1. Introduction
Stochastic processes are essential for describing natural phenomena that are significantly influenced by unpredictable forces [
1]. Occurring across all space–time scales and within diverse scientific disciplines, they model Brownian motion of micrometric particles in fluids, reproduce the irregular or avalanche-like dynamics of electroencephalographic activity, and provide a useful description of stock-market prices, among other examples [
2]. Remarkably, the same stochastic model can exhibit predictive power in seemingly disparate contexts [
3]. This arises from the fact that noise often results from coarse-graining, which smooths out intricate physical details and isolates the essential aspects of a phenomenon [
4]. Observations of noisy data can yield substantial information and serve as a valuable inference tool. For instance, the distance of a system from thermodynamic equilibrium [
5] can be deduced from its fluctuations, as time-reversibility is expected to be encoded in their statistics [
6,
7]. However, while time-symmetry breaking can be evident in fluctuations, it may also become obscured due to the averaging out of crucial degrees of freedom [
8].
This paper focuses on a specific tool within the theory of stochastic processes that is particularly useful in data analysis for discerning the presence or absence of time-reversal symmetry: the stochastic bridge. The term "bridge", in this context, refers to the ensemble of trajectories that start and end at the same specified point within a fixed time interval. This—apparently theoretical—quantity is similar to the so-called “average pulse”, “average avalanche”, or “average fluctuation” shape, which is measured in several experiments, e.g., in the physics of magnetic materials [
9,
10,
11,
12,
13,
14] or in general in materials science [
15,
16,
17,
18,
19,
20], geophysics [
21,
22,
23,
24], biology [
25,
26], neuroscience [
27,
28,
29,
30,
31,
32,
32], and astrophysics [
33,
34]. In several cases, asymmetric shapes have been observed without a physical interpretation because of the lack of a general theory on the subject (the only exception is the asymmetry in the pulse shape observed in some ferromagnetic materials, which has been related to an effective negative mass effect [
35,
36]).
Intuitively, a statistically significant difference between a bridge and its time-reversed counterpart should indicate a violation of the underlying process’s time-reversal symmetry [
37]. To rigorously examine the conditions under which this connection holds, we provide a self-contained development of bridge theory. This approach ensures that readers without prior knowledge of the subject can follow the derivation through to our central result, the symmetry condition presented in Equation (
31). Our derivation relies on the formulation of the effective stochastic differential equation for the bridge, which can also be derived via optimal control (see for instance [
38]). From that, we address the problem of bridge reverse diffusion. Reverse diffusion is a topic that has previously been introduced in physics in the context of stochastic mechanics (an alternative formulation of non-relativistic quantum mechanics [
39]; see [
40] for an introduction and [
41] for a survey of these kinds of approaches) that is currently of significant interest in machine learning due to its fundamental role in techniques for reconstructing data from noise, as well as in generative models [
42,
43,
44]. A careful comparison between the forward and reversed stochastic differential equations allows us to discuss the time-reverse symmetry of the process. We highlight that a natural sufficient condition for the symmetry of the bridge under time-reversal is detailed balance, as has already been discussed in the literature; see [
37] for a enlightening historical summary.
As an instance of the theory, we consider a minimal stochastic model that closely reproduces the observed phenomenology of the asymmetric pulse shape [
45] and at the same time is suitable for an analytic study [
46]. The model, known as a Brownian Gyrator (BG) in the context of stochastic thermodynamics [
47,
48,
49,
50], can be regarded as an extension of the ABBM/CIR process, a paradigmatic model that was introduced to describe the phenomena that manifest “crackling noise” and that can be also considered as a kind of mean field theory for the problem of the dynamics of an elastic manifold in the presence of quenched pinning disorder [
51].
We give, in
Section 2, a recapitulation of the theory of continuous Markov processes (Fokker–Planck and stochastic differential equations). In
Section 3, we describe bridge theory, and in
Section 4, we describe how the same results have been obtained through optimal control theory. In
Section 5, we recall the problem of reverse diffusion, and at the end of that section, we apply the corresponding theory to bridge statistics in order to deduce the necessary condition for bridge time-symmetry (this is our original contribution). Finally, in
Section 6, we list a series of examples and applications. Conclusions are drawn in
Section 7.
2. Continuous Markov Processes
We consider a stochastic process
that is a Markov process, i.e., a process fully characterized by the distribution
and the propagator
, which represent the probability of observing the value
x at time
t without or with, respectively, the knowledge of a value
at a past time
. Every other probability can be evaluated by means of
and
only, since the probabilities of arbitrary
n values at times
are defined as follows:
The Markov property implies that the propagator fulfills the following closed integral relation, known as the Chapman–Kolmogorov (CK) equation:
where
is an intermediate time
, while at the extremes
and
, the relation is trivial, considering that, from the very definition of the propagator,
.
An important class of Markov processes are the continuous processes. The prototype of this class is the Wiener process
, which has a Gaussian propagator
where
whose variance increase linearly from
, resulting in a linear time correlation for the process values, as follows:
(Here, we restrict the discussion to the one-dimensional case, with
and
belonging to the real axis, but the
n-dimensional case can be similarly defined). It is easy to show that the Gaussian propagator of the Wiener process satisfies the CK Equation (
1), essentially because of the stability of the Gaussian distribution under convolution.
However, it is easy to verify that the Wiener propagator satisfies a much more simpler and more useful equation, as follows:
The partial differential Equation (
2) is much more useful than Equation (
1), which is a non-linear integral equation, but it is valid for the Wiener process only. One thus may wonder how we can generalize Equation (
2) to all continuous Markov processes.
A heuristic general definition for this class of processes reads as follows:
where
is a deterministic drift flow,
is an amplitude factor that rules the relative impact of the random noise on the process, and
is the infinitesimal increment of the Wiener process representing the noise. A heuristic derivation of the corresponding forward Fokker–Planck equation follows [
40].
Consider
: the probability of obtaining
y at time
, knowing
x at time
t is equivalent to the probability of obtaining the right Wiener kick
according to Equation (
3):
Subtracting
and dividing for
both sides, we obtain the following in the limit for
on the left hand side of Equation (
4):
For the right side, we obtain, developing for small
:
This gives, in the limit
:
Now, using
and Equation (
2), noting that
we finally transform Equation (
4) as follows:
In this way, we obtained an expression for the time derivative of the propagator of the Markov process at equal times that we can use in the differentiated CK equation in the limit
, as follows:
Using integration by parts (neglecting border terms, which vanish in general), we obtain the so called forward Fokker–Planck (FFP) equation for a (continuous) Markov stochastic process:
This interesting derivation hides an important hypothesis, though. In our derivation, we used the stochastic Equation (
3), which relates the increment of the Markov process
to the Wiener process
. In doing so, we decided that the value of the process at time
does not depend on the values of the Wiener process at times larger than
t. This is the natural way to proceed, as, for instance, when one wants to produce a stochastic sampling of the stochastic trajectories (the Euler–Maruyama method).
In fact, when one tries to provide a more correct foundation in the initial stochastic differential equation (sde), Equation (
3), one discovers that it is well defined only if
b does not depend on
x. The correct procedure by which to give a mathematical meaning to the general sde is to define its integral version and then give a precise definition of the stochastic integral
, which is possible via a partition of the time-integration interval
and an arbitrary choice of the time at which the function
is computed inside the infinitesimal interval
: if one chooses the beginning of the interval
(Ito scheme), the previous derivation gives the correct result. A general choice where
b is computed at point
gives the following
-FFP equation [
52]:
where we use the abbreviated form
. Two interesting cases are the Stratonovich case
, in which the FFP can be rewritten as
and the anti-Ito or isothermal case
, in which the FFP can be rewritten as
Below, unless otherwise specified, we will adopt the Ito integration scheme, that is, the
-FFP equation for
.
The FFP equation is a continuity equation for the probability, since
where
is the current velocity and
is the probability current. The deterministic case
corresponds to the Liouville equation. When
, the new term, sometimes called osmotic velocity
, manifests the effect of random fluctuations.
A stationary solution of the FFP may arise if it exists a stationary distribution
such that
. A special case arises when the current velocity is zero
, that is, when
which is equivalent (see [
2]) to the condition of detailed balance for the stationary process. More precisely, for the stationary process, defined by the stationary one and two point distributions
, and
, the detailed balance condition reads:
stating the invariance of the process under time reversal. (Here, we are considering the simplest case of variables with even parity under time reversal. To better frame the general case, see [
53]).
We recall that the propagator
can also be seen as a function of the conditioning variables
, abbreviated as
, which satisfies the so-called backward Fokker–Planck (BFP) equation, as follows:
En passant, we observe that the FFP and the BFP can be written as
and
, respectively. By simple integration by parts, it is easily seen that the two operators are adjoints, i.e.,
, where the adjoint operation is defined through a scalar product
and
.
In the general dimension
d, one has the Ito sde with independent Wiener processes whose increments are denoted by
,
giving place to the d-dimensional FFP and the corresponding BFP
where, here and below, we use
and
(summation on repeated index is implied, while the position of the index and sub- or superscript is arbitrarily chosen just for clarity of typesetting).
Correspondingly, the current and the osmotic velocities read, respectively, as follows:
while the detailed balance keeps the form given in Equation (
10).
3. Bridge
The bridge of a stochastic process is defined as the set of trajectories passing through a value
at time
and a value
at later time
. A more general definition was introduced in physics by Schrödinger [
37,
54], who proposed the problem of a stochastic process with the starting and final distribution fixed. This problem, known as Schrödinger bridge, was later connected and well framed in terms of optimal transport and control theory (see [
55] for an introduction and [
56] for a survey), giving a better understanding of the clever solution found by Schröedinger. In this paper, we will not address this general problem, and we chose to use only the simplest definition of stochastic bridge, with fixed initial and final values (which correspond to the very special case of degenerate delta-like initial and final distributions).
In the case of a continuous Markov process, the subset of trajectories defining the bridge represents a novel continuous Markov process. Its distribution is defined as follows:
Note that the double conditioning in
is not in contradiction with the Markovianity of the original process, since
. It is possible to derive the differential equation that governs the evolution of
, making use of both the FFP and the BFP of the original problem [
57].
We will show below that
satisfies a modified FFP where the diffusion coefficients are the same
that were used in the original process, while the drift coefficients are modified by the following terms:
The proof proceeds with time-deriving
and then uses Equations (
12) and (
13) as follows:
where
is the normalization, which is constant with respect to the
x derivatives. We note that the first and the third terms in Equation (
16) are enough to build a first drift term, as follows:
Then, we note that
which allows us to rewrite the second and fourth term in Equation (
16) as
Now we can observe that since
we have demonstrated that the terms mentioned before contribute to the sum of a diffusion term for
and a new drift term. All of these together give the new FFP for
:
with initial condition
.
Since the bridge is a Markov process, its propagator connecting to (with ) satisfies the same equation of with an “initial” condition .
The trajectories of the continuous Markov process defined by
and
are described by the Ito sde as follows:
Note that the term
in the sde is the propagator of the unconditioned process whose bridge statistics we want to compute. Usually, we are interested in the bridge statistics for the stationary process, so here and below we assume that
and
.
4. Bridge via Stochastic Control
Here, we recall how to recover the previous result in the framework of stochastic control [
58,
59]. We provide a generalization of the process described in [
38]; see also [
60]. Let us start with a controlled sde:
where the control term
has to be determined by minimizing the cost function
The first integral on the right-hand side is the sum of the running cost of the process (specified by and that of the control. The quadratic form of the running cost for the control will be justifies below, where we show that this choice select the specific sample of free trajectories in order to give the correct final result. We also introduce the arbitrary function , which is a positive definite matrix; we will show what choice is necessary to obtain the final result. The term is a cost on the final position of the trajectories that will be used to impose the arrival point of the bridge.
In order to solve the control problem, one can proceed with the Hamilton–Jacobi–Bellman (HJB) equation, which is derived from the so called dynamical programming principle:
(we omitted the dependence on
in the notation), where
. The minimum is achieved for the value of
given by:
, that is
which, as introduced in the HJB Equation (
21), gives the non-linear partial differential equation
to be solved with the final condition
.
We get rid of the non-linear terms
by means of the transformation
:
and the choice of the Cole–Hopf transformation
with
.
Moreover, from now on, we focus on the case
, since we are interested in the bridge of the process without any running conditioning. In this way, we obtain the following equation for
W:
which is the BFP of the original process.
However, we have to impose the final boundary conditions given by the final cost
. Using the Green function of the equation, which corresponds to the free propagator of the process (since
), we obtain the following:
In order to impose the final position of the bridge
, we can choose an exponential cost
in the limit of
, which in the Laplace approximation gives the following:
This finally results in the control, as follows:
which leads to Equation (
18).
5. Reverse Diffusion
The problem of reverse diffusion can be simply stated as follows [
61]: imagine a continuous Markov stochastic process defined by an initial distribution
and its propagator
, with
, where both the propagator and
solve the FFP (
12). Now, we consider the probability of a value
conditioned to the future value
(i.e.,
):
. We want to answer the question, “Is there an sde (and therefore a corresponding FFP) satisfied by this process?”.
The correct way to answer this question follows. First, one can look for an equation for the joint probability
. This can be done considering the following:
Then, one has to use the BFP equation to express
and the FFP to express
. In this way, we obtain an expression
as a function of partial derivatives in
x of
and
. If now, one reintroduces the joint probability, writing
, one obtains, after algebraic manipulation, an equation for the joint probability, as follows:
Dividing both terms for
one obtains a new equation for
:
Note that Equation (
23) is neither an FFP (because of the signs before the diffusion term) nor a BFP (except in the case of
A and
B bring independent of
x), and we can not directly derive a corresponding sde from that.
A simple interpretation arises from considering a change in the time variable
. Setting
, without loss of generality, the new variable varies between 0 and
, and obviously,
. Naming
allows one to rewrite Equation (
23) as follows:
Again, this equation can be used to define a (non-homogeneous) continuous Markov process with a single time distribution
and a propagator satisfying the same equation:
with
and initial condition
. The trajectories of this continuous Markov process are described by the sde as follows:
where now
is a usual (Ito) Wiener process. The starting condition
of this sde must be chosen with probability
. Obviously, if one considers the reverse of the reversed process, one recovers the forward process.
Note that the new drift can be expressed as
which is the the osmotic velocity of the original process
minus its current velocity
. Note that, on the other hand, the drift of the original (forward) process could also be written in terms of
and
, as
.
We also observe that, at odds with the forward process, in the reverse process we have to fix the starting condition as
and the following evolution is driven by “controlled dynamics”, where a control term (the osmotic velocity) imposing at all times
is added to a reversed drift (opposite sign and
) [
62].
5.1. Stationary Reverse Process
The process defined by Equation (
27) is defined with initial condition
, and the ensemble of the trajectories satisfies
until time
.
is a one-time distribution probability of the (forward) process, i.e., it is governed by a sde with drift
and diffusion term
. If we consider the case of a time-homogeneous process, that is, one where
A and
b do not depend on time, that admits a stationary distribution, we can consider the limit
. In other words, we wonder what the reversed process of a stationary process is. From Equation (
27), its sde is given as follows:
The trajectory of a stationary process that is invariant under time reversal should have the same statistics as its reversed counterpart. This is achieved when the respective sdes coincide, that is, when
which means
that is, when the stationary velocity current is zero. The vanishing of the stationary current implies zero entropy production, consistent with the requirement of time-reversal invariance [
5].
5.2. Reverse Bridge
Now, we proceed to compute the reversed process of the bridge. In this case, we have to consider Equation (
28) with the current and osmotic velocity of the bridge. Since we are considering the bridge of a stationary process defined by a
and
, the distribution to be used in the expressions for the osmotic and current velocity is
while the drift of the bridge sde is
Thus, the drift of the reverse bridge is as follows:
Now, we consider the case
and compute the difference between the forward and reversed drift, which is a measure of the bridge asymmetry with respect to the time reversal
:
that can be also expressed as follows:
where we have inserted the stationary joint distribution
and
is the stationary current velocity of the original process:
Again, note that if the process satisfies detailed balance, both terms in Equation (
32) vanish; see [
37] for a review of the literature where this natural sufficient condition has been previously discussed. A necessary condition is, however, more important: it may happen that, even in the case of zero asymmetry, the two terms are (necessarily) both different from zero but their sum is equal to zero.
6. Examples for Asymmetry of Bridges
In this section, we compute the asymmetry in some simple processes. First, we consider a generic linear process; then, we consider the two-dimensional case, in particular the so-called Brownian Gyrator (BG), which has been introduced as a minimal model for a heat engine at the nano-scale [
47,
48,
49,
50]. In fact, we consider two versions that are not always equivalent: a thermal BG, where the engine is powered by two thermal baths at different temperatures, and a mechanical BG, where a single thermal bath is present, but a non-conservative force feeds the engine. The BG is a Gaussian process and therefore is analytically solvable but can sometimes be overly simplistic. In order to go beyond the Gaussian regime, we also consider a related non-Gaussian process, which is obtained by considering the BG in polar coordinates. In the isotropic case, the stochastic dynamics of the squared modulus are equivalent to those of the CIR [
63] process introduced in finance, which in turn is strictly related to the ABBM model [
64], a paradigm for the “crackling noise” in several natural phenomena [
65], as well as a “mean field” model for depinning dynamics of disordered elastic manifolds [
51]. Since the model can be solved exactly, the analysis of bridge and excursion (which share the same average shape) can be performed analytically [
66]. Hereafter, we consider a non-equilibrium extension.
6.1. Generic Linear Process
We consider the case of the
d-dimensional OU process defined as follows:
where
x is a
d dimensional vector,
W is a
d dimensional Wiener process, and
and
are two constant square matrices, the former having eigenvalues with positive real parts and the latter being positive definite. The two conditions are dictated by typical applications and guarantee the existence of a non-degenerate, asymptotic stationary state.
The propagator of the process is a multivariate Gaussian distribution:
where · is the scalar product between vectors, the matrix
gives the evolution of the mean
, and
gives the covariance of the process
An explicit computation shows that
From the last one can be derived the following implicit equation:
Moreover, it is easy to show that the time evolution of
and
is ruled by the following equations:
We want to compute the bridge asymmetry defined in Equation (
31), which now reads:
Now, let us proceed evaluating the
. It turns out that it reads as follows:
where
. Using the usual property of the scalar product (
), the quadratic terms in
x reduce to
The gradient of this, using
, and
reads
The linear terms, on the other hand, are
whose gradient, since
, is
The remaining terms do not depend on
x and they are wiped out by the
derivatives in the asymmetry expression Equation (
32), which now becomes
Imposing a zero asymmetry for the
d-dimensional OU process amounts to requiring
for every
x,
, and
t, which gives
Equation (
43) implies the Onsager relations; see for instance Section 5.3.6c of [
2], while Equation (
44) means that
is independent from
t, so
On the other hand, the expression for the bridge covariance is
and Equation (
45) implies that
, as expected. In
Appendix A we compute explicitly the asymmetry
for some specific cases of BG in Cartesian coordinates.
6.2. Two-Dimensional BG in Polar Coordinates: An Interesting Non-Linear Process
Let now focus on the two-dimensional case, which, without loss of generality, can be parameterized as follows:
and
We want to describe the process in polar coordinates, defined as follows:
Using Ito’s Lemma, it is easy to derive the sde for the new process
:
The process
, at odds with
, is not Gaussian, as witnessed by the multiplicative (and singular) diffusion coefficients in the sde.
Note that the equation for is always coupled with v. On the other hand, the equation for is coupled to only when the system is anisotropic, that is, for (which makes the noise terms correlated) or for anisotropic potential, that is, for or .
6.3. Mechanical Case
The expression for the asymmetry for a general bridge from to is quite cumbersome. Here, we consider the case of equal temperature , but a non-conservative force, that is .
For
, the correlation between the noise terms is only apparent, since
are two independent Wiener processes, whatever the angle
:
In this case, without loss of generality, we can always perform a rotation of the Cartesian coordinates in order to obtain a drift with equal diagonal terms
(here,
u is the rotated value, not the original value in Equation (
46)), since the square modulus
v is not affected by the rotation, while
changes by an irrelevant additive constant, which can be absorbed in the initial conditions. In this case, the anisotropy is represented only by
u and the equation simplifies as follows:
We call this case the “mechanical Brownian gyrator” because the origin of the probability current is only in the forces (
) and not in the anisotropy of the thermal baths.
It is easy to show that the distribution and the propagator of the process
can be expressed in terms of the Gaussian propagator of the process in Cartesian coordinates with average
and covariance
, as given previously, where now the initial condition coordinates of
are
and
:
(with the factor
coming from the Jacobian
.
Using Equation (
32), the bridge asymmetry for the process
in the isotropic case (
) takes a very simple form:
Considering the case of the bridge from the origin to the origin, that is, for
(the value of
is irrelevant at the origin), the asymmetry reduces to
Note that in this case, the asymmetry for the
v component is zero, despite the dynamics being out of equilibrium. This can be understood by considering the sde of the process in the isotropic case
:
As a consequence of the isotropy of the drift term, the equation for the squared modulus
v does not depend on
(while
is a non-Markovian process slaved to the process
v).
As previously mentioned, Equation (
62) describes a CIR process [
63]
under an appropriate choice of parameters:
which is rigorously equivalent to an ABBM process [
66].
Due to this partial decoupling, the
v component of the stationary current is zero. This can be shown considering the expression of the propagator of the process
from which the stationary distribution of can be computed as the asymptotic limit:
Note that
does not depend on
. Using the expression for
, one obtains the stationary velocity current:
So, while the full process has a non-zero current, the restricted process
is a Markovian process with zero current (hence satisfying detailed balance). The propagator of the
process can be computed marginalizing the full propagator
that does not depend on
, as follows:
where
is the modified Bessel function of the first kind of index 0.
Coming back to the general anisotropic case, we consider the bridge from the origin to the origin, that is,
, and we obtain for the asymmetry the following expression:
where
and
It easy to see that
in Equation (
66) recovers the isotropic case for
.
6.4. Thermal Case
We briefly discuss the much more complicated case in which
, but
. The computations in this case rapidly become cumbersome, so we restrict our discussion to the case
where the potential is anisotropic but the axis of the elliptic equipotential lines are oriented to
in the Cartesian axes.
The expression for the asymmetry
is too long to be shown here, but we can write out the expression for
and small values of
while for small
u,
Note that for
(isotropic potential), the asymmetry is zero, since the two degrees of freedom are uncoupled, and each one equilibrates with the respective thermal bath.
7. Conclusions
We have summarized the core elements for constructing a theory of stochastic bridges and their time-reversal statistics, aiming to derive a general formula that quantifies the asymmetry under time reversal in avalanche-like fluctuations. This asymmetry holds potential for inferring thermodynamic properties from experimental data. Indeed, entropy production, arising from external forcing and energy dissipation in physical systems, manifests in fluctuations as a violation of detailed balance, as consistently observed as an asymmetry in the stochastic bridges we analyzed. Our primary result, Equations (
31) and (
32), suggests the possibility of a cancellation between terms, even when time-reversal symmetry is broken, although we have yet to encounter such examples. Such a formula, when applied to a Brownian gyrator in a generic dimension, Equation (
42), demonstrates that the symmetry of a bridge connecting the origin with itself always corresponds to the condition of detailed balance. As a non-Gaussian process, the BG in
was analyzed in polar coordinates. The problem was split into two cases: the mechanical case (equal temperatures and an applied external torque) and the thermal case (no torque but different temperatures). In the mechanical case, the asymmetry of the polar bridge connecting the origin with itself takes the form given in Equation (
66), which is partially symmetric (the magnitude component is zero) in the case
, even when detailed balance is broken. In the thermal case, on the contrary, the expression for the asymmetry has been made explicit for small departures from equilibrium. At first order, these expressions can become empty only when detailed balance is satisfied (equal temperatures), but no rigorous conclusions are available at this moment.
Finally, we emphasize that “partial information” implies not only limited observability of the asymmetry’s components but also a restriction of the bridge’s definition to a subspace of initial and final conditions. This constraint arises because the boundary conditions can only be imposed on the observable degrees of freedom. The exploration of “partial bridges”—trajectories initiating and terminating within a subspace rather than at single points—will be a subject of our future research.