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

Comparison geometry for substatic manifolds and a weighted Isoperimetric Inequality

Stefano Borghini S. Borghini, Università degli Studi di Trento, via Sommarive 14, 38123 Povo (TN), Italy [email protected]  and  Mattia Fogagnolo M. Fogagnolo, Università degli Studi di Padova , via Trieste 63, 35121 Padova (PD), Italy [email protected]
Abstract.

Substatic Riemannian manifolds with minimal boundary arise naturally in General Relativity as spatial slices of static spacetimes satisfying the Null Energy Condition. Moreover, they constitute a vast generalization of nonnegative Ricci curvature. In this paper we will prove various geometric results in this class, culminating in a sharp, weighted Isoperimetric inequality that quantifies the area minimizing property of the boundary. Its formulation and proof will build on a comparison theory partially stemming from a newly discovered conformal connection with CD(0,1)\mathrm{CD}(0,1) metrics.

Key words and phrases:
Keywords: substatic manifolds, comparison geometry, isoperimetric inequality.

MSC (2020): 49Q10, 53C21, 53E10,

1. Introduction

In this paper we are interested in the study of triples (M,g,f)(M,g,f), where (M,g)(M,g) is a Riemannian manifold of dimension n3n\geq 3 with (possibly empty) compact boundary M\partial M and f:Mf:M\to\mathbb{R} is a smooth function that is positive in the interior of MM and zero on M\partial M, satisfying the following inequality

fRic2f+(Δf)g0,f{\rm Ric}-\nabla^{2}f+(\Delta f)g\geq 0, (1.1)

where Ric{\rm Ric} is the Ricci tensor of the metric gg, 2\nabla^{2} is the Hessian and Δ=tr2\Delta={\rm tr}\nabla^{2} is the Laplace–Beltrami operator with respect to the Levi-Civita connection \nabla of gg. We will refer to such triples (M,g,f)(M,g,f) as substatic triples or simply substatic manifolds. We say that a substatic manifold has horizon boundary if M\partial M is either empty or it is a minimal hypersurface and |f|0\lvert\nabla f\rvert\neq 0 on M\partial M.

Condition (1.1) arises naturally in the study of static spacetimes satisfying the Null Energy Condition, as already observed in [WWZ17]. More precisely, a Lorentzian manifold (L,𝔤)(L,\mathfrak{g}) of the form

L=×M,𝔤=f2dtdt+g,L=\mathbb{R}\times M\,,\qquad\mathfrak{g}\,=\,-f^{2}dt\otimes dt+g,

happens to be a solution to the Einstein Field Equation

Ric𝔤+(Λ12R𝔤)𝔤=T,{\rm Ric}_{\mathfrak{g}}\,+\,\left(\Lambda-\frac{1}{2}{\mathrm{R}}_{\mathfrak{g}}\right)\mathfrak{g}\,=\,T\,,

subject to T(X,X)0T(X,X)\geq 0 for any vector field XX satisfying 𝔤(X,X)=0\mathfrak{g}(X,X)=0, exactly when ff and gg satisfy (1.1). A minimal boundary represents, in this framework, the event horizon of a black hole. For the reader’s sake, we included the computations in Section A.1. The class of substatic manifolds obviously includes the very large and thoroughly studied class of manifolds with nonnegative Ricci curvature, where ff is just constant, and consequently the minimal boundary is empty. However, even considering explicit model warped products only, a whole new zoo of examples arises.

As an example, we recall that the following family of triples (M,g,f)(M,g,f) is in fact a family of substatic triples:

M=I×Σ,g=drdrf2+r2gΣ,f=12Λn(n1)r22mrn2+q2r2n4,M=I\times\Sigma\,,\qquad g=\frac{dr\otimes dr}{f^{2}}+r^{2}g_{\Sigma},\qquad f=\sqrt{1-\frac{2\Lambda}{n(n-1)}r^{2}-\frac{2m}{r^{n-2}}+\frac{q^{2}}{r^{2n-4}}}, (1.2)

where (Σ,gΣ)(\Sigma,g_{\Sigma}) is a closed (n1)(n-1)-dimensional Riemannian manifold satisfying RicgΣ(n2)gΣ{\rm Ric}_{g_{\Sigma}}\geq(n-2)g_{\Sigma}, Λ,q\Lambda,q\in\mathbb{R}, m0m\geq 0 and I[0,+)I\subseteq[0,+\infty) is the maximal interval such that the quantity in square root in (1.2) is nonnegative for all rIr\in I. According to the sign of Λ\Lambda, the case m=q=0m=q=0 corresponds to the space forms. If instead m>0m>0, q=0q=0, one obtains the families of the Schwarzschild, Schwarzschild–de Sitter and Schwarzschild–Anti de Sitter black holes, again with respect to Λ\Lambda being vanishing, positive or negative. If m>0m>0 and q0q\neq 0, one gets the Reissner–Nordström versions of these last spaces. From a physical point of view, Λ\Lambda is the cosmological constant, mm is the mass and qq is the charge of the black hole.

We will always tacitly assume that (M,g)(M,g) is complete as a metric space. This holds true for the models (1.2), provided the absolute value of the charge qq is not too big. For instance, for Λ=0\Lambda=0, the solution has a singularity at r=0r=0 when |q|>m\lvert q\rvert>m.

The main achievement of the present work is the following sharp Isoperimetric Inequality, taking place in a relevant subclass of substatic triples. It is saturated by warped product metrics only, such as the ones in (1.2).

Theorem A (Substatic ff-isoperimetric inequality).

Let (M,g,f)(M,g,f) be a substatic triple of dimension n7n\leq 7, with horizon boundary and one uniform ff-complete end. Assume there exists an exhaustion of nonmimimal outward minimizing hypersurfaces homologous to the boundary. Then, for any bounded domain ΩΣ\Omega_{\Sigma} with smooth boundary ΩΣ=MΣ\partial\Omega_{\Sigma}=\partial M\sqcup\Sigma it holds

|Σ|nn1|M|nn1n[AVR(M,g,f)|𝕊n1|]1n1|ΩΣ|f.\lvert\Sigma\rvert^{\frac{n}{n-1}}-\lvert\partial M\rvert^{\frac{n}{n-1}}\geq n\left[\mathrm{AVR}(M,g,f)\lvert\mathbb{S}^{n-1}\rvert\right]^{\frac{1}{n-1}}\lvert\Omega_{\Sigma}\rvert_{f}. (1.3)

Moreover, in the case AVR(M,g,f)>0\mathrm{AVR}(M,g,f)>0,

  • if MØ\partial M\neq\mathchar 31\relax, the equality holds in (1.3) if and only if M\partial M is connected and (M,g)(M,g) is isometric to

    ([s¯,+)×M,dsdsf(s)2+s2s¯2gM),\left([\overline{s},+\infty)\times\partial M,\,\frac{ds\otimes ds}{f(s)^{2}}+\frac{s^{2}}{\overline{s}^{2}}\,g_{\partial M}\right), (1.4)

    where gMg_{\partial M} is the metric induced by gg on M\partial M and Σ\Sigma is a level set of ss. In particular, f=f(s)f=f(s) is a function of ss alone.

  • If M=Ø\partial M=\mathchar 31\relax, the equality holds in (1.3) if and only if (M,g)(M,g) is isometric to

    ([0,+)×𝕊n1,dsdsf(s)2+s2f(x)2g𝕊n1),\left([0,+\infty)\times\mathbb{S}^{n-1},\,\frac{ds\otimes ds}{f(s)^{2}}+\frac{s^{2}}{f(x)^{2}}\,g_{\mathbb{S}^{n-1}}\right), (1.5)

    where g𝕊n1g_{\mathbb{S}^{n-1}} is the round metric on the (n1)(n-1)-dimensional sphere 𝕊n1\mathbb{S}^{n-1} and xΩΣx\in\Omega_{\Sigma}. In this case, Σ\Sigma is a level set of ss homothetic to the round sphere. The function f=f(s)f=f(s) depends on ss alone also in this case.

The asymptotic assumptions entering in the above statement will be better understood in the next Subsection, in connection with the comparison results presented below. Concerning the quantities appearing in (1.3), we have denoted by |ΩΣ|f\lvert\Omega_{\Sigma}\rvert_{f} the weighted volume ΩΣf𝑑μ\int_{\Omega_{\Sigma}}f\,d\mu, whereas AVR(M,g,f)\mathrm{AVR}(M,g,f) is a suitable substatic generalization of the classical Asymptotic Volume Ratio for nonnegative Ricci curvature, see 1.12. When it is nonzero, inequality 1.3 in particular yields a quantitative information about the minimal boundary being in fact area minimizing, in terms of a suitable weighted volume. Observe that a priori the boundary is not even assumed to be area minimizing at all. From a more analytical point of view, formula (1.3) constitutes a nonstandard weighted isoperimetric inequality, as the perimeter is actually unweighted. The geometric intuition behind it will be given by the end of the following Subsection. One can interpret the very thoroughly recently studied Isoperimetric Inequality in nonnegative Ricci curvature [AFM20, Bre21, Ant+22, BK22, Joh21, CM22, CM22a, Poz23] as a special case of 1.3, obtained when the boundary is empty and ff is constant. The rigidity statement accordingly generalizes the one of the nonnegative Ricci curvature case.

We point out that Theorem A is particularly meaningful and perfectly sharp already in the above recalled Reissner–Nordström and Schwarzschild metrics, consisting in 1.2 for Λ=0\Lambda=0, m>0m>0, |q|<m|q|<m, and more generally in asymptotically flat substatic manifolds. With asymptotically flat we mean that the manifold converges to the Euclidean space (in a very weak sense) and that ff goes to 11 at infinity, see Definition 4.8. From the definition, it follows that an asymptotically flat end is automatically uniform and ff-complete, it possesses a natural exhaustion of coordinate spheres and it is possible to compute AVR(M,g,f)=1{\rm AVR}(M,g,f)=1. Thus, the above statement simplifies significantly.

Corollary 1.1.

Let (M,g,f)(M,g,f) be a substatic triple of dimension n7n\leq 7, with horizon boundary and one asymptotically flat end. Then, for any bounded domain ΩΣ\Omega_{\Sigma} with smooth boundary ΩΣ=MΣ\partial\Omega_{\Sigma}=\partial M\sqcup\Sigma it holds

|Σ|nn1|M|nn1n|𝕊n1|1n1|ΩΣ|f.\lvert\Sigma\rvert^{\frac{n}{n-1}}-\lvert\partial M\rvert^{\frac{n}{n-1}}\geq n\lvert\mathbb{S}^{n-1}\rvert^{\frac{1}{n-1}}\lvert\Omega_{\Sigma}\rvert_{f}. (1.6)

The same rigidity statement as in Theorem A applies in case of equality.

To our knowledge, even in the model cases, inequality 1.6 was never observed before, and does not seem to be inferable from the characterization of classical isoperimetric sets resulting from the work of Brendle [Bre13], or the earlier [Bra97, BM02] about the Schwarzschild case.

1.1. Substatic comparison geometry

Our analysis begins with the aim of working out a satisfactory substatic comparison theory, inspired by the classical nonnegative Ricci case. While, in such case, the model to be compared with is n\mathbb{R}^{n}, or more generally a cone, in the substatic generalization the model should be constituted by the large family of substatic warped products in fact appearing in the rigidity statement of Theorem A.

To pursue our goal, an initial step consists in comparing the mean curvature of geodesic spheres with that of the models. Interestingly, in order to obtain a manageable Riccati equation ruling such comparison, we are led to work in the metric g~=g/f2\tilde{g}=g/f^{2}. This is no accident: the metric g~\tilde{g} happens to fulfil the CD(0,1){\rm CD}(0,1) condition, consisting in a metric subject to

Ricg~+~2ψ+1n1dψdψ0{\rm Ric}_{\tilde{g}}+\widetilde{\nabla}^{2}\psi+\frac{1}{n-1}d\psi\otimes d\psi\geq 0 (1.7)

for some smooth function ψ\psi. To our knowledge, such explicit conformal relation was not pointed out in literature yet. However, a remarkable link is described by Li–Xia [LX17]: they come up with a family of connections with Ricci curvatures interpolating between the tensor in the left-hand side of 1.7 and the tensor in the left-hand side of 1.1. We discuss the CD(0,1)\mathrm{CD}(0,1) conformal change and Li-Xia connections in more details in Section A.3. We also point out that the conformal metric g~\tilde{g} has a natural physical interpretation in the context of static spacetimes, where it is referred to as optical metric. We give some more details on this point at the end of Section A.1.

We will denote by ρ\rho the g~\tilde{g}-distance from a point pMp\in M, or the signed g~\tilde{g}-distance from a smooth strictly mean-convex hypersurface Σ\Sigma homologous to the boundary. We give some more details on this second case, which is slightly less classical but will be crucial for the Willmore-type inequality (1.13) discussed below and in turn for the proof of Theorem A. With homologous to the boundary we mean that there exists a compact domain Ω\Omega with boundary Ω=MΣ\partial\Omega=\partial M\sqcup\Sigma, and by strictly mean-convex we understand that Σ\Sigma has pointwise positive gg-mean curvature H{\rm H} with respect to the normal pointing towards infinity. We always choose the signed distance ρ\rho to be positive in the noncompact region MΩM\setminus\Omega, that is,

ρ(x)={dg~(x,Σ)if xΩ,dg~(x,Σ)if xΩ.\rho(x)\,=\,\begin{cases}{\rm d}_{\tilde{g}}(x,\Sigma)&\hbox{if $x\not\in\Omega$},\\ -{\rm d}_{\tilde{g}}(x,\Sigma)&\hbox{if $x\in\Omega$}.\end{cases} (1.8)

Both in the case of the distance from a point and in the case of the signed distance from a hypersurface, through an analysis of the evolution of the mean curvature of the level sets of ρ\rho, we come up (see Theorem 2.5 and Proposition 2.7) with the following inequality

0<Hf=Δρ+1ff|ρn1η,0\,<\,\frac{{\rm H}}{f}\,=\,\Delta\rho+\frac{1}{f}\langle\nabla f\,|\,\nabla\rho\rangle\,\leq\,\frac{n-1}{\eta}\,, (1.9)

where H{\rm H} denotes the gg-mean curvature of a level set of the g~\tilde{g}-distance ρ\rho, and η\eta denotes an useful auxiliary function that will be called reparametrized distance. It is defined by the first order PDE 2.10, when the distance from a point is concerned, and in 2.25 when ρ\rho is the g~\tilde{g}-distance from a hypersurface. The function η\eta represents the distance along the radial geodesics computed with respect to the metric g¯=f2g=f4g~\overline{g}=f^{2}g=f^{4}\tilde{g}. This third conformal metric will not play a prominent role in the paper, but we will take some advantage from this along the proof of Theorem A. More details on this point and further comments on η\eta (in particular its relation with the weighted connection introduced by Li–Xia) may be found in Remark 2.4.

We remark that 1.9 could be derived also from [Wyl17, Theorem 3.2], rewriting it in the substatic setting thanks to the conformal relation with CD(0,1)\mathrm{CD}(0,1)-metrics. Nevertheless, we have preferred to include a full proof of it, in order to emphasize the role of η\eta and to show the substatic point of view.

A main consequence we draw out of the Laplacian Comparison Theorem above is a Bishop–Gromov Monotonicity Theorem. We state here a version substantially gathering Theorem 2.9 and Theorem 2.11 below.

Theorem B (Substatic Bishop–Gromov).

Let (M,g,f)(M,g,f) be a substatic triple. Suppose that MMM\setminus\partial M is geodesically complete with respect to the metric g~=g/f2\tilde{g}=g/f^{2}. Let ρ\rho be the g~\tilde{g}-distance function from a point or the signed g~\tilde{g}-distance function from a strictly mean-convex hypersurface Σ\Sigma homologous to M\partial M and disjoint from it. Let η\eta be the corresponding reparametrized distance, defined by 2.10 or by 2.25, and let Cutg~{\rm Cut}^{\tilde{g}} be the cut locus of the point/hypersurface. Then, for any k>0k>0, the functions

A(t)=1|𝕊n1|{ρ=t}Cutg~1ηn1𝑑σ,V(t)=1|𝔹n|tk{0ρt}ρk1fηn1𝑑μ,A(t)\,=\,\frac{1}{|\mathbb{S}^{n-1}|}\int_{\{\rho=t\}\setminus{\rm Cut}^{\tilde{g}}}\frac{1}{\eta^{n-1}}d\sigma\,,\qquad V(t)\,=\,\frac{1}{|\mathbb{B}^{n}|t^{k}}\int_{\{0\leq\rho\leq t\}}\frac{\rho^{k-1}}{f\eta^{n-1}}d\mu\,, (1.10)

are well defined and monotonically nonincreasing. Furthermore:

  • if A(t1)=A(t2)A(t_{1})=A(t_{2}) for 0<t1<t20<t_{1}<t_{2}, then the set {t1ρt2}\{t_{1}\leq\rho\leq t_{2}\} is isometric to [t1,t2]×Σ[t_{1},t_{2}]\times\Sigma, for some (n1)(n-1)-dimensional manifold (Σ,g0)(\Sigma,g_{0}), with metric

    g=f2dρdρ+η2g0;g=f^{2}\,d\rho\otimes d\rho+\eta^{2}g_{0}\,;
  • if V(t1)=V(t2)V(t_{1})=V(t_{2}) for 0<t1<t20<t_{1}<t_{2}, then the set {0ρt2}\{0\leq\rho\leq t_{2}\} is isometric to [0,t2]×Σ[0,t_{2}]\times\Sigma, for some (n1)(n-1)-dimensional manifold (Σ,g0)(\Sigma,g_{0}), with metric

    g=f2dρdρ+η2g0;g=f^{2}\,d\rho\otimes d\rho+\eta^{2}g_{0}\,;

    in the case where ρ\rho is the distance from a point xx, then ff and η\eta are functions of ρ\rho only and g0=f(x)2g𝕊n1g_{0}=f(x)^{-2}g_{\mathbb{S}^{n-1}}.

A first CD(0,1)\mathrm{CD}(0,1)-version of Bishop–Gromov monotonicity has been obtained by Wylie–Yeroshkin [WY16]. Various further generalizations then arose; dropping any attempt to be complete, we mention [KL22, LMO22, KS23], and leave the interested reader to the references therein and to Ohta’s monograph [Oht21]. However, even appealing to the conformal change g/f2g/f^{2}, formally relating the CD(0,1)\mathrm{CD}(0,1) to the substatic condition, it does not seem straightforward to deduce a Bishop–Gromov statement in the form above, that will be ruling the Willmore-type inequality 1.13 below and in turn 1.3. To the authors’ knowledge, the rigidity statements contained in Theorem B have not been considered in literature yet.

In the case where ff is constant equal to 11, then both ρ\rho and η\eta coincide with the gg-distance, hence we recover the standard Bishop–Gromov monotonicity for nonnegative Ricci tensor. Remarkably, in contrast with the standard Bishop–Gromov, in our setting we do not have to require the boundary M\partial M to be empty. This is because, since we have f=0f=0 at M\partial M, the boundary becomes an end with respect to the conformal metric g~=g/f2\tilde{g}=g/f^{2}, see Lemma 3.4. As a consequence, we do not have issues when the geodesic spheres {ρ=t}\{\rho=t\} intersect the boundary, simply because the boundary is at infinite g~\tilde{g}-distance so the intersection is always empty. On the other hand, in general g~\tilde{g}-geodesics may have finite length when going towards the ends of MM. To avoid this, we added the assumption of g~\tilde{g}-geodesic completeness in the statement. The main type of ends considered in this paper (the ff-complete ends introduced just below) will be g~\tilde{g}-geodesically complete by definition.

Proceeding in analogy with the nonnegative Ricci curvature case, we are interested in defining a suitable Asymptotic Volume Ratio, motivated by the Bishop–Gromov monotonicity above. In order to get a satisfactory notion, we first aim at understanding basic properties of the ends of substatic manifolds, at least under asymptotic assumptions on the potential of ff. A fundamental tool for this kind of study in the classical theory is Cheeger–Gromoll Splitting Theorem [CG72].

Wylie [Wyl17] in fact exploited the Laplacian Comparison Theorem to prove a splitting theorem in the CD(0,1){\rm CD}(0,1) setting, that will in turn provide surprising pieces of information in the conformal substatic setting. For our main geometric goals, the kind of end that we will be mostly interested in is that of ff-complete ends. We say that an end is ff-complete if for any gg-unit speed curve γ:[0,+)M\gamma:[0,+\infty)\to M going to infinity along the end it holds

limt+ρ(γ(t))=+,0+f(γ(t))𝑑t=+.\lim_{t\to+\infty}\rho(\gamma(t))\,=\,+\infty\,,\qquad\int_{0}^{+\infty}f(\gamma(t))dt\,=\,+\infty\,. (1.11)

The first condition is essentially asking that the end remains an end even with respect to the conformal metric g~=g/f2\tilde{g}=g/f^{2}. In other words, the first condition is equivalent to the requirement that the end is geodesically complete with respect to the metric g~\tilde{g}. The second condition instead is a way to ensure that the reparametrized distance η\eta diverges to ++\infty and is connected to an analogous definition given in [Wyl17] in the CD(0,1){\rm CD}(0,1) framework. Further discussion and comments on this definition can be found after Definition 3.1. It is easy to check that an end is ff-complete whenever there exists a constant cc such that crk<f<crkcr^{-k}<f<cr^{k} for 0<k<10<k<1 at sufficiently large distances, where rr is the gg-distance from a point (see Proposition 3.2).

We provide here the full statement of the substatic Splitting Theorem.

Theorem C (Substatic Splitting Theorem).

Let (M,g,f)(M,g,f) be a substatic triple with ends that are all ff-complete. If there is more than one end, then (M,g)(M,g) is isometric to

(×Σ,f2dsds+gΣ),(\mathbb{R}\times\Sigma,\,f^{2}\,ds\otimes ds+g_{\Sigma})\,,

for some (n1)(n-1)-dimensional Riemannian manifold (Σ,gΣ)(\Sigma,g_{\Sigma}). In particular, if M\partial M is nonempty, then (M,g,f)(M,g,f) has only one end.

Finally, we point out that similar arguments can be performed also for a different kind of ends, known as conformally compact ends, see Theorems 3.7 and 3.8. In particular we will show in Theorem 3.7 that conformally compact substatic manifolds necessarily have connected conformal infinity, generalizing a known result in the literature of static vacuum solutions [CS01]. Static vacuum solutions are in fact substatic triples such that 1.1 is satisfied with equality on the whole space.

Focusing now for simplicity on ff-complete substatic triples that have only one end, one would then be led to define the Asymptotic Volume Ratio as

AVR(M,g,f)=1|𝕊n1|limt+{ρ=t}1ηn1𝑑σ=1|𝔹n|limt+1tn{ρt}ρn1fηn1𝑑μ.{\rm AVR}(M,g,f)\,=\,\frac{1}{|\mathbb{S}^{n-1}|}\lim_{t\to+\infty}\int_{\{\rho=t\}}\frac{1}{\eta^{n-1}}d\sigma\,=\,\frac{1}{|\mathbb{B}^{n}|}\lim_{t\to+\infty}\frac{1}{t^{n}}\int_{\{\rho\leq t\}}\frac{\rho^{n-1}}{f\eta^{n-1}}d\mu\,. (1.12)

with ρ\rho denoting the g/f2g/f^{2}-distance from a mean-convex hypersurface homologous to M\partial M or from a point, if the boundary is empty. The fact that both limits above give the same result is easy to establish. However, we have to make sure that such quantity is independent of the initial hypersurface. We accomplish this task under the assumption of uniformity of the end, meaning that the quotient ηx/ηy\eta_{x}/\eta_{y} of the reparametrized distances with respect to two different points x,yx,y converges uniformly to 11 at infinity. Again, such condition is immediately checked to be fulfilled in the asymptotically flat regime. More generally, it can in fact be inferred under a natural decay condition on the gradient of ff only, see Proposition 4.3.

Exploiting the global features of our Bishop–Gromov Theorem we obtain the following Willmore-like inequality for mean-convex hypersurfaces homologous to M\partial M.

Theorem D (Substatic Willmore inequality).

Let (M,g,f)(M,g,f) be a substatic triple with one uniform ff-complete end. Let Σ\Sigma be a hypersurface homologous to the boundary. Suppose that the mean curvature H{\rm H} of Σ\Sigma with respect to the normal pointing towards infinity satisfies H>0{\rm H}>0 pointwise. Then

Σ[H(n1)f]n1𝑑σAVR(M,g,f)|𝕊n1|.\int_{\Sigma}\left[\frac{{\rm H}}{(n-1)f}\right]^{n-1}d\sigma\,\geq\,{\rm AVR}(M,g,f)\,|\mathbb{S}^{n-1}|\,. (1.13)

Furthermore, if the equality holds, then the noncompact connected component UU of MΣM\setminus\Sigma is isometric to [0,+)×Σ[0,+\infty)\times\Sigma with metric

g=f2dρdρ+η2g0,g\,=\,f^{2}d\rho\otimes d\rho+\eta^{2}g_{0}\,,

where g0g_{0} is a metric on Σ\Sigma.

Notice that if there are multiple ends, the above inequality is trivial, since in this case the Asymptotic Volume Ratio of each end is zero as pointed out in Lemma 4.7.

In the classical nonnegative Ricci curvature setting, the above result was obtained in [AFM20]. However, the more elementary proof we propose displays more resemblances with the alternative argument of Wang [Wan23].

The validity of the above Willmore-type inequality very naturally suggests the isoperimetric inequality 1.3. Indeed, assume that smooth area minimizers exist among hypersurfaces ΣV\Sigma_{V} enclosing a given weighted volume Ωf𝑑μ=V\int_{\Omega}fd\mu=V with M\partial M, for any given value VV. We will call such ΣV\Sigma_{V} ff-isoperimetric. Then, through standard variation formulas, there must exist a Lagrange multiplier λ\lambda\in\mathbb{R} such that

Σ(Hλf)φ𝑑σ=0\int_{\Sigma}({\rm H}-\lambda f)\varphi d\sigma=0 (1.14)

for any φ𝒞c(Σ)\varphi\in\mathscr{C}^{\infty}_{c}(\Sigma), implying that the mean curvature of Σ\Sigma satisfies H/f=λ{\rm H}/f=\lambda. Letting If(V)=|ΣV|I_{f}(V)=\lvert\Sigma_{V}\rvert, one has that (If)(V)(I_{f})^{\prime}(V) is proportional to λ\lambda. If this multiplier happens to positive, it is sharply estimated in terms of If(V)I_{f}(V) by 1.13. The resulting differential inequality leads to

|ΣV|nn1|ΣV0|nn1n(AVR(M,g,f)|𝕊n1|)1n1(VV0)\lvert\Sigma_{V}\rvert^{\frac{n}{n-1}}-\lvert\Sigma_{V_{0}}\rvert^{\frac{n}{n-1}}\geq n(\mathrm{AVR}(M,g,f)\lvert\mathbb{S}^{n-1}\rvert)^{\frac{1}{n-1}}(V-V_{0}) (1.15)

for any V0<VV_{0}<V. Now, if M\partial M happens to be, in addition to minimal, also area-minimizing, 1.15 directly implies 1.3 for ΣV\Sigma_{V} in the limit as V00+V_{0}\to 0^{+}. But ΣV\Sigma_{V} being the best competitor, it holds for any Σ\Sigma as in the claimed statement.

The assumptions of ff-completeness and uniformity at infinity are obviously added in order to count on the validity of Theorem D. The additional requirement of existence of a nonminimal outward minimizing exhaustion (see Section 5) is ultimately added in order to overcome the problem of the possible nonexistence of ff-isoperimetric ΣV\Sigma_{V}’s. We follow the general strategy devised by Kleiner [Kle92], and reinterpreted in the nonnegative Ricci curvature setting in [FM22], that consists in considering ff-isoperimetric sets constrained in mean-convex boxes, in our case given by the exhaustion. However, some new geometric difficulties arise, mostly given by the new portion of boundary M\partial M. They will be overcome by exploiting the fact that such boundary is in turn a priori outermost area-minimizing, a new piece of information that we obtain through an argument involving the Mean Curvature Flow of the outward minimizing exhaustion (see Proposition 5.1), and by discovering that the constrained ff-isoperimetric sets crucially never touch M\partial M, see Theorem 5.2. We will in the end have all the tools at hand to run the argument sketched above, under the usual dimensional threshold ensuring the constrained ff-isoperimetric sets to be regular enough. The strong rigidity statement contained in Theorem A will stem from the fact that in case of equality all of the ΣV\Sigma_{V} must satisfy the equality in 1.13. The rigidity statement of Theorem D will be thus complemented with the additional information given by H=λf{\rm H}=\lambda f, forcing the metric to split as 1.4.

1.2. Further directions

The results presented in this paper raise a number of natural questions, especially out of our main geometric inequalities, Theorem A and Theorem D. The Willmore-type inequality in nonnegative Ricci curvature [AFM20, Theorem 1.1] has been first obtained with a completely different technique, involving the evolution of the initial hypersurface along the level sets of a harmonic potential function. Understanding a version of such a route in the substatic context may have various interests. First of all, it may allow to remove the mean-convexity assumption on Σ\Sigma we have in Theorem D in favour of the absolute value of the mean curvature in 1.13. Secondly, and more interestingly, it would suggest the viability of a suitable version of the analysis through pp-harmonic functions performed in nonnegative curvature in [BFM22], likely leading to a new substatic Minkowski inequality, potentially stronger than our Willmore-type. Moreover, studying the behaviour of such substatic pp-harmonic functions may have implications in the existence of the weak Inverse Mean Curvature Flow [HI01] in the substatic regime, furnishing a vast extension of the important existence results in nonnegative Ricci curvature [MRS22]. Recalling the outward minimizing properties of the evolving hypersurfaces [HI01, Minimizing Hull Property 1.4], the existence of the weak IMCF would imply the a priori existence of the outward minimizing exhaustion requested in Theorem A. In the special case of asymptotically flat static vacuum solutions, the weak IMCF has already been introduced and employed to prove Minkowski-type inequalities in [Wei18, McC18, HW23]. Such inequalities are lower bound on the integral of fHf{\rm H}, and, as such, do not seem related to 1.13.

It would also be rather interesting to explore other approaches for the proof of the ff-Isoperimetric Inequality 1.3 as well, possibly allowing to remove the dimensional threshold. Antonelli–Pasqualetto–Pozzetta–Semola [Ant+22, Theorem 1.1] provided a natural and very strong proof in the nonnegative Ricci curvature case taking advantage of a generalized compactness result [Nar14, AFP22, ANP22] for isoperimetric minimizing sequences in the nonsmooth RCD\mathrm{RCD} setting. This immediately invites to study the nonsmooth counterpart of the substatic condition. A possible key for this may lie in the recent optimal transport equivalent definition of CD(0,1)\mathrm{CD}(0,1) given in [Sak21].

Another, completely different approach one may undertake consists in Brendle’s [Bre21], building on the ABP method applied to a torsion problem with Neumann condition. A substatic version of such approach promises to deal with the PDE considered in [LX17, FP22] in relation with the Heintze-Karcher inequality. We also point out that both these alternative approaches should have consequences in going beyond the dimensional threshold we imposed.

From the comparison geometry point of view, the validity of the Splitting Theorem and of the Bishop–Gromov monotonicity strongly suggests that other classical results, such as the Cheng eigenvalue estimate and Cheng–Yau gradient estimate, should have analogues in the substatic setting. A promising advance in this direction has been obtained in the CD(0,1)\mathrm{CD}(0,1) setting [Fuj22].

It may also be interesting to study compact substatic triples. Important models for this class of manifolds are given by static solutions with positive cosmological constant, most notably the Schwarzschild–de Sitter and Reissner–Nordström–de Sitter spacetimes, corresponding to (1.2) with Λ\Lambda positive and (Σ,gΣ)(\Sigma,g_{\Sigma}) a round sphere. Another natural direction is to investigate what can be said for the more general problem of studying triples (M,g,f)(M,g,f) satisfying

fRic2f+(Δf)gμg,μ.f{\rm Ric}-\nabla^{2}f+(\Delta f)g\geq-\mu g\,,\quad\mu\in\mathbb{R}\,. (1.16)

The case μ=0\mu=0 corresponds to the substatic condition. The case μ0\mu\neq 0 is also of interest: triples that saturate (1.16) for μ0\mu\neq 0 are called VV-static and are connected with the critical point equation and the Besse conjecture, see [FY19, He21] and references therein for more details on these topics. We mention that inequality (1.16) has been considered in [Zen22], where an almost-Schur inequality has been proved and exploited to generalize results in [Che14, LX19].

1.3. Structure of the paper

In Section 2 we compute the evolution of the mean curvature of geodesic spheres, leading to the aforementioned Laplacian Comparison Theorem, formula (1.9) (see Theorem 2.5). Building on it, we prove Theorem B, first for the functional AA (Theorem 2.9) and then for the functional VV (Theorem 2.11). Section 3 is dedicated to the proof of the Splitting Theorem. We first analyze the most important case of ff-complete ends and prove Theorem C (Subsection 3.2), then we discuss analogous results for conformally compact ends as well, see Theorems 3.7 and 3.8. In Section 4 we introduce the notions of uniform ends (Definition 4.1) and Asymptotic Volume Ratio (Definition 4.4) and prove the Willmore Inequality (see Theorem 4.10). Finally, in Section 5 we prove Theorem A. We include an Appendix encompassing the physical motivation for the substatic condition, the conformal relation with the CD(0,1)\mathrm{CD}(0,1) curvature-dimension condition and some additional comments.

Acknowledgements

The work was initiated during the authors’ participation at the conferenceSpecial Riemannian Metrics and Curvature Functionals held at Centro De Giorgi in Pisa in 2022. A substantial part of the work has been carried out during the authors’ attendance to the Thematic Program on Nonsmooth Riemannian and Lorentzian Geometry that took place at the Fields Institute in Toronto. They warmly thank the staff, the organizers and the colleagues for the wonderful atmosphere and the excellent working conditions set up there.

During the preparation of the work, M. F. was supported by the European Union – NextGenerationEU and by the University of Padova under the 2021 STARS Grants@Unipd programme “QuASAR”.

The authors are members of Gruppo Nazionale per l’Analisi Matematica, la Probabilità e le loro Applicazioni (GNAMPA), which is part of the Istituto Nazionale di Alta Matematica (INdAM), and are partially funded by the GNAMPA project “Problemi al bordo e applicazioni geometriche”.

The authors are grateful to Lucas Ambrozio, Gioacchino Antonelli, Luca Benatti, Camillo Brena, Philippe Castillion, Nicola Gigli, Marc Herzlich, Lorenzo Mazzieri, Marco Pozzetta and Eric Woolgar for their interest in their work and for various useful conversations.

2. Riccati comparison and Bishop-Gromov Theorem

2.1. Evolution of the mean curvature

Let (M,g,f)(M,g,f) be a substatic solution. Since ff does not vanish inside MM by definition, the metric g~=g/f2\tilde{g}=g/f^{2} is well defined in MMM\setminus\partial M. Let ρ\rho be the distance function from a point pMMp\in M\setminus\partial M with respect to the metric g~\tilde{g} and consider Riemannian polar coordinates (ρ,θ1,,θn1)(\rho,\theta^{1},\dots,\theta^{n-1}) in a neighborhood UU of pp. In this subsection we focus our computation in UU and we assume that UU does not contain points in the cut locus of pp. This guarantees that ρ\rho is smooth in UU, that |dρ|g~=1|d\rho|_{\tilde{g}}=1 at all points of UU and that the metric g~\tilde{g} has the following form:

g~=dρdρ+g~ij(ρ,θ1,,θn1)dθidθj.\tilde{g}\,=\,d\rho\otimes d\rho+\tilde{g}_{ij}(\rho,\theta^{1},\dots,\theta^{n-1})d\theta^{i}\otimes d\theta^{j}\,. (2.1)

We denote by ~\widetilde{\nabla} the Levi-Civita connection with respect to g~\tilde{g}. It is well known that the Hessian of a distance function satisfies the inequality

|~2ρ|g~2(Δg~ρ)2n1.\big{|}\widetilde{\nabla}^{2}\rho\big{|}^{2}_{\tilde{g}}\geq\frac{(\Delta_{\tilde{g}}\rho)^{2}}{n-1}\,. (2.2)

We now write down this inequality in terms of the original metric. To this end, we first compute

|ρ|2\displaystyle|\nabla\rho|^{2} =1f2,\displaystyle=\,\frac{1}{f^{2}}\,, (2.3)
~2ρ\displaystyle\widetilde{\nabla}^{2}\rho =2ρ+1f(dρdf+dfdρρ|fg),\displaystyle=\,\nabla^{2}\rho\,+\,\frac{1}{f}\left(d\rho\otimes df+df\otimes d\rho-\langle\nabla\rho\,|\,\nabla f\rangle g\right)\,,
Δg~ρ\displaystyle\Delta_{\tilde{g}}\rho =f2Δρ(n2)fρ|f.\displaystyle=\,f^{2}\Delta\rho-(n-2)f\langle\nabla\rho\,|\,\nabla f\rangle\,.

Using these identities, with some computations we can rewrite (2.2) in terms of the original metric as

|2ρ|2(Δρ)2n1n2n11f2ρ|f2+2n11fΔρρ|f+2f4|f|2.|\nabla^{2}\rho|^{2}\,\geq\,\frac{(\Delta\rho)^{2}}{n-1}-\frac{n-2}{n-1}\frac{1}{f^{2}}\langle\nabla\rho\,|\,\nabla f\rangle^{2}+\frac{2}{n-1}\frac{1}{f}\Delta\rho\langle\nabla\rho\,|\,\nabla f\rangle+\frac{2}{f^{4}}|\nabla f|^{2}\,. (2.4)

From this formula onwards we just focus on the original metric gg. From (2.1), with respect to the coordinates (ρ,θ1,,θn1)(\rho,\theta^{1},\dots,\theta^{n-1}), we have

g=f2dρdρ+gij(ρ,θ1,,θn1)dθidθj.g\,=\,f^{2}\,d\rho\otimes d\rho+g_{ij}(\rho,\theta^{1},\dots,\theta^{n-1})d\theta^{i}\otimes d\theta^{j}\,.

We are interested in the evolution of the mean curvature H{\rm H} of the level sets of ρ\rho with respect to the metric gg. We have

H=Δρ|ρ|2ρ(ρ,ρ)|ρ|3=fΔρ12f3f|ρ=fΔρ+f|ρ.{\rm H}\,=\,\frac{\Delta\rho}{|\nabla\rho|}\,-\,\frac{\nabla^{2}\rho(\nabla\rho,\nabla\rho)}{|\nabla\rho|^{3}}\,=\,f\Delta\rho\,-\,\frac{1}{2}f^{3}\langle\nabla f\,|\,\nabla\rho\rangle\,=\,f\Delta\rho\,+\,\langle\nabla f\,|\,\nabla\rho\rangle\,.

On the other hand, using the fact that |ρ|=1/f|\nabla\rho|=1/f and the Bochner formula we compute

6f4|f|22f3Δf=Δ|ρ|2= 2|2ρ|2+ 2Ric(ρ,ρ)+ 2Δρ|ρ.\frac{6}{f^{4}}|\nabla f|^{2}\,-\,\frac{2}{f^{3}}\Delta f\,=\,\Delta|\nabla\rho|^{2}\,=\,2\,|\nabla^{2}\rho|^{2}\,+\,2\,{\rm Ric}(\nabla\rho,\nabla\rho)\,+\,2\,\langle\nabla\Delta\rho\,|\,\nabla\rho\rangle\,.

Combining this with (1.1), some computations lead to

Δρ|ρΔρρ|f1f2f(ρ,ρ)|2ρ|2\langle\nabla\Delta\rho\,|\,\nabla\rho\rangle\,\leq\,\Delta\rho\langle\nabla\rho\,|\,\nabla f\rangle-\frac{1}{f}\nabla^{2}f(\nabla\rho,\nabla\rho)\,-\,|\nabla^{2}\rho|^{2}

We can use this information to find the evolution of H{\rm H}:

H|ρ\displaystyle\langle\nabla{\rm H}\,|\,\nabla\rho\rangle\, f|2ρ|2+2f3|f|2+Δρρ|f\displaystyle\leq\,-f|\nabla^{2}\rho|^{2}\,+\,\frac{2}{f^{3}}|\nabla f|^{2}\,+\,\Delta\rho\langle\nabla\rho\,|\,\nabla f\rangle (2.5)
f(Δρ)2n1+n3n1Δρρ|f+n2n11f(ρ|f)2\displaystyle\leq\,-f\frac{(\Delta\rho)^{2}}{n-1}\,+\,\frac{n-3}{n-1}\Delta\rho\langle\nabla\rho\,|\,\nabla f\rangle\,+\,\frac{n-2}{n-1}\frac{1}{f}\left(\langle\nabla\rho\,|\,\nabla f\rangle\right)^{2} (2.6)
=1n11fH2+1fHρ|f,\displaystyle=\,-\frac{1}{n-1}\frac{1}{f}{\rm H}^{2}\,+\,\frac{1}{f}{\rm H}\langle\nabla\rho\,|\,\nabla f\rangle\,, (2.7)

where in the second inequality we have used estimate (2.4). This formula can be rewritten as

(Hf)|ρ1n1(Hf)2.\bigg{\langle}\nabla\left(\frac{{\rm H}}{f}\right)\,\bigg{|}\,\nabla\rho\bigg{\rangle}\,\leq\,-\frac{1}{n-1}\left(\frac{{\rm H}}{f}\right)^{2}\,.

In other words, we have found the following formula for the evolution of the mean curvature of the level sets of ρ\rho.

Lemma 2.1.

In the notations above, at any point of UU the evolution of the mean curvature H{\rm H} along ρ\rho satisfies

ρ(Hf)1n1H2.\frac{\partial}{\partial\rho}\left(\frac{{\rm H}}{f}\right)\,\leq\,-\frac{1}{n-1}\,{\rm H}^{2}\,. (2.8)

2.2. Bounds on the mean curvature and Laplacian comparison

Let pMMp\in M\setminus\partial M and ρ\rho be the distance function from pp with respect to the metric g~=g/f2\tilde{g}=g/f^{2}. We assume that MMM\setminus\partial M is geodesically complete with respect to the metric g~\tilde{g}. We denote by Cutpg~(M){\rm Cut}^{\tilde{g}}_{p}(M) the cut locus of pp, again with respect to g~\tilde{g}. The function ρ\rho is smooth in U=M(Cutpg~(M){p})U=M\setminus({\rm Cut}^{\tilde{g}}_{p}(M)\cup\{p\}), where Cutpg~(M){\rm Cut}^{\tilde{g}}_{p}(M) is the cut locus of pp with respect to the metric g~\tilde{g}.

For every θ𝕊n1TpM\theta\in\mathbb{S}^{n-1}\subset T_{p}M, we denote by σθ\sigma_{\theta} the geodesic starting from pp in the direction θ\theta and by τ(θ)(0,+]\tau(\theta)\in(0,+\infty] the smallest positive value such that σθ(τ(θ))Cutpg~(M)\sigma_{\theta}(\tau(\theta))\in{\rm Cut}^{\tilde{g}}_{p}(M). We recall from [Car92, Proposition 2.9, Chapter 13] that τ:𝕊n1(0,+]\tau:\mathbb{S}^{n-1}\to(0,+\infty] is a continuous function. Notice that there is a diffeomorphism between UU and the set

{(ρ,θ)(0,+)×𝕊n1:ρ<τ(θ)},\{(\rho,\theta)\in(0,+\infty)\times\mathbb{S}^{n-1}\,:\,\rho<\tau(\theta)\}\,,

hence we can use ρ,θ\rho,\theta as coordinates in UU.

We can now exploit Lemma 2.1 to find a bound for H{\rm H} in UU. To this end, given a positive function η𝒞\eta\in\mathscr{C}^{\infty} we use (2.8) to compute

ρ(fHηn1)=f2H2ρ(Hf)1n1ηρ1n1(f2ηρ).\frac{\partial}{\partial\rho}\left(\frac{f}{{\rm H}}-\frac{\eta}{n-1}\right)\,=\,-\frac{f^{2}}{{\rm H}^{2}}\frac{\partial}{\partial\rho}\left(\frac{{\rm H}}{f}\right)-\frac{1}{n-1}\frac{\partial\eta}{\partial\rho}\,\geq\,\frac{1}{n-1}\left(f^{2}-\frac{\partial\eta}{\partial\rho}\right)\,. (2.9)

We then choose η\eta so that the the right hand side vanishes pointwise. Since ff is smooth, the equation

ηρ=f2\frac{\partial\eta}{\partial\rho}\,=\,f^{2}

can be solved and yields a unique solution once we fix its value on a level set of ρ\rho.

Proposition 2.2.

There exists a unique function η𝒞(U)\eta\in\mathscr{C}^{\infty}(U) satisfying

{ηρ=f2,limρ0+η= 0.\begin{dcases}\frac{\partial\eta}{\partial\rho}\,=\,f^{2}\,,\\ \lim_{\rho\to 0^{+}}\eta\,=\,0\,.\end{dcases} (2.10)
Remark 2.3.

It should be remarked that η\eta is not even necessarily continuous outside UU. In fact, if there are two minimizing geodesics from pp to qCutpg~(M)q\in{\rm Cut}^{\tilde{g}}_{p}(M), the function η\eta may behave differently on the two geodesics, hence the limit of η\eta as we approach qq along the two different geodesics would give different results.

Proof.

For every ε>0\varepsilon>0, consider the function ηε\eta_{\varepsilon} defined by

{ηερ=f2 in {ρ>ε}U,ηε= 0on {ρ=ε}.\begin{dcases}\frac{\partial\eta_{\varepsilon}}{\partial\rho}\,=\,f^{2}&\hbox{ in }\{\rho>\varepsilon\}\cap U\,,\\ \eta_{\varepsilon}\,=\,0&\hbox{on }\{\rho=\varepsilon\}\,.\end{dcases} (2.11)

Since ff is smooth, it is well known (see for instance [Eva10, Section 3.2.4]) that (2.11) can be solved and yields a unique 𝒞2\mathscr{C}^{2} solution. Furthermore, by differentiating the first equation we find that the first derivatives αη\partial_{\alpha}\eta also solve a first order PDE, namely ραη=αf2\partial_{\rho}\partial_{\alpha}\eta=\partial_{\alpha}f^{2}. Since αf2\partial_{\alpha}f^{2} is smooth, it follows that the derivatives αη\partial_{\alpha}\eta are also 𝒞2\mathscr{C}^{2}. Proceeding this way, we deduce that ηε\eta_{\varepsilon} is smooth.

It is now sufficient to pass to the limit as ε0\varepsilon\to 0 using Ascoli–Arzelà. Since the functions ηε\eta_{\varepsilon} (and their derivatives as well) are uniformly continuous and uniformly bounded on any compact domain inside UU, it follows that ηε\eta_{\varepsilon} converge to a smooth function η\eta, defined on the whole UU.

Concerning uniqueness, if there were two different solutions η1,η2\eta_{1},\eta_{2} of (2.10), then the difference η1η2\eta_{1}-\eta_{2} would have derivative equal to zero along the direction /ρ\partial/\partial\rho. Since the limit as ρ0\rho\to 0 is zero, one immediately obtains η1η2=0\eta_{1}-\eta_{2}=0. ∎

Remark 2.4 (The reparametrized distance η\eta).

The function η\eta is also called reparametrized distance. The reason for this terminology is that the radial g~\tilde{g}-geodesics from our point pp, reparametrized with respect to η\eta, are geodesics for the weighted connection

DXY=XY+1fg(X,Y)f.{\rm D}_{X}Y=\nabla_{X}Y+\frac{1}{f}g(X,Y)\nabla f\,.

The significance of such connection is that the Ricci tensor associated to D{\rm D} is nonnegative if and only if the substatic condition is satisfied. More details on this can be found in Section A.3 and in [LX17] (see also [KWY19, WY16] for further discussions in the conformally related CD(0,1){\rm CD(0,1)} setting). Alternatively, as mentioned in the Introduction, η\eta can also be seen to represent the distance along radial g~\tilde{g}-geodesics with respect to the metric g¯=f2g\overline{g}=f^{2}g. In fact, if σ:[0,S]M\sigma:[0,S]\to M is a radial g~\tilde{g}-geodesic with σ(0)=p\sigma(0)=p and σ˙=/ρ\dot{\sigma}=\partial/\partial\rho, we have |σ˙(s)|g~=1|\dot{\sigma}(s)|_{\tilde{g}}=1 and so

dg¯(γ(S),p)=0S|σ˙(s)|g¯𝑑s=0Sf2(σ(s))𝑑s=η(S).{\rm d}_{\overline{g}}(\gamma(S),p)\,=\,\int_{0}^{S}|\dot{\sigma}(s)|_{\overline{g}}ds\,=\,\int_{0}^{S}f^{2}(\sigma(s))ds\,=\,\eta(S)\,.

We are now in a position to prove the following crucial bound on the mean curvature of the level sets of ff. This bound can be naturally interpreted as the Laplacian Comparison for the conformal distance function ρ\rho. This result corresponds to [Wyl17, Theorem 3.2] in the CD(0,1){\rm CD}(0,1) framework.

Theorem 2.5 (Laplacian Comparison).

Let (M,g,f)(M,g,f) be a substatic triple. Suppose that MMM\setminus\partial M is geodesically complete with respect to the metric g~=g/f2\tilde{g}=g/f^{2}. Let ρ\rho be the distance function to a point pMMp\in M\setminus\partial M with respect to the metric g~=g/f2\tilde{g}=g/f^{2} and η\eta be the solution to (2.10). Then the mean curvature H{\rm H} of the level sets of ρ\rho with respect to the metric gg satisfies

0<Hf=Δρ+1ff|ρn1η0\,<\,\frac{{\rm H}}{f}\,=\,\Delta\rho+\frac{1}{f}\langle\nabla f\,|\,\nabla\rho\rangle\,\leq\,\frac{n-1}{\eta} (2.12)

in the classical sense in the open dense set U=M(Cutpg~(M){p})U=M\setminus({\rm Cut}^{\tilde{g}}_{p}(M)\cup\{p\}).

Proof.

Let us first prove the thesis working inside the open set UU. From the definition of η\eta and (2.9), we immediately deduce

ρ(fHηn1) 0.\frac{\partial}{\partial\rho}\left(\frac{f}{{\rm H}}-\frac{\eta}{n-1}\right)\,\geq\,0\,.

In other words, the function f/Hη/(n1)f/{\rm H}-\eta/(n-1) is nondecreasing.

We then estimate its value near the point pp. It is well known that, for small ρ\rho’s, the mean curvature Hg~{\rm H}_{\tilde{g}} of the geodesic balls grows as the geodesic balls in Euclidean space, namely Hg~=(n1)/ρ+o(1/ρ){\rm H}_{\tilde{g}}=(n-1)/\rho+o(1/\rho) when ρ\rho is sufficiently small. Here we have denoted by Hg~{\rm H}_{\tilde{g}} the mean curvature with respect to the metric g~\tilde{g}. This is related to the mean curvature with respect to gg by H=Hg~/f+(n1)ρ|f{\rm H}={\rm H}_{\tilde{g}}/f+(n-1)\langle\nabla\rho\,|\,\nabla f\rangle. Since ρ|f\langle\nabla\rho\,|\,\nabla f\rangle is bounded in a neighborhood of pp, we obtain

H=n1f(p)ρ+o(1/ρ),as ρ0.{\rm H}=\frac{n-1}{f(p)\rho}+o(1/\rho)\,,\quad\hbox{as }\rho\to 0\,.

In particular, f/H0f/{\rm H}\to 0 as ρ0\rho\to 0. Since η0\eta\to 0 as well by definition, we have obtained that f/Hη/(n1)0f/{\rm H}-\eta/(n-1)\to 0 when ρ0\rho\to 0. From the monotonicity of f/Hη/(n1)f/{\rm H}-\eta/(n-1) we then deduce (n1)f/Hη(n-1)f/{\rm H}\geq\eta on the whole UU. Since η\eta is positive on the whole manifold by construction, the conclusion follows. ∎

A first important observation is that there is a more effective version of the above inequality. We now argue that the Laplacian comparison obtained in Theorem 2.5 and Proposition 2.7 gives us a vector with nonnegative divergence, which is

X=fηn1ρ,X\,=\,\frac{f}{\eta^{n-1}}\nabla\rho\,, (2.13)

defined on the open dense set U=M(Cutpg~(M){p})U=M\setminus({\rm Cut}^{\tilde{g}}_{p}(M)\cup\{p\}). In fact:

divX\displaystyle{\rm div}\,X\, =fηn1[Δρ+1ff|ρ](n1)fηnη|ρ\displaystyle=\,\frac{f}{\eta^{n-1}}\left[\Delta\rho\,+\,\frac{1}{f}\langle\nabla f\,|\,\nabla\rho\rangle\right]\,-\,(n-1)\frac{f}{\eta^{n}}\langle\nabla\eta\,|\,\nabla\rho\rangle (2.14)
(n1)fηn(n1)fηn1f2ηρ= 0.\displaystyle\leq\,(n-1)\frac{f}{\eta^{n}}\,-\,(n-1)\frac{f}{\eta^{n}}\frac{1}{f^{2}}\frac{\partial\eta}{\partial\rho}\,=\,0\,. (2.15)

The inequality divX0{\rm div}\,X\leq 0 holds then in the classical sense in the whole UU. It is crucial that this inequality actually holds in the distributional sense in the whole manifold.

Theorem 2.6.

Let (M,g,f)(M,g,f) be a substatic triple. Suppose that MMM\setminus\partial M is geodesically complete with respect to the metric g~=g/f2\tilde{g}=g/f^{2}. Let ρ\rho be the distance function to a point pp with respect to the metric g~=g/f2\tilde{g}=g/f^{2} and η\eta be the solution to (2.10). Then the vector XX defined in (2.13) has nonpositive divergence in the weak sense in the whole M{p}M\setminus\{p\}. Namely, for every nonnegative test function χ𝒞c(M)\chi\in\mathscr{C}_{c}^{\infty}(M) with χ0\chi\equiv 0 in a neighborhood of pp, it holds

MX|χ𝑑σ 0.\int_{M}\langle X\,|\,\nabla\chi\rangle d\sigma\,\geq\,0\,.
Proof.

Let g\sqrt{g} be the volume element of the metric gg with respect to the coordinates (ρ,θ)(\rho,\theta). We first observe that at all points of U=M(Cutpg~(M){p})U=M\setminus({\rm Cut}^{\tilde{g}}_{p}(M)\cup\{p\}) it holds

ρ(gfηn1)\displaystyle\frac{\partial}{\partial\rho}\left(\frac{\sqrt{g}}{f\eta^{n-1}}\right)\, =ρ(fn1g~ηn1)\displaystyle=\,\frac{\partial}{\partial\rho}\left(\frac{f^{n-1}\sqrt{\tilde{g}}}{\eta^{n-1}}\right) (2.16)
=[fn1Hg~ηn1(n1)fn+1ηn+(n1)fn2ηn1fρ]g~\displaystyle=\,\left[\frac{f^{n-1}{\rm H}_{\tilde{g}}}{\eta^{n-1}}\,-\,(n-1)\frac{f^{n+1}}{\eta^{n}}\,+\,(n-1)\frac{f^{n-2}}{\eta^{n-1}}\frac{\partial f}{\partial\rho}\right]\sqrt{\tilde{g}} (2.17)
=fnηn1[H(n1)1f2fρ(n1)fη+(n1)1f2fρ]g~\displaystyle=\,\frac{f^{n}}{\eta^{n-1}}\left[{\rm H}-(n-1)\frac{1}{f^{2}}\frac{\partial f}{\partial\rho}\,-\,(n-1)\frac{f}{\eta}\,+\,(n-1)\frac{1}{f^{2}}\frac{\partial f}{\partial\rho}\right]\sqrt{\tilde{g}} (2.18)
 0,\displaystyle\leq\,0\,, (2.19)

where the last inequality follows from the Laplacian comparison. Recalling that in polar coordinates (ρ,θ)(\rho,\theta) the set UU is diffeomorphic to the set of pairs (ρ,θ)(\rho,\theta) with ρ<τ(θ)\rho<\tau(\theta) for a suitable continuous function τ:𝕊n1(0,+]\tau:\mathbb{S}^{n-1}\to(0,+\infty], we can then compute

MX|χ𝑑μ\displaystyle\int_{M}\langle X\,|\,\nabla\chi\rangle d\mu\, =𝕊n10τ(θ)fηn1ρ|χg𝑑ρ𝑑θ\displaystyle=\,\int_{\mathbb{S}^{n-1}}\int_{0}^{\tau(\theta)}\frac{f}{\eta^{n-1}}\langle\nabla\rho\,|\,\nabla\chi\rangle\sqrt{g}d\rho d\theta (2.20)
=𝕊n10τ(θ)χρgfηn1𝑑ρ𝑑θ\displaystyle=\,\int_{\mathbb{S}^{n-1}}\int_{0}^{\tau(\theta)}\frac{\partial\chi}{\partial\rho}\frac{\sqrt{g}}{f\eta^{n-1}}d\rho d\theta (2.21)
=𝕊n10τ(θ)χρ(gfηn1)𝑑ρ𝑑θ+{θ𝕊n1:τ(θ)<}[χgfηn1](τ(θ),θ)𝑑θ\displaystyle=\,-\int_{\mathbb{S}^{n-1}}\!\int_{0}^{\tau(\theta)}\!\!\chi\,\frac{\partial}{\partial\rho}\left(\frac{\sqrt{g}}{f\eta^{n-1}}\right)\!d\rho d\theta+\!\int_{\{\theta\in\mathbb{S}^{n-1}:\,\tau(\theta)<\infty\}}\left[\chi\frac{\sqrt{g}}{f\eta^{n-1}}\right]\!(\tau(\theta),\theta)d\theta (2.22)
limε0+𝕊n1[χgfηn1](ε,θ)𝑑θ.\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad-\lim_{\varepsilon\to 0^{+}}\int_{\mathbb{S}^{n-1}}\left[\chi\frac{\sqrt{g}}{f\eta^{n-1}}\right]\!(\varepsilon,\theta)d\theta\,. (2.23)

In the last identity, the first integral is nonnegative thanks to (2.19) whereas the second integral is nonnegative by construction. Concerning the third and final integral, notice that η\eta behaves near pp as f2(p)ρf^{2}(p)\rho by definition, hence the integral is easily seen to converge to χ(p)|𝕊n1|/f2n1(p)\chi(p)|\mathbb{S}^{n-1}|/f^{2n-1}(p) as ε0\varepsilon\to 0. Since we are assuming that χ\chi vanishes in a neighborhood of pp, the third integral goes to zero as ε0\varepsilon\to 0. The conclusion follows. ∎

2.3. Evolution of the mean curvature of hypersurfaces

Let Σ\Sigma be a compact smooth hypersurface. We are interested to the case where Σ\Sigma is homologous to the boundary: namely, there exists a compact domain Ω\Omega such that Ω=MΣ\partial\Omega=\partial M\sqcup\Sigma. Let ρ\rho be the g~\tilde{g}-distance from Σ\Sigma in MΩM\setminus\Omega. If Σ\Sigma is smooth, then so is ρ\rho in a collar of Σ\Sigma. Under the usual assumption of g~\tilde{g}-geodesic completeness of MMM\setminus\partial M, it is known from [MM03, Proposition 4.6] that the open set U=(MΩ)CutΣg~(M)U=(M\setminus\Omega)\setminus{\rm Cut}^{\tilde{g}}_{\Sigma}(M) of the points where ρ\rho is smooth is dense in MΩM\setminus\Omega. In particular, there is a function τ:Σ(0,+]\tau:\Sigma\to(0,+\infty] such that the gradient flow of /ρ\partial/\partial\rho gives a diffeomorphism between UU and

{(ρ,x):xΣ,ρ(0,τ(x))}.\{(\rho,x)\,:\,x\in\Sigma\,,\ \rho\in(0,\tau(x))\}\,. (2.24)

It is convenient to estimate the evolution of the geometry of Σ\Sigma with respect to these coordinates. If H>0{\rm H}>0 at the point xΣx\in\Sigma, the evolution of the mean curvature H(0,x){\rm H}(0,x) is quite similar to the one described for geodesic spheres in Subsection 2.1 and 2.2. One can then define the function η𝒞(U)\eta\in\mathscr{C}^{\infty}(U) as the solution to

{ηρ(ρ,x)=f2(ρ,x),η(0,x)=(n1)f(0,x)H(0,x).\begin{dcases}\frac{\partial\eta}{\partial\rho}(\rho,x)\,=\,f^{2}(\rho,x)\,,\\ \eta(0,x)\,=\,(n-1)\frac{f(0,x)}{{\rm H}(0,x)}\,.\end{dcases} (2.25)

As in Proposition 2.2, one shows that η\eta is well defined and smooth in UU. We can then replicate the proof of Theorem 2.5 to find H/f(ρ,x)(n1)/η(ρ,x){\rm H}/f(\rho,x)\leq(n-1)/\eta(\rho,x) for any 0ρ<τ(x)0\leq\rho<\tau(x). In this case, the proof is actually even easier, since f/Hη/(n1)f/{\rm H}-\eta/(n-1) vanishes at xx by construction, without the need of proving it.

Notice that from (2.8) we can get interesting information on the evolution of the mean curvature also in the case where H{\rm H} is nonpositive at a point xΣx\in\Sigma. If H<0{\rm H}<0 at xx, we deduce from (2.8) that H{\rm H} must remain negative for all times, whereas if H=0{\rm H}=0 then H{\rm H} remains nonpositive. Furthermore, even when H{\rm H} is nonpositive one can still define η\eta by (2.25) and find η/(n1)f/H<0\eta/(n-1)\leq f/{\rm H}<0 at the points (ρ,x)(\rho,x). From this fact we can obtain even more information under the assumption that the end is ff-complete. We recall that an end is ff-complete if (1.11) is satisfied, see also Definition 3.1 below for a more extensive discussion about this notion. The main feature of ff-complete ends is that the end is complete with respect to the metric g~\tilde{g} (so that ρ\rho goes to ++\infty) and η\eta grows to ++\infty along the end. Since η(x)\eta(x) is negative when the mean curvature at xx is negative, in particular η\eta must reach the value zero, at which point the bound η/(n1)f/H<0\eta/(n-1)\leq f/{\rm H}<0 fails. This implies that the line ρ(ρ,x)\rho\mapsto(\rho,x) must reach the cut locus before η\eta hits zero. Finally, if H=0{\rm H}=0 then from (2.8) we would get that H{\rm H} remains nonpositive. If at some point H{\rm H} becomes negative, then the previous argument applies and the line ρ(ρ,x)\rho\mapsto(\rho,x) must reach the cut locus. Summarizing, we have obtained the following:

Proposition 2.7.

Let (M,g,f)(M,g,f) be a substatic triple. Suppose that MMM\setminus\partial M is geodesically complete with respect to the metric g~=g/f2\tilde{g}=g/f^{2}. Let ρ\rho be the distance from an hypersurface Σ\Sigma homologous to the boundary with respect to the metric g~=g/f2\tilde{g}=g/f^{2}, and let η\eta be the solution to (2.25). Finally, let xΣx\in\Sigma and consider the evolution H(ρ,x){\rm H}(\rho,x) of the mean curvature in the direction of the end.

  • (i)(i)

    If H(0,x)>0{\rm H}(0,x)>0 then for any 0<ρ<τ(x)0<\rho<\tau(x) it holds

    0<Hf(ρ,x)n1η(ρ,x).0\,<\,\frac{{\rm H}}{f}(\rho,x)\,\leq\,\frac{n-1}{\eta(\rho,x)}\,. (2.26)
  • (ii)(ii)

    If H(0,x)<0{\rm H}(0,x)<0, then for every 0<ρ<τ(x)0<\rho<\tau(x) it holds

    Hf(ρ,x)n1η(ρ,x)< 0.\frac{{\rm H}}{f}(\rho,x)\,\leq\,\frac{n-1}{\eta(\rho,x)}\,<\,0\,. (2.27)

    Furthermore, if the ends of MM are ff-complete then τ(x)<+\tau(x)<+\infty.

  • (iii)(iii)

    If H(0,x)=0{\rm H}(0,x)=0, then H(ρ,x)0{\rm H}(\rho,x)\leq 0 for every 0<ρ<τ(x)0<\rho<\tau(x). Furthermore, if the ends of MM are ff-complete and τ(x)=+\tau(x)=+\infty, then H(ρ,x)=0{\rm H}(\rho,x)=0 for all ρ0\rho\geq 0.

In the following we will focus on the case where the hypersurface Σ\Sigma is homologous to M\partial M and strictly mean-convex, meaning that Σ\Sigma has positive mean curvature H{\rm H} with respect to the normal pointing outside Ω\Omega. In this case, Proposition 2.7-(i)(i) tells us that the bound H/f(n1)/η{\rm H}/f\leq(n-1)/\eta is in place on the whole U=M(ΩCutΣg~)U=M\setminus(\Omega\cup{\rm Cut}^{\tilde{g}}_{\Sigma}). Furthermore, the vector field (2.13), that we recall here for convenience,

X=fηn1ρ,X\,=\,\frac{f}{\eta^{n-1}}\nabla\rho\,, (2.28)

is also well defined on UU. We can now proceed exactly as in the proof of Theorem 2.6 to show that the vector field XX has nonnegative divergence in the weak sense.

Theorem 2.8.

Let (M,g,f)(M,g,f) be a substatic triple. Suppose that MMM\setminus\partial M is geodesically complete with respect to the metric g~=g/f2\tilde{g}=g/f^{2}. Let Σ\Sigma be a strictly mean-convex hypersurface homologous to M\partial M and disjoint from it. Suppose that the mean curvature H{\rm H} of Σ\Sigma with respect to the normal pointing towards infinity satisfies H>0{\rm H}>0 pointwise. Let ρ\rho be the g~\tilde{g}-distance function from Σ\Sigma and η\eta be the solution to (2.25). Then the vector XX defined in (2.28) has nonpositive divergence in the weak sense in the whole MΩM\setminus\Omega. Namely, for every nonnegative test function χ𝒞c(MΩ)\chi\in\mathscr{C}_{c}^{\infty}(M\setminus\Omega) it holds

MΩX|χ𝑑σ 0.\int_{M\setminus\Omega}\langle X\,|\,\nabla\chi\rangle d\sigma\,\geq\,0\,.

2.4. Growth of weighted areas and volumes

In this subsection, we exploit the monotonicity of the mean curvature of the level sets to deduce a Bishop–Gromov-type theorem for the behaviour of areas and volumes of geodesic spheres. We first study the monotonicity of the following functional

A(t)=1|𝕊n1|{ρ=t}Cutg~1ηn1𝑑σ,A(t)\,=\,\frac{1}{|\mathbb{S}^{n-1}|}\int_{\{\rho=t\}\setminus{\rm Cut}^{\tilde{g}}}\frac{1}{\eta^{n-1}}d\sigma\,, (2.29)

where ρ\rho is the distance function from a point or the signed distance from a strictly mean-convex hypersurface homologous to the boundary, with respect to the metric g~\tilde{g}, whereas Cutg~{\rm Cut}^{\tilde{g}} is the cut locus of the point/hypersurface with respect to g~\tilde{g}. It is important that we remove the cut locus from the domain of the integral, as we have observed in Remark 2.3 that the function η\eta is not well defined on it. When ρ\rho is the distance from a point, the function AA can be written in polar coordinates as

A(t)=1|𝕊n1|{θ𝕊n1:τ(θ)>t}g(t,θ)ηn1(t,θ)𝑑θ,A(t)\,=\,\frac{1}{|\mathbb{S}^{n-1}|}\int_{\{\theta\in\mathbb{S}^{n-1}\,:\,\tau(\theta)>t\}}\frac{\sqrt{g}(t,\theta)}{\eta^{n-1}(t,\theta)}d\theta\,, (2.30)

where we recall that τ(θ)\tau(\theta) is the minimum value of ρ\rho such that the point with coordinate (ρ,θ)(\rho,\theta) belongs to the cut locus. An analogous definition can of course be given for the distance from an hypersurface using coordinates (2.24). Notice that τ\tau is a continuous function, hence the domain of the integral in (2.30) is measurable, meaning that the function AA is well defined for all t(0,+)t\in(0,+\infty). The domain of the integral shrinks as tt increases, whereas the integrand is positive and continuous, hence it is easily seen that for all values a(0,+)a\in(0,+\infty) it holds

lim inftaA(t)A(a)lim supta+A(t).\liminf_{t\to a^{-}}A(t)\,\geq\,A(a)\,\geq\,\limsup_{t\to a^{+}}A(t)\,. (2.31)

Furthermore, notice that if the cut locus intersects {ρ=a}\{\rho=a\} in a set with positive n1\mathscr{H}^{n-1}-measure, then the first inequality is strict, that is lim inftaA(t)A(a)\liminf_{t\to a^{-}}A(t)\,\geq\,A(a). We are finally ready to state the first main result of this subsection.

Theorem 2.9.

Let (M,g,f)(M,g,f) be a substatic triple. Suppose that MMM\setminus\partial M is geodesically complete with respect to the metric g~=g/f2\tilde{g}=g/f^{2}. Let ρ\rho be the g~\tilde{g}-distance function from a point or the signed g~\tilde{g}-distance function from a strictly mean-convex hypersurface Σ\Sigma homologous to M\partial M and disjoint from it. Let η\eta be the corresponding reparametrized distance, defined by 2.10 or by 2.25, and let Cutg~{\rm Cut}^{\tilde{g}} be the cut locus of the point/hypersurface. Then the function

A(t)=1|𝕊n1|{ρ=t}Cutg~1ηn1𝑑σA(t)\,=\,\frac{1}{|\mathbb{S}^{n-1}|}\int_{\{\rho=t\}\setminus{\rm Cut}^{\tilde{g}}}\frac{1}{\eta^{n-1}}d\sigma

is monotonically nonincreasing.

Furthermore, if A(t1)=A(t2)A(t_{1})=A(t_{2}) for any 0<t1<t20<t_{1}<t_{2}, then the set U={t1ρt2}U=\{t_{1}\leq\rho\leq t_{2}\} is isometric to [t1,t2]×Σ[t_{1},t_{2}]\times\Sigma with metric

g=f2dρdρ+η2g0,g\,=\,f^{2}d\rho\otimes d\rho+\eta^{2}g_{0}\,,

where g0g_{0} is a metric on the level set Σ\Sigma. In UU the functions ff and η\eta satisfy

1ηηθi=ψηφ1ηn1,1ffθi=ψη+n22φ1ηn1,\frac{1}{\eta}\frac{\partial\eta}{\partial\theta^{i}}\,=\,\psi\eta-\varphi\frac{1}{\eta^{n-1}}\,,\qquad\frac{1}{f}\frac{\partial f}{\partial\theta^{i}}\,=\,\psi\eta+\frac{n-2}{2}\varphi\frac{1}{\eta^{n-1}}\,, (2.32)

where φ,ψ\varphi,\psi are independent of ρ\rho.

Remark 2.10.

If we set f1f\equiv 1 then ρ\rho is the distance function with respect to gg and η=ρ\eta=\rho, hence A(t)=|{ρ=t}Cutg~|/(|𝕊n1|tn1)A(t)=|\{\rho=t\}\setminus{\rm Cut}^{\tilde{g}}|/(|\mathbb{S}^{n-1}|t^{n-1}) and the above monotonicity becomes completely analogous to the standard Bishop–Gromov monotonicity of the areas of geodesic spheres when Ric0{\rm Ric}\geq 0, which concerns the function |{ρt}|/(|𝕊n1|tn1)|\partial\{\rho\leq t\}|/(|\mathbb{S}^{n-1}|t^{n-1}). Clearly, the two functions coincide almost everywhere, except at the values tt such that {ρ=t}Cutg~\{\rho=t\}\cap{\rm Cut}^{\tilde{g}} has nonzero n1\mathscr{H}^{n-1}-measure. Notice that the number of values for which this may happen is necessarily countable. A way to see this is to observe that, as mentioned below formula (2.31), every value such that {ρ=t}Cutg~\{\rho=t\}\cap{\rm Cut}^{\tilde{g}} has nonzero n1\mathscr{H}^{n-1}-measure must correspond to a jump of AA, and these are at most countable since AA has bounded variation (this is shown in the proof below).

Proof.

If the cut locus were empty, the proof of the monotonicity of AA would follow easily by integrating the inequality divX0{\rm div}\,X\leq 0 between two level sets of ρ\rho, where XX is the vector field defined in (2.13), and then applying the Divergence Theorem. In the general case, in order to take into account the lack of smoothness of ρ\rho at the cut locus, we will need a more refined analysis, that we now discuss.

Since ρ\rho is a g~\tilde{g}-distance function, in particular it is locally Lipschitz, hence its gradient is well defined almost everywhere. Furthermore, as highlighted in [BFM22, Proposition 2.1], ρ\rho being Lipschitz also implies that the coarea formula can be applied to the level sets of ρ\rho. In particular, for any 0<a<b<+0<a<b<+\infty we have

abA(t)𝑑t=ab({ρ=t}Cutg~1ηn1𝑑σ)𝑑t={a<ρ<b}1fηn1𝑑μ<+.\int_{a}^{b}A(t)dt\,=\,\int_{a}^{b}\left(\int_{\{\rho=t\}\setminus{\rm Cut}^{\tilde{g}}}\frac{1}{\eta^{n-1}}d\sigma\right)dt\,=\,\int_{\{a<\rho<b\}}\frac{1}{f\eta^{n-1}}d\mu\,<\,+\infty\,.

In the last integral we did not have to specify that we are not integrating on Cutg~{\rm Cut}^{\tilde{g}}, since Cutg~{\rm Cut}^{\tilde{g}} is negligible when integrating on a volume. The above tells us that AA is locally integrable. Consider now a test function χ𝒞c((0,+))\chi\in\mathscr{C}^{\infty}_{c}((0,+\infty)) and let XX be the vector field defined by (2.13). We then compute

M(χρ)|X𝑑μ\displaystyle\int_{M}\langle\nabla(\chi\circ\rho)\,|\,\,X\rangle\,d\mu =Mχ(ρ)X|ρ𝑑μ\displaystyle=\,\int_{M}\chi^{\prime}(\rho)\langle X\,|\,\nabla\rho\rangle d\mu (2.33)
=Mχ(ρ)fηn1|ρ|2𝑑μ\displaystyle=\,\int_{M}\chi^{\prime}(\rho)\frac{f}{\eta^{n-1}}|\nabla\rho|^{2}d\mu (2.34)
=0+{ρ=t}χ(t)fηn1|ρ|𝑑σ𝑑t\displaystyle=\,\int_{0}^{+\infty}\int_{\{\rho=t\}}\chi^{\prime}(t)\frac{f}{\eta^{n-1}}|\nabla\rho|\,d\sigma\,dt (2.35)
=|𝕊n1|0+χ(t)A(t)𝑑t.\displaystyle=\,|\mathbb{S}^{n-1}|\int_{0}^{+\infty}\chi^{\prime}(t)A(t)\,dt\,. (2.36)

On the other hand, Theorem 2.6 (when ρ\rho is the distance from a point) and Theorem 2.8 tells us that the first integral in the above chain of identities is nonnegative whenever the test function χ\chi is nonnegative. More precisely, from (2.23), since χ(0)=0\chi(0)=0, we have

M(χρ)|X𝑑μ=𝕊n10τ(θ)χ(ρ)ρ(gfηn1)𝑑ρ𝑑θ+{θ𝕊n1:τ(θ)<}[χgfηn1](τ(θ),θ)𝑑θ 0,\int_{M}\langle\nabla(\chi\circ\rho)\,|\,X\rangle d\mu\,=-\int_{\mathbb{S}^{n-1}}\!\int_{0}^{\tau(\theta)}\!\!\!\chi(\rho)\,\frac{\partial}{\partial\rho}\!\left(\frac{\sqrt{g}}{f\eta^{n-1}}\right)d\rho d\theta\\ \,+\int_{\{\theta\in\mathbb{S}^{n-1}:\,\tau(\theta)<\infty\}}\!\left[\chi\frac{\sqrt{g}}{f\eta^{n-1}}\right]\!\!(\tau(\theta),\theta)d\theta\,\geq\,0\,, (2.37)

where as usual τ(θ)\tau(\theta) is the minimum value of ρ\rho such that the point with coordinate (ρ,θ)(\rho,\theta) belongs to the cut locus. Combining (2.37) with the above chain of identities we have obtained 0+χ(t)A(t)𝑑t0\int_{0}^{+\infty}\chi^{\prime}(t)A(t)dt\geq 0 for any nonnegative test function. If we knew AA to be weakly differentiable, this would force its weak derivative to be nonpositive thus proving that AA is nonincreasing. However, we have no information on the regularity of the function AA at the moment. In the following we will show that AA has bounded variation, which will be enough to infer the desired monotonicity.

If we fix 0<a<b<+0<a<b<+\infty, for any χ𝒞([a,b])\chi\in\mathscr{C}^{\infty}([a,b]) with χ=1\|\chi\|_{\infty}=1 we have from (2.37) the following bound

M(χρ)|X𝑑μ\displaystyle\int_{M}\langle\nabla(\chi\circ\rho)\,|\,X\rangle d\mu\, {θ𝕊n1:τ(θ)>a}amin{τ(θ),b}ρ(gfηn1)𝑑ρ𝑑θ\displaystyle\leq-\int_{\{\theta\in\mathbb{S}^{n-1}:\,\tau(\theta)>a\}}\!\int_{a}^{\min\{\tau(\theta),b\}}\!\!\frac{\partial}{\partial\rho}\!\left(\frac{\sqrt{g}}{f\eta^{n-1}}\right)d\rho d\theta (2.38)
+{θ𝕊n1:aτ(θ)b}[gfηn1](τ(θ),θ)𝑑θ\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\,+\int_{\{\theta\in\mathbb{S}^{n-1}:\,a\leq\tau(\theta)\leq b\}}\!\left[\frac{\sqrt{g}}{f\eta^{n-1}}\right]\!\!(\tau(\theta),\theta)\,d\theta (2.39)
={θ𝕊n1:τ(θ)>a}[gfηn1](a,θ)𝑑θ{θ𝕊n1:τ(θ)>b}[gfηn1](b,θ)𝑑θ.\displaystyle=\,\int_{\{\theta\in\mathbb{S}^{n-1}:\,\tau(\theta)>a\}}\left[\frac{\sqrt{g}}{f\eta^{n-1}}\right]\!\!(a,\theta)\,d\theta\,-\int_{\{\theta\in\mathbb{S}^{n-1}:\,\tau(\theta)>b\}}\left[\frac{\sqrt{g}}{f\eta^{n-1}}\right]\!\!(b,\theta)\,d\theta\,. (2.40)

In other words, the quantity M(χρ)|X𝑑μ\int_{M}\langle\nabla(\chi\circ\rho)\,|\,X\rangle d\mu, and thus also 0+χ(t)A(t)𝑑t\int_{0}^{+\infty}\chi^{\prime}(t)A(t)\,dt, is bounded from above by a constant that depends on aa, bb, ff and gg, but not on χ\chi. It follows that AA has bounded variation in [a,b][a,b]. As a consequence, the signed finite Radon measure μ\mu on (a,b)(a,b) defined by μ((c,d))=limtdA(t)limtc+A(t)\mu((c,d))=\lim_{t\to d^{-}}A(t)-\lim_{t\to c^{+}}A(t) for any a<c<d<ba<c<d<b, is such that

0+χ(t)A(t)𝑑t=0+χ(t)μ(dt).\int_{0}^{+\infty}\chi^{\prime}(t)A(t)\,dt\,=\,-\int_{0}^{+\infty}\chi(t)\mu(dt)\,.

Since we have already shown that 0+χ(t)A(t)𝑑t0\int_{0}^{+\infty}\chi^{\prime}(t)A(t)\,dt\geq 0 for any nonnegative test function χ\chi, it follows that the measure μ\mu is nonpositive. From the definition of μ\mu and (2.31), we deduce then that the function AA is monotonically nonincreasing in (a,b)(a,b). Since this should hold for any 0<a<b<+0<a<b<+\infty, it must necessarily hold on the whole (0,+)(0,+\infty). This proves that AA is monotonically nonincreasing.

It remains to prove the rigidity statement. If A(t1)=A(t2)A(t_{1})=A(t_{2}), then thanks to the discussion above it follows A(t)=A(t1)A(t)=A(t_{1}) for all t1<t<t2t_{1}<t<t_{2}. As a consequence, for any test function χ\chi supported in [t1,t2][t_{1},t_{2}], we get that the last line in the computation (2.36) vanishes, that is, M(χρ)|X𝑑μ=0\int_{M}\langle\nabla(\chi\circ\rho)\,|\,\,X\rangle\,d\mu=0. On the other hand, this integral can also be computed as in (2.37). From the fact that χ0\chi\geq 0 and from (2.19), we know that the two terms on the right hand side of (2.37) are both nonnegative, hence they must both vanish for all χ𝒞c([0,+))\chi\in\mathscr{C}^{\infty}_{c}([0,+\infty)). This implies that τ(θ)\tau(\theta) never belongs to (t1,t2)(t_{1},t_{2}), meaning that the cut locus does not intersect {t1<ρ<t2}\{t_{1}<\rho<t_{2}\} and that equality is achieved in (2.19). In other words, the following holds

Hf=Δρ+1ff|ρ=n1η in {t1ρt2}.\frac{{\rm H}}{f}\,=\,\Delta\rho+\frac{1}{f}\langle\nabla f\,|\,\nabla\rho\rangle\,=\,\frac{n-1}{\eta}\qquad\hbox{ in }\{t_{1}\leq\rho\leq t_{2}\}\,. (2.41)

This identity in turn triggers the equality in the estimates made in Subsection 2.1, namely

|~2ρ|g~2=(Δg~ρ)2n1,Ric(~ρ,~ρ)=(n1)[1f~2f(~ρ,~ρ)2f2~f|~ρ2],\big{|}\widetilde{\nabla}^{2}\rho\big{|}^{2}_{\tilde{g}}\,=\,\frac{(\Delta_{\tilde{g}}\rho)^{2}}{n-1}\,,\qquad{\rm Ric}\big{(}\widetilde{\nabla}\rho,\widetilde{\nabla}\rho\big{)}\,=\,(n-1)\left[\frac{1}{f}\widetilde{\nabla}^{2}f\big{(}\widetilde{\nabla}\rho,\widetilde{\nabla}\rho\big{)}-\frac{2}{f^{2}}\big{\langle}\widetilde{\nabla}f\,|\,\widetilde{\nabla}\rho\big{\rangle}^{2}\right]\,, (2.42)

where we recall that ~\widetilde{\nabla} is the Levi-Civita connection with respect to g~\tilde{g}. Notice that, since |~ρ|g~=1|\widetilde{\nabla}\rho|_{\tilde{g}}=1, for any vector XX it holds

~2ρ(~ρ,X)=~|~ρ|g~2|Xg~= 0.\widetilde{\nabla}^{2}\rho\big{(}\widetilde{\nabla}\rho,X\big{)}\,=\,\big{\langle}\widetilde{\nabla}|\widetilde{\nabla}\rho|_{\tilde{g}}^{2}\,|\,X\big{\rangle}_{\tilde{g}}\,=\,0\,.

It follows immediately from this and the first equation in (2.42) that, in the coordinates in which g~\tilde{g} has the form (2.1), for any i,j=1,,n1i,j=1,\dots,n-1 it holds

~ij2ρ=Δg~ρn1g~ij=(f2η1ffρ)g~ij.\widetilde{\nabla}^{2}_{ij}\rho\,=\,\frac{\Delta_{\tilde{g}}\rho}{n-1}\tilde{g}_{ij}\,=\,\left(\frac{f^{2}}{\eta}-\frac{1}{f}\frac{\partial f}{\partial\rho}\right)\tilde{g}_{ij}\,.

where the latter identity makes use of (2.3). On the other hand, from the definition of Hessian we have ~ij2ρ=Γijρ=ρg~ij/2\widetilde{\nabla}^{2}_{ij}\rho=-\Gamma_{ij}^{\rho}=\partial_{\rho}\tilde{g}_{ij}/2, hence

g~ijρ= 2(f2η1ffρ)g~ij= 2ρ(logηlogf)g~ij.\frac{\partial\tilde{g}_{ij}}{\partial\rho}\,=\,2\left(\frac{f^{2}}{\eta}-\frac{1}{f}\frac{\partial f}{\partial\rho}\right)\tilde{g}_{ij}\,=\,2\frac{\partial}{\partial\rho}\left(\log\eta-\log f\right)\tilde{g}_{ij}\,.

This identity can be solved explicitly, yielding

g~ij=η2f2(g0)ij,\tilde{g}_{ij}\,=\,\frac{\eta^{2}}{f^{2}}(g_{0})_{ij}\,,

where (g0)ij(g_{0})_{ij} does not depend on ρ\rho. Comparing with (2.1) and recalling g=f2g~g=f^{2}\tilde{g}, we have obtained

g=f2dρdρ+η2g0.g\,=\,f^{2}d\rho\otimes d\rho+\eta^{2}g_{0}\,. (2.43)

The functions ff and η\eta may be functions of both the radial coordinate ρ\rho and of {θ1,,θn1}\{\theta^{1},\dots,\theta^{n-1}\}. Any metric gg having the form (2.43) satisfies the substatic condition with equality in the radial direction, that is:

fRρρρρ2f+(Δf)gρρ= 0.f{\mathrm{R}}_{\rho\rho}\,-\,\nabla^{2}_{\rho\rho}f\,+\,(\Delta f)g_{\rho\rho}\,=\,0\,.

From this identity and the substatic condition we find out that, for any vector X=/ρ+λ/iX=\partial/\partial\rho+\lambda\partial/\partial_{i}, it holds

0\displaystyle 0\, [fRic2f(Δf)g](X,X)\displaystyle\leq\,\left[f{\rm Ric}\,-\,\nabla^{2}f-(\Delta f)g\right](X,X) (2.44)
=λ2[fRiiii2f+(Δf)gii]+ 2λ[fRiρiρ2f+(Δf)giρ].\displaystyle=\,\lambda^{2}\,\left[f{\mathrm{R}}_{ii}\,-\,\nabla^{2}_{ii}f\,+\,(\Delta f)g_{ii}\right]\,+\,2\lambda\,\left[f{\mathrm{R}}_{i\rho}\,-\,\nabla^{2}_{i\rho}f\,+\,(\Delta f)g_{i\rho}\right]\,. (2.45)

Since this inequality holds pointwise for any λ\lambda\in\mathbb{R}, it follows that

fRiρiρ2f+(Δf)giρ= 0.f{\mathrm{R}}_{i\rho}\,-\,\nabla^{2}_{i\rho}f\,+\,(\Delta f)g_{i\rho}\,=\,0\,. (2.46)

Recalling the expression (2.43) for the metric, a direct computation gives us that (2.46) is equivalent to

(n2)iρ2G+iρ2F(n1)ρGiF= 0,(n-2)\partial^{2}_{i\rho}G\,+\,\partial^{2}_{i\rho}F-(n-1)\partial_{\rho}G\partial_{i}F\,=\,0\,, (2.47)

where F=logfF=\log f, G=logηG=\log\eta. On the other hand, since ρη=f2\partial_{\rho}\eta=f^{2}, we have ρG=e2FG\partial_{\rho}G=e^{2F-G}, from which we compute

iρ2G 2iFρG+iGρG= 0.\partial_{i\rho}^{2}G\,-\,2\partial_{i}F\partial_{\rho}G\,+\,\partial_{i}G\partial_{\rho}G\,=\,0\,. (2.48)

We will now combine (2.47) and (2.48) in two different ways.

On the one hand, if we subtract (n1)(n-1) times equation (2.48) from equation (2.47), we obtain

iρ2F+(n1)iFρG=iρ2G+(n1)ρGiG,\partial^{2}_{i\rho}F\,+\,(n-1)\partial_{i}F\partial_{\rho}G\,=\,\partial^{2}_{i\rho}G\,+\,(n-1)\partial_{\rho}G\partial_{i}G\,,

which can be rewritten as

ρ(e(n1)GiF)=ρ(e(n1)GiG).\partial_{\rho}\left(e^{(n-1)G}\partial_{i}F\right)\,=\,\partial_{\rho}\left(e^{(n-1)G}\partial_{i}G\right)\,.

In other words, we have found

e(n1)G(iFiG)=φ(θ),i=1,,n1.e^{(n-1)G}(\partial_{i}F-\partial_{i}G)=\varphi(\theta)\,,\ \ \forall i=1,\dots,n-1\,. (2.49)

On the other hand, subtracting (n/21)(n/2-1) times equation (2.48) from (2.47), we obtain

iρ2FiFρG=n22iρ2G+n22ρGiG,\partial^{2}_{i\rho}F\,-\,\partial_{i}F\partial_{\rho}G\,=\,-\frac{n-2}{2}\partial^{2}_{i\rho}G\,+\,\frac{n-2}{2}\partial_{\rho}G\partial_{i}G\,,

which can be rewritten as

ρ(eGiF)=n22ρ(eGiG),\partial_{\rho}\left(e^{-G}\partial_{i}F\right)\,=\,-\frac{n-2}{2}\partial_{\rho}\left(e^{-G}\partial_{i}G\right)\,,

that gives

eG(iFn22iG)=ψ(θ),i=1,,n1.e^{-G}\left(\partial_{i}F-\frac{n-2}{2}\partial_{i}G\right)=\psi(\theta)\,,\ \ \forall i=1,\dots,n-1\,. (2.50)

Writing (2.49) and (2.50) in terms of η\eta and ff, with straightforward computations we obtain formulas (2.32). ∎

An immediate consequence of the above result is that we can find a bound for the area functional A(t)A(t) by taking its limit as t0t\to 0. In the case where ρ\rho is the g~\tilde{g}-distance from a hypersurface, then A(0)A(0) is in fact well defined and we have A(t)A(0)A(t)\leq A(0). This will be exploited later, in Subsection 4.2 to prove Theorem D.

We focus now briefly on the case where ρ\rho is the g~\tilde{g}-distance from a point xx. As ρ0\rho\to 0 we have η/ρ=f(x)2+o(1)\partial\eta/\partial\rho=f(x)^{2}+o(1), hence η=f(x)2ρ+o(ρ)\eta=f(x)^{2}\rho+o(\rho). Furthermore, dσ=(1+o(1))f(x)n1dσg~=(1+o(1))f(x)n1ρn1dσ𝕊n1d\sigma=(1+o(1))f(x)^{n-1}d\sigma_{\tilde{g}}=(1+o(1))f(x)^{n-1}\rho^{n-1}d\sigma_{\mathbb{S}^{n-1}}. It follows that

limt0+A(t)=1f(x)n1.\lim_{t\to 0^{+}}A(t)\,=\,\frac{1}{f(x)^{n-1}}\,.

As a consequence of the monotonicity of AA we then deduce that, for every t>0t>0, it holds

A(t)1f(x)n1.A(t)\,\leq\,\frac{1}{f(x)^{n-1}}\,. (2.51)

If the equality holds, then the rigidity statement in Theorem 2.9 applies in {ρt}\{\rho\leq t\}.

Building on Theorem 2.9, we can also show the following volumetric version of the Bishop–Gromov monotonicity theorem.

Theorem 2.11.

Let (M,g,f)(M,g,f) be a substatic triple. Suppose that MMM\setminus\partial M is geodesically complete with respect to the metric g~=g/f2\tilde{g}=g/f^{2}. Let ρ\rho be the g~\tilde{g}-distance function from a point or the signed g~\tilde{g}-distance function from a strictly mean-convex hypersurface Σ\Sigma homologous to M\partial M and disjoint from it. Let η\eta be the corresponding reparametrized distance, defined by 2.10 or by 2.25. Then, for any k>0k>0, the function

V(t)=1|𝔹n|tk{0ρt}ρk1fηn1𝑑μV(t)\,=\,\frac{1}{|\mathbb{B}^{n}|t^{k}}\int_{\{0\leq\rho\leq t\}}\frac{\rho^{k-1}}{f\eta^{n-1}}d\mu (2.52)

is well defined and monotonically nonincreasing.

Furthermore, if V(t1)=V(t2)V(t_{1})=V(t_{2}) for 0<t1<t20<t_{1}<t_{2}, then the set U={0ρt2}U=\{0\leq\rho\leq t_{2}\} is isometric to [0,t2]×Σ[0,t_{2}]\times\Sigma with metric

g=f2dρdρ+η2g0,g\,=\,f^{2}d\rho\otimes d\rho+\eta^{2}g_{0}\,,

where g0g_{0} is a metric on the level sets Σ\Sigma. In UU the functions ff and η\eta satisfy

1ηηθi=ψηφ1ηn1,1ffθi=ψη+n22φ1ηn1,\frac{1}{\eta}\frac{\partial\eta}{\partial\theta^{i}}\,=\,\psi\eta-\varphi\frac{1}{\eta^{n-1}}\,,\qquad\frac{1}{f}\frac{\partial f}{\partial\theta^{i}}\,=\,\psi\eta+\frac{n-2}{2}\varphi\frac{1}{\eta^{n-1}}\,, (2.53)

where φ,ψ\varphi,\psi are independent of ρ\rho. If ρ\rho is the distance from a point xx, then g0=f(x)2g𝕊n1g_{0}=f(x)^{-2}g_{\mathbb{S}^{n-1}} and φ=ψ=0\varphi=\psi=0 in the whole UU, that is, both ff and η\eta are functions of ρ\rho in UU.

Remark 2.12.

Recall that η\eta is smooth outside the cut locus. Since the cut locus has finite n1\mathscr{H}^{n-1}-measure, the functional VV in the statement is well posed. When k=nk=n we have recovered the standard Bishop–Gromov monotonicity for volumes of geodesic spheres for Ric0{\rm Ric}\geq 0. Indeed if one sets f=1f=1 in the statement above, so that η=ρ\eta=\rho, one gets V(t)=|{ρt}|/(|𝔹n|tn)V(t)=|\{\rho\leq t\}|/(|\mathbb{B}^{n}|t^{n}).

Proof.

We start by observing that the coarea formula (together with the fact that |ρ|=1/f|\nabla\rho|=1/f) gives the following relation between the functionals AA and VV:

V(t)=n|𝕊n1|tk0t({ρ=τ}τk1ηn1𝑑σ)𝑑τ=ntk0tτk1A(τ)𝑑τ.V(t)\,=\,\frac{n}{|\mathbb{S}^{n-1}|t^{k}}\int_{0}^{t}\left(\int_{\{\rho=\tau\}}\frac{\tau^{k-1}}{\eta^{n-1}}d\sigma\right)d\tau\,=\,\frac{n}{t^{k}}\int_{0}^{t}\tau^{k-1}A(\tau)d\tau\,. (2.54)

From Theorem 2.9, we know that for almost every τt\tau\leq t the area integral A(τ)A(\tau) is well defined and that it is nonincreasing in τ\tau. In the case where ρ\rho is the distance from a point xx, we have observed in (2.51) that A(τ)A(\tau) is bounded by the constant 1/f(x)n11/f(x)^{n-1}. If instead ρ\rho is the distance from a strictly mean-convex hypersurface, then A(0)A(0) is well defined and we have A(τ)A(0)A(\tau)\leq A(0). In both cases, it holds A(τ)CA(\tau)\leq\mathrm{C} for some constant C\mathrm{C}, hence from (2.54), recalling k>0k>0, we compute

V(t)Ctk0tτk1𝑑τ=Ck.V(t)\,\leq\,\frac{\mathrm{C}}{t^{k}}\int_{0}^{t}\tau^{k-1}d\tau\,=\,\frac{\mathrm{C}}{k}\,.

As a consequence, V(t)V(t) is well defined. Furthermore, the monotonicity of AA also implies A(t)A(τ).A(t)\leq A(\tau). Plugging this information in (2.54) gives

V(t)ntkA(t)0tτk1𝑑τ=nkA(t).V(t)\,\geq\,\frac{n}{t^{k}}A(t)\int_{0}^{t}\tau^{k-1}d\tau\,=\,\frac{n}{k}A(t)\,. (2.55)

With this information at hand, we are ready to compute the derivative of V(t)V(t):

V(t)\displaystyle V^{\prime}(t)\, =limε0[V(t+ε)V(t)]/ε\displaystyle=\,\lim_{\varepsilon\to 0}[V(t+\varepsilon)-V(t)]/\varepsilon (2.56)
=limε01ε|𝔹n|(1(t+ε)k1tk){ρt+ε}ρk1fηn1𝑑μ+limε01tk1ε|𝔹n|{tρt+ε}ρk1fηn1𝑑μ\displaystyle=\,\lim_{\varepsilon\to 0}\frac{1}{\varepsilon|\mathbb{B}^{n}|}\left(\frac{1}{(t+\varepsilon)^{k}}-\frac{1}{t^{k}}\right)\int_{\{\rho\leq t+\varepsilon\}}\frac{\rho^{k-1}}{f\eta^{n-1}}d\mu\,+\,\lim_{\varepsilon\to 0}\frac{1}{t^{k}}\frac{1}{\varepsilon|\mathbb{B}^{n}|}\int_{\{t\leq\rho\leq t+\varepsilon\}}\frac{\rho^{k-1}}{f\eta^{n-1}}d\mu (2.57)
=limε0tk(t+ε)kεtkV(t+ε)+limε0ntk1εtt+ετk1A(τ)𝑑τ\displaystyle=\,\lim_{\varepsilon\to 0}\frac{t^{k}-(t+\varepsilon)^{k}}{\varepsilon t^{k}}V(t+\varepsilon)\,+\,\lim_{\varepsilon\to 0}\frac{n}{t^{k}}\frac{1}{\varepsilon}\int_{t}^{t+\varepsilon}\tau^{k-1}A(\tau)d\tau (2.58)
=ktV(t)+ntA(t)\displaystyle=\,-\frac{k}{t}V(t)+\frac{n}{t}A(t) (2.59)
 0.\displaystyle\leq\,0\,. (2.60)

We now prove the rigidity statement. If V(t1)=V(t2)V(t_{1})=V(t_{2}) for two values 0<t1<t20<t_{1}<t_{2}, then retracing our computations we find out that A(τ)=A(t)A(\tau)=A(t) for all 0<τ<t20<\tau<t_{2}. From the rigidity statement of Theorem 2.9 we then deduce that in {0ρt}\{0\leq\rho\leq t\} the metric writes as

g=f2dρdρ+η2g0,g\,=\,f^{2}d\rho\otimes d\rho+\eta^{2}g_{0}\,,

and f,ηf,\eta satisfy formulas (2.53).

Finally, we suppose now that ρ\rho is the distance from a point xx and we prove that f,ηf,\eta must necessarily depend on ρ\rho only. To this end, notice that formulas (2.53) must hold up to ρ=0\rho=0 (that is, up to the point xx), hence at the limit as ρ\rho goes to zero, the derivative f/θi\partial f/\partial\theta^{i} goes to zero. Since η\eta goes to 0 as ρ0\rho\to 0, it follows then from the second formula in (2.53) that φ\varphi must vanish identically.

As a consequence, the first formula in (2.53) can be rewritten as

1η2ηθi=ψ.\frac{1}{\eta^{2}}\frac{\partial\eta}{\partial\theta^{i}}\,=\,\psi\,. (2.61)

In particular, taking the derivative with respect to ρ\rho we deduce that 2(1/η)/ρθ=0\partial^{2}(1/\eta)/\partial\rho\partial\theta=0. In other words, 1/η=α+β1/\eta=\alpha+\beta, where α\alpha is a function of ρ\rho and β\beta is a function of the θi\theta^{i}’s. Substituting this expression for η\eta in (2.61), we deduce that

βθi=ψ\frac{\partial\beta}{\partial\theta^{i}}\,=\,-\psi

On the other hand, taking the difference of the two formulas in (2.53) we have

θi(logηlogf)= 0.\frac{\partial}{\partial\theta^{i}}\left(\log\eta-\log f\right)\,=\,0\,.

In other words, the quantity η/f\eta/f must be a function of ρ\rho. Recalling the decomposition 1/η=α+β1/\eta=\alpha+\beta shown right above, it follows then that 1/f=λ(α+β)1/f=\lambda(\alpha+\beta), where λ\lambda is a function of ρ\rho.

Notice now that, when ρ\rho goes to zero, the limit of ff must go to the value of ff at the point xx, so that in particular the limit of 1/f1/f as ρ0\rho\to 0 must not depend on θi\theta^{i}. It follows that β=0\beta=0, hence ψ=β/θi\psi=-\partial\beta/\partial\theta^{i} vanishes as well. We have proved that both φ\varphi and ψ\psi in (2.53) vanish, hence both ff and η\eta must be functions of the sole ρ\rho in the whole {ρt}\{\rho\leq t\}. Since the metric

g~=dρdρ+η2f2g0\tilde{g}=d\rho\otimes d\rho+\frac{\eta^{2}}{f^{2}}g_{0}

is smooth at the point xx, it follows that (η2/f2)g0(\eta^{2}/f^{2})g_{0} should be close to ρ2g𝕊n1\rho^{2}g_{\mathbb{S}^{n-1}} near xx. From the definition of η\eta it follows η=f(x)2ρ+o(ρ)\eta=f(x)^{2}\rho+o(\rho) close to xx, hence g0=f(x)2g𝕊n1g_{0}=f(x)^{-2}g_{\mathbb{S}^{n-1}} and we conclude the rigidity statement. ∎

3. Wylie’s Splitting Theorem for substatic manifolds

3.1. ff-complete and conformally compact ends

From now on we will study noncompact manifolds with some special behaviour at infinity, focusing mainly on ff-complete ends.

Definition 3.1.

We say that an end is ff-complete if for any gg-unit speed curve γ:[0,+)M\gamma:[0,+\infty)\to M going to infinity along the end it holds

limt+ρ(γ(t))=+,0+f(γ(t))𝑑t=+,\lim_{t\to+\infty}\rho(\gamma(t))\,=\,+\infty\,,\qquad\int_{0}^{+\infty}f(\gamma(t))dt\,=\,+\infty\,, (3.1)

where ρ\rho is the distance from a point with respect to g~=g/f2\tilde{g}=g/f^{2}.

It is clear from the triangle inequality that the definition above does not depend on the point we are taking the distance ρ\rho from. It also would not change if we replace the distance from a point with the distance from a hypersurface.

For all the arguments that follows it would actually be enough to require (3.1) only along g~\tilde{g}-geodesics. More precisely, it is enough to require the end to be g~\tilde{g}-complete and satisfying the second condition in (3.1) along any g~\tilde{g}-geodesic. In fact, the above definition is analogous to the one given in [Wyl17, Definition 6.2] in the CD(0,1){\rm CD}(0,1) framework: there, a triple (M,g~,ψ)(M,\tilde{g},\psi) satisfying the CD(0,1){\rm CD}(0,1) condition is said to be ψ\psi-complete if for any g~\tilde{g}-geodesic σ:[0,+)M\sigma:[0,+\infty)\to M going to infinity along the end it holds

0+e2ψ(σ(t))n1𝑑t=+.\int_{0}^{+\infty}e^{-\frac{2\psi(\sigma(t))}{n-1}}dt\,=\,+\infty\,.

Recalling the relations g~=g/f2\tilde{g}=g/f^{2} and ψ=(n1)logf\psi=-(n-1)\log f between the CD(0,1){\rm CD}(0,1) and substatic setting (see Section A.2), it is easily seen that this integrability condition is equivalent to the second requirement in (3.1). As already observed in [Wyl17], this integrability condition can be interpreted as completeness with respect to the metric f2gf^{2}g or, alternatively, as completeness with respect to the weighted connection introduced in [WY16] and [LX17] (see Section A.3). For what concerns this paper however, the only relevance of the second condition in (3.1) is that it implies that the reparametrized distance η\eta defined in Section 2 goes to infinity along the end. This is easy to show as follows. Let ρ\rho be the g~\tilde{g}-distance to a point or hypersurface and let η\eta be defined by (2.10) or (2.25) depending on whether we are taking the distance from a point or hypersurface. Let σ:[0,+)M\sigma:[0,+\infty)\to M be a g~\tilde{g}-geodesic with σ˙=~ρ\dot{\sigma}=\widetilde{\nabla}\rho and let γ:[0,+)M\gamma:[0,+\infty)\to M be the reparametrization of σ\sigma that has gg-length constant and equal to 11. We then have

η(γ(t))η(γ(0))=0tf2(γ(t))|γ˙(t)|g~𝑑t=0tf(γ(t))|γ˙(t)|g𝑑t=0tf(γ(t))𝑑t,\eta(\gamma(t))-\eta(\gamma(0))\,=\,\int_{0}^{t}f^{2}(\gamma(t))|\dot{\gamma}(t)|_{\tilde{g}}dt\,=\,\int_{0}^{t}f(\gamma(t))|\dot{\gamma}(t)|_{g}dt\,=\,\int_{0}^{t}f(\gamma(t))dt\,,

hence if the second condition in (3.1) holds then η\eta goes to ++\infty.

The family of ff-complete ends includes a number of interesting examples. Most notably, asymptotically flat ends are ff-complete. We say that (M,g,f)(M,g,f) is asymptotically flat if there exists a compact set KK such that MKM\setminus K is diffeomorphic to n\mathbb{R}^{n} minus a ball, the metric gg converges to the Euclidean metric and ff goes to 11 at infinity along the end. A precise definition of asymptotic flatness is given below, see Definition 4.8. A notable example of asymptotically flat substatic solution is the Reissner–Nordström solution, corresponding to (1.2) with Λ=0\Lambda=0. In fact, the family of ff-complete ends is quite more general: for instance, it is sufficient to require a suitable behaviour of ff at infinity, without any assumption on the topology and geometry of the end, as clarified by the following proposition.

Proposition 3.2.

Let (M,g,f)(M,g,f) be a substatic triple and let rr be the gg-distance from a point (or more generally from a compact domain). If there exist a compact set KMK\supset\partial M and constants 0<c<C0<c<C, 0<k<10<k<1 such that

crk<f<Crkcr^{-k}<f<Cr^{k} (3.2)

at all points in MKM\setminus K, then all ends are ff-complete.

Proof.

Let γ:[0,+)M\gamma:[0,+\infty)\to M be a gg-unit speed curve going to infinity along the end and let δ\delta be the gg-distance between γ(0)\gamma(0) and the point pp we are taking the distance rr from (if rr is the distance from a compact domain instead, it is sufficient to choose δ\delta as the maximum distance between γ(0)\gamma(0) and the points of the domain; the rest of the proof is easy to adapt). It is also convenient to assume that γ(0)\gamma(0) and pp belong to KK and that KK is geodesically convex with respect to the metric g~=g/f2\tilde{g}=g/f^{2} (this can of course always be achieved by possibly enlarging KK).

Since γ\gamma has unit speed, we have d(γ(0),γ(t))<td(\gamma(0),\gamma(t))<t, hence by triangle inequality

r(γ(t))<t+δ.r(\gamma(t))<t+\delta\,.

If we then denote by TT the maximum value of tt such that γ(t)K\gamma(t)\in K, estimate (3.2) tells us that for any t>Tt>T it holds

c(t+δ)k<f(γ(t))<C(t+δ)k.c\,(t+\delta)^{-k}\,<\,f(\gamma(t))\,<\,C\,(t+\delta)^{k}\,.

In particular, since 0<k<10<k<1 we have

T+f(γ(t))𝑑t>cT+(t+δ)k𝑑t>climt+(t+δ)1k(T+δ)1k1k=+.\int_{T}^{+\infty}f(\gamma(t))dt\,>\,c\,\int_{T}^{+\infty}(t+\delta)^{-k}dt\,>\,c\lim_{t\to+\infty}\frac{(t+\delta)^{1-k}-(T+\delta)^{1-k}}{1-k}\,=\,+\infty\,.

To conclude, it is sufficient to show that the g~\tilde{g}-distance ρ\rho of γ(t)\gamma(t) from a fixed point (we will take γ(0)\gamma(0) for simplicity) also goes to ++\infty as t+t\to+\infty. For any fixed t>0t>0, let σt\sigma_{t} be the unit-speed g~\tilde{g}-geodesic from γ(0)\gamma(0) to γ(t)\gamma(t). We reparametrize σt\sigma_{t} so that it has speed 11 with respect to the metric gg. With a slight abuse of notation, we still denote by σt\sigma_{t} the reparametrized curve. We will have σt:[0,τ]M\sigma_{t}:[0,\tau]\to M, with τt\tau\geq t. Since we have chosen KK to be g~\tilde{g}-geodesically convex, there will exist a value TtT_{t} such that σt(s)K\sigma_{t}(s)\in K for all sTts\leq T_{t} and σt(s)K\sigma_{t}(s)\not\in K for all s>Tts>T_{t}. Clearly τTt>dg(K,γ(t))\tau-T_{t}>{\rm d}_{g}(K,\gamma(t)). Furthermore, since σt\sigma_{t} restricted to [0,Tt][0,T_{t}] is g~\tilde{g}-minimizing and both σt(0)=γ(0)\sigma_{t}(0)=\gamma(0) and σt(Tt)\sigma_{t}(T_{t}) belong to KK, we have

diamg~(K)>0Tt|σ˙t(s)|g~𝑑s=0Tt1f(σt(s))𝑑s>TtmaxKf.{\rm diam}_{\tilde{g}}(K)\,>\,\int_{0}^{T_{t}}|\dot{\sigma}_{t}(s)|_{\tilde{g}}ds\,=\,\int_{0}^{T_{t}}\frac{1}{f(\sigma_{t}(s))}ds>\frac{T_{t}}{\max_{K}f}\,.

We are now ready to estimate the g~\tilde{g}-distance as follows:

ρ(γ(t))=0τ|σ˙t(s)|g~𝑑s=0τ1f(σt(s))|σ˙t(s)|g𝑑s=0τ1f(σt(s))𝑑s>Ttτr(σt(s))kC𝑑s.\rho(\gamma(t))=\int_{0}^{\tau}|\dot{\sigma}_{t}(s)|_{\tilde{g}}ds\,=\,\int_{0}^{\tau}\frac{1}{f(\sigma_{t}(s))}|\dot{\sigma}_{t}(s)|_{g}ds\,=\,\int_{0}^{\tau}\frac{1}{f(\sigma_{t}(s))}ds\,>\,\int_{T_{t}}^{\tau}\frac{r(\sigma_{t}(s))^{-k}}{\sqrt{C}}ds\,.

Since r(σt(s))<dg(σt(s),γ(0))+δ<s+δr(\sigma_{t}(s))<{\rm d}_{g}(\sigma_{t}(s),\gamma(0))+\delta<s+\delta, we then have

ρ(γ(t))\displaystyle\rho(\gamma(t)) >(τ+δ)1k(Tt+δ)1kC(1k)\displaystyle>\frac{(\tau+\delta)^{1-k}-(T_{t}+\delta)^{1-k}}{\sqrt{C}(1-k)}\, (3.3)
>(Tt+δ+dg(γ(t),K))1k(Tt+δ)1kC(1k)\displaystyle>\,\frac{(T_{t}+\delta+{\rm d}_{g}(\gamma(t),K))^{1-k}-(T_{t}+\delta)^{1-k}}{\sqrt{C}(1-k)} (3.4)
>(dg(γ(t),K))1k(diamg~(K)maxKf)1kC(1k).\displaystyle>\,\frac{\left({\rm d}_{g}(\gamma(t),K)\right)^{1-k}-\left({\rm diam}_{\tilde{g}}(K)\max_{K}f\right)^{1-k}}{\sqrt{C}(1-k)}\,. (3.5)

Since KK is fixed, the distance dg(γ(t),K){\rm d}_{g}(\gamma(t),K) is going to ++\infty, hence ρ(γ(t))+\rho(\gamma(t))\to+\infty as t+t\to+\infty as wished. ∎

Another well studied family of ends is the following:

Definition 3.3.

We say that an end of a substatic triple (M,g,f)(M,g,f) is conformally compact if a neighborhood EE of it is the interior of a compact manifold E¯\overline{E} with boundary E¯\partial\overline{E} and the metric g~=g/f2\tilde{g}=g/f^{2} extends to the boundary of E¯\overline{E} in 𝒞3\mathscr{C}^{3}-fashion. We denote by E=E¯M\partial E_{\infty}=\partial\overline{E}\setminus M the conformal boundary of the end. Finally, we require f1f^{-1} to extend to a 𝒞3\mathscr{C}^{3}-function on E¯\overline{E} in such a way that f1=0f^{-1}=0 and d(f1)0d(f^{-1})\neq 0 on E\partial E_{\infty}.

It is clear that on conformally compact ends the g~\tilde{g}-distance function ρ\rho does not grow to infinity along the end. On the other hand, it is easily seen that η\eta goes to infinity along any ray going into a conformally compact end. An example of this behaviour is given by the Schwarzschild-Anti de Sitter solution.

The two families of ff-complete ends and conformally compact ends enclose the model solutions we are interested in. Furthermore, following [Wyl17], we can prove a splitting theorem for both these types of ends. The proof, given below, makes substantial use of the conformal metric g~\tilde{g}. It is then convenient to remark that our manifold remains complete with respect to this conformal metric.

Lemma 3.4.

Let (M,g,f)(M,g,f) be a substatic triple with ends that are either ff-complete or conformally compact and let M\partial M_{\infty} be the (possibly empty) conformal infinity (namely, the union of the conformal infinities of the conformally compact ends). Then the manifold (MM)M(M\cup\partial M_{\infty})\setminus\partial M is complete with respect to the metric g~=g/f2\tilde{g}=g/f^{2}.

Remark 3.5.

We specify that we are referring here to completeness as a metric space, not to geodesic completeness. Clearly geodesic completeness fails in presence of a conformal boundary, since g~\tilde{g}-geodesics may end at M\partial M_{\infty}.

Proof.

The fact that the manifold is complete near the conformal boundary is immediate from the definition of conformally compact ends. The g~\tilde{g}-completeness of ff-complete ends has already been discussed after Definition 3.1. It remains to prove g~\tilde{g}-completeness near M\partial M. To do this, we show that the boundary components become ends with respect to the conformal metric g~\tilde{g}. Let γ:[0,]\gamma:[0,\ell]\to\mathbb{R} be a gg-unit speed curve with γ(0)M\gamma(0)\in\partial M and γ(t)\gamma(t) in the interior of MM for any t>0t>0. It is enough to show that its g~\tilde{g}-length is infinite. From the mean value theorem we know that for any t(0,]t\in(0,\ell] there exists ξ(0,t)\xi\in(0,t) such that

f(γ(t))=f(γ(t))f(γ(0))=f(γ(ξ))|γ˙(t)t|f|(γ(ξ))t.f(\gamma(t))\,=\,f(\gamma(t))-f(\gamma(0))\,=\,\langle\nabla f(\gamma(\xi))\,|\,\dot{\gamma}(t)\rangle\,t\,\leq\,|\nabla f|(\gamma(\xi))\,t\,.

If KK is a compact collar neighborhood of M\partial M containing γ\gamma, we then compute

0|γ˙(t)|g~𝑑t=0dtf(γ(t))01(maxK|f|)t𝑑t=+,\int_{0}^{\ell}|\dot{\gamma}(t)|_{\tilde{g}}dt\,=\,\int_{0}^{\ell}\frac{dt}{f(\gamma(t))}\,\geq\,\int_{0}^{\ell}\frac{1}{\left(\max_{K}|\nabla f|\right)t}dt=+\infty\,,

as wished. ∎

3.2. Splitting Theorem for ff-complete ends

In [Wyl17], the author proves a Splitting Theorem in the CD(0,1){\rm CD}(0,1) framework. Here we translate this result in the substatic setting, obtaining Theorem C. The proof is essentially the one in [Wyl17], but we prefer to show it for completeness. The strategy is that of exploiting the Laplacian Comparison given by Theorem 2.5 together with standard techniques for the Busemann function, to be coupled with a refined splitting argument.

Proof of Theorem C.

The first part of the proof follows quite closely the usual proof of the classical Splitting Theorem for manifolds with nonpositive Ricci curvature, so we avoid to give all the technical details, that can be found in any standard Riemannian geometry book (see for instance [Pet16, Theorem 7.3.5]).

We are assuming that there are at least two ff-complete ends and we know from Lemma 3.4 that MMM\setminus\partial M is complete with respect to g~\tilde{g}, hence we can take points arbitrarily far away in the two ends and connect them via a g~\tilde{g}-minimizing geodesic. At the limit, we thus produce a globally minimizing geodesic σ:[,+]\sigma:[-\infty,+\infty] going from one end to the other. For a given tt\in\mathbb{R} we then consider the functions

βt(x)=dg~(x,σ(t))t\beta_{t}(x)={\rm d}_{\tilde{g}}(x,\sigma(t))-t

and the Busemann functions

β+(x)=limt+βt(x),β(x)=limtβt(x).\beta_{+}(x)=\lim_{t\to+\infty}\beta_{t}(x)\,,\qquad\beta_{-}(x)=\lim_{t\to-\infty}\beta_{t}(x)\,.

Under the assumption of ff-completeness, we know that the g~\tilde{g}-distance goes to infinity as we approach the end, hence the limits are well defined. Theorem 2.5 tells us that Δβt+(1/f)f|βt1/η\Delta\beta_{t}+(1/f)\langle\nabla f\,|\,\nabla\beta_{t}\rangle\leq 1/\eta for every tt. Again using the fact that we are assuming ff-completeness of the ends, we have η+\eta\to+\infty at infinity. Standard arguments then tell us that β±\beta_{\pm} satisfy the inequality

Δβ±+1ff|β±0\Delta\beta_{\pm}+\frac{1}{f}\langle\nabla f\,|\,\nabla\beta_{\pm}\rangle\leq 0

in the barrier sense. In particular, β+β+\beta_{-}+\beta_{+} satisfies the same elliptic inequality. Furthermore β+β+=0\beta_{-}+\beta_{+}=0 on σ\sigma by construction and a simple application of the triangle inequality shows that β+β+0\beta_{-}+\beta_{+}\geq 0 on the whole MM. In follows then by maximum principle that β++β=0\beta_{+}+\beta_{-}=0. We then conclude that β=β+\beta=\beta_{+} solves

Δβ+1ff|β= 0\Delta\beta+\frac{1}{f}\langle\nabla f\,|\,\nabla\beta\rangle\,=\,0 (3.6)

in the barrier sense. Standard regularity theory tells us that β\beta is in fact smooth, hence solves (3.6) in the classical sense as well.

We denote by ~\widetilde{\nabla} the Levi-Civita connection with respect to the metric g~\tilde{g}. Our next step is to prove that ~β\widetilde{\nabla}\beta is in fact the splitting direction. We first exploit (3.6) to compute

Δg~β=(n1)1f~f|~βg~\Delta_{\tilde{g}}\beta\,=\,-(n-1)\frac{1}{f}\big{\langle}\widetilde{\nabla}f\,\big{|}\,\widetilde{\nabla}\beta\big{\rangle}_{\tilde{g}}

Furthermore, rewriting the substatic condition (1.1) in terms of the new metric, we find

Ricg~(n1)1f~2f2(n1)1f2dfdf{\rm Ric}_{\tilde{g}}\,\geq\,(n-1)\frac{1}{f}\widetilde{\nabla}^{2}f-2(n-1)\frac{1}{f^{2}}df\otimes df

The above formulas can be applied in combination with the Bochner formula to obtain

Δg~|~β|g~2\displaystyle\Delta_{\tilde{g}}\big{|}\widetilde{\nabla}\beta\big{|}_{\tilde{g}}^{2}\, =|~2β|g~2+2Ricg~(~β,~β)+2~Δg~β|~βg~\displaystyle=\,\big{|}\widetilde{\nabla}^{2}\beta\big{|}_{\tilde{g}}^{2}+2{\rm Ric}_{\tilde{g}}\big{(}\widetilde{\nabla}\beta,\widetilde{\nabla}\beta\big{)}+2\big{\langle}\widetilde{\nabla}\Delta_{\tilde{g}}\beta\,\big{|}\,\widetilde{\nabla}\beta\big{\rangle}_{\tilde{g}} (3.7)
 2[|~2β|g~2(Δg~β)2n1](n1)1f~|~β|g~2|~fg~\displaystyle\geq\,2\,\left[\big{|}\widetilde{\nabla}^{2}\beta\big{|}_{\tilde{g}}^{2}-\frac{(\Delta_{\tilde{g}}\beta)^{2}}{n-1}\right]-(n-1)\frac{1}{f}\big{\langle}\widetilde{\nabla}\big{|}\widetilde{\nabla}\beta\big{|}_{\tilde{g}}^{2}\,\big{|}\,\widetilde{\nabla}f\big{\rangle}_{\tilde{g}} (3.8)

A standard estimate for the Hessian tells us that

|~2β|g~2(Δg~β)2n11n1Δg~β|~β|g~2~|~β|g~2|~βg~.\big{|}\widetilde{\nabla}^{2}\beta\big{|}_{\tilde{g}}^{2}-\frac{(\Delta_{\tilde{g}}\beta)^{2}}{n-1}\,\geq\,-\frac{1}{n-1}\frac{\Delta_{\tilde{g}}\beta}{|\widetilde{\nabla}\beta|_{\tilde{g}}^{2}}\big{\langle}\widetilde{\nabla}\big{|}\widetilde{\nabla}\beta\big{|}_{\tilde{g}}^{2}\,\big{|}\,\widetilde{\nabla}\beta\big{\rangle}_{\tilde{g}}\,.

Substituting this in the previous inequality, we get

Δg~|~β|g~2~|~β|g~2|2f~f|~βg~|~β|g~2~β+(n1)1f~fg~\Delta_{\tilde{g}}\big{|}\widetilde{\nabla}\beta\big{|}_{\tilde{g}}^{2}\,\geq\,\bigg{\langle}\widetilde{\nabla}\big{|}\widetilde{\nabla}\beta\big{|}_{\tilde{g}}^{2}\,\bigg{|}\,\frac{2}{f}\frac{\langle\widetilde{\nabla}f\,|\,\widetilde{\nabla}\beta\rangle_{\tilde{g}}}{|\widetilde{\nabla}\beta|_{\tilde{g}}^{2}}\widetilde{\nabla}\beta+(n-1)\frac{1}{f}\widetilde{\nabla}f\bigg{\rangle}_{\tilde{g}}

By Cauchy–Schwarz the vector on the right-hand side is bounded on any compact set. In particular, |~β|g~2|\widetilde{\nabla}\beta|_{\tilde{g}}^{2} satisfies the maximum principle. To conclude from this that |~β|g~|\widetilde{\nabla}\beta|_{\tilde{g}} is constant, we first observe via triangle inequality that β(x)β(y)dg~(x,y)\beta(x)-\beta(y)\leq{\rm d}_{\tilde{g}}(x,y) for any two points xx, yy. This immediately implies that |~β|g~1|\widetilde{\nabla}\beta|_{\tilde{g}}\leq 1 on MM. On the other hand, since β(σ(τ))=τ\beta(\sigma(\tau))=\tau, we conclude that |~β|g~=1|\widetilde{\nabla}\beta|_{\tilde{g}}=1 on σ\sigma. The strong maximum principle then implies that |~β|g~=1|\widetilde{\nabla}\beta|_{\tilde{g}}=1 on the whole manifold. In particular, the previous inequalities must be equalities, namely

Ricg~(~β,~β)\displaystyle{\rm Ric}_{\tilde{g}}\big{(}\widetilde{\nabla}\beta,\widetilde{\nabla}\beta\big{)}\, =(n1)1f~2f(~β,~β)2(n1)1f2~f|~βg~,\displaystyle=\,(n-1)\frac{1}{f}\widetilde{\nabla}^{2}f\big{(}\widetilde{\nabla}\beta,\widetilde{\nabla}\beta\big{)}-2(n-1)\frac{1}{f^{2}}\big{\langle}\widetilde{\nabla}f\,\big{|}\,\widetilde{\nabla}\beta\big{\rangle}_{\tilde{g}}\,, (3.9)
|~2β|g~2\displaystyle\big{|}\widetilde{\nabla}^{2}\beta\big{|}_{\tilde{g}}^{2}\, =(Δg~β)2n1.\displaystyle=\,\frac{(\Delta_{\tilde{g}}\beta)^{2}}{n-1}\,. (3.10)

Furthermore, the fact that |~β|g~=1|\widetilde{\nabla}\beta|_{\tilde{g}}=1 on the whole manifold grants us that we can use β\beta as a coordinate and that the manifold MM is diffeomorphic to ×Σ\mathbb{R}\times\Sigma, for some (n1)(n-1)-dimensional manifold Σ\Sigma. With respect to coordinates {β,θ1,,θn1}\{\beta,\theta^{1},\dots,\theta^{n-1}\}, the conformal metric writes as

g~=dβdβ+g~ijdθidθj.\tilde{g}\,=\,d\beta\otimes d\beta+\tilde{g}_{ij}d\theta^{i}\otimes d\theta^{j}\,.

Again, since |~β|g~=1|\widetilde{\nabla}\beta|_{\tilde{g}}=1, for any vector XX it holds

~2β(~β,X)=~|~β|g~2|Xg~= 0.\widetilde{\nabla}^{2}\beta\big{(}\widetilde{\nabla}\beta,X\big{)}\,=\,\big{\langle}\widetilde{\nabla}\big{|}\widetilde{\nabla}\beta\big{|}_{\tilde{g}}^{2}\,\big{|}\,X\big{\rangle}_{\tilde{g}}\,=\,0\,.

It follows immediately from this and the identity |~2β|g~2=(Δg~β)2/(n1)|\widetilde{\nabla}^{2}\beta|_{\tilde{g}}^{2}=(\Delta_{\tilde{g}}\beta)^{2}/(n-1) that, in the coordinates in which g~\tilde{g} has the form (2.1), for any i,j=1,,n1i,j=1,\dots,n-1 it holds

~ij2β=Δg~βn1g~ij=1ffβg~ij,\widetilde{\nabla}^{2}_{ij}\beta\,=\,\frac{\Delta_{\tilde{g}}\beta}{n-1}\tilde{g}_{ij}\,=\,-\frac{1}{f}\frac{\partial f}{\partial\beta}\tilde{g}_{ij}\,,

where the latter identity makes use of (2.3). On the other hand, from the definition of Hessian we have ~ij2β=Γijβ=βg~ij/2\widetilde{\nabla}^{2}_{ij}\beta=-\Gamma_{ij}^{\beta}=\partial_{\beta}\tilde{g}_{ij}/2, hence

g~ijβ=2ffβg~ij.\frac{\partial\tilde{g}_{ij}}{\partial\beta}\,=\,-\frac{2}{f}\frac{\partial f}{\partial\beta}\tilde{g}_{ij}\,.

This identity can be solved explicitly, yielding

g~ij=1f2(g0)ij,\tilde{g}_{ij}\,=\,\frac{1}{f^{2}}(g_{0})_{ij}\,,

where (g0)ij(g_{0})_{ij} does not depend on ρ\rho. Comparing with (2.1) and recalling g=f2g~g=f^{2}\tilde{g}, we have obtained

g=f2dβdβ+g0.g\,=\,f^{2}d\beta\otimes d\beta+g_{0}\,.

Finally, we remark again that in this proof we did not have to ask for the boundary of MM to be empty since, as observed in Lemma 3.4, with respect to the conformal metric g~\tilde{g} the boundary components become ends, hence they cannot obstruct minimizing geodesics. Therefore, the argument to produce a line between two ff-complete ends goes through and the manifold splits. But then this would imply M=(,+)×Σ\partial M=(-\infty,+\infty)\times\partial\Sigma, which contradicts our initial assumption that the boundary is compact. It follows that the boundary must be empty if there is more than one ff-complete end. ∎

Remark 3.6.

We point out that it is actually possible to obtain a stronger thesis in Theorem C above. In fact, proceeding as in the proof of Theorem 2.9, one can also show that identity (2.46) is in force, and from there deduce that f=f1f2f=f_{1}f_{2}, where f1f_{1} is a function of ss whereas f2f_{2} does not depend on ss. We do not give the details on this computation, which has already been performed in the conformal CD(0,1){\rm CD}(0,1) framework [Wyl17, Proposition 2.2]. Recalling the relation between CD(0,1){\rm CD}(0,1) and substatic discussed in Section A.2, one can easily translate this result in our setting. One may also write down explicitly the substatic condition in the directions tangential to the cross section, to obtain some information on the triple (Σ,g0,f2)(\Sigma,g_{0},f_{2}). Again, in the CD(0,1){\rm CD}(0,1) setting, this has been done in [Wyl17, Proposition 2.3], where it is shown that the triple (Σ,g0,(n1)logf2)(\Sigma,g_{0},-(n-1)\log f_{2}) satisfies the CD(0,1){\rm CD(0,1)} condition (in fact, it is even CD(0,0){\rm CD}(0,0)). It is not immediately clear whether this fact translates nicely in our setting. These refinements of the thesis of Theorem C will not be needed in the rest of the paper.

3.3. Splitting theorem for conformally compact ends

We now discuss conformally compact ends. For such ends, by definition the metric extends to the conformal infinity sufficiently smoothly so that the mean curvature Hg~{\rm H}_{\tilde{g}} of the conformal infinity E\partial E_{\infty} is well defined. On the other hand, the mean curvatures H{\rm H} and Hg~{\rm H}_{\tilde{g}} of a hypersurface with respect to the two different metrics can be seen to be related by

Hg~=fH(n1)f|ν.{\rm H}_{\tilde{g}}\,=\,f{\rm H}-(n-1)\langle\nabla f\,|\,\nu\rangle\,.

Alternatively, setting φ=1/f\varphi=1/f we can write

H=φHg~(n1)~φ|νg~g~.{\rm H}\,=\,\varphi{\rm H}_{\tilde{g}}-(n-1)\big{\langle}\widetilde{\nabla}\varphi\,\big{|}\,\nu_{\tilde{g}}\big{\rangle}_{\tilde{g}}\,.

By definition of conformal compactness we know that φ\varphi extend in a 𝒞3\mathscr{C}^{3} fashion to the conformal boundary by setting φ=0\varphi=0 on E\partial E_{\infty}. In particular |~φ|g~|\widetilde{\nabla}\varphi|_{\tilde{g}} is bounded, which implies that the quantity H/f{\rm H}/f can be extended to zero in a continuous fashion on E\partial E_{\infty}. Taking now as ρ\rho the g~\tilde{g}-distance from E\partial E_{\infty}, recalling the Riccati equation (2.8) we have that

ρ(Hf)1n1H2,Hf|ρ=0=0.\frac{\partial}{\partial\rho}\left(\frac{{\rm H}}{f}\right)\,\leq\,-\frac{1}{n-1}\,{\rm H}^{2}\,,\qquad{\frac{{\rm H}}{f}}_{|_{\rho=0}}=0\,.

The assumption of 𝒞3\mathscr{C}^{3}-regularity of the conformal boundary made in Definition 3.3 was needed precisely to make sense of the ρ\rho-derivative of H{\rm H}. Proceeding exactly as in Subsection 2.3, from this formula on the evolution of the mean curvature we obtain the Laplacian comparison

Hf=Δρ+1ff|ρ 0.\frac{{\rm H}}{f}\,=\,\Delta\rho+\frac{1}{f}\langle\nabla f\,|\,\nabla\rho\rangle\,\leq\,0\,. (3.11)

This is the main ingredient to prove the Splitting Theorem in the conformally compact setting:

Theorem 3.7.

Let (M,g,f)(M,g,f) be a substatic triple with conformally compact ends. Then there is at most one end.

Proof.

Again, a CD(0,1){\rm CD}(0,1)-version of this argument can be found in [Wyl17, Theorem 5.1]. The proof follows closely the one in [Kas83, Theorem B-(1)], where a splitting theorem for compact manifolds is discussed. Suppose by contradiction that the conformal infinity has at least two connected components. Let S,S+S_{-},S_{+} be the two components with least distance. Then there exists a g~\tilde{g}-geodesic σ\sigma minimizing the distance between them.

Let β\beta_{-} (resp. β+\beta_{+}) be the distance from SS_{-} (resp. S+S_{+}) with respect to g~\tilde{g}. The discussion above grants us that both β\beta_{-} and β+\beta_{+} satisfy the Laplacian comparison (3.11). In particular, so does β+β+\beta_{-}+\beta_{+}. Since by construction β+β+\beta_{-}+\beta_{+} reaches its minimum value dist(S,S+){\rm dist}(S_{-},S_{+}) on the geodesic σ\sigma, we then conclude by the strong maximum principle that β+β+\beta_{-}+\beta_{+} is constant and equal to dist(S,S+){\rm dist}(S_{-},S_{+}) on the whole manifold. It follows immediately that β=β+\beta=\beta_{+} satisfies

Δβ+1ff|β= 0\Delta\beta+\frac{1}{f}\langle\nabla f\,|\,\nabla\beta\rangle\,=\,0

in the barrier sense. We are now exactly in the same situation reached in the proof of the Splitting Theorem for ff-complete ends. We can then proceed exactly as after formula (3.6) to conclude that (M,g)(M,g) is isometric to a twisted product

((a,b)×Σ,f2dsds+gΣ).\left((a,b)\times\Sigma,\,f^{2}\,ds\otimes ds+g_{\Sigma}\right)\,.

On the other hand, such a manifold is not conformally compact, as the metric g~=g/f2\tilde{g}=g/f^{2} is degenerate as ss approaches aa or bb. We have thus reached a contradiction, implying that there were not multiple conformally compact ends in the first place. ∎

This result generalizes [CS01, Theorem I.1], where the same thesis is obtained for conformally compact vacuum static solutions with negative cosmological constant. It is interesting to notice that the proof proposed in [CS01] also makes use of the conformal metric g~=g/f2\tilde{g}=g/f^{2}, which is exploited to invoke a spacetime censorship result from [Gal+99, Theorem 2.1].

3.4. Splitting Theorem for mixed ends

For completeness, we include here the case where there are ends with different behaviours. This case was not considered in [Wyl17] but the proof is similar.

Theorem 3.8.

Let (M,g,f)(M,g,f) be a substatic triple with ends that are either conformally compact or ff-complete. If there is at least one ff-complete end, then there cannot be any conformally compact end.

Proof.

This time the proof follows [Kas83, Theorem C-(2)]. Suppose that there is a conformally compact end and an ff-complete end. Then, one constructs a globally minimizing g~\tilde{g}-geodesic σ\sigma starting at a connected component SS of the conformal boundary and reaching infinity. Let β\beta_{-} be the distance from SS and β+\beta_{+} be the Busemann function relative to σ\sigma. As in the previous cases, from the Laplacian comparisons for both β\beta_{-} and β+\beta_{+} and the fact that β+β+\beta_{-}+\beta_{+} achieves its minimum value 0 on σ\sigma, we deduce that β=β+\beta=\beta_{+} satisfies

Δβ+1ff|β= 0\Delta\beta+\frac{1}{f}\langle\nabla f\,|\,\nabla\beta\rangle\,=\,0

in the barrier sense. We now proceed as in the other cases to show that the manifold must be a twisted product

((0,+)×Σ,f2dsds+gΣ).\left((0,+\infty)\times\Sigma,\,f^{2}\,ds\otimes ds+g_{\Sigma}\right)\,.

Again as in the proof of Theorem 3.7, we observe that the end corresponding to s=0s=0 cannot be conformally compact as the metric g~=g/f2\tilde{g}=g/f^{2} becomes degenerate as s0s\to 0. We have thus reached a contradiction, meaning that it is impossible to have an ff-complete end and a conformally compact end at the same time. ∎

This theorem, together with the other results in this Section (Theorem C and Theorem 3.7) strongly narrows the acceptable configurations of ends for a substatic triple. We sum up the topological information we have collected in the following statement.

Corollary 3.9.

Let (M,g,f)(M,g,f) be a substatic triple with ends that are either conformally compact or ff-complete. If there is more than one end, then there are exactly two ends, both ff-complete, and M=Ø\partial M=\mathchar 31\relax.

4. Asymptotic Volume Ratio and Willmore-type inequality

In this section we focus on ff-complete ends and we introduce the notion of asymptotic volume ratio (AVR), in analogy with the classical case of nonpositive Ricci curvature, as the limit of the Bishop–Gromov monotonic quantity. In order to have a well defined AVR, we will need to focus our attention to the special case of uniform ends. Building on the notion of AVR we will finally prove the Willmore-type inequality mentioned in the introduction.

4.1. Uniform ff-complete ends

Here we introduce and comment the notion of uniformity of an ff-complete end. For convenience, instead of working on the whole MM, we focus our attention on the end only. In other words, starting from the next definition and for most of this subsection, instead of working on the whole substatic triple (M,g,f)(M,g,f), we just consider a neighborhood EE of our end and we focus our attention on the restriction (E,g,f)(E,g,f), which we refer to as a substatic ff-complete end. It is easy to show that the definitions and statements below do not depend on the choice of the neighborhood EE of our end.

Definition 4.1.

Let (E,g,f)(E,g,f) be a substatic ff-complete end. We say that (E,g,f)(E,g,f) is uniform, if, for any two compact hypersurfaces Σ1\Sigma_{1}, Σ2\Sigma_{2} contained in the interior of EE and every δ>0\delta>0, there exists a compact set KEK\supset\partial E such that for any two unit speed g~\tilde{g}-geodesics σ1\sigma_{1}, σ2\sigma_{2} minimizing the distance between Σ1\Sigma_{1}, Σ2\Sigma_{2} and a point p=σ1(t1)=σ2(t2)p=\sigma_{1}(t_{1})=\sigma_{2}(t_{2}) outside KK, it holds

|0t1f2(σ1(t))𝑑t0t2f2(σ2(t))𝑑t 1|δ.\left|\frac{\int_{0}^{t_{1}}f^{2}(\sigma_{1}(t))dt}{\int_{0}^{t_{2}}f^{2}(\sigma_{2}(t))dt}\,-\,1\right|\,\leq\,\delta\,. (4.1)

While the definition above is slightly technical, we point out that there are natural cases in which uniformity is guaranteed. We give here a couple of easily described families of uniform ff-complete ends. The following result for instance guarantees us that an end is uniform as long as ff goes to one at infinity.

Proposition 4.2.

Let (E,g,f)(E,g,f) be a substatic end. If f1f\to 1 at infinity, then (E,g,f)(E,g,f) is ff-complete and uniform.

Proof.

The fact that the ends are ff-complete has already been shown in far greater generality in Proposition 3.2. Let now Σ1\Sigma_{1}, Σ2\Sigma_{2} be two hypersurfaces. Since f1f\to 1 at infinity, for every ε\varepsilon the set Kε={|f1|>ε}K_{\varepsilon}=\{|f-1|>\varepsilon\} is compact. In particular 1ε<f<1+ε1-\varepsilon<f<1+\varepsilon outside KεK_{\varepsilon}. We consider now a g~\tilde{g}-geodesically convex compact set KK containing KεK_{\varepsilon} and the two hypersurfaces Σ1\Sigma_{1}, Σ2\Sigma_{2}.

Let σ1\sigma_{1}, σ2\sigma_{2} be two unit speed g~\tilde{g}-geodesics minimizing the distance between Σ1\Sigma_{1}, Σ2\Sigma_{2} and a point p=σ1(t1)=σ2(t2)p=\sigma_{1}(t_{1})=\sigma_{2}(t_{2}) outside KεK_{\varepsilon}. For j=1,2j=1,2, let 0<Tj<tj0<T_{j}<t_{j} be the largest number such that σj(Tj)Kε\sigma_{j}(T_{j})\in K_{\varepsilon}. We then have, for j=1,2j=1,2,

0<0Tjf2(σj(t))𝑑t<TjmaxKf2,(1ε)2(tjTj)<Tjtjf2(σj(t))𝑑t<(1+ε)2(tjTj).0\,<\,\int_{0}^{T_{j}}f^{2}(\sigma_{j}(t))dt\,<\,T_{j}\max_{K}f^{2}\,,\qquad(1-\varepsilon)^{2}(t_{j}-T_{j})\,<\,\int_{T_{j}}^{t_{j}}f^{2}(\sigma_{j}(t))dt\,<\,(1+\varepsilon)^{2}(t_{j}-T_{j})\,.

As a consequence, we estimate:

(1ε)2(t1T1)T2maxKf2+(1ε)2(t2T2)<0t1f2(σ1(t))𝑑t0t2f2(σ2(t))𝑑t<T1maxKf2+(1+ε)2(t1T1)(1ε)2(t2T2).\frac{(1-\varepsilon)^{2}(t_{1}-T_{1})}{T_{2}\max_{K}f^{2}+(1-\varepsilon)^{2}(t_{2}-T_{2})}\,<\,\frac{\int_{0}^{t_{1}}f^{2}(\sigma_{1}(t))dt}{\int_{0}^{t_{2}}f^{2}(\sigma_{2}(t))dt}\,<\,\frac{T_{1}\max_{K}f^{2}+(1+\varepsilon)^{2}(t_{1}-T_{1})}{(1-\varepsilon)^{2}(t_{2}-T_{2})}\,.

Since ff is bounded at infinity, the quantity maxKf2\max_{K}f^{2} is bounded by a constant independent of KK. Furthermore, we have Tj<diamg~KεT_{j}<{\rm diam}_{\tilde{g}}K_{\varepsilon}, for j=1,2j=1,2, by construction. On the other hand, if we take the compact set KK to be much larger than KεK_{\varepsilon}, we can make t1t_{1} and t2t_{2} arbitrarily large. The uniformity estimate (4.1) follows then easily. ∎

Another case in which uniformity is guaranteed is under the assumption that the norm of the gradient of ff decays sufficiently fast.

Proposition 4.3.

Let (E,g,f)(E,g,f) be a substatic ff-complete end and let ρ\rho be the g~\tilde{g}-distance from a point, where g~=g/f2\tilde{g}=g/f^{2}. If for some ε>0\varepsilon>0 there exist a compact set KEK\supset\partial E and a constant C>0C>0 such that

|f|<Cρ1ε|\nabla f|<C\rho^{-1-\varepsilon}

outside KK, then (E,g,f)(E,g,f) is uniform.

Proof.

Fix the compact hypersurfaces Σ1\Sigma_{1}, Σ2\Sigma_{2}, the point xx and the constant ε>0\varepsilon>0. Let KEK\supset\partial E be the compact set such that |f|<Cρ1ε|\nabla f|<C\rho^{-1-\varepsilon} outside KK, where ρ\rho is the g~\tilde{g}-distance from xx. Up to enlarging KK, we can suppose that xx, Σ1\Sigma_{1} and Σ2\Sigma_{2} are inside KK.

Let pp be a point outside KK. For i=1,2i=1,2, consider the unit speed g~\tilde{g}-geodesic σi:[0,ti]M\sigma_{i}:[0,t_{i}]\to M minimizing the distance between Σi\Sigma_{i} and pp and such that σi(0)Σi\sigma_{i}(0)\in\Sigma_{i}, σi(ti)=p\sigma_{i}(t_{i})=p. We compare the value of ff at a point σi(τ)\sigma_{i}(\tau) and at the point pp. Integrating along the geodesic, we find

logf(p)=logf(σi(τ))+τti~logf|σ˙ig~(σi(t))𝑑t,\log f(p)\,=\,\log f(\sigma_{i}(\tau))+\int_{\tau}^{t_{i}}\big{\langle}\widetilde{\nabla}\log f\,\big{|}\,\dot{\sigma}_{i}\big{\rangle}_{\tilde{g}}(\sigma_{i}(t))\,dt\,,

hence

|log(f(σi(τ))f(p))|τti1f|~f|g~(σi(t))𝑑t=τti|f|(σi(t))𝑑t.\left|\log\left(\frac{f(\sigma_{i}(\tau))}{f(p)}\right)\right|\,\leq\,\int_{\tau}^{t_{i}}\frac{1}{f}\big{|}\widetilde{\nabla}f\big{|}_{\tilde{g}}(\sigma_{i}(t))\,dt=\,\int_{\tau}^{t_{i}}|\nabla f|(\sigma_{i}(t))\,dt\,. (4.2)

We now exploit our hypothesis. Assume that the segment σi|[τ,ti]{\sigma_{i}}_{|_{[\tau,t_{i}]}} is outside KK. Then |f|Cρ1ε|\nabla f|\leq C\rho^{-1-\varepsilon} at the points of σi|[τ,ti]{\sigma_{i}}_{|_{[\tau,t_{i}]}}, where ρ\rho is the distance from xx. Notice that for all τtti\tau\leq t\leq t_{i}, σi(t)\sigma_{i}(t) is at distance tt from Σi\Sigma_{i}, so that by triangle inequality, for any yKy\in K it holds

tmaxyΣidg~(x,y)ρ(σi(t))t+maxyΣidg~(x,y).t-\max_{y\in\Sigma_{i}}{\rm d}_{\tilde{g}}(x,y)\leq\rho(\sigma_{i}(t))\leq t+\max_{y\in\Sigma_{i}}{\rm d}_{\tilde{g}}(x,y)\,.

If we then take

t2maxyΣidg~(x,y)t\geq 2\max_{y\in\Sigma_{i}}{\rm d}_{\tilde{g}}(x,y) (4.3)

we get

|f|(σi(t))Cρ1ε(σi(t))C(tmaxyΣidg~(x,y))1εC(t2)1ε=21+εCt1ε.|\nabla f|(\sigma_{i}(t))\,\leq\,C\rho^{-1-\varepsilon}(\sigma_{i}(t))\,\leq\,C(t-\max_{y\in\Sigma_{i}}{\rm d}_{\tilde{g}}(x,y))^{-1-\varepsilon}\,\leq\,C\left(\frac{t}{2}\right)^{-1-\varepsilon}=2^{1+\varepsilon}Ct^{-1-\varepsilon}\,.

If we then suppose that τ2maxyΣidg~(x,y)\tau\geq 2\max_{y\in\Sigma_{i}}{\rm d}_{\tilde{g}}(x,y), we deduce from (4.2) that

|log(f(σi(τ))f(p))| 21+εCτεtiεε\left|\log\left(\frac{f(\sigma_{i}(\tau))}{f(p)}\right)\right|\,\leq\,2^{1+\varepsilon}C\frac{\tau^{-\varepsilon}-t_{i}^{-\varepsilon}}{\varepsilon} (4.4)

that is,

e21+εCετεexp[21+εCtiετεε]f(σi(τ))f(p)exp[21+εCτεtiεε]e21+εCετε.e^{-\frac{2^{1+\varepsilon}C}{\varepsilon}\tau^{-\varepsilon}}\,\leq\,{\rm exp}\left[2^{1+\varepsilon}C\frac{t_{i}^{-\varepsilon}-\tau^{-\varepsilon}}{\varepsilon}\right]\,\leq\,\frac{f(\sigma_{i}(\tau))}{f(p)}\,\leq\,{\rm exp}\left[2^{1+\varepsilon}C\frac{\tau^{-\varepsilon}-t_{i}^{-\varepsilon}}{\varepsilon}\right]\,\leq\,e^{\frac{2^{1+\varepsilon}C}{\varepsilon}\tau^{-\varepsilon}}\,. (4.5)

Let now κ>0\kappa>0 be a large number and consider the compact set

Kκ=Bκg~(K):={yM:dg~(K,y)κ}.K_{\kappa}\,=\,B^{\tilde{g}}_{\kappa}(K)\,:=\,\{y\in M\,:\,{\rm d}_{\tilde{g}}(K,y)\leq\kappa\}\,.

Since ΣiK\Sigma_{i}\subset K, notice that if σi(t)Kκ\sigma_{i}(t)\not\in K_{\kappa} then t>κt>\kappa. It follows that, up to taking κ\kappa large enough, inequality (4.3) holds and in particular |f|(σi(t))21+εCt1ε|\nabla f|(\sigma_{i}(t))\leq 2^{1+\varepsilon}Ct^{-1-\varepsilon} for any tt such that σi(t)Kκ\sigma_{i}(t)\not\in K_{\kappa}.

Conversely, also notice by triangle inequality that if

t>κ+maxyΣi,zKEdg~(y,z),t>\kappa+\max_{y\in\Sigma_{i},z\in\partial K\setminus\partial E}{\rm d}_{\tilde{g}}(y,z)\,,

then σi(t)Kκ\sigma_{i}(t)\not\in K_{\kappa}. Up to taking κ\kappa large enough, we can then also suppose that σi(t)Kκ\sigma_{i}(t)\not\in K_{\kappa} for every t>2κt>2\kappa. In particular, for any τ>2κ\tau>2\kappa we can apply estimate (4.5), obtaining

e2Cεκεf(σi(τ))f(p)e2Cεκεe^{-\frac{2C}{\varepsilon}\kappa^{-\varepsilon}}\,\leq\,\frac{f(\sigma_{i}(\tau))}{f(p)}\,\leq\,e^{\frac{2C}{\varepsilon}\kappa^{-\varepsilon}} (4.6)

We are finally ready to prove uniformity at infinity with respect to the compact set KκK_{\kappa}, for κ\kappa sufficiently large. As already noticed, σi(t)K2κ\sigma_{i}(t)\in K_{2\kappa} for t<2κt<2\kappa, hence

0t1f2(σ1(t))𝑑t0t2f2(σ2(t))𝑑t2κmaxK2κf2+2κt1f2(σ1(t))𝑑t2κminK2κf2+2κt2f2(σ2(t))𝑑t2κmaxK2κf2+(t12κ)f(p)2e2Cεκε2κminK2κf2+(t22κ)f(p)2e2Cεκε.\frac{\int_{0}^{t_{1}}f^{2}(\sigma_{1}(t))dt}{\int_{0}^{t_{2}}f^{2}(\sigma_{2}(t))dt}\,\leq\,\frac{2\kappa\max_{K_{2\kappa}}f^{2}+\int_{2\kappa}^{t_{1}}f^{2}(\sigma_{1}(t))dt}{2\kappa\min_{K_{2\kappa}}f^{2}+\int_{2\kappa}^{t_{2}}f^{2}(\sigma_{2}(t))dt}\,\leq\,\frac{2\kappa\max_{K_{2\kappa}}f^{2}+(t_{1}-2\kappa)f(p)^{2}e^{\frac{2C}{\varepsilon}\kappa^{-\varepsilon}}}{2\kappa\min_{K_{2\kappa}}f^{2}+(t_{2}-2\kappa)f(p)^{2}e^{-\frac{2C}{\varepsilon}\kappa^{-\varepsilon}}}\,.

Notice that t1t_{1} and t2t_{2} are comparable (their difference is bounded via the triangle inequality by the maximum of the distance between points of Σ1\Sigma_{1} and Σ2\Sigma_{2}). Therefore, for any ε~>0\tilde{\varepsilon}>0 arbitrarily small, we can find κ~\tilde{\kappa} much larger than κ\kappa so that, assuming pp is outside Kκ~K_{\tilde{\kappa}} (in particular t1,t2t_{1},t_{2} are also much larger than κ\kappa) it holds

0t1f2(σ1(t))𝑑t0t2f2(σ2(t))𝑑t(1+ε~)e4Cεκε.\frac{\int_{0}^{t_{1}}f^{2}(\sigma_{1}(t))dt}{\int_{0}^{t_{2}}f^{2}(\sigma_{2}(t))dt}\,\leq\,(1+\tilde{\varepsilon})e^{\frac{4C}{\varepsilon}\kappa^{-\varepsilon}}\,.

Up to choosing κ\kappa large enough, we can also make the exponential term in the inequality above as close to 11 as necessary. Of course, exchanging the roles of σ1\sigma_{1} and σ2\sigma_{2} we also find the opposite bound. This proves uniformity. ∎

Equipped with the notion of uniformity of the ends, we are now ready to define the substatic version of the asymptotic volume ratio.

Definition 4.4.

Let (M,g,f)(M,g,f) be a substatic solution and let EE be a uniform ff-complete end. Let ρ\rho be the distance function to a point or a hypersurface with respect to the metric g~=g/f2\tilde{g}=g/f^{2} and η\eta be the solution to (2.10) or (2.25), respectively. The Asymptotic Volume Ratio AVR(E,g,f)\mathrm{AVR}(E,g,f) of EE is defined as

AVR(E,g,f)=1|𝕊n1|limt+{ρ=t}E1ηn1𝑑σ.{\rm AVR}(E,g,f)\,=\,\frac{1}{|\mathbb{S}^{n-1}|}\lim_{t\to+\infty}\int_{\{\rho=t\}\cap E}\frac{1}{\eta^{n-1}}d\sigma\,.

If (M,g,f)(M,g,f) has a unique end EE, we refer to AVR(E,g,f)\mathrm{AVR}(E,g,f) with AVR(M,g,f)\mathrm{AVR}(M,g,f).

The following basic fact motivates the introduction of the notion of uniform ends.

Proposition 4.5.

The substatic Asymptotic Volume Ratio is well-defined on any uniform ff-complete end. In other words, its definition does not depend on the choice of the point/hypersurface we are taking the distance ρ\rho from.

Proof.

Let ρ\rho be the g~\tilde{g}-distance from a point or a hypersurface. We consider the functional V(t)V(t) defined in (2.52), that we recall here for the reader’s convenience:

V(t)=1|𝔹n|tk{0ρt}ρk1fηn1𝑑μ,V(t)\,=\,\frac{1}{|\mathbb{B}^{n}|t^{k}}\int_{\{0\leq\rho\leq t\}}\frac{\rho^{k-1}}{f\eta^{n-1}}d\mu\,,

where k>0k>0 is a constant. A simple application of L’Hôpital’s rule tells us immediately that

limt+V(t)\displaystyle\lim_{t\to+\infty}V(t)\, =1|𝔹n|limt+1tk{ρt}ρk1fηn1𝑑μ\displaystyle=\,\frac{1}{|\mathbb{B}^{n}|}\lim_{t\to+\infty}\frac{1}{t^{k}}\int_{\{\rho\leq t\}}\frac{\rho^{k-1}}{f\eta^{n-1}}d\mu (4.7)
=nk|𝕊n1|limt+{ρ=t}1ηn1𝑑σ\displaystyle=\,\frac{n}{k|\mathbb{S}^{n-1}|}\lim_{t\to+\infty}\int_{\{\rho=t\}}\frac{1}{\eta^{n-1}}d\sigma (4.8)
=nklimt+A(t).\displaystyle=\,\frac{n}{k}\lim_{t\to+\infty}A(t)\,. (4.9)

In order to conclude the proof, it is then enough to show that limt+V(t)\lim_{t\to+\infty}V(t) is independent of the choice of the point/hypersurface. In the rest of the proof, it is convenient to set k=1k=1 in our functional VV. Let η1\eta_{1}, η2\eta_{2} be reparametrized distances with respect to two different points (resp. two different hypersurfaces) and let δ\delta be the distance between the two points (resp. the maximum distance between points of the two hypersurfaces) with respect to the metric g~=g/f2\tilde{g}=g/f^{2}. By triangle inequality we have the inclusion {ρ1tδ}{ρ2t}\{\rho_{1}\leq t-\delta\}\subset\{\rho_{2}\leq t\}, therefore

V1(t)V2(t)\displaystyle V_{1}(t)-V_{2}(t) =1|𝔹n|t{ρ1t}1fη1n1𝑑μ1|𝔹n|t{ρ2t}1fη2n1𝑑μ\displaystyle=\frac{1}{|\mathbb{B}^{n}|t}\int_{\{\rho_{1}\leq t\}}\frac{1}{f\eta_{1}^{n-1}}d\mu-\frac{1}{|\mathbb{B}^{n}|t}\int_{\{\rho_{2}\leq t\}}\frac{1}{f\eta_{2}^{n-1}}d\mu (4.10)
1|𝔹n|t{ρ1t}1fη1n1𝑑μ1|𝔹n|t{ρ1tδ}1fη2n1𝑑μ\displaystyle\leq\frac{1}{|\mathbb{B}^{n}|t}\int_{\{\rho_{1}\leq t\}}\frac{1}{f\eta_{1}^{n-1}}d\mu-\frac{1}{|\mathbb{B}^{n}|t}\int_{\{\rho_{1}\leq t-\delta\}}\frac{1}{f\eta_{2}^{n-1}}d\mu (4.11)
1|𝔹n|t{tδρ1t}1fη1n1𝑑μ+1|𝔹n|t{ρ1tδ}1fη1n1(1η1n1η2n1)𝑑μ.\displaystyle\leq\frac{1}{|\mathbb{B}^{n}|t}\int_{\{t-\delta\leq\rho_{1}\leq t\}}\frac{1}{f\eta_{1}^{n-1}}d\mu+\frac{1}{|\mathbb{B}^{n}|t}\int_{\{\rho_{1}\leq t-\delta\}}\frac{1}{f\eta_{1}^{n-1}}\left(1-\frac{\eta_{1}^{n-1}}{\eta_{2}^{n-1}}\right)d\mu\,. (4.12)

Concerning the first integral, applying again L’Hôpital’s rule we find that its limit is the same as the limit of nA(t)nA(tδ)nA(t)-nA(t-\delta) at t+t\to+\infty, where AA is the usual area functional with respect to the distance ρ1\rho_{1}. Since A(t)A(t) has a finite limit at infinity and δ\delta is fixed, this limit is zero. From the uniformity of the end and the fact that V(t)V(t) is bounded, we deduce that the second integral also goes to zero. Hence, we have found that limt+[V1(t)V2(t)]0\lim_{t\to+\infty}[V_{1}(t)-V_{2}(t)]\leq 0. Switching the roles of V1V_{1} and V2V_{2} we find that the opposite inequality is also in place, hence the limits of V1(t)V_{1}(t) and V2(t)V_{2}(t) are the same, as wished. ∎

Remark 4.6.

As noted in Remark 2.4, η\eta represents the distance along radial g~\tilde{g}-geodesics with respect to the metric g¯=f2g\overline{g}=f^{2}g. Providing a suitable Bishop-Gromov-type Theorem in terms of the g¯\overline{g}-distance in place of η\eta may be useful to cook a notion of Asymptotic Volume Ratio that does not need the notion of uniformity to be well defined.

The following is a basic yet fundamental consequence of the Splitting Theorem C.

Lemma 4.7.

Let (M,g,f)(M,g,f) be a substatic triple with ff-complete ends. If there is more than one uniform ff-complete end, then all ends have vanishing asymptotic volume ratio.

Proof.

Suppose that there is more than one uniform ff-complete end. Then the Splitting Theorem C implies that the manifold splits as a twisted product

(×Σ,f2dsds+gΣ),(\mathbb{R}\times\Sigma,\,f^{2}\,ds\otimes ds+g_{\Sigma})\,,

for some (n1)(n-1)-dimensional Riemannian manifold (Σ,gΣ)(\Sigma,g_{\Sigma}). Let ρ\rho be the g~\tilde{g}-distance from the cross section {s=0}\{s=0\} and η\eta be defined by (2.25), as usual. Notice that the level sets of ρ\rho and η\eta are also level sets of ss, hence in particular the metric induced on any level set of ρ\rho is gΣg_{\Sigma}. It follows then that

AVR(E,g,f)=1|𝕊n1|limt+{ρ=t}E1ηn1𝑑σ=limt+|Σ||𝕊n1|η|{ρ=t}n1.{\rm AVR}(E,g,f)\,=\,\frac{1}{|\mathbb{S}^{n-1}|}\lim_{t\to+\infty}\int_{\{\rho=t\}\cap E}\frac{1}{\eta^{n-1}}d\sigma\,=\,\lim_{t\to+\infty}\frac{|\Sigma|}{|\mathbb{S}^{n-1}|\eta_{|_{\{\rho=t\}}}^{n-1}}\,.

Since the end is ff-complete, we have η+\eta\to+\infty at infinity, hence the above limit vanishes. ∎

In light of the above Lemma, our main geometric inequalities 1.3 and 1.13 will only involve one end and the global AVR(M,g,f)\mathrm{AVR}(M,g,f).

In this framework, we now discuss some cases in which we are able to give more precise estimates for the Asymptotic Volume Ratio. A first simple estimate, in the case where the boundary is empty, is obtained from (2.51). Taking the limit of this formula as t+t\to+\infty, assuming that the boundary is empty (so that the term A(t)A(t) appearing in that formula converges to the asymptotic volume ratio) we find the following

AVR(M,g,f)f(p)n1 1.{\rm AVR}(M,g,f)\,f(p)^{n-1}\,\leq\,1\,.

This must hold for any point pMp\in M. In particular, it follows that if M=Ø\partial M=\mathchar 31\relax and ff is not bounded then the Asymptotic Volume Ratio must vanish.

An important family of substatic manifolds having nonzero AVR is that of asymptotically flat triples, that we now define precisely.

Definition 4.8.

A substatic triple (M,g,f)(M,g,f) is said to be asymptotically flat if

  • (i)(i)

    there exists a compact domain KMK\supset\partial M and a diffeomorphism (called chart at infinity) between MKM\setminus K and n\mathbb{R}^{n} minus a ball.

  • (ii)(ii)

    in the chart at infinity, it holds |gijδij|=o(1)|g_{ij}-\delta_{ij}|=o(1) and |f1|=o(1)|f-1|=o(1) as |x|+|x|\to+\infty.

We remark that the usual definition of asymptotic flatness requires a higher degree of convergence of the metric gg to the Euclidean one. However, the above definition is sufficient to compute precisely the asymptotic volume ratio.

Proposition 4.9.

Let (M,g,f)(M,g,f) be an asymptotically flat substatic triple. Then AVR(M,g,f)=1{\rm AVR}(M,g,f)=1.

Proof.

The fact that the end is ff-complete and uniform follows from Proposition 4.2. Let KK be a compact set as in Definition 4.8 and let S={|x|=R}S=\{|x|=R\} be a large coordinate sphere contained in the chart at infinity. From Proposition 4.5 we know that the asymptotic volume ratio does not depend on the hypersurface we are taking the distance from. It is then convenient to work with the g~\tilde{g}-distance ρ\rho from the coordinate sphere SS. As it follows from (4.9), we can also compute the AVR via the following limit

AVR(M,g,f)=1|𝔹n|limt+1tn{0ρt}ρn1fηn1𝑑μ.{\rm AVR}(M,g,f)\,=\,\frac{1}{|\mathbb{B}^{n}|}\lim_{t\to+\infty}\frac{1}{t^{n}}\int_{\{0\leq\rho\leq t\}}\frac{\rho^{n-1}}{f\eta^{n-1}}d\mu\,. (4.13)

If the radius RR of the coordinate sphere SS is large, then we can assume |f1|<ε|f-1|<\varepsilon and |gijδij|<ε|g_{ij}-\delta_{ij}|<\varepsilon in {|x|>R}\{|x|>R\} for some fixed small ε\varepsilon. It is then easily seen that there exists δ=δ(ε)\delta=\delta(\varepsilon) such that

{R|x|(R+t)(1δ)}{0ρt}{R|x|(R+t)(1+δ)}\{R\leq|x|\leq(R+t)(1-\delta)\}\subset\{0\leq\rho\leq t\}\subset\{R\leq|x|\leq(R+t)(1+\delta)\}

for all tt. Since η\eta grows as f2f^{2} along g~\tilde{g}-geodesics, we have (1ε)2ρ<η<(1+ε)2ρ(1-\varepsilon)^{2}\rho<\eta<(1+\varepsilon)^{2}\rho. It follows that the integral in (4.13) grows as the Euclidean volume of the annulus {R<|x|<R+t}\{R<|x|<R+t\}, or more explicitly as:

[(R+t)nRn]|𝔹n||𝔹n|tn.\left[(R+t)^{n}-R^{n}\right]\,|\mathbb{B}^{n}|\,\cong\,|\mathbb{B}^{n}|t^{n}\,.

The wished result follows easily. ∎

4.2. Willmore inequality

As a consequence of our definition of AVR and the Bishop–Gromov monotonicity of the area functional A(t)A(t) (Theorem 2.9), we obtain the Willmore inequality for hypersurfaces with nonnegative mean curvature of Theorem D. The following statement provides more details about the equality case.

Theorem 4.10 (Willmore inequality).

Let (M,g,f)(M,g,f) be a substatic solution with a uniform ff-complete end. Let Σ\Sigma be a hypersurface that is homologous to the boundary. Suppose that the mean curvature H{\rm H} of Σ\Sigma with respect to the normal pointing towards infinity satisfies H>0{\rm H}>0 pointwise. Then

Σ[H(n1)f]n1𝑑σAVR(M,g,f)|𝕊n1|.\int_{\Sigma}\left[\frac{{\rm H}}{(n-1)f}\right]^{n-1}d\sigma\,\geq\,{\rm AVR}(M,g,f)\,|\mathbb{S}^{n-1}|\,. (4.14)

If the equality holds, then the set U={ρ>0}U=\{\rho>0\} is isometric to [0,+)×Σ[0,+\infty)\times\Sigma with metric

g=f2dρdρ+η2g0,g\,=\,f^{2}d\rho\otimes d\rho+\eta^{2}g_{0}\,,

where g0g_{0} is a metric on the level set Σ\Sigma. Furthermore, in UU the functions ff and η\eta satisfy

η=(α+β)1n1,f2=α˙(n1)(α+β)n2n1\eta\,=\,(\alpha+\beta)^{\frac{1}{n-1}}\,,\qquad f^{2}\,=\,\frac{\dot{\alpha}}{(n-1)(\alpha+\beta)^{\frac{n-2}{n-1}}}

where α\alpha is a function of ρ\rho and β\beta is a function on Σ\Sigma.

Proof.

We recall from Theorem 2.9 that A(t)A(t) is monotonically nonincreasing. Taking the limit as t+t\to+\infty we then get

1|𝕊n1|Σ[H(n1)f]n1𝑑σ=A(0)limt+A(t)=AVR(M,g,f).\frac{1}{|\mathbb{S}^{n-1}|}\int_{\Sigma}\left[\frac{{\rm H}}{(n-1)f}\right]^{n-1}d\sigma\,=\,A(0)\,\geq\,\lim_{t\to+\infty}A(t)\,=\,{\rm AVR}(M,g,f)\,.

This proves the inequality.

Furthermore, if the equality holds, then from the rigidity statement in Theorem 2.9 it follows

1ηηθi=ψηφ1ηn1,1ffθi=ψη+n22φ1ηn1,\frac{1}{\eta}\frac{\partial\eta}{\partial\theta^{i}}\,=\,\psi\eta-\varphi\frac{1}{\eta^{n-1}}\,,\qquad\frac{1}{f}\frac{\partial f}{\partial\theta^{i}}\,=\,\psi\eta+\frac{n-2}{2}\varphi\frac{1}{\eta^{n-1}}\,,

where φ\varphi and ψ\psi are functions on Σ\Sigma. From the first of these equations, in particular we get

θi(1η)=ψφ1ηn.\frac{\partial}{\partial\theta^{i}}\left(\frac{1}{\eta}\right)\,=\,\psi-\varphi\frac{1}{\eta^{n}}\,.

We now focus on this identity near infinity: since our end is ff-complete, we know that η\eta is going to ++\infty. Furthermore, from the uniformity at infinity we can also prove that 1/η1/\eta goes to zero uniformly at infinity. To prove that, it is enough to apply the uniformity at infinity property to the two hypersurfaces Σ\Sigma and Σδ={ρ=δ}\Sigma_{\delta}=\{\rho=\delta\}. Notice that then the function ηδ\eta_{\delta} associated to Σδ\Sigma_{\delta} differs from η\eta just by a constant kk, that is ηδ=η+k\eta_{\delta}=\eta+k. Then uniformity at infinity implies precisely that for any ε\varepsilon there exists a compact set such that

|kη|=|k+ηη1|ε.\left|\frac{k}{\eta}\right|\,=\,\left|\frac{k+\eta}{\eta}-1\right|\,\leq\,\varepsilon\,.

Hence, 1/η1/\eta goes to 0 uniformly at infinity, as wished.

Given ε>0\varepsilon>0, fix R>0R>0 big enough so that 1/η<ε1/\eta<\varepsilon in [R,+)×V[R,+\infty)\times V. If ψ\psi is not everywhere vanishing, then there is an open set VΣV\subset\Sigma such that ψ>δ\psi>\delta in VV (the case ψ<δ\psi<-\delta is done in the exact same way). Therefore we would get

θi(1η)>δ|φ|εnδεnmaxΣ|φ|=δ~\frac{\partial}{\partial\theta^{i}}\left(\frac{1}{\eta}\right)\,>\,\delta-|\varphi|\,\varepsilon^{n}\,\geq\,\delta-\varepsilon^{n}\max_{\Sigma}|\varphi|=\tilde{\delta}

in [R,+)×V[R,+\infty)\times V. Up to taking ε\varepsilon small enough, we can assume that δ~>0\tilde{\delta}>0. But then, for any two points pρ=(ρ,θ1,,θi,,θn)p_{\rho}=(\rho,\theta^{1},\dots,\theta^{i},\dots,\theta^{n}), qρ=(ρ,θ1,,θi+λ,,θn)q_{\rho}=(\rho,\theta^{1},\dots,\theta^{i}+\lambda,\dots,\theta^{n}) belonging to [R,+)×V[R,+\infty)\times V, we would deduce

1η(pρ)=1η(qρ)+0λθi(1η)|(ρ,θ1,,θi+t,,θn)𝑑tλδ~,\frac{1}{\eta}(p_{\rho})\,=\,\frac{1}{\eta}(q_{\rho})+\int_{0}^{\lambda}\frac{\partial}{\partial\theta^{i}}{\left(\frac{1}{\eta}\right)}_{|_{(\rho,\theta^{1},\dots,\theta^{i}+t,\dots,\theta^{n})}}dt\,\geq\,\lambda\tilde{\delta}\,,

which in turn would imply limρ+(1/η)(pρ)λδ~>0\lim_{\rho\to+\infty}(1/\eta)(p_{\rho})\geq\lambda\tilde{\delta}>0, contradicting the fact that 1/η01/\eta\to 0 at infinity. It follows then that

ψ=0.\psi=0\,.

Our constraints on ff and η\eta then become

1ηηθi=φ1ηn1,1ffθi=n22φ1ηn1.\frac{1}{\eta}\frac{\partial\eta}{\partial\theta^{i}}\,=\,-\varphi\frac{1}{\eta^{n-1}}\,,\qquad\frac{1}{f}\frac{\partial f}{\partial\theta^{i}}\,=\,\frac{n-2}{2}\varphi\frac{1}{\eta^{n-1}}\,. (4.15)

The first equation can be rewritten as

θiηn1=(n1)φ.\frac{\partial}{\partial\theta^{i}}\eta^{n-1}\,=\,-(n-1)\varphi\,.

Since φ\varphi does not depend on ρ\rho, it follows then that

ηn1=α+β,\eta^{n-1}\,=\,\alpha+\beta\,,

where α\alpha is a function of ρ\rho and β\beta is a function on Σ\Sigma. Taking the derivative with respect to ρ\rho of this formula and using the fact that ρη=f2\partial_{\rho}\eta=f^{2}, we then deduce

f2=α˙(n1)(α+β)n2n1f^{2}\,=\,\frac{\dot{\alpha}}{(n-1)(\alpha+\beta)^{\frac{n-2}{n-1}}}

It is easy to check that for any ff and η\eta of this form (that is, for any choice of α\alpha and β\beta), formulas (4.15) are satisfied. ∎

5. Isoperimetric Inequality for Substatic manifolds

In this Section, we focus our attention on substatic manifolds (M,g,f)(M,g,f) admitting an exhaustion of outward minimising hypersurfaces homologous to M\partial M. A hypersurface Σ\Sigma homologous to M\partial M is outward minimizing if, denoting by Ω\Omega the compact domain with Ω=ΣM\partial\Omega=\Sigma\sqcup\partial M, we have

P(Ω)P(F)P(\Omega)\leq P(F) (5.1)

for any bounded set FΩF\supset\Omega. We say that a sequence (Sj)j(S_{j})_{j\in\mathbb{N}} of hypersurfaces homologous to M\partial M exhaust MM if, given a compact set KMK\subset M, there exists an element SS in the sequence such that KΩK\subset\Omega, for Ω\Omega satisfying Ω=SM\partial\Omega=S\sqcup\partial M. Conditions ensuring the existence of such an exhaustion are discussed in [FM22].

We start from showing that M\partial M is a priori area minimizing. In showing so, we also derive that M\partial M is outermost, that is, there exist no minimal submanifolds homologous to M\partial M other than the boundary itself. These facts are the first main reason why we require the existence of (nonminimal) outward minimizing sets homologous to M\partial M. Since the following auxiliary result does not need any a priori growth at infinity assumption, we think it may have an independent interest.

Proposition 5.1 (The boundary is outermost area-minimizing).

Let (M,g,f)(M,g,f) be a substatic triple with horizon boundary. Assume that there exists an outward minimizing smooth hypersurface SS homologous to M\partial M. Then, the horizon is outward minimizing, meaning that

|Σ||M|\lvert\Sigma\rvert\geq\lvert\partial M\rvert (5.2)

for any hypersurface Σ\Sigma homologous to M\partial M. Moreover, it is outermost, that is there exists no other minimal hypersurfaces homologous to M\partial M.

Proof.

Let Ω\Omega be such that Ω=SM\partial\Omega=S\sqcup\partial M. Indeed SS being outward minimizing is mean-convex, and consequently the Maximum Principle implies it is disjoint from M\partial M (see e.g. [Lee19, Corollary 4.2]; it is a consequence of the strong comparison principle for quasilinear equations). We flow Ω\Omega by weak Mean Curvature Flow, referring to the notion considered in [Ilm92]. In particular the analysis carried out by White [Whi00] applies. Moreover, observe that the mean curvature of SS is necessarily nonnegative, and in particular Ω\Omega is mean-convex in the sense of [Whi00, Section 3]. Since M\partial M constitutes itself a (steady) MCF, the well-known [Ilm92, Inclusion Property 5.3] ensures that the possibly singular evolving sets ΩtM\partial\Omega_{t}\setminus\partial M remain homologous to the horizon. By [Whi00, Theorem 11.1], Ωt\partial\Omega_{t} must converge smoothly to a minimal hypersurface Σ\Sigma, necessarily homologous to M\partial M. We show that Σ\Sigma can be the horizon only. Indeed, if this were not the case, Σ\Sigma would be detached from M\partial M by the Maximum Principle, and (iii)(iii) in Proposition 2.7 would apply, foliating an outer neighbourhood of Σ\Sigma with hypersurfaces of nonpositive mean curvature. But this is a contradiction, through the Maximum Principle for the mean curvature operator, applied on tangency points, with the smooth mean-convex Mean Curvature Flow smoothly approaching Σ\Sigma. Then, the Mean Curvature Flow of Ω\Omega converges smoothly to M\partial M. Observe that this also implies that no minimal hypersurface homologous to M\partial M contained in Ω\Omega can exist. Indeed, if there were one, it would obviously remain fixed under MCF, and thus Ωt\Omega_{t} converging to M\partial M would eventually go beyond it, contradicting [Ilm92, Inclusion Property 5.3]. The nonminimal outward minimizing sets forming an exhaustion, we in particular proved that M\partial M is outermost.

Finally, recall that the outward minimizing property of the initial set is preserved along the flow, as it can be easily checked applying [Whi00, One-Sided Minimization Theorem 3.5] (see [HI01, Lemma 5.6] for a proof in the smooth flow setting). Then, M\partial M being one-sided limit of outward minimizing hypersurfaces homologous to the boundary, is outward minimizing as well. ∎

As already pointed out in the Introduction, our proof of Theorem A ultimately builds on the application of the Willmore-type inequality (4.14) on hypersurfaces homologous to M\partial M bounding a set that is isoperimetric with respect to the volume weighted by ff. In order to bypass the lack of existence, we will consider constrained isoperimetric sets. We find convenient to extend (M,g)(M,g) over the horizon, letting (N,gN)(N,g_{N}) be the extended Riemannian manifold. This can be obtained through gluing another copy of MM along its boundary, and endowing it with a smooth metric that coincides with gg on the original manifold. The existence of such metric is ensured by [PV20, Theorem A]. Let SS be homologous to M\partial M and disjoint from it, and let ΩM\Omega\subset M have boundary SMS\sqcup\partial M. Extend Ω\Omega too, so to find ΩNN\Omega_{N}\subset N satisfying ΩNM=Ω\Omega_{N}\cap M=\Omega, and let

BεNM(M)={pNM|dN(p,M)ε},B^{N\setminus M}_{\varepsilon}(\partial M)=\{p\in N\setminus M\,|\,\mathrm{d}_{N}(p,\partial M)\leq\varepsilon\}, (5.3)

for ε>0\varepsilon>0 such that BεNM(M)ΩNB^{N\setminus M}_{\varepsilon}(\partial M)\subset\Omega_{N} , where dN\mathrm{d}_{N} is the distance induced by the metric gNg_{N}. We are going to consider sets of finite perimeter EVE_{V} in (N,gN)(N,g_{N}) satisfying

|EVM|f=VP(EV)=inf{P(F)|BεNM(M)FΩN,|FM|f=V}\lvert E_{V}\cap M\rvert_{f}=V\qquad\qquad P(E_{V})=\inf\left\{P(F)\,|\,B^{N\setminus M}_{\varepsilon}(\partial M)\subset F\subset\Omega_{N},\lvert F\cap M\rvert_{f}=V\right\} (5.4)

for V<|Ω|fV<\lvert\Omega\rvert_{f}, where we recall that given EME\subset M we defined

|E|f=Ef𝑑μ.\lvert E\rvert_{f}=\int_{E}fd\mu. (5.5)

The following result gathers the main properties these constrained isoperimetric sets satisfy.

Theorem 5.2 (Existence and structure of constrained ff-isoperimetric sets).

Let (M,g,f)(M,g,f) be a substatic triple with horizon boundary, of dimension n7n\leq 7. Let SS be a strictly mean-convex outward minimizing hypersurface homologous to M\partial M, and let (N,gN),Ω,ΩN(N,g_{N}),\Omega,\Omega_{N} and BεN(M)B^{N}_{\varepsilon}(\partial M) as above, for ε>0\varepsilon>0. Then, for any V<|Ω|fV<\lvert\Omega\rvert_{f}, there exists EVΩNE_{V}\subset\Omega_{N} satisfying 5.4. Moreover,

  • (i)(i)

    EVM=Ø\partial E_{V}\cap\partial M=\mathchar 31\relax. Moreover, (EVM)=ΣM\partial(E_{V}\cap M)=\Sigma\sqcup\partial M, where Σ\Sigma is a 𝒞1,1\mathscr{C}^{1,1}-hypersurface.

  • (ii)(ii)

    The set ΣS\Sigma\setminus S is a smooth hypersurface. Moreover, there exists a positive constant λ\lambda such that H(x)=λf(x){\rm H}(x)=\lambda f(x) for any xΣSx\in\Sigma\setminus S.

  • (iii)(iii)

    We have

    λHf(x)>0\lambda\geq\frac{{\rm H}}{f}(x)>0 (5.6)

    for (n1)(n-1)-almost any xΣx\in\Sigma.

Proof.

The existence of EVE_{V} directly follows from the Direct Method. Indeed, let (Fj)j(F_{j})_{j\in\mathbb{N}} be a minimizing sequence for 5.4. Then, by compactness, up to subsequences it converges to a set EVΩNE_{V}\subset\Omega_{N} in L1L^{1}. In particular, we have

limj+Mf|χEVχFj|𝑑μlimj+supΩf|(EVFj)M|=0.\lim_{j\to+\infty}\int_{M}f\lvert\chi_{E_{V}}-\chi_{F_{j}}\rvert d\mu\leq\lim_{j\to+\infty}\sup_{\Omega}f\lvert(E_{V}\bigtriangleup F_{j})\cap M\rvert=0. (5.7)

So, |EVM|=V\lvert E_{V}\cap M\rvert=V. By the convergence almost everywhere one also deduces that BεNM(M)EVΩNB_{\varepsilon}^{N\setminus M}(\partial M)\subset E_{V}\subset\Omega_{N} is satisfied too. The lower semicontinuity of the perimeter also ensures that the infimum in 5.4 is attained by EVE_{V}.

As far as the regularity of (EVM)\partial(E_{V}\cap M) is concerned, let us first crucially observe that EVME_{V}\cap M is (constrained) isoperimetric in MM endowed with the conformal metric g¯=f2g\overline{g}=f^{2}g with respect to a perimeter and volume with the same weight, namely with respect to

P(E)=Ef1n𝑑σg¯,|E|f=Ef1n𝑑μg¯,{P}(E)=\int_{\partial^{*}E}f^{1-n}d\sigma_{\overline{g}},\quad\lvert E\rvert_{f}=\int_{E}f^{1-n}d\mu_{\overline{g}}, (5.8)

where dσg¯d\sigma_{\overline{g}} and dμg¯d\mu_{\overline{g}} denote the area and volume measure induced by g¯\overline{g} respectively. In particular, away from SS and M\partial M, where g¯\overline{g} becomes singular, we have that classical regularity for the weighted isoperimetric problem applies [Mor03, Section 3.10], and implies the claimed smoothness. In order to prove the global 𝒞1,1\mathscr{C}^{1,1}-regularity, we mainly follow the nice exposition in [MS17, Section 6], taking advantage also of [Mag12, Section 17]. We first show that EVME_{V}\cap M is an almost minimizer for the perimeter. This amounts to say that there exists r0r_{0} such that for every xΩx\in\Omega and every r<r0r<r_{0}

P(EVM)P(F)+CrnP(E_{V}\cap M)\leq P(F)+\mathrm{C}r^{n} (5.9)

holds for any FF such that (EVM)FB(x,r)(E_{V}\cap M)\bigtriangleup F\Subset B(x,r), for some constant C\mathrm{C} independent of xx and rr. Observe that B(x,r)B(x,r) can intersect ΩNΩ\Omega_{N}\setminus\Omega, and this is the main reason we did extend our substatic manifold. Let for simplicity E=EVME=E_{V}\cap M, consider two small enough balls B1B_{1} and B2B_{2} centered on EM\partial E\setminus\partial M with B1,B2MMB_{1},B_{2}\Subset M\setminus\partial M, and let X1X_{1} and X2X_{2} be variation vector fields compactly supported in B1B_{1} and B2B_{2} respectively. Let Eti=ψti(E)E_{t}^{i}=\psi_{t}^{i}(E), where ψti\psi_{t}^{i} is the flow of XiX_{i} at time tt, for i=1,2i=1,2. By [Mag12, Proposition 17.8], we have

|Eti|f=|E|+tEfXi|νE𝑑σ+O(t2)\lvert E_{t}^{i}\rvert_{f}=\lvert E\rvert+t\int_{\partial E}f\langle X_{i}|\nu_{E}\rangle d\sigma+O(t^{2}) (5.10)

as t0t\to 0, where νE\nu_{E} is a unit normal for EE. If f>c>0f>c>0 uniformly on B1B2B_{1}\cup B_{2}, we deduce

||Eti||E||C|t|\left|\lvert E_{t}^{i}\rvert-\lvert E\rvert\right|\geq\mathrm{C}\lvert t\rvert (5.11)

for tt in some small neighbourhood of 0 and for some uniform constant C\mathrm{C}. Moreover, the perimeter satisfies the usual expansion [Mag12, Theorem 17.5]

P(Eti)=P(E)+tEdivEXi𝑑σ+O(t2)P(E^{i}_{t})=P(E)+t\int_{\partial E}\mathrm{div}_{\partial E}X_{i}d\sigma+O(t^{2}) (5.12)

as t0t\to 0, where divE\mathrm{div}_{\partial E} denotes the tangential divergence divEXi=div(Xi)Xi|νE\mathrm{div}_{\partial E}X_{i}=\mathrm{div}(X_{i})-\langle X_{i}|\nu_{E}\rangle. Thus,

||P(Eti)||P(E)||C|t|,\left|\lvert P(E^{i}_{t})\rvert-\lvert P(E)\rvert\right|\leq\mathrm{C}\lvert t\rvert, (5.13)

again for tt in some small neighbourhood of 0 and for some uniform constant CC. We can now conclude the proof of 5.9 for the suitable competitors FF as done for the proof of [MS17, Lemma 6.3]. Namely, letting FF as above, we recover the possibly lost or gained ff-volume δ\delta by the competitor FΩF\cap\Omega by slightly deforming EE inside BiB_{i} with ii chosen so that (FE)Bi=Ø(F\bigtriangleup E)\cap B_{i}=\mathchar 31\relax. Observe that |δ|Crn\lvert\delta\rvert\leq\mathrm{C}r^{n} for some suitable constant. Exploiting 5.11 and 5.13 we get a set F~\tilde{F} with |E|f=|F~|f\lvert E\rvert_{f}=\lvert\tilde{F}\rvert_{f} such that

P(F~)P(FΩ)+C|δ|P(FΩ)+CrnP(\tilde{F})\leq P(F\cap\Omega)+\mathrm{C}\lvert\delta\rvert\leq P(F\cap\Omega)+\mathrm{C}r^{n} (5.14)

for some suitable C>0\mathrm{C}>0, uniform for any r<r0r<r_{0}, with r0r_{0} small enough. Since EE is constrained ff-isoperimetric, we have then

P(E)P(F~)P(F)+P(Ω)P(FΩ)+Crn.P(E)\leq P(\tilde{F})\leq P(F)+P(\Omega)-P(F\cup\Omega)+\mathrm{C}r^{n}. (5.15)

Observe that FΩF\cup\Omega may intersect ΩNΩ\Omega_{N}\setminus\Omega. On the other hand, it is easy to notice that sets with smooth boundary are in fact automatically almost minimizers for the perimeter (see e.g. the derivation of [MS17, (6-9)]), and so P(Ω)P(FΩ)+CrnP(\Omega)\leq P(F\cup\Omega)+\mathrm{C}r^{n}. Plugging it into 5.15 concludes the proof of E=EVME=E_{V}\cap M being almost minimizing. From this crucial property, one deduces that E\partial E is C1,1/2C^{1,1/2} in a neighbourhood of EΩ\partial E\cap\partial\Omega exactly as exposed in [MS17, Proof of Proposition 6.1].

To establish the optimal 𝒞1,1\mathscr{C}^{1,1} regularity, we first take advantage of (ii)(ii), which we proceed to prove. In order to make the comparison with the references easier, along this proof we are going to assume that νE\nu_{E} is the interior unit normal to EE, in the extended manifold. Let YY be a vector field supported in some small ball centered at some point of E\partial E, with a flow ψ\psi such that ψt(E)Ω\psi_{t}(E)\subset\Omega, for tt small enough. Let now XX satisfy the same assumptions, with the additional requirement to be supported around the smooth part of EΩ\partial E\setminus\partial\Omega and such that suppXsuppY=Ø\mathrm{supp}X\cap\mathrm{supp}Y=\mathchar 31\relax. Assume also that the composition of the two flows gives a ff-volume preserving diffeomorphism. Then, the first variation formula gives

0EdivE(Y+X)𝑑σ=EdivEY𝑑σEHX|νE𝑑σ.0\leq\int_{\partial E}\mathrm{div}_{\partial E}(Y+X)d\sigma=\int_{\partial E}\mathrm{div}_{\partial E}Yd\sigma-\int_{\partial E}{\rm H}\langle X|\nu_{E}\rangle d\sigma. (5.16)

Assume for the time being that YY is supported around a point where EΩ\partial E\setminus\partial\Omega is smooth. Then we can integrate by parts also in the integrand involving YY, and obtain

0EHY+X|νE𝑑σ.0\leq-\int_{\partial E}{\rm H}\langle Y+X|\nu_{E}\rangle d\sigma. (5.17)

Repeating the argument with X-X and Y-Y in place of XX and YY, that is possible in the present case since these vector fields are supported away from Ω\partial\Omega, we actually get

0=EHY+X|νE𝑑σ.0=\int_{\partial E}{\rm H}\langle Y+X|\nu_{E}\rangle d\sigma. (5.18)

Moreover, the ff-volume being preserved entails, by 5.10,

0=EfY+X|νE𝑑σ.0=\int_{\partial E}f\langle Y+X|\nu_{E}\rangle d\sigma. (5.19)

Choose now XX and YY so that X=ανEX=\alpha\nu_{E} and Y=βνEY=-\beta\nu_{E} on E\partial E, with α\alpha and β\beta being smooth functions compactly supported on E\partial E. Combining (5.19) with 5.18 with this choice of XX and YY, we obtain

E(Hf)fα𝑑σEfα𝑑σ=E(Hf)fβ𝑑σEfα𝑑σ=E(Hf)fβ𝑑σEfβ𝑑σ.\frac{\int_{\partial E}\left(\frac{{\rm H}}{f}\right)f\alpha d\sigma}{\int_{\partial E}f\alpha d\sigma}=\frac{\int_{\partial E}\left(\frac{{\rm H}}{f}\right)f\beta d\sigma}{\int_{\partial E}f\alpha d\sigma}=\frac{\int_{\partial E}\left(\frac{{\rm H}}{f}\right)f\beta d\sigma}{\int_{\partial E}f\beta d\sigma}. (5.20)

Then, since the support of α\alpha and β\beta on E\partial E can be chosen arbitrarily close to any two points in the smooth part of EΩ\partial E\setminus\partial\Omega, we conclude that there exists λ\lambda\in\mathbb{R} such that H=λf{\rm H}=\lambda f in the smooth part of EΩ\partial E\setminus\partial\Omega. We plug this information into 5.16, for a vector field YY that is now supported in a ball centered on a point of EΩ\partial E\cap\partial\Omega. Coupling with 5.19, this yields

EdivEY𝑑σ+λEfY|νE𝑑σ0.\int_{\partial E}\mathrm{div}_{\partial E}Yd\sigma+\lambda\int_{\partial E}f\langle Y|\nu_{E}\rangle d\sigma\geq 0. (5.21)

Writing 5.21 in local coordinates in a neighbourhood of a point in EΩ\partial E\cap\partial\Omega, where E\partial E is given by the graph of a function u:Bu:B\to\mathbb{R} and Ω\partial\Omega as the graph of a function ψ:B\psi:B\to\mathbb{R} with Bn1B\subset\mathbb{R}^{n-1}, and YY as a normal vector field cut off with a function φ\varphi, it is a routine computation to check that, for some quasilinear elliptic operator LL, it holds

Bφ[Lu+λf(,u())]𝑑μ0,\int_{B}\varphi[-Lu+\lambda f(\cdot,u(\cdot))]d\mu\geq 0, (5.22)

where φ𝒞c(B)\varphi\in\mathscr{C}^{\infty}_{c}(B) and φ0\varphi\geq 0 in {u=ψ}\{u=\psi\}. We address the interested reader to [MS17, Section 6C] and [FGS17, Section 4.1] for details of these computations. In particular, the function uu succumbs to the regularity theory for obstacle problems, that is u𝒞1,1u\in\mathscr{C}^{1,1}, see [FGS17, Theorem 3.8]. As a consequence, E\partial E has a notion of mean curvature defined almost everywhere. Observe that the quasilinear elliptic operator LuLu provides the mean curvature of E\partial E at the point (x,u(x))(x,u(x)), with xBx\in B. We are going to take advantage also of the basic step leading to the regularity result recalled above. Namely, as nicely presented in [FGS17, Proposition 3.2], the variational property 5.22 implies that uu is also solution to the Euler–Lagrange equation

Bφ[Lu+λf(,u())]𝑑μ=Bξ𝑑μ,\int_{B}\varphi[-Lu+\lambda f(\cdot,u(\cdot))]d\mu=\int_{B}\xi\,d\mu, (5.23)

given any φ𝒞c(B)\varphi\in\mathscr{C}^{\infty}_{c}(B), where

0ξ[Lψ+λf(,ψ())]+χ{u=ψ}.0\leq\xi\leq[-L\psi+\lambda f(\cdot,\psi(\cdot))]^{+}\chi_{\{u=\psi\}}. (5.24)

We now proceed to show that λ>0\lambda>0, that EV\partial E_{V} is disjoint from M\partial M and that 5.6 holds, completing thus the proof. Let YY in 5.21 be supported on a neighbourhood of a point ES\partial E\cap S. Integrating by parts the first summand in 5.21, and letting Y=ανEY=\alpha\nu_{E} for some compactly supported nonnegative test function α\alpha, we get

E(λfH)α𝑑σ0.\int_{\partial E}(\lambda f-{\rm H})\alpha\,d\sigma\geq 0. (5.25)

The arbitrariness of α\alpha implies that

λHf(x)for (n1)-almost any xEM.\lambda\geq\frac{{\rm H}}{f}(x)\qquad\text{for $(n-1)$-almost any $x\in\partial E\setminus\partial M$}. (5.26)

Since SS is mean-convex, if the (n1)(n-1)-induced measure of ES\partial E\cap S is strictly positive, then the (weak) mean curvature of such region is strictly positive [MS17, Lemma 6.10], and consequently 5.26 directly implies λ>0\lambda>0 and 5.6. If instead the intersection is (n1)(n-1)-negligible, then 5.23 implies that Lu=λfLu=\lambda f holds in the weak sense outside of M\partial M. In particular, classical regularity theory implies that EM\partial E\setminus\partial M is smooth and that its mean curvature is given by λf\lambda f. Assume by contradiction that λ<0\lambda<0. Hence, by the Maximum Principle, ES=Ø\partial E\cap S=\mathchar 31\relax. Then, SS being outward minimizing acts as a barrier to minimize the perimeter among sets homologous to M\partial M containing EE [FM22, Theorem 2.10]. We call EE^{*} such a minimizer. The boundary of such set is 𝒞1,1\mathscr{C}^{1,1} [SZW91], and obviously it is outward minimizing, so that its (weak) mean curvature is nonnegative. Having assumed that λ<0\lambda<0, and since M\partial M is minimal, we deduce that E\partial E^{*} is minimal itself, and by the Maximum Principle disjoint from M\partial M. However, this is a contradiction with M\partial M being outermost, proved in Proposition 5.1. We established that λ0\lambda\geq 0, and that λ>0\lambda>0 if ES\partial E\cap S has positive (n1)(n-1)-measure. We focus our attention to 5.23 again, considering a neighbourhood of a point where E\partial E meets M\partial M. Crucially observe that by the minimality of M={f=0}\partial M=\{f=0\} the right hand side of 5.23 in this case vanishes, and that thus by classical regularity E\partial E is smooth in a neighbourhood of M\partial M. Since its mean curvature H=λf0{\rm H}=\lambda f\geq 0, the Maximum Principle implies that E\partial E is disjoint from M\partial M. The possibility that λ=0\lambda=0 in the case of negligible intersection ES\partial E\cap S is finally ruled out by Proposition 5.1 again, and 5.6 becomes simply 5.26. ∎

The Willmore-type inequality 1.13 and the description of ff-isoperimetric sets provided by Theorem 5.2 allow to carry out the proof of the Isoperimetric Inequality of Theorem A.

Proof of Theorem A.

If there is more than one end, then by Lemma 4.7 all ends have vanishing asymptotic volume ratio. In this case, 1.3 reduces to the just proved 5.2. Obviously, the same is true for the one-ended case if AVR(M,g,f)=0\mathrm{AVR}(M,g,f)=0. The core of the Theorem then lies in the one-ended case with AVR(M,g,f)>0\mathrm{AVR}(M,g,f)>0.

We first carry out the proof in the more involved and relevant case of nonempty boundary. Let SS be one of the outward minimizing hypersurfaces in the outward minimizing exhaustion, and let Ω\Omega be such that Ω=MS\partial\Omega=\partial M\sqcup S. For V<|Ω|fV<\lvert\Omega\rvert_{f}, consider an ff-isoperimetric set EVE_{V} constrained in Ω\Omega with ff-volume equal to VV, that is, satisfying 5.4. EVE_{V} exists and is subject to the properties described in Theorem 5.2. In particular, EV=MΣV\partial E_{V}=\partial M\sqcup\Sigma_{V}, with ΣV\Sigma_{V} a 𝒞1,1\mathscr{C}^{1,1} hypersurface. Varying VV, we also define If:(0,|Ω|f)(0,+)I_{f}:(0,\lvert\Omega\rvert_{f})\to(0,+\infty) by If(V)=|ΣV|I_{f}(V)=\lvert\Sigma_{V}\rvert, the ff-isoperimetric profile of Ω\Omega. It is argued as in the classical case that IfI_{f} is continuous. Indeed, EV+εE_{V+\varepsilon} converges in L1L^{1} to some E~\tilde{E}, that in particular satisfies |E~|f=V\lvert\tilde{E}\rvert_{f}=V. We have that, for a fixed δ>0\delta>0, by lower semicontinuity, for any ε\varepsilon close enough to zero such that

If(V)P(E~)P(EV+ε)+δ=If(V+ε)+δP(EV)+P(Bε)+δ=If(V)+P(Bε)+δ.I_{f}(V)\leq P(\tilde{E})\leq P(E_{V+\varepsilon})+\delta=I_{f}(V+\varepsilon)+\delta\leq P(E_{V})+P(B_{\varepsilon})+\delta=I_{f}(V)+P(B_{\varepsilon})+\delta. (5.27)

In the above inequality, BεB_{\varepsilon} is chosen so that |FBε|f=V+ε\lvert F\cup B_{\varepsilon}\rvert_{f}=V+\varepsilon or |FBε|f=V+ε\lvert F\setminus B_{\varepsilon}\rvert_{f}=V+\varepsilon, according to the sign of ε\varepsilon. Letting first ε0\varepsilon\to 0, and then δ0+\delta\to 0^{+} establishes the continuity of IfI_{f}. Let now ε>0\varepsilon>0. Let Σε\Sigma^{\varepsilon} be an inward variation of ΣV\Sigma_{V} supported in ΣVS\Sigma_{V}\setminus S such that |EVε|f=Vε\lvert E_{V}^{\varepsilon}\rvert_{f}=V-\varepsilon, where EVεE_{V}^{\varepsilon} is such that EVε=ΣεM\partial E_{V}^{\varepsilon}=\partial\Sigma^{\varepsilon}\sqcup\partial M. We have

lim infε0+Ifnn1(V)Ifnn1(Vε)εlim infε0+|Σ|nn1|Σε|nn1ε\liminf_{\varepsilon\to 0^{+}}\frac{I_{f}^{\frac{n}{n-1}}(V)-I_{f}^{\frac{n}{n-1}}(V-\varepsilon)}{\varepsilon}\geq\liminf_{\varepsilon\to 0^{+}}\frac{\lvert\Sigma\rvert^{\frac{n}{n-1}}-\lvert\Sigma^{\varepsilon}\rvert^{\frac{n}{n-1}}}{\varepsilon} (5.28)

Assume now that Σε\Sigma^{\varepsilon} is obtained through a normal variation field coinciding with φν\varphi\nu on Σ\Sigma, with φ𝒞c(Σ)\varphi\in\mathscr{C}^{\infty}_{c}(\Sigma). Since the first variation of ff-volume is given by the ff-weighted area, the right hand side is computed as

lim infε0+|Σ|nn1|Σε|nn1ε=nn1|ΣV|1n1ΣHφ𝑑σΣfφ𝑑σ=nn1|ΣV|1n1λ,\liminf_{\varepsilon\to 0^{+}}\frac{\lvert\Sigma\rvert^{\frac{n}{n-1}}-\lvert\Sigma^{\varepsilon}\rvert^{\frac{n}{n-1}}}{\varepsilon}=\frac{n}{n-1}\lvert\Sigma_{V}\rvert^{\frac{1}{n-1}}\frac{\int_{\Sigma}{\rm H}\varphi d\sigma}{\int_{\Sigma}f\varphi d\sigma}=\frac{n}{n-1}\lvert\Sigma_{V}\rvert^{\frac{1}{n-1}}\lambda, (5.29)

where H=λf{\rm H}=\lambda f on the support of φ\varphi is due to Theorem 5.2. Letting now WW be the infimum of ΣH/f𝑑σ\int_{\Sigma}{\rm H}/fd\sigma taken among strictly mean-convex smooth hypersurfaces Σ\Sigma homologous to M\partial M, we actually have, by the Substatic Willmore-type inequality 1.13

AVR(M,g,f)|𝕊n1|W1(n1)n1ΣV(Hf)n1𝑑σ1(n1)n1|ΣV|λn1,{\rm AVR}(M,g,f)\lvert\mathbb{S}^{n-1}\rvert\leq W\leq\frac{1}{(n-1)^{n-1}}\int_{\Sigma_{V}}\left(\frac{{\rm H}}{f}\right)^{n-1}d\sigma\leq\frac{1}{(n-1)^{n-1}}\lvert\Sigma_{V}\rvert\lambda^{n-1}, (5.30)

since 0<H/fλ0<{\rm H}/f\leq\lambda by Theorem 5.2 for (n1)(n-1)-almost any point on ΣV\Sigma_{V}. The above inequality holds for the 𝒞1,1\mathscr{C}^{1,1} ΣV\Sigma_{V} since we can approximate it with smooth, strictly mean-convex hypersurfaces through Mean Curvature Flow, see [HI01, Lemma 5.6]. Observe that this is possible since, by Theorem 5.2, ΣV\Sigma_{V} is disjoint from M\partial M. Combining 5.28, 5.29 and 5.30 yields

lim infε0+Ifnn1(V)Ifnn1(Vε)εn[AVR(M,g,f)|𝕊n1|]1n1.\liminf_{\varepsilon\to 0^{+}}\frac{I_{f}^{\frac{n}{n-1}}(V)-I_{f}^{\frac{n}{n-1}}(V-\varepsilon)}{\varepsilon}\geq n\left[{\rm AVR}(M,g,f)\lvert\mathbb{S}^{n-1}\rvert\right]^{\frac{1}{n-1}}. (5.31)

Comparing with the reference warped product ff-isoperimetric profile given by

Jf(V)=nn1n[AVR(M,g,f)|𝕊n1|]1nVn1n,J_{f}(V)=n^{\frac{n-1}{n}}\left[{\rm AVR}(M,g,f)\lvert\mathbb{S}^{n-1}\rvert\right]^{\frac{1}{n}}V^{\frac{n-1}{n}}, (5.32)

whose derivative equals the right-hand side of 5.31, we deduce at once that the continuous function Ifn/(n1)Jfn/(n1)I_{f}^{{n}/{(n-1)}}-J_{f}^{{n}/{(n-1)}} has nonnegative Dini derivative, and is thus monotone nondecreasing. Hence, for V0<VV_{0}<V, we get

Ifnn1(V)n[AVR(M,g,f)|𝕊n1|]1n1V+Ifnn1(V0)n[AVR(M,g,f)|𝕊n1|]1n1V0.I_{f}^{\frac{n}{n-1}}(V)\geq n\left[{\rm AVR}(M,g,f)\lvert\mathbb{S}^{n-1}\rvert\right]^{\frac{1}{n-1}}V+I_{f}^{\frac{n}{n-1}}(V_{0})-n\left[{\rm AVR}(M,g,f)\lvert\mathbb{S}^{n-1}\rvert\right]^{\frac{1}{n-1}}V_{0}. (5.33)

Recall now that If(V0)=|ΣV0|I_{f}(V_{0})=\lvert\Sigma_{V_{0}}\rvert for some ΣV0\Sigma_{V_{0}} homologous to M\partial M. The boundary being minimizing, by Proposition 5.1, implies then that If(V0)|M|I_{f}(V_{0})\geq\lvert\partial M\rvert. Plugging it into 5.33, and then letting V0V_{0} go to 0, leaves us with

Ifnn1(V)|M|nn1n[AVR(M,g,f)|𝕊n1|]1n1V.I_{f}^{\frac{n}{n-1}}(V)-\lvert\partial M\rvert^{\frac{n}{n-1}}\geq n\left[{\rm AVR}(M,g,f)\lvert\mathbb{S}^{n-1}\rvert\right]^{\frac{1}{n-1}}V. (5.34)

By definition of Isoperimetric profile, the above inequality implies 1.3 for any hypersurface homologous to M\partial M inside Ω\Omega, enclosing a set of volume VV. But the volumes being arbitrary and the outward minimizing envelopes forming an exhaustion, the proof of 1.3 is actually complete.

We are left to characterize the situation when some smooth Σ\Sigma homologous to M\partial M fulfils the equality in 1.3. Let VV be the ff-volume subtended by Σ\Sigma. Let Ω=MS\Omega=\partial M\sqcup S, with SS strictly mean-convex outward minimizing and ΣΩ\Sigma\subset\Omega. As above, let IfI_{f} be the ff-isoperimetric profile of Ω\Omega and JfJ_{f} the reference warped product ff-isoperimetric profile defined in 5.32. By approximation, we observe that any other ff-isoperimetric set constrained in Ω\Omega of ff-volume V0V_{0} satisfies the Isoperimetric inequality 1.3. Hence, by 5.33, we have

|M|nn1=Ifnn1(V)Jfnn1(V)Ifnn1(V0)Jfnn1(V0)|M|nn1.\lvert\partial M\rvert^{\frac{n}{n-1}}=I_{f}^{\frac{n}{n-1}}(V)-J_{f}^{\frac{n}{n-1}}(V)\geq I_{f}^{\frac{n}{n-1}}(V_{0})-J_{f}^{\frac{n}{n-1}}(V_{0})\geq\lvert\partial M\rvert^{\frac{n}{n-1}}. (5.35)

As a consequence, any ff-isoperimetric set of volume V0VV_{0}\leq V satisfies the equality in the ff-Isoperimetric inequality 1.3. Observe now, that by approximation with smooth sets in the possibly extended Riemannian manifold (N,gN)(N,g_{N}), this implies that such constrained ff-Isoperimetric sets of volume V0V_{0} are in fact globally ff-Isoperimetric, and consequently the regularity observed in Theorem 5.2 implies that any ΣV0\Sigma_{V_{0}} is smooth. Retracing the steps that lead to 5.33, we have that the smooth hypersurface ΣV0\Sigma_{V_{0}} satisfies the equality in the Willmore-type inequality in Theorem D. This triggers the rigidity stated there, and yields, for ΩV0\Omega_{V_{0}} the domain enclosed between ΣV0\Sigma_{V_{0}} and M\partial M, the isometry between (MΩV0,g)(M\setminus\Omega_{V_{0}},g) and [s0,+)×Σ[{s}_{0},+\infty)\times\Sigma endowed with

g=f2dρdρ+η2gΣV0.g=f^{2}d\rho\otimes d\rho+\eta^{2}g_{\Sigma_{V_{0}}}. (5.36)

In particular, since MM has one end, the hypersurface ΣV0\Sigma_{V_{0}} is necessarily connected. We now observe that, again due to the global ff-isoperimetry of ΣV0\Sigma_{V_{0}}, the value of H/f{\rm H}/f is constant on such hypersurface. But then, retracing the computations that lead to the isometry with 5.36, more precisely coupling 2.41 with 4.15, we deduce that ff and η\eta in 5.36 depend only on ρ\rho. Introduce now a new coordinate ss defined by f2(ρ)dρ=dsf^{2}(\rho)d\rho=ds. Recall that η\eta satisfies ρη=f2\partial_{\rho}\eta=f^{2}, and thus sη=1\partial_{s}\eta=1. Possibly translating the variable ss, we thus have

g=dsdsf2(s)+s2gΣV0,g=\frac{ds\otimes ds}{f^{2}(s)}+s^{2}g_{\Sigma_{V_{0}}}, (5.37)

for any V0VV_{0}\leq V, and ss0s\geq s_{0} for some s0>0s_{0}>0.

Since we have proved that ff is a function of the distance ρ\rho from ΣV0\Sigma_{V_{0}} only, in particular we have shown that ff must be constant on ΣV0\Sigma_{V_{0}}. This must hold for all V0VV_{0}\leq V. Moreover, the (Hausdorff) distance between ΣV0\Sigma_{V_{0}} and M\partial M goes to 0 as V00{V_{0}}\to 0, because otherwise the volume enclosed along the (sub)sequence would be necessarily bounded away from 0. But then, the level sets of ff forming a regular foliation of a neighbourhood of M\partial M, we deduce that ΣV0\Sigma_{V_{0}} must actually be a level set of ff, in particular diffeomorphic to M\partial M, for V0V_{0} small enough. Letting V0V_{0} to zero we thus extend the expression 5.37 to the whole manifold, that is 1.4. The connectednedness of M\partial M is again a consequence of (M,g)(M,g) being one ended.

As far as the characterization of Σ\Sigma is concerned, we already showed that ff is constant on it. If, by contradiction, ss were not constant on Σ\Sigma, then, letting smin=min{s(p):pΣ}s_{\mathrm{min}}=\min\{s(p):p\in\Sigma\} and smax=max{s(p):pΣ}s_{\mathrm{max}}=\max\{s(p):p\in\Sigma\}, Σ\Sigma would lie in the region [smin,smax]×M[s_{\mathrm{min}},s_{\mathrm{max}}]\times\partial M, where, ff being constant, by 1.4 has the metric of a truncated cone. By 5.37 for V0=VV_{0}=V, Σ\Sigma is a totally umbilical constantly mean-curved hypersurface in such cone. Moreover, the constancy of ff on such region makes the substatic condition simplify to nonnegative Ricci curvature. By [MR02, Lemma 3.8], Σ\Sigma could then only be a level set of ss or bound a flat round ball. The first possibility gives a contradiction with the initial assumption that ss were not constant on Σ\Sigma, the second one with Σ\Sigma being homologous to M\partial M. This concludes the proof of Σ\Sigma being a level set of ss, and of Theorem A in the nonempty-boundary case.

We finally discuss the empty-boundary-case. It is immediately checked that the ff-Isoperimetric inequality 1.3 follows with a pure simplification of the proof given above. When a hypersurface Σ\Sigma satisfies with equality the ff-Isoperimetric inequality, arguing as done above for 5.37 we reach for an isometry between (MΩΣ,g)(M\setminus\Omega_{\Sigma},g) and I=[s¯,+)×ΣI=[\overline{s},+\infty)\times\Sigma endowed with

g=dsdsf2(s)+s2gΣ,g=\frac{ds\otimes ds}{f^{2}(s)}+s^{2}g_{\Sigma}, (5.38)

for ΩΣ\Omega_{\Sigma} enclosed by Σ\Sigma. Again, Σ\Sigma must be connected, since MM is one ended by Lemma 4.7. Now, we claim that Σ\Sigma satisfies

n1nΣfH𝑑σ=ΩΣf𝑑μ,\frac{n-1}{n}\int_{\Sigma}\frac{f}{{\rm H}}d\sigma=\int_{\Omega_{\Sigma}}fd\mu, (5.39)

in fact saturating the substatic Heintze-Karcher inequality [LX19, Theorem 1.3] (see also [FP22, Theorem 3.6]) in boundaryless substatic manifolds. The analysis of the equality case worked out in [BFP23, Theorem 3.1-(ii)(ii)] then provides us with an isometry between (ΩΣ,g)(\Omega_{\Sigma},g) and I×𝕊n1I\times\mathbb{S}^{n-1} endowed with

g=dsdsf2(s)+(sf(x))2g𝕊n1,g=\frac{ds\otimes ds}{f^{2}(s)}+\left(\frac{s}{f(x)}\right)^{2}g_{\mathbb{S}^{n-1}}, (5.40)

for xΩΣx\in\Omega_{\Sigma} with Σ\Sigma becoming a level set of ss. Coupled with 5.38 on the complement of Ω\Omega, this yields the desired rigidity statement.

In order to check 5.39, just observe that, since as above one has that Σ\Sigma is ff-isoperimetric, H/f{\rm H}/f is constant on such hypersurface. Moreover, since it satisfies equality in the Willmore-type inequality 1.13, one has

Hf|Σ|1n1=(n1)[AVR(M,g,f)|𝕊n1|]1n1.\frac{{\rm H}}{f}\,\lvert\Sigma\rvert^{\frac{1}{n-1}}=(n-1)\left[\mathrm{AVR}(M,g,f)\lvert\mathbb{S}^{n-1}\rvert\right]^{\frac{1}{n-1}}. (5.41)

Coupling with

|Σ|nn1=n[AVR(M,g,f)|𝕊n1|]1n1|ΩΣ|f,\lvert\Sigma\rvert^{\frac{n}{n-1}}=n\left[\mathrm{AVR}(M,g,f)\lvert\mathbb{S}^{n-1}\rvert\right]^{\frac{1}{n-1}}\lvert\Omega_{\Sigma}\rvert_{f}, (5.42)

it is straightforwardly seen that 5.39 holds, completing the proof. ∎

Appendix A Comments on the substatic condition

A.1. Physical motivation

Here we give a physical interpretation of substatic triples, following [WWZ17, Lemma 3.8]. Let

L=×M,𝔤=f2dtdt+gL=\mathbb{R}\times M\,,\qquad\mathfrak{g}\,=\,-f^{2}dt\otimes dt+g

be a static spacetime satisfying the Einstein Field Equation

Ric𝔤+(Λ12R𝔤)𝔤=T,{\rm Ric}_{\mathfrak{g}}\,+\,\left(\Lambda-\frac{1}{2}{\mathrm{R}}_{\mathfrak{g}}\right)\mathfrak{g}\,=\,T\,,

where TT is the stress-energy tensor and Λ\Lambda\in\mathbb{R} is the cosmological constant. Using standard formulas to express the Ricci tensor of a warped product, we find out that

Ric𝔤(t,t)=fΔf,Ric𝔤(i,j)=Ric(i,j)1f2f(i,j).{\rm Ric}_{\mathfrak{g}}(\partial_{t},\partial_{t})\,=\,f\Delta f\,,\qquad{\rm Ric}_{\mathfrak{g}}(\partial_{i},\partial_{j})\,=\,{\rm Ric}(\partial_{i},\partial_{j})-\frac{1}{f}\nabla^{2}f(\partial_{i},\partial_{j})\,.

In particular a simple computation gives

R𝔤=R2fΔf.{\mathrm{R}}_{\mathfrak{g}}\,=\,{\mathrm{R}}-\frac{2}{f}\Delta f\,.

Putting these pieces of information inside the Einstein Field Equation, we get

Ttt\displaystyle T_{tt}\, =(Λ+R2)f2,\displaystyle=\,\left(-\Lambda+\frac{{\mathrm{R}}}{2}\right)f^{2}\,, (A.1)
Tit\displaystyle T_{it}\, = 0,\displaystyle=\,0\,, (A.2)
Tij\displaystyle T_{ij}\, =Rij1fij2f+(ΛR2+Δff)gij.\displaystyle=\,{\mathrm{R}}_{ij}-\frac{1}{f}\nabla^{2}_{ij}f+\left(\Lambda-\frac{{\mathrm{R}}}{2}+\frac{\Delta f}{f}\right)g_{ij}\,. (A.3)

We now assume that the Null Energy Condition is satisfied. Namely, for any vector X=t+YiiX=\partial_{t}+Y^{i}\partial_{i} with 𝔤(X,X)=0\mathfrak{g}(X,X)=0 (that is, g(Y,Y)=f2g(Y,Y)=f^{2}), we require T(X,X)=Ttt+TijYiYj0T(X,X)=T_{tt}+T_{ij}Y^{i}Y^{j}\geq 0. Using the above identities, this hypothesis tells us

0Ttt+TijYiYj=(Λ+R2)f2+(Ric1f2f+Δffg)(Y,Y)+(ΛR2)g(Y,Y).0\,\leq\,T_{tt}+T_{ij}Y^{i}Y^{j}\,=\,\left(-\Lambda+\frac{{\mathrm{R}}}{2}\right)f^{2}+\left({\rm Ric}-\frac{1}{f}\nabla^{2}f+\frac{\Delta f}{f}g\right)(Y,Y)+\left(\Lambda-\frac{{\mathrm{R}}}{2}\right)g(Y,Y)\,.

Recalling that g(Y,Y)=f2g(Y,Y)=f^{2}, we have obtained

 NEC holds(Ric1f2f+Δffg)(Y,Y)0 for all Y with g(Y,Y)=f2.\hbox{ NEC holds}\ \ \Leftrightarrow\ \ \left({\rm Ric}-\frac{1}{f}\nabla^{2}f+\frac{\Delta f}{f}g\right)(Y,Y)\geq 0\ \hbox{ for all $Y$ with $g(Y,Y)=f^{2}$.}

By rescaling of YY, we then conclude that the Null Energy Condition on static spacetimes is equivalent to

Ric1f2f+Δffg 0.{\rm Ric}-\frac{1}{f}\nabla^{2}f+\frac{\Delta f}{f}g\,\geq\,0\,.

In other words, a static spacetime satisfies the Null Energy Condition if and only if its spacelike slices are substatic.

Finally, we briefly discuss the physical interpretation of the conformal metric g~=g/f2\tilde{g}=g/f^{2}. In the context of static spacetimes, this metric is usually referred to as optical metric and has the property that g~\tilde{g}-geodesics lift to null geodesics in the spacetime metric 𝔤\mathfrak{g}. This follows easily from the fact that the trajectories of null geodesics do not change under a conformal change of metric, hence the null geodesics of 𝔤\mathfrak{g} are the same as the null geodesics of f2𝔤=dtdt+g~f^{2}\mathfrak{g}=-dt\otimes dt+\tilde{g}.

A.2. Relation between CD(0,1){\rm CD}(0,1) and substatic condition

Let (M,g,f)(M,g,f) be a substatic triple and let g~=g/f2\tilde{g}=g/f^{2}. We want to show that (M,g~,ψ)(M,\tilde{g},\psi) satisfies the CD(0,1){\rm CD}(0,1) condition, where ψ=(n1)logf\psi=-(n-1)\log f. To this end, we need to rewrite the substatic condition in terms of the conformal metric. We start from the following formulas:

~2f\displaystyle\widetilde{\nabla}^{2}f\, =2f+1f(2dfdf|f|2g),\displaystyle=\,\nabla^{2}f+\frac{1}{f}\left(2df\otimes df-|\nabla f|^{2}g\right)\,, (A.4)
Δg~f\displaystyle\Delta_{\tilde{g}}f\, =f2Δf(n2)f|f|2,\displaystyle=\,f^{2}\Delta f-(n-2)f|\nabla f|^{2}\,, (A.5)

In particular

1f2f1fΔfg\displaystyle\frac{1}{f}\nabla^{2}f-\frac{1}{f}\Delta fg\, =1f~2f1f2(2dfdf|f|2g)1fΔg~fg~(n2)|f|2g~\displaystyle=\,\frac{1}{f}\widetilde{\nabla}^{2}f-\frac{1}{f^{2}}\left(2df\otimes df-|\nabla f|^{2}g\right)-\frac{1}{f}\Delta_{\tilde{g}}f\,\tilde{g}-(n-2)|\nabla f|^{2}\tilde{g} (A.6)
=1f~2f2f2dfdf1fΔg~fg~(n3)1f2|~f|g~2g~\displaystyle=\,\frac{1}{f}\widetilde{\nabla}^{2}f-\frac{2}{f^{2}}df\otimes df-\frac{1}{f}\Delta_{\tilde{g}}f\tilde{g}-(n-3)\frac{1}{f^{2}}\big{|}\widetilde{\nabla}f\big{|}_{\tilde{g}}^{2}\,\tilde{g} (A.7)

On the other hand, it is well known that the Ricci tensor Ric{\rm Ric} of gg and the Ricci tensor Ricg~{\rm Ric}_{\tilde{g}} of the conformal metric g~\tilde{g} are related as follows

Ric=Ricg~n2f~2f+2(n2)f2dfdf(1fΔg~f+n3f2|~f|g~2)g~.{\rm Ric}\,=\,{\rm Ric}_{\tilde{g}}-\frac{n-2}{f}\widetilde{\nabla}^{2}f+\frac{2(n-2)}{f^{2}}df\otimes df-\left(\frac{1}{f}\Delta_{\tilde{g}}f+\frac{n-3}{f^{2}}|\widetilde{\nabla}f|^{2}_{\tilde{g}}\right)\tilde{g}\,.

Putting together the above formulas, we get

Ric1f2f+1fΔfg\displaystyle{\rm Ric}-\frac{1}{f}\nabla^{2}f+\frac{1}{f}\Delta fg\, =Ricg~n1f~2f+2(n1)f2dfdf\displaystyle=\,{\rm Ric}_{\tilde{g}}-\frac{n-1}{f}\widetilde{\nabla}^{2}f+\frac{2(n-1)}{f^{2}}df\otimes df (A.8)
=Ricg~(n1)~2(logf)+(n1)dlogfdlogf\displaystyle=\,{\rm Ric}_{\tilde{g}}-(n-1)\widetilde{\nabla}^{2}(\log f)+(n-1)d\log f\otimes d\log f (A.9)
=Ricg~+~2ψ+1n1dψdψ.\displaystyle=\,{\rm Ric}_{\tilde{g}}+\widetilde{\nabla}^{2}\psi+\frac{1}{n-1}d\psi\otimes d\psi\,. (A.10)

It follows then that, if (M,g,f)(M,g,f) satisfies the substatic condition, then (M,g~,ψ)(M,\tilde{g},\psi) satisfies the CD(0,1){\rm CD}(0,1) condition.

A.3. Li–Xia connections

In [LX17], Li and Xia consider the family of connections Duαγ{\rm D}^{u\alpha\gamma}, where u𝒞(M)u\in\mathscr{C}^{\infty}(M), α,γ\alpha,\gamma\in\mathbb{R}, defined by

DXuαγY=XY+α[X(u)Y+Y(u)X]+γg(X,Y)u.{\rm D}^{u\alpha\gamma}_{X}Y=\nabla_{X}Y+\alpha\left[X(u)Y+Y(u)X\right]+\gamma\,g(X,Y)\nabla u\,.

They then compute the Ricci tensor Ricuαγ{\rm Ric}^{u\alpha\gamma} induced by a connection Duαγ{\rm D}^{u\alpha\gamma}, showing that it is related to the usual Ricci tensor by

Ricuαγ=Ric[(n1)α+γ]2u+[(n1)α2γ2]dudu+[γΔu+(γ2+(n1)αγ)|u|2]g.{\rm Ric}^{u\alpha\gamma}={\rm Ric}-\left[(n-1)\alpha+\gamma\right]\nabla^{2}u+\left[(n-1)\alpha^{2}-\gamma^{2}\right]du\otimes du+\left[\gamma\Delta u+\left(\gamma^{2}+(n-1)\alpha\gamma\right)|\nabla u|^{2}\right]g\,.

When α=0\alpha=0 and γ=1\gamma=1, in particular we have

Ricu01=Ric2ududu+[Δu+|u|2]g,{\rm Ric}^{u01}\,=\,{\rm Ric}-\nabla^{2}u-du\otimes du+\left[\Delta u+|\nabla u|^{2}\right]g\,,

which can be rewritten as follows by setting u=logfu=\log f:

Ricu01=Ric1f2f+Δffg.{\rm Ric}^{u01}\,=\,{\rm Ric}-\frac{1}{f}\nabla^{2}f+\frac{\Delta f}{f}g\,.

It is then clear that the condition Ricu010{\rm Ric}^{u01}\geq 0 is equivalent to the substatic condition.

Choosing instead α=1/(n1)\alpha=1/(n-1), γ=0\gamma=0, setting v=uv=-u one gets

Ricu1n10=Ric+2v+1n1dvdv,{\rm Ric}^{u\frac{1}{n-1}0}={\rm Ric}+\nabla^{2}v+\frac{1}{n-1}dv\otimes dv\,,

hence Ricu1n100{\rm Ric}^{u\frac{1}{n-1}0}\geq 0 gives the CD(0,1){\rm CD}(0,1) condition. In fact, the connection Du1n10{\rm D}^{u\frac{1}{n-1}0} had already been considered in the work [WY16] that was focused on the CD(0,1){\rm CD}(0,1) case only.

Here we show that the two connections Du01{\rm D}^{u01} and Du1n10{\rm D}^{u\frac{1}{n-1}0} are in fact conformally related: let (M,g)(M,g) be a Riemannian manifold and let ,~\nabla,\widetilde{\nabla} be the Levi-Civita connections corresponding to the metrics gg, g~=g/f2\tilde{g}=g/f^{2}, respectively. It is easy to show that \nabla and ~\widetilde{\nabla} are related a follows

XY=~XY+1f[X(f)Y+Y(f)Xg(X,Y)f]=~XY+X(u)Y+Y(u)Xg(X,Y)u,\nabla_{X}Y\,=\,\widetilde{\nabla}_{X}Y+\frac{1}{f}\left[X(f)Y+Y(f)X-g(X,Y)\nabla f\right]\,=\,\widetilde{\nabla}_{X}Y+X(u)Y+Y(u)X-g(X,Y)\nabla u\,,

hence, setting ψ=(n1)u\psi=-(n-1)u:

DXu01Y\displaystyle{\rm D}_{X}^{u01}Y\, =XY+g(X,Y)u\displaystyle=\,\nabla_{X}Y+g(X,Y)\nabla u (A.11)
=~XY+X(u)Y+Y(u)X\displaystyle=\,\widetilde{\nabla}_{X}Y+X(u)Y+Y(u)X (A.12)
=~XY+1n1[X(ψ)Y+Y(ψ)X].\displaystyle=\,\widetilde{\nabla}_{X}Y+\frac{1}{n-1}\left[X(-\psi)Y+Y(-\psi)X\right]\,. (A.13)

This is then precisely Dψ1n10{\rm D}^{-\psi\frac{1}{n-1}0} using ~\widetilde{\nabla} as the Levi-Civita connection in place of \nabla.

Notice that in this subsection ψ\psi and ff have been introduced as the functions satisfying u=logfu=\log f and ψ=(n1)u\psi=-(n-1)u, hence they are related by ψ=(n1)logf\psi=-(n-1)\log f, in agreement with Subsection A.2.

References

  • [AFM20] Virginia Agostiniani, Mattia Fogagnolo and Lorenzo Mazzieri “Sharp Geometric Inequalities for Closed Hypersurfaces in Manifolds with Nonnegative Ricci Curvature” In Inventiones mathematicae, 2020 DOI: 10.1007/s00222-020-00985-4
  • [AFP22] Gioacchino Antonelli, Mattia Fogagnolo and Marco Pozzetta “The isoperimetric problem on Riemannian manifolds via Gromov–Hausdorff asymptotic analysis” In Communications in Contemporary Mathematics, 2022 DOI: 10.1142/S0219199722500687
  • [ANP22] Gioacchino Antonelli, Stefano Nardulli and Marco Pozzetta “The isoperimetric problem via direct method in noncompact metric measure spaces with lower Ricci bounds” In ESAIM Control Optim. Calc. Var. 28, 2022, pp. Paper No. 57\bibrangessep32 DOI: 10.1051/cocv/2022052
  • [Ant+22] Gioacchino Antonelli, Enrico Pasqualetto, Marco Pozzetta and Daniele Semola “Asymptotic isoperimetry on non collapsed spaces with lower Ricci bounds” Accepted for publication in Math. Ann., 2022 arXiv:2208.03739 [math.DG]
  • [BFM22] Luca Benatti, Mattia Fogagnolo and Lorenzo Mazzieri “Minkowski Inequality on complete Riemannian manifolds with nonnegative Ricci curvature”, 2022 arXiv:2101.06063 [math.DG]
  • [BFP23] Stefano Borghini, Mattia Fogagnolo and Andrea Pinamonti “The equality case in the substatic Heintze-Karcher inequality”, 2023 arXiv:2307.04253 [math.DG]
  • [BK22] Zoltán M. Balogh and Alexandru Kristály “Sharp isoperimetric and Sobolev inequalities in spaces with nonnegative Ricci curvature” In Mathematische Annalen, 2022 DOI: 10.1007/s00208-022-02380-1
  • [BM02] Hubert Bray and Frank Morgan “An isoperimetric comparison theorem for Schwarzschild space and other manifolds” In Proc. Amer. Math. Soc. 130.5, 2002, pp. 1467–1472 DOI: 10.1090/S0002-9939-01-06186-X
  • [Bra97] Hubert Lewis Bray “The Penrose inequality in general relativity and volume comparison theorems involving scalar curvature” Thesis (Ph.D.)–Stanford University ProQuest LLC, Ann Arbor, MI, 1997, pp. 103 URL: http://gateway.proquest.com/openurl?url_ver=Z39.88-2004&rft_val_fmt=info:ofi/fmt:kev:mtx:dissertation&res_dat=xri:pqdiss&rft_dat=xri:pqdiss:9810085
  • [Bre13] Simon Brendle “Constant mean curvature surfaces in warped product manifolds” In Publ. Math. Inst. Hautes Études Sci. 117, 2013, pp. 247–269 DOI: 10.1007/s10240-012-0047-5
  • [Bre21] Simon Brendle “Sobolev Inequalities in Manifolds with Nonnegative Curvature” To appear in Comm. Pure App. Math, 2021
  • [Car92] Manfredo Perdigão Carmo “Riemannian geometry” Translated from the second Portuguese edition by Francis Flaherty, Mathematics: Theory & Applications Birkhäuser Boston, Inc., Boston, MA, 1992, pp. xiv+300 DOI: 10.1007/978-1-4757-2201-7
  • [CG72] Jeff Cheeger and Detlef Gromoll “The splitting theorem for manifolds of nonnegative Ricci curvature” In J. Differential Geometry 6, 1971/72, pp. 119–128 URL: http://projecteuclid.org/euclid.jdg/1214430220
  • [Che14] Xu Cheng “An almost-Schur type lemma for symmetric (2,0)(2,0) tensors and applications” In Pacific J. Math. 267.2, 2014, pp. 325–340 DOI: 10.2140/pjm.2014.267.325
  • [CM22] Fabio Cavalletti and Davide Manini “Isoperimetric inequality in noncompact MCP{MCP} spaces” In Proc. Amer. Math. Soc. 150.8, 2022, pp. 3537–3548 DOI: 10.1090/proc/15945
  • [CM22a] Fabio Cavalletti and Davide Manini “Rigidities of Isoperimetric inequality under nonnegative Ricci curvature”, 2022 arXiv:2207.03423 [math.MG]
  • [CS01] Piotr T. Chruściel and Walter Simon “Towards the classification of static vacuum spacetimes with negative cosmological constant” In J. Math. Phys. 42.4, 2001, pp. 1779–1817 DOI: 10.1063/1.1340869
  • [Eva10] Lawrence C. Evans “Partial differential equations” 19, Graduate Studies in Mathematics American Mathematical Society, Providence, RI, 2010, pp. xxii+749 DOI: 10.1090/gsm/019
  • [FGS17] M. Focardi, F. Geraci and E. Spadaro “The classical obstacle problem for nonlinear variational energies” In Nonlinear Anal. 154, 2017, pp. 71–87 DOI: 10.1016/j.na.2016.10.020
  • [FM22] Mattia Fogagnolo and Lorenzo Mazzieri “Minimising hulls, pp-capacity and isoperimetric inequality on complete Riemannian manifolds” In J. Funct. Anal. 283.9, 2022, pp. Paper No. 109638\bibrangessep49 DOI: 10.1016/j.jfa.2022.109638
  • [FP22] Mattia Fogagnolo and Andrea Pinamonti “New integral estimates in substatic Riemannian manifolds and the Alexandrov theorem” In J. Math. Pures Appl. (9) 163, 2022, pp. 299–317 DOI: 10.1016/j.matpur.2022.05.007
  • [Fuj22] Yasuaki Fujitani “Some functional inequalities under lower Bakry-Émery-Ricci curvature bounds with ε\varepsilon-range”, 2022 arXiv:2211.12310 [math.DG]
  • [FY19] Yi Fang and Wei Yuan “Brown–York mass and positive scalar curvature II: Besse’s conjecture and related problems” In Annals of Global Analysis and Geometry 56 Springer, 2019, pp. 1–15
  • [Gal+99] G.. Galloway, K. Schleich, D.. Witt and E. Woolgar “Topological censorship and higher genus black holes” In Physical Review D 60.10 APS, 1999, pp. 104039
  • [He21] Huiya He “Critical metrics of the volume functional on three-dimensional manifolds” In arXiv preprint arXiv:2101.05621, 2021
  • [HI01] Gerhard Huisken and Tom Ilmanen “The Inverse Mean Curvature Flow and the Riemannian Penrose Inequality” In Journal of Differential Geometry 59.3, 2001, pp. 353–437 URL: https://mathscinet.ams.org/mathscinet-getitem?mr=1916951
  • [HW23] Brian Harvie and Ye-Kai Wang “A rigidity theorem for asymptotically flat static manifolds and its applications”, 2023 arXiv:2305.08570 [math.DG]
  • [Ilm92] Tom Ilmanen “Generalized flow of sets by mean curvature on a manifold” In Indiana Univ. Math. J. 41.3, 1992, pp. 671–705 DOI: 10.1512/iumj.1992.41.41036
  • [Joh21] Florian Johne “Sobolev inequalities on manifolds with nonnegative Bakry-Émery Ricci curvature”, 2021 arXiv:2103.08496 [math.DG]
  • [Kas83] Atsushi Kasue “Ricci curvature, geodesics and some geometric properties of Riemannian manifolds with boundary” In J. Math. Soc. Japan 35.1, 1983, pp. 117–131 DOI: 10.2969/jmsj/03510117
  • [KL22] Kazuhiro Kuwae and Xiang-Dong Li “New Laplacian comparison theorem and its applications to diffusion processes on Riemannian manifolds” In Bull. Lond. Math. Soc. 54.2, 2022, pp. 404–427 DOI: 10.1112/blms.12568
  • [Kle92] Bruce Kleiner “An isoperimetric comparison theorem” In Invent. Math. 108.1, 1992, pp. 37–47 DOI: 10.1007/BF02100598
  • [KS23] Kazuhiro Kuwae and Yohei Sakurai “Comparison geometry of manifolds with boundary under lower NN-weighted Ricci curvature bounds with ε\varepsilon-range” In J. Math. Soc. Japan 75.1, 2023, pp. 151–172 DOI: 10.2969/jmsj/87278727
  • [KWY19] Lee Kennard, William Wylie and Dmytro Yeroshkin “The weighted connection and sectional curvature for manifolds with density” In The Journal of Geometric Analysis 29 Springer, 2019, pp. 957–1001
  • [Lee19] Dan A. Lee “Geometric relativity” 201, Graduate Studies in Mathematics American Mathematical Society, Providence, RI, 2019, pp. xii+361 DOI: 10.1090/gsm/201
  • [LMO22] Yufeng Lu, Ettore Minguzzi and Shin-ichi Ohta “Comparison theorems on weighted Finsler manifolds and spacetimes with ε\varepsilon-range” In Anal. Geom. Metr. Spaces 10.1, 2022, pp. 1–30 DOI: 10.1515/agms-2020-0131
  • [LX17] Junfang Li and Chao Xia “An integral formula for affine connections” In J. Geom. Anal. 27.3, 2017, pp. 2539–2556 DOI: 10.1007/s12220-017-9771-x
  • [LX19] Junfang Li and Chao Xia “An integral formula and its applications on sub-static manifolds” In J. Differential Geom. 113.3, 2019, pp. 493–518 DOI: 10.4310/jdg/1573786972
  • [Mag12] Francesco Maggi “Sets of finite perimeter and geometric variational problems” An introduction to geometric measure theory 135, Cambridge Studies in Advanced Mathematics Cambridge University Press, Cambridge, 2012, pp. xx+454 DOI: 10.1017/CBO9781139108133
  • [McC18] Stephen McCormick “On a Minkowski-like inequality for asymptotically flat static manifolds” In Proceedings of the American Mathematical Society 146.9, 2018, pp. 4039–4046
  • [MM03] Carlo Mantegazza and Andrea Carlo Mennucci “Hamilton-Jacobi equations and distance functions on Riemannian manifolds” In Appl. Math. Optim. 47.1, 2003, pp. 1–25 DOI: 10.1007/s00245-002-0736-4
  • [Mor03] Frank Morgan “Regularity of isoperimetric hypersurfaces in Riemannian manifolds” In Trans. Amer. Math. Soc. 355.12, 2003, pp. 5041–5052 DOI: 10.1090/S0002-9947-03-03061-7
  • [MR02] Frank Morgan and Manuel Ritoré “Isoperimetric regions in cones” In Trans. Amer. Math. Soc. 354.6, 2002, pp. 2327–2339 DOI: 10.1090/S0002-9947-02-02983-5
  • [MRS22] Luciano Mari, Marco Rigoli and Alberto G. Setti “On the 1/H1/H-flow by pp-Laplace approximation: new estimates via fake distances under Ricci lower bounds” In Amer. J. Math. 144.3, 2022, pp. 779–849 DOI: 10.1353/ajm.2022.0016
  • [MS17] Andrea Mondino and Emanuele Spadaro “On an isoperimetric-isodiametric inequality” In Anal. PDE 10.1, 2017, pp. 95–126 DOI: 10.2140/apde.2017.10.95
  • [Nar14] Stefano Nardulli “Generalized existence of isoperimetric regions in non-compact Riemannian manifolds and applications to the isoperimetric profile” In Asian J. Math. 18.1, 2014, pp. 1–28 DOI: 10.4310/AJM.2014.v18.n1.a1
  • [Oht21] Shin-ichi Ohta “Comparison Finsler geometry”, Springer Monographs in Mathematics Springer, Cham, 2021, pp. xxii+316 DOI: 10.1007/978-3-030-80650-7
  • [Pet16] Peter Petersen “Riemannian geometry” 171, Graduate Texts in Mathematics Springer, Cham, 2016, pp. xviii+499 DOI: 10.1007/978-3-319-26654-1
  • [Poz23] Marco Pozzetta “Isoperimetry on manifolds with Ricci bounded below: overview of recent results and methods”, 2023 arXiv:2303.11925 [math.DG]
  • [PV20] Stefano Pigola and Giona Veronelli “The smooth Riemannian extension problem” In Ann. Sc. Norm. Super. Pisa Cl. Sci. (5) 20.4, 2020, pp. 1507–1551
  • [Sak21] Yohei Sakurai “One dimensional weighted Ricci curvature and displacement convexity of entropies” In Math. Nachr. 294.10, 2021, pp. 1950–1967 DOI: 10.1002/mana.201900143
  • [SZW91] Peter Sternberg, William P. Ziemer and Graham Williams C1,1C^{1,1}-regularity of constrained area minimizing hypersurfaces” In J. Differential Equations 94.1, 1991, pp. 83–94 DOI: 10.1016/0022-0396(91)90104-H
  • [Wan23] Xiaodong Wang “Remark on an inequality for closed hypersurfaces in complete manifolds with nonnegative Ricci curvature” In Ann. Fac. Sci. Toulouse Math. (6) 32.1, 2023, pp. 173–178
  • [Wei18] Yong Wei “On the Minkowski-type inequality for outward minimizing hypersurfaces in Schwarzschild space” In Calculus of Variations and Partial Differential Equations 57 Springer, 2018, pp. 1–17
  • [Whi00] Brian White “The size of the singular set in mean curvature flow of mean-convex sets” In J. Amer. Math. Soc. 13.3, 2000, pp. 665–695 DOI: 10.1090/S0894-0347-00-00338-6
  • [WWZ17] Mu-Tao Wang, Ye-Kai Wang and Xiangwen Zhang “Minkowski formulae and Alexandrov theorems in spacetime” In J. Differential Geom. 105.2, 2017, pp. 249–290 DOI: 10.4310/jdg/1486522815
  • [WY16] William Wylie and Dmytro Yeroshkin “On the geometry of Riemannian manifolds with density”, 2016 arXiv:1602.08000 [math.DG]
  • [Wyl17] William Wylie “A warped product version of the Cheeger–Gromoll splitting theorem” In Transactions of the American Mathematical Society 369.9, 2017, pp. 6661–6681
  • [Zen22] Fanqi Zeng “Some almost-Schur type inequalities and applications on sub-static manifolds” In Electron. Res. Arch. 30.8, 2022, pp. 2860–2870 DOI: 10.3934/era.2022145