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

Scattering on singular Yamabe spaces

Sun-Yung Alice Chang*, Stephen E. McKeown**, and Paul Yang*,† In honor of Antonio Córdoba and José Luis Fernández Department of Mathematics, Princeton University, Princeton, NJ 08544, USA Department of Mathematical Sciences, University of Texas at Dallas, 800 W. Campbell Road, Richardson, TX 75080, USA Department of Mathematics, Princeton University, Princeton, NJ 08544, USA [email protected] [email protected] [email protected]
Abstract.

We apply scattering theory on asymptotically hyperbolic manifolds to singular Yamabe metrics, applying the results to the study of the conformal geometry of compact manifolds with boundary. In particular, we define extrinsic versions of the conformally invariant powers of the Laplacian, or GJMS operators, on the boundary of any such manifold, along with associated extrinsic QQ-curvatures. We use the existence and uniqueness of a singular Yamabe metric conformal to define also nonlocal extrinsic fractional GJMS operators on the boundary, and draw other global conclusions about the scattering operator, including a Gauss-Bonnet theorem in dimension four.

2020 Mathematics Subject Classification:
Primary: 53A31, 53C18. Secondary: 53A55, 53C40, 58J50
*Partially supported by NSF grant DMS-1607091. **Partially supported by NSF RTG grant DMS-1502525. †Partially supported by Simons Foundation grant 615589.

1. Introduction

Scattering theory on asymptotically hyperbolic manifolds has been studied with great profit in the case that the metric is Einstein. In this paper we study the scattering problem in the case that the metric has constant scalar curvature. For every compact Riemannian manifold (Xn+1,g¯)(X^{n+1},\bar{g}) with boundary MM, it is known that there is precisely one defining function uu for the boundary so that the singular Yamabe metric g=u2g¯g=u^{-2}\bar{g} has constant scalar curvature, so this study can be seen as bringing scattering theory to bear as a tool in the study of the conformal geometry of compact manifolds with boundary. We apply the methods of [GZ03, FG02, CQY08] and others to the setting of [GW15, Gra17], which used the singular Yamabe problem to study conformal hypersurfaces.

Given an asymptotically hyperbolic (AH) manifold (Xn+1,g)(X^{n+1},g) with boundary Mn=XM^{n}=\partial X, the scattering problem is defined as follows. Let g¯=r2g\bar{g}=r^{2}g be a geodesic compactification of gg, i.e., suppose that |dr|g¯=1|dr|_{\bar{g}}=1 on a neighborhood of MM. Let fC(M)f\in C^{\infty}(M). Consider the equation

(1.1) (Δg+s(ns))v=0,(\Delta_{g}+s(n-s))v=0,

where we assume that s>n2s>\frac{n}{2} and that s(sn)σpp(Δg)s(s-n)\notin\sigma_{pp}(\Delta_{g}) (the set of L2L^{2} eigenvalues of the Laplacian); in our convention, Δg\Delta_{g} is a negative operator. It is shown in [MM87] that a solution vv to this equation has asymptotics

v=rnsF+rsG,v=r^{n-s}F+r^{s}G,

where F,GC(X)F,G\in C^{\infty}(X), at least so long as 2sn2s-n\notin\mathbb{Z}. In [GZ03], it is shown that we can always solve (1.1) with F|M=fF|_{M}=f. The scattering operator S(s):C(M)C(M)S(s):C^{\infty}(M)\to C^{\infty}(M) is then defined by S(s)f=G|MS(s)f=G|_{M}, and this operator extends to be meromorphic on Resn2\operatorname{Re}s\geq\frac{n}{2}, with poles only at ss such that s(ns)s(n-s) is an eigenvalue of Δg\Delta_{g} or at s=n+q2s=\frac{n+q}{2}, qq\in\mathbb{N}.

The case of gg Einstein has been thoroughly studied. In that case, the poles for s=n+q2s=\frac{n+q}{2} with qq odd actually do not exist – S(s)S(s) is regular at these points – while at s=n2+js=\frac{n}{2}+j, the operator S(s)S(s) has a simple pole whose residue is the so-called GJMS operator P2jP_{2j} (with jn/2j\leq n/2 if the boundary dimension nn is even). If nn is even, S(s)1S(s)1 is holomorphic across s=ns=n, and in fact S(n)1=cnQnS(n)1=c_{n}Q_{n}, where QnQ_{n} is the nnth-order QQ-curvature and cnc_{n} is a universal constant. On the other hand, if nn is odd, then S(n)1=0S(n)1=0, which reflects the fact that there is no (locally-defined) QQ-curvature of odd order. However, a nonlocal QQ-like term was defined in [FG02] by taking Qn=kndds|s=nS(s)1Q_{n}=k_{n}\frac{d}{ds}|_{s=n}S(s)1 in case nn is odd. In addition to these results, there have been numerous interesting results linking the scattering operator with the renormalized volume of an AH Einstein manifold.

We recall the definition of the renormalized volume in that setting. Let rr be a geodesic defining function for MM, and consider the expansion

(1.2) r>ε𝑑vg=c0εn+c1ε1n++cn1ε1+log(1ε)+V++o(1).\int_{r>\varepsilon}dv_{g}=c_{0}\varepsilon^{-n}+c_{1}\varepsilon^{1-n}+\cdots+c_{n-1}\varepsilon^{-1}+\mathcal{E}\log\left(\frac{1}{\varepsilon}\right)+V_{+}+o(1).

The quantity V+V_{+} is called the renormalized volume. While a priori it depends on the choice of geodesic defining function rr, it is shown in [HS98, Gra00] that, if nn is odd, then =0\mathcal{E}=0 and V+V_{+} is in fact well-defined independent of rr. If the boundary dimension nn is even, on the other hand, then \mathcal{E} is well-defined independent of rr. Using scattering theory, it was shown in [GZ03, FG02] that in the even case, =MQn𝑑vk\mathcal{E}=\int_{M}Q_{n}dv_{k} (where k=g¯|TMk=\bar{g}|_{TM}). In [CQY08], a formula was given relating the renormalized volume in the even-nn case to an integral over MM of a non-local quantity defined in terms of the scattering operator.

It was shown already in [Gra17] that the renormalized volume can be defined as well in the singular Yamabe case, where rr now is the g¯\bar{g}-distance to the boundary and gg is the corresponding singular Yamabe metric; but in this case, \mathcal{E} is generically nonvanishing in all dimensions and is a conformal invariant, while V+V_{+} is no longer invariant. In fact, in the case n=2n=2, \mathcal{E} is simply a linear combination of the Willmore energy and the Euler characteristic. Thus, in case n>2n>2, \mathcal{E} can be seen as a generalization of the Willmore energy.

We generalize the scattering picture to the singular Yamabe setting, and obtain extrinsic analogues to all of the above results. Unlike the Einstein setting, where the existence and uniqueness of Einstein metrics remain extremely challenging and one is often constrained in practice to use a non-unique formal expansion, the analytic picture in the singular Yamabe case is well understood and perhaps optimal: given any compact Riemannian manifold (Xn+1,g¯)(X^{n+1},\bar{g}) with boundary, there is a unique defining function uu for the boundary so that the AH metric g=u2g¯g=u^{-2}\bar{g} has constant scalar curvature n(n+1)-n(n+1). Any global quantities defined in terms of this singular Yamabe metric are thus well defined for (X,g¯)(X,\bar{g}). This comes with a price, however: the power series expansion of uu at the boundary MM is not smooth beyond order n+1n+1, after which log\log terms appear. Throughout, therefore, we must keep careful track of the expansion of uu and its regularity in a way that has not had to be done in the Einstein case, where smooth formally Einstein metrics have largely been considered.

Also unlike AH Einstein spaces, singular Yamabe metrics have no particular parity properties, and this has numerous consequences: in general, this theory tends to behave more like the even-dimensional-boundary Einstein theory in all dimensions than like the odd-dimensional Einstein theory, whose special behaviors are entirely a product of parity considerations. In some cases, though, even features of the even-dimensional Einstein case are not reproduced. A first consequence, for example, is that S(s)S(s) generically has poles at s=n+q2s=\frac{n+q}{2} for all qq (not only even qq). Therefore, we get residue operators PqSY=ress=n/2+q/2S(s)P_{q}^{SY}=\operatorname{res}_{s=n/2+q/2}S(s) for all qnq\leq n. Of course, since Einstein metrics are the singular Yamabe metrics in their conformal class, our results must in all case reduce to the usual ones in that special case. Similarly, for all nn – not only even – we can define QnSY=cn1S(n)1Q^{SY}_{n}=c_{n}^{-1}S(n)1. Here, QnSYQ^{SY}_{n} is a locally determined curvature quantity along MM, which however depends on g¯\bar{g} and not merely g¯|TM\bar{g}|_{TM}. Similarly, for each qq, PqSYP_{q}^{SY} is a conformally covariant differential operator along MM whose coefficients depend on extrinsic data.

We now state our main results.

Theorem A.

Let (Xn+1,g¯)(X^{n+1},\bar{g}) be a compact Riemannian manifold with boundary (Mn,k)(M^{n},k), and let gg be the singular Yamabe metric associated to g¯\bar{g}. Let L̊\mathring{L} be the tracefree second fundamental form of MM with respect to XX, computed with inward-pointing unit normal vector. Then the scattering operator S(s):C(M)C(M)S(s):C^{\infty}(M)\to C^{\infty}(M) extends to a meromorphic family on the strip 12<s<n+12-\frac{1}{2}<s<n+\frac{1}{2}, regular for Res=n2\operatorname{Re}s=\frac{n}{2}. It has poles only at s=n+q2s=\frac{n+q}{2} (for qq\in\mathbb{N}) and for such ss that s(sn)σpp(Δg)s(s-n)\in\sigma_{pp}(\Delta_{g}), the set of L2L^{2} eigenvalues of Δg\Delta_{g}.

Suppose that qq\in\mathbb{N} satisfies 2qn2\leq q\leq n and that q2n24σpp(Δg)\frac{q^{2}-n^{2}}{4}\notin\sigma_{pp}(\Delta_{g}). Then there is a conformally covariant differential operator PqSY:C(M)C(M)P_{q}^{SY}:C^{\infty}(M)\to C^{\infty}(M), satisfying

cqPqSY=ress=n+q2S(s),c_{q}P_{q}^{SY}=-\operatorname{res}_{s=\frac{n+q}{2}}S(s),

where cqc_{q} is a nonvanishing universal constant, and PqP_{q} has principal part (Δk)q(-\Delta_{k})^{q} if qq is even and, if q>1q>1 is odd, has the same principal part as L̊μνμkνk(Δk)q32\mathring{L}^{\mu\nu}\nabla_{\mu}^{k}\nabla^{k}_{\nu}(\Delta_{k})^{\frac{q-3}{2}}. The operator PqSYP_{q}^{SY} depends only on the jet of g¯\bar{g} at MM.

If g¯~=e2ωg¯\tilde{\bar{g}}=e^{2\omega}\bar{g} is a conformally-related metric with corresponding operators P~qSY\widetilde{P}_{q}^{SY}, then

P~qSYf=e(n+q)ω2PqSY(e(nq)ω2f).\widetilde{P}_{q}^{SY}f=e^{-\frac{(n+q)\omega}{2}}P_{q}^{SY}\left(e^{\frac{(n-q)\omega}{2}}f\right).

Several remarks are in order regarding this statement.

First, note that in the Einstein case considered by [GZ03], the integer qq is assumed to be even, since otherwise PqSYP_{q}^{SY} vanishes identically in that setting. Moreover, the indexing of our operators PqSYP_{q}^{SY} differs from the indexing in that paper by a factor of two – what we call P2P_{2} was there P1P_{1}, etc. This is because in the other case, what are here the odd-order operators vanish. Already in the literature, there is some difference in the numbering of these operators. The operators defined in [GZ03], of course (unlike those here) are the GJMS operators. The operators defined here differ from the GJMS operators by terms depending on the extrinsic geometry of MM in XX.

Next, since in the Einstein case, the odd-order operators vanish, the notation PqP_{q} for qq odd has there developed another meaning that, unfortunately, is not analogous to that used here: see [FG02]. Given the strong analogy between the even-order operators and those in the Einstein case, and the equivalence of definition here of the even- and odd-order operators, it seems there is no very good way around having confusing notation. Observe also that the odd-order operators defined in those settings do not exist in this setting, since their definition depends on the vanishing of the residue at s=n+q2s=\frac{n+q}{2}. A similar remark applies to the definition, below, of QnSYQ_{n}^{SY} for odd nn; again, see [FG02].

As discussed more fully in section 4, the restriction on ss in the statement of the meromorphicity of the scattering operator is due to the non-regularity at order n+2n+2 of the singular Yamabe solution uu. In fact, stronger statements could be made, but the nonregularity makes it somewhat delicate to discuss and define just what this means. Since we are not interested behavior at higher ss in any event, we give the weaker statement.

Finally, in work such as [CdMG11], it is common to use the notation s=n2+γs=\frac{n}{2}+\gamma. The relationship between qq and γ\gamma, then, is q=2γq=2\gamma.

Corollary B.

The operators PqSYP_{q}^{SY} are self-adjoint.

This follows from their definition in terms of the scattering operator.

We now define the QnSYQ_{n}^{SY}-curvature, as in [GZ03], by QnSY=cn1S(n)1Q_{n}^{SY}=c_{n}^{-1}S(n)1. This is well-defined since kerPnSY\mathbb{R}\subseteq\ker P_{n}^{SY}. This extrinsic QQ-curvature quantity follows a conformal transformation law like its intrinsic cousin (which however exists only for even order) and, like it, gives the conformally invariant term in the volume expansion:

Theorem C.

If \mathcal{E} is as in (1.2), interpreted as in [Gra17], then

=MQnSY𝑑vk.\mathcal{E}=\oint_{M}Q_{n}^{SY}dv_{k}.

Furthermore, if g¯~=e2ωg¯\tilde{\bar{g}}=e^{2\omega}\bar{g}, then

enωQ~nSY=QnSY+PnSYω.e^{n\omega}\widetilde{Q}_{n}^{SY}=Q_{n}^{SY}+P_{n}^{SY}\omega.

A corollary then follows immediately from Theorem 3.1 of [Gra17]. To state it, we must say slightly more about the singular Yamabe function. Given g¯\bar{g}, there is a unique defining function uu for MM such that g=u2g¯g=u^{-2}\bar{g} has scalar curvature n(n+1)-n(n+1). As discussed above and in [Gra17], uu has the expansion (in terms of the distance function rr from the boundary MM in XX)

(1.3) u=r+u2r2+u3r3++un+1rn+1+rn+2log(r)+un+2rn+2+o(rn+2),u=r+u_{2}r^{2}+u_{3}r^{3}+\cdots+u_{n+1}r^{n+1}+\mathcal{L}r^{n+2}\log(r)+u_{n+2}r^{n+2}+o(r^{n+2}),

where \mathcal{L} is a locally and extrinsically defined function on MM, conformally invariant of weight (n+1)-(n+1). With this notation fixed, we have the following:

Corollary D.

Let (X,g¯)(X,\bar{g}) be a Riemmanian manifold, and let Ft:MX,0t<δF_{t}:M\hookrightarrow X,0\leq t<\delta, be a smoothly varying variation of MM, where F0F_{0} is the identity. Let Mt=Ft(M)M_{t}=F_{t}(M), and n¯\overline{n} the inward-pointing unit normal to MM in XX. Then

ddt|t=0MtQtSY𝑑vMt=(n+2)(n1)MF˙,n¯g¯𝑑vk,\left.\frac{d}{dt}\right|_{t=0}\oint_{M_{t}}Q^{SY}_{t}dv_{M_{t}}=(n+2)(n-1)\oint_{M}\langle\dot{F},\overline{n}\rangle_{\bar{g}}\mathcal{L}dv_{k},

where \mathcal{L} is as in (1.3).

The results so far stated about PqSYP_{q}^{SY} and QnSYQ_{n}^{SY} have actually appeared previously in the recent literature. They were derived in a somewhat different framework by Gover and Waldron ([GW14, GW15, GW17]), where they are defined in terms of the tractor calculus and so-called Laplace-Robin operators. When the current paper was quite advanced in preparation, the paper [JO21] of Juhl and Orsted was brought to our attention. Among other things, it reinterprets the results of Gover and Waldron in the setting of scattering theory shared by this paper. Our contribution with respect to this part of our material is thus a different treatment, more similar in spirit to [GZ03, FG02, CQY08]. On the other hand, our perspective here is also focused more than those papers on the unique global scattering operator associated to every compact Riemmanian manifold with boundary. In particular, rather than doing only asymptotic analysis at the boundary and thus neglecting the impact of the non-regularity of the singular Yamabe solution, we take account of the logarithmic term. The following results, more global in spirit, are new to this paper.

First, as stated above, S(s)1S(s)1 is smooth across s=ns=n. We now define

(1.4) 𝒮=dds|s=nS(s)1C(M).\mathscr{S}=\left.\frac{d}{ds}\right|_{s=n}S(s)1\in C^{\infty}(M).

This is a function dependent on the global geometry of (X,g¯)(X,\bar{g}). Next, it follows from [GZ03] that for ss\in\mathbb{C} near 1, there exists us=rnsFs+rsGsu_{s}=r^{n-s}F_{s}+r^{s}G_{s} satisfying (Δg+s(ns))us=0(\Delta_{g}+s(n-s))u_{s}=0 and Fs|M=1F_{s}|_{M}=1. Moreover, FsF_{s} and GsG_{s} may be uniquely determined by the requirement that they be holomorphic in ss. We then write

Fs(r)=1+a1(s)r+a2(s)r2++an(s)rn+,F_{s}(r)=1+a_{1}(s)r+a_{2}(s)r^{2}+\cdots+a_{n}(s)r^{n}+\cdots,

where each aja_{j} is a smooth function on MM, and aja_{j} is locally (and extrinsically) determined for 1jn1\leq j\leq n. Finally, we recall the definition in [Gra17] (see also [Gra00]) of the renormalized volume coefficients. Since g=u2g¯g=u^{-2}\bar{g} for a defining function uu for the boundary, we may write

dvg=r1n(1+v(1)r+v(2)r2+)drdvkdv_{g}=r^{-1-n}(1+v^{(1)}r+v^{(2)}r^{2}+\cdots)drdv_{k}

on a collar neighborhood [0,ε)r×M[0,\varepsilon)_{r}\times M near MM in XX. The functions v(j)C(M)v^{(j)}\in C^{\infty}(M) are the renormalized volume coefficients. With these notations in hand, we may state our next result, which is analogous to a theorem proved in [CQY08] (based on [FG02]) in the Einstein setting.

Theorem E.

Suppose (Xn+1,g¯)(X^{n+1},\bar{g}) is a Riemannian manifold with boundary, and that g=u2g¯g=u^{-2}\bar{g} is the associated singular Yamabe metric. Let V(X,g,g¯)V(X,g,\bar{g}) be the renormalized volume of gg when computed with respect to g¯\bar{g}, i.e. the constant term in the expansion (2.2). Then

(1.5) V(X,g,g¯)=𝒮𝑑vk1n(Ma1v(n1)dvk+2Ma2v(n2)dvk++(n1)Man1v(1)dvk+nMandvk),\begin{split}V(X,g,\bar{g})=&-\oint\mathscr{S}dv_{k}\\ &\quad-\frac{1}{n}\left(\oint_{M}a_{1}^{\prime}v^{(n-1)}dv_{k}+2\oint_{M}a_{2}^{\prime}v^{(n-2)}dv_{k}+\cdots+\right.\\ &\quad\left.(n-1)\oint_{M}a^{\prime}_{n-1}v^{(1)}dv_{k}+n\oint_{M}a_{n}^{\prime}dv_{k}\right),\end{split}

where a prime denotes dds|s=n\frac{d}{ds}|_{s=n}.

All the terms on the right-hand side except the first are local.

It was shown in [FG02] that, if gg is in fact Einstein and nn is odd, then this formula holds with only the first term on the right-hand side present. Although, in such a case, gg is also the singular Yamabe metric and thus (1.5) applies, the results are not inconsistent, because in that case, the v(k)v^{(k)} with kk odd and the aka_{k} with kk odd are both identically zero, and one of these is a factor in each of the integrands multiplied above by 1n\frac{1}{n} when nn is odd.

The following theorem, which was proved in [CQY08] in the the special case of even-dimensional Einstein metrics, states (in all dimensions) that the scattering term in (1.5) is a conformal primitive for total QSYQ^{SY}.

Theorem F.

Let g¯\bar{g} be a smooth metric on (Xn+1,Mn)(X^{n+1},M^{n}), and gg the corresponding singular Yamabe metric. As usual, let k=g¯|TMk=\bar{g}|_{TM}. Suppose ωC(X)\omega\in C^{\infty}(X). Then

ddα|α=0M𝒮e2αωg¯𝑑ve2αωk=2cnMQnSYω𝑑vk.\left.\frac{d}{d\alpha}\right|_{\alpha=0}\oint_{M}\mathscr{S}_{e^{2\alpha\omega}\bar{g}}dv_{e^{2\alpha\omega}k}=-2c_{n}\oint_{M}Q^{SY}_{n}\omega dv_{k}.

Finally, for XX of dimension four, we apply Theorem E and the main result of [GG19] to obtain a Gauss-Bonnet theorem in terms of the scattering operator.

Theorem G.

Let (X4,g¯)(X^{4},\bar{g}) be a compact Riemannian manifold with boundary M3=XM^{3}=\partial X, and gg the singular Yamabe metric. Let rr be the g¯\bar{g}-distance function to the boundary, EE the Einstein tensor of XX, WW the Weyl tensor, and 𝒮\mathscr{S} as in (1.4). Then

8π2χ(X)\displaystyle 8\pi^{2}\chi(X) =14X|W|g2𝑑vg12f.p.r>ε|E|g2𝑑vg+M(6𝒮+𝒞)𝑑vk,\displaystyle=\frac{1}{4}\int_{X}|W|_{g}^{2}dv_{g}-\frac{1}{2}f.p.\int_{r>\varepsilon}|E|_{g}^{2}dv_{g}+\oint_{M}\left(-6\mathscr{S}+\mathcal{C}\right)dv_{k},

where

𝒞\displaystyle\mathcal{C} =1136HR+1108HR¯+5108H3+389144H|L̊|k2+14μνL̊μν\displaystyle=-\frac{11}{36}HR+\frac{1}{108}H\overline{R}+\frac{5}{108}H^{3}+\frac{389}{144}H|\mathring{L}|_{k}^{2}+\frac{1}{4}\nabla^{\mu}\nabla^{\nu}\mathring{L}_{\mu\nu}
+236L̊μνR¯μν173L̊μνRμν+112rR¯23L̊3.\displaystyle\quad+\frac{23}{6}\mathring{L}^{\mu\nu}\overline{R}_{\mu\nu}-\frac{17}{3}\mathring{L}^{\mu\nu}R_{\mu\nu}+\frac{1}{12}\partial_{r}\overline{R}-\frac{2}{3}\mathring{L}^{3}.

Here f.p.f.p. denotes the finite part of the integral as ε0\varepsilon\to 0, LL the second fundamental form of MM, H=kμνLμνH=k^{\mu\nu}L_{\mu\nu} its mean curvature, and R,R¯,Rμν,R¯μνR,\overline{R},R_{\mu\nu},\overline{R}_{\mu\nu} the scalar and Ricci curvatures of, respectively, MM and XX.

The following corollary also follows from Corollary 1.4 of that paper, or from the previous result.

Corollary H.

Let (X4,g¯)(X^{4},\bar{g}) be a compact Riemannian manifold with umbilic boundary M3=XM^{3}=\partial X. Then the quantity

V~=M(𝒮+16𝒞)𝑑vk\widetilde{V}=\oint_{M}\left(-\mathscr{S}+\frac{1}{6}\mathcal{C}\right)dv_{k}

is a conformal invariant.

We point out also that having a uniquely defined scattering operator, as stated in Theorem A, allows one to define unique fractional extrinsic GJMS operators P2γSY=S(n2+γ)P_{2\gamma}^{SY}=S(\frac{n}{2}+\gamma) of order 2γ2\gamma as well, as in [CdMG11]. In the intrinsic case, such operators are unique only when a global Einstein metric can be found. As there, these operators are nonlocal.

The paper is organized as follows. In section 2 we review the background necessary for the paper, and also introduce geodesic coordinates for the singular Yamabe setting. These are a useful tool for studying the singular Yamabe metric, and in particular performing computations. In section 3, we develop the existence theory for the class of singular Yamabe GJMS operators of integral order, which entails formal analysis of the scattering operator at the boundary. The analysis is a variation of that carried out in section four of [GZ03]. The result is summarized in Theorem 3.1. In section 4, we turn to the global existence of the scattering operator and the results that follow from it, including Theorems A - F and their corollaries. In section 5, we perform specific computations in low dimensions. This is done both to illuminate and illustrate the theory, and to demonstrate the usefulness of geodesic normal coordinates for carrying out computations with singular Yamabe metrics. We also there prove Theorem G.

Appendix A contains more thorough discussion of some analytic ramifications of the limited regularity of the singular Yamabe metric.

Acknowledgements The authors found an error in their formula for ckc_{k} in section 3 of an earlier draft thanks to an explicit computation in [JO21], and have corrected it here. Stephen McKeown carried out part of the work while a postdoctoral research associate at Princeton University, whom he thanks for the hospitality and support.

2. Background

2.1. The Singular Yamabe Metric

Let (Xn+1,g¯)\left(X^{n+1},\bar{g}\right) be a Riemannian manifold with boundary Mn=XM^{n}=\partial X. The singular Yamabe (or Loewner-Nirenberg) problem is to find a defining function uu for MM so that the conformally related complete metric g=u2g¯g=u^{-2}\bar{g} has constant scalar curvature R(g)=n(n+1)R(g)=-n(n+1). It has long been known that the solution uu exists and is unique (see [LN74, AM88, ACF92]). Moreover, it was shown in [ACF92] that uu is regular in the sense that, if r(x)r(x) is the distance function to MM on XX with respect to g¯\bar{g}, then uu has an asymptotic expansion in powers of rr and rklog(r)jr^{k}\log(r)^{j}, where kn+2k\geq n+2. As discussed in detail in [Gra17], this formal expansion can be obtained term by term by writing out the equation R(g)=n(n+1)R(g)=-n(n+1) in terms of g¯\bar{g}. The equation becomes

(2.1) n(n+1)=n(n+1)|du|g¯22nuΔg¯uu2R¯,n(n+1)=n(n+1)|du|_{\bar{g}}^{2}-2nu\Delta_{\bar{g}}u-u^{2}\overline{R},

where R¯=R(g¯)\overline{R}=R(\bar{g}) is the scalar curvature associated to the metric g¯\bar{g}. Then, differentiating equation (2.1) term by term, one can write the expansion (1.3), where u2,,un+1u_{2},\cdots,u_{n+1} and \mathcal{L} are locally determined smooth functions on MM, while un+2u_{n+2} is globally determined. Each of the locally determined quantities is a universal expression in the intrinsic and extrinsic geometry of MM as a hypersurface of (X,g¯)(X,\bar{g}). In particular, \mathcal{L} is an extrinsic conformal invariant of weight (n+1)-(n+1). The function uu itself is also conformally invariant of weight 11: if g¯˘=e2ωg¯\breve{\bar{g}}=e^{2\omega}\bar{g}, then u˘=eωu\breve{u}=e^{\omega}u. This follows easily from the uniqueness of the singular Yamabe metric.

In [Gra17] (see also [GW17]), it was observed that one can define a renormalized volume for singular Yamabe metrics, and that the volume expansion defines a geometrically interesting energy term \mathcal{E} that in some respects generalizes the Willmore energy. Specifically, with the above notation, consider the quantity Volg({r>ε})\operatorname{Vol}_{g}(\left\{r>\varepsilon\right\}). This can be expanded in powers of ε\varepsilon as follows:

(2.2) Volg({r>ε})=c0εn+c1εn+1++cn1ε1+log1ε+V+o(1).\operatorname{Vol}_{g}\left(\left\{r>\varepsilon\right\}\right)=c_{0}\varepsilon^{-n}+c_{1}\varepsilon^{-n+1}+\cdots+c_{n-1}\varepsilon^{-1}+\mathcal{E}\log\frac{1}{\varepsilon}+V+o(1).

Here each ckc_{k} is determined locally in the sense that it is an integral over MM of locally, extrinsically-determined quantities. The energy \mathcal{E} is a global (extrinsic) conformal invariant of the embedding of MM in XX, and is also the integral over MM of a local term; and VV is globally determined in the sense that it may depend on the geometry of (X,g¯)(X,\bar{g}) far away from MM. When n=2n=2, it was shown in the same paper that \mathcal{E} is a linear combination of the Willmore energy of MM and its Euler characteristic.

The singular Yamabe metric is, among other things, an asymptotically hyperbolic (AH) metric, but there are subtleties that make the application of AH theory to this situation slightly subtle. It will be useful to discuss those subtleties here. We recall the following definition.

Definition 2.1.

An asymptotically hyperbolic space is a compact manifold Xn+1X^{n+1} with boundary MnM^{n}, equipped on the interior X̊\mathring{X} with a metric gg such that, for any defining function φ\varphi for MM in XX, φ2g\varphi^{2}g extends to a metric g¯\bar{g} on XX and |dφ|g¯2|TM1|d\varphi|_{\bar{g}}^{2}|_{TM}\equiv 1. If φ\varphi is smooth, then gg is called CkC^{k} (or smooth) AH if g¯\bar{g} is a CkC^{k} (or smooth) compact metric. The conformal infinity is the conformal class [φ2g|TM][\varphi^{2}g|_{TM}] on MM.

In the most typical AH settings, one is given or constructs an AH metric, and it is this that is considered natural; various compactifications correspond to various defining functions φ\varphi, but there is no canonical defining function and thus no canonical compactification. On the other hand, in the singular Yamabe problem, the problem data is precisely the compactification g¯\bar{g}, which is taken to be smooth. The AH metric, which is the singular Yamabe metric, is canonically obtained from g¯\bar{g}, but is not generically a smooth AH metric since uu is not generically a smooth function; in particular, uu is typically only Cn+1C^{n+1} and polyhomogeneous, and it is easy to show that gg is only a CnC^{n} AH metric.

A very useful result in AH geometry is the normal-form theorem, first proved in [GL91]. We will require it here with more attention than usual to the regularity of the metric, so we here give a statement suited to our needs.

Lemma 2.2.

Let g¯\bar{g} be a smooth metric on the manifold Xn+1X^{n+1} with boundary MM, and k=g¯|TMk=\bar{g}|_{TM}. Let g=u2g¯g=u^{-2}\bar{g} be the corresponding singular Yamabe metric. Then for ε>0\varepsilon>0 sufficiently small there is a unique CnC^{n} diffeomorphism ψ:[0,ε)r^×MX\psi:[0,\varepsilon)_{\hat{r}}\times M\hookrightarrow X onto a collar neighborhood of MM such that ψg=dr^2+hr^r^2\psi^{*}g=\frac{d\hat{r}^{2}+h_{\hat{r}}}{\hat{r}^{2}} with h0=kh_{0}=k.

Moreover, (ψ1)r^C(X̊)Cn+1(X)(\psi^{-1})^{*}\hat{r}\in C^{\infty}(\mathring{X})\cap C^{n+1}(X); we will hereafter denote this function simply r^\hat{r}. It further satisfies rr^Cn+2r\hat{r}\in C^{n+2} and limr0rn+1r^=0\lim_{r\to 0}\partial_{r}^{n+1}\hat{r}=0.

The proof is in appendix A.

If r^\hat{r} is extended to XX as a positive smooth function, we will call the metric g¯^=r^2g\hat{\bar{g}}=\hat{r}^{2}g a geodesic representative associated to g¯\bar{g}. The importance of these metrics for us is that the singular Yamabe function associated to one of them is simply r^\hat{r} itself – that is, for such a metric, the intrinsic distance to the boundary is the solution to the singular Yamabe equation (2.1). This will greatly simplify some of our computations in section 5.

2.2. Scattering on Asymptotically Hyperbolic Spaces

The main results we need come from the paper [GZ03] by Graham and Zworski, which analyzed scattering on asymptotically hyperbolic manifolds using tools from [MM87]. Let (Xn+1,g)(X^{n+1},g) be an AH manifold, with M=XM=\partial X and conformal infinity [h][h]. Let xx be a defining function for the boundary and g¯=x2g\bar{g}=x^{2}g (we do not here assume that xx is a geodesic defining function). Consider the operator (Δg+s(ns))u=0(\Delta_{g}+s(n-s))u=0, where uC(X̊)u\in C^{\infty}(\mathring{X}). It is easy to show, by writing the operator Δg\Delta_{g} in local coordinates, that any solution to the equation must have leading order xnsx^{n-s}, assuming Resn2\operatorname{Re}s\geq\frac{n}{2} and sn2s\neq\frac{n}{2}. Thus, the problem considered in [GZ03] is the following, for Resn2\operatorname{Re}s\geq\frac{n}{2} with sn2s\neq\frac{n}{2}. Let fC(M)f\in C^{\infty}(M) be prescribed. Then consider the problem

(2.3) (Δg+s(ns))u=0u=xnsF+o(xns) if Resn2u=xnsF+xsG+O(xn/2+1) if Res=n2,sn2F,GC(X) with F|M=f.\begin{split}(\Delta_{g}+s(n-s))u&=0\\ u&=x^{n-s}F+o(x^{n-s})\text{ if }\operatorname{Re}s\neq\frac{n}{2}\\ u&=x^{n-s}F+x^{s}G+O(x^{n/2+1})\text{ if }\operatorname{Re}s=\frac{n}{2},s\neq\frac{n}{2}\\ F,G\in C^{\infty}(X)\text{ with }F|_{M}=f.\end{split}

To describe the results of the paper, it will be useful to use sections of the normal density bundles C(M,|NM|s)C^{\infty}(M,|N^{*}M|^{s}) over the boundary, which helpfully parametrize first-order changes in the defining function. Given a choice k[k]k\in[k] of conformal representative of the conformal infinity, and letting xx be any defining function such that x2g|TM=kx^{2}g|_{TM}=k, we can trivialize |NM|s|N^{*}M|^{s} by the global section |dx|s|dx|^{s}, and in particular can identify C(M,|NM|s)C^{\infty}(M,|N^{*}M|^{s}) with C(M)C^{\infty}(M). We will also use the notation (s)\mathcal{E}(-s) for the bundle C(M,|NM|s)C^{\infty}(M,|N^{*}M|^{s}).

Let σ(Δg)\sigma(\Delta_{g}) be the spectrum of the Laplacian of gg. Graham and Zworski proved the following theorem.

Theorem 2.3.

There is a unique family of Poisson operators

𝒫(s):C(M,|NM|ns)C(X̊)\mathcal{P}(s):C^{\infty}(M,|N^{*}M|^{n-s})\to C^{\infty}(\mathring{X})

for Resn/2,sn/2\operatorname{Re}s\geq n/2,s\neq n/2, which is meromorphic in {Res>n2}\left\{\operatorname{Re}s>\frac{n}{2}\right\} with poles only for such ss that s(ns)σ(Δg)s(n-s)\in\sigma(\Delta_{g}), and continuous up to {Res=n/2}{n/2}\left\{\operatorname{Re}s=n/2\right\}\setminus\left\{n/2\right\}, such that

(Δg+s(ns))𝒫(s)=0(\Delta_{g}+s(n-s))\mathcal{P}(s)=0

with expansions

𝒫(s)f\displaystyle\mathcal{P}(s)f =xnsF+xsG if sn/2+0/2\displaystyle=x^{n-s}F+x^{s}G\text{ if }s\notin n/2+\mathbb{N}_{0}/2
𝒫(s)f\displaystyle\mathcal{P}(s)f =xn/2k/2F+Gxn/2+k/2logx if s=n/2+k/2,k,\displaystyle=x^{n/2-k/2}F+Gx^{n/2+k/2}\log x\text{ if }s=n/2+k/2,k\in\mathbb{N},

for F,GC(X)F,G\in C^{\infty}(X) such that F|X=fF|_{\partial X}=f. If s=n2+js=\frac{n}{2}+j, then G|M=2p2kfG|_{M}=-2p_{2k}f, where p2kp_{2k} is a differential operator on MM of order MM having principal part σ2j(p2j)=cjσ2j(Δhj)\sigma_{2j}(p_{2j})=c_{j}\sigma_{2j}(\Delta_{h}^{j}), where cj=(1)j[22jj!(j1)!]1c_{j}=(-1)^{j}[2^{2j}j!(j-1)!]^{-1}.

With this result in hand, we can define the scattering matrix as an operator S(s):C(M,|NM|ns)C(M,|NM|s)S(s):C^{\infty}(M,|N^{*}M|^{n-s})\to C^{\infty}(M,|N^{*}M|^{s}) for Resn/2,2sn0\operatorname{Re}s\geq n/2,2s-n\notin\mathbb{N}_{0}, and s(ns)σ(Δg)s(n-s)\notin\sigma(\Delta_{g}). For such ss, and any fC(M,|NM|ns)f\in C^{\infty}(M,|N^{*}M|^{n-s}), we have by the above

𝒫(s)f=xnsF+xsG,\mathcal{P}(s)f=x^{n-s}F+x^{s}G,

with F|M=fF|_{M}=f. The scattering matrix is defined by S(s)f=G|MS(s)f=G|_{M}. It is shown in [GZ03] that S(s)S(s) extends meromorphically to the entire plane.

The log terms in the theorem arise (when they do arise) for the usual reason seen when the indicial roots of a regular singular ODE are separated by an integer. As the statement makes clear, the log coefficient may vanish for n+q2\frac{n+q}{2} with qq odd, but is always nonvanishing for even qq.

This paper applies the results of [GZ03] to the singular Yamabe metric.

2.3. Notation

Throughout, Xn+1X^{n+1} is a compact manifold with boundary MnM^{n} and smooth metric g¯\bar{g}. The singular Yamabe metric g=u2g¯g=u^{-2}\bar{g} is as above. The distance function to MM on XX with respect to g¯\bar{g} is rr, while r^\hat{r} is as in Lemma 2.2. The induced metric on MM is k=g¯|TMk=\bar{g}|_{TM}. When using coordinates, we use the convention that r=x0r=x^{0}, while x1,,xnx^{1},\cdots,x^{n} restrict to coordinates locally on MM. In index notation, we take 0i,jn0\leq i,j\leq n and 1μ,νn1\leq\mu,\nu\leq n. The second fundamental form of MM with respect to the inward-pointing g¯\bar{g}-unit normal r\frac{\partial}{\partial r} is denoted by LL, and the trace-free part by L̊\mathring{L}. The mean curvature of MM is H=kμνLμνH=k^{\mu\nu}L_{\mu\nu}. Our curvature sign convention is such that Rij=RkijkR_{ij}=R^{k}{}_{ijk}, and the Laplace operator is a negative operator, i.e., the divergence of the gradient.

3. Local Analysis

In this section we analyze formal solutions to the equation (Δg+s(ns))u=0(\Delta_{g}+s(n-s))u=0 for a singular Yamabe metric gg. Let g¯\bar{g} be a smooth metric on Xn+1X^{n+1}, and uu the solution to the singular Yamabe equation (2.1), so that g=u2g¯g=u^{-2}\bar{g} has constant scalar curvature n(n+1)-n(n+1). We again let M=XM=\partial X. Near MM, we write g¯=dr2+hr\bar{g}=dr^{2}+h_{r}, where hrh_{r} is a one-parameter family of metrics on MM and rr is the g¯\bar{g}-distance to MM. We write k=h0=g¯|TMk=h_{0}=\bar{g}|_{TM}.

The following result, which is the primary result of this section, is directly analogous to Proposition 4.2 of [GZ03] for the Einstein case, and the proof is modified accordingly. One difference is that a log term arises here for every integer, whereas in the Einstein cases it arises only for the even integers. The other significant difference is that the metric itself is non-smooth, and via uu has logarithmic terms that must be considered.

Theorem 3.1.

Let gg be the singular Yamabe metric associated to (X,g¯)(X,\bar{g}), and fC(M)f\in C^{\infty}(M). For every qq\in\mathbb{N} with 1qn1\leq q\leq n, and s=n+q2s=\frac{n+q}{2}, there is a formal solution vv to the equation

(Δg+s(ns))v=O(r)(\Delta_{g}+s(n-s))v=O(r^{\infty})

of the form

v=rnq2(F+Grqlogr),v=r^{\frac{n-q}{2}}(F+Gr^{q}\log r),

where FC(X)F\in C^{\infty}(X), GCnq,1ε(X)G\in C^{n-q,1-\varepsilon}(X) is polyhomogeneous, and F|M=fF|_{M}=f. Here FF is uniquely determined mod O(rq)O(r^{q}) and GG is uniquely determined mod O(r)O(r^{\infty}). In addition,

(3.1) G|M=2cqPqSYf,G|_{M}=-2c_{q}P_{q}^{SY}f,

where cq0c_{q}\neq 0 is a universal constant and PqSYP_{q}^{SY} is a differential operator on MM which, if qq is even, has principal part (Δk)q2(-\Delta_{k})^{\frac{q}{2}}, and if q>1q>1 is odd, has the same principal part as L̊μνμν(Δk)q32\mathring{L}^{\mu\nu}\nabla_{\mu}\nabla_{\nu}(\Delta_{k})^{\frac{q-3}{2}}, where L̊\mathring{L} is the tracefree second fundamental form. If q=1q=1, then G=0G=0.

Finally, PqSYP_{q}^{SY} depends only on the jet of g¯\bar{g} at MM, and defines a conformally invariant operator (qn2)(qn2)\mathcal{E}(\frac{q-n}{2})\to\mathcal{E}\left(\frac{-q-n}{2}\right).

Proof of Theorem 3.1.

We wish to formally solve the equation (Δg+s(ns))v=0(\Delta_{g}+s(n-s))v=0. Now u=r+O(r2)u=r+O(r^{2}), so we may write u=ru~u=r\tilde{u} for some u~Cn(X)\tilde{u}\in C^{n}(X) satisfying u~|M=1\tilde{u}|_{M}=1. Thus,

Δgv\displaystyle\Delta_{g}v =r1+nu~1+n(deth)1/2i[r1nu~1n(deth)1/2g¯ijjv]\displaystyle=r^{1+n}\tilde{u}^{1+n}(\det h)^{-1/2}\partial_{i}\left[r^{1-n}\tilde{u}^{1-n}(\det h)^{1/2}\bar{g}^{ij}\partial_{j}v\right]
=r2u~2r2v+(1n)ru~2rv+(1n)r2u~r(u~)rv+12r2u~2hμνhμνrv\displaystyle=r^{2}\tilde{u}^{2}\partial_{r}^{2}v+(1-n)r\tilde{u}^{2}\partial_{r}v+(1-n)r^{2}\tilde{u}\partial_{r}(\tilde{u})\partial_{r}v+\frac{1}{2}r^{2}\tilde{u}^{2}h^{\mu\nu}h^{\prime}_{\mu\nu}\partial_{r}v
+(1n)r2u~hμνμ(u~)ν(v)+r2u~2Δhrv.\displaystyle\quad+(1-n)r^{2}\tilde{u}h^{\mu\nu}\partial_{\mu}(\tilde{u})\partial_{\nu}(v)+r^{2}\tilde{u}^{2}\Delta_{h_{r}}v.

(Here, a prime denotes r\partial_{r}.) Taking v=rnsψv=r^{n-s}\psi, we find

[Δg+s(ns)]v\displaystyle[\Delta_{g}+s(n-s)]v =rns+2u~2r2ψ+2(ns)rns+1u~2rψ\displaystyle=r^{n-s+2}\tilde{u}^{2}\partial_{r}^{2}\psi+2(n-s)r^{n-s+1}\tilde{u}^{2}\partial_{r}\psi
+(ns)(ns1)rnsu~2ψ+(1n)(ns)rnsu~2ψ\displaystyle\quad+(n-s)(n-s-1)r^{n-s}\tilde{u}^{2}\psi+(1-n)(n-s)r^{n-s}\tilde{u}^{2}\psi
+(1n)rns+2u~2rψ+(1n)rns+2u~r(u~)rψ\displaystyle\quad+(1-n)r^{n-s+2}\tilde{u}^{2}\partial_{r}\psi+(1-n)r^{n-s+2}\tilde{u}\partial_{r}(\tilde{u})\partial_{r}\psi
+(1n)(ns)rns+1u~r(u~)ψ+12rns+2u~2hμνhμνrψ\displaystyle\quad+(1-n)(n-s)r^{n-s+1}\tilde{u}\partial_{r}(\tilde{u})\psi+\frac{1}{2}r^{n-s+2}\tilde{u}^{2}h^{\mu\nu}h^{\prime}_{\mu\nu}\partial_{r}\psi
+12(ns)rns+1u~2hμνhμνψ+(1n)rns+2u~hμνμ(u~)ν(ψ)\displaystyle\quad+\frac{1}{2}(n-s)r^{n-s+1}\tilde{u}^{2}h^{\mu\nu}h^{\prime}_{\mu\nu}\psi+(1-n)r^{n-s+2}\tilde{u}h^{\mu\nu}\partial_{\mu}(\tilde{u})\partial_{\nu}(\psi)
+rns+2u~2Δhrψ+s(ns)rnsψ\displaystyle\quad+r^{n-s+2}\tilde{u}^{2}\Delta_{h_{r}}\psi+s(n-s)r^{n-s}\psi
=rns+1[ru~2r2ψ+((n+12s)u~2+(1n)ru~ru~\displaystyle=r^{n-s+1}\left[r\tilde{u}^{2}\partial_{r}^{2}\psi+\left((n+1-2s)\tilde{u}^{2}+(1-n)r\tilde{u}\partial_{r}\tilde{u}\right.\right.
+12ru~2hμνhμν)rψ+(s(sn)u~21r+(1n)(ns)u~ru~\displaystyle\quad+\frac{1}{2}r\tilde{u}^{2}h^{\mu\nu}h^{\prime}_{\mu\nu})\partial_{r}\psi+\left(s(s-n)\frac{\tilde{u}^{2}-1}{r}+(1-n)(n-s)\tilde{u}\partial_{r}\tilde{u}\right.
+12(ns)u~2hμνhμν)ψ+(1n)ru~hμνμu~νψ\displaystyle\quad+\frac{1}{2}(n-s)\tilde{u}^{2}h^{\mu\nu}h^{\prime}_{\mu\nu})\psi+(1-n)r\tilde{u}h^{\mu\nu}\partial_{\mu}\tilde{u}\partial_{\nu}\psi
+ru~2Δhrψ].\displaystyle\quad\left.+r\tilde{u}^{2}\Delta_{h_{r}}\psi\right].

We may conclude that

[Δg+s(ns)]rns=rns+1𝒟s,[\Delta_{g}+s(n-s)]\circ r^{n-s}=r^{n-s+1}\circ\mathcal{D}_{s},

where

(3.2) 𝒟s=ru~2r2+((n+12s)u~2+(1n)ru~ru~+12ru~2hμνhμν)r+s(sn)u~21r+(1n)(ns)u~ru~+12(ns)u~2hμνhμν+(1n)ru~gradhr(u~)+ru~2Δhr.\begin{split}\mathcal{D}_{s}&=r\tilde{u}^{2}\partial_{r}^{2}+\left((n+1-2s)\tilde{u}^{2}+(1-n)r\tilde{u}\partial_{r}\tilde{u}+\frac{1}{2}r\tilde{u}^{2}h^{\mu\nu}h^{\prime}_{\mu\nu}\right)\partial_{r}\\ &\quad+s(s-n)\frac{\tilde{u}^{2}-1}{r}+(1-n)(n-s)\tilde{u}\partial_{r}\tilde{u}+\frac{1}{2}(n-s)\tilde{u}^{2}h^{\mu\nu}h^{\prime}_{\mu\nu}\\ &\quad+(1-n)r\tilde{u}\operatorname{grad}_{h_{r}}(\tilde{u})+r\tilde{u}^{2}\Delta_{h_{r}}.\end{split}

Keeping in mind that u~2=1+O(r)\tilde{u}^{2}=1+O(r) and that μu~=O(r)\partial_{\mu}\tilde{u}=O(r), we observe that

(3.3) 𝒟s(fjrj)=j(j+n2s)rj1fj+O(rj).\mathcal{D}_{s}(f_{j}r^{j})=j(j+n-2s)r^{j-1}f_{j}+O(r^{j}).

This equation is the same as that in [GZ03]; however, we have avoided expressing the metric in normal form here, since g¯\bar{g} is part of the data of the problem. It is also convenient to record that

(3.4) 𝒟s(gjrjlogr)=(2j+n2s)gjrj1+j(j+n2s)gjrj1log(r)+o(rj1).\mathcal{D}_{s}(g_{j}r^{j}\log r)=(2j+n-2s)g_{j}r^{j-1}+j(j+n-2s)g_{j}r^{j-1}\log(r)+o(r^{j-1}).

Suppose n2s0n-2s\notin\mathbb{N}_{0}. Then (3.3) allows us to construct a formal solution. Set f0=F0=ff_{0}=F_{0}=f. For j1j\geq 1 with jnj\leq n, define fjf_{j} by

(3.5) j(j+n2s)fj\displaystyle j(j+n-2s)f_{j} =r1j𝒟s(Fj1)|r=0\displaystyle=-r^{1-j}\mathcal{D}_{s}(F_{j-1})|_{r=0}
Fj\displaystyle F_{j} =Fj1+fjrj.\displaystyle=F_{j-1}+f_{j}r^{j}.

Then setting vj=rnsFjv_{j}=r^{n-s}F_{j}, we clearly have

[Δg+s(ns)]vj=O(rns+j).[\Delta_{g}+s(n-s)]v_{j}=O(r^{n-s+j}).

By induction, we may thus find vn=rnsFnv_{n}=r^{n-s}F_{n} satisfying [Δg+s(ns)]vn=O(r2ns)[\Delta_{g}+s(n-s)]v_{n}=O(r^{2n-s}). However, examining 𝒟s\mathcal{D}_{s} in (3.2), we see that 𝒟s(Fn)\mathcal{D}_{s}(F_{n}) contains a term of the form a(ns)frnlog(r)a(n-s)f\mathcal{L}r^{n}\log(r), for some universal aa\in\mathbb{R}, via the terms u~21r\frac{\tilde{u}^{2}-1}{r} and ru~\partial_{r}\tilde{u}. Here \mathcal{L} is as in (1.3). The induction can therefore be continued only by first adding a term of the form gn+1rn+1log(r)g_{n+1}r^{n+1}\log(r) (see (3.4)) to cancel the rnlog(r)r^{n}\log(r) term before adding fnrnf_{n}r^{n}. Having done this, the induction can be continued to infinite order, by adding monomials and logarithmic terms, as is standard. (In fact, by the polyhomogeneity theorem in [ACF92], higher powers of logarithms may be necessary at high order, but this will be of no concern to us. For a recent very explicit example of a construction with this behavior, see [McK18].)

By Borel’s lemma, this gives us an infinite-order solution vv. Just as in [GZ03], an easy induction shows that for j=2pj=2p even,

f2p=c2p,sP2p,sSYf,c2p,s=(1)pΓ(sn2p)22pp!Γ(sn2),f_{2p}=c_{2p,s}P_{2p,s}^{SY}f,\quad c_{2p,s}=(-1)^{p}\frac{\Gamma\left(s-\frac{n}{2}-p\right)}{2^{2p}p!\Gamma\left(s-\frac{n}{2}\right)},

where P2p,sP_{2p,s} is a differential operator on MM with principal part equal to that of (Δkp)(-\Delta_{k}^{p}).

The analysis of the leading part of the odd-order terms is somewhat more complicated, because f2p+1f_{2p+1} contains derivatives only of order 2p2p, and there are several different contributions to these. Because

(3.6) rΔhr|r=0=Δkr+2L̊μνμν+2nHΔk+l.o.t.s\partial_{r}\Delta_{h_{r}}|_{r=0}=\Delta_{k}\partial_{r}+2\mathring{L}^{\mu\nu}\nabla_{\mu}\nabla_{\nu}+\frac{2}{n}H\Delta_{k}+l.o.t.s

(where l.o.t.sl.o.t.s denotes lower-order terms), it is clear from (3.2) that

f2p+1=c2p+1,sL̊μνμνΔkp1f+d2p+1,sHΔkpf+l.o.t.s.f_{2p+1}=c_{2p+1,s}\mathring{L}^{\mu\nu}\nabla_{\mu}\nabla_{\nu}\Delta_{k}^{p-1}f+d_{2p+1,s}H\Delta_{k}^{p}f+l.o.t.s.

for some constants c2p+1,sc_{2p+1,s} and d2p+1,sd_{2p+1,s}. It is easy to compute, as a base case, that c1,s=0c_{1,s}=0 and d1,s=ns2nd_{1,s}=\frac{n-s}{2n}. Now, as L̊μνμνΔkp1\mathring{L}^{\mu\nu}\nabla_{\mu}\nabla_{\nu}\Delta_{k}^{p-1} and Δk(L̊μνμνΔkp2)\Delta_{k}(\mathring{L}^{\mu\nu}\nabla_{\mu}\nabla_{\nu}\Delta_{k}^{p-2}) have the same principal parts, a straightforward induction shows that

c2p+1,s=12(2p+1)(sn2p12)(2(1)p1c2p2,s+c2p1,s).c_{2p+1,s}=\frac{1}{2(2p+1)\left(s-\frac{n}{2}-p-\frac{1}{2}\right)}(2(-1)^{p-1}c_{2p-2,s}+c_{2p-1,s}).

It is then likewise easy to show by induction that cj,s>0c_{j,s}>0 whenever s>n+j2s>\frac{n+j}{2}, for odd jj.

We now show by induction that d2p+1,s=n+2p2s2n(1)pc2p,sd_{2p+1,s}=\frac{n+2p-2-s}{2n}(-1)^{p}c_{2p,s}. This is clearly true for p=0p=0. We assume it is true for p<pp^{\prime}<p. Let j=2p+1j=2p+1. We next observe that ru~|r=0=12nH\partial_{r}\tilde{u}|_{r=0}=-\frac{1}{2n}H (see (2.6) in [Gra17]), while hμνhμν|r=0=2Hh^{\mu\nu}h^{\prime}_{\mu\nu}|_{r=0}=-2H. As cj3,scj1,s=(j1)(2snj)\frac{c_{j-3,s}}{c_{j-1,s}}=-(j-1)(2s-n-j), we find from (3.2) and (3.5) that

j(2snj)dj,s\displaystyle j(2s-n-j)d_{j,s} =(1)pcj1,s[j12n(12j3n+4s)+ns2n(2sn1)\displaystyle=(-1)^{p}c_{j-1,s}\left[\frac{j-1}{2n}(1-2j-3n+4s)+\frac{n-s}{2n}(2s-n-1)\right.
cj3,sncj1,s+(1)pdj2,scj1,s]\displaystyle\quad-\left.\frac{c_{j-3,s}}{nc_{j-1,s}}+(-1)^{p}\frac{d_{j-2,s}}{c_{j-1,s}}\right]
=(1)pcj1,s2n[(j1)(sn+3sn2s2n2+j(3s2nj))\displaystyle=(-1)^{p}\frac{c_{j-1,s}}{2n}\left[(j-1)(s-n+3sn-2s^{2}-n^{2}+j(3s-2n-j))\right.
+(ns)(2sn1)]\displaystyle\quad+\left.(n-s)(2s-n-1)\right]
=(1)pjcj1,s2n(j+ns1)(2snj).\displaystyle=(-1)^{p}\frac{jc_{j-1,s}}{2n}(j+n-s-1)(2s-n-j).

This yields the claim regarding d2p+1,s=dj,sd_{2p+1,s}=d_{j,s}. We emphasize that d2p+1,sd_{2p+1,s} is smooth across s=n+2p+12s=\frac{n+2p+1}{2}.

For notational consistency, we define P2p+1,sSYP_{2p+1,s}^{SY} such that f2p+1=c2p+1,sPsp+1,sff_{2p+1}=c_{2p+1,s}P_{sp+1,s}f.

Now suppose that 2sn=q2s-n=q, where 1qn1\leq q\leq n, as in the hypothesis. The above construction works until the determination of fqf_{q}; then the coefficient in (3.3) vanishes, and we cannot remove the rq1r^{q-1} term from 𝒟q(Fq1)\mathcal{D}_{q}(F_{q-1}). This may be addressed by adding a term of the form gqrqlog(r)g_{q}r^{q}\log(r), where gqg_{q} is determined by (3.4) to cancel the rq1r^{q-1} term in 𝒟q(Fq1)\mathcal{D}_{q}(F_{q-1}). Note that since qnq\leq n, this happens before the logarithm at order n+1n+1 discussed above. The expansion then continues as before, with additional logarithmic terms appearing at order n+1n+1 (which limits the smoothness of GG). The remainder of the claims follow immediately. As in [GZ03], it is clear that (3.1) holds with PqSY=Pq,n+q2SYP_{q}^{SY}=P_{q,\frac{n+q}{2}}^{SY} and with cq=ress=n+q2cq,sc_{q}=\operatorname{res}_{s=\frac{n+q}{2}}c_{q,s}. In particular,

(3.7) c2p=(1)p[22pp!(p1)!]1c1=0c2p+1=12(2p+1)(2(1)p1c2p2,n+2p+12+c2p1,n+2p+12)>0.\begin{split}c_{2p}&=(-1)^{p}[2^{2p}p!(p-1)!]^{-1}\\ c_{1}&=0\\ c_{2p+1}&=\frac{1}{2(2p+1)}\left(2(-1)^{p-1}c_{2p-2,\frac{n+2p+1}{2}}+c_{2p-1,\frac{n+2p+1}{2}}\right)>0.\end{split}

Since d2p+1,sd_{2p+1,s} does not contribute to the residue of P2p+1,sP_{2p+1,s} at s=n+2p+12s=\frac{n+2p+1}{2}, the principal part of PqSYP_{q}^{SY} is as claimed. ∎

Note that the conformal invariance property of the operators PqSYP_{q}^{SY} can be expressed as follows instead of using bundles: if g¯~=e2ωg¯\tilde{\bar{g}}=e^{2\omega}\bar{g}, then

P~qSYf=e(n+q)ω2PqSY(e(nq)ω2f).\widetilde{P}_{q}^{SY}f=e^{-\frac{(n+q)\omega}{2}}P_{q}^{SY}\left(e^{\frac{(n-q)\omega}{2}}f\right).

4. Global analysis

In section 2, we described the results on the scattering operator and the Poisson operators proved in [GZ03], based on [MM87]. Actually, both of those papers assume that one is dealing with a smooth AH metric, i.e., a metric whose compactifications are smooth. Although such is not true of AH Einstein metrics in odd dimension (even boundary dimension), the existence theory for such metrics (given a conformal infinity) is very difficult and largely open, and both existence and uniqueness are known sometimes to fail. Therefore, papers such as [GZ03] and [FG02] made the reasonable decision to restrict attention to smooth formal Einstein metrics, which are Einstein up to as high an order as possible while remaining smooth. The resulting quantities are not well-defined, if global; but given the non-uniqueness of (some) Einstein metrics, they would not be in any event, in general.

In contrast, the singular Yamabe problem offers both existence and uniqueness for any compact Riemannian manifold with boundary. There is therefore the opportunity to define genuinely well-defined global quantities if one uses the singular Yamabe metric as the AH metric for the scattering problem. On the other hand, this raises slight technical difficulties: as we have seen, the singular Yamabe metric is not generally smooth, but has a logarithmic term, and so technically we cannot simply apply results from [GZ03] or [MM87] in a simplistic fashion. We therefore state the following theorem; since the main goals of our paper are geometric, we defer discussing the proof till appendix A.

Theorem 4.1.

Let (Xn+1,g¯)(X^{n+1},\bar{g}) be a smooth Riemannian manifold with boundary MM, let k=g¯|TMk=\bar{g}|_{TM}, and let gg be the singular Yamabe metric corresponding to g¯\bar{g}. Let rr be the g¯\bar{g}-distance to MM on XX. There is a unique family of Poisson operators

𝒫(s):C(M,|NM|ns)C(X̊)\mathcal{P}(s):C^{\infty}(M,|N^{*}M|^{n-s})\to C^{\infty}(\mathring{X})

for n2Res<n+12\frac{n}{2}\leq\operatorname{Re}s<n+\frac{1}{2}, which is meromorphic in {n+12>Res>n2}\left\{n+\frac{1}{2}>\operatorname{Re}s>\frac{n}{2}\right\} with poles only for such ss that s(ns)σ(Δg)s(n-s)\in\sigma(\Delta_{g}), and continuous up to {Res=n/2}{n/2}\left\{\operatorname{Re}s=n/2\right\}\setminus\left\{n/2\right\}, such that

(Δg+s(ns))𝒫(s)=0(\Delta_{g}+s(n-s))\mathcal{P}(s)=0

with expansions

𝒫(s)\displaystyle\mathcal{P}(s) =rnsF+rsG if sn/2+0/2\displaystyle=r^{n-s}F+r^{s}G\text{ if }s\notin n/2+\mathbb{N}_{0}/2
𝒫(s)f\displaystyle\mathcal{P}(s)f =rn/2q/2F+Grn/2+q/2logr if s=n/2+q/2,q,\displaystyle=r^{n/2-q/2}F+Gr^{n/2+q/2}\log r\text{ if }s=n/2+q/2,q\in\mathbb{N},

where in the first case, both F,GCn(X)F,G\in C^{n}(X) are polyhomogeneous, and in the second case, FC(X)F\in C^{\infty}(X) and GCn(X)G\in C^{n}(X) is polyhomogeneous. If s=n2+js=\frac{n}{2}+j, then G|M=2p2qfG|_{M}=-2p_{2q}f, where p2qp_{2q} is a differential operator on MM of order MM having principal part σ2j(p2j)=cjσ2j(Δkj)\sigma_{2j}(p_{2j})=c_{j}\sigma_{2j}(\Delta_{k}^{j}), where cjc_{j} is as in (3.7).

As in the smooth case, we define the scattering operator for s>n2s>\frac{n}{2}, s(ns)σ(Δg)s(n-s)\notin\sigma(\Delta_{g}) and 2sn2s-n\notin\mathbb{N}, by S(s)f=G|MS(s)f=G|_{M}, where 𝒫(s)f=rnsF+rsG\mathcal{P}(s)f=r^{n-s}F+r^{s}G as above. We obtain the following (see appendix A).

Theorem 4.2.

The scattering operator S(s)S(s) has a meromorphic extension to the strip 12<s<n+12-\frac{1}{2}<s<n+\frac{1}{2}, regular for Res=n2\operatorname{Re}s=\frac{n}{2}. On that strip, Propositions 3.6 - 3.10 of [GZ03] remain true.

The reason for the restriction on ss in both of these statements is the following: as seen in the proof of Proposition 3.1, the logarithmic term in the expansion of the singular Yamabe function uu implies that a term of the form r2n+1slog(r)r^{2n+1-s}\log(r) will appear in the asymptotic expansion of 𝒫(s)f\mathcal{P}(s)f. In order to be able to ignore this phenomenon in analyzing the scattering operator, we require that the higher indicial root, ss, occur before this power, i.e., s<2n+1ss<2n+1-s. The remaining cases could be analyzed, but they would be rather more involved, and in any event are not necessary for what we wish to study here.

We are ready to prove Theorem A.

Proof of Theorem A.

The existence and meromorphicity of the scattering operator is Theorem 4.2. Theorem 3.1, together with Proposition 3.6 of [GZ03], gives the remainder of the claims. ∎

Having defined the scattering operator, we can also finally define our QSYQ^{SY}-curvature as in the introduction:

(4.1) QSY=cn1S(n)1,Q^{SY}=c_{n}^{-1}S(n)1,

where cqc_{q} is as in (3.7). Note that although S(s)S(s) actually has a pole at s=ns=n, the residue has no constant term, so that S(s)1S(s)1 continues holomorphically across s=ns=n. This is discussed in general in [GZ03].

We also can now define the “fractional order GJMS operators” for this setting, following [CdMG11] (where in fact this setting was mentioned, but not discussed in detail). We define P2γSY=S(n2+γ)P_{2\gamma}^{SY}=S(\frac{n}{2}+\gamma). Once again this notation differs by a factor of two from the notation in [CdMG11], which is done in order to be consistent with the integer case. For us, P2γP_{2\gamma} can be viewed as a pseudodifferential operator of order 2γ2\gamma; see p. 107 of [GZ03].

We next prove the following theorem. The proof in the Einstein case is given in [FG02], and in fact the same proof works here. To refresh the reader’s memory and demonstrate that the proof carries over, and because we will want to refer back to it, we reproduce the proof here.

Theorem 4.3.

Let gg be the singular Yamabe metric on (Xn+1,g¯)(X^{n+1},\bar{g}), as above; and let rr be the g¯\bar{g}-distance function to Mn=XM^{n}=\partial X. There is a unique function UC(X̊)U\in C^{\infty}(\mathring{X}) solving

ΔgU=n,-\Delta_{g}U=n,

with asymptotics

U=log(r)+A+Brnlogr,U=\log(r)+A+Br^{n}\log r,

where AC(X)A\in C^{\infty}(X) and BCn(X)B\in C^{n}(X), and where A|M=0A|_{M}=0 and B|M=2cnQSYB|_{M}=-2c_{n}Q^{SY}.

Proof.

By uniqueness, we have 𝒫(n)1=1\mathcal{P}(n)1=1. Now for all ss near nn, we of course have

(4.2) (Δg+s(ns))𝒫(s)1=0,(\Delta_{g}+s(n-s))\mathcal{P}(s)1=0,

and 𝒫(s)1\mathcal{P}(s)1 is a holomorphic family of functions on X̊\mathring{X}.

Set U=dds𝒫(s)|s=nU=-\frac{d}{ds}\mathcal{P}(s)|_{s=n}. Then differentiating (4.2) with respect to ss and taking s=ns=n gives

ΔgU=n.\Delta_{g}U=-n.

On the other hand, near s=ns=n we can write

(4.3) 𝒫(s)1=rnsFs+rsGs.\mathcal{P}(s)1=r^{n-s}F_{s}+r^{s}G_{s}.

This expansion can be extended to s=ns=n, where 𝒫(n)1=1\mathcal{P}(n)1=1, and it was shown in [GZ03] and pointed out in [FG02] that the extension is unique subject to the requirement that Fs,GsF_{s},G_{s} be chosen to depend holomorphically on ss across s=ns=n. (Actually, in the Einstein case considered in [FG02], this argument is needed only for nn even, due to evenness properties available in that case. In our situation, it applies to both even and odd nn.)

It follows from the definition (4.1) of QSYQ^{SY} and the definition of the scattering matrix that Gn|M=cnQSYG_{n}|_{M}=c_{n}Q^{SY}; thus, Fn=1cnQSYrnF_{n}=1-c_{n}Q^{SY}r^{n}.

We can differentiate both sides of (4.3) and conclude

(4.4) U=Fnlog(r)FnGnrnlog(r)rnGn,U=F_{n}\log(r)-F_{n}^{\prime}-G_{n}r^{n}log(r)-r^{n}G_{n}^{\prime},

with =dds{}^{\prime}=\frac{d}{ds}. Since Fs|M1F_{s}|_{M}\equiv 1, the second term vanishes. The result follows from the above asymptotics. ∎

Precisely the same arguments given in section 3 of [FG02] for the even-nn Einstein case now immediately work, given Proposition 3.1 and Theorem 4.3, to produce Theorem C and Corollary D. We do not reproduce those proofs here.

Note that it follows from Corollary D and from [Gra17] that the boundary integral of QSYQ^{SY} is a global conformal invariant. Furthermore, two consequences of the transformation rule in Theorem C is that QSYQ^{SY} is generically nonzero (since PnSYP_{n}^{SY} is generically nontrivial); and, although QSYQ^{SY} is extrinsic in the sense that it depends on g¯\bar{g} and not only k=g¯|TMk=\bar{g}|_{TM}, it is nevertheless the case that given a conformal class [g¯][\bar{g}], QSYQ^{SY} depends only on g¯|TM\bar{g}|_{TM}.

We may now prove Theorem E using an argument from [CQY08].

Proof of Theorem E.

Let UU be as in theorem 4.3. Then by Green’s theorem, we have

(4.5) Volg({r>ε})=1nr>εΔgU𝑑vg=1nr=εUn𝑑vhε=1nε1nMUr|r=εdvhε,\operatorname{Vol}_{g}(\left\{r>\varepsilon\right\})=-\frac{1}{n}\int_{r>\varepsilon}\Delta_{g}Udv_{g}=-\frac{1}{n}\int_{r=\varepsilon}\frac{\partial U}{\partial n}dv_{h_{\varepsilon}}=\frac{1}{n}\varepsilon^{1-n}\oint_{M}\left.\frac{\partial U}{\partial r}\right|_{r=\varepsilon}dv_{h_{\varepsilon}},

since the outward normal to {x=ε}\left\{x=\varepsilon\right\} is εr-\varepsilon\frac{\partial}{\partial r}. Now, by (4.4),

Ur=1r(n1)an1rn1nanrnn(dds|s=nS(s)1)rn12cnQSYrn12ncnQSYrn1log(r)+o(rn1).\begin{split}\frac{\partial U}{\partial r}=&\frac{1}{r}-\cdots-(n-1)a_{n-1}^{\prime}r^{n-1}-na_{n}^{\prime}r^{n}-n\left(\left.\frac{d}{ds}\right|_{s=n}S(s)1\right)r^{n-1}\\ &\quad-2c_{n}Q^{SY}r^{n-1}-2nc_{n}Q^{SY}r^{n-1}\log(r)+o(r^{n-1}).\end{split}

Since by [Gra17] Mv(n)𝑑vk==2cnMQSY𝑑vk\oint_{M}v^{(n)}dv_{k}=\mathcal{E}=2c_{n}\oint_{M}Q^{SY}dv_{k}, the result follows from (4.5) by collecting zeroth-order terms in ε\varepsilon. ∎

Finally, with the above pieces all in place, the proof of Theorem F is identical to that in [CQY08].

5. Explicit Computations

In this section, we compute several of the quantities and operators defined earlier for low dimensions. Throughout, we work near the boundary Mn=Xn+1M^{n}=\partial X^{n+1}, on a collar neighborhood VV with a diffeomorphism ψ:[0,ε)r×MV\psi:[0,\varepsilon)_{r}\times M\to V such that ψg¯=dr2+hr\psi^{*}\bar{g}=dr^{2}+h_{r}. We use Greek indices 1μ,νn1\leq\mu,\nu\leq n on MM, and Latin indices 0i,jn0\leq i,j\leq n on XX. In particular, x0=rx^{0}=r. We set k=h0=g¯|TMk=h_{0}=\bar{g}|_{TM}.

We begin with the following lemma, which is an expansion of the metric in Fermi coordinates. This is well-known to first order. The result to this order is contained in [GG19].

Lemma 5.1.

Let Xn+1X^{n+1} be a smooth manifold with boundary X=Mn\partial X=M^{n}, and g¯\bar{g} a smooth metric on XX. Suppose that, near MM, the metric gg is written as g¯=dr2+hr\bar{g}=dr^{2}+h_{r}, where hrh_{r} is a one-parameter family of metrics on MM. Let k=g¯|TMk=\bar{g}|_{TM}. Then

hμν=kμν2rLμν+r2(LμσLνσR¯0μν0)13r3(¯0R¯0μν04L(μσR¯ν)00σ)+O(r4),h_{\mu\nu}=k_{\mu\nu}-2rL_{\mu\nu}+r^{2}(L_{\mu}^{\sigma}L_{\nu\sigma}-\overline{R}_{0\mu\nu 0})-\frac{1}{3}r^{3}(\overline{\nabla}_{0}\overline{R}_{0\mu\nu 0}-4L^{\sigma}_{(\mu}\overline{R}_{\nu)00\sigma})+O(r^{4}),

where LL is the second fundamental form of MM with respect to g¯\bar{g} and R¯\overline{R} is the curvature tensor of g¯\bar{g}. Moreover, all coefficients are evaluated at r=0r=0, and indices raised with k1k^{-1}.

To compute the local invariants QSYQ^{SY} and PkSYP^{SY}_{k}, we must formally solve the scattering problem (Δg+s(ns))u=0(\Delta_{g}+s(n-s))u=0 for gg the singular Yamabe metric. The straightforward approach to this entails computing uu from g¯\bar{g}, then computing the scattering operator of g=u2g¯g=u^{-2}\bar{g}, and finally formally solving that equation. This approach is extremely tedious, and we will pursue a different one. Let r^\hat{r} be a geodesic defining function with respect to g=u2g¯g=u^{-2}\bar{g}, of the kind constructed in Lemma 2.2, and such that r^r|M=1\frac{\partial\hat{r}}{\partial r}|_{M}=1. Then for g¯^=r^2g\hat{\bar{g}}=\hat{r}^{2}g, which is a CnC^{n} compact metric on XX, the singular Yamabe function u^\hat{u} is equal to r^\hat{r}, the g¯^\hat{\bar{g}}-distance to MM. This drastically simplifies the computation of QSY^\hat{Q^{SY}} and P^kSY\hat{P}^{SY}_{k}. Morever, because (within a conformal class) both depend only on the boundary representative, and because g¯^|TM=g¯|TM\hat{\bar{g}}|_{TM}=\bar{g}|_{TM} by construction, in fact we have in this way computed QSYQ^{SY} and PkSYP_{k}^{SY}. All that remains is to express the computed quantity in terms of invariants of our original metric g¯\bar{g}, and for that task, the following three lemmas are the tools we will need.

The first is completely standard, and is included here only for convenience. The proof is left as an exercise.

Lemma 5.2.

Suppose g¯\bar{g} is a smooth metric on Xn+1X^{n+1}, and that g¯~=e2ωg¯\tilde{\bar{g}}=e^{2\omega}\bar{g}. Let k=g¯|TMk=\bar{g}|_{TM}. Then extrinsic invariants at the boundary MM transform as follows:

H~\displaystyle\widetilde{H} =eω(Hnωr)\displaystyle=e^{-\omega}(H-n\omega_{r})
L~̊μν\displaystyle\mathring{\widetilde{L}}_{\mu\nu} =eωL̊μν\displaystyle=e^{\omega}\mathring{L}_{\mu\nu}
R¯~μν\displaystyle\widetilde{\overline{R}}_{\mu\nu} =R¯μν(n1)¯μν2ω+(n1)ωμων(Δg¯ω(n1)|dω|g¯2)g¯μν\displaystyle=\overline{R}_{\mu\nu}-(n-1)\overline{\nabla}^{2}_{\mu\nu}\omega+(n-1)\omega_{\mu}\omega_{\nu}-(\Delta_{\bar{g}}\omega-(n-1)|d\omega|_{\bar{g}}^{2})\bar{g}_{\mu\nu}
R~μν\displaystyle\widetilde{R}_{\mu\nu} =Rμν(n2)μν2ω+(n2)ωμων(Δkω+(n2)|dω|k2)kμν\displaystyle=R_{\mu\nu}-(n-2)\nabla^{2}_{\mu\nu}\omega+(n-2)\omega_{\mu}\omega_{\nu}-(\Delta_{k}\omega+(n-2)|d\omega|_{k}^{2})k_{\mu\nu}
R¯~\displaystyle\widetilde{\overline{R}} =e2ω(R¯2nΔg¯ωn(n1)|dω|g¯2)\displaystyle=e^{-2\omega}(\overline{R}-2n\Delta_{\bar{g}}\omega-n(n-1)|d\omega|_{\bar{g}}^{2})
R~\displaystyle\widetilde{R} =e2ω(R2(n1)Δkω(n1)(n2)|dω|k2).\displaystyle=e^{-2\omega}(R-2(n-1)\Delta_{k}\omega-(n-1)(n-2)|d\omega|_{k}^{2}).

Here R¯μν\overline{R}_{\mu\nu} and R¯\overline{R} are the Ricci and scalar curvatures for g¯\bar{g}, and RμνR_{\mu\nu} and RR are the Ricci and scalar curvatures for kk; and similarly for R¯~\widetilde{\overline{R}}, etc. Moreover, |dω|g¯2|d\omega|_{\bar{g}}^{2} is the squared g¯\bar{g}-norm of dωd\omega, where dd is the exterior derivative on XX; while |dω|k2|d\omega|_{k}^{2} is the squared kk-norm of d(ω|M)d(\omega|_{M}), where dd is the exterior derivative on MM. Finally, LL is the second fundamental form of MM with respect to the inward-pointing normal r\frac{\partial}{\partial r}, and H=kμνLμνH=k^{\mu\nu}L_{\mu\nu}.

Lemma 5.3.

Let g¯\bar{g} be a smooth metric on Xn+1X^{n+1}, and let uu be the singular Yamabe solution for g¯\bar{g}, so that g=u2g¯g=u^{-2}\bar{g} has constant scalar curvature n(n+1)-n(n+1). Then u=ru~u=r\tilde{u}, where u~=1+O(r)\tilde{u}=1+O(r). Moreover,

ru~|r=0\displaystyle\partial_{r}\tilde{u}|_{r=0} =12nH and\displaystyle=-\frac{1}{2n}H\text{ and}
r2u~|r=0\displaystyle\partial_{r}^{2}\tilde{u}|_{r=0} =13n(R¯+H2)+13(n1)(R|L̊|k2).\displaystyle=-\frac{1}{3n}(\overline{R}+H^{2})+\frac{1}{3(n-1)}(R-|\mathring{L}|_{k}^{2}).

This is proved in [Gra17]. See in particular equation (2.6) and pages 1788-89.

Lemma 5.4.

Let g¯\bar{g} be a smooth metric on Xn+1X^{n+1}, and let gg be the corresponding singular Yamabe metric. Let r^\hat{r} be the geodesic defining function for MM (with respect to gg), as in Lemma 2.2, such that r^r|M=1\frac{\partial\hat{r}}{\partial r}|_{M}=1. Then if g¯^=r2g\hat{\bar{g}}=r^{2}g is the corresponding geodesic compactification of gg, we can write g¯^=e2ωg¯\hat{\bar{g}}=e^{2\omega}\bar{g}, where

ω|M\displaystyle\omega|_{M} =0\displaystyle=0
rω|M\displaystyle\partial_{r}\omega|_{M} =1nH\displaystyle=\frac{1}{n}H
r2ω|M\displaystyle\partial_{r}^{2}\omega|_{M} =1+n2n2H2+12nR¯12(n1)R+12(n1)|L̊|k2\displaystyle=\frac{1+n}{2n^{2}}H^{2}+\frac{1}{2n}\overline{R}-\frac{1}{2(n-1)}R+\frac{1}{2(n-1)}|\mathring{L}|_{k}^{2}
r3ω|M\displaystyle\partial_{r}^{3}\omega|_{M} =1nΔkH+1n2μνL̊μν+1n2L̊μνR¯μν2n2L̊μνRμν\displaystyle=-\frac{1}{n}\Delta_{k}H+\frac{1}{n-2}\nabla^{\mu}\nabla^{\nu}\mathring{L}_{\mu\nu}+\frac{1}{n-2}\mathring{L}^{\mu\nu}\overline{R}_{\mu\nu}-\frac{2}{n-2}\mathring{L}^{\mu\nu}R_{\mu\nu}
+12nrR¯+n2+2n+12n3H3+n+12n2HR¯n+22n(n1)HR\displaystyle\quad+\frac{1}{2n}\partial_{r}\overline{R}+\frac{n^{2}+2n+1}{2n^{3}}H^{3}+\frac{n+1}{2n^{2}}H\overline{R}-\frac{n+2}{2n(n-1)}HR
+3n24n22n(n1)(n20)H|L̊|k2.\displaystyle\quad+\frac{3n^{2}-4n-2}{2n(n-1)(n-20)}H|\mathring{L}|_{k}^{2}.
Proof.

Observe that the singular Yamabe function u=ru~u=r\tilde{u} is a defining function for MM. Thus, taking r0=ur_{0}=u in equation (2.2) of ([Gra00]) gives g¯^=e2ωg¯\hat{\bar{g}}=e^{2\omega}\bar{g} where

2(gradg¯u)(ω)+u|dω|g¯2=1|du|g¯2u.2(\operatorname{grad}_{\bar{g}}u)(\omega)+u|d\omega|_{\bar{g}}^{2}=\frac{1-|du|_{\bar{g}}^{2}}{u}.

We can write this equation as

2r2u~g¯ijiu~jω+2ru~2rω+r2u~2g¯ijiωjω=1u~22ru~u~r2g¯ijiu~ju~.2r^{2}\tilde{u}\bar{g}^{ij}\partial_{i}\tilde{u}\partial_{j}\omega+2r\tilde{u}^{2}\partial_{r}\omega+r^{2}\tilde{u}^{2}\bar{g}^{ij}\partial_{i}\omega\partial_{j}\omega=1-\tilde{u}^{2}-2r\tilde{u}\partial\tilde{u}-r^{2}\bar{g}^{ij}\partial_{i}\tilde{u}\partial_{j}\tilde{u}.

Now, tangential derivatives of both u~\tilde{u} and ω\omega vanish to first order, so we can rewrite this as

2r2u~ru~rω+2ru~2rω+r2u~2r(ω)2=1u~22ru~ru~r2r(u~)2+O(r4).2r^{2}\tilde{u}\partial_{r}\tilde{u}\partial_{r}\omega+2r\tilde{u}^{2}\partial_{r}\omega+r^{2}\tilde{u}^{2}\partial_{r}(\omega)^{2}=1-\tilde{u}^{2}-2r\tilde{u}\partial_{r}\tilde{u}-r^{2}\partial_{r}(\tilde{u})^{2}+O(r^{4}).

Differentiating gives

8ru~r(u~)r(ω)+2r2r(u~)2+2r2u~r2(u~)r(ω)+2r2u~r(u~)r2(ω)+2u~2r(ω)+2ru~2r2(ω)+2ru~2r(ω)2+2r2u~r(u~)r(ω)2+2r2u~2r(ω)r2(ω)=4u~r(u~)4r(u~)22ru~r2(u~)2r2r(u~)r2(u~)+O(r3).\begin{split}8r\tilde{u}\partial_{r}(\tilde{u})\partial_{r}(\omega)+&2r^{2}\partial_{r}(\tilde{u})^{2}+2r^{2}\tilde{u}\partial_{r}^{2}(\tilde{u})\partial_{r}(\omega)+2r^{2}\tilde{u}\partial_{r}(\tilde{u})\partial_{r}^{2}(\omega)\\ &+2\tilde{u}^{2}\partial_{r}(\omega)+2r\tilde{u}^{2}\partial_{r}^{2}(\omega)+2r\tilde{u}^{2}\partial_{r}(\omega)^{2}+2r^{2}\tilde{u}\partial_{r}(\tilde{u})\partial_{r}(\omega)^{2}\\ &+2r^{2}\tilde{u}^{2}\partial_{r}(\omega)\partial_{r}^{2}(\omega)=-4\tilde{u}\partial_{r}(\tilde{u})-4r\partial(\tilde{u})^{2}-2r\tilde{u}\partial_{r}^{2}(\tilde{u})\\ &-2r^{2}\partial_{r}(\tilde{u})\partial_{r}^{2}(\tilde{u})+O(r^{3}).\end{split}

Taking r=0r=0 and applying Lemma 5.3 gives

rω|r=0=1nH.\partial_{r}\omega|_{r=0}=\frac{1}{n}H.

Differentiating again mod O(r2)O(r^{2}), we find

12u~r(u~)r(ω)+12rr(u~)2r(ω)+12ru~r2(u~)r(ω)+12ru~r(u~)r2(ω)+4u~r(u~)r(ω)+4u~2r2(ω)+4ru~r(u~)r2(ω)+2ru~2r3(ω)+2u~2r(ω)2+8ru~r(u~)r(ω)2+8ru~2r(ω)r2(ω)=8r(u~)26u~r2u~14r(u~)r2(u~)2ru~r3(u~)+O(r2).\begin{split}12\tilde{u}\partial_{r}(\tilde{u})\partial_{r}(\omega)+&12r\partial_{r}(\tilde{u})^{2}\partial_{r}(\omega)+12r\tilde{u}\partial_{r}^{2}(\tilde{u})\partial_{r}(\omega)+12r\tilde{u}\partial_{r}(\tilde{u})\partial_{r}^{2}(\omega)\\ &+4\tilde{u}\partial_{r}(\tilde{u})\partial_{r}(\omega)+4\tilde{u}^{2}\partial_{r}^{2}(\omega)+4r\tilde{u}\partial_{r}(\tilde{u})\partial_{r}^{2}(\omega)\\ &+2r\tilde{u}^{2}\partial_{r}^{3}(\omega)+2\tilde{u}^{2}\partial_{r}(\omega)^{2}+8r\tilde{u}\partial_{r}(\tilde{u})\partial_{r}(\omega)^{2}\\ &+8r\tilde{u}^{2}\partial_{r}(\omega)\partial_{r}^{2}(\omega)=-8\partial_{r}(\tilde{u})^{2}-6\tilde{u}\partial_{r}^{2}\tilde{u}-14\partial_{r}(\tilde{u})\partial_{r}^{2}(\tilde{u})\\ &\quad-2r\tilde{u}\partial_{r}^{3}(\tilde{u})+O(r^{2}).\end{split}

and setting r=0r=0 and using Lemma 5.3 gives

r2ω|r=0=1n2nH2+12nR¯12(n1)R+12(n1)|L̊|k2.\partial_{r}^{2}\omega|_{r=0}=\frac{1-n}{2n}H^{2}+\frac{1}{2n}\overline{R}-\frac{1}{2(n-1)}R+\frac{1}{2(n-1)}|\mathring{L}|_{k}^{2}.

Finally, we differentiate again, mod O(r)O(r):

24r(u~)2r(ω)+24u~r2(u~)r(ω)+36u~r(u~)r2(ω)+6u~2r3(ω)+12u~r(u~)r(ω)2+12u~2r(ω)r2(ω)=30r(u~)r2(u~)6r(u~)r2(u~)8u~r2(u~)+O(r).\begin{split}24\partial_{r}(\tilde{u})^{2}\partial_{r}(\omega)+&24\tilde{u}\partial_{r}^{2}(\tilde{u})\partial_{r}(\omega)+36\tilde{u}\partial_{r}(\tilde{u})\partial_{r}^{2}(\omega)+6\tilde{u}^{2}\partial_{r}^{3}(\omega)\\ &+12\tilde{u}\partial_{r}(\tilde{u})\partial_{r}(\omega)^{2}+12\tilde{u}^{2}\partial_{r}(\omega)\partial_{r}^{2}(\omega)=-30\partial_{r}(\tilde{u})\partial_{r}^{2}(\tilde{u})\\ &-6\partial_{r}(\tilde{u})\partial_{r}^{2}(\tilde{u})-8\tilde{u}\partial_{r}^{2}(\tilde{u})+O(r).\end{split}

Taking r=0r=0 and using our prior results gives the claimed formula for r3ω|M\partial_{r}^{3}\omega|_{M}. ∎

We also record some additional elementary formulas. First, it follows from Gauss’s equation that

(5.1) L̊μνR¯0μν0=L̊μνR¯μνL̊μνRμν+n2nH|L̊|k2L̊3,\mathring{L}^{\mu\nu}\overline{R}_{0\mu\nu 0}=\mathring{L}^{\mu\nu}\overline{R}_{\mu\nu}-\mathring{L}^{\mu\nu}R_{\mu\nu}+\frac{n-2}{n}H|\mathring{L}|_{k}^{2}-\mathring{L}^{3},

where L̊3=L̊αβL̊βμL̊μα\mathring{L}^{3}=\mathring{L}_{\alpha}^{\beta}\mathring{L}_{\beta}^{\mu}\mathring{L}_{\mu}^{\alpha}. Next, it is easy to show using Codazzi’s equation that

(5.2) ¯rR¯00|r=0=12rR¯+1nnΔkH+μνL̊μνL̊μνR¯μν+n12nHR¯1+n2nHR1+n2nH|L̊|k2+n212n2H3.\begin{split}\overline{\nabla}_{r}\overline{R}_{00}|_{r=0}=&\frac{1}{2}\partial_{r}\overline{R}+\frac{1-n}{n}\Delta_{k}H+\nabla^{\mu}\nabla^{\nu}\mathring{L}_{\mu\nu}-\mathring{L}^{\mu\nu}\overline{R}_{\mu\nu}\\ &\quad+\frac{n-1}{2n}H\overline{R}-\frac{1+n}{2n}HR-\frac{1+n}{2n}H|\mathring{L}|_{k}^{2}+\frac{n^{2}-1}{2n^{2}}H^{3}.\end{split}

Here, \nabla is the Levi-Civita connection of kk.

We now begin to analyze the singular Yamabe solution.

As mentioned above, we let gg be the singular Yamabe metric corresponding to g¯\bar{g} on Xn+1X^{n+1}, and let r^\hat{r} be the geodesic defining function corresponding to kk, with g¯^=r^2g\hat{\bar{g}}=\hat{r}^{2}g. That is, g=dr^2+h^r^r^2g=\frac{d\hat{r}^{2}+\hat{h}_{\hat{r}}}{\hat{r}^{2}}, where h^r^\hat{h}_{\hat{r}} is a one-parameter family of smooth metric on MM with h^0=k\hat{h}_{0}=k. By Lemma 2.2, r^\hat{r} is CnC^{n}, and thus so is g¯^\hat{\bar{g}}. Also rg¯^r\hat{\bar{g}} is Cn+1C^{n+1}. Now, with respect to g¯^\hat{\bar{g}}, the singular Yamabe function u^\hat{u} one obtains by starting with g¯^\hat{\overline{g}} is r^\hat{r} itself; this follows because g=r^2g¯^g=\hat{r}^{-2}\hat{\bar{g}}, and the singular Yamabe function is unique.

We introduce some further notations related to g¯^\hat{\bar{g}}. The Ricci and scalar curvatures we will denote by R¯ij\overline{R}_{ij} and R¯\overline{R}, respectively. The second fundamental form with respect to g¯^\hat{\bar{g}} we will denote, for convenience, by L^\hat{L}. The tracefree part of this we will call L̊\mathring{L}, since it is a conformal invariant at the boundary and thus identical to the corresponding tensor for g¯\bar{g}, since g¯|TM=g¯^|TM=k\bar{g}|_{TM}=\hat{\bar{g}}|_{TM}=k. We use H^\hat{H} for the mean curvature kμνL^μνk^{\mu\nu}\hat{L}_{\mu\nu}. Meanwhile, we will continue to use RR to denote the intrinsic curvature RkR_{k} of the boundary metric kk.

Now, the equation satisfied by u^\hat{u}, i.e. the singular Yamabe equation, is given by

n(n+1)=n(n+1)|du|g¯^22nu^Δg¯^u^u^2R¯^.n(n+1)=n(n+1)|du|_{\hat{\bar{g}}}^{2}-2n\hat{u}\Delta_{\hat{\bar{g}}}\hat{u}-\hat{u}^{2}\hat{\overline{R}}.

(See (2.1).) Since u^=r^\hat{u}=\hat{r}, we have

0=2nΔg¯^r^+r^R¯^.0=2n\Delta_{\hat{\bar{g}}}\hat{r}+\hat{r}\hat{\overline{R}}.

Since Δg¯^r^=12h^μνh^μν\Delta_{\hat{\bar{g}}}\hat{r}=\frac{1}{2}\hat{h}^{\mu\nu}\hat{h}^{\prime}_{\mu\nu}, this can be re-expressed as

h^μνh^μν=1nr^R¯^.\hat{h}^{\mu\nu}\hat{h}^{\prime}_{\mu\nu}=-\frac{1}{n}\hat{r}\hat{\overline{R}}.

Taking r^=0\hat{r}=0 and applying Lemma 5.1 immediately gives

(5.3) H^=0.\hat{H}=0.

This implies also that L^=L̊\hat{L}=\mathring{L}. We differentiate, using the identity (hμν)=hμαhνβhαβ(h^{\mu\nu})^{\prime}=-h^{\mu\alpha}h^{\nu\beta}h^{\prime}_{\alpha\beta}, to find

(5.4) h^μαh^νβh^αβh^μν+h^μνh^μν′′=1nR¯^1nrr^R¯^.-\hat{h}^{\mu\alpha}\hat{h}^{\nu\beta}\hat{h}^{\prime}_{\alpha\beta}\hat{h}^{\prime}_{\mu\nu}+\hat{h}^{\mu\nu}\hat{h}^{\prime\prime}_{\mu\nu}=-\frac{1}{n}\hat{\overline{R}}-\frac{1}{n}r\partial_{\hat{r}}\hat{\overline{R}}.

Taking r^=0\hat{r}=0 and again using Lemma 5.1 gives

(5.5) 4|L^|k22R¯^00+2|L̊|k2=1nR¯^.-4|\hat{L}|^{2}_{k}-2\hat{\overline{R}}_{00}+2|\mathring{L}|_{k}^{2}=-\frac{1}{n}\hat{\overline{R}}.

(Recall that L^\hat{L} is the second fundamental form.) Now, it is easy to show, for any metric g¯\bar{g} on an (n + 1)-dimensional manifold, that

|L^|k2=|L̊|k2+1nH^2.|\hat{L}|_{k}^{2}=|\mathring{L}|_{k}^{2}+\frac{1}{n}\hat{H}^{2}.

Similarly, we have by Gauss’s formula that

R¯^00=12(R¯^R|L̊|k2+n1nH^2)\hat{\overline{R}}_{00}=\frac{1}{2}\left(\hat{\overline{R}}-R-|\mathring{L}|_{k}^{2}+\frac{n-1}{n}\hat{H}^{2}\right)

(see equation (4.3) in [Gra17], keeping in mind that that paper has a different convention for curvature indices). Applying these equations to g¯^\hat{\bar{g}}, and recalling that H^=0\hat{H}=0, we conclude from (5.5) that at r=0r=0,

(5.6) R¯^=nn1(R|L̊|k2).\hat{\overline{R}}=\frac{n}{n-1}(R-|\mathring{L}|_{k}^{2}).

Note that this result is good for n2n\geq 2, by Lemma 2.2. We differentiate (5.4) again, which we can do if n3n\geq 3 according to Lemma 2.2. We find

2h^μσh^αλh^σλh^αβh^μν3h^μαh^νβh^αβh^μν′′+h^μνh^μν′′′=2nr^R¯^1nr^r^2R¯^.2\hat{h}^{\mu\sigma}\hat{h}^{\alpha\lambda}\hat{h}^{\prime}_{\sigma\lambda}\hat{h}^{\prime}_{\alpha\beta}\hat{h}^{\prime}_{\mu\nu}-3\hat{h}^{\mu\alpha}\hat{h}^{\nu\beta}\hat{h}^{\prime}_{\alpha\beta}\hat{h}_{\mu\nu}^{\prime\prime}+\hat{h}^{\mu\nu}\hat{h}^{\prime\prime\prime}_{\mu\nu}=-\frac{2}{n}\partial_{\hat{r}}\hat{\overline{R}}-\frac{1}{n}\hat{r}\partial_{\hat{r}}^{2}\hat{\overline{R}}.

Taking r^=0\hat{r}=0 and applying Lemma 5.1, we get

4L^34L^μνR¯^0μν02¯^0R¯^00=2nr^R¯^.-4\hat{L}^{3}-4\hat{L}^{\mu\nu}\hat{\overline{R}}_{0\mu\nu 0}-2\hat{\overline{\nabla}}_{0}\hat{\overline{R}}_{00}=-\frac{2}{n}\partial_{\hat{r}}\hat{\overline{R}}.

Now using our previous calculations, (5.1) and (5.2), we find

(5.7) r^R¯^|r^=0=2nn2μνL̊μν+4nn2L̊μνRμν2nn2L̊μνR¯^μν.\partial_{\hat{r}}\hat{\overline{R}}|_{\hat{r}=0}=-\frac{2n}{n-2}\nabla^{\mu}\nabla^{\nu}\mathring{L}_{\mu\nu}+\frac{4n}{n-2}\mathring{L}^{\mu\nu}R_{\mu\nu}-\frac{2n}{n-2}\mathring{L}^{\mu\nu}\hat{\overline{R}}_{\mu\nu}.

Here, \nabla and RR are the connection and Ricci curvature, respectively, of kk. We leave hats off ,R\nabla,R, and L̊\mathring{L} because these are the same for g¯^\hat{\bar{g}} as for g¯\bar{g}.

Having derived the geometric consequences of using a geodesic compactification, we next compute Δg\Delta_{g} for any vC(X)v\in C^{\infty}(X), where (recall) gg is the singular Yamabe metric. We find

Δgv\displaystyle\Delta_{g}v =(detg)1/2i[(detg)1/2gijjv]\displaystyle=(\det g)^{-1/2}\partial_{i}\left[(\det g)^{1/2}g^{ij}\partial_{j}v\right]
=r^1+n(detg¯^)1/2i[r^1n(detg¯^)1/2g¯^ijjv]\displaystyle=\hat{r}^{1+n}(\det\hat{\bar{g}})^{-1/2}\partial_{i}\left[\hat{r}^{1-n}(\det\hat{\bar{g}})^{1/2}\hat{\bar{g}}^{ij}\partial_{j}v\right]
=r^2r^2v+(1n)r^r^v+12r^2h^μνh^μνr^v+r^2Δh^r^v.\displaystyle=\hat{r}^{2}\partial_{\hat{r}}^{2}v+(1-n)\hat{r}\partial_{\hat{r}}v+\frac{1}{2}\hat{r}^{2}\hat{h}^{\mu\nu}\hat{h}^{\prime}_{\mu\nu}\partial_{\hat{r}}v+\hat{r}^{2}\Delta_{\hat{h}_{\hat{r}}}v.

Now if fC(M)f\in C^{\infty}(M) and v=r^nsfv=\hat{r}^{n-s}f, we therefore find

(5.8) Δgv=s(sn)r^nsf+ns2r^ns+1h^μνh^μνf+r^ns+2Δh^r^f.\Delta_{g}v=s(s-n)\hat{r}^{n-s}f+\frac{n-s}{2}\hat{r}^{n-s+1}\hat{h}^{\mu\nu}\hat{h}^{\prime}_{\mu\nu}f+\hat{r}^{n-s+2}\Delta_{\hat{h}_{\hat{r}}}f.

We expand the various quantities that appear in this expression. First, a straightforward computation shows that

(5.9) Δh^r^f=Δkf+2r^[L̊μνμν2f+μL̊μνfν]+O(r^2).\Delta_{\hat{h}_{\hat{r}}}f=\Delta_{k}f+2\hat{r}\left[\mathring{L}^{\mu\nu}\nabla^{2}_{\mu\nu}f+\nabla^{\mu}\mathring{L}_{\mu}^{\nu}f_{\nu}\right]+O(\hat{r}^{2}).

Meanwhile, by iteratively applying the equation (h^μν)=h^μαh^νβh^αβ(\hat{h}^{\mu\nu})^{\prime}=-\hat{h}^{\mu\alpha}\hat{h}^{\nu\beta}\hat{h}^{\prime}_{\alpha\beta} along with Lemma 5.1 and equations (5.3) - (5.7), we find

h^μνh^μν=1n1r^(|L̊|k2R)+2n2r^2(μνL̊μν2L̊μνRμν+L̊μνR¯^μν).\hat{h}^{\mu\nu}\hat{h}^{\prime}_{\mu\nu}=\frac{1}{n-1}\hat{r}(|\mathring{L}|_{k}^{2}-R)+\frac{2}{n-2}\hat{r}^{2}(\nabla^{\mu}\nabla^{\nu}\mathring{L}_{\mu\nu}-2\mathring{L}^{\mu\nu}R_{\mu\nu}+\mathring{L}^{\mu\nu}\hat{\overline{R}}_{\mu\nu}).

Thus, we conclude from (5.8) and (5.9) that

(5.10) [Δg+s(ns)](r^nsf)=r^ns+2[ns2(n1)(|L̊|k2R)+Δkf]+r^ns+3[nsn2(μνL̊μν2L̊μνRμν+L̊μνR¯^μν)+2L̊μνμν2f+2μL̊μνfν]+O(r^ns+4).\begin{split}[\Delta_{g}+s(n-s)](\hat{r}^{n-s}f)=&\hat{r}^{n-s+2}\left[\frac{n-s}{2(n-1)}(|\mathring{L}|_{k}^{2}-R)+\Delta_{k}f\right]\\ &+\hat{r}^{n-s+3}\left[\frac{n-s}{n-2}(\nabla^{\mu}\nabla^{\nu}\mathring{L}_{\mu\nu}-2\mathring{L}^{\mu\nu}R_{\mu\nu}+\mathring{L}^{\mu\nu}\hat{\overline{R}}_{\mu\nu})\right.\\ &+\left.2\mathring{L}^{\mu\nu}\nabla^{2}_{\mu\nu}f+2\nabla^{\mu}\mathring{L}_{\mu}^{\nu}f_{\nu}\right]+O(\hat{r}^{n-s+4}).\end{split}

We now formally solve the equation (Δg+s(ns))v=0(\Delta_{g}+s(n-s))v=0. Let fC(M)f\in C^{\infty}(M) be arbitrary, and set v0=rnsfv_{0}=r^{n-s}f. We will perturb v0v_{0} at increasing orders to formally solve the equation. To do this, we compute the indicial operator Isj:C(M)C(M)I_{s}^{j}:C^{\infty}(M)\to C^{\infty}(M), which we define by

Isj(ψ)=r^(ns+j)[Δg+s(ns)](rns+jψ)|r^=0.I_{s}^{j}(\psi)=\hat{r}^{-(n-s+j)}[\Delta_{g}+s(n-s)](r^{n-s+j}\psi)|_{\hat{r}=0}.

This operator tells us the effect of a perturbation of vv at order r^ns+j\hat{r}^{n-s+j} on [Δg+s(ns)]v[\Delta_{g}+s(n-s)]v. It is easy to compute from (5.8) that

Isj(ψ)=j(n2s+j)ψ.I_{s}^{j}(\psi)=j(n-2s+j)\psi.

Now it follows from (5.10) that [Δg+s(ns)]v0=O(r^ns+2)[\Delta_{g}+s(n-s)]v_{0}=O(\hat{r}^{n-s+2}), so we want to perturb at order r^ns+2\hat{r}^{n-s+2}. Specifically, we wish to solve

Is2ψ2=(ns2(n1)(|L̊|k2R))f+Δkf,I_{s}^{2}\psi_{2}=-\left(\frac{n-s}{2(n-1)}(|\mathring{L}|_{k}^{2}-R)\right)f+\Delta_{k}f,

which gives

ψ2=1(n1)(n+22s)ns4(R|L̊|k2)f12(n+22s)Δkf.\psi_{2}=\frac{1}{(n-1)(n+2-2s)}\frac{n-s}{4}(R-|\mathring{L}|_{k}^{2})f-\frac{1}{2(n+2-2s)}\Delta_{k}f.

We therefore set v2=r^ns(f+r^2ψ2)v_{2}=\hat{r}^{n-s}(f+\hat{r}^{2}\psi_{2}). (We skip v1v_{1} in our numbering since there is no term of order r^ns+1\hat{r}^{n-s+1} in (5.10).) It can easily be shown from (5.10) that, apart from removing the order r^ns+2\hat{r}^{n-s+2} term from (Δg+s(ns))v0(\Delta_{g}+s(n-s))v_{0}, adding this perturbation to v0v_{0} has no other effects before order r^ns+4\hat{r}^{n-s+4}. Thus, the next equation we wish to solve is

Is3ψ3=snn2(μνL̊μν2L̊μνRμν+L̊μνR¯^μν)2L̊μνμν2f2μL̊μνfν.\begin{split}I_{s}^{3}\psi_{3}=&\frac{s-n}{n-2}(\nabla^{\mu}\nabla^{\nu}\mathring{L}_{\mu\nu}-2\mathring{L}^{\mu\nu}R_{\mu\nu}+\mathring{L}^{\mu\nu}\hat{\overline{R}}_{\mu\nu})\\ &-2\mathring{L}^{\mu\nu}\nabla^{2}_{\mu\nu}f-2\nabla^{\mu}\mathring{L}_{\mu}^{\nu}f_{\nu}.\end{split}

Since Is3=3(n+32s)I_{s}^{3}=3(n+3-2s), we obtain

(5.11) ψ3=ns3(n2)(n+32s)(μνL̊μν+2L̊μνRμνL̊μνR¯^μν)f23(n+32s)(L̊μνμν2f+μL̊μνfν).\begin{split}\psi_{3}=&\frac{n-s}{3(n-2)(n+3-2s)}(-\nabla^{\mu}\nabla^{\nu}\mathring{L}_{\mu\nu}+2\mathring{L}^{\mu\nu}R_{\mu\nu}-\mathring{L}^{\mu\nu}\hat{\overline{R}}_{\mu\nu})f\\ &-\frac{2}{3(n+3-2s)}(\mathring{L}^{\mu\nu}\nabla^{2}_{\mu\nu}f+\nabla^{\mu}\mathring{L}_{\mu}^{\nu}f_{\nu}).\end{split}

We set v3=v2+rns+3ψ3v_{3}=v_{2}+r^{n-s+3}\psi_{3}. It then follows from Proposition 3.6 in [GZ03] and Proposition 3.1 that

P^1SY\displaystyle\hat{P}_{1}^{SY} =0\displaystyle=0
P^2SY\displaystyle\hat{P}_{2}^{SY} =Δk+n24(n1)(R|L̊|k2)\displaystyle=-\Delta_{k}+\frac{n-2}{4(n-1)}(R-|\mathring{L}|_{k}^{2})
If n=2n=2,
Q^2SY\displaystyle\hat{Q}_{2}^{SY} =12(R|L̊|k2).\displaystyle=\frac{1}{2}(R-|\mathring{L}|_{k}^{2}).
For n3n\geq 3,
P^3SY\displaystyle\hat{P}_{3}^{SY} =L̊μνμν2f+μL̊μνfν+n34(n2)(μνL̊μν2L̊μνRμν+L̊μνR¯^μν)f.\displaystyle=\mathring{L}^{\mu\nu}\nabla^{2}_{\mu\nu}f+\nabla^{\mu}\mathring{L}_{\mu}^{\nu}f_{\nu}+\frac{n-3}{4(n-2)}\left(\nabla^{\mu}\nabla^{\nu}\mathring{L}_{\mu\nu}-2\mathring{L}^{\mu\nu}R_{\mu\nu}+\mathring{L}^{\mu\nu}\hat{\overline{R}}_{\mu\nu}\right)f.
And if n=3n=3,
Q^3SY\displaystyle\hat{Q}^{SY}_{3} =12μνL̊μνL̊μνRμν+12L̊μνR¯^μν.\displaystyle=\frac{1}{2}\nabla^{\mu}\nabla^{\nu}\mathring{L}_{\mu\nu}-\mathring{L}^{\mu\nu}R_{\mu\nu}+\frac{1}{2}\mathring{L}^{\mu\nu}\hat{\overline{R}}_{\mu\nu}.

Finally, we can use Lemma 5.3 to translate these results into formulas for our general metric g¯\bar{g}. We obtain the following.

Theorem 5.5.

For any dimension,

P2SYf\displaystyle P_{2}^{SY}f =Δkf+n24(n1)(Rk|L̊|k2)f(n2)\displaystyle=-\Delta_{k}f+\frac{n-2}{4(n-1)}(R_{k}-|\mathring{L}|_{k}^{2})f\quad(n\geq 2)
P3SYf\displaystyle P_{3}^{SY}f =L̊μνμν2f\displaystyle=\mathring{L}^{\mu\nu}\nabla^{2}_{\mu\nu}f
+μL̊μνfν+n34(n2)(μνL̊μν2L̊μνRμν+L̊μνR¯^μν+n1nH|L̊|k2)f.\displaystyle\quad+\nabla^{\mu}\mathring{L}_{\mu}^{\nu}f_{\nu}+\frac{n-3}{4(n-2)}\left(\nabla^{\mu}\nabla^{\nu}\mathring{L}_{\mu\nu}-2\mathring{L}^{\mu\nu}R_{\mu\nu}+\mathring{L}^{\mu\nu}\hat{\overline{R}}_{\mu\nu}+\frac{n-1}{n}H|\mathring{L}|_{k}^{2}\right)f.

If n=2n=2, then

Q2SY=12(Rk|L̊|k2).Q_{2}^{SY}=\frac{1}{2}(R_{k}-|\mathring{L}|_{k}^{2}).

For n=3n=3, we have the following.

Q3SY\displaystyle Q^{SY}_{3} =μνL̊μν2L̊μνRμν+L̊μνR¯μν+23H|L̊|k2.\displaystyle=\nabla^{\mu}\nabla^{\nu}\mathring{L}_{\mu\nu}-2\mathring{L}^{\mu\nu}R_{\mu\nu}+\mathring{L}^{\mu\nu}\overline{R}_{\mu\nu}+\frac{2}{3}H|\mathring{L}|_{k}^{2}.

Observe that P2SYP_{2}^{SY} is the conformal Laplacian plus an extrinsic pointwise conformal invariant.

We next wish to compute the coefficients in Theorem E in the cases n=2n=2 and n=3n=3 to express the renormalized volume in terms of the scattering operator. On the basis of (5.11) we have

v3=r^nsf+r^ns+2[ns4(n1)(n+22s)(R|L̊|k2)f12(n+22s)Δkf]+r^ns+3[ns3(n2)(n+32s)(μνL̊μν+2L̊μνRμνL̊μνR¯^μν)23(n+32s)(L̊μνμν2f+μL̊μνfν)].\begin{split}v_{3}=&\hat{r}^{n-s}f+\hat{r}^{n-s+2}\left[\frac{n-s}{4(n-1)(n+2-2s)}(R-|\mathring{L}|_{k}^{2})f-\frac{1}{2(n+2-2s)}\Delta_{k}f\right]\\ &+\hat{r}^{n-s+3}\left[\frac{n-s}{3(n-2)(n+3-2s)}(-\nabla^{\mu}\nabla^{\nu}\mathring{L}_{\mu\nu}+2\mathring{L}^{\mu\nu}R_{\mu\nu}-\mathring{L}^{\mu\nu}\hat{\overline{R}}_{\mu\nu})\right.\\ &-\left.\frac{2}{3(n+3-2s)}(\mathring{L}^{\mu\nu}\nabla_{\mu\nu}^{2}f+\nabla^{\mu}\mathring{L}_{\mu}^{\nu}f_{\nu})\right].\end{split}

Recall that rr (as opposed to r^\hat{r}) is the distance function from MM with respect to g¯\bar{g}. We wish to express v2v_{2} in terms of rr instead of r^\hat{r}, and for this purpose, we want to expand r^α\hat{r}^{\alpha} for a real number α\alpha. Let α\alpha\in\mathbb{R}. Recall that we defined u~\tilde{u} by u=ru~u=r\tilde{u}. Then

r^=eωu=reωu~,\hat{r}=e^{\omega}u=re^{\omega}\tilde{u},

where ω\omega is as in Lemma 5.4. Thus,

r^α=rαeα(ω+logu~).\hat{r}^{\alpha}=r^{\alpha}e^{\alpha(\omega+\log\tilde{u})}.

Now, u~=1+ru~r+12r2u~rr+\tilde{u}=1+r\tilde{u}_{r}+\frac{1}{2}r^{2}\tilde{u}_{rr}+\cdots, where we set u~r=ru~|r=0\tilde{u}_{r}=\partial_{r}\tilde{u}|_{r=0}, etc. A standard calculation shows that

logu~=ru~r+12r2(u~rru~r2)+16r3(u~rrr3u~ru~rr+2u~r3)+.\log\tilde{u}=r\tilde{u}_{r}+\frac{1}{2}r^{2}(\tilde{u}_{rr}-\tilde{u}_{r}^{2})+\frac{1}{6}r^{3}(\tilde{u}_{rrr}-3\tilde{u}_{r}\tilde{u}_{rr}+2\tilde{u}_{r}^{3})+\cdots.

Therefore,

ω+logu~=r(u~r+ωr)+12r2(u~rru~r2+ωrr)+16r3(u~rrr3u~ru~rr+2u~r3+ωrrr)+.\omega+\log\tilde{u}=r(\tilde{u}_{r}+\omega_{r})+\frac{1}{2}r^{2}(\tilde{u}_{rr}-\tilde{u}_{r}^{2}+\omega_{rr})+\frac{1}{6}r^{3}(\tilde{u}_{rrr}-3\tilde{u}_{r}\tilde{u}_{rr}+2\tilde{u}_{r}^{3}+\omega_{rrr})+\cdots.

We obtain

exp[α(ω+logu~)]\displaystyle\exp[\alpha(\omega+\log\tilde{u})] =1+r(αu~r+αωr)+\displaystyle=1+r(\alpha\tilde{u}_{r}+\alpha\omega_{r})+
+r2[α2u~rrα2u~r2+α2ωrr+α22u~r2+α2u~rωr+α22ωr2]\displaystyle\quad+r^{2}\left[\frac{\alpha}{2}\tilde{u}_{rr}-\frac{\alpha}{2}\tilde{u}_{r}^{2}+\frac{\alpha}{2}\omega_{rr}+\frac{\alpha^{2}}{2}\tilde{u}_{r}^{2}+\alpha^{2}\tilde{u}_{r}\omega_{r}+\frac{\alpha^{2}}{2}\omega_{r}^{2}\right]
+r3[α6u~rrrα2u~ru~rr+α3u~r3+α6ωrrr+α22u~ru~rr\displaystyle\quad+r^{3}\left[\frac{\alpha}{6}\tilde{u}_{rrr}-\frac{\alpha}{2}\tilde{u}_{r}\tilde{u}_{rr}+\frac{\alpha}{3}\tilde{u}_{r}^{3}+\frac{\alpha}{6}\omega_{rrr}+\frac{\alpha^{2}}{2}\tilde{u}_{r}\tilde{u}_{rr}\right.
α22u~r3+α22u~rωrr+α22u~rrωrα22u~r2ωr+α22ωrωrr\displaystyle\quad-\frac{\alpha^{2}}{2}\tilde{u}_{r}^{3}+\frac{\alpha^{2}}{2}\tilde{u}_{r}\omega_{rr}+\frac{\alpha^{2}}{2}\tilde{u}_{rr}\omega_{r}-\frac{\alpha^{2}}{2}\tilde{u}_{r}^{2}\omega_{r}+\frac{\alpha^{2}}{2}\omega_{r}\omega_{rr}
+α36u~r3+α32u~r2ωr+α36ωr3]+O(r4).\displaystyle\quad\left.+\frac{\alpha^{3}}{6}\tilde{u}_{r}^{3}+\frac{\alpha^{3}}{2}\tilde{u}_{r}^{2}\omega_{r}+\frac{\alpha^{3}}{6}\omega_{r}^{3}\right]+O(r^{4}).

Supposing now that α=ns+j\alpha=n-s+j (j0j\geq 0), and using Lemmas 5.3 and 5.4, a tedious calculation yields

r^ns+j\displaystyle\hat{r}^{n-s+j} =rns+j+rns+j+1(ns+j2nH)\displaystyle=r^{n-s+j}+r^{n-s+j+1}\left(\frac{n-s+j}{2n}H\right)
+rns+j+2[5n28sn+8jn+3n+3s26js3s+3j2+3j24n2H2\displaystyle\quad+r^{n-s+j+2}\left[\frac{5n^{2}-8sn+8jn+3n+3s^{2}-6js-3s+3j^{2}+3j}{24n^{2}}H^{2}\right.
+ns+j12nR¯ns+j12(n1)R+ns+j12(n1)|L̊|k2]\displaystyle\quad+\left.\frac{n-s+j}{12n}\overline{R}-\frac{n-s+j}{12(n-1)}R+\frac{n-s+j}{12(n-1)}|\mathring{L}|_{k}^{2}\right]
+rns+j+3[snj24nΔkH+ns+j24(n2)μνL̊μν+ns+j24(n2)L̊μνR¯μν\displaystyle\quad+r^{n-s+j+3}\left[\frac{s-n-j}{24n}\Delta_{k}H+\frac{n-s+j}{24(n-2)}\nabla^{\mu}\nabla^{\nu}\mathring{L}_{\mu\nu}+\frac{n-s+j}{24(n-2)}\mathring{L}^{\mu\nu}\overline{R}_{\mu\nu}\right.
(5.12) +snj12(n2)L̊μνRμν+ns+j48nrR¯\displaystyle\quad+\frac{s-n-j}{12(n-2)}\mathring{L}^{\mu\nu}R_{\mu\nu}+\frac{n-s+j}{48n}\partial_{r}\overline{R}
+(ns+j)(4n24sn+4jn+6n+s22js3s+j2+3j+2)48n3H3\displaystyle\quad+\frac{(n-s+j)(4n^{2}-4sn+4jn+6n+s^{2}-2js-3s+j^{2}+3j+2)}{48n^{3}}H^{3}
+(ns+j)(3n2s+2j+2)48n2HR¯(ns+j)(3n2s+2j+3)48n(n1)HR\displaystyle\quad+\frac{(n-s+j)(3n-2s+2j+2)}{48n^{2}}H\overline{R}-\frac{(n-s+j)(3n-2s+2j+3)}{48n(n-1)}HR
+(ns+j)(5n22sn+2jn7n+4s4j4)48n(n1)(n2)H|L̊|k2]+O(rns+j+4).\displaystyle\quad+\left.\frac{(n-s+j)(5n^{2}-2sn+2jn-7n+4s-4j-4)}{48n(n-1)(n-2)}H|\mathring{L}|_{k}^{2}\right]+O(r^{n-s+j+4}).

So for n=2n=2, we get

r^2s=r2s(1+2s4rH+r2[2619s+3s296H2+2s24R¯s212R+2s12|L̊|k2]).\hat{r}^{2-s}=r^{2-s}\left(1+\frac{2-s}{4}rH+r^{2}\left[\frac{26-19s+3s^{2}}{96}H^{2}+\frac{2-s}{24}\overline{R}-\frac{s-2}{12}R+\frac{2-s}{12}|\mathring{L}|_{k}^{2}\right]\right).

Now taking f=1f=1 and n=2n=2, we find

v2\displaystyle v_{2} =r2s+2s4r3sH+r4s[3s219s+2696H2+2s24R¯+s212R+2s12|L̊|k2]\displaystyle=r^{2-s}+\frac{2-s}{4}r^{3-s}H+r^{4-s}\left[\frac{3s^{2}-19s+26}{96}H^{2}+\frac{2-s}{24}\overline{R}+\frac{s-2}{12}R+\frac{2-s}{12}|\mathring{L}|_{k}^{2}\right]
+O(r5s).\displaystyle+O(r^{5-s}).

So

a1(s)\displaystyle a_{1}(s) =2s4H\displaystyle=\frac{2-s}{4}H
a2(s)\displaystyle a_{2}(s) =3s219s+2696H2+2s24R¯+s212R+2s12|L̊|k2.\displaystyle=\frac{3s^{2}-19s+26}{96}H^{2}+\frac{2-s}{24}\overline{R}+\frac{s-2}{12}R+\frac{2-s}{12}|\mathring{L}|_{k}^{2}.

Thus,

a1(2)\displaystyle a_{1}^{\prime}(2) =14H\displaystyle=-\frac{1}{4}H
a2(2)\displaystyle a_{2}^{\prime}(2) =796H2124R¯+112R112|L̊|k2.\displaystyle=-\frac{7}{96}H^{2}-\frac{1}{24}\overline{R}+\frac{1}{12}R-\frac{1}{12}|\mathring{L}|_{k}^{2}.

Now, by ([GG19]), we have

(5.13) v(1)\displaystyle v^{(1)} =1n2nH\displaystyle=\frac{1-n}{2n}H
(5.14) v(2)\displaystyle v^{(2)} =n512(n1)(R|L̊|k2)+n224n2((n3)H22nR¯).\displaystyle=\frac{n-5}{12(n-1)}(R-|\mathring{L}|_{k}^{2})+\frac{n-2}{24n^{2}}\left((n-3)H^{2}-2n\overline{R}\right).

So in this case, we have v(1)=14Hv^{(1)}=-\frac{1}{4}H. Here v(1)v^{(1)} is the first renormalized volume coefficient, not our solution vv to the scattering equation. Thus, using Theorem E, we find that for n=2n=2,

(5.15) V(X,g,g¯)=M(dds|s=2S(s)1)𝑑vk196M(8R4R¯8|L̊|k23H2)𝑑vk.V(X,g,\bar{g})=-\oint_{M}\left(\left.\frac{d}{ds}\right|_{s=2}S(s)1\right)dv_{k}-\frac{1}{96}\oint_{M}(8R-4\overline{R}-8|\mathring{L}|_{k}^{2}-3H^{2})dv_{k}.

Observe that if gg happens to be Einstein and g¯\bar{g} is a geodesic compactification (so that g¯=g¯^\bar{g}=\hat{\bar{g}} – which in this context is smooth), then the last integral vanishes, since L̊\mathring{L} and HH vanish in this case, and 2RR¯=02R-\overline{R}=0 (by (5.6)). Thus, this result is consistent with Theorem 4.1 of [CQY08].

We next turn to n=3n=3. Using (5.12) for j=0,1j=0,1, and 22 sequentially, and putting these into the formula for v3v_{3}, as well as using Lemma 5.2, we get (with f=1f=1)

(5.16) v3=r3s+r4s(3s6H)+r5s[5427s+3s2216H2+3s36R¯+(s3)(1s)12(52s)R+(s3)(s1)12(52s)|L̊|k2]+r6s[s372ΔkHs+124μνL̊μνs+124L̊μνR¯μν+s+112L̊μνRμν+3s144rR¯+(3s)(5615s+s2)1296H3+(3s)(112s)432HR¯+(s3)(2s214s+15)144(52s)HR2s328s2+69s25144(52s)H|L̊|k2]+O(r7s).\begin{split}v_{3}=&r^{3-s}+r^{4-s}\left(\frac{3-s}{6}H\right)\\ &+r^{5-s}\left[\frac{54-27s+3s^{2}}{216}H^{2}+\frac{3-s}{36}\overline{R}+\frac{(s-3)(1-s)}{12(5-2s)}R+\frac{(s-3)(s-1)}{12(5-2s)}|\mathring{L}|_{k}^{2}\right]\\ &+r^{6-s}\left[\frac{s-3}{72}\Delta_{k}H-\frac{s+1}{24}\nabla^{\mu}\nabla^{\nu}\mathring{L}_{\mu\nu}-\frac{s+1}{24}\mathring{L}^{\mu\nu}\overline{R}_{\mu\nu}+\frac{s+1}{12}\mathring{L}^{\mu\nu}R_{\mu\nu}\right.\\ &+\frac{3-s}{144}\partial_{r}\overline{R}+\frac{(3-s)(56-15s+s^{2})}{1296}H^{3}+\frac{(3-s)(11-2s)}{432}H\overline{R}\\ &+\left.\frac{(s-3)(2s^{2}-14s+15)}{144(5-2s)}HR-\frac{2s^{3}-28s^{2}+69s-25}{144(5-2s)}H|\mathring{L}|_{k}^{2}\right]+O(r^{7-s}).\end{split}

It quickly follows that

a1(3)\displaystyle a_{1}^{\prime}(3) =16H\displaystyle=-\frac{1}{6}H
a2(3)\displaystyle a_{2}^{\prime}(3) =124H2136R¯+16R16|L̊|k2\displaystyle=-\frac{1}{24}H^{2}-\frac{1}{36}\overline{R}+\frac{1}{6}R-\frac{1}{6}|\mathring{L}|_{k}^{2}
a3(3)\displaystyle a_{3}^{\prime}(3) =172ΔkH124μνL̊μν124L̊μνR¯μν+112L̊μνRμν1144rR¯\displaystyle=\frac{1}{72}\Delta_{k}H-\frac{1}{24}\nabla^{\mu}\nabla^{\nu}\mathring{L}_{\mu\nu}-\frac{1}{24}\mathring{L}^{\mu\nu}\overline{R}_{\mu\nu}+\frac{1}{12}\mathring{L}^{\mu\nu}R_{\mu\nu}-\frac{1}{144}\partial_{r}\overline{R}
5324H35432HR¯+116HR13144H|L̊|k2.\displaystyle\quad-\frac{5}{324}H^{3}-\frac{5}{432}H\overline{R}+\frac{1}{16}HR-\frac{13}{144}H|\mathring{L}|_{k}^{2}.

Using the formulae (5.13) for the renormalized volume coefficients, we finally obtain

V(X,g,g¯)\displaystyle V(X,g,\bar{g}) =M[dds|s=3S(s)113432HR+51296HR¯+1162H3+25432H|L̊|k2\displaystyle=\oint_{M}\left[-\left.\frac{d}{ds}\right|_{s=3}S(s)1-\frac{13}{432}HR+\frac{5}{1296}H\overline{R}+\frac{1}{162}H^{3}+\frac{25}{432}H|\mathring{L}|_{k}^{2}\right.
172ΔkH+124μνL̊μν+124L̊μνR¯μν+112L̊μνRμν+1144rR¯]dvk.\displaystyle\quad\left.-\frac{1}{72}\Delta_{k}H+\frac{1}{24}\nabla^{\mu}\nabla^{\nu}\mathring{L}_{\mu\nu}+\frac{1}{24}\mathring{L}^{\mu\nu}\overline{R}_{\mu\nu}+\frac{1}{12}\mathring{L}^{\mu\nu}R_{\mu\nu}+\frac{1}{144}\partial_{r}\overline{R}\right]dv_{k}.

Theorem G and Corollary H now follow directly from this computation and the main result of [GG19].

Appendix A Analysis

We present here the proof of Lemma 2.2 and sketch the proof of Theorem 4.1. Both entail adapting standard results for smooth AH metrics to the polyhomogeneous setting with more careful discussion of regularity.

A.I. Normal Form

For (X,g¯)(X,\bar{g}) a smooth manifold with boundary, let rr be the distance function to the boundary. We let Diffb(X)\operatorname{Diff}_{b}(X) be the ring of differential operators generated by vector fields VV that are tangent to the boundary. In local coordinates (r,xμ)(r,x^{\mu}) (1μn1\leq\mu\leq n) near the boundary MM, Diffb(X)\operatorname{Diff}_{b}(X) is generated over CC^{\infty} by rrr\frac{\partial}{\partial r} and xμ\frac{\partial}{\partial x^{\mu}}.

For p2p\geq 2, we let 𝒟p\mathcal{D}_{p} be the conormal functions

𝒟p(X)={uCp(X):LuCp(X) for all LDiffb(X)}.\mathcal{D}_{p}(X)=\left\{u\in C^{p}(X):Lu\in C^{p}(X)\text{ for all }L\in\operatorname{Diff}_{b}(X)\right\}.

The examples of interest to us are smooth functions and those with asymptotic expansions in rr and rqlog(r)r^{q}\log(r), qp+1q\geq p+1, with smooth coefficients. The following lemma is easy.

Lemma A.1.

Let ω𝒟p\omega\in\mathcal{D}_{p}. Then rω𝒟p+1r\omega\in\mathcal{D}_{p+1}.

We prove the following variation on the existence and uniqueness of solutions to first-order scalar noncharacteristic PDEs. We let x0=rx^{0}=r, and locally extend any coordinate chart (x1,,xn)(x^{1},\cdots,x^{n}) on a neighborhood UMU\subseteq M to be coordinates along with rr on a neighborhood U~\widetilde{U} of UU in XX, by the geodesic identification U~[0,ε)r×U\widetilde{U}\approx[0,\varepsilon)_{r}\times U, where \approx denotes diffemorphism. We then let (ξ0,,ξn)(\xi_{0},\cdots,\xi_{n}) be the corresponding natural coordinates on TXT^{*}X on U~\widetilde{U}.

Proposition A.2.

Let FCp(TX×)F\in C^{p}(T^{*}X\times\mathbb{R}) be such that, for any smooth one-form ηΩ1(X),uC(X)\eta\in\Omega^{1}(X),u\in C^{\infty}(X), the function xF(x,η(x),u(x))𝒟p(X)x\mapsto F(x,\eta(x),u(x))\in\mathcal{D}_{p}(X), where p2p\geq 2. Suppose that, in any coordinate system as above, Fξ0|M>0\left.\frac{\partial F}{\partial\xi_{0}}\right|_{M}>0. Finally, suppose given φ,τC(M)\varphi,\tau\in C^{\infty}(M) such that, for every qMq\in M, F(q,τdr+dφ,φ)=0F(q,\tau dr+d\varphi,\varphi)=0. Then there exists a neighborhood 𝒱\mathcal{V} of MM and a unique solution ω\omega to F(x,dω,ω)=0F(x,d\omega,\omega)=0 on 𝒱\mathcal{V} such that ω|M=φ\omega|_{M}=\varphi. Moreover, ω𝒟p\omega\in\mathcal{D}_{p}.

Proof.

Everything except the last statement is standard. We present an adaptation of the usual proof which preserves conormality. We work locally in a coordinate chart. Recall that the ordinary proof relies on converting the PDE to a first-order system of ODEs representing a characteristic flow off of MM, and parametrized by time tt. We first eliminate tt and replace it by the first coordinate rr, which is permissible because rt|t=0>0\frac{\partial r}{\partial t}|_{t=0}>0 by the noncharacteristic hypothesis. We thus begin by considering the following system of ODEs.

dξjdr\displaystyle\frac{d\xi_{j}}{dr} =Fxj+ξjFyFξ0(0jn)\displaystyle=\frac{\frac{\partial F}{\partial x^{j}}+\xi_{j}\frac{\partial F}{\partial y}}{\frac{\partial F}{\partial\xi_{0}}}\quad(0\leq j\leq n)
dxμdr\displaystyle\frac{dx^{\mu}}{dr} =FξμFξ0(1μn)\displaystyle=\frac{\frac{\partial F}{\partial\xi_{\mu}}}{\frac{\partial F}{\partial\xi_{0}}}\quad(1\leq\mu\leq n)
dydr\displaystyle\frac{dy}{dr} =ξjFξjFξ0,\displaystyle=-\xi_{j}\frac{\frac{\partial F}{\partial\xi_{j}}}{\frac{\partial F}{\partial\xi_{0}}},

with initial conditions xμ=ζμx^{\mu}=\zeta^{\mu} (some ζμ\zeta^{\mu}), ξμ(0)=μφ(ζ1,,ζn)\xi_{\mu}(0)=\partial_{\mu}\varphi(\zeta^{1},\cdots,\zeta^{n}), y(0)=φ(ζ1,,ζn)y(0)=\varphi(\zeta^{1},\cdots,\zeta^{n}), and with ξ0(0)=τ\xi_{0}(0)=\tau. Letting z=(x1,,xn,ξ0,,ξn,y)z=(x^{1},\cdots,x^{n},\xi_{0},\cdots,\xi_{n},y), this system may be written

dzdr=G(r,z),\frac{dz}{dr}=G(r,z),

where by hypothesis GG is Cp1C^{p-1}. Moreover, for any multi-index α\alpha, Gxα\frac{\partial G}{\partial x^{\alpha}} is Cp1C^{p-1} (where α\alpha contains no 0’s, i.e., no rr-derivatives); and so likewise is (rr)kG(r\partial_{r})^{k}G for any kk.

The system is equivalent to the integral equation

(A.1) z=z0+0rG(s,z(s))𝑑s,z=z_{0}+\int_{0}^{r}G(s,z(s))ds,

where z0z_{0} contains the initial values written above. This equation, of course, has a C1C^{1} solution by standard theory. Now we view zz as a function both of rr and of (ζ1,,ζn)(\zeta^{1},\cdots,\zeta^{n}). Formally differentiating (A.1), we get

ddrzζμ=z0ζμ+0rGzazaζμ\frac{d}{dr}\frac{\partial z}{\partial\zeta^{\mu}}=\frac{\partial z_{0}}{\partial\zeta^{\mu}}+\int_{0}^{r}\frac{\partial G}{\partial z^{a}}\frac{\partial z^{a}}{\partial\zeta^{\mu}}

(where we use the index aa as an index for the components of zz). Now, the integrand here is still Cp1C^{p-1} by hypothesis; and so by a standard argument, and its accompanying induction (see, e.g., section 13 of [Wal98]), |α|zζα\frac{\partial^{|\alpha|}z}{\partial\zeta^{\alpha}} is CpC^{p}. On the other hand, using the identity

ddr(rdzdr)=rd2zdr2+dzdr\frac{d}{dr}(r\frac{dz}{dr})=r\frac{d^{2}z}{dr^{2}}+\frac{dz}{dr}

and (A.1), along with the easily-verified fact that H=0rG(s,z)𝑑sH=\int_{0}^{r}G(s,z)ds is CpC^{p} along with (rddr)kH\left(r\frac{d}{dr}\right)^{k}H for any kk, we conclude by induction that (rr)k(|α|ζα)z\left(r\frac{\partial}{\partial r}\right)^{k}\left(\frac{\partial^{|\alpha|}}{\partial\zeta^{\alpha}}\right)z is CpC^{p}.

We now have functions x1(r,ζ1,,ζn),,xn(r,ζ1,,ζn)x^{1}(r,\zeta^{1},\cdots,\zeta^{n}),\cdots,x^{n}(r,\zeta^{1},\cdots,\zeta^{n}). We wish to invert the dependence of xx on ζ\zeta. For notational simplicity, we assume by restriction and rescaling if necessary that ζ1,,ζn\zeta^{1},\cdots,\zeta^{n} take values in all of n\mathbb{R}^{n}. Let Φ(r,ζ1,,ζn)=(x1,,xn)\Phi(r,\zeta^{1},\cdots,\zeta^{n})=(x^{1},\cdots,x^{n}), and define Zr0:nnZ_{r_{0}}:\mathbb{R}^{n}\to\mathbb{R}^{n} by Zr0(ζ)=Φ(r0,ζ)Z_{r_{0}}(\zeta)=\Phi(r_{0},\zeta). Now Φ\Phi is CpC^{p}, and DZ0DZ_{0} is the identity, so by restricting r0r_{0} if necessary, we may assume that ZrZ_{r} is invertible. In fact, for each r0r_{0} it is a CC^{\infty} diffeomorphism onto its image, and thus has CC^{\infty} inverse Zr01Z_{r_{0}}^{-1}. That this inverse is CpC^{p} in rr follows from the implicit function theorem.

To show that (rr)kZr1(r\partial_{r})^{k}Z_{r}^{-1} is likewise CpC^{p}, we write

Φ(r,Zr1(x1,,xn))=(x1,,xn),\Phi(r,Z_{r}^{-1}(x^{1},\cdots,x^{n}))=(x^{1},\cdots,x^{n}),

and then differentiate both sides with respect to rr. We obtain

Φr+DζZr(Z1xμxμr+Zr1r)=(x1r,,xnr).\frac{\partial\Phi}{\partial r}+D_{\zeta}Z_{r}\left(\frac{\partial Z^{-1}}{\partial x^{\mu}}\frac{\partial x^{\mu}}{\partial r}+\frac{\partial Z_{r}^{-1}}{\partial r}\right)=\left(\frac{\partial x^{1}}{\partial r},\cdots,\frac{\partial x^{n}}{\partial r}\right).

Since DζZrD_{\zeta}Z_{r} is invertible, we can solve this for Zr1r\frac{\partial Z_{r}^{-1}}{\partial r}, and in particular exhibit rZr1rr\frac{\partial Z_{r}^{-1}}{\partial r} as a combination of terms already known to be CpC^{p}.

Thus, if we set ω(r,x)=y(r,Zr1(x))\omega(r,x)=y(r,Z_{r}^{-1}(x)), then ω\omega is a solution of our PDE – just as in the usual case, it is a solution at r=0r=0, and we show that ddrF(x,dω,ω)=0\frac{d}{dr}F(x,d\omega,\omega)=0. Moreover, ω𝒟p\omega\in\mathcal{D}_{p}. ∎

We now can prove Lemma 2.2.

Proof of Lemma 2.2.

Let rr be the distance to MM with respect to g¯\bar{g}. We wish to find r^\hat{r} so that |dr^|r^2g1|d\hat{r}|_{\hat{r}^{2}g}\equiv 1 on some neighborhood of the boundary. As in the usual proof (e.g. [Gra00]), write r^=ueω=ru~eω,\hat{r}=ue^{\omega}=r\tilde{u}e^{\omega}, where uu is the singular Yamabe function and u~=ur\tilde{u}=\frac{u}{r}. Recall that u~=1+O(r)\tilde{u}=1+O(r). We thus wish to find ω\omega so that ω|M=0\omega|_{M}=0 and so that

2gradg¯(u)(ω)+u|dω|g¯2=1|du|g¯2u.2\operatorname{grad}_{\bar{g}}(u)(\omega)+u|d\omega|_{\bar{g}}^{2}=\frac{1-|du|_{\bar{g}}^{2}}{u}.

In the smooth case, one directly solves this by observing that it is a first-order noncharacteristic equation. However, doing so in this form would not give optimal regularity, which we want. Using the fact that u=ru~u=r\tilde{u}, we can rewrite this as follows:

(A.2) 2rω+2ru~1g¯iju~iωj+rg¯ijωiωj=1u~2r2|du~|g¯22ru~ru~ru~2.2\partial_{r}\omega+2r\tilde{u}^{-1}\bar{g}^{ij}\tilde{u}_{i}\omega_{j}+r\bar{g}^{ij}\omega_{i}\omega_{j}=\frac{1-\tilde{u}^{2}-r^{2}|d\tilde{u}|_{\bar{g}}^{2}-2r\tilde{u}\partial_{r}\tilde{u}}{r\tilde{u}^{2}}.

The right-hand side is Cn1C^{n-1}, and in particular, taking r=0r=0, we may solve for rω|r=0\partial_{r}\omega|_{r=0}. We can then differentiate (A.2) iteratively, and at each stage, rω\partial_{r}\omega is expressed in terms of already-determined quantities. Observe, however, that by (1.3), the right-hand side contains a term of the form 2rnlog(r)2\mathcal{L}r^{n}\log(r). We can handle this by adding a term of the form Arn+1log(r)Ar^{n+1}\log(r) to our formal expansion of ω\omega (AC(M)A\in C^{\infty}(M)), and iterating this procedure, we can find some ω0\omega^{0} of the form

ω0=a1r+a2r2++anrn+Arn+1log(r)+an+1rn+1+,\omega^{0}=a_{1}r+a_{2}r^{2}+\cdots+a_{n}r^{n}+Ar^{n+1}\log(r)+a_{n+1}r^{n+1}+\cdots,

where each ajC(M)a_{j}\in C^{\infty}(M), and such that equation (A.2) is satisfied through order O(rn+1)O(r^{n+1}) by ω0\omega^{0}. Now set ω=ω0+Ω\omega=\omega^{0}+\Omega, and substitute this into (A.2). We obtain the equation

2rΩ+2ru~1g¯iju~iΩj+2rg¯ijωi0Ωj+rg¯ijΩiΩj=O(rn+1),2\partial_{r}\Omega+2r\tilde{u}^{-1}\bar{g}^{ij}\tilde{u}_{i}\Omega_{j}+2r\bar{g}^{ij}\omega^{0}_{i}\Omega_{j}+r\bar{g}^{ij}\Omega_{i}\Omega_{j}=O(r^{n+1}),

where the right-hand side, in particular, is a function in 𝒟n+1\mathcal{D}_{n+1}. Then since rui,rωi0Cnru_{i},r\omega^{0}_{i}\in C^{n}, this is a noncharacteristic first-order equation with CnC^{n} coefficients, and in fact the differential operator is nn-conormal. Thus, by Proposition A.2, there is a unique solution Ω𝒟n\Omega\in\mathcal{D}_{n}, and so we get a unique solution ω=ω0+Ω\omega=\omega^{0}+\Omega to (A.2), and ω𝒟n\omega\in\mathcal{D}_{n} as well. Consequently, r^=ru~eω𝒟n+1\hat{r}=r\tilde{u}e^{\omega}\in\mathcal{D}_{n+1}. We construct the diffeomorphism ψ\psi, as always, by following the flow lines of gradg¯r^\operatorname{grad}_{\bar{g}}\hat{r}. This is a CnC^{n} vector field, so the result follows. ∎

Observe that we actually show rather more than claimed – specifically, that the diffeomorphism is conormal – but the lemma is all we need for our purposes.

A.II. Meromorphic Extension of the Resolvent

Theorem 4.1 is proved in [GZ03] in the context of smooth AH metrics. The proof proceeds by first constructing an infinite-order formal solution ufu_{f} satisfying

(Δg+s(ns))uf=O(r),(\Delta_{g}+s(n-s))u_{f}=O(r^{\infty}),

and then using the following theorem from [MM87] (see also [Gui05]).

Theorem A.3.

Let (X,g)(X,g) be an asymptotically hyperbolic manifold. Then the resolvent R(s)=(Δg+s(ns))1:Lg2(X̊)Lg2(X̊)R(s)=(\Delta_{g}+s(n-s))^{-1}:L^{2}_{g}(\mathring{X})\to L^{2}_{g}(\mathring{X}) for s>ns>n has an extension R(s):rC(X)rsC(X)R(s):r^{\infty}C^{\infty}(X)\to r^{s}C^{\infty}(X) that is holomorphic on Γ\mathbb{C}\setminus\Gamma with Γ\Gamma\subset\mathbb{C} discrete. In particular, R(s)R(s) is meromorphic on the right half-plane.

Thus, we need the same theorem in the case where gg is not smooth AH, but rather polyhomogeneous, and where the codomain of R(s)R(s) may likewise be only polyhomogeneous. Fortunately, the needed modifications to the proof are slight. The proof of [MM87] proceeds by first proving the result on hyperbolic space, and in fact obtaining an explicit formula for the Schwartz kernel of the resolvent there. On a general asymptotically hyperbolic space, the proof proceeds in three steps. First, ordinary elliptic analysis is used to produce an interior solution with poor boundary regularity; then, the “normal operator” of the Laplacian at a fixed point of the boundary is analyzed, and shown to coincide with the Laplacian on hyperbolic space, so that the result there can be used to obtain better regularity. Finally, the indicial operator at each point is used to obtain optimal regularity via formal expansion and Borel’s lemma. The first two steps go through exactly the same if the metric gg is polyhomogeneous and (say) CnC^{n}. And the last step, likewise, will be the same except that logarithmic terms from the metric may appear on the right hand side of the order-by-order construction, and need to be corrected by including logarithmic terms in the solution. The order at which they appear can be computed formally using the indicial operator.

In particular, the construction of the scattering operator proceeds exactly as in [GZ03], except that because smooth compactifications of the metric have a log term appearing of the form rn+1log(r)r^{n+1}\log(r), expansions of 𝒫(s)f\mathcal{P}(s)f starting at rnsr^{n-s} will have a term of the form rns+(n+1)log(r)r^{n-s+(n+1)}\log(r); to avoid the appearance of a log term at or before the second indicial root ss, therefore (which would complicate the analysis), we require s<2ns+1s<2n-s+1, or s<n+12s<n+\frac{1}{2}.

References

  • [ACF92] L. Andersson, P. T. Chruściel, and H. Friederich. On the regularity of solutions to the Yamabe equation and the existence of smooth hyperboloidal initial data for Einstein’s field equations. Comm. Math. Phys., 149(3):587–612, 1992.
  • [AM88] P. Aviles and R. C. McOwen. Complete conformal metrics with negative scalar curvature in compact Riemannian manifolds. Duke Math. J., 56(2):395–398, 1988.
  • [CdMG11] S.-Y. A. Chang and M. d. M. Gonzalez. Fractional Laplacian in conformal geometry. Advances in Mathematics, 226(2):1410–1432, 2011.
  • [CQY08] S.-Y. A. Chang, J. Qing, and P. Yang. Renormalized volumes for conformally compact Einstein manifolds. J. Math. Sci. (N.Y.), 149(6):1755–1769, 2008.
  • [FG02] C. Fefferman and C. R. Graham. qq-curvature and Poincaré metrics. Math. Res. Lett., 9:139–151, 2002.
  • [GG19] C. R. Graham and M. J. Gursky. Chern-Gauss-Bonnet formula for singular Yamabe metrics in dimension four. preprint. arXiv:1902.01562, pages 1–26, 2019.
  • [GL91] C. R. Graham and J. M. Lee. Einstein metrics with prescribed conformal infinity on the ball. Adv. Math., 87:186 – 225, 1991.
  • [Gra00] C. R. Graham. Volume and area renormalizations for conformally compact Einstein metrics. Suppl. Rendiconti Circolo Mat. Palermo, 63:31–42, 2000.
  • [Gra17] C. R. Graham. Volume renormalization for singular Yamabe metrics. Proceedings of the AMS, 145(4):1781–1792, 2017.
  • [Gui05] C. Guillarmou. Meromorophic properties of the resolvent on asymptotically hyperbolic manifolds. Duke Math. J., 129(1):1–37, 2005.
  • [GW14] A. R. Gover and A. Waldron. Boundary calculus for conformally compact manifolds. Indiana Univ. Math. J., 63(1):119–163, 2014.
  • [GW15] A. R. Gover and A. Waldron. Conformal hypersurface geometry via a boundary Loewner-Nirenberg-Yamabe problem. Preprint:arXiv:1506.02723, 2015.
  • [GW17] A. R. Gover and A. Waldron. Renormalized volume. Comm. Math. Phys., 354(3):1205–1244, 2017.
  • [GZ03] C. R. Graham and M. Zworski. Scattering matrix in conformal geometry. Invent. Math., 152:89 – 118, 2003.
  • [HS98] M. Hennington and K. Skenderis. The holographic Weyl anomaly. Journal of High Energy Physics, 7:23, 1998.
  • [JO21] A. Juhl and B. Orsted. Residue families, singular Yamabe problems and extrinsic conformal Laplacians. preprint: arXiv: 2101.09027, pages 1–115, 2021.
  • [LN74] C. Loewner and L. Nirenberg. Partial differential equations invariant under conformal or projective transformations. In Contributions to analysis (a collection of papers dedicated to Lipman Bers), pages 245–272. Academic Press, New York, 1974.
  • [McK18] S. McKeown. Formal theory of cornered asymptotically hyperbolic Einstein metrics. J. Geom. Anal. (to appear), pages 1–53, 2018. arXiv:1708.02390.
  • [MM87] R. R. Mazzeo and R. B. Melrose. Meromorphic extension of the resolvent on complete spaces with asymptotically constant negative curvature. J. Func. Anal., 75:260 – 310, 1987.
  • [Wal98] W. Walter. Ordinary Differential Equations. Springer, New York, 1998.