Preprint
Article

This version is not peer-reviewed.

Existence and Properties of Solution of Nonlinear Differential Equations with Impulses at Variable Times

A peer-reviewed article of this preprint also exists.

Submitted:

02 January 2024

Posted:

03 January 2024

You are already at the latest version

Abstract
In this paper, a class of nonlinear ordinary differential equations with impulses at variable times is considered. The existence and uniqueness of solution are given. At the same time, modifying the classical definitions of continuous dependence and Ga^teaux differentiability, some results on continuous dependence and Ga^teaux differentiable of solution relative to the initial value also are presented in new topology sense. For the autonomous impulsive system, the periodicity of solution is given. As an application, properties of solution for a type of controlled nonlinear ordinary differential equation with impulses at variable times is obtained. These results are foundation to study optimal control problems of systems governed by the differential equations with impulses at variable times.
Keywords: 
;  ;  ;  ;  

1. Introduction

We begin with introducing the problem studied. Let R + [ 0 , + ) , Y ( t ) = { y i ( t ) | i Λ { 1 , 2 , , p } } , f : R + × R n R n , y i : R + R n and J i : R n R n ( i = 1 , 2, ⋯, p) are given maps. Consider the following differential equations with impulses at variable times
x ˙ ( t ) = f ( t , x ( t ) ) , { x ( t ) } Y ( t ) = , t 0 , x ( t + ) = J i ( x ( t ) ) + x ( t ) , { x ( t ) } Y ( t ) = y i ( t ) , t 0 , x ( 0 ) = x 0 .
The main purpose of this study are (i) to provide a sufficient condition to the existence and uniqueness of solution x for the impulsive system (1.1); and (ii) to give the necessary and sufficient condition to the exact times for the solution x meets the curve Y; (iii) to present properties of solution relative to the initial value.
There are some interesting phenomena for the impulsive system (1.1). All of first, it is clear that the system x ˙ ( t ) = x ( t ) + u ( t ) is controllable, but the following impulsive system
x ˙ ( t ) = x ( t ) + u ( t ) , x ( t ) 1 , x ( t + ) = 0 , x ( t ) = 1
is not controllable. Similarly, the system x ˙ ( t ) = x ( t ) is stable, but the impulsive system
x ˙ ( t ) = x ( t ) , x ( t ) 1 , x ( t + ) = 2 , x ( t ) = 1
is not stable when the initial value x ( 0 ) 1 . Let us look at the third example. Denote by x ( · ; 0 , x 0 ) the solution of the following impulsive differential system
x ˙ ( t ) = 2 t , x ( t ) 1 , t > 0 , x ( t + ) = 0 , x ( t ) = 1 , t 0
with the initial value ( 0 , x 0 ) . Then, we have
x t ; 0 , 1 + 1 n = t 2 + 1 + 1 n , t 0 , x ( t ; 0 , 1 ) = 1 , t = 0 , t 2 m , t m , m + 1 , m N .
This implies that the impulsive system (1.1) never has not continuous of solution to respect to the initial value in L 1 . In addition to, we can also use simple cases to show that the implsive system (1.1) may has not global solution.
The motivation to study this problem is as follows. First of all, many physical phenomena and application models are characterized by (1.1). For example, Integrate-and-Fire models derived from the physical oscillation circuit [1,2] is widely used in neuroscience, neuroscience research concerns current-voltage relations at which the states can be reset once the voltage reaches a threshold level [3,4]. Again, in the application, it is crucial to choose appropriate threshold levels for making decisions to trigger or suspend an impulsive intervention: [5] use glucose threshold level guided injections of insulin; [6] use the time that is economic threshold is reached by the amount of pests as the time of impulsive intervention. Secondly, the theory of impulsive differential equations has been an object of increasing interest because of its wide applicability in biology, in medicine and in more and more fields (see [7] and the its references). Significant interest in the investigation of differential equations with impulse effect is explained by the development of equipment in which significant role is played by complex systems [8,9,10]. Particularly, the qualitative theory of the impulsive system (1.1) has not been systematically established. It is natural to ask for the present the qualitative theory of the impulsive system (1.1). We will discuss the existence and uniqueness of global solution and its properties for the nonlinear ordinary differential equations with impulses at variable times (1.1) under weaker conditions. It is worth pointing out that the solutions of differential systems with impulses may experience the pulse phenomena, namely the solutions may hit a given surface finite or infinite number of times causing rhythmical beating. This situation presents difficulties in the investigation of properties of solutions of such systems. In addition to, it is not suitable to the stronger conditions for control problem. Consequently, it is desirable to find weaker conditions that guarantee the absence or presence of pulse phenomena. More generally, it is significant to find conditions that the solution meets a given surface just has k N times ( N denote the natural numbers set).
Before concluding this section, we review the previous literature to qualitative analysis on the impulsive differential equations. In fact, the qualitative analysis on the impulsive differential equations can at least be traced back to the works by N.M. Kruylov and N.N. Bogolyubov [11] in 1937 in their classical monograph to NonlinearMechanics . Mathematical formulation of the differential equation with impulses at fixed times was firstly presented by A.M. Samoilenko and N.A. Perestyuk [12] in 1974. Since then, the qualitative theory for differential equation with impulses at fixed times in finite (or infinite) dimensional spaces has been extensively studied (see [13,14,15,16] and the references therein). For the differential equations with impulses at variable times, A.M. Samoilenko and N.A. Perestyuk [17] in 1981 given the mathematical model
x ˙ ( t ) = f ( t , x ( t ) ) , t τ i ( x ( t ) ) , x ( t + ) = x ( t ) + J i ( x ( t ) ) , t = τ i ( x ( t ) ) .
Later relevant works were made by D.D. Bainov and A.B. Dishliev [18] in 1984, S. Hu [19] in 1989, etc. For more details, one can see the monographs of V.Lakshmikantham [20] in 1989, A.M. Samoilenko [21] in 1995, D.D. Bainov [22] in 1995 and M. Benchohra [23] in 2006, and so on. In a word, these works established the qualitative theory of (1.2) under stronger conditions. Specially, it is not suitable to the stronger conditions for control problem and impulsive differential equation in infinite dimensional spaces. At the same time, when y i , ( i Λ ) is a one to one mapping, x ( t ) = y i ( t ) be equal to t = y i 1 ( x ( t ) ) . Hence, (1.2) can be treated as a simplified case of (1.1). For the linear case of (1.1), Peng etc [24] obtained the existence and uniqueness of solution and its properties.
The rest of the paper is organized as follows. Section 2 presents the main results. In Section 3, 4, 5, the proofs of the three main theorems are given in turns. Periodicity of autonomous impulsive system is presented in Section 6. As an application, variation of solution relative to control is presented in Section 7, which is foundation to study optimal control problems of systems governed by the differential equations with impulses at variable times. Finally, some new phenomena of impulsive differential system are summarized.

2. Main Results

We are going to present our main results in the section. To state the first one, some preliminaries will be introduced. Throughout this paper, we fix T > 0 and approve that T = + . We firstly introduce several definitions. We define the function set: P C Y [ 0 , T ) , R n = { x : [ 0 , T ) R n | x is continuous at t when x ( t ) Y ( t ) , x is left continuous at t and the right limit x ( t + ) exists when x ( t ) Y ( t ) } . For x P C Y [ 0 , T ) , R n , t [ 0 , T ) is called to be an irregular point if x ( t ) Y ( t ) . Otherwise, t is called to be a regular point. One can directly verify that the function set P C Y [ 0 , T ) , R n is not linear. Denoted by B ( z , θ 2 ) the closed ball (in R n ) centered at z and of radius θ 2 > 0 .
Definition 2.1.
A piecewise continuous function x θ is said to be an approximate P C solution of (1.1) if x θ ( · ) x θ ( · ; 0 , x 0 ) satisfies the following integral equation with impulses
x θ ( t ) = x 0 + 0 t f τ , x θ ( τ ) d τ + 0 t i < t , x θ ( t i ) B ( y j ( t i ) , θ 2 ) J j x θ ( t i ) .
Particularly, when θ = 0 , we call x ( · ) x 0 ( · ) P C Y [ 0 , T ) , R n is a P C solution of (1.1).
Meanwhile, we introduced the following basic assumptions.
[F] (1) f : R + × R n R n is measurable in t on R + and locally Lipschitz continuous in x i.e. for any ρ > 0 , there exits L ( ρ ) > 0 such that for all x , y R n with | x | , | y | ρ , we have
| f ( t , x ) f ( t , y ) | L ( ρ ) | x y | f o r a n y t R + .
(2) There exists a constant k ˜ > 0 such that
| f ( t , x ) | k ˜ ( 1 + | x | ) f o r a n y t R + .
(3) f is continuous partially differentiable in x and f x ( · , x ) L l o c 1 R + , R n × n .
[Y](1) y i C R + , R n and y i ( t ) y j ( t ) for all t R + and i j ( i , j Λ ).
(2) y i C 1 ( [ 0 , T ] , R n ) and f t , y i t y ˙ i t ( i Λ ).
[J](1) J i : R n R n is continuous and
Υ i ( t ) y i ( t ) + J i ( y i ( t ) ) y j ( t ) f o r a l l t R + a n d i , j Λ .
(2) J i : R n R n is continuous partially differentiable.
It is clear that when the assumptions [F](1)(2) hold, for any fix ( s , z s ) R + × R n , the differential equation
z ˙ ( t ) = f ( t , z ( t ) ) , t > s , z ( s ) = z s ,
has a unique solution z ( · ; s , z s ) C [ s , + ) , R n given by
z ( t ; s , z s ) = z s + s t f ( τ , z ( τ ; s , z s ) ) d τ .
We define several functions:
F i ( t ; s , z s ) = z ( t ; s , z s ) y i ( t ) , z s y i ( s ) ( i = 1 , 2 , , p ) , t s
and
F i j ( t ; s , Υ i ( s ) ) = z ( t ; s , Υ i ( s ) ) y j ( t ) , Υ i ( s ) y j ( s ) ( i , j = 1 , 2 , , p ) , t s ,
where · , · denotes the inner product in R n .
The first main result is presented as follows.
Theorem 2.2.
Suppose the assumptions [F](1)(2), [Y](1) and [J](1) hold.
(1) The system (1.1) admits a unique P C solution x P C Y R + , R n .
(2) x has exactly irregular point set t i | 0 t 1 < t 2 < < t k < + over R + if and only if there exists l i Λ ( i = 1 , 2 , , k ) such that
F l 1 ( t 1 ; 0 , x 0 ) = 0 , F l i l i + 1 ( t i + 1 ; t i , Υ l i ( t i ) ) = 0 f o r i = 1 , 2 , , k 1 ,
and
F l k j t ; t k , Υ l k ( t k ) > 0 f o r a n y t [ t k , + ) f o r a l l j Λ .
We have to point out that the necessary and sufficient conditions of pulse phenomenon also is given in Theorem 2.2. Not only that, for the existence of solution of the system (1.2), in order to ensure t k = τ k ( x ) monotonous with respect to k in [20], it requires that τ k ( x ) to be smooth and satisfy the corresponding inequality conditions. However, using Theorem 2.2, we can obtain immediately the following result.
Corollary 2.3.
Suppose the assumptions [F](1)(2), [Y](1) and [J](1) hold. If y i is invertible and τ i = y i 1 for any i Λ , then the system (1.2) admits a unique P C solution x P C Y R + , R n .
Now, we state second and third main results. It follows from Theorem 2.2 that for any fixed sufficient small θ > 0 , (1.1) has a unique approximate solution x θ provided that the assumptions [F](1)(2), [Y](1) and [J](1) hold. Let v R n , x θ ( · ; θ , x 0 + θ v ) be an approximate P C solution of the equation (1.1) corresponding to ( θ , x 0 + θ v ) . We note that (1.1) is not well posed. Thus, we can never expect to have the continuity of solution to respect to the initial value. We have to modify the classical definition of continuous and differentiability, respectively.
Definition 2.4.
Let v R n be fixed. The P C solution x ( · ; 0 , x 0 ) of (1.1) is said to have continuous dependence relative to the initial value ( 0 , x 0 ) if the following facts hold:
(i) when x ( t ; 0 , x 0 ) y i ( t ) ( i Λ ), x θ ( t ; θ , x 0 + θ v ) x ( t ; 0 , x 0 ) as θ 0 ;
(ii) for any sufficient small ε > 0 , there exist δ > 0 and I ε [ 0 , T ] such that
x θ t ; θ , x 0 + θ v x t ; 0 , x 0 < ε f o r a n y t I ε ,
when μ [ 0 , T ] I ε < ε , θ < δ , where μ denote the Lebesgue measure.
Definition 2.5.
Let v R n be fixed. The P C solution x · ; 0 , x 0 of (1.1) is said to G a ^ teaux differentiable relative to the initial value ( 0 , x 0 ) if G a ^ teaux derivative φ ( t ) of x t ; 0 , x 0 exists at ( 0 , x 0 ) for all t [ 0 , T ] with x t ; 0 , x 0 y i ( t ) , otherwise,
φ ( t ) = lim s t φ ( s ) ,
where
φ ( t ) = lim ε 0 x ε t ; ε , x 0 + ε v x t ; 0 , x 0 ε w h e n x t ; 0 , x 0 y i ( t ) .
Let we state the following main results.
Theorem 2.6.
Suppose the assumptions [F](1)(2), [Y](1) and [J](1) hold. Then P C solution x ( · ; 0 , x 0 ) of (1.1) have continuous dependence relative to the initial value ( 0 , x 0 ) in the sense of Definition 2.4.
Theorem 2.7.
Suppose the assumptions [F], [Y] and [J] hold. Then P C solution x · ; 0 , x 0 of (1.1) is G a ^ teaux differentiable relative to the initial value ( 0 , x 0 ) in the sense of Definition 2.5. Moreover, its G a ^ teaux derivative φ is a P C solution of the following differential equation with impulses
φ ˙ ( t ) = f x ( t , x ( t ) ) φ ( t ) , t ( 0 , T ] , x ( t ) Y ( t ) = , φ ( t + ) = φ ( t ) + J i ( y i ( t ) ) [ φ ( t ) + h ˙ t ( 0 ) f ( t , y i ( t ) ) ] , x ( t ) Y ( t ) = y i ( t ) , φ ( 0 ) = v f ( 0 , x 0 ) .
Here, h t denote solution of the equation { x ε t ; ε , x 0 + ε v } B y i ( t ) , ε 2 in ε for some i Λ .

3. Proof of Theorem 2.2

Throughout this section, we define the function r : ( 0 , + ) R + given by
r ( T ) = Δ 1 2 inf s , t [ 0 , T ] | y i ( s ) y j ( t ) | , y i ( s ) Υ j ( t ) , y i ( s ) Υ i ( t ) | i , j Λ a n d i j ,
where Υ j is defined by (2.1). It is easy from the assumptions [J](1) and [Y](1) to see Υ i C [ 0 , T ] , R n for all i Λ . Hence, there exists a constant M ( T ) such that
| Υ i ( t ) | M ( T ) f o r a n y t [ 0 , T ] a n d i Λ
and
r ( T ) > 0 f o r a l l T > 0 .
To claim the existence and uniqueness of solution of (1.1), we need the following Lemma.
Lemma 3.1.
If the assumptions [F](1)(2), [Y](1) and [J](1) hold, then for any ( s , ξ ) [ 0 , T ) × { Υ i ( t ) | t [ 0 , T ] , i Λ } , there is a δ > 0 which is independent of ( s , ξ ) such that the following differential equation
ϕ ˙ ( t ) = f ( t , ϕ ( t ) ) , t > s , ϕ ( s ) = ξ ,
has a unique solution ϕ C [ s , s + δ ] , R n and
ϕ ( t ) y i ( t ) r ( T ) 2 f o r   a n y t [ s , s + δ ] a n d i Λ .
Proof. 
It follows from the assumptions [F](1)(2) that (3.4) has a unique solution ϕ C [ s , T ] , R n and
| ϕ ( t ) | | ξ | + s t k ˜ ( 1 + | ϕ ( τ ) | ) d τ .
Using Gronwall’s inequality, we have
| ϕ ( t ) | | ξ | + k ˜ T e k ( t s ) .
Together with (3.2), this means that
| ϕ ( t ) | M ( T ) + k ˜ T e k T M ˜ ( T ; k ˜ ) for   any t [ 0 , T ] .
Consequently, for any t [ 0 , T ] , we have
ϕ ( t ) ξ s t | f ( τ , 0 ) | d τ + L ( M ˜ ( T ; k ˜ ) ) M ˜ ( T ; k ˜ ) | t s | .
Together with (3.3) and
ϕ ( t ) y i ( t ) y i ( t ) ξ | ϕ ( t ) ξ | ,
there exists a constant δ = δ ( T , k ˜ ) > 0 such that (3.5) holds. □
By Lemma 3.1, one can prove that for any T > 0 , the equation (1.1) has a unique P C solution. Thus, the equation (1.1) admits a unique P C solution x ( · ; 0 , x 0 ) on R + .
Next, we discus the number of irregular point for the solution x of (1.1) over R + . It is easy that the solution x of (1.1) has no irregular point over R + if and only if for all i Λ , x ( t ; u , 0 , x 0 ) y i ( t ) for any t R + . Together with (2.4), we have following result.
Lemma 3.1.
If the assumptions [F](1)(2), [Y](1) and [J](1) hold, then the solution x of (1.1) has no irregular point over R + if and only if the following algebraic equations
F i ( t ; 0 , x 0 ) = 0 h a s   n o   s o l u t i o n   o n R + f o r   a l l i Λ .
Now, we prove the necessity on (2) of Theorem 2.2. For convenience, we let x ( · ) = x ( · ; 0 , x 0 ) and t i | 0 t 1 < t 2 < < t k < + stand for the irregular point set of x over R + . Then there firstly exists l 1 Λ such that
x ( t 1 ) = y l 1 ( t 1 ) .
Together with (2.4), we can affirm
F l 1 ( t 1 ; 0 , x 0 ) = 0 .
For the second irregular point t 2 of x, there exists l 2 Λ such that
x ( t 2 ) = x ( t 2 ; t 1 , Υ l 1 ( y l 1 ( t 1 ) ) ) = y l 2 ( t 2 ) .
Together with (2.5), it follows
F l 1 l 2 ( t 2 ; t 1 , Υ l 1 ( y l 1 ( t 1 ) ) = 0 .
Similarly, for the irregular point t k of x, there is a l k Λ such that
F l k 1 l k ( t k ; t k 1 , Υ l k 1 ( t k 1 ) ) = 0 .
Moreover, we can see from Lemma 3.2 that x has not irregular point on ( t k , + ) if and only if
F l k j t ; t k , Υ l k ( t k ) = 0 has   no   solution   on ( t k , + ) for   all j Λ .
Combined with (2.5), it is easy from the assumptions [J] and [Y] to see F l k j · ; t k , Υ l k ( t k ) C ( [ t k , + ) ) and
F l k j t k ; t k , Υ l k ( t k ) = z ( t k ; t k , Υ k k ( t k ) ) y j ( t k ) , Υ l k ( t k ) y j ( t k ) > 0 for   all j Λ .
Therefore, together with (3.7), this means (2.7) and (2.8) hold.
To the sufficient on (2) of Theorem 2.2, suppose t i | 0 t 1 < t 2 < < t k < + satisfies (2.7) and (2.8). Then it is not difficult to claim that t j is the irregular point x of the (1.1) over R + . Simultaneously, one can conclude for (2.8) and Lemma 3.1 that t i | 0 t 1 < t 2 < < t k < + is the irregular point set of x over R + . This completes the proof.

4. Proof of Theorem 2.6

Throughout this section, we fix T > 0 . It follows from Theorem 2.2 that the irregular point to the P C solution x of (1.1) at most a finite number of times on the interval [ 0 , T ] . There are only two possibilities which are Case (1): x has no irregular point on [ 0 , T ] and Case (2): x has at least one irregular point on [ 0 , T ] .
In Case (1), the P C solution x has continuous dependence relative to the initial value in the sense of classical definition, i.e.
| x θ · ; θ , x 0 + θ v x · ; 0 , x 0 | C ( [ 0 , T ] , R n ) 0 as θ 0 .
In Case (2), if x 0 = y i ( 0 ) for some i Λ , we only study the P C solution x · ; 0 + , Υ i ( 0 ) . Consequently, we may assume that x · ; 0 , x 0 meets the warning wall Y ( · ) , k times in [ 0 , T ] and let t ¯ j i be the moments of x · ; 0 , x 0 hits the warning line y i ( · ) ( i Λ , j = 1 , 2, ⋯, k), 0 < t ¯ 1 i < t ¯ 2 j < < t ¯ k r < T . By Theorem 2.2, we can prove that there exists δ ¯ > 0 such that when 0 θ < δ ¯ , the impulsive differential equation (1.1) has a unique approximate P C solution x θ ( · ; θ , x 0 + θ v ) corresponding to the initial value ( θ , x 0 + θ v ) .
Let t j i ( θ ) be the moments of x θ ( · ; θ , x 0 + θ v ) . For sufficient small ε > 0 , it is obvious from the definition of solution (see (2.1)) that there is an δ 1 > 0 with δ 1 < δ ¯ such that for any θ > 0 with θ < δ 1 ,
| x θ t ; θ , x 0 + θ v x t ; 0 , x 0 | < ε f o r a n y t 0 , t ¯ 1 i ε 4 k .
This means
lim θ 0 t 1 i ( θ ) = t ¯ 1 i .
Together with the continuity of J i , we have
lim θ 0 J i x θ t 1 i ( θ ) ; θ , x 0 + θ v = J i x t ¯ 1 i ; 0 , x 0 .
It follows from (2.2) that
lim θ 0 Υ i t 1 i ( θ ) = Υ i t ¯ 1 i ,
where
Υ i t 1 i ( θ ) = x θ t 1 i ( θ ) ; θ , x 0 + θ v + J i x θ t 1 i ( θ ) ; θ , x 0 + θ v .
Consequently, combine with (2.1), there is an δ 2 > 0 with δ 2 < δ 1 such that for any θ > 0 with θ < δ 2 ,
| x θ t ; θ , x 0 + θ v x t ; u , 0 , x 0 | = | x θ t ; t 1 i ( θ ) , Υ i t 1 i ( θ ) x t ; t ¯ 1 i , Υ i t ¯ 1 i | < ε for   any t t ¯ 1 i + ε 4 k , t ¯ 2 j ε 4 k .
Let
Υ i t j i ( θ ) = x θ t j i ( θ ) ; θ , x 0 + θ v + J i x θ t j i ( θ ) ; θ , x 0 + θ v , j > 1 , i Λ .
In general, repeat the above process, one can show that there is an δ j + 1 > 0 with δ j + 1 < δ j such that for any θ > 0 with θ < δ j + 1 ,
| x θ t ; θ , x 0 + θ v x t ; 0 , x 0 | = | x θ t ; t j i ( θ ) , Υ i t j i ( θ ) x t ; t ¯ j i , Υ i t ¯ j i | < ε for   any t t ¯ j i + ε 4 k , t ¯ j + 1 r ε 4 k
and
lim θ 0 t j + 1 r ( θ ) = t ¯ j + 1 r ,
lim θ 0 J r x θ t j + 1 r ( θ ) ; θ , x 0 + θ v = J r x t ¯ j + 1 r ; 0 , x 0 ,
lim θ 0 Υ r t j + 1 r ( θ ) = Υ r t ¯ j + 1 r ,
where
Υ r t j + 1 r ( θ ) = x θ t j + 1 r ( θ ) ; θ , x 0 + θ v + J r x θ t j r ( θ , ) ; θ , x 0 + θ v .
In short, for any sufficient small ε > 0 , there exists an δ > 0 such that
x θ t ; θ , x 0 + θ v x t ; 0 , x 0 < ε for   any t I ε when θ < δ ,
where
I ε = 0 , t ¯ 1 i ε 4 k j = 1 k 1 t ¯ j i + ε 4 k , t ¯ j + 1 r ε 4 k t ¯ k r + ε 4 k , T .
This completes the proof.

5. Proof of Theorem 2.7

Throughout this section, we fix T > 0 . It follows from Theorem 2.6 that there are only two possibilities which are Case ( i ) : x · ; 0 , x 0 has no irregular point on [ 0 , T ] and Case ( i i ) : x · ; 0 , x 0 has at least one irregular point on [ 0 , T ] .
In Case ( i ) , one can directly check that x · ; 0 , x 0 of (1.1) is G a ^ teaux differentiable and its G a ^ teaux derivative φ is a weak solution of the following differential equation
φ ˙ ( t ) = f x ( t , x ( t ; 0 , x 0 ) ) φ ( t ) , t ( 0 , T ] , φ ( 0 ) = v f ( 0 , x 0 ) .
To discuss Case ( i i ) , we define function h t given by
h t ( ε ) denote   solution   of   the   equation H ( ε , t ) = 0 .
Here,
H ( ε , t ) = x ε t ; ε , x 0 + ε v y ˜ ( t , ε ) ,
where y ˜ ( t , ε ) = y ˜ i ( t , ε ) for some i Λ , y ˜ i ( t , ε ) B y i ( t ) , ε 2 . By Theorem 2.6, when x t ; 0 , x 0 = y i ( t ) , there is an δ > 0 such that the definition (5.2) is well for all ε [ 0 , δ ] , that is, h t : [ 0 , δ ] O ( t ) is a function and h t ( 0 ) = t , where O ( t ) denote some neighborhood of t. For convenience, let { t j i | 0 < t 1 i < t 2 j < < t k r < T } denote the irregular point set of x · ; 0 , x 0 on [ 0 , T ] . If y i C 1 ( [ 0 , T ] , R n ) , it follows from Theorem 2.6 and (5.3) that there is an δ > 0 such that
H C [ 0 , δ ] × [ 0 , T ] a n d H ε , h t j i ( ε ) = 0 for   any ε [ 0 , δ ] , i Λ , j = 1 , 2 , , k
and
H t ( ε , t ) = f t , x ε t ; ε , x 0 + ε v y ˜ t ( t , ε ) .
Further, when f t j i , y i t j i y ˙ i t j i ( j = 1 , 2 , , k , i Λ ), we have
H t ε , h t j i ( ε ) = f h t j i ( ε ) , x ε h t j i ( ε ) ; ε , x 0 + ε v y ˙ i ( h t j i ( ε ) ) 0 i n R n , ε [ 0 , δ ] ,
where j = 1 , 2, ⋯, k. Let f = f 1 , f 2 , , f n , y i = y i 1 , y i 2 , , y i n ( i Λ ). Without loss of generality, we suppose
f 1 h t j i ( ε ) , x ε h t j i ( ε ) ; ε , x 0 + ε v y ˙ i 1 ( h t j i ( ε ) ) 0 i n R , ε [ 0 , δ ] , j = 1 , 2 , , k ,
We introduce the following functions
Φ ε ( t , s ) = exp s t f x ( τ , x ε ( τ ; ε , x 0 + ε v ) ) d τ ,
then
Φ ( t , s ) = lim ε 0 Φ ε ( t , s ) = exp s t f x ( τ , x ( τ ; 0 , x 0 ) ) d τ .
We let
Φ ε 1 ( t , s ) and Φ 1 ( t , s ) denote   the   first   line   vector   of Φ ε ( t , s ) and Φ ( t , s ) respectively .
We firstly claim the following lemma.
Lemma 5.1.
Suppose the assumption [F](3) holds. Then h t is differentiable over [ 0 , δ ] for some δ > 0 and its derivative given by
h ˙ t j i ( 0 ) = Φ 1 t 1 i , 0 f ( 0 , x 0 ) v f 1 t 1 i , y i t 1 i y ˙ i 1 t 1 i , j = 1 , h ˙ t j 1 r ( 0 ) Φ 1 t j i , t j 1 r f t j 1 r , y r t j 1 r I + J r y r t j 1 r y ˙ r t j 1 r f 1 t j i , y i t j i y ˙ i 1 t j i , j > 1 .
Here, I is unit matrix.
Proof. 
When t 0 , h t 1 i ( ε ) , it follows from the assumption [F](3) and (2.9), (1.2) that
H ε ( ε , t ) = lim ξ 0 x ε + ξ ( t ; ε + ξ , x 0 + ( ε + ξ ) v ) x ε ( t ; ε , x 0 + ε v ) ξ + ε y ˜ i ( t , ε ) = lim ξ 0 ε + ξ t 0 1 f x ( s , x ε ( s ; ε , x 0 + ε v ) + θ ( x ε + ξ ( s ; ε + ξ , x 0 + ( ε + ξ ) v ) x ε ( s ; ε , x 0 + ε v ) ) ) x ε + ξ ( s ; ε + ξ , x 0 + ( ε + ξ ) v ) x ε ( s ; ε , x 0 + ε v ) ξ d θ d s v f ( ε , x 0 + ε v ) + ε y ˜ i ( t , ε ) .
One can see from (5.7) and the above equality that
H ε ( ε , t ) = Φ ε ( t , ε ) ( v f ( ε , x 0 + ε v ) ) + ε y ˜ i ( t , ε ) .
Combine with (5.6), (5.7) and (5.9), we have
h ˙ t 1 i ( ε ) = Φ ε 1 h t j i ( ε ) , ε v f ( ε , x 0 + ε v ) + ε y ˜ i 1 ( t , ε ) f 1 h t 1 i ( ε ) , x ε h t 1 i ( ε ) ; ε , x 0 + ε v y ˙ i 1 h t 1 i ( ε )
and
h ˙ t 1 i ( 0 ) = Φ 1 t 1 i , 0 v f ( 0 , x 0 ) y ˙ i 1 t 1 i f 1 t 1 i , y i t 1 i ,
In general, when t h t j 1 r ( ε ) , h t j i ( ε ) , it follows from the assumption [F](3) and (2.9), (1.2), (2.1) that
H ε ( ε , t ) = lim ξ 0 x ε + ξ ( t ; ε + ξ , x 0 + ( ε + ξ ) v ) x ε ( t ; ε , x 0 + ε v ) ξ + ε y ˜ i ( t , ε ) = lim ξ 0 x ε + ξ t ; h t j 1 r ( ε + ξ ) , Υ r h t j 1 r ( ε + ξ ) x ε t ; h t j 1 r ( ε ) , Υ r h t j 1 r ( ε ) ξ + ε y ˜ i ( t , ε ) = lim ξ 0 h t j 1 r ( ε + ξ ) t 0 1 f x ( s , x ε ( s ; ε , x 0 + ε v ) + θ ( x ε + ξ ( s ; ε + ξ , x 0 + ( ε + ξ ) v ) x ε ( s ; ε , x 0 + ε v ) ) ) x ε + ξ ( s ; ε + ξ , x 0 + ( ε + ξ ) v ) x ε ( s ; ε , x 0 + ε v ) ξ d θ d s + lim ξ 0 Υ r h t j 1 r ( ε + ξ ) , ε + ξ Υ r h t j 1 r ( ε ) , ε ξ lim ξ 0 h t j 1 r ( ε ) h t j 1 r ( ε + ξ ) f s , x ( s ; ε , x 0 + ε v ) d s ξ + ε y ˜ r ( t , ε ) = lim ξ 0 h t j 1 r ( ε + ξ ) t 0 1 f x ( s , x ε ( s ; ε , x 0 + ε v ) + θ ( x ε + ξ ( s ; ε + ξ , x 0 + ( ε + ξ ) v ) x ε ( s ; ε , x 0 + ε v ) ) ) x ε + ξ ( s ; ε + ξ , x 0 + ( ε + ξ ) v ) x ε ( s ; ε , x 0 + ε v ) ξ d θ d s + I + J r y ˜ r h t j 1 r ( ε ) , ε [ h ˙ t j 1 r ( ε ) t y ˜ r h t j 1 r ( ε ) , ε + ε y ˜ r h t j 1 r ( ε ) , ε ] h ˙ t j 1 r ( ε ) f h t j 1 r ( ε ) , y ˜ r h t j 1 r ( ε ) , ε + ε y ˜ r ( t , ε ) .
We also can infer from (5.7) and the above the equality that
H ε ( ε , t ) = ε y ˜ r ( t , ε ) + Φ ε t , h t j 1 r ( ε ) I + J r y ˜ r h t j 1 r ( ε ) , ε [ h ˙ t j 1 r ( ε ) t y ˜ r h t j 1 r ( ε ) , ε + ε y ˜ r h t j 1 r ( ε ) , ε ] h ˙ t j 1 r ( ε ) Φ ε t , h t j 1 r ( ε ) f h t j 1 r ( ε ) , y ˜ r h t j 1 r ( ε ) , ε .
Together with (5.6) and (5.9), by implicit function theorem, we have
h ˙ t j i ( ε ) = Φ ε 1 h t j i ( ε ) , h t j 1 r ( ε ) I + J r y ˜ r h t j 1 r ( ε ) , ε f 1 h t j i ( ε ) , x ε h t j i ( ε ) ; ε , x 0 + ε v y ˙ i 1 h t 2 j ( ε ) · h ˙ t j 1 r ( ε ) t y ˜ r h t j 1 r ( ε ) , ε + ε y ˜ r h t j 1 r ( ε ) , ε ε y ˜ r 1 ( t , ε ) h ˙ t j 1 r ( ε ) Φ ε 1 h t j i ( ε ) , h t j 1 r ( ε ) f h t j 1 r ( ε ) , y ˜ r h t j 1 r ( ε ) , ε f 1 h t j i ( ε ) , x ε h t j i ( ε ) ; ε , x 0 + ε v y ˙ i 1 h t j i ( ε ) .
Further, this means that
h ˙ t j i ( 0 ) = h ˙ t j 1 r ( 0 ) Φ 1 t j i , t j 1 r f t j 1 r , y r t j 1 r I + J r y r t j 1 r y ˙ r t j 1 r f 1 t j i , y i t j i y ˙ i 1 t j i .
This completes the proof. □
Now, we claim the Case ( i i ) . For t 0 , t 1 i , similarly the Case ( i ) , it is not difficult to check the following result
φ ˙ ( t ) = f x ( t , x ( t ; 0 , x 0 ) ) φ ( t ) , t 0 , t 1 i , φ ( 0 ) = v f ( 0 , x 0 ) ,
Combine with Lemma 5.1, we firstly note that
lim ε 0 x ε h t j i ( ε ) ; ε , x 0 + ε v x t j i ; 0 , x 0 ε = lim ε 0 x ε h t j i ( ε ) ; ε , x 0 + ε v x ε t j i ; ε , x 0 + ε v ε + lim ε 0 x ε t j i ; ε , x 0 + ε v x t j i ; 0 , x 0 ε = φ t j i + h ˙ t j i ( 0 ) f t j i , y i t j i .
Together with the assumption [J](2), it follows from h t j i ( ε ) > t j i that
φ t j i + = lim ε 0 x ε h t j i ( ε ) + ; ε , x 0 + ε v x h t j i ( ε ) ; 0 , x 0 ε = lim ε 0 1 ε [ x ε h t j i ( ε ) ; ε , x 0 + ε v + J i x ε h t j i ( ε ) ; ε , x 0 + ε v x h t j i ( ε ) ; t j i , x t j i ; 0 , x 0 + J i x t j i ; 0 , x 0 ] = lim ε 0 1 ε [ x ε h t j i ( ε ) ; ε , x 0 + ε v + J i x ε h t j i ( ε ) ; ε , x 0 + ε v x t j i ; 0 , x 0 J i x t j i ; 0 , x 0 t j i h t j i ( ε ) f s , x s ; 0 , x 0 d s ] = I + J i y i t j i φ t j i + h ˙ t j i ( 0 ) f t j i , y i t j i h ˙ t j i ( 0 ) f t j i , y i t j i = φ t j i + J i y i t j i φ t j i + h ˙ t j i ( 0 ) f t j i , y i t j i .
When h t j i ( ε ) < t j i , we also have
φ t j i + = lim ε 0 x ε t j i ; ε , x 0 + ε v x t j i + ; 0 , x 0 ε = lim ε 0 1 ε [ x ε h t j i ( ε ) ; ε , x 0 + ε v + J i x ε h t j i ( ε ) ; ε , x 0 + ε v x t j i ; 0 , x 0 J i x t j i ; 0 , x 0 t j i h t j i ( ε ) f s , x ε s ; ε , x 0 + ε v d s ] = φ t j i + J i y i t j i φ t j i + h ˙ t j i ( 0 ) f t j i , y i t j i .
Consequently, we have
φ t j i + = φ t j i + J i y i t j i φ t j i + h ˙ t j i ( 0 ) f t j i , y i t j i , i Λ , j = 1 , 2 , , k .
Therefore, when t t j i , t j + 1 r ( j = 1 , 2, ⋯, k 1 ) or t t k r , T , it follows from the assumption [F](3) and (2.9), (1.2), (2.1), (4.7), (5.9), (5.17) that
φ ( t ) = lim θ 0 x θ ( t ; θ , x 0 + θ v ) x ( t ; 0 , x 0 ) θ = lim θ 0 x θ ( t ; h t j i ( θ ) , Υ r ( h t j i ( θ ) ) ) x ( t ; t j i , Υ r ( t j i ) ) θ = lim θ 0 Υ i ( h t j i ( θ ) ) Υ i ( t j i ) ) θ + lim θ 0 h t j i ( θ ) t 0 1 f x ( s , x ( s ; 0 , x 0 ) + ξ ( x θ ( s ; θ , x 0 + θ v ) x ( s ; 0 , x 0 ) ) ) x θ ( s ; θ , x 0 + θ v ) x ( s ; 0 , x 0 ) θ d ξ d s lim θ 0 1 θ t j i h t j i ( θ ) f ( s , x ( s ; 0 , x 0 ) ) d s = h ˙ t j i ( 0 ) f t j i , y i t j i + lim θ 0 h t j i ( θ ) t 0 1 f x ( s , x ( s ; 0 , x 0 ) + ξ ( x θ ( s ; θ , x 0 + θ v ) x ( s ; 0 , x 0 ) ) ) x θ ( s ; θ , x 0 + θ v ) x ( s ; 0 , x 0 ) θ d ξ d s + I + J i y i t j i φ t j i + h ˙ t j i ( 0 ) f t j i , y i t j i .
Thus, combine with (5.16) and (5.18), we obtain from the above equality that
φ ˙ ( t ) = f x ( t , x ( t ; 0 , x 0 ) ) φ ( t ) , t ( 0 , T ] and t t j i , i Λ , j = 1 , 2 , , k , φ ( 0 ) = v f ( 0 , x 0 ) , φ t j i + = φ t j i + J i y i t j i φ t j i + h ˙ t j i ( 0 ) f t j i , y i t j i , j = 1 , 2 , , k .
This completes the proof of Theorem 2.7.

6. Periodicity of Autonomous Impulsive System

As an application, in this section, we discus the periodicity of solution of the following impulsive differential equation
x ˙ ( t ) = g ( x ( t ) ) , x ( t ) y 1 , t 0 , x ( t + ) = y 2 , x ( t ) = y 1 , t 0 , x ( 0 ) = x 0 ,
where y 1 , y 2 R n and y 1 y 2 . We introduce functions:
G ( t ; s , z s ) = z ( t , s , z s ) y 1 , y 2 y 1 for   any t s 0 .
Here,
z ( t , s , z s ) = z s + s t g ( z ( τ , s , z s ) ) d τ , for   any t s 0 .
For the function G ( · ; 0 , x 0 ) , it is clear that
G ( t ; 0 , x 0 ) = 0 has   no   solution   on R +
or
t 1 is   minimum   solution   of G ( t ; 0 , x 0 ) = 0 o n R + .
Similarly, it is obvious that
G ( t ; t 1 , y 2 ) = 0 has   no   solution   on ( t 1 , + )
or
t 2 is   minimum   solution   of G ( t ; t 1 , y 2 ) = 0 o n ( t 1 , + ) .
Let P C y 1 y 2 R + , R n = { x : [ 0 , + ) R n | x is continuous at t when x ( t ) y 1 , x is left continuous at t and the right limit x ( t + ) exists when x ( t ) = y 1 } . We will check the following main result for the autonomous impulsive system (6.1).
Theorem 6.1.
Suppose g : R n R n is locally Lipschitz continuous in x and there exists a constant k ˜ > 0 such that
| g ( x ) | k ˜ ( 1 + | x | ) f o r   a n y t 0 .
(1) If (6.4) holds, then (6.1) has a unique solution x C R + , R n .
(2) If (6.5) and (6.6) holds, then the solution of (6.1) has a unique irregular point t 1 .
(3) If (6.6) and (6.7) holds, then the solution of (6.1) is a periodic function on [ t 1 , + ) .
Proof. 
Using Theorem 2.2, we directly check that the autonomous impulsive system (6.1) has a unique solution x P C y 1 y 2 R + , R n . Further, the solution only three possibilities which are Case (i): x has not irregular point on R + ; Case (ii): x has a unique irregular point on R + and Case (iii): x has two irregular points on R + at least.
For the Case (i), it follows from (2) of Theorem 2.2 that x has not irregular point on R + if and only if (6.4) holds. This means (6.1) has a unique solution x C R + , R n . Similarly, for the Case (ii), together with (6.5) and (6.6), we can infer also that x has only a unique irregular point t 1 .
For the Case (iii), let t 1 and t 2 denote the smallest two the irregular points of solution x on R + and T = t 2 t 1 . We claim
x ( t + T ) = x ( t ) f o r a n y t [ t 1 , + ) .
By the definitions of t 1 and t 2 (see (6.5) and (6.7)), the solution x of (6.1) has not irregular point on ( t 1 , t 2 ) and satisfies
x ( t ) = y 2 + t 1 t g ( x ( s ) ) d s for   any t ( t 1 , t 2 ] and x ( t 2 ) = x ( t 1 ) = y 1 .
When t ( t 1 , t 2 ] , we have t + T ( t 2 , t 2 + T ] and
x ( t + T ) = y 2 + t 1 + T t + T g ( x ( s ) ) d s = y 2 + t 1 t g ( x ( s + T ) ) d s .
It is easy to see that by assumption conditions of g, there exists ρ > 0 such that | x ( t ) | , | x ( T + t ) | ρ for every t ( t 1 , t 2 ] . Furthermore, it asserts from (6.9) and (6.10) that
| x ( t + T ) x ( t ) | t 1 t | g ( x ( s + T ) ) g ( x ( s ) ) | d s
L ( ρ ) t 1 t | x ( s + T ) x ( s ) | d s .
Together with Gronwall inequality, one can verify that
x ( t + T ) = x ( t ) f o r a n y t ( t 1 , t 2 ] .
Consequently, we can infer (6.8) holds. Thus, this means that the solution x of (6.1) is a periodic function on [ t 1 , + ) with period T. The proof is completed. □

7. Application

As an application, in this section, we will discus the variation of solution relative to control for the following control impulsive differential equation
x ˙ ( t ) = f ( t , x ( t ) ) + B ( t ) u ( t ) , { x ( t ) } Y ( t ) = , t 0 , x ( t + ) = J i ( x ( t ) ) + x ( t ) , { x ( t ) } Y ( t ) = y i ( t ) , t 0 , x ( 0 ) = x 0 ,
where, control function u L l o c 1 R + , R m , B L l o c R + , R n × m .
Using the idea of Theorem 2.2 and Theorem 2.6, for any T > 0 and u L 1 ( 0 , T ) , R m , one can prove the following result.
Theorem 7.1.
Suppose the assumptions [F](1)(2), [Y](1) and [J] hold. Then the system (7.1) has a unique P C solution x ( · ; u ) x ( · ; u , 0 , x 0 ) P C Y [ 0 , T ] , R n given by
x ( t ; u ) = x 0 + 0 t [ f τ , x ( τ ; u ) + B ( τ ) u ( τ ) ] d τ + 0 t i < t , x ( t i ; u ) = y j ( t i ) ) J j x ( t i ; u ) .
Moreover, the solution x ( · ; u ) have continuous dependence relative to the control u in the sense of Definition 2.4.
Not only that, for any fixed sufficient small θ > 0 and fixed v L 1 [ 0 , T ] , R m , (7.1) has a unique approximate solution x θ ( · ) x θ ( · ; u + θ v , 0 , x 0 ) which satisfies
x θ ( t ) = x 0 + 0 t [ f τ , x θ ( τ ) + B ( τ ) ( u ( τ ) + θ v ( τ ) ) ] d τ + 0 t i < t , x θ ( t i ) B ( y j ( t i ) , θ 2 ) J j x θ ( t i ) .
To discus variation of solution relative to control, we introduce the following definitions.
Definition 7.2.
The P C solution x · ; u , 0 , x 0 of (7.1) is said to G a ^ teaux differentiable relative to the control u if G a ^ teaux derivative ψ ( · ) of x t ; u exists at u for all t [ 0 , T ] with x t ; u , 0 , x 0 y i ( t ) , otherwise,
ψ ( t ) = lim s t ψ ( s ) ,
where
ψ ( t ) = lim ε 0 x ε t ; u + ε v , 0 , x 0 x t ; u , 0 , x 0 ε w h e n x t ; u , 0 , x 0 y i ( t ) .
Theorem 7.3.
Suppose the assumptions [F], [Y] and [J] hold and u C [ 0 , T ] , R m , B C [ 0 , T ] , R n × m . The P C solution x ( · ) = x · ; u , 0 , x 0 of (7.1) is G a ^ teaux differentiable relative to the control u in the sense of Definition 7.2. Moreover, its G a ^ teaux derivative ψ is a P C solution of the following differential equation with impulses
ψ ˙ ( t ) = f x ( t , x ( t ) ) ψ ( t ) + B ( t ) v ( t ) , t ( 0 , T ] , x t y i ( t ) , i Λ , ψ ( 0 ) = 0 , ψ ( t + ) = ψ ( t ) + J i ( y i ( t ) ) ψ ( t ) + g ˙ t ( 0 ) ( f t , y i t + B ( t ) u ( t ) , x t = y i ( t ) , i Λ .
Proof. 
There are only two possibilities which are Case (I): x · ; u , 0 , x 0 has no irregular point on [ 0 , T ] and Case (II): x · ; u , 0 , x 0 has at least one irregular point on [ 0 , T ] .
In the Case (I), one can directly check that x · ; u , 0 , x 0 of (7.1) is G a ^ teaux differentiable and its G a ^ teaux derivative ψ is a weak solution of the following differential equation
ψ ˙ ( t ) = f x ( t , x ( t ; u ) ) ψ ( t ) + B ( t ) v ( t ) , t ( 0 , T ] , ψ ( 0 ) = 0 .
To discuss the Case (II), we define function g t given by
g t ( ε ) denote   solution   of   the   equation G ( ε , t ) = 0 .
Here,
G ( ε , t ) = x ε t ; u + ε v , 0 , x 0 y ˜ ( t , ε ) .
By Theorem 7.1, when x t ; u , 0 , x 0 = y i ( t ) , there is an δ > 0 such that for all ε [ 0 , δ ] , g t : [ 0 , δ ] O ( t ) is a function and g t ( 0 ) = t , where O ( t ) denote some neighborhood of t. For convenience, let { t j i | 0 < t 1 i < t 2 j < < t k r < T } denote the irregular point set of x · ; u , 0 , x 0 on [ 0 , T ] . If y i C 1 ( [ 0 , T ] , R n ) , it follows that there is a δ > 0 such that
G t ( ε , t ) = f t , x ε t ; u + ε v , 0 , x 0 + B ( t ) [ u ( t ) + ε v ( t ) ] y ˜ t ( t , ε ) .
Further, when f t j i , y i t j i + B t j i u t j i y ˙ i t j i ( j = 1 , 2 , , k , i Λ ), without loss of generality, we suppose
f 1 g t j i ( ε ) , x ε g t j i ( ε ) ; u + ε v , 0 , x 0 + B 1 g t j i ( ε ) u g t j i ( ε ) y ˙ i 1 ( g t j i ( ε ) ) 0 i n R , i Λ , ε [ 0 , δ ] , j = 1 , 2 , , k ,
where B 1 denote the first line vector of B. We introduce the following functions given by
Ψ ε ( t , s ) = exp s t f x ( τ , x ε ( τ ; u + ε v , 0 , x 0 ) ) d τ ,
then
Ψ ( t , s ) = lim ε 0 Ψ ε ( t , s ) = exp s t f x ( τ , x ( τ ; u , 0 , x 0 ) ) d τ .
We let
Ψ ε 1 ( t , s ) a n d Ψ 1 ( t , s ) d e n o t e t h e f i r s t l i n e v e c t o r o f Ψ ε ( t , s ) a n d Ψ ( t , s ) r e s p e c t i v e l y .
Now we calculate the variation of solution relative to control in the Case (II). For t 0 , t 1 i , similarly the Case (I), it is not difficult to check the following result
ψ ˙ ( t ) = f x ( t , x ( t ; u , 0 , x 0 ) ) ψ ( t ) + B ( t ) v ( t ) , t 0 , t 1 i , ψ ( 0 ) = 0 .
When t 0 , g t 1 i ( ε ) , it follows from the assumption [F](3), (7.3) and (2.9) that
G ε ( ε , t ) = lim ξ 0 x ε + ξ ( t ; u + ( ε + ξ ) v , 0 , x 0 ) x ε ( t ; u + ε v , 0 , x 0 ) ξ + ε y ˜ i ( t , ε ) = lim ξ 0 0 t 0 1 f x ( s , x ε ( s ; u + ε v , 0 , x 0 ) + θ ( x ε + ξ ( s ; u + ( ε + ξ ) v , 0 , x 0 ) x ε ( s ; u + ε v , 0 , x 0 ) ) ) x ε + ξ ( s ; u + ( ε + ξ ) v , 0 , x 0 ) x ε ( s ; u + ε v , 0 , x 0 ) ξ d θ d s 0 t B ( s ) v ( s ) d s + ε y ˜ i ( t , ε ) .
It follows from (7.11) and the above that
G ε ( ε , t ) = 0 t Ψ ε ( t , s ) B ( s ) v ( s ) d s + ε y ˜ i ( t , ε ) .
Using implicit function theorem, combine with (7.10), we have
g ˙ t 1 i ( ε ) = 0 g t 1 i ( ε ) Ψ ε 1 ( g t 1 i ( ε ) , s ) B ( s ) v ( s ) d s + ε y ˜ i 1 ( g t 1 i ( ε ) , ε ) f 1 g t 1 i ( ε ) , x ε g t 1 i ( ε ) ; u + ε v , 0 , x 0 + B 1 g t 1 i ( ε ) u g t 1 i ( ε ) y ˙ i 1 ( g t 1 i ( ε ) ) .
In the above equation, vector product is the inner product operation. In the following operations, the vector product is also the inner product operation. Together with Theorem 7.1, we obtain
g ˙ t 1 i ( 0 ) = 0 t 1 i Ψ 1 t 1 i , s B ( s ) v ( s ) d s f 1 t 1 i , x t 1 i ; u , 0 , x 0 + B 1 t 1 i u t 1 i y ˙ i 1 ( t 1 i ) .
Further,
lim ε 0 x ε g t 1 i ( ε ) ; u + ε v , 0 , x 0 x t 1 i ; u , 0 , x 0 ε = lim ε 0 x ε g t 1 i ( ε ) ; u + ε v , 0 , x 0 x ε t 1 i ; u + ε v , 0 , x 0 ε + lim ε 0 x ε t 1 i ; u + ε v , 0 , x 0 x t 1 i ; u , 0 , x 0 ε = ψ t 1 i + g ˙ t 1 i ( 0 ) f t 1 i , y i t 1 i + B t 1 i u t 1 i .
Together with the assumption [J](2), it follows from (7.16) and (7.17) that when g t 1 i ( ε ) > t 1 i ,
ψ t 1 i + = lim ε 0 x ε g t 1 i ( ε ) + ; u + ε v , 0 , x 0 x g t 1 i ( ε ) ; u , 0 , x 0 ε = lim ε 0 1 ε [ x ε g t 1 i ( ε ) ; u + ε v , 0 , x 0 + J i x ε g t 1 i ( ε ) ; u + ε v , 0 , x 0 x g t 1 i ( ε ) ; u , t 1 i , x t 1 i ; u , 0 , x 0 + J i x t 1 i ; u , 0 , x 0 ] = lim ε 0 1 ε [ x ε g t 1 i ( ε ) ; u + ε v , 0 , x 0 + J i x ε g t 1 i ( ε ) ; u + ε v , 0 , x 0 x t 1 i ; u , 0 , x 0 J i x t 1 i ; u , 0 , x 0 t 1 i g t 1 i ( ε ) f s , x s ; u , 0 , x 0 + B ( s ) u ( s ) d s ] = I + J i y i t 1 i ψ t 1 i + g ˙ t 1 i ( 0 ) f t 1 i , y i t 1 i + B t 1 i u t 1 i g ˙ t 1 i ( 0 ) f t 1 i , y i t 1 i + B t 1 i u t 1 i = ψ t 1 i + J i y i t j i ψ t j i + g ˙ t 1 i ( 0 ) f t 1 i , y i t 1 i + B t 1 i u t 1 i ,
and when g t 1 i ( ε ) < t 1 i , we also have
φ t 1 i + = lim ε 0 x ε t 1 i ; u + ε v , 0 , x 0 x t 1 i + ; u , 0 , x 0 ε = lim ε 0 1 ε [ x ε g t 1 i ( ε ) ; u + ε v , 0 , x 0 + J i x ε g t 1 i ( ε ) ; u + ε v , 0 , x 0 x t 1 i ; u , 0 , x 0 J i x t 1 i ; u , 0 , x 0 ] lim ε 0 1 ε t 1 i g t 1 i ( ε ) f s , x ε s ; u + ε v , 0 , x 0 + B ( s ) ( u ( s ) + ε v ( s ) ) d s = φ t 1 i + J i y i t 1 i φ t 1 i + g ˙ t 1 i ( 0 ) f t 1 i , y i t 1 i + B t 1 i u t 1 i .
Consequently, we have
φ t 1 i + = φ t 1 i + J i y i t 1 i φ t 1 i + g ˙ t 1 i ( 0 ) f t 1 i , y i t 1 i , i Λ .
Generally speaking, we firstly note that
lim ε 0 x ε g t j 1 r ( ε ) ; u + ε v , 0 , x 0 x t j 1 r ; u , 0 , x 0 ε = lim ε 0 x ε g t j 1 r ( ε ) ; u + ε v , 0 , x 0 x ε t j 1 r ; u + ε v , 0 , x 0 ε + lim ε 0 x ε t j 1 r ; u + ε v , 0 , x 0 x t j 1 r ; u , 0 , x 0 ε = φ t j 1 r + g ˙ t j 1 r ( 0 ) f t j 1 r , y r t j 1 r + B t j 1 r u t j 1 r .
Further, when t g t j 1 r ( ε ) , g t j i ( ε ) , one can infers from the assumption [F](3) and (7.3), (2.9), (7.19) that
G ε ( ε , t ) = lim ξ 0 x ε + ξ ( t ; u + ( ε + ξ ) v , 0 , x 0 ) x ε ( t ; u + ε v , 0 , x 0 ) ξ + ε y ˜ i ( t , ε ) = lim ξ 0 1 ξ [ x ε + ξ t ; u + ( ε + ξ ) v , g t j 1 r ( ε + ξ ) , Υ r g t j 1 r ( ε + ξ ) x ε t ; u + ε v , g t j 1 r ( ε ) , Υ r g t j 1 r ( ε ) ] + ε y ˜ i ( t , ε ) = lim ξ 0 g t j 1 r ( ε + ξ ) t 0 1 f x ( s , x ε ( s ; u + ε v , 0 , x 0 ) + θ ( x ε + ξ ( s ; u + ( ε + ξ ) v , 0 , x 0 ) x ε ( s ; u + ε v , 0 , x 0 ) ) ) x ε + ξ ( s ; u + ( ε + ξ ) v , 0 , x 0 ) x ε ( s ; u + ε v , 0 , x 0 ) ξ d θ d s + lim ξ 0 Υ r g t j 1 r ( ε + ξ ) , ε + ξ Υ r g t j 1 r ( ε ) , ε ξ + lim ξ 0 g t j 1 r ( ε + ξ ) t B ( s ) v ( s ) d s lim ξ 0 g t j 1 r ( ε ) g t j 1 r ( ε + ξ ) f s , x ( s ; u + ε v , 0 , x 0 ) + B ( s ) ( u ( s ) + ε v ( s ) ) d s ξ + ε y ˜ r ( t , ε ) = lim ξ 0 g t j 1 r ( ε + ξ ) t 0 1 f x ( s , x ε ( s ; u + ε v , 0 , x 0 ) + θ ( x ε + ξ ( s ; u + ( ε + ξ ) v , 0 , x 0 ) x ε ( s ; u + ε v , 0 , x 0 ) ) ) x ε + ξ ( s ; u + ( ε + ξ ) v , 0 , x 0 ) x ε ( s ; u + ε v , 0 , x 0 ) ξ d θ d s + g t j 1 r ( ε ) t B ( s ) v ( s ) d s + ψ g t j 1 r ( ε ) + ε y ˜ r ( t , ε ) + J r y ˜ r g t j 1 r ( ε ) , ε [ ψ g t j 1 r ( ε ) + g ˙ t j 1 r ( ε ) f g t j 1 r ( ε ) , y ˜ r g t j 1 r ( ε ) , ε + B g t j 1 r ( ε ) u g t j 1 r ( ε ) ] .
Moreover, one can see from (7.11) and the above equality that
G ε ( ε , t ) = ε y ˜ r ( t , ε ) + Ψ ε t , g t j 1 r ( ε ) J r y ˜ r g t j 1 r ( ε ) , ε [ ψ g t j 1 r ( ε ) + g ˙ t j 1 r ( ε ) f g t j 1 r ( ε ) , y ˜ r g t j 1 r ( ε ) , ε + B g t j 1 r ( ε ) u g t j 1 r ( ε ) ] + Ψ ε t , g t j 1 r ( ε ) ψ g t j 1 r ( ε ) + g t j 1 r ( ε ) t Ψ ε t , s B ( s ) v ( s ) d s .
Together with (7.10), by implicit function theorem, we have
g ˙ t j i ( ε ) = Ψ ε 1 g t j i ( ε ) , g t j 1 r ( ε ) J r y ˜ r g t j 1 r ( ε ) , ε f 1 g t j i ( ε ) , x ε g t j i ( ε ) ; u + ε v , 0 , x 0 + B 1 g t j i ( ε ) u g t j i ( ε ) y ˙ i 1 ( g t j i ( ε ) ) · ψ g t j 1 r ( ε ) + g ˙ t j 1 r ( ε ) f g t j 1 r ( ε ) , y ˜ r g t j 1 r ( ε ) , ε + B g t j 1 r ( ε ) u g t j 1 r ( ε ) ε y ˜ r 1 g t j i ( ε ) , ε + Ψ ε 1 g t j i ( ε ) , g t j 1 r ( ε ) ψ g t j 1 r ( ε ) + g t j 1 r ( ε ) g t j i ( ε ) Ψ ε 1 t , s B ( s ) v ( s ) d s f 1 g t j i ( ε ) , x ε g t j i ( ε ) ; u + ε v , 0 , x 0 + B 1 g t j i ( ε ) u g t j i ( ε ) y ˙ i 1 ( g t j i ( ε ) ) .
Further, it follows from the above expression and (7.12), Theorem 7.1 that
g ˙ t j i ( 0 ) = Ψ 1 t j i , t j 1 r J r y r t j 1 r f 1 t j i , x t j i ; u , 0 , x 0 + B 1 t j i u t j i y i 1 t j i · ψ t j 1 r + g ˙ t j 1 r ( 0 ) f t j 1 r , y r t j 1 r + B t j 1 r u t j 1 r Ψ 1 t j i , t j 1 r ψ t j 1 r + t j 1 r t j i Ψ 1 t j i , s B ( s ) v ( s ) d s f 1 t j i , x t j i ; u , 0 , x 0 + B 1 t j i u t j i y ˙ i 1 t j i , i Λ , j = 1 , 2 , , k .
Similar to (7.19), we can obtain
lim ε 0 x ε g t j i ( ε ) ; u + ε v , 0 , x 0 x t j i ; u , 0 , x 0 ε = ψ t j i + g ˙ t j i ( 0 ) f t j i , y i t j i + B t j i u t j i , i Λ , j = 1 , 2 , , k .
Together with the assumption [J](2) and (7.22), (7.21), it follows that when g t j i ( ε ) > t j i ,
φ t j i + = lim ε 0 x ε g t j i ( ε ) + ; u + ε v , , x 0 x g t j i ( ε ) ; u , 0 , x 0 ε = lim ε 0 1 ε [ x ε g t j i ( ε ) ; u + ε v , 0 , x 0 + J i x ε g t j i ( ε ) ; u + ε v , 0 , x 0 x g t j i ( ε ) ; u , t j i , x t j i ; u , 0 , x 0 + J i x t j i ; u , 0 , x 0 ] = lim ε 0 1 ε [ x ε g t j i ( ε ) ; u + ε v , 0 , x 0 + J i x ε g t j i ( ε ) ; u + ε v , 0 , x 0 x t j i ; u , 0 , x 0 J i x t j i ; u , 0 , x 0 t j i g t j i ( ε ) [ f s , x s ; u , 0 , x 0 + B ( s ) u ( s ) ] d s ] = I + J i y i t j i ψ t j i + g ˙ t j i ( 0 ) f t j i , y i t j i + B t j i u t j i g ˙ t j i ( 0 ) f t j i , y i t j i + B t j i u t j i = ψ t j i + J i y i t j i ψ t j i + g ˙ t j i ( 0 ) f t j i , y i t j i + B t j i u t j i ,
and when g t j i ( ε ) < t j i ,
φ t j i + = lim ε 0 x ε t j i ; u + ε v , 0 , x 0 x t j i + ; u , 0 , x 0 ε = lim ε 0 1 ε [ x ε g t j i ( ε ) ; u + ε v , 0 , x 0 + J i x ε g t j i ( ε ) ; u + ε v , 0 , x 0 x t j i ; u , 0 , x 0 J i x t j i ; u , 0 , x 0 t j i g t j i ( ε ) f s , x ε s ; u + ε v , 0 , x 0 + B ( s ) ( u ( s ) + ε v ( s ) ) d s ] = ψ t j i + J i y i t j i ψ t j i + g ˙ t j i ( 0 ) f t j i , y i t j i + B t j i u t j i .
Consequently, we have
ψ t j i + = ψ t j i + J i y i t j i ψ t j i + g ˙ t j i ( 0 ) f t j i , y i t j i + B t j i u t j i
for i Λ , j = 1 , 2 , , k . Therefore, when t t j i , t j + 1 r ( j = 1 , 2, ⋯, k 1 ) or t t k r , T , it follows from the assumption [F](3) and (2.9), (7.3), (2.1), (4.7), (7.21), (7.22) that
ψ ( t ) = lim θ 0 x θ ( t ; u + θ v , 0 , x 0 ) x ( t ; u , 0 , x 0 ) θ = lim θ 0 x θ ( t ; u + θ v , g t j i ( θ ) , Υ i ( g t j i ( θ ) ) ) x ( t ; u , t j i , Υ i ( t j i ) ) θ = lim θ 0 Υ i ( g t j i ( θ ) ) Υ i ( t j i ) ) θ + lim θ 0 g t j i ( θ ) t 0 1 f x ( s , x ( s ; u , 0 , x 0 ) + ξ ( x θ ( s ; u + θ v , 0 , x 0 ) x ( s ; u , 0 , x 0 ) ) ) x θ ( s ; u + θ v , 0 , x 0 ) x ( s ; u , 0 , x 0 ) θ d ξ d s + lim θ 0 g t j i ( θ ) t B ( s ) v ( s ) d s lim θ 0 1 θ t j i g t j i ( θ ) [ f ( s , x ( s ; u , 0 , x 0 ) ) + B ( s ) u ( s ) ] d s = ψ t j i + J i y i t j i ψ t j i + g ˙ t j i ( 0 ) f t j i , y i t j i + B t j i u t j i + t j i t B ( s ) v ( s ) d s + lim θ 0 g t j i ( θ ) t 0 1 f x ( s , x ( s ; u , 0 , x 0 ) + ξ ( x θ ( s ; u + θ v , 0 , x 0 ) x ( s ; u , 0 , x 0 ) ) ) x θ ( s ; u + θ v , 0 , x 0 ) x ( s ; u , 0 , x 0 ) θ d ξ d s .
Thus, it follows from (7.14), (7.18), (7.23) that
ψ ˙ ( t ) = f x ( t , x ( t ; u , 0 , x 0 ) ) ψ ( t ) + B ( t ) v ( t ) , t ( 0 , T ] a n d t t j i , i Λ , j = 1 , 2 , , k , ψ ( 0 ) = 0 , ψ t j i + = I + J i y i t j i ψ t j i + g ˙ t j i ( 0 ) J i y i t j i f t j i , y i t j i + B t j i u t j i , j = 1 , 2 , , k .
This completes the proof of Theorem 7.3. □

8. Conclusions

In this paper, we proposed a class of widely applied impulsive differential systems and gave it’s qualitative theory under some weaker conditions, including the existence and uniqueness, periodicity of the solution, the continuous dependence and differentiability of the solution on the initial value. For the pulse phenomena of solution, it is significant to give the sufficient and necessary conditions. It is very interesting that the pulse may destroy the intrinsical properties of the system, such as the existence, the continuous dependence and differentiability of solution. Moreover, these results given in this paper also lay a theoretical foundation for the optimal control problem given by the impulsive different systems with impulses at variable times and the application of such systems.

Acknowledgments

The authors are grateful to the referees for their helpful comments. This work was partially supported by the National Natural Science Foundation of China (No. 12061021 and No. 11161009).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Bainov, D.; Simeonov, P. Impulsive differential equations: periodic solutions and applications, 1rd ed.; Longman Scientific and Technical: New York, USA, 1993; pp. 39–58. [Google Scholar]
  2. Kobayashi, Y.; Nakano, H.; Saito, T. A simple chaotic circuit with impulsive switch depending on time and state. Nonlinear Dynam. 2006, 44, 73–79. [Google Scholar] [CrossRef]
  3. Touboul, J.; Brette, R. Spiking dynamics of Bidimensional integrate-and-fire neurons. SIAM J. Appl. Dyn. Syst. 2009, 8, 1462–1506. [Google Scholar] [CrossRef]
  4. Izhikevich, E. M. Dynamical systems in neuroscience, 1rd ed.; MIT Press: Massachusetts, USA, 2007. [Google Scholar]
  5. Huang, M.; Li, J.; Song, X.; Guo, H. Modeling impulsive injections of insulin: towards artificial pancreas. SIAM J. Appl. Math. 2012, 72, 1524–1548. [Google Scholar] [CrossRef]
  6. Zhang, Q.; Tang, B.; Cheng, T.; Tang, S. Bifurcation analysis of a generalized impulsive Kolmogorov model with applitions to pest and disease control. SIAM J. Appl. Math. 2020, 80, 1796–1819. [Google Scholar] [CrossRef]
  7. Catll<i>a</i>´, A. J.; Schaeffer, D. G.; Witelski, T. P.; Monson, E. E. On spiking models for synaptic activity and impulsive differential equations. SIAM Rev. 2008, 50, 553–569. [Google Scholar]
  8. Akhmet, M.; Yilmaz, E. impulsive differential equations. In Neural networks with discontinuous/impact activations; Akhmet, M., Yilmaz, E., Eds.; Springer: New York, USA, 2014; pp. 67–83. [Google Scholar]
  9. Lions, P. L.; Perthame, B. Quasi-variational inequalities and ergodic impulse control. SIAM J. Control Optim. 1986, 24, 604–615. [Google Scholar] [CrossRef]
  10. Wang, Y.; Lu, J. Some recent results of analysis and control for impulsive systems. Commun Nonlinear Sci Numer Simulat 2019, 80, 1–15. [Google Scholar] [CrossRef]
  11. Krylov, M.M.; Bogolyubov, N.N. Introduction to nonlinear mechanics. Academiya Nauk Ukrin: SSR, Kiev, 1937. (in Russian) [Google Scholar]
  12. Samoilenko, A. M.; Perestyuk, N. A. The method of averaging in systems with an impulsive action. Ukrainian Mathematical Journal 1974, 26, 342–347. [Google Scholar] [CrossRef]
  13. Ahmed, N. U. Existence of optimal controls for a general class of impulsive systems on Banach spaces. SIAM J. Control Optim. 2003, 42, 669–685. [Google Scholar] [CrossRef]
  14. Peng, Y.; Xiang, X. A class of nonlinear impulsive differential equation and optimal controls on time scales. Discrete Cont. Dyn.-B 2011, 16, 1137–1155. [Google Scholar] [CrossRef]
  15. Nain, A. K.; Vats, R. K.; Kumar, A. Caputo-Hadamard fractional differential equation with impulsive boundary conditions. J. Math. Model. 2021, 9, 93–106. [Google Scholar]
  16. Malik, M.; Kumar, A. Existence and controllability results to second order neutral differential equation with non-instantaneous impulses. Journal of Control and Decision 2020, 7, 286–308. [Google Scholar] [CrossRef]
  17. Samoilenko, A. M.; Perestyuk, N. A. On stability of solutions of impulsive systems. Differentsial’nye Uravneniya 1981, 17, 1995–2002. [Google Scholar]
  18. Bainov, D. D.; Dishliev, A. B. Sutlicient conditions for absence of beating in systems of differential equations with impulses. Appl. Anal. 1984, 18, 67–73. [Google Scholar]
  19. Hu, S.; Lakshmikantham, V. Periodic boundary value problem for second order impulsive differential systems. Nonlinear Analysis TMA 1989, 13, 75–85. [Google Scholar] [CrossRef]
  20. Lakshmikantham, V.; Bainov, D. D.; Simenonov P., S. Theory of Impulsive Differential Equences, 1rd ed.; Wold Scientific: Hong Kong, china, 1989. [Google Scholar]
  21. Samoilenko, A. M. Application of the method of averaging for the investigation of vibrations excited by instantaneous impulses in self-vibrating systems of the second order with a small parameter. Ukrainian Mathematical Journal 1961, 13, 103–108. [Google Scholar]
  22. Bainov, D.; Simeonov, P. Impulsive differential equations, 1rd ed.; Wold Scientific: Singapore, 1995. [Google Scholar]
  23. Benchohra, M.; Henderson, J.; Ntouyas, S. Impulsive differential equations and inclusions, 1rd ed.; Hindawi Publishing Corporation: New York, USA, 2006. [Google Scholar]
  24. Peng, Y.; Wu, K.; Qin, S.; Kang, Y. Properties of solution of linear controlled systems with impulses at variable times. In Proceedings of the 36th Chinese Control Conference, Dalian, China, 26–28 July 2017. [Google Scholar]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.
Copyright: This open access article is published under a Creative Commons CC BY 4.0 license, which permit the free download, distribution, and reuse, provided that the author and preprint are cited in any reuse.
Prerpints.org logo

Preprints.org is a free preprint server supported by MDPI in Basel, Switzerland.

Subscribe

Disclaimer

Terms of Use

Privacy Policy

Privacy Settings

© 2025 MDPI (Basel, Switzerland) unless otherwise stated