Geometric hydrodynamics and infinite-dimensional Newton’s equations

By Boris Khesin, Gerard Misiołek, and Klas Modin

To the memory of Vladimir Arnold and Jerry Marsden, pioneers of geometric hydrodynamics, who left in 2010, ten years ago.

Abstract

We revisit the geodesic approach to ideal hydrodynamics and present a related geometric framework for Newton’s equations on groups of diffeomorphisms and spaces of probability densities. The latter setting is sufficiently general to include equations of compressible and incompressible fluid dynamics, magnetohydrodynamics, shallow water systems and equations of relativistic fluids. We illustrate this with a survey of selected examples, as well as with new results, using the tools of infinite-dimensional information geometry, optimal transport, the Madelung transform, and the formalism of symplectic and Poisson reduction.

The only way to get rid of dragons is to have one of your own.

Evgeny Schwartz, The Dragon

1. Introduction

The Euler equations of hydrodynamics describe motions of an incompressible and inviscid fluid occupying a fixed domain (with or without boundary). In the 1960s V. I. Arnold discovered that these equations are precisely the geodesic equations of a right-invariant metric on the group of diffeomorphisms preserving the volume element of the domain Reference 5. This beautiful observation, combining the early work of Hadamard on geodesic flows on surfaces with the dynamical systems ideas of Poincaré and Kolmogorov and using analogies with classical mechanics of rigid bodies, inspired many researchers—one of the first was J. E. Marsden. Their combined efforts led to remarkable developments, such as formulation of new stability criteria for fluid motions Reference 4Reference 7Reference 23Reference 24, explicit calculation of the associated Hamiltonian structures and first integrals Reference 8Reference 53Reference 54, development of symplectic reduction methods Reference 52Reference 53, introduction of Riemannian geometric techniques to the study of diffeomorphism groups including explicit computations of curvatures, conjugate points, diameters Reference 5Reference 56Reference 57Reference 70, detailed studies of regularity properties of the solution maps of the fluid equations in Lagrangian and Eulerian coordinates Reference 20Reference 21, construction of similar configuration spaces for other partial differential equations of hydrodynamic origin Reference 8Reference 33, etc.

In this paper, based on the research pioneered and developed by Arnold, Marsden and many others, we present a broad geometric framework which includes an infinite-dimensional generalization of the classical Newton’s equations of motion to the setting of diffeomorphism groups and spaces of probability densities. This approach has a wide range of applicability and covers a large class of important equations of mathematical physics. Our goal is twofold. We start by presenting a concise survey of various geodesic and Newton’s equations, thus introducing the reader to the rapidly expanding field of geometric hydrodynamics, and revisiting a few standard examples from the point of view advocated here. We then also include a number of selected new results to illustrate the flexibility and utility of this approach.

We focus primarily on the geometric aspects and emphasize formal procedures leaving until the end analytic issues which in most cases can be resolved using standard methods once an appropriate functional-analytic setting (e.g., Fréchet, Hölder, or Sobolev) is adopted. The corresponding tame Fréchet framework is described in more detail in Appendix B. Our main tools include the Wasserstein metric of optimal transport, the infinite-dimensional analogue of the Fisher–Rao information metric, the Madelung transform, and the formalism of symplectic and Poisson reduction, all of which are defined in the paper. The early sections should be accessible to mathematicians with only general background in geometry. In later sections some acquaintance with the basic material found, for example, in the monographs Reference 7Reference 8Reference 51 will be helpful.

Needless to say, it is not possible to give a comprehensive survey of such a vast area of geometric hydrodynamics in such a limited space, therefore our emphasis on certain topics and the choice of examples are admittedly subjective. (The epigraph to the paper is our take on the Laws of Nature, on the tamed structures discussed below, as well as a counterpoint to the beautiful epigraph in the monograph Reference 14, quoted here in the footnote.⁠Footnote1) We nevertheless hope that this paper provides a flavour of some of the results in this beautiful area pioneered by V. Arnold and J. Marsden.

1

“There once lived a man
who learned how to slay dragons
and gave all he possessed
to mastering the art.

After three years
he was fully prepared but,
alas, he found no opportunity
to practise his skills.”

—Dschuang Dsi

1.1. Geodesics and Newton’s equations: finite-dimensional examples

A curve is a geodesic in a Riemannian manifold if it satisfies the equation of geodesics, namely

where stands for the covariant derivative on and the dot denotes the -derivative. If the Riemannian manifold is flat, then the geodesic equation becomes the familiar in any local Euclidean coordinates on .

From the point of view of classical mechanics, the geodesic equation Equation 1.1 describes motions of a system driven only by its kinetic energy. More general systems may depend also on a potential energy. Indeed, if is a configuration space of some physical system (a Riemannian manifold) and represents its potential energy (a differentiable function), then satisfies Newton’s equations

One of the classical examples of Newton’s equations is the -body system in . Introducing coordinates one can regard as the configuration space of the system. If the bodies have masses , then their kinetic energy is and hence corresponds to a Riemannian metric on of the form . If denotes the gravitational constant, then the potential energy is given by the expression

which becomes infinite on the diagonals . The corresponding Lagrangian function is , while the total energy of the system (its Hamiltonian) is . We shall revisit this system in a fluid dynamical context below.

Another classical example is provided by the C. Neumann problem Reference 63 describing the motion of a single particle on an -sphere under the influence of a quadratic potential energy. Here, the configuration space is the unit sphere in while the phase space is the tangent bundle of the sphere. The potential energy of the system is given by , where and is a positive-definite symmetric matrix. As before, the Lagrangian function is the difference of the kinetic and the potential energies

The C. Neumann system is related to the geodesic flow on the ellipsoid defined by the equation , see, e.g., Moser Reference 62, Sec. 3. The corresponding Hamiltonian system on the cotangent bundle is integrable and, if the eigenvalues of are all different, then the first integrals, expressed in canonical coordinates and , are explicitly given by

where and are the components of and with respect to the eigenbasis of . We will see that an infinite-dimensional analogue of the C. Neumann problem naturally arises in the context of information geometry, while its integrability in infinite dimensions remains an intriguing open problem.

1.2. Three motivating examples from hydrodynamics

We now make a leap from finite to infinite dimensions. Our aim is to show that many well-known PDEs of hydrodynamical pedigree can be cast as Newton’s equations on infinite-dimensional manifolds. Indeed, groups of smooth diffeomorphisms arise naturally as configuration spaces of compressible and incompressible fluids. We begin with three famous examples. Consider a connected compact Riemannian manifold of dimension (for our purposes may be a domain in ) and assume that it is filled with an inviscid fluid (either a gas or a liquid). When the group of diffeomorphisms of is equipped with an metric (essentially, the metric corresponding to the fluid’s kinetic energy, as we shall discuss later) its geodesics describe the motions of noninteracting particles in whose velocity field satisfies the inviscid Burgers equation

When the metric is restricted to the subgroup of diffeomorphisms of that preserve the Riemannian volume form , then its geodesics describe the motions of an ideal (that is, inviscid and incompressible) fluid in whose velocity field satisfies the incompressible Euler equations

Here is the pressure function whose gradient is defined uniquely by the divergence-free condition on the velocity field and can be viewed as a constraining force. (If has a nonempty boundary, then is also required to be tangent to ).

As we shall see below, both of the above equations turn out to be examples of equations of geodesics on diffeomorphism groups with Lagrangians given by the corresponding kinetic energy. However, the Lagrangian in our next example will include also a potential energy. Consider the equations of a compressible (barotropic) fluid describing the evolution of a velocity field and a density function , namely

These equations can be interpreted as Newton’s equations on the full diffeomorphism group of . In this case the pressure is a prescribed function of density and this dependence, called the equation of state, determines the fluid’s potential energy. In sections below we shall also consider general equations with the term replaced by the gradient , where denotes an arbitrary thermodynamical work function; cf. section 4.3.

1.3. Riemannian metrics and their geodesics on spaces of diffeomorphisms and densities

Let us next see how differential geometry of diffeomorphism groups manifests itself in the above equations. Given a Riemannian manifold , we equip the group of all diffeomorphisms of with a (weak) Riemannian metric and a natural fibration.

Namely, assume that the Riemannian volume form has the unit total volume (or total mass) and regard it as a reference density on . Now consider the projection of diffeomorphisms onto the space of (normalized) smooth densities on . The diffeomorphism group is fibered over by means of this projection as follows: the fiber over is the subgroup of -preserving diffeomorphisms, while the fiber over a volume form consists of all diffeomorphisms that push to , or, equivalently, . (Note that diffeomorphisms from act transitively on smooth normalized densities, according to Moser’s theorem.) In other words, two diffeomorphisms and belong to the same fiber if and only if for some diffeomorphism .

Remark 1.1.

It is worth comparing “the functional dimensions” of the fiber and the base . The space of densities can be thought of as the space of functions of variables, where . On the other hand, the group consists of

i)

isometries in dimension (e.g., for it is ),

ii)

symplectic diffeomorphisms in dimension (e.g., for these are Hamiltonian diffeomorphisms, locally described by a function of 2 variables), and

iii)

in dimensions these diffeomorphisms are subject to the only constraint on the Jacobian: (i.e., one equation on functions of variables).

Therefore, in the fibration the fiber is small compared to the base in dimension , the fiber and the base are about the same size in dimension , and the fiber becomes much bigger than the base starting with dimension .

Definition 1.2.

Now define an -metric on by the formula

where is a tangent vector at the point , i.e., a map such that for each , while stands for the pointwise Riemannian product at the point .

One can see that for a flat manifold this is a flat metric on , as it is the -metric on diffeomorphisms regarded as vector functions . This metric is right-invariant for the -action (but not the -action): for , since the change of coordinates leads to the factor in the integrand.

Remark 1.3.

Consider the following optimal mass transport problem: Find a map that pushes the measure forward to another measure of the same total volume and attains the minimum of the -cost functional among all such maps ( denotes here the Riemannian distance function on ). The minimal cost of transport defines the following Kantorovich–Wasserstein distance on the space of densities :

The mass transport problem admits a unique solution for Borel maps and densities (defined up to measure-zero sets), called the optimal map , see, e.g., Reference 12Reference 55Reference 77. In the smooth setting the Kantorovich–Wasserstein distance is generated by a (weak) Riemannian metric on the space of smooth densities Reference 9Reference 64, which we call the Wasserstein–Otto metric and describe in detail in section 2.2. Thus both and can be regarded as infinite-dimensional Riemannian manifolds for the and Wasserstein–Otto metrics, respectively.

Remark 1.4.

Later we will see (following Reference 64) that the corresponding projection is a Riemannian submersion from the diffeomorphism group onto the density space , i.e., the map respecting the above metrics. Recall that for two Riemannian manifolds and a submersion is a smooth map which has a surjective differential and preserves lengths of horizontal tangent vectors to . For a bundle this means that on there is a distribution of horizontal spaces orthogonal to fibers and projecting isometrically to the tangent spaces to . Geodesics on can be lifted to horizontal geodesics in , and the lift is unique for a given initial point in .

Note also that horizontal (i.e., normal to fibers) spaces in the bundle consist of right-translated gradient fields. In short, this follows from the Hodge decomposition for vector fields on : any vector field decomposes uniquely into the sum of a divergence-free field and a gradient field , which are -orthogonal to each other, . The vertical tangent space at the identity coincides with , while the horizontal space is . The vertical space (tangent to a fiber) at a point consists of , divergence-free vector fields right-translated by the diffeomorphism , while the horizontal space is given by the right-translated gradient fields, . The -type metric on horizontal spaces for different points of the same fiber projects isometrically to one and the same metric on the base, due to the -invariance of the metric. Now the Riemannian submersion property follows from the observation that the Wasserstein–Otto metric is Riemannian and is generated by the metric on gradients; see Reference 9.

Example 1.5.

Geodesics in the full diffeomorphism group with respect to the above -metric have a particularly simple description for a flat manifold ; cf. Reference 11Reference 20. In that case the group is locally (a dense subset of) the -space of vector-functions , and hence is flat, while its geodesics are straight lines. If is the velocity field of the flow in defined by , then the geodesic equation becomes , which in turn is equivalent to the inviscid Burgers equation

Furthermore, from the viewpoint of exterior geometry, the Euler equation can be regarded as an equation with a constraining force acting orthogonally to the submanifold of volume-preserving diffeomorphisms and keeping the geodesics confined to that submanifold.

Remark 1.6.

Analytical studies of the differential geometry of the incompressible Euler equations began with the paper of Ebin and Marsden Reference 20 and continued with Reference 21Reference 57Reference 70 and others. The approach via generalized flows was proposed by Brenier Reference 11. Many aspects of this approach to the group of all diffeomorphisms and their relation to the Kantorovich–Wasserstein space of densities and problems of optimal mass transport are discussed in Reference 47Reference 77Reference 82. There is also a finite-dimensional matrix version of the submersion framework and decomposition of diffeomorphsims; see Reference 12. In the finite-dimensional optimal mass transport on discussed in Reference 59 the probability distributions are multivariate Gaussians and the transport maps are linear transformations. The corresponding dynamics turned out to be closely related to many finite-dimensional flows studied in the literature: Toda-lattice, isospectral flows, and an entropy gradient interpretation of the Brockett flow for matrix diagonalization. A sub-Riemannian version of the exterior geometry of with vector fields tangent to a bracket generating distribution in , as well as a nonholonomic version of Moser lemma, is described in Reference 1Reference 36. For a symplectic reduction formulation to the above Riemannian submersion see section 3.2.

1.4. First examples of Newton’s equations on diffeomorphism groups

Example 1.7 (Shallow water equation as a Newton’s equation).

We next proceed to describe Newton’s equations on the diffeomorphism group . To this end we consider the case of a potential on which depends only on the density carried by a diffeomorphism , i.e., the potential for is a pullback for the projection , where as we take a simple quadratic function

on the space of densities. It turns out that with this potential we obtain shallow water equations. There are several equivalent formulations, depending on the functional setting.

Proposition 1.8.

Newton’s equations with respect to the -metric Equation 2.1 and the potential Equation 1.5 take the following forms:

on

where ;

the shallow water equations on

where is the horizontal velocity field and is the water depth;

for the gradient velocity they assume the Hamilton–Jacobi form

Remark 1.9.

The latter form can be regarded as an equation on . Since is a quadratic function, equations Equation 1.7 can be interpreted as a Hamiltonian form of an infinite-dimensional harmonic oscillator with respect to the Wasserstein–Otto metric Equation 2.8. We will prove this theorem in a more general setting of a barotropic fluid (cf. equation Equation 1.4) with an arbitrary potential in section 4.3; here .

Example 1.10 (The -body problem as a Newton’s equation).

Newton’s law of gravitation states that for a body with mass distribution , the associated potential is , where is the gravitational constant. Following the above framework, the potential function on is given by

where is a (suitably defined) inverse Laplacian with appropriate boundary conditions.

The corresponding fluid system is described by

Thus, we have arrived at a fluid dynamics formulation of a continuous Newton mass system under the influence of gravity: a “fluid particle” positioned at experiences a gravitational pull corresponding to the potential . In particular, if we obtain the well-known Green’s function for the Laplacian

We now wish to study weak solutions to these equations where the mass distribution is replaced by an atomic measure

for point masses positioned at . The differential-geometric setting is as follows. We have a Riemannian metric on (the Wasserstein—Otto metric) and a potential function on (the Newton potential). The group acts on the (finite-dimensional) manifold of atomic measures with particles. Clearly, we have . The isotropy subgroup for this action on is

Although the horizontal distribution is not defined rigorously, it is formally given by vector fields with support on . With this notion of horizontality, the projection , given by , is a Riemannian submersion with respect to the weighted Riemannian structure on , given by . For , substituting the atomic measure into the formula Equation 1.8 and using the Green’s function for gives

The resulting finite-dimensional Riemannian structure together with this potential function defines the kinetic and potential energies giving rise to the -body problem.

Remark 1.11.

In the wake of Arnold’s work, various approaches to infinite-dimensional generalizations of Newton’s equations Equation 1.2 have been considered in special settings. Those of perhaps most interest from our point of view were proposed by Smolentsev Reference 72Reference 73, who used diffeomorphism groups to describe the motions of a barotropic fluid, and by Ebin Reference 19, who used a similar framework to study, among others, the incompressible limit of slightly compressible fluids. In the early 1980s, Doebner, Goldin, and Sharp Reference 17Reference 27 began to develop links between representations of diffeomorphism groups, ideal fluids, and nonlinear quantum systems, revisiting in the process the classical transform of Madelung Reference 48Reference 49. More recently, motivated by the problems of optimal transport, von Renesse Reference 79 used it to relate the Schrödinger equations with a variant of Newton’s equations defined on the space of probability measures (see Section 9 below for details). A similar objective, but driven partly by motivation from information geometry and statistics, can be found in a recent paper of Molitor Reference 60.

In what follows we will systematically describe how one can conveniently study various equations of mathematical physics, including all the examples listed in Table 1, from a unified point of view as certain Newton’s equations. Our goal is to present a rigorous infinite-dimensional geometric framework that unifies Arnold’s approach to incompressible and inviscid hydrodynamics and its relatives with various generalizations of Newton’s equations Equation 1.2 such as those mentioned above, to provide a very general setting for systems of hydrodynamical origin on diffeomorphism groups and spaces of probability densities. We will also survey the setting of the Hamiltonian reduction, which establishes a correspondence between various representations of these equations.

Remark 1.12.

More precisely, given a compact -dimensional manifold , we will equip the group of diffeomorphisms and the space of nonvanishing probability densities with the structures of smooth infinite-dimensional manifolds (see Appendix B for details) and study Newton’s equations on these manifolds viewed as the associated configuration spaces.

As a brief preview of what follows, let be the subgroup of diffeomorphisms preserving the Riemannian volume form of . Consider the fibration of the group of all diffeomorphisms over the space of densities

discussed by Moser Reference 61, whose cotangent bundles and are related by a symplectic reduction; cf. Section 3 below. Moser’s construction can be used to introduce two different algebraic objects: the first is obtained by identifying with the left cosets

and the second by identifying it with the right cosets

In this paper we will make use of both identifications.

In order to define Newton’s equations on and , and to investigate their mutual relations, we will choose Riemannian metrics on both spaces so that the natural projections corresponding to Equation 1.9 or Equation 1.10 become (infinite-dimensional) Riemannian submersions. We will consider two such pairs of metrics. In Section 2, using left cosets, we will study a noninvariant -metric on together with the Wasserstein–Otto metric on . In Section 7, using right cosets, we will focus on a right-invariant metric on and the Fisher–Rao information metric on . Extending the results of von Renesse Reference 79, we will then derive in Section 9 various geometric properties of the Madelung transform. This will allow us to represent Newton’s equations on as Schrödinger-type equations for wave functions.

1.5. Other related equations

Newton’s equations for fluids discussed in the present paper are assumed to be conservative systems with a potential force. However, the subject concerning Newton’s equations is broader, and we mention briefly two topics related to nonconservative Newton’s equations for compressible and incompressible fluids that are beyond the scope of this paper.

First, observe that the dissipative term in the viscous Burgers equation

can be viewed as a (linear) friction force while the equation itself can be seen as Newton’s equation on with a nonpotential force. Similarly, observe that the Navier–Stokes equations of a viscous incompressible fluid

can be seen as Newton’s equations on with a nonpotential friction force. There is a large literature treating the Navier–Stokes equations within a stochastic framework where the geodesic setting of the Euler equations is modified by adding a random force which acts on the fluid; see Reference 28Reference 31.

The second topic is related to a recently discovered flexibility and nonuniqueness of weak solutions of the Euler equations. The constructions in Reference 16Reference 69Reference 71 exhibit compactly supported weak solutions describing a moving fluid that comes to rest as . Such constructions can be understood by introducing a special forcing term (sometimes referred to as the “black noise”) into the equations , and require that it is -orthogonal to all smooth functions.” (More precisely, one constructs a family of solutions with increasingly singular and oscillating force and the black noise is a residual forcing observed in the limit; cf. Reference 76.) Using the standard definition of a weak solution, this force is thus not detectable upon multiplication by smooth test functions and hence the existence of such solutions to the Euler equations becomes less surprising. Constructions of similar weak solutions to other PDEs rely on intricate limiting procedures involving possibly more singular and less detectable forces. The study of the geometry of Newton’s equations with black noise on diffeomorphism groups seems to be a promising direction of future research.

Remark 1.13.

We should mention that, in addition to Newton’s equations, there is another class of natural evolution equations on Riemannian manifolds given by the gradient flows

where a given potential function determines velocity rather than acceleration. An interesting example can be found in Reference 64 where the heat flow on is described as the gradient flow of the relative entropy functional providing a geometric interpretation of the second law of thermodynamics; cf. Remark 9.8 on its relation to a Hamiltonian setting.

1.6. An overview and main results

The goal of this paper is twofold. First, we present a survey of the differential geometric approach to several hydrodynamical equations emphasizing the setting of Newton’s equations. Second, we describe new results obtained by implementing this tool.

Here are some highlights of this paper, where the survey topics are intertwined with new contributions: geometry of the Euler equations as geodesic equations along with their Hamiltonian formulation; Riemannian geometry of the spaces of diffeomorphisms and densities and their relation to problems of optimal mass transport; Newton equations in infinite dimensions and their appearance in the geometry of compressible fluids; semidirect product groups in relation to compressible fluids and magnetohydrodynamics; Fisher–Rao geometry on the spaces of densities and diffeomorphisms; geometric properties of the Madelung transform; Casimirs of compressible and incompressible fluids and magnetohydrodynamics. We also recall briefly the symplectic and Poisson reductions in relation to diffeomorphism groups. In more detail:

(1)

Following Reference 72 and Reference 19, we revisit the case of the compressible barotropic Euler equations as a Poisson reduction of Newton’s equations on with the symmetry group and show that the Hamilton–Jacobi equation of fluid mechanics corresponds to its horizontal solutions in section 4.3. We then describe the framework of Newton’s equations for fully compressible (nonbarotropic) fluids in section 6.1 and magnetohydrodynamics in section 6.2.

(2)

After reviewing the semidirect product approach to these equations we relate it to our approach in section A.2. We point out that the Lie–Poisson semidirect product algebra associated with the compressible Euler equations appears naturally in the Poisson reduction setting . We then show that the semidirect product structure is consistent with the symplectic reduction at zero momentum for ; see Section 5 and Appendix A.

(3)

We develop a reduction framework for relativistic fluids in section 6.3 and show how the relativistic Burgers equation arises in this context. We relate it to the relativistic approaches in optimal transport in Reference 13 and ideal hydrodynamics in Reference 32Reference 42.

(4)

Along with the and the Wasserstein–Otto geometries, we also describe the geometry associated with the Sobolev and the Fisher–Rao metrics; see Section 7. We show that infinite-dimensional Neumann systems are (up to time rescaling) Newton’s equations for quadratic potentials in these metrics (in suitable coordinates the Fisher information functional is an example of such a potential); see section 8.2.

(5)

Using the approach presented in this paper, we derive stationary solutions of the Klein–Gordon equation and show that they satisfy a stationary infinite-dimensional Neumann problem; see section 8.3. We also show that the generalized two-component Hunter–Saxton equation is a Newton’s equation in the Fisher–Rao setting; see section 9.5.

(6)

We review the properties of the Madelung transform which relates linear and nonlinear Schrödinger equations to Newton’s equations on and can be used to describe horizontal solutions to Newton’s equations on with -invariant potentials, see Section 9 and Reference 40Reference 79 as well as the so-called Schrödinger smoke Reference 15.

(7)

Finally, we describe the Casimirs for compressible barotropic fluids, compressible and incompressible magnetohydrodynamics; see Section 10.

Notations.

Unless indicated otherwise stands for a compact oriented Riemannian manifold. The spaces of smooth -forms on are denoted by , the spaces of smooth vector fields by , and those of smooth functions by . Given a Riemannian metric on , the symbol stands for the gradient as well as for the covariant derivative of . The Riemannian volume form is denoted by and is assumed to be normalized: . To simplify notation, we will often use typical vector calculus conventions and . The Lie derivative along a vector field will be denoted by . In our computations we will assume all the functionals to be differentiable with variational derivatives belonging to the corresponding smooth duals, unless indicated otherwise.

A Riemannian metric on defines an isomorphism between the tangent and cotangent bundles. For a vector field on we will denote by the corresponding -form on , namely . As usual, the inverse map will be denoted by . The pullback and pushforward maps of a tensor field by a diffeomorphism will be denoted by and , respectively.

Spaces of densities.

The space of smooth probability densities on will be denoted by and will play an important role in the paper. It can be viewed in two ways. Either, as is common among mathematical physicists, an element of is a smooth real-valued function on , constrained to be strictly positive everywhere and of unit mass with respect to the reference volume form on . Or, as is common among differential geometers, the elements of can be viewed as normalized volume forms on . The latter is geometrically more natural since a probability density transforms as a volume form. However, either viewpoint has its pros and cons making various formulas look simpler or more familiar depending on the context. Therefore, we shall retain both conventions in this paper and distinguish between them as follows: a density thought of as a function will be denoted by , whereas the corresponding volume form will be denoted by . Notice that if is a diffeomorphism, then the equality corresponds to , where is the Jacobian with respect to .

2. Wasserstein–Otto geometry

2.1. Newton’s equations on

In this section we describe Newton’s equations on the full diffeomorphism group. Following Reference 5Reference 20, we first introduce a (weak) Riemannian structure.⁠Footnote2

2

A rigorous infinite-dimensional setting for diffeomorphisms and densities will be given in Appendix B. Here, for simplicity, we emphasize only the underlying geometric structure, leaving aside technical issues.

Definition 2.1.

The -metric on is given by

or, equivalently, after a change of variables

where , , and is a vector field on .

Newton’s equation on is a second order differential equation of the form

where is a potential energy function and is the covariant derivative of the -metric. We are interested in the case in which potential energy depends on implicitly via the associated density, i.e.,

where and is a given functional. We always assume that for each has a variational derivative given as a smooth function .

A more explicit form of Equation 2.3 is given by the following theorem.

Theorem 2.2 (Reference 72Reference 73).

Newton’s equations on for the metric Equation 2.1 and a potential function Equation 2.4 can be written as

In reduced variables and the equations assume the form

The right-hand side of the equations in Equation 2.5 is a result of a direct calculation, which we state in a separate lemma.

Lemma 2.3.

If is of the form Equation 2.4, then

where .

Proof.

This lemma is essentially the divergence theorem on the infinite-dimensional space. The proof in terms of variations of diffeomorphisms and densities mimics the finite-dimensional one.

Since stands here for the gradient of in the -metric 2.1 and is the flow of the vector field , we have

where we used that .

Proof of Theorem 2.2.

The equations in Equation 2.5 follow directly from Lemma 2.3 and the fact that the covariant derivative with respect to the -metric is just the pointwise covariant derivative on . The reduced equations in Equation 2.6 are derived in the Hamiltonian setting in section 3.2.

The following special class of solutions to Newton’s equations is of particular interest.

Proposition 2.4.

The gradient fields form an invariant set of solutions of the reduced equations Equation 2.6. Expressed in and , these solutions fulfill the Hamilton–Jacobi equations

Proof.

This follows from a direct computation using the identity A geometric explanation for the appearance of the Hamilton–Jacobi equation will be given in the next section.

An important point we want to emphasize in this survey is that a large number of interesting systems in mathematical physics originate as Newton’s equations on corresponding to different choices of potential functions. A partial list of examples discussed here is given in Table 2. We will also describe other systems on including the MHD equations or the relativistic as well as the fully compressible Euler equations.

We have already seen two different formulations of Newton’s equations: the second order (Lagrangian) representation in Equation 2.5 and the reduced first order (Eulerian) respresentation in Equation 2.6. In order to obtain all the equations listed in Table 2 we will need two further formulations: one defined on the space of densities and another defined on the space of wave functions. We begin with the former, postponing wave functions until Section 9.

2.2. Riemannian submersion over densities

The space of smooth probability densities on is an open subset of the affine subspace of all smooth function (or -forms) that integrate to one. It can be given the structure of an infinite-dimensional manifold whose tangent bundle is trivial

where

Definition 2.5.

The left coset projection between the space of diffeomorphisms and the space of probability densities is given by

or, equivalently, by pushforward of the reference volume form .

This projection relates the -metric Equation 2.1 and the following metric on the space of densities.

Definition 2.6.

The Wasserstein–Otto metric is a Riemannian metric on given by

where is defined by the transport equation

and is a tangent vector at .

The Riemannian distance defined by the metric Equation 2.8 on is precisely the Kantorovich–Wasserstein distance of optimal transport; see Reference 9, Reference 64, Reference 47, or Reference 77.

Theorem 2.7 (Reference 64).

The projection Equation 2.7 is an (infinite-dimensional) Riemannian submersion with respect to the -metric on and the Wasserstein–Otto metric on . Namely, given a horizontal ⁠Footnote3 vector one has

3

That is, such that for all .

where .

An illustration of this theorem is given in Figure 1. The proof is based on two lemmas. Recall that the left coset projection is the pushforward action of on . The corresponding isotropy group is the subgroup of volume-preserving diffeomorphisms

so that if is a left coset in , then if and only if there exists such that .

The first lemma states in particular that the action of on is transitive.

Lemma 2.8.

Let be the left coset projection Equation 2.7. Then

is a principal bundle. Consequently, the quotient space of left cosets is isomorphic to .

Proof.

Surjectivity of is a consequence of Moser’s lemma Reference 61. The fact that defines an infinite-dimensional principal bundle in the category of tame Fréchet manifolds (cf. Appendix B) follows from a standard argument using the Nash–Moser–Hamilton theorem; cf. Reference 30.

The second lemma states that the -metric on is compatible with the principal bundle structure above.

Lemma 2.9.

The -metric Equation 2.1 is right-invariant with respect to the action, namely

for any and .

Proof.

Since the result follows at once from Equation 2.2.

In Reference 5 Arnold used the -metric Equation 2.1 to show that its geodesic equation on , when expressed in Eulerian coordinates, yields the classical Euler equations of an ideal fluid. This marked the beginning of geometric and topological hydrodynamics; cf. Reference 8 or section 3.1.

The Riemannian submersion framework described above concerns objects that are extrinsic to Arnold’s (intrinsic) point of view. More precisely, rather than restricting to the vertical directions tangent to the fibre , we consider the horizontal directions in the total space and use the fact that any structure on which is invariant under the right action of induces a corresponding structure on by Lemma 2.8.

We are now ready to prove the main result of this subsection.

Proof of Theorem 2.7.

Given , let and with as before. Then

The kernel of is and defines a vertical distribution. On the other hand, the horizontal distribution is

Indeed, if , then from Equation 2.1 we have

and it follows that is an isometry. Its inverse is

where . From Equation 2.1 we now compute

where the last equality follows from the definition of . Thus, the projection is a Riemannian submersion.

Remark 2.10.

If is a smooth functional on with a variational derivative for every , then from the above expression we have

which gives the following formula for the gradient of in the Wasserstein–Otto metric

since every vector tangent to has zero mean. In particular, if is the relative entropy , then and since

we recover the formula , i.e., the Wasserstein gradient flow of entropy corresponds to the heat flow on the space of densities, cf. Reference 64.

3. Hamiltonian setting

The point of view of incompressible hydrodynamics as a Hamiltonian system on the cotangent bundle of described by Arnold Reference 5 turned out to be remarkably useful in applications involving invariants and stability (this is reviewed in section 3.1). In the next sections we develop the framework for Newton’s equations (adding a potential energy term to the kinetic energy which yields geodesics) on the group of all diffeomorphisms (rather than volume-preserving ones).

3.1. Hamiltonian framework for the incompressible Euler equations

In Reference 5 Arnold suggested using the following general framework on an arbitrary group describing a geodesic flow with respect to a suitable one-sided invariant Riemannian metric on this group. (Similar ideas can be traced back to S. Lie and H. Poincaré Reference 46Reference 66.)

Let a (possibly infinite-dimensional) Lie group be the configuration space of some physical system. The tangent space at the identity element is the corresponding Lie algebra . Fix a positive definite quadratic form (the “energy”) on and right translate it to the tangent space at any point (this is “translational symmetry” of the energy). In this way the energy defines a right-invariant Riemannian metric on the group. The geodesic flow on with respect to this energy metric represents extremals of the least action principle; i.e., actual motions of the physical system.

The operator defining the energy (and called the inertia operator) allows one to rewrite the Euler equation on the dual space . The Euler equation on turns out to be Hamiltonian with respect to the natural Lie–Poisson structure on the dual space. The corresponding Hamiltonian function is the energy quadratic form lifted from the Lie algebra to its dual space by the same identification: , where . Now the Euler equation on corresponding to the right-invariant metric on the group is given by

as an evolution of a point ; see, e.g., Reference 8. Here is the operator of the coadjoint representation of the Lie algebra on its dual : for any elements and .

Applied to the group of volume-preserving diffeomorphisms on , this framework provides an infinite-dimensional Riemannian setting for the Euler equations Equation 1.3 of an ideal fluid in . Namely, the right-invariant energy metric is given here by the -inner product on divergence-free vector fields on , that constitute the Lie algebra . The equations Equation 3.1 in this particular setting then correspond to the incompressible Euler equations Equation 1.3. The approach also provides the following Hamiltonian framework for classical hydrodynamics.

Theorem 3.1 (See, e.g., Reference 8).
a)

The dual space to the Lie algebra is , the space of cosets of -forms on modulo exact -forms. The coadjoint action of is given by change of coordinates in a -form, while the coadjoint action of is given by the Lie derivative along a vector field . It is well-defined on the cosets in .

b)

The inertia operator is defined by assigning to a given divergence-free vector field the coset in .

c)

The incompressible Euler equations Equation 1.3 on the dual space have the form

where and .

The proof follows from the fact that the map provides an isomorphism of the space of divergence-free vector fields and the space of closed -forms on , i.e., , since . The dual space is and the pairing is given by

For more details we refer to Reference 8.

Remark 3.2.

Equation Equation 3.2 can be rewritten in terms of a representative -form and a differential of a (pressure) function

which is a more familiar form of the Euler equations of an ideal fluid.

Note that each coset contains a unique coclosed -form which is related to a divergence-free vector field by means of the metric on , namely . Such a choice of a representative defines the (pressure) function uniquely modulo a constant since is prescribed for each time .

3.2. Hamiltonian formulation and Poisson reduction

Newton’s equations Equation 2.5 can be viewed as a canonical Hamiltonian system on . To write down this system, we identify each cotangent space with the dual of the space of vector fields . The (smooth part of the) latter space consists of differential -forms with values in the space of densities

where the tensor product is taken over the ring . The natural pairing between and is given by

(when , we will sometimes omit the subscript). This pairing does not depend on the Riemannian metric on .

Remark 3.3.

The spaces discussed and maps between them are summarized in the commutative diagram in Figure 2. The right column of the diagram describes a natural projection from the diffeomorphism group to the space of normalized smooth densities on with fibers that consist of all those diffeomorphisms which push a given reference density to any other density. As we discussed, this projection is a Riemannian submersion for a (noninvariant) -metric on and the (Kantorovich–)Wasserstein–Otto metric on used in the optimal mass transport; see Reference 64. The symplectic viewpoint on the Riemannian submersion leads naturally to the Hamiltonian description of the corresponding equations and the appearance of the momentum map, as we discuss below and in Appendix A.

The same diagram arises for a different Riemannian submersion, when is equipped with a right-invariant homogeneous Sobolev -metric and with the Fisher–Rao (information) metric, which plays an important role in geometric statistics, as we discuss in Section 7; cf. Reference 38.

Consider the standard Lagrangian on in the kinetic-minus-potential energy form

As usual, the passage to the Hamiltonian formulation on is obtained through the Legendre transform which in this case is given by

(In this section it is more convenient to work with the volume form instead of the density function .)

Lemma 3.4.

The Hamiltonian corresponding to the Lagrangian is

Proof.

In the above notation for , we have

The result follows since is quadratic and is independent of .

We can now turn to Newton’s equations on .

Theorem 3.5.

The Hamiltonian form of the equations Equation 2.5 is

where .

Proof.

In canonical coordinates on the Hamiltonian equations take the form

where are the canonical momenta satisfying .

Given a Hamiltonian on and a variation we have

and thus , where . Differentiating with respect to the variable, we obtain

As before, writing the Hamiltonian in Equation 3.4 as where and letting be a variation of generated by the field , we find

Thus and the equation for becomes

Finally, a straightforward computation using Equation 3.4 gives

which concludes the proof.

Rewriting the system Equation 3.5 in terms of and provides an example of Poisson reduction with respect to as the symmetry group. From Equation 3.3 we obtain a formula for the cotangent action of this group on , namely

Theorem 3.6 (Poisson reduction).

The quotient space is isomorphic to . The isomorphism is given by the projection . Furthermore, is a Poisson map with respect to the canonical Poisson structure on and the Poisson structure on given by

In Appendix B we provide an alternative construction in the setting of Fréchet spaces.

Proof.

From Equation 3.6 and Lemma 2.8 it follows that

with the projection given by . The fact that is a Poisson map (in fact, a Poisson submersion) follows from the consideration in Appendix A.

Remark 3.7.

The bracket Equation 3.7 is the classical Lie–Poisson structure on the dual of the semidirect product .

Corollary 3.8.

Let be a Hamiltonian function on satisfying

Then for some function . In reduced variables and , the Hamiltonian equations assume the form

where .

Proof.

Using the Poisson form of the Hamiltonian equations with , we obtain from Equation 3.7 the weak form of the equations

for any and . Rewriting the right-hand side as

completes the proof.

The following is the Hamiltonian analogue of Proposition 2.4.

Proposition 3.9.

The product manifold

is a Poisson submanifold of .

Proof.

From Equation 3.8 we find that the momenta form an invariant set in for any choice of Hamiltonian .

It turns out that the submanifold in Proposition 3.9 is symplectic, as we shall discuss below.

3.3. Newton’s equations on

Poisson reduction with respect to the cotangent action of on leads to reduced dynamics on the Poisson manifold (cf. Theorem 3.6). This Poisson manifold is a union of symplectic leaves one of which can be identified with equipped with the canonical symplectic structure. Indeed, the latter turns out to be the symplectic quotient corresponding to the zero-momentum leaf; see Appendix A. Here we identify as a symplectic submanifold of .

Lemma 3.10.

The (smooth part of the) cotangent bundle of is

Furthermore, can be regarded as a symplectic leaf in the Poisson manifold via the mapping

Proof.

Since the space of zero-mean functions is a subspace of , it follows that

That is a symplectic leaf in now follows from Proposition 3.9 since the mapping is bijective and Poisson. (Even more geometrically, one can regard as the zero-momentum symplectic reduction leaf as outlined in Appendix A.)

Next, we turn to Newton’s equations on for the Wasserstein–Otto metric Equation 2.8.

Corollary 3.11.

The Hamiltonian on corresponding to Newton’s equations on with respect to the Wasserstein–Otto metric Equation 2.8 is

and the Hamiltonian equations are

Solutions of Equation 3.9 correspond to horizontal solutions of Newton’s equations Equation 2.5 on or, equivalently, to zero-momentum solutions of the reduced equations Equation 3.8 with Hamiltonian

Proof.

Given the Hamiltonian Equation 3.4 on , the result follows directly from Theorem 3.6 and the zero-momentum reduction result Theorem A.4 in Appendix A.

4. Wasserstein–Otto examples

In this section we provide and study examples of Newton’s equations on with respect to the metric Equation 2.1 and -invariant potentials. We also derive the corresponding Poisson reduced equations on (cf. section 3.2) and symplectic reduced equations on corresponding to Newton’s equations for the Wasserstein–Otto metric Equation 2.8 (cf. section 3.3).

4.1. Inviscid Burgers equation

We start with the simplest case when the potential function is zero. The corresponding Newton’s equations are the geodesic equations on .

Proposition 4.1.

Newton’s equations with respect to the -metric Equation 2.1 and with zero potential admit the following formulations:

the geodesic equations on

the inviscid Burgers equations on

where ;

the Poisson reduced equations on

where ;

the symplectically reduced equations on

corresponding to the Hamiltonian form of the geodesic equations for the Wasserstein–Otto metric Equation 2.8.

Observe that the system in Equation 4.1 consists of the Hamilton–Jacobi equation for the kinetic energy Hamiltonian on together with the transport equation for .

Proof.

The results follow directly from Theorem 2.2, Theorem 3.6, and Corollary 3.11 after setting .

4.2. Classical mechanics and Hamilton–Jacobi equations

Let  be a smooth potential function on and consider the corresponding potential function on the space of densities

where .

Proposition 4.2.

Newton’s equations with respect to the -metric Equation 2.1 and the potential in Equation 4.2 admit the following formulations:

the geodesic equations with potential on

the inviscid Burgers equations with potential on

where ;

the Poisson reduced equations on

where ;

the symplectically reduced equations on

Observe that the system Equation 4.3 consists of the Hamilton–Jacobi equation for the classical Hamiltonian together with the transport equation for .

Proof.

Since the proposition follows by combining Theorem 2.2, Theorem 3.6, and Corollary 3.11.

4.3. Barotropic fluid equations

The motion of barotropic fluids is characterized by a functional relation between the pressure and the fluid’s density. The corresponding equations on a Riemannian manifold expressed in terms of the velocity field and the density function have the form

The function relates and the pressure function . This relation depends on the properties of the fluid and is called the barotropic equation of state. Note that the equations of barotropic gas dynamics are usually specified by a particular choice (where, e.g., corresponds to the standard approximation for atmospheric air).

To connect these objects with our framework, we let be a function describing the internal energy of a barotropic fluid per unit mass and consider a general potential

where . The relation between pressure and the internal energy is given by

We also define the thermodynamical work function as

We have which helps explain the idea of introducing the work function in that the force in Equation 4.4 becomes a pure gradient (here is understood as the gradient of a function on ). This can be arranged if the internal energy depends functionally on . As we have seen in the general form of Equation 2.6 the internal work function is more fundamental than the pressure function in the following sense: when the internal energy depends on the derivatives of , it may not be possible to find the pressure as a differential operator on unlike the the work function.

Proposition 4.3.

Newton’s equations for the -metric Equation 2.1 and the potential Equation 4.5 admit the following formulations:

on

the barotropic compressible fluid equations Equation 4.4 on for the velocity field and the density function ;

the Poisson reduced equations on

where and is the work function Equation 4.6;

the symplectically reduced form of the barotropic compressible fluid equations on

Proof.

The energy function of a compressible barotropic fluid with velocity and density is

where the first term corresponds to fluid’s kinetic energy and the second is the potential energy under the barotropic assumption. Introducing the momentum variable , we obtain a Hamiltonian on of the form

It is clear that . Furthermore, we have

Substituting into Equation 3.8 we arrive at the system

To obtain the compressible Euler equations, we rewrite the second term in the first equation using Equation 4.6 as

to get

Differentiating in the time variable

and substituting into Equation 4.8, we obtain

Using the identities and , we now recover the compressible Euler equations Equation 4.4, that is

To describe these equations as Newton’s equations Equation 3.5 on , we consider the potential given in Equation 4.5 with . Note that this potential is of the form Equation 2.4. From Theorem 2.2 we find Newton’s equations corresponding to the compressible Euler equations

From Corollary 3.11 we get the symplectically reduced form on .

5. Semidirect product reduction

In this section we recall one standard approach to the equations of compressible fluid dynamics using semidirect products; see Reference 52Reference 78. Recall from the earlier sections that the barotropic Euler equations can be viewed as a mechanical system on the configuration space with the symmetry group . On the other hand, such a system can be also obtained by a semidirect product construction as a so-called Lie–Poisson system provided that the configuration space is extended so that it coincides with the given symmetry group. We unify these approaches in section A.2: any Lie–Poisson system on a semidirect product can be viewed as a Newton system with a smaller symmetry group. We begin with two standard examples.

5.1. Barotropic fluids via semidirect products

In order to describe a barotropic fluid Equation 4.4 as a Lie–Poisson system, one can introduce the semidirect product group as a space of pairs equipped with the group structure

which is smooth in the Fréchet topology; cf. Appendix B.

The Lie algebra is also a semidirect product with a commutator given by

The corresponding (smooth) dual space is whose elements are pairs with and , where is a fixed volume form and is a -form on . The pairing between and is given by

The Lie algebra structure of determines the Lie–Poisson structure on , and the corresponding Poisson bracket at is given by the formula Equation 3.7. It is sometimes called the compressible fluid bracket. (We refer to section A.2 for a general setting of semidirect products and explicit formulas.) Notice that is strictly bigger than , since now can be any function (it does not have to be a probability density).

In order to define a dynamical system on consider a smooth function (relating pressure to fluid’s density , as in section 4.3) of the form and define the following energy function on

Lifting to the dual with the help of the inertia operator of the Riemannian metric, we obtain the Hamiltonian on ,

Observe that, by construction, the associated Hamiltonian system on the cotangent bundle is right-invariant with respect to the action of .

Theorem 5.1 (Reference 52Reference 78).

The barotropic fluid equations Equation 4.4 correspond to the Lie–Poisson system on with the Poisson bracket of type Equation 3.7 and the Hamiltonian Equation 5.3.

While the general barotropic equations described above are valid for any smooth initial velocity field, one is often interested only in potential solutions of the system. These are obtained from initial conditions of the form , where is a smooth function on . As we have already seen, such solutions retain their gradient form for all times and the equations can be viewed as the Hamilton–Jacobi equations; see Equation 1.7. Potential solutions of this type arise naturally in the context of the Madelung transform; see Section 9.

Remark 5.2.

The semidirect product framework is a natural setting whenever the physical model contains a quantity transported by the flow; e.g., the continuity equation Equation 4.4. However, while the Hamiltonian point of view works similarly to the case of incompressible fluids, the Lagrangian approach with semidirect product Lie algebras encounters drawbacks; cf. Reference 33. These are mostly related to the fact that the Lagrangian is not quadratic and cannot be directly interpreted as a kinetic energy yielding geodesics on the group (for some attempts to bypass this problem using the Maupertuis principle, see Reference 72; for a geodesic formulation in an extended phase space, see Reference 67). Furthermore, there is no physical interpretation of the action of the full semidirect product on its dual space: the particle reparametrization symmetry is related only to the action of the first (diffeomorphisms) but not of the second (functions) factor in the product . One advantage of our point of view in section 4.3 using Newton’s equations is that it resolves such issues.

5.2. Incompressible magnetohydrodynamics

An approach based on semidirect products is also possible in the case of the equations of self-consistent magnetohydrodynamics (MHD). We start with the incompressible case and discuss the compressible case in detail in section 6.2. The underlying system describes an ideal fluid whose divergence-free velocity is governed by the Euler equations (see section 3.1 for Lagrangian and Hamiltonian formulations). Assume next that the fluid has infinite conductivity and carries a (divergence-free) magnetic field . Transported by the flow (i.e., frozen in the fluid), acts reciprocally (via the Lorenz force) on the velocity field and the resulting MHD system on a three-dimensional Riemannian manifold takes the form

A natural configuration space for the system Equation 5.4 is the semidirect product of the group of volume-preserving diffeomorphisms and the dual of the space of divergence-free vector fields on a threefold . The corresponding Lie algebra is the semidirect product of and its dual. The group product and the algebra commutator are given by the formulas Equation 5.1 and Equation 5.2, respectively.

More generally, the configuration space of incompressible magnetohydrodynamics on a manifold of arbitrary dimension is the semidirect product group (which for reduces to ) with its Lie algebra . Since the dual of is the space of closed -forms on , we have . Magnetic fields in can be viewed as either closed -forms or fields that are related to by . This latter point of view will be useful also for the description of compressible magnetohydrodynamics.

The corresponding Poisson bracket on is given by the formula Equation 3.7 interpreted accordingly.

Finally, as the Hamiltonian function we take the sum of the kinetic and magnetic energies of the fluid, i.e.,

(here the Riemannian metric defines the inertia operator and hence the quadratic form on all spaces , and ; see, e.g., Reference 8). The Hamiltonian on is

Theorem 5.3 (Reference 8Reference 78).

The incompressible MHD equations Equation 5.4 correspond to the Lie–Poisson system on for the Hamiltonian Equation 5.5.

An analogue of this equation for compressible fluids in an -dimensional manifold will be discussed in section 6.2.

6. More general Lagrangians

6.1. Fully compressible fluids

For general compressible (nonbarotropic) inviscid fluids, the equation of state includes pressure as a function of both density and specific entropy (defined as a smooth function on representing entropy per unit mass; cf. Dolzhansky Reference 18, Sect. 3.2). Thus, the equations of motion describe the evolution of three quantities: the velocity of the fluid , its density , and the specific entropy , namely

The purpose of this section is to show that under natural assumptions this system also describes Newton’s equations on but with potential function of more general form than Equation 2.4. In a nutshell, a proper phase space for this equation is the reduction of over a subgroup which is smaller than . In view of the results in section A.2 the full compressible Euler equations are a semidirect product representation of a Newton system on whose symmetry group is a proper subgroup of .

Theorem 6.1.

The fully compressible system Equation 6.1 is obtained using an embedding into the Lie–Poisson space , where (cf. Proposition A.5) from Newton’s equations on with Lagrangian

where is a potential function (of density and entropy density for some fixed initial entropy density of the form

and where the internal energy and pressure are related by

From the point of view of symplectic reduction in section A.2 the symmetry subgroup is given by . Our aim is to embed in . (Notice that while the quotient might not be manifold, it can be viewed as an invariant set consisting of coadjoint orbits in the dual space of an appropriate semidirect product Lie algebra, as discussed in section A.2.) To achieve this embedding, we need to compute the momentum map for the cotangent lifted action of on .

Lemma 6.2.

The momentum map for the cotangent action of on

is

Proof.

From Section 3 we already know the momentum map for the action on . This is the same as the action on . For diagonal actions we then just get a sum as stated in the lemma.

Proof of Theorem 6.1.

Any Hamiltonian system on the Poisson space has the form

where . The Hamiltonian corresponding to the Lagrangian Equation 6.2 is the same as in Equation 4.7 except that the potential energy depends now also on . By Lemma 6.2 the first equation then becomes

The variational derivatives are given by

Using

we then recover from Equation 6.3 the fully compressible Euler equations Equation 6.1.

Observe that an invariant subset of solutions is given by those solutions with momenta , where . They can be regarded as analogues of potential solutions of the barotropic fluid equations. We thereby obtain a canonical set of equations on ,

with the restricted Hamiltonian

We point out that the group corresponding to is associated with a multicomponent version of the Madelung transform relating compressible fluids and the NLS-type equations; cf. the details in Section 9 and see also Reference 40. Applying the multicomponent Madelung transform , one can also rewrite the fully compressible system on the space of rank- spinors .

Remark 6.3.

Solutions of barotropic fluid equations are contained in the solution space of the fully compressible Euler equations as “horizontal-within-horizontal” solutions in the following sense. Let the initial entropy function have the form for some function . Then

where the last equality follows from the evolution equation for . From the equation for we obtain

Thus, the entropy remains in the form so that we obtain a barotropic flow with the pressure function . From a geometric point of view these solutions correspond to a special symplectic leaf in .

6.2. Compressible magnetohydrodynamics

Next, we turn to a description of compressible inviscid magnetohydrodynamics. A compressible fluid of infinite conductivity carries a magnetic field acting reciprocally on the fluid. The corresponding equations on a Riemannian 3-manifold have the form

where is the velocity and is density of the fluid, while is the magnetic vector field. Note that these equations reduce to the incompressible MHD equations Equation 5.4 when density is a constant.

As mentioned before, it is more natural to think of magnetic fields as closed -forms. This becomes apparent when the equations are generalized to a compressible setting or to other dimensions. (For instance, a non-volume-preserving diffeomorphism violates the divergence-free constraint of a magnetic vector field but preserves closedness of differential forms.) In fact, let denote the space of smooth closed differential -forms on an -manifold . The diffeomorphism group acts on by pushforward and the (smooth) dual of is the quotient .

The cotangent lift of the left action of to is given by

Observe that this is well-defined since pushforward commutes with the exterior differential.

Lemma 6.4.

The momentum map associated with the cotangent action in Equation 6.5 is given by

where the vector field is uniquely defined by .

As expected, the map is independent of the choice of and a representative . In what follows it will be convenient to replace by —resulting in a different vector field but without affecting the momentum map.

Proof.

The infinitesimal action of a vector field on is . Since it is a cotangent lifted action, the momentum map is given by

Now, if , then

Consider a Lagrangian on given by the fluid’s kinetic and potential energies with an additional term involving the action on the magnetic field , namely

where , and . As in Lemma 3.4 the corresponding Hamiltonian is

where . Letting denote the isotropy subgroup for the action of , the (right) symmetry group of the Hamiltonian Equation 6.6 is

The corresponding Lie algebra consists of vector fields such that

If is even-dimensional and is nondegenerate, then the pair is a symplectic manifold and the Lie algebra consists of symplectic vector fields that also preserve the first integral .

Next, we proceed to carry out Poisson reduction; i.e., to compute the reduced equations on . In contrast to the case studied in Section 3, there is no simple way to identify , and so it will be convenient to use the semidirect product reduction framework developed in section A.2. (Similarly to section 6.1, the quotient might not be manifold, but it can be regarded as an invariant set formed by coadjoint orbits in the dual space of an appropriate semidirect product Lie algebra; see section A.2. For now one can regard these considerations as taking place at a “smooth point” of the quotient.) To this end, consider the semidirect product algebra and its dual

We have a natural embedding of in via the map and the corresponding Hamiltonian on is

Theorem 6.5.

The Poisson reduced form on

of the Euler–Lagrange equations for the Hamiltonian Equation 6.6 is

where the field is defined by and the momentum variable is . For a threefold these equations correspond to the equations of the compressible inviscid magnetohydrodynamics Equation 6.4 where the magnetic field is related to the closed -form by .

Proof.

In general, if acts on a space from the left with the momentum map , then the Poisson reduced system is

In our case, and the momentum map is

The rest of the proof follows from direct calculations.

Corollary 6.6.

The equations Equation 6.7 admit special horizontal solutions corresponding to momenta of the form

These solutions can be expressed in the variables

as a canonical Hamiltonian system for the Hamiltonian

where and .

Proof.

The horizontal solutions correspond to the submanifold , where is the momentum map associated with the subgroup . We refer to Appendix A for details on symplectic reduction. The Hamiltonian Equation 6.8 is just the restriction of to the special momenta.

6.3. Relativistic inviscid Burgers equation

In this section we present a relativistic version of the Otto calculus, motivated by the treatment in Reference 13. We show that it leads to a relativistic Lagrangian on , and we employ Poisson reduction of section 3.2 to obtain the relativistic hydrodynamics equations.

As in the classical case, we consider a path in the space of diffeomorphisms as a family of free relativistic particles. Given the action is then given by

It is natural to think of this action as the restriction to a fixed reference frame of the corresponding action functional on the Lorentzian manifold equipped with the Lorentzian metric

More explicitly, this extended action is given by

where is the volume form associated with .⁠Footnote4 In contrast with the classical case, the action Equation 6.10 is left-invariant under the subgroup of Lorentz transformations in the following sense: if , then

4

While in classical mechanics the action stands for the length square, note that in the classical limit, i.e., for small velocities, , so that formula Equation 6.10 leads to the classical action.

Returning to Equation 6.9, the associated Lagrangian on is

Since the Lagrangian is right-invariant with respect to , we can carry out Poisson reduction of the corresponding Hamiltonian system on as described above.

Brenier in Reference 13 used such an approach to derive a relativistic heat equation. We are now in a position to use it for relativistic hydrodynamics.

Theorem 6.7.

The relativistic Lagrangian Equation 6.11 on induces a Poisson reduced system on . The Hamiltonian is given by

and the governing equations are

where .

Proof.

The reduced Lagrangian for Equation 6.11 is given by

The momentum variable is given by the Legendre transformation

with the inverse

The corresponding Hamiltonian is

so that

and the result follows from Corollary 3.8.

Remark 6.8.

As we formally recover the classical inviscid Burgers equation in section 4.1. Indeed, assuming is small in comparison with , a Taylor expansion of the right-hand side of Equation 6.12 gives

As we also have as we recover the classical inviscid Burgers equation.

Remark 6.9.

In order to obtain the equations of relativistic hydrodynamics, one needs to incorporate internal energy via the reduced Hamiltonian on given by

where is the internal energy function; cf. Reference 42 and Reference 32. This gives a relativistic version of the classical barotropic equations in section 4.3.

7. Fisher–Rao geometry

7.1. Newton’s equations on

We now focus on another important Riemannian structure on . This structure is induced by the Sobolev -inner product on vector fields and has the same relation to the Fisher–Rao metric on as the -metric on to the Wasserstein–Otto metric on .

Definition 7.1.

Let be a compact Riemannian manifold with volume form . For any and we set

where is the Laplacian on vector fields and is a quadratic form depending only on the vertical (divergence-free) component of .

Remark 7.2.

From the point of view of the geometry of (and for most of our applications) only the first term on the right-hand side of Equation 7.1 is relevant. However, it is convenient to work with the above metric on , in particular, because of its relation to a number of familiar equations; cf. Reference 38Reference 58 and below. Note also the following analogy between the Wasserstein and the Fisher–Rao structures: while the noninvariant -metric induces a factorization of where one of the factors solves the optimal mass transport problem, the invariant metric Equation 7.1 induces a different factorization of which solves an optimal information transport problem; cf. Reference 58.

Consider a potential function of the form

where is a potential functional on . (In this section it is convenient to work with volume forms instead of .) It is interesting to compare the present setting with that of section 2.1, where the potential function on was defined using pushforwards rather than pullbacks. As a result one works with the left cosets rather than with the right cosets; cf. Remark 7.9

Theorem 7.3.

Newton’s equations of the metric Equation 7.1 on with a potential function Equation 7.2 have the form

where the inertia operator is given by

Proof.

The derivation of the equation in the case of zero potential can be found in Reference 58. Modifications needed here follow from the calculation

where is the flow of the vector field in .

We proceed with a Hamiltonian formulation. As in section 3.2 we will identify cotangent spaces with .

Proposition 7.4.

The Hamiltonian form of Newton’s equations Equation 7.3 on  is

where .

Proof.

This follows simply by pulling back by the equations in Equation 7.3 and applying the identity

Remark 7.5.

Observe that if the potential function is zero, then the equation in Equation 7.5 expresses conservation of the momentum associated with the right invariance of the metric.

7.2. Riemannian submersion over densities

We turn to the geometry of the fibration of with respect to the metric Equation 7.1.

Definition 7.6.

The right coset projection between diffeomorphisms and smooth probability densities is given by

As before, it turns out that the projection Equation 7.6 is a Riemannian submersion if the base space is equipped with a suitable metric.

Definition 7.7.

The Fisher–Rao metric is the Riemannian metric on given by

where represents a tangent vector at .

Theorem 7.8.

The right coset projection Equation 7.6 is a Riemannian submersion with respect to the metric Equation 7.1 on and the Fisher–Rao metric on . In particular, if is horizontal, i.e.,

then where .

Proof.

See Reference 58, Thm. 4.9.

Note also that it follows from the Hodge decomposition that the horizontal distribution on consists of elements of the form ; cf. Reference 58 for details.

Remark 7.9.

Let us summarize the definition of the two metrics on that we discussed so far. The Wasserstein–Otto metric (cf. section 2.2) is defined as follows:

Note that it depends on the Riemannian structure on . The Fisher–Rao metric, on the other hand, is given by the “universal formula”,

and is independent of the Riemannian structure on .

Note also that the setting of Theorem 7.8 is quite different from that of Theorem 2.7. In the latter, the Riemannian metric on is right-invariant with respect to and automatically descends to the quotient from the right, namely . In the former, the metric is right-invariant with respect to and descends to the quotient from the left, namely . Thus, in Theorem 7.8 the right-invariance property is retained after taking the quotient, and therefore the Fisher–Rao metric on remains right-invariant with respect to the action of (corresponding to right translation of the fibers), which is easy to verify.

Proposition 7.10.

The gradient of a smooth function with respect to the Fisher–Rao metric is

where is a Lagrange multiplier such that .

Proof.

Let , and let be a representative of the variational derivative in . We have

which yields the explicit form of the gradient.

We end this subsection by recalling a particularly remarkable property of the Fisher–Rao metric. Let be the unit sphere in the pre-Hilbert space .

Theorem 7.11.

The square root map

is, up to a factor , a Riemannian isometry between equipped with the Fisher–Rao metric in Equation 7.7 and the (geodesically convex) subset

of the sphere .

This result was first obtained by Friedrich Reference 25 and later independently in Reference 38 in the Euler–Arnold framework of diffeomorphism groups.

7.3. Newton’s equations on

Recall that in section 3.3 the Hamiltonian equations on were obtained by symplectic reduction of a -invariant system on . In the setting with the right coset projection Equation 7.6 and the metric Equation 7.1, the situation is quite different, since the Riemannian metric is not left-invariant with respect to (otherwise, interchanging pushforwards and pullbacks would give a completely dual theory). Nevertheless, there is a zero momentum reduction on the Hamiltonian side corresponding to the Riemannian submersion structure described in section 7.2.

Proposition 7.12.

The exact momenta, i.e., tensor products of the form

form an invariant set for the system Equation 7.5.

Proof.

Substituting Equation 7.9 in Equation 7.5, we get

where and

From Equation 7.4 we find that solutions of the form define (up to a constant) , so that

Using , we then obtain

which proves the assertion.

Theorem 7.13.

Newton’s equations with respect to the Fisher–Rao metric 7.7 on and a potential have the form

where is a multiplier subject to . Furthermore, the Lagrangian and Hamiltonian are and respectively. The corresponding Hamiltonian equations have the form

Solutions of Equation 7.12 correspond to potential solutions (cf. Proposition 7.12) of Newton’s equations Equation 7.5 on .

Proof.

The result follows directly from the proof of Proposition 7.12 by setting and .

8. Fisher–Rao examples

8.1. The CH equation and Fisher–Rao geodesics

The periodic CH equation (also known in the literature as the HS equation) is a nonlinear evolution equation of the form

where . It was derived in Reference 37 as an Euler–Arnold equation on the group of diffeomorphisms of the circle equipped with the right-invariant Sobolev metric given at the identity by the inner product

The CH equation is known to be bi-Hamiltonian and to admit smooth, as well as cusped, soliton-type solutions. It may be viewed as describing a director field in the presence of an external (e.g., magnetic) force. The associated Cauchy problem has been studied extensively in the literature; cf. Reference 29Reference 37Reference 68. Many of its geometric properties can also be found in Reference 75. The following result was proved in Reference 58.

Proposition 8.1.

The CH equation Equation 8.1 is a (right-reduced) Newton’s equation Equation 7.3 with vanishing potential on . Geodesics of the Fisher–Rao metric Equation 7.7 on correspond to horizontal solutions of the CH equation described by the equations

(As in Theorem 7.13, the relation between , , and is given by , where is the Lagrangian flow of , and with the constant choosen so that

Observe that the Euler–Arnold equation of the metric Equation 7.1 can be naturally viewed as a higher-dimensional generalization of the equation Equation 8.1; see Reference 58. Furthermore, in the one-dimensional case horizontal solutions of this equation can be written in terms of the derivative . In higher dimensions we similarly have

Proposition 8.2.

The geodesic equations of the Fisher–Rao metric Equation 7.7 on reduce to the following equations on

where and .

Proof.

The equations follow directly from Equation 7.10 with . Note that this equation preserves the zero-mean condition for , while the constant appearing in the right-hand side of this equation, as well as the one in Proposition 8.1, is conditioned by the equation’s solvability.

8.2. The infinite-dimensional Neumann problem

The C. Neumann problem (1856) describes the Newtonian motion of a point on the -dimensional sphere under the influence of a quadratic potential; see section 1.1. It is known to be equivalent (up to a change of the time parameter) to the geodesic equations on an ellipsoid in with the induced metric; see, e.g., Reference 62Reference 63.

Here we describe a natural infinite-dimensional generalization of the C. Neumann problem. Consider the infinite-dimensional unit sphere

in the pre-Hilbert space and the quadratic potential function

We seek a curve that minimizes the action functional for the Lagrangian

Proposition 8.3.

Newton’s equations associated with the infinite-dimensional Neumann problem with potential Equation 8.2 have the form

where is a Lagrange multiplier subject to the constraint . In fact, we have .

Proof.

This is a simple consequence of the integration by parts formula.

Our next objective is to show that the infinite-dimensional Neumann problem on corresponds to Newton’s equations on with respect to the Fisher–Rao metric and a natural choice of the potential function. The latter is given by the Fisher information functional

where the density is .

Lemma 8.4.

The gradient of with respect to the Fisher–Rao metric can be computed from either of the two expressions

Proof.

Using the identities and , we can rewrite the Fisher information functional as

Differentiating the first of these expressions in the direction of the vector yields

Similarly, differentiating the second yields

The result now follows from Proposition 7.10.

Proposition 8.5.

Newton’s equations Equation 7.11 on with respect the Fisher–Rao metric and the Fisher–Rao potential Equation 8.4 are

where is a Lagrange multiplier for the constraint . The map establishes an isomorphism with the infinite-dimensional Neumann problem Equation 8.3.

Proof.

The form of the equation on follows from Theorem7.13. It is straighforward to check that . The result then follows from isometric properties of the square root map Equation 7.8.

Remark 8.6.

Of particular interest are the stationary solutions to the Neumann problem Equation 8.3, i.e., those with , in which case is a normalized eigenvector of the Laplacian with eigenvalue . If , then . Consequently, the stationary solutions correspond to the principal axes of the corresponding infinite-dimensional ellipsoid .

It is also possible to obtain quasi-stationary solutions this way. Indeed, assume that the eigenspace of is at least two-dimensional (for example, when ). If are two orthogonal eigenvectors with eigenvalue , then it is straightforward to check that a solution originating from with initial velocity for is given by

8.3. The Klein–Gordon equation

The Klein–Gordon equation

describes spin-less scalar particles of mass . It is invariant under Lorentz transformations and can be viewed as a relativistic quantum equation. To see how it relates to the Neumann problem of the previous subsection, let denote the space-time manifold equipped with the Minkowski metric of signature and consider a quadratic functional

which is the -norm of the the Minkowski gradient .

Proposition 8.7.

For the space-time manifold solutions of the infinite-dimensional Neumann problem with potential on the hypersurface

satisfy the Klein–Gordon equation Equation 8.5 with mass parameter .

Proof.

This is a calculation analogous to that in Remark 8.6.

9. Geometric properties of the Madelung transform

In this section we recall several results concerning the Madelung transform which provides a link between geometric hydrodynamics and quantum mechanics; see Reference 39Reference 40. It was introduced in the 1920s by E. Madelung Reference 49 in an attempt to give a hydrodynamical formulation of the Schrödinger equation. Using the setting developed in previous sections, one can now present a number of surprising geometric properties of this transform.

Definition 9.1.

Let and be real-valued functions on with . The Madelung transform is defined by

where is a parameter (Planck’s constant).⁠Footnote5

5

In the publications Reference 39Reference 40 the convention is used.

Observe that is a complex extension of the square root map described in Theorem 7.8. Heuristically, the functions and can be interpreted as the absolute value and argument of the complex-valued function as in polar coordinates. Throughout this section we assume that is a compact simply connected manifold.

9.1. Madelung transform as a symplectomorphism

Let denote the complex projective space of smooth complex-valued functions on . Its elements can be represented as cosets of the -sphere of smooth functions, where if and only if for some . A tangent vector at a coset is a linear coset of the form . Following the geometrization of quantum mechanics by Kibble Reference 41, a natural symplectic structure on the projective space is

The projective space of nonvanishing complex functions is a submanifold of . It turns out that the Madelung transform induces a symplectomorphism between and the cotangent bundle of probability densities ; see Figure 3.

Namely, we have the following.

Theorem 9.2 (Reference 40).

The Madelung transform Equation 9.1 induces a map

which is a symplectomorphism (in the Fréchet topology of smooth functions) with respect to the canonical symplectic structure of and the symplectic structure Equation 9.2 of .

The Madelung transform was shown to be a symplectic submersion from to the unit sphere of nonvanishing wave functions by von Renesse Reference 79. The stronger symplectomorphism property stated in Theorem 9.2 is deduced using the projectivization .

9.2. Examples: linear and nonlinear Schrödinger equations

Let be a wavefunction on and consider the family of Schrödinger equations (or Gross–Pitaevsky equations) with Planck’s constant and mass of the form

where and . Setting , we obtain the linear Schrödinger equation with potential , while setting yields a family of nonlinear Schrödinger equations (NLS); typical choices are or .

From the point of view of geometric quantum mechanics (cf. Reference 41), equation Equation 9.4 is Hamiltonian with respect to the symplectic structure Equation 9.2, which is compatible with the complex structure of . The Hamiltonian associated with Equation 9.4 is

where is a primitive of .

Observe that the -norm of a wave function satisfying the Schrödinger equation Equation 9.4 is conserved in time. Furthermore, the equation is equivariant with respect to phase change and hence it descends to the projective space .

Proposition 9.3 (cf. Reference 48Reference 79).

The Madelung transform maps the family of Schrödinger Hamiltonians Equation 9.5 to a family of Hamiltonians on given by

at the density . In particular, if , we recover Newton’s equations Equation 3.9 on for the potential function

where is Fisher’s information functional Equation 8.4. The extension Equation 2.6 to a fluid equation on is

Remark 9.4.

For the linear Schrödinger equation Equation 9.4, where , notice that the “classical limit” immediately follows from Equation 9.6: as , we recover classical mechanics and the Hamilton–Jacobi equation as presented in section 4.2.

Remark 9.5.

In this section we have seen how the Schrödinger equation can be expressed as a compressible fluid equation via Madelung’s transform. Conversely, the classical equations of hydrodynamics can be formulated as nonlinear Schrödinger equations (since the Madelung transform is a symplectomorphism, so any Hamiltonian on induces a corresponding Hamiltonian on ). In particular, potential solutions of the compressible Euler equations of a barotropic fluid Equation 4.4 can be expressed as solutions to an NLS equation with Hamiltonian

where is the specific internal energy of the fluid. The choice gives a Schrödinger-type formulation for potential solutions of Burgers equation describing geodesics of the Wasserstein–Otto metric Equation 2.8 on . We thus have a geometric framework that connects optimal transport for cost functions with potentials, the Euler equations of compressible hydrodynamics, and the NLS-type equations described above.

Remark 9.6.

Another relevant development is the Schrödinger Bridge problem, which seeks the most likely probability law for a diffusion process in the probability space, that matches marginals at two endpoints in time, as we discuss in the next section: one can interpret it as a stochastic perturbation of Wasserstein–Otto geodesics on the density space for given endpoints. The Madelung transform allows one to translate questions about the Schrödinger equation to questions about probability laws; cf. Reference 83. More recently, the Madelung transform for quantum-classical hybrid systems has been studied in Reference 26.

9.3. The Madelung and Hopf–Cole transforms

There is a real version of the complex Madelung transform.

Definition 9.7.

Let and be real-valued functions on with , and let be a positive constant. The (symmetrized) Hopf–Cole transform is the mapping defined by

In Reference 43 it is shown that this map, along with its generalizations, has the property that its inverse takes the constant symplectic structure on to (a multiple of) the standard symplectic structure on . Note that the choice corresponds to the standard Madelung transform Equation 9.1: the function becomes a complex-valued wave function so that the symplectic properties of can be viewed as an extension of those of the Madelung map .

Consider the (viscous) Burgers equation

The second component of Equation 9.7 with maps the potential solutions , which satisfy the Hamilton–Jacobi equation

to the solutions of the heat equation .

Similarly, the Hopf–Cole map can be used to transform certain barotropic-type systems to heat equations. This can be verified directly in the example of section 9.2: setting Planck’s constant to be in the Schrödinger equation Equation 9.4 with , , and gives the forward and the backward heat equations

The corresponding barotropic fluid system, which is readily obtained from Equation 9.6 with , reads

This is again a Newton system on but in this case the potential function is corrected by the Fisher functional with the minus sign (instead of the plus sign as in Proposition 9.3). Equipped with the two-point boundary conditions and , the horizontal solutions of Equation 9.8 correspond to the solutions of a dynamical formulation of the Schrödinger bridge problem, as surveyed in Reference 45. In this way one can study nonconservative systems with viscosity in a symplectic setting. It is interesting to incorporate the incompressible Navier–Stokes equations into this framework. This would require a two-component version of the map in Reference 43 related to the two-component Madelung transform in the Schrödinger’s smoke example in section 9.4.

Remark 9.8.

Equation Equation 9.8 displays yet another relation to the heat flow connected to an invariant submanifold of . Consider the submanifold

A straightforward calculation shows that is an invariant submanifold for Equation 9.8 and the evolution on is given by a system of decoupled equations

Furthermore, since is the variational derivative of the entropy functional , it follows that the substitution (or ) corresponds to the momentum in the direction of negative entropy. This is related to the observation in Reference 64 that the heat flow is the -Wasserstein gradient flow of the entropy functional.

9.4. Example: Schrödinger’s smoke

While the Madelung transform provides a link between quantum mechanics and compressible hydrodynamics, in this section we describe how incompressible hydrodynamics is related to the so-called incompressible Schrödinger equation. The approach described here was developed in computer graphics by Chern et al. in Reference 15 to obtain a fast algorithm that could be used to visualize realistic smoke motion.

It is clear that the standard Madelung transform is not adequate to describe incompressible hydrodynamics since the group of volume-preserving diffeomorphisms lies in the kernel of the Madelung projection (any trajectory along projects to the constant wave function ). Instead, one has to consider the multi-component Madelung transform; cf. Reference 40. For simplicity, we use two components although one can easily extend the constructions below to the case of several components.

Consider the diagonal action of on and the associated momentum map given by

Fix two densities and consider the group intersection . This intersection is itself a group, which can be thought of as the subsgroup of, e.g., consisting of diffeomorphisms that also preserve the ratio function on . As in section A.2 we (formally) consider the quotient

This quotient is Poisson (assuming that it is a manifold or considering it “at a regular point”), and it can be regarded as a Poisson submanifold of the dual of the semidirect product algebra . (The same type of quotients appeared in the constructions related to compressible fluids section 6.1 and compressible MHD section 6.2.) Given a Hamiltonian on the governing equations are

where . The zero-momentum symplectic reduction, corresponding to momenta of the form , yields a canonical system

on for the Hamiltonian

Next, we turn to the incompressible case. Imposing the holonomic constraint for the equations on leads to a constrained Hamiltonian system

where is a Lagrange multiplier.

The induced cotangent constraint on is obtained by

which implies that the vector field is divergence-free. Therefore, solutions of Equation 9.9 correspond to zero-momentum solutions of the incompressible fluid equations on with the Hamiltonian

In particular, the choice

yields special solutions to the incompressible Euler equations; see section 3.1 (and, if the constraints are dropped, special solutions to the inviscid Burgers equation in section 4.1).

Schrödinger’s smoke is an approximation to the zero-momentum incompressible Euler solutions, where the Hamiltonian corresponding to Equation 9.10 is replaced by a sum of two independent Hamiltonian systems

This approximation corresponds to dropping the cross-terms in the original kinetic energy and adding the Fisher information functionals as potentials for and . Applying the two-component Madelung transform

and setting gives the incompressible Schrödinger equation

where, as before, the pressure function is a Lagrange multiplier for the pointwise constraint . Notice that the resulting equation is a wave-map equation on ; cf., e.g., Reference 74.

Remark 9.9.

It has been claimed that numerical solutions to the incompressible Schrödinger equations (ISE) yield realistic visualization of the dynamics of smoke; see Reference 15. However, it is an open question in what sense (or, in which regime) these solutions are approximations to solutions of the incompressible Euler equations.

9.5. Madelung transform as a Kähler morphism

We now assume that both the cotangent bundle and the projective space are equipped with suitable Riemannian structures. Consider first the bundle . Its elements can be described as 4-tuples where , , and are subject to the constraint

Definition 9.10.

The Sasaki (or Sasaki–Fisher–Rao) metric on is the cotangent lift of the Fisher–Rao metric Equation 7.7, namely

On the projective space we define the infinite-dimensional Fubini–Study metric

Theorem 9.11 (Reference 40).

The Madelung transform Equation 9.3 with is an isometry, up to a factor , between the spaces equipped with Equation 9.11 and equipped with Equation 9.12.

Since the Fubini–Study metric together with the complex structure of defines a Kähler structure, it follows that also admits a natural Kähler structure which corresponds to the canonical symplectic structure. Note that an almost complex structure on , which is related via the Madelung transform to the Wasserstein–Otto metric, does not integrate to a complex structure; cf. Reference 60. In fact, it was shown in Reference 40 that the corresponding complex structure becomes integrable (and considerably simpler) when the Fisher–Rao metric is used in place of the Wasserstein–Otto metric. It would be interesting to write down the Kähler potentials for all metrics compatible with the corresponding complex structure on and identify those that are invariant under the action of the diffeomorphism group.

Example 9.12.

The 2-component Hunter–Saxton (2HS) equation is the following system

where and are time-dependent periodic functions on the real line. It can be viewed as a high-frequency limit of the two-component Camassa–Holm equation; cf. Reference 81.

It turns out that Equation 9.13 describes the geodesic flow of a right-invariant -type metric on the semidirect product of the group of circle diffeomorphisms that fix a prescribed point and the space of -valued maps of a circle. Furthermore, there is an isometry between subsets of the group and the unit sphere in the space of wave functions ; see Reference 44. In Reference 40 it is proved that the 2HS equation Equation 9.13 with initial data satisfying is equivalent to the geodesic equation of the Sasaki–Fisher–Rao metric Equation 9.11 on and the Madelung transformation induces a Kähler map to geodesics in equipped with the Fubini–Study metric.

Note also that (subject to the -invariant condition ) the 2-component Hunter–Saxton equation Equation 9.13 reduces to the standard Hunter–Saxton equation. This is a consequence of the fact that horizontal geodesics on with the Sasaki–Fisher–Rao metric descend to geodesics on with the Fisher–Rao metric.

10. Casimirs in hydrodynamics

In this section we start by surveying results on Casimirs for inviscid incompressible fluids, and then continue with compressible and magnetic hydrodynamics. Recall that a Casimir on the dual of a Lie algebra is a function that is invariant under the coadjoint action of the corresponding group . Note that Casimirs are first integrals for Hamiltonian dynamics on for any choice of Hamiltonian functions.

10.1. Casimirs for ideal fluids

The Hamiltonian description of the dynamics of an ideal fluid gives some insight into the nature of its first integrals. Recall that the Euler equation is a Hamiltonian system on the dual space with respect to the Poisson–Lie structure and with the fluid energy as the Hamiltonian; see section 3.1. In this setting we have

Proposition 10.1 (Reference 8Reference 65).

For the group the following functionals are Casimirs on the dual space (the space of cosets

If , then the functional

is a Casimir function on .

If , then the functionals

are Casimir functions on for any measurable function .

Here, the quotient of a -form and the volume form is a function, which being composed with can be integrated against the volume form over .

Proof.

First, we have to check that and are well-defined functionals on . Note that for any exact -form we have and . Similarly, we find that each of the functionals and depends on a coset but not on a representative, e.g., . Furthermore, the group acts on by change of coordinates for any . Since both and are defined in a coordinate-free way, they are invariant under this action.

Corollary 10.2.

For a velocity field satisfying the incompressible Euler equations in the functionals and computed for the -forms (related to by the Riemannian metric on are first integrals in odd and even dimension, respectively.

Proof.

The Euler equations for in the Lie algebra become Hamiltonian when rewritten for with respect to the standard Lie–Poisson bracket on the dual space . For this Hamiltonian system the trajectories always remain tangent to coadjoint orbits of . By Proposition 10.1, the functions and are constant on coadjoint orbits and hence are constant along the Euler trajectories.

Remark 10.3.

The functionals and are Casimirs of the Lie–Poisson bracket on ; i.e., they yield conservation laws for any Hamiltonian equation on this space. In particular, both and are first integrals of the Euler equations for an arbitrary metric on . They express “kinematic symmetries” of the hydrodynamical system, while the energy is an invariant related to the system’s “dynamics.”

Example 10.4.

If is a domain in , then the function

is a first integral of the Euler equations, where the -form is related to the velocity field by means of the Euclidean metric. The last integral has a natural geometric meaning of the helicity of the vector field defined by .

Example 10.5.

Similarly, if is a domain in , we find infinitely many first integrals of the Euler equations, namely

where is the vorticity function on .

Remark 10.6.

While the functions and on are Casimirs, generally speaking, they do not form a complete set of invariants of the coadjoint representation.

In the two-dimensional case the complete set of invariants includes a measured Reeb graph of the vorticity function and circulation data of the field on the surface ; see Reference 35. In the three-dimensional case the invariant is shown to be unique among -Casimirs Reference 22, while there are more invariants of ergodic nature (such as pairwise linkings of the trajectories of the vorticity field) that are not continuous functionals Reference 6.

10.2. Casimirs for barotropic fluids

In many respects the behaviour of barotropic compressible fluids is similar to that of incompressible fluids (while the fully compressible fluids resemble thermodynamical rather than mechanical systems). In particular, their Hamiltonian description suggests similar sets of Casimir invariants of motion. While the incompressible Euler equations on a manifold are geodesic equations on the group and hence a Hamiltonian system on the corresponding dual space , the equations of compressible barotropic fluids Equation 4.4 are known to be related to the semidirect product group ; see section 5.1. Its Lie algebra is and the corresponding dual space was described in section 3.2.

The equations of barotropic fluids are Hamiltonian equations on with the Lie–Poisson bracket given by the formula Equation 3.7 and the invariants of the corresponding coadjoint action, i.e., the Casimir functions, are the first integrals of the equations of motion.

Recall that the smooth part of the dual of the semidirect product algebra can be identified with via the pairing

In what follows we restrict to the subset of corresponding to everywhere positive densities on . It turns out that the equations of incompressible fluid also have an infinite number of conservation laws in the even-dimensional case and possess at least one first integral in the odd-dimensional case; see section 10.1 and Reference 8Reference 65.

The following proposition shows that Casimir functions for a barotropic fluid are similar to the ones for an incompressible fluid.

Proposition 10.7 (Reference 33Reference 65).

Let and . If , then the functional

is a Casimir function on .

If , then for any measurable function the functional

is a Casimir function on .

Proof.

The proof is based on the fact that the coadjoint action of the group on the dual space is given by

Thus, and transform according to the rules and , and it is now straightforward to check that the functionals and are invariant under such transformations. Indeed, up to the change of coordinates by a diffeomorphism , the -form changes within its coset and the functionals and are well defined on the cosets.

The above argument shows that, in a certain sense, a barotropic fluid “becomes incompressible” when viewed in a coordinate system which “moves with the flow.” The Hamiltonian approach makes it possible to apply Casimir functions to study stability of barotropic fluids and inviscid MHD systems: their dynamics are confined to coadjoint orbits of the corresponding groups and Casimir functions can be used to describe the corresponding conditional extrema of the Hamiltonians.

10.3. Casimirs for magnetohydrodynamics

We start with the three-dimensional incompressible magnetohydrodynamics described in section 5.2; cf. equations Equation 5.4. In this case the configuration space of a magnetic fluid is the semidirect product of the volume preserving diffeomorphism group and the dual space of the Lie algebra of divergence-free vector fields on a -manifold . The semidirect product algebra is , and its action is given by formula Equation 5.2. The corresponding dual space is

and the Poisson brackets on are given by Equation 3.7, interpreted accordingly.

Proposition 10.8 (Reference 8Reference 34).

Let be a manifold with , and let and . Then the magnetic helicity

and the cross-helicity

are Casimir functions on .

The condition ensures that any magnetic field has a vector potential . It turns out that these are the only Casimirs for incompressible magnetohydrodynamics—any other sufficiently smooth Casimir is a function of these two; cf. Reference 22.

Consider now the setting of compressible magnetohydrodynamics on a Riemannian manifold of arbitrary dimension; see Equation 6.7. Recall also from section 6.2 that the semidirect product group associated with the compressible MHD equations is

The corresponding Lie algebra is

with dual

where is the space of closed -forms referred to as magnetic -forms. Recall that if is a threefold, then a magnetic vector field and a magnetic -form are related by . We again confine our constructions to positive densities .

Proposition 10.9.

Let , , and . If , then the generalized cross-helicity functional

is a Casimir function on .

If and , so that for some -form , then

is a Casimir function on .

If , then for any measurable function the functional

is a Casimir function on .

If is a vector field on defined by , then the functional can be equivalently written as

In the three-dimensional () and incompressible () cases it reduces to the cross-helicity functional of Proposition 10.8.

Proof.

The coadjoint action is

where the vector field is defined by the condition . Since both and are transported by , the only nontrivial functional to check is the generalized cross-helicity .

For this purpose we first note that since is closed, then so is . Hence, the change of variables formula gives

where the last term on the right-hand side vanishes by Stokes’ theorem while the -form vanishes pointwise on . The latter holds since evaluating this form on any linearly independent vectors tangent to is equivalent to evaluating on any linearly dependent set of tangent vectors containing , which is evidently zero.

Remark 10.10.

In two and three dimensions such Casimirs in different terms were described in Reference 34. Other differential-geometric invariants of hydrodynamical equations include Ertel-type invariants Reference 80, local invariants Reference 2Reference 3, invariants of Lagrangian type Reference 10, and many others.

Appendix A. Symplectic and Poisson reductions

A.1. Symplectic reduction

In sections 3.2 and 3.3 we described Poisson reduction on with respect to the cotangent action of . This lead to reduced dynamics on the Poisson manifold (Theorem 3.6). Furthermore, any Hamiltonian system descends to symplectic leaves and with the canonical symplectic structure is one of the symplectic leaves of . In this appendix we shall describe symplectic reduction which leads to the same manifold —the symplectic quotient corresponding to the cotangent bundle equipped with the canonical symplectic structure (for a more thorough treatment; see Reference 50).

As before, let be the Lie algebra of . Recall that the dual space is naturally isomorphic to ; see Theorem 3.1.

Lemma A.1.

The (smooth) dual can be identified with the quotient space

where is taken over smooth functions on . The cotangent left action of on is Hamiltonian. The associated momentum map is given by

where and . The momentum map is equivariant, i.e.,

for all .

Proof.

From the Hodge decomposition it follows that for all if and only if for some . This proves Equation A.1.

From the standard Lie–Poisson theory (see, e.g., Reference 50Reference 51) we find that the momentum map for acting on is given by . Since is a subgroup of , it follows from Equation A.1 that the momentum map must be Equation A.2.

Regarding the equivariance statement, we have

as required.

Lemma A.2.

The zero momentum level set

is invariant under the action of , i.e., for any and one has .

Proof.

We have so that if , then

Next, assume that and write for some . Since it follows that must be exact, i.e., . Thus, , and so is exact. The fact that is invariant under follows from the equivariance property in Lemma A.1, since for all . This concludes the proof.

To identify the symplectic structure of the quotient, we shall first identify the momentum map associated with the action of on . In what follows we will use the notation for the density corresponding to the density function .

Lemma A.3.

The associated momentum map for the left cotangent action of on is given by

Proof.

The smooth dual of is with the natural pairing

Since is a subspace of , it follows that

The infinitesimal left action of is and the momentum map is then given by

By Cartan’s formula we obtain

which proves the lemma.

The main result of this section is

Theorem A.4.

The zero momentum symplectic quotient

is isomorphic, as a symplectic manifold, to and the symplectomorphism is given by

Thus can be viewed as a symplectic leaf of the Poisson manifold . Theorem A.4 is an infinite-dimensional variant of the following general result. For a homogeneous space the zero momentum reduction space is symplectomorphic to through the mapping

where is the momentum map for the natural action of on ; see Reference 51.

Proof.

Let so that . If , then

By the Moser–Hamilton result in Lemma 2.8 it follows that in the Fréchet category we have

Thus, the symplectic quotient is naturally identified with a subbundle of the Poisson manifold in Theorem 3.6. By conservation of momentum this subbundle is invariant under the flow of any Hamiltonian. To prove that it is a symplectic leaf, it suffices to show that the map corresponding to Equation A.3

is a diffeomorphism and Poisson. The former follows from the fact that the kernel of on is trivial. It thus remains to show that

for any .

We have

and

Combining Equation A.4 and Equation A.5 we get

This concludes the proof.

A.2. Reduction and momentum map for semidirect product groups

We exhibit here geometric structures behind the semidirect product reduction generalizing the considerations of sections 5.1 and 5.2. The main point of this appendix is that the semidirect product approach is just a convenient way of presenting various Newton’s systems on for which the symmetry group is a proper subset of : this way various quotient spaces appear as invariant sets in the vector space which is the dual of an appropriate Lie algebra.

Let be a subgroup of . Suppose that acts from the left on a linear space (a left representation of ). For instance, for compressible fluids in section 6.1 and compressible MHD in section 6.2, the space was taken to be the spaces of functions or the dual of the space of divergence-free vector fields , while can be a subgroup of volume-preserving diffeomorphisms . However the consideration below is more general.

The quotient space of left cosets is acted upon from the left by . Assume now that the quotient is a manifold and it can be embedded as an orbit in , while denotes the embedding. Since the action of on induces a linear left dual action on we can construct the semidirect product . Let be the dual of the corresponding semidirect product algebra .

Proposition A.5.

The quotient is naturally embedded via a Poisson map in the Lie–Poisson space .

Proof.

The Poisson embedding is given by

where we used the identifications

and . Recall that the Lie algebra of is the space of vector fields on whose dual is . The action of on is given by

where and is the momentum map associated with the cotangent lifted action of on . The corresponding infinitesimal action of is

Since the second component is only acted upon by (or ) but not (or ), it follows from the embedding of as an orbit in that we have a natural Poisson action of (or ) on via the Poisson embedding Equation A.6. Notice that the momentum map of (or ) acting on is tautological (i.e., the identity) this follows from the fact that the Hamiltonian vector field on for is given by Equation A.7.

We now return to the standard symplectic reduction (without semidirect products). The dual of the subalgebra is naturally identified with the affine cosets of such that

The momentum map of the subgroup acting on by is then given by , since the momentum map of acting on is the identity. If , i.e., , then is in the zero momentum coset. Since we also have , this gives us an embedding as a symplectic leaf in . The restriction to this leaf is called the zero-momentum symplectic reduction.

Turning next to the semidirect product reduction, we now have Poisson embeddings of in and of in . The combined embedding of as a symplectic leaf in is given by the map

This implies that we have a Hamiltonian action of (or ) on the zero-momentum symplectic leaf inside , which in turn lies inside .

Since provides a natural symplectic action on and since is an orbit in we have, by restriction, a natural action of on . Furthermore, since the momentum map associated with the group acting on is the identity, the Poisson embedding map Equation A.6 is the momentum map for acting on . Thus, the momentum map of acting on is given by Equation A.8.

The above consideration leads to the Madelung transform.

Theorem A.6 (Reference 40).

Semidirect product reduction and Poisson embedding for the subgroup coincides with the inverse of the Madelung transform defined in Section 9.

Appendix B. Tame Fréchet manifolds

A natural functional-analytic setting for the results presented in this paper is that of tame Fréchet spaces; cf. Hamilton Reference 30. An alternative setting for groups of diffeomorphisms deals with Sobolev completions (or any reasonably strong Banach topology) of the corresponding function spaces Reference 20. If , then the Sobolev completions of the diffeomorphism groups and are smooth Hilbert manifolds but not Banach Lie groups since, e.g., the left multiplication and the inversion maps are not even uniformly continuous in the topology.

B.1. Tame Fréchet structures on diffeomorphism groups

On the other hand, both and can be equipped with the structure of tame Fréchet Lie groups. In this setting becomes a closed tame Lie subgroup of which can be viewed as a tame principal bundle over the quotient space of either left or right cosets. Furthermore, the tangent bundle over is also a tame manifold. However, since the dual of a Fréchet space, which itself is not a Banach space, is never a Fréchet space, to avoid working with currents on it is expedient to restrict to a suitable subset of the (full) cotangent bundle over .

More precisely, consider the tensor product of the cotangent bundle and the vector bundle of -forms on and define another bundle over whose fibre over is the space of smooth sections of the pullback bundle over . We will refer to this object as (the smooth part of) the cotangent bundle of and denote it also by . We will write and . Throughout the paper we will assume that derivatives of various Hamiltonian functions can be viewed as maps to the smooth cotangent bundle of the phase space.

Lemma B.1.

is a tame Fréchet manifold, and the map

is an isomorphism of tame Fréchet manifolds.

Proof.

Recall that is an open subset of , and observe that is the inverse image of under the smooth tame projection between tame Fréchet manifolds and . The argument is routine: the space is trivialized by the fiber mapping since is a diffeomorphism, while the fact that the fiber mapping is smooth and tame with a smooth tame inverse follows since is a tame Fréchet Lie group (all group operations are smooth tame maps).

Let and . As before in Equation 3.3 we have the pairing

between the fibers and .

Our goal in this section is to describe Poisson reduction of with respect to the right action of as a smooth tame principal bundle. We will use the Poisson bivector for the canonical symplectic structure on , which we identify with its right trivialization as in Lemma B.1. By construction, each element of can be viewed as a tensor product of a -form and a volume form on . Choose and note that for each the Poisson bivector on is a bilinear form on defined by

Lemma B.2.

The bivector induces a smooth tame vector bundle isomorphism

which at any point is given by

for any and .

Proof.

First, observe that one can identify the tangent and cotangent bundles of with and , respectively. The formula in Equation B.1 can be verified by a direct calculation from

for any and using integration by parts and the assumption that has no boundary. Smoothness of follows from the fact that all the operations in Equation B.2 are smooth tame maps. The inverse of is given by

Again, all the operations involve smooth tame maps, which implies that the inverse is also smooth.

Remark B.3.

That is a symplectomorphism corresponds to the fact that

is a symplectic manifold with canonical symplectic structure

The space is a tame Fréchet manifold, since so are both and .

Next, consider the Poisson bivector defined on the tame Fréchet manifold by

for the density .

Lemma B.4.

The bivector induces a smooth tame vector bundle homomorphism

which at any point is given by

for any and .

Proof.

The proof follows the same steps as the proof of Lemma B.2 with the adjustment that now is only a homomorphism, rather than an isomorphism, of vector bundles.

Remark B.5.

The Hamiltonian equations on and on , as discussed in the previous sections, can now be written as

Notice that corresponds to a Poisson structure, but not to a symplectic structure as does; is not invertible whereas is.

The next theorem is the main result of this section.

Theorem B.6.

The following diagram

is a smooth tame principal bundle. The projection is a Poisson submersion with respect to the Poisson structure on . Solutions to the Hamiltonian equations for a -invariant Hamiltonian on project to solutions of the Hamiltonian equations for the (unique) Hamiltonian on satisfying .

Proof.

First, consider the map from to the product and observe that it is a smooth tame vector bundle isomorphism, as in Lemma B.1. The cotangent action of on acts on the first component by composition and is clearly also a smooth tame map. Furthermore, we have

The fact that is a smooth tame principal bundle over with fiber follows from the Nash–Moser–Hamilton theorem; cf., e.g., Reference 30, Thm. III.2.5.3. Consequently, is a smooth tame principal bundle over with fiber .

The projection is a Poisson submersion and smooth solutions are mapped to smooth solutions: this follows from Lemmas B.2 and B.4 together with a straightforward calculation showing that whenever .

Remark B.7.

We point out that the situation is more complicated if one works with Banach spaces such as Sobolev or Hölder . In those settings the results in Lemma B.1, Lemma B.2, Lemma B.4 and Theorem B.6 need not hold. For example, the bundle projection in Theorem B.6 typically fails to be Lipschitz continuous in the topology.

B.2. Short-time existence of compressible Euler equations

We include here a local existence result that applies to all the examples in Section 4. To this end consider the compressible Euler equations on a compact manifold in the form

where is the thermodynamical work function defined in Equation 4.6. The equations discussed previously can be captured by different choices of the functions and . If is strictly increasing, then short-time solutions of these equations can be obtained using standard techniques; see, e.g., Reference 30, Thm. III.2.1.2 for a result for the shallow water equations Equation 1.6 corresponding to .

Theorem B.8.

For any , and any smooth function such that , there exists a unique smooth solution of the equations Equation B.3 satisfying , and defined in some open neighborhood of .

Proof.

The basic idea is to transform Equation B.3 so that its linearization becomes a symmetric linear system. This can be achieved by a substitution where is a new density function and is the solution of the scalar initial value problem

Compactness of together with the assumption ensure that the right-hand side of Equation B.4 is Lipschitz continuous in the interval given by the range of . Thus, there is a smooth solution whose range covers the range of . Since and , this solution is strictly increasing.

It follows that the corresponding linearized equations form a symmetric linear system in a neighborhood of the density and thus admit a unique tame solution by the general theory of symmetric systems. Applying the Nash–Moser–Hamilton theorem completes the proof.

Using the results in section B.1, it is possible to deduce from the above theorem short-time existence results for each of the equations considered in Section 4: the Newton systems on , the Poisson systems on , or the canonical Hamiltonian systems on .

Acknowledgments

The authors are grateful to the anonymous referee for many helpful remarks.

About the authors

Boris Khesin is professor of mathematics at the University of Toronto, Canada. His research interests are in geometric and topological hydrodynamics, infinite-dimensional groups, and Hamiltonian and integrable systems.

Gerard Misiołek is professor of mathematics at the University of Notre Dame. He specializes in geometric analysis and partial differential equations.

Klas Modin is professor of mathematics at Chalmers University of Technology and the University of Gothenburg in Sweden. His research interests include Hamiltonian PDEs, shape analysis, and two-dimensional turbulence, often in combination with computational mathematics.

Table of Contents

  1. Abstract
  2. Untitled Section
  3. 1. Introduction
    1. 1.1. Geodesics and Newton’s equations: finite-dimensional examples
    2. 1.2. Three motivating examples from hydrodynamics
    3. 1.3. Riemannian metrics and their geodesics on spaces of diffeomorphisms and densities
    4. Definition 1.2.
    5. Example 1.5.
    6. 1.4. First examples of Newton’s equations on diffeomorphism groups
    7. Example 1.7 (Shallow water equation as a Newton’s equation).
    8. Proposition 1.8.
    9. Example 1.10 (The -body problem as a Newton’s equation).
    10. 1.5. Other related equations
    11. 1.6. An overview and main results
    12. Spaces of densities.
  4. 2. Wasserstein–Otto geometry
    1. 2.1. Newton’s equations on
    2. Definition 2.1.
    3. Theorem 2.2 (7273).
    4. Lemma 2.3.
    5. Proposition 2.4.
    6. 2.2. Riemannian submersion over densities
    7. Definition 2.5.
    8. Definition 2.6.
    9. Theorem 2.7 (64).
    10. Lemma 2.8.
    11. Lemma 2.9.
  5. 3. Hamiltonian setting
    1. 3.1. Hamiltonian framework for the incompressible Euler equations
    2. Theorem 3.1 (See, e.g., 8).
    3. 3.2. Hamiltonian formulation and Poisson reduction
    4. Lemma 3.4.
    5. Theorem 3.5.
    6. Theorem 3.6 (Poisson reduction).
    7. Corollary 3.8.
    8. Proposition 3.9.
    9. 3.3. Newton’s equations on
    10. Lemma 3.10.
    11. Corollary 3.11.
  6. 4. Wasserstein–Otto examples
    1. 4.1. Inviscid Burgers equation
    2. Proposition 4.1.
    3. 4.2. Classical mechanics and Hamilton–Jacobi equations
    4. Proposition 4.2.
    5. 4.3. Barotropic fluid equations
    6. Proposition 4.3.
  7. 5. Semidirect product reduction
    1. 5.1. Barotropic fluids via semidirect products
    2. Theorem 5.1 (5278).
    3. 5.2. Incompressible magnetohydrodynamics
    4. Theorem 5.3 (878).
  8. 6. More general Lagrangians
    1. 6.1. Fully compressible fluids
    2. Theorem 6.1.
    3. Lemma 6.2.
    4. 6.2. Compressible magnetohydrodynamics
    5. Lemma 6.4.
    6. Theorem 6.5.
    7. Corollary 6.6.
    8. 6.3. Relativistic inviscid Burgers equation
    9. Theorem 6.7.
  9. 7. Fisher–Rao geometry
    1. 7.1. Newton’s equations on
    2. Definition 7.1.
    3. Theorem 7.3.
    4. Proposition 7.4.
    5. 7.2. Riemannian submersion over densities
    6. Definition 7.6.
    7. Definition 7.7.
    8. Theorem 7.8.
    9. Proposition 7.10.
    10. Theorem 7.11.
    11. 7.3. Newton’s equations on
    12. Proposition 7.12.
    13. Theorem 7.13.
  10. 8. Fisher–Rao examples
    1. 8.1. The CH equation and Fisher–Rao geodesics
    2. Proposition 8.1.
    3. Proposition 8.2.
    4. 8.2. The infinite-dimensional Neumann problem
    5. Proposition 8.3.
    6. Lemma 8.4.
    7. Proposition 8.5.
    8. 8.3. The Klein–Gordon equation
    9. Proposition 8.7.
  11. 9. Geometric properties of the Madelung transform
    1. Definition 9.1.
    2. 9.1. Madelung transform as a symplectomorphism
    3. Theorem 9.2 (40).
    4. 9.2. Examples: linear and nonlinear Schrödinger equations
    5. Proposition 9.3 (cf. 4879).
    6. 9.3. The Madelung and Hopf–Cole transforms
    7. Definition 9.7.
    8. 9.4. Example: Schrödinger’s smoke
    9. 9.5. Madelung transform as a Kähler morphism
    10. Definition 9.10.
    11. Theorem 9.11 (40).
    12. Example 9.12.
  12. 10. Casimirs in hydrodynamics
    1. 10.1. Casimirs for ideal fluids
    2. Proposition 10.1 (865).
    3. Corollary 10.2.
    4. Example 10.4.
    5. Example 10.5.
    6. 10.2. Casimirs for barotropic fluids
    7. Proposition 10.7 (3365).
    8. 10.3. Casimirs for magnetohydrodynamics
    9. Proposition 10.8 (834).
    10. Proposition 10.9.
  13. Appendix A. Symplectic and Poisson reductions
    1. A.1. Symplectic reduction
    2. Lemma A.1.
    3. Lemma A.2.
    4. Lemma A.3.
    5. Theorem A.4.
    6. A.2. Reduction and momentum map for semidirect product groups
    7. Proposition A.5.
    8. Theorem A.6 (40).
  14. Appendix B. Tame Fréchet manifolds
    1. B.1. Tame Fréchet structures on diffeomorphism groups
    2. Lemma B.1.
    3. Lemma B.2.
    4. Lemma B.4.
    5. Theorem B.6.
    6. B.2. Short-time existence of compressible Euler equations
    7. Theorem B.8.
  15. Acknowledgments
  16. About the authors

Figures

Table 1.

Examples of Newton’s equations.

Wasserstein–Otto geometry Fisher–Rao geometry
Newton’s equations on
• Inviscid Burgers equation (§4.1) -Camassa–Holm equation (§8.1)
• Classical mechanics (§4.2) • Optimal information transport (§7)
• Barotropic inviscid fluid (§4.3)
• Fully compressible fluid (§6.1)
• Magnetohydrodynamics (§6.2)
Newton’s equations on
• Hamilton–Jacobi equation (§4.2) -dim Neumann problem (§8.2)
• Linear Schrödinger equation (§9.2) • Klein–Gordon equation (§8.3)
• Nonlinear Schrödinger (§9.2) • 2-component Hunter–Saxton (§9.5)
Table 2.

Various PDEs as Newton’s equations on .

Equation on Potential Section
inviscid Burgers§4.1
Hamilton–Jacobi,§4.2
shallow-water§1.4
barotropic compressible Euler,§4.3
linear Schrödinger§9.2
nonlinear Schrödinger (NLS)§9.2
Figure 1.

Illustration of the Riemannian submersion in Theorem 2.7. Horizontal geodesics on (potential solutions) are transversal to the fibres and project to geodesics on . Note that the point in denoted by corresponds to the reference volume form , while corresponds to the volume density .

Graphic without alt text
Figure 2.

Relation between various phase space representations of Newton’s equations on and .

Figure 3.

The bottom arrow is the isometry of the density space with the Fisher–Rao metric and a part of the infinite-dimensional sphere, while the top arrow corresponds to the Madelung transform.

Mathematical Fragments

Equation (1.1)
Equation (1.2)
Equation (1.3)
Equation (1.4)
Example 1.7 (Shallow water equation as a Newton’s equation).

We next proceed to describe Newton’s equations on the diffeomorphism group . To this end we consider the case of a potential on which depends only on the density carried by a diffeomorphism , i.e., the potential for is a pullback for the projection , where as we take a simple quadratic function

on the space of densities. It turns out that with this potential we obtain shallow water equations. There are several equivalent formulations, depending on the functional setting.

Proposition 1.8.

Newton’s equations with respect to the -metric 2.1 and the potential Equation 1.5 take the following forms:

on

where ;

the shallow water equations on

where is the horizontal velocity field and is the water depth;

for the gradient velocity they assume the Hamilton–Jacobi form

Example 1.10 (The -body problem as a Newton’s equation).

Newton’s law of gravitation states that for a body with mass distribution , the associated potential is , where is the gravitational constant. Following the above framework, the potential function on is given by

where is a (suitably defined) inverse Laplacian with appropriate boundary conditions.

The corresponding fluid system is described by

Thus, we have arrived at a fluid dynamics formulation of a continuous Newton mass system under the influence of gravity: a “fluid particle” positioned at experiences a gravitational pull corresponding to the potential . In particular, if we obtain the well-known Green’s function for the Laplacian

We now wish to study weak solutions to these equations where the mass distribution is replaced by an atomic measure

for point masses positioned at . The differential-geometric setting is as follows. We have a Riemannian metric on (the Wasserstein—Otto metric) and a potential function on (the Newton potential). The group acts on the (finite-dimensional) manifold of atomic measures with particles. Clearly, we have . The isotropy subgroup for this action on is

Although the horizontal distribution is not defined rigorously, it is formally given by vector fields with support on . With this notion of horizontality, the projection , given by , is a Riemannian submersion with respect to the weighted Riemannian structure on , given by . For , substituting the atomic measure into the formula 1.8 and using the Green’s function for gives

The resulting finite-dimensional Riemannian structure together with this potential function defines the kinetic and potential energies giving rise to the -body problem.

Remark 1.12.

More precisely, given a compact -dimensional manifold , we will equip the group of diffeomorphisms and the space of nonvanishing probability densities with the structures of smooth infinite-dimensional manifolds (see Appendix B for details) and study Newton’s equations on these manifolds viewed as the associated configuration spaces.

As a brief preview of what follows, let be the subgroup of diffeomorphisms preserving the Riemannian volume form of . Consider the fibration of the group of all diffeomorphisms over the space of densities

discussed by Moser Reference 61, whose cotangent bundles and are related by a symplectic reduction; cf. Section 3 below. Moser’s construction can be used to introduce two different algebraic objects: the first is obtained by identifying with the left cosets

and the second by identifying it with the right cosets

In this paper we will make use of both identifications.

In order to define Newton’s equations on and , and to investigate their mutual relations, we will choose Riemannian metrics on both spaces so that the natural projections corresponding to 1.9 or 1.10 become (infinite-dimensional) Riemannian submersions. We will consider two such pairs of metrics. In Section 2, using left cosets, we will study a noninvariant -metric on together with the Wasserstein–Otto metric on . In Section 7, using right cosets, we will focus on a right-invariant metric on and the Fisher–Rao information metric on . Extending the results of von Renesse Reference 79, we will then derive in Section 9 various geometric properties of the Madelung transform. This will allow us to represent Newton’s equations on as Schrödinger-type equations for wave functions.

Definition 2.1.

The -metric on is given by

or, equivalently, after a change of variables

where , , and is a vector field on .

Equation (2.3)
Equation (2.4)
Theorem 2.2 (Reference 72Reference 73).

Newton’s equations on for the metric Equation 2.1 and a potential function Equation 2.4 can be written as

In reduced variables and the equations assume the form

Lemma 2.3.

If is of the form Equation 2.4, then

where .

Proposition 2.4.

The gradient fields form an invariant set of solutions of the reduced equations Equation 2.6. Expressed in and , these solutions fulfill the Hamilton–Jacobi equations

Definition 2.5.

The left coset projection between the space of diffeomorphisms and the space of probability densities is given by

or, equivalently, by pushforward of the reference volume form .

Definition 2.6.

The Wasserstein–Otto metric is a Riemannian metric on given by

where is defined by the transport equation

and is a tangent vector at .

Theorem 2.7 (Reference 64).

The projection Equation 2.7 is an (infinite-dimensional) Riemannian submersion with respect to the -metric on and the Wasserstein–Otto metric on . Namely, given a horizontal ⁠Footnote3 vector one has

3

That is, such that for all .

where .

Lemma 2.8.

Let be the left coset projection Equation 2.7. Then

is a principal bundle. Consequently, the quotient space of left cosets is isomorphic to .

Equation (3.1)
Theorem 3.1 (See, e.g., Reference 8).
a)

The dual space to the Lie algebra is , the space of cosets of -forms on modulo exact -forms. The coadjoint action of is given by change of coordinates in a -form, while the coadjoint action of is given by the Lie derivative along a vector field . It is well-defined on the cosets in .

b)

The inertia operator is defined by assigning to a given divergence-free vector field the coset in .

c)

The incompressible Euler equations Equation 1.3 on the dual space have the form

where and .

Equation (3.3)
Lemma 3.4.

The Hamiltonian corresponding to the Lagrangian is

Theorem 3.5.

The Hamiltonian form of the equations Equation 2.5 is

where .

Equation (3.6)
Theorem 3.6 (Poisson reduction).

The quotient space is isomorphic to . The isomorphism is given by the projection . Furthermore, is a Poisson map with respect to the canonical Poisson structure on and the Poisson structure on given by

Corollary 3.8.

Let be a Hamiltonian function on satisfying

Then for some function . In reduced variables and , the Hamiltonian equations assume the form

where .

Proposition 3.9.

The product manifold

is a Poisson submanifold of .

Corollary 3.11.

The Hamiltonian on corresponding to Newton’s equations on with respect to the Wasserstein–Otto metric Equation 2.8 is

and the Hamiltonian equations are

Solutions of 3.9 correspond to horizontal solutions of Newton’s equations Equation 2.5 on or, equivalently, to zero-momentum solutions of the reduced equations Equation 3.8 with Hamiltonian

Proposition 4.1.

Newton’s equations with respect to the -metric Equation 2.1 and with zero potential admit the following formulations:

the geodesic equations on

the inviscid Burgers equations on

where ;

the Poisson reduced equations on

where ;

the symplectically reduced equations on

corresponding to the Hamiltonian form of the geodesic equations for the Wasserstein–Otto metric Equation 2.8.

Equation (4.2)
Proposition 4.2.

Newton’s equations with respect to the -metric Equation 2.1 and the potential in Equation 4.2 admit the following formulations:

the geodesic equations with potential on

the inviscid Burgers equations with potential on

where ;

the Poisson reduced equations on

where ;

the symplectically reduced equations on

Equation (4.4)
Equation (4.5)
Equation (4.6)
Equation (4.7)
Equation (4.8)
Equation (5.1)
Equation (5.2)
Equation (5.3)
Equation (5.4)
Equation (5.5)
Equation (6.1)
Theorem 6.1.

The fully compressible system Equation 6.1 is obtained using an embedding into the Lie–Poisson space , where (cf. Proposition A.5) from Newton’s equations on with Lagrangian

where is a potential function (of density and entropy density for some fixed initial entropy density of the form

and where the internal energy and pressure are related by

Lemma 6.2.

The momentum map for the cotangent action of on

is

Equation (6.3)
Equation (6.4)
Equation (6.5)
Equation (6.6)
Theorem 6.5.

The Poisson reduced form on

of the Euler–Lagrange equations for the Hamiltonian Equation 6.6 is

where the field is defined by and the momentum variable is . For a threefold these equations correspond to the equations of the compressible inviscid magnetohydrodynamics Equation 6.4 where the magnetic field is related to the closed -form by .

Corollary 6.6.

The equations Equation 6.7 admit special horizontal solutions corresponding to momenta of the form

These solutions can be expressed in the variables

as a canonical Hamiltonian system for the Hamiltonian

where and .

Equation (6.9)
Equation (6.10)
Equation (6.11)
Theorem 6.7.

The relativistic Lagrangian Equation 6.11 on induces a Poisson reduced system on . The Hamiltonian is given by

and the governing equations are

where .

Definition 7.1.

Let be a compact Riemannian manifold with volume form . For any and we set

where is the Laplacian on vector fields and is a quadratic form depending only on the vertical (divergence-free) component of .

Equation (7.2)
Theorem 7.3.

Newton’s equations of the metric Equation 7.1 on with a potential function Equation 7.2 have the form

where the inertia operator is given by

Proposition 7.4.

The Hamiltonian form of Newton’s equations Equation 7.3 on  is

where .

Definition 7.6.

The right coset projection between diffeomorphisms and smooth probability densities is given by

Definition 7.7.

The Fisher–Rao metric is the Riemannian metric on given by

where represents a tangent vector at .

Theorem 7.8.

The right coset projection Equation 7.6 is a Riemannian submersion with respect to the metric Equation 7.1 on and the Fisher–Rao metric on . In particular, if is horizontal, i.e.,

then where .

Remark 7.9.

Let us summarize the definition of the two metrics on that we discussed so far. The Wasserstein–Otto metric (cf. section 2.2) is defined as follows:

Note that it depends on the Riemannian structure on . The Fisher–Rao metric, on the other hand, is given by the “universal formula”,

and is independent of the Riemannian structure on .

Note also that the setting of Theorem 7.8 is quite different from that of Theorem 2.7. In the latter, the Riemannian metric on is right-invariant with respect to and automatically descends to the quotient from the right, namely . In the former, the metric is right-invariant with respect to and descends to the quotient from the left, namely . Thus, in Theorem 7.8 the right-invariance property is retained after taking the quotient, and therefore the Fisher–Rao metric on remains right-invariant with respect to the action of (corresponding to right translation of the fibers), which is easy to verify.

Proposition 7.10.

The gradient of a smooth function with respect to the Fisher–Rao metric is

where is a Lagrange multiplier such that .

Theorem 7.11.

The square root map

is, up to a factor , a Riemannian isometry between equipped with the Fisher–Rao metric in Equation 7.7 and the (geodesically convex) subset

of the sphere .

Proposition 7.12.

The exact momenta, i.e., tensor products of the form

form an invariant set for the system Equation 7.5.

Equation (7.10)
Theorem 7.13.

Newton’s equations with respect to the Fisher–Rao metric 7.7 on and a potential have the form

where is a multiplier subject to . Furthermore, the Lagrangian and Hamiltonian are and respectively. The corresponding Hamiltonian equations have the form

Solutions of 7.12 correspond to potential solutions (cf. Proposition 7.12) of Newton’s equations Equation 7.5 on .

Equation (8.1)
Proposition 8.1.

The CH equation Equation 8.1 is a (right-reduced) Newton’s equation Equation 7.3 with vanishing potential on . Geodesics of the Fisher–Rao metric Equation 7.7 on correspond to horizontal solutions of the CH equation described by the equations

(As in Theorem 7.13, the relation between , , and is given by , where is the Lagrangian flow of , and with the constant choosen so that

Equation (8.2)
Proposition 8.3.

Newton’s equations associated with the infinite-dimensional Neumann problem with potential Equation 8.2 have the form

where is a Lagrange multiplier subject to the constraint . In fact, we have .

Equation (8.4)
Remark 8.6.

Of particular interest are the stationary solutions to the Neumann problem Equation 8.3, i.e., those with , in which case is a normalized eigenvector of the Laplacian with eigenvalue . If , then . Consequently, the stationary solutions correspond to the principal axes of the corresponding infinite-dimensional ellipsoid .

It is also possible to obtain quasi-stationary solutions this way. Indeed, assume that the eigenspace of is at least two-dimensional (for example, when ). If are two orthogonal eigenvectors with eigenvalue , then it is straightforward to check that a solution originating from with initial velocity for is given by

Equation (8.5)
Definition 9.1.

Let and be real-valued functions on with . The Madelung transform is defined by

where is a parameter (Planck’s constant).⁠Footnote5

5

In the publications Reference 39Reference 40 the convention is used.

Equation (9.2)
Theorem 9.2 (Reference 40).

The Madelung transform Equation 9.1 induces a map

which is a symplectomorphism (in the Fréchet topology of smooth functions) with respect to the canonical symplectic structure of and the symplectic structure Equation 9.2 of .

Equation (9.4)
Equation (9.5)
Proposition 9.3 (cf. Reference 48Reference 79).

The Madelung transform maps the family of Schrödinger Hamiltonians Equation 9.5 to a family of Hamiltonians on given by

at the density . In particular, if , we recover Newton’s equations Equation 3.9 on for the potential function

where is Fisher’s information functional Equation 8.4. The extension Equation 2.6 to a fluid equation on is

Definition 9.7.

Let and be real-valued functions on with , and let be a positive constant. The (symmetrized) Hopf–Cole transform is the mapping defined by

Equation (9.8)
Remark 9.8.

Equation Equation 9.8 displays yet another relation to the heat flow connected to an invariant submanifold of . Consider the submanifold

A straightforward calculation shows that is an invariant submanifold for Equation 9.8 and the evolution on is given by a system of decoupled equations

Furthermore, since is the variational derivative of the entropy functional , it follows that the substitution (or ) corresponds to the momentum in the direction of negative entropy. This is related to the observation in Reference 64 that the heat flow is the -Wasserstein gradient flow of the entropy functional.

Equation (9.9)
Equation (9.10)
Definition 9.10.

The Sasaki (or Sasaki–Fisher–Rao) metric on is the cotangent lift of the Fisher–Rao metric Equation 7.7, namely

On the projective space we define the infinite-dimensional Fubini–Study metric

Example 9.12.

The 2-component Hunter–Saxton (2HS) equation is the following system

where and are time-dependent periodic functions on the real line. It can be viewed as a high-frequency limit of the two-component Camassa–Holm equation; cf. Reference 81.

It turns out that 9.13 describes the geodesic flow of a right-invariant -type metric on the semidirect product of the group of circle diffeomorphisms that fix a prescribed point and the space of -valued maps of a circle. Furthermore, there is an isometry between subsets of the group and the unit sphere in the space of wave functions ; see Reference 44. In Reference 40 it is proved that the 2HS equation 9.13 with initial data satisfying is equivalent to the geodesic equation of the Sasaki–Fisher–Rao metric Equation 9.11 on and the Madelung transformation induces a Kähler map to geodesics in equipped with the Fubini–Study metric.

Proposition 10.1 (Reference 8Reference 65).

For the group the following functionals are Casimirs on the dual space (the space of cosets

If , then the functional

is a Casimir function on .

If , then the functionals

are Casimir functions on for any measurable function .

Proposition 10.8 (Reference 8Reference 34).

Let be a manifold with , and let and . Then the magnetic helicity

and the cross-helicity

are Casimir functions on .

Lemma A.1.

The (smooth) dual can be identified with the quotient space

where is taken over smooth functions on . The cotangent left action of on is Hamiltonian. The associated momentum map is given by

where and . The momentum map is equivariant, i.e.,

for all .

Theorem A.4.

The zero momentum symplectic quotient

is isomorphic, as a symplectic manifold, to and the symplectomorphism is given by

Equation (A.4)
Equation (A.5)
Proposition A.5.

The quotient is naturally embedded via a Poisson map in the Lie–Poisson space .

Equation (A.6)
Equation (A.7)
Equation (A.8)
Lemma B.1.

is a tame Fréchet manifold, and the map

is an isomorphism of tame Fréchet manifolds.

Lemma B.2.

The bivector induces a smooth tame vector bundle isomorphism

which at any point is given by

for any and .

Lemma B.4.

The bivector induces a smooth tame vector bundle homomorphism

which at any point is given by

for any and .

Theorem B.6.

The following diagram

is a smooth tame principal bundle. The projection is a Poisson submersion with respect to the Poisson structure on . Solutions to the Hamiltonian equations for a -invariant Hamiltonian on project to solutions of the Hamiltonian equations for the (unique) Hamiltonian on satisfying .

Equation (B.3)
Equation (B.4)

References

Reference [1]
A. A. Agrachev and M. Caponigro, Controllability on the group of diffeomorphisms (English, with English and French summaries), Ann. Inst. H. Poincaré Anal. Non Linéaire 26 (2009), no. 6, 2503–2509, DOI 10.1016/j.anihpc.2009.07.003. MR2569905,
Show rawAMSref \bib{AgCa2009}{article}{ author={Agrachev, A. A.}, author={Caponigro, M.}, title={Controllability on the group of diffeomorphisms}, language={English, with English and French summaries}, journal={Ann. Inst. H. Poincar\'{e} Anal. Non Lin\'{e}aire}, volume={26}, date={2009}, number={6}, pages={2503--2509}, issn={0294-1449}, review={\MR {2569905}}, doi={10.1016/j.anihpc.2009.07.003}, }
Reference [2]
S. C. Anco and A. Dar, Classification of conservation laws of compressible isentropic fluid flow in spatial dimensions, Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci. 465 (2009), no. 2108, 2461–2488, DOI 10.1098/rspa.2009.0072. MR2519707,
Show rawAMSref \bib{AnDa2009}{article}{ author={Anco, Stephen C.}, author={Dar, Amanullah}, title={Classification of conservation laws of compressible isentropic fluid flow in $n>1$ spatial dimensions}, journal={Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci.}, volume={465}, date={2009}, number={2108}, pages={2461--2488}, issn={1364-5021}, review={\MR {2519707}}, doi={10.1098/rspa.2009.0072}, }
Reference [3]
S. C. Anco and G. M. Webb, Hierarchies of new invariants and conserved integrals in inviscid fluid flow, Phys. Fluids 32 (2020), 086104.
Reference [4]
V. I. Arnold, Variational principle for three-dimensional steady-state flows of an ideal fluid, Prikl. Mat. Mekh. 29 (1965).
Reference [5]
V. Arnold, Sur la géométrie différentielle des groupes de Lie de dimension infinie et ses applications à l’hydrodynamique des fluides parfaits (French), Ann. Inst. Fourier (Grenoble) 16 (1966), no. fasc. 1, 319–361. MR202082,
Show rawAMSref \bib{Ar1966}{article}{ author={Arnold, V.}, title={Sur la g\'{e}om\'{e}trie diff\'{e}rentielle des groupes de Lie de dimension infinie et ses applications \`a l'hydrodynamique des fluides parfaits}, language={French}, journal={Ann. Inst. Fourier (Grenoble)}, volume={16}, date={1966}, number={fasc. 1}, pages={319--361}, issn={0373-0956}, review={\MR {202082}}, }
Reference [6]
V. I. Arnold, The asymptotic Hopf invariant and its applications, Selecta Math. Soviet. 5 (1986), no. 4, 327–345. Selected translations. MR891881,
Show rawAMSref \bib{Ar2014}{article}{ author={Arnold, V. I.}, title={The asymptotic Hopf invariant and its applications}, note={Selected translations}, journal={Selecta Math. Soviet.}, volume={5}, date={1986}, number={4}, pages={327--345}, issn={0272-9903}, review={\MR {891881}}, }
Reference [7]
V. I. Arnold, Mathematical methods of classical mechanics, 2nd ed., Graduate Texts in Mathematics, vol. 60, Springer-Verlag, New York, 1989. Translated from the Russian by K. Vogtmann and A. Weinstein, DOI 10.1007/978-1-4757-2063-1. MR997295,
Show rawAMSref \bib{Ar1989}{book}{ author={Arnold, V. I.}, title={Mathematical methods of classical mechanics}, series={Graduate Texts in Mathematics}, volume={60}, edition={2}, note={Translated from the Russian by K. Vogtmann and A. Weinstein}, publisher={Springer-Verlag, New York}, date={1989}, pages={xvi+508}, isbn={0-387-96890-3}, review={\MR {997295}}, doi={10.1007/978-1-4757-2063-1}, }
Reference [8]
V. I. Arnold and B. A. Khesin, Topological methods in hydrodynamics, Applied Mathematical Sciences, vol. 125, Springer-Verlag, New York, 1998. Extended 2nd ed., Springer Nature Switzerland, 2021. MR1612569,
Show rawAMSref \bib{ArKh1998}{book}{ author={Arnold, Vladimir I.}, author={Khesin, Boris A.}, title={Topological methods in hydrodynamics}, series={Applied Mathematical Sciences}, volume={125}, note={Extended 2nd ed., Springer Nature Switzerland, 2021}, publisher={Springer-Verlag, New York}, date={1998}, pages={xvi+374}, isbn={0-387-94947-X}, review={\MR {1612569}}, }
Reference [9]
J.-D. Benamou and Y. Brenier, A computational fluid mechanics solution to the Monge-Kantorovich mass transfer problem, Numer. Math. 84 (2000), no. 3, 375–393, DOI 10.1007/s002110050002. MR1738163,
Show rawAMSref \bib{BeBr2000}{article}{ author={Benamou, Jean-David}, author={Brenier, Yann}, title={A computational fluid mechanics solution to the Monge-Kantorovich mass transfer problem}, journal={Numer. Math.}, volume={84}, date={2000}, number={3}, pages={375--393}, issn={0029-599X}, review={\MR {1738163}}, doi={10.1007/s002110050002}, }
Reference [10]
N. Besse and U. Frisch, Geometric formulation of the Cauchy invariants for incompressible Euler flow in flat and curved spaces, J. Fluid Mech. 825 (2017), 412–478, DOI 10.1017/jfm.2017.402. MR3692802,
Show rawAMSref \bib{BeFr2017}{article}{ author={Besse, Nicolas}, author={Frisch, Uriel}, title={Geometric formulation of the Cauchy invariants for incompressible Euler flow in flat and curved spaces}, journal={J. Fluid Mech.}, volume={825}, date={2017}, pages={412--478}, issn={0022-1120}, review={\MR {3692802}}, doi={10.1017/jfm.2017.402}, }
Reference [11]
Y. Brenier, The least action principle and the related concept of generalized flows for incompressible perfect fluids, J. Amer. Math. Soc. 2 (1989), no. 2, 225–255, DOI 10.2307/1990977. MR969419,
Show rawAMSref \bib{Br1989}{article}{ author={Brenier, Yann}, title={The least action principle and the related concept of generalized flows for incompressible perfect fluids}, journal={J. Amer. Math. Soc.}, volume={2}, date={1989}, number={2}, pages={225--255}, issn={0894-0347}, review={\MR {969419}}, doi={10.2307/1990977}, }
Reference [12]
Y. Brenier, Polar factorization and monotone rearrangement of vector-valued functions, Comm. Pure Appl. Math. 44 (1991), no. 4, 375–417, DOI 10.1002/cpa.3160440402. MR1100809,
Show rawAMSref \bib{Br1991}{article}{ author={Brenier, Yann}, title={Polar factorization and monotone rearrangement of vector-valued functions}, journal={Comm. Pure Appl. Math.}, volume={44}, date={1991}, number={4}, pages={375--417}, issn={0010-3640}, review={\MR {1100809}}, doi={10.1002/cpa.3160440402}, }
Reference [13]
Y. Brenier, Extended Monge-Kantorovich theory, Optimal transportation and applications (Martina Franca, 2001), Lecture Notes in Math., vol. 1813, Springer, Berlin, 2003, pp. 91–121, DOI 10.1007/978-3-540-44857-0_4. MR2006306,
Show rawAMSref \bib{Br2003c}{article}{ author={Brenier, Yann}, title={Extended Monge-Kantorovich theory}, conference={ title={Optimal transportation and applications}, address={Martina Franca}, date={2001}, }, book={ series={Lecture Notes in Math.}, volume={1813}, publisher={Springer, Berlin}, }, date={2003}, pages={91--121}, review={\MR {2006306}}, doi={10.1007/978-3-540-44857-0\_4}, }
Reference [14]
Th. Bröcker, Differentiable germs and catastrophes, Cambridge University Press, Cambridge-New York-Melbourne, 1975. Translated from the German, last chapter and bibliography by L. Lander; London Mathematical Society Lecture Note Series, No. 17. MR0494220,
Show rawAMSref \bib{BrLa1978}{book}{ author={Br\"{o}cker, Th.}, title={Differentiable germs and catastrophes}, note={Translated from the German, last chapter and bibliography by L. Lander; London Mathematical Society Lecture Note Series, No. 17}, publisher={Cambridge University Press, Cambridge-New York-Melbourne}, date={1975}, pages={vi+179}, review={\MR {0494220}}, }
Reference [15]
A. Chern, F. Knöppel, U. Pinkall, P. Schröder, and S. Weissmann, Schrödinger’s smoke, ACM Trans. Graph. 35 (2016), 77:1–77:13.
Reference [16]
C. De Lellis and L. Székelyhidi Jr., High dimensionality and h-principle in PDE, Bull. Amer. Math. Soc. (N.S.) 54 (2017), no. 2, 247–282, DOI 10.1090/bull/1549. MR3619726,
Show rawAMSref \bib{LeSz2017}{article}{ author={De Lellis, Camillo}, author={Sz\'{e}kelyhidi, L\'{a}szl\'{o}, Jr.}, title={High dimensionality and h-principle in PDE}, journal={Bull. Amer. Math. Soc. (N.S.)}, volume={54}, date={2017}, number={2}, pages={247--282}, issn={0273-0979}, review={\MR {3619726}}, doi={10.1090/bull/1549}, }
Reference [17]
H.-D. Doebner and G. A. Goldin, On a general nonlinear Schrödinger equation admitting diffusion currents, Phys. Lett. A 162 (1992), no. 5, 397–401, DOI 10.1016/0375-9601(92)90061-P. MR1152014,
Show rawAMSref \bib{DoGo1992}{article}{ author={Doebner, H.-D.}, author={Goldin, Gerald A.}, title={On a general nonlinear Schr\"{o}dinger equation admitting diffusion currents}, journal={Phys. Lett. A}, volume={162}, date={1992}, number={5}, pages={397--401}, issn={0375-9601}, review={\MR {1152014}}, doi={10.1016/0375-9601(92)90061-P}, }
Reference [18]
F. V. Dolzhansky, Fundamentals of geophysical hydrodynamics, Encyclopaedia of Mathematical Sciences, vol. 103, Springer, Heidelberg, 2013. Translated from the 2011 Russian 2nd edition by Boris Khesin; Mathematical Physics, 4, DOI 10.1007/978-3-642-31034-8. MR2987817,
Show rawAMSref \bib{Do2013}{book}{ author={Dolzhansky, Felix V.}, title={Fundamentals of geophysical hydrodynamics}, series={Encyclopaedia of Mathematical Sciences}, volume={103}, note={Translated from the 2011 Russian 2nd edition by Boris Khesin; Mathematical Physics, 4}, publisher={Springer, Heidelberg}, date={2013}, pages={xiv+272}, isbn={978-3-642-31033-1}, isbn={978-3-642-31034-8}, review={\MR {2987817}}, doi={10.1007/978-3-642-31034-8}, }
Reference [19]
D. G. Ebin, Motion of a slightly compressible fluid, Proc. Nat. Acad. Sci. U.S.A. 72 (1975), 539–542, DOI 10.1073/pnas.72.2.539. MR397780,
Show rawAMSref \bib{Eb1975}{article}{ author={Ebin, David G.}, title={Motion of a slightly compressible fluid}, journal={Proc. Nat. Acad. Sci. U.S.A.}, volume={72}, date={1975}, pages={539--542}, issn={0027-8424}, review={\MR {397780}}, doi={10.1073/pnas.72.2.539}, }
Reference [20]
D. G. Ebin and J. Marsden, Groups of diffeomorphisms and the motion of an incompressible fluid, Ann. of Math. (2) 92 (1970), 102–163, DOI 10.2307/1970699. MR271984,
Show rawAMSref \bib{EbMa1970}{article}{ author={Ebin, David G.}, author={Marsden, Jerrold}, title={Groups of diffeomorphisms and the motion of an incompressible fluid}, journal={Ann. of Math. (2)}, volume={92}, date={1970}, pages={102--163}, issn={0003-486X}, review={\MR {271984}}, doi={10.2307/1970699}, }
Reference [21]
D. G. Ebin, G. Misiołek, and S. C. Preston, Singularities of the exponential map on the volume-preserving diffeomorphism group, Geom. Funct. Anal. 16 (2006), no. 4, 850–868, DOI 10.1007/s00039-006-0573-8. MR2255383,
Show rawAMSref \bib{EbMiPr2006}{article}{ author={Ebin, D. G.}, author={Misio\l ek, G.}, author={Preston, S. C.}, title={Singularities of the exponential map on the volume-preserving diffeomorphism group}, journal={Geom. Funct. Anal.}, volume={16}, date={2006}, number={4}, pages={850--868}, issn={1016-443X}, review={\MR {2255383}}, doi={10.1007/s00039-006-0573-8}, }
Reference [22]
A. Enciso, D. Peralta-Salas, and F. Torres de Lizaur, Helicity is the only integral invariant of volume-preserving transformations, Proc. Natl. Acad. Sci. USA 113 (2016), no. 8, 2035–2040, DOI 10.1073/pnas.1516213113. MR3474089,
Show rawAMSref \bib{EnPeLi2016}{article}{ author={Enciso, Alberto}, author={Peralta-Salas, Daniel}, author={Torres de Lizaur, Francisco}, title={Helicity is the only integral invariant of volume-preserving transformations}, journal={Proc. Natl. Acad. Sci. USA}, volume={113}, date={2016}, number={8}, pages={2035--2040}, issn={0027-8424}, review={\MR {3474089}}, doi={10.1073/pnas.1516213113}, }
Reference [23]
S. Friedlander and A. Shnirelman, Instability of steady flows of an ideal incompressible fluid, Mathematical fluid mechanics, Adv. Math. Fluid Mech., Birkhäuser, Basel, 2001, pp. 143–172. MR1865052,
Show rawAMSref \bib{FrSh2001}{article}{ author={Friedlander, Susan}, author={Shnirelman, Alexander}, title={Instability of steady flows of an ideal incompressible fluid}, conference={ title={Mathematical fluid mechanics}, }, book={ series={Adv. Math. Fluid Mech.}, publisher={Birkh\"{a}user, Basel}, }, date={2001}, pages={143--172}, review={\MR {1865052}}, }
Reference [24]
S. Friedlander, W. Strauss, and M. Vishik, Nonlinear instability in an ideal fluid (English, with English and French summaries), Ann. Inst. H. Poincaré Anal. Non Linéaire 14 (1997), no. 2, 187–209, DOI 10.1016/S0294-1449(97)80144-8. MR1441392,
Show rawAMSref \bib{FrStVi1997}{article}{ author={Friedlander, Susan}, author={Strauss, Walter}, author={Vishik, Misha}, title={Nonlinear instability in an ideal fluid}, language={English, with English and French summaries}, journal={Ann. Inst. H. Poincar\'{e} Anal. Non Lin\'{e}aire}, volume={14}, date={1997}, number={2}, pages={187--209}, issn={0294-1449}, review={\MR {1441392}}, doi={10.1016/S0294-1449(97)80144-8}, }
Reference [25]
T. Friedrich, Die Fisher-Information und symplektische Strukturen (German), Math. Nachr. 153 (1991), 273–296, DOI 10.1002/mana.19911530125. MR1131949,
Show rawAMSref \bib{Fr1991}{article}{ author={Friedrich, Thomas}, title={Die Fisher-Information und symplektische Strukturen}, language={German}, journal={Math. Nachr.}, volume={153}, date={1991}, pages={273--296}, issn={0025-584X}, review={\MR {1131949}}, doi={10.1002/mana.19911530125}, }
Reference [26]
F. Gay-Balmaz and C. Tronci, Madelung transform and probability densities in hybrid quantum-classical dynamics, Nonlinearity 33 (2020), no. 10, 5383–5424, DOI 10.1088/1361-6544/aba233. MR4151413,
Show rawAMSref \bib{GaTr2020}{article}{ author={Gay-Balmaz, Fran\c {c}ois}, author={Tronci, Cesare}, title={Madelung transform and probability densities in hybrid quantum-classical dynamics}, journal={Nonlinearity}, volume={33}, date={2020}, number={10}, pages={5383--5424}, issn={0951-7715}, review={\MR {4151413}}, doi={10.1088/1361-6544/aba233}, }
Reference [27]
G. A. Goldin and D. H. Sharp, Diffeomorphism groups and local symmetries: some applications in quantum physics, Symmetries in science, III (Vorarlberg, 1988), Plenum, New York, 1989, pp. 181–205. MR1120322,
Show rawAMSref \bib{GoSh1989}{article}{ author={Goldin, Gerald A.}, author={Sharp, David H.}, title={Diffeomorphism groups and local symmetries: some applications in quantum physics}, conference={ title={Symmetries in science, III}, address={Vorarlberg}, date={1988}, }, book={ publisher={Plenum, New York}, }, date={1989}, pages={181--205}, review={\MR {1120322}}, }
Reference [28]
D. A. Gomes, A variational formulation for the Navier-Stokes equation, Comm. Math. Phys. 257 (2005), no. 1, 227–234, DOI 10.1007/s00220-004-1263-8. MR2163576,
Show rawAMSref \bib{Go2005}{article}{ author={Gomes, Diogo Aguiar}, title={A variational formulation for the Navier-Stokes equation}, journal={Comm. Math. Phys.}, volume={257}, date={2005}, number={1}, pages={227--234}, issn={0010-3616}, review={\MR {2163576}}, doi={10.1007/s00220-004-1263-8}, }
Reference [29]
G. Gui, Y. Liu, and M. Zhu, On the wave-breaking phenomena and global existence for the generalized periodic Camassa-Holm equation, Int. Math. Res. Not. IMRN 21 (2012), 4858–4903, DOI 10.1093/imrn/rnr214. MR2993438,
Show rawAMSref \bib{GuLiZh2011}{article}{ author={Gui, Guilong}, author={Liu, Yue}, author={Zhu, Min}, title={On the wave-breaking phenomena and global existence for the generalized periodic Camassa-Holm equation}, journal={Int. Math. Res. Not. IMRN}, date={2012}, number={21}, pages={4858--4903}, issn={1073-7928}, review={\MR {2993438}}, doi={10.1093/imrn/rnr214}, }
Reference [30]
R. S. Hamilton, The inverse function theorem of Nash and Moser, Bull. Amer. Math. Soc. (N.S.) 7 (1982), no. 1, 65–222, DOI 10.1090/S0273-0979-1982-15004-2. MR656198,
Show rawAMSref \bib{Ha1982}{article}{ author={Hamilton, Richard S.}, title={The inverse function theorem of Nash and Moser}, journal={Bull. Amer. Math. Soc. (N.S.)}, volume={7}, date={1982}, number={1}, pages={65--222}, issn={0273-0979}, review={\MR {656198}}, doi={10.1090/S0273-0979-1982-15004-2}, }
Reference [31]
D. D. Holm, Variational principles for stochastic fluid dynamics, Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci. 471 (2015), no. 2176, 20140963, 19, DOI 10.1098/rspa.2014.0963. MR3325187,
Show rawAMSref \bib{Ho2015}{article}{ author={Holm, Darryl D.}, title={Variational principles for stochastic fluid dynamics}, journal={Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci.}, volume={471}, date={2015}, number={2176}, pages={20140963, 19}, issn={1364-5021}, review={\MR {3325187}}, doi={10.1098/rspa.2014.0963}, }
Reference [32]
D. D. Holm and B. A. Kupershmidt, Relativistic fluid dynamics as a Hamiltonian system, Phys. Lett. A 101 (1984), no. 1, 23–26, DOI 10.1016/0375-9601(84)90083-5. MR735726,
Show rawAMSref \bib{HoKu1984}{article}{ author={Holm, Darryl D.}, author={Kupershmidt, Boris A.}, title={Relativistic fluid dynamics as a Hamiltonian system}, journal={Phys. Lett. A}, volume={101}, date={1984}, number={1}, pages={23--26}, issn={0375-9601}, review={\MR {735726}}, doi={10.1016/0375-9601(84)90083-5}, }
Reference [33]
D. D. Holm, J. E. Marsden, and T. S. Ratiu, The Euler-Poincaré equations and semidirect products with applications to continuum theories, Adv. Math. 137 (1998), no. 1, 1–81, DOI 10.1006/aima.1998.1721. MR1627802,
Show rawAMSref \bib{HoMaRa1998}{article}{ author={Holm, Darryl D.}, author={Marsden, Jerrold E.}, author={Ratiu, Tudor S.}, title={The Euler-Poincar\'{e} equations and semidirect products with applications to continuum theories}, journal={Adv. Math.}, volume={137}, date={1998}, number={1}, pages={1--81}, issn={0001-8708}, review={\MR {1627802}}, doi={10.1006/aima.1998.1721}, }
Reference [34]
D. D. Holm, J. E. Marsden, T. Ratiu, and A. Weinstein, Nonlinear stability of fluid and plasma equilibria, Phys. Rep. 123 (1985), no. 1-2, 116, DOI 10.1016/0370-1573(85)90028-6. MR794110,
Show rawAMSref \bib{HoMaRaWe1985}{article}{ author={Holm, Darryl D.}, author={Marsden, Jerrold E.}, author={Ratiu, Tudor}, author={Weinstein, Alan}, title={Nonlinear stability of fluid and plasma equilibria}, journal={Phys. Rep.}, volume={123}, date={1985}, number={1-2}, pages={116}, issn={0370-1573}, review={\MR {794110}}, doi={10.1016/0370-1573(85)90028-6}, }
Reference [35]
A. Izosimov, B. Khesin, and M. Mousavi, Coadjoint orbits of symplectic diffeomorphisms of surfaces and ideal hydrodynamics (English, with English and French summaries), Ann. Inst. Fourier (Grenoble) 66 (2016), no. 6, 2385–2433. MR3580175,
Show rawAMSref \bib{IzKhMo2016}{article}{ author={Izosimov, Anton}, author={Khesin, Boris}, author={Mousavi, Mehdi}, title={Coadjoint orbits of symplectic diffeomorphisms of surfaces and ideal hydrodynamics}, language={English, with English and French summaries}, journal={Ann. Inst. Fourier (Grenoble)}, volume={66}, date={2016}, number={6}, pages={2385--2433}, issn={0373-0956}, review={\MR {3580175}}, }
Reference [36]
B. Khesin and P. Lee, A nonholonomic Moser theorem and optimal transport, J. Symplectic Geom. 7 (2009), no. 4, 381–414. MR2551999,
Show rawAMSref \bib{KhLe2009}{article}{ author={Khesin, Boris}, author={Lee, Paul}, title={A nonholonomic Moser theorem and optimal transport}, journal={J. Symplectic Geom.}, volume={7}, date={2009}, number={4}, pages={381--414}, issn={1527-5256}, review={\MR {2551999}}, }
Reference [37]
B. Khesin, J. Lenells, and G. Misiołek, Generalized Hunter-Saxton equation and the geometry of the group of circle diffeomorphisms, Math. Ann. 342 (2008), no. 3, 617–656, DOI 10.1007/s00208-008-0250-3. MR2430993,
Show rawAMSref \bib{KhLeMi2008}{article}{ author={Khesin, Boris}, author={Lenells, Jonatan}, author={Misio\l ek, Gerard}, title={Generalized Hunter-Saxton equation and the geometry of the group of circle diffeomorphisms}, journal={Math. Ann.}, volume={342}, date={2008}, number={3}, pages={617--656}, issn={0025-5831}, review={\MR {2430993}}, doi={10.1007/s00208-008-0250-3}, }
Reference [38]
B. Khesin, J. Lenells, G. Misiołek, and S. C. Preston, Geometry of diffeomorphism groups, complete integrability and geometric statistics, Geom. Funct. Anal. 23 (2013), no. 1, 334–366, DOI 10.1007/s00039-013-0210-2. MR3037902,
Show rawAMSref \bib{KhLeMiPr2013}{article}{ author={Khesin, B.}, author={Lenells, J.}, author={Misio\l ek, G.}, author={Preston, S. C.}, title={Geometry of diffeomorphism groups, complete integrability and geometric statistics}, journal={Geom. Funct. Anal.}, volume={23}, date={2013}, number={1}, pages={334--366}, issn={1016-443X}, review={\MR {3037902}}, doi={10.1007/s00039-013-0210-2}, }
Reference [39]
B. Khesin, G. Misiolek, and K. Modin, Geometric hydrodynamics via Madelung transform, Proc. Natl. Acad. Sci. USA 115 (2018), no. 24, 6165–6170, DOI 10.1073/pnas.1719346115. MR3829316,
Show rawAMSref \bib{KhMiMo2018}{article}{ author={Khesin, Boris}, author={Misiolek, Gerard}, author={Modin, Klas}, title={Geometric hydrodynamics via Madelung transform}, journal={Proc. Natl. Acad. Sci. USA}, volume={115}, date={2018}, number={24}, pages={6165--6170}, issn={0027-8424}, review={\MR {3829316}}, doi={10.1073/pnas.1719346115}, }
Reference [40]
B. Khesin, G. Misiołek, and K. Modin, Geometry of the Madelung transform, Arch. Ration. Mech. Anal. 234 (2019), no. 2, 549–573, DOI 10.1007/s00205-019-01397-2. MR3995046,
Show rawAMSref \bib{KhMiMo2019}{article}{ author={Khesin, Boris}, author={Misio\l ek, Gerard}, author={Modin, Klas}, title={Geometry of the Madelung transform}, journal={Arch. Ration. Mech. Anal.}, volume={234}, date={2019}, number={2}, pages={549--573}, issn={0003-9527}, review={\MR {3995046}}, doi={10.1007/s00205-019-01397-2}, }
Reference [41]
T. W. B. Kibble, Geometrization of quantum mechanics, Comm. Math. Phys. 65 (1979), no. 2, 189–201. MR528191,
Show rawAMSref \bib{Ki1979}{article}{ author={Kibble, T. W. B.}, title={Geometrization of quantum mechanics}, journal={Comm. Math. Phys.}, volume={65}, date={1979}, number={2}, pages={189--201}, issn={0010-3616}, review={\MR {528191}}, }
Reference [42]
L. D. Landau and E. M. Lifshitz, Fluid mechanics, Translated from the Russian by J. B. Sykes and W. H. Reid. Course of Theoretical Physics, Vol. 6, Pergamon Press, London-Paris-Frankfurt; Addison-Wesley Publishing Co., Inc., Reading, Mass., 1959. MR0108121,
Show rawAMSref \bib{LaLi1959}{book}{ author={Landau, L. D.}, author={Lifshitz, E. M.}, title={Fluid mechanics}, series={Translated from the Russian by J. B. Sykes and W. H. Reid. Course of Theoretical Physics, Vol. 6}, publisher={Pergamon Press, London-Paris-Frankfurt; Addison-Wesley Publishing Co., Inc., Reading, Mass.}, date={1959}, pages={xii+536}, review={\MR {0108121}}, }
Reference [43]
F. Léger and W. Li, Hopf-Cole transformation via generalized Schrödinger bridge problem, J. Differential Equations 274 (2021), 788–827, DOI 10.1016/j.jde.2020.10.029. MR4188994,
Show rawAMSref \bib{LeLi2019}{article}{ author={L\'{e}ger, Flavien}, author={Li, Wuchen}, title={Hopf-Cole transformation via generalized Schr\"{o}dinger bridge problem}, journal={J. Differential Equations}, volume={274}, date={2021}, pages={788--827}, issn={0022-0396}, review={\MR {4188994}}, doi={10.1016/j.jde.2020.10.029}, }
Reference [44]
J. Lenells, Spheres, Kähler geometry and the Hunter-Saxton system, Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci. 469 (2013), no. 2154, 20120726, 19, DOI 10.1098/rspa.2012.0726. MR3061310,
Show rawAMSref \bib{Le2013b}{article}{ author={Lenells, Jonatan}, title={Spheres, K\"{a}hler geometry and the Hunter-Saxton system}, journal={Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci.}, volume={469}, date={2013}, number={2154}, pages={20120726, 19}, issn={1364-5021}, review={\MR {3061310}}, doi={10.1098/rspa.2012.0726}, }
Reference [45]
C. Léonard, A survey of the Schrödinger problem and some of its connections with optimal transport, Discrete Contin. Dyn. Syst. 34 (2014), no. 4, 1533–1574, DOI 10.3934/dcds.2014.34.1533. MR3121631,
Show rawAMSref \bib{Le2014}{article}{ author={L\'{e}onard, Christian}, title={A survey of the Schr\"{o}dinger problem and some of its connections with optimal transport}, journal={Discrete Contin. Dyn. Syst.}, volume={34}, date={2014}, number={4}, pages={1533--1574}, issn={1078-0947}, review={\MR {3121631}}, doi={10.3934/dcds.2014.34.1533}, }
Reference [46]
S. Lie, Theorie der Transformationsgruppen: Abschnitt 2, Technische Informationsbibliothek (TIB), 1890.
Reference [47]
J. Lott, Some geometric calculations on Wasserstein space, Comm. Math. Phys. 277 (2008), no. 2, 423–437, DOI 10.1007/s00220-007-0367-3. MR2358290,
Show rawAMSref \bib{Lo2008}{article}{ author={Lott, John}, title={Some geometric calculations on Wasserstein space}, journal={Comm. Math. Phys.}, volume={277}, date={2008}, number={2}, pages={423--437}, issn={0010-3616}, review={\MR {2358290}}, doi={10.1007/s00220-007-0367-3}, }
Reference [48]
E. Madelung, Quantentheorie in hydrodynamischer form, Zeitschr. Phys. 40 (1927), 322–326.
Reference [49]
E. Madelung, Die mathematischen Hilfsmittel des Physikers (German), Siebente Auflage. Die Grundlehren der mathematischen Wissenschaften, Band 4, Springer-Verlag, Berlin, 1964. MR0164476,
Show rawAMSref \bib{Ma1964}{book}{ author={Madelung, Erwin}, title={Die mathematischen Hilfsmittel des Physikers}, language={German}, series={Siebente Auflage. Die Grundlehren der mathematischen Wissenschaften, Band 4}, publisher={Springer-Verlag, Berlin}, date={1964}, pages={xx+536}, review={\MR {0164476}}, }
Reference [50]
J. E. Marsden, G. Misiołek, J.-P. Ortega, M. Perlmutter, and T. S. Ratiu, Hamiltonian reduction by stages, Lecture Notes in Mathematics, vol. 1913, Springer, Berlin, 2007. MR2337886,
Show rawAMSref \bib{MaMiOrPeRa2007}{book}{ author={Marsden, Jerrold E.}, author={Misio\l ek, Gerard}, author={Ortega, Juan-Pablo}, author={Perlmutter, Matthew}, author={Ratiu, Tudor S.}, title={Hamiltonian reduction by stages}, series={Lecture Notes in Mathematics}, volume={1913}, publisher={Springer, Berlin}, date={2007}, pages={xvi+519}, isbn={978-3-540-72469-8}, review={\MR {2337886}}, }
Reference [51]
J. E. Marsden and T. S. Ratiu, Introduction to mechanics and symmetry: A basic exposition of classical mechanical systems, 2nd ed., Texts in Applied Mathematics, vol. 17, Springer-Verlag, New York, 1999, DOI 10.1007/978-0-387-21792-5. MR1723696,
Show rawAMSref \bib{MaRa1999}{book}{ author={Marsden, Jerrold E.}, author={Ratiu, Tudor S.}, title={Introduction to mechanics and symmetry}, series={Texts in Applied Mathematics}, volume={17}, edition={2}, subtitle={A basic exposition of classical mechanical systems}, publisher={Springer-Verlag, New York}, date={1999}, pages={xviii+582}, isbn={0-387-98643-X}, review={\MR {1723696}}, doi={10.1007/978-0-387-21792-5}, }
Reference [52]
J. E. Marsden, T. Ratiu, and A. Weinstein, Reduction and Hamiltonian structures on duals of semidirect product Lie algebras, Fluids and plasmas: geometry and dynamics (Boulder, Colo., 1983), Contemp. Math., vol. 28, Amer. Math. Soc., Providence, RI, 1984, pp. 55–100, DOI 10.1090/conm/028/751975. MR751975,
Show rawAMSref \bib{MaRaWe1984b}{article}{ author={Marsden, Jerrold E.}, author={Ratiu, Tudor}, author={Weinstein, Alan}, title={Reduction and Hamiltonian structures on duals of semidirect product Lie algebras}, conference={ title={Fluids and plasmas: geometry and dynamics}, address={Boulder, Colo.}, date={1983}, }, book={ series={Contemp. Math.}, volume={28}, publisher={Amer. Math. Soc., Providence, RI}, }, date={1984}, pages={55--100}, review={\MR {751975}}, doi={10.1090/conm/028/751975}, }
Reference [53]
J. Marsden and A. Weinstein, Reduction of symplectic manifolds with symmetry, Rep. Mathematical Phys. 5 (1974), no. 1, 121–130, DOI 10.1016/0034-4877(74)90021-4. MR402819,
Show rawAMSref \bib{MaWe1974}{article}{ author={Marsden, Jerrold}, author={Weinstein, Alan}, title={Reduction of symplectic manifolds with symmetry}, journal={Rep. Mathematical Phys.}, volume={5}, date={1974}, number={1}, pages={121--130}, issn={0034-4877}, review={\MR {402819}}, doi={10.1016/0034-4877(74)90021-4}, }
Reference [54]
J. Marsden and A. Weinstein, Coadjoint orbits, vortices, and Clebsch variables for incompressible fluids, Phys. D 7 (1983), no. 1-3, 305–323, DOI 10.1016/0167-2789(83)90134-3. Order in chaos (Los Alamos, N.M., 1982). MR719058,
Show rawAMSref \bib{MaWe1983}{article}{ author={Marsden, Jerrold}, author={Weinstein, Alan}, title={Coadjoint orbits, vortices, and Clebsch variables for incompressible fluids}, note={Order in chaos (Los Alamos, N.M., 1982)}, journal={Phys. D}, volume={7}, date={1983}, number={1-3}, pages={305--323}, issn={0167-2789}, review={\MR {719058}}, doi={10.1016/0167-2789(83)90134-3}, }
Reference [55]
R. J. McCann, Polar factorization of maps on Riemannian manifolds, Geom. Funct. Anal. 11 (2001), no. 3, 589–608, DOI 10.1007/PL00001679. MR1844080,
Show rawAMSref \bib{Mc2001}{article}{ author={McCann, Robert J.}, title={Polar factorization of maps on Riemannian manifolds}, journal={Geom. Funct. Anal.}, volume={11}, date={2001}, number={3}, pages={589--608}, issn={1016-443X}, review={\MR {1844080}}, doi={10.1007/PL00001679}, }
Reference [56]
D. Mumford and P. W. Michor, On Euler’s equation and ‘EPDiff’, J. Geom. Mech. 5 (2013), no. 3, 319–344, DOI 10.3934/jgm.2013.5.319. MR3110104,
Show rawAMSref \bib{MiMu2013}{article}{ author={Mumford, David}, author={Michor, Peter W.}, title={On Euler's equation and `EPDiff'}, journal={J. Geom. Mech.}, volume={5}, date={2013}, number={3}, pages={319--344}, issn={1941-4889}, review={\MR {3110104}}, doi={10.3934/jgm.2013.5.319}, }
Reference [57]
G. Misiołek, Conjugate points in , Proc. Amer. Math. Soc. (1996), 977–982.
Reference [58]
K. Modin, Generalized Hunter-Saxton equations, optimal information transport, and factorization of diffeomorphisms, J. Geom. Anal. 25 (2015), no. 2, 1306–1334, DOI 10.1007/s12220-014-9469-2. MR3319973,
Show rawAMSref \bib{Mo2015}{article}{ author={Modin, Klas}, title={Generalized Hunter-Saxton equations, optimal information transport, and factorization of diffeomorphisms}, journal={J. Geom. Anal.}, volume={25}, date={2015}, number={2}, pages={1306--1334}, issn={1050-6926}, review={\MR {3319973}}, doi={10.1007/s12220-014-9469-2}, }
Reference [59]
K. Modin, Geometry of matrix decompositions seen through optimal transport and information geometry, J. Geom. Mech. 9 (2017), no. 3, 335–390, DOI 10.3934/jgm.2017014. MR3666687,
Show rawAMSref \bib{Mo2017}{article}{ author={Modin, Klas}, title={Geometry of matrix decompositions seen through optimal transport and information geometry}, journal={J. Geom. Mech.}, volume={9}, date={2017}, number={3}, pages={335--390}, issn={1941-4889}, review={\MR {3666687}}, doi={10.3934/jgm.2017014}, }
Reference [60]
M. Molitor, On the relation between geometrical quantum mechanics and information geometry, J. Geom. Mech. 7 (2015), no. 2, 169–202, DOI 10.3934/jgm.2015.7.169. MR3356591,
Show rawAMSref \bib{Mo2015b}{article}{ author={Molitor, Mathieu}, title={On the relation between geometrical quantum mechanics and information geometry}, journal={J. Geom. Mech.}, volume={7}, date={2015}, number={2}, pages={169--202}, issn={1941-4889}, review={\MR {3356591}}, doi={10.3934/jgm.2015.7.169}, }
Reference [61]
J. Moser, On the volume elements on a manifold, Trans. Amer. Math. Soc. 120 (1965), 286–294, DOI 10.2307/1994022. MR182927,
Show rawAMSref \bib{Mo1965}{article}{ author={Moser, J\"{u}rgen}, title={On the volume elements on a manifold}, journal={Trans. Amer. Math. Soc.}, volume={120}, date={1965}, pages={286--294}, issn={0002-9947}, review={\MR {182927}}, doi={10.2307/1994022}, }
Reference [62]
J. Moser, Integrable Hamiltonian systems and spectral theory, Lezioni Fermiane. [Fermi Lectures], Scuola Normale Superiore, Pisa, 1983. MR854865,
Show rawAMSref \bib{Mo1983}{book}{ author={Moser, J.}, title={Integrable Hamiltonian systems and spectral theory}, series={Lezioni Fermiane. [Fermi Lectures]}, publisher={Scuola Normale Superiore, Pisa}, date={1983}, pages={iv+85}, review={\MR {854865}}, }
Reference [63]
C. Neumann, De problemate quodam mechanico, quod ad primam integralium ultraellipticorum classem revocatur: Dissertatio inauguralis, Dalkowski, 1856.
Reference [64]
F. Otto, The geometry of dissipative evolution equations: the porous medium equation, Comm. Partial Differential Equations 26 (2001), no. 1-2, 101–174, DOI 10.1081/PDE-100002243. MR1842429,
Show rawAMSref \bib{Ot2001}{article}{ author={Otto, Felix}, title={The geometry of dissipative evolution equations: the porous medium equation}, journal={Comm. Partial Differential Equations}, volume={26}, date={2001}, number={1-2}, pages={101--174}, issn={0360-5302}, review={\MR {1842429}}, doi={10.1081/PDE-100002243}, }
Reference [65]
V. Y. Ovsienko, B. A. Khesin, and Y. V. Chekanov, Integrals of the Euler equations of multidimensional hydrodynamics and superconductivity, J. Sov. Math. 59 (1992), 1096–1101. MR1015703
Reference [66]
H. Poincaré, Sur une forme nouvelle des équations de la mécanique, C.R. Acad. Sci. 132 (1901), 369–371.
Reference [67]
S. C. Preston, The geometry of barotropic flow, J. Math. Fluid Mech. 15 (2013), no. 4, 807–821, DOI 10.1007/s00021-013-0142-5. MR3127021,
Show rawAMSref \bib{Pr2013}{article}{ author={Preston, Stephen C.}, title={The geometry of barotropic flow}, journal={J. Math. Fluid Mech.}, volume={15}, date={2013}, number={4}, pages={807--821}, issn={1422-6928}, review={\MR {3127021}}, doi={10.1007/s00021-013-0142-5}, }
Reference [68]
C. Qu, Y. Zhang, X. Liu, and Y. Liu, Orbital stability of periodic peakons to a generalized -Camassa-Holm equation, Arch. Ration. Mech. Anal. 211 (2014), no. 2, 593–617, DOI 10.1007/s00205-013-0672-2. MR3149067,
Show rawAMSref \bib{QuZhLiLi2014}{article}{ author={Qu, Changzheng}, author={Zhang, Ying}, author={Liu, Xiaochuan}, author={Liu, Yue}, title={Orbital stability of periodic peakons to a generalized $\mu $-Camassa-Holm equation}, journal={Arch. Ration. Mech. Anal.}, volume={211}, date={2014}, number={2}, pages={593--617}, issn={0003-9527}, review={\MR {3149067}}, doi={10.1007/s00205-013-0672-2}, }
Reference [69]
V. Scheffer, An inviscid flow with compact support in space-time, J. Geom. Anal. 3 (1993), no. 4, 343–401, DOI 10.1007/BF02921318. MR1231007,
Show rawAMSref \bib{Sc1993}{article}{ author={Scheffer, Vladimir}, title={An inviscid flow with compact support in space-time}, journal={J. Geom. Anal.}, volume={3}, date={1993}, number={4}, pages={343--401}, issn={1050-6926}, review={\MR {1231007}}, doi={10.1007/BF02921318}, }
Reference [70]
A. I. Shnirelman, The geometry of the group of diffeomorphisms and the dynamics of an ideal incompressible fluid (Russian), Mat. Sb. (N.S.) 128(170) (1985), no. 1, 82–109, 144. MR805697,
Show rawAMSref \bib{Sh1985}{article}{ author={Shnirelman, A. I.}, title={The geometry of the group of diffeomorphisms and the dynamics of an ideal incompressible fluid}, language={Russian}, journal={Mat. Sb. (N.S.)}, volume={128(170)}, date={1985}, number={1}, pages={82--109, 144}, issn={0368-8666}, review={\MR {805697}}, }
Reference [71]
A. Shnirelman, On the non-uniqueness of weak solution of the Euler equations, Journées “Équations aux Dérivées Partielles” (Saint-Jean-de-Monts, 1996), École Polytech., Palaiseau, 1996, pp. Exp. No. XVIII, 10. MR1417742,
Show rawAMSref \bib{Sh1996}{article}{ author={Shnirelman, A.}, title={On the non-uniqueness of weak solution of the Euler equations}, conference={ title={Journ\'{e}es ``\'{E}quations aux D\'{e}riv\'{e}es Partielles''}, address={Saint-Jean-de-Monts}, date={1996}, }, book={ publisher={\'{E}cole Polytech., Palaiseau}, }, date={1996}, pages={Exp. No. XVIII, 10}, review={\MR {1417742}}, }
Reference [72]
N. K. Smolencev, The Maupertuis principle (Russian), Sibirsk. Mat. Zh. 20 (1979), no. 5, 1092–1098, 1167. MR559071,
Show rawAMSref \bib{Sm1979}{article}{ author={Smolencev, N. K.}, title={The Maupertuis principle}, language={Russian}, journal={Sibirsk. Mat. Zh.}, volume={20}, date={1979}, number={5}, pages={1092--1098, 1167}, issn={0037-4474}, review={\MR {559071}}, }
Reference [73]
N. K. Smolentsev, Diffeomorphism groups of compact manifolds (Russian, with Russian summary), Sovrem. Mat. Prilozh. 37, Geometriya (2006), 3–100, DOI 10.1007/s10958-007-0471-0; English transl., J. Math. Sci. (N.Y.) 146 (2007), no. 6, 6213–6312. MR2568571,
Show rawAMSref \bib{Sm2007}{article}{ author={Smolentsev, N. K.}, title={Diffeomorphism groups of compact manifolds}, language={Russian, with Russian summary}, journal={Sovrem. Mat. Prilozh.}, date={2006}, number={37, Geometriya}, pages={3--100}, issn={1512-1712}, translation={ journal={J. Math. Sci. (N.Y.)}, volume={146}, date={2007}, number={6}, pages={6213--6312}, issn={1072-3374}, }, review={\MR {2568571}}, doi={10.1007/s10958-007-0471-0}, }
Reference [74]
D. Tataru, The wave maps equation, Bull. Amer. Math. Soc. (N.S.) 41 (2004), no. 2, 185–204, DOI 10.1090/S0273-0979-04-01005-5. MR2043751,
Show rawAMSref \bib{Ta2004}{article}{ author={Tataru, Daniel}, title={The wave maps equation}, journal={Bull. Amer. Math. Soc. (N.S.)}, volume={41}, date={2004}, number={2}, pages={185--204}, issn={0273-0979}, review={\MR {2043751}}, doi={10.1090/S0273-0979-04-01005-5}, }
Reference [75]
F. Tiğlay and C. Vizman, Generalized Euler-Poincaré equations on Lie groups and homogeneous spaces, orbit invariants and applications, Lett. Math. Phys. 97 (2011), no. 1, 45–60, DOI 10.1007/s11005-011-0464-2. MR2802316,
Show rawAMSref \bib{TiVi2011}{article}{ author={Ti\u {g}lay, Feride}, author={Vizman, Cornelia}, title={Generalized Euler-Poincar\'{e} equations on Lie groups and homogeneous spaces, orbit invariants and applications}, journal={Lett. Math. Phys.}, volume={97}, date={2011}, number={1}, pages={45--60}, issn={0377-9017}, review={\MR {2802316}}, doi={10.1007/s11005-011-0464-2}, }
Reference [76]
B. Tsirelson, Scaling limit, noise, stability, Lectures on probability theory and statistics, Lecture Notes in Math., vol. 1840, Springer, Berlin, 2004, pp. 1–106, DOI 10.1007/978-3-540-39982-7_1. MR2079671,
Show rawAMSref \bib{Ts2004}{article}{ author={Tsirelson, Boris}, title={Scaling limit, noise, stability}, conference={ title={Lectures on probability theory and statistics}, }, book={ series={Lecture Notes in Math.}, volume={1840}, publisher={Springer, Berlin}, }, date={2004}, pages={1--106}, review={\MR {2079671}}, doi={10.1007/978-3-540-39982-7\_1}, }
Reference [77]
C. Villani, Optimal transport: Old and new, Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 338, Springer-Verlag, Berlin, 2009, DOI 10.1007/978-3-540-71050-9. MR2459454,
Show rawAMSref \bib{Vi2009}{book}{ author={Villani, C\'{e}dric}, title={Optimal transport}, series={Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]}, volume={338}, subtitle={Old and new}, publisher={Springer-Verlag, Berlin}, date={2009}, pages={xxii+973}, isbn={978-3-540-71049-3}, review={\MR {2459454}}, doi={10.1007/978-3-540-71050-9}, }
Reference [78]
S. M. Višik and F. V. Dolžanskiĭ, Analogues, connected with Lie groups, of the Euler-Poisson equations and the equations of magnetohydrodynamics (Russian), Dokl. Akad. Nauk SSSR 238 (1978), no. 5, 1032–1035. MR0474423,
Show rawAMSref \bib{ViDo1978}{article}{ author={Vi\v {s}ik, S. M.}, author={Dol\v {z}anski\u {\i }, F. V.}, title={Analogues, connected with Lie groups, of the Euler-Poisson equations and the equations of magnetohydrodynamics}, language={Russian}, journal={Dokl. Akad. Nauk SSSR}, volume={238}, date={1978}, number={5}, pages={1032--1035}, issn={0002-3264}, review={\MR {0474423}}, }
Reference [79]
M.-K. von Renesse, An optimal transport view of Schrödinger’s equation, Canad. Math. Bull. 55 (2012), no. 4, 858–869, DOI 10.4153/CMB-2011-121-9. MR2994690,
Show rawAMSref \bib{Re2012}{article}{ author={von Renesse, Max-K.}, title={An optimal transport view of Schr\"{o}dinger's equation}, journal={Canad. Math. Bull.}, volume={55}, date={2012}, number={4}, pages={858--869}, issn={0008-4395}, review={\MR {2994690}}, doi={10.4153/CMB-2011-121-9}, }
Reference [80]
G. Webb, Magnetohydrodynamics and Fluid Dynamics: Action Principles and Conservation Laws, Springer, 2018.
Reference [81]
H. Wu and M. Wunsch, Global existence for the generalized two-component Hunter-Saxton system, J. Math. Fluid Mech. 14 (2012), no. 3, 455–469, DOI 10.1007/s00021-011-0075-9. MR2964744,
Show rawAMSref \bib{HaWu2011}{article}{ author={Wu, Hao}, author={Wunsch, Marcus}, title={Global existence for the generalized two-component Hunter-Saxton system}, journal={J. Math. Fluid Mech.}, volume={14}, date={2012}, number={3}, pages={455--469}, issn={1422-6928}, review={\MR {2964744}}, doi={10.1007/s00021-011-0075-9}, }
Reference [82]
L. Younes, Shapes and diffeomorphisms, Applied Mathematical Sciences, vol. 171, Springer-Verlag, Berlin, 2010, DOI 10.1007/978-3-642-12055-8. MR2656312,
Show rawAMSref \bib{Yo2010}{book}{ author={Younes, Laurent}, title={Shapes and diffeomorphisms}, series={Applied Mathematical Sciences}, volume={171}, publisher={Springer-Verlag, Berlin}, date={2010}, pages={xviii+434}, isbn={978-3-642-12054-1}, review={\MR {2656312}}, doi={10.1007/978-3-642-12055-8}, }
Reference [83]
J.-C. Zambrini, Variational processes and stochastic versions of mechanics, J. Math. Phys. 27 (1986), no. 9, 2307–2330, DOI 10.1063/1.527002. MR854761,
Show rawAMSref \bib{Za1986}{article}{ author={Zambrini, J.-C.}, title={Variational processes and stochastic versions of mechanics}, journal={J. Math. Phys.}, volume={27}, date={1986}, number={9}, pages={2307--2330}, issn={0022-2488}, review={\MR {854761}}, doi={10.1063/1.527002}, }

Article Information

MSC 2010
Primary: 35Q35 (PDEs in connection with fluid mechanics), 58B20 (Riemannian, Finsler and other geometric structures), 76-02 (Research exposition)
Author Information
Boris Khesin
Department of Mathematics, University of Toronto, Toronto, Canada
MathSciNet
Gerard Misiołek
Department of Mathematics, University of Notre Dame, Notre Dame, Indiana
Klas Modin
Department of Mathematics, Chalmers University of Technology, Gothenburg, Sweden; and University of Gothenburg, Gothenburg, Sweden
ORCID
MathSciNet
Additional Notes

Part of this work was done while the first author held the Pierre Bonelli Chair at the IHES. He was also partially supported by an NSERC research grant and a Simons Fellowship.

Part of this work was done while the second author held the Ulam Chair Visiting Professorship in University of Colorado at Boulder.

The third author was supported by the Swedish Foundation for International Cooperation in Research and Higher Eduction (STINT) grant No. PT2014-5823, by the Swedish Research Council (VR) grant No. 2017-05040, and by the Knut and Alice Wallenberg Foundation grant No. WAF2019.0201.

Journal Information
Bulletin of the American Mathematical Society, Volume 58, Issue 3, ISSN 1088-9485, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on and published on .
Copyright Information
Copyright 2021 American Mathematical Society
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/bull/1728
  • MathSciNet Review: 4273106
  • Show rawAMSref \bib{4273106}{article}{ author={Khesin, Boris}, author={Misio\l ek, Gerard}, author={Modin, Klas}, title={Geometric hydrodynamics and infinite-dimensional Newton's equations}, journal={Bull. Amer. Math. Soc.}, volume={58}, number={3}, date={2021-07}, pages={377-442}, issn={0273-0979}, review={4273106}, doi={10.1090/bull/1728}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.