Rigidity of low index solutions on
via a Frankel theorem
for the Allen–Cahn equation
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 with low Morse index [Urb90]. The only embedded surfaces with index at most five are the equatorial spheres , with index one, and the Clifford tori , 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 , although much of the discussion could remain unchanged in higher-dimensional spheres or general closed manifolds.) This is the Allen–Cahn functional , which depends on a small parameter . (For the sake of this exposition we may use the potential function .) This models the phase transition in a two-phase liquid. Its critical points solve the elliptic Allen–Cahn equation
| (1) |
we write for the set of (classical) solutions of this equation. As their values tend to congregate near , 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 and critical points , one can form a sequence of varifolds which converge weakly to a stationary integral varifold . When the Morse index of the is bounded along the sequence, then the combined work of Tonegawa–Wickramasekera [TW12] and Guaraco [Gua18] establishes that
| (2) |
where is a smooth embedded minimal surface and 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 . (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 parameters, these produce a critical point which realises the -th width of the min-max spectrum of the Allen–Cahn functional. With fixed and varying , these form a monotone increasing sequence,
| (3) |
One is naturally led to draw comparisons with the sequence of widths attained by Almgren–Pitts methods: . Nurser [Nur16] computed the first terms in this sequence in the round . The first four have values , and are realised by equatorial spheres, and the next three widths are , 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 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 is realised by a function vanishing on a Clifford torus , with the same symmetries as . This result, stated in Corollary 2 below, is a direct consequence of the following rigidity theorem.
Theorem 1.
Given any , there is so that for all the following holds. If a solution on has and , then its nodal set is either an equatorial sphere or a Clifford torus, and 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 ; this allows conclusions about the symmetry of .
Corollary 2.
There is so that if then any on with 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 be a closed manifold with positive Ricci curvature. Frankel’s theorem states that any two minimal hypersurfaces in must intersect. This is not quite true for the nodal sets of two solutions on ; 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 be a closed manifold of dimension with . Let be two solutions of (5) on . If
-
(A)
either is separating and is connected,
-
(B)
or are both connected,
then
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) |
where and . By [AA81, HL71] these are known to have . Moreover they are homogeneous, and are stabilised by . 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 for the tubular neighbourhood of with size .)
Theorem 4.
Let and . There exist so that for all there is a unique solution on with and , up to rotation and change of sign. Moreover is -symmetric, up to conjugation.
In the more symmetric case where and , the information about the nodal set can be recovered. We state this corollary in a slightly different way.
Corollary 5.
Let and . Let , and be so that . For large , the nodal set of is a rigid copy of and 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 equipped with the standard round metric. For every the -Allen–Cahn equation is the semilinear partial differential equation
| (5) |
where is a regular non-negative even potential with a positive, non-degenerate maximum at the origin and two minima at , where . (For example one may take .)
This is the Euler–Lagrange equation of the Allen–Cahn functional, . We write for the space of (classical) solutions of this equation. There is a standard way to associate to each solution a varifold , meaning a Radon measure on the Grassmann bundle . Given any , this is defined by
| (6) |
where is the level set of through and 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 be a positive sequence with as , and be a corresponding sequence with for all . We assume uniform bounds for the energy and that the Morse index of these solutions is fixed, namely there are and so that for all ,
| (7) |
A subsequence of the (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 , and
| (8) |
with integer multiplicity . (Here and throughout we write for the unit density varifold associated to .) Using different methods, Hiesmayr [Hie18] and Gaspar [Gas20] both proved that
| (9) |
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 has positive Ricci curvature, the results of Chodosh–Mantoulidis [CM20] give that the convergence is with multiplicity one, meaning and as . They also show that
| (10) |
Combining this with the lower semicontinuity of the Morse index, we find that for large enough ,
| (11) |
We are especially interested in the situation where the Morse index of the 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 ,
-
•
either and is an equatorial sphere,
-
•
or and is a Clifford torus.
By [HL71] they respectively have and .
2.2. Action by rotations
The orthogonal group acts on the space of minimal surfaces in , and preserves their index and nullity. Given a surface , let as usual and . The former is a closed Lie subgroup of denoted , and , the dimension of .
Let be minimal and be the Jacobi operator on . We write for the number of strictly negative eigenvalues of , counted with multiplicity. The kernel of is spanned by the so-called Jacobi fields. Those that stem from the action of are also called Killing Jacobi fields. They span a linear subspace of of dimension ; this is the Killing nullity. (Here we follow the conventions of Hsiang–Lawson [HL71].) For any minimal surface ,
| (12) |
We are especially interested in surfaces for which 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) |
Let , with isotropy subgroup . The surface is called homogeneous if acts transitively on it. This is equivalently expressed by saying that is an orbit of the action of on . The equatorial sphere and the Clifford torus both satisfy this property; in fact they are the only homogeneous minimal surfaces in [HL71].
These notions have natural analogues in the setting of the Allen–Cahn equation. The group gives a right action on functions defined on by pre-composition, where sends a function to . Given a function on we write and . Again we have . (To see this fix any ; by elliptic regularity any critical point has . Consider the orbit map . Upon quotienting by the stabiliser this defines a homeomorphism onto its image. In particular and have the same dimension, namely . Note that a rigorous rederivation of the analogous orbit-stabiliser identity we quoted above for minimal surfaces would go along the same lines, working in a suitable Banach space of embeddings into .)
Let be the semilinear operator corresponding to (5). The linearisation of around a function is (up to multiplication by ) the linear operator . (This describes the second variation of at via integration by parts.) This operator has finite-dimensional kernel, whose dimension is denoted . We also write for the number of strictly negative eigenvalues of counted with multiplicity. The action of preserves the Allen–Cahn functional, for all . It also maps to itself, and for all , and . The invariance of under also means that the action generates functions in the kernel of , which span a space of dimension , the Killing nullity of . As above we have and
| (14) |
(The fact that is not hard to see. To derive this rigorously requires working again with the orbit map we used above for the orbit-stabiliser identity. This differentiates to the map , where is the Killing vector field corresponding to . Writing for the Lie algebra corresponding to , we obtain a linear isomorphism between and the tangent space to the orbit of . In particular there are precisely Killing vector fields so that forms a linearly independent family.)
2.3. Consequences of multiplicity one convergence
Let and be a sequence of critical points in with , where is an embedded minimal surface. Using either a simple calculation as in [CG19] or Lemma 2 as justification, one finds that for large ,
| (15) |
Assume additionally that is integrable through rotations, that is . Stringing together (11), (14) and (15), we find that eventually
| (16) |
We now specialise this to the setting of low index in , and show that if then either or , provided is large enough. To this end assume that and along the sequence. By (9) and Urbano’s classification, converges to an equatorial sphere ; moreover by [CM20] the convergence is with multiplicity one, that is as . From the above we have that for large , . At the same time by (10). As is unstable, we find .
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 have . 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 , the functions obtained from min-max methods must be symmetric solutions vanishing on a Clifford torus. Moreover as . By [MN14], the Clifford tori are the only minimal surfaces in whose area has this value (or a fraction thereof), so that , after extracting a subsequence if necessary. Arguing as above one obtains index bounds for the , 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 be a closed manifold with dimension 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 and be two solutions of (5) on . If 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 do not intersect in general. Indeed, Guaraco–Marques–Neves [GNM19, Expl. 1] give a short construction of solutions near equatorial spheres in 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 on , we write for its nodal set, and say that it is separating if has exactly two connected components.
Proposition 2.
Let and be two solutions of (5) on . If is separating, then either or
| (17) |
Proof.
We argue by contradiction, and assume that . Divide the complement of into its connected components, where . (There are at most countably many of these, and the proof is the same whether is finite or infinite.) We decompose by conditioning on the , and write . As is separating, only of these is non-empty, and say. Now on the one hand we may assume that on , flipping its sign if necessary. On the other hand, and thus . Therefore on . A simple consequence of the maximum principle—see [GNM19, Cor. 7.4]—means that on . This actually holds on the whole , as on . Let be one of the remaining components. If in then , so we may assume . Then on while on . Arguing as above we find on . As was arbitrary, we conclude that on , which contradicts Proposition 1 because neither function is constant equal . ∎
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 is separating and is connected then
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 are connected then .
Proof.
Again we argue by contradiction, assuming that . Split according to the sign of . As is connected, only one of these can be non-empty, and say. After perhaps flipping the sign of , we find that in by [GNM19, Cor. 7.4]. The mirror argument shows that also in . On the remaining region , which means is established on the entirety of ; 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 , which given and is defined to be
| (18) |
Lemma 6.
Proof.
Using the classical theory of parabolic PDE, we find that the solution of (18) is unique, smooth and exists for all time. To obtain the monotonicity of the flow, note that satisfies the following parabolic equation: . As initially , the parabolic maximum principle forces for all . Moreover, this is strict unless , which happens precisely if is a solution of (5). Together with classical parabolic Schauder estimates, this monotonicity guarantees the convergence of as , with limit the smooth function . Moreover is a classical solution of (5). As the set works as a barrier for the flow, the function must be its least element.
It remains to see the stability of ; there are various ways to obtain this. For example, assume that is unstable and let be its first eigenfunction, with eigenvalue . A quick computation reveals that for a suitably small , is a supersolution of (5). Indeed . If we were to apply the arguments above, we would find that solving the parabolic Allen–Cahn equation with initial datum yields a strictly decreasing solution. Perhaps after adjusting to a smaller value so that , this acts as an upper barrier and makes 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 . This is because, as a quick computation shows, . This is non-negative precisely when and is constant.
Let be two distinct solutions of (5), and assume that is not constant equal . By the maximum principle on . (If then this would imply that is constant equal one.) Let be the first eigenfunction of the linearised operator , with eigenvalue . The same computation as in the proof of Lemma 6 shows that for sufficiently small , is a subsolution of (5). Take small enough that still . Then solving (18) with this function as an initial datum we obtain a strictly increasing family of functions with . By Lemma 6 this converges to a stable solution when we let . The characterisation of given there shows that , while the fact that means that is constant equal . ∎
4. Symmetry of solutions
4.1. Stabilisers and weak convergence
The Lie group can be endowed with a bi-invariant metric, which induces a distance function . Write for the set of closed subgroups of . When endowed with the topology induced by the Hausdorff distance, the couple forms a compact metric space. Moreover, we have the following useful result. (We state this for an arbitrary compact Lie group , although for our purposes or would be sufficient. Moreover given a closed subgroup , we let be its open tubular neighbourhood of size .)
Lemma 1 ([MZ42]).
Let be a compact Lie group, and be a closed subgroup of . There is so that every subgroup with is conjugate to a subgroup of .
Let be a sequence of positive scalars and be a sequence of solutions of (5), with respectively . We assume that they additionally have uniformly bounded Allen–Cahn energy and for all , whence as . We abbreviate
| (19) |
Moreover, we may assume throughout that is large enough that (11) holds, ensuring that . We use Lemma 1 to show that eventually the stabilisers and are conjugate subgroups of .
First we point that out that defines a natural action on the space of two-varifolds . A rotation maps to its push-forward . (Here we implicitly identify with the isometry it induces on , as we have been doing.) This action is continuous in the varifold topology. To see this, let be arbitrary, and consider two sequences and , the latter being in the varifold topology. Given , one has , both of which tend to zero as . We define the stabiliser and orbit of in the usual way. For an embedded surface and all , it holds that and we need not distinguish between the stabiliser and orbit of as an embedded surface or as a varifold.
Let , and the varifold be defined as in (6). Via a simple change of variable we find
| (20) |
To see this, let be arbitrary and . Assume without loss of generality that , ensuring that the level sets and are regular near and respectively. The map induces on sends to . Given any , we may thus compute to be equal :
| (21) |
Although the compared actions are on the right and left respectively,it follows from (20) that , and specialised to our sequence for all .
Lemma 2.
Let and be as described above, with as . For large , is conjugate to a subgroup of .
Proof.
Consider a sequence with . Upon extracting a subsequence we may assume that it converges to some . As we get on the one hand . On the other hand by construction, so and . Via another extraction argument, we find that given any there is so that when . By Lemma 1, is conjugate to a subgroup of . ∎
Therefore eventually and . Combining this with (11), we find that for large
| (22) |
Thus is conjugate to ; further given any there is so that for , there is with and .
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 so that for all there is a unique function with , up to change of sign. Moreover is invariant under .
Proof.
It is convenient to write . The complement of has two connected components , and we write and . These two regions are isometric via the involution . Let and be the first eigenvalue of the Laplacian on with Dirichlet eigenvalues. By [BO86], provided
| (23) |
there is a unique solution on respectively of the system
| (24) |
As this system is invariant under the action of , the solutions must further be -invariant. Moreover they are symmetric under the involution above, that is for all with . By elliptic regularity are respectively smooth up to and including the boundary of . We define the function by setting
| (25) |
By the symmetry of under the involution, this function is smooth and solves (5) on . Thus there exists at least one solution of (5) which vanishes precisely on . Conversely, if an arbitrary solution had then its restrictions to would both have a sign, and thus is forced to coincide with up to change of sign. ∎
We conclude with a proof of the main result: Theorem 1.
Proof.
Let be a solution with and . Let a small be given, and be small enough in terms of that (22) holds and there is with and , where . As the nodal set converges to with respect to Hausdorff distance we may moreover assume that . Let ; the energy, index and nullities are unaffected by this operation and is stabilised by . Therefore its nodal set is union of orbits of . Appealing to [CM20] for example, it must be connected and so be of the form for some which tends to zero as . Let be the symmetric solution around the Clifford torus from Lemma 3. The Frankel-type property of Corollary 4 forces and . (In fact here both nodal sets are connected and separating.) By Lemma 3 up to a change of sign , 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 , is Lemma 3. Even then, in case the statement and its proof remain valid with no changes. However the less symmetric case where 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 and . Given any there is so that for all there is a solution on with and .
Applying [BO86] here gives that any other solution of (5) on with must coincide with 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 small enough the nodal set of the converge smoothly to the limit surface . The uniqueness is then a direct consequence of [BO86], applied in the two regions making up . 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 be closed. Let and be two solutions of (5) on . If , then .
Proof.
Write . The cases where or are trivial, and we leave them aside. Divide the complement of into its connected components, say , where . The two functions have a sign in the interior of each , and vanish on . By [BO86] they are equal up to a change of sign; however this needs to be chosen consistently across all regions of . Call two regions adjacent if . By [HL89] the nodal set may be decomposed like , where is the set of regular points, that is those so that for some , is a -dimensional submanifold, and is a countably –rectifiable set. If are adjacent then contains a regular point say, and there is so that . The regularity of the boundary near allows the application of [BO86, Lem. 1] to deduce that on . It follows that the respective signs of on determines their signs on , and vice-versa. Now let be two connected components, which are not necessarily adjacent. There is a path with endpoints and . Using a perturbation analogous to that used in [SW16, Lem. A.1] one may arrange for . The curve runs through finitely many regions of . List them as , which are pairwise adjacent in this order. Therefore the sign of on determines their sign on and vice-versa; this concludes the proof. ∎
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.