This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

A bifurcation phenomenon for the critical Laplace and pp-Laplace equation in the ball

Francesca Dalbono
Matteo Franca
Andrea Sfecci
Dipartimento di Matematica e Informatica, Università degli Studi di Palermo, Via Archirafi 34, 90123 Palermo - Italy. email: francesca.dalbono@unipa.it Dipartimento di Matematica, Università degli Studi di Bologna, Piazza di porta san Donato 5 - 40126, Bologna - Italy. email: matteo.franca4@unibo.it Dipartimento di Matematica e Geoscienze, Università degli Studi di Trieste, Via Valerio 12/1, 34127, Trieste - Italy. email: asfecci@units.it.
(Received: date / Accepted: date)
Abstract

In this paper we show that the number of radial positive solutions of the following critical problem

{Δpu(x)+λ𝒦(|x|)u(x)|u(x)|q2=0,u(x)>0|x|<1,u(x)=0|x|=1,\begin{cases}\Delta_{p}u(x)+\lambda{\mathcal{K}}(|x|)\,u(x)\,|u(x)|^{q-2}=0\,,\\ \vskip 3.0pt plus 1.0pt minus 1.0pt\cr u(x)>0&\quad|x|<1,\\ \vskip 3.0pt plus 1.0pt minus 1.0pt\cr u(x)=0&\quad|x|=1,\end{cases}

where q=npnpq=\frac{np}{n-p}, 2nn+2p2\frac{2n}{n+2}\leq p\leq 2 and xnx\in{\mathbb{R}}^{n}, undergoes a bifurcation phenomenon. Namely, the problem admits one solution for any λ>0\lambda>0 if 𝒦{\mathcal{K}} is steep enough at 0, while it admits no solutions for λ\lambda small and two solutions for λ\lambda large if 𝒦{\mathcal{K}} is too flat at 0.

The existence of the second solution is new, even in the classical Laplace case. The proofs use Fowler transformation and dynamical systems tools.

Mathematics Subject Classification (MSC): 35J92; Secondary: 35J62, 35B33, 35B09, 34C45.
Keywords: Scalar curvature equation, bifurcation phenomena, radial solutions, order of flatness, Fowler transformation, invariant manifold, phase plane analysis

1 Introduction

In this paper we focus on positive radial solutions for the generalized scalar curvature equation

Δpu+𝒦(|x|)u|u|q2=0,\Delta_{p}u+{\mathcal{K}}(|x|)\,u|u|^{q-2}=0, (1)

where Δpu=div(u|u|p2)\Delta_{p}u=\textrm{div}(\nabla u|\nabla u|^{p-2}) denotes the pp-Laplace operator, xnx\in\mathbb{R}^{n}, 2n2+np22\,\frac{n}{2+n}\leq p\leq 2,   1<p<n1<p<n, and qq is the Sobolev critical exponent

q=p=npnp.q=p^{*}=\frac{np}{n-p}. (2)

The function 𝒦:[0,+[]0,+[{\mathcal{K}}:[0,+\infty[{}\to{}]0,+\infty[{} is assumed to be C1C^{1}, for simplicity. We are interested in crossing solutions, which means solutions of the problem

{Δpu(x)+𝒦(|x|)u(x)|u(x)|q2=0u(x)>0|x|<u(x)=0|x|=.\begin{cases}\Delta_{p}u(x)+{\mathcal{K}}(|x|)\,u(x)\,|u(x)|^{q-2}=0\\ \vskip 3.0pt plus 1.0pt minus 1.0pt\cr u(x)>0&\quad|x|<{\mathcal{R}}\\ \vskip 3.0pt plus 1.0pt minus 1.0pt\cr u(x)=0&\quad|x|={\mathcal{R}}.\end{cases} (3)

In particular, we will show that the existence of radial solutions of (3) depends on the behavior of 𝒦{\mathcal{K}} in a neighborhood of zero, and on the length of the radius {\mathcal{R}}.

Since we exclusively concentrate our study on radial solutions, we can reduce equation (1) to the following singular ordinary differential equation

(rn1u(r)|u(r)|p2)+rn1𝒦(r)u|u|q2=0,(r^{n-1}\,u^{\prime}(r)|u^{\prime}(r)|^{p-2})^{\prime}+r^{n-1}\,{\mathcal{K}}(r)\,u|u|^{q-2}=0\,, (4)

where “ \,{}^{\prime}\, ” denotes the differentiation with respect to r=|x|r=|x|, and, with a slight abuse of notation, u(r)=u(x)u(r)=u(x).

We say that a solution u(r)u(r) of (4) is regular if and only if limr0u(r)=d>0\lim_{r\to 0}u(r)=d>0 for a suitable d>0d>0: in this case it will be denoted by u(r;d)u(r;d). It is well known that u(r;d)u(r;d) exists and it is unique for any d>0d>0, cf. e.g. [25, 26, 33, 34]. Note that u(0;d)=0u^{\prime}(0;d)=0.

As an alternative, from a standard scaling argument, we can find a counterpart for the eigenvalue equation where we fix =1{\mathcal{R}}=1 for definiteness and we multiply the potential 𝒦{\mathcal{K}} by a parameter λ\lambda, i.e.

{Δpu(x)+λ𝒦(|x|)u(x)|u(x)|q2=0u(x)>0|x|<1u(x)=0|x|=1.\begin{cases}\Delta_{p}u(x)+\lambda{\mathcal{K}}(|x|)\,u(x)\,|u(x)|^{q-2}=0\\ \vskip 3.0pt plus 1.0pt minus 1.0pt\cr u(x)>0&\quad|x|<1\\ \vskip 3.0pt plus 1.0pt minus 1.0pt\cr u(x)=0&\quad|x|=1.\end{cases} (5)

Indeed, if we set

w(s)=u(r),s=rλ1/p,w(s)=u(r),\qquad s=\frac{r}{\lambda^{1/p}}, (6)

and 𝒦(s)=𝒦(sλ1/p)\mathscr{K}(s)={\mathcal{K}}(s\lambda^{1/p}), then uu solves equation (4) if and only if ww solves

(sn1w(s)|w(s)|p2)+λsn1𝒦(s)w|w|q2=0,(s^{n-1}\,w^{\prime}(s)|w^{\prime}(s)|^{p-2})^{\prime}+\lambda s^{n-1}\,\mathscr{K}(s)\,w|w|^{q-2}=0,

and w(1)=0w(1)=0 if and only if u()=0u({\mathcal{R}})=0, where :=λ1/p{\mathcal{R}}:=\lambda^{1/p}.

Hence, the number of solutions uu of (3) as {\mathcal{R}} varies coincides with the number of solutions uu of (5) as λ\lambda varies.

The scalar curvature equation (1) has been extensively studied in the literature due to its significance in a broad variety of applications, such as Riemannian geometry, astrophysics, quantum mechanic, chemistry, theory of non-Newtonian fluids and elasticity (cf. [21] for more detailed references on application of pp-Laplace equations and e.g. [7, 8] for application to Riemannian geometry in the p=2p=2 case).

In many phenomena, positivity of solutions has a physical relevance.

Non-linear eigenvalue problems similar to (5) are nowadays a classical topic, see e.g. [5] and the more recent [6, 15, 16, 28, 40] where the reaction term 𝒦(r)uq1{\mathcal{K}}(r)u^{q-1} is replaced by a sum of a linear and a term either critical or supercritical with respect to the Sobolev exponent. We address the interested reader to the introduction of [5, 15, 16] for a discussion of several possible diagrams appearing as different reaction terms are considered.

A key role in our analysis is played by the following hypothesis

(𝐇)\boldsymbol{({\rm H}_{\ell})}

There are some constants A,B,>0A,B,\ell>0 such that

𝒦(r)=A+Br+h(r)andlimr0|h(r)|+r|h(r)|r=0.{\mathcal{K}}(r)=A+Br^{\ell}+h(r)\qquad\mbox{and}\qquad\lim_{r\to 0}\frac{|h(r)|+r|h^{\prime}(r)|}{r^{\ell}}=0\,.

The existence and the multiplicity of the solutions of (3) and (5) depend crucially on \ell, which, in literature, is often referred to as the order of flatness of the function 𝒦{\mathcal{K}} at r=0r=0.

Problem (3) subject to condition (𝐇)\boldsymbol{({\rm H}_{\ell})} has been already investigated in the early 1990s by Bianchi-Egnell in [3] and by Lin-Lin in [39], who determined the existence of the critical value 2=n2\ell^{*}_{2}=n-2 in the Laplacian setting p=2p=2. In fact, Bianchi-Egnell and Lin-Lin have been able to prove the following result.

Theorem A [39].Assume that 𝒦{\mathcal{K}} satisfies (𝐇)\boldsymbol{({\rm H}_{\ell})} and consider 𝐩=𝟐\mathbb{\bf p=2} in (3).

  • (i)(i)

    If <n2\ell<n-2, then problem (3) admits a radial solution for every >0{\mathcal{R}}>0.

  • (ii)ii)

    If n2\ell\geq n-2, then there exists a sufficiently small constant r0>0r_{0}>0 such that problem (3) does not admit any radial solution when <r0{\mathcal{R}}<r_{0};

  • (iii(iii)

    If n2<nn-2\leq\ell<n, then there exists a sufficiently large constant R0>0{R}_{0}>0 such that problem (3) admits a radial solution for every R0{\mathcal{R}}\geq{R}_{0}.

In fact Bianchi and Egnell in [3] focused on the R=1R=1 case. In particular, following a shooting approach based on ordinary differential equations, they constructed and glued together two regular solutions of (4), one shooting from the zero initial condition and the other shooting from infinity.

The restriction <n\ell<n has often been adopted in the Laplacian literature to ensure the existence of a solution to problem (3). For instance, it appears in [37, Theorem 0.19], where the scalar curvature 𝒦{\mathcal{K}} is required to be strictly decreasing in a left neighbourhood of {\mathcal{R}}. The upper constraint <n\ell<n can be also found in the recent work [38], dealing with the σk\sigma_{k}-Nirenberg problem on the standard sphere 𝕊n\mathbb{S}^{n} for 2k<n/22\leq k<n/2. Among the very few examples of existence results for the Laplacian scalar curvature problem in the absence of upper bound conditions on \ell, we refer to the very recent papers [10, 43], where hypothesis (𝐇)\boldsymbol{({\rm H}_{\ell})} is combined with a (not so easily verifiable) topological global index formula on the critical points of 𝒦{\mathcal{K}}.

Note that in the pp-Laplace context, the condition <n\ell<n generalizes to <np1\ell<\frac{n}{p-1}, cf. [22].

We emphasize that Theorem A, besides its intrinsic interest, has been a key starting point in proving the existence of Ground States with fast decay, i.e. solutions u(r)u(r) positive for any r>0r>0 and decaying as r(n2)r^{-(n-2)} at infinity.

In fact, it can be shown that if 𝒦{\mathcal{K}} is increasing close to r=0r=0 and decreasing close to r=r=\infty we might expect to find Ground States with fast decay: in the 90s there was a flourishing of papers giving sufficient conditions for existence and non-existence of these solutions. A possible strategy is indeed to combine Theorem A, or similar results, with the use of Kelvin inversion, which transfers the information on regular solutions to fast decay solutions, see e.g. [3, 12]. Roughly speaking, one can expect that if 𝒦{\mathcal{K}} is steep enough at 0 (i.e. <n\ell<n) and at infinity 𝒦(r)a+br{\mathcal{K}}(r)\sim a+br^{-{\ell}} with <n\ell<n and a>0a>0, b>0b>0, then there is a Ground State with fast decay, while if these conditions are violated, one can construct a counterexample to Theorem A, see [3, Theorem 0.3]. For a generalization to the pp-Laplace setting, we also refer to [22], which follows a different strategy since Kelvin inversion is not available in that context.

We think it is worthwhile to point out that if A=0A=0 in (𝐇)\boldsymbol{({\rm H}_{\ell})}, then the problem becomes easier: roughly speaking, its solutions behave as the solutions of the A>0A>0, \ell-subcritical case, i.e. (3) admits a radial solution for any >0{\mathcal{R}}>0 (or, equivalently, (5) admits a radial solution for any λ>0\lambda>0), for any >0\ell>0. Again, combining this result with Kelvin inversion one might obtain the existence of Ground States with fast decay. This idea was extensively used in the 90s in many papers to handle the Laplacian problem, starting, probably, from [35], [9][44], and then it was developed and adapted to related problems, such as existence of Ground States with fast decay with a prescribed number of sign changes, see e.g. [45] and [33], dealing with the Laplacian and pp-Laplacian setting, respectively.

The aim of this paper is to improve Theorem A in 3 main directions.

Firstly, we extend the results to the pp-Laplace context, proving the existence of the generalized flatness order’s threshold

p=npp1,\ell^{*}_{p}=\frac{n-p}{p-1}\,,

which coincides with 2=n2\ell^{*}_{2}=n-2 of Theorem A for p=2p=2. As far as we are aware, these are the first results in this direction, although we have to require the probably technical condition 2n2+np2\frac{2n}{2+n}\leq p\leq 2. The possibility to remove this restriction will be the object of further investigations.

Secondly, we are able to remove the condition n\ell\leq n (i.e. np1\ell\leq\frac{n}{p-1} in the pp-Laplace context), by requiring that 𝒦(r){\mathcal{K}}(r) is increasing for any r>0r>0.

Thirdly, and probably more importantly, we are able to prove the existence of a second solution when >n2\ell>n-2 (i.e. >p\ell>\ell^{*}_{p} in the pp-Laplace context), thus completing the bifurcation diagram even in the p=2p=2 case.

Let us state our assumptions and the main results of the paper.

(𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)}

The function 𝒦(r){\mathcal{K}}(r) is increasing for any r0r\geq 0, strictly in some interval.

(𝐖𝒔)\boldsymbol{({\rm W}_{s})}

The function 𝒦(et){\mathcal{K}}({\rm e}^{t}) is uniformly continuous in [0,+[[0,+\infty[{} and there are K¯>K¯>0\overline{K}>\underline{K}>0 such that K¯<𝒦(r)<K¯\underline{K}<{\mathcal{K}}(r)<\overline{K} for any r>0r>0.

Note that if r𝒦(r)r{\mathcal{K}}^{\prime}(r) is bounded in [1,+[[1,+\infty[ then 𝒦(et){\mathcal{K}}({\rm e}^{t}) is uniformly continuous in [0,+[[0,+\infty[{}.

Theorem 1.

Assume that 𝒦{\mathcal{K}} satisfies (𝐇)\boldsymbol{({\rm H}_{\ell})}.

  • If <p\ell<\ell^{*}_{p}, then problem (3) admits a radial solution for every >0{\mathcal{R}}>0.

  • If p\ell\geq\ell^{*}_{p}, then there exists 0>0{\mathcal{R}}_{0}>0 such that problem (3) does not admit any radial solution when <0{\mathcal{R}}<{\mathcal{R}}_{0}.

Theorem 2.

Assume that 𝒦{\mathcal{K}} satisfies (𝐇)\boldsymbol{({\rm H}_{\ell})} and (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)}.

If =p\ell=\ell^{*}_{p}, then there exists 0>0{\mathcal{R}}_{0}>0 such that problem (3)

  • does not admit any radial solution when <0{\mathcal{R}}<{\mathcal{R}}_{0};

  • admits at least a radial solution when >0{\mathcal{R}}>{\mathcal{R}}_{0}.

If >p\ell>\ell^{*}_{p}, then there exists 0>0{\mathcal{R}}_{0}>0 such that problem (3)

  • does not admit any radial solution when <0{\mathcal{R}}<{\mathcal{R}}_{0};

  • admits at least a radial solution when =0{\mathcal{R}}={\mathcal{R}}_{0};

  • admits at least 22 radial solutions when >0{\mathcal{R}}>{\mathcal{R}}_{0}.

In fact, Theorem 2 holds also if we drop the global assumption (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)}, but we strengthen the requirement on 𝒦(0){\mathcal{K}}^{\prime}(0) by introducing the upper bound on the order of flatness np1\ell\leq\frac{n}{p-1} at zero, in the spirit of Theorem A. Unfortunately, we need to ask for some further very weak technical conditions on 𝒦{\mathcal{K}} for rr large.

Theorem 3.

Assume that 𝒦{\mathcal{K}} satisfies (𝐇)\boldsymbol{({\rm H}_{\ell})} with pnp1\ell^{*}_{p}\leq\ell\leq\frac{n}{p-1}. Assume further that either (𝐖𝐬)\boldsymbol{({\rm W}_{s})} holds or there is ρ1>0\rho_{1}>0 such that 𝒦(r)𝒦(ρ1){\mathcal{K}}(r)\geq{\mathcal{K}}(\rho_{1}) when rρ1r\geq\rho_{1}, then we get the same conclusion as in Theorem 2.

Using the change of variable (6), we can rewrite Theorems 12, and 3 as follows.

Corollary 4.

Assume that 𝒦{\mathcal{K}} satisfies (𝐇)\boldsymbol{({\rm H}_{\ell})}.

  • If <p\ell<\ell^{*}_{p}, then problem (5) admits a radial solution for every λ>0\lambda>0.

  • If p\ell\geq\ell^{*}_{p}, then there exists λ0>0\lambda_{0}>0 such that problem (5) does not admit any radial solution when λ<λ0\lambda<\lambda_{0}.

Corollary 5.

Assume that 𝒦{\mathcal{K}} satisfies (𝐇)\boldsymbol{({\rm H}_{\ell})} and (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)}.

If =p\ell=\ell^{*}_{p}, then there exists λ0>0\lambda_{0}>0 such that problem (5)

  • does not admit any radial solution when λ<λ0\lambda<\lambda_{0};

  • admits at least a radial solution when λ>λ0\lambda>\lambda_{0}.

If >p\ell>\ell^{*}_{p}, then there exists λ0>0\lambda_{0}>0 such that problem (5)

  • does not admit any radial solution when λ<λ0\lambda<\lambda_{0};

  • admits at least a radial solution when λ=λ0\lambda=\lambda_{0};

  • admits at least two radial solutions when λ>λ0\lambda>\lambda_{0}.

Again, according to Theorem 3, we can drop the global condition (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)} and get the same result by restricting the interval in which \ell varies and by imposing some further weak asymptotic conditions on 𝒦(r){\mathcal{K}}(r) for rr large.

Corollary 6.

Assume that 𝒦{\mathcal{K}} satisfies (𝐇)\boldsymbol{({\rm H}_{\ell})} with pnp1\ell^{*}_{p}\leq\ell\leq\frac{n}{p-1}; assume further that either (𝐖𝐬)\boldsymbol{({\rm W}_{s})} holds or there is ρ1>0\rho_{1}>0 such that 𝒦(r)𝒦(ρ1){\mathcal{K}}(r)\geq{\mathcal{K}}(\rho_{1}) for any rρ1r\geq\rho_{1}. Then, we get the same conclusions as in Corollary 5.

In fact via Theorem 3 and Corollary 6 we are also able to extend Theorem A to the case where =np1\ell=\frac{n}{p-1}, which was not covered by [3, 39], with the addition of a very weak technical condition which, roughly speaking, is satisfied unless 𝒦(r){\mathcal{K}}(r) is subject to wild oscillations for rr large, or converges to 0 for rr large.

Moreover, our methods are considerably different from those of Lin-Lin in [39]. Our proofs are based on the Fowler transformation, which discloses the geometrical aspect of our problem by converting the singular ordinary differential equation (4) into an equivalent dynamical system, cf. (11), following the way paved by [32, 31, 30] and later on by [1, 2, 23]. Then, we develop a detailed phase plane analysis, involving invariant manifold theory for non-autonomous systems, energy estimates, comparisons of the non-autonomous planar system with suitable autonomous ones and a Grönwall’s argument.

1.1 On the proofs of the theorems

We briefly sketch the plan of our proof. Let

J={d>0u(r;d) is a crossing solution},J=\{d>0\mid u(r;d)\;\textrm{ is a crossing solution}\}, (7)

and denote by R(d)R(d) the first zero of u(r;d)u(r;d) when dJd\in J: our argument relies on a study of the properties of the function R:J]0,+[R:J\to{}]0,+\infty[{}.

Using some classical results (see [14, 36] for the Laplacian case, and [26, 34] for pp-Laplacian extensions), we know that J=]0,+[J={\color[rgb]{1,0,0}}]0,+\infty[{} under assumption (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)}. Then, using a standard transversality argument, it is straightforwardly proved that R(d)R(d) is continuous, see Proposition 22 below.

Then, it is not difficult to show, see, e.g. [33, Proposition 2.4] that

limd0R(d)=+.\lim_{d\to 0}R(d)=+\infty\,. (8)

Refer to caption     Refer to caption

Refer to caption

Figure 1: A sketch of the graph of the function R(d)R(d), when (𝐇)\boldsymbol{({\rm H}_{\ell})} and (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)} are assumed. In the critical case =p\ell=\ell^{*}_{p}, we have three possible alternatives.

The focus and the original part of our study consists in analyzing the asymptotic behaviour of the solutions with large initial data dd, which is determined by the parameter \ell in (𝐇)\boldsymbol{({\rm H}_{\ell})}. In particular, we have the following bifurcation phenomenon, when (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)} is assumed

{limd+R(d)=0, if 0<<p,lim infd+R(d)>0, if =p,limd+R(d)=+, if >p,\begin{cases}\displaystyle\lim_{d\to+\infty}R(d)=0\,,&\mbox{ if }\displaystyle{0<\ell<\ell^{*}_{p}}\,,\\ \displaystyle\liminf_{d\to+\infty}R(d)>0\,,&\mbox{ if }\displaystyle{\ell=\ell^{*}_{p}}\,,\\ \displaystyle\lim_{d\to+\infty}R(d)=+\infty\,,&\mbox{ if }\displaystyle{\ell>\ell^{*}_{p}}\,,\\ \end{cases} (9)

see Propositions 37 and 43 below. This asymptotic analysis will allow us to draw the diagrams in Figure 1.

Since the number of solutions of problem (3) is the number of points of the preimage R1()R^{-1}({\mathcal{R}}), the proof of Theorem 2 is immediately given.

Actually, using a truncation argument, see Remark 25, we are able to get the first estimate in (9), also when assumption (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)} is dropped, cf. Proposition 37. Hence, the proof of Theorem 1 in the subcritical case follows.

In fact, in the p=2p=2 case, a part of (9) has been already shown in [39, Theorem 1.6], see also [7, Remark 4.1]:

{limd+R(d)=0 if <2,r0>0:R(d)r0,d>0 if 2.\begin{cases}\displaystyle\lim_{d\to+\infty}R(d)=0&\mbox{\, if \,}\ell<\ell^{*}_{2}\,,\\ \vskip 3.0pt plus 1.0pt minus 1.0pt\cr\exists\,r_{0}>0\,:\,R(d)\geq r_{0}\,,\,\forall\,d>0&\mbox{\, if \,}\ell\geq\ell^{*}_{2}.\end{cases}

Refer to caption     Refer to caption

Refer to caption

Figure 2: A sketch of the graph of the function R(d)R(d), in the setting of Theorem 3. In the critical case =p\ell=\ell^{*}_{p}, we have three possible alternatives.

Finally, we adapt our analysis to the context where the global requirement (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)} is replaced by the local requirement np1\ell\leq\frac{n}{p-1}, and we reprove the existence of the second solution in the supercritical case. As shown in Lemma 45 below, the restriction np1\ell\leq\frac{n}{p-1}, and the technical requirement that either (𝐖𝒔)\boldsymbol{({\rm W}_{s})} holds or there is ρ1>0\rho_{1}>0 such that 𝒦(r)𝒦(ρ1){\mathcal{K}}(r)\geq{\mathcal{K}}(\rho_{1}) when rρ1r\geq\rho_{1} are needed just in order to ensure that there is D^0\hat{D}\geq 0 such that ]D^,+[J{}]\hat{D},+\infty[{}\subset J, with JJ as in (7).

Then, using classical arguments, we see that if D^J\hat{D}\not\in J, then

limdD^R(d)=+,\lim_{d\to\hat{D}}R(d)=+\infty,

and, adapting the computation performed when (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)} holds, we prove that R(d)R(d) satisfies (9) if np1\ell\leq\frac{n}{p-1}, see Propositions 44 and 46 below. As a consequence, we are able to draw the diagrams for R(d)R(d) in Figure 2, from which Theorem 3 follows.

Notice that if D^>0\hat{D}>0, then u(r;D^)u(r;\hat{D}) is a Ground State, that is a solution of (4), positive for any r>0r>0.

The paper is organized as follows. In §2 we introduce the Fowler transformation, i.e. the change of variables (2) which turns (4) into the planar, non-autonomous dynamical system (11); then we recall some basic tools in this context. In particular, we review some aspects of invariant manifold theory for non-autonomous systems, and we introduce the unstable leaves Wu(τ)W^{u}(\tau) of (11) which correspond to regular solutions of (4). In §3 we establish some standard properties of the function R(d)R(d), such as continuity and asymptotic properties close to d=0d=0 and to d=D^d=\hat{D}, when u(r;D^)u(r;\hat{D}) is a Ground State. The core of our argument is in §4, where we prove the asymptotic properties of R(d)R(d) as dd tends to infinity: in §4.1 we focus on the subcritical case <p\ell<\ell^{*}_{p}, and in §4.2 we consider the critical and supercritical setting p\ell\geq\ell^{*}_{p}. In §5 we briefly conclude the proofs of the main results of the paper.

2 Fowler transformation,
basic notation and preliminaries

Let us introduce a change of variables known as Fowler transformation, which allows to transform (4) into a two-dimensional dynamical system. We define

α=npp,β=n(p1)p,\displaystyle\alpha=\displaystyle{\frac{n-p}{p}},\qquad\beta=\displaystyle{\frac{n(p-1)}{p}},
x=u(r)rαy=u(r)|u(r)|p2rβr=et.\displaystyle x=u(r)r^{\alpha}\qquad y=u^{\prime}(r)|u^{\prime}(r)|^{p-2}r^{\beta}\qquad r={\rm e}^{t}. (10)

This change of variable is known from the ’30s, see [17], and it has been generalized to the pp-Laplacian case by Bidaut-Véron [4] and some years later (independently) by Franca, see e.g. [18, 19, 20, 22, 23].
According to (2), we can rewrite (4) as the following dynamical system:

(x˙y˙)=(α00α)(xy)+(y|y|2pp1K(t)x|x|q2),\left(\begin{array}[]{c}\dot{x}\\ \vskip 6.0pt plus 2.0pt minus 2.0pt\cr\dot{y}\end{array}\right)=\left(\begin{array}[]{cc}\alpha&0\\ \vskip 6.0pt plus 2.0pt minus 2.0pt\cr 0&-\alpha\end{array}\right)\left(\begin{array}[]{c}x\\ \vskip 6.0pt plus 2.0pt minus 2.0pt\cr y\end{array}\right)+\left(\begin{array}[]{c}y\,|y|^{\frac{2-p}{p-1}}\\ \vskip 6.0pt plus 2.0pt minus 2.0pt\cr-K(t)\,x|x|^{q-2}\end{array}\right), (11)

where K(t)=𝒦(et)K(t)={\mathcal{K}}({\rm e}^{t}), q=p=npnpq=p^{*}=\frac{np}{n-p} as in (2), and “\cdot” denotes the differentiation with respect to tt.

Remark 7.

Note that system (11) is C1C^{1} if and only if 2n2+np2\frac{2n}{2+n}\leq p\leq 2.

Given the initial data τ\tau\in{\mathbb{R}} and 𝑸2{\boldsymbol{Q}}\in{\mathbb{R}}^{2}, we will denote by

ϕ(t;τ,𝑸)=(x(t;τ,𝑸),y(t;τ,𝑸))\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}})=\big{(}x(t;\tau,{\boldsymbol{Q}}),y(t;\tau,{\boldsymbol{Q}})\big{)} (12)

the trajectory of (11) such that ϕ(τ;τ,𝑸)=𝑸\boldsymbol{\phi}(\tau;\tau,{\boldsymbol{Q}})={\boldsymbol{Q}}.

Define the energy function

(x,y;t):=αxy+p1p|y|pp1+K(t)|x|qq.{\mathcal{H}}(x,y;t):=\alpha xy\,+\,\frac{p-1}{p}\,|y|^{\frac{p}{p-1}}\,+\,K(t)\,\frac{|x|^{q}}{q}\,. (13)

If we evaluate \mathcal{H} along a solution (x(t),y(t))(x(t),y(t)) of (11), we obtain the associated Pohozaev type energy (x(t),y(t);t)\mathcal{H}(x(t),y(t);t), whose derivative with respect to tt satisfies

ddt(x(t),y(t);t)=K˙(t)|x(t)|qq.\frac{d}{dt}{\mathcal{H}}\big{(}x(t),y(t);t\big{)}=\dot{K}(t)\,\frac{|x(t)|^{q}}{q}\,. (14)

We also need to consider the autonomous system obtained by freezing the tt-dependence of K()K(\cdot). In particular, fixed τ{±}\tau\in{\mathbb{R}}\cup\{\pm\infty\}, we consider the frozen autonomous system:

(x˙y˙)=(α00α)(xy)+(y|y|2pp1K(τ)x|x|q2),\left(\begin{array}[]{c}\dot{x}\\ \vskip 6.0pt plus 2.0pt minus 2.0pt\cr\dot{y}\end{array}\right)=\left(\begin{array}[]{cc}\alpha&0\\ \vskip 6.0pt plus 2.0pt minus 2.0pt\cr 0&-\alpha\end{array}\right)\left(\begin{array}[]{c}x\\ \vskip 6.0pt plus 2.0pt minus 2.0pt\cr y\end{array}\right)+\left(\begin{array}[]{c}y\,|y|^{\frac{2-p}{p-1}}\\ \vskip 6.0pt plus 2.0pt minus 2.0pt\cr-K(\tau)\,x|x|^{q-2}\end{array}\right), (15)

where 𝒦{\mathcal{K}} is assumed to have finite limit, whenever τ=±\tau=\pm\infty.

If we differentiate the energy (;τ)\mathcal{H}(\cdot;\tau) of system (15) along a solution (x(t),y(t))(x(t),y(t)) of the dynamical system (11), we get

ddt(x(t),y(t);τ)=(K(τ)K(t))x(t)|x(t)|q2x˙(t).\frac{d}{dt}{\mathcal{H}}\big{(}x(t),y(t);\tau\big{)}=\big{(}K(\tau)-K(t)\big{)}\,x(t)|x(t)|^{q-2}\,\dot{x}(t)\,. (16)

Let us fix τ{±}\tau\in{\mathbb{R}}\cup\{\pm\infty\} in (15), let τ0\tau_{0}\in{\mathbb{R}} and 𝑸2{\boldsymbol{Q}}\in{\mathbb{R}}^{2}; we denote by

ϕ𝝉(t;τ0,𝑸)=(xτ(t;τ0,𝑸),yτ(t;τ0,𝑸))\boldsymbol{\phi_{\tau}}(t;\tau_{0},{\boldsymbol{Q}})=\big{(}x_{\tau}(t;\tau_{0},{\boldsymbol{Q}}),y_{\tau}(t;\tau_{0},{\boldsymbol{Q}})\big{)} (17)

the trajectory of (15), such that ϕ𝝉(τ0;τ0,𝑸)=𝑸\boldsymbol{\phi_{\tau}}(\tau_{0};\tau_{0},{\boldsymbol{Q}})={\boldsymbol{Q}}.

System (15) exhibits an equilibrium point in {x>0,y<0}\{x>0\,,y<0\}, of coordinates

𝑬(τ)=(Ex(τ),Ey(τ))=((αpK(τ))1qp,(αqK(τ))p1qp).{\boldsymbol{E}}(\tau)=(E_{x}(\tau),E_{y}(\tau))=\left(\left(\frac{\alpha^{p}}{K(\tau)}\right)^{\frac{1}{q-p}}\,,\,-\left(\frac{\alpha^{q}}{K(\tau)}\right)^{\frac{p-1}{q-p}}\right). (18)

It is also well known (see, among others, [22]) that, for any fixed τ{±}\tau\in{\mathbb{R}}\cup\{\pm\infty\}, system (15) admits a homoclinic orbit (see Figure 3)

𝚪𝝉={(x,y)(x,y;τ)=0,x>0},\boldsymbol{\Gamma_{\tau}}=\{(x,y)\mid{\mathcal{H}}(x,y;\tau)=0\,,x>0\}\,, (19)

recalling that the case τ=+\tau=+\infty requires a boundedness restriction on 𝒦{\mathcal{K}}. Taking into account that the flows of (11) and (15) are ruled by their linear part near the origin, we easily deduce that the origin is a saddle-type critical point for (11) and (15). Finally, notice that 𝑬(τ){\boldsymbol{E}}(\tau) is a centre and it lies in the interior of the region enclosed by 𝚪𝝉\boldsymbol{\Gamma_{\tau}}.

According to [23, 27], we also know the exact expression of the homoclinic trajectories ϕ=(𝒙,𝒚)\boldsymbol{\phi^{*}}=(\boldsymbol{x^{*}},\boldsymbol{y^{*}}) of (15). In particular, the one corresponding to the regular Ground State uu^{*} such that u(0)=du^{*}(0)=d satisfies

x(t)=d[et+C(d)etp1]npp,x^{*}(t)=d\left[{\rm e}^{-t}+C(d)\,{\rm e}^{\frac{t}{p-1}}\right]^{-\frac{n-p}{p}}\,, (20)

where C(d)>0C(d)>0 is a computable constant, see [27].
Moreover, fixed 𝑷𝚪\boldsymbol{P}\in\boldsymbol{\Gamma_{-\infty}} there exists a constant c𝑷>0c_{\boldsymbol{P}}>0 such that

sup{ϕ(θ+τ;τ,𝑸)eαθθ0,τ0,𝑸𝚪~𝑷}c𝑷,\sup\{\|\boldsymbol{\phi_{-\infty}}(\theta+\tau;\tau,\boldsymbol{Q})\|{\rm e}^{-\alpha\theta}\mid\theta\leq 0,\;\tau\leq 0,\;\boldsymbol{Q}\in\boldsymbol{\widetilde{\Gamma}_{-\infty}^{P}}\}\leq c_{\boldsymbol{P}}, (21)

where 𝚪~𝑷:=ϕ(],0];0,𝑷)\boldsymbol{\widetilde{\Gamma}_{-\infty}^{P}}:=\boldsymbol{\phi_{-\infty}}({}]-\infty,0];0,\boldsymbol{P}).

Let us now list some immediate consequences of assumption (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)}.

Remark 8.

Assume K(τ2)>K(τ1)K(\tau_{2})>K(\tau_{1}), then the homoclinic orbit 𝚪𝛕𝟐\boldsymbol{\Gamma_{\tau_{2}}} belongs to the region enclosed by 𝚪𝛕𝟏\boldsymbol{\Gamma_{\tau_{1}}}, cf. Figure 3.

Refer to caption
Refer to caption
Figure 3: On the left, the energy levels of {\mathcal{H}} at fixed time τ\tau, with the homoclinic orbit 𝚪𝝉\boldsymbol{\Gamma_{\tau}}. On the right, the position of the set 𝚪𝝉𝟐\boldsymbol{\Gamma_{\tau_{2}}} with respect to the set 𝚪𝝉𝟏\boldsymbol{\Gamma_{\tau_{1}}} in the case K(τ2)>K(τ1)K(\tau_{2})>K(\tau_{1}).
Lemma 9.

Assume (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)}. If u(r;d)u(r;d) is a regular solution, then the corresponding trajectory ϕ(;d)=(x(;d),y(;d))\boldsymbol{\phi}(\cdot;d)=(x(\cdot;d),y(\cdot;d)) of system (11) satisfies

limt(ϕ(t;d);t)=0,\lim_{t\to-\infty}{\mathcal{H}}(\boldsymbol{\phi}(t;d);t)=0\,,

and

(ϕ(t;d);t)0 for every tT(d){\mathcal{H}}(\boldsymbol{\phi}(t;d);t)\geq 0\quad\mbox{ for every }t\leq T(d)\,

(the equality occurs only when K(t)=K()K(t)=K(-\infty) holds), where T(d):=sup{τ:x(t;d)>0,tτ}T(d):=\sup\{\tau\in{\mathbb{R}}:\,x(t;d)>0\,,\forall t\leq\tau\} is the first zero of x(;d)x(\cdot;d). Moreover,

(ϕ(t;d);τ)(ϕ(t;d);t)0 for every t<min{τ,T(d)}{\mathcal{H}}(\boldsymbol{\phi}(t;d);\tau)\geq{\mathcal{H}}(\boldsymbol{\phi}(t;d);t)\geq 0\quad\mbox{ for every }t<\min\{\tau,T(d)\}

(the first equality occurs only when K(t)=K(τ)K(t)=K(\tau) holds).

Proof.

The result immediately follows from (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)}, (14), and (13). ∎

As an immediate consequence we have the following.

Lemma 10.

Assume (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)} and fix d>0d>0. Then, for every tT(d)t\leq T(d), the trajectory ϕ(t;d)\boldsymbol{\phi}(t;d) of system (11) either lies in the exterior of the region enclosed by 𝚪𝐭\boldsymbol{\Gamma_{t}}, or it lies on 𝚪𝐭\boldsymbol{\Gamma_{t}} if K(t)=K()K(t)=K(-\infty).

The next part of the section is devoted to explore the dynamical system (11) through an invariant manifold approach. From [23, 29, 31], and [11, §13.4], we know that the existence of the unstable manifold is ensured by the following condition:

(𝐖𝒖)\boldsymbol{({\rm W}_{u})}

The function KK is bounded and uniformly continuous in ],0]{}]-\infty,0] and limtK(t)=K()0\displaystyle{{\lim_{t\to-\infty}}K(t)=K(-\infty)\geq 0} is finite.

Notice that (𝐖𝒖)\boldsymbol{({\rm W}_{u})} is always satisfied when (𝐇)\boldsymbol{({\rm H}_{\ell})} holds.

Correspondingly, the existence of the stable manifold is guaranteed by (𝐖𝒔)\boldsymbol{({\rm W}_{s})}. The unstable manifold will play a key role in this paper, since we will see via Remark 17 that regular solutions u(r;d)u(r;d) correspond to trajectories of (11) converging to the origin as tt\to-\infty, i.e. leaving from the unstable manifold.

Let B(𝟎,δ)B(\boldsymbol{0},\delta) denote the open ball of radius δ\delta, centered at the origin. Following [31, Theorem 2.1], which is in fact a rewording of [29, Theorem 2.25], and its reformulation in[24, Appendix], we find the next result.

Theorem 11.

Assume (𝐖𝐮)\boldsymbol{({\rm W}_{u})}, then there is δ>0\delta>0 such that for any τ0\tau\leq 0 the set

Mlocu(τ):={𝑸ϕ(t;τ,𝑸)B(𝟎,δ)for any tτ}M^{u}_{\textrm{loc}}(\tau):=\{{\boldsymbol{Q}}\mid\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}})\in B(\boldsymbol{0},\delta)\;\textrm{for any $t\leq\tau$}\}

is a C1C^{1} embedded manifold tangent to the xx-axis at the origin if 2n2+n<p2\frac{2n}{2+n}<p\leq 2, and to the line y=K()nxy=-\frac{K(-\infty)}{n}\,x if p=2n2+np=\frac{2n}{2+n}. Further, let {\mathcal{L}} be a segment transversal to Mlocu(τ)M^{u}_{\textrm{loc}}(\tau), then Mlocu(τ)={𝐐𝐮(τ)}M^{u}_{\textrm{loc}}(\tau)\cap{\mathcal{L}}=\{\boldsymbol{Q^{u}}(\tau)\} is a singleton, and 𝐐𝐮(τ)\boldsymbol{Q^{u}}(\tau) is uniformly continuous in τ\tau.

Assume (𝐖𝐬)\boldsymbol{({\rm W}_{s})}, then there is δ>0\delta>0 such that for any τ0\tau\geq 0 the set

Mlocs(τ):={𝑸ϕ(t;τ,𝑸)B(𝟎,δ)for any tτ}M^{s}_{\textrm{loc}}(\tau):=\{{\boldsymbol{Q}}\mid\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}})\in B(\boldsymbol{0},\delta)\;\textrm{for any $t\geq\tau$}\}

is a C1C^{1} embedded manifold tangent to the yy-axis at the origin if 2n2+np<2\frac{2n}{2+n}\leq p<2 and to the line y=(n2)xy=-(n-2)x if p=2p=2. Further, let {\mathcal{L}} be a segment transversal to Mlocs(τ)M^{s}_{\textrm{loc}}(\tau), then Mlocs(τ)={𝐐𝐬(τ)}M^{s}_{\textrm{loc}}(\tau)\cap{\mathcal{L}}=\{\boldsymbol{Q^{s}}(\tau)\} is a singleton, and 𝐐𝐬(τ)\boldsymbol{Q^{s}}(\tau) is uniformly continuous in τ\tau.

This result may also be obtained by using the simpler approach developed in [11, §13.4], with the exception of the part concerning the uniform continuity of 𝑸𝒖()\boldsymbol{Q^{u}}(\cdot) and 𝑸𝒔()\boldsymbol{Q^{s}}(\cdot). Actually, [31, Theorem 2.1] ensures that 𝑸𝒖(τ)\boldsymbol{Q^{u}}(\tau) and 𝑸𝒔(τ)\boldsymbol{Q^{s}}(\tau) are C1C^{1} if K(t)K(t) is C1C^{1}.

We notice that Mlocu(τ)M^{u}_{\textrm{loc}}(\tau) is split by the origin in two connected components. Since we are just interested in positive solutions, we denote by Wlocu(τ)W^{u}_{\textrm{loc}}(\tau) and Wlocu,(τ)W^{u,-}_{\textrm{loc}}(\tau) the components lying respectively in x>0x>0 and x<0x<0. Similarly, Mlocs(τ)M^{s}_{\textrm{loc}}(\tau) is split by the origin in two connected components, say Wlocs(τ)W^{s}_{\textrm{loc}}(\tau) and Wlocs,(τ)W^{s,-}_{\textrm{loc}}(\tau), lying in x>0x>0 and x<0x<0, respectively.
Following again either [29, Theorem 2.25] or [11, §13.4] and in particular Theorem 4.5, we deduce some crucial properties concerning the uniform exponential asymptotic behaviour of the trajectories intersecting the manifolds Wlocu(τ)W^{u}_{\textrm{loc}}(\tau) and Wlocs(τ)W^{s}_{\textrm{loc}}(\tau).

Theorem 12.

Assume (𝐖𝐮)\boldsymbol{({\rm W}_{u})} and (𝐖𝐬)\boldsymbol{({\rm W}_{s})}, then the sets

Wlocu(τ1):={𝑸Mlocu(τ1)x(t;τ1,𝑸)>0for any tτ1}{(0,0)},W^{u}_{\textrm{loc}}(\tau_{1}):=\{{\boldsymbol{Q}}\in M^{u}_{\textrm{loc}}(\tau_{1})\mid x(t;\tau_{1},{\boldsymbol{Q}})>0\;\textrm{for any $t\leq\tau_{1}$}\}\cup\{(0,0)\},
Wlocs(τ2):={𝑸Mlocs(τ2)x(t;τ2,𝑸)>0for any tτ2}{(0,0)}W^{s}_{\textrm{loc}}(\tau_{2}):=\{{\boldsymbol{Q}}\in M^{s}_{\textrm{loc}}(\tau_{2})\mid x(t;\tau_{2},{\boldsymbol{Q}})>0\;\textrm{for any $t\geq\tau_{2}$}\}\cup\{(0,0)\}

are C1C^{1} embedded manifold for any τ10τ2\tau_{1}\leq 0\leq\tau_{2}. Furthermore, if 𝐐𝐮Wlocu(τ1)\boldsymbol{Q^{u}}\in W^{u}_{\textrm{loc}}(\tau_{1}) and 𝐐𝐬Wlocs(τ2)\boldsymbol{Q^{s}}\in W^{s}_{\textrm{loc}}(\tau_{2}), then the limits

limtϕ(t;τ1,𝑸𝒖)eαt,limt+ϕ(t;τ2,𝑸𝒔)eαt{\lim_{t\to-\infty}}\|\boldsymbol{\phi}(t;\tau_{1},\boldsymbol{Q^{u}})\|{\rm e}^{-\alpha t}\,,\qquad{\lim_{t\to+\infty}}\|\boldsymbol{\phi}(t;\tau_{2},\boldsymbol{Q^{s}})\|{\rm e}^{\alpha t}

are positive and finite. Further, there is c0=c0(δ)c_{0}=c_{0}(\delta) such that

sup{ϕ(θ+τ1;τ1,𝑸𝒖)eαθθ0,τ10,𝑸𝒖Wlocu(τ1)}c0(δ),\sup\{\|\boldsymbol{\phi}(\theta+\tau_{1};\tau_{1},\boldsymbol{Q^{u}})\|{\rm e}^{-\alpha\theta}\mid\theta\leq 0,\;\tau_{1}\leq 0,\;\boldsymbol{Q^{u}}\in W^{u}_{\textrm{loc}}(\tau_{1})\}\leq c_{0}(\delta),
sup{ϕ(θ+τ2;τ2,𝑸𝒔)eαθθ0,τ20,𝑸𝒔Wlocs(τ2)}c0(δ).\sup\{\|\boldsymbol{\phi}(\theta+\tau_{2};\tau_{2},\boldsymbol{Q^{s}})\|\,{\rm e}^{\alpha\theta}\,\,\,\mid\theta\geq 0,\;\tau_{2}\geq 0,\;\boldsymbol{Q^{s}}\in W^{s}_{\textrm{loc}}(\tau_{2})\}\leq c_{0}(\delta).

Using the flow of (11), with a standard argument, we pass from the local manifolds Wlocu(τ1)W^{u}_{\textrm{loc}}(\tau_{1}) and Wlocs(τ2)W^{s}_{\textrm{loc}}(\tau_{2}), defined for τ10τ2\tau_{1}\leq 0\leq\tau_{2}, to the global manifolds Wu(τ)W^{u}(\tau) and Ws(τ)W^{s}(\tau) defined for any τ\tau\in{\mathbb{R}}. In fact, rephrasing [24, Appendix], we set

Wu(τ):=T0{ϕ(τ;T,𝑸)𝑸Wlocu(T)},Ws(τ):=T0{ϕ(τ;T,𝑸)𝑸Wlocs(T)}.\begin{split}W^{u}(\tau):=&\displaystyle{\bigcup_{T\leq 0}}\big{\{}\boldsymbol{\phi}(\tau;T,{\boldsymbol{Q}})\mid{\boldsymbol{Q}}\in W^{u}_{\textrm{loc}}(T)\big{\}},\\ \vskip 3.0pt plus 1.0pt minus 1.0pt\cr W^{s}(\tau):=&\displaystyle{\bigcup_{T\geq 0}\big{\{}\boldsymbol{\phi}(\tau;T,{\boldsymbol{Q}})\mid{\boldsymbol{Q}}\in W^{s}_{\textrm{loc}}(T)\big{\}}}.\end{split} (22)

We observe that Wu(τ)W^{u}(\tau) and Ws(τ)W^{s}(\tau), the unstable and stable leaves respectively, are C1C^{1} immersed manifolds which can be characterized as follows

Wu(τ):={𝑸limtϕ(t;τ,𝑸)=(0,0),x(t;τ,𝑸)0 when t0},Ws(τ):={𝑸limt+ϕ(t;τ,𝑸)=(0,0),x(t;τ,𝑸)0 when t0}.\begin{split}W^{u}(\tau):=&\Big{\{}{\boldsymbol{Q}}\mid\lim_{t\to-\infty}\,\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}})=(0,0),\;x(t;\tau,{\boldsymbol{Q}})\geq 0\textrm{ when $t\ll 0$}\Big{\}}\,,\\ W^{s}(\tau):=&\Big{\{}{\boldsymbol{Q}}\mid\lim_{t\to+\infty}\,\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}})=(0,0),\;x(t;\tau,{\boldsymbol{Q}})\geq 0\textrm{ when $t\gg 0$}\Big{\}}\,.\end{split} (23)

By construction, if 𝑸Wu(τ){\boldsymbol{Q}}\in W^{u}(\tau), then ϕ(t;τ,𝑸)Wu(t)\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}})\in W^{u}(t) for any tt\in{\mathbb{R}}.

From Theorem 11 and the smooth dependence of the flow of (11) on initial data, we get the smoothness property of the unstable and stable manifold.

Remark 13.

Assume (𝐖𝐮)\boldsymbol{({\rm W}_{u})}, then Wu(τ)W^{u}(\tau) depends continuously on τ\tau. Namely, let {\mathcal{L}} be a segment which intersects Wu(τ0)W^{u}(\tau_{0}) transversely in a point 𝐐(τ0){\boldsymbol{Q}}(\tau_{0}), then there is a neighborhood II of τ0\tau_{0} such that Wu(τ)W^{u}(\tau) intersects {\mathcal{L}} in a point 𝐐(τ){\boldsymbol{Q}}(\tau) for any τI\tau\in I, and 𝐐(){\boldsymbol{Q}}(\cdot) is continuous. Actually, it has the same regularity as (11), so it is C1C^{1} if (11) is C1C^{1} in tt. Analogously if (𝐖𝐬)\boldsymbol{({\rm W}_{s})} holds, then Ws(τ)W^{s}(\tau) depends continuously on τ\tau, and smoothly if (11) is smooth in tt.

Let us denote by Wu()W^{u}(-\infty) the unstable manifold 𝚪\boldsymbol{\Gamma_{-\infty}}, defined in (19).
According to [1] and [24, §2.2], the smoothness property of Wu(τ)W^{u}(\tau) observed in Remark 13 can be extended to τ0=\tau_{0}=-\infty.

Remark 14.

Assume (𝐖𝐮)\boldsymbol{({\rm W}_{u})}, then Wu(τ)W^{u}(\tau) depends smoothly on τ{}\tau\in{\mathbb{R}}\cup\{-\infty\}.
In particular, let {\mathcal{L}} be a segment transversal to 𝚪\boldsymbol{\Gamma_{-\infty}} and let 𝐐()\boldsymbol{Q}_{{\mathcal{L}}}(-\infty) be the intersection point between {\mathcal{L}} and 𝚪\boldsymbol{\Gamma_{-\infty}}; follow Wu(τ)W^{u}(\tau) from the origin towards x>0x>0, then it intersects {\mathcal{L}} transversely in a point, say 𝐐(τ){\boldsymbol{Q}}_{{\mathcal{L}}}(\tau), for any τN\tau\leq-N, for a suitable sufficiently large N=N()>0N=N({\mathcal{L}})>0; furthermore, the function 𝐐(){\boldsymbol{Q}}_{{\mathcal{L}}}(\cdot) is C1C^{1} and limτ𝐐(τ)=𝐐()\lim_{\tau\to-\infty}\boldsymbol{Q}_{{\mathcal{L}}}(\tau)=\boldsymbol{Q}_{{\mathcal{L}}}(-\infty).
We stress that 𝐐(τ){\boldsymbol{Q}}_{{\mathcal{L}}}(\tau) is uniquely defined as the first intersection between Wu(τ)W^{u}(\tau) and {\mathcal{L}}, although it might not be the unique intersection, especially if {\mathcal{L}} is too large.

Let {\mathcal{L}} be a segment transversal to 𝚪\boldsymbol{\Gamma_{-\infty}}, take τN\tau\leq-N and denote by W~u(τ)\tilde{W}^{u}_{{\mathcal{L}}}(\tau) the compact and connected branch of Wu(τ)W^{u}(\tau) between the origin and 𝑸(τ){\boldsymbol{Q}}_{{\mathcal{L}}}(\tau). Analogously, denote by W~u()\tilde{W}^{u}_{{\mathcal{L}}}(-\infty) the compact connected branch of 𝚪=Wu()\boldsymbol{\Gamma_{-\infty}}=W^{u}(-\infty) between the origin and 𝑸(){\boldsymbol{Q}}_{{\mathcal{L}}}(-\infty).

Lemma 15.

Assume (𝐖𝐮)\boldsymbol{({\rm W}_{u})}. Let {\mathcal{L}} and NN be as in Remark 14, then there is c=c()c=c({\mathcal{L}}) such that

sup{ϕ(θ+τ;τ,𝑸)eαθ𝑸W~u(τ),θ0,τN}c().\sup\{\|\boldsymbol{\phi}(\theta+\tau;\tau,{\boldsymbol{Q}})\|{\rm e}^{-\alpha\theta}\mid{\boldsymbol{Q}}\in\tilde{W}^{u}_{{\mathcal{L}}}(\tau),\;\theta\leq 0,\;\tau\leq-N\}\leq c({\mathcal{L}}). (24)
Proof.

Let δ>0\delta>0 be the fixed constant defined in Theorem 11. If {\mathcal{L}} is close enough to the origin so that 𝑸()<δ\|\boldsymbol{Q}_{{\mathcal{L}}}(-\infty)\|<\delta, then by construction W~u(τ)Wlocu(τ)\tilde{W}^{u}_{{\mathcal{L}}}(\tau)\subset W^{u}_{\textrm{loc}}(\tau) for every τN\tau\leq-N, see (22), and (24) follows straightforwardly from Theorem 12.

Now, let {\mathcal{L}} be a generic segment transversal to 𝚪\boldsymbol{\Gamma_{-\infty}} satisfying 𝑸()δ\|\boldsymbol{Q}_{{\mathcal{L}}}(-\infty)\|\geq\delta, and let M=M(δ,)>0M=M(\delta,{\mathcal{L}})>0 be such that

ϕ(M;0,𝑸())=δ/2.\|\boldsymbol{\phi_{-\infty}}(-M;0,\boldsymbol{Q}_{{\mathcal{L}}}(-\infty))\|=\delta/2\,.

Fix a point 𝑸^\boldsymbol{\hat{Q}} belonging to W~u()B(𝟎,δ2)\tilde{W}^{u}_{{\mathcal{L}}}(-\infty)\setminus B(\boldsymbol{0},\frac{\delta}{2}), then there is t𝑸^[M,0]t_{\boldsymbol{\hat{Q}}}\in[-M,0] such that ϕ(t𝑸^;0,𝑸())=𝑸^\boldsymbol{\phi_{-\infty}}(t_{\boldsymbol{\hat{Q}}};0,\boldsymbol{Q}_{{\mathcal{L}}}(-\infty))=\boldsymbol{\hat{Q}}. By construction, ϕ(τM;τ,𝑸^)δ2\|\boldsymbol{\phi_{-\infty}}(\tau-M;\tau,\boldsymbol{\hat{Q}})\|\leq{\frac{\delta}{2}} for any τ<0,\tau<0, since ϕ(τM;τ,𝑸^)=ϕ(M;0,𝑸^)=ϕ(M+t𝑸^;0,𝑸())\boldsymbol{\phi_{-\infty}}(\tau-M;\tau,\boldsymbol{\hat{Q}})=\boldsymbol{\phi_{-\infty}}(-M;0,\boldsymbol{\hat{Q}})=\boldsymbol{\phi_{-\infty}}(-M+t_{\boldsymbol{\hat{Q}}};0,\boldsymbol{Q}_{{\mathcal{L}}}(-\infty)).

Using Remark 14 combined with the compactness of W~u(τ)Wlocu(τ)\tilde{W}^{u}_{{\mathcal{L}}}(\tau)\setminus W^{u}_{\textrm{loc}}(\tau), continuous dependence on initial data and parameters of the flow of (11), and possibly choosing a larger NN, we can assume that 𝑹=ϕ(τM;τ,𝑸)\boldsymbol{R}=\boldsymbol{\phi}(\tau-M;\tau,{\boldsymbol{Q}}) is such that

𝑹<δ, so that𝑹Wlocu(τM)\|\boldsymbol{R}\|<\delta,\quad\textrm{ so that}\quad\boldsymbol{R}\in W^{u}_{\textrm{loc}}(\tau-M) (25)

whenever τ<N\tau<-N and 𝑸W~u(τ)Wlocu(τ){\boldsymbol{Q}}\in\tilde{W}^{u}_{{\mathcal{L}}}(\tau)\setminus W^{u}_{\textrm{loc}}(\tau); notice that MM does not depend on 𝑸{\boldsymbol{Q}} and τ\tau, but depends on {\mathcal{L}}. The existence of 𝑹\boldsymbol{R} as in (25) is trivial for any 𝑸Wlocu(τ){\boldsymbol{Q}}\in W^{u}_{\textrm{loc}}(\tau), so (25) holds for any 𝑸W~u(τ){\boldsymbol{Q}}\in\tilde{W}^{u}_{{\mathcal{L}}}(\tau).

Thus, Theorem 12 ensures the existence of c0=c0(δ)>0c_{0}=c_{0}(\delta)>0 such that

sup{ϕ(s+τM;τ,𝑸)eαs𝑸W~u(τ),s0,τN}c0.\sup\{\|\boldsymbol{\phi}(s+\tau-M;\tau,\boldsymbol{Q})\|{\rm e}^{-\alpha s}\mid{\boldsymbol{Q}}\in\tilde{W}^{u}_{{\mathcal{L}}}(\tau),\;s\leq 0,\;\tau\leq-N\}\leq c_{0}.

Hence, ϕ(θ+τ;τ,𝑸)eαθc0eαM\|\boldsymbol{\phi}(\theta+\tau;\tau,\boldsymbol{Q})\|{\rm e}^{-\alpha\theta}\leq c_{0}{\rm e}^{\alpha M}, for any θM\theta\leq-M and 𝑸W~u(τ){\boldsymbol{Q}}\in\tilde{W}^{u}_{{\mathcal{L}}}(\tau), which proves  (24) when θM\theta\leq-M.

To prove (24) for Mθ0-M\leq\theta\leq 0, it is enough to recall that W~u(τ)\tilde{W}^{u}_{{\mathcal{L}}}(\tau) is compact, M>0M>0 is fixed and the flow of (11) is continuous; then the lemma follows by choosing some c()c0eαMc({\mathcal{L}})\geq c_{0}{\rm e}^{\alpha M}. ∎

The following lemma better describes the behaviour of the solutions ϕ(;τ,𝑸)\boldsymbol{\phi}(\cdot;\tau,{\boldsymbol{Q}}) departing from a point 𝑸Wu(τ){\boldsymbol{Q}}\in W^{u}(\tau), as τ\tau\to-\infty. Roughly speaking, we can say that such trajectories mime the autonomous dynamical system (15) frozen at τ=\tau=-\infty.

Lemma 16.

Assume (𝐖𝐮)\boldsymbol{({\rm W}_{u})}. Let {\mathcal{L}} and 𝐐(){\boldsymbol{Q}}_{\mathcal{L}}(-\infty) be as in Remark 14. Then,

limτsupθ0ϕ(θ+τ;τ,𝑸(τ))ϕ(θ;0,𝑸())=0.\lim_{\tau\to-\infty}\,\sup_{\theta\leq 0}\left\|\boldsymbol{\phi}\big{(}\theta+\tau;\tau,{\boldsymbol{Q}}_{\mathcal{L}}(\tau)\big{)}-\boldsymbol{\phi}_{-\infty}\big{(}\theta;0,{\boldsymbol{Q}}_{\mathcal{L}}(-\infty)\big{)}\right\|=0\,. (26)
Proof.

For brevity, we set

𝝃𝝉(θ)=ϕ(θ+τ;τ,𝑸(τ))and𝝃(θ)=ϕ(θ;0,𝑸()).\boldsymbol{\xi_{\tau}}(\theta)=\boldsymbol{\phi}(\theta+\tau;\tau,{\boldsymbol{Q}}_{\mathcal{L}}(\tau))\quad\text{and}\quad\boldsymbol{\xi^{*}}(\theta)=\boldsymbol{\phi_{-\infty}}(\theta;0,{\boldsymbol{Q}}_{\mathcal{L}}(-\infty))\,.

We argue by contradiction, and assume that there exist ε>0\varepsilon>0 and two sequences (θn)n,(τn)n],0](\theta_{n})_{n}\,,\,(\tau_{n})_{n}\subset{}]-\infty,0], such that τn\tau_{n}\to-\infty satisfying

𝝃𝝉𝒏(θn)𝝃(θn)>ε.\|\boldsymbol{\xi_{\tau_{n}}}(\theta_{n})-\boldsymbol{\xi^{*}}(\theta_{n})\|>\varepsilon\,. (27)

Combining Lemma 15 with (21), we easily find that

𝝃𝝉𝒏(θn)𝝃(θn)𝝃𝝉𝒏(θn)+𝝃(θn)2c()eαθn,τnN.\|\boldsymbol{\xi_{\tau_{n}}}(\theta_{n})-\boldsymbol{\xi^{*}}(\theta_{n})\|\leq\|\boldsymbol{\xi_{\tau_{n}}}(\theta_{n})\|+\|\boldsymbol{\xi^{*}}(\theta_{n})\|\leq 2c({\mathcal{L}}){\rm e}^{\alpha\theta_{n}},\quad\forall\tau_{n}\leq-N.

If (θn)n(\theta_{n})_{n} is unbounded, then we easily get a contradiction with (27).

Thus, we can assume that there is M>0M>0 such that (θn)n[M,0](\theta_{n})_{n}\subset[-M,0]. Using continuous dependence on initial data and parameters, we see that for any ε>0\varepsilon>0 there exists δ(ε)>0\delta(\varepsilon)>0 such that if 𝑸(τ)𝑸()<δ\|{\boldsymbol{Q}}_{\mathcal{L}}(\tau)-{\boldsymbol{Q}}_{\mathcal{L}}(-\infty)\|<\delta, then 𝝃𝝉(θ)𝝃(θ)<ε\|\boldsymbol{\xi_{\tau}}(\theta)-\boldsymbol{\xi^{*}}(\theta)\|<\varepsilon for any θ[M,0]\theta\in[-M,0] whenever τ<N\tau<-N.

Afterwards, from Remark 14, we find N1=N1(δ)>N>0N_{1}=N_{1}(\delta)>N>0 large enough so that 𝑸(τ)𝑸()<δ\|{\boldsymbol{Q}}_{\mathcal{L}}(\tau)-{\boldsymbol{Q}}_{\mathcal{L}}(-\infty)\|<\delta for any τ<N1\tau<-N_{1}, so that eventually we get 𝝃𝝉(θ)𝝃(θ)<ε\|\boldsymbol{\xi_{\tau}}(\theta)-\boldsymbol{\xi^{*}}(\theta)\|<\varepsilon for any θ[M,0]\theta\in[-M,0] whenever τ<N1\tau<-N_{1}; but this is in contradiction with (27), since we are assuming that (θn)n[M,0](\theta_{n})_{n}\subset[-M,0] and τn\tau_{n}\to-\infty. The Lemma is thus proved. ∎

We denote by

ϕ(t;d)=(x(t;d),y(t;d))\boldsymbol{\boldsymbol{\phi}}(t;d)=(x(t;d),y(t;d))

the trajectory of (11) corresponding to the regular solution u(r;d)u(r;d) of (4).

According to [21, 22], all the regular solutions correspond to trajectories in the unstable leaf.

Remark 17.

Assume that 𝒦C1{\mathcal{K}}\in C^{1} and (𝐖𝐮)\boldsymbol{({\rm W}_{u})} holds. Then,

u(r;d) is a regular solutionϕ(τ;d)Wu(τ) for every τ.u(r;d)\mbox{ is a regular solution}\quad\Longleftrightarrow\quad\boldsymbol{\phi}(\tau;d)\in W^{u}(\tau)\mbox{ for every }\tau\in{\mathbb{R}}.

Further, fixed τ0\tau_{0}\in{\mathbb{R}}, the function 𝓠𝛕𝟎:[0,+[Wu(τ0){\boldsymbol{{\cal Q}}}_{\boldsymbol{\tau_{0}}}:[0,+\infty[{}\to W^{u}(\tau_{0}) defined by 𝓠𝛕𝟎(d):=ϕ(τ0;d){\boldsymbol{{\cal Q}}}_{\boldsymbol{\tau_{0}}}(d):=\boldsymbol{\phi}(\tau_{0};d) is a continuous (bijective) parametrization of Wu(τ0)W^{u}(\tau_{0}). In particular, 𝓠𝛕𝟎(0)=(0,0){\boldsymbol{{\cal Q}}}_{\boldsymbol{\tau_{0}}}(0)=(0,0). Therefore, for any 𝐐Wu(τ0){\boldsymbol{Q}}\in W^{u}(\tau_{0}) there is a unique d(𝐐)0d({\boldsymbol{Q}})\geq 0 such that ϕ(t;d(𝐐))ϕ(t;τ0,𝐐)\boldsymbol{\phi}(t;d({\boldsymbol{Q}}))\equiv\boldsymbol{\phi}(t;\tau_{0},{\boldsymbol{Q}}) for any tt\in{\mathbb{R}}.

In fact, the dependence of 𝓠𝛕𝟎{\boldsymbol{{\cal Q}}}_{\boldsymbol{\tau_{0}}} with respect to the parameter τ0\tau_{0} is C1C^{1}.

The existence of a bijective parametrization of Wu(τ)W^{u}(\tau) can be obtained extending to the pp-Laplacian setting the argument developed in the proof of [12, Lemma 2.10], written in the case p=2p=2. The smoothness of 𝓠𝝉𝟎(d)=ϕ(τ0;d){\boldsymbol{{\cal Q}}}_{\boldsymbol{\tau_{0}}}(d)=\boldsymbol{\phi}(\tau_{0};d) with respect to τ0\tau_{0} follows from the smoothness of the flow of (11).

According to [41, Lemma 3.7], we can easily prove the monotonicity properties of regular solutions. In particular,

Remark 18.

Any regular solution u(r;d)u(r;d) of (4) is decreasing until its first zero.

Fix {\mathcal{L}} as in Remark 14, so that 𝑸(τ){\boldsymbol{Q}}_{{\mathcal{L}}}(\tau) is well defined for any τ<N()\tau<-N({\mathcal{L}}). From Remark 17, we can define the function d:],N()[]0,+[d_{\mathcal{L}}:{}]-\infty,-N({\mathcal{L}})[{}\to{}]0,+\infty[{} by setting d(τ):=d(𝑸(τ))d_{\mathcal{L}}(\tau):=d({\boldsymbol{Q}}_{{\mathcal{L}}}(\tau)) for any τ<N()\tau<-N({\mathcal{L}}). In particular, ϕ(t;d(τ))ϕ(t;τ,𝑸(τ))\boldsymbol{\boldsymbol{\phi}}(t;d_{\mathcal{L}}(\tau))\equiv\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}}_{{\mathcal{L}}}(\tau)).

Now, we need the following weak version of the implicit function theorem, cf. [13, Theorem 15.1].

Theorem 19.

Let A(x,y):2A(x,y):{\mathbb{R}}^{2}\to{\mathbb{R}} be a function continuous along with its partial derivative Ax(x,y)\frac{\partial A}{\partial x}(x,y). Let A(x0,y0)=0A(x_{0},y_{0})=0 and Ax(x0,y0)0\frac{\partial A}{\partial x}(x_{0},y_{0})\neq 0, then we can find δ>0\delta>0 and exactly one continuous function x(y):[y0δ,y0+δ]x(y):[y_{0}-\delta,y_{0}+\delta]\to{\mathbb{R}} such that x(y0)=x0x(y_{0})=x_{0} and A(x(y),y)=0A(x(y),y)=0 for any y[y0δ,y0+δ]y\in[y_{0}-\delta,y_{0}+\delta].

Our next aim consists in showing the invertibility and monotonicity of dd_{{\mathcal{L}}}.

Lemma 20.

Assume (𝐖𝐮)\boldsymbol{({\rm W}_{u})}. Let {\mathcal{L}}, N=N()N=N({\mathcal{L}}) be as in Remark 14. Then, there is D=D(,N)D=D({\mathcal{L}},N) such that the function d:],N[]D,+[d_{\mathcal{L}}:{}]-\infty,-N[{}\to{}]D,+\infty[{}, defined by the property ϕ(t;d(τ))ϕ(t;τ,𝐐(τ))\boldsymbol{\boldsymbol{\phi}}(t;d_{\mathcal{L}}(\tau))\equiv\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}}_{{\mathcal{L}}}(\tau)), is continuous, bijective, monotone decreasing, and its inverse τ:]D,+[],N[\tau_{\mathcal{L}}:{}]D,+\infty[{}\to{}]-\infty,-N[{} is continuous.

Furthermore, limτd(τ)=+\displaystyle{\lim_{\tau\to-\infty}d_{\mathcal{L}}(\tau)}=+\infty, and so limd+τ(d)=\displaystyle{\lim_{d\to+\infty}\tau_{\mathcal{L}}(d)}=-\infty.

Proof.

As noticed in Remark 14, the function d:],N[]0,+[d_{\mathcal{L}}:{}]-\infty,-N[{}\to{}]0,+\infty[{} is well defined, due to the transversality of the first crossing between Wu(τ)W^{u}(\tau) and {\mathcal{L}}, for τ<N\tau<-N. We show now that dd_{\mathcal{L}} admits a continuous inverse, by constructing it via Theorem 19.

Denote by ¯\bar{{\mathcal{L}}} the straight line containing the segment {\mathcal{L}}, and let 𝒟(𝑸)\mathcal{D}({\boldsymbol{Q}}) be the smooth function which evaluates the directed distance from ¯\bar{{\mathcal{L}}} to 𝑸{\boldsymbol{Q}}, so that 𝒟(𝑸)=0\mathcal{D}({\boldsymbol{Q}})=0 if and only if 𝑸¯{\boldsymbol{Q}}\in\bar{{\mathcal{L}}}. Recalling the parametrization 𝓠𝝉{\boldsymbol{{\cal Q}}}\boldsymbol{{}_{\tau}} of Wu(τ)W^{u}(\tau) introduced in Remark 17, we define A(τ,d):=𝒟(𝓠(d)𝝉)A(\tau,d):=\mathcal{D}({\boldsymbol{{\cal Q}}}\boldsymbol{{}_{\tau}}(d)). Let us consider a couple (τ0,d0)(\tau_{0},d_{0}) with τ0<N\tau_{0}<-N and d0=d(τ0)d_{0}=d_{\mathcal{L}}(\tau_{0}). In particular, 𝓠(d0)𝝉𝟎=𝑸(τ0){\boldsymbol{{\cal Q}}}\boldsymbol{{}_{\tau_{0}}}(d_{0})={\boldsymbol{Q}}_{{\mathcal{L}}}(\tau_{0}) and so A(τ0,d0)=𝒟(𝑸(τ0))=0A(\tau_{0},d_{0})=\mathcal{D}({\boldsymbol{Q}}_{{\mathcal{L}}}(\tau_{0}))=0. From the smoothness property of Wu(τ)W^{u}(\tau) given in Remark 13, we can compute

Aτ(τ0,d0)\displaystyle\frac{\partial A}{\partial\tau}(\tau_{0},d_{0}) =𝒟(𝓠(d0)𝝉𝟎),τ𝓠(d0)𝝉𝟎=\displaystyle=\left\langle\nabla\mathcal{D}({\boldsymbol{{\cal Q}}}\boldsymbol{{}_{\tau_{0}}}(d_{0}))\,,\,\frac{\partial}{\partial\tau}{\boldsymbol{{\cal Q}}}\boldsymbol{{}_{\tau_{0}}}(d_{0})\right\rangle= (28)
=𝒟(𝓠(d0)𝝉𝟎),ϕ˙(τ0;τ0,𝓠(d0)𝝉𝟎)0,\displaystyle=\left\langle\nabla\mathcal{D}({\boldsymbol{{\cal Q}}}\boldsymbol{{}_{\tau_{0}}}(d_{0}))\,,\,\boldsymbol{\dot{\phi}}(\tau_{0};\tau_{0},{\boldsymbol{{\cal Q}}}\boldsymbol{{}_{\tau_{0}}}(d_{0}))\right\rangle\neq 0, (29)

since 𝒟(𝑸)\nabla\mathcal{D}({\boldsymbol{Q}}) is orthogonal to ¯\bar{\mathcal{L}} and ϕ˙(τ0;τ0,𝓠(d0)𝝉𝟎)\boldsymbol{\dot{\phi}}(\tau_{0};\tau_{0},{\boldsymbol{{\cal Q}}}\boldsymbol{{}_{\tau_{0}}}(d_{0})) is transversal to {\mathcal{L}}.

From the previous formula we also find that Aτ\frac{\partial A}{\partial\tau} is continuous, so we can apply Theorem 19, thus finding locally a continuous function τ(d)\tau_{\mathcal{L}}(d) such that A(τ(d),d)0A(\tau_{\mathcal{L}}(d),d)\equiv 0; so by construction τ(d)\tau_{\mathcal{L}}(d) is the local inverse of dd_{\mathcal{L}}. Since the previous argument can be performed for every couple (τ0,d0)(\tau_{0},d_{0}) satisfying d0=d(τ0)d_{0}=d_{\mathcal{L}}(\tau_{0}) with τ0],N[\tau_{0}\in{}]-\infty,-N[{}, we can conclude that the image of dd_{\mathcal{L}} is an open interval 𝒰\mathcal{U}, and that τ(d)\tau_{\mathcal{L}}(d) is a global inverse.

We now show that limτd(τ)=+\lim_{\tau\to-\infty}d_{\mathcal{L}}(\tau)=+\infty.

Let c>0c>0 be such that {(x,y)x>c}{\mathcal{L}}\subset\{(x,y)\mid x>c\}. Then, for every τ<N\tau<-N the point 𝑸(τ)=(Qx(τ),Qy(τ)){\boldsymbol{Q}}_{{\mathcal{L}}}(\tau)=(Q_{x}(\tau),Q_{y}(\tau)) is such that Qx(τ)>cQ_{x}(\tau)>c. Let us consider the solution u(r;d(τ))u(r;d_{\mathcal{L}}(\tau)) corresponding to the trajectory ϕ(t;d(τ))ϕ(t;τ,𝑸(τ))\boldsymbol{\boldsymbol{\phi}}(t;d_{\mathcal{L}}(\tau))\equiv\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}}_{{\mathcal{L}}}(\tau)). According to Lemma 16, u(r;d(τ))u(r;d_{\mathcal{L}}(\tau)) is positive for r[0,eτ]r\in[0,{\rm e}^{\tau}], and from Remark 18, we deduce that

d(τ)=u(0;d(τ))u(eτ;d(τ))=Qx(τ)eατ>ceατ+d_{\mathcal{L}}(\tau)=u(0;d_{\mathcal{L}}(\tau))\geq u({\rm e}^{\tau};d_{\mathcal{L}}(\tau))=Q_{x}(\tau)\,{\rm e}^{-\alpha\tau}>c\,{\rm e}^{-\alpha\tau}\to+\infty (30)

as τ\tau\to-\infty. Finally, we find that dd_{\mathcal{L}} is decreasing and there exists D=D(,N)D=D({\mathcal{L}},N) such that d(],N[)=]D,+[d_{\mathcal{L}}({}]-\infty,-N[{})={}]D,+\infty[{}. ∎

3 Basic properties of the function 𝑹(𝒅)\boldsymbol{R(d)}.

Using the Pohozaev identity, see e.g.[20, 21, 34, 42], it is possible to obtain the following classical result.

Proposition 21.

Assume (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)}, then all the regular solutions u(r;d)u(r;d) of (4) have a zero at r=R(d)r=R(d). Hence, all the corresponding trajectories ϕ(t;d)\boldsymbol{\phi}(t;d) of (11) are such that y(t;d)<0<x(t;d)y(t;d)<0<x(t;d) when t<T(d):=lnR(d)t<T(d):=\ln R(d) and x(T(d);d)=0>y(T(d);d)x(T(d);d)=0>y(T(d);d).

For the proof, we refer to [14, 36] for the Laplacian operator and to [26, 34] for pp-Laplacian extensions. Then, following [21, Theorem 4.2], we can prove the continuity of the function R(d)R(d).

Proposition 22.

Assume (𝐖𝐮)\boldsymbol{({\rm W}_{u})}, then the set JJ introduced in (7) is open and the function R:JR:J\to{\mathbb{R}} is continuous.

We need two further asymptotic results, which can already be found in literature in slightly different contexts.

Proposition 23.

Assume (𝐖𝐮)\boldsymbol{({\rm W}_{u})}, and infJ=0\inf J=0, then limd0+R(d)=+\lim_{d\to 0^{+}}R(d)=+\infty.

We refer to [33, Proposition 2.4] for the proof. With some effort, we can generalize the previous proposition a bit.

Proposition 24.

Assume (𝐖𝐮)\boldsymbol{({\rm W}_{u})} and the existence of D~>0\tilde{D}>0, with D~J\tilde{D}\notin J which is an accumulation point for JJ. Then,

limdD~dJR(d)=+.\lim_{\scriptsize\begin{array}[]{l}d\!\to\!{\tilde{D}}\\ d\in J\end{array}}\!\!R(d)=+\infty.
Proof.

According to Remark 18, u(r,D~)u(r,\tilde{D}) is positive and decreasing for any r0r\geq 0, since D~J\tilde{D}\notin J. Let δ>0\delta>0 be as in Theorem 11, and choose τ00\tau_{0}\ll 0 so that

W~u(τ0):={𝓠𝝉𝟎(d)d[0,D~+1]}Wlocu(τ0)B(𝟎,δ),\tilde{W}^{u}(\tau_{0}):=\{{\boldsymbol{{\cal Q}}}_{\boldsymbol{\tau_{0}}}(d)\mid d\in[0,\tilde{D}+1]\}\subset W^{u}_{\textrm{loc}}(\tau_{0})\subset B(\boldsymbol{0},\delta),

where 𝓠𝝉𝟎(d){\boldsymbol{{\cal Q}}}_{\boldsymbol{\tau_{0}}}(d) is as in Remark 17. By construction, R(d)>eτ0R(d)>{\rm e}^{\tau_{0}} for any 0<dD~+10<d\leq\tilde{D}+1.

Assume by contradiction that there are dnJd_{n}\in J, dnD~d_{n}\to\tilde{D}, and MM\in{\mathbb{R}} such that R(dn)eMR(d_{n})\leq{\rm e}^{M}. Let us denote by ε=12min{x(t;D~)t[τ0,M]}\varepsilon=\frac{1}{2}\min\{x(t;\tilde{D})\mid t\in[\tau_{0},M]\}.

From the continuity of the parametrization 𝓠𝝉𝟎(){\boldsymbol{{\cal Q}}}_{\boldsymbol{\tau_{0}}}(\cdot), for any σ>0\sigma>0 we can find n¯=n¯(σ)>0\bar{n}=\bar{n}(\sigma)>0 so that 𝓠𝝉𝟎(dn)𝓠𝝉𝟎(D~)<σ\|{\boldsymbol{{\cal Q}}}_{\boldsymbol{\tau_{0}}}(d_{n})-{\boldsymbol{{\cal Q}}}_{\boldsymbol{\tau_{0}}}(\tilde{D})\|<\sigma when n>n¯n>\bar{n}. Then, using the continuity of the flow of (11), we can choose σ=σ(ε)>0\sigma=\sigma(\varepsilon)>0 small enough (and n¯>0\bar{n}>0 large enough) so that

ϕ(t;dn)ϕ(t;D~)=ϕ(t;τ0,𝓠𝝉𝟎(dn))ϕ(t;τ0,𝓠𝝉𝟎(D~))<ε\|\boldsymbol{\phi}(t;d_{n})-\boldsymbol{\phi}(t;\tilde{D})\|=\|\boldsymbol{\phi}(t;\tau_{0},{\boldsymbol{{\cal Q}}}_{\boldsymbol{\tau_{0}}}(d_{n}))-\boldsymbol{\phi}(t;\tau_{0},{\boldsymbol{{\cal Q}}}_{\boldsymbol{\tau_{0}}}(\tilde{D}))\|<\varepsilon

for any t[τ0,M]t\in[\tau_{0},M], whenever n>n¯n>\bar{n}. Hence, we get

x(t;dn)x(t;D~)|x(t;dn)x(t;D~)|>2εε=ε,x(t;d_{n})\geq x(t;\tilde{D})-|x(t;d_{n})-x(t;\tilde{D})|>2\varepsilon-\varepsilon=\varepsilon,

for any t[τ0,M]t\in[\tau_{0},M] and n>n¯n>\bar{n}. Hence x(t;dn)>0x(t;d_{n})>0 for any tMt\leq M and, consequently, R(dn)>eMR(d_{n})>{\rm e}^{M} for any n>n¯n>\bar{n}: a contradiction. The Proposition is thus proved. ∎

Let us observe also that u(r;D~)u(r;\tilde{D}) is a Ground State and converges to 0 as r+r\to+\infty.

4 The asymptotic estimate of 𝑹(𝒅)\boldsymbol{R(d)} for 𝒅\boldsymbol{d} large.

Now we proceed to study the behavior of R(d)R(d) for dd large. We emphasize that the asymptotic estimates of the first zero for large initial data are mainly based on the crucial assumption (𝐇)\boldsymbol{({\rm H}_{\ell})}.

Our first step is to locate 𝑸(τ){\boldsymbol{Q}}_{{\mathcal{L}}}(\tau) through the function {\mathcal{H}}, see Proposition 26 and Remark 27. Then, we use a Grönwall’s argument to get a lower and an upper bound of the time T(τ,)T(\tau,{\mathcal{L}}) taken by ϕ(t;τ,𝑸(τ))\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}}_{{\mathcal{L}}}(\tau)) to cross the yy negative semi-axis, see Lemmas 31 and 33. Finally, we prove the asymptotic estimates of R(d)R(d) in Proposition 35 (for the <p\ell<\ell^{*}_{p} case) and in Propositions 43 and 46 (for the p\ell\geq\ell^{*}_{p} case).

We emphasize that assumption (𝐇)\boldsymbol{({\rm H}_{\ell})} implies that the function 𝒦{\mathcal{K}} is strictly increasing in a neighborhood of r=0r=0, so we can use the following truncation argument.

Remark 25.

From (𝐇)\boldsymbol{({\rm H}_{\ell})} we see that there exists T^0<0\hat{T}_{0}<0 such that K(T^0)<A+1K(\hat{T}_{0})<A+1 and

0<12Bet<K˙(t)<2Bet for every tT^0.0<\frac{1}{2}B\ell{\rm e}^{\ell t}<\dot{K}(t)<2B\ell{\rm e}^{\ell t}\mbox{ for every }t\leq\hat{T}_{0}. (31)

So, we can find a function K^:\hat{K}:{\mathbb{R}}\to{\mathbb{R}} of class C1C^{1} satisfying K^=K\hat{K}=K in the interval ],T^0]{}]-\infty,\hat{T}_{0}], dK^dt(t)>0\frac{d\!{\hat{K}}}{dt}(t)>0 for every tt\in{\mathbb{R}}, and

A+1=K^(+):=limt+K^(t).A+1=\hat{K}(+\infty):=\lim_{t\to+\infty}\hat{K}(t)\in{\mathbb{R}}\,.

In particular, K^\hat{K} satisfies (𝐇)\boldsymbol{({\rm H}_{\ell})}, (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)} and (𝐖𝐬)\boldsymbol{({\rm W}_{s})}. Such a truncation argument will permit us to assume implicitly the validity of the previous hypotheses, when we will look for properties of system (11) in a neighborhood of t=t=-\infty in the presence of the only hypothesis (𝐇)\boldsymbol{({\rm H}_{\ell})}.

Now, we provide some estimates of the energy (𝑸(τ);τ){\mathcal{H}}({\boldsymbol{Q}}_{\mathcal{L}}(\tau);\tau), where 𝑸(τ){\boldsymbol{Q}}_{\mathcal{L}}(\tau) is as in Remark 14.

Proposition 26.

Assume (𝐇)\boldsymbol{({\rm H}_{\ell})}. Let {\mathcal{L}} be a small enough segment, transversal to 𝚪\boldsymbol{\Gamma_{-\infty}}, and consider the point 𝐐(τ){\boldsymbol{Q}}_{\mathcal{L}}(\tau), with τ<N()\tau<-N({\mathcal{L}}). Then, there is a constant c~()>0\tilde{c}({\mathcal{L}})>0 such that

(𝑸(τ);τ)=c~()eτ+o(eτ) as τ.{\mathcal{H}}({\boldsymbol{Q}}_{\mathcal{L}}(\tau);\tau)=\tilde{c}({\mathcal{L}})\,{\rm e}^{\ell\tau}+o({\rm e}^{\ell\tau})\qquad\textrm{ as $\tau\to-\infty$}. (32)
Proof.

We write for brevity 𝝃𝝉(t)=(xτ(t),yτ(t))=ϕ(t+τ;τ,𝑸(τ))\boldsymbol{\xi_{\tau}}(t)=(x_{\tau}(t),y_{\tau}(t))=\boldsymbol{\phi}(t+\tau;\tau,{\boldsymbol{Q}}_{\mathcal{L}}(\tau)) and 𝝃(t)=(x(t),y(t))=ϕ(t;0,𝑸())\boldsymbol{\xi^{*}}(t)=(x^{*}(t),y^{*}(t))=\boldsymbol{\phi_{-\infty}}(t;0,{\boldsymbol{Q}}_{\mathcal{L}}(-\infty)). Notice that, according to Lemma 16, xτx_{\tau} is positive in ],0]]-\infty,0].

Since 𝑸(τ)Wu(τ){\boldsymbol{Q}}_{\mathcal{L}}(\tau)\in W^{u}(\tau), from Lemma 9 combined with Remark 17 we know that limt(𝝃𝝉(t);t+τ)=0\displaystyle{{\lim_{t\to-\infty}}{\mathcal{H}}(\boldsymbol{\xi_{\tau}}(t);t+\tau)=0}.

Hence, in the spirit of [22, pag. 357], by (14) and Remark 25 we find

(𝑸(τ);τ)=0K˙(t+τ)xτ(t)qq𝑑t==0[Be(t+τ)+h(et+τ)et+τ]xτ(t)qq𝑑t=eτ(I1+I2),\begin{split}{\mathcal{H}}({\boldsymbol{Q}}_{\mathcal{L}}(\tau);\tau)=&\int_{-\infty}^{0}\dot{K}(t+\tau)\frac{x_{\tau}(t)^{q}}{q}\,dt=\\ =&\int_{-\infty}^{0}\left[B\ell{\rm e}^{\ell(t+\tau)}+h^{\prime}({\rm e}^{t+\tau}){\rm e}^{t+\tau}\right]\frac{x_{\tau}(t)^{q}}{q}\,dt={\rm e}^{\ell\tau}\left(I_{1}+I_{2}\right)\,,\end{split} (33)

where

I1:=Bq0etxτ(t)q𝑑t and I2:=1q0h(et+τ)et+τe(t+τ)etxτ(t)q𝑑t.I_{1}:=\frac{B\ell}{q}\int_{-\infty}^{0}{\rm e}^{\ell t}x_{\tau}(t)^{q}\,dt\quad\mbox{ and }\quad I_{2}:=\frac{1}{q}\int_{-\infty}^{0}\frac{h^{\prime}({\rm e}^{t+\tau}){\rm e}^{t+\tau}}{{\rm e}^{\ell(t+\tau)}}\,{\rm e}^{\ell t}x_{\tau}(t)^{q}\,dt.

We can rewrite I1I_{1} in the following equivalent form:

I1=Bq0etx(t)q𝑑t+Bq0et[xτ(t)qx(t)q]𝑑t.I_{1}=\frac{B\ell}{q}\int_{-\infty}^{0}{\rm e}^{\ell t}x^{*}(t)^{q}\,dt+\frac{B\ell}{q}\int_{-\infty}^{0}{\rm e}^{\ell t}[x_{\tau}(t)^{q}-x^{*}(t)^{q}]\,dt.

Recalling (26) and the fact that both xτ(t)x_{\tau}(t), x(t)x^{*}(t) converge to 0 exponentially as tt\to-\infty as observed in Lemma 15 and in (20), we apply the Lebesgue theorem to deduce the existence of ω1(t)\omega_{1}(t) such that

I1=c~()+ω1(τ) with limτω1(τ)=0,I_{1}=\tilde{c}({\mathcal{L}})+\omega_{1}(\tau)\hskip 8.53581pt\mbox{ with }\hskip 8.53581pt\lim_{\tau\to-\infty}\omega_{1}(\tau)=0\,, (34)

where c~():=Bq0etx(t)q𝑑t>0\tilde{c}({\mathcal{L}}):=\frac{B\ell}{q}\int_{-\infty}^{0}{\rm e}^{\ell t}x^{*}(t)^{q}\,dt>0 is finite, due to (20).

Analogously, from (𝐇)\boldsymbol{({\rm H}_{\ell})}, we find ω2(t)\omega_{2}(t) such that

I2=ω2(τ) with limτω2(τ)=0.I_{2}=\omega_{2}(\tau)\hskip 8.53581pt\mbox{ with }\hskip 8.53581pt\lim_{\tau\to-\infty}\omega_{2}(\tau)=0\,. (35)

The thesis follows by plugging (34) and (35) in (33). ∎

For a fixed a>0a>0, let (a){\mathcal{L}}(a) denote the segment of y=ay=-a lying between the negative yy semi-axis and the isocline x˙=0\dot{x}=0 of system (11), i.e.

(a)={(s,a)0sX~(a)}, where X~(a):=1αa1p1.{\mathcal{L}}(a)=\{(s,-a)\mid 0\leq s\leq\tilde{X}(a)\},\quad\textrm{ where $\tilde{X}(a):=\tfrac{1}{\alpha}\,{a^{\frac{1}{p-1}}}$}\,. (36)

Recalling the definition of equilibrium point 𝑬(){\boldsymbol{E}}(-\infty) given in (18), we observe that for any 0<a<|Ey()|0<a<|E_{y}(-\infty)| the homoclinic orbit 𝚪\boldsymbol{\Gamma_{-\infty}} intersects (a){\mathcal{L}}(a) transversely. From Lemma 16 and Proposition 26, we get the following asymptotic result.

Remark 27.

Assume (𝐇)\boldsymbol{({\rm H}_{\ell})}. Follow Wu(τ)W^{u}(\tau) from the origin towards x>0x>0. Then, for any 0<a<|Ey()|0<a<|E_{y}(-\infty)| we can find N(a)>0N(a)>0 such that Wu(τ)W^{u}(\tau) intersects (a){\mathcal{L}}(a) transversely in a point, say 𝐐(τ,a):=𝐐(a)(τ)=(Qx(τ,a),a){\boldsymbol{Q}}(\tau,a):={\boldsymbol{Q}}_{{\mathcal{L}}(a)}(\tau)=(Q_{x}(\tau,a),-a), whenever τ<N(a)\tau<-N(a). From (36) we see that

y˙(τ;τ,𝑸(τ,a))>0>x˙(τ;τ,𝑸(τ,a)),for any τN(a).\dot{y}(\tau;\tau,{\boldsymbol{Q}}(\tau,a))>0>\dot{x}(\tau;\tau,{\boldsymbol{Q}}(\tau,a)),\quad\mbox{for any }\tau\leq-N(a).

Moreover, there is a constant c(a)>0c(a)>0 such that

(𝑸(τ,a);τ)=c(a)eτ+o(eτ) as τ.{\mathcal{H}}({\boldsymbol{Q}}(\tau,a);\tau)=c(a)\,{\rm e}^{\ell\tau}+o({\rm e}^{\ell\tau})\qquad\textrm{ as $\tau\to-\infty$}\,. (37)

In the next part of this section we study of the asymptotic behavior of the solutions, under the more restrictive assumptions (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)}, (𝐇)\boldsymbol{({\rm H}_{\ell})} and (𝐖𝒔)\boldsymbol{({\rm W}_{s})}. Finally, (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)} and (𝐖𝒔)\boldsymbol{({\rm W}_{s})} will be removed using the truncation argument suggested by Remark 25.

Lemma 28.

Assume (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)}. Let u(r;d)u(r;d) be a regular solution of (4) and let ϕ(t;d)\boldsymbol{\phi}(t;d) be the corresponding trajectory of (11). Then, there are Tx(d)<T(d)T_{x}(d)<T(d) such that x(t;d)>0x(t;d)>0 when t<T(d)t<T(d) and it becomes null at t=T(d)t=T(d); furthermore, x˙(t;d)>0\dot{x}(t;d)>0 when t<Tx(d)t<T_{x}(d), and x˙(t;d)<0\dot{x}(t;d)<0 when Tx(d)<tT(d)T_{x}(d)<t\leq T(d).

Proof.

The existence of T(d)T(d) follows from Proposition 21. From Lemma 9, we know that (ϕ(t;d);t)0{\mathcal{H}}(\boldsymbol{\phi}(t;d);t)\geq 0 for any tT(d)t\leq T(d). By a simple calculation, (𝑸;t)<0{\mathcal{H}}({\boldsymbol{Q}};t)<0 for every 𝑸=(Qx,Qy){\boldsymbol{Q}}=(Q_{x},Q_{y}) in the isocline x˙=0\dot{x}=0 with 0<QxEx(t)0<Q_{x}\leq E_{x}(t), so we easily deduce the existence of Tx(d)<T(d)T_{x}(d)<T(d) such that x˙(Tx(d);d)=0\dot{x}(T_{x}(d);d)=0 and x(Tx(d);d)>Ex(Tx(d))x(T_{x}(d);d)>E_{x}(T_{x}(d)). ∎

Assume (𝐇)\boldsymbol{({\rm H}_{\ell})} and (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)}. Fixed 0<a<|Ey()|0<a<|E_{y}(-\infty)|, consider (a){\mathcal{L}}(a) and N(a)>0N(a)>0 as in Remark 27 so that 𝑸(τ,a)Wu(τ)(a){\boldsymbol{Q}}(\tau,a)\in W^{u}(\tau)\cap{\mathcal{L}}(a) is well defined for any τ<N(a)\tau<-N(a).

From Proposition 21, there exists T(τ,a)>τT(\tau,a)>\tau such that ϕ(t;τ,𝑸(τ,a))\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}}(\tau,a)) crosses the yy negative semi-axis at t=T(τ,a)t=T(\tau,a) in a point, say (see Figure 4)

𝑹(τ,a)=(0,Ry(τ,a))=ϕ(T(τ,a);τ,𝑸(τ,a)).\boldsymbol{R}(\tau,a)=(0,R_{y}(\tau,a))=\boldsymbol{\phi}(T(\tau,a);\tau,{\boldsymbol{Q}}(\tau,a))\,. (38)

According to Remark 25, it is not restrictive to assume that K˙(t)>0\dot{K}(t)>0 for any tτN(a)t\leq\tau\leq-N(a) (provided that we choose N(a)<T^0-N(a)<\hat{T}_{0} in Remark 27). So, from Lemma 9 we deduce that (𝑸(τ,a);τ)>0{\mathcal{H}}(\boldsymbol{Q}(\tau,a);\tau)>0.

Refer to caption
Figure 4: The position of the points 𝑸(τ,a)\boldsymbol{Q}(\tau,a), 𝑹(τ,a)\boldsymbol{R}(\tau,a), and 𝑹1(τ,a)\boldsymbol{R}^{1}(\tau,a).

Our next purpose is to provide suitable estimates from above and from below of T(τ,a)T(\tau,a). From Lemma 28, ϕ(t;τ,𝑸(τ,a))\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}}(\tau,a)) is a graph on the xx-axis when t[τ,T(τ,a)]t\in[\tau,T(\tau,a)]. In particular, we can find a function ψ:[0,Qx(τ,a)]],0[\psi:[0,Q_{x}(\tau,a)]\to{}]-\infty,0[\, such that the image of ϕ(;τ,𝑸(τ,a)):[τ,T(τ,a)]2\boldsymbol{\phi}(\cdot;\tau,{\boldsymbol{Q}}(\tau,a)):[\tau,T(\tau,a)]\to{\mathbb{R}}^{2} can be parametrized as (x,ψ(x))(x,\psi(x)) for x[0,Qx(τ,a)]x\in[0,Q_{x}(\tau,a)].

Let us now consider the trajectory ϕ𝝉(t;τ,𝑸(τ,a))\boldsymbol{\phi_{\tau}}(t;\tau,{\boldsymbol{Q}}(\tau,a)) of the system (15) frozen at t=τt=\tau. Notice that its graph is contained in the level set

{(x,y)(x,y;τ)=(𝑸(τ,a);τ)>0},\{(x,y)\mid{\mathcal{H}}(x,y;\tau)={\mathcal{H}}(\boldsymbol{Q}(\tau,a);\tau)>0\}, (39)

which lies in the exterior of the homoclinic orbit 𝚪𝝉\boldsymbol{\Gamma_{\tau}} defined in (19).

Hence, there exists T1(τ,a)>τT^{1}(\tau,a)>\tau such that ϕ𝝉(;τ,𝑸(τ,a))\boldsymbol{\phi_{\tau}}(\cdot;\tau,{\boldsymbol{Q}}(\tau,a)) lies in the 4th4^{th} quadrant when t[τ,T1(τ,a)[t\in[\tau,T^{1}(\tau,a)[{} and it crosses transversely the yy negative semi-axis at t=T1(τ,a)t=T^{1}(\tau,a) in a point, say (see Figure 4)

𝑹1(τ,a)=(0,Ry1(τ,a))=ϕ𝝉(T1(τ,a);τ,𝑸(τ,a)).\boldsymbol{R}^{1}(\tau,a)=(0,R^{1}_{y}(\tau,a))=\boldsymbol{\phi_{\tau}}(T^{1}(\tau,a);\tau,{\boldsymbol{Q}}(\tau,a))\,. (40)
Remark 29.

Assume (𝐖𝐮)\boldsymbol{({\rm W}_{u})}, then the functions T(τ,a)T(\tau,a) and T1(τ,a)T^{1}(\tau,a) are continuous in their domain since the flow on the negative yy-axis is transversal.

From the study of the trajectories of the autonomous system (15), we can deduce that

x˙τ(t;τ,𝑸(τ,a))<0<y˙τ(t;τ,𝑸(τ,a))for any t[τ,T1(τ,a)],\dot{x}_{\tau}(t;\tau,{\boldsymbol{Q}}(\tau,a))<0<\dot{y}_{\tau}(t;\tau,{\boldsymbol{Q}}(\tau,a))\quad\mbox{for any }t\in[\tau,T^{1}(\tau,a)], (41)

and, consequently, we can find a strictly decreasing function ψτ:[0,Qx(τ,a)][a,Ry1(τ,a)]\psi_{\tau}:[0,Q_{x}(\tau,a)]\to[-a,R^{1}_{y}(\tau,a)] such that the image of ϕ𝝉(;τ,𝑸(τ,a)):[τ,T1(τ,a)]2\boldsymbol{\phi_{\tau}}(\cdot;\tau,{\boldsymbol{Q}}(\tau,a)):[\tau,T^{1}(\tau,a)]\to{\mathbb{R}}^{2} can be parametrized as (x,ψτ(x))(x,\psi_{\tau}(x)) for x[0,Qx(τ,a)]x\in[0,Q_{x}(\tau,a)].

Lemma 30.

Assume (𝐇)\boldsymbol{({\rm H}_{\ell})} and that K˙(t)>0\dot{K}(t)>0 for any tt\in{\mathbb{R}}. Fix 0<a<|Ey()|0<a<|E_{y}(-\infty)|, and let N(a)>0N(a)>0 be as in Remark 27. Then,

T(τ,a)<T1(τ,a),for any τ<N(a).T(\tau,a)<T^{1}(\tau,a)\,,\qquad\mbox{for any }\tau<-N(a).
Proof.

For brevity, let us write ϕ(t)=(x(t),y(t))=ϕ(t;τ,𝑸(τ,a))\boldsymbol{\phi}(t)=(x(t),y(t))=\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}}(\tau,a)) and ϕ𝝉(t)=(xτ(t),yτ(t))=ϕ𝝉(t;τ,𝑸(τ,a))\boldsymbol{\phi_{\tau}}(t)=(x_{\tau}(t),y_{\tau}(t))=\boldsymbol{\phi_{\tau}}(t;\tau,{\boldsymbol{Q}}(\tau,a)).

From (16) combined with Lemma 28, we see that (ϕ();τ){\mathcal{H}}(\boldsymbol{\phi}(\cdot);\tau) is strictly increasing in the interval [τ,T(τ,a)][\tau,T(\tau,a)]. In particular, recalling (39), we have

(ϕ(t);τ)>(𝑸(τ,a);τ)=(ϕ𝝉(s);τ){\mathcal{H}}(\boldsymbol{\phi}(t);\tau)>{\mathcal{H}}(\boldsymbol{Q}(\tau,a);\tau)={\mathcal{H}}(\boldsymbol{\phi_{\tau}}(s);\tau)

for every t]τ,T(τ,a)]t\in{}]\tau,T(\tau,a)] and s[τ,T1(τ,a)]s\in[\tau,T^{1}(\tau,a)]. Thus, we get

ψ(x)<ψτ(x) for any 0x<Qx(τ,a),\psi(x)<\psi_{\tau}(x)\qquad\mbox{ for any }0\leq x<Q_{x}(\tau,a)\,, (42)

and, as an immediate consequence,

Ry(τ,a)<Ry1(τ,a).R_{y}(\tau,a)<R^{1}_{y}(\tau,a)\,.

The statement of the lemma holds if we prove that x(t)<xτ(t)x(t)<x_{\tau}(t) for any t]τ,T(τ,a)]t\in{}]\tau,T(\tau,a)]. Observe first that ϕ(τ)=ϕ𝝉(τ)\boldsymbol{\phi}(\tau)=\boldsymbol{\phi_{\tau}}(\tau). Then, from (11) and (15) we find that x(τ)x(\tau) equals xτ(τ)x_{\tau}(\tau) along with its first and second derivatives, however

d3dt3[x(τ)xτ(τ)]=y¨(τ)y¨τ(τ)p1|y(τ)|2pp1=K˙(τ)x(τ)q1p1|y(τ)|2pp1<0.\frac{d^{3}}{dt^{3}}[x(\tau)-x_{\tau}(\tau)]=\frac{\ddot{y}(\tau)-\ddot{y}_{\tau}(\tau)}{p-1}|y(\tau)|^{\frac{2-p}{p-1}}=-\dot{K}(\tau)\frac{x(\tau)^{q-1}}{p-1}|y(\tau)|^{\frac{2-p}{p-1}}<0\,.

Hence, x(t)<xτ(t)x(t)<x_{\tau}(t) when tt is in a suitable right neighborhood of τ\tau. Let

T^:=sup{t>τx(s)<xτ(s),s]τ,t]}.\hat{T}:=\sup\{t>\tau\mid x(s)<x_{\tau}(s)\,,\,\forall s\in{}]\tau,t]\,\}\,.

If T^>T(τ,a)\hat{T}>T(\tau,a), the lemma is proved. So, we assume by contradiction that τ<T^T(τ,a)\tau<\hat{T}\leq T(\tau,a). Then, we have x¯=x(T^)=xτ(T^)\bar{x}=x(\hat{T})=x_{\tau}(\hat{T}), ϕ(T^)=(x¯,ψ(x¯))\boldsymbol{\phi}(\hat{T})=(\bar{x},\psi(\bar{x})) and ϕ𝝉(T^)=(x¯,ψτ(x¯))\boldsymbol{\phi_{\tau}}(\hat{T})=(\bar{x},\psi_{\tau}(\bar{x})). Hence, recalling (42),

x˙(T^)\displaystyle\dot{x}(\hat{T}) =αx(T^)|y(T^)|1p1=αx¯|ψ(x¯)|1p1\displaystyle=\alpha x(\hat{T})-|y(\hat{T})|^{\frac{1}{p-1}}=\alpha\bar{x}-|\psi(\bar{x})|^{\frac{1}{p-1}}
<αx¯|ψτ(x¯)|1p1=αxτ(T^)|yτ(T^)|1p1=x˙τ(T^).\displaystyle<\alpha\bar{x}-|\psi_{\tau}(\bar{x})|^{\frac{1}{p-1}}=\alpha x_{\tau}(\hat{T})-|y_{\tau}(\hat{T})|^{\frac{1}{p-1}}=\dot{x}_{\tau}(\hat{T}).

So x>xτx>x_{\tau} in a left neighborhood of T^\hat{T}, giving a contradiction. ∎

In the following Lemma we assume KK bounded. This assumption will be removed via Remark 25.

Lemma 31.

Assume (𝐇)\boldsymbol{({\rm H}_{\ell})}, (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)}, and (𝐖𝐬)\boldsymbol{({\rm W}_{s})}. For any ε>0\varepsilon>0 we define

a=a(ε):=(αqε(1+ε)K¯)p1qp<|Ey(+)|=(αqK(+))p1qp.a=a(\varepsilon):=\left(\frac{\alpha^{q}\varepsilon}{(1+\varepsilon)\overline{K}\ }\right)^{\frac{p-1}{q-p}}<|E_{y}(+\infty)|=\left(\frac{\alpha^{q}}{K(+\infty)}\right)^{\frac{p-1}{q-p}}\,. (43)

Then, there is 𝒯(ε)\mathcal{T}(\varepsilon) such that, for any τ<𝒯(ε)\tau<\mathcal{T}(\varepsilon),

T(τ,a)\displaystyle T(\tau,a) >τ+1αln(a|Ry(τ,a)|)\displaystyle>\tau+\frac{1}{\alpha}\ln\left(\frac{a}{|R_{y}(\tau,a)|}\right) (44)
T1(τ,a)\displaystyle T^{1}(\tau,a) <τ+(1+ε)αln(a|Ry1(τ,a)|).\displaystyle<\tau+\frac{(1+\varepsilon)}{\alpha}\ln\left(\frac{a}{|R_{y}^{1}(\tau,a)|}\right)\,. (45)
Proof.

Let 𝒯(ε)=N(a(ε))\mathcal{T}(\varepsilon)=-N(a(\varepsilon)) be the value provided by Remark 27, and fix τ<𝒯(ε)\tau<\mathcal{T}(\varepsilon). We consider again the trajectories ϕ(t)=(x(t),y(t))=ϕ(t;τ,𝑸(τ,a))\boldsymbol{\phi}(t)=(x(t),y(t))=\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}}(\tau,a)) and ϕ𝝉(t)=(xτ(t),yτ(t))=ϕ𝝉(t;τ,𝑸(τ,a))\boldsymbol{\phi_{\tau}}(t)=(x_{\tau}(t),y_{\tau}(t))=\boldsymbol{\phi_{\tau}}(t;\tau,{\boldsymbol{Q}}(\tau,a)).

Observe that y˙>0\dot{y}>0 along the horizontal segment (a){\mathcal{L}}(a). as well as in the interior of the bounded set enclosed by (a){\mathcal{L}}(a), 𝚪𝝉\boldsymbol{\Gamma_{\tau}} and the yy negative semi-axis. Using this fact, according to Lemma 28 and  (41), we see that

x˙(t)<0<y˙(t),\displaystyle\dot{x}(t)<0<\dot{y}(t)\,, x˙τ(s)<0<y˙τ(s),\displaystyle\dot{x}_{\tau}(s)<0<\dot{y}_{\tau}(s)\,,
ay(t)<0,\displaystyle-a\leq y(t)<0\,, ayτ(s)<0,\displaystyle-a\leq y_{\tau}(s)<0\,,

for any t[τ,T(τ,a)]t\in[\tau,T(\tau,a)] and any s[τ,T1(τ,a)]s\in[\tau,T^{1}(\tau,a)]. Moreover, for both the trajectories we get

x˙<0αx<|y|1p1xq1>|y|q1p1αq1.\dot{x}<0\quad\Rightarrow\quad\alpha x<|y|^{\frac{1}{p-1}}\quad\Rightarrow\quad-x^{q-1}>-\frac{|y|^{\frac{q-1}{p-1}}}{\alpha^{q-1}}\,. (46)

Then, since ϕ(t)=(x(t),y(t))\boldsymbol{\phi}(t)=(x(t),y(t)) solves (11), we find

αy(t)K(t)αq1|y(t)|q1p1<y˙(t)=αy(t)K(t)x(t)q1<αy(t),-\alpha y(t)-\frac{K(t)}{\alpha^{q-1}}|y(t)|^{\frac{q-1}{p-1}}<\dot{y}(t)=-\alpha y(t)-K(t)x(t)^{q-1}<-\alpha y(t)\,, (47)

for any t[τ,T(τ,a)[t\in[\tau,T(\tau,a)[{}. Analogously, since ϕ𝝉(t)=(xτ(t),yτ(t))\boldsymbol{\phi_{\tau}}(t)=(x_{\tau}(t),y_{\tau}(t)) solves (15),

αyτ(t)K(τ)αq1|yτ(t)|q1p1<y˙τ(t)=αyτ(t)K(τ)xτ(t)q1<αyτ(t),-\alpha y_{\tau}(t)-\frac{K(\tau)}{\alpha^{q-1}}|y_{\tau}(t)|^{\frac{q-1}{p-1}}<\dot{y}_{\tau}(t)=-\alpha y_{\tau}(t)-K(\tau)x_{\tau}(t)^{q-1}<-\alpha y_{\tau}(t)\,, (48)

for any t[τ,T1(τ,a)[t\in[\tau,T^{1}(\tau,a)[{}.

Note that (43) guarantees that

K¯αq|y|qpp1ε1+ε,for every y[a,0].\frac{\overline{K}}{\alpha^{q}}|y|^{\frac{q-p}{p-1}}\leq\frac{\varepsilon}{1+\varepsilon}\,,\qquad\mbox{for every }y\in[-a,0].

Thus, we get

αyK(t)αq1|y|q1p1αy(1K¯αq|y|qpp1)α1+εy,\begin{split}-\alpha y-\frac{K(t)}{\alpha^{q-1}}|y|^{\frac{q-1}{p-1}}&\geq-\alpha y\left(1-\frac{\overline{K}}{\alpha^{q}}|y|^{\frac{q-p}{p-1}}\right)\geq-\frac{\alpha}{1+\varepsilon}y\,,\end{split} (49)

for every tt\in{\mathbb{R}} and y[a,0]y\in[-a,0]. In particular, from (47), (48), and (49) we find

α1+εy(t)<y˙(t)<αy(t),α1+εyτ(s)<y˙τ(s)<αyτ(s),\begin{split}-\frac{\alpha}{1+\varepsilon}y(t)<\dot{y}(t)<-\alpha y(t)\,,&\qquad-\frac{\alpha}{1+\varepsilon}y_{\tau}(s)<\dot{y}_{\tau}(s)<-\alpha y_{\tau}(s)\,,\end{split} (50)

for every t[τ,T(τ,a)[t\in[\tau,T(\tau,a)[{} and s[τ,T1(τ,a)[s\in[\tau,T^{1}(\tau,a)[{}.

Consequently, y˙(t)αy(t)<1\frac{\dot{y}(t)}{-\alpha y(t)}<1 for any t[τ,T(τ,a)[t\in[\tau,T(\tau,a)[; hence, recalling the definition of 𝑹\boldsymbol{R} in (38), we obtain

T(τ,a)τ=τT(τ,a)𝑑t>τT(τ,a)y˙(t)αy(t)𝑑t==1αln(y(T(τ,a))y(τ))=1αln(a|Ry(τ,a)|).\begin{split}T(\tau,a)-\tau&=\int_{\tau}^{T(\tau,a)}dt>\int_{\tau}^{T(\tau,a)}\frac{\dot{y}(t)}{-\alpha y(t)}dt=\\ &=-\frac{1}{\alpha}\ln\left(\frac{y(T(\tau,a))}{y(\tau)}\right)=\frac{1}{\alpha}\ln\left(\frac{a}{|R_{y}(\tau,a)|}\right)\,.\end{split}

Analogously, from (50) we also find (1+ε)y˙τ(s)αyτ(s)>1\frac{(1+\varepsilon)\dot{y}_{\tau}(s)}{-\alpha y_{\tau}(s)}>1 for any s[τ;T1(τ,a)[s\in[\tau;T^{1}(\tau,a)[; hence recalling the definition of 𝑹1\boldsymbol{R}^{1} in (40), we get

T1(τ,a)τ=τT1(τ,a)𝑑s<τT1(τ,a)(1+ε)y˙τ(s)αyτ(s)𝑑s==1+εαln(yτ(T1(τ,a))yτ(τ))=1+εαln(a|Ry1(τ,a)|).\begin{split}T^{1}(\tau,a)-\tau&=\int_{\tau}^{T^{1}(\tau,a)}ds<\int_{\tau}^{T^{1}(\tau,a)}\frac{(1+\varepsilon)\dot{y}_{\tau}(s)}{-\alpha y_{\tau}(s)}ds=\\ &=-\frac{1+\varepsilon}{\alpha}\ln\left(\frac{y_{\tau}(T^{1}(\tau,a))}{y_{\tau}(\tau)}\right)=\frac{1+\varepsilon}{\alpha}\ln\left(\frac{a}{|R_{y}^{1}(\tau,a)|}\right)\,.\end{split}

By a simple integration of (50), we obtain a crucial inequality.

Remark 32.

Assume (𝐇)\boldsymbol{({\rm H}_{\ell})}, (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)}, and (𝐖𝐬)\boldsymbol{({\rm W}_{s})}. Then, for every ε>0\varepsilon>0, there is 𝒯(ε)\mathcal{T}(\varepsilon) such that, for any τ<𝒯(ε)\tau<\mathcal{T}(\varepsilon), the trajectory ϕ(t;τ,𝐐(τ,a))=(x(t),y(t))\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}}(\tau,a))=(x(t),y(t)) satisfies

aeα(tτ)\displaystyle a{\rm e}^{-\alpha(t-\tau)} <|y(t)|<aeα1+ε(tτ),for every t]τ,T(τ,a)[.\displaystyle<|y(t)|<a\,{\rm e}^{-\frac{\alpha}{1+\varepsilon}(t-\tau)}\,,\qquad\qquad\mbox{for every }t\in]\tau,T(\tau,a)[\,. (51)
Lemma 33.

Assume (𝐇)\boldsymbol{({\rm H}_{\ell})}, (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)}, and (𝐖𝐬)\boldsymbol{({\rm W}_{s})}. Let ε>0\varepsilon>0 and a=a(ε)>0a=a(\varepsilon)>0 be as in Lemma 31. Then, there are 𝒯(ε)<0\mathcal{T}(\varepsilon)<0 and a constant C=C(a(ε),ε)C=C(a(\varepsilon),\varepsilon) such that

T1(τ,a(ε))<C(a(ε),ε)+|τ|((1+ε)p1),for any τ<𝒯(ε).T^{1}(\tau,a(\varepsilon))<C(a(\varepsilon),\varepsilon)+|\tau|\left((1+\varepsilon)\,\frac{\ell}{\ell_{p}^{*}}-1\right)\,,\hskip 11.38109pt\mbox{for any }\tau<\mathcal{T}(\varepsilon). (52)
Proof.

From Remark 27, (13) and (39), we deduce the following estimates on the yy-coordinate of the point 𝑹1(τ,a)\boldsymbol{R}^{1}(\tau,a) introduced in (40):

p1p|Ry1(τ,a)|pp1=(𝑹1(τ,a);τ)=(𝑸(τ,a);τ)=c(a)eτ+o(eτ),\frac{p-1}{p}|R^{1}_{y}(\tau,a)|^{\frac{p}{p-1}}={\mathcal{H}}(\boldsymbol{R}^{1}(\tau,a);\tau)={\mathcal{H}}({\boldsymbol{Q}}(\tau,a);\tau)=c(a){\rm e}^{\ell\tau}+o\left({\rm e}^{\ell\tau}\right)\,,

for any τ<𝒯(ε)\tau<\mathcal{T}(\varepsilon), possibly choosing a larger |𝒯(ε)||\mathcal{T}(\varepsilon)|. So, there is c1(a)=(c(a)pp1)p1p>0c^{1}(a)=\left(c(a)\frac{p}{p-1}\right)^{\frac{p-1}{p}}>0, which is independent of τ\tau, such that

|Ry1(τ,a)|=c1(a)ep1pτ+o(ep1pτ),for any τ<𝒯(ε).|R^{1}_{y}(\tau,a)|=c^{1}(a){\rm e}^{\frac{p-1}{p}\ell\tau}+o\left({\rm e}^{\frac{p-1}{p}\ell\tau}\right)\,,\qquad\mbox{for any }\tau<\mathcal{T}(\varepsilon). (53)

Then, by (45) in Lemma 31, setting C(a,ε)=1+εαln(2ac1(a))C(a,\varepsilon)=\frac{1+\varepsilon}{\alpha}\ln\left(\frac{2a}{c^{1}(a)}\right), we find

T1(τ,a)<τ+1+εαln(2ac1(a)ep1pτ)==τ+C(a,ε)+|τ|1+εαp1p==C(a,ε)+|τ|(1+εαp1p1),\begin{split}T^{1}(\tau,a)&<\tau+\frac{1+\varepsilon}{\alpha}\ln\left(\frac{2a}{c^{1}(a){\rm e}^{\frac{p-1}{p}\ell\tau}}\right)=\\ \vskip 3.0pt plus 1.0pt minus 1.0pt\cr&=\tau+C(a,\varepsilon)+|\tau|\ \ell\ \frac{1+\varepsilon}{\alpha}\ \frac{p-1}{p}=\\ \vskip 3.0pt plus 1.0pt minus 1.0pt\cr&=C(a,\varepsilon)+|\tau|\left(\ell\ \frac{1+\varepsilon}{\alpha}\ \frac{p-1}{p}-1\right)\,,\end{split} (54)

for any τ<𝒯(ε)\tau<\mathcal{T}(\varepsilon). Since

p=npp1=αpp1,\ell_{p}^{*}=\frac{n-p}{p-1}=\frac{\alpha p}{p-1}\,, (55)

we conclude. ∎

4.1 The case <𝒑\boldsymbol{\ell<\ell_{p}^{*}}.

Proposition 34.

Assume (𝐇)\boldsymbol{({\rm H}_{\ell})}, (𝐖𝐬)\boldsymbol{({\rm W}_{s})} and that K˙(t)>0\dot{K}(t)>0 for any tt\in{\mathbb{R}}. If <p\ell<\ell_{p}^{*}, then there exists a>0a>0 such that

limτT(τ,a)=.\lim_{\tau\to-\infty}T(\tau,a)=-\infty\,.
Proof.

We simply need to choose ε>0\varepsilon>0 in Lemma 33 satisfying the assumption (1+ε)<p\ell(1+\varepsilon)<\ell_{p}^{*}. Then, we set a=a(ε)a=a(\varepsilon) as in (43), so that Lemma 33 permits us to conclude the proof recalling that T(τ,a)<T1(τ,a)T(\tau,a)<T_{1}(\tau,a) by Lemma 30. ∎

Proposition 35.

Assume (𝐇)\boldsymbol{({\rm H}_{\ell})}, (𝐖𝐬)\boldsymbol{({\rm W}_{s})} and that K˙(t)>0\dot{K}(t)>0 for any tt\in{\mathbb{R}}. If <p\ell<\ell_{p}^{*}, then there exists D>0D>0 such that ]D,+[J{}]D,+\infty[{}\subset J and limd+R(d)=0\lim_{d\to+\infty}R(d)=0.

Proof.

Let a=a(ε)>0a=a(\varepsilon)>0 be as in Proposition 34. Then, we can recover the constant N(a)>0N(a)>0 provided by Remark 27, and the function d(a)d_{{\mathcal{L}}(a)} defined in Lemma 20 such that ϕ(t;d(a)(τ)):=ϕ(t;τ,𝑸(τ,a))\boldsymbol{\phi}(t;d_{{\mathcal{L}}(a)}(\tau)):=\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}}(\tau,a)). From Lemma 20 there exists D=D(a)D=D(a) such that the inverse function τ(a):]D,+[],N(a)[\tau_{{}_{{\mathcal{L}}(a)}}:{}]D,+\infty[{}\to{}]-\infty,-N(a)[{} is continuous and satisfies limd+τ(a)(d)=\lim_{d\to+\infty}\tau_{{}_{{\mathcal{L}}(a)}}(d)=-\infty. In particular, ]D,+[J\,]D,+\infty[\,\subset J.

Then, from Proposition 34 and Remark 29, we find

limd+T(τ(a)(d),a)=limτT(τ,a)=.\lim_{d\to+\infty}T(\tau_{{}_{{\mathcal{L}}(a)}}\!(d),a)=\lim_{\tau\to-\infty}T(\tau,a)=-\infty.

Therefore, the first zero R(d)R(d) of u(r;d)u(r;d) satisfies

limd+R(d)=limd+eT(τ(a)(d),a)=0,\lim_{d\to+\infty}R(d)=\lim_{d\to+\infty}{\rm e}^{T(\tau_{{}_{{\mathcal{L}}(a)}}(d),a)}=0\,,

thus concluding the proof. ∎

Since the previous results focus their attention on a neighborhood of τ=\tau=-\infty, recalling Remark 25 we can remove the hypothesis (𝐖𝒔)\boldsymbol{({\rm W}_{s})} and the monotonicity assumption on KK.

Proposition 36.

Assume (𝐇)\boldsymbol{({\rm H}_{\ell})}. If <p\ell<\ell_{p}^{*}, then there exists a>0a>0 such that

limτT(τ,a)=.\lim_{\tau\to-\infty}T(\tau,a)=-\infty\,.

Now, repeating the argument of the proof of Proposition 35 combined with the truncation argument of the proof of Proposition 36, we obtain the asymptotic behavior of R(d)R(d) for large values of dd, which will allow us to prove the part of Theorem 1 concerning the <p\ell<\ell_{p}^{*} case.

Proposition 37.

Let assumption (𝐇)\boldsymbol{({\rm H}_{\ell})} hold with <p\ell<\ell_{p}^{*}. Then, there exists D^>0\hat{D}>0 such that ]D^,+[J{}]\hat{D},+\infty[{}\subset J and limd+R(d)=0\displaystyle{\lim_{d\to+\infty}R(d)=0}.

4.2 The case 𝒑\boldsymbol{\ell\geq\ell_{p}^{*}}.

In this subsection we focus our attention on the opposite case p\ell\geq\ell_{p}^{*}. We invite the reader to take in mind Lemma 20, i.e. the intersection time τ(d)\tau_{{\mathcal{L}}}(d)\to-\infty if and only if d+d\to+\infty .

Remark 38.

Recalling the definition of α\alpha in (2) we see that if ε<1/α\varepsilon<1/\alpha then

p<11+εqαp1.\ell_{p}^{*}<\frac{1}{1+\varepsilon}\ \frac{q\alpha}{p-1}\,. (56)
Lemma 39.

Let assumptions (𝐇)\boldsymbol{({\rm H}_{\ell})}, (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)}, and (𝐖𝐬)\boldsymbol{({\rm W}_{s})} hold with p\ell\geq\ell^{*}_{p}. Let us fix ε<1/α\varepsilon<1/\alpha and define a=a(ε)a=a(\varepsilon) as in (43). Assume that there exists a sequence (τn)n(\tau_{n})_{n} with τn\tau_{n}\to-\infty, satisfying limn+T(τn,a)=\displaystyle{\lim_{n\to+\infty}T(\tau_{n},a)}=-\infty. Then, if nn is sufficiently large,

(𝑹(τn,a);T(τn,a))2(𝑸(τn,a);τn).{\mathcal{H}}({\boldsymbol{R}}(\tau_{n},a);T(\tau_{n},a))\leq 2{\mathcal{H}}({\boldsymbol{Q}}(\tau_{n},a);\tau_{n})\,.
Proof.

Since both τn\tau_{n} and T(τn,a)T(\tau_{n},a) converge to -\infty, according to Remark 25, we can set K˙(t)<2Bet2Bept\dot{K}(t)<2B\ell{\rm e}^{\ell t}\leq 2B\ell{\rm e}^{\ell^{*}_{p}t} for every t[τn,T(τn,a)]t\in[\tau_{n},T(\tau_{n},a)]. Then, using (14), (46), Remark 32 and (56), we get

(𝑹(τn,a);T(τn,a))(𝑸(τn,a);τn)==τnT(τn,a)K˙(t)|x(t)|qq𝑑t<<2BqαqτnT(τn,a)ept|y(t)|qp1𝑑t<<2Bqαqaqp1epτn0T(τn,a)τne(pα1+εqp1)s𝑑s<<2Bqαqaqp1epτn0+e(p11+εqαp1)s𝑑s==2Bqαqaqp1111+εqαp1pepτn=:C(ε)epτn,\begin{split}{\mathcal{H}}({\boldsymbol{R}}(\tau_{n},a);T(\tau_{n},a))&\,-{\mathcal{H}}({\boldsymbol{Q}}(\tau_{n},a);\tau_{n})\,=\\ &=\int_{\tau_{n}}^{T(\tau_{n},a)}\,\dot{K}(t)\frac{|x(t)|^{q}}{q}\,dt\,<\\ &<\frac{2B\ell}{q\alpha^{q}}\int_{\tau_{n}}^{T(\tau_{n},a)}{\rm e}^{\ell^{*}_{p}t}|y(t)|^{\frac{q}{p-1}}\,dt\,<\\ &<\frac{2B\ell}{q\alpha^{q}}a^{\frac{q}{p-1}}{\rm e}^{\ell^{*}_{p}\tau_{n}}\,\int_{0}^{T(\tau_{n},a)-\tau_{n}}{\rm e}^{\left(\ell^{*}_{p}-\frac{\alpha}{1+\varepsilon}\,\frac{q}{p-1}\right)s}\,ds\,<\\ &<\frac{2B\ell}{q\alpha^{q}}a^{\frac{q}{p-1}}{\rm e}^{\ell^{*}_{p}\tau_{n}}\int_{0}^{+\infty}{\rm e}^{\left(\ell^{*}_{p}-\frac{1}{1+\varepsilon}\,\frac{q\alpha}{p-1}\right)s}\,ds\,=\\ &=\frac{2B\ell}{q\alpha^{q}}a^{\frac{q}{p-1}}\,\frac{1}{\frac{1}{1+\varepsilon}\,\frac{q\alpha}{p-1}-\ell^{*}_{p}}\,{\rm e}^{\ell^{*}_{p}\tau_{n}}=:C_{\mathcal{H}}(\varepsilon)\ {\rm e}^{\ell^{*}_{p}\tau_{n}}\,,\\ \end{split}

when nn is sufficiently large.

We argue by contradiction, and assume that there exists a subsequence of τn\tau_{n}, still called τn\tau_{n} for simplicity, which satisfies

(𝑹(τn,a);T(τn,a))>2(𝑸(τn,a);τn),{\mathcal{H}}({\boldsymbol{R}}(\tau_{n},a);T(\tau_{n},a))\,>2{\mathcal{H}}({\boldsymbol{Q}}(\tau_{n},a);\tau_{n})\,,

so that, for sufficiently large nn,

(𝑹(τn,a);T(τn,a))<2[(𝑹(τn,a);T(τn,a))(𝑸(τn,a);τn)]2C(ε)epτn.\begin{split}{\mathcal{H}}({\boldsymbol{R}}(\tau_{n},a);T(\tau_{n},a))\,&<2\left[{\mathcal{H}}({\boldsymbol{R}}(\tau_{n},a);T(\tau_{n},a))-{\mathcal{H}}({\boldsymbol{Q}}(\tau_{n},a);\tau_{n})\right]\,\leq\\ &\leq 2C_{\mathcal{H}}(\varepsilon)\ {\rm e}^{\ell^{*}_{p}\tau_{n}}\,.\end{split}

Then, recalling the definition of {\mathcal{H}} in (13) and the definition of 𝑹\boldsymbol{R} in (38), we get

|Ry(τn,a)|<C^(ε)epp1pτn,|R_{y}(\tau_{n},a)|<\hat{C}_{\mathcal{H}}(\varepsilon)\ {\rm e}^{\ell^{*}_{p}\frac{p-1}{p}\tau_{n}}\,,

where C^(ε)=[2C(ε)pp1]p1p\hat{C}_{\mathcal{H}}(\varepsilon)=\left[2C_{\mathcal{H}}(\varepsilon)\,\frac{p}{p-1}\right]^{\frac{p-1}{p}}. Hence, setting C~(ε):=1αln(aC^(ε))\widetilde{C}_{\mathcal{H}}(\varepsilon):={\frac{1}{\alpha}}\ln\left(\frac{a}{\hat{C}_{\mathcal{H}}(\varepsilon)}\right), from (44) and (55) it follows that

T(τn,a)>τn+1αln(aC^(ε)epp1pτn)==C~(ε)+τn(1pp1pα)=C~(ε),\begin{split}T(\tau_{n},a)\,&>\tau_{n}+\frac{1}{\alpha}\ln\left(\frac{a}{\hat{C}_{\mathcal{H}}(\varepsilon)}{\rm e}^{-\ell^{*}_{p}\frac{p-1}{p}\tau_{n}}\right)\,=\\ &=\widetilde{C}_{\mathcal{H}}(\varepsilon)+\tau_{n}\left(1-\ell^{*}_{p}\,\frac{p-1}{p\alpha}\right)=\widetilde{C}_{\mathcal{H}}(\varepsilon)\,,\end{split}

which contradicts our assumption limn+T(τn,a)=\displaystyle{\lim_{n\to+\infty}T(\tau_{n},a)=-\infty}. ∎

Lemma 40.

Let assumption (𝐇)\boldsymbol{({\rm H}_{\ell})} hold with p\ell\geq\ell_{p}^{*}. Fix ε<1/α\varepsilon<1/\alpha and define

a=a(ε):=(αqε[K()+1](1+ε))p1qp.a=a(\varepsilon):=\left(\frac{\alpha^{q}\varepsilon}{[K(-\infty)+1](1+\varepsilon)}\right)^{\frac{p-1}{q-p}}\,. (57)

Let us assume that there exist a sequence (τn)n(\tau_{n})_{n} with τn\tau_{n}\to-\infty, such that limn+T(τn,a)=\displaystyle{\lim_{n\to+\infty}T(\tau_{n},a)}=-\infty. Then, if nn is sufficiently large,

(𝑹(τn,a);T(τn,a))2(𝑸(τn,a);τn).{\mathcal{H}}({\boldsymbol{R}}(\tau_{n},a);T(\tau_{n},a))\leq 2{\mathcal{H}}({\boldsymbol{Q}}(\tau_{n},a);\tau_{n})\,.
Proof.

Arguing as in Remark 25, we can modify system (11), replacing KK with K^\hat{K} and notice that we can apply Lemma 39 in this case. From the hypothesis, we deduce that, for nn large, we have τn<T(τn,a)<T^0\tau_{n}<T(\tau_{n},a)<\hat{T}_{0}. Since K^\hat{K} and the original KK coincide on ],T^0]{}]-\infty,\hat{T}_{0}], we easily conclude. ∎

Proposition 41.

Assume (𝐇)\boldsymbol{({\rm H}_{\ell})} with p\ell\geq\ell_{p}^{*}. Fix ε<1/α\varepsilon<1/\alpha, and define a=a(ε)a=a(\varepsilon) as in (57). Then,

lim infτT(τ,a)>,\liminf_{\tau\to-\infty}T(\tau,a)>-\infty\,,

where we set T(τ,a):=+T(\tau,a):=+\infty, if the corresponding trajectory ϕ(t;τ,𝐐(τ,a))\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}}(\tau,a)) does not cross the negative yy-axis.

Proof.

We argue by contradiction assuming the existence of a sequence (τn)n(\tau_{n})_{n}, with τn\tau_{n}\to-\infty and satisfying limn+T(τn,a)=\lim_{n\to+\infty}T(\tau_{n},a)=-\infty.

Since we are focusing our attention on a neighborhood of t=t=-\infty, we can again suitably modify KK in the function K^\hat{K} as suggested in Remark 25, in order to ensure the validity of the hypotheses of Lemma 39.

Hence, we can assume, without loss of generality, that (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)} and (𝐖𝒔)\boldsymbol{({\rm W}_{s})} hold, too. So, according to (14), the energy {\mathcal{H}} is increasing along the trajectories. As a consequence, from Lemma 40 and Remark 27, we get

12c(a)eτn(𝑸(τn,a);τn)\displaystyle\tfrac{1}{2}c(a){\rm e}^{\ell\tau_{n}}\leq{\mathcal{H}}({\boldsymbol{Q}}(\tau_{n},a);\tau_{n}) (𝑹(τn,a);T(τn,a))\displaystyle\leq{\mathcal{H}}({\boldsymbol{R}}(\tau_{n},a);T(\tau_{n},a))
2(𝑸(τn,a);τn)3c(a)eτn,\displaystyle\leq 2{\mathcal{H}}({\boldsymbol{Q}}(\tau_{n},a);\tau_{n})\leq 3c(a){\rm e}^{\ell\tau_{n}},

for nn sufficiently large. So, recalling the definition of {\mathcal{H}} in (13) and the definition of 𝑹\boldsymbol{R} in (38), we can find a positive constant cˇ1(a)\check{c}_{1}(a) such that

|Ry(τn,a)|cˇ1(a)ep1pτn.|R_{y}(\tau_{n},a)|\leq\check{c}_{1}(a){\rm e}^{\frac{p-1}{p}\,\ell\tau_{n}}\,.

Moreover, we can use the estimate in Remark 32 to get

aeα(T(τn,a)τn)|Ry(τn,a)|,a{\rm e}^{-\alpha(T(\tau_{n},a)-\tau_{n})}\leq|R_{y}(\tau_{n},a)|\,,

for nn sufficiently large. Hence, we have

aeα(T(τn,a)τn)cˇ1(a)ep1pτn,a{\rm e}^{-\alpha(T(\tau_{n},a)-\tau_{n})}\leq\check{c}_{1}(a){\rm e}^{\frac{p-1}{p}\,\ell\tau_{n}}\,,\qquad

leading to

T(τn,a)\displaystyle T(\tau_{n},a) 1αln(acˇ1(a))+τn[11αp1p]\displaystyle\geq\frac{1}{\alpha}\ln\left(\frac{a}{\check{c}_{1}(a)}\right)+\tau_{n}\left[1-\frac{1}{\alpha}\,\frac{p-1}{p}\,\ell\right]
=Cˇ1(ε)+τn[1p]Cˇ1(ε).\displaystyle=\check{C}_{1}(\varepsilon)+\tau_{n}\left[1-\frac{\ell}{\ell_{p}^{*}}\right]\geq\check{C}_{1}(\varepsilon)\,.

which is in contradiction with limn+T(τn,a)=\lim_{n\to+\infty}T(\tau_{n},a)=-\infty. The proposition is thus proved. ∎

When >p\ell>\ell_{p}^{*} (with the strict inequality), we find a more precise estimate.

Proposition 42.

Assume (𝐇)\boldsymbol{({\rm H}_{\ell})} with >p\ell>\ell_{p}^{*}. Fix ε>0\varepsilon>0 with ε<1/α\varepsilon<1/\alpha and define a=a(ε)a=a(\varepsilon) as in (57).

Assume that there is N~(a)>0\widetilde{N}(a)>0 such that, for every τ<N~(a)\tau<-\widetilde{N}(a), there exists a time T(τ,a)T(\tau,a)\in{\mathbb{R}} at which the trajectory ϕ(t;τ,𝐐(τ,a))\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}}(\tau,a)) crosses the yy negative semi-axis, and x˙(t;τ,𝐐(τ,a))<0\dot{x}(t;\tau,{\boldsymbol{Q}}(\tau,a))<0 for any τtT(τ,a)\tau\leq t\leq T(\tau,a). Then

limτT(τ,a)=+.\lim_{\tau\to-\infty}T(\tau,a)=+\infty\,.
Proof.

Let Tˇ:=lim infτT(τ,a){+}\check{T}:=\liminf_{\tau\to-\infty}T(\tau,a)\in{\mathbb{R}}\cup\{+\infty\}, given by Proposition 41. Without loss of generality, we choose the value T^0\hat{T}_{0} provided by Remark 25 so that T^0Tˇ\hat{T}_{0}\leq\check{T}. Recalling Remark 38, we first fix δ]0,1[\delta\in{}]0,1[{} satisfying

p<1δ1+εqαp1,\ell_{p}^{*}<\frac{1-\delta}{1+\varepsilon}\ \frac{q\alpha}{p-1}\,, (58)

and then τ¯δ<N~(a)\overline{\tau}_{\delta}<-\widetilde{N}(a) such that

τ<δτ<T^01Tˇ1<T(τ,a),for any τ<τ¯δ.\tau<\delta\tau<\hat{T}_{0}-1\leq\check{T}-1<T(\tau,a),\quad\mbox{for any }\tau<\overline{\tau}_{\delta}.

Since >p\ell>\ell_{p}^{*}, we can introduce 1\ell_{1}\leq\ell such that

p<11δ1+εqαp1<11+εqαp1.\ell_{p}^{*}<\ell_{1}\leq\frac{1-\delta}{1+\varepsilon}\ \frac{q\alpha}{p-1}<\frac{1}{1+\varepsilon}\ \frac{q\alpha}{p-1}\,. (59)

Consider the trajectory of system (11) departing from 𝑸(τ,a){\boldsymbol{Q}}(\tau,a) at the time τ\tau and the points

𝑺(τ,a)=ϕ(δτ;τ,𝑸(τ,a)),𝑹(τ,a)=ϕ(T(τ,a);τ,𝑸(τ,a)).{\boldsymbol{S}}(\tau,a)={\boldsymbol{\phi}}(\delta\tau;\tau,{\boldsymbol{Q}}(\tau,a))\,,\qquad{\boldsymbol{R}}(\tau,a)={\boldsymbol{\phi}}(T(\tau,a);\tau,{\boldsymbol{Q}}(\tau,a))\,.

When we focus our attention on the trajectory ϕ(;τ,𝑸(τ,a)){\boldsymbol{\phi}}(\cdot;\tau,{\boldsymbol{Q}}(\tau,a)) restricted to the interval [τ,δτ]],T^0[[\tau,\delta\tau]\subset{}]-\infty,\hat{T}_{0}[{}, recalling the truncation argument in Remark 25, we can assume that both (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)} and (𝐖𝒔)\boldsymbol{({\rm W}_{s})} hold. So, we can argue as in the proof of Lemma 39, to obtain 𝒯(ε,δ)τ¯δ\mathcal{T}(\varepsilon,\delta)\leq\overline{\tau}_{\delta} such that

(𝑺(τ,a);δτ)\displaystyle{\mathcal{H}}({\boldsymbol{S}}(\tau,a);\delta\tau) (𝑸(τ,a);τ)=\displaystyle\,-{\mathcal{H}}({\boldsymbol{Q}}(\tau,a);\tau)=
=τδτK˙(t)|x(t)|qq𝑑t\displaystyle=\int_{\tau}^{\delta\tau}\,\dot{K}(t)\frac{|x(t)|^{q}}{q}\,dt\leq
2Bqαqτδτe1t|y(t)|qp1𝑑t\displaystyle\leq\frac{2B\ell}{q\alpha^{q}}\int_{\tau}^{\delta\tau}{\rm e}^{\ell_{1}t}|y(t)|^{\frac{q}{p-1}}\,dt\leq
2Bqαqaqp1e1τ0+e(1α1+εqp1)s𝑑s=C(ε)e1τ,\displaystyle\leq\frac{2B\ell}{q\alpha^{q}}a^{\frac{q}{p-1}}{\rm e}^{\ell_{1}\tau}\int_{0}^{+\infty}{\rm e}^{\left(\ell_{1}-\frac{\alpha}{1+\varepsilon}\,\frac{q}{p-1}\right)s}\,ds=C_{\mathcal{H}}(\varepsilon)\ {\rm e}^{\ell_{1}\tau}, (60)

for any τ<𝒯(ε,δ)\tau<\mathcal{T}(\varepsilon,\delta). Moreover, from (51) and x˙(δτ)<0\dot{x}(\delta\tau)<0, we have

x(δτ)<1α|y(δτ)|1p1<1αa1p1e1δ1+εαp1τfor any τ<𝒯(ε,δ).x(\delta\tau)<\frac{1}{\alpha}|y(\delta\tau)|^{\frac{1}{p-1}}<\frac{1}{\alpha}\,a^{\frac{1}{p-1}}\,{\rm e}^{\frac{1-\delta}{1+\varepsilon}\frac{\alpha}{p-1}\tau}\quad\mbox{for any }\tau<\mathcal{T}(\varepsilon,\delta). (61)

Assume by contradiction that there is M>0M>0 and a sequence (τn)n(\tau_{n})_{n} such that τn\tau_{n}\to-\infty and T(τn,a)MT(\tau_{n},a)\leq M.

Let us now focus our attention on [δτn,T(τn,a)]],M][\delta\tau_{n},T(\tau_{n},a)]\subset{}]-\infty,M]. We remark that the validity of (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)} and (𝐖𝒔)\boldsymbol{({\rm W}_{s})} is not guaranteed anymore in [δτn,T(τn,a)][\delta\tau_{n},T(\tau_{n},a)]; however x˙<0\dot{x}<0, so x(t)<x(δτn)x(t)<x(\delta\tau_{n}) in this interval. Hence, we can compute

(𝑹(τn,a);T(τn,a))(𝑺(τn,a);δτn)=δτnT(τn,a)K˙(t)|x(t)|qq𝑑t1qδτnT(τn,a)max{K˙(t),0}|x(δτn)|q𝑑t:=Δ(τn)qαq.{\mathcal{H}}({\boldsymbol{R}}(\tau_{n},a);T(\tau_{n},a))\,-{\mathcal{H}}({\boldsymbol{S}}(\tau_{n},a);\delta\tau_{n})=\int_{\delta\tau_{n}}^{T(\tau_{n},a)}\,\dot{K}(t)\frac{|x(t)|^{q}}{q}\,dt\leq\\ \leq\frac{1}{q}\int_{\delta\tau_{n}}^{T(\tau_{n},a)}\max\{\dot{K}(t),0\}|x(\delta\tau_{n})|^{q}\,dt:=\frac{\Delta(\tau_{n})}{q\alpha^{q}}\,. (62)

Then, using (61), we find

Δ(τn)\displaystyle\Delta(\tau_{n}) [Mmax{K˙(t),0}𝑑t]aqp1e1δ1+εqαp1τn=\displaystyle\leq\,\left[\int_{-\infty}^{M}\max\{\dot{K}(t),0\}\,dt\right]a^{\frac{q}{p-1}}\,{\rm e}^{\frac{1-\delta}{1+\varepsilon}{\frac{q\alpha}{p-1}}\tau_{n}}=
:=K(M)aqp1e1δ1+εqαp1τn.\displaystyle:=\mathcal{I}_{K}(M)\,a^{\frac{q}{p-1}}\,{\rm e}^{\frac{1-\delta}{1+\varepsilon}{\frac{q\alpha}{p-1}}\tau_{n}}\,.

Plugging this last inequality in (62) and using (59), we obtain

(𝑹(τn,a);T(τn,a))(𝑺(τn,a);δτn)K(M)aqp1e1τn,\begin{split}{\mathcal{H}}({\boldsymbol{R}}(\tau_{n},a);T(\tau_{n},a))&\,-{\mathcal{H}}({\boldsymbol{S}}(\tau_{n},a);\delta\tau_{n})\leq\mathcal{I}_{K}(M)a^{\frac{q}{p-1}}{\rm e}^{\ell_{1}\tau_{n}}\,,\end{split} (63)

for nn sufficiently large.

Hence, summing (60) and (63) with the estimate in Remark 27, since 1\ell\geq\ell_{1}, we get the existence of a constant C~(ε)>0\tilde{C}_{\mathcal{H}}(\varepsilon)>0 satisfying

p1p|Ry(τn,a)|pp1=(𝑹(τn,a),T(τn,a))C~(ε)e1τn.\frac{p-1}{p}|R_{y}(\tau_{n},a)|^{\frac{p}{p-1}}={\mathcal{H}}({\boldsymbol{R}}(\tau_{n},a),T(\tau_{n},a))\leq\tilde{C}_{\mathcal{H}}(\varepsilon){\rm e}^{\ell_{1}\tau_{n}}\,.

Finally, since y˙<αy\dot{y}<-\alpha y in [τn,T(τn,a)[[\tau_{n},T(\tau_{n},a)[{}, and so 1αy˙y<1-\frac{1}{\alpha}\,\frac{\dot{y}}{y}<1, for a certain constant C~H(ε)\widetilde{C}_{H}(\varepsilon), we obtain

T(τn,a)>τn1ατnT(τn,a)y˙(s)y(s)𝑑s==τn1αa|Ry(τn,a)|dyy=τn+1αln(a|Ry(τn,a)|)C~H(ε)+τn(11p),\begin{split}T(\tau_{n},a)&>\tau_{n}-\frac{1}{\alpha}\int_{\tau_{n}}^{T(\tau_{n},a)}\frac{\dot{y}(s)}{y(s)}\,ds\,=\\ &=\tau_{n}-\frac{1}{\alpha}\int_{a}^{|R_{y}(\tau_{n},a)|}\frac{dy}{y}=\tau_{n}+\frac{1}{\alpha}\ln\left(\frac{a}{|R_{y}(\tau_{n},a)|}\right)\geq\\ &\geq\widetilde{C}_{H}(\varepsilon)+\tau_{n}\left(1-\frac{\ell_{1}}{\ell_{p}^{*}}\right)\,,\end{split}

when nn is large. Recalling (59), we get limn+T(τn,a)=+\lim_{n\to+\infty}T(\tau_{n},a)=+\infty, giving a contradiction. The Proposition is thus proved. ∎

Proposition 43.

Assume (𝐇)\boldsymbol{({\rm H}_{\ell})} and (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)}. Then, J=]0,+[J={}]0,+\infty[{} and

  • if =p\ell=\ell_{p}^{*}, then lim infd+R(d)>0\displaystyle{\liminf_{d\to+\infty}R(d)>0},

  • if >p\ell>\ell_{p}^{*}, then limd+R(d)=+\displaystyle{\lim_{d\to+\infty}R(d)=+\infty}.

Proof.

As stated in Proposition 21, J=]0,+[J={}]0,+\infty[{} is a well-known consequence of assumption (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)}.

Let ε<1α\varepsilon<\frac{1}{\alpha}, define a=a(ε)a=a(\varepsilon) as in (57), and set N=N(a)N=N(a) as in Remark 27. So, Wu(τ)W^{u}(\tau) crosses transversely (a){\mathcal{L}}(a) in 𝑸(τ,a){\boldsymbol{Q}}(\tau,a) for any τ<N\tau<-N.

From Proposition 21 and Lemma 28, we see that for every τ<N\tau<-N there is T(τ,a)T(\tau,a) such that x˙(t;τ,𝑸(τ,a))<0\dot{x}(t;\tau,{\boldsymbol{Q}}(\tau,a))<0 for any τtT(τ,a)\tau\leq t\leq T(\tau,a), and ϕ(t;τ,𝑸(τ,a))\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}}(\tau,a)) crosses the yy negative semi-axis at t=T(τ,a)t=T(\tau,a).

Now, recalling Lemma 20, we consider the function d(a):],N[]D,+[d_{{\mathcal{L}}(a)}:{}]-\infty,-N[{}\to{}]D,+\infty[{} such that u(r;d(a)(τ))u(r;d_{{\mathcal{L}}(a)}(\tau)) is the solution of (4) corresponding to ϕ(t;τ,𝑸(τ,a))\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}}(\tau,a)) via (2), and its inverse τ(a)\tau_{{\mathcal{L}}(a)} satisfying limd+τ(a)(d)=\lim_{d\to+\infty}\tau_{{\mathcal{L}}(a)}(d)=-\infty.

If p\ell\geq\ell_{p}^{*}, Proposition 41 gives lim infτT(τ,a)>\liminf_{\tau\to-\infty}T(\tau,a)>-\infty. Arguing as in the proof of Proposition 35, we have

lim infd+R(d)=lim infd+eT(τ(a)(d),a)=lim infτeT(τ,a)>0.\liminf_{d\to+\infty}R(d)=\liminf_{d\to+\infty}{\rm e}^{T(\tau_{{\mathcal{L}}(a)}(d),a)}=\liminf_{\tau\to-\infty}{\rm e}^{T(\tau,a)}>0\,. (64)

If >p\ell>\ell_{p}^{*}, we are able to apply Proposition 42 and get

limd+R(d)=limd+eT(τ(a)(d),a)=limτeT(τ,a)=+.\lim_{d\to+\infty}R(d)=\lim_{d\to+\infty}{\rm e}^{T(\tau_{{\mathcal{L}}(a)}(d),a)}=\lim_{\tau\to-\infty}{\rm e}^{T(\tau,a)}=+\infty\,. (65)

The proposition is thus proved. ∎

Theorem 2 follows from Proposition 43, cf. §5 for more details.

In order to prove the second part of Theorem 1, we need to remove the monotonicity assumption from Proposition 43. Note that if (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)} does not hold, we cannot ensure the existence of points of JJ in a neighborhood of ++\infty, when p\ell\geq\ell_{p}^{*}. However, we can prove a weaker result.

Proposition 44.

Assume (𝐇)\boldsymbol{({\rm H}_{\ell})}, and define

R~(d)={R(d)if dJ,+if dJ.\tilde{R}(d)=\begin{cases}R(d)&\mbox{if }d\in J\,,\\ +\infty&\mbox{if }d\notin J\,.\end{cases}

If p\ell\geq\ell_{p}^{*}, then there is 0>0{\mathcal{R}}_{0}>0 such that R~(d)0\tilde{R}(d)\geq{\mathcal{R}}_{0} for any d0d\geq 0.

Proof.

Fix 0<ε<1/α0<\varepsilon<1/\alpha, a=a(ε)a=a(\varepsilon) as in (57) and set N(a)N(a) as in Remark 27. Let T^0\hat{T}_{0} be as in Remark 25. From Proposition 41 we deduce that

there is TT^0 such that T(τ,a)T for any τ<N(a).\mbox{there is $T^{\prime}\leq\hat{T}_{0}$ such that $T(\tau,a)\geq T^{\prime}$ for any }\tau<-N(a)\,. (66)

Now, let us fix τ0<N(a)\tau_{0}<-N(a) and denote by W~u(τ0)\tilde{W}^{u}(\tau_{0}) the branch of the unstable manifold Wu(τ0)W^{u}(\tau_{0}) between the origin and 𝑸(τ0,a){\boldsymbol{Q}}(\tau_{0},a). From Remark 17 we see that there is D>0D^{*}>0 such that the function 𝓠𝝉𝟎(d){\boldsymbol{{\cal Q}}}_{\boldsymbol{\tau_{0}}}(d) restricted to d[0,D]d\in[0,D^{*}] gives a parametrization of W~u(τ0)\tilde{W}^{u}(\tau_{0}), i.e. 𝓠𝝉𝟎:[0,D]W~u(τ0){\boldsymbol{{\cal Q}}}_{\boldsymbol{\tau_{0}}}:[0,D^{*}]\to\tilde{W}^{u}(\tau_{0}) is a continuous bijective function.

Using Remark 14 and Lemma 16, reducing τ0\tau_{0} if necessary, we see that if 𝑸W~u(τ0){\boldsymbol{Q}}\in\tilde{W}^{u}(\tau_{0}) then ϕ(t;τ0,𝑸)\boldsymbol{\phi}(t;\tau_{0},{\boldsymbol{Q}}) is close to the corresponding trajectory in Wu()W^{u}(-\infty) when tτ0t\leq\tau_{0}; in particular x(t;τ0,𝑸)>0x(t;\tau_{0},{\boldsymbol{Q}})>0 if tτ0t\leq\tau_{0}. Hence R~(d)eτ0\tilde{R}(d)\geq{\rm e}^{\tau_{0}} for any dDd\leq D^{*}.

Let now d>Dd>D^{*}. Recalling Remark 17, we have 𝓠𝝉𝟎(d)Wu(τ0)W~u(τ0){\boldsymbol{{\cal Q}}}_{\boldsymbol{\tau_{0}}}(d)\in W^{u}(\tau_{0})\setminus\tilde{W}^{u}(\tau_{0}), and using Lemma 20 we see that there is a unique τ(a)(d)<τ0\tau_{{\mathcal{L}}(a)}(d)<\tau_{0} such that the trajectory ϕ(t;τ0,𝓠𝝉𝟎(d))\boldsymbol{\phi}(t;\tau_{0},{\boldsymbol{{\cal Q}}}_{\boldsymbol{\tau_{0}}}(d)) crosses transversely (a){\mathcal{L}}(a) at t=τ(a)(d)<τ0<N(a)t=\tau_{{\mathcal{L}}(a)}(d)<\tau_{0}<-N(a). Then, from (66) we see that

T(τ(a)(d),a)Tfor any d>D.T(\tau_{{\mathcal{L}}(a)}(d),a)\geq T^{\prime}\qquad\mbox{for any }d>D^{*}\,.

Hence R~(d)eT\tilde{R}(d)\geq{\rm e}^{T^{\prime}} for any d>Dd>D^{*}.

Then, setting 0=min{eτ0,eT}{\mathcal{R}}_{0}=\min\{{\rm e}^{\tau_{0}},{\rm e}^{T^{\prime}}\}, the Proposition is proved. ∎

Now, we reprove Proposition 43 by replacing the global assumption (𝐇)\boldsymbol{({\rm H}_{\uparrow}\!\!\!)} with the local assumption np1\ell\leq\frac{n}{p-1}. Motivated by [3, 7, 39], and inspired by [18] and [22], we obtain the following result.

Lemma 45.

Assume (𝐇)\boldsymbol{({\rm H}_{\ell})} with np1\ell\leq\frac{n}{p-1}; assume further either (𝐖𝐬)\boldsymbol{({\rm W}_{s})} or that there is ρ1>0\rho_{1}>0 such that 𝒦(r)𝒦(ρ1){\mathcal{K}}(r)\geq{\mathcal{K}}(\rho_{1}) when rρ1r\geq\rho_{1}. Then, there is D^>0\hat{D}>0 such that ]D^,+[J{}]\hat{D},+\infty[{}\subset J.

Notice that if (𝐇)\boldsymbol{({\rm H}_{\ell})} with np1\ell\leq\frac{n}{p-1} holds and limr+𝒦(r)\lim_{r\to+\infty}{\mathcal{K}}(r) exists and it is positive, either bounded or unbounded, or if K(t)K(t) is asymptotically periodic, then Lemma 45 applies.

The proof of Lemma 45 is rather technical and it is postponed to Section 4.2.1, as well as the related proof of the following adapted version of Proposition 43.

Proposition 46.

Assume (𝐇)\boldsymbol{({\rm H}_{\ell})} with p<np1\ell_{p}^{*}<\ell\leq\frac{n}{p-1}; assume further either (𝐖𝐬)\boldsymbol{({\rm W}_{s})} or that there is ρ1>0\rho_{1}>0 such that 𝒦(r)𝒦(ρ1){\mathcal{K}}(r)\geq{\mathcal{K}}(\rho_{1}) when rρ1r\geq\rho_{1}. Then

limd+R(d)=+.\displaystyle{\lim_{d\to+\infty}R(d)=+\infty}\,.

4.2.1 Proof of Lemma 45 and Proposition 46.

We start by proving Lemma 45 under assumption (𝐖𝒔)\boldsymbol{({\rm W}_{s})}: the following arguments are preliminary to the proof under this hypothesis. The alternative case where it is assumed that 𝒦(r)𝒦(ρ1){\mathcal{K}}(r)\geq{\mathcal{K}}(\rho_{1}) is then obtained as a Corollary.

Let us assume (𝐖𝒔)\boldsymbol{({\rm W}_{s})}, so that we can construct the stable manifold Ws(τ)W^{s}(\tau), see §2 and, in particular, (23).

To develop our construction we need to define several sets, and we invite the reader to follow the argument on Figure 5. Assume (𝐇)\boldsymbol{({\rm H}_{\ell})} and (𝐖𝒔)\boldsymbol{({\rm W}_{s})}, then for any τ\tau\in{\mathbb{R}} we set

¯(x,y)\displaystyle\overline{{\mathcal{H}}}(x,y) =αxy+p1p|y|pp1+K¯|x|qq,\displaystyle=\alpha xy+\frac{p-1}{p}|y|^{\frac{p}{p-1}}+\overline{K}\,\frac{|x|^{q}}{q}\,,
¯(x,y)\displaystyle\underline{{\mathcal{H}}}(x,y) =αxy+p1p|y|pp1+K¯|x|qq.\displaystyle=\alpha xy+\frac{p-1}{p}|y|^{\frac{p}{p-1}}+\underline{K}\,\frac{|x|^{q}}{q}\,.

We define Γ¯:={(x,y)¯(x,y)=0x0}\overline{\Gamma}:=\{(x,y)\mid\overline{{\mathcal{H}}}(x,y)=0\;x\geq 0\} and Γ¯:={(x,y)¯(x,y)=0x0}\underline{\Gamma}:=\{(x,y)\mid\underline{{\mathcal{H}}}(x,y)=0\;x\geq 0\}. Notice that both Γ¯\overline{\Gamma} and Γ¯\underline{\Gamma} are the image of closed regular curves; we denote by F¯\overline{F} and by F¯\underline{F} the bounded sets enclosed by Γ¯\overline{\Gamma} and Γ¯\underline{\Gamma}, respectively: notice that F¯F¯\overline{F}\subset\underline{F}. We denote by 𝑮¯=(G¯x,G¯y)\boldsymbol{\overline{G}}=(\overline{G}_{x},\overline{G}_{y}) the (transversal) intersection between Γ¯\overline{\Gamma} and the isocline x˙=0\dot{x}=0 such that G¯x>0\overline{G}_{x}>0. Then, we denote by 𝑮¯\boldsymbol{\underline{G}} the intersection between the line x=G¯xx=\overline{G}_{x} and Γ¯\underline{\Gamma} contained in x˙<0\dot{x}<0 and by 𝒢\mathcal{G} the vertical segment between 𝑮¯\boldsymbol{\underline{G}} and 𝑮¯\boldsymbol{\overline{G}}. Moreover, we denote by \partial\mathcal{B}^{\uparrow} the branch of Γ¯\overline{\Gamma} between the origin and 𝑮¯\boldsymbol{\overline{G}} contained in x˙0\dot{x}\leq 0 and by \partial\mathcal{B}^{\downarrow} the branch of Γ¯\underline{\Gamma} between the origin and 𝑮¯\boldsymbol{\underline{G}} contained in x˙0\dot{x}\leq 0. Finally, we denote by \mathcal{B} the compact set enclosed by 𝒢\mathcal{G}, \partial\mathcal{B}^{\uparrow} and \partial\mathcal{B}^{\downarrow}, see Figure 5.

We emphasize that if ϕ(t)=(x(t),y(t))\boldsymbol{\phi}(t)=(x(t),y(t)) is a trajectory of (11), we find, according to (16),

ddt¯(ϕ(t))\displaystyle\frac{d}{dt}\,\overline{{\mathcal{H}}}(\boldsymbol{\phi}(t)) =(K¯K(t))x(t)|x(t)|q2x˙(t),\displaystyle=\big{(}\,\overline{K}-K(t)\big{)}\,x(t)\,|x(t)|^{q-2}\,\,\dot{x}(t)\,,
ddt¯(ϕ(t))\displaystyle\frac{d}{dt}\,\underline{{\mathcal{H}}}(\boldsymbol{\phi}(t)) =(K¯K(t))x(t)|x(t)|q2x˙(t).\displaystyle=\big{(}\,\underline{K}-K(t)\big{)}\,x(t)\,|x(t)|^{q-2}\,\,\dot{x}(t)\,.

Using this fact, we easily obtain the following crucial remark

Remark 47.

Assume (𝐇)\boldsymbol{({\rm H}_{\ell})} and (𝐖𝐬)\boldsymbol{({\rm W}_{s})}, then the flow of (11) on 𝒢˘:=𝒢{𝐆¯,𝐆¯}\breve{\mathcal{G}}:=\mathcal{G}\setminus\{\boldsymbol{\overline{G}},\boldsymbol{\underline{G}}\} aims towards the interior of \mathcal{B} for any tt\in{\mathbb{R}}, while on (){(0,0)}(\partial\mathcal{B}^{\uparrow}\cup\partial\mathcal{B}^{\downarrow})\setminus\{(0,0)\} aims towards the exterior of \mathcal{B} for any tt\in{\mathbb{R}}.

Refer to caption   Refer to caption

Figure 5: The constructions needed in the proof of Lemma 45

Let us now fix τ\tau\in{\mathbb{R}}. From Remark 47 and the construction in §2, we see that Ws(τ)W^{s}(\tau) intersects 𝒢˘\breve{\mathcal{G}} (not necessarily transversely), see [22, Lemma 2.8]. Follow Ws(τ)W^{s}(\tau) from the origin towards x>0x>0: denote by 𝑸𝒔(τ)\boldsymbol{Q^{s}}(\tau) the first intersection with 𝒢˘\breve{\mathcal{G}} and by W~s(τ)\tilde{W}^{s}(\tau) the connected branch of Ws(τ)W^{s}(\tau) between the origin and 𝑸𝒔(τ)\boldsymbol{Q^{s}}(\tau).

Moreover, we introduce the set

W^s(τ):={𝑸Ws(τ)x˙(t;τ,𝑸)<0 for any tτ}W~s(τ).\hat{W}^{s}(\tau):=\{{\boldsymbol{Q}}\in W^{s}(\tau)\cap\mathcal{B}\mid\dot{x}(t;\tau,{\boldsymbol{Q}})<0\;\textrm{ for any }t\geq\tau\}\supset\tilde{W}^{s}(\tau)\,. (67)

In particular, using Remark 47, we have

𝑸W^s(τ)\displaystyle{\boldsymbol{Q}}\in\hat{W}^{s}(\tau) ϕ(t;τ,𝑸)W^s(t) for every tτ,\displaystyle\Rightarrow\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}})\in\hat{W}^{s}(t)\mbox{ for every }t\geq\tau\,, (68)
𝑸W~s(τ)\displaystyle{\boldsymbol{Q}}\in\tilde{W}^{s}(\tau) ϕ(t;τ,𝑸)W~s(t) for every tτ.\displaystyle\Rightarrow\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}})\in\tilde{W}^{s}(t)\mbox{ for every }t\geq\tau\,. (69)

In what follows, we recall the argument in [22, pp. 357–360]. Let us consider the autonomous systems (11), where K(t)K(t) is replaced by K¯\overline{K}, respectively K¯\underline{K}, and we denote by

ϕ¯(t;τ,𝑸)=(x¯(t;τ,𝑸),y¯(t;τ,𝑸)), resp. ϕ¯(t;τ,𝑸)=(x¯(t;τ,𝑸),y¯(t;τ,𝑸))\boldsymbol{\overline{\phi}}(t;\tau,{\boldsymbol{Q}})=(\overline{x}(t;\tau,{\boldsymbol{Q}}),\overline{y}(t;\tau,{\boldsymbol{Q}}))\,,\mbox{ resp. }\boldsymbol{\underline{\phi}}(t;\tau,{\boldsymbol{Q}})=(\underline{x}(t;\tau,{\boldsymbol{Q}}),\underline{y}(t;\tau,{\boldsymbol{Q}}))

the trajectories of these systems starting at time τ\tau from the point 𝑸{\boldsymbol{Q}}. From (20), recalling that in the region x˙<0\dot{x}<0 the branch of Γ¯\underline{\Gamma} lies under the corresponding branch of Γ¯\overline{\Gamma}, see Figure 5, we can find C>c>0C>c>0 such that

cenpp(p1)tx¯(t;0,𝑮¯)x¯(t;0,𝑮¯)Cenpp(p1)t, for any t0,c\,{\rm e}^{-\frac{n-p}{p(p-1)}t}\leq\underline{x}(t;0,\boldsymbol{\underline{G}})\leq\overline{x}(t;0,\boldsymbol{\overline{G}})\leq C\,{\rm e}^{-\frac{n-p}{p(p-1)}t}\,,\mbox{ for any $t\geq 0$,}

or, equivalently,

cenpp(p1)(tτ)x¯(t;τ,𝑮¯)x¯(t;τ,𝑮¯)Cenpp(p1)(tτ), for any tτ.c\,{\rm e}^{-\frac{n-p}{p(p-1)}(t-\tau)}\leq\underline{x}(t;\tau,\boldsymbol{\underline{G}})\leq\overline{x}(t;\tau,\boldsymbol{\overline{G}})\leq C\,{\rm e}^{-\frac{n-p}{p(p-1)}(t-\tau)}\,,\mbox{ for any }t\geq\tau\,. (70)

Such estimates permit us to provide analogous ones for the solutions of the non-autonomous system (11).

Lemma 48.

Assume (𝐇)\boldsymbol{({\rm H}_{\ell})} and (𝐖𝐬)\boldsymbol{({\rm W}_{s})}. Let τ\tau\in{\mathbb{R}} and 𝐐^W^s(τ)𝒢\boldsymbol{\hat{Q}}\in\hat{W}^{s}(\tau)\cap\mathcal{G}, then

cenpp(p1)(tτ)x(t;τ,𝑸^)Cenpp(p1)(tτ),for any tτ.c\,{\rm e}^{-\frac{n-p}{p(p-1)}(t-\tau)}\leq x(t;\tau,\boldsymbol{\hat{Q}})\leq C\,{\rm e}^{-\frac{n-p}{p(p-1)}(t-\tau)}\,,\qquad\mbox{for any }t\geq\tau\,. (71)
Proof.

We will prove that

x¯(t;τ,𝑮¯)x(t;τ,𝑸^)x¯(t;τ,𝑮¯), for any tτ.\underline{x}(t;\tau,\boldsymbol{\underline{G}})\leq x(t;\tau,\boldsymbol{\hat{Q}})\leq\overline{x}(t;\tau,\boldsymbol{\overline{G}})\,,\mbox{ for any }t\geq\tau\,.

Then, the conclusion will follow from (70).

We give the proof of the first inequality, the other being similar. Let T=sup{tτ:x¯(s;τ,𝑮¯)x(s;τ,𝑸^) for any s[τ,t]}T=\sup\{t\geq\tau\,:\,\underline{x}(s;\tau,\boldsymbol{\underline{G}})\leq x(s;\tau,\boldsymbol{\hat{Q}})\mbox{ for any }s\in[\tau,t]\,\}. Assume by contradiction that TT\in{\mathbb{R}}. Then, x¯(T;τ,𝑮¯)=x(T;τ,𝑸^)\underline{x}(T;\tau,\boldsymbol{\underline{G}})=x(T;\tau,\boldsymbol{\hat{Q}}) and since 𝑸^W^s(τ)\boldsymbol{\hat{Q}}\in\hat{W}^{s}(\tau), from (68) and Remark 47 the trajectory remains in the interior of \mathcal{B} and in particular we get y¯(T;τ,𝑮¯)<y(T;τ,𝑸^)\underline{y}(T;\tau,\boldsymbol{\underline{G}})<y(T;\tau,\boldsymbol{\hat{Q}}). So, from (11) and (15), we deduce that x¯˙(T;τ,𝑮¯)<x˙(T;τ,𝑸^)\dot{\underline{x}}(T;\tau,\boldsymbol{\underline{G}})<\dot{x}(T;\tau,\boldsymbol{\hat{Q}}), providing x¯(t;τ,𝑮¯)<x(t;τ,𝑸^)\underline{x}(t;\tau,\boldsymbol{\underline{G}})<x(t;\tau,\boldsymbol{\hat{Q}}) in a right neighborhood of TT, leading to a contradiction. ∎

From Lemma 48, we obtain the following result, which has already been proved in [22, Lemma 3.1] focusing the attention on the point 𝑸𝒔(τ)W~s(τ)\boldsymbol{Q^{s}}(\tau)\in\tilde{W}^{s}(\tau), but we repeat the argument here to correct some typos.

Lemma 49.

Assume (𝐖𝐬)\boldsymbol{({\rm W}_{s})} and (𝐇)\boldsymbol{({\rm H}_{\ell})} with np1\ell\leq\frac{n}{p-1}. Then, there is N1>0N_{1}>0 such that (𝐐;τ)<0{\mathcal{H}}({\boldsymbol{Q}};\tau)<0 for any 𝐐W^s(τ){\boldsymbol{Q}}\in\hat{W}^{s}(\tau) and τ<N1\tau<-N_{1}.

Proof.

Let T^0\hat{T}_{0} be as in Remark 25, so that (31) holds.

Take any τT^0\tau\leq\hat{T}_{0} and consider a point 𝑸W^s(τ){\boldsymbol{Q}}\in\hat{W}^{s}(\tau). We denote, for brevity, the trajectory ϕ(t;τ,𝑸)\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}}) as ϕ(t)=(x(t),y(t))\boldsymbol{\phi}(t)=(x(t),y(t)).

We note that ϕ\boldsymbol{\phi} must cross the line 𝒢\mathcal{G} at a time smaller than τ\tau. So, let us denote by T𝒢τT_{\mathcal{G}}\leq\tau the largest value with such a property. In particular, by construction, x(T𝒢)=G¯xx(T_{\mathcal{G}})=\overline{G}_{x}, 0<x(t)<G¯x0<x(t)<\overline{G}_{x} and x˙(t)<0\dot{x}(t)<0 for every t>T𝒢t>T_{\mathcal{G}}. Using Lemma 48, we have

cenpp(p1)(tT𝒢)x(t)Cenpp(p1)(tT𝒢), for any tT𝒢.c\,{\rm e}^{-\frac{n-p}{p(p-1)}(t-T_{\mathcal{G}})}\leq x(t)\leq C\,{\rm e}^{-\frac{n-p}{p(p-1)}(t-T_{\mathcal{G}})}\,,\mbox{ for any }t\geq T_{\mathcal{G}}\,. (72)

Since T𝒢τT^0T_{\mathcal{G}}\leq\tau\leq\hat{T}_{0} we see that xx is decreasing in the interval ]T^0,+[]T𝒢,+[]\hat{T}_{0},+\infty[{}\subset]T_{\mathcal{G}},+\infty[{}; so using (72),

T^0+K˙(t)x(t)qq𝑑t\displaystyle\int_{\hat{T}_{0}}^{+\infty}\dot{K}(t)\frac{x(t)^{q}}{q}\,dt =K(T^0)x(T^0)qq+T^0+K(t)ddt[x(t)qq]𝑑t\displaystyle=-K(\hat{T}_{0})\frac{x(\hat{T}_{0})^{q}}{q}+\int_{\hat{T}_{0}}^{+\infty}K(t)\frac{d}{dt}\left[-\frac{x(t)^{q}}{q}\right]\,dt\,\geq
K(T^0)x(T^0)qqCqqK(T^0)eq(np)p(p1)(T^0T𝒢)=\displaystyle\geq-K(\hat{T}_{0})\frac{x(\hat{T}_{0})^{q}}{q}\,\geq\,-\frac{C^{q}}{q}\,K(\hat{T}_{0})\,{\rm e}^{-\frac{q(n-p)}{p(p-1)}(\hat{T}_{0}-T_{\mathcal{G}})}=
=CqqK(T^0)enp1(T^0T𝒢).\displaystyle=-\frac{C^{q}}{q}\,K(\hat{T}_{0})\,{\rm e}^{-\frac{n}{p-1}(\hat{T}_{0}-T_{\mathcal{G}})}\,.

Moreover, using (31) and (72),

τT^0K˙(t)x(t)qq𝑑t\displaystyle\int_{\tau}^{\hat{T}_{0}}\dot{K}(t)\frac{x(t)^{q}}{q}\,dt Bcq2qτT^0eteq(np)p(p1)(tT𝒢)𝑑t=\displaystyle\geq\frac{B\ell c^{q}}{2q}\int_{\tau}^{\hat{T}_{0}}{\rm e}^{\ell t}\,{\rm e}^{-\frac{q(n-p)}{p(p-1)}(t-T_{\mathcal{G}})}\,dt=
=Bcq2qenp1T𝒢τT^0e(np1)t𝑑t.\displaystyle=\frac{B\ell c^{q}}{2q}{\rm e}^{\frac{n}{p-1}T_{\mathcal{G}}}\int_{\tau}^{\hat{T}_{0}}{\rm e}^{-\left(\frac{n}{p-1}-\ell\right)\,t}\,dt\,.

Summing up, setting b1=CqqK(T^0)enp1T^0b_{1}=\frac{C^{q}}{q}\,K(\hat{T}_{0})\,{\rm e}^{-\frac{n}{p-1}\hat{T}_{0}} and b2=Bcq2qb_{2}=\frac{B\ell c^{q}}{2q}, from (14), we get

(𝑸;τ)\displaystyle{\mathcal{H}}({\boldsymbol{Q}};\tau) =τ+K˙(t)|x(t)|qq𝑑t=\displaystyle=-\int_{\tau}^{+\infty}\dot{K}(t)\frac{|x(t)|^{q}}{q}\,dt\,=
=T^0+K˙(t)x(t)qq𝑑tτT^0K˙(t)x(t)qq𝑑t\displaystyle=-\int_{\hat{T}_{0}}^{+\infty}\dot{K}(t)\frac{x(t)^{q}}{q}\,dt-\int_{\tau}^{\hat{T}_{0}}\dot{K}(t)\frac{x(t)^{q}}{q}\,dt\,\leq
enp1T𝒢[b1b2τT^0e(np1)t𝑑t].\displaystyle\leq{\rm e}^{\frac{n}{p-1}T_{\mathcal{G}}}\left[b_{1}-b_{2}\int_{\tau}^{\hat{T}_{0}}{\rm e}^{-\left(\frac{n}{p-1}-\ell\right)\,t}\,dt\right]\,.

So, since np1\ell\leq\frac{n}{p-1}, the integral diverges as τ\tau\to-\infty, thus giving the proof. ∎

Let us assume now (𝐇)\boldsymbol{({\rm H}_{\ell})} and (𝐖𝒔)\boldsymbol{({\rm W}_{s})}. Recalling the definition of 𝑬{\boldsymbol{E}} given in (18), fix

0<a<(αqK¯)p1qpinf{|Ey(t)|:t}0<a<\left(\frac{\alpha^{q}}{\overline{K}}\right)^{\frac{p-1}{q-p}}\leq\inf\{|E_{y}(t)|\,:\,t\in{\mathbb{R}}\} (73)

and consider the segment (a){\mathcal{L}}(a) defined as in (36).

From Remark 27, there is N(a)>0N(a)>0 such that Wu(τ)W^{u}(\tau) intersects transversely (a){\mathcal{L}}(a) in a point denoted by 𝑸(τ,a){\boldsymbol{Q}}(\tau,a), for every τ<N(a)\tau<-N(a). Moreover, since a>G¯ya>\overline{G}_{y}, a subsegment of (a){\mathcal{L}}(a) joins \partial\mathcal{B}^{\downarrow} with \partial\mathcal{B}^{\uparrow} (transversely), and, consequently, from the previous argument, W~s(τ)\tilde{W}^{s}(\tau) intersects (a){\mathcal{L}}(a), too, for every τ\tau\in{\mathbb{R}}.

Lemma 50.

Assume (𝐖𝐬)\boldsymbol{({\rm W}_{s})} and (𝐇)\boldsymbol{({\rm H}_{\ell})} with np1\ell\leq\frac{n}{p-1}. Then, there is N^(a)>0\hat{N}(a)>0 such that for any τ<N^(a)\tau<-\hat{N}(a) the trajectory ϕ(t;τ,𝐐(τ,a))\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}}(\tau,a)) corresponds to a crossing solution, i.e. there is T=T(τ,a)>τT=T(\tau,a)>\tau such that x(t;τ,𝐐(τ,a))>0x(t;\tau,{\boldsymbol{Q}}(\tau,a))>0 for any t<Tt<T and it becomes null at t=Tt=T. Further x˙(t;τ,𝐐(τ,a))<0\dot{x}(t;\tau,{\boldsymbol{Q}}(\tau,a))<0 for any τtT(τ,a)\tau\leq t\leq T(\tau,a).

Proof.

Let us consider the compact set 𝒞=𝒞(a,τ)\mathcal{C}=\mathcal{C}(a,\tau), see Figure 5, delimited by W~s(τ)\tilde{W}^{s}(\tau), the yy negative semi-axis and the line (a){\mathcal{L}}(a), for any τ\tau\in{\mathbb{R}}. Then, from Remark 27, we can find N^(a)>N1\hat{N}(a)>N_{1}, with N1N_{1} provided by Lemma 49, such that

(𝑸(τ,a);τ)>c(a)2eτ>0for every τ<N^(a).{\mathcal{H}}({\boldsymbol{Q}}(\tau,a);\tau)>\frac{c(a)}{2}{\rm e}^{\ell\tau}>0\,\qquad\mbox{for every }\tau<-\hat{N}(a). (74)

Hence, from Lemma 49, we get 𝑸(τ,a)𝒞(a,τ){\boldsymbol{Q}}(\tau,a)\in\mathcal{C}(a,\tau) for any τ<N^(a)\tau<-\hat{N}(a).

Fix τ<N^(a)\tau<-\hat{N}(a); then by construction ϕ(t;τ,𝑸(τ,a))𝒞(a,t)\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}}(\tau,a))\in\mathcal{C}(a,t), when tt is in a sufficiently small right neighborhood of τ\tau, see Remark 13.

So, there is T(τ,a)>τT(\tau,a)>\tau such that ϕ(t;τ,𝑸(τ,a))𝒞(a,t)\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}}(\tau,a))\in\mathcal{C}(a,t) for any τtT(τ,a)\tau\leq t\leq T(\tau,a) and it leaves 𝒞(a,t)\mathcal{C}(a,t) when t>T(τ,a)t>T(\tau,a), or ϕ(t;τ,𝑸(τ,a))𝒞(a,t)\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}}(\tau,a))\in\mathcal{C}(a,t) for any tτt\geq\tau (i.e. T(τ,a)=+T(\tau,a)=+\infty).

From an analysis of the phase portrait we see that

x˙(t;τ,𝑸(τ,a))<0<y˙(t;τ,𝑸(τ,a))when τtT(τ,a).\dot{x}(t;\tau,{\boldsymbol{Q}}(\tau,a))<0<\dot{y}(t;\tau,{\boldsymbol{Q}}(\tau,a))\,\quad\textrm{when $\tau\leq t\leq T(\tau,a)$}\,. (75)

Assume first that T(τ,a)T(\tau,a)\in{\mathbb{R}}, then either ϕ(t;τ,𝑸(τ,a))\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}}(\tau,a)) crosses the yy negative semi-axis at t=T(τ,a)t=T(\tau,a) proving the thesis, or 𝑹:=ϕ(T(τ,a);τ,𝑸(τ,a))W~s(T(τ,a))\boldsymbol{R}:=\boldsymbol{\phi}(T(\tau,a);\tau,{\boldsymbol{Q}}(\tau,a))\in\tilde{W}^{s}(T(\tau,a)).

So, assume the latter and observe that 𝑹W^s(T(τ,a))\boldsymbol{R}\in\hat{W}^{s}(T(\tau,a)), thus 𝑸(τ,a)Ws(τ){\boldsymbol{Q}}(\tau,a)\in{W}^{s}(\tau). Then, x˙(t;T(τ,a),𝑹)<0\dot{x}(t;T(\tau,a),\boldsymbol{R})<0 for any t>T(τ,a)t>T(\tau,a), and by construction x˙(t;τ,𝑸(τ,a))<0\dot{x}(t;\tau,{\boldsymbol{Q}}(\tau,a))<0 for any τtT(τ,a)\tau\leq t\leq T(\tau,a). So, since the trajectories ϕ(;τ,𝑸(τ,a))\boldsymbol{\phi}(\,\cdot\,;\tau,{\boldsymbol{Q}}(\tau,a)) and ϕ(;T(τ,a),𝑹)\boldsymbol{\phi}(\,\cdot\,;T(\tau,a),\boldsymbol{R}) coincide, we conclude that x˙(t;τ,𝑸(τ,a))<0\dot{x}(t;\tau,{\boldsymbol{Q}}(\tau,a))<0 for any tτt\geq\tau, thus giving us 𝑸(τ,a)W^s(τ){\boldsymbol{Q}}(\tau,a)\in\hat{W}^{s}(\tau).

Hence, we get a contradiction comparing (74) with Lemma 49.

We consider now the remaining case: T(τ,a)=+T(\tau,a)=+\infty. Recalling (75), the only reasonable conclusion is that ϕ(t;τ,𝑸(τ,a))(0,0)\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}}(\tau,a))\to(0,0) as t+t\to+\infty, i.e. 𝑸(τ,a)Ws(τ){\boldsymbol{Q}}(\tau,a)\in{W}^{s}(\tau). Whence, from (75) we find 𝑸(τ,a)W^s(τ){\boldsymbol{Q}}(\tau,a)\in\hat{W}^{s}(\tau), and get again the contradiction comparing (74) with Lemma 49.

So, the lemma is proved. ∎

Taking advantage of this result, we can extend it to a wider class of functions.

Lemma 51.

Assume (𝐇)\boldsymbol{({\rm H}_{\ell})} with np1\ell\leq\frac{n}{p-1}, and that there is ρ1>0\rho_{1}>0 such that 𝒦(r)𝒦(ρ1){\mathcal{K}}(r)\geq{\mathcal{K}}(\rho_{1}) when rρ1r\geq\rho_{1}. Then, the same conclusion as in Lemma 50 holds.

Proof.

Recalling that K(t):=𝒦(et)K(t):={\mathcal{K}}({\rm e}^{t}), by assumption there is τ1\tau_{1}\in\mathbb{R} such that K(t)K(τ1)K(t)\geq K(\tau_{1}) when tτ1t\geq\tau_{1}. Let us define

Km(t):={K(t)if tτ1,K(τ1)if tτ1.K_{m}(t):=\begin{cases}K(t)&\textrm{if }t\leq\tau_{1}\,,\\ K(\tau_{1})&\textrm{if }t\geq\tau_{1}\,.\end{cases}

Let

ϕ𝒎(t;τ,𝑸(τ,a))=(xm(t;τ,𝑸(τ,a)),ym(t;τ,𝑸(τ,a)))\boldsymbol{\phi_{m}}(t;\tau,{\boldsymbol{Q}}(\tau,a))=\big{(}x_{m}(t;\tau,{\boldsymbol{Q}}(\tau,a)),y_{m}(t;\tau,{\boldsymbol{Q}}(\tau,a))\big{)}

be the trajectory of the modified system (11) where K(t)K(t) is replaced by Km(t)K_{m}(t). Observe that Km(t)K_{m}(t) satisfies (𝐖𝒔)\boldsymbol{({\rm W}_{s})} and we can still apply Remark 13, but Ws(τ)W^{s}(\tau) depends continuously and not smoothly on τ\tau.

Since the original system and the modified one coincide for tτ1t\leq\tau_{1}, then their unstable manifolds are the same in this interval.

Concerning the modified system, let K¯m>suptKm(t)=suptτ1K(t)\overline{K}_{m}>\sup_{t\in{\mathbb{R}}}K_{m}(t)=\sup_{t\leq\tau_{1}}K(t) and select the point 𝑮¯𝒎=(G¯x,m,G¯y,m)\boldsymbol{\overline{G}_{m}}=(\overline{G}_{x,m},\overline{G}_{y,m}) as above. Then, from Remark 27 we see that for any 0<a<(αqK¯m)p1qp0<a<\left(\frac{\alpha^{q}}{\overline{K}_{m}}\right)^{\frac{p-1}{q-p}}, cf. (73), there is N=N(a)>|τ1|>0N={N}(a)>|\tau_{1}|>0, such that Wu(τ)W^{u}(\tau) crosses transversely (a){\mathcal{L}}(a) in 𝑸(τ,a){\boldsymbol{Q}}(\tau,a) for any τ<N\tau<-{N}.

From Lemma 50 applied to the modified system, we can find N^(a)N(a)\hat{N}(a)\geq N(a) such that for any τ<N^(a)\tau<-\hat{N}(a) there are Tm(τ,a)T_{m}(\tau,a)\in{\mathbb{R}} and 𝑹𝒎=(0,Rm)\boldsymbol{R_{m}}=(0,R_{m}) such that ϕ𝒎(Tm(τ,a);τ,𝑸(τ,a))=𝑹𝒎\boldsymbol{\phi_{m}}(T_{m}(\tau,a);\tau,{\boldsymbol{Q}}(\tau,a))=\boldsymbol{R_{m}} and both xm(t;τ,𝑸(τ,a))>0{x}_{m}(t;\tau,{\boldsymbol{Q}}(\tau,a))>0 and x˙m(t;τ,𝑸(τ,a))<0<y˙m(t;τ,𝑸(τ,a))\dot{x}_{m}(t;\tau,{\boldsymbol{Q}}(\tau,a))<0<\dot{y}_{m}(t;\tau,{\boldsymbol{Q}}(\tau,a)) hold for any τtTm(τ,a)\tau\leq t\leq T_{m}(\tau,a), see (75).

Since ϕ𝒎(t;τ,𝑸(τ,a))ϕ(t;τ,𝑸(τ,a))\boldsymbol{\phi_{m}}(t;\tau,{\boldsymbol{Q}}(\tau,a))\equiv\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}}(\tau,a)) for any tτ1t\leq\tau_{1}, if Tm(τ,a)τ1T_{m}(\tau,a)\leq\tau_{1}, the thesis is achieved. So, we assume Tm(τ,a)>τ1T_{m}(\tau,a)>\tau_{1}; since KmK_{m} is constant then (ϕ𝒎(t;τ,𝑸(τ,a));τ1){\mathcal{H}}(\boldsymbol{\phi_{m}}(t;\tau,{\boldsymbol{Q}}(\tau,a));\tau_{1}) is constant when tτ1t\geq\tau_{1}, see (16), then

(ϕ𝒎(t;τ,𝑸(τ,a));τ1)(𝑹𝒎;τ1)=pp1|Rm|p/(p1)>0,{\mathcal{H}}(\boldsymbol{\phi_{m}}(t;\tau,{\boldsymbol{Q}}(\tau,a));\tau_{1})\equiv{\mathcal{H}}(\boldsymbol{R_{m}};\tau_{1})=\tfrac{p}{p-1}|R_{m}|^{p/(p-1)}>0\,,

for any t[τ1,Tm(τ,a)]t\in[\tau_{1},T_{m}(\tau,a)].

Let us set 𝑺=(Sx,Sy):=ϕ(τ1;τ,𝑸(τ,a))=ϕ𝒎(τ1;τ,𝑸(τ,a))\boldsymbol{S}=(S_{x},S_{y}):=\boldsymbol{\phi}(\tau_{1};\tau,{\boldsymbol{Q}}(\tau,a))=\boldsymbol{\phi_{m}}(\tau_{1};\tau,{\boldsymbol{Q}}(\tau,a)), then from the previous estimate we get (𝑺;τ1)=(𝑹𝒎;τ1)>0{\mathcal{H}}(\boldsymbol{S};\tau_{1})={\mathcal{H}}(\boldsymbol{R_{m}};\tau_{1})>0. Notice that we can rewrite the image of ϕ𝒎(t;τ,𝑸(τ,a))\boldsymbol{\phi_{m}}(t;\tau,{\boldsymbol{Q}}(\tau,a)) as a graph in xx, i.e. there is a decreasing smooth function fmf_{m} such that

:={ϕ𝒎(t;τ,𝑸(τ,a))τ1tTm(τ,a)}={(x,y)y=fm(x), 0xSx}.\begin{split}\mathcal{F}:=&\{\boldsymbol{\phi_{m}}(t;\tau,{\boldsymbol{Q}}(\tau,a))\mid\tau_{1}\leq t\leq T_{m}(\tau,a)\}=\{(x,y)\mid y\!=\!f_{m}(x),\;0\leq x\leq S_{x}\}.\end{split}

Let us define

T(τ,a)=sup{sτ1x˙(t;τ,𝑸(τ,a))<0<x(t;τ,𝑸(τ,a)) for any τ1ts};T(\tau,a)=\sup\{s\geq\tau_{1}\mid\dot{x}(t;\tau,{\boldsymbol{Q}}(\tau,a))<0<x(t;\tau,{\boldsymbol{Q}}(\tau,a))\,\textrm{ for any $\tau_{1}\leq t\leq s$}\};

from (16) we have

(ϕ(t;τ,𝑸(τ,a));τ1)(𝑺;τ1)=(𝑹𝒎;τ1)>0,{\mathcal{H}}(\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}}(\tau,a));\tau_{1})\geq{\mathcal{H}}(\boldsymbol{S};\tau_{1})={\mathcal{H}}(\boldsymbol{R_{m}};\tau_{1})>0\,,

for every τ1tT(τ,a)\tau_{1}\leq t\leq T(\tau,a). Therefore we see that the trajectory ϕ(t;τ,𝑸(τ,a))\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}}(\tau,a)) is forced to stay in the unbounded set

Λ={(x,y)0xSx,y<0,(x,y;τ1)(Sx,Sy;τ1)}={(x,y)0xSx;yfm(x)},\begin{split}\Lambda=&\{(x,y)\mid 0\leq x\leq S_{x},\;y<0,\;{\mathcal{H}}(x,y;\tau_{1})\geq{\mathcal{H}}(S_{x},S_{y};\tau_{1})\}\\ =&\{(x,y)\mid 0\leq x\leq S_{x}\,;y\leq f_{m}(x)\},\end{split}

whenever τ1tT(τ,a)\tau_{1}\leq t\leq T(\tau,a). Observe now that the maximum of x˙(x,y)=αx|y|1/(p1)\dot{x}(x,y)=\alpha x-|y|^{1/(p-1)} within Λ\Lambda is obtained in \mathcal{F} which is compact; further it has to be negative, so there is C>0C>0 such that x˙(t;τ,𝑸(τ,a))C\dot{x}(t;\tau,{\boldsymbol{Q}}(\tau,a))\leq-C when τ1tT(τ,a)\tau_{1}\leq t\leq T(\tau,a). Then it is easy to check that ϕ(t;τ,𝑸(τ,a))\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}}(\tau,a)) is bounded when τ1tT(τ,a)\tau_{1}\leq t\leq T(\tau,a). So, from elementary considerations, we see that ϕ(t;τ,𝑸(τ,a))\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}}(\tau,a)) crosses the yy negative semiaxis at t=T(τ,a)<τ1+Sx/Ct=T(\tau,a)<\tau_{1}+S_{x}/C.

The previous lemmas permit us to complete the proof of Lemma 45.

Proof of Lemma 45.

We fix aa as in (73), and consider the segment (a){\mathcal{L}}(a) defined as in (36). From Remark 27, there is N(a)>0N(a)>0 such that Wu(τ)W^{u}(\tau) intersects (a){\mathcal{L}}(a) in 𝑸(τ,a){\boldsymbol{Q}}(\tau,a) for every τ<N(a)\tau<-N(a).

Using Lemma 20, we consider the decreasing continuous function d(a):],N(a)[]D,+[d_{{\mathcal{L}}(a)}:{}]-\infty,-N(a)[{}\to{}]D,+\infty[{} such that the solution u(r,d(a)(τ))u(r,d_{{\mathcal{L}}(a)}(\tau)) of (4) corresponds to the trajectory ϕ(t;τ,𝑸(τ,a))\boldsymbol{\phi}(t;\tau,{\boldsymbol{Q}}(\tau,a)) of (11).

Then, either Lemma 50 or 51 applies providing the value N^(a)N(a)\hat{N}(a)\geq N(a) such that the solution u(r,d(a)(τ))u(r,d_{{\mathcal{L}}(a)}(\tau)) is a crossing solution for every τ<N^(a)\tau<-\hat{N}(a). Setting D^\hat{D} such that d(a)(],N^(a)[)=]D^,+[d_{{\mathcal{L}}(a)}({}]-\infty,-\hat{N}(a)[{})={}]\hat{D},+\infty[{}, we get ]D^,+[J{}]\hat{D},+\infty[{}\subset J. The lemma is thus proved. ∎

Proof of Proposition 46.

Take ε<1/α\varepsilon<1/\alpha such that a=a(ε)a=a(\varepsilon) defined in (57) satisfies the inequality (73). Define the segment (a){\mathcal{L}}(a) as in (36), and N^(a)\hat{N}(a), T(,a)T(\cdot,a) as in Lemma 50 or 51. Set D^\hat{D} as in the proof of Lemma 45.

Recalling Lemmas 20 and 45, we have τ(a)(]D^,+[)=],N^(a)[\tau_{{\mathcal{L}}(a)}({}]\hat{D},+\infty[{})={}]-\infty,-\hat{N}(a)[{} and limd+τ(a)(d)=\lim_{d\to+\infty}\tau_{{\mathcal{L}}(a)}(d)=-\infty. Moreover, the function T(,a)T(\cdot,a) is well defined in ],N^(a)[{}]-\infty,-\hat{N}(a)[{}, and it is continuous, cf. Remark 29.

Hence, we are able to apply Proposition 42 to infer (65).

5 Proof of the theorems

Since all the preliminaries are well-established, we conclude our paper by giving the explicit proof of our main theorems.

Proof of Theorem 1.

If <p\ell<\ell_{p}^{*}, from Proposition 37 there is D^>0\hat{D}>0 such that ]D^,+[J{}]\hat{D},+\infty[{}\subset J and limd+R(d)=0\lim_{d\to+\infty}R(d)=0. So, there is D~0\tilde{D}\geq 0 such that D~J\tilde{D}\not\in J and ]D~,+[J]\tilde{D},+\infty[\subset J; whence from Propositions 23 and 24 we have limdD~+R(d)=+\lim_{d\to\tilde{D}^{+}}R(d)=+\infty. Further, since RR is continuous in JJ, see Proposition 22, we find R(J)=]0,+[R(J)={}]0,+\infty[{}: i.e. for every >0{\mathcal{R}}>0 there is dJd\in J such that R(d)=R(d)={\mathcal{R}}, which amounts to say that u(r;d)u(r;d) solves problem (3).

The second assertion follows immediately from Proposition 44. ∎

Proof of Theorem 2.

The proof takes advantage of Propositions 23 and 43.

If =p\ell=\ell_{p}^{*}, we distinguish two alternatives: either R(]0,+[)=[0,+[R({}]0,+\infty[{})=[{\mathcal{R}}_{0},+\infty[{} where 0{\mathcal{R}}_{0} is an internal minimum of RR or R(]0,+[)=]0,+[R({}]0,+\infty[{})={}]{\mathcal{R}}_{0},+\infty[{} where 0=lim infd+R(d)>0{\mathcal{R}}_{0}=\liminf_{d\to+\infty}R(d)>0. Then, the proof is concluded.

On the other hand, if >p\ell>\ell_{p}^{*}, from limd0R(d)=limd+R(d)=+\lim_{d\to 0}R(d)=\lim_{d\to+\infty}R(d)=+\infty, we deduce that the function RR has an internal minimum 0{\mathcal{R}}_{0}, and the pre-image R1()R^{-1}({\mathcal{R}}) has at least two elements for every >0{\mathcal{R}}>{\mathcal{R}}_{0}, thus giving the multiplicity result. ∎

Remark 52.

We want to underline that in the critical case =p\ell=\ell_{p}^{*} we are not able to discern which of the alternatives analyzed in the proof of Theorem 2 holds, and, consequently, we are not able to say if there is a solution for =0{\mathcal{R}}={\mathcal{R}}_{0}.

Proof of Theorem 3.

From Proposition 44 and Lemma 45, the set JJ contains a nontrivial interval, and there exists 0:=infdJR(d)>0{\mathcal{R}}_{0}:=\inf_{d\in J}R(d)>0. Then, the proof follows the lines of the one of Theorem 2, profiting from Propositions 2223, 24 and 46. ∎

Acknowledgements

Francesca Dalbono was partially supported by the PRIN Project 2017JPCAPN “Qualitative and quantitative aspects of nonlinear PDEs” and by FFR 2022-2023 from University of Palermo.

All the authors are members of INdAM-GNAMPA.

References

  • [1] Bamón R., Flores I., del Pino M., Ground states of semilinear elliptic equations: a geometric approach, Ann. Inst. H. Poincaré Anal. Non Linéaire, 17 (2000), 551-581.
  • [2] Battelli F., Johnson R., Singular ground states for the scalar curvature equation in n{\mathbb{R}}^{n}, Differential Integral Equations, 14 (2001), 141-158.
  • [3] Bianchi G., Egnell H., An ODE approach to the equation Δu+Kun+2n2=0\Delta u+Ku^{\frac{n+2}{n-2}}=0, in n{\mathbb{R}}^{n}, Math. Z., 210 (1992), 137-166.
  • [4] Bidaut-Véron M.F., Local and global behavior of solutions of quasilinear equations of Emden-Fowler type, Arch. Rational Mech. Anal., 107 (1989), 293-324.
  • [5] H. Brezis, L. Nirenberg, Positive solutions of nonlinear elliptic equations involving critical Sobolev exponents, Comm. Pure Appl. Math., 36 (4) (1983), 437-477
  • [6] C. Budd, J. Norbury, Semilinear elliptic equation and supercritical growth, J. Differential Equations 68, (1987) n. 2, 169-167.
  • [7] Chen C.C., Lin C.S., Estimates of the conformal scalar curvature equation via the method of moving planes, Comm. Pure Appl. Math., 50 (1997), 971-1017.
  • [8] Chen C.C., Lin C.S. Blowing up with infinite energy of conformal metrics on 𝕊n\mathbb{S}^{n}, Comm. Partial Differential Equations, 24 (1999), 785-799.
  • [9] Cheng K.S., Chern J.L., Existence of positive entire solutions of some semilinear elliptic equations, J. Differerential Equations, 98 (1992), 169–180.
  • [10] Chtioui H., Hajaiej H., Soula M., The scalar curvature problem on four-dimensional manifolds, Commun. Pure Appl. Anal., 19 (2020), 723-746.
  • [11] Coddington E., Levinson N., Theory of Ordinary Differential Equations. Mc Graw Hill, New York, 1955.
  • [12] Dalbono F., Franca M., Nodal solutions for supercritical Laplace equations, Commun. Math. Phys., 347 (2016), 875-901.
  • [13] Deimling K., Nonlinear functional analysis. Springer, Berlin, 1985.
  • [14] Ding, W.Y., Ni, W.M., On the elliptic equation Δu+Kun+2n2=0\Delta u+Ku^{\frac{n+2}{n-2}}=0 and related topics, Duke Math. J., 52 (1985), 485-506.
  • [15] J. Dolbeault, I. Flores, Geometry of Phase space and solutions of semilinear elliptic equations in a ball, Trans. Am. Math. Soc. 359, (2007) n. 9, 4073-4087
  • [16] (MR3373577) [10.1016/j.na.2015.04.015] I. Flores and M. Franca, Phase plane analysis for radial solutions to supercritical quasilinear elliptic equations in a ball, Nonlinear Anal., 125 (2015), 128-149.
  • [17] Fowler R.H., Further studies of Emden’s and similar differential equations, Q. J. Math., 2 (1931), 259-288.
  • [18] Franca M., Non-autonomous quasilinear elliptic equations and Ważewski’s principle, Topol. Methods Nonlinear Anal., 23, (2004), 213-238.
  • [19] Franca M., Corrigendum and addendum to “Non-autonomous quasilinear elliptic equations and Ważewski’s principle”, Topol. Methods Nonlinear Anal., 56 (2020), 1-30.
  • [20] Franca M., Classification of positive solution of pp-Laplace equation with a growth term, Arch. Math. (Brno), 40, (2004), 415-434.
  • [21] Franca M., Fowler transformation and radial solutions for quasilinear elliptic equations. I. The subcritical and the supercritical case, Can. Appl. Math. Q. 16 (2008), 123-159.
  • [22] Franca M., Structure theorems for positive radial solutions of the generalized scalar curvature equation, Funkcial. Ekvac., 52 (2009), 343-369.
  • [23] Franca M., Johnson R., Ground states and singular ground states for quasilinear partial differential equations with critical exponent in the perturbative case, Adv. Nonlinear Stud., 4 (2004), 93-120.
  • [24] Franca M., Sfecci A., Entire solutions of superlinear problems with indefinite weights and Hardy potentials, J. Dynam. Differential Equations, 30 (2018), 1081-1118.
  • [25] Franchi B., Lanconelli E., Serrin J., Existence and uniqueness of nonnegative solutions of quasilinear equations in n{\mathbb{R}}^{n}, Adv. Math., 118 (1996) 177-243.
  • [26] García-Huidobro M., Manásevich R, Yarur C.S., On the structure of positive radial solutions to an equation containing a pp-Laplacian with weight, J. Differential Equations, 223 (2006), 51-95.
  • [27] Gazzola F., Critical exponents which relate embedding inequalities with quasilinear elliptic problems, Discrete Contin. Dyn. Syst., suppl. (2003), 327-335.
  • [28] Z. Guo, J. Wei, Global solution branch and Morse index estimates of a semilinear elliptic equation with super-critical exponent, Trans. Amer. Math. Soc., 363, (2011) n. 9, 4777-4799.
  • [29] Johnson R., Concerning a theorem of Sell, J. Differential Equations, 30 (1978), 324-339.
  • [30] Johnson R., Pan X.B., Yi Y.F., Singular ground states of semilinear elliptic equations via invariant manifold theory, Nonlinear Anal., 20 (1993), 1279-1302.
  • [31] Johnson R., Pan X.B., Yi Y.F., The Melnikov method and elliptic equation with critical exponent, Indiana Univ. Math. J., 43 (1994), 1045-1077.
  • [32] Jones C., Küpper T., On the infinitely many solutions of a semilinear elliptic equation, SIAM J. Math. Anal., 17 (1986), 803-835.
  • [33] Kabeya Y., Yanagida E., Yotsutani S., Existence of nodal fast-decay solutions to div(|u|m2u)+K(|x|)|u|q1u=0{\rm div}(|\nabla u|^{m-2}\nabla u)+K(|x|)|u|^{q-1}u=0 in n\mathbb{R}^{n}, Differential Integral Equations, 9 (1996), 981-1004.
  • [34] Kawano N., Ni W.M., Yotsutani S., A generalized Pohozaev identity and its applications, J. Math. Soc. Japan, 42 (1990), 541-564.
  • [35] Kawano N., Satsuma J., Yotsutani S., Existence of positive entire solutions of an Emden-type elliptic equation, Funkcial. Ekvac., 31 (1988), 121–145.
  • [36] Kusano T., Naito M., Oscillation theory of entire solutions of second order superlinear elliptic equations, Funkcial. Ekvac., 30 (1987), 269-282.
  • [37] Li Y.Y., Prescribing scalar curvature on 𝕊n\mathbb{S}^{n} and related problems, Part II: Existence and compactness, Comm. Pure Appl. Math., 49 (1996), 541-597.
  • [38] Li Y.Y., Nguyen L., Wang B., The axisymmetric σk\sigma_{k}-Nirenberg problem, J. Funct. Anal., 281 (2021), Paper N. 109198, 60 pp.
  • [39] Lin C.S., Lin S.S., Positive radial solutions for Δu+Kun+2n2=0\Delta u+Ku^{\frac{n+2}{n-2}}=0 in n{\mathbb{R}}^{n} and related topics, Appl. Anal., 38 (1990), 121-159.
  • [40] F. Merle, L.A. Peletier, Positive solutions of elliptic equations involving supercritical growth, Proc. Roy. Soc. Edinburgh sect A, 118, (1991), 49-62.
  • [41] Ni W.M., On the elliptic equation Δu+K(x)un+2n2=0\Delta u+K(x)u^{\frac{n+2}{n-2}}=0, its generalizations, and applications in geometry, Indiana Univ. Math. J., 31 (1982), 493-529.
  • [42] Ni W.M., Serrin J., Nonexistence theorems for quasilinear partial differential equations, Rend. Circ. Mat. Palermo (2), 8 (1985), 171-185.
  • [43] Sharaf K., A note on a second order PDE with critical nonlinearity, Electron. J. Qual. Theory Differ. Equ., 10 (2019), 16 pp.
  • [44] Yanagida E., Yotsutani S., Classification of the structure of positive radial solutions to Δu+K(|x|)up=0\Delta u+K(|x|)u^{p}=0 in n\mathbb{R}^{n}, Arch. Rational Mech. Anal., 124 (1993), 239-259.
  • [45] Yanagida E., Yotsutani S., Existence of nodal fast-decay solutions to Δu+K(|x|)|u|p1u=0\Delta u+K(|x|)|u|^{p-1}u=0 in n\mathbb{R}^{n}, Nonlinear Anal., 22 (1994), 1005-1015.