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

On the topology of stable minimal hypersurfaces in a homeomorphic S4S^{4}

Chao Li Courant Institute, New York University, 251 Mercer St, New York, NY 10012, USA [email protected]  and  Boyu Zhang Department of Mathematics, The University of Maryland at College Park, Maryland, 20742, USA [email protected]
Abstract.

We construct stable minimal hypersurfaces with simple topology in certain compact 44-manifolds XX with boundary, where XX embeds into a smooth manifold homeomorphic to S4S^{4}. For example, if XX is equipped with a Riemannian metric gg with positive scalar curvature, we prove the existence of a stable minimal hypersurface MM that is diffeomorphic to either S3S^{3} or a connected sum of S2×S1S^{2}\times S^{1}’s, ruling out spherical space forms in its prime decomposition. These results imply new theorems on the topology of black holes in four dimensions. The proof involves techniques from geometric measure theory and 44-manifold topology.

1. Introduction

Given a compact Riemannian manifold XX with nontrivial topology, a fundamental question is to construct closed minimal surfaces with controlled topology in XX. The classical result by Sacks-Uhlenbeck [17] produces branched minimally immersed S2S^{2}. When XX is three dimensional, deep results by Meeks-Yau [16] and Meeks-Simon-Yau [15] enable us to minimize area in homotopy and isotopy classes, respectively. In particular, one obtains area minimizing surfaces with controlled topology. These results have profound applications in geometry and topology.

It is generally impossible to control the topology of a kk-dimensional minimal submanifold in (Xn+1,g)(X^{n+1},g) when k>2k>2. Indeed, when k>2k>2, White [26, 27] proved that the least area mapping in a given homotopy class generally does not give a smooth immersion. The Federer-Fleming compactness theorem constructs area-minimizing currents in an integral homology class. These geometric objects enjoy much better regularity properties: for example, when k=n6k=n\leq 6, an area-minimizing hypersurface MnM^{n} in (Xn+1,g)(X^{n+1},g) is smooth. On the other hand, just knowing that MM is in a given homology class does not place much restrictions on its topological type.

The primary scope of this paper is to construct stable (or locally minimizing) minimal hypersurfaces with controlled topology in certain 44-manifolds under natural curvature conditions. Let (X4,g)(X^{4},g) be a smooth orientable 44-manifold. To motivate the discussion, recall that positive Ricci curvature of gg implies the nonexistence of two-sided stable minimal hypersurfaces. On the other hand, if gg has positive scalar curvature (abbreviated as PSC in the sequel), then the Schoen-Yau descent argument implies that a two-sided stable minimal M3M^{3} is Yamabe positive. Therefore, by Schoen-Yau [19], Gromov-Lawson [9] and Perelman, M3M^{3} is diffeomorphic to a connected sum of spherical space forms and S2×S1S^{2}\times S^{1}’s. Conversely, any such M3M^{3} may arise as a stable (or locally area-minimizing) hypersurface in certain PSC (X4,g)(X^{4},g).

In our first result, we prove that one obtains significantly better topology control of stable minimal hypersurfaces in a PSC 44-manifold (X4,g)(X^{4},g), provided that X4X^{4} is itself topologically simple. We introduce our important topological assumption on XX:

Assumption (\ast).

XX is a connected compact manifold with boundary, and there exists a smooth embedding ι:XS4\iota:X\to S^{4} such that S4ι(X)S^{4}\setminus\iota(X) has at least two connected components and at least one of them is simply connected.

Remark 1.1.

In fact, if we replace S4S^{4} by any smooth manifold that is homeomorphic to S4S^{4}, the results in this paper still hold without any change. Since there is no known example of a smooth manifold that is homeomorphic but not diffeomorphic to S4S^{4}, we only state Assumption (\ast)1 with S4S^{4} to simplify notation. If needed, one may replace S4S^{4} in Assumption (\ast)1 with any smooth manifold that is homeomorphic to S4S^{4}.

Since S4S^{4} is simply connected, each connected component of S4ι(X)S^{4}\setminus\iota(X) is bounded by exactly one component of (ι(X))\partial(\iota(X)). Thus, any such XX has at least two boundary components. Also, XX is necessarily orientable. Examples of such XX include the complement on S4S^{4} of finitely many open domains D1,,DnD_{1},\cdots,D_{n} whose closures are disjoint and at least one of {Dj}\{D_{j}\} is simply connected. In particular, S3×[0,1]S^{3}\times[0,1] and (#lS2×S1)×[0,1]\left(\#^{l}S^{2}\times S^{1}\right)\times[0,1] satisfy (\ast)1.

Theorem 1.2.

Suppose (X4,g)(X^{4},g) is a smooth Riemannian manifold satisfying (\ast)1, such that gg has positive scalar curvature and X\partial X is gg-weakly mean convex. Then (X4,g)(X^{4},g) contains a two-sided embedded stable minimal hypersurface that is diffeomorphic to S3S^{3} or a connected sum of S2×S1S^{2}\times S^{1}’s.

In other words, Theorem 1.2 rules out all nontrivial spherical components in the prime decomposition of MM. This was unknown even in the simplest case when X=S3×[0,1]X=S^{3}\times[0,1]. We will see in Section 5 that there exists (X4,g)(X^{4},g) satisfying the assumptions of Theorem 1.2 where the only stable minimal hypersurfaces are diffeomorphic to S2×S1S^{2}\times S^{1}. We note here that the assumption that X\partial X has at least two connected components is only used to guarantee that each any such component YY is homologically nontrivial, and this assumption may be removed if we know the existence of a stable minimal hypersurface in the class [Y][Y] for some other reasons.

In four dimensions, there is an interesting notion of curvature – the bi-Ricci curvature – that interpolates between the Ricci and the scalar curvature.

Definition 1.3.

Given a Riemannian manifold (Xn,g)(X^{n},g), pXp\in X and orthonormal e1,e2TpXe_{1},e_{2}\in T_{p}X, the bi-Ricci curvature pp of {e1,e2}\{e_{1},e_{2}\} is defined as

BiRic(e1,e2):=Ric(e1)+Ric(e2)sec(e1,e2).\operatorname{BiRic}(e_{1},e_{2}):=\operatorname{Ric}(e_{1})+\operatorname{Ric}(e_{2})-\sec(e_{1},e_{2}).

Equivalently, BiRicg(e1,e2)=Rgsecg(e3,e4)\operatorname{BiRic}_{g}(e_{1},e_{2})=R_{g}-\sec_{g}(e_{3},e_{4}) for an orthonormal basis {e1,,e4}\{e_{1},\cdots,e_{4}\}. Here RgR_{g} denotes the scalar curvature of gg. Observe that positive bi-Ricci curvature implies PSC. Manifolds with positive bi-Ricci curvature were first considered by Shen-Ye [22]. Bi-Ricci curvature and its generalizations have been studied systematically by Brendle-Hirsch-Johne [1] recently, and have deep applications in the stable Bernstein problem for minimal hypersurfaces [3, 14]. In [22], it is proved that a two-sided stable minimal hypersurface M3M^{3} in (X4,g)(X^{4},g) with positive bi-Ricci curvature has positive Ricci curvature in the spectral sense, and hence (by the resolved Poincaré conjecture) is diffeomorphic to a spherical space form. In our second result, we prove the existence of such MM that is actually diffeomorphic to S3S^{3}, provided that X4X^{4} satisfies (\ast)1.

Theorem 1.4.

Suppose (X4,g)(X^{4},g) satisfies all assumptions in Theorem 1.2, and additionally BiRicg>0\operatorname{BiRic}_{g}>0. Then (X4,g)(X^{4},g) contains a two-sided embedded stable minimal hypersurface that is diffeomorphic to S3S^{3}.

We expect such existence results to be applicable for further investigations on 44-manifolds.

1.1. Strategy of proof

In [28], White illustrated how to obtain some mild controls on the topology of solutions to the Plateau problem in a ball. His basic observation is that stable minimal hypersurfaces at extremal positions tend to bound topologically simple regions. We adopt this general principle in our construction. Given (X4,g)(X^{4},g) satisfying (\ast)1, let DD be a simply connected component of S4ι(X)S^{4}\setminus\iota(X), and YY be the boundary component of XX such that ι(Y)=D\iota(Y)=\partial D. The key is to consider, among stable minimal hypersurfaces homologous to YY, the one MM that is the nearest to YY. Through a novel minimization argument in general covering spaces (that are possibly neither compact or normal), we establish some conditions on the fundamental group of the region between YY and MM (see Proposition 2.1). This part of the proof is very general (e.g. does not depend on the dimension or the topological assumptions of XX), and we expect it to be useful for other problems.

The second part of the argument relies heavily on techniques in 44-manifold topology. Since XX satisfies (\ast)1, Proposition 2.1 implies that MM admits a locally flat embedding into S4S^{4} that bounds a simply connected region. The embedding problem of three-manifolds into S4S^{4} has been extensively investigated, and a collection of classical results may be found in a recent survey by Hillman [12]. These arguments are extremely well suited to study the case when M=(#S3/Γi)#(#l(S2×S1))M=(\#S^{3}/\Gamma_{i})\#(\#^{l}(S^{2}\times S^{1})), which holds in this setting because of the Yamabe positivity property. It follows from classical arguments that the only possible nontrivial spherical prime factor in MM is the Poincaré homology sphere. We then use Floer-theoretic invariants and a deep result of Taubes [24] to rule it out.

1.2. A theorem on the topology of black holes

Consider an asymptotically flat manifold (Xn,g)(X^{n},g) with nonnegative scalar curvature, and let EE be an end of XX. The outermost apparent horizon MM of EE, if non-empty, is defined as the outermost minimal hypersurface in EE. Hawking’s classical black hole topology theorem [11] states that when n=3n=3, MM is a disjoint union of S2S^{2}. Generalizations in higher dimensions due to Cai, Galloway and Schoen [2, 8, 7] conclude that if MM is smooth, then it is a stable minimal hypersurface, and is Yamabe positive.

Our proof of Theorem 1.2 reveals the topology of stable minimal hypersurfaces which is outermost with respect to a mean convex barrier. Thus it directly applies to studying the topology of apparent horizons in asymptotically flat four-manifolds, provided that X4X^{4} satisfies (\ast)1.

Theorem 1.5.

Suppose (X4,g)(X^{4},g) is a smooth asymptotically flat manifold with nonnegative scalar curvature, and XX is diffeomorphic to the interior of a manifold that satisfies (\ast)1. Then the outermost apparent horizon of each asymptotically flat end of XX is diffeomorphic to a disjoint union of S3S^{3} or a connected sum of S2×S1S^{2}\times S^{1}’s.

In fact, our proof only requires that the trapped region (the unbounded component separated by an apparent horizon on an asymptotically flat end) is a subset of 4\mathbb{R}^{4}, for which the topological assumption (\ast)1 is a natural sufficient condition. In particular, if X4X^{4} is diffeomorphic to the complement of finitely many compact sets on S4S^{4}, its apparent horizon of each asymptotically flat end is S3S^{3} or connected sums of S2×S1S^{2}\times S^{1}. We expect this conclusion to be sharp: Dahl-Larsson [5] constructed examples of apparent horizons on a subset of some asymptotically flat 4\mathbb{R}^{4} that are diffeomorphic to the disjoint union of ll copies of S2×S1S^{2}\times S^{1} for each l1l\geq 1. We note here that Theorem 1.5 also applies to other ALF or ALG four dimensional gravitational instantons with nonnegative scalar curvature, provided that XX is diffeomorphic to the interior of a manifold that satisfies (\ast)1. These include certain manifolds admitting a fibered end with fibers diffeomorphic to S2×S1S^{2}\times S^{1}. On the other hand, the topology of apparent horizons may be more complicated (e.g. lens spaces L(p,q)L(p,q)) if XX fails to satisfy (\ast)1, as illustrated by the beautiful black lenses examples constructed by Khuri-Rainone [13]. Let us remark that all these examples are simply connected 44-manifolds, which (after suitable compactifications) are homeomorphic to connected sums of S2×S2S^{2}\times S^{2}, P2\mathbb{C}P^{2} and P2-\mathbb{C}P^{2}. It would be interesting to investigate whether other spherical space forms (e.g. the Poincaré homology sphere) may appear as the outermost apparent horizon for a four dimensional gravitational instanton with nonnegative scalar curvature.

1.3. Organization of the paper

The paper is organized as follows. In Section 2, we establish general topological constraints for minimal hypersurfaces at extremal positions. The key result is Proposition 2.1, which relies on a very general existence result for homologically area-minimizing hypersurfaces in any covering space of a compact manifold with boundary. In Section 3, we focus on four-dimensional manifolds and use topological techniques to rule out nontrivial spherical space forms in the prime decomposition of the stable minimal hypersurface, finishing the proof of Theorem 1.2 and Theorem 1.4. Section 4 is devoted to the black hole topology theorem. Finally, in Section 5, we discuss an extension of our results for solutions to the Plateau problem in a homeomorphic D4D^{4}, examples of PSC embeddings of connected sums of lens spaces into S4S^{4}, and the Dahl-Larsson example of S2×S1S^{2}\times S^{1} apparent horizons. Some natural questions are also discussed and posed.

Acknowledgement

The authors are grateful to Claude LeBrun, Christos Mantoulidis, Rick Schoen and Brian White for stimulating conversations on various topics in this paper. We thank Daniel Ruberman for explaining the Zeeman construction to us, and Marcus Khuri for discussions on black hole topology and for patiently answering our questions on [13]. C.L. is supported by an NSF grant (DMS-2202343), a Simons Junior Faculty Fellowship and a Sloan Fellowship. B.Z. is supported by an NSF grant (DMS-2405271) and a travel grant from the Simons Foundation.

2. Stable minimal hypersurfaces of extremal positions

In this section, we obtain some preliminary topological constraints on stable minimal hypersurfaces in a compact Riemannian manifold (Xn+1,g)(X^{n+1},g) with weakly mean convex boundary. Assume n6n\leq 6 for the regularity of area-minimizing hypersurfaces 111The conclusions of this section should hold in higher dimensions as well, thanks to Simon’s maximum principle [23]. We do not pursue this direction.. Let YY be a connected component of X\partial X. Suppose that

I:=inf{n(M):Mn is a hypersurface in X homologous to Y}>0.I:=\inf\{\mathcal{H}^{n}(M):M^{n}\text{ is a hypersurface in }X\text{ homologous to }Y\}>0.

This holds when [Y]Hn(X,)[Y]\in H^{n}(X,\mathbb{Z}) is nonzero, for example when X\partial X has at least two connected components. Since X\partial X is weakly mean convex, II can be realized by a (possibly disconnected) smooth stable minimal hypersurface MnXn+1M^{n}\subset X^{n+1}. Consider

𝒮={MnXn+1:M is an embedded stable minimal hypersurface homologous to Y,n(M)n(Y)}.\mathcal{S}=\{M^{n}\subset X^{n+1}:M\text{ is an embedded stable minimal hypersurface}\\ \text{ homologous to }Y,\mathcal{H}^{n}(M)\leq\mathcal{H}^{n}(Y)\}.

By standard curvature estimates for stable minimal hypersurfaces [18, 20], 𝒮\mathcal{S} is compact in the CC^{\infty} topology. For each M𝒮M\in\mathcal{S}, let ΩM\Omega_{M} be the (n+1)(n+1)-dimensional manifold such that ΩM=YM\partial\Omega_{M}=Y-M (regarded as currents). The key result of this section is the Proposition 2.1. This is a nontrivial extension of [28, Theorem 1].

Proposition 2.1.

Given a compact oriented Riemannian manifold (Xn+1,g)(X^{n+1},g) as above. There exists M𝒮M\in\mathcal{S} such that either M=YM=Y, or ΩM\Omega_{M} is connected and satisfies

n+1(ΩM)=min{n+1(ΩM):M𝒮}.\mathcal{H}^{n+1}(\Omega_{M})=\min\{\mathcal{H}^{n+1}(\Omega_{M^{\prime}}):M^{\prime}\in\mathcal{S}\}.

Moreover, for every yYy\in Y, the inclusion i:YΩMi:Y\to\Omega_{M} induces a surjective mapping

i:π1(Y,y)π1(ΩM,y).i_{*}:\pi_{1}(Y,y)\to\pi_{1}(\Omega_{M},y).

Note that since YY is connected, the surjectivity of ii_{*} does not depend on the choice of the base point yy.

The rest of this section is devoted to the proof of Proposition 2.1. Since 𝒮\mathcal{S} is compact in CC^{\infty} convergence, there exists M𝒮M\in\mathcal{S} such that n+1(ΩM)=min{n+1(ΩM):M𝒮}\mathcal{H}^{n+1}(\Omega_{M})=\min\{\mathcal{H}^{n+1}(\Omega_{M^{\prime}}):M^{\prime}\in\mathcal{S}\}. Assume that MYM\neq Y. Then by the strong maximum principle, MM is contained in the interior of XX. Also, observe we necessarily have that ΩM\Omega_{M} is connected: otherwise discarding all connected components of ΩM\Omega_{M} except for the one containing YY yields an ΩM\Omega_{M^{\prime}} with smaller volume.

Suppose, for the sake of contradiction, that ii_{*} is not surjective. Take yYy\in Y, take the connected covering space π:(Ω~M,y~)(ΩM,y)\pi:(\tilde{\Omega}_{M},\tilde{y})\to(\Omega_{M},y) such that

π(π1(Ω~M,y~))=i(π1(Y,y)).\pi_{*}(\pi_{1}(\tilde{\Omega}_{M},\tilde{y}))=i_{*}(\pi_{1}(Y,y)).

Denote by g~=πg\tilde{g}=\pi^{*}g the covering metric and Y0Y_{0} the connected component of π1(Y)\pi^{-1}(Y) that contains y~\tilde{y}. Since ii_{*} is not surjective, π1(Y)\pi^{-1}(Y) has more than one connected components. Our basic idea is to prove that [Y0]Hn(Ω~M)[Y_{0}]\in H_{n}(\tilde{\Omega}_{M}) is nontrivial, and that we may minimize area in [Y0][Y_{0}] to find another area-minimizing hypersurface in Ω~M\tilde{\Omega}_{M}, so that its projection in ΩM\Omega_{M} gives a stable minimal hypersurface in ΩM\Omega_{M} that bounds a region with strictly smaller volume, contradicting the choice of MM.

However, a key difficulty is that generally, a minimizing sequence in a complete noncompact manifold does not necessarily converge. Note that we cannot assume that the covering space Ω~M\tilde{\Omega}_{M} is normal, and we cannot expect our area-minimizing hypersurface to be connected. To proceed, we establish some topological properties of the covering space of a compact manifold with boundary, and we carefully carry out the construction of an area-minimizing hypersurface. This argument seems novel and we expect it to be useful for other applications.

We start by introducing some notations. If XX is a manifold with boundary, let int(X)\operatorname{int}(X) denote the interior of XX.

Definition 2.2.

Suppose XX is a smooth oriented (n+1)(n+1)–manifold with boundary, aHn(X)a\in H_{n}(X), and (γ,γ)(X,X)(\gamma,\partial\gamma)\subset(X,\partial X) is a smooth oriented properly embedded compact 1-manifold. Let ωγ\omega_{\gamma} be a Thom form of γ\gamma. Define the intersection number of aa and γ\gamma to be a,[ωγ],\langle a,[\omega_{\gamma}]\rangle, where [ωγ]HdRn(int(X))Hn(X,)[\omega_{\gamma}]\in H^{n}_{dR}(\operatorname{int}(X))\cong H^{n}(X,\mathbb{R}) denotes the cohomology class of ωγ\omega_{\gamma}.

Remark 2.3.

If a=[M]a=[M] for a smooth embedded compact nn-manifold MM, then the intersection number of aa and γ\gamma equals Mωγ\int_{M}\omega_{\gamma}. If we further assume that MM intersects γ\gamma transversely, then the integral is equal to the signed counting of the number of intersection points between MM and aa.

Remark 2.4.

Alternatively, the intersection number of aa and γ\gamma can be defined without using differential forms as follows. Let [γ]H1(X,X)[\gamma]\in H_{1}(X,\partial X) be the fundamental class of γ\gamma. The inclusion map induces an isomorphism between Hn(int(X))H_{n}(\operatorname{int}(X)) and Hn(X)H_{n}(X), so we may view aa as an element of Hn(int(X))H_{n}(\operatorname{int}(X)). Poincaré duality gives an isomorphism

PD:Hn(int(X))Hc1(int(X)),PD:H_{n}(\operatorname{int}(X))\to H_{c}^{1}(\operatorname{int}(X)),

where HcH_{c}^{*} denotes the compactly supported cohomology. By the definition of Hc1(int(X))H_{c}^{1}(\operatorname{int}(X)), there exists a compact set Cint(X)C\subset\operatorname{int}(X) such that PD(a)PD(a) is represented by an element

aH1(int(X),int(X)C).a^{\prime}\in H^{1}(\operatorname{int}(X),\operatorname{int}(X)\setminus C).

The intersection number of aa and γ\gamma is equal to the pairing of aa^{\prime} with the image of [γ][\gamma] under the map

H1(X,X)H1(X,XC)H1(int(X),int(X)C),H_{1}(X,\partial X)\to H_{1}(X,X\setminus C)\xrightarrow{\cong}H_{1}(\operatorname{int}(X),\operatorname{int}(X)\setminus C),

where the first arrow is induced by inclusion, and the second map is the excision isomorphism.

The key topological property that enables us to construct homologically minimizing hypersurfaces in Ω~M\tilde{\Omega}_{M} is the following Lemma.

Lemma 2.5.

Suppose 0aHn(Ω~M)0\neq a\in H_{n}(\tilde{\Omega}_{M}). Then there exists a smooth oriented properly embedded compact 1-manifold (γ,γ)(Ω~M,Ω~M)(\gamma,\partial\gamma)\subset(\tilde{\Omega}_{M},\partial\tilde{\Omega}_{M}) such that the intersection number of γ\gamma and aa is non-zero.

Remark 2.6.

We emphasize that Lemma 2.5 holds for all covering spaces of a compact manifold with nonempty boundary, and it relies on the existence of boundary in an essential way. For example, if MM is a closed nn–manifold, then [M]Hn(×M)[M]\in H_{n}(\mathbb{R}\times M) has intersection number zero with every compact 11–submanifold of ×M\mathbb{R}\times M.

Proof of Lemma 2.5.

If Ω~M\tilde{\Omega}_{M} is compact, then the desired result follows from Poincaré duality. In the following, we assume that Ω~M\tilde{\Omega}_{M} is non-compact. Recall that g~\tilde{g} denotes the pull-back metric on Ω~M\tilde{\Omega}_{M} via the covering map.

Let xx be a fixed point in int(Ω~M)\operatorname{int}(\tilde{\Omega}_{M}), and let Bρ(x)B_{\rho}(x) be the geodesic ball with radius ρ\rho centered at xx. For generic ρ\rho, we have Bρ(x)\partial B_{\rho}(x) is a properly embedded, compact, smooth submanifold of Ω~M\tilde{\Omega}_{M} with codimension 11. Since every homology class is represented by finitely many singular simplices, for ρ\rho sufficiently large, the homology class aa is contained in the image of Hn(Bρ(x))Hn(Ω~M)H_{n}(B_{\rho}(x))\to H_{n}(\tilde{\Omega}_{M}). From now, let ρ\rho be a fixed number that is both sufficiently large and generic so that the above properties hold.

Since Bρ(x)\partial B_{\rho}(x) has only finitely many connected components, the set Ω~MBρ(x)\tilde{\Omega}_{M}\setminus B_{\rho}(x) has at most finitely many connected components. Let C1,,CnC_{1},\dots,C_{n} be the closures of the connected components of Ω~MBρ(x)\tilde{\Omega}_{M}\setminus B_{\rho}(x). Then each CiC_{i} is a manifold with corners, where the codimension of each corner stratum is at most 22.

Claim.

For each non-compact CiC_{i}, we have CiΩ~M\partial C_{i}\cap\partial\tilde{\Omega}_{M}\neq\emptyset.

Proof of the claim.

Assume there exists a CiC_{i} that is non-compact and CiΩ~M=\partial C_{i}\cap\partial\tilde{\Omega}_{M}=\emptyset. Then CiC_{i} is a manifold with boundary, and CiBρ(x)\partial C_{i}\subset\partial B_{\rho}(x).

Since ΩM\Omega_{M} is a compact connected Riemannian manifold with a non-empty boundary, it has the following two properties. These are straightforward extensions of the Hopf–Rinow theorem to manifolds with boundary:

  1. (1)

    Every geodesic [0,ϵ)int(ΩM)[0,\epsilon)\to\operatorname{int}(\Omega_{M}) either extends to a geodesic [0,+)int(ΩM)[0,+\infty)\to\operatorname{int}(\Omega_{M}) or to a geodesic [0,l]ΩM[0,l]\to\Omega_{M} such that [0,l)[0,l) is mapped to int(ΩM)\operatorname{int}(\Omega_{M}) and ll is mapped to ΩM\partial\Omega_{M}.

  2. (2)

    For every pint(ΩM)p\in\operatorname{int}(\Omega_{M}), there exists qΩMq\in\partial\Omega_{M} and a geodesic γ\gamma from pp to qq, such that γΩM\gamma\perp\partial\Omega_{M} at qq, and the length of γ\gamma equals the distance between pp and ΩM\partial\Omega_{M}.

By Property (1) above, a geodesic on CiC_{i} either extends indefinitely or intersects Ci\partial C_{i} in finite time. Fix a constant RR such that R>diamΩM.R>\operatorname{diam}\Omega_{M}. Since CiC_{i} is non-compact, there exists pCip\in C_{i} such that the distance from pp to Ci\partial C_{i} is at least RR. Therefore, every geodesic starting at pp can be extended to a geodesic in CiC_{i} with length at least RR. On the other hand, let π(p)\pi(p) be the image of pp on ΩM\Omega_{M}. By Property (2) above, there exists a geodesic on ΩM\Omega_{M} from π(p)\pi(p) to ΩM\partial\Omega_{M} with length at most diamΩM\operatorname{diam}\Omega_{M}. It lifts to a geodesic on Ω~M\tilde{\Omega}_{M}, which starts at pp and ends on Ω~M\partial\tilde{\Omega}_{M} and has length at most diamΩM<R\operatorname{diam}\Omega_{M}<R. This yields a contradiction. ∎

Now we finish the proof of the lemma using the claim. Let Ω^Ω~M\hat{\Omega}\subset\tilde{\Omega}_{M} be the union of Bρ(X)B_{\rho}(X) and all the CiC_{i}’s which are compact. Then Ω^\hat{\Omega} is a compact manifold with corners, and aa is in the image of H3(Ω^)H3(Ω~M)H_{3}(\hat{\Omega})\to H_{3}(\tilde{\Omega}_{M}).

Let a^H3(Ω^)\hat{a}\in H_{3}(\hat{\Omega}) be a preimage of aa. Since a0a\neq 0, we know that a^0\hat{a}\neq 0. Note that Ω^\hat{\Omega} is a compact smooth manifold with boundary after smoothing the corners. By Poincaé duality, there exists a properly embedded smooth 1-manifold (γ^,γ^)(Ω^,Ω^)(\hat{\gamma},\partial\hat{\gamma})\subset(\hat{\Omega},\partial\hat{\Omega}) such that the intersection number of a^\hat{a} with γ^\hat{\gamma} is non-zero, and we may perturb γ^\hat{\gamma} such that γ\partial\gamma does not intersect the corners of Ω^\partial\hat{\Omega}.

For every qγ^q\in\partial\hat{\gamma} with qΩ~Mq\notin\partial\tilde{\Omega}_{M}, there exists a unique CiC_{i} such that CiC_{i} is non-compact and qCiq\in\partial C_{i}. By the above claim, there exists qCiΩ~Mq^{\prime}\in\partial C_{i}\cap\partial\tilde{\Omega}_{M}. Let γq\gamma_{q} be a properly embedded arc in CiC_{i} that connects qq and qq^{\prime}. After perturbing γ\gamma and γq\gamma_{q} near qq, we may further assume that γγq\gamma\cup\gamma_{q} is smooth near qq.

Let

γ=γ^(qγq),\gamma=\hat{\gamma}\cup(\cup_{q}\gamma_{q}),

where the union takes over all qq such that qγ^,qΩ~Mq\in\partial\hat{\gamma},q\notin\partial\tilde{\Omega}_{M}. If n2n\geq 2, then after a generic perturbation, γ\gamma is a properly embedded 11–manifold in Ω~M\tilde{\Omega}_{M}. If n=1n=1, then after a generic perturbation γ\gamma is a properly immersed 11–manifold with transverse self-intersections, and we can resolve the self-intersection of γ\gamma to obtain an properly embedded 11–manifold in Ω~M\tilde{\Omega}_{M} with the same homology class.

Since all γq\gamma_{q} are disjoint from int(Ω^)\operatorname{int}(\hat{\Omega}), a Thom form of γ\gamma on Ω~M\tilde{\Omega}_{M} restricts to a Thom form of γ^\hat{\gamma} on Ω^\hat{\Omega}. So the intersection number of aa and γ\gamma in Ω~M\tilde{\Omega}_{M} is equal to the intersection number of a^\hat{a} and γ^\hat{\gamma} in Ω^\hat{\Omega}, which is non-zero by the definition of γ^\hat{\gamma}. Hence the lemma is proved. ∎

We are now ready to continue the proof of Proposition 2.1.

Proof of Proposition 2.1, continued.

Consider the minimization problem

I~=inf{n(N):N[Y0]}.\tilde{I}=\inf\{\mathcal{H}^{n}(N):N\in[Y_{0}]\}.

Let (γ0,γ0)(Ω~M,Ω~M)(\gamma_{0},\partial\gamma_{0})\subset(\tilde{\Omega}_{M},\partial\tilde{\Omega}_{M}) be the compact embedded curve constructed in Lemma 2.5.

Let ω0\omega_{0} be the Thom form of γ0\gamma_{0} as in Definition 2.2. By definition, ω0\omega_{0} is compactly supported in a neighborhood of γ0\gamma_{0}. Then any nn-cycle N[Y0]N\in[Y_{0}] satisfies that Nω00\int_{N}\omega_{0}\neq 0.

For ρ1\rho\gg 1 consider the g~\tilde{g}-geodesic ball Bρ(y~)B_{\rho}(\tilde{y}), and let Sρ(y~)S_{\rho}(\tilde{y}) be the g~\tilde{g}-geodesic sphere. Choose ρ\rho large enough such that [Y0]0Hn(Bρ(y~))[Y_{0}]\neq 0\in H_{n}(B_{\rho}(\tilde{y})). Perturbing ρ\rho a bit if necessary, we assume that Sρ(y~)S_{\rho}(\tilde{y}) meets Ω~M\partial\tilde{\Omega}_{M} transversely. Deform the metric g~\tilde{g} to g~\tilde{g}^{\prime} in a small neighborhood of Sρ(y~)S_{\rho}(\tilde{y}), such that Sρ(y~)S_{\rho}(\tilde{y}) is g~\tilde{g}^{\prime} strictly mean convex and meets Ω~M\partial\tilde{\Omega}_{M} orthogonally. We then may minimize the n\mathcal{H}^{n} volume (with respect to g~\tilde{g}^{\prime}) in the nontrivial homology [Y0][Y_{0}] in (Bρ(y~),g~)(B_{\rho}(\tilde{y}),\tilde{g}^{\prime}), and obtain a possibly disconnected area minimizing hypersurface NρN_{\rho}. Since

Nρω00,\int_{N_{\rho}}\omega_{0}\neq 0,

each NρN_{\rho} has a non-trivial intersection with the compact set suppω0\operatorname{supp}\omega_{0}. Consider Nρ0N_{\rho}^{0} the union of all connected components of NρN_{\rho} that intersect suppω0\operatorname{supp}\omega_{0}. Since Nρ0N_{\rho}^{0} has uniformly bounded n\mathcal{H}^{n}-volume, standard curvature estimates imply that they subsequentially (which we do not relabel) CC^{\infty} graphically converge to a limit N0N^{0} (possibly with integer multiplicity at this moment). Note that N0N^{0} is compact: since Ω~M\tilde{\Omega}_{M} has bounded geometry, by the monotonicity formula, there exist r0>0r_{0}>0 and ε0>0\varepsilon_{0}>0 depending only on ΩM\Omega_{M} such that if xN0x\in N^{0}, then the intersection of N0N^{0} and Br0(x)B_{r_{0}}(x) has n\mathcal{H}^{n}-volume at least ε0\varepsilon_{0}.

Therefore we conclude that {Nρ0}ρ\{N_{\rho}^{0}\}_{\rho} is also compactly supported, and the convergence to N0N^{0} in fact holds as currents. In particular, Nρ0N_{\rho}^{0} is homologous to N0N^{0} for sufficiently large ρ\rho.

If N0N^{0} is homologous to Y0Y_{0}, we are done. Otherwise, consider the homology class [Y0][N0][Y_{0}]-[N_{0}]. By Lemma 2.5, there exists a properly embedded compact curve (γ1,γ1)(Ω~M,Ω~M)(\gamma_{1},\partial\gamma_{1})\subset(\tilde{\Omega}_{M},\partial\tilde{\Omega}_{M}) that has a non-zero intersection number with [Y0][N0][Y_{0}]-[N_{0}]. Denote by ω1\omega_{1} the Thom class of γ1\gamma_{1}. Then for all sufficiently large ρ\rho, the minimizing sequence NρNρ0N_{\rho}-N_{\rho}^{0} satisfies that

NρNρ0ω10.\int_{N_{\rho}-N_{\rho}^{0}}\omega_{1}\neq 0.

Let Nρ1N_{\rho}^{1} be the union of the connected components of NρNρ0N_{\rho}-N_{\rho}^{0} that intersect the support of ω1\omega_{1}. By passing to a further subsequence (which again we do not relabel), the same argument as above finds a compact minimizing hypersurface N1N_{1} as the limit of {Nρ1}\{N_{\rho}^{1}\}.

Inductively, assuming that we have constructed N0,,NkN_{0},\cdots,N_{k}. If N0++NkN_{0}+\cdots+N_{k} (taking the sum as currents) is not homologous to Y0Y_{0}, we apply the above argument and construct a compact minimizing hypersurface Nk+1N_{k+1} from the minimizing sequence {Nρ(Nρ0++Nρk)}\{N_{\rho}-(N_{\rho}^{0}+\cdots+N_{\rho}^{k})\}. Note that since Ω~M\tilde{\Omega}_{M} is the covering space of a compact manifold, any minimal hypersurface has a uniform lower bound on its n\mathcal{H}^{n}-volume. This implies that this construction terminates in finitely many steps, since {Nρ}\{N_{\rho}\} has a uniform n\mathcal{H}^{n}-volume upper bound. We thus conclude that there exists a compact area minimizing hypersurface NN (possibly disconnected) in the homology class [Y0][Y_{0}]. Note that NN is not contained entirely in Ω~M\partial\tilde{\Omega}_{M}, as the only nn-currents entirely contained in Ω~M\partial\tilde{\Omega}_{M} that are homologous to [Y0][Y_{0}] are in the form Y0+kΩ~MY_{0}+k\partial\tilde{\Omega}_{M} for some integer kk (if Ω~M\partial\tilde{\Omega}_{M} is non-compact, then we must have k=0k=0). Since Y0Y_{0} is isometric to YY and π1(Y)\pi^{-1}(Y) has at least two connected components, the mass of a current of the form Y0+kΩ~MY_{0}+k\partial\tilde{\Omega}_{M} is at least n(Y0)\mathcal{H}^{n}(Y_{0}). We know NY0N\neq Y_{0} (as otherwise we would have picked M=YM=Y), so NN cannot be contained entirely in Ω~M\partial\tilde{\Omega}_{M}.

Let N^=π(N)\hat{N}=\pi(N). If N^\hat{N} is embedded, then it is homologous to YY in ΩM\Omega_{M}, and bounds a region with smaller volume than n+1(ΩM)\mathcal{H}^{n+1}(\Omega_{M}), contradicting the choice of MM. If N^\hat{N} is immersed, then it still represents the homology class [Y][Y] in Hn(ΩM,)H_{n}(\Omega_{M},\mathbb{Z}). We may then minimize the nn-dimensional volume in an open neighborhood of YY in ΩM\Omega_{M}, among hypersurfaces homologous to YY that does not intersect N^\hat{N} (note that immersed minimal hypersurfaces are weakly mean convex as self-intersections form an angle that is strictly less than π\pi). This produces another stable minimal hypersurface M𝒮M^{\prime}\in\mathcal{S} such that n+1(ΩM)<n+1(ΩM)\mathcal{H}^{n+1}(\Omega_{M^{\prime}})<\mathcal{H}^{n+1}(\Omega_{M}), contradiction. ∎

Remark 2.7.

If the covering (Ω~M,y~)(Y,y)(\tilde{\Omega}_{M},\tilde{y})\to(Y,y) is normal and the minimizer is connected, then we may also find the homologically minimizing hypersurface NN by translating (with deck transformations) a minimizing sequence to intersect a fixed compact set. This is the case treated by White [28].

Remark 2.8.

Using Lemma 2.5, we actually proved the following existence result for stable minimal hypersurfaces. Let (Ωn+1,g)(\Omega^{n+1},g) be a compact manifold with nonempty boundary, and Ω\partial\Omega is gg-weakly mean convex. Let (Ω~,g~)(\tilde{\Omega},\tilde{g}) be a covering space. Then any nonzero homology class αHn(Ω~,)\alpha\in H_{n}(\tilde{\Omega},\mathbb{Z}) can be represented by a compact area-minimizing hypersurface.

Remark 2.9.

Although not used in this paper, we observe that an analogous statement also holds for minimal hypersurfaces with obstacles. That is, without assuming that X\partial X is weakly mean convex, we may still consider the minimization problem

inf{n(Ω):Ω is an open set of X containing Y}.\inf\{\mathcal{H}^{n}(\partial\Omega):\Omega\text{ is an open set of }X\text{ containing }Y\}.

Wang [25, Corollary 3.3] proved that local minimizers of this problem enjoys a C1C^{1} compactness property. Thus, we may consider, among all open sets Ω\Omega containing YY that locally minimizes n(Ω)\mathcal{H}^{n}(\partial\Omega), the one with the smallest volume. The proof of Proposition 2.1 carries over to this situation verbatim and implies that π1(Y)π1(Ω)\pi_{1}(Y)\to\pi_{1}(\Omega) is surjective.

Remark 2.10.

On the other hand, the proof of Proposition 2.1 does not seem to easily extend to the case of stable constant mean curvature hypersurfaces, or more generally prescribed mean curvature surfaces. One key issue is that, in the universal cover Ω~M\tilde{\Omega}_{M}, other connected components of π1(Y)\pi^{-1}(Y) have a reversed bound of mean curvature, when regarded as a barrier for the minimization problem in the homology class of Y0Y_{0}.

3. Eliminating space forms

Now we focus on the case when n+1=4n+1=4 and prove Theorem 1.2. The proof of Theorem 1.4 is similar but simpler. Given (X4,g)(X^{4},g) satisfying the assumptions of Theorem 1.2, let DD be a simply connected component of S4ι(X)S^{4}\setminus\iota(X), and YY be the boundary component of XX such that ι(Y)=D\iota(Y)=\partial D. We apply Proposition 2.1 to find an embedded stable minimal hypersurface MM and the connected region ΩM\Omega_{M}, such that i:π1(Y)π1(ΩM)i_{*}:\pi_{1}(Y)\to\pi_{1}(\Omega_{M}) is surjective. If M=YM=Y, we define ΩM=Y\Omega_{M}=Y.

Denote by A=ι(ΩM)YDA=\iota(\Omega_{M})\cup_{Y}D. Since π1(D)=1\pi_{1}(D)=1 and ii_{*} is surjective, it follows from the van Kampen theorem that π1(A)=1\pi_{1}(A)=1. By the definitions of ΩM\Omega_{M} and AA, we know that ι(M)A\iota(M)\subset\partial A.

On the other hand, since MM is a stable minimal hypersurface in (X4,g)(X^{4},g) with Rg>0R_{g}>0, MM itself admits a PSC metric. Therefore, each connected component of MM is diffeomorphic to a connected sum of spherical space forms and S2×S1S^{2}\times S^{1}’s. The next basic lemma further implies that, in fact, each connected component of A\partial A individually bounds a simply connected domain of S4S^{4}.

Lemma 3.1.

Suppose AS4A\subset S^{4} is a simply connected domain with smooth boundary, and let MM be a connected component of A\partial A. Then S4MS^{4}\setminus M has two connected components. Let A,BA^{\prime},B be the closures of the components of S4MS^{4}\setminus M such that AAA\subset A^{\prime}. Then AA^{\prime} is also simply connected.

Proof.

Let D1,,DkD_{1},\dots,D_{k} be the closures of the connected components of S4AS^{4}\setminus A. Then A\partial A is the disjoint union of D1,,Dk\partial D_{1},\dots,\partial D_{k}. We first show that every Di\partial D_{i} is connected. Assume Di\partial D_{i} is not connected, let M1,M2M_{1},M_{2} be two connected components of Di\partial D_{i}, let pp be a point in the interior of DiD_{i}, let qq be a point in the interior of AA. Then there exist arcs γ1,γ2\gamma_{1},\gamma_{2} from p,qp,q such that γi\gamma_{i} intersects A\partial A transversely at one point in each MiM_{i} (i=1,2i=1,2). As a consequence, the arcs γ1\gamma_{1} and γ2\gamma_{2} combine to define a loop in S4S^{4} that intersects each of M1M_{1} and M2M_{2} transversely at one point. This contradicts the fact that H1(S4)=0H_{1}(S^{4})=0.

Let D^i=Di/Di\hat{D}_{i}=D_{i}/\partial D_{i}. By definition, D^i\hat{D}_{i} is the quotient space of DiD_{i} by collapsing Di\partial D_{i} to a point. The space D^i\hat{D}_{i} may not be a manifold. Let S{1,,k}S\subset\{1,\dots,k\} be non-empty, let AS=A(iSDi)A_{S}=A\cup(\cup_{i\in S}D_{i}). We claim that

(1) π1(AS)iS(π1(D^i)),\pi_{1}(A_{S})\cong*_{i\in S}\,(\pi_{1}(\hat{D}_{i})),

where the right-hand side is the free product.

We prove (1) by induction on the number of elements of SS. If S=S=\emptyset, the statement is trivial. Suppose (1) holds for all sets SS with ll elements, and consider a set SS^{\prime} of the form S=S{i}S^{\prime}=S\cup\{i\} where SS has ll elements and iSi\notin S. Then AS=ASDiA_{S^{\prime}}=A_{S}\cup D_{i}. By the van Kampen theorem,

π1(AS)π1(AS)π1(Di)π1(Di).\pi_{1}(A_{S^{\prime}})\cong\pi_{1}(A_{S})*_{\pi_{1}(\partial D_{i})}\pi_{1}(D_{i}).

Since the inclusion-induced map π1(Di)π1(AS)\pi_{1}(\partial D_{i})\to\pi_{1}(A_{S}) factors through π1(A)\pi_{1}(A), it is the trivial map. Therefore, we have

π1(AS)π1(AS)π1(Di)π1(Di)π1(AS)(π1(Di)/π1(Di))π1(AS)π1(D^i),\pi_{1}(A_{S^{\prime}})\cong\pi_{1}(A_{S})*_{\pi_{1}(\partial D_{i})}\pi_{1}(D_{i})\cong\pi_{1}(A_{S})*(\pi_{1}(D_{i})/\pi_{1}(\partial D_{i}))\cong\pi_{1}(A_{S})*\pi_{1}(\hat{D}_{i}),

where π1(Di)/π1(Di)\pi_{1}(D_{i})/\pi_{1}(\partial D_{i}) denotes the quotient of π1(Di)\pi_{1}(D_{i}) by the normal subgroup generated by π1(Di)\pi_{1}(\partial D_{i}). Hence (1) is proved.

Now we can prove the lemma. Let S={1,,k}S=\{1,\dots,k\}, we have S4=ASS^{4}=A_{S}, and hence (1) shows that π1(D^i)=1\pi_{1}(\hat{D}_{i})=1 for all ii. Applying (1) again, we conclude that π1(AS)=1\pi_{1}(A_{S})=1 for all SS. Since the set AA^{\prime} in the statement of the lemma has the form ΩS\Omega_{S} where SS contains all but one element of {1,,k}\{1,\dots,k\}, we know that AA^{\prime} is simply connected. ∎

By Lemma 3.1 and the previous discussions, there exists a connected stable minimal surface MM in XX such that ι(M)\iota(M) bounds a simply connected domain AA in S4S^{4}. Theorem 1.2 then follows from the following result in topology. We will establish a stronger statement that not only determines the topology of MM but also finds the homeomorphism type of the domain AA.

Proposition 3.2.

Suppose MM is a connected smooth submanifold of S4S^{4} such that one of the connected components of S4MS^{4}\setminus M is simply connected. Also assume that MM has the form

M=(#i(S3/Γi))#(#kS1×S2)M=\big{(}\#_{i}(S^{3}/\Gamma_{i})\big{)}\#\big{(}\#^{k}S^{1}\times S^{2}\big{)}

where Γi\Gamma_{i} acts freely and isometrically on S3S^{3}. Then M#kS1×S2M\cong\#^{k}S^{1}\times S^{2}, and the closure of the simply connected component of S4MS^{4}\setminus M is homeomorphic to kD2×S2\natural^{k}D^{2}\times S^{2}.

The next lemma uses classical arguments to prove some basic properties of the homology groups of the complement of MM. For its proof, we only need to assume that one of the components of S4MS^{4}\setminus M has trivial H1H_{1}. For later reference, we also allow MM to be a locally flat (but not necessarily smooth) submanifold.

Lemma 3.3.

Suppose MM is a locally flat connected oriented 33-submanifold in S4S^{4}. Let AA, BB denote the closures of the two components of the complement of MM. If H1(A)=0H_{1}(A)=0, then

  1. (1)

    H1(M)H_{1}(M) has no torsion.

  2. (2)

    The map H2(M)H2(A)H_{2}(M)\to H_{2}(A) induced by the inclusion is an isomorphism.

  3. (3)

    H3(A)=0H_{3}(A)=0.

Proof.

(1) We follow an argument from [12]. Consider the following commutative diagram

H2(A){H_{2}(A)}H2(A,M){H_{2}(A,M)}H2(S4){H_{2}(S^{4})}H2(S4,B){H_{2}(S^{4},B)}

where the maps are induced by the inclusions of spaces. Since H2(S4)=0H_{2}(S^{4})=0, and since by excision, the map H2(A,M)H2(S4,B)H_{2}(A,M)\to H_{2}(S^{4},B) is an isomorphism, we conclude that the map H2(A)H2(A,M)H_{2}(A)\to H_{2}(A,M) is zero. Now consider the homology long exact sequence

H2(A)0H2(A,M)H1(M)H1(A)0,H_{2}(A)\xrightarrow{0}H_{2}(A,M)\to H_{1}(M)\to H_{1}(A)\cong 0,

we have H1(M)H2(A,M)H_{1}(M)\cong H_{2}(A,M). By Lefschetz duality, we have H2(A,M)H2(A)H_{2}(A,M)\cong H^{2}(A). By the universal coefficient theorem, the torsion of H2(A)H^{2}(A) is isomorphic to the torsion of H1(A)H_{1}(A), which is zero. So H1(M)H_{1}(M) has no torsion.

(2) By the Mayer-Vietoris sequence

0H3(S4)H2(M)H2(A)H2(B)H2(S4)0,0\cong H_{3}(S^{4})\to H_{2}(M)\to H_{2}(A)\oplus H_{2}(B)\to H_{2}(S^{4})\cong 0,

we have H2(M)H2(A)H2(B)H_{2}(M)\cong H_{2}(A)\oplus H_{2}(B), so the map H2(M)H2(A)H_{2}(M)\to H_{2}(A) induced by the inclusion is surjective. On the other hand, by the exact sequence

H3(A,M)H2(M)H2(A)H_{3}(A,M)\to H_{2}(M)\to H_{2}(A)

and Lefschetz duality H3(A,M)H1(A)0H_{3}(A,M)\cong H^{1}(A)\cong 0, we know that the map H2(M)H2(A)H_{2}(M)\to H_{2}(A) is injective. Hence it is an isomorphism.

(3) By Lefschetz duality, H3(A)H1(A,M)H_{3}(A)\cong H^{1}(A,M). By the cohomology long exact sequence for (A,M)(A,M), we have an exact sequence

H0(A)H0(M)H1(A,M)H1(A)0,H^{0}(A)\xrightarrow{\cong}H^{0}(M)\to H^{1}(A,M)\to H^{1}(A)\cong 0,

so H1(A,M)0H^{1}(A,M)\cong 0. ∎

We continue with the proof of Proposition 3.2. Let AA, BB be the closures of the two components of the complement of MM such that π1(A)\pi_{1}(A) is trivial. By the assumptions of Proposition 3.2, the manifold MM has the form (#i(S3/Γi))#(#kS1×S2)\big{(}\#_{i}(S^{3}/\Gamma_{i})\big{)}\#\big{(}\#^{k}S^{1}\times S^{2}\big{)}. Write M1=#i(S3/Γi)M_{1}=\#_{i}(S^{3}/\Gamma_{i}), M2=#kS1×S2M_{2}=\#^{k}S^{1}\times S^{2}. Following the usual convention, M1M_{1} or M2M_{2} is defined to be S3S^{3} if there is no factor in the connected sum expression.

In the connected sum decomposition of MM, there are kk factors of S1×S2S^{1}\times S^{2}. Let A^\hat{A} be obtained from AA by attaching kk 3-handles to A=M\partial A=M, such that the kk attaching spheres are given by {pt}×S2\{pt\}\times S^{2} in each component of S1×S2S^{1}\times S^{2}.

Lemma 3.4.

A^\partial\hat{A} is homeomorphic to M1M_{1}, and A^\hat{A} is contractible.

Proof.

A^\partial\hat{A} is obtained from A\partial{A} by removing a tubular neighborhood of each attaching sphere and gluing 2 copies of D2D^{2} for each removed neighborhood. So A^\partial\hat{A} is homeomorphic to M1M_{1}.

Now we show that H1(A^)H2(A^)H3(A^)0H_{1}(\hat{A})\cong H_{2}(\hat{A})\cong H_{3}(\hat{A})\cong 0.

Attaching 3-handles does not modify H1H_{1}, so we have H1(A^)H1(A)0H_{1}(\hat{A})\cong H_{1}(A)\cong 0.

Consider the exact sequence

(2) H3(A)H3(A^)H3(A^,A)H2(A)H2(A^)H2(A^,A).H_{3}(A)\to H_{3}(\hat{A})\to H_{3}(\hat{A},A)\to H_{2}(A)\to H_{2}(\hat{A})\to H_{2}(\hat{A},A).

By excision, H2(A^,A)(H2(D3,D3))k0H_{2}(\hat{A},A)\cong(H_{2}(D^{3},\partial D^{3}))^{k}\cong 0. Let FF be the closure of A^A\hat{A}\setminus A in A^\hat{A}. We have a commutative diagram

H3(A^,A){H_{3}(\hat{A},A)}H2(A){H_{2}(A)}H3(F,F){H_{3}(F,\partial F)}H2(F){H_{2}(\partial F)}

The vertical arrow from H3(F,F)H_{3}(F,\partial F) to H3(A^,A)H_{3}(\hat{A},A) is an isomorphism because of excision. A direct computation shows that the horizontal arrow from H3(F,F)H_{3}(F,\partial F) to H2(F)H_{2}(\partial F) is an isomorphism. The vertical arrow from H2(F)H_{2}(\partial F) to H2(A)H_{2}(A) is the composition of

H2(F)H2(M)H2(A),H_{2}(\partial F)\to H_{2}(M)\to H_{2}(A),

where the first map is an isomorphism by Lemma 3.3(1), the second map is an isomorphism by Lemma 3.3(2). So the boundary map H3(A^,A)H2(A)H_{3}(\hat{A},A)\to H_{2}(A) is an isomorphism. By Lemma 3.3(3), H3(A)0H_{3}(A)\cong 0. So the exact sequence (2) becomes

0H3(A)H3(A^)H3(A^,A)H2(A)H2(A^)H2(A^,A)0.0\cong H_{3}(A)\to H_{3}(\hat{A})\to H_{3}(\hat{A},A)\xrightarrow{\cong}H_{2}(A)\to H_{2}(\hat{A})\to H_{2}(\hat{A},A)\cong 0.

Therefore, H3(A^)H2(A^)0H_{3}(\hat{A})\cong H_{2}(\hat{A})\cong 0.

Since A^\hat{A} is a connected 4-manifold with non-empty boundary, we have Hi(A^)=0H_{i}(\hat{A})=0 for all i4i\geq 4.

Since attaching 3-handles does not modify π1\pi_{1}, we know that A^\hat{A} is simply connected.

Therefore, by the Hurewicz theorem, we know that πi(A^)\pi_{i}(\hat{A}) is trivial for all ii, so A^\hat{A} is contractible. ∎

The next lemma proves the first part of Proposition 3.2.

Lemma 3.5.

M1S3M_{1}\cong S^{3}.

Proof.

By Part (1) of Lemma 3.3, we know that H1(M1)H_{1}(M_{1}) has no torsion. Therefore, M1=aP#b(P)M_{1}=aP\#b(-P), where a,ba,b are non-negative integers, P=Σ(2,3,5)P=\Sigma(2,3,5) is the Poincaré homology sphere, and (P)(-P) is the manifold PP with the reversed orientation. Now we follow an argument from [4]. By Lemma 3.4, M1M_{1} represents the trivial element of the homology cobordism group Θ3\Theta_{\mathbb{Z}}^{3}. Since there exist homomorphisms from Θ3\Theta_{\mathbb{Z}}^{3} to \mathbb{Z} that takes the Poincaré homology sphere to non-zero integers (some examples of such homomorphisms are the Froyshov invariant, the d-invariant, and Manolescu’s β\beta-invariant), we have a=ba=b. By [24, Theorem 1.7], we know that aP#a(P)aP\#a(-P) does not bound contractible smooth 4-manifolds unless a=0a=0. Hence we have M1S3M_{1}\cong S^{3}. ∎

By Lemma 3.5, we have M#kS1×S2M\cong\#^{k}S^{1}\times S^{2}. Now we prove the second part of Proposition 3.2.

Lemma 3.6.

Let M,A,BM,A,B be as above. Then AA is homeomorphic to kD2×S2\natural^{k}D^{2}\times S^{2}.

Proof.

By the previous results, A=M#kS1×S2\partial A=M\cong\#^{k}S^{1}\times S^{2}. Let Hk=kD3×S1H_{k}=\natural^{k}D^{3}\times S^{1}. Let ZZ be the smooth manifold obtained by gluing HkH_{k} with AA along a diffeomorphism of the boundaries. Since π1(Hk)π1(Hk)\pi_{1}(\partial H_{k})\to\pi_{1}(H_{k}) is surjective and π1(A)\pi_{1}(A) is trivial, the van Kampen theorem implies that π1(Z)=1\pi_{1}(Z)=1.

We claim that H2(Z)H_{2}(Z) is trivial. Suppose H2(Z)0H_{2}(Z)\neq 0, then there exists two closed oriented smooth 2-dimensional submanifolds M1M_{1}, M2M_{2} of ZZ such that M1M_{1} and M2M_{2} have a non-zero algebraic intersection number. Note that HkH_{k} is a (closed) regular neighborhood of kS1\vee_{k}S^{1} in ZZ, so one can isotope M1M_{1} and M2M_{2} so that they are disjoint from HkH_{k}. Therefore, M1M_{1} and M2M_{2} are included in AA and have a non-zero algebraic intersection number. Since AA is embedded in S4S^{4}, this is impossible.

By Freedman’s classification theorem [6, Theorem 1.5], ZZ is homeomorphic to S4S^{4}.

There is a standard embedding of HkH_{k} in 4\mathbb{R}^{4} such that the closure of its complement is a punctured kD2×S2\natural^{k}D^{2}\times S^{2}. Since ZZ is a smooth 44–manifold and HkH_{k} is a regular neighborhood of an embedded 11–dimensional complex, and since ZZ is simply connected, every smooth embedding of HkH_{k} in ZZ is smoothly isotopic to the standard embedding in a coordinate chart. Therefore, the complement of int(Hk)\operatorname{int}(H_{k}) in ZZ is diffeomorphic to (kD2×S2)#Z(\natural^{k}D^{2}\times S^{2})\#Z, which is homeomorphic to kD2×S2\natural^{k}D^{2}\times S^{2}. ∎

This finishes the proof of Proposition 3.2. Now we carry out the proof of Theorem 1.4.

Proof of Theorem 1.4.

Let M^3(X4,g)\hat{M}^{3}\subset(X^{4},g) be a two-sided stable minimal hypersurface as given by Proposition 2.1 and assume BiRicg>0\operatorname{BiRic}_{g}>0. By Lemma 3.1 and the discussions above it, each connected component of ι(M^)\iota(\hat{M}) separates S4S^{4} into two connected components, and at least one of them is simply connected. Let MM be a connected component of M^\hat{M}. Denote by λRic\lambda_{\operatorname{Ric}} the smallest eigenvalue of the Ricci tensor of MM. Then stability of MM implies that there exists a smooth function WλRicW\geq-\lambda_{\operatorname{Ric}} such that

(3) λ1(ΔW)λ>0\lambda_{1}(-\Delta-W)\geq\lambda>0

on MM, here λ\lambda is a positive lower bound of the BiRicci curvature of gg. Since the universal cover of (M,g|M)(M,g|_{M}) also satisfies (3), [22, Section 2] implies that this universal cover has diameter bounded by c/λc/\sqrt{\lambda} for some constant c>0c>0. In particular, π1(M)\pi_{1}(M) is finite and hence MM is diffeomorphic to S3/ΓS^{3}/\Gamma for some discrete subgroup Γ\Gamma of SO(4)SO(4). Thus, Lemma 3.3 implies that H1(M)H_{1}(M) is torsion free, and hence MM is diffeomorphic to either S3S^{3} or the Poincaré homology sphere. However, the latter cannot happen by Proposition 3.2. ∎

4. A black hole topology theorem

In this section we apply the proof of Theorem 1.2 and Theorem 1.4 to obtain topological control of apparent horizons (quasilocal black hole boundaries) of certain time-symmetric 44-dimensional initial data to the Einstein equations satisfying the dominant energy condition. To see its relations to our proof of Theorem 1.2, recall that such an initial data (Xn+1,g)(X^{n+1},g) necessarily satisfies that Rg0R_{g}\geq 0, and if Mn=Xn+1M^{n}=\partial X^{n+1} is an apparent horizon, then MnXn+1M^{n}\subset X^{n+1} and Xn+1X^{n+1} contains no interior minimal hypersurfaces. By [2, 7, 8], MnM^{n} is a stable minimal hypersurface and is necessarily Yamabe positive (that is, it is conformal to a PSC manifold). We are ready to prove Theorem 1.5.

Proof of Theorem 1.5.

Let EE be an asymptotically flat end of (X4,g)(X^{4},g) and suppose MM is the apparent horizon in EE. Then MM is Yamabe positive, and thus we have that

M=(#i(S3/Γi))#(#kS2×S1).M=\left(\#_{i}(S^{3}/\Gamma_{i})\right)\#\left(\#^{k}S^{2}\times S^{1}\right).

Since EE is asymptotically flat, it admits a mean convex foliation of three spheres outside a compact subset. Let YY be a leaf in this foliation, and denote by XX^{\prime} the region bounded between MM and YY. Note that XX^{\prime} satisfies the assumptions of Theorem 1.2: it is a subset of XX, which is assumed to have an embedding ι\iota into S4S^{4}, and moreover its complement in S4S^{4} contains a homeomorphic D4D^{4} (bounded by ι(Y)\iota(Y) by the locally flat Schoenflies theorem). Since by assumption, XX^{\prime} contains no interior stable minimal hypersurface that is homologous to YY, the proof of Proposition 2.1 implies that i:π1(Y)π1(X)i_{*}:\pi_{1}(Y)\to\pi_{1}(X^{\prime}) is surjective.

Consequently, we may apply the results in Section 3 to conclude that each connected component of MM is diffeomorphic to S3S^{3} or a connected sum of S2×S1S^{2}\times S^{1}’s. ∎

We remark that, other than assuming that XX is diffeomorphic to the interior of a manifold satisfying (\ast)1, the proof above is quite flexible. For instance, it suffices to assume that EE admits a mean convex foliation by mean convex three spheres outside a compact subset. In particular, any metric on EE that is C1C^{1} asypmtotic to the Euclidean metric admits such a foliation.

Moreover, we may extend this result to complete noncompact manifolds X4X^{4} with an end EE admitting a mean convex foliation of its infinity by YY, such that ι:XS4\iota:X\to S^{4} satisfies that S4ι(X)S^{4}\setminus\iota(X) has a simply connected component bounded by YY. For instance, for all k>0k\in\mathbb{Z}_{>0}, Y=#kS2×S1Y=\#^{k}S^{2}\times S^{1} bounds the simply connected region kS2×D2\natural^{k}S^{2}\times D^{2}. Such manifolds appear as certain ALF or ALG gravitational instantons. An analogous argument proves that the outermost minimal hypersurface on this end is diffeomorphic to a disjoint union of S3S^{3} or connected sums of S2×S1S^{2}\times S^{1}. We will see an explicit example of such a manifold in Section 5.3.

5. Extensions, examples and further questions

This section is devoted to several related discussions to our main results. We describe an extension of our results to construct locally minimizing hypersurfaces with controlled topology in a homeomorphic B4B^{4} with an S2S^{2} boundary. We also present a few examples and counterexamples on our main topological assumption (\ast)1: we construct a PSC embedding of the connected sums of lens spaces into a PSC S4S^{4}, illustrating the potential topological complexities of stable minimal hypersurfaces; we also discuss in detail the Dahl-Larsson construction [5] of S2×S1S^{2}\times S^{1} horizons.

5.1. Topology of the solution to Plateau problems in D4D^{4}

Consider a smooth homeomorphic closed 4-ball D4D^{4} equipped with a Riemannian metric gg. Let Σ2S3=D4\Sigma^{2}\subset S^{3}=\partial D^{4} be an embedded S2S^{2}. Assuming that D4\partial D^{4} is weakly mean convex, we may use it as a barrier and solve the Plateau problem with boundary Σ\Sigma. It is a very interesting question which 33-manifolds with boundary may appear as such a Plateau solution - we believe that this is unknown even for the standard round ball in 4\mathbb{R}^{4} (but with an arbitrary Σ=S2\Sigma=S^{2} in its boundary). We apply our results in Section 2 to obtain some preliminary topology control of at least one such minimal hypersurface.

Theorem 5.1.

Let XX be a compact smooth manifold that is homeomorphic to D4D^{4}, Σ2X\Sigma^{2}\subset\partial X is an S2S^{2}. Suppose gg is a Riemannian metric on XX such that X\partial X is gg-weakly mean convex. Then Σ\Sigma bounds a stable minimal hypersurface MM satisfying Tor(H1(M))=0\operatorname{Tor}(H_{1}(M))=0.

Proof.

We proceed as in [28]. It is known that D4Σ\partial D^{4}\setminus\Sigma has two components Y1,Y2Y_{1},Y_{2} and each YiY_{i}, i=1,2i=1,2, is diffeomorphic to D3D^{3}. Consider the space of stable minimal hypersurfaces 𝒮\mathcal{S} in (X,g)(X,g) with boundary Σ\Sigma and 3\mathcal{H}^{3}-volume bounded by 3(X)\mathcal{H}^{3}(\partial X). Then 𝒮\mathcal{S} is compact by standard curvature estimates. For an element M𝒮M\in\mathcal{S}, let ΩMX\Omega_{M}\subset X denote the region bounded by MM and Y1Y_{1}. Then

min{4(ΩM):M𝒮}\min\{\mathcal{H}^{4}(\Omega_{M}):M\in\mathcal{S}\}

is achieved by N𝒮N\in\mathcal{S}.

By a similar proof as in Proposition 2.1 (in this case, we may simply apply White’s proof [28] since Y1Y_{1} is simply connected and each M𝒮M\in\mathcal{S} is connected, see Remark 2.7), we conclude that ΩN\Omega_{N} is simply connected. Thus, applying Lemma 3.3 to ΩN=MY1\partial\Omega_{N}=M\cup Y_{1}, we conclude that H1(M)H1(MY1)H_{1}(M)\cong H_{1}(M\cup Y_{1}) is torsion free. ∎

Inspired by Theorem 1.2 and Theorem 1.4, it would be very interesting to further narrow down the possibilities of the topology of MM, perhaps with some additional curvature assumptions on gg.

Question 5.2.

Under some additional curvature assumptions on gg, prove that MM has more restrictions on its topology. For example, is it true that if gg has nonnegative Ricci curvature and X\partial X is gg-convex, we can find a Plateau minimal hypersurface MM that is diffeomorphic to D3D^{3} or (#k(S2×S1))D3\left(\#^{k}(S^{2}\times S^{1})\right)\setminus D^{3}?

5.2. A PSC embedding of connected sums of lens spaces

Given a PSC metric gg on S4S^{4}, we are interested in which 33-manifolds may appear as a stable minimal hypersurface in (S4,g)(S^{4},g). If M3(S4,g)M^{3}\subset(S^{4},g) is a stable minimal hypersurface, then necessarily M=(S3/Γi)#(#k(S2×S1))M=(S^{3}/\Gamma_{i})\#\left(\#^{k}(S^{2}\times S^{1})\right). On the other hand, it is known that not all 33-manifolds in this form admits a smooth embedding into S4S^{4}. One necessary condition that dates back to Hantzsche [10] from 1938 states that for such an embedding to exist, Tor(H1(M))=GG\operatorname{Tor}(H_{1}(M))=G\oplus G for some abelian group GG (we again refer the readers to the survey article by Hillman [12] for the collection of classical results of the embedding problem). The most basic example of such a 33-manifold with nontrivial Tor(H1(M))\operatorname{Tor}(H_{1}(M)) is the connected sum of two lens spaces, L(p,q)#L(p,q)L(p,q)\#-L(p,q) when pp is odd, as constructed by Zeeman [29]. Below we show that such manifold does admit a PSC embedding into a PSC S4S^{4}.

Let S3={(z,w)2|z|2+|w|2=1}S^{3}=\{(z,w)\in\mathbb{C}^{2}\mid|z|^{2}+|w|^{2}=1\}. Let p0p\neq 0 and p,qp,q be coprime. Consider the action of /p\mathbb{Z}/p\mathbb{Z} on S3S^{3} generated by

f:(z,w)(e1p2πiz,eqp2πiw).f:(z,w)\mapsto(e^{\frac{1}{p}\cdot{2\pi i}}z,e^{\frac{q}{p}\cdot 2\pi i}w).

Then the action is free, and the quotient manifold is the lens space L(p,q)L(p,q). Let \sim denote the equivalence relation on S3S^{3} induced by the action by ff. Denote by τ:S3S3\tau:S^{3}\to S^{3} the conjugation map defined by τ(z,w)=(z¯,w¯)\tau(z,w)=(\bar{z},\bar{w}).

The next lemma reinterprets a construction of Schubert [21] so that we can keep track of the curvature in Zeeman’s construction.

Lemma 5.3.

τ\tau induces an involution on L(p,q)L(p,q). The quotient space with respect to this involution is S3S^{3}, and the quotient map is a double branched cover. When pp is odd, the branching locus is a smooth knot.

Proof.

It is straightforward to check that fτ=τf1f\circ\tau=\tau\circ f^{-1}, so τ\tau induces an involution on (S3/)=L(p,q)(S^{3}/\sim)=L(p,q).

Let T1={(z,w)S3|z||w|}T_{1}=\{(z,w)\in S^{3}\mid|z|\geq|w|\}, T2={(z,w)S3|w||z|}T_{2}=\{(z,w)\in S^{3}\mid|w|\geq|z|\}. Then f(T1)=T1f(T_{1})=T_{1}, f(T2)=T2f(T_{2})=T_{2}. A fundamental domain of the action of ff on T1T_{1} is

F1:={(z,w)S3arg(z)[0,2π/p],|z||w|}[0,2π/p]×D2,F_{1}:=\{(z,w)\in S^{3}\mid arg(z)\in[0,2\pi/p],|z|\geq|w|\}\cong[0,2\pi/p]\times D^{2},

where D2D^{2} is identified with the unit disk in \mathbb{C}, and the above diffeomorphism is defined by

(z,w)(arg(z),2w).(z,w)\mapsto(arg(z),\sqrt{2}w).

The gluing map on F1[0,2π/p]×D2\partial F_{1}\cong\partial[0,2\pi/p]\times D^{2} is given by

(0,x)(2π/p,eqp2πix)(0,x)\sim(2\pi/p,e^{\frac{q}{p}\cdot 2\pi i}x)

for all xD2x\in D^{2}, so T1/T_{1}/\sim is a solid torus S1×D2S^{1}\times D^{2}.

We write down a diffeomorphism between T1/T_{1}/\sim and S1×D2S^{1}\times D^{2} explicitly. Identify S1S^{1} with the unit circle in \mathbb{C}. Let φ¯:F1S1×D2\bar{\varphi}:F_{1}\to S^{1}\times D^{2} be defined by

φ¯(θ,x)=(eipθ,eiqθx),\bar{\varphi}(\theta,x)=(e^{ip\theta},e^{-iq\theta}x),

then φ¯\bar{\varphi} induces a diffeomorphism φ:(T1/)S1×D2\varphi:(T_{1}/\sim)\to S^{1}\times D^{2}.

The map τ\tau induces an involution on T1/T_{1}/\sim. The map is described on the fundamental domain as

τ|T1/:[0,2π/p]×D2/\displaystyle\tau|_{T_{1}/\sim}:[0,2\pi/p]\times D^{2}/\sim [0,2π/p]×D2/\displaystyle\to[0,2\pi/p]\times D^{2}/\sim
(θ,x)\displaystyle(\theta,x) (2π/pθ,eqp2πix¯).\displaystyle\mapsto(2\pi/p-\theta,e^{\frac{q}{p}\cdot 2\pi i}\bar{x}).

The fixed point set of τ|T1/\tau|_{T_{1}/\sim} consists of two arcs, represented in the fundamental domain by

(4) {π/p}×{reqpπir[1,1]}and{0}×{rr[1,1]}.\{\pi/p\}\times\{re^{\frac{q}{p}\cdot\pi i}\mid r\in[-1,1]\}\quad\text{and}\quad\{0\}\times\{r\mid r\in[-1,1]\}.

Note that

φ(τ|T1/)φ1(eipθ,eiqθx)\displaystyle\varphi\circ(\tau|_{T_{1}/\sim})\circ\varphi^{-1}(e^{ip\theta},e^{-iq\theta}x) =(eip(2π/pθ),eiq((2π/pθ)eqp2πix¯)\displaystyle=(e^{ip(2\pi/p-\theta)},e^{-iq((2\pi/p-\theta)}e^{\frac{q}{p}\cdot 2\pi i}\bar{x})
=(eipθ,eiqθx¯).\displaystyle=(e^{-ip\theta},e^{iq\theta}\bar{x}).

So the action of φ(τ|T1/)φ1\varphi\circ(\tau|_{T_{1}/\sim})\circ\varphi^{-1} on S1×D2S^{1}\times D^{2} is the product of two reflections on S1S^{1} and D2D^{2}. Therefore, the quotient space is homeomorphic to D3D^{3} and the quotient map is a double branched cover.

Similarly, the involution τ\tau defines a double branched cover from T2/T_{2}/\sim to D3D^{3}. As a result, the quotient map L(p,q)L(p,q)/τL(p,q)\to L(p,q)/\tau is a double branched cover and we have L(p,q)/τS3L(p,q)/\tau\cong S^{3}.

The branching locus is the image of four arcs on L(p,q)L(p,q), which are given by (4) and a similar formula on T2/T_{2}/\sim. To write down the formula for the branching locus on T2/T_{2}/\sim, let

F2:={(z,w)S3arg(w)[0,2π/p],|w||z|}D2×[0,2π/p],F_{2}:=\{(z,w)\in S^{3}\mid arg(w)\in[0,2\pi/p],|w|\geq|z|\}\cong D^{2}\times[0,2\pi/p],

be a fundamental domain of T2/T_{2}/\sim, where the diffeomorphism is given by (z,w)(2z,arg(w))(z,w)\mapsto(\sqrt{2}z,\arg(w)). Let qq^{\prime} be an integer such that pqq1p\mid qq^{\prime}-1. Then the branching locus on T2/T_{2}/\sim are given by the following two arcs on F2F_{2}:

(5) {reqpπir[1,1]}×{π/p}and{rr[1,1]}×{0}.\{re^{\frac{q^{\prime}}{p}\cdot\pi i}\mid r\in[-1,1]\}\times\{\pi/p\}\quad\text{and}\quad\{r\mid r\in[-1,1]\}\times\{0\}.

The endpoints of the four arcs have the following preimages in S3S^{3}:

  1. (1)

    The endpoints of {π/p}×{reqpπir[1,1]}\{\pi/p\}\times\{re^{\frac{q}{p}\cdot\pi i}\mid r\in[-1,1]\} in (4) lift to 22(eπi/p,±eqpπi)\frac{\sqrt{2}}{2}(e^{\pi i/p},\pm e^{\frac{q}{p}\cdot\pi i}).

  2. (2)

    The endpoints of {0}×{rr[1,1]}\{0\}\times\{r\mid r\in[-1,1]\} in (4) lift to 22(1,±1)\frac{\sqrt{2}}{2}(1,\pm 1).

  3. (3)

    The endpoints of {reqpπir[1,1]}×{π/p}\{re^{\frac{q^{\prime}}{p}\cdot\pi i}\mid r\in[-1,1]\}\times\{\pi/p\} in (5) lift to 22(±eqpπi,eπi/p)\frac{\sqrt{2}}{2}(\pm e^{\frac{q^{\prime}}{p}\cdot\pi i},e^{\pi i/p}).

  4. (4)

    The endpoints of {rr[1,1]}×{0}\{r\mid r\in[-1,1]\}\times\{0\} in (5) lift to 22(±1,1)\frac{\sqrt{2}}{2}(\pm 1,1).

Since pp is odd, we may assume without loss of generality that both qq^{\prime} and qq are odd. Then 2pqq12p\mid qq^{\prime}-1. In the following, we abuse notation and extend the action of ff to 2\mathbb{C}^{2} by the same formula. Then

f(q1)/2(eπi/p,eqpπi)=(eqpπi,eqpπieq12qp2πi)=(eqpπi,eπi/p),f^{(q^{\prime}-1)/2}(e^{\pi i/p},e^{\frac{q}{p}\cdot\pi i})=(e^{\frac{q^{\prime}}{p}\cdot\pi i},e^{\frac{q}{p}\cdot\pi i}e^{\frac{q^{\prime}-1}{2}\frac{q}{p}\cdot 2\pi i})=(e^{\frac{q^{\prime}}{p}\cdot\pi i},e^{\pi i/p}),
f(p1)/2(eπi/p,eqpπi)=(eppπi,eqpπiep12qp2πi)=(1,1),f^{(p-1)/2}(e^{\pi i/p},-e^{\frac{q}{p}\cdot\pi i})=(e^{\frac{p}{p}\cdot\pi i},-e^{\frac{q}{p}\cdot\pi i}e^{\frac{p-1}{2}\frac{q}{p}\cdot 2\pi i})=(-1,1),
f(p+q)/2(1,1)=(ep+qpπi,eq(p+q)pπi)=(eqpπi,eπi/p).f^{(p+q^{\prime})/2}(1,-1)=(e^{\frac{p+q^{\prime}}{p}\cdot\pi i},-e^{\frac{q(p+q^{\prime})}{p}\cdot\pi i})=(-e^{\frac{q^{\prime}}{p}\cdot\pi i},e^{\pi i/p}).

As a result, the four arcs above connect to form a smooth knot in L(p,q)L(p,q). ∎

We are now ready to state our main construction. We show that Zeeman’s construction can be performed preserving the PSC conditions.

Proposition 5.4.

Let p,qp,q be coprime integers and pp is odd. There exists an embedding of M=L(p,q)#L(p,q)M=L(p,q)\#-L(p,q) into S4S^{4} equipped with a Riemannian metric gg, such that gg has positive scalar curvature, and the induced metric g|Mg|_{M} has positive scalar curvature.

Proof.

Let N=L(p,q)N=L(p,q) be a lens space with pp odd, and let τ¯:L(p,q)L(p,q)\bar{\tau}:L(p,q)\to L(p,q) be the involution given by Lemma 5.3. Equip L(p,q)L(p,q) with the standard round metric g0g_{0} inherited from the round S3S^{3}. Lemma 5.3 implies that τ¯\bar{\tau} is an isometry of g0g_{0}. Let xL(p,q)x\in L(p,q) be a point on the branching locus of τ\tau, and for small ρ>0\rho>0, Dρ(x)D_{\rho}(x) be the geodesic ball at xx. We now fix ρ>0\rho>0 sufficiently small, so we may apply the classical result of Gromov-Lawson [9] (the point {x}\{x\} has codimension 33 in NN), and deform the metric g0g_{0} in Dρ{x}D_{\rho}\setminus\{x\} to another PSC metric g1g_{1}, such that g1g_{1} is a product metric in Dρ/2{x}D_{\rho/2}\setminus\{x\}, while keeping τ\tau an isometry of g1g_{1}. Let

Mτ=(NDρ(x))×[0,1]/(y,0)(τ(y),1)M_{\tau}=\left(N\setminus D_{\rho}(x)\right)\times[0,1]/(y,0)\sim(\tau(y),1)

be the mapping torus of τ\tau on NDρ(x)N\setminus D_{\rho}(x). The product metric g1+dt2g_{1}+dt^{2} on (NDρ(x))×[0,1](N\setminus D_{\rho}(x))\times[0,1] induces a metric g^\hat{g} with positive scalar curvature and is product near Mτ=S2×S1\partial M_{\tau}=S^{2}\times S^{1}. Thus, we may glue onto MτM_{\tau} a S2×D2S^{2}\times D^{2} with positive scalar curvature, equipped with a product metric near its boundary. The gluing is defined by a fiber preserving diffeomorphism φ\varphi of S2×S1S^{2}\times S^{1}. A classical result of Zeeman222In fact, Zeeman’s theorem states that the same result holds for every cyclic branched cover of S3S^{3} along a knot. [29] states that by suitably choosing φ\varphi, the result

Mτφ(S2×D2)M_{\tau}\cup_{\varphi}(S^{2}\times D^{2})

is diffeomorphic to the standard S4S^{4}. Note that the metrics glue smoothly into gg with positive scalar curvature.

Taking any t[0,1]t\in[0,1] in the [0,1][0,1]–coordinate of MτM_{\tau}, we obtain a smooth PSC embedding of

L(p,q)Dρ(x)Mτ(S4,g)L(p,q)\setminus D_{\rho}(x)\hookrightarrow M_{\tau}\hookrightarrow(S^{4},g)

such that the induced metric is product near (L(p,q)Dρ(x))\partial(L(p,q)\setminus D_{\rho}(x)). Therefore, by taking the boundary of a small tubular neighborhood of L(p,q)Dρ(x)L(p,q)\setminus D_{\rho}(x) (see, e.g. [9, Theorem 5.7]) – which is diffeomorphic to M=L(p,q)#L(p,q)M=L(p,q)\#-L(p,q) – we obtain a smooth embedding of MM into (S4,g)(S^{4},g) such that the induced metric has positive scalar curvature. ∎

Proposition 5.4 motivates the following natural question.

Question 5.5.

Construct a stable minimal embedding of M=L(p,q)#L(p,q)M=L(p,q)\#-L(p,q) into some PSC Riemannian (S4,g)(S^{4},g).

The example constructed in Proposition 5.4 is minimal (in fact, the metric is a product nearby) everywhere except for the boundary of the tubular neighborhood of (L(p,q)Dρ(x))\partial(L(p,q)\setminus D_{\rho}(x)), where the mean curvature is large. With some trivial modifications of this construction, MM can be made everywhere mean convex.

5.3. The Dahl-Larsson example

In [5], Dahl-Larsson constructed, for all n3n\geq 3, an asymptotically flat manifold (Xn,g)(X^{n},g) with zero scalar curvature such that its apparent horizon is diffeomorphic to the unit normal bundle of a codimension 3\geq 3 submanifold NN. We briefly discuss their discussion in the case we consider in our paper.

Let δ\delta be the flat metric on 4\mathbb{R}^{4}, γ4\gamma\subset\mathbb{R}^{4} be an embedded curve. Consider the Green’s function

G(x)=γ|xy|2𝑑yG(x)=\int_{\gamma}|x-y|^{-2}dy

with poles along γ\gamma. It is well-known that GG satisfies the asymptotics

G(x)A|x|δ2,|x|;G(x)Bdistδ(x,γ),xγ,G(x)\sim\frac{A}{|x|_{\delta}^{2}},~{}~{}|x|\to\infty;\quad G(x)\sim\frac{B}{\operatorname{dist}_{\delta}(x,\gamma)},~{}~{}x\to\gamma,

For some constants A,BA,B. For ε>0\varepsilon>0, consider the conformally deformed metric

gε=(1+εG)2δ.g_{\varepsilon}=(1+\varepsilon G)^{2}\delta.

The manifold (4,gε)(\mathbb{R}^{4},g_{\varepsilon}) is complete, noncompact and has two ends: E1E_{1} near the infinity of 4\mathbb{R}^{4}, and E2E_{2} near γ\gamma. For each ε>0\varepsilon>0, gεg_{\varepsilon} has zero scalar curvature since Δδuε=0\Delta_{\delta}u_{\varepsilon}=0. gεg_{\varepsilon} on E1E_{1} is asymptotically flat. gδg_{\delta} on E2E_{2} enjoys a similar asymptotic decay. To see this, let ss be the arclength coordinate of γ\gamma, tt the distance function to γ\gamma. Then locally we may write the metric δ=ds2+dt2+t2gs,t\delta=ds^{2}+dt^{2}+t^{2}g_{s,t}, where gs,tg_{s,t} is a metric on the unit S2S^{2} converging smoothly to the round metric as s,t0s,t\to 0. Setting r=logtr=-\log t, we see that

gε\displaystyle g_{\varepsilon} =(1+εer)2(ds2+e2rdr2+e2rgs,r)\displaystyle=(1+\varepsilon e^{r})^{2}(ds^{2}+e^{-2r}dr^{2}+e^{-2r}g_{s,r})
=(er+ε)2(e2rds2+dr2+gS2)+o(1).\displaystyle=(e^{-r}+\varepsilon)^{2}\left(e^{2r}ds^{2}+dr^{2}+g_{S^{2}}\right)+o(1).

Therefore, gεg_{\varepsilon} is asymptotic to the product metric on S2×2S^{2}\times\mathbb{H}^{2} on E2E_{2}. In particular, near the infinity of E2E_{2}, it admits a mean convex foliation by S2×S1S^{2}\times S^{1}.

Theorem 5.6 ([5, Theorem 1.1]).

For sufficiently small ε>0\varepsilon>0, it holds that the outermost apparent horizon MεM_{\varepsilon} of (4,gε)(\mathbb{R}^{4},g_{\varepsilon}) in E1E_{1} is diffeomorphic to S2×S1S^{2}\times S^{1}. In fact MεM_{\varepsilon} is the graph of a smooth function on the unit normal bundle in normal coordinates for γ\gamma.

The proof of Theorem 5.6 is delicate. They constructed mean convex barriers and forced the apparent horizon in E1E_{1} to lie between the tubular hypersurfaces around γ\gamma of δ\delta-distance CinnerεC_{inner}\varepsilon and CouterεC_{outer}\varepsilon, provided that ε>0\varepsilon>0 is sufficiently small.

We observe that the same proof in fact also proves the following simpler statement:

Proposition 5.7.

For sufficiently small ε>0\varepsilon>0, the outermost minimal hypersurface on E2E_{2} is diffeomorphic to S2×S1S^{2}\times S^{1}, and is the graph of a smooth function on the unit normal bundle in normal coordinates for γ\gamma.

The Dahl-Larsson example naturally leads to the following question.

Question 5.8.

For each integer l>1l>1, construct an asymptotically flat manifold (4γ,g)(\mathbb{R}^{4}\setminus\gamma,g) with Rg0R_{g}\geq 0 such that the outermost minimal hypersurface is diffeomorphic to #l(S2×S1)\#^{l}(S^{2}\times S^{1}).

References

  • [1] Simon Brendle, Sven Hirsch, and Florian Johne, A generalization of Geroch’s conjecture, Comm. Pure Appl. Math. 77 (2024), no. 1, 441–456. MR 4666629
  • [2] Mingliang Cai and Gregory J. Galloway, On the topology and area of higher-dimensional black holes, Classical Quantum Gravity 18 (2001), no. 14, 2707–2718. MR 1846368
  • [3] Otis Chodosh, Chao Li, Paul Minter, and Douglas Stryker, Stable minimal hypersurfaces in 𝐑5\mathbf{R}^{5}, arXiv:2401.01492 (2025).
  • [4] Otis Chodosh, Davi Maximo, and Anubhav Mukherjee, Complete Riemannian 4-manifolds with uniformly positive scalar curvature, arXiv:2407.05574 (2024).
  • [5] Mattias Dahl and Eric Larsson, Outermost apparent horizons diffeomorphic to unit normal bundles, Asian J. Math. 23 (2019), no. 6, 1013–1040. MR 4136488
  • [6] Michael Hartley Freedman, The topology of four-dimensional manifolds, Journal of Differential Geometry 17 (1982), no. 3, 357–453.
  • [7] Gregory J. Galloway, Rigidity of marginally trapped surfaces and the topology of black holes, Comm. Anal. Geom. 16 (2008), no. 1, 217–229. MR 2411473
  • [8] Gregory J. Galloway and Richard Schoen, A generalization of Hawking’s black hole topology theorem to higher dimensions, Comm. Math. Phys. 266 (2006), no. 2, 571–576. MR 2238889
  • [9] Mikhael Gromov and H. Blaine Lawson, Jr., Spin and scalar curvature in the presence of a fundamental group. I, Ann. of Math. (2) 111 (1980), no. 2, 209–230. MR 569070
  • [10] W. Hantzsche, Einlagerung von Mannigfaltigkeiten in euklidische Räume, Math. Z. 43 (1938), no. 1, 38–58. MR 1545714
  • [11] Stephen W. Hawking and George F. R. Ellis, The large scale structure of space-time, anniversary ed., Cambridge Monographs on Mathematical Physics, Cambridge University Press, Cambridge, 2023, With a foreword by Abhay Ashtekar. MR 4615777
  • [12] J. A. Hillman, Locally flat embeddings of 3-manifolds in S4S^{4}, arXiv:2408.10535 (2024).
  • [13] Marcus A. Khuri and Jordan F. Rainone, Black lenses in Kaluza-Klein matter, Phys. Rev. Lett. 131 (2023), no. 4, Paper No. 041402, 7. MR 4630791
  • [14] Laurent Mazet, Stable minimal hypersurfaces in 𝐑6\mathbf{R}^{6}, arXiv:2405.14676 (2024).
  • [15] William Meeks, III, Leon Simon, and Shing Tung Yau, Embedded minimal surfaces, exotic spheres, and manifolds with positive Ricci curvature, Ann. of Math. (2) 116 (1982), no. 3, 621–659. MR 678484
  • [16] William H. Meeks, III and Shing Tung Yau, Topology of three-dimensional manifolds and the embedding problems in minimal surface theory, Ann. of Math. (2) 112 (1980), no. 3, 441–484. MR 595203
  • [17] J. Sacks and K. Uhlenbeck, The existence of minimal immersions of 22-spheres, Ann. of Math. (2) 113 (1981), no. 1, 1–24. MR 604040
  • [18] R. Schoen, L. Simon, and S. T. Yau, Curvature estimates for minimal hypersurfaces, Acta Math. 134 (1975), no. 3-4, 275–288. MR 423263
  • [19] R. Schoen and Shing Tung Yau, Existence of incompressible minimal surfaces and the topology of three-dimensional manifolds with nonnegative scalar curvature, Ann. of Math. (2) 110 (1979), no. 1, 127–142. MR 541332
  • [20] Richard Schoen and Leon Simon, Regularity of stable minimal hypersurfaces, Comm. Pure Appl. Math. 34 (1981), no. 6, 741–797. MR 634285
  • [21] Horst Schubert, Knoten mit zwei brücken, Mathematische Zeitschrift 65 (1956), no. 1, 133–170.
  • [22] Ying Shen and Rugang Ye, On stable minimal surfaces in manifolds of positive bi-Ricci curvatures, Duke Math. J. 85 (1996), no. 1, 109–116. MR 1412440
  • [23] Leon Simon, A strict maximum principle for area minimizing hypersurfaces, J. Differential Geom. 26 (1987), no. 2, 327–335. MR 906394
  • [24] Clifford Henry Taubes, Gauge theory on asymptotically periodic 44-manifolds, Journal of Differential Geometry 25 (1987), no. 3, 363–430.
  • [25] Zhihan Wang, Min-max minimal hypersurfaces with obstacle, Calc. Var. Partial Differential Equations 61 (2022), no. 5, Paper No. 175, 26. MR 4447247
  • [26] Brian White, Existence of least-area mappings of NN-dimensional domains, Ann. of Math. (2) 118 (1983), no. 1, 179–185. MR 707165
  • [27] by same author, Mappings that minimize area in their homotopy classes, J. Differential Geom. 20 (1984), no. 2, 433–446. MR 788287
  • [28] by same author, On the topological type of minimal submanifolds, Topology 31 (1992), no. 2, 445–448. MR 1167182
  • [29] E Christopher Zeeman, Twisting spun knots, Transactions of the American Mathematical Society 115 (1965), 471–495.