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

Generic nondegeneracy for solutions of the Allen-Cahn equation under a volume constraint in closed manifolds

Gustavo de Paula Ramos Instituto de Matemática e Estatística, Universidade de São Paulo, Rua do Matão, 1010, 05508-090, São Paulo, SP, Brazil. [email protected] http://www.ime.usp.br/ gpramos
Abstract.

Let MnM^{n} be a connected closed smooth manifold with n2n\geq 2. We adapt the techniques in [MP09] and [GM11] to prove the generic nondegeneracy for solutions of the Van der Waals-Allen-Cahn-Hilliard equation under a volume constraint in MM.

Keywords. nondegenerate critical points, Allen-Cahn equation, generic result.

2010 Mathematics Subject Classification. 58E05, 35J20.

1. Introduction and main result

Let (Mn,g)(M^{n},g) be a connected closed smooth Riemannian manifold, where n2n\geq 2. Let W:W\colon\mathbb{R}\to\mathbb{R} be a function of class C2C^{2}. Fix ν,ϵ>0\nu,\epsilon>0. A pair (u,λ)Hg(M)×(u,\lambda)\in H_{g}(M)\times\mathbb{R} is a solution for the Van der Waals-Allen-Cahn Hilliard equation under volume constraint ν\nu when

(PW,ν,ϵ,gP_{W,\nu,\epsilon,g}) {ϵ2Δgu+W(u)=λMudμg=ν,\begin{cases}-\epsilon^{2}\Delta_{g}u+W^{\prime}(u)=\lambda\\ \int_{M}u\mathop{}\!\mathrm{d}\mu_{g}=\nu\end{cases},

where μg\mu_{g} is a measure induced by gg defined on the Borel subsets of MM and Hg(M)H_{g}(M) is a convenient Sobolev space of functions defined in section 2.

In [BNAP20], the authors establish lower bounds on the number of solutions for (PW,ν,ϵ,g)\eqref{eqn:nonlinear-problem} in function of topological invariants of MM for sufficiently small ν,ϵ>0\nu,\epsilon>0 and under specific hypotheses on the potential function WW. In particular: if (PW,ν,ϵ,g)\eqref{eqn:nonlinear-problem} only admits nondegenerate solutions, then Morse theory may be applied to prove that it admits at least PM(1)P_{M}(1) solutions, where PM(t)P_{M}(t) is the Poincaré polynomial of MM.

Our main result is that under suitable growth conditions for WW^{\prime} and W′′W^{\prime\prime}, this is indeed the case generically with respect to (ϵ,g)]0,[×k(\epsilon,g)\in]0,\infty[\times\mathcal{M}^{k}, where 1k<1\leq k<\infty and k\mathcal{M}^{k} is the space of Riemannian metrics of class CkC^{k} on MM:

Theorem 1.1.

Fix g0kg_{0}\in\mathcal{M}^{k}. Suppose that (1) and (2) hold. Then

𝒟W,ν={(ϵ,g)]0,[×k: any solution (u,λ)Hg0(M)× for (PW,ν,ϵ,g) is nondegenerate}\mathcal{D}^{*}_{W,\nu}=\left\{(\epsilon,g)\in]0,\infty[\times\mathcal{M}^{k}\mathrel{\mathop{\mathchar 58\relax}}\text{ any solution }(u,\lambda)\in H_{g_{0}}(M)\times\mathbb{R}\right.\\ \left.\text{ for }\eqref{eqn:nonlinear-problem}\text{ is nondegenerate}\right\}

is an open dense subset of ]0,[×k]0,\infty[\times\mathcal{M}^{k}.

This result is obtained by the application of an abstract transversality theorem through an appropriate adaptation of the techniques in [MP09] and [GM11] to the context of this article.

More precisely, we say that a solution (u,λ)Hg(M)×(u,\lambda)\in H_{g}(M)\times\mathbb{R} for (PW,ν,ϵ,g)\eqref{eqn:nonlinear-problem} is nondegenerate when the only pair (v,Λ)Hg(M)×(v,\Lambda)\in H_{g}(M)\times\mathbb{R} which solves the linearized problem

(QW,ϵ,g,uQ_{W,\epsilon,g,u}) {ϵ2Δgv+W′′(u)v=ΛMvdμg=0\begin{cases}-\epsilon^{2}\Delta_{g}v+W^{\prime\prime}(u)v=\Lambda\\ \int_{M}v\mathop{}\!\mathrm{d}\mu_{g}=0\end{cases}

is the trivial one (v,Λ)=(0,0)(v,\Lambda)=(0,0).

In fact, this notion coincides with the Morse theoretic notion of a nondegenerate critical point for the functional JW,ϵ,g:Hg(M)×J_{W,\epsilon,g}\colon H_{g}(M)\times\mathbb{R}\to\mathbb{R} given by

JW,ϵ,g(u,λ)=Mϵ22g(u,u)+W(u)λudμgλν.J_{W,\epsilon,g}(u,\lambda)=\int_{M}\frac{\epsilon^{2}}{2}g(\nabla u,\nabla u)+W(u)-\lambda u\mathop{}\!\mathrm{d}\mu_{g}-\lambda\nu.

Indeed, JW,ϵ,gJ_{W,\epsilon,g} is a functional of class C2C^{2} for which (v,Λ)(v,\Lambda) is a solution for (QW,ϵ,g,u)\eqref{eqn:linearized-problem} if, and only if, Mvdμg=0\int_{M}v\mathop{}\!\mathrm{d}\mu_{g}=0 and (v,Λ)kerHess(JW,ϵ,g)(u,λ)(v,\Lambda)\in\ker\mathrm{Hess}(J_{W,\epsilon,g})_{(u,\lambda)}. Therefore, (u,λ)(u,\lambda) is a nondegenerate solution for (PW,ν,ϵ,g)\eqref{eqn:nonlinear-problem} precisely when (u,λ)(u,\lambda) is a nondegenerate critical point of JW,ϵ,gJ_{W,\epsilon,g} such that Mudμg=ν\int_{M}u\mathop{}\!\mathrm{d}\mu_{g}=\nu.

For Differential Geometry, interest for the Van der Waals-Allen-Cahn-Hilliard equation under a volume constraint is justified by the results of [PR03], where Pacard and Ritoré showed that one can approach constant mean curvature hypersurfaces by the nodal sets of critical points for JW,ϵ,g,λJ_{W,\epsilon,g,\lambda} as ϵ0+\epsilon\to 0^{+}. If we consider critical points without the volume constraint, these sets approach a minimal hypersurface.

Acknowledgement

The author thanks Paolo Piccione for suggesting the topic and discussing drafts of this article.

2. Preliminaries

Basic constructions

Fix 1k<1\leq k<\infty. Denote by 𝒮k\mathcal{S}^{k} the Banach space of symmetric 2-covectors on MM of class CkC^{k}. The space k\mathcal{M}^{k} of Riemannian metrics on MM of class CkC^{k} is an open convex cone in 𝒮k\mathcal{S}^{k}.

Consider any (ϵ,g)]0,[×k(\epsilon,g)\in]0,\infty[\times\mathcal{M}^{k}. (ϵ,g)(\epsilon,g) induces the following inner products on C(M)C^{\infty}(M):

u,vg=Mg(u,v)+uvdμg;\left\langle u,v\right\rangle_{g}=\int_{M}g\left(\nabla u,\nabla v\right)+uv\mathop{}\!\mathrm{d}\mu_{g};
Eϵ,g(u,v):=Mϵ2g(u,v)+uvdμg.E_{\epsilon,g}(u,v)\mathrel{\mathop{\mathchar 58\relax}}=\int_{M}\epsilon^{2}g\left(\nabla u,\nabla v\right)+uv\mathop{}\!\mathrm{d}\mu_{g}.

Hg(M)H_{g}(M), Hϵ,g(M)H_{\epsilon,g}(M) are, respectively, the Hilbert spaces endowed with ,g\left\langle\cdot,\cdot\right\rangle_{g}, Eϵ,gE_{\epsilon,g} obtained as completions of C(M)C^{\infty}(M). Similarly: given 1q<1\leq q<\infty, Lgq(M)L^{q}_{g}(M) is the Banach space obtained as completion of C(M)C^{\infty}(M) with respect to the norm

uq,g:=(M|u|qdμg)1/q.\mathinner{\!\left\lVert u\right\rVert}_{q,g}\mathrel{\mathop{\mathchar 58\relax}}=\left(\int_{M}\mathinner{\!\left\lvert u\right\rvert}^{q}\mathop{}\!\mathrm{d}\mu_{g}\right)^{1/q}.

One may check that the norms induced by ,g\left\langle\cdot,\cdot\right\rangle_{g}, Eϵ,gE_{\epsilon,g} on C(M)C^{\infty}(M) are equivalent. In particular, this implies Hg(M)=Hϵ,g(M)H_{g}(M)=H_{\epsilon,g}(M) as sets and that the canonical inclusion Hg(M)Hϵ,g(M)H_{g}(M)\to H_{\epsilon,g}(M) is an isomorphism of Banach spaces. The same holds for the canonical inclusion Hg(M)Hg(M)H_{g^{\prime}}(M)\to H_{g}(M) for any gkg^{\prime}\in\mathcal{M}^{k}. For details, we refer the reader to [Heb00, Proposition 2.2].

Considered setting

Suppose that

(1) K1>0t,|W(t)|K1(1+|t|p1);\exists K_{1}>0\ \forall t\in\mathbb{R},\ \mathinner{\!\left\lvert W^{\prime}(t)\right\rvert}\leq K_{1}(1+\mathinner{\!\left\lvert t\right\rvert}^{p-1});
(2) K2>0t,|W′′(t)|K2(1+|t|p2);\exists K_{2}>0\ \forall t\in\mathbb{R},\ \mathinner{\!\left\lvert W^{\prime\prime}(t)\right\rvert}\leq K_{2}(1+\mathinner{\!\left\lvert t\right\rvert}^{p-2});

for a certain p]2,pn[p\in\left]2,p_{n}\right[, where pn=p_{n}=\infty for n=2n=2, pn=(2n)/(n2)p_{n}=(2n)/(n-2) for n3n\geq 3.

Fix g0kg_{0}\in\mathcal{M}^{k}. Consider any (ϵ,g)]0,[×k(\epsilon,g)\in]0,\infty[\times\mathcal{M}^{k}. Due to the Kondrakov theorem, the canonical inclusion iϵ,g:Hϵ,g(M)Lgp(M)i_{\epsilon,g}\colon H_{\epsilon,g}(M)\to L_{g}^{p}(M) is a compact operator. Set p:=p/(p1)p^{\prime}\mathrel{\mathop{\mathchar 58\relax}}=p/(p-1). We define Aϵ,gA_{\epsilon,g} as the adjoint of iϵ,gi_{\epsilon,g} while considering the canonical Banach space isomorphisms (Lgp(M))Lgp(M)(L^{p}_{g}(M))^{\prime}\simeq L^{p^{\prime}}_{g}(M) and Hϵ,g(M)(Hϵ,g(M))H_{\epsilon,g}(M)\simeq(H_{\epsilon,g}(M))^{\prime}:

Definition 2.1.

Aϵ,g=iϵ,g:Lgp(M)Hϵ,g(M).A_{\epsilon,g}=i_{\epsilon,g}^{*}\colon L_{g}^{p^{\prime}}(M)\to H_{\epsilon,g}(M).

Remark 2.2.

Aϵ,gA_{\epsilon,g} is a compact self-adjoint operator and Eϵ,g(Aϵ,gu,v)=MuvdμgE_{\epsilon,g}\left(A_{\epsilon,g}u,v\right)=\int_{M}uv\mathop{}\!\mathrm{d}\mu_{g} for any u,vHg(M)u,v\in H_{g}(M).

For details on lemmas 2.3 and 2.4, we refer the reader to[MP09, Lemmas 2.1, 2.3].

Lemma 2.3.

E:]0,[×kBil(Hg0(M))E\colon]0,\infty[\times\mathcal{M}^{k}\to\mathrm{Bil}\left(H_{g_{0}}(M)\right) is a map of class C1C^{1}, where E(ϵ,g):=Eϵ,gE(\epsilon,g)\mathrel{\mathop{\mathchar 58\relax}}=E_{\epsilon,g}. In particular,

dE(ϵ,g)[η,h](u,v)=2ϵηMg(u,v)dμg+ϵ2Mbg,h(u,v)dμg++12M(trgh)uvdμg,\mathop{}\!\mathrm{d}E_{(\epsilon,g)}[\eta,h](u,v)=2\epsilon\eta\int_{M}g\left(\nabla u,\nabla v\right)\mathop{}\!\mathrm{d}\mu_{g}+\epsilon^{2}\int_{M}b_{g,h}\left(\nabla u,\nabla v\right)\mathop{}\!\mathrm{d}\mu_{g}+\\ +\frac{1}{2}\int_{M}\left(\mathrm{tr}_{g}h\right)uv\mathop{}\!\mathrm{d}\mu_{g},

where bg,hb_{g,h} is a symmetric 2-covector on MM of class CkC^{k} given locally by

(bg,h)ij=(trgh)gij/2giqhqlglj.\left(b_{g,h}\right)_{ij}=\left(\mathrm{tr}_{g}h\right)g^{ij}/2-g^{iq}h_{ql}g^{lj}.
Lemma 2.4.

A:]0,[×kB(Lg0p(M),Hg0(M))A\colon]0,\infty[\times\mathcal{M}^{k}\to B\left(L_{g_{0}}^{p^{\prime}}(M),H_{g_{0}}(M)\right) is a map of class C1C^{1}, where A(ϵ,g):=Aϵ,gA(\epsilon,g)\mathrel{\mathop{\mathchar 58\relax}}=A_{\epsilon,g}. In particular,

dE(ϵ,g)[η,h](Aϵ,gu,v)+Eϵ,g(dA(ϵ,g)[η,h]u,v)=12M(trgh)uvdμg.\mathop{}\!\mathrm{d}E_{(\epsilon,g)}[\eta,h]\left(A_{\epsilon,g}u,v\right)+E_{\epsilon,g}\left(\mathop{}\!\mathrm{d}A_{(\epsilon,g)}[\eta,h]u,v\right)=\frac{1}{2}\int_{M}\left(\mathrm{tr}_{g}h\right)uv\mathop{}\!\mathrm{d}\mu_{g}.

W:W^{\prime}\colon\mathbb{R}\to\mathbb{R} is a function of class C1C^{1} with suitable growth conditions, so Hg0(M)uW(u)Lg0p(M)H_{g_{0}}(M)\ni u\mapsto W^{\prime}(u)\in L_{g_{0}}^{p^{\prime}}(M) is a Nemytskii operator of class C1C^{1}. For details on this argument, we recommend the reference [Kav93]. This implies:

Lemma 2.5.

The Nemytskii operator BW:Hg0(M)×Lg0p(M)B_{W}\colon H_{g_{0}}(M)\times\mathbb{R}\to L_{g_{0}}^{p^{\prime}}(M) given by BW(u,λ)=λ+uW(u)B_{W}(u,\lambda)=\lambda+u-W^{\prime}(u) is a map of class C1C^{1}. In particular,

d(BW)(u,λ)[v,Λ]=Λ+vvW′′(u).\mathop{}\!\mathrm{d}\left(B_{W}\right)_{(u,\lambda)}[v,\Lambda]=\Lambda+v-vW^{\prime\prime}(u).

In the next definition, we identify the space of constant real-valued functions on MM with \mathbb{R}:

Definition 2.6.

Let FW:]0,[×k×(Hg0(M))×Hg0(M)×F_{W}\colon]0,\infty[\times\mathcal{M}^{k}\times(H_{g_{0}}(M)\setminus\mathbb{R})\times\mathbb{R}\to H_{g_{0}}(M)\times\mathbb{R} be given by

FW(ϵ,g,u,λ)=(uAϵ,gBW(u,λ),Mudμg).F_{W}\left(\epsilon,g,u,\lambda\right)=\left(u-A_{\epsilon,g}\circ B_{W}(u,\lambda),\int_{M}u\mathop{}\!\mathrm{d}\mu_{g}\right).

Using remark 2.2, we can prove that the set of solutions (u,λ)(Hg0(M))×(u,\lambda)\in(H_{g_{0}}(M)\setminus\mathbb{R})\times\mathbb{R} for (PW,ν,ϵ,g)\eqref{eqn:nonlinear-problem} is a level-set of FWF_{W}:

Remark 2.7.

(u,λ)Hg0(M)×(u,\lambda)\in H_{g_{0}}(M)\times\mathbb{R} is a solution for (PW,ν,ϵ,g)\eqref{eqn:nonlinear-problem} if, and only if, FW(ϵ,g,u,λ)=(0,ν)F_{W}(\epsilon,g,u,\lambda)=(0,\nu).

Lemma 2.8.

FW:]0,[×k×(Hg0(M))×Hg0(M)×F_{W}\colon]0,\infty[\times\mathcal{M}^{k}\times(H_{g_{0}}(M)\setminus\mathbb{R})\times\mathbb{R}\to H_{g_{0}}(M)\times\mathbb{R} is a map of class C1C^{1}. In particular,

d(FW)(ϵ,g,u,λ)[η,h,v,Λ]==(vAϵ,gd(BW)(u,λ)[v,Λ]dA(ϵ,g)[η,h]BW(u,λ),,M12(trgh)u+vdμg)\mathop{}\!\mathrm{d}\left(F_{W}\right)_{\left(\epsilon,g,u,\lambda\right)}[\eta,h,v,\Lambda]=\\ =\left(v-A_{\epsilon,g}\circ\mathop{}\!\mathrm{d}\left(B_{W}\right)_{(u,\lambda)}[v,\Lambda]-\mathop{}\!\mathrm{d}A_{(\epsilon,g)}[\eta,h]\circ B_{W}(u,\lambda),\right.\\ \left.,\int_{M}\frac{1}{2}\left(\mathrm{tr}_{g}h\right)u+v\mathop{}\!\mathrm{d}\mu_{g}\right)

3. Proof of main result

Consider the following abstract transversality theorem:

Theorem 3.1.

[HHL05, Theorem 5.4] Let X,Y,ZX,Y,Z be real Banach spaces and U,VU,V be respective open subsets of X,YX,Y. Let F:V×UZF\colon V\times U\to Z be a map of class CmC^{m}, where m1m\geq 1. Let z0imFz_{0}\in\mathrm{im}\ F. Suppose that

  1. (1)

    Given yVy\in V, F(y,):xF(x,y)F(y,\cdot)\colon x\mapsto F(x,y) is a Fredholm map of index l<ml<m, i.e., dF(y,)x:XZ\mathop{}\!\mathrm{d}F(y,\cdot)_{x}\colon X\to Z is a Fredholm operator of index ll for any xUx\in U;

  2. (2)

    z0z_{0} is a regular value of FF, i.e., dF(y0,x0):Y×XZ\mathop{}\!\mathrm{d}F_{\left(y_{0},x_{0}\right)}\colon Y\times X\to Z is surjective for any (y0,x0)F1(z0)(y_{0},x_{0})\in F^{-1}(z_{0});

  3. (3)

    Let ι:F1(z0)Y×X\iota\colon F^{-1}(z_{0})\to Y\times X be the canonical embedding and πY:Y×XY\pi_{Y}\colon Y\times X\to Y be the projection of the first coordinate. Then πYι:F1(z0)Y\pi_{Y}\circ\iota\colon F^{-1}(z_{0})\to Y is σ\sigma-proper, i.e., F1(z0)=s=1CsF^{-1}(z_{0})=\bigcup_{s=1}^{\infty}C_{s}, where given s=1,2,s=1,2,..., CsC_{s} is a closed subset of F1(z0)F^{-1}(z_{0}) and πYι|Cs\left.\pi_{Y}\circ\iota\right|_{C_{s}} is proper.

Then the set {yV:z0 is a regular value of F(y,)}\{y\in V\mathrel{\mathop{\mathchar 58\relax}}z_{0}\text{ is a regular value of }F(y,\cdot)\} is an open dense subset of VV.

The first step to prove our main result is the lemma that follows, in which we restrict ourselves to nonconstant solutions:

Lemma 3.2.

Fix g0kg_{0}\in\mathcal{M}^{k}. Suppose that (1) and (2) hold. Then

𝒟W,ν={(ϵ,g)]0,[×k: any solution (u,λ)(Hg0(M))× for (PW,ν,ϵ,g) is nondegenerate}\mathcal{D}_{W,\nu}=\left\{(\epsilon,g)\in]0,\infty[\times\mathcal{M}^{k}\mathrel{\mathop{\mathchar 58\relax}}\text{ any solution }(u,\lambda)\in(H_{g_{0}}(M)\setminus\mathbb{R})\times\mathbb{R}\right.\\ \left.\text{ for }\eqref{eqn:nonlinear-problem}\text{ is nondegenerate}\right\}

is an open dense subset of ]0,[×k]0,\infty[\times\mathcal{M}^{k}.

Its proof consists of a direct application of the abstract transversality theorem. Specifically, we consider X=Z=Hg0(M)×X=Z=H_{g_{0}}(M)\times\mathbb{R}, Y=V=]0,[×𝒮kY=V=]0,\infty[\times\mathcal{S}^{k}, U=(Hg0(M))×U=(H_{g_{0}}(M)\setminus\mathbb{R})\times\mathbb{R}, F=FWF=F_{W} and z0=(0,ν)z_{0}=(0,\nu). We verify that its hypotheses hold in section 4.

After analysing the constant solutions for (PW,ν,ϵ,g)\eqref{eqn:nonlinear-problem}, we refine lemma 3.2 to prove our main result:

Proof of Theorem 1.1.

Let (ϵ,g)]0,[×k(\epsilon,g)\in]0,\infty[\times\mathcal{M}^{k} and 𝒰\mathcal{U} be a neighborhood of (ϵ,g)(\epsilon,g) in ]0,[×k]0,\infty[\times\mathcal{M}^{k}.

𝒟W,ν𝒰\mathcal{D}^{*}_{W,\nu}\cap\mathcal{U} is not empty. Indeed, let (ϵ¯,g¯)𝒟W,ν𝒰(\overline{\epsilon},\overline{g})\in\mathcal{D}_{W,\nu}\cap\mathcal{U}. If (Pν,ϵ¯,g¯,W)\left(P_{\nu,\overline{\epsilon},\overline{g},W}\right) does not admit constant solutions, then (ϵ¯,g¯)𝒟W,ν𝒰(\overline{\epsilon},\overline{g})\in\mathcal{D}^{*}_{W,\nu}\cap\mathcal{U}. Otherwise, the volume constraint shows that the unique constant solution is ν/μg¯(M)\nu/\mu_{\overline{g}}(M). This is a degenerate solution if, and only if, (QW,ϵ,g,u)\eqref{eqn:linearized-problem} admits a nontrivial solution. This only happens when there exists j=1,2,j=1,2,... such that

(3) ϵ¯2=W′′(ν/μg¯(M))αj(g¯),\overline{\epsilon}^{2}=-\frac{W^{\prime\prime}(\nu/\mu_{\overline{g}}(M))}{\alpha_{j}(\overline{g})},

where g¯={αj(g¯):j=1,2,}\mathcal{E}_{\overline{g}}=\mathinner{\left\{\alpha_{j}\left(\overline{g}\right)\mathrel{\mathop{\mathchar 58\relax}}j=1,2,...\right\}} is the set of nonzero eigenvalues of Δg¯-\Delta_{\overline{g}}. g¯\mathcal{E}_{\overline{g}} is a discrete subset of ]0,[]0,\infty[, so there exists ϵ^>0\hat{\epsilon}>0 such that (ϵ^,g¯)𝒟W,ν𝒰(\hat{\epsilon},\overline{g})\in\mathcal{D}_{W,\nu}\cap\mathcal{U} and (3) does not hold for any positive integer jj. This implies (ϵ^,g¯)𝒟W,ν(\hat{\epsilon},\overline{g})\in\mathcal{D}^{*}_{W,\nu}.

𝒟W,ν\mathcal{D}^{*}_{W,\nu} is an open subset of ]0,[×k]0,\infty[\times\mathcal{M}^{k}. Indeed, let (ϵ^,g^)𝒟W,ν(\hat{\epsilon},\hat{g})\in\mathcal{D}^{*}_{W,\nu}. If (PW,ν,ϵ^,g^)(P_{W,\nu,\hat{\epsilon},\hat{g}}) does not admit constant solutions, the result is a corollary of lemma 3.2. Otherwise, note that kgW′′(ν/μg(M))\mathcal{M}^{k}\ni g\mapsto W^{\prime\prime}(\nu/\mu_{g}(M))\in\mathbb{R} and kgαj(g)\mathcal{M}^{k}\ni g\mapsto\alpha_{j}(g)\in\mathbb{R} are continuous maps for any positive integer jj, so (ϵ^,g^)(\hat{\epsilon},\hat{g}) admits a neighborhood 𝒱\mathcal{V} in ]0,[×k]0,\infty[\times\mathcal{M}^{k} in which the constant solutions are nondegenerate. To conclude, 𝒱𝒟W,ν\mathcal{V}\cap\mathcal{D}_{W,\nu} is a neighborhood of (ϵ,g)(\epsilon,g) in ]0,[×k]0,\infty[\times\mathcal{M}^{k} for which the respective Allen-Cahn equation does not admit degenerate solutions. ∎

4. Technical steps

For a pair (ϵ,g)]0,[×k(\epsilon,g)\in]0,\infty[\times\mathcal{M}^{k}, let FW,ϵ,g:(Hg0(M))×Hg0×F_{W,\epsilon,g}\colon(H_{g_{0}}(M)\setminus\mathbb{R})\times\mathbb{R}\to H_{g_{0}}\times\mathbb{R} be given by FW,ϵ,g(u,λ)=FW(ϵ,g,u,λ).F_{W,\epsilon,g}(u,\lambda)=F_{W}(\epsilon,g,u,\lambda). We adopt similar notation when fixing other variables.

In lemma 4.2, we shall verify that the first hypothesis of theorem 3.1 holds. With that objective in mind, consider the following preliminary result:

Lemma 4.1.

Let gkg\in\mathcal{M}^{k}, Cg:Hg0(M)C_{g}\colon H_{g_{0}}(M)\to\mathbb{R} be given by Cg(v)=MvdμgC_{g}(v)=\int_{M}v\mathop{}\!\mathrm{d}\mu_{g} and Tg:Hg0(M)×Hg0(M)×T_{g}\colon H_{g_{0}}(M)\times\mathbb{R}\to H_{g_{0}}(M)\times\mathbb{R} be given by Tg(v,Λ)=(v,Cg(v))T_{g}(v,\Lambda)=\left(v,C_{g}(v)\right). Then TgT_{g} is a Fredholm operator of index 0.

Proof.

CgC_{g} is a linear functional, so codimkerCg=1\mathrm{codim}\ \ker C_{g}=1 in Hg0(M)H_{g_{0}}(M). This implies

codimTg(kerCg×)=2\mathrm{codim}\ T_{g}(\ker C_{g}\times\mathbb{R})=2

in Hg0(M)×H_{g_{0}}(M)\times\mathbb{R}.

Tg(kerCg×)Tg((1,0))=0,T_{g}(\ker C_{g}\times\mathbb{R})\cap T_{g}(\mathbb{R}(1,0))=0,

so

codim[Tg(kerCg×)+Tg((1,0))]=codimTg(kerCg×)1=1.\mathrm{codim}\ [T_{g}(\ker C_{g}\times\mathbb{R})+T_{g}(\mathbb{R}(1,0))]=\mathrm{codim}\ T_{g}(\ker C_{g}\times\mathbb{R})-1=1.
Hg0(M)×=(kerCg×)((1,0)),H_{g_{0}}(M)\times\mathbb{R}=(\ker C_{g}\times\mathbb{R})\oplus(\mathbb{R}(1,0)),

so codimimTg=1.\mathrm{codim}\ \mathrm{im}\ T_{g}=1. kerTg={0}×\ker T_{g}=\mathinner{\left\{0\right\}}\times\mathbb{R}, so TgT_{g} is a Fredholm operator of index 0. ∎

Lemma 4.2.

Given (ϵ,g)]0,[×k(\epsilon,g)\in]0,\infty[\times\mathcal{M}^{k}, FW,ϵ,gF_{W,\epsilon,g} is a Fredholm map of index 0.

Proof.

Fix (u,λ)Hg0(M)×(u,\lambda)\in H_{g_{0}}(M)\times\mathbb{R} and let KW,ϵ,g,u,λ:Hg0(M)×Hg0(M)×K_{W,\epsilon,g,u,\lambda}\colon H_{g_{0}}(M)\times\mathbb{R}\to H_{g_{0}}(M)\times\mathbb{R} be given by

KW,ϵ,g,u,λ(v,Λ)=(Aϵ,gd(BW)(u,λ)[v,Λ],0).K_{W,\epsilon,g,u,\lambda}(v,\Lambda)=\left(A_{\epsilon,g}\circ\mathop{}\!\mathrm{d}\left(B_{W}\right)_{(u,\lambda)}[v,\Lambda],0\right).

d(FW,ϵ,g)(u,λ)=TgKW,ϵ,g,u,λ\mathop{}\!\mathrm{d}\left(F_{W,\epsilon,g}\right)_{(u,\lambda)}=T_{g}-K_{W,\epsilon,g,u,\lambda}, where TgT_{g} was defined in lemma 4.1. Therefore, it suffices to prove that KW,ϵ,g,u,λK_{W,\epsilon,g,u,\lambda} is a compact operator to conclude that d(FW,ϵ,g)(u,λ)\mathop{}\!\mathrm{d}\left(F_{W,\epsilon,g}\right)_{(u,\lambda)} is a Fredholm operator with index 0. This is indeed the case, because Aϵ,gA_{\epsilon,g} is a compact operator. ∎

Let us examine the second hypothesis of the abstract transversality theorem. Let (ϵ,g,u,λ)FW1(0,ν)(\epsilon,g,u,\lambda)\in F_{W}^{-1}(0,\nu). To conclude that d(FW)(ϵ,g,u,λ)\mathop{}\!\mathrm{d}\left(F_{W}\right)_{(\epsilon,g,u,\lambda)} is surjective, it suffices to show that

(4) [imd(FW,ϵ,g)(u,λ)]imd(FW,ϵ,u,λ)g,\left[\mathrm{im}\ \mathop{}\!\mathrm{d}\left(F_{W,\epsilon,g}\right)_{(u,\lambda)}\right]^{\perp}\subset\mathrm{im}\ \mathop{}\!\mathrm{d}\left(F_{W,\epsilon,u,\lambda}\right)_{g},

which we shall prove in lemma 4.4.

The following defines an inner product on Hg0(M)×H_{g_{0}}(M)\times\mathbb{R}:

(u1,t1),(u2,t2)ϵ,g=Eϵ,g(u1,u2)+t1t2.\left\langle\left(u_{1},t_{1}\right),\left(u_{2},t_{2}\right)\right\rangle_{\epsilon,g}^{\prime}=E_{\epsilon,g}\left(u_{1},u_{2}\right)+t_{1}t_{2}.

This allows us to establish the characterization:

Remark 4.3.

Let (ϵ,g)]0,[×k(\epsilon,g)\in]0,\infty[\times\mathcal{M}^{k} and (u,λ),(v,Λ)Hg0(M)×(u,\lambda),\ (v,\Lambda)\in H_{g_{0}}(M)\times\mathbb{R}. Then

(v,Λ)[imd(FW,ϵ,g)(u,λ)] if, and only if, (v,Λ) is a solution for (QW,ϵ,g,u).(v,\Lambda)\in\left[\mathrm{im}\ \mathop{}\!\mathrm{d}\left(F_{W,\epsilon,g}\right)_{(u,\lambda)}\right]^{\perp}\text{ if, and only if, }(v,-\Lambda)\text{ is a solution for }\eqref{eqn:linearized-problem}.

We use this characterization to prove inclusion (4):

Lemma 4.4.

Let (ϵ,g,u,λ)FW1(0,ν)(\epsilon,g,u,\lambda)\in F_{W}^{-1}(0,\nu). Let (v,Λ)Hg0(M)×(v,-\Lambda)\in H_{g_{0}}(M)\times\mathbb{R} be a solution for (QW,ϵ,g,u)\eqref{eqn:linearized-problem}. If

d(FW,ϵ,u,λ)g[h],(v,Λ)ϵ,g=0\left\langle\mathop{}\!\mathrm{d}\left(F_{W,\epsilon,u,\lambda}\right)_{g}[h],(v,\Lambda)\right\rangle_{\epsilon,g}^{\prime}=0

for all h𝒮kh\in\mathcal{S}^{k}, then (v,Λ)=(0,0)(v,\Lambda)=(0,0).

Proof.

Due to lemmas 2.3, 2.4 and 2.8, the equation on the statement is rewritten

(5) Mϵ2bg,h(u,v)+(trgh)2[(W(u)λ)v+Λu]dμg=0,\int_{M}\epsilon^{2}b_{g,h}\left(\nabla u,\nabla v\right)+\frac{\left(\mathrm{tr}_{g}h\right)}{2}\left[\left(W^{\prime}(u)-\lambda\right)v+\Lambda u\right]\mathop{}\!\mathrm{d}\mu_{g}=0,

where we recall that bg,hb_{g,h} is a symmetric 2-covector on MM of class CkC^{k} given locally by

(bg,h)ij=(trgh)gij/2giqhqlglj.\left(b_{g,h}\right)_{ij}=\left(\mathrm{tr}_{g}h\right)g^{ij}/2-g^{iq}h_{ql}g^{lj}.

An argument with normal coordinates centered at arbitrary xMx\in M which considers specific perturbations of gg proves that

(6) g(u,v)=bg,h(u,v)=0Lg01(M)g\left(\nabla u,\nabla v\right)=b_{g,h}\left(\nabla u,\nabla v\right)=0\in L^{1}_{g_{0}}(M)

for any h𝒮kh\in\mathcal{S}^{k}. For details, see [GM11, Lemma 12].

Taking h=φgh=\varphi g for arbitrary φC(M)\varphi\in C^{\infty}(M) shows that (5) and (6) imply

(7) (W(u)λ)v+Λu=0Lg01(M).\left(W^{\prime}(u)-\lambda\right)v+\Lambda u=0\in L^{1}_{g_{0}}(M).

On one hand: integrating the equation above (7) yields

(8) M(W(u)λ)vdμg=Λν.\int_{M}\left(W^{\prime}(u)-\lambda\right)v\mathop{}\!\mathrm{d}\mu_{g}=-\Lambda\nu.

On the other hand: taking into account (6) and the fact that uu is a weak solution for ϵ2Δgu+W(u)=λ-\epsilon^{2}\Delta_{g}u+W^{\prime}(u)=\lambda,

MW(u)vdμg=Mλvdμg.\int_{M}W^{\prime}(u)v\mathop{}\!\mathrm{d}\mu_{g}=\int_{M}\lambda v\mathop{}\!\mathrm{d}\mu_{g}.

ν0\nu\neq 0, so the last equation and (8) imply Λ=0\Lambda=0. Due to (7), Λ=0\Lambda=0 implies

(W(u)λ)v=0Lg01(M).\left(W^{\prime}(u)-\lambda\right)v=0\in L^{1}_{g_{0}}(M).

If λW(u)0\lambda-W^{\prime}(u)\equiv 0, then uu is a weak solution for ϵ2Δgu=0-\epsilon^{2}\Delta_{g}u=0 – which only happens with a constant uu. We do not consider constant solutions, so λW(u)\lambda-W^{\prime}(u) does not vanish identically. Due to proposition 4.5, u,vu,v are functions of class C1C^{1}. Therefore, λW(u)\lambda-W^{\prime}(u) is a continuous function which does not vanish identically.

In particular, vv vanishes in a nonempty open subset of MM. In this context, we can use strong unique continuation ([PRS08, Theorem A.5]) in problem (QW,ϵ,g,u)\eqref{eqn:linearized-problem} to conclude that v=0Hg0(M)v=0\in H_{g_{0}}(M). ∎

Proposition 4.5.

Fix (ϵ,g)]0,[×k(\epsilon,g)\in]0,\infty[\times\mathcal{M}^{k} and α]0,1[\alpha\in]0,1[. If (u,λ)Hg0(M)×(u,\lambda)\in H_{g_{0}}(M)\times\mathbb{R} is a solution for (PW,ν,ϵ,g)\eqref{eqn:nonlinear-problem}, then uC1,α(M)u\in C^{1,\alpha}(M). If it also holds that (v,Λ)Hg0(M)×(v,\Lambda)\in H_{g_{0}}(M)\times\mathbb{R} is a solution for (QW,ϵ,g,u)\eqref{eqn:linearized-problem}, then vC1,α(M)v\in C^{1,\alpha}(M).

Proof.

Regularity is a local problem, so we fix a coordinate system ρ:Ωn\rho\colon\Omega\to\mathbb{R}^{n} where the gijg^{ij}s are bounded and Ω\Omega is the coordinate open subset of MM. Let u~:ρ(Ω)\tilde{u}\colon\rho(\Omega)\to\mathbb{R} be the local expression of uu. Note that u~\tilde{u} is a weak solution for

ϵ2i(gijju~)+ϵ2bi(iu~)+W(u~)=λ,-\epsilon^{2}\partial_{i}\left(g^{ij}\partial_{j}\tilde{u}\right)+\epsilon^{2}b^{i}\left(\partial_{i}\tilde{u}\right)+W^{\prime}\left(\tilde{u}\right)=\lambda,

where bi=j(gij)+gijΓkjkb^{i}=\partial_{j}\left(g^{ij}\right)+g^{ij}\Gamma_{kj}^{k} for any i=1,,ni=1,...,n.

A slight adaptation of [Str10, Lemma B.3] shows that u~Lq(ρ(Ω))\tilde{u}\in L^{q}\left(\rho(\Omega)\right) for every q<q<\infty. Arguing as in [Jos13, Theorem 12.2.2], one may show that given q>1q>1, W(u~)Lq(ρ(Ω))W^{\prime}\left(\tilde{u}\right)\in L^{q}\left(\rho(\Omega)\right) implies uH2,q(ρ(Ω))u\in H^{2,q}\left(\rho(\Omega)\right). To conclude, we use the Sobolev Embedding Theorem. ∎

The third hypothesis is proved analogously as[GM11, Lemma 11]:

Lemma 4.6.

πYι:FW1(0,ν)Y\pi_{Y}\circ\iota\colon F_{W}^{-1}(0,\nu)\to Y is σ\sigma-proper, where πY\pi_{Y} and ι\iota are defined in theorem 3.1.

Proof.

Given s=1,2,s=1,2,...; let

Cs=([1/s,s]×s¯×I(0,s)B(,1/s)¯×[s,s])FW1(0,ν),C_{s}=\left([1/s,s]\times\overline{\mathscr{B}_{s}}\times\overline{I(0,s)\setminus B(\mathbb{R},1/s)}\times[-s,s]\right)\cap F_{W}^{-1}(0,\nu),

where s,I(0,s)\mathscr{B}_{s},I(0,s) are respective open balls in 𝒮k,Hg0(M)\mathcal{S}^{k},H_{g_{0}}(M) centered at 0 with radius ss and

B(,1/s)={uHg0(M):infvu+vHg0<1/s}.B(\mathbb{R},1/s)=\mathinner{\left\{u\in H_{g_{0}}(M)\mathrel{\mathop{\mathchar 58\relax}}\inf_{v\in\mathbb{R}}\mathinner{\!\left\lVert u+v\right\rVert}_{H_{g_{0}}}<1/s\right\}}.

Fix a positive integer ss. Let us prove that πYι|Cs\left.\pi_{Y}\circ\iota\right|_{C_{s}} is a proper map. Let {(ϵn,gn,un,λn)}nCs\mathinner{\left\{\left(\epsilon_{n},g_{n},u_{n},\lambda_{n}\right)\right\}}_{n}\subset C_{s} be a sequence such that limngn=gk\lim_{n}g_{n}=g\in\mathcal{M}^{k}, limnϵn=ϵ[1/s,s]\lim_{n}\epsilon_{n}=\epsilon\in[1/s,s] and given nn, (un,λn)\left(u_{n},\lambda_{n}\right) is a solution for (PW,ν,ϵn,gn)(P_{W,\nu,\epsilon_{n},g_{n}}).

We claim that ((un,λn))n\left(\left(u_{n},\lambda_{n}\right)\right)_{n} has a convergent subsequence. Due to the Kondrakov theorem, the canonical inclusion iϵ,g,t:Hϵ,g0(M)Lg0t(M)i_{\epsilon,g,t}\colon H_{\epsilon,g_{0}}(M)\to L_{g_{0}}^{t}(M) is a compact operator for any t]2,pn[t\in\left]2,p_{n}\right[, so (un)n\left(u_{n}\right)_{n} converges in Lg0t(M)L_{g_{0}}^{t}(M) up to subsequence to a certain uLg0t(M)u\in L_{g_{0}}^{t}(M). (λn)n\left(\lambda_{n}\right)_{n} is bounded, so it converges up to a subsequence to a certain λ[s,s]\lambda\in[-s,s]. Arguing as in lemma 4.2, we see that limnAϵ,gBW(un,λn)=Aϵ,gBW(u,λ)\lim_{n}A_{\epsilon,g}\circ B_{W}\left(u_{n},\lambda_{n}\right)=A_{\epsilon,g}\circ B_{W}\left(u,\lambda\right). We can use the Mean Value Inequality and lemma 2.4 to prove that, in fact, limnAϵn,gnBW(un,λn)=Aϵ,gBW(u,λ)\lim_{n}A_{\epsilon_{n},g_{n}}\circ B_{W}\left(u_{n},\lambda_{n}\right)=A_{\epsilon,g}\circ B_{W}\left(u,\lambda\right). ∎

References

  • [BNAP20] Vieri Benci, Stefano Nardulli, Luis Eduardo Osorio Acevedo, and Paolo Piccione. Lusternik-schnirelman and morse theory for the van der waals-cahn-hilliard equation with volume constraint, 2020.
  • [GM11] M. Ghimenti and A. M. Micheletti. Non degeneracy for solutions of singularly perturbed nonlinear elliptic problems on symmetric riemannian manifolds, 2011.
  • [Heb00] E. Hebey. Nonlinear Analysis on Manifolds: Sobolev Spaces and Inequalities. American Mathematical Society, 2000.
  • [HHL05] Dan Henry, Jack Kenneth Hale, and Pereira Antônio Luiz. Perturbation of the boundary in boundary-value problems of partial differential equations. Cambridge University Press, 2005.
  • [Jos13] Jürgen Jost. Partial Differential Equations. Springer, 2013.
  • [Kav93] O. Kavian. Introduction à la théorie des points critiques: et applications aux problèmes elliptiques. Springer, 1993.
  • [MP09] A. M. Micheletti and A. Pistoia. Generic properties of singularly perturbed nonlinear elliptic problems on riemannian manifold. Advanced Nonlinear Studies, 9:803–813, 2009.
  • [PR03] Frank Pacard and Manuel Ritoré. From constant mean curvature hypersurfaces to the gradient theory of phase transitions. J. Differential Geom., 64(3):359–423, 01 2003.
  • [PRS08] Stefano Pigola, Marco Rigoli, and Alberto G. Setti. Vanishing and Finiteness Results in Geometric Analysis A Generalization of the Bochner Technique. Birkhäuser Basel, 2008.
  • [Str10] Michael Struwe. Variational methods: applications to nonlinear partial differential equations and Hamiltonian systems. Springer, 2010.