This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

\marginsize

2.4cm2.4cm2.4cm2.4cm\titlecontentssection[20pt] \contentspush\thecontentslabel. \contentspage

Mountain pass type solutions for a generalized Frenkel-Kontorova model

Wen-Long Li Wen-Long Li, School of Mathematics, Sun Yat-Sen University, Guangzhou, 510275, P. R. China [email protected]
Abstract.

We study a generalized Frenkel-Kontorova model and obtain periodic and heteroclinic mountain pass solutions. Heteroclinic mountain pass solution in the second laminations is new to the generalized Frenkel-Kontorova model. Our proof follows that of Bolotin and Rabinowitz for an Allen-Cahn equation, which is different with heat flow method for finding critical point of Frenkel-Kontorova model in the literature. The proofs depend on suitable choices of functionals and working spaces. We also study the multiplicity of these mountain pass solutions.

Key words and phrases:
Mountain pass solution; periodic solution; heteroclinic solution; Frenkel-Kontorova model
2020 Mathematics Subject Classification:
Primary: 49J35; Secondary: 74G22, 74G35

1.  Introduction

In this paper, we study a generalized Frenkel-Kontorova (or FK) model. To introduce the FK model, we need some notations. Let 𝐢,𝐣,𝐤\mathbf{i},\mathbf{j},\mathbf{k} (resp. i,j,ki,j,k), etc. denote the elements of n\mathbb{Z}^{n} (resp. \mathbb{Z}) and define 𝐢:=j=1n|𝐢j|\left\lVert\mathbf{i}\right\rVert:=\sum_{j=1}^{n}|\mathbf{i}_{j}|. Fix rr\in\mathbb{N} and set B𝟎r={𝐤n|𝐤r}B_{\mathbf{0}}^{r}=\{\mathbf{k}\in\mathbb{Z}^{n}\,|\,\left\lVert\mathbf{k}\right\rVert\leq r\}. Assume that sC2(B𝟎r,)s\in C^{2}(\mathbb{R}^{B_{\mathbf{0}}^{r}},\mathbb{R}) satisfies (cf. [16]):

  1. (S1)

    s(u+1B𝟎r)=s(u)s(u+1_{B_{\mathbf{0}}^{r}})=s(u), where 1B𝟎r1_{B_{\mathbf{0}}^{r}} is the constant function 11 on B𝟎rB_{\mathbf{0}}^{r};

  2. (S2)

    ss is bounded from below and coercive in the following sence,

    lim|u(𝐤)u(𝐣)|s(u)=, for 𝐤,𝐣B𝟎r with 𝐤𝐣=1;\lim_{|u(\mathbf{k})-u(\mathbf{j})|\to\infty}s(u)=\infty,\textrm{ for $\mathbf{k},\mathbf{j}\in B_{\mathbf{0}}^{r}$ with $\left\lVert\mathbf{k}-\mathbf{j}\right\rVert=1$;}
  3. (S3)

    𝐤,𝐣s0\partial_{\mathbf{k},\mathbf{j}}s\leq 0 for 𝐤,𝐣B𝟎r\mathbf{k},\mathbf{j}\in B_{\mathbf{0}}^{r} with 𝐤𝐣\mathbf{k}\neq\mathbf{j}, while 𝟎,𝐣s<0\partial_{\mathbf{0},\mathbf{j}}s<0 for 𝐣=1\left\lVert\mathbf{j}\right\rVert=1;

  4. (S4)

    there is some constant CC such that |𝐢,𝐤s|C|\partial_{\mathbf{i},\mathbf{k}}s|\leq C for all 𝐢,𝐤B𝟎r\mathbf{i},\mathbf{k}\in B_{\mathbf{0}}^{r}.

For unu\in\mathbb{R}^{\mathbb{Z}^{n}}, set

S𝐣(u)=s(τ𝐣nnτ𝐣11u|B𝟎r),S_{\mathbf{j}}(u)=s(\tau_{-\mathbf{j}_{n}}^{n}\cdots\tau_{-\mathbf{j}_{1}}^{1}u|_{B_{\mathbf{0}}^{r}}),

where τkj:nn\tau_{-k}^{j}:\mathbb{R}^{\mathbb{Z}^{n}}\to\mathbb{R}^{\mathbb{Z}^{n}} is defined by (τkju)(𝐢)=u(𝐢+k𝐞j)(\tau_{-k}^{j}u)(\mathbf{i})=u(\mathbf{i}+k\mathbf{e}_{j}) with 𝐞j=(0,,1,,0)\mathbf{e}_{j}=(0,\cdots,1,\cdots,0), i.e., the jjth component is 11 and others 0. With these local potentials S𝐣S_{\mathbf{j}}, we can define the formal sum

𝐣nS𝐣(u)\sum_{\mathbf{j}\in\mathbb{Z}^{n}}S_{\mathbf{j}}(u)

and its Euler-Lagrange equation

𝐣n𝐢S𝐣(u)=𝐣:𝐣𝐢r𝐢S𝐣(u)=0, for all 𝐢n.\sum_{\mathbf{j}\in\mathbb{Z}^{n}}\partial_{\mathbf{i}}S_{\mathbf{j}}(u)=\sum_{\mathbf{j}:\left\lVert\mathbf{j}-\mathbf{i}\right\rVert\leq r}\partial_{\mathbf{i}}S_{\mathbf{j}}(u)=0,\quad\textrm{ for all }\mathbf{i}\in\mathbb{Z}^{n}. (1.1)

(1.1) is the equation of our generalized FK moldel.

FK model was first proposed in 1938 ([10]), since then it “has become one of the fundamental and universal tools of low-dimensional nonlinear physics” ([6, p. VII, line 16]). FK model is constituted by a chain of atoms subjected to a periodic potential and is described by the following equation:

d2udt2(i)[u(i+1)+u(i1)2u(i)]+V(u(i))=0,for all i.\frac{\mathrm{d}^{2}u}{\mathrm{d}t^{2}}(i)-[u(i+1)+u(i-1)-2u(i)]+V^{\prime}(u(i))=0,\quad\textrm{for all $i\in\mathbb{Z}$.}

Here VC2(,)V\in C^{2}(\mathbb{R},\mathbb{R}) is 11-periodic. Equilibrium or stationary state of FK model is a function u:u:\mathbb{Z}\to\mathbb{R} satisfying

[u(i+1)+u(i1)2u(i)]+V(u(i))=0,for all i.-[u(i+1)+u(i-1)-2u(i)]+V^{\prime}(u(i))=0,\quad\textrm{for all $i\in\mathbb{Z}$.} (1.2)

Our generalized FK model (1.1) is a generalization of (1.2) by setting

s(u|B01)=18{[u(1)u(0)]2+[u(1)u(0)]2}+V(u(0))s(u|_{B_{0}^{1}})=\frac{1}{8}\left\{[u(1)-u(0)]^{2}+[u(-1)-u(0)]^{2}\right\}+V(u(0))

and Sj(u)=s(σju|B01)S_{j}(u)=s(\sigma_{-j}u|_{B_{0}^{1}}), where (σju)()=u(+j)(\sigma_{-j}u)(\cdot)=u(\cdot+j). So solutions of (1.1) are also called equilibrium or stationary state of the generalized FK model.

In 1983, Aubry and Le Daeron ([1]) studied minimal solutions of (1.2) and obtained the classification of minimal solutions. Minimal solutions are one of the important classes of equilibrium state. For (1.2), a function uu is said to be minimal if

j(Sj(u+v)Sj(u))0\sum_{j\in\mathbb{Z}}\left(S_{j}(u+v)-S_{j}(u)\right)\geq 0 (1.3)

for any vv with {i|v(i)0}\{i\in\mathbb{Z}\,|\,v(i)\neq 0\} a finite set. Aubry and Le Daeron found that minimal solution uu did not cross with any of its translation u(j)+lu(\cdot-j)+l, which led to an oriented homeomorphism map of a circle and then a rotation number. Using rotation number Aubry and Le Daeron made the classification of minimal solutions. Now their results are called Aubry-Mather theory because Mather ([13]) obtained similar results for monotone twist maps of annulus.

After the establishment of Aubry-Mather theory, Moser [15] attempted to generalize this theory to elliptic PDE. He found that for higher dimensional space, minimal solution might cross with its translation. So he posed another property, i.e., without self-intersections on minimal solution. In other words, Moser asked uu satisfied one and only one of the following inequality holds:

u(xj𝐞k)+l>u,or u(xj𝐞k)+l=u,or u(xj𝐞k)+l<u.u(x-j\mathbf{e}_{k})+l>u,\quad\textrm{or }\,\,u(x-j\mathbf{e}_{k})+l=u,\quad\textrm{or }\,\,u(x-j\mathbf{e}_{k})+l<u. (1.4)

Moser and then Bangert ([2]) studied a class of elliptic PDE and they obtained similar results of Aubry-Mather theory. Now their results are called Moser-Bangert theory ([17]). Bolotin, Rabinowitz, Stredulinsky ([17, 3, 4, 5]) studied an Allen-Cahn equation, which belonged to the elliptic PDE of Moser and Bangert. They used variational methods to construct more homoclinic and heteroclinic solutions of the Allen-Cahn equation other than Moser and Bangert’s.

In [16, 14, 11], Birkhoff minimizers (corresponding to minimal and without self-intersections solutions in Moser-Bangert theory) have been established and multitransition solutions was constructed in [12]. In this paper, we shall use the methods of [3, 4] to establish a new type of solution, mountain pass solution. Noting that in [16, 14, 11, 12], (1.1) was studied without the assumption (S4) except in [16]. Our results can be seen as a new proof and a refinement of some results of [16] (see also [8, 9]). Note that we only consider the case of rotation vector αn\alpha\in\mathbb{Q}^{n}. In [16], Mountain Pass Theorem was also used to establish critical point. But to prove Mountain Pass Theorem (cf. [16, Lemma 8.6]), Mramor and Rink asked the functional to be a Morse function. If the functional is a Morse function, they obtained a ghost circle which contained a periodic mountain pass solution. When the functional is not a Morse function, using a limiting progress, Mramor and Rink established a ghost circle that contained a stationary solution. If a gap of periodic minimal and Birkhoff solutions is not filled up by minimal solutions, the above stationary solution should be not minimal. The proofs of this paper are more direct than that of [16]. We also establish heteroclinic mountain pass solution in the second laminations (please see [14] for the definition of second laminations) while Mramor and Rink’s result only holds for the “first” lamination.

But we point out that in [16], the authors obtained non-minimal solution for rotation vector αnn\alpha\in\mathbb{R}^{n}\setminus\mathbb{Q}^{n} such that the Aubry-Mather set had gap, provided that ghost circle was not consists of minimizers. In [8, 9], the authors showed that there was some critical point in the gap of ground states of some FK model for any rotation vector αn\alpha\in\mathbb{R}^{n} such that gap (in the “first” lamination) condition held. Our result does not cover these cases and we limit ourselves in the case that αn\alpha\in\mathbb{Q}^{n}. We also prove the multiplicity of mountain pass solutions which is new to this generalized FK model. Other FK type models (cf. [8, 9] and references there in) may be studied using the method of the present paper and will be considered in the future.

This paper is organized as follows. We introduce some definitions and lemmas in Section 2. In Section 3, periodic mountain pass solution is established and it is proved that there are infinitely many solutions of this type. Heteroclinic mountain pass solution is considered and the multiplicity is studied in Section 4. In Appendix A, we present the detailed proofs of some properties of Section 3. A heat flow method for proving the existence of mountain pass solution is also included in Appendix A.

2.  Preliminary

We review some definitions and some lemmas of [16, 14, 11, 12]. Assume ss satisfies (S1)-(S3) in this section. For functions u,vnu,v\in\mathbb{R}^{\mathbb{Z}^{n}}, v<uv<u means v(𝐢)<u(𝐢)v(\mathbf{i})<u(\mathbf{i}) for all 𝐢n\mathbf{i}\in\mathbb{Z}^{n}, and similarly one define =,>,,,=,>,\geq,\leq, etc. The following lemmas provide important comparison results.

Lemma 2.1 (cf. [14, Lemma 2.6]).

For u,vnu,v\in\mathbb{R}^{\mathbb{Z}^{n}} and an arbitrary finite set BnB\subset\mathbb{Z}^{n}, we have

𝐣BS𝐣(max(u,v))+𝐣BS𝐣(min(u,v))𝐣BS𝐣(u)+𝐣BS𝐣(v).\sum_{\mathbf{j}\in B}S_{\mathbf{j}}(\max(u,v))+\sum_{\mathbf{j}\in B}S_{\mathbf{j}}(\min(u,v))\leq\sum_{\mathbf{j}\in B}S_{\mathbf{j}}(u)+\sum_{\mathbf{j}\in B}S_{\mathbf{j}}(v).
Lemma 2.2 (cf. [11, (3.1)]).

If {un}n\{u_{n}\}_{n\in\mathbb{N}} satisfies vunwv\leq u_{n}\leq w for fixed v,wnv,w\in\mathbb{R}^{\mathbb{Z}^{n}}, then there is a subsequence of {un}n\{u_{n}\}_{n\in\mathbb{N}} converging pointwise.

Lemma 2.3 (cf. [14, Lemma 2.5]; [16, Lemma 4.5]).

Assume that uu and vv are solutions of (1.1) and uvu\leq v. Then either u<vu<v or u=vu=v.

A function uu is said to have bounded action if there exists C>0C>0, such that |u(𝐤)u(𝐣)|C|u(\mathbf{k})-u(\mathbf{j})|\leq C for all 𝐤,𝐣n\mathbf{k},\mathbf{j}\in\mathbb{Z}^{n} with 𝐤𝐣=1\left\lVert\mathbf{k}-\mathbf{j}\right\rVert=1 (cf. [14, p. 1525, line -3], [11, p. 1112, line -8]).

Lemma 2.4 (cf. [14, Lemma 2.4], [11, Lemma 2.11]).

Assume u,vnu,v\in\mathbb{R}^{\mathbb{Z}^{n}} have bounded action with bounded constant CC. Then there exists a constant L=L(C,r)>0L=L(C,r)>0 such that for any finite set BnB\subset\mathbb{Z}^{n},

|𝐣BS𝐣(u)𝐣BS𝐣(v)|L𝐣B¯|(uv)(𝐣)|.|\sum_{\mathbf{j}\in B}S_{\mathbf{j}}(u)-\sum_{\mathbf{j}\in B}S_{\mathbf{j}}(v)|\leq L\sum_{\mathbf{j}\in\bar{B}}|(u-v)(\mathbf{j})|.

Here the closure of a set BB is defined by B¯=𝐣B{𝐤n|𝐤𝐣r}\bar{B}=\cup_{\mathbf{j}\in B}\{\mathbf{k}\in\mathbb{Z}^{n}\,|\,\left\lVert\mathbf{k}-\mathbf{j}\right\rVert\leq r\}.

Similar to (1.3) and (1.4), we introduce the following definition.

Definition 2.5.
  • (cf. [16, Definition 2.3]) A function u:nu:\mathbb{Z}^{n}\to\mathbb{R} is said to be minimal for potentials S𝐣S_{\mathbf{j}} (or for potential ss) if for every finite subset BnB\subset\mathbb{Z}^{n} and every v:nv:\mathbb{Z}^{n}\to\mathbb{R} with support, denoted by supp(v)supp(v), included in intr(B)int_{r}(B),

    𝐣B(S𝐣(u+v)S𝐣(u))0,\sum_{\mathbf{j}\in B}(S_{\mathbf{j}}(u+v)-S_{\mathbf{j}}(u))\geq 0,

    where the support of vv is supp(v):={𝐢n|v(𝐢)0}supp(v):=\{\mathbf{i}\in\mathbb{Z}^{n}\,|\,v(\mathbf{i})\neq 0\} and interior of BB is intr(B)={𝐢B|𝐢+B𝟎rB}int_{r}(B)=\{\mathbf{i}\in B\,|\,\mathbf{i}+B_{\mathbf{0}}^{r}\subset B\}.

  • (cf. [14, Definition 2.1], [17, p.3, line 25]) A function uu is said to be Birkhoff if {τjku|j and 1kn}\{\tau_{j}^{k}u\,|\,j\in\mathbb{Z}\textrm{ and }1\leq k\leq n\} is totally ordered, i.e., for all jj\in\mathbb{Z} and 1kn1\leq k\leq n, it follows that

    τjku<u, or τjku=u, or τjku>u.\tau_{j}^{k}u<u,\textrm{\quad or \quad}\tau_{j}^{k}u=u,\textrm{\quad or \quad}\tau_{j}^{k}u>u.

For 𝐩=(𝐩1,,𝐩n)n\mathbf{p}=(\mathbf{p}_{1},\cdots,\mathbf{p}_{n})\in\mathbb{N}^{n}, let

n/(𝐩n):={u:n|u(𝐢+𝐩j𝐞j)=u(𝐢),for any j{1,,n} and 𝐢n}.\begin{split}&\mathbb{R}^{\mathbb{Z}^{n}/(\mathbf{p}\mathbb{Z}^{n})}\\ :=&\{u:\mathbb{Z}^{n}\to\mathbb{R}\,|\,u(\mathbf{i}+\mathbf{p}_{j}\cdot\mathbf{e}_{j})=u(\mathbf{i}),\quad\textrm{for any $j\in\{1,\cdots,n\}$ and $\mathbf{i}\in\mathbb{Z}^{n}$}\}.\end{split}

If 𝐩=(1,,1)n\mathbf{p}=(1,\cdots,1)\in\mathbb{N}^{n}, we use n/n\mathbb{R}^{\mathbb{Z}^{n}/\mathbb{Z}^{n}} to replace n/(𝐩n).\mathbb{R}^{\mathbb{Z}^{n}/(\mathbf{p}\mathbb{Z}^{n})}. Similarly for 𝐪=(𝐪2,,𝐪n)n1\mathbf{q}=(\mathbf{q}_{2},\cdots,\mathbf{q}_{n})\in\mathbb{N}^{n-1}, one define ×n1/(𝐪n1)\mathbb{R}^{\mathbb{Z}\times\mathbb{Z}^{n-1}/(\mathbf{q}\mathbb{Z}^{n-1})}, which consists of functions that is periodic in 𝐢2,,𝐢n\mathbf{i}_{2},\cdots,\mathbf{i}_{n} with periods 𝐪2,,𝐪n\mathbf{q}_{2},\cdots,\mathbf{q}_{n}.

2.1. Periodic minimal and Birkhoff solutions

For un/nu\in\mathbb{R}^{\mathbb{Z}^{n}/\mathbb{Z}^{n}}, define J0(u):=S𝟎(u)J_{0}(u):=S_{\mathbf{0}}(u), c0:=infun/nJ0(u)c_{0}:=\inf_{u\in\mathbb{R}^{\mathbb{Z}^{n}/\mathbb{Z}^{n}}}J_{0}(u) and 0:={un/n|J0(u)=c0}\mathcal{M}_{0}:=\{u\in\mathbb{R}^{\mathbb{Z}^{n}/\mathbb{Z}^{n}}\,|\,J_{0}(u)=c_{0}\}. It was proved in [11] that 0()\mathcal{M}_{0}(\neq\emptyset) was ordered and consisted of minimal and Birkhoff solutions of (1.1). Replacing n/n\mathbb{R}^{\mathbb{Z}^{n}/\mathbb{Z}^{n}} by n/(𝐩n)\mathbb{R}^{\mathbb{Z}^{n}/(\mathbf{p}\mathbb{Z}^{n})} and minimizing the corresponding functional, we do not obtain more periodic solutions, as stated in the following.

For 𝐩=(𝐩1,,𝐩n)n\mathbf{p}=(\mathbf{p}_{1},\cdots,\mathbf{p}_{n})\in\mathbb{N}^{n}, let

𝕋0𝐩:={0,,𝐩11}×{0,,𝐩21}××{0,,𝐩n1}.\mathbb{T}^{\mathbf{p}}_{0}:=\{0,\cdots,\mathbf{p}_{1}-1\}\times\{0,\cdots,\mathbf{p}_{2}-1\}\times\cdots\times\{0,\cdots,\mathbf{p}_{n}-1\}.

and Γ0𝐩:=n/(𝐩n)\Gamma_{0}^{\mathbf{p}}:=\mathbb{R}^{\mathbb{Z}^{n}/(\mathbf{p}\mathbb{Z}^{n})}. For uΓ0𝐩u\in\Gamma_{0}^{\mathbf{p}}, define

J0𝐩(u):=𝐣𝕋0𝐩S𝐣(u).J^{\mathbf{p}}_{0}(u):=\sum_{\mathbf{j}\in\mathbb{T}^{\mathbf{p}}_{0}}S_{\mathbf{j}}(u). (2.1)

The following lemma was proved in [11] by Moser’s method (cf. [15], see also [17, Proposition 2.2]).

Lemma 2.6 (cf. [11, Proposition 3.1]).

Let 𝐩n\mathbf{p}\in\mathbb{N}^{n} and c0𝐩:=infuΓ0𝐩J0𝐩(u).c_{0}^{\mathbf{p}}:=\inf_{u\in\Gamma_{0}^{\mathbf{p}}}J_{0}^{\mathbf{p}}(u). Then 0𝐩:={uΓ0𝐩|J0𝐩(u)=c0𝐩}.\mathcal{M}_{0}^{\mathbf{p}}:=\{u\in\Gamma_{0}^{\mathbf{p}}\,|\,J_{0}^{\mathbf{p}}(u)=c_{0}^{\mathbf{p}}\}\neq\emptyset. Moreover, 0𝐩=0\mathcal{M}_{0}^{\mathbf{p}}=\mathcal{M}_{0} and c0𝐩=(i=1n𝐩i)c0c_{0}^{\mathbf{p}}=(\prod_{i=1}^{n}\mathbf{p}_{i})c_{0}.

Suppose that 0\mathcal{M}_{0} constitutes a lamination, or in other words, there is a gap in 0\mathcal{M}_{0}, i.e.,

there are v0,w00 with v0<w0 such that v0,w0 are adjacent.\textrm{there are $v_{0},w_{0}\in\mathcal{M}_{0}$ with $v_{0}<w_{0}$ such that $v_{0},w_{0}$ are adjacent}. (0*_{0})

Here adjacent means there does not exist u0u\in\mathcal{M}_{0} such that v0uw0v_{0}\leq u\leq w_{0}. In [11], heteroclinic minimal and Birkhoff solutions are constructed under condition (0*_{0}).

2.2. Heteroclinic minimal and Birkhoff solutions in 𝐢1\mathbf{i}_{1}

To construct heteroclinic minimal and Birkhoff solutions, assume that (0*_{0}) holds. Let 𝐓i=i𝐞1\mathbf{T}_{i}=i\mathbf{e}_{1}. Set Γ^1(v0,w0):={u×n1/n1|v0uw0}\hat{\Gamma}_{1}(v_{0},w_{0}):=\{u\in\mathbb{R}^{\mathbb{Z}\times\mathbb{Z}^{n-1}/\mathbb{Z}^{n-1}}\,|\,v_{0}\leq u\leq w_{0}\}. For uΓ^1(v0,w0)u\in\hat{\Gamma}_{1}(v_{0},w_{0}), define J1;p,q(u):=i=pq[J0(τi1u)c0]J_{1;p,q}(u):=\sum_{i=p}^{q}[J_{0}(\tau_{-i}^{1}u)-c_{0}], then it was proved in [11, Proposition 3.2] that J1;p,q(u)K1J_{1;p,q}(u)\geq-K_{1} for some K1=K1(v0,w0)0K_{1}=K_{1}(v_{0},w_{0})\geq 0. Thus we can define

J1(u):=lim infpqJ1;p,q(u),J_{1}(u):=\liminf_{p\to-\infty\atop q\to\infty}J_{1;p,q}(u), (2.2)

and we have (by [11, Lemma 3.3])

J1;p,q(u)J1(u)+2K1.J_{1;p,q}(u)\leq J_{1}(u)+2K_{1}. (2.3)

Set

Γ1(v0,w0):={uΓ^1(v0,w0)|limi|(uv0)(𝐓i)|=0,limi|(uw0)(𝐓i)|=0}.\begin{split}\Gamma_{1}(v_{0},w_{0}):=\{u\in\hat{\Gamma}_{1}(v_{0},w_{0})\,|\,&\lim_{i\to-\infty}|(u-v_{0})(\mathbf{T}_{i})|=0,\\ &\lim_{i\to\infty}|(u-w_{0})(\mathbf{T}_{i})|=0\}.\end{split}

For uΓ1(v0,w0)u\in\Gamma_{1}(v_{0},w_{0}), as was proved in [11, Proposition 3.4], if J1(u)<J_{1}(u)<\infty, then

J1(u)=limpqJ1;p,q(u),i.e., J1(u)=i[J0(τi1u)c0].J_{1}(u)=\lim_{p\to-\infty\atop q\to\infty}J_{1;p,q}(u),\quad\textrm{i.e., }J_{1}(u)=\sum_{i\in\mathbb{Z}}[J_{0}(\tau_{-i}^{1}u)-c_{0}].

In other words, lim inf\liminf becomes lim\lim in the definition of J1(u)J_{1}(u). Set

c1(v0,w0):=infuΓ1(v0,w0)J1(u).c_{1}(v_{0},w_{0}):=\inf_{u\in\Gamma_{1}(v_{0},w_{0})}J_{1}(u).

Then, as was proved in [11], c1(v0,w0)c_{1}(v_{0},w_{0}) is attained and

1(v0,w0):={uΓ1(v0,w0)|J1(u)=c1(v0,w0)}\mathcal{M}_{1}(v_{0},w_{0}):=\{u\in\Gamma_{1}(v_{0},w_{0})\,|\,J_{1}(u)=c_{1}(v_{0},w_{0})\}

is an ordered set and consists of heteroclinic minimal and Birkhoff solutions of (1.1). Moreover, we have

Lemma 2.7 (cf. [12, Proposition 2.13]).

Suppose (0*_{0}) holds and uΓ^1(v0,w0)u\in\hat{\Gamma}_{1}(v_{0},w_{0}) with J1(u)<J_{1}(u)<\infty. If uu satisfies (1.1) for 𝐢1R\mathbf{i}_{1}\geq R (resp. 𝐢1R\mathbf{i}_{1}\leq-R), then |(uϕ)(𝐓i)|0|(u-\phi)(\mathbf{T}_{i})|\to 0 as ii\to\infty (resp. |(uϕ)(𝐓i)|0|(u-\phi)(\mathbf{T}_{i})|\to 0 as i)i\to-\infty), where RR\in\mathbb{N} and ϕ=v0\phi=v_{0} or w0w_{0}.

Similar to Section 2.1, varying the periods of function in Γ1(v0,w0)\Gamma_{1}(v_{0},w_{0}) cannot produce more minimal and Birkhoff solution. To see this, for 𝐪=(𝐪2,,𝐪n)n1\mathbf{q}=(\mathbf{q}_{2},\cdots,\mathbf{q}_{n})\in\mathbb{N}^{n-1} let

𝕋1𝐪:={0,,𝐪21}××{0,,𝐪n1}.\mathbb{T}_{1}^{\mathbf{q}}:=\{0,\cdots,\mathbf{q}_{2}-1\}\times\cdots\times\{0,\cdots,\mathbf{q}_{n}-1\}.

Set

Γ1𝐪(v0,w0):={u×n1/(𝐪n1)|v0uw0,limi𝐣{i}×𝕋1𝐪|(uv0)(𝐣)|=0,limi𝐣{i}×𝕋1𝐪|(uw0)(𝐣)|=0}.\begin{split}\Gamma_{1}^{\mathbf{q}}(v_{0},w_{0}):=\{u\in\mathbb{R}^{\mathbb{Z}\times\mathbb{Z}^{n-1}/(\mathbf{q}\mathbb{Z}^{n-1})}\,|\,&v_{0}\leq u\leq w_{0},\\ &\lim_{i\to-\infty}\sum_{\mathbf{j}\in\{i\}\times\mathbb{T}_{1}^{\mathbf{q}}}|(u-v_{0})(\mathbf{j})|=0,\\ &\lim_{i\to\infty}\sum_{\mathbf{j}\in\{i\}\times\mathbb{T}_{1}^{\mathbf{q}}}|(u-w_{0})(\mathbf{j})|=0\}.\end{split} (2.4)

For uΓ1𝐪(v0,w0)u\in\Gamma_{1}^{\mathbf{q}}(v_{0},w_{0}), define

J1;p,q𝐪(u):=i=pqJ1,i𝐪(u):=i=pq𝐣{i}×𝕋1𝐪[S𝐣(u)c0],J_{1;p,q}^{\mathbf{q}}(u):=\sum_{i=p}^{q}J_{1,i}^{\mathbf{q}}(u):=\sum_{i=p}^{q}\sum_{\mathbf{j}\in\{i\}\times\mathbb{T}_{1}^{\mathbf{q}}}\left[S_{\mathbf{j}}(u)-c_{0}\right],

and

J1𝐪(u):=lim infpqJ1;p,q𝐪(u).J_{1}^{\mathbf{q}}(u):=\liminf_{p\to-\infty\atop q\to\infty}J_{1;p,q}^{\mathbf{q}}(u).

Similar to (2.2) J1𝐪(u)J_{1}^{\mathbf{q}}(u) is well-defined and it satisfies

Lemma 2.8 (cf. [11, Proposition 3.4]).

For uΓ1𝐪(v0,w0)u\in\Gamma_{1}^{\mathbf{q}}(v_{0},w_{0}), if J1𝐪(u)<J_{1}^{\mathbf{q}}(u)<\infty,

J1𝐪(u)=limpqJ1;p,q𝐪(u),i.e., J1𝐪(u)=iJ1,i𝐪(u).J_{1}^{\mathbf{q}}(u)=\lim_{p\to-\infty\atop q\to\infty}J_{1;p,q}^{\mathbf{q}}(u),\quad\textrm{i.e., }J_{1}^{\mathbf{q}}(u)=\sum_{i\in\mathbb{Z}}J_{1,i}^{\mathbf{q}}(u).
Remark 2.9.

Suppose that v,w0v,w\in\mathcal{M}_{0} satisfy vv0<w0wv\leq v_{0}<w_{0}\leq w and v,wv,w may be not adjacent. A careful reading of the proof of [11, Proposition 3.4] tells us that for

u{u×n1/(𝐪n1)|vuw,limi𝐣{i}×𝕋1𝐪|(uv0)(𝐣)|=0,limi𝐣{i}×𝕋1𝐪|(uw0)(𝐣)|=0},\begin{split}u\in\{u\in\mathbb{R}^{\mathbb{Z}\times\mathbb{Z}^{n-1}/(\mathbf{q}\mathbb{Z}^{n-1})}\,|\,&v\leq u\leq w,\\ &\lim_{i\to-\infty}\sum_{\mathbf{j}\in\{i\}\times\mathbb{T}_{1}^{\mathbf{q}}}|(u-v_{0})(\mathbf{j})|=0,\\ &\lim_{i\to\infty}\sum_{\mathbf{j}\in\{i\}\times\mathbb{T}_{1}^{\mathbf{q}}}|(u-w_{0})(\mathbf{j})|=0\},\end{split}

J1𝐪(u)J_{1}^{\mathbf{q}}(u) is well-defined and Lemma 2.8 holds.

Similar to Lemma 2.6, we have

Lemma 2.10 (cf. [11, Proposition 3.20]).

Let 𝐪=(𝐪2,,𝐪n)n1\mathbf{q}=(\mathbf{q}_{2},\cdots,\mathbf{q}_{n})\in\mathbb{N}^{n-1} and c1𝐪(v0,w0):=infuΓ1𝐪(v0,w0)J1𝐪(u)c_{1}^{\mathbf{q}}(v_{0},w_{0}):=\inf_{u\in\Gamma_{1}^{\mathbf{q}}(v_{0},w_{0})}J_{1}^{\mathbf{q}}(u). Then 1𝐪:={uΓ1𝐪(v0,w0)|J1𝐪(u)=c1𝐪(v0,w0)}.\mathcal{M}_{1}^{\mathbf{q}}:=\{u\in\Gamma_{1}^{\mathbf{q}}(v_{0},w_{0})\,|\,J_{1}^{\mathbf{q}}(u)=c_{1}^{\mathbf{q}}(v_{0},w_{0})\}\neq\emptyset. Moreover, 1𝐪(v0,w0)=1(v0,w0)\mathcal{M}_{1}^{\mathbf{q}}(v_{0},w_{0})=\mathcal{M}_{1}(v_{0},w_{0}), and c1𝐪(v0,w0)=(i=2n𝐪i)c1(v0,w0)c_{1}^{\mathbf{q}}(v_{0},w_{0})=(\prod_{i=2}^{n}\mathbf{q}_{i})c_{1}(v_{0},w_{0}).

In analogy with (0*_{0}), assume

there are v1,w11(v0,w0) with v1<w1 such that v1,w1 are adjacent.\textrm{there are }v_{1},w_{1}\in\mathcal{M}_{1}(v_{0},w_{0})\textrm{ with }v_{1}<w_{1}\textrm{ such that $v_{1},w_{1}$ are adjacent}. (1*_{1})

In Section 3 we shall establish the existence of periodic mountain pass solution in the gap of v0,w0v_{0},w_{0} while in Section 4, we shall construct heteroclinic mountain pass solution in the gap of v1,w1v_{1},w_{1}.

3.  Mountain pass solutions in the gap of 0\mathcal{M}_{0}

Assume that ss satisfies (S1)-(S4) in this and the following two sections. We establish periodic mountain pass solution of (1.1) in this section. Firstly we introduce the working space and the corresponding functional. For 𝐩n\mathbf{p}\in\mathbb{N}^{n}, set

Λ0𝐩:={un/(𝐩n)|uΛ0𝐩2:=𝐣𝕋0𝐩|u(𝐣)|2<}.\Lambda_{0}^{\mathbf{p}}:=\left\{u\in\mathbb{R}^{\mathbb{Z}^{n}/(\mathbf{p}\mathbb{Z}^{n})}\,\Big{|}\,\left\lVert u\right\rVert^{2}_{\Lambda_{0}^{\mathbf{p}}}:=\sum_{\mathbf{j}\in\mathbb{T}_{0}^{\mathbf{p}}}|u(\mathbf{j})|^{2}<\infty\right\}.

It is easy to see that (Λ0𝐩,Λ0𝐩)(\Lambda_{0}^{\mathbf{p}},\left\lVert\cdot\right\rVert_{\Lambda_{0}^{\mathbf{p}}}) is a Banach space. Define J0𝐩J^{\mathbf{p}}_{0} as in Section 2.1 and assume that (0*_{0}) holds. For uΛ0𝐩u\in\Lambda_{0}^{\mathbf{p}}, set I0𝐩(u):=J0𝐩(u+v0)I^{\mathbf{p}}_{0}(u):=J^{\mathbf{p}}_{0}(u+v_{0}). Then since sC2(B𝟎r,)s\in C^{2}(\mathbb{R}^{B_{\mathbf{0}}^{r}},\mathbb{R}), I0𝐩C1(Λ0𝐩,)I^{\mathbf{p}}_{0}\in C^{1}(\Lambda_{0}^{\mathbf{p}},\mathbb{R}) and

(I0𝐩)(u)v=𝐣𝕋0𝐩𝐤:𝐤𝐣r𝐤S𝐣(u+v0)v(𝐤)=𝐣𝕋0𝐩v(𝐣)𝐤:𝐤𝐣r𝐤S𝐣(u+v0),\begin{split}(I_{0}^{\mathbf{p}})^{\prime}(u)v=&\sum_{\mathbf{j}\in\mathbb{T}^{\mathbf{p}}_{0}}\sum_{\mathbf{k}:\left\lVert\mathbf{k}-\mathbf{j}\right\rVert\leq r}\partial_{\mathbf{k}}S_{\mathbf{j}}(u+v_{0})v(\mathbf{k})\\ =&\sum_{\mathbf{j}\in\mathbb{T}^{\mathbf{p}}_{0}}v(\mathbf{j})\sum_{\mathbf{k}:\left\lVert\mathbf{k}-\mathbf{j}\right\rVert\leq r}\partial_{\mathbf{k}}S_{\mathbf{j}}(u+v_{0}),\end{split} (3.1)

where (I0𝐩)(I_{0}^{\mathbf{p}})^{\prime} is the Fréchet derivative of I0𝐩I_{0}^{\mathbf{p}}. If (I0𝐩)(u)=0(I_{0}^{\mathbf{p}})^{\prime}(u)=0, then

𝐤:𝐤𝐣r𝐤S𝐣(u+v0)=0\sum_{\mathbf{k}:\left\lVert\mathbf{k}-\mathbf{j}\right\rVert\leq r}\partial_{\mathbf{k}}S_{\mathbf{j}}(u+v_{0})=0

hold for all 𝐣𝕋0𝐩\mathbf{j}\in\mathbb{T}^{\mathbf{p}}_{0}. Hence by the periodicities of uu and v0v_{0}, u+v0u+v_{0} is a solution of (1.1).

3.1. Periodic mountain pass solution

Consider the semiflow Φt0:Λ0𝐩Λ0𝐩\Phi_{t}^{0}:\Lambda_{0}^{\mathbf{p}}\rightarrow\Lambda_{0}^{\mathbf{p}}, which is defined by

{tΦt0(u)(𝐢)=𝐣:𝐣𝐢r𝐢S𝐣(Φt0(u)+v0),for t>0,Φ00(u)(𝐢)=u(𝐢).\left\{\begin{array}[]{ll}-\partial_{t}\Phi_{t}^{0}(u)(\mathbf{i})&=\sum_{\mathbf{j}:\left\lVert\mathbf{j}-\mathbf{i}\right\rVert\leq r}\partial_{\mathbf{i}}S_{\mathbf{j}}(\Phi_{t}^{0}(u)+v_{0}),\quad\quad\textrm{for }t>0,\\ \Phi_{0}^{0}(u)(\mathbf{i})&=u(\mathbf{i}).\end{array}\right. (3.2)

Set W(u)(𝐢)=𝐣:𝐣𝐢r𝐢S𝐣(u+v0)W(u)(\mathbf{i})=\sum_{\mathbf{j}:\left\lVert\mathbf{j}-\mathbf{i}\right\rVert\leq r}\partial_{\mathbf{i}}S_{\mathbf{j}}(u+v_{0}), then W(u)Λ0𝐩W(u)\in\Lambda_{0}^{\mathbf{p}} for any uΛ0𝐩u\in\Lambda_{0}^{\mathbf{p}}. For u,vΛ0𝐩u,v\in\Lambda_{0}^{\mathbf{p}},

W(u)W(v)Λ0𝐩2=𝐢𝕋0𝐩|W(u)(𝐢)W(v)(𝐢)|2=𝐢𝕋0𝐩|𝐣:𝐣𝐢r[𝐢S𝐣(u+v0)𝐢S𝐣(v+v0)]|2=𝐢𝕋0𝐩|𝐣:𝐣𝐢r[01ddt𝐢S𝐣(v+t(uv)+v0)dt]|2=𝐢𝕋0𝐩|𝐣:𝐣𝐢r𝐥:𝐥𝐣r01𝐢,𝐥S𝐣(v+t(uv)+v0)dt[u(𝐥)v(𝐥)]|2𝐢𝕋0𝐩𝐣:𝐣𝐢r𝐥:𝐥𝐣r(01𝐢,𝐥S𝐣(v+t(uv)+v0)dt)2𝐣:𝐣𝐢r𝐥:𝐥𝐣r[u(𝐥)v(𝐥)]2C2C(r)𝐢𝕋0𝐩𝐣:𝐣𝐢r𝐥:𝐥𝐣r[u(𝐥)v(𝐥)]2C2C(r)C1(r)uvΛ0𝐩2\begin{split}&\left\lVert W(u)-W(v)\right\rVert_{\Lambda_{0}^{\mathbf{p}}}^{2}\\ =&\sum_{\mathbf{i}\in\mathbb{T}^{\mathbf{p}}_{0}}|W(u)(\mathbf{i})-W(v)(\mathbf{i})|^{2}\\ =&\sum_{\mathbf{i}\in\mathbb{T}^{\mathbf{p}}_{0}}\Big{|}\sum_{\mathbf{j}:\left\lVert\mathbf{j}-\mathbf{i}\right\rVert\leq r}[\partial_{\mathbf{i}}S_{\mathbf{j}}(u+v_{0})-\partial_{\mathbf{i}}S_{\mathbf{j}}(v+v_{0})]\Big{|}^{2}\\ =&\sum_{\mathbf{i}\in\mathbb{T}^{\mathbf{p}}_{0}}\left|\sum_{\mathbf{j}:\left\lVert\mathbf{j}-\mathbf{i}\right\rVert\leq r}\left[\int_{0}^{1}\frac{\mathrm{d}}{\mathrm{d}t}\partial_{\mathbf{i}}S_{\mathbf{j}}(v+t(u-v)+v_{0})\mathrm{d}t\right]\right|^{2}\\ =&\sum_{\mathbf{i}\in\mathbb{T}^{\mathbf{p}}_{0}}\left|\sum_{\mathbf{j}:\left\lVert\mathbf{j}-\mathbf{i}\right\rVert\leq r\atop\mathbf{l}:\left\lVert\mathbf{l}-\mathbf{j}\right\rVert\leq r}\int_{0}^{1}\partial_{\mathbf{i},\mathbf{l}}S_{\mathbf{j}}(v+t(u-v)+v_{0})\mathrm{d}t\cdot[u(\mathbf{l})-v(\mathbf{l})]\right|^{2}\\ \leq&\sum_{\mathbf{i}\in\mathbb{T}^{\mathbf{p}}_{0}}\sum_{\mathbf{j}:\left\lVert\mathbf{j}-\mathbf{i}\right\rVert\leq r\atop\mathbf{l}:\left\lVert\mathbf{l}-\mathbf{j}\right\rVert\leq r}\left(\int_{0}^{1}\partial_{\mathbf{i},\mathbf{l}}S_{\mathbf{j}}(v+t(u-v)+v_{0})\mathrm{d}t\right)^{2}\sum_{\mathbf{j}:\left\lVert\mathbf{j}-\mathbf{i}\right\rVert\leq r\atop\mathbf{l}:\left\lVert\mathbf{l}-\mathbf{j}\right\rVert\leq r}[u(\mathbf{l})-v(\mathbf{l})]^{2}\\ \leq&C^{2}\cdot C(r)\sum_{\mathbf{i}\in\mathbb{T}^{\mathbf{p}}_{0}}\sum_{\mathbf{j}:\left\lVert\mathbf{j}-\mathbf{i}\right\rVert\leq r\atop\mathbf{l}:\left\lVert\mathbf{l}-\mathbf{j}\right\rVert\leq r}[u(\mathbf{l})-v(\mathbf{l})]^{2}\\ \leq&C^{2}\cdot C(r)\cdot C_{1}(r)\left\lVert u-v\right\rVert_{\Lambda_{0}^{\mathbf{p}}}^{2}\end{split} (3.3)

where CC is the constant in (S4) and C(r),C1(r)C(r),C_{1}(r) are constants depending only on rr. By Cauchy-Lipschitz-Picard Theorem (please see e.g., [7, Theorem 7.3]), Φt0\Phi_{t}^{0} is well-defined and is C1C^{1} in tt. For Φt0\Phi_{t}^{0}, we have the following comparison result.

Proposition 3.1.

Assume u1,u2Λ0𝐩u_{1},u_{2}\in\Lambda_{0}^{\mathbf{p}}. If u1u2u_{1}\leq u_{2} and u1u2u_{1}\neq u_{2}, then Φt0(u1)<Φt0(u2)\Phi_{t}^{0}(u_{1})<\Phi_{t}^{0}(u_{2}) for all t>0t>0.

The proof of Proposition 3.1 follows from [16, Theorem 6.2] with slight modifications. For the reader’s convenience, we provide the proof of Proposition 3.1 in Appendix A. Result similar to Proposition 3.1 also appears in [9]. In [9], one need a “transitive” condition ([9, p. 2414, line 7]). In our settings, (S3) ensures this condition.

As in [4], we choose a subset of Λ0𝐩\Lambda_{0}^{\mathbf{p}} to prove the deformation lemma. Set

𝒢0𝐩={uΛ0𝐩| 0uw0v0}.\mathcal{G}^{\mathbf{p}}_{0}=\left\{u\in\Lambda_{0}^{\mathbf{p}}\,|\,0\leq u\leq w_{0}-v_{0}\right\}.

It is easy to see that 𝒢0𝐩\mathcal{G}^{\mathbf{p}}_{0} is a compact set with respect to the norm Λ0𝐩\left\lVert\cdot\right\rVert_{\Lambda_{0}^{\mathbf{p}}}, as shown in the following proposition.

Proposition 3.2.

𝒢0𝐩\mathcal{G}^{\mathbf{p}}_{0} is compact with respect to the norm Λ0𝐩\left\lVert\cdot\right\rVert_{\Lambda_{0}^{\mathbf{p}}}.

Proof.

Assume (uk)𝒢0𝐩(u_{k})\subset\mathcal{G}^{\mathbf{p}}_{0}. By the definition of Λ0𝐩\left\lVert\cdot\right\rVert_{\Lambda_{0}^{\mathbf{p}}}, we only need to prove that (uk)(u_{k}) is compact with respect to pointwise convergence, which follows from Lemma 2.2. ∎

Remark 3.3.

Proposition 3.2 will be used in the proof of deformation lemma (please see Lemma 3.4 (v) below). In the proof of [3, Proposition 3.6], Bolotin and Rabinowitz obtained “compactness” by verifying the corresponding functional satisfied Palais-Smale condition. In our settings, the compactness condition is directly obtained.

We have the following deformation lemma.

Lemma 3.4.

For the semiflow Φt0\Phi_{t}^{0} defined in (3.2), we have:

  1. (i)

    Φt0(u)=u\Phi_{t}^{0}(u)=u if (I0𝐩)(u)=0(I_{0}^{\mathbf{p}})^{\prime}(u)=0;

  2. (ii)

    I0𝐩(Φt0(u))I0𝐩(u)I_{0}^{\mathbf{p}}\left(\Phi_{t}^{0}(u)\right)\leq I_{0}^{\mathbf{p}}(u);

  3. (iii)

    Φt0𝒢0𝐩𝒢0𝐩\Phi_{t}^{0}\mathcal{G}^{\mathbf{p}}_{0}\subset\mathcal{G}^{\mathbf{p}}_{0};

  4. (iv)

    For any u𝒢0𝐩u\in\mathcal{G}^{\mathbf{p}}_{0}, there is a sequence (ti)(t_{i})\subset\mathbb{R} with tit_{i}\to\infty as ii\to\infty such that Φti0(u)U\Phi_{t_{i}}^{0}(u)\to U pointwise for some U𝒢0𝐩U\in\mathcal{G}^{\mathbf{p}}_{0}, and I0𝐩(U)=limtI0𝐩(Φt0(u))I_{0}^{\mathbf{p}}(U)=\lim_{t\to\infty}I_{0}^{\mathbf{p}}(\Phi_{t}^{0}(u)), and U+v0U+v_{0} is a solution of (1.1);

  5. (v)

    If 𝒦c:={u𝒢0𝐩|I0𝐩(u)=c,(I0𝐩)(u)=0}=\mathcal{K}_{c}:=\{u\in\mathcal{G}^{\mathbf{p}}_{0}\,|\,I_{0}^{\mathbf{p}}(u)=c,(I_{0}^{\mathbf{p}})^{\prime}(u)=0\}=\emptyset, there is an ϵ>0\epsilon>0 such that Φ10((I0𝐩)c+ϵ)(I0𝐩)cϵ\Phi_{1}^{0}((I_{0}^{\mathbf{p}})^{c+\epsilon})\subset(I_{0}^{\mathbf{p}})^{c-\epsilon}, where (I0𝐩)t:={u𝒢0𝐩|I0𝐩(u)t}(I_{0}^{\mathbf{p}})^{t}:=\{u\in\mathcal{G}^{\mathbf{p}}_{0}\,|\,I_{0}^{\mathbf{p}}(u)\leq t\}.

In [16, Lemma 8.6], the authors established Mountain Pass Theorem by imposing a condition that the functional was a Morse function. This condition is used to prove a similar property of Lemma 3.4 (v). We prove Lemma 3.4 in Appendix A. Now set

0𝐩={hC([0,1],𝒢0𝐩)|h(0)=0,h(1)=w0v0}\mathcal{H}^{\mathbf{p}}_{0}=\left\{h\in C\left([0,1],\mathcal{G}^{\mathbf{p}}_{0}\right)\,|\,h(0)=0,h(1)=w_{0}-v_{0}\right\}

and

d0𝐩=infh0𝐩maxθ[0,1]I0𝐩(h(θ)).d_{0}^{\mathbf{p}}=\inf_{h\in\mathcal{H}^{\mathbf{p}}_{0}}\max_{\theta\in[0,1]}I_{0}^{\mathbf{p}}(h(\theta)).
Proposition 3.5.

d0𝐩>c0𝐩d_{0}^{\mathbf{p}}>c_{0}^{\mathbf{p}}.

Proof.

By Lemma 2.6, for any uΛ0𝐩u\in\Lambda_{0}^{\mathbf{p}},

I0𝐩(u)c0𝐩.I_{0}^{\mathbf{p}}(u)\geq c_{0}^{\mathbf{p}}. (3.4)

So d0𝐩c0𝐩d_{0}^{\mathbf{p}}\geq c_{0}^{\mathbf{p}}. Suppose, by contradiction, d0𝐩=c0𝐩d_{0}^{\mathbf{p}}=c_{0}^{\mathbf{p}}. Then there exist hj0𝐩h_{j}\in\mathcal{H}^{\mathbf{p}}_{0} and σj(0,1)\sigma_{j}\in(0,1) such that

maxθ[0,1]I0𝐩(hj(θ))c0𝐩as j\max_{\theta\in[0,1]}I_{0}^{\mathbf{p}}\left(h_{j}(\theta)\right)\rightarrow c_{0}^{\mathbf{p}}\quad\quad\textrm{as }j\to\infty (3.5)

and

hj(σj)(𝟎)=12(w0v0)(𝟎).h_{j}\left(\sigma_{j}\right)(\mathbf{0})=\frac{1}{2}\left(w_{0}-v_{0}\right)(\mathbf{0}). (3.6)

By (3.4)-(3.5), we have

I0𝐩(hj(σj))c0𝐩as j.I_{0}^{\mathbf{p}}\left(h_{j}\left(\sigma_{j}\right)\right)\rightarrow c_{0}^{\mathbf{p}}\quad\quad\textrm{as }j\to\infty. (3.7)

Since hj(σj)𝒢0𝐩h_{j}(\sigma_{j})\in\mathcal{G}^{\mathbf{p}}_{0}, a compact set by Proposition 3.2, hj(σj)h_{j}(\sigma_{j}) has a subsequence (still denoted by hj(σj)h_{j}(\sigma_{j})) which converges in Λ0𝐩\Lambda_{0}^{\mathbf{p}} to U𝒢0𝐩U\in\mathcal{G}^{\mathbf{p}}_{0}. Since I0𝐩I_{0}^{\mathbf{p}} is continuous on Λ0𝐩\Lambda_{0}^{\mathbf{p}}, by (3.4) and (3.7),

c0𝐩I0𝐩(U)=limjI0𝐩(hj(σj))=c0𝐩.c_{0}^{\mathbf{p}}\leq I_{0}^{\mathbf{p}}(U)=\lim_{j\rightarrow\infty}I_{0}^{\mathbf{p}}\left(h_{j}\left(\sigma_{j}\right)\right)=c_{0}^{\mathbf{p}}.

Then U0𝐩=0U\in\mathcal{M}_{0}^{\mathbf{p}}=\mathcal{M}_{0}. Hence U+v0=v0U+v_{0}=v_{0} or U+v0=w0U+v_{0}=w_{0}. But by (3.6),

U(𝟎)=12(w0v0)(𝟎),U(\mathbf{0})=\frac{1}{2}\left(w_{0}-v_{0}\right)(\mathbf{0}),

a contradiction. So d0𝐩>c0𝐩d_{0}^{\mathbf{p}}>c_{0}^{\mathbf{p}}. ∎

Theorem 3.6.

d0𝐩d_{0}^{\mathbf{p}} is a critical value of I0𝐩I_{0}^{\mathbf{p}} on Λ0𝐩\Lambda_{0}^{\mathbf{p}} with a corresponding critical point u𝐩u_{\mathbf{p}} satisfying 0<u𝐩<w0v00<u_{\mathbf{p}}<w_{0}-v_{0} and u𝐩+v0u_{\mathbf{p}}+v_{0} is a solution of (1.1).

Proof.

Suppose, by contradiction, 𝒦d0𝐩=\mathcal{K}_{d_{0}^{\mathbf{p}}}=\emptyset. Lemma 3.4 (v) implies that there exists ϵ>0\epsilon>0 such that

Φ10((I0𝐩)d0𝐩+ϵ)(I0𝐩)d0𝐩ϵ.\Phi_{1}^{0}((I_{0}^{\mathbf{p}})^{d_{0}^{\mathbf{p}}+\epsilon})\subset(I_{0}^{\mathbf{p}})^{d_{0}^{\mathbf{p}}-\epsilon}. (3.8)

By the definition of d0𝐩d_{0}^{\mathbf{p}}, there is an h0𝐩h\in\mathcal{H}^{\mathbf{p}}_{0} satisfying

maxθ[0,1]I0𝐩(h(θ))d0𝐩+ϵ.\max_{\theta\in[0,1]}I_{0}^{\mathbf{p}}(h(\theta))\leq d_{0}^{\mathbf{p}}+\epsilon.

Then by Lemma 3.4 (i) and (iii), we have Φ10h0𝐩\Phi_{1}^{0}\circ h\in\mathcal{H}^{\mathbf{p}}_{0}. But by (3.8),

maxθ[0,1]I0𝐩(Φ10h(θ))d0𝐩ϵ,\max_{\theta\in[0,1]}I_{0}^{\mathbf{p}}\left(\Phi_{1}^{0}\circ h(\theta)\right)\leq d_{0}^{\mathbf{p}}-\epsilon,

which is impossible by the definition of d0𝐩d_{0}^{\mathbf{p}}. Thus 𝒦d0𝐩\mathcal{K}_{d_{0}^{\mathbf{p}}}\neq\emptyset and d0𝐩d_{0}^{\mathbf{p}} is a critical value with a corresponding critical point u𝐩𝒢0𝐩u_{\mathbf{p}}\in\mathcal{G}^{\mathbf{p}}_{0}, so 0u𝐩w0v00\leq u_{\mathbf{p}}\leq w_{0}-v_{0}. Since u𝐩+v0u_{\mathbf{p}}+v_{0} is a solution of (1.1), by Lemma 2.3, v0<u𝐩+v0<w0v_{0}<u_{\mathbf{p}}+v_{0}<w_{0}. ∎

Remark 3.7.

Another proof of Theorem 3.6 by heat flow method is provided in Appendix A. We prefer the above argument because it is more intuitive.

As in [4], we prove that d0𝐩d_{0}^{\mathbf{p}} is indeed a mountain pass critical value as follows. Set

¯0𝐩={h¯C([0,1],Λ0𝐩)|h¯(0)=0,h¯(1)=w0v0}\bar{\mathcal{H}}^{\mathbf{p}}_{0}=\{\bar{h}\in C([0,1],\Lambda_{0}^{\mathbf{p}})\,|\,\bar{h}(0)=0,\bar{h}(1)=w_{0}-v_{0}\}

and

d¯0𝐩=infh¯¯0𝐩maxθ[0,1]I0𝐩(h¯(θ)).\bar{d}_{0}^{\mathbf{p}}=\inf_{\bar{h}\in\bar{\mathcal{H}}^{\mathbf{p}}_{0}}\max_{\theta\in[0,1]}I_{0}^{\mathbf{p}}\left(\bar{h}(\theta)\right).

So d¯0𝐩\bar{d}_{0}^{\mathbf{p}} is a classical mountain pass critical value. We have:

Proposition 3.8.

d0𝐩=d¯0𝐩d_{0}^{\mathbf{p}}=\bar{d}_{0}^{\mathbf{p}}.

Proof.

Obviously d0𝐩d¯0𝐩d_{0}^{\mathbf{p}}\geq\bar{d}_{0}^{\mathbf{p}}. To prove the converse inequality, for any h¯¯0𝐩\bar{h}\in\bar{\mathcal{H}}^{\mathbf{p}}_{0}, set

h(θ)=max{min(h¯(θ),w0v0),0}.h(\theta)=\max\{\min\left(\bar{h}(\theta),w_{0}-v_{0}\right),0\}.

Then h0𝐩h\in\mathcal{H}^{\mathbf{p}}_{0} and

I0𝐩(h(θ))+c0𝐩I0𝐩(h(θ))+I0𝐩(min(min(h¯(θ),w0v0),0))I0𝐩(min(h¯(θ),w0v0))+I0𝐩(0)=I0𝐩(min(h¯(θ),w0v0))+c0𝐩,\begin{split}&I_{0}^{\mathbf{p}}(h(\theta))+c_{0}^{\mathbf{p}}\\ \leq&I_{0}^{\mathbf{p}}(h(\theta))+I_{0}^{\mathbf{p}}(\min\left(\min\left(\bar{h}(\theta),w_{0}-v_{0}\right),0\right))\\ \leq&I_{0}^{\mathbf{p}}(\min\left(\bar{h}(\theta),w_{0}-v_{0}\right))+I_{0}^{\mathbf{p}}(0)\\ =&I_{0}^{\mathbf{p}}(\min\left(\bar{h}(\theta),w_{0}-v_{0}\right))+c_{0}^{\mathbf{p}},\end{split}

where the first inequality follows from Lemma 2.6 and the second follows from Lemma 2.1. Similarly, we have

I0𝐩(min(h¯(θ),w0v0))+c0𝐩I0𝐩(min(h¯(θ),w0v0))+I0𝐩(max(h¯(θ),w0v0))I0𝐩(h¯(θ))+I0𝐩(w0v0)=I0𝐩(h¯(θ))+c0𝐩,\begin{split}&I_{0}^{\mathbf{p}}(\min\left(\bar{h}(\theta),w_{0}-v_{0}\right))+c_{0}^{\mathbf{p}}\\ \leq&I_{0}^{\mathbf{p}}(\min\left(\bar{h}(\theta),w_{0}-v_{0}\right))+I_{0}^{\mathbf{p}}(\max\left(\bar{h}(\theta),w_{0}-v_{0}\right))\\ \leq&I_{0}^{\mathbf{p}}(\bar{h}(\theta))+I_{0}^{\mathbf{p}}(w_{0}-v_{0})\\ =&I_{0}^{\mathbf{p}}(\bar{h}(\theta))+c_{0}^{\mathbf{p}},\end{split}

so I0𝐩(h(θ))I0𝐩(h¯(θ))I_{0}^{\mathbf{p}}(h(\theta))\leq I_{0}^{\mathbf{p}}\left(\bar{h}(\theta)\right) for each θ\theta. ∎

Remark 3.9.

As is well-known, mountain pass solutions have Morse index 11. Thus part of the arguments in [16] can be simplified. For instance, the proofs of [16, Theorem 8.4, Lemma 8.6].

There is another candidate of solution, the maximum of I0𝐩I_{0}^{\mathbf{p}} on 𝒢0𝐩\mathcal{G}_{0}^{\mathbf{p}}, in the gap of 0\mathcal{M}_{0}.

Proposition 3.10.

I0𝐩I_{0}^{\mathbf{p}} attains the maximum, say u^𝐩\hat{u}_{\mathbf{p}} on 𝒢0𝐩\mathcal{G}_{0}^{\mathbf{p}}. If 𝐣𝕋0𝐩\mathbf{j}\in\mathbb{T}^{\mathbf{p}}_{0} such that 0<u^𝐩(𝐣)<(w0v0)(𝐣)0<\hat{u}_{\mathbf{p}}(\mathbf{j})<(w_{0}-v_{0})(\mathbf{j}), then

𝐤:𝐤𝐣r𝐤S𝐣(u^𝐩+v0)=0,\sum_{\mathbf{k}:\left\lVert\mathbf{k}-\mathbf{j}\right\rVert\leq r}\partial_{\mathbf{k}}S_{\mathbf{j}}(\hat{u}_{\mathbf{p}}+v_{0})=0, (3.9)

i.e., u^𝐩\hat{u}_{\mathbf{p}} satisfies (1.1) at 𝐣\mathbf{j}.

Proof.

By the definition of 𝒢0𝐩\mathcal{G}_{0}^{\mathbf{p}} and Proposition 3.2, there is u^𝐩𝒢0𝐩\hat{u}_{\mathbf{p}}\in\mathcal{G}_{0}^{\mathbf{p}} such that

I0𝐩(u^𝐩)=supu𝒢0𝐩I0𝐩(u).I_{0}^{\mathbf{p}}(\hat{u}_{\mathbf{p}})=\sup_{u\in\mathcal{G}_{0}^{\mathbf{p}}}I_{0}^{\mathbf{p}}(u).

If 𝐣𝕋0𝐩\mathbf{j}\in\mathbb{T}^{\mathbf{p}}_{0} such that 0<u^𝐩(𝐣)<(w0v0)(𝐣)0<\hat{u}_{\mathbf{p}}(\mathbf{j})<(w_{0}-v_{0})(\mathbf{j}), then choose t0>0t_{0}>0 such that 0<u^𝐩(𝐣)+t<(w0v0)(𝐣)0<\hat{u}_{\mathbf{p}}(\mathbf{j})+t<(w_{0}-v_{0})(\mathbf{j}) hold for all |t|<t0|t|<t_{0}. For 𝐤𝕋0𝐩\mathbf{k}\in\mathbb{T}^{\mathbf{p}}_{0}, define

v(𝐤)={t,𝐤=𝐣,0,𝐤𝐣,v(\mathbf{k})=\left\{\begin{array}[]{ll}t,&\mathbf{k}=\mathbf{j},\\ 0,&\mathbf{k}\neq\mathbf{j},\end{array}\right.

and extend vv to be a period function in Λ0𝐩\Lambda_{0}^{\mathbf{p}}. Then by (3.1),

0=(I0𝐩)(u^𝐩)v=t𝐤:𝐤𝐣r𝐤S𝐣(u^𝐩+v0),0=(I_{0}^{\mathbf{p}})^{\prime}(\hat{u}_{\mathbf{p}})v=t\sum_{\mathbf{k}:\left\lVert\mathbf{k}-\mathbf{j}\right\rVert\leq r}\partial_{\mathbf{k}}S_{\mathbf{j}}(\hat{u}_{\mathbf{p}}+v_{0}), (3.10)

where the first equality follows from u^𝐩\hat{u}_{\mathbf{p}} is a maximum point of I0𝐩I_{0}^{\mathbf{p}} on 𝒢0𝐩\mathcal{G}_{0}^{\mathbf{p}}. Since (3.10) holds for any |t|<t0|t|<t_{0}, we have (3.9). ∎

Remark 3.11.

If 𝐣𝕋0𝐩\mathbf{j}\in\mathbb{T}^{\mathbf{p}}_{0} such that 0=u^𝐩(𝐣)0=\hat{u}_{\mathbf{p}}(\mathbf{j}), then choose t0>0t_{0}>0 such that 0<u^𝐩(𝐣)+t<(w0v0)(𝐣)0<\hat{u}_{\mathbf{p}}(\mathbf{j})+t<(w_{0}-v_{0})(\mathbf{j}) hold for all 0<t<t00<t<t_{0}. The argument in Proposition 3.10 shows that 𝐤:𝐤𝐣r𝐤S𝐣(u^𝐩+v0)0\sum_{\mathbf{k}:\left\lVert\mathbf{k}-\mathbf{j}\right\rVert\leq r}\partial_{\mathbf{k}}S_{\mathbf{j}}(\hat{u}_{\mathbf{p}}+v_{0})\leq 0.

If u^𝐩\hat{u}_{\mathbf{p}} obtained in Proposition 3.10 is a solution of (1.1), then by Lemma 2.3, 0<u^𝐩<w0v00<\hat{u}_{\mathbf{p}}<w_{0}-v_{0}. Conversely, if 0<u^𝐩<w0v00<\hat{u}_{\mathbf{p}}<w_{0}-v_{0}, by Proposition 3.10, u^𝐩\hat{u}_{\mathbf{p}} is a solution of (1.1). Unfortunately, we do not know whether u^𝐩\hat{u}_{\mathbf{p}} obtained in Proposition 3.10 is a solution of (1.1) for general 𝐩\mathbf{p}.

But for 𝐩=(1,,1)=:𝟏\mathbf{p}=(1,\cdots,1)=:\mathbf{1}, we see 0<u^𝟏<w0v00<\hat{u}_{\mathbf{1}}<w_{0}-v_{0}, thus u^𝟏\hat{u}_{\mathbf{1}} is a solution of (1.1). In fact, as one can easily see, u^𝐩\hat{u}_{\mathbf{p}} is same to the mountain pass solution of Theorem 3.6.

For 𝐩(1,,1)\mathbf{p}\neq(1,\cdots,1), u^𝐩\hat{u}_{\mathbf{p}} may be coincide with u^𝟏\hat{u}_{\mathbf{1}} and it may not give more solutions. To see this, we examine the classical Frenkel-Kontorova model.

Example.

Let n=2n=2 and 𝐩(1)=(1,0)\mathbf{p}(1)=(1,0), 𝐩(2)=(1,1)\mathbf{p}(2)=(1,1). Set

S𝟎(u):=s(u|B𝟎1):=sin(2πu(𝟎))+116𝐣=1[u(𝐣)u(𝟎)]2.S_{\mathbf{0}}(u):=s(u|_{B_{\mathbf{0}}^{1}}):=\sin(2\pi u(\mathbf{0}))+\frac{1}{16}\sum_{\|\mathbf{j}\|=1}[u(\mathbf{j})-u(\mathbf{0})]^{2}.

Then infuΛ0𝐩(1)S𝟎(u)=1\inf_{u\in\Lambda_{0}^{\mathbf{p}(1)}}S_{\mathbf{0}}(u)=-1 is attained at k14k-\frac{1}{4} (kk\in\mathbb{Z}) and supuΛ0𝐩(1)S𝟎(u)=1\sup_{u\in\Lambda_{0}^{\mathbf{p}(1)}}S_{\mathbf{0}}(u)=1 is attained at k+14k+\frac{1}{4} (kk\in\mathbb{Z}). Assume the gap pair is v0=14v_{0}=-\frac{1}{4} and w0=34w_{0}=\frac{3}{4}. So I0𝐩(1)(u):=S𝟎(u14)I_{0}^{\mathbf{p}(1)}(u):=S_{\mathbf{0}}(u-\frac{1}{4}) and infuΛ0𝐩(1)I0𝐩(1)(u)=1\inf_{u\in\Lambda_{0}^{\mathbf{p}(1)}}I_{0}^{\mathbf{p}(1)}(u)=-1 is attained at kk (kk\in\mathbb{Z}) and supuΛ0𝐩(1)I0𝐩(1)(u)=1\sup_{u\in\Lambda_{0}^{\mathbf{p}(1)}}I_{0}^{\mathbf{p}(1)}(u)=1 is attained at k+12k+\frac{1}{2} (kk\in\mathbb{Z}).

On the other hand,

I0𝐩(2)(u):=S𝟎(u14)+S𝟎(τ11u14)=14[u((1,0))u((0,0))]2cos(2πu((0,0)))cos(2πu((1,0))).\begin{split}I_{0}^{\mathbf{p}(2)}(u):=&S_{\mathbf{0}}(u-\frac{1}{4})+S_{\mathbf{0}}(\tau^{1}_{-1}u-\frac{1}{4})\\ =&\frac{1}{4}[u((1,0))-u((0,0))]^{2}-\cos(2\pi u((0,0)))-\cos(2\pi u((1,0))).\end{split}

In the gap of v0,w0v_{0},w_{0}, there is one locally maximum point u12u\equiv\frac{1}{2}, and there are two mountain pass critical points u¯\bar{u} and τ11u¯\tau^{1}_{-1}\bar{u} but no other locally maximum point exists. Please see Figure 1 for the graph of I0𝐩(2)I_{0}^{\mathbf{p}(2)}.

Refer to caption
Figure 1. Graph of the function I0𝐩(2)I_{0}^{\mathbf{p}(2)}

3.2. Multiplicity of periodic mountain pass solutions

We shall prove that varying 𝐩\mathbf{p} will produce more periodic mountain pass solutions. This is different with periodic minimal and Birkhoff solutions (cf. Section 2.1, Lemma 2.6). Toward this end, for kk\in\mathbb{N}, assume 𝐩(k)=(k,1,,1)n\mathbf{p}(k)=(k,1,\cdots,1)\in\mathbb{N}^{n}. By Theorem 3.6, there exists uk𝒢0𝐩(k)u_{k}^{*}\in\mathcal{G}_{0}^{\mathbf{p}(k)} such that uk+v0u_{k}^{*}+v_{0} is a mountain pass solution. Since 𝒢0𝐩(1)𝒢0𝐩(k)\mathcal{G}_{0}^{\mathbf{p}(1)}\subset\mathcal{G}_{0}^{\mathbf{p}(k)}, u1𝒢0𝐩(k)u_{1}^{*}\in\mathcal{G}_{0}^{\mathbf{p}(k)}. It is possible that uk=u1u_{k}^{*}=u_{1}^{*}. But we shall show this cannot happen for infinite many kk. First we have

Proposition 3.12.

There is a constant M0M_{0}, independent of kk, such that

0<d0𝐩(k)c0𝐩(k)M0.0<d_{0}^{\mathbf{p}(k)}-c_{0}^{\mathbf{p}(k)}\leq M_{0}.
Proof.

To estimate d0𝐩(k)d_{0}^{\mathbf{p}(k)} we need to construct a suitable hk0𝐩(k)h_{k}\in\mathcal{H}_{0}^{\mathbf{p}(k)}. To do so, we first define the following χk\chi_{k} for k2k\geq 2.

If kk is even, set

χk(t,0)={t,0t1,1,1tk+52.\chi_{k}(t,0)=\left\{\begin{array}[]{ll}{t,}&{0\leq t\leq 1},\\ {1,}&{1\leq t\leq\frac{k+5}{2}}.\end{array}\right.

For 1ik211\leq i\leq\frac{k}{2}-1, set

χk(t,i)={0,0ti+12,2t12i,i+12ti+1,1,i+1tk+52.\chi_{k}(t,i)=\left\{\begin{array}[]{ll}{0,}&{0\leq t\leq i+\frac{1}{2}},\\ {2t-1-2i,}&{i+\frac{1}{2}\leq t\leq i+1},\\ {1,}&{i+1\leq t\leq\frac{k+5}{2}}.\end{array}\right.

Lastly set

χk(t,k2)={0,0tk+12,112(k+52t),k+12tk+52,\chi_{k}(t,\frac{k}{2})=\left\{\begin{array}[]{ll}{0,}&{0\leq t\leq\frac{k+1}{2}},\\ {1-\frac{1}{2}(\frac{k+5}{2}-t),}&{\frac{k+1}{2}\leq t\leq\frac{k+5}{2}},\end{array}\right.

and χk(t,i)=χk(t,i)\chi_{k}(t,i)=\chi_{k}(t,-i) for k2<i<0-\frac{k}{2}<i<0. Extend χk(t,)\chi_{k}(t,\cdot) as a kk-periodic function on \mathbb{Z}.

If kk is odd, then k1k-1 is even and thus χk1(t,i)\chi_{k-1}(t,i) is well-defined for k12<ik12-\frac{k-1}{2}<i\leq\frac{k-1}{2} and 0tk+520\leq t\leq\frac{k+5}{2} by the previous paragraph. Now define χk(t,i):=χk1(t,i)\chi_{k}(t,i):=\chi_{k-1}(t,i) for 0ik120\leq i\leq\frac{k-1}{2} and 0tk+520\leq t\leq\frac{k+5}{2}. Let χk(t,i):=χk(t,i)\chi_{k}(t,i):=\chi_{k}(t,-i) for k12i<0-\frac{k-1}{2}\leq i<0 and then extend χk(t,)\chi_{k}(t,\cdot) as a kk-periodic function on \mathbb{Z}.

For the above χk\chi_{k}, set

ϕk(θ,i)=χk(θ(k+5)2,i).\phi_{k}(\theta,i)=\chi_{k}\left(\frac{\theta(k+5)}{2},i\right). (3.11)

Letting hk(θ)(𝐢)=ϕk(θ,𝐢1)[w0(𝐢)v0(𝐢)]h_{k}(\theta)(\mathbf{i})=\phi_{k}(\theta,\mathbf{i}_{1})[w_{0}(\mathbf{i})-v_{0}(\mathbf{i})] gives hkH0𝐩(k)h_{k}\in H_{0}^{\mathbf{p}(k)}. Notice that for θ[0,1]\theta\in[0,1], hk(θ)0h_{k}(\theta)\neq 0, w0v0w_{0}-v_{0} on {𝐢n| 0𝐢1k,𝐢2=𝐢3==𝐢n=0}\{\mathbf{i}\in\mathbb{Z}^{n}\,|\,0\leq\mathbf{i}_{1}\leq k,\,\mathbf{i}_{2}=\mathbf{i}_{3}=\cdots=\mathbf{i}_{n}=0\} for at most two points, and for any 𝐢\mathbf{i}, hk(θ)(𝐢)h_{k}(\theta)(\mathbf{i}) is monotone nondecreasing in θ\theta. Thus we have

maxθ[0,1]I0𝐩(k)(hk(θ))c0𝐩(k)M0\max_{\theta\in[0,1]}I^{\mathbf{p}(k)}_{0}\left(h_{k}(\theta)\right)-c_{0}^{\mathbf{p}(k)}\leq M_{0}

for some M0M_{0} independent of kk. Hence for k2k\geq 2, d0𝐩(k)c0𝐩(k)M0d_{0}^{\mathbf{p}(k)}-c_{0}^{\mathbf{p}(k)}\leq M_{0}. Enlarging M0M_{0} if necessary shows that d0𝐩(k)c0𝐩(k)M0d_{0}^{\mathbf{p}(k)}-c_{0}^{\mathbf{p}(k)}\leq M_{0} hold for all kk\in\mathbb{N}, thus Proposition 3.12 is proved. ∎

Now we prove the multiplicity of periodic mountain pass solutions.

Theorem 3.13.

The set {uk}k=1\{u_{k}^{*}\}_{k=1}^{\infty} is infinite.

Proof.

Suppose, by contradiction, {uk}k=1={uki}i=1m\{u_{k}^{*}\}_{k=1}^{\infty}=\{u_{k_{i}}^{*}\}_{i=1}^{m} and 1=k1<<km1=k_{1}<\cdots<k_{m}. For ll\in\mathbb{N}, we obtain ulkm=ukju_{lk_{m}}^{*}=u_{k_{j}}^{*} for some j{1,,m}j\in\{1,\dots,m\}. Hence

d0𝐩(lkm)=I0𝐩(lkm)(ulkm)=I0𝐩(lkm)(ukj)=γkmd0𝐩(kj)γkmmin1imd0𝐩(ki)lmin1imd0𝐩(ki).\begin{split}d_{0}^{\mathbf{p}(lk_{m})}=&I^{\mathbf{p}(lk_{m})}_{0}(u_{lk_{m}}^{*})=I^{\mathbf{p}(lk_{m})}_{0}(u_{k_{j}}^{*})\\ =&\gamma_{k_{m}}d_{0}^{\mathbf{p}(k_{j})}\geq\gamma_{k_{m}}\min_{1\leq i\leq m}d_{0}^{\mathbf{p}(k_{i})}\\ \geq&l\min_{1\leq i\leq m}d_{0}^{\mathbf{p}(k_{i})}.\end{split}

Proposition 3.5 implies d0𝐩(ki)>c0𝐩(ki)d_{0}^{\mathbf{p}(k_{i})}>c_{0}^{\mathbf{p}(k_{i})}, 1im1\leq i\leq m and then

min1imd0𝐩(ki)min1imc0𝐩(ki)>0.\min_{1\leq i\leq m}d_{0}^{\mathbf{p}(k_{i})}-\min_{1\leq i\leq m}c_{0}^{\mathbf{p}(k_{i})}>0.

Thus

d0𝐩(lkm)c0𝐩(lkm)γkmmin1imd0𝐩(ki)lc0𝐩(km)γkmmin1imd0𝐩(ki)lmin1imc0𝐩(ki)l(min1imd0𝐩(ki)min1imc0𝐩(ki))\begin{split}d_{0}^{\mathbf{p}(lk_{m})}-c_{0}^{\mathbf{p}(lk_{m})}\geq&\gamma_{k_{m}}\min_{1\leq i\leq m}d_{0}^{\mathbf{p}(k_{i})}-lc_{0}^{\mathbf{p}(k_{m})}\\ \geq&\gamma_{k_{m}}\min_{1\leq i\leq m}d_{0}^{\mathbf{p}(k_{i})}-l\min_{1\leq i\leq m}c_{0}^{\mathbf{p}(k_{i})}\\ \geq&l\left(\min_{1\leq i\leq m}d_{0}^{\mathbf{p}(k_{i})}-\min_{1\leq i\leq m}c_{0}^{\mathbf{p}(k_{i})}\right)\\ \to&\infty\end{split}

as ll\to\infty, contrary to Proposition 3.12. ∎

If 𝐩(k)=(k,1,,1)\mathbf{p}(k)=(k,1,\cdots,1) is replaced by (1,,k,,1)(1,\cdots,k,\cdots,1), similar arguments of this section give more periodic mountain pass solutions. What happens if we change the coordinate systems (𝐞1,,𝐞n)(\mathbf{e}_{1},\cdots,\mathbf{e}_{n})? we learn from [11, Lemma 5.4] that if ωi=j=1nαij𝐞j\omega_{i}=\sum_{j=1}^{n}\alpha_{ij}\mathbf{e}_{j} with αij\alpha_{ij}\in\mathbb{Z} and the vectors ωi\omega_{i} are linearly independent, then we do not obtain more periodic minimal and Birkhoff solutions. Different with minimal and Birkhoff solutions, changing coordinate systems produces more periodic mountain pass solutions. To see this, for simplicity set ω1=𝐞1+𝐞2\omega_{1}=\mathbf{e}_{1}+\mathbf{e}_{2}, ω2=𝐞1+𝐞2\omega_{2}=-\mathbf{e}_{1}+\mathbf{e}_{2}, and ωj=𝐞j\omega_{j}=\mathbf{e}_{j} for j=3,,nj=3,\cdots,n. Set 𝐩(k,i)=(1,,1)+(k1)𝐞i\mathbf{p}(k,i)=(1,\cdots,1)+(k-1)\mathbf{e}_{i}. Denote by 𝒮\mathcal{S} the set of critical points of I0𝐩(k,i)I_{0}^{\mathbf{p}(k,i)} on Λ0𝐩(k,i)\Lambda_{0}^{\mathbf{p}(k,i)} in the coordinate sysetems (𝐞1,,𝐞n)(\mathbf{e}_{1},\cdots,\mathbf{e}_{n}). Let uku_{k}^{*} be the critical point that are kk-periodic in ω1\omega_{1} and 11-periodic in ω2,,ωn\omega_{2},\cdots,\omega_{n}. Assume ukn/nu_{k}^{*}\not\in\mathbb{R}^{\mathbb{Z}^{n}/\mathbb{Z}^{n}}. Notice that by Proposition 3.12 and Theorem 3.13 there are infinitely many functions of this type. We have:

Proposition 3.14.

uk𝒮u^{*}_{k}\not\in\mathcal{S}.

Proof.

Suppose, by contradiction, uk=uu^{*}_{k}=u for some u𝒮u\in\mathcal{S}. Since u𝒮u\in\mathcal{S}, there is some jj\in\mathbb{N} such that

{u(𝐢+j𝐞p)=u(𝐢),u(𝐢+𝐞i)=u(𝐢),for all ip.\left\{\begin{array}[]{ll}u\left(\mathbf{i}+j\mathbf{e}_{p}\right)=u(\mathbf{i}),&\\ u(\mathbf{i}+\mathbf{e}_{i})=u(\mathbf{i}),&\hbox{for all $i\neq p$.}\end{array}\right. (3.12)
  • If p1,2p\neq 1,2, then

    u(𝐢+𝐞p)=uk(𝐢+𝐞p)=uk(𝐢+ωp)=uk(𝐢)=u(𝐢),u(\mathbf{i}+\mathbf{e}_{p})=u_{k}^{*}(\mathbf{i}+\mathbf{e}_{p})=u_{k}^{*}(\mathbf{i}+\omega_{p})=u_{k}^{*}(\mathbf{i})=u(\mathbf{i}), (3.13)

    Combining (3.12) and (3.13) gives un/nu\in\mathbb{R}^{\mathbb{Z}^{n}/\mathbb{Z}^{n}}. But this contradicts u=ukn/nu=u_{k}^{*}\not\in\mathbb{R}^{\mathbb{Z}^{n}/\mathbb{Z}^{n}}.

  • If p=2p=2, then

    u(𝐢+𝐞2)=u(𝐢+𝐞2𝐞1)=uk(𝐢+𝐞2𝐞1)=uk(𝐢+ξ2)=uk(𝐢)=u(𝐢).\begin{split}u(\mathbf{i}+\mathbf{e}_{2})=&u(\mathbf{i}+\mathbf{e}_{2}-\mathbf{e}_{1})=u_{k}^{*}(\mathbf{i}+\mathbf{e}_{2}-\mathbf{e}_{1})\\ =&u_{k}^{*}(\mathbf{i}+\mathbf{\xi}_{2})=u_{k}^{*}(\mathbf{i})=u(\mathbf{i}).\end{split}

    Again un/nu\in\mathbb{R}^{\mathbb{Z}^{n}/\mathbb{Z}^{n}}, a contradiction.

  • If p=1p=1, then

    u(𝐢𝐞1)=u(𝐢𝐞1+𝐞2)=u(𝐢+ξ2)=uk(𝐢+ξ2)=uk(𝐢)=u(𝐢),\begin{split}u(\mathbf{i}-\mathbf{e}_{1})=&u(\mathbf{i}-\mathbf{e}_{1}+\mathbf{e}_{2})=u(\mathbf{i}+\mathbf{\xi}_{2})\\ =&u^{*}_{k}(\mathbf{i}+\mathbf{\xi}_{2})=u^{*}_{k}(\mathbf{i})=u(\mathbf{i}),\end{split}

    again a contradiction.

Certain classes of sets which consist of solutions of (1.1) attracts researchers’ attention. When the elements in the set have good order property, the set becomes a foliation or lamination. With the periodic mountain pass solutions in hand, one may wonder: is there a possibility that the periodic mountain pass solutions constitute a foliation or lamination? Unfortunately, the answer is “negative”. (In fact, we construct some periodic solutions that are cross. These periodic solutions are suspected to be mountain pass type. Please see Remark 3.21 below.) To this end, we need the following definition.

Definition 3.15.

We say uu touches vv from below (resp. above) if uvu\leq v (resp. vuv\leq u), uvu\neq v and there exists 𝐢\mathbf{i} such that u(𝐢)=v(𝐢)u(\mathbf{i})=v(\mathbf{i}). We say uu intersects vv if there are 𝐢\mathbf{i}, 𝐣n\mathbf{j}\in\mathbb{Z}^{n} such that [u(𝐢)v(𝐢)][u(𝐣)v(𝐣)]<0[u(\mathbf{i})-v(\mathbf{i})][u(\mathbf{j})-v(\mathbf{j})]<0.

Remark 3.16.

Note that if uvu\leq v are solutions of (1.1), by Lemma 2.3 uu will not touch vv from below.

Set

^0𝐩={h0𝐩|h is monotone nondecreasing in θ,i.e.,h(θ1)h(θ2) for any θ1>θ2}.\begin{split}\hat{\mathcal{H}}_{0}^{\mathbf{p}}=\{h\in\mathcal{H}^{\mathbf{p}}_{0}\,|\,&h\text{ is monotone nondecreasing in }\theta,i.e.,\\ &h(\theta_{1})\geq h(\theta_{2})\textrm{ for any }\theta_{1}>\theta_{2}\}.\end{split}

and

d^0𝐩=infh^0𝐩maxθ[0,1]I0𝐩(h(θ)).\hat{d}_{0}^{\mathbf{p}}=\inf_{h\in\hat{\mathcal{H}}_{0}^{\mathbf{p}}}\max_{\theta\in[0,1]}I_{0}^{\mathbf{p}}(h(\theta)).

Then d^0𝐩d0𝐩\hat{d}_{0}^{\mathbf{p}}\geq d_{0}^{\mathbf{p}}. Similar to Theorem 3.6, d^0𝐩\hat{d}_{0}^{\mathbf{p}} is a critical value of I0𝐩I_{0}^{\mathbf{p}} with a corresponding critical point in 𝒢0𝐩\mathcal{G}^{\mathbf{p}}_{0}. By Proposition 3.1 and Lemma 3.4, Φt0(^0𝐩)^0𝐩\Phi_{t}^{0}(\hat{\mathcal{H}}_{0}^{\mathbf{p}})\subset\hat{\mathcal{H}}_{0}^{\mathbf{p}}, where Φt0:Λ0𝐩(k)Λ0𝐩(k)\Phi_{t}^{0}:\Lambda^{\mathbf{p}(k)}_{0}\to\Lambda^{\mathbf{p}(k)}_{0} is defined as in (3.2). For 𝐩(k)=(k,1,,1)n\mathbf{p}(k)=(k,1,\cdots,1)\in\mathbb{N}^{n}, set h^0𝐩(k)h\in\hat{\mathcal{H}}_{0}^{\mathbf{p}(k)}. For t>0t>0, define

θ¯t:=sup{θ[0,1]|Φt0h(θ)u0, but Φt0h(θ)u0}θ¯t:=inf{θ[0,1]|Φt0h(θ)u0, but Φt0h(θ)u0},\begin{split}&\underline{\theta}_{t}:=\sup\{\theta\in[0,1]\,|\,\Phi_{t}^{0}h(\theta)\leq u_{0},\textrm{ but }\Phi_{t}^{0}h(\theta)\neq u_{0}\}\\ \leq&\overline{\theta}_{t}:=\inf\{\theta\in[0,1]\,|\,\Phi_{t}^{0}h(\theta)\geq u_{0},\textrm{ but }\Phi_{t}^{0}h(\theta)\neq u_{0}\},\end{split}

where u0Λ0𝐩(1)u_{0}\in\Lambda_{0}^{\mathbf{p}(1)} satisfying 0<u0<w0v00<u_{0}<w_{0}-v_{0} and u0+v0u_{0}+v_{0} is a solution of (1.1). By the periodicity of u0u_{0} and Φt0h(θ)\Phi_{t}^{0}h(\theta), Φt0h(θ)u0\Phi_{t}^{0}h(\theta)\leq u_{0} means

  1. (a)

    Φt0h(θ)<u0\Phi_{t}^{0}h(\theta)<u_{0}, or

  2. (b)

    Φt0h(θ)\Phi_{t}^{0}h(\theta) touches u0u_{0} from below, or

  3. (c)

    Φt0h(θ)=u0\Phi_{t}^{0}h(\theta)=u_{0}.

For θ=θ¯t\theta=\underline{\theta}_{t}, (a) will not hold. We need the following lemma.

Lemma 3.17.

For θ<θ¯t\theta<\underline{\theta}_{t}, either Φt0h(θ)<u0\Phi_{t}^{0}h(\theta)<u_{0} or Φt0h(θ)\Phi_{t}^{0}h(\theta) touches u0u_{0} from below.

Proof.

For θ<θ¯t\theta<\underline{\theta}_{t}, h(θ)h(θ¯t)h(\theta)\leq h(\underline{\theta}_{t}) and then Φt0h(θ)Φt0h(θ¯t)u0\Phi_{t}^{0}h(\theta)\leq\Phi_{t}^{0}h(\underline{\theta}_{t})\leq u_{0}. We have the following three cases.

  1. (1)

    Assume h(θ)<h(θ¯t)h(\theta)<h(\underline{\theta}_{t}). Then Φt0h(θ)<Φt0h(θ¯t)u0\Phi_{t}^{0}h(\theta)<\Phi_{t}^{0}h(\underline{\theta}_{t})\leq u_{0} and we are throuth.

  2. (2)

    Assume h(θ)h(\theta) touches h(θ¯t)h(\underline{\theta}_{t}) from below. Then Φt0h(θ)<Φt0h(θ¯t)u0\Phi_{t}^{0}h(\theta)<\Phi_{t}^{0}h(\underline{\theta}_{t})\leq u_{0} and we are throuth.

  3. (3)

    Assume h(θ)=h(θ¯t)h(\theta)=h(\underline{\theta}_{t}). Then Φt0h(θ)=Φt0h(θ¯t)u0\Phi_{t}^{0}h(\theta)=\Phi_{t}^{0}h(\underline{\theta}_{t})\leq u_{0}.

    1. (a)

      If Φt0h(θ¯t)=u0\Phi_{t}^{0}h(\underline{\theta}_{t})=u_{0}, then so is Φt0h(θ)\Phi_{t}^{0}h(\theta), contradicting the definition of θ¯t\underline{\theta}_{t}.

    2. (b)

      If Φt0h(θ¯t)\Phi_{t}^{0}h(\underline{\theta}_{t}) touches u0u_{0} from below, then so is Φt0h(θ)\Phi_{t}^{0}h(\theta).

Thus either Φt0h(θ)<u0\Phi_{t}^{0}h(\theta)<u_{0} or Φt0h(θ)\Phi_{t}^{0}h(\theta) touches u0u_{0} from below. ∎

Since hh is continous, θ¯t\underline{\theta}_{t} is attained. Similar results hold for θ¯t\overline{\theta}_{t}. Moreover, θ¯t,θ¯t\underline{\theta}_{t},\overline{\theta}_{t} have the following monotone property.

Proposition 3.18.

Assume h^0𝐩(k)h\in\hat{\mathcal{H}}_{0}^{\mathbf{p}(k)}.

  1. (1)

    If hu0h\neq u_{0} for all θ[0,1]\theta\in[0,1], then the map tθ¯tt\mapsto\underline{\theta}_{t} (resp. tθ¯tt\mapsto\overline{\theta}_{t}) is monotone nondecreasing (monotone non-increasing).

  2. (2)

    Assume there is some θ0(0,1)\theta_{0}\in(0,1) such that h(θ)=u0h(\theta)=u_{0}. If such θ0\theta_{0} is unique, then θ¯t=θ0=θ¯t\underline{\theta}_{t}=\theta_{0}=\overline{\theta}_{t} for all t>0t>0. If h(θ)=u0h(\theta)=u_{0} for θ[a,b](0,1)\theta\in[a,b]\subset(0,1) and [a,b][a,b] is the maximal interval owning this property, then θ¯t=a\underline{\theta}_{t}=a and θ¯t=b\overline{\theta}_{t}=b.

Proof.

By the definition of θ¯t,θ¯t\underline{\theta}_{t},\overline{\theta}_{t}, Proposition 3.18 (2) is obvious. Thus we only prove Proposition 3.18 (1). Assume t1<t2t_{1}<t_{2}. For μ(0,θ¯t1)\mu\in(0,\underline{\theta}_{t_{1}}), by Lemma 3.4 (i), Lemma 3.17 and Proposition 3.1, {u𝒢0𝐩(k)| 0uu0}\{u\in\mathcal{G}^{\mathbf{p}(k)}_{0}\,|\,0\leq u\leq u_{0}\} is invariant under Φt0\Phi_{t}^{0} and

Φt+t10h(μ)<u0,for t>0.\Phi_{t+t_{1}}^{0}h(\mu)<u_{0},\quad\textrm{for }t>0.

If we take t=t2t1t=t_{2}-t_{1}, then μ<θ¯t2\mu<\underline{\theta}_{t_{2}}. Hence θ¯t1θ¯t2\underline{\theta}_{t_{1}}\leq\underline{\theta}_{t_{2}}. θ¯t1θ¯t2\overline{\theta}_{t_{1}}\geq\overline{\theta}_{t_{2}} can be proved similarly. ∎

Remark 3.19.

We suspect that in the case (1) of Proposition 3.18, either θ¯t\underline{\theta}_{t} is strictly increasing or θ¯tθ0\underline{\theta}_{t}\equiv\theta_{0} for some θ0(0,1)\theta_{0}\in(0,1). To support this, suppose t1<t2t_{1}<t_{2}.

  • If Φt10h(θ¯t1)\Phi_{t_{1}}^{0}h(\underline{\theta}_{t_{1}}) touches u0u_{0} from below, then θ¯t1<θ¯t2\underline{\theta}_{t_{1}}<\underline{\theta}_{t_{2}}. Indeed, by Proposition 3.1, Φt+t10h(θ¯t1)<u0\Phi_{t+t_{1}}^{0}h(\underline{\theta}_{t_{1}})<u_{0} for t>0t>0. In particular, Φt20(h(θ¯t1))<u0\Phi_{t_{2}}^{0}(h(\underline{\theta}_{t_{1}}))<u_{0}. If θ¯t1=θ¯t2\underline{\theta}_{t_{1}}=\underline{\theta}_{t_{2}} then u0>Φt20(h(θ¯t2))u_{0}>\Phi_{t_{2}}^{0}(h(\underline{\theta}_{t_{2}})), a contradiction.

  • If Φt10h(θ¯t1)=u0\Phi_{t_{1}}^{0}h(\underline{\theta}_{t_{1}})=u_{0}, then θ¯t1=θ¯t2\underline{\theta}_{t_{1}}=\underline{\theta}_{t_{2}}. Indeed, suppose, by contradiction θ¯t1<θ¯t2\underline{\theta}_{t_{1}}<\underline{\theta}_{t_{2}}. Then by Lemma 3.17, either Φt20h(θ¯t1)<u0\Phi_{t_{2}}^{0}h(\underline{\theta}_{t_{1}})<u_{0} or Φt20h(θ¯t1)\Phi_{t_{2}}^{0}h(\underline{\theta}_{t_{1}}) touches u0u_{0} from below. Both cases contradict Φt10h(θ¯t1)=u0\Phi_{t_{1}}^{0}h(\underline{\theta}_{t_{1}})=u_{0}.

But we do not know if there are t1<t2t_{1}<t_{2} such that Φt10h(θ¯t1)\Phi_{t_{1}}^{0}h(\underline{\theta}_{t_{1}}) touches u0u_{0} from below and Φt20h(θ¯t2)=u0\Phi_{t_{2}}^{0}h(\underline{\theta}_{t_{2}})=u_{0}.

Theorem 3.20.

Let u0Λ0𝐩(1)u_{0}\in\Lambda_{0}^{\mathbf{p}(1)} such that u0+v0u_{0}+v_{0} is a solution of (1.1) and 0<u0<w0v00<u_{0}<w_{0}-v_{0}. Then for kk\in\mathbb{N} sufficiently large, there exists a ukΛ0𝐩(k)u_{k}\in\Lambda^{\mathbf{p}(k)}_{0} with uk+v0u_{k}+v_{0} is a solution of (1.1) and uku_{k} intersects u0u_{0}.

Proof.

By Proposition 3.12 and its proof, for kk\in\mathbb{N} there are h𝐩(k)^0𝐩(k)h_{\mathbf{p}(k)}\in\hat{\mathcal{H}}_{0}^{\mathbf{p}(k)} and M0>0M_{0}>0 such that

max0θ1I0𝐩(k)(h𝐩(k)(θ))kc0M0.\max_{0\leq\theta\leq 1}I^{\mathbf{p}(k)}_{0}\left(h_{\mathbf{p}(k)}(\theta)\right)-kc_{0}\leq M_{0}.

By Proposition 3.5, d0𝐩(1)>c0𝐩(1)=c0d_{0}^{\mathbf{p}(1)}>c_{0}^{\mathbf{p}(1)}=c_{0}. Then for large kk,

M0<k(d0𝐩(1)c0𝐩(1))=I0𝐩(k)(u0)kc0.M_{0}<k(d_{0}^{\mathbf{p}(1)}-c_{0}^{\mathbf{p}(1)})=I^{\mathbf{p}(k)}_{0}\left(u_{0}\right)-kc_{0}. (3.14)

Since θ¯t,θ¯t\underline{\theta}_{t},\overline{\theta}_{t} are monotone, we have

limtθ¯t=:θ¯θ¯:=limtθ¯t.\lim_{t\rightarrow\infty}\underline{\theta}_{t}=:\underline{\theta}\leq\overline{\theta}:=\lim_{t\rightarrow\infty}\overline{\theta}_{t}.

Lemma 3.4 (iv) shows that there are a uΛ0𝐩(k)u^{-}\in\Lambda_{0}^{\mathbf{p}(k)} and a sequence tmt_{m}\to\infty as mm\to\infty such that

u=limmΦtm0(h𝐩(k)(θ¯))u^{-}=\lim_{m\rightarrow\infty}\Phi_{t_{m}}^{0}(h_{\mathbf{p}(k)}(\underline{\theta}))

and u+v0u^{-}+v_{0} is a solution of (1.1). Note that by (3.14) and Lemma 3.4 (ii),

I0𝐩(k)(u)kc0=limmI0𝐩(k)(Φtm0(h𝐩(k)(θ¯)))kc0I0𝐩(k)(h𝐩(k)(θ¯))kc0M0<I0𝐩(k)(u0)kc0,\begin{split}I^{\mathbf{p}(k)}_{0}(u^{-})-kc_{0}=&\lim_{m\to\infty}I^{\mathbf{p}(k)}_{0}(\Phi_{t_{m}}^{0}(h_{\mathbf{p}(k)}(\underline{\theta})))-kc_{0}\\ \leq&I^{\mathbf{p}(k)}_{0}(h_{\mathbf{p}(k)}(\underline{\theta}))-kc_{0}\\ \leq&M_{0}\\ <&I^{\mathbf{p}(k)}_{0}(u_{0})-kc_{0},\end{split} (3.15)

thus I0𝐩(k)(u)<I0𝐩(k)(u0)I^{\mathbf{p}(k)}_{0}(u^{-})<I^{\mathbf{p}(k)}_{0}(u_{0}) and then uu0u^{-}\neq u_{0}. We claim that uu^{-} intersects u0u_{0}. Noticing that u+v0,u0+v0u^{-}+v_{0},u_{0}+v_{0} are solutions of (1.1), if uu^{-} does not intersect u0u_{0}, then by Lemma 2.3 either (i) u>u0u^{-}>u_{0} or (ii) u<u0u^{-}<u_{0}. If (i) holds, by the definition of θ¯tm\overline{\theta}_{t_{m}}, there is some 𝐢n\mathbf{i}\in\mathbb{Z}^{n}, such that Φtm0(h𝐩(k)(θ¯))(𝐢)Φtm0(h𝐩(k)(θ¯tm))(𝐢)=u0(𝐢)\Phi_{t_{m}}^{0}(h_{\mathbf{p}(k)}(\overline{\theta}))(\mathbf{i})\leq\Phi_{t_{m}}^{0}(h_{\mathbf{p}(k)}(\overline{\theta}_{t_{m}}))(\mathbf{i})=u_{0}(\mathbf{i}). But by the periodicity of Φtm0(h𝐩(k)(θ¯))\Phi_{t_{m}}^{0}(h_{\mathbf{p}(k)}(\underline{\theta})) and u0u_{0}, for mm large enough,

Φtm0(h𝐩(k)(θ¯))Φtm0(h𝐩(k)(θ¯))>u0,\Phi_{t_{m}}^{0}(h_{\mathbf{p}(k)}(\overline{\theta}))\geq\Phi_{t_{m}}^{0}(h_{\mathbf{p}(k)}(\underline{\theta}))>u_{0},

a contradiction. If (ii) is satisfied, by the definition of θ¯tm\underline{\theta}_{t_{m}}, there exists 𝐢n\mathbf{i}\in\mathbb{Z}^{n} such that Φtm0(h𝐩(k)(θ¯tm))(𝐢)=u0(𝐢)\Phi_{t_{m}}^{0}(h_{\mathbf{p}(k)}(\underline{\theta}_{t_{m}}))(\mathbf{i})=u_{0}(\mathbf{i}). But

Φtm0(h𝐩(k)(θ¯tm))Φtm0(h𝐩(k)(θ¯))<u0\Phi_{t_{m}}^{0}\left(h_{\mathbf{p}(k)}\left(\underline{\theta}_{t_{m}}\right)\right)\leq\Phi_{t_{m}}^{0}\left(h_{\mathbf{p}(k)}\left(\underline{\theta}\right)\right)<u_{0}

for large mm, again a contradiction. Thus Theorem 3.20 is proved by setting uk=uu_{k}=u^{-}. ∎

Remark 3.21.

In the proof of Theorem 3.20, one can obtain another solution, say u++v0u^{+}+v_{0} by taking limit for a subsequence of Φt0(h𝐩(k)(θ¯))+v0\Phi_{t}^{0}(h_{\mathbf{p}(k)}(\overline{\theta}))+v_{0}. Of course, it may happen that u+=uu^{+}=u^{-}. We cannot figure out whether uu^{-} (or u+u^{+}) is a mountain pass critical point from the proof of Theorem 3.20, but please see [4, Remark 3.3] for more discussions.

Since {uk}k=1\{u_{k}\}_{k=1}^{\infty} obtained by Theorem 3.20 are lying in the gap of 0,w0v00,w_{0}-v_{0}, by Lemma 2.2 we can extract a subsequence converging to a function uΓ^(v0,w0)u\in\hat{\Gamma}(v_{0},w_{0}) such that u+v0u+v_{0} is a solution of (1.1).

Corollary 3.22.

There is a subsequence of τjk1uk\tau_{-j_{k}}^{1}u_{k} converging to a function UΓ^(v0,w0)U\in\hat{\Gamma}(v_{0},w_{0}) with U(𝟎)u0(𝟎)U(\mathbf{0})\geq u_{0}(\mathbf{0}) and U+v0U+v_{0} is a solution of (1.1), where u0u_{0} is as in Theorem 3.20.

Proof.

By Theorem 3.20, without loss of generality uku_{k} intersects u0u_{0} for all kk\in\mathbb{N}. Thus we have τjk1uk(𝟎)>u0(𝟎)\tau_{-j_{k}}^{1}u_{k}\left(\mathbf{0}\right)>u_{0}\left(\mathbf{0}\right) for some jkj_{k}\in\mathbb{Z}. Let Uk:=τjk1ukU_{k}:=\tau^{1}_{-j_{k}}u_{k}, then Uk+v0U_{k}+v_{0} are solutions of (1.1). Since 0Ukw0v00\leq U_{k}\leq w_{0}-v_{0}, by Lemma 2.2, there is a function U×n1/n1U\in\mathbb{R}^{\mathbb{Z}\times\mathbb{Z}^{n-1}/\mathbb{Z}^{n-1}} such that UkUU_{k}\to U pointwise (up to a subsequence). So U(𝟎)u0(𝟎)U(\mathbf{0})\geq u_{0}(\mathbf{0}) and U+v0U+v_{0} is a solution of (1.1). ∎

Since U+v0Γ^(v0,w0)U+v_{0}\in\hat{\Gamma}(v_{0},w_{0}), either J1(U+v0)=J_{1}(U+v_{0})=\infty or J1(U+v0)<J_{1}(U+v_{0})<\infty.

  1. (1)

    Suppose J1(U+v0)=J_{1}(U+v_{0})=\infty. Then the solution U+v0U+v_{0} is different from any known solutions. For instance, U+v0U+v_{0} is not a minimal and Birkhoff solution in [14, 11], and is not a multitransition solution in [12] since the functional J1J_{1} at any of these solutions is finite.

  2. (2)

    Now assume J1(U+v0)<J_{1}(U+v_{0})<\infty. By Lemma 2.7, there are ϕ,ψ{v0,w0}\phi,\psi\in\{v_{0},w_{0}\} satisfying

    |(U+v0ϕ)(𝐓i)|0,\displaystyle|(U+v_{0}-\phi)(\mathbf{T}_{i})|\rightarrow 0, i;\displaystyle\quad i\rightarrow-\infty; (3.16)
    |(U+v0ψ)(𝐓i)|0,\displaystyle|(U+v_{0}-\psi)(\mathbf{T}_{i})|\rightarrow 0, i.\displaystyle\quad i\rightarrow\infty.
    1. (a)

      If ϕ=ψ\phi=\psi, then U+v0U+v_{0} is a homoclinic solution.

    2. (b)

      Otherwise suppose ϕψ\phi\neq\psi and assume there is no subsequence of τjk1uk+v0\tau_{-j_{k}}^{1}u_{k}+v_{0} converging to homoclinic solution. Without loss of generality, set ϕ=v0\phi=v_{0} and ψ=w0\psi=w_{0}. Then U+v0U+v_{0} is a heteroclinic solution. In this case, it is interesting that we can construct another solution of (1.1). Indeed, for ϵ>0\epsilon>0, by (3.16) there exists i0(ϵ)i_{0}(\epsilon)\in\mathbb{N} satisfying

      |(U+v0v0)(𝐓i)|ϵ,\displaystyle\left|(U+v_{0}-v_{0})(\mathbf{T}_{i})\right|\leq\epsilon, ii0(ϵ),\displaystyle\quad i\leq-i_{0}(\epsilon), (3.17)
      |(U+v0w0)(𝐓i)|ϵ,\displaystyle\left|(U+v_{0}-w_{0})(\mathbf{T}_{i})\right|\leq\epsilon, ii0(ϵ).\displaystyle\quad i\geq i_{0}(\epsilon).

      Since UkUU_{k}\to U pointwise as kk\to\infty, there is a k0=k0(ϵ)k_{0}=k_{0}(\epsilon) such that for kk0k\geq k_{0},

      |(Uk+v0v0)(𝐓i0(ϵ))|\displaystyle\left|(U_{k}+v_{0}-v_{0})(\mathbf{T}_{-i_{0}(\epsilon)})\right| 2ϵ,\displaystyle\leq 2\epsilon,
      |(Uk+v0w0)(𝐓i0(ϵ))|\displaystyle\left|(U_{k}+v_{0}-w_{0})(\mathbf{T}_{i_{0}(\epsilon)})\right| 2ϵ.\displaystyle\leq 2\epsilon.

      Thus we have Uk(𝐓i0(ϵ))<u0(𝐓i0(ϵ))U_{k}(\mathbf{T}_{-i_{0}(\epsilon)})<u_{0}(\mathbf{T}_{-i_{0}(\epsilon)}) and u0(𝐓i0(ϵ))<Uk(𝐓i0(ϵ))u_{0}(\mathbf{T}_{i_{0}(\epsilon)})<U_{k}(\mathbf{T}_{i_{0}(\epsilon)}) provided ϵ\epsilon sufficiently small. But noticing that UkU_{k} is kk-periodic in 𝐢1\mathbf{i}_{1}, we obtain

      |(Uk+v0v0)(𝐓ki0(ϵ))|2ϵ\left|(U_{k}+v_{0}-v_{0})(\mathbf{T}_{k-i_{0}(\epsilon)})\right|\leq 2\epsilon

      and Uk(𝐓ki0(ϵ))<u0(𝐓ki0(ϵ))U_{k}(\mathbf{T}_{k-i_{0}(\epsilon)})<u_{0}(\mathbf{T}_{k-i_{0}(\epsilon)}) with ki0(ϵ)>i0(ϵ)k-i_{0}(\epsilon)>i_{0}(\epsilon). Hence there is a qk(i0(ϵ),ki0(ϵ))q_{k}\in(i_{0}(\epsilon),k-i_{0}(\epsilon)) such that

      Uk(𝐓i)\displaystyle U_{k}(\mathbf{T}_{i}) u0(𝐓i),i0(ϵ)iqk1;\displaystyle\geq u_{0}(\mathbf{T}_{i}),\quad i_{0}(\epsilon)\leq i\leq q_{k}-1; (3.18)
      Uk(𝐓qk)\displaystyle U_{k}(\mathbf{T}_{q_{k}}) <u0(𝐓qk).\displaystyle<u_{0}(\mathbf{T}_{q_{k}}).

      Let Wk:=τqk1UkW_{k}:=\tau^{1}_{-q_{k}}U_{k}. Then proceeding as for UU, we have that WkWW_{k}\rightarrow W pointwise for some WW satisfying W+v0Γ^1(v0,w0)W+v_{0}\in\hat{\Gamma}_{1}(v_{0},w_{0}) and W(𝟎)u0(𝟎)W(\mathbf{0})\leq u_{0}(\mathbf{0}). If J1(W+v0)=J_{1}(W+v_{0})=\infty, then similar to (1), W+v0W+v_{0} is a new solution which is different with U+v0U+v_{0} since J1(U+v0)<J_{1}(U+v_{0})<\infty. So we assume J1(W+v0)<J_{1}(W+v_{0})<\infty. We claim

      qk as k.q_{k}\to\infty\textrm{\quad as \quad}k\to\infty. (3.19)

      Suppose (3.19) holds for the moment. Then (3.18) implies

      W(𝐢)u0(𝐢),𝐢1<0.W(\mathbf{i})\geq u_{0}(\mathbf{i}),\quad\mathbf{i}_{1}<0. (3.20)

      Hence applying Lemma 2.7 shows that

      |(W+v0w0)(𝐓i)|0,i.|(W+v_{0}-w_{0})(\mathbf{T}_{i})|\rightarrow 0,\quad i\rightarrow-\infty.

      Since we assume there is no subsequence of τjk1uk+v0\tau_{-j_{k}}^{1}u_{k}+v_{0} converging to homoclinic solution, so W+v0W+v_{0} is a heteroclinic solution from w0w_{0} to v0v_{0}.

      What is left is to show (3.19). Suppose, by contradiction, that qkq_{k} is bounded. Then up to a subsequence, we can assume qkq>i0(ϵ)q_{k}\equiv q>i_{0}(\epsilon) and W=τq1UW=\tau^{1}_{-q}U. Thus

      U(𝐓i)u0(𝐓i),i0(ϵ)iq1;\displaystyle U(\mathbf{T}_{i})\geq u_{0}(\mathbf{T}_{i}),\quad i_{0}(\epsilon)\leq i\leq q-1;
      U(𝐓q)u0(𝐓q).\displaystyle U(\mathbf{T}_{q})\leq u_{0}(\mathbf{T}_{q}).

      But by (3.17) for ϵ\epsilon small enough, U(𝐓q)>u0(𝐓q)U(\mathbf{T}_{q})>u_{0}(\mathbf{T}_{q}), a contradiction. Thus (3.19) is proved.

Summarizing the above discussion, we have the following Table 1, where (3.21) is

there is no subsequence of τjk1uk\tau_{-j_{k}}^{1}u_{k} converging to homoclinic solution. (3.21)
Table 1. Limits of τjk1uk\tau_{-j_{k}}^{1}u_{k}
J1(U+v0)J_{1}(U+v_{0}) U+v0U+v_{0}
(1): ==\infty A new solution
(2): <<\infty (2a): Homoclinic solution
(2b): Heteroclinic solution (say, from v0v_{0} to w0w_{0}) and (3.21) holds J1(V+v0)J_{1}(V+v_{0}) V+v0V+v_{0}
==\infty A new solution
<<\infty Heteroclinic solution
(from w0w_{0} to v0v_{0})

4.  Mountain pass solutions in the gap of 1(v0,w0)\mathcal{M}_{1}(v_{0},w_{0})

We construct heteroclinic mountain pass solution in this section. The difference is that the sum in the definition of I0𝐩I_{0}^{\mathbf{p}} of (2.1) involves only finite many terms, but in this section a new functional, I1𝐪I_{1}^{\mathbf{q}}, will be a sum of infinitely many terms. Suppose that (0*_{0}) and (1*_{1}) hold. For 𝐪=(𝐪2,,𝐪n)n1\mathbf{q}=(\mathbf{q}_{2},\cdots,\mathbf{q}_{n})\in\mathbb{N}^{n-1}, set

Λ1𝐪={u×n1/(𝐪n1)|uΛ1𝐪:=u1+u2:=𝐣×𝕋1𝐪|u(𝐣)|+𝐣×𝕋1𝐪|u(𝐣)|2<}.\begin{split}\Lambda_{1}^{\mathbf{q}}=\{u\in\mathbb{R}^{\mathbb{Z}\times\mathbb{Z}^{n-1}/(\mathbf{q}\mathbb{Z}^{n-1})}\,|\,\left\lVert u\right\rVert_{\Lambda_{1}^{\mathbf{q}}}:=&\left\lVert u\right\rVert_{\ell^{1}}+\left\lVert u\right\rVert_{\ell^{2}}\\ :=&\sum_{\mathbf{j}\in\mathbb{Z}\times\mathbb{T}_{1}^{\mathbf{q}}}|u(\mathbf{j})|+\sqrt{\sum_{\mathbf{j}\in\mathbb{Z}\times\mathbb{T}_{1}^{\mathbf{q}}}|u(\mathbf{j})|^{2}}<\infty\}.\end{split} (4.1)

Obviously (Λ1𝐪,Λ1𝐪)(\Lambda_{1}^{\mathbf{q}},\left\lVert\cdot\right\rVert_{\Lambda_{1}^{\mathbf{q}}}) is a Banach space with norm Λ1𝐪\left\lVert\cdot\right\rVert_{\Lambda_{1}^{\mathbf{q}}}. In fact, this norm is 1(×𝕋1𝐪)+2(×𝕋1𝐪)\left\lVert\cdot\right\rVert_{\ell^{1}(\mathbb{Z}\times\mathbb{T}_{1}^{\mathbf{q}})}+\left\lVert\cdot\right\rVert_{\ell^{2}(\mathbb{Z}\times\mathbb{T}_{1}^{\mathbf{q}})} on 1(×𝕋1𝐪)2(×𝕋1𝐪)\ell^{1}(\mathbb{Z}\times\mathbb{T}_{1}^{\mathbf{q}})\cap\ell^{2}(\mathbb{Z}\times\mathbb{T}_{1}^{\mathbf{q}}).

Remark 4.1.

The norm 2\|\cdot\|_{\ell^{2}} will be used in a similar result of Lemma 3.4 (v) of the deformation lemma. In the proof of Lemma 3.4 (v), 2\|\cdot\|_{\ell^{2}} can be replaced by equivalent norm 1\|\cdot\|_{\ell^{1}} since there are only finite many terms. But for infinitely many term in the definition of I1𝐪I_{1}^{\mathbf{q}}, 2\|\cdot\|_{\ell^{2}} cannot be replaced by 1\|\cdot\|_{\ell^{1}} any more. Noticing that to show I1𝐪I_{1}^{\mathbf{q}} is C1C^{1}, one need the norm 1\|\cdot\|_{\ell^{1}}, which cannot be replaced by 2\|\cdot\|_{\ell^{2}}. Please see the proof of Proposition 4.3.

Note that

w1(𝐢)v1(𝐢)w0(𝟎)v0(𝟎)1w_{1}(\mathbf{i})-v_{1}(\mathbf{i})\leq w_{0}(\mathbf{0})-v_{0}(\mathbf{0})\leq 1 (4.2)

and

[v1(𝐢),w1(𝐢))[v1(𝐣),w1(𝐣))=for 𝐢1𝐣1,\Big{[}v_{1}(\mathbf{i}),w_{1}(\mathbf{i})\Big{)}\cap\Big{[}v_{1}(\mathbf{j}),w_{1}(\mathbf{j})\Big{)}=\emptyset\quad\textrm{for }\mathbf{i}_{1}\neq\mathbf{j}_{1},

so

𝐣×𝕋1𝐪|w1(𝐣)v1(𝐣)|C(𝐪)\sum_{\mathbf{j}\in\mathbb{Z}\times\mathbb{T}_{1}^{\mathbf{q}}}|w_{1}(\mathbf{j})-v_{1}(\mathbf{j})|\leq C(\mathbf{q})

for some constant C(𝐪)C(\mathbf{q}) depending only on 𝐪\mathbf{q}. Thus by (4.2),

𝐣×𝕋1𝐪|w1(𝐣)v1(𝐣)|2𝐣×𝕋1𝐪|w1(𝐣)v1(𝐣)|C(𝐪)\sum_{\mathbf{j}\in\mathbb{Z}\times\mathbb{T}_{1}^{\mathbf{q}}}|w_{1}(\mathbf{j})-v_{1}(\mathbf{j})|^{2}\leq\sum_{\mathbf{j}\in\mathbb{Z}\times\mathbb{T}_{1}^{\mathbf{q}}}|w_{1}(\mathbf{j})-v_{1}(\mathbf{j})|\leq C(\mathbf{q})

and hence w1v1Λ1𝐪<\left\lVert w_{1}-v_{1}\right\rVert_{\Lambda_{1}^{\mathbf{q}}}<\infty, i.e., w1v1Λ1𝐪w_{1}-v_{1}\in\Lambda_{1}^{\mathbf{q}}.

For uΛ1𝐪u\in\Lambda_{1}^{\mathbf{q}}, define

I1;p,q𝐪(u):=J1;p,q𝐪(u+v1),I_{1;p,q}^{\mathbf{q}}(u):=J_{1;p,q}^{\mathbf{q}}(u+v_{1}),

where J1;p,q𝐪J_{1;p,q}^{\mathbf{q}} is defined in Section 2.2. For simplicity, set c1=c1(v0,w0)c_{1}=c_{1}(v_{0},w_{0}) and c1𝐪=c1𝐪(v0,w0)c_{1}^{\mathbf{q}}=c_{1}^{\mathbf{q}}(v_{0},w_{0}). Since uΛ1𝐪u\in\Lambda_{1}^{\mathbf{q}} implies |u(𝐢)|j0|u(\mathbf{i})|\leq j_{0} for some j0=j0(u)j_{0}=j_{0}(u)\in\mathbb{N}, we have v0j0u+v1w0+j0v_{0}-j_{0}\leq u+v_{1}\leq w_{0}+j_{0} and

limi𝐣{i}×𝕋1𝐪|(u+v1)(𝐣)v0(𝐣)|=0,limi𝐣{i}×𝕋1𝐪|(u+v1)(𝐣)w0(𝐣)|=0.\begin{split}\lim_{i\to-\infty}\sum_{\mathbf{j}\in\{i\}\times\mathbb{T}^{\mathbf{q}}_{1}}|(u+v_{1})(\mathbf{j})-v_{0}(\mathbf{j})|&=0,\\ \lim_{i\to\infty}\sum_{\mathbf{j}\in\{i\}\times\mathbb{T}^{\mathbf{q}}_{1}}|(u+v_{1})(\mathbf{j})-w_{0}(\mathbf{j})|&=0.\end{split}

Thus by Lemma 2.8 and Remark 2.9,

I1𝐪(u):=lim infpqI1;p,q𝐪(u)I_{1}^{\mathbf{q}}(u):=\liminf_{p\to-\infty\atop q\to\infty}I_{1;p,q}^{\mathbf{q}}(u)

is well-defined and if I1𝐪(u)<I_{1}^{\mathbf{q}}(u)<\infty, then

I1𝐪(u)=limpqI1;p,q𝐪(u), i.e., I1𝐪(u)=iI1,i𝐪(u).I_{1}^{\mathbf{q}}(u)=\lim_{p\to-\infty\atop q\to\infty}I_{1;p,q}^{\mathbf{q}}(u),\quad\textrm{ i.e., }\quad I_{1}^{\mathbf{q}}(u)=\sum_{i\in\mathbb{Z}}I_{1,i}^{\mathbf{q}}(u). (4.3)

Since we use a modified Mountain Pass Theorem to show the existence of critical point, the functional should be well-defined from Λ1𝐪\Lambda_{1}^{\mathbf{q}} to \mathbb{R} and be C1C^{1}. Fortunately, this is the case, as the following two propositions show.

Proposition 4.2.

For any uΛ1𝐪u\in\Lambda_{1}^{\mathbf{q}}, I1𝐪(u)<I_{1}^{\mathbf{q}}(u)<\infty and thus (4.3) holds.

Proof.

Assume uΛ1𝐪u\in\Lambda_{1}^{\mathbf{q}}, then there exists j0=j0(u)j_{0}=j_{0}(u)\in\mathbb{N} such that v0j0v1,v1+uw0+j0v_{0}-j_{0}\leq v_{1},v_{1}+u\leq w_{0}+j_{0}. Thus v1,v1+uv_{1},v_{1}+u have bounded action. By Lemma 2.4, there exists some L=L(u,r)(=L(j0,r))L=L(u,r)(=L(j_{0},r)), such that

|I1;p,q𝐪(u)I1;p,q𝐪(0)|L𝐣[p,q]×𝕋1𝐪¯|u(𝐣)|LC(r)uΛ1𝐪,|I_{1;p,q}^{\mathbf{q}}(u)-I_{1;p,q}^{\mathbf{q}}(0)|\leq L\sum_{\mathbf{j}\in\overline{[p,q]\times\mathbb{T}_{1}^{\mathbf{q}}}}|u(\mathbf{j})|\leq L\cdot C(r)\left\lVert u\right\rVert_{\Lambda_{1}^{\mathbf{q}}},

where C(r)C(r) is a constant depending on rr. Thus

I1;p,q𝐪(u)I1;p,q𝐪(0)+LC(r)uΛ1𝐪(I1𝐪(0)+2K1)+LC(r)uΛ1𝐪=(c1𝐪+2K1)+LC(r)uΛ1𝐪<,\begin{split}I_{1;p,q}^{\mathbf{q}}(u)\leq&I_{1;p,q}^{\mathbf{q}}(0)+L\cdot C(r)\left\lVert u\right\rVert_{\Lambda_{1}^{\mathbf{q}}}\\ \leq&(I_{1}^{\mathbf{q}}(0)+2K_{1})+L\cdot C(r)\left\lVert u\right\rVert_{\Lambda_{1}^{\mathbf{q}}}\\ =&(c_{1}^{\mathbf{q}}+2K_{1})+L\cdot C(r)\left\lVert u\right\rVert_{\Lambda_{1}^{\mathbf{q}}}\\ <&\infty,\end{split} (4.4)

where the second inequality follows from (2.3) with K1=K1(𝐪,u,v0,w0)K_{1}=K_{1}(\mathbf{q},u,v_{0},w_{0}). Then I1𝐪(u)<I_{1}^{\mathbf{q}}(u)<\infty and thus (4.3) follows. ∎

Proposition 4.3.

We have I1𝐪C1(Λ1𝐪,)I_{1}^{\mathbf{q}}\in C^{1}(\Lambda_{1}^{\mathbf{q}},\mathbb{R}). If (I1𝐪)(u)=0(I_{1}^{\mathbf{q}})^{\prime}(u)=0, i.e., uu is a critical point of I1𝐪I_{1}^{\mathbf{q}}, then u+v1u+v_{1} is a solution of (1.1).

Proof.

Firstly, we prove that I1𝐪I_{1}^{\mathbf{q}} is Gateaux differentiable. For u,vΛ1𝐪u,v\in\Lambda_{1}^{\mathbf{q}}, and t[1,1]{0}t\in[-1,1]\setminus\{0\},

|𝐣{i}×𝕋1𝐪S𝐣(u+tv+v1)S𝐣(u+v1)t|L𝐣{i}×𝕋1𝐪¯|v(𝐣)|=:Mi.\Big{|}\sum_{\mathbf{j}\in\{i\}\times\mathbb{T}_{1}^{\mathbf{q}}}\frac{S_{\mathbf{j}}(u+tv+v_{1})-S_{\mathbf{j}}(u+v_{1})}{t}\Big{|}\leq L\sum_{\mathbf{j}\in\overline{\{i\}\times\mathbb{T}_{1}^{\mathbf{q}}}}|v(\mathbf{j})|=:M_{i}.

Since vΛ1𝐪v\in\Lambda_{1}^{\mathbf{q}}, iMi<\sum_{i\in\mathbb{Z}}M_{i}<\infty. Thus we have

(I1𝐪)(u)v=i𝐣{i}×𝕋1𝐪𝐤:𝐤𝐣r𝐤S𝐣(u+v1)v(𝐤)=i𝐣{i}×𝕋1𝐪v(𝐣)𝐤:𝐤𝐣r𝐣S𝐤(u+v1).\begin{split}(I_{1}^{\mathbf{q}})^{\prime}(u)v=&\sum_{i\in\mathbb{Z}}\sum_{\mathbf{j}\in\{i\}\times\mathbb{T}_{1}^{\mathbf{q}}}\sum_{\mathbf{k}:\left\lVert\mathbf{k}-\mathbf{j}\right\rVert\leq r}\partial_{\mathbf{k}}S_{\mathbf{j}}(u+v_{1})\cdot v(\mathbf{k})\\ =&\sum_{i\in\mathbb{Z}}\sum_{\mathbf{j}\in\{i\}\times\mathbb{T}_{1}^{\mathbf{q}}}v(\mathbf{j})\sum_{\mathbf{k}:\left\lVert\mathbf{k}-\mathbf{j}\right\rVert\leq r}\partial_{\mathbf{j}}S_{\mathbf{k}}(u+v_{1}).\end{split} (4.5)

To show that (I1𝐪)(I_{1}^{\mathbf{q}})^{\prime} is continuous, set umuu_{m}\to u in Λ1𝐪\Lambda_{1}^{\mathbf{q}} as mm\to\infty. Then

|[(I1𝐪)(um)(I1𝐪)(u)]v|=|i𝐣{i}×𝕋1𝐪v(𝐣)𝐤:𝐤𝐣r[𝐣S𝐤(um+v1)𝐣S𝐤(u+v1)]|=|i𝐣{i}×𝕋1𝐪v(𝐣)𝐤:𝐤𝐣r01ddt𝐣S𝐤(u+t(umu)+v1)dt|=|i𝐣{i}×𝕋1𝐪v(𝐣)𝐤:𝐤𝐣r01𝐥:𝐥𝐤r𝐣,𝐥S𝐤(u+t(umu)+v1)dt(umu)(𝐥)|Ci𝐣{i}×𝕋1𝐪|v(𝐣)|𝐤:𝐤𝐣r𝐥:𝐥𝐤r|(umu)(𝐥)|.\begin{split}&|[(I_{1}^{\mathbf{q}})^{\prime}(u_{m})-(I_{1}^{\mathbf{q}})^{\prime}(u)]v|\\ =&\Big{|}\sum_{i\in\mathbb{Z}}\sum_{\mathbf{j}\in\{i\}\times\mathbb{T}_{1}^{\mathbf{q}}}v(\mathbf{j})\sum_{\mathbf{k}:\left\lVert\mathbf{k}-\mathbf{j}\right\rVert\leq r}[\partial_{\mathbf{j}}S_{\mathbf{k}}(u_{m}+v_{1})-\partial_{\mathbf{j}}S_{\mathbf{k}}(u+v_{1})]\Big{|}\\ =&\Big{|}\sum_{i\in\mathbb{Z}}\sum_{\mathbf{j}\in\{i\}\times\mathbb{T}_{1}^{\mathbf{q}}}v(\mathbf{j})\sum_{\mathbf{k}:\left\lVert\mathbf{k}-\mathbf{j}\right\rVert\leq r}\int_{0}^{1}\frac{\mathrm{d}}{\mathrm{d}t}\partial_{\mathbf{j}}S_{\mathbf{k}}(u+t(u_{m}-u)+v_{1})\mathrm{d}t\Big{|}\\ =&\Big{|}\sum_{i\in\mathbb{Z}}\sum_{\mathbf{j}\in\{i\}\times\mathbb{T}_{1}^{\mathbf{q}}}v(\mathbf{j})\sum_{\mathbf{k}:\left\lVert\mathbf{k}-\mathbf{j}\right\rVert\leq r}\int_{0}^{1}\sum_{\mathbf{l}:\left\lVert\mathbf{l}-\mathbf{k}\right\rVert\leq r}\partial_{\mathbf{j},\mathbf{l}}S_{\mathbf{k}}(u+t(u_{m}-u)+v_{1})\mathrm{d}t\,\cdot(u_{m}-u)(\mathbf{l})\Big{|}\\ \leq&C\sum_{i\in\mathbb{Z}}\sum_{\mathbf{j}\in\{i\}\times\mathbb{T}_{1}^{\mathbf{q}}}|v(\mathbf{j})|\sum_{\mathbf{k}:\left\lVert\mathbf{k}-\mathbf{j}\right\rVert\leq r}\sum_{\mathbf{l}:\left\lVert\mathbf{l}-\mathbf{k}\right\rVert\leq r}|(u_{m}-u)(\mathbf{l})|.\end{split} (4.6)

So (I1𝐪)(um)(I1𝐪)(u)(I_{1}^{\mathbf{q}})^{\prime}(u_{m})\to(I_{1}^{\mathbf{q}})^{\prime}(u) as mm\to\infty. By (4.5), if uu is a critical point of I1𝐪I_{1}^{\mathbf{q}}, then u+v1u+v_{1} is a solution of (1.1). So the proof of Proposition 4.3 is complete. ∎

Now following Section 3, let us define semiflow Φt1:Λ1𝐪Λ1𝐪\Phi_{t}^{1}:\Lambda_{1}^{\mathbf{q}}\to\Lambda_{1}^{\mathbf{q}} as follows:

{tΦt1(u)(𝐢)=𝐣:𝐣𝐢r𝐢S𝐣(Φt1(u)+v1),for t>0,Φ01(u)(𝐢)=u(𝐢).\left\{\begin{array}[]{ll}-\partial_{t}\Phi_{t}^{1}(u)(\mathbf{i})&=\sum_{\mathbf{j}:\left\lVert\mathbf{j}-\mathbf{i}\right\rVert\leq r}\partial_{\mathbf{i}}S_{\mathbf{j}}(\Phi_{t}^{1}(u)+v_{1}),\quad\quad\textrm{for }t>0,\\ \Phi_{0}^{1}(u)(\mathbf{i})&=u(\mathbf{i}).\end{array}\right.

Set W(u)(𝐢):=𝐣:𝐣𝐢r𝐢S𝐣(u+v1)W(u)(\mathbf{i}):=\sum_{\mathbf{j}:\left\lVert\mathbf{j}-\mathbf{i}\right\rVert\leq r}\partial_{\mathbf{i}}S_{\mathbf{j}}(u+v_{1}). Similar proof of (3.3) shows that W(u)W(v)2Cuv2\|W(u)-W(v)\|_{\ell^{2}}\leq C\|u-v\|_{\ell^{2}} for some C=C(r)C=C(r). The proof of W(u)W(v)1Cuv1\|W(u)-W(v)\|_{\ell^{1}}\leq C\|u-v\|_{\ell^{1}} is easier (cf. (4.6)). So Φt1\Phi_{t}^{1} is well-defined and is C1C^{1} in tt. Moreover, a new version of Proposition 3.1 is obtained.

Set

𝒢1𝐪={uΛ1𝐪| 0uw1v1}.\mathcal{G}^{\mathbf{q}}_{1}=\left\{u\in\Lambda_{1}^{\mathbf{q}}\,|\,0\leq u\leq w_{1}-v_{1}\right\}.

Note that if u×n1/(𝐪n1)u\in\mathbb{R}^{\mathbb{Z}\times\mathbb{Z}^{n-1}/(\mathbf{q}\mathbb{Z}^{n-1})} and 0uw1v10\leq u\leq w_{1}-v_{1},

uΛ1𝐪w1v1Λ1𝐪<.\left\lVert u\right\rVert_{\Lambda_{1}^{\mathbf{q}}}\leq\left\lVert w_{1}-v_{1}\right\rVert_{\Lambda_{1}^{\mathbf{q}}}<\infty. (4.7)

In other words, {u×n1/(𝐪n1)| 0uw1v1}𝒢1𝐪\{u\in\mathbb{R}^{\mathbb{Z}\times\mathbb{Z}^{n-1}/(\mathbf{q}\mathbb{Z}^{n-1})}\,|\,0\leq u\leq w_{1}-v_{1}\}\subset\mathcal{G}^{\mathbf{q}}_{1}.

Proposition 4.4.

𝒢1𝐪\mathcal{G}_{1}^{\mathbf{q}} is compact with respect to the norm Λ1𝐪\left\lVert\cdot\right\rVert_{\Lambda_{1}^{\mathbf{q}}}.

Proof.

For any un𝒢1𝐪u_{n}\in\mathcal{G}_{1}^{\mathbf{q}}, by Lemma 2.2 and (4.7), there is a u𝒢1𝐪u\in\mathcal{G}_{1}^{\mathbf{q}} such that unuu_{n}\to u (maybe up to a subsequence) pointewise as nn\to\infty and |(unu)(𝐣)|(w1v1)(𝐣)|(u_{n}-u)(\mathbf{j})|\leq(w_{1}-v_{1})(\mathbf{j}). Note that

ukuΛ1𝐪=𝐣[N,N]×𝕋1𝐪|(uku)(𝐣)|+𝐣([N,N])×𝕋1𝐪|(uku)(𝐣)|+(𝐣[N,N]×𝕋1𝐪|(uku)(𝐣)|2+𝐣([N,N])×𝕋1𝐪|(uku)(𝐣)|2)12𝐣[N,N]×𝕋1𝐪|(uku)(𝐣)|+𝐣([N,N])×𝕋1𝐪|(uku)(𝐣)|+(𝐣[N,N]×𝕋1𝐪|(uku)(𝐣)|2)12+(𝐣([N,N])×𝕋1𝐪|(uku)(𝐣)|2)12\begin{split}\left\lVert u_{k}-u\right\rVert_{\Lambda_{1}^{\mathbf{q}}}=&\sum_{\mathbf{j}\in[-N,N]\times\mathbb{T}_{1}^{\mathbf{q}}}|(u_{k}-u)(\mathbf{j})|+\sum_{\mathbf{j}\in(\mathbb{Z}\setminus[-N,N])\times\mathbb{T}_{1}^{\mathbf{q}}}|(u_{k}-u)(\mathbf{j})|\\ \quad&+\left(\sum_{\mathbf{j}\in[-N,N]\times\mathbb{T}_{1}^{\mathbf{q}}}|(u_{k}-u)(\mathbf{j})|^{2}+\sum_{\mathbf{j}\in(\mathbb{Z}\setminus[-N,N])\times\mathbb{T}_{1}^{\mathbf{q}}}|(u_{k}-u)(\mathbf{j})|^{2}\right)^{\frac{1}{2}}\\ \leq&\sum_{\mathbf{j}\in[-N,N]\times\mathbb{T}_{1}^{\mathbf{q}}}|(u_{k}-u)(\mathbf{j})|+\sum_{\mathbf{j}\in(\mathbb{Z}\setminus[-N,N])\times\mathbb{T}_{1}^{\mathbf{q}}}|(u_{k}-u)(\mathbf{j})|\\ \quad&+\left(\sum_{\mathbf{j}\in[-N,N]\times\mathbb{T}_{1}^{\mathbf{q}}}|(u_{k}-u)(\mathbf{j})|^{2}\right)^{\frac{1}{2}}+\left(\sum_{\mathbf{j}\in(\mathbb{Z}\setminus[-N,N])\times\mathbb{T}_{1}^{\mathbf{q}}}|(u_{k}-u)(\mathbf{j})|^{2}\right)^{\frac{1}{2}}\end{split}

and

𝐣([N,N])×𝕋1𝐪|(uku)(𝐣)|+(𝐣([N,N])×𝕋1𝐪|(uku)(𝐣)|2)12𝐣([N,N])×𝕋1𝐪(w1v1)(𝐣)+(𝐣([N,N])×𝕋1𝐪|(w1v1)(𝐣)|2)120\begin{split}&\sum_{\mathbf{j}\in(\mathbb{Z}\setminus[-N,N])\times\mathbb{T}_{1}^{\mathbf{q}}}|(u_{k}-u)(\mathbf{j})|+\left(\sum_{\mathbf{j}\in(\mathbb{Z}\setminus[-N,N])\times\mathbb{T}_{1}^{\mathbf{q}}}|(u_{k}-u)(\mathbf{j})|^{2}\right)^{\frac{1}{2}}\\ \leq&\sum_{\mathbf{j}\in(\mathbb{Z}\setminus[-N,N])\times\mathbb{T}_{1}^{\mathbf{q}}}(w_{1}-v_{1})(\mathbf{j})+\left(\sum_{\mathbf{j}\in(\mathbb{Z}\setminus[-N,N])\times\mathbb{T}_{1}^{\mathbf{q}}}|(w_{1}-v_{1})(\mathbf{j})|^{2}\right)^{\frac{1}{2}}\\ \to&0\end{split}

as NN\to\infty. Thus for any ϵ>0\epsilon>0, one can choose NN sufficiently large such that

𝐣([N,N])×𝕋1𝐪|(uku)(𝐣)|+(𝐣([N,N])×𝕋1𝐪|(uku)(𝐣)|2)12ϵ,\sum_{\mathbf{j}\in(\mathbb{Z}\setminus[-N,N])\times\mathbb{T}_{1}^{\mathbf{q}}}|(u_{k}-u)(\mathbf{j})|+\left(\sum_{\mathbf{j}\in(\mathbb{Z}\setminus[-N,N])\times\mathbb{T}_{1}^{\mathbf{q}}}|(u_{k}-u)(\mathbf{j})|^{2}\right)^{\frac{1}{2}}\leq\epsilon,

so

limkukuΛ1𝐪limk𝐣[N,N]×𝕋1𝐪|(uku)(𝐣)|+(𝐣[N,N]×𝕋1𝐪|(uku)(𝐣)|2)12+ϵ=ϵ.\begin{split}&\lim_{k\to\infty}\left\lVert u_{k}-u\right\rVert_{\Lambda_{1}^{\mathbf{q}}}\\ \leq&\lim_{k\to\infty}\sum_{\mathbf{j}\in[-N,N]\times\mathbb{T}_{1}^{\mathbf{q}}}|(u_{k}-u)(\mathbf{j})|+\left(\sum_{\mathbf{j}\in[-N,N]\times\mathbb{T}_{1}^{\mathbf{q}}}|(u_{k}-u)(\mathbf{j})|^{2}\right)^{\frac{1}{2}}+\epsilon\\ =&\epsilon.\end{split}

Since ϵ\epsilon is arbitrary, Proposition 4.4 is proved. ∎

Set

1𝐪={hC([0,1],𝒢1𝐪)|h(0)=0,h(1)=w1v1}.\mathcal{H}^{\mathbf{q}}_{1}=\left\{h\in C\left([0,1],\mathcal{G}^{\mathbf{q}}_{1}\right)\,|\,h(0)=0,h(1)=w_{1}-v_{1}\right\}.

Hence one have new versions of Lemma 3.4, Proposition 3.5 and Theorem 3.6. Thus we obtain a mountain pass critical point u𝐪u_{\mathbf{q}} and then a heteroclinic mountain pass solution u𝐪+v1u_{\mathbf{q}}+v_{1} satisfying v1<u𝐪+v1<w1v_{1}<u_{\mathbf{q}}+v_{1}<w_{1} for any 𝐪n1\mathbf{q}\in\mathbb{N}^{n-1}. Next we study the multiplicity of heteroclinic mountain pass solutions. It sufficies to prove a silimlar result of Proposition 3.12. To this end, for each kk\in\mathbb{N}, let 𝐪(k)=(k,1,,1)n1\mathbf{q}(k)=(k,1,\cdots,1)\in\mathbb{N}^{n-1}.

Set

d1𝐪(k)=infh1𝐪(k)maxθ[0,1]I1𝐪(k)(h(θ)).d_{1}^{\mathbf{q}(k)}=\inf_{h\in\mathcal{H}_{1}^{\mathbf{q}(k)}}\max_{\theta\in[0,1]}I_{1}^{\mathbf{q}(k)}(h(\theta)).

Then d1𝐪(k)d_{1}^{\mathbf{q}(k)} is a mountain pass critical value of I1𝐪(k)I_{1}^{\mathbf{q}(k)} on 𝒢1𝐪(k)\mathcal{G}_{1}^{\mathbf{q}(k)} with a corresponding mountain pass critical point UkU_{k} such that 0<Uk<w1v10<U_{k}<w_{1}-v_{1} with I1𝐪(k)(Uk)=d1𝐪(k)>c1𝐪(k)I_{1}^{\mathbf{q}(k)}(U_{k})=d_{1}^{\mathbf{q}(k)}>c_{1}^{\mathbf{q}(k)} and Uk+v1U_{k}+v_{1} is a solution of (1.1). We have:

Proposition 4.5.

There is a constant M1M_{1}, independent of kk, such that

0<d1𝐪(k)c1𝐪(k)M1.0<d_{1}^{\mathbf{q}(k)}-c_{1}^{\mathbf{q}(k)}\leq M_{1}.
Proof.

For uΛ1𝐪(k)u\in\Lambda_{1}^{\mathbf{q}(k)}, let I1(u):=𝐣×{0}n1[S𝐣(u+v1)c0]I_{1}(u):=\sum_{\mathbf{j}\in\mathbb{Z}\times\{0\}^{n-1}}[S_{\mathbf{j}}(u+v_{1})-c_{0}], then

I1𝐪(k)(u)=i=0k1I1(τi2u).I_{1}^{\mathbf{q}(k)}(u)=\sum_{i=0}^{k-1}I_{1}(\tau^{2}_{-i}u).

Set hk(θ)=ϕk(θ,𝐢2)(w1v1),h_{k}(\theta)=\phi_{k}\left(\theta,\mathbf{i}_{2}\right)\left(w_{1}-v_{1}\right), where ϕk\phi_{k} is defined in (3.11). To prove Proposition 4.5, it suffices to show that

maxθ[0,1]I1𝐪(k)(hk(θ))c1𝐪(k)M1\max_{\theta\in[0,1]}I_{1}^{\mathbf{q}(k)}\left(h_{k}(\theta)\right)-c_{1}^{\mathbf{q}(k)}\leq M_{1} (4.8)

holds for some M1M_{1} independent of kk. Note

I1𝐪(k)(hk(θ))\displaystyle I_{1}^{\mathbf{q}(k)}\left(h_{k}(\theta)\right) =𝐣×[0,k1]×{0}n2[S𝐣(hk(θ)+v1)c0]\displaystyle=\sum_{\mathbf{j}\in\mathbb{Z}\times[0,k-1]\times\{0\}^{n-2}}[S_{\mathbf{j}}(h_{k}(\theta)+v_{1})-c_{0}]
=𝐣×[a,a+k1]×{0}n2[S𝐣(hk(θ)+v1)c0]\displaystyle=\sum_{\mathbf{j}\in\mathbb{Z}\times[a,a+k-1]\times\{0\}^{n-2}}[S_{\mathbf{j}}(h_{k}(\theta)+v_{1})-c_{0}]

for any aa\in\mathbb{Z}. For any θ[0,1]\theta\in[0,1], let

𝒜:={𝐢n|0𝐢2k1,𝐢3==𝐢n=0,ϕk(θ,𝐢2)0 or ϕk(θ,𝐢2)1}.\begin{split}\mathcal{A}:=\{\mathbf{i}\in\mathbb{Z}^{n}\,|\,&0\leq\mathbf{i}_{2}\leq k-1,\mathbf{i}_{3}=\cdots=\mathbf{i}_{n}=0,\\ &\phi_{k}\left(\theta,\mathbf{i}_{2}\right)\neq 0\text{ or }\phi_{k}\left(\theta,\mathbf{i}_{2}\right)\neq 1\}.\end{split}

By the construction of ϕk\phi_{k}, 𝒜\mathcal{A} consists of at most two regions, say RiR_{i} (i=1,2i=1,2) of the form Ri=×{ai}×{0}n2R_{i}=\mathbb{Z}\times\{a_{i}\}\times\{0\}^{n-2} with a1a2a_{1}\leq a_{2}. Therefore

hk=0 or w1v1 on i=0k1(×{i}×{0}n2)\(R1R2).\textrm{$h_{k}=0$ or $w_{1}-v_{1}$ on $\bigcup_{i=0}^{k-1}(\mathbb{Z}\times\{i\}\times\{0\}^{n-2})\backslash\left(R_{1}\cup R_{2}\right)$}. (4.9)

Then

I1𝐪(k)(hk(θ))c1𝐪(k)=i[a,a+k1]{𝐣×{i}×{0}[S𝐣(hk(θ)+v1)c0]c1}=i[a,a+k1]{𝐣×{i}×{0}[S𝐣(hk(θ)+v1)c0]𝐣×{i}×{0}[S𝐣(v1)c0]}=i[a,a+k1]𝐣×{i}×{0}[S𝐣(hk(θ)+v1)S𝐣(v1)].\begin{split}&I_{1}^{\mathbf{q}(k)}\left(h_{k}(\theta)\right)-c_{1}^{\mathbf{q}(k)}\\ =&\sum_{i\in[a,a+k-1]}\left\{\sum_{\mathbf{j}\in\mathbb{Z}\times\{i\}\times\{0\}}[S_{\mathbf{j}}(h_{k}(\theta)+v_{1})-c_{0}]-c_{1}\right\}\\ =&\sum_{i\in[a,a+k-1]}\left\{\sum_{\mathbf{j}\in\mathbb{Z}\times\{i\}\times\{0\}}[S_{\mathbf{j}}(h_{k}(\theta)+v_{1})-c_{0}]-\sum_{\mathbf{j}\in\mathbb{Z}\times\{i\}\times\{0\}}[S_{\mathbf{j}}(v_{1})-c_{0}]\right\}\\ =&\sum_{i\in[a,a+k-1]}\sum_{\mathbf{j}\in\mathbb{Z}\times\{i\}\times\{0\}}[S_{\mathbf{j}}(h_{k}(\theta)+v_{1})-S_{\mathbf{j}}(v_{1})].\end{split} (4.10)

Then by (4.9), the cardinality of

{i|𝐣×{i}×{0}S𝐣(hk(θ)+v1)S𝐣(v1)0}\{i\in\mathcal{B}\,|\,\sum_{\mathbf{j}\in\mathbb{Z}\times\{i\}\times\{0\}}S_{\mathbf{j}}(h_{k}(\theta)+v_{1})-S_{\mathbf{j}}(v_{1})\neq 0\}

is at most a finite number, denoted by C1C_{1}, independent of kk, where

:={i=a1ra1r+k1{i}i=a1ra2+r{i},if a2a12r;i=a1ra1r+k1{i}(i=a1ra1+r{i}i=a2ra2+r{i}),if 2r<a2a1 and a1+ka2>2r;i=a2ra2r+k1{i}i=a2ra1+r+k{i},if 2r<a2a1 and a1+ka22r.\mathcal{B}:=\left\{\begin{array}[]{ll}{\cup_{i=a_{1}-r}^{a_{1}-r+k-1}\{i\}}\setminus{\cup_{i=a_{1}-r}^{a_{2}+r}\{i\}},&\textrm{if $a_{2}-a_{1}\leq 2r$;}\\ {\cup_{i=a_{1}-r}^{a_{1}-r+k-1}\{i\}}\setminus{(\cup_{i=a_{1}-r}^{a_{1}+r}\{i\}\cup\cup_{i=a_{2}-r}^{a_{2}+r}\{i\})},&\textrm{if $2r<a_{2}-a_{1}$ and $a_{1}+k-a_{2}>2r$;}\\ {\cup_{i=a_{2}-r}^{a_{2}-r+k-1}\{i\}}\setminus{\cup_{i=a_{2}-r}^{a_{1}+r+k}\{i\}},&\textrm{if $2r<a_{2}-a_{1}$ and $a_{1}+k-a_{2}\leq 2r$.}\end{array}\right.

Noticing that for any ii\in\mathbb{Z},

|𝐣×{i}×{0}[I1(hk(θ))c1]|=|𝐣×{i}×{0}S𝐣(hk(θ)+v1)S𝐣(v1)|=|𝐣×{i}×{0}01𝐤:𝐤𝐣r𝐤S𝐣(v1+thk(θ))dthk(θ)(𝐤)|L𝐤×{i}×{0}¯|hk(θ)(𝐤)|L𝐤×{i}×{0}¯[w1(𝐤)v1(𝐤)]LCw1v1Λ1𝐪=:M.\begin{split}&|\sum_{\mathbf{j}\in\mathbb{Z}\times\{i\}\times\{0\}}[I_{1}(h_{k}(\theta))-c_{1}]|\\ =&|\sum_{\mathbf{j}\in\mathbb{Z}\times\{i\}\times\{0\}}S_{\mathbf{j}}(h_{k}(\theta)+v_{1})-S_{\mathbf{j}}(v_{1})|\\ =&|\sum_{\mathbf{j}\in\mathbb{Z}\times\{i\}\times\{0\}}\int_{0}^{1}\sum_{\mathbf{k}:\left\lVert\mathbf{k}-\mathbf{j}\right\rVert\leq r}\partial_{\mathbf{k}}S_{\mathbf{j}}(v_{1}+th_{k}(\theta))\mathrm{d}t\cdot h_{k}(\theta)(\mathbf{k})|\\ \leq&L\sum_{\mathbf{k}\in\overline{\mathbb{Z}\times\{i\}\times\{0\}}}|h_{k}(\theta)(\mathbf{k})|\\ \leq&L\sum_{\mathbf{k}\in\overline{\mathbb{Z}\times\{i\}\times\{0\}}}[w_{1}(\mathbf{k})-v_{1}(\mathbf{k})]\\ \leq&LC\left\lVert w_{1}-v_{1}\right\rVert_{\Lambda_{1}^{\mathbf{q}}}\\ =:&M.\end{split} (4.11)

Here L=L(w0v0,r)L=L(w_{0}-v_{0},r), C=C(r)C=C(r) are constants and ×{i}×{0}¯=𝐣×{i}×{0}{𝐤n|𝐤𝐣r}\overline{\mathbb{Z}\times\{i\}\times\{0\}}=\cup_{\mathbf{j}\in\mathbb{Z}\times\{i\}\times\{0\}}\{\mathbf{k}\in\mathbb{Z}^{n}\,|\,\left\lVert\mathbf{k}-\mathbf{j}\right\rVert\leq r\}. The first inequality in (4.11) needs to be explained. Since v1+thk(θ)v_{1}+th_{k}(\theta) have bounded action with bounded constant w0v0w_{0}-v_{0} for all t[0,1]t\in[0,1], thus by (S1), there exists L=L(w0v0,r)L=L(w_{0}-v_{0},r) such that the first inequality in (4.11) holds (cf. [14, the proof of Lemma 2.4]).

By (4.4) and (4.7), there is an M(𝐪)>0M(\mathbf{q})>0 such that I1𝐪(u)c1𝐪M(𝐪)I_{1}^{\mathbf{q}}(u)-c_{1}^{\mathbf{q}}\leq M(\mathbf{q}) for all u𝒢1𝐪.u\in\mathcal{G}_{1}^{\mathbf{q}}. So without loss of generality, assume k>4r+1k>4r+1 and Mmax(M(𝐪(1)),,M(𝐪(4r)))M\geq\max(M(\mathbf{q}(1)),\cdots,M(\mathbf{q}(4r))). We have the following three cases.

  • If a2a12ra_{2}-a_{1}\leq 2r,

    |I1𝐪(k)(u)kc1|=|i=a1ra2+r[I1(τi2u)c1]+i[I1(τi2u)c1]|M(a2+2r+1a1)+MC1M(4r+1+C1).\begin{split}&|I_{1}^{\mathbf{q}(k)}(u)-kc_{1}|\\ =&|\sum_{i=a_{1}-r}^{a_{2}+r}[I_{1}(\tau^{2}_{-i}u)-c_{1}]+\sum_{i\in\mathcal{B}}[I_{1}(\tau^{2}_{-i}u)-c_{1}]|\\ \leq&M\cdot(a_{2}+2r+1-a_{1})+M\cdot C_{1}\\ \leq&M\cdot(4r+1+C_{1}).\end{split}
  • If 2r<a2a12r<a_{2}-a_{1} and a1+ka2>2ra_{1}+k-a_{2}>2r, i.e., a1(a22r,a2+2r)a_{1}\not\in(a_{2}-2r,a_{2}+2r),

    |I1𝐪(k)(u)kc1|=|i=a1ra1+r[I1(τi2u)c1]+i=a2ra2+r[I1(τi2u)c1]+i[I1(τi2u)c1]|2M(2r+1)+MC1=2M(2r+1+C1).\begin{split}&|I_{1}^{\mathbf{q}(k)}(u)-kc_{1}|\\ =&|\sum_{i=a_{1}-r}^{a_{1}+r}[I_{1}(\tau_{-i}^{2}u)-c_{1}]+\sum_{i=a_{2}-r}^{a_{2}+r}[I_{1}(\tau_{-i}^{2}u)-c_{1}]+\sum_{i\in\mathcal{B}}[I_{1}(\tau^{2}_{-i}u)-c_{1}]|\\ \leq&2M\cdot(2r+1)+M\cdot C_{1}\\ =&2M\cdot(2r+1+C_{1}).\end{split}
  • If 2r<a2a12r<a_{2}-a_{1} and a1+ka22ra_{1}+k-a_{2}\leq 2r,

    |I1𝐪(k)(u)kc1|=|i=a2ra1+k+r[I1(τi2u)c1]+i[I1(τi2u)c1]|M(a1+k+2r+1a2)+MC1M(4r+1+C1).\begin{split}&|I_{1}^{\mathbf{q}(k)}(u)-kc_{1}|\\ =&|\sum_{i=a_{2}-r}^{a_{1}+k+r}[I_{1}(\tau_{-i}^{2}u)-c_{1}]+\sum_{i\in\mathcal{B}}[I_{1}(\tau^{2}_{-i}u)-c_{1}]|\\ \leq&M\cdot(a_{1}+k+2r+1-a_{2})+M\cdot C_{1}\\ \leq&M\cdot(4r+1+C_{1}).\end{split}

Thus (4.8) follows by setting M1=(4r+2+2C1)MM_{1}=(4r+2+2C_{1})M. ∎

Proceeding as in Theorem 3.13 we obtain infinitely many heteroclinic mountain pass solutions. When we want to go further as in Section 3 to see that heteroclinic mountain pass solutions do not constitute a foliation or laminaion, we encounter more difficulties. For instance, in the definition of θ¯t\underline{\theta}_{t}, Φt0h(θ)u0\Phi_{t}^{0}h(\theta)\leq u_{0} means either Φt0h(θ)\Phi_{t}^{0}h(\theta) touches u0u_{0} from below or Φt0h(θ)=u0\Phi_{t}^{0}h(\theta)=u_{0}. But for unbounded domain ×𝕋1𝐪\mathbb{Z}\times\mathbb{T}^{\mathbf{q}}_{1}, besides the above two possibilities, Φt1h(θ)u0Λ1𝐪\Phi_{t}^{1}h(\theta)\leq u_{0}\in\Lambda_{1}^{\mathbf{q}} may lead to Φt1h(θ)<u0\Phi_{t}^{1}h(\theta)<u_{0} and

Φt1h(θ)(𝐢)u0(𝐢)as |𝐢1|0.\Phi_{t}^{1}h(\theta)(\mathbf{i})\to u_{0}(\mathbf{i})\quad\textrm{as }|\mathbf{i}_{1}|\to 0. (4.12)

Notice that (4.12) always holds since (w1v1)(𝐢)0(w_{1}-v_{1})(\mathbf{i})\to 0 as |𝐢1|0|\mathbf{i}_{1}|\to 0. Thus if one want to show a result similar to Theorem 3.20, one need a new idea to exclude the third possibility in the proof of Theorem 3.20.

Remark 4.6.

We can obtain another heteroclinic mountain pass solution lying in the gap of 1(w0,v0)\mathcal{M}_{1}(w_{0},v_{0}) (please see [11] for the definition) and we can construct more heteroclinic mountain pass solutions by the methods of Sections 3 and 4 for higher dimension (cf. [11, Section 5]). Thus in the gap of the second laminations (in the sense of [14]) we have heteroclinic mountain pass solutions.

Remark 4.7.

Throughout this paper, only the minimal and Birkhoff solutions corresponding to rotation vector 𝟎\mathbf{0} are considered (for the definition of rotation vector, please see [11]). One can generalize the above results to minimal and Birkhoff solutions corresponding to rotation vector αn\alpha\in\mathbb{Q}^{n} and obtain corresponding mountain pass solutions.

Appendix A Appendix

In this appendix, we prove Proposition 3.1, Lemma 3.4 and Theorem 3.6 of Sections 3. First is the proof of Proposition 3.1, which follows [16, Theorem 6.2] with slight modifications.

Proof of Proposition 3.1. Define v(t):=Φt0(u2)Φt0(u1)v(t):=\Phi^{0}_{t}(u_{2})-\Phi^{0}_{t}(u_{1}). So v(0)0v(0)\geq 0, v(0)0v(0)\neq 0 and vv satisfies the following linear ODE:

v˙(t)(𝐢)\displaystyle\dot{v}(t)(\mathbf{i})
=\displaystyle= W(Φt0(u2))(𝐢)+W(Φt0(u1))(𝐢)\displaystyle-W(\Phi^{0}_{t}(u_{2}))(\mathbf{i})+W(\Phi^{0}_{t}(u_{1}))(\mathbf{i})
=\displaystyle= 01ddt~(𝐣:𝐢𝐣r𝐢S𝐣(Φt0(u1)+t~(Φt0(u2)Φt0(u1))+v0))dt~\displaystyle\int_{0}^{1}\frac{\mathrm{d}}{\mathrm{d}\tilde{t}}\left(-\sum_{\mathbf{j}:\|\mathbf{i}-\mathbf{j}\|\leq r}\partial_{\mathbf{i}}S_{\mathbf{j}}(\Phi^{0}_{t}(u_{1})+\tilde{t}(\Phi^{0}_{t}(u_{2})-\Phi^{0}_{t}(u_{1}))+v_{0})\right)\mathrm{d}\tilde{t}
=\displaystyle= 𝐣:𝐢𝐣r𝐤:𝐣𝐤r(01𝐢,𝐤S𝐣(Φt0(u1)+t~(Φt0(u2)Φt0(u1))+v0)dt~)v(t)(𝐤)\displaystyle\sum_{\mathbf{j}:\|\mathbf{i}-\mathbf{j}\|\leq r}\sum_{\mathbf{k}:\|\mathbf{j}-\mathbf{k}\|\leq r}\left(\int_{0}^{1}-\partial_{\mathbf{i},\mathbf{k}}S_{\mathbf{j}}(\Phi^{0}_{t}(u_{1})+\tilde{t}(\Phi^{0}_{t}(u_{2})-\Phi^{0}_{t}(u_{1}))+v_{0})\mathrm{d}\tilde{t}\right)v(t)(\mathbf{k})
=\displaystyle= :(H(t)v(t))(𝐢).\displaystyle:(H(t)v(t))(\mathbf{i}).

Similar calculation of (3.3) implies H(t):Λ0𝐩Λ0𝐩H(t):\Lambda_{0}^{\mathbf{p}}\to\Lambda_{0}^{\mathbf{p}} is Lipschitz. By (S3)-(S4), there is an M>0M>0 such that the operators H~(t):=H(t)+MId:dd\tilde{H}(t):=H(t)+M\cdot\mathrm{Id}:\mathbb{R}^{\mathbb{Z}^{d}}\rightarrow\mathbb{R}^{\mathbb{Z}^{d}} are positive: v0v\geq 0 implies H~(t)v0\tilde{H}(t)v\geq 0.

Note moreover that both the H(t)H(t) and the H~(t)\tilde{H}(t) are uniformly bounded operators, whence the ODEs v˙=H(t)v\dot{v}=H(t)v and w˙=H~(t)w\dot{w}=\tilde{H}(t)w define well-posed initial value problems. More importantly, v(t)v(t) solves v˙=H(t)v\dot{v}=H(t)v if and only if w(t):=eMtv(t)w(t):=e^{Mt}v(t) solves w˙=H~(t)w\dot{w}=\tilde{H}(t)w. We will now prove that for every t>0t>0 and every 𝐢\mathbf{i}, w(t)(𝐢)>0w(t)(\mathbf{i})>0. Then, obviously, v(t)(𝐢)>0v(t)(\mathbf{i})>0 as well, which then completes the proof of Proposition 3.1.

To prove the claim on w(t)w(t), we solve the initial value problem for w˙=H~(t)w\dot{w}=\tilde{H}(t)w by Picard iteration, that is we write

w(t)=(n=0H~(n)(t))w(0),w(t)=\left(\sum_{n=0}^{\infty}\tilde{H}^{(n)}(t)\right)w(0),

where the H~(n)(t)\tilde{H}^{(n)}(t) are defined inductively by

H~(0)(t)=id and H~(n)(t):=0tH~(t~)H~(n1)(t~)𝑑t~ for n1.\tilde{H}^{(0)}(t)=id\quad\text{ and }\quad\tilde{H}^{(n)}(t):=\int_{0}^{t}\tilde{H}(\tilde{t})\circ\tilde{H}^{(n-1)}(\tilde{t})d\tilde{t}\quad\text{ for }n\geq 1.

Observe that the positivity of H~(t)\tilde{H}(t) implies that the H~(n)(t)\tilde{H}^{(n)}(t) are positive as well. Because w(0)=v(0)0w(0)=v(0)\geq 0 and v(0)0v(0)\neq 0, we can therefore estimate, for any 𝐢,𝐤d\mathbf{i},\mathbf{k}\in\mathbb{Z}^{d} with 𝐢𝐤=1\left\lVert\mathbf{i}-\mathbf{k}\right\rVert=1,

w(t)(𝐢)\displaystyle w(t)(\mathbf{i}) =(n=0H~(n)(t)w(0))(𝐢)(0tH~(t¯)w(0)dt¯)(𝐢)\displaystyle=\left(\sum_{n=0}^{\infty}\tilde{H}^{(n)}(t)w(0)\right)(\mathbf{i})\geq\left(\int_{0}^{t}\tilde{H}(\bar{t})w(0)\mathrm{d}\bar{t}\right)(\mathbf{i}) (A.1)
(0t01𝐢,𝐤S𝐢[Φt¯0(u1)+t~(Φt¯0(u2)Φt¯0(u1))+v0]dt~dt¯)w(0)(𝐤).\displaystyle\geq\left(\int_{0}^{t}\int_{0}^{1}-\partial_{\mathbf{i},\mathbf{k}}S_{\mathbf{i}}\left[\Phi^{0}_{\bar{t}}(u_{1})+\tilde{t}(\Phi^{0}_{\bar{t}}(u_{2})-\Phi^{0}_{\bar{t}}(u_{1}))+v_{0}\right]\mathrm{d}\tilde{t}\mathrm{d}\bar{t}\right)w(0)(\mathbf{k}).

Now choose a 𝐤d\mathbf{k}\in\mathbb{Z}^{d} such that w(0)(𝐤)>0w(0)(\mathbf{k})>0 and recall that 𝐢,𝐤S𝐢<0\partial_{\mathbf{i},\mathbf{k}}S_{\mathbf{i}}<0. Then from (A.1) it follows that if 𝐢𝐤=1\left\lVert\mathbf{i}-\mathbf{k}\right\rVert=1, then for all t>0t>0, w(t)(𝐢)>0w(t)(\mathbf{i})>0.

To generalize to the case that 𝐢𝐤1\left\lVert\mathbf{i}-\mathbf{k}\right\rVert\neq 1, let us choose a sequence of lattice points 𝐣(0)=𝐤,,𝐣(N)=𝐢\mathbf{j}(0)=\mathbf{k},\cdots,\mathbf{j}(N)=\mathbf{i} such that 𝐣(n)𝐣(n1)=1\left\|\mathbf{j}(n)-\mathbf{j}(n-1)\right\|=1 and N=𝐢𝐤N=\|\mathbf{i}-\mathbf{k}\|. Then, by induction, w(ntN)(𝐣(n))>0w\left(\frac{nt}{N}\right)(\mathbf{j}(n))>0 for any n{0,,N}n\in\{0,\cdots,N\}. Thus, if w(0)(𝐤)>0w(0)(\mathbf{k})>0 and t>0t>0, then w(t)(𝐢)>0w(t)(\mathbf{i})>0. ∎


Next we prove Lemma 3.4.

Proof of Lemma 3.4. (i). Suppose (I0𝐩)(u)=0(I^{\mathbf{p}}_{0})^{\prime}(u)=0. Then u+v0u+v_{0} is a solution of (1.1), thus 𝐣:𝐣𝐢r𝐢S𝐣(u+v0)=0\sum_{\mathbf{j}:\left\lVert\mathbf{j}-\mathbf{i}\right\rVert\leq r}\partial_{\mathbf{i}}S_{\mathbf{j}}(u+v_{0})=0 for all 𝐢n\mathbf{i}\in\mathbb{Z}^{n}. So uu is a solution of the initial problem:

{tΦt0(u)(𝐢)=𝐣:𝐣𝐢r𝐢S𝐣(Φt0(u)+v0),for t>0,Φ00(u)(𝐢)=u(𝐢).\left\{\begin{array}[]{ll}-\partial_{t}\Phi_{t}^{0}(u)(\mathbf{i})&=\sum_{\mathbf{j}:\left\lVert\mathbf{j}-\mathbf{i}\right\rVert\leq r}\partial_{\mathbf{i}}S_{\mathbf{j}}(\Phi_{t}^{0}(u)+v_{0}),\quad\quad\textrm{for }t>0,\\ \Phi_{0}^{0}(u)(\mathbf{i})&=u(\mathbf{i}).\end{array}\right.

By the uniqueness of the solution of the above initial problem, Φt0(u)=u\Phi_{t}^{0}(u)=u.

(ii). By (3.1),

ddtI0𝐩(Φt0(u))\displaystyle\frac{\mathrm{d}}{\mathrm{d}t}I_{0}^{\mathbf{p}}\left(\Phi_{t}^{0}(u)\right) (A.2)
=\displaystyle= 𝐣𝕋0𝐩𝐤:𝐤𝐣r𝐤S𝐣(Φt0(u)+v0)tΦt0(u)(𝐤)\displaystyle\sum_{\mathbf{j}\in\mathbb{T}^{\mathbf{p}}_{0}}\sum_{\mathbf{k}:\left\lVert\mathbf{k}-\mathbf{j}\right\rVert\leq r}\partial_{\mathbf{k}}S_{\mathbf{j}}(\Phi_{t}^{0}(u)+v_{0})\cdot\partial_{t}\Phi_{t}^{0}(u)(\mathbf{k})
=\displaystyle= 𝐣𝕋0𝐩𝐤:𝐤𝐣r𝐤S𝐣(Φt0(u)+v0)[𝐥:𝐥𝐤r𝐤S𝐥(Φt0(u)+v0)]\displaystyle\sum_{\mathbf{j}\in\mathbb{T}^{\mathbf{p}}_{0}}\sum_{\mathbf{k}:\left\lVert\mathbf{k}-\mathbf{j}\right\rVert\leq r}\partial_{\mathbf{k}}S_{\mathbf{j}}(\Phi_{t}^{0}(u)+v_{0})\cdot\Big{[}-\sum_{\mathbf{l}:\left\lVert\mathbf{l}-\mathbf{k}\right\rVert\leq r}\partial_{\mathbf{k}}S_{\mathbf{l}}(\Phi_{t}^{0}(u)+v_{0})\Big{]}
=\displaystyle= 𝐣𝕋0𝐩[𝐤:𝐤𝐣r𝐣S𝐤(Φt0(u)+v0)]2\displaystyle-\sum_{\mathbf{j}\in\mathbb{T}^{\mathbf{p}}_{0}}\Big{[}\sum_{\mathbf{k}:\left\lVert\mathbf{k}-\mathbf{j}\right\rVert\leq r}\partial_{\mathbf{j}}S_{\mathbf{k}}(\Phi_{t}^{0}(u)+v_{0})\Big{]}^{2}
\displaystyle\leq 0.\displaystyle 0.

Thus (ii) holds.

(iii). By (i), Φt0(0)=0\Phi_{t}^{0}(0)=0 and Φt0(w0v0)=w0v0\Phi_{t}^{0}(w_{0}-v_{0})=w_{0}-v_{0}. Thus by Proposition 3.1, 0=Φt0(0)Φt0(u)Φt0(w0v0)=w0v00=\Phi_{t}^{0}(0)\leq\Phi_{t}^{0}(u)\leq\Phi_{t}^{0}(w_{0}-v_{0})=w_{0}-v_{0} for any u𝒢0𝐩u\in\mathcal{G}^{\mathbf{p}}_{0}. So (iii) follows.

(iv). By (A.2) and Lemma 2.6, I0𝐩(Φt0(u))I_{0}^{\mathbf{p}}(\Phi_{t}^{0}(u)) is non-increasing with a lower bound c0𝐩c_{0}^{\mathbf{p}}. So I0𝐩(Φt0(u))I_{0}^{\mathbf{p}}\left(\Phi_{t}^{0}(u)\right) has a limit as tt\to\infty. Since

I0𝐩(u)I0𝐩(Φt0(u))=0t𝐣𝕋0𝐩[𝐤:𝐤𝐣r𝐣S𝐤(Φt~0(u)+v0)]2dt~,I_{0}^{\mathbf{p}}(u)-I_{0}^{\mathbf{p}}\left(\Phi_{t}^{0}(u)\right)=\int_{0}^{t}\sum_{\mathbf{j}\in\mathbb{T}^{\mathbf{p}}_{0}}\Big{[}\sum_{\mathbf{k}:\left\lVert\mathbf{k}-\mathbf{j}\right\rVert\leq r}\partial_{\mathbf{j}}S_{\mathbf{k}}(\Phi^{0}_{\tilde{t}}(u)+v_{0})\Big{]}^{2}\mathrm{d}\tilde{t},

we have

𝐣𝕋0𝐩[𝐤:𝐤𝐣r𝐣S𝐤(Φti0(u)+v0)]20\sum_{\mathbf{j}\in\mathbb{T}^{\mathbf{p}}_{0}}\Big{[}\sum_{\mathbf{k}:\left\lVert\mathbf{k}-\mathbf{j}\right\rVert\leq r}\partial_{\mathbf{j}}S_{\mathbf{k}}(\Phi^{0}_{t_{i}}(u)+v_{0})\Big{]}^{2}\to 0

for some sequence tit_{i}\to\infty. Since Φti0(u)𝒢0𝐩\Phi^{0}_{t_{i}}(u)\in\mathcal{G}^{\mathbf{p}}_{0}, a compact set by Proposition 3.2, there is a U𝒢0𝐩U\in\mathcal{G}^{\mathbf{p}}_{0} such that Φti0(u)U\Phi^{0}_{t_{i}}(u)\to U in Λ0𝐩\Lambda_{0}^{\mathbf{p}} along a subsequence of tit_{i}. Note that 𝐣S𝐤\partial_{\mathbf{j}}S_{\mathbf{k}} is continuous, thus

𝐣𝕋0𝐩[𝐤:𝐤𝐣r𝐣S𝐤(U+v0)]2=0,\sum_{\mathbf{j}\in\mathbb{T}^{\mathbf{p}}_{0}}\Big{[}\sum_{\mathbf{k}:\left\lVert\mathbf{k}-\mathbf{j}\right\rVert\leq r}\partial_{\mathbf{j}}S_{\mathbf{k}}(U+v_{0})\Big{]}^{2}=0,

which implies

𝐤:𝐤𝐢r𝐢S𝐤(U+v0)=0,for 𝐢n,\sum_{\mathbf{k}:\left\lVert\mathbf{k}-\mathbf{i}\right\rVert\leq r}\partial_{\mathbf{i}}S_{\mathbf{k}}(U+v_{0})=0,\quad\textrm{for }\mathbf{i}\in\mathbb{Z}^{n},

i.e., U+v0U+v_{0} is a solution of (1.1).

(v). Firstly we claim that: there exists a constant ϵ>0\epsilon>0 such that if Kc=K_{c}=\emptyset, then

(I0𝐩)(u)(Λ0𝐩)2ϵ,for u(I0𝐩)cϵc+ϵ.\left\|(I_{0}^{\mathbf{p}})^{\prime}(u)\right\|_{(\Lambda_{0}^{\mathbf{p}})^{\prime}}\geq\sqrt{2\epsilon},\quad\textrm{for }u\in(I_{0}^{\mathbf{p}})_{c-\epsilon}^{c+\epsilon}. (A.3)

Here (Λ0𝐩)(\Lambda_{0}^{\mathbf{p}})^{\prime} is the dual space of the Banach space Λ0𝐩\Lambda_{0}^{\mathbf{p}} and (I0𝐩)t2t1:={u𝒢0𝐩|t2I0𝐩(u)t1}(I_{0}^{\mathbf{p}})^{t_{1}}_{t_{2}}:=\{u\in\mathcal{G}_{0}^{\mathbf{p}}\,|\,t_{2}\leq I_{0}^{\mathbf{p}}(u)\leq t_{1}\}. Indeed, if the claim fails then for any kk\in\mathbb{N}, there are (uk)(I0𝐩)c1/kc+1/k(u_{k})\subset(I_{0}^{\mathbf{p}})_{c-1/k}^{c+1/k} satisfying

(I0𝐩)(uk)(Λ0𝐩)0\left\|(I_{0}^{\mathbf{p}})^{\prime}(u_{k})\right\|_{(\Lambda_{0}^{\mathbf{p}})^{\prime}}\to 0

as kk\to\infty. But since (uk)𝒢0𝐩(u_{k})\subset\mathcal{G}^{\mathbf{p}}_{0}, a compact set by Proposition 3.2, we may assume that ukU𝒢0𝐩u_{k}\to U\in\mathcal{G}^{\mathbf{p}}_{0} as kk\to\infty and I0𝐩(U)=cI_{0}^{\mathbf{p}}(U)=c. Thus UKcU\in K_{c}, which contradicts Kc=K_{c}=\emptyset. So (A.3) holds.

Now we prove (v). Suppose, by contradiction, that there is a u(I0𝐩)cϵc+ϵu\in(I_{0}^{\mathbf{p}})_{c-\epsilon}^{c+\epsilon} such that Φ10(u)(I0𝐩)cϵ\Phi_{1}^{0}(u)\not\in(I_{0}^{\mathbf{p}})^{c-\epsilon}, then cϵ<I0𝐩(Φt0(u))c+ϵc-\epsilon<I_{0}^{\mathbf{p}}(\Phi_{t}^{0}(u))\leq c+\epsilon for all t[0,1]t\in[0,1]. Since

(I0𝐩)(u)(Λ0𝐩)=supvΛ0𝐩1|(I0𝐩)(u)v|=supvΛ0𝐩1|𝐣𝕋0𝐩𝐤:𝐤𝐣r𝐤S𝐣(u+v0)v(𝐤)|supvΛ0𝐩1𝐣𝕋0𝐩|v(𝐣)𝐤:𝐤𝐣r𝐣S𝐤(u+v0)|supvΛ0𝐩1(𝐣𝕋0𝐩|v(𝐣)|2)12(𝐣𝕋0𝐩|𝐤:𝐤𝐣r𝐣S𝐤(u+v0)|2)12=(𝐣𝕋0𝐩|𝐤:𝐤𝐣r𝐣S𝐤(u+v0)|2)12,\begin{split}&\left\|(I_{0}^{\mathbf{p}})^{\prime}(u)\right\|_{(\Lambda_{0}^{\mathbf{p}})^{\prime}}\\ =&\sup_{\left\lVert v\right\rVert_{\Lambda_{0}^{\mathbf{p}}}\leq 1}|(I_{0}^{\mathbf{p}})^{\prime}(u)v|\\ =&\sup_{\left\lVert v\right\rVert_{\Lambda_{0}^{\mathbf{p}}}\leq 1}\Big{|}\sum_{\mathbf{j}\in\mathbb{T}^{\mathbf{p}}_{0}}\sum_{\mathbf{k}:\left\lVert\mathbf{k}-\mathbf{j}\right\rVert\leq r}\partial_{\mathbf{k}}S_{\mathbf{j}}(u+v_{0})v(\mathbf{k})\Big{|}\\ \leq&\sup_{\left\lVert v\right\rVert_{\Lambda_{0}^{\mathbf{p}}}\leq 1}\sum_{\mathbf{j}\in\mathbb{T}^{\mathbf{p}}_{0}}\Big{|}v(\mathbf{j})\cdot\sum_{\mathbf{k}:\left\lVert\mathbf{k}-\mathbf{j}\right\rVert\leq r}\partial_{\mathbf{j}}S_{\mathbf{k}}(u+v_{0})\Big{|}\\ \leq&\sup_{\left\lVert v\right\rVert_{\Lambda_{0}^{\mathbf{p}}}\leq 1}\left(\sum_{\mathbf{j}\in\mathbb{T}^{\mathbf{p}}_{0}}\left|v(\mathbf{j})\right|^{2}\right)^{\frac{1}{2}}\left(\sum_{\mathbf{j}\in\mathbb{T}^{\mathbf{p}}_{0}}\left|\sum_{\mathbf{k}:\left\lVert\mathbf{k}-\mathbf{j}\right\rVert\leq r}\partial_{\mathbf{j}}S_{\mathbf{k}}(u+v_{0})\right|^{2}\right)^{\frac{1}{2}}\\ =&\left(\sum_{\mathbf{j}\in\mathbb{T}^{\mathbf{p}}_{0}}\left|\sum_{\mathbf{k}:\left\lVert\mathbf{k}-\mathbf{j}\right\rVert\leq r}\partial_{\mathbf{j}}S_{\mathbf{k}}(u+v_{0})\right|^{2}\right)^{\frac{1}{2}},\end{split} (A.4)

we have

I0𝐩(Φ10(u))=I0𝐩(u)+01ddtI0𝐩(Φt0(u))dt=I0𝐩(u)+01(I0𝐩)(Φt0(u))tΦt0(u)dt=I0𝐩(u)01𝐣𝕋0𝐩[𝐤:𝐤𝐣r𝐣S𝐤(Φt0(u)+v0)]2dt(c+ϵ)2ϵ=cϵ,\begin{split}&I_{0}^{\mathbf{p}}(\Phi_{1}^{0}(u))\\ =&I_{0}^{\mathbf{p}}(u)+\int_{0}^{1}\frac{\mathrm{d}}{\mathrm{d}t}I_{0}^{\mathbf{p}}(\Phi_{t}^{0}(u))\mathrm{d}t\\ =&I_{0}^{\mathbf{p}}(u)+\int_{0}^{1}(I_{0}^{\mathbf{p}})^{\prime}(\Phi_{t}^{0}(u))\partial_{t}\Phi_{t}^{0}(u)\mathrm{d}t\\ =&I_{0}^{\mathbf{p}}(u)-\int_{0}^{1}\sum_{\mathbf{j}\in\mathbb{T}^{\mathbf{p}}_{0}}\Big{[}\sum_{\mathbf{k}:\left\lVert\mathbf{k}-\mathbf{j}\right\rVert\leq r}\partial_{\mathbf{j}}S_{\mathbf{k}}(\Phi_{t}^{0}(u)+v_{0})\Big{]}^{2}\mathrm{d}t\\ \leq&(c+\epsilon)-2\epsilon\\ =&c-\epsilon,\end{split}

contrary to the existence of uu. Here the third equality follows from (A.2) and the first inequality follows from (A.3)-(A.4). ∎


Now a heat flow method is used to give the following (cf. [5, Proposition 2.12]):

Another proof of Theorem 3.6. For ϵ>0\epsilon>0, let hC([0,1],0𝐩(I0𝐩)d0𝐩+ϵ)h\in C([0,1],\mathcal{H}_{0}^{\mathbf{p}}\cap(I_{0}^{\mathbf{p}})^{d_{0}^{\mathbf{p}}+\epsilon}). Set ht:=Φt0hh_{t}:=\Phi_{t}^{0}\circ h, then for any t0t\geq 0, htC([0,1],0𝐩(I0𝐩)d0𝐩+ϵ)h_{t}\in C([0,1],\mathcal{H}_{0}^{\mathbf{p}}\cap(I_{0}^{\mathbf{p}})^{d_{0}^{\mathbf{p}}+\epsilon}). We claim that:

there is a θ[0,1]\theta_{\infty}\in[0,1] satisfying I0𝐩(ht(θ))d0𝐩I_{0}^{\mathbf{p}}\left(h_{t}\left(\theta_{\infty}\right)\right)\geq d_{0}^{\mathbf{p}} for all t0t\geq 0.

Indeed, for any t0t\geq 0 there exists θt[0,1]\theta_{t}\in[0,1] such that I0𝐩(ht(θt))d0𝐩I_{0}^{\mathbf{p}}\left(h_{t}\left(\theta_{t}\right)\right)\geq d_{0}^{\mathbf{p}}. We can extract a subsequence of θt\theta_{t}, say θtk\theta_{t_{k}} converging to some θ[0,1]\theta_{\infty}\in[0,1] as tkt_{k}\to\infty. If I0𝐩(hτ(θ))<d0𝐩I_{0}^{\mathbf{p}}\left(h_{\tau}\left(\theta_{\infty}\right)\right)<d_{0}^{\mathbf{p}} for some τ>0\tau>0, then I0𝐩(hτ(θtk))<d0𝐩I_{0}^{\mathbf{p}}\left(h_{\tau}\left(\theta_{t_{k}}\right)\right)<d_{0}^{\mathbf{p}} for large tkt_{k}. Then enlarging tkt_{k} if necessary such that tk>τt_{k}>\tau, by Lemma 3.4 (ii), I0𝐩(htk(θtk))I0𝐩(hτ(θtk))<d0𝐩I_{0}^{\mathbf{p}}\left(h_{t_{k}}\left(\theta_{t_{k}}\right)\right)\leq I_{0}^{\mathbf{p}}\left(h_{\tau}\left(\theta_{t_{k}}\right)\right)<d_{0}^{\mathbf{p}}, which is a contradiction. Thus the claim holds.

By Lemma 3.4 (iv), there is a subsequence of ht(θ)h_{t}(\theta_{\infty}) converging to some vϵ𝒢0𝐩v_{\epsilon}\in\mathcal{G}_{0}^{\mathbf{p}} such that vϵ+v0v_{\epsilon}+v_{0} is a solution of (1.1) and

d0𝐩+ϵI0𝐩(vϵ)=limtI0𝐩(ht(θ))d0𝐩.d_{0}^{\mathbf{p}}+\epsilon\geq I_{0}^{\mathbf{p}}\left(v_{\epsilon}\right)=\lim_{t\rightarrow\infty}I_{0}^{\mathbf{p}}\left(h_{t}\left(\theta_{\infty}\right)\right)\geq d_{0}^{\mathbf{p}}.

By Proposition 3.2, letting ϵ0\epsilon\to 0 (up to a subsequence) completes the proof of Theorem 3.6. ∎

Acknowledgments

The author wishes to express his gratitude to Professor Zhi-Qiang Wang (Utah State University) for helpful discussion and for giving me a lot of encouragement. The author is supported by the Fundamental Research Funds for the Central Universities (no. 34000-31610274).

References

  • [1] S. Aubry, P. Y. Le Daeron, The discrete Frenkel-Kontorova model and its extensions. I. Exact results for the ground-states. Phys. D 8 (1983), no. 3, 381-422.
  • [2] V. Bangert, On minimal laminations of the torus. Ann. Inst. H. Poincaré Anal. Non Linéaire. 6 (1989), no. 2, 95-138.
  • [3] S. Bolotin, P. H. Rabinowitz, A note on heteroclinic solutions of mountain pass type for a class of nonlinear elliptic PDE’s. Contributions to nonlinear analysis, 105-114, Progr. Nonlinear Differential Equations Appl., 66, Birkhäuser, Basel, 2006.
  • [4] S. Bolotin, P. H. Rabinowitz, On the multiplicity of periodic solutions of mountain pass type for a class of semilinear PDE’s. J. Fixed Point Theory Appl. 2 (2007), no. 2, 313-331.
  • [5] S. Bolotin, P. H. Rabinowitz, Hybrid mountain pass homoclinic solutions of a class of semilinear elliptic PDEs. Ann. Inst. H. Poincaré Anal. Non Linéaire 31 (2014), no. 1, 103-128.
  • [6] O. M. Braun, Y. S. Kivshar, The Frenkel-Kontorova model. Concepts, methods, and applications. Texts and Monographs in Physics. Springer-Verlag, Berlin, 2004.
  • [7] H. Brezis, Functional analysis, Sobolev spaces and partial differential equations. Universitext. Springer, New York, 2011.
  • [8] R. de la Llave, E. Valdinoci, Critical points inside the gaps of ground state laminations for some models in statistical mechanics. J. Stat. Phys. 129 (2007), no. 1, 81-119.
  • [9] R. de la Llave, E. Valdinoci, Ground states and critical points for generalized Frenkel-Kontorova models in d\mathbb{Z}^{d}. Nonlinearity 20 (2007), no. 10, 2409-2424.
  • [10] J. Frenkel, T. Kontorova, On the theory of plastic deformation and twinning. Acad. Sci. U.S.S.R. J. Phys. 1 (1939), 137-149.
  • [11] Wen-Long Li, Xiaojun Cui, Heteroclinic solutions for a Frenkel-Kontorova model by minimization methods of Rabinowitz and Stredulinsky. J. Differential Equations 268 (2020), no. 3, 1106-1155.
  • [12] Wen-Long Li, Xiaojun Cui, Multitransition solutions for a generalized Frenkel-Kontorova model. arXiv:2004.04868.
  • [13] J. Mather, Existence of quasi-periodic orbits for twist homeomorphisms of the annulus. Topology 21 (1982), no. 4, 457-467.
  • [14] Xue-Qing Miao, Wen-Xin Qin, Ya-Nan Wang, Secondary invariants of Birkhoff minimizers and heteroclinic orbits. Journal of Differential Equations 260 (2016), no. 2, 1522-1557.
  • [15] J. Moser, Minimal solutions of variations problems on a torus. Ann. Inst. H. Poincaré Anal. Non Linéaire 3 (1986), 229-272.
  • [16] B. Mramor, B. Rink, Ghost circles in lattice Aubry-Mather theory. J. Differential Equations 252 (2012), no. 4, 3163-3208.
  • [17] P. H. Rabinowitz, E. W. Stredulinsky, Extensions of Moser-Bangert Theory: Locally Minimal Solutions. Progress in Nonlinear Differential Equations and Their Applications 81, Birkha¨\ddot{a}user Basel, 2011.