and Lyapunov Approaches for Discrete-time Peak Computation Problems
Abstract
In this paper, we propose a method to solve discrete-time peak computation problems (DPCPs for short). DPCPs are optimization problems that consist of maximizing a function over the reachable values set of a discrete-time dynamical system. The optimal value of a DPCP can be rewritten as the supremum of the sequence of optimal values. Previous results provide general techniques for computing the supremum of a real sequence from a well-chosen pair of a strictly increasing continuous function on and a positive scalar in . In this paper, we exploit the specific structure of the optimal value of the DPCP to construct such a pair from classical tools from stability theory: certificate and Lyapunov functions.
1 Introduction
In this paper, we propose to solve discrete-time peak computation problems in finite-dimensional state spaces. A discrete-time peak computation problem can be viewed as a particular maximization problem for which the constraint on the decision variable is to belong to the reachable values set of a given discrete-time dynamical system. The objective function of this optimization problem is a function of the states. This particular maximization problem can model various situations in engineering, economics, physics, biology… For example, in population dynamics, for a predator-prey system, one may be interested in the maximal size of each species, even if the system asymptotically oscillates between several equilibria. This problem boils down to maximizing each coordinate of the state variable separately over the system modeling the population dynamics.
Besides its importance in analyzing critical concrete situations, the literature on discrete-time peak computation problems in the communities of dynamical systems analysis and optimization is quite thin. Discrete-time peak computation problems appear for linear systems, where the objective function is the Euclidean norm in [APS18]. This problem also appears in the analysis of stable linear systems with general quadratic objective functions in [Adj21] and in [Adj25a]. A more constrained version of the problem appears in [AG24]. The additional constraint for the state variable is to live in a given set. Discrete-time peak computation problems have a continuous-time counterpart (see e.g., [MHS20]). Note that in [MHS20], the authors address the problem of verifying safety properties (see, e.g., [BYG17] for verification problems on discrete-time systems) for dynamical systems and formulate the safety verification problem as a continuous-time peak computation problem. Safety verification problems aim to prove that the trajectories of a system are contained in a safe set. A safety verification problem can be translated into a peak computation problem when the safe set is a sublevel of sufficiently regular function. In this case, the objective function of the peak computation problem is the function associated with the sublevel set. A safety analysis from the peak computation point of view is also proposed for switched systems in [Adj17] and for program analysis in [AGM15]. For the safety analysis of dynamical systems, one can think that a reachability analysis (see e.g., [RKML06] and references therein), equivalent to a feasibility analysis in our context, is sufficient to prove some properties of the dynamical system. However, maximality in our context is synonymous with criticality. The optimal value in this situation can be viewed as the value that penalizes the system the most with respect to some criteria and defines the extreme conditions under which the system could enter.
In [MHS20], the authors are interested in computing an upper bound of the peak optimal value for the continuous-time case. The techniques used are based on Sums-Of-Squares (SOS) and Linear Matrix Inequalities (LMI) formulations (the interested reader can consult [Las15] for those subjects). The same goal motivates the authors of [APS18]. They limit their study to linear systems and norm objective functions they directly use semidefinite programming (see [BTN01] for further details on the subject). The purpose of the current paper is to develop a method for computing the exact value of a discrete-time peak computation problem. This purpose is shared by the authors of [AG24]. However, this latter cannot be fairly compared because of the additional constraint. Moreover, the studied systems are linear or piecewise linear, and the objective functions are linear.
The goal of the current paper is to provide a theoretical generalization of the works in [Adj21] and [Adj25a]. In [Adj21] and in [Adj25a], the resolution of discrete-time peak computation problems is restricted to stable affine systems and quadratic objective functions. In this paper, we consider wider classes of stable systems and general objective functions. The techniques developed in this paper combine the method proposed in [Adj25b] with the Lyapunov approach of [Adj25a]. In [Adj25b], the method aims to find the maximal term of a real sequence using homeomorphisms of convergent positive geometric sequences. In [Adj25a], homeomorphisms and geometric sequences are constructed from quadratic Lyapunov functions. In this paper, we also propose a construction of homeomorphisms and geometric sequences from bounds. The use of such classical tools of dynamical systems stability theory (the interested reader can consult [Ela05, Chap. 4] for discrete-time stability notions) to solve an optimization problem appears to be unnatural. This is a consequence of a simple observation: the existence of a maximizer for a real sequence depends on its asymptotic behavior. More precisely, the existence of a maximizer relies on the finiteness of the limit superior and on the existence of a term strictly greater than the limit superior. Hence, tools for studying the asymptotic behavior of the system help in the existence of a peak and its characterization. Although general stability properties can be difficult to obtain, asymptotic stability properties can be easily presented and understood from bounds (e.g., [ATP10] or [NT04]), that is, the norms of the states are globally bounded from above by a function. However, verifying that the function is an upper bound over the norms of the states must be done for all integers, which can be difficult. In contrast, Lyapunov functions are functional certificates for system stability. Some results guarantee the existence of Lyapunov functions: a system is globally asymptotically stable if and only if there exists a Lyapunov function (see, e.g., [Kha02]). Being a Lyapunov function is a timeless property that depends only on the dynamics. In a few advantageous situations, Lyapunov functions can be computed using numerical optimization solvers (see the survey [GH15]).
The paper is organized as follows. Section 2 presents the discrete-time peak computation problems addressed in this paper. In Section 2, we also present a running example that illustrates the techniques developed in this paper. Section 3 recalls the useful results obtained in [Adj25b]. In Section 3, we also provide basic definitions of comparison functions. Then, in Section 4, we develop an approach based on bounds. In Section 5 we develop the approach using general Lyapunov functions. In these two sections, we explain how to use and Lyapunov functions to solve a discrete-time peak computation problem. We also apply these techniques to the running example presented in Section 2. Section 6 concludes the paper and discusses future research.
Notations: stands for the set of reals, the set of nonnegative reals, the set of nonzero reals and the set of strictly positive reals. The vector space of vectors of reals is denoted by . The set of natural integers is denoted by whereas denotes the set of nonzero natural integers. The norm stands for the Euclidean norm. Finally, denotes the closed ball of radius centered at zero i.e., .
2 Problem formulation
2.1 Problem formulation
Let us consider a discrete-time dynamical system on where
-
•
is the nonempty subset of of initial conditions;
-
•
is a nonzero self-map on that updates the state-variable.
The system is autonomous; it evolves without any control, disturbance, or external input. In this paper, we are interested in the reachable values that maximize a given function. Then, given , we define the discrete-time peak computation problem as follows:
(DPCP) |
where is the -fold composition of if is not null, and the identity, if .
Like any optimization problem, to solve a problem of form (DPCP), we must compute
-
•
its optimal value i.e., ;
-
•
its optimal solutions i.e., couples such that
An optimal solution of is thus composed of an initial condition leading to a peak and its associated date of peak realization.
Without loss of generality, we can assume that . Indeed, for the problem , the optimal value is equal to the sum of the optimal values of and . Moreover, the optimal solutions of and are equal.
The optimal value of can be described using a sequence of optimal values. Indeed, we can introduce a static classical optimization problem where is fixed:
() |
Then, we introduce the sequence defined, for all , by:
(1) |
and its supremum:
(2) |
From these notations, is equal to the optimal value of whereas is the optimal value of the static optimization problem (). Solving boils down to compute , an integer such that and an optimal solution of . We concentrate our efforts on the computation of the integer as the computation of the optimal solution of relies on an optimization solver (when a suitable one is available). The approach proposed in this paper consists in computing a stopping integer for .
Definition 1 (Stopping integer).
An integer is a stopping integer for if, for the sequence defined in Equation (1):
If there is no ambiguity, we will simply say that is a stopping integer for .
Obviously, any stopping integer for is greater than and is the smallest stopping integer for . Following [Adj25b], if the set is nonempty, a stopping integer for can deduced from the formula:
(3) |
where is such that
-
1.
is strictly increasing and continuous on ;
-
2.
;
-
3.
for all , ;
-
4.
there exists such that .
In this paper, we extend the techniques developed in [Adj25a] for stable affine systems and quadratic objective functions to nonlinear stable systems and nonquadratic objective functions. In [Adj25a], we construct from quadratic Lyapunov functions. In this paper, we propose to construct from any stability certificate and any Lyapunov function associated with the system . The interested reader can consult [Kel15] and references therein for complementary readings about those tools from stability theory.
2.2 Practical limitations
From a theoretical perspective, static optimization problems () are classical optimization problems; consequently, their study and resolution are relegated to optimization theory. For example, when is concave and is convex, the problem () is said to be convex, and numerous methods exist (see e.g., [BTN01, Ber15]), or when is polynomial and a basic semialgebraic set (see e.g., [AL11, Las15]), the problem () falls into the class of polynomial optimization problems.
In practice, the most critical challenge for the computation of lies before the optimization procedure. Indeed, when is nonlinear, it is difficult to obtain a closed form for . For example, for logistic sequences, closed forms are only known for particular cases (see e.g., [GRR10]). Some specific difference equations or systems require intensive efforts to obtain closed forms (see e.g., [Els12, HR18, Ste14] and references therein). We could use instead at least an exact formulation of when is an infinite set. However, its computation is limited in practice. For example, in [MHL15], the authors approximate the polynomial image of a basic semialgebraic set using a sequence of semidefinite programs. The authors obtain the convergence in sense of the sequence of approximations to the polynomial image of the set.
Hopefully, when is nonlinear, there exist some situations making the computation of ”easier” in practice. For example, if is a finite set or is conjugate to a linear dynamics. Conjugate111In the literature, the term topologically conjugate is employed, but in this context, the bijection and its inverse are required to be continuous. In our case, continuity is not mandatory from an optimization perspective. means that there exists a bijection and a linear map identified with its matrix such that for all , . In this case, we simply have for all for all , .
In this paper, the theoretical part does not require finiteness concerning or . We only require the finiteness of . In our numerical illustrations, to concentrate on the novelty and avoid entering too deep into optimization tools, we will only consider finite initial condition sets . The consideration of more general initial condition sets in practice will be addressed in future works.
The developments made in this paper require real-valued sequences; hence the only assumption made in this paper is as follows:
Assumption 1.
For all is finite.
A simple situation where Assumption 1 holds when is a bounded set, maps from bounded sets to bounded sets, and is upper semicontinuous.
2.3 Running example
Throughout this paper, we experiment with our main tools on a nonlinear system found in [LHK14a]. The dynamics is defined as:
(4) |
To write a discrete-time peak computation problem, we also need an initial conditions set and an objective function. Before discussing the choice of the initial conditions set and the objective functions, it should be noted that starting outside a specific ball makes the norm of the state variable blow up. First, by a simple calculus, we obtain, for all :
Then, by introducing the following two functions:
(5) |
we get for all . Now as for all nonzero and since the assertion ( and ) is equivalent to , we conclude that if and only if . Hence, we introduce:
(6) |
Finally, we have for all , .
In future examples, an initial conditions set will be a subset of satisfying
(7) |
The exact definition of will depend on the statements that we want to highlight.
An objective function is also required. In the examples, we will analyze the behavior of each coordinate separately; thus, the objective functions are the coordinate functions:
(8) |
For an initial condition set satisfying the condition (7), we will define two discrete-time peak computation problems and . The optimal value of is written as
(9) |
The optimal value of is written similarly:
(10) |
The condition (7) for ensures that for all and all , and . This guarantees that Assumption 1 holds for the sequences
(11) |
3 Preliminary results
Before considering the main results of this paper, we recall some useful theoretical results. First, in Subsection 3.1, we bring details about the formula (3) obtained in [Adj25b]. Second, in Subsection 3.2, we introduce useful material about comparison functions, which constitute the underlying tool behind and Lyapunov stability.
3.1 Results on the supremum of a real sequence
The developments made in [Adj25b] concern general real sequences, and in this subsection, we consider for which we aim to compute:
(12) |
3.1.1 Notations and basic results
We introduce the subset of consisting of sequences bounded from above:
For an element of , we are interested in computing the supremum of its terms.
Proposition 1.
.
For , we define and we introduce, for , the two sets of ranks:
(13) |
and we define
(14) |
We recall that the infimum of the empty set is equal to . Hence, if and only if . Furthermore, if is nonempty, then .
Proposition 2.
Let . The following assertions hold:
-
1.
;
-
2.
;
-
3.
If then .
3.1.2 A formula based on pairs of strictly increasing continuous functions and convergent geometric sequences
In this subsection, we present how to construct the formula (3). All details can be found in [Adj25b]. The key tool to compute the supremum of a sequence is the existence of a particular sequence that is an upper bound on . The particularity comes from the fact that is strictly increasing and continuous on and . We then introduce the following set of real functions:
It should be noted that we can define a function that is strictly increasing and continuous on on a set containing and then extend it as we want on the entire real line. Therefore, sometimes, we will partially define elements of on a set greater than but not everywhere on the real line.
We also introduce, for the following sets:
and
Finally, we introduce for and :
As all elements of are strictly increasing and continuous on , they have an inverse on . For the sake of simplicity, we will simply denote by the inverse of . Moreover, the fact that the elements of are strictly increasing and continuous allows for simple properties.
Proposition 3.
Let . Then :
-
1.
For all , we have ;
-
2.
For all , and ;
-
3.
If there exists such that then .
We can also fully characterize the set from the nonemptiness of .
Proposition 4.
if and only if .
In the formula (3), the natural logarithm is defined when , which is the same as . Then, we must ensure that the function satisfies for some . This provides the notion of usefulness.
Definition 2 (Useful strictly increasing continuous functions).
Let such that . A function is said to be useful for if the set
(15) |
is nonempty. By extension, is useful for if is useful for .
From the first statement of Proposition 3, if is useful for , there exists such that i.e., .
The following result aims to prove that there exists an optimal useful affine function for when .
Theorem 1 (Existence of a useful optimal affine function).
Let such that . Then, there exist , and such that:
-
1.
;
-
2.
for all ;
-
3.
.
Corollary 1.
Let such that . Then, there exists a pair useful for such that .
Proposition 5.
Let . Then, if and only if there exists useful for .
Theorem 1 provides a theoretical result and to have in hand an optimal function useful for is very unlikely to happen. However, if we have any function , we can produce a stopping integer for which is, in fact, an upper bound of . To compute it, for , we introduce the function defined for all , and by:
(16) |
3.1.3 Properties of and the computation of the supremum of a sequence
The function defined in (16) has useful properties for computing a stopping integer for as shown in Theorem 2. To obtain the results of Theorem 2, we need auxiliary results.
Proposition 6.
Let us take a pair useful for and let . The following statements are true:
-
1.
is well-defined and strictly positive if and null if and ;
-
2.
;
-
3.
;
-
4.
For all , , ;
-
5.
Let . If then .
Proposition 7.
Let such that . The following properties hold:
-
1.
Let and . Let such that . Then .
-
2.
Let . Let . Let . Then .
The first statement of Proposition 7 can be read as follows: we must choose the smallest possible to reduce the approximation gap. The second statement has a similar meaning in a functional sense: if we have several functions in , we must consider the minimum functions to reduce the overapproximation gap.
Proposition 8.
Let and . Then for all , . Moreover, if is useful for , we have:
Theorem 2 (Solution for (12)).
Let with . Let be useful for and let . Then :
Theorem 3 (Formula optimality).
For all , we have the following formula:
and the can be replaced by when .
We conclude this subsection with a summary in the form of an algorithm (Algorithm 1) to solve the problem (12).
Algorithm 1 terminates as is supposed to be useful i.e., making finite for all . We can also improve when the variable is updated. Corollary 1 ensures that a useful function exists.
Returning to the discrete-time peak computation problem (DPCP), to compute a stopping integer for (defined in (1)), we need to construct a couple useful for .
Problem 1.
Construct a couple useful for when .
The next sections of the paper focus on classical tools from dynamical systems stability theory, such as certificate and Lyapunov functions. To properly define these tools, we need to introduce the concept of comparison functions.
3.2 Elements of comparison functions
The global asymptotic stability of discrete-time systems can be studied by means of comparison functions (see e.g., [GGLW14]). For a comprehensive literature review on comparison functions, interested readers can refer to [Kel14]. In our context, as suggested in Subsection 3.1, the problem (DPCP) can be studied by analyzing the limit superior of the sequence defined in (1). Comparison functions will be used to study its limit superior.
Definition 3 (/ functions classes).
A function for some belongs to if is strictly increasing and continuous on , and . If moreover, and , is said to belong to .
Definition 4 ( functions class).
A function belongs to if it is continuous, decreasing, and .
Definition 5 ( functions class).
A function belongs to if for all , belongs to and if for all , belongs to .
The class is included in but functions are strictly positive except at 0, where they are null. Hence, is useful for if at least one term is strictly positive. Thus, in this subsection, Assumption 2 must be satisfied
Assumption 2.
There exist and satisfying .
Let us consider the set of positive definite functions on a subset containing 0:
that is the set of nonnegative functions on null at 0 and nowhere else.
We also need useful results linking positive definite functions and comparison functions. A proof can be found in [Kha02] (extensions can be found in [Adj24]).
Proposition 9 (Lemma 4.3/Appendix C.4 [Kha02]).
Let containing 0 in its interior. Let be a continuous function in . Let such that . Then, for all , we have
(17) |
where:
(18) |
and .
If, moreover, and tends to as tends to , then and the inequality (17) holds for all .
Even if represents the Euclidean norm, the result holds for any norm in as they are equivalent.
4 Solving problem 1 from bounds
In this section, we return to the resolution of defined in (DPCP). We recall that the data are the following:
-
•
an initial conditions set (nonempty) ;
-
•
a nonzero map ;
-
•
a function such that .
We recall that we want to compute where . To use the results of Section 3 with the sequence , we suppose that Assumption 1 holds.
To simplify the notation of the supremum of an objective function over a nonempty set , we introduce the following new notation:
and the set of functions that are bounded from above on :
With these notations, and Assumption 1 is the same as for all , .
In Subsection 4.1, we show how to solve Problem 1 from a upper bound of the images of the reachable values of the dynamical system . Then, with some additional assumptions, we show that we can also solve Problem 1 from the classical upper bound. Classical means that the upper bound is a upper bound of the norm of the reachable values. The existence of such an upper bound is equivalent to the global asymptotic stability of the system. In Subsection 4.2, we apply the theoretical results obtained in Subsection 4.1 to the running example introduced in Subsection 2.3.
4.1 Theoretical constructions from functions
We propose to combine upper bounds with the celebrated result obtained by Sontag (recalled in Lemma 1) to construct a function such that .
Lemma 1 (Proposition 7 of [Son98]).
For all , there exists such that for all , .
In Subsection 3.1, we saw the importance of studying the limit superior of a real sequence to compute its maximal term. In general, this study is difficult. Here, we can exploit the particular structure of the sequence and the fact that its terms originate from the reachable values of a dynamical system. The limit superior of can be determined from the upper bounds. Moreover, this functional approach provides natural functions .
Theorem 4 (Relaxed construction).
Proof.
1. For all , the function is strictly increasing and then for all , for all , . Taking the supremum over , we have . We conclude by taking the as tends to and recalling that tends to 0 (as tends to as for all ). Finally, under Assumption 2, there exists , such that .
2. This is a direct consequence of Sontag’s result (Lemma 1 here). Indeed, there exist such that for all and for all , . Then . Then, as and are increasing and continuous, can also be written . The function is thus the composition of , and that belong to (the third as ) and thus belongs to . Finally, under Assumption 2, there exists such that and so is useful in this case.
∎
Classically i.e., when , the inequality (for ) is true for all and all if and only if the system is said to be uniformly global asymptotic stability (UGAS) (see for example [NT04, Prop. 1] or [ATP10, Th. 2]. This implies that tends to 0 as tends to . Thus, with a general function , we obtain one-sided information on the asymptotic behavior of the dynamics. However, from Proposition 9 when is continuous and is bounded, the classical certificate of stability is sufficient to construct a function such that .
Theorem 5 (Classical construction).
Proof.
It suffices to obtain an inequality of the form of (19) from the inequality (20) and then to use Theorem 4.
If is continuous, then from Proposition 9, as is continuous, belongs to (as ) and satisfies tends to as tends to . Hence, we have with . Then, from inequality (20), we obtain for all . Thus the inequality (19) holds with and that verifies (as bounded and non reduced to ). We conclude from Theorem 4.
∎
Remark 1.
The approach using the functional upper bound detailed in (19) relies on Sontag’s result, which is not completely constructive. Therefore, it would be complicated to construct a function from and appearing in (20) in practice for general systems. Hopefully, we can develop the latter approach to solve, for well-chosen initial condition sets , the problems presented in Subsection 2.3.
4.2 Application to the running example of Subsection 2.3
First, we recall that we want to compute, for a subset satisfying the condition (7), and , which are respectively the supremum of the sequences and defined in (9) and in (10). We have to construct for all , a function such that . Actually, we will see that we can choose the same function for and and we can construct this function from a upper bound. From this , we can define for all . Next, we explicitly define two finite initial condition sets, . For each, we define two vectors with numerical values. Then, we apply Algorithm 1 to compute the supremum of and for the two finite initial condition sets.
4.2.1 Theoretical developments of the running example using a upper bound function.
We observe, since:
that if and only if . Let and . Then, for and for all and :
Then we have for all and all , where:
The function is clearly in and is clearly nonnegative. As is not reduced to , is strictly positive.
Assumption 2 holds for both and . Indeed, if is a nonzero vector such that converges to 0, then for all , is strictly positive for some integer . A proof is provided in the appendix. The existence of bound implies that for all , converges to 0.
As mentioned previously, in general, the functions of Sontag’s result (Lemma 1 here) are not known explicitly. As we have obtained an inequality of the form (4) where appears explicitly, we have identified and we can directly deduce a function such that . The same function can be chosen for and and we simply write defined as follows:
(21) |
As , we have for , . It is straightforward to see that
Now since, for all , we dispose of an element of , we can solve using and Algorithm 1. Very simple calculi lead to the formula for all and all such that :
4.2.2 Numerical applications.
As discussed earlier, we consider two different finite initial condition sets:
-
1.
where and ;
-
2.
where and .
To simplify the notations, we write for all , and instead of and . Recall that, for all , and :
As and are now explicit, we can refine the definition of the function . We write for the function associated with and for the one associated with . Thus, as
we get:
and we have, for all and all such that :
and for all and all such that :
Then, we are able to apply Algorithm 1 to compute and the maximal integer solution for and .
The case where and
To determine the number of images required for the search of the maximum, we need to compute . However, this is only possible if and only if . So we have to wait for the first integer such that is strictly positive. The coordinates of and are negative, so we move on and compute:
The vectors still have negative coordinates, and we compute
The vectors and have strictly positive coordinates, and we can compute our stopping integers:
This means that the stopping integer in Algorithm 1 is set to 7 for and to 8 for and
Following Algorithm 1, we compute the until reaches the stopping integer. As during the process, we have not found greater values than , we conclude that and for all .
The case where and
Contrary to the case , as and , we can compute stopping integers at :
This leads to the stopping integer in Algorithm 1 being set to 2 for and 10 for . Hence
Now, we compute the images of and by :
As , we can update the stopping integer for and:
Following the fifth statement of Proposition 6, this is slightly smaller than but its integer part is still equal to 2 for . The second coordinate of and of are negative, then we cannot update the stopping integer for . Next, we have
As is strictly less than and the stopping integer for is equal to 2, it follows that and . Since and are negative then nothing changes for . We continue to iterate for until . We have
thus , which was the maximal previous value. We then update the value of the stopping integer of Algorithm 1 for and obtain
The new stopping integer is equal to 8 for . The next values are smaller than , which implies that and .
5 Solving Problem 1 from Lyapunov functions
Classically, Lyapunov functions are used as certificates of stability (see [Ela05] or [KP01]). They bring a constructive approach to prove the stability of the dynamical system through converse Lyapunov theorems, which proved that stability notions are equivalent to the existence of Lyapunov functions [Kha02]. For some linear, polynomial, or piecewise polynomial systems, Lyapunov functions can be computed using numerical optimization solvers based on linear or semidefinite programming (e.g., see [GH15] and references therein).
Classical quadratic Lyapunov functions solve Problem 1 when is affine and stable, and is quadratic (see [Adj25a]). In this section, we prove that this approach can be generalized to all discrete-time systems for which a Lyapunov function exists.
5.1 Definition and useful facts
We begin by recalling the definition of classical Lyapunov functions for discrete-time systems. We briefly present the equivalent definitions of the decrease condition required to be a Lyapunov function. For topological notions, we recall that is the Euclidean norm. However, the context of finite-dimensional space makes the results independent of the choice of the norm.
For , we denote by the set of closed subsets of containing 0 but non reduced to 0 and stable by i.e.,
Following Lazar [Laz06], we define local Lyapunov functions on stable sets. We slightly relax the notion of stable sets, called classically positive or forward invariant sets. Indeed, classically, positive invariant sets are our stable sets with an auxiliary condition: 0 belongs to the interior of the stable set. This condition is not considered stability problems in this paper.
Definition 6 (Classical Lyapunov functions).
A function is said to be a (local) Lyapunov function for on if and only if:
-
1.
There exist such that, for all :
-
2.
There exists such that for all :
Example 1.
In [LHK14a], the authors propose the following function:
(22) |
as a (local) Lyapunov function for the dynamics of our running example presented in Subsection 2.3.
Next, we prove that is a Lyapunov function for on any for which . In this example, we prove the first statement of Definition 6. Let be in . Thus, from the discussion of Subsection 2.3, we have for all nonzero in . Then, for all in , . We conclude that there exist ( and ) such that .
We will prove the second statement by computing the local Lyapunov ratio operator of .
First, due to the equivalence of norms in and since any element of is increasing if the first statement holds for , it holds for all norms on . Second, combining the results and discussions found in [GGLW14, Remark 5], [LHK14b, Def. 5, Th.4] and [Kel14, Lemma 25, Corollary 1], the condition that there exists such that can be equivalently replaced by one of those statements:
-
1.
there exists positive definite and continuous such that ;
-
2.
there exists positive definite, continuous and such that is positive definite satisfying .
First, we remark that from the first statement of Definition 6, is positive definite on any closed ball for which the radius belongs to the domain of . If we require to be continuous and contains 0 in its interior, from Proposition 9, the first statement of Definition 6 is actually equivalent to . Second, if there exists a classical Lyapunov function for , must be null at zero222Actually, if for a nonzero vector, we could replace by and the lower and the upper bounds on by respectively and . Finally, from the second statement of Definition 6, we can define a type of norm operator relative to a Lyapunov function. We propose a general definition of local positive definite ratio operators and then specify the results for local Lyapunov functions.
Definition 7 (Local positive definite ratio operators).
Let such that and . The local positive definite ratio operator of on is the following map:
(23) |
Note that the local positive definite ratio operator is always nonnegative and can take infinite values.
In Proposition 10, we present some useful properties of the local positive definite ratio operator .
Proposition 10.
Let such that and . Then, for all :
-
1.
where ;
-
2.
for all , ;
-
3.
for all , ;
-
4.
there exists such that for all , if and only if .
Proof.
1. This is a straightforward application of the definitions.
2. The result clearly holds if is equal to . Then, if is finite, directly applying the definition of and the positivity of lead to the result.
3. The result holds when is equal to . Now, suppose that is finite. If , then for all . It follows that for all . Finally, for all and . Then and the inequality holds. If for some the inequality obviously holds. We then suppose that and for all are strictly positive. We now apply induction. The result obviously holds for . Assume that it holds for some . Let such that . It follows that . Then . As , and belong to and are not equal to 0. Finally, using the induction hypothesis. We obtain the inequality by taking the supremum over .
4. It follows readily from the definition. ∎
The fourth statement of Proposition 10 is not difficult to prove. However, it is worth noting that if there exists such that for all , , then . For our applications, this latter inequality is important. Indeed, will be used as an element for a function to be determined. Since also belongs to , following Proposition 7, we will have . This will be highlighted in Paragraph 5.3.2.
Corollary 2.
Let be a self-map on . Suppose that admits a local Lyapunov function on . Then:
-
1.
;
-
2.
for all , ;
-
3.
for all , .
Proof.
The first statement follows from the fourth statement of Proposition 10. The second statement is a reformulation of the second statement of Proposition 10, for which the inequality holds for as is finite. The finiteness of also extends the third statement of Proposition 10.
∎
Example 2 (The local positive definite ratio operator of at on closed balls).
Now, let us consider . To go further into our computations, we must compute the local positive definite ratio operator on of at the Lyapunov candidate defined in (22). For sake of simplicity, we write instead of .
First, we suppose that . It follows that . Moreover, as , we also have and so, by assumption, and . Finally, in the case where , we have
The function is strictly decreasing on and strictly increasing on and for all , , hence for all :
When , we get the equality . Futhermore, since is increasing on , .
Now, let consider and such that . Then we have and if and if . Then:
Let us introduce the set
(24) |
We thus are interested in the supremum of for all . As is strictly increasing on , there exists a unique real root denoted by for the polynomial function of degree 3, . As , we have . We do not require the exact value of . In fact, is nonempty if and only if . Indeed, if there exists , then, as strictly increases, and . Conversely, it is obvious that is nonempty if . Note that if is empty, the supremum over it is making for all . Suppose that is nonempty. As , is strictly increasing on and for all :
Finally, we obtain for all :
(25) |
We conclude that satisfies for all :
-
1.
for all , ;
-
2.
there exists such that for all ,
and is a Lyapunov function for on all where .
5.2 Theoretical constructions from Lyapunov functions
In this subsection, we explain how to construct a solution for Problem 1 from a Lyapunov function. We recall that a solution of Problem 1 is an ( defined in (1)) useful when . First, we establish a direct result derived from Theorem 4 without any continuity condition on .
We formulate a simple lemma asserting that the supremum of a local Lyapunov function is finite over nonempty bounded sets contained in a suitable stable set. The result would be evident if the local Lyapunov function is (upper semi)continuous, but continuity is not mandatory in Definition 6.
Lemma 2.
Let . Assume that admits a local Lyapunov function on . Then, for all nonempty bounded sets , .
Proof.
Let be a self-map on and be a local Lyapunov function on . Let be a nonempty bounded set contained in . By definition, there exists , such that for all . Then, as is strictly increasing, for some as is closed. We conclude that is defined at and thus .
∎
Remark 2.
Lemma 2 justifies the closedness of elements on . As a matter of fact, the definition of being a Lyapunov function may allow functions as functions . If such a function is defined on the set where is a closed set, then for all , . This implies that is compact, and then for some . On the contrary, let . If the function is defined on the set . As , so is not defined on .
Theorem 6 (Direct Lyapunov construction).
Proof.
Let and . As and , then, and, from (26), . Moreover, as is a Lyapunov function, from Lemma 2, is finite and we have . The inequality involving follows by taking the supremum over . As is a Lyapunov function on , . This implies that the first statement as tends to 0. As , then and from Lemma 2, as is bounded, . We obtain the second statement from this.
∎
Remark 3.
Actually, the function is a function and as is bounded and not reduced to , belongs to and verifies . Theorem 4 thus applies. As we do not depend on Sontag’s lemma, we have an explicit description of the element .
If we assume that is continuous, the inequality (26) is no longer required. To recover a link between and a Lyapunov function, it suffices to use Proposition 9.
Theorem 7 (Continuous Lyapunov construction).
Assume that is bounded and not reduced to . Suppose that is continuous and there exists a Lyapunov function for on such that . Finally, suppose that there exists satisfying for all and for some . Let verifying for all . We define:
Then . If Assumption 2 holds, then is useful for .
Proof.
Let and . As and , and thus . As is Lyapunov, we then have . As , it follows that . By assumption, for some . We also have . Then, exists. Since on with increasing, the inequality holds. Taking the supremum over leads to . As , and thus is useful if Assumption 2 holds.
∎
Remark 4.
If we impose the strongest condition: for some (instead of ), we can directly use Theorem 6. Indeed, if this stronger condition holds, we can deduce for all .
5.3 Numerical applications to our running example of Subsection 2.3
We propose to illustrate the theoretical results obtained in Subsection 5.2 on the running example presented in Subsection 2.3. First, we explicit the function . Then, we compare the results obtained from the Lyapunov approach with those obtained with upper bounds. Finally, we complete the numerical experiments with new data not treated with the approach.
5.3.1 Theoretical developments of the running example using
We can return to the discrete-time peak computation problems presented in Subsection 2.3. Recall that the objective functions are the coordinate functions and . It is evident that for all where . Recall that defined in (22) is proved in Example 2 to be a Lyapunov function for on all where . It follows that for all where .
Now let . Suppose that is such that and there exists with . As the computation depends on , we write . Then, by definition of , . By assumption on , . Now, following Theorem 7,
(27) |
satisfies for all . This means that for all
Finally, it is easy to see that
Since , we only apply to positive terms and .
5.3.2 Comparison with upper bounds results
In Subsection 4.2.2, we considered the case where with . We recall that is equivalent to , which is the same as . Then, let and . We have for all , and thus . We conclude that defined in (27) and defined in (21) are equal. Moreover, in Subsection 4.2.2, we chose whereas we choose here . We proved that in Example 22 in the case where , that . From the first statement of Proposition 7, we have for all such that , .
For example, we go back the set where and . We have with . Then . However, in Example 2, the value of and
This directly provides the stopping integer equal to 2 and saves computations for . We also have
Again, we save computations with respect to .
The case where and is less significative as 5.7341 is close to . We have with . Then . However, in Example 2, the value of and:
We also have
5.3.3 Numerical applications
In Subsection 4.2.2, we are limited to where . Considering such a radius , we obtain an upper bound of the form . Now, we can consider a greater radius . We cannot exceed since the norm explodes starting with vectors of norm greater than . As in Subsection 4.2.2, we consider two initial condition sets: the first is a singleton for which the norm of the vector is greater than . This will explain the difference from the cases developed in Subsection 4.2.2. In the second case, we start with two vectors with strictly negative coordinates, the norms of which are greater than . Let us define
-
1.
where ;
-
2.
where and .
Again to simplify the notations, we write for all , and instead of and . Recall that, for all , and :
We can refine the definition of and . We write for the function associated with and for the one associated with . First, the smallest satisfying is equal to . The smallest satisfying is equal to . In both cases, the radius exceeds , then and . To know the exact value of and , we have to check whether and . Since and , we conclude that and . Thus, we get:
and we have, for all and all such that :
and for all and all such that :
Then, we are able to apply Algorithm 1 to compute and the maximal integer solution for and .
The case where
As we can compute directly a first stopping integer for and:
This means that we must compare the first 20 terms of . For , we cannot compute a stopping integer because .
Then, we compute (actually already computed since we need this value to compute ). Since
we can update the stopping integer for and we have:
This means that we finally compare the first 17 terms of . The second coordinate is still negative; thus, we cannot compute a stopping integer. Now, as
we have
Then comparing the 16 first terms of leads to the conclusion that and .
For , we have to wait until to obtain . Then, we have:
and
We must compare the 88 first terms of to be guaranteed to pass the maximizer rank. However, the terms after are strictly smaller than (even the positive ones) and we conclude that and .
The case where and
As both vectors have negative coordinates, we must wait for the first integers , such that and . We have, , , and . As and , we can compute:
and
meaning that and .
Then, we compute the next images of and by . We find that and thus we update the stopping integer for and we find:
The new stopping integer for is set to 233. The next values of are smaller than and thus we conclude that and .
The same situation holds for . Indeed, and we have
The new stopping integer for is set to 316. The next values of are smaller than and thus we conclude that and .
6 Concluding remarks
In this paper, we solved specific discrete-time peak computation problems. In general, discrete-time peak computation problems consist in searching, for a given objective function, for the reachable value of a given discrete-time system that maximizes the objective function. The optimal value of this maximization problem can be viewed as the supremum of a sequence of optimal values. The solution proposed in this paper is thus based on previous results, allowing us to compute the maximum of a real sequence. The computational method requires a particular sequence greater than the analyzed sequence. This particular sequence is the image of a positive convergent geometric sequence by a strictly increasing continuous function on . In this paper, we use the structure of the problem (i.e., the dynamical system) to compute the strictly increasing continuous and the positive convergent geometric sequence. The construction of these elements requires the stability of the discrete-time system, as they are constructed from certificates and Lyapunov functions. We illustrate the techniques on an example from the literature.
Three main questions remain unsolved and are left for future work. First, how can we solve a discrete-time peak computation problem for which the underlying system diverges? Second, for nonlinear systems, how can discrete-time peak computation problems be managed with an infinite bounded initial conditions set? Finally, how can the number of unuseful computations be reduced?
References
- [Adj17] A. Adjé. Proving properties on PWA systems using copositive and semidefinite programming. In Numerical Software Verification: 9th International Workshop, NSV 2016, Toronto, ON, Canada, July 17-18, 2016, Revised Selected Papers 9, pages 15–30. Springer, 2017.
- [Adj21] A. Adjé. Quadratic maximization of reachable values of affine systems with diagonalizable matrix. J. Optim. Theory Appl., 189(1):136–163, 2021.
- [Adj24] A. Adjé. A parametric optimization point-of-view of comparison functions. arXiv preprint arXiv:2408.14440, 2024.
- [Adj25a] A. Adjé. Quadratic maximization of reachable values of stable discrete-time affine systems. To appear in Optimization, pages 1–52, 2025.
- [Adj25b] Assalé Adjé. On the maximization of real sequences, 2025.
- [AG24] A. A. Ahmadi and O. Günlük. Robust-to-dynamics optimization. Mathematics of Operations Research, 2024.
- [AGM15] A. Adjé, P.-L. Garoche, and V. Magron. Property-based polynomial invariant generation using sums-of-squares optimization. In Static Analysis: 22nd International Symposium, SAS 2015, Saint-Malo, France, September 9-11, 2015, Proceedings 22, pages 235–251. Springer, 2015.
- [AL11] M. F. Anjos and J.B Lasserre. Handbook on semidefinite, conic and polynomial optimization, volume 166. Springer Science & Business Media, 2011.
- [APS18] U. M. Ahiyevich, S. E. Parsegov, and P. S. Shcherbakov. Upper bounds on peaks in discrete-time linear systems. Automation and Remote Control, 79:1976–1988, 2018.
- [ATP10] A. Loria A.R. Teel, D. Nešić and E. Panteley. Summability characterizations of uniform exponential and asymptotic stability of sets for difference inclusions. Journal of Difference Equations and Applications, 16(2-3):173–194, 2010.
- [Ber15] D. Bertsekas. Convex optimization algorithms. Athena Scientific, 2015.
- [BTN01] A. Ben-Tal and A. Nemirovski. Lectures on modern convex optimization: analysis, algorithms, and engineering applications. SIAM, 2001.
- [BYG17] C. Belta, B. Yordanov, and E. A. Gol. Formal methods for fiscrete-time dynamical systems, volume 89. Springer, 2017.
- [Ela05] S. Elaydi. An introduction to difference equations. 2005.
- [Els12] E.M. Elsayed. Solutions of rational difference systems of order two. Mathematical and Computer Modelling, 55(3):378–384, 2012.
- [GGLW14] R. Geiselhart, R. H. Gielen, M. Lazar, and F. R. Wirth. An alternative converse Lyapunov theorem for discrete-time systems. Systems & Control Letters, 70:49–59, 2014.
- [GH15] P. Giesl and S. Hafstein. Review on computational methods for Lyapunov functions. Discrete and Continuous Dynamical Systems-B, 20(8):2291–2331, 2015.
- [GRR10] M Ranferi Gutiérrez, M A Reyes, and H C Rosu. A note on verhulst’s logistic equation and related logistic maps. Journal of Physics A: Mathematical and Theoretical, 43(20):205204, apr 2010.
- [HR18] Yacine Halim and Julius Fergy Rabago. On the solutions of a second-order difference equation in terms of generalized padovan sequences. Mathematica Slovaca, 68, 06 2018.
- [Kel14] C. M. Kellett. A Compendium of comparison function results. Math. Control. Signals Syst., 26(3):339–374, 2014.
- [Kel15] Christopher M. Kellett. Classical converse theorems in lyapunov’s second method. Discrete and Continuous Dynamical Systems - B, 20(8):2333–2360, 2015.
- [Kha02] H.K. Khalil. Nonlinear systems. Pearson Education. Prentice Hall, 3rd edition, 2002.
- [KP01] W. G. Kelley and A. C. Peterson. Difference equations: an introduction with applications. Academic press, 2001.
- [Las15] J.-B. Lasserre. An introduction to polynomial and semi-algebraic optimization. Cambridge Texts in Applied Mathematics. Cambridge University Press, 2015.
- [Laz06] M. Lazar. Model predictive control of hybrid systems: stability and robustness. Phd thesis, T.U. Eindhoven, Sept 2006.
- [LHK14a] H. Li, S. Hafstein, and C. M. Kellett. Computation of lyapunov functions for discrete-time systems using the yoshizawa construction. In 53rd IEEE conference on decision and control, pages 5512–5517. IEEE, 2014.
- [LHK14b] H. Li, S. Hafstein, and C. M. Kellett. Computation of Lyapunov functions for discrete-time systems using the Yoshizawa construction. In 53rd IEEE Conference on Decision and Control, pages 5512–5517, 2014.
- [MHL15] Victor Magron, Didier Henrion, and Jean-Bernard Lasserre. Semidefinite approximations of projections and polynomial images of semialgebraic sets. SIAM Journal on Optimization, 25(4):2143–2164, 2015.
- [MHS20] J. Miller, D. Henrion, and M. Sznaier. Peak estimation recovery and safety analysis. IEEE Control Systems Letters, 5(6):1982–1987, 2020.
- [NT04] D. Nešić and A. R. Teel. Matrosov theorem for parameterized families of discrete-time systems. Automatica, 40(6):1025–1034, 2004.
- [RKML06] S. V. Rakovic, E. C. Kerrigan, D. Q. Mayne, and J. Lygeros. Reachability analysis of discrete-time systems with disturbances. IEEE Transactions on Automatic Control, 51(4):546–561, 2006.
- [Son98] E. D. Sontag. Comments on integral variants of ISS. Systems & Control Letters, 34(1):93–100, 1998.
- [Ste14] Stevo Stevic. Representation of solutions of bilinear difference equations in terms of generalized fibonacci sequences. Electron. J. Qual. Theory Differ. Equ, 67(1):15, 2014.
Appendix
Lemma 3.
Let us consider the map in (4). Let be a nonzero vector such that tends to 0 as tends to . There exists such that and .
Proof.
Actually, it suffices to prove that for all in the open unit ball, there exists , such that and . Indeed, if converges to 0, then eventually enters the open unit ball. The worst situation is when has only negative coordinates before entering the open unit ball.
Therefore, let us consider not being 0 and such that . First, we consider . If the result obviously holds. Therefore, we suppose that . If (with if ), as , we have . Now, suppose that . Since , we have . Recall that for all where , . If , the result follows. Thus, if , we get since and and .
Now, we prove the result for . The result trivially holds if . So, we suppose that . If ( if ), then, as , we get . Now, assume that . Then, it leads to . If , the results holds. Assuming that , then from the fact that , and , we get .
∎