FIG 6 - uploaded by Gregory Eyink
Content may be subject to copyright.
͑ a ͒ H. L. Dryden, the distinguished American aero- dynamicist. His wind-tunnel experiments were a critical influence on Onsager’s theories of 3D turbulence. ͑ b ͒ A smoke visualization of turbulent flow in a wind tunnel, where the grid is at the top and the mean flow is downward. ͑ a ͒ courtesy of the National Archives and Records Administration, and ͑ b ͒ with permission of Thomas Corke and Hassan Nagib. 

͑ a ͒ H. L. Dryden, the distinguished American aero- dynamicist. His wind-tunnel experiments were a critical influence on Onsager’s theories of 3D turbulence. ͑ b ͒ A smoke visualization of turbulent flow in a wind tunnel, where the grid is at the top and the mean flow is downward. ͑ a ͒ courtesy of the National Archives and Records Administration, and ͑ b ͒ with permission of Thomas Corke and Hassan Nagib. 

Context in source publication

Context 1
... Z ͑ r ͒ , ͑ ␴ ͒ , are Lagrange multipliers to enforce constraints ͑ 15 ͒ , ͑ 17 ͒ , and ͑ 19 ͒ , respectively. The stream function satisfies the generalized mean-field equation − ⌬␺ ̄ ͑ r ͒ = 1 d ␴␴ exp ͕ − ␤ ̄ ͓ ␴␺ ̄ ͑ r ͒ − ␮ ͑ ␴ ͔͖͒ ; ͑ 21 ͒ see Miller ͑ 1990 ͒ and Robert ͑ 1990 ͒ . This theory is an application to 2D Euler of the method worked out by Lynden-Bell ͑ 1967 ͒ to describe gravitational equilibrium after “violent relaxation” in stellar systems. The Robert-Miller theory solves the problems discussed by Onsager, in the passage quoted above, with respect to the point-vortex assumption. The new theory incorporates infinitely many conservation laws of 2D Euler, although in our opinion that is not the critical difference. In fact, the point-vortex model, in the gener- ality considered by Onsager, also has infinitely many conserved quantities, i.e., the total number of vortices of a given circulation. 8 More importantly, the Robert- Miller theory includes information about the area of the vorticity level sets, which is lacking in the point-vortex model. As remarked by Miller et al. ͑ 1992 ͒ , the Joyce- Montgomery mean-field equation is formally recovered in a “dilute-vorticity limit” in which the area of the level sets shrinks to zero keeping the net circulation fixed. This corresponds well with the conditions suggested by Onsager for the validity of the point-vortex model that “vorticity is mostly concentrated in small regions.” The second main assumption invoked in Onsager’s theory is the ergodicity of the point vortex dynamics. This is a standard assumption invoked in justifying Gibbsian statistical theory. It has, however, proved to be false! Khanin ͑ 1982 ͒ showed that a part of the phase space of the system of N point vortices in the infinite plane consists of integrable tori. His proof used the fact that the three-vortex system is exactly integrable ͑ Novikov, 1975 ͒ . By adding additional vortices successively at further and further distances and using the fact that these additional vortices only weakly perturb the previous system, one can apply Kolmogorov-Arnold-Moser theory iteratively to establish integrability of the N -vortex system. Of course, statistical mechanics does not require strict ergodicity because macroscopic ob- servables are nearly constant over the energy surface. Thus any reasonable mixing over the energy surface will suffice to justify the use of a microcanonical ensemble. Of more serious concern are the possible slow time scales of this mixing. Onsager also worried about this point when he wrote to Lin that: “I still have to find out whether the processes anticipated by these considerations are rapid enough to play a dominant role in the evolution of vortex sheets, and just how the conservation of momentum will modify the conclusions.” Lundgren and Pointin ͑ 1977 ͒ performed numerical simulations of the point-vortex model with initial conditions corresponding to several local clusters of vortices at some distance from each other. The equilibrium theory predicts their final coalescence into a single large super- vortex. Instead, it was found that the clusters individu- ally reach a “local equilibrium,” not coalescing over the time scale of the simulation. Lundgren and Pointin argued theoretically that the vortices will eventually reach the equilibrium, single-vortex state. Similar metastable states of several large vortices have been seen in experiments with magnetically confined, pure electron col- umns and dubbed “vortex crystals” by Fine et al. ͑ 1995 ͒ and Jin and Dubin ͑ 2000 ͒ . These states have been explained by a regional maximum-entropy theory in which entropy is maximized assuming a fixed number of the strong vortices ͑ Jin and Dubin, 1998, 2000 ͒ . Clearly, Onsager’s ergodicity hypothesis is nontrivial and open to question. Despite these caveats, equilibrium theories of large- scale vortices have had some notable successes. Onsager himself considered decaying wake turbulence in an “in- finite vortex trail,” as he wrote to Lin. Indeed pp. 28–31 of Folder 11:129 contain detailed calculations, 9 similar to those in Lamb ͑ 1932 ͒ , Chap. 156, pp. 224 and 225. Analytical solutions of the mean-field Poisson-Boltzmann equation for vortex street geometries were later discovered ͑ Chow et al. , 1998; Kuvshinov and Shep, 2000 ͒ . Final states of freely decaying 2D Navier-Stokes simulations at high Reynolds number, started from fully turbulent initial conditions, have also been found to be in remarkable agreement with the predictions of the Joyce-Montgomery or sinh-Poisson mean-field equation ͑ Montgomery et al. , 1992, 1993 ͒ . Similar simulations started from a single band of vorticity, periodically modulated to induce Kelvin-Helmholtz instability, show good agreement with the generalized Robert-Miller theory ͑ Sommeria et al. , 1991 ͒ . In the limit of a thin initial band, the original Joyce-Montgomery mean-field theory is found to give identical results and agrees well with the simulations. Furthermore, the process is much as Onsager anticipated when he wrote to Lin: “ 1⁄4 the sheet will roll up and possibly contract into concentrated vortices in some places, and at the same time the remaining sections of the sheet will be stretched into feeble, more or less haphazard distributed discontinuities of velocity.” For further comparisons of mean-field equations with results of numerical simulations, see Yin et al. ͑ 2003 ͒ . A number of natural phenomena have been tentatively described by equilibrium vortex models of the sort proposed by Onsager. A fascinating example that was mentioned earlier is the Great Red Spot of Jupiter. For some recent work on this topic, see Turkington et al. ͑ 2001 ͒ and Bouchet and Sommeria ͑ 2002 ͒ . The second half of Onsager ͑ 1949d ͒ , titled “Turbu- lence,” deals with three-dimensional and fully developed turbulence. The second halves of the Pauling and Lin notes also discuss 3D turbulence. The Gibbsian statistical theory discussed in the first half of these documents does not describe a turbulent cascade process. As Onsager wrote at the end of the first section of Onsager ͑ 1949d ͒ on two dimensions, “How soon will the vortices discover that there are three dimensions rather than two? The latter question is important because in three dimensions a mechanism for complete dissipation of all kinetic energy, even without the aid of viscosity, is available.” Of course, it is no surprise that equilibrium statistical mechanics is inapplicable to a dissipative, irreversible process such as turbulence. More startling is Onsager’s conclusion that turbulent motion remains dissipative even in the limit as molecular viscosity tends to zero. In the Pauling note of March 1945, he had already made a similar assertion: “The energy is gradually divided up among ρ degrees of freedom, only for sufficiently large k គ the viscosity disposes of it for good; but it does not seem to matter much just how large this k គ is.” This remark was repeated at greater length in the Lin note of June 1945 as well: “We anticipate a mechanism of dissipation in which the role of the viscosity is altogether second- ary, as suggested by G. I. Taylor: a smaller viscosity is automatically compensated by a reduced micro- scale of the motion, in such a way that most of the vorticity will belong to the micro-motion, but only a small fraction of the energy.” Again, in the abstract of his APS talk in November, he wrote: “In actual liquids this subdivision of energy is in- tercepted by the action of viscosity, which destroys the energy more rapidly the greater the wave number. However, various experiments indicate that the viscosity has a negligible effect on the primary process; hence one may inquire about the laws of turbulent dissipation in an ideal fluid.” For good measure, similar remarks were made no less than four times in the published paper ͑ Onsager, 1949d ͒ . Considering the economy Onsager routinely prized in stating his results, it would appear that explaining the inviscid mechanism of energy dissipation in 3D turbulence was a chief preoccupation of Onsager’s work on statistical hydrodynamics. We can ask what evidence may have pushed Onsager in that direction. One reference in the 1949 paper was Dryden’s review article ͑ Dryden, 1943 ͒ on the statistical theory of turbulence. At the time, Dryden ͑ see Fig. 6 ͒ was a researcher in aerodynamics at the National Bu- reau of Standards in Washington, D.C. Starting in 1929, he published a series of papers on the measurement of turbulence in wind tunnels. A problem he had studied was the decay of nearly homogeneous and isotropic turbulence behind a wire-mesh screen. Dryden used hot- wire anemometry techniques to take accurate measurements of turbulence levels v in the tunnel, where v denotes the velocity fluctuation away from the mean. This permitted him to determine the rate of decay of the turbulent kinetic energy 10 Q as Q = − 1 ͑ d / dt ͒ v 2 , where d / dt denotes the convective derivative. Let V = ͑ v ͒ be the root-mean-square velocity fluctuation and L the spatial correlation length of the velocity, usually called the integral length scale. By simple dimensional ...

Citations

... These models are constrained by the nonvanishing of energy dissipation at infinite Reynolds numbers. Indeed, according to Onsager's theory of 1949, it implies that velocity differences (equation 15) should have at least some singularities such that h + 1 ≤ 1/3 (Eyink and Sreenivasan 2006;Eyink 2018). Therefore, the singularity spectrum models should have singularity exponents equal or less than ...
Preprint
Full-text available
The multifractal theory of turbulence is used to investigate the energy cascade in the Northwestern Atlantic ocean. The statistics of singularity exponents of velocity gradients computed from in situ measurements are used to show that the anomalous scaling of the velocity structure functions at depths between 50 ad 500 m has a linear dependence on the exponent characterizing the strongest velocity gradient, with a slope that decreases with depth. Since the distribution of exponents is asymmetric about the mode at all depths, we use an infinitely divisible asymmetric model of the energy cascade, the log-Poisson model, to derive the functional dependence of the anomalous scaling with dissipation. Using this model we can interpret the vertical change of the linear slope as a change in the energy cascade.
... In a Bose-Einstein condensate (BEC) under planar confinement, vortex bending is suppressed and vortex motion can become effectively two dimensional (2D) [1]. 2D quantum vortex systems support a rich phenomenology [2], including vortex clusters [3], the Kosterlitz-Thouless phase [4,5], and negative-temperature states [6][7][8][9][10][11][12][13]. A vortex closely bound with a vortex of opposite circulation (an antivortex) in a BEC form a vortex dipole that carries linear fluid momentum [14]. ...
Preprint
A quantum vortex dipole, comprised of a closely bound pair of vortices of equal strength with opposite circulation, is a spatially localized travelling excitation of a planar superfluid that carries linear momentum, suggesting a possible analogy with ray optics. We investigate numerically and analytically the motion of a quantum vortex dipole incident upon a step-change in the background superfluid density of an otherwise uniform two-dimensional Bose-Einstein condensate. Due to the conservation of fluid momentum and energy, the incident and refracted angles of the dipole satisfy a relation analogous to Snell's law, when crossing the interface between regions of different density. The predictions of the analogue Snell's law relation are confirmed for a wide range of incident angles by systematic numerical simulations of the Gross-Piteavskii equation. Near the critical angle for total internal reflection, we identify a regime of anomalous Snell's law behaviour where the finite size of the dipole causes transient capture by the interface. Remarkably, despite the extra complexity of the surface interaction, the incoming and outgoing dipole paths obey Snell's law.
... In two-dimensional (2D) classical fluids, giant coherent vortex structures can emerge from microscopic vortex excita- tions [1][2][3], as end states of turbulent fluid dynamics involving an inverse energy cascade [4,5]. The Great Red Spot in Jupiter's atmosphere [6][7][8]is a well-known example. ...
... The Great Red Spot in Jupiter's atmosphere [6][7][8]is a well-known example. Onsager explained this spontaneous formation of large-scale vortices from an equilibrium statistical mechanics point of view by studying a point-vortex model in a bounded domain [1,9]. The phenomenon stems from the bounded phase space, which supports negative temperature states that favor the spontaneous clustering of like-sign vortices. ...
... In order to support negative temperature states, a bounded domain is necessary. A well-defined thermodynamic limit for the vortex system in a bounded domain is found by a careful choice of scaling [1,67,68]. In the clustered phase, the energy H ∼ N 2 is due to the sum over all vortex pairs. ...
Article
Clustering of like-sign vortices in a planar bounded domain is known to occur at negative temperature, a phenomenon that Onsager demonstrated to be a consequence of bounded phase space. In a confined superfluid, quantized vortices can support such an ordered phase, provided they evolve as an almost isolated subsystem containing sufficient energy. A detailed theoretical understanding of the statistical mechanics of such states thus requires a microcanonical approach. Here we develop an analytical theory of the vortex clustering transition in a neutral system of quantum vortices confined to a two-dimensional disk geometry, within the microcanonical ensemble. The choice of ensemble is essential for identifying the correct thermodynamic limit of the system, enabling a rigorous description of clustering in the language of critical phenomena. As the system energy increases above a critical value, the system develops global order via the emergence of a macroscopic dipole structure from the homogeneous phase of vortices, spontaneously breaking the Z2 symmetry associated with invariance under vortex circulation exchange, and the rotational SO(2) symmetry due to the disk geometry. The dipole structure emerges characterized by the continuous growth of the macroscopic dipole moment which serves as a global order parameter, resembling a continuous phase transition. The critical temperature of the transition, and the critical exponent associated with the dipole moment, are obtained exactly within mean-field theory. The clustering transition is shown to be distinct from the final state reached at high energy, known as supercondensation. The dipole moment develops via two macroscopic vortex clusters and the cluster locations are found analytically, both near the clustering transition and in the supercondensation limit. The microcanonical theory shows excellent agreement with Monte Carlo simulations, and signatures of the transition are apparent even for a modest system of 100 vortices, accessible in current Bose-Einstein condensate experiments.
... On a longer (secular) timescale, " collisions " 1 between stars or between point vortices come into play and drive the system towards a statistical equilibrium state described by the Boltzmann distribution. This statistical equilibrium state was conjectured by Ogorodnikov [8] in the case of stellar systems and by On- sager [9, 10] and Montgomery and Joyce [11] in the case of 2D point vortices. Actually, for collisional stellar systems such as globular clusters the relaxation towards the Boltzmann statistical equilibrium state is hampered by the evaporation of stars [12] and by the gravothermal catastrophe [13, 14]. ...
Article
We present a brief derivation of the kinetic equation describing the secular evolution of point vortices in two-dimensional hydrodynamics, by relying on a functional integral formalism. We start from Liouville's equations which describe the exact dynamics of a two-dimensional system of point vortices. At the order , the evolution of the system is characterised by the first two equations of the BBGKY hierarchy involving the system's 1-body distribution function and its 2-body correlation function. Thanks to the introduction of auxiliary fields, these two evolution constraints may be rewritten as a functional integral. When functionally integrated over the 2-body correlation function, this rewriting leads to a new constraint coupling the 1-body distribution function and the two auxiliary fields. Once inverted, this constraint provides, through a new route, the closed non-linear kinetic equation satisfied by the 1-body distribution function. Such a method sheds new lights on the origin of these kinetic equations complementing the traditional derivation methods.
... The idea comes back to Onsager [23], and has been mostly developed during the nineties after the work of Miller-Robert-Sommeria [17,27], see also Refs. [12,6,3] and references therein. Importantly, the theory predicts that the contribution of small scale fluctuations to the total energy are negligible in the two-dimensional case. ...
... The dynamics in Eqs. (6), (13) and (14) can also be written in terms of Ψ and Φ: ...
... Provided that the velocity and height fields remain differentiable, the shallow water dynamics in in Eqs. (6), (13) and (14) conserves the total energy ...
Article
Full-text available
The aim of this paper is to use large deviation theory in order to compute the entropy of macrostates for the microcanonical measure of the shallow water system. The main prediction of this full statistical mechanics computation is the energy partition between a large scale vortical flow and small scale fluctuations related to inertia-gravity waves. We introduce for that purpose a discretized model of the continuous shallow water system, and compute the corresponding statistical equilibria. We argue that microcanonical equilibrium states of the discretized model in the continuous limit are equilibrium states of the actual shallow water system. We show that the presence of small scale fluctuations selects a subclass of equilibria among the states that were previously computed by phenomenological approaches that were neglecting such fluctuations. In the limit of weak height fluctuations, the equilibrium state can be interpreted as two subsystems in thermal contact: one subsystem corresponds to the large scale vortical flow, the other subsystem corresponds to small scale height and velocity fluctuations. It is shown that either a non-zero circulation or rotation and bottom topography are required to sustain a non-zero large scale flow at equilibrium. Explicit computation of the equilibria and their energy partition is presented in the quasi-geostrophic limit for the energy-enstrophy ensemble. The possible role of small scale dissipation and shocks is discussed. A geophysical application to the Zapiola anticyclone is presented.
... The mathematical underpinnings of the dynamics of pointlike vortices in classical fluids were established in the nineteenth century [1][2][3][4][5]. In the 1940s, Lars Onsager realized that the equations of motion describing such vortices in two dimensions are mathematically equivalent to Hamilton's equations of motion for particles moving in one spatial dimension and that a large collection of them could be treated with the machinery of statistical mechanics [6, 7]. His motivation was to develop understanding of fluid turbulence by describing the statistical properties of the turbulent fluid in terms of a collection of point-vortex particles. ...
Article
Full-text available
We have studied numerically the Hamiltonian dynamics of two same-sign point vortices in an effectively two-dimensional, harmonically trapped Bose-Einstein condensate. We have found in the phase space of the system an impenetrable wall that divides the dynamics into two distinct and exhaustive types. In the two-dimensional position-coordinate space, the first type corresponds to intersecting single-vortex orbits and the second type to orbits that have no points in common. The two types are also easily distinguished in the two-dimensional space spanned by the radial and angular velocities of the vortices: in the first type, both single-vortex orbits are the same simple loop in this two-dimensional space, whereas in the second type the two orbits constitute two nonintersecting loops. The phase-space-dividing wall is distinct from the bifurcation curve of rigidly rotating states found by Navarro et al. [Phys. Rev. Lett. 110, 225301 (2013)].
... Twodimensional systems have attracted particular interest due to a prediction of an inverse energy cascade [28, 29] from small to large spatial scales which originates from the theory of classical fluid turbulence. Using a statistical model of point-vortices, Onsager [30] predicted for such systems emergence of large-scale vortex structures, such as as those seen in geophysical systems [31] . Similarly , in 2D QT, the inverse cascade is anticipated to lead to the clustering of like-sign vortices into large-scale Onsager vortex structures. ...
Article
Full-text available
We study computationally dynamics of quantised vortices in two-dimensional superfluid Bose-Einstein condensates confined in highly oblate power-law traps. We have found that the formation of large scale Onsager vortex clusters prevalent in steep-walled traps is suppressed in condensates confined by harmonic potentials. However, the shape of the trapping potential does not appear to adversely affect the evaporative heating efficiency of the vortex gas. Instead, the suppression of Onsager vortex formation in harmonic traps can be understood in terms of the energy of the vortex configurations. Furthermore, we find that the vortex-antivortex pair annihilation that underpins the vortex evaporative heating mechanism requires the interaction of at least three vortices. We conclude that experimental observation of Onsager vortices should be the most apparent in flat or inverted-bottom traps.
... The energy necessary for active motion can be supplied by an external spatio-temporal modulated field or by energy input from a local " self-generated " force [5] ( small objects can swim by generating concentration and other gradients around them [15, 16]). Active Brownian motion can be encountered in a large variety of phenomena, including protein diffusion [17], the motion of swimming bacteria [18, 19] and artificial nanoscale swimming devices [16, 20], turbulent flows [21, 22] or even in the processes leading to collective opinion formation [23]. Self-generated forces can also be optically tuned through photo-chemical [24] or photo-thermal [25, 26] mechanisms. ...
Article
Full-text available
A two-dimensional periodic optical force field, which combines conservative dipolar forces with vortices from radiation pressure, is proposed in order to influence the diffusion properties of optically susceptible nano-particles. The different deterministic flow patterns are identified. In the low noise limit, the diffusion coefficient is computed from a Mean First Passage Time (MFPT) and the Most Probable Escape Paths (MPEP) are identified for those flow patterns which possess an stable stationary point. Numerical simulations of the associated Langevin equations show remarkable agreement with the analytically deduced expressions. Modifications of the force field are proposed so that a wider range of phenomena could be tested.
... It allows to reduce the study of the large scale organization to a few parameters, without describing the full complexity of the dynamics involving a huge number of degrees of freedom. The original idea to use statistical mechanics arguments to describe selforganization of 2D flows comes from L. Onsager himself in the framework of point vortex models [3,4]. A statistical mechanics theory for the continuous Euler dynamics has been proposed by Miller, Robert, Sommeria [5-8] (MRS hereafter), which has led to several successful applications to geophysical flows [9][10][11][12]. ...
... It allows to reduce the study of the large scale organization to a few parameters, without describing the full complexity of the dynamics involving a huge number of degrees of freedom. The original idea to use statistical mechanics arguments to describe selforganization of 2D flows comes from L. Onsager himself in the framework of point vortex models [3,4]. A statistical mechanics theory for the continuous Euler dynamics has been proposed by Miller, Robert, Sommeria [5][6][7][8] (MRS hereafter), which has led to several successful applications to geophysical flows [9][10][11][12]. ...
Article
Full-text available
Understanding the relaxation of a system towards equilibrium is a longstanding problem in statistical mechanics. Here we address the role of long-range interactions in this process by considering a class of two-dimensional or geophysical flows where the interaction between fluid particles varies with the distance as $\sim$$r^{$\alpha$--2}$ with $\alpha$ \textgreater{} 0. Previous studies in the Euler case $\alpha$ = 2 had shown convergence towards a variety of quasi-stationary states by changing the initial state. Unexpectedly, all those regimes are recovered by changing $\alpha$ with a prescribed initial state. For small $\alpha$, a coarsening process leads to the formation of a sharp interface between two regions of homogenized $\alpha$-vorticity; for large $\alpha$, the flow is attracted to a stable dipolar structure through a filamentation process.
... It is difficult and quite open to establish mathematically rigorous conditions for justifying the application of statistical mechanics. For example, according to literatures (see, e.g., Eyink & Sreenivasan 2006, and references therein), ergodicity, being sufficient, may not be trivially satisfied but may neither be necessary; and, the mixing time scale could be hard to estimate for evaluating the closeness of physical relevance of the equilibrium ensemble. However there is a trivial bottom line that is assumed to be met, that is, all modes should be directly or indirectly connected by forming the interacting triads to define a system. ...
Article
Full-text available
Left- and right-handed helical modes' statistical absolute equilibria appear \textit{separately}. If both chiral sectors present, one can dominate around its positive pole, which is relevant to the nearly maximally helical (force free for magnetic field) states of turbulence. Pure magnetodynamics (PMD, or electron magnetohydrodynamics --- EMHD), pure hydrodynamics (PHD), and, single-fluid and two-fluid MHDs are studied. Relevant documented data and issues of cascade properties, and, helical and nonhelical dynamos are revisited. We also discuss new scenarios, such as PMD inverse magnetic helicity and forward energy cascades, and, the continuous transition from completely-inverse to partly-inverse-and-partly-forward and to completely-forward energy transfers in PHD.