Startseite Existence Results for Solutions to Nonlinear Dirac Systems on Compact Spin Manifolds
Artikel Open Access

Existence Results for Solutions to Nonlinear Dirac Systems on Compact Spin Manifolds

  • Xu Yang EMAIL logo
Veröffentlicht/Copyright: 18. Oktober 2017

Abstract

In this article, we study the existence of solutions for the Dirac system

{ D u = H v ( x , u , v ) on  M , D v = H u ( x , u , v ) on  M ,

where M is an m-dimensional compact Riemannian spin manifold, u,vC(M,ΣM) are spinors, D is the Dirac operator on M, and the fiber preserving map H:ΣMΣM is a real-valued superquadratic function of class C1 with subcritical growth rates. Two existence results of nontrivial solutions are obtained via Galerkin-type approximations and linking arguments.

MSC 2010: 58J05; 58E05; 53C27

1 Introduction and Main Results

Let (M,g) be an m-dimensional compact oriented Riemannian manifold equipped with a spin structure ρ:Pspin(M)Pso(M), and let ΣM=Pspin(M)×σΣm denote the complex spinor bundle on M, which is a complex vector bundle of rank 2[m/2] endowed with the spinorial Levi-Civita connection and a pointwise Hermitian scalar product ,. We always assume m2 in this paper. Write the point of ΣM as (x,ϕ), where xM and ϕΣxM. The Dirac operator is an elliptic differential operator of order one,

D = D g : C ( M , Σ M ) C ( M , Σ M ) ,

locally given by

D ϕ = i = 1 m e i e i ϕ

for ϕC(M,ΣM) and a local g-orthogonal frame {ei}i=1m of the tangent bundle TM. It is an unbounded essential self-adjoint operator in L2(M,ΣM) with the domain C(M,ΣM), and its spectrum consists of an unbounded sequence of real numbers; these mean that the closure of Dg, also denoted by Dg, is a self-adjoint operator in L2(M,ΣM) with the domain H1(M,ΣM) and has the same spectrum as the original Dg (cf. [10, 17]).

Write a point of the fiber product bundle ΣMΣM as (x,u,v), where xM and u,vΣxM. We consider the nonlinear Dirac system

(1.1) { D u = H v ( x , u , v ) on  M , D v = H u ( x , u , v ) on  M ,

where u,vC1(M,ΣM) are spinors and the fiber preserving map H:ΣMΣM is a real-valued superquadratic function of class C1 with subcritical growth rates. Problem (1.1) is the Euler–Lagrange equation of the functional

(1.2) 𝔏 ( u , v ) = M { D u , v - H ( x , u , v ) } 𝑑 x ,

where dx is the Riemann volume measure on M with respect to the metric g and , is the compatible metric on ΣM.

Let

ψ = ( u , v ) , L = ( 0 D D 0 ) , L ψ := ( D v , D u ) .

Then (1.2) becomes

𝔏 ( ψ ) = M { 1 2 L ψ , ψ - H ( x , ψ ) } 𝑑 x .

Problem (1.1) can be viewed as a spinorial analogue of other strongly indefinite variational problems such as infinite dynamical systems [7, 6] and elliptic systems [4, 8]. A typical way to deal with such problems is the min-max method of Benci and Rabinowitz [20], including the mountain pass theorem, linking arguments and so on. Other approaches make use of the homological method, Morse theory and Rabinowitz–Floer homology as in [1, 5, 16]. Recently, Takeshi Isobe [13] established existence results via Galerkin-type approximations and linking arguments. Our aim in this article is to study problem (1.1) with such a method. For the nonlinear Dirac equations on a general compact spin manifold, some results were recently obtained in, e.g., [2, 3, 14, 15, 18, 19, 12, 11, 24].

For the nonlinearity H, we make the following hypotheses: HC0(ΣMΣM,) is C1 in the fiber direction. Real constants 2<p,q<2:=2mm-1 satisfy

(1.3) max { p - 1 , q - 1 , 2 } < μ min { p , q } ,
(1.4) ( 1 - 1 p ) max { p μ , q μ } < 1 2 + 1 2 m ,
(1.5) ( 1 - 1 q ) max { p μ , q μ } < 1 2 + 1 2 m ,
(1.6) p - 1 p q μ < 1 and q - 1 q p μ < 1 .

Moreover, we consider the following hypotheses:

  1. There exists 0<α<1 such that H is α-Hölder continuous in the direction of the base. Moreover, there exists a constant c1>0 such that

    | H u ( x , u , v ) | c 1 ( | u | p - 1 + | v | ( p - 1 ) q p + 1 ) ,
    | H v ( x , u , v ) | c 1 ( | v | q - 1 + | u | ( q - 1 ) p q + 1 ) .

  2. There exists R1>0 such that

    0 < μ H ( x , u , v ) H u , u + H v , v

    for all (x,u,v)ΣMΣM with |(u,v)|R1.

  3. H(x,u,v)0 for all (x,u,v)ΣMΣM.

  4. H(x,u,v)=o(|(u,v)|2) as |(u,v)|0 uniformly for xM.

  5. H(x,-ψ)=H(x,ψ) for any (x,ψ)ΣMΣM.

Notice that H(x,u,v)=|u|p+|v|q satisfies these conditions.

Our main result is the following.

Theorem 1.1.

If the above H satisfies (H1)(H4), then the Dirac system (1.1) possesses at least one solution ψC1(M,ΣM)×C1(M,ΣM) with L(ψ)>0.

Furthermore, for odd nonlinearities we have the following multiplicity result.

Theorem 1.2.

If the above H satisfies (H1), (H2) and (H5), then there exists a sequence of solutions

{ ψ k } k = 1 C 1 ( M , Σ M ) × C 1 ( M , Σ M )

to (1.1) with L(ψk) as k.

These two theorems will be proved by Galerkin-type approximations, linking arguments and the Fountain theorem.

2 Preliminaries

2.1 Spectrum of Operator L

Recall that the operator D is essentially self-adjoint in L2(M,ΣM) with domain 𝒟(D)=C(M,ΣM) of all smooth sections of the spinor bundle on M. In particular, there exists a complete orthogonal basis {ψk} which are contained in C(M,ΣM). Similarly, we can follow the lines of [10] to prove that the operator L=(0DD0) has the same properties as D. This will be completed in several steps. Consider the product Hilbert space L2(M,ΣM)×L2(M,ΣM) with inner product

(2.1) ( ψ 1 , ψ 2 ) 2 = M ψ 1 , ψ 2 𝑑 x = ( u 1 , u 2 ) 2 + ( v 1 , v 2 ) 2

for ψi=(ui,vi)L2(M,ΣM)×L2(M,ΣM), where

( u 1 , u 2 ) 2 = M u 1 , u 2 𝑑 x

is the L2-inner product on spinors. Then L is a symmetric operator in L2(M,ΣM)×L2(M,ΣM) with domain 𝒟(L)=C(M,ΣM)×C(M,ΣM). Without occurrences of confusions we also denote by 2 the norm induced by the inner product in (2.1).

Now we start to prove the essential self-adjointness of the operator L. Let L be the adjoint to L. In the domain D(L) we introduce the norm

N ( ψ ) = ψ 2 2 + L ψ 2 2 ,

where 2 denotes the norm in the Hilbert space L2(M,ΣM)×L2(M,ΣM). Denote by L¯ the closure of L. We shall prove L¯=L.

Lemma 2.1.

If C(M,ΣM)×C(M,ΣM)D(L) is dense with respect to the N-norm, then L is essentially self-adjoint.

Proof.

We prove the lemma in three steps. Step 1: Prove D(L¯)=D(D¯)×D(D¯).

For ψ=(u,v)𝒟(L¯), by the definition of L¯ we have a sequence

ψ n = ( u n , v n ) 𝒟 ( L )

such that ψn-ψ20 and Lψn-L¯ψ20. These imply

u n - u 2 0 , v n - v 2 0 , D v n - ( L ¯ ψ ) 1 2 0 , D u n - ( L ¯ ψ ) 2 2 0 ,

where L¯ψ=((L¯ψ)1,(L¯ψ)2)L2(M,ΣM)×L2(M,ΣM). These mean

u , v 𝒟 ( D ¯ ) , D ¯ v = ( L ¯ ψ ) 1 , D ¯ u = ( L ¯ ψ ) 2 ,

and therefore 𝒟(L¯)𝒟(D¯)×𝒟(D¯) and L¯ψ=(D¯v,D¯u) for all ψ=(u,v)𝒟(L¯).

Conversely, let (ξ,η)𝒟(D¯)×𝒟(D¯). Then there exist sequences ξn𝒟(D) and ηn𝒟(D) such that

ξ n - ξ 2 0 , D ξ n - D ¯ ξ 2 0 , η n - η 2 0 , D η n - D ¯ η 2 0 .

So in L2(M,ΣM)×L2(M,ΣM) we have

ψ n = ( ξ n , η n ) ψ = ( ξ , η ) , L ψ n = ( D η n , D ξ n ) ( D ¯ η , D ¯ ξ ) .

These mean that ψ𝒟(L¯) and L¯ψ=(D¯v,D¯u)). That is, 𝒟(D¯)×𝒟(D¯)𝒟(L¯).

Step 2: Prove D(L)=D(D)×D(D).

Let L be the adjoint to L. Then 𝒟(L) consists of those

ψ = ( u , v ) L 2 ( M , Σ M ) × L 2 ( M , Σ M )

for which there exist ϕ=(a,b)L2(M,ΣM)×L2(M,ΣM) such that

( D g , u ) 2 + ( D f , v ) 2 = ( L φ , ψ ) 2 = ( φ , ϕ ) 2 = ( f , a ) 2 + ( g , b ) 2

for all φ=(f,g)C(M,ΣM)×C(M,ΣM). It easily follows that 𝒟(L)=𝒟(D)×𝒟(D).

Step 3. Since D is essentially self-adjoint, we have D¯=D, and hence 𝒟(L)=𝒟(D¯)×𝒟(D¯)=𝒟(L¯) by Steps 1 and 2. ∎

Since C(M,ΣM) is dense in 𝒟(D) with respect to the N-norm (cf. [10, p. 94]) and 𝒟(L)=𝒟(D)×𝒟(D), we obtain that C(M,ΣM)×C(M,ΣM) is dense in 𝒟(L) with respect to the N-norm. So Lemma 2.1 leads to the following result.

Lemma 2.2.

Let (Mm,g) be a compact spin Riemann manifold. Then the operator L is essentially self-adjoint in L2(M,ΣM)×L2(M,ΣM).

Properties of the operator D are given by the following proposition.

Proposition 2.3.

Let (Mm,g) be a compact spin Riemann manifold. Then the spectrum of L satisfies

σ p ( L ) = σ ( L ¯ ) , σ c ( L ¯ ) = σ r ( L ¯ ) = ,

and hence

σ ( L ) = σ ( L ¯ ) = σ p ( L ) = σ p ( L ¯ ) .

Proof.

Note that σ(L)=σ(L¯). Lemma 2.1 implies that L¯ is a self-adjoint operator, and thus L¯ has no residual spectrum, i.e., σr(L¯)=. We have

(2.2) σ ( L ) = σ ( L ¯ ) = σ p ( L ¯ ) σ c ( L ¯ ) .

If λ is an eigenvalue of L¯, i.e., λσp(L¯), then there exists a spinor field ψL2(M,ΣM)×L2(M,ΣM) with L¯(ψ)=λψ. The regularity theorem for elliptic differential operators then implies that ψ is smooth. Hence, we have ψ𝒟(L), i.e., the point spectrum of L coincides with the point spectrum of the closure:

σ p ( L ) = σ p ( L ¯ ) .

Let us prove that the residual spectrum of L is empty, i.e.,

(2.3) σ r ( L ) = .

In fact, following the arguments in [10, p. 99], let us suppose that σr(L) contains a number λ. Then there exists a spinor field 0φL2(M,ΣM)×L2(M,ΣM) such that

( L ( ψ - λ ψ ) , φ ) 2 = 0 for all  ψ C ( M , Σ M ) × C ( M , Σ M ) .

Choosing ψ with support in a chart and transferring this equation to the Euclidean space, we obtain an elliptic differential operator P (=L-λ) as well as a function φL2(m,Σm)×L2(m,Σm) such that (P(ψ),φ)2=0 for all ψCc(m). By the regularity theorem for elliptic operators, φ is smooth. In this case, the equation (P(ψ),φ)2=0 can be written as (ψ,(L-λ)φ)2=0. This in turn implies Lφ=λφ and φ𝒟(L), i.e., λσp(L). This contradiction affirms (2.3).

From (2.2) and (2.3) we obtain

(2.4) σ p ( L ) σ c ( L ) = σ ( L ) = σ ( L ¯ ) = σ p ( L ¯ ) σ c ( L ¯ ) .

Recall that the approximation spectrumσα(A) of an arbitrary operator A:𝒟(A) (here (A) denotes the range of A) in a Hilbert space is defined by

{ λ x n 𝒟 ( A )  such that  x n = 1 , A ( x n ) - λ x n 0 , n } .

It is easily seen that we have always

σ p ( A ) σ c ( A ) σ α ( A ) σ ( A ) .

Applying this to our L gives σp(L)σc(L)σα(L)σ(L), and thus we arrive at

(2.5) σ α ( L ) = σ α ( L ¯ ) = σ ( L ) = σ ( L ¯ )

by (2.4). It remains to be shown that

(2.6) σ α ( L ) = σ p ( L ) .

Clearly, σp(L)σα(L) by (2.5). Let us prove σα(L)σp(L). Take λσα(L¯). Then there is a sequence of spinor fields ψnC(M,ΣM)×C(M,ΣM) such that

(2.7) ψ n 2 = 1 and L ( ψ n ) - λ ψ n 2 0 .

The latter implies

L ( ψ n ) - λ ψ n 2 2 = D v n - λ u n 2 2 + D u n - λ v n 2 2 0 .

By [10, () on p. 101] we have a constant c>0 such that

(2.8) u H 1 2 + ( min R 4 - c - 1 ) u 2 2 D u 2 2 u H 1 2 + ( max R 4 + c - 1 ) u 2 2

for any spinor u, where R is the scalar curvature of the Riemannian manifold M. Since

1 = ψ n 2 2 = u n 2 2 + v n 2 2 ,

we deduce

D v n 2 2 = D v n - λ u n 2 2 + | λ | 2 u n 2 2 + 2 λ ( D v n - λ u n , u n ) 2
D v n - λ u n 2 2 + | λ | 2 u n 2 2 + 2 | λ | u n 2 D v n - λ u n 2
D v n - λ u n 2 2 + | λ | 2 + 2 | λ | D v n - λ u n 2 ,
D u n 2 2 = D u n - λ v n 2 2 + | λ | 2 v n 2 2 + 2 λ ( D u n - λ v n , v n ) 2
D u n - λ v n 2 2 + | λ | 2 + 2 | λ | D u n - λ v n 2 .

These and (2.8) lead to

u n H 1 2 + v n H 1 2 D u n 2 2 + D v n 2 2 + | min R 4 - c - 1 | ( u n 2 2 + v n 2 2 )
D u n - λ v n 2 2 + 2 | λ | D u n - λ v n 2 + D v n - λ u n 2 2
+ 2 | λ | D v n - λ u n 2 + | min R 4 - c - 1 | + 2 | λ | 2 ,

and therefore ψn=(un,vn) is a bounded sequence in the Sobolev space H1(M,ΣM)×H1(M,ΣM) by (2.7). Moreover, the embedding H1(M,ΣM)×H1(M,ΣM)L2(M,ΣM)×L2(M,ΣM) is compact by the Rellich lemma. Passing to a subsequence, if necessary, we can assume that ψn converges to a spinor field ψ0 in L2(M,ΣM)×L2(M,ΣM), and this immediately implies Lψ0=λψ0. Note that ψ02=limnψn2=1 by (2.7). It follows that λ is an eigenvalue of L, and hence that σα(L)σp(L). Equation (2.6) is proved. ∎

Since L(ψn)22=Dun22+Dvn22 and the closure D¯=D of the Dirac operator D is defined on the subspace H1(M,ΣM)L2(M,ΣM), we immediately obtain the following proposition.

Proposition 2.4.

Under the above notation, the closure L¯=L of L is defined on the subspace

H 1 ( M , Σ M ) × H 1 ( M , Σ M ) L 2 ( M , Σ M ) × L 2 ( M , Σ M ) .

Proposition 2.5.

There exists a complete orthonormal basis ψ1,ψ2, of the Hilbert space of

L 2 ( M , Σ M ) × L 2 ( M , Σ M )

consisting of eigenspinors of the operator L, which are contained in C(M,ΣM)×C(M,ΣM), i.e., L(ψn)=λ¯ψn. Moreover, limn|λ¯n|= and all corresponding eigenvalues have finite multiplicity.

Proof.

Let λ¯σ(L). Since σ(L)=σp(L) by Proposition 2.3, there is a nonzero

ψ = ( u , v ) L 2 ( M , Σ M ) × L 2 ( M , Σ M )

such that L(u,v)=λ¯(u,v), which implies that Du=λ¯v and Dv=λ¯u. So

D ( u + v ) = λ ¯ ( u + v ) and D ( u - v ) = - λ ¯ ( u - v ) .

Note that either u+v0 or u-v0. We get that either λ¯σ(D) or -λ¯σ(D). That is, σ(L)σ(D)(-σ(D)). Then the conclusion follows from [10, proposition on p. 102]. Of course, we can also deduce the claim as in the proof of [10, second proposition on p. 101]. In fact, by (2.8) we have

ψ H 1 2 + ( min R 4 - c - 1 ) ψ 2 2 = u H 1 2 + ( min R 4 - c - 1 ) u 2 2 + v H 1 2 + ( min R 4 - c - 1 ) v 2 2
D u 2 2 + D v 2 2 = L ψ 2 2

for any ψ=(u,v)H1(M,ΣM)×H1(M,ΣM). As in the arguments in [10, p. 101], we derive from this that for each λσ(L¯) the operator

( L - λ ) - 1 : L 2 ( M , Σ M ) × L 2 ( M , Σ M ) L 2 ( M , Σ M ) × L 2 ( M , Σ M )

is compact. Then the conclusions can be derived as in [10, p. 102]. ∎

2.2 Sobolev Spaces

Recall that the closure of the operator D, still denoted by D, is an unbounded self-adjoint operator in L2(M,ΣM) with domain H1(M,ΣM). Moreover, there exists a complete orthonormal basis η1,η2, of the Hilbert space L2(M,ΣM) which are contained in C(M,ΣM) consisting of the eigenspinors of the operator D, i.e. Dηk=λkηk. Moreover, |λk| (as k) and all corresponding eigenvalues λ1,λ2, have finite multiplicity.

For s0, we define the unbounded operator |D|s:L2(M,ΣM)L2(M,ΣM) by

| D | s η = k = 1 | λ k | s a k η k ,

where η=k=1akηkL2(M,ΣM). Denote by Hs(M,ΣM) the domain of it. Then

η = k = 1 a k η k H s ( M , Σ M )

if and only if

k = 1 | λ k | 2 s | a k | 2 < .

Note that Hs(M,ΣM) coincides with the usual L2-Sobolev space of order s, Ws,2(M,ΣM). The inner product on Hs(M,ΣM) is defined by

( η , ζ ) s , 2 := ( | D | s η , | D | s ζ ) 2 + ( η , ζ ) 2 ,

and the induced norm is denoted by s,2, i.e.,

η s , 2 = ( η , η ) s , 2 1 2 .

As usual, we write

η p = ( M | η | p ) 1 p

for the Lp-norm of ηLp(M,ΣM). The Sobolev space H1/2(M,ΣM) is the largest Sobolev space where the integral Mη,Dη𝑑x is well defined. By the Sobolev embedding theorem, we have the continuous embedding H1/2(M,ΣM)Lp(M,ΣM) for 1p2. Moreover, it is also compact if 1p<2.

To treat the Dirac system (1.1) from a variational point of view, it is necessary to give a suitable functional analytic framework. A suitable function space to work with the functional 𝔏 is the Sobolev space E1/2:=H1/2(M,ΣM)×H1/2(M,ΣM) with norm

ψ E 1 / 2 = ( u 1 / 2 , 2 2 + v 1 / 2 , 2 2 ) 1 2

for ψ=(u,v)E1/2.

By Proposition 2.5, the closure of the operator L, still denoted by L, is an unbounded self-adjoint operator in L2(M,ΣM)×L2(M,ΣM) with domain H1(M,ΣM)×H1(M,ΣM). Moreover, the spectrum of L consists of eigenvalues satisfying

- λ ¯ k - λ ¯ 1 - < 0 < λ ¯ 1 + λ ¯ k + + , k .

The complete orthonormal basis {ψk} of L2(M,ΣM)×L2(M,ΣM) consisting of the eigenspinors of L is decomposed into three parts:

{ ψ k } k = 1 = { ψ k - } k = 1 { ψ k 0 } k = 1 l { ψ k + } k = 1 ,

where Lψk-=λ¯k-ψk-, Lψk+=λ¯k+ψk+ and Lψk0=0, k=1,,l=dimker(L)<. By the elliptic regularity,

{ ψ k } k = 1 C ( M , Σ M ) × C ( M , Σ M ) .

Set

E - := span { ψ k - } k = 1 ¯ , E 0 := span { ψ k 0 } k = 1 l , E + := span { ψ k + } k = 1 ¯ ,

where the closure is taken in the E1/2-topology. We then have the orthogonal decomposition of the Hilbert space:

E 1 2 = E - E 0 E + .

Under the growth hypothesis (H1), using Young’s inequality, we derive

(2.9) | H ( x , u , v ) | c 2 ( | u | p + | v | q ) + c 3 ,

which implies that the functional 𝔏 is well defined on E1/2. By the compactness of the inclusion

E 1 2 L p ( M , Σ M ) × L q ( M , Σ M ) ,

we can define the functional :E1/2 by

( u , v ) = M H ( x , u , v ) 𝑑 x .

Proposition 2.6 ([11, Proposition 2.1]).

Assume that H satisfies condition (H1). Then the functional H is of class C1 and its derivative is given by

( u , v ) ( h 1 , h 2 ) = M { H u , h 1 + H v , h 2 } 𝑑 x = M H ψ , h 𝑑 x

for all ψ=(u,v),h=(h1,h2)E1/2. Moreover, H:E1/2E12* is a compact operator.

It follows that for ψ=(u,v),h=(h1,h2)E1/2 we have

d 𝔏 ( ψ ) , h = M D v , h 1 + D u , h 2 d x - M H u , h 1 + H v , h 2 d x
= M L ψ , h 𝑑 x - M H ψ , h 𝑑 x .

Thus the critical points of 𝔏 correspond to solutions of the Dirac system (1.1) as asserted.

3 Palais–Smale Condition for 𝔏n

Let F be a C1 functional on a Banach space H. Recall that a sequence {xn}H is called a (PS)c-sequence if F(xn)c as n and dF(xn)H*0 as n. If all (PS)c -sequences converge in H, we say that F satisfies the (PS)c condition. As in [13], the following modified version of the condition is needed.

Definition 3.1.

Let F:H be as above, and let {H(n)}n=1 be a sequence of closed subspaces in H satisfying H(n)H(n+1) for all n. Set Fn:=FH(n), the restriction of F to H(n). A sequence {xn}H is called a (PS)c*-sequence with respect to {H(n)} if xnH(n) for all n and

  1. F(xn)c as n,

  2. dFn(xn)H(n)*0 as n.

If all (PS)c* sequences converge in H, we say that F satisfies the (PS)c* condition with respect to {H(n)}.

Write E11/2=E-E0 and E21/2=E+. Then E1/2=E11/2E21/2. For nl=dimker(L), we define

E 2 1 2 ( n ) := span ( { ψ k - } k = 1 n - l { ψ k 0 } k = 1 l ) E 2 1 2

equipped with the E1/2-norm.

Lemma 3.2.

Suppose H satisfies (H1) and (H2). Then for any cR the functional L satisfies the (PS)c*-condition with respect to {E21/2(n)}.

Proof.

Write

𝔏 n = 𝔏 | E 2 1 / 2 ( n ) ,

and let {ψn}={(un,vn)}E1/2 be a (PS)c*-sequence with respect to {E21/2(n)}, i.e., ψnE21/2(n) and it satisfies

(3.1) 𝔏 ( ψ n ) c as  n

and

(3.2) d 𝔏 n ( ψ n ) E 2 1 / 2 ( n ) * 0 as  n .

Claim 1: {ψn}E1/2 is bounded.

Condition (H2) implies that there are constants c4,c5>0 such that

(3.3) H ( x , u , v ) c 4 ( | u n | μ + | v n | μ ) - c 5 .

(see [9] for a proof). By (3.1)–(3.3), for large n we have

C + ψ n E 1 / 2 2 𝔏 ( ψ n ) - d 𝔏 n ( ψ n ) , ψ n
= M H ψ , ψ n 𝑑 x - 2 M H ( x , ψ n ) 𝑑 x
( μ - 2 ) M H ( x , ψ n ) 𝑑 x - C
(3.4) c 4 ( μ - 2 ) M | u n | μ + | v n | μ d x - c 5 ( μ - 2 ) - C .

Hereafter, C denotes various positive constants which do not depend on n. Clearly, (3.4) implies that

u n μ μ + v n μ μ C ( 1 + ψ n E 1 / 2 )

for large n, and so

(3.5) u n μ μ C ( 1 + ψ n E 1 / 2 ) and v n μ μ C ( 1 + ψ n E 1 / 2 ) .

Write ψn=ψn-+ψn0+ψn+ according to the decomposition E1/2=E-E0E+. Then we have

ψ + E 1 / 2 2 = ( u + , v + ) E 1 / 2 2
= M ( | | D | 1 2 u + | 2 + | | D | 1 2 v + | 2 + | u + | 2 + | v + | 2 ) 𝑑 x
= M D u + , u + + D v + , v + d x + M ( | u + | 2 + | v + | 2 ) 𝑑 x
M L ψ + , ψ + 𝑑 x + 1 λ 1 + M L ψ + , ψ + 𝑑 x
= ( 1 + 1 λ 1 + ) M L ψ + , ψ + 𝑑 x .

Set

C + = 1 2 ( 1 + 1 λ 1 + ) - 1 .

Then

(3.6) C + ψ + E 1 / 2 2 1 2 M L ψ + , ψ + 𝑑 x .

By (3.2), for large n we have

| d 𝔏 n ( ψ n ) , ψ n + | = | M ψ n + , L ψ n 𝑑 x - M ψ n + , H ψ 𝑑 x | ψ n + E 1 / 2 .

This and (H1) lead to

| M ψ n + , L ψ n 𝑑 x | | M ψ n + , H ψ 𝑑 x | + ψ n + E 1 / 2
| M u n + , H u 𝑑 x | + | M v n + , H v 𝑑 x | + ψ n + E 1 / 2
M | u n + | | H u | 𝑑 x + M | v n + | | H v | 𝑑 x + ψ n + E 1 / 2
(3.7) M c 1 ( | u n | p - 1 + | v n | ( p - 1 ) q p + 1 ) | u n + | 𝑑 x + M c 1 ( | v n | q - 1 + | u n | ( q - 1 ) p q + 1 ) | v n + | + ψ n + E 1 / 2 .

Note that

1 < μ μ - p + 1 < 2 and 1 < μ p μ p - ( p - 1 ) q < 2

by conditions (1.4) and (1.6). We derive

H 1 2 ( M , Σ M ) L μ μ - p + 1 , H 1 2 ( M , Σ M ) L μ p μ p - ( p - 1 ) q ,

and therefore

(3.8) M | u n | p - 1 | u n + | 𝑑 x C u n μ p - 1 u n + 1 2 , 2 ,
(3.9) M | v n | ( p - 1 ) q p | u n + | 𝑑 x C v n μ ( p - 1 ) q p u n + 1 2 , 2 .

By using conditions (1.5) and (1.6), an analogous reasoning yields

(3.10) M | v n | q - 1 | v n + | 𝑑 x C v n μ q - 1 v n + 1 2 , 2 ,
(3.11) M | u n | ( q - 1 ) p q | v n + | 𝑑 x C u n μ ( q - 1 ) p q v n + 1 2 , 2 .

Moreover, it also holds that

(3.12) M | u n + | 𝑑 x C u n + 1 2 , 2 , M | v n + | 𝑑 x C v n + 1 2 , 2 .

By (3.6) and (3.7)–(3.12), we deduce

ψ n + E 1 / 2 2 C M L ψ n + , ψ n + 𝑑 x
= C | M L ψ n + , ψ n 𝑑 x |
ψ n + E 1 / 2 + C ( u n μ p - 1 + v n μ ( p - 1 ) q p + 1 ) u n + 1 2 , 2 + C ( v n μ q - 1 + u n μ ( q - 1 ) p q + 1 ) v n + 1 2 , 2 .

Hence, this and (3.5) lead to

(3.13) ψ n + E 1 / 2 2 C ( ψ n E 1 / 2 p - 1 μ + ψ n E 1 / 2 ( p - 1 ) q μ p + 1 ) ψ n E 1 / 2 + C ( ψ n E 1 / 2 q - 1 μ + ψ n E 1 / 2 ( q - 1 ) p μ q + 1 ) ψ n E 1 / 2 + ψ n E 1 / 2 .

As in (3.6), for any ψ-E- we also have

(3.14) 1 2 M ψ - , L ψ - 𝑑 x - C - ψ - E 1 / 2 2 ,

where C-=12(1-(λ1-)-1)-1>0. Then similar arguments lead to

(3.15) ψ n - E 1 / 2 2 C ( ψ n E 1 / 2 p - 1 μ + ψ n E 1 / 2 ( p - 1 ) q μ p + 1 ) ψ n E 1 / 2 + ( ψ n E 1 / 2 q - 1 μ + ψ n E 1 / 2 ( q - 1 ) p μ q + 1 ) ψ n E 1 / 2 + ψ n E 1 / 2 .

Note that all norms on the finite dimension space E0 are equivalent. We obtain

(3.16) ψ n 0 E 1 / 2 2 C ψ n 0 2 2 C ψ n μ 2 C ( 1 + ψ n E 1 / 2 2 μ ) .

Adding (3.13), (3.15) and (3.16) yields

ψ n E 1 / 2 2 C ( ψ n E 1 / 2 p - 1 μ + ψ n E 1 / 2 ( p - 1 ) q μ p + 1 ) ψ n E 1 / 2 + C ( ψ n E 1 / 2 q - 1 μ + ψ n E 1 / 2 ( q - 1 ) p μ q + 1 ) ψ n E 1 / 2
(3.17) + 2 ψ n E 1 / 2 + C ψ n E 1 / 2 2 μ + C .

By the assumptions on p, q and μ above (H1), it is easily checked that the total exponent of each term in the right-hand side of (3) is less than 2. It follows that the sequence {ψn} is bounded in E1/2. Claim 1 is proved.

Claim 2: {ψn}={(un,vn)} has a convergent subsequence in E1/2.

Passing to a subsequence, we may assume that for some ψ=(u,v)E1/2,

ψ n ψ weakly in  E 1 2 ,
ψ n ψ strongly in  L p ( M , Σ M ) × L p ( M , Σ M ) .

From (3.2) and the boundedness of {ψn} in E1/2 we deduce

o ( 1 ) = d 𝔏 n ( ψ n ) , ψ n + - ψ +
= M L ( ψ n ) , ψ n + - ψ + 𝑑 x - M H ψ ( x , ψ n ) , ψ n + - ψ + 𝑑 x
(3.18) = M L ( ψ n ) , ψ n + - ψ + 𝑑 x - M H u ( x , ψ n ) , u n + - u + 𝑑 x - M H v ( x , ψ n ) , v n + - v + 𝑑 x .

It follows from (H1) that

| M H u ( x , ψ n ) , u n + - u + 𝑑 x | M | H u ( x , ψ n ) | | u n + - u + | 𝑑 x
M c 1 ( | u n | p - 1 + | v n | ( p - 1 ) q p + 1 ) | u n + - u + | 𝑑 x
c 1 ( u n + - u + 1 + u n p ( p - 1 ) u n + - u + p + v n μ ( p - 1 ) q p u n + - u + μ p μ p - ( p - 1 ) q )
(3.19) = o ( 1 ) as  n .

Similarly, we can also arrive at

(3.20) | M H v ( x , ψ n ) , v n + - v + 𝑑 x | = o ( 1 ) as  n .

Then (3.18)–(3.20) yield

| M L ( ψ n ) , ψ n + - ψ + 𝑑 x | 0 as  n ,

and thus

(3.21) ψ n + - ψ + E 1 / 2 2 C + - 1 2 M L ( ψ n + - ψ + ) , ψ n + - ψ + 𝑑 x = o ( 1 ) as  n

by (3.6). Note that dimE0=l< implies

(3.22) ψ n 0 - ψ 0 E 1 / 2 0 as  n .

In order to prove ψn-ψ- let Pn:E1/2E21/2(n) denote the orthogonal projection. Note that Pnφφ in E1/2 for any φE1/2. Hence,

o ( 1 ) = d 𝔏 n ( ψ n ) , ψ n + - P n ψ +
(3.23) = M L ( ψ n ) , ψ n + - P n ψ + 𝑑 x - M H ψ ( x , ψ n ) , ψ n + - P n ψ + 𝑑 x .

As in (3.19) and (3.20), we can show that the second term of (3.23) converges to 0 as n. Moreover, by (3.21) and (3.22) we obtain

o ( 1 ) = M L ( ψ n ) , ψ n - P n ψ 𝑑 x = M L ( ψ n - ) , ψ n - - P n ψ - 𝑑 x + o ( 1 ) ,

and therefore

o ( 1 ) = M L ( ψ n ) , ψ n - - P n ψ - 𝑑 x = M L ( ψ n - ) , ψ n - - ψ - 𝑑 x + o ( 1 ) .

These and (3.14) lead to

(3.24) ψ n - - ψ - E 1 / 2 2 - C - - 1 2 M L ( ψ n - - ψ - ) , ψ n - - ψ - 𝑑 x = o ( 1 )

as n. Combing (3.21) with (3.22) and (3.24), we deduce that ψnψ in E1/2. This proves Claim 2, and so the (PS)c-condition is verified. ∎

Remark 3.3.

By essentially the same argument, it can be shown that for any n and c, the functional 𝔏n satisfies the (PS)c-condition on E21/2(n).

4 Proofs of the Main Results

4.1 Proof of Theorem 1.1

Taking e+E+ such that e+E1/2=1, we define for r>0, R>0 and ρ>0,

Q r , R := { ψ - + ψ 0 + s e + ψ - E - , ψ 0 E 0 ,  0 s r , ψ - + ψ 0 E 1 / 2 R }

and

S ρ + := { ψ E + ψ E 1 / 2 = ρ } .

We need to give a uniform estimate for approximate critical levels of 𝔏. To this end, let us estimate the quantities

α r , R := sup { 𝔏 ( ψ ) ψ Q r , R }

and

β ρ := inf { 𝔏 ( ψ ) ψ S ρ + } .

Lemma 4.1.

Suppose that H satisfies (H2) and (H3). Then there exist r>0 and R>0 such that αr,R0.

Proof.

For ψ=(u,v)=ψ-+ψ0+se+Qr,R, a direct computation gives

𝔏 ( ψ ) = 1 2 M ψ - , L ψ - 𝑑 x + s 2 2 M e + , L e + 𝑑 x - M H ( x , ψ - + ψ 0 + s e + ) 𝑑 x .

Set

C e + := 1 2 M e + , L e + 𝑑 x > 0 .

We can derive from (3.3) that

(4.1) M H ( x , ψ ) 𝑑 x c 4 M | u | μ + | v | μ d x - c 5 .

Moreover, by Hölder’s inequality we have

M | ψ - | 2 𝑑 x + M | ψ 0 | 2 𝑑 x + s 2 M | e + | 2 𝑑 x = M | ψ | 2 𝑑 x
= M | u | 2 + | v | 2 d x
( vol ( M ) 1 - 2 μ ) ( M | u | μ d x ) 2 μ + ( vol ( M ) 1 - 2 μ ) ( M | v | μ d x ) 2 μ .

This and (4.1) lead to

(4.2) 𝔏 ( ψ ) 1 2 M ψ - , L ψ - 𝑑 x + C e + s 2 - C ( ψ - 2 2 + ψ 0 2 2 + s 2 e + 2 2 ) μ 2 + c 5 .

Since all norms on the finite dimension space E0 are equivalent, it follows from (4.2) and (3.14) that

𝔏 ( ψ ) - C - ψ - E 1 / 2 2 + C e + s 2 - C ( ψ - 2 2 + ψ 0 2 2 + s 2 e + 2 2 ) μ 2 + c 5
(4.3) - C - ψ - E 1 / 2 2 + C e + s 2 - C ( ψ 0 E 1 / 2 μ + s μ ) + c 5 .

Note that μ>2. We can choose r>0 so that Ce+s2-Csμ+c50 for sr. Set

M := max 0 s r ( C e + s 2 - C s μ + c 5 ) > 0 .

Then we can choose R>0 such that

ψ - + ψ 0 E 1 / 2 R C - ψ - E 1 / 2 2 + C ψ 0 E 1 / 2 μ M .

There are three cases:

  1. 0sr and ψ-+ψ0E1/2=R;

  2. s=0 and ψ-+ψ0E1/2R;

  3. s=r and ψ-+ψ0E1/2R.

For the first case, our arguments above have showed that 𝔏(ψ)0. In the second case, it follows from (H3) that

𝔏 ( ψ ) = 1 2 M ψ - , L ψ - 𝑑 x - M H ( x , ψ ) 𝑑 x 0 .

Inequality (4.3) and the choices of r and R lead to 𝔏(ψ)0 in the final case.

Summing up the three cases, we complete the proof. ∎

Lemma 4.2.

Assume that H satisfies (H1), (H3) and (H4). Then for r>0 in Lemma 4.1, there exists 0<ρ<r such that βρ>0.

Proof.

By (2.9), (H3) and (H4), there exists c6>0 such that

(4.4) 0 H ( x , u , v ) C + 2 ( | u | 2 + | v | 2 ) + c 6 ( | u | p + | v | q )

for any (x,u,v)ΣMΣM. It follows from this, (3.5) and (4.4) that

𝔏 ( ψ ) = 1 2 M ψ , L ψ 𝑑 x - M H ( x , ψ ) 𝑑 x
C + 2 ψ E 1 / 2 2 - c 6 M ( | u | p + | v | q ) 𝑑 x
C + 2 ( u 1 2 , 2 2 + v 1 2 , 2 2 ) - c 7 ( u 1 2 , 2 p + v 1 2 , 2 q )
= C + 2 ( ρ 1 2 + ρ 2 2 ) - c 7 ( ρ 1 p + ρ 2 q )

for ψ=(u,v)Sρ+, where

ρ 1 = u 1 2 , 2 , ρ 2 = v 1 2 , 2 ,

and thus they satisfy ρ2=ρ12+ρ22 because of

ψ E 1 / 2 2 = u 1 2 , 2 2 + v 1 2 , 2 2 .

Note that p>2 and q>2. We can choose ρ(0,r) such that

C + 2 ( ρ 1 2 + ρ 2 2 ) - c 7 ( ρ 1 p + ρ 2 q ) > 0 .

It follows that the conclusion of the lemma holds for this ρ. ∎

Completing the proof of Theorem 1.1.

Let r>0 and R>0 be as in Lemma 4.1. Then Qr,R(n):=Qr,RE21/2(n) is homeomorphic to the 2(n+1)-dimensional disc. Define

𝒞 ( n ) := { Φ C 0 ( Q r , R ( n ) , E 2 1 2 ( n ) ) Φ | Q r , R ( n ) = I Q r , R ( n ) } .

For 0<ρ<r, by using the topological degree theory, it is easily checked that Sρ+ and Qr,R(n) link, i.e., the following hold (see [21]):

  1. Sρ+Qr,R(n)=.

  2. For any Φ𝒞(n), there holds Φ(Qr,R(n))Sρ+.

Define the min-max value

c ( n ) := inf Φ 𝒞 ( n ) max 𝔏 ( Φ ( Q r , R ( n ) ) ) .

Then Lemma 4.2 and (ii) above imply

(4.5) β ρ c ( n ) .

Observe that the inclusion I:Qr,R(n)E21/2(n) belongs to 𝒞(n) and that Qr,RE1/2 is bounded. We deduce

(4.6) c ( n ) max 𝔏 ( Q r , R ( n ) ) sup 𝔏 ( Q r , R ) <

because 𝔏 maps bounded sets in E1/2 into bounded sets in by (H1) and the Sobolev embedding.

Now by Remark 3.3, Lemma 4.1 and (4.5), we may use the standard deformation argument (cf. [22]) to show that c(n) is a critical value of 𝔏n and thus that there exists ψnE21/2(n) such that

d 𝔏 n ( ψ n ) = 0 and 𝔏 n ( ψ n ) = c ( n ) .

Note that (4.5) and (4.6) imply c(n) to be bounded. After taking a subsequence if necessary, we may assume that c(n)c for some βρc<. Therefore, {ψn} becomes a (PS)c*-sequence with respect to {E21/2(n)}. It follows from Lemma 3.2 that 𝔏 satisfies the (PS)c-condition with respect to {E21/2(n)}. Furthermore, taking a subsequence if necessary, we may assume that {ψn} converges to ψ in E21/2(n). It is clear that ψ satisfies 𝔏(ψ)=c. Since the boundedness of {ψn} in E1/2 implies that of

{ d 𝔏 ( ψ n ) E 1 / 2 } ,

for any φE21/2(n) we deduce

| d 𝔏 ( ψ n ) , φ - P n φ | d 𝔏 ( ψ n ) E 1 / 2 φ - P n φ E 1 / 2 0

as n. This and

d 𝔏 ( ψ n ) , φ = d 𝔏 ( ψ n ) , P n φ + d 𝔏 ( ψ n ) , φ - P n φ
= d 𝔏 ( ψ n ) , φ - P n φ

lead to d𝔏(ψ),φ=0 for any φE1/2, i.e., ψ is a critical point of 𝔏 with 𝔏(ψ)βρ>0. The elliptic regularity theory also implies ψC1(M,ΣM)×C1(M,ΣM) (cf. [2]), and thus ψ is a nontrivial classical solution of (1.1). Theorem 1.1 is proved. ∎

4.2 Proof of Theorem 1.2

We begin with the Fountain theorem for semi-definite functionals [23].

Theorem 4.3 (Fountain theorem).

Let H be a Banach space with basis {ej}j=1. For d1, we have H=H1+H2, where

H 1 = span { e j } j = 1 d 𝑎𝑛𝑑 H 2 = span { e j } j = d ¯ .

Suppose FC1(H,R) is an even functional. For ρ>0, let

B ρ = { u H 1 u ρ } 𝑎𝑛𝑑 Γ = { γ C 0 ( B ρ , H ) γ ( - u ) = - γ ( u ) for all  u B ρ , γ | B ρ = I B ρ } .

Define

c := inf γ Γ max 𝔏 ( γ ( B ρ ) ) .

Suppose that F satisfies the (PS)c-condition and there exists 0<r<ρ such that

b := inf { 𝔉 ( u ) u H 2 , u = r } > a := max { 𝔉 ( u ) u H 1 , u = ρ } .

Then c is a critical value of F with cb.

In order to apply the Fountain theorem for the functional

𝔏 n = 𝔏 | E 2 1 / 2 ( n ) ,

for each j1 we define

E 2 1 2 ( n , j ) = span ( { ψ k - } k = 1 n - l { ψ k 0 } k = 1 l ) span { ψ k + } k = 1 j

and

E 2 1 2 + ( n , j ) = span { ψ k + } k = j ¯ .

Then

E 2 1 2 ( n ) = E 2 1 2 ( n , j ) + E 2 1 2 + ( n , j ) ,

and 𝔏n is even on E21/2(n) by (H5).

Lemma 4.4.

Suppose (H2) is satisfied. Then for each j1 there exists ρ(j)>0 such that Ln(ψ)0, provided ψE21/2(n,j) and ψE1/2ρ(j).

Proof.

For ψ=ψ-+ψ++ψ0E21/2(n,j), by (3.3) and (3.14) we have

𝔏 n ( ψ ) = 1 2 L ψ , ψ - H ( x , ψ ) d x
1 2 L ψ - , ψ - 𝑑 x + 1 2 L ψ + , ψ + 𝑑 x - c 4 | u | μ + | v | μ d x + c 5
(4.7) - C - ψ - E 1 / 2 2 + λ ( j ) 2 ψ + 2 2 - c 4 ( | u | μ + | v | μ ) 𝑑 x + c 5 ,

where λ(j) is the largest eigenvalue of L on E21/2(n,j), and thus

1 2 L ψ + , ψ + 𝑑 x λ ( j ) 2 ψ + 2 2 .

Since u2Cuμ and v2Cvμ by Hölder’s inequality, and

a λ + b λ + c λ 3 ( a + b + c ) λ

for all nonnegative numbers a, b, c and λ1, we get

u + 2 μ + u 0 2 μ + u - 2 μ 3 ( u + 2 2 + u 0 2 2 + u - 2 2 ) μ 2
= 3 u 2 μ C u μ μ ,
v + 2 μ + v 0 2 μ + v - 2 μ 3 ( v + 2 2 + v 0 2 2 + v - 2 2 ) μ 2
= 3 v 2 μ C v μ μ .

From these and (4.7) we deduce

𝔏 n ( ψ ) - C - ψ - E 1 / 2 2 + λ ( j ) 2 u + 2 2 + λ ( j ) 2 v + 2 2 C u - 2 μ - C u 0 2 μ - C u + 2 μ - C v - 2 μ - C v 0 2 μ - C v + 2 μ + C .

Note that dim(E21/2(n,j)E+)=j<. Any two norms on E21/2(n,j)E+ are equivalent. In particular, for the L2 and the H1/2-norm on it we have constants C1, C2 (independent of n) such that

C 1 u + 1 2 , 2 u + 2 C 2 u + 1 2 , 2 , C 1 v + 1 2 , 2 v + 2 C 2 v + 1 2 , 2 .

It follows from these that

𝔏 n ( ψ ) - C - ψ - E 1 / 2 2 + λ ( j ) 2 C 2 2 u + 1 2 , 2 2 + λ ( j ) 2 C 2 2 v + 1 2 , 2 2 - C u - 2 μ - C u 0 2 μ - C C 1 μ u + 1 2 , 2 μ
- C v - 2 μ - C v 0 2 μ - C C 1 μ v + 1 2 , 2 μ + C .

Note that we have assumed μ>2. It is not hard to prove that there exists ρ(j)>0 such that

λ ( j ) 2 C 2 2 u + 1 2 , 2 2 + λ ( j ) 2 C 2 2 v + 1 2 , 2 2 - C C 1 μ u + 1 2 , 2 μ - C C 1 μ v + 1 2 , 2 μ 0

for all ψE21/2(n,j) with ψE1/2ρ(j). Hence, 𝔏n(ψ)0. ∎

Note that E21/2+(n,j) does not depend on n. For 1p<2, let us define

β j , p := sup { ψ p ψ E 2 1 2 + ( n , j ) , ψ E 1 / 2 = 1 } .

Lemma 4.5.

β j , p 0 as j.

Proof.

By the definition of βj,p, for each j we can find ψjE21/2+(n,j) such that ψjE1/2=1 and 12βj,p<ψjp. Using the compactness of the embedding E1/2Lp×Lp, we may assume (after taking a subsequence if necessary) ψjψ in E1/2 and ψjψ in Lp×Lp for some ψE1/2. We can also assume ψjψ in L2×L2 since the embedding E1/2L2×L2 is compact. For a fixed

ψ ^ { ψ k - } k = 1 { ψ k 0 } k = 1 l { ψ k + } k = 1 ,

from

ψ j E 2 1 / 2 + ( n , j ) = span { ψ k + } k = j ¯

we derive that (ψj,ψ^)2=0 for sufficiently large j, and so limj(ψj,ψ^)2=0. These show that ψj0 in L2×L2. Hence ψ=0, and therefore

1 2 β j , p < ψ j p 0 .

Lemma 4.6.

Under conditions (H1) and (H2), there exists r(j)>0 such that

b ( j ) := inf { 𝔏 n ( ψ ) ψ E 2 1 2 + ( n , j ) , ψ E 1 / 2 = r ( j ) }

as j.

Proof.

By (2.9) and (3.6), we have for ψE21/2+(n,j) with ψE1/2=r(j),

𝔏 n ( ψ ) = 1 2 M ψ , L ψ 𝑑 x - M H ( x , ψ ) 𝑑 x
C + ψ E 1 2 2 - c 2 M ( | u | p + | v | q ) 𝑑 x - c 3
C + ψ E 1 2 2 - c 2 ψ p p - c 2 ψ q q - c 3
C + r ( j ) 2 - c 2 r ( j ) p β j , p p - c 2 r ( j ) q β j , q q - c 3 .

Without loss of generality, we may assume p>q, and proceed in two cases.

Case 1. If 0<r(j)<1, then

𝔏 n ( ψ ) C + r ( j ) 2 - c 2 r ( j ) q β j , p p - c 2 r ( j ) q β j , q q - c 3 .

Since the function

r ( j ) C + r ( j ) 2 - c 2 r ( j ) q β j , p p - c 2 r ( j ) q β j , q q - c 3

attains the maximum

( C + - 2 C + q ) ( 2 C + c 2 q ( β j , p p + β j , q q ) ) 2 q - 2 - c 3

at

r ( j ) = ( 2 C + q c 2 ( β j , p p + β j , q q ) ) 1 q - 2 ,

we may use Lemma 4.5 to obtain

b ( j ) ( C + - 2 C + q ) ( 2 C + c 2 q ( β j , p p + β j , q q ) ) 2 q - 2 - c 3

as j.

Case 2. If r(j)>1, we have

𝔏 n ( ψ ) C + r ( j ) 2 - c 2 r ( j ) p β j , p p - c 2 r ( j ) p β j , q q - c 3 .

As in Case 1, we may use Lemma 4.5 to prove

b ( j ) ( C + - 2 C + p ) ( 2 C + c 2 p ( β j , p p + β j , q q ) ) 2 p - 2 - c 3

as j. The lemma is proved. ∎

Proof of Theorem 1.2.

Note that ρ(j) in Lemma 4.4 can be replaced by a larger number. By taking ρ(j) large if necessary, we may assume that r(j)<ρ(j) for each j. Then from Remark 3.3, Lemma 4.4, Lemma 4.6 and Theorem 4.3 we may derive that for large j (independent of n) the min-max value

c ( n , j ) := inf γ Γ ( n , j ) max 𝔏 n ( γ ( B ( n , j ) ) )

is a critical value of 𝔏n, where

B ( n , j ) = { ψ E 2 1 2 ( n , j ) ψ E 1 / 2 = ρ ( j ) }

and

Γ ( n , j ) = { γ C 0 ( B ( n , j ) , E 2 1 2 ( n ) ) γ ( - ψ ) = - γ ( ψ ) , γ | B ( n , j ) = I | B ( n , j ) } .

Let

B ( j ) = { ψ E 1 1 2 span { ψ k + } k = 1 j ψ E 1 / 2 ρ ( j ) } .

Observe that 𝔏 is bounded on each bounded subset in E1/2. We obtain

b ( j ) c ( n , j ) max 𝔏 n ( B ( n , j ) ) sup 𝔏 ( B ( j ) ) = : d ( j ) < .

Furthermore, passing to a subsequence if necessary, we may assume

c ( n , j ) c ( j ) [ b ( j ) , d ( j ) ] as  n .

Note that c(n,j) is a critical value of 𝔏n. We have ψn,jE21/2(n) such that d𝔏n(ψn,j)=0 and 𝔏n(ψn,j))=c(n,j). Moreover, since 𝔏 satisfies the (PS)c(j)-condition with respect to {E21/2(n)}, after taking a subsequence if necessary, there exists ψjE1/2 such that ψn,jψj in E1/2 as n. As in the proof of Theorem 1.1, we can prove that ψj is a critical point of 𝔏 with 𝔏(ψj)=c(j). Finally, by noting that b(j)c(j) and that b(j) by Lemma 4.6, the proof is complete. ∎


Communicated by Yiming Long


Award Identifier / Grant number: 10971014

Award Identifier / Grant number: 11271044

Funding statement: This work was partially supported by the National Natural Science Foundation of China through grants 10971014 and 11271044.

References

[1] A. Abbondandolo, A new cohomology for the Morse theory of strongly indefinite functionals on Hilbert spaces, Topol. Methods Nonlinear Anal. 9 (1997), no. 2, 325–382. 10.12775/TMNA.1997.017Suche in Google Scholar

[2] B. Ammann, A variational problem in conformal spin geometry, J. Geom. Phys. 62 (2003), no. 2, 213–223. Suche in Google Scholar

[3] B. Ammann, The smallest Dirac eigenvalue in a spin-conformal class and cmc immersions, Comm. Anal. Geom. 17 (2009), no. 3, 429–479. 10.4310/CAG.2009.v17.n3.a2Suche in Google Scholar

[4] S. Angenent and R. van der Vorst, A superquadratic indefinite elliptic system and its Morse–Conley–Floer homology, Math. Z. 231 (1999), no. 2, 203–248. 10.1007/PL00004731Suche in Google Scholar

[5] A. Bahri and P.-L. Lions, Solutions of superlinear elliptic equations and their Morse indices, Comm. Pure Appl. Math. 45 (1992), no. 9, 1205–1215. 10.1002/cpa.3160450908Suche in Google Scholar

[6] T. Bartsch and Y. Ding, Periodic solutions of superlinear beam and membrane equations with perturbations from symmetry, Nonlinear Anal. 44 (2001), no. 6, 727–748. 10.1016/S0362-546X(99)00302-8Suche in Google Scholar

[7] T. Bartsch and Y. Ding, Homoclinic solutions of an infinite-dimensional Hamiltonian system, Math. Z. 240 (2002), no. 2, 289–310. 10.1007/s002090100383Suche in Google Scholar

[8] D. G. de Figueiredo and P. L. Felmer, On superquadratic elliptic systems, Trans. Amer. Math. Soc. 343 (1994), no. 1, 99–116. 10.1007/978-3-319-02856-9_26Suche in Google Scholar

[9] P. L. Felmer, Periodic solutions of “superquadratic” Hamiltonian systems, J. Differential Equations 102 (1993), no. 1, 188–207. 10.1006/jdeq.1993.1027Suche in Google Scholar

[10] T. Friedrich, Dirac Operators in Riemannian Geometry, Grad. Stud. Math. 25, American Mathematical Society, Providence, 2000. 10.1090/gsm/025Suche in Google Scholar

[11] W. Gong and G. Lu, Existence results for coupled Dirac systems via Rabinowitz–Floer theory, preprint (2015), https://arxiv.org/abs/1511.06829. Suche in Google Scholar

[12] W. Gong and G. Lu, On Dirac equation with a potential and critical Sobolev exponent, Commun. Pure Appl. Anal. 14 (2017), no. 6, 2231–2263. 10.3934/cpaa.2015.14.2231Suche in Google Scholar

[13] T. Isobe, Existence results for solutions to nonlinear Dirac equations on compact spin manifolds, Manuscripta Math. 135 (2011), no. 3–4, 329–360. 10.1007/s00229-010-0417-6Suche in Google Scholar

[14] T. Isobe, Nonlinear Dirac equations with critical nonlinearities on compact Spin manifolds, J. Funct. Anal. 260 (2011), no. 1, 253–307. 10.1016/j.jfa.2010.09.008Suche in Google Scholar

[15] T. Isobe, A perturbation method for spinorial Yamabe type equations on Sm and its application, Math. Ann. 355 (2013), no. 4, 1255–1299. 10.1007/s00208-012-0818-9Suche in Google Scholar

[16] W. Kryszewski and A. Szulkin, An infinite-dimensional Morse theory with applications, Trans. Amer. Math. Soc. 349 (1997), no. 8, 3181–3234. 10.1090/S0002-9947-97-01963-6Suche in Google Scholar

[17] H. B. Lawson, Jr. and M.-L. Michelsohn, Spin Geometry, Princeton Math. Ser. 38, Princeton University Press, Princeton, 1989. Suche in Google Scholar

[18] A. Maalaoui, Rabinowitz–Floer homology for superquadratic Dirac equations on compact spin manifolds, J. Fixed Point Theory Appl. 13 (2013), no. 1, 175–199. 10.1007/s11784-013-0116-5Suche in Google Scholar

[19] A. Maalaoui and V. Martino, The Rabinowitz–Floer homology for a class of semilinear problems and applications, J. Funct. Anal. 269 (2015), no. 12, 4006–4037. 10.1016/j.jfa.2015.09.015Suche in Google Scholar

[20] P. H. Rabinowitz, Periodic solutions of Hamiltonian systems, Comm. Pure Appl. Math. 31 (1978), no. 2, 157–184. 10.21236/ADA096642Suche in Google Scholar

[21] J. T. Schwartz, Nonlinear Functional Analysis, Gordon & Breach Science Publishers, New York, 1969. Suche in Google Scholar

[22] M. Struwe, Variational Methods, 4th ed., Ergeb. Math. Grenzgeb. (3) 34, Springer, Berlin, 2008. Suche in Google Scholar

[23] M. Willem, Minimax Theorems, Progr. Nonlinear Differential Equations Appl. 24, Birkhäuser, Boston, 1996. 10.1007/978-1-4612-4146-1Suche in Google Scholar

[24] X. Yang, R. Jin and G. Lu, Solutions of Dirac equations on compact spin manifolds via saddle point reduction, J. Fixed Point Theory Appl. 19 (2017), no. 1, 215–229. 10.1007/s11784-016-0350-8Suche in Google Scholar

Received: 2017-03-11
Revised: 2017-07-23
Accepted: 2017-09-18
Published Online: 2017-10-18
Published in Print: 2018-02-01

© 2017 Walter de Gruyter GmbH, Berlin/Boston

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

Heruntergeladen am 14.9.2025 von https://www.degruyterbrill.com/document/doi/10.1515/ans-2017-6034/html
Button zum nach oben scrollen