Home Global structure of positive and sign-changing periodic solutions for the equations with Minkowski-curvature operator
Article Open Access

Global structure of positive and sign-changing periodic solutions for the equations with Minkowski-curvature operator

  • Ruyun Ma EMAIL logo , Zhongzi Zhao EMAIL logo and Xiaoxiao Su
Published/Copyright: March 29, 2024

Abstract

We show the existence of unbounded connected components of 2π-periodic positive solutions for the equations with one-dimensional Minkowski-curvature operator u 1 u 2 = λ a ( x ) f ( u , u ) , x R , where λ > 0 is a parameter, a C ( R , R ) is a 2π-periodic sign-changing function with 0 2 π a ( x ) d x < 0 , f C ( R × R , R ) satisfies a generalized regular-oscillation condition. Moreover, for the special case that f does not contain derivative term, we also investigate the global structure of 2π-periodic odd/even sign-changing solutions set under some parity conditions. The proof of our main results are based upon bifurcation techniques.

MR(2010): 34B15; 34C23; 34C25

1 Introduction

In this paper, we investigate the global structure of 2π-periodic solutions set for

(1.1) u 1 u 2 = λ a ( x ) f ( u , u ) , x R ,

where a C ( R , R ) is a 2π-periodic function, f C ( R × R , R ) . Equation (1.1) is associated with the mean curvature operator in the flat Minkowski space

L N + 1 = { ( x , t ) : x R N , t R }

endowed with the Lorentzian metric

i = 1 N = ( d x i ) 2 ( d t ) 2 ,

where (x, t) are the canonical coordinates in R N + 1 . The initial motivation for the study of equation

(1.2) M u div u 1 | u | 2 = λ g ( x , u ) in Ω

comes from the fact that the relativistic mean curvature operator is an essential object in Geometry and Physics, where Ω R N ( N 1 ) is an open, bounded domain with smooth boundary Ω. More precisely, M appears naturally in the Riemannian Geometry-where it is involved in the determination of the maximal or constant mean curvature hypersurfaces in the Lorentz-Minkowski space (see, e.g., Cheng and Yau [1], Bartnik and Simon [2], Kiessling [3], Corsato et al. [4]).

Existence and multiplicity of solutions for (1.2) with Dirichlet boundary condition have been extensively studied by many authors, see Bereanu, Jebelean and Mawhin [5], Ma, Gao and Lu [6] and the references therein. All spacelike solutions of (1.2) with Dirichlet boundary condition are uniformly bounded by the quantity 1 2 d ( Ω ) , with d(Ω) the diameter of Ω, see Coelho et al. [7]–[9].

However, there are two new difficulties had to be faced in studying the positive periodic solutions of (1.1):

  1. the set of positive periodic solutions may be unbounded;

  2. if λ = 0 in (1.1), any constant c > 0 is a positive periodic solution.

Recently, Boscaggin et al. [10], [11] investigated the existence of positive periodic solutions to the equation

(1.3) u 1 u 2 + λ a ( x ) g ( u ) = 0 .

Based on the Mawhin’s coincidence degree theory, they obtained the following

Theorem A.

(See [10, Theorem 3.4].) Assume that

(a*) a is a T-periodic and locally integrable function with 0 T a ( x ) d x < 0 ;

(a *) there exist m ≥ 1 intervals I 1 + , , I m + , closed and pairwise disjoint in the quotient space R / T Z , such that

a ( x ) 0 , for a.e. x I i + , a ( x ) 0 on I i + , for i = 1 , , m ,

a ( x ) 0 , for a.e. x ( R / T Z ) \ i = 1 m I i + ;

( g * ) g : R + R + is a continuous function, g(0) = 0, g(s) > 0 for s > 0, and lim sup s 0 + g ( s ) s < + ;

( g ) lim s + , w 1 g ( w s ) g ( s ) = 1 .

Then, there exists λ * > 0 such that for every λ > λ *, there exists at least a positive T-periodic solution to (1.3).

Remark 1.1

  1. Condition (a*) guarantees that the periodic linear eigenvalue problem associated with (1.3) has a positive principal eigenvalue λ 1 > 0 in addition to λ 0 = 0.

  2. In [10], (g ) is referred to as a regular-oscillation condition: for instance, the function g(u) = u p satisfies (g ) for every p > 1. Condition (g ) guarantees that the connected component branching from (λ 0, 0) and the one branching from (λ 1, 0) are separate.

  3. Theorem A provides no any information about the existence of positive T-periodic solutions for λλ *.

  4. Theorem A cannot deal with the case of m = ∞ in (a *).

  5. Theorem A cannot deal with the case where g in (1.3) contains the derivative term.

The purpose of this paper is to investigate the global structure of positive 2π-periodic solutions set for (1.1), where m in (a *) is allowed to be equal to infinity and nonlinearity may depend on u′.

It is worth underlining that the solvability of the equations with derivative dependence of the form (1.1) has been explicitly raised as an open question in Mawhin [12]. On the other hand, the study of global behavior of the positive periodic solutions set is very useful for computing the numerical solutions of (1.1) as it can be used to guide the numerical work.

Assume that

  1. a C ( R , R ) is a 2π-periodic sign-changing function that satisfies 0 2 π a ( x ) d x < 0 ;

  2. f(0, p) = 0 for any p ∈ [−1, 1], f(s, p) > 0 for s > 0 and p ∈ [−1, 1];

  3. lim s 0 + f ( s , p ) s f 0 ( 0 , + ) uniformly for p ∈ [−1, 1];

  4. lim ( s , ω , ν ) ( + , 1,0 ) f ( ω s , ν s ) f ( s , 0 ) = 1 .

Let Y { u C ( R , R ) u is 2 π periodic } and

H { u C 1 ( R , R ) u is 2 π periodic } .

Then Y and H are the Banach spaces endowed with the norm

u = sup x [ 0,2 π ] | u ( x ) | and u H = sup x [ 0,2 π ] | u ( x ) | + sup x [ 0,2 π ] | u ( x ) | ,

respectively. Let

S + { ( λ , u ) R × H : u satisfies ( 1.1 ) , u 0 and u 0 } ̄ R × H .

Remark 1.2.

If aY and f C ( R × R , R ) , then any solution uC 1[0, 2π] of

(1.4) u 1 u 2 = λ a ( x ) f ( u , u ) , x ( 0,2 π ) , u ( 0 ) = u ( 2 π ) , u ( 0 ) = u ( 2 π )

can be extended to a 2π-periodic function u ̃ defined on the whole of R .

By the same argument in the proof of [13, Theorem 2.1], with obvious changes, we may deduce that

(1.5) u = λ a ( x ) u , x ( 0,2 π ) , u ( 0 ) = u ( 2 π ) , u ( 0 ) = u ( 2 π )

has two simple principal eigenvalues λ 0 = 0 and λ 1 > 0. The eigenfunctions corresponding to λ 0 and λ 1 are a constant and a positive function respectively.

Theorem 1.1.

Assume that (H1)–(H4) hold. Then there exists an unbounded connected component C + S + , such that

  1. C + joins ( λ 1 f 0 , 0 ) with infinity in λ direction;

  2. for any compact interval J ⊂ [0, + ∞), there exists a positive constant γ* < 1 such that for any ( λ , u ) C + ( J × H ) ,

    u γ * ;

  3. Proj R C + = [ λ * , ) for some λ * ( 0 , λ 1 f 0 ] .

Corollary 1.1.

Assume that (H1)–(H4) hold. Then there exists λ * ( 0 , λ 1 f 0 ] such that (1.1) has at least one positive 2π-periodic solution for λ > λ 1 f 0 , while it has no positive 2π-periodic solution for λ ∈ (0, λ *].

Remark 1.3

  1. It is the first time that a generalized regular-oscillation condition, i.e. (H4), has been introduced. A bounded connected component of positive solutions connecting two eigenvalues is called a “mushroom”. López-Gómez and Molina-Meyer [14] and Brown [15] proved the existence of “mushrooms” under appropriate conditions. It is worth underlining that, the generalized regular-oscillation condition guarantees that the “mushroom” can not occur in Theorem 1.1.

  2. (H3) implies that f satisfies a generalized regular-oscillation condition at 0, i.e.,

    lim ( s , ω , ν ) ( 0,1,0 ) f ( ω s , ν s ) f ( s , 0 ) = 1 .

  3. We allow m = ∞ in (a *).

Next, for the special case of (1.1) that f does not contain derivative term, we also study the global structure of 2π-periodic sign-changing solutions set to the equation

(1.6) u 1 u 2 = λ a ( x ) f ( u ) , x R ,

where a C ( R , R ) is a 2π-periodic function and f C ( R , R ) .

If λ k , k ≥ 2, is the eigenvalue of (1.5), then the multiplicity of λ k may be even. For example, if a ≡ 1, then (1.5) has a list of eigenvalues λ k = (k − 1)2, k = 1, 2, 3, …. The eigenspace corresponding to λ k , k ≥ 2, is

M k = span { sin k x , cos k x } , x [ 0,2 π ] .

Therefore, we cannot directly apply Rabinowitz global bifurcation theorem to (1.6), see [16], [17].

In order to overcome the above difficulties, we have to work in

E 1 { u H u ( x ) = u ( x ) , x R }

and

E 2 { u H u ( x ) = u ( x ) , x R } ,

which are the Banach spaces endowed with the norm

u H = sup x [ 0,2 π ] | u ( x ) | + sup x [ 0,2 π ] | u ( x ) | .

This idea of constructing invariant subspaces is similar to the one in Marlin [18] and Coron [19].

We will use the following assumptions

  1. a C ( R , R ) is a 2π-periodic function that satisfies a(−x) = a(x), x R ;

  2. f : R R is an odd continuous function;

  3. s f ( s ) 0 , s R \ { 0 } .

Theorem 1.2.

Assume that (H5)–(H6) hold and

(1.7) lim s 0 f ( s ) s = + .

Then for any k ∈ {2, 3, …}, ν ∈ {+, − },

  1. there exists an unbounded connected component C k ν R × E 1 of 2π-periodic sign-changing solutions of (1.6) such that, C k ν joins (0,0) with infinity in λ direction, and any solution ( λ , u ) C k ν satisfies ‖u < π;

  2. if f satisfies (H7), then there exists an unbounded connected component D k ν R × E 2 of 2π-periodic sign-changing solutions of (1.6) such that, D k ν joins (0,0) with infinity in λ direction, and any solution ( λ , u ) D k ν satisfies ‖u < 2π.

Corollary 1.2.

Assume that (H5)–(H6) and (1.7) hold. Then for any λ ∈ (0, + ∞), k ∈ {2, 3, …}, ν ∈ {+, − },

  1. Equation (1.6) has 2π-periodic odd solution u k ν satisfies that u k + has exactly 2k − 3 zeros in (0, 2π) and is positive near 0+, u k has exactly 2k − 3 zeros in (0, 2π) and is negative near 0+;

  2. if f satisfies (H7), then (1.6) has 2π-periodic even solution v k ν satisfies that v k + has exactly 2k − 2 zeros in (0, 2π) and is positive near 0, v k has exactly 2k − 2 zeros in (0, 2π) and is negative near 0.

Due to f is sublinear at 0, the global bifurcation techniques cannot be used directly in the case. Therefore, we refer to [20] and apply some properties of the superior limit of a certain infinity collection of connected sets.

Remark 1.4.

For other results of boundary value problems where the nonlinearity is odd, see Naito and Tanaka [21], Garcia-Huidobro and Ubilla [22].

Remark 1.5.

Bereanu and Torres [23], using critical point theory, proved that the existence of an infinite number of solutions for the Neumann problem associated to some prescribed mean curvature problems in a FLRW spacetime under some appropriate assumptions. However, they provided no any information about the nodal properties of these solutions.

2 Preliminary results

Let ϕ ( s ) s 1 s 2 and

W w H : max [ 0,2 π ] | w | < 1 and ϕ ( w ) C 1 [ 0,2 π ] .

Obviously, 1 ∈ W. A 2π-periodic solution of (1.1) is a function uW satisfying (1.1).

Proposition 2.1.

Let uW be a 2π-periodic solution of (1.1). Then uC 2[0, 2π].

Proof.

Since ϕ(u′) ∈ C 1[0, 2π], it follows that there exists yC[0, 2π], such that

ϕ ( u ( x ) ) = y ( x ) , x [ 0,2 π ] .

The boundary condition u(0) = u(2π) implies that

u ( x 0 ) = 0 for some x 0 [ 0,2 π ] .

Thus

ϕ ( u ( x ) ) = ϕ ( u ( x 0 ) ) + x 0 x y ( τ ) d τ = x 0 x y ( τ ) d τ , x [ 0,2 π ] ,

and subsequently,

u ( x ) = ψ x 0 x y ( τ ) d τ , x [ 0,2 π ] ,

where ψ = ϕ −1 which can be explicitly given by

ψ ( s ) = s 1 + s 2 .

Let P ( x ) x 0 x y ( τ ) d τ . Obviously ψ ( ) , P ( ) C 1 ( R , R ) . Then

u ( x ) = ψ ( P ( x ) ) P ( x ) , x [ 0,2 π ] .

Therefore, uC 2[0, 2π].□

Rabinowitz [17] considered the equation

(2.1) v = K ( λ ̄ , v ) ,

where λ ̄ R , vX, a real Banach space with the norm ‖ ⋅‖ and K : R × X X is compact and continuous. In addition, K ( λ ̄ , v ) = λ ̄ L v + H ( λ ̄ , v ) , where H ( λ ̄ , v ) is ◦(‖v‖) for v near 0 uniformly on bounded λ ̄ intervals and L is a compact linear map on X. Let μ be a simple characteristic value of L. Let

S { ( λ ̄ , v ) R × X : v satisfies ( 2.1 ) , and v 0 } ̄ R × X .

Lemma 2.1.

(See [17, Theorem 1.3].) If μr(L) is of odd multiplicity, then S possesses a maximal subcontinuum C μ such that (μ, 0) ∈ C μ , and C μ either

  1. meets infinity in R × X , or

  2. meets ( μ ̂ , 0 ) , where μ μ ̂ r ( L ) .

Let l X (the dual of X) be corresponding eigenvectors of L T , the transpose of L. Denote ⟨⋅, ⋅⟩ the duality between X and X′. Let B ϵ denote open balls in R × X of radius ϵ centered at (μ, 0). For ξ , η R , define

K ξ , η = { ( λ ̄ , v ) R × X : | λ ̄ μ | < ξ , | l , v | > η v } .

Lemma 2.2.

(See [17, Lemma 1.24].) There exists a ζ 0 > 0 such that for all ζ < ζ 0, ( C μ { ( μ , 0 ) } ) B ζ K ξ , η . If ( λ ̄ , v ) ( C μ { ( μ , 0 ) } ) B ζ , then v = α v ̄ + w , where |α| > ηv‖, v ̄ is the eigenfunction of L corresponding to μ. Moreover, | λ ̄ μ | = ( 1 ) , w = ◦ (|α|) for α near 0.

Definition 2.1.

(See [24].) Let X be a Banach space and { C n n = 1,2 , } be a family of subsets of X. Then the superior limit D of C n is defined by

D lim sup n C n = x X { n i } N and x n i C n i such that x n i x .

Lemma 2.3.

(See [20, Lemma 2.2].) Let X be a Banach space, and let { C n } be a family of connected subsets of X. Assume that

  1. there exist z n C n , n = 1,2 , , and z* ∈ X, such that z n z*;

  2. lim n r n = , where r n = sup { x : x C n } ;

  3. for every R > 0, n = 1 C n B R is a relatively compact set of X, where

B R = { x X x R } .

Then there exists an unbounded connected component C in D and z * C .

Let us define h : R R by setting

h ( s ) = ( 1 s 2 ) 3 / 2 if | s | 1 , 0 if | s | > 1 ,

and consider the equation

(2.2) u = λ ρ ( x , u , u ) h ( u ) ,

where ρ : R 3 R is assumed to be continuous and 2π-periodic with respect to the first variable.

Lemma 2.4.

(See [25, Proposition 2].) If (λ, u) is a 2π-periodic solution of (2.2), then uH satisfies

u < 1 .

Lemma 2.5.

A function uH is a 2π-periodic solution of

(2.3) u 1 u 2 = λ ρ ( x , u , u )

if and only if uH is a 2π-periodic solution of (2.2).

Proof.

It is clear that a 2π-periodic solution uH of (2.3) is a 2π-periodic solution of (2.2) as well. Conversely, suppose that uH is a 2π-periodic solution of (2.2). Then ‖u′‖ < 1 can be obtained by Lemma 2.4. Consequently, uH is a 2π-periodic solution of (2.3).□

From Remark 1.2, any positive solution uC 1[0, 2π] of

(2.4) u 1 u 2 = λ ρ ( x , u , u ) , x ( 0,2 π ) , u ( 0 ) = u ( 2 π ) , u ( 0 ) = u ( 2 π )

can be extended to a 2π-periodic function u ̃ defined on the whole of R , and function u ̃ is a positive 2π-periodic solution of (2.3).

Lemma 2.6.

If (λ, u) is a solution of (2.4) and u has a double zero, then u ≡ 0.

Proof.

Assume that there exists a 2π-periodic solution ( λ ̃ , u ) , λ ̃ > 0 , of (2.4) and u has a double zero. Let x * ∈ [0, 2π] be a double zero of u. Then it is easy to check that u satisfies the Cauchy problem

u = λ ̃ ρ ( x , u , u ) h ( u ) F ( x , u , u ) , u ( x * ) = u ( x * ) = 0 .

It is easy to deduce that the function F(x, s, p) is locally Lipschitz with respect to x, s and p in some neighbourhood U of (x *, 0, 0),

U { ( x , s , p ) : ( x , s , p ) [ 0,2 π ] × R × ( 1,1 ) , | s | + | p | + | x x * | < δ 0 } ,

for some constant δ 0 > 0. Thus, by the existence and uniqueness theorem of initial value problem [26, Page 127, Theorem III], it concludes that u ≡ 0 in a small neighbourhood U ̃ of x *. Denote U ̃ = [ a , b ] [ 0,2 π ] . If a = 0 and b = 2π, then u ≡ 0 in [0, 2π]. It suffices to repeat the above analysis if a ≠ 0 or b ≠ 2π.□

Lemma 2.7.

For any fixed λ > 0, if {u i } is a sequence of solutions of (2.4) satisfying lim i u i ( x ) = 0 uniformly for x ∈ [0, 2π], then lim i u i ( x ) = 0 uniformly for x ∈ [0, 2π].

Proof.

Integrating the first equation of (2.4) from 0 to x for any x ∈ [0, 2π], we obtain that

u i ( x ) + u i ( 0 ) = λ 0 x ρ ( τ , u i ( τ ) , u i ( τ ) ) h ( u i ( τ ) ) d τ ,

which together with lim i u i ( x ) = 0 uniformly for x ∈ [0, 2π],

lim i ( u i ( x ) u i ( 0 ) ) = 0

uniformly for x ∈ [0, 2π]. This suggests that lim i u i ( x ) = c uniformly for x ∈ [0, 2π], where c R is a constant. If c ≠ 0, then

lim i ( u i ( x ) u i ( 0 ) ) = lim i 0 x u i ( τ ) d τ = c x , x [ 0,2 π ] ,

and

lim i u i ( x ) = lim i u i ( 0 ) + c x 0 , x ( 0,2 π ] .

This is a contradiction.□

Lemma 2.8.

For any 2π-periodic sign-changing solution uH of (1.6), if u is odd, then

u < π .

Proof.

If u(0) = u(2π) = ‖u, then ‖u = 0. The conclusion is clearly correct. If u(0) = u(2π) ≠ ‖u, we know that there is at least one point x 0 ∈ (0, 2π) such that u′(x 0) = 0. Suppose instead |u(x 0)| = ‖u.

  1. Assume that u(x 0) > 0. Then there exist 0 ≤ x 1 < x 0 < x 2 ≤ 2π such that u(x 1) = u(x 2) = 0 and u(x) > 0, x ∈ (x 1, x 2). From Lemma 2.4, |u′(x)| < 1 for any x ∈ [0, 2π] and

    u ( x 0 ) u ( x 1 ) x 1 x 0 | u ( τ ) | d τ < x 0 x 1 ,

    u ( x 2 ) u ( x 0 ) x 0 x 2 | u ( τ ) | d τ > x 0 x 2 ,

    u ( x 0 ) < 0 + ( x 2 x 1 ) 2 < π .

  2. Assume that u(x 0) < 0. Then there exist 0 ≤ x 1 < x 0 < x 2 ≤ 2π such that u(x 1) = u(x 2) = 0 and u(x) < 0, x ∈ (x 1, x 2). Similarly,

    | u ( x 0 ) | < π .

Lemma 2.9.

Assume that f : R R is odd and f(s) ≥ 0, s ∈ [0, ∞). Then any 2π-periodic sign-changing solution uH of (1.6) satisfies

u < 2 π .

Proof.

Let x 1, x 2 ∈ [0, 2π] such that

u ( x 1 ) = max x [ 0,2 π ] u ( x ) , u ( x 2 ) = min x [ 0,2 π ] u ( x ) .

Then u(x 2) < 0 < u(x 1) and moreover,

u ( x 1 ) u ( x 2 ) = x 2 x 1 u ( τ ) d τ 0 2 π | u ( τ ) | d τ 2 π ,

which implies

u = u ( x 1 ) u ( x 2 ) + 2 π < 2 π if | u ( x 1 ) | | u ( x 2 ) | ,

u = u ( x 2 ) 2 π u ( x 1 ) < 2 π if | u ( x 1 ) | < | u ( x 2 ) | .

3 Proof of Theorem 1.1

Let σ ∈ (0, λ 1) be a fixed constant. By a simple transformation, we know that η 1 = λ 1 + σ > 0 is the simple eigenvalue with positive eigenfunction φ 1 of the indefinite weight linear eigenvalue problem

(3.1) u + σ a ( x ) u = η a ( x ) u , x ( 0,2 π ) , u ( 0 ) = u ( 2 π ) , u ( 0 ) = u ( 2 π ) .

Define the operator L : D(L) → Y by

(3.2) L u = u + σ a ( x ) u ,

where D ( L ) = { u C 2 ( R , R ) | u satisfies u ( x ) = u ( x + 2 π ) , u ( x ) = u ( x + 2 π ) , x R } H . Let ζ C ( R × R , R ) be such that

f ( u , u ) = f 0 u + ζ ( u , u ) .

Note that

lim | s | 0 ζ ( s , p ) s = 0 uniformly for p [ 1,1 ] .

Let us consider

(3.3) L u = ( λ f 0 + σ ) a u + λ a f 0 + ζ ( u , u ) u h ̃ ( u ) + ζ ( u , u ) u u

as a bifurcation problem from the trivial solution u ≡ 0, where h ̃ ( s ) = h ( s ) 1 , s R .

Let G(x, s) be the Green function of −u″ + σau = 0 with the periodic boundary condition. Then (3.3) can be equivalently written as

u = ( λ f 0 + σ ) L u + H ( λ , x , u ) ,

where

L u = 0 2 π G ( x , s ) a ( s ) u ( s ) d s

and

H ( λ , x , u ) = λ 0 2 π G ( x , s ) f 0 + ζ ( u ( s ) , u ( s ) ) u ( s ) h ̃ ( u ( s ) ) + ζ ( u ( s ) , u ( s ) ) u ( s ) a ( s ) u ( s ) d s

Then it is well known that L : Y H is linear completely continuous. From Lemma 2.7, we have that

a ( x ) u ( x ) ( ( 1 u ( x ) 2 ) 3 / 2 1 ) u H 0 as u H 0

uniformly for x R . Obviously, the operator H : R × R × H H is completely continuous, and for any x R ,

lim u H 0 H ( λ , x , u ) H u H = 0

uniformly in λ of any bounded set.

Let

S { ( λ , u ) R × H : u satisfies ( 3.3 ) , and u 0 } ̄ R × H .

Let S + denote the set of functions in H which are positive in R , and set S = −S +, and S = S +S . It is clear that S + and S are disjoint and open in H. The results of Rabinowitz [16], [17] for (3.3) can be stated as follows: there exists a continuum C S bifurcating from ( λ 1 f 0 , 0 ) , such that either it is unbounded or passes through ( λ ̄ f 0 , 0 ) for some λ ̄ , eigenvalue of problem (1.5) with λ ̄ λ 1 .

From [17, Lemma 1.24], if ( λ , u ) C and is near ( λ 1 f 0 , 0 ) , u = αφ 1 + w with w = ◦ (|α|). Since S ± is open and φ 1S, then

C \ { ( λ 1 / f 0 , 0 ) } B ε ( ( λ 1 / f 0 , 0 ) ) R × S

for all positive ɛ small enough, where

B ε ( ( λ 1 / f 0 , 0 ) ) = { ( λ , u ) R × H : u H + | λ λ 1 / f 0 | < ε } .

Lemma 2.6 implies that

C \ { ( λ 1 / f 0 , 0 ) } R × S = .

Otherwise, there exists ( λ , u ) ( C \ { ( λ 1 / f 0 , 0 ) } R × S ) . Then Lemma 2.6 shows that u ≡ 0. So λ = λ ̄ / f 0 for some λ ̄ λ 1 . Then there exists a sequence ( μ i , u i ) C \ { ( λ 1 / f 0 , 0 ) } R × S such that ( μ i , u i ) ( λ ̄ / f 0 , 0 ) in R × H as i → +∞. This contradicts the fact that λ 1 > 0 is the unique eigenvalue with positive eigenfunction. Consequently, one must has that C ( R × S { ( λ 1 / f 0 , 0 ) } ) . Case (ii) in the Lemma 2.1 does not happen. So C is unbounded.

Furthermore, by [27, Theorem 2], there are two continua C + and C , consisting of the bifurcation branch C , which satisfy either C + and C are both unbounded or C + C { ( λ 1 / f 0 , 0 ) } . We have known that u = αφ 1 + w for ( λ , u ) C \ { ( λ 1 / f 0 , 0 ) } near (λ 1/f 0, 0). Since αφ 1S ± if 0 α R ± or R , we have that

C ± \ { ( λ 1 / f 0 , 0 ) } B ε ( ( λ 1 / f 0 , 0 ) ) R × S ±

for all positive ɛ small enough. Similar to the above argument, we may deduce that C ± \ { ( λ 1 / f 0 , 0 ) } cannot leave R × S ± outside of a neighborhood of (λ 1/f 0, 0). Therefore, we have that C ± ( R × S ± { ( λ 1 / f 0 , 0 ) } ) . It follows that both C + and C are unbounded. Otherwise, without less of generality, we may suppose that C is bounded. Then there exists ( λ * , u * ) C + C such that (λ *, u *) ≠ (λ 1/f 0, 0) and u *S +S . This contradicts the definitions of S + and S .

Next, we will analyze the properties of C + .

Lemma 3.1.

Let (H1)–(H4) hold and let Λ ⊂ (0, + ∞) be a closed and bounded interval. Then there exists a constant D* > 0 such that for any ( λ , u ) C + and λ ∈ Λ,

u D * .

Proof.

Assume on the contrary that there exists a sequence

{ ( μ i , y i ) } C + ( Λ × H )

satisfies

(3.4) y i 1 y i 2 = μ i a ( x ) f ( y i ( x ) , y i ( x ) ) , x ( 0,2 π ) , y i ( 0 ) = y i ( 2 π ) , y i ( 0 ) = y i ( 2 π )

and ‖y i → ∞. Set v i ( x ) = y i ( x ) y i and D i = ‖y i . Since y i 1 , we easily find v i 0 . As a consequence, v i → 1 uniformly in [0, 2π], since

| v i ( x ) 1 | = | v i ( x ) v i ( x ̂ i ) | 0 2 π | v i ( τ ) | d τ for all x [ 0,2 π ] ,

where x ̂ i [ 0,2 π ] is such that y i ( x ̂ i ) = y i = D i . Integrate the equation in (3.4) on [0, 2π], we have

μ i f ( D i , 0 ) 0 2 π a ( x ) d x = μ i 0 2 π a ( x ) f ( y i ( x ) , y i ( x ) ) f ( D i , 0 ) d x

and

0 < 0 2 π a ( x ) d x a L 1 sup x [ 0,2 π ] f ( D i v i , D i v i ) f ( D i , 0 ) 1 ,

where a L 1 0 2 π a ( x ) d x . Using (H4), a contradiction follows.□

Lemma 3.2.

Let (H1)–(H4) hold. Then there exists λ* > 0 such that

C + ( ( 0 , λ * ) × H ) = .

Proof.

Assume on the contrary that there exists

{ ( μ j , y j ) } C + ( ( 0 , + ) × H )

such that (μ j , y j ) → (0, c) for some constants c.

Claim. There exist two positive constants 0 < d *D* such that d *cD*.

Assume on the contrary that there exists a sequence

{ ( μ i , y i ) } C + ( ( 0 , + ) × H )

satisfies

y i 1 y i 2 = μ i a ( x ) f ( y i ( x ) , y i ( x ) ) , x ( 0,2 π ) , y i ( 0 ) = y i ( 2 π ) , y i ( 0 ) = y i ( 2 π )

and μ i → 0, y i → 0 uniformly for x ∈ [0, 2π]. Set v i ( x ) = y i ( x ) y i and d i = ‖y i . Passing to the absolute value we have

| v i ( x ) | | v i ( x ) | 1 | v i ( x ) | 2 μ i 0 2 π a ( x ) · f ( y i ( x ) , y i ( x ) ) y i ( x ) · v i ( x ) d x

for all x ∈ [0, 2π]. Therefore, using f 0 ∈ (0, + ∞) and the fact that ‖v i ≤ 1, we obtain that v i 0 uniformly. Similarly, we have

0 < 0 2 π a ( x ) d x a L 1 sup x [ 0,2 π ] f ( d i v i , d i v i ) f ( d i , 0 ) 1 .

Using (H3) and (b) in Remark 1.3, a contradiction easily follows. The existence of D* is shown in Lemma 3.1. Therefore, there exists two positive constant 0 < d *D* such that d *cD*.

From the claim, (μ j , y j ) satisfies

(3.5) 1 μ j y j 1 y j 2 = a ( x ) f ( y j ( x ) , y j ( x ) ) , x ( 0,2 π ) , y j ( 0 ) = y j ( 2 π ) , y j ( 0 ) = y j ( 2 π )

and y j c d * , D * , y j 0 uniformly for x ∈ [0, 2π]. Integrate the equation in (3.5) over the interval, we have

0 = 0 2 π a ( x ) f ( y j , y j ) d x .

Let j → ∞, using the Lebesgue dominated convergence theorem, it follows that

0 = f ( c , 0 ) 0 2 π a ( x ) d x < 0

since f(c, 0) > 0. This is a contradiction.□

Lemma 3.3.

Let (H1)–(H4) hold and let J = Λ * , Λ * [ 0 , + ) be a closed and bounded interval. Then there exists a positive constant γ* < 1 such that for any ( λ , y ) C + ( J × H ) ,

(3.6) | y ( x ) | γ * , x [ 0,2 π ] .

Proof.

Let x 0 ∈ [0, 2π] such that | y ( x 0 ) | = max x [ 0,2 π ] | y ( x ) | = γ < 1 . We can prove it in the following two cases.

Case 1. y′(0) = y′(2π) = 0. If y′(x 0) > 0, without loss of generality, we may assume that x 1 ∈ (x 0, 2π] such that y′(x 1) = 0 and y′(x) > 0 for x ∈ [x 0, x 1), then

γ 1 γ 2 y ( x 0 ) 1 ( y ( x 0 ) ) 2 = λ x 0 x 1 a ( x ) f ( y , y ) d x λ M * x 0 x 1 | a ( x ) | d x Λ * M * a L 1 ,

where M * = max s [ 0 , D * ] , p [ 1,1 ] f ( s , p ) and D* defined in Lemma 3.1. By a simple calculation, we can obtain

(3.7) γ 2 Λ * M * a L 1 Λ * M * a L 1 + 1 < 1 .

If y′(x 0) < 0, without loss of generality, we may assume that x 2 ∈ [0, x 0) such that y′(x 2) = 0 and y′(x) < 0 for x ∈ (x 2, x 0], then

γ 1 γ 2 y ( x 0 ) 1 ( y ( x 0 ) ) 2 = λ x 2 x 0 a ( x ) f ( y , y ) d x λ M * x 2 x 0 | a ( x ) | d x Λ * M * a L 1 .

Similarly,

γ 2 Λ * M * a L 1 Λ * M * a L 1 + 1 < 1 .

Case 2. y′(0) = y′(2π) ≠ 0. Since y(0) = y(2π), we know that y′ changes the sign on the interval [0, 2π]. Then y′ has at least two zeros on the interval (0, 2π).

If y′(x 0) = y′(0) > 0, without loss of generality, we may assume that x 1 ∈ (0, 2π] such that y′(x 1) = 0 and y′(x) > 0 for x ∈ [0, x 1), then we have (3.7). If y′(x 0) = y′(2π) < 0, without loss of generality, we may assume that x 2 ∈ [0, x 0) such that y′(x 2) = 0 and y′(x) < 0 for x ∈ (x 2, x 0], then we can obtain (3.7).

If x 0 ∈ (0, 2π) and y′(x 0) > 0, we may assume that x 1 ∈ (0, 2π] such that y′(x 1) = 0 and y′(x) > 0 for x ∈ [0, x 1); If x 0 ∈ (0, 2π) and y′(x 0) < 0, we may assume that x 2 ∈ [0, x 0) such that y′(x 2) = 0 and y′(x) < 0 for x ∈ (x 2, x 0]. The same method guarantees that (3.7) is true.

In conclusion, choose γ * = Λ * M * a L 1 Λ * M * a L 1 + 1 and the Lemma 3.3 is proved.□

Proof of Theorem 1.1.

The existence of C + is proved in this section. It is easy to verify that (i)-(iii) in Theorem 1.1 are the direct results of Lemma 3.1–3.3.□

4 Proof of Theorem 1.2

  1. Let Y 1 be a subspace of Y defined by

    Y 1 { u Y u satisfies u ( x ) = u ( x ) , x R } .

    Let E 1 be a subspace of H defined by

    (4.1) E 1 { u H u satisfies u ( x ) = u ( x ) , x R } .

    Then Y 1 and E 1 are the Banach spaces endowed with the norm ‖u and ‖u H , respectively.

    Proposition 4.1

    Let f : R R be an odd function, i.e.

    f ( x ) = f ( x ) for all x R .

    Then f(E 1) ⊆ E 1.

    Proof.

    In fact, for uE 1, we have

    u ( x ) = u ( x ) , x R ,

    which means

    f ( u ( x ) ) = f ( u ( x ) ) , x R .

    Since f : R R is an odd function, we may deduce

    f ( u ( x ) ) = f ( u ( x ) ) = f ( u ( x ) ) , x R .

    So f(u) ∈ E 1.□

    Let L 1 : D(L 1) → Y be the linear operator of E 1 defined by

    L 1 u = L u ,

    where D(L 1)≔D(L) ∩ E 1. The eigenvalue of L 1 is λ k + σ, k ≥ 2, and the eigenfunction corresponding to λ k + σ is ϕ k E 1. Let uD(L 1) such that L 1 uY. Then L 1 uY 1. Thus, L 1 1 : Y 1 E 1 is compact.

  2. Let Y 2 be a subspace of Y defined by

    Y 2 { u Y | u satisfies u ( x ) = u ( x ) , x R } .

    Let E 2 be a subspace of H defined by

    E 2 { u H | u satisfies u ( x ) = u ( x ) , x R } .

    Then Y 2 and E 2 are the Banach spaces endowed with the norm ‖u and ‖u H , respectively.

Proposition 4.2.

Let f : R R be an odd function, i.e.

f ( x ) = f ( x ) for all x R .

Then f(E 2) ⊆ E 2.

Proof.

In fact, for uE 2, we have

u ( x ) = u ( x ) , x R ,

which means

f ( u ( x ) ) = f ( u ( x ) ) , x R .

So f(u) ∈ E 2.□

Let L 2 : D(L 2) → Y be the linear operator of E 2 defined by

L 2 u = L u ,

where D(L 2) = D(L) ∩ E 2. The eigenvalue of L 2 is λ k + σ, k ≥ 2, and the eigenfunction corresponding to λ k + σ is ψ k E 2. Let uD(L 2) such that L 2 uY. Then L 2 uY 2. Thus, L 2 1 : Y 2 E 2 is compact.

Define f [ n ] : R R by

(4.2) f [ n ] ( s ) = f ( s ) , | s | 1 n , , n f 1 n s , | s | 0 , 1 n .

Then f [ n ] C ( R , R ) with f [n] is odd, and ( f [ n ] ) 0 lim s 0 f [ n ] ( s ) s = n f ( 1 / n ) > 0 . By (1.7), it follows that lim n ( f [ n ] ) 0 = + .

Since aY and f [ n ] C ( R , R ) , combined with Lemma 2.4 and 2.5, we can consider the auxiliary family of the equations

(4.3) u = λ a ( x ) f [ n ] ( u ) h ( u ) , x R ,

where

h ( s ) = ( 1 s 2 ) 3 / 2 , | s | 1 , 0 , | s | > 1 .

Let ζ [ n ] C ( R , R ) be such that

f [ n ] ( u ) = ( f [ n ] ) 0 u + ζ [ n ] ( u ) = n f ( 1 / n ) u + ζ [ n ] ( u ) .

Note that

lim | s | 0 ζ [ n ] ( s ) s = 0 .

Let h ̃ ( s ) = h ( s ) 1 , s R . A simple calculation would give the following

lim | s | 0 h ̃ ( s ) s = 0 .

4.1 Existence of odd sign-changing solutions

From f [ n ] ( u ) = ( f [ n ] ) 0 u + ζ [ n ] ( u ) , (4.3) is equivalent to

u + σ a u = σ a u + λ a ( f [ n ] ) 0 u + ζ [ n ] ( u ) [ h ( u ) 1 ] + λ ζ [ n ] ( u ) a u + λ ( f [ n ] ) 0 a u , x R .

Thus, we consider

(4.4) L 1 u = λ ( f [ n ] ) 0 + σ a u + λ ( f [ n ] ) 0 + ζ [ n ] ( u ) u h ̃ ( u ) + ζ [ n ] ( u ) u a u

as a bifurcation problem from the trivial solution u ≡ 0. Since uE 1 means that

u is even and ζ [ n ] ( u ) is odd ,

which together with the fact a is even imply that

( f [ n ] ) 0 + ζ [ n ] ( u ) u h ̃ ( u ) + ζ [ n ] ( u ) u a u = ζ [ n ] ( u ( x ) ) h ( u ) a ( x ) Y 1 .

Let G(x, s) be the Green function of −u″ + σau = 0 with the periodic boundary condition. Then problem (4.4) can be equivalently written as

(4.5) u = ( λ ( f [ n ] ) 0 + σ ) L 1 u + H 1 ( λ , x , u ) ,

where

L 1 u = 0 2 π G ( x , s ) a ( s ) u ( s ) d s

and

H 1 ( λ , x , u ) = λ 0 2 π G ( x , s ) ( f [ n ] ) 0 + ζ [ n ] ( u ( s ) ) u ( s ) h ̃ ( u ( s ) ) + ζ [ n ] ( u ( s ) ) u ( s ) a ( s ) u ( s ) d s .

Then it is well known that L 1 : E 1 E 1 is linear completely continuous. From Lemma 2.7, we have that

a ( x ) u ( x ) ( ( 1 u ( x ) 2 ) 3 / 2 1 ) u H 0 as u H 0

uniformly for x R . Obviously, the operator H 1 : R × R × E 1 E 1 is completely continuous, and for any x R ,

lim u H 0 H ( λ , x , u ) H u H = 0

uniformly in λ of any bounded set.

For k ∈ {2, 3, …}, let S k + denote the set of functions in E 1 which have exactly 2k − 3 interior nodal (i.e. nondegenerate) zeros in (0, 2π) and are positive near x = 0+, and set S k = S k + , and S k = S k + S k . They are disjoint and open in E 1. Finally, let Φ k ± = R × S k ± and Φ k = R × S k .

Let

S 1 [ n ] { ( λ , u ) R × E 1 : u satisfies ( 4.4 ) , and u 0 } ̄ R × E 1 .

From Lemma 2.1 (Rabinowitz’s global bifurcation theorem), for each integer k ≥ 2, ν ∈ {+, − }, there exists a continuum ( C [ n ] ) k ν of S 1 [ n ] bifurcating from ( λ k ( f [ n ] ) 0 , 0 ) , such that either it is unbounded or passes through ( λ j ( f [ n ] ) 0 , 0 ) for some jk.

For the sake of completeness we recall that the following framework which was developed in [17, Section 2]. If ( λ , u ) ( C [ n ] ) k ν and is near ( λ k ( f [ n ] ) 0 , 0 ) , then ( λ ̄ , v ) is near (μ, 0). According to Lemma 2.2, v = αϕ k + w with w = ◦ (|α|). Since S k ± is open and ϕ k S k , then

( C [ n ] ) k ν \ λ k ( f [ n ] ) 0 , 0 B ε λ k ( f [ n ] ) 0 , 0 R × S k

for all positive ɛ small enough, where

B ε λ k ( f [ n ] ) 0 , 0 = ( λ , u ) R × E 1 : u H + | λ λ k ( f [ n ] ) 0 | < ε .

Lemma 2.6 implies that

( C [ n ] ) k ν \ λ k ( f [ n ] ) 0 , 0 R × S k = .

Otherwise, there exists ( λ , u ) ( C [ n ] ) k ν \ ( λ k ( f [ n ] ) 0 , 0 ) R × S k . Then Lemma 2.6 shows that u ≡ 0. So λ = λ j ( f [ n ] ) 0 for some jk. Then there exists a sequence ( μ i , u i ) ( C [ n ] ) k ν \ ( λ k ( f [ n ] ) 0 , 0 ) R × S k such that ( μ i , u i ) ( λ k ( f [ n ] ) 0 , 0 ) R × H as i → +∞. This contradicts the facts of u i S k and jk. Consequently, one must has that ( C [ n ] ) k ν ( R × S k ( λ k ( f [ n ] ) 0 , 0 ) ) . Case (ii) in the Lemma 2.1 does not happen. So ( C [ n ] ) k ν is unbounded.

Furthermore, by [27, Theorem 2], there are two continua ( C [ n ] ) k + and ( C [ n ] ) k , consisting of the bifurcation branch ( C [ n ] ) k ν , which satisfy either ( C [ n ] ) k + and ( C [ n ] ) k are both unbounded or ( C [ n ] ) k + ( C [ n ] ) k ( λ k ( f [ n ] ) 0 , 0 ) . We have known that u = αϕ k + w for ( λ , u ) ( C [ n ] ) k ν \ ( λ k ( f [ n ] ) 0 , 0 ) near ( λ k ( f [ n ] ) 0 , 0 ) . Since αϕ k S ± if 0 α R ± or R , we have that

( C [ n ] ) k ± \ λ k ( f [ n ] ) 0 , 0 B ε λ k ( f [ n ] ) 0 , 0 R × S k ±

for all positive ɛ small enough. Similar to the above argument, we may deduce that ( C [ n ] ) k ± \ ( λ k ( f [ n ] ) 0 , 0 ) cannot leave R × S k ± outside of a neighborhood of ( λ k ( f [ n ] ) 0 , 0 ) . Therefore, we have that ( C [ n ] ) k ± ( R × S k ± ( λ k ( f [ n ] ) 0 , 0 ) ) . It follows that both ( C [ n ] ) k + and ( C [ n ] ) k are unbounded. Otherwise, without less of generality, we may suppose that ( C [ n ] ) k is bounded. Then there exists ( λ * , u * ) ( C [ n ] ) k + ( C [ n ] ) k such that ( λ * , u * ) ( λ k ( f [ n ] ) 0 , 0 ) and u * S k + S k . This contradicts the definitions of S k + and S k .

Proof of Theorem 1.2 (i).

Let us verify that ( C [ n ] ) k ν satisfies all of the conditions of Lemma 2.3. Since

lim n λ k ( f [ n ] ) 0 = lim n λ k n f ( 1 / n ) = 0 .

Condition (a) in Lemma 2.3 is satisfied with z* = (0, 0). Obviously

r n = sup | λ | + u H | ( λ , u ) ( C [ n ] ) k ν = ,

and accordingly, (b) holds. (c) Can be deduced directly from the Arzelà–Ascoli Theorem and the definition of f [n]. Therefore, the superior limit of ( C [ n ] ) k ν , i.e. D , contains an unbounded connected component C k ν with ( 0,0 ) C k ν .

Let ( μ n , u n ) C k ν satisfy

μ n + u n H .

Then u n E 1 is odd. From Lemma 2.8, ‖u n < π. Therefore,

sup λ ( λ , u n ) C k ν = .

The proof is completed.□

Proof of Corollary 1.2 (i).

From the proof of Theorem 1.2 (i), we have

( λ , u ) C k ν Φ k ν R × E 1 .

Then for any λ ∈ (0, ∞), k ∈ {2, 3, …}, ν ∈ {+, −}, (1.6) has odd 2π-periodic solutions u k ν satisfies that u k + has exactly 2k − 3 zeros in (0, 2π) and is positive near 0+, u k has exactly 2k − 3 zeros in (0, 2π) and is negative near 0+.□

4.2 Existence of even sign-changing solutions

Let us consider

(4.6) L 2 v = λ ( f [ n ] ) 0 + σ a v + λ ( f [ n ] ) 0 + ζ [ n ] ( v ) v h ̃ ( v ) + ζ [ n ] ( v ) v a v

as a bifurcation problem from the trivial solution v ≡ 0.

For any k ∈ {2, 3, …}, let S ̃ k + denote the set of functions in E 2 which have exactly 2k − 2 interior nodal (i.e. nondegenerate) zeros in (0, 2π) and are positive near x = 0, and set S ̃ k = S ̃ k + , and S ̃ k = S ̃ k + S ̃ k . They are disjoint and open in E 2. Finally, let Ψ k ± = R × S ̃ k ± and Ψ k = R × S ̃ k .

Similarly, for each integer k ≥ 2, ν ∈ {+, − }, there exists a continuum ( D [ n ] ) k ν Ψ k ν of solutions of (4.6) joining ( λ k ( f [ n ] ) 0 , 0 ) to infinity in Ψ k ν . Moreover, ( D [ n ] ) k ν / ( λ k ( f [ n ] ) 0 , 0 ) Ψ k ν .

Proof of Theorem 1.2 (ii).

From Lemma 2.3, the superior limit of ( D [ n ] ) k ν , i.e. D , contains an unbounded connected component D k ν with ( 0,0 ) D k ν .

Since f satisfies (H7) and v n S ̃ k ν is a sign-changing solution, then we can obtain ‖v n < 2π by Lemma 2.9. Therefore,

sup λ ( λ , v n ) D k ν = .

The proof is completed.□

Proof of Corollary 1.2 (ii).

From the proof of Theorem 1.2 (ii), we have

( λ , v ) D k ν Ψ k ν R × E 2 .

Therefore, for any λ ∈ (0, ∞), k ∈ {2, 3, …}, ν ∈ {+, − }, (1.6) has even 2π-periodic solutions v k ν satisfies that v k + has exactly 2k − 2 zeros in (0, 2π) and is positive near 0, v k has exactly 2k − 2 zeros in (0, 2π) and is negative near 0.□


Corresponding authors: Ruyun Ma, School of Mathematics and Statistics, Xidian University, Xi’an 710071, P.R. China; and Department of Mathematics, Northwest Normal University, Lanzhou, 730070, P.R. China, E-mail: ; and Zhongzi Zhao, School of Mathematics and Statistics, Xidian University, Xi’an 710071, P.R. China, E-mail:

Funding source: National Natural Science Foundation of China

Award Identifier / Grant number: 12061064,

Award Identifier / Grant number: 1231040

Funding source: Shaanxi Fundamental Science Research Project for Mathematics and Physics

Award Identifier / Grant number: 22JSY018

Acknowledgments

We are very grateful to the anonymous referees for their valuable suggestions.

  1. Research ethics: Not applicable.

  2. Author contributions: The authors claim that the research was realized in collaboration with the same responsibility. All authors read and approved the last of the manuscript.

  3. Competing interests: All of the authors of this article claim that together they have no competing interests on each other.

  4. Research funding: This work was supported by National Natural Science Foundation of China (No.12061064, No.1231040) and Shaanxi Fundamental Science Research Project for Mathematics and Physics (No.22JSY018).

  5. Data availability: Data sharing not applicable to this article as no datasets were generated.

References

[1] S.-Y. Cheng and S.-T. Yau, “Maximal spacelike hypersurfaces in the Lorentz-Minkowski spaces,” Ann. Math., vol. 104, no. 2, pp. 407–419, 1976. https://doi.org/10.2307/1970963.Search in Google Scholar

[2] R. Bartnik and L. Simon, “Spacelike hypersurfaces with prescribed boundary values and mean curvature,” Commun. Math. Phys., vol. 87, no. 1, pp. 131–152, 1982. https://doi.org/10.1007/bf01211061.Search in Google Scholar

[3] M. K.-H. Kiessling, “On the quasilinear elliptic PDE in physics and geometry,” Commun. Math. Phys., vol. 314, no. 2, pp. 509–523, 2012. https://doi.org/10.1007/s00220-012-1502-3.Search in Google Scholar

[4] C. Corsato, F. Obersnel, and P. Omari, “The Dirichlet problem for gradient dependent prescribed mean curvature equations in the Lorentz-Minkowski space,” Georgian Math. J., vol. 24, no. 1, pp. 113–134, 2017. https://doi.org/10.1515/gmj-2016-0078.Search in Google Scholar

[5] C. Bereanu, P. Jebelean, and J. Mawhin, “Radial solutions for some nonlinear problems involving mean curvature operators in Euclidean and Minkowski spaces,” Proc. Am. Math. Soc., vol. 137, no. 1, pp. 171–178, 2009. https://doi.org/10.1090/s0002-9939-08-09612-3.Search in Google Scholar

[6] R. Ma, H. Gao, and Y. Lu, “Global structure of radial positive solutions for a prescribed mean curvature problem in a ball,” J. Funct. Anal., vol. 270, no. 7, pp. 2430–2455, 2016. https://doi.org/10.1016/j.jfa.2016.01.020.Search in Google Scholar

[7] I. Coelho, C. Corsato, F. Obersnel, and P. Omari, “Positive solutions of the Dirichlet problem for the one-dimensional Minkowski-curvature equation,” Adv. Nonlinear Stud., vol. 12, no. 3, pp. 621–638, 2012. https://doi.org/10.1515/ans-2012-0310.Search in Google Scholar

[8] I. Coelho, C. Corsato, and S. Rivetti, “Positive radial solutions of the Dirichlet problem for the Minkowski-curvature equation in a ball,” Topol. Methods Nonlinear Anal., vol. 44, no. 1, pp. 23–39, 2014. https://doi.org/10.12775/tmna.2014.034.Search in Google Scholar

[9] C. Corsato, F. Obersnel, P. Omari, and S. Rivetti, “Positive solutions of the Dirichlet problem for the prescribed mean curvature equation in Minkowski space,” J. Math. Anal. Appl., vol. 405, no. 1, pp. 227–239, 2013. https://doi.org/10.1016/j.jmaa.2013.04.003.Search in Google Scholar

[10] A. Boscaggin and G. Feltrin, “Positive periodic solutions to an indefinite Minkowski-curvature equation,” J. Differ. Equ., vol. 269, no. 7, pp. 5595–5645, 2020. https://doi.org/10.1016/j.jde.2020.04.009.Search in Google Scholar

[11] A. Boscaggin, G. Feltrin, and F. Zanolin, “Pairs of positive periodic solutions of nonlinear ODEs with indefinite weight: a topological degree approach for the super-sublinear case,” Proc. R. Soc. Edinburgh, Sect. A, vol. 146, no. 3, pp. 449–474, 2016. https://doi.org/10.1017/s0308210515000621.Search in Google Scholar

[12] J. Mawhin, “Radial solutions of Neumann problem for periodic perturbations of the mean extrinsic curvature operator,” Milan J. Math., vol. 79, no. 1, pp. 95–112, 2011. https://doi.org/10.1007/s00032-011-0148-5.Search in Google Scholar

[13] R. Ma, C. Gao, and X. Han, “On linear and nonlinear fourth-order eigenvalue problems with indefinite weight,” Nonlinear Anal., vol. 74, no. 18, pp. 6965–6969, 2011. https://doi.org/10.1016/j.na.2011.07.017.Search in Google Scholar

[14] J. López-Gómez and M. Molina-Meyer, “Bounded components of positive solutions of abstract fixed point equations: mushrooms, loops and isolas,” J. Differ. Equ., vol. 209, no. 2, pp. 416–441, 2005. https://doi.org/10.1016/j.jde.2004.07.018.Search in Google Scholar

[15] K. J. Brown, “Local and global bifurcation results for a semilinear boundary value problem,” J. Differ. Equ., vol. 239, no. 2, pp. 296–310, 2007. https://doi.org/10.1016/j.jde.2007.05.013.Search in Google Scholar

[16] P. H. Rabinowitz, “Nonlinear Sturm-Liouville problems for second order ordinary differential equations,” Commun. Pure Appl. Math., vol. 23, no. 6, pp. 939–961, 1970. https://doi.org/10.1002/cpa.3160230606.Search in Google Scholar

[17] P. H. Rabinowitz, “Some global results for nonlinear eigenvalue problems,” J. Funct. Anal., vol. 7, no. 3, pp. 487–513, 1971. https://doi.org/10.1016/0022-1236(71)90030-9.Search in Google Scholar

[18] J. A. Marlin, “Periodic motions of coupled simple pendulums with periodic disturbances,” Int. J. Non-Linear Mech., vol. 3, no. 4, pp. 439–447, 1968. https://doi.org/10.1016/0020-7462(68)90030-9.Search in Google Scholar

[19] J. M. Coron, “Periodic solutions of a nonlinear wave equation without monotonicity,” Math. Ann., vol. 262, no. 2, pp. 273–285, 1983. https://doi.org/10.1007/bf01455317.Search in Google Scholar

[20] R. Ma and Y. An, “Global structure of positive solutions for nonlocal boundary value problems involving integral conditions,” Nonlinear Anal., vol. 71, no. 10, pp. 4364–4376, 2009. https://doi.org/10.1016/j.na.2009.02.113.Search in Google Scholar

[21] Y. Naito and S. Tanaka, “On the existence of multiple solutions of the boundary value problem for nonlinear second-order differential equations,” Nonlinear Anal., vol. 56, no. 6, pp. 919–935, 2004. https://doi.org/10.1016/j.na.2003.10.020.Search in Google Scholar

[22] M. Garcia-Huidobro and P. Ubilla, “Multiplicity of solutions for a class of nonlinear second-order equations,” Nonlinear Anal., vol. 28, no. 9, pp. 1509–1520, 1997. https://doi.org/10.1016/s0362-546x(96)00014-4.Search in Google Scholar

[23] C. Bereanu and P. J. Torres, “A variational approach for the Neumann problem in some FLRW spacetimes,” Adv. Nonlinear Stud., vol. 19, no. 2, pp. 413–423, 2019. https://doi.org/10.1515/ans-2018-2030.Search in Google Scholar

[24] G. T. Whyburn, Topological Analysis, Princeton, Princeton University Press, 1958.Search in Google Scholar

[25] P. Omari, “A remark on the lower and upper solution method for quasilinear equations modeling relativistic oscillators,” Unpublished note, 2013, pp. 1–5. Available at: https://www.researchgate.net/publication/377896643_A_remark_on_the_lower_and_upper_solution_method_for_quasilinear_equations_modeling_relativistic_oscillators?channel=doilinkId=65bcc1821e1ec12eff6cc16cshowFulltext=rue).Search in Google Scholar

[26] W. Walter, Ordinary Differential Equations, New York, Springer-Verlag, 1998.10.1007/978-1-4612-0601-9Search in Google Scholar

[27] E. N. Dancer, “On the structure of solutions of non-linear eigenvalue problems,” Indiana Univ. Math. J., vol. 23, no. 11, pp. 1069–1076, 1974. https://doi.org/10.1512/iumj.1974.23.23087.Search in Google Scholar

Received: 2023-07-14
Accepted: 2024-02-29
Published Online: 2024-03-29

© 2024 the author(s), published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 13.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ans-2023-0130/html
Scroll to top button