
A new bifurcation from simple eigenvalue theorem is proved for general nonlinear functional equations. It is shown that in this bifurcation scenario, the bifurcating solutions are on a curve which is tangent to the line of trivial solutions, while in typical bifurcations the curve of bifurcating solutions is transversal to the line of trivial ones. The stability of bifurcating solutions can be determined, and examples from partial differential equations are shown to demonstrate such bifurcations.
Citation: Ping Liu, Junping Shi. A degenerate bifurcation from simple eigenvalue theorem[J]. Electronic Research Archive, 2022, 30(1): 116-125. doi: 10.3934/era.2022006
[1] | J. F. Toland . Path-connectedness in global bifurcation theory. Electronic Research Archive, 2021, 29(6): 4199-4213. doi: 10.3934/era.2021079 |
[2] | Gheorghe Moza, Mihaela Sterpu, Carmen Rocşoreanu . An analysis of two degenerate double-Hopf bifurcations. Electronic Research Archive, 2022, 30(1): 382-403. doi: 10.3934/era.2022020 |
[3] | Mengxin He, Zhong Li . Dynamic behaviors of a Leslie-Gower predator-prey model with Smith growth and constant-yield harvesting. Electronic Research Archive, 2024, 32(11): 6424-6442. doi: 10.3934/era.2024299 |
[4] | Yujia Xiang, Yuqi Jiao, Xin Wang, Ruizhi Yang . Dynamics of a delayed diffusive predator-prey model with Allee effect and nonlocal competition in prey and hunting cooperation in predator. Electronic Research Archive, 2023, 31(4): 2120-2138. doi: 10.3934/era.2023109 |
[5] | Bin Long, Shanshan Xu . Persistence of the heteroclinic loop under periodic perturbation. Electronic Research Archive, 2023, 31(2): 1089-1105. doi: 10.3934/era.2023054 |
[6] | Zhuo Ba, Xianyi Li . Period-doubling bifurcation and Neimark-Sacker bifurcation of a discrete predator-prey model with Allee effect and cannibalism. Electronic Research Archive, 2023, 31(3): 1405-1438. doi: 10.3934/era.2023072 |
[7] | Jie Xia, Xianyi Li . Bifurcation analysis in a discrete predator–prey model with herd behaviour and group defense. Electronic Research Archive, 2023, 31(8): 4484-4506. doi: 10.3934/era.2023229 |
[8] | Xin Du, Quansheng Liu, Yuanhong Bi . Bifurcation analysis of a two–dimensional p53 gene regulatory network without and with time delay. Electronic Research Archive, 2024, 32(1): 293-316. doi: 10.3934/era.2024014 |
[9] | Dong Li, Xiaxia Wu, Shuling Yan . Periodic traveling wave solutions of the Nicholson's blowflies model with delay and advection. Electronic Research Archive, 2023, 31(5): 2568-2579. doi: 10.3934/era.2023130 |
[10] | Qinghua Zhu, Meng Li, Fang Han . Hopf bifurcation control of the ML neuron model with Hc bifurcation type. Electronic Research Archive, 2022, 30(2): 615-632. doi: 10.3934/era.2022032 |
A new bifurcation from simple eigenvalue theorem is proved for general nonlinear functional equations. It is shown that in this bifurcation scenario, the bifurcating solutions are on a curve which is tangent to the line of trivial solutions, while in typical bifurcations the curve of bifurcating solutions is transversal to the line of trivial ones. The stability of bifurcating solutions can be determined, and examples from partial differential equations are shown to demonstrate such bifurcations.
In many mathematical models, it is required to finding the solutions of a stationary problem, which can be formulated as an equation
F(λ,u)=0, | (1.1) |
where F is a nonlinear smooth mapping defined on (λ,u)∈R×X and mapped to Y, λ is a parameter, and X,Y are Banach spaces. Often the system Eq (1.1) has a trivial state u=u0 for all parameter values λ, and it may have other nontrivial solutions near (λ0,u0) for some λ0. Such λ0 is called a bifurcation point for Eq (1.1), and the bifurcating nontrivial solutions near a bifurcation point are often with significance for the models as they represent states breaking from the symmetric or uniform ones.
If the Fréchet derivative Fu(λ0,u0) of F at (λ0,u0) is invertible, then (λ0,u0) is not a bifurcation point from the Implicit Function Theorem [1,2,3]. Hence a necessary condition for the bifurcation to occur is that Fu(λ0,u0) is not invertible. The most useful bifurcation occurs when that 0 is a simple eigenvalue of the linearized operator Fu(λ0,u0), that is
(F1) dimN(Fu(λ0,u0))=codimR(Fu(λ0,u0))=1, and N(Fu(λ0,u0))=span{w0},
where N(Fu) and R(Fu) are the null space and the range of linear operator Fu. Crandall and Rabinowitz [2] prove the following celebrated "bifurcation from a simple eigenvalue" theorem (see [2,Theorem 1.7]). Here is an expanded version of the theorem for our purpose:
Theorem 1.1. Let U be a neighborhood of (λ0,u0) in R×X, and let F:U→Y be a twice continuously differentiable mapping. Assume that F(λ,u0)=0 for (λ,u0)∈U. At (λ0,u0), F satisfies (F1) and (F3) Fλu(λ0,u0)[w0]∉R(Fu(λ0,u0)).
Let Z be any complement of span{w0} in X. Thenthe solution set of (1.1) near (λ0,u0) consistsprecisely of the curves Γ0={(λ,u0)} and Γ1={(λ(s),u(s)):s∈|s|<δ}, where λ:I→R, z:I→Zare C1 functionssuch that u(s)=u0+sw0+sz(s), λ(0)=λ0, z(0)=0, and
λ′(0)=−⟨l,Fuu(λ0,u0)[w0]2⟩2⟨l,Fλu(λ0,u0)[w0]⟩, | (1.2) |
where l∈Y∗satisfying N(l)=R(Fu(λ0,u0)). If F also satisfies
(F4) Fuu(λ0,u0)[w0]2∉R(Fu(λ0,u0)),
then we have λ′(0)≠0, and it is called a transcritical bifurcation; If F satisfies
(F4′) Fuu(λ0,u0)[w0]2∈R(Fu(λ0,u0)),
and in addition F∈C3, then λ′(0)=0 and
λ″(0)=−⟨l,Fuuu(λ0,u0)[w0]3⟩+3⟨l,Fuu(λ0,u0)[w0,θ1]⟩3⟨l,Fλu(λ0,u0)[w0]⟩, | (1.3) |
where θ1 satisfies
Fuu(λ0,u0)[w0]2+Fu(λ0,u0)[θ1]=0. | (1.4) |
If λ′(0)=0 and λ″(0)≠0, then it is called a pitchfork bifurcation.
Note that the classification of transcritical and pitchfork bifurcations using (F4) and (F4′) was first used in reference [4], although this has been widely used in finite dimensional dynamical systems [5].
The transversality condition (F3) holds in most practical situations, but there are also important exceptional cases for which (F3) fail. In reference [6], we considered a degenerate bifurcation scenario in which (F3) is not satisfied. In this case, we prove that, under some higher order transversality conditions on F, the local solution set of Eq (1.1) near the bifurcation point (λ0,u0) consists of the line of trivial solutions, and two other solution curves. First we recall a degenerate version of Theorem 1.1, which can be used to obtain more than two intersecting solution curves near the bifurcation point.
Theorem 1.2. ([6,Theorem 2.3]) Let U be a neighborhood of (λ0,u0) in R×X, and let F∈C3(U,Y).Assume that F(λ,u0)=0 for (λ,u0)∈U and at(λ0,u0), F satisfies (F1), (F4′), and
(F3′) Fλu(λ0,u0)[w0]∈R(Fu(λ0,u0)).
Let X=N(Fu(λ0,u0))⊕Z be a fixed splitting of X, and let l∈Y∗ such that R(Fu(λ0,u0))={v∈Y:⟨l,v⟩=0}.Denote by θ2∈Z the unique solution of
Fλu(λ0,u0)[w0]+Fu(λ0,u0)[θ2]=0, | (1.5) |
and recall θ1 to be the unique solution of Eq (1.4). We assume that the matrix (all derivatives areevaluated at (λ0,u0))
H=H(λ0,u0)=(H11H12H12H22) | (1.6) |
is non-degenerate, i.e., det(H)≠0, where Hij is given by
H11=⟨l,Fλλu[w0]+2Fλu[θ2]⟩, | (1.7) |
H12=12⟨l,Fλuu[w0]2+Fλu[θ1]+2Fuu[w0,θ2]⟩, | (1.8) |
H22=13⟨l,Fuuu[w0]3+3Fuu[w0,θ1]⟩. | (1.9) |
1) If H is definite, i.e., det(H)>0, then the solution set of Eq (1.1) near (λ,u)=(λ0,u0) is the line Γ0={(λ,u0)}.
2) If H is indefinite, i.e., det(H)<0, then the solution set of Eq (1.1) near (λ,u)=(λ0,u0) is the union of C1curves intersecting at (λ0,u0), including the line of trivial solutions Γ0={(λ,u0)}and two other curves Γi={(λi(s),ui(s)):|s|<δ} (i=1,2) for someδ>0, with
λi(s)=λ0+μis+sαi(s),ui(s)=u0+ηisw0+sβi(s), |
where (μ1,η1) and (μ2,η2) are non-zerolinear independent solutions of the equation
H11μ2+2H12μη+H22η2=0, | (1.10) |
αi(0)=α′i(0)=0, βi(s)∈Z, andβi(0)=β′i(0)=0, i=1,2.
In this paper, we prove another bifurcation result when the transversality condition (F3) fails. In this case, under the complement (F3′) of (F3), and as well as (F4), we show that the solution set of Eq (1.1) near the bifurcation point (λ0,u0) consists of the line of trivial solution, and another curve of nontrivial solutions which is tangent to the line of trivial ones.
Theorem 1.3. Let U be a neighborhood of (λ0,u0) in R×X, and let F∈C2(U,Y). Assume that F(λ,u0)=0 for any λ∈R. At(λ0,u0), F satisfies (F1), (F3′) and (F4).Then the solution set of Eq (1.1) near (λ,u)=(λ0,u0) is the union of two C2 curves which are tangent to each other at (λ0,u0), including the line of trivial solutions Γ0={(λ,u0)} and Γ1={(λ,u(λ)):λ∈I} for some δ>0, where I=(λ0−δ,λ0+δ), u(λ)=u0+t(λ)w0+g(λ,t(λ)), t:I→V and g:I×V→Z are continuously differentiable functions, t(λ0)=t′(λ0)=0, and g(λ,0)=gλ(λ0,0)=gt(λ0,0)=0, where V⊂R is a neighborhood of t=0. Moreover if F∈C4(U,Y), then t(λ) is C3 and
t″(λ0)=−2⟨l,Fλλu(λ0,u0)[w0]+2Fλu(λ0,u0)[θ2]⟩⟨l,Fuu(λ0,u0)[w0]2⟩, | (1.11) |
where θ2 is defined by Eq (1.5).
If F satisfies Fλλu(λ0,u0)[w0]+3Fλu(λ0,u0)[θ2]∉R(Fu(λ0,u0), then the solution set of Eq (1.1) near (λ0,u0) in Theorem 1.3 is the union of a line and a parabola-like curve which is tangent to the line. The simplest example for Theorem 1.3 is the function f:R2→R defined by f(λ,u)=u(u−λ2), and its solution set of f(λ,u)=0 near (λ,u)=(0,0) is the union of the line u=0 and the curve of nontrivial solutions u=λ2 which are tangent to each other at (0,0) (see the lower left panel of Figure 1).
Theorem 1.1, Theorem 1.2 and Theorem 1.3 together provide a complete classification of bifurcation scenarios for Eq (1.1) under the conditions (F3) or (F3′), and (F4) or (F4′), while F(λ,u0)≡0 and (F1) are assumed:
(a) Transcritical: (F3) and (F4), a crossing curve of nontrivial solutions (Theorem 1.1);
(b) Pitchfork: (F3) and (F4′), a crossing curve of nontrivial solutions bending leftward or rightward (Theorem 1.1);
(c) Tangential: (F3′) and (F4), a tangential curve of nontrivial solutions bending upward or downward (Theorem 1.3); and
(d) Double transcritical: (F3′) and (F4′), two crossing curves of nontrivial solutions (Theorem 1.2).
Because of the condition (F3′), the tangential and the double transcritical bifurcations are called degenerate ones. Figure 1 shows examples of each types of bifurcations using simple mappings Fi:R×R→R, and these mappings can be regarded as normal forms of these bifurcations.
Another bifurcation result with solution set being two tangential curves was proved in reference [7] but it is under the assumption that the kernel N(Fu(λ0,u0)) is of two-dimensional and it is of saddle-node bifurcation type. It is applied to a nonlinear Schrödinger system with quadratic nonlinearity [8], where one of the two tangential curves is indeed vertical in a form {(λ0,kφ1):k∈R}. Here one of the two tangential curves is horizontal in a form {(λ,u0):λ∈R}.
We also remark that transversality conditions like (F3) or (F4) are not needed for global bifurcation theorems concerning the topological structure of the solution continuum. Indeed in the celebrated Rabinowitz global bifurcation theorem [9] (see also extensions in [10,11]), only the odd algebraic multiplicity was assumed, thus all four scenarios shown in Figure 1 can occur as local pictures for the global bifurcation diagrams in reference [9,10,11].
We prove Theorem 1.3 in Section 2, and we consider the stability of the bifurcating solutions obtained in Theorem 1.3 in Section 3. Finally in Section 4 we show some examples to apply Theorem 1.3. Throughout the paper, we use the same labeling of conditions such as (F1) and (F2) on F as in our previous work [4,6,12], and we use the convention that (Fi′) stands for the negation of (Fi) for i∈N. We use ||⋅|| as the norm of Banach space X, ⟨⋅,⋅⟩ as the duality pair of a Banach space X and its dual space X∗. For a linear operator L, we use N(L) as the null space of L and R(L) as the range space of L, and we use L[w] to denote the image of w under the linear mapping L. For a multilinear operator L, we use L[w1,w2,⋯,wk] to denote the image of (w1,w2,⋯,wk) under L, and when w1=w2=⋯=wk, we use L[w1]k instead of L[w1,w1,⋯,w1]. For a nonlinear operator F, we use Fu as the partial derivative of F with respect to argument u.
First we recall an important lemma from our previous work [12]. First is the well-known Lyapunov-Schmidt reduction under the condition (F1) which is standard from many textbooks in nonlinear analysis (see for example [1,3,13]).
Lemma 2.1. Supposethat F:R×X→Y is a Cp (p≥1) mapping such that F(λ0,u0)=0, and F satisfies (F1) at (λ0,u0). Thenthe equation F(λ,u)=0 for (λ,u) near (λ0,u0) can be reduced to
⟨l,F(λ,u0+tw0+g(λ,t))⟩=0, |
where t∈(−δ,δ), λ∈(λ0−δ,λ0+δ) where δ is a small constant, l∈Y∗ such that ⟨l,v⟩=0 if and only if v∈R(Fu(λ0,u0)), and g is a Cp function into Z such that g(λ0,0)=0 and Z is acomplement of N(Fu(λ0,u0)) in X.
Proof of Theorem 1.3. We denote the projection from Y into R(Fu(λ0,u0)) by Q. Applying Lemma 2.1 to F in Theorem 1.3 at (λ0,u0), we have that the function g(λ,t) in Lemma 2.1 is obtained from (see [12]),
f1(λ,t)≡Q∘F(λ,u0+tw0+g(λ,t))=0. | (2.1) |
Since u0 is a trivial solution for all λ near λ0, that is, F(λ,u0)≡0, then by Lemma 2.1 we have g(λ,0)≡0, hence gλ(λ0,0)=gλλ(λ0,0)=0. Differentiating f1 and evaluating at (λ,t)=(λ0,0), we obtain
0=∇f1=(Q∘(Fλ+Fu[gλ]),Q∘Fu[w0+gt]). | (2.2) |
Since Fu[w0]=0 and gt∈Z, and Fu(λ0,c0u∗)|Z is an isomorphism, then gt(λ0,0)=0.
Next we calculate the second derivatives of f1:
∂2f1∂λ∂t(λ0,0)=Q∘(Fλu(λ0,u0)[w0+gt(λ0,0)]+Fuu(λ0,u0)[w0+gt(λ0,0),gλ(λ0,0)]+Fu(λ0,u0)[gλt(λ0,0)])=Fλu(λ0,u0)[w0]+Fu(λ0,u0)[gλt(λ0,0)]=0, |
thus gλt(λ0,0)=θ2 from (F3'), where θ2 is defined as in Eq (1.5). We define the bifurcation function
f(λ,t)=⟨l,F(λ,u0+tw0+g(λ,t))⟩. | (2.3) |
From the assumptions, f is C2 in U and f(λ,0)=0. Next we apply the Implicit Function Theorem to the equation h(λ,t)=0 where the function h(λ,t) is defined by
h(λ,t)={1tf(λ,t),if t≠0,ft(λ,0),if t=0. | (2.4) |
Then h(λ,0)=0 from the assumption that F(λ,u0)=0, and from (F4), we have
ht(λ0,0)=limt→01t(h(λ0,t)−h(λ0,0))=limt→01t(1tf(λ0,t)−ft(λ0,0))=limt→01t2(f(λ0,t)−f(λ0,0)−ft(λ0,0)t)=12ftt(λ0,0)=12⟨l,Fuu(λ0,u0)[w0+gt(λ0,0)]2+Fu(λ0,u0)[gtt(λ0,0)]⟩=12⟨l,Fuu(λ0,u0)[w0]2⟩≠0. |
By the Implicit Function Theorem, there exists a unique continuously differentiable function t=t(λ)∈R satisfying h(λ,t(λ))=0 and t(λ0)=0, and
F(λ,u0+t(λ)w0+g(λ,t(λ)))=0. | (2.5) |
Now we assume that F∈C4(U,Y). Let u(λ)=u0+t(λ)w0+g(λ,t(λ)). Then we have
F(λ,u(λ))=0. | (2.6) |
Differentiating Eq (2.6) with respect to λ twice and evaluating at λ=λ0, we obtain that
Fλλ+2Fλu[uλ]+Fuu[uλ]2+Fu[uλλ]=0. | (2.7) |
Here all partial derivatives are evaluated as λ=λ0. By applying l∈Y∗ to Eq (2.7), we have t′(λ0)=0 from (F3′) and (F4), and we also have uλ=0 and uλλ=t″(λ0)w0. Differentiating Eq (2.6) with respect to λ three times and evaluating at λ=λ0, we obtain that
3Fλu[uλλ]+Fu[uλλλ]=0, | (2.8) |
which implies that uλλλ=3t″(λ0)θ2. Finally differentiating Eq (2.6) with respect to λ four times and evaluating at λ=λ0, we obtain that
6Fλλu[uλλ]+4Fλu[uλλλ]+3Fuu[uλλ]2+Fu[uλλλλ]=0. | (2.9) |
By applying l∈Y∗ to Eq (2.9), we can obtain Eq (1.11).
In this section, we consider the stability of the bifurcating solutions obtained in Theorem 1.3. First similar to [14,Corollary 1.13], we have the
Proposition 3.1. Let X, Y, U, F, Z, λ0, w0, θ2 be the same as in Theorem 1.3, and let all assumptions in Theorem 1.3 on F be satisfied.In addition we assume that X⊂Y, and the inclusion mapping i:X→Y is continuous. Let Γ1={(λ,u(λ)):|λ−λ0|<δ} be the solution curve in Theorem 1.3.Then there exist ε>0, C2 functions γ:(λ0−ε,λ0+ε)→R, μ:(λ0−ε,λ0+ε)→R, v:(λ0−ε,λ0+ε)→X and ω:(λ0−ε,λ0+ε)→X such that
Fu(λ,u0)[v(λ)]=γ(λ)v(λ)forλ∈(λ0−ε,λ0+ε), | (3.1) |
Fu(λ,u(λ))[ω(λ)]=μ(λ)ω(λ)forλ∈(λ0−ε,λ0+ε), | (3.2) |
where γ(λ0)=μ(λ0)=0, v(λ0)=ω(λ0)=w0, v(λ)−w0∈Z and ω(λ)−w0∈Z.
We have the following result on the stabilities of the bifurcating solution (λ,u(λ)) obtained in Theorem 1.3.
Proposition 3.2. Let the assumptions of Proposition 3.1 hold, and let γ(λ) and μ(λ) be the functions defined in Proposition 3.1.In addition, we assume that
w0∉R(Fu(λ0,u0)),wherew0(≠0)∈N(Fu(λ0,u0)). | (3.3) |
Thenγ′(λ0)=μ′(λ0)=0 and
γ″(λ0)=⟨l,Fλλu(λ0,u0)[w0]+2Fλu(λ0,u0)[θ2]⟩⟨l,w0⟩, | (3.4) |
μ″(λ0)=−⟨l,Fλλu(λ0,u0)[w0]+2Fλu(λ0,u0)[θ2]⟩⟨l,w0⟩. | (3.5) |
Proof. We differentiate Eq (3.1) to obtain
Fλu(λ,u0)[v(λ)]+Fu(λ,u0)[v′(λ)]=γ′(λ)v(λ)+γ(λ)v′(λ). | (3.6) |
Setting λ=λ0 and applying l to the equation, we get γ′(λ0)=0 and v′(λ0)=θ2 by (F3′). Differentiating Eq (3.6) again, we have
Fλλu(λ,u0)[v(λ)]+2Fλu(λ,u0)[v′(λ)]+Fu(λ,u0)[v″(λ)]=γ″(λ)v(λ)+2γ′(λ)v′(λ)+γ(λ)v″(λ). | (3.7) |
Setting λ=λ0, we get
Fλλu(λ0,u0)[w0]+2Fλu(λ0,u0)[θ2]+Fu(λ0,u0)[v″(λ0)]=γ″(λ0)v(λ0). | (3.8) |
Thus by applying l to Eq (3.8), we obtain Eq (3.4).
On the other hand, we differentiate Eq (3.2) to obtain
Fλu(λ,u(λ))[ω(λ)]+Fuu(λ,u(λ))[u′(λ),ω(λ)]+Fu(λ,u(λ))[ω′(λ)]=μ′(λ)ω(λ)+μ(λ)ω′(λ). | (3.9) |
Setting λ=λ0, we have
Fλu(λ0,u0)[w0]+Fu(λ0,u0)[ω′(λ0)]=μ′(λ0)w0, | (3.10) |
we get μ′(λ0)=0 and ω′(λ0)=θ2. We differentiate Eq (3.9) again and set λ=λ0, and we have
Fλλu(λ0,u0)[w0]+2Fλu(λ0,u0)[θ2]+t″(λ0)Fuu(λ0,u0)[w0]2+Fu(λ0,u0)[ω″(λ0)]=μ″(λ0)w0, | (3.11) |
by u′(λ0)=0 and u″(λ0)=t″(λ0)w0. Thus by applying l to Eq (3.11) and using Eq (1.11), we obtain Eq (3.5).
Proposition 3.2 implies that the trivial solution u=u0 on Γ0 and the non-trivial solution u(λ) on Γ1 in Theorem 1.3 both have the same stability before and after the bifurcation point. Thus there is no exchange of stability occurring in the tangential bifurcation described in Theorem 1.3. Furthermore, the stability of the trivial solution u0 on Γ0 and the one of the non-trivial solution u(λ) on Γ1 in Theorem 1.3 are always opposite: while one is stable, the other is unstable, or vice versa, if γ(λ0)=μ(λ0)=0 is the principal eigenvalue of Fu(λ0,u0).
We show that the tangential bifurcations described in Theorem 1.3 occurs for the following semilinear elliptic equations.
Example 4.1.
{Δu+u(u−λ2)=0,x∈Ω,∂u∂n=0,x∈∂Ω, | (4.1) |
where λ is a positive parameter, Ω is a bounded region with smooth boundary in Rn for n≥1.
It is easy to see (λ,0) is a trivial solution of Eq (4.1). Define a nonlinear mapping F:R×X→Y by
F(λ,u)=Δu+u(u−λ2), | (4.2) |
where X={u∈W2,p(Ω):∂u∂n=0} and Y=Lp(Ω). It is easy to verify that Fu(0,0)[ϕ]=Δϕ, we have N(Fu(0,0))=span{1}, R(Fu(0,0))={y∈Y:∫Ωydx=0}. And Fλu(0,0)[1]=0, θ2=0, Fuu(0,0)[1]2=2, Fλλu(0,0)[1]=−2, so (F1), (F3′), (F4) are satisfied. We can apply Theorem 1.3 to F. Then the solution set of Eq (4.1) near (λ,u)=(0,0) is the union of two C1 curves which are tangent to each other at (0,0), including the line of trivial solutions Γ0={(λ,0)} and Γ1={(λ,u(λ)):|λ|<δ} for some δ>0, where u(λ) is a continuously differentiable function, u(λ)=t(λ)+g(λ,t(λ)), t(0)=t′(0)=0 and t″(0)=2. Furthermore from Proposition 3.2, we have γ″(0)=−μ″(0)=−2. It implies the trivial solution (λ,0) is stable and the nontrivial solution (λ,u(λ)) is unstable. Note that this example is rather trivial as u(λ)=λ2 is a constant solution of Eq (4.1).
Example 4.2.
{Δu+2λ√μ1u−λ2u+λu2=0,x∈Ω,u=0,x∈∂Ω, | (4.3) |
where λ is a positive parameter, Ω is a bounded region with smooth boundary in Rn for n≥1, and μ1 is the principal eigenvalue of −Δ on H10(Ω).
For any λ>0, u=0 is a trivial solution of (4.3). Define a nonlinear mapping F:R×X→Y by
F(λ,u)=Δu+2λ√μ1u−λ2u+λu2, | (4.4) |
where X=W2,p(Ω)∩W1,p0(Ω) and Y=Lp(Ω). We show that λ=√μ1 is a bifurcation point for the trivial solution u=0. We can verify that Fu(√μ1,0)[ϕ]=Δϕ+μ1ϕ, we have N(Fu(√μ1,0))=span{φ1}, R(Fu(√μ1,0))={y∈Y:∫Ωyφ1dx=0}, where φ1>0 is the principal eigenfunction of −Δ on H10(Ω) corresponding to μ1. Moreover we can verify that Fλu(√μ1,0)[φ1]=0, θ2=0, Fuu(√μ1,0)[φ1]2=2φ21, Fλλu(√μ1,0)[φ1]=−2φ1. so the conditions (F1), (F3′), (F4) are satisfied. We can apply Theorem 1.3 to F at λ=√μ1. The solution set of Eq (4.3) near (λ,u)=(√μ1,0) is the union of two C1 curves which are tangent to each other at (√μ1,0), including the line of trivial solutions Γ0={(λ,0):λ>0} and Γ1={(λ,u(λ)):|λ−√μ1|<δ} for some δ>0, where u(λ)=t(λ)φ1+g(λ,t(λ)) is smooth, t(√μ1)=t′(√μ1)=0, t″(√μ1)=2A>0 where A=∫Ωφ21/∫Ωφ31>0, and g(λ,0)=gλ(√μ1,0)=gt(√μ1,0)=0. Thus Eq (4.3) has a positive solution u(λ)≈A(λ−√μ1)2φ1 for any 0<|λ−√μ1|<δ. Furthermore from Proposition 3.2, we have γ″(0)=−μ″(0)=−2. It implies the trivial solution (λ,0) is stable and the nontrivial solution (λ,u(λ)) is unstable when λ≠√μ1.
P. Liu is partially supported by NSFC grant 11571086, and J. Shi is partially supported by NSF grant DMS-1853598.
The authors declare there is no conflicts of interest.
[1] | K. C. Chang, Methods in nonlinear analysis, Springer Monographs in Mathematics, 2005. |
[2] |
M. G. Crandall, P. H. Rabinowitz, Bifurcation from simple eigenvalues, J. Funct. Anal., 8 (1971), 321–340. https://doi.org/10.1016/0022-1236(71)90015-2 doi: 10.1016/0022-1236(71)90015-2
![]() |
[3] | K. Deimling. Nonlinear functional analysis, Springer-Verlag, 1985. https://doi.org/10.1007/978-3-662-00547-7 |
[4] |
J. P. Shi, Persistence and bifurcation of degenerate solutions, J. Funct. Anal., 169 (1999), 494–531. https://doi.org/10.1006/jfan.1999.3483 doi: 10.1006/jfan.1999.3483
![]() |
[5] | J. Guckenheimer, P. Holmes, Nonlinear oscillations, dynamical systems, and bifurcations of vector fields, Springer-Verlag, 1983. |
[6] |
P. Liu, J. P. Shi, Y. Wang, Bifurcation from a degenerate simple eigenvalue, J. Funct. Anal., 264 (2013), 2269–2299. https://doi.org/10.1016/j.jfa.2013.02.010 doi: 10.1016/j.jfa.2013.02.010
![]() |
[7] |
P. Liu, J. P. Shi, Y. Wang, A double saddle-node bifurcation theorem, Commun. Pure Appl. Anal., 12 (2013), 2923–2933. https://doi.org/10.3934/cpaa.2013.12.2923 doi: 10.3934/cpaa.2013.12.2923
![]() |
[8] |
L. Zhao, F. Zhao, J. Shi, Higher dimensional solitary waves generated by second-harmonic generation in quadratic media, Calc. Var. Partial Differ. Equations, 54 (2015), 2657–2691. https://doi.org/10.1007/s00526-015-0879-1 doi: 10.1007/s00526-015-0879-1
![]() |
[9] |
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
![]() |
[10] |
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
![]() |
[11] |
E. N. Dancer, Bifurcation from simple eigenvalues and eigenvalues of geometric multiplicity one, Bull. London Math. Soc., 34 (2002), 533–538. https://doi.org/10.1112/S002460930200108X doi: 10.1112/S002460930200108X
![]() |
[12] |
P. Liu, J. Shi, Y. Wang, Imperfect transcritical and pitchfork bifurcations, J. Funct. Anal., 251 (2007), 573–600. https://doi.org/10.1016/j.jfa.2007.06.015 doi: 10.1016/j.jfa.2007.06.015
![]() |
[13] | L.Nirenberg, Topics in nonlinear functional analysis, American Mathematical Society, 2001. |
[14] |
M. G. Crandall, P. H. Rabinowitz, Bifurcation, perturbation of simple eigenvalues and linearized stability, Arch. Rational Mech. Anal., 52 (1973), 161–180. https://doi.org/10.1007/BF00282325 doi: 10.1007/BF00282325
![]() |
1. | Claudia García, Luis Vega, Steady solutions for the Schrödinger map equation, 2024, 49, 0360-5302, 505, 10.1080/03605302.2024.2358500 | |
2. | Rachidi B. Salako, Yixiang Wu, Shuwen Xue, Exact multiplicity, bifurcation curves, and asymptotic profiles of endemic equilibria of a cross-diffusive epidemic model, 2025, 434, 00220396, 113226, 10.1016/j.jde.2025.113226 |