1. Introduction
In the early days of quantum mechanics, before the arrival of the path-integral formalism, Koopman and von Neumann (KvN) [
1,
2,
3], motivated by improving their understanding of classical ergodicity, reformulated classical mechanics on a symplectic manifold
M as quantum mechanics on
with Hamiltonian linear in momenta and complex wave function on
M given by the square root of the classical density function modulo phase, whose fate remains an interesting open problem in KvN mechanics. Over half a century later, an equivalent path-integral formulation was given by E. Gozzi, M. Reuter, and W. D. Thacker (GRT) [
4,
5].
On the other hand, the Alexandrov–Kontsevich–Schwarz–Zaboronsky (AKSZ) formalism [
6] provides a natural framework for the deformation quantization of graded geometries appearing in the context of symplectic mechanics and gauge field theories. It was initially applied for the quantization of topological systems, but is also suitable to the description of dynamical systems with local degrees of freedom, provided these systems are described in terms of an exterior Cartan-integrable system; see, e.g., [
7,
8,
9] for discussions.
In this paper, we will show that the AKSZ formalism enables us to reframe in a very natural way the path-integral formulation of classical mechanics starting from the operatorial formulation by Koopman and von Neumann (KvN) [
1,
2,
3], and was further developed in [
4,
5]. For that purpose, we will be studying the dynamics of a classical system whose phase space corresponds to the symplectic manifold
and focus on its cotangent bundle
, which is also symplectic. We will further subject
to a system of first-class constraints that will provide us with a natural extension of the Koopman–von Neumann (KvN) reformulation of classical mechanics on
M. More precisely, as we shall see, the link between KvN and AKSZ consists of the fact that the rewriting due to a series of works by Gozzi, Reuter, and Thacker (GRT) [
4,
5] (and thereafter by Gozzi and Regini [
10]) of the pull-back operation on
M along a symplectomorphism generated by a Hamiltonian vector field during a fixed time
t, by means of time slicing of the path integral over particle configurations on
, can be recovered by gauge fixing of a one-dimensional AKSZ sigma model with the target
In [
10], the authors worked out all the different transformations of the fields appearing in the path-integral formulation of classical mechanics, which we will briefly discuss in
Section 2, to show that they form the cotangent bundle
of the reversed-parity tangent bundle of the phase space
M. In our case, we show that this identification is quite natural from the viewpoint of a one-dimensional sigma model, opening up the possibility of generalizing the construction by replacing
in (
1) with a general group
G. Most importantly, our observation would facilitate novel ways of looking at the symmetries and conservation laws of classical mechanics from the viewpoint of the rich geometric and topological structure of the AKSZ sigma models. The possibility of a connection between KvN formalism and geometric quantization as envisaged previously in [
11,
12] can now be further explained from the vantage point of AKSZ sigma models.
Our plan for this paper is the following: In
Section 2, we begin with a brief review of the Koopman–von Neumann (KvN) formulation [
1,
2,
3] of classical mechanics, as well its path-integral reformulation due to Gozzi, Reuter, and Thacker (GRT) in [
4,
5,
10].
Section 3 will then elaborate on the AKSZ action for the worldline of a particle. Finally, in
Section 4, we show the exact connection between the AKSZ sigma model and the GRT model and conclude with a summary of our results and future outlook in
Section 5.
2. KvN and Classical Path-Integral Formulation
Classical mechanics and quantum mechanics are developed on the basis of two completely different mathematical paradigms. Contrary to the geometric approach of classical mechanics, the description of quantum mechanics is more algebraic in nature
1. A state in classical mechanics can be viewed as a point on a symplectic manifold, the phase space, which is by definition endowed with a Lie bracket, the Poisson bracket. Any individual observable can then be described as some real-valued function on this symplectic manifold, associated with a Hamiltonian vector field, generating individual flows on this manifold. For example, the flow corresponding to the Hamiltonian
H would describe the time evolution of the system. On the opposite side, the algebraic language of quantum mechanics revolves around the construction of a Hilbert space. Each physical state is described by a ray in the Hilbert space, and observables are the self-adjoint linear operators defined on the Hilbert space. The Lie algebra structure appears by taking commutators between different observables; i.e., it comes through via the associative product defined by the composition of operators acting on the Hilbert space.
During the era of 1930s, several attempts were made to reconcile these two languages. Perhaps with this early motivation, Koopman and von Neumann reformulated classical mechanics to associate it with a Hilbert space of complex and square-integrable functions similar to its quantum mechanical counterparts. Analogous to quantum mechanics, one can also associate complex classical wave functions with a classical mechanical system, of course, with some caveats on which we will not focus in the current context and instead refer to [
18].
Without losing any generality, let us start with an one-dimensional system with phase space density
, which can be interpreted as the probability density of finding a particle at point
q with momentum
p exactly at time
t with the measure
. Liouville’s theorem states that this density has the same property as an incompressible fluid that the phase space volume
remains constant and the corresponding continuity equation becomes
One can now use Hamilton’s equations
to show that
One can easily generalize the discussion to a dynamical system with
n degrees of freedom in configuration space, corresponding to a phase space of dimension
. Defining the Liouville operator as
one can rewrite (
4) as
In the following, we will use the notation
for the dynamical variables in phase space. The basic postulates of Koopman and von Neumann formalism are the following:
- (1)
The existence of a complex function
, which obeys the same dynamical equation as
, i.e.,
- (2)
is
normalizable; i.e., its norm with respect to the following scalar product is finite
Equation (
7) can then be thought of as the analogue of Schrödinger’s equation in quantum mechanics. The Hilbert space spanned by the functions
can then be considered as the Hilbert space for classical mechanics. The postulate of the scalar product ensures a proper definition of the Hilbert space and imposes the norm squared of the states to be
With this definition of scalar product, one can further show that
is a Hermitian operator
2 such that
The Hermitian character of
ensures that
remains conserved during the evolution. Therefore, one can now consistently interpret
as the density probability function, and note that the Liouville theorem (
6) can be derived starting from the postulate (
7) of KvN mechanics itself. As evident from (
6), although the classical wave function
is complex, the evolution of its phase is completely independent from its modulus, unlike the situation in quantum mechanics. We will keep the implications of this observation and further comparisons of KvN mechanics and quantum mechanics for the excellent review in [
18], and move on to the path-integral approach. More details of the following discussion can be found in [
10,
18].
In [
19], the author prescribes a simple way to introduce a path-integral formulation for classical mechanics. Unlike quantum mechanics, where each path is weighted by a probability
with
S being the action of the path considered, in classical mechanics, only the classical path between two fixed end points is allowed to have weight 1, while all the others are weighted to zero. Nevertheless, Hamilton’s variational principle considers all these virtual paths as well, only one being realized in the classical world as the one that extremizes the action. One can think of the classical analogue of the propagator, i.e., the probability of finding a classical particle at a point
in phase space at some time
t, if it was initially at the point
at the time
, as follows:
where, by
, we denote the classical solution of the Hamiltonian equations of motion
given the initial condition
, where
is the inverse of the symplectic matrix.
Slicing up the time interval
into
equal intervals
, and denoting the time in each interval as
with
and
, one can write the delta distribution in (
12) as
Using (
13) and having in mind the limit where
, so that the interval
goes to zero, each of the delta distributions above (
) can be rewritten as
where we have made use of the standard formula
, where
. At this level of formality, the absolute value of the determinant is dropped. Collecting all these definitions together and taking the
limit
3, one can rewrite
in a form of path integral in phase space, where the symbol
indicates a functional definition for the product of the infinite number of delta function coming from (
14) in the limit
.
One can then exponentiate both factors under the path integral in (
16) by introducing
variables
and a total of
anti-commuting variables
through the simple relations
The final result is that the propagator in classical mechanics can be represented as the path integral
with the Lagrangian
being
The first two terms provide one with a symplectic structure. The rest give the extended Hamiltonian
as [
10],
Hence, starting from the original
dimensional phase space with coordinates
, one arrives at an
dimensional extended phase space with coordinates
, where each of the paths in the path-integral formulation is weighted by a factor of
, which, by construction, reproduces all the standard results of classical mechanics. In the series of works pioneered by E. Gozzi in [
19], the authors have explicitly searched for the geometric meaning of this
dimensional space, which at this point seems like an abstraction over the usual notions of the symplectic formulation of classical mechanics. In the following sections, we will show how this apparent abstraction of the extended
dimensional phase space can be understood through a one-dimensional AKSZ sigma model. Together with the equations of motion derived from
and the transformations of each of these new fields under symplectic diffeomorphisms of
, the authors in [
10,
18] correctly concluded that the phase space spanned by the
variables
is
, where
M is the symplectic manifold coordinatized by the original
variables
. We remark that, here and in the rest of the paper, we work in Darboux coordinates, only allowing for canonical transformations instead of all the possible diffeomorphisms of
M.
4. Recovering the GRT Formulation
In order to make contact with the Lagrangian (
19) and the other results of [
4] (
Section 3) reviewed in
Section 2, it appears that one must consider the cotangent bundle of the phase space of our original system, i.e.,
, where
M is the original symplectic manifold with local coordinates
, as this would account for the classical fields
, where
are conjugated to
in
. In other words, we have the decomposition
On top of that, the symplectic potential on
is taken to be canonical,
, which does lead to the kinetic term
. To account for the interactions in the action, as a result of a gauged fixed AKSZ action as described above, we find that one should use a BFV–BRST charge of the form
where we recall the notation
, and where the dots denote terms that are independent of
and that ensure that
.
This suggests that the constrained system described by the sought-for BRST charge
is determined by choosing the constraints
where
which verify
These constraints are all first class, in accordance with our working assumption, and the structure functions are nothing but the first derivatives of the components of the Hamiltonian vector field of the Hamiltonian
H. A direct computation shows that the BRST charge
with the previously defined constraints and structure functions, does indeed satisfy
.
Therefore, we have shown that the action of [
4] (
Section 3) reproduced in the exponential in Equation (
18) is recovered from a gauge fixed AKSZ model in one dimension, whose target space is associated with the system of first-class constraints
on
(namely, the Lagrangian (
19) is recovered by plugging (
52) in the gauge-fixed action (
46) for a first-class constrained system). This is the main result of the present paper.
Let us discuss these constraints. The easiest ones are , which identify the constraint surface as a submanifold of the phase space . In fact, does not specify further the constraint surface, as it vanishes already on . Recall however that in the presence of first-class constraints, one is interested in the reduced phase space, that is, the quotient of the constraint surface by the action of the distribution generated by the first-class constraints. This is where becomes relevant for us, as quotienting M by its action yields the set of classical trajectories (the flows generated by the Hamiltonian H) as the reduced phase space of our model.
Constraints from the Shifted Tangent Bundle
Let us re-derive this constrained system from a different perspective. Suppose we are given a symplectic manifold M and an Hamiltonian . The latter defines an action on M of , which viewed a Lie group with addition as its multiplication rule, whose fundamental vector field is thus the associated Hamiltonian vector field . The integral curves of this vector field, which are nothing but the classical trajectories of this mechanical system, correspond to the orbits of on M. Therefore, the set of classical solutions can be identified with the quotient , the set of the aforementioned orbits.
In order to recover the space of classical solutions from a one-dimensional AKSZ sigma model, to be identified with the previous model, we should find a BFV description of this space, i.e., identify the symplectic -manifold of degree 0 encoding the space of classical trajectories as the result of a coisotropic Weinstein reduction. In other words, let us look for a constrained system, with only first-class constraints, such that the orbits of the gauge symmetry generated by the latter on the constraint surface is isomorphic to the set of classical trajectories.
We have recalled that the space of classical solutions can be thought of as the set of orbits of the
-action generated by the Hamiltonian
H, on the phase space
M. Therefore, we should find a way to recover the latter as a constraint surface, in another symplectic manifold. One simple manner to do so is to consider the cotangent bundle
, with coordinates
, where
are coordinates on
M and
the associated momenta, i.e., the coordinates along the fibers of
. The original manifold
M can then be recovered as a constraint surface defined by
or more geometrically, by identifying
M as the zero section
of its cotangent bundle
. We can lift the
action to
, where it becomes Hamiltonian, generated by the cotangent lift of
. To summarize, this reasoning leads us to considering the same system of first-class constraints as we proposed before, that is,
The first ones,
, identify
M as the constraint surface in
, while the last one,
, corresponds to the
-action lifted to
.
At this point, we can make two observations. First, the Hamiltonian vector fields associated with
obviously form an integrable distribution, as they span the tangent bundle of
M at any point, and hence the Lie algebroid associated with it is simply
. Second, the
-action also defines a Lie algebroid, as any action of a Lie algebra on a manifold does,
5 denoted
, whose underlying vector bundle is the trivial one,
, and which has only a non-trivial anchor in the guise of the fundamental vector
generating the action of
. Both
and
are Lie algebroids over
M, and hence, so is their direct (or Whitney) sum, which we shall denote with
Any Lie algebroid famously gives rise to a
-manifold [
37], so in our case,
is a graded manifold with coordinates
of degree 0, corresponding to coordinates on
M, and degree 1 coordinates
and
corresponding to coordinates along the fibers of
and
, respectively, giving rise to ghosts on the worldline. The cohomological vector field making
into a
-manifold reads
in this coordinate system. The cotangent bundle of this
-manifold
defines the symplectic
-manifold encoding the BFV description of the classical trajectories we discussed, in accordance with the results and observations of [
38]. Recalling from (
22) that the momentum of
is
, the cotangent lift of
is given by
which exactly reproduces the BFV–BRST charge (
52) defining the AKSZ sigma model, with target space
, and whose gauge fixing reproduces the GRT one [
4], as we have shown.
5. Conclusions
This paper offers a way to rethink classical mechanics as a gauge fixed AKSZ sigma model. The way we showed this is to start from the GRT reformulation of KvN classical mechanics.
In the GRT reformulation of classical mechanics, one considers a simple classical system evolving in phase space along a Hamiltonian flow. The authors of [
4,
5,
10] proposed a path-integral prescription for the classical system, for which the price to pay is the introduction of additional fields. They showed that, for the classical motion of the system with
n degrees of freedom in configuration space, one has to introduce a total of
fields to ensure the consistency of the path integral and to reproduce the expected classical trajectories in the phase space of dimension
. As observed in [
10], these
variables span
, where
M is the
dimensional phase space for the system under consideration.
In this paper, we considered the worldline of a particle constrained by a set of first class constraints, and wrote the AKSZ action corresponding to that constrained particle. We showed that the gauge-fixed version of the AKSZ action, for a suitable choice of target space and constraints spelled out in
Section 4, reproduces the action that dictates the GRT path-integral formulation of classical mechanics.
We then reinterpreted our AKSZ sigma model as the BFV description of the constrained system that was designed to reproduce the GRT formulation of a classical, unconstrained, dynamical system in a phase space M. The reduced phase space of this constraint system on , which consists of the set of classical trajectories of the original mechanical system encoded by M and the Hamiltonian H, is recovered by taking the quotient of M by the distribution associated with the Lie algebroid . In particular, the last factor accounts for the flow generated by the Hamiltonian H of the original system. These observations confirm our claim that a classical system, whose phase space corresponds to the symplectic manifold M, is equivalent to a gauge fixed one-dimensional AKSZ sigma model with target space .
This simple yet intriguing mapping between a classical system and a gauge-fixed AKSZ sigma model opens up interesting avenues of research. One direct application of this mapping would be to start with a constrained classical system, and look for its AKSZ counterpart. In particular, one could consider the case of first-class constraints generated by the action of a Lie group G on M, whose BFV–BRST description leads to an AKSZ model with target space , where is the Lie algebra of G. In light of the previous treatment of an unconstrained classical system, one could expect that the relevant target space be of the form , where C is the constraint surface defined by the first-class constraints.
As another direction of research, one can study higher-dimensional sigma models and look for an effective classical mechanical system equivalent to it. Another interesting avenue would be to understand the connection between geometric quantization (see [
13,
14,
15] for original references, and, e.g., [
16,
17]) and KvN mechanics, as shown by [
11,
12] in more detail now in the light of AKSZ sigma models.