TY - JOUR AU - Nieto, Juan, J AB - Abstract This paper is concerned with a kind of first-order singular differential system with impulses. Based on the Schaefer fixed-point theorem, some new verifiable algebraic criteria are given to ensure the controllability of bounded solutions for the considered system. The results obtained in this paper not only achieve the controllability of the singular differential system with impulses for the first time, but also complement the previous researches on singular differential system with impulses. Consequently, the results established are essentially new. Finally, the effectiveness of the obtained results are illustrated via a numerical example. 1. Introduction During the past years, impulsive differential equations have been studied by many authors. Some classical tools have been widely used to get the solutions of impulsive differential equations, such as fixed point theorems in cones, topological degree theory (including continuation method and coincidence degree theory), the method of lower and upper solutions and the critical point theory. For the theory and more related classical results, we assemble here some relevant results, such as Alzabut (2008), Georgieva et al. (2012), Liu (2007), Luo & Jing (2008), Saker & Alzabut (2007), Stamov & Alzabut (2011), Stamov et al. (2012) and Yang & Shen (2007) and monographs Benchohra et al. (2006), Lakshmikantham et al. (1989), Samoilenko & Perestyuk (1995) and Zavalishchin & Sesekin (1997). For example, Yang & Shen (2007) studied nonlinear boundary value problems for a class of first-order functional differential equations as follows $$\begin{align*} \begin{cases} x^{\prime}(t)=f(t,x(t),x(\theta(t))),& t\in J=[0,T], t\neq t_k,\\ \varDelta x(t_k)=I_k(x(t_k)),&k=1,2,...,m,\\ x(0)=px(T), x(t)=(0),&t\in[-r,0], \end{cases} \end{align*}$$ where |$f\in C(J\times{\mathbb{R}}^2,{\mathbb{R}})$|⁠, |$\theta (t)\in C(J,J^+)$|⁠, |$J^+=[-r,T]$|⁠, |$r>0$|⁠, |$0 < t_1 < t_2 < \cdot \cdot \cdot < t_m < T$|⁠, |$I_k\in C({\mathbb{R}},{\mathbb{R}})$|⁠, |$\varDelta x(t_k)=x(t^+_k)-x(t^-_k)$|⁠, |$k=1,2,...,m$|⁠. By using the lower and upper method and monotone iterative technique, the authors established new existence results of extreme solutions. Singular equations appear in a great deal of physical models and differential equations with singularities arise naturally in the study of the motion of particles under the influence of gravitational or electrostatic forces. Singular equation means that the equation becomes infinite at some value of the state variable. For example, in Mawhin (1993), the singularity models in which the restoring force caused by a compressed perfect gas, and in Pishkenari et al. (2008), Rützel et al. (2003) and Yang et al. (2005), the singular term can be regarded as a generalized Lennard-Jones potential or Van der Waals force and it is widely found in molecular dynamics to model the interaction between atomic particles. Compared with the classical first-order singular differential equations or impulsive differential equations, the study on first-order singular differential equations with impulses has been less considered in the literature, see Agarwal et al. (2005), Chu & Nieto (2007), Kong (2017), Kong & Luo (2018), Nieto & Uzal (2018a)and Nieto & Uzal (2018b). For example, in Agarwal et al. (2005), based on Schauder’s fixed point theorem, Agarwal et al. established the existence of at least one positive solution for the following first singular boundary value problems with impulses of the form: $$\begin{align*} \begin{cases} y^{\prime}=f(t,y), & \textrm{a.e. on}\ t\in J^{\prime} = J \setminus \{t_1,t_2,...,t_p\},\\ \varDelta y(t_k) = I_k(y(t_k)), &k = 1, 2,...,p,\\ y(0) = 0, \end{cases} \end{align*}$$ where |$J = [0,T]$|⁠, |$I_k: {\mathbb{R}} \rightarrow{\mathbb{R}}$| is continuous for each |$k =1, 2,...,p$| and the nonlinearity |$f$| may be singular in the independent variable and may also be singular at |$y = 0$|⁠. On the other hand, control theory is an area of application-oriented mathematics, which deals with basic principles underlying the analysis and design of control systems. It is well known that the study of controllability plays an important role in the control theory and engineering (Aeyels, 1984; Rugh, 1993; Wang, 1999). The concept of controllability leads to some important conclusions regarding the behaviour of linear and nonlinear dynamical systems. Most of the practical systems are nonlinear in nature, and hence the study of nonlinear systems is important. For the basic theory on evolution system, the reader is referred to Tanabe’s book (Tanabe, 1979). As one of the most fundamental concepts in modern control theory, controllability has been studied by many scholars in the past several decades. For linear control systems, their controllability was completely solved in 1960s, which became a classical result. But for nonlinear control systems, although their controllability has been investigated extensively since 1970, and considerable significant progress has been obtained by introducing some powerful methods, including the well-known differential geometric method (Isidori, 1995), and the fixed point methods (Balachandran & Dauer, 1987). In the recent years, the study of impulsive control systems has received an increasing interest. There have been some researches undertaken dealing with the fundamental issues such as controllability for impulsive systems. See, to name a few, Arthi & Balachandran (2012), Arthi & Balachandran (2014), Fu (2003), Liu (2005), Radhakrishnan & Balachandran (2011) and Sakthivel et al. (2009). However, to the best of our knowledge, little attention has been devoted to the study of controllability of the singular differential systems with impulses. This may be due to the fact that the singular forces make the controllability of the impulsive differential systems more difficult and complex, and the mechanism on which how the controllability is influenced by the singularities and impulses associated with the first-order differential systems is far away from clarity. Therefore, at this stage, it is crucial and necessary to further research the dynamical relationship between the two models and fill this gap partially. Motivated by the above works and discussions, in this paper, we study the controllability for the following first-order singular differential equation with impulses of the form: $$\begin{equation} \begin{cases} x^{\prime}(t)= -a(t)x(t)+ f(t,x_t)+B(t)u(t), t\in J=[0,b], t\neq t_k, \\ \varDelta x(t_k)= x(t^+_k) -x(t^-_k)= I_k(x(t^-_k)), k = 1,2,...,m, \\ x(t)=\phi(t), t\in [-\tau,0], \end{cases} \end{equation}$$ (1.1) where the state variable |$x(\cdot )$| takes values in Banach space |$X$| with the norm |$|\cdot |$|⁠. The control function |$u(\cdot )$| is given in |$L^2(J,U)$|⁠, a Banach space of admissible control functions with |$U$| as a Banach space, |$a$| is the infinitesimal generator of a strongly continuous semigroup of bounded linear operators |$T(t)$| in |$X$|⁠, |$B$| is a bounded linear functions from |$U$| into |$X$|⁠, |$f: J\times D \rightarrow X$| can be singular at |$x=0$|⁠, |$D = \big \{\varphi : [-\tau ,0] \rightarrow X, \varphi (t)\textrm{is continuous everywhere except a finite number} \textrm{of points} \widetilde{t} \textrm{at which} \varphi (\widetilde{t}^-), \varphi (\widetilde{t}^+)\ \textrm{exist and}\ \varphi (\widetilde{t}^-)=\varphi (\widetilde{t})\big \}$|⁠, |$I_k: X \rightarrow X $| is continuous, and |$\varDelta x(t_k) =x(t^+_k)-x(t^-_k)$|⁠, for all |$k = 1,2,...,m$|⁠, |$0 = t_0 < t_1 < t_2 <\cdot \cdot \cdot < t_m k, t\in J\ a.e. \end{gather*}$$ (H6) There exist integrable functions |$h: J \rightarrow [0,\infty )$| such that |$ |f(t,\phi )|\leq h(t)\psi (\|\phi \|_{PC}), $| for almost all |$t \in J$|⁠, |$\phi \in PC([-\tau ,0],X)$|⁠, where |$\psi : (0,\infty ) \rightarrow (0,\infty )$| is a continuous nondecreasing function with $$\begin{align*} M_1\int_{0}^{b}h(s)\textrm{d}s< \int_{c}^{\infty} \frac{\textrm{d}s}{ \psi(s)}, \end{align*}$$ where |$c= M_1\|\phi \|_{PC}+ M_1Nb +\sum ^m_{k=1}M_1 \eta _k$|⁠, and $$\begin{align*} N:= B^+M_2 \left[ |\hat{x} |+M_1\|\phi \|_{PC} +M_1\int_{0}^{b}h(s)\psi(\|x_s\|_{PC})\textrm{d}s +\sum^m_{k=1}M_1\eta_k \right]. \end{align*}$$ Define $$\begin{equation} \begin{aligned} &\eta_1= \frac{\beta^-_{\eta_1} -B^+ M_2\Big( |\hat{x}|+M_1\|\phi\|_{PC} +M_1 b \alpha^+_{\eta_2} +\sum^m_{k=1}M_1 \eta_k\Big)}{a^+}, \\ &\eta_2=\frac{\alpha^+_{\eta_2} +B^+ M_2\Big( |\hat{x}|+M_1\|\phi\|_{PC} +M_1 b \alpha^+_{\eta_2} +\sum^m_{k=1}M_1 \eta_k\Big)}{a^-}, \end{aligned} \end{equation}$$ (2.1) where |$\alpha ^+_{\eta _2}=\sup _{t\in J}|\alpha _{\eta _2}(t)|$| and |$\beta ^-_{\eta _1}=\inf _{t\in J}|\beta _{\eta _1}(t)|$|⁠. Lemma 2.3 Assume that |$0<\eta _1< \eta _2$|⁠. Then the solution |$x(t)$| of (1.1) satisfies $$\begin{align*} 0<\eta_1 \leq x(t) \leq \eta_2,\ \textrm{for all}\ t\in [0, b]. \end{align*}$$ Proof. Let |$[0,b_1) \subseteq [0,b]$| be an interval such that |$x(t)>0$| for all |$t\in [0,b_1)$|⁠. Firstly, we prove $$\begin{equation} 00, t\in [0, b_1). \end{equation}$$ (2.3) Otherwise, there exists |$\underline{t}\in [0,b_1)$| such that |$x(\underline{t})=\eta _1$| and |$x(t)\geq \eta _1$| for all |$t\in [0, \underline{t})$|⁠. Then from (2.2), we can see that |$\eta _1< x(t)< \eta _2$| for all |$t\in [0, \underline{t})$|⁠. Calculating the derivative of |$x(t)$|⁠, by (H5) and (2.1), we have that $$\begin{align*} 0\geq&\ x^{\prime}(\underline{t}) = -a(\underline{t})x(\underline{t}) +f(\underline{t},x_{\underline{t}}) +B(\underline{t})u(\underline{t})\\ =&\ -a(\underline{t})x(\underline{t}) +f(\underline{t},x_{\underline{t}})\\ &+B(\underline{t}) \cdot W^{-1}\left[ \hat{x}-T(b)\phi(0)-\int_{0}^{b}T(b-s)f(s,x_s)\textrm{d}s-\sum^m_{k=1}T(b-t_k)I_k(x(t^-_k)) \right](\underline{t}) \\> & -a^+ \eta_1 + \beta_{\eta_1}(\underline{t}) -B^+ M_2\left( |\hat{x}|+M_1\|\phi\|_{PC} +M_1 b \alpha^+_{\eta_2} +\sum^m_{k=1}M_1 \eta_k\right)\\ > & -a^+ \eta_1 + \beta^-_{\eta_1} -B^+ M_2\left( |\hat{x}|+M_1\|\phi\|_{PC} +M_1 b \alpha^+_{\eta_2} +\sum^m_{k=1}M_1 \eta_k\right)\\ =& -a^+ \cdot \frac{\beta^-_{\eta_1} -B^+ M_2\left( |\hat{x}|+M_1\|\phi\|_{PC} +M_1 b \alpha^+_{\eta_2} +\sum^m_{k=1}M_1 \eta_k\right)}{a^+}\\ &+ \beta^-_{\eta_1} -B^+ M_2\left( |\hat{x}|+M_1\|\phi\|_{PC} +M_1 b \alpha^+_{\eta_2} +\sum^m_{k=1}M_1 \eta_k\right) =0, \end{align*}$$ which also leads to a contradiction. Then, (2.3) is satisfied. Therefore, the proof is completed. Now, we are ready to present our main result. Theorem 2.4 If the conditions (H1)–(H6) are satisfied, then (1.1) is controllable on |$J$|⁠. 3. Proof of Theorem 2.4 In this section, we divide the proof into the following two lemmas. Lemma 3.1 If the conditions (H1)–(H6) are satisfied, then the mild solution of (1.1) is bounded, i.e., there exist positive constants |$0 0$|⁠, the compactness implies that the continuity in the uniform operator topology. Let |$\sigma _2-\sigma _1\rightarrow 0$|⁠, then we can see that $$\begin{align*} \big|(Fy )(\sigma_1)-(Fy )(\sigma_2)\big|\rightarrow 0. \end{align*}$$ Thus, |$F$| maps |$B_k$| into an equicontinuous family of functions. For the case |$\sigma _1 < \sigma _2 < 0$| or |$\sigma _1 < 0 < \sigma _2$|⁠, we can hand them similarly. Step 2. We show |$\overline{FB_k}$| is compact. Since we have shown |$FB_k$| is equicontinuous collection, it suffices by the Arzela–Ascoli theorem to show that |$F$| maps |$B_k$| into a precompact set in |$X$|⁠. Let |$0 < t \leq b$| be fixed and |$\epsilon $| a real number satisfying |$0 <\epsilon < t$|⁠. For |$y \in B_k$|⁠, we define $$\begin{align*} (F_\epsilon y )(t)=& \int_{0}^{t-\epsilon}T(t-\theta) B(t)W^{-1}\Bigg[\hat{x} -T(b)\phi (0) -\int_{0}^{b}T(b-s)f(s,y_s+\widehat{\phi}_s)\textrm{d}s\\ &{\hskip102pt}-\sum^m_{k=1}T(b-t_k)I_k\big(y (t^-_k)+\widehat{\phi} (t^-_k)\big) \Bigg](\theta)\textrm{d}\theta\\ & +\int_{0}^{t-\epsilon} T(t-s) f(s,y_s+\widehat{\phi}_s)\textrm{d}s+\sum_{0