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

\AtAppendix

Rigidity of low index solutions on 𝐒3\mathbf{S}^{3}
via a Frankel theorem
for the Allen–Cahn equation

Fritz Hiesmayr University College London, 25 Gordon Street, London WC1H 0AY f.hiesmayr@ucl.ac.uk
Abstract.

We prove a rigidity theorem in the style of Urbano for the Allen–Cahn equation on the three-sphere: the critical points with Morse index five are symmetric functions that vanish on a Clifford torus. Moreover they realise the fifth width of the min-max spectrum for the Allen–Cahn functional. We approach this problem by analysing the nullity and symmetries of these critical points. We then prove a suitable Frankel-type theorem for their nodal sets, generally valid in manifolds with positive Ricci curvature. This plays a key role in establishing the conclusion, and further allows us to derive ancillary rigidity results in spheres with larger dimension.

1. Introduction

F. Urbano gave a description of the minimal surfaces in the round 𝐒3\mathbf{S}^{3} with low Morse index [Urb90]. The only embedded surfaces with index at most five are the equatorial spheres S𝐒3S\subset\mathbf{S}^{3}, with index one, and the Clifford tori T𝐒3T\subset\mathbf{S}^{3}, with index five. Our aim here is to prove an analogous result in the setting of the Allen–Cahn equation.

Consequences of Urbano’s theorem have found applications in numerous contexts, most notably the proof of the Willmore conjecture by Marques–Neves [MN14]. Their proof relied on the development of the so-called Almgren–Pitts min-max methods. In recent years an alternative min-max theory has grown, whose underlying idea is to first find critical points of a certain semilinear elliptic functional. (For the sake of this exposition we work in 𝐒3\mathbf{S}^{3}, although much of the discussion could remain unchanged in higher-dimensional spheres or general closed manifolds.) This is the Allen–Cahn functional Eϵ(u)=𝐒3ϵ2|u|2+ϵ1W(u)E_{\epsilon}(u)=\int_{\mathbf{S}^{3}}\frac{\epsilon}{2}\lvert\nabla u\rvert^{2}+\epsilon^{-1}W(u), which depends on a small parameter ϵ>0\epsilon>0. (For the sake of this exposition we may use the potential function W(t)=14(1t2)2W(t)=\frac{1}{4}(1-t^{2})^{2}.) This models the phase transition in a two-phase liquid. Its critical points solve the elliptic Allen–Cahn equation

(1) ϵΔuϵ1W(u)=0;\epsilon\Delta u-\epsilon^{-1}W^{\prime}(u)=0;

we write 𝒵ϵ\mathcal{Z}_{\epsilon} for the set of (classical) solutions of this equation. As ϵ0\epsilon\to 0 their values tend to congregate near ±1\pm 1, separated by a thin transition layer around a minimal surface.

A number of authors have made contributions to this alternative min-max theory; we give a condensed summary of only those results most pertinent here. Hutchinson–Tonegawa [HT00] proved that given a sequence of positive ϵj0\epsilon_{j}\to 0 and critical points uj𝒵ϵju_{j}\in\mathcal{Z}_{\epsilon_{j}}, one can form a sequence of varifolds VjV_{j} which converge weakly to a stationary integral varifold VV. When the Morse index of the uju_{j} is bounded along the sequence, then the combined work of Tonegawa–Wickramasekera [TW12] and Guaraco [Gua18] establishes that

(2) VjQ|Σ| as j,V_{j}\to Q\lvert\Sigma\rvert\text{ as $j\to\infty$},

where Σ𝐒3\Sigma\subset\mathbf{S}^{3} is a smooth embedded minimal surface and Q𝐙>0Q\in\mathbf{Z}_{>0} is its so-called multiplicity. (Both of these rely fundamentally on the deep results of Wickramasekera [Wic14] on the regularity of stable codimension one stationary varifolds.) Following this, Chodosh–Mantoulidis [CM20] proved that the convergence is actually with multiplicity one, that is Q=1Q=1. (This is the main element not available in higher dimensions, although it does hold in three-manifolds with positive Ricci curvature.)

It was Guaraco [Gua18] and Gaspar–Guaraco [GG18] who respectively implemented one-and multi-parameter min-max methods in the Allen–Cahn setting. When working with p𝐙>0p\in\mathbf{Z}_{>0} parameters, these produce a critical point which realises the pp-th width cϵ(p)c_{\epsilon}(p) of the min-max spectrum of the Allen–Cahn functional. With fixed ϵ>0\epsilon>0 and varying pp, these form a monotone increasing sequence,

(3) 0<cϵ(1)cϵ(p)cϵ(p+1).0<c_{\epsilon}(1)\leq\cdots\leq c_{\epsilon}(p)\leq c_{\epsilon}(p+1)\leq\cdots.

One is naturally led to draw comparisons with the sequence of widths attained by Almgren–Pitts methods: (ω(p)p𝐙>0)(\omega(p)\mid p\in\mathbf{Z}_{>0}). Nurser [Nur16] computed the first terms in this sequence in the round 𝐒3\mathbf{S}^{3}. The first four have values ω(1)==ω(4)=4π\omega(1)=\cdots=\omega(4)=4\pi, and are realised by equatorial spheres, and the next three widths are ω(5)=ω(6)=ω(7)=2π2\omega(5)=\omega(6)=\omega(7)=2\pi^{2}, and are realised by Clifford tori. Recently Caju–Gaspar–Guaraco–Matthiesen [CGGM20] proved an analogue for the first four min-max widths for the Allen–Cahn functional, namely they are equal and are realised by symmetric functions vanishing on equatorial spheres. Moreover, they point out that for topological reasons the next width cϵ(5)c_{\epsilon}(5) must be distinct from the first four. (The lower bounds for the phase transition spectrum of [GG18] give an alternative justification for this.)

We prove that in fact cϵ(5)c_{\epsilon}(5) is realised by a function vanishing on a Clifford torus TT, with the same symmetries as TT. This result, stated in Corollary 2 below, is a direct consequence of the following rigidity theorem.

Theorem 1.

Given any C>1C>1, there is ϵ0>0\epsilon_{0}>0 so that for all 0<ϵ<ϵ00<\epsilon<\epsilon_{0} the following holds. If a solution u𝒵ϵu\in\mathcal{Z}_{\epsilon} on 𝐒3\mathbf{S}^{3} has indexu5\operatorname{index}u\leq 5 and Eϵ(u)CE_{\epsilon}(u)\leq C, then its nodal set is either an equatorial sphere or a Clifford torus, and uu is a symmetric solution around it.

Its proof is given in the last section. We proceed by studying the nullity and comparing it to the Killing nullity of uu; this allows conclusions about the symmetry of uu.

Corollary 2.

There is ϵ0>0\epsilon_{0}>0 so that if 0<ϵ<ϵ00<\epsilon<\epsilon_{0} then any u𝒵ϵu\in\mathcal{Z}_{\epsilon} on 𝐒3\mathbf{S}^{3} with Eϵ(u)=cϵ(5)E_{\epsilon}(u)=c_{\epsilon}(5) is a symmetric solution vanishing on a Clifford torus.

These are not the first rigidity results for the Allen–Cahn equation. Besides the aforementioned [CGGM20] concerning the ground states, there is also recent work of Guaraco–Marques–Neves [GNM19] near non-degenerate minimal surfaces. In [GNM19] the authors adapt the curvature estimates of [CM20] to produce a non-trivial Jacobi field on the limit minimal surface. We sidestep the technical difficulties tied to such an approach, and instead start by proving a result inspired by Frankel’s theorem [Fra61, PW03].

Let (M,g)(M,g) be a closed manifold with positive Ricci curvature. Frankel’s theorem states that any two minimal hypersurfaces in MM must intersect. This is not quite true for the nodal sets Z(uϵi)={uϵi=0}Z(u_{\epsilon}^{i})=\{u_{\epsilon}^{i}=0\} of two solutions uϵi𝒵ϵu_{\epsilon}^{i}\in\mathcal{Z}_{\epsilon} on MM; see for example [GNM19, Expl. 1]. However, we prove that under mild topological hypotheses on the nodal sets a Frankel-type result actually does hold. The following theorem combines the statements of Corollary 4 and Proposition 5.

Theorem 3.

Let (M,g)(M,g) be a closed manifold of dimension n+12n+1\geq 2 with Ric>0\operatorname{Ric}>0. Let uϵ1,uϵ2±1u_{\epsilon}^{1},u_{\epsilon}^{2}\neq\pm 1 be two solutions of (5) on MM. If

  1. (A)

    either Z(uϵ1)Z(u_{\epsilon}^{1}) is separating and Z(uϵ2)Z(u_{\epsilon}^{2}) is connected,

  2. (B)

    or Z(uϵ1),Z(uϵ2)Z(u_{\epsilon}^{1}),Z(u_{\epsilon}^{2}) are both connected,

then Z(uϵ1)Z(uϵ2).Z(u_{\epsilon}^{1})\cap Z(u_{\epsilon}^{2})\neq\emptyset.

This Frankel property is of pivotal importance to our approach of the rigidity problem, and is one of the main reasons for the brevity of our arguments. Moreover it allows us to obtain ancillary rigidity results in spheres with dimension larger than three, the caveat being that in absence of [CM20] the convergence must be assumed to be with multiplicity one. We illustrate this with the Clifford-type minimal hypersurfaces

(4) Tp,q=𝐒p(p/n)×𝐒q(q/n)𝐒n+1,T_{p,q}=\mathbf{S}^{p}(\sqrt{p/n})\times\mathbf{S}^{q}(\sqrt{q/n})\subset\mathbf{S}^{n+1},

where p,q𝐙>0p,q\in\mathbf{Z}_{>0} and n=p+q2n=p+q\geq 2. By [AA81, HL71] these are known to have ν(Tp,q)=nullityTp,q=(p+1)(q+1)\nu(T_{p,q})=\operatorname{nullity}T_{p,q}=(p+1)(q+1). Moreover they are homogeneous, and are stabilised by SO(p)×SO(q)SO(n+2)SO(p)\times SO(q)\subset SO(n+2). The rigidity result valid near these surfaces is similar to Theorem 1, except that here no claim is made about the precise location of the nodal set. (Here and in the remainder we write (Tp,q)δ𝐒n+1(T_{p,q})_{\delta}\subset\mathbf{S}^{n+1} for the tubular neighbourhood of Tp,qT_{p,q} with size δ>0\delta>0.)

Theorem 4.

Let p,q𝐙>0p,q\in\mathbf{Z}_{>0} and n=p+q2n=p+q\geq 2. There exist 0<ϵ0,δ<10<\epsilon_{0},\delta<1 so that for all 0<ϵ<ϵ00<\epsilon<\epsilon_{0} there is a unique solution u𝒵ϵu\in\mathcal{Z}_{\epsilon} on 𝐒n+1\mathbf{S}^{n+1} with Z(u)(Tp,q)δZ(u)\subset(T_{p,q})_{\delta} and (1δ)n(Tp,q)Eϵ(u)(1+δ)n(Tp,q)(1-\delta)\mathcal{H}^{n}(T_{p,q})\leq E_{\epsilon}(u)\leq(1+\delta)\mathcal{H}^{n}(T_{p,q}), up to rotation and change of sign. Moreover uu is SO(p)×SO(q)SO(p)\times SO(q)-symmetric, up to conjugation.

In the more symmetric case where p=qp=q and n=2pn=2p, the information about the nodal set can be recovered. We state this corollary in a slightly different way.

Corollary 5.

Let p𝐙>0p\in\mathbf{Z}_{>0} and n=2pn=2p. Let ϵj0\epsilon_{j}\to 0, and uj𝒵ϵju_{j}\in\mathcal{Z}_{\epsilon_{j}} be so that V(ϵj,uj)|Tp,p|V(\epsilon_{j},u_{j})\to\lvert T_{p,p}\rvert. For large jj, the nodal set of uju_{j} is a rigid copy of Tp,pT_{p,p} and uju_{j} is a symmetric solution around it.

Theorem 4 and Corollary 5 are proved in essentially the same way as the main result, Theorem 1. We give the necessary modifications to the higher-dimensional setting in Appendix A but do not repeat the overlapping arguments. In general, beyond the restrictive assumptions on the limit minimal surface—homogeneity and integrability through rotations—our arguments are quite flexible, and apply in closed manifolds with positive Ricci curvature, of dimension three or larger.

Acknowledgements.

The author wishes to express his thanks to Costante Bellettini for his support and invaluable discussions, as well as to Lucas Ambrozio, Neshan Wickramasekera and Jason Lotay for helpful comments; in fact this work was prompted by a question of Lucas Ambrozio. The author’s research is supported by the EPSRC under grant EP/S005641/1.

2. Preliminary results

Consider 𝐒3\mathbf{S}^{3} equipped with the standard round metric. For every ϵ>0\epsilon>0 the ϵ\epsilon-Allen–Cahn equation is the semilinear partial differential equation

(5) ϵΔuϵ1W(u)=0,\epsilon\Delta u-\epsilon^{-1}W^{\prime}(u)=0,

where W:𝐑𝐑W:\mathbf{R}\to\mathbf{R} is a regular non-negative even potential with a positive, non-degenerate maximum at the origin and two minima at ±1\pm 1, where W(±1)=0W(\pm 1)=0. (For example one may take W(t)=(1t2)2/4W(t)=(1-t^{2})^{2}/4.)

This is the Euler–Lagrange equation of the Allen–Cahn functional, Eϵ(u)=𝐒3ϵ2|u|2+ϵ1W(u)E_{\epsilon}(u)=\int_{\mathbf{S}^{3}}\frac{\epsilon}{2}\lvert\nabla u\rvert^{2}+\epsilon^{-1}W(u). We write 𝒵ϵ\mathcal{Z}_{\epsilon} for the space of (classical) solutions of this equation. There is a standard way to associate to each solution u=uϵu=u_{\epsilon} a varifold Vϵ=V(ϵ,u)V_{\epsilon}=V(\epsilon,u), meaning a Radon measure on the Grassmann bundle Gr2(𝐒3)Gr_{2}(\mathbf{S}^{3}). Given any φC(Gr2(𝐒3))\varphi\in C(Gr_{2}(\mathbf{S}^{3})), this is defined by

(6) Vϵ(φ)=12σ𝐒3ϵ|u(X)|2φ(X,TX{u=u(X)})d3(X)V_{\epsilon}(\varphi)=\frac{1}{2\sigma}\int_{\mathbf{S}^{3}}\epsilon\lvert\nabla u(X)\rvert^{2}\varphi(X,T_{X}\{u=u(X)\})\mathop{}\!\mathrm{d}\mathcal{H}^{3}(X)

where {u=u(X)}\{u=u(X)\} is the level set of uu through XX and σ=11W(t)/2dt\sigma=\int_{-1}^{1}\sqrt{W(t)/2}\mathop{}\!\mathrm{d}t is a normalising constant. (We use very little in the way of varifold theory in the remainder, though one may consult [Sim84] for unfamiliar vocabulary.)

2.1. Index and nullity

Let (ϵjj𝐍)(\epsilon_{j}\mid j\in\mathbf{N}) be a positive sequence with ϵj0\epsilon_{j}\to 0 as jj\to\infty, and (ujj𝐍)(u_{j}\mid j\in\mathbf{N}) be a corresponding sequence with uj𝒵ϵju_{j}\in\mathcal{Z}_{\epsilon_{j}} for all jj. We assume uniform bounds for the energy and that the Morse index of these solutions is fixed, namely there are C>0C>0 and k𝐙0k\in\mathbf{Z}_{\geq 0} so that for all j𝐍j\in\mathbf{N},

(7) Eϵj(uj)C and indexuj=k.E_{\epsilon_{j}}(u_{j})\leq C\text{ and }\operatorname{index}u_{j}=k.

A subsequence of the Vj=V(ϵj,uj)V_{j}=V(\epsilon_{j},u_{j}) (which we extract without relabelling) converges to a stationary integral varifold, as was shown by Hutchinson–Tonegawa [HT00]. Relying on the regularity theory developed by Wickramasekera [Wic14], the combined work of Tonegawa–Wickramasekera [TW12] and Guaraco [Gua18] established that this limit is supported in a smooth embedded minimal surface Σ𝐒3\Sigma\subset\mathbf{S}^{3}, and

(8) VjQ|Σ|V_{j}\to Q\lvert\Sigma\rvert

with integer multiplicity Q𝐙>0Q\in\mathbf{Z}_{>0}. (Here and throughout we write |Σ|\lvert\Sigma\rvert for the unit density varifold associated to Σ\Sigma.) Using different methods, Hiesmayr [Hie18] and Gaspar [Gas20] both proved that

(9) indexΣk;\operatorname{index}\Sigma\leq k;

the former building on previous work of Tonegawa [Ton05] in the stable case, the latter by adapting computations of Le [Le11, Le15].

As the three-sphere 𝐒3\mathbf{S}^{3} has positive Ricci curvature, the results of Chodosh–Mantoulidis [CM20] give that the convergence is with multiplicity one, meaning Q=1Q=1 and Vj|Σ|V_{j}\to\lvert\Sigma\rvert as jj\to\infty. They also show that

(10) lim supj(indexuj+nullityuj)indexΣ+nullityΣ.\limsup_{j\to\infty}(\operatorname{index}u_{j}+\operatorname{nullity}u_{j})\leq\operatorname{index}\Sigma+\operatorname{nullity}\Sigma.

Combining this with the lower semicontinuity of the Morse index, we find that for large enough jj,

(11) nullityujnullityΣ.\operatorname{nullity}u_{j}\leq\operatorname{nullity}\Sigma.

We are especially interested in the situation where the Morse index of the uju_{j} is low, specifically where it is five at the most. In this range the classification of minimal surfaces of Urbano [Urb90] shows that for the limit minimal surface Σ\Sigma,

  • either indexΣ=1\operatorname{index}\Sigma=1 and Σ=S\Sigma=S is an equatorial sphere,

  • or indexΣ=5\operatorname{index}\Sigma=5 and Σ=T\Sigma=T is a Clifford torus.

By [HL71] they respectively have nullityS=3\operatorname{nullity}S=3 and nullityT=4\operatorname{nullity}T=4.

2.2. Action by rotations

The orthogonal group SO(4)SO(4) acts on the space of minimal surfaces in 𝐒3\mathbf{S}^{3}, and preserves their index and nullity. Given a surface Σ𝐒3\Sigma\subset\mathbf{S}^{3}, let as usual StabΣ={PSO(4)P(Σ)=Σ}\operatorname{Stab}\Sigma=\{P\in SO(4)\mid P(\Sigma)=\Sigma\} and OrbΣ={P(Σ)𝐒3PSO(4)}\operatorname{Orb}\Sigma=\{P(\Sigma)\subset\mathbf{S}^{3}\mid P\in SO(4)\}. The former is a closed Lie subgroup of SO(4)SO(4) denoted GΣG_{\Sigma}, and dimStabΣ+dimOrbΣ=6\dim\operatorname{Stab}\Sigma+\dim\operatorname{Orb}\Sigma=6, the dimension of SO(4)SO(4).

Let Σ\Sigma be minimal and LΣL_{\Sigma} be the Jacobi operator on Σ\Sigma. We write indexΣ𝐙0\operatorname{index}\Sigma\in\mathbf{Z}_{\geq 0} for the number of strictly negative eigenvalues of LΣL_{\Sigma}, counted with multiplicity. The kernel of LΣL_{\Sigma} is spanned by the so-called Jacobi fields. Those that stem from the action of SO(4)SO(4) are also called Killing Jacobi fields. They span a linear subspace of L2(Σ)L^{2}(\Sigma) of dimension ν(Σ)=dimOrbΣ\nu(\Sigma)=\dim\operatorname{Orb}\Sigma; this is the Killing nullity. (Here we follow the conventions of Hsiang–Lawson [HL71].) For any minimal surface Σ𝐒3\Sigma\subset\mathbf{S}^{3},

(12) ν(Σ)nullityΣ.\nu(\Sigma)\leq\operatorname{nullity}\Sigma.

We are especially interested in surfaces for which ν\nu equals the full nullity. Surfaces with this property are sometimes said to be integrable through rotations. The equatorial sphere and the Clifford torus both have this property, and quoting [HL71] they respectively have

(13) nullityS=ν(S)=3 and nullityT=ν(T)=4.\operatorname{nullity}S=\nu(S)=3\text{ and }\operatorname{nullity}T=\nu(T)=4.

Let Σ𝐒3\Sigma\subset\mathbf{S}^{3}, with isotropy subgroup GΣSO(4)G_{\Sigma}\subset SO(4). The surface is called homogeneous if GΣG_{\Sigma} acts transitively on it. This is equivalently expressed by saying that Σ\Sigma is an orbit of the action of GΣG_{\Sigma} on 𝐒3\mathbf{S}^{3}. The equatorial sphere and the Clifford torus both satisfy this property; in fact they are the only homogeneous minimal surfaces in 𝐒3\mathbf{S}^{3} [HL71].

These notions have natural analogues in the setting of the Allen–Cahn equation. The group SO(4)SO(4) gives a right action on functions defined on 𝐒3\mathbf{S}^{3} by pre-composition, where PSO(4)P\in SO(4) sends a function uu to uPu\circ P. Given a function uu on 𝐒3\mathbf{S}^{3} we write Stabu={PSO(4)uP=u}\operatorname{Stab}u=\{P\in SO(4)\mid u\circ P=u\} and Orbu={uPPSO(4)}\operatorname{Orb}u=\{u\circ P\mid P\in SO(4)\}. Again we have dimStabu+dimOrbu=6\dim\operatorname{Stab}u+\dim\operatorname{Orb}u=6. (To see this fix any α(0,1)\alpha\in(0,1); by elliptic regularity any critical point has uC2,α(𝐒3)u\in C^{2,\alpha}(\mathbf{S}^{3}). Consider the orbit map PSO(4)uPC2,α(𝐒3)P\in SO(4)\mapsto u\circ P\in C^{2,\alpha}(\mathbf{S}^{3}). Upon quotienting by the stabiliser Gu=StabuG_{u}=\operatorname{Stab}u this defines a homeomorphism onto its image. In particular SO(4)/GuSO(4)/G_{u} and Orbu\operatorname{Orb}u have the same dimension, namely 6dimGu6-\dim G_{u}. Note that a rigorous rederivation of the analogous orbit-stabiliser identity we quoted above for minimal surfaces Σ𝐒3\Sigma\subset\mathbf{S}^{3} would go along the same lines, working in a suitable Banach space of embeddings into 𝐒3\mathbf{S}^{3}.)

Let MϵM_{\epsilon} be the semilinear operator corresponding to (5). The linearisation of MϵM_{\epsilon} around a function uu is (up to multiplication by ϵ\epsilon) the linear operator Lϵ,u=Δϵ2W′′(u)L_{\epsilon,u}=\Delta-\epsilon^{-2}W^{\prime\prime}(u). (This describes the second variation of Eϵ-E_{\epsilon} at uu via integration by parts.) This operator has finite-dimensional kernel, whose dimension is denoted nullityu𝐙0\operatorname{nullity}u\in\mathbf{Z}_{\geq 0}. We also write indexu𝐙0\operatorname{index}u\in\mathbf{Z}_{\geq 0} for the number of strictly negative eigenvalues of Lϵ,uL_{\epsilon,u} counted with multiplicity. The action of SO(4)SO(4) preserves the Allen–Cahn functional, Eϵ(uP)=Eϵ(u)E_{\epsilon}(u\circ P)=E_{\epsilon}(u) for all PSO(4)P\in SO(4). It also maps 𝒵ϵ\mathcal{Z}_{\epsilon} to itself, and for all PSO(4)P\in SO(4), indexuP=indexu\operatorname{index}u\circ P=\operatorname{index}u and nullityuP=nullityu\operatorname{nullity}u\circ P=\operatorname{nullity}u. The invariance of EϵE_{\epsilon} under SO(4)SO(4) also means that the action generates functions in the kernel of Lϵ,uL_{\epsilon,u}, which span a space of dimension ν(u)𝐙0\nu(u)\in\mathbf{Z}_{\geq 0}, the Killing nullity of uu. As above we have ν(u)=dimOrbu\nu(u)=\dim\operatorname{Orb}u and

(14) ν(u)nullityu.\nu(u)\leq\operatorname{nullity}u.

(The fact that ν(u)=dimOrbu=6dimGu\nu(u)=\dim\operatorname{Orb}u=6-\dim G_{u} is not hard to see. To derive this rigorously requires working again with the orbit map PSO(4)uPC2,α(𝐒3)P\in SO(4)\to u\circ P\in C^{2,\alpha}(\mathbf{S}^{3}) we used above for the orbit-stabiliser identity. This differentiates to the map A𝔰𝔬(4)u,ξAA\in\mathfrak{so}(4)\mapsto-\langle\nabla u,\xi_{A}\rangle, where ξA\xi_{A} is the Killing vector field corresponding to AA. Writing 𝔤u𝔰𝔬(4)\mathfrak{g}_{u}\subset\mathfrak{so}(4) for the Lie algebra corresponding to GuG_{u}, we obtain a linear isomorphism between 𝔰𝔬(4)/𝔤u\mathfrak{so}(4)/\mathfrak{g}_{u} and the tangent space to the orbit of uu. In particular there are precisely ν=6dim𝔤u=6dimGu\nu=6-\dim\mathfrak{g}_{u}=6-\dim G_{u} Killing vector fields so that u,ξ1,,u,ξν\langle\nabla u,\xi_{1}\rangle,\dots,\langle\nabla u,\xi_{\nu}\rangle forms a linearly independent family.)

2.3. Consequences of multiplicity one convergence

Let ϵj0\epsilon_{j}\to 0 and (ujj𝐍)(u_{j}\mid j\in\mathbf{N}) be a sequence of critical points in 𝐒3\mathbf{S}^{3} with V(ϵj,uj)|Σ|V(\epsilon_{j},u_{j})\to\lvert\Sigma\rvert, where Σ𝐒3\Sigma\subset\mathbf{S}^{3} is an embedded minimal surface. Using either a simple calculation as in [CG19] or Lemma 2 as justification, one finds that for large jj,

(15) ν(Σ)ν(uj).\nu(\Sigma)\leq\nu(u_{j}).

Assume additionally that Σ\Sigma is integrable through rotations, that is ν(Σ)=nullityΣ\nu(\Sigma)=\operatorname{nullity}\Sigma. Stringing together (11), (14) and (15), we find that eventually

(16) nullityuj=ν(uj).\operatorname{nullity}u_{j}=\nu(u_{j}).

We now specialise this to the setting of low index in 𝐒3\mathbf{S}^{3}, and show that if indexuj5\operatorname{index}u_{j}\leq 5 then either indexuj=1\operatorname{index}u_{j}=1 or 55, provided jj is large enough. To this end assume that Eϵj(uj)CE_{\epsilon_{j}}(u_{j})\leq C and indexuj4\operatorname{index}u_{j}\leq 4 along the sequence. By (9) and Urbano’s classification, V(ϵj,uj)V(\epsilon_{j},u_{j}) converges to an equatorial sphere S𝐒3S\subset\mathbf{S}^{3}; moreover by [CM20] the convergence is with multiplicity one, that is V(ϵj,uj)|S|V(\epsilon_{j},u_{j})\to\lvert S\rvert as jj\to\infty. From the above we have that for large jj, nullityuj=ν(uj)=3\operatorname{nullity}u_{j}=\nu(u_{j})=3. At the same time 4indexuj+nullityujindexuj+34\geq\operatorname{index}u_{j}+\operatorname{nullity}u_{j}\geq\operatorname{index}u_{j}+3 by (10). As uju_{j} is unstable, we find indexuj=1\operatorname{index}u_{j}=1.

We apply this observation to the context of Corollary 2, where no assumption is made about the index of the solutions. Instead there only imposes that the uj𝒵ϵju_{j}\in\mathcal{Z}_{\epsilon_{j}} have Eϵj(uj)=cϵj(5)E_{\epsilon_{j}}(u_{j})=c_{\epsilon_{j}}(5). This notwithstanding, critical points obtained through five-parameter min-max arguments as in [GG18] have index at most five. Taking the conclusions of Theorem 1 for granted, and as cϵj(4)<cϵj(5)c_{\epsilon_{j}}(4)<c_{\epsilon_{j}}(5), the functions obtained from min-max methods must be symmetric solutions vanishing on a Clifford torus. Moreover cϵj(5)2π2c_{\epsilon_{j}}(5)\to 2\pi^{2} as jj\to\infty. By [MN14], the Clifford tori are the only minimal surfaces in 𝐒3\mathbf{S}^{3} whose area has this value (or a fraction thereof), so that V(ϵj,uj)|T|V(\epsilon_{j},u_{j})\to\lvert T\rvert, after extracting a subsequence if necessary. Arguing as above one obtains index bounds for the uju_{j}, which prove that they too must be of the rigid form described in Theorem 1.

3. A Frankel-type result for the Allen–Cahn equation

Let (M,g)(M,g) be a closed manifold with dimension n+1n+1 and positive Ricci curvature. We formulate two types of results for the Allen–Cahn equation in the style of Frankel’s theorem [Fra61]: the first concerning the solutions themselves, and the second their nodal sets.

Proposition 1.

Let ϵ>0\epsilon>0 and uϵ1uϵ2u_{\epsilon}^{1}\neq u_{\epsilon}^{2} be two solutions of (5) on MM. If uϵ1uϵ2u_{\epsilon}^{1}\leq u_{\epsilon}^{2} then one of the two is constant.

We postpone the proof of this for now, and move on to the Frankel-type result for nodal sets. Let us remark first that the nodal sets of two solutions uϵ1,uϵ2𝒵ϵu_{\epsilon}^{1},u_{\epsilon}^{2}\in\mathcal{Z}_{\epsilon} do not intersect in general. Indeed, Guaraco–Marques–Neves [GNM19, Expl. 1] give a short construction of solutions near equatorial spheres in 𝐒n+1\mathbf{S}^{n+1} for which this fails. (In their example one critical point vanishes on the equatorial sphere, and the other vanishes along two hypersurfaces lying on either side of it, a small distance away.) One must therefore impose natural hypotheses on the nodal sets to ensure that they meet. Given a function uu on MM, we write Z(u)={u=0}Z(u)=\{u=0\} for its nodal set, and say that it is separating if MZ(u)M\setminus Z(u) has exactly two connected components.

Proposition 2.

Let ϵ>0\epsilon>0 and uϵ1,uϵ2±1u^{1}_{\epsilon},u^{2}_{\epsilon}\neq\pm 1 be two solutions of (5) on MM. If Z(uϵ1)Z(u_{\epsilon}^{1}) is separating, then either Z(uϵ1)Z(uϵ2)Z(u_{\epsilon}^{1})\cap Z(u_{\epsilon}^{2})\neq\emptyset or

(17) Z(uϵ2){uϵ1>0} and Z(uϵ2){uϵ1<0}.Z(u_{\epsilon}^{2})\cap\{u_{\epsilon}^{1}>0\}\neq\emptyset\text{ and }Z(u_{\epsilon}^{2})\cap\{u_{\epsilon}^{1}<0\}\neq\emptyset.
Proof.

We argue by contradiction, and assume that Z(uϵ2){uϵ1>0}Z(u_{\epsilon}^{2})\subset\{u_{\epsilon}^{1}>0\}. Divide the complement of Z(uϵ2)Z(u_{\epsilon}^{2}) into its connected components, MZ(uϵ2)=j=0NUjM\setminus Z(u_{\epsilon}^{2})=\cup_{j=0}^{N}U_{j} where N𝐙0{}N\in\mathbf{Z}_{\geq 0}\cup\{\infty\}. (There are at most countably many of these, and the proof is the same whether NN is finite or infinite.) We decompose {uϵ1<0}\{u_{\epsilon}^{1}<0\} by conditioning on the UjU_{j}, and write {uϵ1<0}=j=0N{uϵ1<0}Uj\{u_{\epsilon}^{1}<0\}=\cup_{j=0}^{N}\{u_{\epsilon}^{1}<0\}\cap U_{j}. As Z(uϵ1)Z(u_{\epsilon}^{1}) is separating, only of these is non-empty, and {uϵ1<0}U0\{u_{\epsilon}^{1}<0\}\subset U_{0} say. Now on the one hand we may assume that uϵ2<0u_{\epsilon}^{2}<0 on U0U_{0}, flipping its sign if necessary. On the other hand, U0{uϵ1>0}\partial U_{0}\subset\{u_{\epsilon}^{1}>0\} and thus (U0{uϵ1<0}){uϵ1<0}\partial(U_{0}\cap\{u_{\epsilon}^{1}<0\})\subset\partial\{u_{\epsilon}^{1}<0\}. Therefore uϵ1=0u_{\epsilon}^{1}=0 on (U0{uϵ1<0})\partial(U_{0}\cap\{u_{\epsilon}^{1}<0\}). A simple consequence of the maximum principle—see [GNM19, Cor. 7.4]—means that uϵ2<uϵ1u_{\epsilon}^{2}<u_{\epsilon}^{1} on U0{uϵ10}U_{0}\cap\{u_{\epsilon}^{1}\leq 0\}. This actually holds on the whole U0U_{0}, as uϵ2<0uϵ1u_{\epsilon}^{2}<0\leq u_{\epsilon}^{1} on U0{uϵ10}U_{0}\setminus\{u_{\epsilon}^{1}\leq 0\}. Let UjU_{j} be one of the remaining components. If uϵ20u_{\epsilon}^{2}\leq 0 in UjU_{j} then uϵ2uϵ1u_{\epsilon}^{2}\leq u_{\epsilon}^{1}, so we may assume uϵ20u_{\epsilon}^{2}\geq 0. Then uϵ2=0u_{\epsilon}^{2}=0 on Uj\partial U_{j} while uϵ1>0u_{\epsilon}^{1}>0 on U¯j\overline{U}_{j}. Arguing as above we find uϵ2<uϵ1u_{\epsilon}^{2}<u_{\epsilon}^{1} on U¯j\overline{U}_{j}. As UjU_{j} was arbitrary, we conclude that uϵ2<uϵ1u_{\epsilon}^{2}<u_{\epsilon}^{1} on MM, which contradicts Proposition 1 because neither function is constant equal ±1\pm 1. ∎

Remark 3.

Technically [GNM19, Cor. 7.4] is stated on regular domains, but an inspection of the proof reveals this hypothesis to not be necessary.

The following is an immediate consequence.

Corollary 4.

If Z(uϵ1)Z(u^{1}_{\epsilon}) is separating and Z(uϵ2)Z(u_{\epsilon}^{2})\neq\emptyset is connected then Z(uϵ1)Z(uϵ2).Z(u^{1}_{\epsilon})\cap Z(u^{2}_{\epsilon})\neq\emptyset.

The next proposition can be derived independently, using a similar argument. Either Corollary 4 or Proposition 5 would be sufficient for our purposes; we include both for the sake of completeness.

Proposition 5.

If Z(uϵ1),Z(uϵ2)Z(u^{1}_{\epsilon}),Z(u^{2}_{\epsilon})\neq\emptyset are connected then Z(uϵ1)Z(uϵ2)Z(u^{1}_{\epsilon})\cap Z(u_{\epsilon}^{2})\neq\emptyset.

Proof.

Again we argue by contradiction, assuming that Z(uϵ1)Z(uϵ2)=Z(u_{\epsilon}^{1})\cap Z(u_{\epsilon}^{2})=\emptyset. Split Z=Z(uϵ1)=ZZ+Z=Z(u_{\epsilon}^{1})=Z_{-}\cup Z_{+} according to the sign of uϵ2u_{\epsilon}^{2}. As ZZ is connected, only one of these can be non-empty, and Z=ZZ=Z_{-} say. After perhaps flipping the sign of uϵ1u_{\epsilon}^{1}, we find that uϵ2<uϵ1u_{\epsilon}^{2}<u_{\epsilon}^{1} in {uϵ10}\{u_{\epsilon}^{1}\leq 0\} by [GNM19, Cor. 7.4]. The mirror argument shows that also uϵ2<uϵ1u_{\epsilon}^{2}<u_{\epsilon}^{1} in {uϵ20}\{u_{\epsilon}^{2}\geq 0\}. On the remaining region uϵ2<0<uϵ1u_{\epsilon}^{2}<0<u_{\epsilon}^{1}, which means uϵ2<uϵ1u_{\epsilon}^{2}<u_{\epsilon}^{1} is established on the entirety of MM; the conclusion follows from Proposition 1. ∎

We now turn to the proof of Proposition 1. For this we use the parabolic Allen–Cahn equation with initial data a bounded function u0C1(M)u_{0}\in C^{1}(M), which given ϵ>0\epsilon>0 and T(0,]T\in(0,\infty] is defined to be

(18) {ut=ϵΔuϵ1W(u)in M×[0,T),u(0,)=u0on M.\begin{cases}u_{t}=\epsilon\Delta u-\epsilon^{-1}W^{\prime}(u)&\text{in $M\times[0,T)$},\\ u(0,\cdot)=u_{0}&\text{on $M$}.\end{cases}
Lemma 6.

Let u0C1(M)u_{0}\in C^{1}(M) be a weak subsolution of (5) with |u0|1\lvert u_{0}\rvert\leq 1. Then the solution to (18) exists for all time, and 1u(s,)u(t,)1-1\leq u(s,\cdot)\leq u(t,\cdot)\leq 1 for all 0st0\leq s\leq t. As tt\to\infty, u(t,)u(t,\cdot) converges to a smooth function u+u_{+}. Moreover u+u_{+} is a stable solution of (5), and can be characterised as the least element of {v𝒵ϵvu0}\{v\in\mathcal{Z}_{\epsilon}\mid v\geq u_{0}\}.

Proof.

Using the classical theory of parabolic PDE, we find that the solution uu of (18) is unique, smooth and exists for all time. To obtain the monotonicity of the flow, note that utu_{t} satisfies the following parabolic equation: tut=ϵΔutϵ1W′′(u)ut\partial_{t}u_{t}=\epsilon\Delta u_{t}-\epsilon^{-1}W^{\prime\prime}(u)u_{t}. As initially ut(0,)0u_{t}(0,\cdot)\geq 0, the parabolic maximum principle forces ut0u_{t}\geq 0 for all t0t\geq 0. Moreover, this is strict unless ut=0u_{t}=0, which happens precisely if u0u_{0} is a solution of (5). Together with classical parabolic Schauder estimates, this monotonicity guarantees the convergence of u(t,)u(t,\cdot) as tt\to\infty, with limit the smooth function u+u_{+}. Moreover u+u_{+} is a classical solution of (5). As the set {v𝒵ϵvu0}\{v\in\mathcal{Z}_{\epsilon}\mid v\geq u_{0}\} works as a barrier for the flow, the function u+u_{+} must be its least element.

It remains to see the stability of u+u_{+}; there are various ways to obtain this. For example, assume that u+u_{+} is unstable and let φ1>0\varphi_{1}>0 be its first eigenfunction, with eigenvalue λ1<0\lambda_{1}<0. A quick computation reveals that for a suitably small θ(0,1)\theta\in(0,1), u+θφ1<u+u_{+}-\theta\varphi_{1}<u_{+} is a supersolution of (5). Indeed ϵΔ(u+θφ1)ϵ1W(u+θφ1)=φ1(θλ1ϵ10θW′′′(u+tφ1)tφ1dt)\epsilon\Delta(u_{+}-\theta\varphi_{1})-\epsilon^{-1}W^{\prime}(u_{+}-\theta\varphi_{1})=\varphi_{1}(\theta\lambda_{1}-\epsilon^{-1}\int_{0}^{\theta}W^{{}^{\prime\prime\prime}}(u_{+}-t\varphi_{1})t\varphi_{1}\mathop{}\!\mathrm{d}t). If we were to apply the arguments above, we would find that solving the parabolic Allen–Cahn equation with initial datum u+θφ1u_{+}-\theta\varphi_{1} yields a strictly decreasing solution. Perhaps after adjusting θ\theta to a smaller value so that u+θφ1>u0u_{+}-\theta\varphi_{1}>u_{0}, this acts as an upper barrier and makes u(t,)u+u(t,\cdot)\to u_{+} absurd. ∎

We use this to prove Proposition 1.

Proof.

In a manifold with positive Ricci curvature, the only stable solutions of (5) are the constant functions with values one of {1,0,1}\{-1,0,1\}. This is because, as a quick computation shows, δ2Eϵ(u)(|u|,|u|)=ϵM|2u|2||u||2+Ric(u,u)\delta^{2}E_{\epsilon}(u)(\lvert\nabla u\rvert,\lvert\nabla u\rvert)=-\epsilon\int_{M}\lvert\nabla^{2}u\rvert^{2}-\lvert\nabla\lvert\nabla u\rvert\rvert^{2}+\operatorname{Ric}(\nabla u,\nabla u). This is non-negative precisely when |u|0\lvert\nabla u\rvert\equiv 0 and uu is constant.

Let u1u2u_{1}\leq u_{2} be two distinct solutions of (5), and assume that u1u_{1} is not constant equal ±1\pm 1. By the maximum principle u1<u2u_{1}<u_{2} on MM. (If u1=0u_{1}=0 then this would imply that u2u_{2} is constant equal one.) Let φ1>0\varphi_{1}>0 be the first eigenfunction of the linearised operator Lϵ,u1L_{\epsilon,u_{1}}, with eigenvalue λ1<0\lambda_{1}<0. The same computation as in the proof of Lemma 6 shows that for sufficiently small θ(0,1)\theta\in(0,1), u1+θφ1u_{1}+\theta\varphi_{1} is a subsolution of (5). Take θ>0\theta>0 small enough that still u1+θφ1<u2u_{1}+\theta\varphi_{1}<u_{2}. Then solving (18) with this function as an initial datum we obtain a strictly increasing family u(t,)u(t,\cdot) of functions with u1<u(t,)u2u_{1}<u(t,\cdot)\leq u_{2}. By Lemma 6 this converges to a stable solution u+u_{+} when we let tt\to\infty. The characterisation of u+u_{+} given there shows that u+u2u_{+}\leq u_{2}, while the fact that Ric>0\operatorname{Ric}>0 means that u+u_{+} is constant equal 11. ∎

4. Symmetry of solutions

4.1. Stabilisers and weak convergence

The Lie group SO(4)SO(4) can be endowed with a bi-invariant metric, which induces a distance function dd. Write 𝒢\mathcal{G} for the set of closed subgroups of SO(4)SO(4). When endowed with the topology induced by the Hausdorff distance, the couple (𝒢,dH)(\mathcal{G},d_{H}) forms a compact metric space. Moreover, we have the following useful result. (We state this for an arbitrary compact Lie group GG, although for our purposes G=SO(4)G=SO(4) or SO(n+2)SO(n+2) would be sufficient. Moreover given a closed subgroup HGH\subset G, we let (H)τ={PGdist(P,H)<τ}(H)_{\tau}=\{P\in G\mid\operatorname{dist}(P,H)<\tau\} be its open tubular neighbourhood of size τ>0\tau>0.)

Lemma 1 ([MZ42]).

Let GG be a compact Lie group, and HH be a closed subgroup of GG. There is τ>0\tau>0 so that every subgroup H𝒢H^{\prime}\in\mathcal{G} with H(H)τH^{\prime}\subset(H)_{\tau} is conjugate to a subgroup of HH.

Let ϵj0\epsilon_{j}\to 0 be a sequence of positive scalars and (ujj𝐍)(u_{j}\mid j\in\mathbf{N}) be a sequence of solutions of (5), with respectively uj𝒵ϵju_{j}\in\mathcal{Z}_{\epsilon_{j}}. We assume that they additionally have uniformly bounded Allen–Cahn energy and indexuϵj=5\operatorname{index}u_{\epsilon_{j}}=5 for all jj, whence V(ϵj,uj)|T|V(\epsilon_{j},u_{j})\to\lvert T\rvert as jj\to\infty. We abbreviate

(19) Gj=Stabuj and GT=StabT.G_{j}=\operatorname{Stab}u_{j}\text{ and }G_{T}=\operatorname{Stab}T.

Moreover, we may assume throughout that jj is large enough that (11) holds, ensuring that nullityuj4\operatorname{nullity}u_{j}\leq 4. We use Lemma 1 to show that eventually the stabilisers GjG_{j} and GTG_{T} are conjugate subgroups of SO(4)SO(4).

First we point that out that SO(4)SO(4) defines a natural action on the space of two-varifolds 𝐕2(𝐒3)\operatorname{\mathbf{V}}_{2}(\mathbf{S}^{3}). A rotation PSO(4)P\in SO(4) maps V𝐕2(𝐒3)V\in\operatorname{\mathbf{V}}_{2}(\mathbf{S}^{3}) to its push-forward P#VP_{\#}V. (Here we implicitly identify PSO(4)P\in SO(4) with the isometry it induces on 𝐒3\mathbf{S}^{3}, as we have been doing.) This action is continuous in the varifold topology. To see this, let (P,V)SO(4)×𝐕2(𝐒3)(P,V)\in SO(4)\times\operatorname{\mathbf{V}}_{2}(\mathbf{S}^{3}) be arbitrary, and consider two sequences PkPP_{k}\to P and VkVV_{k}\to V, the latter being in the varifold topology. Given φC(Gr2(𝐒3))\varphi\in C(Gr_{2}(\mathbf{S}^{3})), one has (Pk#VkP#V)φ=Pk#(VkV)φ+(Pk#VP#V)φ(P_{k\#}V_{k}-P_{\#}V)\varphi=P_{k\#}(V_{k}-V)\varphi+(P_{k\#}V-P_{\#}V)\varphi, both of which tend to zero as kk\to\infty. We define the stabiliser and orbit of VV in the usual way. For an embedded surface Σ𝐒3\Sigma\subset\mathbf{S}^{3} and all PSO(4)P\in SO(4), it holds that P#|Σ|=|P(Σ)|P_{\#}\lvert\Sigma\rvert=\lvert P(\Sigma)\rvert and we need not distinguish between the stabiliser and orbit of Σ\Sigma as an embedded surface or as a varifold.

Let u𝒵ϵu\in\mathcal{Z}_{\epsilon}, and the varifold V(ϵ,u)V(\epsilon,u) be defined as in (6). Via a simple change of variable we find

(20) P#V(ϵ,u)=V(ϵ,uP1) for all PSO(4).P_{\#}V(\epsilon,u)=V(\epsilon,u\circ P^{-1})\text{ for all $P\in SO(4)$.}

To see this, let X𝐒3X\in\mathbf{S}^{3} be arbitrary and Y=P(X)Y=P(X). Assume without loss of generality that |u|(X)=|(uP1)|(Y)>0\lvert\nabla u\rvert(X)=\lvert\nabla(u\circ P^{-1})\rvert(Y)>0, ensuring that the level sets {u=u(X)}\{u=u(X)\} and {uP1=(uP1)(Y)}\{u\circ P^{-1}=(u\circ P^{-1})(Y)\} are regular near XX and YY respectively. The map PP induces on Gr2(𝐒3)Gr_{2}(\mathbf{S}^{3}) sends (X,TX{u=u(X)})(X,T_{X}\{u=u(X)\}) to (Y,TY{uP1=(uP1)(Y)}(Y,T_{Y}\{u\circ P^{-1}=(u\circ P^{-1})(Y)\}. Given any φC(Gr2(𝐒3))\varphi\in C(Gr_{2}(\mathbf{S}^{3})), we may thus compute [P#V(ϵ,u)](φ)=V(ϵ,u)(φP)[P_{\#}V(\epsilon,u)](\varphi)=V(\epsilon,u)(\varphi\circ P) to be equal V(ϵ,uP1)(φ)V(\epsilon,u\circ P^{-1})(\varphi):

(21) 12σ𝐒3ϵ|u|2(X)(φP)(X,TX{u=u(X)})d3(X)=12σ𝐒3ϵ|(uP1)|2(Y)φ(Y,TY{uP1=(uP1)(Y)}d3(Y).\frac{1}{2\sigma}\int_{\mathbf{S}^{3}}\epsilon\lvert\nabla u\rvert^{2}(X)(\varphi\circ P)(X,T_{X}\{u=u(X)\})\mathop{}\!\mathrm{d}\mathcal{H}^{3}(X)\\ =\frac{1}{2\sigma}\int_{\mathbf{S}^{3}}\epsilon\lvert\nabla(u\circ P^{-1})\rvert^{2}(Y)\varphi(Y,T_{Y}\{u\circ P^{-1}=(u\circ P^{-1})(Y)\}\mathop{}\!\mathrm{d}\mathcal{H}^{3}(Y).

Although the compared actions are on the right and left respectively,it follows from (20) that StabuStabV(ϵ,u)\operatorname{Stab}u\subset\operatorname{Stab}V(\epsilon,u), and specialised to our sequence GjStabV(ϵj,uj)G_{j}\subset\operatorname{Stab}V(\epsilon_{j},u_{j}) for all jj.

Lemma 2.

Let ϵj0\epsilon_{j}\to 0 and uj𝒵ϵju_{j}\in\mathcal{Z}_{\epsilon_{j}} be as described above, with V(ϵj,uj)|T|V(\epsilon_{j},u_{j})\to\lvert T\rvert as jj\to\infty. For large jj, GjG_{j} is conjugate to a subgroup of GT=SO(2)×SO(2)G_{T}=SO(2)\times SO(2).

Proof.

Consider a sequence (Pjj𝐍)(P_{j}\mid j\in\mathbf{N}) with PjStabujP_{j}\in\operatorname{Stab}u_{j}. Upon extracting a subsequence we may assume that it converges to some PSO(4)P\in SO(4). As V(ϵj,uj)|T|V(\epsilon_{j},u_{j})\to\lvert T\rvert we get on the one hand Pj#V(ϵj,uj)P#|T|P_{j\#}V(\epsilon_{j},u_{j})\to P_{\#}\lvert T\rvert. On the other hand Pj#V(ϵj,uj)=V(ϵj,ujPj1)=V(ϵj,uj)P_{j\#}V(\epsilon_{j},u_{j})=V(\epsilon_{j},u_{j}\circ P_{j}^{-1})=V(\epsilon_{j},u_{j}) by construction, so P#|T|=|T|P_{\#}\lvert T\rvert=\lvert T\rvert and PGT=StabTP\in G_{T}=\operatorname{Stab}T. Via another extraction argument, we find that given any τ>0\tau>0 there is J(τ)𝐍J(\tau)\in\mathbf{N} so that Gj(GT)τG_{j}\subset(G_{T})_{\tau} when jJ(τ)j\geq J(\tau). By Lemma 1, GjG_{j} is conjugate to a subgroup of GTG_{T}. ∎

Therefore eventually dimStabuj2\dim\operatorname{Stab}u_{j}\leq 2 and ν(uj)4\nu(u_{j})\geq 4. Combining this with (11), we find that for large jj

(22) ν(uj)=nullityuj=4.\nu(u_{j})=\operatorname{nullity}u_{j}=4.

Thus GjG_{j} is conjugate to GTG_{T}; further given any τ>0\tau>0 there is J(τ)𝐍J(\tau)\in\mathbf{N} so that for jJ(τ)j\geq J(\tau), there is PjSO(4)P_{j}\in SO(4) with d(Pj,I)<τd(P_{j},I)<\tau and Pj1GjPj=GTP_{j}^{-1}G_{j}P_{j}=G_{T}.

4.2. Proof of Theorem 1

Applying the results of [BO86] in the present setting, one obtains the following; see also [CGGM20] for an alternative construction of the symmetric critical point.

Lemma 3.

There is ϵ0=ϵ0(T)>0\epsilon_{0}=\epsilon_{0}(T)>0 so that for all 0<ϵ<ϵ00<\epsilon<\epsilon_{0} there is a unique function uT,ϵ=u𝒵ϵu_{T,\epsilon}=u\in\mathcal{Z}_{\epsilon} with {u=0}=T\{u=0\}=T, up to change of sign. Moreover uT,ϵu_{T,\epsilon} is invariant under GTG_{T}.

Proof.

It is convenient to write 𝐒3={(X,Y)𝐑2×𝐑2|X|2+|Y|2=1}\mathbf{S}^{3}=\{(X,Y)\in\mathbf{R}^{2}\times\mathbf{R}^{2}\mid\lvert X\rvert^{2}+\lvert Y\rvert^{2}=1\}. The complement of TT has two connected components U±U_{\pm}, and we write U+={(X,Y)𝐒3|X|<|Y|}U_{+}=\{(X,Y)\in\mathbf{S}^{3}\mid\lvert X\rvert<\lvert Y\rvert\} and U={(X,Y)𝐒3|X|>|Y|}U_{-}=\{(X,Y)\in\mathbf{S}^{3}\mid\lvert X\rvert>\lvert Y\rvert\}. These two regions are isometric via the involution (X,Y)𝐒3(Y,X)(X,Y)\in\mathbf{S}^{3}\mapsto(Y,X). Let U=U±U=U_{\pm} and λ1(Δ;U)\lambda_{1}(\Delta;U) be the first eigenvalue of the Laplacian on UU with Dirichlet eigenvalues. By [BO86], provided

(23) 0<ϵ<{W′′(0)/λ1(Δ;U)}1/2,0<\epsilon<\{W^{\prime\prime}(0)/\lambda_{1}(\Delta;U)\}^{1/2},

there is a unique solution u±u_{\pm} on U±U_{\pm} respectively of the system

(24) {ϵΔuϵ1W(u)=0in U,u>0in U,u=0on U.\begin{cases}\epsilon\Delta u-\epsilon^{-1}W^{\prime}(u)=0&\text{in $U$},\\ u>0&\text{in $U$},\\ u=0&\text{on $\partial U$}.\end{cases}

As this system is invariant under the action of GTG_{T}, the solutions u±u_{\pm} must further be GTG_{T}-invariant. Moreover they are symmetric under the involution above, that is u+(X,Y)=u(Y,X)u_{+}(X,Y)=u_{-}(Y,X) for all (X,Y)𝐒3(X,Y)\in\mathbf{S}^{3} with |X|<|Y|\lvert X\rvert<\lvert Y\rvert. By elliptic regularity u±u_{\pm} are respectively smooth up to and including the boundary of U±U_{\pm}. We define the function uu by setting

(25) u(X,Y)={u+(X,Y)if |X||Y|,u(X,Y)if |X|>|Y|.u(X,Y)=\begin{cases}u_{+}(X,Y)&\text{if $\lvert X\rvert\leq\lvert Y\rvert$,}\\ -u_{-}(X,Y)&\text{if $\lvert X\rvert>\lvert Y\rvert$}.\end{cases}

By the symmetry of u±u_{\pm} under the involution, this function is smooth and solves (5) on 𝐒3\mathbf{S}^{3}. Thus there exists at least one solution of (5) which vanishes precisely on TT. Conversely, if an arbitrary solution v𝒵ϵv\in\mathcal{Z}_{\epsilon} had Z(vϵ)=TZ(v_{\epsilon})=T then its restrictions to U±{U_{\pm}} would both have a sign, and thus vv is forced to coincide with uu up to change of sign. ∎

We conclude with a proof of the main result: Theorem 1.

Proof.

Let uϵ𝒵ϵu_{\epsilon}\in\mathcal{Z}_{\epsilon} be a solution with Eϵ(uϵ)CE_{\epsilon}(u_{\epsilon})\leq C and indexuϵ=5\operatorname{index}u_{\epsilon}=5. Let a small τ>0\tau>0 be given, and ϵ>0\epsilon>0 be small enough in terms of τ,D\tau,D that (22) holds and there is PϵSO(4)P_{\epsilon}\in SO(4) with d(Pϵ,I)<τd(P_{\epsilon},I)<\tau and Pϵ1GϵPϵ=GTP_{\epsilon}^{-1}G_{\epsilon}P_{\epsilon}=G_{T}, where Gϵ=StabuϵG_{\epsilon}=\operatorname{Stab}u_{\epsilon}. As the nodal set Z(uϵ)={uϵ=0}Z(u_{\epsilon})=\{u_{\epsilon}=0\} converges to TT with respect to Hausdorff distance we may moreover assume that Z(uϵ)(T)τZ(u_{\epsilon})\subset(T)_{\tau}. Let vϵ=uϵPϵ𝒵ϵv_{\epsilon}=u_{\epsilon}\circ P_{\epsilon}\in\mathcal{Z}_{\epsilon}; the energy, index and nullities are unaffected by this operation and vϵv_{\epsilon} is stabilised by GTG_{T}. Therefore its nodal set is union of orbits of GTG_{T}. Appealing to [CM20] for example, it must be connected and so be of the form Z(vϵ)={dist(,T)=δ}Z(v_{\epsilon})=\{\operatorname{dist}(\cdot,T)=\delta\} for some δ\delta which tends to zero as τ,ϵ0\tau,\epsilon\to 0. Let uT,ϵu_{T,\epsilon} be the symmetric solution around the Clifford torus from Lemma 3. The Frankel-type property of Corollary 4 forces δ=0\delta=0 and Z(vϵ)=Z(uT,ϵ)Z(v_{\epsilon})=Z(u_{T,\epsilon}). (In fact here both nodal sets are connected and separating.) By Lemma 3 up to a change of sign vϵ=uT,ϵv_{\epsilon}=u_{T,\epsilon}, which concludes the proof. ∎

Appendix A Modifications in higher dimensions

The only result that needs to be slightly altered in higher dimensions, when working with the hypersurface Tp,q𝐒n+1T_{p,q}\subset\mathbf{S}^{n+1}, is Lemma 3. Even then, in case p=qp=q the statement and its proof remain valid with no changes. However the less symmetric case where pqp\neq q calls for a more complicated construction; Caju–Gaspar [CG19] prove the following. (Their result is valid more broadly; we give a modified statement specialised to the present context.)

Lemma 1 ([CG19, Thm. 1.1]).

Let p,q>0p,q>0 and n=p+q3n=p+q\geq 3. Given any δ>0\delta>0 there is ϵ0>0\epsilon_{0}>0 so that for all 0<ϵ<ϵ00<\epsilon<\epsilon_{0} there is a solution up,q,ϵ𝒵ϵu_{p,q,\epsilon}\in\mathcal{Z}_{\epsilon} on 𝐒n+1\mathbf{S}^{n+1} with Z(up,q,ϵ)(Tp,q)δZ(u_{p,q,\epsilon})\subset(T_{p,q})_{\delta} and (1δ)n(Tp,q)Eϵ(up,q,ϵ)(1+δ)n(Tp,q)(1-\delta)\mathcal{H}^{n}(T_{p,q})\leq E_{\epsilon}(u_{p,q,\epsilon})\leq(1+\delta)\mathcal{H}^{n}(T_{p,q}).

Applying [BO86] here gives that any other solution vϵv_{\epsilon} of (5) on 𝐒n+1\mathbf{S}^{n+1} with Z(vϵ)=Z(up,q,ϵ)Z(v_{\epsilon})=Z(u_{p,q,\epsilon}) must coincide with up,q,ϵu_{p,q,\epsilon} up to a possible change of sign. There are two ways to justify this here. The first is via the results of  [CM20], which show that for ϵ>0\epsilon>0 small enough the nodal set of the up,q,ϵu_{p,q,\epsilon} converge smoothly to the limit surface Tp,qT_{p,q}. The uniqueness is then a direct consequence of [BO86], applied in the two regions making up 𝐒n+1Z(up,q,ϵ)\mathbf{S}^{n+1}\setminus Z(u_{p,q,\epsilon}). For an argument that does not rely open the convergence of the nodal sets, one may combine [HL89] with [BO86] to obtain the following general lemma.

Lemma 2.

Let (M,g)(M,g) be closed. Let ϵ>0\epsilon>0 and uϵ1,uϵ2u_{\epsilon}^{1},u_{\epsilon}^{2} be two solutions of (5) on MM. If Z(uϵ1)=Z(uϵ2)Z(u_{\epsilon}^{1})=Z(u_{\epsilon}^{2}), then uϵ1=±uϵ2u_{\epsilon}^{1}=\pm u_{\epsilon}^{2}.

Proof.

Write Z=Z(uϵ1)=Z(uϵ2)Z=Z(u_{\epsilon}^{1})=Z(u_{\epsilon}^{2}). The cases where Z=Z=\emptyset or Z=MZ=M are trivial, and we leave them aside. Divide the complement of ZZ into its connected components, say MZ=j=0NUjM\setminus Z=\cup_{j=0}^{N}U_{j}, where N𝐙0{}N\in\mathbf{Z}_{\geq 0}\cup\{\infty\}. The two functions uϵ1,uϵ2u_{\epsilon}^{1},u_{\epsilon}^{2} have a sign in the interior of each UjU_{j}, and vanish on Uj\partial U_{j}. By [BO86] they are equal up to a change of sign; however this needs to be chosen consistently across all regions of MZM\setminus Z. Call two regions Uj,UkU_{j},U_{k} adjacent if n1(UjUj)0\mathcal{H}^{n-1}(\partial U_{j}\cap\partial U_{j})\neq 0. By [HL89] the nodal set may be decomposed like Z=RSZ=R\cup S, where RR is the set of regular points, that is those XZX\in Z so that for some ρ>0\rho>0, Bρ(X)ZB_{\rho}(X)\cap Z is a C1C^{1} (n1)(n-1)-dimensional submanifold, and SS is a countably (n2)(n-2)–rectifiable set. If Uj,UkU_{j},U_{k} are adjacent then UjUk\partial U_{j}\cap\partial U_{k} contains a regular point XX say, and there is ρ>0\rho>0 so that UjUkBρ(X)=RBρ(X)\partial U_{j}\cap\partial U_{k}\cap B_{\rho}(X)=R\cap B_{\rho}(X). The regularity of the boundary near XX allows the application of [BO86, Lem. 1] to deduce that u1ν,u2ν0\frac{\partial u_{1}}{\partial\nu},\frac{\partial u_{2}}{\partial\nu}\neq 0 on UjUkBρ(X)\partial U_{j}\cap\partial U_{k}\cap B_{\rho}(X). It follows that the respective signs of u1,u2u_{1},u_{2} on UjU_{j} determines their signs on UkU_{k}, and vice-versa. Now let Uj,UkU_{j},U_{k} be two connected components, which are not necessarily adjacent. There is a path γ:[0,1]M\gamma:[0,1]\to M with endpoints γ(0)Uj\gamma(0)\in U_{j} and γ(1)Uk\gamma(1)\in U_{k}. Using a perturbation analogous to that used in [SW16, Lem. A.1] one may arrange for γ([0,1])S=\gamma([0,1])\cap S=\emptyset. The curve γ\gamma runs through finitely many regions of MZM\setminus Z. List them as Uj1=Uj,Uj2,,UjD=UkU_{j_{1}}=U_{j},U_{j_{2}},\dots,U_{j_{D}}=U_{k}, which are pairwise adjacent in this order. Therefore the sign of u1,u2u_{1},u_{2} on UjU_{j} determines their sign on UkU_{k} and vice-versa; this concludes the proof. ∎

For the remaining steps in the proof of Theorem 4 one may thus follow the arguments we used for Theorem 1.

References

  • [AA81] William K. Allard and Frederick J. Almgren, Jr. On the radial behavior of minimal surfaces and the uniqueness of their tangent cones. Ann. of Math. (2), 113(2):215–265, 1981.
  • [BO86] Haïm Brezis and Luc Oswald. Remarks on sublinear elliptic equations. Nonlinear Anal., 10(1):55–64, 1986.
  • [CG19] Rayssa Caju and Pedro Gaspar. Solutions of the Allen–Cahn equation on closed manifolds in the presence of symmetry. Preprint, 2019. https://arxiv.org/abs/1906.05938.
  • [CGGM20] Rayssa Caju, Pedro Gaspar, Marco Guaraco, and Henrik Matthiesen. Ground states of semilinear elliptic equations. Preprint, 2020. https://arxiv.org/abs/2006.10607.
  • [CM20] Otis Chodosh and Christos Mantoulidis. Minimal surfaces and the Allen-Cahn equation on 3-manifolds: index, multiplicity, and curvature estimates. Ann. of Math. (2), 191(1):213–328, 2020.
  • [Fra61] Theodore Frankel. Manifolds with positive curvature. Pacific J. Math., 11:165–174, 1961.
  • [Gas20] Pedro Gaspar. The second inner variation of energy and the Morse index of limit interfaces. J. Geom. Anal., 30(1):69–85, 2020.
  • [GG18] Pedro Gaspar and Marco Guaraco. The Allen–Cahn equation on closed manifolds. Calc. Var. Partial Differential Equations, 57(4):1–42, 2018.
  • [GNM19] Marco Guaraco, Andre Neves, and Fernando Marques. Multiplicity one and strictly stable Allen–Cahn minimal hypersurfaces. Preprint, 2019. http://arxiv.org/abs/192.08997.
  • [Gua18] Marco A. M. Guaraco. Min-max for phase transitions and the existence of embedded minimal hypersurfaces. J. Differential Geom., 108(1):91–133, 2018.
  • [Hie18] Fritz Hiesmayr. Spectrum and index of two-sided Allen-Cahn minimal hypersurfaces. Comm. Partial Differential Equations, 43(11):1541–1565, 2018.
  • [HL71] Wu-yi Hsiang and H. Blaine Lawson, Jr. Minimal submanifolds of low cohomogeneity. J. Differential Geometry, 5:1–38, 1971.
  • [HL89] Robert Hardt and Simon Leon. Nodal sets for solutions of elliptic equations. J. Differ. Geom., 30(2):505–522, 1989.
  • [HT00] John Hutchinson and Yoshihiro Tonegawa. Convergence of phase interfaces in the van der Waals-Cahn-Hilliard theory. Calc. Var. Partial Differential Equations, 10(1):49–84, 2000.
  • [Le11] Nam Le. On the second inner variation of the Allen–Cahn functional and its applications. Indiana Univ. Math. J., 60:1843–1856, 2011.
  • [Le15] Nam Le. On the second inner variation of Allen–Cahn type energies and applications to local minimizers. J. Math. Pures Appl., 103:1317–1345, 2015.
  • [MN14] Fernando C. Marques and André Neves. Min-max theory and the Willmore conjecture. Ann. of Math. (2), 179(2):683–782, 2014.
  • [MZ42] Deane Montgomery and Leo Zippin. A theorem on Lie groups. Bull. Amer. Math. Soc., 48:448–452, 1942.
  • [Nur16] Charles Nurser. Low min-max widths of the round three-sphere. PhD thesis, Imperial College London, 2016.
  • [PW03] Peter Petersen and Frederick Wilhelm. On Frankel’s theorem. Canad. Math. Bull., 46(1):130–139, 2003.
  • [Sim84] Leon Simon. Lectures on geometric measure theory. Proceedings of the Center for Mathematical Analysis. Australian National University, 1984.
  • [SW16] Leon Simon and Neshan Wickramasekera. A frequency function and singular set bounds for branched minimal immersions. Comm. Pure Appl. Math., 69:1213–1258, 2016.
  • [Ton05] Yoshihiro Tonegawa. On stable critical points for a singular perturbation problem. Communications in Analysis and Geometry, 13(2):439–459, 2005.
  • [TW12] Yoshihiro Tonegawa and Neshan Wickramasekera. Stable phase interfaces in the van der Waals–Cahn–Hilliard theory. J. Reine Angew. Math., 2012(668):191–210, 2012.
  • [Urb90] Francisco Urbano. Minimal surfaces with low index in the three-dimensional sphere. Proc. Amer. Math. Soc., 108(4):989–992, 1990.
  • [Wic14] Neshan Wickramasekera. A general regularity theory for stable codimension 1 integral varifolds. Ann. of Math. (2), 179(3):843–1007, 2014.