Loading [MathJax]/jax/output/SVG/jax.js
Special Issues

Path-connectedness in global bifurcation theory

  • A celebrated result in bifurcation theory is that, when the operators involved are compact, global connected sets of non-trivial solutions bifurcate from trivial solutions at non-zero eigenvalues of odd algebraic multiplicity of the linearized problem. This paper presents a simple example in which the hypotheses of the global bifurcation theorem are satisfied, yet all the path-connected components of the connected sets that bifurcate are singletons. Another example shows that even when the operators are everywhere infinitely differentiable and classical bifurcation occurs locally at a simple eigenvalue, the global continua may not be path-connected away from the bifurcation point. A third example shows that the non-trivial solutions which bifurcate at non-zero eigenvalues, irrespective of multiplicity when the problem has gradient structure, may not be connected and may not contain any paths except singletons.

    Citation: J. F. Toland. Path-connectedness in global bifurcation theory[J]. Electronic Research Archive, 2021, 29(6): 4199-4213. doi: 10.3934/era.2021079

    Related Papers:

    [1] J. F. Toland . Path-connectedness in global bifurcation theory. Electronic Research Archive, 2021, 29(6): 4199-4213. doi: 10.3934/era.2021079
    [2] Muwafaq Salih, Árpád Száz . Generalizations of some ordinary and extreme connectedness properties of topological spaces to relator spaces. Electronic Research Archive, 2020, 28(1): 471-548. doi: 10.3934/era.2020027
    [3] Xiaochen Mao, Weijie Ding, Xiangyu Zhou, Song Wang, Xingyong Li . Complexity in time-delay networks of multiple interacting neural groups. Electronic Research Archive, 2021, 29(5): 2973-2985. doi: 10.3934/era.2021022
    [4] Sik Lee, Sang-Eon Han . Semi-separation axioms associated with the Alexandroff compactification of the $ MW $-topological plane. Electronic Research Archive, 2023, 31(8): 4592-4610. doi: 10.3934/era.2023235
    [5] Bo Lu, Xiaofang Jiang . Reduced and bifurcation analysis of intrinsically bursting neuron model. Electronic Research Archive, 2023, 31(10): 5928-5945. doi: 10.3934/era.2023301
    [6] Yun Ni, Jinqing Zhan, Min Liu . Topological design of continuum structures with global stress constraints considering self-weight loads. Electronic Research Archive, 2023, 31(8): 4708-4728. doi: 10.3934/era.2023241
    [7] Ming Zhao, Yudan Ma, Yunfei Du . Global dynamics of an amensalism system with Michaelis-Menten type harvesting. Electronic Research Archive, 2023, 31(2): 549-574. doi: 10.3934/era.2023027
    [8] Yanjiao Li, Yue Zhang . Dynamic behavior on a multi-time scale eco-epidemic model with stochastic disturbances. Electronic Research Archive, 2025, 33(3): 1667-1692. doi: 10.3934/era.2025078
    [9] Xueyong Zhou, Xiangyun Shi . Stability analysis and backward bifurcation on an SEIQR epidemic model with nonlinear innate immunity. Electronic Research Archive, 2022, 30(9): 3481-3508. doi: 10.3934/era.2022178
    [10] 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
  • A celebrated result in bifurcation theory is that, when the operators involved are compact, global connected sets of non-trivial solutions bifurcate from trivial solutions at non-zero eigenvalues of odd algebraic multiplicity of the linearized problem. This paper presents a simple example in which the hypotheses of the global bifurcation theorem are satisfied, yet all the path-connected components of the connected sets that bifurcate are singletons. Another example shows that even when the operators are everywhere infinitely differentiable and classical bifurcation occurs locally at a simple eigenvalue, the global continua may not be path-connected away from the bifurcation point. A third example shows that the non-trivial solutions which bifurcate at non-zero eigenvalues, irrespective of multiplicity when the problem has gradient structure, may not be connected and may not contain any paths except singletons.



    Krasnosel'skii [17] considered non-linear eigenvalues in the form

    λx=Lx+R(λ,x),λR,xX, (1.1a)

    where X is a real Banach space, the linear operator L:XX is compact, and the non-linear R:R×XX is continuous, compact, and satisfies

    R(λ,x)x0 as 0x0 uniformly for λ in bounded sets. (1.1b)

    Since R is continuous, (1.1b) implies that R(λ,0)=0, and hence x=0 is a solution of (1.1a), for all λR. Let T={(λ,0):λR} denote the set of trivial solutions of (1.1a) and S the set of solutions that are not trivial. In all that follows, L is linear and compact and R is nonlinear, continuous and compact. The first observation is that under these hypotheses S may be empty.

    Example 1.1. Let X=R2, L(x,y)=(x+y,y) and R(x,y)=(0,x3). Then L is linear, (1.1b) holds, and equation (1.1a) is satisfied if and only if

    (λ1)x=y and (λ1)y=x3,

    which implies that x((λ1)2+x2)=0. Hence x=0 and, by the first equation, y=0, which shows S=.

    According to Krasnosel'skii [17,Ch. Ⅳ,p. 181], a point λ0R is a bifurcation point for (1.1a) if there exists a sequence {(λn,xn)}S such that

    λnλ0 in R and 0xn0 as n.

    In this definition there is no mention of curves, paths or connected sets in S, bifurcating from T at (λ0,0). (A path in S is a set {γ(t):t[0,1]} where γ:[0,1]SR×X is continuous and a path is non-trivial if γ is not constant. A curve is a smooth 1-dimensional manifold.)

    In this paper it is shown by example that even when the operators in (1.1) are smooth, if they are not real-analytic the non-trivial solution sets predicted by classical theories may not be path-connected, and indeed may contain no paths at all.

    A necessary condition for bifurcation. The following necessary criterion for λ0 to be a bifurcation point, when L is compact and R satisfies (1.1b) in a Banach space X, is well known [17,Ch. Ⅳ,§2,Lem. 2.1]. A real number λ00 is a bifurcation point only if λ0 is an eigenvalue of L. If X is finite-dimensional and λ0=0 is a bifurcation point, then 0 is an eigenvalue of L. From Example 1.1 a real eigenvalue of L need not be a be bifurcation point.

    Multiplicities. The geometric multiplicity of an eigenvalue λ0 of L is the dimension of the eigenspace ker(λ0IL), and its algebraic multiplicity is the dimension of kNker(λ0IL)k. When the algebraic multiplicity is one, and either λ00 or X is finite-dimensional, λ0 is called simple. Both the multiplicities of all non-zero eigenvalues of compact operators are finite.

    Bifurcation results by classical analysis. Many seemingly different bifurcation phenomena were studied in ad hoc situations before being recognised by Crandall & Rabinowitz [6] as special cases of the following overarching result, which is a corollary of the Implicit Function Theorem.

    Theorem 1.2. Bifurcation from a simple eigenvalue [6] Suppose that λ0 is a simple eigenvalue of L, that R:R×XX is continuously differentiable, and that xλR exists and is continuous in a neighbourhood of (λ0,0). Then there is an injective, continuous function γ:(1,1)R×X such that γ(0)=(λ0,0) and a neighbourhood U of (λ0,0) such that US={γ(s):s(1,0)(0,1)}. If xxR is also continuous in the neighbourhood of (λ0,0), then γ is C1.

    Remark. In Example 1.1 the nonlinearity R satisfies the hypotheses in Theorem 1.2, and the only eigenvalue of L is λ0=1 which has geometric multiplicity 1, but λ0=0 is not a bifurcation point. However Theorem 1.2 does not apply because the algebraic multiplicity of λ0 is 2. Henceforth the word multiplicity always refers to algebraic multiplicity.

    Bifurcation results by topological methods. In his D.Sc thesis (Kiev 1950 [16]), Krasnosel'skii used Leray-Schauder degree theory to prove, under the hypotheses of (1.1), that every non-zero eigenvalue of L with odd multiplicity is a bifurcation point [15,Thm. 2], [17,Ch. Ⅳ,Thm. 2.1]. Then, in 1971, Rabinowitz reached the ground-breaking conclusion that this bifurcation is not local: indeed, under Krasnosel'skii's hypotheses, he showed a global connected set of non-trivial solutions bifurcates in R×X from T at an eigenvalue of odd multiplicity.

    Theorem 1.3. Global bifurcation at eigenvalues of odd multiplicity. [18,Thm. 1.3]. Suppose L and R are as in (1.1) and λ0 is a non-zero eigenvalue of L of odd multiplicity. Then λ0 is a bifurcation point and there exists a connected subset C of S such that (λ0,0)¯C and either C is unbounded in R×X or there exists (λ1,0)¯C where λ1λ0 is also an eigenvalue of odd multiplicity of L. (If X is finite-dimensional, the result holds when λ0=0 is an eigenvalue of odd multiplicity.)

    Related Results. Dancer [9] used topological obstruction arguments to extend the topological account of bifurcation at simple eigenvalues and eigenvalues of geometric multiplicity one. Krasnosel'skii [17,Ch. Ⅳ.5,p.232 ff.] also studied bifurcation at eigenvalues of even multiplicity, when the nonlinearity R is non-degenerate in a certain sense. Under his hypotheses, the methods introduced by Rabinowitz [18] lead easily to global bifurcation at eigenvalues of even multiplicity for certain nonlinearities [21]. The methods of Section 2 are relevant in that context also.

    Bifurcation results by variational methods. To justify linearisation without reference to multiplicity of eigenvalues, Krasnoselskii [17,§Ⅵ] developed a variational approach to bifurcation theory in Hilbert space. (For an up-to-date account in Banach spaces, see [19].) Let X be a real Hilbert space with inner product , and let h:XR be differentiable with derivative Dh[x]:XR at xX. Then Dh[x] is a bounded linear operator on X and, by the Riesz Representation Theorem, there exists a unique xX such Dh[x]y=x,y for all yX. Hence h(x)=x defines an operator h:XX, called the gradient of h, and an operator H:XX is said to have gradient structure if H=h for some differentiable h:XR.

    It is easily seen that a bounded linear operator L:XX has gradient structure if and only if Lx,y=x,Ly for all x,yX. In other words L is a gradient if and only if it is self adjoint, in which case Lx=(x) where (x)=12Lx,x,xX. Note that when L is self-adjoint, (LλI)2x=0 implies

    (LλI)x2=(LλI)x,(LλI)x=(LλI)2x,x=0,

    and hence algebraic multiplicity and geometric multiplicity of eigenvalues coincide for self-adjoint operators. Obviously the identity operator I on X has gradient structure I=ι where ι(x)=12x2. Finally, a function g:XR is weakly continuous if g(xk)g(x0) in R for every weakly convergent sequence xkx0 in X. Vainberg proved [22,Thm. 8.2] that g is weakly continuous when its gradient is a compact operator. In this setting, a special case of (1.1a) is

    λx=Lx+R(x),λR,xX, (1.2a)

    where X is a real Hilbert space, L:XX is self-adjoint, and R satisfies (1.1b) with gradient structure independent of λ:

    R(x)=r(x), where r is weakly continuous. (1.2b)

    Krasnosel'skii proved [17,Ch. Ⅵ,§6,Thm. 2.2,p. 332] that for this class of problems bifurcation occurs at all non-zero eigenvalues of L, independent of multiplicity. The following version of his theorem replaces his hypothesis that "R is uniform differentiable" near 0 with Vainberg's characterisation [22,Thm. 4.2,p. 45] of uniform differentiability in terms of the bounded uniform continuity of its Fréchet derivative.

    Theorem 1.4. Variational theory of bifurcation at any eigenvalue. If R in (1.2) has Fréchet derivative xdR[x] bounded and uniformly continuous in a neighbourhood of 0 in X, all eigenvalues λ00 of L are bifurcation points. (When X is finite-dimensional, the condition λ00 is not needed.)

    Remark 1.5. While Rabinowitz's theory of global bifurcation yields connected sets CS bifurcating from T at eigenvalues of odd multiplicity, Böhme's example [3,§6] showed that no such connectedness is guaranteed by Theorem 1.4.

    Bifurcation results by real-analyticity. So far in this summary of bifurcation theory, Theorem 1.2 is the only result which guarantees the existence of a path of non-trivial solutions of equation (1.1), and even then it is localized to a neighbourhood (λ0,0), where λ0 is the bifurcation point. However, in his Ph.D thesis (Cambridge 1972) Dancer [7,8] showed, among many other things, that when the operators in (1.1a) are real-analytic (infinitely differentiable and equal to the sum of the Taylor series in a neighbourhood of every point), there bifurcates from a simple eigenvalue a global path of solutions which, at every point, has a local real-analytic re-parametrization. More precisely, Dancer showed that the global continuum, which by Theorem 1.3 bifurcates from the trivial solutions at a simple eigenvalue, contains a path K={(Λ(s),κ(s)):s[0,)}R×X with the following properties.

    (i) Λ(0)=λ0R, κ(0)=0X and K{(λ0,0)} is a real-analytic curve in a neighbourhood of (λ0,0) [4,Thm. 8.3.1].

    (ii) K is either unbounded or forms a closed loop in R×X.

    (iii) For each s(0,) there exists ρ:(1,1)R (a re-parametrisation) which is continuous, injective, and

    ρ(0)=s and t(Λ(ρ(t)),κ(ρ(t))) is analytic ont(1,1).

    This does not imply that K is locally a smooth curve. (The map σ:(1,1)R2 given by σ(t)=(t2,t3) is real-analytic and its image is two curves forming a cusp.) Nor does it preclude the possibility of secondary bifurcation points on K. In particular, since (Λ,κ):[0,)R×X is not required to be globally injective; self-intersection of K (as in a figure eight) is not ruled out.

    (iv) Secondary bifurcation points and points where the bifurcating branch is not smooth, if any, are isolated.

    Under these hypotheses the real-analytic implicit function theorem [4,§4.5] can be used as in the proof of Theorem 1.2 to obtain a real-analytic curve of solutions which intersects the trivial solutions at (λ0,0). Dancer used the theory of real-analytic varieties to show that this observation has a global extension: when the operators are real-analytic there bifurcates from a simple eigenvalue a global path of solutions which is a real-analytic curve except possibly at a discrete set of points. This path is unique in the sense that it has a pre-determined continuation through secondary bifurcation points, or even through points where it intersects higher-dimensional manifolds of solutions. See [4] for an introductory account.

    The following three examples are designed to illustrate how the situation may differ from Dancer's theory when the hypotheses of Theorems 1.2, 1.3 and 1.4 are satisfied by operators that are infinitely differentiable but not real-analytic. The main conclusion is that the non-trivial solution set S may contain global connected sets while having no path-connected components except singletons. Since no two non-trivial solutions in such a connected set can be joined by a path in S, this possibility has serious implications for applications. The paper ends with a simple criterion for a connected set to contain a path joining two of its points. Thus Theorem 6.5 gives an insight into the lack of path-connectedness in connected sets.

    In the first two examples of (1.1), X=R, L=0, λ0=0 is the only eigenvalue of L and is simple, and R=r:R2R, where r is at least continuously differentiable and satisfies (1.1b). Under these hypotheses (1.1a) has the form

    λx=r(λ,x),(λ,x)R2. (2.1)

    The first example concerns the global connected sets of solutions of (2.1) that, by Theorem 1.3, bifurcate at the simple eigenvalue λ0=0, although Theorem 1.2 does not apply because λxr is not continuous at (0,0).

    Example A. There is a C1-function r:R2R which is infinitely differentiable on R2{(0,0)} for which the non-trivial solution set S of (2.1) has no path-connected components except singletons. However by Theorem 1.3 it has an unbounded global connected set of non-trivial solutions which bifurcates at λ0=0.

    The second example illustrates the possible behaviour of solutions which bifurcate at λ0=0 when simultaneously Theorem 1.2 yields the local bifurcation of a smooth curve, and Theorem 1.3 yields global bifurcation of an unbounded connected set, of non-trivial solutions of (2.1).

    Example B. For an infinitely differentiable function r:R2R let ¯C denote the closure of the connected sets of non-trivial solutions of (2.1) which by Theorems 1.2 and 1.3 bifurcate at λ0=0. In this example ¯C is the union LC+C of three disjoint connected sets: L is the smooth curve {0}×(12,12) and C± are closed, unbounded, connected sets in the first and third quadrants, respectively, (±12,0)C± and all path-connected components of C+C are singletons. (The only non-trivial paths in ¯C are subsets of the closure of L.)

    Although Böhme [3] showed the non-trivial solution set of (1.2) given by Theorem 1.4 when R has gradient structure need not be connected, he did not exclude the possibility of it having non-trivial connected components. The next example shows that in any case all the path-connected components may be singletons.

    Example C. In this example of problem (1.2), X=R2, R=r where r:R2R is infinitely differentiable, and L is the zero operator which has only one eigenvalue, namely 0 with multiplicity 2. Then (1.2a) has the form

    λ(x,y)=r(x,y),(x,y)R2,λR, (2.2)

    and the existence of non-trivial solutions with (λ,(x,y)) near (0,(0,0)) is given by Theorem 1.4. Example C shows, in addition to not forming a connected set, that all path-connected components of the non-trivial solution set may be singletons.

    The construction of these examples depends crucially on classical results of Whitney in analysis and Knaster in point-set topology.

    Theorem. (Whitney) For any closed set GRn there is an infinitely differentiable, globally Lipschitz continuous function h such that G={xRn:h(x)=0}, and all the derivatives of h are zero at every point of G.

    Proof. Let u:[0,)[0,1] be a C-function with

    u(t)=1,t[0,1/2];u(t)(0,1),t(1/2,1);u(t)=0,t[1,).

    For a closed set G, let the open set RnG be the union of a countable collection of open balls {Brj(aj):jN}, with radius rj(0,1) centred at ajRn, and put

    uj(x)=u(|xaj|rj),xRn.

    Then uj is infinitely differentiable on Rn and positive on Brj(aj). Now, see [10,§2.7], let Link(Rn,R) denote the linear space of all k-linear maps from (Rn)kR and let Ak denote the norm of A Link(Rn,R). Then Dkuj(x) Link(Rn,R) where Dkuj(x) is the kth derivative of uj at xRn, and Dkuj(x)=0 for xG and jN. Moreover, since uj is supported on ¯Brj(aj),

    γj=max{Dkuj(x)k:0kj,xRn}< for all jN.

    Therefore, since both series are convergent, uniformly in x, in their respective spaces,

    Dkh(x)=jNDkuj(x)γj2jLink(Rn,R) when h(x)=jNuj(x)γj2jR,xRn,

    and h:Rn[0,) is C, G={xRn:h(x)=0}, and Dkh(x)=0 for all kN and xG.

    A deep result in point-set topology due to Knaster (1922) concerns the possible structure of compact connected sets in metric spaces.

    Definition. A continuum, which is a compact, connected set in a metric space, is indecomposable if it is not a union of two proper sub-continua, and hereditarily indecomposable if every sub-continuum is indecomposable. (See [2,5,12,13].)

    Theorem. (Knaster) [14] In R2 there exists a hereditarily indecomposable Q.

    Remark 3.1. Since a non-trivial path in Q would be a decomposable sub-continuum, there are no non-trivial paths in Q. In other words, although Q is compact and connected in R2, all its path-connected components are singletons.

    A hereditarily indecomposable continuum which is snake-like (Definition 6.4) is called a pseudo-arc and all pseudo-arcs are homeomorphic [1,Thm. 1]. Since, by construction, Knaster's Q is snake-like, it is in this sense the unique pseudo-arc. But all that is important here is that Q is compact, connected and has no paths.

    Preliminaries. Let Q be a pseudo-arc and without loss of generality suppose

    Q[0,π]×[14,14],Q({0}×[14,14]) and Q({π}×[14,14]).

    Now let P={(λ,xsinλ)R2:(λ,x)Q}. Then P[0,π]×[14,14],

    P({0}×[14,14])={(0,0)},P({π}×[14,14])={(π,0)},

    and P is a connected set (being the continuous image of a connected set) which contains no non-constant paths (since Q is hereditarily indecomposable).

    Since P is connected, by Proposition 6.1 and Corollary 6.3, for any ϵ>0 there exists an ordered set, {pϵi:1inϵ}P such that

    pϵ1=(0,0),pϵnϵ=(π,0),andpϵipϵi+1<ϵ for all 1inϵ1,

    and the union Lϵ, of the straight line segments which join the points in order, is a continuous, piecewise-linear, non-self-intersecting path joining (0,0) to (π,0). Now define subsets of R2 by

    Pk=P+(kπ,0),Lϵk=Lϵ+(kπ,0),˜P=kZPk,˜Lϵ=kZLϵk, (3.1)

    and note that ˜Lϵ is an unbounded, piecewise linear, connected set which separates the plane, and each point of ˜Lϵ is within distance ϵ of a point of ˜P.

    Now let ˜P±c denote the connected components of R2˜P which contain the half spaces {(λ,x):λR,±x>14}, respectively.

    Lemma 3.2. In the plane, ˜PR×[14,14] is an unbounded, connected subset of a double cone centred on the horizontal axis with opening angle θ<π/6. Moreover ˜P contains no non-trivial paths, (0,0)˜P, and ˜P+c˜Pc=.

    Proof. From the definition, (0,0)˜P and ˜PR×[14,14] is unbounded. Since P={(λ,xsinλ)R2:(λ,x)Q} and |x|14, ˜P lies in a cone with opening angle less than 2arctan(14)<π/6. Moreover ˜P is connected because Pk is connected and PkPk+1={((k+1)π,0)} for all k. Since each Pk contains no paths, a non-trivial path in ˜P must contain points (λi,xi) with λi in the open intervals (kiπ,(ki+1)π), i=1,2, where k1k2. However, this implies that these Pki contain non-trivial paths, which is false. Hence ˜P contains no non-trivial paths.

    Now suppose ˜P+c˜Pc. Then, since ˜P+c˜Pc is open and connected, it is path-connected. Therefore there exists a path γ˜P+c˜Pc joining (0,2) to (0,2) with, since γ is continuous, γ[0,1][K,K]×[K,K] for some K>0. Since, for all ϵ>0, ˜Lϵ in (3.1) separates the plane, there exists

    qϵγ˜Lϵ[K,K]×[K,K], and pϵ˜P with pϵqϵ<ϵ.

    Therefore, by compactness, for a sequence 0<ϵj0,

    qϵjq0γ˜P,

    which is false since γ˜P+c˜Pc. Hence ˜P+c˜Pc=.

    General considerations. For 0<α<β<, let

    C(α,β)={(λ,x):0<αλ<x<βλ or 0>αλ>x>βλ}{(0,0)},

    a double cone in the first and third quadrants. Then there exists ω:R2R with the following properties:

    (a) ω(λ,x)=0 if |x|α|λ|/2, in particular, ω=0 on C(α,β);

    (b) λω(λ,x)0 on R2;

    (c) ω(λ,0)=λ,λR;

    (d) ω is infinitely differentiable on R2{(0,0)};

    (e) ω is globally Lipschitz continuous R2.

    To see this, let ϖ:RR be an infinitely differentiable even function which is non-increasing on [0,) with ϖ(0)=1 and ϖ(r)=0 for all rα/2. Then, for xR, let

    ω(λ,x)=λϖ(xλ),λ0,ω(0,x)=0.

    That ω satisfies (a)-(d) follows immediately from the definition and the properties of ϖ. Moreover, the partial derivatives at (λ,x) are

    xω(λ,x)=ϖ(xλ),λω(λ,x)=ϖ(xλ)(xλ)ϖ(xλ),λ0, (4.1a)
    xω(0,x)=λω(0,x)=0 when λ=0 and x0, (4.1b)
    xω(0,0)=0,λω(0,0)=1, (4.1c)

    since ω(λ,x)=0 when |λ||2|x|/α. For future reference note that

    xλ(xω(λ,x))=ϖ(xλ)(xλ)ϖ(xλ)(xλ)2ϖ(xλ),λ0. (4.1d)

    Since ϖ(r)=0,|r|α/2, the partial derivatives of ω are uniformly bounded in R2{(0,0)}, and property (e) follows.

    Remark 4.1. It follows from (4.1b) and (4.1c) that λω is not continuous at (0,0) and from (4.1a) and (4.1c) that xω is not continuous at (0,0). However, (λ,x)xω(λ,x) is continuously differentiable on R2. Note also, from the intermediate value theorem, that for any ρ(0,1) there exists s(0,α/2) such that ϖ(s)sϖ(s)s2ϖ(s)=ρ and hence, from (4.1d) with x=sλ, λ0, that

    xλ(xω(λ,x))|(λ,sλ)=λx(xω(λ,x))|(λ,sλ)ρ and (λ,sλ)(0,0), as λ0.

    Thus, although λx(xω)(0,0)=xλ(xω)(0,0)=0, the mixed partial derivatives λx(xω(λ,x)) and xλ(xω(λ,x)) are not continuous at (0,0).

    Let D+ and D denote the two disjoint components of the complement of C(α,β) which contain the positive and negative λ-axes respectively.

    Definition (H). A set GC(α,β) satisfies hypothesis (H) if it is closed and connected, its intersections with both half planes, {λ0} and {λ0} are unbounded, and H+H=, where H± are the connected components of R2G with D±H±.

    Lemma 4.2. If G satisfies (H), then ω0 on H+ and ω0 on H

    Proof. This is immediate from properties (a) and (b) of ω.

    Lemma 4.3. When G satisfies (H) there is a locally Lipschitz continuous function g:R2R which is infinitely differentiable on R2{(0,0)}, with the property that g(λ,0)=λ for all λR, and g(λ,x)=0if and only if(λ,x)G.

    Proof. Since G is closed, by Whitney's lemma there exists a non-negative, infinitely differentiable function h:R2[0,) such that h(λ,x)=0 if and only if (λ,x)G, and every derivative of h is zero at every point of G. Let ˆh:R2R be defined by

    ˆh(λ,x)={h(λ,x),(λ,x)Hh(λ,x),otherwise}, with H± defined in Definition (H).

    In particular, ˆh(λ,x)=±h(λ,x), (λ,x)H±, ˆh is infinitely differentiable on R2, and ˆh(λ,x)=0 if and only if (λ,x)G. Now with ω satisfying (a)-(e) above, let

    g(λ,x)=x2ˆh(λ,x)+ω(λ,x).

    It follows from (4.1) that g is infinitely differentiable on R2{(0,0)} and by Lemma 4.2 g satisfies the conclusions of the Lemma.

    Proposition 4.4. For G satisfying (H), there is a continuously differentiable function r:R2R with rC(R2{(0,0)}), such that |r(λ,x)|/|x|0as0|x|0 uniformly for λ in bounded intervals, and G{(0,0)} is the set of non-trivial solutions of λx=r(λ,x).

    Proof. For G satisfying (H) and the corresponding function g in Lemma 4.3, let

    r(λ,x)=x(λg(λ,x)),(λ,x)R2.

    Then the smoothness of ˆh and the properties of ω in (4.1) imply that g is infinitely differentiable on R2{(0,0)} and, by Remark 4.1, xg is continuously differentiable on R2 with |r(λ,x)|/|x|0 as 0|x|0 uniformly for λ in bounded intervals. Moreover, by construction, non-trivial solutions of (2.1) are the zeros of g with x0. So, by Lemma 4.3, G{(0,0)} is the set of non-trivial solutions of λx=r(λ,x) in R2. This completes the proof.

    Remark. Since, from Remark 4.1, the mixed partial derivative λxr is not continuous at (0,0), Theorem 1.2 does not apply to equation λx=r(λ,x) in this situation.

    Construction of Example A of (2.1). Let ˜P be the unbounded connected set defined in (3.1) and let G denote ˜P rotated counter-clockwise about the origin through an angle π/4. By Lemma 3.2, G is connected, contains no non-trivial paths, and satisfies (H) with α=tan(π/6) and β=tan(π/3). With this choice of G, Example A is a special case of Proposition 4.4.

    Construction of Example B of (2.1). This example shows that the global connected set C given by Theorem 1.3 need not be path connected even when all the operators involved are infinitely differentiable on R2 and locally, by Theorem 1.2, a smooth curve of solutions bifurcates from the trivial solutions at a simple eigenvalue.

    With ˜P defined in (3.1) let three disjoint connected sets be defined by

    L={0}×(1/2,1/2),C+=(0,1/2)+(˜P([0,)×R))[0,)×[1/4,3/4],C=(0,1/2)+(˜P((,0]×R))(,0]×[3/4,1/4].

    Then L is the smooth curve {0}×(12,12), and C± are closed, unbounded, connected sets in the first and third quadrants respectively with (0,±12)C± and all path-connected components of C+C are singletons. Let ¯C be their union

    ¯C=LC+C.

    Now let E and E+ denote the connected components of R2¯C which contains (,0]×{0} and [0,)×{0}, respectively and note, from the argument for Lemma 3.2, that E+E=. By Whitney's result there exists a non-negative, infinitely differentiable function h on R2 which is zero only on the closed set ¯C, and at each point of ¯C all the derivatives of h are zero. Let

    ˜h(λ,x)={h(λ,x),(λ,x)Eh(λ,x),otherwise},

    so that ˜h0 on E+.

    Now let ˜ω:RR be an infinitely differentiable even function with ˜ω(0)=1, ˜ω decreasing on [0,1/4] and ˜ω(x)=0 when |x|1/4, let ˜g(λ,x)=x2˜h(λ,x)+λ˜ω(x). Finally let r(λ,x)=x(λ˜g(λ,x)). Then the set of non-trivial solutions of λx=r(λ,x) coincide with the non-trivial solution set of ˜g(λ,x)=0 which is the set ¯C{(0,0)}. This completes the justification of Example B.

    Example C is a simplified version of Böhme's example [3] with added structure to ensure that all path-connected sets of non-trivial solutions are singletons.

    According to Bing [1,Ex. 2,p. 48] there exists a hereditarily indecomposable continuum, H say, which separates the plane. Let Ω be a non-empty bounded component of R2H and Ω its boundary. Then ΩH, since points which are not in H (which is closed) are interior points of their connected component in R2H.

    Without loss of generality, suppose that in the (ς,τ)-plane

    Ω[14π,14π]×[a,a] and ¯Ω[14π,14π]×{±a},a>0. (5.1)

    Denote by S the strip [π,π]×R and, with a<p<2a, consider two parallel columns of copies of Ω, arranged periodically with period 2p in the τ direction, centred on the lines ς=±π/2, and with height 2a, as illustrated in Figure 1. The copies of Ω in the right column are translates through (π,p) of those on the left. (Apart from being open, connected and satisfying (5.1), nothing is known about the shape of Ω, so the diagram is for illustration only.) Let ˆΩ denote the union of all the copies of Ω in this arrangement. The key to what follows is the property of ˆΩ that, for all τR, the set {ς:(ς,τ)ˆΩ} has strictly positive measure.

    Figure 1.  Schematic diagram of ˆΩ.

    Now by Whitney's result there exists ψ:R2R which is infinitely differentiable, ψ>0 on R2ˆΩ, and ψ and all its derivatives are zero on ˆΩ. There is no loss of generality in assuming that ψ is 2p-periodic in τ and equals 1 in the two strips [7π/8,π]×R and [π,7π/8]×R. Now, for (ς,τ)S, let

    ψ(ς,τ)=ψ(ς,τ) when (ς,τ)ˆΩ, and ψ(ς,τ)=0 otherwise,ψ+(ς,τ)=ψ(ς,τ) when (ς,τ)SˆΩ, and ψ+(ς,τ)=0 otherwise.

    Next define infinitely differentiable functions κ± which are 2p-periodic in τR by

    κ±(τ)=ππψ±(ς,τ)dς,τR,

    where κ(τ)<0<κ+(τ),τR, and let

    φ(ς,τ)=κ+(τ)ψ(ς,τ)κ(τ)ψ+(ς,τ). (5.2)

    Then φ(ς,τ)=κ(τ)>0 when |ςπ|<π/8, φ is infinitely differentiable, ˆΩ is the zero set of φ, and by (5.2)

    ππφ(ς,τ)dς=0 for all τR. (5.3)

    If Φ:SR is defined by

    Φ(ς,τ)=ςπφ(s,τ)ds,(ς,τ)S, (5.4)

    then by (5.3), for τR,

    Φ(π,τ)=Φ(π,τ)=0,Φς(ς,τ)=κ(τ),|ςπ|<π/8, and kΦςk(π,τ)=kΦςk(π,τ)=0 for all k2.

    With this in mind, an infinitely differentiable function r:R2R can be defined by putting r(0,0)=0 and, for (x,y)=ρ(cosϑ,sinϑ) in polar coordinates, let

    r(x,y)=ˆr(ρ,ϑ):=exp(1ρ2)Φ(ϑ,1ρ),ρ>0,ϑ[π,π]. (5.5)

    Then, since (2.2) is of the form

    (12λ(x,y)2r(x,y))=0,

    its non-trivial solutions satisfy

    ρ(12λρ2ˆr(ρ,ϑ))=0,ϑ(12λρ2ˆr(ρ,ϑ))=0ρ>0,ϑ[π,π].

    By (5.4) and (5.5), the second equation implies that

    φ(ϑ,1ρ)=0,ρ>0, which means that (ϑ,1ρ)ˆΩ. (5.6)

    Therefore, from (5.6) and the construction of ˆΩ, it follows that in this example any non-trivial solutions (λ,(x,y))R×R2 of (2.2) has

    (x,y)=ρ(cosϑ,sinϑ)) where (ϑ,1ρ)ˆΩ.

    Since ˆΩ is the union of an infinite set of disjoint translates of Ω, see (5.1), and since ΩH which is a hereditarily indecomposable continuum, all path-connected components of the non-trivial solution set of (2.2) are singletons. However, non-trivial solutions of (2.2) with (λ,(x,y)) near to (0,(0,0)) exist, by Theorem 1.4.

    Proposition 6.1. In a metric space (M,d) let G={Gα:αA} be an open cover of a connected set A. Then for x,yA there is a set {Gα1,,Gαn}G with

    xGα1,yGαnandGαiGαjif and only if|ij|1. (6.1)

    Proof. Fix xA and let BA be the set of yA such that (6.1) holds for an ordered finite subset of G. Then B because xB, and if yB then by (6.1) zB for all zGαnA. So B is open in A. Now suppose z is in the closure of B in A. Then, since G covers A, there exists GG such that zG, and there exists yB with yG. Since yBG, there exists {G˜α1,,G˜αm}G such that (6.1) holds. Let k be the smallest element of {1,,m} for which G˜αkG. Then {G˜αj,1jk}{G} satisfies (6.1) with z instead of y. So zB, whence B is both closed and open in A. Hence B=A since A is connected and B.

    Corollary 6.2. For ϵ>0 and x,yA, where A is connected in (M,d), there is a set {x1,,xn}A with

    x1=x,xn=y,Bϵ(xi)Bϵ(xj)if and only if|ij|1,1i,jn, (6.2)

    where Bϵ(a)M is the open ball with radius ϵ centred at a.

    Proof. For given ϵ>0 and x,yA, by Proposition 6.1 with G={Bϵ(a):aA}, there exist {aj:1jp}A with

    xBϵ(a1),yBϵ(ap) and Bϵ(ai)Bϵ(aj) if and only if |ij|1.

    Let q=max{j1:Bϵ(x)Bϵ(aj)} and put

    y1=x,y2=aq,yj=aj+q2, for 2jr where r=pq+2.

    Then Bϵ(yi)Bϵ(yj) if and only |ij|1, y1=x and yBϵ(yr).

    Now let n=m+1 where m=min{jr:Bϵ(y)Bϵ(yj)}, and put xi=yi,1in1, xn=y, to achieve the required result.

    Corollary 6.3. When (M,d) is a normed linear space, let Li={txi+(1t)xi+1:t[0,1]}, 1in1, be straight line segments joining the centres of consecutive balls in Corollary 6.2. Then

    LiLi+1={xi+1},1in1,LiLj=,i+1<jn1,1in2.

    Consequently, L:=n1i=1Li is a continuous, piecewise-linear, non-self-intersecting path joining x to y.

    Proof. First suppose that zLiLi+1 and zxi+1. Then

    z=(1s)xi+1+sxi+2=txi+(1t)xi+1,s,t(0,1],

    whence t(xixi+1)=s(xi+2xi+1). So st because xixi+2. If s<t,

    2ϵxixi+2=(1(s/t))xi+2xi+1<2ϵ,

    a contradiction, and if t<s,

    2ϵxixi+2=(1(t/s))xixi+1<2ϵ,

    which is also false. This proves that LiLi+1={xi+1} for all i.

    Suppose zLiLj for i1 and i+1<jn1. Then, by (6.2),

    xixi+1<2ϵ,xjxj+1<2ϵ,xixj2ϵ,xi+1xj+12ϵ,

    and

    z=sxi+(1s)xi+1=txj+(1t)xj+1,s,t[0,1],=(1s)xi+sxi+1=(1t)xj+txj+1,s=1s,t=1t.

    Therefore xi+1+s(xixi+1)=xj+1+t(xjxj+1), which implies

    2ϵxi+1xj+1sxixi+1+txjxj+1<2ϵ(s+t),

    and hence s+t>1. Also xixj=s(xixi+1)+t(xj+1xj) and hence

    2ϵxixjsxixi+1+txj+1xj<2ϵ(s+t),

    from which it follows that s+t>1, equivalently, s+t<1, which is a contradiction. Since different line segments Li joining centres of balls do not intersect, their union L is a continuous, piecewise-linear, non-self-intersecting path joining x1 to xn.

    Definition 6.4. In a metric space a linear chain G is an ordered, finite collection of open sets with GiGj if and only if |ij|1. The Gi, which may not be connected, are the links of G and an ϵ-linear chain is a linear chain with links of diameter less that ϵ. If, for all ϵ>0, a set A can be covered by an ϵ-linear chain, A is said to be snake-like. A snake-like hereditarily indecomposable compact connected set is called a pseudo-arc.

    A criterion for the existence of paths in connected sets in Banach spaces. Suppose in Corollaries 6.2 and 6.3 that the metric space (M,d) is a Banach space (V,), that AV is closed and connected, and that closed bounded subsets of A are compact. For fixed xyA and any ϵ>0 let

    ϵ:=inf{n1i=1xi+1xi, where x1,,xn satisfies (6.2)}xy>0.

    Theorem 6.5. If ϵ is bounded as ϵ0, there is a path in A joining x to y.

    The proof depends on infinite-dimensional versions of two well-known theorems.

    Theorem 6.6. [20,p. 179] (Ascoli-Arzelà) When X is a separable topological space, Y is a metric space and {hk} is an equi-continuous sequence of functions from X to Y with the property that the closure of {hk(x):kN} is compact in Y for each xX, there is a subsequence {hkj} and a continuous function h such that hkj(x)h(x) pointwise, and uniformly on every compact subset of X.

    Theorem 6.7. [11,Ch.Ⅴ §2.6] (Mazur) In a Banach space the closed convex hull ¯co(K) of a compact set K is compact.

    Proof of Theorem 6.5. Let 0<ϵk0 and for each k let Lϵk be a piecewise linear, non-self-intersecting path in V joining x to y, as in Corollary 6.3, with length γk where

    xϵkixγk=n1i=1xϵki+1xϵkiϵk+ϵk,xϵkiA.

    Since, for all k, the xϵkis are in A which is closed, and since xϵkix is bounded independent of i and k, there is a closed bounded subset A of A with xϵkiA for all i and k. By hypothesis, A is compact. Therefore, since each Lϵk is a union of straight-lines joining points of A, the paths Lϵk¯co(A), where ¯co(A) is compact in the Banach space V by Mazur's Theorem 6.7. Moreover, by hypothesis each Lϵk is rectifiable with length γk bounded above independent of k. Since Lϵk is piecewise linear it can be parameterised by arc-length s, Lϵk={fk(s):s[0,γk]} say, where fk:[0,γk]¯co(A)V, fk(s)=1 almost everywhere, fk(0)=x and fk(γk)=y.

    Now let hk(t)=fk(γkt),t[0,1], so that Lϵk={hk(t):t[0,1]}¯co(A) and {hk:kN} is uniformly bounded and equi-continuous on [0,1], because hk(t)=γk for almost all t[0,1] and γk is bounded.

    Since [0,1]R and ¯co(A)V are both compact, it follows from the Ascoli-Arzelà Theorem 6.6, with X=[0,1] and Y=¯co(A), that a subsequence {hkj} converges uniformly on [0,1] to a continuous h:[0,1]¯co(A). Since, for s[0,1] and jN, there exists xϵkjiA with xϵkjihkj(s)ϵkj,

    dist (h(s),A)h(s)hkj(s)+dist (hkj(s),A)h(s)hkj(s)+xϵkjihkj(s)h(s)hkj(s)+ϵkj0 as j.

    Since A is closed, h:[0,1]AA and h is continuous. Finally note that x=hkj(0)h(0), y=hkj(1)h(1). So h defines a path in A joining x to y.



    [1] Concerning hereditarily indecomposable continua. Pacific J. Math. (1951) 1: 43-51.
    [2] Snake-like continua. Duke Math. J. (1951) 18: 653-663.
    [3] Die Lösung der Verzweigungsgleichungen für nichtlineare Eigenwertprobleme. Math. Zeit. (1972) 127: 105-126.
    [4] (2003) Analytic Theory of Global Bifurcation,. Princeton and Oxford: Princeton University Press.
    [5] Chainable continua and indecomposability. Pacific J. Math. (1959) 9: 653-659.
    [6] Bifurcation from simple eigenvalues. J. Funct. Anal. (1971) 8: 321-340.
    [7] Bifurcation theory for analytic operators. Proc. Lond. Math. Soc. (1973) 26: 359-384.
    [8] Global structure of the solutions of non-linear real analytic eigenvalue problems. Proc. Lond. Math. Soc. (1973) 27: 747-765.
    [9] Bifurcation from simple eigenvalues and eigenvalues of geometric multiplicity one. Bull. Lond. Math. Soc. (2002) 34: 533-538.
    [10] J. J. Duistermaat and J. A. C. Kolk, Multidimensional Real Analysis I., Cambridge Studies in Advanced Mathematics 86. Cambridge University Press, Cambridge, 2004. doi: 10.1017/CBO9780511616716
    [11] N. Dunford and J. T. Schwartz, Linear Operators, Part Ⅰ: General Theory, , With the assistance of W. G. Bade and R. G. Bartle. Pure and Applied Mathematics, Vol. 7 Interscience Publishers, Inc., New York; Interscience Publishers, Ltd., London, 1958.
    [12] The pseudo-circle is unique,. Trans. Amer. Math. Soc. (1970) 149: 45-64.
    [13] A brief historical view of continuum theory. Topology Appl. (2006) 153: 1530-1539.
    [14] Un continu dont tout sous-continu est indécomposable. Fund. Math. (1922) 3: 247-286.
    [15] On a topological method in the problem of eigenfunctions of nonlinear operators. Dokl. Akad. Nauk SSSR (N.S.) (1950) 74: 5-7.
    [16] Some problems in nonlinear analysis. Amer. Math. Soc. Trans. Ser. (1958) 10: 345-409.
    [17] . A. Krasnolsel'skii, Topological Methods in the Theory of Nonlinear Eigenvalue Problems, Pergamon Press, Oxford, 1963. (Original in Russian: Topologicheskiye Metody v Teorii Nelineinykh Integral'nykh Uravnenii., Gostekhteoretizdat, Moscow, 1956.)
    [18] Some global results for nonlinear eigenvalue problems. J. Funct. Anal. (1971) 7: 487-513.
    [19] Some aspects of nonlinear eigenvalue problems. Rocky Mountain J. Math. (1973) 3: 161-202.
    [20] H. L. Royden, Real Analysis, Third Edition, McMillan, New York, 1988.
    [21] Global bifurcation for k-set contractions without multiplicity assumptions. Quart. J. Math. Oxford Ser. (1976) 27: 199-216.
    [22] M. M. Vainberg, Variational Methods for the Study of Nonlinear Operators, With a chapter on Newton's method by L. V. Kantorovich and G. P. Akilov. Translated and supplemented by Amiel Feinstein Holden-Day, Inc., San Francisco, Calif.-London-Amsterdam, 1964.
  • This article has been cited by:

    1. Vladimir Kozlov, On the first subharmonic bifurcations in a branch of Stokes water waves, 2024, 379, 00220396, 676, 10.1016/j.jde.2023.10.038
  • Reader Comments
  • © 2021 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2061) PDF downloads(129) Cited by(1)

Figures and Tables

Figures(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog