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

Low index capillary minimal surfaces
in Riemannian 33-manifolds

Eduardo Longa Departamento de Matemática, Instituto de Matemática e Estatística, Universidade de São Paulo, R. do Matão 1010, São Paulo, SP 05508-900, Brazil [email protected]
Abstract.

We prove a local rigidity result for infinitesimally rigid capillary surfaces in some Riemannian 33-manifolds with mean convex boundary. We also derive bounds on the genus, number of boundary components and area of any compact two-sided capillary minimal surface with low index under certain assumptions on the curvature of the ambient manifold and of its boundary.

Key words and phrases:
Minimal surface, capillary surface, rigidity, scalar curvature
2010 Mathematics Subject Classification:
53A10, 53C42
The author was partially supported by grant 2017/22704-0, São Paulo Research Foundation (FAPESP).

1. Introduction

Minimal surfaces are critical points for the area functional under suitable constraints. If we only consider closed surfaces, no constraints are necessary. If the surfaces have a fixed boundary, this leads to the so called Plateau problem, first studied by Lagrange [11] and Meusnier [14]. However, there is a third situation that can be investigated: when we consider compact surfaces whose boundaries are allowed to move freely in the boundary of the ambient manifold. Not surprisingly, critical points of the area functional under this constraint are called free boundary minimal surfaces. The boundary of such surfaces meet the boundary of the ambient manifold at a 9090^{\circ} angle.

Although the first works dealing with this type of surfaces date back to 1938 with R. Courant (see [8] and [9]), in the last decade there have been incredible developments in this field, with the employment of new techniques and the emergence of interesting conceptual links. Among the main contributors to this topic of research we could cite Fraser, Chen and Pang, with their work on free boundary surfaces on positively curved ambients [6], Ambrozio with his work on rigidity of mean-convex manifolds [2], and additionally Carlotto and Sharp (jointly with Ambrozio), with their works on compactness analysis and index estimates for free boundary minimal hypersurfaces ([3] and [4]).

In this paper we are interested in a natural generalisation of free boundary minimal surfaces, namely, capillary minimal surfaces. These are critical points of a certain energy functional, which will be presented in Section 2. As will be deduced later, they can be characterised as minimal surfaces whose boundary meet the ambient boundary at a constant angle — the capillary angle.

Capillary surfaces in 3\mathbb{R}^{3} model the configuration of liquids in containers in the absence of gravity. In fact, the interface between the fluid and the air is a surface with boundary that (locally) minimises the energy functional. This energy depends on the area of the interface, the area wetted by the fluid in the container and the angle of contact between the surface and the boundary of the container. More general situations have also been considered in the literature, like the influence of gravity and the density of the fluid in the equilibrium shape. We refer the interested reader to the book of Finn [10] for an extensive survey on this subject and the derivation of the equations that describe such surfaces.

Like in the free boundary case, questions relating the topology and the geometry of surfaces raise a lot of attention from geometers. For instance, given an ambient manifold (M3,g)(M^{3},g) of a particular shape, what are the possible topological types of surfaces Σ\Sigma that admit a capillary CMC or minimal embedding into MM? Is it possible to characterise the geometry of the allowed types? The first result in this direction was obtained by Nitsche [15], who proved that any immersed capillary disc in the unit ball of 3\mathbb{R}^{3} must be either a spherical cap or a flat disc. Later, Ros and Souam [18] extended this result to capillary discs in balls of 33-dimensional space forms. Recently, Wang and Xia [21] analysed the problem in an arbitrary dimension and proved that any stable immersed capillary hypersurface in a ball in space forms is totally umbilical. There are many interesting uniqueness results in other types of domains, like slabs [1], wedges [16, 7], cylinders [13] and cones [17].

We are interested in the same questions raised above, but in a more general setting. Namely, we only impose curvature assumptions on the ambient 33-manifold and look for restrictions in the topology of the possible immersed (or embedded) capillary minimal surfaces.

When dealing with the free boundary case, Ambrozio introduced the following functional in the space of compact and properly immersed surfaces in a Riemannian 33-manifold MM:

I(Σ)=12infMRM|Σ|+infMHM|Σ|,\displaystyle\operatorname{I}(\Sigma)=\frac{1}{2}\inf_{M}R_{M}|\Sigma|+\inf_{\partial M}H_{\partial M}|\partial\Sigma|,

where RMR_{M} is the scalar curvature of MM, HMH_{\partial M} is the mean curvature of M\partial M, |Σ||\Sigma| denotes the area of Σ\Sigma and |Σ||\partial\Sigma| denotes the length of Σ\partial\Sigma.

As a first result, we modify Ambrozio’s functional to take care of the capillary case and we show the following theorem, generalising Proposition 6 in [2]:

Theorem A.

Let (M3,g)(M^{3},g) be a Riemannian 33-manifold with nonempty boundary, and assume that RMR_{M} and HMH_{\partial M} are bounded from below. If Σ\Sigma is a compact two sided capillary stable minimal surface, immersed in MM with contact angle θ(0,π)\theta\in(0,\pi), then

Iθ(Σ):=12infMRM|Σ|+1sinθinfMHM|Σ|2πχ(Σ),\displaystyle\operatorname{I}_{\theta}(\Sigma):=\frac{1}{2}\inf_{M}R_{M}|\Sigma|+\frac{1}{\sin\theta}\inf_{\partial M}H_{\partial M}|\partial\Sigma|\leq 2\pi\chi(\Sigma),

where χ(Σ)\chi(\Sigma) denotes the Euler characteristic of Σ\Sigma. Moreover, equality occurs if and only if Σ\Sigma satisfies the following properties:

  • (i)

    Σ\Sigma is totally geodesic in MM and the geodesic curvature of Σ\partial\Sigma in M\partial M is equal to (cotθ)infHM(\cot\theta)\inf H_{\partial M};

  • (ii)

    the scalar curvature RMR_{M} is constant along Σ\Sigma and equal to infRM\inf R_{M}, and the mean curvature HMH_{\partial M} is constant along Σ\partial\Sigma and equal to infHM\inf H_{\partial M};

  • (iii)

    Ric(N)=0\operatorname{Ric}(N)=0 and II(ν¯,ν¯)=0\operatorname{\mathrm{I\!I}}(\overline{\nu},\overline{\nu})=0, where NN is a unit normal for Σ\Sigma and II(ν¯,ν¯)\operatorname{\mathrm{I\!I}}(\overline{\nu},\overline{\nu}) denotes the second fundamental form of M\partial M in the direction of a unit conormal for Σ\partial\Sigma in M\partial M.

In particular, (i), (ii) and (iii) imply that Σ\Sigma has constant Gaussian curvature equal to infRM/2\inf R_{M}/2 and Σ\partial\Sigma has constant geodesic curvature equal to infHM/sinθ\inf H_{\partial M}/\sin\theta.

A compact two-sided capillary minimal surface, properly embedded in (M,g)(M,g) with contact angle θ(0,π)\theta\in(0,\pi), that satisfies conditions (i), (ii) and (iii) of Theorem A will be called infinitesimally rigid. Given one such surface Σ\Sigma, there is a way to obtain a vector field ZZ in Σ\Sigma such that g(Z,N)=1g(Z,N)=1 and ZZ is tangent to M\partial M along Σ\partial\Sigma. Let us also denote by ZZ an extension to MM which is tangent to M\partial M along the entire boundary of MM. Let ϕ=ϕ(x,t)\phi=\phi(x,t) the local flow of ZZ and fix a number α\alpha between 0 and 11. We show the existence of a local foliation around Σ\Sigma, employing the same techniques as Ambrozio in Proposition 10 in [2]:

Theorem B.

Let (M3,g)(M^{3},g) be a Riemannian 33-manifold with nonempty boundary, and assume that RMR_{M} and HMH_{\partial M} are bounded from below. Let Σ\Sigma be a compact two-sided capillary minimal surface, properly embedded in MM with contact angle θ(0,π)\theta\in(0,\pi). If Σ\Sigma is infinitesimally rigid, then there exists ε>0\varepsilon>0 and a map wC2,α(Σ×(ε,ε))w\in C^{2,\alpha}(\Sigma\times(-\varepsilon,\varepsilon)) such that for every t(ε,ε)t\in(-\varepsilon,\varepsilon), the set

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

is a capillary CMC surface with contact angle θ\theta and mean curvature H(t)H(t). Moreover, for each xΣx\in\Sigma and t(ε,ε)t\in(-\varepsilon,\varepsilon),

w(x,0)=0,Σ(w(,t)t)dA=0,andwt(x,0)=1.\displaystyle w(x,0)=0,\quad\int_{\Sigma}(w(\cdot,t)-t)\,\mathrm{d}A=0,\quad\text{and}\quad\frac{\partial w}{\partial t}(x,0)=1.

In particular, taking a smaller ε>0\varepsilon>0 if necessary, {Σt}t(ε,ε)\{\Sigma_{t}\}_{t\in(-\varepsilon,\varepsilon)} is a capillary CMC foliation of a neighbourhood of Σ\Sigma in MM.

It is also possible to show the existence of a capillary minimal foliation, where the contact angles now vary from leaf to leaf. Although it will not be used subsequently, we believe it may be of independent interest (see Remark 3.4).

Theorem C.

Let (M3,g)(M^{3},g) be a Riemannian 33-manifold with nonempty boundary, and assume that RMR_{M} and HMH_{\partial M} are bounded from below. Let Σ\Sigma be a compact two-sided capillary minimal surface, properly embedded in MM with contact angle θ(0,π)\theta\in(0,\pi). If Σ\Sigma is infinitesimally rigid, then there exists ε>0\varepsilon>0 and a map wC2,α(Σ×(ε,ε))w\in C^{2,\alpha}(\Sigma\times(-\varepsilon,\varepsilon)) such that for every t(ε,ε)t\in(-\varepsilon,\varepsilon), the set

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

is a capillary minimal surface with contact angle θ(t)(0,π)\theta(t)\in(0,\pi). Moreover, for each xΣx\in\Sigma and t(ε,ε)t\in(-\varepsilon,\varepsilon),

w(x,0)=0,Σ(w(,t)t)dA=0,andwt(x,0)=1.\displaystyle w(x,0)=0,\quad\int_{\Sigma}(w(\cdot,t)-t)\,\mathrm{d}A=0,\quad\text{and}\quad\frac{\partial w}{\partial t}(x,0)=1.

In particular, taking a smaller ε>0\varepsilon>0 if necessary, {Σt}t(ε,ε)\{\Sigma_{t}\}_{t\in(-\varepsilon,\varepsilon)} is a capillary minimal foliation of a neighbourhood of Σ\Sigma in MM.

Then, we use Theorem B to show that, under some hypotheses, a dichotomy occurs: either the contact angle is equal to π/2\pi/2 or a very special situation takes place. More precisely, we have:

Theorem D.

Let (M3,g)(M^{3},g) be a Riemannian 33-manifold with nonempty and weakly mean-convex boundary, and assume that RMR_{M} is bounded from below. Let Σ\Sigma be an energy-minimising and infinitesimally rigid surface, properly embedded in MM with contact angle θ(0,π)\theta\in(0,\pi). Assume that one of the following hypothesis holds:

  • (a)

    each component of Σ\partial\Sigma is locally length-minimising in M\partial M; or

  • (b)

    infMHM=0\inf_{\partial M}H_{\partial M}=0.

Then either θ=π/2\theta=\pi/2 or Σ\Sigma is a flat and totally geodesic cylinder, MM is flat and M\partial M is totally geodesic around Σ\Sigma. In the first case, there is a neighbourhood of Σ\Sigma in MM that is isometric to (Σ×(ε,ε),gΣ+dt2)(\Sigma\times(-\varepsilon,\varepsilon),g_{\Sigma}+\mathrm{d}t^{2}), where (Σ,gΣ)(\Sigma,g_{\Sigma}) has constant Gaussian curvature infRM/2\inf R_{M}/2 and Σ\partial\Sigma has constant geodesic curvature infHM\inf H_{\partial M} in Σ\Sigma.

A situation when θπ/2\theta\neq\pi/2 may happen in Theorem D is the following. Let P1P_{1} and P2P_{2} be two non parallel planes in 3\mathbb{R}^{3}, intersecting along a line \ell, and let QQ be a plane in 3\mathbb{R}^{3} which is parallel to \ell and intersects both P1P_{1} and P2P_{2} at the same angle. Fix S0S_{0} to be the (closed) wedge determined by P1P_{1} and P2P_{2} which contains both QP1Q\cap P_{1} and QP2Q\cap P_{2}, and let S=S0S=S_{0}\setminus\ell (see Figure 1). Now fix TT a translation of 3\mathbb{R}^{3} by a vector parallel to the line \ell and let MM be the quotient of SS by the group GG generated by TT. If we define Σ\Sigma to be the quotient of QSQ\cap S by GG, then Σ\Sigma is an infinitesimally rigid cylinder intersecting M\partial M at a constant angle, (a) and (b) hold, MM is flat and M\partial M is totally geodesic, as we wanted. One question remains: is Σ\Sigma energy-minimising? We believe so, but we did not find a proof.

Refer to caption
Figure 1. An arrangement of 33 planes in 3\mathbb{R}^{3} seen from a plane orthogonal to the line \ell.

Next, we prove two other inequalities relating the geometry and the topology of capillary minimal surfaces of low index. This result generalises Theorem 1.2 in [6] to the capillary case. We note that item (i) below also generalises the free boundary case.

Theorem E.

Let (M3,g)(M^{3},g) be a Riemannian 33-manifold with nonempty boundary. Suppose that Σ\Sigma is a compact orientable two-sided capillary minimal surface of genus gg and with k1k\geq 1 boundary components, immersed in MM with contact angle θ(0,π)\theta\in(0,\pi).

  • (i)

    Suppose that (M,g)(M,g) has nonnegative Ricci curvature and weakly mean-convex boundary. If Σ\Sigma has index 11 then

    ΣkgdL<2π[9(1)g2(g+k)].\displaystyle\int_{\partial\Sigma}k_{g}\,\mathrm{d}L<2\pi\left[9-(-1)^{g}-2(g+k)\right].

    In particular, if the total geodesic curvature of Σ\partial\Sigma (in Σ\Sigma) is nonnegative (which happens if Σ\Sigma is free boundary and M\partial M is weakly convex, for instance), then

    • (a)

      g+k3g+k\leq 3 if gg is even;

    • (b)

      g+k4g+k\leq 4 if gg is odd.

  • (ii)

    Suppose that the scalar curvature of (M,g)(M,g) and the mean curvature of M\partial M are bounded from below. If Σ\Sigma has index 11, then

    Iθ(Σ)=12infMRM|Σ|+1sinθinfMHM|Σ|<2π[7(1)gk].\displaystyle\operatorname{I}_{\theta}(\Sigma)=\frac{1}{2}\inf_{M}R_{M}|\Sigma|+\frac{1}{\sin\theta}\inf_{\partial M}H_{\partial M}|\partial\Sigma|<2\pi\left[7-(-1)^{g}-k\right].
  • (iii)

    Suppose that (M,g)(M,g) has scalar curvature RMR0>0R_{M}\geq R_{0}>0 and weakly mean-convex boundary.

    • (a)

      If Σ\Sigma is stable, then it is a disc and |Σ|4πR0|\Sigma|\leq\frac{4\pi}{R_{0}}.

    • (b)

      If Σ\Sigma has index 11, then |Σ|4π[7(1)gk]R0|\Sigma|\leq\frac{4\pi\left[7-(-1)^{g}-k\right]}{R_{0}}.

Remark 1.1.

The scalar curvature in [6] is one half of ours. This is why a factor 4π4\pi appears in item (iii) of Theorem E instead of the 2π2\pi of Theorem 1.2 in [6].

Acknowledgements

The author would like to thank Paolo Piccione for his constant support and encouragement during the period when this article was written and revised, and for countless fruitful mathematical conversations. He also expresses sincere gratitude to Lucas Ambrozio for valuable comments on this work. Additionally, the author thanks Izabella Freitas and Jackeline Conrado for the figures in this paper.

2. Preliminaries: variational problem and stability

The purpose of this section is to formally introduce the concept of capillary CMC and minimal hypersurfaces. Despite the fact that in section 3 we will be dealing only with capillary surfaces in Riemannian 33-manifolds, there is no significant simplification in introducing the main concepts only in dimension 22. So, the general situation will be addresed in the sequel.

Let (Mn+1,g)(M^{n+1},g) be a Riemannian manifold with nonempty boundary. Let Σn\Sigma^{n} be a smooth compact manifold with nonempty boundary, and let φ:ΣM\varphi:\Sigma\to M be a smooth immersion of Σ\Sigma into MM. We say that φ\varphi is a proper immersion if φ(Σ)M=φ(Σ)\varphi(\Sigma)\cap\partial M=\varphi(\partial\Sigma).

Henceforth, we assume that φ\varphi is two-sided. Fix a unit normal vector field NN for Σ\Sigma along φ\varphi and denote by ν\nu the outward unit conormal for Σ\partial\Sigma in Σ\Sigma. Moreover, let N¯\overline{N} the outward pointing unit normal for M\partial M and let ν¯\overline{\nu} the unit normal for Σ\partial\Sigma in M\partial M such that the bases {N,ν}\{N,\nu\} and {N¯,ν¯}\{\overline{N},\overline{\nu}\} determine the same orientation in (TΣ)(T\partial\Sigma)^{\perp}. See Figure 2 to gain some intuition.

A smooth function Φ:Σ×(ε,ε)M\Phi:\Sigma\times(-\varepsilon,\varepsilon)\to M is called a proper variation of φ\varphi is the maps φt:ΣM\varphi_{t}:\Sigma\to M, defined by φt(x)=Φ(x,t)\varphi_{t}(x)=\Phi(x,t), are proper immersions for all t(ε,ε)t\in(-\varepsilon,\varepsilon), and if φ0=φ\varphi_{0}=\varphi.

Let us fix a proper variation Φ\Phi of φ\varphi. The variational vector field associated to Φ\Phi is the vector field ξΦ=ξ:ΣTM\xi_{\Phi}=\xi:\Sigma\to TM along φ\varphi defined by

ξ(x)=Φt(x,0),xΣ.\displaystyle\xi(x)=\frac{\partial\Phi}{\partial t}(x,0),\quad x\in\Sigma.

We now define some important functionals related to the variation Φ\Phi. The area functional A:(ε,ε)A:(-\varepsilon,\varepsilon)\to\mathbb{R} is given by

A(t)=ΣdAφtg,\displaystyle A(t)=\int_{\Sigma}\mathrm{d}A_{\varphi_{t}^{*}g},

where dAφtg\mathrm{d}A_{\varphi_{t}^{*}g} denotes the area element of (Σ,φtg)(\Sigma,\varphi_{t}^{\ast}g). Even if n=dimΣ>2n=\dim\Sigma>2, it is customary to refer to this as the area functional.

The volume functional V:(ε,ε)V:(-\varepsilon,\varepsilon)\to\mathbb{R} is defined by

V(t)=Σ×[0,t]Φ(dV),\displaystyle V(t)=\int_{\Sigma\times[0,t]}\Phi^{*}(\mathrm{d}V),

where dV\mathrm{d}V is the volume element of MM. We say that the variation Φ\Phi is volume preserving if V(t)=0V(t)=0 for every t(ε,ε)t\in(-\varepsilon,\varepsilon).

We also consider the wetting area functional W:(ε,ε)W:(-\varepsilon,\varepsilon)\to\mathbb{R}:

W(t)=Σ×[0,t]Φ(dAM),\displaystyle W(t)=\int_{\partial\Sigma\times[0,t]}\Phi^{*}(\mathrm{d}A_{\partial M}),

where dAM\mathrm{d}A_{\partial M} denotes the area element of M\partial M.

Finally, we define the energy functional. In order to do so, let us fix an angle θ(0,π)\theta\in(0,\pi). Then EΦ,θ=E:(ε,ε)E_{\Phi,\theta}=E:(-\varepsilon,\varepsilon)\to\mathbb{R} is given by

E(t)=A(t)(cosθ)W(t).\displaystyle E(t)=A(t)-(\cos\theta)W(t).

The following proposition contains the formulae for the first variation of the energy and volume, whose proof can be found in [20] and [5]:

Proposition 2.1.

Let Φ\Phi be a proper variation of the immersion φ:ΣM\varphi:\Sigma\to M. Then the following formulae hold:

E(0)\displaystyle E^{\prime}(0) =ΣHfdA+Σg(ξ,ν(cosθ)ν¯)dL and\displaystyle=-\int_{\Sigma}Hf\,\mathrm{d}A+\int_{\partial\Sigma}g(\xi,\nu-(\cos\theta)\overline{\nu})\,\mathrm{d}L\quad\text{ and }
V(0)\displaystyle V^{\prime}(0) =ΣfdA,\displaystyle=\int_{\Sigma}f\,\mathrm{d}A,

where f=g(ξ,N)f=g(\xi,N), HH is the mean curvature of Σ\Sigma with respect to NN, dA\mathrm{d}A is the area element of Σ\Sigma induced by φ\varphi and dL\mathrm{d}L is the line element of Σ\partial\Sigma induced by φ\varphi.

We say that the immersion φ\varphi is a capillary CMC immersion if E(0)=0E^{\prime}(0)=0 for every volume preserving variation of φ\varphi. If E(0)=0E^{\prime}(0)=0 for every variation of φ\varphi, we call φ\varphi a capillary minimal immersion. When there is only one immersion under consideration and there is no risk of confusion, we just say that Σ\Sigma is a capillary CMC or minimal hypersurface. See Figure 2 for an example in an Euclidean cone.

Refer to caption
Figure 2. The four fundamental vector fields in a capillary minimal surface in a cone.

Notice that Σ\Sigma is a capillary CMC hypersurface if and only if Σ\Sigma has constant mean curvature and g(N,N¯)=cosθg(N,\overline{N})=\cos\theta along Σ\partial\Sigma; this last condition means that Σ\partial\Sigma meets M\partial M at an angle of θ\theta. Similarly, Σ\Sigma is a capillary minimal hypersurface when Σ\Sigma is a minimal hypersurface and Σ\partial\Sigma meets M\partial M at an angle of θ\theta. When θ=π/2\theta=\pi/2, we use the term free boundary CMC (or minimal) hypersurface.

For a capillary CMC or minimal hypersuface Σ\Sigma with contact angle θ(0,π)\theta\in(0,\pi), the orthonormal bases {N,ν}\{N,\nu\} and {N¯,ν¯}\{\overline{N},\overline{\nu}\} are related by the following equations:

{N¯=(cosθ)N+(sinθ)νν¯=(sinθ)N+(cosθ)ν\displaystyle\begin{cases}\overline{N}=(\cos\theta)N+(\sin\theta)\nu\\ \overline{\nu}=-(\sin\theta)N+(\cos\theta)\nu\end{cases}

This will be important in the calculations.

Proposition 2.2.

[1, Proposition 2.1] Let φ:ΣnMn+1\varphi:\Sigma^{n}\to M^{n+1} be a proper immersion which is transversal to M\partial M. For any smooth function f:Σf:\Sigma\to\mathbb{R} satisfying ΣfdA=0\int_{\Sigma}f\,\mathrm{d}A=0, there exists a volume preserving variation Φ\Phi of φ\varphi such that f=g(ξΦ,N)f=g(\xi_{\Phi},N), where ξ=Φt|t=0\xi=\frac{\partial\Phi}{\partial t}|_{t=0} is the variational vector field associated to Φ\Phi. If we don’t assume that ΣfdA=0\int_{\Sigma}f\,\mathrm{d}A=0, then the result still holds, but now Φ\Phi doesn’t need to be volume preserving.

We now discuss the notion of stability. Firstly, let us fix some notation. For a proper immersion φ:ΣM\varphi:\Sigma\to M, let A(X,Y)=g(XN,Y)A(X,Y)=g(-\nabla_{X}N,Y) the second fundamental form of Σ\Sigma, and denote by ΔΣ\Delta_{\Sigma} the Laplacian of Σ\Sigma with respect to the metric induced by φ\varphi. Moreover, let II(v,w)=g(vN¯,w)\operatorname{\mathrm{I\!I}}(v,w)=g(\nabla_{v}\overline{N},w) be the second fundamental form of M\partial M with respect to N¯-\overline{N}.

The following proposition gives a formula for the second variation of energy. The proof is long and can be found in the Appendix of [18].

Proposition 2.3.

Let φ:ΣM\varphi:\Sigma\to M be a capillary CMC immersion and let f:Σf:\Sigma\to\mathbb{R} be a smooth function which satisfies ΣfdA=0\int_{\Sigma}f\,\mathrm{d}A=0. Let Φ\Phi be a volume preserving proper variation of φ\varphi such that f=g(ξΦ,N)f=g(\xi_{\Phi},N) (see Proposition 2.2). Then,

E′′(0)\displaystyle E^{\prime\prime}(0) =Σ[ΔΣf+(Ric(N)+A2)f]fdA+Σ(fνqf)fdL\displaystyle=-\int_{\Sigma}\left[\Delta_{\Sigma}f+(\operatorname{Ric}(N)+\lVert A\rVert^{2})f\right]f\,\mathrm{d}A+\int_{\partial\Sigma}\left(\frac{\partial f}{\partial\nu}-qf\right)f\,\mathrm{d}L
=ΣfLΣ(f)dA+Σ(fνqf)fdL,\displaystyle=-\int_{\Sigma}fL_{\Sigma}(f)\,\mathrm{d}A+\int_{\partial\Sigma}\left(\frac{\partial f}{\partial\nu}-qf\right)f\,\mathrm{d}L,

where

q=1sinθII(ν¯,ν¯)+(cotθ)A(ν,ν)\displaystyle q=\frac{1}{\sin\theta}\operatorname{\mathrm{I\!I}}(\overline{\nu},\overline{\nu})+(\cot\theta)A(\nu,\nu)

and LΣ=ΔΣ+(Ric(N)+A2)L_{\Sigma}=\Delta_{\Sigma}+(\operatorname{Ric}(N)+\lVert A\rVert^{2}) is the Jacobi operator of Σ\Sigma. If φ\varphi is a capillary minimal immersion and ff is any smooth function defined on Σ\Sigma, then the same formula holds, but now Φ\Phi doesn’t need to be volume preserving.

The capillary CMC immersion φ:ΣM\varphi:\Sigma\to M (or just Σ\Sigma) is called stable if E′′(0)0E^{\prime\prime}(0)\geq 0 for any volume preserving variation of φ\varphi. If φ\varphi is a capillary minimal immersion, we call it stable whenever E′′(0)0E^{\prime\prime}(0)\geq 0 for every variation of φ\varphi.

Alternatively, let ={fH1(Σ):ΣfdA=0}\mathcal{F}=\{f\in H^{1}(\Sigma):\int_{\Sigma}f\,\mathrm{d}A=0\}, where H1(Σ)H^{1}(\Sigma) is the first Sobolev space of Σ\Sigma. The index form Q:H1(Σ)×H1(Σ)Q:H^{1}(\Sigma)\times H^{1}(\Sigma)\to\mathbb{R} of Σ\Sigma is given by

(1) Q(f,h)=Σ[g(f,h)(Ric(N)+A2)fh]dAΣqfhdL.\displaystyle Q(f,h)=\int_{\Sigma}\left[g(\nabla f,\nabla h)-(\operatorname{Ric}(N)+\lVert A\rVert^{2})fh\right]\,\mathrm{d}A-\int_{\partial\Sigma}qfh\,\mathrm{d}L.

Then φ\varphi is a capillary CMC stable immersion if and only if Q(f,f)0Q(f,f)\geq 0 for every ff\in\mathcal{F}. If φ\varphi is a capillary minimal immersion, then it is stable precisely when Q(f,f)0Q(f,f)\geq 0 for every fH1(Σ)f\in H^{1}(\Sigma).

The stability index of a capillary CMC (resp. minimal) hypersurface Σ\Sigma is the dimension of the largest vector subspace of \mathcal{F} (resp. H1(Σ)H^{1}(\Sigma)) restricted to which the bilinear form QQ is negative definite. Thus, stable hyperfurfaces are those which have index equal to zero. Intuitively, this means that no variation of a capillary minimal stable hypersurface decreases area up to second order.

It is possible to analitically compute the stability index of a capillary CMC (resp. minimal) hypersurface Σ\Sigma in terms of the spectrum of its Jacobi operator with a Robin boundary condition. Indeed, the boundary condition

fν=qf\displaystyle\frac{\partial f}{\partial\nu}=qf

is elliptic for the Jacobi operator LΣL_{\Sigma}. Therefore, there exists a nondecreasing and divergent sequence λ1λ2λk\lambda_{1}\leq\lambda_{2}\leq\cdots\leq\lambda_{k}\nearrow\infty of eigenvalues for the problem

(2) {LΣ(f)+λf=0on Σfν=qfon Σ\displaystyle\begin{cases}L_{\Sigma}(f)+\lambda f=0\quad&\text{on }\Sigma\\ \frac{\partial f}{\partial\nu}=qf\quad&\text{on }\partial\Sigma\end{cases}

The index of Σ\Sigma is then the number of negative eigenvalues of the problem above associated to (smooth) eigenfunctions that lie in \mathcal{F} (resp. H1(Σ)H^{1}(\Sigma)) (see more details in [6] and [19]).

3. Proof of the Theorems

This section is devoted to prove the Theorems mentioned in the Introduction.

Recall that, given a Riemannian 33-manifold (M3,g)(M^{3},g) with nonempty boundary and a fixed angle θ(0,π)\theta\in(0,\pi), we can define the following functional in the space of compact properly immeresd surfaces in MM:

Iθ(Σ)=12infMRM|Σ|+1sinθinfMHM|Σ|,\displaystyle\operatorname{I}_{\theta}(\Sigma)=\frac{1}{2}\inf_{M}R_{M}|\Sigma|+\frac{1}{\sin\theta}\inf_{\partial M}H_{\partial M}|\partial\Sigma|,

provided that the scalar curvature RMR_{M} and the mean curvature HMH_{\partial M} are bounded from below.

We first show that if Σ\Sigma is capillary minimal stable, then Iθ(Σ)\operatorname{I}_{\theta}(\Sigma) can be bounded above by the topology of Σ\Sigma. We also characterise the equality case.

We begin with a simple lemma.

Lemma 3.1.

Let (M3,g)(M^{3},g) be a Riemannian 33-manifold with nonempty boundary, and let Σ\Sigma be a two-sided capillary CMC surface embedded in MM with contact angle θ(0,π)\theta\in(0,\pi) and mean curvature equal to HH. Then:

II(ν¯,ν¯)+(cosθ)A(ν,ν)+(sinθ)kg=HM+Hcosθ\displaystyle\operatorname{\mathrm{I\!I}}(\overline{\nu},\overline{\nu})+(\cos\theta)A(\nu,\nu)+(\sin\theta)k_{g}=H_{\partial M}+H\cos\theta
Proof.

Let TT be a unit tangent vector field along Σ\partial\Sigma. Since Σ\Sigma is CMC, we know that A(ν,ν)=HA(T,T)A(\nu,\nu)=H-A(T,T). So,

(cosθ)A(ν,ν)+(sinθ)kg\displaystyle(\cos\theta)A(\nu,\nu)+(\sin\theta)k_{g} Hcosθ=(cosθ)A(T,T)+(sinθ)kg\displaystyle-H\cos\theta=-(\cos\theta)A(T,T)+(\sin\theta)k_{g}
=g(TT,(cosθ)N+(sinθ)ν)\displaystyle=g(-\nabla_{T}T,(\cos\theta)N+(\sin\theta)\nu)
=g(TT,N¯)\displaystyle=g(-\nabla_{T}T,\overline{N})
=II(T,T).\displaystyle=\operatorname{\mathrm{I\!I}}(T,T).

But {T,ν¯}\{T,\overline{\nu}\} is an orthonormal referential for T(M)T(\partial M) along Σ\partial\Sigma. Thus,

II(ν¯,ν¯)+(cosθ)A(ν,ν)+(sinθ)kg=II(ν¯,ν¯)+II(T,T)+Hcosθ=HM+Hcosθ,\displaystyle\operatorname{\mathrm{I\!I}}(\overline{\nu},\overline{\nu})+(\cos\theta)A(\nu,\nu)+(\sin\theta)k_{g}=\operatorname{\mathrm{I\!I}}(\overline{\nu},\overline{\nu})+\operatorname{\mathrm{I\!I}}(T,T)+H\cos\theta=H_{\partial M}+H\cos\theta,

as we wanted. ∎

Proof of Theorem A.

Since Σ\Sigma is capillary minimal stable, we know that Q(f,f)0Q(f,f)\geq 0 for any fH1(Σ)f\in H^{1}(\Sigma), where QQ is the index form given by (1). In particular, taking f1f\equiv 1 yields

(3) Σ(Ric(N)+A2)dA+ΣqdL0.\displaystyle\int_{\Sigma}(\operatorname{Ric}(N)+\lVert A\rVert^{2})\,\mathrm{d}A+\int_{\partial\Sigma}q\,\mathrm{d}L\leq 0.

By the Gauss equation, we have:

Ric(N)=12(RM+HΣ2A2)KΣ,\displaystyle\operatorname{Ric}(N)=\frac{1}{2}(R_{M}+H_{\Sigma}^{2}-\lVert A\rVert^{2})-K_{\Sigma},

where HΣ0H_{\Sigma}\equiv 0 and KΣK_{\Sigma} are the mean and Gaussian curvatures of Σ\Sigma. Plugging this into (3) and using Gauss-Bonnet theorem for the term involving KΣK_{\Sigma}, we obtain:

0\displaystyle 0 12Σ(RM+A2)dA(2πχ(Σ)ΣkgdL)+1sinθΣ[II(ν¯,ν¯)+(cotθ)A(ν,ν)]dL\displaystyle\geq\frac{1}{2}\int_{\Sigma}(R_{M}+\lVert A\rVert^{2})\,\mathrm{d}A-\left(2\pi\chi(\Sigma)-\int_{\partial\Sigma}k_{g}\,\mathrm{d}L\right)+\frac{1}{\sin\theta}\int_{\partial\Sigma}\left[\operatorname{\mathrm{I\!I}}(\overline{\nu},\overline{\nu})+(\cot\theta)A(\nu,\nu)\right]\,\mathrm{d}L
=12Σ(RM+A2)dA+1sinθΣ[II(ν¯,ν¯)+(cotθ)A(ν,ν)+(sinθ)kg]dL2πχ(Σ)\displaystyle=\frac{1}{2}\int_{\Sigma}(R_{M}+\lVert A\rVert^{2})\,\mathrm{d}A+\frac{1}{\sin\theta}\int_{\partial\Sigma}\left[\operatorname{\mathrm{I\!I}}(\overline{\nu},\overline{\nu})+(\cot\theta)A(\nu,\nu)+(\sin\theta)k_{g}\right]\,\mathrm{d}L-2\pi\chi(\Sigma)
12ΣRMdA+1sinθΣ[II(ν¯,ν¯)+(cotθ)A(ν,ν)+(sinθ)kg]dL2πχ(Σ).\displaystyle\geq\frac{1}{2}\int_{\Sigma}R_{M}\,\mathrm{d}A+\frac{1}{\sin\theta}\int_{\partial\Sigma}\left[\operatorname{\mathrm{I\!I}}(\overline{\nu},\overline{\nu})+(\cot\theta)A(\nu,\nu)+(\sin\theta)k_{g}\right]\,\mathrm{d}L-2\pi\chi(\Sigma).

Using Lemma 3.1 with H=0H=0, we have:

0\displaystyle 0 12ΣRMdA+1sinθΣHMdL2πχ(Σ)\displaystyle\geq\frac{1}{2}\int_{\Sigma}R_{M}\,\mathrm{d}A+\frac{1}{\sin\theta}\int_{\partial\Sigma}H_{\partial M}\,\mathrm{d}L-2\pi\chi(\Sigma)
12infMRM|Σ|+1sinθinfMHM|Σ|2πχ(Σ),\displaystyle\geq\frac{1}{2}\inf_{M}R_{M}|\Sigma|+\frac{1}{\sin\theta}\inf_{\partial M}H_{\partial M}|\partial\Sigma|-2\pi\chi(\Sigma),

proving the first assertion of the theorem.

Let us assume now that equality holds. It is immediate to see that Σ\Sigma must be totally geodesic, item (b) holds and Q(1,1)=0Q(1,1)=0. The latter implies that Q(1,f)=0Q(1,f)=0 for any fH1(Σ)f\in H^{1}(\Sigma). Indeed, if c>0c>0, then

0Q(1cf,1cf)=Q(1,1)2cQ(1,f)+c2Q(f,f)=2cQ(1,f)+c2Q(f,f).\displaystyle 0\leq Q(1-cf,1-cf)=Q(1,1)-2cQ(1,f)+c^{2}Q(f,f)=-2cQ(1,f)+c^{2}Q(f,f).

Dividing by cc and rearranging, we obtain

Q(1,f)c2Q(f,f).\displaystyle Q(1,f)\leq\frac{c}{2}Q(f,f).

Now we let c0c\searrow 0 and conclude that Q(1,f)0Q(1,f)\leq 0. The reverse inequality is analogous. This proves the claim. Thus, we know that

Q(1,f)=ΣRic(N)fdA+ΣqfdL=0\displaystyle Q(1,f)=\int_{\Sigma}\operatorname{Ric}(N)f\,\mathrm{d}A+\int_{\partial\Sigma}qf\,\mathrm{d}L=0

for any fH1(Σ)f\in H^{1}(\Sigma). This implies that Ric(N)=0\operatorname{Ric}(N)=0 and q=0q=0. Since we already know that A=0A=0, it follows that II(ν¯,ν¯)=0\operatorname{\mathrm{I\!I}}(\overline{\nu},\overline{\nu})=0, proving item (c).

It remains to calculate the geodesic curvature k¯g\overline{k}_{g} of Σ\Sigma in M\partial M. We have

k¯g:=g(TT,ν¯)=g(TT,(sinθ)N+(cosθ)ν)=(sinθ)A(T,T)+(cosθ)kg=(cosθ)kg.\displaystyle\overline{k}_{g}:=g(-\nabla_{T}T,\overline{\nu})=g(-\nabla_{T}T,-(\sin\theta)N+(\cos\theta)\nu)=-(\sin\theta)A(T,T)+(\cos\theta)k_{g}=(\cos\theta)k_{g}.

But Lemma 3.1 implies that (sinθ)kg=HM(\sin\theta)k_{g}=H_{\partial M}. So, k¯g=(cotθ)HM\overline{k}_{g}=(\cot\theta)H_{\partial M}, proving (a). The converse statement can be easily proved. ∎

We are amost ready to prove Theorem B. Recall that ZZ is a vector field in MM which is tangent to M\partial M along the entire boundary of MM and which satisfies g(Z,N)=1g(Z,N)=1 along Σ\Sigma. Let φ=φ(x,t)\varphi=\varphi(x,t) be the local flow of ZZ and fix α(0,1)\alpha\in(0,1).

We begin with a lemma, whose proof can be found in the appendices of [2] and [12].

Lemma 3.2.

Let (M3,g)(M^{3},g) be a Riemannian 33-manifold with nonempty boundary, and let φ:ΣM\varphi:\Sigma\to M a two-sided proper embedding of a compact surface Σ\Sigma with boundary. Fix a proper variation Φ:Σ×(ε,ε)M\Phi:\Sigma\times(-\varepsilon,\varepsilon)\to M of φ\varphi and let Σt=Φ(Σ×{t})\Sigma_{t}=\Phi(\Sigma\times\{t\}). We use the subscript tt to denote all terms related to Σt\Sigma_{t}, e.g. NtN_{t} is the unit normal, HtH_{t} is the mean curvature, etc. Let ρt=g(Nt,ξΦ,t)\rho_{t}=g(N_{t},\xi_{\Phi,t}) and let θt\theta_{t} be the contact angle between Σt\Sigma_{t} and M\partial M. Then the following formulae hold:

Htt\displaystyle\frac{\partial H_{t}}{\partial t} =LΣt(ρt)dHt(ξΦ,t)on Σt,\displaystyle=L_{\Sigma_{t}}(\rho_{t})-\mathrm{d}H_{t}(\xi_{\Phi,t}^{\top})\quad\text{on }\Sigma_{t},
(cosθt)νt\displaystyle\frac{\partial(\cos\theta_{t})}{\partial\nu_{t}} =(sinθt)ρtνt+(cosθt)At(νt,νt)ρt+II(ν¯t,ν¯t)ρt+dθt(Wt)on Σt,\displaystyle=-(\sin\theta_{t})\frac{\partial\rho_{t}}{\partial\nu_{t}}+(\cos\theta_{t})A_{t}(\nu_{t},\nu_{t})\rho_{t}+\operatorname{\mathrm{I\!I}}(\overline{\nu}_{t},\overline{\nu}_{t})\rho_{t}+\mathrm{d}\theta_{t}(W_{t})\quad\text{on }\partial\Sigma_{t},

where ()(\cdot)^{\top} denotes the orthogonal projection onto TΣtT\Sigma_{t}, LΣtL_{\Sigma_{t}} is the Jacobi operator of Σt\Sigma_{t} and WtW_{t} is a certain vector field along Σt\partial\Sigma_{t}. In particular, if each Σt\Sigma_{t} is capillary CMC with contact angle θ(0,π)\theta\in(0,\pi), then

(4) dHtdt\displaystyle\frac{\mathrm{d}H_{t}}{\mathrm{d}t} =LΣt(ρt)on Σt\displaystyle=L_{\Sigma_{t}}(\rho_{t})\quad\text{on }\Sigma_{t}
(5) ρtνt\displaystyle\frac{\partial\rho_{t}}{\partial\nu_{t}} =[1sinθII(ν¯t,ν¯t)+(cotθ)At(νt,νt)]ρton Σt\displaystyle=\left[\frac{1}{\sin\theta}\operatorname{\mathrm{I\!I}}(\overline{\nu}_{t},\overline{\nu}_{t})+(\cot\theta)A_{t}(\nu_{t},\nu_{t})\right]\rho_{t}\quad\text{on }\partial\Sigma_{t}
Proof of Theorem B.

Given a function uC2,α(Σ)u\in C^{2,\alpha}(\Sigma), let Σu={ϕ(x,u(x)):xΣ}\Sigma_{u}=\{\phi(x,u(x)):x\in\Sigma\}. If the norm of uu is small enough, then Σu\Sigma_{u} is a properly embedded surface in MM. As in the statement of Lemma 3.2 we use the subscript uu to denote the terms related to Σu\Sigma_{u}.

Now consider the Banach spaces

E\displaystyle E ={uC2,α(Σ):ΣudA=0},\displaystyle=\left\{u\in C^{2,\alpha}(\Sigma):\int_{\Sigma}u\,\mathrm{d}A=0\right\},
F\displaystyle F ={uC0,α(Σ):ΣudA=0}.\displaystyle=\left\{u\in C^{0,\alpha}(\Sigma):\int_{\Sigma}u\,\mathrm{d}A=0\right\}.

For small ε>0\varepsilon>0 and δ>0\delta>0, define the map Ψ:BE(0,δ)×(ε,ε)F×C1,α(Σ)\Psi:B_{E}(0,\delta)\times(-\varepsilon,\varepsilon)\to F\times C^{1,\alpha}(\partial\Sigma) by

Ψ(u,t)=(Hu+t1|Σ|ΣHu+tdA,g(Nu+t,N¯u+t)cosθ)\displaystyle\Psi(u,t)=\left(H_{u+t}-\frac{1}{|\Sigma|}\int_{\Sigma}H_{u+t}\,\mathrm{d}A,g(N_{u+t},\overline{N}_{u+t})-\cos\theta\right)

where BE(0,δ)B_{E}(0,\delta) denotes the open ball of radius δ\delta centered at the origin of EE. We claim that DΨ(0,0):E×D\Psi(0,0):E\times\mathbb{R} is an isomorphism when restricted to {0}×E\{0\}\times E. To prove this, let vEv\in E. Notice that the map Φ:Σ×(ε,ε)(x,s)ϕ(x,sv(x))M\Phi:\Sigma\times(-\varepsilon,\varepsilon)\ni(x,s)\mapsto\phi(x,sv(x))\in M is a proper variation of Σ\Sigma whose variational vector field is

ξΦ(x)=s|s=0ϕ(x,sv(x))=v(x)ϕs(x,0)=v(x)Z(x),xΣ.\displaystyle\xi_{\Phi}(x)=\left.{\frac{\partial}{\partial s}}\right|_{s=0}\phi(x,sv(x))=v(x)\frac{\partial\phi}{\partial s}(x,0)=v(x)Z(x),\quad x\in\Sigma.

So, using the formulae from Lemma 3.2 and the fact that the Jacobi operator of Σ\Sigma is just the Laplacian, since Σ\Sigma is infinitesimally rigid, we obtain:

(6) DΨ(0,0)(v,0)\displaystyle D\Psi(0,0)\cdot(v,0) =dds|s=0Ψ(sv,0)=(ΔΣv1|Σ|ΣΔΣvdA,(sinθ)vν)\displaystyle=\left.{\frac{\mathrm{d}}{\mathrm{d}s}}\right|_{s=0}\Psi(sv,0)=\left(\Delta_{\Sigma}v-\frac{1}{|\Sigma|}\int_{\Sigma}\Delta_{\Sigma}v\,\mathrm{d}A,-(\sin\theta)\frac{\partial v}{\partial\nu}\right)
(7) =(ΔΣv1|Σ|ΣvνdA,(sinθ)vν).\displaystyle=\left(\Delta_{\Sigma}v-\frac{1}{|\Sigma|}\int_{\partial\Sigma}\frac{\partial v}{\partial\nu}\,\mathrm{d}A,-(\sin\theta)\frac{\partial v}{\partial\nu}\right).

If this is equal to zero, then vv is harmonic and satisfies a Neumann condition on the boundary of Σ\Sigma. This implies that vv is constant, and this constant must be equal to zero since vEv\in E, which proves that DΨ(0,0)D\Psi(0,0) is injective. The surjectivity follows from classical results for Neumann type boundary conditions for the Laplace operator.

The final step is to apply the implicit function theorem: for some smaller ε>0\varepsilon>0, there exists a function (ε,ε)tu(t)BE(0,δ)(-\varepsilon,\varepsilon)\ni t\mapsto u(t)\in B_{E}(0,\delta) such that u(0)=0u(0)=0 and Ψ(u(t),t)=(0,0)\Psi(u(t),t)=(0,0) for any t(ε,ε)t\in(-\varepsilon,\varepsilon). This means precisely that the surfaces

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

have constant mean curvature and meet M\partial M along their boundaries at an angle of θ\theta.

Now let w:Σ×(ε,ε)w:\Sigma\times(-\varepsilon,\varepsilon)\to\mathbb{R} be given by w(x,t)=u(t)(x)+tw(x,t)=u(t)(x)+t. Then w(x,0)=0w(x,0)=0 and w(,t)t=u()BE(0,δ)w(\cdot,t)-t=u(\cdot)\in B_{E}(0,\delta). Define G:Σ×(ε,ε)MG:\Sigma\times(-\varepsilon,\varepsilon)\to M by G(x,t)=ϕ(x,w(x,t))G(x,t)=\phi(x,w(x,t)) and observe that this is a proper variation of Σ\Sigma whose variational vector field is

ξG(x)=Gt(x,0)=wt(x,0)ϕt(x,0)=wt(x,0)Z(x),xΣ.\displaystyle\xi_{G}(x)=\frac{\partial G}{\partial t}(x,0)=\frac{\partial w}{\partial t}(x,0)\frac{\partial\phi}{\partial t}(x,0)=\frac{\partial w}{\partial t}(x,0)Z(x),\quad x\in\Sigma.

Since

0=Ψ(u(t),t)=(Hw(,t)1|Σ|ΣHw(,t)dA,g(Nw(,t),N¯w(,t))cosθ)\displaystyle 0=\Psi(u(t),t)=\left(H_{w(\cdot,t)}-\frac{1}{|\Sigma|}\int_{\Sigma}H_{w(\cdot,t)}\,\mathrm{d}A,g(N_{w(\cdot,t)},\overline{N}_{w(\cdot,t)})-\cos\theta\right)

for all t(ε,ε)t\in(-\varepsilon,\varepsilon), differentiating this equation with respect to tt implies that wt|t=0\left.\frac{\partial w}{\partial t}\right|_{t=0} satisfies the Neumann problem on Σ\Sigma (see equation (6)). Therefore, it must be constant. But

Σ(w(x,t)t)dA(x)=Σu(t)dA=0,t(ε,ε)\displaystyle\int_{\Sigma}(w(x,t)-t)\,\mathrm{d}A(x)=\int_{\Sigma}u(t)\,\mathrm{d}A=0,\quad t\in(-\varepsilon,\varepsilon)

since u(t)Eu(t)\in E. By differentiating this equation with respect to tt and evaluating it at t=0t=0 we obtain that Σwt|t=0dA=|Σ|\int_{\Sigma}\left.\frac{\partial w}{\partial t}\right|_{t=0}\,\mathrm{d}A=|\Sigma|. Thus, wt|t=0=1\left.\frac{\partial w}{\partial t}\right|_{t=0}=1 as we wanted to show.

Finally, since G(x,0)=ϕ(x,0)=xG(x,0)=\phi(x,0)=x for every xΣx\in\Sigma, ξG=Z\xi_{G}=Z is transverse to Σ\Sigma and Σ\Sigma is properly embedded, by taking a smaller ε>0\varepsilon>0 if necessary, we may assume that GG parametrises a foliation of a neighbourhood of Σ\Sigma in MM. This concludes the proof. ∎

The proof of Theorem C is similar to the proof of Theorem B. This way, we only sketch it, indicating the main differences.

Proof of Theorem C.

As in the previous proof, for uC2,α(Σ)u\in C^{2,\alpha}(\Sigma), let Σu={ϕ(x,u(x)):xΣ}\Sigma_{u}=\{\phi(x,u(x)):x\in\Sigma\}.

Now consider the Banach spaces

E\displaystyle E ={uC2,α(Σ):ΣudA=0},\displaystyle=\left\{u\in C^{2,\alpha}(\Sigma):\int_{\Sigma}u\,\mathrm{d}A=0\right\},
G\displaystyle G ={uC1,α(Σ):ΣudL=0}.\displaystyle=\left\{u\in C^{1,\alpha}(\partial\Sigma):\int_{\partial\Sigma}u\,\mathrm{d}L=0\right\}.

For small ε>0\varepsilon>0 and δ>0\delta>0, define the map Λ:BE(0,δ)×(ε,ε)G×C0,α(Σ)\Lambda:B_{E}(0,\delta)\times(-\varepsilon,\varepsilon)\to G\times C^{0,\alpha}(\Sigma) by

Λ(u,t)=(g(Nu+t,N¯u+t)1|Σ|Σg(Nu+t,N¯u+t)dL,Hu+t),\displaystyle\Lambda(u,t)=\left(g(N_{u+t},\overline{N}_{u+t})-\frac{1}{|\partial\Sigma|}\int_{\partial\Sigma}g(N_{u+t},\overline{N}_{u+t})\,\mathrm{d}L,H_{u+t}\right),

where BE(0,δ)B_{E}(0,\delta) denotes the open ball of radius δ\delta centered at the origin of EE. Then, using the formulae from Lemma 3.2 and the fact that the Jacobi operator of Σ\Sigma is just the Laplacian, since Σ\Sigma is infinitesimally rigid, we have:

(8) DΛ(0,0)(v,0)\displaystyle D\Lambda(0,0)\cdot(v,0) =dds|s=0Ψ(sv,0)=((sinθ)vν+sinθ|Σ|ΣvνdL,ΔΣv)\displaystyle=\left.{\frac{\mathrm{d}}{\mathrm{d}s}}\right|_{s=0}\Psi(sv,0)=\left(-(\sin\theta)\frac{\partial v}{\partial\nu}+\frac{\sin\theta}{|\partial\Sigma|}\int_{\partial\Sigma}\frac{\partial v}{\partial\nu}\,\mathrm{d}L,\Delta_{\Sigma}v\right)

for any vEv\in E. It is easy to show that DΛ(0,0)D\Lambda(0,0) is an isomorphism when restricted to E×{0}E\times\{0\}. Now, just apply the implicit function theorem and proceed as in the proof of Theorem B. ∎

Before proving Theorem D, let us clarify that a closed curve γ\gamma in M\partial M is locally length-minimising when every nearby closed curve in M\partial M has length greater than or equal to the length of γ\gamma.

Let us start with a lemma.

Lemma 3.3.

Let (M3,g)(M^{3},g) be a Riemannian 33-manifold with nonempty boundary and let φ:ΣM\varphi:\Sigma\to M be a two-sided capillary minimal proper immersion of a compact surface Σ\Sigma, with contact angle equal to θ(0,π)\theta\in(0,\pi). If φ\varphi is energy-minimising for the angle θ\theta, then, changing the sign of its unit normal, it is energy-minimising for the angle πθ\pi-\theta.

Proof.

Given a proper variation Φ:Σ×(ε,ε)M\Phi:\Sigma\times(-\varepsilon,\varepsilon)\to M of φ\varphi, define Φ~:Σ×(ε,ε)M\tilde{\Phi}:\Sigma\times(-\varepsilon,\varepsilon)\to M to be the variation given by Φ~(x,t)=Φ(x,t)\tilde{\Phi}(x,t)=\Phi(x,-t). Denote by WΦW_{\Phi} and WΦ~W_{\tilde{\Phi}} the wetting areas of Φ\Phi and Φ~\tilde{\Phi}, respectively. We claim the WΦ~(t)=WΦ(t)W_{\tilde{\Phi}}(t)=-W_{\Phi}(-t) for every t(ε,ε)t\in(-\varepsilon,\varepsilon). To prove this, define f:Σ×(ε,ε)f:\partial\Sigma\times(-\varepsilon,\varepsilon)\to\mathbb{R} by

f=Φ(dAM)(T,ds),\displaystyle f=\Phi^{*}(\mathrm{d}A_{\partial M})(T,\mathrm{d}s),

where TT is a unit tangent vector field along Σ\partial\Sigma and ds\mathrm{d}s is the standard vector field in \mathbb{R}. Then, for t>0t>0, we have

WΦ~(t)\displaystyle W_{\tilde{\Phi}}(t) =Σ×[0,t]f(x,s)dV(x,s)=Σ0tf(x,s)dsdx\displaystyle=\int_{\partial\Sigma\times[0,t]}f(x,-s)\,\mathrm{d}V(x,s)=\int_{\partial\Sigma}\int_{0}^{t}f(x,-s)\,\mathrm{d}s\,\mathrm{d}x
=Σ0tf(x,u)dudx\displaystyle=-\int_{\partial\Sigma}\int_{0}^{-t}f(x,u)\,\mathrm{d}u\,\mathrm{d}x
=Σ×[0,t]f(x,u)dV(x,u)\displaystyle=-\int_{\partial\Sigma\times[0,-t]}f(x,u)\,\mathrm{d}V(x,u)
=WΦ(t).\displaystyle=-W_{\Phi}(-t).

The computation is analogous for t<0t<0. This proves the claim. To prove the lemma, we show that EΦ~,πθ(t)E_{\tilde{\Phi},\pi-\theta}(t) is equal to EΦ,θ(t)E_{\Phi,\theta}(-t):

EΦ~,πθ(t)\displaystyle E_{\tilde{\Phi},\pi-\theta}(t) =A(Φ~(Σ,t))(cos(πθ))WΦ~(t)\displaystyle=A(\tilde{\Phi}(\Sigma,t))-(\cos(\pi-\theta))W_{\tilde{\Phi}}(t)
=A(Φ(Σ,t))(cosθ)WΦ(t)\displaystyle=A(\Phi(\Sigma,-t))-(\cos\theta)W_{\Phi}(-t)
=EΦ,θ(t).\displaystyle=E_{\Phi,\theta}(-t).

Thus, we have

EΦ~,πθ(t)=EΦ,θ(t)EΦ,θ(0)=EΦ~,πθ(0)\displaystyle E_{\tilde{\Phi},\pi-\theta}(t)=E_{\Phi,\theta}(-t)\geq E_{\Phi,\theta}(0)=E_{\tilde{\Phi},\pi-\theta}(0)

for any variation Φ\Phi and any tt, which implies that φ\varphi is energy-minimising for the angle πθ\pi-\theta. ∎

Proof of Theorem D.

Let GG be a parametrisation of the capillary CMC foliation {Σt}t(ε,ε)\{\Sigma_{t}\}_{t\in(-\varepsilon,\varepsilon)} given by Theorem B. For each t(ε,ε)t\in(-\varepsilon,\varepsilon), let ρt=g(G,Nt)\rho_{t}=g(\partial_{G},N_{t}), where as usual, we use the subscript tt for terms related to Σt\Sigma_{t}. Then, by equations (4) and (5) in Lemma 3.2, we have

(9) H(t)\displaystyle H^{\prime}(t) =Δtρt+(Ric(Nt)+At2)ρton Σt,\displaystyle=\Delta_{t}\rho_{t}+(\operatorname{Ric}(N_{t})+\lVert A_{t}\rVert^{2})\rho_{t}\quad\text{on }\Sigma_{t},
(10) ρtνt\displaystyle\frac{\partial\rho_{t}}{\partial\nu_{t}} =qtρt=[1sinθII(ν¯t,ν¯t)+(cotθ)At(νt,νt)]ρton Σt.\displaystyle=q_{t}\rho_{t}=\left[\frac{1}{\sin\theta}\operatorname{\mathrm{I\!I}}(\overline{\nu}_{t},\overline{\nu}_{t})+(\cot\theta)A_{t}(\nu_{t},\nu_{t})\right]\rho_{t}\quad\text{on }\partial\Sigma_{t}.

Notice that ρ0=g(Z,N)=1\rho_{0}=g(Z,N)=1, so we can assume ρt>0\rho_{t}>0 for any t(ε,ε)t\in(-\varepsilon,\varepsilon), maybe for some smaller ε>0\varepsilon>0. Using the Gauss equation, we rewrite equation (9) as

H(t)ρt1=ρt1Δtρt+12(RM,t+H(t)2+At2)Kt,\displaystyle H^{\prime}(t)\rho_{t}^{-1}=\rho_{t}^{-1}\Delta_{t}\rho_{t}+\frac{1}{2}(R_{M,t}+H(t)^{2}+\lVert A_{t}\rVert^{2})-K_{t},

where RM,tR_{M,t} denotes the scalar curvature of MM restricted to Σt\Sigma_{t} and KtK_{t} denotes the Gaussian curvature of Σt\Sigma_{t}. Now recall that H(t)H(t) is constant on Σt\Sigma_{t}, so, integration by parts gives

H(t)Σρt1dAt\displaystyle H^{\prime}(t)\int_{\Sigma}\rho_{t}^{-1}\,\mathrm{d}A_{t} =Σρt2ρt2dAt+Σρt1ρtνtdLt\displaystyle=\int_{\Sigma}\rho_{t}^{-2}\lVert\nabla\rho_{t}\rVert^{2}\,\mathrm{d}A_{t}+\int_{\partial\Sigma}\rho_{t}^{-1}\frac{\partial\rho_{t}}{\partial\nu_{t}}\,\mathrm{d}L_{t}
+12Σ(RM,t+H(t)2+At2)dAtΣKtdAt\displaystyle+\frac{1}{2}\int_{\Sigma}(R_{M,t}+H(t)^{2}+\lVert A_{t}\rVert^{2})\,\mathrm{d}A_{t}-\int_{\Sigma}K_{t}\,\mathrm{d}A_{t}
=Σρt2ρt2dAt+ΣqtdLt\displaystyle=\int_{\Sigma}\rho_{t}^{-2}\lVert\nabla\rho_{t}\rVert^{2}\,\mathrm{d}A_{t}+\int_{\partial\Sigma}q_{t}\,\mathrm{d}L_{t}
+12Σ(RM,t+H(t)2+At2)dAtΣKtdAt\displaystyle+\frac{1}{2}\int_{\Sigma}(R_{M,t}+H(t)^{2}+\lVert A_{t}\rVert^{2})\,\mathrm{d}A_{t}-\int_{\Sigma}K_{t}\,\mathrm{d}A_{t}
ΣqtdLt+12Σ(RM,t+H(t)2)dAtΣKtdAt.\displaystyle\geq\int_{\partial\Sigma}q_{t}\,\mathrm{d}L_{t}+\frac{1}{2}\int_{\Sigma}(R_{M,t}+H(t)^{2})\,\mathrm{d}A_{t}-\int_{\Sigma}K_{t}\,\mathrm{d}A_{t}.

We now apply Gauss-Bonnet Theorem and Lemma 3.1 to obtain

H(t)Σρt1dAt\displaystyle H^{\prime}(t)\int_{\Sigma}\rho_{t}^{-1}\,\mathrm{d}A_{t} Σ(qt+kg,t)dLt+12Σ(RM,t+H(t)2)dAt2πχ(Σt)\displaystyle\geq\int_{\partial\Sigma}(q_{t}+k_{g,t})\,\mathrm{d}L_{t}+\frac{1}{2}\int_{\Sigma}(R_{M,t}+H(t)^{2})\,\mathrm{d}A_{t}-2\pi\chi(\Sigma_{t})
=1sinθΣ(HM,t+H(t)cosθ)dLt+12Σ(RM,t+H(t)2)dAt2πχ(Σt)\displaystyle=\frac{1}{\sin\theta}\int_{\partial\Sigma}(H_{\partial M,t}+H(t)\cos\theta)\,\mathrm{d}L_{t}+\frac{1}{2}\int_{\Sigma}(R_{M,t}+H(t)^{2})\,\mathrm{d}A_{t}-2\pi\chi(\Sigma_{t})
12ΣRM,tdAt+1sinθΣHM,tdLt+H(t)|Σt|cotθ2πχ(Σt).\displaystyle\geq\frac{1}{2}\int_{\Sigma}R_{M,t}\,\mathrm{d}A_{t}+\frac{1}{\sin\theta}\int_{\partial\Sigma}H_{\partial M,t}\,\mathrm{d}L_{t}+H(t)|\partial\Sigma_{t}|\cot\theta-2\pi\chi(\Sigma_{t}).

Since each Σt\Sigma_{t} is homeomorphic to Σ\Sigma, χ(Σt)=χ(Σ)\chi(\Sigma_{t})=\chi(\Sigma). Next, we use the fact that Σ\Sigma is infinitesimally rigid to apply the identity Iθ(Σ)=2πχ(Σ)\operatorname{I}_{\theta}(\Sigma)=2\pi\chi(\Sigma):

H(t)Σρt1dAt\displaystyle H^{\prime}(t)\int_{\Sigma}\rho_{t}^{-1}\,\mathrm{d}A_{t} 12infMRM|Σt|+1sinθinfΣHM|Σt|+H(t)|Σt|cotθ2πχ(Σt)\displaystyle\geq\frac{1}{2}\inf_{M}R_{M}|\Sigma_{t}|+\frac{1}{\sin\theta}\inf_{\partial\Sigma}H_{\partial M}|\partial\Sigma_{t}|+H(t)|\partial\Sigma_{t}|\cot\theta-2\pi\chi(\Sigma_{t})
=12infMRM(|Σt||Σ|)+1sinθinfΣHM(|Σt||Σ|)+H(t)|Σt|cotθ.\displaystyle=\frac{1}{2}\inf_{M}R_{M}(|\Sigma_{t}|-|\Sigma|)+\frac{1}{\sin\theta}\inf_{\partial\Sigma}H_{\partial M}(|\partial\Sigma_{t}|-|\partial\Sigma|)+H(t)|\partial\Sigma_{t}|\cot\theta.

By hypothesis, infHM0\inf H_{\partial M}\geq 0. If each component of Σ\partial\Sigma is length-minimising, then the second term in the right hand side is nonnegative, and if infHM=0\inf H_{\partial M}=0, then it is equal to zero. In either case, we have

H(t)Σρt1dA\displaystyle H^{\prime}(t)\int_{\Sigma}\rho_{t}^{-1}\,\mathrm{d}A 12infMRM(|Σt||Σ|)+H(t)|Σt|cotθ\displaystyle\geq\frac{1}{2}\inf_{M}R_{M}(|\Sigma_{t}|-|\Sigma|)+H(t)|\partial\Sigma_{t}|\cot\theta
=12infMRM0tdds|Σs|ds+H(t)|Σt|cotθ.\displaystyle=\frac{1}{2}\inf_{M}R_{M}\int_{0}^{t}\frac{\mathrm{d}}{\mathrm{d}s}|\Sigma_{s}|\,\mathrm{d}s+H(t)|\partial\Sigma_{t}|\cot\theta.

By the first variation of area,

dds|Σs|\displaystyle\frac{\mathrm{d}}{\mathrm{d}s}|\Sigma_{s}| =ΣH(s)ρsdAs+Σg(νs,tG)dLs\displaystyle=-\int_{\Sigma}H(s)\rho_{s}\,\mathrm{d}A_{s}+\int_{\partial\Sigma}g(\nu_{s},\partial_{t}G)\,\mathrm{d}L_{s}
(11) =H(s)ΣρsdAscotθΣρsdLs.\displaystyle=-H(s)\int_{\Sigma}\rho_{s}\,\mathrm{d}A_{s}-\cot\theta\int_{\partial\Sigma}\rho_{s}\,\mathrm{d}L_{s}.

Thus, we obtain the following differential inequality:

H(t)Σρt1dA12infMRM0t[H(s)ΣρsdAs+cotθΣρsdLs]ds+H(t)|Σt|cotθ.\displaystyle H^{\prime}(t)\int_{\Sigma}\rho_{t}^{-1}\,\mathrm{d}A\geq-\frac{1}{2}\inf_{M}R_{M}\int_{0}^{t}\left[H(s)\int_{\Sigma}\rho_{s}\,\mathrm{d}A_{s}+\cot\theta\int_{\partial\Sigma}\rho_{s}\,\mathrm{d}L_{s}\right]\mathrm{d}s+H(t)|\partial\Sigma_{t}|\cot\theta.

Let us rewrite this as

(12) H(t)R0ψ(t)0tH(s)ξ(s)ds+cotθ[H(t)|Σt|+R0ψ(t)0tη(s)ds],\displaystyle H^{\prime}(t)\geq\frac{R_{0}}{\psi(t)}\int_{0}^{t}H(s)\xi(s)\,\mathrm{d}s+\cot\theta\left[H(t)|\partial\Sigma_{t}|+\frac{R_{0}}{\psi(t)}\int_{0}^{t}\eta(s)\,\mathrm{d}s\right],

where

  • R0=12infMRMR_{0}=-\frac{1}{2}\inf_{M}R_{M}

  • ψ(t)=Σρt1dAt\psi(t)=\int_{\Sigma}\rho_{t}^{-1}\,\mathrm{d}A_{t}

  • ξ(t)=ΣρtdAt\xi(t)=\int_{\Sigma}\rho_{t}\,\mathrm{d}A_{t}

  • η(t)=ΣρtdLt\eta(t)=\int_{\partial\Sigma}\rho_{t}\,\mathrm{d}L_{t}

It is time to apply Lemma 3.3. Changing the sign of the unit normal NN of Σ\Sigma in the begining of the proof, if necessary, we may assume that θ(0,π/2]\theta\in(0,\pi/2] if R00R_{0}\geq 0 and that θ[π/2,π)\theta\in[\pi/2,\pi) if R0<0R_{0}<0. In either case, inequality (12) implies that H(0)0H^{\prime}(0)\geq 0. If H(0)>0H^{\prime}(0)>0 then H(t)>0H(t)>0 for all small positive tt. The first variation formula for the energy (see Proposition 2.1) would then give that E(t)<E(0)E(t)<E(0):

E(t)E(0)=0tE(s)ds=0tH(s)(ΣρsdAs)ds<0,\displaystyle E(t)-E(0)=\int_{0}^{t}E^{\prime}(s)\,\mathrm{d}s=-\int_{0}^{t}H(s)\left(\int_{\Sigma}\rho_{s}\,\mathrm{d}A_{s}\right)\,\mathrm{d}s<0,

which is absurd, since Σ\Sigma is energy-minimising. So H(0)=0H^{\prime}(0)=0. Now let α(t)\alpha(t) denote the real function in the right hand side of inequality (12). A simple computation yields α(0)=η(0)ψ(0)R0cotθ\alpha^{\prime}(0)=\frac{\eta(0)}{\psi(0)}R_{0}\cot\theta. This is positive if and only if R00R_{0}\neq 0 and θπ/2\theta\neq\pi/2. In this case, we would have H′′(0)α(0)>0H^{\prime\prime}(0)\geq\alpha^{\prime}(0)>0, which is also absurd by the same reason. So either θ=π/2\theta=\pi/2, and we are done, or R0=0R_{0}=0.

Let us analyse the case R0=0R_{0}=0. Inequality (12) takes the form H(t)(cotθ)|Σt|H(t)H^{\prime}(t)\geq(\cot\theta)|\partial\Sigma_{t}|H(t), and since cotθ0\cot\theta\geq 0, this implies that H(t)0H(t)\geq 0 for t0t\geq 0 and H(t)0H(t)\leq 0 for t0t\leq 0. By the first variation of energy again, E(t)E(0)E(t)\leq E(0) for any t(ε,ε)t\in(-\varepsilon,\varepsilon). Thus, E(t)E(0)E(t)\equiv E(0) because Σ\Sigma is energy-minimising. So, H(t)0H(t)\equiv 0 by the same formula. In particular, every Σt\Sigma_{t} is capillary minimal stable. Applying Theorem A we obtain that, for every tt,

2πχ(Σt)Iθ(Σt)=1sinθinfMHM|Σt|1sinθinfMHM|Σ|=Iθ(Σ)=2πχ(Σ)=2πχ(Σt).\displaystyle 2\pi\chi(\Sigma_{t})\geq\operatorname{I}_{\theta}(\Sigma_{t})=\frac{1}{\sin\theta}\inf_{\partial M}H_{\partial M}|\partial\Sigma_{t}|\geq\frac{1}{\sin\theta}\inf_{\partial M}H_{\partial M}|\partial\Sigma|=\operatorname{I}_{\theta}(\Sigma)=2\pi\chi(\Sigma)=2\pi\chi(\Sigma_{t}).

This way, every inequality above is an equality, which shows that every Σt\Sigma_{t} is infinitesimally rigid. In particular, Σt\partial\Sigma_{t} has constant geodesic curvature equal to (cotθ)infMHM(\cot\theta)\inf_{\partial M}H_{\partial M} in M\partial M. Since we are supposing either condition (a) or (b) in the theorem, we conclude that either infMHM=0\inf_{\partial M}H_{\partial M}=0 or each component of Σ\partial\Sigma is a geodesic in M\partial M. Hence, either θ=π/2\theta=\pi/2 and we are done, or infMHM=0\inf_{\partial M}H_{\partial M}=0.

Assuming R0=0R_{0}=0 and infMHM=0\inf_{\partial M}H_{\partial M}=0 we are going to show that MM is flat in a neighbourhood of Σ\Sigma and that M\partial M is totally geodesic in this neighbourhood.

Claim: The unit normals NtN_{t} to Σt\Sigma_{t} define a parallel vector field.

Firstly, notice that each map ρt\rho_{t} is constant because it harmonic with Neumann boundary condition by equations (9) and (10). Now, let (x1,x2)(x_{1},x_{2}) be local coordinates on Σ\Sigma and write Ei=xiGE_{i}=\partial_{x_{i}}G. Since each Σt\Sigma_{t} is totally geodesic, we know that EiNt=0\nabla_{E_{i}}N_{t}=0. Moreover, since NtN_{t} is of unit length, tGNt\nabla_{\partial_{t}G}N_{t} is tangent to Σt\Sigma_{t}. Thus,

g(tGNt,Ei)\displaystyle g(\nabla_{\partial_{t}G}N_{t},E_{i}) =tg(Nt,Ei)g(Nt,tGEi)\displaystyle=\partial_{t}g(N_{t},E_{i})-g(N_{t},\nabla_{\partial_{t}G}E_{i})
=g(Nt,EitG)\displaystyle=-g(N_{t},\nabla_{E_{i}}\partial_{t}G)
=xig(Nt,tG)+g(EiNt,tG)\displaystyle=-\partial_{x_{i}}g(N_{t},\partial_{t}G)+g(\nabla_{E_{i}}N_{t},\partial_{t}G)
=xiρt\displaystyle=-\partial_{x_{i}}\rho_{t}
=0.\displaystyle=0.

This proves the Claim.

The Gauss equation for Σt\Sigma_{t} shows that K¯(x1Gx2G)=K(x1Gx2G)=0\overline{K}(\partial_{x_{1}}G\wedge\partial_{x_{2}}G)=K(\partial_{x_{1}}G\wedge\partial_{x_{2}}G)=0, where KK and K¯\overline{K} denote the sectional curvatures of Σ\Sigma and of MM. This holds since Σt\Sigma_{t} is flat and totally geodesic. Now, since NtN_{t} is parallel, g(R(,)Nt,)=0g(R(\cdot,\cdot)N_{t},\cdot)=0. The symmetries of the curvature tensor then imply that R0R\equiv 0, that is, MM is flat around Σ\Sigma.

Finally, let us show that M\partial M is totally geodesic near Σ\Sigma. By infinitesimal rigidity we know that II(νt,νt)0\operatorname{\mathrm{I\!I}}(\nu_{t},\nu_{t})\equiv 0 for every t(ε,ε)t\in(-\varepsilon,\varepsilon). If we denote by TtT_{t} a unit normal for Σt\partial\Sigma_{t}, it is straighforward to check that II(Tt,νt)0\operatorname{\mathrm{I\!I}}(T_{t},\nu_{t})\equiv 0 and II(Tt,Tt)0\operatorname{\mathrm{I\!I}}(T_{t},T_{t})\equiv 0. Since {νt,Tt}\{\nu_{t},T_{t}\} is an orthonormal referential for T(M)T(\partial M) around Σ\partial\Sigma, the bilinearity of II\operatorname{\mathrm{I\!I}} implies that II0\operatorname{\mathrm{I\!I}}\equiv 0, i.e. M\partial M is totally geodesic in the union of all Σt\partial\Sigma_{t}, t(ε,ε)t\in(-\varepsilon,\varepsilon).

The last statement of the theorem (when θ=π/2\theta=\pi/2) follows from Theorem 7 in [2]. ∎

Remark 3.4.

Theorem D can be proved using the capillary minimal foliation with variable contact angle given by Theorem C instead of the capillary CMC foliation of Theorem B.

Finally, we move on to Theorem E, which gives conditions under which we can control the topology of low index surfaces in some Riemannian 33-manifolds.

As usual, we start with a lemma, whose proof can be found in [6].

Lemma 3.5.

Let Σ¯\overline{\Sigma} be a closed orientable Riemannian surface of genus g0g\geq 0, and let h:Σ¯h:\overline{\Sigma}\to\mathbb{R} be any nonnegative smooth function. Then there exists a conformal map f¯:Σ¯𝕊2\overline{f}:\overline{\Sigma}\to\mathbb{S}^{2} such that Σ¯f¯hdA=0\int_{\overline{\Sigma}}\overline{f}h\,\mathrm{d}A=0 and f¯\overline{f} has degree less than or equal to [g+32]\left[\frac{g+3}{2}\right], where [x][x] denotes the integer part of xx.

Proof of Theorem E.

We analyse each one of the three cases.

  • (i)

    Let {E1,E2,E3}\{E_{1},E_{2},E_{3}\} a local orthonormal frame on Σ\Sigma such that E1E_{1} and E2E_{2} are tangent to Σ\Sigma and E3E_{3} coincides with the unit normal NN to Σ\Sigma. Since Σ\Sigma is minimal, the Gauss equation gives that

    KΣ=K¯(E1E2)12A2.\displaystyle K_{\Sigma}=\overline{K}(E_{1}\wedge E_{2})-\frac{1}{2}\lVert A\rVert^{2}.

    So,

    Ric(N)+2KΣ\displaystyle\operatorname{Ric}(N)+2K_{\Sigma} =K¯(E1,N)+K¯(E2,N)+K¯(E1,E2)A2\displaystyle=\overline{K}(E_{1},N)+\overline{K}(E_{2},N)+\overline{K}(E_{1},E_{2})-\lVert A\rVert^{2}
    (13) =Ric(E1)+Ric(E2)A2.\displaystyle=\operatorname{Ric}(E_{1})+\operatorname{Ric}(E_{2})-\lVert A\rVert^{2}.

    Now let h0h\geq 0 be the first eighenfunction of Problem (2). By gluing a disc on each boundary component of Σ\Sigma, we may view Σ\Sigma as a compact domain of a closed orientable Riemannian surface Σ¯\overline{\Sigma}. So, Lemma 3.5 furnishes a conformal map f¯:Σ¯𝕊2\overline{f}:\overline{\Sigma}\to\mathbb{S}^{2} such that Σ¯f¯hdA=0\int_{\overline{\Sigma}}\overline{f}h\,\mathrm{d}A=0 and deg(f¯)[g+32]\deg(\overline{f})\leq\left[\frac{g+3}{2}\right]. Let ff denote the restriction of f¯\overline{f} to Σ\Sigma. If we write f=(f1,f2,f3)f=(f_{1},f_{2},f_{3}), then we know that each component fif_{i} is orthogonal to the first eigenfunction hh. Since Σ\Sigma has index 11, we have:

    Q(fi,fi)=Σ[fi2(Ric(N)+A2)fi2]dAΣqfi2dL0.\displaystyle Q(f_{i},f_{i})=\int_{\Sigma}\left[\lVert\nabla f_{i}\rVert^{2}-(\operatorname{Ric}(N)+\lVert A\rVert^{2})f_{i}^{2}\right]\,\mathrm{d}A-\int_{\partial\Sigma}qf_{i}^{2}\,\mathrm{d}L\geq 0.

    Summing over ii and since i=13|fi|2=1\sum_{i=1}^{3}|f_{i}|^{2}=1, we get

    Σ[f2(Ric(N)+A2)]dAΣqdL0.\displaystyle\int_{\Sigma}\left[\lVert\nabla f\rVert^{2}-(\operatorname{Ric}(N)+\lVert A\rVert^{2})\right]\,\mathrm{d}A-\int_{\partial\Sigma}q\,\mathrm{d}L\geq 0.

    By the conformality of f¯\overline{f},

    (14) Σf2dA<Σ¯f¯2dA=2Area(f¯(Σ¯))=2Area(𝕊2)deg(f¯)8π[g+32].\displaystyle\int_{\Sigma}\lVert\nabla f\rVert^{2}\,\mathrm{d}A<\int_{\overline{\Sigma}}\lVert\nabla\overline{f}\rVert^{2}\,\mathrm{d}A=2\operatorname{Area}(\overline{f}(\overline{\Sigma}))=2\operatorname{Area}(\mathbb{S}^{2})\deg(\overline{f})\leq 8\pi\left[\frac{g+3}{2}\right].

    Therefore,

    Σ(Ric(N)+A2)dA+ΣqdL<8π[g+32].\displaystyle\int_{\Sigma}(\operatorname{Ric}(N)+\lVert A\rVert^{2})\,\mathrm{d}A+\int_{\partial\Sigma}q\,\mathrm{d}L<8\pi\left[\frac{g+3}{2}\right].

    We now use equation (13) and the hypotheses of curvature Ric0\operatorname{Ric}\geq 0 and HM0H_{\partial M}\geq 0 to continue:

    2ΣKΣdAΣ(Ric(E1)+Ric(E2)2KΣ)dA<8π[g+32]ΣqdL.\displaystyle-2\int_{\Sigma}K_{\Sigma}\,\mathrm{d}A\leq\int_{\Sigma}(\operatorname{Ric}(E_{1})+\operatorname{Ric}(E_{2})-2K_{\Sigma})\,\mathrm{d}A<8\pi\left[\frac{g+3}{2}\right]-\int_{\partial\Sigma}q\,\mathrm{d}L.

    By Gauss-Bonnet theorem,

    2[2π(22gk)ΣkgdL]<8π[g+32]ΣqdL.\displaystyle-2\left[2\pi(2-2g-k)-\int_{\partial\Sigma}k_{g}\,\mathrm{d}L\right]<8\pi\left[\frac{g+3}{2}\right]-\int_{\partial\Sigma}q\,\mathrm{d}L.

    Applying Lemma 3.1 and rearanging, we arive at

    ΣkgdL<2π[9(1)g2(g+k)],\displaystyle\int_{\partial\Sigma}k_{g}\,\mathrm{d}L<2\pi\left[9-(-1)^{g}-2(g+k)\right],

    as we wanted. The particular case mentioned in the theorem follows immediately.

  • (ii)

    To prove this item, we use the identity

    (15) Ric(N)=12(RM+HΣ2A2)KΣ.\displaystyle\operatorname{Ric}(N)=\frac{1}{2}(R_{M}+H_{\Sigma}^{2}-\lVert A\rVert^{2})-K_{\Sigma}.

    Let f:Σ𝕊2f:\Sigma\to\mathbb{S}^{2} be the same function of item (i). As before,

    Σ[f2(Ric(N)+A2)]dAΣqdL0.\displaystyle\int_{\Sigma}\left[\lVert\nabla f\rVert^{2}-(\operatorname{Ric}(N)+\lVert A\rVert^{2})\right]\,\mathrm{d}A-\int_{\partial\Sigma}q\,\mathrm{d}L\geq 0.

    Using equation (15) and inequality (14), we obtain

    12Σ(RM+A2)dAΣKΣdA+ΣqdL<8π[g+32].\displaystyle\frac{1}{2}\int_{\Sigma}(R_{M}+\lVert A\rVert^{2})\,\mathrm{d}A-\int_{\Sigma}K_{\Sigma}\,\mathrm{d}A+\int_{\partial\Sigma}q\,\mathrm{d}L<8\pi\left[\frac{g+3}{2}\right].

    Now we use Gauss-Bonnet therorem and Lemma 3.1 to write

    12Σ(RM+A2)dA+1sinθΣHMdL<8π[g+32]+2π(22gk).\displaystyle\frac{1}{2}\int_{\Sigma}(R_{M}+\lVert A\rVert^{2})\,\mathrm{d}A+\frac{1}{\sin\theta}\int_{\partial\Sigma}H_{\partial M}\,\mathrm{d}L<8\pi\left[\frac{g+3}{2}\right]+2\pi(2-2g-k).

    Finally, we use the curvature assumptions to get

    12infMRM|Σ|+1sinθinfMHM|Σ|<2π[7(1)gk],\displaystyle\frac{1}{2}\inf_{M}R_{M}|\Sigma|+\frac{1}{\sin\theta}\inf_{\partial M}H_{\partial M}|\partial\Sigma|<2\pi\left[7-(-1)^{g}-k\right],

    as we wanted.

  • (iii)

    Item (a) follows from Theorem A, since under the curvature hypotheses Σ\Sigma must be a disc. Item (b) follows from from item (ii) of the current theorem (again, Σ\Sigma must be a disc).

References

  • [1] A. Ainouz and R. Souam, Stable Capillary Hypersurfaces in a Half-Space or a Slab, Indiana Univ. Math. J. 65(3)(2016), 813–831.
  • [2] L. Ambrozio, Rigidity of area-minimizing free boundary surfaces in mean convex three-manifolds, J. Geom. Anal. 25 (2015), 1001–1017.
  • [3] L. Ambrozio, A. Carlotto and B. Sharp, Compactness analysis for free boundary minimal hypersurfaces, Calc. Var. PDE 57(2018), 1–39.
  • [4] L. Ambrozio, A. Carlotto and B. Sharp, Index estimates for free boundary minimal hypersurfaces, Math. Ann. 370(3–4)(2018), 1063–1078.
  • [5] J. L. Barbosa, M. do Carmo and J. Eschenburg, Stability of hypersurfaces of constant mean curvature in Riemannian manifolds, Math. Z. 197(1988), 123–128.
  • [6] J. Chen, A. Fraser and C. Pang, Minimal immersions of compact bordered Riemann surfaces with free boundary, Trans. Amer. Math. Soc. 367(4)(2014), 2487–2507.
  • [7] J. Choe and M. Koiso, Stable capillary hypersurfaces in a wedge, Pacific J. Math. 280(1)(2016), 1–15.
  • [8] R. Courant, The Existence of a Minimal Surface of Least Area Bounded by Prescribed Jordan Arcs and Prescribed Surfaces, Proc. Natl. Acad. Sci. USA 24(2)(1938), 97–101.
  • [9] R. Courant, Dirichlet’s Principle, Conformal Mapping, and Minimal Surfaces, Springer-Verlag New York, 1950.
  • [10] R. Finn, Equilibrium Capillary Surfaces, Volume 284 of Grundlehren der mathematischen Wissenschaften, Springer-Verlag New York, 1986.
  • [11] J. L. Lagrange, Essai d’une nouvelle méthode pour déterminer les maxima et les minima des formules intégrales indéfinies, Miscellanea Taurinensia 2(1)(1760), 173–199.
  • [12] C. Li, A polyhedron comparison theorem for 33-manifolds with positive scalar curvature, Invent. Math. 219(2020), 1–37.
  • [13] R. López, Stability and bifurcation of a capillary surface on a cylinder, SIAM J. Appl. Math. 77(1)(2017), 108–127.
  • [14] J. B. Meusnier, Mémoire sur la courbure des surfaces, Mém. Mathém. Phys. Acad. Sci. Paris, prés. par div. Savans 10(1776), Presented in 1776, 477–510.
  • [15] J. C. C. Nitsche, Stationary partitioning of convex bodies, Arch. Rational Mech. Anal. 89 (1985), 1–19.
  • [16] S. H. Park, Every ring type spanner in a wedge is spherical, Math. Ann. 332(3)(2005), 475–482.
  • [17] M. Ritoré and C. Rosales, Existence and characterization of regions minimizing perimeter under a volume constraint inside Euclidean cones, Trans. Am. Math. Soc. 356(11)(2004), 4601–4622.
  • [18] A. Ros and R. Souam, On stability of capillary surfaces in a ball, Pacific J. Math. 178(2) (1997), 345–361.
  • [19] R. Schoen, Minimal submanifolds in higher codimension, Mat. Contemp. 30(2006), 169–199.
  • [20] M. Spivak, A Comprehensive Introduction to Differential Geometry, Volume 4, Publish Or Perish, 1999.
  • [21] G. Wang and C. Xia, Uniqueness of stable capillary hypersurfaces in a ball, Math. Ann. 374 (2019), 1845–1882.