Wi<br />

Figure 4.2: <strong>Newtonian</strong> (squares) <strong>and</strong> <strong>viscoelastic</strong> (Wi = 14.3, circles) spectra of<br />

kinetic energy, normalized with ɛ 2/3<br />

I . The dotted line corresponds to Kolmogorov<br />

K41 scaling EK(k) ∼ k −5/3 . Lower inset: viscous dissipation ɛν normalized<br />

to the energy input ɛI as a function of Wi; the point size is of the order of the<br />

versus Wi.<br />

statistical uncertainty. Upper inset: drag coefficient D = ɛIL/E 3/2<br />

K<br />

Moreover the effect of polymers on <strong>turbulence</strong> is local in scales, i.e. scales larger<br />

than ℓL are essentially not affected by the presence of polymers.<br />

Since the total kinetic energy EK is dominated by large scales, we do not<br />

observe a significant variation of the "drag" coefficient. As is customary in homogeneous<br />

isotropic configurations, the drag coefficient is here defined:<br />

D = ɛIL/E 3/2<br />

K<br />

(4.12)<br />

The inset in fig. 4.2 shows that from Wi = 0 to Wi = 24 we measure fluctuations<br />

of the drag of about 2% which are within the statistical uncertainty. This is at<br />

variance with the results reported in [71], where, however, it must be noticed<br />

that a different <strong>viscoelastic</strong> model was used, with a different large scale forcing<br />

mechanism.<br />

67


68 4. Small-scale statistics of <strong>viscoelastic</strong> <strong>turbulence</strong><br />

4.4 Acceleration statistics<br />

Since the turbulent cascade survives at scales smaller than ℓL, we expect to observe<br />

some features of <strong>Newtonian</strong> <strong>turbulence</strong> in <strong>viscoelastic</strong> flows. One of the<br />

usual characteristics of small-scale <strong>turbulence</strong> is the highly intermittent acceleration,<br />

which displays fluctuations much larger than the rms value, accounting for<br />

large probabilities of extreme events (see sec. 1.2.4 <strong>and</strong> references therein).<br />

In the following we report numerical results about the statistics of acceleration<br />

in <strong>viscoelastic</strong> turbulent flows.<br />

4.4.1 Fluctuations of <strong>viscoelastic</strong> accelerations<br />

The numerically computed acceleration displays an intermittent behaviour, with<br />

fluctuations up to 15 times the root mean square value, as shown in fig. 4.3, where<br />

its probability distribution at different values of elasticity above Wi ∗ is plotted,<br />

together with the corresponding <strong>Newtonian</strong> case for comparison.<br />

In presence of polymers, the rms value arms is reduced with respect to the<br />

<strong>Newtonian</strong> case (inset of fig. 4.3), again in agreement with experimental observations<br />

[79]. For sufficiently large values of the Weissenberg number, arms becomes<br />

almost independent of Wi <strong>and</strong>, at the largest value Wi = 24, it is about 80%<br />

of the <strong>Newtonian</strong> case. This is consistent with a reduced value of energy flux at<br />

small scales shown in fig. 4.2: indeed by compensating arms with the dimensional<br />

estimation ɛ 3/4<br />

ν ν −1/4 this becomes virtually independent of Wi (the fluctuations<br />

being of about 5%). This implies that, at least in the present range of parameters,<br />

the Heisenberg-Yaglom relation a 2 ∼ ν −1/2 ɛ 3/2 holds in the <strong>viscoelastic</strong> case as<br />

well, provided ɛ is substituted with ɛν.<br />

The probability density functions shown in fig. 4.3 indicate that relative fluctuations<br />

of turbulent acceleration are not affected by the presence of polymers.<br />

Indeed, the pdf of the rescaled quantity a/arms is found to be Wi- independent in<br />

all the range of Wi investigated.<br />

This is a remarkable result which, besides its intrinsic interest, has an important<br />

practical consequence: the fact that the shape of the acceleration pdf is not<br />

affected by polymers implies that it is possible to model small-scale statistics in<br />

<strong>viscoelastic</strong> <strong>turbulence</strong> below the Lumley scale by means of the same models used<br />

for <strong>Newtonian</strong> <strong>fluids</strong>, by simply changing global quantities such as the energy flux.<br />

68


4.4. Acceleration statistics 69<br />

P(a/a rms )<br />

1<br />

0.1<br />

10 -2<br />

10 -2<br />

10 -2<br />

10 -2<br />

10 -3<br />

10 -3<br />

10 -3<br />

10 -3<br />

10 -4<br />

10 -4<br />

10 -4<br />

10 -4<br />

10 -5<br />

10 -5<br />

10 -5<br />

10 -5<br />

1.6<br />

1.4<br />

1.2<br />

0 10 20<br />

Wi<br />

-15 -10 -5 0 5 10 15<br />

a/a rms<br />

Figure 4.3: Acceleration probability distributions for Wi = 0, 12, 24; the inner<br />

dotted curve shows a Gaussian for comparison. Inset: total arms, normalized with<br />

ɛ 3/4<br />

I ν−1/4 , vs Wi.<br />

4.4.2 Role of the back-reaction<br />

In a <strong>viscoelastic</strong> flow the acceleration is, by definition, the right-h<strong>and</strong> side of<br />

eq. (4.3):<br />

a ≡ du<br />

2ν<br />

= −∇p + ν∆u + R · ∇R (4.13)<br />

dt τ<br />

where R has been made dimensionless by rescaling with the equilibrium length<br />

R0 <strong>and</strong> η = 1. From this expression it is evident that a is determined by three<br />

different contributions: pressure gradients, viscous <strong>and</strong> elastic contributions. As<br />

in the <strong>Newtonian</strong> case, the forcing term, acting on the largest scales, can be disregarded.<br />

In fully developed <strong>turbulence</strong> the viscous contribution is negligible <strong>and</strong> one<br />

may ask which is the dominant contribution in the strong feedback regime (Wi ><br />

Wi∗ ) of the <strong>viscoelastic</strong> flow. Figure 4.4 (inset) shows that the pressure gradient<br />

contribution<br />

ap = 〈(∇p) 2 〉 1/2<br />

(4.14)<br />

which is the only term in the <strong>Newtonian</strong> limit Wi = 0, is always dominant in<br />

the range of Wi investigated. Nevertheless, the contribution of polymers is not<br />

69


70 4. Small-scale statistics of <strong>viscoelastic</strong> <strong>turbulence</strong><br />

p(cos(θ))<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

-1 -0.5 0 0.5 1<br />

1<br />

0.5<br />

0<br />

cos(θ)<br />

0 5 10 15 20 25<br />

Figure 4.4: Probability density functions of the cosine of the angle θ between<br />

the vectors −∇p <strong>and</strong> R · ∇R for Wi = 9.5 (squares), Wi = 14.3 (circles)<br />

<strong>and</strong> Wi = 24.0 (triangles). Inset: Leading contributions to the acceleration, normalized<br />

with the total rms acceleration, as a function of Wi: pressure gradient<br />

ap/arms (squares) <strong>and</strong> elastic stress contribution ael/arms (circles).<br />

negligible: at the largest Wi = 24 the rms value of the elastic acceleration<br />

ael = 2ν<br />

τ 〈(R · ∇R)2 〉 1/2<br />

Wi<br />

is almost 50% of the total acceleration arms.<br />

According to the above observations, we therefore have:<br />

(4.15)<br />

a � −∇p + 2ν<br />

R · ∇R (4.16)<br />

τ<br />

The fact that ap � arms means that, increasing Wi, the flow develops strong correlations<br />

between the pressure gradient <strong>and</strong> the elastic component in eq. (4.13). As<br />

a measure of this correlation we have computed the angle θ between the vectors<br />

−∇p <strong>and</strong> R · ∇R. Figure 4.4 displays the pdf of cos(θ) <strong>and</strong> shows that, indeed,<br />

increasing Wi, the two vectors tend to be anticorrelated (i.e. cos(θ) = −1)<br />

with higher <strong>and</strong> higher probability. This anticorrelation means that pressure gradients<br />

<strong>and</strong> back-reaction somehow act in opposite directions in the build up of the<br />

70


4.5. Summary 71<br />

acceleration, <strong>and</strong> polymers provide a partial balancing of strong fluctuations of<br />

pressure.<br />

4.5 Summary<br />

In this chapter it has been presented a study on the small-scale statistics of turbulent<br />

polymer solutions within the uniaxial model of <strong>viscoelastic</strong> flow.<br />

We have found numerical evidence for a coil-stretch transition at Wi ∗ above<br />

which polymers affect the small scales of the turbulent flow. At large scales the<br />

velocity field has been found unaffected by the presence of polymers.<br />

For Wi > Wi ∗ , the energy flux is partially removed by polymer elasticity<br />

at the Lumley scale. This effect saturates <strong>and</strong> the turbulent cascade proceeds to<br />

scales smaller than the Lumley scale.<br />

As a consequence, small-scale statistics display features typical of <strong>Newtonian</strong><br />

<strong>turbulence</strong>. In particular, the acceleration of the <strong>viscoelastic</strong> fluid is reduced with<br />

respect to the <strong>Newtonian</strong> case, but its rescaled probability distribution is left unchanged.<br />

We remark that the above results have been obtained within the linear uniaxial<br />

model which is very attractive both analytically <strong>and</strong> numerically for its simplicity.<br />

Yet, it is worth reminding that such model has some limitations, e. g. it does<br />

not allow to capture possible feedback effects in the coiled state. Therefore, it<br />

would be extremely interesting to compare our findings with the outcome of more<br />

realistic <strong>viscoelastic</strong> models.<br />

71


Chapter 5<br />

Elastic <strong>turbulence</strong> in<br />

two-dimensional flows<br />

It is known that the presence of polymers can change the stability diagram of laminar<br />

flows [80, 81], due to elastic instabilities absent in purely <strong>Newtonian</strong> <strong>fluids</strong><br />

[82, 83, 84].<br />

A notable effect recently observed experimentally is the onset of elastic <strong>turbulence</strong><br />

(see sec. 2.4) in the limit of very low Reynolds numbers, provided elasticity<br />

is large enough [55]. In this regime, the polymer solution displays features typical<br />

of turbulent flows (e. g. a broad range of active scales). Consequently, elastic<br />

<strong>turbulence</strong> has been proposed as an efficient technique for mixing in very low Re<br />

flows as, e. g., microchannel flows [56, 58, 85].<br />

Despite its wide technological interest, elastic <strong>turbulence</strong> is still poorly understood<br />

from a theoretical point of view; recent predictions are based on simplified<br />

versions of <strong>viscoelastic</strong> models <strong>and</strong> on the analogy with MHD equations [39, 74].<br />

This chapter reports an investigation of elastic <strong>turbulence</strong> in the <strong>viscoelastic</strong><br />

Kolmogorov flow, in two dimensions. The basic phenomenology observed in<br />

laboratory experiments is reproduced by means of numerical simulations.<br />

At sufficiently high Weissenberg numbers, the resistance of the low-Reynolds<br />

flow is found to significantly increase, in close analogy with what happens in<br />

<strong>Newtonian</strong> <strong>turbulence</strong>. Above a threshold elasticity, a transition to chaotic states is<br />

indeed observed <strong>and</strong> characterized by means of Eulerian Lyapunov exponents; at<br />

intermediate Wi, a regime of elastic waves is detected. Power spectra of velocity<br />

fluctuations display a wide range of excited scales when elasticity is large enough,<br />

which confirms the onset of a chaotic state with many degrees of freedom, closely<br />

resembling usual fluid <strong>turbulence</strong>. Mixing properties are discussed in terms of<br />

Lagrangian Lyapunov exponents.<br />

73


74 5. Elastic <strong>turbulence</strong> in two-dimensional flows<br />

5.1 Viscoelastic model<br />

The dynamics of the dilute polymer solution can be conveniently described by the<br />

Oldroyd-B <strong>viscoelastic</strong> model (sec. 2.2.1):<br />

∂tu + (u · ∇)u = −∇p + ν∆u +<br />

2η ν<br />

∇ · σ + f (5.1)<br />

τ<br />

∂tσ + (u · ∇)σ = (∇u) T 2(σ − 1)<br />

· σ + σ · (∇u) − (5.2)<br />

τ<br />

where u is an incompressible, two-dimensional, velocity field <strong>and</strong> we recall that<br />

the symmetric positive definite matrix σ is the conformation tensor of polymer<br />

molecules, its trace trσ being a measure of their square elongation.<br />

In order to avoid complications induced by boundary conditions, we limit in<br />

the present study to the configuration of the periodic Kolmogorov flow [86]. Another<br />

advantage of using Kolmogorov flow is that its stability properties are known<br />

for both the <strong>Newtonian</strong> [87] <strong>and</strong> the <strong>viscoelastic</strong> [84] case. With the constant<br />

forcing f = (F cos(y/L), 0), the system of equations (5.1)-(5.2) has a laminar<br />

Kolmogorov fixed point given by<br />

u = (U0 cos(y/L), 0) (5.3)<br />

�<br />

2 U<br />

1 + τ<br />

σ =<br />

2 0<br />

2L2 sin2 (y/L) −τ U0<br />

�<br />

sin (y/L)<br />

2L (5.4)<br />

sin (y/L) 1<br />

−τ U0<br />

2L<br />

with F = [νU0(1 + η)]/L 2 [84].<br />

The laminar flow fixes characteristic scales of length L, velocity U0 <strong>and</strong> time<br />

T = L/U0. In terms of these variables, the Reynolds <strong>and</strong> Weissenberg numbers<br />

are, respectively, defined as:<br />

Re ≡<br />

Wi ≡ τU0<br />

L<br />

U0L<br />

ν(1 + η)<br />

The ratio of these numbers defines the elasticity of the flow El ≡ Wi/Re.<br />

5.1.1 Elastic instabilities<br />

(5.5)<br />

(5.6)<br />

A remarkable feature of <strong>viscoelastic</strong> <strong>fluids</strong> is the existence of purely elastic instabilities.<br />

In particular, it has been found that elastic stresses nonlinearly depending<br />

on the fluid velocity can destabilize the flow of a polymer solution [80, 83].<br />

In the present case, the Kolmogorov flow is unstable with respect to largescale<br />

perturbations, i. e. with wavelenght much larger than L. In the <strong>Newtonian</strong><br />

74


5.1. Viscoelastic model 75<br />

Figure 5.1: Stability diagram of the <strong>viscoelastic</strong> Kolmogorov flow, as predicted by<br />

multiscale analysis (solid line) <strong>and</strong> computed by numerical solution of the exact<br />

linearized equations (triangles). In the region denoted by U the flow is unstable,<br />

in that denoted by S it is stable. Inside the area denoted by CSL the flow is stable<br />

with respect to large-scale perturbations, but not with respect to generic ones [84].<br />

case, the instability arises at Rec = √ 2 [87]. For the <strong>viscoelastic</strong> case, recent<br />

analytical <strong>and</strong> numerical investigations have found the complete stability diagram<br />

in the Re-Wi plane [84] (see fig. 5.1).<br />

We recall that linear stability analysis shows that for sufficiently large values<br />

of elasticity, the Kolmogorov flow displays purely elastic instabilities, even<br />

at vanishing Reynolds numbers. The idea of elastic <strong>turbulence</strong> comes from the<br />

observation that, above the elastic instability threshold, the flow can develop a<br />

disordered secondary flow which persists in the limit of vanishing Re [57].<br />

5.1.2 Numerical setup<br />

We have numerically integrated the evolution equations for the vorticity ω, which<br />

can be derived by taking the curl of eq. (5.1), <strong>and</strong> for the conformation tensor<br />

σ, eq. (5.2), with periodic boundary conditions, by means of a pseudo-spectral<br />

method at resolution 512 2 .<br />

It is well known that direct numerical integration of <strong>viscoelastic</strong> models is<br />

limited by instabilities associated with the loss of positiveness of the conformation<br />

tensor [88]. These instabilities are particularly important at high elasticity<br />

<strong>and</strong> limit the possibility to investigate the elastic turbulent regime by direct integration<br />

of equations (5.1)-(5.2). To overcome this problem, we have implemented<br />

an algorithm based on a Cholesky decomposition of the conformation matrix that<br />

ensures symmetry <strong>and</strong> positive definiteness [89]. In extreme synthesis, this con-<br />

75


76 5. Elastic <strong>turbulence</strong> in two-dimensional flows<br />

sists in writing σ as the product of a lower triangular matrix with its transpose:<br />

σ = LL T<br />

(5.7)<br />

which guarantees, by construction, the positive definiteness of σ, <strong>and</strong> in deriving<br />

a transport equation for the tensor L. In 2D the equations of motion for the three<br />

nonzero elements of L can be derived from those for the elements of σ sequentially,<br />

starting with ℓ11 = √ σ11 <strong>and</strong> then proceeding to ℓ21 <strong>and</strong> ℓ22. The matrix σ<br />

will remain positive definite as long as the diagonal elements of L are greater than<br />

zero. A possibility to satisfy this constraint, is to perform a logarithmic transformation<br />

for the diagonal elements:<br />

˜ℓii = ln(ℓii) ; ˜ ℓij = ℓij if i �= j (5.8)<br />

Exponentiation, after numerical integration of ˜ ℓij, ensures positive definite diagonal<br />

elements.<br />

To further control numerical instabilities, the simulations have been performed<br />

adding a small diffusivity κ for polymers to eq. (5.2), the corresponding Schmidt<br />

number Sc ≡ ν/κ being always greater than 5 × 10 2 .<br />

5.2 Momentum budget<br />

One of the first indications of a strongly nonlinear state in the flow is the growth of<br />

its resistance, compared to the laminar flow at the same Re. This can be quantified<br />

in terms of the total stress Πr+Πp, where Πr = 〈uxuy〉 is the usual Reynolds stress<br />

<strong>and</strong> Πp = 2νητ −1 〈σxy〉 is the polymeric stress. Together with the viscous stress<br />

Πν = ν∂y〈ux〉 they contribute to the momentum budget in eq. (5.1).<br />

In the laminar fixed point (in which Πr = 0) the momentum budget gives for<br />

the amplitudes FL = νU0/L + (2νη/τ)Σ0, with Σ0 = τU0/(2L) the amplitude<br />

of σxy at the fixed point (5.4).<br />

In the turbulent state, we have that mean velocity <strong>and</strong> polymer conformation<br />

tensor are accurately described by sinusoidal profiles: 〈ux〉 = U cos(y/L) <strong>and</strong><br />

〈σxy〉 = −Σ sin(y/L) [53]. The ratio of the total turbulent stress Πr + Πp to the<br />

laminar stress Πlam is then expressed in terms of the ratio of amplitudes<br />

r = Π<br />

Πlam<br />

= U2(1 + η)<br />

FLη<br />

+ 2ν(1 + η)Σ<br />

τFL<br />

(5.9)<br />

where the first term comes from the Reynolds stress <strong>and</strong> the second from the elastic<br />

stress. Figure 5.2 shows the behaviour of the relative stress r as a function of<br />

the Weissenberg number Wi. We see that at Wi � 10 there is a transition from the<br />

laminar regime to a "turbulent" regime in which Π > Πlam. This turbulent stress<br />

76


5.3. Transition to <strong>turbulence</strong> 77<br />

r<br />

2.2<br />

2<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0 10 20 30 40 50 60 70<br />

Figure 5.2: Relative turbulent stress as a function of Wi for a set of simulations<br />

with ν = 0.769, η = 0.3, L = 1/4 <strong>and</strong> τ = 4. The elasticity is El = 64 <strong>and</strong> the<br />

maximum Reynolds number is Re = 1.<br />

originates from elasticity, since the contribution of the first term in eq. (5.9) always<br />

remains smaller than 10 −2 . This is the hallmark of elastic <strong>turbulence</strong>, where<br />

the elastic stress has the role played by the Reynolds stress in usual <strong>Newtonian</strong><br />

<strong>turbulence</strong>.<br />

Let us observe that, despite the qualitative behaviour is the same, this result is<br />

not in quantitative agreement with the measurements reported in [55]. However, it<br />

should be noticed that a different flow geometry, namely a swirling flow between<br />

two parallel disks, in three dimensions, was considered in that case.<br />

Wi<br />

5.3 Transition to <strong>turbulence</strong><br />

In order to get more insight in the flow properties above the elastic instability<br />

threshold, we now proceed to an analysis of the transition towards the turbulentlike<br />

states, presenting the main numerical evidences.<br />

Inspection of fig. 5.3 immediately allows to visualize the progressive destabilization<br />

of the flow at growing elasticity. Two snapshots of the two-dimensional<br />

vorticity field at two different Wi are shown. The first simulation at Wi = 16 is<br />

slightly above the elastic instability threshold. The flow in this regime is still not<br />

fully turbulent (i. e. the energy spectrum is exponential) <strong>and</strong> a secondary flow in<br />

the form of thin filaments is clearly observable on the underlying periodic structure<br />

of the basic flow. At higher <strong>and</strong> higher values of elasticity the vorticity pattern<br />

77


78 5. Elastic <strong>turbulence</strong> in two-dimensional flows<br />

becomes more <strong>and</strong> more irregular in space <strong>and</strong> time, with chaotic motion of filaments.<br />

At Wi = 64 we observe a largely disordered pattern in which the basic<br />

flow is hardly distinguishable.<br />

Figure 5.3: Snapshots of the vorticity field at Wi = 16 (left) <strong>and</strong> Wi = 64 (right).<br />

Colors code intensity.<br />

5.3.1 Eulerian Lyapunov exponent<br />

The onset of a chaotic dynamics can be highlighted by measuring the Eulerian<br />

Lyapunov exponent λE. This is the average logarithmic rate of separation of two<br />

realizations of the time-evolving fields, initially differing by a small perturbation.<br />

In order to measure this quantity we consider the dynamical system defined<br />

by the equations of motion for the fields ω <strong>and</strong> ˜ ℓij. The distance between the two<br />

realizations is computed according to the norm<br />

∆(t) = 1<br />

2<br />

�<br />

d 2 �<br />

x |δω(x, t)| 2 + |δ˜ ℓ11(x, t)| 2 + |δ˜ ℓ12(x, t)| 2 + |δ˜ ℓ22(x, t)| 2<br />

�<br />

where the difference fields are defined as<br />

δω(x, t) = 1 � �<br />

(2) (1) √ ω (x, t) − ω (x, t)<br />

2<br />

δ˜ ℓij(x, t) = 1<br />

�<br />

√ ˜ℓ (2)<br />

2<br />

ij (x, t) − ˜ ℓ (1)<br />

ij<br />

�<br />

(x, t)<br />

(5.10)<br />

(5.11)<br />

(5.12)<br />

The behaviour of λE at fixed Re = 1 <strong>and</strong> varying Weissenberg number is<br />

shown in fig. 5.4. Above Wi � (10 ÷ 20), the Eulerian Lyapunov exponent be-<br />

78


5.3. Transition to <strong>turbulence</strong> 79<br />

λ E<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

-0.2<br />

0 50 100 150 200 250<br />

Figure 5.4: Eulerian Lyapunov exponent as a function of the Weissenberg number.<br />

comes positive, meaning that the nonlinearities are producing a chaotic evolution;<br />

for larger elasticities λE monotonically grows with Wi.<br />

Wi<br />

5.3.2 Intermediate regime: elastic waves<br />

Close to the elastic instability, as already mentioned, a secondary flow appears<br />

in the form of thin filaments superimposed to the basic flow. These small scale<br />

structures, present in all the fields ω <strong>and</strong> σij evolved according to the Oldroyd-B<br />

model (5.1)-(5.2), move along the x direction (see figures 5.5, 5.6).<br />

They are elastic waves reminiscent of Alfvén waves propagating in presence of<br />

a large scale magnetic field in a plasma. The possibility to observe elastic waves in<br />

polymer solutions was theoretically predicted within the simplified uniaxial model<br />

[74], but these had never been observed before.<br />

The filaments are most likely produced by the stretching of polymers in the<br />

areas of maximal shear rate of the flow, corresponding to the zeroes of the cosine<br />

velocity profile, <strong>and</strong> then transported at a velocity comparable with that of the<br />

mean flow, which "pulls" their leading edge, as it is clearly seen in fig. 5.5.<br />

When the Weissenberg number grows above approximately Wi = 20, the increasing<br />

complexity of the flow patterns suggests nontrivial interactions between<br />

the waves, which are not anymore passive objects transported by the flow field.<br />

The filaments are less <strong>and</strong> less well defined in this range of Wi, until a spatially<br />

disordered state sets in, as shown in fig. 5.3 for Wi = 64.<br />

79


80 5. Elastic <strong>turbulence</strong> in two-dimensional flows<br />

(a) (b) (c) (d)<br />

Figure 5.5: Snapshots of the vorticity field ω [(a)] <strong>and</strong> the component σ11 of the<br />

conformation tensor [(c)] at Re = 1, Wi = 16, together with corresponding<br />

plots [(b), (d)] of the space-time traces of waves moving along the horizontal<br />

direction x. The waves are selected from ω <strong>and</strong> σ11 by fixing a vertical coordinate<br />

y; here yω = 3/8 <strong>and</strong> yσ11 = 7/16 (in units of the box size). Time is plotted<br />

versus the position of the wave along the horizontal direction; due to periodic<br />

boundary conditions, waves leaving the box on the left re-enter on the right. Lines<br />

of constant slope mean that the wave velocity is constant. Black <strong>and</strong> red colors<br />

code minimum <strong>and</strong> maximum intensities, respectively.<br />

-2<br />

0<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

z<br />

0 50 100 150<br />

t<br />

200 250 300 350 0<br />

600<br />

500<br />

400<br />

300<br />

200 x<br />

100<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

z<br />

0 50 100 150<br />

t<br />

200 250 300 350 0<br />

600<br />

500<br />

400<br />

300<br />

200 x<br />

100<br />

Figure 5.6: Wave propagation: peaks of intensity from the space-time traces<br />

shown in fig. 5.5 at three succesive later times for ω (left) <strong>and</strong> σ11 (right)<br />

.<br />

5.3.3 Turbulent energy spectrum<br />

In the regime of highest Weissenberg numbers, the flow develops active modes at<br />

all the scales. Figure 5.7 shows power spectra of velocity fluctuations averaged<br />

over several configurations such as the one shown in fig. 5.3.<br />

A power law behaviour E(k) ∼ k −α is clearly observable with spectral exponent<br />

α slightly larger than 3. This result constitutes a striking evidence of the<br />

80


5.4. Mixing: Lagrangian Lyapunov exponent 81<br />

emergence of turbulent features in the chaotic flow, <strong>and</strong> it is in reasonable agreement<br />

with laboratory experiments [55, 57], where an exponent close to 3.5 is<br />

found, <strong>and</strong> with the theoretical predictions based on the simplified uniaxial model<br />

[74].<br />

E(k)<br />

10 -2<br />

10 -4<br />

10 -6<br />

10 -8<br />

10 100<br />

k<br />

Wi=128<br />

Wi=64<br />

Figure 5.7: Velocity fluctuation spectra at Wi = 64 <strong>and</strong> Wi = 128.<br />

5.4 Mixing: Lagrangian Lyapunov exponent<br />

One of the most promising applications of elastic <strong>turbulence</strong> is for efficient mixing<br />

at very low Reynolds number, which is of paramount importance in many<br />

industrial problems such as, for instance, the engineering of microfluidic devices.<br />

Indeed, laboratory experiments in a curvilinear channel have demonstrated that<br />

very viscous polymer solutions in the elastic <strong>turbulence</strong> regime are very efficient<br />

for small scale mixing [56]. Mixing efficiency of polymer solutions has been<br />

also studied in other setups, including microchannels [85] <strong>and</strong> two-dimensional,<br />

magnetically driven, flows [58].<br />

Since in the elastic <strong>turbulence</strong> regime the flow is smooth, i. e. the energy spectrum<br />

is steeper that k −3 , a suitable characterization of mixing is given in terms of<br />

the Lagrangian Lyapunov exponent λL [90]. This is defined as the mean rate of<br />

separation of two infinitesimally close particles transported by the flow.<br />

Figure 5.8 shows the behaviour of the Lyapunov exponent as a function of Wi<br />

at fixed Re = 1. We observe that λL grows following approximately a power law<br />

λL ∼ Wi 0.31 in a rather large range of Wi. We remark that this behaviour is in<br />

some way opposite to that at high Reynolds numbers, where it has been found<br />

that the injection of polymers reduces the degree of chaoticity to λL < 1/τ [40].<br />

81<br />

k -3


82 5. Elastic <strong>turbulence</strong> in two-dimensional flows<br />

λ L<br />

0.5<br />

0.4<br />

0.3<br />

50<br />

6<br />

4<br />

2<br />

0<br />

100<br />

Wi<br />

-2<br />

G(γ)<br />

0<br />

γ<br />

200<br />

2<br />

4<br />

400<br />

Figure 5.8: Lagrangian Lyapunov exponent at varying Weissenberg number<br />

(Re = 1); the straight dashed line has slope 0.31. Inset: Cramer function.<br />

Of course this cannot be true in the present situation as λL = 0 in absence of<br />

polymers <strong>and</strong> thus, at least in some range, chaoticity must increases with Wi.<br />

In the inset of fig. 5.8 we also plot the Cramer function G(γ) which is defined<br />

from the probability density function of Lyapunov exponent fluctuations<br />

Pt(γ) ∼ exp(−tG(γ)) [90]. As it is evident, increasing Wi, not only the degree<br />

of mean chaoticity grows, but fluctuations become larger. In particular the<br />

distribution of γ becomes asymmetric with a larger relative probability of positive<br />

fluctuations. It is remarkable that the same qualitative behaviour is observed in the<br />

case of high-Reynolds <strong>Newtonian</strong> <strong>turbulence</strong>, where the distribution of Lyapunov<br />

fluctuations becomes progressively more asymmetric with increasing Re. Again,<br />

this is another indication that, in elastic <strong>turbulence</strong>, elasticity (i. e. Wi) has the<br />

same role of non-linearity (i. e. Re) of usual fluid <strong>turbulence</strong>.<br />

5.5 Summary<br />

In this chapter it has been reported a study on elastic <strong>turbulence</strong> in a two-dimensional<br />

Kolmogorov flow with periodic boundary conditions. To our knowledge, this<br />

is the first observation of such a phenomenon by means of numerical simulations.<br />

We remark that the onset of elastic <strong>turbulence</strong> in this simple configuration extends<br />

previous results, in which curvilinear flows were used [55, 57], to the case of<br />

straight streamlines, <strong>and</strong> highlights the irrelevance of material boundaries for the<br />

problem.<br />

82


5.5. Summary 83<br />

We found evidence of a significant increase in the flow resistance, above the<br />

elastic instability threshold, which resembles the behaviour of usual <strong>Newtonian</strong><br />

<strong>turbulence</strong>.<br />

It has been investigated the transition from laminar to chaotic states happening<br />

at sufficiently high elasticity in terms of Eulerian Lyapunov exponents, <strong>and</strong> an<br />

intermediate regime has been detected, in which elastic waves appear, before the<br />

onset of turbulent-like features. In the turbulent regime the flow exhibits a wide<br />

range of excited scales <strong>and</strong> a power law spectrum of velocity fluctuations. The<br />

value of the scaling exponent indicates that the flow is smooth even at the highest<br />

Weissenberg numbers.<br />

To characterize mixing properties, we also computed the Lagrangian Lyapunov<br />

exponent, which was found to grow with Wi according to a power law<br />

behaviour.<br />

It would be very interesting to extend the work to a more realistic threedimensional<br />

setup, in order to better compare with experimental results <strong>and</strong> gain<br />

more insight in the basic physical mechanisms underlying the phenomenon.<br />

83


Chapter 6<br />

Turbulence <strong>and</strong> coarsening in<br />

<strong>binary</strong> <strong>mixtures</strong><br />

Phase ordering dynamics becomes very complex when the fluid mixture is externally<br />

driven <strong>and</strong> the underst<strong>and</strong>ing of its phenomenonology is, in that case, still<br />

incomplete [62, 91, 92]. Beyond their theoretical interest, phase separating <strong>binary</strong><br />

<strong>fluids</strong> under flow embody a great technological interest [93] for their distinctive<br />

rheological properties.<br />

This problem has been extensively investigated in shear flows [63, 94, 95, 96,<br />

97, 98] where coarsening becomes highly anisotropic: the single-phase domain<br />

growth accelerates in the shear direction while in the transversal one the growth<br />

is arrested [63, 97, 98] or strongly slowed down [95, 96].<br />

Less clear is the case in which the mixture is stirred by a turbulent flow [93,<br />

99, 100, 101, 102, 103]. Here, phase separation may be completely suppressed<br />

[100, 101], or a dynamical steady state with domains of finite length <strong>and</strong> well<br />

defined phases may develop [62, 92, 102, 103]. A similar phenomenology has<br />

been experimentally observed in stirred immiscible <strong>fluids</strong> [104].<br />

This chapter reports the study of the phase ordering process in the presence of<br />

an external forcing acting on the velocity field, in two dimensions. For both active<br />

<strong>and</strong> passive <strong>mixtures</strong> it is found that, for sufficiently strong stirring, coarsening is<br />

arrested [105].<br />

The nature of the nonequilibrium steady state, characterized by the continuous<br />

rupture <strong>and</strong> formation of domains, is discussed. Previous investigations focused<br />

on passive <strong>binary</strong> <strong>mixtures</strong> (when the feedback of the phase ordering on fluid<br />

velocity is neglected) in r<strong>and</strong>om flows [106] <strong>and</strong> in chaotic flows [107] (in a Lagrangian<br />

sense, i. e. two initially very close particles separate exponentially in time<br />

[108]). Here we consider the phase ordering dynamics of active two-dimensional<br />

<strong>binary</strong> <strong>mixtures</strong> in which the fluid is driven by chemical potential inhomogeneities<br />

[62, 92].<br />

85


86 6. Turbulence <strong>and</strong> coarsening in <strong>binary</strong> <strong>mixtures</strong><br />

By means of direct numerical simulations we show that coarsening arrest is a<br />

generic <strong>and</strong> robust phenomenon, whose existence can be understood by an energy<br />

conservation argument.<br />

Moreover, we show that in the passive limit Lagrangian chaos is not necessary<br />

for coarsening arrest.<br />

6.1 Phase separation under stirring<br />

In the presence of stirring, the main question concerns the competition between<br />

thermodynamic forces, driving the phase segregation, <strong>and</strong> fluid motion, leading to<br />

mixing <strong>and</strong> the domain’s breakup.<br />

For very high flow intensities phase separation can be completely suppressed<br />

[99, 100, 101] due to mixing of the components <strong>and</strong> inhibition of interface formation.<br />

In active <strong>mixtures</strong> with very low viscosities such a phenomenon may be<br />

self-induced by the feedback [66, 68, 109]: the fluid responds vigorously to local<br />

chemical potential variations <strong>and</strong> remixes the components.<br />

On the other h<strong>and</strong>, stirring may lower the critical temperature [91, 102, 103,<br />

110]. However, by performing a deeper quench, phase separation in a nontrivial<br />

statistically stationary state may still develop [62, 92].<br />

6.1.1 Evolution equations<br />

Being interested in deep quenching into the coexistence region, here we work at<br />

zero temperature, as in references [106, 107]. Hence, thermal fluctuations of the<br />

local composition field will be neglected. We consider a symmetric (50% − 50%)<br />

mixture of two incompressible <strong>fluids</strong> of equal density ρ = 1 <strong>and</strong> viscosity ν.<br />

Such a bi-component system is described by a scalar order parameter θ(x, t),<br />

the local fraction of the two <strong>fluids</strong>. The associated L<strong>and</strong>au-Ginzburg free energy<br />

reads [60]:<br />

�<br />

F[θ]=<br />

�<br />

dx − 1<br />

2 θ2 + 1<br />

4 θ4 + ξ2<br />

2 |∇θ|2<br />

�<br />

, (6.1)<br />

being ξ the equilibrium correlation length, which provides a measure of the interface<br />

width. The dynamics is then governed by the Cahn-Hilliard (CH) equation<br />

∂tθ + v · ∇θ = Γ∆ δF<br />

δθ<br />

= Γ∆µ (6.2)<br />

where µ = −θ+θ 3 −ξ 2 ∆θ is the chemical potential, <strong>and</strong> Γ is a mobility coefficient<br />

that we assume constant <strong>and</strong> independent of θ.<br />

86


6.1. Phase separation under stirring 87<br />

Hydrodynamics enters in eq. (6.2) through the convective term. The order<br />

parameter is transported by the two-dimensional velocity field v, which evolves<br />

according to the Navier-Stokes (NS) equation:<br />

∂tv + v · ∇v = ν∆v − ∇p − θ∇µ + f , (6.3)<br />

where p is the pressure. The fluid is forced by the external mechanical force f <strong>and</strong><br />

by local chemical potential variations −θ∇µ. This latter term can be rewritten<br />

as −ξ 2 ∆θ∇θ plus a gradient term which can be absorbed into the pressure [62].<br />

Therefore eq. (6.3) formally reduces to the 2D magnetohydrodynamics equation<br />

for the velocity field. Actually phase ordering <strong>and</strong> MHD share many phenomenological<br />

properties [62].<br />

6.1.2 Numerical analysis<br />

We numerically integrate the coupled equations (6.2)-(6.3) by means of a st<strong>and</strong>ard<br />

pseudo-spectral code implemented on a two-dimensional periodic box of size 2π×<br />

2π with 5122 collocation points.<br />

Statistical analysis of domain sizes is obtained by considering the characteristic<br />

length, defined as<br />

L(t) ≡ 〈(1 − θ 2 )〉 −1<br />

(6.4)<br />

where 〈...〉 denotes a spatial average. Observe that this measure is constructed with<br />

the total contour length, since it picks values of the order parameter at the interface<br />

between the two bulk phases, where the argument (1 − θ2 ) is nonzero. Let us remark<br />

that measures based on the correlation function, C(r, t) = 〈θ(x, t)θ(x+r)〉,<br />

or on the spherically averaged wavenumber weighted with the structure factor,<br />

� R �<br />

dkkq −1/q<br />

S(k,t)<br />

i. e. L(t) = R (being S(k, t) = 〈| dkS(k,t)<br />

ˆ θ(k, t)| 2 〉|k|=k) give equivalent<br />

results. A discussion on the different definitions can be found in ref. [111]. At<br />

large times, finite size effects spoil the scaling laws.<br />

The initial condition for the order parameter is a high temperature configuration<br />

with θ set as white noise in space. In the presented results time is rescaled<br />

with the diffusive time tm = ξ2 /Γ.<br />

87


88 6. Turbulence <strong>and</strong> coarsening in <strong>binary</strong> <strong>mixtures</strong><br />

6.2 Active <strong>mixtures</strong><br />

At the highest level of complexity, the description of phase separating <strong>binary</strong> <strong>fluids</strong><br />

requires a full hydrodynamic treatment in terms of a transported active field able<br />

to react on the flow, as discussed above.<br />

In this section we report numerical results on the case of active <strong>mixtures</strong> <strong>and</strong><br />

derive the balance equation governing its statistically steady state in the presence<br />

of external stirring.<br />

6.2.1 Unstirred case<br />

Starting from the initial configuration with the fluid at rest (v = 0), after a few<br />

diffusive time scales tm, sharp interfaces appear <strong>and</strong> phase separation proceeds<br />

through domain coarsening. At long times, the domains length L - the only characteristic<br />

scale of the system, provided L ≫ ξ - grows in time as a power law<br />

[59, 60].<br />

In 2D different regimes have been predicted <strong>and</strong> observed [66]: L(t) ∼ t 1/3 , as<br />

in <strong>fluids</strong> at rest, for high viscosity; L(t) ∼ t 2/3 for lower viscosities. At intermediate<br />

values of viscosity it is still unclear whether there is only one characteristic<br />

scale [66]; for ν ≪ 1 <strong>and</strong> low mobility Γ ≪ 1 mixing may overwhelm phase<br />

demixing [68, 109].<br />

In the following we will limit our analysis to the turbulent, low viscous, regime<br />

where the scaling exponent 2/3 is expected. This exponent can be dimensionally<br />

derived by balancing the inertial term v · ∇v with θ∇µ in (6.3), <strong>and</strong> assuming<br />

that L(t) is the only length scale of the system.<br />

The scaling behaviour of L(t) implies the following ones for the kinetic energy<br />

<strong>and</strong> the enstrophy [65]:<br />

E = 〈v2 〉<br />

2<br />

Z = 〈ω2 〉<br />

2<br />

∼ t−2/3<br />

∼ t−5/3<br />

(6.5)<br />

(6.6)<br />

where ω = ∇ × v is the vorticity. Figure 6.1 shows that the scaling predictions<br />

are well reproduced by our DNS. We remark that in the absence of stirring, phase<br />

separation is accelerated by the presence of hydrodynamics.<br />

6.2.2 Stirred case<br />

We now consider the presence of an external mechanical forcing acting on the<br />

velocity field. As is customary in turbulent simulations, energy is injected by<br />

88


6.2. Active <strong>mixtures</strong> 89<br />

L(t)<br />

10 2<br />

10 1<br />

10 0<br />

(a)<br />

10 1<br />

t 2/3<br />

10 2<br />

t<br />

10 3<br />

10 4<br />

/2, /2<br />

Figure 6.1: (a) L(t) vs t obtained by DNS of (6.2) <strong>and</strong> (6.3) with ξ = 0.015,<br />

ν = 10 −3 , without external forcing. (b) Kinetic energy E = 〈v 2 〉/2 (bottom) <strong>and</strong><br />

enstrophy Z = 〈ω 2 〉/2 (top) vs t in the same run.<br />

means of a r<strong>and</strong>om, time uncorrelated, homogeneous, <strong>and</strong> isotropic Gaussian process<br />

with amplitude F which is restricted to a few Fourier modes around kf. The<br />

injection scale is, clearly, identified by ℓf ∼ 2π/kf. The δ correlation in time<br />

allows for controlling the kinetic energy input ɛin=F 2 nf, where nf is the number<br />

of excited Fourier modes.<br />

Equations (6.2)-(6.3) are integrated starting with v = 0. In fig. 6.2 we show<br />

typical snapshots of the order parameter at different flow intensities, <strong>and</strong> for two<br />

r<strong>and</strong>om forcing with ℓf ≈ 26ξ <strong>and</strong> ℓf ≈ 84ξ. As it is evident, the stronger the<br />

stirring intensity the smaller the typical domain length.<br />

This is confirmed by the temporal evolution of L(t) at varying the external<br />

forcing (fig. 6.3). After an initial growth characterized by the 2/3 scaling exponent,<br />

L(t) stabilizes at a value L ∗ that decreases with the stirring intensity. Both<br />

the kinetic energy E(t) <strong>and</strong> the enstrophy Z(t) (not shown here) stabilize at corresponding<br />

values. Therefore a well defined statistically steady state is reached.<br />

A closer inspection of fig. 6.2 reveals qualitative differences in domain shapes.<br />

When L ∗ is larger than the forcing scale ℓf (left) the domains are almost isotropic,<br />

while in the case L ∗ < ℓf (right) the underlying velocity field reveals itself through<br />

the filamental structure of the domains.<br />

Nevertheless, coarsening process is always arrested confirming the robustness<br />

of the phenomenon. Stirring always selects a scale through the competition between<br />

the thermodynamic forces <strong>and</strong> the stretching induced by local shears that<br />

deform <strong>and</strong> break the domains.<br />

Estimating the shear rate as γ = urms/L ∗ we find that, for the case with L ∗ <<br />

ℓf, L ∗ ∼ γ −0.29 (see inset of fig. 6.3), in fairly good agreement with experiments<br />

<strong>and</strong> simulations in pure shear flows [63, 98]. However we should mention that in<br />

our settings, homogeneous <strong>and</strong> isotropic flows, there is not a well defined rate γ<br />

89<br />

10 2<br />

10 0<br />

10 -2<br />

10 -4<br />

(b)<br />

10 1<br />

10 2<br />

t -5/3<br />

t -2/3<br />

t<br />

10 3<br />

10 4


90 6. Turbulence <strong>and</strong> coarsening in <strong>binary</strong> <strong>mixtures</strong><br />

Figure 6.2: Snapshots of θ at time t = 4000 at varying the forcing intensity F with<br />

ℓf =26ξ (left) <strong>and</strong> ℓf =84ξ (right). Black/white codes θ = ±1. Other parameters<br />

are as in fig. 6.1.<br />

as in genuine shear flows. The definition adopted here is a dimensional estimation<br />

of the shear rate at the arrest scale. In the case L ∗ ≥ ℓf no clear scaling behaviour<br />

is observed.<br />

The existence of a stationary state can be understood in terms of conservation<br />

laws. Due to the presence of two inviscid quadratic invariants, E <strong>and</strong> Z, the single<br />

fluid 2D NS equation (i. e. eq. (6.3) without the feedback term) is characterized by<br />

a double cascade [26]: E flows towards the large scales (r>ℓf) <strong>and</strong> Z towards the<br />

small ones (r


6.3. Passive <strong>mixtures</strong> 91<br />

L(t)<br />

10 1<br />

10 0<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

10 -2<br />

10 1<br />

L *<br />

10 -1<br />

u rms /L *<br />

10 0<br />

10 2<br />

Figure 6.3: L vs t at varying F , from top F = 0 (thick curve) <strong>and</strong> F =<br />

0.05, 0.10, 0.15, . . ., 0.30. Data refer to DNS with ℓf = 84ξ (the case with<br />

ℓf = 26ξ is qualitatively similar). The straight line displays the scaling t 2/3 .<br />

Inset: L ∗ vs urms/L ∗ ; the straight line has slope −0.29, point size is of the order<br />

of the statistical error.<br />

overcome the feedback term, the kinetic energy dissipation induced by the |∇µ| 2<br />

term is no more effective. Indeed, when ℓf is much larger than ξ <strong>and</strong> F is very<br />

high, the coupling term becomes negligible <strong>and</strong> we observe the single-fluid phenomenology<br />

with an inverse energy cascade.<br />

6.3 Passive <strong>mixtures</strong><br />

Let us now turn to the case in which the coupling term in eq. (6.3) is switched<br />

off <strong>and</strong> consequently the order parameter is passively transported by the velocity<br />

field. This case has been already considered in [106, 107]. In order to obtain<br />

a statistically stationary state, as customary we added a large scale friction term<br />

−αv to the Navier-Stokes equation [112].<br />

The velocity field in eq. (6.2) is rescaled by a factor β; this is a numerically<br />

convenient way to change the velocity intensity <strong>and</strong> to study the effect of stirring<br />

on coarsening. For β = 0 equation (6.2) recovers the Cahn-Hilliard equation in a<br />

fluid at rest for which L(t) ∼ t 1/3 .<br />

t<br />

91<br />

10 3


92 6. Turbulence <strong>and</strong> coarsening in <strong>binary</strong> <strong>mixtures</strong><br />

10 1<br />

L(t)<br />

10 0<br />

10 1<br />

10 2<br />

t<br />

10 3<br />

(a)<br />

10 4<br />

L(t)<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

6<br />

4<br />

2<br />

L *<br />

1 10<br />

β<br />

Figure 6.4: Results of DNS in the passive case. (a) L vs t with (from top) β =<br />

0.25, 0.5, 1, 2, 4. Forcing parameters are kf = 70 <strong>and</strong> F = 3 × 10 −5 . The straight<br />

line has slope 1/3. (b) Same as (a), in linear scale, for a frozen velocity field<br />

(see text), from top β = 1, 2, 4, 10. Inset: L ∗ vs β; the straight line corresponds to<br />

L ∗ ∼ β −0.28 , point size is of the order of the statistical error. DNS were performed<br />

with hyperdissipation −ν2∆ 2 v with ν2 = 10 −7 , <strong>and</strong> friction coefficient α = 0.1.<br />

The parameters of eq. (6.2) are ξ = 0.015, Γ = 0.02.<br />

For β > 0 we observe the following phenomenology [fig. 6.4(a)]. For small<br />

values of β (weak stirring) we did not find a clear evidence of coarsening arrest.<br />

This is likely due to finite size effects hiding the phenomenon, i. e. L ∗ becomes<br />

comparable with, or even larger than, the box size. For β large enough (strong<br />

stirring), the existence of an arrest scale L ∗ , that decreases with β, is well evident.<br />

6.3.1 Role of Lagrangian chaos<br />

Previous studies stressed the importance of Lagrangian chaos in the coarsening<br />

arrest phenomenon [107]. Now, in order to elucidate this point, we discuss a<br />

nonchaotic example.<br />

It is well known (see, e. g., [108]) that two-dimensional stationary flows do not<br />

generate chaotic trajectories. We have thus integrated eq. (6.2) in a frozen configuration<br />

of the turbulent velocity field: v(x, t) = v(x). As shown in fig. 6.4(b),<br />

domain growth is strongly weakened <strong>and</strong> finally arrested, even in this nonchaotic<br />

flow.<br />

For moderate velocity intensities, L(t) still grows in time, but with a much<br />

slower scaling law than the dimensional prediction for <strong>fluids</strong> at rest, t 1/3 . This<br />

slowing down is probably due to a different growth mechanism: after an initial<br />

transient, a slow process of droplet passage among close domains is indeed observed<br />

(see fig. 6.5). However, for high enough intensities a complete stabilization<br />

of the domain length is realized.<br />

92<br />

10 4<br />

t<br />

(b)<br />

2 10 4


6.3. Passive <strong>mixtures</strong> 93<br />

Figure 6.5: Snapshots at late times of a small portion of the composition field θ<br />

in a frozen turbulent velocity field; β = 4. Other parameters are as in fig. 6.4(b).<br />

Time grows from left to right with an interval of order 10 (in units of tm). The<br />

circle highlights a droplet being exchanged between two close domains.<br />

This suggests that the main ingredient for coarsening arrest is the presence of<br />

local shears that overwhelm the surface tension driving force. The dependence of<br />

L ∗ on the shear rate, here naturally defined as β, is shown in the inset of fig. 6.4(b).<br />

We find a power law behaviour with exponent −0.28 very close to the one observed<br />

in the active case <strong>and</strong> in shear flows [63, 98], while in the chaotic case no<br />

clear scaling is observed.<br />

To further support the marginal role of Lagrangian chaos in coarsening arrest,<br />

we report the results obtained in a stationary regular cellular flow:<br />

vx = U sin(Kx) cos(Ky) (6.8)<br />

vy = −U cos(Kx) sin(Ky) (6.9)<br />

where U fixes the velocity amplitude <strong>and</strong> K the characteristic scale.<br />

As shown in fig. 6.6 (left), for large intensities the order parameter is frozen<br />

into a r<strong>and</strong>om chessboard pattern with a finite length. At lower intensities a<br />

growth much slower than in the absence of the flow is still visible [fig. 6.6 (right)]<br />

coming from a slow droplet migration from one cell to another. At large U’s the<br />

shear between the counter-rotating vortexes overwhelms the demixing induced by<br />

the thermodynamic forces, breaking the domains which freeze into the cells.<br />

93


94 6. Turbulence <strong>and</strong> coarsening in <strong>binary</strong> <strong>mixtures</strong><br />

L(t)<br />

10<br />

10<br />

1<br />

10 1<br />

5<br />

1<br />

2 10 4 4 10 4 6 10 4 8 10 4 10 5<br />

Figure 6.6: (left) Snapshots of θ at t= 6×10 4 for the cellular flow with different<br />

U <strong>and</strong> K = 8, here ξ = 0.018, Γ = 0.1. (right) L vs t, from top to bottom<br />

U =0, 0.125, 0.25, 0.5, 1.0, 2.0, 4.0. Inset: the same in linear scale.<br />

6.4 Summary<br />

In this chapter it has been shown that in the presence of an external stirring the<br />

coarsening process is slowed down both for active <strong>and</strong> passive <strong>mixtures</strong>.<br />

The nonequilibrium steady state emerging at sufficiently strong intensities of<br />

the forcing acting on the velocity field has been explained by means of an energy<br />

conservation argument.<br />

We have also demonstrated that the phenomenon of coarsening arrest, first<br />

predicted in [62, 92], does not necessarily require a chaotic flow, as suggested in<br />

[107], but is a consequence of the competition between thermodynamic forces <strong>and</strong><br />

stretching induced by local shears. Our investigation in both active <strong>and</strong> passive<br />

<strong>mixtures</strong> shows that this behaviour is robust.<br />

Moreover we found numerical evidence that the dependence of the arrest scale<br />

on the shear rate follows a power law behaviour with an exponent close to the one<br />

measured in experiments <strong>and</strong> numerical simulations in pure shear flows [63, 98].<br />

Our results might suggest the existence of a mechanism independent of the nature<br />

of the flow in the coarsening arrest.<br />

Further numerical <strong>and</strong> experimental investigations, with the aim of clarifying<br />

the dependence of the arrest scale on the flow properties, would be extremely<br />

interesting.<br />

94<br />

10 2<br />

10 3<br />

t<br />

10 4<br />

10 5


Conclusions<br />

In this thesis it has been presented a numerical <strong>and</strong> theoretical study on <strong>turbulence</strong><br />

in <strong>viscoelastic</strong> <strong>fluids</strong> <strong>and</strong> <strong>binary</strong> <strong>mixtures</strong>. Concerning the first subject, two items<br />

have been considered: the study of small-scale statistics of <strong>viscoelastic</strong> <strong>turbulence</strong><br />

<strong>and</strong> the destabilization of a laminar flow by means of polymer additives. In the<br />

case of <strong>binary</strong> <strong>fluids</strong> the interplay between the coarsening process <strong>and</strong> <strong>turbulence</strong><br />

has been addressed.<br />

The effects of polymer injection on the dynamics of the turbulent cascade<br />

have been studied by means of direct numerical simulations of a simplified uniaxial<br />

model of <strong>viscoelastic</strong> flow, in a three-dimensional homogeneous isotropic<br />

configuration. It has been found evidence of a coil-stretch transition, above which<br />

polymers affect the small scales of the turbulent flow by partially removing the<br />

energy flux. Nevertheless, this effect saturates <strong>and</strong> high-wavenumber activity survives<br />

to the addition of polymers, even for large elasticity; as a consequence the<br />

energy flux to small scales remains finite <strong>and</strong> the small-scale statistics, such as acceleration<br />

probability density function, retain some characteristics of <strong>Newtonian</strong><br />

flows.<br />

The phenomenon of elastic <strong>turbulence</strong> has been investigated in a two-dimensional<br />

flow with periodic boundary conditions. At growing values of elasticity a<br />

transition to chaotic <strong>and</strong>, eventually, turbulent states has been numerically observed.<br />

Control of numerical instabilities, which are known to severely limit<br />

simulation of <strong>viscoelastic</strong> <strong>fluids</strong>, has been achieved by developing an algorithm<br />

based on a Cholesky decomposition of the conformation tensor, which preserves<br />

its positive definiteness. The appearence of turbulent-like features has been detected<br />

by measuring the flow resistance <strong>and</strong> Eulerian Lyapunov exponents. Close<br />

to the elastic instability threshold, an intermediate regime of elastic waves was<br />

observed. Power spectra of velocity fluctuations were found to display a power<br />

law behaviour, at very large elasticities. Mixing has been studied by means of<br />

Lagrangian Lyapunov exponents, finding that the mean chaoticity degree of the<br />

flow grows at growing Weissenberg number <strong>and</strong> the behaviour of its fluctuations<br />

resembles that of st<strong>and</strong>ard <strong>Newtonian</strong> <strong>fluids</strong> in a turbulent regime.<br />

The phase ordering process of both active <strong>and</strong> passive <strong>binary</strong> <strong>fluids</strong> has been<br />

95


96 Conclusions<br />

studied by numerical integration of coupled Cahn-Hilliard <strong>and</strong> Navier-Stokes equations<br />

in two dimensions, when the fluid velocity field is externally forced. In particular,<br />

it has been considered the comptetition between the thermodynamic forces<br />

leading to phase separation <strong>and</strong> the mixing induced by a turbulent flow. For both<br />

active <strong>and</strong> passive <strong>mixtures</strong>, it has been found that, for sufficiently strong stirring,<br />

coarsening is arrested. The existence of a statistically steady state characterized<br />

by continuously breaking <strong>and</strong> reforming domains of a single phase, has been understood<br />

with an energy conservation argument. Moreover, in the passive limit<br />

it has been demonstrated through numerical experiments that Lagrangian chaos<br />

is not necessary for the phenomenon of coarsening arrest, since this is due to the<br />

relative strength of local shears <strong>and</strong> segregation forces.<br />

96


Bibliography<br />

[1] U. Frisch, Turbulence: the legacy of A. N. Kolmogorov, Cambridge University<br />

Press (1995).<br />

[2] A. S. Monin <strong>and</strong> A. M. Yaglom, Statistical fluid mechanics I, The Mit Press<br />

(1979).<br />

[3] A. S. Monin <strong>and</strong> A. M. Yaglom, Statistical fluid mechanics II, The Mit<br />

Press (1981).<br />

[4] G. K. Batchelor, The theory of homogeneous <strong>turbulence</strong>, Cambridge University<br />

Press (1953).<br />

[5] H. Tennekes <strong>and</strong> J. L. Lumley, A first course in <strong>turbulence</strong>, The Mit Press<br />

(1972).<br />

[6] L. D. L<strong>and</strong>au <strong>and</strong> E. M. Lifshitz, Fluid mechanics, Pergamon Press (1959).<br />

[7] L. F. Richardson, Weather prediction by numerical process, Cambridge<br />

University Press (1922).<br />

[8] H. L. Grant, R. W. Stewart <strong>and</strong> A. Moilliet, J. Fluid Mech. 12, 241 (1962).<br />

[9] F. Anselmet, Y. Gagne, E. J. Hopfinger, R. A. Antonia, J. Fluid Mech. 140,<br />

63 (1984).<br />

[10] J. Maurer, P. Tabeling, G. Zocchi Europhys. Lett. 26, 31 (1994).<br />

[11] B. Castaing, Y. Gagne, E. J. Hopfinger, Physica D 46, 177 (1990).<br />

[12] J. Herweijer, W. Van de Water, Phys. Rev. Lett. 74, 4651 (1995).<br />

[13] Y. Malécot, C. Auriault, Y. Gagne, O. Chanal, B. Chabaud <strong>and</strong> B. Castaing,<br />

Eur. Phys. J. B 16, 549 (2000).<br />

97


98 BIBLIOGRAPHY<br />

[14] Parisi <strong>and</strong> U. Frisch, in Turbulence <strong>and</strong> predictability in geophysical fluid<br />

dynamics, Proceed. Inern. School of Physics ’E. Fermi’, 1983, Varenna,<br />

Italy, 84, edited by M. Ghil, R. Benzi <strong>and</strong> G. Parisi, North-Holl<strong>and</strong> (Amsterdam,<br />

1985).<br />

[15] R. Benzi, G. Paladin, G. Parisi, A. Vulpiani, J. Phys. A 17, 3521 (1984).<br />

[16] P. A. Vaillancourt <strong>and</strong> M. K. Yau, Bull. Am. Meteorol. Soc. 81, 285 (2000).<br />

[17] S. B. Pope, Annu. Rev. Fluid Mech. 26, 23 (1994).<br />

[18] A. La Porta, G. A. Voth, A. M. Crawford, J. Alex<strong>and</strong>er <strong>and</strong> E. Bodenschatz,<br />

Nature 409, 1017 (2001).<br />

[19] C. Beck, Phys. Lett. A 287, 240 (2001).<br />

[20] C. Beck, Europhys. Lett. 64, 151 (2003).<br />

[21] A. K. Aringazin, M. I. Mazhitov, Phys. Lett. A 313, 284 (2003).<br />

[22] T. Arimitsu, N. Arimitsu, cond-math/0301516.<br />

[23] T. Gotoh <strong>and</strong> R. H. Kraichnan, nlin.CD/0305040.<br />

[24] L. Biferale, G. Boffetta, A. Celani, B. J. Devenish, A. Lanotte <strong>and</strong> F. Toschi,<br />

Phys. Rev. Lett. 93, 064502 (2004).<br />

[25] D. Biskamp, <strong>Non</strong>linear magnetohydrodynamics, Cambridge University<br />

Press (1993).<br />

[26] R. H. Kraichnan, Phys. Fluids 10, 1417 (1967).<br />

[27] G. K.Batchelor, Phys. Fluids 12, II-233 (1969).<br />

[28] R. H. Kraichnan <strong>and</strong> D. Montgomery, Rep. Prog. Phys. 43, 548 (1980).<br />

[29] M. Chertkov, Phys. Fluids 10, 3017 (1998).<br />

[30] G. Boffetta, A. Celani, M. Vergassola, Phys. Rev. E 61, R29 (2000).<br />

[31] R. B. Bird, R. C. Armstrong, O. Hassager, C. F. Curtiss, Dynamics of polymeric<br />

liquids, Wiley (New York, 1987).<br />

[32] R. H. Nadolink, W. W. Haigh, ASME Appl. Mech. Rev. 48, 351 (1995).<br />

[33] B. H. Zimm, J. Chem. Phys. 24, 269 (1956).<br />

98


BIBLIOGRAPHY 99<br />

[34] T. T. Perkins, S. R. Quake, D. E. Smith <strong>and</strong> S. Chu, Science 264, 822<br />

(1994).<br />

[35] M. Doi, S. F. Edwards, The theory of polymer dynamics, Clarendon Press<br />

(1986).<br />

[36] P. Flory, Principles of polymer chemistry, Cornell University Press (1971).<br />

[37] J. L. Lumley, J. Polym. Sci. Part D 7, 263 (1973).<br />

[38] E. Balkovsky, A. Fouxon <strong>and</strong> V. Lebedev, Phys. Rev. Lett. 84, 4765 (2000).<br />

[39] E. Balkovsky, A. Fouxon <strong>and</strong> V. Lebedev, Phys. Rev. E 64, 056301 (2001).<br />

[40] G. Boffetta, A. Celani, S. Musacchio, Phys. Rev. Lett. 91, 034501 (2003).<br />

[41] E. J. Hinch, Phys. Fluids 20, S22 (1977).<br />

[42] S. R. Quake, H. Babcock <strong>and</strong> S. Chu, Nature 388, 151 (1997).<br />

[43] A. Peterlin, J. Polym. Sci. B, Poly. Lett. B4, 287 (1966).<br />

[44] E. De Angelis, C. M. Casciola, V. S. L’Vov, R. Piva <strong>and</strong> I. Procaccia,<br />

Phys. Rev. E 67, 056312 (2003).<br />

[45] B. A. Toms, in Proceedings of the International Congress on Rheology II,<br />

135, North-Holl<strong>and</strong> (Amsterdam, 1949).<br />

[46] J. L. Lumley, Annu. Rev. Fluid Mech. 1, 367 (1969).<br />

[47] M. Tabor, P. G. de Gennes, Europhys. Lett. 2, 519 (1986).<br />

[48] V. S. L’Vov, A. Pomyalov, I. Procaccia, V. Tiberkevich, Phys. Rev. Lett. 92,<br />

244503 (2004).<br />

[49] R. Benzi, V. S. L’Vov, I. Procaccia, V. Tiberkevich, Europhys. Lett. 68, 825<br />

(2004).<br />

[50] A. Gyr, H. W. Bewersdorff Drag reduction of turbulent flows by additives,<br />

Kluwer Academic Publishers (1995).<br />

[51] K. R. Sreenivasan, C. M. White, J. Fluid. Mech. 409, 149 (2000).<br />

[52] Y. Amarouchene, H. Kellay, Phys. Rev. Lett. 89, 104502 (2002).<br />

[53] G. Boffetta, A. Celani, A. Mazzino, Phys. Rev. E 71, 036307 (2005).<br />

99


100 BIBLIOGRAPHY<br />

[54] A. Groisman, V. Steinberg, Europhys. Lett. 43, 165 (1998).<br />

[55] A. Groisman, V. Steinberg, Nature 405, 53 (2000).<br />

[56] A. Groisman, V. Steinberg, Nature 410, 905 (2001).<br />

[57] A. Groisman, V. Steinberg, New J. Phys. 6, 29 (2004).<br />

[58] P. E. Arratia, G. A. Voth, J. P. Gollub, Phys. Fluids 17, 053102 (2005).<br />

[59] H. Furukawa, Adv. Phys. 34, 6, 703 (1985).<br />

[60] A. J. Bray, Adv. Phys. 51, 481 (2002).<br />

[61] E. D. Siggia, Phys. Rev. A 20, 595 (1979).<br />

[62] R. Ruiz <strong>and</strong> D. R. Nelson, Phys. Rev. A 23, 3224 (1981).<br />

[63] T. Hashimoto, K. Matsuzaka, E. Moses <strong>and</strong> A. Onuki, Phys. Rev. Lett. 74,<br />

126 (1995).<br />

[64] J. W. Cahn <strong>and</strong> J. H. Hilliard, J. Chem. Phys. 28, 258 (1958).<br />

[65] V. M. Kendon, M. E. Cates, I. Pagonabarraga, J. C. Desplat <strong>and</strong> P. Bl<strong>and</strong>on,<br />

J. Fluid Mech. 440, 147 (2001).<br />

[66] A. J. Wagner <strong>and</strong> J. M. Yeomans, Phys. Rev. Lett. 80, 1429 (1998).<br />

[67] H. Furukawa, Phys. Rev. A 31, 1103 (1985).<br />

[68] H. Tanaka <strong>and</strong> T. Araki, Phys. Rev. Lett. 81, 389 (1998).<br />

[69] P. S. Virk, AIChE Journal 21, 625 (1975).<br />

[70] A. Liberzon, M. Guala, B. Lüthi, W. Kinzelbach <strong>and</strong> A. Tsinober,<br />

Phys. Fluids 17, 031707 (2005).<br />

[71] E. De Angelis, C. M. Casciola, R. Benzi <strong>and</strong> R. Piva, J. Fluid Mech. 531, 1<br />

(2005).<br />

[72] S. Berti, A. Bistagnino, G. Boffetta, A. Celani, S. Musacchio, Europhys.<br />

Lett. 76, 63 (2006).<br />

[73] G. I. Ogilvie <strong>and</strong> M. R. E. Proctor, J. Fluid Mech. 476, 389 (2003).<br />

[74] A. Fouxon, V. Lebedev, Phys. Fluids 15, 2060 (2003).<br />

100


BIBLIOGRAPHY 101<br />

[75] S. Chen, G. D. Doolen, R. H. Kraichnan <strong>and</strong> Z. S. She, Phys. Fluids A 5,<br />

458 (1993).<br />

[76] S. S. Girimaji <strong>and</strong> B. Pope, J. Fluid Mech. 220, 427 (1990).<br />

[77] S. Gerashchenko, C. Chevallard <strong>and</strong> V. Steinberg, Europhys. Lett. 71, 221<br />

(2005).<br />

[78] R. Benzi, E. De Angelis, R. Govindarajan <strong>and</strong> I. Procaccia, Phys. Rev. E<br />

68, 016308 (2003).<br />

[79] A. Crawford, N. Mordant, A. La Porta <strong>and</strong> E. Bodenschatz, in Advances in<br />

Turbulence IX, Proceedings of the Ninth European Turbulence Conference,<br />

edited by I, P. Castro, P. E. Hancock <strong>and</strong> T. G. Thomas, CIMNE (Barcelona,<br />

2002).<br />

[80] R. G. Larson, Rheol. Acta 31, 213 (1992).<br />

[81] A. Groisman <strong>and</strong> V. Steinberg, Phys. Rev. Lett. 77, 1480 (1996).<br />

[82] R. G. Larson, E. S. G. Shaqfeh <strong>and</strong> S. J. Muller, J. Fluid Mech. 218, 573<br />

(1990).<br />

[83] E. S. G. Shaqfeh, Annu. Rev. Fluid Mech. 28, 129 (1999).<br />

[84] G. Boffetta, A. Celani, A. Mazzino, A. Puliafito <strong>and</strong> M. Vergassola, J. Fluid<br />

Mech. 523, 161 (2005).<br />

[85] T. Burghelea, E. Segre, I. Bar-Joseph, A. Groisman <strong>and</strong> V. Steinberg,<br />

Phys. Rev. E 69, 066305 (2004).<br />

[86] V. I. Arnold <strong>and</strong> L. Meshalkin, Uspekhi Mat. Nauk 15, 247 (1960).<br />

[87] L. Meshalkin <strong>and</strong> Y. G. Sinai, J. Appl. Math. Mech. 25, 1700 (1961).<br />

[88] R. Sureshkumar <strong>and</strong> A. N. Beris, J. <strong>Non</strong>-<strong>Newtonian</strong> Fluid Mech. 60, 53<br />

(1995).<br />

[89] T. Vaithianathan <strong>and</strong> L. R. Collins, J. Comp. Phys. 187, 1 (2003).<br />

[90] G. Paladin <strong>and</strong> A. Vulpiani, Phys. Rep. 156, 147 (1987).<br />

[91] A. Onuki, J. Phys.: Condens. Matter 9, 6119 (1997).<br />

[92] J. A. Aronovitz <strong>and</strong> D. R. Nelson, Phys. Rev. A 29, 2012 (1984).<br />

[93] A. Onuki, Phys. Lett. A 101, 286 (1984).<br />

101


102 BIBLIOGRAPHY<br />

[94] M. E. Cates, V. M. Kendon, P. Bladon <strong>and</strong> J. C. Desplat, Faraday Discuss.<br />

112 1 (1999).<br />

[95] F. Corberi, G. Gonnella <strong>and</strong> A. Lamura, Phys. Rev. Lett. 81, 3852 (1998).<br />

[96] F. Corberi, G. Gonnella <strong>and</strong> A. Lamura, Phys. Rev. E 61, 6621 (2000).<br />

[97] L. Berthier, Phys. Rev. E 63, 051503 (2001).<br />

[98] Z. Shou <strong>and</strong> A. Chakrabarti, Phys. Rev. E 61, R2200 (2000).<br />

[99] D. J. Pine, N. Easwar, J. V. Maher <strong>and</strong> W. I. Goldburg, Phys. Rev. A 29,<br />

308 (1984).<br />

[100] C. K. Chan, W. I. Goldburg <strong>and</strong> J. V. Maher, Phys. Rev. A 35, 1756 (1987).<br />

[101] P. Tong, W. I. Goldburg, J. Stavans <strong>and</strong> A. Onuki, Phys. Rev. Lett. 62, 2668<br />

(1989).<br />

[102] K. Y. Min, J. Stavans, R. Piazza <strong>and</strong> W. I. Goldburg, Phys. Rev. Lett. 63,<br />

1070 (1989).<br />

[103] K. Y. Min <strong>and</strong> W. I. Goldburg, Phys. Rev. Lett. 70, 469 (1993).<br />

[104] F. J. Muzzio, M. Tjahjadi, <strong>and</strong> J. M. Ottino, Phys. Rev. Lett. 67, 54 (1991).<br />

[105] S. Berti, G. Boffetta, M. Cencini, A. Vulpiani, Phys. Rev. Lett. 95, 224501<br />

(2005).<br />

[106] A. M. Lacasta, J. M. Sancho <strong>and</strong> F. Sagués, Phys. Rev. Lett. 75, 1791<br />

(1995).<br />

[107] L. Berthier, J. L. Barrat <strong>and</strong> J. Kurchan, Phys. Rev. Lett. 86, 2014 (2001).<br />

[108] A. Crisanti, M. Falcioni, G. Paladin <strong>and</strong> A. Vulpiani, Riv. Nuovo Cim. 14,<br />

1 (1991).<br />

[109] H. Tanaka, Phys. Rev. Lett. 72, 3690 (1994).<br />

[110] G. Satten <strong>and</strong> D. Ronis, Phys. Rev. Lett. 55, 91 (1985).<br />

[111] H. Furukawa, Phys. Rev. E 61, 1423 (2000).<br />

[112] K. Nam, E. Ott, T. M. Antonsen <strong>and</strong> P. N. Guzdar, Phys. Rev. Lett. 84, 5134<br />

(2000).<br />

102


Acknowledgements<br />

I would like to thank my supervisors for introducing me to the study of <strong>turbulence</strong><br />

in complex <strong>fluids</strong>, <strong>and</strong> for their friendship: Guido Boffetta for guidance <strong>and</strong><br />

huge support in the work, but also for being one of the first letting me discover<br />

the beauty of the mountains of Northwestern Italy; Antonio Celani for support,<br />

illuminating observations <strong>and</strong> the extreme clarity of his arguments.<br />

The works about <strong>viscoelastic</strong> turbulent flows were done together with Andrea<br />

Bistagnino <strong>and</strong> Stefano Musacchio; in relation to the work on small-scale statistics<br />

I also acknowledge useful <strong>and</strong> interesting discussions with Alex Liberzon <strong>and</strong><br />

Beat Lüthi. The study about phase separation in turbulent <strong>binary</strong> <strong>fluids</strong> was done<br />

in collaboration with Massimo Cencini <strong>and</strong> Angelo Vulpiani.<br />

I also would like to mention some other people I had the opportunity to work<br />

<strong>and</strong> discuss with: Filippo De Lillo, Davide Dezzani, Cristóbal López, Alberto<br />

Puliafito <strong>and</strong> Davide Vergni.<br />

I am grateful to everybody at ISAC-CNR, Torino, for their hospitality in my<br />

first year of PhD; people at INLN-CNRS, Nice, for hospitality during the period<br />

I worked there; Agnese Seminara <strong>and</strong> Dario Vincenzi for their kind hospitality in<br />

Nice; people at IMEDEA, Palma de Mallorca, <strong>and</strong> Francesco Visconti for hospitality<br />

during the period I spent in Spain.<br />

I would like to thank my friend Gabriele Giovanetti from Rome, for sharing the<br />

first times in Torino <strong>and</strong> the company he formerly worked for, which generously<br />

allowed us to discover the "gastronomia piemontese" by offering us a number of<br />

great dinners.<br />

A special thanks to all the friends I met in Torino, the scientists, the "alpinists",<br />

<strong>and</strong> those who are both. Among these, a huge thanks goes to Floriana Gargiulo,<br />

together with the promise to climb all the mountains of her beloved Valle di Susa...<br />

"a sarà düra!"<br />

Last but not least, I would like to thank all the friends from Garbatella, Rome,<br />

for never giving up doing what they do, <strong>and</strong> my family for all the support they<br />

give me.<br />

103

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!