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

Scalar curvature rigidity of domains in a 3-dimensional warped product

Xiaoxiang Chai Department of Mathematics, POSTECH, Pohang, Gyeongbuk, South Korea xxchai@kias.re.kr, xxchai@postech.ac.kr  and  Gaoming Wang Yau Mathematical Sciences Center, Tsinghua University, Beijing, 100084, China gmwang@tsinghua.edu.cn
Abstract.

A warped product with a spherical factor and a logarithmically concave warping function satisfies a scalar curvature rigidity of the Llarull type. We develop a scalar curvature rigidity of the Llarull type for a general class of domains in a three dimensional spherical warped product. In the presence of rotational symmetry, we identify this class of domains as those satisfying a boundary condition analogous to the logarithmic concavity of the warping function.

Key words and phrases:
Scalar curvature, mean curvature, stable capillary surface, prescribed mean curvature, rigidity, foliation.
1991 Mathematics Subject Classification:
53C24, 49Q20

1. Introduction

Llarull proved a scalar curvature rigidity theorem for the standard nn-spheres.

Theorem 1.1 ([Lla98]).

Let gg be a smooth metric on the n-sphere with the metric comparison gg¯g\geqslant\bar{g} and the scalar curvature comparison Rgn(n1)R_{g}\geqslant n(n-1). Then g=g¯g=\bar{g}.

An interesting feature of this scalar curvature rigidity for spheres comparing to that of torus [SY79a], the Euclidean space [SY79b] and the hyperbolic space [MO89] is the requirement of a metric comparison gg¯g\geqslant\bar{g}. A counterexample without the requirement gg¯g\geqslant\bar{g} was given in [BM11] (for half nn-sphere fixing the geometry of the boundary), but gg¯g\geqslant\bar{g} can be weakened, see [Lis10]. Llarull also showed that the condition gg¯g\geqslant\bar{g} can be formulated more generally as the existence of a distance non-increasing map F:(M,g)(𝕊n,g¯)F:(M,g)\to(\mathbb{S}^{n},\bar{g}) of non-zero degree.

Recently, there were efforts in extending Llarull’s theorem to a spherical warped product

(1.1) (M¯n,g¯):=([t,t+]×Sn1,dt2+ψ(t)2g𝕊n1) with t<t+, (logψ)′′<0,(\bar{M}^{n},\bar{g}):=([t_{-},t_{+}]\times S^{n-1},\mathrm{d}t^{2}+\psi(t)^{2}g_{\mathbb{S}^{n-1}})\text{ with }t_{-}<t_{+},\text{ }(\log\psi)^{\prime\prime}<0,

by spinors [CZ24], [BBHW24], [WX23b], by μ\mu-bubbles [Gro21], [HLS23] and by spacetime harmonic functions [HKKZ].

We are interested in the Llarull type theorems of domains in the spherical warped product (1.2). Although the form (1.2) can also be considered as a domain in a larger spherical warped product, our focus will be on such domains with boundaries that are not necessarily given by tt-level sets. Previously, this direction has been explored by Lott [Lot21], Wang-Xie [WX23a] and Chai-Wan [CW24], which all involved spinors. Gromov first suggested the use of stable capillary minimal surface in studying the scalar curvature rigidity of Euclidean balls (see Section 5.8.1 of [Gro21]; Spin-Extremality of Doubly Punctured Balls) and Li [Li20] in three-dimensional Euclidean dihedral rigidity. In this article, we make use of the stable capillary surfaces with prescribed (varying) contact angle and prescribed mean curvature, or in the terminology of Gromov [Gro21], (part of) the boundary of a stable capillary μ\mu-bubble. This is also a further development of the previous work [CW23] and the recent work of Ko-Yao [KY24] in the Euclidean case.

We consider three dimensions, fix a metric gS2g_{S^{2}} of positive Gauss curvature on the 22-sphere S2S^{2}, and the following

(1.2) (M¯3,g¯):=([t,t+]×S2,dt2+ψ(t)2gS2) with t<t+, (logψ)′′<0.(\bar{M}^{3},\bar{g}):=([t_{-},t_{+}]\times S^{2},\mathrm{d}t^{2}+\psi(t)^{2}g_{{S}^{2}})\text{ with }t_{-}<t_{+},\text{ }(\log\psi)^{\prime\prime}<0.

We reserve g𝕊2g_{\mathbb{S}^{2}} for the standard round metric on the 2-sphere S2S^{2}. We use txt_{x} to indicate the tt coordinate and pxp_{x} to denote the S2S^{2} coordinate of a point xM[t,t+]×S2x\in M\subset[t_{-},t_{+}]\times S^{2}. Let K(p)K(p) be the Gauss curvature at p(S2,gS2)p\in(S^{2},g_{S^{2}}).

Let P±={t±}×S2P_{\pm}=\{t_{\pm}\}\times{S}^{2}, we always assume that MM lies between P±P_{\pm} such that P±MP_{\pm}\cap\partial M is non-empty. If P±MP_{\pm}\cap\partial M contains only a point, we set this point to be p±p_{\pm}. Let sM=M\(P+P)\partial_{s}M=\partial M\backslash(P_{+}\cup P_{-}) and X¯\bar{X} be the unit outward normal of sM\partial_{s}M in MM with respect to g¯\bar{g}. We fix

(1.3) h¯(t)=2ψ(t)/ψ(t)\bar{h}(t)=2\psi^{\prime}(t)/\psi(t)

and γ¯\bar{\gamma} to be the dihedral angles formed by sM\partial_{s}M and Σt=({t}×S2)M\Sigma_{t}=(\{t\}\times{S}^{2})\cap M that is given by

(1.4) cosγ¯=g¯(X¯,t).\cos\bar{\gamma}=\bar{g}(\bar{X},\partial_{t}).

We call the tt-suplevel set Ωt+\Omega_{t}^{+} and tt-sublevel set Ωt\Omega_{t}^{-}. We also need to fix some more conventions for the direction of the unit normal, the sign of the mean curvatures and the dihedral angles. Let Σ\Sigma be a surface with boundary on sM\partial_{s}M and separates P+MP_{+}\cap\partial M and PMP_{-}\cap\partial M, we always fix the direction of the unit normal NN of Σ\Sigma to be the direction which points inside of the region bounded by Σ,\Sigma, P+MP_{+}\cap\partial M and sM\partial_{s}M. The mean curvature is then the trace of the second fundamental form N\nabla N. We fix γΣ\gamma_{\Sigma} to be the contact angle formed by Σ\Sigma and sM\partial_{s}M, that is, cosγΣ=X,N\cos\gamma_{\Sigma}=\langle X,N\rangle. For the mean curvature of sM\partial_{s}M, it is always computed with respect to the outward unit normal. The geometric quantity on (M,g¯)(M,\bar{g}) comes with a bar unless otherwise specified (see Figure 2.1).

As is well known, the warped product metric (1.2) is conformal to a direct product metric. Indeed, let s=t1ψ(τ)dτs=\int^{t}\tfrac{1}{\psi(\tau)}\mathrm{d}\tau, then ds=1ψ(t)dt\mathrm{d}s=\tfrac{1}{\psi(t)}\mathrm{d}t and

(1.5) dt2+ψ(t)2gS2=ψ(t)2ds2+ψ(t)2gS2=ψ(t)2(ds2+gS2)\mathrm{d}t^{2}+\psi(t)^{2}g_{{S}^{2}}=\psi(t)^{2}\mathrm{d}s^{2}+\psi(t)^{2}g_{{S}^{2}}=\psi(t)^{2}(\mathrm{d}s^{2}+g_{{S}^{2}})

where t=t(s)t=t(s) is implicitly given by s=t1ψ(τ)dτs=\int^{t}\tfrac{1}{\psi(\tau)}\mathrm{d}\tau.

Now we state our scalar curvature rigidity result. (For a quick feel, we refer to Theorem 1.4.)

Theorem 1.2.

Assume that g¯\bar{g} is a Riemannian metric on M¯3\bar{M}^{3} as in (1.2). Assume that MM is a domain in M¯\bar{M} such that sM\partial_{s}M is convex with respect to the conformally related metric ds2+gS2\mathrm{d}s^{2}+g_{S^{2}} where ss is given in (1.5) and gg is a smooth metric on MM. If gg satisfies the metric comparison

(1.6) gg¯,g\geqslant\bar{g},

and the scalar curvature comparison

(1.7) RgRg¯,R_{g}\geqslant R_{\bar{g}},

and the mean curvature comparison

(1.8) HsMH¯sMH_{\partial_{s}M}\geqslant\bar{H}_{\partial_{s}M}

along sM\partial_{s}M, and further that P±MP_{\pm}\cap M and ψ(t)\psi(t) satisfy either of the following conditions:

  1. (1)

    ψ(t±)>0\psi(t_{\pm})>0, P±MP_{\pm}\cap\partial M only contains a point where M\partial M is smooth under both metrics gg and g¯\bar{g};

  2. (2)

    ψ(t±)>0\psi(t_{\pm})>0, P±MP_{\pm}\cap\partial M only contains a point where M\partial M where the tangent cone of (M,g)(M,g) at p±p_{\pm} is Euclidean, and the tangent cone of (M,g¯)(M,\bar{g}) at p±p_{\pm} is a round (solid) circular cone and that (|tt±|1Σt,|tt±|2g¯)(|t-t_{\pm}|^{-1}\Sigma_{t},|t-t_{\pm}|^{-2}\bar{g}) converges to its axial section as tt±t\to t_{\pm};

  3. (3)

    ψ(t)=a|tt±|+o(|tt±|2)\psi(t)=a|t-t_{\pm}|+o(|t-t_{\pm}|^{2}) as tt±t\to t_{\pm} where a>0a>0 is a constant such that the tangent cone of (M,g¯)(M,\bar{g}) at p±p_{\pm} is of non-negative Ricci curvature and (M,g)(M,g) admits a metric tangent cone at p±p_{\pm};

  4. (4)

    ψ(t)>0\psi(t)>0 on [t,t+][t_{-},t_{+}], P±MP_{\pm}\cap\partial M is a disk, the mean curvatures ±HP±M±h¯P±M=±h¯(t±)\pm H_{P_{\pm}\cap\partial M}\geqslant\pm\bar{h}_{P_{\pm}\cap\partial M}=\pm\bar{h}(t_{\pm}) and the dihedral angles ±γP+M±γ¯|P±M\pm\gamma_{P_{+}\cap\partial M}\geqslant\pm\bar{\gamma}|_{P_{\pm}\cap\partial M} formed by sM\partial_{s}M and P±MP_{\pm}\cap\partial M,

then g=g¯g=\bar{g}.

Remark 1.3.

The mean curvature comparisons can be reformulated as HMH¯MH_{\partial M}\geqslant\bar{H}_{\partial M} on M\partial M if all mean curvatures are computed with respect to the outward unit normal. For convergence of sequences of Riemannian manifolds and notions of tangent cones at a point of a Riemannian manifold, we refer to [BBI01, Chapter 8]. We believe that the extra condition on the tangent cone of (M,g¯)(M,\bar{g}) at p±p_{\pm} can be dropped in item (2).

We use a special case of (M,g¯)(M,\bar{g}) to illustrate the convexity of sM\partial_{s}M with respect to the metric ds2+gS2\mathrm{d}s^{2}+g_{S^{2}} given in (1.5). Assume that gS2g_{S^{2}} is just the standard round metric, that is,

(1.9) gS2=g𝕊2=dr2+sin2rdθ2, r[0,π], θ𝕊1,g_{S^{2}}=g_{\mathbb{S}^{2}}=\mathrm{d}r^{2}+\sin^{2}r\mathrm{d}\theta^{2},\text{ }r\in[0,\pi],\text{ }\theta\in\mathbb{S}^{1},

using the polar coordinates and MM is given by

M={(t,r,θ): t[t,t+], rρ(t)<π2, θ𝕊1}M=\{(t,r,\theta):\text{ }t\in[t_{-},t_{+}],\text{ }r\leqslant\rho(t)<\tfrac{\pi}{2},\text{ }\theta\in\mathbb{S}^{1}\}

for some positive function ρ(t)\rho(t) on [t,t+][t_{-},t_{+}]. In this case, the prescribed contact angle γ¯\bar{\gamma} only depends on tt. It is easy to check that the convexity of sM\partial_{s}M with respect to ds2+g𝕊2\mathrm{d}s^{2}+g_{\mathbb{S}^{2}} is equivalent to dγ¯ds<0\tfrac{\mathrm{d}\bar{\gamma}}{\mathrm{d}s}<0. Also, we find that

(1.10) dγ¯dt<0\tfrac{\mathrm{d}\bar{\gamma}}{\mathrm{d}t}<0

using (1.5) or that the angles are conformally invariant. The mean curvature of a tt-level set is given by

h¯(t):=2ψ1dψdt=2d(logψ)dt.\bar{h}(t):=2\psi^{-1}\tfrac{\mathrm{d}\psi}{\mathrm{d}t}=2\tfrac{\mathrm{d}(\log\psi)}{\mathrm{d}t}.

The logarithmic concavity of ψ\psi is equivalent to the more geometric statement that the mean curvature of the tt-level set is monotonically decreasing as tt increases. The condition dγ¯dt<0\tfrac{\mathrm{d}\bar{\gamma}}{\mathrm{d}t}<0 can be viewed as a boundary analog of the logarithmic concavity. Geometrically, (1.10) says that the dihedral angles (1.4) formed by Σt\Sigma_{t} and sM\partial_{s}M monotonically decreases along the t\partial_{t} direction with respect to the metric g¯\bar{g}. This condition answers a question raised by Gromov at the end of [Gro21, Section 5.8.1]. See also Chai-Wan [CW24, Theorem 1.1].

We can have various scalar curvature rigidity assuming different combinations of geometric structures at P±P_{\pm}. The following is a simple corollary of Theorem 1.2 where both P±MP_{\pm}\cap\partial M take the structure as specified in item (1) of Theorem 1.2.

Theorem 1.4.

Let MM be a smooth domain in [t,t+]×S2[t_{-},t_{+}]\times{S}^{2} with the metric dt2+ψ(t)2gS2\mathrm{d}t^{2}+\psi(t)^{2}g_{S^{2}} where (logψ)′′<0(\log\psi)^{\prime\prime}<0, ψ(t)>0\psi(t)>0 on [t,t+][t_{-},t_{+}]. Assume that M\partial M is convex with respect to the conformally related metric ds2+gS2\mathrm{d}s^{2}+g_{S^{2}} where ss is given in (1.5) and gg is another metric on MM which satisfies the comparisons of:

  1. (1)

    the scalar curvatures RgRg¯R_{g}\geqslant R_{\bar{g}},

  2. (2)

    the mean curvatures HgHg¯H_{g}\geqslant H_{\bar{g}} of the boundary M\partial M in MM,

  3. (3)

    and the metrics gg¯g\geqslant\bar{g},

then g=g¯g=\bar{g}.

Our approach toward items (1)-(3) is by construction of surfaces of prescribed mean curvature and prescribed contact angles near t=tt=t_{-} which serves as barriers, a concept which now introduce. The existence of barriers enables us to find a stable capillary μ\mu-bubble, if fact, P±MP_{\pm}\cap\partial M are natural barriers if we are in item (4) of Theorem 1.2.

Definition 1.5.

We say that a surface Σ+\Sigma_{+} (Σ\Sigma_{-}) whose boundary separates (P+M)\partial(P_{+}\cap\partial M) and (PM)\partial(P_{-}\cap\partial M) is an upper (lower) barrier if HΣ+h¯|Σ+H_{\Sigma_{+}}\geqslant\bar{h}|_{\Sigma_{+}} (HΣh¯|ΣH_{\Sigma_{-}}\leqslant\bar{h}|_{\Sigma_{-}}) and  γΣ+γ¯|Σ+M\gamma_{\Sigma_{+}}\geqslant\bar{\gamma}|_{\partial\Sigma_{+}\cap\partial M} (γΣγ¯|ΣM\gamma_{\Sigma_{-}}\leqslant\bar{\gamma}|_{\partial\Sigma_{-}\cap\partial M}) along Σ+\partial\Sigma_{+} (Σ\partial\Sigma_{-}). We call Σ+\Sigma_{+} and Σ\Sigma_{-} are a set of barriers if Σ+\Sigma_{+} and Σ\Sigma_{-} are respectively an upper barrier and a lower barrier and Σ+\Sigma_{+} is closer to P+P_{+} than Σ\Sigma_{-}.

Our construction of barriers near t=t±t=t_{\pm} are purely local, and in this way, our proof of items (1)-(3) can be reduced to the last item of Theorem 1.2, that is, the case with a set of barriers. In particular, if both P±MP_{\pm}\cap\partial M take the structure as specified in item (3) of Theorem 1.2, we obtain a genuine generalization of Theorem 1.1, since in the case of round metric, ψ(t)=sint\psi(t)=\sin t, t[0,π]t\in[0,\pi] is allowed to take zero values at t=0t=0 and t=πt=\pi. Hu-Liu-Shi [HLS23] (see also Gromov [Gro21]) used a μ\mu-bubble approach for Theorem 1.1. As indicated earlier, we use the capillary version of the μ\mu-bubble approach. However, our method differs from theirs in a technical manner when handling

ψ(t)=sint=t+o(|t|)\psi(t)=\sin t=t+o(|t|)

near t=0t=0. (Similarly, near t=πt=\pi.) They constructed a family of perturbations on the function 2ψ/ψ2\psi^{\prime}/\psi while our strategy is to perform a careful tangent cone analysis near t=0t=0 or t=πt=\pi. As a result of this new strategy, we are able to generalize the Llarull Theorem 1.1 to the case where the background metric g¯\bar{g} are equipped with antipodal conical points.

Theorem 1.6.

Let n=3n=3 and (M¯,g¯)(\bar{M},\bar{g}) be a three dimensional warped product given in (1.1) such that

ψ(t±)=a±|tt±|+o(|tt±|), 0<a±1,\psi(t_{\pm})=a_{\pm}|t-t_{\pm}|+o(|t-t_{\pm}|),\text{ }0<a_{\pm}\leqslant 1,

If gg is another smooth metric on M¯\bar{M} with possible cone singularity at only t=t±t=t_{\pm} which satisfies the comparisons of metrics gg¯g\geqslant\bar{g} and scalar curvatures RgRg¯R_{g}\geqslant R_{\bar{g}}, then g=g¯g=\bar{g}.

Theorem 1.6 directly follows from the proof of item (3) of Theorem 1.2 with only slight changes and we omit its proof. See Remark 3.13. Note that the condition 0<a±10<a_{\pm}\leqslant 1 ensures that the Ricci curvature of the tangent cone with respect to g¯\bar{g} at t=t±t=t_{\pm} is non-negative. The scalar curvature rigidity of the Llarull type for a±>1a_{\pm}>1 is an interesting question. One could also compare Theorem 1.6 with [CLZ24] where conical singularities with respect to the metric gg are allowed at multiple points on SnS^{n}.

Some of the essential difficulties of items (1) and (2) are already present in [CW23, Theorem 1.2 (2) and (3)]. In light of this, we only give a sketch for their proof in Section 4 , and refer relevant details to [CW23].

It is possible that the inequalities in (logψ)′′<0(\log\psi)^{\prime\prime}<0 and the convexity of sM\partial_{s}M can be weakened in some cases. For instance, we can consider a convex disk MM in (,)×S2(-\infty,\infty)\times S^{2} with the direct product metric dt2+gS2\mathrm{d}t^{2}+g_{S^{2}}. In this case, logψ\log\psi vanishes. The Llarull type rigidity Theorem 1.4 is still valid for this MM.

Now one could naturally ask what are other shapes of point singularities, in particular, asymptotics of ψ\psi near t=t±t=t_{\pm}, such that Theorem 1.2 remain valid. However, it is a quite intricate matter to which we do not have an answer at the moment. It is also desirable to find a proof for higher dimensional analogs of our results using the stable capillary μ\mu-bubbles. This seems to be a promising direction to investigate being aware of the recent works [CWXZ24, WWZ24].

The article is organized as follows:

In Section 2, we introduce basics of stable capillary μ\mu-bubble and we use it to show item (4) of Theorem 1.2.

In Section 3, we use the tangent cone analysis at t=tt=t_{-} to construct barriers and reduce item (3) of Theorem 1.2 to item (4).

In Section 4, we revisit our constructions in [CW23] and use the techniques developed there to show items (2) and (1) of Theorem 1.2.

Acknowledgments X. Chai was supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIT) (No. RS-2024-00337418) and an NRF grant No. 2022R1C1C1013511. G. Wang was supported by the China Postdoctoral Science Foundation (No. 2024M751604).

2. Stable capillary μ\mu-bubble

In this section, we introduce the functional (2.1) whose minimiser is a stable capillary μ\mu-bubble. We introduce the barrier condition which combining with a maximum principle ensures the existence of a regular minimiser to (2.1). By a rigidity analysis on the second variation of (2.1), we conclude the proof of item (4) of Theorem 1.2.

2.1. Notations

We set up some notations. Let EME\subset M be be a set such that EM\partial E\cap M is a regular surface with boundary which we name it Σ\Sigma. We set

  • NN, unit normal vector of Σ\Sigma pointing inside EE;

  • ν\nu, unit normal vector of Σ\partial\Sigma in Σ\Sigma pointing outside of Σ\Sigma;

  • η\eta, unit normal vector of Σ\partial\Sigma in M\partial M pointing outside of EM\partial E\cap\partial M;

  • XX: unit normal vectors of M\partial M in MM pointing outside of MM;

  • γ\gamma: the contact angle formed by Σ\Sigma and M\partial M and the magnitude of the angle is given by cosγ=X,N\cos\gamma=\langle X,N\rangle,

  • Y,Z=g(Y,Z)\langle Y,Z\rangle=g(Y,Z), the inner product of vectors YY and ZZ with respect to the metric gg;

  • Y,Zg¯=g¯(Y,Z)\langle Y,Z\rangle_{\bar{g}}=\bar{g}(Y,Z), the inner product of vectors YY and ZZ with respect to the metric g¯\bar{g}.

See Figure 2.1. We use a bar on every quantity to denote that the quantity is computed with respect to the metric g¯\bar{g} given in (1.2).

Refer to caption
γ¯\bar{\gamma}
Σ\Sigma
NN
t=t+t=t_{+}
Refer to caption
XX
Refer to caption
ν\nu
Refer to caption
η\eta
Refer to caption
MM
sM\partial_{s}M
Figure 2.1. Notations.

2.2. Functional and first variation

We fix h¯=2ψ/ψ\bar{h}=2\psi^{\prime}/\psi and γ¯\bar{\gamma} to be given by (1.4). We define the functional

(2.1) I(E)=2(EintM)Eh¯EMcosγ¯,I(E)=\mathcal{H}^{2}(\partial^{\ast}E\cap\operatorname{int}M)-\int_{E}\bar{h}-\int_{\partial^{\ast}E\cap\partial M}\cos\bar{\gamma},

where E\partial^{\ast}E denotes the reduced boundary of EE and the variational problem

(2.2) =inf{I(E): E},\mathcal{I}=\inf\{I(E):\text{ }E\in\mathcal{E}\},

where \mathcal{E} is the collection of contractible open subsets EE^{\prime} such that P+EP_{+}\subset E^{\prime}. Let Σ\Sigma be a surface with boundary Σ\partial\Sigma such that Σ\partial\Sigma separates P±P_{\pm}. Then Σ\Sigma separates MM into two components and the component closer to P+P_{+} is just EE. We reformulate the functional (2.1) in terms of Σ\Sigma. We define

(2.3) F(Σ)=I(E)=|Σ|Eh¯EMcosγ¯.F(\Sigma)=I(E)=|\Sigma|-\int_{E}\bar{h}-\int_{\partial E\cap\partial M}\cos\bar{\gamma}.

Let ϕt\phi_{t} be a family of immersions ϕt:ΣM\phi_{t}:\Sigma\to M such that ϕt(Σ)M\phi_{t}(\partial\Sigma)\subset\partial M and ϕ0(Σ)=Σ\phi_{0}(\Sigma)=\Sigma. Let Σt=ϕt(Σ)\Sigma_{t}=\phi_{t}(\Sigma) and EtE_{t} be the corresponding component separated by Σt\Sigma_{t}. Let YY be the vector field ϕtt\tfrac{\partial\phi_{t}}{\partial t}. Define 𝒜(t)=F(Σt)\mathcal{A}(t)=F(\Sigma_{t}) and f=Y,Nf=\langle Y,N\rangle, then by the first variation

(2.4) 𝒜(0)=Σf(Hh¯)+ΣY,νηcosγ¯.\mathcal{A}^{\prime}(0)=\int_{\Sigma}f(H-\bar{h})+\int_{\partial\Sigma}\langle Y,\nu-\eta\cos\bar{\gamma}\rangle.

We know that if Σ\Sigma is regular, then it is of mean curvature h¯\bar{h} and meets M\partial M at a prescribed angle γ¯\bar{\gamma}. And EE is called a capillary μ\mu-bubble. The second variation at such Σ\Sigma is

(2.5) 𝒜′′(0)=Q(f,f):=Σ(fΔf+(|A|2+Ric(N)+Nh¯)f2)+Σf(fνqf).\mathcal{A}^{\prime\prime}(0)=Q(f,f):=-\int_{\Sigma}(f\Delta f+(|A|^{2}+\operatorname{Ric}(N)+\partial_{N}\bar{h})f^{2})+\int_{\partial\Sigma}f(\tfrac{\partial f}{\partial\nu}-qf).

where fC(Σ)f\in C^{\infty}(\Sigma) and

(2.6) q:=1sinγ¯AM(η,η)cotγ¯A(ν,ν)+1sin2γ¯ηcosγ¯.q:=\tfrac{1}{\sin\bar{\gamma}}A_{\partial M}(\eta,\eta)-\cot\bar{\gamma}A(\nu,\nu)+\tfrac{1}{\sin^{2}\bar{\gamma}}\partial_{\eta}\cos\bar{\gamma}.

We define two operators

(2.7) L=Δ(|A|2+Ric(N)+Nh¯) in Σ,L=-\Delta-(|A|^{2}+\operatorname{Ric}(N)+\partial_{N}\bar{h})\text{ in }\Sigma,

and

B=νq on Σ.B=\tfrac{\partial}{\partial\nu}-q\text{ on }\partial\Sigma.

The surface Σ\Sigma is called stable if

(2.8) Q(f,f)0Q(f,f)\geqslant 0

for all fC(M)f\in C^{\infty}(M). The second variation (2.5) is closely related to the variation of Hh¯H-\bar{h} and cosγcosγ¯\cos\gamma-\cos\bar{\gamma}. Indeed, let f=Y,Nf=\langle Y,N\rangle, we have that the first variation of Hh¯H-\bar{h} is

(2.9) Y(Hh¯)\displaystyle\nabla_{Y}(H-\bar{h}) =Lf+Y(Hh¯)\displaystyle=Lf+\nabla_{Y^{\top}}(H-\bar{h})
(2.10) =Δf(|A|2+Ric(N)+Nh¯)f+Y(Hh¯).\displaystyle=-\Delta f-(|A|^{2}+\operatorname{Ric}(N)+\partial_{N}\bar{h})f+\nabla_{Y^{\top}}(H-\bar{h}).

And the first variation of the angle difference X,Ncosγ¯\langle X,N\rangle-\cos\bar{\gamma} is

(2.11) Y(cosγcosγ¯)=\displaystyle\nabla_{Y}(\cos\gamma-\cos\bar{\gamma})= sinγ¯fν\displaystyle-\sin\bar{\gamma}\tfrac{\partial f}{\partial\nu}
(2.12) +(AM(η,η)\displaystyle+(A_{\partial M}(\eta,\eta) cosγ¯A(ν,ν)+1sinγ¯ηcosγ¯)f+Y(X,Ncosγ¯).\displaystyle-\cos\bar{\gamma}A(\nu,\nu)+\tfrac{1}{\sin\bar{\gamma}}\partial_{\eta}\cos\bar{\gamma})f+\nabla_{Y^{\top}}(\langle X,N\rangle-\cos\bar{\gamma}).

For Σ\Sigma, Schoen-Yau [SY79b] rewrote the term |A|2+Ric(N)|A|^{2}+\operatorname{Ric}(N) as

(2.13) |A|2+Ric(N)=12(Rg2K+|A|2+H2)|A|^{2}+\operatorname{Ric}(N)=\tfrac{1}{2}(R_{g}-2K+|A|^{2}+H^{2})

where KK is the Gauss curvature of Σ\Sigma. Along the boundary Σ\partial\Sigma, we have the rewrite (see [RS97, Lemma 3.1] or [Li20, (4.13)])

(2.14) 1sinγ¯AM(η,η)cosγ¯A(ν,ν)=Hcotγ¯+HMsinγ¯κ\tfrac{1}{\sin\bar{\gamma}}A_{\partial M}(\eta,\eta)-\cos\bar{\gamma}A(\nu,\nu)=-H\cot\bar{\gamma}+\tfrac{H_{\partial M}}{\sin\bar{\gamma}}-\kappa

where κ\kappa is the geodesic curvature of Σ\partial\Sigma in Σ\Sigma.

2.3. Analysis of stability

Starting from now on, we assume that Σ\Sigma is a regular stable capillary μ\mu-bubble in (M,g)(M,g) which satisfies the assumptions of item (4) of Theorem 1.2.

Lemma 2.1.

Let Σ\Sigma be a regular stable capillary μ\mu-bubble, then Σ\Sigma is a tt-level set.

Proof.

First, we note that the second variation 𝒜′′(0)0\mathcal{A}^{\prime\prime}(0)\geqslant 0 as in (2.5). First, using Schoen-Yau’s rewrite (2.13) we see that

(2.15) |A|2+Ric(N)+Nh¯\displaystyle|A|^{2}+\operatorname{Ric}(N)+\partial_{N}\bar{h}
(2.16) =\displaystyle= 12(R2K+|A|2+H2)+Nh¯\displaystyle\tfrac{1}{2}(R-2K+|A|^{2}+H^{2})+\partial_{N}\bar{h}
(2.17) =\displaystyle= 12(R2K+|A0|2+H22+H2)+Nh¯\displaystyle\tfrac{1}{2}(R-2K+|A^{0}|^{2}+\tfrac{H^{2}}{2}+H^{2})+\partial_{N}\bar{h}
(2.18) =\displaystyle= 12(R+32h¯2+2Nh¯)K+12|A0|2,\displaystyle\tfrac{1}{2}(R+\tfrac{3}{2}\bar{h}^{2}+2\partial_{N}\bar{h})-K+\tfrac{1}{2}|A^{0}|^{2},

where A0A^{0} is the traceless part of the second fundamental form. Similarly using (2.14), we see

q=Hcotγ¯+HMsinγ¯κ+1sin2γ¯ηcosγ¯.q=-H\cot\bar{\gamma}+\tfrac{H_{\partial M}}{\sin\bar{\gamma}}-\kappa+\tfrac{1}{\sin^{2}\bar{\gamma}}\partial_{\eta}\cos\bar{\gamma}.

We obtain by letting f1f\equiv 1 in the (2.8) (also using (2.5) and (2.6)),

(2.19) 2πχ(Σ)\displaystyle 2\pi\chi(\Sigma) =ΣK+Σκ\displaystyle=\int_{\Sigma}K+\int_{\partial\Sigma}\kappa
(2.20) Σ[12(R+32h¯2+2Nh¯)+12|A0|2]+Σ(HMsinγ¯h¯cotγ¯+1sin2γ¯ηcosγ¯)\displaystyle\geqslant\int_{\Sigma}\left[\tfrac{1}{2}(R+\tfrac{3}{2}\bar{h}^{2}+2\partial_{N}\bar{h})+\tfrac{1}{2}|A^{0}|^{2}\right]+\int_{\partial\Sigma}\left(\tfrac{H_{\partial M}}{\sin\bar{\gamma}}-\bar{h}\cot\bar{\gamma}+\tfrac{1}{\sin^{2}\bar{\gamma}}\partial_{\eta}\cos\bar{\gamma}\right)
(2.21) Σ12(R+32h¯2+2Nh¯)+Σ(HMsinγ¯h¯cotγ¯+1sin2γ¯ηcosγ¯)\displaystyle\geqslant\int_{\Sigma}\tfrac{1}{2}\left(R+\tfrac{3}{2}\bar{h}^{2}+2\partial_{N}\bar{h}\right)+\int_{\partial\Sigma}\left(\tfrac{H_{\partial M}}{\sin\bar{\gamma}}-\bar{h}\cot\bar{\gamma}+\tfrac{1}{\sin^{2}\bar{\gamma}}\partial_{\eta}\cos\bar{\gamma}\right)
(2.22) Σ12(Rg¯+32h¯2+2Nh¯)+Σ(H¯Msinγ¯h¯cotγ¯+1sin2γ¯ηcosγ¯),\displaystyle\geqslant\int_{\Sigma}\tfrac{1}{2}\left(R_{\bar{g}}+\tfrac{3}{2}\bar{h}^{2}+2\partial_{N}\bar{h}\right)+\int_{\partial\Sigma}\left(\tfrac{\bar{H}_{\partial M}}{\sin\bar{\gamma}}-\bar{h}\cot\bar{\gamma}+\tfrac{1}{\sin^{2}\bar{\gamma}}\partial_{\eta}\cos\bar{\gamma}\right),

where in the last line we have incorporated the comparisons RgRg¯R_{g}\geqslant R_{\bar{g}} in MM and HMH¯MH_{\partial M}\geqslant\bar{H}_{\partial M} on M\partial M.

Now we estimate Rg¯+32h¯2+2Nh¯R_{\bar{g}}+\tfrac{3}{2}\bar{h}^{2}+2\partial_{N}\bar{h}. We have that

Nh¯=g¯(N,g¯h¯)|N|g¯|g¯h¯|g¯=|N|g¯h¯,\partial_{N}\bar{h}=\bar{g}(N,\nabla^{\bar{g}}\bar{h})\geqslant-|N|_{\bar{g}}|\nabla^{\bar{g}}\bar{h}|_{\bar{g}}=|N|_{\bar{g}}\bar{h}^{\prime},

since gg¯g\geqslant\bar{g}, so

1=|N|g|N|g¯,1=|N|_{g}\geqslant|N|_{\bar{g}},

and we get

Nh¯h¯.\partial_{N}\bar{h}\geqslant\bar{h}^{\prime}.

So

Rg¯+32h¯2+2Nh¯Rg¯+32h¯2+2h¯.R_{\bar{g}}+\tfrac{3}{2}\bar{h}^{2}+2\partial_{N}\bar{h}\geqslant R_{\bar{g}}+\tfrac{3}{2}\bar{h}^{2}+2\bar{h}^{\prime}.

For any point xΣx\in\Sigma, the right hide side is just 2K(px)ψ2(tx)\tfrac{2K(p_{x})}{\psi^{2}(t_{x})}. Recall that x=(tx,px)x=(t_{x},p_{x}) is the coordinate of xΣ¯tx\in\bar{\Sigma}_{t}. This is by a direct calculation of the scalar curvature of the warped product metric (1.2). So

(2.23) Rg¯+32h¯2+2Nh¯2K(px)ψ2(tx).R_{\bar{g}}+\tfrac{3}{2}\bar{h}^{2}+2\partial_{N}\bar{h}\geqslant\tfrac{2K(p_{x})}{\psi^{2}(t_{x})}.

Let g^=ds2+gS2\hat{g}=\mathrm{d}s^{2}+g_{{S}^{2}} and it is conformally related to g¯\bar{g} via (1.5). Let X^\hat{X} be the unit outward normal of sM\partial_{s}M in MM and H^sM\hat{H}_{\partial_{s}M} be the mean curvature of sM\partial_{s}M in MM with respect to g^\hat{g}. Since g¯\bar{g} is conformal to g^\hat{g}, by a well known formula of conformal change of mean curvature,

(2.24) H¯sM=1φ(s)(H^sM+2X^logφ).\bar{H}_{\partial_{s}M}=\tfrac{1}{\varphi(s)}(\hat{H}_{\partial_{s}M}+2\partial_{\hat{X}}\log\varphi).

Similarly, the mean curvature h¯\bar{h} of Σt\Sigma_{t} in MM is

(2.25) h¯(t)=2φ(s)2φ(s).\bar{h}(t)=\tfrac{2}{\varphi(s)^{2}}\varphi^{\prime}(s).

Hence, by (2.24), (2.25), (1.5) and that g^(s,X^)=cosγ¯\hat{g}(\partial_{s},\hat{X})=\cos\bar{\gamma},

H¯Msinγ¯h¯cotγ¯+1sin2γ¯ηcosγ¯=1ψ(tx)sinγ¯(H^sMψηγ¯).\tfrac{\bar{H}_{\partial M}}{\sin\bar{\gamma}}-\bar{h}\cot\bar{\gamma}+\tfrac{1}{\sin^{2}\bar{\gamma}}\partial_{\eta}\cos\bar{\gamma}=\tfrac{1}{\psi(t_{x})\sin\bar{\gamma}}(\hat{H}_{\partial_{s}M}-\partial_{\psi\eta}\bar{\gamma}).

Inserting (2.23) and the above in (2.22) yields

2πχ(Σ)ΣK(px)ψ(tx)2dσ+Σ1ψ(tx)sinγ¯(H^sMψηγ¯)dλ,2\pi\chi(\Sigma)\geqslant\int_{\Sigma}\tfrac{K(p_{x})}{\psi(t_{x})^{2}}\mathrm{d}\sigma+\int_{\partial\Sigma}\tfrac{1}{\psi(t_{x})\sin\bar{\gamma}}(\hat{H}_{\partial_{s}M}-\partial_{\psi\eta}\bar{\gamma})\mathrm{d}\lambda,

where we have written the area element dσ\mathrm{d}\sigma and line length element dλ\mathrm{d}\lambda explicitly in the metric gg. The rest of the proof is deferred to the next Lemma 2.2. ∎

Lemma 2.2.

Assume that gg¯g\geqslant\bar{g}. If Σ\Sigma is a surface in MM whose boundary Σ\partial\Sigma is a simple smooth curve that separates (P+M)\partial(P_{+}\cap\partial M) and (PM)\partial(P_{-}\cap\partial M), then

(2.26) ΣK(px)ψ(tx)2dσ+Σ1ψ(tx)sinγ¯(H^sMψηγ¯)dλ2π,\int_{\Sigma}\tfrac{K(p_{x})}{\psi(t_{x})^{2}}\mathrm{d}\sigma+\int_{\partial\Sigma}\tfrac{1}{\psi(t_{x})\sin\bar{\gamma}}(\hat{H}_{\partial_{s}M}-\partial_{\psi\eta}\bar{\gamma})\mathrm{d}\lambda\geqslant 2\pi,

where equality occurs if and only if Σ\Sigma is a tt-level set.

Proof.

It suffices to prove (2.26) for ψ1\psi\equiv 1 since gg¯=ψ2g^g\geqslant\bar{g}=\psi^{2}\hat{g}. In this case, s=ts=t, we just use tt. We also suppress the subscript sM\partial_{s}M for clarity in this proof. In addition, we assume that every tt-coordinate of Σ\Sigma is strictly less than t+t_{+}, since we can increase t+t_{+} a little. Let e1e_{1} be the unit tangent vector of Σt\partial\Sigma_{t} with respect to g^\hat{g} and e2e_{2} be the unit outward normal of Σt\partial\Sigma_{t} in sM\partial_{s}M with respect to g^\hat{g}. Let TT (resp. T^\hat{T}) be the unit tangent vector of Σ\partial\Sigma with respect to gg (resp. g^\hat{g}). We recall an ingenious inequality of [KY24, Lemma 3.2],

(2.27) H^ηγ¯T,e21γ¯+(H^2γ¯)e1,\hat{H}-\nabla_{\eta}\bar{\gamma}\geqslant\langle T,e_{2}\nabla_{1}\bar{\gamma}+(\hat{H}-\nabla_{2}\bar{\gamma})e_{1}\rangle,

where ,\langle\cdot,\cdot\rangle denotes the inner product with respect to g^\hat{g}.

Let λ\lambda be an arc-length parameter of Σ\partial\Sigma with respect to gg, then the length element of Σ\partial\Sigma with respect to g^\hat{g} is given by |T|g^dλ=:dλ^|T|_{\hat{g}}\mathrm{d}\lambda=:\mathrm{d}\hat{\lambda} and T=|T|g^T^T=|T|_{\hat{g}}\hat{T}. Therefore,

(2.28) Σ1sinγ¯(H^ηγ¯)dλ\displaystyle\int_{\partial\Sigma}\tfrac{1}{\sin\bar{\gamma}}(\hat{H}-\partial_{\eta}\bar{\gamma})\mathrm{d}\lambda
(2.29) \displaystyle\geqslant Σ1sinγ¯T,1γ¯e2+(H^2γ¯)e1dλ\displaystyle\int_{\partial\Sigma}\tfrac{1}{\sin\bar{\gamma}}\langle T,\partial_{1}\bar{\gamma}e_{2}+(\hat{H}-\partial_{2}\bar{\gamma})e_{1}\rangle\mathrm{d}\lambda
(2.30) =\displaystyle= Σ1sinγ¯T^,1γ¯e2+(H^2γ¯)e1dλ^.\displaystyle\int_{\partial\Sigma}\tfrac{1}{\sin\bar{\gamma}}\langle\hat{T},\nabla_{1}\bar{\gamma}e_{2}+(\hat{H}-\nabla_{2}\bar{\gamma})e_{1}\rangle\mathrm{d}\hat{\lambda}.

Let η^\hat{\eta} be the unit outward normal of Σ\partial\Sigma in sM\partial_{s}M with respect to g^\hat{g}. Recall the notation A^=A^sM\hat{A}=\hat{A}_{\partial_{s}M}, and we know that the components of A^\hat{A} satisfy

(2.31) 1γ¯=A^12, 2γ¯=A^22, H^2γ¯=A^11,\partial_{1}\bar{\gamma}=\hat{A}_{12},\text{ }\partial_{2}\bar{\gamma}=\hat{A}_{22},\text{ }\hat{H}-\partial_{2}\bar{\gamma}=\hat{A}_{11},

in the frame {e1,e2}\{e_{1},e_{2}\}. So

T^,1γ¯e2+(H^2γ¯)e1=T^,A^12e2+A^11e1=A^1T^.\langle\hat{T},\partial_{1}\bar{\gamma}e_{2}+(\hat{H}-\partial_{2}\bar{\gamma})e_{1}\rangle=\langle\hat{T},\hat{A}_{12}e_{2}+\hat{A}_{11}e_{1}\rangle=\hat{A}_{1\hat{T}}.

Since the orthonormal frames {T^,η^}\{\hat{T},\hat{\eta}\} and {e1,e2}\{e_{1},e_{2}\} have the same orientation, we can set T^=a1e1+a2e2\hat{T}=a_{1}e_{1}+a_{2}e_{2} and so η^=a2e1+a1e2\hat{\eta}=-a_{2}e_{1}+a_{1}e_{2} for some a1a_{1} and a2a_{2} which satisfy a12+a22=1a_{1}^{2}+a_{2}^{2}=1. It then follows that

(2.32) A^1T^=(A^H^g^)(e2,η^)=:W^(e2,η^),\hat{A}_{1\hat{T}}=-(\hat{A}-\hat{H}\hat{g})(e_{2},\hat{\eta})=:-\hat{W}(e_{2},\hat{\eta}),

by considering also that H^=A^11+A^22\hat{H}=\hat{A}_{11}+\hat{A}_{22}. Hence

Σ1sinγ¯(H^ηγ¯)dλΣ1sinγ¯W^(e2,η^)dλ^.\int_{\partial\Sigma}\tfrac{1}{\sin\bar{\gamma}}(\hat{H}-\partial_{\eta}\bar{\gamma})\mathrm{d}\lambda\geqslant-\int_{\partial\Sigma}\tfrac{1}{\sin\bar{\gamma}}\hat{W}(e_{2},\hat{\eta})\mathrm{d}\hat{\lambda}.

Suppose that Σ(t+)\partial\Sigma(t_{+}) (we set Σ(t+±)=Σt±\Sigma(t_{+\pm})=\Sigma_{t_{\pm}} to avoid double subscripts) and Σ\partial\Sigma enclose a region SS in sM\partial_{s}M. By the divergence theorem,

(2.33) Σ1sinγ¯W^(e2,η^)dλ^\displaystyle-\int_{\partial\Sigma}\tfrac{1}{\sin\bar{\gamma}}\hat{W}(e_{2},\hat{\eta})\mathrm{d}\hat{\lambda}
(2.34) =\displaystyle= S^iS(W^ij1sinγ¯e2,ej)dσ^Σ(t+)1sinγ¯(A^H^g^)(e2,e2)dλ^\displaystyle-\int_{S}\hat{\nabla}^{S}_{i}\left(\hat{W}_{ij}\langle\tfrac{1}{\sin\bar{\gamma}}e_{2},e_{j}\rangle\right)\mathrm{d}\hat{\sigma}-\int_{\partial\Sigma(t_{+})}\tfrac{1}{\sin\bar{\gamma}}(\hat{A}-\hat{H}\hat{g})(e_{2},e_{2})\mathrm{d}\hat{\lambda}
(2.35) =:\displaystyle=: S^iS(A^ijH^g^ij)1sinγ¯e2,ejdσ^SW^ij^iS(1sinγ¯e2),ejdσ^I3\displaystyle-\int_{S}\hat{\nabla}^{S}_{i}(\hat{A}_{ij}-\hat{H}\hat{g}_{ij})\langle\tfrac{1}{\sin\bar{\gamma}}e_{2},e_{j}\rangle\mathrm{d}\hat{\sigma}-\int_{S}\hat{W}_{ij}\langle\hat{\nabla}_{i}^{S}(\tfrac{1}{\sin\bar{\gamma}}e_{2}),e_{j}\rangle\mathrm{d}\hat{\sigma}-I_{3}
(2.36) =:\displaystyle=: I1I2I3,\displaystyle-I_{1}-I_{2}-I_{3},

where ^S\hat{\nabla}^{S} is the induced connection on SS with respect to the metric g^\hat{g}. For I1I_{1}, we use Gauss-Codazzi equation,

I1=SRicg^(X^,1sinγ¯e2)dσ^=SK(px)cosγ¯dσ^=SK(px)t,X^dσ^,I_{1}=-\int_{S}\operatorname{Ric}_{\hat{g}}(\hat{X},\tfrac{1}{\sin\bar{\gamma}}e_{2})\mathrm{d}\hat{\sigma}=-\int_{S}K(p_{x})\cos\bar{\gamma}\mathrm{d}\hat{\sigma}=-\int_{S}K(p_{x})\langle\partial_{t},\hat{X}\rangle\mathrm{d}\hat{\sigma},

where the fact that g^=dt2+gS2\hat{g}=\mathrm{d}t^{2}+g_{S^{2}} was also used in determining the Ricci curvature. For I3I_{3}, we use (2.32), we see

I3=Σ(t+)1sinγ¯A^11dλ^=Σ(t+)κ^(px,t)dλ^,I_{3}=\int_{\partial\Sigma(t_{+})}\tfrac{1}{\sin\bar{\gamma}}\hat{A}_{11}\mathrm{d}\hat{\lambda}=\int_{\partial\Sigma(t_{+})}\hat{\kappa}(p_{x},t)\mathrm{d}\hat{\lambda},

where κ^(p,t)\hat{\kappa}(p,t) is the geodesic curvature of Σt\partial\Sigma_{t} in Σt\Sigma_{t} at (p,t)Σt(p,t)\in\partial\Sigma_{t}. It seems tricky to calculate ^i(1sinγ¯e2)\hat{\nabla}_{i}\left(\tfrac{1}{\sin\bar{\gamma}}e_{2}\right) in I2I_{2} directly at a first sight, but we can convert to terms that are easier. By the identity,

1sinγ¯e2=tcosγ¯sinγ¯ν^=tcosγ¯sin2γ¯X^+cos2γ¯sin2γ¯t\tfrac{1}{\sin\bar{\gamma}}e_{2}=\partial_{t}-\tfrac{\cos\bar{\gamma}}{\sin\bar{\gamma}}\hat{\nu}=\partial_{t}-\tfrac{\cos\bar{\gamma}}{\sin^{2}\bar{\gamma}}\hat{X}+\tfrac{\cos^{2}\bar{\gamma}}{\sin^{2}\bar{\gamma}}\partial_{t}

at xsMx\in\partial_{s}M, we see

(2.37) W^ij^i(1sinγ¯e2),ej\displaystyle\hat{W}_{ij}\langle\hat{\nabla}_{i}(\tfrac{1}{\sin\bar{\gamma}}e_{2}),e_{j}\rangle
(2.38) =\displaystyle= W^ij^i(tcosγ¯sin2γ¯X^+cos2γ¯sin2γ¯t),ej\displaystyle\hat{W}_{ij}\left\langle\hat{\nabla}_{i}\left(\partial_{t}-\tfrac{\cos\bar{\gamma}}{\sin^{2}\bar{\gamma}}\hat{X}+\tfrac{\cos^{2}\bar{\gamma}}{\sin^{2}\bar{\gamma}}\partial_{t}\right),e_{j}\right\rangle
(2.39) =\displaystyle= cosγ¯sin2γ¯W^ijA^ij+W^iji(cos2γ¯sin2γ¯)t,ej\displaystyle-\tfrac{\cos\bar{\gamma}}{\sin^{2}\bar{\gamma}}\hat{W}_{ij}\hat{A}_{ij}+\hat{W}_{ij}\partial_{i}\left(\tfrac{\cos^{2}\bar{\gamma}}{\sin^{2}\bar{\gamma}}\right)\langle\partial_{t},e_{j}\rangle
(2.40) =\displaystyle= cosγ¯sin2γ¯W^ijA^ij2W^i2cosγ¯sin3γ¯iγ¯t,e2\displaystyle-\tfrac{\cos\bar{\gamma}}{\sin^{2}\bar{\gamma}}\hat{W}_{ij}\hat{A}_{ij}-2\hat{W}_{i2}\tfrac{\cos\bar{\gamma}}{\sin^{3}\bar{\gamma}}\partial_{i}\bar{\gamma}\langle\partial_{t},e_{2}\rangle

at xx. It is now a tedious task to check from the definition of W^\hat{W}, t,e2=sinγ¯\langle\partial_{t},e_{2}\rangle=-\sin\bar{\gamma} and (2.31) that the above vanishes. Therefore, I2=0I_{2}=0 and to sum up, we have shown that

(2.41) Σ1sinγ¯(H^ηγ¯)dλSK(px)t,X^dσ^+Σ(t+)κ^(px,t)dλ^.\int_{\partial\Sigma}\tfrac{1}{\sin\bar{\gamma}}(\hat{H}-\partial_{\eta}\bar{\gamma})\mathrm{d}\lambda\geqslant-\int_{S}K(p_{x})\langle\partial_{t},\hat{X}\rangle\mathrm{d}\hat{\sigma}+\int_{\partial\Sigma(t_{+})}\hat{\kappa}(p_{x},t)\mathrm{d}\hat{\lambda}.

Now we set the region enclosed by Σ(t+)\Sigma(t_{+}), Σ\Sigma and sM\partial_{s}M to be Ω\Omega. Let G^=Ricg^12Rg^g^\hat{G}=\operatorname{Ric}_{\hat{g}}-\frac{1}{2}R_{\hat{g}}\hat{g}. Using the divergence free property of G^\hat{G} (twice-contracted Gauss-Codazzi equation), t\partial_{t}, and the divergence theorem,

0=Ω^i(G^ij(t)j)=ΩG^(t,X^)=12ΩRg^t,X^,=ΩK(px)t,X^,0=\int_{\Omega}\hat{\nabla}_{i}(\hat{G}_{ij}(\partial_{t})_{j})=\int_{\partial\Omega}\hat{G}(\partial_{t},\hat{X})=-\frac{1}{2}\int_{\partial\Omega}R_{\hat{g}}\langle\partial_{t},\hat{X}\rangle,=-\int_{\partial\Omega}K(p_{x})\langle\partial_{t},\hat{X}\rangle,

where X^\hat{X} now also denotes the unit outward normal of Ω\partial\Omega with respect to g^\hat{g} and (t)j(\partial_{t})_{j} denotes the jj-th component of the vector field t\partial_{t}. Note that Ω=SΣ(t+)Σ\partial\Omega=S\cup\Sigma(t_{+})\cup\Sigma and so

0=Σ(t+)K(px)t,tΣK(px)t,N^+SK(px)t,X^,0=\int_{\Sigma(t_{+})}K(p_{x})\langle\partial_{t},\partial_{t}\rangle-\int_{\Sigma}K(p_{x})\langle\partial_{t},\hat{N}\rangle+\int_{S}K(p_{x})\langle\partial_{t},\hat{X}\rangle,

and it follows from that K(px)>0K(p_{x})>0 that

(2.42) ΣK(px)ΣK(px)t,N^=Σ(t+)K(px)+SK(px)t,X^.\int_{\Sigma}K(p_{x})\geqslant\int_{\Sigma}K(p_{x})\langle\partial_{t},\hat{N}\rangle=\int_{\Sigma(t_{+})}K(p_{x})+\int_{S}K(p_{x})\langle\partial_{t},\hat{X}\rangle.

Finally, it follows from (2.41) and (2.42) that

ΣK(px)+Σ1sinγ¯(H^ηγ¯)dλΣ(t+)K(px)+Σ(t+)κ^(px,t)dλ^.\int_{\Sigma}K(p_{x})+\int_{\partial\Sigma}\tfrac{1}{\sin\bar{\gamma}}(\hat{H}-\partial_{\eta}\bar{\gamma})\mathrm{d}\lambda\geqslant\int_{\Sigma(t_{+})}K(p_{x})+\int_{\partial\Sigma(t_{+})}\hat{\kappa}(p_{x},t)\mathrm{d}\hat{\lambda}.

An application of the Gauss-Bonnet theorem on the right hand finishes the proof of (2.26). The equality case is easy to trace. ∎

Remark 2.3.

The convexity of sM\partial_{s}M in (M,ds2+gS2)(M,\mathrm{d}s^{2}+g_{S_{2}}) with ds2+gS2\mathrm{d}s^{2}+g_{S_{2}} given in (1.5) is used in the inequality (2.27).

2.4. Infinitesimally rigid surface

The surface Σ\Sigma is a stable capillary μ\mu-bubble has more consequences than the mere Lemma 2.1. We can conclude that Σ\Sigma is a so-called infinitesimally rigid surface. See Definition 2.4.

All inequalities are in fact equalities by Lemma 2.2 and tracing the equalities in (2.22), we arrive that

(2.43) Rg=Rg¯,N=N¯,|A0|=0 in ΣR_{g}=R_{\bar{g}},N=\bar{N},|A^{0}|=0\text{ in }\Sigma

and

(2.44) HM=H¯M along Σ.H_{\partial M}=\bar{H}_{\partial M}\text{ along }\partial\Sigma.

It then follows from the equality case of Lemma 2.2 that

(2.45) tx=t0 at all xΣ¯t_{x}=t_{0}\text{ at all }x\in\bar{\Sigma}

for some constant t0[t,t+]t_{0}\in[t_{-},t_{+}]. Because Σ\Sigma is stable (equivalently Q(f,f)0Q(f,f)\geqslant 0), so the eigenvalue problem

(2.46) {Lf=μf in ΣBf=0 on Σ\left\{\begin{array}[]{ll}Lf&=\mu f\text{ in }\Sigma\\ Bf&=0\text{ on }\partial\Sigma\end{array}\right.

has a non-negative first eigenvalue μ10\mu_{1}\geqslant 0.

The analysis now is similar to [FCS80]. Letting f1f\equiv 1 in (2.8), using (2.43), (2.44) and (2.45), we get

(2.47) Q(1,1)=\displaystyle Q(1,1)= Σ[K12(R+32h¯2+2Nh¯)]\displaystyle\int_{\Sigma}\left[K-\tfrac{1}{2}(R+\tfrac{3}{2}\bar{h}^{2}+2\partial_{N}\bar{h})\right]
(2.48) +Σ[κ(HMsinγ¯h¯cotγ¯1sinγ¯γ¯η)]=0.\displaystyle\quad+\int_{\partial\Sigma}\left[\kappa-(\tfrac{H_{\partial M}}{\sin\bar{\gamma}}-\bar{h}\cot\bar{\gamma}-\tfrac{1}{\sin\bar{\gamma}}\tfrac{\partial\bar{\gamma}}{\partial\eta})\right]=0.

And so the first eigenvalue μ1\mu_{1} is zero, and the constant 1 is its corresponding eigenfunction.

By (2.43) and (2.18), the stability operator LL reduces to

L=Δ(K(px)ψ(t0)2K);L=-\Delta-\left(\tfrac{K(p_{x})}{\psi(t_{0})^{2}}-K\right);

by considering (2.44) and that tx=t0t_{x}=t_{0} , the boundary stability operator BB reduces to

B=ν(κ^(px,t0)ψ(t0)κ).B=\partial_{\nu}-\left(\frac{\hat{\kappa}(p_{x},t_{0})}{\psi(t_{0})}-\kappa\right).

Putting f=1f=1 and μ1=0\mu_{1}=0 in the eigenvalue problem (2.46), we get

(2.49) K=K(px)ψ2(t0) in Σ, κ=κ^(px,t0)ψ(t0) on Σ.K=\tfrac{K(p_{x})}{\psi^{2}(t_{0})}\text{ in }\Sigma,\text{ }\kappa=\tfrac{\hat{\kappa}(p_{x},t_{0})}{\psi(t_{0})}\text{ on }\partial\Sigma.

Now we summarize the properties of Σ\Sigma in the definition of an infinitesimally rigid surface.

Definition 2.4.

We say that Σ\Sigma is infinitesimally rigid if it satisfies (2.43), (2.44), (2.45) and (2.49).

2.5. Capillary foliation of constant Hh¯H-\bar{h}

See for instance the works [Ye91], [BBN10] and [Amb15] on constructing CMC foliations. First, we construct a foliation with prescribed angles γ¯\bar{\gamma} whose leaf is of constant Hh¯H-\bar{h}. Let ϕ(x,t)\phi(x,t) be a local flow of a vector field YY which is tangent to M\partial M and transverse to Σ\Sigma and that Y,N=1\langle Y,N\rangle=1.

In the following theorem, we only require that the scalar curvature of (M,g)(M,g) and the mean curvature of M\partial M are bounded below.

Theorem 2.5.

Suppose (M,g)(M,g) is a three manifold with boundary, if Σ\Sigma is an infinitesimally rigid surface, then there exists ε>0\varepsilon>0 and a function w(x,t)w(x,t) on Σ×(ε,ε)\Sigma\times(-\varepsilon,\varepsilon) such that for each t(ε,ε)t\in(-\varepsilon,\varepsilon), the surface

(2.50) Σt={ϕ(x,w(x,t)):xΣ}\Sigma_{t}=\{\phi(x,w(x,t)):x\in\Sigma\}

is a surface of constant Hh¯H-\bar{h} intersecting M\partial M with prescribed angle γ¯\bar{\gamma}. Moreover, for every xΣx\in\Sigma and every t(ε,ε)t\in(-\varepsilon,\varepsilon),

(2.51) w(x,0)=0, Σ(w(x,t)t)=0 and tw(x,t)|t=0=1.w(x,0)=0,\text{ }\int_{\Sigma}(w(x,t)-t)=0\text{ and }\tfrac{\partial}{\partial t}w(x,t)|_{t=0}=1.
Proof.

Given a function in the Hölder space C2,α(Σ)C1,α(Σ¯)C^{2,\alpha}(\Sigma)\cap C^{1,\alpha}(\bar{\Sigma}) (0<α<10<\alpha<1), we consider

Σu={ϕ(x,u(x)):xΣ},\Sigma_{u}=\{\phi(x,u(x)):x\in\Sigma\},

which is a properly embedded surface if the norm of uu is small enough. We use the subscript uu to denote the quantities associated with Σu\Sigma_{u}.

Consider the space

𝒴={uC2,α(Σ)C1,α(Σ¯):Σu=0}\mathcal{Y}=\left\{u\in C^{2,\alpha}(\Sigma)\cap C^{1,\alpha}(\bar{\Sigma}):\int_{\Sigma}u=0\right\}

and

𝒵={uC0,α(Σ):Σu=0}.\mathcal{Z}=\left\{u\in C^{0,\alpha}(\Sigma):\int_{\Sigma}u=0\right\}.

Given small δ>0\delta>0 and ε>0\varepsilon>0, we define the map

Φ:(ε,ε)×B(0,δ)𝒵×C0,α(Σ)\Phi:(-\varepsilon,\varepsilon)\times B(0,\delta)\to\mathcal{Z}\times C^{0,\alpha}(\partial\Sigma)

given by

(2.52) Φ(t,u)\displaystyle\Phi(t,u)
(2.53) =\displaystyle= ((Ht+uh¯t+u)1|Σ|Σ(Ht+uh¯t+u),Xt+u,Nt+ucosγ¯t+u).\displaystyle\left((H_{t+u}-\bar{h}_{t+u})-\tfrac{1}{|\Sigma|}\int_{\Sigma}(H_{t+u}-\bar{h}_{t+u}),\langle X_{t+u},N_{t+u}\rangle-\cos\bar{\gamma}_{t+u}\right).

Here, B(0,δ)B(0,\delta) is a ball of radius δ>0\delta>0 centered at the zero function in 𝒴\mathcal{Y}. For each vΣv\in\Sigma, the map

f:(x,s)Σ×(ε,ε)ϕ(x,sv(x))Mf:(x,s)\in\Sigma\times(-\varepsilon,\varepsilon)\to\phi(x,sv(x))\in M

gives a variation with

fs|s=0=sϕ(x,sv(x))|s=0=vN.\tfrac{\partial f}{\partial s}|_{s=0}=\tfrac{\partial}{\partial s}\phi(x,sv(x))|_{s=0}=vN.

Since Σ\Sigma is infinitesimally rigid and using also (2.10) and (2.12), we obtain that

DΦ(0,0)(0,v)=ddsΦ(0,sv)|s=0=(Δv+1|Σ|ΣΔv,sinγ¯vν).D\Phi_{(0,0)}(0,v)=\tfrac{\mathrm{d}}{\mathrm{d}s}\Phi(0,sv)|_{s=0}=\left(-\Delta v+\tfrac{1}{|\Sigma|}\int_{\partial\Sigma}\Delta v,-\sin\bar{\gamma}\tfrac{\partial v}{\partial\nu}\right).

It follows from the elliptic theory for the Laplace operator with Neumann type boundary conditions that DΦ(0,0)D\Phi(0,0) is an isomorphism when restricted to 0×𝒴0\times\mathcal{Y}.

Now we apply the implicit function theorem: For some smaller ε\varepsilon, there exists a function u(t)B(0,δ)𝒳u(t)\in B(0,\delta)\subset\mathcal{X}, t(ε,ε)t\in(-\varepsilon,\varepsilon) such that u(0)=0u(0)=0 and Φ(t,u(t))=Φ(0,0)=(0,0)\Phi(t,u(t))=\Phi(0,0)=(0,0) for every tt. In other words, the surfaces

Σt+u(t)={ϕ(x,t+u(t)):xΣ}\Sigma_{t+u(t)}=\{\phi(x,t+u(t)):x\in\Sigma\}

are of constant Hh¯H-\bar{h} with prescribed angles γ¯\bar{\gamma}.

Let w(x,t)=t+u(t)(x)w(x,t)=t+u(t)(x) where (x,t)Σ×(ε,ε)(x,t)\in\Sigma\times(-\varepsilon,\varepsilon). By definition, w(x,0)=0w(x,0)=0 for every xΣx\in\Sigma and w(,t)t=u(t)B(0,δ)𝒳w(\cdot,t)-t=u(t)\in B(0,\delta)\subset\mathcal{X} for every t(ε,ε)t\in(-\varepsilon,\varepsilon). Observe that the map sϕ(x,w(x,s))s\mapsto\phi(x,w(x,s)) gives a variation of Σ\Sigma with variational vector field is given by

ϕtws|s=0=ws|s=0Y.\tfrac{\partial\phi}{\partial t}\tfrac{\partial w}{\partial s}|_{s=0}=\tfrac{\partial w}{\partial s}|_{s=0}Y.

Since for every tt we have that

(2.54) 0=\displaystyle 0= Φ(t,u(t))\displaystyle\Phi(t,u(t))
(2.55) =\displaystyle= ((Hw(,t)h¯w(,t))1|Σ|Σ(Hw(,t)h¯w(,t)),Xt+u,Nt+ucosγ¯t+u),\displaystyle\left((H_{w(\cdot,t)}-\bar{h}_{w(\cdot,t)})-\tfrac{1}{|\Sigma|}\int_{\Sigma}(H_{w(\cdot,t)}-\bar{h}_{w(\cdot,t)}),\langle X_{t+u},N_{t+u}\rangle-\cos\bar{\gamma}_{t+u}\right),

by taking the derivative at t=0t=0 we conclude that

wt|t=0Y,N=wt|t=0\langle\tfrac{\partial w}{\partial t}|_{t=0}Y,N\rangle=\tfrac{\partial w}{\partial t}|_{t=0}

satisfies the homogeneous Neumann problem. Therefore, it is constant on Σ\Sigma. Since

Σ(w(x,t)t)=Σu(x,t)=0\int_{\Sigma}(w(x,t)-t)=\int_{\Sigma}u(x,t)=0

for every tt, by taking derivatives at t=0t=0 again, we conclude that

Σwt|t=0=|Σ|.\int_{\Sigma}\tfrac{\partial w}{\partial t}|_{t=0}=|\Sigma|.

Hence, wt|t=0=1\tfrac{\partial w}{\partial t}|_{t=0}=1. Taking ε\varepsilon small, we see that ϕ(x,w(x,t))\phi(x,w(x,t)) parameterize a foliation near Σ\Sigma. ∎

Theorem 2.6.

There exists a continuous function Ψ(t)\Psi(t) such that

ddt(exp(0tΨ(τ)dτ)(Hh¯))0.\tfrac{\mathrm{d}}{\mathrm{d}t}\left(\exp(-\int_{0}^{t}\Psi(\tau)\mathrm{d}\tau)(H-\bar{h})\right)\leqslant 0.
Proof.

Let ψ:Σ×IM\psi:\Sigma\times I\to M parameterize the foliation, Y=ψtY=\tfrac{\partial\psi}{\partial t}, vt=Y,Ntv_{t}=\langle Y,N_{t}\rangle. Then

(2.56) ddt(Hh¯)=Δtvt+(Ric(Nt)+|At|2)vt+vtNth¯ in Σt,-\tfrac{\mathrm{d}}{\mathrm{d}t}(H-\bar{h})=\Delta_{t}v_{t}+(\operatorname{Ric}(N_{t})+|A_{t}|^{2})v_{t}+v_{t}\nabla_{N_{t}}\bar{h}\text{ in }\Sigma_{t},

and

(2.57) vtνt=[cotγ¯At(νt,νt)+1sinγ¯AM(ηt,ηt)+1sin2γ¯ηtcosγ¯]vt.\tfrac{\partial v_{t}}{\partial\nu_{t}}=[-\cot\bar{\gamma}A_{t}(\nu_{t},\nu_{t})+\tfrac{1}{\sin\bar{\gamma}}A_{\partial M}(\eta_{t},\eta_{t})+\tfrac{1}{\sin^{2}\bar{\gamma}}\nabla_{\eta_{t}}\cos\bar{\gamma}]v_{t}.

By shrinking the interval if needed, we assume that vt>0v_{t}>0 for tIt\in I. By multiplying of (2.56) and integrate on Σt\Sigma_{t}, we deduce by integration by parts and applying the Schoen-Yau rewrite (2.13) that

(2.58) (Hh¯)Σt1vt\displaystyle-(H-\bar{h})^{\prime}\int_{\Sigma_{t}}\tfrac{1}{v_{t}}
(2.59) =\displaystyle= ΣtΔtvtvt+(Ric(Nt)+|At|2+Nth¯)\displaystyle\int_{\Sigma_{t}}\tfrac{\Delta_{t}v_{t}}{v_{t}}+(\operatorname{Ric}(N_{t})+|A_{t}|^{2}+\nabla_{N_{t}}\bar{h})
(2.60) =\displaystyle= Σt1vtvtνt+12Σt(Rg+|At|2+Ht2+2Nth¯)ΣtKΣt+Σt|vt|2vt2.\displaystyle\int_{\partial\Sigma_{t}}\tfrac{1}{v_{t}}\tfrac{\partial v_{t}}{\partial\nu_{t}}+\tfrac{1}{2}\int_{\Sigma_{t}}(R_{g}+|A_{t}|^{2}+H_{t}^{2}+2\nabla_{N_{t}}\bar{h})-\int_{\Sigma_{t}}K_{\Sigma_{t}}+\int_{\Sigma_{t}}\tfrac{|\nabla v_{t}|^{2}}{v_{t}^{2}}.

Let χ=A12h¯σ\chi=A-\tfrac{1}{2}\bar{h}\sigma, we have that

(2.61) |At|2\displaystyle|A_{t}|^{2}
(2.62) =\displaystyle= |χ+12h¯σ|2\displaystyle|\chi+\tfrac{1}{2}\bar{h}\sigma|^{2}
(2.63) =\displaystyle= |χ|2+χ,h¯σ+12h¯2,\displaystyle|\chi|^{2}+\langle\chi,\bar{h}\sigma\rangle+\tfrac{1}{2}\bar{h}^{2},
(2.64) =\displaystyle= |χ0|2+12(trσχ)2+h¯trσχ+12h¯2,\displaystyle|\chi^{0}|^{2}+\tfrac{1}{2}(\operatorname{tr}_{\sigma}\chi)^{2}+\bar{h}\operatorname{tr}_{\sigma}\chi+\tfrac{1}{2}\bar{h}^{2},

where χ0\chi^{0} is the traceless part of χ\chi. Also,

H2=(trσχ+h¯)2=(trσχ)2+2trσχh¯+h¯2.H^{2}=(\operatorname{tr}_{\sigma}\chi+\bar{h})^{2}=(\operatorname{tr}_{\sigma}\chi)^{2}+2\operatorname{tr}_{\sigma}\chi\bar{h}+\bar{h}^{2}.

So

(2.65) (Hh¯)Σt1vt\displaystyle-(H-\bar{h})^{\prime}\int_{\Sigma_{t}}\tfrac{1}{v_{t}}
(2.66) =\displaystyle= Σt1vtvtνt+12Σt(Rg+|At|2+Ht2+2Nth¯)ΣtKΣt+Σt|vt|2vt2\displaystyle\int_{\partial\Sigma_{t}}\tfrac{1}{v_{t}}\tfrac{\partial v_{t}}{\partial\nu_{t}}+\tfrac{1}{2}\int_{\Sigma_{t}}(R_{g}+|A_{t}|^{2}+H_{t}^{2}+2\nabla_{N_{t}}\bar{h})-\int_{\Sigma_{t}}K_{\Sigma_{t}}+\int_{\Sigma_{t}}\tfrac{|\nabla v_{t}|^{2}}{v_{t}^{2}}
(2.67) =\displaystyle= Σt1vtvtνt+12Σt(Rg+32h¯2+2Nth¯)\displaystyle\int_{\partial\Sigma_{t}}\tfrac{1}{v_{t}}\tfrac{\partial v_{t}}{\partial\nu_{t}}+\tfrac{1}{2}\int_{\Sigma_{t}}(R_{g}+\tfrac{3}{2}\bar{h}^{2}+2\nabla_{N_{t}}\bar{h})
(2.68) +12Σt|χ0|2+32(trσχ)2+3h¯trσχΣtKΣt+Σt|vt|2vt2\displaystyle+\tfrac{1}{2}\int_{\Sigma_{t}}|\chi^{0}|^{2}+\tfrac{3}{2}(\operatorname{tr}_{\sigma}\chi)^{2}+3\bar{h}\operatorname{tr}_{\sigma}\chi-\int_{\Sigma_{t}}K_{\Sigma_{t}}+\int_{\Sigma_{t}}\tfrac{|\nabla v_{t}|^{2}}{v_{t}^{2}}
(2.69) \displaystyle\geqslant Σt1vtvtνt+ΣtK(px)ψ2(tx)+32(Hh¯)Σth¯ΣtKΣt,\displaystyle\int_{\partial\Sigma_{t}}\tfrac{1}{v_{t}}\tfrac{\partial v_{t}}{\partial\nu_{t}}+\int_{\Sigma_{t}}\tfrac{K(p_{x})}{\psi^{2}(t_{x})}+\tfrac{3}{2}(H-\bar{h})\int_{\Sigma_{t}}\bar{h}-\int_{\Sigma_{t}}K_{\Sigma_{t}},

where in the last line we have also used the bound (2.23). Now we use (2.57) and also the rewrite (2.14), we see that

(2.70) (Hh¯)Σt1vt\displaystyle-(H-\bar{h})^{\prime}\int_{\Sigma_{t}}\tfrac{1}{v_{t}}
(2.71) \displaystyle\geqslant Σt[cotγ¯At(νt,νt)+1sinγ¯AM(ηt,ηt)+1sin2γ¯ηtcosγ¯]\displaystyle\int_{\partial\Sigma_{t}}[-\cot\bar{\gamma}A_{t}(\nu_{t},\nu_{t})+\tfrac{1}{\sin\bar{\gamma}}A_{\partial M}(\eta_{t},\eta_{t})+\tfrac{1}{\sin^{2}\bar{\gamma}}\nabla_{\eta_{t}}\cos\bar{\gamma}]
(2.72) +ΣtK(px)ψ2(tx)+32(Hh¯)Σth¯ΣtKΣt\displaystyle\quad+\int_{\Sigma_{t}}\tfrac{K(p_{x})}{\psi^{2}(t_{x})}+\tfrac{3}{2}(H-\bar{h})\int_{\Sigma_{t}}\bar{h}-\int_{\Sigma_{t}}K_{\Sigma_{t}}
(2.73) \displaystyle\geqslant Σt[κΣtH(t)cotγ¯+1sinγ¯HM+1sin2γ¯ηtcosγ¯]\displaystyle\int_{\partial\Sigma_{t}}[-\kappa_{\partial\Sigma_{t}}-H(t)\cot\bar{\gamma}+\tfrac{1}{\sin\bar{\gamma}}H_{\partial M}+\tfrac{1}{\sin^{2}\bar{\gamma}}\nabla_{\eta_{t}}\cos\bar{\gamma}]
(2.74) +ΣtK(px)ψ2(tx)+32(Hh¯)Σth¯ΣtKΣt\displaystyle\quad+\int_{\Sigma_{t}}\tfrac{K(p_{x})}{\psi^{2}(t_{x})}+\tfrac{3}{2}(H-\bar{h})\int_{\Sigma_{t}}\bar{h}-\int_{\Sigma_{t}}K_{\Sigma_{t}}
(2.75) =\displaystyle= (ΣtKΣt+ΣtκΣt)+[ΣtK(px)ψ2(tx)+Σt(1sinγ¯HMh¯cotγ¯+1sin2γ¯ηtcosγ¯)]\displaystyle-\left(\int_{\Sigma_{t}}K_{\Sigma_{t}}+\int_{\partial\Sigma_{t}}\kappa_{\partial\Sigma_{t}}\right)+\left[\int_{\Sigma_{t}}\tfrac{K(p_{x})}{\psi^{2}(t_{x})}+\int_{\partial\Sigma_{t}}\left(\tfrac{1}{\sin\bar{\gamma}}H_{\partial M}-\bar{h}\cot\bar{\gamma}+\tfrac{1}{\sin^{2}\bar{\gamma}}\nabla_{\eta_{t}}\cos\bar{\gamma}\right)\right]
(2.76) +32(Hh¯)Σth¯(Hh¯)Σtcotγ¯.\displaystyle\quad+\tfrac{3}{2}(H-\bar{h})\int_{\Sigma_{t}}\bar{h}-(H-\bar{h})\int_{\partial\Sigma_{t}}\cot\bar{\gamma}.

It follows from Lemma 2.2 and the proof of Lemma 2.1 that the second term in the big bracket is bounded below by 2π2\pi. Using also the Gauss-Bonnet theorem on the first term in the bracket, we see that

(2.77) (Hh¯)Σt1vt(Hh¯)(32Σth¯Σtcotγ¯).-(H-\bar{h})^{\prime}\int_{\Sigma_{t}}\tfrac{1}{v_{t}}\geqslant(H-\bar{h})(\tfrac{3}{2}\int_{\Sigma_{t}}\bar{h}-\int_{\partial\Sigma_{t}}\cot\bar{\gamma}).

Let

(2.78) Ψ(t)=(Σt1vt)1(Σtcotγ¯32Σth¯),\Psi(t)=\left(\int_{\Sigma_{t}}\tfrac{1}{v_{t}}\right)^{-1}(\int_{\partial\Sigma_{t}}\cot\bar{\gamma}-\tfrac{3}{2}\int_{\Sigma_{t}}\bar{h}),

then note that we have assume that vt>0v_{t}>0 near t=0t=0, so Hh¯H-\bar{h} satisfies the ordinary differential inequality

(2.79) (Hh¯)Ψ(t)(Hh¯)0.(H-\bar{h})^{\prime}-\Psi(t)(H-\bar{h})\leqslant 0.

We see then

ddt(exp(0tΨ(τ)dτ)(Hh¯))0.\tfrac{\mathrm{d}}{\mathrm{d}t}\left(\exp\left(-\int_{0}^{t}\Psi(\tau)\mathrm{d}\tau\right)(H-\bar{h})\right)\leqslant 0.

So the function exp(0tΨ(τ)dτ)(Hh¯)\exp(-\int_{0}^{t}\Psi(\tau)\mathrm{d}\tau)(H-\bar{h}) is non-increasing. ∎

2.6. From local foliation to rigidity

Let Σt\Sigma_{t} be the constant mean curvature surfaces with prescribed contact angles γ¯\bar{\gamma} with M\partial M.

Proposition 2.7.

Every Σt\Sigma_{t} constructed in Theorem 2.5 is infinitesimally rigid.

Proof.

Let Ωt\Omega_{t} be the component of M\ΣtM\backslash\Sigma_{t} whose closure contains P+MP_{+}\cap\partial M. We abuse the notation and define

F(t)=|Σt|Ωth¯Ωtcosγ¯.F(t)=|\Sigma_{t}|-\int_{\Omega_{t}}\bar{h}-\int_{\partial\Omega_{t}}\cos\bar{\gamma}.

By the first variation formula (2.4),

F(t2)F(t1)=t1t2dtΣt(Hh¯)vt.F(t_{2})-F(t_{1})=\int_{t_{1}}^{t_{2}}\mathrm{d}t\int_{\Sigma_{t}}(H-\bar{h})v_{t}.

By Theorem 2.6,

Hh¯0 if t0; Hh¯0 if t0,H-\bar{h}\leqslant 0\text{ if }t\geqslant 0;\text{ }H-\bar{h}\geqslant 0\text{ if }t\leqslant 0,

which in turn implies that

F(t)0 if t0; F(t)0 if t0.F(t)\leqslant 0\text{ if }t\geqslant 0;\text{ }F(t)\leqslant 0\text{ if }t\leqslant 0.

However, Ωt\Omega_{t} is a minimiser to the functional (2.1), hence

F(t)F(0).F(t)\equiv F(0).

It then follows every Σt\Sigma_{t} is a minimiser, hence infinitesimally rigid. ∎

We can conclude the proof of item (4) of Theorem 1.2.

Proof of item (4) of Theorem 1.2.

We note easily by the assumptions of item (4) of Theorem 1.2 that Σ±=P±M\Sigma_{\pm}=P_{\pm}\cap\partial M are a set of barriers (see Definition 1.5), by the maximum principle, there exists a minimiser EE to (2.2) such that EE is either empty or E\sM\partial E\backslash\partial_{s}M or lies entirely away from P±P_{\pm}. Without loss of generality, we assume that Σ=EintM\Sigma=\partial E\cap\operatorname{int}M non-empty. By [DPM15], Σ\Sigma is a regular stable surface of prescribed mean curvature h¯\bar{h} and prescribed contact angle γ¯\bar{\gamma}. Moreover, the second variation 𝒜′′(0)0\mathcal{A}^{\prime\prime}(0)\geqslant 0 in (2.5) for any smooth family Σs\Sigma_{s} such that Σ0=Σ\Sigma_{0}=\Sigma.

Let Y=ddtϕ(x,w(x,t))Y=\tfrac{\mathrm{d}}{\mathrm{d}t}\phi(x,w(x,t)) where ϕ\phi and ww are as Theorem 2.5, we show first that NtN_{t} is conformal. It suffices to show that YY^{\bot} is conformal.

Since every Σt\Sigma_{t} is infinitesimally rigid by Proposition 2.7, from (2.46) and (2.48), we know that Y,Nt\langle Y,N_{t}\rangle is a constant. Let i\partial_{i}, i=1,2i=1,2 be vector fields induced by local coordinates on Σ\Sigma, i\partial_{i} also extends to a neighborhood of Σ\Sigma via the diffeomorphism ϕ\phi. We have iY,N=0\nabla_{\partial_{i}}\langle Y,N\rangle=0. Note that Σt\Sigma_{t} are umbilical with constant mean curvature h¯\bar{h}, so

iN12h¯i\nabla_{\partial_{i}}N\equiv\tfrac{1}{2}\bar{h}\partial_{i}

and

(2.80) 0\displaystyle 0 =iY,N\displaystyle=\nabla_{\partial_{i}}\langle Y,N\rangle
(2.81) =iY,N+Y,iN\displaystyle=\langle\nabla_{\partial_{i}}Y,N\rangle+\langle Y,\nabla_{\partial_{i}}N\rangle
(2.82) =iY,N+12h¯Y,i.\displaystyle=\langle\nabla_{\partial_{i}}Y,N\rangle+\tfrac{1}{2}\bar{h}\langle Y,\partial_{i}\rangle.

On the other hand,

(2.83) 0\displaystyle 0 =iY,N=Yi,N\displaystyle=\langle\nabla_{\partial_{i}}Y,N\rangle=\langle\nabla_{Y}\partial_{i},N\rangle
(2.84) =Yi,Ni,YN\displaystyle=Y\langle\partial_{i},N\rangle-\langle\partial_{i},\nabla_{Y}N\rangle
(2.85) =i,YN\displaystyle=-\langle\partial_{i},\nabla_{Y}N\rangle
(2.86) =i,YNi,YN\displaystyle=-\langle\partial_{i},\nabla_{Y^{\top}}N\rangle-\langle\partial_{i},\nabla_{Y^{\bot}}N\rangle
(2.87) =12h¯Y,ii,YN.\displaystyle=-\tfrac{1}{2}\bar{h}\langle Y^{\top},\partial_{i}\rangle-\langle\partial_{i},\nabla_{Y^{\bot}}N\rangle.

Combining the two equations above, we conclude that YN=0\nabla_{Y^{\bot}}N=0 which implies that Σ\Sigma foliates a warped product under the diffeomorphism ϕ\phi (parameterized by tt). Considering that the induced metric on Σ\Sigma agrees with the induced metric from g¯\bar{g}, we conclude that g=g¯g=\bar{g}. ∎

3. Construction of barriers (I)

In this section, we prove item (3) of Theorem 1.2. Our strategy is to construct a surface Σ\Sigma_{-} (Σ+\Sigma_{+}) which serves as a lower (upper) barrier, and to use item (4) of Theorem 1.2 to finish the proof. This section is occupied by such a construction of Σ\Sigma_{-}.

3.1. Setting up coordinates and notations

For convenience, we set t=0t_{-}=0. As before, for any t>0t>0, we set Σt\Sigma_{t} to be the tt-level set of tt and Ωt\Omega_{t} to be the tt-sublevel set, that is, all points of MM which lie below Σt\Sigma_{t}. Since both (M,g)(M,g) and (M,g¯)(M,\bar{g}) has cone structures near where t=0t_{-}=0 where each cross-section of the cone is a topological disk and it collapses to a point which we denote by p0p_{0}.

In the following subsections, we construct graphical perturbations Σt,t2u\Sigma_{t,t^{2}u} of Σt\Sigma_{t}. Let Σt,t2u\Sigma_{t,t^{2}u} be the surface which consists of points x+t2u(x,t)Nt(x)x+t^{2}u(x,t)N_{t}(x) where NtN_{t} is the unit normal of Σt\Sigma_{t} with respect to the metric gg at xΣtx\in\Sigma_{t}. The boundary Σt,t2u\partial\Sigma_{t,t^{2}u} might not lie in sM\partial_{s}M, we can compensate this by expanding or shrinking Σt,t2u\Sigma_{t,t^{2}u} a little, and we still denote the resulting surface Σt,t2u\Sigma_{t,t^{2}u}.

We use a tt subscript on every geometric quantity on Σt\Sigma_{t} and a t,t2ut,t^{2}u subscript on every geometric quantity on Σt,t2u\Sigma_{t,t^{2}u}. We will explicitly indicate when there was confusion or change.

Refer to caption
P0P_{0}
t=0t=0
Σt\Sigma_{t}
Σt,t2u\Sigma_{t,t^{2}u}
Refer to caption
Figure 3.1. Construction of Σt,t2u\Sigma_{t,t^{2}u}.

3.2. Capillary foliation with constant Hh¯H-\bar{h}

We assume that (M,g)(M,g) and (M,g¯)(M,\bar{g}) have isometric tangent cones at p0p_{0} and we construct a foliation of constant Hh¯H-\bar{h} with prescribed angles γ¯\bar{\gamma} near p0p_{0}. In fact, later in Subsection 3.3, it is shown that this is the only case.

By the first variation formula of the mean curvatures

(3.1) Ht,t2uHt=Δtut2(Ric(Nt)+|At|2)u+O(t),H_{t,t^{2}u}-H_{t}=-\Delta_{t}u-t^{2}(\operatorname{Ric}(N_{t})+|A_{t}|^{2})u+O(t),

where Δt\Delta_{t} is the Laplacian with respect to the induced rescaled metric t2g|Σtt^{-2}g|_{\Sigma_{t}}. Note that Ric(Nt)=O(t1)\operatorname{Ric}(N_{t})=O(t^{-1}) by the fact the tangent cone is dt2+a2t2gS2\mathrm{d}t^{2}+a^{2}t^{2}g_{S^{2}}. By the Taylor expansion of the function h¯\bar{h}, we see that

(3.2) h¯t,t2uh¯t=h¯(t)t2u=t2uNth¯+O(t).\bar{h}_{t,t^{2}u}-\bar{h}_{t}=\bar{h}^{\prime}(t)t^{2}u=t^{2}u\nabla_{N_{t}}\bar{h}+O(t).

So

(3.3) (Ht,t2uh¯t,t2u)(Hth¯t)=Δtut2(Ric(Nt)+|At|2+Nth¯)u+O(t).(H_{t,t^{2}u}-\bar{h}_{t,t^{2}u})-(H_{t}-\bar{h}_{t})=-\Delta_{t}u-t^{2}(\operatorname{Ric}(N_{t})+|A_{t}|^{2}+\nabla_{N_{t}}\bar{h})u+O(t).

Note that both Hth¯tH_{t}-\bar{h}_{t} and Ht,t2uh¯t,t2uH_{t,t^{2}u}-\bar{h}_{t,t^{2}u} are finite and |At|2+Nth¯=O(t1)|A_{t}|^{2}+\nabla_{N_{t}}\bar{h}=O(t^{-1}) considering that (M,g)(M,g) and (M,g¯)(M,\bar{g}) have isometric tangent cones at p0p_{0}.

Remark 3.1.

We elaborate a bit more on (3.1) and its O(t)O(t) remainder term. Since the metric gg is close to dt2+ψ(t)2gS2\mathrm{d}t^{2}+\psi(t)^{2}g_{S^{2}} when t0+t\to 0^{+}, we calculate the expansions with respect to the rescaled metric t2gt^{-2}g when computing for small t>0t>0. This is similar to [Ye91]. Then we rescale back and we obtain (3.1). The term O(t)O(t) involves products of |At||A_{t}| which is of order t1t^{-1} with terms of order at most O(1)O(1). That is why the remainder is only of order O(t)O(t) instead of O(t2)O(t^{2}).

Also, the variation of angles give

(3.4) t1[Xt,t2u,Nt,t2uXt,Nt]\displaystyle t^{-1}[\langle X_{t,t^{2}u},N_{t,t^{2}u}\rangle-\langle X_{t},N_{t}\rangle]
(3.5) =\displaystyle= sinγuνt+t(cosγA(t1νt,t1νt)+AM(ηt,ηt))u+O(t2),\displaystyle-\sin\gamma\tfrac{\partial u}{\partial\nu_{t}}+t(-\cos\gamma A(t^{-1}\nu_{t},t^{-1}\nu_{t})+A_{\partial M}(\eta_{t}{,}\eta_{t}))u+O(t^{2}),

where νt\nu_{t} is the outward unit normal of Σt\partial\Sigma_{t} in Σt\Sigma_{t} with respect to the rescaled induced metric t2g|Σtt^{-2}g|_{\Sigma_{t}} (note that t1νtt^{-1}\nu_{t} is of unit length with respect to gg). Other geometric quantities are not rescaled. By the variation of the prescribed angle γ¯\bar{\gamma},

(3.6) t1(cosγ¯t,t2ucosγ¯t)=tusinγ¯η¯tcosγ¯+O(t2).t^{-1}(\cos\bar{\gamma}_{t,t^{2}u}-\cos\bar{\gamma}_{t})=-\tfrac{tu}{\sin\bar{\gamma}}\partial_{\bar{\eta}_{t}}\cos\bar{\gamma}+O(t^{2}).

So

(3.7) t1[(Xt,t2u,Nt,t2ucosγ¯t,t2u)(Xt,Ntcosγ¯t)]\displaystyle t^{-1}[(\langle X_{t,t^{2}u},N_{t,t^{2}u}\rangle-\cos\bar{\gamma}_{t,t^{2}u})-(\langle X_{t},N_{t}\rangle-\cos\bar{\gamma}_{t})]
(3.8) =\displaystyle= sinγuνt\displaystyle-\sin\gamma\tfrac{\partial u}{\partial\nu_{t}}
(3.9) +t(cosγA(t1νt,t1νt)+AM(ηt,ηt)+1sinγ¯η¯tcosγ¯)u+O(t2).\displaystyle\quad+t(-\cos\gamma A(t^{-1}\nu_{t},t^{-1}\nu_{t})+A_{\partial M}(\eta_{t}{,}\eta_{t})+\tfrac{1}{\sin\bar{\gamma}}\partial_{\bar{\eta}_{t}}\cos\bar{\gamma})u+O(t^{2}).
Remark 3.2.

The term A(t1νt,t1νt)=O(t1)A(t^{-1}\nu_{t},t^{-1}\nu_{t})=O(t^{-1}), however, we observe that limt0γ¯t=π/2\lim_{t\to 0}\bar{\gamma}_{t}=\pi/2, and AM(ηt,ηt)=O(1)A_{\partial M}(\eta_{t},\eta_{t})=O(1) since AM(η¯t,η¯t)=O(1)A_{\partial M}(\bar{\eta}_{t},\bar{\eta}_{t})=O(1). Or we can calculate with respect to the rescaling metric as in Remark 3.1.

Since gg and g¯\bar{g} has isometric tangent cone at p0p_{0}, we see that the limit of the surface (Σt,t2g|Σt)(\Sigma_{t},t^{-2}g|_{\Sigma_{t}}) as t0t\to 0 is (Σ,a2gS2)(\Sigma,a^{2}g_{{S}^{2}}) where Σ\Sigma is a scaling copy of a geodesic disk of radius ρ(0)=limt0ρ(t)>0\rho(0)=\lim_{t\to 0}\rho(t)>0 in the standard 2-sphere. Consider the spaces

𝒴={uC2,α(Σ)C1,α(Σ¯):Σu=0}\mathcal{Y}=\left\{u\in C^{2,\alpha}(\Sigma)\cap C^{1,\alpha}(\bar{\Sigma}):\int_{\Sigma}u=0\right\}

and

(3.10) 𝒵={uC0,α(Σ):Du=0}.\mathcal{Z}=\left\{u\in C^{0,\alpha}(\Sigma):\int_{D}u=0\right\}.

Given small δ>0\delta>0 and ε>0\varepsilon>0, we define the map

(3.11) Φ:(ε,ε)×B(0,δ)𝒵×C1,α(Σ)\Phi:(-\varepsilon,\varepsilon)\times B(0,\delta)\to\mathcal{Z}\times C^{1,\alpha}(\partial\Sigma)

given by Φ(t,u)=(Φ1(t,u),Φ2(t,u))\Phi(t,u)=(\Phi_{1}(t,u),\Phi_{2}(t,u)) where Φi\Phi_{i}, i=1,2i=1,2 are given by

(3.12) Φ1(t,u)=\displaystyle\Phi_{1}(t,u)= (Ht,t2uh¯t,t2u)1|Σ|Σ(Ht,t2uh¯t,t2u),\displaystyle(H_{t,t^{2}u}-\bar{h}_{t,t^{2}u})-\frac{1}{|\Sigma|}\int_{\Sigma}(H_{t,t^{2}u}-\bar{h}_{t,t^{2}u}),
(3.13) Φ2(t,u)=\displaystyle\Phi_{2}(t,u)= t1(Xt,t2u,Nt,t2ucosγ¯t,t2u)\displaystyle t^{-1}(\langle X_{t,t^{2}u},N_{t,t^{2}u}\rangle-\cos\bar{\gamma}_{t,t^{2}u})

for t0t\neq 0. Here B(0,δ)𝒴B(0,\delta)\subset\mathcal{Y} is an open ball with radius δ\delta in the C2,αC^{2,\alpha} norm and the integration on Σ\Sigma is with respect to the metric g𝕊2g_{\mathbb{S}^{2}}. We extend Φ(t,u)\Phi(t,u) to t=0t=0 by taking limits, that is,

(3.14) Φ(0,u)=limt0Ψ(t,u).\Phi(0,u)=\lim_{t\to 0}\Psi(t,u).

We have the following proposition.

Proposition 3.3.

For each t[0,ε)t\in[0,\varepsilon) with ε\varepsilon small enough, we can find ut=u(,t)C2,α(Σ)C1,α(Σ¯)u_{t}=u(\cdot,t)\in C^{2,\alpha}(\Sigma)\cap C^{1,\alpha}(\bar{\Sigma}) such that Σu(,t)=0\int_{\Sigma}u(\cdot,t)=0 and

Φ(t,ut)=(0,0).\Phi(t,u_{t})=(0,0).

In particular, each of the surfaces Σt,t2u\Sigma_{t,t^{2}u} have constant λt:=Ht,t2uh¯t,t2u\lambda_{t}:=H_{t,t^{2}u}-\bar{h}_{t,t^{2}u} and prescribed angles γt,t2u=γ¯t,t2u\gamma_{t,t^{2}u}=\bar{\gamma}_{t,t^{2}u}. Moreover, λt0\lambda_{t}\leqslant 0 for all small t[0,ε)t\in[0,\varepsilon).

Before proving this proposition, we give a variational lemma.

Lemma 3.4.

Suppose that (Ω,g^)(\Omega,\hat{g}) is a compact manifold with piecewise smooth boundary Ω\partial\Omega and Σ\Sigma is a relatively open, smooth subset of Ω\partial\Omega. Let gsg_{s} be a smooth family of metrics indexed by s[0,ε)s\in[0,\varepsilon) such that gsg^g_{s}\to\hat{g} as s0s\to 0, let hs=gsg^h_{s}=g_{s}-\hat{g}. We now omit the subscript on hsh_{s}. Let ν\nu be the unit outward normal of Ω\partial\Omega in (Ω,g)(\Omega,g), HgH_{g} and AgA_{g} be the mean curvatures and the second fundamental form of Ω\partial\Omega in (Ω,g)(\Omega,g) computed with respect to the unit normal pointing outward of Ω\Omega, and γ\gamma be the dihedral angles formed by Σ\Sigma and Ω\Σ\partial\Omega\backslash\Sigma with respect to the metric gg. We put a hat at appropriate places for the geometric quantities with respect to g^\hat{g}.

Then

(3.15) 2[Σ(HgHg^)+Σ1sinγg^(cosγg^cosγg)]\displaystyle 2\left[-\int_{\Sigma}(H_{g}-H_{\hat{g}})+\int_{\partial\Sigma}\tfrac{1}{\sin\gamma_{\hat{g}}}(\cos\gamma_{\hat{g}}-\cos\gamma_{g})\right]
(3.16) =\displaystyle= Ω((RgRg^)+Ricg^,hg^)+2Ω\Σ(HgHg^)+Ωh,Ag^+O(s2).\displaystyle\int_{\Omega}((R_{g}-R_{\hat{g}})+\langle\operatorname{Ric}_{\hat{g}},h\rangle_{\hat{g}})+2\int_{\partial\Omega\backslash\Sigma}(H_{g}-H_{\hat{g}})+\int_{\partial\Omega}\langle h,A_{\hat{g}}\rangle+O(s^{2}).

Here, we have used O(s2)O(s^{2}) to denote a remainder term comparable to |h|g^2+|h|g^|^h|g^+|^h|g^2|h|_{\hat{g}}^{2}+|h|_{\hat{g}}|\hat{\nabla}h|_{\hat{g}}+|\hat{\nabla}h|_{\hat{g}}^{2}.

Proof.

From the variational formulas of the scalar curvature and the mean curvature, we have

(3.17) RgRg^=Ricg^,hg^divg^(d(trg^h)divg^h)+O(s2),R_{g}-R_{\hat{g}}=-\langle\operatorname{Ric}_{\hat{g}},h\rangle_{\hat{g}}-\operatorname{div}_{\hat{g}}(\mathrm{d}(\operatorname{tr}_{\hat{g}}h)-\operatorname{div}_{\hat{g}}h)+O(s^{2}),

and

(3.18) 2(HgHg^)=(d(trg^h)divg^h)(ν^)divσYh,Ag^σ+O(s2)2(H_{g}-H_{\hat{g}})=(\mathrm{d}(\operatorname{tr}_{\hat{g}}h)-\operatorname{div}_{\hat{g}}h)(\hat{\nu})-\operatorname{div}_{\sigma}Y-\langle h,A_{\hat{g}}\rangle_{\sigma}+O(s^{2})

where YY is the tangential component dual to the 1-form h(,ν^)h(\cdot,\hat{\nu}). For the explicit form of the remainder terms, refer to [BM11, Proposition 4] and [MP21].

We integrate the variation of the mean curvature (3.18) on the boundary Ω\partial\Omega with respect to the metric g^\hat{g}, we see

Ω[(d(trg^h)divg^h)(ν^)divσ^Yh,Ag^]=2Ω(HgHg^)+O(s2).\int_{\partial\Omega}[(\mathrm{d}(\operatorname{tr}_{\hat{g}}h)-\operatorname{div}_{\hat{g}}h)(\hat{\nu})-\operatorname{div}_{\hat{\sigma}}Y-\langle h,A_{\hat{g}}\rangle]=2\int_{\partial\Omega}(H_{g}-H_{\hat{g}})+O(s^{2}).

By the divergence theorem and the variation of the scalar curvature,

Ω(d(trg^h)divg^h)(g^)=Ω[(RgRg^)Ricg^,hg^]+O(s2).\int_{\partial\Omega}(\mathrm{d}(\operatorname{tr}_{\hat{g}}h)-\operatorname{div}_{\hat{g}}h)(\hat{g})=\int_{\Omega}[-(R_{g}-R_{\hat{g}})-\langle\operatorname{Ric}_{\hat{g}},h\rangle_{\hat{g}}]+O(s^{2}).

For the term Ωdivσ^Y\int_{\partial\Omega}\operatorname{div}_{\hat{\sigma}}Y, we follow [MP21, (3.18)] and obtain

Ωdivσ^Y=Σdivg^Y+Ω\Σdivσ^Y=2Σ1sinγ^(cosγ^cosγ)+O(s2).\int_{\partial\Omega}\operatorname{div}_{\hat{\sigma}}Y=\int_{\Sigma}\operatorname{div}_{\hat{g}}Y+\int_{\partial\Omega\backslash\Sigma}\operatorname{div}_{\hat{\sigma}}Y=2\int_{\partial\Sigma}\tfrac{1}{\sin\hat{\gamma}}(\cos\hat{\gamma}-\cos\gamma)+O(s^{2}).

Collecting all the formulas in the proof proves the lemma. ∎

Lemma 3.5 implies the following by taking the difference of two families of metrics.

Corollary 3.5.

Assume (Ω,g^)(\Omega,\hat{g}) is the manifold from Lemma 3.5, for two family of metrics {gi}i=1,2\{g_{i}\}_{i=1,2} close to g^\hat{g} indexed both by a small parameter ss, we have

(3.19) 2[Σ(Hg2Hg1)+Σ1sinγ^(cosγg1cosγg2)]\displaystyle 2\left[-\int_{\Sigma}(H_{g_{2}}-H_{g_{1}})+\int_{\partial\Sigma}\tfrac{1}{\sin\hat{\gamma}}(\cos\gamma_{g_{1}}-\cos\gamma_{g_{2}})\right]
(3.20) =\displaystyle= Ω((Rg2Rg1)+Ricg^,g2g1g^)+2Ω\Σ(Hg2Hg1)+Ωg2g1,Ag^+O(s2).\displaystyle\int_{\Omega}((R_{g_{2}}-R_{g_{1}})+\langle\operatorname{Ric}_{\hat{g}},g_{2}-g_{1}\rangle_{\hat{g}})+2\int_{\partial\Omega\backslash\Sigma}(H_{g_{2}}-H_{g_{1}})+\int_{\partial\Omega}\langle g_{2}-g_{1},A_{\hat{g}}\rangle+O(s^{2}).

Now we are ready to prove Proposition 3.3.

Proof of Proposition 3.3.

The proof is similar to [CW23]. We bring up only the main differences.

Because the right hand of both (3.3) and (3.9) converge to Δu\Delta u and uν\tfrac{\partial u}{\partial\nu} (up to a constant) respectively, so we can first follow [CW23, Proposition 4.2] to construct a foliation {Σt,t2u}t[0,ε)\{\Sigma_{t,t^{2}u}\}_{t\in[0,\varepsilon)} near p0p_{0} with constant Hh¯H-\bar{h} and γt,t2u=γ¯t,t2u\gamma_{t,t^{2}u}=\bar{\gamma}_{t,t^{2}u} along Σt,t2u\partial\Sigma_{t,t^{2}u}, and then [CW23, Lemma 4.3] to obtain that

(3.21) λt|Σt|=Σt(Hth¯t)+Σt1sinγt(cosγ¯tcosγt)+O(t3).-\lambda_{t}|\Sigma_{t}|=\int_{\Sigma_{t}}(H_{t}-\bar{h}_{t})+\int_{\partial\Sigma_{t}}\tfrac{1}{\sin\gamma_{t}}(\cos\bar{\gamma}_{t}-\cos\gamma_{t})+O(t^{3}).

Now we show that limt0λt0\lim_{t\to 0}\lambda_{t}\leqslant 0.

We consider the rescaled set t1Ωtt^{-1}\Omega_{t} with two rescaled metrics t2gt^{-2}g and t2g¯t^{-2}\bar{g}. Since g¯=dt2+ψ(t)2gS2\bar{g}=\mathrm{d}t^{2}+\psi(t)^{2}g_{{S}^{2}} and ψ(t)=at+o(t)\psi(t)=at+o(t), it is easy to see that (t1Ωt,t2g¯)(t^{-1}\Omega_{t},t^{-2}\bar{g}) converges to a truncated metric cone Λ=(0,1]×D\Lambda=(0,1]\times D with the metric ϱ:=ds2+a2s2gS2\varrho:=\mathrm{d}s^{2}+a^{2}s^{2}g_{{S}^{2}} where s(0,1]s\in(0,1] and (D,a2gS2)(D,a^{2}g_{{S}^{2}}) is some convex disk in a 2-sphere (S2,a2gS2)({S}^{2},a^{2}g_{{S}^{2}}). We set Ds={s}×DD_{s}=\{s\}\times D. Since gg and g¯\bar{g} has isometric tangent cone at p0p_{0}, (t1Ωt,t2g)(t^{-1}\Omega_{t},t^{-2}g) converges to (Λ,ϱ)(\Lambda,\varrho) as well. Therefore, we can view g1=t2gg_{1}=t^{-2}g and g2=t2g¯g_{2}=t^{-2}\bar{g} (indexed by tt) as two metrics on Λ\Lambda getting closer to ϱ\varrho as t0t\to 0. We rescale (3.21) by a factor of t2t^{-2}, we obtain

λt|Σt|t2=Σt(Hth¯t)t2+Σt1sinγt(cosγ¯tcosγt)t2+O(t)-\lambda_{t}|\Sigma_{t}|t^{-2}=\int_{\Sigma_{t}}(H_{t}-\bar{h}_{t})t^{-2}+\int_{\partial\Sigma_{t}}\tfrac{1}{\sin\gamma_{t}}(\cos\bar{\gamma}_{t}-\cos\gamma_{t})t^{-2}+O(t)

which is equivalent to

λt|D|g1=D(Hg2Hg1)+D1sinγt(cosγ¯tcosγt)+O(t).-\lambda_{t}|D|_{g_{1}}=\int_{D}(H_{g_{2}}-H_{g_{1}})+\int_{\partial D}\tfrac{1}{\sin\gamma_{t}}(\cos\bar{\gamma}_{t}-\cos\gamma_{t})+O(t).

In the above the integration done is with respect to the metric g1g_{1} and HgiH_{g_{i}} are the mean curvature of {1}×D\{1\}\times D in (Λ,gi)(\Lambda,g_{i}) computed with respect to the normal pointing inside of Λ\Lambda.

All the comparisons in item (3) of Theorem 1.2 carry over to the rescaled metrics g1g_{1} and g2g_{2} on Λ\Lambda, and that (Λ,ϱ)(\Lambda,\varrho) has non-negative Ricci curvature by the assumptions of item (3) of Theorem 1.2. We use Corollary 3.5 and arrive that λtO(t)\lambda_{t}\leqslant O(t), that is,

limt0λ(t)0.\lim_{t\to 0}\lambda(t)\leqslant 0.

Since λt\lambda_{t} satisfies the differential inequality (2.79) and considering the asymptotics u(,t)=1+O(t)u(\cdot,t)=1+O(t), cotγ¯=O(t)\cot\bar{\gamma}=O(t) and h¯=2/t+O(1)\bar{h}=2/t+O(1) in (2.78), we see that λt0\lambda_{t}\leqslant 0 for all t(0,ε)t\in(0,\varepsilon). ∎

Remark 3.6.

The Ricci curvature in Corollary 3.5 blows up near {0}×D\{0\}\times D, however, because we are integrating with respect to the metric ϱ\varrho, the volume near {0}×D\{0\}\times D is small. Also, the difference g2g1g_{2}-g_{1} is small. So the blowing up of the Ricci curvature will not cause an issue.

3.3. Barrier construction with non-isometric tangent cones

Since g¯=dt2+ψ(t)2gS2\bar{g}=\mathrm{d}t^{2}+\psi(t)^{2}g_{{S}^{2}}, the manifold (M,g¯)(M,\bar{g}) is topologically a cone near t=0t=0 and it is a point at t=0t=0. According to the assumptions of item (3) of Theorem 1.2, (M,g)(M,g) at p0p_{0} also locally resembles a cone, that is,

(3.22) g=ds2+s2g0+g1,g=\mathrm{d}s^{2}+s^{2}g_{0}+g_{1},

where ss is a parameter, g0g_{0} is a metric on a two dimensional disk DD and g1g_{1} is small compare to ds2+s2g0\mathrm{d}s^{2}+s^{2}g_{0}. In other words, the tangent cone at p0p_{0} is a cone with the metric ds2+s2g0\mathrm{d}s^{2}+s^{2}g_{0}.

Now we can also identify MM near p0p_{0} as (0,ε)×D(0,\varepsilon)\times D and tt as a function on (0,ε)×D(0,\varepsilon)\times D. Let (s,x)(0,ε)×D(s,x)\in(0,\varepsilon)\times D, we see that τ:=s/t\tau:=s/t as a function on MM only depends on xDx\in D. So we view τ\tau as a function on DD. Since gg¯g\geqslant\bar{g} on MM, we have that τ(x)1\tau(x)\geqslant 1. Now we discuss the case that τ(x)1\tau(x)\equiv 1 on DD.

Lemma 3.7.

If τ1\tau\equiv 1 on DD, then g0=a2gS2g_{0}=a^{2}g_{{S}^{2}}. That is, (M,g)(M,g) and (M,g¯)(M,\bar{g}) have isometric tangent cones at p0p_{0}.

Proof.

Since τ1\tau\equiv 1, so we can rescale (M,g¯)(M,\bar{g}) and (M,g)(M,g) by the same scale to obtain a cone 𝒞=(0,)×D\mathcal{C}=(0,\infty)\times D but with two different metrics χ1=dt2+a2t2gS2\chi_{1}=\mathrm{d}t^{2}+a^{2}t^{2}g_{{S}^{2}} and χ2=dt2+t2g0\chi_{2}=\mathrm{d}t^{2}+t^{2}g_{0}. For s>0s>0, set Ds={s}×D𝒞D_{s}=\{s\}\times D\subset\mathcal{C}. Since the metric comparison, the mean curvature and the scalar curvature comparison are preserved by rescaling, so g0a2gS2g_{0}\geqslant a^{2}g_{{S}^{2}}, the scalar curvature Rχ2Rχ1R_{\chi_{2}}\geqslant R_{\chi_{1}} and the mean curvature of 𝒞\partial\mathcal{C} at D1\partial D_{1} satisfies Hχ2Hχ1H_{\chi_{2}}\geqslant H_{\chi_{1}}.

Since both χi\chi_{i}, i=1,2i=1,2 are warped product metrics, the comparison Rχ2Rχ1R_{\chi_{2}}\geqslant R_{\chi_{1}} reduces to Gaussian curvature comparison K2K1=a2K_{2}\geqslant K_{1}=a^{-2} of (D1,g0)(D_{1},g_{0}) and (D1,a2gS2)(D_{1},a^{2}g_{{S}^{2}}) by a direct computation of scalar curvature (or Gauss equation). Let κi\kappa_{i} be the geodesic curvatures of D1\partial D_{1} with respect to χi|D1\chi_{i}|_{D_{1}}. By direct calculation, the second fundamental form A𝒞(i)A^{(i)}_{\partial\mathcal{C}} of 𝒞\partial\mathcal{C} in the direction t\partial_{t} vanishes with respect to both metrics χi\chi_{i} and the second fundamental form AD1(i)A_{D_{1}}^{(i)} of D1D_{1} in 𝒞\mathcal{C} with respect to χi\chi_{i} agree. It then follows from Hχ2Hχ1H_{\chi_{2}}\geqslant H_{\chi_{1}} and (2.14) that κ2κ1\kappa_{2}\geqslant\kappa_{1}.

To summarize, we have comparisons on D1D_{1} that g0a2gS2g_{0}\geqslant a^{2}g_{{S}^{2}}, K2K1K_{2}\geqslant K_{1} and κ2κ1\kappa_{2}\geqslant\kappa_{1} along D1\partial D_{1}. By Gauss-Bonnet theorem, g0a2gS2g_{0}\equiv a^{2}g_{{S}^{2}} on D1D_{1} and it follows that χ1χ2\chi_{1}\equiv\chi_{2}. Therefore, (M,g)(M,g) and (M,g¯)(M,\bar{g}) have isometric tangent cones at p0p_{0}. ∎

By the above lemma, the case τ1\tau\equiv 1 is the case which implies isometric tangent cones of (M,g)(M,g) and (M,g¯)(M,\bar{g}) at p0p_{0}. This is the case we have already addressed in Subsection 3.2. Without loss of generality, we assume that τ1\tau\mathrel{\not\equiv}1.

We first consider the difference of Hh¯H-\bar{h} of the perturbation for DsD_{s}. We now represent h¯\bar{h} at DsD_{s} and its value at the graphical perturbations of DsD_{s} by ζ\zeta to avoid notational confusion. By the first variation of the mean curvatures,

(3.23) (Hs,s2uζs,s2u)(Hsζs)\displaystyle(H_{s,s^{2}u}-\zeta_{s,s^{2}u})-(H_{s}-\zeta_{s})
(3.24) =\displaystyle= Δsus2(Ric(Ns)+|As|2+s2(ζs,s2uζs))u+O(s),\displaystyle-\Delta_{s}u-s^{2}(\operatorname{Ric}(N_{s})+|A_{s}|^{2}+s^{-2}(\zeta_{s,s^{2}u}-\zeta_{s}))u+O(s),

where Δs\Delta_{s} is the Laplacian with respect to the metric s2g|Dss^{-2}g|_{D_{s}}.

Remark 3.8.

We have {(s1Ds,s2g|Ds)}s>0\{(s^{-1}D_{s},s^{-2}g|_{D_{s}})\}_{s>0} converges to (D,g0)(D,g_{0}) as s0s\to 0 by the metric (3.22) near p0p_{0}, and to indicate that the limit carries the metric g0g_{0}, we use D0D_{0} instead of DD only.

Lemma 3.9.

We have that

s2(|As|2s2(ζs,s2uζs))=(22τ)+O(s).s^{2}(|A_{s}|^{2}-s^{-2}(\zeta_{s,s^{2}u}-\zeta_{s}))=(2-2\tau)+O(s).
Proof.

Since {(s1Λs,s2g)}s>0\{(s^{-1}\Lambda_{s},s^{-2}g)\}_{s>0} converges to a truncated radial cone and {(s1Ds,s2g|Ds)}s>0\{(s^{-1}D_{s},s^{-2}g|_{D_{s}})\}_{s>0} converges to the section of the radial cone with unit distance to p0p_{0}, so the section has second fundamental form 2-2 and by rescaling,

|As|2=2s2+O(s1)|A_{s}|^{2}=2s^{-2}+O(s^{-1})

as s0s\to 0.

At a point p=(s,x)Dsp=(s,x)\in D_{s}, the value of tt is given by t=sτ(x)t=s\tau(x) where xx is the projection of pp to the second coordinate. Since τ\tau as a function on MM only depends on xx, we see that the value of the function tt at the graphical perturbation s+s2us+s^{2}u of DsD_{s} is given by (s+s2u)τ(s+s^{2}u)\tau. Since h¯(t)=2t1+O(1)\bar{h}(t)=2t^{-1}+O(1), so

ζs,s2uζs=2(s+s2u)τ2sτ+O(1)=2τs2(s2u)+O(1).\zeta_{s,s^{2}u}-\zeta_{s}=\tfrac{2}{(s+s^{2}u)\tau}-\tfrac{2}{s\tau}+O(1)=-\tfrac{2\tau}{s^{2}}(s^{2}u)+O(1).

Hence

s2(|As|2+s2(ζs,s2uζs))=(22τ)+O(s),s^{2}(|A_{s}|^{2}+s^{-2}(\zeta_{s,s^{2}u}-\zeta_{s}))=(2-2\tau)+O(s),

which proves the lemma. ∎

Let f=lims0s2(Ric(Ns)+|As|2+s2(ζs,s2uζs))f=\lim_{s\to 0}s^{2}(\operatorname{Ric}(N_{s})+|A_{s}|^{2}+s^{-2}(\zeta_{s,s^{2}u}-\zeta_{s})) which is a function on the limit D0D_{0}, so

(3.25) lims0[(Hs,s2uζs,s2u)(Hsζs)]=Δ0ufu,\lim_{s\to 0}[(H_{s,s^{2}u}-\zeta_{s,s^{2}u})-(H_{s}-\zeta_{s})]=-\Delta_{0}u-fu,

where Δ0\Delta_{0} is the Laplacian of D0D_{0}. Recall that Ric(Ns)=O(s1)\operatorname{Ric}(N_{s})=O(s^{-1}), so

f=22τ on D0.f=2-2\tau\text{ on }D_{0}.

Let αs\alpha_{s} be the dihedral angles formed by M\partial M and DsD_{s}, and αs,s2u\alpha_{s,s^{2}u} be the angles formed by M\partial M and the graphical perturbation of DsD_{s}.

Lemma 3.10.

The dihedral angles αs\alpha_{s} formed by M\partial M and DsD_{s} approach π/2\pi/2 as s0s\to 0.

Proof.

Since {(s1Λs,s2g)}s>0\{(s^{-1}\Lambda_{s},s^{-2}g)\}_{s>0} converges to a truncated radial cone, {(s1Ds,s2g|Ds)}s>0\{(s^{-1}D_{s},s^{-2}g|_{D_{s}})\}_{s>0} converges to the section of the radial cone with unit distance to p0p_{0}, and this section is orthogonal to the radial direction in the limit,  so the intersection angles of M\partial M and DsD_{s} approaches π/2\pi/2 as s0s\to 0. ∎

Lemma 3.11.

We have that AM(η,η)=O(1)A_{\partial M}(\eta,\eta)=O(1).

Proof.

The lemma can be deduced from that η\eta is approximately the radial direction s\partial_{s} as s0s\to 0, the scaling property of AMA_{\partial M} and the following lemma. ∎

Lemma 3.12.

Let (S,σ)(S,\sigma) be a 2-surface with boundary and (C=[0,)×S,ds2+s2σ)(C=[0,\infty)\times S,\mathrm{d}s^{2}+s^{2}\sigma) be the cone over (S,σ)(S,\sigma). Then the second fundamental form of C\partial C in CC in the direction t\partial_{t} vanishes.

Proof.

Let ZZ be a tangent vector field over Σ\Sigma, then by direct calculation tZ=Xt=s1Z\nabla_{\partial_{t}}Z=\nabla_{X}\partial_{t}=s^{-1}Z. So tZ,t=0\langle\nabla_{\partial_{t}}Z,\partial_{t}\rangle=0 since on CC the metric is dt2+t2σ\mathrm{d}t^{2}+t^{2}\sigma. Due to the same reason, the unit normal vector ZZ of C\partial C in MM is tangent to Σ\Sigma, so the claim is proved. ∎

We are interested in the difference between αs,s2u\alpha_{s,s^{2}u} and the value of γ¯\bar{\gamma} which to avoid confusion we denote by βs\beta_{s} (βs,s2u\beta_{s,s^{2}u}) at (the graphical perturbation s2us^{2}u of) DsD_{s}. Using the relation of ss and tt, β=γ¯s/τ,s2u/τ\beta=\bar{\gamma}_{s/\tau,s^{2}u/\tau}. By the expansion of angles (see (3.5)), we see

cosαs,s2ucosαs=sinαsuνs+s(cosαsA(s1νs,s1νs)+AM(ηs,ηs))u+O(s2).\cos\alpha_{s,s^{2}u}-\cos\alpha_{s}=-\sin\alpha_{s}\tfrac{\partial u}{\partial\nu_{s}}+s(-\cos\alpha_{s}A(s^{-1}\nu_{s},s^{-1}\nu_{s})+A_{\partial M}(\eta_{s}{,}\eta_{s}))u+O(s^{2}).

And

s1(cosβs,s2ucosβs)=suτ1ηs/τcosγ¯s/τ,s2u/τ+O(s2)s^{-1}(\cos\beta_{s,s^{2}u}-\cos\beta_{s})=su\tau^{-1}\nabla_{\eta_{s/\tau}}\cos\bar{\gamma}_{s/\tau,s^{2}u/\tau}+O(s^{2})

Since each Σt\Sigma_{t} is stable capillary minimal surface under the metric g¯\bar{g}, so we know that

1sinγ¯ηtcosγ¯=cosγ¯A(νt,νt)+AM(ηt,ηt).\tfrac{1}{\sin\bar{\gamma}}\nabla_{\eta_{t}}\cos\bar{\gamma}=-\cos\bar{\gamma}A(\nu_{t},\nu_{t})+A_{\partial M}(\eta_{t}{,}\eta_{t}).

Based on the above asymptotic analysis and Lemmas 3.10 and 3.11, we see

(3.26) lims0[s1(cosαs,s2ucosαs)s1(cosβs,s2ucosβs)]=uν0\lim_{s\to 0}[s^{-1}(\cos\alpha_{s,s^{2}u}-\cos\alpha_{s})-s^{-1}(\cos\beta_{s,s^{2}u}-\cos\beta_{s})]=-\tfrac{\partial u}{\partial\nu_{0}}

on D0\partial D_{0} where ν0\nu_{0} is the outward normal of D0\partial D_{0} in D0D_{0}. By the elliptic strong maximum principle, the operator

(Δ0f,ν0):C2,α(D0)C1,α(D¯0)C0,α(D0)×C0,α(D0)(-\Delta_{0}-f,-\tfrac{\partial}{\partial\nu_{0}}):C^{2,\alpha}(D_{0})\cap C^{1,\alpha}(\bar{D}_{0})\to C^{0,\alpha}(D_{0})\times C^{0,\alpha}(\partial D_{0})

is an isomorphism since f0f\leqslant 0 in D0D_{0} due to Lemma 3.9 and τ1\tau\gneqq 1. In other words, we can specify the limits

(3.27) lims0[(Hs,s2uζs,s2u)(Hsζs)]\displaystyle\lim_{s\to 0}[(H_{s,s^{2}u}-\zeta_{s,s^{2}u})-(H_{s}-\zeta_{s})]
(3.28)  and lims0[s1(cosαs,s2ucosαs)s1(cosβs,s2ucosβs)]\displaystyle\quad\text{ and }\lim_{s\to 0}[s^{-1}(\cos\alpha_{s,s^{2}u}-\cos\alpha_{s})-s^{-1}(\cos\beta_{s,s^{2}u}-\cos\beta_{s})]

by choosing a suitable uC2,α(D0)C1,α(D¯0)u\in C^{2,\alpha}(D_{0})\cap C^{1,\alpha}(\bar{D}_{0}).

We have these facts: by Lemma 3.10, both αs\alpha_{s} and βs\beta_{s} tend to π/2\pi/2 as s0s\to 0, so lims0s1(αsβs)\lim_{s\to 0}s^{-1}(\alpha_{s}-\beta_{s}) is a function on D0\partial D_{0}; Hsζs=(22τ)s1+O(1)H_{s}-\zeta_{s}=(2-2\tau)s^{-1}+O(1);

(3.29)  Hs,s2uζs,s2u=(22τ)s1+O(1)\text{ }H_{s,s^{2}u}-\zeta_{s,s^{2}u}=(2-2\tau)s^{-1}+O(1)

for small s>0s>0 with a remainder term depending on uu. Hence, we can specify a function uu to counter-effect the O(1)O(1) remainder term in HsζsH_{s}-\zeta_{s} and make the remainder term in (3.29) strictly negative. That is, we can specify a function uu such that

(3.30) lims0(Hs,s2uζs,s2u(22τ)s1)\displaystyle\lim_{s\to 0}(H_{s,s^{2}u}-\zeta_{s,s^{2}u}-(2-2\tau)s^{-1}) =u0 in D0,\displaystyle=u_{0}\text{ in }D_{0},
(3.31) lims0s1(cosαs,s2ucosβs,s2u)\displaystyle\quad\lim_{s\to 0}s^{-1}(\cos\alpha_{s,s^{2}u}-\cos\beta_{s,s^{2}u}) <0 along D0,\displaystyle<0\text{ along }\partial D_{0},

for some negative function u0C0,α(D¯0)u_{0}\in C^{0,\alpha}(\bar{D}_{0}). Recall the definitions of ζ\zeta, τ\tau, β\beta, and by continuity, there exists a surface ΣM\Sigma_{-}\subset M satisfying

Hh¯<0 in Σ and α>γ¯ along Σ.H-\bar{h}<0\text{ in }\Sigma_{-}\text{ and }\alpha>\bar{\gamma}\text{ along }\partial\Sigma_{-}.

This surface Σ\Sigma_{-} is a lower barrier in the sense of Definition 1.5.

Now we can prove item (3) of Theorem 1.2.

Proof of item (3) of Theorem 1.2.

Assume that gg and g¯\bar{g} do not have isometric tangent cone at p0p_{0}, then we can construct a barrier Σ\Sigma_{-} such that Hh¯<0H-\bar{h}<0 in Σ\Sigma_{-} and the angle α>γ¯\alpha>\bar{\gamma} along Σ\partial\Sigma_{-}. But due to item (4) of Theorem 1.2, this is not possible. So gg and g¯\bar{g} have isometric tangent cones at p0p_{0}, then by the construction of the foliation in Theorem 3.3, again we have a barrier near t=0t=0, but the barrier condition is now not strict. We can extend the rigidity g=g¯g=\bar{g} in item (4) of Theorem 1.2 beyond the barrier and to all of MM. ∎

Remark 3.13.

By considering only the mean curvature, this provides an alternative proof of Theorem 1.1 in dimension 3. Moreover, we allow conical metrics of (𝕊3,g)(\mathbb{S}^{3},g) at two antipodal points.

Remark 3.14.

During the construction of barriers in the case of non-isometric cones, the Gauss-Bonnet theorem is only used in Lemma 3.7.

4. Construction of barriers (II)

In this section, we prove items (2) and (1) of Theorem 1.2. Our method is similar to the previous work [CW23].

4.1. Proof of item (2) of Theorem 1.2

For convenience, we set t=0t_{-}=0. We will construct a lower barrier near t=0t=0. As before, for any t>0t>0, we set Σt\Sigma_{t} to be the tt-level set of tt and Ωt\Omega_{t} to be the tt-sublevel set, that is, all points of MM which lie below Σt\Sigma_{t}. We see from the assumption on the tangent cone at pp_{-} that the sequence {(t1M,t2g¯)}t>0\{(t^{-1}M,t^{-2}\bar{g})\}_{t>0} converges to some right circular cone C¯\bar{C} in 3\mathbb{R}^{3} equipped with a flat metric g3g_{\mathbb{R}^{3}} as t0t\to 0. Then {(t1M,t2g)}t>0\{(t^{-1}M,t^{-2}g)\}_{t>0} converges to the same cone C¯\bar{C} but with a different constant metric g0g_{0}. The cone (C¯,g0)(\bar{C},g_{0}) can be represented as a cone in the Euclidean space (3,g3)(\mathbb{R}^{3},g_{\mathbb{R}^{3}}).

We have the existence of a barrier if (M,g)(M,g) and (M,g¯)(M,\bar{g}) have non-isometric tangent cones at p0p_{0}.

Lemma 4.1.

Let MM be given as in item (2) of Theorem 1.2. If the tangent cones of (M,g)(M,g) and (M,g¯)(M,\bar{g}) at pp_{-} are not isometric, assume that the mean curvature comparison and the metric comparison hold near pp_{-}, then there exists a surface Σ\Sigma_{-} satisfying

Hh¯<0 in Σ and γΣ>γ¯ along ΣH-\bar{h}<0\text{ in }\Sigma_{-}\text{ and }\gamma_{\Sigma_{-}}>\bar{\gamma}\text{ along }\partial\Sigma_{-}

as the above. This surface Σ\Sigma_{-} is a barrier in the sense of Definition 1.5.

Proof.

First, we note that the mean curvature comparison and the metric comparison (we only need boundary metric comparison) are preserved in the limits. Because that the tangent cone of (M,g¯)(M,\bar{g}) at pp_{-} is a round (solid) circular cone and that (t1Σt,t2g¯)(t^{-1}\Sigma_{t},t^{-2}\bar{g}) converges to its axial section as t0t\to 0, it follows from the angle comparison of [CW23, Proposition 4.9] that there exists a plane PP in C¯\bar{C} such that the the dihedral angles formed by C¯\partial\bar{C} and PP in the metric g0g_{0} are everywhere larger than γ¯(t)\bar{\gamma}(t_{-}).

We gain a lot of freedom to construct the barrier from the strict comparison of angles. The rest of the argument is analogous to [CW23, Proposition 4.10]. ∎

Remark 4.2.

Note that the scalar curvature comparison is not needed here.

Proof of item (2) of Theorem 1.2.

First, the tangent cones of (M,g)(M,g) and (M,g¯)(M,\bar{g}) at pp_{-} must be isometric. Indeed, by Lemma 4.1 and item (4) of Theorem 1.2, the barrier constructed in Lemma 4.1 cannot have Hh¯<0H-\bar{h}<0 in Σ\Sigma_{-} and γΣ<γ¯\gamma_{\Sigma_{-}}<\bar{\gamma} hold strictly along Σ\partial\Sigma_{-}.

By following Subsection 3.2, we can construct graphical perturbations Σt,t2u\Sigma_{t,t^{2}u} of Σt\Sigma_{t} which satisfy Proposition 3.3. For every sufficiently small t>0t>0, Σt,t2u\Sigma_{t,t^{2}u}  is a barrier in the sense of Definition 1.5, we conclude that g=g¯g=\bar{g} for the region bounded by Σt,t2u\Sigma_{t,t^{2}u} and P+MP_{+}\cap\partial M for every t>0t>0 from item (4) of Theorem 1.2. Hence, we finish the proof of item (2) of Theorem 1.2. ∎

4.2. Proof of item (1) of Theorem 1.2

This part is a slightly extension of the argument in Section 5 in our previous paper [CW23] and Ko-Yao’s paper [KY24]. So we only sketch the key steps here and refer to the above papers for more details.

Again, we set s=0s_{-}=0 and we construct a lower barrier near s=0s=0. Suppose MM is given by

M={(s,p)[0,ε)×S2:sf(p)},M=\{(s,p)\in[0,\varepsilon)\times{S}^{2}:s\geq f(p)\},

near p=Op_{-}=O, where ff is a smooth function such that f(p)=0f(p_{-})=0 and Hessf\mathrm{Hess}f is positive definite at pp_{-} under metric gS2g_{S^{2}}. Note that we need to assume ψ(0)0\psi(0)\neq 0, otherwise, the manifold MM will have a cusp at point O=(0,0,0)O=(0,0,0). For simplicity, we assume ψ(t(0))=1\psi(t(0))=1, where t(s)t(s) is determined by (1.5).

To better illustrate the situation, we can choose the coordinate (x1,x2)(x_{1},x_{2}) on S2{S}^{2} such that the expansion of metric g¯\bar{g} in (1.5) at OO is given by

g¯=ds2+dx12+dx22+O(s)+O(|x|2),\bar{g}=ds^{2}+dx_{1}^{2}+dx_{2}^{2}+O(s)+O(|x|^{2}),

where |x|=x12+x22|x|=\sqrt{x_{1}^{2}+x_{2}^{2}}.

For simplicity, we denote g¯0=ds2+dx12+dx22\bar{g}_{0}=ds^{2}+dx_{1}^{2}+dx_{2}^{2} as the linearised part of g¯\bar{g} at OO. After a suitable rotation, we can write g=g0+sh+O(s2)g=g_{0}+sh+O(s^{2}) for some constant metric g0g_{0} defined as

(4.1) g0=a33ds2+(a11dx12+a22dx22)+2a13dx1ds+2a23dx2ds,g_{0}=a_{33}ds^{2}+(a_{11}dx_{1}^{2}+a_{22}dx_{2}^{2})+2a_{13}dx_{1}ds+2a_{23}dx_{2}ds,

where the matrix

[a110a130a22a23a13a23a33]\begin{bmatrix}a_{11}&0&a_{13}\\ 0&a_{22}&a_{23}\\ a_{13}&a_{23}&a_{33}\end{bmatrix}

is positive definite and satisfies a11,a22,a331a_{11},a_{22},a_{33}\geq 1.

We assume the manifold MM can be written as

M={(s,x1,x2):s[0,ε),sζ(x1,x2)},M=\{(s,x_{1},x_{2}):s\in[0,\varepsilon),s\leq\zeta(x_{1},x_{2})\},

where ζ(x1,x2)=c11x12+2c12x1x2+c22x22+O(|x|3)\zeta(x_{1},x_{2})=c_{11}x_{1}^{2}+2c_{12}x_{1}x_{2}+c_{22}x_{2}^{2}+O(|x|^{3}) is a smooth function with c11,c22,c11c22c122>0c_{11},c_{22},c_{11}c_{22}-c_{12}^{2}>0. Here, we have used the fact that Hessf\mathrm{Hess}f is positive definite at pp_{-} under metric gS2g_{S^{2}}.

We write aija^{ij} as the inverse matrix of aija_{ij}, and define several constants as

B=\displaystyle B={} a33((a11c22+a22c11)2+(a11a22)2c122)\displaystyle\sqrt{a^{33}((\sqrt{a_{11}}c_{22}+\sqrt{a_{22}}c_{11})^{2}+(\sqrt{a_{11}}-\sqrt{a_{22}})^{2}c_{12}^{2})}
b11=\displaystyle b_{11}={} a33B1c111(a11(c11c22c122)+a11a22(c112+c122))\displaystyle a^{33}B^{-1}c_{11}^{-1}(a_{11}(c_{11}c_{22}-c_{12}^{2})+\sqrt{a_{11}a_{22}}(c_{11}^{2}+c_{12}^{2}))
b12=\displaystyle b_{12}={} b21=a33B1a11a22(c11+c22)\displaystyle b_{21}=a^{33}B^{-1}\sqrt{a_{11}a_{22}}(c_{11}+c_{22})
b22=\displaystyle b_{22}={} a33B1c221(a22(c11c22c122)+a11a22(c122+c222)).\displaystyle a^{33}B^{-1}c_{22}^{-1}(a_{22}(c_{11}c_{22}-c_{12}^{2})+\sqrt{a_{11}a_{22}}(c_{12}^{2}+c_{22}^{2})).

and consider the function Gλ,sG_{\lambda,s} defined by

Gλ,s(x1,x2)=\displaystyle G_{\lambda,s}(x_{1},x_{2})={} c11(b11(1+λ)1)x12+c22(b22(1+λ)1)x22\displaystyle c_{11}(b_{11}(1+\lambda)-1)x_{1}^{2}+c_{22}(b_{22}(1+\lambda)-1)x_{2}^{2}
+2c12(b12(1+λ)1)x1x2s2.\displaystyle+2c_{12}(b_{12}(1+\lambda)-1)x_{1}x_{2}-s^{2}.

and the surface Σλ,s\Sigma_{\lambda,s} is defined by

Σλ,s={(Gλ,s(x),x):x2 and Gλ,s(x)ζ(|x|2)}.\Sigma_{\lambda,s}=\left\{(G_{\lambda,s}(x),x):x\in\mathbb{R}^{2}\text{ and }G_{\lambda,s}(x)\geq\zeta(|x|^{2})\right\}.

We use an ellipse EλE_{\lambda} to parameterize Σλ,s\Sigma_{\lambda,s} where Eλ2E_{\lambda}\subset\mathbb{R}^{2} is given by

Eλ:={x^2:c11b11x^12+c22b22x^22+2c12b12x^1x^2<11+λ}.E_{\lambda}:=\{\hat{x}\in\mathbb{R}^{2}:c_{11}b_{11}\hat{x}_{1}^{2}+c_{22}b_{22}\hat{x}_{2}^{2}+2c_{12}b_{12}\hat{x}_{1}\hat{x}_{2}<\frac{1}{1+\lambda}\}.

Then, the surface Σλ,s\Sigma_{\lambda,s} can be written as a map EλΣλ,sE_{\lambda}\rightarrow\Sigma_{\lambda,s} such that

Σλ,s(x^):=(Gλ,s(Φλ,s(x^)),Φλ,s(x^))\Sigma_{\lambda,s}(\hat{x}):=(G_{\lambda,s}(\Phi_{\lambda,s}(\hat{x})),\Phi_{\lambda,s}(\hat{x}))

where Φλ,s:Eλ2\Phi_{\lambda,s}:E_{\lambda}\rightarrow\mathbb{R}^{2} satisfies

Φλ,s(x^)=sx^+O(s3).\Phi_{\lambda,s}(\hat{x})=s\hat{x}+O(s^{3}).

We also use Σs=Σ0,s\Sigma_{s}=\Sigma_{0,s} for short. We have the following result by the argument in [CW23].

Refer to caption
Σs\Sigma_{s^{\prime}}
Refer to caption
Σλ,s\Sigma_{\lambda,s}
Σs\Sigma_{s}
Refer to caption
Σs,u\Sigma_{s^{\prime},u}
Figure 4.1. Construction of Σλ,s\Sigma_{\lambda,s} and Σs,u\Sigma_{s,u}.
Proposition 4.3.

Suppose the metric gg can be written as g=g0+sh+O(s2)g=g_{0}+sh+O(s^{2}) where g0g_{0} is the constant metric defined in (4.1) and hh is a bounded symmetric two-tensor. Then, we have

cosγλ,s(x^)=\displaystyle\cos\gamma_{\lambda,s}(\hat{x})={} cosγ¯λ,s(x^)4λs2[(x^1c11b11+x^2c12b12)2a11a33+(x^1c12b12+x^2c22b22)2a22a33]\displaystyle\cos\bar{\gamma}_{\lambda,s}(\hat{x})-4\lambda s^{2}\left[\frac{(\hat{x}_{1}c_{11}b_{11}+\hat{x}_{2}c_{12}b_{12})^{2}}{a_{11}a^{33}}+\frac{(\hat{x}_{1}c_{12}b_{12}+\hat{x}_{2}c_{22}b_{22})^{2}}{a_{22}a^{33}}\right]
+s2O(λ2)+A(x^)s3+L(h)s3+O(s4).\displaystyle+s^{2}O(\lambda^{2})+A(\hat{x})s^{3}+L(h)s^{3}+O(s^{4}).

for any x^Eλ\hat{x}\in E_{\lambda}. Here, A(x^)A(\hat{x}) is a bounded term (not related to ss and hh) which is also odd symmetric with respect to x^\hat{x}, L(h)L(h) is a bounded term (not related to ss) relying on hh linearly.

Sketch of the proof.

We use the same argument for Proposition 5.1 in [KY24] and also track the term s3s^{3} to get the following expansions

cosg0(Σλ,s,M)=\displaystyle\cos\measuredangle_{g_{0}}(\Sigma_{\lambda,s},\partial M)={} 12(x^βcαβbαβ(1+λ))2aααa33s2+A(x^)s3+O(s4)\displaystyle 1-\frac{2(\hat{x}_{\beta}c_{\alpha\beta}b_{\alpha\beta}(1+\lambda))^{2}}{a_{\alpha\alpha}a^{33}}s^{2}+A(\hat{x})s^{3}+O(s^{4})
cosγ¯λ,s(x^)=\displaystyle\cos\bar{\gamma}_{\lambda,s}(\hat{x})={} 12s2[(x^1c11+x^2c12)2+(x^1c12+x^2c22)2]+A(x^)s3+O(s4),\displaystyle 1-2s^{2}\left[(\hat{x}_{1}c_{11}+\hat{x}_{2}c_{12})^{2}+(\hat{x}_{1}c_{12}+\hat{x}_{2}c_{22})^{2}\right]+A(\hat{x})s^{3}+O(s^{4}),

where we assume c21=c12c_{21}=c_{12}. Together with the remaining computation for Proposition 5.1 in [KY24] and the Corollary 5.5 in [CW23] (see the proof for Corollary 5.17 in [CW23]), we can establish the result. ∎

As a corollary of Proposition 4.3, we can easily establish the following results for sinγs\sin\gamma_{s},

(4.2) sinγs(x^)=\displaystyle\sin\gamma_{s}(\hat{x})={} sinγ¯s(x^)+O(s2)\displaystyle\sin\bar{\gamma}_{s}(\hat{x})+O(s^{2})
(4.3) =\displaystyle={} 2s(x^1c11+x^2c12)2+(x^1c12+x^2c22)2+O(s2),\displaystyle 2s\sqrt{(\hat{x}_{1}c_{11}+\hat{x}_{2}c_{12})^{2}+(\hat{x}_{1}c_{12}+\hat{x}_{2}c_{22})^{2}}+O(s^{2}),

and the following proposition.

Proposition 4.4.

Suppose the conditions in Proposition 4.3 hold. Then, for any λ>0\lambda>0, we can find s0>0s_{0}>0 (might rely on λ\lambda) such that for any s<s0s<s_{0}, we have

(4.4) γλ,s(x^)>γ¯λ,s(x^)\gamma_{\lambda,s}(\hat{x})>\bar{\gamma}_{\lambda,s}(\hat{x})

for any x^Eλ\hat{x}\in\partial E_{\lambda}.

We need to analyze the asymptotic behavior of mean curvature. We define the following mean curvatures:

Hλ,sg(x^):=\displaystyle H_{\lambda,s}^{g}(\hat{x}):={} Mean curvature of Σλ,s at Σλ,s(x^) under metric g,\displaystyle\text{Mean curvature of $\Sigma_{\lambda,s}$ at $\Sigma_{\lambda,s}(\hat{x})$ under metric $g$},
Hλ,s,Mg(x^):=\displaystyle H_{\lambda,s,\partial M}^{g}(\hat{x}):={} Mean curvature of M at (φ(|Φλ,s(x^)|2),Φλ,s(x^)) under metric g.\displaystyle\text{Mean curvature of $\partial M$ at $(\varphi(|\Phi_{\lambda,s}(\hat{x})|^{2}),\Phi_{\lambda,s}(\hat{x}))$ under metric $g$}.

Using the same computation for Corollary 5.2 in [KY24], we have

Proposition 4.5.

Suppose the metric gg can be written as g=g0+sh+O(s2)g=g_{0}+sh+O(s^{2}) where g0g_{0} is a constant metric defined in (4.1), and hh is a bounded symmetric two-tensor. Then, we have the following formula for the behavior of mean curvature

(4.5) Hsg(x^)=Hs,Mg(x^)Hs,Mg¯(x^)2(c11+c22)+2Ba11a22a33+sL(x^)+O(s2),H^{g}_{s}(\hat{x})=H^{g}_{s,\partial M}(\hat{x})-H^{\bar{g}}_{s,\partial M}(\hat{x})-2(c_{11}+c_{22})+\frac{2B}{\sqrt{a_{11}a_{22}a^{33}}}+sL(\hat{x})+O(s^{2}),

for any x^E\hat{x}\in E. Here, we write Hsg=H0,sgH^{g}_{s}=H^{g}_{0,s} and Hs,Mg=H0,s,MgH^{g}_{s,\partial M}=H^{g}_{0,s,\partial M} for short.

Now, we consider

H0:=lims0Hsg(x^),H_{0}:=\lim_{s\rightarrow 0}H_{s}^{g}(\hat{x}),

which is well-defined by (4.5) (the limit does not depend on the choice of x^\hat{x}.)

We have two subcases to consider.

If H0<h¯(0)H_{0}<\bar{h}(0), then we can use the continuation of Hλ,sgH_{\lambda,s}^{g} with respect to λ\lambda and ss, together with Proposition 4.4, we can show the following results (cf. Proposition 5.10 in [CW23]).

Proposition 4.6.

Suppose the metric gg can be written as g=g0+sh+O(s2)g=g_{0}+sh+O(s^{2}) where g0g_{0} is a constant metric defined in (4.1), and hh is a bounded symmetric two-tensor. If H0<h¯(0)H_{0}<\bar{h}(0), we can choose some λ>0,s>0\lambda>0,s>0 small such that Hλ,sg(x^)>h¯(Σλ,s(x^))H_{\lambda,s}^{g}(\hat{x})>\bar{h}(\Sigma_{\lambda,s}(\hat{x})) for any x^Eλ\hat{x}\in\partial E_{\lambda} and γλ,s(x^)<γ¯λ,s(x^)\gamma_{\lambda,s}(\hat{x})<\bar{\gamma}_{\lambda,s}(\hat{x}) for each x^Eλ\hat{x}\in\partial E_{\lambda}.

Now, we focus on the case H0=h¯(0)H_{0}=\bar{h}(0). In particular, it implies a11=a22=1a_{11}=a_{22}=1 and HMg(O)=HMg¯(O)H^{g}_{\partial M}(O)=H^{\bar{g}}_{\partial M}(O).

Then, we need to construct a foliation near OO. We define the vector field Ys(x^):=sΣs(x^)Y_{s}(\hat{x}):=\frac{\partial}{\partial s}\Sigma_{s}(\hat{x}). Given uC1,α(E¯)C2,α(E)u\in C^{1,\alpha}(\bar{E})\cap C^{2,\alpha}(E) where E=E0E=E_{0}, we can define the perturbation surface Σs,u\Sigma_{s,u} by

Σs,u:={Σs+uYs(x^),Ns(x^)(x^):x^E}\Sigma_{s,u}:=\left\{\Sigma_{s+\frac{u}{\left<Y_{s}(\hat{x}),N_{s}(\hat{x})\right>}}(\hat{x}):\hat{x}\in E\right\}

where Ns(x^)N_{s}(\hat{x}) is the unit normal vector field of Σs\Sigma_{s}.

We write E=E0E=E_{0}. Using the variational formula for mean curvatures and contact angles, we have

Hs,s3uh¯s,s3us=\displaystyle\frac{H_{s,s^{3}u}-\bar{h}_{s,s^{3}u}}{s}={} ΔsEu+Hsh¯ss+O(s),\displaystyle-\Delta_{s}^{E}u+\frac{H_{s}-\bar{h}_{s}}{s}+O(s),
cosγs,s3ucosγ¯s,s3us3=\displaystyle\frac{\cos\gamma_{s,s^{3}u}-\cos\bar{\gamma}_{s,s^{3}u}}{s^{3}}={} 2s(x^1c11+x^2c12)2+(x^1c12+x^2c22)2uνsE\displaystyle-2s\sqrt{(\hat{x}_{1}c_{11}+\hat{x}_{2}c_{12})^{2}+(\hat{x}_{1}c_{12}+\hat{x}_{2}c_{22})^{2}}\frac{\partial u}{\partial\nu_{s}^{E}}
+(AM(ηs,ηs)cosγsA(νs,νs)\displaystyle+(A_{\partial M}(\eta_{s},\eta_{s})-\cos\gamma_{s}A(\nu_{s},\nu_{s})
A¯M(η¯s,η¯s))u+cosγscosγ¯ss3+O(s),\displaystyle-\bar{A}_{\partial M}(\bar{\eta}_{s},\bar{\eta}_{s}))u+\frac{\cos\gamma_{s}-\cos\bar{\gamma}_{s}}{s^{3}}+O(s),

where ΔsE\Delta_{s}^{E} denotes the Laplacian-Beltrami operator on EE under the metric 1s2Σs(g)\frac{1}{s^{2}}\Sigma^{*}_{s}(g), and νsE\nu_{s}^{E} is the unit normal vector field of E\partial E under the metric 1s2Σs(g)\frac{1}{s^{2}}\Sigma^{*}_{s}(g). Here, we have used (4.3).

By using the same argument for Proposition 5.27 in [CW23], together with the asymptotic behavior of mean curvature, for each s(0,ε)s\in(0,\varepsilon) sufficiently small, we can find us()=u(,s)u_{s}(\cdot)=u(\cdot,s) such that the mean curvature Hs,s3usH_{s,s^{3}u_{s}} is h¯s,s3u+sλ(s)\bar{h}_{s,s^{3}u}+s\lambda(s) where λ(s)\lambda(s) is a function only depends on ss, the contact angle γs,s3us=γ¯s,s3us\gamma_{s,s^{3}u_{s}}=\bar{\gamma}_{s,s^{3}u_{s}}, and uu satisfies the following

lims0(u(x^,s)+u(x^,s))=0\lim_{s\rightarrow 0}(u(\hat{x},s)+u(-\hat{x},s))=0

for any x^E\hat{x}\in E. A finer analysis of λ(s)\lambda(s) will give λ(s)<0\lambda(s)<0 for ss sufficiently small (cf. Proposition 5.28 in [CW23]), and it leads to the following.

Proposition 4.7.

We can construct a surface Σ\Sigma_{-} near OO such that the mean curvature of Σ\Sigma_{-} is not greater than h¯\bar{h} and it has prescribed contact angle γ¯\bar{\gamma} with M\partial M.

Proof of item (1) of Theorem 1.2.

We can use Proposition 4.6 or Proposition 4.7 depending on the value of H0H_{0} to construct a barrier surface Σ\Sigma_{-} with mean curvature not greater than h¯\bar{h} and prescribed contact angle γ¯\bar{\gamma} with M\partial M. Then, we can use item (4) of Theorem 1.2 to extend the rigidity to all of MM. ∎

References

  • [Amb15] Lucas C. Ambrozio. Rigidity of area-minimizing free boundary surfaces in mean convex three-manifolds. J. Geom. Anal., 25(2):1001–1017, 2015.
  • [BBHW24] Christian Bär, Simon Brendle, Bernhard Hanke, and Yipeng Wang. Scalar curvature rigidity of warped product metrics. SIGMA Symmetry Integrability Geom. Methods Appl., 20:Paper No. 035, 26, 2024.
  • [BBI01] Dmitri Burago, Yuri Burago, and Sergei Ivanov. A course in Metric Geometry, volume 33 of Graduate Studies in Mathematics. American Mathematical Society, Providence, Rhode Island, 2001.
  • [BBN10] Hubert Bray, Simon Brendle, and Andre Neves. Rigidity of area-minimizing two-spheres in three-manifolds. Comm. Anal. Geom., 18(4):821–830, 2010.
  • [BM11] Simon Brendle and Fernando C. Marcques. Scalar Curvature Rigidity of Geodesic Balls in sns^{n}. Journal of Differential Geometry, 88(3):379–394, 2011.
  • [CLZ24] Jianchun Chu, Man-Chun Lee, and Jintian Zhu. Llarull’s theorem on punctured sphere with L{L}\infty metric. arXiv: 2405.19724, 2024.
  • [CW23] Xiaoxiang Chai and Gaoming Wang. Scalar curvature comparison of rotationally symmetric sets. arXiv/2304.13152, 2023.
  • [CW24] Xiaoxiang Chai and Xueyuan Wan. Scalar curvature rigidity of domains in a warped product. arXiv:2407.10212 [math], 2024.
  • [CWXZ24] Simone Cecchini, Jinmin Wang, Zhizhang Xie, and Bo Zhu. Scalar curvature rigidity of the four-dimensional sphere. arXiv: 2402.12633v2, 2024.
  • [CZ24] Simone Cecchini and Rudolf Zeidler. Scalar and mean curvature comparison via the Dirac operator. Geom. Topol., 28(3):1167–1212, 2024.
  • [DPM15] G. De Philippis and F. Maggi. Regularity of free boundaries in anisotropic capillarity problems and the validity of Young’s law. Arch. Ration. Mech. Anal., 216(2):473–568, 2015.
  • [FCS80] Doris Fischer-Colbrie and Richard Schoen. The structure of complete stable minimal surfaces in 3-manifolds of non-negative scalar curvature. Communications on Pure and Applied Mathematics, 33(2):199–211, 1980.
  • [Gro21] Misha Gromov. Four Lectures on Scalar Curvature. arXiv:1908.10612 [math], 2021.
  • [HKKZ] Sven Hirsch, Demetre Kazaras, Marcus Khuri, and Yiyue Zhang. Rigid comparison geometry for riemannian bands and open incomplete manifolds. (arXiv:2209.12857).
  • [HLS23] Yuhao Hu, Peng Liu, and Yuguang Shi. Rigidity of 3D spherical caps via μ\mu-bubbles. Pacific J. Math., 323(1):89–114, 2023.
  • [KY24] Dongyeong Ko and Xuan Yao. Scalar curvature comparison and rigidity of 33-dimensional weakly convex domains. arXiv:2410.20548, 2024.
  • [Li20] Chao Li. A polyhedron comparison theorem for 3-manifolds with positive scalar curvature. Invent. Math., 219(1):1–37, 2020.
  • [Lis10] Mario Listing. Scalar curvature on compact symmetric spaces. arXiv:1007.1832 [math], 2010.
  • [Lla98] Marcelo Llarull. Sharp estimates and the Dirac operator. Math. Ann., 310(1):55–71, 1998.
  • [Lot21] John Lott. Index theory for scalar curvature on manifolds with boundary. Proc. Amer. Math. Soc., 149(10):4451–4459, 2021.
  • [MO89] Maung Min-Oo. Scalar curvature rigidity of asymptotically hyperbolic spin manifolds. Math. Ann., 285(4):527–539, 1989.
  • [MP21] Pengzi Miao and Annachiara Piubello. Mass and Riemannian Polyhedra. arXiv:2101.02693 [gr-qc], 2021.
  • [RS97] Antonio Ros and Rabah Souam. On stability of capillary surfaces in a ball. Pacific J. Math., 178(2):345–361, 1997.
  • [SY79a] R. Schoen and Shing Tung Yau. Existence of incompressible minimal surfaces and the topology of three-dimensional manifolds with nonnegative scalar curvature. Ann. of Math. (2), 110(1):127–142, 1979.
  • [SY79b] Richard Schoen and Shing Tung Yau. On the proof of the positive mass conjecture in general relativity. Comm. Math. Phys., 65(1):45–76, 1979.
  • [WWZ24] Jinmin Wang, Zhichao Wang, and Bo Zhu. Scalar-mean rigidity theorem for compact manifolds with boundary. arXiv: 2409.14503, 2024.
  • [WX23a] Jinmin Wang and Zhizhang Xie. Dihedral rigidity for submanifolds of warped product manifolds. arXiv: 2303.13492, 2023.
  • [WX23b] Jinmin Wang and Zhizhang Xie. Scalar curvature rigidity of degenerate warped product spaces. arXiv: 2306.05413, 2023.
  • [Ye91] Rugang Ye. Foliation by constant mean curvature spheres. Pacific Journal of Mathematics, 147(2):381–396, 1991.