Preprint – December 2020
Bound sets for a class of -Laplacian operators
Abstract.
We provide an extension of the Hartman–Knobloch theorem for periodic solutions of vector differential systems to a general class of -Laplacian differential operators. Our main tool is a variant of the Manásevich–Mawhin continuation theorem developed for this class of operator equations, together with the theory of bound sets. Our results concern the case of convex bound sets for which we show some new connections using a characterisation of sublevel sets due to Krantz and Parks. We also extend to the -Laplacian vector case a classical theorem of Reissig for scalar periodically perturbed Liénard equations.
Key words and phrases:
Periodic solutions, continuation theorems, -Laplacian operators, bound sets, Nagumo–Hartman condition, Liénard and Rayleigh systems.1991 Mathematics Subject Classification:
34B15, 34C25, 47H11.1. Introduction
In the present paper, we study the existence of periodic solutions to differential systems involving a -Laplacian differential operator, with the aim of extending the classical theorem of Hartman–Knobloch for the differential system
(1.1) |
In his original work [18] (see also [19, ch. XII]), Hartman considered the two-point boundary value problem (BVP) associated with (1.1), assuming a (at most quadratic) growth condition of in (the so-called “Bernstein–Nagumo–Hartman condition”, cf. [9, 32]) and the hypothesis that the vector field is “repulsive” according to
for every with | (1.2) | |||
and such that . |
In the special case of a Newtonian force without friction , condition (1.2) reads as
(1.3) |
Under the above assumptions, one find that the BVP has at least one solution in the closed ball with center and radius (or in the open one , if the inequalities in (1.2) or (1.3) are strict). The result was then obtained in the context of the periodic BVP by Knobloch in [22], under the same geometric conditions on the vector field.
In the Seventies, from these results a line of research initiated and received a great interest, as shown by a series of works [4, 5, 29, 49]. In particular, assumptions (1.2) and (1.3), referring to the boundary of a ball, were extended to a broader families of domains. With this respect, very general conditions were proposed by Mawhin [29] and Bebernes [4], by introducing the concept of “bounding function”, namely a Lyapunov-like function of class which plays the same role as , so that (1.2) can be written as
for every with | (1.4) | |||
and such that , |
where we have denoted with the gradient of and by its Hessian matrix. In this case, the open ball is replaced by the sublevel set .
Assumptions of the form (1.2) or (1.4) (typically with the strict inequality) imply that there are no solutions “tangent from the interior” to the boundary of a given open bounded set (i.e., an open ball or a sublevel set in the above examples), namely the vector field satisfies Ważewski-type conditions at the boundary (see [20], as well as [51] and the references therein). This in turn lead to the concept of bound set introduced in [14], which has found many relevant applications to first and second order differential systems. Extensions of these methods have been obtained by several authors in various directions such as Floquet BVPs, Carathéodory conditions, non-smooth bounding functions and non-smooth boundaries, differential inclusions, extensions to Banach spaces, PDEs, integro-differential equations, differential equations with impulses, nonlocal BVPs. A constantly growing literature in this area shows the persistent influence of the bound sets and bounding functions approach; see [2, 3, 6, 7, 37, 38, 39, 44] and the references therein.
The aim of our work is to propose further applications of the bound set technique to the study of the periodic BVP associated with the second-order -Laplacian differential operator
(1.5) |
where is a homeomorphism such that , and is a continuous vector field which is -periodic in the -variable. The study of nonlinear scalar differential equations with a -Laplacian operator, or more generally of the form
for in a domain of , is a classical topic in the area of PDEs, see [50] (cf. also [46] and the references therein). In the setting of ODEs differential systems, a systematic investigation of the periodic problem was initiated by Manásevich and Mawhin in [26]. It is noteworthy that the analysis of systems has an independent interest also from the point of view of the analysis of differential equations in the complex plane or in the complex space (cf. [35]).
Our contribution pursues a line of research started by Mawhin [34] and by Mawhin and Ureña [41], dealing with the -Laplacian operator with , with the aim of further extending Hartman–Knobloch theorem to a broader class of nonlinear differential operators. A partial result in this direction was obtained in [10, Theorem 5.2] when the domain is a ball. Here, we extend also this result to more general domains.
The plan of the paper is as follows. In Section 2, we introduce the main theoretical tools for the solvability of the periodic BVP associated with (1.5), which are based on the Manásevich–Mawhin continuation theorems [26] in the versions stated in [10] for a general homeomorphism . This leads to Theorem 2.3 and Theorem 2.4, after the introduction of the concept of bound sets and Nagumo–Hartman condition for equation (1.5). As a next step, following [14, 29, 30, 31], we introduce a family of bounding functions and obtain the corresponding Theorem 2.5 and Theorem 2.6. Section 3 is devoted to the application of these latter results. In particular, for a set and the equation
we find the existence of a -periodic solution with values in , provided that
-
, for every ;
-
, for every and with ;
-
, for every and .
The class of -Laplacians for which the above result holds require that , , with a positive scalar function. As shown in [10], maps of this form are not necessarily monotone and include all the -Laplacian vector differential operators, thus extending Hartman–Knobloch theorem for frictionless vector fields along the line of [34]. The case of a vector field depending on is studied in Section 3.2 extending [29, Corollary 6.3] for vector Rayleigh systems to the -Laplacians. Another result for vector -Laplacian Liénard systems is given in Section 3.3 extending [45, Theorem 3.3] as well as a classical theorem of Reissig [47, Theorem 3].
We observe that conditions and appear explicitly or implicitly in some classical results for equation (1.1), see for instance [4, 17, 29]. Indeed, consider condition (1.4) already assumed in the above quoted papers in the special case when . From
(1.6) |
(with ), we obtain that the quadratic form is bounded from below and hence positive semi-definite, so that holds. Notice also that follows reading (1.4) for . Conditions and hint some kind of convexity properties for the sublevel sets of , however, at the best of our knowledge, there are no explicit proofs relating these conditions with the convexity of the set . In general, the convexity of the sublevel set is not guaranteed by the sole conditions and , as it could be shown by an example of a sublevel set made by the union of two disjoint balls. On the other hand, if we assume that is connected, then the convexity is achieved thanks to an analysis due to Krantz and Parks in [24] (see also [23]). For the reader’s convenience, we propose a different proof of this result in Appendix A. We exploit this characterization of the convexity (see Lemma 3.1) in the proof of Corollary 3.2. Clearly, for a general depending on , or if we have some information on the a priori bounds on (say ), we are not allowed to conclude with the positive semi-definiteness of from (1.6) because only the with will be involved in (1.6). An investigation of non-convex bound sets was proposed in [1] (see also [58] for some remarks about this problem).
Throughout the paper, we denote by the Euclidean norm in and by the associated standard inner product.
2. Main results
We consider the vector differential equation
(2.1) |
where is a homeomorphism such that , and is a continuous vector field which is -periodic in the -variable.
We study the problem of the existence of -periodic solutions for (2.1), which, equivalently, can be reduced to the search of solutions of (2.1) satisfying the boundary condition
(2.2) |
By a solution of (2.1) we mean a continuously differentiable function with continuously differentiable and satisfying (2.1) for all . With this respect, it is useful to introduce the space of the continuously differentiable functions satisfying the boundary condition (2.2), endowed with the -norm
where, for a continuous function , we denote
the classical -norm.
Equivalently, instead of (2.1), we can consider the differential system in
(2.3) |
and look for a (continuously differentiable) solution . Every -periodic solution of (2.1) corresponds to the -periodic solution of (2.3) with . Note that a solution of (2.1) defined on is bounded in the -norm if and only if the corresponding solution of (2.3) is bounded in the uniform norm of .
For the existence of -periodic solutions of (2.1), we apply the Manásevich–Mawhin continuation theorems [26] in the versions elaborated in [10] where weaker conditions on are assumed.
Theorem 2.1.
Theorem 2.1 corresponds to [26, Theorem 4.1] (see also [10, Theorem 4.6]). We recall also another version which is strictly related to the classical Mawhin continuation theorem [27, 33], in an appropriate form for the -Laplacian operators given in [26, Theorem 3.1] (see also [10, Theorem 3.11]).
Theorem 2.2.
We propose now some applications of the above continuation theorems which are motivated by the theory of bound sets (cf. [14, 30, 31]). We shall focus our attention to the case of Theorem 2.1. Similar applications can be given starting from Theorem 2.2.
The bound set approach represents a general method to verify the abstract condition of the non-existence of solutions in , given in (respectively ), by introducing a more concrete condition of the non-existence of solutions tangent to the boundary of an open and bounded set . In this setting, we also introduce some Nagumo–Hartman conditions (cf. [18, 41]), which are classical in this framework (cf. [32]) and allow to find a priori bounds for .
Let be an open and bounded set. Following [29] (see also [58]), we say that the system
is a Nagumo equation with respect to (with constant ) if there exists a constant such that, for every and for every solution of problem , with for all , it holds that .
If the Nagumo condition is satisfied, from Theorem 2.1 we deduce the next result.
Theorem 2.3.
Let be a continuous function such that
where is an autonomous vector field. Suppose that there exists an open bounded set such that
-
the system is a Nagumo equation with respect to ;
-
for each there is no solution of such that for all and for some ;
-
.
Then, problem (2.1)-(2.2) has at least a solution such that , for all .
In the sequel, we refer to as a bound set condition. As previously observed, this condition of non-tangency of the solutions at the boundary of replaces condition of Theorem 2.1.
Proof.
Our argument borrows the classical scheme in [29]. We are going to apply Theorem 2.1. According to the Nagumo condition , there exists a constant such that all the solutions of with values in have the derivative bounded by . Hence, we define the set of functions
which is open and bounded in (cf. [14, 31]). In order to check that condition holds, observe that if (by contradiction) a solution of satisfies , then , hence for all . Then, by the Nagumo condition, . By the definition of and the assumption , we cannot have for all and therefore there exists such that . This situation is not possible in view of the bound set condition .
To check we simply observe that and thus apply . This concludes the proof. ∎
The concept of Nagumo equation was introduced by Mawhin in [29] and further developed in [32] as a generalization of the classical Nagumo–Hartman condition [18, 19]. This latter condition was originally expressed as a growth restriction on the vector field in order to provide an a priori bound on for the solutions of the second-order vector differential equation
(2.4) |
with uniformly bounded.
For the periodic boundary value problem associated with (2.4), the Nagumo–Hartman condition reads as follows: given ,
-
there exists a continuous function such that
-
if , there are non-negative constants such that
Under these assumptions it holds that for every there exists a constant (depending on ) such that every -periodic solution of (2.4) satisfying is such that (cf. [4, 22, 29]). Observe that condition implies that the solution satisfies the constraint for all .
In the setting of Theorem 2.3, for the special case , the condition should be applied with in place of (and uniformly with respect to ).
As observed in [29, Proposition 5.2], the concept of Nagumo equation is more general as it covers some second-order differential systems for which - are not satisfied. Extensions of the Nagumo–Hartman conditions to more general differential operators have been obtained in more recent years, see, for instance, [41] dealing with the vector -Laplacian, and [40] for more general scalar nonlinear differential operators. In previous works, it has been provided precise growth assumptions for the vector field , generalizing to the -Laplacian type operator the classical conditions -, for the second order linear differential operator (cf. [41, conditions -]). Typically these modified Hartman–Nagumo conditions involve a growth assumption on with respect to the -variable which is related to the exponent in the -Laplacian operator. For a general -Laplacian the situation appears more complicated. Indeed, the following example shows that, for any homeomorphism of the real line having a power-growth at infinity, we can determine a suitable growth-rate in such that the Hartman–Nagumo condition is not satisfied.
Example 2.1.
Let be an increasing homeomorphism such that . Suppose also that , for some . Then, the differential equation
(2.5) |
has bounded solutions , with as well as unbounded. Indeed, setting we find that solves the equation with . Next, we obtain and thus
is a solution of (2.5) such that , , with bounded in and for . The example can be modified in order to consider the case of periodic solutions as well. Clearly, if at , then taking we provide an example where there are bounded solutions with unbounded derivatives.
Due to the great generality of our differential operator , we prefer to not propose a specific growth assumption, like in the above quoted papers. On the other hand, in Section 3, we provide some specific application where conditions - can be checked by a direct inspection.
A slightly variant of Theorem 2.3 can be obtained using Theorem 2.2 and referring the concept of Nagumo equation to the system
Theorem 2.4.
In the sequel, the following lemma will be used as a technical step to provide the desired bounds on in the context of condition . We denote by the unit sphere in .
Lemma 2.1.
Let be a family of absolutely continuous and -periodic functions such that
-
•
there exists such that for all and there exists such that ;
-
•
there exist and such that for every .
Then, there exists such that for every .
Proof.
Let with . Let be such that . We set and as in the first hypothesis. Therefore, we have
Setting (which does not depend on ), the conclusion holds. ∎
We discuss now a technique introduced and developed in [14, 29, 30, 31] to verify the bound set condition. It consists in controlling locally the solutions of at the boundary of , by means of suitable Lyapunov-like functionals which are usually called bounding functions (see also [30] for an introduction to this topic). To this end we give the following definition (see also [11, 57, 58]).
Definition 2.1.
Let be an open and bounded set. Assume that for each , there exist an open ball of center and radius and a function such that and
In this case, the family is called a set of bounding functions for .
We are in position to present an application of the method of bounding functions to the periodic problem associated with (2.1) for a homeomorphism having the following form
(2.6) |
where is a continuous function. As shown in [10] this case includes most nonlinear differential operators considered in the literature, in particular the vector -Laplacians. Moreover, observe also that (2.6) does not imply that the operator is monotone (cf. [10, Section 5]).
The following result holds.
Lemma 2.2.
Let be a homeomorphism of of the form (2.6). Let be a continuous function and let be a family of bounding functions of class for an open bounded set . Suppose that
-
for every , and
Then, the bound set condition holds with respect to problem .
Proof.
By contradiction, let us suppose that is not valid. Therefore, for some there exists a solution of such that for all and for some . Due to the boundary condition , without loss of generality, we can suppose . We consider the point , the ball and the function according to Definition 2.1. Thus, there exists an open neighborhood of such that for all and , so that is a point of maximum for the function , . By the chain rule, we have for all . If , then and therefore
(2.7) |
On the other hand, if then both and are maximum points for the differentiable function which now is defined on a neighborhood of both points. Since and for in a right neighborhood of and in a left neighborhood of , then and, using the boundary condition , we obtain again (2.7).
As a next step, we introduce the auxiliary function
Observe that . This is trivial if , because . If , then , by (2.7). By differentiating in a neighborhood of we obtain
Hence, from and (2.7) we find that . This implies that there exists an interval such that for all . From the definition of and , we must have for all and hence
Using the fact that for all , we obtain that the map
is strictly positive on . We conclude that is strictly increasing on a right neighborhood of and thus for all . This contradicts the fact that is a maximum point for . The proof is thus complete. ∎
At last, by combining Theorem 2.3 with Lemma 2.2, we are in position to state our main result which combines the abstract continuation theorem with the bounding function technique.
Theorem 2.5.
Let be a continuous function such that
where is an autonomous vector field. Suppose that there exists an open bounded set satisfying condition . Assume that admits a family of bounding functions of class satisfying . Assume also . Then, problem (2.1)-(2.2) has at least a solution such that , for all .
Using the continuation Theorem 2.2 we can provide a variant of Theorem 2.5 which reads as follows. Clearly the next result is strongly connected with the classical one in [29] for .
Theorem 2.6.
Variants of the results presented in this section could be provided for the more general case of a vector field satisfying the Carathéodory conditions. For the linear differential operator , this approach has been already considered in [40] (see also [2, 3, 52, 53]). In the Carathéodory case, pointwise estimates must be replaced by suitable control of the solutions in a neighborhood of the boundary of the set . For simplicity in the exposition, we will not investigate this framework.
3. Applications
In this section, we present some applications of our main results contained in Section 2. In Section 3.1 we consider frictionless vector fields , while in Section 3.2 and Section 3.3 we analyse more general situations: a -Laplacian Rayleigh-type equation and a -Laplacian Liénard equation, respectively.
3.1. The case
We start with the simplest case , namely, we consider the periodic boundary value problem
(3.1) |
where is an homeomorphism of the form (2.6) (considered in the previous section) and is a continuous function.
In our first result, we apply the concept of field of outer normals at the boundary of a convex body. The use of outer normals in the theory of bound sets is classical and dates back to the Seventies (cf. [4, 16]). We refer to [37, 38, 39] for a recent renewal of considerations of this technique for boundary value problems of non-local type and also for the rich bibliography.
Recall that if is a convex body (i.e. the closure of an open bounded convex set of ) then for each there exists a normal cone such that for each we have that
A family of vectors is called a field of outer normals for if for each (cf. [12]). We notice that if then we also have for all with .
The following result holds.
Corollary 3.1.
Proof.
We fix a point and observe that from basic properties of convex sets (cf. [11, p. 107] and [25]) we have for all . Thus, we define
so that and . In this manner, is clearly satisfied as
To prove condition we define and observe that and . Hence
for every , and ; then holds.
To conclude the proof we have still to check , that is (3.1) is a Nagumo equation with respect to . To this end, let and be a solution of with values in the bounded set . Then, is uniformly bounded and consequently is bounded as well. Hence, there are two positive constants such that every solution of satisfies
Let , and consider the auxiliary function , . By Rolle’s theorem there exists such that . Therefore, for the conditions in Lemma 2.1 are satisfied with , , . We conclude hat there exists a constant such that
holds for every possible solution of . As a consequence of the above inequality, we get that , and so , is uniformly bounded. This proves Finally, we conclude with an application of Theorem 2.5. ∎
In the next result the domain is the sublevel set of a functional. Preliminarily, we make some remarks.
For given function and , we denote by the sublevel set and similarly and . Assume that is of class and that, for some , the set is nonempty and, moreover,
(3.2) |
In this case, it turns out that is a regularly closed set, namely for we have and, moreover, as well as .
To prove this claim, we observe that in any case, we have and . On the other hand, if for every we consider the scalar function we find that and so that in any neighborhood of there are points of from which we obtain that . Finally, from we also conclude that and thus the claim is proved.
Generally speaking, the condition , for all , is not enough to guarantee the same property, as shown by the example , , and . In this case, .
The next lemma, which is borrowed and adapted from [24] (see also the survey [23]), is useful in this context as it allows to study the case of sublevel domains with a minimal set of assumptions.
Lemma 3.1.
Let be a -function satisfying (3.2) for some with nonempty, bounded and connected. Then, (as well as ) is convex if and only if
-
for all , , for all with .
Lemma 3.1 can be proved by combining the results in [24, ch. 6, pp. 195–196] or [23]. In Appendix A we provide a sketch of the proof. Clearly it is not restrictive to consider the case , as done in the sequel.
The following result provides a variant of Corollary 3.1 for convex sets with smooth boundary, observing that the involved domain is a sublevel set (see [4, 29] for classical results in this direction for the linear differential operator).
Corollary 3.2.
Proof.
We define and therefore . By the preliminary observation we also have and . Let be a connected component of . Setting for all , we have that is a family of bounding functions for . Next, we consider the homotopy
and observe that
for every , , , and such that . Therefore, condition holds. Next, from Lemma 3.1 we observe that is convex and therefore
(3.3) |
To prove the above formula, let us fix a point . We claim that
(3.4) |
Indeed, setting for , we have that and for all . Hence, . From (3.4), via the convex homotopy we find that
An alternative manner to prove (3.3) is to apply [12, Theorem 3] to the vector field on the convex set .
Finally, we conclude with an application of Theorem 2.5 which implies the existence of a solution in and hence in . ∎
Theorem 3.1 (Hartman–Knobloch).
Proof.
We can further state another straightforward consequence of Corollary 3.1.
Theorem 3.2 (Poincaré–Miranda).
As is well known, the inequalities in (3.5) can be reversed for first order systems; on the other hand, for second-order systems (as in our case) reversing the inequalities is not possible unless further growth (non-resonance) assumptions on the vector field are imposed, as one can see from the trivial example (see, for instance, [28] for a result in this direction).
It is worth noticing that Theorem 3.2 is related to the theory of (well ordered) lower and upper solutions for a differential system in , as it represents the case of vector valued lower solution and vector valued upper solution with and . In order to deal with non-constant lower and upper solutions in , one could adapt to our case the technique of “curvature bound sets” in [13, 14] or “non autonomous bounding Lyapunov functions” in [4]; see also [17] for a similar approach.
3.2. A -Laplacian Rayleigh-type equation
Up to now we have considered only a trivial case of the Nagumo–Hartman condition, namely when the vector field does not depend on . In the next example, inspired by a case studied in [29], we treat a more general situation.
Let us consider the following -periodic boundary value problem associated with a -Laplacian Rayleigh-type equation
(3.6) |
which is a natural generalisation of the system considered in [29]. In this framework, we state the following result, which is in the spirit of Corollary 3.2. It can be seen as an extension of [29, Corollary 6.3] (for ).
Theorem 3.3.
Let be a homeomorphism of the form (2.6) and let be a continuous function. Let be continuous and such that has either the same or the opposite direction of . Assume also that there exists of class such that
-
•
is bounded.
Let be a -function such that for all . Let be nonempty and bounded. Suppose that for every and
-
•
;
-
•
, for all with .
Then, problem (3.6) has at least a solution with values in .
Proof.
The proof follows the same steps of Corollary 3.2. Accordingly we just point out the main modifications which are needed. Having defined and as above, we consider the homotopy
For every , , , and such that , it holds that
where we have used the fact that is parallel to . Therefore, condition holds. The verification of the degree conditions is the same as in the proof of Corollary 3.2 using the convexity of .
In order to conclude the proof, we need only to check that the Nagumo condition is satisfied with respect to the set . To this end, let and be a solution of with values in the bounded set . Then, is uniformly bounded and consequently
is uniformly bounded too, by a constant . Consider now the system
From this we have
where is a constant which bounds . An integration on yields to
and thus a bound for in is achieved. An application of Lemma 2.1 for , arguing as in the proof of Corollary 3.1, ensures an a priori bound for . Finally, Theorem 2.5 implies the existence of a solution in and hence in . ∎
As in Section 3.1 we have the following straightforward corollaries.
Corollary 3.3 (Hartman–Knobloch).
3.3. A -Laplacian Liénard equation
All the preceding examined examples depend on the fact that first we find a set with no solutions tangent to the boundary from the interior, and next we provide some a priori bound on using Nagumo-type conditions. There are however some situations in which the particular form of the equations allows to find a priori bounds on (for some ) independently on . In such cases, the set can be found using some “sign-conditions” on the nonlinearity. As a possible example in this direction and our third application, we deal with the following -periodic boundary value problem associated with a -Laplacian Liénard equation
(3.7) |
In this setting, we can state the following.
Theorem 3.4.
Let be a homeomorphism such that
-
for every , and for every there exists such that , for all .
Let be a -function. Let be a continuous function such that
-
there exists such that , for every and with .
Then, problem (3.7) has at least a solution.
Proof.
We aim to apply Theorem 2.1. Accordingly, we introduce the parameter-dependent problem
(3.8) |
where
(3.9) |
We divide the proof in some steps.
Step 1. A priori bound of . Let be a solution of problem (3.12). By integrating the scalar product between and the equation in (3.8), we have
(3.10) |
Next, we observe that condition implies the existence of such that , for every and . Therefore, from (3.10), we deduce that
We fix and, by hypothesis , we obtain that
which is the desired bound.
Step 2. A priori bound of . Let be a solution of problem (3.12). First, we prove that there exists such that . Indeed, if it is not true, for all and from (3.10) (and since ), we deduce that
a contradiction. Consequently, for every , we immediately obtain
and thus , as desired. As a consequence, the open ball is (trivially) a bound set for system (3.8).
Step 3. Conclusion. From
we have that is bounded in . Therefore, an application of Lemma 2.1 for , arguing as in the proof of Corollary 3.1, ensures an a priori bound for .
For , we have . Therefore, the thesis follows from Theorem 2.3. ∎
Remark 3.1.
Our result is related to a classical theorem by Reissig [47, Theorem 3] for the classical scalar generalized Liénard equation with a periodic forcing term
where the existence of a -periodic solution is proved by assuming for every with and . No special assumption besides continuity on is considered. Extensions of this and other related results for a -Laplacian differential operator with or without singularity (including the Minkowski operator for the relativistic acceleration) have been obtained in [8, 36] in the scalar case. We show now how the proof of Theorem 3.4 can be easily adapted to treat the case of a periodic forcing term with zero mean value, namely we deal with
(3.11) |
Theorem 3.5.
Let be a homeomorphism satisfying . Let be a -function. Let be a continuous function with . Let be a continuous function such that
-
there exists such that , for all .
Moreover, suppose that at least one of the following three condition holds:
-
as uniformly in ;
-
for every there exists such that , for all with and ;
-
there exists such that, for every , , for every with .
Then, problem (3.11) has at least a solution.
Proof.
The proof is similar to the one of Theorem 3.4 and thus we only focus on the main modifications requested. We introduce the parameter-dependent problem
(3.12) |
where is defined as in (3.9). Since , it is convenient to fix a -periodic continuously differentiable function such that . We divide the proof in some steps.
Step 1. A priori bound of . Let be a solution of problem (3.12). By integrating the scalar product between and the equation in (3.12), we have
(3.13) | ||||
Next, from hypothesis and the fact that , we deduce that
We fix and, by hypothesis , we obtain that
and so
which is the desired bound.
Step 2. A priori bound of . Without loss of generality, we can consider ; indeed, an easy computation shows that for the only -periodic solution is the trivial one. Let be a solution of problem (3.12).
Assume condition . From (3.13) we have
and so, dividing by ,
Let . By there exists such that for all such that . We immediately conclude that there exists such that , otherwise a contradiction can be easily obtained. Then, arguing as in the proof of Theorem 3.4, we have that .
Assume now condition or condition . Let us also denote by the mean value of . Let (arbitrary). An integration of the scalar product between and the equation in (3.12) gives
which equivalently reads as
(3.14) |
Let hold. If , we set and so . Then, there exists such that . On the other hand, we have that is bounded. Indeed, let be such that , thus
and so . Therefore, from and , we have that there exists such that . Clearly, the same inequality holds if . Next, it is easy to prove that
and thus . Let hold. We prove that for every there exists such that . Indeed, let . From (3.14), we have
and, proceeding by contradiction, we easily reach the claim. Then, arguing as above, we have that .
As a consequence, the open ball is (trivially) a bound set for system (3.8).
Step 3. Conclusion. From
we have that is bounded in and we conclude as above. ∎
Remark 3.2.
Theorem 3.5 in the variant provides an extension of [45, Theorem 3.3], where the result has been proved for and a conservative vector field satisfying the more restrictive condition for every , with , and .
Conditions and are classical ones in this context for vector second-order systems of Liénard-type (cf. [15, 28, 43, 56]). Observe that both conditions are satisfied in the one-dimensional case if we assume the sign-condition for every and with . Thus, Theorem 3.5 extends to a large class of -Laplacian differential systems the classical theorem of Reissig mentioned above.
Remark 3.3.
Some comments on the hypotheses of Theorem 3.5 are in order. First of all, we notice that, in dimension , condition implies the validity of and , which are equivalent each other. On the other hand, the function
(3.15) |
satisfies and , but not . In this case, for every and thus follows.
We focus on the case . In order to show that , , are independent, we are going to present three examples for which exactly one of the hypotheses is valid. Every example is given for , however it can be easily generalized to treat the -dimensional case for a general . Moreover, in each example condition can be straightforwardly checked.
Example 1. Let . Let us consider the function
We observe that
which tends to as . Therefore, holds. Moreover, we have
Let , , and . Therefore,
Hence, does not hold. Let again , , then for sufficiently large. Therefore, does not hold.
Example 2. Let us consider the function
where is defined as in (3.15). We observe that
which tends to as . Therefore, does not hold. Moreover,
whenever and is bounded. Therefore, holds.
For , , we have . Hence, does not hold.
Example 3. Let us consider the function
Let , . The quantity
tends to as . Therefore, and (with ) do not hold. For every , we notice that
for every with . Therefore, holds.
Remark 3.4.
We observe that one could add to the list of hypotheses – in Theorem 3.5 another classical condition, that is the generalized Villari condition (cf. [26, p. 381]), which, in our context, reads as follows:
-
•
there exists such that for some , for every with for some .
It is worth noticing that condition with the strict inequalities is a special case of the generalized Villari condition. On the other hand, in Example 2 and Example 3 of Remark 3.3, if we consider the maps multiplied by a scalar function which vanishes outside a large open ball, the examples continue to satisfy and, respectively, , but the generalized Villari condition does not hold.
Appendix A Remarks on the convexity of a sublevel set
In this appendix, we propose an alternative proof of Lemma 3.1 based on a classical result about convex sets, namely the Tietze–Nakajima theorem [42, 54], that we recall for reader’s convenience (see also [21, 48] for more general versions of the result). A set is locally convex if each point of has a neighborhood whose intersection with is convex (cf. [25, Section 17]). Then, the following result holds (cf. [25, 55] for the proof).
Theorem A.1 (Tietze–Nakajima).
A closed connected locally convex set in a Euclidean space is convex.
We now give the proof of Lemma 3.1.
Proof of Lemma 3.1.
Let be a -function satisfying (3.2) for some with nonempty, bounded and connected. Let us assume that
-
for all , , for all with .
holds and we will prove that is convex. By Tietze–Nakajima theorem it is sufficient to verify that is locally convex. Accordingly, let be an arbitrary but fixed point in and we aim to prove that there exists a neighborhood of such that is convex. Recalling that by assumption (3.2), we define the vector . Let , that is the subspace (of dimension ) of orthogonal to the vector . Since every can be uniquely written as for , we can equivalently consider the function
Thus, our goal reduces to prove that there exists a neighborhood of in the -space such that is convex, where
We observe that and thus, by the implicit function theorem, there exist a neighborhood of of the form and a continuously differentiable map such that
(A.1) |
for every . Here, is an open neighborhood of the origin in . Without loss of generality, we assume also that for all . Hence, it will be sufficient to verify that the set
is convex. Let and set if and arbitrary with otherwise. Then, to check the (local) convexity we can restrict ourselves to the intersection of with the -dimensional subspace of spanned by and . We set
Then, for every , we have
where we have set
Moreover, for every , it holds that
By hypothesis we conclude that for all , which implies that the subgraph of restricted to is convex.
Conversely, assume that is convex. If condition is not satisfied there exists such that for some vector with and . Arguing as in the previous part of the proof, we introduce the vector and, using the implicit function theorem we find a neighborhood of of the form and a continuously differentiable map such that (A.1) holds for every . As a next step, we consider the intersection of with the -dimensional subspace of spanned by and . By the convexity of it follows that the set
is convex. On the other hand, setting as above and , we obtain
Hence, for , we have that , which implies , and that , which implies . Therefore, the map is strictly convex in a neighborhood of the origin for a suitable . This clearly contradicts the convexity of the set . The proof is completed. ∎
References
- [1] P. Amster, J. Haddad, A Hartman-Nagumo type condition for a class of contractible domains, Topol. Methods Nonlinear Anal. 41 (2013) 287–304.
- [2] J. Andres, M. Kožušníková, L. Malaguti, Bound sets approach to boundary value problems for vector second-order differential inclusions, Nonlinear Anal. 71 (2009) 28–44.
- [3] J. Andres, L. Malaguti, M. Pavlačková, Strictly localized bounding functions for vector second-order boundary value problems, Nonlinear Anal. 71 (2009) 6019–6028.
- [4] J. W. Bebernes, A simple alternative problem for finding periodic solutions of second order ordinary differential systems, Proc. Amer. Math. Soc. 42 (1974) 121–127.
- [5] J. W. Bebernes, K. Schmitt, Periodic boundary value problems for systems of second order differential equations, J. Differential Equations 13 (1973) 32–47.
- [6] I. Benedetti, N. V. Loi, L. Malaguti, V. Taddei, Nonlocal diffusion second order partial differential equations, J. Differential Equations 262 (2017) 1499–1523.
- [7] I. Benedetti, L. Malaguti, V. Taddei, Nonlocal solutions of parabolic equations with strongly elliptic differential operators, J. Math. Anal. Appl. 473 (2019) 421–443.
- [8] C. Bereanu, J. Mawhin, Existence and multiplicity results for some nonlinear problems with singular -Laplacian, J. Differential Equations 243 (2007) 536–557.
- [9] C. Fabry, Nagumo conditions for systems of second-order differential equations, J. Math. Anal. Appl. 107 (1985) 132–143.
- [10] G. Feltrin, F. Zanolin, An application of coincidence degree theory to cyclic feedback type systems associated with nonlinear differential operators, Topol. Methods Nonlinear Anal. 50 (2017) 683–726.
- [11] M. L. C. Fernandes, F. Zanolin, Repelling conditions for boundary sets using Liapunov-like functions. I. Flow-invariance, terminal value problem and weak persistence, Rend. Sem. Mat. Univ. Padova 80 (1988) 95–116.
- [12] A. Fonda, P. Gidoni, Generalizing the Poincaré-Miranda theorem: the avoiding cones condition, Ann. Mat. Pura Appl. 195 (2016) 1347–1371.
- [13] R. E. Gaines, J. Mawhin, Ordinary differential equations with nonlinear boundary conditions, J. Differential Equations 26 (1977) 200–222.
- [14] R. E. Gaines, J. L. Mawhin, Coincidence degree, and nonlinear differential equations, vol. 568 of Lecture Notes in Math., Springer-Verlag, Berlin-New York, 1977.
- [15] M. García-Huidobro, R. Manásevich, J. R. Ward, Periodic solutions and asymptotic behavior in Liénard systems with -Laplacian operators, Differential Integral Equations 22 (2009) 979–998.
- [16] G. B. Gustafson, K. Schmitt, A note on periodic solutions for delay-differential systems, Proc. Amer. Math. Soc. 42 (1974) 161–166.
- [17] P. Habets, K. Schmitt, Nonlinear boundary value problems for sytems of differential equations, Arch. Math. (Basel) 40 (1983) 441–446.
- [18] P. Hartman, On boundary value problems for systems of ordinary, nonlinear, second order differential equations, Trans. Amer. Math. Soc. 96 (1960) 493–509.
- [19] P. Hartman, Ordinary differential equations, John Wiley & Sons, Inc., New York-London-Sydney, 1964.
- [20] J. L. Kaplan, A. Lasota, J. A. Yorke, An application of the Waṡewski retract method to boundary value problems, Zeszyty Nauk. Uniw. Jagielloń. Prace Mat. (1974) 7–14.
- [21] V. L. Klee, Jr., Convex sets in linear spaces, Duke Math. J. 18 (1951) 443–466.
- [22] H.-W. Knobloch, On the existence of periodic solutions for second order vector differential equations, J. Differential Equations 9 (1971) 67–85.
- [23] S. G. Krantz, Convexity in real analysis, Real Anal. Exchange 36 (2010/11) 1–27.
- [24] S. G. Krantz, H. R. Parks, The geometry of domains in space, Birkhäuser Advanced Texts: Basler Lehrbücher., Birkhäuser Boston, Inc., Boston, MA, 1999.
- [25] S. R. Lay, Convex sets and their applications, Pure and Applied Mathematics, A Wiley-Interscience Publication, John Wiley & Sons, Inc., New York, 1982.
- [26] R. Manásevich, J. Mawhin, Periodic solutions for nonlinear systems with -Laplacian-like operators, J. Differential Equations 145 (1998) 367–393.
- [27] J. Mawhin, Équations intégrales et solutions périodiques des systèmes différentiels non linéaires, Acad. Roy. Belg. Bull. Cl. Sci. 55 (1969) 934–947.
- [28] J. Mawhin, An extension of a theorem of A. C. Lazer on forced nonlinear oscillations, J. Math. Anal. Appl. 40 (1972) 20–29.
- [29] J. Mawhin, Boundary value problems for nonlinear second-order vector differential equations, J. Differential Equations 16 (1974) 257–269.
- [30] J. Mawhin, Functional analysis and boundary value problems, in: Studies in ordinary differential equations, vol. 14 of Stud. in Math., Math. Assoc. of America, Washington, D.C., 1977, pp. 128–168.
- [31] J. Mawhin, Topological degree methods in nonlinear boundary value problems, vol. 40 of CBMS Regional Conference Series in Mathematics, American Mathematical Society, Providence, R.I., 1979.
- [32] J. Mawhin, The Bernstein-Nagumo problem and two-point boundary value problems for ordinary differential equations, in: Qualitative theory of differential equations, Vol. I, II (Szeged, 1979), vol. 30 of Colloq. Math. Soc. János Bolyai, North-Holland, Amsterdam-New York, 1981, pp. 709–740.
- [33] J. Mawhin, Topological degree and boundary value problems for nonlinear differential equations, in: Topological methods for ordinary differential equations (Montecatini Terme, 1991), vol. 1537 of Lecture Notes in Math., Springer, Berlin, 1993, pp. 74–142.
- [34] J. Mawhin, Some boundary value problems for Hartman-type perturbations of the ordinary vector -Laplacian, Nonlinear Anal. 40 (2000) 497–503.
- [35] J. Mawhin, Periodic solutions for quasilinear complex-valued differential systems involving singular -Laplacians, Rend. Istit. Mat. Univ. Trieste 44 (2012) 75–87.
- [36] J. Mawhin, Resonance problems for some non-autonomous ordinary differential equations, in: Stability and bifurcation theory for non-autonomous differential equations, vol. 2065 of Lecture Notes in Math., Springer, Heidelberg, 2013, pp. 103–184.
- [37] J. Mawhin, K. Szymańska-Dȩbowska, Convex sets and second order systems with nonlocal boundary conditions at resonance, Proc. Amer. Math. Soc. 145 (2017) 2023–2032.
- [38] J. Mawhin, K. Szymańska-Dȩbowska, Bound sets and two-point boundary value problems for second order differential systems, Math. Bohem. 144 (2019) 373–392.
- [39] J. Mawhin, K. Szymańska-Dȩbowska, Convexity, topology and nonlinear differential systems with nonlocal boundary conditions: a survey, Rend. Istit. Mat. Univ. Trieste 51 (2019) 125–166.
- [40] J. Mawhin, H. B. Thompson, Nagumo conditions and second-order quasilinear equations with compatible nonlinear functional boundary conditions, Rocky Mountain J. Math. 41 (2011) 573–596.
- [41] J. Mawhin, A. J. Ureña, A Hartman-Nagumo inequality for the vector ordinary -Laplacian and applications to nonlinear boundary value problems, J. Inequal. Appl. 7 (2002) 701–725.
- [42] S. Nakajima (Matsumura), Über konvexe Kurven und Flächen, Tohoku Math. J. 29 (1928) 227–230.
- [43] P. Omari, F. Zanolin, Periodic solutions of Liénard equations, Rend. Sem. Mat. Univ. Padova 72 (1984) 203–230.
- [44] M. Pavlačková, V. Taddei, Bounding function approach for impulsive Dirichlet problems with upper-Carathéodory right-hand side, Electron. J. Differential Equations (2019) Paper No. 12, 18 pp.
- [45] S. Peng, Z. Xu, On the existence of periodic solutions for a class of -Laplacian system, J. Math. Anal. Appl. 325 (2007) 166–174.
- [46] P. Pucci, J. Serrin, The maximum principle, vol. 73 of Progress in Nonlinear Differential Equations and their Applications, Birkhäuser Verlag, Basel, 2007.
- [47] R. Reissig, Extension of some results concerning the generalized Liénard equation, Ann. Mat. Pura Appl. 104 (1975) 269–281.
- [48] R. Sacksteder, E. G. Straus, F. A. Valentine, A generalization of a theorem of Tietze and Nakajima on local convexity, J. London Math. Soc. 36 (1961) 52–56.
- [49] K. Schmitt, Periodic solutions of systems of second-order differential equations, J. Differential Equations 11 (1972) 180–192.
- [50] J. Serrin, Local behavior of solutions of quasi-linear equations, Acta Math. 111 (1964) 247–302.
- [51] R. Srzednicki, Ważewski method and Conley index, in: Handbook of differential equations, Elsevier/North-Holland, Amsterdam, 2004, pp. 591–684.
- [52] V. Taddei, Two-points boundary value problems for Carathéodory second order equations, Arch. Math. (Brno) 44 (2008) 93–103.
- [53] V. Taddei, F. Zanolin, Bound sets and two-point boundary value problems for second order differential equations, Georgian Math. J. 14 (2007) 385–402.
- [54] H. Tietze, Über Konvexheit im kleinen und im großen und über gewisse den Punkten einer Menge zugeordnete Dimensionszahlen, Math. Z. 28 (1928) 697–707.
- [55] F. A. Valentine, Convex sets, McGraw-Hill Series in Higher Mathematics, McGraw-Hill Book Co., New York-Toronto-London, 1964.
- [56] F. Zanolin, On forced periodic oscillations in dissipative Liénard systems, Rend. Sem. Mat. Univ. Padova 69 (1983) 51–62.
- [57] F. Zanolin, Bound sets, periodic solutions and flow-invariance for ordinary differential equations in : some remarks, Rend. Istit. Mat. Univ. Trieste 19 (1987) 76–92.
- [58] F. Zanolin, On the periodic boundary value problem for forced nonlinear second order vector differential equations, Riv. Mat. Pura Appl. 1 (1987) 105–124.