Abstract

We establish the existence of solutions to systems of second-order dynamic equations on time scales with the right member 𝑓, a Δ-Carathéodory function. First, we consider the case where the nonlinearity 𝑓 does not depend on the Δ-derivative, 𝑥Δ(𝑡). We obtain existence results for Strum-Liouville and for periodic boundary conditions. Finally, we consider more general systems in which the nonlinearity 𝑓 depends on the Δ-derivative and satisfies a linear growth condition with respect to 𝑥Δ(𝑡). Our existence results rely on notions of solution-tube that are introduced in this paper.

1. Introduction

In this paper, we establish existence results for the following systems of second-order dynamic equations on time scales:𝑥ΔΔ(𝑡)=𝑓(𝑡,𝑥(𝜎(𝑡))),Δ-a.e.𝑡𝕋𝜅20,𝑥(BC),(1.1)𝑥ΔΔ(𝑡)=𝑓𝑡,𝑥(𝜎(𝑡)),𝑥Δ(𝑡),Δ-a.e.𝑡𝕋𝜅20,𝑎0𝑥(𝑎)𝑥Δ(𝑎)=𝑥0,𝑎1𝑥(𝑏)+𝛾1𝑥Δ(𝜌(𝑏))=𝑥1.(1.2) Here, 𝕋 is a compact time scale where 𝑎=min𝕋, 𝑏=max𝕋, and 𝕋𝜅20 is defined in (2.4). The map 𝑓𝕋𝜅0×𝑛𝑛 is Δ-Carathéodory (see Definition 2.9), and (BC) denotes one of the following boundary conditions:𝑥𝑥(𝑎)=𝑥(𝑏),Δ(𝑎)=𝑥Δ(𝜌(𝑏)),(1.3)𝑎0𝑥(𝑎)𝛾0𝑥Δ(𝑎)=𝑥0,𝑎1𝑥(𝑏)+𝛾1𝑥Δ(𝜌(𝑏))=𝑥1,(1.4) where 𝑎0,𝑎1,𝛾0,𝛾10, max{𝑎0,𝛾0}>0, and max{𝑎1,𝛾1}>0.

Problem (1.1) was mainly treated in the case where it has only one equation (𝑛=1) and 𝑓 is continuous. In particular, the existence of a solution of (1.1) was established by Akın [1] for the Dirichlet boundary condition and by Stehlík [2] for the periodic boundary condition. Equation (1.1) with nonlinear boundary conditions was studied by Peterson et al. [3]. In all those results, the method of lower and upper solutions was used. See also [4, 5] and the references therein for other results on the problem (1.1) when 𝑛=1.

Very few existence results were obtained for the system (1.1) when 𝑛>1. Recently, Henderson et al. [6] and Amster et al. [7] established the existence of solutions of (1.1) with Sturm-Liouville and nonlinear boundary conditions, respectively, assuming that 𝑓 is a continuous function satisfying the following condition:𝑅>0suchthat𝑥,𝑓(𝑡,𝑥)>0if𝑥=𝑅.(1.5)

The fact that the right member in the system (1.2) depends also on the Δ-derivative, 𝑥Δ, increases considerably the difficulty of this problem. So, it is not surprising that there are almost no results for this problem in the literature. Atici et al. [8] studied this problem in the particular case, where there is only one equation (𝑛=1) and 𝑓 is positive, continuous and satisfies a monotonicity condition. Assuming a growth condition of Wintner type and using the method of lower and upper solutions, they obtained the existence of a solution.

The system (1.2) with the Dirichlet boundary condition was studied by Henderson and Tisdell [9] in the general case where 𝑛>1. They considered 𝑓 a continuous function and 𝕋 a regular time scale (i.e., 𝜌(𝑡)<𝑡<𝜎(𝑡) or 𝕋=[𝑎,𝑏]). They established the existence of a solution of (1.2) under the following assumptions: (A1) there exists 𝑅>𝑚𝑎𝑥{𝑥0,𝑥1} such that 2𝑥,𝑓(𝑡,𝑥,𝑦)+𝑦2>0 if 𝑥=𝑅, 2𝑥,𝑦𝜇(𝑡)𝑦2,(A2) there exist 𝑐,𝑑0 such that 𝑑(𝜌(𝑏)𝑎)<1 and 𝑓(𝑡,𝑥,𝑦)𝑐+𝑑𝑦 if 𝑥𝑅.

In the third section of this paper, we establish an existence theorem for the system (1.1). To this aim, we introduce a notion of solution-tube of (1.1) which generalizes to systems the notions of lower and upper solutions introduced in [1, 2]. This notion generalizes also condition (1.5) used by Henderson et al. [6] and Amster et al. [7]. Our notion of solution-tube is in the spirit of the notion of solution-tube for systems of second-order differential equations introduced in [10]. Our notion is new even in the case of systems of second-order difference equations.

In the last section of this paper, we study the system (1.2). Again, we introduce a notion of solution-tube of (1.2) which generalizes the notion of lower and upper solutions used by Atici et al. [8]. This notion generalizes also condition (1.5) and the notion of solution-tube of systems of second-order differential equations introduced in [10]. In addition, we assume that 𝑓 satisfies a linear growth condition. It is worthwhile to mention that the time scale 𝕋 does not need to be regular, and we do not require the restriction 𝑑(𝜌(𝑏)𝑎)<1 as in assumption (A2) used in [9].

Moreover, we point out that the right members of our systems are not necessarily continuous. Indeed, we assume that the weaker condition: 𝑓 is a Δ-Carathéodory function. This condition is interesting in the case where the points of 𝕋 are not all right scattered. We obtain the existence of solutions to (1.1) and to (1.2) in the Sobolev space 𝑊Δ2,1(𝕋,𝑛). To our knowledge, it is the first paper applying the theory of Sobolev spaces with topological methods to obtain solutions to (1.1) and (1.2). Solutions of second-order Hamiltonian systems on time scales were obtained in a Sobolev space via variational methods in [11]. Finally, let us mention that our results are new also in the continuous case and for systems of second-order difference equations.

2. Preliminaries and Notations

For sake of completeness, we recall some notations, definitions, and results concerning functions defined on time scales. The interested reader may consult [12, 13] and the references therein to find the proofs and to get a complete introduction to this subject.

Let 𝕋 be a compact time scale with 𝑎=min𝕋<𝑏=max𝕋. The forward jump operator 𝜎𝕋𝕋 (resp., the backward jump operator 𝜌𝕋𝕋) is defined by 𝜎(𝑡)=inf{𝑠𝕋𝑠>𝑡}if𝑏𝑡<𝑏,if𝑡=𝑏,resp.,𝜌(𝑡)=sup{𝑠𝕋𝑠<𝑡}if𝑎𝑡>𝑎,if.𝑡=𝑎(2.1) We say that 𝑡<𝑏 is right scattered (resp., 𝑡>𝑎 is left scattered) if 𝜎(𝑡)>𝑡 (resp., 𝜌(𝑡)<𝑡) otherwise, we say that 𝑡 is right dense (resp., left dense). The set of right-scattered points of 𝕋 is at most countable, see [14]. We denote it by 𝑅𝕋𝑡={𝑡𝕋𝑡<𝜎(𝑡)}=𝑖,𝑖𝐼(2.2) for some 𝐼. The graininess function 𝜇𝕋[0,) is defined by 𝜇(𝑡)=𝜎(𝑡)𝑡. We denote 𝕋𝜅]=𝕋(𝜌(𝑏),𝑏,𝕋0=𝕋{𝑏}.(2.3) So, 𝕋𝜅=𝕋 if 𝑏 is left dense, otherwise 𝕋𝜅=𝕋0. Since 𝕋𝜅 is also a time scale, we denote𝕋𝜅2=(𝕋𝜅)𝜅,𝕋𝜅20=𝕋𝜅2{𝑏}.(2.4)

In 1990, Hilger [15] introduced the concept of dynamic equations on time scales. This concept provides a unified approach to continuous and discrete calculus with the introduction of the notion of delta-derivative 𝑥Δ(𝑡). This notion coincides with 𝑥(𝑡) (resp., Δ𝑥(𝑡)) in the case where the time scale 𝕋 is an interval (resp., the discrete set {0,1,,𝑁}).

Definition 2.1. A map 𝑓𝕋𝑛 is Δ-differentiable at 𝑡𝕋𝜅 if there exists 𝑓Δ(𝑡)𝑛 (called the Δ-derivative of 𝑓 at 𝑡) such that for all 𝜀>0, there exists a neighborhood 𝑈 of 𝑡 such that 𝑓(𝜎(𝑡))𝑓(𝑠)𝑓Δ||||(𝑡)(𝜎(𝑡)𝑠)𝜀𝜎(𝑡)𝑠𝑠𝑈.(2.5) We say that 𝑓 is Δ-differentiable if 𝑓Δ(𝑡) exists for every 𝑡𝕋𝜅.
If 𝑓 is Δ-differentiable and if 𝑓Δ is Δ-differentiable at 𝑡𝕋𝜅2, we call 𝑓ΔΔ(𝑡)=(𝑓Δ)Δ(𝑡) the second Δ-derivative of 𝑓 at 𝑡.

Proposition 2.2. Let 𝑓𝕋𝑛 and 𝑡𝕋𝜅. (i)If 𝑓 is Δ-differentiable at 𝑡, then 𝑓 is continuous at 𝑡.(ii)If 𝑓 is continuous at 𝑡𝑅𝕋, then 𝑓Δ(𝑡)=𝑓(𝜎(𝑡))𝑓(𝑡).𝜇(𝑡)(2.6)(iii)The map 𝑓 is Δ-differentiable at 𝑡𝕋𝜅𝑅𝕋 if and only if 𝑓Δ(𝑡)=lim𝑠𝑡𝑓(𝑡)𝑓(𝑠).𝑡𝑠(2.7)

Proposition 2.3. If 𝑓𝕋𝑛 and 𝑔𝕋𝑚 are Δ-differentiable at 𝑡𝕋𝜅, then (i)if 𝑛=𝑚, (𝛼𝑓+𝑔)Δ(𝑡)=𝛼𝑓Δ(𝑡)+𝑔Δ(𝑡) for every 𝛼,(ii)if 𝑚=1, (𝑓𝑔)Δ(𝑡)=𝑔(𝑡)𝑓Δ(𝑡)+𝑓(𝜎(𝑡))𝑔Δ(𝑡)=𝑓(𝑡)𝑔Δ(𝑡)+𝑔(𝜎(𝑡))𝑓Δ(𝑡),(iii)if 𝑚=1 and 𝑔(𝑡)𝑔(𝜎(𝑡))0, then 𝑓𝑔Δ(𝑡)=𝑔(𝑡)𝑓Δ(𝑡)𝑓(𝑡)𝑔Δ(𝑡),𝑔(𝑡)𝑔(𝜎(𝑡))(2.8)(iv)if 𝑊𝑛 is open and 𝑊 is differentiable at 𝑓(𝑡)𝑊 and 𝑡𝑅𝕋, then (𝑓)Δ(𝑡)=(𝑓(𝑡)),𝑓Δ(𝑡).

We denote 𝐶(𝕋,𝑛) the space of continuous maps on 𝕋, and we denote 𝐶1(𝕋,𝑛) the space of continuous maps on 𝕋 with continuous Δ-derivative on 𝕋𝜅. With the norm 𝑥0=max{𝑥(𝑡)𝑡𝕋} (resp., 𝑥1=max{𝑥0,max{𝑥Δ(𝑡)𝑡𝕋𝜅}}), 𝐶(𝕋,𝑛) (resp., 𝐶1(𝕋,𝑛)) is a Banach space.

We study the second Δ-derivative of the norm of a map.

Lemma 2.4. Let 𝑥𝕋𝑛 be Δ-differentiable. (1)On {𝑡𝕋𝜅2𝑥(𝜎(𝑡))>0 and 𝑥ΔΔ(𝑡) exists}, 𝑥(𝑡)ΔΔ𝑥(𝜎(𝑡)),𝑥ΔΔ(𝑡).𝑥(𝜎(𝑡))(2.9)(2) On {𝑡𝕋𝜅2𝑅𝕋𝑥(𝜎(𝑡))>0 and 𝑥ΔΔ(𝑡) exists}, 𝑥(𝑡)ΔΔ=𝑥(𝑡),𝑥ΔΔ(𝑡)+𝑥Δ(𝑡)2𝑥(𝑡)𝑥(𝑡),𝑥Δ(𝑡)2𝑥(𝑡)3.(2.10)

Proof. Denote 𝐴={𝑡𝕋𝜅2𝑥(𝜎(𝑡))>0 and 𝑥ΔΔ(𝑡) exists}. By Proposition 2.3, on the set 𝐴𝑅𝕋, we have 𝑥(𝑡)Δ=𝑥(𝑡),𝑥Δ(𝑡),𝑥(𝑡)𝑥(𝑡)ΔΔ=𝑥(𝑡),𝑥ΔΔ(𝑡)+𝑥Δ(𝑡)2𝑥(𝑡)𝑥(𝑡),𝑥Δ(𝑡)2𝑥(𝑡)3𝑥(𝜎(𝑡)),𝑥ΔΔ(𝑡).𝑥(𝜎(𝑡))(2.11) If 𝑡𝐴 is such that 𝑡<𝜎(𝑡)=𝜎2(𝑡), then by Propositions 2.2 and 2.3, we have 𝑥(𝑡)ΔΔ=𝑥(𝜎(𝑡))Δ𝑥(𝑡)Δ=𝜇(𝑡)𝑥(𝜎(𝑡)),𝑥Δ(𝜎(𝑡))𝜇(𝑡)𝑥(𝜎(𝑡))𝑥(𝜎(𝑡))𝑥(𝑡)𝜇(𝑡)2=𝑥(𝜎(𝑡)),𝑥Δ(𝑡)+𝜇(𝑡)𝑥ΔΔ(𝑡)𝜇(𝑡)𝑥(𝜎(𝑡))𝑥(𝜎(𝑡)),𝑥(𝑡)+𝜇(𝑡)𝑥Δ(𝑡)𝜇(𝑡)2+𝑥(𝜎(𝑡))𝑥(𝑡)𝜇(𝑡)2=𝑥(𝜎(𝑡)),𝑥ΔΔ(𝑡)𝑥(𝜎(𝑡))𝑥(𝜎(𝑡)),𝑥(𝑡)𝜇(𝑡)2+𝑥(𝜎(𝑡))𝑥(𝑡)𝜇(𝑡)2𝑥(𝜎(𝑡)),𝑥ΔΔ(𝑡).𝑥(𝜎(𝑡))(2.12) If 𝑡𝐴 is such that 𝑡<𝜎(𝑡)<𝜎2(𝑡), then 𝑥(𝑡)ΔΔ=𝑥(𝜎(𝑡))Δ𝑥(𝑡)Δ=𝜎𝜇(𝑡)𝑥2(𝑡)𝑥(𝜎(𝑡))𝜇(𝜎(𝑡))𝜇(𝑡)𝑥(𝜎(𝑡))𝑥(𝑡)𝜇(𝑡)2𝜎𝑥(𝜎(𝑡)),𝑥2(𝑡)𝑥(𝜎(𝑡))𝜇(𝜎(𝑡))𝜇(𝑡)𝑥(𝜎(𝑡))𝑥(𝜎(𝑡))𝑥(𝑡)𝜇(𝑡)2=𝑥(𝜎(𝑡)),𝑥Δ(𝜎(𝑡))𝜇(𝑡)𝑥(𝜎(𝑡))𝑥(𝜎(𝑡))𝑥(𝑡)𝜇(𝑡)2,(2.13) and we conclude as in the previous case.

Let 𝜖>0. The exponential function 𝑒𝜖(,𝑡0) is defined by𝑒𝜖𝑡,𝑡0=exp[𝑡0,𝑡)𝕋𝜉𝜖,(𝜇(𝑠))Δ𝑠(2.14) where 𝜉𝜖()=𝜖,if=0,log(1+𝜖),if>0.(2.15) It is the unique solution to the initial value problem 𝑥Δ𝑡(𝑡)=𝜖𝑥(𝑡),𝑥0=1.(2.16)

Here is a result on time scales, analogous to Gronwall's inequality. The reader may find the proof of this result in [13].

Theorem 2.5. Let 𝛼>0, 𝜖>0, and 𝑦𝐶(𝕋,). If 𝑦(𝑡)𝛼+[𝑎,𝑡)𝕋𝜖𝑦(𝑠)Δ𝑠forevery𝑡𝕋,(2.17) then 𝑦(𝑡)𝛼𝑒𝜖(𝑡,𝑎)forevery𝑡𝕋.(2.18)

We recall some notions and results related to the theory of Δ-measure and Δ-Lebesgue integration introduced by Bohner and Guseinov in [12]. The reader is also referred to [14] for expressions of the Δ-measure and the Δ-integral in terms of the classical Lebesgue measure and the classical Lebesgue integral, respectively.

Definition 2.6. A set 𝐴𝕋 is said to be Δ-measurable if for every set 𝐸𝕋, 𝑚1(𝐸)=𝑚1(𝐸𝐴)+𝑚1(𝐸(𝕋𝐴)),(2.19) where 𝑚1(𝐸)=inf𝑚𝑘=1𝑑𝑘𝑐𝑘𝐸𝑚𝑘=1𝑐𝑘,𝑑𝑘with𝑐𝑘,𝑑𝑘𝕋if𝑏𝐸,if𝑏𝐸.(2.20) The Δ-measure on (𝑚1)={𝐴𝕋𝐴isΔ-measurable}, denoted by 𝜇Δ, is the restriction of 𝑚1 to (𝑚1). So, (𝕋,(𝑚1),𝜇Δ) is a complete measurable space.

Proposition 2.7 (see [14]). Let 𝐴𝕋, then 𝐴 is Δ-measurable if and only if 𝐴 is Lebesgue measurable. Moreover, if 𝑏𝐴, 𝜇Δ(𝐴)=𝑚(𝐴)+𝑡𝑖𝐴𝑅𝕋𝜎𝑡𝑖𝑡𝑖,(2.21) where 𝑚 is the Lebesgue measure.

The notions of Δ-measurable and Δ-integrable functions 𝑓𝕋𝑛 can be defined similarly to the general theory of Lebesgue integral.

Let 𝐸𝕋 be a Δ-measurable set and 𝑓𝕋𝑛 a Δ-measurable function. We say that 𝑓𝐿1Δ(𝐸,𝑛) provided 𝐸𝑓(𝑠)Δ𝑠<.(2.22) The set 𝐿1Δ(𝕋0,𝑛) is a Banach space endowed with the norm 𝑓𝐿1Δ=𝕋0𝑓(𝑠)Δ𝑠.(2.23) Here is an analog of the Lebesgue dominated convergence Theorem which can be proved as in the general theory of Lebesgue integration theory.

Theorem 2.8. Let {𝑓𝑘}𝑘 be a sequence of functions in 𝐿1Δ(𝕋0,𝑛). Assume that there exists a function 𝑓𝕋0𝑛 such that 𝑓𝑘(𝑡)𝑓(𝑡)Δ-a.e. 𝑡𝕋0, and there exists a function 𝑔𝐿1Δ(𝕋0) such that 𝑓𝑘(𝑡)𝑔(𝑡)Δ-a.e. 𝑡𝕋0 and for every 𝑘, then 𝑓𝑘𝑓 in 𝐿1Δ(𝕋0,𝑛).

In our existence results, we will consider Δ-Carathéodory functions.

Definition 2.9. A function 𝑓𝕋0×𝑚𝑛 is Δ-Carathéodory if the following conditions hold: (i)𝑡𝑓(𝑡,𝑥) is Δ-measurable for every 𝑥𝑚,(ii)𝑥𝑓(𝑡,𝑥) is continuous for Δ-almost every 𝑡𝕋0,(iii)for every 𝑟>0, there exists 𝑟𝐿1Δ(𝕋0,[0,)) such that 𝑓(𝑡,𝑥)𝑟(𝑡) for Δ-almost every 𝑡𝕋0 and for every 𝑥𝑚 such that 𝑥𝑟.

In this context, there is also a notion of absolute continuity introduced in [16].

Definition 2.10. A function 𝑓𝕋𝑛 is said to be absolutely continuous on 𝕋 if for every 𝜀>0, there exists a 𝛿>0 such that if {[𝑎𝑘,𝑏𝑘)}𝑚𝑘=1 with 𝑎𝑘,𝑏𝑘𝕋 is a finite pairwise disjoint family of subintervals satisfying 𝑚𝑘=1𝑏𝑘𝑎𝑘<𝛿,then𝑚𝑘=1𝑏𝑓𝑘𝑎𝑓𝑘<𝜀.(2.24)

Proposition 2.11 (see [17]). If 𝑓𝕋𝑛 is an absolutely continuous function, then the Δ-measure of the set {𝑡𝕋0𝑅𝕋𝑓(𝑡)=0and𝑓Δ(𝑡)0} is zero.

Proposition 2.12 (see [17]). If 𝑔𝐿1Δ(𝕋0,𝑛) and 𝑓𝕋𝑛 is the function defined by 𝑓(𝑡)=[𝑎,𝑡)𝕋𝑔(𝑠)Δ𝑠,(2.25) then 𝑓 is absolutely continuous and 𝑓Δ(𝑡)=𝑔(𝑡)Δ-almost everywhere on 𝕋0.

Proposition 2.13 (see [16]). A function 𝑓𝕋 is absolutely continuous on 𝕋 if and only if 𝑓 is Δ-differentiable Δ-almost everywhere on 𝕋0, 𝑓Δ𝐿1Δ(𝕋0) and [𝑎,𝑡)𝕋𝑓Δ(𝑠)Δ𝑠=𝑓(𝑡)𝑓(𝑎),forevery𝑡𝕋.(2.26)

We also recall a notion of Sobolev space, see [18],𝑊Δ1,1(𝕋,𝑛)=𝑥𝐿1Δ𝕋0,𝑛𝑔𝐿1Δ𝕋0,𝑛suchthat𝕋0𝑥(𝑠)𝜑Δ(𝑠)Δ𝑠=𝕋0𝑔(𝑠)𝜑(𝜎(𝑠))Δ𝑠forevery𝜑𝐶10,rd,(𝕋)(2.27) where 𝐶10,rd(𝕋)={𝜑𝕋𝜑(𝑎)=0=𝜑(𝑏),𝜑isΔ-dierentiableand𝜑Δiscontinuousatright-densepointsof𝕋anditsleft-sidedlimitsexistatleft-densepointsof𝕋}.(2.28) A function 𝑥𝑊Δ1,1(𝕋,𝑛) can be identified to an absolutely continuous map.

Proposition 2.14 (see [18]). Suppose that 𝑥𝑊Δ1,1(𝕋,) with some 𝑔𝐿1Δ(𝕋0,) satisfying (2.27), then there exists 𝑦𝕋 absolutely continuous such that 𝑦=𝑥,𝑦Δ=𝑔Δ-a.e.on𝕋0.(2.29) Moreover, if 𝑔 is 𝐶rd(𝕋𝜅,), then there exists 𝑦𝐶1rd(𝕋,) such that 𝑦=𝑥Δ-a.e.on𝕋0,𝑦Δ=𝑔Δ-a.e.on𝕋𝜅.(2.30)

Sobolev spaces of higher order can be defined inductively as follows: 𝑊Δ2,1(𝕋,𝑛)=𝑥𝑊Δ1,1(𝕋,𝑛)𝑥Δ𝑊Δ1,1(𝕋𝜅,𝑛).(2.31) With the norm 𝑥𝑊Δ1,1=𝑥𝐿1Δ+𝑥Δ𝐿1Δ (resp., 𝑥𝑊Δ2,1=𝑥𝐿1Δ+𝑥Δ𝐿1Δ+𝑥ΔΔ𝐿1Δ), 𝑊Δ1,1(𝕋,𝑛) (resp., 𝑊Δ2,1(𝕋,𝑛)) is a Banach space.

Remark 2.15. By Proposition 2.7, we know that 𝜇Δ({𝑡})>0 for every 𝑡𝑅𝕋. From this fact and the previous proposition, one realizes that there is no interest to look for solutions to (1.1) and (1.2) in 𝑊Δ2,1(𝕋,𝑛) and to consider Δ-Carathéodory maps 𝑓 in the case where the time scale is such that 𝜇Δ(𝕋0𝑅𝕋)=0. In particular, this is the case for difference equations. Let us point out that we consider more general time scales. Nevertheless, the results that we obtained are new in both cases.

As in the classical case, some embeddings have nice properties.

Proposition 2.16 (see [18]). The inclusion 𝑗1𝑊Δ2,1(𝕋,𝑛)𝐶1(𝕋,𝑛) is continuous.

Proposition 2.17. The inclusion 𝑗0𝑊Δ2,1(𝕋,𝑛)𝐶(𝕋,𝑛) is a continuous, compact, linear operator.

Proof. Arguing as in the proof of the Arzelà-Ascoli Theorem, we can show that the inclusion 𝑖𝐶1(𝕋,𝑛)𝐶(𝕋,𝑛) is linear, continuous, and compact. The conclusion follows from the previous proposition since 𝑗0=𝑖𝑗1.

We obtain a maximum principle in this context. To this aim, we use the following result.

Lemma 2.18 (see [19]). Let 𝑓𝕋 be a function with a local maximum at 𝑡0(𝑎,𝑏)𝕋. If 𝑓ΔΔ(𝜌(𝑡0)) exists, then 𝑓ΔΔ(𝜌(𝑡0))0 provided 𝑡0 is not at the same time left dense and right scattered.

Theorem 2.19. Let 𝑟𝑊Δ2,1(𝕋) be a function such that 𝑟ΔΔ(𝑡)>0Δ-almost everywhere on {𝑡𝕋𝜅20𝑟(𝜎(𝑡))>0}. If one of the following conditions holds: (i)𝑎0𝑟(𝑎)𝛾0𝑟Δ(𝑎)0 and 𝑎1𝑟(𝑏)+𝛾1𝑟Δ(𝜌(𝑏))0 (where 𝑎0, 𝑎1, 𝛾0, and 𝛾1 are defined as in (1.4)),(ii)𝑟(𝑎)=𝑟(𝑏) and 𝑟Δ(𝑎)𝑟Δ(𝜌(𝑏)), then 𝑟(𝑡)0, for every 𝑡𝕋.

Proof. If the conclusion is false, there exists 𝑡0𝕋 such that 𝑟𝑡0=max𝑡𝕋𝑟(𝑡)>0.(2.32)
In the case where 𝑎𝜌(𝑡0)<𝑡0<𝑏, then 𝑟ΔΔ(𝜌(𝑡0)) exists since 𝜇Δ({𝜌(𝑡0)})=𝑡0𝜌(𝑡0)>0 and 𝑟𝑊Δ2,1(𝕋). By Lemma 2.18, 𝑟ΔΔ(𝜌(𝑡0))0, which is a contradiction since 𝑟(𝜎(𝜌(𝑡0)))=𝑟(𝑡0)>0.
If 𝑎<𝜌(𝑡0)=𝑡0=𝜎(𝑡0)<𝑏, there exists 𝑡1>𝑡0 such that 𝑟(𝜎(𝑡))>0 for every 𝑡(𝑡0,𝑡1)𝕋. On the other hand, since 𝑟(𝑡0) is a maximum, 𝑟Δ(𝑡0)=0, and there exists 𝑠(𝑡0,𝑡1) such that 𝑟Δ(𝑠)0, thus, 0𝑟Δ(𝑠)𝑟Δ𝑡0=[𝑡0,𝑠)𝕋𝑟ΔΔ(𝜏)Δ𝜏>0,(2.33) by hypothesis and Proposition 2.13, which is a contradiction. Observe that the same argument applies if 𝑎=𝜌(𝑡0)=𝑡0=𝜎(𝑡0) and 𝑟Δ(𝑡0)=0. Notice also that if 𝜌(𝑡0)=𝑡0=𝜎(𝑡0)=𝑏 and 𝑟Δ(𝑡0)=0, we get a contradiction with an analogous argument for a suitable 𝑡1<𝑡0.
If 𝑎𝜌(𝑡0)=𝑡0<𝜎(𝑡0)<𝑏 and 𝑟(𝑡0)=𝑟(𝜎(𝑡0)), we argue as in the first case replacing 𝑡0 by 𝜎(𝑡0) to obtain a contradiction.
In the case where 𝑎<𝜌(𝑡0)=𝑡0<𝜎(𝑡0)𝑏 and 𝑟(𝑡0)>𝑟(𝜎(𝑡0)), then 𝑟Δ(𝑡0)<0. Since 𝑡𝑟Δ(𝑡) is continuous, there exists 𝛿>0 such that 𝑟Δ(𝑡)<0 on an interval (𝑡0𝛿,𝑡0), which contradicts the maximality of 𝑟(𝑡0).
Observe that 𝑎=𝑡0 and 𝑟Δ(𝑎)<0 could happen if 𝑎=𝑡0=𝜎(𝑡0) or if 𝑎=𝑡0<𝜎(𝑡0) and 𝑟(𝑎)>𝑟(𝜎(𝑎)). In this case, we get a contradiction if 𝑟 satisfies condition (i). On the other hand, if 𝑟 satisfies condition (ii), then 𝑟(𝑏)=𝑟(𝑎)>0 and 𝑟Δ(𝜌(𝑏))<0. So 𝜌(𝑏)=𝑏 since 𝑟(𝑏)𝑟(𝜌(𝑏)). This contradicts the maximality of 𝑟(𝑏).
On the other hand, the case where 𝑡0=𝑏 and 𝑟 satisfies condition (i) can be treated similarly to the previous case.

Finally, we define the Δ-differential operator associated to the problems that we will consider 𝐿𝑊2,1Δ,𝐵(𝕋,𝑛)𝐿1Δ𝕋𝜅0,𝑛denedby𝐿(𝑥)(𝑡)=𝑥ΔΔ(𝑡)𝑥(𝜎(𝑡)),(2.34) where 𝑊2,1Δ,𝐵(𝕋,𝑛)={𝑥𝑊Δ2,1(𝕋,𝑛)𝑥(BC)} with (BC) denoting the periodic boundary condition (1.3) or the Sturm-Liouville boundary condition (1.4).

Proposition 2.20. The operator 𝐿 is invertible and 𝐿1 is affine and continuous.

Proof. If (BC) denotes the Sturm-Liouville boundary condition (1.4), consider the associated homogeneous boundary condition 𝑎0𝑥(𝑎)𝛾0𝑥Δ𝑎(𝑎)=0,1𝑥(𝑏)+𝛾1𝑥Δ(𝜌(𝑏))=0.(2.35) Denote 𝑊2,1Δ,𝐵0(𝕋,𝑛)=𝑥𝑊Δ2,1(𝕋,𝑛)𝑥satises(1.3)if(BC)denotes(1.3),𝑥𝑊Δ2,1(𝕋,𝑛)𝑥satises(2.35)if(BC)denotes(1.4).(2.36) Notice that 𝑊2,1Δ,𝐵0(𝕋,𝑛) is a Banach space. Define 𝐿0𝑊2,1Δ,𝐵0(𝕋,𝑛)𝐿1Δ𝕋𝜅0,𝑛by𝐿0(𝑥)(𝑡)=𝑥ΔΔ(𝑡)𝑥(𝜎(𝑡)).(2.37) It is obvious that 𝐿0 is linear and continuous. It follows directly from Theorem 2.19 that 𝐿0 is injective.
If (BC) denotes (1.4) (resp., (1.3)), let 𝐺(𝑡,𝑠) be the Green function given in [13, Theorem  4.70] (resp., [13, Theorem  4.89]). Arguing as in [13, Theorem  4.70] (resp., [13, Theorem  4.89]), one can verify that for any 𝐿1Δ(𝕋𝜅0,𝑛), 𝑥(𝑡)=[𝑎,𝜌(𝑏))𝕋𝐺(𝑡,𝑠)(𝑠)Δ𝑠(2.38) is a solution of 𝐿0(𝑥)=. So, 𝐿0 is bijective and, hence, invertible with 𝐿01 continuous by the inverse mapping theorem.
Finally, if (BC) denotes (1.3), since 𝐿=𝐿0, we have the conclusion. On the other hand, if (BC) denotes (1.4), let 𝑦 be given in [13, Theorem  4.67] such that 𝑦ΔΔ(𝑡)𝑦(𝜎(𝑡))=0on𝕋𝜅2,𝑎0𝑦(𝑎)𝛾0𝑦Δ(𝑎)=𝑥0,𝑎1𝑦(𝑏)+𝛾1𝑦Δ(𝜌(𝑏))=𝑥1,(2.39) then 𝐿1=𝑦+𝐿01.

Remark 2.21. We could have considered the operator 𝐿𝑊2,1Δ,𝐵(𝕋,𝑛)𝐿1Δ𝕋𝜅0,𝑛denedby𝐿(𝑥)(𝑡)=𝑥ΔΔ(𝑡),(2.40) when the boundary condition is (1.4) with suitable constants 𝑎0, 𝑎1, 𝛾0, 𝛾1 such that 𝐿 is injective. For simplicity, we prefer to use only the operator 𝐿.

3. Nonlinearity Not Depending on the Delta-Derivative

In this section, we establish existence results for the problem 𝑥ΔΔ(𝑡)=𝑓(𝑡,𝑥(𝜎(𝑡))),Δ-a.e.𝑡𝕋𝜅20,𝑥(BC),(3.1) where (BC) denotes the periodic boundary condition 𝑥𝑥(𝑎)=𝑥(𝑏),Δ(𝑎)=𝑥Δ(𝜌(𝑏)),(3.2) or the Sturm-Liouville boundary condition 𝑎0𝑥(𝑎)𝛾0𝑥Δ(𝑎)=𝑥0,𝑎1𝑥(𝑏)+𝛾1𝑥Δ(𝜌(𝑏))=𝑥1,(3.3) where 𝑎0,𝑎1,𝛾0,𝛾10, max{𝑎0,𝛾0}>0, and max{𝑎1,𝛾1}>0. We look for solutions in 𝑊Δ2,1(𝕋,𝑛).

We introduce the notion of solution-tube for the problem (3.1).

Definition 3.1. Let (𝑣,𝑀)𝑊Δ2,1(𝕋,𝑛)×𝑊Δ2,1(𝕋,[0,)). We say that (𝑣,𝑀) is a solution-tube of (3.1) if (i)𝑥𝑣(𝜎(𝑡)),𝑓(𝑡,𝑥)𝑣ΔΔ(𝑡)𝑀(𝜎(𝑡))𝑀ΔΔ(𝑡) for Δ-almost every 𝑡𝕋𝜅20 and for every 𝑥𝑛 such that 𝑥𝑣(𝜎(𝑡))=𝑀(𝜎(𝑡)),(ii)𝑣ΔΔ(𝑡)=𝑓(𝑡,𝑣(𝜎(𝑡))) and 𝑀ΔΔ(𝑡)0 for Δ-almost every 𝑡𝕋𝜅20 such that 𝑀(𝜎(𝑡))=0,(iii)(a)if (BC) denotes (3.2), then 𝑣(𝑎)=𝑣(𝑏), 𝑀(𝑎)=𝑀(𝑏), and 𝑣Δ(𝜌(𝑏))𝑣Δ(𝑎)𝑀Δ(𝜌(𝑏))𝑀Δ(𝑎), (b)if (BC) denotes (3.3), 𝑥0(𝑎0𝑣(𝑎)𝛾0𝑣Δ(𝑎))𝑎0𝑀(𝑎)𝛾0𝑀Δ(𝑎), 𝑥1(𝑎1𝑣(𝑏)+𝛾1𝑣Δ(𝜌(𝑏)))𝑎1𝑀(𝑏)+𝛾1𝑀Δ(𝜌(𝑏)). We denote 𝑇(𝑣,𝑀)=𝑥𝑊Δ2,1(𝕋,𝑛.)𝑥(𝑡)𝑣(𝑡)𝑀(𝑡)𝑡𝕋(3.4)

We state the main theorem of this section.

Theorem 3.2. Let 𝑓𝕋𝜅0×𝑛𝑛 be a Δ-Carathéodory function. If (𝑣,𝑀)𝑊Δ2,1(𝕋,𝑛)×𝑊Δ2,1(𝕋,[0,)) is a solution-tube of (3.1), then the system (3.1) has a solution 𝑥𝑊Δ2,1(𝕋,𝑛)𝑇(𝑣,𝑀).

In order to prove this result, we consider the following modified problem:𝑥ΔΔ(𝑡)𝑥(𝜎(𝑡))=𝑓𝑡,𝑥(𝜎(𝑡))𝑥(𝜎(𝑡)),Δ-a.e.𝑡𝕋𝜅20,𝑥(BC),(3.5) where𝑥(𝑠)=𝑀(𝑠)𝑥𝑣(𝑠)(𝑥𝑣(𝑠))+𝑣(𝑠)if𝑥𝑥𝑣(𝑠)>𝑀(𝑠),otherwise.(3.6)

We define the operator 𝐹𝐶(𝕋,𝑛)𝐿1Δ(𝕋𝜅0,𝑛) by 𝐹(𝑥)(𝑡)=𝑓𝑡,𝑥(𝜎(𝑡))𝑥(𝜎(𝑡)).(3.7)

Proposition 3.3. Under the assumptions of Theorem 3.2, the operator 𝐹 defined above is continuous and bounded.

Proof. First of all, we show that the set 𝐹(𝐶(𝕋,𝑛)) is bounded. Fix 𝑅>𝑣0+𝑀0. Let 𝑅𝐿1Δ(𝕋𝜅0,[0,)) be given by Definition 2.9(iii). Thus, for every 𝑥𝐶(𝕋,𝑛), 𝐹(𝑥)(𝑠)𝑅(𝑠)+𝑅=(𝑠)Δ-a.e.𝑠𝕋𝜅0.(3.8)
To prove the continuity of 𝐹, we consider {𝑥𝑘}𝑘 a sequence of 𝐶(𝕋,𝑛) converging to 𝑥𝐶(𝕋,𝑛). We already know that for every 𝑘, 𝑥𝐹𝑘(𝑠)(𝑠)Δ-a.e.𝑠𝕋𝜅0.(3.9) One can easily check that 𝑥𝑛(𝑡)𝑥(𝑡) for all 𝑡𝕋. It follows from Definition 2.9(ii) that 𝐹𝑥𝑘(𝑠)𝐹(𝑥)(𝑠)Δ-a.e.𝑠𝕋𝜅0.(3.10) Theorem 2.8 implies that 𝐹(𝑥𝑘)𝐹(𝑥) in 𝐿1Δ(𝕋𝜅0,𝑛).

Lemma 3.4. Under the assumptions of Theorem 3.2, every solution 𝑥 of (3.5) is in 𝑇(𝑣,𝑀).

Proof. Since 𝑥𝑊Δ2,1(𝕋,𝑛) (resp., 𝑣𝑊Δ2,1(𝕋,𝑛), 𝑀𝑊Δ2,1(𝕋,)), 𝑥ΔΔ(𝑡) (resp., 𝑣ΔΔ(𝑡), 𝑀ΔΔ(𝑡)), there exists Δ-almost everywhere on 𝕋𝜅2. Denote 𝐴=𝑡𝕋𝜅20.𝑥(𝜎(𝑡))𝑣(𝜎(𝑡))>𝑀(𝜎(𝑡))(3.11) By Lemma 2.4(1), (𝑥(𝑡)𝑣(𝑡)𝑀(𝑡))ΔΔ𝑥(𝜎(𝑡))𝑣(𝜎(𝑡)),𝑥ΔΔ(𝑡)𝑣ΔΔ(𝑡)𝑥(𝜎(𝑡))𝑣(𝜎(𝑡))𝑀ΔΔ(𝑡)Δ-a.e.on𝐴.(3.12) We claim that (𝑥(𝑡)𝑣(𝑡)𝑀(𝑡))ΔΔ>0Δ-a.e.on𝐴.(3.13) Indeed, we deduce from the fact that (𝑣,𝑀) is a solution-tube of (3.5) and from (3.12) that Δ-almost everywhere on {𝑡𝐴𝑀(𝜎(𝑡))>0}, (𝑥(𝑡)𝑣(𝑡)𝑀(𝑡))ΔΔ𝑥(𝜎(𝑡))𝑣(𝜎(𝑡)),𝑓𝑡,𝑥(𝜎(𝑡))𝑥(𝜎(𝑡))+𝑥(𝜎(𝑡))𝑣ΔΔ(𝑡)𝑥(𝜎(𝑡))𝑣(𝜎(𝑡))𝑀ΔΔ=(𝑡)𝑥(𝜎(𝑡))𝑣(𝜎(𝑡)),𝑓𝑡,𝑥(𝜎(𝑡))𝑣ΔΔ(𝑡)𝑀(𝜎(𝑡))+𝑥(𝜎(𝑡))𝑣(𝜎(𝑡))𝑀(𝜎(𝑡))𝑀ΔΔ>(𝑡)𝑀(𝜎(𝑡))𝑀ΔΔ(𝑡)𝑀(𝜎(𝑡))𝑀ΔΔ(𝑡)=0,(3.14) and Δ-almost everywhere on {𝑡𝐴𝑀(𝜎(𝑡))=0}, (𝑥(𝑡)𝑣(𝑡)𝑀(𝑡))ΔΔ𝑥(𝜎(𝑡))𝑣(𝜎(𝑡)),𝑓𝑡,𝑥(𝜎(𝑡))𝑥(𝜎(𝑡))+𝑥(𝜎(𝑡))𝑣ΔΔ(𝑡)𝑥(𝜎(𝑡))𝑣(𝜎(𝑡))𝑀ΔΔ=(𝑡)𝑥(𝜎(𝑡))𝑣(𝜎(𝑡)),𝑓(𝑡,𝑣(𝜎(𝑡)))𝑣ΔΔ(𝑡)𝑥(𝜎(𝑡))𝑣(𝜎(𝑡))+𝑥(𝜎(𝑡))𝑣(𝜎(𝑡))𝑀ΔΔ(𝑡)>𝑀ΔΔ(𝑡)0.(3.15)
Observe that if 𝑥(𝑎)𝑣(𝑎)𝑀(𝑎)>0, 𝑥(𝑎)𝑣(𝑎)Δ𝑥(𝑎)𝑣(𝑎),𝑥Δ(𝑎)𝑣Δ(𝑎).𝑥(𝑎)𝑣(𝑎)(3.16) Indeed, this follows from Proposition 2.3 when 𝑎=𝜎(𝑎). For 𝑎<𝜎(𝑎), 𝑥(𝑎)𝑣(𝑎)Δ=𝑥(𝜎(𝑎))𝑣(𝜎(𝑎))𝑥(𝑎)𝑣(𝑎)𝜇(𝑎)𝑥(𝑎)𝑣(𝑎),𝑥(𝜎(𝑎))𝑣(𝜎(𝑎))(𝑥(𝑎)𝑣(𝑎))=𝜇(𝑎)𝑥(𝑎)𝑣(𝑎)𝑥(𝑎)𝑣(𝑎),𝑥Δ(𝑎)𝑣Δ(𝑎).𝑥(𝑎)𝑣(𝑎)(3.17) Similarly, if 𝑥(𝑏)𝑣(𝑏)𝑀(𝑏)>0, 𝑥(𝜌(𝑏))𝑣(𝜌(𝑏))Δ𝑥(𝑏)𝑣(𝑏),𝑥Δ(𝜌(𝑏))𝑣Δ(𝜌(𝑏)).𝑥(𝑏)𝑣(𝑏)(3.18)
If (BC) denotes (3.2), 𝑥(𝑎)𝑣(𝑎)𝑀(𝑎)=𝑥(𝑏)𝑣(𝑏)𝑀(𝑏).(3.19) We deduce from (3.16), (3.18), and Definition 3.1 that 𝑥(𝑎)𝑣(𝑎)𝑀(𝑎)=𝑥(𝑏)𝑣(𝑏)𝑀(𝑏)0,(3.20) or 𝑥(𝜌(𝑏))𝑣(𝜌(𝑏))Δ𝑀Δ(𝜌(𝑏))𝑥(𝑎)𝑣(𝑎)Δ𝑀Δ(𝑎)𝑥Δ(𝜌(𝑏))𝑣Δ(𝜌(𝑏)),𝑥(𝑏)𝑣(𝑏)𝑥(𝑏)𝑣(𝑏)𝑥Δ(𝑎)𝑣Δ(𝑎),𝑥(𝑎)𝑣(𝑎)𝑀𝑥(𝑎)𝑣(𝑎)Δ(𝜌(𝑏))𝑀Δ=(𝑎)𝑣Δ(𝑎)𝑣Δ(𝜌(𝑏)),𝑥(𝑎)𝑣(𝑎)𝑀𝑥(𝑎)𝑣(𝑎)Δ(𝜌(𝑏))𝑀Δ(𝑎)𝑣Δ(𝑎)𝑣Δ𝑀(𝜌(𝑏))Δ(𝜌(𝑏))𝑀Δ(𝑎)0.(3.21)
If (BC) denotes (3.3), we deduce from (3.16), (3.18), and Definition 3.1 that 𝑥(𝑎)𝑣(𝑎)𝑀(𝑎)0,(3.22) or 𝑎0(𝑥(𝑎)𝑣(𝑎)𝑀(𝑎))𝛾0𝑥(𝑎)𝑣(𝑎)Δ𝑀Δ(𝑎)𝑥(𝑎)𝑣(𝑎),𝑎0(𝑥(𝑎)𝑣(𝑎))𝛾0𝑥Δ(𝑎)𝑣Δ(𝑎)𝑥(𝑎)𝑣(𝑎)𝑎0𝑀(𝑎)+𝛾0𝑀Δ𝑥(𝑎)0𝑎0𝑣(𝑎)𝛾0𝑣Δ(𝑎)𝑎0𝑀(𝑎)+𝛾0𝑀Δ(𝑎)0,(3.23) and similarly 𝑥(𝑏)𝑣(𝑏)𝑀(𝑏)0,(3.24) or 𝑎1(𝑥(𝑏)𝑣(𝑏)𝑀(𝑏))+𝛾1𝑥(𝜌(𝑏))𝑣(𝜌(𝑏))Δ𝑀Δ(𝜌(𝑏))0.(3.25)
Finally, it follows from (3.13), (3.20), (3.21), (3.22), (3.23), (3.24), (3.25), and Theorem 2.19 applied to 𝑟(𝑡)=𝑥(𝑡)𝑣(𝑡)𝑀(𝑡) that 𝑥(𝑡)𝑣(𝑡)𝑀(𝑡) for every 𝑡𝕋.

Now, we can prove the main theorem of this section.

Proof of Theorem 3.2. A solution of (3.5) is a fixed point of the operator 𝑇𝐶(𝕋,𝑛)𝐶(𝕋,𝑛)denedby𝑇=𝑗0𝐿1𝐹,(3.26) where 𝐿 and 𝑗0 are defined in (2.34) and Proposition 2.17, respectively. By Propositions 2.17, 2.20, and 3.3, the operator 𝑇 is compact. So, the Schauder fixed point theorem implies that 𝑇 has a fixed point and, hence, Problem (3.5) has a solution 𝑥. By Lemma 3.4, this solution is in 𝑇(𝑣,𝑀). Thus, 𝑥 is a solution of (3.1).

In the particular case where 𝑛=1, as corollary of Theorem 3.2, we obtain a generalization of results of Akın [1] and Stehlík [2] for the Dirichlet and the periodic boundary conditions, respectively.

Corollary 3.5. Let 𝑓𝕋𝜅0× be a Δ-Carathéodory function. Assume that there exists 𝛼,𝛽𝑊Δ2,1(𝕋,) such that (i)𝛼(𝑡)𝛽(𝑡) for Δ-almost every 𝑡𝕋,(ii)𝛼ΔΔ(𝑡)𝑓(𝑡,𝛼(𝜎(𝑡))) and 𝛽ΔΔ(𝑡)𝑓(𝑡,𝛽(𝜎(𝑡))) for Δ-almost every 𝑡𝕋𝜅2,(iii)(a)if (BC) denotes (3.2), then 𝛼(𝑎)=𝛼(𝑏), 𝛼Δ(𝑎)𝛼Δ(𝜌(𝑏)), 𝛽(𝑎)=𝛽(𝑏), and 𝛽Δ(𝑎)𝛽Δ(𝜌(𝑏)),(b)if (BC) denotes (3.3), then 𝑎0𝛼(𝑎)𝛾0𝛼Δ(𝑎)𝑥0𝑎0𝛽(𝑎)𝛾0𝛽Δ(𝑎) and 𝑎1𝛼(𝑏)+𝛾1𝛼Δ(𝜌(𝑏))𝑥1𝑎1𝛽(𝑏)+𝛾1𝛽Δ(𝜌(𝑏)), then (3.1) has a solution 𝑥𝑊Δ2,1(𝕋,) such that 𝛼(𝑡)𝑥(𝑡)𝛽(𝑡) for every 𝑡𝕋.

Proof. Observe that ((𝛼+𝛽)/2,(𝛽𝛼)/2) is a solution-tube of (3.1). The conclusion follows from Theorem 3.2.

Theorem 3.2 generalizes also a result established by Henderson et al. [6] for systems of second-order dynamic equations on time scales. Let us mention that they considered a continuous map 𝑓 and they assumed a strict inequality in (3.27).

Corollary 3.6. Let 𝑓𝕋𝜅0×𝑛𝑛 be a Δ-Carathéodory function. Assume that there exists a constant 𝑅>0 such that 𝑥,𝑓(𝑡,𝑥)0Δ-a.e.𝑡𝕋𝜅20,𝑥suchthat𝑥=𝑅.(3.27) Moreover, if (𝐵𝐶) denotes (3.3), assume that 𝑥0𝑎0𝑅 and 𝑥1𝑎1𝑅, then the system (3.1) has a solution 𝑥𝑊Δ2,1(𝕋,𝑛) such that 𝑥(𝑡)𝑅 for every 𝑡𝕋.

Here is an example in which one cannot find a solution-tube of the form (0,𝑅).

Example 3.7. Consider the system 𝑥ΔΔ(𝑡)=𝑘(𝑡)(𝑥(𝜎(𝑡))𝜎(𝑡)𝑐1)Δ-a.e.𝑡𝕋𝜅20,𝑥(𝑎)=0,𝑥(𝑏)=0,(3.28) where 𝑘𝐿1Δ(𝕋0,𝑛{0}) and 𝑐𝑛{0} is such that 𝑐max{|𝑎|,|𝑏|}1. One can check that (𝑣,𝑀) is a solution-tube of (3.28) with 𝑣(𝑡)=𝑡𝑐, 𝑀(𝑡)1. By Theorem 3.2, this problem has at least one solution 𝑥 such that 𝑥(𝑡)𝑡𝑐1. Observe that there is no 𝑅>0 such that (3.27) is satisfied.

4. Nonlinearity Depending on 𝑥Δ

In this section, we study more general systems of second-order dynamic equations on time scales. Indeed, we allow the nonlinearity 𝑓 to depend also on 𝑥Δ. We consider the problem 𝑥ΔΔ(𝑡)=𝑓𝑡,𝑥(𝜎(𝑡)),𝑥Δ(𝑡),Δ-a.e.𝑡𝕋𝜅20,𝑎0𝑥(𝑎)𝑥Δ(𝑎)=𝑥0,𝑎1𝑥(𝑏)+𝛾1𝑥Δ(𝜌(𝑏))=𝑥1,(4.1) where 𝑎0,𝑎1,𝛾10 and max{𝑎1,𝛾1}>0.

We also introduce a notion of solution-tube for this problem.

Definition 4.1. Let (𝑣,𝑀)𝑊Δ2,1(𝕋,𝑛)×𝑊Δ2,1(𝕋,(0,)). We say that (𝑣,𝑀) is a solution-tube of (4.1) if (i)for Δ-almost every 𝑡{𝑡𝕋𝜅20𝑡=𝜎(𝑡)}, 𝑥𝑣(𝑡),𝑓(𝑡,𝑥,𝑦)𝑣ΔΔ(𝑡)+𝑦𝑣Δ(𝑡)2𝑀(𝑡)𝑀ΔΔ𝑀(𝑡)+Δ(𝑡)2,(4.2) for every (𝑥,𝑦)2𝑛 such that 𝑥𝑣(𝑡)=𝑀(𝑡) and 𝑥𝑣(𝑡),𝑦𝑣Δ(𝑡)=𝑀(𝑡)𝑀Δ(𝑡),(ii)for every 𝑡{𝑡𝕋𝜅20𝑡<𝜎(𝑡)}, 𝑥𝑣(𝜎(𝑡)),𝑓(𝑡,𝑥,𝑦)𝑣ΔΔ(𝑡)𝑀(𝜎(𝑡))𝑀ΔΔ(𝑡),(4.3) for every (𝑥,𝑦)2𝑛 such that 𝑥𝑣(𝜎(𝑡))=𝑀(𝜎(𝑡)),(iii)𝑥0(𝑎0𝑣(𝑎)𝑣Δ(𝑎))𝑎0𝑀(𝑎)𝑀Δ(𝑎), 𝑥1(𝑎1𝑣(𝑏)+𝛾1𝑣Δ(𝜌(𝑏)))𝑎1𝑀(𝑏)+𝛾1𝑀Δ(𝜌(𝑏)).

If 𝕋 is the real interval [𝑎,𝑏], condition (ii) of the previous definition becomes useless, and we get the notion of solution-tube introduced by the first author in [10] for a system of second-order differential equations.

Here is the main result of this section.

Theorem 4.2. Let 𝑓𝕋𝜅0×2𝑛𝑛 be a Δ-Carathéodory function. Assume that (H1) there exists (𝑣,𝑀)𝑊Δ2,1(𝕋,𝑛)×𝑊Δ2,1(𝕋,(0,)) a solution-tube of (4.1),(H2)there exist constants 𝑐,𝑑>0 such that 𝑓(𝑡,𝑥,𝑦)𝑐+𝑑𝑦 for Δ-almost every 𝑡𝕋𝜅0 and for every (𝑥,𝑦)2𝑛 such that 𝑥𝑣(𝑡)𝑀(𝑡), then (4.1) has a solution 𝑥𝑊Δ2,1(𝕋,𝑛)𝑇(𝑣,𝑀).

To prove this existence result, we consider the following modified problem:𝑥ΔΔ(𝑡)𝑥(𝜎(𝑡))=𝑔𝑡,𝑥(𝜎(𝑡)),𝑥Δ(𝑡),Δ-a.e.𝑡𝕋𝜅20,𝑎0𝑥(𝑎)𝑥Δ(𝑎)=𝑥0,𝑎1𝑥(𝑏)+𝛾1𝑥Δ(𝜌(𝑏))=𝑥1,(4.4) where 𝑔𝕋𝜅0×2𝑛𝑛 is defined by 𝑔(𝑡,𝑥,𝑦)=𝑀(𝜎(𝑡))𝑓𝑥𝑣(𝜎(𝑡))𝑡,𝑥(𝜎(𝑡)),̃𝑦(𝑡)+𝑥(𝜎(𝑡))1𝑀(𝜎(𝑡))𝑣𝑥𝑣(𝜎(𝑡))ΔΔ𝑀(𝑡)+ΔΔ(𝑡)𝑥𝑣(𝜎(𝑡))(𝑥𝑣(𝜎(𝑡)))if𝑓𝑥𝑣(𝜎(𝑡))>𝑀(𝜎(𝑡)),𝑡,𝑥(𝜎(𝑡)),̃𝑦(𝑡)𝑥(𝜎(𝑡)),otherwise,(4.5) where 𝑥(𝜎(𝑡)) is defined as in (3.6), 𝑀̃𝑦(𝑡)=̂𝑦(𝑡)+Δ(𝑡)𝑥𝑣(𝜎(𝑡)),̂𝑦(𝑡)𝑣Δ(𝑡)𝑥𝑣(𝜎(𝑡))𝑥𝑣(𝜎(𝑡))𝑥𝑣(𝜎(𝑡))if𝑀𝑡=𝜎(𝑡),𝑥𝑣(𝜎(𝑡))>𝑀(𝜎(𝑡)),̂𝑦(𝑡)+Δ(𝑡)𝐾𝑀(𝜎(𝑡))1𝑦𝑣Δ(𝑡)(𝑥𝑣(𝜎(𝑡)))if𝑡=𝜎(𝑡),𝑥𝑣(𝜎(𝑡))𝑀(𝜎(𝑡)),𝑦𝑣Δ(𝑡)>𝐾,̂𝑦(𝑡)if𝑦𝑡<𝜎(𝑡),otherwise,𝐾̂𝑦(𝑡)=𝑦𝑣Δ(𝑡)𝑦𝑣Δ(𝑡)+𝑣Δ(𝑡)if𝑦𝑣Δ(𝑡)>𝐾,𝑦,otherwise,(4.6) with 𝐾>0 a constant which will be fixed later.

Remark 4.3. (1) Remark that 𝑥(𝜎(𝑡))𝑣(𝜎(𝑡))𝑀(𝜎(𝑡)),̃𝑦(𝑡)2𝐾+𝑣Δ||𝑀(𝑡)+Δ||.(𝑡)(4.7)
(2) If 𝑥𝑣(𝜎(𝑡))>𝑀(𝜎(𝑡)), 𝑥(𝜎(𝑡))𝑣(𝜎(𝑡))=𝑀(𝜎(𝑡)),𝑥𝑥(𝜎(𝑡))𝑣(𝜎(𝑡)),Δ(𝑡)𝑣Δ(𝑡)=𝑀(𝜎(𝑡))𝑀Δ(𝑡),for𝑡=𝜎(𝑡).(4.8)
(3) If 𝑥𝑣(𝜎(𝑡))>𝑀(𝜎(𝑡)) and 𝑡=𝜎(𝑡), 𝑥Δ(𝑡)𝑣Δ(𝑡)2𝑥=Δ(𝑡)𝑣Δ(𝑡)2+𝑀Δ(𝑡)2𝑥𝑥(𝑡)𝑣(𝑡),Δ(𝑡)𝑣Δ(𝑡)2𝑥(𝑡)𝑣(𝑡)2.(4.9)
(4) Since 𝑓 is Δ-Carathéodory, by (1), there exists 𝐿1Δ(𝕋𝜅0,) such that for every 𝑥,𝑦𝑛, 𝑔(𝑡,𝑥,𝑦)(𝑡)Δ-a.e.𝑡𝕋𝜅0.(4.10)
We associate to 𝑔 the operator 𝐺𝐶1(𝕋,𝑛)𝐿1Δ(𝕋𝜅0,𝑛) defined by 𝐺(𝑥)(𝑡)=𝑔𝑡,𝑥(𝜎(𝑡)),𝑥Δ.(𝑡)(4.11)

Proposition 4.4. Let 𝑓𝕋𝜅0×2𝑛𝑛 be a Δ-Carathéodory function. Assume that (H1) is satisfied, then 𝐺 is continuous.

Proof. Let {𝑥𝑘} be a sequence of 𝐶1(𝕋,𝑛) converging to 𝑥𝐶1(𝕋,𝑛). It is clear that 𝑥𝑘(𝑡)𝑥𝑥(𝑡),Δ𝑘𝑥(𝑡)Δ(𝑡).(4.12)
On {𝑡𝕋𝑡<𝜎(𝑡)}, we have 𝑔𝑡,𝑥𝑘(𝜎(𝑡)),𝑥Δ𝑘(𝑡)𝑔𝑡,𝑥(𝜎(𝑡)),𝑥Δ,(𝑡)(4.13) since 𝑓 is Δ-Carathéodory.
Similarly, Δ-almost everywhere on {𝑡𝕋𝜅0𝑡=𝜎(𝑡),𝑥(𝜎(𝑡))𝑣(𝜎(𝑡))𝑀(𝜎(𝑡))}, 𝑔𝑡,𝑥𝑘(𝜎(𝑡)),𝑥Δ𝑘(𝑡)𝑔𝑡,𝑥(𝜎(𝑡)),𝑥Δ,(𝑡)(4.14) since 𝑥Δ𝑘𝑥(𝑡)Δ(𝑡) and, for 𝑘 sufficiently large, 𝑥𝑘(𝜎(𝑡))𝑣(𝜎(𝑡))𝑀(𝜎(𝑡)).
Denote 𝑆={𝑡𝕋𝜅0𝑡=𝜎(𝑡)and𝑥(𝜎(𝑡))𝑣(𝜎(𝑡))=𝑀(𝜎(𝑡))} and 𝐼𝑡={𝑘𝑥𝑘(𝜎(𝑡))𝑣(𝜎(𝑡))𝑀(𝜎(𝑡))}. As before, it is easy to check that Δ-almost everywhere on {𝑡𝑆card𝐼𝑡=}, 𝑔𝑡,𝑥𝑘(𝜎(𝑡)),𝑥Δ𝑘(𝑡)𝑔𝑡,𝑥(𝜎(𝑡)),𝑥Δ(𝑡)as𝑘𝐼𝑡goestoinnity.(4.15) On the other hand, Proposition 2.11 implies that 𝑥(𝜎(𝑡))𝑣(𝜎(𝑡)),𝑥Δ(𝑡)𝑣Δ(𝑡)=𝑀(𝜎(𝑡))𝑀Δ(𝑡)Δ-a.e.𝑡𝑆.(4.16) So, Δ-almost everywhere on {𝑡𝑆card(𝐼𝑡)=}, 𝑥Δ𝑘𝑥(𝑡)=Δ𝑘+𝑀(𝑡)Δ𝑥(𝑡)𝑘𝑥(𝜎(𝑡))𝑣(𝜎(𝑡)),Δ𝑘(𝑡)𝑣Δ(𝑡)𝑥𝑘(𝑥𝜎(𝑡))𝑣(𝜎(𝑡))𝑘(𝜎(𝑡))𝑣(𝜎(𝑡))𝑥𝑘(𝑥𝜎(𝑡))𝑣(𝜎(𝑡))Δ+𝑀(𝑡)Δ𝑥(𝑡)𝑥(𝜎(𝑡))𝑣(𝜎(𝑡)),Δ(𝑡)𝑣Δ(𝑡)𝑥(𝜎(𝑡))𝑣(𝜎(𝑡))𝑥(𝜎(𝑡))𝑣(𝜎(𝑡))=𝑥𝑥(𝜎(𝑡))𝑣(𝜎(𝑡))Δ𝑀(𝑡)+Δ(𝑡)𝐾𝑀(𝜎(𝑡))1𝑥Δ(𝑡)𝑣Δ(𝑡)(𝑥(𝜎(𝑡))𝑣(𝜎(𝑡)))if𝑥Δ(𝑡)𝑣Δ𝑥(𝑡)>𝐾,Δ(𝑡)if𝑥Δ(𝑡)𝑣Δ=𝑥(𝑡)𝐾,Δ(𝑡),(4.17) as 𝑘𝐼𝑡 goes to infinity. Thus, Δ-almost everywhere on {𝑡𝑆card(𝐼𝑡)=}, 𝑔𝑡,𝑥𝑘(𝜎(𝑡)),𝑥Δ𝑘(𝑡)𝑔𝑡,𝑥(𝜎(𝑡)),𝑥Δ(𝑡)as𝑘𝐼𝑡goestoinnity.(4.18)
By Remark 4.3(4), we have 𝑔𝑡,𝑥𝑘(𝜎(𝑡)),𝑥Δ𝑘(𝑡)(𝑡)Δ-a.e.𝑡𝕋𝜅0.(4.19) Theorem 2.8 implies that 𝐺𝑥𝑘𝐺(𝑥)in𝐿1Δ𝕋𝜅0,𝑛.(4.20)

Lemma 4.5. Assume (H1), then 𝑥𝑇(𝑣,𝑀) for every solution 𝑥 of (4.4).

Proof. Observe that 𝑥ΔΔ(𝑡) (resp., 𝑣ΔΔ(𝑡), 𝑀ΔΔ(𝑡)) exists Δ-almost everywhere on 𝕋𝜅2, since 𝑥𝑊Δ2,1(𝕋,𝑛) (resp., 𝑣𝑊Δ2,1(𝕋,𝑛), 𝑀𝑊Δ2,1(𝕋,)). Denote 𝐴=𝑡𝕋𝜅20.𝑥(𝜎(𝑡))𝑣(𝜎(𝑡))>𝑀(𝜎(𝑡))(4.21) Observe that by (H1) and Remark 4.3(2), for Δ-almost every 𝑡𝐴, 𝑥(𝜎(𝑡))𝑣(𝜎(𝑡)),𝑥ΔΔ(𝑡)𝑣ΔΔ(𝑡)=𝑥(𝜎(𝑡))𝑣(𝜎(𝑡)),𝑔𝑡,𝑥(𝜎(𝑡)),𝑥Δ(𝑡)+𝑥(𝜎(𝑡))𝑣ΔΔ(𝑡)=𝑥(𝜎(𝑡))𝑣(𝜎(𝑡)),𝑓𝑡,𝑥𝑥(𝜎(𝑡)),Δ(𝑡)𝑣ΔΔ(𝑡)+𝑀ΔΔ(𝑡)(𝑥(𝜎(𝑡))𝑣(𝜎(𝑡))𝑀(𝜎(𝑡)))+𝑥(𝜎(𝑡))𝑣(𝜎(𝑡))𝑀(𝜎(𝑡))>𝑥(𝜎(𝑡))𝑣(𝜎(𝑡)),𝑓𝑡,𝑥𝑥(𝜎(𝑡)),Δ(𝑡)𝑣ΔΔ(𝑡)+𝑀ΔΔ𝑀(𝑡)(𝑥(𝜎(𝑡))𝑣(𝜎(𝑡))𝑀(𝜎(𝑡)))ΔΔ(𝑡)𝑥(𝜎(𝑡))𝑣(𝜎(𝑡))if𝑀𝑡<𝜎(𝑡),ΔΔ𝑀(𝑡)𝑥(𝜎(𝑡))𝑣(𝜎(𝑡))+Δ(𝑡)2𝑥Δ(𝑡)𝑣Δ(𝑡)2if𝑡=𝜎(𝑡).(4.22)
This inequality with Lemma 2.4(1) imply that for Δ-almost every 𝑡{𝑡𝐴𝑡<𝜎(𝑡)}, (𝑥(𝑡)𝑣(𝑡)𝑀(𝑡))ΔΔ>0.(4.23) Also, (4.22), Lemma 2.4(2), and Remark 4.3(3) imply that for Δ-almost every 𝑡{𝑡𝐴𝑡=𝜎(𝑡)}, (𝑥(𝑡)𝑣(𝑡)𝑀(𝑡))ΔΔ>𝑀Δ(𝑡)2𝑥Δ(𝑡)𝑣Δ(𝑡)2+𝑥Δ(𝑡)𝑣Δ(𝑡)2𝑥(𝑡)𝑣(𝑡)𝑥(𝑡)𝑣(𝑡),𝑥Δ(𝑡)𝑣Δ(𝑡)2𝑥(𝑡)𝑣(𝑡)3=𝑥Δ(𝑡)𝑣Δ(𝑡)2𝑥Δ(𝑡)𝑣Δ(𝑡)2+𝑥𝑥(𝑡)𝑣(𝑡)𝑥(𝑡)𝑣(𝑡),Δ(𝑡)𝑣Δ(𝑡)2𝑥(𝑡)𝑣(𝑡),𝑥Δ(𝑡)𝑣Δ(𝑡)2𝑥(𝑡)𝑣(𝑡)3=0if𝑥Δ(𝑡)𝑣Δ𝐾(𝑡)𝐾,12𝑥Δ(𝑡)𝑣Δ(𝑡)2𝑥Δ(𝑡)𝑣Δ(𝑡)2𝑥(𝑡)𝑣(𝑡)𝑥(𝑡)𝑣(𝑡),𝑥Δ(𝑡)𝑣Δ(𝑡)2𝑥(𝑡)𝑣(𝑡)3if𝑥Δ(𝑡)𝑣Δ(𝑡)>𝐾,0.(4.24)
Let us denote 𝑟(𝑡)=𝑥(𝑡)𝑣(𝑡)𝑀(𝑡). Inequalities (4.23) and (4.24) imply that 𝑟ΔΔ(𝑡)>0 for Δ-almost every 𝑡{𝑡𝕋𝜅20𝑟(𝜎(𝑡))>0}. Arguing as in the proof of Lemma 3.4, we can show that 𝑟(𝑎)0or𝑎0𝑟(𝑎)𝑟Δ(𝑎)0,𝑟(𝑏)0or𝑎1𝑟(𝑏)+𝛾1𝑟Δ(𝜌(𝑏))0.(4.25) Theorem 2.19 implies that 𝑥𝑇(𝑣,𝑀).

We can now prove the existence theorem of this section.

Proof of Theorem 4.2. We first show that for every solution 𝑥 of (4.4), there exists a constant 𝐾>0 such that 𝑥Δ(𝑡)𝑣Δ(𝑡)𝐾forevery𝑡𝕋𝜅.(4.26) By (H2), Proposition 2.13 and Lemma 4.5, for any 𝑥 solution of (4.4), we have for Δ-almost every 𝑡𝕋𝜅, 𝑥Δ(𝑡)𝑥Δ(𝑎)+[𝑎,𝑡)𝕋𝑥ΔΔ(𝑠)Δ𝑠=𝑥0𝑎0𝑥(𝑎)+[𝑎,𝑡)𝕋𝑓𝑥𝑠,𝑥(𝜎(𝑠)),Δ(𝑠)Δ𝑠𝑐0+[𝑎,𝑡)𝕋𝑥𝑐+𝑑Δ(𝑠)Δ𝑠𝑐0+[𝑎,𝑡)𝕋𝑥𝑐+𝑑Δ(𝑠)𝑣Δ(𝑠)+𝑣Δ||𝑀(𝑠)+Δ||(𝑠)Δ𝑠𝑐0+[𝑎,𝑡)𝕋𝑐+𝑑𝑥Δ(𝑠)𝑣Δ(𝑠)+𝑣Δ(||𝑀𝑠)+Δ(||𝑠)Δ𝑠𝑐1+[𝑎,𝑡)𝕋𝑑𝑥Δ(𝑠)Δ𝑠,(4.27) where 𝑐0=𝑥0+𝑎0𝑐(𝑀(𝑎)+𝑣(𝑎)),1=𝑐0+[𝑎,𝑏)𝕋𝑐+𝑑2𝑣Δ||𝑀(𝑠)+Δ||(𝑠)Δ𝑠.(4.28) Fix 𝐾𝑣Δ0+𝑐1𝑒𝑑(,𝑎)0. By Theorem 2.5, 𝑥Δ(𝑡)𝑣Δ(𝑡)𝑥Δ(𝑡)+𝑣Δ(𝑡)𝑣Δ(𝑡)+𝑐1𝑒𝑑(𝑡,𝑎)𝐾𝑡𝕋𝜅.(4.29)
Consider the operator 𝑇=𝑗1𝐿1𝐺𝐶1(𝕋,𝑛)𝐶1(𝕋,𝑛),(4.30) where 𝐿 and 𝑗1 are defined, respectively, in (2.34) and Proposition 2.16. By Propositions 2.16, 2.20, and 4.4, 𝑇 is continuous. Moreover, 𝑇 is compact. Indeed, by Remark 4.3(4), there exists 𝐿1Δ(𝕋𝜅0,[0,)) such that for every 𝑧𝑇(𝐶1(𝕋,𝑛)), there exists 𝑥𝐶1(𝕋,𝑛) such that 𝑧=𝑇(𝑥) and 𝐺(𝑥)(𝑠)(𝑠)Δ-a.e.𝑡𝕋𝜅0.(4.31) Since 𝑗1 and 𝐿1 are continuous and affine, they map bounded sets in bounded sets. Thus, there exists a constant 𝑘0>0 such that 𝑧1𝑘0.(4.32) Moreover, 𝑧𝑊Δ2,1(𝕋,𝑛) and 𝐿(𝑧)(𝑠)=𝑧ΔΔ(𝑠)𝑧(𝜎(𝑠))=𝐺(𝑥)(𝑠)Δ-a.e.𝑠𝕋𝜅2.(4.33) So, for every 𝑡<𝜏 in 𝕋𝜅, 𝑧Δ(𝑡)𝑧Δ(𝜏)[𝑡,𝜏)𝕋𝑧(𝜎(𝑠))+𝐺(𝑥)(𝑠)Δ𝑠[𝑡,𝜏)𝕋𝑘0+(𝑠)Δ𝑠.(4.34) Thus, 𝑇(𝐶1(𝕋,𝑛)) is bounded and equicontinuous in 𝐶1(𝕋,𝑛). By an analog of the Arzelà-Ascoli Theorem for our context, 𝑇(𝐶1(𝕋,𝑛)) is relatively compact in 𝐶1(𝕋,𝑛).
By the Schauder fixed point theorem, 𝑇 has a fixed point 𝑥 which is a solution of (4.4). By Lemma 4.5, 𝑥𝑇(𝑣,𝑀). Also, 𝑥 satisfies (4.26). Hence, 𝑥 is also a solution of (4.1).
Here is an example in which one cannot find a solution-tube of the form (0,𝑅). Moreover, Assumption (A2) stated in the introduction and assumed in [9] is not satisfied.

Example 4.6. Let 𝕋=[0,1]{2}[3,4] and consider the system 𝑥ΔΔ(𝑡)=𝑥(𝜎(𝑡))𝑙(𝑡)+𝑥Δ(𝑡)+𝑘(𝑡)Δ-a.e.𝑡𝕋0,𝑥Δ(0)=𝑥0,𝑥Δ(4)=𝑥1,(4.35) where 𝑥0,𝑥1𝑛 are such that 𝑥01, 𝑥11, 𝑙𝐿1Δ(𝕋0,[0,)), and 𝑘𝐿1Δ(𝕋0,𝑛) such that 𝑙(1)1, 𝑙(2)1, 𝑘(𝑡)2, and 𝑙(𝑡)𝑟 for Δ-almost every 𝑡𝕋0 for some 𝑟>0.
Take 𝑣(𝑡)0 and 𝑀(𝑡)=5𝑡if[]𝑡𝑡0,2𝕋,if[]𝑡3,4𝕋.(4.36) So, 𝑣𝑊Δ2,1(𝕋,𝑛), 𝑀𝑊Δ2,1(𝕋,(0,)), and 𝑀Δ(𝑡)=1if[],0𝑡0,1if1𝑡=2,if[],𝑀𝑡3,4ΔΔ0(𝑡)=if[1𝑡0,1),if0𝑡{1,2},if[].𝑡3,4(4.37) One has 𝑥0𝑀Δ(0) and 𝑥1𝑀Δ(4). Observe that Δ-almost everywhere on [0,1)[3,4] if 𝑥=𝑀(𝑡) and 𝑥,𝑦=𝑀(𝑡)𝑀Δ(𝑡), one has 𝑦1 and 𝑥,𝑓(𝑡,𝑥,𝑦)+𝑦2=𝑥2(𝑙(𝑡)+𝑦)+𝑥,𝑘(𝑡)+𝑦2𝑀(𝑡)(𝑀(𝑡)𝑘(𝑡))+11=𝑀(𝑡)𝑀ΔΔ𝑀(𝑡)+Δ(𝑡)2.(4.38) If 𝑡{1,2} and 𝑥=𝑀(𝜎(𝑡))=3, 𝑥,𝑓(𝑡,𝑥,𝑦)=𝑥2(𝑙(𝑡)+𝑦)+𝑥,𝑘(𝑡)9𝑙(𝑡)3𝑘(𝑡)3=𝑀(𝜎(𝑡))𝑀ΔΔ(𝑡).(4.39) So, (𝑣,𝑀) is a solution-tube of (4.35). Moreover, 𝑓(𝑡,𝑥,𝑦)2+5𝑟+5𝑦Δ-a.e.𝑡𝕋0,all(𝑥,𝑦)suchthat𝑥𝑀(𝑡).(4.40) Theorem 4.2 implies that (4.35) has at least one solution 𝑥 such that 𝑥(𝑡)𝑀(𝑡). Observe that if 𝑥00 or 𝑥10, this problem has no solution-tube of the form (0,𝑅) with 𝑅 a positive constant since Definition 4.1(iii) would not be satisfied. This explains why Henderson and Tisdell [9] did not consider the Neumann boundary condition. Notice also that the restriction 𝑑(𝜌(𝑏)𝑎) in (A2) is not satisfied in this example.

Acknowledgments

The authors would like to thank, respectively, CRSNG-Canada and FQRNT-Québec for their financial support.