1. Introduction
The classical notions of asymptotic and exponential stability in dynamical systems theory involve system trajectories approaching a Lyapunov stable equilibrium state over an infinite time horizon. However, in many engineering and scientific applications, it is necessary for the system to reach a stable equilibrium within finite time rather than asymptotically. Achieving finite time stability in deterministic systems requires the closed-loop system dynamics to be non-Lipschitz, which can result in nonuniqueness of solutions in backward time. Nonetheless, it is possible to preserve forward time uniqueness of solutions for finite time convergent systems.
Sufficient conditions for forward time uniqueness in deterministic systems without assumption of the Lipschitz continuity of the system dynamics are given in [
1]. For stochastic systems, refs. [
2,
3,
4] provide criteria that guarantee the forward time uniqueness of solutions without requiring Lipschitz continuity. These works also show that if the dynamics are continuous and forward time uniqueness holds, then the system trajectories are almost surely continuous with respect to the system initial conditions, even in the absence of the Lipschitz continuity of the drift and diffusion functions characterizing the stochastic dynamical system.
The concept of
finite time stability, that is, convergence of the system trajectories to a Lyapunov stable equilibrium in finite time, was first introduced by Roxin [
5] and further developed in [
6,
7] for time-invariant deterministic systems, as well as in [
8,
9,
10] for time-varying deterministic systems. In particular, Lyapunov and converse Lyapunov theorems for finite time stability were established using Lyapunov functions that satisfy scalar differential inequalities with fractional powers, and it was shown that the regularity of the Lyapunov function depends on the properties of the settling time function, which characterizes the finite time convergence behavior of the system.
Even though extensions of finite time stability for stochastic systems have been addressed in the literature [
11,
12,
13,
14], several of these results contain errors. Specifically, as pointed out in [
15], several definitions and the main result in [
11] are incorrect. Moreover, the authors in [
13] used the results of [
11] to develop partial-state stabilization in finite time, propagating the errors of [
11] in their work, whereas the proof of the main theorem of [
12] used Jensen’s inequality incorrectly, thus invalidating their result. And finally, ref. [
14] failed to provide a bound on the expectation of the settling time function, which is crucial in providing a complete theory for finite time stability and stabilization for stochastic dynamical systems.
In this paper, we correct these oversights to present a self-contained theory for finite time stability in probability and build upon the framework established in [
16,
17] to address the problem of optimal finite time stabilization for
stochastic nonlinear systems. Specifically, we ensure finite time stability in probability for the closed-loop system using a Lyapunov function that satisfies a scalar differential inequality that involves fractional powers. This Lyapunov function is shown to correspond to the steady-state solution of the stochastic Hamilton–Jacobi–Bellman equation, ensuring both finite time stability in probability and optimal performance. Finally, we also develop connections of our approach to inverse optimal control [
18,
19] by constructing a family of finite time stabilizing stochastic feedback laws that minimize a derived cost functional.
2. Mathematical Preliminaries
We will start by reviewing some basic results on nonlinear stochastic dynamical systems [
20,
21,
22,
23]. First, however, we require some notations and definitions. The notation, definitions, and mathematical preliminaries in this section are adopted from [
17]. A
probability space is a mathematical construct that provides a model for a random experiment and consists of the triple
. The
sample space is the set of all possible outcomes of the experiment. The
event space is a collection of subsets of the sample space, where each subset represents an event. The event space
has the algebraic structure of a
σ-algebra. The pair
is a measurable space, and the function
defines a
probability measure on the
-algebra
, assigning a probability to each event in the event space
. A
complete probability space is one in which the
-algebra
includes all the subsets of sets with a probability measure zero.
A Borel set is a set in a topological space that is derived from open (or closed) sets through the repeated operation of countable unions, countable intersections, and relative complements. The Borel σ-algebra on , denoted by , is the smallest -algebra containing all the open sets in . An -valued random vector x is a measurable function , that is, for every Borel set , its preimage is . A random variable is a scalar random vector. In general, given a random vector x and a -algebra , we say that x is -measurable if for all . Given an integrable random vector x and a sub--algebra , and denote, respectively, the expectation of the random vector x and the conditional expectation of x given under the probability measure .
A continuous-time stochastic process is a collection of random vectors defined on the probability space , and it is indexed by the set of nonnegative real numbers that are greater than or equal to 0. Occasionally, we write for to denote the explicit dependence of the random variable on the outcome . For every fixed time , the random variable assigns a vector to every outcome , and for every fixed , the mapping generates a sample path (or sample trajectory) of the stochastic process , where, for convenience, we write to denote the stochastic process .
A (continuous-time)
filtration on
is a collection of sub-
-fields of
, and it is indexed by
, such that
,
. A filtration is
complete if every
contains the set of all sets that are contained in a
-null set. The stochastic process
is
progressively measurable with respect to
if, for every
, the map
defined on
is
-measurable. The stochastic process
is said to be
adapted with respect to
, or it is simply
-adapted if
is
-measurable for every
. An adapted stochastic process with right continuous (or left continuous) sample paths is progressively measurable [
24].
The stochastic process
, where
, is a random variable, is a
martingale with respect to the filtration
if it is
-adapted,
, and
. If we replace the equality in
with “≤” (respectively, “≥”), then
is a
supermartingale (respectively,
submartingale). A random variable
is called a
stopping time of the filtration
if
for all
. A stopping time
is a
bounded stopping time if the event
, where
is a constant, has a probability of one. (For an additional discussion on stochastic processes, filtrations, martingales, and stopping times, see [
24]).
In this paper, we consider stochastic dynamical systems
of the form
where (
1) is a
stochastic differential equation. The stochastic process
represents the system state,
is a random system initial condition vector, and
is a
d-dimensional
Brownian motion process. For every
, the random variable
takes values in the state space
. The Borel measurable mappings
and
satisfy
and
, and they are known as the system
drift and
diffusion functions. The stochastic differential equation (
1) is interpreted as a way of expressing the integral equation
where the first integral in (
2) is a Lebesgue integral and the second is an Itô integral [
25].
Let
be a fixed complete filtered probability space,
be a
-adapted Brownian motion, and
be a
-measurable initial condition. A
solution to (
1) is an
-valued
-adapted process
with continuous sample paths such that the integrals in (
2) exist and (
2) holds almost surely for all
. For a Brownian motion disturbance and initial condition given in a prescribed probability space, the solution to (
2) is known as a strong solution [
26]. In this paper, we focus on strong solutions, and we will simply use the term “solution” to refer to a strong solution. A solution to (
1) is
unique if for any two solutions
and
that satisfy (
1),
for all
almost surely.
For a stochastic system of the form given by (
1) with solution
, the
(infinitesimal) generator of
is an operator acting on the function
, and it is defined as ([
22])
where
denotes the expectation given that
x is a fixed point in
. The set of functions
for which the limit in (
3) exists is denoted by
. If
has compact support, where
denotes the space of functions with
r-continuous derivatives, then
and
, where
, and
[
22]. Note that the differential operator
introduced in (
4) is defined for every
, and it is characterized by the system drift and diffusion functions. With a minor abuse in terminology, we will refer to the differential operator
as the (infinitesimal) generator of the system
.
If
, then it follows from
Itô’s formula [
26] that the stochastic process
satisfies
If the terms appearing in (
5) are integrable and the Itô integral in (
5) is a martingale, then it follows from (
5) that
A more general version of (
6) that holds for
t is replaced by a stopping time
is known as
Dynkin’s formula [
22].
We say that a function
is of
polynomial growth if there exist positive constants
C and
m such that
For
, where
denotes the set of natural numbers, we write
if
V and all its partial derivatives up to order
r are of polynomial growth. As shown in [
27],
is a sufficient condition for the integrability of the terms in (
5). In this case, it can be shown that (
3) implies (
4). In this paper, we assume all Lyapunov functions
V are of polynomial growth with
.
Given a function , we say that V is positive definite with respect to if and . If , then V is positive definite with respect to the origin, or V is simply positive definite. Moreover, we say a function V is nonnegative definite if . We say that V is negative definite if is positive definite. In addition, we say that V is radially unbounded if .
As discussed in the introduction, in order to achieve convergence in finite time for stochastic dynamical systems, the drift and diffusion functions characterizing the system dynamics need to be non-Lipschitzian giving rise to nonuniqueness of solutions in backward time ([
20], Lemma 5.3). Uniqueness of solutions in forward time, however, can be preserved in the case of finite time convergence. The next result establishes the existence and uniqueness of solutions for the stochastic differential equation (
1) with non-Lipchitzian drift and diffusion functions. For the statement of this result,
, denotes the
open ball centered at
with radius
in the Euclidean norm.
Theorem 1 ([
4]).
Consider the nonlinear stochastic dynamical system (
1)
with initial condition such that , and assume that the following conditions hold: - (i)
Continuity. and are continuous.
- (ii)
Linear growth.
There exists a constant such that, for all ,- (iii)
For every , there exist a strictly increasing, continuous, and concave function , as well as a constant such that, for all , Then, there exists a unique solution to (
1).
Furthermore,
Assumption 1. For the remainder of this paper, we assume that the conditions for existence and uniqueness given in Theorem 1 are satisfied for system (1). 3. Finite Time Stability for Stochastic Dynamical Systems
In this section, we introduce the notion of finite time stability in probability and present sufficient conditions for finite time stability of (
1) using a Lyapunov function that satisfies a scalar differential inequality involving fractional powers. First, however, we require some additional definitions and results. We denote the solution
to (
1) with the initial condition
at time
t by
, and we define
and
, with
denoting the sample trajectory of (
1). Thus, for every
, there exists a trajectory defined for all
and
satisfying the dynamical process (
1) with initial condition
. For simplicity of exposition, we write
for
and
for
omitting their dependence on
.
The following definitions introduce the notions of a stochastic settling time and finite time stability in probability for stochastic dynamical systems. Here, we assume that the initial condition is a constant, and hence, whenever we write , we mean that is a constant vector. In this case, we will find it convenient to introduce the notation and to denote the probability and expected value, respectively, given that the initial condition is the fixed point almost surely.
Definition 1. Consider the nonlinear stochastic dynamical system given by (
1).
The stochastic settling time
is a stopping time with respect to , and is defined as Note that if , then . For simplicity of exposition, we write for , omitting its dependence on .
Definition 2. Consider the nonlinear stochastic dynamical system given by (
1)
and let be as in Definition 1, then the zero solution to (
1)
is finite time stable in probability
if the following statements hold: - (i)
The zero solution to (
1)
is Lyapunov stable in probability,
that is, for every , we have or, equivalently, for every and there exists such that, for all , - (ii)
For every , there exists such that if , then
The zero solution to (
1)
is globally finite time stable in probability
if for all .
Proposition 1. Consider the nonlinear stochastic dynamical system (
1).
Assume that the zero solution to (
1)
is globally finite time stable in probability. Let be as in Definition 1, then, for all , Proof. Since (
1) is globally finite time stable,
and
for all
and almost all
. Defining
as
and using the fact that
is sample continuous in
t, it follows that
which implies
is sample continuous in
t. Since
is
-adapted, it follows that
is
-adapted. Clearly,
satisfies (
2) for all
.
Next, using the fact that
and
, it follows that, for all
,
and hence, it follows that (
2) is satisfied for all
. Now, since the solution to (
1) is unique,
is the only solution to (
1) with
. Finally, the result follows by noting that this holds for almost all
. □
Proposition 1 implies that, for every
and
, there exists
such that, for all
,
and
which are equivalent. Hence, it follows from Proposition 1 that if the zero solution
to (
1) is globally finite time stable in probability, then it is globally asymptotically stable in probability. Thus, global finite time stability in probability is a stronger notion than global asymptotic stability in probability.
The following theorem based on the results appearing in [
14,
28] gives sufficient conditions for stochastic finite time stability using a Lyapunov function involving a scalar differential inequality. For completeness, we give a self-contained proof of this result as it forms the foundation for all later developments in this paper. For the statement of this theorem,
denotes the indicator function of the set
, that is,
Theorem 2. Let be an open subset containing the origin. Consider the nonlinear stochastic dynamical system (
1)
and assume that there exist a two-times continuously differentiable function and constants and such thatThen, the zero solution to (
1)
is finite time stable in probability. If, in addition, , V is radially unbounded, and (
19)
and (
20)
hold on , then the zero solution to (1) is globally finite time stable in probability. Moreover, there exists a stochastic settling time such that Proof. Conditions (
18)–(
20) imply Lyapunov stability in probability by Theorem 2 of [
17]. Thus, for every
and
such that
, there exists
such that, for all
,
Next, note that finite time stability in probability holds trivially for
. For all
, define the stopping times
,
,
, and
, where
,
and
.
Since
is two-times continuously differentiable for all
such that
, it follows from Itô’s formula [
22] that, for all
and
,
The process
is
-measurable and
-adapted due to the measurability of the mappings involved and the properties of the processes
. Now, since
V is continuously differentiable and
D is continuous, it follows that, for all
and
,
and hence, by Corollary 3.2.6 of [
22], the Itô integral in (
23) is a martingale. Now, (
23) and the properties of the expectation of a martingale yield
Note that the expectations in (
25) are well defined because
is a stopping time for exiting the bounded set
and the integrands are continuous in
.
Next, it follows from (
19), (
20), and (
25) that
Since (
26) holds for all
, it also holds for
, implying that
We claim that
To see this, note that, since the stopping times
are increasing with
k,
exists or is infinite for almost all
. Next, since
is sample continuous,
for all
, which yields
Now, for almost all
, (
27) holds trivially; moreover, for some
, if
, then (
28) implies (
27). Alternatively, if
, then assume that,
ad absurdum,
. It follows from the definition of
that
which implies
Next, since
is sample continuous in
t, we conclude
which contradicts
, and hence, (
27) holds for almost all
.
Now, it follows from Fatou’s lemma [
29] that
which implies that
proving finite time stability in probability.
To show global finite time stability in probability, for every
, define the stopping times
and
, where
and
, such that
. Using a similar argument as the one used to obtain (
26) with
, replacing
yields
Since (
31) holds for all
, it also holds for
, which implies
Next, we claim that
Since the stopping times
are increasing with
l,
exists or is infinite for almost all
. Now, using an analogous argument as the one used to obtain (
28) gives
Next, for almost all
, (
33) implies (
32). Alternatively, for some
, assume that,
ad absurdum,
. Since (
18)−(
20) hold and
V is radially unbounded,
applies by Theorem 2 of [
17]. Thus, there exists
such that
for all
, and hence,
which implies
Now, since
is sample continuous in
t, we conclude
which contradicts
, and hence, (
32) holds for almost all
.
Finally, it follows from Fatou’s lemma [
29] that
and hence,
which proves global finite time stability in probability. □
4. Stochastic Optimal Finite Time Stabilization
In the first part of this section, we provide connections between the Lyapunov functions and nonquadratic cost evaluation. Specifically, we consider the problem of evaluating a nonlinear–nonquadratic performance measure that depends on the solution of the stochastic nonlinear dynamical system given by (
1). In particular, we show that the nonlinear–nonquadratic performance measure
where
and
,
, satisfies (
1) with
, can be evaluated in a convenient form so long as (
1) is related to an underlying Lyapunov function that proves finite time stability in the probability of (
1).
The following theorem generalizes Theorem 6 of [
17] to finite time stability.
Theorem 3. Consider the nonlinear stochastic dynamical system given by (
1)
with the nonlinear–nonquadratic performance measure (
36),
where is the solution to (
1).
Furthermore, assume that there exist a two-times continuously differentiable radially unbounded function and constants and such thatwhere . Then, the zero solution to (1) is globally finite time stable in probability. Moreover, there exists a stochastic settling time such thatand Proof. Global finite time stability in probability along with the existence of a stochastic settling time
such that (
42) holds are a direct consequence of (
37)–(
39) and
V being radially unbounded by Theorem 2. Using the fact that
V is continuous and
yields
almost surely.
Next, we show that the stochastic process
is a martingale. To see this, first note that the process
is
-measurable and
-adapted because of the measurability of the mappings involved and the properties of the process
. Now, using Tonelli’s theorem [
29] it follows that, for all
,
for some positive constants
and
, and hence, by Corollary 3.2.6 of [
22] the Itô integral
is a martingale. To arrive at (
45) we used the fact that
, the linear growth condition (
8), and the finiteness of the expected value of the supremum of the moments of the system state (
10). Note that the supremum in (
45) exists because of the continuity of the sample paths of
.
Next, the measurability of the solution of (
1) implies that
is
-adapted and
exists since
and (
10) holds. Now, using Itô’s formula [
22] and (
39) we have, for every
and almost every
,
Since the Itô integral in (
46) is a martingale, (
46) implies
which shows that the process
is a nonnegative supermartingale. An analogous argument implies that the process
is also a nonnegative supermartingale for any
.
Next, using the fact that supermartingales have decreasing expectations gives
which shows that the stochastic process
is uniformly integrable [
22] (p. 323). Thus, by Doob’s martingale convergence theorem [
22],
converges in
, and hence,
Furthermore, it follows from (
41) and Itô’s formula that, for all
,
Taking the expected value operator on both sides of (
50) and using the martingale property of the stochastic integral in (
50) yields
Now, taking the limit as
and using (
49) yields
Finally, note that
where the interchanging of the limit with the expectation operator in the second equality in (
52) follows from the Lebesgue monotone convergence theorem [
30] by noting that, by (
39) and (
41),
, and hence,
, is monotone increasing, and thus, converges pointwise to
. □
Next, we use the framework developed in Theorem 3 to obtain a characterization of the stochastic optimal feedback controllers that guarantee closed-loop, finite time stabilization in probability. Specifically, sufficient conditions for optimality are given in a form that corresponds to a steady-state version of the stochastic Hamilton–Jacobi–Bellman equation. To address the problem of characterizing finite time stabilizing feedback controllers, consider the nonlinear controlled stochastic dynamical system
where the stochastic process
represents the controller input, and
is the set of admissible control inputs. For every
, the random variables
and
take values in the state space
and the control space
, respectively. The mappings
and
satisfy (
8) and (
9), with
and
, uniformly in
u, and
and
.
We assume that every
is an
-valued Markov control process. An input process
is a
Markov control process if there exists a function
such that
. Note that the class of Markov controls encompasses both time-varying inputs (i.e., possibly
open-loop control input processes) and state-dependent inputs (i.e., possibly a
state feedback control input, where
is a
feedback control law). Note that, given the feedback control law
, the
closed-loop system (
53) takes the form
The following theorem generalizes Theorem 7 of [
17] to optimal finite time stabilization.
Theorem 4. Consider the nonlinear stochastic dynamical system given by (
53)
with nonlinear-nonquadratic performance measurewhere is the solution to (53) with control input . Furthermore, assume that there exist a two-times continuously differentiable function , constants and , and a feedback control law such thatwhere andThen, with the feedback control law , the closed-loop system (
54)
is globally finite time stable in probability. Moreover, there exists a stochastic settling time such thatandIn addition, the feedback control law minimizes (
55)
in the sense thatwhere denotes the set of controllers given bywhere and . Proof. Global finite time stability in probability along with the existence of a stochastic settling time
such that (
64) holds are a direct consequence of (
56)−(
59) by applying Theorem 2 to the closed-loop system (
54).
To show that
, note that since
is Borel measurable,
is
-progressively measurable and (59) and (61) imply that
Thus,
Now, using an analogous argument as in the proof of Theorem 3, it follows that
for
, and hence,
.
Next, let
, and note that, by Itô’s lemma [
22],
It now follows using an analogous argument as the one in the proof of Theorem 3 that the stochastic integral in (
69) is a martingale, and hence,
where
exists since
and (
10) holds. Next, taking the limit as
yields
Since
, the control law satisfies
, and hence, it follows from (
71) that
Now, combining (62) and (
72) yields
Next, note that, for every
,
and, since
,
. Thus, it follows from the dominated convergence theorem [
31] that
Finally, combining (
65), (
73), and (
75) yields
which proves (
66). □
Observe that (61) represents the steady-state form of the stochastic Hamilton–Jacobi–Bellman equation. To see this, recall that the general form of the stochastic Hamilton–Jacobi–Bellman equation is given by
which serves as the fundamental condition for optimal control in stochastic, time-dependent systems over either finite or infinite horizons [
21]. When the system is time-invariant and considered over an infinite time horizon, the value function becomes stationary, that is,
; as a result, (
77) simplifies to (61) and (62). These equations ensure optimality within the class of admissible control policies
. Notably, it is not necessary to explicitly characterize this admissible set
, and the optimal feedback control law
is independent of the specific initial state
.
To guarantee that the closed-loop system given by (
54) is globally finite time stable in probability, Theorem 4 requires that the value function
V satisfy Conditions (
56), (57), and (59). These requirements ensure that
V serves as a Lyapunov function for the closed-loop system. However, these Lyapunov conditions are not necessary for establishing optimality. In particular, if
V is twice continuously differentiable and
, and if the control signal
lies in the admissible set
, then satisfaction of (61) and (62) leads to the fulfillment of (
65) and (
66). It is also crucial to emphasize that, in contrast to the deterministic framework [
16], p. 857, establishing optimality in the stochastic setting necessitates an additional condition, namely the transversality condition
as stated in (
67). (For more detailed discussions on this requirement, the reader is referred to [
32], p. 323, [
33], p. 125, and [
34], p. 139.)