Home Numerical approach to the controllability of fractional order impulsive differential equations
Article Open Access

Numerical approach to the controllability of fractional order impulsive differential equations

  • Avadhesh Kumar , Ramesh K. Vats , Ankit Kumar and Dimplekumar N. Chalishajar EMAIL logo
Published/Copyright: September 19, 2020
Become an author with De Gruyter Brill

Abstract

In this manuscript, a numerical approach for the stronger concept of exact controllability (total controllability) is provided. The proposed control problem is a nonlinear fractional differential equation of order α(1,2] with non-instantaneous impulses in finite-dimensional spaces. Furthermore, the numerical controllability of an integro-differential equation is briefly discussed. The tool for studying includes the Laplace transform, the Mittag-Leffler matrix function and the iterative scheme. Finally, a few numerical illustrations are provided through MATLAB graphs.

MSC 2010: 34A08; 34K45; 93B05; 34G20

1 Introduction

Sometimes, integer-order differential equation becomes inadequate to model some physical phenomena such as in anomalous diffusion. Such phenomena give rise to fractional order differential equations. The main advantage of studying fractional order systems is that they allow greater degrees of freedom in the model. Differential equations of fractional order appear more often in diverse areas of science and engineering, such as image processing, signal processing, bio-engineering, viscoelasticity, fluid flow and control theory [1,2,3,4,5,6]. For the fundamental understanding of fractional calculus and related numerical methods, one can refer to [7,8,9,10,11].

On the other hand, some phenomena are characterized by rapid changes. The first kind of changes takes place over a relatively short time compared to the overall duration of the entire process. Mathematical models in these cases are developed using impulsive differential equations. In the second kind, the changes are not negligibly short in duration and these changes begin impulsively at some points and remain active over certain time intervals. The mathematical model of these situations gives rise to a differential equation with non-instantaneous impulses.

The study of non-instantaneous impulsive differential equations has significant applications in different areas, for example, in hemodynamical equilibrium and the theory of rocket combustion. An excellent application of non-instantaneous impulse is the introduction of insulin into the bloodstream. It produces an abrupt change in the bloodstream. The consequent absorption is a gradual process that remains active over a finite time span.

Recently, many researchers have shown their interest in existence, uniqueness of solutions, stability and controllability of impulsive problems with non-instantaneous impulses [12,13,14,15,16]. Hernández and Regan [17] studied mild and classical solutions for the impulsive differential equation with non-instantaneous impulses. Wang and Fečkan [18] have shown existence, uniqueness and stability of solutions of such a general class of first-order impulsive differential equations. Later, Muslim et al. [12] investigated existence, uniqueness of solutions and stability of second-order differential equations with non-instantaneous impulses. However, the controllability of the non-instantaneous impulsive control system is the less treated topic as compared to the existence and uniqueness of solutions.

In the setting of controllability, the control system is an interconnection of components forming a system configuration that will result in a desired system response. Controllability is one of the structural properties of dynamical systems. It provides the ability to move a system around entire configuration space using only certain feasible manipulations. It deals with whether or not the state of a state-space dynamic system can be controlled from the input. Many authors dealt with controllability problems that can be found in many recently published papers [19,20,21,22].

Recently, Wang et al. [19] discussed controllability of fractional non-instantaneous impulsive differential inclusions. However, Wang et al. achieved exact controllability by only applying control in the last subinterval of time. But, Wang et al. did not propose any computational scheme for the steering control. In this manuscript, the control is applied for each subinterval of time, due to which the concept of total controllability arises. Moreover, none of the research papers have so far discussed the numerical approach for the controllability of the non-instantaneous impulsive differential equation of order α(1,2]. Therefore, this manuscript is devoted to the study of numerical controllability for the following fractional order nonlinear differential equation with non-instantaneous impulses in a space n:

(1.1)DtαCu(t)=Au(t)+Bw(t)+g(t,u(t)),t(si,ti+1],i=0,1,,m,u(t)=Ψi1(t,u(ti)),t(ti,si],i=1,2,,m,u(t)=Ψi2(t,u(ti)),t(ti,si],i=1,2,,m,u(0)=u0,u(0)=v0,

where α(1,2] and u(t) is a state function with time interval 0=s0=t0<t1<s1<t2,,tm<sm<tm+1=T<. Let A be a coefficient matrix of system (1.1). The control function w()L2(J1i=0m[si,ti+1],m). Let B be a bounded linear operator from m to n. Consider the state function uC((ti,ti+1],n),i=0,1,,m, and there exist u(ti) and u(ti+),i=1,2,,m with u(ti)=u(ti). The functions Ψi1(t,u(ti)) and Ψi2(t,u(ti)) represent non-instantaneous impulses during the intervals (ti,si],i=1,2,,m. Ψi1,Ψi2 and g are suitable functions. These functions will be explained briefly in the subsequent sections.

The manuscript proceeds as follows. In Sections 1–3, the introduction, notations, results and required assumptions are given, which will be required for the later sections. In Section 4, controllability of problem (1.1) is investigated by the iterative scheme. Later, controllability of the integro-differential equation is briefly mentioned in Section 5. In Section 6, a few numerical examples are given to show the application of the obtained results.

2 Preliminaries and assumptions

In this section, some useful definitions related to fractional calculus are briefly reviewed. Also, some necessary properties concerned with the Mittag-Leffler function are discussed.

Let L1(+,n) and C1(+,n),+=(0,) be the space of all integrable functions and continuously differentiable functions, respectively. Furthermore, n be the n-dimensional Euclidean space, where n.

Definition 2.1

[24] The Riemann-Liouville fractional integral of order α>0 is defined by

Jtαg(t)=1Γ(α)0t(ts)α1g(s)ds,

where g(t)L1(+,n) and Γ() is the gamma function.

Definition 2.2

[24] If g(t)L1(+,n), then the Riemann-Liouville fractional derivative of order α(1,2] is defined by

DtαRLg(t)=d2dt2Jt2αg(t),

where DtαRLg(t)L1(+,n).

Definition 2.3

[24] The Caputo fractional derivative of order α(1,2] is defined by

DtαCg(t)=Jt2αd2dt2g(t),

where g(t)L1(+,n)C1(+,n).

The Laplace transform of the Riemann-Liouville fractional derivatives is defined by

{DtαRLg(t)}=sαG(s)k=1nsk1g(αk)(0+).

The Laplace transform of the Caputo fractional derivatives is defined by

{DtαCg(t)}=sαG(s)k=1nsαkg(k1)(0+).

In particular, if α(1,2], then

{DtαCg(t)}=sαG(s)g(0+)sα1g(0+)sα2.

2.1 Mittag-Leffler function

The Mittag-Leffler function is a generalization of the exponential function, and it plays an important role in the solution of the fractional differential equations.

Definition 2.4

[23] A function of the complex variable z defined by

(2.1)Eα(z)=k=0zkΓ(αk+1)

is called the one-parameter Mittag-Leffler function.

In particular, when α=1, we obtain

E1(z)=k=0zkΓ(k+1)=k=0zkk!=ez,

i.e., the classical exponential function.

An extension of the one-parameter Mittag-Leffler function is given by the following two-parameter function.

Definition 2.5

[23] A function of the complex variable z defined by

(2.2)Eα,β(z)=k=0zkΓ(αk+β),α,β>0

is called the two-parameter Mittag-Leffler function.

For β=1, from (2.2) we obtain (2.1). Moreover, we have the following identities:

E1,1(z)=k=0zkΓ(k+1)=k=0zkk!=ez,E0,1(z)=k=0zk=11z,E2,1(z2)=k=0(1)kz2kΓ(2k+1)=cosz,E2,2(z2)=k=0(1)kz2kΓ(2k+2)=sinzz.

Definition 2.6

[23] A function of the matrix An×n defined by

Eα,β(A)=k=0AkΓ(αk+β),α,β>0

is called the two-parameter Mittag-Leffler matrix function.

Let us consider the following fractional impulsive differential equation of order α(1,2] without control term Bw(t):

(2.3)DtαCu(t)=Au(t)+g(t),t(si,ti+1],i=0,1,,m,u(t)=Ψi1(t,u(ti)),t(ti,si],i=1,2,,m,u(t)=Ψi2(t,u(ti)),t(ti,si],i=1,2,,m,u(0)=u0,u(0)=v0,

where un and An×n is a coefficient matrix of system (2.3). Let g:J1n, J1=i=0m[si,ti+1] be a continuous function. The solution of system (2.3) is obtained by applying the Laplace transform method. It is given by the following integral equation [23,24]:

(2.4)u(t)=Φ0(tsi)Ψi1(si,u(ti))+Φ1(tsi)Ψi2(si,u(ti))+sitΦ(ts)g(s)ds,t[si,ti+1],i=0,1,2,,m,

where

Φ0(t)=Eα(Atα),Φ1(t)=tEα,2(Atα),Φ(t)=tα1Eα,α(Atα)

and

Ψi1(0,)=u0,Ψi2(0,)=v0.

3 Controllability for the linear system

We consider the linear impulsive differential equation of order α(1,2] with control function u(t) as follows:

(3.1)DtαCu(t)=Au(t)+Bw(t),t(si,ti+1],i=0,1,,m,u(t)=Ψi1(t,u(ti)),t(ti,si],i=1,2,,m,u(t)=Ψi2(t,u(ti)),t(ti,si],i=1,2,,m,u(0)=u0,u(0)=v0,

where un and control function w()L2(J1i=0m[si,ti+1],m). Let B be a bounded linear operator from m to n. The solution of the linear impulsive system (3.1) is given by the following integral equation:

(3.2)u(t)=Φ0(tsi)Ψi1(si,u(ti))+Φ1(tsi)Ψi2(si,u(ti))+sitΦ(ts)Bw(s)ds,t[si,ti+1],i=0,1,2,,m.

Definition 3.1

(Exact controllability) [22] System (3.1) is said to be exactly controllable on [0,T], if for the initial state u(0)n and arbitrary final state uf(tm+1)n, there exists a control wL2(J1,m) such that solution (3.2) satisfies u(tm+1)=uf(tm+1).

Definition 3.2

(Total controllability) [22] System (3.1) is said to be totally controllable on [0,T], if for the initial state u(0)n and arbitrary final state uf(ti+1)n of each sub-interval [si,ti+1], there exists a control wL2(J1,m) such that solution (3.2) satisfies u(ti+1)=uf(ti+1), where i=0,1,,m.

Remark

Total controllability Exact controllability.

Lemma 3.3

[21,25] The linear system (3.1) is controllable on[0,T]iff the controllability Gramian

Msiti+1=siti+1Φ(ti+1s)BBΦ(ti+1s)ds
is nonsingular, wheres0=0,ti+1=T,i=0,1,2,,m.

In order to prove the controllability for the nonlinear system (1.1), the following assumptions are taken:

(A1) g:J1×nn, J1=i=0m[si,ti+1] is a continuous function and there exist positive constants λ1 and λ2 such that

g(t,u)g(t,v)λ1uv

and g(t,u)λ2, for every u,vn,tJ1.

(A2) There exist positive constants CΨi1andCΨi2,i=1,2,,m such that CΨi1=maxtIiΨi1(t,)andCΨi2=maxtIiΨi2(t,), where Ii[ti,si].

(A3) ΨikC(Ii×n,n) and there are positive constants LΨik,i=1,2,,m,k=1,2, such that Ψik(t,u)Ψik(t,v)LΨikuv, tIiandu,vn.

Let =PC([0,T],n) be the space of piecewise continuous functions.

PC([0,T],n){u:[0,T]n:uC((tk,tk+1],n),k=0,1,,m, and there exist u(tk) and u(tk+),k=1,2,,m}. It can be easily proved that PC([0,T],n), for all t[0,T], is a Banach space endowed with the supremum norm. For the sake of notational convenience, let us define

KΦ0=supt[0,T]Φ0(t);KΦ1=supt[0,T]Φ1(t);KΦ=supt[0,T]Φ(t);KM=(Msiti+1)1;KB=B.

Lemma 3.4

If all the assumption (A2) is fulfilled, then the control function for problem (3.1) has an estimatew(t)δ,t[si,ti+1],i=0,1,2,,m, where

δ=KBKΦKM(uf(ti+1)+KΦ0CΨi1+KΦ1CΨi2).

Proof

The control function for t[si,ti+1],i=0,1,2,,m, is defined as follows:

(3.3)w(t)=BΦ(ti+1t)(Msiti+1)1uf(ti+1)Φ0(ti+1si)(Ψi1(si,u(ti)))Φ1(ti+1si)(Ψi2(si,u(ti))),

where uf(ti+1) is the arbitrary final state of each sub-interval [si,ti+1],i=0,1,,m.

By solution (3.2), the final state at t=ti+1 is

u(ti+1)=Φ0(ti+1si)(Ψi1(si,u(ti)))+Φ1(ti+1si)(Ψi2(si,u(ti)))+siti+1Φ(ti+1s)Bw(s)ds=Φ0(ti+1si)(Ψi1(si,u(ti)))+Φ1(ti+1si)(Ψi2(si,u(ti)))+siti+1Φ(ti+1s)BBΦ(ti+1t)ds×(Msiti+1)1uf(ti+1)Φ0(ti+1si)(Ψi1(si,u(ti)))Φ1(ti+1si)(Ψi2(si,u(ti)))=Φ0(ti+1si)(Ψi1(si,u(ti)))+Φ1(ti+1si)(Ψi2(si,u(ti)))+(Msiti+1)(Msiti+1)1uf(ti+1)Φ0(ti+1si)(Ψi1(si,u(ti)))Φ1(ti+1si)(Ψi2(si,u(ti)))=uf(ti+1).

Hence, control function (3.3) is suitable for problem (3.1), for every t[si,ti+1] and i=0,1,2,,m. The estimate of control function w(t) is given by:

w(t)BΦ(ti+1t)Msiti+11uf(ti+1)+Φ0(ti+1si)Ψi1(si,u(ti))+Φ1(ti+1si)Ψi2(si,u(ti))KBKΦKM(uf(ti+1)+KΦ0CΨi1+KΦ1CΨi2).

Hence, the required estimate for control (3.3) is obtained.□

4 Controllability for the nonlinear system

Steering of a dynamical control system from an arbitrary initial state to an arbitrary final state on each sub-interval [tk,tk+1] using the set of admissible controls is called a totally controllable system. In this section, total controllability of system (1.1) is investigated through the iterative scheme.

Theorem 4.1

If all the assumptions (A1)–(A3) are satisfied and the linear system (3.1) is controllable, then the nonlinear system (1.1) is totally controllable on[0,T].

Proof

In order to prove the controllability results, we adopt the successive approximation technique. Let us define the iterative scheme as follows:

(4.1)u0(t)=u0un+1(t)=Φ0(tsi)Ψi1(si,un(ti))+Φ1(tsi)Ψi2(si,un(ti))+sitΦ(ts)Bwn(s)ds+sitΦ(ts)g(s,un(s))ds,t[si,ti+1],i=0,1,2,,m,

where

(4.2)wn(t)=BΦ(ti+1t)(Msiti+1)1uf(ti+1)Φ0(ti+1si)Ψi1(si,un(ti))Φ1(ti+1si)Ψi2(si,un(ti))siti+1Φ(ti+1s)g(s,un(s))ds

and n=0,1,2,.

Since the initial vector u0 is given, the sequence {un(t)} can be easily obtained. We will show that the sequence {un(t)} is Cauchy in . Moreover, we observe that

(4.3)wn(t)BΦ(ti+1t)Msiti+11uf(ti+1)+Φ0(ti+1si)Ψi1(si,un(ti))+Φ1(ti+1si)Ψi2(si,un(ti))+siti+1Φ(ti+1s)g(s,un(s))dsKBKΦKM(uf(ti+1)+KΦ0CΨi1+KΦ1CΨi2+λ2KΦT)

and

(4.4)wn(t)wn1(t)BΦ(ti+1t)Msiti+11Φ0(ti+1si)Ψi1(si,un(ti))Ψi1(si,un1(ti))+Φ1(ti+1si)Ψi2(si,un(ti))Ψi2(si,un1(ti))+siti+1Φ(ti+1s)g(s,un(s))g(s,un1(s))dsKBKΦKM[KΦ0LΨi1+KΦ1LΨi2+λ1KΦT]un(t)un1(t)Ωun(t)un1(t),

where Ω=KBKΦKM[KΦ0LΨi1+KΦ1LΨi2+λ1KΦT]. Furthermore, we have

(4.5)un+1(t)un(t)Φ0(tsi)Ψi1(si,un(ti))Ψi1(si,un1(ti))+Φ1(tsi)Ψi2(si,un(ti))Ψi2(si,un1(ti))+sitΦ(ts)Bwn(s)wn1(s)ds+sitΦ(ts)g(s,un(s))g(s,un1(s))ds[KΦ0LΨi1+KΦ1LΨi2+KBKΦΩT+λ1KΦT]un(t)un1(t)Θun(t)un1(t),

where Θ=[KΦ0LΨi1+KΦ1LΨi2+KBKΦΩT+λ1KΦT]. Moreover, it can be observed that

(4.6)u1(t)u0(t)Φ0(tsi)Ψi1(si,u0(ti))+Φ1(tsi)Ψi2(si,u0(ti))+u0+sitΦ(ts)Bw0(s)ds+sitΦ(ts)g(s,u0(s))ds[KΦ0CΨi1+KΦ1CΨi2+u0+(KBKΦΩT+λ1KΦT)][KΦ0CΨi1+KΦ1CΨi2+u0+(KBKΦΩ+λ1KΦ)T]KT,K>0.

By using inequality (4.6) and the method of induction, the estimate for inequality (4.5) is as follows:

(4.7)un+1(t)un(t)KΘnTn+1n!.

The right-hand side in the aforementioned estimate (4.7) can be made arbitrarily small by choosing sufficiently large value of n. This implies that {un(t)} is a Cauchy sequence in . Since is a Banach space, the sequence {un(t)} converges uniformly to a continuous function u(t) on [0,T]. It is followed by taking limit as n on both sides of (4.1) and (4.2). Thus, we have

(4.8)u(t)=Φ0(tsi)Ψi1(si,u(ti))+Φ1(tsi)Ψi2(si,u(ti))+sitΦ(ts)Bw(s)ds+sitΦ(ts)g(s,u(s))ds,t[si,ti+1],i=0,1,2,,m,

where

(4.9)w(t)=BΦ(ti+1t)(Msiti+1)1uf(ti+1)Φ0(ti+1si)Ψi1(si,u(ti))Φ1(ti+1si)Ψi2(si,u(ti))siti+1Φ(ti+1s)g(s,u(s))ds.

Since, the control u(t) steers the system from the initial state u0 to final state uf(ti+1) at time ti+1, system (1.1) is said to be totally controllable on [0,T].□

5 Controllability for an integro-differential equation

In this section, a control system represented by an integro-differential equation in space n is considered as follows:

(5.1)DtαCu(t)=Au(t)+Bw(t)+g(t,u(t))+0tξ(ts)h(s,u(s))ds,ti=0m(si,ti+1],u(t)=Ψi1(t,u(ti)),ti=1m(ti,si],u(t)=Ψi2(t,u(ti)),ti=1m(ti,si],u(0)=u0,u(0)=v0,

where α(1,2] and u(t) is a state function with time interval 0=s0=t0<t1<s1<t2,,tm<sm<tm+1=T<. Let An×n be a coefficient matrix of system (5.1). The control function w()L2(J1,m) and B:mn is a bounded linear operator.

In order to prove the controllability of the integro-differential Eq. (5.1), the following conditions are required:

(A4) The real-valued function ξ is piece-wise continuous on [0,T] and there exists a positive constant ϒ such that ϒ=0T|ξ(s)|ds.

(A5) h:J1×nn, J1=i=0m[si,ti+1] is a continuous function and there exist positive constants λ3 and λ4 such that

h(t,u)h(t,v)λ3uv

and h(t,u)λ4, for every u,vn,tJ1.

Theorem 5.1

If all the assumptions (A1)–(A5) are satisfied and the linear system (3.1) is controllable, then the nonlinear integro-differential system (5.1) is totally controllable on[0,T].

Proof

Let us define an iterative scheme for the integro-differential system (5.1) as follows:

(5.2)u0(t)=u0un+1(t)=Φ0(tsi)Ψi1(si,un(ti))+Φ1(tsi)Ψi2(si,un(ti))+sitΦ(ts)Bwn(s)ds+sitΦ(ts)g(s,un(s))ds+sitΦ(ts)0sξ(sη)h(η,un(η))dηds,t[si,ti+1],i=0,1,2,,m,

where

(5.3)wn(t)=BΦ(ti+1t)Msiti+11uf(ti+1)Φ0(ti+1si)Ψi1(si,un(ti))Φ1(ti+1si)Ψi2(si,un(ti))siti+1Φ(ti+1s)g(s,un(s))dssiti+1Φ(ti+1s)0sξ(sη)h(η,un(η))dηds

and n=0,1,2,.

Furthermore, the proof is similar to Theorem 4.1. Therefore, it is omitted.□

6 Application

In this section, we will consider the forced string problem and apply the results obtained in the previous section. The Mittag-Leffler matrix function will be evaluated by using Roberto Garrappa’s MATLAB algorithm.

Example 1

Let us consider the linear fractional order impulsive system without control

(6.1)Dt5/4Cu1(t)=u1(t)+u2(t),t(0,1]32,2,Dt5/4Cu2(t)=2u1(t)+3u2(t),t(0,1]32,2,u(t)=sin(t)3t+2,t1,32,u(t)=cos(t)3,t1,32

with initial conditions u1(0)u2(0)=02 and u1(0)u2(0)=13.

Comparing (6.1) with (1.1), we have u(t)=u1(t)u2(t),Ψ11=sin(t)3t+2,Ψ12=cos(t)3,B=00 and A=1123. Let the final state uf(2)=u1(2)u2(2)=1010 (Figure 1).

Figure 1 The trajectory of the impulsive linear system (6.1) starts from the initial state 02\left[\begin{array}{c}0\\ 2\end{array}\right] and does not reach the final state 10−10\left[\begin{array}{c}10\\ -10\end{array}\right] on [0,2]{[}0,2].
Figure 1

The trajectory of the impulsive linear system (6.1) starts from the initial state 02 and does not reach the final state 1010 on [0,2].

After introducing the control parameter w(t) in (6.1), we have the impulsive linear control system as follows (Figure 2):

(6.2)Dt5/4Cu1(t)=u1(t)+u2(t)+w(t),t(0,1]32,2,Dt5/4Cu2(t)=2u1(t)+3u2(t),t(0,1]32,2,u(t)=sin(t)3t+2,t1,32,u(t)=cos(t)3,t1,32.

Here, B=10. The controllability Gramian matrix for system (6.2) is

M322=0.42270.16550.16550.0796.

It is clear that Gramian matrix M322 is nonsingular. Therefore, system (6.2) is exactly controllable on [0,2]. The control which steers the initial state u(0)=02 of system (6.2) to the arbitrary desired final state uf(2)=1010 during [0,2] is given by

w(t)=BΦ(2t)M3221uf(2)Φ0232Ψ1132,u(1)Φ1232Ψ1232,u(1).
Figure 2 The trajectory of the impulsive linear system (6.2) starts from the initial state 02\left[\begin{array}{c}0\\ 2\end{array}\right] and reaches the final state 10−10\left[\begin{array}{c}10\\ -10\end{array}\right] on [0,2]{[}0,2].
Figure 2

The trajectory of the impulsive linear system (6.2) starts from the initial state 02 and reaches the final state 1010 on [0,2].

Example 2

Consider the nonlinear fractional order impulsive system without control

(6.3)Dt5/4Cu1(t)=u1(t)+u2(t)+cosu1(t),t(0,1]32,2,Dt5/4Cu2(t)=2u1(t)+3u2(t)+sinu2(t),t(0,1]32,2,u(t)=5cos(t)t5,t1,32,u(t)=5sin(t)1,t1,32

with initial conditions u1(0)u2(0)=55 and u1(0)u2(0)=01.

Comparing (6.3) with (1.1), we have u(t)=u1(t)u2(t),Ψ11=5cos(t)t5,Ψ12=5sin(t)1,B=00,g(t,u(t))=cosu1(t)sinu2(t). Let the final state uf(2)=u1(2)u2(2)=05 (Figure 3).

Figure 3 The trajectory of the impulsive linear system (6.3) starts from the initial state 5−5\left[\begin{array}{c}5\\ -5\end{array}\right] and does not reach the final state 05\left[\begin{array}{c}0\\ 5\end{array}\right] on [0,2]{[}0,2].
Figure 3

The trajectory of the impulsive linear system (6.3) starts from the initial state 55 and does not reach the final state 05 on [0,2].

After introducing the control parameter w(t) in (6.1), we have the impulsive nonlinear control system as follows (Figure 4):

(6.4)Dt5/4Cu1(t)=u1(t)+u2(t)+cosu1(t)+w1(t),t(0,1]32,2,Dt5/4Cu2(t)=2u1(t)+3u2(t)+sinu2(t)+w2(t),t(0,1]32,2,u(t)=5cos(t)t5,t1,32,u(t)=5sin(t)1,t1,32.

Here, B=11. The controllability Gramian matrix for system (6.4) is

M322=0.60810.91600.91601.4029.

It is clear that Gramian matrix M322 is nonsingular. Therefore, under assumptions (A1)–(A3), the nonlinear system (6.4) is exactly controllable on [0,2] (Figure 5).

Figure 4 The trajectory of the impulsive nonlinear system (6.4) starts from the initial state 5−5\left[\begin{array}{c}5\\ -5\end{array}\right] and reaches the final state 10−10\left[\begin{array}{c}10\\ -10\end{array}\right] on [0,2]{[}0,2].
Figure 4

The trajectory of the impulsive nonlinear system (6.4) starts from the initial state 55 and reaches the final state 1010 on [0,2].

Figure 5 The trajectory of the nonlinear system (6.4) starts from the initial state 5−5\left[\begin{array}{c}5\\ -5\end{array}\right] and reaches the final states 06\left[\begin{array}{c}0\\ 6\end{array}\right] and −22\left[\begin{array}{c}-2\\ 2\end{array}\right] in the intervals [0,1]\mathrm{[0,1]} and [32,2]\left[\tfrac{3}{2}\left,2\right], respectively.
Figure 5

The trajectory of the nonlinear system (6.4) starts from the initial state 55 and reaches the final states 06 and 22 in the intervals [0,1] and [32,2], respectively.

But, in order to get total controllability of the nonlinear system (6.4), we have the Gramian matrix

(6.5)M=M01,t[0,1],M322,t32,2,

where M01=7.294115.610815.610833.8369 and M322=0.60810.91600.91601.4029. It is clear that both matrices M01 and M322 are nonsingular. Moreover, the nonlinear function g(t,u(t)) and non-instantaneous impulses Ψ11(t,) and Ψ12(t,) satisfy conditions (A1)–(A3). Hence, by Theorem 4.1, the nonlinear impulsive fractional order system (6.4) is totally controllable (Figure 6).

Figure 6 The trajectory of the nonlinear system (6.4) steers the initial state 5−5\left[\begin{array}{c}5\\ -5\end{array}\right] to the final states 06\left[\begin{array}{c}0\\ 6\end{array}\right] and −22\left[\begin{array}{c}-2\\ 2\end{array}\right] in the intervals [0,1]{[}0,1] and [32,2]\left[\tfrac{3}{2}\left,2\right], respectively.
Figure 6

The trajectory of the nonlinear system (6.4) steers the initial state 55 to the final states 06 and 22 in the intervals [0,1] and [32,2], respectively.

7 Conclusion

In this manuscript, the total controllability of fractional order nonlinear differential equations with non-instantaneous impulses is investigated through the iterative scheme. The total controllability conditions for the nonlinear system are examined by imposing that the linear system is controllable and the nonlinear function satisfies some suitable assumptions. The computation of controlled state and steering control for the linear and nonlinear fractional order impulsive system is proposed by using the Mittag-Leffler matrix function and the Gramian matrix. In the future, for a better understanding of controllability, the computational scheme can be applied to the nonlinear control problems in finite dimensional spaces.

Acknowledgments

Ramesh K. Vats is grateful to Council of Scientific and Industrial Research (CSIR), Government of India, for providing financial assistance under research project no. 25(0268)/17/EMR-II.

References

[1] Q. Yang and D. Chen, Fractional calculus in image processing: a review, Fract. Calc. Appl. Anal. 19 (2016), no. 5, 1222–1249.10.1515/fca-2016-0063Search in Google Scholar

[2] R. Magin, M. D. Ortigueira, I. Podlubny, and J. J. Trujillo, On the fractional signals and systems, Signal Proc. 91 (2011), 350–371.10.1016/j.sigpro.2010.08.003Search in Google Scholar

[3] R. L. Magin, Fractional Calculus in Bioengineering, Begell House Redding, 2006.Search in Google Scholar

[4] R. L. Bagley and P. J. Torvik, On the fractional calculus model of viscoelastic behavior, J. Rheol. 30 (1986), 133–155.10.1122/1.549887Search in Google Scholar

[5] T. Wenchang, P. Wenxiao, and X. Mingyu, A note on unsteady flows of a viscoelastic fluid with the fractional Maxwell model between two parallel plates, Int. J. Nonlinear Mech. 38 (2003), 645–650.10.1016/S0020-7462(01)00121-4Search in Google Scholar

[6] K. Balachandran and V. Govindaraj, Numerical controllability of fractional dynamical systems, Optimization 63 (2014), 1267–1279.10.1080/02331934.2014.906416Search in Google Scholar

[7] I. Podlubny, Fractional Differential Equations, Academic Press, New York, 1999.Search in Google Scholar

[8] A. A. Kilbas, H. M. Srivastava, and J. J. Trujillo, Theory and Applications of Fractional Differential Equations, Elsevier, Amsterdam, 2006.Search in Google Scholar

[9] K. S. Miller and B. Ross, An Introduction to the Fractional Calculus and Fractional Differential Equations, Wiley, New York, 1993.Search in Google Scholar

[10] M. Zayernouri and G. E. Karniadakis, Fractional special collocation method, SIAM J. Sci. Comput. 36 (2014), no. 1, A40–A62.10.1137/130933216Search in Google Scholar

[11] M. Zayernouri and G. E. Karniadakis, Fractional Sturm-Liouville eigen-problems: theory and numerical approximation, J. Comput. Phys. 252 (2013), 495–517.10.1016/j.jcp.2013.06.031Search in Google Scholar

[12] M. Muslim, A. Kumar, and M. Fečkan, Existence, uniqueness and stability of solutions to second order nonlinear differential equations with non-instantaneous impulses, J. King Saud Univ. – Sci. 30 (2016), no. 2, 204–213.10.1016/j.jksus.2016.11.005Search in Google Scholar

[13] S. Hristova and R. Terzieva, Lipschitz stability of differential equations with non-instantaneous impulses, Adv. Difference Equ. 2016 (2016), 322, 10.1186/s13662-016-1045-6.Search in Google Scholar

[14] R. Agarwal, S. Hristova, and D. Ó Regan, Practical stability of differential equations with non-instantaneous impulses, Differ. Equ. Appl. 9 (2017), no. 4, 413–432.10.7153/dea-2017-09-29Search in Google Scholar

[15] R. Agarwal, D. Ó Regan, and S. Hristova, Stability by Lyapunov like functions of nonlinear differential equations with non-instantaneous impulses, J. Appl. Math. Comput. 53 (2017), no. 1–2, 147–168.10.1007/s12190-015-0961-zSearch in Google Scholar

[16] R. Agarwal, S. Hristova, and D. Ó Regan, Non-Instantaneous Impulses in Differential Equations, Springer, New York City, 2017.10.1007/978-3-319-66384-5Search in Google Scholar

[17] E. Hernández and D. Ó Regan, On a new class of abstract impulsive differential equations, Proc. Amer. Math. Soc. 141 (2013), 1641–1649.10.1090/S0002-9939-2012-11613-2Search in Google Scholar

[18] J. Wang and M. Fečkan, A general class of impulsive evolution equations, Topol. Methods Nonlinear Anal. 46 (2015), 915–933.10.12775/TMNA.2015.072Search in Google Scholar

[19] J. Wang, A. G. Ibrahim, M. Fečkan, and Y. Zhou, Controllability of fractional non-instantaneous impulsive differential inclusions without compactness, IMA J. Math. Control Inform. 36 (2017), 443–460.10.1093/imamci/dnx055Search in Google Scholar

[20] M. Muslim and A. Kumar, Controllability of fractional differential equation of order α ∈ [1,2] with non-instantaneous impulses, Asian J. Control 20 (2018), no. 2, 935–942.10.1002/asjc.1604Search in Google Scholar

[21] A. Kumar, M. Muslim, and R. Sakthivel, Controllability of the second-order nonlinear differential equations with non-instantaneous impulses, J. Dyn. Control Syst. 24 (2018), 325–342.10.1007/s10883-017-9376-5Search in Google Scholar

[22] D. Chalishajar and A. Kumar, Total controllability of the second order semi-linear differential equation with infinite delay and non-instantaneous impulses, Math. Comput. Appl. 23 (2018), no. 3, 32, 10.3390/mca23030032.Search in Google Scholar

[23] T. Kaczorek, Selected Problems of Fractional Systems Theory, Springer-Verlag, Berlin, 2011.10.1007/978-3-642-20502-6Search in Google Scholar

[24] C. A. Monje, Y. Q. Chen, B. M. Vinagre, X. Xue, and V. Feliu, Fractional-Order Systems and Controls: Fundamentals and Applications, Springer, London, 2010.10.1007/978-1-84996-335-0Search in Google Scholar

[25] K. Balachandran, V. Govindaraj, L. Rodriguez-Germa, and J. J. Trujillo, Controllability of nonlinear higher order fractional dynamical systems, Nonlinear Dyn. 71 (2013), 605–612.10.1007/s11071-012-0612-ySearch in Google Scholar

Received: 2020-02-11
Revised: 2020-05-25
Accepted: 2020-06-16
Published Online: 2020-09-19

© 2020 Avadhesh Kumar et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Regular Articles
  2. Stability of an additive-quadratic-quartic functional equation
  3. Two new forms of ordered soft separation axioms
  4. Coefficient inequalities for a subclass of Bazilevič functions
  5. Equivalence of the existence of best proximity points and best proximity pairs for cyclic and noncyclic nonexpansive mappings
  6. Generalized parabolic Marcinkiewicz integrals associated with polynomial compound curves with rough kernels
  7. Jordan centralizer maps on trivial extension algebras
  8. On soft pc-separation axioms
  9. Direct and strong converse inequalities for approximation with Fejér means
  10. On perturbed quadratic integral equations and initial value problem with nonlocal conditions in Orlicz spaces
  11. On θ-generalized demimetric mappings and monotone operators in Hadamard spaces
  12. On the domain of implicit functions in a projective limit setting without additional norm estimates
  13. Scalar linear impulsive Riemann-Liouville fractional differential equations with constant delay-explicit solutions and finite time stability
  14. The special atom space and Haar wavelets in higher dimensions
  15. A strong convergence theorem for a zero of the sum of a finite family of maximally monotone mappings
  16. Existence and uniqueness of mild solutions for a fractional differential equation under Sturm-Liouville boundary conditions when the data function is of Lipschitzian type
  17. The new investigation of the stability of mixed type additive-quartic functional equations in non-Archimedean spaces
  18. Numerical approach to the controllability of fractional order impulsive differential equations
  19. Two modifications of the inertial Tseng extragradient method with self-adaptive step size for solving monotone variational inequality problems
  20. Further results on Ulam stability for a system of first-order nonsingular delay differential equations
  21. Existence results for nonlinear coupled system of integral equations of Urysohn Volterra-Chandrasekhar mixed type
  22. Structure of n-quasi left m-invertible and related classes of operators
  23. Hyers-Ulam stability of a nonautonomous semilinear equation with fractional diffusion
  24. Hasimoto surfaces for two classes of curve evolution in Minkowski 3-space
  25. Applications of some operators on supra topological spaces
  26. An iterative algorithm for the system of split mixed equilibrium problem
  27. Almost graded multiplication and almost graded comultiplication modules
  28. Strong convergence of an inertial extrapolation method for a split system of minimization problems
  29. On the non-hypercyclicity of scalar type spectral operators and collections of their exponentials
  30. Exponential spline method for singularly perturbed third-order boundary value problems
  31. Existence results of noninstantaneous impulsive fractional integro-differential equation
  32. Review Articles
  33. On a characterization of exponential, Pearson and Pareto distributions via covariance and pseudo-covariance
Downloaded on 13.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/dema-2020-0015/html
Scroll to top button