In this work, we study the existence of one-sign solutions without signum condition for the following problem:
{−Δu=λa(x)f(u),x∈RN, u(x)→0,as|x|→+∞,
where N≥3, λ is a real parameter and a∈Cαloc(RN,R) for some α∈(0,1) is a weighted function, f∈Cα(R,R), and there exist two constants s2<0<s1, such that f(s1)=f(s2)=f(0)=0 and sf(s)>0 for s∈R∖{s1,0,s2}. Furthermore, we consider the exact multiplicity of one-sign solutions for above problem under more strict hypotheses. We use bifurcation techniques and the approximation of connected components to prove our main results.
Citation: Wenguo Shen. Bifurcation and one-sign solutions for semilinear elliptic problems in RN[J]. AIMS Mathematics, 2023, 8(5): 10453-10467. doi: 10.3934/math.2023530
[1] | Ruyun Ma, Dongliang Yan, Liping Wei . Global bifurcation of sign-changing radial solutions of elliptic equations of order 2m in annular domains. AIMS Mathematics, 2020, 5(5): 4909-4916. doi: 10.3934/math.2020313 |
[2] | Wenguo Shen . Unilateral global interval bifurcation and one-sign solutions for Kirchhoff type problems. AIMS Mathematics, 2024, 9(7): 19546-19556. doi: 10.3934/math.2024953 |
[3] | Changmu Chu, Yuxia Xiao, Yanling Xie . Infinitely many sign-changing solutions for a semilinear elliptic equation with variable exponent. AIMS Mathematics, 2021, 6(6): 5720-5736. doi: 10.3934/math.2021337 |
[4] | Haohao Jia, Feiyao Ma, Weifeng Wo . Large positive solutions to an elliptic system of competitive type with nonhomogeneous terms. AIMS Mathematics, 2021, 6(8): 8191-8204. doi: 10.3934/math.2021474 |
[5] | H. Al Jebawy, H. Ibrahim, Z. Salloum . On oscillating radial solutions for non-autonomous semilinear elliptic equations. AIMS Mathematics, 2024, 9(6): 15190-15201. doi: 10.3934/math.2024737 |
[6] | Changling Xu, Hongbo Chen . A two-grid P20-P1 mixed finite element scheme for semilinear elliptic optimal control problems. AIMS Mathematics, 2022, 7(4): 6153-6172. doi: 10.3934/math.2022342 |
[7] | Chunjuan Hou, Zuliang Lu, Xuejiao Chen, Xiankui Wu, Fei Cai . Superconvergence for optimal control problems governed by semilinear parabolic equations. AIMS Mathematics, 2022, 7(5): 9405-9423. doi: 10.3934/math.2022522 |
[8] | Liangying Miao, Zhiqian He . Reversed S -shaped connected component for second-order periodic boundary value problem with sign-changing weight. AIMS Mathematics, 2020, 5(6): 5884-5892. doi: 10.3934/math.2020376 |
[9] | Diane Denny . Existence of a solution to a semilinear elliptic equation. AIMS Mathematics, 2016, 1(3): 208-211. doi: 10.3934/Math.2016.3.208 |
[10] | Andrey Muravnik . Nonclassical dynamical behavior of solutions of partial differential-difference equations. AIMS Mathematics, 2025, 10(1): 1842-1858. doi: 10.3934/math.2025085 |
In this work, we study the existence of one-sign solutions without signum condition for the following problem:
{−Δu=λa(x)f(u),x∈RN, u(x)→0,as|x|→+∞,
where N≥3, λ is a real parameter and a∈Cαloc(RN,R) for some α∈(0,1) is a weighted function, f∈Cα(R,R), and there exist two constants s2<0<s1, such that f(s1)=f(s2)=f(0)=0 and sf(s)>0 for s∈R∖{s1,0,s2}. Furthermore, we consider the exact multiplicity of one-sign solutions for above problem under more strict hypotheses. We use bifurcation techniques and the approximation of connected components to prove our main results.
Consider the following semilinear elliptic problem
{−Δu=λa(x)f(u),x∈RN,u(x)→0,as|x|→+∞, | (1.1) |
where λ is a real parameter, N≥3, and a∈Cαloc(RN,R) for some α∈(0,1) is a weighted function which can be sign-changing and f∈Cα(R,R), and f(s)s>0 for any s≠0. Edelson and Rumbos [1] have shown that problem (1.1) with f(u)≡u has a positive, simple and principal eigenvalue λ1 and the positive principle eigenfunction ϕ satisfies the asymptotic decay law lim|x|→+∞|x|N−2ϕ1(x)=c for some constant c(where a satisfied the following condition (A1)). Edelson and Rumbos [1,2] have also studied the existence of positive solution and the existence of global branches of minimal solutions of the problem (1.1) by the Schauder-Tychonoff fixed point theorem and the Dancer global bifurcation theorems [3]. By using the Rabinowitz global bifurcation method [4], Edelson and Furi [5] have shown the existence of positive minimal solution of the problem (1.1). In 2017, Dai [6] have established a global bifurcation result for the problem (1.1).
By [6], set
M(Ω):={a∈Cαloc(Ω,R):{x∈Ω:a(x)>0}≠∅}. |
For any u∈C∞c(Ω) with Ω⊆RN, we define
‖u‖1=(∫Ω|∇u|2dx)1/2. |
Denote by D1,2(Ω) the completion of C∞c(Ω) with respect to the norm ‖u‖1. Denote by S(RN) the set of all measurable real functions defined on RN. Two functions in S(RN) are considered as the same element of S(RN) when they are equal almost everywhere. Let L2(RN;|a|)=:{u∈S(RN):∫RN|a|u2dx<+∞}. For u∈D1,2(RN),u≠0, define the Rayleigh quotient
R(u)=∫RN|∇u|2dx∫RNau2dx. |
Dai et al. [6] also assumed that a satisfied the following condition:
(A1) Let a∈M(RN). Assume that p,P∈C(RN,R) are positive, radially symmetric and satisfies
0<p(|x|)≤a(x)≤P(|x|),∀x∈RN |
and
∫RN|x|2−NP(|x|)dx<+∞. |
Furthermore, if P satisfies the following more strong condition (with r=|x|)
∫∞0rN−1P(r)dr<+∞. | (1.2) |
Dai [5] established the following spectrum structure:
Lemma 1.1 (see [6, Theorem 1.1]). Let (A1) hold. Then there exists an orthonormal basis {φk}+∞1 of L2(RN;|a|) and a sequence of positive real numbers {λk}+∞1with λk→+∞ as k→+∞, such that
0<λ1<λ2≤⋅⋅⋅≤λk≤⋅⋅⋅, |
{−Δφk=λa(x)φk,x∈RN,φk∈D1,2(RN)∩C2,αloc(RN),φk(x)→0,as|x|→+∞. | (1.3) |
Moreover, one has that
λ1=minu≠0,u∈D1,2(RN)R(u)=R(φ1) |
and
λk=R(φk)=maxu≠0,u∈span{φ1,⋅⋅⋅,φk−1}R(u)=minu≠0,u⊥{φ1,⋅⋅⋅,φk−1}R(u)=mindimW=k,W⊂D1,2(RN)maxu∈WR(u). |
For k≥2. λ1 is simple and principal eigenvalue. Furthermore, if P satisfies (1.2), λ1 is the unique positive principal eigenvalue and
lim|x|→+∞|x|N−2φ1(x)=c |
for some constant c.
By the Rabinowitz global bifurcation theorem [2, Theorem 1.3] and the Dancer unilateral global bifurcation theorem [7, Theorem 2], Dai [6] obtained [6, Theorem 1.3]. The signum condition f(s)s>0 for s≠0 plays an important role in the [6, Theorem 1.3].
Of course, the natural question is that of what would happen without signum condition for the problem (1.1). Recently, Dai [8,9] studied the global behavior of the components of positive solutions for the Schr¨odinger equation and one-sign solutions for the p-Laplacian without the signum condition, respectively.
Motivated by the above interesting and important studies, we shall show that the branches bifurcating from infinity and the trivial solution line for the problem (1.1) are disjoint and the existence results of radial nodal solutions to problem (1.1) without signum condition.
We now present the following assumptions on f :
(A2) f∈C(R,R), and there exist two constants s2<0<s1, such that f(s1)=f(s2)=f(0)=0 and sf(s)>0 for s∈R∖{s1,0,s2}.
(A3) There exist two constants γ1>0 and γ2<0 such that
lims→s−1f(s)s1−s=γ1,lims→s+2f(s)s−s2=γ2. |
Let
f0=lim|s|→0f(s)s,f∞=lim|s|→+∞f(s)s. |
The main results of this section are the following interesting results:
Theorem 1.1. Let (A1)–(A3) hold.
(a) If f0,f∞∈(0,∞), then the problem (1.1) has at least two solutions u+∞ and u−∞ for λ∈(min{λ1f0,λ1f∞},max{λ1f0,λ1f∞}] such that u+∞ is positive in RN and u−∞ is negative in RN; the problem (1.1) has at least four solutions u+∞ and u−∞, u+0 and u−0 for λ∈(max{λ1f0,λ1f∞},∞) such that u+∞, u+0 are positive in RN and u−∞, u−0 are negative in RN. Moreover, if P satisfies (1.2), we have that
lim|x|→+∞|x|N−2u+i(x)=ci1 |
for all x∈RN and some constants ci1≠0, where i=0,∞. Do the same for u−i.
(b) If f0∈(0,∞) and f∞=∞, then the problem (1.1) has at least two solutions u+∞ and u−∞ for λ∈(0,λ1f0] such that u+∞ is positive in RN and u−∞ is negative in RN; the problem (1.1) has at least four solutions u+∞ and u−∞, u+0 and u−0 for λ∈(λ1f0,∞) such that u+∞, u+0 are positive in RN and u−∞, u−0 are negative in RN. Moreover, if P satisfies (1.2), we have that
lim|x|→+∞|x|N−2u+i(x)=ci1 |
for all x∈RN and some constants ci1≠0, where i=0,∞. Do the same for u−i.
(c) If f0=∞ and f∞∈(0,∞), then the problem (1.1) has at least two solutions u+∞ and u−∞ for λ∈(0,λ1f∞] such that u+∞ is positive in RN and u−∞ is negative in RN; the problem (1.1) has at least four solutions u+∞ and u−∞, u+0 and u−0 for λ∈(λ1f∞,+∞) such that u+∞, u+0 are positive in RN and u−∞, u−0 are negative in RN. Moreover, if P satisfies (1.2), we have
lim|x|→+∞|x|N−2u+i(x)=ci1 |
for all x∈RN and some constants ci1≠0, where i=0,∞. Do the same for u−i.
(d) If f0=∞ and f∞=∞, then the problem (1.1) has at least four solutions u+∞ and u−∞, u+0 and u−0 for λ∈(0,∞) such that u+∞, u+0 are positive in RN and u−∞, u−0 are negative in RN. Moreover, if P satisfies (1.2), we have that
lim|x|→+∞|x|N−2u+i(x)=ci1 |
for all x∈RN and some constants ci1≠0, where i=0,∞. Do the same for u−i.
Furthermore, we can get the exact multiplicity of one-sign solutions for problem (1.1) under more strict hypotheses.
(A4) f(s)≡0 for any s∉[s2,s1], f(s) is C1 with respect to s∈[s2,s1], and such that f(s)/s is decreasing in [0,s1] and is increasing in [s2,0].
The following are the main results of this section.
Theorem 1.2. Let (A1), (A2) and (A4) hold. Assume that f0∈(0,∞). Then,
(ⅰ) the problem (1.1) has exactly two solutions u+(λ,⋅) and u−(λ,⋅) for λ∈(λ1f0,+∞), such that 0<u+(λ,⋅)≤s1 and s2≤u−(λ,⋅)<0 in RN;
(ⅱ) all one-sign solutions of problem (1.1) lie on two smooth curves
Cν={(λ,u±(λ,⋅)):λ∈(λ1f0,+∞)}, |
C+ and C− join at (λ1/f0,0);
(ⅲ) u+(λ,⋅)(u−(λ,⋅)) is increasing (decreasing) with respect to λ.
(ⅳ) If P satisfies (1.2), we have that
lim|x|→+∞|x|N−2u+1(x)=c1andlim|x|→+∞|x|N−2u−1(x)=c2 |
for all x∈RN and some constants c1,c2≠0.
Theorem 1.3. Let (A1), (A2) and (A4) hold. Assume that f0=∞. Then,
(ⅰ) the problem (1.1) has exactly two solutions u+(λ,⋅) and u−(λ,⋅) for λ∈(0,+∞), such that 0<u+(λ,⋅)≤s1 and s2≤u−(λ,⋅)<0 in RN;
(ⅱ) all one-sign solutions of problem (1.1) lie on two smooth curves
Cν={(λ,u±(λ,⋅)):λ∈(λ1f0,+∞)}, |
C+ and C− join at (λ1/f0,0);
(ⅲ) u+(λ,⋅)(u−(λ,⋅)) is increasing (decreasing) with respect to λ.
(ⅳ) If P satisfies (1.2), we have that
lim|x|→+∞|x|N−2u+1(x)=c1andlim|x|→+∞|x|N−2u−1(x)=c2 |
for all x∈RN and some constants c1,c2≠0.
The rest of this paper is arranged as follows: Section 2, provides some preliminaries. In Section 3, we prove Theorem 1.1, which considers the existence of one-sign solutions for the problem (1.1) without signum condition. In Section 4, we consider exact multiplicity of one-sign solutions for the problem (1.1) and give the proof of Theorems 1.2 and 1.3.
Let
E={u∈C(RN,R):supx∈RN|u(x)|<+∞} |
with the norm
‖u‖=supx∈RN|u(x)|,forallu∈E. |
Clearly, E is a Banach space. Let P+={u∈E|u>0,forallx∈RN} and set P−=−P+ and P=P+∪P−.
By [6], we can show that u is a one-sign C2+α solution of problem (1.1) if and only if u is a solution of the operator equation
u=L(f(u))=λ∫RNΓN(x−y)a(y)f(u(y))dy, | (2.1) |
where ΓN(x−y)=1N(N−2)ωN|x−y|2−N, ωN being the volume of the unit ball in RN. Dai [6] also can show that L:E→E is completely continuous.
Consider the following problem
{−Δu=λa(x)u(x)+g(λ,x,u),x∈RN,u(x)→0,as|x|→+∞. | (2.2) |
Suppose g∈C(RN×E×R,E) satisfies
lim|s|→0g(x,s,λ)s=0 | (2.3) |
uniformly for x∈RN and λ on bounded sets.
Similar the proof of [6, Theorem 1.3], we can obtain that the following result:
Theorem 2.1. Assume that (A1) and (2.3) hold. The pair (λ1,0) is a bifurcation point of the problem (2.2) and there are two distinct unbounded continuum C+ and C− in R×E of solutions of the problem (2.2) emanating from (λ1,0). Moreover, we have
Cν⊂((R×Pν)∪{(λ1,0)}), |
where ν∈{+,−}.
By (2.1), Eq (2.2) is equivalent to
u=λLu+H(λ,u)=G(λ,u), |
where H(λ,u)=o(‖u‖) at u=0 uniformly on bounded λ intervals, H(λ,u)=L[g(x,u,λ)]. G(λ,u):R×E→E is completely continuous.
In order to prove Theorem 1.1, we also need to establish the unilateral global bifurcation result of the problem (2.2) from infinity under assumption
lims→+∞g(x,s,λ)s=0 | (2.4) |
uniformly for x∈RN and λ on bounded sets.
By Rabinowitz [10], we have the following theorem.
Theorem 2.2. Let (A1) and (2.4) hold. There exists a connected component Dν of of solutions of the problem (2.2), containing λ1×{∞}. Moreover, if Λ⊂R is an interval such that Λ∖{λ1} doesn't contain any other eigenvalue of problem (1.3), and M is a neighborhood of λ1×{∞} whose projection on R lies in Λ and whose projection on E is bounded away from 0, then either
1o Dν−M is bounded in R×E in which case Dν−M meets R={(λ,0)|λ∈R} or
2o Dν−M is unbounded.
If 2o occurs and Dν−M has a bounded projection on R, then Dν−M meets λj×{∞} for some j≠1, and ν∈{+,−}.
To prove our main results, we need the following results:
Lemma 2.1. (see [9]) Let X be a normal space and let {Cn|n=1,2,...} be a sequence of unbounded connected subsets of X. Assume that:
(i) there exists z∗∈lim infn→+∞Cn with ‖z∗‖<+∞;
(ii) for every R>0, (∪+∞n=1Cn)∩¯BR is a relatively compact set of X, where
BR={x∈X|‖x‖≤R}. |
Then, D:=lim supn→∞Cn is unbounded, closed and connected.
In order to treat the problems with non-asymptotic nonlinearity at ∞, we shall need the following lemmas.
Lemma 2.2. (see [9]) Let (X,ρ) be a metric space. If {Ci}i∈N is a sequence of sets whose limit superior is L and there exists a homeomorphism T:X→X such that for every R>0, (∪+∞i=1T(Ci))∩¯BR is a relatively compact set, then for each ϵ>0 there exists an m such that for every n>m,Cn⊂Vϵ(L), where Vϵ(L) denotes the set of all points p with ρ(p,x)<ϵ for any x∈L.
Now, in order to study the exact multiplicity of one-sign solutions for (1.1), let E=R×E, Φ(λ,u)=u−G(λ,u) and
S=¯{(λ,u)∈E:Φ(λ,u)=0,u≠0}R×E. |
For λ∈R and 0<s<+∞, define an open neighborhood of (λ1,0) in E as follows:
Bs(λ1,0)={(λ,u)∈E:‖u‖+|λ−λ1|<s}. |
Let E0 be a closed subset of E satisfying E=span{ψ1}⊕E0, where ψ1 is an eigenfunction corresponding to λ1 with ‖ψ1‖=1. According to the Hahn-Banach theorem, we have l∈E∗ satisfying
l(ψ1)=1andE0={u∈E:l(u)=0}, |
where E∗ denotes the dual space of E. For any 0<ε<+∞ and 0<η<1, define
Kε,η={(λ,u)∈E:|λ−λ1|<ε,|l(u)|>η‖u‖}. |
Obviously, Kε,η is an open subset of E, Kε,η=K+ε,η∪K−ε,η, with K+ε,η={(λ,u)∈E:|λ−λ1|<ε,l(u)>η‖u‖}, K−ε,η=−K+ε,η, which are disjoint and open in E.
Similar to that of [11, Lemma 6.4.1], we can show the following lemma.
Lemma 2.3. Let η∈(0,1), there is δ0>0 such that for each δ:0<δ<δ0, it holds that
((S∖{(λ1,0)})∩Bδ(λ1,0))⊆Kε,η. |
And there exist s∈R and a unique y∈E0 such that
v=sψ1+yand|s|>η‖v‖ |
for each (λ,v)∈((S∖{(λ1,0)})∩Bδ(λ1,0)). Further, λ=λ1+o(1) and y=o(s) as s→0 for these solutions (λ,v).
Remark 2.1. From Lemma 2.3, we can see that D=D+∪D− near (λ1,0) is given by a curve (λ(s),u(s))=(λ1+o(1),sψ1+o(s)) for s near 0. Moreover, we can distinguish between two portions of this curve by s≥0 and s≤0.
When f0∈(0,∞), let ζ(u)∈C(R,R) be such that
f(u)=f0u+ζ(u) |
with
lim|u|→0ζ(u)u=0. |
Let us consider
{−Δu=λa(x)f0u(x)+λa(x)ζ(u),x∈RN,u(x)→0,as|x|→+∞ | (3.1) |
as a bifurcation problem from the trivial solution u≡0.
Applying Theorem 2.1 to (3.1), we have the following result.
Remark 3.1. There is an unbounded continuum Cν of solutions of the problem (1.1) emanating from (λ1f0,0), such that Cν⊂((R×Pν)∪{(λ1f0,0)}), where ν∈{+,−}.
We now analyze the global behavior of C++ and C−.
Lemma 3.1. Let (A1)–(A3) hold. Then
(ⅰ) for (λ,u)∈(C+∪(C−), we have that s2<u(x)<s1 for all x∈RN;
(ⅱ) for (λ,u)∈(D+∪D−), we have that either supx∈RNu(x)>s1 or infx∈RNu(x)<s2.
Proof. Suppose on the contrary that there exists (λ,u)∈(C+∪C−∪D+∪D−) such that either supx∈RNu(x)=s1 or infx∈RNu(x)=s2.
We only treat the case of supx∈RNu(x)=s1 because the proof for the case of infx∈RNu(x)=s2 can be given similarly.
We claim that there exists 0<m<+∞ such that f(s)≤m(s1−s) for any s∈[0,s1]. Clearly, the claim is true for the case of s=0 or s=s1 by virtue of (A2).
For any ϵ∈(0,γ1), it follows from (A3) that there exists δ>0, such that
f(s)<(γ1+ϵ)(s1−s) |
for any s∈(s1−δ,s1). From (A2), it arrives
maxs∈[0,s1−δ]f(s)s1−s=ρ>0. |
So, the claim is verified by choosing m=max{ρ,γ1+ϵ}.
Now, let us consider the following problem
{−Δ(s1−u)+λa(x)m(s1−u)=λa(x)[m(s1−u)−f(u)],x∈RN,u(x)→0,as|x|→+∞. |
It is obvious that f(s)≤m(s1−s) for any s∈[0,s1] implies
{−Δ(s1−u)+λa(x)m(s1−u)≥0,x∈RN,u(x)→0,as|x|→+∞. |
The strong maximum principle of [12] implies that s1>u in RN. This is a contradiction.
Lemma 3.2. Let (A1)–(A3) hold. Then
(λ1f0,∞)⊆ProjR(C+),(λ1f0,∞)⊆ProjR(C−). |
Proof. We show that the projection of C+ on R is unbounded. It is sufficient to show that the set {(λ,u)∈C+|λ∈[0,d]} is bounded for any fixed d∈(0,+∞). Suppose on the contrary that there exists (λn,un)∈C+,n∈N, such that λn→μ≤d, ‖un‖→+∞ as n→+∞. Let vn=un/‖un‖. Then, vn should be the solutions of problem
{−Δvn=λa(x)f(un)‖un‖,x∈RN,vn(x)→0,as|x|→+∞. | (3.2) |
By Lemma 3.1 (ⅰ), we have that
f(un)≤maxs∈[0,s1]|f(s)|. |
By (A2), one can obtain that f(s)∈C([0,s1]).
Thus, we can show
limn→+∞f(un)‖un‖=0. | (3.3) |
By the compactness of L(.), it follows from (3.2) that vn→v0≡0 as n→+∞. This contradicts ‖v0‖=1.
This together with the fact that C+ joins (λ1f0,0) to infinity yields that
(λ1f0,∞)⊆ProjR(C+). |
Similarly, we can show that
(λ1f0,∞)⊆ProjR(C−). |
In the following we will investigate the other one-sign solutions of problem (1.1).
When f∞∈(0,∞), let ξ(u)∈C(R,R) be such that
f(u)=f∞u+ξ(u) |
with
lim|u|→∞ξ(u)u=0. |
Let us consider
{−Δu=λa(x)f∞u(x)+λa(x)ξ(u),x∈RN,u(x)→0,as|x|→+∞ | (3.4) |
as a bifurcation problem from infinity. We add the points {(λ,∞)|λ∈R} to space R×E.
Applying Theorem 2.2 to (3.4), we have the following result.
Remark 3.2. There exists an unbounded continua Dν of solutions of (1.1), emanating from (λ1f∞,∞), such that Dν⊂((R×Pν)∪{(λ1f∞,∞)}), where ν∈{+,−}.
We now analyze the global behavior of D++ and D−.
Lemma 3.3. Let (A1)–(A3) hold. Then
(λ1f∞,∞)⊆ProjR(D+),(λ1f∞,∞)⊆ProjR(D−). |
Proof. We show that Dν−M has an unbounded projection on R.
Applying Theorem 2.2 to (3.4), one can obtain that (10) of Theorem 2.2 does not occur by Lemma 3.1 (ii). So Dν−M is unbounded.
Now, we show that the case of Dν−M meeting λj×{∞} for some j>1 is impossible, where λj denotes the jth eigenvalue of the problem (1.3). Assume on the contrary that Dν−M meets λj×{∞} for some j>1. So there exists a neighborhood ˜N⊂˜M of λj×{∞} such that u must change sign for any (λ,u)∈(Dν−M)∩(˜N∖(λj×{∞})), where ˜M is a neighborhood of λj×{∞} which satisfies the assumptions of Theorem 2.2, which contradicts Dν⊂((R×Pν)∪{(λ1f∞,∞)}).
Thus,
(λ1f∞,∞)⊆ProjR(D+). |
Similarly, we have
(λ1f∞,∞)⊆ProjR(D−). |
Now, we give Proof of Theorem 1.1
Proof of Theorem 1.1.
(a) Since problem (1.1) has a unique solution u≡0 for λ≡0, we get
(C+∪C−∪D+∪D−)⊂{(μ,z)∈R×E|μ≥0}. |
By Lemmas 3.1–3.3, we conclude the desired results.
We only derive the rate of decay of u+∞ since the proof for the other case is completely analogous.
By (2.1), we have
u+∞=λ∫RNΓn(x−y)a(y)f(u+∞(y))dy, |
where ΓN(x−y)=1N(N−2)ωN|x−y|2−N, ωN being the volume of the unit ball in RN.
By f0,f∞∈(0,∞), there exist some constant ϱ>0 such that |f(s)|≤ϱ|s| for any s∈R. Then, we have that
u+∞=λ∫RNΓn(x−y)a(y)f(u+∞(y))dy≤ϱλ∫RNΓn(x−y)a(y)u+∞(y)dy. |
Since u+∞ is bounded, one can get u+∞≤c3 for some constants c3>0. By condition (1.2), it follows that
∫RNa(y)u+∞(y)dy≤c3∫RNP(y)dy≤c3∫+∞0rN−1P(r)dr<+∞. | (3.5) |
By (3.5), for any ε>0, there exists a R>0 such that for all x∈RN
|x|N−2∫ΩRΓN(x−y)a(y)u+∞(y)dy<ε4,∫ΩRa(y)u+∞(y)dy<ε4cN | (3.6) |
and
lim|x|→+∞|x|N−2∫BRΓN(x−y)a(y)u+∞(y)dy=cN∫BRa(y)u+∞(y)dy, | (3.7) |
where ΩR={y∈RN:|y|>R}, BR={y∈RN:|y|<R}.
Furthermore, by (3.6), (3.7), and proof of [6, Theorem 1.1: p. 5948-5949], one can obtain
lim|x|→+∞|x|N−2∫RNΓN(x−y)a(y)u+∞(y)dy=cN∫RNa(y)u+∞(y)dy. |
where cN=1N(N−2)ωN.
By (3.5), it follows that
lim|x|→+∞|x|N−2u+∞(x)=λ⋅lim|x|→+∞|x|N−2∫RNΓn(x−y)a(y)f(u+∞(y))dy≤λϱ⋅lim|x|→+∞|x|N−2∫RNΓn(x−y)a(y)u+∞(y)dy≤cNλϱ∫RNa(y)u+∞(y)dy<∞. |
Therefore, we have that
lim|x|→+∞|x|N−2u+i(x)=ci1 |
for all x∈RN and some constants ci1≠0, where i=0,∞. Do the same for u−i.
(b) Inspired by the idea of [13], we define the cut-off function of f as the following
f[n](s):={f(s),s∈[−n,n],2n2−f(n)n(s−n)+f(n),s∈(n,2n),2n2+f(−n)n(s+n)+f(−n),s∈(−2n,−n),ns,s∈(−∞,−2n]∪[2n,+∞). |
We consider the following problem
{−Δu=λa(x)f[n](u),x∈RN,u(x)→0,as|x|→+∞. | (3.8) |
Clearly, we can see that limn→+∞f[n](s)=f(s), (f[n])0=f0 and (f[n])∞=n.
By Remarks 3.1 and 3.2, there are two distinct unbounded continuum Cν and Dν[n] of solutions of the problem (3.8) emanating from (λ1f0,0), and (λ1n,∞)) respectively, such that they are disjoint, unbounded in the direction of λ and
Cν⊂(R×Pν)∪{(λ1f0,0)},Dν[n]⊂(R×Pν)∪{(λ1n,∞)}, |
where ν=+,−.
By Lemma 2.2, one derives that for each ϵ>0 there exists an N, such that for n>N, Dν[n]⊂Vϵ(Dν), where Dν=lim supn→+∞Dν[n]. So it achieves
(λ1n,+∞)⊆ProjR(Dν[n])⊆ProjRVϵ(Dν). |
It follows (λ1n+ϵ,+∞)⊆ProjR(Dν). The arbitrariness of ϵ>0 and n imply
(0,∞)⊆ProjR(Dν). |
Similar the proof of Lemma 3.2, we have that
(λ1f0,∞)⊆ProjR(Cν). |
Similar the proof of (a), we have that
lim|x|→+∞|x|N−2u+i(x)=ci1 |
for all x∈RN and some constants ci1≠0, where i=0,∞. Do the same for u−i.
(c) Define
f[n](s):={ns,s∈[−1n,1n],[f(2n)−1](ns−2)+f(2n),s∈(1n,2n),−[f(−2n)+1](ns+2)+f(−2n),s∈(−2n,−1n),f(s),s∈(−∞,−2n]∪[2n,+∞). |
We consider the following problem
{−Δu=λa(x)f[n](u),x∈RN,u(x)→0,as|x|→+∞. | (3.9) |
Clearly, we can see that limn→+∞f[n](s)=f(s), (f[n])0=n and (f[n])∞=f∞.
By Remarks 3.1 and 3.2, there are two distinct unbounded continuum Cν and Dν[n] of solutions of the problem (3.9) emanating from (λ1n,0), and (λkf∞,∞)) respectively, such that they are disjoint, unbounded in the direction of λ and
Cν[n]⊂(R×Pν)∪{(λ1n,0)},Dν⊂(R×Pν)∪{(λ1f∞,∞)}, |
where ν=+,−.
Taking z∗=(0,0), and applying Lemma 2.1 again, one derives that Cν=lim supn→+∞Cν[n] is unbounded and connected, moreover, z∗∈Cν.
We claim that Cν∩(R×Pν)={(0,0)}).
Suppose on the contrary that there exists a sequence (λn,un)∈Cν∖{(0,0)}=lim supn→∞Cν[n]∖{(0,0)} such that limn→∞λn=μ≠0 and limn→∞‖un‖=0. Hence, for any N0∈N, there exists n0≥N0 such that (λn,un)∈Cν[n0]. By (3.9), it follows that λn0=λ1n for n0≥N0. From the arbitrary of N0, it implies that n0→∞, i.e., μ=0, which contradicts the assumption of μ≠0.
Lemma 3.1(ⅰ) implies that the projection of C+k,0 on R is unbounded.
Furthermore, we have
(0,∞)⊆ProjR(Cνk,0). |
Similar the proof of Lemma 3.3, we have that
(λ1f∞,∞)⊆ProjR(Dν). |
(d) Similar the proof of (b) and (c), respectively, we have that
(0,∞)⊆ProjR(Dν) |
and
(0,∞)⊆ProjR(Cν). |
To prove Theorems 1.2 and 1.3, by Dai and Han [14], Afrouzi and Rasouli [15], we first give the definition of linearly stable solution for the problem (1.1). For any ϕ∈E and positive solution u of problem (1.1), we can calculate that the linearized eigenvalue problem of (1.1) at the direction ϕ is
{−Δϕ−λa(x)f′(u)ϕ=μϕ,x∈RN,ϕ(x)→0,as|x|→+∞, | (4.1) |
where f′(u)ϕ denotes the Fr′echet derivative of f about u at the direction ϕ. A solution u of problem (1.1) is stable if all eigenvalues of problem (4.1) are positive, otherwise it is unstable. We define the Morse index M(u) of u to problem (1.1) to be the number of negative eigenvalues of problem (4.1). A solution u of problem (1.1) is degenerate if 0 is an eigenvalue of problem (4.1), otherwise it is non-degenerate. The following lemma is our main stability result for the one-sign solution.
Lemma 4.1. Under the assumptions of Theorem 1.2 (a), then any positive or negative solution u of problem (1.1) is stable and non-degenerate, and Morse index M(u)=0.
Proof. Without loss of generality, let u be a positive solution of problem (1.1), and let (μ1,ϕ1) be the corresponding principal eigenpair of problem (4.1) with ϕ1>0 in RN. Notice that u and ϕ1 satisfy
{−Δu=λa(x)f(u),x∈RN,u(x)→0,as|x|→+∞ | (4.2) |
and
{−Δϕ1−λa(x)f′(u)ϕ1=μϕ1,x∈RN,ϕ1(x)→0,as|x|→+∞. | (4.3) |
Multiplying the first equation of problem (4.3) by u and the first equation of problem (4.2) by ϕ1, subtracting and integrating, we obtain
μ1∫RNϕ1udx=λ∫RNa(x)ϕ1(f(u)−f′(u)u)dx. |
By some simple computations, we can show that it follows from (A4) that f(s)−f′(s)s≥0 for any s≥0. Since u>0 and ϕ1>0 in RN, we have μ1>0 and the positive solution u must be stable.
Proof of Theorem 1.2
(a) By Theorem 2.1, we can obtain the problem (1.1) possesses at least two one-sign solutions u+ and u− such that u+>0 and u−<0 in RN. In order to prove exact multiplicity of one-sign solutions for (1.1). Define F:R×E→R by
F(λ,u)=−Δu−λa(x)f(u). |
From Lemma 4.1, we know that any one-sign solution (λ,u) of problem (1.1) is stable. Therefore, at any one-sign solution (λ∗,u∗) for the problem (1.1), we can apply Implicit Function Theorem to F(λ,u)=0, and all the solutions of F(λ,u)=0 near (λ∗,u∗) are on a curve (λ,u(λ)) with |λ−λ∗|≤ε for some small ε>0. Furthermore, by virtue of Remark 2.2, the unbounded continua D+ and D− are all curves.
To complete the proof, it suffices to show that u+(λ,⋅)(u−(λ,⋅)) is increasing (decreasing) with respect to λ. We only prove the case of u+(λ,⋅). The proof of u−(λ,⋅) can be given similarly. Since u+(λ,⋅) is differentiable with respect to λ (as a consequence of Implicit Function Theorem), taking the derivative of the first equation of problem (4.2) by λ, one can obtain that
−Δ(du+dλ)=λa(x)f′(u+)du+dλ+a(x)f′(u+). | (4.4) |
Multiplying the first equation of problem (4.4) by u and the first equation of problem (4.2) by du+dλ, subtracting and integrating, we obtain
∫RN[λa(x)(f′(u+)u+−f(u+))du+dλ+f(u+)u+]dx=0. |
(A2) implies f(s)s≥0 for any s∈R. So we get (f′(u+)u+−f(u+))du+dλ≤0 by (A1). While (A4) shows that f′(u+)u+−f(u+)≤0. Therefore, we have du+dλ≥0.
Next, we only prove the case of the uniqueness of positive solution of problem (1.1) since the proof of the uniqueness of negative solution of problem (1.1) is similar. Suppose on the contrary that there exist two solutions v+1 and v+2 corresponding to λ with v+1∈D+ of the problem (1.1) for λ∈(λ1/f0,+∞). For ε>0, take (λ−ε,v+λ−ε),(λ+ε,v+λ+ε)∈D+, then vλ±ε→v+1 as ε→0. By the monotonicity of v+2 with respect to λ, we get v+λ−ε≤v+2≤v+λ+ε. Then v+2=v+1.
Similar the proof of (a) of Theorem 1.1, we have that
lim|x|→+∞|x|N−2u+1(x)=c1andlim|x|→+∞|x|N−2u−1(x)=c2 |
for all x∈RN and some constants c1,c2≠0.
Proof of Theorem 1.3. By Theorem 4.1 and Lemma 2.1, there is a distinct unbounded continuum Dν(ν∈{+,−}) of solutions of the problem (1.2) emanating from (0,0). Furthermore, we have that the problem (1.1) possesses at least two one-sign solutions u+ and u− such that u+>0 and u−<0 in (0,+∞). In view of the argument of Theorem 1.2, the desired conclusion can be obtained immediately.
In this study, we have proved the existence of one-sign solutions without signum condition for the semilinear elliptic problems (1.1) by the bifurcation techniques and the approximation of connected components. We also obtained the exact multiplicity of one-sign solutions for the problem (1.1) under more strict hypotheses. Our fifindings can be applied to further other differential equations with various other boundary conditions, high dimensional case, and so on as future work.
The author declares no conflicts of interest in this paper.
[1] |
A. L. Edelson, A. J. Rumbos, Linear and semilinear eigenvalue problems in Rn, Comm. Part. Diff. Eq., 18 (1993), 215–240. https://doi.org/10.1080/03605309308820928 doi: 10.1080/03605309308820928
![]() |
[2] | A. J. Rumbos, A. L. Edelson, Bifurcation properties of semilinear elliptic equations in Rn, Differ. Integral Equ., 7 (1994), 399–410. https://projecteuclid.org/7.2/1369330436 |
[3] |
E. N. Dancer, Global solution branches for positive mappings, Arch. Ration. Mech. An., 52 (1973), 181–192. https://doi.org/10.1007/BF00282326 doi: 10.1007/BF00282326
![]() |
[4] |
A. L. Edelson, M. Furi, Global solution branches for semilinear equations in Rn, Nonlinear Anal., 28 (1997), 1521–1532. https://doi.org/10.1016/S0362-546X(96)00018-1 doi: 10.1016/S0362-546X(96)00018-1
![]() |
[5] |
P. H. Rabinowitz, Some global results for nonlinear eigenvalue problems, J. Funct. Anal., 7 (1971), 487–513. https://doi.org/10.1016/0022-1236(71)90030-9 doi: 10.1016/0022-1236(71)90030-9
![]() |
[6] |
G. Dai, J. Yao, F. Li, Spectrum and bifurcation for semilinear elliptic problems in RN, J. Differ. Equations, 263 (2017), 5939–5967. http://dx.doi.org/10.1016/j.jde.2017.07.004 doi: 10.1016/j.jde.2017.07.004
![]() |
[7] |
E. N. Dancer, On the structure of solutions of non-linear eigenvalue problems, Indiana Univ. Math. J., 23 (1974), 1069–1076. https://doi.org/10.1512/iumj.1974.23.23087 doi: 10.1512/iumj.1974.23.23087
![]() |
[8] |
G. Dai, Bifurcation and standing wave solutions for a quasilinear Schr¨odinger equation, P. Roy. Soc. Edinb., 149 (2019), 939–968. https://doi.org/10.1017/prm.2018.59 doi: 10.1017/prm.2018.59
![]() |
[9] |
G. Dai, Bifurcation and one-sign solutions of the p-Laplacian involving a nonlinearity with zeros, Discrete Cont. Dyn. Syst., 36 (2016), 5323–5345. https://doi.org/10.48550/arXiv.1511.06756 doi: 10.48550/arXiv.1511.06756
![]() |
[10] |
P. H. Rabinowitz, On bifurcation from infinity, J. Funct. Anal., 14 (1973), 462–475. https://doi.org/10.1016/0022-0396(73)90061-2 doi: 10.1016/0022-0396(73)90061-2
![]() |
[11] | J. Loˊpez-G\acute{o}mez, Spectral theory and nonlinear functional analysis, Chapman and Hall/CRC, Boca Raton, 2001. |
[12] |
M. Montenego, Strong maximum principles for super-solutions of quasilinear elliptic equations, Nonlinear Anal., 37 (1999), 431–448. https://doi.org/10.1016/S0362-546X(98)00057-1 doi: 10.1016/S0362-546X(98)00057-1
![]() |
[13] | A. Ambrosetti, R. M. Calahorrano, F. R. Dobarro, Global branching for discontinuous problems, Comment. Math. Univ. Ca., 31 (1990), 213–222. https://zbmath.org/0732.35101 |
[14] |
G. Dai, X. Han, Exact multiplicity of one-sign solutions for a class of quasilinear eigenvalue problems, J. Math. Res. Appl., 34 (2014), 84–88. https://doi.org/10.3770/j.issn:2095-2651.2014.01.008 doi: 10.3770/j.issn:2095-2651.2014.01.008
![]() |
[15] |
G. A. Afrouzi, S. H. Rasouli, Stability properties of non-negative solutions to a non-autonomous p-Laplacian equation, Chaos Soliton. Fract., 29 (2006), 1095–1099. https://doi.org/10.1016/j.chaos.2005.08.165 doi: 10.1016/j.chaos.2005.08.165
![]() |