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

Calabi-Yau type theorem for complete manifolds with nonnegative scalar curvature

Jintian Zhu Institute for Theoretical Sciences, Westlake University, 600 Dunyu Road, 310030, Hangzhou, Zhejiang, People’s Republic of China [email protected]
Abstract.

In this paper, we are able to prove an analogy of the Calabi-Yau theorem for complete Riemannian manifolds with nonnegative scalar curvature which are aspherical at infinity. The key tool is an existence result for arbitrarily large bounded regions with weakly mean-concave boundary in Riemannian manifolds with sublinear volume growth. As an application, we use the same tool to show that a complete contractible Riemannian 33-manifold with positive scalar curvature and sublinear volume growth is necessarily homeomorphic to 3\mathbb{R}^{3}.

1. Introduction

Let (M,g)(M,g) be a complete Riemannian manifold. Recall that (M,g)(M,g) has sublinear volume growth if it holds

lim infr+volg(Br(p))r=0 for some point pM,\liminf_{r\to+\infty}\frac{\operatorname{vol}_{g}(B_{r}(p))}{r}=0\mbox{ for some point }p\in M,

where the choice of the point pp plays no role in this definition. It is well-known that there are interplays between nonnegative curvatures and the underlying topology on complete Riemannian manifolds with sublinear volume growth. The famous Calabi-Yau theorem [Cal75, Yau76] states that if (M,g)(M,g) has nonnegative Ricci curvature, then it has sublinear volume growth if and only if it is closed. Very recently, such purely geometric way of characterizing compactness for complete manifolds under nonnegative-curvature condition plays a crucial role in the author’s establishment of various rigidity results on complete manifolds with nonnegative scalar curvature (see the author’s previous works [Zhu23a, Zhu23+]).

The goal of this paper is to establish an analogy of the Calabi-Yau theorem with much weaker condition of nonnegative scalar curvature. First we point out that additional conditions are necessary in order to obtain such kind of result since there are indeed open complete Riemannian manifolds with nonnegative scalar curvature even having finite volume (see Appendix A).

As a natural choice, we consider the class of manifolds which are aspherical at infinity. Recall that a manifold MM is said to be aspherical at infinity means that for each positive integer i2i\geq 2 and compact subset KMK\subset M there is a larger compact subset K~K\tilde{K}\supset K such that the inclusion map πi(MK~)πi(MK)\pi_{i}(M\setminus\tilde{K})\to\pi_{i}(M\setminus K) is the zero map. Our main theorem is

Theorem 1.1.

Let (Mn,g)(M^{n},g), 2n42\leq n\leq 4, be a complete Riemannian manifold with nonnegative scalar curvature outside a compact subset, which is aspherical at infinity. Then it has sublinear volume growth if and only if it is closed.

We emphasize that the extra aspherical condition at infinity is natural in the sense that every open surface appears to be aspherical at infinity. In other words, Theorem 1.1 is indeed a generalization of the 2D Calabi-Yau theorem in the context of scalar curvature.

Given the Calabi-Yau type Theorem 1.1 and also the aspherical splitting theorem established in [HZ23+, Theorem 1.7], there may be an underlying principle saying that nonnegative scalar curvature on complete open manifolds with extra natural topological condition has similar behaviours with nonnegative Ricci curvature on complete open manifolds. Here natural means that open surfaces automatically satisfy these topological conditions.

The main strategy to prove Theorem 1.1 is the following existence lemma for an arbitrarily large bounded region with weakly mean-concave boundary in complete open Riemannian manifolds with sublinear volume growth. The proof is based on the μ\mu-bubble method raised by Gromov [Gro23]. Recently, Gromov’s μ\mu-bubble method has led to applications in many other geometric problems (refer to [Gro20+, CL20+, LUY21+, Zhu21, CL23, CRZ23, Zhu23b, CLMS24+] and references therein).

Lemma 1.2.

Let (Mn,g)(M^{n},g), 2n72\leq n\leq 7, be an open complete Riemannian manifold with sublinear volume growth. Then for any bounded region KK we can find a larger bounded region Ω^K\hat{\Omega}\supset K such that Ω\partial\Omega is weakly mean-concave with respect to the unit outer normal of Ω\partial\Omega in Ω\Omega.

We note that it follows from the previous works [CL20, Song23] that there is at least one embedded minimal hypersurface in a complete Riemannian manifold with sublinear volume growth, but the compactness of the minimal hypersurface cannot be guaranteed therein. In this work, the pass from a minimal hypersurface to a μ\mu-bubble successfully handle the compactness issue.

With Lemma 1.2 we are able to provide an alternative proof of the Calabi-Yau theorem combined with the splitting theorem [CK92, Theorem 2], which has the obvious advantage that the Bishop-Gromov volume comparison need not be used, and this explains the reason why our method can be generalized to the weaker setting of scalar curvature. With the same philosophy, we just reduce Theorem 1.1 to the validity of the following splitting theorem:

Proposition 1.3.

Let (Mn,g)(M^{n}_{\infty},g), 3n73\leq n\leq 7, be a complete and non-compact Riemannian manifold with compact and weakly mean-convex boundary M\partial M_{\infty}. If (M,g)(M_{\infty},g) has nonnegative scalar curvature and any embedded hypersurface representing a non-trivial \mathbb{Q}-homology class cannot admit any metric with positive scalar curvature, then (M,g)(M_{\infty},g) must split into the Riemannian product (M,g|M)×[0,+)(\partial M_{\infty},g|_{\partial M_{\infty}})\times[0,+\infty).

In the same spirit of the Calabi-Yau theorem, we also mention the Gromov-Lawson theorem below [GL83, Theorem 8.11].

Theorem 1.4 (Gromov-Lawson).

Let Σ\Sigma be a complete non-compact Riemannian surface with the property that, for some fixed constant α>1/2\alpha>1/2,

Iα(φ):=Σ|φ|2+αKφ2dσ0 for all φC0(Σ),I_{\alpha}(\varphi):=\int_{\Sigma}|\nabla\varphi|^{2}+\alpha K\varphi^{2}\,\mathrm{d}\sigma\geq 0\mbox{ for all }\varphi\in C_{0}^{\infty}(\Sigma),

where KK is the Gaussian curvature of Σ\Sigma. Then Σ\Sigma has infinite volume.

We remark that the nonnegative curvature condition is weakened from the pointwise sense to some spectrum sense in above Gromov-Lawson theorem, which seems even not to be well understood in the context of Ricci curvature. Concerning our Theorem 1.1 it is very natural to ask the following

Question 1.5.

Let α\alpha be a positive constant greater than 1/41/4. Let (Mn,g)(M^{n},g), n3n\geq 3, be a complete Riemannian manifold with

M|gφ|2+αR(g)φ2dμg0 for all φ0C0(M),\int_{M}|\nabla_{g}\varphi|^{2}+\alpha R(g)\varphi^{2}\,\mathrm{d}\mu_{g}\geq 0\mbox{ for all }\varphi\neq 0\in C_{0}^{\infty}(M),

which is aspherical at infinity. If (M,g)(M,g) has finite volume, does it have to be compact?

For further application of Lemma 1.2 we mention the following

Question 1.6.

Is any complete contractible 33-manifold with positive scalar curvature necessarily homeomorphic to 3\mathbb{R}^{3}?

This question was considered by Wang in his works [Wang19+, Wang23+], where he proved that if (M,g)(M,g) is a complete contractible 33-manifold with positive scalar curvature, then its fundamental group at infinity must be trivial. As a special case of Question 1.6, we are able to prove

Theorem 1.7.

Let (M3,g)(M^{3},g) be a complete and contractible Riemannian 33-manifold with positive scalar curvature and sublinear volume growth. Then MM is homeomorphic to 3\mathbb{R}^{3}.

Acknowledgement

The author is grateful to Prof. Chao Li for inspiring conversations. He also thianks Dr. Jian Wang and Dr. Liman Chen for helpful suggestions. The author is partially supported by National Key R&D Program of China with grant no. 2020YFA0712800 and 2023YFA1009900 as well as the start-up fund from Westlake University.

2. The proof

Proof of Lemma 1.2.

By enlarging the region KK we may assume it to be a smooth connected and bounded region such that every component of MKM\setminus K is unbounded. In the following, we just focus on one fixed component of MKM\setminus K, denoted by EE. Two things need to be handled in the search of a bounded and weakly mean-concave region Ω^\hat{\Omega} containing KK using variational method, including the obstacle issue caused by the existence of inner boundary E\partial E and a possible loss of compactness due to the non-compactness of EE.

The obstacle issue is overcome by inserting a separation band VV, from which we determine certain geometric quantities for later use in setting the μ\mu-bubble problem. Fix a bounded open neighborhood UMU\subset M of KK and we take the separation band VV to be (U¯K)E(\overline{U}\setminus K)\cap E. For convenience, we make the illustration of the separation band VV in Figure 1, where the boundary V\partial V can be divided into two parts E\partial E and U\partial U respectively. To clarify we point out that both E\partial E and U\partial U can be disconnected in its worst case.

Refer to caption
Figure 1. The separation band VV

Let us introduce two positive constants associated to the separation band VV. The first constant is designed to describe the infimum area that an arbitrary homologically non-trivial hypersurface can have. For the definition we denote 𝒞V\mathcal{C}_{V} to be the collection

{integer multiplicity rectifiable (n1)-currents T with sptTVand T=0 such that T is homologically non-trivial}\left\{\begin{array}[]{c}\mbox{integer multiplicity rectifiable $(n-1)$-currents $T$ with $\operatorname{spt}T\subset V$}\\ \mbox{and $\partial T=0$ such that $T$ is homologically non-trivial}\end{array}\right\}

and take

cgap=inf{𝕄g(T):T𝒞V}.c_{gap}=\inf\left\{\mathbb{M}_{g}(T):T\in\mathcal{C}_{V}\right\}.

Here we recommend the audience to consult the book [Leo83] for the precise definitions for current TT and mass 𝕄g(T)\mathbb{M}_{g}(T). For our purpose, it is enough to consider currents as oriented hypersurfaces and mass as their areas. The second constant is used to measure the height of VV in the sense of area, which is defined by

cheight=inf{gn1(Σ)|ΣV is a connected minimal hypersurfacewith ΣV intersecting both E and U}.c_{height}=\inf\left\{\mathcal{H}^{n-1}_{g}(\Sigma)\left|\begin{array}[]{c}\mbox{$\Sigma\subset V$ is a connected minimal hypersurface}\\ \mbox{with $\partial\Sigma\subset\partial V$ intersecting both $\partial E$ and $\partial U$}\end{array}\right.\right\}.

Now let us verify the positivity of the constants cgapc_{gap} and cheightc_{height} as claimed. To see that c1c_{1} is positive, we take a conformal metric gˇ\check{g} of gg such that

  • c2ggˇc2gc^{-2}g\leq\check{g}\leq c^{2}g for some positive constant cc;

  • and (V,gˇ)(V,\check{g}) is a compact Riemannian manifold with convex boundary.

From geometric measure theory we can find a smooth hypersurface Σˇ\check{\Sigma} which attains the least gˇ\check{g}-area among all the currents in 𝒞V\mathcal{C}_{V}. In particular, we have

cgapc1ngˇn1(Σˇ)>0.c_{gap}\geq c^{1-n}\cdot\mathcal{H}_{\check{g}}^{n-1}(\check{\Sigma})>0.

For the positivity of c2c_{2} we take a smooth hypersurface Σsep\Sigma_{sep} separating E\partial E and U\partial U. Let ΣV\Sigma\subset V be any connected minimal hypersurface with ΣV\partial\Sigma\subset\partial V intersecting both E\partial E and U\partial U. Clearly, Σ\Sigma must have non-empty intersection with Σsep\Sigma_{sep} and so it follows from the monotonicity formula that

gn1(Σ)c0(V,g,dist(Σsep,V))>0,\mathcal{H}_{g}^{n-1}(\Sigma)\geq c_{0}(V,g,\operatorname{dist}(\Sigma_{sep},\partial V))>0,

which yields cheight>0c_{height}>0.

At this stage we are ready to set the μ\mu-bubble problem to overcome the non-compactness issue and to find the desired bounded and weakly mean-concve region Ω^\hat{\Omega}. Recall that (M,g)(M,g) has sublinear volume growth. By definition we can find a sequence of positive constants ri+r_{i}\to+\infty such that for some point pp we have

volg(Bri(p))ri0 as i.\frac{\operatorname{vol}_{g}(B_{r_{i}}(p))}{r_{i}}\to 0\mbox{ as }i\to\infty.

Fix a smooth and proper function ρ:E[0,+)\rho:E\to[0,+\infty) with ρ1(0)=E\rho^{-1}(0)=\partial E and Lipρ2\operatorname{Lip}\rho\leq 2 as well as |ρ()dist(,E)|1|\rho(\cdot)-\operatorname{dist}(\cdot,\partial E)|\leq 1. Then it is not difficult to verify

ri1volg(ρ1([0,ri2]))0 as i.r_{i}^{-1}\operatorname{vol}_{g}\left(\rho^{-1}\left(\left[0,\frac{r_{i}}{2}\right]\right)\right)\to 0\mbox{ as }i\to\infty.

From the co-area formula we have

ri4ri2gn1({ρ=τ})dτ={ri/4ρri/2}|dρ|gdgn0 as i.\int_{\frac{r_{i}}{4}}^{\frac{r_{i}}{2}}\mathcal{H}_{g}^{n-1}(\{\rho=\tau\})\,\mathrm{d}\tau=\int_{\{r_{i}/4\leq\rho\leq r_{i}/2\}}|\mathrm{d}\rho|_{g}\,\mathrm{d}\mathcal{H}^{n}_{g}\to 0\mbox{ as }i\to\infty.

Then it follows that for each ii we can find a regular value ri/4<ti<ri/2r_{i}/4<t_{i}<r_{i}/2 of ρ\rho such that gn1({ρ=ti})0 as i.\mathcal{H}^{n-1}_{g}(\{\rho=t_{i}\})\to 0\mbox{ as }i\to\infty. In particular, we can fix tit_{i} large enough such that V{ρ<ti}V\subset\{\rho<t_{i}\} and

gn1({ρ=ti})<min{cgap,cheight}.\mathcal{H}^{n-1}_{g}(\{\rho=t_{i}\})<\min\{c_{gap},c_{height}\}.

Just take another regular value si>tis_{i}>t_{i} of ρ\rho casually and denote WW to be ρ1([0,si])\rho^{-1}([0,s_{i}]). In the following, we take η:[0,si)(,0]\eta:[0,s_{i})\to(-\infty,0] to be a smooth function such that η0\eta\leq 0 everywhere, η0\eta\equiv 0 in [0,ti][0,t_{i}] and η(s)\eta(s)\to-\infty as ssis\to s_{i}. The prescribed mean curvature function is now defined by h=ηρh=\eta\circ\rho.

Consider the class

𝒞={Caccioppoli sets ΩM such thatEcΩ and ΩEcW}\mathcal{C}=\left\{\begin{array}[]{c}\mbox{Caccioppoli sets $\Omega\subset M$ such that}\\ \mbox{$E^{c}\subset\Omega$ and $\Omega\setminus E^{c}\Subset W$}\end{array}\right\}

and the functional

𝒜h(Ω)=gn1(Ω)ΩEchdgn,\mathcal{A}^{h}(\Omega)=\mathcal{H}^{n-1}_{g}(\partial^{*}\Omega)-\int_{\Omega\setminus E^{c}}h\,\mathrm{d}\mathcal{H}^{n}_{g},

where Ω\partial^{*}\Omega is denoted to be the reduced boundary from [Giu77, Definition 3.3]. Through a standard argument from geometric measure theory we can find a minimizer Ω𝒞\Omega^{*}\in\mathcal{C} such that

Ah(Ω)=infΩ𝒞𝒜h(Ω)A^{h}(\Omega^{*})=\inf_{\Omega\in\mathcal{C}}\mathcal{A}^{h}(\Omega)

and that ΩE̊\Omega^{*}\cap\mathring{E} is a region in E̊\mathring{E} with smooth boundary ΩE̊\partial\Omega^{*}\cap\mathring{E}, where E̊\mathring{E} is denoted to be the interior of EE. Denote U1,,UkU^{*}_{1},\ldots,U^{*}_{k} to be the unbounded components of MΩM\setminus\Omega^{*} and Ω^E\hat{\Omega}_{E} to be the unique unbounded component of M(iUi)M\setminus\left(\cup_{i}U^{*}_{i}\right) (the uniqueness comes from the connectedness of KK). It is not difficult to verify MΩ^EEM\setminus\hat{\Omega}_{E}\subset E. We claim that Ω^E\partial\hat{\Omega}_{E} is smooth and weakly mean-concave with respect to the outer unit normal as the boundary of Ω^E\hat{\Omega}_{E}. Let Σ^\hat{\Sigma} be an arbitrary component of Ω^E\partial\hat{\Omega}_{E}. Notice that Σ^\hat{\Sigma} is a common boundary of two unbounded connected regions Ω^E\hat{\Omega}_{E} and some UiU^{*}_{i}, then it has to be homologically non-trivial since we can construct a line intersecting Σ^\hat{\Sigma} only once. Notice that the set Ecρ1([0,ti])E^{c}\cup\rho^{-1}([0,t_{i}]) also belongs to the class 𝒞\mathcal{C} and it follows from a direct comparison that

gn1(Σ^)𝒜h(Ω)gn1({ρ=ti})<min{cgap,cheight}.\mathcal{H}^{n-1}_{g}(\hat{\Sigma})\leq\mathcal{A}^{h}(\Omega^{*})\leq\mathcal{H}^{n-1}_{g}(\{\rho=t_{i}\})<\min\{c_{gap},c_{height}\}. (2.1)

In particular, by definition of cgapc_{gap} we conclude that Σ^(EV)\hat{\Sigma}\cap(E\setminus V) is non-empty. From the first variation formula of 𝒜h\mathcal{A}^{h} it follows that Σ^E̊\hat{\Sigma}\cap\mathring{E} as the boundary of Ω\Omega^{*} has mean curvature h|Σ^E̊h|_{\hat{\Sigma}\cap\mathring{E}} with respect to the outer unit normal. In particular, Σ^V\hat{\Sigma}\cap V is a minimal hypersurface. If Σ^\hat{\Sigma} intersects with E\partial E, then we can find a connected minimal hypersurface Σ^0\hat{\Sigma}_{0} among components of Σ^V\hat{\Sigma}\cap V satisfying Σ^0V\partial\hat{\Sigma}_{0}\subset\partial V and that Σ^\hat{\Sigma} intersects both \partial_{-} and +\partial_{+}. This implies gn1(Σ^)gn1(Σ^0)cheight\mathcal{H}^{n-1}_{g}(\hat{\Sigma})\geq\mathcal{H}^{n-1}_{g}(\hat{\Sigma}_{0})\geq c_{height}, which contradicts to (2.1). Therefore, Σ^\hat{\Sigma} does not touch E\partial E and so it is smooth everywhere. The mean-concavity of Σ^\hat{\Sigma} comes directly from the non-positivity of the function hh after realizing that the outer unit normals of Σ^\hat{\Sigma} with respect to Ω\Omega^{*} and Ω^\hat{\Omega} coincide.

Finally let us take all unbounded components of MKM\setminus K into consideration. After labeling them as E1,E2,,ElE_{1},E_{2},\ldots,E_{l} we can find regions Ω^E1,,Ω^El\hat{\Omega}_{E_{1}},\ldots,\hat{\Omega}_{E_{l}} with smooth weakly mean-concave boundary from above discussion. The desired bounded region is given by

Ω^=i=1lΩ^i,\hat{\Omega}=\bigcap_{i=1}^{l}\hat{\Omega}_{i},

which is obviously a bounded region containing KK with weakly mean-concave boundary. ∎

To be self-contained we would like to provide an alternative proof for the Calabi-Yau theorem, which will be used in the proof of Proposition 1.3. Of course, due to the use of geometric measure theory the dimension is assumed to be no greater than seven.

Theorem 2.1 (Calabi-Yau).

Let (Mn,g)(M^{n},g), n7n\leq 7, be a complete Riemannian manifold with nonnegative Ricci curvature outside a compact subset. Then (M,g)(M,g) has sublinear volume growth if and only if MM is compact.

An alternative proof without using volume comparison.

Denote KK to be a smooth compact subset such that (M,g)(M,g) has nonnegative Ricci curvature outside KK. Suppose that (M,g)(M,g) is non-compact but has sublinear volume growth. Then it follows from Lemma 1.2 that there is a bounded smooth region Ω^\hat{\Omega} containing KK with weakly mean-concave boundary. Notice that Ω^\partial\hat{\Omega} is weakly mean-convex as the boundary of MΩ^M\setminus\hat{\Omega} with respect to the corresponding outer unit normal. Since MM is non-compact, there is at least one unbounded component of MΩ^M\setminus\hat{\Omega} denoted by EE. After applying [CK92, Theorem 2] to the end EE we conclude that EE must be isometric to the Riemannian product (E,g|E)×[0,+)(\partial E,g|_{\partial E})\times[0,+\infty), which obviously has linear volume growth. This leads to a contradiction! ∎

In the following, let us establish the splitting result which is involved in the proof of our main theorem.

Proof of Proposition 1.3.

Here we take a similar argument from the author’s previous work [Zhu23b] based on Gromov’s μ\mu-bubble. Let ρ:M[0,+)\rho:M_{\infty}\to[0,+\infty) be a smooth proper function satisfying ρ1(0)=M\rho^{-1}(0)=\partial M_{\infty} and Lipρ<1\operatorname{Lip}\rho<1. From [Zhu23b, Lemma 2.3] we can construct a smooth function hϵ:[0,1nϵ)(,0]h_{\epsilon}:[0,\frac{1}{n\epsilon})\to(-\infty,0] for any 0<ϵ<10<\epsilon<1 such that

  • hϵh_{\epsilon} satisfies

    nn1hϵ2+2hϵ=n(n1)ϵ2 on [12n,1nϵ)\frac{n}{n-1}h_{\epsilon}^{2}+2h_{\epsilon}^{\prime}=n(n-1)\epsilon^{2}\mbox{ on }\left[\frac{1}{2n},\frac{1}{n\epsilon}\right)

    and there is a universal constant C=C(n)C=C(n) such that

    |nn1hϵ2+2hϵ|LCϵ2.\left|\frac{n}{n-1}h_{\epsilon}^{2}+2h_{\epsilon}\right|_{L^{\infty}}\leq C\epsilon^{2}.
  • hϵ<0h_{\epsilon}<0 and

    limt1nϵhϵ(t)=.\lim_{t\to\frac{1}{n\epsilon}}h_{\epsilon}(t)=-\infty.
  • as ϵ0\epsilon\to 0, hϵh_{\epsilon} converges smoothly to the zero function on any closed interval.

Now we are ready to set appropriate μ\mu-bubble problems. Take ϵ\epsilon such that 1nϵ\frac{1}{n\epsilon} appears to be a regular value of ρ\rho and denote Vϵ=ρ1([0,1nϵ])V_{\epsilon}=\rho^{-1}([0,\frac{1}{n\epsilon}]). Consider the class

𝒞ϵ={Caccioppoli sets Ω such that MΩVϵ}\mathcal{C}_{\epsilon}=\{\mbox{Caccioppoli sets $\Omega$ such that $M_{\infty}\setminus\Omega\Subset V_{\epsilon}$}\}

and the functional

𝒜ϵ(Ω)=gn1(Ω)MΩhϵρdgn.\mathcal{A}_{\epsilon}(\Omega)=\mathcal{H}^{n-1}_{g}(\partial^{*}\Omega)-\int_{M_{\infty}\setminus\Omega}h_{\epsilon}\circ\rho\,\mathrm{d}\mathcal{H}^{n}_{g}.

It follows from [Zhu21, Proposition 2.1] or [CL20+, Proposition 12] that we can find a smooth minimizer Ωϵ\Omega^{*}_{\epsilon} of the functional 𝒜ϵ\mathcal{A}_{\epsilon} among the class 𝒞ϵ\mathcal{C}_{\epsilon}. From a direct comparison we see

gn1(Ωϵ)gn1(Ωϵ)MΩϵhϵρdgngn1(M).\mathcal{H}^{n-1}_{g}(\partial\Omega^{*}_{\epsilon})\leq\mathcal{H}^{n-1}_{g}(\partial\Omega^{*}_{\epsilon})-\int_{M_{\infty}\setminus\Omega^{*}_{\epsilon}}h_{\epsilon}\circ\rho\,\mathrm{d}\mathcal{H}^{n}_{g}\leq\mathcal{H}^{n-1}_{g}(\partial M_{\infty}).

The first variation formula of 𝒜ϵ\mathcal{A}_{\epsilon} yields that the mean curvature of Ωϵ\partial\Omega^{*}_{\epsilon} as the boundary of Ωϵ\Omega^{*}_{\epsilon} with respect to the unit outer normal ν\nu^{*} is

HΩϵ=(hϵρ)|Ωϵ.H_{\partial\Omega^{*}_{\epsilon}}=-(h_{\epsilon}\circ\rho)|_{\partial\Omega^{*}_{\epsilon}}.

From the second variation formula of 𝒜ϵ\mathcal{A}_{\epsilon} we have

Ωϵ|ϕ|212(R(g)R(g|Ωϵ)+|A|2+H22ν(hϵρ))ϕ2dn1g0\begin{split}\int_{\partial\Omega^{*}_{\epsilon}}|\nabla\phi|^{2}-\frac{1}{2}&\Big{(}R(g)-R(g|_{\partial\Omega^{*}_{\epsilon}})+|A|^{2}\\ &\qquad\qquad+H^{2}-2\partial_{\nu^{*}}(h_{\epsilon}\circ\rho)\Big{)}\phi^{2}\,\mathrm{d}\mathcal{H}^{n-1}_{g}\geq 0\end{split} (2.2)

for every ϕ\phi in C0(Ωϵ)C_{0}^{\infty}(\partial\Omega^{*}_{\epsilon}). Using the facts R(g)0R(g)\geq 0 and

|A|2+H2nn1(hϵρ)2|A|^{2}+H^{2}\geq\frac{n}{n-1}(h_{\epsilon}\circ\rho)^{2}

as well as ν(hϵρ)|dhϵ|ρ\partial_{\nu^{*}}(h_{\epsilon}\circ\rho)\leq|\mathrm{d}h_{\epsilon}|\circ\rho, we can write (2.2) as

Ωϵ|ϕ|2+12R(g|Ω)ϕ2dgn1Ωϵ((nn1hϵ2+2hϵ)ρ)|Ωϵϕ2dgn1.\begin{split}\int_{\partial\Omega^{*}_{\epsilon}}|\nabla\phi|^{2}+&\frac{1}{2}R(g|_{\partial\Omega^{*}})\phi^{2}\,\mathrm{d}\mathcal{H}^{n-1}_{g}\\ &\geq\int_{\partial\Omega^{*}_{\epsilon}}\left.\left(\left(\frac{n}{n-1}h_{\epsilon}^{2}+2h_{\epsilon}^{\prime}\right)\circ\rho\right)\right|_{\partial\Omega^{*}_{\epsilon}}\phi^{2}\,\mathrm{d}\mathcal{H}^{n-1}_{g}.\end{split} (2.3)

We claim that there is at least one component Σϵ\Sigma^{*}_{\epsilon} of Ωϵ\partial\Omega_{\epsilon}^{*} having non-empty intersection with the compact subset Ko:=ρ1([0,12n])K_{o}:=\rho^{-1}([0,\frac{1}{2n}]). Otherwise, Ωϵ\Omega^{*}_{\epsilon} has to be disjoint with KoK_{o} and in particular we can take VV to be the component of MΩϵM_{\infty}\setminus\Omega^{*}_{\epsilon} containing KoK_{o}. Let Σϵ\Sigma^{*}_{\epsilon} be some common boundary component of Ωϵ\Omega^{*}_{\epsilon} and VV. It is easy to construct a ray γ:([0,+),0)(M,M)\gamma:\big{(}[0,+\infty),0\big{)}\to(M_{\infty},\partial M_{\infty}) intersecting only once with Σϵ\Sigma^{*}_{\epsilon}, which implies that Σϵ\Sigma^{*}_{\epsilon} represents a non-trivial \mathbb{Q}-homology class. From our assumption Σϵ\Sigma^{*}_{\epsilon} cannot admit any metric with positive scalar curvature. On the other hand, since Σϵ\Sigma^{*}_{\epsilon} as part of Ωϵ\partial\Omega^{*}_{\epsilon} is disjoint from KoK_{o}, we conclude

Σϵ|ϕ|2+12R(g|Ω)ϕ2dgn1>0 for any non-zero ϕC0(Σ^ϵ).\int_{\Sigma^{*}_{\epsilon}}|\nabla\phi|^{2}+\frac{1}{2}R(g|_{\partial\Omega^{*}})\phi^{2}\,\mathrm{d}\mathcal{H}^{n-1}_{g}>0\mbox{ for any non-zero }\phi\in C_{0}^{\infty}(\hat{\Sigma}^{*}_{\epsilon}).

In particular, the first eigenvalue of the conformal Laplacian of Σϵ\Sigma^{*}_{\epsilon} is positive and we can construct a conformal metric of Σϵ\Sigma^{*}_{\epsilon} with positive scalar curvature, which leads to a contradiction.

Now we analyze the limiting behavior of Ωϵ\Omega^{*}_{\epsilon} as ϵ0\epsilon\to 0. Notice that the functional 𝒜ϵ\mathcal{A}_{\epsilon} converges to the area functional 𝒜\mathcal{A} as ϵ0\epsilon\to 0. From geometric measure theory up to a subsequence Ωϵ\Omega^{*}_{\epsilon} converges to a (possibly empty) Caccioppoli set Ω0\Omega^{*}_{0} whose boundary is locally area-minimizing. On the other hand, all the hypersurfaces Σϵ\Sigma^{*}_{\epsilon} intersect with a fixed compact subset KoK_{o} and they have a uniform area bound

gn1(Σϵ)gn1(Ωϵ)Hgn1(M).\mathcal{H}^{n-1}_{g}(\Sigma^{*}_{\epsilon})\leq\mathcal{H}^{n-1}_{g}(\partial\Omega^{*}_{\epsilon})\leq H^{n-1}_{g}(\partial M_{\infty}).

Fixed a point qϵq^{*}_{\epsilon} in ΣϵKo\Sigma^{*}_{\epsilon}\cap K_{o}, the curvature estimate [ZZ20, Theorem 3.6] yields that the pointed hypersurface (Σϵ,qϵ)(\Sigma^{*}_{\epsilon},q^{*}_{\epsilon}) converges smoothly to a pointed minimal hypersurface (Σ0,q0)(\Sigma^{*}_{0},q^{*}_{0}) up to a subsequence, which is part of the boundary Ω0\partial\Omega^{*}_{0}. It follows from [Zhu23b, Proposition 3.2] (with the original condition having non-zero degree to TnT^{n} replaced by non-existence of positive scalar curvature) that the limit hypersurface Σ0\Sigma^{*}_{0} must have vanishing Ricci curvature. With the uniform area bound passing to the limit we see gn1(Σ0)gn1(M)\mathcal{H}^{n-1}_{g}(\Sigma^{*}_{0})\leq\mathcal{H}^{n-1}_{g}(\partial M_{\infty}). Combined with the Calabi-Yau theorem (see Theorem 2.1 above) we conclude that the limit hypersurface Σ^0\hat{\Sigma}^{*}_{0} must be compact. As a consequence, Σ0\Sigma^{*}_{0} represents a non-zero \mathbb{Q}-homology class and it is area-minimizing as a boundary component of Ω0\Omega^{*}_{0}. Now it follows from the foliation argument as in [Zhu20, Proposition 3.4] that (M,g)(M_{\infty},g) splits into the Riemannian product (M,g|M)×[0,+)(\partial M_{\infty},g|_{\partial M_{\infty}})\times[0,+\infty). ∎

Now we are ready to prove the main theorem.

Proof of Theorem 1.1.

The theorem follows immediately from the Calabi-Yau theorem in dimension two, and so we only need to deal with the case when 3n43\leq n\leq 4. In these cases we are going to deduce some contradiction by assuming that (M,g)(M,g) is non-compact but has sublinear volume growth.

First let us find the bounded region to use Lemma 1.2. Recall that MM has nonnegative scalar curvature outside a compact subset K0K_{0}. Without loss of generality we can assume K0K_{0} to be connected. Since MM is aspherical at infinity, we are able to find smooth compact sets Kn2K1K0K_{n-2}\supset\cdots\supset K_{1}\supset K_{0} such that

(πi(MKni)πi(MKni1))=0 for all 2in1.(\pi_{i}(M\setminus K_{n-i})\to\pi_{i}(M\setminus K_{n-i-1}))=0\mbox{ for all }2\leq i\leq n-1.

In the same way we just assume all KiK_{i} to be connected.

By our assumption in the beginning, (M,g)(M,g) has sublinear volume growth. Applying Lemma 1.2 to the compact set Kn2K_{n-2} we can find a larger bounded region ΩKn2\Omega\supset K_{n-2} with weakly mean-concave boundary Ω\partial\Omega with respect to the outer unit normal. By adding bounded components of MΩM\setminus\Omega and then passing to the component of UU containing Kn2K_{n-2}, we can further assume UU to be connected.

In order to apply Proposition B.2 we set Xi=MKniX_{i}=M\setminus K_{n-i} for 3in3\leq i\leq n and set X2=MΩ¯X_{2}=M\setminus\overline{\Omega}. Now we need to verify the conditions listed in Proposition B.2. From our construction it is clear that πi(Xi)πi(Xi+1)\pi_{i}(X_{i})\to\pi_{i}(X_{i+1}) is the zero map for all 2in12\leq i\leq n-1, so it remains to show the injectivity of Hn1(X2,)Hn1(Xn,)H_{n-1}(X_{2},\mathbb{Q})\to H_{n-1}(X_{n},\mathbb{Q}) and we consider the exact sequence

Hn(Xn,X2,)Hn1(X2,)Hn1(Xn,).H_{n}(X_{n},X_{2},\mathbb{Q})\to H_{n-1}(X_{2},\mathbb{Q})\to H_{n-1}(X_{n},\mathbb{Q}).

It suffices to show Hn(Xn,X2,)=0H_{n}(X_{n},X_{2},\mathbb{Q})=0. To see this we notice that U¯K0\overline{U}\setminus K_{0} consists of non-compact manifolds with boundary, which satisfies the well-known fact Hn(U¯K0,U)=0H_{n}(\overline{U}\setminus K_{0},\partial U)=0. From the excision we have

Hn(Xn,X2)=Hn(MK0,MU¯)=Hn(U¯K0,U)=0.H_{n}(X_{n},X_{2})=H_{n}(M\setminus K_{0},M\setminus\overline{U})=H_{n}(\overline{U}\setminus K_{0},\partial U)=0.

In particular, we have Hn(Xn,X2,)=0H_{n}(X_{n},X_{2},\mathbb{Q})=0.

Now denote MM_{\infty} to be MΩM\setminus\Omega and it follows from Proposition B.2 that any embedded hypersurface representing a non-zero \mathbb{Q}-homology class cannot admit any metric with positive scalar curvature. Then it follows from Proposition 1.3 that (M,g)(M_{\infty},g) splits into the Riemannian product

(M,g|M)×[0,+),(\partial M_{\infty},g|_{\partial M_{\infty}})\times[0,+\infty),

which leads to a contradiction to the fact that (M,g)(M,g) has sublinear volume growth. ∎

Finally let us prove Theorem 1.7.

Proof of Theorem 1.7.

To show that MM is homeomorphic to 3\mathbb{R}^{3} it suffices to prove that MM is simply-connected at infinity (see [Sta72] for instance). That is, for any compact subset KK we can find a larger compact subset K~\tilde{K} such that the inclusion map i:π1(MK~)π1(MK)i_{*}:\pi_{1}(M\setminus\tilde{K})\to\pi_{1}(M\setminus K) is the zero map. Let us argue by contradiction and suppose that there is a compact subset KK such that any disk bounded by a loop γ\gamma outside KK intersects KK.

The strategy is to show that we can find arbitrarily large bounded region which has spherical boundary. From the contractibility of MM we see that MM is non-compact. Since (M,g)(M,g) has sublinear volume growth, from Lemma 1.2 we can find a bounded smooth region Ω\Omega containing KK whose boundary is weakly mean-concave. Denote E1,E2,,ElE_{1},\,E_{2},\ldots,E_{l} to be the unbounded components of MΩM\setminus\Omega. Set the μ\mu-bubble problem on each EiE_{i} in the same way as in the proof of Proposition 1.3. Since (Ei,g)(E_{i},g) has positive scalar curvature, for ϵ\epsilon small enough we can find a smooth region Ei,E_{i,\infty} from the μ\mu-bubble problem such that EiEi,E_{i}\setminus E_{i,\infty} is bounded and that all Ei,\partial E_{i,\infty} satisfies

Ei,|ϕ|2+12R(g|Ei,)ϕ2dg2>0\int_{\partial E_{i,\infty}}|\nabla\phi|^{2}+\frac{1}{2}R(g|_{\partial E_{i,\infty}})\phi^{2}\,\mathrm{d}\mathcal{H}^{2}_{g}>0

for any non-zero ϕC0(Σ^ϵ)\phi\in C_{0}^{\infty}(\hat{\Sigma}^{*}_{\epsilon}). Taking the test function ϕ1\phi\equiv 1 on each component of Ei,\partial E_{i,\infty} and using the Gauss-Bonnet formula we conclude that Ei,\partial E_{i,\infty} consists of 22-spheres. Now we take K~\tilde{K} to be MiEi,M\setminus\cup_{i}E_{i,\infty}.

We claim that any loop γ\gamma in MK~M\setminus\tilde{K} can shrink to a point in MKM\setminus K. Recall that MM is contractible. So the loop γ\gamma bounds a disk DD in MM. After perturbation we may assume that DD is transversal to K~\partial\tilde{K}. Let us take the component Σ\Sigma of DK~D\setminus\tilde{K} containing γ\gamma. Then Σ\Sigma is simply the disk DD with finitely many disjoint sub-disks DjD_{j} removed. Since Dj\partial D_{j} is contained in the spherical boundary K~\partial\tilde{K}, we can find disks DjK~D_{j}^{*}\subset\partial\tilde{K} with Dj=Dj\partial D_{j}^{*}=\partial D_{j}. As a consequence, the set

Σ(jDj)\Sigma\cup\left(\bigcup_{j}D_{j}^{*}\right)

provides a disk outside KK with boundary γ\gamma, which leads to a contradiction to our assumption in the beginning. ∎

Appendix A Finite-volume complete manifolds with nonnegative scalar curvature

Lemma A.1.

There is a complete metric gg on n\mathbb{R}^{n}, n3n\geq 3, with nonnegative scalar curvature such that (M,g)(M,g) has finite volume.

Proof.

Let us consider a conformally flat metric g=u4n2geucg=u^{\frac{4}{n-2}}g_{euc} with uu a smooth positive function on n\mathbb{R}^{n} to be determined. To ensure (n,g)(\mathbb{R}^{n},g) having nonnegative scalar curvature we just need to guarantee Δu0\Delta u\leq 0 concerning the formula

Δu=cnR(g)un+2n2.-\Delta u=c_{n}R(g)u^{\frac{n+2}{n-2}}.

The desired function uu is constructed as follows. We start with a function

v=(rlnr)n22.v=\left(r\ln r\right)^{-\frac{n-2}{2}}.

A straight-forward computation gives

Δv=n22(rlnr)n+22(n22ln2rn2).\Delta v=-\frac{n-2}{2}\left(r\ln r\right)^{-\frac{n+2}{2}}\left(\frac{n-2}{2}\ln^{2}r-\frac{n}{2}\right).

In particular, there is an absolute constant r0r_{0} such that Δv<0\Delta v<0 when rr0r\geq r_{0}. Denote v0=v(r0)v_{0}=v(r_{0}). To do composition we have to construct a function ζ:[0,+)[0,v0/2]\zeta:[0,+\infty)\to[0,v_{0}/2] satisfying

  • ζ(t)t\zeta(t)\equiv t in a neighborhood of 0 and ζ(t)const.\zeta(t)\equiv const. when tv0/2t\geq v_{0}/2;

  • ζ(t)0\zeta^{\prime}(t)\geq 0 and ζ′′(t)0\zeta^{\prime\prime}(t)\leq 0 for all t0t\geq 0.

Such function can be constructed from integration. Take a nonnegative monotone-decreasing function η:[0,+)[0,1]\eta:[0,+\infty)\to[0,1] such that η1\eta\equiv 1 in [0,v0/4][0,v_{0}/4] and η0\eta\equiv 0 in [v0/2,+)[v_{0}/2,+\infty). It suffices to define

ζ(t)=0tη(s)ds.\zeta(t)=\int_{0}^{t}\eta(s)\,\mathrm{d}s.

Let u=ζvu=\zeta\circ v. Note that uu is defined on the whole n\mathbb{R}^{n} since it is constant in the r0r_{0}-ball. It is direct to compute

Δu=ζ′′|v|2+ζΔv.\Delta u=\zeta^{\prime\prime}|\nabla v|^{2}+\zeta^{\prime}\Delta v.

When rr0r\geq r_{0} it follows from Δv<0\Delta v<0 and the construction of ζ\zeta that Δu0\Delta u\leq 0. When rr0r\leq r_{0} we simply have Δu0\Delta u\equiv 0 due to its constancy.

It remains to verify the completeness and the finite volume of (n,g)(\mathbb{R}^{n},g). To see the completeness we compute

dist(O,)=0+u2n2drr0+1rlnrdr=lnlnr|r0+=+.\begin{split}\operatorname{dist}(O,\infty)&=\int_{0}^{+\infty}u^{\frac{2}{n-2}}\,\mathrm{d}r\\ &\geq\int_{r_{0}}^{+\infty}\frac{1}{r\ln r}\,\mathrm{d}r=\left.\ln\ln r\right|_{r_{0}}^{+\infty}=+\infty.\end{split}

On the other hand, the volume can be computed as

vol(n,g)=nu2nn2dxωnr0n(v02)2nn2+nωnr0+1rlnnrdr<+.\begin{split}\operatorname{vol}(\mathbb{R}^{n},g)&=\int_{\mathbb{R}^{n}}u^{\frac{2n}{n-2}}\,\mathrm{d}x\\ &\leq\omega_{n}r_{0}^{n}\left(\frac{v_{0}}{2}\right)^{\frac{2n}{n-2}}+n\omega_{n}\int_{r_{0}}^{+\infty}\frac{1}{r\ln^{n}r}\,\mathrm{d}r\\ &<+\infty.\end{split}

This completes the proof. ∎

Appendix B \mathbb{Q}-homology vanishing theorem

The \mathbb{Q}-homology vanishing conjecture was raised by Gromov [Gro23, page 96] as following

Conjecture B.1.

Let MnM^{n} be a closed manifold admitting positive scalar curvature. For any continuous map f:MXf:M\to X mapping MM into an aspherical topological space XX, we have f([M])=0f_{*}([M])=0 in Hn(X,)H_{n}(X,\mathbb{Q}).

Based on the work [CL20, LM23] the author and his collaborator [HZ23+] proved for 3n53\leq n\leq 5 that if (M,g)(M,g) is a closed (n1)(n-1)-manifold with positive scalar curvature and XX is an aspherical manifold, then for any continuous map f:MXf:M\to X we have f([M])=0Hn1(X,)f_{*}([M])=0\in H_{n-1}(X,\mathbb{Q}). In this paper, we need to use the following variant.

Proposition B.2.

Let n=3n=3 or 44, and XnX^{n} be an nn-manifold associated with a finite open exhaustion X2X3Xn=XX_{2}\subset X_{3}\subset\cdots\subset X_{n}=X satisfying

  • Hn1(X2,)Hn1(X,)H_{n-1}(X_{2},\mathbb{Q})\to H_{n-1}(X,\mathbb{Q}) is injective;

  • πi(Xi)πi(Xi+1)\pi_{i}(X_{i})\to\pi_{i}(X_{i+1}) is the zero map for all 2in12\leq i\leq n-1.

Assume that (Mn1,g)(M^{n-1},g) is a closed manifold with positive scalar curvature. Then for any continuous map f:MX2f:M\to X_{2} we must have f([M])=0Hn1(X2,)f_{*}([M])=0\in H_{n-1}(X_{2},\mathbb{Q}).

Proof.

When n=3n=3, MM can only be a topological 22-sphere or the projective space P2\mathbb{R}P^{2}. Since π2(X2)π2(X)\pi_{2}(X_{2})\to\pi_{2}(X) is the zero map, we obtain f([M])=0H2(M,)f_{*}([M])=0\in H_{2}(M,\mathbb{Q}) in both cases and the conclusion comes from the injectivity of the map H2(X2,)H2(X,)H_{2}(X_{2},\mathbb{Q})\to H_{2}(X,\mathbb{Q}). When n=4n=4, the argument is similar but slightly more complicated. By lifting we may assume MM to be orientable and so it follows from the classification result of orientable closed 33-manifolds with positive scalar curvature that MM is diffeomorphic to the connected sum

(𝕊3/Γ1)##(𝕊3/Γk)#l(𝕊2×𝕊1).(\mathbb{S}^{3}/\Gamma_{1})\#\cdots\#(\mathbb{S}^{3}/\Gamma_{k})\#l(\mathbb{S}^{2}\times\mathbb{S}^{1}).

Passing to some finite cover M~\tilde{M} we see that M~\tilde{M} is a connected sum of finitely many 𝕊2×𝕊1\mathbb{S}^{2}\times\mathbb{S}^{1}s. Since π2(X2)π2(X3)\pi_{2}(X_{2})\to\pi_{2}(X_{3}) is the zero map, we can break M~\tilde{M} into spherical 33-cycles in X3X_{3}. From the zero map π3(X3)π3(X)\pi_{3}(X_{3})\to\pi_{3}(X) we conclude f~([M~])=0H3(X,)\tilde{f}_{*}([\tilde{M}])=0\in H_{3}(X,\mathbb{Q}), where f~\tilde{f} is denoted to be the composition of the map ff and the covering map M~M\tilde{M}\to M. Once again, from the injectivity of the map H3(X2,)H3(X,)H_{3}(X_{2},\mathbb{Q})\to H_{3}(X,\mathbb{Q}) we obtain f~([M~])=0H3(X2,)\tilde{f}_{*}([\tilde{M}])=0\in H_{3}(X_{2},\mathbb{Q}) and so f([M])=0H3(X2,)f_{*}([M])=0\in H_{3}(X_{2},\mathbb{Q}). ∎

References

  • [Cal75] Eugenio Calabi. On manifolds with nonnegative Ricci curvature II. Notices of the American Mathematical Society, vol. 22, 1975.
  • [CL20] Gregory Chambers and Yevgeny Liokumovich. Existence of minimal hypersurfaces in complete manifolds of finite volume. Invent. Math. 219 (2020), no. 1, 179–217.
  • [CL20+] Otis Chodosh and Chao Li. Generalized soap bubbles and the topology of manifolds with positive scalar curvature. preprint, arXiv: 2008.11888v4, to appear in Ann. of Math. (2)
  • [CL23] Otis Chodosh and Chao Li. Stable anisotropic minimal hypersurfaces in 𝐑4\mathbf{R}^{4}. Forum Math. Pi 11 (2023), Paper No. e3, 22 pp.
  • [CLMS24+] Otis Chodosh, Chao Li, Paul Minter, and Douglas Stryker. Stable minimal hypersurfaces in 𝐑5\mathbf{R}^{5}. preprint, arXiv:2401.01492
  • [CRZ23] Simone Cecchini, Daniel Räde, and Rudolf Zeidler. Nonnegative scalar curvature on manifolds with at least two ends. J. Topol. 16 (2023), no. 3, 855–876.
  • [CK92] Christopher B. Croke and Bruce Kleiner. A warped product splitting theorem. Duke Math. J. 67 (1992), no.3, 571–574.
  • [Giu77] Enrico Giusti. Minimal surfaces and functions of bounded variation. Notes on Pure Math., 10, Australian National University, Department of Pure Mathematics, Canberra, 1977, xi+185 pp. ISBN: 0-7081-1294-3
  • [Gro20+] Misha Gromov. No metrics with positive scalar curvature on aspherical 5-manifolds. preprint, arXiv: 2009.05332
  • [Gro23] Misha Gromov. Four lectures on scalar curvature. World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, 2023, 1–514. ISBN: 978-981-124-998-3; 978-981-124-935-8; 978-981-124-936-5
  • [GL83] Misha Gromov and H. Blaine Jr. Lawson. Positive scalar curvature and the Dirac operator on complete Riemannian manifolds. Inst. Hautes Études Sci. Publ. Math. (1983), no. 58, 83–196.
  • [HZ23+] Shihang He and Jintian Zhu. A note on rational homology vanishing theorem for hypersurfaces in aspherical manifolds. preprint, arXiv: 2311.14008
  • [Leo83] Leon Simon. Lectures on geometric measure theory. Proc. Centre Math. Anal. Austral. Nat. Univ., 3 Australian National University, Centre for Mathematical Analysis, Canberra, 1983, vii+272 pp.
  • [LUY21+] Martin Lesourd, Ryan Unger, and Shing-Tung Yau. The positive mass theorem with arbitrary ends. preprint, arXiv: 2103.02744
  • [LM23] Y. Liokumovich and D. Maximo, Waist inequality for 3-manifolds with positive scalar curvature. Perspectives in scalar curvature. Vol. 2, 799–831. World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, [2023], ⓒ 2023.
  • [Song23] Antoine Song. A dichotomy for minimal hypersurfaces in manifolds thick at infinity. Ann. Sci. Éc. Norm. Supér. (4) 56 (2023), no. 4, 1085–1134.
  • [Sta72] John Stallings. Group theory and three dimensional manifolds. Yale University Press, 1972.
  • [Wang19+] Jian Wang. Contractible 3-manifolds and positive scalar curvature (I). preprint, arXiv: 1901.04605, to appear in J. Differential Geom.
  • [Wang23+] Jian Wang. Contractible 3-manifolds and positive scalar curvature (II). preprint, arXiv: 1906.04128v3, to appear in JEMS
  • [Yau76] Shing-Tung Yau. Some function-theoretic properties of complete Riemannian manifold and their applications to geometry. Indiana Univ. Math. J. 25 (1976), no. 7, 659–670.
  • [ZZ20] Xin Zhou and Jonathan Zhu. Existence of hypersurfaces with prescribed mean curvature I—generic min-max. Camb. J. Math. 8 (2020), no.2, 311–362.
  • [Zhu20] Jintian Zhu. Rigidity of area-minimizing 2-spheres in n-manifolds with positive scalar curvature. Proc. Amer. Math. Soc. 148 (2020), no. 8, 3479–3489.
  • [Zhu21] Jintian Zhu. Width estimate and doubly warped product. Trans. Amer. Math. Soc. 374 (2021), no. 2, 1497–1511.
  • [Zhu23a] Jintian Zhu. Positive mass theorem with arbitrary ends and its application. Int. Math. Res. Not. IMRN (2023), no. 11, 9880–9900.
  • [Zhu23b] Jintian Zhu. Rigidity results for complete manifolds with nonnegative scalar curvature. J. Differential Geom. 125 (3): 623–644
  • [Zhu23+] Jintian Zhu. Riemannian Penrose inequality without horizon in dimension three. preprint, arXiv:2304.01769, to appear in Trans. Amer. Math. Soc.