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 manifolds with positive intermediate curvature

Liam Mazurowski Department of Mathematics, Lehigh University, Bethlehem, Pennsylvania, 18015, United States [email protected] Tongrui Wang School of Mathematical Sciences, Shanghai Jiao Tong University, Minhang District, Shanghai 200240, China [email protected]  and  Xuan Yao Department of Mathematics, Cornell University, Ithaca, New York, 14853, United States [email protected]
Abstract.

We formulate a conjecture relating the topology of a manifold’s universal cover with the existence of metrics with positive mm-intermediate curvature. We prove the result for manifolds of dimension n{3,4,5}n\in\{3,4,5\} and for most choices of mm when n=6n=6. As a corollary, we show that a closed, aspherical 6-manifold cannot admit a metric with positive 44-intermediate curvature.

1. Introduction

A manifold MM is called aspherical if the universal cover of MM is contractible. Equivalently, MM is aspherical if πk(M)=0\pi_{k}(M)=0 for k2k\geq 2. Thus an aspherical manifold MM is a K(π,1)K(\pi,1) space for π=π1(M)\pi=\pi_{1}(M). Schoen and Yau [16] conjectured that a closed aspherical manifold cannot carry a metric of positive scalar curvature; also see Gromov [6]. This K(π,1)K(\pi,1) conjecture is known to be true when n=3n=3 by work of Schoen-Yau [15] and Gromov-Lawson [7], when n=4n=4 by Chodosh-Li [3], and when n=5n=5 by Chodosh-Li [3] and independently by Gromov [10].

Theorem 1.

Assume that MnM^{n} is a closed K(π,1)K(\pi,1) manifold of dimension n{3,4,5}n\in\{3,4,5\}. Then MM does not admit a metric with positive scalar curvature.

Let TnT^{n} be the nn-dimensional torus. The Geroch conjecture states that TnT^{n} does not admit a metric of positive scalar curvature. This can be seen as a special case of the K(π,1)K(\pi,1) conjecture. The Geroch conjecture was resolved by Schoen-Yau [15] for n7n\leq 7 and by Gromov-Lawson [8] in all dimensions. In a different direction, Brendle-Hirsch-Johne [2] have generalized the Geroch conjecture to show that a closed manifold Mn=Nnm×TmM^{n}=N^{n-m}\times T^{m} does not admit a metric with positive mm-intermediate curvature when n7n\leq 7. Here the mm-intermediate curvature (see Definition 8) interpolates between the Ricci curvature (m=1m=1) and the scalar curvature (m=n1m=n-1).

Theorem 2 (Brendle-Hirsch-Johne [2]).

A closed manifold Mn=Nnm×TmM^{n}=N^{n-m}\times T^{m} does not admit a metric with positive mm-intermediate curvature when n7n\leq 7.

In this paper, we propose the following conjecture which relates the topology of a manifold’s universal cover with the existence of metrics with positive mm-intermediate curvature. We note that both Theorem 1 and Theorem 2 occur as special cases of this conjecture.

Conjecture 3.

Let MnM^{n} be a closed manifold of dimension 3n73\leq n\leq 7. Assume that the universal cover M¯\overline{M} of MM satisfies

Hn(M¯,)=Hn1(M¯,)==Hnm+1(M¯,)=0.H_{n}(\overline{M},\mathbb{Z})=H_{n-1}(\overline{M},\mathbb{Z})=\ldots=H_{n-m+1}(\overline{M},\mathbb{Z})=0.

Then MM does not admit a metric with positive mm-intermediate curvature.

As our main result, we show that Conjecture 3 is true for n{3,4,5}n\in\{3,4,5\} and for most choices of mm when n=6n=6.

Theorem 4 (Main Theorem).

Suppose MnM^{n} is a closed manifold and that the universal cover M¯\overline{M} of MM satisfies

Hn(M¯,)=Hn1(M¯,)==Hnm+1(M¯,)=0.H_{n}(\overline{M},\mathbb{Z})=H_{n-1}(\overline{M},\mathbb{Z})=\ldots=H_{n-m+1}(\overline{M},\mathbb{Z})=0.

Then MM does not admit a metric of positive mm-intermediate curvature in any of the following cases: n{3,4,5}n\in\{3,4,5\} and m{1,2,,n1}m\in\{1,2,\dots,n-1\}; n=6n=6 and m{1,2,3,4}m\in\{1,2,3,4\}; n7n\geq 7 and m=1m=1.

In dimension n=6n=6, the 55-intermediate curvature is equivalent to the scalar curvature, and positive 44-intermediate curvature is a slight strengthening of positive scalar curvature. As a corollary of the n=6n=6 and m=4m=4 case of our main theorem, we see that a closed, aspherical 66-manifold cannot admit a metric with positive 44-intermediate curvature. This provides some evidence for the K(π,1)K(\pi,1) conjecture in dimension 6.

Corollary 5.

A closed, aspherical 66-manifold does not admit a metric with positive 4-intermediate curvature.

Finally, we note that Chodosh-Li-Liokumovich [4] have proven a mapping version of the K(π,1)K(\pi,1) conjecture and a reformulation of the K(π,1)K(\pi,1) conjecture as a classification theorem. We show that corresponding results also hold in our setting.

Theorem 6.

Suppose MnM^{n} is a closed manifold and that the universal cover M¯\overline{M} of MM satisfies

Hn(M¯,)=Hn1(M¯,)==Hnm+1(M¯,)=0.H_{n}(\overline{M},\mathbb{Z})=H_{n-1}(\overline{M},\mathbb{Z})=\ldots=H_{n-m+1}(\overline{M},\mathbb{Z})=0.

Assume further that NnN^{n} is a closed manifold with a non-zero degree map to MnM^{n}. Then NN does not admit a metric of positive mm-intermediate curvature in any of the following cases: n{3,4,5}n\in\{3,4,5\} and m{1,2,,n1}m\in\{1,2,\dots,n-1\}; n=6n=6 and m{1,2,3,4}m\in\{1,2,3,4\}; n7n\geq 7 and m=1m=1.

Theorem 7.

Suppose MnM^{n} is a closed manifold which admits positive mm-intermediate curvature. Suppose further that

  1. (1)

    n=5n=5, m=2m=2, and π2(M5)=0\pi_{2}(M^{5})=0; or

  2. (2)

    n=6n=6, m=4m=4, and π2(M6)=π3(M6)=π4(M6)=0\pi_{2}(M^{6})=\pi_{3}(M^{6})=\pi_{4}(M^{6})=0.

Then a finite covering of NN is homeomorphic to SnS^{n} or connected sums of Sn1×S1S^{n-1}\times S^{1}.

1.1. Further Discussion and Proof Ideas

Schoen and Yau’s proof [15] of the Geroch conjecture uses minimal hypersurfaces and an inductive descent argument. A key step in the argument is to show that if (M,g)(M,g) has positive scalar curvature and Σ\Sigma is an area minimizing hypersurface in MM then Σ\Sigma admits a conformal metric with positive scalar curvature. In fact, the necessary conformal factor is a suitable power of the first eigenfunction for the second variation of area.

Brendle, Hirsch, and Johne [2] proved a generalization of the Geroch conjecture to closed manifolds of the form Mn=Nnm×TmM^{n}=N^{n-m}\times T^{m}. To do so, they introduced a new curvature condition called mm-intermediate curvature.

Definition 8.

Suppose (Mn,g)(M^{n},g) is Riemannian manifold. Given a collection of orthonormal vectors {e1,,em}\{e_{1},\cdots,e_{m}\} at a point pMp\in M, let {e1,em,em+1,,en}\{e_{1},\cdots e_{m},e_{m+1},\ldots,e_{n}\} be an extension to an orthonormal basis of TpMT_{p}M. The mm-intermediate curvature CmC_{m} of the orthonormal vectors {e1,,em}\{e_{1},\cdots,e_{m}\} is defined by

Cm(e1,,em):=p=1mq=m+1nRM(ep,eq,ep,eq).C_{m}(e_{1},\cdots,e_{m}):=\sum_{p=1}^{m}\sum_{q=m+1}^{n}R_{M}(e_{p},e_{q},e_{p},e_{q}).

We say that (M,g)(M,g) has positive mm-intermediate curvature if for any choice of orthonormal vectors {e1,,em}\{e_{1},\cdots,e_{m}\} at any point pMp\in M, we have Cm(e1,,em)>0C_{m}(e_{1},\cdots,e_{m})>0.

We note that the 1-intermediate curvature is precisely the Ricci curvature and that the (n1)(n-1)-intermediate curvature is equal to the scalar curvature up to a constant factor. The 2-intermediate curvature had previously been introduced and studied by Shen and Ye [17] as bi-Ricci curvature.

Brendle, Hirsch, and Johne’s main result says that Mn=Nnm×TmM^{n}=N^{n-m}\times T^{m} cannot admit a metric with positive mm-intermediate curvature when n7n\leq 7. Their proof is based on the fact that certain weighted minimal slicings do not exist in manifolds with positive mm-intermediate curvature. In the construction of their weighted slicings, each successive weight ρk+1\rho_{k+1} is obtained from the previous weight ρk\rho_{k} by multiplying by the first eigenfunction for the second variation of ρk\rho_{k}-weighted area. This is reminiscent of the Schoen-Yau descent argument. In dimension n>7n>7, Xu [19] constructed interesting counter-examples to Brendle-Hirsch-Johne’s generalization of the Geroch Conjecture, which is the reason why we only expect Conjecture 3 to be true for 3n73\leq n\leq 7.

Recall that the Geroch conjecture is a particular case of the K(π,1)K(\pi,1) conjecture. In their 1987 paper, Schoen and Yau [16] put forth an outline for how to prove the K(π,1)K(\pi,1) conjecture in dimension n=4n=4. Chodosh and Li’s proof [3] is inspired by this outline. Let MnM^{n} be a closed, aspherical manifold. Assume for contradiction that MM has a metric with positive scalar curvature. Working in the universal cover M¯\overline{M} of MM, Chodosh and Li construct a geodesic line σ\sigma and a closed null-homologous submanifold Λn2\Lambda^{n-2} which is far away from σ\sigma but linked with σ\sigma. Let Σ1\Sigma_{1} be the area minimizing filling of Λ\Lambda. Using a clever choice of prescribed mean curvature functional, Chodosh and Li construct a μ\mu-bubble Σ2Σ1\Sigma_{2}\subset\Sigma_{1} such that Λ\Lambda is homologous to Σ2\Sigma_{2} in Σ1\Sigma_{1} and Σ2\Sigma_{2} lies far away from σ\sigma. When n=4n=4, the μ\mu-bubble Σ2\Sigma_{2} is 2-dimensional and it is possible to show with a second variation argument that each component of Σ2\Sigma_{2} has uniformly bounded diameter. Thus, by a general property of universal covers, it is possible to fill each component of Σ2\Sigma_{2} within a fixed sized neighborhood in M¯\overline{M}. In this manner, one obtains a new filling of Λ\Lambda which does not intersect σ\sigma and this is a contradiction.

When n=5n=5, new difficulties arise because the μ\mu-bubble Σ2\Sigma_{2} is 3-dimensional and may not have a diameter bound. To overcome this, Chodosh and Li developed the so-called slice-and-dice argument. They first slice Σ2\Sigma_{2} open along suitable μ\mu-bubbles Σ3,1,,Σ3,k\Sigma_{3,1},\ldots,\Sigma_{3,k} to obtain a manifold Σ^2\hat{\Sigma}_{2} with simple 2 dimensional homology. Then they dice Σ^2\hat{\Sigma}_{2} into pieces with small diameter using free boundary μ\mu-bubbles. These free boundary μ\mu-bubbles are two dimensional, so it is possible to use a second variation argument together with the Gauss-Bonnet theorem to show that each dicing surface is a disk. Crucially, the boundary of a disk is connected, and this gives a simple combinatorial structure to the output of the slice and dice procedure. This can then be used to construct a filling of Σ2\Sigma_{2} within a fixed sized neighborhood in M¯\overline{M} as before.

Our proof of Theorem 4 is based on the argument of Chodosh and Li. The main observation is that positive mm-intermediate curvature should imply a diameter bound on the slice Σm1\Sigma_{m-1} in low dimensions. A key tool in this regard is the generalized Bonnet-Myers theorem discovered by Shen and Ye [18]. We will apply the generalized Bonnet-Myers theorem with a weight equal to a product of eigenfunctions for the stability operators of certain weighted area and μ\mu-bubble functionals. We remark that the case n=6n=6 and m=3m=3 is very delicate and requires the full strength of the Shen-Ye generalized Bonnet-Myers theorem as well as the use of a carefully chosen weighted μ\mu-bubble functional.

To handle the case n=6n=6 and m=4m=4, we adapt the slice and dice procedure of Chodosh and Li. Since the dicing submanifolds are now 3-dimensional, we can no longer use the Gauss-Bonnet theorem to determine their topology. Nevertheless, we can prove the following Frankel type theorem which implies that the dicing submanifolds have connected boundary:

Theorem 9.

Let Σ\Sigma be a 33-dimensional compact Riemannian manifold with boundary. Assume Σ\Sigma admits a positive function ff which satisfies

Ric(v,v)f1Δf+12|lnf|20\operatorname{Ric}(v,v)-f^{-1}\Delta f+\frac{1}{2}|\nabla\ln f|^{2}\geq 0

for all unit vectors vv. Moreover, suppose that ηf=fHΣ\partial_{\eta}f=-fH_{\partial\Sigma} where η\eta is the unit outward normal to Σ\partial\Sigma. Then Σ\partial\Sigma is connected.

The fact that the dicing submanifolds have connected boundary is then enough to carry through the remainder of the argument.

Remark 10.

The hypothesis of Shen and Ye’s theorem is dimension dependent, and we were not able to verify the hypothesis on the slice Σm1\Sigma_{m-1} when n=7n=7 and m{2,3,4}m\in\{2,3,4\}.

1.2. Organization

In Section 2, we first introduce some convenient terminology and then we recall basic facts about μ\mu-bubbles, including general existence and stability results. We prove a Frankel type theorem for positive conformal Ricci curvature in Section 3. Combining the weighted slicing techniques, Shen-Ye’s diameter estimate, and the slice-and-dice procedure of Chodosh-Li, we prove the Main Theorem in Section 4. In Section 5, we generalize our results to a mapping version and prove a refinement of the Main Theorem into a positive result.

1.3. Acknowledgments

We would like to thank Xin Zhou for his helpful discussions and guidance on this project. L.M. acknowledges the support of an AMS-Simons travel grant.

2. Preliminaries

In this section, we discuss some preliminary results that will be needed in the proof of our main theorem. First, we introduce some convenient terminology for manifolds where the homology of the universal cover vanishes in certain degrees.

Definition 11.

We say a closed manifold MnM^{n} is mm-acyclic, m{1,2,n1}m\in\{1,2\cdots,n-1\}, if the universal cover M¯\overline{M} of MM satisfies

Hnm+1(M¯;)==Hn(M¯;)=0.H_{n-m+1}(\overline{M};\mathbb{Z})=\cdots=H_{n}(\overline{M};\mathbb{Z})=0.

2.1. Topological Preliminaries

In this subsection, we discuss some topological results about the universal cover of a closed manifold.

Lemma 12.

Suppose (Mn,g)(M^{n},g) is an mm-acyclic closed manifold for some m{1,2,,n1}m\in\{1,2,\cdots,n-1\}. Then there exists a geodesic line on (M¯,g¯)(\overline{M},\bar{g}), where g¯\bar{g} is the lifted Riemannian metric.

Proof.

Since Hn(M¯;)=0H_{n}(\overline{M};\mathbb{Z})=0, M¯\overline{M} is non-compact. The proof follows easily as in [3, Lemma 6]. ∎

More generally, we recall the following construction of Chodosh and Li [3].

Proposition 13.

Assume that M¯\overline{M} is the universal cover of a closed Riemannian manifold MnM^{n}. Further suppose that Hn(M¯,)=Hn1(M¯,)=0H_{n}(\overline{M},\mathbb{Z})=H_{n-1}(\overline{M},\mathbb{Z})=0. Then for any L>0L>0, there exists a geodesic line σ\sigma in M¯\overline{M} together with a closed, null-homologous manifold Λn2\Lambda^{n-2} embedded in M¯\overline{M} such that

  • (i)

    Λ\Lambda is linked with σ\sigma, i.e., if Λ=Σ\Lambda=\partial\Sigma then Σ\Sigma has non-zero algebraic intersection number with σ\sigma,

  • (ii)

    d(Λ,σ)Ld(\Lambda,\sigma)\geq L.

Proof.

This follows from the argument in [3, Section 2]. While Chodosh and Li consider the universal cover of a closed aspherical manifold, it is easy to see that their arguments apply to the universal cover M¯\overline{M} of an arbitrary closed manifold MnM^{n} provided that Hn(M¯,)=Hn1(M¯,)=0H_{n}(\overline{M},\mathbb{Z})=H_{n-1}(\overline{M},\mathbb{Z})=0. ∎

Proposition 14.

Suppose MnM^{n} is mm-acyclic. For any r>0r>0, there exists R(r)R(r) with the following property. For any knm+1k\geq n-m+1 and any kk-cycle α\alpha with αBr(p)\alpha\subset B_{r}(p), we have α=β\alpha=\partial\beta with βBR(p)\beta\subset B_{R}(p).

Proof.

The proof is exactly the same as in Chodosh-Li [3]; see also He-Zhu [12]. ∎

2.2. Preliminaries on μ\mu-bubbles

Next we recall some basic facts about μ\mu-bubbles, including general existence and stability results.

For n7n\leq 7, consider a Riemannian manifold (Mn,g)(M^{n},g) with boundary M=M+M\partial M=\partial_{-}M\sqcup\partial_{+}M, where neither of ±M\partial_{\pm}M is empty. Suppose uu is a smooth positive function on MM, and hh is a smooth function defined on the interior of MM so that h±h\to\pm\infty as x±Mx\to\partial_{\pm}M respectively. Given a Caccioppoli set Ω0M\Omega_{0}\subset M with smooth boundary and containing +M\partial_{+}M, consider the μ\mu-bubble functional:

(1) 𝒜(Ω):=Ωu𝑑aM(χΩχΩ0)hu𝑑v,\displaystyle\mathcal{A}(\Omega):=\int_{\partial^{*}\Omega}uda-\int_{M}(\chi_{\Omega}-\chi_{\Omega_{0}})hudv,

for all Caccioppoli set ΩM\Omega\subset M with ΩΔΩ0int(M)\Omega\Delta\Omega_{0}\subset\subset\operatorname{int}(M), where Ωint(M)\partial^{*}\Omega\subset\operatorname{int}(M) is the reduced boundary of Ω\Omega in int(M)\operatorname{int}(M). We call the 𝒜\mathcal{A}-minimizer Ω\Omega in this class a μ\mu-bubble.

The existence and regularity of a minimizer of 𝒜\mathcal{A} among all Caccioppoli sets was claimed by Gromov [9] and proven rigorously by Zhu [20].

Proposition 15 (Gromov[9], Zhu [20]).

There exists a smooth minimizer Ω\Omega for 𝒜\mathcal{A} such that ΩΔΩ0int(Mn)\Omega\Delta\Omega_{0}\subset\subset\operatorname{int}(M^{n}), for 3n73\leq n\leq 7.

In Chodosh-Li’s proof [3] of the K(π,1)K(\pi,1) conjecture in dimension 44 and 55, they generalized the μ\mu-bubble techniques to a free boundary version. In their setting, they assume that (Mn,g)(M^{n},g) is a Riemannian manifold with co-dimension 22 corners in the sense that any boundary point has a neighborhood diffeomorphic to one of the following: {xn:xn0}\{x\in\mathbb{R}^{n}:x_{n}\geq 0\} or {xn:xn1,xn0}\{x\in\mathbb{R}^{n}:x_{n-1},x_{n}\geq 0\}. Furthermore, M=±M0M\partial M=\partial_{\pm}M\cup\partial_{0}M, ±M\partial_{\pm}M meets 0M\partial_{0}M orthogonally, and ±M0M\partial_{\pm}M\cap\partial_{0}M consists of smooth co-dimension 22 closed submanifolds. In the free boundary setting, we assume that

H0M+u1Mu,ν0M=0.H_{\partial_{0}M}+u^{-1}\langle\nabla_{M}u,\nu_{\partial_{0}M}\rangle=0.

For the functional 𝒜\mathcal{A} defined as before, Chodosh-Li used similar arguments as in Zhu’s work [20] and proved the following results.

Proposition 16 (Chodosh-Li [3]).

Suppose n7n\leq 7, there exists Ω\Omega with Ωint(M)0M\partial\Omega\subset\operatorname{int}(M)\cup\partial_{0}M minimizing 𝒜\mathcal{A} among such regions. The boundary Ω\partial\Omega is smooth and meets 0M\partial_{0}M orthogonally. We have

H=u1Mu,νΩ+hH=-u^{-1}\langle\nabla_{M}u,\nu_{\partial\Omega}\rangle+h

along Ω\partial\Omega, where HH is the mean curvature of Ω\partial\Omega with respect to the unit outer normal νΩ\nu_{\partial\Omega}. Finally, if Σ\Sigma is a component of Ω\partial\Omega, then for any ψC1(Σ)\psi\in C^{1}(\Sigma), we have

0\displaystyle 0 d2dt2|t=0𝒜(Ωt)\displaystyle\leq\left.\frac{d^{2}}{dt^{2}}\right|_{t=0}\mathcal{A}(\Omega^{t})
=ΣuψΔΣψ(AΣ2+RicM(ν,ν))uψ2ψΣu,Σψ\displaystyle=\int_{\Sigma}-u\psi\Delta_{\Sigma}\psi-\left(\|A_{\Sigma}\|^{2}+\operatorname{Ric}_{M}(\nu,\nu)\right)u\psi^{2}-\psi\langle\nabla_{\Sigma}u,\nabla_{\Sigma}\psi\rangle
+ψ2M2u(ν,ν)ψ2u1Mu,ν2uψ2Mh,νdn1\displaystyle\qquad\qquad+\psi^{2}\nabla^{2}_{M}u(\nu,\nu)-\psi^{2}u^{-1}\langle\nabla_{M}u,\nu\rangle^{2}-u\psi^{2}\langle\nabla_{M}h,\nu\rangle~{}d\mathcal{H}^{n-1}
+ΣuψψηA0M(νΣ,νΣ)uψ2dn2,\displaystyle\qquad\qquad+\int_{\partial\Sigma}u\psi\frac{\partial\psi}{\partial\eta}-A_{\partial_{0}M}(\nu_{\partial\Sigma},\nu_{\partial\Sigma})u\psi^{2}~{}d\mathcal{H}^{n-2},

where ν=νΩ\nu=\nu_{\partial\Omega} is the unit outer normal of Ω\Omega, η\eta is the unit co-normal of Σ\Sigma along Σ\partial\Sigma, and {Ωt}\{\Omega^{t}\} is a smooth family of regions with Ω0=Ω\Omega^{0}=\Omega and normal speed ψ\psi at t=0t=0.

2.3. Diameter Estimates

The classical Bonnet-Myers theorem gives a diameter bound for manifolds with positive Ricci curvature. Shen and Ye discovered the following diameter bound for stable minimal hypersurfaces in manifolds with positive 2-intermediate curvature.

Theorem 17 (Shen-Ye [17]).

Let MnM^{n} be a complete Riemannian manifold of dimension n{3,4,5}n\in\{3,4,5\}. Assume that the 22-intermediate curvature of MM is at least κ>0\kappa>0. Let Σ\Sigma be a stable minimal hypersurface in MM. Then for every pΣp\in\Sigma, one has

d(p,Σ)c(n)πκ,d(p,\partial\Sigma)\leq\sqrt{c(n)}\frac{\pi}{\sqrt{\kappa}},

where c(n)c(n) is a dimensional constant.

Shen and Ye also proved the following generalized Bonnet-Myers theorem which will be very useful for proving diameter bounds for certain weighted minimal slicings. We state the 3-dimensional version first since it is simpler and more powerful.

Theorem 18 (Shen-Ye [18]).

Let N3N^{3} be a complete Riemannian manifold. Assume there is a function f>0f>0 on NN such that

(2) RicN(e,e)f1ΔNf+12|Nlnf|2κ>0\operatorname{Ric}_{N}(e,e)-f^{-1}\Delta_{N}f+\frac{1}{2}|\nabla_{N}\ln f|^{2}\geq\kappa>0

for all unit tangent vectors ee. Then

diam(N)2πκ.\operatorname{diam}(N)\leq\sqrt{2}\frac{\pi}{\sqrt{\kappa}}.

In particular, NN is compact.

Definition 19.

Following Shen-Ye, we will say a manifold satisfying (2) admits positive conformal Ricci curvature.

In higher dimensions, the theorem must be modified as follows.

Theorem 20 (Shen-Ye [18]).

Let NkN^{k} be a complete Riemannian manifold of dimension k4k\geq 4. Assume there is a function f>0f>0 on NN and τ,ε>0\tau,\varepsilon>0 so that

RicN(e,e)τf1ΔNf+[τ(k14+ε)τ2]|Nlnf|2κ>0\operatorname{Ric}_{N}(e,e)-\tau f^{-1}\Delta_{N}f+\left[\tau-\left(\frac{k-1}{4}+\varepsilon\right)\tau^{2}\right]|\nabla_{N}\ln f|^{2}\geq\kappa>0

for all unit tangent vectors ee. Then the diameter of NN is bounded above in terms of kk, κ\kappa, and ε\varepsilon.

3. Frankel Type Theorems

Recall that Frankel’s theorem says that if MM has positive Ricci curvature and minimal boundary then M\partial M is connected. In this section, we are going to prove a Frankel type theorem for manifolds with positive conformal Ricci curvature. As motivation, we first prove the following version of Frankel’s theorem for bi-Ricci curvature. This theorem fits into the general theme, first observed by Shen-Ye [17], that stable minimal hypersurfaces in manifolds with positive bi-Ricci curvature behave as if they had positive Ricci curvature. Curiously, in contrast to the diameter bounds, the following theorem does not require any dimension restriction.

Theorem 21.

Let Mn+1M^{n+1} be a compact manifold with boundary. Assume that MM has positive bi-Ricci curvature and minimal boundary. Assume that Σn\Sigma^{n} is a two-sided, stable, free boundary minimal hypersurface in MM. Then Σ\partial\Sigma is connected.

Proof.

Let ν\nu be the unit normal vector to Σ\Sigma in MM and let η\eta be the unit outward co-normal along Σ\partial\Sigma. The second variation formula says that for any smooth function ψ\psi on Σ\Sigma we have

0\displaystyle 0 Σ|Σψ2|(|AΣ|2+Ric(ν,ν))ψ2dvΣAM(ν,ν)ψ2𝑑a\displaystyle\leq\int_{\Sigma}|\nabla_{\Sigma}\psi^{2}|-(|A_{\Sigma}|^{2}+\operatorname{Ric}(\nu,\nu))\psi^{2}\,dv-\int_{\partial\Sigma}A_{\partial M}(\nu,\nu)\psi^{2}\,da
=ΣψJΣψ𝑑v+ΣψψηAM(ν,ν)ψ2da,\displaystyle=-\int_{\Sigma}\psi J_{\Sigma}\psi\,dv+\int_{\partial\Sigma}\psi\frac{\partial\psi}{\partial\eta}-A_{\partial M}(\nu,\nu)\psi^{2}\,da,

where JΣ=ΔΣ+|AΣ|2+Ric(ν,ν)J_{\Sigma}=\Delta_{\Sigma}+|A_{\Sigma}|^{2}+\operatorname{Ric}(\nu,\nu) is the Jacobi operator. Let f>0f>0 be a first eigenfunction so that

{JΣf+λf=0,ηf=AM(ν,ν)f,\begin{cases}J_{\Sigma}f+\lambda f=0,\\ \partial_{\eta}f=A_{\partial M}(\nu,\nu)f,\end{cases}

with λ0\lambda\geq 0. We are going to consider a weighted length functional with weight ff as in Shen-Ye [17].

More precisely, given a curve cc in Σ\Sigma, define

L(c)=0f|c˙|𝑑t.L(c)=\int_{0}^{\ell}f|\dot{c}|\,dt.

We now restrict the calculations in Σ\Sigma and follow [17] to compute the first and second variation of LL. Assume that c(t)c(t) is a unit speed curve in Σ\Sigma. Let c(s,t)c(s,t) be a variation with c(0,t)=c(t)c(0,t)=c(t). We compute

Ls=0s(fc˙,c˙1/2)𝑑t=0f,cs|c˙|+f/sc˙,c˙|c˙|dt.\displaystyle\frac{\partial L}{\partial s}=\int_{0}^{\ell}\frac{\partial}{\partial s}(f\langle\dot{c},\dot{c}\rangle^{1/2})\,dt=\int_{0}^{\ell}\langle\nabla f,\frac{\partial c}{\partial s}\rangle|\dot{c}|+f\frac{\langle\nabla_{\partial/\partial s}\dot{c},\dot{c}\rangle}{|\dot{c}|}\,dt.

Therefore setting V(t)=cs(0,t)V(t)=\frac{\partial c}{\partial s}(0,t) we get the first variation formula

Ls|s=0=0f,V+fc˙V,c˙dt.\frac{\partial L}{\partial s}\bigg{|}_{s=0}=\int_{0}^{\ell}\langle\nabla f,V\rangle+f\langle\nabla_{\dot{c}}V,\dot{c}\rangle\,dt.

Assuming cc is a critical point among curves with endpoints constrained to lie in Σ\partial\Sigma, we can test this against VV with V(0)Tc(0)ΣV(0)\in T_{c(0)}\partial\Sigma and V()Tc()ΣV(\ell)\in T_{c(\ell)}\partial\Sigma to get

0\displaystyle 0 =0f,V+ftV,c˙fV,c˙c˙dt\displaystyle=\int_{0}^{\ell}\langle\nabla f,V\rangle+f\frac{\partial}{\partial t}\langle V,\dot{c}\rangle-f\langle V,\nabla_{\dot{c}}\dot{c}\rangle\,dt
=0f,V+ddt[fV,c˙]f,c˙V,c˙fV,c˙c˙dt\displaystyle=\int_{0}^{\ell}\langle\nabla f,V\rangle+\frac{d}{dt}\left[f\langle V,\dot{c}\rangle\right]-\langle\nabla f,\dot{c}\rangle\langle V,\dot{c}\rangle-f\langle V,\nabla_{\dot{c}}\dot{c}\rangle\,dt
=fV,c˙|0+0f,Vf,c˙V,c˙fV,c˙c˙dt\displaystyle=f\langle V,\dot{c}\rangle\bigg{|}_{0}^{\ell}+\int_{0}^{\ell}\langle\nabla f,V\rangle-\langle\nabla f,\dot{c}\rangle\langle V,\dot{c}\rangle-f\langle V,\nabla_{\dot{c}}\dot{c}\rangle\,dt
=fV,c˙|0+0(f),VfV,c˙c˙dt.\displaystyle=f\langle V,\dot{c}\rangle\bigg{|}_{0}^{\ell}+\int_{0}^{\ell}\langle(\nabla f)^{\perp},V\rangle-f\langle V,\nabla_{\dot{c}}\dot{c}\rangle\,dt.

Here (f)(\nabla f)^{\perp} is the part of f\nabla f orthogonal to c˙\dot{c}. Therefore cc satisfies the weighted geodesic equation

c˙c˙=f1(f),\nabla_{\dot{c}}{\dot{c}}=f^{-1}(\nabla f)^{\perp},

and cc meets Σ\partial\Sigma orthogonally at both endpoints. Next we compute the 2nd variation to get

2Ls2|s=0\displaystyle\frac{\partial^{2}L}{\partial s^{2}}\bigg{|}_{s=0} =0csf,V+f,cscs+2f,Vcsct,ct+fscsct,ctfcsct,ct2dt\displaystyle=\int_{0}^{\ell}\langle\nabla_{\frac{\partial c}{\partial s}}\nabla f,V\rangle+\langle\nabla f,\nabla_{\frac{\partial c}{\partial s}}\frac{\partial c}{\partial s}\rangle+2\langle\nabla f,V\rangle\langle\nabla_{\frac{\partial c}{\partial s}}\frac{\partial c}{\partial t},\frac{\partial c}{\partial t}\rangle+f\frac{\partial}{\partial s}\langle\nabla_{\frac{\partial c}{\partial s}}\frac{\partial c}{\partial t},\frac{\partial c}{\partial t}\rangle-f\langle\nabla_{\frac{\partial c}{\partial s}}\frac{\partial c}{\partial t},\frac{\partial c}{\partial t}\rangle^{2}\,dt
=0HessfΣ(V,V)+f,cscs+2f,Vctcs,ct+fsctcs,ctfctcs,ct2dt\displaystyle=\int_{0}^{\ell}\operatorname{Hess}^{\Sigma}_{f}(V,V)+\langle\nabla f,\nabla_{\frac{\partial c}{\partial s}}\frac{\partial c}{\partial s}\rangle+2\langle\nabla f,V\rangle\langle\nabla_{\frac{\partial c}{\partial t}}\frac{\partial c}{\partial s},\frac{\partial c}{\partial t}\rangle+f\frac{\partial}{\partial s}\langle\nabla_{\frac{\partial c}{\partial t}}\frac{\partial c}{\partial s},\frac{\partial c}{\partial t}\rangle-f\langle\nabla_{\frac{\partial c}{\partial t}}\frac{\partial c}{\partial s},\frac{\partial c}{\partial t}\rangle^{2}\,dt
=0HessfΣ(V,V)+f,cscs+fcsctcs,ct+fctcs,csctdt\displaystyle=\int_{0}^{\ell}\operatorname{Hess}^{\Sigma}_{f}(V,V)+\langle\nabla f,\nabla_{\frac{\partial c}{\partial s}}\frac{\partial c}{\partial s}\rangle+f\langle\nabla_{\frac{\partial c}{\partial s}}\nabla_{\frac{\partial c}{\partial t}}\frac{\partial c}{\partial s},\frac{\partial c}{\partial t}\rangle\,+f\langle\nabla_{\frac{\partial c}{\partial t}}\frac{\partial c}{\partial s},\nabla_{\frac{\partial c}{\partial s}}\frac{\partial c}{\partial t}\rangle\,dt
+02f,Vctcs,ctfctcs,ct2dt\displaystyle\qquad\quad+\int_{0}^{\ell}2\langle\nabla f,V\rangle\langle\nabla_{\frac{\partial c}{\partial t}}\frac{\partial c}{\partial s},\frac{\partial c}{\partial t}\rangle-f\langle\nabla_{\frac{\partial c}{\partial t}}\frac{\partial c}{\partial s},\frac{\partial c}{\partial t}\rangle^{2}\,dt
=0HessfΣ(V,V)+f,cscs+fctcscs,ctfRΣ(cs,ct,cs,ct)dt\displaystyle=\int_{0}^{\ell}\operatorname{Hess}^{\Sigma}_{f}(V,V)+\langle\nabla f,\nabla_{\frac{\partial c}{\partial s}}\frac{\partial c}{\partial s}\rangle+f\langle\nabla_{\frac{\partial c}{\partial t}}\nabla_{\frac{\partial c}{\partial s}}\frac{\partial c}{\partial s},\frac{\partial c}{\partial t}\rangle-fR^{\Sigma}(\frac{\partial c}{\partial s},\frac{\partial c}{\partial t},\frac{\partial c}{\partial s},\frac{\partial c}{\partial t})\,dt
+0fctcs,ctcs+2f,Vctcs,ctfctcs,ct2dt\displaystyle\qquad\quad+\int_{0}^{\ell}f\langle\nabla_{\frac{\partial c}{\partial t}}\frac{\partial c}{\partial s},\nabla_{\frac{\partial c}{\partial t}}\frac{\partial c}{\partial s}\rangle+2\langle\nabla f,V\rangle\langle\nabla_{\frac{\partial c}{\partial t}}\frac{\partial c}{\partial s},\frac{\partial c}{\partial t}\rangle-f\langle\nabla_{\frac{\partial c}{\partial t}}\frac{\partial c}{\partial s},\frac{\partial c}{\partial t}\rangle^{2}\,dt
=0HessfΣ(V,V)+f,cscs+ftcscs,ctfcscs,ctctfRΣ(cs,ct,cs,ct)dt\displaystyle=\int_{0}^{\ell}\operatorname{Hess}^{\Sigma}_{f}(V,V)+\langle\nabla f,\nabla_{\frac{\partial c}{\partial s}}\frac{\partial c}{\partial s}\rangle+f\frac{\partial}{\partial t}\langle\nabla_{\frac{\partial c}{\partial s}}\frac{\partial c}{\partial s},\frac{\partial c}{\partial t}\rangle-f\langle\nabla_{\frac{\partial c}{\partial s}}\frac{\partial c}{\partial s},\nabla_{\frac{\partial c}{\partial t}}\frac{\partial c}{\partial t}\rangle-fR^{\Sigma}(\frac{\partial c}{\partial s},\frac{\partial c}{\partial t},\frac{\partial c}{\partial s},\frac{\partial c}{\partial t})\,dt
+0fctcs,ctcs+2f,Vctcs,ctfctcs,ct2dt.\displaystyle\qquad\quad+\int_{0}^{\ell}f\langle\nabla_{\frac{\partial c}{\partial t}}\frac{\partial c}{\partial s},\nabla_{\frac{\partial c}{\partial t}}\frac{\partial c}{\partial s}\rangle+2\langle\nabla f,V\rangle\langle\nabla_{\frac{\partial c}{\partial t}}\frac{\partial c}{\partial s},\frac{\partial c}{\partial t}\rangle-f\langle\nabla_{\frac{\partial c}{\partial t}}\frac{\partial c}{\partial s},\frac{\partial c}{\partial t}\rangle^{2}\,dt.

If cc minimizes LL then this will be non-negative for all admissible variations.

Now assume for contradiction that Σ\Sigma has two distinct boundary components Σ1\Sigma_{1} and Σ2\Sigma_{2}. Let cc be a unit speed curve which minimizes LL over all curves connecting Σ1\Sigma_{1} to Σ2\Sigma_{2}. We select e1=c˙e_{1}=\dot{c}, e2e_{2}, \ldots, ene_{n} to be an orthonormal frame along cc. Then

c˙ej,c˙=ej,c˙c˙=ej,f1f\langle\nabla_{\dot{c}}e_{j},\dot{c}\rangle=-\langle e_{j},\nabla_{\dot{c}}\dot{c}\rangle=-\langle e_{j},f^{-1}\nabla f\rangle

for j=2,,nj=2,\ldots,n. Actually, we can further select e2,,ene_{2},\ldots,e_{n} to be parallel in the normal bundle of cc so that c˙ej=ej,f1fc˙\nabla_{\dot{c}}e_{j}=-\langle e_{j},f^{-1}\nabla f\rangle\dot{c}.

We select variations cj(s,t)c_{j}(s,t) with Vj=ejV_{j}=e_{j} and cj(s,0)Σ1c_{j}(s,0)\in\Sigma_{1} and cj(s,)Σ2c_{j}(s,\ell)\in\Sigma_{2} for j=2,,nj=2,\ldots,n. Plugging these into the second variation formula and summing over jj we get

0\displaystyle 0 0ΔΣfHessfΣ(c˙,c˙)+j=2nf,cjscjs+fj=2ntcjscjs,c˙fRicΣ(c˙,c˙)dt\displaystyle\leq\int_{0}^{\ell}\Delta_{\Sigma}f-\operatorname{Hess}^{\Sigma}_{f}(\dot{c},\dot{c})+\sum_{j=2}^{n}\langle\nabla f,\nabla_{\frac{\partial c_{j}}{\partial s}}\frac{\partial c_{j}}{\partial s}\rangle+f\sum_{j=2}^{n}\frac{\partial}{\partial t}\langle\nabla_{\frac{\partial c_{j}}{\partial s}}\frac{\partial c_{j}}{\partial s},\dot{c}\rangle-f\operatorname{Ric}^{\Sigma}(\dot{c},\dot{c})\,dt
+0fj=2ncjscjs,c˙c˙+2j=2nf,ejc˙ej,c˙fj=2nc˙ej,c˙2+fj=2nc˙ej,c˙ejdt\displaystyle\qquad\quad+\int_{0}^{\ell}-f\sum_{j=2}^{n}\langle\nabla_{\frac{\partial c_{j}}{\partial s}}\frac{\partial c_{j}}{\partial s},\nabla_{\dot{c}}\dot{c}\rangle+2\sum_{j=2}^{n}\langle\nabla f,e_{j}\rangle\langle\nabla_{\dot{c}}e_{j},\dot{c}\rangle-f\sum_{j=2}^{n}\langle\nabla_{\dot{c}}e_{j},\dot{c}\rangle^{2}+f\sum_{j=2}^{n}\langle\nabla_{\dot{c}}e_{j},\nabla_{\dot{c}}e_{j}\rangle\,dt
=0ΔΣfHessfΣ(c˙,c˙)+j=2nf,cjscjs+j=2nt[fcjscjs,c˙]j=2nf,c˙cjscjs,c˙dt\displaystyle=\int_{0}^{\ell}\Delta_{\Sigma}f-\operatorname{Hess}^{\Sigma}_{f}(\dot{c},\dot{c})+\sum_{j=2}^{n}\langle\nabla f,\nabla_{\frac{\partial c_{j}}{\partial s}}\frac{\partial c_{j}}{\partial s}\rangle+\sum_{j=2}^{n}\frac{\partial}{\partial t}\left[f\langle\nabla_{\frac{\partial c_{j}}{\partial s}}\frac{\partial c_{j}}{\partial s},\dot{c}\rangle\right]-\sum_{j=2}^{n}\langle\nabla f,\dot{c}\rangle\langle\nabla_{\frac{\partial c_{j}}{\partial s}}\frac{\partial c_{j}}{\partial s},\dot{c}\rangle\,dt
+0fj=2ncjscjs,f1(f)2j=2nf,ejej,f1ffj=2nej,f1f2dt\displaystyle\qquad\quad+\int_{0}^{\ell}-f\sum_{j=2}^{n}\langle\nabla_{\frac{\partial c_{j}}{\partial s}}\frac{\partial c_{j}}{\partial s},f^{-1}(\nabla f)^{\perp}\rangle-2\sum_{j=2}^{n}\langle\nabla f,e_{j}\rangle\langle e_{j},f^{-1}\nabla f\rangle-f\sum_{j=2}^{n}\langle e_{j},f^{-1}\nabla f\rangle^{2}\,dt
+0fRicΣ(c˙,c˙)+j=2nf1f,ej2dt\displaystyle\qquad\quad+\int_{0}^{\ell}-f\operatorname{Ric}^{\Sigma}(\dot{c},\dot{c})+\sum_{j=2}^{n}f^{-1}\langle\nabla f,e_{j}\rangle^{2}\,dt
=0ΔΣfHessfΣ(c˙,c˙)+j=2nt[fcjscjs,c˙]2f1|(f)|2fRicΣ(c˙,c˙)dt.\displaystyle=\int_{0}^{\ell}\Delta_{\Sigma}f-\operatorname{Hess}^{\Sigma}_{f}(\dot{c},\dot{c})+\sum_{j=2}^{n}\frac{\partial}{\partial t}\left[f\langle\nabla_{\frac{\partial c_{j}}{\partial s}}\frac{\partial c_{j}}{\partial s},\dot{c}\rangle\right]-2f^{-1}|(\nabla f)^{\perp}|^{2}-f\operatorname{Ric}^{\Sigma}(\dot{c},\dot{c})\,dt.

Now observe that

HessfΣ(c˙,c˙)=c˙f,c˙=d2fdt2f,c˙c˙=d2fdt2f1|(f)|2.\operatorname{Hess}_{f}^{\Sigma}(\dot{c},\dot{c})=\langle\nabla_{\dot{c}}\nabla f,\dot{c}\rangle=\frac{d^{2}f}{dt^{2}}-\langle\nabla f,\nabla_{\dot{c}}\dot{c}\rangle=\frac{d^{2}f}{dt^{2}}-f^{-1}|(\nabla f)^{\perp}|^{2}.

Hence we obtain

0\displaystyle 0 0f|AΣ|2fRic(ν,ν)fRicΣ(c˙,c˙)d2fdt2+j=2nt[fcjscjs,c˙]f1|(f)|2dt\displaystyle\leq\int_{0}^{\ell}-f|A_{\Sigma}|^{2}-f\operatorname{Ric}(\nu,\nu)-f\operatorname{Ric}^{\Sigma}(\dot{c},\dot{c})-\frac{d^{2}f}{dt^{2}}+\sum_{j=2}^{n}\frac{\partial}{\partial t}\left[f\langle\nabla_{\frac{\partial c_{j}}{\partial s}}\frac{\partial c_{j}}{\partial s},\dot{c}\rangle\right]-f^{-1}|(\nabla f)^{\perp}|^{2}\,dt
=0f[Ric(ν,ν)+Ric(c˙,c˙)R(ν,c˙,ν,c˙)]𝑑t0d2fdt2𝑑t0f1|(f)|2𝑑t\displaystyle=-\int_{0}^{\ell}f\left[\operatorname{Ric}(\nu,\nu)+\operatorname{Ric}(\dot{c},\dot{c})-R(\nu,\dot{c},\nu,\dot{c})\right]\,dt-\int_{0}^{\ell}\frac{d^{2}f}{dt^{2}}\,dt-\int_{0}^{\ell}f^{-1}|(\nabla f)^{\perp}|^{2}\,dt
0f[|AΣ|2+j=2n(AΣ(c˙,c˙)AΣ(e˙j,ej)AΣ(c˙,ej)2)]𝑑t\displaystyle\qquad\quad-\int_{0}^{\ell}f\left[|A_{\Sigma}|^{2}+\sum_{j=2}^{n}(A_{\Sigma}(\dot{c},\dot{c})A_{\Sigma}(\dot{e}_{j},e_{j})-A_{\Sigma}(\dot{c},e_{j})^{2})\right]\,dt
f(c())HΣ2(c())f(c(0))HΣ1(c(0))\displaystyle\qquad\quad-f(c(\ell))H_{\Sigma_{2}}(c(\ell))-f(c(0))H_{\Sigma_{1}}(c(0))\phantom{\int}
<ηf(c())ηf(c(0))f(c())HΣ2(c())f(c(0))HΣ1(c(0)).\displaystyle<-\partial_{\eta}f(c(\ell))-\partial_{\eta}f(c(0))-f(c(\ell))H_{\Sigma_{2}}(c(\ell))-f(c(0))H_{\Sigma_{1}}(c(0)).\phantom{\int}

Here we used the fact that Ric(ν,ν)+Ric(c˙,c˙)R(ν,c˙,ν,c˙)>0\operatorname{Ric}(\nu,\nu)+\operatorname{Ric}(\dot{c},\dot{c})-R(\nu,\dot{c},\nu,\dot{c})>0 by the assumption on the bi-Ricci curvature, and the fact that

|AΣ|2+j=2n(AΣ(c˙,c˙)AΣ(ej,ej)AΣ(c˙,ej)2)0|A_{\Sigma}|^{2}+\sum_{j=2}^{n}(A_{\Sigma}(\dot{c},\dot{c})A_{\Sigma}(e_{j},e_{j})-A_{\Sigma}(\dot{c},e_{j})^{2})\geq 0

since Σ\Sigma is minimal; see [17, Equation 15]. Finally, it remains to note that

ηf=AM(ν,ν)f=fHΣ\partial_{\eta}f=A_{\partial M}(\nu,\nu)f=-fH_{\partial\Sigma}

since M\partial M is minimal and Σ\partial\Sigma meets M\partial M orthogonally. Thus the final term in the previous chain of inequalities is equal to 0 and we get our contradiction. ∎

Next, we prove a Frankel type theorem for manifolds with positive conformal Ricci curvature.

Theorem 22.

Let Σ\Sigma be a 33-dimensional compact Riemannian manifold with boundary. Assume Σ\Sigma admits a positive function ff which satisfies

Ric(v,v)f1Δf+12|lnf|20\operatorname{Ric}(v,v)-f^{-1}\Delta f+\frac{1}{2}|\nabla\ln f|^{2}\geq 0

for all unit vectors vv. Moreover, suppose that ηf=fHΣ\partial_{\eta}f=-fH_{\partial\Sigma} where η\eta is the unit outward normal to Σ\partial\Sigma. Then Σ\partial\Sigma is connected.

Proof.

The argument is similar to the previous one but with a slight improvement coming from a modified choice of the variations. Assume for contradiction that Σ\partial\Sigma has two distinct connected components Σ1\Sigma_{1} and Σ2\Sigma_{2}. We let cc be a unit speed curve which minimizes the ff-weighted length from Σ1\Sigma_{1} to Σ2\Sigma_{2}. Again we let e1=c˙,e2,,ene_{1}=\dot{c},e_{2},\ldots,e_{n} be an orthonormal frame along cc such that e2,,ene_{2},\ldots,e_{n} are parallel in the normal bundle of cc. However, this time we choose variations cj(s,t)c_{j}(s,t) with Vj=f1/2ejV_{j}=f^{-1/2}e_{j} and cj(s,0)Σ1c_{j}(s,0)\in\Sigma_{1} and cj(s,)Σ2c_{j}(s,\ell)\in\Sigma_{2} for j=2,,nj=2,\ldots,n.

Applying the second variation formula to each cjc_{j} and then summing over jj, we deduce that

0\displaystyle 0 0f1Δff1Hessf(c˙,c˙)+j=2nf,cjscjs+fj=2ntcjscjs,c˙Ric(c˙,c˙)dt\displaystyle\leq\int_{0}^{\ell}f^{-1}\Delta f-f^{-1}\operatorname{Hess}_{f}(\dot{c},\dot{c})+\sum_{j=2}^{n}\langle\nabla f,\nabla_{\frac{\partial c_{j}}{\partial s}}\frac{\partial c_{j}}{\partial s}\rangle+f\sum_{j=2}^{n}\frac{\partial}{\partial t}\langle\nabla_{\frac{\partial c_{j}}{\partial s}}\frac{\partial c_{j}}{\partial s},\dot{c}\rangle-\operatorname{Ric}(\dot{c},\dot{c})\,dt
+0fj=2ncjscjs,c˙c˙+2j=2nf,f1/2ejc˙(f1/2ej),c˙fj=2nc˙(f1/2ej),c˙2dt\displaystyle\qquad\quad+\int_{0}^{\ell}-f\sum_{j=2}^{n}\langle\nabla_{\frac{\partial c_{j}}{\partial s}}\frac{\partial c_{j}}{\partial s},\nabla_{\dot{c}}\dot{c}\rangle+2\sum_{j=2}^{n}\langle\nabla f,f^{-1/2}e_{j}\rangle\langle\nabla_{\dot{c}}(f^{-1/2}e_{j}),\dot{c}\rangle-f\sum_{j=2}^{n}\langle\nabla_{\dot{c}}(f^{-1/2}e_{j}),\dot{c}\rangle^{2}\,dt
+0fj=2nc˙(f1/2ej),c˙(f1/2ej)dt.\displaystyle\qquad\quad+\int_{0}^{\ell}f\sum_{j=2}^{n}\langle\nabla_{\dot{c}}(f^{-1/2}e_{j}),\nabla_{\dot{c}}(f^{-1/2}e_{j})\rangle\,dt.

This simplifies to give

0\displaystyle 0 0f1Δff1Hessf(c˙,c˙)+j=2nf,cjscjs+j=2nt[fcjscjs,c˙]j=2nf,c˙cjscjs,c˙dt\displaystyle\leq\int_{0}^{\ell}f^{-1}\Delta f-f^{-1}\operatorname{Hess}_{f}(\dot{c},\dot{c})+\sum_{j=2}^{n}\langle\nabla f,\nabla_{\frac{\partial c_{j}}{\partial s}}\frac{\partial c_{j}}{\partial s}\rangle+\sum_{j=2}^{n}\frac{\partial}{\partial t}\left[f\langle\nabla_{\frac{\partial c_{j}}{\partial s}}\frac{\partial c_{j}}{\partial s},\dot{c}\rangle\right]-\sum_{j=2}^{n}\langle\nabla f,\dot{c}\rangle\langle\nabla_{\frac{\partial c_{j}}{\partial s}}\frac{\partial c_{j}}{\partial s},\dot{c}\rangle\,dt
+0fj=2ncjscjs,f1(f)2f2j=2nf,ejej,ff2j=2nej,f2dt\displaystyle\qquad\quad+\int_{0}^{\ell}-f\sum_{j=2}^{n}\langle\nabla_{\frac{\partial c_{j}}{\partial s}}\frac{\partial c_{j}}{\partial s},f^{-1}(\nabla f)^{\perp}\rangle-2f^{-2}\sum_{j=2}^{n}\langle\nabla f,e_{j}\rangle\langle e_{j},\nabla f\rangle-f^{-2}\sum_{j=2}^{n}\langle e_{j},\nabla f\rangle^{2}\,dt
+0Ric(c˙,c˙)+f2j=2nf,ej2+n14c˙,lnf2dt\displaystyle\qquad\quad+\int_{0}^{\ell}-\operatorname{Ric}(\dot{c},\dot{c})+f^{-2}\sum_{j=2}^{n}\langle\nabla f,e_{j}\rangle^{2}+\frac{n-1}{4}\langle\dot{c},\nabla\ln f\rangle^{2}\,dt
=0f1Δff1Hessf(c˙,c˙)+j=2nt[fcjscjs,c˙]2f2|(f)|2Ric(c˙,c˙)+n14c˙,lnf2dt.\displaystyle=\int_{0}^{\ell}f^{-1}\Delta f-f^{-1}\operatorname{Hess}_{f}(\dot{c},\dot{c})+\sum_{j=2}^{n}\frac{\partial}{\partial t}\left[f\langle\nabla_{\frac{\partial c_{j}}{\partial s}}\frac{\partial c_{j}}{\partial s},\dot{c}\rangle\right]-2f^{-2}|(\nabla f)^{\perp}|^{2}-\operatorname{Ric}(\dot{c},\dot{c})+\frac{n-1}{4}\langle\dot{c},\nabla\ln f\rangle^{2}\,dt.

As before, we have

Hessf(c˙,c˙)=c˙f,c˙=d2fdt2f,c˙c˙=d2fdt2f1|(f)|2.\operatorname{Hess}_{f}(\dot{c},\dot{c})=\langle\nabla_{\dot{c}}\nabla f,\dot{c}\rangle=\frac{d^{2}f}{dt^{2}}-\langle\nabla f,\nabla_{\dot{c}}\dot{c}\rangle=\frac{d^{2}f}{dt^{2}}-f^{-1}|(\nabla f)^{\perp}|^{2}.

Therefore, combined with n14c˙,lnf2=12c˙,lnf212|lnf|2\frac{n-1}{4}\langle\dot{c},\nabla\ln f\rangle^{2}=\frac{1}{2}\langle\dot{c},\nabla\ln f\rangle^{2}\leq\frac{1}{2}|\nabla\ln f|^{2}, we obtain

0\displaystyle 0 0f1Δff1d2fdt2+j=2nt[fcjscjs,c˙]f2|(f)|2Ric(c˙,c˙)+12|lnf|2dt.\displaystyle\leq\int_{0}^{\ell}f^{-1}\Delta f-f^{-1}\frac{d^{2}f}{dt^{2}}+\sum_{j=2}^{n}\frac{\partial}{\partial t}\left[f\langle\nabla_{\frac{\partial c_{j}}{\partial s}}\frac{\partial c_{j}}{\partial s},\dot{c}\rangle\right]-f^{-2}|(\nabla f)^{\perp}|^{2}-\operatorname{Ric}(\dot{c},\dot{c})+\frac{1}{2}|\nabla\ln f|^{2}\,dt.

Next, note that

f1d2fdt2=d2dt2(lnf)+[d(lnf)dt]2=d2dt2(lnf)+f2f,c˙2.f^{-1}\frac{d^{2}f}{dt^{2}}=\frac{d^{2}}{dt^{2}}(\ln f)+\left[\frac{d(\ln f)}{dt}\right]^{2}=\frac{d^{2}}{dt^{2}}(\ln f)+f^{-2}\langle\nabla f,\dot{c}\rangle^{2}.

It follows that

0\displaystyle 0 0f1Δff2f,c˙2f2|(f)|2+12|lnf|2Ric(c˙,c˙)d2dt2(lnf)+j=2nt[fcjscjs,c˙]dt\displaystyle\leq\int_{0}^{\ell}f^{-1}\Delta f-f^{-2}\langle\nabla f,\dot{c}\rangle^{2}-f^{-2}|(\nabla f)^{\perp}|^{2}+\frac{1}{2}|\nabla\ln f|^{2}-\operatorname{Ric}(\dot{c},\dot{c})-\frac{d^{2}}{dt^{2}}(\ln f)+\sum_{j=2}^{n}\frac{\partial}{\partial t}\left[f\langle\nabla_{\frac{\partial c_{j}}{\partial s}}\frac{\partial c_{j}}{\partial s},\dot{c}\rangle\right]\,dt
=0f1Δf12|lnf|2Ric(c˙,c˙)d2dt2(lnf)+j=2nt[fcjscjs,c˙]dt\displaystyle=\int_{0}^{\ell}f^{-1}\Delta f-\frac{1}{2}|\nabla\ln f|^{2}-\operatorname{Ric}(\dot{c},\dot{c})-\frac{d^{2}}{dt^{2}}(\ln f)+\sum_{j=2}^{n}\frac{\partial}{\partial t}\left[f\langle\nabla_{\frac{\partial c_{j}}{\partial s}}\frac{\partial c_{j}}{\partial s},\dot{c}\rangle\right]\,dt
<0d2dt2(lnf)+j=2nt[fcjscjs,c˙]dt.\displaystyle<\int_{0}^{\ell}-\frac{d^{2}}{dt^{2}}(\ln f)+\sum_{j=2}^{n}\frac{\partial}{\partial t}\left[f\langle\nabla_{\frac{\partial c_{j}}{\partial s}}\frac{\partial c_{j}}{\partial s},\dot{c}\rangle\right]\,dt.

Here we used the first assumption of the theorem to get the final inequality. Now we can apply the fundamental theorem of calculus to get

0\displaystyle 0 <η(lnf)(c())η(lnf)(c(0))HΣ2(c())HΣ1(c(0))\displaystyle<-\partial_{\eta}(\ln f)(c(\ell))-\partial_{\eta}(\ln f)(c(0))-H_{\Sigma_{2}}(c(\ell))-H_{\Sigma_{1}}(c(0))
=ηff(c())ηff(c(0))HΣ2(c())HΣ1(c(0)).\displaystyle=-\frac{\partial_{\eta}f}{f}(c(\ell))-\frac{\partial_{\eta}f}{f}(c(0))-H_{\Sigma_{2}}(c(\ell))-H_{\Sigma_{1}}(c(0)).

The second assumption of the theorem implies that the previous line is equal to 0, and again we’ve reached a contradiction. ∎

4. Proof of the Main Theorem

In this section, we prove our main result Theorem 4. We will proceed case by case based on the value of mm.

4.1. The Ricci Curvature Case

Note that the 11-intermediate curvature is just the Ricci curvature. Therefore, the m=1m=1 case of Conjecture 3 is a well-known corollary of the Bonnet-Myers theorem. In this case, we do not need any restriction on the ambient dimension nn.

Theorem 23.

Let MnM^{n} be a closed manifold. Assume that the universal cover M¯\overline{M} of MM satisfies Hn(M¯,)=0H_{n}(\overline{M},\mathbb{Z})=0. Then MM does not admit a metric of positive Ricci curvature.

Proof.

We prove the contrapositive. Assume that MM admits a metric of positive Ricci curvature. Let M¯\overline{M} be the universal cover of MM. By the Bonnet-Myers theorem, M¯\overline{M} is a closed, orientable nn-dimensional manifold. Therefore Hn(M¯,)0H_{n}(\overline{M},\mathbb{Z})\neq 0. ∎

4.2. The Scalar Curvature Case

We now show that the m=n1m=n-1 case of Conjecture 3 is equivalent to the K(π,1)K(\pi,1) conjecture. Since the (n1)(n-1)-intermediate curvature is equivalent to scalar curvature, it suffices to prove the following simple fact.

Theorem 24.

Let MnM^{n} be a closed manifold. Then the universal cover M¯\overline{M} of MM satisfies

(3) Hn(M¯,)=Hn1(M¯,)==H2(M¯,)=0H_{n}(\overline{M},\mathbb{Z})=H_{n-1}(\overline{M},\mathbb{Z})=\ldots=H_{2}(\overline{M},\mathbb{Z})=0

if and only if MM is aspherical.

Proof.

First suppose that MM is aspherical. Then M¯\overline{M} is contractible and so it is immediate that (3) holds. Conversely, suppose that M¯\overline{M} satisfies (3). Since M¯\overline{M} is an nn-dimensional manifold, (3) implies that Hk(M¯,)=0H_{k}(\overline{M},\mathbb{Z})=0 for all k2k\geq 2. As M¯\overline{M} is simply connected, it then follows from the Hurewicz theorem [11, Theorem 4.32] that πk(M¯)=0\pi_{k}(\overline{M})=0 for all k1k\geq 1. The Whitehead theorem [11, Theorem 4.5] now implies that M¯\overline{M} is contractible and so MM is aspherical. ∎

Since the K(π,1)K(\pi,1) conjecture is known to be true for n{3,4,5}n\in\{3,4,5\}, we have the following immediate corollary.

Corollary 25.

Conjecture 3 is true for n{3,4,5}n\in\{3,4,5\} and m=n1m=n-1.

4.3. The Codimension Two Case

In this section, we prove the m=2m=2 case of Conjecture 3 for n{3,4,5,6}n\in\{3,4,5,6\}. Our argument combines a construction of Chodosh-Li [3] with a diameter bound of Shen-Ye [17, 18].

Theorem 26.

Let MnM^{n} be a closed manifold of dimension n{3,4,5,6}n\in\{3,4,5,6\}, Assume that the universal cover M¯\overline{M} of MM satisfies Hn(M¯,)=Hn1(M¯,)=0H_{n}(\overline{M},\mathbb{Z})=H_{n-1}(\overline{M},\mathbb{Z})=0. Then MM does not admit a metric of 22-positive intermediate curvature.

Proof.

Let MM be as in the statement of the theorem. Assume for the sake of contradiction that MM admits a metric with positive 2-intermediate curvature. By scaling, we can suppose that the 2-intermediate curvature of MM is at least 22. For a fixed large constant L>0L>0, we can apply Proposition 13 to find a geodesic line σ\sigma in M¯\overline{M} and a closed manifold Λn2\Lambda^{n-2} embedded in M¯\overline{M} such that Λ\Lambda is linked with σ\sigma and d(σ,Λ)Ld(\sigma,\Lambda)\geq L. Now let Σ1\Sigma_{1} be the area minimizing minimal hypersurface in M¯\overline{M} with Σ1=Λ\partial\Sigma_{1}=\Lambda. Then Σ1\Sigma_{1} must intersect σ\sigma and so there is a point pΣ1p\in\Sigma_{1} with d(p,Σ1)Ld(p,\partial\Sigma_{1})\geq L. For n{3,4,5}n\in\{3,4,5\}, this contradicts the Shen-Ye [17] diameter estimate for stable minimal hypersurfaces in manifolds with positive 2-intermediate curvature (Theorem 17) provided LL is chosen large enough.

When n=6n=6, we need to argue more carefully to get the diameter bound. After establishing the diameter bound, one then gets a contradiction in the same way. Choose τ<1\tau<1 close enough to 11 so that (1τ)|RicM(v,v)|<1(1-\tau)|\operatorname{Ric}_{M}(v,v)|<1 for all unit vectors vv. Let Σ1\Sigma_{1} be constructed as above, and let f>0f>0 be the first eigenfunction of the stability operator so that

ΔΣ1f+|AΣ1|2f+RicM¯(ν,ν)f0.\Delta_{\Sigma_{1}}f+|A_{\Sigma_{1}}|^{2}f+\operatorname{Ric}_{\overline{M}}(\nu,\nu)f\leq 0.

Since Σ1\Sigma_{1} is five dimensional, Theorem 20 says that if

RicΣ1(v,v)τf1ΔΣ1f+[τ(1+ε)τ2]|Σ1lnf|2κ>0\operatorname{Ric}_{\Sigma_{1}}(v,v)-\tau f^{-1}\Delta_{\Sigma_{1}}f+\left[\tau-(1+\varepsilon)\tau^{2}\right]|\nabla_{\Sigma_{1}}\ln f|^{2}\geq\kappa>0

for all unit vectors vTpΣ1v\in T_{p}\Sigma_{1}, then the diameter of Σ1\Sigma_{1} is bounded above in terms of nn, κ\kappa, and ε\varepsilon. Importantly, we can choose ε=ε(τ)\varepsilon=\varepsilon(\tau) close enough to 0 that τ(1+ε)τ20\tau-(1+\varepsilon)\tau^{2}\geq 0.

Thus it is enough to show that

RicΣ1(v,v)τf1ΔΣ1fκ>0\operatorname{Ric}_{\Sigma_{1}}(v,v)-\tau f^{-1}\Delta_{\Sigma_{1}}f\geq\kappa>0

for all unit vectors vTpΣ1v\in T_{p}\Sigma_{1}. Let e1,e2,e3,e4,e5,e6e_{1},e_{2},e_{3},e_{4},e_{5},e_{6} be an orthonormal basis for TpM¯T_{p}\overline{M} for which e2,e3,e4,e5,e6e_{2},e_{3},e_{4},e_{5},e_{6} is an orthonormal basis for TpΣ1T_{p}\Sigma_{1}. We compute

RicΣ1\displaystyle\operatorname{Ric}_{\Sigma_{1}} (e2,e2)τf1ΔΣ1f\displaystyle(e_{2},e_{2})-\tau f^{-1}\Delta_{\Sigma_{1}}f\phantom{\int}
RicΣ1(e2,e2)+τ|AΣ1|2+τRicM¯(e1,e1)\displaystyle\geq\operatorname{Ric}_{\Sigma_{1}}(e_{2},e_{2})+\tau|A_{\Sigma_{1}}|^{2}+\tau\operatorname{Ric}_{\overline{M}}(e_{1},e_{1})
=RicM¯(e2,e2)+τRicM¯(e1,e1)RM¯(e1,e2,e1,e2)+τ|AΣ1|2+j=36[A22Σ1AjjΣ1(A2jΣ1)2]\displaystyle=\operatorname{Ric}_{\overline{M}}(e_{2},e_{2})+\tau\operatorname{Ric}_{\overline{M}}(e_{1},e_{1})-R_{\overline{M}}(e_{1},e_{2},e_{1},e_{2})+\tau|A_{\Sigma_{1}}|^{2}+\sum_{j=3}^{6}\left[A^{\Sigma_{1}}_{22}A^{\Sigma_{1}}_{jj}-(A^{\Sigma_{1}}_{2j})^{2}\right]
2+(τ1)RicM¯(e1,e1)+τ|AΣ1|2+j=36[A22Σ1AjjΣ1(A2jΣ1)2]\displaystyle\geq 2+(\tau-1)\operatorname{Ric}_{\overline{M}}(e_{1},e_{1})+\tau|A_{\Sigma_{1}}|^{2}+\sum_{j=3}^{6}\left[A^{\Sigma_{1}}_{22}A^{\Sigma_{1}}_{jj}-(A^{\Sigma_{1}}_{2j})^{2}\right]
1+τ|AΣ1|2+j=36[A22Σ1AjjΣ1(A2jΣ1)2].\displaystyle\geq 1+\tau|A_{\Sigma_{1}}|^{2}+\sum_{j=3}^{6}\left[A^{\Sigma_{1}}_{22}A^{\Sigma_{1}}_{jj}-(A^{\Sigma_{1}}_{2j})^{2}\right].

It remains to show that the second fundamental form terms give a non-negative contribution. We will drop the superscript Σ1\Sigma_{1} in what follows. Since Σ1\Sigma_{1} is minimal, we have

j=36[A22AjjA2j2]\displaystyle\sum_{j=3}^{6}\left[A_{22}A_{jj}-A_{2j}^{2}\right] =i=36j=36AiiAjjj=36A2j2=23i<j6AiiAjjj=36Ajj2j=36A2j2.\displaystyle=-\sum_{i=3}^{6}\sum_{j=3}^{6}A_{ii}A_{jj}-\sum_{j=3}^{6}A_{2j}^{2}=-2\sum_{3\leq i<j\leq 6}A_{ii}A_{jj}-\sum_{j=3}^{6}A_{jj}^{2}-\sum_{j=3}^{6}A_{2j}^{2}.

Also we have

|A|2\displaystyle|A|^{2} A222+j=36Ajj2+2j=36A2j2\displaystyle\geq A_{22}^{2}+\sum_{j=3}^{6}A_{jj}^{2}+2\sum_{j=3}^{6}A_{2j}^{2}
=(j=36Ajj)2+j=36Ajj2+2j=36A2j2\displaystyle=\left(\sum_{j=3}^{6}A_{jj}\right)^{2}+\sum_{j=3}^{6}A_{jj}^{2}+2\sum_{j=3}^{6}A_{2j}^{2}
=23i<j6AiiAjj+2j=36Ajj2+2j=36A2j2.\displaystyle=2\sum_{3\leq i<j\leq 6}A_{ii}A_{jj}+2\sum_{j=3}^{6}A_{jj}^{2}+2\sum_{j=3}^{6}A_{2j}^{2}.

So assuming τ1/2\tau\geq 1/2 we get

(4) τ|A|2+j=36[A22AjjA2j2]\displaystyle\tau|A|^{2}+\sum_{j=3}^{6}\left[A_{22}A_{jj}-A_{2j}^{2}\right] (2τ2)3i<j6AiiAjj+(2τ1)j=36Ajj2.\displaystyle\geq(2\tau-2)\sum_{3\leq i<j\leq 6}A_{ii}A_{jj}+(2\tau-1)\sum_{j=3}^{6}A_{jj}^{2}.

Finally, note that

j=36Ajj2=133i<j6(Aii2+Ajj2)233i<j6|AiiAjj|.\displaystyle\sum_{j=3}^{6}A_{jj}^{2}=\frac{1}{3}\sum_{3\leq i<j\leq 6}(A_{ii}^{2}+A_{jj}^{2})\geq\frac{2}{3}\sum_{3\leq i<j\leq 6}|A_{ii}A_{jj}|.

Combining this with (4), we obtain

τ|A|2+j=36[A22AjjA2j2]0\tau|A|^{2}+\sum_{j=3}^{6}\left[A_{22}A_{jj}-A_{2j}^{2}\right]\geq 0

as long as

22τ2τ1<23,\frac{2-2\tau}{2\tau-1}<\frac{2}{3},

which we can always ensure by choosing τ\tau close enough to 1. This completes the proof.

4.4. The Case n=5n=5 and m=3m=3

Next we show that Conjecture 3 is true for n=5n=5 and m=3m=3. This will complete the proof of Theorem 4 for n{3,4,5}n\in\{3,4,5\}.

Theorem 27.

Let M5M^{5} be a closed manifold. Assume that the universal cover M¯\overline{M} of MM satisfies

H5(M¯,)=H4(M¯,)=H3(M¯,)=0.H_{5}(\overline{M},\mathbb{Z})=H_{4}(\overline{M},\mathbb{Z})=H_{3}(\overline{M},\mathbb{Z})=0.

Then MM does not admit a metric with positive 3-intermediate curvature.

Let MM be as in the statement of the theorem. Assume for the sake of contradiction that MM admits a metric with positive 33-intermediate curvature. By scaling, we can suppose the 3-intermediate curvature of MM is at least 11. For a fixed large L>0L>0, we apply Proposition 13 to find a geodesic line σ\sigma in M¯\overline{M} and a closed manifold Λn2\Lambda^{n-2} embedded in M¯\overline{M} such that Λ\Lambda is linked with σ\sigma and d(σ,Λ)2Ld(\sigma,\Lambda)\geq 2L. Let Σ1\Sigma_{1} be the area minimizing minimal hypersurface in M¯\overline{M} with Σ1=Λ\partial\Sigma_{1}=\Lambda. Let η\eta be the unit normal to Σ1\Sigma_{1}. Since Σ1\Sigma_{1} is area minimizing, there exists a function u>0u>0 on Σ1\Sigma_{1} which satisfies ΔΣ1u+(|AΣ1|2+RicM¯(η,η))u0\Delta_{\Sigma_{1}}u+(|A_{\Sigma_{1}}|^{2}+\operatorname{Ric}_{\overline{M}}(\eta,\eta))u\leq 0.

Next, we argue as in Chodosh-Li [3] to construct a suitable function hh on Σ1\Sigma_{1}.Define

ρ0(x)=dM¯(x,σ),xM¯.\rho_{0}(x)=d_{\overline{M}}(x,\sigma),\quad x\in\overline{M}.

Then let ρ1\rho_{1} be a smooth approximation to ρ0\rho_{0} which satisfies ρ1(x)2\|\nabla\rho_{1}(x)\|\leq 2 for all xM¯x\in\overline{M}, ρ1<1\rho_{1}<1 on σ\sigma, and ρ1>L+4π+1\rho_{1}>L+4\pi+1 on Σ1=Λ\partial\Sigma_{1}=\Lambda. Let ρ2\rho_{2} be the restriction of ρ1\rho_{1} to Σ1\Sigma_{1}. Now choose ε>0\varepsilon>0 sufficiently small so that LεL-\varepsilon, L+2π(1+ε)+ε2L+2\pi(1+\varepsilon)+\varepsilon^{2}, and L+4π+εL+4\pi+\varepsilon are all regular values of ρ2\rho_{2}. We can further ensure that Σ1BM¯(σ,L/4){ρ2Lε}\Sigma_{1}\cap B_{\overline{M}}(\sigma,L/4)\subset\{\rho_{2}\leq L-\varepsilon\}. More precisely, assume LL is chosen so that L/4<(Lϵ1)/2L/4<(L-\epsilon-1)/2. Then for qΣ1BM¯(σ,L/4)q\in\Sigma_{1}\cap B_{\overline{M}}(\sigma,L/4) we have

ρ2(q)=ρ1(q)supσρ1+(L/4)ρ1LLϵ,\rho_{2}(q)=\rho_{1}(q)\leq\sup_{\sigma}\rho_{1}+(L/4)\cdot\|\nabla\rho_{1}\|_{L^{\infty}}\leq L-\epsilon,

so Σ1BM¯(σ,L/4){ρ2Lε}\Sigma_{1}\cap B_{\overline{M}}(\sigma,L/4)\subset\{\rho_{2}\leq L-\varepsilon\}. Now define

ρ(x)=ρ2(x)L2π2π+ε,xΣ1.\rho(x)=\frac{\rho_{2}(x)-L-2\pi}{2\pi+\varepsilon},\quad x\in\Sigma_{1}.

Then we have Σ1ρ(x)1π\|\nabla_{\Sigma_{1}}\rho(x)\|\leq\frac{1}{\pi} for all xΣ1x\in\Sigma_{1}. We define

Σ~1={1ρ1},\displaystyle\tilde{\Sigma}_{1}=\{-1\leq\rho\leq 1\},
+Σ~1={ρ=1},\displaystyle\partial_{+}\tilde{\Sigma}_{1}=\{\rho=-1\},
Σ~1={ρ=1},\displaystyle\partial_{-}\tilde{\Sigma}_{1}=\{\rho=1\},
Ω0={1ρ<ε}.\displaystyle\Omega_{0}=\{-1\leq\rho<\varepsilon\}.

By construction, ±Σ~1\partial_{\pm}\tilde{\Sigma}_{1} and Ω0\partial\Omega_{0} are all smooth hypersurfaces in Σ~1\tilde{\Sigma}_{1}, and Ω0\Omega_{0} contains +Σ~1\partial_{+}\tilde{\Sigma}_{1}, and Σ1Σ~1=\partial\Sigma_{1}\cap\tilde{\Sigma}_{1}=\emptyset. Finally, we define

h(x)=tan(πρ(x)2).h(x)=-\tan\left(\frac{\pi\rho(x)}{2}\right).

Then hh is a smooth function on the interior of Σ~1\tilde{\Sigma}_{1} which satisfies h(x)±h(x)\to\pm\infty as x±Σ~1x\to\partial_{\pm}\tilde{\Sigma}_{1}. Moreover, we have

Σ1h(x)=π2sec2(πρ(x)2)Σ1ρ(x)12(1+tan2(πρ(x)2))=12(1+h2)\|\nabla_{\Sigma_{1}}h(x)\|=\frac{\pi}{2}\sec^{2}\left(\frac{\pi\rho(x)}{2}\right)\|\nabla_{\Sigma_{1}}\rho(x)\|\leq\frac{1}{2}\left(1+\tan^{2}\left(\frac{\pi\rho(x)}{2}\right)\right)=\frac{1}{2}\left(1+h^{2}\right)

for all xΣ~1x\in\tilde{\Sigma}_{1}. It follows that

(5) 1+h22Σ1h01+h^{2}-2\|\nabla_{\Sigma_{1}}h\|\geq 0

on Σ~1\tilde{\Sigma}_{1}. Proposition 15 implies that there exists a minimizer ΩΣ~1\Omega\subset\tilde{\Sigma}_{1} of the μ\mu-bubble functional

𝒜(Ω)=Ωu𝑑3Σ~1(χΩχΩ0)hu𝑑4\mathcal{A}(\Omega)=\int_{\partial\Omega}u\,d\mathcal{H}^{3}-\int_{\tilde{\Sigma}_{1}}(\chi_{\Omega}-\chi_{\Omega_{0}})hu\,d\mathcal{H}^{4}

such that ΩΔΩ0\Omega\operatorname{\Delta}\Omega_{0} is compactly contained in the interior of Σ~1\tilde{\Sigma}_{1}.

Let Σ2\Sigma_{2} be a connected component of Ω\partial\Omega, and consider the slicing Σ2Σ1M¯\Sigma_{2}\subset\Sigma_{1}\subset\overline{M} together with the functions uu and hh.

We are going to show that diam(Σ2)\operatorname{diam}(\Sigma_{2}) is bounded from above by a universal constant that does not depend on LL. Let ν\nu be the unit normal vector to Σ2\Sigma_{2} in Σ1\Sigma_{1} pointing out of Ω\Omega. According to Proposition 16, the mean curvature of Σ2\Sigma_{2} satisfies

HΣ2=hu1Σ1u,ν.H_{\Sigma_{2}}=h-u^{-1}\langle\nabla_{\Sigma_{1}}u,\nu\rangle.

In the next proposition, we re-arrange the second variation formula into a more convenient form.

Proposition 28.

For any ψC1(Σ2)\psi\in C^{1}(\Sigma_{2}) we have

0Σ2|Σ2ψ|2u(|AΣ1|2+RicM¯(η,η)+|AΣ2|2+RicΣ1(ν,ν)12HΣ2212+ΔΣ2uu)ψ2u\displaystyle 0\leq\int_{\Sigma_{2}}|\nabla_{\Sigma_{2}}\psi|^{2}u-\left(|A_{\Sigma_{1}}|^{2}+\operatorname{Ric}_{\overline{M}}(\eta,\eta)+|A_{\Sigma_{2}}|^{2}+\operatorname{Ric}_{\Sigma_{1}}(\nu,\nu)-\frac{1}{2}H_{\Sigma_{2}}^{2}-\frac{1}{2}+\frac{\Delta_{\Sigma_{2}}u}{u}\right)\psi^{2}u
Proof.

The second variation formula (Proposition 16) and HΣ2=hu1Σ1u,νH_{\Sigma_{2}}=h-u^{-1}\langle\nabla_{\Sigma_{1}}u,\nu\rangle give that

0Σ2|Σ2ψ|2u(|AΣ2|2+RicΣ1(ν,ν))ψ2uΣ1h,νψ2uhΣ1u,νψ2+(ΔΣ1uΔΣ2u)ψ2.0\leq\int_{\Sigma_{2}}|\nabla_{\Sigma_{2}}\psi|^{2}u-(|A_{\Sigma_{2}}|^{2}+\operatorname{Ric}_{\Sigma_{1}}(\nu,\nu))\psi^{2}u-\langle\nabla_{\Sigma_{1}}h,\nu\rangle\psi^{2}u-h\langle\nabla_{\Sigma_{1}}u,\nu\rangle\psi^{2}+(\Delta_{\Sigma_{1}}u-\Delta_{\Sigma_{2}}u)\psi^{2}.

We now simplify this as in Chodosh-Li [3]. First, observe that

12HΣ22ψ2u=12u1Σ1u,ν2ψ2hΣ1u,νψ2+12h2ψ2u.\frac{1}{2}H_{\Sigma_{2}}^{2}\psi^{2}u=\frac{1}{2}u^{-1}\langle\nabla_{\Sigma_{1}}u,\nu\rangle^{2}\psi^{2}-h\langle\nabla_{\Sigma_{1}}u,\nu\rangle\psi^{2}+\frac{1}{2}h^{2}\psi^{2}u.

Using this to eliminate the hΣ1u,νψ2h\langle\nabla_{\Sigma_{1}}u,\nu\rangle\psi^{2} term in the previous inequality, we see that

0\displaystyle 0 Σ2|Σ2ψ|2u(|AΣ2|2+RicΣ1(ν,ν)12HΣ22)ψ2u+(ΔΣ1uΔΣ2u)ψ2\displaystyle\leq\int_{\Sigma_{2}}|\nabla_{\Sigma_{2}}\psi|^{2}u-(|A_{\Sigma_{2}}|^{2}+\operatorname{Ric}_{\Sigma_{1}}(\nu,\nu)-\frac{1}{2}H_{\Sigma_{2}}^{2})\psi^{2}u+(\Delta_{\Sigma_{1}}u-\Delta_{\Sigma_{2}}u)\psi^{2}
12Σ2(u1Σ1u,ν2+h2+2Σ1h,ν)ψ2u.\displaystyle\qquad\qquad-\frac{1}{2}\int_{\Sigma_{2}}\left(\langle u^{-1}\nabla_{\Sigma_{1}}u,\nu\rangle^{2}+h^{2}+2\langle\nabla_{\Sigma_{1}}h,\nu\rangle\right)\psi^{2}u.

Since u1Σ1u,ν20\langle u^{-1}\nabla_{\Sigma_{1}}u,\nu\rangle^{2}\geq 0 and 1+h2+2Σ1h,ν01+h^{2}+2\langle\nabla_{\Sigma_{1}}h,\nu\rangle\geq 0 by (5), it follows that

0Σ2|Σ2ψ|2u(|AΣ2|2+RicΣ1(ν,ν)12HΣ2212)ψ2u+(ΔΣ1uΔΣ2u)ψ2.\displaystyle 0\leq\int_{\Sigma_{2}}|\nabla_{\Sigma_{2}}\psi|^{2}u-\left(|A_{\Sigma_{2}}|^{2}+\operatorname{Ric}_{\Sigma_{1}}(\nu,\nu)-\frac{1}{2}H_{{\Sigma_{2}}}^{2}-\frac{1}{2}\right)\psi^{2}u+(\Delta_{\Sigma_{1}}u-\Delta_{\Sigma_{2}}u)\psi^{2}.

Finally, recalling that ΔΣ1u+(|AΣ1|2+RicM¯(η,η))u0\Delta_{\Sigma_{1}}u+(|A_{\Sigma_{1}}|^{2}+\operatorname{Ric}_{\overline{M}}(\eta,\eta))u\leq 0, we get

0Σ2|Σ2ψ|2u(|AΣ1|2+RicM¯(η,η)+|AΣ2|2+RicΣ1(ν,ν)12HΣ2212+ΔΣ2uu)ψ2u,0\leq\int_{\Sigma_{2}}|\nabla_{\Sigma_{2}}\psi|^{2}u-\left(|A_{\Sigma_{1}}|^{2}+\operatorname{Ric}_{\overline{M}}(\eta,\eta)+|A_{\Sigma_{2}}|^{2}+\operatorname{Ric}_{\Sigma_{1}}(\nu,\nu)-\frac{1}{2}H_{\Sigma_{2}}^{2}-\frac{1}{2}+\frac{\Delta_{\Sigma_{2}}u}{u}\right)\psi^{2}u,

as needed. ∎

Now we can argue that the diameter of Σ2{\Sigma_{2}} is bounded.

Proposition 29.

The diameter diam(Σ2)\operatorname{diam}({\Sigma_{2}}) of Σ2\Sigma_{2} is bounded uniformly from above.

Proof.

Note that Proposition 28 implies that there is a function w>0w>0 on Σ2{\Sigma_{2}} which satisfies

(6) divΣ2(uΣ2w)(|AΣ2|2+RicΣ1(ν,ν)+|AΣ1|2+RicM¯(η,η)12HΣ2212+ΔΣ2uu)wu.\operatorname{div}_{\Sigma_{2}}(u\nabla_{\Sigma_{2}}w)\leq-\left(|A_{\Sigma_{2}}|^{2}+\operatorname{Ric}_{\Sigma_{1}}(\nu,\nu)+|A_{\Sigma_{1}}|^{2}+\operatorname{Ric}_{\overline{M}}(\eta,\eta)-\frac{1}{2}H_{{\Sigma_{2}}}^{2}-\frac{1}{2}+\frac{\Delta_{\Sigma_{2}}u}{u}\right)wu.

Our strategy is to apply Theorem 18 with N=Σ2N={\Sigma_{2}} and f=uwf=uw.

Fix a point pΣ2p\in{\Sigma_{2}} and let {e1=η,e2=ν,e3,e4,e5}\{e_{1}=\eta,e_{2}=\nu,e_{3},e_{4},e_{5}\} be an orthonormal frame at pp such that {e2,e3,e4,e5}\{e_{2},e_{3},e_{4},e_{5}\} is an orthonormal basis for Σ1{\Sigma_{1}} and {e3,e4,e5}\{e_{3},e_{4},e_{5}\} is an orthonormal basis for Σ2{\Sigma_{2}}. We compute that

RicΣ2(e3,e3)ΔΣ2(uw)uw+12|Σ2log(uw)|2\displaystyle\operatorname{Ric}_{\Sigma_{2}}(e_{3},e_{3})-\frac{\Delta_{\Sigma_{2}}(uw)}{uw}+\frac{1}{2}|\nabla_{\Sigma_{2}}\log(uw)|^{2}
=RicΣ2(e3,e3)divΣ2(uΣ2w)uwdivΣ2(wΣ2u)uw+12|Σ2uu+Σ2ww|2\displaystyle\qquad\qquad=\operatorname{Ric}_{\Sigma_{2}}(e_{3},e_{3})-\frac{\operatorname{div}_{\Sigma_{2}}(u\nabla_{\Sigma_{2}}w)}{uw}-\frac{\operatorname{div}_{\Sigma_{2}}(w\nabla_{\Sigma_{2}}u)}{uw}+\frac{1}{2}\left|\frac{\nabla_{\Sigma_{2}}u}{u}+\frac{\nabla_{\Sigma_{2}}w}{w}\right|^{2}
RicΣ2(e3,e3)divΣ2(uΣ2w)uwΔΣ2uu.\displaystyle\qquad\qquad\geq\operatorname{Ric}_{\Sigma_{2}}(e_{3},e_{3})-\frac{\operatorname{div}_{\Sigma_{2}}(u\nabla_{\Sigma_{2}}w)}{uw}-\frac{\Delta_{\Sigma_{2}}u}{u}.

Therefore, by inequality (6), we obtain

RicΣ2(e3,e3)ΔΣ2(uw)uw+12|Σ2log(uw)|2\displaystyle\operatorname{Ric}_{\Sigma_{2}}(e_{3},e_{3})-\frac{\Delta_{\Sigma_{2}}(uw)}{uw}+\frac{1}{2}|\nabla_{\Sigma_{2}}\log(uw)|^{2}
RicM¯(e1,e1)+RicΣ1(e2,e2)+RicΣ2(e3,e3)+|AΣ1|2+|AΣ2|212HΣ2212.\displaystyle\qquad\qquad\geq\operatorname{Ric}_{\overline{M}}(e_{1},e_{1})+\operatorname{Ric}_{\Sigma_{1}}(e_{2},e_{2})+\operatorname{Ric}_{\Sigma_{2}}(e_{3},e_{3})+|A_{\Sigma_{1}}|^{2}+|A_{\Sigma_{2}}|^{2}-\frac{1}{2}H_{\Sigma_{2}}^{2}-\frac{1}{2}.

It remains to get a lower bound on the right hand side.

According to [2, Lemma 3.8], we have

RicM¯(e1,e1)+RicΣ1(e2,e2)+RicΣ2(e3,e3)=C3(e1,e2,e3)+\displaystyle\operatorname{Ric}_{\overline{M}}(e_{1},e_{1})+\operatorname{Ric}_{\Sigma_{1}}(e_{2},e_{2})+\operatorname{Ric}_{\Sigma_{2}}(e_{3},e_{3})=C_{3}(e_{1},e_{2},e_{3})+\mathcal{B}

where

=p=23q=p+15(AΣ1(ep,ep)AΣ1(eq,eq)AΣ1(ep,eq)2)+q=45(AΣ2(e3,e3)AΣ2(eq,eq)AΣ2(e3,eq)2).\mathcal{B}=\sum_{p=2}^{3}\sum_{q=p+1}^{5}\big{(}A_{\Sigma_{1}}(e_{p},e_{p})A_{\Sigma_{1}}(e_{q},e_{q})-A_{{\Sigma_{1}}}(e_{p},e_{q})^{2}\big{)}+\sum_{q=4}^{5}\big{(}A_{\Sigma_{2}}(e_{3},e_{3})A_{\Sigma_{2}}(e_{q},e_{q})-A_{\Sigma_{2}}(e_{3},e_{q})^{2}\big{)}.

Now, since Σ1{\Sigma_{1}} is minimal, [2, Lemma 3.11] implies that

|AΣ1|2+p=23q=p+15(AΣ1(ep,ep)AΣ1(eq,eq)AΣ1(ep,eq)2)0.|A_{\Sigma_{1}}|^{2}+\sum_{p=2}^{3}\sum_{q=p+1}^{5}\big{(}A_{\Sigma_{1}}(e_{p},e_{p})A_{\Sigma_{1}}(e_{q},e_{q})-A_{{\Sigma_{1}}}(e_{p},e_{q})^{2}\big{)}\geq 0.

Moreover, we have

|AΣ2|212HΣ22+q=45(AΣ2(e3,e3)AΣ2(eq,eq)AΣ2(e3,eq)2)\displaystyle|A_{\Sigma_{2}}|^{2}-\frac{1}{2}H_{\Sigma_{2}}^{2}+\sum_{q=4}^{5}\big{(}A_{\Sigma_{2}}(e_{3},e_{3})A_{\Sigma_{2}}(e_{q},e_{q})-A_{\Sigma_{2}}(e_{3},e_{q})^{2}\big{)}
=12(A33Σ2)2+12(A44Σ2)2+12(A55Σ2)2A44Σ2A55Σ2+(A34Σ2)2+(A35Σ2)2+2(A45Σ2)20.\displaystyle\qquad=\frac{1}{2}(A^{\Sigma_{2}}_{33})^{2}+\frac{1}{2}(A^{\Sigma_{2}}_{44})^{2}+\frac{1}{2}(A^{\Sigma_{2}}_{55})^{2}-A^{\Sigma_{2}}_{44}A^{\Sigma_{2}}_{55}+(A^{\Sigma_{2}}_{34})^{2}+(A^{\Sigma_{2}}_{35})^{2}+2(A^{\Sigma_{2}}_{45})^{2}\geq 0.

Hence we obtain

RicΣ2(e3,e3)ΔΣ2(uw)uw+12|Σ2log(uw)|2\displaystyle\operatorname{Ric}_{\Sigma_{2}}(e_{3},e_{3})-\frac{\Delta_{\Sigma_{2}}(uw)}{uw}+\frac{1}{2}|\nabla_{\Sigma_{2}}\log(uw)|^{2}
C3(e1,e2,e3)++|AΣ1|2+|AΣ2|212HΣ2212\displaystyle\qquad\qquad\geq C_{3}(e_{1},e_{2},e_{3})+\mathcal{B}+|A_{\Sigma_{1}}|^{2}+|A_{\Sigma_{2}}|^{2}-\frac{1}{2}H_{\Sigma_{2}}^{2}-\frac{1}{2}
C3(e1,e2,e3)1212.\displaystyle\qquad\qquad\geq C_{3}(e_{1},e_{2},e_{3})-\frac{1}{2}\geq\frac{1}{2}.

Since e3e_{3} was an arbitrary unit tangent vector to Σ2{\Sigma_{2}} and Σ2\Sigma_{2} is 3-dimensional, Shen-Ye’s generalized Bonnet-Myers theorem (Theorem 18) now gives the conclusion. ∎

We can now finish the proof of Theorem 27. Since Σ2=0\partial{\Sigma_{2}}=0 and diam(Σ2)\operatorname{diam}({\Sigma_{2}}) is uniformly bounded and H3(M¯,)=0H_{3}(\overline{M},\mathbb{Z})=0, it follows from Proposition 14 that Σ2{\Sigma_{2}} bounds within an RR-neighborhood for some constant RR that depends only on M¯\overline{M}. Applying this argument to each component of Ω\partial\Omega, we see that Ω\partial\Omega bounds within its RR-neighborhood in M¯\overline{M}. Since d(Ω,σ)L/4d(\partial\Omega,\sigma)\geq L/4 and Ω\partial\Omega is linked with σ\sigma by construction, this is a contradiction provided we select L>2RL>2R. This completes the proof.

4.5. The Case n=6n=6 and m=4m=4

Next we prove the n=6n=6 and m=4m=4 case of Theorem 4. Our argument combines the slice and dice procedure of Chodosh-Li [3], the diameter bound of Shen-Ye [17], and also the Frankel type theorems we developed in Section 3.

Theorem 30.

Let M6M^{6} be a closed 6-dimensional manifold. Assume the universal cover M¯\overline{M} of MM satisfies H6(M¯,)=H5(M¯,)=H4(M¯,)=H3(M¯,)=0H_{6}(\overline{M},\mathbb{Z})=H_{5}(\overline{M},\mathbb{Z})=H_{4}(\overline{M},\mathbb{Z})=H_{3}(\overline{M},\mathbb{Z})=0. Then MM does not admit a metric of positive 44-intermediate curvature.

For n=6n=6, m=4m=4, let Σ2Σ1M¯6\Sigma_{2}\subset\Sigma_{1}\subset\overline{M}^{6} be constructed as in the previous subsection, which are now of dimension 44 and 55. We consider the following weighted functional for 33-dimensional Σ3Σ2\Sigma_{3}\subset{\Sigma_{2}}:

𝒜2(Σ3):=Σ3uw.\mathcal{A}_{2}(\Sigma_{3}):=\int_{\Sigma_{3}}uw.

We will use the 𝒜2\mathcal{A}_{2}-functional to slice Σ2{\Sigma_{2}} into a 44-manifold Σ^2\hat{{\Sigma}}_{2} with simple third homology. Firstly, we prove that each boundary component of Σ^2\hat{{\Sigma}}_{2} has finite diameter. Without loss of generality, we assume the 44-intermediate curvature of MM is at least 32\frac{3}{2}.

Lemma 31.

Suppose Σ3\Sigma_{3} is a connected (two-sided) stable critical point of 𝒜2\mathcal{A}_{2}. Then diamΣ32π\operatorname{diam}\Sigma_{3}\leq 2\pi and H1(Σ3)H_{1}(\Sigma_{3}) is finite.

Proof.

By the first variation formula, we have

HΣ3=Σ2log(uw),ω,H_{\Sigma_{3}}=-\langle\nabla_{{\Sigma_{2}}}\log(uw),\omega\rangle,

where HΣ3H_{\Sigma_{3}} is the mean curvature of Σ3\Sigma_{3} with respect to the unit normal ω\omega of Σ3\Sigma_{3} in Σ2{\Sigma_{2}}. As before, let {ei}i=16\{e_{i}\}_{i=1}^{6} be an orthonormal basis of TpM¯T_{p}\overline{M} so that ejΣje_{j}\perp\Sigma_{j} for j=1,2,3j=1,2,3, i.e. e1=η,e2=ν,e3=ωe_{1}=\eta,e_{2}=\nu,e_{3}=\omega.

By the second variation formula for 𝒜2\mathcal{A}_{2}, there exists a positive function vC(Σ3)v\in C^{\infty}(\Sigma_{3}) such that

v1ΔΣ3vAΣ32+RicΣ2(ω,ω)Σ22log(uw)(ω,ω)+Σ3log(uw),Σ3logv.-v^{-1}\Delta_{\Sigma_{3}}v\geq\|A_{\Sigma_{3}}\|^{2}+\operatorname{Ric}_{{\Sigma_{2}}}(\omega,\omega)-\nabla_{{\Sigma_{2}}}^{2}\log(uw)(\omega,\omega)+\langle\nabla_{\Sigma_{3}}\log(uw),\nabla_{\Sigma_{3}}\log v\rangle.

We let f=uwvf=uwv, and our aim is to prove

RicΣ3(e,e)f1ΔΣ3f+12|Σ3logf|212>0,\operatorname{Ric}_{\Sigma_{3}}(e,e)-f^{-1}\Delta_{\Sigma_{3}}f+\frac{1}{2}|\nabla_{\Sigma_{3}}\log f|^{2}\geq\frac{1}{2}>0,

for any unit vector eTΣ3e\in T\Sigma_{3}.

Suppose f2f_{2} is a smooth function defined on Σ2{\Sigma_{2}}. Then we have

(7) ΔΣ2f2=ΔΣ3f2+Σ22f2(ω,ω)Σ2f2,ωHΣ3.\displaystyle\Delta_{{\Sigma_{2}}}f_{2}=\Delta_{\Sigma_{3}}f_{2}+\nabla_{{\Sigma_{2}}}^{2}f_{2}(\omega,\omega)-\langle\nabla_{{\Sigma_{2}}}f_{2},\omega\rangle H_{\Sigma_{3}}.

By (6), we have

(uw)1\displaystyle(uw)^{-1} ΔΣ2(uw)\displaystyle\Delta_{{\Sigma_{2}}}(uw)
\displaystyle\leq AΣ12RicM¯(η,η)AΣ22RicΣ1(ν,ν)+12HΣ22+Σ2logu,Σ2logw+12,\displaystyle-\|A_{{\Sigma_{1}}}\|^{2}-\operatorname{Ric}_{\overline{M}}(\eta,\eta)-\|A_{{\Sigma_{2}}}\|^{2}-\operatorname{Ric}_{{\Sigma_{1}}}(\nu,\nu)+\frac{1}{2}H_{{\Sigma_{2}}}^{2}+\langle\nabla_{{\Sigma_{2}}}\log u,\nabla_{{\Sigma_{2}}}\log w\rangle+\frac{1}{2},
=\displaystyle= AΣ12RicM¯(η,η)AΣ22RicΣ1(ν,ν)+12HΣ22+12\displaystyle-\|A_{{\Sigma_{1}}}\|^{2}-\operatorname{Ric}_{\overline{M}}(\eta,\eta)-\|A_{{\Sigma_{2}}}\|^{2}-\operatorname{Ric}_{{\Sigma_{1}}}(\nu,\nu)+\frac{1}{2}H_{{\Sigma_{2}}}^{2}+\frac{1}{2}
+Σ2logu,ωΣ2logw,ω+Σ3logu,Σ3logw.\displaystyle+\langle\nabla_{{\Sigma_{2}}}\log u,\omega\rangle\langle\nabla_{{\Sigma_{2}}}\log w,\omega\rangle+\langle\nabla_{\Sigma_{3}}\log u,\nabla_{\Sigma_{3}}\log w\rangle.

Using (7) and the first variation formula, we can further write

(uw)1ΔΣ3(uw)=\displaystyle(uw)^{-1}\Delta_{\Sigma_{3}}(uw)= (uw)1ΔΣ2(uw)Σ22log(uw)(ω,ω)\displaystyle(uw)^{-1}\Delta_{{\Sigma_{2}}}(uw)-\nabla_{{\Sigma_{2}}}^{2}\log(uw)(\omega,\omega)
\displaystyle\leq Σ22log(uw)(ω,ω)+Σ2logu,ωΣ2logw,ω+Σ3logu,Σ3logw\displaystyle-\nabla_{{\Sigma_{2}}}^{2}\log(uw)(\omega,\omega)+\langle\nabla_{{\Sigma_{2}}}\log u,\omega\rangle\langle\nabla_{{\Sigma_{2}}}\log w,\omega\rangle+\langle\nabla_{\Sigma_{3}}\log u,\nabla_{\Sigma_{3}}\log w\rangle
AΣ12RicM¯(η,η)AΣ22RicΣ1(ν,ν)+12HΣ22+12.\displaystyle-\|A_{{\Sigma_{1}}}\|^{2}-\operatorname{Ric}_{\overline{M}}(\eta,\eta)-\|A_{{\Sigma_{2}}}\|^{2}-\operatorname{Ric}_{{\Sigma_{1}}}(\nu,\nu)+\frac{1}{2}H_{{\Sigma_{2}}}^{2}+\frac{1}{2}.

We now have that

(uwv)1ΔΣ3(uwv)=\displaystyle(uwv)^{-1}\Delta_{\Sigma_{3}}(uwv)= (uw)1ΔΣ3(uw)+v1ΔΣ3v+2Σ3log(uw),Σ3logv\displaystyle(uw)^{-1}\Delta_{\Sigma_{3}}(uw)+v^{-1}\Delta_{\Sigma_{3}}v+2\langle\nabla_{\Sigma_{3}}\log(uw),\nabla_{\Sigma_{3}}\log v\rangle
\displaystyle\leq Σ2logu,ωΣ2logw,ω+Σ3logu,Σ3logw+Σ3log(uw),Σ1logv\displaystyle\langle\nabla_{{\Sigma_{2}}}\log u,\omega\rangle\langle\nabla_{{\Sigma_{2}}}\log w,\omega\rangle+\langle\nabla_{\Sigma_{3}}\log u,\nabla_{\Sigma_{3}}\log w\rangle+\langle\nabla_{{\Sigma_{3}}}\log(uw),\nabla_{{\Sigma_{1}}}\log v\rangle
AΣ12RicM¯(η,η)AΣ22RicΣ1(ν,ν)AΣ32RicΣ2(ω,ω)+12HΣ22+12\displaystyle-\|A_{{\Sigma_{1}}}\|^{2}-\operatorname{Ric}_{\overline{M}}(\eta,\eta)-\|A_{{\Sigma_{2}}}\|^{2}-\operatorname{Ric}_{{\Sigma_{1}}}(\nu,\nu)-\|A_{\Sigma_{3}}\|^{2}-\operatorname{Ric}_{{\Sigma_{2}}}(\omega,\omega)+\frac{1}{2}H_{{\Sigma_{2}}}^{2}+\frac{1}{2}
\displaystyle\leq 12(Σ2logu,ω+Σ2logw,ω)2+Σ3logu,Σ3logw+Σ3log(uw),Σ1logv\displaystyle\frac{1}{2}\left(\langle\nabla_{{\Sigma_{2}}}\log u,\omega\rangle+\langle\nabla_{{\Sigma_{2}}}\log w,\omega\rangle\right)^{2}+\langle\nabla_{\Sigma_{3}}\log u,\nabla_{\Sigma_{3}}\log w\rangle+\langle\nabla_{{\Sigma_{3}}}\log(uw),\nabla_{{\Sigma_{1}}}\log v\rangle
AΣ12RicM¯(η,η)AΣ22RicΣ1(ν,ν)AΣ32RicΣ2(ω,ω)+12HΣ22+12\displaystyle-\|A_{{\Sigma_{1}}}\|^{2}-\operatorname{Ric}_{\overline{M}}(\eta,\eta)-\|A_{{\Sigma_{2}}}\|^{2}-\operatorname{Ric}_{{\Sigma_{1}}}(\nu,\nu)-\|A_{\Sigma_{3}}\|^{2}-\operatorname{Ric}_{{\Sigma_{2}}}(\omega,\omega)+\frac{1}{2}H_{{\Sigma_{2}}}^{2}+\frac{1}{2}
=\displaystyle= Σ3logu,Σ3logw+Σ3log(uw),Σ1logv+12HΣ22+12HΣ32+12\displaystyle\langle\nabla_{\Sigma_{3}}\log u,\nabla_{\Sigma_{3}}\log w\rangle+\langle\nabla_{{\Sigma_{3}}}\log(uw),\nabla_{{\Sigma_{1}}}\log v\rangle+\frac{1}{2}H_{{\Sigma_{2}}}^{2}+\frac{1}{2}H^{2}_{\Sigma_{3}}+\frac{1}{2}
AΣ12RicM¯(η,η)AΣ22RicΣ1(ν,ν)AΣ32RicΣ2(ω,ω).\displaystyle-\|A_{{\Sigma_{1}}}\|^{2}-\operatorname{Ric}_{\overline{M}}(\eta,\eta)-\|A_{{\Sigma_{2}}}\|^{2}-\operatorname{Ric}_{{\Sigma_{1}}}(\nu,\nu)-\|A_{\Sigma_{3}}\|^{2}-\operatorname{Ric}_{{\Sigma_{2}}}(\omega,\omega).

We used the first variation in the last equality.

Summarizing the above computations, we have

RicΣ3(e4,e4)\displaystyle\operatorname{Ric}_{\Sigma_{3}}(e_{4},e_{4}) (uwv)1ΔΣ3(uwv)+12|Σ3log(uwv)|2\displaystyle-(uwv)^{-1}\Delta_{\Sigma_{3}}(uwv)+\frac{1}{2}|\nabla_{\Sigma_{3}}\log(uwv)|^{2}
\displaystyle\geq AΣ12+AΣ22+AΣ3212HΣ2212HΣ3212\displaystyle\|A_{{\Sigma_{1}}}\|^{2}+\|A_{{\Sigma_{2}}}\|^{2}+\|A_{\Sigma_{3}}\|^{2}-\frac{1}{2}H_{{\Sigma_{2}}}^{2}-\frac{1}{2}H_{\Sigma_{3}}^{2}-\frac{1}{2}
+RicΣ3(e4,e4)+RicΣ2(ω,ω)+RicΣ1(ν,ν)+RicM¯(η,η)\displaystyle+\operatorname{Ric}_{\Sigma_{3}}(e_{4},e_{4})+\operatorname{Ric}_{{\Sigma_{2}}}(\omega,\omega)+\operatorname{Ric}_{{\Sigma_{1}}}(\nu,\nu)+\operatorname{Ric}_{\overline{M}}(\eta,\eta)
=\displaystyle= AΣ12+AΣ22+AΣ3212HΣ2212HΣ3212\displaystyle\|A_{{\Sigma_{1}}}\|^{2}+\|A_{{\Sigma_{2}}}\|^{2}+\|A_{\Sigma_{3}}\|^{2}-\frac{1}{2}H_{{\Sigma_{2}}}^{2}-\frac{1}{2}H_{\Sigma_{3}}^{2}-\frac{1}{2}
+C4(e1,e2,e3,e4)+p=24q=p+16(AΣ1(ep,ep)AΣ1(eq,eq)AΣ1(ep,eq)2)\displaystyle+C_{4}(e_{1},e_{2},e_{3},e_{4})+\sum_{p=2}^{4}\sum_{q=p+1}^{6}\left(A_{{\Sigma_{1}}}(e_{p},e_{p})A_{{\Sigma_{1}}}(e_{q},e_{q})-A_{{\Sigma_{1}}}(e_{p},e_{q})^{2}\right)
+p=34q=p+16(AΣ2(ep,ep)AΣ2(eq,eq)AΣ2(ep,eq)2)\displaystyle+\sum_{p=3}^{4}\sum_{q=p+1}^{6}\left(A_{{\Sigma_{2}}}(e_{p},e_{p})A_{{\Sigma_{2}}}(e_{q},e_{q})-A_{{\Sigma_{2}}}(e_{p},e_{q})^{2}\right)
+q=56(AΣ3(e4,e4)AΣ3(eq,eq)AΣ3(e4,eq)2)\displaystyle+\sum_{q=5}^{6}\left(A_{\Sigma_{3}}(e_{4},e_{4})A_{\Sigma_{3}}(e_{q},e_{q})-A_{\Sigma_{3}}(e_{4},e_{q})^{2}\right)
\displaystyle\geq C4(e1,e2,e3,e4)121.\displaystyle C_{4}(e_{1},e_{2},e_{3},e_{4})-\frac{1}{2}\geq 1.

To get the last line above, we used the following estimates. First, by [2, Lemma 3.11], we have

AΣ12+p=24p=q+16(AΣ1(ep,ep)AΣ1(eq,eq)AΣ1(ep,eq)2)0.\|A_{\Sigma_{1}}\|^{2}+\sum_{p=2}^{4}\sum_{p=q+1}^{6}\left(A_{\Sigma_{1}}(e_{p},e_{p})A_{\Sigma_{1}}(e_{q},e_{q})-A_{\Sigma_{1}}(e_{p},e_{q})^{2}\right)\geq 0.

Moreover, via direct computations we have

AΣ22\displaystyle\|A_{\Sigma_{2}}\|^{2} 12HΣ22+p=34q=p+16(AΣ2(ep,ep)AΣ2(eq,eq)AΣ2(ep,eq)2)\displaystyle-\frac{1}{2}H_{\Sigma_{2}}^{2}+\sum_{p=3}^{4}\sum_{q=p+1}^{6}\left(A_{\Sigma_{2}}(e_{p},e_{p})A_{\Sigma_{2}}(e_{q},e_{q})-A_{\Sigma_{2}}(e_{p},e_{q})^{2}\right)
=12p=36(AppΣ2)2+p=34q=p+16(ApqΣ2)2+2(A56Σ2)2A55Σ2A66Σ20,\displaystyle=\frac{1}{2}\sum_{p=3}^{6}(A^{\Sigma_{2}}_{pp})^{2}+\sum_{p=3}^{4}\sum_{q=p+1}^{6}(A_{pq}^{\Sigma_{2}})^{2}+2(A_{56}^{\Sigma_{2}})^{2}-A^{\Sigma_{2}}_{55}A_{66}^{\Sigma_{2}}\geq 0,

and similarly

AΣ32\displaystyle\|A_{\Sigma_{3}}\|^{2} 12HΣ32+q=56(AΣ3(e4,e4)AΣ3(eq,eq)AΣ3(e4,eq)2)\displaystyle-\frac{1}{2}H_{\Sigma_{3}}^{2}+\sum_{q=5}^{6}\left(A_{\Sigma_{3}}(e_{4},e_{4})A_{\Sigma_{3}}(e_{q},e_{q})-A_{\Sigma_{3}}(e_{4},e_{q})^{2}\right)
=12p=46(AppΣ3)2+q=56(A4qΣ3)2+2(A56Σ3)2A55Σ3A66Σ30.\displaystyle=\frac{1}{2}\sum_{p=4}^{6}(A_{pp}^{\Sigma_{3}})^{2}+\sum_{q=5}^{6}(A_{4q}^{\Sigma_{3}})^{2}+2(A_{56}^{\Sigma_{3}})^{2}-A_{55}^{\Sigma_{3}}A_{66}^{\Sigma_{3}}\geq 0.

By the above computation, we know Σ3\Sigma_{3} admits a metric with positive conformal Ricci curvature, and hence it has finite diameter; see Theorem 18. As a direct corollary, the fundamental group of Σ3\Sigma_{3} is finite and hence H1(Σ3)H_{1}(\Sigma_{3}) is finite. ∎

We need the following slicing Lemma by Bamler-Li-Mantoulidis [1], which is a generalization of Chodosh-Li [3, Lemma 20].

Lemma 32.

There are Σ3,1,Σ3,2,,Σ3,kΣ2\Sigma_{3,1},\Sigma_{3,2},\dots,\Sigma_{3,k}\subset{\Sigma_{2}} pairwise disjoint two-sided stable critical points of 𝒜2\mathcal{A}_{2} so that the manifold with boundary Σ^2:=Σ2(i=1kΣ3,i)\hat{\Sigma}_{2}:={\Sigma_{2}}\setminus(\cup_{i=1}^{k}\Sigma_{3,i}) is connected and has H3(Σ^2)H3(Σ^2)H_{3}(\partial\hat{{\Sigma}}_{2})\to H_{3}(\hat{{\Sigma}}_{2}) surjective.

Proof.

Suppose we have constructed Σ3,1,Σ3,2,,Σ3,j\Sigma_{3,1},\Sigma_{3,2},\cdots,\Sigma_{3,j} which are pairwise disjoint two-sided stable critical points of 𝒜2\mathcal{A}_{2} and Mj:=Σ2(i=1jΣ3,i)M_{j}:=\Sigma_{2}\setminus(\cup_{i=1}^{j}\Sigma_{3,i}) is connected. If the inclusion map i:H3(Mj)H3(Mj)i:H_{3}(\partial M_{j})\to H_{3}(M_{j}) is not surjective, then there exists Σ3,j+1\Sigma_{3,j+1}, a closed connected stable two-sided critical point of 𝒜2\mathcal{A}_{2}, so the induction proceeds. To see that this process eventually terminates, we refer to Bamler-Li-Mantoulidis [1, Lemma 2.5]. ∎

Note that, by Poincaré duality and Lemma 31, we have

H2(Σ3;)=H1(Σ3;)=Hom(H1(Σ3;);)=0.H_{2}(\Sigma_{3};\mathbb{Z})=H^{1}(\Sigma_{3};\mathbb{Z})=\text{Hom}(H_{1}(\Sigma_{3};\mathbb{Z});\mathbb{Z})=0.

Given ΩΣ^2\Omega\subset\hat{{\Sigma}}_{2}, we write Ω\partial\Omega for its topological boundary, and we assume Ω\partial\Omega consists of smooth properly embedded surfaces in Σ^2\hat{{\Sigma}}_{2}.

Lemma 33.

A connected component of Σ^2Ω\hat{{\Sigma}}_{2}\setminus\Omega contains exactly one component of Ω\partial\Omega.

Proof.

Assuming the contrary, as in Chodosh-Li [3], there exists σ\sigma an embedded S1S^{1} such that [σ][\sigma] is not torsion in H1(Σ^2)H_{1}(\hat{{\Sigma}}_{2}). The long exact sequence in homology for (Σ^2,Σ^2)(\hat{{\Sigma}}_{2},\partial\hat{{\Sigma}}_{2}) yields:

H3(Σ^2)H3(Σ^2)H3(Σ^2,Σ^2)H2(Σ^2)=0.H_{3}(\partial\hat{{\Sigma}}_{2})\to H_{3}(\hat{{\Sigma}}_{2})\to H_{3}(\hat{{\Sigma}}_{2},\partial\hat{{\Sigma}}_{2})\to H_{2}(\partial\hat{{\Sigma}}_{2})=0.

The final term vanishes since Σ^2\partial\hat{{\Sigma}}_{2} consisits of components with vanishing second homology group. Combining with Lemma 32, we conclude that H3(Σ^2,Σ^2)=0H_{3}(\hat{{\Sigma}}_{2},\partial\hat{{\Sigma}}_{2})=0. Poincaré duality implies that H1(Σ^2)=0H^{1}(\hat{{\Sigma}}_{2})=0, and so the universal coefficient theorem implies that H1(Σ^2)H_{1}(\hat{{\Sigma}}_{2}) is torsion. This is a contradiction. ∎

Next, we proceed to the dice-procedure as in Chodosh-Li [3]. We consider the following μ\mu-bubble functional:

𝒜3(Ω)=ΩuwΩ(χΩχΩ0)uwh,\mathcal{A}_{3}(\Omega)=\int_{\partial^{*}\Omega}uw-\int_{\Omega}(\chi_{\Omega}-\chi_{\Omega_{0}})uwh,

where hh satisfies

1+h2(x)2|h|0,1+h^{2}(x)-2|\nabla h|\geq 0,

and is to be specified later.

Proposition 34.

Suppose Υ\Upsilon is a component of a stable, free boundary μ\mu-bubble of 𝒜3\mathcal{A}_{3}. Then there exists a conformal factor such that the corresponding conformal Ricci curvature on Υ\Upsilon is positive and the boundary is minimal after the conformal change.

Proof.

By the first variation, the mean curvature of Υ\Upsilon with respect to the (outer) unit normal νΥ\nu_{\Upsilon} of Υ\Upsilon in Σ2{\Sigma_{2}} is given by

HΥ=Σ2log(uw),νΥ+h.H_{\Upsilon}=-\langle\nabla_{{\Sigma_{2}}}\log(uw),\nu_{\Upsilon}\rangle+h.

For any smooth function ψ\psi on Υ\Upsilon, the second variation formula gives

0\displaystyle 0\leq Υ|Υψ|2uw(AΥ2+RicΣ2(νΥ,νΥ))ψ2uwΣ2h,νΥψ2uwhΣ2(uw),νΥψ2\displaystyle\int_{\Upsilon}|\nabla_{\Upsilon}\psi|^{2}uw-\left(\|A_{\Upsilon}\|^{2}+\operatorname{Ric}_{{\Sigma_{2}}}(\nu_{\Upsilon},\nu_{\Upsilon})\right)\psi^{2}uw-\langle\nabla_{{\Sigma_{2}}}h,\nu_{\Upsilon}\rangle\psi^{2}uw-h\langle\nabla_{{\Sigma_{2}}}(uw),\nu_{\Upsilon}\rangle\psi^{2}
+Υ(ΔΣ2(uw)ΔΥ(uw))ψ2ΥAΣ^2(ω,ω)ψ2uw,\displaystyle+\int_{\Upsilon}\left(\Delta_{{\Sigma_{2}}}(uw)-\Delta_{\Upsilon}(uw)\right)\psi^{2}-\int_{\partial\Upsilon}A_{\partial\hat{\Sigma}_{2}}(\omega,\omega)\psi^{2}uw,

where ω\omega is the unit normal of Σ^2\partial\hat{\Sigma}_{2}. Since Υ\partial\Upsilon meets Σ^2\partial\hat{\Sigma}_{2} orthogonally, ω\omega is also the unit outer normal of Υ\partial\Upsilon in Υ\Upsilon.

Since the components of Σ^2\partial\hat{\Sigma}_{2} are μ\mu-bubbles of 𝒜2\mathcal{A}_{2}, we see from the first variation formula that

HΣ^2=Σ2log(uw),ω,H_{\partial\hat{\Sigma}_{2}}=-\langle\nabla_{{\Sigma_{2}}}\log(uw),\omega\rangle,

where HΣ^2H_{\partial\hat{\Sigma}_{2}} is the mean curvature of Σ^2\partial\hat{\Sigma}_{2} with respect to ω\omega. Noting Υ\partial\Upsilon meets Σ^2\partial\hat{\Sigma}_{2} orthogonally, we can denote by HΥH_{\partial\Upsilon} the mean curvature of Υ\partial\Upsilon with respect to ω\omega, and write

HΣ^2=HΥ+AΣ^2(ω,ω).H_{\partial\hat{\Sigma}_{2}}=H_{\partial\Upsilon}+A_{\partial\hat{\Sigma}_{2}}(\omega,\omega).

Since ω\omega lies in the tangent space of Υ\Upsilon, we rewrite the first variation formula of Σ^2\partial\hat{\Sigma}_{2} at the intersection points with Υ\partial\Upsilon as

HΣ^2=Υlog(uw),ω.H_{\partial\hat{\Sigma}_{2}}=-\langle\nabla_{\Upsilon}\log(uw),\omega\rangle.

Now we have that

Υ|Υψ|2uw\displaystyle\int_{\Upsilon}|\nabla_{\Upsilon}\psi|^{2}uw ΥAΣ^2(ω,ω)ψ2uw\displaystyle-\int_{\partial\Upsilon}A_{\partial\hat{\Sigma}_{2}}(\omega,\omega)\psi^{2}uw
=\displaystyle= ΥuwψΔΥψψΥψ,Υ(uw)\displaystyle\int_{\Upsilon}-uw\psi\Delta_{\Upsilon}\psi-\psi\langle\nabla_{\Upsilon}\psi,\nabla_{\Upsilon}(uw)\rangle
+Υ(Υψ,ωuw+Υ(uw),ωψΥ(uw),ωψAΣ^2(ω,ω)ψuw)ψ\displaystyle+\int_{\partial\Upsilon}\left(\langle\nabla_{\Upsilon}\psi,\omega\rangle uw+\langle\nabla_{\Upsilon}(uw),\omega\rangle\psi-\langle\nabla_{\Upsilon}(uw),\omega\rangle\psi-A_{\partial\hat{\Sigma}_{2}}(\omega,\omega)\psi uw\right)\psi
=\displaystyle= ΥuwψΔΥψψΥψ,Υ(uw)\displaystyle\int_{\Upsilon}-uw\psi\Delta_{\Upsilon}\psi-\psi\langle\nabla_{\Upsilon}\psi,\nabla_{\Upsilon}(uw)\rangle
+Υ(Υ(uwψ),ω+(HΣ^2AΣ^2(ω,ω))uwψ)ψ\displaystyle+\int_{\partial\Upsilon}\left(\langle\nabla_{\Upsilon}(uw\psi),\omega\rangle+\left(H_{\partial\hat{\Sigma}_{2}}-A_{\partial\hat{\Sigma}_{2}}(\omega,\omega)\right)uw\psi\right)\psi
=\displaystyle= ΥuwψΔΥψψΥψ,Υ(uw)+Υ(Υ(uwψ),ω+HΥuwψ)ψ.\displaystyle\int_{\Upsilon}-uw\psi\Delta_{\Upsilon}\psi-\psi\langle\nabla_{\Upsilon}\psi,\nabla_{\Upsilon}(uw)\rangle+\int_{\partial\Upsilon}\left(\langle\nabla_{\Upsilon}(uw\psi),\omega\rangle+H_{\partial\Upsilon}uw\psi\right)\psi.

Summarizing the above computations, we can rewrite the second variation formula of Υ\Upsilon as

0\displaystyle 0\leq ΥuwψΔΥψψΥψ,Υ(uw)(AΥ2+RicΣ2(νΥ,νΥ))ψ2uw\displaystyle\int_{\Upsilon}-uw\psi\Delta_{\Upsilon}\psi-\psi\langle\nabla_{\Upsilon}\psi,\nabla_{\Upsilon}(uw)\rangle-\left(\|A_{\Upsilon}\|^{2}+\operatorname{Ric}_{{\Sigma_{2}}}(\nu_{\Upsilon},\nu_{\Upsilon})\right)\psi^{2}uw
ΥΣ2h,νΥψ2uwhΣ2(uw),νΥψ2+(ΔΣ2(uw)ΔΥ(uw))ψ2\displaystyle\int_{\Upsilon}-\langle\nabla_{{\Sigma_{2}}}h,\nu_{\Upsilon}\rangle\psi^{2}uw-h\langle\nabla_{{\Sigma_{2}}}(uw),\nu_{\Upsilon}\rangle\psi^{2}+\left(\Delta_{{\Sigma_{2}}}(uw)-\Delta_{\Upsilon}(uw)\right)\psi^{2}
+Υ(Υ(uwψ),ω+HΥuwψ)ψ.\displaystyle+\int_{\partial\Upsilon}\left(\langle\nabla_{\Upsilon}(uw\psi),\omega\rangle+H_{\partial\Upsilon}uw\psi\right)\psi.

The first variation formula implies the following equality

12HΥ2ψ2uw=12(uw)1Σ2(uw),νΥ2ψ2hΣ2(uw),νΥψ2+12h2ψ2uw.\frac{1}{2}H_{\Upsilon}^{2}\psi^{2}uw=\frac{1}{2}(uw)^{-1}\langle\nabla_{{\Sigma_{2}}}(uw),\nu_{\Upsilon}\rangle^{2}\psi^{2}-h\langle\nabla_{{\Sigma_{2}}}(uw),\nu_{\Upsilon}\rangle\psi^{2}+\frac{1}{2}h^{2}\psi^{2}uw.

As in Chodosh-Li [3], we use the above equality and the fact that

12|Σ2h|+h201-2|\nabla_{{\Sigma_{2}}}h|+h^{2}\geq 0

to further simplify the second variation formula as follows:

0\displaystyle 0\leq ΥuwψΔΥψψΥψ,Υ(uw)(AΥ2+RicΣ2(νΥ,νΥ)12HΥ212)ψ2uw\displaystyle\int_{\Upsilon}-uw\psi\Delta_{\Upsilon}\psi-\psi\langle\nabla_{\Upsilon}\psi,\nabla_{\Upsilon}(uw)\rangle-\left(\|A_{\Upsilon}\|^{2}+\operatorname{Ric}_{{\Sigma_{2}}}(\nu_{\Upsilon},\nu_{\Upsilon})-\frac{1}{2}H_{\Upsilon}^{2}-\frac{1}{2}\right)\psi^{2}uw
+Υ(ΔΣ2(uw)ΔΥ(uw))ψ212Σ2log(uw),νΥ2ψ2uw+Υ(Υ(uwψ),ω+HΥuwψ)ψ.\displaystyle+\int_{\Upsilon}\left(\Delta_{{\Sigma_{2}}}(uw)-\Delta_{\Upsilon}(uw)\right)\psi^{2}-\frac{1}{2}\langle\nabla_{{\Sigma_{2}}}\log(uw),\nu_{\Upsilon}\rangle^{2}\psi^{2}uw+\int_{\partial\Upsilon}\left(\langle\nabla_{\Upsilon}(uw\psi),\omega\rangle+H_{\partial\Upsilon}uw\psi\right)\psi.

Thus there exists a smooth positive function v2v_{2} defined on Υ\Upsilon satisfying

(8) v21ΔΥv2\displaystyle-v_{2}^{-1}\Delta_{\Upsilon}v_{2}\geq Υlogv2,Υlog(uw)(uw)1(ΔΣ2(uw)ΔΥ(uw))\displaystyle\langle\nabla_{\Upsilon}\log v_{2},\nabla_{\Upsilon}\log(uw)\rangle-(uw)^{-1}\left(\Delta_{{\Sigma_{2}}}(uw)-\Delta_{\Upsilon}(uw)\right)
+12Σ2log(uw),νΥ2+AΥ2+RicΣ2(νΥ,νΥ)12HΥ212.\displaystyle+\frac{1}{2}\langle\nabla_{{\Sigma_{2}}}\log(uw),\nu_{\Upsilon}\rangle^{2}+\|A_{\Upsilon}\|^{2}+\operatorname{Ric}_{{\Sigma_{2}}}(\nu_{\Upsilon},\nu_{\Upsilon})-\frac{1}{2}H_{\Upsilon}^{2}-\frac{1}{2}.

and it satisfies the following boundary condition on Υ\partial\Upsilon

(9) Υ(uwv2),ω+HΥuwv2=0.\displaystyle\langle\nabla_{\Upsilon}(uwv_{2}),\omega\rangle+H_{\partial\Upsilon}uwv_{2}=0.

The boundary condition implies that Υ\partial\Upsilon is minimal under the conformal change (uwv2)2g(uwv_{2})^{2}g.

By Proposition 28, we have

(10) (uw)1\displaystyle(uw)^{-1} ΔΣ2(uw)\displaystyle\Delta_{{\Sigma_{2}}}(uw)
\displaystyle\leq Σ2logu,Σ2logwAΣ12RicM¯(η,η)AΣ22RicΣ1(ν,ν)+12HΣ22+12.\displaystyle\langle\nabla_{{\Sigma_{2}}}\log u,\nabla_{{\Sigma_{2}}}\log w\rangle-\|A_{\Sigma_{1}}\|^{2}-\operatorname{Ric}_{\overline{M}}(\eta,\eta)-\|A_{{\Sigma_{2}}}\|^{2}-\operatorname{Ric}_{\Sigma_{1}}(\nu,\nu)+\frac{1}{2}H_{{\Sigma_{2}}}^{2}+\frac{1}{2}.

Combining (8) and (10), we obtain

(uwv2)1\displaystyle(uwv_{2})^{-1} ΔΥ(uwv2)\displaystyle\Delta_{\Upsilon}(uwv_{2})
=\displaystyle= (uw)1ΔΥ(uw)+v21ΔΥ(v2)+2Υlog(uw),Υlogv2\displaystyle(uw)^{-1}\Delta_{\Upsilon}(uw)+v_{2}^{-1}\Delta_{\Upsilon}(v_{2})+2\langle\nabla_{\Upsilon}\log(uw),\nabla_{\Upsilon}\log v_{2}\rangle
\displaystyle\leq (uw)1ΔΣ2(uw)+Υlog(uw),Υlogv2AΥ2RicΣ2(ν3,ν3)+12HΥ2+1212Σ2log(uw),νΥ2\displaystyle(uw)^{-1}\Delta_{{\Sigma_{2}}}(uw)+\langle\nabla_{\Upsilon}\log(uw),\nabla_{\Upsilon}\log v_{2}\rangle-\|A_{\Upsilon}\|^{2}-\operatorname{Ric}_{{\Sigma_{2}}}(\nu_{3},\nu_{3})+\frac{1}{2}H_{\Upsilon}^{2}+\frac{1}{2}-\frac{1}{2}\langle\nabla_{{\Sigma_{2}}}\log(uw),\nu_{\Upsilon}\rangle^{2}
\displaystyle\leq Σ2logu,νΥΣ2logw,νΥ+Υlogu,Υlogw12Σ2log(uw),νΥ2+Υlog(uw),Υlogv2\displaystyle\langle\nabla_{{\Sigma_{2}}}\log u,\nu_{\Upsilon}\rangle\langle\nabla_{{\Sigma_{2}}}\log w,\nu_{\Upsilon}\rangle+\langle\nabla_{\Upsilon}\log u,\nabla_{\Upsilon}\log w\rangle-\frac{1}{2}\langle\nabla_{{\Sigma_{2}}}\log(uw),\nu_{\Upsilon}\rangle^{2}+\langle\nabla_{\Upsilon}\log(uw),\nabla_{\Upsilon}\log v_{2}\rangle
AΣ12AΣ22AΥ2RicM¯(η,η)RicΣ1(ν,ν)RicΣ2(νΥ,νΥ)\displaystyle-\|A_{\Sigma_{1}}\|^{2}-\|A_{{\Sigma_{2}}}\|^{2}-\|A_{\Upsilon}\|^{2}-\operatorname{Ric}_{\overline{M}}(\eta,\eta)-\operatorname{Ric}_{\Sigma_{1}}(\nu,\nu)-\operatorname{Ric}_{{\Sigma_{2}}}(\nu_{\Upsilon},\nu_{\Upsilon})
+12HΣ22+12HΥ2+1,\displaystyle+\frac{1}{2}H_{{\Sigma_{2}}}^{2}+\frac{1}{2}H_{\Upsilon}^{2}+1,
\displaystyle\leq Υlog(uw),Υlogv2+Υlogu,ΥlogwAΣ12AΣ22AΥ2\displaystyle\langle\nabla_{\Upsilon}\log(uw),\nabla_{\Upsilon}\log v_{2}\rangle+\langle\nabla_{\Upsilon}\log u,\nabla_{\Upsilon}\log w\rangle-\|A_{\Sigma_{1}}\|^{2}-\|A_{{\Sigma_{2}}}\|^{2}-\|A_{\Upsilon}\|^{2}
RicM¯(η,η)RicΣ(ν,ν)RicΣ2(νΥ,νΥ)+12HΣ22+12HΥ2+1.\displaystyle-\operatorname{Ric}_{\overline{M}}(\eta,\eta)-\operatorname{Ric}_{\Sigma}(\nu,\nu)-\operatorname{Ric}_{{\Sigma_{2}}}(\nu_{\Upsilon},\nu_{\Upsilon})+\frac{1}{2}H_{{\Sigma_{2}}}^{2}+\frac{1}{2}H_{\Upsilon}^{2}+1.

Supposing e4e_{4} is a unit vector in TΥT\Upsilon, we have

RicΥ(e4,e4)\displaystyle\operatorname{Ric}_{\Upsilon}(e_{4},e_{4}) (uwv2)1ΔΥ(uwv2)+12|Υlog(uwv2)|2\displaystyle-(uwv_{2})^{-1}\Delta_{\Upsilon}(uwv_{2})+\frac{1}{2}|\nabla_{\Upsilon}\log(uwv_{2})|^{2}
\displaystyle\geq 12|Υlog(uwv2)|2Υlog(uw),Υlogv2Υlogu,Υlogw\displaystyle\frac{1}{2}|\nabla_{\Upsilon}\log(uwv_{2})|^{2}-\langle\nabla_{\Upsilon}\log(uw),\nabla_{\Upsilon}\log v_{2}\rangle-\langle\nabla_{\Upsilon}\log u,\nabla_{\Upsilon}\log w\rangle
+AΣ12+AΣ22+AΥ212HΣ2212HΥ21\displaystyle+\|A_{\Sigma_{1}}\|^{2}+\|A_{{\Sigma_{2}}}\|^{2}+\|A_{\Upsilon}\|^{2}-\frac{1}{2}H_{{\Sigma_{2}}}^{2}-\frac{1}{2}H_{\Upsilon}^{2}-1
+RicM¯(η,η)+RicΣ1(ν,ν)+RicΣ2(νΥ,νΥ)+RicΥ(e4,e4)\displaystyle+\operatorname{Ric}_{\overline{M}}(\eta,\eta)+\operatorname{Ric}_{\Sigma_{1}}(\nu,\nu)+\operatorname{Ric}_{{\Sigma_{2}}}(\nu_{\Upsilon},\nu_{\Upsilon})+\operatorname{Ric}_{\Upsilon}(e_{4},e_{4})
\displaystyle\geq AΣ12+AΣ22+AΥ212HΣ2212HΥ21\displaystyle\|A_{{\Sigma_{1}}}\|^{2}+\|A_{\Sigma_{2}}\|^{2}+\|A_{\Upsilon}\|^{2}-\frac{1}{2}H_{{\Sigma_{2}}}^{2}-\frac{1}{2}H_{\Upsilon}^{2}-1
+RicΥ(e4,e4)+RicΣ2(ω,ω)+RicΣ1(ν,ν)+RicM¯(η,η)\displaystyle+\operatorname{Ric}_{\Upsilon}(e_{4},e_{4})+\operatorname{Ric}_{{\Sigma_{2}}}(\omega,\omega)+\operatorname{Ric}_{\Sigma_{1}}(\nu,\nu)+\operatorname{Ric}_{\overline{M}}(\eta,\eta)
=\displaystyle= AΣ12+AΣ22+AΥ212HΣ2212HΥ21\displaystyle\|A_{{\Sigma_{1}}}\|^{2}+\|A_{\Sigma_{2}}\|^{2}+\|A_{\Upsilon}\|^{2}-\frac{1}{2}H_{{\Sigma_{2}}}^{2}-\frac{1}{2}H_{\Upsilon}^{2}-1
+C4(e1,e2,e3,e4)+p=24q=p+16(AΣ1(ep,ep)AΣ1(eq,eq)AΣ1(ep,eq)2)\displaystyle+C_{4}(e_{1},e_{2},e_{3},e_{4})+\sum_{p=2}^{4}\sum_{q=p+1}^{6}\left(A_{\Sigma_{1}}(e_{p},e_{p})A_{\Sigma_{1}}(e_{q},e_{q})-A_{\Sigma_{1}}(e_{p},e_{q})^{2}\right)
+p=34q=p+16(AΣ2(ep,ep)AΣ2(eq,eq)AΣ2(ep,eq)2)\displaystyle+\sum_{p=3}^{4}\sum_{q=p+1}^{6}\left(A_{{\Sigma_{2}}}(e_{p},e_{p})A_{{\Sigma_{2}}}(e_{q},e_{q})-A_{{\Sigma_{2}}}(e_{p},e_{q})^{2}\right)
+q=56(AΥ(e4,e4)AΥ(eq,eq)AΥ(e4,eq)2)\displaystyle+\sum_{q=5}^{6}\left(A_{\Upsilon}(e_{4},e_{4})A_{\Upsilon}(e_{q},e_{q})-A_{\Upsilon}(e_{4},e_{q})^{2}\right)
\displaystyle\geq C4(e1,e2,e3,e4)112.\displaystyle C_{4}(e_{1},e_{2},e_{3},e_{4})-1\geq\frac{1}{2}.

We have completed the proof. ∎

Corollary 35.

Suppose Υ\Upsilon is a component of a free boundary μ\mu-bubble of 𝒜3\mathcal{A}_{3}. Then it has connected boundary and bounded diameter.

Proof.

Given the previous proposition, the diameter estimate follows from Theorem 18, and the fact that the boundary is connected follows from Theorem 22 and (9). ∎

We can summarize the dice procedure as follows.

Proposition 36.

Suppose pΣ^2p\in\hat{\Sigma}_{2} is a fixed interior point, and assume further that Bε(p)Σ^2B_{\varepsilon}(p)\subset\subset\hat{\Sigma}_{2}. There exists a finite number kk and open connected domains {Ωi}i=1k\{\Omega_{i}\}_{i=1}^{k},

Bε(p)=Ω1Ω2,Ωk=Σ^2,B_{\varepsilon}(p)=\Omega_{1}\subset\Omega_{2},\subset\cdots\subset\Omega_{k}=\hat{\Sigma}_{2},

with the following properties:

  1. (1)

    dΣ^2(Ωi+1,Ωi)23πd_{\hat{\Sigma}_{2}}(\partial\Omega_{i+1},\partial\Omega_{i})\geq\frac{2}{3}\pi;

  2. (2)

    Each component of Ωi+1Ωi\Omega_{i+1}\setminus\Omega_{i} has diameter at most 10π10\pi;

  3. (3)

    Any component ΥΩj\Upsilon\subset\partial\Omega_{j} has diameter diamΥπ\operatorname{diam}\Upsilon\leq\pi;

  4. (4)

    Each component of Ωj\partial\Omega_{j} is either a closed manifold with finite fundamental group or a compact manifold with connected boundary in Σ^2\partial\hat{\Sigma}_{2}.

Proof.

The proof follows from induction. Suppose we have constructed Ω1,Ω2,,Ωj\Omega_{1},\Omega_{2},\cdots,\Omega_{j} satisfying the above assumptions and ΩjΣ^2\Omega_{j}\neq\hat{\Sigma}_{2}. Smooth d(,Ωj)d(\cdot,\Omega_{j}) to a function ρ\rho such that d(,Ωj)ρ32d(,Ωj)d(\cdot,\Omega_{j})\leq\rho\leq\frac{3}{2}d(\cdot,\Omega_{j}) and ρ|Ωj=0\rho|_{\Omega_{j}}=0. Take hh to be

h(x)=tan(13(ρπ)π2),h(x)=-\tan\left(\frac{1}{3}(\rho-\pi)-\frac{\pi}{2}\right),

and note that

1+h22|h|0.1+h^{2}-2|\nabla h|\geq 0.

Minimize 𝒜^3\hat{\mathcal{A}}_{3} to get a μ\mu-bubble Ωj+1\Omega_{j+1}.

By definition of hh, Ωj+1\Omega_{j+1} satisfies condition (1)(1). By Lemma 33 and Lemma 31, it satisfies condition (2)(2). By Corollary 35 and Lemma 31, it also satisfies condition (3)(3) and (4)(4). Since condition (1)(1) is satisfied by the sets we constructed, the induction process must terminate after finitely many steps. ∎

We now have all the ingredients we need to prove the last case of the Main Theorem 4.

Proof of Theorem 30.

Let MM be as in the statement of the theorem. Assume for the sake of contradiction that MM admits a metric with positive 44-intermediate curvature. We can assume the 44-intermediate curvature of MM is at least 32\frac{3}{2}.

Arguing as in the proof of Theorem 27, we obtain 55-dimensional Σ1\Sigma_{1} and 44-dimensional Σ2\Sigma_{2}. By Lemma 31, Lemma 32 and Proposition 36, there exists a set of disjoint embedded closed 33 manifolds Σ3,1,Σ3,2,Σ3,kΣ2\Sigma_{3,1},\Sigma_{3,2}\cdots,\Sigma_{3,k}\subset\Sigma_{2} with diamΣ3,iπ\operatorname{diam}\Sigma_{3,i}\leq\pi. Moreover, there exists a set of embedded compact 33-manifolds Υ1,Υ2,,ΥlΣ2\Upsilon_{1},\Upsilon_{2},\cdots,\Upsilon_{l}\subset\Sigma_{2} such that, diamΥjπ\operatorname{diam}\Upsilon_{j}\leq\pi, Υj\partial\Upsilon_{j} is connected and contained in i=1kΣ3,i\cup_{i=1}^{k}\Sigma_{3,i}, and the interiors of Υj\Upsilon_{j} are pairwise disjoint with each other and disjoint with each Σ3,i\Sigma_{3,i}. Also, each component K1,K2,,KsK_{1},K_{2},\cdots,K_{s} of

Σ2((i=1kΣ3,i)(j=1sΥj))\Sigma_{2}\setminus\left(\left(\cup_{i=1}^{k}\Sigma_{3,i}\right)\cup\left(\cup_{j=1}^{s}\Upsilon_{j}\right)\right)

has diameter bounded by 10π10\pi.

We write the boundary components of KjK_{j} as Υ^j1,Υ^j2,,Υ^jn(j)\hat{\Upsilon}_{j}^{1},\hat{\Upsilon}^{2}_{j},\cdots,\hat{\Upsilon}_{j}^{n(j)}. By Proposition 14, there exists a positive R>0R>0, independent of LL, such that we can fill-in Υ^ji\hat{\Upsilon}_{j}^{i} by K^ji\hat{K}_{j}^{i} with extrinsic diameter at most RR. Then

KjK^j1K^jn(j)K_{j}-\hat{K}_{j}^{1}-\cdots-\hat{K}_{j}^{n(j)}

is a cycle with extrinsic diameter at most 2R+10π2R+10\pi. By Proposition 14, there exists a 55-chain Γ~j\tilde{\Gamma}_{j} with extrinsic diameter bounded by R~\tilde{R}, which is independent of LL, such that

Γ~j=KjK^j1K^jn(j).\partial\tilde{\Gamma}_{j}=K_{j}-\hat{K}_{j}^{1}-\cdots-\hat{K}_{j}^{n(j)}.

Note that

Σ2j=1sΓ~j=j=1si=1n(j)K^ji.\Sigma_{2}-\sum_{j=1}^{s}\partial\tilde{\Gamma}_{j}=\sum_{j=1}^{s}\sum_{i=1}^{n(j)}\hat{K}_{j}^{i}.

Since Υi\Upsilon_{i} has connected boundary, there exists an index u(i,j){1,2,,k}u(i,j)\in\{1,2,\cdots,k\} so that Υ^ji\hat{\Upsilon}_{j}^{i} only intersects with Σ3,u(i,j)\Sigma_{3,u(i,j)} but not any of the other components of Σ^2\partial\hat{\Sigma}_{2}. We group the K^ji\hat{K}_{j}^{i} by u(i,j)=au(i,j)=a for a{1,2,,k}a\in\{1,2,\cdots,k\}. Then

{i,j:u(i,j)=a}K^ji\sum_{\{i,j:u(i,j)=a\}}\hat{K}_{j}^{i}

is a cycle of diameter at most 2R+π2R+\pi. Furthermore,

[a=1k{i,j:u(i,j)=a}K^ji]=0.\partial\left[\sum_{a=1}^{k}\sum_{\{i,j:u(i,j)=a\}}\hat{K}_{j}^{i}\right]=0.

Therefore, by Proposition 14, there exists R^>0\hat{R}>0, independent of LL, such that there exists a 55-chain Θa\Theta_{a} with extrinsic diameter at most R^\hat{R} satisfying

Θa={i,j:u(i,j)=a}K^ji.\partial\Theta_{a}=\sum_{\{i,j:u(i,j)=a\}}\hat{K}_{j}^{i}.

In conclusion, we have

Σ2=[j=1sΓ~j+a=1kΘa],\Sigma_{2}=\partial\left[\sum_{j=1}^{s}\tilde{\Gamma}_{j}+\sum_{a=1}^{k}\Theta_{a}\right],

where each term in the sum has uniform bounded diameter as LL\to\infty. A contradiction is achieved. ∎

4.6. The Case n=6n=6 and m=3m=3

Finally, we prove the case where n=6n=6 and m=3m=3. This will complete the proof of Theorem 4. This case is somewhat subtle and requires a delicate analysis.

Theorem 37.

Let M6M^{6} be a closed manifold. Assume that the universal cover M¯\overline{M} of MM satisfies H6(M¯,)=H5(M¯,)=H4(M¯,)=0H_{6}(\overline{M},\mathbb{Z})=H_{5}(\overline{M},\mathbb{Z})=H_{4}(\overline{M},\mathbb{Z})=0. Then MM does not admit a metric with positive 3-intermediate curvature.

Let MM be as in the statement of the theorem. Assume for the sake of contradiction that MM admits a metric with positive 33-intermediate curvature. By scaling, we can suppose the 3-intermediate curvature of MM is at least 33. For a fixed large L>0L>0, we apply Proposition 13 to find a geodesic line σ\sigma in M¯\overline{M} and a closed manifold Λn2\Lambda^{n-2} embedded in M¯\overline{M} such that Λ\Lambda is linked with σ\sigma and d(σ,Λ)2Ld(\sigma,\Lambda)\geq 2L. Let Σ1\Sigma_{1} be the area minimizing minimal hypersurface in M¯\overline{M} with Σ1=Λ\partial\Sigma_{1}=\Lambda. Let η\eta be the unit normal to Σ1\Sigma_{1}. Since Σ1\Sigma_{1} is area minimizing, there exists a function u>0u>0 on Σ1\Sigma_{1} which satisfies ΔΣ1u+(|AΣ1|2+RicM¯(η,η))u0\Delta_{\Sigma_{1}}u+(|A_{\Sigma_{1}}|^{2}+\operatorname{Ric}_{\overline{M}}(\eta,\eta))u\leq 0.

We are now going to construct a μ\mu-bubble in Σ1\Sigma_{1}. The construction depends on a choice of several parameters a,τ,ε,αa,\tau,\varepsilon,\alpha which depend only on MM and not on LL. First, we select

0.9748390101<a<10.9748\approx\frac{\sqrt{390}}{10}-1<a<1

close enough to 1 so that the following two conditions hold:

  • (i)

    (1a2)|RM(X,Y,X,Y)|1100(1-a^{2})|R_{M}(X,Y,X,Y)|\leq\frac{1}{100} for all unit tangent vectors X,YX,Y to MM;

  • (ii)

    The matrix

    (1a916121a38121a1a038014)\begin{pmatrix}\frac{1}{a}-\frac{9}{16}&\frac{1}{2}-\frac{1}{a}&\frac{\sqrt{3}}{8}\\ \\ \frac{1}{2}-\frac{1}{a}&\frac{1}{a}&0\\ \\ \frac{\sqrt{3}}{8}&0&\frac{1}{4}\end{pmatrix}

    is positive definite.

To see that (ii) is possible, note that the determinant of this matrix is 116+116a-\frac{1}{16}+\frac{1}{16a} which is positive for a<1a<1. Also when a=1a=1, by direct computation, the matrix

(7161238121038014)\begin{pmatrix}\frac{7}{16}&-\frac{1}{2}&\frac{\sqrt{3}}{8}\\ \\ -\frac{1}{2}&1&0\\ \\ \frac{\sqrt{3}}{8}&0&\frac{1}{4}\end{pmatrix}

has eigenvalues

λ1=132(27+217)>0,λ2=132(27217)>0,λ3=0.\lambda_{1}=\frac{1}{32}(27+\sqrt{217})>0,\quad\lambda_{2}=\frac{1}{32}(27-\sqrt{217})>0,\quad\lambda_{3}=0.

Therefore, by continuity, a suitable choice of aa is possible. Next, we select a<τ<1a<\tau<1 and α>0\alpha>0 so that the following condition holds:

  • (iii)

    The matrix

    (31634τ+τaτ2τa38τ2τaτaα0380τ34)\begin{pmatrix}\frac{3}{16}-\frac{3}{4}\tau+\frac{\tau}{a}&\frac{\tau}{2}-\frac{\tau}{a}&\frac{\sqrt{3}}{8}\\ \\ \frac{\tau}{2}-\frac{\tau}{a}&\frac{\tau}{a}-\alpha&0\\ \\ \frac{\sqrt{3}}{8}&0&\tau-\frac{3}{4}\end{pmatrix}

    is positive definite.

Since this matrix reduces to the one in condition (ii) when τ=1\tau=1 and α=0\alpha=0, such a choice is again possible by continuity. Finally, since τ<1\tau<1, we can select ε>0\varepsilon>0 so that

  • (iv)

    τ(34+ε)τ2τ4.\displaystyle\tau-\left(\frac{3}{4}+\varepsilon\right)\tau^{2}\geq\frac{\tau}{4}.

The reason for selecting this choice of parameters will become apparent in the proof.

Next, consider the differential equation

{k(t)=1α2k(t)2,k(0)=0.\begin{cases}k^{\prime}(t)=-1-\frac{\alpha}{2}k(t)^{2},\\ k(0)=0.\end{cases}

The solution to this ODE is

k(t)=2αarctan(α2t)k(t)=-\frac{2}{\sqrt{\alpha}}\arctan\left(\sqrt{\frac{\alpha}{2}}t\right)

for t(β,β)t\in(-\beta,\beta) where β=π22α\beta=\frac{\pi}{2}\sqrt{\frac{2}{\alpha}}. Note that k(t)k(t)\to\infty as tβt\to-\beta and that k(t)k(t)\to-\infty as tβt\to\beta.

Define

ρ0(x)=dM¯(x,σ),xM¯.\rho_{0}(x)=d_{\overline{M}}(x,\sigma),\quad x\in\overline{M}.

Then let ρ1\rho_{1} be a smooth approximation to ρ0\rho_{0} which satisfies ρ1(x)2\|\nabla\rho_{1}(x)\|\leq 2 for all xM¯x\in\overline{M}, ρ1<1\rho_{1}<1 on σ\sigma, and ρ1>L+β+1\rho_{1}>L+\beta+1 on Σ1=Λ\partial\Sigma_{1}=\Lambda. Let ρ2\rho_{2} be the restriction of ρ1\rho_{1} to Σ1\Sigma_{1}. Now choose δ>0\delta>0 sufficiently small so that Lβ+δL-\beta+\delta, L+2δL+2\delta, and L+β+δL+\beta+\delta are all regular values of ρ2\rho_{2}. We can further ensure that Σ1BM¯(σ,L/4){ρ2Lβ1}\Sigma_{1}\cap B_{\overline{M}}(\sigma,L/4)\subset\{\rho_{2}\leq L-\beta-1\}. Now define

ρ(x)=ρ2(x)Lδ,xΣ1.\rho(x)=\rho_{2}(x)-L-\delta,\quad x\in\Sigma_{1}.

Then we have Σ1ρ(x)2\|\nabla_{\Sigma_{1}}\rho(x)\|\leq 2 for all xΣ1x\in\Sigma_{1}. We define

Σ~1={βρβ},\displaystyle\tilde{\Sigma}_{1}=\{-\beta\leq\rho\leq\beta\},
+Σ~1={ρ=β},\displaystyle\partial_{+}\tilde{\Sigma}_{1}=\{\rho=-\beta\},
Σ~1={ρ=β},\displaystyle\partial_{-}\tilde{\Sigma}_{1}=\{\rho=\beta\},
Ω0={βρ<δ}.\displaystyle\Omega_{0}=\{-\beta\leq\rho<\delta\}.

By construction, ±Σ~1\partial_{\pm}\tilde{\Sigma}_{1} and Ω0\partial\Omega_{0} are all smooth hypersurfaces in Σ~1\tilde{\Sigma}_{1}, and Ω0\Omega_{0} contains +Σ~1\partial_{+}\tilde{\Sigma}_{1}, and Σ1Σ~1=\partial\Sigma_{1}\cap\tilde{\Sigma}_{1}=\emptyset. Finally, we define

h(x)=k(ρ(x)).h(x)=k(\rho(x)).

Then hh is a smooth function on the interior of Σ~1\tilde{\Sigma}_{1} which satisfies h(x)±h(x)\to\pm\infty as x±Σ~1x\to\partial_{\pm}\tilde{\Sigma}_{1}. Moreover, we have

Σ1h(x)=|k(ρ(x))|Σ1ρ(x)2(1+α2k(ρ(x))2)=2+αh2,\|\nabla_{\Sigma_{1}}h(x)\|=|k^{\prime}(\rho(x))|\cdot\|\nabla_{\Sigma_{1}}\rho(x)\|\leq 2\left(1+\frac{\alpha}{2}k(\rho(x))^{2}\right)=2+\alpha h^{2},

for all xΣ~1x\in\tilde{\Sigma}_{1}. It follows that

(11) 2+αh2Σ1h02+\alpha h^{2}-\|\nabla_{\Sigma_{1}}h\|\geq 0

on Σ~1\tilde{\Sigma}_{1}. Proposition 15 (with uu replaced by uau^{a}) implies that there exists a minimizer ΩΣ~1\Omega\subset\tilde{\Sigma}_{1} of the μ\mu-bubble functional

𝒜(Ω)=Ωua𝑑4Σ~1(χΩχΩ0)uah𝑑5\mathcal{A}(\Omega)=\int_{\partial\Omega}u^{a}\,d\mathcal{H}^{4}-\int_{\tilde{\Sigma}_{1}}(\chi_{\Omega}-\chi_{\Omega_{0}})u^{a}h\,d\mathcal{H}^{5}

such that ΩΔΩ0\Omega\operatorname{\Delta}\Omega_{0} is compactly contained in the interior of Σ~1\tilde{\Sigma}_{1}. Let Σ2\Sigma_{2} be a component of Ω\partial\Omega.

According to Proposition 16 with uu replaced by uau^{a} (see also Mazet [14] Section 4.2), the first variation satisfies

HΣ2=hau1Σ1u,ν,H_{\Sigma_{2}}=h-au^{-1}\langle\nabla_{\Sigma_{1}}u,\nu\rangle,

and the second variation formula implies

0\displaystyle 0 Σ2ua[|Σ2ψ|2(|AΣ2|2+RicΣ1(ν,ν))ψ2au2Σ1u,ν2ψ2\displaystyle\leq\int_{\Sigma_{2}}u^{a}\bigg{[}|\nabla_{\Sigma_{2}}\psi|^{2}-(|A_{\Sigma_{2}}|^{2}+\operatorname{Ric}_{\Sigma_{1}}(\nu,\nu))\psi^{2}-au^{-2}\langle\nabla_{\Sigma_{1}}u,\nu\rangle^{2}\psi^{2}
+au1(ΔΣ1uΔΣ2uHΣ2Σ1u,ν)ψ2Σ1h,νψ2]\displaystyle\qquad\qquad+au^{-1}(\Delta_{\Sigma_{1}}u-\Delta_{\Sigma_{2}}u-H_{\Sigma_{2}}\langle\nabla_{\Sigma_{1}}u,\nu\rangle)\psi^{2}-\langle\nabla_{\Sigma_{1}}h,\nu\rangle\psi^{2}\bigg{]}

for all ψC1(Σ2)\psi\in C^{1}(\Sigma_{2}). It follows that there exists a function w>0w>0 on Σ2\Sigma_{2} which satisfies

divΣ2(uaΣ2w)\displaystyle\operatorname{div}_{\Sigma_{2}}(u^{a}\nabla_{\Sigma_{2}}w) [(|AΣ2|2+RicΣ1(ν,ν))a(|AΣ1|2+RicM¯(η,η))]uaw\displaystyle\leq\bigg{[}-(|A_{\Sigma_{2}}|^{2}+\operatorname{Ric}_{\Sigma_{1}}(\nu,\nu))-a(|A_{\Sigma_{1}}|^{2}+\operatorname{Ric}_{\overline{M}}(\eta,\eta))\bigg{]}u^{a}w
+[Σ1h,νau2Σ1u,ν2aHΣ2u1Σ1u,νaΔΣ2uu]uaw.\displaystyle\qquad+\bigg{[}-\langle\nabla_{\Sigma_{1}}h,\nu\rangle-au^{-2}\langle\nabla_{\Sigma_{1}}u,\nu\rangle^{2}-aH_{\Sigma_{2}}u^{-1}\langle\nabla_{\Sigma_{1}}u,\nu\rangle-a\frac{\Delta_{\Sigma_{2}}u}{u}\bigg{]}u^{a}w.

Now we turn our attention to the diameter bound.

Proposition 38.

The diameter of Σ2\Sigma_{2} is uniformly bounded independently of LL.

Proof.

We are going to apply the Shen-Ye generalized Bonnet-Myers theorem with test function uawu^{a}w. Fix an orthonormal basis {e1,e2,e3,e4,e5,e6}\{e_{1},e_{2},e_{3},e_{4},e_{5},e_{6}\} for M¯\overline{M} so that e1=ηe_{1}=\eta and e2=νe_{2}=\nu. By condition (iv), it suffices to show that

RicΣ2(e3,e3)τΔΣ2(uaw)uaw+τ4|Σ2ln(uaw)|2κ>0\operatorname{Ric}_{\Sigma_{2}}(e_{3},e_{3})-\tau\frac{\Delta_{\Sigma_{2}}(u^{a}w)}{u^{a}w}+\frac{\tau}{4}|\nabla_{\Sigma_{2}}\ln(u^{a}w)|^{2}\geq\kappa>0

for some κ\kappa which does not depend on LL. We have

RicΣ2(e3,e3)τΔΣ2(uaw)uaw+τ4|Σ2ln(uaw)|2\displaystyle\operatorname{Ric}_{\Sigma_{2}}(e_{3},e_{3})-\tau\frac{\Delta_{\Sigma_{2}}(u^{a}w)}{u^{a}w}+\frac{\tau}{4}|\nabla_{\Sigma_{2}}\ln(u^{a}w)|^{2}
=RicΣ2(e3,e3)τdivΣ2(uaΣ2w)uawτdivΣ2(wΣ2ua)uaw+τ4|Σ2uaua+Σ2ww|2.\displaystyle\qquad=\operatorname{Ric}_{\Sigma_{2}}(e_{3},e_{3})-\tau\frac{\operatorname{div}_{\Sigma_{2}}(u^{a}\nabla_{\Sigma_{2}}w)}{u^{a}w}-\tau\frac{\operatorname{div}_{\Sigma_{2}}(w\nabla_{\Sigma_{2}}u^{a})}{u^{a}w}+\frac{\tau}{4}\left|\frac{\nabla_{\Sigma_{2}}u^{a}}{u^{a}}+\frac{\nabla_{\Sigma_{2}}w}{w}\right|^{2}.

Now observe that

τdivΣ2(wΣ2ua)uaw+τ4|Σ2uaua+Σ2ww|2\displaystyle-\tau\frac{\operatorname{div}_{\Sigma_{2}}(w\nabla_{\Sigma_{2}}u^{a})}{u^{a}w}+\frac{\tau}{4}\left|\frac{\nabla_{\Sigma_{2}}u^{a}}{u^{a}}+\frac{\nabla_{\Sigma_{2}}w}{w}\right|^{2}
=τ[Σ2uaua,Σ2ww+a(a1)|Σ2u|2u2+aΔΣ2uu]+τ4|Σ2uaua+Σ2ww|2\displaystyle\qquad\quad=-\tau\left[\left\langle\frac{\nabla_{\Sigma_{2}}u^{a}}{u^{a}},\frac{\nabla_{\Sigma_{2}}w}{w}\right\rangle+a(a-1)\frac{|\nabla_{\Sigma_{2}}u|^{2}}{u^{2}}+a\frac{\Delta_{\Sigma_{2}}u}{u}\right]+\frac{\tau}{4}\left|\frac{\nabla_{\Sigma_{2}}u^{a}}{u^{a}}+\frac{\nabla_{\Sigma_{2}}w}{w}\right|^{2}
aτΔΣ2uuτΣ2uaua,Σ2ww+τ4|Σ2uaua+Σ2ww|2\displaystyle\qquad\quad\geq-a\tau\frac{\Delta_{\Sigma_{2}}u}{u}-\tau\left\langle\frac{\nabla_{\Sigma_{2}}u^{a}}{u^{a}},\frac{\nabla_{\Sigma_{2}}w}{w}\right\rangle+\frac{\tau}{4}\left|\frac{\nabla_{\Sigma_{2}}u^{a}}{u^{a}}+\frac{\nabla_{\Sigma_{2}}w}{w}\right|^{2}
=aτΔΣ2uu+τ4|Σ2uauaΣ2ww|2aτΔΣ2uu,\displaystyle\qquad\quad=-a\tau\frac{\Delta_{\Sigma_{2}}u}{u}+\frac{\tau}{4}\left|\frac{\nabla_{\Sigma_{2}}u^{a}}{u^{a}}-\frac{\nabla_{\Sigma_{2}}w}{w}\right|^{2}\geq-a\tau\frac{\Delta_{\Sigma_{2}}u}{u},

where we used the fact that a<1a<1. Thus we have

RicΣ2(e3,e3)τΔΣ2(uaw)uaw+τ4|Σ2ln(uaw)|2\displaystyle\operatorname{Ric}_{\Sigma_{2}}(e_{3},e_{3})-\tau\frac{\Delta_{\Sigma_{2}}(u^{a}w)}{u^{a}w}+\frac{\tau}{4}|\nabla_{\Sigma_{2}}\ln(u^{a}w)|^{2}
RicΣ2(e3,e3)+τ|AΣ2|2+τRicΣ1(e2,e2)+aτ|AΣ1|2+aτRicM¯(e1,e1)\displaystyle\qquad\quad\geq\operatorname{Ric}_{\Sigma_{2}}(e_{3},e_{3})+\tau|A_{\Sigma_{2}}|^{2}+\tau\operatorname{Ric}_{\Sigma_{1}}(e_{2},e_{2})+a\tau|A_{\Sigma_{1}}|^{2}+a\tau\operatorname{Ric}_{\overline{M}}(e_{1},e_{1})
+τΣ1h,e2+aτHΣ2u1Σ1u,e2+aτu2Σ1u,e22\displaystyle\qquad\qquad\qquad+\tau\langle\nabla_{\Sigma_{1}}h,e_{2}\rangle+a\tau H_{\Sigma_{2}}u^{-1}\langle\nabla_{\Sigma_{1}}u,e_{2}\rangle+a\tau u^{-2}\langle\nabla_{\Sigma_{1}}u,e_{2}\rangle^{2}
=RicΣ2(e3,e3)+RicΣ1(e2,e2)+RicM¯(e1,e1)+\displaystyle\qquad\quad=\operatorname{Ric}_{\Sigma_{2}}(e_{3},e_{3})+\operatorname{Ric}_{\Sigma_{1}}(e_{2},e_{2})+\operatorname{Ric}_{\overline{M}}(e_{1},e_{1})+\mathcal{R}
+τ|AΣ2|2+aτ|AΣ1|2+(τ1)q=36[A22Σ1AqqΣ1(A2qΣ1)2]\displaystyle\qquad\qquad\qquad+\tau|A_{\Sigma_{2}}|^{2}+a\tau|A_{\Sigma_{1}}|^{2}+(\tau-1)\sum_{q=3}^{6}\left[A^{\Sigma_{1}}_{22}A^{\Sigma_{1}}_{qq}-(A^{\Sigma_{1}}_{2q})^{2}\right]
+τΣ1h,e2+aτHΣ2u1Σ1u,e2+aτu2Σ1u,e22,\displaystyle\qquad\qquad\qquad+\tau\langle\nabla_{\Sigma_{1}}h,e_{2}\rangle+a\tau H_{\Sigma_{2}}u^{-1}\langle\nabla_{\Sigma_{1}}u,e_{2}\rangle+a\tau u^{-2}\langle\nabla_{\Sigma_{1}}u,e_{2}\rangle^{2},

where \mathcal{R} denotes a sum of 9 terms of the form (τ1)RM¯(X,Y,X,Y)(\tau-1)R_{\overline{M}}(X,Y,X,Y) or (aτ1)RM¯(X,Y,X,Y)(a\tau-1)R_{\overline{M}}(X,Y,X,Y) and therefore satisfies ||12|\mathcal{R}|\leq\frac{1}{2} by condition (i) on aa.

Define the quantity

𝒦:=\displaystyle\mathcal{K}:= RicΣ2(e3,e3)+RicΣ1(e2,e2)+RicM(e1,e1)\displaystyle\operatorname{Ric}_{\Sigma_{2}}(e_{3},e_{3})+\operatorname{Ric}_{\Sigma_{1}}(e_{2},e_{2})+\operatorname{Ric}_{M}(e_{1},e_{1})
+τ|AΣ2|2+aτ|AΣ1|2+(τ1)q=36[A22Σ1AqqΣ1(A2qΣ1)2]\displaystyle\quad+\tau|A_{\Sigma_{2}}|^{2}+a\tau|A_{\Sigma_{1}}|^{2}+(\tau-1)\sum_{q=3}^{6}\left[A^{\Sigma_{1}}_{22}A^{\Sigma_{1}}_{qq}-(A^{\Sigma_{1}}_{2q})^{2}\right]
+τΣ1h,e2+aτHΣ2u1Σ1u,e2+aτu2Σ1u,e22.\displaystyle\quad+\tau\langle\nabla_{\Sigma_{1}}h,e_{2}\rangle+a\tau H_{\Sigma_{2}}u^{-1}\langle\nabla_{\Sigma_{1}}u,e_{2}\rangle+a\tau u^{-2}\langle\nabla_{\Sigma_{1}}u,e_{2}\rangle^{2}.

By the bound on \mathcal{R}, to prove the proposition, it suffices to show that 𝒦1\mathcal{K}\geq 1. According to [2, Lemma 3.8], we have

(12) 𝒦=C3(e1,e2,e3)+1+2+τΣ1h,e2,\displaystyle\mathcal{K}=C_{3}(e_{1},e_{2},e_{3})+\mathcal{B}_{1}+\mathcal{B}_{2}+\tau\langle\nabla_{\Sigma_{1}}h,e_{2}\rangle,

where

1=aτ|AΣ1|2+τq=36[A22Σ1AqqΣ1(A2qΣ1)2]+q=46[A33Σ1AqqΣ1(A3qΣ1)2]\mathcal{B}_{1}=a\tau|A_{\Sigma_{1}}|^{2}+\tau\sum_{q=3}^{6}\left[A^{\Sigma_{1}}_{22}A^{\Sigma_{1}}_{qq}-(A^{\Sigma_{1}}_{2q})^{2}\right]+\sum_{q=4}^{6}\left[A^{\Sigma_{1}}_{33}A^{\Sigma_{1}}_{qq}-(A^{\Sigma_{1}}_{3q})^{2}\right]

and

2=τ|AΣ2|2+q=46[A33Σ2AqqΣ2(ApqΣ2)2]+aτHΣ2u1Σ1u,e2+aτu2Σ1u,e22.\mathcal{B}_{2}=\tau|A_{\Sigma_{2}}|^{2}+\sum_{q=4}^{6}\left[A^{\Sigma_{2}}_{33}A^{\Sigma_{2}}_{qq}-(A^{\Sigma_{2}}_{pq})^{2}\right]+a\tau H_{\Sigma_{2}}u^{-1}\langle\nabla_{\Sigma_{1}}u,e_{2}\rangle+a\tau u^{-2}\langle\nabla_{\Sigma_{1}}u,e_{2}\rangle^{2}.

Next we focus on obtaining lower bounds for 1\mathcal{B}_{1} and 2\mathcal{B}_{2}.

Lemma 39.

The quantity 1\mathcal{B}_{1} is non-negative.

Proof.

Dropping the subscript and superscript Σ1\Sigma_{1}’s, we have

1=aτ|A|2+τq=36[A22AqqA2q2]+q=46[A33AqqA3q2].\displaystyle\mathcal{B}_{1}=a\tau|A|^{2}+\tau\sum_{q=3}^{6}\left[A_{22}A_{qq}-A_{2q}^{2}\right]+\sum_{q=4}^{6}\left[A_{33}A_{qq}-A_{3q}^{2}\right].

Since Σ1\Sigma_{1} is minimal and aτ>a2>1/2a\tau>a^{2}>1/2, it follows that

1\displaystyle\mathcal{B}_{1} aτ(A222+A332+A442+A552+A662)+τ(A22A33+A22A44+A22A55+A22A66)\displaystyle\geq a\tau(A_{22}^{2}+A_{33}^{2}+A_{44}^{2}+A_{55}^{2}+A_{66}^{2})+\tau(A_{22}A_{33}+A_{22}A_{44}+A_{22}A_{55}+A_{22}A_{66})
+(A33A44+A33A55+A33A66)\displaystyle\qquad+(A_{33}A_{44}+A_{33}A_{55}+A_{33}A_{66})
=(aτ12)(A222+A332+A442+A552+A662)+12(A22+A33+A44+A55+A66)2\displaystyle=(a\tau-\frac{1}{2})(A_{22}^{2}+A_{33}^{2}+A_{44}^{2}+A_{55}^{2}+A_{66}^{2})+\frac{1}{2}(A_{22}+A_{33}+A_{44}+A_{55}+A_{66})^{2}
+(τ1)(A22A33+A22A44+A22A55+A22A66)A44A55A44A66A55A66\displaystyle\qquad+(\tau-1)(A_{22}A_{33}+A_{22}A_{44}+A_{22}A_{55}+A_{22}A_{66})-A_{44}A_{55}-A_{44}A_{66}-A_{55}A_{66}
=(aτ12)(A222+A332+A442+A552+A662)\displaystyle=(a\tau-\frac{1}{2})(A_{22}^{2}+A_{33}^{2}+A_{44}^{2}+A_{55}^{2}+A_{66}^{2})
+(τ1)(A22A33+A22A44+A22A55+A22A66)A44A55A44A66A55A66.\displaystyle\qquad+(\tau-1)(A_{22}A_{33}+A_{22}A_{44}+A_{22}A_{55}+A_{22}A_{66})-A_{44}A_{55}-A_{44}A_{66}-A_{55}A_{66}.

Again, since Σ1\Sigma_{1} is minimal, we have

A222+A33212(A22+A33)2=12(A44+A55+A66)2.A_{22}^{2}+A_{33}^{2}\geq\frac{1}{2}(A_{22}+A_{33})^{2}=\frac{1}{2}(A_{44}+A_{55}+A_{66})^{2}.

Thus we obtain

1\displaystyle\mathcal{B}_{1} 2(1τ)(A222+A332+A442+A552+A662)+(τ1)(A22A33+A22A44+A22A55+A22A66)\displaystyle\geq 2(1-\tau)(A_{22}^{2}+A_{33}^{2}+A_{44}^{2}+A_{55}^{2}+A_{66}^{2})+(\tau-1)(A_{22}A_{33}+A_{22}A_{44}+A_{22}A_{55}+A_{22}A_{66})
+32(2τ52+aτ)(A442+A552+A662)+(2τ72+aτ)(A44A55+A44A66+A55A66)\displaystyle\qquad+\frac{3}{2}(2\tau-\frac{5}{2}+a\tau)(A_{44}^{2}+A_{55}^{2}+A_{66}^{2})+(2\tau-\frac{7}{2}+a\tau)(A_{44}A_{55}+A_{44}A_{66}+A_{55}A_{66})
32(2τ52+aτ)(A442+A552+A662)+(2τ72+aτ)(A44A55+A44A66+A55A66).\displaystyle\geq\frac{3}{2}(2\tau-\frac{5}{2}+a\tau)(A_{44}^{2}+A_{55}^{2}+A_{66}^{2})+(2\tau-\frac{7}{2}+a\tau)(A_{44}A_{55}+A_{44}A_{66}+A_{55}A_{66}).

This will be non-negative as long as

32(2τ52+aτ)722τaτ.\frac{3}{2}(2\tau-\frac{5}{2}+a\tau)\geq\frac{7}{2}-2\tau-a\tau.

This is equivalent to

5τ+52τa294,5\tau+\frac{5}{2}\tau a\geq\frac{29}{4},

which holds since τ>a\tau>a and a>390101a>\frac{\sqrt{390}}{10}-1. ∎

Lemma 40.

The quantity 2\mathcal{B}_{2} satisfies 2αh2\mathcal{B}_{2}\geq\alpha h^{2}.

Proof.

Using HΣ2=hau1Σ1u,ηH_{\Sigma_{2}}=h-au^{-1}\langle\nabla_{\Sigma_{1}}u,\eta\rangle, and dropping the subscript and superscript Σ2\Sigma_{2}’s, we have

2αh2=τ|A|2+q=46[A33Aqq(Apq)2]+τH(hH)+τa(Hh)2αh2.\displaystyle\mathcal{B}_{2}-\alpha h^{2}=\tau|A|^{2}+\sum_{q=4}^{6}\left[A_{33}A_{qq}-(A_{pq})^{2}\right]+\tau H(h-H)+\frac{\tau}{a}(H-h)^{2}-\alpha h^{2}.

Now we analyze this quadratic form as in Mazet [14]. Write

A=(H4+Φ33A34A35A36A43H4+Φ44A45A46A53A54H4+Φ55A56A63A64A65H4+Φ66).A=\begin{pmatrix}\frac{H}{4}+\Phi_{33}&A_{34}&A_{35}&A_{36}\\ A_{43}&\frac{H}{4}+\Phi_{44}&A_{45}&A_{46}\\ A_{53}&A_{54}&\frac{H}{4}+\Phi_{55}&A_{56}\\ A_{63}&A_{64}&A_{65}&\frac{H}{4}+\Phi_{66}\end{pmatrix}.

Then, since τ12\tau\geq\frac{1}{2}, we have

τ|A|2+q=46[A33Aqq(Apq)2]+τH(hH)+τa(Hh)2αh2\displaystyle\tau|A|^{2}+\sum_{q=4}^{6}\left[A_{33}A_{qq}-(A_{pq})^{2}\right]+\tau H(h-H)+\frac{\tau}{a}(H-h)^{2}-\alpha h^{2}
τH24+τ|Φ|2+A33(HA33)+τH(hH)+τa(Hh)2αh2\displaystyle\qquad\geq\tau\frac{H^{2}}{4}+\tau|\Phi|^{2}+A_{33}(H-A_{33})+\tau H(h-H)+\frac{\tau}{a}(H-h)^{2}-\alpha h^{2}
=τH24+τ|Φ|2+(H4+Φ33)(34HΦ33)+τH(hH)+τa(Hh)2αh2\displaystyle\qquad=\tau\frac{H^{2}}{4}+\tau|\Phi|^{2}+(\frac{H}{4}+\Phi_{33})(\frac{3}{4}H-\Phi_{33})+\tau H(h-H)+\frac{\tau}{a}(H-h)^{2}-\alpha h^{2}
=(31634τ+τa)H2+(τaα)h2+τ|Φ|2Φ332+(τ2τa)Hh+12HΦ33.\displaystyle\qquad=\left(\frac{3}{16}-\frac{3}{4}\tau+\frac{\tau}{a}\right)H^{2}+\left(\frac{\tau}{a}-\alpha\right)h^{2}+\tau|\Phi|^{2}-\Phi_{33}^{2}+\left(\tau-\frac{2\tau}{a}\right)Hh+\frac{1}{2}H\Phi_{33}.

Let E={(x3,x4,x5,x6)4:x3+x4+x5+x6=0}E=\{(x_{3},x_{4},x_{5},x_{6})\in\mathbb{R}^{4}:x_{3}+x_{4}+x_{5}+x_{6}=0\}. Then any vector in EE can be written in the form

12(0011)y1+16(0211)y2+112(3111)z.\frac{1}{\sqrt{2}}\begin{pmatrix}0\\ 0\\ 1\\ -1\end{pmatrix}y_{1}+\frac{1}{\sqrt{6}}\begin{pmatrix}0\\ 2\\ -1\\ -1\end{pmatrix}y_{2}+\frac{1}{\sqrt{12}}\begin{pmatrix}3\\ -1\\ -1\\ -1\end{pmatrix}z.

Expressing Φ\Phi in these coordinates, the previous quadratic form is at least

(31634τ+τa)H2+(τaα)h2+(τ34)z2+(τ2τa)Hh+34Hz.\left(\frac{3}{16}-\frac{3}{4}\tau+\frac{\tau}{a}\right)H^{2}+\left(\frac{\tau}{a}-\alpha\right)h^{2}+\left(\tau-\frac{3}{4}\right)z^{2}+\left(\tau-\frac{2\tau}{a}\right)Hh+\frac{\sqrt{3}}{4}Hz.

Condition (iii) on aa, τ\tau, and α\alpha ensures that this quadratic form in HH, hh, and zz is positive definite. The lemma follows. ∎

Combining (12) with Lemma 39, Lemma 40, the assumption on the intermediate curvature, and (11) we now obtain

𝒦3+αh2τΣ1h1.\mathcal{K}\geq 3+\alpha h^{2}-\tau\|\nabla_{\Sigma_{1}}h\|\geq 1.

Proposition 38 follows. ∎

Finally, we can finish the proof of Theorem 37. Since Σ2=0\partial{\Sigma_{2}}=0 and diam(Σ2)\operatorname{diam}({\Sigma_{2}}) is uniformly bounded and H4(M¯,)=0H_{4}(\overline{M},\mathbb{Z})=0, it follows from Proposition 14 that Σ2{\Sigma_{2}} bounds within an RR-neighborhood for some constant RR that depends only on M¯\overline{M}. Applying this argument to each component of Ω\partial\Omega, we see that Ω\partial\Omega bounds within its RR-neighborhood in M¯\overline{M}. Since d(Ω,σ)L/4d(\partial\Omega,\sigma)\geq L/4 and Ω\partial\Omega is linked with σ\sigma by construction, this is a contradiction provided we select L>2RL>2R. This completes the proof.

5. Mapping Version and Classification

In this section we prove a mapping version and a refinement of Theorem 4 into a positive result. We start with the lifting Lemma proven by Chodosh-Li-Liokumovich [4].

Lemma 41 (Chodosh-Li-Liokumovich).

Suppose that X,NX,N are closed oriented manifolds and f:NXf:N\to X has non-zero degree. Letting X¯\bar{X} denote the universal covering of XX, there exists a connected cover N^N\hat{N}\to N and a lift f^:N^X¯\hat{f}:\hat{N}\to\bar{X} such that f^\hat{f} is proper and degf^=degf\deg\hat{f}=\deg f.

Remark 42.

It can be checked that f^\hat{f} is a Lipschitz map.

Corollary 43 (Mapping Version).

Suppose MnM^{n} is a closed manifold and that the universal cover M¯\overline{M} of MM satisfies

Hn(M¯,)=Hn1(M¯,)==Hnm+1(M¯,)=0.H_{n}(\overline{M},\mathbb{Z})=H_{n-1}(\overline{M},\mathbb{Z})=\ldots=H_{n-m+1}(\overline{M},\mathbb{Z})=0.

Assume further NnN^{n} is a closed manifold with non-zero degree map to MnM^{n}, then NN does not admit a metric of positive mm-intermediate curvature in any of the following cases: n{3,4,5}n\in\{3,4,5\} and m{1,2,,n1}m\in\{1,2,\dots,n-1\}; n=6n=6 and m{1,2,3,4}m\in\{1,2,{\color[rgb]{0,.5,.5}\definecolor[named]{pgfstrokecolor}{rgb}{0,.5,.5}3},4\}; nn is arbitrary and m=1m=1.

Proof.

Since f^\hat{f} is proper with non-zero degree and since it is a Lipschitz map, all the diameter estimate arguments are true in the connected cover N^\hat{N} given by Lemma 41, the result then follows. ∎

We now summarize the filling estimates for n=6,m=4n=6,m=4 in previous sections as follows.

Theorem 44 (Filling Estimates).

Suppose (N6,g)(N^{6},g) is a closed manifold with positive 44-intermediate curvature no less than 32\frac{3}{2}. Fix a connected Riemannian cover N^\hat{N}:

Consider a closed embedded 44-manifold Λ^\hat{\Lambda} such that [Λ^]=0H4(N^;)[\hat{\Lambda}]=0\in H_{4}(\hat{N};\mathbb{Z}). Then there exists a 55-chain Σ^BL(Λ^)\hat{\Sigma}\subset B_{L}(\hat{\Lambda}) and a 44-dimensional submanifold Γ^\hat{\Gamma} such that

Σ^=Λ^Γ^\partial\hat{\Sigma}=\hat{\Lambda}-\hat{\Gamma}

as chains.

Furthermore, there are 44-chains K^1,,K^s\hat{K}_{1},\cdots,\hat{K}_{s} with diameter bounded by LL and 33-cycles {Υ^jl}\{\hat{\Upsilon}_{j}^{l}\} where j=1,2,,sj=1,2,\cdots,s and l=1,2,,n(j)l=1,2,\cdots,n(j) such that

Γ^=j=1sK^j,K^j=l=1n(j)Υ^jl,\hat{\Gamma}=\sum_{j=1}^{s}\hat{K}_{j},\quad\partial\hat{K}_{j}=\sum_{l=1}^{n(j)}\hat{\Upsilon}_{j}^{l},

where Υjl\Upsilon_{j}^{l} has diameter bounded by LL, and the above equalities hold as chains.

Finally, there is an integer qq and a function

u:{(j,l):j=1,2,,s,l=1,,n(j)}{1,2,,q}u:\{(j,l):j=1,2,\cdots,s,l=1,\cdots,n(j)\}\to\{1,2,\cdots,q\}

such that for any t{1,2,,q}t\in\{1,2,\cdots,q\},

diam((j,l)u1(t)Υ^jl)L\operatorname{diam}\left(\cup_{(j,l)\in u^{-1}(t)}\hat{\Upsilon}_{j}^{l}\right)\leq L

and

u(j,l)=u1(r)Υ^jl=0\sum_{u(j,l)=u^{-1}(r)}\hat{\Upsilon}_{j}^{l}=0

as chains.

Theorem 45.

Suppose NnN^{n} is a closed manifold which admits positive mm-intermediate curvature. Further suppose that either

  1. (1)

    n=5n=5, m=2m=2, and π2(N5)=0\pi_{2}(N^{5})=0; or

  2. (2)

    n=6n=6, m=4m=4, and π2(N6)=π3(N6)=π4(N6)=0\pi_{2}(N^{6})=\pi_{3}(N^{6})=\pi_{4}(N^{6})=0.

Then a finite covering of NN is homeomorphic to SnS^{n} or connected sum of Sn1×S1S^{n-1}\times S^{1}.

Proof.

The idea is to prove the fundamental group of NN is virtually free, and then with the aid of Theorem 1.3 in Gadgil-Seshadri [5], we obtain the desired result.

Case 1: n=5,m=2n=5,m=2.

By Ma [13, Theorem 1.6], π1(N5)\pi_{1}(N^{5}) is virtually free. Thus there exists a finite connected cover N^\hat{N} of NN such that π1(N^)\pi_{1}(\hat{N}) is free. Since π2(N)=0\pi_{2}(N)=0, by Gadgil- Seshadri [5, Theorem 1.3], the proof then follows.

Case 2: n=6,m=4n=6,m=4. Since π2(N)=π3(N)=π4(N)=0\pi_{2}(N)=\pi_{3}(N)=\pi_{4}(N)=0, by the Hurewicz Theorem, the universal covering N~\tilde{N} of NN has trivial 44th homology group. Then as a direct Corollary of Theorem 44 and Proposition 14, there exists an L=L(N,g)>0L=L(N,g)>0 such that for a closed embedded 44-submanifold in N~\tilde{N} is null-homologous in its LL-neighborhood. By Chodosh-Li-Liokumovich [4, Proposition 8], for any point pN~p\in\tilde{N}, each connected component of a level set of d(p,)d(p,\cdot) has diameter bounded by 20L20L. By Chosdosh-Li-Liokumovich [4, Corollary 14], π1(N)\pi_{1}(N) is virtually free. Since π2(N)=π3(N)=0\pi_{2}(N)=\pi_{3}(N)=0, by Gadgil- Seshadri [5, Theorem 1.3], the proof then follows. ∎

With Lemma 41, we have the following mapping version classification.

Corollary 46.

Suppose NnN^{n} is a closed manifold admits positive mm-intermediate curvature, assume further there exists a non-zero degree map f:NXf:N\to X, where XX is a closed manifold satisfying

  1. (1)

    when n=5n=5, m=2m=2, π2(X5)=0\pi_{2}(X^{5})=0;

  2. (2)

    when n=6n=6, m=4m=4, π2(X6)=π3(X6)=π4(X6)=0\pi_{2}(X^{6})=\pi_{3}(X^{6})=\pi_{4}(X^{6})=0.

Then a finite covering of NN is homeomorphic to SnS^{n} or connected sum of Sn1×S1S^{n-1}\times S^{1}.

References

  • [1] Richard H Bamler, Chao Li, and Christos Mantoulidis. Decomposing 4-manifolds with positive scalar curvature. Advances in Mathematics, 430:109231, 2023.
  • [2] Simon Brendle, Sven Hirsch, and Florian Johne. A generalization of geroch’s conjecture. Communications on Pure and Applied Mathematics, 77(1):441–456, 2024.
  • [3] Otis Chodosh and Chao Li. Generalized soap bubbles and the topology of manifolds with positive scalar curvature. Annals of Mathematics, 199(2):707–740, 2024.
  • [4] Otis Chodosh, Chao Li, and Yevgeny Liokumovich. Classifying sufficiently connected PSC manifolds in 4 and 5 dimensions. Geometry & Topology, 27(4):1635–1655, 2023.
  • [5] Siddartha Gadgil and Harish Seshadri. On the topology of manifolds with positive isotropic curvature. Proceedings of the American Mathematical Society, 137(5):1807–1811, 2009.
  • [6] Mikhael Gromov. Large riemannian manifolds. In Curvature and Topology of Riemannian Manifolds: Proceedings of the 17th International Taniguchi Symposium held in Katata, Japan, Aug. 26–31, 1985, pages 108–121. Springer, 2006.
  • [7] Mikhael Gromov and H Blaine Lawson. Positive scalar curvature and the Dirac operator on complete Riemannian manifolds. Publications Mathématiques de l’IHÉS, 58:83–196, 1983.
  • [8] Mikhael Gromov and H Blaine Lawson Jr. Spin and scalar curvature in the presence of a fundamental group. I. Annals of Mathematics, pages 209–230, 1980.
  • [9] Misha Gromov. Four lectures on scalar curvature. arXiv preprint arXiv:1908.10612, 2019.
  • [10] Misha Gromov. No metrics with positive scalar curvatures on aspherical 5-manifolds. arXiv preprint arXiv:2009.05332, 2020.
  • [11] Allen Hatcher. Algebraic Topology. Cambridge University Press, 2002.
  • [12] Shihang He and Jintian Zhu. A note on rational homology vanishing theorem for hypersurfaces in aspherical manifolds. arXiv preprint arXiv:2311.14008, 2024.
  • [13] Junyu Ma. Urysohn 1-width for 4 and 5 manifolds with positive biRicci curvature. arXiv preprint arXiv:2402.18141, 2024.
  • [14] Laurent Mazet. Stable minimal hypersurfaces in 6\mathbb{R}^{6}. arXiv preprint arXiv:2405.14676, 2024.
  • [15] Richard Schoen and Shing-Tung Yau. Existence of incompressible minimal surfaces and the topology of three dimensional manifolds with non-negative scalar curvature. Annals of Mathematics, 110(1):127–142, 1979.
  • [16] Richard Schoen and Shing-Tung Yau. The structure of manifolds with positive scalar curvature. In Directions in partial differential equations, pages 235–242. Elsevier, 1987.
  • [17] Ying Shen and Rugang Ye. On stable minimal surfaces in manifolds of positive bi-Ricci curvatures. Duke Math. J., 85(1):109–116, 1996.
  • [18] Ying Shen and Rugang Ye. On the geometry and topology of manifolds of positive bi-Ricci curvature. arXiv preprint dg-ga/9708014, 1997.
  • [19] Kai Xu. Dimension constraints in some problems involving intermediate curvature. Transactions of the American Mathematical Society, 2024.
  • [20] Jintian Zhu. Width estimate and doubly warped product. Transactions of the American Mathematical Society, 374(2):1497–1511, 2021.