A Nehari manifold method for nonvariational problems

Abstract

The aim of this paper is to extend the Nehari manifold method from the variational setting to the nonvariational framework of fixed point equations. This is achieved by constructing a radial energy functional that generalizes the standard one from the variational case. Furthermore, the solutions obtained through our method are localized in conical annular sets, which leads to the existence of multiple solutions. The abstract results are illustrated by two representative applications.

Authors

Radu Precup
Faculty of Mathematics and Computer Science and Institute of Advanced Studies in Science and Technology, Babes-Bolyai, University, Cluj-Napoca, Romania
Tiberiu Popoviciu Institute of Numerical Analysis, Romanian Academy, Romania

Andrei Stan
Tiberiu Popoviciu Institute of Numerical Analysis, Romanian Academy, Romania
Department of Mathematics, Babes-Bolyai University, Cluj-Napoca, Romania

Keywords

Critical point, Nehari manifold, Birkhoff-Kellogg invariant-direction, cone, p-Laplace operator, positive solution, multiple solutions

Paper coordinates

Precup R, Stan A, A Nehari manifold method for nonvariational problems, 2025. 

About this paper

Journal
Preprint
Publisher Name
Print ISSN
Online ISSN

google scholar link

Paper (preprint) in HTML form

A Nehari manifold method for nonvariational problems

Abstract

The aim of this paper is to extend the Nehari manifold method from the variational setting to the nonvariational framework of fixed point equations. This is achieved by constructing a radial energy functional that generalizes the standard one from the variational case. Furthermore, the solutions obtained through our method are localized in conical annular sets, which leads to the existence of multiple solutions. The abstract results are illustrated by two representative applications.

keywords:
Nehari manifold method, nonvariational problems, fixed point, radial energy, pp-Laplacian equation

1 Introduction and preliminaries

The Nehari manifold method is a well-known technique for finding nonzero critical points of functionals. Given a C1C^{1} functional EE defined on a Banach space X,X, the Nehari manifold is defined as:

𝒩={uX{0}:E(u),uX=0},\mathcal{N}=\left\{u\in X\setminus\left\{0\right\}:\ \left\langle E^{\prime}\left(u\right),u\right\rangle_{X}=0\right\},

where ,X\left\langle\cdot,\cdot\right\rangle_{X} denotes the duality between XX and its dual X.X^{\ast}. It is clear that any nonzero critical point of EE lies in 𝒩.\mathcal{N}. Therefore, one may expect that it is easier to find nonzero critical points by restricting the search to this smaller set, as fewer or weaker conditions may be required compared to working in the entire space X.X. This is the case, for example, with the Palais-Smale compactness condition. Once the Nehari manifold is defined, the method proceeds by minimizing EE on 𝒩.\mathcal{N}. This approach is effective when EE has a specific geometric structure, namely, for each direction vX,v\in X, |v|=1,\left|v\right|=1, the function

0<tE(tv)0<t\ \mapsto\ E\left(tv\right)

has a unique critical point at some tv>0.t_{v}>0. In this case, the derivative satisfies

ddtE(tv)|t=tv=E(tvv),vX=0,\frac{d}{dt}\left.E\left(tv\right)\right|_{t=t_{v}}=\left\langle E^{\prime}\left(t_{v}v\right),v\right\rangle_{X}=0,

so u=tvvu=t_{v}v\in 𝒩.\mathcal{N}. Conversely, if u𝒩,u\in\mathcal{N}, then

0=E(u),uX=E(|u|u|u|),u|u|X,0=\left\langle E^{\prime}\left(u\right),u\right\rangle_{X}=\left\langle E^{\prime}\left(\left|u\right|\frac{u}{\left|u\right|}\right),\frac{u}{\left|u\right|}\right\rangle_{X},

showing that t=|u|t=\left|u\right| is a critical point of the function E(tu|u|),E\left(t\frac{u}{\left|u\right|}\right), i.e., tu|u|=|u|.t_{\frac{u}{\left|u\right|}}=\left|u\right|. Thus, under this geometric condition, the Nehari manifold can be characterized as

𝒩={tvv:vX,|v|=1}.\mathcal{N}=\left\{t_{v}v:\ v\in X,\ \left|v\right|=1\right\}.

This method has received increasing attention in recent years and has been successfully applied to various types of problems [4, 18, 8, 5, 25]. In the recent papers [23, 20], this technique has been refined to locate critical points within cones, which is useful when searching for solutions with additional properties such as positivity. Also, in [21], the method has been combined with Schauder’s fixed point theorem to obtain a localized solution for a system.

Compared to other techniques in critical point theory—such as general linking methods, and particularly the mountain pass theorem—the Nehari manifold method essentially requires only a radial behavior of the energy functional. Motivated by this observation, in this paper we propose a similar approach within the framework of fixed point theory, specifically for equations of non-variational type. To extend the technique to a fixed point equation of the form u=T(u),u=T\left(u\right), we begin by introducing a three-variable functional (u,v,w)\mathcal{F}\left(u,v,w\right) which is used to define a radial energy type function \mathcal{E} in any direction v.v. That is, for any unit vector vv and t>0,t>0, we define:

(v)(t)=0t(T(sv),sv,v)𝑑s.\mathcal{E}\left(v\right)\left(t\right)=\int_{0}^{t}\mathcal{F}\left(T\left(sv\right),sv,v\right)ds.

For each direction v,v, the derivative of (v)\mathcal{E}\left(v\right) with respect to tt is given by

ddt(v)(t)=(T(tv),tv,v).\frac{d}{dt}\mathcal{E}\left(v\right)\left(t\right)=\mathcal{F}\left(T\left(tv\right),tv,v\right).

This expression suggests that, for some direction vv, the quantity

t(T(tv),tv,v)t\mapsto\mathcal{F}\left(T\left(tv\right),tv,v\right)

can be interpreted as the (radial) potential of the energy functional (v)\mathcal{E}(v) at the moment tt. This interpretation is natural since, as shown in the paper, in the particular case of equations possessing a variational structure, i.e., the operator TT is of potential type in the sense that T=IET=I-E^{\prime}, for some functional EE, then the value of (v)(t)\mathcal{E}(v)(t) corresponds to E(tv)E(tv). In this situation, the classical Nehari manifold 𝒩\mathcal{N} is recovered, and existing methods apply. We emphasize that even in this case, our results significantly refine the classical ones by localizing the critical point and by imposing a weaker condition on the radial geometry of the energy function, which, in our case, requires only a unique maximum point in each direction rather than a unique critical point which is a maximum.

Under suitable conditions on the operator TT and the functional \mathcal{F}, and through a surprising combination with the Birkhoff–Kellogg invariant-direction theorem, we develop a Nehari-type method for establishing the existence of fixed points of TT within a conical annular region of a Banach space.

To summarize, the novelties of this paper are the following:

  • (a):

    First and most importantly, we extend the Nehari manifold method from the variational setting to a nonvariational framework. This is achieved by constructing a radial energy that corresponds to the usual energy in the variational case.

  • (b):

    The solutions obtained by this method are localized in conical annular sets. By applying the method repeatedly on different annular sets, we obtain the existence of multiple solutions.

  • (c):

    For a given operator, we may construct multiple radial energies by choosing different radial potentials \mathcal{F}. Thus, even in the variational case, the energy corresponds to a particular choice of \mathcal{F}.

The structure of the paper is as follows. In Section 2 we present the main abstract result, introducing the functionals \mathcal{F} and \mathcal{E}, together with the associated Nehari manifold. Section 3 is devoted to applications: first, we discuss a nonvariational pp-Laplace problem illustrating our theoretical findings; second, we construct an example of \mathcal{F} and TT such that, for each direction vv, the associated radial energy functional (v)\mathcal{E}(v) has a unique maximum but two critical points, contrary to the common setting in the literature, to which, however, our result still applies.

We conclude this section with a series of basic well-known results from literature. The first result is Vitali’s theorem (see, e.g., [19, Lemma 9.1]) used in Lemma 6.

Theorem 1.

Let ΩN\Omega\subset\mathbb{R}^{N} be a bounded open set and let (uk)(u_{k}) be a sequence of functions from Lq(Ω;n)L^{q}(\Omega;\mathbb{R}^{n}) (1q<1\leq q<\infty) with uk(x)u(x)u_{k}(x)\to u(x) as kk\to\infty, for a.e. xΩx\in\Omega. Then uLq(Ω;n)u\in L^{q}(\Omega;\mathbb{R}^{n}) and ukuu_{k}\to u in Lq(Ω;n)L^{q}(\Omega;\mathbb{R}^{n}) as kk\to\infty, if and only if for each ε>0\varepsilon>0 there exists a δε>0\delta_{\varepsilon}>0 such that

D|uk|q𝑑x<ε\int_{D}|u_{k}|^{q}\,dx<\varepsilon

for all kk and every DΩD\subset\Omega with μ(D)<δε\mu(D)<\delta_{\varepsilon}.

The next two results are essential in Theorems 4 and 5, respectively. The first one is the version in cones due to Krasnoselskii and Ladyzenskii [16] (see also [11, p.139], [12]), of the classical theorem of Birkhoff and Kellogg invariant-direction theorem [2] (see also [11, Theorem 6.6]) regarding the existence of a ‘nonlinear’ eigenvalue and eigenvector for compact maps in Banach spaces.

Theorem 2 (Krasnoselskii and Ladyzenskii).

Let XX be a real Banach space, UXU\subset X be an open bounded set with 0U0\in U, KXK\subset X a cone, and T:KU¯KT:K\cap\overline{U}\rightarrow K a completely continuous operator. If

infxKU|T(x)|>0,\inf_{x\in K\cap\partial U}\left|T\left(x\right)\right|>0,

then, there exist λ0>0\lambda_{0}>0 and x0KUx_{0}\in K\cap\partial U such that

x0=λ0T(x0).x_{0}=\lambda_{0}T\left(x_{0}\right).

The second result is the well-known ”nonlinear alternative” from compact mappings (see, e.g., [11, Theorem 5.2]).

Theorem 3 (Nonlinear alternative).

Let XX be a real Banach space, KXK\subset X a convex set, and UKU\subset K an open bounded set with 0U0\in U. Suppose that T:U¯KT:\overline{U}\to K is a completely continuous operator. Then at least one of the following two alternatives holds:

  1. [(a)]

  2. 1.

    TT has a fixed point in U¯\overline{U},

  3. 2.

    there exist xUx\in\partial U and λ(0,1)\lambda\in(0,1) such that

    x=λT(x).x=\lambda T(x).

2 Main results

2.1 Existence, localization and multiplicity in conical annular sets

Let (X,||)\left(X,\lvert\cdot\rvert\right) be a Banach space, KXK\subset X be a nondegenerate cone (also called a wedge), i.e., a closed, convex set such that λKK\lambda K\subset K for all λ+\lambda\in\mathbb{R}_{+}, and K{0}K\setminus\{0\}\neq\emptyset. Note that, in particular, one may take K=XK=X. For any ρ>0,\rho>0, denote

Kρ:={vK:|v|=ρ}.K_{\rho}:=\left\{v\in K:~\left|v\right|=\rho\right\}.

Let

T:KKand :K×K×K1T\colon K\rightarrow K\ \ \ \text{and\ \ \ }\mathcal{F}:K\times K\times K_{1}\rightarrow\mathbb{R}

be a completely continuous operator and a continuous functional, respectively.

For each vK1,v\in K_{1}, we associate to \mathcal{F} and TT the ”radial energy functional” in direction v,v,

(v):(0,),(v)(t)=0t(T(sv),sv,v)𝑑s.\mathcal{E}(v):(0,\infty)\rightarrow\mathbb{R},\ \ \ \mathcal{E}(v)(t)=\int_{0}^{t}\mathcal{F}(T(sv),sv,v)ds.

It is straightforward to verify that

ddt(v)(t)=(T(tv),tv,v).\frac{d}{dt}\mathcal{E}(v)(t)=\mathcal{F}(T(tv),tv,v).

Let r,Rr,R be two real numbers such that 0r<R0\leq r<R\leq\infty, and define

KrR:={uK:r|u|R}.K_{rR}:=\left\{u\in K\,:\,r\leq|u|\leq R\right\}.

Clearly, if r=0r=0, then

K0R={uK:|u|R};K_{0R}=\left\{u\in K\,:\,|u|\leq R\right\};

if R=R=\infty, then

Kr={uK:|u|r};K_{r\infty}=\left\{u\in K\,:\,|u|\geq r\right\};

and if 0=r,R=0=r,\,R=\infty, then

K0=K.K_{0\infty}=K.

Our main assumption on the functional \mathcal{E} is the following:

(h1)

There exist r0>0r_{0}>0 and R0<R_{0}<\infty with rr0<R0Rr\leq r_{0}<R_{0}\leq R such that, for every vK1v\in K_{1}, the mapping (v)\mathcal{E}(v) attains a unique maximum tvt_{v} on the interval [r,R][r,R], and this maximum satisfies r0<tv<R0r_{0}<t_{v}<R_{0}.

In particular, one may set

r0=rif r>0and R0=Rif R<.r_{0}=r\ \ \text{if }r>0\ \ \ \text{and\ \ \ }R_{0}=R\ \ \text{if }R<\infty.
Lemma 1.

Under the assumption (h1), the mapping

vK1tvv\in K_{1}\mapsto t_{v}

is continuous.

Proof.

Let vkK1v_{k}\in K_{1} with vkv0v_{k}\to v^{0} as kk\to\infty. Clearly, v0K1v^{0}\in K_{1}, and by assumption (h1), the sequence (tvk)\left(t_{v_{k}}\right) is bounded. In order to prove that tvktv0t_{v_{k}}\to t_{v^{0}}, it suffices to show that any convergent subsequence of (tvk)\left(t_{v_{k}}\right) has limit tv0t_{v^{0}} (see, e.g., [17, Lemma 1.1]). Thus, let (tvk)\left(t_{v_{k}}\right) be a subsequence converging to some t0[r0,R0]t^{0}\in[r_{0},R_{0}]. Using (h1) we have that

(vk)(tvk)(vk)(t)for all t[r,R],\mathcal{E}\left(v_{k}\right)\left(t_{v_{k}}\right)\geq\mathcal{E}\left(v_{k}\right)\left(t\right)\ \ \ \text{for all\ }t\in\left[r,R\right],

that is

0tvk(T(svk),svk,vk)𝑑s0t(T(svk),svk,vk)𝑑s(t[r,R]).\int_{0}^{t_{v_{k}}}\mathcal{F}(T(sv_{k}),sv_{k},v_{k})ds\geq\int_{0}^{t}\mathcal{F}(T(sv_{k}),sv_{k},v_{k})ds\ \ \left(t\in\left[r,R\right]\right).

Passing to the limit we obtain

0t0(T(sv0),sv0,v0)𝑑s0t(T(sv0),sv0,v0)𝑑s(t[r,R]),\int_{0}^{t^{0}}\mathcal{F}\left(T(sv^{0}),sv^{0},v^{0}\right)ds\geq\int_{0}^{t}\mathcal{F}(T(sv^{0}),sv^{0},v^{0})ds\ \ \left(t\in\left[r,R\right]\right),

so t0t^{0} is a maximum on [r,R]\left[r,R\right] of (v0).\mathcal{E}(v^{0}). Again from (h1), we have t0=tv0.t^{0}=t_{v^{0}}. Since the convergent subsequence (tvk)\left(t_{v_{k}}\right) was arbitrarily chosen, we infer that the entire sequence (tvk)\left(t_{v_{k}}\right) converges to tv0,t_{v^{0}}, which proves our statement. ∎

In what follows we denote

U:={tv:vK1, 0t<tv}\displaystyle U:=\left\{tv:\ v\in K_{1},\ 0\leq t<t_{v}\right\}

and

Ub:={tvv:vK1}.U_{b}:=\left\{t_{v}v:\ v\in K_{1}\right\}.

We call UbU_{b} the Nehari-type manifold associated with the functional \mathcal{F}, operator TT and the conical annular region Kr,RK_{r,R}.

Remark 1.

Under assumption (h1), any point uUbu\in U_{b} satisfies

rr0<|u|<R0R.r\leq r_{0}<\left|u\right|<R_{0}\leq R.

Thus, any fixed point of TT that belongs to UbU_{b} lies in the conical annular set KrRK_{rR}. Moreover, assuming that 0<r<R<0<r<R<\infty, if uUbu\in U_{b} is a fixed point of TT, then the associated energy in the direction v=u|u|v=\frac{u}{|u|}, that is (u|u|)(t)\mathcal{E}\left(\frac{u}{|u|}\right)(t), attains its maximum over the interval [r,R][r,R] at the unique point t=tu|u|t=t_{\frac{u}{|u|}}. Therefore, looking for a fixed point of the operator TT within the Nehari-type manifold UbU_{b}, we not only obtain a fixed point lying in a conical annular region, but also one that maximizes the energy functional along its own direction.

Remark 2.

Under assumption (h1), one has

(T(u),u,u|u|)=0for all uUb.\mathcal{F}\left(T\left(u\right),u,\frac{u}{|u|}\right)=0\ \ \ \text{for all }u\in U_{b}.

Indeed, if uUb,u\in U_{b}, then u=tvvu=t_{v}v for v=u|u|K1.v=\frac{u}{|u|}\in K_{1}. Also, tvt_{v} being the maximum of (v)\mathcal{E}\left(v\right) on [r,R]\left[r,R\right] located in the open interval (r,R),\left(r,R\right), one has

ddt(v)(tv)=(T(tvv),tvv,v)=0.\frac{d}{dt}\mathcal{E}\left(v\right)\left(t_{v}\right)=\mathcal{F}(T(t_{v}v),t_{v}v,v)=0.
Lemma 2.

Under the assumption (h1), the set UU is an open bounded subset of K,K, 0U,0\in U, and its boundary U\partial U relative to KK is Ub.U_{b}.

Proof.

Clearly, by definition, we have 0U0\in U.

(i) Boundedness. Let uUu\in U. Then u=tvu=tv, where v=u|u|K1v=\frac{u}{|u|}\in K_{1} and t<tvt<t_{v}. Thus, by assumption (h1), we have

|u|t<tv<R0,|u|\leq t<t_{v}<R_{0},

and consequently,

UKR0,U\subset K_{R_{0}},

which shows that UU is bounded.

(ii) Openness. To prove that UU is open we show that its complement

Uc:={tv:vK1,ttv}U_{c}:=\left\{tv:\ v\in K_{1},\ t\geq t_{v}\right\}

is closed. Let ukUcu_{k}\in U_{c} be a sequence such that

ukuK.u_{k}\to u\in K.

We show that uUcu\in U_{c}. By the definition of UcU_{c}, we have

uk=tkvk,u_{k}=t_{k}v_{k},

where tktvkt_{k}\geq t_{v_{k}} and |vk|=1|v_{k}|=1. From the convergence tkvkut_{k}v_{k}\rightarrow u, it follows that that tk|u|.t_{k}\rightarrow\left|u\right|. Since

tktvkr0>0,t_{k}\geq t_{v_{k}}\geq r_{0}>0, (1)

one has |u|r0,\left|u\right|\geq r_{0}, so uK{0}.u\in K\setminus\left\{0\right\}.

Next, observe that vk1|u|u=:v0v_{k}\rightarrow\frac{1}{\left|u\right|}u=:v^{0} and based on Lemma 1, it follows that tvktv0.t_{v_{k}}\rightarrow t_{v^{0}}. Passing to the limit in (1), we deduce that |u|tv0.\left|u\right|\geq t_{v^{0}}. Thus

u=|u|v0tv0v0u=\left|u\right|v^{0}\geq t_{v^{0}}v^{0}

where v0K1.v^{0}\in K_{1}. This shows that uUcu\in U_{c} as desired, and therefore UcU_{c} is closed.

(iii) U=Ub\partial U=U_{b}. In order to prove that U=Ub,\partial U=U_{b}, we need to show that

U¯=D:={tv: 0ttv,vK1}.\overline{U}=D:=\left\{tv:\ 0\leq t\leq t_{v},\ v\in K_{1}\right\}. (2)

Let uU¯u\in\overline{U} be any point. Then uu is the limit of a sequence ukUu_{k}\in U, where uk=tkvku_{k}=t_{k}v_{k} with

vkK1, 0tk<tvkand tkvku.v_{k}\in K_{1},\ \ 0\leq t_{k}<t_{v_{k}}\ \ \text{and\ \ }t_{k}v_{k}\rightarrow u.

Taking to a subsequence we may assume that tkt0[0,R0].t_{k}\rightarrow t^{0}\in\left[0,R_{0}\right]. If t0=0,t^{0}=0, then u=0D.u=0\in D. Otherwise, if t0>0,t^{0}>0, we deduce that vk1t0u=:v0v_{k}\rightarrow\frac{1}{t^{0}}u=:v^{0} and v0K1.v^{0}\in K_{1}. Also, based on Lemma 1, one has tvktv0,t_{v_{k}}\rightarrow t_{v^{0}}, and from tk<tvkt_{k}<t_{v_{k}} we deduce that t0tv0.t^{0}\leq t_{v^{0}}. Hence u=t0v0,u=t^{0}v^{0}, where 0t0tv00\leq t^{0}\leq t_{v^{0}} and v0K1,v^{0}\in K_{1}, that is uD.u\in D. Thus U¯D.\overline{U}\subset D.

For the converse inclusion, take any uD.u\in D. Then

u=tv for some vK1and twith 0ttv.u=tv\ \ \ \text{ for some\ \ }v\in K_{1}\ \ \text{and\ }t\ \text{with\ \ }0\leq t\leq t_{v}.

If t=0,t=0, then u=0U¯.u=0\in\overline{U}. For t>0,t>0, we may choose an increasing sequence (tk)\left(t_{k}\right) with 0<tk<ttv0<t_{k}<t\leq t_{v}  and tkt.t_{k}\rightarrow t. Clearly uk:=tkvUu_{k}:=t_{k}v\in U and uktv=u.u_{k}\rightarrow tv=u. Hence uU¯.u\in\overline{U}. Therefore DU¯D\subset\overline{U} finishing the proof of (2).

Finally, from the representations of U¯\overline{U} and U,U, we see that U=U¯U=Ub.\partial U=\overline{U}\setminus U=U_{b}.

We now proceed to state the main result of the paper, a theorem of existence and localization of a fixed point in a conical annular region. Consider the following set of conditions:

(h2)

One has

infuUb|T(u)|>0.\inf_{u\in U_{b}}\left|T(u)\right|>0.
(h3)

The functional \mathcal{F} is such that if

(tu,u,u|u|)=0for some t>0 and uK{0},\mathcal{F}\left(tu,u,\frac{u}{|u|}\right)=0\quad\text{for some }t>0\text{ and }u\in K\setminus\{0\},

then t=1t=1.

(h4)

The functional \mathcal{F} is such that if

(tu,u,u|u|)=0for some t>0 and uK{0},\mathcal{F}\left(tu,u,\frac{u}{|u|}\right)=0\quad\text{for some }t>0\text{ and }u\in K\setminus\{0\},

then t1t\geq 1.

(h5)

The functional \mathcal{F} is such that if

(tu,u,u|u|)=0for some t>0 and uK{0},\mathcal{F}\left(tu,u,\frac{u}{|u|}\right)=0\quad\text{for some }t>0\text{ and }u\in K\setminus\{0\},

then t1t\leq 1.

(h6)

The operator TT is compressive on UU, i.e.,

|Tu||u|for all uUb.\left|Tu\right|\leq\left|u\right|\ \ \ \text{for all }u\in U_{b}.
(h7)

The operator TT is expansive on UU, i.e.,

|Tu||u|for all uUb.\left|Tu\right|\geq\left|u\right|\ \ \ \text{for all }u\in U_{b}.
Theorem 4.

Let condition (h1) holds. If either the set of conditions (h2) and (h3); or the set (h2), (h4) and (h6); or the set (h2), (h5) and (h7) is satisfied, then TT has a fixed point uu in Ub.U_{b}. In addition, uu maximizes the radial energy functional (v)\mathcal{E}(v) on the interval (r,R),\left(r,R\right), along its own direction v=1|u|u.v=\frac{1}{\left|u\right|}u.

Proof.

(a) Let conditions (h2) and (h3) hold. Hypothesis (h2) makes applicable the Birkhoff-Kellogg theorem for the operator TT on the set U.U. Hence, it is guaranteed the existence of a number t>0t>0 and uU=Ubu\in\partial U=U_{b} (from Lemma 2) such that T(u)=tu.T(u)=tu. Since uUbu\in U_{b}, using Remarks 2, we have

0=(T(u),u,u|u|)=(tu,u,u|u|),0=\mathcal{F}\left(T(u),u,\frac{u}{|u|}\right)=\mathcal{F}\left(tu,u,\frac{u}{|u|}\right), (3)

so by assumption (h3), t=1t=1, i.e., uu is a fixed point for TT.

(b) Let conditions (h2), (h4) and (h6) hold. As above, there exist t>0t>0 and uUbu\in U_{b} such that T(u)=tuT(u)=tu, and relation (3) holds. Then, by assumption (h4), we have t1t\geq 1. If t>1t>1, then

|T(u)|=t|u|>|u|,|T(u)|=t|u|>|u|,

which contradicts assumption (h6). Hence, t=1t=1, and therefore uu is a fixed point of TT.

(c) Let conditions (h2), (h5) and (h7) hold. As in step (b), there exist t>0t>0 and uUbu\in U_{b} such that T(u)=tuT(u)=tu, and relation (3) holds. Now, from assumption (h5), we have t1t\leq 1. If t<1t<1, then

|T(u)|=t|u|<|u|,|T(u)|=t|u|<|u|,

which contradicts assumption (h7). Hence, t=1t=1, and thus T(u)=uT(u)=u. ∎

Using the nonlinear alternative instead of the Birkhoff-Kellogg theorem, we obtain the following existence result.

Theorem 5.

Assume conditions (h1) and (h5) hold true. Then the operator TT has a fixed point in uU¯.u\in\overline{U}.

Proof.

From the nonlinear alternative we have that either

(a)

TT has a fixed point in U¯,\overline{U},

or

(b)

T(u)=tuT(u)=tu for some uUbu\in U_{b} and t>1t>1.

The second case of the alternative does not occur, since otherwise we would have

0=(T(u),u,u|u|)=(tu,u,u|u|),0=\mathcal{F}\left(T\left(u\right),u,\frac{u}{|u|}\right)=\mathcal{F}\left(tu,u,\frac{u}{|u|}\right),

whence by (h5), we obtain t1,t\leq 1, which is a contradiction. ∎

Remark 3.

Clearly, the fixed point uu given by Theorem 5 can be the origin. However, if T(0)0,T\left(0\right)\neq 0, then uu is not trivial.

Remark 4.

All the previous results remain valid if, in condition (h1), the maximum is replaced by the minimum, that is, the following condition holds:

(h1)

There exist r0>0r_{0}>0 and R0<R_{0}<\infty with rr0<R0Rr\leq r_{0}<R_{0}\leq R such that for each vK1v\in K_{1}, the mapping (v)\mathcal{E}(v) has a unique minimum tvt_{v} on [r,R]\left[r,R\right] and r0<tv<R0.r_{0}<t_{v}<R_{0}.

Indeed, (h1) implies (h1) for ,-\mathcal{F}, while all the other conditions (h2)-(h7) are not affected by this change of sign.

Noting that the definition of the Nehari-type manifold UbU_{b} and all the associated conditions are given with respect to a fixed pair (r,R),(r,R), we can expect to repeat them for different such pairs, thereby obtaining a finite or infinite number of fixed points of the operator TT in the cone K.K. This can happen if the mapping \ \mathcal{E} has oscillations. Thus, we have

Theorem 6.
(10)

If there are pairs of numbers (ri,Ri),\left(r_{i},R_{i}\right), i=1,2,,ni=1,2,...,n\ such that

ri<Riri+1fori=1,2,,n1r_{i}<R_{i}\leq r_{i+1}\ \ {\ \text{f}or\ }i=1,2,...,n-1

and Theorem 4 applies to each of these pairs, then the operator TT has nn fixed points uiKu_{i}\in K with

ri<|ui|<Ri,i=1,2,,n.r_{i}<\left|u_{i}\right|<R_{i},\ \ \ i=1,2,...,n.
(20)

If there is a sequence of pairs (ri,Ri),i=1,2,\left(r_{i},R_{i}\right),\ i=1,2,...\ such that

ri<Riri+1fori=1,2,r_{i}<R_{i}\leq r_{i+1}\ \ {\ \text{f}or\ }i=1,2,...

and Theorem 4 applies to each of these pairs, then the operator TT has a sequence of fixed points uiKu_{i}\in K with

ri<|ui|<Ri,i=1,2,.r_{i}<\left|u_{i}\right|<R_{i},\ \ \ i=1,2,....
(30)

If there is a sequence of pairs (ri,Ri),i=1,2,\left(r_{i},R_{i}\right),\ i=1,2,...\ such that

ri+1<Ri+1rifori=1,2,r_{i+1}<R_{i+1}\leq r_{i}\ \ {\ \text{f}or\ }i=1,2,...

and Theorem 4 applies to each of these pairs, then the operator TT has a sequence of fixed points uiKu_{i}\in K with

ri<|ui|<Ri,i=1,2,.r_{i}<\left|u_{i}\right|<R_{i},\ \ \ i=1,2,....

2.2 Recover the classical variational framework

In Section 2.1, a Nehari-type method has been introduced in a nonvariational framework, starting from a fixed point problem

T(u)=uT\left(u\right)=u

associated with an operator T.T. In this respect, an energy-type function and a Nehari-type manifold have been defined. The natural question is whether we can recover the classical Nehari method within the variational framework.

To give an answer, let XX^{\ast} be the dual of XX, ,\langle\cdot,\cdot\rangle denote the duality between XX^{\ast} and X,X, and let the norms on XX and XX^{\ast} be denoted by the same symbol ||.\left|\cdot\right|. We shall denote by JJ the duality mapping corresponding to a normalization function θ,\theta, i.e. the set-valued operator J:X𝒫(X)J:X\rightarrow{\mathcal{P}}(X^{\ast}) defined by

Jx={xX:x,x=θ(|x|)|x|,|x|=θ(|x|)},xX.Jx=\big\{x^{\ast}\in X^{\ast}:\ \langle x^{\ast},x\rangle=\theta\left(|x|\right)\left|x\right|,\ |x^{\ast}|=\theta\left(|x|\right)\big\},\quad x\in X.

Recall that by a normalization function we mean a continuous strictly increasing function θ:++\theta:\mathbb{R}_{+}\rightarrow\mathbb{R}_{+} with θ(0)=0\theta\left(0\right)=0 and limθt(t)=.{}_{t\rightarrow\infty}\theta\left(t\right)=\infty. We assume that θ(1)=1.\theta\left(1\right)=1. Obviously, one has

J(λx)=θ(λ)J(x)J\left(\lambda x\right)=\theta\left(\lambda\right)J(x)

for every xXx\in X and λ.\lambda\in\mathbb{R}. We assume that JJ is single-valued, bijective, and that both JJ and its inverse J1J^{-1} are continuous.

Furthermore, assume that EC1(X)E\in C^{1}(X) is a functional satisfying, without loss of generality, E(0)=0E(0)=0. We can easily see that for any t>0t>0 and uXu\in X, one has

E(tu)=E(tu)E(0)=01E(θtu),tu𝑑θ=0tE(su),u𝑑s.E(tu)=E(tu)-E(0)=\int_{0}^{1}\langle E^{\prime}(\theta tu),\,tu\rangle d\theta=\int_{0}^{t}\langle E^{\prime}(su),\,u\rangle ds.

Our goal is to obtain critical point of E,E, that is to solve the equation

E(u)=0.E^{\prime}\left(u\right)=0.

Take K=XK=X. We consider the operator T:KKT:K\rightarrow K defined by

T(u):=J1(JuE(u)).T\left(u\right):=J^{-1}\left(Ju-E^{\prime}\left(u\right)\right).

Obviously, the critical points of the functional EE are the fixed points of the operator T.T.

Let us consider the functional :K×K×K1\mathcal{F}:K\times K\times K_{1}\rightarrow\mathbb{R} (recall that K1={uK:|x|=1}K_{1}=\{u\in K\,:\,|x|=1\}),

(u,v,w):=JvJu,w.\mathcal{F}\left(u,v,w\right):=\left\langle Jv-Ju,w\right\rangle.

Then,

(v)(t)\displaystyle\mathcal{E}\left(v\right)\left(t\right) :=0t(T(sv),sv,v)𝑑s\displaystyle:=\int_{0}^{t}\mathcal{F}\left(T\left(sv\right),sv,v\right)ds
=0tJ(sv)JT(sv),v𝑑s\displaystyle=\int_{0}^{t}\left\langle J\left(sv\right)-JT\left(sv\right),v\right\rangle ds
=0tE(sv),v𝑑s\displaystyle=\int_{0}^{t}\left\langle E^{\prime}\left(sv\right),v\right\rangle ds
=E(tv).\displaystyle=E(tv).

From this, we observe that the classical energy functional EE is recovered through \mathcal{E}. Consequently, the classical Nehari manifold method (see [24]) is also retrieved; both in terms of the maximized mapping

tE(tu),t\mapsto E(tu),

and in the classical Nehari manifold 𝒩\mathcal{N} given by

𝒩={uX:E(u),u=0}.\mathcal{N}=\left\{u\in X:\langle E^{\prime}(u),u\rangle=0\right\}.
Corollary 1.

Under the previous conditions, if in addition

J1(JuE(u))(K)K,J^{-1}\left(Ju-E^{\prime}\left(u\right)\right)(K)\subset K,

then EE has a critical point uu in the Nehari manifold Ub,U_{b}, with

uK,r<|u|<R.u\in K,\ \ r<\left|u\right|<R.

In addition, uu maximizes the radial energy functional

t(v)(t)=E(tv),t\mapsto\mathcal{E}(v)\left(t\right)=E(tv),

on the interval (r,R),\left(r,R\right), along its own direction v=1|u|u.v=\frac{1}{\left|u\right|}u.

Proof.

We only need to check conditions (h2) and (h3).

Check of (h2): Let uUbu\in U_{b} and denote

w:=T(u)=J1(JuE(u)).w:=T\left(u\right)=J^{-1}\left(Ju-E^{\prime}\left(u\right)\right).

Then |u|r0\left|u\right|\geq r_{0} and

Jw=JuE(u),Jw=Ju-E^{\prime}\left(u\right),

whence, since E(u),u=0,\left\langle E^{\prime}\left(u\right),u\right\rangle=0, one has

Jw,u=JuE(u),u=Ju,u=θ(|u|)|u||Jw||u|=θ(|w|)|u|.\left\langle Jw,u\right\rangle=\left\langle Ju-E^{\prime}(u),u\right\rangle=\left\langle Ju,u\right\rangle=\theta\left(\left|u\right|\right)\left|u\right|\leq\left|Jw\right|\left|u\right|=\theta\left(\left|w\right|\right)\left|u\right|.

From this, we have θ(|w|)θ(|u|)\theta(|w|)\geq\theta(|u|). Thus, by the monotonicity of the function θ\theta, it follows that

|w||u|r0.|w|\geq|u|\geq r_{0}.

Therefore,

infuUb|T(u)|r0>0.\inf_{u\in U_{b}}\left|T\left(u\right)\right|\geq r_{0}>0.

Check of (h3): Assume that

(tu,u,u|u|)=0,\mathcal{F}\left(tu,u,\frac{u}{|u|}\right)=0,

for some t>0t>0 and uK{0}.u\in K\setminus\{0\}. Then

0=JuJ(tu),u|u|,0=\left\langle Ju-J\left(tu\right),\frac{u}{|u|}\right\rangle,

which is equivalent to

0=Juφ(t)Ju,u=(1φ(t))φ(|u|)|u|.0=\left\langle Ju-\varphi\left(t\right)Ju,u\right\rangle=\left(1-\varphi\left(t\right)\right)\varphi\left(\left|u\right|\right)\left|u\right|.

Hence, φ(t)=1\varphi(t)=1, and therefore t=1t=1, according to our assumption that φ(1)=1\varphi(1)=1.

Thus Theorem 4 applies and gives the result. ∎

3 Applications

3.1 A nonvariational Dirichlet problem with pp-Laplacian

To illustrate the theoretical results, we first consider the following nonvariational Dirichlet problem for a pp-Laplace equation

{(|u|p2u)(t)=f(u(t),u(t)),t(0,1),u(t)0u(0)=u(1)=0,\begin{cases}-\left(\left|u^{\prime}\right|^{p-2}u^{\prime}\right)^{\prime}(t)=f(u(t),u^{\prime}(t)),\quad t\in\left(0,1\right),\\ u(t)\geq 0\\ u(0)=u(1)=0,\end{cases} (4)

where 1<p<1<p<\infty, and f:+×+f:\mathbb{R}_{+}\times\mathbb{R}\rightarrow\mathbb{R}_{+} is a continuous function.

Consider the Banach space X:=W01,p(0,1)X:=W_{0}^{1,p}\left(0,1\right) endowed with the usual norm |u|1,p:=|u|p|u|_{1,p}:=|\nabla u|_{p}, where

|w|p=(01|w|p𝑑x)1p(wLp(0,1)).|w|_{p}=\left(\int_{0}^{1}|w|^{p}dx\right)^{\frac{1}{p}}\quad(w\in L^{p}(0,1)).

It is well known (see, e.g., [7, Chapter 1.2]) that W01,p(0,1)W_{0}^{1,p}(0,1) is a uniformly convex and reflexive Banach space with W01,q(0,1)W_{0}^{-1,q}(0,1) its dual, where qq is the conjugate of pp, i.e., 1p+1q=1\frac{1}{p}+\frac{1}{q}=1. Let ,\langle\cdot,\cdot\rangle denote the duality pairing between W01,q(0,1)W_{0}^{-1,q}(0,1) and W01,p(0,1)W_{0}^{1,p}(0,1). Then, for any vLq(0,1)W01,q(0,1)v\in L^{q}(0,1)\subset W_{0}^{-1,q}(0,1), we have

v,u=01v(t)u(t)𝑑t,\langle v,u\rangle=\int_{0}^{1}v(t)u(t)dt,

for all uW01,p(0,1)u\in W_{0}^{1,p}(0,1) (see, e.g., [3, Proposition 8.14]).

In what follows, J:W01,p(0,1)W01,q(0,1)J\colon W^{1,p}_{0}(0,1)\to W^{-1,q}_{0}(0,1) denotes the duality mapping of W01,p(0,1)W^{1,p}_{0}(0,1) corresponding to the normalization function θ(t)=tp1\theta(t)=t^{p-1} (t0t\geq 0). It is known that

Ju=(|u|p2u),uW01,p(0,1),Ju=-\left(|u^{\prime}|^{p-2}u^{\prime}\right)^{\prime},\quad u\in W^{1,p}_{0}(0,1),

and that JJ is bijective and continuous, while its inverse J1J^{-1} is strongly monotone, bounded (maps bounded sets into bounded sets), and continuous (see, e.g., [7, Theorem 8]).

Let λp\lambda_{p} denote the first eigenvalue of the Euler-Lagrange equation

Ju=λ|u|1,pp2uin (0,1),u(0)=u(1)=0.Ju=\lambda|u|_{1,p}^{p-2}u\quad\text{in }\,(0,1),\quad u(0)=u(1)=0.

Then (see, e.g., [7, Remark 6]),

λp=minuW01,p(0,1){0}|u|1,pp|u|pp,\lambda_{p}=\min_{u\in W_{0}^{1,p}(0,1)\setminus\{0\}}\frac{|u|_{1,p}^{p}}{|u|_{p}^{p}},

that is, cp:=λp1/pc_{p}:=\lambda_{p}^{-1/p} is the smallest constant such that, for all uW01,p(0,1)u\in W_{0}^{1,p}(0,1), we have

|u|pcp|u|1,p.|u|_{p}\leq c_{p}\,|u|_{1,p}. (5)

Given the continuous embedding of W01,p(0,1)W_{0}^{1,p}(0,1) in C[0,1]C[0,1] (see e.g., [3]), one has

|u(t)||u|1,p(t[0,1]),\left|u(t)\right|\leq|u|_{1,p}\ \ \ \left(t\in\left[0,1\right]\right), (6)

for every uW01,p(0,1)u\in W_{0}^{1,p}(0,1).

In W01,p(0,1)W^{1,p}_{0}(0,1), we consider the cone

K={uW01,p(0,1):\displaystyle K=\Big\{u\in W^{1,p}_{0}(0,1)\,: u0,u is concave,u(t)=u(1t)\displaystyle u\geq 0,\,u\text{ is concave},\,u(t)=u(1-t)
and u(t)ϕ(t)|u|1,p for all t[0,1/2]},\displaystyle\text{ and }u(t)\geq\phi(t)\,|u|_{1,p}\text{ for all }t\in[0,1/2]\Big\},

where ϕ\phi is given by

ϕ:[0,1/2]+, ϕ(t)=0t(12s)1p1𝑑s.\phi:[0,1/2]\rightarrow\mathbb{R}_{+},\text{ \quad$\phi(t)=\int_{0}^{t}(1-2s)^{\frac{1}{p-1}}ds$.} (7)

For 0<r<R<0<r<R<\infty, we assume that the following conditions hold:

(H1)

For each x+x\in\mathbb{R}_{+}, the function f(x,)f(x,\cdot) is decreasing on +\mathbb{R}_{+}, and for each yy\in\mathbb{R}, the function f(,y)f(\cdot,y) is increasing on +\mathbb{R}_{+}. Moreover, ff is even in its second variable, that is,

f(x,y)=f(x,y),f(x,y)=f(x,-y),

for all x+x\in\mathbb{R}_{+} and yy\in\mathbb{R}.

(H2)

Let

f0(x)=f(x,0) and f(x)=limyf(x,y).f_{0}(x)=f(x,0)\quad\text{ and }\quad f_{\infty}(x)=\lim_{y\to\infty}f(x,y).

The functions f0f_{0} and ff_{\infty} satisfy

f0(r)<rp1cp and f(Rϕ(β))>Rp12Φ,f_{0}(r)<\frac{r^{p-1}}{c_{p}}\quad\text{ and }\quad f_{\infty}(R\phi(\beta))>\frac{R^{p-1}}{2\Phi}, (8)

where β(0,12)\beta\in\left(0,\frac{1}{2}\right) is a fixed value, and

Φ:=β1/2ϕ(s)𝑑s.\Phi:=\int_{\beta}^{1/2}\phi(s)ds.
(H3)

For each x(0,)x\in(0,\infty) and yy\in\mathbb{R}, the mapping

tf(tx,ty)tp1 t\mapsto\frac{f(tx,ty)}{t^{p-1}}\,\text{ }

is strictly increasing on (0,R].(0,R].

In order to apply Theorem 4, we choose the functional :K×K×K1\mathcal{F}\colon K\times K\times K_{1}\to\mathbb{R} to be

(u,v,w)=JvJu,w.\mathcal{F}(u,v,w)=\langle Jv-Ju,w\rangle.

Clearly \mathcal{F} is continuous as JJ is continuous.

We observe that the problem (4) allows for a fixed point formulation

u=T(u),u=T(u),

where T:KW01,p(0,1)T\colon K\to W^{1,p}_{0}(0,1) is given by

T(u)=J1Nf(u,u),T(u)=J^{-1}N_{f}(u,u^{\prime}), (9)

and

Nf:K×Lp(0,1)Lq(0,1),Nf(u,u~)(t)=f(u(t),u~(t)),N_{f}\colon K\times L^{p}(0,1)\to L^{q}(0,1),\quad N_{f}(u,\tilde{u})(t)=f(u(t),\tilde{u}(t)), (10)

is the Nemytskii operator associated with the function ff.

In the subsequent, we will present a series of auxiliary results that will be used to prove the invariance of the operator TT over KK. When we say that a function uu is symmetric, we understand symmetric with respect to 12\frac{1}{2}.

The first result concerns the symmetry and nonnegativity of the solution to a pp-Laplace equation.

Lemma 3.

Let hLq(0,1)h\in L^{q}(0,1) be nonnegative a.e. on (0,1)(0,1) and symmetric. Then so is J1hJ^{-1}h.

Proof.

Clearly J1hJ^{-1}h is well defined (recall that Lq(0,1)W01,qL^{q}(0,1)\subset W^{-1,q}_{0}(0,1)). Nonnegativity follows directly from the comparison principle for the pp-Laplace operator (see [1, Lemma 1.3]). Let uh=J1hu_{h}=J^{-1}h and define w(t)=uh(1t)w(t)=u_{h}(1-t). A simple computation shows that Jw=hJw=h. Thus, by the uniqueness of the solution to the pp-Laplace equation, we deduce that uh=wu_{h}=w, which proves that J1hJ^{-1}h is symmetric. ∎

Our second result concerns the preservation of concavity for a pp-Laplace equation. Before stating it, let us define

φ:,φ(x)=|x|p2x=sgn(x)|x|p1.\varphi\colon\mathbb{R}\to\mathbb{R},\quad\varphi(x)=|x|^{p-2}x=\operatorname{sgn}(x)\,|x|^{p-1}.

Clearly, φ\varphi is a homeomorphism from \mathbb{R} onto \mathbb{R}, and its inverse φ1\varphi^{-1} is increasing on \mathbb{R}. Moreover, the problem (4) can be written in terms of φ\varphi as

(φ(u))=f(u,u),u(0)=u(1)=0.-\left(\varphi(u^{\prime})\right)^{\prime}=f(u,u^{\prime}),\quad u(0)=u(1)=0.
Lemma 4.

For any hLq(0,1)h\in L^{q}(0,1) that is symmetric with respect to 12\frac{1}{2}, we have the representation

(J1h)(t)=0t(φ1μ)(s)𝑑s,\left(J^{-1}h\right)(t)=\int_{0}^{t}(\varphi^{-1}\circ\mu)(s)ds, (11)

where

μ(t)=t12h(s)𝑑s.\mu(t)=\int_{t}^{\frac{1}{2}}h(s)ds.

If in addition hh is nonnegative a.e. on (0,1)(0,1) and nondecreasing a.e. on (0,12)\left(0,\frac{1}{2}\right), then J1hJ^{-1}h is concave on (0,1)(0,1).

Proof.

Denote

w(t)=0t(φ1μ)(s)𝑑s,w(t)=\int_{0}^{t}(\varphi^{-1}\circ\mu)(s)\,ds,

Since μ\mu is continuous (see, e.g., [3, Theorem 8.2]), it follows that wC1[0,1]w\in C^{1}[0,1]. Thus,

φ(w)=μ,\varphi(w^{\prime})=\mu,

and by [3, Lemma 8.2], we have that

φ(w)=h.-\varphi(w^{\prime})^{\prime}=h.

By the uniqueness of the solution to the pp-Laplace equation (see, e.g., [7, Remark 3]), we obtain (11) as the unique representation of J1hJ^{-1}h.

Assume now that hh is nonnegative a.e. on (0,1)(0,1) and nondecreasing a.e. on (0,12)\left(0,\frac{1}{2}\right). We observe that μ=h0\mu^{\prime}=-h\leq 0 a.e. on (0,1)(0,1); hence μ\mu is nonincreasing on (0,1)(0,1) (see, e.g., [10, Chapter 3]). Since

(J1h)=φ1μ,(J^{-1}h)^{\prime}=\varphi^{-1}\circ\mu,

and the right-hand side is nonincreasing on (0,1)(0,1) (as the composition of an increasing and a nonincreasing function), it follows that J1hJ^{-1}h is concave on (0,1)(0,1). ∎

In the following, we establish a Harnack type inequality.

Lemma 5.

Let hLq(0,1)h\in L^{q}(0,1) be nonnegative a.e. on (0,1)(0,1), symmetric with respect to 12\frac{1}{2}, and nondecreasing a.e. on (0,12)\left(0,\frac{1}{2}\right). Then, for all tt\in (0,12)\left(0,\frac{1}{2}\right), one has

(J1h)(t)ϕ(t)|J1h|1,p,\left(J^{-1}h\right)(t)\geq\phi(t)\left|J^{-1}h\right|_{1,p}, (12)

where ϕ\phi is given in (7).

Proof.

Let uh=J1hu_{h}=J^{-1}h. Since uhu_{h} is symmetric with respect to 12\frac{1}{2} (see Lemma 3), one has (recall that uhC1[0,1]u_{h}\in C^{1}[0,1]),

uh(12)=0.u_{h}^{\prime}\left(\frac{1}{2}\right)=0.

It is not difficult to prove that

uh(t)(12t)1p1uh(0),u_{h}^{\prime}(t)\geq(1-2t)^{\frac{1}{p-1}}u_{h}^{\prime}(0), (13)

for all t(0,12).t\in\left(0,\frac{1}{2}\right). To see this, let

σ:(0,12),σ(t)=uh(t)p1(12t)uh(0)p1.\sigma\colon\left(0,\frac{1}{2}\right)\to\mathbb{R},\quad\sigma(t)=u_{h}^{\prime}(t)^{p-1}-(1-2t)u_{h}^{\prime}(0)^{p-1}.

Since uhu_{h} is concave and uh(12)=0u_{h}^{\prime}\left(\frac{1}{2}\right)=0, we deduce that uh(t)0u_{h}^{\prime}(t)\geq 0 for all t(0,12)t\in\left(0,\frac{1}{2}\right). Thus,

uh(t)p1=Juh,u_{h}^{\prime}(t)^{p-1}=-Ju_{h},

which implies

(uh(t)p1)=h.\left(u_{h}^{\prime}(t)^{p-1}\right)^{\prime}=-h.

Therefore,

σ(t)=h(t)+2uh(0)p1.\sigma^{\prime}(t)=-h(t)+2u_{h}^{\prime}(0)^{p-1}.

which is nonincreasing a.e. on (0,12)\left(0,\frac{1}{2}\right). Following [14, Remark 2.5], we conclude that σ\sigma is concave. As σ(0)=σ(12)=0\sigma(0)=\sigma\left(\frac{1}{2}\right)=0, we have that σ(t)0\sigma(t)\geq 0 for all t(0,12)t\in\left(0,\frac{1}{2}\right), i.e., (13) holds.

Since uh(0)=0u_{h}(0)=0, integrating (13) from 0 to tt (t12t\leq\frac{1}{2}), we obtain

uh(t)uh(0)0t(12s)1p1𝑑s=ϕ(t)uh(0).u_{h}(t)\geq u^{\prime}_{h}(0)\int_{0}^{t}(1-2s)^{\frac{1}{p-1}}ds=\phi(t)u^{\prime}_{h}(0).

Finally, the conclusion follows from the above inequality and

|uh|1,p=(01|u(s)|p𝑑s)1p=(201/2u(s)p𝑑s)1p(u(0)p)1p=uh(0).|u_{h}|_{1,p}=\left(\int_{0}^{1}|u^{\prime}(s)|^{p}ds\right)^{\frac{1}{p}}=\left(2\int_{0}^{1/2}u^{\prime}(s)^{p}ds\right)^{\frac{1}{p}}\leq\left(u^{\prime}(0)^{p}\right)^{\frac{1}{p}}=u_{h}^{\prime}(0).

Now we prove that the Nemytskii’s operator NfN_{f} is well-defined, bounded (maps bounded sets into bounded sets) and continuous.

Lemma 6.

Assume condition (H1) holds true. Then the Nemytskii operator NfN_{f} is well-defined, bounded (maps bounded sets into bounded sets) and continuous from K×Lp(0,1)K\times L^{p}(0,1) to Lq(0,1)L^{q}(0,1).

Proof.

From (H1), we have

0f(x,y)f(x,0),0\leq f(x,y)\leq f(x,0), (14)

for all x+x\in\mathbb{R}_{+} and yy\in\mathbb{R}. Indeed, for x+x\in\mathbb{R}_{+} and yy\in\mathbb{R}, the evenness of ff in its second variable gives f(x,y)=f(x,|y|)f(x,y)=f(x,|y|). Since f(x,)f(x,\cdot) is decreasing on +\mathbb{R}_{+}, we obtain f(x,|y|)f(x,0)f(x,|y|)\leq f(x,0). The left-hand side of (14) follows from the definition of ff.

Well-definedness. Let uKu\in K and u~Lp(0,1).\tilde{u}\in L^{p}(0,1). Since ff is continuous, it follows that Nf(u,u~)N_{f}(u,\tilde{u}) is measurable (see, e.g., [19, Proposition 9.1]). Moreover, as f(u,0)C[0,1]f(u,0)\in C[0,1], using (14) we obtain

01f(u,u~)q𝑑x01f(u,0)q𝑑x=|f(u,0)|qq,\int_{0}^{1}f(u,\tilde{u})^{q}\,dx\leq\int_{0}^{1}f(u,0)^{q}\,dx=|f(u,0)|_{q}^{q},

which shows that Nf(u,u~)Lq(0,1)N_{f}(u,\tilde{u})\in L^{q}(0,1).

Boundedness. Let uKu\in K and u~Lp(0,1).\tilde{u}\in L^{p}(0,1). Using (14), to prove the boundedness of NfN_{f}, it suffices to show that the operator uf(u,0)u\mapsto f(u,0) maps bounded sets into bounded sets from W01,p(0,1)W^{1,p}_{0}(0,1) to Lq(0,1)L^{q}(0,1). This however follows immediately from the continuous embedding of W01,p(0,1)W^{1,p}_{0}(0,1) into C[0,1]C[0,1] and the continuity of ff.

Continuity. Let (uk,u~k)K×Lp(0,1)(u_{k},\tilde{u}_{k})\in K\times L^{p}(0,1) be such that (uk,u~k)(u,u~)(u_{k},\tilde{u}_{k})\to(u,\tilde{u}) in W01,p(0,1)×Lp(0,1)W^{1,p}_{0}(0,1)\times L^{p}(0,1) as kk\to\infty. Then, there exists a subsequence (still denoted by (uk,u~k)(u_{k},\tilde{u}_{k})) with the property that (see, e.g., [3, Theorem 4.9])

(uk(x),u~k(x))(u(x),u~(x)) a.e. on (0,1).(u_{k}(x),\tilde{u}_{k}(x))\to(u(x),\tilde{u}(x))\,\,\text{ a.e. on }(0,1).

Passing again to a subsequence if necessarily, we may assume that

Nf(uk,u~k)(x)Nf(u,u~)(x) a.e. on (0,1).N_{f}(u_{k},\tilde{u}_{k})(x)\to N_{f}(u,\tilde{u})(x)\,\,\text{ a.e. on }(0,1).

Since ukuu_{k}\to u in W01,p(0,1)W^{1,p}_{0}(0,1), using (6), there exists M>0M>0 such that |uk(t)|M|u_{k}(t)|\leq M, for all t[0,1]t\in[0,1] and kk\in\mathbb{N}. Denote M1:=supx[0,M]|f(x,0)|M_{1}:=\sup_{x\in[0,M]}|f(x,0)|.

Let ε>0\varepsilon>0 and set δ=εqM1\delta=\frac{\varepsilon^{q}}{M_{1}}. Then, for any D[0,1]D\subset[0,1] with meas(D)<δmeas(D)<\delta, using (14), we estimate

|Nf(uk,u~k)|Lq(D|f(uk(x),0)|q𝑑x)1qmeas(D)1qM1δ1qM1=ε.|N_{f}(u_{k},\tilde{u}_{k})|_{L^{q}}\leq\left(\int_{D}|f(u_{k}(x),0)|^{q}dx\right)^{\frac{1}{q}}\leq meas(D)^{\frac{1}{q}}M_{1}\leq\delta^{\frac{1}{q}}M_{1}=\varepsilon.

By Vitali’s theorem, Nf(uk,u~k)N_{f}(u_{k},\tilde{u}_{k}) converges to Nf(u,u~)N_{f}(u,\tilde{u}) in Lq(0,1)L^{q}(0,1). Since the limit is the same regardless of the chosen subsequence, we conclude that the entire sequence Nf(uk,u~k)N_{f}(u_{k},\tilde{u}_{k}) converges to Nf(u,u~)N_{f}(u,\tilde{u}) in Lq(0,1)L^{q}(0,1) (see, e.g., [17, Lemma 1.1]). ∎

Now we are ready to show that T(K)KT(K)\subset K.

Lemma 7.

Assume condition (H1) holds true. Then,

T(K)K.T(K)\subset K.
Proof.

Let uKu\in K. Then, by Lemma 6, we have Nf(u,u)Lq(0,1)N_{f}(u,u^{\prime})\in L^{q}(0,1). Since u0u\geq 0, it follows that f(u,u)0f(u,u^{\prime})\geq 0. The symmetry of uu implies that uu^{\prime} is antisymmetric, i.e.,

u(t)=u(1t)a.e. on (0,1).u^{\prime}(t)=-u^{\prime}(1-t)\quad\text{a.e.\ on $(0,1)$}.

Thus, condition (H1) ensures that Nf(u,u)N_{f}(u,u^{\prime}) is symmetric. Moreover, since uu is nondecreasing on (0,12)\left(0,\frac{1}{2}\right) and uu^{\prime} is nonincreasing a.e. on (0,12)\left(0,\frac{1}{2}\right), condition (H1) also ensures that Nf(u,u)N_{f}(u,u^{\prime}) is nondecreasing a.e. on (0,12)\left(0,\frac{1}{2}\right). Now applying Lemmas 3, 4 and 5, we find that J1f(u,u)=T(u)J^{-1}f(u,u^{\prime})=T(u) is nonegative, symmetric, concave and satisfies the Harnack inequality (12). Therefore, T(u)KT(u)\in K. ∎

Next, we prove the complete continuity of the operator TT.

Lemma 8.

Under assumption (H1), the operator TT is completely continuous from KK to KK.

Proof.

We observe that

T=J1PNfI,T=J^{-1}\circ P\circ N_{f}\circ I,

where

P:KK×Lp(0,1),P(u)=(u,u),P\colon K\to K\times L^{p}(0,1),\quad P(u)=(u,u^{\prime}),

and II is the embedding operator

I:Lq(0,1)W01,q(0,1),I(u)=u,.I\colon L^{q}(0,1)\to W^{-1,q}_{0}(0,1),\quad I(u)=\langle u,\cdot\rangle.

Since, by the Rellich–Kondrachov theorem (see, e.g., [3]), W01,p(0,1)W^{1,p}_{0}(0,1) embeds compactly into Lp(0,1)L^{p}(0,1), it follows that Lq(0,1)L^{q}(0,1) embeds compactly into W01,q(0,1)W^{-1,q}_{0}(0,1), as the dual spaces of Lp(0,1)L^{p}(0,1) and W01,p(0,1)W^{1,p}_{0}(0,1), respectively. This follows from the Schauders’s theorem on compact operators (see, [3, Theorem 6.4], [22, Theorem 4.19] or [15, Chapter 9]). Additionally, it is clear that II is continuous and bounded. Since J1J^{-1}, PP, NfN_{f}, and II are continuous and bounded, and one of these operators is compact, it follows that TT is completely continuous, as the composition of all these four operators.

Theorem 7.

Under conditions (H1)-(H3), the problem (4) admits a solution uKu\in K such that

r<|u|1,p<R.r<|u|_{1,p}<R.

Moreover, uu maximizes the energy function (v)\mathcal{E}(v) on the interval (r,R),\left(r,R\right), along its own direction v=1|u|u.v=\frac{1}{\left|u\right|}u.

Proof.

Check of (h1). We will show that for any vK1v\in K_{1}, we have that

ddt(v)(t)|t=r>0,ddt(v)(t)|t=R<0,\frac{d}{dt}\mathcal{E}(v)(t)|_{t=r}>0,\quad\frac{d}{dt}\mathcal{E}(v)(t)|_{t=R}<0, (15)

and the mapping

tddt(v)(t),t\mapsto\frac{d}{dt}\mathcal{E}(v)(t), (16)

has a unique zero. It is easily to see that whenever (15) and (16) hold, then the function t(v)(t)t\mapsto\mathcal{E}(v)(t) has a unique maximum, as required by the abstract result.

For any vK1v\in K_{1} and t>0t>0, one has

ddt(v)(t)\displaystyle\frac{d}{dt}\mathcal{E}(v)(t) =tp1|v|1,pp01f(tv(s),tv(s))v(s)𝑑s\displaystyle=t^{p-1}|v|_{1,p}^{p}-\int_{0}^{1}f\left(tv(s),tv^{\prime}(s)\right)v(s)ds
=tp1(|v|1,pp01f(tv(s),tv(s))tp1v(s)𝑑s).\displaystyle=t^{p-1}\left(|v|_{1,p}^{p}-\int_{0}^{1}\frac{f\left(tv(s),tv^{\prime}(s)\right)}{t^{p-1}}v(s)ds\right).

Using (14) and the monotonicity of ff in the first variable, we obtain

f(rv(s),rv(s))f(rv(s),0)f(r,0)=f0(r),s(0,1),f(rv(s),rv^{\prime}(s))\leq f(rv(s),0)\leq f(r,0)=f_{0}(r),\quad s\in(0,1), (17)

where the latter inequality follows from (6), which ensures that v(s)1v(s)\leq 1 for all s[0,1]s\in[0,1].

Thus, (17) yields

01f(rv(s),rv(s))rp1v(t)𝑑sf0(r)rp101v(s)𝑑s<1cp|v|p1.\int_{0}^{1}\frac{f\left(rv(s),rv^{\prime}(s)\right)}{r^{p-1}}v(t)ds\leq\frac{f_{0}\left(r\right)}{r^{p-1}}\int_{0}^{1}v(s)ds<\frac{1}{c_{p}}|v|_{p}\leq 1.

Consequently,

ddt(v)(t)|t=r=rp1(101f(rv(s),rv(s))rp1v(s)𝑑s)>0,\frac{d}{dt}\mathcal{E}(v)(t)|_{t=r}=r^{p-1}\left(1-\int_{0}^{1}\frac{f\left(rv(s),rv^{\prime}(s)\right)}{r^{p-1}}v(s)ds\right)>0,

for all vK1v\in K_{1}.

To obtain the second inequality in (15), observe that for any vK1v\in K_{1}, the Harnack inequality, together with the monotonicity of vv on (0,12)\left(0,\frac{1}{2}\right), gives

v(t)ϕ(t)for all t(0,12) and v(t)ϕ(β)for all t(β,12).v(t)\geq\phi(t)\,\,\,\text{for all }t\in\left(0,\frac{1}{2}\right)\quad\text{ and }\quad v(t)\geq\phi(\beta)\,\,\,\text{for all }t\in\left(\beta,\frac{1}{2}\right). (18)

From (18) we obtain

β1v(s)𝑑sΦ.\int_{\beta}^{1}v(s)\,ds\geq\Phi. (19)

Next, by (H1) we have

f(x,y)f(x),x+,y.f(x,y)\geq f_{\infty}(x),\quad x\in\mathbb{R}_{+},\ y\in\mathbb{R}. (20)

Therefore, the monotonicity of ff in its first variable, together with (18) and (20), ensures that

f(Rv(s),Rv(s))f(Rv(s))f(Rv(β))f(Rϕ(β)),f(Rv(s),Rv^{\prime}(s))\geq f_{\infty}(Rv(s))\geq f_{\infty}(Rv(\beta))\geq f_{\infty}(R\phi(\beta)), (21)

for all s(β,12)s\in\left(\beta,\frac{1}{2}\right). Combining the second inequality from (H2), (19) and (21), we estimate

01f(Rv(s),Rv(s))Rp1v(s)𝑑s\displaystyle\int_{0}^{1}\frac{f(Rv(s),Rv^{\prime}(s))}{R^{p-1}}v(s)ds =2012f(Rv(s),Rv(s))Rp1v(s)𝑑s\displaystyle=2\int_{0}^{\frac{1}{2}}\frac{f(Rv(s),Rv^{\prime}(s))}{R^{p-1}}v(s)ds
2f(Rϕ(β))Rp1β12v(s)𝑑s\displaystyle\geq 2\frac{f(R\phi(\beta))}{R^{p-1}}\int_{\beta}^{\frac{1}{2}}v(s)ds
2f(Rϕ(β))Rp1Φ.\displaystyle\geq 2\frac{f(R\phi(\beta))}{R^{p-1}}\Phi.
>1.\displaystyle>1.

Thus,

ddt(v)(t)|t=R=Rp1(101f(Rv(s),Rv(s))Rp1v(s)𝑑s)<0,\frac{d}{dt}\mathcal{E}(v)(t)|_{t=R}=R^{p-1}\left(1-\int_{0}^{1}\frac{f\left(Rv(s),Rv^{\prime}(s)\right)}{R^{p-1}}v(s)ds\right)<0,

for all vK1v\in K_{1}, so both inequalities in (15) are valid.

To verify (16), we note that by (H3), for each vK1v\in K_{1}, the mapping

t01f(tv(s),tv(s))tp1v(s)𝑑st\mapsto\int_{0}^{1}\frac{f\left(tv(s),tv^{\prime}(s)\right)}{t^{p-1}}v(s)ds

is strictly increasing. Therefore, by (15), we have that

101f(rv(s),rv(s))rp1v(s)𝑑s>0 and 101f(Rv(s),Rv(s))Rp1v(s)𝑑s<0,1-\int_{0}^{1}\frac{f\left(rv(s),rv^{\prime}(s)\right)}{r^{p-1}}v(s)ds>0\quad\text{ and }\quad 1-\int_{0}^{1}\frac{f\left(Rv(s),Rv^{\prime}(s)\right)}{R^{p-1}}v(s)ds<0,

so the equation

101f(tv(s),tv(s))tp1v(s)𝑑s=0,1-\int_{0}^{1}\frac{f\left(tv(s),tv^{\prime}(s)\right)}{t^{p-1}}v(s)ds=0,

has a unique solution. This however implies that, for all vK1,v\in K_{1}, the mapping

tddt(v)(t)=tp1(101f(tv(s),tv(s))tp1v(s)𝑑s),t\mapsto\frac{d}{dt}\mathcal{E}(v)(t)=t^{p-1}\left(1-\int_{0}^{1}\frac{f\left(tv(s),tv^{\prime}(s)\right)}{t^{p-1}}v(s)ds\right),

has a unique zero, as desired.

Check of (h2). Let uUbu\in U_{b}. Then, by definition, we have

0=(T(u),u,u|u|1,p)=JuJ(T(u)),u|u|1,p,0=\mathcal{F}\left(T(u),u,\frac{u}{|u|_{1,p}}\right)=\left\langle Ju-J(T(u)),\frac{u}{|u|_{1,p}}\right\rangle,

which gives,

J(T(u)),u=Ju,u.\langle J(T(u)),u\rangle=\langle Ju,u\rangle.

One has,

θ(|u|1,p)|u|1,p=Ju,u=J(T(u)),u|J(T(u))|1,q|u|1,p=θ(|T(u)|1,p)|u|1,p,\theta(|u|_{1,p})|u|_{1,p}=\langle Ju,u\rangle=\langle J(T(u)),u\rangle\leq|J(T(u))|_{-1,q}\,|u|_{1,p}=\theta(|T(u)|_{1,p})|u|_{1,p},

whence

θ(|u|1,p)θ(|T(u)|1,p).\theta(|u|_{1,p})\leq\theta(|T(u)|_{1,p}).

By the monotonicity of θ\theta we obtain

|T(u)|1,p|u|1,pr,|T(u)|_{1,p}\geq|u|_{1,p}\geq r,

which shows that (h2) is satisfied.

Check of (h3). Since θ(t)=tp1\theta(t)=t^{p-1}, as shown in Corollary 1, the functional \mathcal{F} satisfies (h3).

The conclusion follows from Theorem 4, as TT is completely continuous and invariant over KK (see Lemmas 7 and 8), and conditions (h1)–(h3) are satisfied. ∎

Remark 5.

If the inequalities in (15) are reversed, i.e.,

ddt(v)(t)|t=r<0,ddt(v)(t)|t=R>0,\left.\frac{d}{dt}\mathcal{E}(v)(t)\right|_{t=r}<0,\quad\left.\frac{d}{dt}\mathcal{E}(v)(t)\right|_{t=R}>0, (22)

then Theorem 4 is still applicable. Indeed, in this case, the point tvt_{v} is a minimum point of the mapping t(v)(t)t\mapsto\mathcal{E}(v)(t), so condition (h1) is satisfied instead of (h1). As noted in Remark 4, the theory still applies and yields the existence of a point uKu\in K such that r<|u|1,p<Rr<|u|_{1,p}<R, which, in this case, minimizes the associated energy functional.

It is straightforward to verify that if

f(rϕ(β))(rϕ(β))p1>12Φϕ(β)p1andf0(R)Rp1<1cp,\frac{f_{\infty}(r\phi(\beta))}{(r\phi(\beta))^{p-1}}>\frac{1}{2\Phi\,\phi(\beta)^{p-1}}\quad\text{and}\quad\frac{f_{0}(R)}{R^{p-1}}<\frac{1}{c_{p}}, (23)

are assumed instead of (15), then (22) holds.

Example 1.

A typical example of a function ff that satisfies conditions (H1)–(H3) is

f(x,y)=xp(1+11+|y|),x+,y.f(x,y)=x^{p}\left(1+\frac{1}{1+|y|}\right),\quad x\in\mathbb{R}_{+},\,y\in\mathbb{R}.

Clearly, ff satisfies (H1). Moreover, f0(x)=|x|pf_{0}(x)=|x|^{p} and f(x)=2|x|pf_{\infty}(x)=2|x|^{p}, and they satisfy

limx0f0(x)xp1=limx0f(x)xp1=0,\lim_{x\to 0}\frac{f_{0}(x)}{x^{p-1}}=\lim_{x\to 0}\frac{f_{\infty}(x)}{x^{p-1}}=0,

and

limxf0(x)xp1=limxf(x)xp1=.\lim_{x\to\infty}\frac{f_{0}(x)}{x^{p-1}}=\lim_{x\to\infty}\frac{f_{\infty}(x)}{x^{p-1}}=\infty.

Therefore, (H2)–(H3) hold for sufficiently small rr and sufficiently large RR, respectively.

3.2 Radial functional energy with two critical points on each direction

In this subsection, we construct an explicit example of an operator TT and a functional \mathcal{F} such that, for each direction vv, the associated radial energy functional (v)\mathcal{E}(v) has two critical points: one corresponding to a global maximum point and the other to a local minimum. We remark that, for such a problem, a an approach similar to Theorem 7 is not applicable, since there we require that require the mapping t(v)(t)t\mapsto\mathcal{E}(v)(t) has a global maximizer, which is the unique critical point (relation (16)). However, the abstract result, Theorem 4, can still be applied, as it only requires (v)\mathcal{E}(v) to have a unique maximum point.

This example is illustrative since many results based on the Nehari manifold method typically require that, for each direction vv, the mapping tE(tv)t\mapsto E(tv) has a unique critical point, which is a global maximum point (see, e.g., [24, Chapter 3], [9, Theorem 2.1]). Therefore, our results can indeed be applied to a larger class of problems.

We consider the fixed point problem

u(t)=01k(t,s)f(u(s))ds=:T(u)(t),t(0,1),\displaystyle u(t)=\int_{0}^{1}k(t,s)\,f(u(s))\,ds=:T(u)(t),\quad t\in(0,1), (24)

where the kernel k:[0,1]2+k\colon[0,1]^{2}\to\mathbb{R}_{+} is given by

k(t,s)={t(1s),ts,s(1t),st,k(t,s)=\begin{cases}t(1-s),&t\leq s,\\ s(1-t),&s\leq t,\end{cases}

and f(x)=a2x2+a1x+a0f(x)=a_{2}x^{2}+a_{1}x+a_{0} with coefficients

a2=102,a1=2.5×103,a0=1.a_{2}=10^{-2},\quad a_{1}=2.5\times 10^{-3},\quad a_{0}=1.

In order to apply the abstract result, let X=C[0,1]X=C[0,1], endowed with the supremum norm

|u|=maxt[0,1]|u(t)|,|u|_{\infty}=\max_{t\in[0,1]}\,|u(t)|,

and let the cone KXK\subset X be given by

K={uC[0,1]:u0andmint[14,34]u(t)14|u|}.K=\left\{u\in C[0,1]:u\geq 0\ \text{and}\ \min_{t\in\left[\frac{1}{4},\frac{3}{4}\right]}u(t)\geq\frac{1}{4}|u|_{\infty}\right\}.

The functional :K×K×K1\mathcal{F}:K\times K\times K_{1}\to\mathbb{R} is chosen to be

(u,v,w)=01v(t)w(t)𝑑t01u(t)w(t)𝑑t.\mathcal{F}(u,v,w)=\int_{0}^{1}v(t)w(t)\,dt-\int_{0}^{1}u(t)w(t)\,dt.

Standard arguments (see, e.g., [13]) show that TT is completely continuous from XX to XX. It follows immediately that the cone KK is invariant under the operator TT, i.e., T(K)KT(K)\subset K. Indeed, first note that since a0,a1,a2>0a_{0},a_{1},a_{2}>0 then for any u0u\geq 0 we have T(u)0T(u)\geq 0. Next, let uKu\in K and σ[14,34].\sigma\in\left[\frac{1}{4},\frac{3}{4}\right]. Since (see, e.g., [13]),

k(t,s)14Φ(s) for all t[14,34]ands[0,1],k(t,s)\geq\frac{1}{4}\Phi(s)\,\,\,\text{ for all }t\in\left[\frac{1}{4},\frac{3}{4}\right]\,\,\text{and}\,\,s\in[0,1],

and

k(t,s)Φ(s) for all t,s[0,1],k(t,s)\leq\Phi(s)\,\,\,\text{ for all }\,t,s\in[0,1],

where Φ(s)=s(1s)\Phi(s)=s(1-s), we have

T(u)(σ)\displaystyle T(u)(\sigma) =01k(σ,s)f(u(s))𝑑s\displaystyle=\int_{0}^{1}k(\sigma,s)f(u(s))ds
1401Φ(s)f(u(s))𝑑s\displaystyle\geq\frac{1}{4}\int_{0}^{1}\Phi(s)f(u(s))ds
1401k(t,s)f(u(s))𝑑s\displaystyle\geq\frac{1}{4}\int_{0}^{1}k(t,s)f(u(s))ds
=T(u)(t),\displaystyle=T(u)(t),

for all t[0,1]t\in[0,1]. Thus, T(u)KT(u)\in K.

Next, in order to provide an explicit expression for the radial energy functional \mathcal{E}, we introduce the following notation

αu(k)=0101k(t,s)u(s)ku(t)𝑑s𝑑t,\alpha_{u}(k)=\int_{0}^{1}\int_{0}^{1}k(t,s)\,u(s)^{k}\,u(t)\,ds\,dt,

for all uKu\in K and k0k\geq 0. Let vK1v\in K_{1} and σ>0\sigma>0. Then, we have

(T(σv),σv,v)\displaystyle\mathcal{F}(T(\sigma v),\sigma v,v) =σ|v|L2201T(σv)(t)v(t)𝑑t\displaystyle=\sigma|v|_{L^{2}}^{2}-\int_{0}^{1}T(\sigma v)(t)\,v(t)\,dt (25)
=σ|v|L2201(01k(t,s)f(σv(s))𝑑s)v(t)𝑑t.\displaystyle=\sigma|v|_{L^{2}}^{2}-\int_{0}^{1}\left(\int_{0}^{1}k(t,s)f(\sigma v(s))\,ds\right)v(t)\,dt.
=a2σ2αv(2)+σ(|v|L22a1αv(1))a0αv(0).\displaystyle=-a_{2}\sigma^{2}\alpha_{v}(2)+\sigma\left(|v|_{L^{2}}^{2}-a_{1}\alpha_{v}(1)\right)-a_{0}\alpha_{v}(0).

Therefore, from (25), one has

(v)(t)\displaystyle\mathcal{E}(v)(t) =0t(T(σv),σv,v)𝑑σ\displaystyle=\int_{0}^{t}\mathcal{F}(T(\sigma v),\sigma v,v)d\sigma (26)
=a2t33αv(2)+t22(|v|L22a1αv(1))a0αv(0)t.\displaystyle=-a_{2}\frac{t^{3}}{3}\alpha_{v}(2)+\frac{t^{2}}{2}\left(|v|_{L^{2}}^{2}-a_{1}\alpha_{v}(1)\right)-a_{0}\alpha_{v}(0)t. (27)

In what follows, we show that conditions (h1)-(h3) of Theorem 4 are satisfied in the entire cone KK, that is,

r=0 and R=+.r=0\quad\text{ and }\quad R=+\infty.

Check of (h1). Let vK1.v\in K_{1}. Then, one has

|v|L22=01v(s)2𝑑s1/43/4v(s)2𝑑s1/43/4(14|v|)2𝑑s=132,\displaystyle|v|_{L^{2}}^{2}=\int_{0}^{1}v(s)^{2}ds\geq\int_{1/4}^{3/4}v(s)^{2}ds\geq\int_{1/4}^{3/4}\left(\frac{1}{4}|v|_{\infty}\right)^{2}ds=\frac{1}{32},
|v|L22=01v(s)2𝑑s|v|=1,\displaystyle|v|_{L^{2}}^{2}=\int_{0}^{1}v(s)^{2}ds\leq|v|_{\infty}=1,
αv(k)=0101k(t,s)v(s)kv(t)𝑑s𝑑t|v|k+10101k(t,s)𝑑s𝑑t=112,\displaystyle\alpha_{v}(k)=\int_{0}^{1}\int_{0}^{1}k(t,s)\,v(s)^{k}\,v(t)\,ds\,dt\leq|v|_{\infty}^{k+1}\int_{0}^{1}\int_{0}^{1}k(t,s)ds\,dt=\frac{1}{12},
αv(k)=0101k(t,s)v(s)kv(t)𝑑s𝑑t14k+1|u|k+11/43/41/43/4k(t,s)𝑑s𝑑t=13214k+1.\displaystyle\alpha_{v}(k)=\int_{0}^{1}\int_{0}^{1}k(t,s)\,v(s)^{k}\,v(t)\,ds\,dt\geq\frac{1}{4^{k+1}}|u|_{\infty}^{k+1}\int_{1/4}^{3/4}\int_{1/4}^{3/4}k(t,s)ds\,dt=\frac{1}{32}\frac{1}{4^{k+1}}.

From these, we can deduce the subsequent robust estimates

a2αv(2)[4.88×106, 8.34×104],\displaystyle a_{2}\alpha_{v}(2)\in\left[4.88\times 10^{-6},\,8.34\times 10^{-4}\right], (28)
|v|L22a1αv(1)[3×102, 1],\displaystyle|v|_{L^{2}}^{2}-a_{1}\alpha_{v}(1)\in\left[3\times 10^{-2},\,1\right],
a0αv(0)[7.81×103, 8.34×102].\displaystyle a_{0}\alpha_{v}(0)\in\left[7.81\times 10^{-3},\,8.34\times 10^{-2}\right].

Our next goal is to show that, for any choice of the parameters

b2[4.88×106, 8.34×104],b1[3×102, 1],\displaystyle b_{2}\in\left[4.88\times 10^{-6},\,8.34\times 10^{-4}\right],\,\,b_{1}\in\left[3\times 10^{-2},\,1\right], (29)
b0[7.81×103, 8.34×102],\displaystyle b_{0}\in\left[7.81\times 10^{-3},\,8.34\times 10^{-2}\right],

the cubic function

g(t):=b2t33+b1t22b0tg(t):=-\,b_{2}\frac{t^{3}}{3}+b_{1}\frac{t^{2}}{2}-b_{0}t

admits exactly one maximum point on the positive semi-axis (t>0t>0). To prove this, we look for the critical points of gg. Differentiating gives

g(t)=b2t2+b1tb0,g^{\prime}(t)=-b_{2}t^{2}+b_{1}t-b_{0},

so critical points are solutions of the quadratic equation

0=b2t2+b1tb0.0=-b_{2}t^{2}+b_{1}t-b_{0}.

The discriminant satisfies

b124b2b09×1042.79×104>4104>0,b_{1}^{2}-4b_{2}b_{0}\geq 9\times 10^{-4}-2.79\times 10^{-4}>4\cdot 10^{-4}>0, (30)

hence there are two distinct positive roots,

t±=b1±b124b2b02b2>0.t_{\pm}=\frac{b_{1}\pm\sqrt{b_{1}^{2}-4b_{2}b_{0}}}{2b_{2}}>0.

Notice that by (29) and (30), we have

t+=b1+b124b2b02b2>3×1021.7×103>10,t_{+}=\frac{b_{1}+\sqrt{b_{1}^{2}-4b_{2}b_{0}}}{2b_{2}}>\frac{3\times 10^{-2}}{1.7\times 10^{-3}}>10, (31)

and

t+=b1+b124b2b02b2<1b2<107.t_{+}=\frac{b_{1}+\sqrt{b_{1}^{2}-4b_{2}b_{0}}}{2b_{2}}<\frac{1}{b_{2}}<10^{7}. (32)

One clearly has,

g(0)=0,g(t)<0for sufficiently small t>0,andg(t)as t.g(0)=0,\quad g(t)<0\ \text{for sufficiently small }t>0,\quad\text{and}\qquad g(t)\to-\infty\ \text{as }t\to\infty.

Hence, if

g(t+)>0andg(t)<0,g(t_{+})>0\quad\text{and}\quad g(t_{-})<0, (33)

then t+t_{+} is the unique global maximizer, while tt_{-} corresponds to a local minimizer.

To verify (33), observe that since

b2t±2+b1t±b0=0,-b_{2}t_{\pm}^{2}+b_{1}t_{\pm}-b_{0}=0,

we have

g(t±)\displaystyle g(t_{\pm}) =13t±(b2t±2+32b1t±3b0)\displaystyle=\tfrac{1}{3}t_{\pm}\left(-b_{2}t_{\pm}^{2}+\tfrac{3}{2}b_{1}t_{\pm}-3b_{0}\right)
=13t±(b2t±2+b1t±b0+12b1t±2b0)\displaystyle=\tfrac{1}{3}t_{\pm}\left(-b_{2}t_{\pm}^{2}+b_{1}t_{\pm}-b_{0}+\tfrac{1}{2}b_{1}t_{\pm}-2b_{0}\right)
=13t±(12b1t±2b0)\displaystyle=\tfrac{1}{3}t_{\pm}\left(\tfrac{1}{2}b_{1}t_{\pm}-2b_{0}\right)
=16t±(b1t±4b0).\displaystyle=\tfrac{1}{6}t_{\pm}\left(b_{1}t_{\pm}-4b_{0}\right).

Therefore, establishing g(t+)>0g(t_{+})>0 and g(t)<0g(t_{-})<0 reduces to proving

b1t+4b0>0andb1t4b0<0,b_{1}t_{+}-4b_{0}>0\quad\text{and}\quad b_{1}t_{-}-4b_{0}<0,

respectively. The first inequality is immediate, since by (30), we have

b1t+4b0\displaystyle b_{1}t_{+}-4b_{0} =12b2(b12+b1b124b2b08b0b2)\displaystyle=\frac{1}{2b_{2}}\Big(b_{1}^{2}+b_{1}\sqrt{b_{1}^{2}-4b_{2}b_{0}}-8b_{0}b_{2}\Big)
12b2(9×104+61048×8.342×106)\displaystyle\geq\frac{1}{2b_{2}}\Big(9\times 10^{-4}+610^{-4}-8\times 8.34^{2}\times 10^{-6}\Big)
12b2(15×1045.57×104)>0.\displaystyle\geq\frac{1}{2b_{2}}\Big(15\times 10^{-4}-5.57\times 10^{-4}\Big)>0.

For the second one, denote γ=4b0b2.\gamma=4b_{0}b_{2}. Then,

b1t4b0\displaystyle b_{1}t_{-}-4b_{0} =12b2(b12b1b12γ2γ)\displaystyle=\frac{1}{2b_{2}}\left(b_{1}^{2}-b_{1}\sqrt{b_{1}^{2}-\gamma}-2\gamma\right)
=12b2(b12γb1b12γγ)\displaystyle=\frac{1}{2b_{2}}\left(b_{1}^{2}-\gamma-b_{1}\sqrt{b_{1}^{2}-\gamma}-\gamma\right)
=12b2(b12γ(b12γb1)γ)\displaystyle=\frac{1}{2b_{2}}\left(\sqrt{b_{1}^{2}-\gamma}\left(\sqrt{b_{1}^{2}-\gamma}-b_{1}\right)-\gamma\right)
<0,\displaystyle<0,

where the latter inequality follows since γ>0\gamma>0 and b12γb1<0\sqrt{b_{1}^{2}-\gamma}-b_{1}<0. Hence, g(t)<0,g(t_{-})<0, as desired.

From the above results and given the uniform estimates in (28) (which are independent of vK1v\in K_{1}), we deduce that for each vK1v\in K_{1}, the mapping

t(v)(t),t\mapsto\mathcal{E}(v)(t),

has exactly one maximum point on the positive semi-axis, that is tv=t+t_{v}=t_{+}, which by (31) and (32) is bounded away from both zero and infinity, more exactly, r0=10r_{0}=10 and R0=107.R_{0}=10^{7}. Hence, condition (h1) is verified.

Check of (h2). Note that for any nonnegative function uC[0,1]u\in C[0,1], we have

|T(u)|maxt[0,1]a001k(t,s)𝑑s=a0maxt[0,1]t(1t)2=a04.|T(u)|_{\infty}\geq\max_{t\in[0,1]}a_{0}\int_{0}^{1}k(t,s)\,ds=a_{0}\max_{t\in[0,1]}\frac{t(1-t)}{2}=\frac{a_{0}}{4}.

Since every uUbu\in U_{b} is nonnegative, it follows that

infuUb|T(u)|a04>0,\inf_{u\in U_{b}}|T(u)|_{\infty}\geq\frac{a_{0}}{4}>0,

and therefore condition (h2) is satisfied.

Check of (h3). Assume that

(tu,u,u|u|)=0,\mathcal{F}\left(tu,u,\frac{u}{|u|_{\infty}}\right)=0,

for some t>0t>0 and uK{0}.u\in K\setminus\{0\}. Then,

1|u|(|u|L22t|u|L22)=0,\frac{1}{|u|_{\infty}}\left(|u|^{2}_{L^{2}}-t|u|^{2}_{L^{2}}\right)=0,

which yields t=1t=1, so (h3) holds.

Consequently, applying Theorem 4, we deduce that TT has a fixed point in UbU_{b}.

Declarations

Data availability This manuscript has no associated data.

Conflict of interest

The authors have no conflict of interest to declare.

Related Posts

No results found.