From Second-Order Differential Geometry to
Stochastic Geometric Mechanics
Abstract
Classical geometric mechanics, including the study of symmetries, Lagrangian and Hamiltonian mechanics, and the Hamilton-Jacobi theory, are founded on geometric structures such as jets, symplectic and contact ones. In this paper, we shall use a partly forgotten framework of second-order (or stochastic) differential geometry, developed originally by L. Schwartz and P.-A. Meyer, to construct second-order counterparts of those classical structures. These will allow us to study symmetries of stochastic differential equations (SDEs), to establish stochastic Lagrangian and Hamiltonian mechanics and their key relations with second-order Hamilton-Jacobi-Bellman (HJB) equations. Indeed, stochastic prolongation formulae will be derived to study symmetries of SDEs and mixed-order Cartan symmetries. Stochastic Hamilton’s equations will follow from a second-order symplectic structure and canonical transformations will lead to the HJB equation. A stochastic variational problem on Riemannian manifolds will provide a stochastic Euler-Lagrange equation compatible with HJB one and equivalent to the Riemannian version of stochastic Hamilton’s equations. A stochastic Noether’s theorem will also follow. The inspirational example, along the paper, will be the rich dynamical structure of Schrödinger’s problem in optimal transport, where the latter is also regarded as a Euclidean version of hydrodynamical interpretation of quantum mechanics.
AMS 2020 Mathematics Subject Classification: 70L10, 35F21, 70H20, 49Q22.
Keywords and Phrases: Stochastic Hamiltonian mechanics, stochastic Lagrangian mechanics, Hamilton-Jacobi-Bellman equations, stochastic Hamilton’s equations, stochastic Euler-Lagrange equation, stochastic Noether’s theorem, Schrödinger’s problem, second-order differential geometry.
1 Introduction
Hamilton-Jacobi (HJ) partial differential equations and the associated theory lie at the center of classical mechanics [1, 7, 63, 35]. Motivated by Hamilton’s approach to geometrical optic where the action represents the time needed by a particle to move between two points and a variational principle due to Fermat, Jacobi extended this approach to Lagrangian and Hamiltonian mechanics. Jacobi designed a concept of “complete” solution of HJ equations allowing him to recover all solutions simply by substitutions and differentiations. Although, in general, it is more complicated to solve than a system of ODEs like Hamilton’s ones, HJ equations proved to be powerful tools of integration of classical equations of motion. In addition, Jacobi’s approach suggested him to ask what diffeomorphisms of the cotangent bundle, the geometric arena of canonical equations, preserve the structure of these first order equations. Those are called today symplectic or canonical transformations and Jacobi’s method of integration is precisely one of them.
It is not always recognized as it should be that HJ equations were also fundamental in the construction of quantum mechanics. The reading of Schrödinger [78], Fock [30], Dirac [20] and others until Feynman [28] makes abundantly clear that most of new ideas in the field made use of HJ equations for the classical system to be “quantized”, or some quantum deformation of them. There are at least two ways to express this deformation. On the one hand, one can exponentiate the wave function, call its complex exponent and look for the equation solved by (see [35]). When the system is a single particle in a scalar potential, one obtain the classical HJ equation with an additional Laplacian term and factor , representing the regularization expected from the quantization of the system. This complex factor is symptomatic of the basic quantum probability problem, at least for pure states. In a nutshell, it is the reason why Feynman’s diffusions, in his path integral approach, do not exist. On the other hand, there is an hydrodynamical interpretation of quantum mechanics, founded on Madelung transform, a polar representation of the wave function whose real part is the square root of a probability density. The argument solves another deformation of HJ equation. The geometry of this transform has been thoroughly investigated recently, highlighting its relations with optimal transport theory [49, 86].
However, the probabilistic content of quantum mechanics, especially for pure states, remained a vexing mathematical mystery right from its beginning, despite several interesting (but unsuccessful) attempts [69]. The current consensus is that regular probability theory and stochastic analysis have little or nothing to teach us about it. And, in particular, that all that can be saved from Feynman path integral theory is Wiener’s measure and perturbations of it by potential terms. This is the “Euclidean approach”, one of the starting points of mathematical quantum field theory.
In 1931, however, Schrödinger suggested in a paper almost forgotten until the eighties [79] (but insightfully commented by the probabilist S. Bernstein [11]) the existence of a completely different Euclidean approach to quantum dynamics. In short, a stochastic variational boundary value problem for probability densities characterizes optimal diffusions on a given time interval as having a density product of two positive solutions of time adjoint heat equations. This idea, revived and elaborated from 1986 [88], is known today as “Schrödinger’s problem” in the community of optimal transport, where it has proved to provide, among other results, very efficient regularization of fundamental problems of this field [58]. In fact, Schrödinger’s problem hinted toward the existence of a stochastic dynamical theory of processes, considerably more general than its initial quantum motivation. In it, various regularizations associated with the tools of stochastic calculus should play the role of those involved in quantum mechanics in Hilbert space, where the looked-for measures do not exist.
The variational side of the stochastic theory has been developed in the last decades, inspired by number of results in stochastic optimal control [37, 29] and stochastic optimal transport [67]. In this context, the crucial role of (second-order) Hamilton-Jacobi-Bellman (HJB) equation has been known for a long time. It provides the proper regularization of the (first-order) HJ equation needed to construct well defined stochastic dynamical theories. In contrast, for instance, with the notion of viscosity solution, whose initial target was the study of the classical PDE, HJB equation becomes central, there, as natural stochastic deformation of this one, compatible with Itô’s calculus. It is worth mentioning that in any fields like AI or reinforcement learning, where HJB equations play a fundamental role [75], it is natural to expect that such a stochastic dynamical framework, built on them, should present some interest.
The geometric, and especially, Hamiltonian side of the dynamical theory had resisted until now and constitutes the main contribution of this paper. It is our hope that it will be useful far beyond its initial motivation referred to, afterward, as its “inspirational examples”. In this sense, it can clearly be interpreted as a general contribution to stochastic geometric mechanics. More precisely, we are trying to answer the following questions:
-
•
Do we have any geometric interpretation of the Hamilton-Jacobi-Bellman equation? That is, can we derive the HJB equation from some sort of canonical transformations?
-
•
Can we formulate some variational problem that leads to a Euler-Lagrange equation which is equivalent to the HJB equation?
-
•
More systematically, can we develop some counterpart of Lagrangian and Hamiltonian mechanics that are associated with the HJB equation?
The first question indicates that canonical transformations should be somehow second-order, so that the corresponding symplectic and contact structures are also second-order. Meanwhile, the stochastic generalization of optimal control and optimal transport suggests that the variational problem of the second question should be formulated in stochastic sense. Combining these hints, the third question amounts to seeking a new theory of geometric mechanics that integrates stochastics and second-order together.
The cornerstone of stochastic analysis, the well-known Itô’s formula, tells us that the generator of a diffusion process is a second-order differential operator. This provides a very natural way to connect the stochastics with the second-order. That is, in order to build a stochastic or second-order counterpart of geometric mechanics, we need to encode the rule of Itô’s formula into the geometric structures.
There is a theory named second-order differential geometry (“stochastic differential geometry” is also used by some authors but we would like to keep the original terminology), which was devised by L. Schwartz and P.-A. Meyer around 1980 [80, 81, 82, 64, 65], and later on developed by Belopolskaya and Dalecky [10], Gliklikh [34], Emery [25], etc. See [26] for a survey of this aspect. Compared with the theory of stochastic analysis on manifolds (or geometric stochastic analysis) developed by Itô [44, 45], Malliavin [62], Bismut [12] and Elworthy [24] etc., which focus on Stratonovich stochastic differential equations on classical geometric structures, like Riemannian manifolds, frame bundles and Lie groups, so that the Leibniz’s rule is preserved, Schwartz’ second-order differential calculus alter the underlying geometric structures to include second-order Itô correction terms, and provide a broader picture even though it loses Leibniz’s rule and is much less known.
In this paper, we will adopt the viewpoint of Schwartz–Meyer and enlarge their picture to develop a theory of stochastic geometric mechanics. We first give an equivalent and more intuitive description for the second-order tangent bundle by equivalent classes of diffusions, via Nelson’s mean derivatives. And then we generalize this idea to construct stochastic jets, from which stochastic prolongation formulae are proved and the stochastic counterpart of Cartan symmetries is studied. The second-order cotangent bundle is also studied, which helps us to establish stochastic Hamiltonian mechanics. We formulate the stochastic Hamilton’s equations, a system of stochastic equations on the second-order cotangent bundle in terms of mean derivatives. By introducing the second-order symplectic structure and the mixed-order contact structure, we derive the second-order HJB equations via canonical transformations. Finally, we set up a stochastic variational problem on the space of diffusion processes, also in terms of mean derivatives. Two kinds of stochastic principle of least action are built: stochastic Hamilton’s principle and stochastic Maupertuis’s principle. Both of them yield a stochastic Euler-Lagrange equation. The equivalence between the stochastic Euler-Lagrange equation and the HJB equation is proved, which exactly leads to the equivalence between our stochastic variational problem and Schrödinger’s problem in optimal transport. Last but not least (actually vital), a stochastic Noether’s theorem is proved. It says that every symmetry of HJB equation corresponds to a martingale that is exactly a conservation law in the stochastic sense. It should be observed, however, that the Schwartz-Meyer approach, together with the one of Bismut [12], has also inspired a distinct, Stratonovich-type stochastic Hamiltonian framework [53] leading to a stochastic HJ equation [54], without relations with Schrödinger’s problem or optimal transport.
The key results of the present paper and the dependence among them are briefly expressed in the following diagram:
The organization of this paper is the following.
Chapter 2 is a summary on the theory of stochastic differential equations on manifolds, in the perspective appropriate to our goal. In particular, diffusions will be characterized by their mean and quadratic mean derivatives as in Nelson’s stochastic mechanics [69] although the resulting dynamical content of our theory will have very little to do with his. In this way, we are able to rewrite Itô SDEs on manifolds as ODE-like equations that have better geometric nature. The notion of second-order tangent bundle answers to the question: the drift parts of Itô SDEs are sections of what?
Chapter 3 is devoted to the notion of Stochastic jets. In the same way as tangent vector on are defined as equivalence classes of smooth curves through a given point and then generalized to higher-order cases to produce the notion of jets, the stochastic tangent vector is defined as equivalence classes of diffusions so that the stochastic tangent bundle is isomorphic to the elliptic subbundle of the second-order tangent bundle. Stochastic jets are also constructed. This provides an intrinsic definition of SDEs under consideration.
Chapter 4 illustrates the use of the above geometric formulation of SDEs for the study of their symmetries. Prolongations of -valued diffusions are defined as new processes with values on the stochastic tangent bundle. Among all deterministic space-time transformations, bundle homomorphisms will be the only subclass to transform diffusions to diffusions. Total mean and quadratic derivative are defined in conformity with the rules of Itô’s calculus. The prolongation of diffusions allows to define symmetries of SDEs and their infinitesimal versions. Stochastic prolongation formulae are derived for infinitesimal symmetries, which yield determining equations for Itô SDEs.
In Chapter 5, the second-order cotangent bundle, as dual bundle of second-order tangent bundle, is defined and analyzed. The properties of second-order differential operator, pushforwards and pullbacks are described. When time is involved, i.e., the base manifold is the product manifold , the corresponding bundles are mixed-order tangent and cotangent bundles, where “mixed-order” means they are second-order in space but first-order in time. More about this topic, like mixed-order pushforwards and pullbacks, pushforwards and pullbacks by diffusions, and Lie derivatives, can be found in Appendix A. An generalized notion to stochastic Cartan distribution and its symmetries are discussed in Appendix B based on the mixed-order contact structure.
The point of Chapter 6 is to use the tools developed before in the construction of the stochastic Hamiltonian mechanics which is one of the main goals of the paper. One of our inspirational example will be the one underlying the dynamical content of Schrödinger’s problem. By analogy with Poincaré -form in the cotangent bundle of classical mechanics and its associated symplectic form, one can construct counterparts in the second-order cotangent bundle. Using the canonical second-order symplectic form on second-order cotangent bundles, one defines second-order symplectomorphisms. The generalization of classical Hamiltonian vector fields becomes second-order operators, for a given real-valued Hamiltonian function on the second-order cotangent bundle. The resulting stochastic Hamiltonian system involves pairs of extra equations compared with their classical versions. Bernstein’s reciprocal processes inspired by Schrödinger’s problem are described in this framework, corresponding to a large class of second-order Hamiltonians on Riemannian manifolds. A mixed-order contact structure describes time-dependent stochastic Hamiltonian systems. The last section of this chapter is devoted to canonical transformations preserving the form of stochastic Hamilton’s equations. The corresponding generating function satisfies the Hamilton-Jacobi-Bellman equation.
Chapter 7 treats the stochastic version of classical Lagrangian mechanics on Riemannian manifolds. Itô’s stochastic deformation of the classical notion of parallel displacements are recalled. Another one, called damped parallel displacement in the mathematical literature, involving the Ricci tensor, is also indicated. Each of these displacements corresponds to a mean covariant derivative along diffusions. The action functional is defined as expectation of Lagrangian and the stochastic Euler-Lagrange equation involves the damped mean covariant derivative. The dynamics of Schrödinger’s problem is, again, used as illustration. The equivalence between stochastic Hamilton’s equations on Riemannian manifolds and the stochastic Euler-Lagrange one as well as the HJB equation are derived via the Legendre transform. Relations with stochastic control are also mentioned. The chapter ends with the stochastic Noether’s theorem. The stochastic version of Maupertuis principle, as the twin of stochastic Hamilton’s principle, is left into Appendix C.
We end the introduction with a list of notations and abbreviations frequently used in the paper, for reader’s convenience.
1.1 List of main notations and abbreviations
HJB equation | Hamilton-Jacobi-Bellman equation |
MDE | Mean differential equations |
2nd-order | Second-order |
SDE | Stochastic differential equations |
S-EL equation | Stochastic Euler-Lagrange equation |
S-H equations | Stochastic Hamilton’s equations |
A general second-order differential operator or second-order vector field | |
Generator of the diffusion | |
Stratonovich stochastic differential | |
Exterior differential on the manifold , or Itô stochastic differential | |
d | Linear operator extended from the exterior differential on the tangent bundle |
Second-order differential on | |
Mixed-order differential on | |
Horizontal differential on the tangent bundle or cotangent bundle | |
Vertical differential on | |
Mean derivatives | |
Total mean derivatives | |
Mean covariant derivative and damped mean covariant derivative | |
, | Connection Laplacian and Laplace-de Rham operator |
Second-order pushforward and pullback of a smooth map | |
Mixed-order pushforward and pullback of | |
Christoffel symbols or stochastic parallel displacement | |
Damped parallel displacement | |
Various sets of -valued diffusions starting at time | |
Stochastic tangent vectors and stochastic jets | |
Lie derivatives | |
Linear connection, Levi-Civita connection, covariant derivative, or gradient on | |
Hessian operator | |
Vertical gradient on | |
Probability space with -field and probability measure | |
, | Past (nondecreasing) filtration and future (nonincreasing) filtration |
Differential operator with respect to coordinate | |
Differential operator with respect to coordinate | |
Second-order differential operator with respect to coordinates and | |
Differential operator with respect to coordinate | |
, | Riemann curvature tensor and Ricci -tensor |
Second-order tangent bundle and second-order elliptic tangent bundle | |
Stochastic tangent bundle | |
Second-order cotangent bundle | |
A general vector field | |
Canonical coordinates on | |
Canonical coordinates on | |
Pushforward and pullback of the diffusion | |
A horizontal diffusion valued on a general bundle or on | |
A horizontal diffusion valued on |
2 Stochastic differential equations on manifolds
In this chapter, we will study several types of stochastic differential equations on manifolds which are weakly equivalent to Itô SDEs. We start with a -dimensional smooth manifold and a probability space , and equip the latter with a filtration , i.e., a family of nondecreasing sub--fields of . We call a past filtration. Unless otherwise specified, the manifold will not be endowed with any structures other than the smooth structure. In some cases, it will be endowed with a linear connection, a Riemannian metric, or a Levi-Civita connection.
Recall from [40, Definition 1.2.1] that by an -valued (forward) -semimartingale, we mean a -adapted continuous -valued process , where and is a -stopping time satisfying , such that is a real-valued -semimartingale on for all . The stopping time is called the lifetime of . If we adopt the convention to introduce the one-point compactification of by , then the process can be extended to the whole time line by setting for all . The point is often called the cemetery point in the context of Markovian theory.
2.1 Itô SDEs on manifolds
Given time-dependent vector field on , one can introduce a Stratonovich SDE in local coordinates, which has the same form as in Euclidean space [40, Section 1.2]. The form of Stratonovich SDEs on is invariant under changes of coordinates, as Stratonovich stochastic differentials obey the Leibniz’s rule.
However, for Itô stochastic differentials this is not the case because of Itô’s formula. Hence, we cannot directly write an Euclidean form of Itô SDE on in local coordinates, since it is no longer invariant under changes of coordinates. Indeed, a change of coordinates will always produce an additional term. To balance this term, a common way is to add a correction term to the drift part of the Euclidean form of Itô SDE, by taking advantage of a linear connection. More precisely, under local coordinates , we consider the following Itô SDE [34, Section 7.1, 7.2]:
(2.1) |
where is the family of Christoffel symbols for a given linear connection on . When conditioning on and taking as normal coordinates at , (2.1) turns to the Euclidean form, since at ,
(2.2) |
If we denote
Then, clearly is a symmetric and positive semi-definite -tensor field. We also introduce formally a modified drift which has the following coordinate expression
(2.3) |
We change the coordinate chart from to with . Since each transforms as a vector, we apply the change-of-coordinate formula for Christoffel symbols (e.g., [50, Proposition III.7.2]) to derive that
It follows that the coefficients of the modified drift in (2.3) transform as
(2.4) |
Therefore, is not a vector field as it does not pointwisely transform as a vector.
Finally, using Itô’s formula, we derive the transformation of (2.1) as follows:
where the bracket on the right hand side (RHS) of the first equality denotes the quadratic variation. This shows that equation (2.1) is indeed invariant under changes of coordinates.
Remark 2.1.
One can regard as an -valued vector field on . In this way, the pair is called an Itô vector field in [34, Chapter 7], while the pair is called an Itô equation therein.
Now we present the definition of weak solutions to (2.1).
Definition 2.2 (Weak solutions to Itô SDEs).
Given a linear connection on , a weak solution of the Itô SDE (2.1) is a triple , , , where
-
(i)
is a probability space, and is a past (i.e., nondecreasing) filtration of satisfying the usual conditions,
-
(ii)
is a continuous, -adapted -valued process with -stopping time , is an -dimensional -Brownian motion, and
-
(iii)
for every , and any coordinate chart of , it holds under the conditional probability that almost surely in the event ,
Definition 2.3 (Uniqueness in law).
We say that uniqueness in the sense of probability law holds for the Itô SDE (2.1) if, for any two weak solutions , , , and , , with the same initial data, i.e., , the two processes and have the same law.
Note that it is possible to change and in the Itô SDE (2.1) but keep the same weak solution in law. In other words, the form of (2.1) does not univocally correspond to its weak solution in law. For this reason, we will reformulate SDEs in a fashion that makes them look more like ODEs and have better geometric nature. Moreover, we will see that it is the pair that univocally corresponds to the weak solution of (2.1).
2.2 Mean derivatives and mean differential equations on manifolds
In this part, we will recall the definitions of Nelson’s mean derivatives and extend them to -valued processes. In Nelson’s stochastic mechanics [69], the probability space is equipped with two different filtrations. The first one is just an usual nondecreasing filtration , a past filtration. The second is a family of nonincreasing sub--fields of , which is denoted by and called a future filtration. For an -valued process , its forward mean derivative and forward quadratic mean derivative are defined by conditional expectations as follows:
Their backward versions, i.e., the backward mean derivative and backward quadratic mean derivative, are defined as follows:
In our present paper, we will only focus on the “forward” case, so that only the past filtration will be invoked. The “backward” case is analogous and every part of this paper can have its “backward” counterpart (cf. [89]).
Denote by (and ) the fiber bundle of symmetric (and respectively, symmetric positive semi-definite) -tensors on . Now we define quadratic mean derivatives for -valued semimartingales, cf. [34, Chapter 9].
Definition 2.4 (Quadratic mean derivatives).
The (forward) quadratic mean derivative of the -valued semimartingale is a -valued process on , whose value at time in any coordinate chart and in the event is given by
(2.5) |
where the limits are assumed to exist in .
More generally, we can define the (forward) quadratic mean derivative for two -valued semimartingales and in local coordinates by
Due to Itô’s formula for semimartingales, does transform as a -tensor and is obviously symmetric, so that the definition is independent of the choice of . However, the formal limit under any coordinates , no longer transforms as a vector, as can be guessed from (2.4). In order to turn it into a vector we need to specify a coordinate system. A natural choice is the normal coordinate system. For this purpose, we endow with a linear connection , which determines a normal coordinate system near each point on .
Definition 2.5 (-mean derivatives).
Given a linear connection on , the (forward) -mean derivative of the -valued semimartingale is a -valued process on , whose value at time is defined in normal coordinates on the normal neighborhood of and under the conditional probability as follows:
where the limits are assumed to exist in .
As we force to be vector-valued by definition, its coordinate expression under any other coordinate system can be calculated via Leibniz’s rule. Let us stress that the notation should not be confused with the one of covariant derivatives in geometry.
Now we formally take forward mean derivatives in Itô SDE (2.1), and note that the correction term in the modified drift involving Christoffel symbols vanishes by (2.2). Then, we get an ODE-like system:
(2.6) |
We call equations (2.6) a system of mean differential equations (MDEs). Note that both MDEs (2.6) and Itô SDE (2.1) rely on linear connections on .
Definition 2.6 (Solutions to MDEs).
Given a linear connection on , a solution of MDEs (2.6) is a triple , , , where
-
(i)
is a probability space, and is a past filtration of satisfying the usual conditions,
-
(ii)
is a continuous, -adapted -valued semimartingale with lifetime a -stopping time , and
-
(iii)
the -mean derivative and quadratic mean derivative of exist and satisfy (2.6).
2.3 Second-order operators and martingale problems
Definition 2.7 (Second-order operators).
A second-order operator on is a linear operator , which has the following expression in a coordinate chart ,
(2.7) |
where is a symmetric -tensor field, and the expression is required to be invariant under changes of coordinates. If is positive semi-definite, then we say the second-order operator is elliptic; if is positive definite, we say is nondegenerate elliptic.
There is a coordinate-free definition of second-order operators. A linear map is called a second-order derivation at , if there is a symmetric -tensor at such that for all . Then, a second-order operator is nothing but a smooth field of second-order derivations. From this, we see that for in (2.7), , , and
(2.8) |
We call the squared field operator (originally “opérateur carré du champ”) associated with . We also denote . Clearly, for a classical vector field , by Leibniz’s rule.
It is easy to verify from the coordinate-change invariance that the coefficients ’s and ’s transform under the change of coordinates from to by the following rule (e.g., [43, Section V.4]),
(2.9) |
The formal generator of Itô SDE (2.1) is given by,
(2.10) |
which is a time-dependent second-order elliptic operator due to the change-of-coordinate formula (2.4).
Denote by the subspace of consisting of all paths always staying in or eventually stopped at . That is, if and only if there exists such that for and for . Let be the -field generated by Borel cylinder sets. Let be the coordinate mapping. For each , define a sub--field by . Then is a past filtration of and is a -stopping time.
Definition 2.8 (Martingale problems on manifolds, [40, Definition 1.3.1]).
Given a time-dependent second-order elliptic operator , a solution to the martingale problem associated with is a triple , , , where
-
(i)
is a probability space, and is a past filtration of satisfying the usual conditions,
-
(ii)
is an -valued -semimartingale, and
-
(iii)
for every , the process , , is a real-valued continuous -martingale.
The process is called an -valued -diffusion process with generator (or simply an -diffusion).
The uniqueness in the sense of probability law for both MDEs and martingale problems can be defined in a similar fashion to Definition 2.3. Note that unlike Itô SDEs or MDEs, the definition for martingale problems does not rely on linear connections.
When provided with a linear connection on , one can see, in the same way as in Stroock and Varadhan’s theory (e.g., [48, Section 5.4]), that the existence of a solution to the martingale problem associated with in (2.10) is equivalent to the existence of a weak solution to the Itô SDE (2.1), and also equivalent to the existence of a solution to MDEs (2.6); their uniqueness in law of are also equivalent.
2.4 The second-order tangent bundle
As we have seen, the modified drift in (2.3) is not a vector field. Is a section (and, in the affirmative, of what)? In fact, it is not a section of any bundle, as its changes-of-coordinate formula (2.4) involves . But if we look at the formal generator in (2.10), or the pair of its coefficients, then we can construct a bundle whose structure group is governed by the changes-of-coordinate formulae (2.9), so that the sections are just second-order operators.
We denote by the space of all symmetric -tensors on , and by the subspace of it consisting of all positive semi-definite -tensors. Also denote by the space of all linear maps from to .
Definition 2.9 (The second-order tangent bundle).
(i). [34, Definition 7.14] The Itô group is the Cartesian product (but not direct product of groups) equipped with the following binary operation:
for all , .
(ii). The left group action of on is defined by
(2.11) |
for all , , .
(iii). The second-order tangent bundle is the fiber bundle with base space , typical fiber , and structure group .
(iv). The fiber at is called second-order tangent space to at . An element is called a second-order tangent vector at . A (global or local) section of is called a second-order vector field.
(v). Denote by the subbundle of consisting of all elements , , with a positive semi-definite -tensors. Let . We call the second-order elliptic tangent bundle.
Remark 2.10.
(i). We indulge in some abuse of notions. For example, the second-order vector fields should not be confused with the semisprays which are sections of the double tangent bundle (e.g., [77, Section 1.4], [51, Section IV.3]).
(ii). Some authors just defined second-order vector fields as second-order operators as in Definition 2.7 ([25, Definition 6.3] or [34, Definition 2.74]). As soon as we choose a frame for , it will be clear that second-order vector fields are identified with second-order operators.
(iii). The authors in [10, 34] define a bundle which has the Itô group as its structure group and has the pair of coefficients in Itô SDE (2.1) as its section. They name it Itô’s bundle and denote it as . The difference is that, in our formulation, the pair of coefficients of the generator of Itô SDE (2.1) is a section of second-order elliptic tangent bundle . The advantage of the bundle is that it is a natural generalization of tangent bundle to second-order and has a good geometric interpretation, as we will see in Proposition 3.2.
(iv). Note that the typical fiber of is a vector space of dimension . But is not a vector bundle, since its structure group is not a linear group (subgroup of general linear group). The typical fiber of is , which is not even a vector space, so that is not a vector bundle either. Indeed, we may call them quadratic bundles, just as the way they call Itô’s bundle in [10, Chapter 4].
(v). The Itô’s bundle defined in [34, Definition 7.17] is the fiber bundle over manifold , with fiber and structure group which acts on the fiber from the left by
for all , , . For the same reason as or , Itô’s bundle is not a vector bundle. There is a bundle homomorphism over from to , which maps in fibers from to , , by . It is easy to see that this bundle homomorphism is also a subjective submersion. If we identify with , then is a subgroup of . We define the Stratonovich’s bundle to be the reduction of to the structure group , that is, the fiber bundle over , with fiber and structure group which acts on the fiber from the left by
Unlike or , Stratonovich’s bundle is indeed a vector bundle, and the tangent bundle is a vector subbundle of . It can be expected that Stratonovich’s bundle is a natural bundle to formulate Stratonovich SDEs. But, in this paper, we mainly focus on Itô SDEs and their generators.
It is natural to regard the differential operators
(2.12) |
as a local frame of over the local chart on . In the sequel, we will usually shorten them by
We make the convention that for all . A second-order vector field is expressed in terms of this local frame by
In this way, every second-order vector field can be regarded as a second-order operator and vice versa. In particular, the generator of an -valued diffusion process , for example the generator (2.10) of the Itô SDE, is a time-dependent second-order vector field, so that we can rewrite as .
The tangent bundle is a subbundle (but not a vector subbunddle) and also an embedded submanifold of , as the bundle monomorphism
(2.13) |
is also an embedding. However, there is no canonical bundle epimorphism from to which is a left inverse of and linear in fiber. We call such a bundle epimorphism a fiber-linear bundle projection from to . The choice of such a bundle epimorphism is exactly the choice of a linear connection on . More precisely, we have the following connection correspondence properties, the first of which can also be found in [34, Section 2.9].
Proposition 2.11 (Connection correspondence).
Any linear connection on induces a fiber-linear bundle projection from to . Conversely, any fiber-linear bundle projection from to induces a torsion-free linear connection on .
Remark 2.12.
The connection correspondence is similar to the correspondence between horizontal subbundles of the tangent bundle of a vector bundle and connections on this vector bundle, cf. [77, Section 3.1].
Proof.
Let be the Christoffel symbols of a linear connection on . Define a projection by
(2.14) |
Clearly, is linear in fiber and . Conversely, let be a fiber-linear bundle projection. Then, on each coordinate chart around , there exists a diffeomorphism , such that
The family of diffeomorphisms determines a spray and then a torsion-free linear connection on (see, e.g., [51, Section IV.3]). The torsion-freeness follows from the symmetry of ’s. ∎
Observe that a group action of on can be separated from (2.11), which is given by . Thus the second component of each element can be regarded as a -tensor. Recall that we denote by the bundle of -tensors on , then there is a canonical bundle epimorphism
(2.15) |
whose kernel is the image of . Conversely, we also have a similar connection correspondence property for , as in Proposition 2.11. That is, a linear connection on induces a fiber-linear bundle monomorphism from to , which is a right inverse of and given by
(2.16) |
where is the second covariant derivative [74, Subsection 2.2.2.3] (which is also called the Hessian operator when acting on smooth functions [47]). In other words, , where is the symmetrization operator on .
Combining (2.13) and (2.14) together, we have the following short exact sequence:
(2.17) |
Proposition 2.11 and (2.15), (2.16) imply that when a linear connection is given, the sequence is also split, in the fiber-wise sense. The induced decomposition
(2.18) |
where both the first direct sum and the isomorphism are in the fiber-wise sense (but not bundle isomorphism and Whitney sum), while the second direct sum is the Whitney sum, and is given by
(2.19) |
for . A similar short exact sequence as (2.17) holds with and in place of and , respectively.
Now we introduce a subclass of semimartingales on manifolds which contains diffusions. We call the -valued process an Itô process, if there exists a -adapted continuous -valued process satisfying for each , such that for every , , is a real-valued -martingale, where . We call the process the random generator of . A similar notion “Brownian semimartingale” is also used in the literature (e.g., [22]). If is a diffusion with generator , then it is an Itô process with random generator . The difference between Itô processes and diffusions is that the randomness of the random generator of the former can not only appear on the base manifold , but also on the fibers.
Then, we can define forward mean derivatives in a coordinate-free way, without relying on linear connections.
Definition 2.13 (Mean derivatives).
For an -valued Itô process , we define its (forward) mean derivatives at time by
where is the random generator of .
Comparing with forward mean derivatives defined in local coordinates before, we have the following relations. The proof follows the lines of [34, Lemma 9.4].
Lemma 2.14.
Given an -valued Itô process and a coordinate chart centered at .
(i). In the event , has the coordinate expression (2.5) and
(ii). Given a linear connection on , we have, under the conditional probability , that
(2.20) |
For a time-dependent second-order vector field , we can take MDEs (2.6) to set up a new type of MDEs by using the mean derivatives as follows:
(2.22) |
Then, similarly to Definitions 2.6 and 2.3, we may also define solutions and uniqueness in law for MDEs (2.22). We call a solution of (2.22) an integral process of . Note that the system (2.22) does not rely on linear connections. The equivalence of the well-posedness of (2.22) and the martingale problem in Definition 2.8 is easy to verify. When a linear connection is specified, the system (2.22) and martingale problem associated with in (2.10) are both equivalent to the Itô SDE (2.1) and MDEs (2.6).
3 Stochastic jets
In classical differential geometry, a tangent vector to a manifold may be defined as an equivalence class of curves passing through a given point, where two curves are equivalent if they have the same derivative at that point [55, Chapter 3]. This idea can be generalized to higher-order cases, which leads to the notion of jets. The jet structures allow us to translate a system of differential equations to a system of algebraic equations, and make it more intuitive to study the symmetries of systems of differential equations.
In this chapter we shall generalize these ideas to the stochastic case. We will first give an equivalent description to the second-order elliptic tangent bundle by constructing an equivalence relation on diffusions. Then we will define the stochastic jets and figure out the “jet-like” bundle structure involved in the space of stochastic jets. Finally, we shall see that the bundle structure is the appropriate platform to formulate SDEs intrinsically. In the next chapter, we will apply stochastic jets to study stochastic symmetries.
3.1 The stochastic tangent bundle
Recall that a tangent vector can be represented as a equivalence classes of smooth curves that have the same velocity at the base point. This leads to the following equivalent definition of tangent bundle :
(3.1) |
where is the set of all smooth curves on that pass through at time , and the equivalence relation is defined as are equivalent if and only if for every real-valued smooth function defined in neighborhood . If we replace smooth curves by diffusion processes, and time derivatives by mean derivatives, then we get the following definition.
Definition 3.1 (The stochastic tangent bundle).
Two -valued diffusion processes , are said to be stochastically equivalent at , if, almost surely, and for all . The equivalence class containing is called the stochastic tangent vector of at and is denoted by . When , we denote in short. Let be the set of all -valued diffusion processes starting from at time . The stochastic tangent bundle of is the set
Note that since are -valued diffusion processes, and are real-valued Itô processes, and hence their mean derivatives exists.
At this stage, we have not yet touched the jet-like formulation even though we used the jet-like notation . Indeed, if one follows strictly the definition of jet bundles over the trivial bundle , it is more rational to use the time line as “source” and the manifold as “target” (cf. [77, Example 4.1.16]). But here we just assign the “target” to the manifold , because, roughly speaking, one can talk about the velocity of a smooth curve at a moment , but not about the generator of a diffusion at a moment . Instead, we can talk about the generator of a diffusion at a position . Later on, we will define the “bona fide” stochastic jet space which possess the time line as “source” and the manifold as “target”.
Similarly to the one-to-one correspondence between tangent space and space of equivalence classes of smooth curves, we have the following:
Proposition 3.2.
There is a one-to-one correspondence between the stochastic tangent bundle and the second-order elliptic tangent bundle .
Proof.
For an -valued diffusion process , , we denote by its generator. Then the map defines a one-to-one correspondence between and . The inverse map is , where is a section of (i.e., an elliptic second-order operator) smoothly extending the element , and is a diffusion processes having as its generator. ∎
Therefore, the stochastic tangent bundle admit a smooth structure which makes it to be a smooth manifold diffeomorphic to , and hence it is a bona fide fiber bundle over . In the sequel, we will identify with without ambiguity. And the projection map from to will be denoted by , that is, for any .
Definition 3.3 (Canonical coordinate system on ).
Let be an coordinate system on . The induced canonical coordinate chart on is defined by
where , and .
Our slightly ambiguous notations and are chosen so as to avoid the worse one .
When a linear connection is provided, we can also define the coordinates via the -mean derivative instead of , as follows:
Then, also forms a coordinate system on , which we call the -canonical coordinate system. It follows from relation (2.20) that
(3.2) |
Using the identification of elements and via Proposition 3.2, as well as their relations with the element , via (2.19), we have , and . In this way the fiber-linear bundle projection of (2.14) maps, under the canonical coordinates on , as follows:
(3.3) |
so that . Therefore, is a partial coordinate system on that coincides with when restricted on . Moreover, the decomposition in (2.19) yields the following expressions for second-order vector fields:
(3.4) |
Similarly to Definition 3.1, we define a -dependent equivalence relation as follows:
Definition 3.4.
Two -valued diffusion processes , are said to be -stochastically equivalent at , if, almost surely, and . The equivalence class containing is called the -tangent vector of at and is denoted by . When , we denote for short.
Then, similarly to Proposition 3.2, one can show that the tangent bundle can be identified with the following set of equivalent classes of diffusions:
(3.5) |
via . Under this identification, it follows from (2.21) that . Clearly, if we regard all smooth curves as special diffusions, then the partition determined by (3.1) is the restriction of the one determined by (3.5) to the set of all smooth curves.
Remark 3.5.
In presence of a linear connection on , one can easily follow Definition 3.1 and Proposition 3.2 with in place of , to verify the one-to-one correspondence between the set of equivalent classes and the Whitney sum , which brings back to the fiber-wise isomorphism (2.18). But since such kind of correspondence need to specify beforehand a linear connection, we still endow with the structure of instead of that of in this paper, although the latter is also feasible and may provide easier calculations.
3.2 The stochastic jet space
In classical jet theory, for the trivial bundle , there is a one-to-one correspondence between 1-jets and tangent vectors, and there is a canonical diffeomorphism between the first-order jet bundle and [77, Example 4.1.16].
Now using similar ideas, we will introduce the “bona fide” stochastic jet space. The key is to modify the definition of stochastic tangent vectors, to involve the time line as the “source” as well as to randomize the initial datum of the diffusion processes. Intuitively, an -valued diffusion process can be regarded as a random “section” of the trivial “bundle” which is merely continuous in time and depends on the sample point .
For a metric space , we denote by the quotient space of all -valued random elements, by the following equivalence relation: two random elements are equivalent if and only if they are identical almost surely. We endow with the topology of the following -essential metric (cf. [68, Section 43]):
Definition 3.6.
Two -valued diffusion processes , starting at time , are said to be stochastically equivalent at , if, almost surely, and . The equivalence class containing is called the stochastic jet of at , denoted by . Let be the set of all -valued diffusion processes starting at time . Then the stochastic jet space of is the set
The functions and , called stochastic source and target projections, are defined by
and
In the above definition, since , we have a.s., that is, is the projection of .
To characterize the relation between and (or ), we need the following definitions.
Definition 3.7 (Horizontal subspace).
Let be a fiber bundle. The horizontal subspace of is defined by
An element of the horizontal subspace of is then of the form , where is a section of and . Such an element will be denoted by . By the correspondence of and , one can easily get the following equivalent definition for ,
The correspondence is given explicitly by
where is an -valued diffusion with generator and with a.s..
Proposition 3.8.
The stochastic jet space is trivial. More precisely, we have the homeomorphism
given by , for any , where is the shift operator on , that is, .
Proof.
The homeomorphism is given by . The homeomorphism is given by , whose inverse map is . ∎
Definition 3.9 (Stochastic fibered space).
(i) Given a fiber bundle with total space , base space and typical fiber manifold , the stochastic fibered space associated with it is the triplet where
is the natural projection given by , and is a subspace of , with denoting the fiber of over . The fiber bundle is called model bundle of . There is a family of projections from the stochastic fiber manifold to its model bundle , defined by
(ii) A global section of is called a random global section. A random local section is a map defined on some measurable subset and such that, for almost all , is a local section of , where .
Note that a random global section is a random local section defined on all .
It follows from Proposition 3.8 that the stochastic jet space is a stochastic fibered space, whose associated model bundle is . Just like the first-order jet bundle which is diffeomorphic to , the model bundle is itself a jet bundle and also has two bundle structures, with base space and , respectively. The corresponding source and target projections are defined, respectively by
and
Moreover, we will denote the natural projection from to by . This projection map is indeed a bundle homomorphism from to , whose projection is the natural projection from to , denoted by .
Similarly to Proposition 3.8, we have the following diffeomorphisms for the model bundle :
which is given by
(3.6) |
for any , where is the generator of as a section of (i.e., a time-dependent elliptic second-order differential operator). Furthermore, the proof of Proposition 3.2 allows us to find simply the inverse maps, especially for the second diffeomorphism. That is, for any ,
(3.7) |
where is a section of such that , and is a diffusion process having as its generator.
The “stochastic target” of , i.e., the trivial bundle , is another example of stochastic fibered spaces. Its model bundle is the trivial bundle . The graph of an -valued stochastic process defined on a random time interval is a random (local) section of . The projection of on the targets from to is denoted by .
We may summarize how all these maps fit together by the following diagram:
When a linear connection is specified on , one can easily obtain, similarly to (3.6), the following homeomorphism:
and the following diffeomorphisms:
where the first two diffeomorphisms are given by
and the last one is due to the classical theory.
3.3 Intrinsic formulation of SDEs
With the classical machinery of jet structures, it is possible to translate differential equations into algebraic equations on jet bundle [77]. In this section, we follow this way to formulate intrinsic SDEs.
For a subset of the model bundle and , we denote by the intersection of with the fiber .
Definition 3.10.
A stochastic differential equation on is a closed embedded submanifold of the model jet bundle with . A (local) solution of the stochastic differential equation is a triple , , , where
-
(i)
is a probability space, and is a past filtration of satisfying the usual conditions,
-
(ii)
is a -adapted -valued diffusion process over , where is a -stopping time, and
-
(iii)
almost surely for every .
Remark 3.11.
(i). The condition that is just for convenience, in order to set the initial time at .
(ii). There is an equivalent way to formulate the solution of a stochastic differential equation . That is, a (local) solution is a pair , where is a probability measure on and is a -stopping time, such that for -almost surely , for every .
This definition does not look like the traditional definition of a stochastic differential equation, but we can see the relationship between the two by using coordinates. Since is a embedded submanifold of , it admits a local defining function in a neighborhood of each of its points [55, Proposition 5.16]. That is, for a coordinate chart of the point , there is a function where , such that and 0 is a regular value of . Then, the condition before leaves the neighborhood reads in local coordinates as
(3.8) |
which defines a general MDE (in terms of mean derivatives). The use of a submanifold is therefore a way to distinguish the definition of the equation from a definition of its solutions.
As an example, the system of MDEs (2.22) can be rewritten to the form (3.8) by setting the defining function
(3.9) |
So far we have not done anything but reformulate the basic problem of finding solutions of systems of stochastic differential equations in a more geometrical form, ideally suited to our investigation into symmetry groups thereof.
4 Stochastic symmetries
The symmetry group of a system of differential equations is the largest local group of transformations acting on the independent and dependent variables of the system with the property that it transform solutions of the system into other solutions [72]. In the stochastic case, we can proceed analogously.
All methods of this chapter work in the local case, that is, the vector fields are not necessarily complete and the bundle homomorphisms could be only locally defined.
4.1 Prolongations of diffusions and bundle homomorphisms
Definition 4.1 (Prolongations of diffusions).
Let be an -valued diffusion process defined on a stopping time interval . The prolongation of is a -valued process defined by, for the shift operator,
Note that . Thus the graph of the prolongation process is nothing but the random section of the stochastic jet space . It is easy to see that if is an -valued diffusion process, then is a -valued diffusion process.
Given two smooth manifolds and , a bundle homomorphism from to is a projectable (or fiber-preserving) smooth map, which means it maps fibers of to fibers of . Hence, there exist two smooth maps and such that . This leads to which is the original definition of bundle homomorphisms. We denote and say that projects to .
The following lemma shows that a bundle homomorphisms has the property that it always transforms diffusions into diffusions. One can find a proof of it in Lemma 4.8 or Corollary A.5.
Lemma 4.2.
Given a bundle homomorphism from to , where is a diffeomorphism, for every -valued diffusion process , the image of its graph (or its corresponding random local section) by , i.e.,
is almost surely the graph of a well-defined -valued diffusion process given by
(4.1) |
As observed in Remark A.6, among all (deterministic) smooth maps from to , the class of bundle homomorphisms is the only subclass that maps diffusions to diffusions.
Definition 4.3 (Pushforwards of diffusions by bundle homomorphisms).
We call the diffusion of Lemma 4.2 the pushforward of by , and write . When and is a bundle endomorphism on , we also call the transform of by .
We now introduce the idea of stochastic prolongation whereby a bundle homomorphism may be extended to act upon the model jet bundle.
Definition 4.4 (Stochastic prolongations of bundle homomorphisms).
Let be a bundle homomorphism from to projecting to a diffeomorphism . The stochastic prolongation of is the map defined by
(4.2) |
It is easy to see from (4.1) that if , then . Therefore, the map is well defined. By letting , definition (4.2) can be rewritten in a more evident way:
(4.3) |
The following properties are easy to check.
Corollary 4.5.
(i) The map is a bundle homomorphism projecting to .
(ii) The map is a bundle homomorphism projecting to .
(iii) . Let and be two bundle endomorphisms on projecting to diffeomorphisms. Then .
By virtue of (4.3) and Corollary 4.5.(i), we may write , where is the smooth map given by
(4.4) |
We can also consider the pushforward of the -valued process by the bundle homomorphism .
Corollary 4.6.
Given a bundle homomorphism projecting to a diffeomorphism on , and an -valued diffusion process , we have
Now we need to investigate the coordinate representation of , in stochastic analysis terms. Before that, we introduce the stochastic version of the notion of total derivatives.
Definition 4.7 (Total mean derivatives).
Let be a smooth real-valued function on . The total mean derivative and total quadratic mean derivative of are the unique smooth functions and defined on , with the property that if is a representative diffusion process of , then
There is an abuse of notations in the above definition. Indeed, the left-hand sides (LHSs) of the above two equations both involve subscripts , but their RHS’s do not depend on . Those two equations need to be understood as that functions taking their values on the point equal to the RHS’s.
It is easy to check that the definitions of total mean derivatives are independent of the choice of representative diffusions. By Itô’s formula, we have the following coordinate representation for total mean derivatives in the local chart on ,
(4.5) | ||||
If a linear connection is specified, we can use (3.4) to rewrite as follows:
(4.6) |
Lemma 4.8.
Let us be given a bundle homomorphism from to projecting to a diffeomorphism and an -valued diffusion process . If , then in local coordinates around and around ,
Proof.
Assume that the diffusion can be represented in local coordinates by
where is an -dimensional Brownian motion, so that
Let . Then
Define
Then [70, Theorem 8.5.7] says that is an -dimensional -Brownian motion, as by a change of variable , we have
Therefore,
Recall that . Using Itô’s formula, we have
It follows that
This completes the proof. ∎
We denote the induced local coordinates on by . Then clearly, . Now take . Then
(4.7) | |||
(4.8) |
4.2 Symmetries of SDEs
As an important application of the prolongations of diffusions and bundle homomorphisms, we now study the symmetries of stochastic differential equations. As in classical Lie’s theory of symmetries of ODEs, a symmetry of a stochastic differential equation is a space-time transformation that maps solutions to solutions. But this is not sufficient for the stochastic case. As we have mentioned in Section 4.1, the only smooth transformation on mapping diffusions to diffusions are bundle endomorphisms. Moreover, a solution of stochastic differential equation is always accompanied by a filtration, which will also be altered under space-time transformations. Thus, we have the following definition:
Definition 4.9 (Symmetries).
Given a stochastic differential equation , a symmetry of is a bundle automorphism on projecting to such that if is a solution of , then so is .
Using the definitions of stochastic differential equations and pushforwards, we have the following equivalent characterization of symmetries.
Lemma 4.10.
Let be a stochastic differential equation on . A bundle automorphism on is a symmetry of , if and only if, whenever we have , or equivalently, .
Recall that the infinitesimal version of bundle homomorphisms are the so called projectable or fiber-preserving vector fields. More precisely, a vector field on is called -projectable, if the (local) flow (or one-parameter group action) generated by consists of (local) bundle endomorphisms on (cf. [72, Example 2.22] or [77, Proposition 3.2.15]). For such a vector field, we define its prolongation to be the infinitesimal generator of the prolongated flow.
Definition 4.11 (Stochastic prolongations of projectable vector fields).
Let be a -projectable vector field on , with corresponding (local) flow . Then, the stochastic prolongation of , denoted by , will be a vector field on the model jet bundle , defined as the infinitesimal generator of the corresponding prolonged flow . In other words, is a vector field on defined by
for any .
Now we can define infinitesimal versions of symmetries.
Definition 4.12 (Infinitesimal symmetries).
Let be a stochastic differential equation on . An infinitesimal symmetry of is a -projectable vector field on whose stochastic prolongation is tangent to .
The following properties follow straightforwardly from definitions.
Lemma 4.13.
Given a stochastic differential equation on , let be a complete -projectable vector field on and be its flow. Then
(i) is an infinitesimal symmetry of if and only if for every local defining function of ;
(ii) is an infinitesimal symmetry of if and only if for each , is a symmetry of .
4.3 Stochastic prolongation formulae
We consider a coordinate chart on the model jet bundle , which is induced by the coordinate chart on . A -projectable vector field on has the following local coordinate representation
(4.9) |
Its prolongation is a vector field of the form
Now we use Lemma 4.8 to compute the coefficients ’s and ’s.
Theorem 4.14.
Suppose is complete and -projectable and has the local representation (4.9). Then in the canonical coordinates , the coefficient functions of its prolongation are given by the following formulae:
(4.10) | ||||
(4.11) |
Proof.
Let be the flow generated by . Since is complete and -projectable, each is a bundle endomorphism on projecting to a diffeomorphism on . Let . Note that , and
Let be a representative diffusion of . Then by Lemma 4.2 and Definition 4.4, a representative diffusion of is
Now we apply Lemma 4.8 and take derivatives with respect to . Since commutes with the total mean derivative as is clear from the coordinate representation, we have
Also,
In the induced coordinate system , the last two formulae read as (4.10) and (4.11), respectively. ∎
Stochastic analogs of contact structure on and Cartan symmetries will be discussed in Appendix B. It turns out that the infinitesimal symmetry of the mixed-order Cartan distribution is equivalent to stochastic prolongation formulae of Theorem 4.14.
Corollary 4.15.
Proof.
Remark 4.16.
In [31], the author proved a result similar to Corollary 4.15, with the following equation instead of equation (4.12):
(4.13) |
By multiplying both sides of (4.13) with , and using the symmetry for index , one gets easily (4.12). So our determining equations for infinitesimal symmetries are more general than those of [31]. Basically, the paper [31] concerns symmetries of the Itô equation , while we consider symmetries of the diffusion with generator , or equivalently, a weak formulation of SDE. The former symmetries belong to the latter obviously, but not vice versa.
Now given a linear connection on , we define the -dependent versions of Definitions 4.1, 4.4 and 4.11. More precisely, for a diffusion on , we define its -prolongation to be a -valued diffusion given by . For a bundle homomorphism from projecting to a diffeomorphism , the -prolongation of is the map defined by . The -prolongation of , denoted by , is defined to be the infinitesimal generator of the corresponding prolonged flow , so that is a vector field on and has the form
for of the form (4.9). If we denote so that , we have
Corollary 4.17.
Under the canonical coordinates , the coefficient of the -prolongation are given by:
where is the curvature tensor.
5 The second-order cotangent bundle
5.1 Second-order covectors
Definition 5.1 (Second-order cotangent space).
The second-order cotangent space at is the dual vector space of , denoted by . The pairing of and is denoted by or . Elements of are called second-order covectors at . The disjoint union is called the stochastic cotangent bundle of . The natural projection map from to is denoted by . A (local or global) smooth section of is called a second-order covector field or a second-order form.
Dual to the left action (2.11) of on fibers of , will act on those of from the right.
Lemma 5.2.
The stochastic cotangent bundle is the fiber bundle dual to , with structure group acting on the typical fiber from the right by
for all , , .
The notion of second-order forms should not be confused with the classical one of 2-forms. There are two basic examples of second-order forms, say, and , where and are given smooth functions on . They are defined as follows: for ,
(5.1) |
where is the squared field operator defined in (2.8). These notations go back to L. Schwartz [82] and P.A. Meyer [65] (see also [25, Chapters VI and VII]), where the term is called the second differential of , and the term is called the symmetric product of and . Note that in these original references, there is a factor at the RHS of the definition of . Here we drop this factor. Obviously, when restricted to , the second differential is just the differential but the symmetric product vanishes.
The definition of the symmetric product yields two properties: is symmetric in and ; and if one of and vanishes. These lead to a more general definition for symmetric products of two 1-forms. More precisely, let , then there exist smooth functions and on such that and . By the preceding property, the second-order covector does not depend on the choice of and , and we will denote it by . Now if are second-order forms, then their symmetric product is defined pointwisely through . More formally, we have
Definition 5.3 (Symmetric product, [25, Chapter VI]).
There exists a unique fiber-linear bundle homomorphism from to , which is called the symmetric product, such that for all , .
It is easy to verify from (5.1) that the local frame, dual to (2.12), for over the local chart is given by (see also [25, Chapter VI])
We adopt the convention that for all . Under this frame, a second-order covector has a local expression
(5.2) |
where is symmetric in . The coordinates induce a canonical coordinate system on , denoted by and defined by
(5.3) |
for in (5.2). Since the coefficients do transform like a covector, as indicated in Lemma 5.2, it will cause no ambiguity to retain as canonical coordinates on . As in classical geometric mechanics [1, 38], we still call the coordinates the conjugate momenta. And we shall call the second-order coordinates the conjugate diffusivities.
The pairing of and the second-order vector field in (2.7) is then
It follows from (5.1) and (2.8) that for smooths functions and on ,
More generally, for 1-forms and with local expressions and , the symmetric product has local expression
(5.4) |
Dual to the tangent case, there is indeed a canonical bundle epimorphism , given by
In particular . In local coordinates, reads as
The map is well defined since is a covector. Clearly, is also a surjective submersion, so that is a fiber bundle over . Occasionally, we will use the notation to indicate the base manifold .
However, there is no canonical bundle monomorphism from to which is a left inverse of and linear in fiber. We call such a bundle epimorphism a fiber-linear bundle injection from to . Similarly to Proposition 2.11, we also have a connection correspondence property. Namely, if we are given a linear connection on , then it induces a fiber-linear bundle injection from to by
(5.5) |
or in local coordinates . Any fiber-linear bundle injection from to induces a torsion-free linear connection on .
Denote by the subbundle of consisting of all -tensors on . Then the symmetric product , when restricting to , is a bundle monomorphism whose image is the kernel of . Conversely, still by the connection correspondence, a linear connection induces a fiber-linear bundle epimorphism from to which is a right inverse of and is given by
We introduce the -dependent coordinates by for in (5.2), i.e.,
(5.6) |
Then and in particular
The coordinates form a coordinate system on , which we call the -canonical coordinate system. The coordinates also form a coordinate system on when restricted to it. We will call the coordinates the tensorial conjugate diffusivities.
To sum up, we have the following short exact sequence which is split when a linear connection is provided:
(5.7) |
It is easy to check that the bundle homomorphisms , , and are dual to , , and in (2.13), (2.14), (2.15) and (2.16), respectively, so that the short exact sequence (5.7) is dual to (2.17). Similarly to (2.18), we have the following decomposition if a linear connection is given,
with fiber-wise isomorphism and first direct sum , which is given by
In particular,
(5.8) |
Similarly to the classical cotangent space, the second-order cotangent space may be defined via germs. To be precise, we denote by the set of all germs of smooth functions at , and define a equivalence relation between germs: are equivalent if and only if they have the same Taylor expansion at higher than order zero and up to order two. Then, one can easily check that there is a one-to-one correspondence between and the quotient space of by this equivalence relation. Along this way, we can also observe the following diffeomorphism:
(5.9) |
by mapping to , where is the classical second-order jet bundle of . This is similar to is diffeomorphic to the first-order jet bundle (e.g., [32, Example 2.5.11 ] or [77, Example 4.1.15 ]). We denote the natural projection maps from to and from to by and , respectively.
The relations and projection maps are integrated into the following commutative diagram:
Remark 5.4.
(i). As in Remark 3.5, given a linear connection , we can obtain a one-to-one correspondence between and by mapping to . One can find in [19] an application of the jet-like structure on and higher-order bundles to Martin Hairer’s theory of regularity structures [36].
(ii). As we have seen, the product is the model bundle of the stochastic jet space , while the product is diffeomorphic to the second-order jet bundle . So, in a way, we can say that the “stochastic” and the “second-order” are dual to each other. This stochastic–second-order duality is somehow analogous to the particle–wave duality in quantum mechanics.
5.2 Second-order tangent and cotangent maps
Definition 5.5 (Second-order tangent and cotangent maps, [25, Chapter VI]).
Let and be two smooth manifolds, be a smooth map. The second-order tangent map of at is a linear map defined by
The second-order cotangent map of at is a linear map dual to , that is,
The restrictions of to coincide with the classical tangent map . But this is not the case for when restricting to , since for , is still a linear map on . A manifestation of these phenomena may be seen through local coordinates in the following lemma.
Lemma 5.6.
Let and be local coordinate charts around and , respectively. If
Then
Now if , then all vanish and thereby so do ’s. Thus, . This makes clear that . But if , then ’s vanish and
while . Hence, .
Definition 5.7 (Second-order pushforwards and pullbacks).
Let be smooth map. The second-order pushforward by is a bundle homomorphism defined by
Given a second-order form on , the second-order pullback of by is a second-order form on defined by
Let be a diffeomorphism. The second-order pullback by is a bundle isomorphism defined by
Given a second-order vector field on , the second-order pushforward of by is a second-order vector field on defined by
Clearly, is the usual pushforward, but . The following properties are straightforward.
Lemma 5.8.
Let , be two smooth maps. Let be a second-order vector field on and be two smooth functions on .
(i) .
(ii) If is a diffeomorphism, then .
(iii) , .
5.3 Mixed-order tangent and cotangent bundles
In this section, we will extend the notions of the previous two sections to the product manifold .
Definition 5.9.
The mixed-order tangent bundle of is the product bundle ([77, Definition 1.4.1]) . The mixed-order cotangent bundle of is the product bundle . A section of the mixed-order tangent or cotangent bundle is called a mixed-order vector field or mixed-order form, respectively.
The mixed-order tangent and cotangent bundles are dual to each other. The mixed-order tangent (or cotangent) bundle is the bundle that mixes the first-order tangent (or cotangent) bundle in time and the second-order one in space (this is why we use the terminology “mixed-order”). It also matches the fundamental principle of stochastic analysis, whose Itô’s logo is .
For an -valued diffusion with (time-dependent) generator , we call the operator its extended generator. This extended generator is a mixed-order vector field on . Also note that the extended generator of can be characterized by the property that for every , the process
is a real-valued continuous -martingale. In general, a mixed-order vector field has the following local expression:
To give an example of mixed-order forms, we consider a smooth function on , and define in local coordinates
Then is a mixed-order form, and we call it the mixed differential of . Clearly, the pairing of the mixed differential and a mixed-order vector field is .
Given a bundle homomorphism from , we define its mixed-order tangent map at by
Its mixed-order cotangent map at is defined as the linear map dual to . If, moreover, is a bundle isomorphism, its mixed-order pushforward and pullback, denoted by and , respectively, can be defined in a similar manner to Definition 5.7. We leave their detailed but cumbersome definitions and properties to Appendix A.1.
6 Stochastic Hamiltonian mechanics
6.1 Horizontal diffusions
In this section, we consider a general fiber bundle over a manifold , with fiber dimension . We first introduce a special class of diffusions on this fiber bundle, which we call horizontal diffusions. They are defined in a similar fashion as the horizontal subspaces in Definition 3.7. Roughly speaking, a horizontal diffusion process on is a diffusion that is random only “horizontally”, but not on fibers.
Definition 6.1 (Horizontal diffusions on fiber bundles).
Let be a fiber bundle. A -valued diffusion process is said to be horizontal, if there exists an -valued diffusion process and a smoothly time-dependent section of , such that a.s. for all .
The process in the above definition is just the projection of , for a.s.. Since the projection map is smooth, is still a diffusion process.
Now we are going to define a subclass of “integral processes” for second-order vector fields on by making use of horizontal diffusions. We use for an adapted coordinate system on (see [77, Definition 1.1.5]), where we use Greek alphabet to label the coordinates of fibers.
Given a second-order vector field with local expression
(6.1) |
where are smooth functions in the local chart of , by a horizontal integral process of in (6.1) we mean an -valued horizontal diffusion process such that is an integral process of in the sense of (2.22), that is, it is determined by the system
(6.2) |
where the expression means that the family of coordinate functions acts on , and so on. Set for some time-dependent section of and -valued diffusion . Denote . By Itô’s formula, the system (6.2) can be written as
(6.3) |
If has full support for all , then the last three equations in (6.3) translate into a system of (possibly degenerate) parabolic equations on ,
(6.4) |
Therefore, under suitable assumptions for the coefficients , equation (6.4) is solvable, at least locally, by some time-dependent local section over a time interval . Then, plugging into the first two equations of (6.3), we can find and hence . We call an projective integral process of .
6.2 The second-order symplectic structure on and stochastic Hamilton’s equations
It is well known that the classical cotangent bundle has a natural symplectic structure, given by the canonical symplectic form , where are the canonical local coordinates on induced by local coordinates on . Clearly is closed, because it is exact as , where is called the Poincaré (or tautological) 1-form.
Now we need to define a similar structure on the second-order cotangent bundle , which is a second-order counterpart of the symplectic structure. Firstly, we adapt the coordinate-free definition of the tautological 1-form to the second-order case.
Definition 6.2.
The second-order tautological form is a second-order form on defined by
Under the induced coordinate system on defined in (5.3), the second-order tautological form has the following coordinate representation
(6.5) |
We introduce the canonical second-order symplectic form on by writing . Although we do not define the exterior differential for second-order forms, we can still take formally on both sides of (6.5), using Leibniz’s rule and the composition rule (cf. [66, Section 6.(e)]), and forcing and . Then, we get
(6.6) |
We call the pair a second-order symplectic manifold. The complete axiom system for a second-order differential system is beyond the scope of this paper.
Remark 6.3.
In the formal expression , , the two differential operators at LHS are different. The second is still de Rham’s exterior differential on , while the first needs to be understood as the exterior differential on by regarding the first differential as a function on . Thus the complete expression should be . Along this way, the differential operator can be extended to a linear transform that maps 1-forms to 2nd-order forms and satisfies Leibniz’s rule, see [25, Theorem 7.1]. We shall denote the linear operator extended from by d in order to distinguish. In local coordinates, it acts on a 1-form by , so that and . When a linear connection is specified, which covers (5.8).
As in the classical case, we have the following property for the second-order tautological form.
Lemma 6.4.
The second-order tautological form is the unique second-order form on with the property that, for every second-order form on , .
Proof.
Recall that, in Definition 5.7, we have defined the second-order pullbacks of second-order forms. Now, given a smooth map and a second-order 2-form on , we may also define the second-order pullback of by by allowing to be exchangeable with the symmetric product as well as the wedge product . Then, as a corollary of Lemma 6.4, we have
Definition 6.5.
Let and be the canonical second-order symplectic forms on and , respectively. A bundle homomorphism is called second-order symplectic or a second-order symplectomorphism if .
Theorem 6.6.
Let be a diffeomorphism. The second-order pullback by is second-order symplectic; in fact , where is the second-order tautological form on .
Proof.
For , and ,
where we used the fact that in the fourth line. ∎
Clearly, the counterparts of Hamiltonian vector fields on are now second-order vector fields on . Remark that for a second-order vector field on , the form take values in the cotangent bundle .
Definition 6.7.
Let be a given smooth function. A second-order vector field on satisfying
(6.7) |
is called a second-order Hamiltonian vector field of . We call the triple a second-order Hamiltonian system. The function is called the second-order Hamiltonian of the system.
According to (6.7), the 2nd-order vector field satisfies
(6.8) |
The condition (6.7) cannot uniquely determine . It is easy to verify that is of the general form
(6.9) |
where the coefficients are smooth functions on local chart satisfying
such that the local expression (6.9) is invariant under the canonical change of coordinates on induced by a change of coordinates on , governed by the structure group in Lemma 5.2.
Given such a second-order Hamiltonian vector field of , its horizontal integral process is a -valued horizontal diffusion X determined by the following MDEs on ,
(6.10) |
or, in coordinates,
where are canonical coordinates on . The first and third equations has been conjectured in [89] as stochastic Hamilton’s equations in the Euclidean space, since they have the same form as classical Hamilton’s equations (e.g., [1, Proposition 3.3.2]) except that mean derivative replaces classical time derivative.
At first glance, one may think that the system (6.10) is underdetermined, as there are fewer equations than unknowns (the number of unknowns is equal to the fiber dimension of ). Besides, we haven not yet given (6.10) initial or terminal data. These will become clear after we make the following observations. Firstly, the first two equations of (6.10) constitute MDEs that are equivalent to an Itô SDE for in weak sense, as we have seen in Section 2.4. So should be assigned an initial value, say,
(6.11) |
where is a given probability measure on . Secondly, in the third and fourth equations of (6.10), only the “drift” information of and is clear. To overcome the lack of information, we need to assign and terminal values, say,
(6.12) |
where is a given second-order form. Therefore, the third and fourth equations are understood as backward SDEs, whose drifts rely on diffusion coefficients via the last equation. The system (6.10) together with boundary values (6.11) and (6.12) could be understood as a (coupled) forward-backward system of SDEs [87] (where “backward” is taken in a different sense from ours in Chapter 2).
Notice that those forward-backward SDEs are not necessarily solvable (see [87, Proposition 7.5.2] for an example). In order to solve (6.10)–(6.12), we have to take the horizontal condition into consideration, and make some compatibility assumption. More precisely, we set and
(6.13) |
for some time-dependent second-order form on , and denote and , so that . Assume that for each , has full support. Then, by applying Itô’s formula, in the same way as in (6.4), the system (6.10) reduces to
(6.14) |
Next, by taking partial derivative on both sides of the first equation of (6.14) and comparing with the next two, we find the following sufficient condition for the last two equations of (6.14):
(6.15) |
or equivalent, for the terminal value ,
(6.16) |
Equation (6.15) implies that in (6.13) is “exact”, in the sense that for the time-dependent 1-form , where d is the extended differential operator defined in Remark 6.3. Similarly, equation (6.16) implies that for 1-form . The second equality of (6.15) (or (6.16)), called Onsager reciprocity or Maxwell relations [1, Section 5.3], implies that the 1-form (or ) is closed. We will refer to equation (6.15) or (6.16) as second-order Maxwell relations.
Under the 2nd-order Maxwell relations, the original stochastic Hamilton’s system (6.10) turns to the following MDE-PDE coupled system.
(6.17) |
The boundary values in (6.11) and (6.12) now read
(6.18) |
We first use the terminal value in (6.18), which satisfies (6.16), to solve the last two PDEs in (6.17). This gives and hence the 2nd-order form . Then we plug and into the first two MDEs and solve them with initial distribution in (6.18). This yields in law the -valued diffusion as a projective integral process of .
We call system (6.10) or (6.17) the stochastic Hamilton’s equations (S-H equations in short). The second-order Maxwell relations are sufficient for the component of in (6.13) to solve the last two equations of (6.10), so we refer to it as an integrability condition of (6.10). When restricting settings to Riemannian manifolds, the S-H equations (6.10) can be simplified to a global Hamiltonian-type system on , as we will see in Subsection 7.4.2.
Lemma 6.8.
Let be a time-dependent 2nd-order Hamiltonian, and be a horizontal integral process of . Then, the total mean derivative of along is
Proof.
In particular, when is time-independent, we have
(6.19) |
which is also a consequence of (6.8). Equivalently, is harmonic with respect to the horizontal integral process . In this case, we can say that is stochastically conserved, or is a stochastic conserved quantity. In particular, the expectation is a constant.
6.3 Two inspirational examples
Let be a Riemannian manifold with Riemannian metric . Assume for simplicity that is compact. Let be the Levi-Civita connection on with Christoffel symbols . In this section, we will consider two types of processes on , to provide some intuition of our stochastic Hamiltonian formalism.
6.3.1 Diffusion processes on Riemannian manifolds
Consider a second-order Hamiltonian on with the following coordinate expression:
(6.20) |
where is a given smooth vector field on and a smooth function on . One can easily verify that the expression at RHS of (6.20) is indeed invariant under changes of coordinates. We consider the S-H equations (6.17) subject to boundary conditions and , where is a given probability distribution and a given smooth function on .
By the first two equations of system (6.17), the projection diffusion satisfies the following MDEs,
(6.21) |
subject to the initial distribution ; or equivalently (according to the end of Section 2.4), it can be rewritten as the following Itô SDE in weak sense,
(6.22) |
where is the positive definite square root -tensor of , i.e., , denotes an -valued standard Brownian motion. Note that equations (6.21) are independent of coordinates , so they form a closed system on the base manifold and can be solved independently. Indeed, the solution is a diffusion on with generator .
Now we consider the last two equations of (6.17). The LHS of the third equation reads
where denotes the pairing of vectors and covectors, is the Laplace-Beltrami operator and the gradient, with respect to . In order to find the solution of the third equation of (6.17), we consider the following linear backward parabolic equation (where “backward” has a meaning different from that in Section 2.2)
(6.23) |
with terminal value . We let
(6.24) |
and use (6.23) and (6.15) to derive
(6.25) |
which agree with the third equation of (6.17).
Example 6.9 (Brownian motions).
When and , the 2nd-order Hamiltonian is , the solution process is a standard Brownian motion on with initial distribution . Such 2nd-order Hamiltonian can be regarded as a “stochastic deformation” of the trivial classical Hamiltonian . Indeed, is the -canonical lift of that will be defined in forthcoming Section 6.6. Therefore, we may regard Brownian motions as “stochastization” or “stochastic deformation” of trivially constant curves on the base manifold .
We are going to describe in the next example a dynamical approach to diffusions, elaborated afterwards (Section 7.3), inspired by Schrödinger.
6.3.2 Reciprocal processes and diffusion bridges on Riemannian manifolds
With the same coefficients and boundary data in Subsection 6.3.1, we consider the S-H system (6.17) with the following second-order Hamiltonian on :
(6.26) |
subject to boundary conditions and . Here, and are called, respectively, vector and scalar potentials in classical mechanics. Again, it is easy to verify that the expression at RHS of (6.26) is indeed invariant under changes of coordinates.
The LHS of the third equation in (6.17) now reads
In order to find the solution of the third equation of (6.17), we first consider the positive solution of following backward parabolic equation on
(6.27) |
with terminal value , where denotes the Riemannian inner product with respect to . If we let , then it is easy to verify that satisfies the following Hamilton-Jacobi-Bellman (HJB) equation
(6.28) |
with terminal value , where denotes the Riemannian norm with respect to . Now we let
(6.29) |
and use (6.28) and (6.15) to derive, in a way similar to (6.25),
which agree with the third equation of (6.17). Therefore, the projection diffusion of the system (6.17) satisfies the following MDEs,
(6.30) |
subject to the initial distribution ; or equivalently (according to the end of Section 2.4), it can be rewritten as the following Itô SDE in weak sense,
(6.31) |
where is the positive definite square root -tensor of , i.e., , denotes an -valued standard Brownian motion.
The solution process of (6.31) is called a Bernstein process [11, 18] (or the reciprocal process derived from the -valued diffusion in (6.22) [46]). The time marginal distribution of satisfies a Born-type formula (see, e.g., [88, Corollary 3.3.1] or [17, Equations (2.9), (4.6) and (4.8)]), where satisfies the adjoint equation of (6.27). The terminal law of can be determined in the following way: we first solve (6.27) to get , and then find out the initial value for via and solve the equation for to get , finally the terminal law of is given by . In particular, when and for , the solution of (6.31) is the Markovian bridge of the diffusion conditioning on ending point [13].
Remark 6.10.
(i). The derivation of the reciprocal process (6.31) from the diffusion (6.22) was the way chosen by Jamison [46], inspired by Schrödinger’s original problem [79]. No geometry or dynamical equations like HJB equation (6.28) was involved by him. Like here, Jamison’s construction was involving only the past (nondecreasing) filtration. The dynamical content dates back to [88, 17, 15], where a reciprocal process was constructed from the only data of a Hamiltonian operator as required by Schrödinger’s original problem, and the future (nonincreasing) filtration was also used to study the time-reversed dynamics. Cf. also Example 6.12 and Section 7.3.
(ii). Equations (6.30) suggest that the transformation from coordinates to coordinates is not invertible. More precisely, the coordinates are transformed from but the coordinates are only related to . Besides, these two equations have nothing to do with the coordinates . However, if we look at the -canonical coordinates for (6.30), then
which indicates that the transform from to is invertible. These will help us establish stochastic Lagrangian mechanics and second-order Legendre transforms, in forthcoming Chapter 7.
(iii). As observed in Section 2.2, every result presented here has a backward version (in the sense of backward mean derivatives with respect to the future filtration ). Indeed, two forward-backward SDE systems for Bernstein diffusions on Euclidean space were derived in [16]: one is under the past filtration and coincides with ours, whereas the other one is under the future filtration.
There are some special cases which are of independent interests and have been considered in the literature.
Example 6.11 (Brownian (free) reciprocal processes and Brownian bridges).
Consider the case where , . In this case, is a Brownian motion on , so we call a Brownian reciprocal process. In particular, the Brownian bridge from to of time length is driven by the Itô SDE (6.31) where , and satisfies the backward heat equation (6.27) with and final value . See also [40, Theorem 5.4.4]. Thus, Brownian bridges are understood as stochastic Hamiltonian flows of the 2nd-order Hamiltonian , compared with geodesics as Hamiltonian flows of the classical Hamiltonian (cf. [1, Theorem 3.7.1]). Here, the 2nd-order Hamiltonian is the -canonical lift of . We can also say that Brownian bridges are “stochastization” or “stochastic deformation” of geodesics, cf. Example 6.9. Relations between geodesics and Brownian motions have attracted many studies. For example, one can find various interpolation relations between geodesics and Brownian motions in [4, 61].
Example 6.12 (Euclidean quantum mechanics [15, 2, 3]).
It is insightful to consider the case and . The Riemannian metric under consideration is the flat Euclidean one. To catch sight of the analogy with quantum mechanics, we involve the reduced Planck constant into the second-order Hamiltonian of (6.26), so that
The system (6.10) then reads as
Note that the first three equations form a sub-system and can be solved separately, as they are independent of the coordinates ’s. Equation (6.27) and its adjoint now reduce to the following -dependent backward and forward heat equations, respectively,
which together with the Born-type formula display the strong analogy to quantum mechanics [88]. The function solves the following -dependent HJB equation
The first three equations then can be solved by letting . The first and third equations imply a Newton-type equation
(6.33) |
This is indeed the equation of motion of the Euclidean version of quantum mechanics, which was the original motivation of Schrödinger in his well-known problem to be discussed in Section 7.3. See [15, p. 158] and [89, Eq. (4.17)] for more. Note that [15, 89] used to denote the physical scalar potential and used the relation and to formulate the HJB equation from backward heat equation in the case of nondecreasing (past) filtration.
There are two special cases of which more will be studied later.
(i). When and , i.e., , we call its projective integral process the (forward) stochastic harmonic oscillator. It is a stochastization of the classical harmonic oscillator with Hamiltonian [1, Example 5.2.3]. Likewise, here is the canonical lift , see Section 6.6.
(ii). When and , i.e., , we call it the (forward) Euclidean harmonic oscillator.
6.4 The mixed-order contact structure on
In the later sections we will investigate time-dependent systems. The proper space for consideration is now . Recall in (5.9) that , where the latter is the second-order jet bundle of .
In classical differential geometry, the first-order jet bundle can be equipped with an exact contact structure in several ways [1, Section 5.1]. Among others, the canonical symplectic form on corresponds to a contact structure on via , which is indeed exact as for . Another commonly used contact structure is the Poincaré-Cartan form for a given function . It is also exact as where . The advantage of the Poincaré-Cartan form, compared with the contact form , is that it can be related to the (time-dependent) Hamiltonian vector field on of . More precisely, the vector field , treated as a vector field on and called the characteristic vector field of , is the unique vector field satisfying and .
Now we proceed in a similar way for the second-order jet bundle . Define
Then . We call the pair a second-order contact manifold and the pair a mixed-order exact contact manifold. In local coordinates, has the same expression as in (6.6), but we stress that it is now a second-order form on . The form has the local expression
This makes clear that is a mixed-order form on .
A time-dependent second-order Hamiltonian is a smooth function on . The second-order Hamiltonian vector field of is now a time-dependent second-order vector field on , its horizontal integral process share the same equations as (6.10) or (6.17), only with explicitly depending on time. Define a mixed-order vector field on by
where is a second-order Hamiltonian vector field of the form (6.9). We call the extended second-order Hamiltonian vector field of .
We define the second-order counterpart of Poincaré-Cartan form by
and call it the mixed-order Poincaré-Cartan form on . It is exact in the sense that , where .
The following lemma gives the relations between and .
Lemma 6.13.
The class of extended second-order Hamiltonian vector fields is the unique class of mixed-order vector fields on satisfying
Proof.
Firstly, we show that satisfies the two equalities. The second equality is trivial. For the first one, we pick a mixed-order vector field on ; then,
To prove the uniqueness, it suffices to show that any mixed-order vector field on satisfying is a multiplier of . Suppose that has the local expression
Then, it follows that
The vanishing of each coefficient gives
Therefore, . ∎
6.5 Canonical transformations and Hamilton-Jacobi-Bellman equations
Let us study the second-order analogs of canonical transformations and their generating functions. To do so, we need to find a change of coordinates from to that preserves the form of stochastic Hamilton’s equations (6.10) (with time-dependent 2nd-order Hamiltonian). More precisely, we have the following definition of canonical transformations between mixed-order contact structures, which is adapted from those between classical contact structures in [9].
Definition 6.14.
Let and be two second-order contact manifolds corresponding to second-order tautological forms and . A bundle isomorphism is called a canonical transformation if its projection is a bundle isomorphism from to projecting to , and there is a function such that
(6.34) |
where .
The map in the definition is also a bundle isomorphism from to projecting to . Hence, we may assume for all , where is a smooth map from to . For each , we define a map by . We also introduce an injection by . Then, we have .
Lemma 6.15.
The map is second-order symplectic for each if and only if there is a mixed-order form on such that
In particular, condition (6.34) implies that each is a second-order symplectomorphism.
Proof.
The sufficiency follows from
For the necessity, we observe that
So we can write , where is a mixed-order form which does not involve . This leads to . The result follows. ∎
The following lemma gives some equivalent statements to the condition (6.34).
Lemma 6.16.
Condition (6.34) is equivalent to the following:
(i) is mixed-order closed;
(ii) for all , ;
(iii) for all , ;
where .
Proof.
Definition 6.17.
Let be canonical. If we can locally write
(6.35) |
for , then we call a generating function for the canonical transformation .
We use for local coordinates on and for those on . Recall that . Then, using (A.4), the relation (6.35) reads in coordinates as
Balancing the coefficient of , we get
By Lemma 6.16, the new Hamiltonian function after transformation is related to the old Hamiltonian by . Let us further assume that we can choose coordinates in which and are independent, so that the independent variables in (6.35) are . Then, relation (6.35) means
(6.36) |
which implies that the generating function of the canonical transformation satisfies
(6.37) |
The expressions for and are due to the mixed differential term in , and correspond to the relation (6.15).
Remark 6.18.
Unlike the canonical transformations of classical Hamiltonian systems which have four types of generating functions related via classical Legendre transform (see [35, Section 9.1]), here we can only have the type using as independent variables but not others. This can be attributed to the ill-behaveness of the 2nd-order analog of Legendre transform, as indicated in Remark 6.10.(iii). However, if the configuration space is a Riemannian manifold, stochastic Hamiltonian mechanics can be simplified to share the same phase space as classical Hamiltonian mechanics, so that we can also have four types of generating functions. See Subsection 7.4.2 for details and examples of canonical transformations.
The Hamilton-Jacobi-Bellman (HJB) equation can be introduced as a special case of a time-dependent canonical transformation (6.37). In the case where and the new Hamiltonian vanishes formally, we denote by the corresponding generating function . It follows from (6.37) that solves the Hamilton-Jacobi-Bellman equation,
(6.38) |
We will refer to equation (6.38) as the HJB equation associated with second-order Hamiltonian , and a solution of (6.38) as a second-order Hamilton’s principal function of .
More generally, we have
Theorem 6.19.
Let be a second-order Hamiltonian vector field on and let . Then, the following statements are equivalent:
(i) for every -valued diffusion satisfying
the -valued process is a horizontal integral process of ;
(ii) satisfies the Hamilton-Jacobi-Bellman equation
(6.39) |
for some function depending only on .
Proof.
Let and set , , . Then
(6.40) |
These imply that the last equation of the system (6.17) holds. Since
the first two equations in (6.10) or (6.17) hold. Hence, to turn the process into a horizontal integral process of , it is sufficient and necessary to make sure that the third equation in (6.17) holds. Plugging the first equation of (6.40) into the third equation, it reads as
A straightforward reinterpretation yields
The result follows. ∎
Remark 6.20.
If solves the HJB equation (6.39), then solve (6.38) with a primitive function of . As a matter of fact, one can always integrate the time-dependent function into the 2nd-order Hamiltonian function such that the HJB equation (6.39) has the same form as (6.38). More precisely, if we let , then Theorem 6.19 also holds with and zero function in place of and , respectively. A similar argument holds for S-H equations (6.10). Indeed, adding a function depending only on time to a 2nd-order Hamiltonian does not change its S-H equations.
Example 6.21.
The function considered in Section 6.3 satisfies the Hamilton-Jacobi-Bellman equation (6.28), which is exactly with the second-order Hamiltonian given in (6.26). Hence, this theorem yields that the process is a horizontal integral process of , which coincides with (6.32). The Euclidean case for such argument has been discovered in [15, p. 180] or [89, Eq. (4.20)].
6.6 Second-order Hamiltonian functions from classical ones
In the presence of a linear connection on , we are able to reduce (or produce) second-order Hamiltonian functions to (from) classical ones.
Let be given a second-order Hamiltonian function . We make use of the fiber-linear bundle injection in (5.5) to define a classical Hamiltonian by
(6.42) |
In canonical coordinates, it maps as . If we introduce a family of auxiliary variables by
(6.43) |
Then, we can write
We say reduces to under the connection , or is the -reduction of .
Clearly, the way to lift from a classical Hamiltonian to a second-order Hamiltonian function that reduces to under is not unique. But there is a canonical reduction when we are provided with a symmetric -tensor field (not necessarily Riemannian), given by
(6.44) |
Then, is the -reduction of , and
(6.45) |
We call the -canonical lift of . If is a Riemannian metric and is the associated Levi-Civita connection, then we simply call the -canonical lift of . If there is a classical Hamiltonian such that the second-order Hamiltonian is the - (or -) canonical lift of , we say is - (or -) canonical.
As an example, the second-order Hamiltonian of (6.26) is -canonical and reduces to .
Furthermore, for the canonical transformation in Definition 6.14, we can reduce its associated function to a classical function via (6.42). As a consequence of (6.34), the projection of , i.e., the map satisfies where . It follows that is a classical canonical transformation [1, Definition 5.2.6].
We will go back to this issue in Section 7.4 where the second-order Legendre transform will be developed. In particular, we will show there that for the canonical 2nd-order Hamiltonian in (6.44), the corresponding 2nd-order Hamilton’s equations (6.17) can be rewritten on the cotangent bundle in a global fashion, see Theorem 7.22.
7 Stochastic Lagrangian mechanics
In this chapter, we specify a Riemannian metric for the manifold , and a -compatible linear connection . Note that such and always exist but are not unique in general.
We will denote by and the Riemannian norm and inner product, respectively. Also, denote by the inverse metric tensor of , and the Christoffel symbols of . We observe that is a -tensor field. Denote by the Riemann curvature tensor and the Ricci -tensor.
7.1 Mean covariant derivatives
Definition 7.1 (Vector fields and 1-forms along diffusions).
Let be diffusion on . By a vector field along , we mean a -valued process , such that for all . Similarly, by a 1-form along , we mean a -valued process , such that for all .
Clearly, for a time-dependent vector field on , the restriction of on , i.e., , is a vector field along . In this case, we call a vector field restricted on . In this way, vector fields restricted on are just -valued horizontal diffusions projecting to . Similarly for 1-forms.
Definition 7.2 (Parallelisms along diffusions).
Let . A vector field along is said to be parallel along if the following Stratonovich SDE in local coordinates holds,
(7.1) |
A 1-form along is said to be parallel along if
Definition 7.3 (Stochastic parallel displacements).
Given a diffusion and a (random) vector , the stochastic parallel displacement of along is the extension of to a parallel vector field along , that is, satisfies the SDE (7.1) with initial condition . We denote and . The stochastic parallel displacement of a (random) covector along is defined in a similar fashion.
Definition 7.4 (Damped parallel displacements).
Let . Given a (random) vector and covector , the damped parallel displacement of along is the extension of to a vector field along that satisfies the SDE
(7.2) |
The damped parallel displacement of along is the extension of to a vector field along that satisfies the SDE
(7.3) |
We denote , , and , .
If and are restrictions on , that is, and , then equations (7.2) and (7.3) can be rewritten, respectively, as
where we mean by the 1-form . The Stratonovich stochastic differentials can be transformed into Itô ones. For example, (7.3) is equivalent to
(7.4) |
Remark 7.5.
Lemma 7.6.
Proof.
We only prove Assertion (ii), as (i) is similar. Since Stratonovich stochastic differentials obey Leibniz’s rule, we have
This proves the first statement of (ii). The second statement of (ii) follows by letting . ∎
Definition 7.7 (Mean covariant derivatives along diffusions).
Given a diffusion on . Let and be time-dependent vector field and 1-form along , respectively. The (forward) mean covariant derivative of with respect to is a time-dependent vector field along , defined by
(7.5) |
The damped mean covariant derivative of with respect to is a time-dependent vector field along with instead of in (7.5). Similarly, we can define and .
Lemma 7.8.
(i). Let and be vector field and 1-form along . If is parallel along , then
(7.6) |
If satisfies the SDE (7.3), then (7.6) holds true with instead of .
(ii). Let be a vector field restricted on . Then
(iii). Let be a 1-form restricted on . Then
(iv). Let and be a vector field and a 1-form restricted on . Then
Proof.
(i). By Lemma 7.6.(i), we have
This proves the first statement of (i). The second statement of (i) follows by a similar argument with in place of and in place of .
(ii). It suffices to derive the expression for . Suppose that is a 1-form satisfying the SDE (7.3) and the diffusion satisfies . Then, we apply Itô’s formula to and make use of (2.20) and (7.4). We get
Hence, the result (i) implies
The arbitrariness of yields (ii).
(iii). Similar to (ii).
If , then
and similarly,
(7.7) |
where is the connection Laplacian, and is the Laplace-de Rham operator on forms. The relation is due to the Weitzenböck identity [74, Theorem 9.4.1]. We remark here that the operator acting on vector fields is also called Laplace-de Rham operator in [21].
In the context of fluid dynamics, the operator , with a vector field, is often referred to as material derivative or hydrodynamic derivative. So the mean covariant derivative and its damped variant can be regarded as stochastic deformations of material derivative.
7.2 A stochastic stationary-action principle
In this section, we will establish a type of stochastic stationary-action principle: the stochastic Hamilton’s principle. Another version for systems with conserved energy, the stochastic Maupertuis’s principle, can be found in Appendix C.
In contrast to second-order Hamiltonians, not all real-valued functions on can be used as second-order Lagrangians in stochastic Lagrangian mechanics. This has been hinted in Section 6.3, as we have mentioned in Remark 6.10. For this reason, we will produce a class of second-order Lagrangians from classical Lagrangians, via the fiber-linear bundle projection in (3.3) and the -canonical coordinates in (3.2).
Definition 7.9.
By an admissible second-order Lagrangian, we mean a function such that there exists a classical Lagrangian satisfying . We call the -lift of .
In local coordinates, the -lift of is expressed as
(7.8) |
Let . Our stochastic variational problem consists in finding the extrema (maxima or minima) of the stochastic action functional
(7.9) |
over a suitable domain of diffusions on , where is an admissible second-order Lagrangian lifted from .
In order to formulate a well-posed stochastic variational problem in an economical way, we assume that the manifold is compact and the metric is geodesically complete (which will be used to characterize the variations of diffusions in Lemma 7.13), and that the connection is the associated Levi-Civita connection. The geodesic completeness can be ensured, for example, if is connected (see, e.g., [55, p. 346]). Whenever the metric is given, the associated Levi-Civita connection is uniquely determined, due to the fundamental theorem of Riemannian geometry [50, Theorem IV.2.2]. We will refer to such a geodesically complete Riemannian metric as a reference metric tensor.
For a fixed point and a probability distribution on , we define an admissible class of diffusions by
(7.10) |
where denotes the set all -valued diffusion processes starting from at and with final distribution , i.e., .The action functional is now defined on the set , that is, .
Note that the admissible class is similar to the Wiener space, so that a candidate for its “tangent space” is Cameron–Martin space. Denote by the Hilbert space of absolutely continuous curves such that . Let be the subspace consisting of all satisfying .
Definition 7.10.
Let . For a curve , the vector field along given by is called a tangent vector to at . The tangent space to at is the set of all such tangent vectors, that is,
Definition 7.11.
By a variation (or deformation) of a diffusion along , we mean a one-parameter family of diffusions , where for each , satisfies the following ODE
(7.11) |
The diffusion is called a stationary (or critical) point of , if the first variation vanishes at , i.e.,
(7.12) |
Remark 7.12.
(i). The variations of diffusions on manifolds, via differential equation (7.11), is standard in stochastic analysis on path spaces of Riemannian manifolds. See for example [22, Eq. (2.3)] and [39, Theorem 4.1], where it is shown that Wiener measure is quasi-invariant under such variations. This kind of variations has some equivalent constructions. For instance, the previous two references also provided an approach by lifting to the frame bundle and projecting to the Euclidean space (a stochastic analog of Cartan’s development), while Malliavin and Fang [27] provided an alternative perspective via Bismut connection.
The following lemma is the key for establishing stochastic Hamilton’s principle. The first statement shows that the variation is well defined on the path space . The second one describes the infinitesimal changes of with respect to the variation parameter . The proof of the latter is based on a geodesic approximation technique, which is originally due to Itô [44].
Lemma 7.13.
Given and . We have
(i) for each , ; and
(ii) for all ,
(7.13) |
where , is the (classical) covariant derivative with respect to the parameter .
Proof.
(i). Let and be the anti-development ([40, Definition 2.3.1]) of and , respectively, with fixed initial frame . Equivalently, for example, is an -valued diffusion related to by the following SDEs [40, Section 2.3]
Applying the fact that (e.g., [50, Proposition 1.5]) and the condition , we have
(7.14) |
and consequently, . Meanwhile, it follows from [27, Section 3.5] (or [22, Theorem 5.1], [39, Section 3]) that
where is the curvature form on the orthogonal frame bundle , taking values in , and the frame is viewed as an isomorphism from to . It follows that . For the reason similar to (7.14), we have . The result follows. See [22, Theorem 8.3] for a more elaborate proof.
(ii). Fix . Let be a division of the time interval , and let be one of the variation parameter interval . Denote . Consider the polygonal curve , which is an approximation of made of minimizing geodesic segments joining with for all . This is attainable by the geodesic completeness. We will construct an approximation scheme for the variational processes ’s.
For , we construct the approximation of as follows. We extend each , , to a geodesic
Let be the polygonal curve consisting of minimizing geodesic segments joining with for all .
Then, we construct for , , by induction. Suppose , , has been defined. Then, in particular, we have a curve . Extend each , , to a geodesic by
Let be the polygonal curve consisting of minimizing geodesic segments joining with for all . In a similar way, we can define for , .
Now we have a family of polygonal curves , which satisfies and
As for each and , is a geodesic, the vector field
is a Jacobi field along . This leads to the following Jacobi equation
(7.15) |
with boundary values
(7.16) |
Since the connection is torsion-free, we can exchange the covariant derivative and standard derivative to have
(7.17) |
On the other hand, Taylor’s theorem yields
(7.18) |
Combining (7.15)–(7.18), we have
A standard limit theorem yields the result (ii). ∎
Remark 7.14.
(i). The constraint in (7.10) looks strong. A possibly better viewpoint is to force all diffusions under consideration to have the same nondegenerate diffusion tensor , i.e., . Then, the inverse of defines a Riemannian metric , cf. [43, Section V.4]. As can be seen from the first part of the above proof, the constraint of fixing the diffusion tensor is a natural one in the literature of variational calculus on the path space. An intuitive reason for this constraint is to assure that the induced measures are equivalent, which is necessary for equation (7.11) to be well-posed, cf. [22]. The assumption of Levi-Civita connection may be relaxed to that the connection is -compatible and torsion skew symmetric [22, Definition 8.1], in which case the second assertion of this lemma needs to add the effect of torsion.
For a smooth function on , we denote by the differential of with respect to the coordinates . Since , is treated as a 1-form on and
(7.19) |
We call the vertical differential of . Regarding the differential with respect to the coordinates , we introduce the horizontal differential which depends on the connection , by
(7.20) |
It is easy to check that both definitions (7.19) and (7.20) are invariant under change of coordinates. In fact, by the classical theory [77, Section 3.5 and Example 4.6.7], we know that the connection can uniquely determine a -valued 1-form on horizontal over , which is given in local coordinates by
Hence, the horizontal differential is , where is the total differential of . Given a vector field on , is a smooth function on . Then, it is easy to check that
(7.21) |
The following integration-by-parts formula will be used. Its proof is straightforward from definitions of stochastic integrals and mean derivatives, cf. [17, Lemma 4.4].
Lemma 7.15.
Let be a real-valued continuous semimartingale such that exists, let be a real-valued continuous process on , of finite variation. Then
Now we are in position to present the stochastic version of Hamilton’s principle.
Theorem 7.16 (Stochastic Hamilton’s principle).
Let be a regular Lagrangian on . A diffusion is a stationary point of , if and only if satisfies the following stochastic Euler-Lagrange (S-EL) equation
(7.22) |
where is the damped mean covariant derivative with respect to .
We remark that since , the operator in (7.22) is just the one of (7.7). The unknown in (7.22) is the process , so the conditions and , indicated in the assumption , can be regarded as boundary conditions of (7.22).
Proof.
Remark 7.17.
(i). For a special class of Lagrangians in the Euclidean context, the stochastic Euler-Lagrange equation (7.22) has been established in [17, Subsection 5.1] where they called it stochastic Newton equation, see also [89]. For general Lagrangians on Riemmannian manifolds, equation (7.22) is new (to the authors’ best knowledge). See Section 7.3 for discussions of a special case.
(ii). The second author and his collaborator formulated a weak stochastic Euler-Lagrange equation in [52]. They mean by “weak” that their stochastic Euler-Lagrange equation holds in the sense of stochastic integrals. The main differences between their formulation and ours is that we get rid of the stochastic integral (martingale) part in our equation since we use mean derivatives instead of stochastic differentials.
7.3 An inspirational example: Schrödinger’s problem
The inspirational example of stochastic Hamiltonian mechanics presented in Section 6.3 also provides an example of our stochastic Lagrangian mechanics. Consider the following Lagrangian defined on :
(7.25) |
where is a given time-dependent vector field on . It actually relates to the 2nd-order Hamiltonian in (6.26) via the 2nd-order Legendre transform, which will be considered in Section 7.4. For such Lagrangian, we can directly figure out the relation between stochastic Euler-Lagrange equation (7.22) and Hamilton-Jacobi-Bellman equation. We denote by the set all -valued diffusion processes over time interval .
Theorem 7.18 (S-EL & HJB).
Proof.
For a function on , we will denote by the exterior differential of on , i.e., with respect to coordinates . Condition (7.26) can be rewritten in local coordinates as
(7.28) |
Then, it is clear that
(7.29) |
Since , we use Leibniz’s rule to derive
(7.30) |
Now we take the differential with respect to to the HJB equation (7.27). Obviously,
For the second term,
For the third term, we use again . Then, we have
For the fourth term, in the same way we have
Combining these together and applying (7.26)–(7.30) as well as (7.7), we obtain
The result follows. ∎
Remark 7.19.
Theorem 7.18 strongly suggests some relations between stochastic Lagrangian (and also Hamiltonian) mechanics and Schrödinger’s problem in the reintepretation of optimal transport. In the setting of the latter (see, e.g., [18, 58, 59]), there is a given reversible positive measure on the path space , called reference measure, as well as two probability distributions . Schrödinger’s problem aims to minimize the following relative entropy
(7.31) |
over all probability measures on such that are the initial and final time marginal distributions of , i.e., and , where is the time marginal distribution of and is the coordinate mapping. Denote, respectively, by and , the coordinate process under the measure and . Then, Girsanov theorem implies that [56, Theorem 1] a necessary condition for the finite entropy condition is , -a.s.. Furthermore, if is a diffusion measure, i.e., is a diffusion process, then a similar application of Girsanov theorem yields that a necessary condition for is that is also a diffusion measure and there exists a time-dependent vector field such that
The solution of Schrödinger’s problem, i.e., minimizing (7.31), is related to the reference measure by a time-symmetric version of Doob’s -transform [58, Section 3]. Its coordinate process is sometimes called a Schrödinger bridge or Schrödinger process. When the reference measure is Markovian, i.e., the law of a Markov process, the solution process is also called a reciprocal [11, 46] or Bernstein process [18, 16].
If the manifold is endowed with a Riemannian metric , and the reference coordinate process has generator
for some time-dependent vector field on , then the density of the minimizer of (7.31) solves the following Kolmogorov forward equation
(7.32) |
Moreover, an analog of Benamou-Brenier formula was derived (see [58]). Consider the problem of minimizing the average action
(7.33) |
among all pairs , where is is a measurable path in , is a measurable time-dependent vector field and the following constraints are satisfied (in the weak sense of PDEs):
(7.34) |
The relation between in (7.33) and in (7.31) is just that is the time marginal of , namely,
(7.35) |
The minimizer of (7.33) is the pair where solves (7.32) and solves (6.28).
These results are summarized in the following equivalent relations:
(7.36) |
Now if the coordinate process under the reference measure is a nondegenerate -valued diffusion in which is diffusion-homogeneous, then assigning such a reference measure amounts to assigning a pair , where is a positive-definite symmetric -tensor, i.e., a Riemannian metric tensor. More precisely, we let be the generator of . Since is nondegenerate and diffusion-homogeneous, is a time-independent nondegenerate symmetric -tensor field. Let be the inverse of , so that is a Riemannian metric tensor. We then equip the Riemannian manifold with the associated Levi-Civita connection . The isomorphism (2.19) implies that
where is the time-dependent vector field given by , and and are the gradient and Laplace-Beltrami operator with respect to , respectively.
We set that is a diffusion measure and , -a.s., which is a necessary condition for . Then, by (3.4), the generator of is given by
From (7.34) and (7.35), one can see that and the action (7.33) equals to
(7.37) |
So the minimizing problem turns into minimizing the action (7.37) over all diffusion measures with , and , -a.s.. If and , this brings us back to our stochastic variational problem, that is, to minimize the action functional in (7.9) over , with Lagrangian . Note that in this case, since is Dirac, the relative entropy in (7.31) and are always infinite, while their difference can be finite as in (7.36). Moreover, by Theorem 7.16 and 7.18, a necessary condition for to be the minimizer of is that satisfies (7.26) and (7.27), which coincides with (7.32).
Remark 7.20.
(i). Compared to the Lagrangian (7.25) used here for addressing Schrödinger’s problem, there is another type of Lagrangians used in the Euclidean version of quantum mechanics in [17, Eq. (5.4)]. The latter has an additional term of divergence of , which helps to express part of the action functional as a Stratonovich integral. The stochastic Euler-Lagrange equation (7.22) applied to their Lagrangians recovers the equations of motion in [17, Theorem 5.3].
(ii). In the seminal paper [73], F. Otto provided a geometric perspective for numerous PDEs by introducing a Riemannian structure in the Wasserstein space. It is known as Otto’s calculus. A similar idea can ascend to V.I. Arnold, who established a geometric framework for hydrodynamics by studying the Riemannian nature of the infinite-dimensional group of diffeomorphisms [8]. The recent paper [33] formulated Schrödinger’s problem via Otto calculus, where the equation of motion is given by an infinite-dimensional Newton equation, cf. [49, 86] on related matters. All these works can be called a “geometrization” of (stochastic) dynamics. In contrast, the present framework can be called a “stochastization” of geometric mechanics. The difference and relations between our framework and theirs are similar to those between two ways of producing HJ equations for quantum mechanics mentioned in the introduction. More precisely, while (second-order) HJB equations play a key role in our framework, various HJ equations with density-dependent potential terms were derived by them (see [33, Corollary 23] and [49, Proposition 2.4]).
7.4 Second-order Legendre transform
7.4.1 From to and back
Let us fix a linear connection on . Here, for simplicity, we consider time-independent Hamiltonians and Lagrangians.
We first produce second-order Lagrangians from second-order Hamiltonians. To this end, we first reduce the second-order Hamiltonian to a classical one. Given a time-independent second-order Hamiltonian , its -reduction is the classical Hamiltonian , as in (6.42). If is hyperregular (see [1, Section 3.6]), then its fiber derivative , which is given in canonical coordinates by , is a diffeomorphism and defines the classical Legendre transform [1, Section 3.6]:
(7.38) |
where is a family of auxiliary variables introduced in (6.43). Then we lift to an admissible second-order Lagrangian as in Definition 7.9, that is, . Combining (7.38) with (7.8), the relation between and is
(7.39) |
We call (7.39) the second-order Legendre transform. In particular, if we restrict the admissible 2nd-order Lagrangian to the subbundle of with coordinate constraint for some symmetric -tensor field (which is just the condition in (7.10)), and let be -canonical, then by (6.45), we have
(7.40) |
Consequently, we can find the relation between 2nd-order Hamilton’s principal functions and action functionals. By (6.41) and (7.40),
One concludes, from Dynkin’s formula, that for an -valued diffusion ,
and
where is the conditional expectation . These mean that the action functional is the expectation of 2nd-order Hamilton’s principal function (up to an undetermined constant), while the 2nd-order Hamilton’s principal function is the conditional expectation version of action functional.
Conversely, let us be given an admissible 2nd-order Lagrangian which is the -lift of a classical Lagrangian . If is hyperregular, then its fiber derivative
(7.41) |
which is written in coordinates as , is a diffeomorphism and defines the classical inverse Legendre transform:
(7.42) |
We replace coordinates by , due to (3.2). Now, given a symmetric -tensor field , we lift to the -canonical in (6.44). The relation between and is
(7.43) |
where is the tensorial conjugate diffusivities defined in (5.6). We call (7.43) the -canonical inverse 2nd-order Legendre transform. When is Riemannian and is the associated Levi-Civita connection, we call (7.43) the -canonical inverse 2nd-order Legendre transform. In particular, when restricting onto the subbundle of with coordinate constraint , we have
(7.44) |
Following the procedure in classical mechanics [1, Definition 3.5.11], for a given classical Lagrangian , we define a function by , and the classical energy by . Notice that in local coordinates, and .
Example 7.21.
It is easy to check that the -lift of the classical Lagrangian in (7.25) is the second-order Legendre transform of the second-order Hamiltonian in (6.26). And conversely, the latter is the -canonical inverse 2nd-order Legendre transform of the former. The classical energy associated with this Lagrangian is given by
(7.45) |
Each term at RHS corresponds to a kinetic energy, a vector potential energy and a scalar potential energy, respectively.
7.4.2 Stochastic Hamiltonian mechanics on Riemannian manifolds
Given a reference metric tensor , i.e., a geodesically complete Riemannian metric as in Section 7.2, let be the associated Levi-Civita connection. If a 2nd-order Hamiltonian is the -canonical lift of a classical Hamiltonian , namely, as in (6.44), then the stochastic Hamilton’s equations (6.17) can reduce to a simpler Hamilton-type system on , which is exactly equivalent to the stochastic Euler-Lagrange equation (7.22) via the classical Legendre transform (7.41) and (7.42).
Similarly to (7.19) and (7.20), we introduce, for a smooth function on , the vertical gradient and horizontal differential which are given in local coordinates by
Both are invariant under change of coordinates. Still by the classical theory, the connection can uniquely determine a -valued 1-form on horizontal over , given by
Hence, we have . Given a 1-form on , is a smooth function on . Then, it is easy to verify that
(7.46) |
Theorem 7.22.
Given a smooth function .
(i). Let be the -canonical lift of . Let be the horizontal integral process of stochastic Hamilton’s equations (6.17) corresponding to and . Define a -valued horizontal diffusion by . Then solves the following system on ,
(7.47) |
subject to , where is the damped mean covariant derivative with respect to . In this case, we refer to the system (7.47) as the -canonical reduction of (6.17), or global stochastic Hamilton’s equations.
(ii). If is hyperregular, then the global stochastic Hamilton’s equations (7.47) are equivalent to the stochastic Euler-Lagrange equation (7.22) via the classical Legendre transform and .
(iii). Let . Then the following statements are equivalent:
(a) for every -valued diffusion satisfying
(7.48) |
the -valued process solves the global stochastic Hamilton’s equations (7.47);
(b) satisfies the following Hamilton-Jacobi-Bellman equation
(7.49) |
for some function depending only on .
Proof.
Remark 7.23.
(i). Assertions (ii) and (iii) of Theorem 7.22 generalize Theorem 7.18, since from the Legendre transform we observe that the S-EL equation (7.22) is related to HJB equation (7.49) via equation (7.29). However, assertion (iii) is a special case of Theorem 6.19, since HJB equation (7.49) is just the one in (6.39) with the -canonical lift of , due to the observation that .
(ii). The advantage of Theorem 7.22 is that it formulates stochastic Hamiltonian mechanics in a global way similar to stochastic Lagrangian mechanics, while its disadvantage is that it depends on the choice of Riemannian structures. However, unlike stochastic Hamiltonian mechanics of Chapter 6, neither global S-H equations (7.47) nor HJB equation (7.49) encodes any new symplectic or contact structures, as the Hamiltonian functions therein are still classical.
(iii). By a direct calculation similar to (7.50), one easily obtains the following local version of stochastic Euler-Lagrange equation (7.22):
(7.51) |
This local version is related to stochastic Hamilton’s equations (6.10) via the canonical 2nd-order Legendre transform (7.43).
(iv). Similarly to Remark 6.20, if we let , then Theorem 7.22 holds with and zero function in place of and . We will refer to equation (7.49) with as the HJB equation associated with Hamiltonian , or the HJB equation associated with the Lagrangian related to via the Legendre transform (when is hyperregular).
On Riemannian manifolds, canonical transformations of Section 6.5 can also be reduced to tangent bundles. We consider a bundle isomorphism from to , projecting to a time-change map . The transformation is a map from coordinates to satisfying . Both base manifolds and are equipped with some Riemannian metrics and the corresponding Levi-Civita connections.
By the inverse 2nd-order Legendre transform (7.44) and the integrability condition (6.15), the action functional in (7.9) can be rewritten as
where denotes the Stratonovich stochastic differential and . We denote simply , and . Then . Now we make a change of coordinates from to satisfying , and denote that and . We have
where the function plays the role of the 2nd-order Hamiltonian in new coordinate system.
As in Section 6.5, the general condition for a transformation to be canonical is to preserve the form of stochastic Hamilton’s system (7.47). This is equivalent to preserve the form of stochastic stationary-action principle (7.12), according to Theorem 7.22.(ii). It follows from that
Since the underlying process has zero variation at the endpoints, both equalities will be satisfied if the integrands are related by the following SDE:
(7.52) |
where is a function of phase space coordinates or or any mixture of them and called the generating function. Note that in contrast with the classical theory of canonical transformation and also (6.36), here equation (7.52) for canonical transformations is a stochastic differential equation, instead of equation for forms.
Consider the type one generating function , that is, is given as a function of the old and new generalized position coordinates (cf. [35, Section 9.1]). Then using Itô’s formula , and vanishing the coefficients of every (stochastic) differentials , and in (7.52), we get
which recovers (6.37). By taking (i.e., no time-change) and requiring the new Hamiltonian to be identically zero, and writing as the last equation turns into the following HJB equation
where are regarded as coordinates on the product manifold equipped with the direct-sum Riemannian metric and its corresponding Levi-Civita connection, and are the Laplacian on and , respectively, so that is the Laplacian on under the aforementioned connection.
In contrast to the mixed-order contact approach to canonical transformations of Section 6.5, since the changes of coordinates proceed on , one can easily formulate four types of generating functions that are related to each other through classical Legendre transforms in the same way as in classical mechanics [35, Section 9.1]. For example, the type two generating function takes the form , for which we have
(7.53) |
In this case, since and are no longer independent variables, Riemannian structures on and should be related by the transformation. In view of this, we only consider point transformations, a subclass of canonical transformations. That is, we assume to be the form
for some diffeomorphisms ’s and . The second equation of (7.53) implies
So we equip with the (time-dependent) pushforward Riemannian metric of on by , and with the Levi-Civita connection.
Example 7.24 (Canonical transformations for one-dimensional Bernstein’s reciprocal processes).
Consider the scalar case of Example 6.11, that is, the -valued Brownian reciprocal process with 2nd-order Hamiltonian . The equations of motion are , (cf. (6.33)). In the following, we will consider two canonical transformations which transform Brownian reciprocal processes to reciprocal processes derived from diffusions with linear potentials and quadratic potentials, respectively.
(i). Consider the time-dependent change of coordinates from to (without time-change) induced by . By (7.53),
(7.54) |
For the latent 2nd-order coordinates, we have
Hence, by the last equation of (7.53), the new 2nd-order Hamiltonian is
which is still of the form (6.26), with and . The equations of motion under new coordinates are and . By Remark 6.20, share the same equations of motion with . In other words, (7.54) transforms Brownian reciprocal processes to reciprocal processes derived from diffusions with linear potentials. This example is taken from [60, Theorem 4.1.(1)], where the authors used (7.54) to transform free heat equations to heat equations with linear potentials. We refer readers to [60] for more applications of canonical transformations of contact Hamiltonian systems to Euclidean quantum mechanics in Example 6.12.
(ii). Consider the following change of coordinates from to (with time-change)
(7.55) |
Clearly, the map is induced by the type three generating function via relations and . The relation between the latent coordinates and is
(7.56) |
The new 2nd-order Hamiltonian satisfies . Hence, combining with (7.55) and (7.56), we obtain
This differs with the 2nd-order Hamiltonian of Euclidean harmonic oscillators in Example 6.12.(ii), i.e., , by a term depending only on time. So by virtue of Remark 6.20, and share the same equations of motion , . Therefore, (7.55) transforms free reciprocal processes to Euclidean harmonic oscillators.
Example 7.25 (Canonical transformations for vanishing potentials).
Let be Riemannian. Take for some function . Then
Since the transformation on base manifold is identity, it does not change the Riemannian metric, and
(i). We consider the Hamiltonian , whose corresponding 2nd-order Hamiltonian has a diffusion with generator for solution process (see Subsection 6.3.1). Then, the new Hamiltonian is
If we choose solving the backward PDE (6.23), then has a diffusion process with generator for solution. In particular, such a canonical transformation can transform a diffusion process with a scalar potential into a free motion.
(ii). Consider the Hamiltonian , whose corresponding 2nd-order Hamiltonian has a Schrödinger’s bridge with vector potential and scalar potential for solution process. Then, the new Hamiltonian is
To transform into the standard form whose solution is a Schrödinger’s bridge with vector potential , we only need to assume that solves HJB equation (6.28). In particular, such a canonical transformation transforms a Schrödinger’s bridge with a scalar potential into a free one.
Regarding the classical energy introduced in the end of Subsection 7.4.1, for a given classical Lagrangian , we introduce its generalized (or deformed) energy by
where is the solution of the Hamilton-Jacobi-Bellman equation (7.49) associated with (with ). The term stands for the stochastic deformation.
7.4.3 Small-noise limits
In this part, we will see, informally, how our stochastic framework degenerates into classical mechanics as the noise goes to zero. Let be a small parameter which we refer to as diffusivity. The limit when is called the small-noise limit.
Let be the small-noise version of the admissible class (7.10), that is, with the constraint . The -dependent stochastic variational problem is to minimize the action functional in (7.9) among all . Then, the same procedure as Section 7.2 yields the following -dependent stochastic Euler-Lagrange equation,
(7.57) |
which is an equivalent condition for to be a stationary point of . Here is the damped mean covariant derivative with respect to so that
Now as , since , tends to some deterministic curve (in a suitable probabilistic sense), and tends to . Thus, we can write informally
The -dependent stochastic variational problem tends to the following deterministic variational problem
(7.58) |
And the -dependent stochastic Euler-Lagrange equation (7.57) tends to
(7.59) |
where, is the material derivative along . This is the classical Euler-Lagrange equation in global form, cf. [85, p. 153].
We introduce the following -dependent version of the -canonical lift (6.44):
Let be a horizontal integral process of stochastic Hamilton’s equations (6.10) corresponding to and . Since as , converges to a -valued process. And since and , the limit -valued process satisfies classical Hamilton’s equations,
(7.60) |
Let . Then, solves the system of global stochastic Hamilton’s equations (7.47), with , and in place of , and , respectively, subject to . As goes to 0, this system tend to the following deterministic system,
(7.61) |
This is indeed the global form of (7.60) which is equivalent to the global Euler-Lagrange equation (7.59) via the classical Legendre transform.
The corresponding -dependent Hamilton-Jacobi-Bellman equation is now
which, as , goes to the classical Hamilton-Jacobi equation
The latter corresponds to (7.59)–(7.61) via classical Hamilton-Jacobi theory (e.g., [1, Chapter 5]).
We list here some previous works that have independent interests in the above small-noise limits, in some special cases. The time-asymptotic large deviation for Brownian bridges of Example 6.11 was studied in [41]. The second author of the present paper and his collaborator proved in [76] a large deviation result for one-dimensional Bernstein bridges which are solution processes of Euclidean quantum mechanics in Example 6.12. The paper [57] proved that the -limit of Schrödinger’s problem in Section 7.3 with small variance is the Monge-Kantorovich problem. The latter is the optimal transport problem associated with the classical variational problem (7.58) [85, Chapter 7]. See [67, Section 2.3] for more on small-noise limits of stochastic optimal transport.
Remark 7.26.
There are various terminologies in other areas related to the small-noise limit. In thermodynamics [42], stands for the Boltzmann constant which relates to the diffusion coefficient via Einstein relation, as consistent with Schrödinger’s original statistical problem [79]; when applied to quantum mechanics as in Example 6.12, the small-noise limit is called the semiclassical limit and the parameter stands for the reduced Planck constant ; when/if applied to hydrodynamics (cf. [6, 14]), it is often called the vanishing viscosity limit and stands for the kinematic viscosity . The latter may be expected to solve Kolmogorov’s conjecture that the “stochastization” of dynamical systems is related to hydrodynamic PDEs as viscosity vanishes [8]. In physics, diffusivity, Planck constant and viscosity are indeed related to each other [84].
7.4.4 Relations to stochastic optimal control
Following the way of converting problems of classical calculus of variations into optimal control problems (see [29]), we can regard the stochastic variational problem of Section 7.2 as a stochastic optimal control problem.
Assume that is compact (for simplicity). Consider a stochastic control model in which the state evolves according to an -valued diffusion governed by a system of MDEs on the time interval , of the form
(7.62) |
or equivalently, by an Itô SDE of the form
where is the positive definite square root -tensor of , i.e., , is an -valued standard Brownian motion and, most importantly, is a -valued process called the control process. There are no control constraints for as it is admissible in the sense of [29, Definition 2.1]. As endpoint condition, we require that .
The control problem on a finite time interval is to choose to minimize
(7.63) |
among all pairs satisfying the system (7.62) and the endpoint condition, where is a given smooth function on . The real-valued smooth function on is called running cost function and the payoff functional. The problem is called a stochastic Bolza problem. In the case , this stochastic control problem is of the same form as our stochastic variational problem of Section 7.2. For this reason, we call the latter stochastic control problem to be in Lagrange form. By an argument similar to Theorem 7.16, one can derive the same equation as (7.22), but with boundary conditions and .
The starting point of dynamic programming is to regard the infimum of being minimized as a function of the initial data:
Then, Bellman’s principle of dynamic programming [29, Section III.7] states that for ,
Divide the equation by , let , and then use Dynkin’s formula. We get the dynamic programming equation
(7.64) |
subjected to terminal data . By (4.5) and (7.62),
We let
where the supremum can be ignored if is convex, so that is exactly the canonical inverse 2nd-order Legendre transform in (7.43). Then, the dynamic programming equation (7.64) can be written as the HJB equation (6.38), cf. [29, Section IV.3].
There is also a stochastic version of Pontryagin’s maximum principle [87, Theorem 3.3.2]. The crucial objects in stochastic Pontryagin’s principle are first- and second-order adjoint processes, and , respectively. Corresponding to the stochastic control problem (7.62)–(7.63), its adjoint processes and satisfy the following backward SDEs [87, Section 3.3.2] (where “backward” is again in a different sense from ours in Chapter 2),
(7.65) |
and
(7.66) |
which are called first- and second-order adjoint equation, respectively. The unknowns in (7.65) and (7.66) are the pairs and , respectively. Suppose that and for time-dependent 2nd-order form that satisfies 2nd-order Maxwell relations (6.15). Then
Plugging them into (7.65) and (7.66), we get
(7.67) | ||||
These coincide with the corresponding equations in the S-H system (6.10) for 2nd-order Hamiltonian . The first equality of (7.67) also recovers (7.51).
7.5 Stochastic variational symmetries
Definition 7.27.
Given an action functional as in (7.9), a bundle automorphism on projecting to is called a variational symmetry of if, whenever is a subinterval of , we have . A -projectable vector field on is called an infinitesimal variational symmetry of , if its flow consists of variational symmetries of .
Lemma 7.28.
The -projectable vector field of the form (4.9) is an infinitesimal variational symmetry of if and only if
is a martingale, for all .
Proof.
As in the proof of Theorem 4.14, we let be the flow generated by , and denote . Then, by a change of variable ,
Since for all and each , , we have that the difference
is a martingale (depending on ). Taking derivatives with respect to and evaluating at for the above equality, and recalling that , we can obtain the desired result. ∎
Definition 7.29.
Given a smooth function . A -projectable vector field on is called an infinitesimal -divergence symmetry of , if
for all and .
Proposition 7.30.
A vector field of the form (4.9) is an infinitesimal -divergence symmetry of if and only if
Corollary 7.31.
Let be a hyperregular Lagrangian. Let be a vector field of the form (4.9). Given a smooth function , define the -extension of by
(7.68) |
which is a vector field on . Suppose that satisfies
for the solution of the Hamilton-Jacobi-Bellman equation (7.49) associated with (for ). Then, is an infinitesimal -divergence symmetry of if and only if is an infinitesimal symmetry of equation (7.49).
Proof.
By the classical jet bundle theory, we know that is an infinitesimal symmetry of Hamilton-Jacobi-Bellman equation (7.49) if and only if [72, Theorem 2.31]
(7.69) |
where
with coefficients given by [72, Theorem 2.36 or Example 2.38]
Moreover, the jet coordinates satisfy
where we recall from equation (7.29) and Remark 7.23, and also that . Plugging these into (7.69) and using the fact that and due to classical Legendre transform, we have
(7.70) |
where, in the last equality, we used the fact that to derive . The result then follows from Proposition 7.30. ∎
Theorem 7.32 (Stochastic Noether’s theorem).
Proof.
Recall that and . By applying Lemma 7.8.(iv) and (7.22), as well as the fact that , we have
Then, we use HJB equation (7.49) (with ) and the classical Legendre transform to derive
Combining these with the S-EL equation (7.22) and the criterion (7.70) for symmetries of the HJB equation (7.27), we have
The result follows. ∎
Remark 7.33.
(i). In stochastic Hamiltonian formalism, (7.71) reads as .
(ii). The stochastic conservation law (6.19) of a time-independent -canonical 2nd-order Hamiltonian can be regarded as a special case of the above stochastic Noether’s theorem. Indeed, consider the infinitesimal unit time translation , i.e., , , . Then, the criterion (7.70) reduces to , which means that is time-independent. The resulting stochastic conservation law is .
Applying stochastic Noether’s theorem to Schrödinger’s problem of Section 7.3, we have the following corollary. Its Euclidean case with zero vector potential (i.e., ) has already been formulated in [83].
Corollary 7.34 (Stochastic Noether’s theorem for Schrödinger’s problem).
Let be the Lagrangian given in (7.25). Suppose that the vector field in (7.68) is an infinitesimal symmetry of Hamilton-Jacobi-Bellman equation (7.27) with . Then the following stochastic conservation law holds for the coordinate process of the solution of Schrödinger’s problem in (7.33),
where is the classical energy given in (7.45) and is the solution of (7.27).
Appendix A Mixed-order tangent and cotangent bundles
A.1 Mixed-order tangent and cotangent maps
Clearly, the mixed-order tangent bundle is a subbundle of the totally second-order tangent bundle , and contains the tangent bundle as a subbundle. Similar properties hold for the mixed-order cotangent bundle.
It is easy to verify that the mixed-order tangent bundle can be characterized as follows:
We also define the stochastic analog of the vertical bundle as
Then, it is easy to see that .
Given a smooth map , we can define its second-order pushforward as in Definitions 5.5 and 5.7, so that is a bundle homomorphism from to . In general, neither maps the mixed-order tangent bundle to the mixed-order tangent bundle, nor maps the vertical bundle to the vertical bundle. But if is projectable, then it does.
Lemma A.1.
Let and be two smooth manifolds and be connected. Let be a smooth map. Then the following statements are equivalent:
(i) is a bundle homomorphism from to ;
(ii) ;
(iii) .
Proof.
We first prove that (i) implies both (ii) and (iii). Suppose that is a bundle homomorphism projecting to . Then, and hence, for any ,
If , then and thus . This implies . If , then , it follows and therefore .
Next we prove either (ii) or (iii) implies (i). Choose local coordinates around and around . Suppose has a local expression . Let having the following local expression:
(A.1) |
Then, Lemma 5.6 yields
If (ii) holds, then . It then follows
(A.2) |
Since is arbitrary, we know that for all . Then, by the connectness of , is independent of . This implies that is a bundle homomorphism. Now assume that has a local expression in (A.1) with . If (iii) holds, then . This amounts to (A.2) together with
Again, the arbitrariness of yields that for all . Thus, is a bundle homomorphism. ∎
It is easy to deduce from the proof that if is a bundle homomorphism from to , then is a bundle homomorphism from to .
When is a diffeomorphism, we can also consider the second-order pullback map which is a bundle homomorphism from to . But when we restrict to the mixed-order cotangent bundle , there are difficulties. We can check that even if is a bundle homomorphism, does not necessarily map into . The reason is basically that the restrictions of second-order pullbacks to the cotangent bundle do not coincide with usual pullbacks. To overcome this, we consider the dual map of . This motivates the following definition, which contrasts with Definitions 5.5 and 5.7.
Definition A.2 (Mixed-order pushforward and pullback).
Let be a bundle homomorphism from to . The mixed-order tangent map of at is the linear map defined by
The mixed-order cotangent map of at is the linear map dual to , that is,
The mixed-order pushforward by is the bundle homomorphism defined by
Given a mixed-order form on , the mixed-order pullback of by is the mixed-order form on defined by
If, moreover, is a bundle isomorphism, then the mixed-order pullback by is the bundle isomorphism defined by
Given a mixed-order vector field on , the mixed-order pushforward of by is the mixed-order vector field on defined by
Clearly, the mixed-order pushforward is nothing but . Write . Then, in local coordinates, acts on of (A.1) as follows:
(A.3) |
And acts on the mixed-order cotangent vector by
(A.4) |
By virtue of these local expressions, one easily deduce that
And in turn, these verify the linearity of and . The following property is easy to check.
Lemma A.3.
Let be a bundle isomorphism from to and be a mixed-order vector field. Let be a smooth functions on . Then .
A.2 Pushforwards of generators
A smooth map can be associated naturally with a bundle homomorphism that projects to the identity on . In this case, the pushforward of a diffusion by is just . The stochastic prolongations of the bundle homomorphism is then
Corollary A.4.
Let be a diffeomorphism. If a diffusion on has a generator , then the process is a diffusion on , with generator .
Proof.
Assume . For every , , by the assumption, we have
is a real-valued continuous -martingale. This proves that has generator . ∎
This corollary together with the identification between and in (3.6) and (3.7), give rise to the relation between prolongations and pushforwards as follows:
so that .
The following corollary is an extension of Corollary A.4 and a straightforward consequence of Lemma 4.8. Here, we will present another proof, using notions of Appendix A.1.
Corollary A.5.
Let be a bundle isomorphism from to projecting to . If is a diffusion on with respect to and has a extended generator where is a time-dependent second-order vector field, then the pushforward is a diffusion on with respect to , with extended generator
Proof.
Assume that and . For every , Lemma A.3 yields that the process
is a continuous -martingale. Denote . By substituting which can be done because is an isomorphism, and using the change of variable , and recalling that , the process
is a continuous -martingale. The result follows. ∎
Remark A.6.
(i) As a consequence, the generator of the pushforward is given in local coordinates by
This coincides with Lemma 4.8.
(ii) This corollary together with Lemma A.1 indicates that the bundle homomorphisms from to are the only (deterministic) smooth maps between them that map diffusions to diffusions. Indeed, if a smooth map from to pushes forward a diffusion to another diffusion, then a similar argument as in Corollary A.5 implies that would map the extended generator of the former diffusion to that of the latter, whereas Lemma A.1 says such must be the second-order pushforward of some bundle homomorphism.
(iii) In particular, if is a smooth map from to and is a diffusion on with generator , then is a diffusion on with respect to the same filtration, with generator .
A.3 Pushforwards and pullbacks by diffusions
Definition A.7 (Pushforwards and pullbacks by diffusions).
Let be an -valued diffusion process. Let be a coordinate chart on . The pushforward map from to is defined in the local coordinate by
(A.5) |
The pullback map from to is defined by
(A.6) |
Remark A.8.
Recall that in classical differential geometry, the pushforward by a smooth curve on is a map given by . While if we look at the graph of as a section of the trivial bundle , denoted by , then the pushforward map by is . For this reason, it would be more appropriate to call and in Definition A.7 the pushforward and pullback by graph of , or by random section corresponding to , instead of by itself. But we avoid that for simplicity.
One can see from the definition that the pushforward maps the time vector to the value of the extended generator of at . There is an informal way to look at the pullback map : one first replace all ’s by ’s in the brackets at LHS of (A.6) and obtain
then substituting and , and following Itô’s calculus,
and getting rid of the martingale part, we get the RHS of (A.6).
The following corollary is straightforward. We will see that pushforward and pullback maps by diffusions are also closely related to the concept of “total derivatives”.
Corollary A.9.
(i). Let be an -valued diffusion process. For all and ,
(A.7) |
(ii). If , is a smooth function on and a smooth function on , then
(iii). Let be -valued diffusion processes satisfying a.s.. Then, a.s. if and only if a.s.. In particular, if , then if and only if .
(iv). Let be a bundle homomorphism from to projecting to , and be an -valued diffusion process. Then .
(v). Let be a smooth function from to , and be an -valued diffusion process. Then .
Proof.
Assertions (i), (ii) and (iii) are easy to deduce from the definitions. We prove (iv) using local expressions. Assume that and denote . Recall that . Then
The result follows. ∎
A.4 Lie derivatives
Definition A.10 (Lie derivatives).
Let be a vector field on and be its flow. Let be a second-order vector field and be a second-order form on . The Lie derivative of with respect to is a second-order vector field on , denoted by , and defined by
The Lie derivative of with respect to is a second-order form on , denoted by , and defined by
For sufficient small , is defined in a neighborhood of and is the inverse of . So the difference quotients in the above definitions of Lie derivatives make sense. It is easy to verify that the derivatives exist for each , and is a smooth second-order vector field, is a smooth second-order covector field. Likewise, the restrictions of to and coincide with the classical Lie derivatives. In the following, we will seek properties of . Some of them can be found in [66, Section 6.(d)].
Lemma A.11.
Let be a vector field and be a smooth function. Let and be a second-order vector field and second-order form, respectively. Then
(i) , where the RHS denotes the commutator of and as linear operators;
(ii) ;
(iii) ;
(iv) ;
(v) .
Remark A.12.
Note that the commutator is a second-order vector field. Indeed, if and have coordinate expressions and , then the following local expression for is easy to verify,
Proof.
(i) For a function ,
Then, a similar argument to the derivation of classical Lie derivatives yields
(ii) .
(iii) For a second-order vector field ,
(iv) Use (iii) to derive
(v) Again using (iii) we have . ∎
Corollary A.13.
(i) .
(ii) .
(iii) commutes with the symmetric product operator .
Proof.
Given a vector field on , the Lie derivative can also be defined for second-order vector fields and second-order forms on , as in Definition A.10, without any changes. But when restricting to the mixed-order vector fields and mixed-order forms, it is necessary that the flow in Definition A.10 consists of bundle homomorphisms on , so that its mixed-order pushforwards and pullbacks are well defined. This feeding back to the vector field amounts to is -projectable. In this case, we just replace the second-order pushforwards and pullbacks in Definition A.10 by mixed-order pushforwards and pullbacks, to define the Lie derivative for mixed-order vector fields and mixed-order forms on .
Now let be a -projectable vector field on . Then, Lemma A.11.(i)–(iv) still holds for smooth functions on , mixed-order vector fields and mixed-order forms on . The assertion (v) will hold with the mixed differential instead of the second-order differential, that is, . Moreover, if and have coordinate expressions and where only depends on time, then the Lie derivative has the following expression:
Appendix B The mixed-order contact structure on
B.1 Mixed-order total derivatives and mixed-order contact forms
We denote by the pullback bundle (see [77, Definition 1.4.5]) of by . It is a fiber bundle over .
Definition B.1 (Mixed-order holonomic lift).
Let , , and . The mixed-order holonomic lift of by is defined to be
The set of all mixed-order holonomic lifts is denoted by , that is,
Since depends only upon the mean derivatives of at , the holonomic lift of a tangent vector is completely determined by and does not depend on the choice of the representative diffusion . In particular, the set is well defined and is clearly a subbundle of .
Lemma B.2.
The fiber bundle can be written as the Whitney sum of two subbundles
Proof.
Suppose that . Then , and
It follows easily from the definition of pushforward (A.5) that . Hence, and
The result follows. ∎
The decomposition of may then be found by letting
Definition B.3.
A section of the bundle is called a mixed-order total derivative. The specific section
is called the coordinate mixed-order total derivative, and is denoted by .
The coordinate mixed-order total derivative is just the total mean derivative in Definition 4.7. The dual construction is the mixed-order contact cotangent vector, which may be described as being in the kernel of .
Definition B.4.
An element is called a mixed-order contact cotangent vector if . The set of all mixed-order contact cotangent vectors is denoted by , that is,
It is straightforward to check that the vanishing of does not depend on the particular choice of the representative diffusion . The dual relation between and in (A.7) implies that the mixed-order contact and holonomic elements annihilate each other.
To express a mixed-order contact cotangent vector in coordinates, let us consider
(B.1) |
Using the definition (A.6) we get
There are two basic nontrivial solutions of the above equation, say,
Plugging these solutions in (B.1), we get two basic types of mixed-order contact cotangent vectors
Thus, every mixed-order contact cotangent vector in is a linear combination of these basic mixed-order contact cotangent vectors.
Lemma B.5.
The fiber bundle can be written as the Whitney sum of two subbundles
Proof.
Suppose that . Then, , and the definition of pullback yields
Since , it follows that
This ends the proof. ∎
The decomposition of may then be found by letting
Definition B.6.
A section of the bundle is called a mixed-order contact form. The following specific sections
are called basic mixed-order contact forms.
It follows from the construction that the set of basic mixed-order contact forms defines a local frame of the bundle .
Remark B.7.
Corollary B.8.
Let be a coordinate chart on . Let be a -valued diffusion process. In local coordinates, the pushforward map from to is given by
The pullback map from to is given by
Corollary B.9.
Let be a section of . Then is a mixed-order contact form if and only if for every and every ,
Proof.
We first let be a mixed-order contact form and let . Then
(B.2) |
To prove the converse, we suppose
Fix a particular index with . Let such that , and
Then,
It follows from the arbitrariness of that for all . Similarly, all , and vanish. Consequently, . As in (B.2), we have . Hence, is a mixed-order contact form. ∎
Corollary B.10.
Let be a -valued diffusion process. Then , with an -valued diffusion process, if and only if for every mixed-order contact form on .
Proof.
We first suppose with an -valued diffusion process. Then, for a mixed-order contact form ,
To prove the converse, it suffices to show, in local coordinates, that
This can be done as soon as we let be a basic mixed-order contact form. For example, let , then
which leads to . ∎
B.2 The mixed-order Cartan distribution and its symmetries
The model bundle is a trivial bundle over in its own right, and so we may consider its mixed-order tangent bundle .
Definition B.11.
The bundle endomorphisms of is defined by
where and .
Definition B.12 (Mixed-order Cartan distribution).
The mixed-order Cartan distribution is the kernel of the vector bundle homomorphism over
and is denoted by .
Note that is a subbundle of . It follows from the above two definitions that
Hence, for each ,
Similarly to the proof of Lemma B.2, we can decompose an element as
(B.3) |
where and .
From the duality relations it also follows that is the annihilator of , or in other words, the basic mixed-order contact forms are local defining forms for the mixed-order contact distribution . A typical element may be written in coordinates as
(B.4) |
From this it is easy to deduce .
Definition B.13.
A symmetry of the mixed-order Cartan distribution on is a bundle automorphism of which satisfies .
It follows by duality that symmetries of the mixed-order Cartan distribution are those bundle automorphisms which satisfy . For this reason, is also called a mixed-order contact transformation. Similarly, may be characterized by the fact that whenever is a mixed-order contact form then so is .
Proposition B.14.
Let be a bundle homomorphism from to that projects to a diffeomorphism . Then if and only if where is a bundle homomorphism from to that projects to .
Proof.
First, we prove the sufficiency. Let . According to (B.3), we decompose by with and . Then, since by Corollaries 4.6 and A.9.(iv), where is the pushforward of by , we have
Besides, since is a bundle homomorphism projecting to by Corollary 4.5.(ii), we have . Then,
which yields . This proves .
For the necessity, we first prove that is bundle homomorphism from to by showing , by virtue of Lemma A.1. Let . Set , where and for some diffusion . Since projects to ,
while . Thus, . Since , we set . Then . Hence, and so . This leads to and so that is bundle homomorphism from to . Denote the projection of onto a map from to by . It follows that
Since is surjective, we obtain , so that is a bundle homomorphism from to projecting to . We shall write and .
Corollary B.15.
Let be a bundle automorphism on projecting to a diffeomorphism . Then is a symmetry of if and only if where is a bundle automorphism on that projects to .
Proof.
If is a symmetry, then and . By Proposition B.14, and for some bundle endomorphisms and on that projects to and , respectively. Then, Corollary 4.5.(iii) implies that and hence . For the same reason, . Thus, is a bundle automorphism on . Conversely, if and is a bundle automorphism, then , which yields and hence is a bundle automorphism on . ∎
B.3 Infinitesimal symmetries
Definition B.16.
An infinitesimal symmetry of the mixed-order Cartan distribution is a -projectable vector field on with the property that, whenever the mixed-order vector field belongs to , then so does the mixed-order vector field .
Like in the classical case, an infinitesimal symmetry of the mixed-order Cartan distribution may also be called an infinitesimal mixed-order contact transformation. By duality, is such an infinitesimal symmetry precisely when is a contact form for every mixed-order contact form .
The following lemma is a consequence of the definition of Lie derivatives.
Lemma B.17.
Let be a -projectable vector field on with flow . Then, is an infinitesimal symmetry of the mixed-order Cartan distribution if and only if for each , the diffeomorphism is a symmetry of the mixed-order Cartan distribution.
The following result is the infinitesimal version of Corollary B.15. It can be deduced directly from Lemma B.17 and Corollary B.15. But here we give a computational proof based on the Lie derivative of mixed-order contact forms.
Theorem B.18.
Let be a -projectable vector field on . Then, is an infinitesimal symmetry of the mixed-order Cartan distribution if and only if is the prolongation of a -projectable vector field on .
Proof.
Let the vector field having the following local expression:
where only depends on time due to the projectability of . We then derive the Lie derivative of the basic mixed-order contact forms and as follows:
and
Thus, the mixed-order forms and are mixed-order contact forms if and only if
(B.5) | |||
(B.6) | |||
(B.7) |
Now (B.5) means that ’s only depend on the variables on , so that the vector field is also -projectable. The two equations (B.6) and (B.7) are just restatements of the prolongation formulae in Theorem 4.14. ∎
Appendix C Stochastic Maupertuis’s principle
Based on Definition 7.11, if we further consider the variation caused by time-change, as in classical mechanics (cf. [1, Definition 3.8.4] or the so called -variation in [35, Section 8.6]), then we need to impose the constraint of constant energy. So the path space in (7.10) is modified to
where is a regular value of .
Definition C.1.
Given and , by a variation of the pair along , we mean a family of pairs where , , such that for each , , , and for each , , satisfies the ODE
(C.1) |
Define a functional by
The pair is called a stationary point of , if
As in Lemma 7.13, it is easy to deduce from (C.1) that for each so that . Moreover, formula (7.13) still holds for all , with in place of .
Lemma C.2.
Keep the notations in Definition C.1. Then, in normal coordinates we have
Proof.
Theorem C.3 (Stochastic Maupertuis’s principle).
Let be a regular Lagrangian on . Let such that . Then, the pair is a stationary point of if and only if satisfy the stochastic Euler-Lagrange equation (7.22).
Proof.
Since all diffusions in have the same average energy , we have
Denote . As in (7.23),
We apply (7.24) and notice that in the present situation we do not have in general. Hence,
One the other hand, since for all , and . It follows from Lemma C.2 that
Therefore,
By the definition of the energy , we know that
The result follows. ∎
Data Availability
Our manuscript has no associated data.
Acknowledgements.
We would like to thank Prof. Ana Bela Cruzeiro and Prof. Marc Arnaudon for their careful reading and helpful discussions, which helped us a lot especially in improving Section 6.2 and 7.2. We also would like to thank Prof. Maosong Xiang for his helpful suggestions and kind experience-sharing. This paper is supported by FCT, Portugal, project PTDC/MAT-STA/28812/2017, “Schrödinger’s problem and optimal transport: a multidisciplinary perspective (SchröMoka)”.
References
- Abraham and Marsden [1978] R. Abraham and J.E. Marsden. Foundations of mechanics. Addison-Wesley Publishing Company, 2nd ed., 1978.
- Albeverio et al. [1989] S. Albeverio, K. Yasue, and J.-C. Zambrini. Euclidean quantum mechanics: analytical approach. In Annales de l’IHP Physique théorique, vol. 50, pp. 259–308, 1989.
- Albeverio et al. [2006] S. Albeverio, J. Rezende, and J.-C. Zambrini. Probability and quantum symmetries. II. The theorem of Noether in quantum mechanics. Journal of mathematical physics, 47(6):062107, 2006.
- Angst et al. [2015] J. Angst, I. Bailleul, and C. Tardif. Kinetic Brownian motion on Riemannian manifolds. Electronic Journal of Probability, 20:1–40, 2015.
- Arnaudon and Thalmaier [1998] M. Arnaudon and A. Thalmaier. Complete lifts of connections and stochastic Jacobi fields. Journal de mathématiques pures et appliquées, 77(3):283–315, 1998.
- Arnaudon et al. [2014] M. Arnaudon, X. Chen, and A.B. Cruzeiro. Stochastic Euler-Poincaré reduction. Journal of Mathematical Physics, 55(8):081507, 2014.
- Arnold [1989] V.I. Arnold. Mathematical methods of classical mechanics, vol. 60. Springer-Verlag New York, 2nd ed., 1989.
- Arnold and Khesin [2021] V.I. Arnold and B.A. Khesin. Topological methods in hydrodynamics, vol. 125. Springer Nature Switzerland, 2nd ed., 2021.
- Asorey et al. [1983] M. Asorey, J.F. Carinena, and L.A. Ibort. Generalized canonical transformations for time-dependent systems. Journal of mathematical physics, 24(12):2745–2750, 1983.
- Belopolskaya and Dalecky [1990] Y.I. Belopolskaya and Y.L. Dalecky. Stochastic equations and differential geometry. Kluwer Academic Publishers, 1990.
- Bernstein [1932] S. Bernstein. Sur les liaisons entre les grandeurs aléatoires. Verh. Internat. Math.-Kongr., Zurich, Band I, 1932.
- Bismut [1981] J.-M. Bismut. Mécanique aléatoire, vol. 866. Springer-Verlag Berlin Heidelberg, 1981.
- Çetin and Danilova [2016] U. Çetin and A. Danilova. Markov bridges: SDE representation. Stochastic Processes and their Applications, 126(3):651–679, 2016.
- Chen et al. [2023] X. Chen, A.B. Cruzeiro, and T.S. Ratiu. Stochastic variational principles for dissipative equations with advected quantities. Journal of Nonlinear Science, 33(1):5, 2023.
- Chung and Zambrini [2003] K.L. Chung and J.-C. Zambrini. Introduction to random time and quantum randomness, vol. 1. World Scientific, new ed., 2003.
- Cruzeiro and Vuillermot [2015] A.B. Cruzeiro and P.-A. Vuillermot. Forward-backward stochastic differential equations generated by Bernstein diffusions. Stochastic Analysis and Applications, 33(1):91–109, 2015.
- Cruzeiro and Zambrini [1991] A.B. Cruzeiro and J.-C. Zambrini. Malliavin calculus and Euclidean quantum mechanics. I. Functional calculus. J. Funct. Anal., 96(1):62–95, 1991.
- Cruzeiro et al. [2000] A.B. Cruzeiro, L. Wu, and J.-C. Zambrini. Bernstein processes associated with a Markov process. In Stochastic Analysis and Mathematical Physics, pp. 41–72. Springer Science & Business Media, 2000.
- Dahlqvist et al. [2019] A. Dahlqvist, J. Diehl, and B.K. Driver. The parabolic Anderson model on Riemann surfaces. Probability Theory and Related Fields, 174(1):369–444, 2019.
- Dirac [1933] P.A.M. Dirac. The Lagrangian in quantum mechanics. Physik. Zeitschrift der Sowjetunion, Band 3(Heft 1):64–72, 1933.
- Dohrn and Guerra [1979] D. Dohrn and F. Guerra. Geodesic correction to stochastic parallel displacement of tensors. In Stochastic Behavior in Classical and Quantum Hamiltonian Systems, pp. 241–249. Springer, 1979.
- Driver [1992] B.K. Driver. A Cameron–Martin type quasi-invariance theorem for Brownian motion on a compact Riemannian manifold. Journal of functional analysis, 110(2):272–376, 1992.
- Dynkin [1968] E.B. Dynkin. Diffusion of tensors. In Dokl. Akad. Nauk SSSR, vol. 179, pp. 1264–1267. Russian Academy of Sciences, 1968.
- Elworthy [1982] K.D. Elworthy. Stochastic differential equations on manifolds, vol. 70. Cambridge University Press, 1982.
- Emery [1989] M. Emery. Stochastic calculus in manifolds. Springer-Verlag Berlin Heidelberg, 1989.
- Emery [2007] M. Emery. An invitation to second-order stochastic differential geometry. hal-00145073, 2007.
- Fang and Malliavin [1993] S. Fang and P. Malliavin. Stochastic analysis on the path space of a Riemannian manifold: I. Markovian stochastic calculus. J. Funct. Anal., 118(1):249–274, 1993.
- Feynman [1948] R.P. Feynman. Space-time approach to non-relativistic quantum mechanics. Rev. Mod. Phys., 118:367–387, 1948.
- Fleming and Soner [2006] W.H. Fleming and H.M. Soner. Controlled Markov processes and viscosity solutions, vol. 25. Springer Science & Business Media, 2nd ed., 2006.
- Fock [1978] V.A. Fock. Fundamentals of quantum mechanics. Mir Publishers, 2nd ed., 1978.
- Gaeta and Quintero [1999] G. Gaeta and N.R. Quintero. Lie-point symmetries and stochastic differential equations. Journal of Physics A: Mathematical and General, 32(48):8485–8505, 1999.
- Geiges [2008] H. Geiges. An introduction to contact topology, vol. 109. Cambridge University Press, 2008.
- Gentil et al. [2020] I. Gentil, C. Léonard, and L. Ripani. Dynamical aspects of the generalized Schrödinger problem via Otto calculus–a heuristic point of view. Revista Matemática Iberoamericana, 36(4):1071–1112, 2020.
- Gliklikh [2011] Y.E. Gliklikh. Global and stochastic analysis with applications to mathematical physics. Springer-Verlag London Limited, 2011.
- Goldstein et al. [2002] H. Goldstein, C. Poole, and J. Safko. Classical mechanics. Pearson Education, 3rd ed., 2002.
- Hairer [2014] M. Hairer. A theory of regularity structures. Inventiones mathematicae, 198(2):269–504, 2014.
- Haussmann [1986] U.G. Haussmann. A stochastic maximum principle for optimal control of diffusions, vol. 151. Longman Scientific and Technical, 1986.
- Holm et al. [2009] D.D. Holm, T. Schmah, and C. Stoica. Geometric mechanics and symmetry: from finite to infinite dimensions, vol. 12. Oxford University Press, 2009.
- Hsu [1995] E.P. Hsu. Quasi-invariance of the Wiener measure on the path space over a compact Riemannian manifold. J. Funct. Anal., 134:417–450, 1995.
- Hsu [2002] E.P. Hsu. Stochastic analysis on manifolds, vol. 38. American Mathematical Society, 2002.
- Hsu [1990] P. Hsu. Brownian bridges on Riemannian manifolds. Probability theory and related fields, 84(1):103–118, 1990.
- Huang and Zambrini [2023] Q. Huang and J.-C. Zambrini. Stochastic geometric mechanics in nonequilibrium thermodynamics: Schrödinger meets Onsager. Journal of Physics A: Mathematical and Theoretical, 56(13):134003, 2023.
- Ikeda and Watanabe [1989] N. Ikeda and S. Watanabe. Stochastic differential equations and diffusion processes, vol. 24. North-Holland Publishing Company, 2nd ed., 1989.
- Itô [1962] K. Itô. The Brownian motion and tensor fields on Riemannian manifold. In Proceedings of the international congress of mathematicians 1962, pp. 536–539. Almqvist & Wiksells, 1962.
- Itô [1975] K. Itô. Stochastic parallel displacement. In Probabilistic methods in differential equations, pp. 1–7. Springer, 1975.
- Jamison [1975] B. Jamison. The Markov processes of Schrödinger. Z. Wahrscheinlichkeitstheorie verw. Gebiete, 32(4):323–331, 1975.
- Jost [2017] J. Jost. Riemannian geometry and geometric analysis. Springer International Publishing, 7th ed., 2017.
- Karatzas and Shreve [1991] I. Karatzas and S. Shreve. Brownian motion and stochastic calculus, vol. 113. Springer-Verlag New York, 1991.
- Khesin et al. [2021] B. Khesin, G. Misiołek, and K. Modin. Geometric hydrodynamics and infinite-dimensional Newton’s equations. Bulletin of the American Mathematical Society, 58(3):377–442, 2021.
- Kobayashi and Nomizu [1963] S. Kobayashi and K. Nomizu. Foundations of differential geometry, vol. 1. Interscience Publishers, 1963.
- Lang [1999] S. Lang. Fundamentals of differential geometry, vol. 191. Springer Science & Business Media New York, 1999.
- Lassalle and Zambrini [2016] R. Lassalle and J.-C. Zambrini. A weak approach to the stochastic deformation of classical mechanics. Journal of Geometric Mechanics, 8(2):221, 2016.
- Lázaro-Camí and Ortega [2008] J.-A. Lázaro-Camí and J.-P. Ortega. Stochastic Hamiltonian dynamical systems. Report on Mathematical Physics, 61(1):65–122, 2008.
- Lázaro-Camí and Ortega [2009] J.-A. Lázaro-Camí and J.-P. Ortega. The stochastic Hamilton-Jacobi equation. Journal of Geometric Mechanics, 1(3):295, 2009.
- Lee [2013] J.M. Lee. Introduction to smooth manifolds, vol. 218. Springer Science+Business Media New York, 2nd ed., 2013.
- Léonard [2012a] C. Léonard. Girsanov theory under a finite entropy condition. In Séminaire de Probabilités XLIV, pp. 429–465. Springer, 2012a.
- Léonard [2012b] C. Léonard. From the Schrödinger problem to the Monge–Kantorovich problem. Journal of Functional Analysis, 262(4):1879–1920, 2012b.
- Léonard [2014] C. Léonard. A survey of the Schrödinger problem and some of its connections with optimal transport. Discret. Contin. Dyn. Syst., 34(4):1533–1574, 2014.
- Léonard et al. [2014] C. Léonard, S. Rœlly, and J.-C. Zambrini. Reciprocal processes: A measure-theoretical point of view. Probability Surveys, 11:237–269, 2014.
- Lescot and Zambrini [2007] P. Lescot and J.-C. Zambrini. Probabilistic deformation of contact geometry, diffusion processes and their quadratures. In Seminar on Stochastic Analysis, Random Fields and Applications V, vol. 59, pp. 203–226. Springer, 2007.
- Li [2016] X.-M. Li. Limits of random differential equations on manifolds. Probability Theory and Related Fields, 166(3):659–712, 2016.
- Malliavin [1997] P. Malliavin. Stochastic analysis, vol. 313. Springer-Verlag Berlin Heidelberg, 1997.
- Marsden and Ratiu [1999] J.E. Marsden and T.S. Ratiu. Introduction to mechanics and symmetry: a basic exposition of classical mechanical systems, vol. 17. Springer Science & Business Media, 2nd ed., 1999.
- Meyer [1979] P.-A. Meyer. Formes differentielles d’ordre . Publication IRMA, Université Louis Pasteur, Strasbourg, 80, 1979.
- Meyer [1981a] P.-A. Meyer. A differential geometric formalism for the Itô calculus. In Stochastic Integrals, vol. 851 of LNM, pp. 256–270. Springer, 1981a.
- Meyer [1981b] P.-A Meyer. Géométrie stochastique sans larmes. In Séminaire de Probabilités XV 1979/80, pp. 44–102. Springer, 1981b.
- Mikami [2021] T. Mikami. Stochastic Optimal Transportation: Stochastic Control with Fixed Marginals. Springer Nature, 2021.
- Munkres [1975] J.R. Munkres. Topology. Prentice Hall. Inc, 2nd ed., 1975.
- Nelson [2001] E. Nelson. Dynamical theories of Brownian motion, vol. 106. Princeton University Press, 2nd ed., 2001.
- Øksendal [2010] B. Øksendal. Stochastic differential equations: an introduction with applications. Springer-Verlag Berlin Heidelberg, 2010.
- Olver [1995] P.J. Olver. Equivalence, invariants and symmetry. Cambridge University Press, 1995.
- Olver [1998] P.J. Olver. Applications of Lie groups to differential equations, vol. 107. Springer-Verlag New York, 2nd ed., 1998.
- Otto [2001] F. Otto. The geometry of dissipative evolution equations: the porous medium equation. Commun. Partial Differ. Equ., 26(1-2):107–174, 2001.
- Petersen [2016] P. Petersen. Riemannian geometry, vol. 171. Springer International Publishing AG, 3rd ed., 2016.
- Peyré et al. [2019] G. Peyré, L. Chizat, F.-X. Vialard, and J. Solomon. Quantum entropic regularization of matrix-valued optimal transport. European Journal of Applied Mathematics, 30(6):1079–1102, 2019.
- Privault et al. [2016] N. Privault, X. Yang, and J.-C. Zambrini. Large deviations for Bernstein bridges. Stochastic Processes and their Applications, 126(5):1285–1305, 2016.
- Saunders [1989] D.J. Saunders. The geometry of jet bundles, vol. 142. Cambridge University Press, 1989.
- Schrödinger [1926] E. Schrödinger. Quantization as a problem of proper values (part I). Annalen der Physik, 1926.
- Schrödinger [1932] E. Schrödinger. Sur la théorie relativiste de l’électron et l’interprétation de la mécanique quantique. Annales de l’institut Henri Poincaré, 2(4):269–310, 1932.
- Schwartz [1980] L. Schwartz. Semi-martingales sur des variétés, et martingales conformes sur des variétés analytiques complexes, vol. 780. Springer-Verlag Berlin Heidelberg, 1980.
- Schwartz [1982] L. Schwartz. Géométrie différentielle du 2 ème ordre, semi-martingales et équations différentielles stochastiques sur une variété différentielle. In Séminaire de Probabilités XVI, 1980/81 Supplément: Géométrie Différentielle Stochastique, pp. 1–148. Springer, 1982.
- Schwartz [1984] L. Schwartz. Semimartingales and their stochastic calculus on manifolds. Gaetan Morin Editeur Ltee, 1984.
- Thieullen and Zambrini [1997] M. Thieullen and J.-C. Zambrini. Probability and quantum symmetries I: The theorem of Noether in Schrödinger’s Euclidean quantum mechanics. Ann. Inst. Henri Poincaré, 67(3):297–338, 1997.
- Trachenko and Brazhkin [2021] K. Trachenko and V.V. Brazhkin. The quantum mechanics of viscosity. Physics Today, 74(12):66–67, 2021.
- Villani [2009] C. Villani. Optimal transport: old and new, vol. 338. Springer-Verlag Berlin Heidelberg, 2009.
- von Renesse [2012] M.-K. von Renesse. An optimal transport view of Schrödinger’s equation. Canadian mathematical bulletin, 55(4):858–869, 2012.
- Yong and Zhou [1999] J. Yong and X.Y. Zhou. Stochastic controls: Hamiltonian systems and HJB equations, vol. 43. Springer-Verlag New York, Inc, 1999.
- Zambrini [1986] J.-C. Zambrini. Variational processes and stochastic versions of mechanics. Journal of Mathematical Physics, 27(9):2307–2330, 1986.
- Zambrini [2015] J.-C. Zambrini. The research program of stochastic deformation (with a view toward geometric mechanics). In Stochastic analysis: a series of lectures, vol. 68 of Progress in Probability, pp. 359–393. Springer Basel, 2015.