Skip to main content
Log in

On the index of a free-boundary minimal surface in Riemannian Schwarzschild-AdS

  • Published:
Annals of Global Analysis and Geometry Aims and scope Submit manuscript

Abstract

We consider the index of a certain non-compact free-boundary minimal surface with boundary on the rotationally symmetric minimal sphere in the Schwarzschild-AdS geometry with \(m>0\). As in the Schwarzschild case, we show that in dimensions \(n\ge 4\), the surface is stable, whereas in dimension three, the stability depends on the value of the mass \(m>0\) and the cosmological constant \(\Lambda <0\) via the parameter \(\mu :=m\sqrt{-\Lambda /3}\). We show that while for \(\mu \ge \tfrac{5}{27}\) the surface is stable, there exist positive numbers \(\mu _0\) and \(\mu _1\), with \(\mu _1<\tfrac{5}{27}\), such that for \(0<\mu <\mu _0\), the surface is unstable, while for all \(\mu \ge \mu _1\), the index is at most one.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

References

  1. Barbosa, E., Espinar, J.: On free boundary minimal hypersurfaces in the Riemannian Schwarzschild space. J. Geom. Anal. 31(12), 12548–12567 (2021)

    Article  MathSciNet  MATH  Google Scholar 

  2. Brendle, S., Wang, M.-T.: A Gibbons-Penrose inequality for surfaces in Schwarzschild spacetime. Commun. Math. Phys. 330, 33–43 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  3. Chruściel, P.T., Pollack, D.: Singular Yamabe metrics and initial data with exactly Kottler–Schwarzschild–de Sitter Ends. Ann. Henri Poincaré 9(4), 639–654 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  4. Cruz, N., Olivares, M., Villanueva, J.R.: The geodesic structure of the Schwarzschild anti-de Sitter black hole. Class. Quantum Gravity 22(6), 1167–1190 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  5. Fischer-Colbrie, D.: On complete minimal surfaces with finite Morse index in three-manifolds. Invent. Math. 82(1), 121–132 (1985)

    Article  MathSciNet  MATH  Google Scholar 

  6. Montezuma, R.: On free-boundary minimal surfaces in the Riemannian Schwarzschild manifold. Bull. Braz. Math. Soc. (N.S.) 52(4), 1055–1071 (2021)

    Article  MathSciNet  MATH  Google Scholar 

  7. O’Neill, B.: Semi-Riemannian Geometry. Pure and Applied Mathematics, vol. 103. Academic Press, New York (1983)

    Google Scholar 

  8. Schoen, R.M., Yau, S.-T.: On the proof of the Positive Mass Conjecture in General Relativity. Commun. Math. Phys. 65(1), 45–76 (1979)

    Article  MathSciNet  MATH  Google Scholar 

  9. Solanki, R.: Kottler spacetime in isotropic static coordinates. Class. Quantum Gravity 39(1), 19 (2022)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

We thank Peter McGrath for comments on a draft version. We also thank an anonymous referee for thoughtful comments and insightful suggestions that improved the presentation.

Author information

Authors and Affiliations

Authors

Contributions

All authors contributed to and reviewed the manuscript.

Corresponding author

Correspondence to Justin Corvino.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: Curvature calculations

We record the Ricci curvature for warped products \((B\times F, g_B + \psi ^2 g_F)\), where \(\psi \) is a positive function on B. In verifying curvature calculations in the paper, we have applied this to metrics which have the forms: (i) \(h(r) \, \text {d}r^2 + r^2\mathring{g}_{{\mathbb {S}}^{n-1}}\), with base \((I, h(r)\, \text {d}r^2)\) on an interval \(I\subset \mathbb (0, +\infty )\) and fibre \({\mathbb {S}}^{n-1}\); (ii) \(\text {d}\ell ^2 + (r(\ell ))^2 \mathring{g}_{{\mathbb {S}}^{n-1}}\) with base \((I, \text {d}s^2)\); and (iii) Lorentzian warped products \(-f(r)\, \text {d}t^2 + g\), with base \((M^n, g)\) and fibre \({\mathbb {R}}\). We work out (i) as an example below.

We suppose X and Y are (lifts) of tangent vectors to B, and V and W are lifts of tangent vectors to F, \(d= \textrm{dim}(F)\), and where we note that for a function like \(\psi \) on the base, then with \(g=g_B + \psi ^2 g_F\), \(\textrm{Hess}_g\psi (X, Y)= \textrm{Hess}_{g_B}\psi (X, Y)\), and \(g_B( \mathrm {grad_{g_B}}\psi , \mathrm {grad_{g_B}}\psi )= g( \mathrm {grad_{g}}\psi , \mathrm {grad_{g}}\psi )=:\langle \text {d}\psi , d\psi \rangle \). The Ricci curvature is determined from the following [7, Corollary 43]:

$$\begin{aligned} \textrm{Ric}_g(X, Y)&= \textrm{Ric}_{g_B}(X, Y) - \frac{ d \;\textrm{Hess}_{g_B}\psi (X, Y)}{\psi } \nonumber \\ \textrm{Ric}_g(X, V)&= 0 \nonumber \\ \textrm{Ric}_g(V, W)&= \textrm{Ric}_{g_F}(V, W) - g( V, W) \left( \frac{ \Delta _{g_B} \psi }{\psi } + (d-1) \frac{ \langle \text {d}\psi , \text {d}\psi \rangle }{\psi ^2}\right) . \end{aligned}$$
(A.1)

We begin with the Levi-Civita connection \(\nabla \) for

$$\begin{aligned} g&= h(r) \, \text {d}r^2 + r^2\mathring{g}_{{\mathbb {S}}^{n-1}} \qquad \qquad \qquad \qquad (n\ge 2)\\&= h(r) \, \text {d}r^2 + r^2\, ( \text {d}\phi ^2 +\sin ^2\phi \;\mathring{g}_{{\mathbb {S}}^{n-2}})\qquad \qquad (n\ge 3)\\&= h(r) \, \text {d}r^2 + r^2\, (d\phi ^2 + \sin ^2\phi (d\theta ^2 + \sin ^2\theta \mathring{g}_{{\mathbb {S}}^{n-3}})) \qquad (n\ge 4). \end{aligned}$$

For \(n\ge 4\), we recursively employ corresponding angular coordinates on \({\mathbb {S}}^{n-3}\), and in these coordinates, the metric components and Ricci components comprise diagonal matrices. We let \(\vartheta \) denote the general angular coordinate, and we abbreviate \(\mathring{g}_{{\mathbb {S}}^{n-1}}=:\mathring{g}\). The nonzero Christoffel symbols \(\Gamma ^k_{rj}\) are

$$\begin{aligned} \Gamma ^r_{rr} = \frac{ h'(r) }{2 h(r)}, \qquad \Gamma ^\vartheta _{r\vartheta } = \frac{1}{r} \end{aligned}$$

i.e.

$$\begin{aligned} \nabla _{\frac{\partial }{\partial r}} \tfrac{\partial }{\partial r} = \tfrac{ h'(r) }{2 h(r)} \tfrac{\partial }{\partial r}, \qquad \nabla _{\frac{\partial }{\partial r}} \tfrac{\partial }{\partial \vartheta } = \tfrac{ 1 }{r} \tfrac{\partial }{\partial \vartheta }. \end{aligned}$$

The nonzero \(\Gamma ^r_{jk}\) for j and k angular indices are

$$\begin{aligned} \Gamma ^r_{\vartheta \vartheta }= -\frac{r\mathring{g}_{\vartheta \vartheta }}{h(r)} = -\frac{g_{\vartheta \vartheta }}{rh(r)}, \end{aligned}$$

i.e. the vector-valued second fundamental form \({I\hspace{-0.1cm}I}\) of the spherical fibre is

$$\begin{aligned} {I\hspace{-0.1cm}I}(V, W)= - \frac{g(V, W)}{r h(r)} \frac{\partial }{\partial r} = - \frac{g(V, W)}{r } \textrm{grad}(r)=- \frac{g(V, W)}{r \sqrt{h(r)}} {\textbf{n}} \end{aligned}$$

where \({\textbf{n}}= \frac{ 1}{\sqrt{h(r)}} \frac{\partial }{\partial r}\) is a unit normal to a spherical fibre.

The Christoffel symbols \(\Gamma ^k_{ij}\) with all angular indices are just the Christoffel symbols of the round sphere (which are invariant under a constant scaling, i.e. independent of radius). If the angular coordinates are numbered \(\vartheta _1=\phi \), \(\vartheta _2=\theta , \ldots , \vartheta _{n-1}\), such that for \(j\ge 2\), \(g_{\vartheta _j \vartheta _j} = r^2 \sin ^2 \vartheta _1 \cdots \sin ^2\vartheta _{j-1}\), then \(\Gamma _{\vartheta _j\vartheta _j}^{\vartheta _i}=0\) for \(i\ge j\), whereas for \(1\le i <j\), \(\Gamma _{\vartheta _j\vartheta _j}^{\vartheta _i}=-\sin \vartheta _i\cos \vartheta _i \sin ^2\vartheta _{i+1} \cdots \sin ^2\vartheta _{j-1}\), while for \(i<j\), \(\Gamma _{\vartheta _i \vartheta _j}^{\vartheta _j}=\cot \vartheta _i\).

Observe that \(\textrm{Hess}_g r(\tfrac{\partial }{\partial r}, \frac{\partial }{\partial r}) = - \Gamma _{rr}^r= -\frac{ h'(r) }{2\,h(r)}=\textrm{Hess}_{g_B} r(\tfrac{\partial }{\partial r}, \frac{\partial }{\partial r}) \), the Hessian of r with respect to \(g_B:=h(r) \, \text {d}r^2\).

As for the Ricci curvature components, we apply the above-warped product formulas, with the warping function \(\psi =r\). Since the base is one-dimensional, its curvature vanishes, and

$$\begin{aligned} \textrm{Ric}_{rr} = - \frac{n-1}{r} \textrm{Hess}_g r( \tfrac{\partial }{\partial r}, \tfrac{\partial }{\partial r})= \frac{(n-1) h'(r)}{2rh(r)}. \end{aligned}$$

Now, on \({\mathbb {S}}^{n-1}\) we have \(\textrm{Ric}(\mathring{g}) = (n-2)\mathring{g}\). Thus the remaining nonzero components of Ricci are

$$\begin{aligned} \textrm{Ric}_{\vartheta \vartheta }&= (n-2) \mathring{g}_{\vartheta \vartheta } - r^2 \mathring{g}_{\vartheta \vartheta } \left( -\frac{ h'(r) }{2r (h(r))^2} + (n-2) \frac{1}{r^2h(r)} \right) \\ {}&= \mathring{g}_{\vartheta \vartheta }\left( (n-2)\Big ( 1-\frac{1}{h(r)} \Big ) + \frac{ rh'(r) }{2(h(r))^2} \right) . \end{aligned}$$

The scalar curvature of g is given by

$$\begin{aligned} R(g)&= \frac{(n-1) h'(r)}{2r(h(r))^2}+\frac{ (n-1) (n-2)}{r^2} \left( 1- \frac{1}{h(r)}\right) + \frac{ (n-1) h'(r) }{2r (h(r))^2}\\&= \frac{(n-1) h'(r)}{r(h(r))^2}+\frac{ (n-1) (n-2)}{r^2} \left( 1- \frac{1}{h(r)}\right) . \end{aligned}$$

Example A.1

We apply these formulas to \(g_{m, \Lambda }= \tfrac{\text {d}r^2}{1-\frac{2\Lambda }{n(n-1)} r^2-\frac{2\,m}{r^{n-2}}} + r^2 \mathring{g}_{{\mathbb {S}}^{n-1}}\). In this case we have \((h(r))^{-1} = 1-\tfrac{2\Lambda }{n(n-1)} r^2-\tfrac{2\,m}{r^{n-2}}, \) and so

$$\begin{aligned} 1-\frac{1}{h(r)}&= \frac{2\Lambda }{n(n-1)} r^2+\frac{2m}{r^{n-2}}\\ \frac{2}{rh(r)} \Gamma ^r_{rr} = \frac{h'(r)}{r(h(r))^2}&= \frac{4\Lambda }{n(n-1)} -\frac{2(n-2)m}{r^n}. \end{aligned}$$

Thus

$$\begin{aligned} \textrm{Ric}_{rr}&= \frac{n-1}{2} h(r) \left( \frac{4\Lambda }{n(n-1)} -\frac{2(n-2)m}{r^n} \right) \\ \textrm{Ric}_{\vartheta \vartheta }&= \mathring{g}_{\vartheta \vartheta } \left( (n-2) \left( \frac{2\Lambda }{n(n-1)} r^2+\frac{2m}{r^{n-2}}\right) + \frac{2\Lambda }{n(n-1)}r^2 -\frac{(n-2)m}{r^{n-2}}\right) \\&= r^2\mathring{g}_{\vartheta \vartheta } \left( \frac{2\Lambda }{n} +\frac{(n-2)m}{r^{n}}\right) \\ R(g)&= (n-1) \Big (\frac{2\Lambda }{n(n-1)} -\frac{(n-2)m}{r^n}\Big ) + (n-1) \left( \frac{2\Lambda }{n} +\frac{(n-2)m}{r^{n}}\right) \\&=2\Lambda . \end{aligned}$$

A unit normal \(\nu \) along the surface \(x^n=0\) is given by \(\nu =\frac{1}{r} \frac{\partial }{\partial \phi }\), and so we have \(\textrm{Ric}_g(\nu , \nu ) = \frac{2\Lambda }{n} + \frac{(n-2)m}{r^n}\). Furthermore, the second fundamental form of the spherical fibre at fixed r satisfies, with \(\eta = -{\textbf{n}}= - \frac{ 1}{\sqrt{h(r)}} \frac{\partial }{\partial r}\),

$$\begin{aligned} \langle {I\hspace{-0.1cm}I}(\nu , \nu ), \eta \rangle = \frac{\sqrt{1-\tfrac{2\Lambda }{n(n-1)} r^2-\tfrac{2m}{r^{n-2}}}}{r}. \end{aligned}$$

This remains bounded as r increases when \(\Lambda <0\), and tends to 0 as \(r\nearrow \infty \) for \(\Lambda =0\).

Appendix B: Schwarzschild test function

Consider the Schwarzschild metric \(g_m^S = (1+\tfrac{m}{2\rho })^4 (d\rho ^2 + \rho ^2 \mathring{g}_{{\mathbb {S}}^2})\) in isotropic coordinates \(x\in {\mathbb {R}}^3{\setminus } \{ 0\}\), \(\rho = |x|\) (Euclidean length). We let \(\Sigma \) be the free-boundary minimal surface \(x^3=0\) (\(\phi =\frac{\pi }{2}\)) and \(\rho \ge \tfrac{m}{2}\). For any \(\alpha >\max ( 1, \tfrac{m}{2})\), let \(\Sigma ^\alpha \subset \Sigma \) be the subset with \(\alpha \le \rho \le \alpha ^2\), and let \(\Sigma '_R\subset \Sigma \) be given by \(\tfrac{m}{2} \le \rho \le R\).

If u is a radial eigenfunction of \({\mathcal {L}}_{\Sigma }\) on \(\Sigma '_R\) with eigenvalue \(\lambda \), we let \(w= \sqrt{\rho } \; u\), to obtain

$$\begin{aligned} w''(\rho ) + w(\rho ) \left( \tfrac{1}{4\rho ^2} + (1+ \tfrac{m}{2\rho })^{-2} \tfrac{m}{\rho ^3} + \lambda (1+ \tfrac{m}{2\rho })^{4}\right) =0. \end{aligned}$$

The Neumann condition on u at \(\rho = \tfrac{m}{2}\) translates to \(w'(\tfrac{m}{2})= m^{-1} w(\tfrac{m}{2})\). We let \(w_\lambda \) be the solution to the above equation with initial conditions \(w(\tfrac{m}{2})=1\) and \(w'(\tfrac{m}{2})=m^{-1}\). For \(\lambda =0\) one can solve to get \(w_0(\rho )= \sqrt{\frac{2\rho }{m}} \left( 1- \frac{ 2\rho -m}{2\rho +m} \log \sqrt{\frac{2\rho }{m}} \right) \), as observed in [6]. This is eventually negative, so that 0 is not an eigenvalue on \((\Sigma '_R, g_m^S)\) for R large (recall Lemma 3.4). That said, as in [6], as \(\lambda \nearrow 0\), \(w_\lambda \) must have a zero, just as \(w_0\) does, by perturbation, and thus \(\Sigma \) is unstable. We recall a different argument here, using a log cut-off, cf. [8, p. 54].

The idea is to modify the test function \(\psi =1\), which is not in \(L^2(\Sigma )\), and likewise does not have the appropriate boundary conditions on \(\Sigma '_R\). It is easy to show that a simple linear cut-off over \(\Sigma ^\alpha \) will not suffice, so we use the following test function:

$$\begin{aligned} \psi _\alpha (x) = {\left\{ \begin{array}{ll} 1, \qquad \qquad \quad |x|< \alpha \\ \frac{\ln (\alpha ^2 |x|^{-1})}{\ln \alpha }, \quad x\in \Sigma ^{\alpha } \\ 0, \qquad \qquad \quad |x|>\alpha ^2 \end{array}\right. }. \end{aligned}$$

This test function is in \(W^{1,2}(\Sigma )\), and on \(\Sigma ^{\alpha }\),

$$\begin{aligned} |\nabla \psi _\alpha |^2_{g_m^S} =( 1+ \tfrac{m}{2|x|})^{-4}|d\psi _\alpha |^2_{ g_{{\mathbb {E}}^2}}=( 1+ \tfrac{m}{2|x|})^{-4} \tfrac{1}{|x|^2 (\ln \alpha )^2}. \end{aligned}$$

Thus we estimate

$$\begin{aligned} \int \limits _{\Sigma ^\alpha } \Big (&|\nabla \psi _\alpha |^2_{g_m^S} -\textrm{Ric}_{g_m^S} (\nu , \nu )\psi _\alpha ^2\Big ) \; \text {d}A_{g_m^S}\\&= \int \limits _0^{2\pi } \int \limits _{\alpha }^{\alpha ^2} ( 1+ \tfrac{m}{2 \rho })^{-4} \tfrac{1}{\rho ^2 (\ln \alpha )^2}\rho \; d\rho \; \text {d}\theta -\int \limits _{\Sigma ^\alpha } \textrm{Ric}_{g_m^S} (\nu , \nu )\psi _\alpha ^2 \; \text {d}A_{g_m^S}\\&\le 2\pi \frac{\ln \rho }{(\ln \alpha )^2} \Big |_{\rho =\alpha }^{\rho =\alpha ^2} -\int \limits _{\Sigma ^\alpha } \textrm{Ric}_{g_m^S} (\nu , \nu )\psi _\alpha ^2 \; \text {d}A_{g_m^S}\\&= \frac{2\pi }{\ln \alpha } -\int \limits _{\Sigma ^\alpha } \textrm{Ric}_{g_m^S} (\nu , \nu )\psi _\alpha ^2 \; \text {d}A_{g_m^S}. \end{aligned}$$

We conclude

$$\begin{aligned} {\mathcal {Q}}^{\Sigma '_{\alpha ^2}}(\psi _\alpha )&=\frac{2\pi }{\ln \alpha } -\int \limits _{\Sigma ^\alpha } \textrm{Ric}_{g_m^S} (\nu , \nu )\psi _\alpha ^2 \; \text {d}A_{g_m^S}- \int \limits _{\Sigma '_\alpha } \textrm{Ric}_{g_m^S} (\nu , \nu ) \; \text {d}A_{g_m^S} \\ {}&{\mathop {\longrightarrow }\limits ^{\alpha \nearrow \infty }}\int _{\Sigma }-\textrm{Ric}_{g_m^S} (\nu , \nu ) \; \text {d}A_{g_m^S}. \end{aligned}$$

As \(\textrm{Ric}_{g_m^S}(\nu , \nu )>0\) along \(\Sigma \), for \(\alpha \) large enough, \({\mathcal {Q}}^{\Sigma '_{\alpha ^2}}(\psi _\alpha )<0\), as desired.

Appendix C: On eigenfunctions on \(\Sigma \)

In this section we discuss an application of the arguments in [5] to the setting we consider. We take \(\Sigma \subset (M,g)\) to be a complete non-compact free-boundary minimal surface in (Mg), with smooth unit normal field \(\nu \), and with free boundary \(\partial \Sigma \) lying on a totally geodesic hypersurface S. We also consider an exhaustion \(\Sigma _R\) of \(\Sigma \) by smooth regions, with \(\partial \Sigma _R= \partial \Sigma \sqcup \partial \Sigma _R^+\), and let \(\eta \) be the outward-pointing unit conormal along the boundary of \(\Sigma _R\). The Robin boundary condition along \(\partial \Sigma \) reduces to the Neumann condition. Recall the eigenvalue equation \({\mathcal {L}}_\Sigma u + \lambda u=0\). We start with a simple lemma.

Lemma C.1

Suppose \(W\subset L^2(\Sigma )\) is a finite-dimensional subspace such that for all \(\varphi \in C^{\infty }_c(\Sigma ) \cap W^{\perp }\), \({\mathcal {Q}}^\Sigma (\varphi )\ge 0\). Then \(\textrm{ind}(\Sigma )\le \textrm{dim}(W)\).

Proof

If \(\textrm{ind}(\Sigma _R)>\textrm{dim}(W)\), there is a non-trivial test function \(\varphi \) which is a linear combination of eigenfunctions with negative eigenvalues for \(\Sigma _R\) (Neumann on \(\partial \Sigma \), Dirichlet on \(\partial \Sigma _R^+\) and extended to vanish outside \(\Sigma _R\)), and is in the kernel of the \(L^2(\Sigma )\)-orthogonal projection map onto W, i.e. \(\varphi \) is \(L^2(\Sigma )\)-orthogonal to W. But since \(\mathcal Q^\Sigma (\varphi )<0\), this is a contradiction. Thus \(\textrm{ind}(\Sigma ) \le \textrm{dim}(W)\). \(\square \)

The next proposition addresses the converse question, for which we make two additional assumptions, both of which hold in case \(g= g_{m, \Lambda }\) with \(m>0\) (recall the proof of Proposition 3.9 for (ii)):

  1. (i)

    For some constant \(c>0\), \(q:=\textrm{Ric}_g(\nu ,\nu )+| I\hspace{-0.1cm}I^\Sigma (\nu , \nu )|^2_g\) satisfies \(|q|\le c\).

  2. (ii)

    Given any \(\ell \in {\mathbb {Z}}_+\) and \(u_1, \ldots , u_\ell \in W^{1,2}(\Sigma )\), there is a sequence \(r_j\nearrow \infty \) such that for each \(1\le i\le \ell \), \(\int \limits _{\partial \Sigma ^+_{r_j}} (u_i^2 + |du_i|^2_g)\; d\sigma _g\rightarrow 0\).

Under these conditions, the next proposition is a corollary of the proof of Proposition 2 in [5], utilizing the uniform bound on |q|, which was not assumed in [5].

Proposition C.2

Suppose \(\Sigma \subset (M,g)\) is a FBMS with free boundary \(\partial \Sigma \) lying on a totally geodesic hypersurface S, and that conditions (i) and (ii) above hold. If \(\textrm{ind}(\Sigma )=k\) is finite, then there is a k-dimensional subspace \(W\subset W^{1,2}(\Sigma )\) spanned by \(L^2(\Sigma )\)-orthonormal eigenfunctions \(f_1, \ldots , f_k\) with negative eigenvalues \(\lambda _1, \ldots , \lambda _k\) and Neumann boundary condition along \(\partial \Sigma \), such that for any \(\varphi \in W^{1,2}(\Sigma )\) which is \(L^2(\Sigma )\)-orthogonal to W, we have \(\mathcal Q^\Sigma (\varphi )\ge 0\). Moreover, \({\mathcal {Q}}^{\Sigma }\) is negative definite on W, and \({\mathcal {Q}}^{\Sigma }(f_i)=\lambda _i\). As such, if \(\widetilde{W} \subset W^{1,2}(\Sigma )\) is a subspace such that \({\mathcal {Q}}^{\Sigma }\) is negative definite on \(\widetilde{W}\), then \(\textrm{dim}(\widetilde{W})\le \textrm{ind}(\Sigma )\).

Proof

We will follow closely the proof of [5, Proposition 2]. As S is totally geodesic, \({\mathcal {Q}}^\Sigma (\varphi )= \int \limits _\Sigma ( |\nabla ^\Sigma \varphi |^2_g - q \varphi ^2) \; dA_{g}\), which converges for all \(\varphi \in W^{1,2}(\Sigma )\) by the bound on q. In case \(\Sigma \) is stable, \(\mathcal Q^{\Sigma }(\varphi )\ge 0\) for all \(\varphi \in C^\infty _c(\Sigma )\), so that by approximation we have \({\mathcal {Q}}^\Sigma (\varphi )\ge 0\) for all \(\varphi \in W^{1,2}(\Sigma )\) (where we use that |q| is bounded). Thus \(W=\{0\}\) in case \(k=0\).

We now suppose \(\textrm{ind}(\Sigma )=k\in {\mathbb {Z}}_+\). The first part of the proof of Proposition 1 in [5] shows that for large R, say \(R\ge R_0\), \(\Sigma {\setminus } \Sigma _R\) is stable (of course in the setting we consider above in \(g_{m, \Lambda }\) with \(\Lambda <0\), this is obvious). Just as in [5], for each R there is a cut-off function \(\eta =\eta _R\) which is zero on \(\Sigma _R\) and identically 1 outside \(\Sigma _{2R}\), with \(0\le \eta \le 1\) and \(|\nabla ^\Sigma \eta |_g^2 \le \frac{5(1-\eta ^2)}{R^2}\). We apply the stability inequality on \(\Sigma \setminus \Sigma _R\) to \(\eta \varphi \) for \(\varphi \in C^\infty _c(\Sigma )\) to find

$$\begin{aligned} 0&\le \int \limits _\Sigma ( |\nabla ^\Sigma (\eta \varphi )|^2_g-q\eta ^2 \varphi ^2)\; \text {d}A_{g} \\&= \int \limits _\Sigma (\eta ^2 |\nabla ^\Sigma \varphi |^2_g + 2 \eta \varphi \langle \nabla ^\Sigma \eta , \nabla ^\Sigma \varphi \rangle + \varphi ^2|\nabla ^\Sigma \eta |^2_g -q\eta ^2 \varphi ^2)\; \text {d}A_{g}\\&\le \int \limits _\Sigma (\eta ^2 |\nabla ^\Sigma \varphi |^2_g + 2 | \nabla ^\Sigma \eta |^2_g | \nabla ^\Sigma \varphi |_g^2 + \varphi ^2|\nabla ^\Sigma \eta |^2_g +(\tfrac{1}{2} -q) \eta ^2\varphi ^2)\; dA_{g}.\\ \end{aligned}$$

We add \({\mathcal {Q}}^\Sigma (\varphi ) = \int _\Sigma ( |\nabla ^\Sigma \varphi |^2_g -q \varphi ^2)\; \text {d}A_{g}\) to each side and rearrange, as in [5],

$$\begin{aligned} \int \limits _\Sigma (1&-\eta ^2) |\nabla ^\Sigma \varphi |^2_g \; dA_{g} \\&\le {\mathcal {Q}}^{\Sigma }(\varphi )+ \int \limits _\Sigma ( 2 | \nabla ^\Sigma \eta |^2_g | \nabla ^\Sigma \varphi |_g^2 + \varphi ^2|\nabla ^\Sigma \eta |^2_g +(\tfrac{1}{2} \eta ^2 + (1-\eta ^2) q) \varphi ^2)\; \text {d}A_{g}. \end{aligned}$$

Hence we can use the above bound on \(|\nabla ^{\Sigma } \eta |_g\) to see that for \(R\ge \max (R_0, \sqrt{20})\),

$$\begin{aligned} \int \limits _{\Sigma _R} |\nabla ^\Sigma \varphi |^2_g \; \text {d}A_{g}&\le \int \limits _\Sigma (1 -\eta ^2) |\nabla ^\Sigma \varphi |^2_g \; dA_{g} \le 2 {\mathcal {Q}}^{\Sigma }(\varphi )+ (1 + 2c) \int \limits _\Sigma \varphi ^2\; \text {d}A_{g}. \end{aligned}$$
(C.1)

By approximation, \({\mathcal {Q}}^\Sigma (\varphi ) \ge - ( \tfrac{1}{2} + c) \int \limits _\Sigma \varphi ^2\; \text {d}A_{g}\) holds for all \(\varphi \in W^{1,2}(\Sigma )\).

Since \(k=\textrm{ind}(\Sigma )\), then for R large, say \(R\ge R_1\), there are precisely k negative eigenvalues (with multiplicity) for \(\Sigma _R\), with Neumann boundary condition on \(\partial \Sigma \), and Dirichlet on \(\partial \Sigma _R^+\). We take \(f_{i, R}\), \(1\le i\le k\), to be \(L^2(\Sigma _R)\)-orthonormal eigenfunctions with respective eigenvalues \(\lambda _{i, R}<0\), and extend \(f_{i,R}\) by zero to \(\Sigma \). By the mini-max characterization of eigenvalues, we have that the maximum of these negative eigenvalues decreases with R, so that there is a \(\varepsilon _0>0\) so that \(\lambda _{i,R} \le -\varepsilon _0<0\) for all \(R\ge R_1\). On the other hand, from the last paragraph above we can infer that the eigenvalues are uniformly bounded below: \(\lambda _{i,R}\ge -c_0:=-(\tfrac{1}{2} +c)\).

From the stability inequality, with \(\eta =\eta _R\) for \(R\ge R_0\) and \(\rho \ge R_1\),

$$\begin{aligned} 0&\le \int \limits _\Sigma \Big (\eta ^2 |\nabla ^\Sigma f_{i, \rho }|^2_g + 2 \eta f_{i, \rho } \langle \nabla ^\Sigma \eta , \nabla ^\Sigma f_{i, \rho } \rangle + f_{i, \rho }^2|\nabla ^\Sigma \eta |^2_g -q\eta ^2 f_{i, \rho }^2\Big )\; \text {d}A_{g}\\&= \int \limits _\Sigma \Big (\eta ^2 |\nabla ^\Sigma f_{i,\rho }|^2_g + \tfrac{1}{2} \langle \nabla ^\Sigma ( \eta ^2), \nabla ^\Sigma (f_{i, \rho }^2) \rangle + f_{i,\rho }^2|\nabla ^\Sigma \eta |^2_g -q\eta ^2 f_{i, \rho }^2\Big )\; \text {d}A_{g}\\&= \int \limits _\Sigma \Big (\eta ^2 |\nabla ^\Sigma f_{i,\rho }|^2_g - \eta ^2( f_{i,\rho } \Delta _\Sigma f_{i,\rho } + | \nabla ^\Sigma f_{i, \rho }|_g^2) + f_{i,\rho }^2|\nabla ^\Sigma \eta |^2_g -q\eta ^2 f_{i, \rho }^2\Big )\; \text {d}A_{g}, \end{aligned}$$

which we can rearrange to obtain

$$\begin{aligned} -\lambda _{i,\rho } \int \limits _\Sigma \eta ^2 f_{i,\rho }^2 \; \text {d}A_{g} \le \int \limits _\Sigma f_{i,\rho }^2|\nabla ^\Sigma \eta |^2_g\; dA_{g}\le 5R^{-2} . \end{aligned}$$

Using the upper bound on the eigenvalues we find

$$\begin{aligned} \int \limits _{\Sigma \setminus \Sigma _{2R}} f_{i,\rho }^2\; \text {d}A_{g}\le 5\varepsilon _0^{-1}R^{-2}. \end{aligned}$$
(C.2)

Using \(\varphi =f_{i, \rho }\) in (C.1), we find

$$\begin{aligned} \int \limits _{\Sigma _R} |\nabla ^\Sigma f_{i,\rho }|^2_g \; \text {d}A_{g} \le 2 {\mathcal {Q}}^{\Sigma }(f_{i,\rho })+ (1 + 2c) \le -2\varepsilon _0+(1 + 2c), \end{aligned}$$

and thus \(\{f_{i,\rho }: 1\le i\le k, \; \rho \ge R_1\}\) is bounded in \(W^{1,2}(\Sigma _R)\). By Rellich’s lemma (along with Banach-Alaoglu and Riesz Representation), we may arrange a sequence \(\rho _j\nearrow \infty \) such that for each \(1\le i\le k\), \(f_{i, \rho _j}\) converges in \(L^2(\Sigma _{R})\) and weakly in \(W^{1,2}(\Sigma _{R})\) for each \(R\ge R_0\), and the limit is given by a function \(f_i\in W^{1,2}_{\textrm{loc}}(\Sigma )\). Using (C.2), we see that \(f_i\in L^2(\Sigma )\) and that \(f_{i, \rho _j}\) converges to \(f_i\) in \(L^2(\Sigma )\). As such, \(f_1, \ldots , f_k\) is an \(L^2(\Sigma )\)-orthonormal collection. We can also arrange by boundedness of the eigenvalues that for each i, \(\lambda _{i, \rho _j}\) converges, say to \(\lambda _i\in [-c_0, -\varepsilon _0]\).

Moreover, by the uniform bound on \(\int _{\Sigma _R} |\nabla ^\Sigma f_{i, \rho _j}|^2_g \; \text {d}A_{g}\) and the lower semicontinuity of the norm under weak convergence, we see that \(f_i\in W^{1,2}(\Sigma )\). By the weak convergence, we have for any test function \(\psi \in C^\infty _c(\Sigma )\), and for j large enough so that \(\psi \) is supported in \(\Sigma _{\rho _j}\),

$$\begin{aligned} 0&=\int \limits _{\Sigma } ( \langle \nabla ^\Sigma \psi , \nabla ^\Sigma f_{i, \rho _j} \rangle -q\psi f_{i, \rho _j}-\lambda _{i, \rho _j} \psi f_{i, \rho _j})\; \text {d}A_{g}\\ {}&\qquad \longrightarrow \int \limits _{\Sigma } ( \langle \nabla ^\Sigma \psi , \nabla ^\Sigma f_{i} \rangle -q\psi f_{i}-\lambda _i \psi f_i)\; \text {d}A_{g}. \end{aligned}$$

This is the weak formulation of the boundary-value problem \(\mathcal L_\Sigma f_i + \lambda _i f_i=0\), \(\frac{\partial f_i}{\partial \eta } =0\) on \(\partial \Sigma \). By regularity for the elliptic boundary-value problem, we have that \(f_i\) is a classical solution.

That \({\mathcal {Q}}^{\Sigma }(f_i)=\lambda _i\) now follows by (ii) and Cauchy-Schwarz:

$$\begin{aligned} \int \limits _{\Sigma _{r_j}}( |\nabla ^\Sigma f_i|^2_g - qf_i^2)\; dA_{g}= \lambda _i \int \limits _{\Sigma _{r_j}} f_i^2 \; \text {d}A_{g} + \int \limits _{\partial \Sigma _{r_j}^+} f_i \frac{\partial f_i}{\partial \eta }\; \text {d}\sigma _{g}\rightarrow \lambda _i. \end{aligned}$$

That \({\mathcal {Q}}^\Sigma \Big (\sum \limits _{i=1}^k c_i f_i \Big ) = \sum \limits _{i=1}^k \lambda _i c_i^2\) then follows by a similar argument.

Finally, let W be the span of \(f_1, \ldots , f_k\), and suppose \(\varphi \in W^{1,2}(\Sigma )\cap W^\perp \). We follow [5] to show \({\mathcal {Q}}^\Sigma (\varphi )\ge 0\). Let \(\varepsilon >0\), and let \(\varphi _{\ell }\in C^\infty _c(\Sigma )\) approach \(\varphi \) in \(W^{1,2}(\Sigma )\). For sufficiently large \(\ell \), \(|{\mathcal {Q}}^\Sigma (\varphi )-\mathcal Q^{\Sigma }(\varphi _{\ell })|< \varepsilon \), and for each \(1\le i\le k\), \(|\int \limits _\Sigma \varphi _{\ell } f_i dA_{g_\Sigma }|<\sqrt{\varepsilon }\). Fix such an \(\ell \), and consider j so large such that \(\varphi _{\ell }\) is supported in \(\Sigma _{\rho _j}\). Decompose \(\varphi _{\ell } = \sum \limits _{i=1}^k {a_{ij\ell }} f_{i, \rho _j} + \psi _{\ell }\), where \(\psi _{\ell }\perp f_{i, \rho _j}\) in \(L^2(\Sigma )\). Then since \(\varphi _{\ell }\) and \(f_{i, \rho _j}\) vanish outside \(\Sigma _{\rho _j}\), \(\psi _{\ell }\) does as well, and so \({\mathcal {Q}}^{\Sigma }(\psi _{\ell })\ge 0\) and

$$\begin{aligned} {\mathcal {Q}}^{\Sigma }(\varphi _{\ell })&= {\mathcal {Q}}^{\Sigma } \Big ( \sum \limits _{i=1}^k a_{ij\ell } f_{i, \rho _j}\Big ) + \mathcal Q^{\Sigma } (\psi _{\ell }) + 2 \sum \limits _{i=1}^k a_{ij\ell } \mathcal B^\Sigma (f_{i, \rho _j}, \psi _{\ell }) \ge \sum \limits _{i=1}^k \lambda _{i, \rho _j} a^2_{ij\ell } . \end{aligned}$$

Furthermore, \(a_{ij\ell }= \int \limits _{\Sigma } \varphi _{\ell } f_{i, \rho _j} \text {d}A_{g_\Sigma } {\mathop {\longrightarrow }\limits ^{j\nearrow \infty }} \int \limits _{\Sigma } \varphi _{\ell } f_{i} \text {d}A_{g_\Sigma },\) so that for large enough j, \(|a_{ij\ell }|< \sqrt{\varepsilon }\). By the uniform bound on the eigenvalues \(\lambda _{i, \rho _j}\), we see for some constant \(C>0\), \({\mathcal {Q}}^{\Sigma }(\varphi )\ge - C \varepsilon \). Since \(\varepsilon >0\) was arbitrary, we conclude \({\mathcal {Q}}^{\Sigma } (\varphi )\ge 0\). \(\square \)

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Corvino, J., Karangozishvili, E. & Ozbay, D. On the index of a free-boundary minimal surface in Riemannian Schwarzschild-AdS. Ann Glob Anal Geom 64, 22 (2023). https://doi.org/10.1007/s10455-023-09925-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10455-023-09925-w

Keywords

Mathematics Subject Classification

Navigation