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

Positive mass theorem for asymptotically flat manifolds with isolated conical singularities

Xianzhe Dai Department of Mathematics, University of Californai, Santa Barbara CA93106, USA [email protected] Yukai Sun Key Laboratory of Pure and Applied Mathematics, School of Mathematical Sciences, Peking University, Beijing, 100871, P. R. China [email protected]  and  Changliang Wang School of Mathematical Sciences and Institute for Advanced Study, Tongji University, Shanghai 200092, China [email protected]
Abstract.

We prove the positive mass theorem for asymptotical flat (AF for short) manifolds with finitely many isolated conical singularities. We do not impose the spin condition. Instead we use the conformal blow up technique which dates back to Schoen’s final resolution of the Yamabe conjecture.

Key words and phrases:
Positive mass theorem, Conical singularity, Conformal change, Blow up

1. Introduction

The famous Positive Mass Theorem states that an asymptotically flat (AF for short; also called asymptotically Euclidean) smooth manifold with nonnegative scalar curvature must have nonnegative ADM mass, and the mass is zero iff the manifold is the Euclidean space. This result was first proven by R. Schoen and S.-T. Yau [25] for manifolds of dimension between 33 and 77 using minimal surface method (for the recent progress of Schoen and Yau’s argument in higher dimension, see [26]). E. Witten [28] proved the Positive Mass Theorem for spin manifolds of arbitrary dimension by using the Dirac operator.

It is a natural problem to generalize the Positive Mass Theorem to singular Riemannian manifolds (smooth manifolds endowed with non-smooth metrics), partially motived by the investigation of weak notions of nonnegative scalar curvature, and the study of stability aspect of positive mass theorem. There has been a lot of interesting works in this problem, see e.g. [7, 11, 13, 14, 17, 19, 20, 27]. These deal with metric singularities occurring on smooth manifolds, and are usually assumed to be continuous and have W1,pW^{1,p}-regularity for pnp\geq n.

It was observed in [3, Section 4.2] (see also Proposition 2.3 in [27]) that the negative mass Schwarzchild metric can be viewed as an AF metric with a r23r^{\frac{2}{3}}-horn singularity, which is scalar flat but has negative mass. A model rbr^{b}-horn metric is given as

(1.1) g¯b:=dr2+r2bgN,for someb>0,\overline{g}_{b}:=dr^{2}+r^{2b}g^{N},\ \ \text{for some}\ \ b>0,

on a product manifold (0,1)×N(0,1)\times N, where gNg^{N} is a Riemannian metric on closed manifold NN, and near a rbr^{b}-horn singularity the metric is asymptotic to the model rbr^{b}-horn metric as r0r\to 0. In particular, for b=1b=1, it is a conical singularity. These singular metrics are not continuous at the singular point, which is not even a topological manifold point if the cross section NN is not a sphere. In contrast to the negative mass Schwarzchild metric, the authors in [5] proved Positive Mass Theorem for AF spin manifolds with isolated rbr^{b}-horn singularities for b1b\geq 1, and in particular for AF spin manifolds with isolated conical singularities. In [17], Li and Mantoulidis proved Positive Mass Theorem for metrics with conical singularities along a codimension two submanifold (aka edge metrics) on smooth AF manifolds. In [12], Ju and Viaclovsky proved a positive mass theorem for AF manifolds with isolated orbifold singularities.

Conical singularities also occur naturally in the study of the near horizon geometry of black holes. Motivated from this consideration, a positive mass theorem for conical singularity under the nonnegative Ricci curvature is established in [18] (modulo some analytic details).

In this paper, we prove Positive Mass Theorem for general AF manifolds with isolated conical singularity, i.e., without imposing the spin condition.

Theorem 1.1.

Let (Mn,g)(M^{n},g) be an AF manifold with finitely many isolated conical singularity and n3n\geq 3. If the scalar curvature is nonnegative on the smooth part, then the ADM mass m(g)m(g) is nonnegative. Furthermore, the mass m(g)=0m(g)=0 if and only if (Mn,g)(M^{n},g) is isometric to (n,gn)(\mathbb{R}^{n},g_{\mathbb{R}^{n}}).

Our result also holds for multiple AF ends, as we indicate later. In 3-dimension, by estimating capacity and Willmore functional of the cross sections, as an application of mass-capacity inequalities in [22], Miao [21] obtained the nonnegativity of the mass of an AF manifold with rbr^{b}-horn singularity with b>23b>\frac{2}{3}, provided that the relative homology group H2(MBr0(o),Br0(o))=0H_{2}(M\setminus B_{r_{0}}(o),\partial B_{r_{0}}(o))=0, here oo is the singular point and Br0(o)B_{r_{0}}(o) is the geodesic ball centered at oo with radius r0r_{0}.

The precise definition of AF manifold with conical singularity is given in Section 2. These singular manifolds are not complete. This is one of main differences from the classical Positive Mass Theorem on complete AF manifolds. Some positive mass theorems on incomplete AF manifolds have been established in [15, 16], provided that the scalar curvature on the AF end is nonnegative and satisfies some quantitative positive lower bound on a bounded domain near the AF infinity.

The non-negativity of the mass in Theorem 1.1 is proved in Section 4. The basic strategy is to construct an AF manifold with boundary, and then apply the results about the positive mass theorem for manifolds with boundary, e.g. in [8, 9, 10, 22] and others. A straightforward way to construct an AF manifold with boundary is to remove a conical neighborhood of the singular point from an AF manifold with a conical singularity. This idea works well for rbr^{b}-horn singularity with b>1b>1, as we have done in [5]. However, in the case of b<1b<1, the mean curvature of the boundary with respect to the normal pointing to AF end would be too large (relative to a certain negative power of the area of the boundary) to apply positive mass theorems for manifolds with boundary. As mentioned above, the negative mass Schwarzchild metric (b=23b=\frac{2}{3}) gives a counter-example to the positive mass theorem with rbr^{b}-horn singularity for b<1b<1. In the case of conical singularity (b=1b=1), this simple construction may not give desired AF manifold with boundary either, since the required upper bound for the mean curvature seems not easy to be estimated.

Instead, in the case of conical singularity, before cutting off a part of the manifold, we first apply a conformal blow up technique to the AF metric with a conical singularity, following Ju and Viaclovsky’s approach for orbifold singularity in [12]. More precisely, we first conformally deform the AF metric with conical singularity by using a harmonic function, which has asymptotic order like Green’s function (i.e. r2nr^{2-n}) near the conical singularity, and is asymptotic to 11 near the AF infinity. One may think the harmonic function as the Green function with its pole at the conical singularity. This conformal deformation flips over a cone, converting the conical singularity to the infinite end of the cone, and keeps the AF end up to a higher order perturbation. Then we remove the conical end, and obtain an AF manifold with boundary whose mean curvature is negative, and so we can apply Hirsch and Miao’s results about positive mass theorem with boundary [10] to prove the non-negativity of the mass.

Note that by conformally blowing up the conical singularity, we obtain a Riemannian manifold with two ends, one is an AF end, and another is an asymptotically conical end, i.e. this manifold is an AF manifold with an additional complete end. Therefore, we can also apply results about the positive mass theorems on AF manifolds with arbitrary ends in [16, 29] (for dimensions between 3 and 7) and [4] (for spin manifolds) to prove the nonnegativity of the mass in Theorem 1.1.

In a related work [1], K. Akutagawa and B. Botvinnik studied the Yamabe problem for cylindrical manifolds, and obtained positive mass theorem for certain AF manifold with two specific conical singularities, which is constructed by a conformal blow up of a cylinder. The conformal blow up technique dates back to Schoen’s solution of Yamabe problem [24], and has been a very useful and powerful tool in the study of various problems related to scalar curvature. The Schwarzschild metric g=(1+mrn2)4n2gng=(1+\frac{m}{r^{n-2}})^{\frac{4}{n-2}}g_{\mathbb{R}^{n}} on n\{0}\mathbb{R}^{n}\backslash\{0\}, for m>0m>0, can be viewed as the conformal blowup of the origin by the Green’s function 1+mrn21+\frac{m}{r^{n-2}}, which is asymptotic to 11 as rr\to\infty and has the asymptotical order r2nr^{2-n} as r0r\to 0.

The key in our argument is to find a suitable harmonic function. In contrast to the case of orbifold singularity, we need to solve for a harmonic function with certain prescribed asymptotic behavior on the AF manifold with conical singularity, which may not be lifted to a smooth manifold. To deal with the conical singularity, we employ weighted Sobolev spaces introduced in [5], which have two weights, one for the conical singularity and the other for the AF infinity.

The rigidity result in Theorem 1.1 is proved in Section 5. We first prove that the mass m(g)=0m(g)=0 and Scg0{\rm Sc}_{g}\geq 0 imply the AF metric with conical singularity must be Ricci flat, by deforming the metric in a compact domain away from conical singularity, and following a conformal change. In the case of orbifold singularity, as shown in [12], by applying the Bishop-Gromov’s volume inequality for orbifolds, the Euclidean volume growth at AF infinity then implies that the metric has to be flat. It is more delicate and involved in the case of conical singularity, since the Bishop-Gromov’s volume inequality may not hold for AF manifolds with conical singularity. We need to get the rigidity for the cross section of the cone so that we can get more information about the regularity of the metric. For that, we solve harmonic coordinates on AF manifolds with conical singularity, and study its asymptotic behavior near the conical singularity, which can be linked to the first nonzero eigenvalue on the cross section. Then we could apply Obata’s rigidity theorem to obtain that the cross section must be isometric to the standard sphere. This then implies that the metric is continuous and has W1,nW^{1,n}-regularity, and so the rigidity result in [11] can be applied.

The paper is organized as follows. In Section 2, we introduce asymptotically flat (AF) manifolds with isolated conical singularities and their ADM mass and collect some basic facts on cones. Section 3 is devoted to the analysis on AF manifolds with isolated conical singularities. We review the weighted Sobolev spaces introduced in [5] and discuss the corresponding elliptic estimates for the Laplace operators, their Fredholm properties, and surjectivity. This is used in Section 4 and 5 to obtain asymptotically harmonic function with certain prescribed asymptotic behavior at the conical singularity or AF infinity. In Section 4 and 5 we prove the positive mass theorem for conical singularity (Theorem 1.1).

Acknowledgement: The authors are grateful to Jeff Viaclovsky for suggesting that something like this could work. Xianzhe Dai is partially supported by the Simons Foundation. Yukai Sun is partially funded by the National Key R&D Program of China Grant 2020YFA0712800. Changliang Wang is partially supported by the Fundamental Research Funds for the Central Universities and Shanghai Pilot Program for Basic Research.

2. AF manifolds with isolated conical singularities

In this section, we give the precise definition of what we call asymptotically flat (AF for short) manifolds with finitely many isolated conical singularities, and the definition of the mass (at infinity) for these singular manifolds.

Definition 2.1.

We say (M0n,g,d,o)(M^{n}_{0},g,d,o) is a compact Riemannian manifold with smooth boundary and a single conical singularity at oM0M0o\in M_{0}\setminus\partial M_{0}, if

  1. (i)

    dd is a metric on M0M_{0} and (M0,d)(M_{0},d) is a compact metric space with smooth boundary,

  2. (ii)

    gg is a smooth Riemannian metric on the regular part M0{o}M_{0}\setminus\{o\}, dd is the induced metric by the Riemannian metric gg on M0{o}M_{0}\setminus\{o\} ,

  3. (iii)

    there exists a neighborhood UoU_{o} of oo in MMM\setminus\partial M, such that Uo{o}(0,1)×NU_{o}\setminus\{o\}\simeq(0,1)\times N for a smooth compact manifold NN, and on Uo{o}U_{o}\setminus\{o\} the metric g=g¯+hg=\overline{g}+h, where

    g¯=dr2+r2gN,\overline{g}=dr^{2}+r^{2}g^{N},

    gNg^{N} is a smooth Riemannian metric on NN, rr is a coordinate on (0,1)(0,1), r=0r=0 corresponding the singular point oo, and hh satisfies

    (2.1) |¯kh|g¯=O(rαk),asr0,|\overline{\nabla}^{k}h|_{\overline{g}}=O(r^{\alpha-k}),\ \ \text{as}\ \ r\rightarrow 0,

    for some α>0\alpha>0 and k=0,1k=0,1 and 22, where ¯\overline{\nabla} is the Levi-Civita connection of g¯\overline{g}.

Definition 2.2.

We say (Mn,g,o)(M^{n},g,o) is an asymptotically flat manifold with a single isolated conical singularity at oo, if Mn=M0MM^{n}=M_{0}\cup M_{\infty} satisfies

  1. (i)

    (M0,g|M0{o},o)(M_{0},g|_{M_{0}\setminus\{o\}},o) is a compact Riemannian manifold with smooth boundary and a single conical singularity at oo defined as in Definition 2.1,

  2. (ii)

    MM_{\infty} is diffeomorphic to nBR(0)\mathbb{R}^{n}\setminus B_{R}(0) for some R>0R>0, and under this diffeomorphism the smooth Riemannian metric gg on MM_{\infty} satisfies

    g=gn+O(ρτ),|(gn)ig|gn=O(ρτi),asρ+,g=g_{\mathbb{R}^{n}}+O(\rho^{-\tau}),\ \ |(\nabla^{g_{\mathbb{R}^{n}}})^{i}g|_{g_{\mathbb{R}^{n}}}=O(\rho^{-\tau-i}),\ \ \text{as}\ \ \rho\rightarrow+\infty,

    for i=1,2i=1,2 and 33, where τ>n22\tau>\frac{n-2}{2} is the asymptotical order, gn\nabla^{g_{\mathbb{R}^{n}}} is the Levi-Civita connection of the Euclidean metric gng_{\mathbb{R}^{n}}, and ρ\rho is the Euclidean distance to a base point.

Remark 2.3.

For the sake of simplicity of notations, in Definition 2.2 we only defined AF manifolds with a single conical singularity and one AF end, and we will only focus on this case in this paper. AF manifolds with finitely many isolated conical singularities and finitely many AF ends can be defined similarly, and all results in this paper can be easily extended to the case of finitely many isolated conical singularities and finitely many AF ends.

Now we give the definition of the ADM mass of such Riemannian manifold.

Definition 2.4.

Let (Mn,g,o)(M^{n},g,o) be an asymptotically flat manifold with a single isolated conical singularity at oo. The mass m(g)m(g) is defined as:

m(g)=limR1ωnSR(igjijgii)𝑑xj,m(g)=\lim_{R\to\infty}\frac{1}{\omega_{n}}\int_{S_{R}}(\partial_{i}g_{ji}-\partial_{j}g_{ii})\ast dx^{j},

where {xi}\{\frac{\partial}{\partial x^{i}}\} is an orthonormal basis of gng_{\mathbb{R}^{n}} and the \ast operator is the Hodge star operator on the Euclidean space, the indices i,ji,j run over MnM^{n} and SRS_{R} is the sphere of radius RR on n\mathbb{R}^{n} and ωn\omega_{n} is the volume of the unit sphere in n\mathbb{R}^{n}.

Finally we collect some basic facts for our model cone. Let (Nn1,gN)(N^{n-1},g^{N}) be a compact Riemannian manifold, and (C(N),g¯)=(+×N,dr2+r2gN)(C(N),\overline{g})=(\mathbb{R}_{+}\times N,dr^{2}+r^{2}g^{N}) be the Riemannian cone over (Nn1,gN)(N^{n-1},g^{N}), where rr is the coordinate on +\mathbb{R}_{+}. We call (N,gN)(N,g^{N}) the cross section of the cone. Let {e1,,en1}\{e_{1},\cdots,e_{n-1}\} be a local orthonormal frame of TNTN with respect to gNg^{N}, e¯i=1rei\overline{e}_{i}=\frac{1}{r}e_{i} for all i=1,,n1i=1,\cdots,n-1, and r=r\partial_{r}=\frac{\partial}{\partial r}. Then {e¯1,,e¯n1,r}\{\overline{e}_{1},\cdots,\overline{e}_{n-1},\partial_{r}\} is a local orthonormal frame of TC(N)TC(N) with respect to g¯\overline{g}. Let ¯\overline{\nabla} denote the Levi-Civita connection of g¯\overline{g}. By Koszul’s formula, one easily obtains

(2.2) {¯e¯ir=1re¯i,¯re¯i=[r,e¯i]+¯e¯ir=1re¯i+1re¯i=0,¯rr=0,¯e¯ie¯j=e¯igNe¯j1rδijr,1i,jn1.\begin{cases}&\overline{\nabla}_{\overline{e}_{i}}\partial_{r}=\frac{1}{r}\overline{e}_{i},\\ &\overline{\nabla}_{\partial_{r}}\overline{e}_{i}=[\partial_{r},\overline{e}_{i}]+\overline{\nabla}_{\overline{e}_{i}}\partial_{r}=-\frac{1}{r}\overline{e}_{i}+\frac{1}{r}\overline{e}_{i}=0,\\ &\overline{\nabla}_{\partial_{r}}\partial_{r}=0,\\ &\overline{\nabla}_{\overline{e}_{i}}\overline{e}_{j}=\nabla^{g^{N}}_{\overline{e}_{i}}\overline{e}_{j}-\frac{1}{r}\delta_{ij}\partial_{r},\end{cases}\ \ \forall 1\leq i,j\leq n-1.

Let R¯\overline{R} denote the Riemann curvature tensor of g¯\overline{g}, and RNR^{N} denote the Riemann curvature tensor of gNg^{N} on NN. In the local frame {e¯1,,e¯n1,r}\{\overline{e}_{1},\cdots,\overline{e}_{n-1},\partial_{r}\}, the only non-vanishing components of R¯\overline{R} are

(2.3) R¯ijkl=r2(RijklN+gikNgjlNgilNgjkN), 1i,j,k,ln1.\overline{R}_{ijkl}=r^{2}\left(R^{N}_{ijkl}+g^{N}_{ik}g^{N}_{jl}-g^{N}_{il}g^{N}_{jk}\right),\ \ 1\leq i,j,k,l\leq n-1.

In particular,

(2.4) R¯(r,X,Y,Z)=0,X,Y,ZΓ(TC(N)),\overline{R}(\partial_{r},X,Y,Z)=0,\ \ \forall X,Y,Z\in\Gamma(TC(N)),

and so

Ricg¯(X,Y)\displaystyle{\operatorname{Ric}}_{\overline{g}}(X,Y) =\displaystyle= RicgN(X,Y)(n2)gN(X,Y),X,YΓ(TN);\displaystyle{\operatorname{Ric}}_{g^{N}}(X,Y)-(n-2)g^{N}(X,Y),\ \ \forall X,Y\in\Gamma(TN);
Ricg¯(r,)\displaystyle{\operatorname{Ric}}_{\overline{g}}(\partial_{r},\cdot) =\displaystyle= 0,\displaystyle 0,

and

(2.5) Scg¯=ScgN(n1)(n2)r2.{\rm Sc}_{\overline{g}}=\frac{{\rm Sc}_{g^{N}}-(n-1)(n-2)}{r^{2}}.

The Laplace operator Δg¯\Delta_{\overline{g}} on (C(N),g¯)(C(N),\overline{g}) is given by

(2.6) Δg¯=r2+n1rr+1r2ΔgN,\Delta_{\overline{g}}=\partial_{r}^{2}+\frac{n-1}{r}\partial_{r}+\frac{1}{r^{2}}\Delta_{g^{N}},

where ΔgN\Delta_{g^{N}} is Laplace operator on the cross section (N,gN)(N,g^{N}).

3. Analysis on AF manifolds with isolated conical singularities

3.1. Weighted Sobolev spaces

Let (Mn,g,o)(M^{n},g,o) be a AF manifold with a single conical singularity at oo as defined in Definition 2.2. Choose three cut-off functions 0χ1,χ2,χ310\leq\chi_{1},\chi_{2},\chi_{3}\leq 1 satisfying:

(3.1) χ1(x)={1,dist(x,o)<ϵ,0,dist(x,o)>2ϵ,\chi_{1}(x)=\begin{cases}1,&{\rm dist}(x,o)<\epsilon,\\ 0,&{\rm dist}(x,o)>2\epsilon,\end{cases}
(3.2) χ2(x)={1,dist(x,o)>2R,0,dist(x,o)<R,\chi_{2}(x)=\begin{cases}1,&{\rm dist}(x,o)>2R,\\ 0,&{\rm dist}(x,o)<R,\end{cases}

and

(3.3) χ3=1χ1χ2,\chi_{3}=1-\chi_{1}-\chi_{2},

where ϵ>0\epsilon>0 is chosen sufficiently small such that the ball B2ϵ(o)B_{2\epsilon}(o), centered at singular point oo with radius 2ϵ2\epsilon, is contained in the asymptotically conical neighborhood UoU_{o} in Definition 2.1, and R>0R>0 is chosen sufficiently large such that MBR(o)MM\setminus B_{R}(o)\subset M_{\infty}.

For each 1p<+,k1\leq p<+\infty,k\in\mathbb{N} and δ,β\delta,\beta\in\mathbb{R}, the weighted Sobolev space Wδ,βk,p(M)W^{k,p}_{\delta,\beta}(M) is defined to be the completion of C0(M{o})C^{\infty}_{0}(M\setminus\{o\}) with respect to the weighted Sobolev norm given by

(3.4) uWδ,βk,p(M)p:=Mi=0k(rp(δi)n|iu|pχ1+ρp(βi)n|iu|pχ2+|iu|pχ3)dvolg,\|u\|^{p}_{W^{k,p}_{\delta,\beta}(M)}:=\int_{M}\sum^{k}_{i=0}\left(r^{-p(\delta-i)-n}|\nabla^{i}u|^{p}\chi_{1}+\rho^{-p(\beta-i)-n}|\nabla^{i}u|^{p}\chi_{2}+|\nabla^{i}u|^{p}\chi_{3}\right)d{\rm vol}_{g},

where rr is the radial coordinate on the conical neighborhood UoU_{o} in Definition 2.1 and ρ\rho is the Euclidean distance function to a base point on MM_{\infty} in Definition 2.2, and |iu||\nabla^{i}u| is the norm of ithi^{th} covariant derivative of uu with respect to gg.

Note that by the definition of the weighted Sobolev norms, we clearly have

(3.5) {δδ,ββ,Wδ,βk,pWδ,βk,p,\begin{cases}\delta^{\prime}\geq\delta,\cr\beta^{\prime}\leq\beta,\end{cases}\Rightarrow\ \ W^{k,p}_{\delta^{\prime},\beta^{\prime}}\subset W^{k,p}_{\delta,\beta},

and

(3.6) rνχ1Wδ,βk,p(M)ν>δ,ρμχ2Wδ,βk,p(M)μ<β.r^{\nu}\chi_{1}\in W^{k,p}_{\delta,\beta}(M)\Longleftrightarrow\nu>\delta,\quad\rho^{\mu}\chi_{2}\in W^{k,p}_{\delta,\beta}(M)\Longleftrightarrow\mu<\beta.

3.2. Elliptic estimate for the Laplace operator

Throughout the paper, we always use λj\lambda_{j} to denote the eigenvalues of the Laplace operator ΔgN\Delta_{g^{N}} on the cross section of the model cone (C(N),g¯)(C(N),\overline{g}), and EjE_{j} to denote the space of the corresponding eigenfunctions. We set

(3.7) νj±:=(n2)±(n2)2+4λj2.\nu_{j}^{\pm}:=\frac{-(n-2)\pm\sqrt{(n-2)^{2}+4\lambda_{j}}}{2}.
Definition 3.1.

We say δ\delta\in\mathbb{R} is critical at conical point if δ=νj+\delta=\nu_{j}^{+} or δ=νj\delta=\nu_{j}^{-} for some j0j\in\mathbb{Z}_{\geq 0}, where νj±\nu^{\pm}_{j} is defined as in (3.7).

We view the Euclidean space (n,gn)(\mathbb{R}^{n},g_{\mathbb{R}^{n}}) as a cone over the round sphere 𝕊n1\mathbb{S}^{n-1} with constant sectional curvature 11, whose Laplace operator has eigenvalues: j(n2+j)j(n-2+j), jj\in\mathbb{N}. Replacing λj\lambda_{j} by these eigenvalues on round sphere, we make following definition.

Definition 3.2.

We say β\beta\in\mathbb{R} is critical at infinity if β=k\beta=k or β=2nk\beta=2-n-k for some k0k\in\mathbb{Z}_{\geq 0}.

Definition 3.3.

We say (δ,β)2(\delta,\beta)\in\mathbb{R}^{2} is critical if δ\delta is critical at conical point or β\beta is critical at infinity.

By using scaling technique, the usual interior elliptic estimates, and the asymptotic control of metric gg near conical point in Definition 2.1 and near infinity in Definition 2.2, one can obtain the following weighted elliptic estimate. This is treated in detail in [5] for the Dirac operator and carries over except minor modifications.

Proposition 3.4.

Let (Mn,g,o)(M^{n},g,o) be an AF manifold with a single conical singularity at oo. For any 1p<+1\leq p<+\infty, and δ,β\delta,\beta\in\mathbb{R}, if uLδ,βp(M)u\in L^{p}_{\delta,\beta}(M), and ΔuLδ2,β2p(M)\Delta u\in L^{p}_{\delta-2,\beta-2}(M), then

(3.8) uWδ,β2,p(M)C(ΔuLδ2,β2p(M)+uLδ,βp(M))\|u\|_{W^{2,p}_{\delta,\beta}(M)}\leq C\left(\|\Delta u\|_{L^{p}_{\delta-2,\beta-2}(M)}+\|u\|_{L^{p}_{\delta,\beta}(M)}\right)

holds for some constant C=C(g,n,p)C=C(g,n,p) independent of uu.

To obtain Fredholm property for the Laplace operator, the following refined weighted elliptic estimate is crucial.

Proposition 3.5.

Let (Mn,g,o)(M^{n},g,o) be an AF manifold with a single conical singularity at oo. If (δ,β)2(\delta,\beta)\in\mathbb{R}^{2} is not critical as in Definition 3.3, there exists a constant C=C(g,n)C=C(g,n) and a compact set BM{o}B\subset M\setminus\{o\} such that for any uLδ,β2(M)u\in L^{2}_{\delta,\beta}(M) with ΔuLδ2,β22(M)\Delta u\in L^{2}_{\delta-2,\beta-2}(M),

(3.9) uWδ,β2,2(M)C(ΔuLδ2,β22(M)+uL2(B)).\|u\|_{W^{2,2}_{\delta,\beta}(M)}\leq C\left(\|\Delta u\|_{L^{2}_{\delta-2,\beta-2}(M)}+\|u\|_{L^{2}(B)}\right).

Proposition 3.5 is an immediate consequence of Proposition 3.4 and the following Lemma:

Lemma 3.6.

Let (Mn,g,o)(M^{n},g,o) be an AF manifold with a single conical singularity at oo. If (δ,β)2(\delta,\beta)\in\mathbb{R}^{2} is not critical as in Definition 3.3, there exists a constant CC and a compact set BM{o}B\subset M\setminus\{o\} such that for any uLδ,β2(M)u\in L^{2}_{\delta,\beta}(M) with ΔuLδ2,β22(M)\Delta u\in L^{2}_{\delta-2,\beta-2}(M),

(3.10) uLδ,β2(M)C(ΔuLδ2,β22(M)+uL2(B)).\|u\|_{L^{2}_{\delta,\beta}(M)}\leq C\left(\|\Delta u\|_{L^{2}_{\delta-2,\beta-2}(M)}+\|u\|_{L^{2}(B)}\right).

To prove Lemma 3.6, it is suffices to deal with model cone metric g¯\overline{g} on the conical neighborhood of the singular point oo and gng_{\mathbb{R}^{n}} near the AF infinity. The estimate near the AF infinity has been done in [23], by doing a spectral decomposition with respect to Δg𝕊n1\Delta_{g_{\mathbb{S}^{n-1}}}, and reducing the problem to estimating coefficient functions, which satisfy certain second order ODEs. The Euclidean metric gng_{\mathbb{R}^{n}} can be viewed as a model cone metric with standard sphere as the cross section. The proof in [23] can be adapted to that near the tip of a cone, and then the estimate near the conical singularity follows. In [5], we have derived in detail the analogous estimate for the Dirac operator. So we omit the details of the proof, and refer to [23] and [5].

3.3. Fredholm property of Laplace operator

We consider the unbounded operator

Δδ,β:Dom(Δδ,β)\displaystyle\Delta_{\delta,\beta}:{\rm Dom}(\Delta_{\delta,\beta}) \displaystyle\rightarrow Lδ2,β22(M)\displaystyle L^{2}_{\delta-2,\beta-2}(M)
φ\displaystyle\varphi \displaystyle\mapsto Δφ,\displaystyle\Delta\varphi,

whose domain Dom(Δδ,β){\rm Dom}(\Delta_{\delta,\beta}) is dense subset of Lδ,β2(M)L^{2}_{\delta,\beta}(M) consisting of function uu such that ΔuLδ2,β22(M)\Delta u\in L^{2}_{\delta-2,\beta-2}(M) in the sense of distributions.

As Lemma 4.8 in [5], we have:

Lemma 3.7.

The unbounded operator Δδ,β\Delta_{\delta,\beta} is closed.

The usual L2L^{2} pairing (,)L2(M)(\cdot,\cdot)_{L^{2}(M)} identifies the topological dual space of Lδ,β2L^{2}_{\delta,\beta} with Lδn,βn2L^{2}_{-\delta-n,-\beta-n}. By this identification, the adjoint operator (Δδ,β)\left(\Delta_{\delta,\beta}\right)^{*} of Δδ,β\Delta_{\delta,\beta} is given as

(Δδ,β):Dom((Δδ,β))\displaystyle\left(\Delta_{\delta,\beta}\right)^{*}:{\rm Dom}\left((\Delta_{\delta,\beta})^{*}\right) \displaystyle\rightarrow Lδn,βn2\displaystyle L^{2}_{-\delta-n,-\beta-n}
u\displaystyle u \displaystyle\mapsto Δu,\displaystyle\Delta u,

where the domain: Dom((Δδ,β)){\rm Dom}\left(\left(\Delta_{\delta,\beta}\right)^{*}\right) is the dense subset of Lδ+2n,β+2n2L^{2}_{-\delta+2-n,-\beta+2-n} consisting of functions uu such that ΔuLδn,βn2\Delta u\in L^{2}_{-\delta-n,-\beta-n} in the distributional sense.

By applying the refined weighted elliptic estimate in Proposition 3.5, as Proposition 4.9 in [5], we have the following Fredholm property.

Proposition 3.8.

If (δ,β)2(\delta,\beta)\in\mathbb{R}^{2} is not critical as in Definition 3.3, then the operator Δδ,β\Delta_{\delta,\beta} is Fredholm, namely,

  1. (1)(1)

    Ran(Δδ,β){\rm Ran}(\Delta_{\delta,\beta}) is closed,

  2. (2)(2)

    dim(Ker(Δδ,β))<+{\rm dim}\left({\rm Ker}(\Delta_{\delta,\beta})\right)<+\infty,

  3. (3)(3)

    dim(Ker((Δδ,β)))<+{\rm dim}\left({\rm Ker}((\Delta_{\delta,\beta})^{*})\right)<+\infty.

For the proof one can see that of Proposition 4.9 in [5].

3.4. Solving the Laplace equation

Let (Mn,g,o)(M^{n},g,o) be an AF manifold with a single conical singularity at oo as defined in 2.2. The equation Δu=f\Delta u=f has been solved near infinity in [23]. Here we similarly solve it near the conical point oo. The strategy is similar as solving the Dirac equation near the cone point in [5], and so we omit proofs of some results in this section and refer to the analogous results for Dirac operator in Section 4 in [5].

We use BrB_{r} to denote the ball with radius rr centered at the conical singularity oo. In the notion of weighted Sobolev norms and spaces over BrB_{r}, the subscript β\beta will be neglected, and they will be written as Wδ2,2(Br)\|\cdot\|_{W^{2,2}_{\delta}(B_{r})} and Wδ2,2(Br)W^{2,2}_{\delta}(B_{r}), since there is no asymptotic control near infinity need to be concerned over the finite ball BrB_{r}.

For the model cone metric g¯=dr2+r2gN\overline{g}=dr^{2}+r^{2}g^{N}, by solving the Dirichlet problem for the equation Δg¯u=f\Delta_{\overline{g}}u=f on a compact exhaustion of B2r0{o}B_{2r_{0}}\setminus\{o\}, and using refined weighted elliptic estimate in Lemma 3.6, one can solve Δg¯u=f\Delta_{\overline{g}}u=f on B2r0B_{2r_{0}} and prove the following:

Lemma 3.9.

For δ\delta\in\mathbb{R}, which is noncritical at conical point, and a small number r0>0r_{0}>0, there is a bounded operator

Gg¯:Lδ22(B2r0)Wδ2,2(B2r0)G_{\overline{g}}:L^{2}_{\delta-2}(B_{2r_{0}})\to W^{2,2}_{\delta}(B_{2r_{0}})

such that Δg¯Gg¯=id\Delta_{\overline{g}}\circ G_{\overline{g}}=id.

A perturbation argument extends this result to a more general setting, namely, for an asymptotically conical metric g=g¯+hg=\overline{g}+h where hh satisfies (2.1), and we have the following:

Proposition 3.10.

For δ\delta\in\mathbb{R}, which is noncritical at cone point as in Definition 3.1, and a small number r0r_{0}, there is a bounded operator Gg:Lδ22(B2r0)Wδ2,2(B2r0)G_{g}:L^{2}_{\delta-2}(B_{2r_{0}})\mapsto W^{2,2}_{\delta}(B_{2r_{0}}) such that ΔgGg=id\Delta_{g}\circ G_{g}=id.

As an application of Proposition 3.10, one can solve harmonic functions near the conical point, which are asymptotic to harmonic functions, rνj±ϕjr^{\nu^{\pm}_{j}}\phi_{j}, with respect to the model cone metric g¯\overline{g}.

Corollary 3.11.

Given jj\in\mathbb{N} and ϕEj\phi\in E_{j}, there are functions j,ϕ±\mathcal{H}^{\pm}_{j,\phi} that are harmonic near the conical point and can be written as

j,ϕ±=rνj±ϕ+v±\mathcal{H}^{\pm}_{j,\phi}=r^{\nu^{\pm}_{j}}\phi+v_{\pm}

with v+v_{+} in Wη2,2W^{2,2}_{\eta} for any η<νj++α\eta<\nu^{+}_{j}+\alpha and vv_{-} in Wη2,2W^{2,2}_{\eta} for any η<νj+α\eta<\nu^{-}_{j}+\alpha.

The solution of the equation Δu=f\Delta u=f may not be unique, and the difference of two solutions is a harmonic function. For the model cone metric g¯=dr2+r2gN\overline{g}=dr^{2}+r^{2}g^{N}, the Laplace operator Δg¯\Delta_{\overline{g}} is given in (2.6), and the harmonic functions are linear combinations of rνj±ϕjr^{\nu^{\pm}_{j}}\phi_{j}, where νj±\nu^{\pm}_{j} are given in (3.7) and ϕjEj\phi_{j}\in E_{j}, i.e. ΔgNϕj=λjϕj\Delta_{g^{N}}\phi_{j}=\lambda_{j}\phi_{j}. Based on this observation, by using (3.6) and Lemma 3.9, one can prove the following:

Lemma 3.12.

Suppose Δg¯u=f\Delta_{\overline{g}}u=f with uu in Lδ2(B2r0)L_{\delta}^{2}(B_{2r_{0}}) and ff in Lδ22(B2r0)L^{2}_{\delta^{\prime}-2}(B_{2r_{0}}) for non-critical exponents δ<δ\delta<\delta^{\prime} and a small number r0r_{0}. Then there is an element vv of Lδ2(B2r0)L^{2}_{\delta^{\prime}}(B_{2r_{0}}) such that uvu-v is a linear combination of functions: rνj±ϕjr^{\nu_{j}^{\pm}}\phi_{j} with ϕjEj\phi_{j}\in E_{j} and δ<νj±<δ\delta<\nu^{\pm}_{j}<\delta^{\prime};

By using Corollary 3.11, the property for model cone metrics in Lemma 3.12 can be extended to asymptotically conical metrics gg as following:

Proposition 3.13.

Suppose Δgu=f\Delta_{g}u=f with uu in Lδ2(B2r0)L_{\delta}^{2}(B_{2r_{0}}) and ff in Lδ22(B2r0)L^{2}_{\delta^{\prime}-2}(B_{2r_{0}}) for non-critical exponents δ<δ\delta<\delta^{\prime} . Then, up to making B2r0B_{2r_{0}} smaller, there is an element vv of Lδ2(B2r0)L^{2}_{\delta^{\prime}}(B_{2r_{0}}) such that uvu-v is a linear combination of the following functions: j,ϕj±\mathcal{H}^{\pm}_{j,\phi_{j}} with ϕj\phi_{j} in EjE_{j} and δ<νj±<δ\delta<\nu^{\pm}_{j}<\delta^{\prime}, where j,ϕj±\mathcal{H}^{\pm}_{j,\phi_{j}} are functions, harmonic near the conical singularity, obtained in Corollary 3.11.

Now we are ready to prove the surjectivity of Δδ,β\Delta_{\delta,\beta} for certain noncritical indices (δ,β)(\delta,\beta), which enables us to solve the equation Δgu=f\Delta_{g}u=f.

First, by using Fredholm property in Proposition 3.8, we can prove the following surjectivity. For the proof, we refer to proofs of Corollary 2 in [23] and Proposition 4.15 in [5].

Proposition 3.14.

Let (Mn,g,o)(M^{n},g,o) be an AF manifold with a single conical singularity at oo. For any noncritical (δ,β)(\delta,\beta) satisfying δ2n2\delta\leq\frac{2-n}{2} and β2n2\beta\geq\frac{2-n}{2}, the map

(3.11) Δδ,β:Dom(Δδ,β)Lδ2,β22(M)\Delta_{\delta,\beta}:{\rm Dom}\left(\Delta_{\delta,\beta}\right)\rightarrow L^{2}_{\delta-2,\beta-2}(M)

is surjective.

Moreover, the map

(3.12) Δ2n2,2n2:Dom(Δ2n2,2n2)Ln+22,n+222(M)\Delta_{\frac{2-n}{2},\frac{2-n}{2}}:{\rm Dom}\left(\Delta_{\frac{2-n}{2},\frac{2-n}{2}}\right)\rightarrow L^{2}_{-\frac{n+2}{2},-\frac{n+2}{2}}(M)

is an isomorphism.

Then by applying Propositions 3.13 and 3.14, we can extend the region of indices (δ,β)(\delta,\beta) for which Δδ,β\Delta_{\delta,\beta} is surjective, and obtain the following:

Proposition 3.15.

Let (Mn,g,o)(M^{n},g,o) be an AF manifold with a single conical singularity at oo. We have that

(3.13) Δδ,β:Dom(Δδ,β)Lδ2,β22(M)\Delta_{\delta,\beta}:{\rm Dom}\left(\Delta_{\delta,\beta}\right)\rightarrow L^{2}_{\delta-2,\beta-2}(M)

is surjective for noncritical δ<0\delta<0 and β>2n\beta>2-n.

Proof.

By Proposition 3.14, it suffices to show that Δδ,β\Delta_{\delta,\beta} is surjective for 2n2<δ<0\frac{2-n}{2}<\delta<0 and 2n<β<2n22-n<\beta<\frac{2-n}{2}.

For arbitrary noncritical δ(2n2,0)\delta^{\prime}\in\left(\frac{2-n}{2},0\right) and β(2n,2n2)\beta^{\prime}\in\left(2-n,\frac{2-n}{2}\right), we take an arbitrary function fLδ2,β22(M)L2n22,2n222(M)f\in L^{2}_{\delta^{\prime}-2,\beta^{\prime}-2}(M)\subset L^{2}_{\frac{2-n}{2}-2,\frac{2-n}{2}-2}(M). Proposition 3.14 then implies that there exists uL2n2,2n22(M)u\in L^{2}_{\frac{2-n}{2},\frac{2-n}{2}}(M) such that Δu=f\Delta u=f. Proposition 3.13 then implies that uLδ,β2u\in L^{2}_{\delta^{\prime},\beta^{\prime}}, since there is no critical index at conical point in (δ,0)\left(\delta^{\prime},0\right) and no critical index at infinity in (β,2n2)\left(\beta^{\prime},\frac{2-n}{2}\right). Therefore, Δδ,β\Delta_{\delta^{\prime},\beta^{\prime}} is surjective, and this completes the proof. ∎

4. Nonnegativity of mass

In this section, we prove the non-negativity of the ADM mass for AF manifolds with a single conical singularity with nonnegative scalar curvature on the regular part, by following the approach of Ju and Viaclovsky in the case of orbifold singularity in [12]. For that, we first solve a harmonic function as following:

Lemma 4.1.

Let (Mn,g,o)(M^{n},g,o) be a nn-dimensional AF manifold with a single conical singularity at oo. There exists a harmonic function uu on MM which satisfies u>1u>1 and admits the expansion

(4.1) u={r2n+o(r2n+α),asr0,1+Aρ2n+o(ρ2nϵ),asρ,u=\begin{cases}r^{2-n}+o(r^{2-n+\alpha^{\prime}}),&\text{as}\ \ r\to 0,\\ 1+A\rho^{2-n}+o(\rho^{2-n-\epsilon}),&\text{as}\ \ \rho\to\infty,\end{cases}

for 0<α<min{1,α}0<\alpha^{\prime}<\min\{1,\alpha\} and 0<ϵ<min{1,τ}0<\epsilon<\min\{1,\tau\}, and some constant A>0A>0. Here α\alpha and τ\tau are asymptotic orders of the metric gg in Definitions 2.1 and 2.2.

Proof.

Choose a cut-off function ϕ\phi satisfying

ϕ={1,onB12(o),0,onMB1(o).\phi=\begin{cases}1,&\text{on}\ \ B_{\frac{1}{2}}(o),\\ 0,&\text{on}\ \ M\setminus B_{1}(o).\end{cases}

Let u0=ϕr2nu_{0}=\phi r^{2-n}, which is a smooth function supported in a neighborhood of the conically singular point oo. Then because Δg¯r2n=0\Delta_{\overline{g}}r^{2-n}=0, by the asymptotic control of gg near conically singular point oo, we have

Δgu0={O(rn+α),asr0,0,onMB1(o).\Delta_{g}u_{0}=\begin{cases}O(r^{-n+\alpha}),&\text{as}\ \ r\to 0,\\ 0,&\text{on}\ \ M\setminus B_{1}(o).\end{cases}

By (3.6), this implies that

Δgu0Lδ2,β22(M),δ<2n+α,β>2n.\Delta_{g}u_{0}\in L^{2}_{\delta-2,\beta-2}(M),\quad\forall\delta<2-n+\alpha,\ \ \beta>2-n.

Then by applying Proposition 3.15, we obtain vLδ,β2v\in L^{2}_{\delta,\beta} such that

Δgv=Δgu0.\Delta_{g}v=\Delta_{g}u_{0}.

So by setting u1=u0vu_{1}=u_{0}-v, we have Δgu1=0\Delta_{g}u_{1}=0.

Now we derive the asymptotic behavior of vv near conically singular point and AF infinity. Near the singular point oo, u1=r2nvu_{1}=r^{2-n}-v, and so Δgu1=0\Delta_{g}u_{1}=0 implies

Δgv=Δgr2n=(ΔgΔg¯)r2n=O(rn+α)Lδ2p(B12(o)),δ<2n+αandp>1,\Delta_{g}v=\Delta_{g}r^{2-n}=\left(\Delta_{g}-\Delta_{\overline{g}}\right)r^{2-n}=O(r^{-n+\alpha})\in L^{p}_{\delta-2}(B_{\frac{1}{2}}(o)),\ \ \forall\delta<2-n+\alpha\ \ \text{and}\ \ \forall p>1,

by (3.6). Then the weighted elliptic estimate in Proposition 3.4 implies that vWδ2,2(B12(o))v\in W^{2,2}_{\delta}(B_{\frac{1}{2}}(o)) for δ<2n+α\delta<2-n+\alpha. Consequently, the weighted Sobolev inequality in Proposition 3.4 in [6] implies that vLδ2nn2(B12(o))v\in L^{\frac{2n}{n-2}}_{\delta}(B_{\frac{1}{2}}(o)), and by applying weighed elliptic estimate in Proposition 3.4 (with p=2nn2p=\frac{2n}{n-2}) again, we obtain vWδ2,2nn2(B12(o))v\in W_{\delta}^{2,\frac{2n}{n-2}}(B_{\frac{1}{2}}(o)) for δ<2n+α\delta<2-n+\alpha. By repeating this process, we can obtain that vWδ2,p(B12(o))v\in W^{2,p}_{\delta}(B_{\frac{1}{2}}(o)) for all p>1p>1 and δ<2n+α\delta<2-n+\alpha, and so, by Proposition 3.4 in [6], we have

(4.2) v=o(r2n+α),asr0,v=o(r^{2-n+\alpha^{\prime}}),\ \ \text{as}\ \ r\to 0,

for 0<α<α0<\alpha^{\prime}<\alpha.

Near the AF infinity, u1=vLβ2(M)u_{1}=v\in L^{2}_{\beta}(M_{\infty}) for β>2n\beta>2-n, and Δgv=Δgu1=0Lβ2(M)\Delta_{g}v=\Delta_{g}u_{1}=0\in L^{2}_{\beta^{\prime}}(M_{\infty}) for 1n<β<2n1-n<\beta^{\prime}<2-n. Thus, Proposition 4 in [23] implies that

v=Aρ2n+v,vLβ2(M),v=A\rho^{2-n}+v^{\prime},\ \ v^{\prime}\in L^{2}_{\beta^{\prime}}(M_{\infty}),

for 1n<β<2n1-n<\beta^{\prime}<2-n, and some constant AA. Moreover,

Δgv=Δg(Aρ2n)=(ΔgnΔg)(Aρ2n)=O(ρnτ)Lβ2p(M),\Delta_{g}v^{\prime}=-\Delta_{g}(A\rho^{2-n})=(\Delta_{g_{\mathbb{R}^{n}}}-\Delta_{g})(A\rho^{2-n})=O(\rho^{-n-\tau})\in L^{p}_{\beta^{\prime}-2}(M_{\infty}),

for (2nτ<)1n<β<2n(2-n-\tau<)1-n<\beta^{\prime}<2-n, and all p>1p>1. Then the weighted elliptic estimate in Proposition 3.4 implies vWβ2,2(M)v^{\prime}\in W^{2,2}_{\beta^{\prime}}(M_{\infty}). Then by a similar weighted elliptic bootstrapping argument as above, we can obtain vWβ2,p(M)v^{\prime}\in W^{2,p}_{\beta^{\prime}}(M_{\infty}) for all p>1p>1 and 1n<β<2n1-n<\beta^{\prime}<2-n, and so

(4.3) v=o(ρ2nϵ),asρ,v^{\prime}=o(\rho^{2-n-\epsilon}),\ \ \text{as}\ \ \rho\to\infty,

for 0<ϵ<10<\epsilon<1.

In summary, we obtain a harmonic function u1u_{1} admitting the asymptotic behavior:

(4.4) u1={r2n+o(r2n+α),asr0,Aρ2n+o(ρ2nϵ),asρ,u_{1}=\begin{cases}r^{2-n}+o(r^{2-n+\alpha^{\prime}}),&\text{as}\ \ r\to 0,\\ A\rho^{2-n}+o(\rho^{2-n-\epsilon}),&\text{as}\ \ \rho\to\infty,\end{cases}

for 0<α<α0<\alpha^{\prime}<\alpha and 0<ϵ<min{1,τ}0<\epsilon<\min\{1,\tau\}.

We let u:=1+u1u:=1+u_{1}. Then uu is a harmonic function satisfying the asymptotic control in (4.1). Finally, by using the strong maximum principle and the asymptotic behavior in (4.1), we obtain u1u\geq 1, and in particular, A>0A>0. Then by applying the strong maximum principle again, we can obtain u>1u>1. ∎

Now we are ready to prove the following:

Theorem 4.2.

Let (Mn,g,o),n3,(M^{n},g,o),n\geq 3, be a nn-dimensional AF manifold with a single conical singularity at oo. If the scalar curvature Scg0{\rm Sc}_{g}\geq 0, then the ADM mass m(g)0m(g)\geq 0.

Proof.

For the harmonic function obtained in Lemma 4.1, and any δ>0\delta>0, we define

(4.5) uδ=δu+(1δ),u_{\delta}=\delta u+(1-\delta),

which satisfies Δguδ=0\Delta_{g}u_{\delta}=0, uδ>1u_{\delta}>1, and admits the asymptotic expansion

(4.6) uδ={δr2n+o(r2n+α),asr0,1+δAr2n+o(ρ2nϵ),asρ,u_{\delta}=\begin{cases}\delta r^{2-n}+o(r^{2-n+\alpha^{\prime}}),&\text{as}\ \ r\to 0,\\ 1+\delta Ar^{2-n}+o(\rho^{2-n-\epsilon}),&\text{as}\ \ \rho\to\infty,\end{cases}

for 0<α<α0<\alpha^{\prime}<\alpha and 0<ϵ<10<\epsilon<1. Here the constant A>0A>0.

Then we do a conformal change for the metric gg, and let

(4.7) gδ:=(uδ)4n2gon M.g_{\delta}:=\left(u_{\delta}\right)^{\frac{4}{n-2}}g\ \ \text{on }\ \ M.

By the asymptotic expansion of uδu_{\delta} at infinity as in (4.6), gδg_{\delta} is asymptotically flat of order min{τ,n2}\min\{\tau,n-2\}. Near the conically singular point oo, the metric gg is given as

(4.8) g=dr2+r2gN+h,g=dr^{2}+r^{2}g^{N}+h,

where hh satisfies (2.1), and so

(4.9) gδ=(uδ)4n2g=δ4n2r4(1+o(rα))(dr2+r2gN+h),asr0.g_{\delta}=\left(u_{\delta}\right)^{\frac{4}{n-2}}g=\delta^{\frac{4}{n-2}}r^{-4}\left(1+o(r^{\alpha^{\prime}})\right)\left(dr^{2}+r^{2}g^{N}+h\right),\ \ \text{as}\ \ r\to 0.

Now we do a change of variable, by letting s=δ2n21rs=\delta^{\frac{2}{n-2}}\frac{1}{r}, and rewrite gδg_{\delta} as

(4.10) gδ=(1+o(sα))(ds2+s2gN+h~),g_{\delta}=(1+o(s^{-\alpha^{\prime}}))(ds^{2}+s^{2}g^{N}+\tilde{h}),

where |h~|ds2+s2gN=O(sα)|\tilde{h}|_{ds^{2}+s^{2}g^{N}}=O(s^{-\alpha}) as ss\to\infty. Therefore,

(4.11) gδ=ds2+s2gN+o(sα),ass,g_{\delta}=ds^{2}+s^{2}g^{N}+o(s^{-\alpha^{\prime}}),\ \ \text{as}\ \ s\to\infty,

since 0<α<α0<\alpha^{\prime}<\alpha. Note that the metric gδg_{\delta} is asymptotically conical, and it tends to the infinity end of a cone, as ss\to\infty (i.e. r0r\to 0). Moreover, the scalar curvature of gδg_{\delta} is given by

(4.12) Scgδ=4(n1)n2(uδ)n+2n2(Δguδ+n24(n1)Scguδ)=(uδ)4n2Scg0.{\rm Sc}_{g_{\delta}}=\frac{4(n-1)}{n-2}\left(u_{\delta}\right)^{-\frac{n+2}{n-2}}\left(-\Delta_{g}u_{\delta}+\frac{n-2}{4(n-1)}{\rm Sc}_{g}u_{\delta}\right)=\left(u_{\delta}\right)^{-\frac{4}{n-2}}{\rm Sc}_{g}\geq 0.

For any fixed δ>0\delta>0, we can choose a hypersurface Σδ,s0:={s=s0}\Sigma_{\delta,s_{0}}:=\{s=s_{0}\} for some sufficiently large s0s_{0} such that Σδ,s0\Sigma_{\delta,s_{0}} is strictly mean concave with respect to the normal pointing to the AF end of (M,gδ)(M,g_{\delta}). Let Mδ=M{ss0}M_{\delta}=M\setminus\{s\geq s_{0}\} be a AF manifold with strictly mean concave boundary. By applying Theorem in [10], we obtain

(4.13) m(Mδ,gδ)0.m(M_{\delta},g_{\delta})\geq 0.

On the other hand,

(4.14) m(Mδ,gδ)=m(M,g)+4(n2)δA.m(M_{\delta},g_{\delta})=m(M,g)+4(n-2)\delta A.

Because this is true for any constant δ>0\delta>0, and AA is a fixed constant, we must have

(4.15) m(M,g)0.m(M,g)\geq 0.

Remark 4.3.

Note that (M,gδ)(M,g_{\delta}) is a complete Riemannian manifolds with two ends, one is an AF end and another is an asymptotically conical end. In the case of the dimension 3n73\leq n\leq 7, the results of the positive mass theorems on manifold with arbitrary ends in [16, 29] imply the mass m(M,gδ)0m(M,g_{\delta})\geq 0. Then our conclusion follows as well.

5. Rigidity

In this section, we prove that when the mass is zero, then MM is isometric to n\mathbb{R}^{n}. We first prove that MM is Ricci flat, then show that it must be the Euclidean space.

First of all, we state the following Sobolev inequality on AF manifold with a single conical singularity, which can be obtained by combining the Sobolev inequality on cone (see, Lemma 3.2 in [6]) and that on the Euclidean space. It will be used in the proof of Lemma 5.2.

Lemma 5.1.

Let (Mn,g,o)(M^{n},g,o) be a nn-dimensional AF conical singularity at oo. There is a Sobolev constant C>0C>0 such that

(5.1) (M|f|2nn2𝑑volg)n22nC(M|f|2𝑑volg)12\left(\int_{M}|f|^{\frac{2n}{n-2}}d{\rm vol}_{g}\right)^{\frac{n-2}{2n}}\leq C\left(\int_{M}|\nabla f|^{2}d{\rm vol}_{g}\right)^{\frac{1}{2}}

holds for all fC0(M̊)f\in C^{\infty}_{0}(\mathring{M}).

Then as analysis preparations for proving the rigidity result in Theorem 1.1 (stated in Theorem 5.4 below), we solve Schrödinger equation with compactly supported potential function on AF manifolds with a conical singularity in Lemma 5.2; and solve harmonic functions, which are asymptotic to Euclidean coordinate near the AF infinity in Lemma 5.3. We also study their asymptotic behaviors near the conical singularity, which are critical in the proof of Theorem 5.4.

Lemma 5.2.

Let (Mn,g,o)(M^{n},g,o) be an AF manifold with a conical singularity at oo. There exists a constant ϵ0=ϵ(g)\epsilon_{0}=\epsilon(g), such that if ff is a smooth function with compact support in M̊\mathring{M} and fLn2(M)<ϵ0\|f_{-}\|_{L^{\frac{n}{2}}(M)}<\epsilon_{0}, then the equation

(5.2) {Δgu+fu=0,u1 as x,\begin{cases}-\Delta_{g}u+fu=0,\\ u\to 1\mbox{ as }x\to\infty,\end{cases}

has a positive solution admitting the asymptotic as:

(5.3) u={1+Aρn2+u1asρ,B+a(y)rν1+u2asr0,u=\begin{cases}1+\frac{A}{\rho^{n-2}}+u_{1}\ \ &\text{as}\ \ \rho\to\infty,\\ B+a(y)r^{\nu_{1}}+u_{2}\ \ &\text{as}\ \ r\to 0,\end{cases}

where AA and BB are constants, and B1B\geq 1,

(5.4) ν1:=2n+(n2)2+4λ12,\nu_{1}:=\frac{2-n+\sqrt{(n-2)^{2}+4\lambda_{1}}}{2},

here λ1\lambda_{1} is the first non-zero eigenvalue of the Laplacian ΔgN\Delta_{g^{N}} and a(y)a(y) is a corresponding eigenfunction, i.e. ΔgNa(y)=λ1a(y)\Delta_{g^{N}}a(y)=-\lambda_{1}a(y), and u1u_{1} and u2u_{2} have asymptotic as

(5.5) |iu1|=o(ρβi),asρ+,for 2nτ<β<2n,i=0,1,|\nabla^{i}u_{1}|=o(\rho^{\beta-i}),\ \ \text{as}\ \ \rho\to+\infty,\ \ \text{for}\ \ 2-n-\tau<\beta<2-n,\ \ i=0,1,

and

(5.6) |iu2|=o(rδi),asr0,forν1<δ<ν1+α,i=0,1.|\nabla^{i}u_{2}|=o(r^{\delta-i}),\ \ \text{as}\ \ r\to 0,\ \ \text{for}\ \ \nu_{1}<\delta<\nu_{1}+\alpha,\ \ i=0,1.
Proof.

Existence of a positive solution: Let v=1uv=1-u. Then equation (5.2) becomes

(5.7) Δgvfv=f.\Delta_{g}v-fv=-f.

On a compact subset Ar,ρ:=Bρ(o)Br(o)A_{r,\rho}:=B_{\rho}(o)\setminus B_{r}(o), consider the Dirichlet problem

(5.8) {Δgvr,ρfvr,ρ=f,inAr,ρ,vr,ρ=0,onAr,ρ.\begin{cases}\Delta_{g}v_{r,\rho}-fv_{r,\rho}=-f,&\text{in}\;A_{r,\rho},\\ v_{r,\rho}=0,&\text{on}\;\partial A_{r,\rho}.\end{cases}

By Fredholm alternative, if the homogeneous equation

(5.9) {Δgvr,ρfvr,ρ=0,inAr,ρvr,ρ=0,onAr,ρ\begin{cases}\Delta_{g}v_{r,\rho}-fv_{r,\rho}=0,&\text{in}\;A_{r,\rho}\\ v_{r,\rho}=0,&\text{on}\;\partial A_{r,\rho}\end{cases}

has only zero solution, then equation (5.8) has a unique solution. Suppose ω\omega is a solution of equation (5.9). Multiplying ω\omega to both sides of the equation in (5.9) and integrating by parts, by the Hölder inequality with p=n2,q=nn2p=\frac{n}{2},q=\frac{n}{n-2} and the Sobolev inequality in Lemma 5.1, we have

Ar,ρ|ω|2dvolg\displaystyle\int_{A_{r,\rho}}|\nabla\omega|^{2}d\operatorname{vol}_{g} =\displaystyle= Ar,ρfω2dvolgAr,ρfω2dvolg\displaystyle-\int_{A_{r,\rho}}f\omega^{2}d\operatorname{vol}_{g}\leq\int_{A_{r,\rho}}f_{-}\omega^{2}d\operatorname{vol}_{g}
\displaystyle\leq (Ar,ρfn2dvolg)2n(Ar,ρω2nn2dvolg)n2n\displaystyle\left(\int_{A_{r,\rho}}f_{-}^{\frac{n}{2}}d\operatorname{vol}_{g}\right)^{\frac{2}{n}}\left(\int_{A_{r,\rho}}\omega^{\frac{2n}{n-2}}d\operatorname{vol}_{g}\right)^{\frac{n-2}{n}}
\displaystyle\leq c1(Ar,ρfn2dvolg)2n(Ar,ρ|ω|2dvolg)\displaystyle c_{1}\left(\int_{A_{r,\rho}}f_{-}^{\frac{n}{2}}d\operatorname{vol}_{g}\right)^{\frac{2}{n}}\left(\int_{A_{r,\rho}}|\nabla\omega|^{2}d\operatorname{vol}_{g}\right)

Thus if fLn2(M)<1c1\|f_{-}\|_{L^{\frac{n}{2}}(M)}<\frac{1}{c_{1}}, then ω=0\omega=0. Therefore, equation (5.8) has a unique solution vr,ρv_{r,\rho}. Multiplying vr,ρv_{r,\rho} to both sides of (5.8), using Hölder inequality and Sobolev inequality again,

Ar,ρ|vr,ρ|2dvolg\displaystyle\int_{A_{r,\rho}}|\nabla v_{r,\rho}|^{2}d\operatorname{vol}_{g} \displaystyle\leq Ar,ρfvr,ρ2dvolg+Ar,ρfvr,ρdvolg\displaystyle\int_{A_{r,\rho}}f_{-}v^{2}_{r,\rho}d\operatorname{vol}_{g}+\int_{A_{r,\rho}}fv_{r,\rho}d\operatorname{vol}_{g}
\displaystyle\leq c1(Ar,ρfn2dvolg)2n(Ar,ρ|vr,ρ|2dvolg)\displaystyle c_{1}\left(\int_{A_{r,\rho}}f_{-}^{\frac{n}{2}}d\operatorname{vol}_{g}\right)^{\frac{2}{n}}\left(\int_{A_{r,\rho}}|\nabla v_{r,\rho}|^{2}d\operatorname{vol}_{g}\right)
+(Ar,ρ|f|2nn+2dvolg)n+22n(Ar,ρvr,ρ2nn2dvolg)n22n\displaystyle+\left(\int_{A_{r,\rho}}|f|^{\frac{2n}{n+2}}d\operatorname{vol}_{g}\right)^{\frac{n+2}{2n}}\left(\int_{A_{r,\rho}}v_{r,\rho}^{\frac{2n}{n-2}}d\operatorname{vol}_{g}\right)^{\frac{n-2}{2n}}
\displaystyle\leq c1(Ar,ρfn2dvolg)2n(Ar,ρ|vr,ρ|2dvolg)\displaystyle c_{1}\left(\int_{A_{r,\rho}}f_{-}^{\frac{n}{2}}d\operatorname{vol}_{g}\right)^{\frac{2}{n}}\left(\int_{A_{r,\rho}}|\nabla v_{r,\rho}|^{2}d\operatorname{vol}_{g}\right)
+c1(Ar,ρ|f|2nn+2dvolg)n+22n(Ar,ρ|vr,ρ|2dvolg)12\displaystyle+c_{1}\left(\int_{A_{r,\rho}}|f|^{\frac{2n}{n+2}}d\operatorname{vol}_{g}\right)^{\frac{n+2}{2n}}\left(\int_{A_{r,\rho}}|\nabla v_{r,\rho}|^{2}d\operatorname{vol}_{g}\right)^{\frac{1}{2}}

As a consequence, there is a constant c2c_{2} depending on (M,g,f)(M,g,f) such that vr,ρL2nn2<c2\|v_{r,\rho}\|_{L^{\frac{2n}{n-2}}}<c_{2} and vr,ρL2<c2\|\nabla v_{r,\rho}\|_{L^{2}}<c_{2}. The standard theory of elliptic equations concludes that vr,ρv_{r,\rho} have uniformly bounded C2,αC^{2,\alpha} norms. By Arzela-Ascoli theorem we may pass to a limit and conclude that equation (5.7) has a solution vv. Then u:=1+vu:=1+v is a solution of (5.2).

We claim that uu is a positive solution. First, we show that uu is nonnegative. Otherwise, Ω:={xMu(x)<0}\Omega_{-}:=\{x\in M\mid u(x)<0\} is nonempty open set, and its boundary Ω=u1(0)\partial\Omega_{-}=u^{-1}(0) is a (n1)(n-1)-dimensional manifold, except on a closed set of lower dimension. We consider the Dirichlet boundary problem:

(5.10) {Δgufu=0,inΩ,u=0,onΩ.\begin{cases}\Delta_{g}u-fu=0,&\text{in}\;\Omega_{-},\\ u=0,&\text{on}\;\partial\Omega_{-}.\end{cases}

Similarly as showing that (5.9) has only zero solution, we can show that (5.10) has also only zero solution, i.e. u=0u=0 in Ω\Omega_{-}. This contradicts with the construction of Ω\Omega_{-}. Thus, uu must be nonnegative. Then by applying the strong maximum principle, one can see that uu must be positive everywhere.

Asymptotics of the solution: Because ff has compact support away from the singular point oo, there exits sufficiently small r0>0r_{0}>0 such that Δgvr,ρ=0\Delta_{g}v_{r,\rho}=0 on Ar,r0=Br0(o)Br(o)A_{r,r_{0}}=B_{r_{0}}(o)\setminus B_{r}(o) for all 0<r<r00<r<r_{0}. For a small ϵ>0\epsilon>0, there exists a sufficient large constant C0>0C_{0}>0 such that |vr,ρ|C0rn2ϵ|v_{r,\rho}|\leq\frac{C_{0}}{r^{n-2-\epsilon}} on Br0(o)\partial B_{r_{0}}(o) for all r>0r>0, since vr,ρv_{r,\rho} have uniformly bounded C2,αC^{2,\alpha}-norm. Note that C0C_{0} only depends on r0r_{0} and ϵ\epsilon. Moreover,

(5.11) Δg(C0r2n+ϵ)\displaystyle\Delta_{g}\left(C_{0}r^{2-n+\epsilon}\right) =\displaystyle= Δg¯(C0r2n+ϵ)+(ΔgΔg¯)(r2n+ϵ)\displaystyle\Delta_{\overline{g}}\left(C_{0}r^{2-n+\epsilon}\right)+(\Delta_{g}-\Delta_{\overline{g}})\left(r^{2-n+\epsilon}\right)
(5.12) =\displaystyle= ϵ(2n+ϵ)C0rn+ϵ+O(rn+ϵ+α),\displaystyle\epsilon(2-n+\epsilon)C_{0}r^{-n+\epsilon}+O(r^{-n+\epsilon+\alpha}),

and so we can choose r0r_{0} sufficiently small so that Δg(Cr2n+ϵ)0\Delta_{g}\left(Cr^{2-n+\epsilon}\right)\leq 0 on Br0(o)B_{r_{0}}(o). Thus, we have

(5.13) {Δg(C0r2n+ϵ±vr,ρ)0,inAr,r0,C0r2n+ϵ±vr,ρ0,onAr,r0.\begin{cases}\Delta_{g}\left(C_{0}r^{2-n+\epsilon}\pm v_{r,\rho}\right)\leq 0,&\text{in}\ \ A_{r,r_{0}},\\ C_{0}r^{2-n+\epsilon}\pm v_{r,\rho}\geq 0,&\text{on}\ \ \partial A_{r,r_{0}}.\end{cases}

Then maximum principle implies

(5.14) C0r2n+ϵ±vr,ρ0,onAr,r0,0<r<r0.C_{0}r^{2-n+\epsilon}\pm v_{r,\rho}\geq 0,\ \ \text{on}\ \ A_{r,r_{0}},\ \ \forall 0<r<r_{0}.

In other words, |vr,r0|C0r2n+ϵ|v_{r,r_{0}}|\leq C_{0}r^{2-n+\epsilon} on Ar,r0A_{r,r_{0}} for all 0<r<r00<r<r_{0}. By taking limit, we conclude that a solution vv of (5.7) satisfies

(5.15) |v|C0r2n+ϵonBr0(o).|v|\leq C_{0}r^{2-n+\epsilon}\ \ \text{on}\ \ B_{r_{0}}(o).

Similarly, we can show that for sufficiently large ρ0>0\rho_{0}>0 and C1>0C_{1}>0, the solution vv of (5.7) satisfies

(5.16) |v|C1ρ2n+ϵonMBρ0(o).|v|\leq C_{1}\rho^{2-n+\epsilon}\ \ \text{on}\ \ M\setminus B_{\rho_{0}}(o).

Consequently, we obtain that vLδ,β2v\in L^{2}_{\delta,\beta} for 2n<δ<2n+ϵ<02-n<\delta<2-n+\epsilon<0 and 2n+ϵ<β<02-n+\epsilon<\beta<0, by (3.6). Note that

(5.17) Δgv=fvfLδ2,β22,forδ>0andβ<2n,\Delta_{g}v=fv-f\in L^{2}_{\delta^{\prime}-2,\beta^{\prime}-2},\ \ \text{for}\ \ \delta^{\prime}>0\ \ \text{and}\ \ \beta^{\prime}<2-n,

since ff vanishes near the conically singular point and near the AF infinity. Then because there is no critical index between 2n2-n and 0 at either cone point or infinity, by applying Proposition 3.13 and the maximal principle, there are constants A1A_{1} and B1B_{1} such that

(5.18) v={A1ρn2+u1,asρ,B1+a(y)rν1+u2,asr0,v=\begin{cases}\frac{A_{1}}{\rho^{n-2}}+u_{1},&\text{as}\ \ \rho\to\infty,\\ B_{1}+a(y)r^{\nu_{1}}+u_{2},&\text{as}\ \ r\to 0,\end{cases}

where

(5.19) ν1:=2n+(n2)2+4λ12,\nu_{1}:=\frac{2-n+\sqrt{(n-2)^{2}+4\lambda_{1}}}{2},

here λ1\lambda_{1} is the first non-zero eigenvalue of the Laplacian ΔgN\Delta_{g^{N}} and a(y)a(y) is a corresponding eigenfunction, i.e. ΔgNa(y)=λ1a(y)\Delta_{g^{N}}a(y)=-\lambda_{1}a(y), u1Lβ2u_{1}\in L^{2}_{\beta^{\prime}} for some β<2n\beta^{\prime}<2-n, and u2Lδ2u_{2}\in L^{2}_{\delta^{\prime}} for some δ>ν1\delta^{\prime}>\nu_{1}. Note that B10B_{1}\geq 0, since vv is a nonnegative function on M{o}M\setminus\{o\}.

Moreover, because Δgv=0\Delta_{g}v=0 on Br0(o)B_{r_{0}}(o), we have:

Δgu2=Δg(B1+a(y)rν1)=(ΔgΔg¯)(a(y)rν1)=O(rν12+α)Lδ2p(Br0(o)),\Delta_{g}u_{2}=-\Delta_{g}(B_{1}+a(y)r^{\nu_{1}})=-(\Delta_{g}-\Delta_{\overline{g}})(a(y)r^{\nu_{1}})=O(r^{\nu_{1}-2+\alpha})\in L^{p}_{\delta-2}(B_{r_{0}}(o)),

for all p>1p>1 and ν1<δ<ν1+α\nu_{1}<\delta<\nu_{1}+\alpha by (3.6). Then by applying the weighted elliptic estimate in Proposition 3.4 (with p=2p=2), we obtain u2Wδ2,2(Br0(o))u_{2}\in W^{2,2}_{\delta}(B_{r_{0}}(o)) for ν1<δ<ν1+α\nu_{1}<\delta<\nu_{1}+\alpha. The weighted Sobolev embedding further implies u2Lδ2nn2(Br0(o))u_{2}\in L^{\frac{2n}{n-2}}_{\delta}(B_{r_{0}}(o)) for ν1<δ<ν1+α\nu_{1}<\delta<\nu_{1}+\alpha. Applying Proposition 3.4 (with p=2nn2p=\frac{2n}{n-2}) again gives u2Wδ2,2nn2(Br0(o))u_{2}\in W^{2,\frac{2n}{n-2}}_{\delta}(B_{r_{0}}(o)). By repeating this process, we can obtain that u2Wδ2,p(Br0(o))u_{2}\in W^{2,p}_{\delta}(B_{r_{0}}(o)) for all p>1p>1 and ν1<δ<ν1+α\nu_{1}<\delta<\nu_{1}+\alpha. Then weighted Sobolev embedding implies

(5.20) |iu2|=o(rδi),asr0,|\nabla^{i}u_{2}|=o(r^{\delta-i}),\ \ \text{as}\ \ r\to 0,

for ν1<δ<ν1+α\nu_{1}<\delta<\nu_{1}+\alpha and i=0,1i=0,1.

Similarly, we can obtain that u1u_{1} has the asymptotic at infinity as in (5.5).

Thus,

(5.21) u=1+v={1+Aρn2+u1asρ,B+a(y)rν1+u2asr0,u=1+v=\begin{cases}1+\frac{A}{\rho^{n-2}}+u_{1}\ \ &\text{as}\ \ \rho\to\infty,\\ B+a(y)r^{\nu_{1}}+u_{2}\ \ &\text{as}\ \ r\to 0,\end{cases}

where u1u_{1} and u2u_{2} have asymptotic as in (5.5) and (5.6) respectively, and B=1+B11B=1+B_{1}\geq 1. This completes the proof of the lemma. ∎

Lemma 5.3.

Let (Mn,g,o)(M^{n},g,o) be a AF manifold with a single conical singularity at oo, and {xi}i=1n\{x_{i}\}^{n}_{i=1} is the Euclidean coordinate on MM_{\infty} given by the diffeomorphism in Definition 2.2. There exist harmonic functions yiy_{i}, 1in1\leq i\leq n, on MM admitting the following asymptotic:

(5.22) yi={c+r2n+(n2)2+4λ12ϕi+O(rν), near the conical point o,xiui, near the AF infinity,y_{i}=\begin{cases}c+r^{\frac{2-n+\sqrt{(n-2)^{2}+4\lambda_{1}}}{2}}\phi_{i}+O(r^{\nu^{\prime}}),&\mbox{ near the conical point }o,\\ x_{i}-u_{i},&\mbox{ near the AF infinity},\end{cases}

where λ1\lambda_{1} is the first nonzero eigenvalue of the Laplacian ΔgN\Delta_{g^{N}} on the cross section (N,gN)(N,g^{N}) and ϕi\phi_{i} is a corresponding eigenfunction, i.e. ΔgNϕi=λ1ϕi\Delta_{g^{N}}\phi_{i}=-\lambda_{1}\phi_{i}, and

ν>2n+(n2)2+4λ12>0,\nu^{\prime}>\frac{2-n+\sqrt{(n-2)^{2}+4\lambda_{1}}}{2}>0,

and cc\in\mathbb{R} is a constant. Moreover, uiu_{i} satisfy

(5.23) |ui|+ρ|ui|+ρ2|2ui|=o(ρ1τ),asρ+,|u_{i}|+\rho|\partial u_{i}|+\rho^{2}|\partial^{2}u_{i}|=o(\rho^{1-\tau^{\prime}}),\ \ \text{as}\ \ \rho\to+\infty,

with τ=τϵ>2n2\tau^{\prime}=\tau-\epsilon>\frac{2-n}{2} for sufficiently small ϵ>0\epsilon>0.

Proof.

Let χ\chi be a cut-off function supported in MM_{\infty} and χ=1\chi=1 outside of a compact subset of MM_{\infty}. Then

Δ(χxi)={O(ρτ1),asρ+,0,near the conical pointo,\Delta(\chi x_{i})=\begin{cases}O(\rho^{-\tau-1}),&\text{as}\ \ \rho\to+\infty,\\ 0,&\text{near the conical point}\ \ o,\end{cases}

and so by (3.6)

Δ(χxi)Lδ2,β22(M),for 2n<δ<0,β>τ+1>2n.\Delta(\chi x_{i})\in L^{2}_{\delta-2,\beta-2}(M),\ \ \text{for}\ \ 2-n<\delta<0,\,\beta>-\tau+1>2-n.

Therefore, Proposition 3.15 implies that there exists uiLδ,β2(M)u_{i}\in L^{2}_{\delta,\beta}(M) for 2n<δ<0,β>τ+1>2n2-n<\delta<0,\,\beta>-\tau+1>2-n such that

Δui=Δ(χxi).\Delta u_{i}=\Delta(\chi x_{i}).

By letting yi=χxiuiy_{i}=\chi x_{i}-u_{i}, we obtain harmonic functions yiy_{i}, for 1in1\leq i\leq n.

Then note that near the conical point oo, yi=uiy_{i}=-u_{i}, and similarly as in the proof of Lemma 5.2 in [5], applying Proposition 3.13 gives the asymptotic of yiy_{i} near the conical point as in (5.22). In addition, similarly as in the proof of Lemma 5.1 in [5], applying a Nash-Moser iteration argument gives the asymptotic of uiu_{i} near infinity as in (5.23). ∎

Now we are ready to prove the rigidity result in Theorem 1.1 as following.

Theorem 5.4.

Let (Mn,g,o)(M^{n},g,o) be an AF manifold with a single conical singularity at oo. If the scalar curvature Scg0{\rm Sc}_{g}\geq 0 and the ADM mass m(g)=0m(g)=0, then (Mn,g)(M^{n},g) is isometric to the Euclidean space (n,gn)(\mathbb{R}^{n},g_{\mathbb{R}^{n}}).

Proof.

First we show that if Scg0{\rm Sc}_{g}\geq 0 and m(g)=0m(g)=0 then gg is Ricci flat. For that, we take a variation gt=g+thg_{t}=g+th of the metric gg, where hh is an arbitrary compactly supported 22-tensor on M̊\mathring{M}. Note that for each tt such that |t||t| is sufficiently small, gtg_{t} is still an AF manifold with a single conical singularity at oo. Then we consider the following equation:

(5.24) {Δgtu+n24(n1)(ScgtScg)u=0,onM̊,u1, as x.\begin{cases}-\Delta_{g_{t}}u+\frac{n-2}{4(n-1)}\left({\rm Sc}_{g_{t}}-{\rm Sc}_{g}\right)u=0,&\text{on}\ \ \mathring{M},\\ u\to 1,&\mbox{ as }x\to\infty.\end{cases}

For each tt such that |t||t| is sufficiently small, by Lemma 5.2, the equation (5.24) has a positive solution utu_{t}, since hh is compactly supported and so is ScgtScg{\rm Sc}_{g_{t}}-{\rm Sc}_{g}, and (ScgtScg)Ln2(M)\left\|\left({\rm Sc}_{g_{t}}-{\rm Sc}_{g}\right)\right\|_{L^{\frac{n}{2}}(M)} can be arbitrary small as |t||t| sufficiently small. Therefore, for each tt such that |t||t| is sufficiently small, by the asymptotic of utu_{t} as in (5.3), g~t:=ut4n2gt\tilde{g}_{t}:=u^{\frac{4}{n-2}}_{t}g_{t} is a AF metric with a conically singular point at oo as in Definition 2.2. Because utu_{t} is a solution of (5.24), we have

Scg~t=4(n1)n2(ut)n+2n2(Δgtut+n24(n1)Scgtut)=(ut)4n2Scg0.{\rm Sc}_{\tilde{g}_{t}}=\frac{4(n-1)}{n-2}(u_{t})^{-\frac{n+2}{n-2}}\left(-\Delta_{g_{t}}u_{t}+\frac{n-2}{4(n-1)}{\rm Sc}_{g_{t}}u_{t}\right)=(u_{t})^{-\frac{4}{n-2}}{\rm Sc}_{g}\geq 0.

Thus, Theorem 4.2 implies that the mass m(g~t)0m(\tilde{g}_{t})\geq 0, and so t=0t=0 is a interior minimum point of the function m(g~t)m(\tilde{g}_{t}), since u01,g0=gu_{0}\equiv 1,g_{0}=g and m(g)=0m(g)=0. As a result,

0=dm(g~t)dt|t=0=MhjkRicjkdvolg,\left.0=\frac{dm(\tilde{g}_{t})}{dt}\right|_{t=0}=\int_{M}h^{jk}{\operatorname{Ric}}_{jk}d\operatorname{vol}_{g},

which implies that gg is Ricci flat, since hh is arbitrary.

Now we prove that the cross section (N,gN)(N,g^{N}) of the conical neighborhood of the singular point is the standard sphere. The idea is to compute the mass, m(g)m(g), using the harmonic coordinate, yiy_{i}, obtained in Lemma 5.3, with the help of Ricci-flatness of gg that we have already obtained. As a consequence, m(g)=0m(g)=0 then implies that dyidy_{i} are parallel 1-forms, and this then enables us to apply Obata’s rigidity theorem for (N,gN)(N,g^{N}).

Note that for an asymptotically conical metric, gg, as in Definition 2.1, its Ricci tensor, Ricg\operatorname{Ric}_{g}, satisfies

Ricg(X,Y)=RicgN(X,Y)(n2)gN(X,Y)+O(rα),asr0,\operatorname{Ric}_{g}(X,Y)=\operatorname{Ric}_{g^{N}}(X,Y)-(n-2)g^{N}(X,Y)+O(r^{\alpha}),\ \ \text{as}\ \ r\to 0,

for all X,YX,Y tangent to the cross section NN. Recall that the decay order α>0\alpha>0. Thus, Ricg=0{\operatorname{Ric}}_{g}=0, implies RicgN=(n2)gN\operatorname{Ric}_{g^{N}}=(n-2)g^{N} on the cross section. As a result, the Lichnerowicz eigenvalue estimate implies that the first nonzero eigenvalue of ΔgN\Delta_{g^{N}}, λ1n1\lambda_{1}\geq n-1.

Let yi,1in,y_{i},1\leq i\leq n, be harmonic functions obtained in Lemma 5.3. By the asymptotic of yiy_{i} at infinity as in (5.23), the argument as in the proof of Theorem 4.4 in [2] implies that we can compute the mass, m(g)m(g), using yiy_{i} as coordinate on MM_{\infty}. In addition, by Bochner formula, recall that gg is Ricci flat, we have

(5.25) Δg(12|dyi|2)=|dyi|2.\displaystyle\Delta_{g}\left(\frac{1}{2}|dy_{i}|^{2}\right)=|\nabla dy_{i}|^{2}.

Then by integrating by part, we have

ωnm(g)\displaystyle\omega_{n}m(g) =\displaystyle= limρ[i=1nBρνρ(12|dyi|2)]\displaystyle\lim_{\rho\to\infty}\left[\sum_{i=1}^{n}\int_{\partial B_{\rho}}\partial_{\nu_{\rho}}\left(\frac{1}{2}|dy_{i}|^{2}\right)\right]
=\displaystyle= limr0[i=1nBrνr(12|dyi|2)]+limr0,ρ[i=1nAr,ρ(|dyi|2)]\displaystyle\lim_{r\to 0}\left[\sum_{i=1}^{n}\int_{\partial B_{r}}\partial_{\nu_{r}}\left(\frac{1}{2}|dy_{i}|^{2}\right)\right]+\lim_{r\to 0,\rho\to\infty}\left[\sum_{i=1}^{n}\int_{A_{r,\rho}}\left(|\nabla dy_{i}|^{2}\right)\right]
=\displaystyle= i=1nM(|dyi|2)\displaystyle\sum_{i=1}^{n}\int_{M}\left(|\nabla dy_{i}|^{2}\right)

where νρ\nu_{\rho} and νr\nu_{r} are the outer normal vector of Bρ\partial B_{\rho} and Br\partial B_{r} respectively, and ωn\omega_{n} is the volume of the unit sphere in n\mathbb{R}^{n}. Here in the last step, we use that the asymptotic behavior of yiy_{i} near the conical point as in (5.22) and the fact λ1n1\lambda_{1}\geq n-1. As a result, m(g)=0m(g)=0 implies that dyi=0\nabla dy_{i}=0, i.e. dyidy_{i} is parallel. Then |dyi|=1|dy^{i}|=1, since dui=O(ρτ)du_{i}=O(\rho^{-\tau}). Thus, by considering the asymptotic behavior of dyidy_{i} near the conical point, the exponent in (5.22)(\ref{eqn-asymptotic-of-harmonic-coordinate}):

(5.26) 2n+(n2)2+4λ12=1,\frac{2-n+\sqrt{(n-2)^{2}+4\lambda_{1}}}{2}=1,

because otherwise |dyi||dy_{i}| will tend to either zero or infinity as r0r\to 0, i.e. as approaching to the conical point. Therefore, by solving (5.26), we obtain that λ1=n1\lambda_{1}=n-1 is an eigenvalue of ΔgN\Delta_{g^{N}}. Then we apply the Obata’s rigidity theorem to conclude that (N,gN)(N,g^{N}) must be the standard sphere, since we have shown that RicgN=(n2)gNRic_{g^{N}}=(n-2)g^{N}. As a consequence, the conically singular point oo is actually a manifold point, and our metric gg is continuous and has W1,nW^{1,n}-regularity near the conical singularity point, and so it satisfies the regularity assumption in [11]. By Theorem 1.1 in [11], we conclude that (Mn,g)(n,gn)(M^{n},g)\cong(\mathbb{R}^{n},g_{\mathbb{R}^{n}}). ∎

Remark 5.5.

In [5], for the proof of rigidity, after we get the manifold is Ricci flat, then we can construct the harmonic coordinate to conclude that the manifold is the Euclidean space as we do in the above.

References

  • [1] K. Akutagawa and B. Botvinnik. Yamabe metrics on cylindrical manifolds. Geom. Funct. Anal., 13(2):259–333, 2003.
  • [2] Robert Bartnik. The mass of an asymptotically flat manifold. Comm. Pure Appl. Math., 39(5):661–693, 1986.
  • [3] Hubert L. Bray and Jeffery L. Jauregui. A geometric theory of zero area singularities in general relativity. Asian J. Math., 17(3):525–560, 2013.
  • [4] Simone Cecchini and Rudolf Zeidler. The positive mass theorem and distance estimates in the spin setting. arXiv:2108.11972, 2021.
  • [5] Xianzhe Dai, Yukai Sun, and Changliang Wang. Positive mass theorem for asymptotically flat spin manifolds with isolated conical singularities. arXiv:2310.13285, 2023.
  • [6] Xianzhe Dai and Changliang Wang. Perelman’s W-functional on manifolds with conical singularities. Math. Res. Lett., 27(3):665–685, 2020.
  • [7] James D.E. Grant and Nathalie Tassotti. A positive mass theorem for low-regularity riemannian metrics. arXiv:1408.6425, pages 1–21, 2014.
  • [8] Marc Herzlich. A Penrose-like inequality for the mass of Riemannian asymptotically flat manifolds. Comm. Math. Phys., 188(1):121–133, 1997.
  • [9] Marc Herzlich. Minimal surfaces, the Dirac operator and the Penrose inequality. Séminaire de Théorie Spectrale et Géométrie, Univ. Grenoble I, Saint-Martin-d’Hères, 20:9–16, 2002.
  • [10] Sven Hirsch and Pengzi Miao. A positive mass theorem for manifolds with boundary. Pac. J. Math, 306(1):185–201, 2020.
  • [11] Wenshuai Jiang, Weimin Sheng, and Huaiyu Zhang. Removable singularity of positive mass theorem with continuous metrics. Math. Z., 302(2):839–874, 2022.
  • [12] Tao Ju and Jeff Viaclovsky. Conformally prescribed scalar curvature on orbifolds. Comm. Math. Phys., 398(2):877–923, 2023.
  • [13] Dan Lee. A positive mass theorem for Lipschitz metrics with small singular sets. Proc. Am. Math. Soc., 14:3997–4004, 2013.
  • [14] Dan A. Lee and Philippe G. LeFloch. The positive mass theorem for manifolds with distributional curvature. Comm. Math. Phys., 339(1):99–120, 2015.
  • [15] Dan A. Lee, Martin Lesourd, and Ryan Unger. Density and positive mass theorems for incomplete manifolds. Calc. Var. Partial Differential Equations, 62(7):Paper No. 194, 23, 2023.
  • [16] Martin Lesourd, Ryan Unger, and Shing-Tung Yau. The positive mass theorem with arbitrary ends. To appear in J. Differential Geometry, 2021.
  • [17] Chao Li and Christos Mantoulidis. Positive scalar curvature with skeleton singularities. Math. Ann., 374:99–131, 2019.
  • [18] James Lucietti. All higher-dimensional Majumdar-Papapetrou black holes. Ann. Henri Poincaré, 22(7):2437–2450, 2021.
  • [19] Donovan McFeron and Gábor Székelyhidi. On the positive mass theorem for manifolds with corners. Comm. Math. Phys., 313(2):425–443, 2012.
  • [20] Pengzi Miao. Positive mass theorem on manifolds admitting corners along a hypersurface. Adv. Theor. Math. Phys., 6(6):1163–1182, 2002.
  • [21] Pengzi Miao. Implications of some mass-capacity inequalities. arXiv:2307.06428v2, 2023.
  • [22] Pengzi Miao. Mass, capacitary functions, and the mass-to-capacity ratio. Peking Mathematical Journal, 2023.
  • [23] Vincent Minerbe. A mass for ALF manifolds. Comm. Math. Phys., 289(3):925–955, 2009.
  • [24] Richard Schoen. Conformal deformation of a Riemannian metric to constant scalar curvature. J. Differential Geometry, 20:479–495, 1984.
  • [25] Richard Schoen and Shing Tung Yau. On the proof of the positive mass conjecture in general relativity. Comm. Math. Phys., 65:45–76, 1979.
  • [26] Richard Schoen and Shing Tung Yau. Positive scalar curvature and minimal hypersurface singularities. Surveys in Differential Geometry, XXIV:441–480, 2021.
  • [27] Yuguang Shi and Luen-Fai Tam. Scalar curvature and singular metrics. Pacific J. Math., 293(2):427–470, 2018.
  • [28] Edward Witten. A new proof of the positive energy theorem. Comm. Math. Phys., 80:381–402, 1981.
  • [29] Jintian Zhu. Positive mass theorem with arbitrary ends and its application. Int. Math. Res. Not. IMRN, (11):9880–9900, 2023.