Energy Dissipation in Fractal-Forced Flow. Abstract

Similar documents
Energy and enstrophy dissipation in steady state 2d turbulence

Scaling of space time modes with Reynolds number in two-dimensional turbulence

Small-Scale Statistics and Structure of Turbulence in the Light of High Resolution Direct Numerical Simulation

Bounds on Surface Stress Driven Flows

Global regularity of a modified Navier-Stokes equation

Chapter 7 The Time-Dependent Navier-Stokes Equations Turbulent Flows

Turbulence: Basic Physics and Engineering Modeling

The 3D Navier-Stokes Problem

PSEUDO-COMPRESSIBILITY METHODS FOR THE UNSTEADY INCOMPRESSIBLE NAVIER-STOKES EQUATIONS

On the (multi)scale nature of fluid turbulence

Mathematical Hydrodynamics

Global well-posedness of the primitive equations of oceanic and atmospheric dynamics

Euler Equations: local existence

Linearly forced isotropic turbulence

ASYMPTOTIC STRUCTURE FOR SOLUTIONS OF THE NAVIER STOKES EQUATIONS. Tian Ma. Shouhong Wang

Energy Spectrum of Quasi-Geostrophic Turbulence Peter Constantin

36. TURBULENCE. Patriotism is the last refuge of a scoundrel. - Samuel Johnson

DIRECTION OF VORTICITY AND A REFINED BLOW-UP CRITERION FOR THE NAVIER-STOKES EQUATIONS WITH FRACTIONAL LAPLACIAN

A RECURRENCE THEOREM ON THE SOLUTIONS TO THE 2D EULER EQUATION

Turbulent energy density and its transport equation in scale space

An Introduction to Theories of Turbulence. James Glimm Stony Brook University

Week 6 Notes, Math 865, Tanveer

Logarithmic Bounds for Infinite Prandtl Number Rotating Convection

Locality of Energy Transfer

New Discretizations of Turbulent Flow Problems

A LOWER BOUND ON BLOWUP RATES FOR THE 3D INCOMPRESSIBLE EULER EQUATION AND A SINGLE EXPONENTIAL BEALE-KATO-MAJDA ESTIMATE. 1.

THE POINCARÉ RECURRENCE PROBLEM OF INVISCID INCOMPRESSIBLE FLUIDS

Logarithmic scaling in the near-dissipation range of turbulence

Euler equation and Navier-Stokes equation

Introduction Efficiency bounds Source dependence Saturation Shear Flows Conclusions STIRRING UP TROUBLE

Logarithmic bounds for infinite Prandtl number rotating convection

O. A Survey of Critical Experiments

CONNECTIONS BETWEEN A CONJECTURE OF SCHIFFER S AND INCOMPRESSIBLE FLUID MECHANICS

Lecture 14. Turbulent Combustion. We know what a turbulent flow is, when we see it! it is characterized by disorder, vorticity and mixing.

On the local existence for an active scalar equation in critical regularity setting

Fundamentals of Fluid Dynamics: Elementary Viscous Flow

Coupled systems of two-dimensional turbulence

Separation of Variables in Linear PDE: One-Dimensional Problems

Probing Fundamental Bounds in Hydrodynamics Using Variational Optimization Methods

SPACE AVERAGES AND HOMOGENEOUS FLUID FLOWS GEORGE ANDROULAKIS AND STAMATIS DOSTOGLOU

On the well-posedness of the Prandtl boundary layer equation

3D Large-Scale DNS of Weakly-Compressible Homogeneous Isotropic Turbulence with Lagrangian Tracer Particles

EXACT SOLUTIONS TO THE NAVIER-STOKES EQUATION FOR AN INCOMPRESSIBLE FLOW FROM THE INTERPRETATION OF THE SCHRÖDINGER WAVE FUNCTION

The Euler Equations and Non-Local Conservative Riccati Equations

From Boltzmann Equations to Gas Dynamics: From DiPerna-Lions to Leray

On non negative solutions of some quasilinear elliptic inequalities

CHAPTER 7 SEVERAL FORMS OF THE EQUATIONS OF MOTION

Intermittency, Fractals, and β-model

hal , version 6-26 Dec 2012

Characteristics of Linearly-Forced Scalar Mixing in Homogeneous, Isotropic Turbulence

Validation of an Entropy-Viscosity Model for Large Eddy Simulation

Existence of minimizers for the pure displacement problem in nonlinear elasticity

Global well-posedness for the critical 2D dissipative quasi-geostrophic equation

Dissipative quasi-geostrophic equations with L p data

Rough solutions of the Stochastic Navier-Stokes Equation

Computational Fluid Dynamics 2

1 Introduction. J.-L. GUERMOND and L. QUARTAPELLE 1 On incremental projection methods

Multiscale Computation of Isotropic Homogeneous Turbulent Flow

Stability of Shear Flow


A new approach to the study of the 3D-Navier-Stokes System

Universität des Saarlandes. Fachrichtung 6.1 Mathematik

Convergence of Time Averaged Statistics of Finite Element Approximations of the Navier-Stokes equations

Isotropic homogeneous 3D turbulence

This note presents an infinite-dimensional family of exact solutions of the incompressible three-dimensional Euler equations

and finally, any second order divergence form elliptic operator

On the decay of two-dimensional homogeneous turbulence

Another particular instance includes the space B 1/3

The behaviour of high Reynolds flows in a driven cavity

SHARP BOUNDARY TRACE INEQUALITIES. 1. Introduction

Equilibrium analysis for a mass-conserving model in presence of cavitation

Max Planck Institut für Plasmaphysik

On the Ladyzhenskaya Smagorinsky turbulence model of the Navier Stokes equations in smooth domains. The regularity problem

Turbulence is a ubiquitous phenomenon in environmental fluid mechanics that dramatically affects flow structure and mixing.

Buoyancy Fluxes in a Stratified Fluid

1. Comparison of stability analysis to previous work

THE L 2 -HODGE THEORY AND REPRESENTATION ON R n

A Sufficient Condition for the Kolmogorov 4/5 Law for Stationary Martingale Solutions to the 3D Navier-Stokes Equations

The Kolmogorov Law of turbulence

arxiv: v1 [math.ap] 14 Apr 2009

Cascade of Random Rotation and Scaling in a Shell Model Intermittent Turbulence

General introduction to Hydrodynamic Instabilities

SOME VIEWS ON GLOBAL REGULARITY OF THE THIN FILM EQUATION

New Identities for Weak KAM Theory

OPTIMAL CONVERGENCE RATES FOR THE COMPRESSIBLE NAVIER-STOKES EQUATIONS WITH POTENTIAL FORCES

Borel Summability in PDE initial value problems

A Pressure-Correction Scheme for Rotational Navier-Stokes Equations and Its Application to Rotating Turbulent Flows

Lecture 4: The Navier-Stokes Equations: Turbulence

Everywhere differentiability of infinity harmonic functions

Dimensionality influence on energy, enstrophy and passive scalar transport.

The Navier-Stokes equations I consider an incompressible fluid continuum R 3, of constant density ( 0 = 1) I conservation of mass/incompressibility re

Takens embedding theorem for infinite-dimensional dynamical systems

Author(s) Huang, Feimin; Matsumura, Akitaka; Citation Osaka Journal of Mathematics. 41(1)

Optimal L p (1 p ) rates of decay to linear diffusion waves for nonlinear evolution equations with ellipticity and dissipation

Simulating Drag Crisis for a Sphere Using Skin Friction Boundary Conditions

Modelling of turbulent flows: RANS and LES

On an uniqueness theorem for characteristic functions

Geometric Multigrid Methods

Minimal periods of semilinear evolution equations with Lipschitz nonlinearity

7. Basics of Turbulent Flow Figure 1.

Transcription:

APS/13-QED Energy Dissipation in Fractal-Forced Flow Alexey Chesidov Department of Mathematics, University of Michigan, Ann Arbor, MI 48109 Charles R. Doering Department of Mathematics & Michigan Center for Theoretical Physics University of Michigan, Ann Arbor, MI 48109 Niola P. Petrov Department of Mathematics University of Olahoma, Norman, OK 73019 (Dated: July 9, 006) Abstract The rate of energy dissipation in solutions of the body-forced 3-d incompressible Navier-Stoes equations is rigorously estimated with a focus on its dependence on the nature of the driving force. For square integrable body forces the high Reynolds number (low viscosity) upper bound on the dissipation is independent of the viscosity, consistent with the existence of a conventional turbulent energy cascade. On the other hand when the body force is not square integrable, i.e., when the Fourier spectrum of the force decays sufficiently slowly at high wavenumbers, there is significant direct driving at a broad range of spatial scales. Then the upper limit for the dissipation rate may diverge at high Reynolds numbers, consistent with recent experimental and computational studies of fractal-forced turbulence. PACS numbers: 47.10.ad, 47.10.A-, 47.7.E-, 47.7.-i, 47.7.Gs, 47.7.Jv, 0.30.Jr, 0.30.Sa Keywords: Navier-Stoes equations, energy dissipation, turbulence, turbulent cascade Electronic address: achesid@umich.edu Electronic address: doering@umich.edu Electronic address: npetrov@ou.edu 1

I. INTRODUCTION A fundamental principle of modern hydrodynamic turbulence theory is that nonlinear interactions between Fourier modes of the velocity field can transfer energy from directlyforced large spatial scales, through the so-called inertial range, down to a small dissipation length scale where viscosity effectively consumes inetic energy and transforms it into heat. This turbulent cascade process has been intensively studied experimentally, numerically, and theoretically (at various levels of mathematical rigor) since the first half of the twentieth century. See, e.g., the boo by Frisch [1] for an introduction and entry into the vast literature on this subject, which is still the focus of much current research. One profound consequence of the cascade mechanism is the so-called dissipative anomaly wherein a finite and non-vanishing residual energy dissipation persists in the singular limit of vanishing viscosity, i.e., in the infinite Reynolds number limit. This phenomenon is quantitatively described as Kolmogorov scaling of the energy dissipation, namely β ɛl U 3 = O(Re 0 ) as Re (1) where ɛ is the total energy dissipation rate per unit mass, l is an integral (large) length scale in the flow characterizing the domain or a large scale in the forcing and flow, U is a turbulent velocity scale, and Re = U l/ν is the Reynolds number with ν denoting the inematic viscosity. Sreenivasan has collected together relevant data illustrating Kolmogorov scaling in experiments [] and direct numerical simulations [3]. Moreover, given precise definitions of all the quantities involved, this β Re 0 Kolmogorov scaling has been shown to be an upper bound for (wea) solutions of the 3-d incompressible Navier-Stoes equations driven by sufficiently smooth in particular, square integrable body forces [4 7]. While the cascade picture of turbulence requires that energy be predominantly injected in a relatively narrow range of spatial scales, some researchers have recently performed experimental and computational studies of fractal-generated turbulence. These are flows driven by spatially broadband fractal forces with certain scaling properites that inject energy directly at a wide range of scales most notably at small scales that could otherwise only be excited by the cascade. Such forcing can impose a self-similar structure on the flow that is independent of the turbulent energy cascade. If such forcing can be achieved experimentally then one can observe, and in principle control, the balance between the energy that has been directly injected and the energy transfered by the nonlinear mode interactions.

Indeed, Queiros-Conde and Vassilicos [8] performed experiments by forcing fluid past a fractal object, an obstacle that was structurally self-similar over several scales. Staicu et al [9] experimentally measured energy spectra and structure functions in the wae of grids of fractal dimensions.05,.17, and.40 in a wind tunnel, concluding that there may be a direct relation between the scaling properties of the fractal object and the turbulence it creates. This is more easily investigated in direct numerical simulations where details of the flow field is directly observable. Mazzi and Vassilicos [10] performed direct numerical simulations of stationary homogeneous and isotropic turbulence in a fluid in a 3-d periodic box of size l driven by a velocity-dependent fractal body force f(x, t) with Fourier components of the form: ( F (l ) ζ û (t) ˆf (t) = û (t) + i û(t) ), 0 < < F, û (t) () 0, > F where û are the velocity field s Fourier components, ˆf = 0 whenever û = 0, and û 0 0. The scaling exponent ζ is intended to characterize the fractal properties of the stirrer or obstacle, and the maximum wavenumber F is to be thought of as the inverse of the spatial size of the smallest parts of the fractal stirrer. Mazzi and Vassilicos used numerical values for which the fractal forcing extended down to scales 1 F on the order of the Kolmogorov dissipation length η (ν 3 /ɛ) 1/4. They observed that the bul energy dissipation rate did not exhibit Kolmogorov scaling β Re 0, but rather β Re 1 corresponding to ɛ U 4 /ν. Biferale et al [11, 1] performed numerical simulations of the 3-d Navier-Stoes equations with a stochastic body force that was white-noise in time but with a power law spectrum of spatial scales ζ. They investigated small scale turbulent fluctuations and concluded that the statistics displayed two distinct qualitative behaviors. When the spatial spectrum of the forcing decayed sufficiently fast, the small scale fluctuations were universal in the sense that they were independent of the details of the force spectrum. This regime corresponds to conventional cascade dynamics. When the spatial spectrum of the forcing decayed more slowly, however, the small scale fluctuations were force-dominated with the cascade being overwhelmed by the direct excitation from the driving. Interestingly, they reported that this transition occurs at a value ζ = 3 of the scaling exponent corresponding to the boundary between (spatially) square integrable and rougher forcing functions without square summable Fourier coefficients. 3

In this paper we derive rigorous upper bounds on the bul energy dissipation ɛ in an incompressible Newtonian fluid driven by a variety of body forces including forces that are not square integrable. This wor generalizes the previous analysis for square integrable body forces [4, 5, 7] to include fractal forces that drive the flow directly at a broad range of scales. In accord with the findings of Biferale et al we find that the case of square integrable forcing is a borderline situation: β Re 0 when the body forces are square integrable (or smoother), but the estimates increase for rougher driving so that the dissipation coefficient β may increase asymptotically as Re. For the roughest forcing functions that mae sense mathematically, i.e., forcing functions with Fourier coefficients satisfying ˆf <, we find that β Re 1, the scaling observed by Mazzi et al. The rest of this paper is organized as follows. The following Section II lays out the mathematical setting for the analysis and gives the definitions of the physically relevant quantities of interest. In Section III we study the case of time-independent body forces, and the subsequent Section IV deals with velocity-dependent forces lie (). The concluding Section V contains a brief summary and some closing remars. For completeness and to mae the paper self-contained, we include some mathematical details in an appendix. II. STATEMENT OF THE PROBLEM AND DEFINITIONS Consider the incompressible 3-d Navier-Stoes equations on a periodic domain x [0, l] 3 : t u + (u )u + p = ν u + f (3) where u(x, t) is the divergence-free velocity field, p(x, t) is the pressure, f(x, t) is the applied body-force, ν > 0 is the inematic viscosity, and u t=0 = u 0 (x) is the initial condition. We will tae the body force to be a specified time independent (divergence-free) function f(x) in Section III, or given by a velocity-dependent expression lie () in Section IV. We write Fourier expansions as u(x, t) = û (t)e ix where û (t) = 1 l 3 l 3 e ix u(x, t)d 3 x. (4) Without loss of generality, in every case we will tae the applied body force and initial data to have spatial mean zero so that = > 0 in all sums. 4

A field u(x, t) H α if u(, t) H α < where we define the Sobolev norms H α by u(, t) H α u (l) α û (t) = 1 l 3 l 3 ( l ) α u(x, t) d 3 x. (5) The index α can be positive or negative and the function spaces H α are nested according to H α H α for α > α. The case α = 0 corresponds to the usual L norm (with the volume normalization) and we write u H 0 = u L = u (6) Fractal forces are defined as those with power-law Fourier coefficients, ˆf = C ζ. For such a function to belong to the Sobelov space H α its exponent must satistfy ζ < α 3, i.e., the Fourier coefficients must decay as ˆf α 3 δ for some δ > 0. We define time averages of functions g(t) according to T g = lim T 0 g(t)dt (7) and for simplicity in this paper we presume that this limit exists for all quantities of interest. The bul (volume and time) average of a function h(x, t) is denoted by h = 1 l 3 l 3 h(x, )d 3 x. (8) The root means square velocity U of a solution u(x, t) of the Navier-Stoes equations is U = u 1/ = u 1/, (9) and the bul energy dissipation rate (per unit mass) is defined by ɛ = ν u = ν l u H 1. (10) When a solution u(x, t) satisfies the energy equality (i.e., when the energy is absolutely continuous, which holds for every regular solution), the energy dissipation rate satisfies ɛ = f u. (11) That is, the power supplied by the driving force is balanced by the viscous dissipation. Wea solutions to these 3-d Navier-Stoes equations exist for f H 1, and then in general the relation in (11) is only an inequality, i.e., ɛ f u [13 15]. This fact does not affect our results, however, because we will just derive upper limits on ɛ. Moreover, 5

the assumption of the existence of the long times averages is not necessary if the limit is replaced by lim sup T. With that definition the estimates we derive are fully applicable to wea solutions. Using the definitions above, the Reynolds number is identified Re = Ul/ν and the dissipation coefficient as β = ɛl/u 3. In the scenario described here both Re (or U) and β (or ɛ) are formally emergent quantities, not directly controllable but determined rather as functions of l and ν and functionals of u 0 and f. These bul averaged quantities generally depend on u 0, but the relationships derived below are uniformly valid for all solutions regardless of initial data so we will drop any further reference to them. In practice one assumes that the parameters of the force, e.g., its amplitude, can be tuned to achieve any desired Reynolds number. Then β may be considered a function of Re. The overdamped highly viscous limit is Re 0 and the vanishing viscosity limit is explored as Re. Some very general statements can be made for the overdamped limit. Poincare s inequality implies that so for any forcing ɛ 4π ν l U, (1) β 4π Re. (13) This Reynolds number scaling is sharp: as will be seen below, for a wide variety of forces there exists a constant c 4π (generally depending on the details of the forcing) such that β c Re as Re 0. (14) This scaling, β Re 1, is characteristic of large scale laminar flows where the typical rate of strain is proportional to U/l and the typical stress is proportional to νu/l. For higher Reynolds numbers the lower estimate in (13) can generally not be improved. That is, at arbitrarily high Re there are forces that can sustain the flow with β Re 1. Those flows which may be unstable are necessarily sufficiently laminar to exclude any characterization as being turbulent. The upper bound on β, however, necessarily increases above Re 1 as Re. For turbulent flows with an effective energy cascade the dissipation becomes independent of ν as Re, i.e., β Re 0, as evidenced by experiments and direct numerical simulations. But for sufficiently broadband forcing β may increase indefinitely in this limit, and the tas of the next two sections is to place rigorous upper bounds on β as a function of Re for flows driven by fractal forces. 6

III. STEADY H α BODY FORCES In this section we generalize the approach introduced by Doering & Foias [4] an approach that was inspired by previous wor of Foias and coworers [16 18] to cases where the time independent force f(x) H α with α [0, 1]. For α 0 the force f L and the β Re 0 upper bound, corresponding to the usual energy cascade, is effective [4]. We do not consider values of α > 1, for then even wea solutions of the Navier-Stoes equations are not nown to exist. While the analysis in this section is not restricted to strictly fractal forces, the results apply nevertheless to those with power-law Fourier coefficients ˆf ζ where ζ = α 3 δ for any δ > 0. Write the steady body force as f(x) = F φ(l 1 x), (15) where F is the amplitude of the force, the H α norm of f, and the shape function φ is a dimensionless divergence-free field on the unit 3-torus normalized according to φ H α = 1. (16) Using Cauchy-Schwarz and the interpolation inequality (A6) with s = α, r = 1, t = 0, we estimate 1 f u dx l 3 l 3 [ ] 1/ [ ] 1/ ˆf û (l) α ˆf (l) α û = f H α u H α = F u H α F u α H 1 u 1 α. Then taing time average and applying Hölder s inequality, Note that (1 α) α ɛ f u F u α H 1 u 1 α F = F = F ( ν ) α ( ν ) α l l u H 1 ( ν l ) α ɛ α ( u (1 α) α ) ( ) α α ( u H u (1 α) 1 α ( ) α u (1 α) α ) α. [0, 1 ] so Jensen s inequality (A) ensures that the last term in the last line above is bounded by U 1 α. Hence ɛ l α α ν α α F α U (1 α) α. (17) 7

On the other hand we can also estimate F from above independently in terms of U, ν and l. Multiply the Navier-Stoes equation (3) by a sufficiently smooth time-independent, divergence-free function ψ(l 1 x) on the unit 3-torus satisfying φ ψ > 0. (It s easy to produce such fields ψ, for example as a finite Fourier mode Galerin truncation of φ.) Integrating by parts, taing time averages, and applying Hölder and Cauchy-Schwarz, F φ ψ = u ( ψ) u ν u ψ ψ L u + ν ψ u 1/. (18) Hence F [ ] 1 φ ψ ψ U L l + ψ νu H l, (19) where = l is the dimensionless gradient on the unit 3-torus. Plugging this estimate for F into the bound (17) for ɛ we deduce ( β Re α α C1 + C Re 1) α, (0) where the coefficients C j depend only on the shape function φ and the multiplier function ψ but not on the parameters of the problem, i.e., the force strength F, the viscosity ν, or the outer length scale l. Specifically, C 1 = φ H α ψ L φ ψ and C = For Re 1 the upper bound (0) scales where the exponent φ H α ψ H φ ψ. (1) β Re α α () α [0, 1]. If α = 0, i.e., when the force f α L, we recover the classical estimate corresponding to Kolmogorov scaling β 1 (3) that holds as well when α < 0 [4]. In the other borderline case α = 1, β Re. (4) And as advertised, when Re 1 the overdamped laminar scaling β Re 1 (5) emerges for all α 1. 8

IV. A TIME DEPENDENT FRACTAL FORCE Following Mazzi & Vassilicos [10], consider a fractal forcing function of the form: ( ˆf (t) = F (l ) ζ δ û û + i û ), (6) û where F is the strength coefficient, ζ [ 3, 1 ] and δ > 0 and ˆf 0 whenever û = 0. The Navier-Stoes equations (3) driven by this velocity-dependent time-varying force constitute an autonomous system. We assume initial data u 0 (x) 0, that a (statistically) steady flow is subsequently sustained, and that for t > 0 each û (t) = 0 only on a measure-zero set of times. The scaling exponent ζ = 1 corresponds to the case where the forcing is in H 1 at each instant of time for all δ > 0, while ζ = 3 (or less) is L (or smoother) forcing. Start by writing ɛ = f u = ˆf û = F 1 (l ) 3/+δ (l )ζ+3/ û. (7) Applying the Cauchy-Schwarz inequality, where ( ɛ F C (l ) ζ+3 û )1/. (8) C ( ) 1/ 1. (9) (l ) 3+δ Note that the (3-d) sum defining C converges iff δ > 0. Indeed, C = O(δ 1/ ) as δ 0. Hölder s inequality then implies ( ) ζ/ 1/4 ( ) ζ/+3/4 ɛ CF û (l ) û (30) ( ) ɛl = CF U ζ 1/ ζ/+3/4. (31) ν Solving for ɛ, ( ɛ C 4 4 4ζ+ 1 ζ F 1 ζ U ζ 1 ν ) ζ+3 ζ 1. (3) l Now the challenge is to eliminate F in favor of U, l and ν. 9

To derive an upper bound on F we will estimate the bul average of the (3) dotted into the time-dependent test function ψ(x, t) with the Fourier coefficients: ˆψ (t) = û û (l ) 4 δ (33) for û 0, with ˆψ = 0 when û = 0, and δ > 0. We consider the resulting terms one by one. First note that the pressure term ψ p = 0 since ψ = 0. The advection term is estimated ψ (u u) = u ( ψ) u ψ L u (34) where ψ L l 1 l ˆψ (t) l 1 (l ) 3 δ = C l (35) and the pure number C is finite for all δ > 0; C = O(δ 1 ) as δ 0. The force term is f ψ = F (l ) ζ 4 δ δ = C F (36) where the sum for C converges uniformly for all non-negative δ and δ and all ζ 1/. By asserting equality in (36) above we use the assumption that û > 0 for almost all t > 0. Next, the viscous term is νu ψ ν u = ν ψ u ν ψ H U = C (37) l where ( ) 1/ C = l 4 δ is uniformly bounded for all δ 0. Finally, observe that (38) ψ t u = d û (l ) dt 4 δ. (39) The time average of the time derivative of a quantity vanishes if the quantity is uniformly bounded in time. Because ( ) 1/ ( û (t) (l ) 4 δ û (t) (l ) 8 δ ) 1/ (40) where û (t) = u(, t) is uniformly bounded in time for these H 1 (or smoother) forces, the sum above converges for all δ 0 and we conclude that ψ t u = 0. 10

Hence the bul average of ψ dotted into the Navier-Stoes equations yields F C C U l + C C νu l (41) with absolute constants C and C, and C depending only on δ > 0. Inserting into (3), β Re 3+ζ 1 ζ ( c1 + c Re 1) 4 1 ζ (4) where c 1 = CC C and c = CC C. (43) As before, when Re 0, this result produces the laminar scaling β Re 1, (44) for all relevant values of the force s scaling exponent. When Re, however, the dissipation may be as large as β Re 3+ζ 1 ζ (45) with exponent 0 3+ζ 1 ζ 1 as ζ varies from 3 to 1. It is worthwhile noting that the coefficients c 1 and c depend on δ > 0 (and δ > 0, introduced for convenience) but not at all on the force parameters F and ζ or on ν or l and that the coefficients c 1 (δ) and c (δ) diverge as δ 0 because C(δ) defined in (9) diverges as δ 0. V. SUMMARY & DISCUSSION In this paper we generalized the analysis that was previously employed for square integrable (or smoother) steady forces [4] and velocity-dependent forces [7] to derive bounds on the energy dissipation in the case of broad-band and fractally-forced flow described by the incompressible 3-d Navier-Stoes equations. When a steady body-force f(x) H α with α [0, 1], we showed that the dimensionless dissipation factor β(re) is limited according to ( 4π Re 1 β Re α α C1 + C Re α ) α. (46) For velocity-dependent fractal forces of the form (6) with ˆf ζ δ, ζ [ 3, 1 ], and δ > 0, we deduced that 4π Re 1 β Re 3+ζ 1 ζ ( c1 (δ) + c (δ)re 1) 4 1 ζ. (47) 11

These scalings are sharp as Re 0, displaying the laminar behavior β Re 1. As Re, both upper estimates are β Re 0 for square integrable forcing, i.e., α = 0 and ζ = 3. And in the extreme limits α = 1 and ζ = 1, both estimates give β Re1. We remar that the scalings in (46) and (47) are clearly consistent with each other when it is recognized that forces with ˆf ζ δ are in H α when ζ = α 3. In terms of dimensional physical quantities, we have estimated the energy dissipation rate (per unit mass) ɛ in terms of the rms velocity U, l and ν. Laminar dissipation corresponds to ɛ νu /l while the turbulent cascade is characterized by ɛ U 3 /l and the roughest fractal forces may allows ɛ U 4 /ν. But for a specified form of the body force it is natural to consider ɛ and U as functions of the forcing amplitude F, l and ν [19]. When the force is specified, rather than the rms velocity, it is well nown (and easy to show) that the Stoes flow driven by the given force sets an upper limit for the dissipation rate; any other flow necessarily dissipates less energy. In terms of the F, l and ν, the maximal Stoes flow dissipation is ɛ F l /ν which may be interpreted as a laminar flow bound. It is interesting to note that in the extreme limits of H 1 forcing in (17) and ζ = 1 in (3), the scaling in this laminar upper limit is reproduced explicitly. Acnowledgments This research was funded in part by NSF grants PHY-044859, DMS-0405903 and PHY- 055534. CRD acnowledges additional support from a Humboldt Research Award, and NP is grateful for support from the Michigan Center for Theoretical Physics. Part of this wor was completed in the stimulating atmosphere of the Geophysical Fluid Dynamics Program at Woods Hole Oceanographic Institution. 1

APPENDIX A: INEQUALITIES For convenience, in this appendix we collect the mathematical estimates used here: (a) Jensen s inequality: If the real-valued function of a real variable θ(x) is convex, then for each real-valued function g θ( g ) θ g, (A1) where stands for averaging. In particular, for any nonnegative function g and any real number p [0, 1], g p g p. (A) (b) Hölder s inequality: ( φ(x) ψ(x) dx ) 1/p ( φ(x) p dx ψ(x) q dx) 1/q (A3) valid for all φ L p and ψ L q, where p and q 1 and 1 + 1 = 1. For an p q lp sequence (a ) and an l q sequence (b ) (where p and q are related as above) the discrete analogue of (A3) reads ( ) 1/p ( ) 1/q a b a p b q. (A4) An important case of (A4) (for p = q = ) is the Cauchy-Schwarz inequality φψ φ 1/ ψ 1/. (A5) (c) Interpolation inequalities between Sobolev spaces: Let 0 r < s < t and u H t. Note the algebraic identities r t s t r + t s r t r = s, t s t r + s r t r = 1. These interpolation estimates are the result of applying Hölder s inequality (A4) in Fourier space: u s = (l) s û = [ (l) r t s û ] [ t r (l) t s r û ] t r [ ] t s [ ] s r t r t r (l) r û (l) t û = u t s t r r u s r t r t. (A6) 13

[1] U. Frisch, Turbulence: The Legacy of A. N. Kolmogorov (Cambridge University Press, Cambridge, 1995). [] K. R. Sreenivasan, Phys. Fluids 7, 1048 (1984). [3] K. R. Sreenivasan, Phys. Fluids 10, 58 (1998). [4] C. R. Doering and C. Foias, J. Fluid Mech. 467, 89 (00). [5] C. R. Doering, B. Echardt, and J. Schumacher, J. Fluid Mech. 494, 75 (003). [6] N. P. Petrov, L. Lu, and C. R. Doering, J. Turbul. 6, Paper 17, 17 pp. (electronic) (005). [7] C. R. Doering and N. P. Petrov, in Progress in Turbulence. Edited by J. Peine, A. Kittel, S. Barth, M. Oberlac (Springer, Berlin, 005), pp. 11 18. [8] D. Queiros-Conde and J. C. Vassilicos, in Intermittency in Turbulent Flows (Cambridge, 1999). Edited by J. C. Vassilicos (Cambridge University Press, Cambridge, 001), pp. 136 167. [9] A. Staicu, B. Mazzi, J. C. Vassilicos, and W. van de Water, Phys. Rev. E (3) 67, 066306, 8 (003). [10] B. Mazzi and J. C. Vassilicos, J. Fluid Mech. 50, 65 (004). [11] L. Biferale, A. S. Lanotte, and F. Toschi, Phys. Rev. Lett. 9, 094503, 4 (004). [1] L. Biferale, M. Cencini, A. S. Lanotte, M. Sbragaglia, and F. Toschi, New J. Phys. 6, 37.1 (004). [13] P. Constantin and C. Foias, Navier-Stoes Equations (University of Chicago Press, Hyde Par, 1988). [14] C. R. Doering and J. D. Gibbon, Applied Analysis of the Navier-Stoes Equations (Cambridge University Press, Cambridge, 1995). [15] C. Foias, O. Manley, R. Rosa, and R. Temam, Navier-Stoes Equations and Turbulence (Cambridge University Press, Cambridge, 001). [16] C. Foias, O. P. Manley, and R. Temam, Phys. Lett. A 174, 10 (1993). [17] C. Foias, in Harmonic Analysis and Nonlinear Differential Equations (Riverside, CA, 1995) (Amer. Math. Soc., Providence, RI, 1997), vol. 08 of Contemp. Math., pp. 151 180. [18] C. Foias, O. P. Manley, R. M. S. Rosa, and R. Temam, C. R. Acad. Sci. Paris Sér. I Math. 333, 499 (001). [19] S. Childress, R. R. Kerswell, and A. D. Gilbert, Phys. D 158, 105 (001). 14