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

Extrinsic Curvature and Conformal Gauss-Bonnet for Four-Manifolds with Corner

Stephen E. McKeown Department of Mathematical Sciences, FO 35; University of Texas at Dallas; 800 W. Campbell Road; Richardson, TX 75080; USA [email protected]
Abstract.

This paper defines two new extrinsic curvature quantities on the corner of a four-dimensional Riemannian manifold with corner. One of these is a pointwise conformal invariant, and the conformal transformation of the other is governed by a new linear second-order pointwise conformally invariant partial differential operator. The Gauss-Bonnet theorem is then stated in terms of these quantities.

2010 Mathematics Subject Classification:
Primary 53C40, 53C18; Secondary 58J99.
Research partially supported by NSF RTG Grant DMS-1502525.

1. Introduction

The extension of the Gauss-Bonnet theorem to manifolds of higher even dimension than two was one of the great achievements of mid-twentieth century geometry. (See [Fen40, All42, AW43, Che44, Che45] and the classic exposition [Spi99, Volume V].) Nevertheless, the Pfaffian is a distinctly more complicated object than the Gaussian curvature, and mining the theorem for its geometric consequences has long been a challenging task.

One successful application in dimension four has been the use of the theorem to identify and analyze conformally invariant differential operators and associated curvature quantities. It was first observed in [BG94] that the Gauss-Bonnet formula on a closed Riemannian four-manifold (X,g)(X,g) could be rewritten as

4π2χ(X)=X(18|Wg|g2+12Q)𝑑vg,4\pi^{2}\chi(X)=\oint_{X}\left(\frac{1}{8}|W_{g}|_{g}^{2}+\frac{1}{2}Q\right)dv_{g},

where WW is the Weyl curvature tensor and QQ is the so-called QQ-curvature (see [BØ91]). The first term in the integrand yields a pointwise conformal invariant, so this furnishes a proof that XQ𝑑vg\oint_{X}Qdv_{g} is a global conformal invariant. One can also calculate that, under the conformal change g~=e2ωg\tilde{g}=e^{2\omega}g, the QQ-curvature transforms pointwise by

(1.1) e4ωQ~=Q+P4ω,e^{4\omega}\widetilde{Q}=Q+P_{4}\omega,

where P4P_{4} is the Paneitz operator, a linear elliptic fourth-order PDO that satisfies the pointwise conformal invariance property P~4=e4ωP4\widetilde{P}_{4}=e^{-4\omega}P_{4}. Thus, this form of the Gauss-Bonnet formula offers a tractable link between conformal geometry and topology. The relationship between the Gauss-Bonnet formula and QQ-curvature motivated much important geometric analysis, including Alexakis’s structure theorem on global conformal invariants ([Ale12]), the conformally invariant sphere theorem [CGY03], and the topological result in [Gur99], among numerous others.

Motivated by the work of Branson and Gilkey, Chang and Qing in [CQ97] showed that the Chern-Gauss-Bonnet formula for compact Riemannian four-manifolds with boundary can also be written in a form that interacts in a particularly nice way with conformal geometry, and in so doing, they identified important new boundary curvature quantities. To be specific, let (X4,g)(X^{4},g) be such a manifold with boundary M3=XM^{3}=\partial X and induced metric h=g|TMh=g|_{TM}. The result of Chang and Qing was that the Gauss-Bonnet formula in this setting can be written

(1.2) 4π2χ(X)=X(18|Wg|g2+12Q)𝑑vg+M(+T)𝑑vh,4\pi^{2}\chi(X)=\int_{X}\left(\frac{1}{8}|W_{g}|_{g}^{2}+\frac{1}{2}Q\right)dv_{g}+\oint_{M}(\mathcal{L}+T)dv_{h},

where \mathcal{L} is an extrinsic pointwise conformal invariant of weight 3-3 along the boundary, and TT is an extrinsic third-order curvature quantity along MM which, under conformal transformations, transforms by

(1.3) e3ωT~=T+P3ω,e^{3\omega}\widetilde{T}=T+P_{3}\omega,

where P3:C(X)C(M)P_{3}:C^{\infty}(X)\to C^{\infty}(M) is a linear third-order differential operator that is pointwise conformally invariant of weight 3-3. Both the curvature quantity TT and the formula (1.2) have since proved important in conformal geometry. For example, they were used in ([CQY08]) to give an analysis of the renormalized volume on asymptotically hyperbolic Einstein manifolds; were used in [Ndi08] and [Ndi09] to study Escobar-Yamabe type problems for QQ-curvature on manifolds with boundary; yielded the important defect theorem for total QQ-curvature in [CQY00]; and have proven important in the analysis of compactness for conformally compact Einstein metrics (see [CGQ18]). A Sobolev trace inequality was proved in [AC17]. There have been numerous further applications besides. (See also [Juh09].)

In light of the importance of the T/P3T/P_{3} pair, there has been interest in finding generalizations to other situations. For example, Case and Luo in [CL19], along with Gover and Peterson in [GP18], defined fifth-order conformally covariant boundary operators and associated fifth-order extrinsic boundary curvatures (as well as various lower orders) such that the operators parametrized conformal changes in the curvatures, whose total boundary integral was conformally invariant and which, when paired with the sixth-order GJMS operator (the sixth-order generalization of the Paneitz operator) yield a symmetric pair. The former paper also proves a Sobolev trace theorem.

In this paper, I extend the result of [CQ97] in a different direction: I consider four-manifolds with corners of codimension two, and in particular, derive an order-two conformal curvature/operator pair along the corner. I then show that the Gauss-Bonnet formula may be written on such a manifold in terms of these quantities and a pointwise conformal invariant. The proof is based on the Allendoerfer-Weil formula of [AW43]. One forthcoming application is discussed below; other expected applications are to extend theorems and arguments in the conformal geometry of four-manifolds to situations with higher boundary codimension.

To state the results cleanly, we let XX be a compact four-manifold with corners, having two embedded boundary hypersurfaces M3M^{3} and N3N^{3}, whose common boundary Σ2=MN\Sigma^{2}=M\cap N is a smooth closed surface. (Nothing prevents XX from having more than one corner component, or a mixture of closed and boundaried boundary hypersurfaces. The modifications necessary in that case will be obvious, but it will be notationally much cleaner to consider this simpler situation.) We assume that MM and NN meet transversely, at an angle which may vary but which lies at each point between 0 and π\pi.

For any three-dimensional hypersurface, we let hh denote the induced metric, μ\mu denote an inward-pointing unit vector, LL the second fundamental form computed with respect to this vector, and H=hμνLμνH=h^{\mu\nu}L_{\mu\nu} the mean curvature. For MM and NN respectively, we will decorate these various terms with a subscript to indicate the hypersurface intended; so, for example, we will have hM,μM,LMh_{M},\mu_{M},L_{M}, and HMH_{M}.

The QQ-curvature is defined on XX by

6Q=ΔgR+R23RijRij,6Q=-\Delta_{g}R+R^{2}-3R^{ij}R_{ij},

where RR here denotes the scalar or Ricci curvatures of gg and the Laplacian is a negative operator. The TT-curvature along a hypersurface, meanwhile, is defined by

T\displaystyle T =112μ(Rg)L̊μνRμνg+L̊μνRμνh12H|L̊|h2+23L̊3\displaystyle=-\frac{1}{12}\mu(R_{g})-\mathring{L}^{\mu\nu}R_{\mu\nu}^{g}+\mathring{L}^{\mu\nu}R_{\mu\nu}^{h}-\frac{1}{2}H|\mathring{L}|_{h}^{2}+\frac{2}{3}\mathring{L}^{3}
+16HRh127H313ΔhH.\displaystyle\quad+\frac{1}{6}HR_{h}-\frac{1}{27}H^{3}-\frac{1}{3}\Delta_{h}H.

The \mathcal{L}-curvature along a hypersurface is

=L̊μνRμνg2L̊μνRμνh+23H|L̊|h2L̊3.\mathcal{L}=\mathring{L}^{\mu\nu}R_{\mu\nu}^{g}-2\mathring{L}^{\mu\nu}R_{\mu\nu}^{h}+\frac{2}{3}H|\mathring{L}|_{h}^{2}-\mathring{L}^{3}.

For ease of reference, we also record the formula

P3gu=12μΔguΔhμ(u)HΔhuL̊μνμhνhu13Hμuμ+(16Rg12Rh12|L̊|h2+13H2)μ(u).\begin{split}P_{3}^{g}u=&\frac{1}{2}\mu\Delta_{g}u-\Delta_{h}\mu(u)-H\Delta_{h}u-\mathring{L}^{\mu\nu}\nabla_{\mu}^{h}\nabla_{\nu}^{h}u-\frac{1}{3}H^{\mu}u_{\mu}\\ &+\left(\frac{1}{6}R_{g}-\frac{1}{2}R_{h}-\frac{1}{2}|\mathring{L}|_{h}^{2}+\frac{1}{3}H^{2}\right)\mu(u).\end{split}

The corresponding quantities on MM and NN will be denoted by TM,MT_{M},\mathcal{L}_{M}, etc.

Now define θ0C(Σ)\theta_{0}\in C^{\infty}(\Sigma) by allowing θ0\theta_{0}, at each point pΣp\in\Sigma, to be the angle at which MM and NN meet at pp. That is, θ0\theta_{0} is defined by 0<θ0<π0<\theta_{0}<\pi and cosθ0=μM,μNg\cos\theta_{0}=-\langle\mu_{M},\mu_{N}\rangle_{g}. Let k=g|TΣk=g|_{T\Sigma}, and let νM,νN\nu_{M},\nu_{N} be the inward-pointing unit normal vectors to Σ\Sigma in MM and NN, respectively. Let IIMII_{M}, IINII_{N} be the second fundamental forms of Σ\Sigma regarded as a submanifold of MM or NN, and computed with respect to νM,νN\nu_{M},\nu_{N}. Let KK be the Gaussian curvature of Σ\Sigma. Finally, let ηM,ηN\eta_{M},\eta_{N} be the mean curvatures of Σ\Sigma regarded as a hypersurface in MM or NN, respectively.

Define UC(Σ)U\in C^{\infty}(\Sigma) by

(1.4) U=(πθ0)K14cot(θ0)(ηM2+ηN2)+12csc(θ0)ηMηN13(νMHM+νNHN).U=(\pi-\theta_{0})K-\frac{1}{4}\cot(\theta_{0})(\eta_{M}^{2}+\eta_{N}^{2})+\frac{1}{2}\csc(\theta_{0})\eta_{M}\eta_{N}-\frac{1}{3}(\nu_{M}H_{M}+\nu_{N}H_{N}).

Now define GC(Σ)G\in C^{\infty}(\Sigma) by

(1.5) G=12cot(θ0)(|II̊M|k2+|II̊N|k2)csc(θ0)II̊αβMII̊Nαβ.G=\frac{1}{2}\cot(\theta_{0})(|\mathring{II}_{M}|_{k}^{2}+|\mathring{II}_{N}|_{k}^{2})-\csc(\theta_{0})\mathring{II}^{M}_{\alpha\beta}\mathring{II}_{N}^{\alpha\beta}.

Finally, define a corner differential operator P2b:C(X)C(Σ)P_{2}^{b}:C^{\infty}(X)\to C^{\infty}(\Sigma) by

(1.6) P2bu=(θ0π)Δku+νMμMu+νNμNu+cot(θ0)(ηMνMu+ηNνNu)csc(θ0)(ηNνMu+ηMνNu)+13(HMνMu+HNνNu).\begin{split}P_{2}^{b}u=&(\theta_{0}-\pi)\Delta_{k}u+\nu_{M}\mu_{M}u+\nu_{N}\mu_{N}u\\ &+\cot(\theta_{0})(\eta_{M}\nu_{M}u+\eta_{N}\nu_{N}u)-\csc(\theta_{0})(\eta_{N}\nu_{M}u+\eta_{M}\nu_{N}u)\\ &+\frac{1}{3}(H_{M}\nu_{M}u+H_{N}\nu_{N}u).\end{split}

We then have the following results.

Theorem 1.1.

Let (X,g)(X,g) be a compact Riemannian manifold with codimension-two corner Σ2=MN\Sigma^{2}=M\cap N, where MM and NN are the two embedded hypersurfaces composing X\partial X. We suppose that M,NM,N meet transversely and at positive angle less than π\pi (which angle may vary along Σ\Sigma). Then under the conformal change g~=e2ωg\tilde{g}=e^{2\omega}g, we find

(1.7) G~\displaystyle\widetilde{G} =e2ωG,\displaystyle=e^{-2\omega}G,
(1.8) P~2bu\displaystyle\widetilde{P}_{2}^{b}u =e2ωP2bu,\displaystyle=e^{-2\omega}P_{2}^{b}u,
and
(1.9) e2ωU~\displaystyle e^{2\omega}\widetilde{U} =U+P2bω\displaystyle=U+P_{2}^{b}\omega
Theorem 1.2.

Let X,M,Q,Σ,X,M,Q,\Sigma, and gg satisfy the same hypotheses as Theorem 1.1. Then

(1.10) 4π2χ(X4)=X(18|Wg|g2+12Q)𝑑vg+M(M+TM)𝑑vhM+N(N+TN)𝑑vhN+Σ(G+U)𝑑vk.\begin{split}4\pi^{2}\chi(X^{4})=&\int_{X}\left(\frac{1}{8}|W_{g}|_{g}^{2}+\frac{1}{2}Q\right)dv_{g}+\int_{M}(\mathcal{L}_{M}+T_{M})dv_{h_{M}}\\ &+\int_{N}(\mathcal{L}_{N}+T_{N})dv_{h_{N}}+\oint_{\Sigma}(G+U)dv_{k}.\end{split}

Both theorems are proved in Section 2.2.

In a companion paper in preparation, Matthew Gursky, Aaron Tyrrell, and I will use this to obtain a Gauss-Bonnet formula for asymptotically hyperbolic Einstein manifolds with minimal surfaces meeting the infinite boundary. This generalizes the result of [And03, CQY08] to the setting of [GW99].

2. Background and Proof

2.1. The Allendoerfer-Weil Formula

We begin with the theorem of Allendoerfer-Weil, proved in [AW43]. The proof will simply involve a rewriting of what their formula says, so we take some time to introduce their notation, or a slightly simplified version of it. The setting of their theorem is a Riemannian polyhedron, or a compact Riemannian manifold with corners of arbitrary codimension, having non-reentrant angles.

Throughout the paper, the Laplacian Δg\Delta_{g} will be a negative operator, and the sign of the Riemman tensor will be such that Rij=gklRikljR_{ij}=g^{kl}R_{iklj}.

Let (Mn,g)(M^{n},g) be any Riemannian nn-manifold, and MrM^{r} an embedded rr-dimensional submanifold, with induced metric γ\gamma. Near a point in MrM^{r}, we choose coordinates {xi}i=1n\left\{x^{i}\right\}_{i=1}^{n} on MM, such that {xμ}μ=1r\left\{x^{\mu}\right\}_{\mu=1}^{r} restrict to coordinates on MrM^{r}. Let σ\sigma be a normal vector to MrM^{r} in MnM^{n}. Following [AW43] (here and throughout this subsection), we define locally along MrM^{r} the quantity

Λμν(σ)=σiΓμνi.\Lambda_{\mu\nu}(\sigma)=-\sigma_{i}\Gamma_{\mu\nu}^{i}.

Thus, Λ\Lambda is essentially the negative of the second fundamental form in the σ\sigma direction.

Now for 02pr0\leq 2p\leq r and for ζMr\zeta\in M^{r}, define

Φr,p(ζ,σ)=(1)p122pp!(r2p)!(detγ)1εμ1μrεν1νrRμ1μ2ν1ν2Rμ2p1μ2pν2p1ν2pΛμ2p+1ν2p+1(σ)Λμrνr(σ).\begin{split}\Phi_{r,p}(\zeta,\sigma)=&(-1)^{p}\frac{1}{2^{2p}p!(r-2p)!}(\det\gamma)^{-1}\varepsilon^{\mu_{1}\ldots\mu_{r}}\varepsilon^{\nu_{1}\ldots\nu_{r}}R_{\mu_{1}\mu_{2}\nu_{1}\nu_{2}}\cdots\\ &\cdot R_{\mu_{2p-1}\mu_{2p}\nu_{2p-1}\nu_{2p}}\Lambda_{\mu_{2p+1}\nu_{2p+1}}(\sigma)\cdots\Lambda_{\mu_{r}\nu_{r}}(\sigma).\end{split}

Here, RR is the curvature tensor of gg on MnM^{n}, while ε\varepsilon is the Levi-Civita symbol. (Note that the factor (1)p(-1)^{p} is not present in the original paper [AW43], but is here because I am using the opposite sign convention for the curvature tensor.)

For ζMr\zeta\in M^{r}, let Sζnr1S_{\zeta}^{n-r-1} be the unit normal sphere bundle in NζMrTζMnN_{\zeta}M^{r}\subset T_{\zeta}M^{n}. Let ξSζnr1\xi\in S_{\zeta}^{n-r-1}, and let dξd\xi be the area element on Sζnr1S_{\zeta}^{n-r-1}. We define

Ψ(ζ|Mr)=πn/2Γ(n2)2p=0r2Φr,p(ζ,ξ)(n2)(n4)(n2p)dξ,\Psi(\zeta|M^{r})=\frac{\pi^{-n/2}\Gamma\left(\frac{n}{2}\right)}{2}\sum_{p=0}^{\lfloor\frac{r}{2}\rfloor}\frac{\Phi_{r,p}(\zeta,\xi)}{(n-2)(n-4)\cdots(n-2p)}d\xi,

which is a volume form on Sζnr1S_{\zeta}^{n-r-1}. Now, for ζMr\zeta\in M^{r}, let Γ(ζ)Sζnr1\Gamma(\zeta)\subset S_{\zeta}^{n-r-1} be the set of unit normal vectors that have negative inner product with all inward-pointing normal vectors (including normal vectors tangent to the boundary). This is a spherical cell, bounded by great spheres. We call Γ(ζ)\Gamma(\zeta) the outer angle at ζ\zeta.

Meanwhile, on the interior of MnM^{n}, we define ΨΩn(Mn)\Psi\in\Omega^{n}(M^{n}) by

(2.1) Ψ(z)=(2π)n/212n(n/2)!det(g)1εi1inεj1jnRi1i2j1j2Rin1injn1jndvg,\Psi(z)=(-2\pi)^{-n/2}\frac{1}{2^{n}(n/2)!}\det(g)^{-1}\varepsilon^{i_{1}\ldots i_{n}}\varepsilon^{j_{1}\ldots j_{n}}R_{i_{1}i_{2}j_{1}j_{2}}\cdots R_{i_{n-1}i_{n}j_{n-1}j_{n}}dv_{g},

if nn is even, and by Ψ=0\Psi=0 if nn is odd.

Let PnP^{n} be a compact Riemannian manifold with corners of arbitrary codimension; we may write P=r=0n1λ=1NrPλr\partial P=\cup_{r=0}^{n-1}\cup_{\lambda=1}^{N_{r}}P_{\lambda}^{r}, where PλrP_{\lambda}^{r} is an rr-dimensional manifold with corners, and if λλ\lambda\neq\lambda^{\prime}, then PλrPλrP_{\lambda}^{r}\cap P_{\lambda^{\prime}}^{r} contains only manifolds (with corner) of dimension less than rr.

Theorem 2.1 ([AW43], Theorem II).

With PnP^{n} and {Pλr}\{P_{\lambda}^{r}\} as above, we have

(2.2) (1)nχ(Pn)=PnΨ+r=0n1λPλrΓ(ζ)Ψ(ζ|Pλr)𝑑vγλr(ζ).(-1)^{n}\chi^{\prime}(P^{n})=\int_{P^{n}}\Psi+\sum_{r=0}^{n-1}\sum_{\lambda}\int_{P_{\lambda}^{r}}\int_{\Gamma(\zeta)}\Psi(\zeta|P_{\lambda}^{r})dv_{\gamma_{\lambda}^{r}}(\zeta).

The quantity χ(Pn)\chi^{\prime}(P^{n}) is the interior Euler characteristic, i.e., the Euler characteristic computed using only fully interior cells. For a four-manifold with corners, it coincides with the usual Euler characteristic. (See [Che45] and pp. 154-55 of [Lef56]. The distinction is important in [AW43] because that paper considers Riemannian polyhedra more general than manifolds with corner.)

2.2. Proofs of Theorems

To prove the Gauss-Bonnet formula, we rewrite the terms appearing in (2.2) in terms of more geometric quantities.

Proof of Theorem 1.2.

First, let us set some notation. Recall that we are working on a compact four-manifold X4X^{4} with X=M3N3\partial X=M^{3}\cup N^{3}, where M,NM,N are embedded submanifolds with boundary and M=MN=N=Σ2\partial M=M\cap N=\partial N=\Sigma^{2}, a closed embedded submanifold of (co)dimension two. We will use indices 1i,j41\leq i,j\leq 4 on XX, with 1μ,ν31\leq\mu,\nu\leq 3 on MM or NN, and 1α,β21\leq\alpha,\beta\leq 2 on Σ\Sigma.

Let us begin by briefly computing Ψ(z)\Psi(z); although the answer is of course the best-known among the quantities above, it will be useful to work through the computation and to get the constants right. For any metric gg, a well-known formula tells us that

(detg)1εi1inεj1jn=det(gikjl)k,l=1n.(\det g)^{-1}\varepsilon^{i_{1}\ldots i_{n}}\varepsilon^{j_{1}\ldots j_{n}}=\det\left(g^{i_{k}j_{l}}\right)_{k,l=1}^{n}.

Taking n=4n=4, a straightforward calculation shows that

εi1i4εj1j4Ri1i2j1j2gRi3i4j3j4g=4Rg216RgijRijg+4RgijklRijklg.\varepsilon^{i_{1}\ldots i_{4}}\varepsilon^{j_{1}\ldots j_{4}}R_{i_{1}i_{2}j_{1}j_{2}}^{g}R_{i_{3}i_{4}j_{3}j_{4}}^{g}=4R_{g}^{2}-16R_{g}^{ij}R_{ij}^{g}+4R^{ijkl}_{g}R_{ijkl}^{g}.

Now, again for n=4n=4, we define the Schouten tensor PijgP_{ij}^{g} by

Pijg=12Rij112Rggij,P_{ij}^{g}=\frac{1}{2}R_{ij}-\frac{1}{12}R_{g}g_{ij},

and define JJ by

J=gijPijg.J=g^{ij}P^{g}_{ij}.

Recall next that the Riemann curvature tensor of gg satisfies

Rijklg=Wijklg+Pilgjk+PjkgilPikgjlPjlgik,R_{ijkl}^{g}=W_{ijkl}^{g}+P_{il}g_{jk}+P_{jk}g_{il}-P_{ik}g_{jl}-P_{jl}g_{ik},

where WW is the Weyl tensor. Putting all these and (2.1) together, we finally obtain

Ψ(z)=(2π)2[18|Wg|g2|Pg|g2+Jg2]dvg.\Psi(z)=(2\pi)^{-2}\left[\frac{1}{8}|W_{g}|_{g}^{2}-|P_{g}|_{g}^{2}+J_{g}^{2}\right]dv_{g}.

But Q=2(Jg2|Pg|g2)16ΔgRgQ=2(J_{g}^{2}-|P_{g}|_{g}^{2})-\frac{1}{6}\Delta_{g}R_{g}, so we obtain

(2.3) XΨ(z)=14π2X(18|Wg|g2+12Q+112ΔgRg)𝑑vg.\int_{X}\Psi(z)=\frac{1}{4\pi^{2}}\int_{X}\left(\frac{1}{8}|W_{g}|_{g}^{2}+\frac{1}{2}Q+\frac{1}{12}\Delta_{g}R_{g}\right)dv_{g}.

We next turn to computing Ψ3,0\Psi_{3,0} along M3XM^{3}\subset X. (We choose MM just for definiteness – the formula of course will also hold along NN.) Let, for now, h=g|TMh=g|_{TM}. For σNM\sigma\in NM and X,YTMX,Y\in TM all at the same point, let LM(σ)(X,Y)=XgY,σgL_{M}(\sigma)(X,Y)=\langle\nabla_{X}^{g}Y,\sigma\rangle_{g} be the second fundamental form (which of course does not depend on the extension of YY chosen). Thus, Λij(σ)=L(σ)ij\Lambda_{ij}(\sigma)=-L(\sigma)_{ij}. Now, recall that for any 3×33\times 3 symmetric matrix AA,

det(A)=16[tr(A)33tr(A)|A|2+2tr(A3)].\det(A)=\frac{1}{6}\left[\operatorname{tr}(A)^{3}-3\operatorname{tr}(A)|A|^{2}+2\operatorname{tr}(A^{3})\right].

(This may be proved, for example, by diagonalizing.) We thus get, along MM,

Ψ3,0(ζ,σ)\displaystyle\Psi_{3,0}(\zeta,\sigma) =16det(h)εi1i2i3εj1j2j3Λi1j1(σ)Λi2j2(σ)Λi3j3(σ)\displaystyle=\frac{1}{6\det(h)}\varepsilon^{i_{1}i_{2}i_{3}}\varepsilon^{j_{1}j_{2}j_{3}}\Lambda_{i_{1}j_{1}}(\sigma)\Lambda_{i_{2}j_{2}}(\sigma)\Lambda_{i_{3}j_{3}}(\sigma)
=1det(h)det(Λij(σ))\displaystyle=\frac{1}{\det(h)}\det(\Lambda_{ij}(\sigma))
=det(h1L(σ))\displaystyle=-\det(h^{-1}L(\sigma))
=16[trh(L)3+3trh(L)|L|h2+2trh(L3)]\displaystyle=\frac{1}{6}\left[-\operatorname{tr}_{h}(L)^{3}+3\operatorname{tr}_{h}(L)|L|_{h}^{2}+2\operatorname{tr}_{h}(L^{3})\right]
=16H|L̊|h213trh(L̊3)127H3,\displaystyle=\frac{1}{6}H|\mathring{L}|_{h}^{2}-\frac{1}{3}\operatorname{tr}_{h}(\mathring{L}^{3})-\frac{1}{27}H^{3},

where H=hμνLμνH=h^{\mu\nu}L_{\mu\nu}, L̊\mathring{L} is the tracefree part of LL, and the last step follows by expanding the previous line.

Next we consider Ψ3,1\Psi_{3,1} along M3M^{3}. Recalling that Λ=L\Lambda=-L, we apply Gauss’s formula and find

Φ3,1(ζ,σ)\displaystyle\Phi_{3,1}(\zeta,\sigma) =14[hμ1ν1hμ2ν2hμ3ν3hμ1ν1hμ3ν2hμ2ν3hμ1ν2hμ2ν1hμ3ν3\displaystyle=\frac{1}{4}\left[h^{\mu_{1}\nu_{1}}h^{\mu_{2}\nu_{2}}h^{\mu_{3}\nu_{3}}-h^{\mu_{1}\nu_{1}}h^{\mu_{3}\nu_{2}}h^{\mu_{2}\nu_{3}}-h^{\mu_{1}\nu_{2}}h^{\mu_{2}\nu_{1}}h^{\mu_{3}\nu_{3}}\right.
+hμ1ν2hμ3ν1hμ2ν3+hμ1ν3hμ2ν1hμ3ν2hμ1ν3hμ3ν1hμ2ν2]Rμ1μ2ν1ν2hLμ3ν3\displaystyle\quad\left.+h^{\mu_{1}\nu_{2}}h^{\mu_{3}\nu_{1}}h^{\mu_{2}\nu_{3}}+h^{\mu_{1}\nu_{3}}h^{\mu_{2}\nu_{1}}h^{\mu_{3}\nu_{2}}-h^{\mu_{1}\nu_{3}}h^{\mu_{3}\nu_{1}}h^{\mu_{2}\nu_{2}}\right]R_{\mu_{1}\mu_{2}\nu_{1}\nu_{2}}^{h}L_{\mu_{3}\nu_{3}}
14εμ1μ2μ3εν1ν2ν3(Lμ1ν2Lμ2ν1Lμ3ν3Lμ1ν1Lμ2ν2Lμ3ν3)\displaystyle\quad-\frac{1}{4}\varepsilon^{\mu_{1}\mu_{2}\mu_{3}}\varepsilon^{\nu_{1}\nu_{2}\nu_{3}}(L_{\mu_{1}\nu_{2}}L_{\mu_{2}\nu_{1}}L_{\mu_{3}\nu_{3}}-L_{\mu_{1}\nu_{1}}L_{\mu_{2}\nu_{2}}L_{\mu_{3}\nu_{3}})
=14[HRh+LμνRμνhHRh+3LμνRμνh]+3det(L)\displaystyle=\frac{1}{4}\left[-HR_{h}+L^{\mu\nu}R_{\mu\nu}^{h}-HR_{h}+3L^{\mu\nu}R_{\mu\nu}^{h}\right]+3\det(L)
=L̊μνRμνh16HRh+trh(L̊3)+19H312H|L̊|h2.\displaystyle=\mathring{L}^{\mu\nu}R_{\mu\nu}^{h}-\frac{1}{6}HR_{h}+\operatorname{tr}_{h}(\mathring{L}^{3})+\frac{1}{9}H^{3}-\frac{1}{2}H|\mathring{L}|_{h}^{2}.

(Throughout, LL is computed with respect to the unit normal vector σ\sigma – of which, for a three-manifold embedded in a four-manifold, there are only two choices.) Therefore,

Ψ(ζ|M3)\displaystyle\Psi(\zeta|M^{3}) =π2Γ(2)2(Φ3,0+12Φ3,1)dξ\displaystyle=\frac{\pi^{-2}\Gamma(2)}{2}(\Phi_{3,0}+\frac{1}{2}\Phi_{3,1})d\xi
=12π2[12L̊μν(ξ)Rμνh112H(ξ)Rh112H(ξ)|L̊(ξ)|h2\displaystyle=\frac{1}{2\pi^{2}}\left[\frac{1}{2}\mathring{L}^{\mu\nu}(\xi)R_{\mu\nu}^{h}-\frac{1}{12}H(\xi)R_{h}-\frac{1}{12}H(\xi)|\mathring{L}(\xi)|_{h}^{2}\right.
+154H(ξ)3+16trh(L̊(ξ)3)]dξ.\displaystyle\quad+\left.\frac{1}{54}H(\xi)^{3}+\frac{1}{6}\operatorname{tr}_{h}(\mathring{L}(\xi)^{3})\right]d\xi.

The measure dξd\xi here is actually a measure on the 0-sphere, but we include the ξ\xi’s to remind the reader of the dependence of this formula on a choice of normal vector. The only element ξΓ(ζ)\xi\in\Gamma(\zeta) is μM-\mu_{M}; consequently, we find

(2.4) MΓ(ζ)Ψ(ζ|M)𝑑vh(ζ)=14π2M(TM+M+13ΔhMHM+112μM(Rg))𝑑vhM.\int_{M}\int_{\Gamma(\zeta)}\Psi(\zeta|M)dv_{h}(\zeta)=\frac{1}{4\pi^{2}}\int_{M}(T_{M}+\mathcal{L}_{M}+\frac{1}{3}\Delta_{h_{M}}H_{M}+\frac{1}{12}\mu_{M}(R_{g}))dv_{h_{M}}.

Of course the integral along NN is analogous.

We now consider the integrand along Σ2\Sigma^{2}. Let k=g|TΣk=g|_{T\Sigma}. For a normal vector σNΣ\sigma\in N\Sigma, we let LΣ(σ)L_{\Sigma}(\sigma) be the corresponding second fundamental form. (When appropriate in index notation, we will write the Σ\Sigma up instead of down.) That is, L(σ)(X,Y)=XgY,σgL(\sigma)(X,Y)=\langle\nabla_{X}^{g}Y,\sigma\rangle_{g}. We write IIII for the vector second fundamental form, which is to say, II(X,Y)=(XgY)II(X,Y)=(\nabla_{X}^{g}Y)^{\perp}. We also write IIM(X,Y)=L(νM)(X,Y)II_{M}(X,Y)=L(\nu_{M})(X,Y) and IIN(X,Y)=L(νN)(X,Y)II_{N}(X,Y)=L(\nu_{N})(X,Y).

First, we find

Φ2,0(ζ,σ)\displaystyle\Phi_{2,0}(\zeta,\sigma) =12det(k)εα1α2εβ1β2Λα1β1(σ)Λα2β2(σ)\displaystyle=\frac{1}{2\det(k)}\varepsilon^{\alpha_{1}\alpha_{2}}\varepsilon^{\beta_{1}\beta_{2}}\Lambda_{\alpha_{1}\beta_{1}}(\sigma)\Lambda_{\alpha_{2}\beta_{2}}(\sigma)
=det(k1LΣ(σ))\displaystyle=\det(k^{-1}L_{\Sigma}(\sigma))
=12(HΣ(σ)2|LΣ(σ)|k2).\displaystyle=\frac{1}{2}(H_{\Sigma}(\sigma)^{2}-|L_{\Sigma}(\sigma)|_{k}^{2}).

Here HΣ(σ)=kαβLΣ(σ)αβH_{\Sigma}(\sigma)=k^{\alpha\beta}L_{\Sigma}(\sigma)_{\alpha\beta}.

Next, we find

Φ2,1(ζ,σ)\displaystyle\Phi_{2,1}(\zeta,\sigma) =14det(k)εα1α2εβ1β2Rα1α2β1β2g\displaystyle=-\frac{1}{4\det(k)}\varepsilon^{\alpha_{1}\alpha_{2}}\varepsilon^{\beta_{1}\beta_{2}}R_{\alpha_{1}\alpha_{2}\beta_{1}\beta_{2}}^{g}
=1det(k)R1212g\displaystyle=-\frac{1}{\det(k)}R_{1212}^{g}

By Gauss’s equation, we have

R1212g\displaystyle R_{1212}^{g} =R1212kII(1,2),II(1,2)+II(1,1),II(2,2)\displaystyle=R_{1212}^{k}-\langle II(\partial_{1},\partial_{2}),II(\partial_{1},\partial_{2})\rangle+\langle II(\partial_{1},\partial_{1}),II(\partial_{2},\partial_{2})\rangle
=R1212II12iII12i+II11iII22i.\displaystyle=R_{1212}-II_{12i}II_{12}^{i}+II_{11i}II_{22}^{i}.

Now let μM\mu_{M} be the unit normal to MM in XX, and let νM\nu_{M} be the unit normal to Σ\Sigma in MM. Together, these vectors form an orthonormal basis at each point for NΣN\Sigma. Thus, we find

Φ2,1(ζ,σ)\displaystyle\Phi_{2,1}(\zeta,\sigma) =Kdet(LΣ(νM))det(LΣ(μM))\displaystyle=K-\det(L_{\Sigma}(\nu_{M}))-\det(L_{\Sigma}(\mu_{M}))
=K12(HΣ(νM)2+HΣ(μM)2|LΣ(νM)|k2|LΣ(μM)|k2.\displaystyle=K-\frac{1}{2}(H_{\Sigma}(\nu_{M})^{2}+H_{\Sigma}(\mu_{M})^{2}-|L_{\Sigma}(\nu_{M})|_{k}^{2}-|L_{\Sigma}(\mu_{M})|_{k}^{2}.

(Recall that KK is the Gaussian curvature of Σ\Sigma.) Thus,

Ψ(ζ|Σ2)\displaystyle\Psi(\zeta|\Sigma^{2}) =12π2[Φ2,0+12Φ2,1]\displaystyle=\frac{1}{2\pi^{2}}\left[\Phi_{2,0}+\frac{1}{2}\Phi_{2,1}\right]
=12π2[12HΣ(ξ)212|LΣ(ξ)|k2+12K14HΣ(νM)214HΣ(μM)2\displaystyle=\frac{1}{2\pi^{2}}\left[\frac{1}{2}H_{\Sigma}(\xi)^{2}-\frac{1}{2}|L_{\Sigma}(\xi)|_{k}^{2}+\frac{1}{2}K-\frac{1}{4}H_{\Sigma}(\nu_{M})^{2}-\frac{1}{4}H_{\Sigma}(\mu_{M})^{2}\right.
+14|LΣ(νM)|k2+14|LΣ(μM)|k2]dξ.\displaystyle\quad\left.+\frac{1}{4}|L_{\Sigma}(\nu_{M})|_{k}^{2}+\frac{1}{4}|L_{\Sigma}(\mu_{M})|_{k}^{2}\right]d\xi.

Here, dξd\xi is the measure on the circle S1S^{1}, viewed as the fiber of the unit normal bundle S1NΣS^{1}N\Sigma to Σ\Sigma at ζ\zeta.

Now, any ξS1NΣ\xi\in S^{1}N\Sigma can be written ξ=cos(θ)νM+sin(θ)μM\xi=\cos(\theta)\nu_{M}+\sin(\theta)\mu_{M}. The outer angle Γ(ζ)\Gamma(\zeta) can then be written

Γ(ζ)={cos(θ)νM+sin(θ)μM:θ0(ζ)+π2<θ<3π2}.\Gamma(\zeta)=\left\{\cos(\theta)\nu_{M}+\sin(\theta)\mu_{M}:\theta_{0}(\zeta)+\frac{\pi}{2}<\theta<\frac{3\pi}{2}\right\}.

The contribution to the Gauss-Bonnet formula from Σ2\Sigma_{2} will therefore be

IΣ:=14π2Σθ0(ζ)+π23π2(K12HΣ(νM)2+12HΣ(μM)2+12|LΣ(νM)|k2+12|LΣ(μM)|k2+HΣ(cos(θ)νM+sin(θ)μM)2|LΣ(cos(θ)νM+sin(θ)μM)|k2)dθdvk(ζ)\begin{split}I_{\Sigma}:=&\frac{1}{4\pi^{2}}\oint_{\Sigma}\int_{\theta_{0}(\zeta)+\frac{\pi}{2}}^{\frac{3\pi}{2}}\left(K-\frac{1}{2}H_{\Sigma}(\nu_{M})^{2}+\frac{1}{2}H_{\Sigma}(\mu_{M})^{2}+\frac{1}{2}|L_{\Sigma}(\nu_{M})|_{k}^{2}\right.\\ &+\frac{1}{2}|L_{\Sigma}(\mu_{M})|_{k}^{2}+H_{\Sigma}(\cos(\theta)\nu_{M}+\sin(\theta)\mu_{M})^{2}\\ &\left.-|L_{\Sigma}(\cos(\theta)\nu_{M}+\sin(\theta)\mu_{M})|_{k}^{2}\right)d\theta dv_{k}(\zeta)\end{split}

Observe that, by bilinearity of gg

HΣ(cos(θ)νM+sin(θ)μM)2\displaystyle H_{\Sigma}(\cos(\theta)\nu_{M}+\sin(\theta)\mu_{M})^{2} =cos2(θ)HΣ(νM)+2sin(θ)cos(θ)HΣ(νM)HΣ(μM)\displaystyle=\cos^{2}(\theta)H_{\Sigma}(\nu_{M})+2\sin(\theta)\cos(\theta)H_{\Sigma}(\nu_{M})H_{\Sigma}(\mu_{M})
+sin2(θ)HΣ(μM),\displaystyle\quad+\sin^{2}(\theta)H_{\Sigma}(\mu_{M}),
and
|LΣ(cos(θ)νM+sin(θ)μM)|k2\displaystyle|L_{\Sigma}(\cos(\theta)\nu_{M}+\sin(\theta)\mu_{M})|_{k}^{2} =cos2(θ)|LΣ(νM)|k2\displaystyle=\cos^{2}(\theta)|L_{\Sigma}(\nu_{M})|_{k}^{2}
+2sin(θ)cos(θ)LΣ(νM)αβLΣ(μM)αβ\displaystyle\quad+2\sin(\theta)\cos(\theta)L_{\Sigma}(\nu_{M})_{\alpha\beta}L_{\Sigma}(\mu_{M})^{\alpha\beta}
+sin2(θ)|LΣ(μM)|k2.\displaystyle\quad+\sin^{2}(\theta)|L_{\Sigma}(\mu_{M})|_{k}^{2}.

Now, θ0+π23π2cos2(θ)𝑑θ=12(πθ0)+sin(2θ0)4\int_{\theta_{0}+\frac{\pi}{2}}^{\frac{3\pi}{2}}\cos^{2}(\theta)d\theta=\frac{1}{2}(\pi-\theta_{0})+\frac{\sin(2\theta_{0})}{4}, θ0+π2sin2(θ)𝑑θ=12(πθ0)sin(2θ0)4, and\int_{\theta_{0}+\frac{\pi}{2}}\sin^{2}(\theta)d\theta=\frac{1}{2}(\pi-\theta_{0})-\frac{\sin(2\theta_{0})}{4},\text{ and}, and θ0+π23π2sin(θ)cos(θ)𝑑θ=1414cos(2θ0)\int_{\theta_{0}+\frac{\pi}{2}}^{\frac{3\pi}{2}}\sin(\theta)\cos(\theta)d\theta=\frac{1}{4}-\frac{1}{4}\cos(2\theta_{0}). Also,

LΣ(ξ)αβLΣ(τ)αβ=L̊Σ(ξ)αβL̊Σ(τ)αβ+12HΣ(ξ)HΣ(τ).L_{\Sigma}(\xi)_{\alpha\beta}L_{\Sigma}(\tau)^{\alpha\beta}=\mathring{L}_{\Sigma}(\xi)_{\alpha\beta}\mathring{L}_{\Sigma}(\tau)^{\alpha\beta}+\frac{1}{2}H_{\Sigma}(\xi)H_{\Sigma}(\tau).

Next, we can write νN=cos(θ0)νM+sin(θ0)μM\nu_{N}=\cos(\theta_{0})\nu_{M}+\sin(\theta_{0})\mu_{M}, so

(2.5) μM=csc(θ0)νNcot(θ0)νM.\mu_{M}=\csc(\theta_{0})\nu_{N}-\cot(\theta_{0})\nu_{M}.

Finally, we have L̊Σ(νM)=II̊M\mathring{L}_{\Sigma}(\nu_{M})=\mathring{II}_{M}, and L̊Σ(νN)=II̊N\mathring{L}_{\Sigma}(\nu_{N})=\mathring{II}_{N}, while HΣ(νM)=ηMH_{\Sigma}(\nu_{M})=\eta_{M} and HΣ(νN)=ηNH_{\Sigma}(\nu_{N})=\eta_{N}. Applying all these identities, we get

(2.6) IΣ\displaystyle I_{\Sigma} =14π2Σ((πθ0)K14cot(θ0)(ηM2+ηN2)\displaystyle=\frac{1}{4\pi^{2}}\oint_{\Sigma}\left((\pi-\theta_{0})K-\frac{1}{4}\cot(\theta_{0})(\eta_{M}^{2}+\eta_{N}^{2})\right.
(2.7) +12csc(θ0)ηMηN+12cot(θ0)(|II̊M|k2+|II̊N|k2)csc(θ0)II̊αβMII̊Nαβ)dvk\displaystyle\quad+\left.\frac{1}{2}\csc(\theta_{0})\eta_{M}\eta_{N}+\frac{1}{2}\cot(\theta_{0})(|\mathring{II}_{M}|_{k}^{2}+|\mathring{II}_{N}|_{k}^{2})-\csc(\theta_{0})\mathring{II}_{\alpha\beta}^{M}\mathring{II}^{\alpha\beta}_{N}\right)dv_{k}
(2.8) =14π2Σ(U+G+13(νMHM+νNHN))𝑑vk,\displaystyle=\frac{1}{4\pi^{2}}\oint_{\Sigma}\left(U+G+\frac{1}{3}\left(\nu_{M}H_{M}+\nu_{N}H_{N}\right)\right)dv_{k},

where U,GU,G are as in (1.4),(1.5).

Now applying (2.3), (2.4), (2.6), and Green’s theorem yields (1.10), and thus the theorem. ∎

One observes that, in order to write the Chern-Gauss-Bonnet formula in terms of quantities QQ and TT that would satisfy (1.1), (1.3), it was necessary for Chang and Qing to add divergence terms both to the interior and boundary integrals (as well as a correction term at the boundary, via Green’s theorem). On the other hand, as Theorem 1.1 and the above calculation shows, there is no need to add any divergence to the corner integral (though we still need the Green term); this is fortuitous, as there is no obvious second-order divergence to add!

Proof of Theorem 1.1.

Suppose that g~=e2ωg\tilde{g}=e^{2\omega}g, where ωC(X)\omega\in C^{\infty}(X). Recall the following standard formulas along Σ\Sigma:

K~\displaystyle\widetilde{K} =e2ω(KΔkω)\displaystyle=e^{-2\omega}(K-\Delta_{k}\omega)
η~M\displaystyle\widetilde{\eta}_{M} =eω(ηM2νMω)\displaystyle=e^{-\omega}(\eta_{M}-2\nu_{M}\omega)
η~N\displaystyle\widetilde{\eta}_{N} =eω(ηN2νNω)\displaystyle=e^{-\omega}(\eta_{N}-2\nu_{N}\omega)
II~̊M\displaystyle\mathring{\widetilde{II}}_{M} =eωII̊M\displaystyle=e^{\omega}\mathring{II}_{M}
II~̊N\displaystyle\mathring{\widetilde{II}}_{N} =eωII̊N\displaystyle=e^{\omega}\mathring{II}_{N}
H~M\displaystyle\widetilde{H}_{M} =eω(H3μMω)\displaystyle=e^{-\omega}(H-3\mu_{M}\omega)
H~N\displaystyle\widetilde{H}_{N} =eω(H3μNω)\displaystyle=e^{-\omega}(H-3\mu_{N}\omega)
θ~0\displaystyle\tilde{\theta}_{0} =θ0.\displaystyle=\theta_{0}.

Using these and equation (1.5), (1.7) follows easily.

We show (1.9) explicitly. Using (1.4), we find

U~\displaystyle\widetilde{U} =(πθ0)e2ω(KΣΔkω)14cot(θ0)e2ω(ηM24ηMνMω+4ηM(ω)2\displaystyle=(\pi-\theta_{0})e^{-2\omega}(K_{\Sigma}-\Delta_{k}\omega)-\frac{1}{4}\cot(\theta_{0})e^{-2\omega}\left(\eta_{M}^{2}-4\eta_{M}\nu_{M}\omega+4\eta_{M}(\omega)^{2}\right.
+ηN24ηNνNω+4ηN(ω)2)\displaystyle\quad+\left.\eta_{N}^{2}-4\eta_{N}\nu_{N}\omega+4\eta_{N}(\omega)^{2}\right)
+12csc(θ0)e2ω(ηMηN2ηMνNω2ηNνMω+4ηM(ω)ηN(ω))\displaystyle\quad+\frac{1}{2}\csc(\theta_{0})e^{-2\omega}\left(\eta_{M}\eta_{N}-2\eta_{M}\nu_{N}\omega-2\eta_{N}\nu_{M}\omega+4\eta_{M}(\omega)\eta_{N}(\omega)\right)
13eωνM(eω(HM3μMω))13eωνN(eω(HN3μNω));\displaystyle\quad-\frac{1}{3}e^{-\omega}\nu_{M}(e^{-\omega}(H_{M}-3\mu_{M}\omega))-\frac{1}{3}e^{-\omega}\nu_{N}(e^{-\omega}(H_{N}-3\mu_{N}\omega));
Now applying (2.5) to the last two terms, we find
U~\displaystyle\widetilde{U} =e2ω[(πθ0)KΣcot(θ0)(ηM2+ηN2)+2csc(θ0)ηMηN13(νMHM+νNHN)\displaystyle=e^{-2\omega}\left[(\pi-\theta_{0})K_{\Sigma}-\cot(\theta_{0})(\eta_{M}^{2}+\eta_{N}^{2})+2\csc(\theta_{0})\eta_{M}\eta_{N}-\frac{1}{3}(\nu_{M}H_{M}+\nu_{N}H_{N})\right.
+cot(θ0)ηMνMωcot(θ0)νM(ω)2+cot(θ0)ηNνNωcot(θ0)νN(ω)2\displaystyle\quad+\cot(\theta_{0})\eta_{M}\nu_{M}\omega-\cot(\theta_{0})\nu_{M}(\omega)^{2}+\cot(\theta_{0})\eta_{N}\nu_{N}\omega-\cot(\theta_{0})\nu_{N}(\omega)^{2}
csc(θ0)ηMνNωcsc(θ0)ηNνMω+2csc(θ0)νM(ω)νN(ω)+(θ0π)Δkω\displaystyle\quad-\csc(\theta_{0})\eta_{M}\nu_{N}\omega-\csc(\theta_{0})\eta_{N}\nu_{M}\omega+2\csc(\theta_{0})\nu_{M}(\omega)\nu_{N}(\omega)+(\theta_{0}-\pi)\Delta_{k}\omega
+13HMνMω+νMμMωcsc(θ0)νM(ω)νN(ω)+cot(θ0)νM(ω)2\displaystyle\quad+\frac{1}{3}H_{M}\nu_{M}\omega+\nu_{M}\mu_{M}\omega-\csc(\theta_{0})\nu_{M}(\omega)\nu_{N}(\omega)+\cot(\theta_{0})\nu_{M}(\omega)^{2}
+13HNνNω+νNμNωcsc(θ0)νN(ω)νM(ω)+cot(θ0)νN(ω)2\displaystyle\quad+\frac{1}{3}H_{N}\nu_{N}\omega+\nu_{N}\mu_{N}\omega-\csc(\theta_{0})\nu_{N}(\omega)\nu_{M}(\omega)+\cot(\theta_{0})\nu_{N}(\omega)^{2}
=e2ω(U+P2bω).\displaystyle=e^{-2\omega}\left(U+P_{2}^{b}\omega\right).

It is interesting in the above to note that nonlinear terms coming from the “Green terms” νMHM+νNHN\nu_{M}H_{M}+\nu_{N}H_{N} in UU precisely cancel those arising from the Allendoerfer-Weil formula.

We now need show only (1.8). This will follow from (1.9) and the linearity of P2bP_{2}^{b}, as in proposition 3.1 of [CQ97]. Let ϕC(X)\phi\in C^{\infty}(X). Suppose g^=e2(ω+ϕ)g=e2ϕg~\hat{g}=e^{2(\omega+\phi)}g=e^{2\phi}\tilde{g}. Then

P2b(ω+ϕ)+U\displaystyle P_{2}^{b}(\omega+\phi)+U =e2(ω+ϕ)U^\displaystyle=e^{2(\omega+\phi)}\widehat{U}
P~2bϕ+U~\displaystyle\widetilde{P}_{2}^{b}\phi+\widetilde{U} =U^e2ϕ\displaystyle=\widehat{U}e^{2\phi}

Multiply the second equation by e2ωe^{2\omega} and subtract to get

P2b(ω+ϕ)+Ue2ωP~2bϕe2ωU~=0.P_{2}^{b}(\omega+\phi)+U-e^{2\omega}\widetilde{P}_{2}^{b}\phi-e^{2\omega}\widetilde{U}=0.

The result now follows by (1.9) and linearity. ∎

Acknowledgments I am indebted to Alice Chang, Paul Yang, Matthew Gursky, and Baris Coskunuzer for helpful conversations. This work was carried out at Princeton University and the University of Texas at Dallas, and I thank both institutions for their support and for the excellent environments for doing math that they have provided. The work at Princeton was supported also by NSF RTG DMS-1502525.

References

  • [AC17] A. G. Ache and S.-Y. A. Chang. Sobolev trace inequalities of order four. Duke Math. J., 166(14):2719–2748, 2017.
  • [Ale12] S. Alexakis. The decomposition of global conformal invariants. Number 182 in Annals of Mathematics Studies. Princeton University Press, Princeton, New Jersey, 2012.
  • [All42] C. B. Allendoerfer. The Euler number of a Riemannian manifold. Amer. J. Math, 62:243, 1942.
  • [And03] M. T. Anderson. Boundary regularity, uniqueness and non-uniqueness for AH Einstein metrics on 4-manifolds. Adv. Math, 179:205–249, 2003.
  • [AW43] C. B. Allendoerfer and A. Weil. The Gauss-Bonnet theorem for Riemannian polyhedra. Transactions of the AMS, 53(1):101–129, 1943.
  • [BG94] T. Branson and P. Gilkey. The functional determinant of a 4-dimensional boundary value problem. Transactions of the AMS, 344:479–531, 1994.
  • [BØ91] T.P. Branson and B. Ørsted. Explicit functional determinants in four dimensions. Proceedings of the AMS, 113:669–682, 1991.
  • [CGQ18] S.-Y. A. Chang, Y. Ge, and J. Qing. Compactness of conformally compact Einstein 4-manifolds II. preprint:arXiv 1811.02112, pages 1–28, 2018.
  • [CGY03] S.-Y. A. Chang, M. J. Gursky, and P. Yang. A conformally invariant sphere theorem in four dimensions. Publ. Math. de l’IHES, 98(1):105–143, 2003.
  • [Che44] S.-S. Chern. A simple intrinsic proof of the Gauss-Bonnet formula for closed Riemannian manifolds. Annals of Mathematics, pages 747–752, 1944.
  • [Che45] S.-S. Chern. On the curvatura integra in a Riemannian manifold. Annals of Mathematics, 46(4):674–684, 1945.
  • [CL19] J. S. Case and W. Luo. Boundary operators associated with the sixth-order GJMS operator. International Math. Res. Notices, doi:doi.org/10.10983/imrn/rnz121 ; arXiv:1810.07027:1–54, 2019.
  • [CQ97] S.-Y. A. Chang and J. Qing. The zeta functional determinants on manifolds with boundary I–the formula. Journal of Functional Analysis, 147:327–362, 1997.
  • [CQY00] S.-Y. A. Chang, J. Qing, and P. C. Yang. On the Chern-Gauss-Bonnet integral for conformal metrics on 4\mathbb{R}^{4}. Duke Math. J., 103(3):523–544, 2000.
  • [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.
  • [Fen40] W. Fenchel. On total curvatures of Riemannian manifolds. J. London Math. Soc., 15:15, 1940.
  • [GP18] A. R. Gover and L. J. Peterson. Conformal boundary operators, T-curvatures, and conformal fractional Laplacians of odd order. arXiv: 1802.08366, pages 1–58, 2018.
  • [Gur99] M. J. Gursky. The principal eigenvalue of a conformally invariant differential operator, with an application to semilinear elliptic PDE. Comm. Math. Phys., 207:13–143, 1999.
  • [GW99] C. R. Graham and E. Witten. Conformal anomaly of submanifold observables in AdS/CFT correspondence. Nuclear Phys. B, 546(1-2):52–64, 1999.
  • [Juh09] A. Juhl. Families of Conformally Covariant Differential Operators, Q-Curvature and Holography, volume 275 of Progress in Mathematics. Birkhäuser, Basel, 2009.
  • [Lef56] S. Lefschetz. Topology. Chelsea, New York, 2nd edition, 1956.
  • [Ndi08] C. B. Ndiaye. Conformal metrics with constant Q-curvature for manifolds with boundary. Communications in Analysis and Geometry, 16(5):1049 – 1124, 2008.
  • [Ndi09] C. B. Ndiaye. Constant T-curvature conformal metrics on 4-manifolds with boundary. Pacific Journal of Mathematics, 240(1):151 – 184, 2009.
  • [Spi99] M. Spivak. A Comprehensive Introduction to Differential Geometry. Publish or Perish, Houston, 3rd edition, 1999.