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

Decompositions of the space of Riemannian metrics
on a compact manifold with boundary

Shota Hamanaka supported in doctoral program in Chuo University, 2020.
Abstract

In this paper, for a compact manifold MM with non-empty boundary M\partial M, we give a Koiso-type decomposition theorem, as well as an Ebin-type slice theorem, for the space of all Riemannian metrics on MM endowed with a fixed conformal class on M\partial M. As a corollary, we give a characterization of relative Einstein metrics.

1 Introduction

The study of the differential structure of the space \mathscr{M} of all Riemannian metrics on a closed manifold is one of important studies in geometry. In [10], Ebin particularly has proved a slice theorem for the pullback action of the diffeomorphism group on \mathscr{M}. In [17], Koiso has also extended it to an Inverse Limit Hilbert (ILH for brevity)-version. Moreover, he has also studied the conformal action on \mathscr{M}, and consequently has proved the following decomposition theorem:

Theorem 1.1 (Koiso’s decomposition theorem [18, Corollary 2.9] ).

Let MnM^{n} be a closed nn-manifold (n3)(n\geq 3), \mathscr{M} the space of all Riemannian metrics on MM and Diff(M)\mathrm{Diff}(M) the diffeomorphism group of MM. Set also

C+(M):={fC(M)|f>0onM},C^{\infty}_{+}(M):=\bigl{\{}f\in C^{\infty}(M)\bigm{|}f>0~{}\mathrm{on}~{}M\bigr{\}},
𝔖ˇ:={g|Vol(M,g)=1,Rg=const,Rgn1Spec(Δg)},\check{\mathfrak{S}}:=\Biggl{\{}g\in\mathscr{M}\Biggm{|}\mathrm{Vol}(M,g)=1,~{}R_{g}=\mathrm{const},~{}\frac{R_{g}}{n-1}\notin\mathrm{Spec}(-\Delta_{g})\Biggr{\}},

where Vol(M,g)\mathrm{Vol}(M,g), RgR_{g} and Spec(Δg)\mathrm{Spec}(-\Delta_{g}) denote respectively the volume of (M,g)(M,g), the scalar curvature of gg and the set of all non-zero eigenvalues of the (non-negative) Laplacian Δg-\Delta_{g} of gg. Note that these four spaces become naturally ILH-manifolds. For any g=fg¯(fC+,g¯𝔖ˇ)g=f\bar{g}~{}(f\in C^{\infty}_{+},~{}\bar{g}\in\check{\mathfrak{S}}) and any smooth deformation {g(t)}t(ϵ,ϵ)\{g(t)\}_{t\in(-\epsilon,\epsilon)} of gg for sufficiently small ϵ>0\epsilon>0, then there exist uniquely smooth deformations {f(t)}t(ϵ,ϵ)(C+(M))\{f(t)\}_{t\in(-\epsilon,\epsilon)}(\subset C^{\infty}_{+}(M)) of ff, {ϕ(t)}t(ϵ,ϵ)(Diff(M)\{\phi(t)\}_{t\in(-\epsilon,\epsilon)}(\subset\mathrm{Diff}(M) ) of the identity idMid_{M} and {g(t)}t(ϵ,ϵ)(𝔖ˇ\{g(t)\}_{t\in(-\epsilon,\epsilon)}(\subset\check{\mathfrak{S}} ) of g¯\bar{g} with δg(g¯(0))=0\delta_{g}(\bar{g}^{{}^{\prime}}(0))=0 such that

g(t)=f(t)ϕ(t)g¯(t).g(t)=f(t)\phi(t)^{*}\bar{g}(t).

Here, δg(g¯(0))\delta_{g}(\bar{g}^{{}^{\prime}}(0)) denotes the divergence i(g¯(0))i-~{}\nabla^{i}(\bar{g}^{{}^{\prime}}(0))_{i} with respect to gg.

Note that the above decomposition can be replaced by

g(t)=(f(t)ϕ(t))ϕ(t)g¯(t)withδg(f(0)g¯(0)+f(0)g¯(0))=0.g(t)=(f(t)\circ\phi(t))\phi(t)^{*}\bar{g}(t)~{}\mathrm{with}~{}\delta_{g}\big{(}f^{{}^{\prime}}(0)\bar{g}(0)+f(0)\bar{g}^{{}^{\prime}}(0)\big{)}=0.

This theorem has often played an important role in studying gemoetric structures related to several variational problems on a closed manifold. Hence, extending this to on a manifold with boundary seems to be also important.

From now on, we throughout assume that MM is a compact connected oriented smooth nn-manifold (n3)(n\geq 3) with smooth boundary M\partial M. Let \mathscr{M} be the space of all Riemannian metrics on M.M. In order to obtain a corresponding Koiso-type decomposition theorem on MM with M\partial M to Theorem 1.1 on a closed manifold, we need to fix a suitable boundary condition for each metric gg on MM. From the variational view point of the Einstein-Hilbert functional, a candidate of such boundary conditions may be the following: For a fixed metric g0g_{0} on MM with zero mean curvature Hg0=0H_{g_{0}}=0 along M\partial M, we fix the boundary condition for each gg on MM as [g|M]=[g0|M][g|_{\partial M}]=[g_{0}|_{\partial M}] and Hg=0H_{g}=0 along M\partial M (see Fact 2.1 and [2],[3]). Here, [g|M][g|_{\partial M}] denotes the conformal class of g|Mg|_{\partial M}. However, one can notice that this boundary condition is not enough to get a Koiso-type decomposition theorem, even more an Ebin-type slice theorem. Here, we will fix a slightly stronger boundary condition below, which still has a naturality(see Fact 2.1(1)).

Fix a Riemannian metric g0g_{0} on MM with Hg0=0H_{g_{0}}=0 along M\partial M and set its conformal class C:=[g0]C:=[g_{0}] on MM. νg0\nu_{g_{0}} denotes the outer unit normal vector field along M\partial M with respect to g0g_{0}. When two metrics gg and g~\tilde{g} on MM have the same 1-jets jx1g=jx1g~j^{1}_{x}g=j^{1}_{x}\tilde{g} for all xM,x\in\partial M, we write it as jM1g=jM1g~.j^{1}_{\partial M}g=j^{1}_{\partial M}\tilde{g}. Set also

C+(M)N:={fC+(M)|νg0(f)|M=0},C^{\infty}_{+}(M)_{N}:=\bigl{\{}f\in C^{\infty}_{+}(M)\bigm{|}\nu_{g_{0}}(f)|_{\partial M}=0\bigr{\}},
C0:={g|g=fg0onMforsomefC+(M),Hg=0onM},\mathscr{M}_{C_{0}}:=\bigl{\{}g\in\mathscr{M}\bigm{|}g=fg_{0}~{}\mathrm{on}~{}\partial M~{}\mathrm{for~{}some}~{}f\in C^{\infty}_{+}(M),~{}H_{g}=0~{}\mathrm{on}~{}\partial M\bigr{\}},
C01:={g|jM1g=jM1(fg0)forsomefC+(M)N},\mathscr{M}_{C^{1}_{0}}:=\bigl{\{}g\in\mathscr{M}\bigm{|}j^{1}_{\partial M}g=j^{1}_{\partial M}(fg_{0})~{}\mathrm{for~{}some}~{}f\in C^{\infty}_{+}(M)_{N}\bigr{\}},
𝔖C0(1):={gC0(1)|Vol(M,g)=1,Rg=const},\mathfrak{S}_{C^{(1)}_{0}}:=\bigl{\{}g\in\mathscr{M}_{C^{(1)}_{0}}\bigm{|}\mathrm{Vol}(M,g)=1,~{}R_{g}=\mathrm{const}\bigr{\}},
𝔖ˇC0(1):={g𝔖C0(1)|Rgn1Spec(Δg;Neumann)},\check{\mathfrak{S}}_{C^{(1)}_{0}}:=\Biggl{\{}g\in\mathfrak{S}_{C^{(1)}_{0}}\Biggm{|}\frac{R_{g}}{n-1}\notin\mathrm{Spec}(-\Delta_{g};\mathrm{Neumann})\Biggr{\}},
DiffC0:={ϕDiff(M)|ϕg0=fg0onMforsomefC+(M)},\mathrm{Diff}_{C_{0}}:=\bigl{\{}\phi\in\mathrm{Diff}(M)\bigm{|}\phi^{*}g_{0}=fg_{0}~{}\mathrm{on}~{}\partial M~{}\mathrm{for~{}some}~{}f\in C^{\infty}_{+}(M)\bigr{\}},

where Spec(Δg;Neumann)\mathrm{Spec}(-\Delta_{g};\mathrm{Neumann}) denote the set of all non-zero eigenvalues of Δg-\Delta_{g} with the Neumann boundary condition respectively. Note that Hg=0H_{g}=0 along M\partial M for all gC01g\in\mathscr{M}_{C^{1}_{0}}. Our main result is the following theorem:

Main Theorem.

For any g=fg¯(fC+(M)N,g¯𝔖ˇC01g=f\bar{g}~{}(f\in C^{\infty}_{+}(M)_{N},~{}\bar{g}\in\check{\mathfrak{S}}_{C^{1}_{0}} ) and any smooth deformation {g(t)}t(ϵ,ϵ)(C01)\{g(t)\}_{t\in(-\epsilon,\epsilon)}(\subset\mathscr{M}_{C^{1}_{0}}) of gg for sufficiently small ϵ>0\epsilon>0 , there exist smooth deformations {f(t)}t(ϵ,ϵ)(C+(M)N)\{f(t)\}_{t\in(-\epsilon,\epsilon)}(\subset C^{\infty}_{+}(M)_{N}) of f,f, {ϕ(t)}t(ϵ,ϵ)(DiffC0)\{\phi(t)\}_{t\in(-\epsilon,\epsilon)}(\subset\mathrm{Diff}_{C_{0}}) of idMid_{M} and {g¯(t)}t(ϵ,ϵ)(𝔖ˇC01)\{\bar{g}(t)\}_{t\in(-\epsilon,\epsilon)}(\subset\check{\mathfrak{S}}_{C^{1}_{0}}) of g¯\bar{g} with δg(g¯(0))=0\delta_{g}(\bar{g}^{{}^{\prime}}(0))=0 such that

g(t)=f(t)ϕ(t)g¯(t).g(t)=f(t)\phi(t)^{*}\bar{g}(t).

The rest of this paper is organized as follows. In Section 2, we state a Slice theorem for a manifold with boundary with a fixed conformal class on the boundary and prove it. In Section 3, we prepare some necessary lemmas for the proof of Main Theorem. Finally, combining them with Slice theorem, we prove Main Theorem. In Section 4, we give two applications of Main Theorem.

Acknowledgement

I would like to thank my supervisor Kazuo Akutagawa for suggesting the initial direction for my study, his good advice and support.

2 Preliminaries and a slice theorem

Let MM be a compact connected oriented smooth nn-dimensional manifold with non-empty smooth boundary M\partial M. Fix a Riemannian metric g0g_{0} with Hg0=0H_{g_{0}}=0. Here, Hg0H_{g_{0}} denotes the mean curvature of M\partial M with respect to g0g_{0}. And set C:=[g0]C:=[g_{0}] its conformal class on MM. For a given positive definite symmetric (0,2)-type tensor field TT on MM, we will write T||MC||MT||_{\partial M}\in C||_{\partial M} when T=fg0forsomefC+(M)onMT=f\cdot g_{0}~{}\mathrm{for~{}some}~{}f\in C^{\infty}_{+}(M)~{}\mathrm{on}~{}\partial M. Note that this condition equivalent to ιT=ι(fg0)\iota^{*}T=\iota^{*}(fg_{0}) for some fC+(M)f\in C^{\infty}_{+}(M), where ι:MM\iota~{}:\partial M\rightarrow M is the natural inclusion. Moreover, we denote T||M1C0||M1T||^{1}_{\partial M}\in C_{0}||^{1}_{\partial M} when jM1T=jM1(fg0)forsomefC+(M)Nj^{1}_{\partial M}T=j^{1}_{\partial M}(f\cdot g_{0})~{}\mathrm{for~{}some}~{}f\in C^{\infty}_{+}(M)_{N}. With this understood, we set C:={g|g||MC||M}\mathscr{M}_{C}:=\bigl{\{}g\in\mathscr{M}\bigm{|}g||_{\partial M}\in C||_{\partial M}\bigr{\}}. Note also that DiffC0:={ϕDiff(M)|(ϕg0)||MC||M}.\mathrm{Diff}_{C_{0}}:=\bigl{\{}\phi\in\mathrm{Diff}(M)\bigm{|}(\phi^{*}g_{0})||_{\partial M}\in C||_{\partial M}\bigr{\}}. and C01={g|g||M1C||M1}C0={g|g||MC||M,Hg=0onM}.\mathscr{M}_{C^{1}_{0}}=\bigl{\{}g\in\mathscr{M}\bigm{|}g||^{1}_{\partial M}\in C||^{1}_{\partial M}\bigr{\}}\subsetneq\mathscr{M}_{C_{0}}=\bigl{\{}g\in\mathscr{M}\bigm{|}g||_{\partial M}\in C||_{\partial M},~{}H_{g}=0~{}\mathrm{on}~{}\partial M\bigr{\}}.

Remark 2.1.

In the case that M=\partial M=\emptyset (that is, MM is a closed manifold), it is well known that a Riemannian metric on MM is Einstein if and only if it is a critical point of the normalized Einstein-Hilbert functional \mathcal{E} on the space :\mathscr{M}:

:,g(g):=MRgdvgVolg(M)n2n,\mathcal{E}~{}:~{}\mathscr{M}\rightarrow\mathbb{R},~{}~{}g\mapsto\mathcal{E}(g):=\frac{\mathop{\text{\LARGE$\int_{\text{\normalsize$\scriptstyle M$}}$}}\nolimits R_{g}dv_{g}}{\mathrm{Vol}_{g}(M)^{\frac{n-2}{n}}},

where Rg,dvg,Volg(M)R_{g},~{}dv_{g},~{}\mathrm{Vol}_{g}(M) denote respectively the scalar curvature, the volume measure of gg and the volume of (M,g)(M,g). However, if we consider the analogue of the case of \mathcal{E} on compact nn-manifold MM with non-empty boundary, then the set of critical points of \mathcal{E} on the space \mathscr{M} is empty (see Fact 2.1 below). Hence, in this case, we need to fix a suitable boundary condition for all metrics, and then \mathcal{E} must be restricted to a subspace of \mathscr{M}.

When M,\partial M\neq\emptyset, set the several subspaces of \mathscr{M} below:

C|:={g|[g|M]=C|M},\mathscr{M}_{C|_{\partial}}:=\bigl{\{}g\in\mathscr{M}\bigm{|}[g|_{\partial M}]=C|_{\partial M}\bigr{\}},
Cconst|:={gC||cs.t.Hg=conM},\mathscr{M}_{C_{\mathrm{const}}|_{\partial}}:=\bigl{\{}g\in\mathscr{M}_{C|_{\partial}}\bigm{|}\exists c\in\mathbb{R}~{}\mathrm{s.t.}~{}H_{g}=c~{}\mathrm{on}~{}\partial M\bigr{\}},
0:={g|Hg=0onM},\mathscr{M}_{0}:=\bigl{\{}g\in\mathscr{M}\bigm{|}H_{g}=0~{}\mathrm{on}~{}\partial M\bigr{\}},
C0|:=C|0={gC||Hg=0onM}.\mathscr{M}_{C_{0}|_{\partial}}:=\mathscr{M}_{C|_{\partial}}\cap\mathscr{M}_{0}=\bigl{\{}g\in\mathscr{M}_{C|_{\partial}}\bigm{|}H_{g}=0~{}\mathrm{on}~{}\partial M\bigr{\}}.

A metric gg\in\mathscr{M} is called a relative metric if g0.g\in\mathscr{M}_{0}. By Fact 2.1 below, it is reasonable to restrict the functional \mathcal{E} to the subspace C0\mathscr{M}_{C_{0}} as well as C0|\mathscr{M}_{C_{0}|_{\partial}} and 0.\mathscr{M}_{0}.

Fact 2.1 ([2, Proposition 2.1][3, Remark 1, Theorem 1.1]).

Let M,,M,~{}\mathcal{E}, and \mathscr{M} be the same as the above. Then the following holds:

(1) gCrit(|C0(1))g\in\mathrm{Crit}(\mathcal{E}|_{\mathscr{M}_{C^{(1)}_{0}}}) if and only if gg is an Einstein metric with Hg=0H_{g}=0 (namely, a relative Einstein metric) and g||M(1)C||M(1).g||^{(1)}_{\partial M}\in C||^{(1)}_{\partial M}.

(2) gCrit(|C0|)g\in\mathrm{Crit}(\mathcal{E}|_{\mathscr{M}_{C_{0}|_{\partial}}}) if and only if gg is a relative Einstein metric and [g|M]=C|M[g|_{\partial M}]=C|_{\partial M}.

(3) gCrit(|0)g\in\mathrm{Crit}(\mathcal{E}|_{\mathscr{M}_{0}}) if and only if gg is an Einstein metric with totally geodesic boundary.

(4) Crit()=,Crit(|C)=,Crit(C|)=,Crit(Cconst|)=.\mathrm{Crit}(\mathcal{E})=\emptyset,~{}~{}\mathrm{Crit}(\mathcal{E}|_{\mathscr{M}_{C}})=\emptyset,~{}~{}\mathrm{Crit}(\mathcal{E}_{\mathscr{M}_{C|_{\partial}}})=\emptyset,~{}~{}\mathrm{Crit}(\mathcal{E}_{\mathscr{M}_{C_{\mathrm{const}}|_{\partial}}})=\emptyset. Here, for instance, Crit()\mathrm{Crit}(\mathcal{E}) and Crit(|C)\mathrm{Crit}(\mathcal{E}|_{\mathscr{M}_{C}}) denote respectively the set of all critical metrics of \mathcal{E} and the set of those of its restriction to C\mathscr{M}_{C}.

From now on, we will consentrate on the spaces C\mathscr{M}_{C} and C0.\mathscr{M}_{C_{0}}. For a smooth fibre bundle FF , we denote by Hs(F)H^{s}(F) the space of all Ws,2W^{s,2} -sections. (Note that L2L^{2} -norm does not depend on the choice of Riemannian metric, hence, we fix a reference metric to define these function spaces.) The Sobolev embedding theorem states that Hs(F)Ck(F)H^{s}(F)\hookrightarrow C^{k}(F) is continuous if s>n/2+ks>n/2+k, see for instance [4]. By the Sobolev embedding, if s>n/2s>n/2, Hs(M×M)H^{s}(M\times M) (the set of all HsH^{s}-maps from MM to itself) is a Hilbert manifold. Pick s>4+n2s>4+\frac{n}{2} and let CsDiff:={ηCs(M×M)|η1Cs(M×M)}C^{s}\mathrm{Diff}:=\bigl{\{}\eta\in C^{s}(M\times M)\bigm{|}\eta^{-1}\in C^{s}(M\times M)\bigr{\}} and let Diffs:=Hs(M×M)C1Diff\mathrm{Diff}^{s}:=H^{s}(M\times M)\cap C^{1}\mathrm{Diff}. From the Sobolev embedding, Diffs\mathrm{Diff}^{s} is open in Hs(M×M)H^{s}(M\times M) and hence it is also a Hilbert manifold. Let DiffC0s:={ηDiffs|(ηg0)||MC||M}={ηDiffs|ηg0=fg0onMforsomefHs3/2(C+(M))}\mathrm{Diff}^{s}_{C_{0}}:=\bigl{\{}\eta\in\mathrm{Diff}^{s}\bigm{|}(\eta^{*}g_{0})||_{\partial M}\in C||_{\partial M}\bigr{\}}=\bigl{\{}\eta\in\mathrm{Diff}^{s}\bigm{|}\eta^{*}g_{0}=fg_{0}~{}\mathrm{on}~{}\partial M~{}\mathrm{for~{}some}~{}f\in H^{s-3/2}(C^{\infty}_{+}(M))\bigr{\}}. Then DiffC0s\mathrm{Diff}^{s}_{C_{0}} is a Hilbert submanifold of Diffs2\mathrm{Diff}^{s-2}. We denote by idMDiffC0s(Diffs)id_{M}\in\mathrm{Diff}^{s}_{C_{0}}(\subset\mathrm{Diff}^{s}) the identity map.

We set s:=Hs(S2TM)C0\mathscr{M}^{s}:=H^{s}(S^{2}T^{*}M)\cap C^{0}\mathscr{M}, where S2TMS^{2}T^{*}M and C0C^{0}\mathscr{M} the tensor field consisting of all symmetric (0,2)-tensors on MM and the set of all C0C^{0} metrics respectively. Then s\mathscr{M}^{s} is a Hilbert manifold modeled on H2(S2TM)H^{2}(S^{2}T^{*}M) (by the Sobolev embedding). And we define a closed Hilbert submanifold of s1\mathscr{M}^{s-1} as Cs:={gs|g||MC||M}={gs|g=fg0forsomefHs1/2(C+(M))onM}\mathscr{M}^{s}_{C}:=\bigl{\{}g\in\mathscr{M}^{s}\bigm{|}g||_{\partial M}\in C||_{\partial M}\bigr{\}}=\bigl{\{}g\in\mathscr{M}^{s}\bigm{|}g=fg_{0}~{}\mathrm{for~{}some}~{}f\in H^{s-1/2}(C^{\infty}_{+}(M))~{}\mathrm{on}~{}\partial M\bigr{\}}. Additionally, we set C0s:={gs|g=fg0onMforsomefHs1/2(C+(M))andHg=0onM}\mathscr{M}^{s}_{C_{0}}:=\bigl{\{}g\in\mathscr{M}^{s}\bigm{|}g=fg_{0}~{}\mathrm{on}~{}\partial M~{}\mathrm{for~{}some}~{}f\in H^{s-1/2}(C^{\infty}_{+}(M))~{}\mathrm{and}~{}H_{g}=0~{}\mathrm{on}~{}\partial M\bigr{\}} and C01s:={gs|g||M1C||M1}={gs|jM1g=jM1(fg0)forsomefHs1/2(C+(M)N)}\mathscr{M}^{s}_{C^{1}_{0}}:=\bigl{\{}g\in\mathscr{M}^{s}\bigm{|}g||^{1}_{\partial M}\in C||^{1}_{\partial M}\bigr{\}}=\bigl{\{}g\in\mathscr{M}^{s}\bigm{|}j^{1}_{\partial M}g=j^{1}_{\partial M}(fg_{0})~{}\mathrm{for~{}some}~{}f\in H^{s-1/2}(C^{\infty}_{+}(M)_{N})\bigr{\}}. Then C0s\mathscr{M}^{s}_{C_{0}} and C01\mathscr{M}_{C^{1}_{0}} are Hilbert submanifolds of s2\mathscr{M}^{s-2}. See [10], [23] for more detail about these spaces. Note that, for ηDiffC0s\eta\in\mathrm{Diff}^{s}_{C_{0}} and gC0(1),ηgC0(1).g\in\mathscr{M}_{C^{(1)}_{0}},~{}\eta^{*}g\in\mathscr{M}_{C^{(1)}_{0}}. Moreover, for gg\in\mathscr{M}, we denote IgI_{g} be the isotropy subgroup of gg in DiffC0s\mathrm{Diff}^{s}_{C_{0}} , that is,

Ig:={ηDiffC0s|ηg=g}.I_{g}:=\bigl{\{}\eta\in\mathrm{Diff}^{s}_{C_{0}}\bigm{|}\eta^{*}g=g\bigr{\}}.

In this section, we shall prove the following theorem:

Theorem 2.1 (Slice theorem for manifold with boundary).

Let s>n2+4s>\frac{n}{2}+4 and

A:DiffC0s+1×CsCsA~{}:~{}\mathrm{Diff}^{s+1}_{C_{0}}\times\mathscr{M}^{s}_{C}\longrightarrow\mathscr{M}^{s}_{C}

be an usual action by pullback. Then for each γC\gamma\in\mathscr{M}_{C} there exsits a submanifold 𝒮Cs\mathcal{S}\subset\mathscr{M}^{s}_{C} containing γ\gamma ,which is diffeomorphic to a ball in a separable real Hilbert space, such that

(1) ηIγA(η,𝒮)=𝒮,\eta\in I_{\gamma}\Rightarrow A(\eta,\mathcal{S})=\mathcal{S},

(2) ηDiffC0s+1,A(η,𝒮)𝒮ηIγ\eta\in\mathrm{Diff}^{s+1}_{C_{0}}~{},~{}A(\eta,\mathcal{S})\cap\mathcal{S}\neq\emptyset\Rightarrow\eta\in I_{\gamma} and

(3)There exists a local section:

χ:(DiffC0s+1/Iγ)UDiffC0s+1\chi~{}:~{}\bigl{(}\mathrm{Diff}^{s+1}_{C_{0}}/I_{\gamma}\supset\bigr{)}U\longrightarrow\mathrm{Diff}^{s+1}_{C_{0}}

defined in a neighborhood UU of the identity coset such that if

F:U×𝒮Cs;(u,t)A(χ(u),t),F~{}:~{}U\times\mathcal{S}\longrightarrow\mathscr{M}^{s}_{C}~{};~{}(u,t)\mapsto A\bigl{(}\chi(u),t\bigr{)},

then FF is a homeomorphism onto a neighborhood of γ.\gamma. Moreover, the same statement holds when we replace Cs\mathscr{M}^{s}_{C} by C0(1)s\mathscr{M}^{s}_{C^{(1)}_{0}}.

First, we get the following lemma:

Lemma 2.1.

Under the identification TidMDiffs+1Hs+1(TM),T_{id_{M}}\mathrm{Diff}^{s+1}\cong H^{s+1}(TM),

(1)TidMDiffC0s+1={XHs+1(TM)|ρHs1/2(M),ig0Xj+g0jXi=ρg0,g0(X,νg0)=0onM},\begin{split}(1)~{}T_{id_{M}}\mathrm{Diff}^{s+1}_{C_{0}}=\bigl{\{}X&\in H^{s+1}(TM)\bigm{|}\exists\rho\in H^{s-1/2}(M),\\ &{}^{g_{0}}\nabla_{i}X_{j}+^{g_{0}}\nabla_{j}X_{i}=\rho g_{0},~{}g_{0}(X,\nu_{g_{0}})=0~{}\mathrm{on}~{}\partial M\bigr{\}},\end{split}

where g0{}^{g_{0}}\nabla and XiX_{i} are respectively the Levi-Civita connection with respect to g0g_{0} and the ii-th component gijXjg_{ij}X^{j} of X=(Xj)X=(X^{j}) in terms of some local coordinates. For gCsg\in\mathscr{M}^{s}_{C},

(2)TgCs={hHs(S2TM)|ρHs1/2(M),h=ρgonM}.(2)~{}~{}~{}~{}~{}T_{g}\mathscr{M}^{s}_{C}=\bigl{\{}h\in H^{s}(S^{2}T^{*}M)\bigm{|}\exists\rho\in H^{s-1/2}(M),h=\rho g~{}\mathrm{on}~{}\partial M\bigr{\}}.

Here, TidMDiffC0s+1T_{id_{M}}\mathrm{Diff}^{s+1}_{C_{0}} and TgCsT_{g}\mathscr{M}^{s}_{C} represent the tangent spaces respectively. Moreover, for gC0sg\in\mathscr{M}^{s}_{C_{0}}, we also have

(3)TgC0s={hHs(S2TM)|ρHs1/2(M),h=ρgonM,DgH(h)=0onM},\begin{split}(3)~{}~{}~{}~{}T_{g}\mathscr{M}^{s}_{C_{0}}=\bigl{\{}h\in H^{s}(S^{2}T^{*}M)\bigm{|}&\exists\rho\in H^{s-1/2}(M),h=\rho g~{}\mathrm{on}~{}\partial M,\\ &D_{g}H(h)=0~{}\mathrm{on}~{}\partial M\bigr{\}},\end{split}

where DgHD_{g}H denotes the derivative of the mean curvatrue function HH at g.g. And, for gC01s,g\in\mathscr{M}^{s}_{C^{1}_{0}},

(4)TgC01s={hHs(S2TM)|ρHs1/2(M),jM1h=jM1(ρg)}.(4)~{}~{}~{}~{}T_{g}\mathscr{M}^{s}_{C^{1}_{0}}=\bigl{\{}h\in H^{s}(S^{2}T^{*}M)\bigm{|}\exists\rho\in H^{s-1/2}(M),j^{1}_{\partial M}h=j^{1}_{\partial M}(\rho g)\bigr{\}}.
Proof.

From the definition, (2), (3) and (4) are obvious. For proving (1), we note that the derivative of a diffeomorphism-action via pull-back on metrics gg at idMid_{M} coincides with the Lie derivative of gg under the above identification (see for instance Lemma 6.2 in [10]). Then the first condition comes from the conformal condition on M\partial M and the second from the fact that any diffeomorphisms map M\partial M to itself. And (3) follows in the same way.

Here, we also note that M\partial M may have some connected components. (Note that the number of components is finite since MM is compact.) In fact, we assume that M=i=1kΣi\partial M=\coprod^{k}_{i=1}\Sigma_{i}, where Σi\Sigma_{i} is a connected component of M\partial M and k1.k\in\mathbb{Z}_{\geq 1}. Then, under the above identification, ηt(Σi)=Σiforallt(ϵ,ϵ)\eta_{t}(\Sigma_{i})=\Sigma_{i}~{}\mathrm{for~{}all}~{}t\in(-\epsilon,\epsilon) for sufficiently small ϵ>0\epsilon>0 as explained below. Here, ηtTidMDiffs+1\eta_{t}\in T_{id_{M}}\mathrm{Diff}^{s+1} is the corresponding curve to a tangent vector. Since MM is compact manifold, we can take some open neighborhoods of each Σi\Sigma_{i}, UiU_{i} such that UiUj=U_{i}\cap U_{j}=\emptyset for all iji\neq j. Consider W(M,U):={fC0(M,M)|f(Σi)Ui}W(\partial M,U):=\bigl{\{}f\in C^{0}(M,M)\bigm{|}f(\Sigma_{i})\subset U_{i}\bigr{\}}, then this is an open subset of C0(M,M)C^{0}(M,M) with respect to the compact-open topology. Hence, this is an open neighborhood of idMid_{M} in C0(M,M)C^{0}(M,M). Since η0=idM\eta_{0}=id_{M} and tηtt\mapsto\eta_{t} is continuous, ηtW(M,U)\eta_{t}\in W(\partial M,U) for all tt with |t|<<1|t|<<1. In paticular, ηt(Σi)=Σi\eta_{t}(\Sigma_{i})=\Sigma_{i} for all ii and tt with |t|<<1|t|<<1. ∎

From [23, Section 9], Hs(TM)H^{s}(TM) is linearly isomorphic to a closed subspace of finite direct sum of Hs(Dn,)H^{s}(D_{n},\mathbb{R}), where DnD_{n} is a closed nn-dimensional disc. Therefore we obtain the following lemma in exactly the same way as in [10, Section 3].

Lemma 2.2 ([10, Section 3]).

DiffC0\mathrm{Diff}_{C_{0}} is a topological group. Furthermore, for all σDiffC0:=sn/2+5DiffC0s\sigma\in\mathrm{Diff}_{C_{0}}:=\bigcap_{s\geq n/2+5}\mathrm{Diff}^{s}_{C_{0}}, the left(right) action Lη(Rη):DiffC0sDiffC0sL_{\eta}(R_{\eta})~{}:~{}\mathrm{Diff}^{s}_{C_{0}}\rightarrow\mathrm{Diff}^{s}_{C_{0}} is smooth.

Remark 2.2.

It is well known that any C1C^{1}-diffeomorphism, which is an isometry of a smooth metric, is smooth (see [10][16][20]). Therefore, from the Sobolev embedding, IγDiffC0I_{\gamma}\subset\mathrm{Diff}_{C_{0}} and IγI_{\gamma} is the same for any such that s>n2+1s>\frac{n}{2}+1.

The following lemma follows from Section 5 in [10] and the fact that DiffC0s\mathrm{Diff}^{s}_{C_{0}} is a submanifold of Diffs2\mathrm{Diff}^{s-2}.

Lemma 2.3 ([10, Section 5]).

(1) The natural inclusion

i:IγDiffC0si~{}:~{}I_{\gamma}\longrightarrow\mathrm{Diff}^{s}_{C_{0}}

is smooth.

(2) ii is embedding, that is for all gIγg\in I_{\gamma} , its derivative DgiD_{g}i is injective and its image is closed in TgDiffC0s.T_{g}\mathrm{Diff}^{s}_{C_{0}}.

(3) The composition map:

c:Iγ×DiffC0sDiffC0s;(g,η)gηc~{}:~{}I_{\gamma}\times\mathrm{Diff}^{s}_{C_{0}}\longrightarrow\mathrm{Diff}^{s}_{C_{0}};~{}(g,\eta)\longmapsto g\circ\eta

is smooth.

(4) Let S:=ηDiffCsTidMRη()S:=\bigcup_{\eta\in\mathrm{Diff}^{s}_{C}}T_{id_{M}}R_{\eta}(\mathscr{I}) , then SS is a CC^{\infty} involutive subbundle of TDiffC0s,T\mathrm{Diff}^{s}_{C_{0}}, that is,

X,YS[X,Y]S,X,Y\in S\Rightarrow[X,Y]\in S,

where \mathscr{I} is the Lie algebra of IγI_{\gamma}.

Using lemma 2.3 and the Frobenius’s theorem(see [19, Chapter 6, Theorem 2]), we obtain a Banach manifold structure on DiffC0s/Iγ\mathrm{Diff}^{s}_{C_{0}}/I_{\gamma} (see [10, Proposition 5.8, 5.9]). And we can show the exsitence of a local section by using this Banach-coordinate-charts:

Lemma 2.4 ([10, Proposition 5.10]).

For any IγηDiffC0s/Iγ,I_{\gamma}\eta\in\mathrm{Diff}^{s}_{C_{0}}/I_{\gamma}, there exists a local section π:DiffC0sDiffC0s/Iγ\pi:~{}\mathrm{Diff}^{s}_{C_{0}}\longrightarrow\mathrm{Diff}^{s}_{C_{0}}/I_{\gamma} defined on a neighborhood IγηI_{\gamma}\eta.

Proof.

Using the above Banach-coordinate-charts, we can construct a local section exactly same as Proposition 5.10 in [10]. ∎

For any γC\gamma\in\mathscr{M}_{C}, σC0(1)\sigma\in\mathscr{M}_{C^{(1)}_{0}}, we define

ψγ:DiffC0s+1Cs;ψγ(η):=ηγ\psi_{\gamma}:\mathrm{Diff}^{s+1}_{C_{0}}\longrightarrow\mathscr{M}^{s}_{C}~{};~{}\psi_{\gamma}(\eta):=\eta^{*}\gamma

and

ψσ0:DiffC0s+1C0(1)s;ψσ(η):=ησ.\psi^{0}_{\sigma}:\mathrm{Diff}^{s+1}_{C_{0}}\longrightarrow\mathscr{M}^{s}_{C^{(1)}_{0}}~{};~{}\psi_{\sigma}(\eta):=\eta^{*}\sigma.

Then, by the definition of IγI_{\gamma}, it naturally induces ϕγ:DiffC0s/IγCs\phi_{\gamma}:~{}\mathrm{Diff}^{s}_{C_{0}}/I_{\gamma}\longrightarrow\mathscr{M}^{s}_{C} and ϕσ0:DiffC0s/IσC0(1)s\phi^{0}_{\sigma}:~{}\mathrm{Diff}^{s}_{C_{0}}/I_{\sigma}\longrightarrow\mathscr{M}^{s}_{C^{(1)}_{0}}. Moreover, from the existence of a local section and the definition, ϕγ,ϕσ0\phi_{\gamma},~{}\phi^{0}_{\sigma} are smooth injective maps.

Here, we will show that ϕγ:DiffC0s+1/IγCs\phi_{\gamma}:~{}\mathrm{Diff}^{s+1}_{C_{0}}/I_{\gamma}\rightarrow\mathscr{M}^{s}_{C} and ϕσ0:DiffC0s+1/IσC0(1)s\phi^{0}_{\sigma}:~{}\mathrm{Diff}^{s+1}_{C_{0}}/I_{\sigma}\rightarrow\mathscr{M}^{s}_{C^{(1)}_{0}} are immersions (i.e., its derivation is injective and has closed image).

Remark 2.3.

We can similarly define ϕγ:Diffs/Iγs\phi_{\gamma}:~{}\mathrm{Diff}^{s}/I_{\gamma}\rightarrow\mathscr{M}^{s}, but it is not an immersion in general.

Next, we will show that the image of DηψγD_{\eta}\psi_{\gamma} is closed in Tψγ(η)CsT_{\psi_{\gamma}(\eta)}\mathscr{M}^{s}_{C} and imege of Dη(ψσ0)D_{\eta}(\psi^{0}_{\sigma}) is closed in Tψσ0(η)C0(1)sT_{\psi^{0}_{\sigma}(\eta)}\mathscr{M}^{s}_{C^{(1)}_{0}}.

As mentioned in the proof of lemma 2.1, under the identification as in lemma 2.1, DidMψγ(X)=X(γ)(XTidMDiffC0s+1)D_{id_{M}}\psi_{\gamma}(X)=\mathscr{L}_{X}(\gamma)~{}(X\in T_{id_{M}}\mathrm{Diff}_{C_{0}}^{s+1}), and DidMψσ0(X)=X(σ),(XTidMDiffC0s+1)D_{id_{M}}\psi^{0}_{\sigma}(X)=\mathscr{L}_{X}(\sigma),~{}(X\in T_{id_{M}}\mathrm{Diff}_{C_{0}}^{s+1}). For simplicity, we put the Lie derivative with respect to γ\gamma as α\alpha, that is α():=(γ).\alpha(\bullet):=\mathscr{L}_{\bullet}(\gamma).

For XHs+1(M),YiHs(M),ZijHs1(M)(i,j=1,,n),X\in H^{s+1}(M),~{}Y_{i}\in H^{s}(M),~{}Z_{ij}\in H^{s-1}(M)~{}(i,j=1,\dots,n), we set

A(X,Y,Z)s+1:={uTidMDiffC0s+1|u=X,iu=Yi,ij2u=ZijonM},A^{s+1}_{(X,Y,Z)}:=\bigl{\{}u\in T_{id_{M}}\mathrm{Diff}^{s+1}_{C_{0}}\bigm{|}u=X,\nabla_{i}u=Y_{i},\nabla^{2}_{ij}u=Z_{ij}~{}\mathrm{on}~{}\partial M\bigr{\}},
B(X,Y,Z)s:={ηTγCs|η(,)=γ(Y(),)+γ(,Y()),Tγr1,2(η)=4(δδu)onM},\begin{split}B^{s}_{(X,Y,Z)}:=\bigl{\{}\eta\in T_{\gamma}\mathscr{M}^{s}_{C}\bigm{|}\eta(\bullet,*)&=\gamma\bigl{(}Y(\bullet),*\bigr{)}+\gamma\bigl{(}\bullet,Y(*)\bigr{)},\\ &-~{}_{\gamma}Tr_{1,2}(\nabla\eta)=4\bigl{(}\delta\delta^{*}u\bigr{)}~{}\mathrm{on}~{}\partial M\bigr{\}},\end{split}

where Tγr1,2{}_{\gamma}Tr_{1,2}, δ\delta and δ\delta^{*} denote respectively the (1,2)-contraction with respect to γ,\gamma, the divergence operator of γ\gamma : δ=γTr1,2\delta=-_{\gamma}Tr_{1,2}\circ\nabla and its formal adjoint : δ=Sym\delta^{*}=Sym\circ\nabla. And “ uu ” in the definition of B(X,Y,Z)sB^{s}_{(X,Y,Z)} is an element of TidMDiffC0s+1T_{id_{M}}\mathrm{Diff}^{s+1}_{C_{0}} such that iu|M=Yiandij2u|M=Zij.\nabla_{i}u|_{\partial M}=Y_{i}~{}\mathrm{and}~{}\nabla^{2}_{ij}u|_{\partial M}=Z_{ij}. Thus, (δδu)|M\bigl{(}\delta\delta^{*}u\bigr{)}|_{\partial M} is determined by them. And these are Hilbert spaces.

Since the composition map αα\alpha^{*}\alpha is elliptic, we obtain the following boundary estimate (see [12, Theorem 6.6]):

There exists a positive constant CC such that for any uA(X,Y,Z)s+1u\in A^{s+1}_{(X,Y,Z)},

||u||kC(||(αα)u||k2+||u||k2+||X~||k)||u||_{k}\leq C\bigl{(}\bigl{|}\bigl{|}(\alpha^{*}\alpha)u\bigr{|}\bigr{|}_{k-2}+||u||_{k-2}+\bigl{|}\bigl{|}\tilde{X}\bigr{|}\bigr{|}_{k}\bigr{)}

where X~\tilde{X} is an extension to MM of X.X.

For uTid.DiffC0s+1u\in T_{id.}\mathrm{Diff}^{s+1}_{C_{0}} and hTγCsh\in T_{\gamma}\mathscr{M}^{s}_{C} , from the Green’s formula,

(αu,h)=(u,αh)+2(u,h(ν,))M,(\alpha u,h)=(u,\alpha^{*}h)+2\bigl{(}u^{\flat},h(\nu,\bullet)\bigr{)}_{\partial M},

where (,)and(,)M(,)~{}\mathrm{and}~{}(,)_{\partial M} are L2L^{2}-inner product with respect to γandγ|M\gamma~{}\mathrm{and}~{}\gamma|_{\partial M} . And ν\nu is the outer unit normal vector along M\partial M with respect to γ.\gamma.

Since hTid.DiffC0s+1h\in T_{id.}\mathrm{Diff}^{s+1}_{C_{0}}, from lemma 2.1, the second term vanishes. Thus we get

(αu,h)=(u,αh).(\alpha u,h)=(u,\alpha^{*}h).

Therefore for all u,vTid.DiffC0s+1u,v\in T_{id.}\mathrm{Diff}^{s+1}_{C_{0}},

((αα)u,v)=(u,(αα)v).\bigl{(}(\alpha^{*}\alpha)u,v\bigr{)}=\bigl{(}u,(\alpha^{*}\alpha)v\bigr{)}. (1)

From the closed range theorem([12, Lemma 5.10]), Proposition 6.8 and 6.9 in [10], we get the following:

Lemma 2.5.

α(A(X,Y,Z)s+1)\alpha(A^{s+1}_{(X,Y,Z)}) is a closed subspace of B(X,Y,Z)sB^{s}_{(X,Y,Z)} and there is an orthogonal decomposition:

B(X,Y,Z)s=Im(α|A(X,Y,Z)s+1)Ker(α|B(X,Y,Z)s).B^{s}_{(X,Y,Z)}=\mathrm{Im}\bigl{(}\alpha|_{A^{s+1}_{(X,Y,Z)}}\bigr{)}\oplus\mathrm{Ker}\bigl{(}\alpha^{*}|_{B^{s}_{(X,Y,Z)}}\bigr{)}.

From lemma 2.5, we get that

TγCs=Im(α|Tid.DiffC0s+1)+Ker(α|TγCs).T_{\gamma}\mathscr{M}^{s}_{C}=\mathrm{Im}\bigl{(}\alpha|_{T_{id.}\mathrm{Diff}^{s+1}_{C_{0}}}\bigr{)}+\mathrm{Ker}\bigl{(}\alpha^{*}|_{T_{\gamma}\mathscr{M}^{s}_{C}}\bigr{)}.

On the other hand, from the equation (1),

Im(α|Tid.DiffC0s+1)Ker(α|TγCs).\mathrm{Im}\bigl{(}\alpha|_{T_{id.}\mathrm{Diff}^{s+1}_{C_{0}}}\bigr{)}\perp\mathrm{Ker}\bigl{(}\alpha^{*}|_{T_{\gamma}\mathscr{M}^{s}_{C}}\bigr{)}.

Hence we get an orthogonal decomposition

TγCs=Im(α|Tid.DiffC0s+1)Ker(α|TγCs).T_{\gamma}\mathscr{M}^{s}_{C}=\mathrm{Im}\bigl{(}\alpha|_{T_{id.}\mathrm{Diff}^{s+1}_{C_{0}}}\bigr{)}\oplus\mathrm{Ker}\bigl{(}\alpha^{*}|_{T_{\gamma}\mathscr{M}^{s}_{C}}\bigr{)}.

In paticular, ImDid.ψ\mathrm{Im}~{}D_{id.}\psi is closed subspace of TγCsT_{\gamma}\mathscr{M}^{s}_{C}. Therefore, for each ηDiffC0s+1\eta\in\mathrm{Diff}^{s+1}_{C_{0}}, ImDηψγ\mathrm{Im}~{}D_{\eta}\psi_{\gamma} is a closed subspace of Tψγ(η)CsT_{\psi_{\gamma}(\eta)}\mathscr{M}^{s}_{C}.

Similarly, (however we consider up to upper order boundary datum,) we also get the decomposition

TγC0(1)s=Im(α|Tid.DiffC0s+1)Ker(α|TγC0(1)s).T_{\gamma}\mathscr{M}^{s}_{C^{(1)}_{0}}=\mathrm{Im}\bigl{(}\alpha|_{T_{id.}\mathrm{Diff}^{s+1}_{C_{0}}}\bigr{)}\oplus\mathrm{Ker}\bigl{(}\alpha^{*}|_{T_{\gamma}\mathscr{M}^{s}_{C^{(1)}_{0}}}\bigr{)}.

and that ImDηψσ0\mathrm{Im}~{}D_{\eta}\psi^{0}_{\sigma} is a closed subspace of Tψσ0(η)C0(1)sT_{\psi^{0}_{\sigma}(\eta)}\mathscr{M}^{s}_{C^{(1)}_{0}} for all ηDiffC0s+1\eta\in\mathrm{Diff}^{s+1}_{C_{0}}.

Moreover, we can show that D[η]ϕγD_{[\eta]}\phi_{\gamma} and D[η]ϕσ0D_{[\eta]}\phi^{0}_{\sigma} are injective in the same way as Proposition 6.11 in [10]. Consequently, we obtain the following:

Lemma 2.6.
ϕγ:DiffC0s+1/IγCsandϕσ0:DiffC0s+1/IγC0(1)s\phi_{\gamma}~{}:~{}\mathrm{Diff}^{s+1}_{C_{0}}/I_{\gamma}\longrightarrow\mathscr{M}^{s}_{C}~{}\mathrm{and}~{}\phi^{0}_{\sigma}~{}:~{}\mathrm{Diff}^{s+1}_{C_{0}}/I_{\gamma}\longrightarrow\mathscr{M}^{s}_{C^{(1)}_{0}}

are smooth injective immersions.

Moreover, we can prove the following in the same way as Proposition 6.13 in [10]:

Lemma 2.7.

Let s>n2+4s>\frac{n}{2}+4 , then

ϕγ:DiffC0s+1/IγCs\phi_{\gamma}~{}:~{}\mathrm{Diff}^{s+1}_{C_{0}}/I_{\gamma}\longrightarrow\mathscr{M}^{s}_{C}

is a homeomorphism mapping DiffC0s+1/Iγ\mathrm{Diff}^{s+1}_{C_{0}}/I_{\gamma} onto a closed subspace of Cs\mathscr{M}^{s}_{C} . Therefore, in paticular, ϕγ\phi_{\gamma} is an embedding. The same statement also holds when we replace respectively Cs\mathscr{M}^{s}_{C} and ϕ\phi by C0(1)s\mathscr{M}^{s}_{C^{(1)}_{0}} and ϕ0\phi^{0}.

Remark 2.4.

The connectedness of MM was used in the proof of lemma 2.7 (see the proof of Proposition 6.13 in [10]).

Proof of the Theorem 2.1.

The proof is same as in [10]. Because we can show in the same way for C0(1)\mathscr{M}_{C^{(1)}_{0}}, we only descrie about C\mathscr{M}_{C}.

For γC\gamma\in\mathscr{M}_{C} , we set Oγs:=ϕγ(DiffC0s+1/Iγ)O^{s}_{\gamma}:=\phi_{\gamma}\bigl{(}\mathrm{Diff}^{s+1}_{C_{0}}/I_{\gamma}\bigr{)} the orbit of the action AA through γ\gamma (where ϕγ\phi_{\gamma} is in the lemma 2.7). From Lemma 2.7, this is a closed submanifold of Cs.\mathscr{M}^{s}_{C}. Moreover, we define its normal vector bundle:

ν:={VTCs|Oγs|(W,V)γ=0,WTOγs},\nu:=\bigl{\{}V\in T\mathscr{M}^{s}_{C}|_{O^{s}_{\gamma}}\bigm{|}(W,V)_{\gamma}=0~{},~{}\forall W\in TO^{s}_{\gamma}\bigr{\}},

where (,)γ:=M,γdvγ.(*,*)_{\gamma}:=\mathop{\text{\LARGE$\int_{\text{\normalsize$\scriptstyle M$}}$}}\nolimits\langle*,*\rangle_{\gamma}dv_{\gamma}.

Our first step is constructing the normal bundle ν\nu of OγsO^{s}_{\gamma} in Cs.\mathscr{M}^{s}_{C}. As stated in [10], this Riemannian metric is strong on H0,H^{0}, but is not on Hs(s1).H^{s}~{}(s\geq 1). Thus we do not know automatically that ν\nu is a CC^{\infty} subbundle of TCs|Oγs.T\mathscr{M}^{s}_{C}|_{O^{s}_{\gamma}}.

To show this, we shall find a CC^{\infty} surjective vector-bundle-map:

P:TCs|OγsTOγsP~{}:~{}T\mathscr{M}^{s}_{C}|_{O^{s}_{\gamma}}\longrightarrow TO^{s}_{\gamma}

such that KerP=ν\mathrm{Ker}~{}P=\nu (see [19, Chapter3, Section3]).

Since, from the proof of the lemma 2.6,

TγCs=ImαKerα.T_{\gamma}\mathscr{M}^{s}_{C}=\mathrm{Im}~{}\alpha\oplus\mathrm{Ker}~{}\alpha^{*}.

Hence, from the definition of Oγs,O^{s}_{\gamma},

Imα=TγOγs.\mathrm{Im}~{}\alpha=T_{\gamma}O^{s}_{\gamma}.

Thus

νγ(Oγs)=Ker(α|TγCs).\nu_{\gamma}(O^{s}_{\gamma})=\mathrm{Ker}\bigl{(}\alpha^{*}|_{T_{\gamma}\mathscr{M}^{s}_{C}}\bigr{)}.

Moreover, since the weak Riemannian metric (,)γ(~{},~{})_{\gamma} is invariant under the action of DiffCs+1\mathrm{Diff}^{s+1}_{C} ([10, Section 4]), νηγ(Oγs)=η(Kerα).\nu_{\eta^{*}\gamma}(O^{s}_{\gamma})=\eta^{*}\bigl{(}\mathrm{Ker}~{}\alpha^{*}\bigr{)}. On the orbit of γ,\gamma, we define

P:=α(αα)1α:TγCs|OγsTγOγsP:=\alpha\circ(\alpha^{*}\circ\alpha)^{-1}\circ\alpha^{*}~{}:~{}T_{\gamma}\mathscr{M}^{s}_{C}|_{O^{s}_{\gamma}}\longrightarrow T_{\gamma}O^{s}_{\gamma}

Thus, as in the same way in [10, Theorem 7.1], we can show that this PP satisfy the above properties.

Next, we shall construct the slice 𝒮\mathcal{S} of γ.\gamma. To do this, we consider the exponential map of (,)γ,(~{},~{})_{\gamma}, exp:TCsCs.\exp:~{}T\mathscr{M}^{s}_{C}\longrightarrow\mathscr{M}^{s}_{C}. Thus we know the following fact:

Fact 2.2 ([10, Section 4]).

This is a smooth map and exp|ν:νCsexp|_{\nu}:~{}\nu\longrightarrow\mathscr{M}^{s}_{C} is a diffeomorphism mapping a neighborhood of the zero section of ν\nu to a neighborhood of OγsO^{s}_{\gamma} in Cs.\mathscr{M}^{s}_{C}. Moreover, since AA is continuous and expexp and the action of η\eta are commutative, there are a neighborhood UU of γ\gamma in OγsO^{s}_{\gamma} and a neighborhood VV of 0 in νγ\nu_{\gamma} such that

νW:={η(v)|vV,ηχ(U)}.\nu\supset W:=\bigl{\{}\eta^{*}(v)\bigm{|}v\in V~{},~{}\eta\in\chi(U)\bigr{\}}.

Then exp|Wexp|_{W} is a diffeomorphism mapping WW onto a neighborhood of γ.\gamma. Moreover, if necessary, we shall take UU and VV small enough so that exp(W)Oγs=U.\exp(W)\cap O^{s}_{\gamma}=U.

Consider the strong inner product (,)γs(~{},~{})_{\gamma}^{s} on Hs(S2T),H^{s}(S^{2}T^{*}), defined as at the end of Section 4 in [10]. Now let ρs\rho_{s} be the metric defined on Cs\mathscr{M}^{s}_{C} by (,)γs.(~{},~{})^{s}_{\gamma}. Let BγrB^{r}_{\gamma} be the open ball about γ\gamma of radius rr with respect to ρs.\rho_{s}. Then, for some positive δ,\delta, exp(W)Bγ2δ.\exp(W)\supset B^{2\delta}_{\gamma}. Pick U1U,ϵ1<ϵ(V1V)U_{1}\subset U,~{}\epsilon_{1}<\epsilon~{}(\leadsto V_{1}\subset V) so that if W1:={η(v)|vV1,ηχ(U1)},W_{1}:=\bigl{\{}\eta^{*}(v)\bigm{|}v\in V_{1}~{},~{}\eta\in\chi\bigl{(}U_{1}\bigr{)}\bigr{\}}, then exp(W1)Bγδ.\exp(W_{1})\subset B^{\delta}_{\gamma}.

Then we set

𝒮:=exp(V1)\mathcal{S}:=\exp(V_{1})

and this 𝒮\mathcal{S} has the three properties of a slice (These are checked in the same way in [10, Section 7 and the proof of Theorem 7.1]).

3 Main Results

Before starting the proof of Main theorem, we shall line up some basic definitions below:

Definition 3.1 ([21]).

(1) A topological space EE is called ILH-space if EE is an inverse limit of Hilbert spaces {Ei}i1,\{E_{i}\}_{i\in\mathbb{Z}_{\geq 1}}, such that EjEi(ij)E_{j}\subset E_{i}~{}(i\leq j) and each inclusions are bounded linear operators.

(2) A topological space XX is called CkC^{k}-ILH-manifold modeled on EE if XX has the following (a) and (b):

(a) XX is an inverse limit of CkC^{k}-Hilbert manifolds {Xi}i1\{X_{i}\}_{i\in\mathbb{Z}_{\geq 1}} modeled on EiE_{i} such that XjXi(ij),X_{j}\subset X_{i}~{}(i\leq j),

(b) For each xXx\in X and i,i, there is an open neighborhood XiUi(x)X_{i}\supset U_{i}(x) and homeomorphism ψi\psi_{i} from Ui(x)U_{i}(x) onto an open subset ViV_{i} in EiE_{i} which gives a CkC^{k}-coordinate around xx in XiX_{i} and satisfies Uj(x)Ui(x)(ij),ψi+1(y)=ψi(y)U_{j}(x)\subset U_{i}(x)~{}(i\leq j),~{}\psi_{i+1}(y)=\psi_{i}(y) for all yUi+1(x).y\in U_{i+1}(x).

(3) Let XX be a CkC^{k}-ILH-manifold( k1k\geq 1 ) and TXiTX_{i} the tangent budle of Xi.X_{i}. The inverse limit of {TXi}\{TX_{i}\} is called ILH-tangent bundle of XX.

(4) Let X,YX,Y be CkC^{k}-ILH-manifolds. A mapping  ϕ:XY\phi:~{}X\rightarrow Y is called ClC^{l}-ILH-differentiable ( lkl\leq k ) if ϕ\phi is an inverse limit of ClC^{l}-differentiable maping {ϕi}i1\{\phi_{i}\}_{i\in\mathbb{Z}_{\geq 1}} (that is, for each i,i, there exists j(i)j(i) and ClC^{l}-map ϕi:Xj(i)Yi\phi_{i}:X_{j(i)}\rightarrow Y_{i} such that ϕi(x)=ϕi+1(x)\phi_{i}(x)=\phi_{i+1}(x) for all xXj(i+1).)x\in X_{j(i+1)}.)

(5) XX is a ILH-manifold if XX is a CkC^{k}-ILH-manifold for all k0.k\geq 0.

(6) Let X,YX,Y be ILH\mathrm{ILH}-manifolds. A mapping ϕ:XY\phi:X\rightarrow Y is called ILH-differentiable if ϕ:XY\phi:X\rightarrow Y is CkC^{k}-ILH-differentiable for all k0.k\geq 0.

(7) Let TxXiT_{x}X_{i} be the tangent space of XiX_{i} at xx and TxXT_{x}X the inverse limit of {TxXi}.\{T_{x}X_{i}\}. Let

Drϕi(x):l=1rTxXj(i)Tϕ(x)YiD^{r}\phi_{i}(x):~{}\prod^{r}_{l=1}T_{x}X_{j(i)}\longrightarrow T_{\phi(x)}Y_{i}

be the rr-th (Fréchet) derivative of ϕi\phi_{i} at x.x. Then, {Drϕi(x)}i1\{D^{r}\phi_{i}(x)\}_{i\in\mathbb{Z}_{\geq 1}} has the inverse limit

limDrϕi(x):l=1rTxXTϕ(x)Y.\lim_{\leftarrow}D^{r}\phi_{i}(x):~{}\prod^{r}_{l=1}T_{x}X\longrightarrow T_{\phi(x)}Y.

It is called rr-th derivative of ϕ\phi and we denote it by Drϕ(x).D^{r}\phi(x).

Let MM and g0g_{0} be the same as in Section 2 and use the same notations there. As in the closed case, :=limr\mathscr{M}:=\lim_{\leftarrow}\mathscr{M}^{r} , Diff(M):=limDiffr\mathrm{Diff}(M):=\lim_{\leftarrow}\mathrm{Diff}^{r} naturally become ILH-manifolds and the pullback-action A:Diff(M)×A:~{}\mathrm{Diff}(M)\times\mathscr{M}\longrightarrow\mathscr{M} is ILH-differentiable. Moreover, for a fixed metric g0g_{0} on MM, since each Cr\mathscr{M}^{r}_{C} are a submanifold of r1\mathscr{M}^{r-1}, C:=limCr\mathscr{M}_{C}:=\lim_{\leftarrow}\mathscr{M}^{r}_{C} is an ILH-submanifold of \mathscr{M} and the inclusion C\mathscr{M}_{C}\hookrightarrow\mathscr{M} is CC^{\infty}-differentiable. And, DiffC0:=limDiffC0r\mathrm{Diff}_{C_{0}}:=\lim_{\leftarrow}\mathrm{Diff}^{r}_{C_{0}} is an ILH-submanifold of Diff(M)\mathrm{Diff}(M) and the inclusion DiffC0Diff(M)\mathrm{Diff}_{C_{0}}\hookrightarrow\mathrm{Diff}(M) is CC^{\infty}-differentiable. Similarly, C0(1):=limC0(1)r\mathscr{M}_{C^{(1)}_{0}}:=\lim_{\leftarrow}\mathscr{M}^{r}_{C^{(1)}_{0}} is an ILH-submanifold of \mathscr{M} and the inclusion C0(1)\mathscr{M}_{C^{(1)}_{0}}\hookrightarrow\mathscr{M} is CC^{\infty} -differenciable. Note that the pull-back action A:DiffC0×C0(1)C0(1)A:\mathrm{Diff}_{C_{0}}\times\mathscr{M}_{C^{(1)}_{0}}\rightarrow\mathscr{M}_{C^{(1)}_{0}} is also CC^{\infty}-differentiable. By the Sobolev embedding for fibre bundles over manifold with boundary, we obtain the following (see [4]):

Lemma 3.1.

Let E,FE,~{}F be vector bundles over MM and let f:EFf:~{}E\rightarrow F be a CC^{\infty}-differentiable which preserves each fibers.

Let s>n2s>\frac{n}{2} .Then the bundle map induced by ff

ϕ:Hs(E)Hs(F);ϕ(α):=fα\phi:~{}H^{s}(E)\rightarrow H^{s}(F)~{};~{}\phi(\alpha):=f\circ\alpha

is CC^{\infty}-differentiable.

Proof.

Same as Lemma 1.1 in [18]. See also [23, Theorem 11.3]. ∎

Using this lemma and that Cr\mathscr{M}^{r}_{C} is a submanifold of \mathscr{M}, the following hold in the same as in [18]:

Lemma 3.2 ([18, Proposition1.2 , Corollary1.3 , Corollary1.4]).

Let s>n2s>\frac{n}{2} .

(1)  D:Cs+1×Hs+1(Tqp)Hs(Tq+1p);(g,ξ)gξD:~{}\mathscr{M}^{s+1}_{C}\times H^{s+1}(T^{p}_{q})\rightarrow H^{s}(T^{p}_{q+1})~{};~{}(g,\xi)\mapsto\nabla_{g}\xi

is CC^{\infty} -differentiable, where TqpT^{p}_{q} is the type (p,q)(p,q) tensor bundle and g\nabla_{g} is the Levi-Civita connection with respect to gg .

(2)  Cs+1×Hs+2(M)Hs(M);(g,f)Δgf:=gdf.\mathscr{M}^{s+1}_{C}\times H^{s+2}(M)\rightarrow H^{s}(M)~{};~{}(g,f)\mapsto\Delta_{g}f:=\nabla_{g}df.

is CC^{\infty} -differentiable.

(3) Mappings listed below are CC^{\infty} -differentiable:

Cs+2Hs(S2T);gRicg(theRiccicurvatureofg).\mathscr{M}^{s+2}_{C}\rightarrow H^{s}(S^{2}T^{*})~{};~{}g\mapsto{\rm Ric_{g}}~{}(\mathrm{the~{}Ricci~{}curvature~{}of}~{}g).
Cs+2Hs(M);gRg(thescalarcurvatureofg).\mathscr{M}^{s+2}_{C}\rightarrow H^{s}(M)~{};~{}g\mapsto R_{g}~{}(\mathrm{the~{}scalar~{}curvature~{}of}~{}g).
Cs+2Hs+1/2(M);gHg(themeancurvatureofgalongM).\mathscr{M}^{s+2}_{C}\rightarrow H^{s+1/2}(\partial M)~{};~{}g\mapsto H_{g}~{}(\mathrm{the~{}mean~{}curvature~{}of}~{}g~{}\mathrm{along}~{}\partial M).
Remark 3.1.

Since C0\mathscr{M}_{C_{0}} is an ILH submanifold of \mathscr{M}, the same statements hold in the above lemma replaced C\mathscr{M}_{C} by C0(1)\mathscr{M}_{C^{(1)}_{0}}.

Let r>n2+4r>\frac{n}{2}+4. We define

C,1r:={gCr|Volg(M)=1}\mathscr{M}^{r}_{C,1}:=\bigl{\{}g\in\mathscr{M}^{r}_{C}\bigm{|}{\rm Vol}_{g}(M)=1\bigr{\}},  C,1:=rC,1r\mathscr{M}_{C,1}:=\bigcap_{r}\mathscr{M}^{r}_{C,1},

𝔖C0(1)r:={gC,1rC0(1)r|Rg=const}\mathfrak{S}^{r}_{C^{(1)}_{0}}:=\bigl{\{}g\in\mathscr{M}^{r}_{C,1}\cap\mathscr{M}^{r}_{C^{(1)}_{0}}\bigm{|}R_{g}={\rm const}\bigr{\}},  𝔖C0(1):=r𝔖C0(1)r\mathfrak{S}_{C^{(1)}_{0}}:=\bigcap_{r}\mathfrak{S}^{r}_{C^{(1)}_{0}},

𝔖ˇC0(1):={g𝔖C0(1)|Rgn1Spec(Δg;Neumann)}\check{\mathfrak{S}}_{C^{(1)}_{0}}:=\Biggl{\{}g\in\mathfrak{S}_{C^{(1)}_{0}}\Biggm{|}\frac{R_{g}}{n-1}\notin\mathrm{Spec}(-\Delta_{g};\mathrm{Neumann})\Biggr{\}}.

For g¯,gC,1r,\bar{g},g\in\mathscr{M}^{r}_{C,1}, we define

Φg(g¯)r:Hg¯r(M)(1,1)g¯Hr32(M)\Phi^{r}_{g~{}(\bar{g})}:~{}H^{r}_{\bar{g}}(M)\longrightarrow\langle(1,-1)\rangle^{\perp_{\bar{g}}}\oplus H^{r-\frac{3}{2}}(\partial M)

by

fΦg(g¯)r(f):=((n1)(Δg)2fRgΔgfM{(n1)(Δg)2f+RgΔgf}dvg¯+Mνg{(n1)ΔgfRgf}dsg¯|M,νg{(n1)ΔgfRgf}|M,νg(f)|M),\begin{split}f&\mapsto\Phi^{r}_{g~{}(\bar{g})}(f):=\Biggl{(}(n-1)(\Delta_{g})^{2}f-R_{g}-\Delta_{g}f-\int_{M}\bigl{\{}(n-1)(\Delta_{g})^{2}f+R_{g}\Delta_{g}f\bigr{\}}dv_{\bar{g}}\\ &\quad+\int_{\partial M}\nu_{g}\{-(n-1)\Delta_{g}f-R_{g}f\}ds_{\bar{g}|_{\partial M}},~{}\nu_{g}\{-(n-1)\Delta_{g}f-R_{g}f\}\bigm{|}_{\partial M},~{}\nu_{g}(f)\bigm{|}_{\partial M}\Biggr{)},\end{split}

where Hr(TM),Hr(TM)H^{r}(TM),H^{r}(T\partial M) are defined by fixed g¯,\bar{g}, νg\nu_{g} is the outer unit normal vector along M\partial M with respect to g,g, Hg¯r(M):={fHr(M)|Mf𝑑vg¯=0},H^{r}_{\bar{g}}(M):=\bigl{\{}f\in H^{r}(M)\bigm{|}\int_{M}f~{}dv_{\bar{g}}=0\bigr{\}}, and

(1,1)g¯:={(u,v)Hr4(M)Hr52(M)|((u,v),(1,1))L2(g¯)=0}.\langle(1,-1)\rangle^{\perp_{\bar{g}}}:=\bigl{\{}(u,v)\in H^{r-4}(M)\oplus H^{r-\frac{5}{2}}(\partial M)\bigm{|}\bigl{(}(u,v),(1,-1)\bigr{)}_{L^{2}(\bar{g})}=0\bigr{\}}. Here, dsg¯|Mds_{\bar{g}|_{\partial M}} also denotes the volume measure with respect to g¯|M.\bar{g}|_{\partial M}.

From Lemma 3.2 and the trace theorem([13, Appendix B]), (g,f)Φgr(f)(g,f)\mapsto\Phi^{r}_{g}(f) is a CC^{\infty} -differentiable map from C,1r×Hg¯r(M)\mathscr{M}^{r}_{C,1}\times H^{r}_{\bar{g}}(M) to Hg0r4(M)Hr52(M)Hr32(M).H^{r-4}_{g_{0}}(M)\oplus H^{r-\frac{5}{2}}(\partial M)\oplus H^{r-\frac{3}{2}}(\partial M).

And we define

𝒦Cr:={gC,1r|g¯C,1s.t.Φg(g¯)risanisomorphism}.\mathcal{K}^{r}_{C}:=\bigl{\{}g\in\mathscr{M}^{r}_{C,1}\bigm{|}\exists\bar{g}\in\mathscr{M}_{C,1}~{}~{}{\rm s.t.}~{}~{}\Phi^{r}_{g~{}(\bar{g})}~{}\mathrm{is~{}an~{}isomorphism}\bigr{\}}.

Then, the following holds:

Lemma 3.3.

𝒦Cr\mathcal{K}^{r}_{C} is an open subset of C,1r.\mathscr{M}^{r}_{C,1}.

Proof.
gΦgrg\mapsto\Phi^{r}_{g}

is a diffeomorphism mapping C,1r\mathscr{M}^{r}_{C,1} to (Hg¯r(M),(1,1)g¯Hr32(M))\mathcal{L}\bigl{(}H^{r}_{\bar{g}}(M),\langle(1,-1)\rangle^{\perp_{\bar{g}}}\oplus H^{r-\frac{3}{2}}(\partial M)\bigr{)} .

On the other hand, the set of all isomorphisms is open in

(Hg¯r(M),(1,1)g¯Hr32(M))\mathcal{L}\bigl{(}H^{r}_{\bar{g}}(M),\langle(1,-1)\rangle^{\perp_{\bar{g}}}\oplus H^{r-\frac{3}{2}}(\partial M)\bigr{)}

with respect to the operator-norm. Hence 𝒦Cr\mathcal{K}^{r}_{C} is an open subset of C,1r\mathscr{M}^{r}_{C,1}. ∎

Lemma 3.4.

𝔖ˇC0=C0𝒦Cr𝔖C0r\check{\mathfrak{S}}_{C_{0}}=\mathscr{M}_{C_{0}}\cap\mathcal{K}^{r}_{C}\cap\mathfrak{S}^{r}_{C_{0}} and 𝔖ˇC01=C01𝒦Cr𝔖C01r.\check{\mathfrak{S}}_{C^{1}_{0}}=\mathscr{M}_{C^{1}_{0}}\cap\mathcal{K}^{r}_{C}\cap\mathfrak{S}^{r}_{C^{1}_{0}}.

Proof.

Since the proof for 𝔖ˇC01\check{\mathfrak{S}}_{C^{1}_{0}} is exactly the same as for 𝔖ˇC0\check{\mathfrak{S}}_{C_{0}}, we will only prove for 𝔖ˇC0.\check{\mathfrak{S}}_{C_{0}}.

( \subset )

Fix g𝔖ˇC0g\in\check{\mathfrak{S}}_{C_{0}} .

surjectivity ;

Firstly, we consider the case that Rg0R_{g}\neq 0.

Given (F,G,H)(1,1)gHr32(M)(F,G,H)\in\langle(1,-1)\rangle^{g}\oplus H^{r-\frac{3}{2}}(\partial M) , we consider two boundary value problem:

Δgu=F,νg(u)|M=G,-\Delta_{g}u=F~{},~{}\nu_{g}(u)\Bigm{|}_{\partial M}=G, (2)
ΔgvRgn1v=u,νg(v)|M=H,-\Delta_{g}v-\frac{R_{g}}{n-1}v=u~{},~{}\nu_{g}(v)\bigm{|}_{\partial M}=H, (3)

where uHr2(M),vHgr(M).u\in H^{r-2}(M),~{}v\in H^{r}_{g}(M).

We firstly consider (3). For a fixed positive constant α>0,\alpha\in\mathbb{R}_{>0}, we consider

Δgv+αv=u,νg(v)|M=H.-\Delta_{g}v+\alpha v=u~{},~{}\nu_{g}(v)\bigm{|}_{\partial M}=H.

Thus we can show that there is a unique solution by using standard variational argument.

Let

L~g:Hr(M)Hr2(M)Hr32(M)\tilde{L}_{g}:~{}H^{r}(M)\longrightarrow H^{r-2}(M)\oplus H^{r-\frac{3}{2}}(\partial M)
;u(Δgu+αu,νg(u)|M);~{}u\mapsto\bigl{(}-\Delta_{g}u+\alpha u~{},~{}\nu_{g}(u)\bigm{|}_{\partial M}\bigr{)}

be the operatopr corresponding to the above equation, then this is an elliptic operator in the sence of Definition 20.1.1 in [13].

Therefore, from Theorem 20.1.2 in [13], L~g\tilde{L}_{g} is a Fredholm operator. Hence dimKerL~g<,\mathrm{dim}~{}\mathrm{Ker}\tilde{L}_{g}<\infty, dimCokerL~g<\mathrm{dim}~{}\mathrm{Coker}\tilde{L}_{g}<\infty and ImL~g\mathrm{Im}\tilde{L}_{g} is closed. Moreover, because of the existence and uniqueness of the above boundary value problem, ind(L~g)=dimKerL~gdimCokerL~g=0\mathrm{ind}(\tilde{L}_{g})=\mathrm{dim}~{}\mathrm{Ker}\tilde{L}_{g}-\mathrm{dim}~{}\mathrm{Coker}\tilde{L}_{g}=0 .

We shall back to (3). We consider the corresponding operator:

Lg:Hr(M)Hr2(M)Hr32(M)L_{g}:~{}H^{r}(M)\longrightarrow H^{r-2}(M)\oplus H^{r-\frac{3}{2}}(\partial M)
;u(ΔguRgn1u,νg(u)|M),;~{}u\mapsto\bigl{(}-\Delta_{g}u-\frac{R_{g}}{n-1}u~{},~{}\nu_{g}(u)\bigm{|}_{\partial M}\bigr{)},

then, this operator is also an elliptic operator. From Theorem 20.1.8 in [13], ind(Lg)=ind(L~g)\mathrm{ind}(L_{g})=\mathrm{ind}(\tilde{L}_{g}) and since g𝔖ˇC0,g\in\check{\mathfrak{S}}_{C_{0}}, KerLg={0}.\mathrm{Ker}L_{g}=\{0\}. Hence dimCokerLg=0.\mathrm{dim}~{}\mathrm{Coker}L_{g}=0. Therefore LgL_{g} is surjective.

Next, we consider (2). The ellipticity only depends on its principal symbol, thus (2) is also elliptic (Exactly speaking, the operator corresponding to (2) is elliptic). Let

Lˇg:Hr(M)Hr2(M)Hr32(M)\check{L}_{g}:~{}H^{r}(M)\longrightarrow H^{r-2}(M)\oplus H^{r-\frac{3}{2}}(\partial M)
;u(Δgu,νg(u)|M);~{}u\mapsto\bigl{(}-\Delta_{g}u~{},~{}\nu_{g}(u)\bigm{|}_{\partial M}\bigr{)}

be the corresponding operator, then from Theorem 20.1.8 in [13] and the above things,

0=ind(L~g)=ind(Lˇg).0=\mathrm{ind}(\tilde{L}_{g})=\mathrm{ind}(\check{L}_{g}).

Since KerLˇg=(={constantfunctions}),\mathrm{Ker}\check{L}_{g}=\mathbb{R}~{}(=\{\mathrm{constant~{}functions}\}), dimKerLˇg=1.\mathrm{dim}~{}\mathrm{Ker}\check{L}_{g}=1.

Thus dimCokerLˇg=1.\mathrm{dim}\mathrm{Coker}\check{L}_{g}=1. On the other hand, from the Green’s formula,

(F,G)ImLˇgMF𝑑vgMG𝑑sg|M=0.(F,G)\in\mathrm{Im}~{}\check{L}_{g}\Rightarrow\int_{M}F~{}dv_{g}-\int_{\partial M}G~{}ds_{g|_{\partial M}}=0.

Hence CokerLˇg(1,1).\mathrm{Coker}\check{L}_{g}\cong\langle(1,-1)\rangle. Thus ImLˇg=(1,1)g.\mathrm{Im}\check{L}_{g}=\langle(1,-1)\rangle^{\perp_{g}}. Therefore Φg(g)r\Phi^{r}_{g~{}(g)} is surjective if Rg0.R_{g}\neq 0.

Next, we consider the case that Rg=0.R_{g}=0. The above observation implies that (2) has a unique solution up to constants. That is, if we take a solution u(F,G)u(F,G) of (2), then u(F,G)+Cu(F,G)+C ( CC is arbitrary constant) is also solution of this equation. Hence, for given HH, we take a constant CC so that

Mu(F,G)+Cdvg=MH𝑑sg|M.\int_{M}u(F,G)+C~{}dv_{g}=-\int_{\partial M}H~{}ds_{g|_{\partial M}}.

Then, from the above observation, there exists a solution vv of (3). Therefore Φg(g)r\Phi^{r}_{g~{}(g)} is also surjective if Rg=0R_{g}=0.

injectivity ;

Let Φgr(u)=0\Phi^{r}_{g}(u)=0 , then

(n1)(Δg)2u+RgΔgu=0,(n-1)(\Delta_{g})^{2}u+R_{g}\Delta_{g}u=0, (4)
νg{(n1)ΔguRgu}|M=0,\nu_{g}\bigl{\{}-(n-1)\Delta_{g}u-R_{g}u\bigr{\}}\bigm{|}_{\partial M}=0, (5)
νg(u)|M=0.\nu_{g}(u)\bigm{|}_{\partial M}=0. (6)

Multiply the contents in { } of (5) by the left hand side of (4) and integration it over M.M. Thus, by integration by parts, (4) and (5),

(n1)ΔguRgu=const.-(n-1)\Delta_{g}u-R_{g}u=\mathrm{const}.

On the other hand,

M(n1)ΔguRgudvg=M(n1)νg(u)dsg|MRgMudvg.\mathop{\text{\LARGE$\int_{\text{\normalsize$\scriptstyle M$}}$}}\nolimits-(n-1)\Delta_{g}u-R_{g}u~{}dv_{g}=\mathop{\text{\LARGE$\int_{\text{\normalsize$\scriptstyle\partial M$}}$}}\nolimits(n-1)\nu_{g}(u)~{}ds_{g|_{\partial M}}-R_{g}\mathop{\text{\LARGE$\int_{\text{\normalsize$\scriptstyle M$}}$}}\nolimits u~{}dv_{g}.

Hence, from (6) and uHgr(M)u\in H^{r}_{g}(M), the first term of the right hand side is zero. Thus, since uHgru\in H^{r}_{g} , the second term also vanish. Therefore, we have

(n1)ΔguRgu=0.-(n-1)\Delta_{g}u-R_{g}u=0. (7)

Hence, if Rg0R_{g}\neq 0, then u=0u=0.

On the other hand, if Rg=0R_{g}=0, we multiply the both sides of (7) by uu, integral over M,M, use the integration by parts and get u=const.u=\mathrm{const}. But, since uHgr(M)u\in H^{r}_{g}(M), u0.u\equiv 0. Therefore Φg(g)r\Phi^{r}_{g~{}(g)} is injective.

( \supset ) This inclusion is obvious. ∎

Lemma 3.5.

𝔖ˇC01.\check{\mathfrak{S}}_{C^{1}_{0}}\neq\emptyset.

Proof.

Since dimM3,\mathrm{dim}M\geq 3, we can construct a negative constant scalar curvature metric gg (see [8]). Thus, since the eigenvalues of Δg-\Delta_{g} are positive([4, Theorem 4.4]), from Lemma 3.4, g𝔖ˇC01.g\in\check{\mathfrak{S}}_{C^{1}_{0}}.

Lemma 3.6.

𝔖C0(1)r𝒦Cr\mathfrak{S}^{r}_{C^{(1)}_{0}}\cap\mathcal{K}^{r}_{C} is an ILH-submanifold of C,1r.\mathscr{M}^{r}_{C,1}.

Proof.

For each g𝔖C0r𝒦Crg\in\mathfrak{S}^{r}_{C_{0}}\cap\mathcal{K}^{r}_{C} , we define a map

Ψ:C,1r(1,1)g¯Hr32(M)\Psi:~{}\mathscr{M}^{r}_{C,1}\longrightarrow\langle(1,-1)\rangle^{\perp_{\bar{g}}}\oplus H^{r-\frac{3}{2}}(\partial M)

by

g(ΔgRg+MΔgRg𝑑vg0+Mνg(Rg)𝑑sg0|M,νg(Rg)|M,2n1Hg|M).g\mapsto\Bigl{(}-\Delta_{g}R_{g}+\int_{M}\Delta_{g}R_{g}~{}dv_{g_{0}}+\int_{\partial M}\nu_{g}(R_{g})~{}ds_{g_{0}\bigm{|}_{\partial M}},\nu_{g}(R_{g})|_{\partial M},\frac{2}{n-1}H_{g}|_{\partial M}\Bigr{)}.

From Lemma 3.2, this is a CC^{\infty}-differentiable map.

We note that Ψ1(0)=𝔖C0r\Psi^{-1}(0)=\mathfrak{S}^{r}_{C_{0}} . In fact, if g𝔖C0rg\in\mathfrak{S}^{r}_{C_{0}}, then Hg=0H_{g}=0 along M\partial M since gC0rg\in\mathscr{M}^{r}_{C_{0}}. And it is clear that the first two terms are zero if g𝔖C0rg\in\mathfrak{S}^{r}_{C_{0}}, hence the inclusion of Ψ1(0)𝔖C0r\Psi^{-1}(0)\supset\mathfrak{S}^{r}_{C_{0}} holds.

On the other hand, for gΨ1(0)g\in\Psi^{-1}(0), then, since the first and second components are both zero,

ΔgRg=const.\Delta_{g}R_{g}=\mathrm{const}~{}.

But, since the second component is zero and from the Green’s formula, this constant must be zero. Thus, by multiplying ΔgRg\Delta_{g}R_{g} by RgR_{g} and integrating it over M,M, from integration by parts and the fact that the second component is zero,

0=M|gRg|g2dvg.0=\int_{M}\bigl{|}\nabla_{g}R_{g}\bigr{|}^{2}_{g}~{}dv_{g}.

Hence Rg=const.R_{g}=\mathrm{const}. Since the third component is zero, we obtain gC0r.g\in\mathscr{M}^{r}_{C_{0}}.

The derivative of Ψ\Psi at g𝔖C0r𝒦Crg\in\mathfrak{S}^{r}_{C_{0}}\cap\mathcal{K}^{r}_{C} is calculated as follows(see [5, Theorem 1.174]):

DgΨ(h)\displaystyle D_{g}\Psi(h)
=(Δg(Δgtrgh+δgδghh,Ricgg)MΔg(Δgtrghδgδgh+h,Ricgg)dvg0\displaystyle=\Biggl{(}-\Delta_{g}\bigl{(}-\Delta_{g}tr_{g}h+\delta_{g}\delta_{g}h-\langle h,{\rm Ric_{g}}\rangle_{g}\bigr{)}-\int_{M}\Delta_{g}\bigl{(}\Delta_{g}tr_{g}h-\delta_{g}\delta_{g}h+\langle h,{\rm Ric_{g}}\rangle_{g}\bigr{)}dv_{g_{0}}
Mνg(Δgtrghδgδgh+h,Ricgg)𝑑sg0|M,νg(Δgtrgh+δgδghh,Ricgg)|M,\displaystyle-\int_{\partial M}\nu_{g}\bigl{(}\Delta_{g}tr_{g}h-\delta_{g}\delta_{g}h+\langle h,{\rm Ric_{g}}\rangle_{g}\bigr{)}ds_{g_{0}|_{\partial M}},\nu_{g}\bigl{(}-\Delta_{g}tr_{g}h+\delta_{g}\delta_{g}h-\langle h,{\rm Ric_{g}}\rangle_{g}\bigr{)}|_{\partial M},
1n1([d(trgh)δgh](νg)δg|M(h(,νg))gM(Πg,h))|M),\displaystyle\frac{1}{n-1}\bigl{(}[d(tr_{g}h)-\delta_{g}h](\nu_{g})-\delta_{g|_{\partial M}}(h(\bullet,\nu_{g}))-g_{\partial M}(\Pi_{g},h)\bigr{)}|_{\partial M}\Biggr{)},

where Πg\Pi_{g} denotes the second fundamental form of (M,g).(\partial M,g). Take the variation h=fg(fHgr(M)).h=fg~{}(f\in H^{r}_{g}(M)). Then, since g𝒦Cr,g\in\mathcal{K}^{r}_{C}, we get DgΨD_{g}\Psi is surjective. Therefore, from the Inverse function theorem ([22]), 𝔖C0r𝒦Cr\mathfrak{S}^{r}_{C_{0}}\cap\mathcal{K}^{r}_{C} is a submanifold of C,1r\mathscr{M}^{r}_{C,1} and the tangent space at g𝔖C0r𝒦Crg\in\mathfrak{S}^{r}_{C_{0}}\cap\mathcal{K}^{r}_{C} is KerDgΨ\mathrm{Ker}D_{g}\Psi.

On the other hand, since 𝔖C01r𝒦Cr\mathfrak{S}^{r}_{C^{1}_{0}}\cap\mathcal{K}^{r}_{C} is an ILH submanifold of 𝔖C0r𝒦Cr\mathfrak{S}^{r}_{C_{0}}\cap\mathcal{K}^{r}_{C}, 𝔖C01r𝒦Cr\mathfrak{S}^{r}_{C^{1}_{0}}\cap\mathcal{K}^{r}_{C} is an ILH submanifold of C,1r.\mathscr{M}^{r}_{C,1}.

Lemma 3.7.

Let C+r(M)N:={fHr(M)|f>0onM,νg0(f)=0onM}C^{r}_{+}(M)_{N}:=\bigl{\{}f\in H^{r}(M)\bigm{|}f>0~{}\mathrm{on}~{}M,~{}\nu_{g_{0}}(f)=0~{}\mathrm{on}~{}\partial M\bigr{\}} and be a map

χr:C+r(M)N×(𝔖C01r𝒦Cr)C01r;(f,g)fg.\chi^{r}:~{}C^{r}_{+}(M)_{N}\times\bigl{(}\mathfrak{S}^{r}_{C^{1}_{0}}\cap\mathcal{K}^{r}_{C}\bigr{)}\longrightarrow\mathscr{M}^{r}_{C^{1}_{0}}~{};~{}(f,g)\mapsto f\cdot g.

Then χr\chi^{r} is CC^{\infty}-differentiable. Moreover, if g𝔖ˇC01,g\in\check{\mathfrak{S}}_{C^{1}_{0}}, then D(f,g)χrD_{(f,g)}\chi^{r} is an isomorphism.

Proof.

It is clear that this is a CC^{\infty}-differentiable map. In fact,

D(f,g)χr(ϕ,h)=fh+ϕg.D_{(f,g)}\chi^{r}(\phi,h)=fh+\phi g.

injectivity ;

Let fh+ϕg=0fh+\phi g=0 ,then since KerDgΨh=ϕfg=:ϕ~g,\mathrm{Ker}D_{g}\Psi\in h=-\frac{\phi}{f}g=:\tilde{\phi}g, ϕ~KerΦgr.\tilde{\phi}\in\mathrm{Ker}\Phi^{r}_{g}. On the other hand, since g𝒦Cr,g\in\mathcal{K}^{r}_{C}, ϕ~=0.\tilde{\phi}=0. Hence, since f0,f\neq 0, ϕ=0,h=0.\phi=0~{},~{}h=0.

surjectivity ;

We shall show it by contradiction.

If D(f,g)χrD_{(f,g)}\chi^{r} is not surjective, then h¯(ImD(f,g)χr)g{0}\exists\bar{h}\in\bigl{(}\mathrm{Im}D_{(f,g)}\chi^{r}\bigr{)}^{\perp_{g}}\setminus\{0\}

(since

ImD(f,g)χr=fTg(𝔖C01r𝒦Cr)+Hr(M)g\mathrm{Im}D_{(f,g)}\chi^{r}=fT_{g}\bigl{(}\mathfrak{S}^{r}_{C^{1}_{0}}\cap\mathcal{K}^{r}_{C}\bigr{)}+H^{r}(M)g

is a closed subspace in TfgC01rT_{fg}\mathscr{M}^{r}_{C^{1}_{0}} ).

We define an operator on (Hr(M)g)g\bigl{(}H^{r}(M)g\bigr{)}^{\perp_{g}} (which is a closed subspace in TfgC01rT_{fg}\mathscr{M}^{r}_{C^{1}_{0}})

Kg(h):=Δgtrgh+δgδghh,Ricgg.K_{g}(h):=-\Delta_{g}tr_{g}h+\delta_{g}\delta_{g}h-\langle h,{\rm Ric_{g}}\rangle_{g}.

From the Green’s formula,

(Kgh,f)M(h,Kgf)M=(νg(f),trgh)M(f,νg(trgh))M+(h(νg,),gf)M+(δgh(νg),f)M,\begin{split}\bigl{(}K_{g}h,f\bigr{)}_{M}-\bigl{(}h,K^{*}_{g}f\bigr{)}_{M}=\bigl{(}&\nu_{g}(f),tr_{g}h\bigr{)}_{\partial M}-\bigl{(}f,\nu_{g}(tr_{g}h)\bigr{)}_{\partial M}\\ &\quad+\bigl{(}-h(\nu_{g},\bullet),\nabla_{g}f\bigr{)}_{\partial M}+\bigl{(}-\delta_{g}h(\nu_{g}),f\bigr{)}_{\partial M},\end{split}

where KgK^{*}_{g} is the formal adjoint of Kg:K_{g}:

Kgf:=(Δgf)g+ggffRicg.K^{*}_{g}f:=-\bigl{(}\Delta_{g}f\bigr{)}g+\nabla_{g}\nabla_{g}f-f{\rm Ric_{g}}.

Since KgK_{g} is defined on (Hr(M)g)g\bigl{(}H^{r}(M)g\bigr{)}^{\perp_{g}} and hTfgC01r,trgh0h\in T_{fg}\mathscr{M}^{r}_{C^{1}_{0}},~{}tr_{g}h\equiv 0 on M.\partial M. Therefore h=trghng0=0h=\frac{tr_{g}h}{n}g_{0}=0 on M.\partial M. Hence the first three terms in the right hand side of the above equation vanish on M\partial M. Also, since hTfgC01r,DfgH(h)=0,h\in T_{fg}\mathscr{M}^{r}_{C^{1}_{0}},D_{fg}H(h)=0, where DfgHD_{fg}H denotes the derivative of the mean curvature HH at fg.fg. And

DfgH(h)=12([d(trfgh)δfgh](νfg)δ(fg)|M(h(,νfg))(fg)|M(Πfg,h)),D_{fg}H(h)=\frac{1}{2}\bigl{(}[d(tr_{fg}h)-\delta_{fg}h](\nu_{fg})-\delta_{(fg)|_{\partial M}}(h(\bullet,\nu_{fg}))-(fg)|_{\partial M}(\Pi_{fg},h)\bigr{)},

(cf. [3, Claim 3.1][8, Section 2]) where Πfg\Pi_{fg} denotes the second fundamental form on (M,fg)(\partial M,fg). Since hTfgC01rh\in T_{fg}\mathscr{M}^{r}_{C^{1}_{0}} and trgh=0tr_{g}h=0, we obtain δfgh(νfg)=0.\delta_{fg}h(\nu_{fg})=0. Take a point pM.p\in\partial M. Since h|M=(n1trfgh)g|M=0h|_{\partial M}=(n^{-1}tr_{fg}h)g|_{\partial M}=0, δgh(νg)=dhi0(xi)\delta_{g}h(\nu_{g})=dh_{i0}(x^{i}) at p,p, where xix^{i} and hi0h_{i0} denote respectively a local normal coordinates at pp with respect to gg such that νg(p)=x0(p)\nu_{g}(p)=\frac{\partial}{\partial x^{0}}(p) and the (i,0)(i,0)-th components of hh with respect to (xi).(x^{i}). Let (yj)(y^{j}) be a local normal coordinates at pp with respect to fgfg such that νfg(p)=y0(p)\nu_{fg}(p)=\frac{\partial}{\partial y^{0}}(p) and h~ij\tilde{h}_{ij} the components of hh with respect to (yj).(y^{j}). Then, dhi0(xi)(p)=f1/2(fh~i0)yi(p)=(f3/2h~i0yi+νgfh~00+α=1nfxαh~α0)(p)=f3/2δfgh(νfg)(p)=0.dh_{i0}(x^{i})(p)=f^{1/2}\frac{\partial(f\tilde{h}_{i0})}{\partial y^{i}}(p)=\bigl{(}f^{3/2}\frac{\partial\tilde{h}_{i0}}{\partial y^{i}}+\nu_{g}f\tilde{h}_{00}+\sum^{n}_{\alpha=1}\frac{\partial f}{\partial x^{\alpha}}\tilde{h}_{\alpha 0}\bigr{)}(p)=f^{3/2}\delta_{fg}h(\nu_{fg})(p)=0. Hence δgh(νg)=0onM\delta_{g}h(\nu_{g})=0~{}\mathrm{on}~{}\partial M and we get

(Kgh,f)M=(h,Kgf),h(Hr(M)Ng)g,fHr2(M).\bigl{(}K_{g}h,f\bigr{)}_{M}=\bigl{(}h,K^{*}_{g}f\bigr{)}~{},~{}\forall h\in\bigl{(}H^{r}(M)_{N}g\bigr{)}^{\perp_{g}},\forall f\in H^{r-2}(M).

Thus, from the proof of Lemma 2.6 (and the fact that the principal symbol of KgK_{g} is surjective), we can get an orthogonal decomposition(cf. the proof of Lemma 2.4 in [18]):

TfgCr=Hr(M)gKerKgImKg.T_{fg}\mathscr{M}^{r}_{C}=H^{r}(M)g\oplus\mathrm{Ker}K_{g}\oplus\mathrm{Im}K^{*}_{g}.

From the hypothesis fh¯(Tg(𝔖C0rKgr))g,f\bar{h}\in\Bigl{(}T_{g}\bigl{(}\mathfrak{S}^{r}_{C_{0}}\cap K^{r}_{g}\bigr{)}\Bigr{)}^{\perp_{g}}, fh¯(Hr(M)g)g,f\bar{h}\in\bigl{(}H^{r}(M)g\bigr{)}^{\perp_{g}}, and from the above decomposition, fh¯ImKgf\bar{h}\in\mathrm{Im}K^{*}_{g} (since, if fgKerKgfg\in\mathrm{Ker}K_{g}, then it must be fgKerΨgr=Tg(𝔖C0r𝒦Cr)fg\in\mathrm{Ker}\Psi^{r}_{g}=T_{g}(\mathfrak{S}^{r}_{C_{0}}\cap\mathcal{K}^{r}_{C})). Let fh¯=Kg(ψ)f\bar{h}=K^{*}_{g}(\psi) , then

fh¯=(Δgψ)g+ggψψRicg.f\bar{h}=-\bigl{(}\Delta_{g}\psi\bigr{)}g+\nabla_{g}\nabla_{g}\psi-\psi{\rm Ric_{g}}.

But, since fh¯(Hr(M)g)gf\bar{h}\in\bigl{(}H^{r}(M)g\bigr{)}^{\perp_{g}},

0=trg(fh¯)=(n1)ΔgψRgψ.0=tr_{g}(f\bar{h})=-(n-1)\Delta_{g}\psi-R_{g}\psi.

Thus we can see that the image of KgK_{g} is included in Hr2(M)NH^{r-2}(M)_{N}. Then, from the above equation and g𝔖ˇC01g\in\check{\mathfrak{S}}_{C^{1}_{0}}, we obtain ψ0\psi\equiv 0. Hence fh¯0f\bar{h}\equiv 0. This contradicts that fh¯0f\bar{h}\neq 0 . Therefore D(f,g)χrD_{(f,g)}\chi^{r} is surjective.

We will show that the image of KgK_{g} is included in Hr2(M)NH^{r-2}(M)_{N} in the following. As in the proof of lemma 2.6, there is a decomposition of TfgC01rT_{fg}\mathscr{M}^{r}_{C^{1}_{0}}

TfgC0r=Im(α|Tid.DiffC0r+1)Ker(α|TgC0r),T_{fg}\mathscr{M}^{r}_{C_{0}}=\mathrm{Im}\bigl{(}\alpha|_{T_{id.}\mathrm{Diff}^{r+1}_{C_{0}}}\bigr{)}\oplus\mathrm{Ker}\bigl{(}\alpha^{*}|_{T_{g}\mathscr{M}^{r}_{C_{0}}}\bigr{)},

where α\alpha is the Lie derivative of gg and α\alpha^{*} is the divergence operator with respect to gg. Therefore, we can write that h=h1+h2,h1Im(α|Tid.DiffC0r+1),h2Ker(α|TgC01r)h=h_{1}+h_{2},~{}h_{1}\in\mathrm{Im}\bigl{(}\alpha|_{T_{id.}\mathrm{Diff}^{r+1}_{C_{0}}}\bigr{)},~{}h_{2}\in\mathrm{Ker}\bigl{(}\alpha^{*}|_{T_{g}\mathscr{M}^{r}_{C^{1}_{0}}}\bigr{)}.

We firstly consider h1h_{1}. Let h1=Xg,XHr+1(TM)h_{1}=\mathcal{L}_{X}g,~{}X\in H^{r+1}(TM). Since KgK_{g} is the first derivative of the functional gRgg\mapsto R_{g}, RgR_{g} is diffeomorphism invariant and the derivative of the pull-back action of diffeomorphism on gg is g\mathcal{L}_{\bullet}g (as mentioning in the proof of lemma 2.1),

Kgh1=0onM.K_{g}h_{1}=0~{}\mathrm{on}~{}M.

Hence νg0(Kgh1)=0onM\nu_{g_{0}}(K_{g}h_{1})=0~{}\mathrm{on}~{}\partial M.

Finally, we consider h2Kerδgh_{2}\in\mathrm{Ker}~{}\delta_{g}. Then δgδgh2=0onM\delta_{g}\delta_{g}h_{2}=0~{}\mathrm{on}~{}M. Since h2TfgC0rh_{2}\in T_{fg}\mathscr{M}^{r}_{C_{0}}, we can write jM1h2=jM1(ρg0)j^{1}_{\partial M}h_{2}=j^{1}_{\partial M}(\rho\cdot g_{0}) for some ρHr1/2(C(M))\rho\in H^{r-1/2}(C^{\infty}(M)). Note that νg0ρ=0\nu_{g_{0}}\rho=0 on M.\partial M. In fact, since νg0g00,\nabla_{\nu_{g_{0}}}g_{0}\equiv 0, (νg0ρ)g0=νg0h2(\nabla_{\nu_{g_{0}}}\rho)g_{0}=\nabla_{\nu_{g_{0}}}h_{2} on M\partial M. Hence, νg0ρ=(1/n)g0ijνg0(h2)ij\nabla_{\nu_{g_{0}}}\rho=(1/n)g_{0}^{ij}\nabla_{\nu_{g_{0}}}(h_{2})_{ij} on M,\partial M, where g0ijg_{0}^{ij} and (h2)ij(h_{2})_{ij} denote respectively the components of g1g^{-1} and hh in terms of some local coordinates. On the other hand, g0ijνg0(h2)ij=νg0(g0ij)(h2)ij=0g_{0}^{ij}\nabla_{\nu_{g_{0}}}(h_{2})_{ij}=-\nabla_{\nu_{g_{0}}}(g_{0}^{ij})(h_{2})_{ij}=0 on M\partial M since trg0h2=0tr_{g_{0}}h_{2}=0 on MM and h2|M(=1nρ|M)=0h_{2}|_{\partial M}~{}(=\frac{1}{n}\rho|_{\partial M})=0. Therefore, on M\partial M,

νg0(h2,Ricgg)=νg0(ρg0),Ricgg+ρg0,νg0RicggonM,\nabla_{\nu_{g_{0}}}\bigl{(}\langle h_{2},{\rm Ric_{g}}\rangle_{g}\bigr{)}=\langle\nabla_{\nu_{g_{0}}}(\rho g_{0}),{\rm Ric_{g}}\rangle_{g}+\langle\rho g_{0},\nabla_{\nu_{g_{0}}}{\rm Ric_{g}}\rangle_{g}~{}\mathrm{on}~{}\partial M,

where ,g\langle,\rangle_{g} denote the natural inner product on the (0,2)-tensor bundle of MM induced by g.g. Here, as mentioning above, ρ=0onM\rho=0~{}\mathrm{on}~{}\partial M. And νg0(ρg0)=(νg0(ρ))g0+ρνg0g0=0onM\nabla_{\nu_{g_{0}}}(\rho g_{0})=(\nu_{g_{0}}(\rho))g_{0}+\rho\nabla_{\nu_{g_{0}}}g_{0}=0~{}\mathrm{on}~{}\partial M since νg0(ρ)=0onM\nu_{g_{0}}(\rho)=0~{}\mathrm{on}~{}\partial M and νg0g00onM\nabla_{\nu_{g_{0}}}g_{0}\equiv 0~{}\mathrm{on}~{}M. Consequently, we obtain νg0(Kgh2)=0onM.\nu_{g_{0}}(K_{g}h_{2})=0~{}\mathrm{on}~{}\partial M.

We get the following lemma in the same way as Lemma 2.8 in [17]:

Lemma 3.8.

Let EE and FF be bector bundles over MM associated with the frame bundle. Any ηDiffC0\eta\in\mathrm{Diff}_{C_{0}} defines a natural linear map (by pullback)

η:Hk(E)Hk(E)(kn/2+2).\eta^{*}:~{}H^{k}(E)\longrightarrow H^{k}(E)~{}~{}(k\geq n/2+2).

Let rn/2+2,AHr(E)r\geq n/2+2,~{}A\subset H^{r}(E) be an open subset and let ϕ:AHr(F)\phi:~{}A\rightarrow H^{r}(F) be a CC^{\infty} -differentiable map which commutes with the action of DiffC0\mathrm{Diff}_{C_{0}} . Put As:=AHs(E)(sr)A^{s}:=A\cap H^{s}(E)~{}(s\geq r) . Then ϕ(As)Hs(F)\phi\bigl{(}A^{s}\bigr{)}\subset H^{s}(F) and ϕ|As:AsHs(F)\phi|_{A^{s}}:~{}A^{s}\rightarrow H^{s}(F) is CC^{\infty} -differentiable.

Theorem 3.1.

𝔖ˇC01\check{\mathfrak{S}}_{C^{1}_{0}} is an ILH-submanifold of C01\mathscr{M}_{C^{1}_{0}} and the map

χ:C+(M)N×𝔖ˇC01C01;(f,g)fg\chi:~{}C^{\infty}_{+}(M)_{N}\times\check{\mathfrak{S}}_{C^{1}_{0}}\longrightarrow\mathscr{M}_{C^{1}_{0}}~{};~{}(f,g)\mapsto f\cdot g

is a local ILH-diffeomorphism into C01\mathscr{M}_{C^{1}_{0}}.

Proof.

It can be proved exactly the same as the proof of Theorem 2.5 in [18] using lemma 3.7 and 3.8. And note that rC+r(M)N=C+(M)N\bigcap_{r}C^{r}_{+}(M)_{N}=C^{\infty}_{+}(M)_{N}. ∎

Since DiffC0\mathrm{Diff}_{C_{0}} and C0(1)\mathscr{M}_{C^{(1)}_{0}} are submanifold of Diff(M)\mathrm{Diff}(M) and \mathscr{M} respectively, from Lemma 3.8 and Theorem 2.1, we can obtain the following CC^{\infty}-version of the Slice theorem exactly same as in [17]:

Theorem 3.2 (CC^{\infty}-version of Theorem 2.1).

For all gC0(1)g\in\mathscr{M}_{C^{(1)}_{0}} there exists an ILH-submanifold 𝒮gC0(1)\mathcal{S}_{g}\subset\mathscr{M}_{C^{(1)}_{0}} containing γ\gamma so that the following holds:

(1) ηIgη𝒮g=𝒮g,\eta\in I_{g}\Rightarrow\eta^{*}\mathcal{S}_{g}=\mathcal{S}_{g},

(2) ηDiffC0,η(𝒮g)𝒮gηIg\eta\in\mathrm{Diff}_{C_{0}}~{},~{}\eta^{*}\bigl{(}\mathcal{S}_{g}\bigr{)}\cap\mathcal{S}_{g}\neq\emptyset\Rightarrow\eta\in I_{g} and

(3) There exists a local section defined on an open neighborhood of [Ig]:[I_{g}]:

χ:(DiffC0/Ig)UDiffC0\exists\chi~{}:~{}\bigl{(}\mathrm{Diff}_{C_{0}}/I_{g}\supset\bigr{)}U\longrightarrow\mathrm{Diff}_{C_{0}}

such that

F:U×𝒮gC0(1);(u,t)χ(u)tF~{}:~{}U\times\mathcal{S}_{g}\longrightarrow\mathscr{M}_{C^{(1)}_{0}}~{};~{}(u,t)\mapsto\chi(u)^{*}t

is an ILH-diffeomorphism mapping onto an open nighborhood of g.g.

Consequently, from this Slice theorem and Theorem 3.1, we can prove Main Theorem in Section 1.

Proof of Main Theorem.

From Theorem 3.1, we can decompose as

g(t)=f(t)g~(t),g(t)=f(t)\tilde{g}(t),

where f(t)f(t) is a deformation of ff in C+(M)NC^{\infty}_{+}(M)_{N} and g~(t)\tilde{g}(t) is a deformation of g¯\bar{g} in 𝔖ˇC01\check{\mathfrak{S}}_{C^{1}_{0}}. Moreover, from Theorem 3.2, g~(t)\tilde{g}(t) can be decomposed as

g~(t)=ϕ(t)g¯(t)withδg¯(0)=0.\tilde{g}(t)=\phi(t)^{*}\bar{g}(t)~{}\mathrm{with}~{}\delta\bar{g}^{{}^{\prime}}(0)=0.

Since the scalar curvature is invariant under the action of diffeomorphisms,

Rg~(t)=Rg¯(t)const.R_{\tilde{g}(t)}=R_{\bar{g}(t)}\equiv\mathrm{const}.

Theorem 3.3.

For any g=fg¯(fC+(M)N,g¯𝔖ˇC01)g=f\bar{g}~{}(f\in C^{\infty}_{+}(M)_{N},~{}\bar{g}\in\check{\mathfrak{S}}_{C^{1}_{0}}) and any smooth deformation {g(t)}t(ϵ,ϵ)(C01)\{g(t)\}_{t\in(-\epsilon,\epsilon)}(\subset\mathscr{M}_{C^{1}_{0}}) of gg for sufficiently small ϵ>0,\epsilon>0, there exists uniquely a smooth deformation {f(t)}t(ϵ,ϵ)(C+(M)N)\{f(t)\}_{t\in(-\epsilon,\epsilon)}(\subset C^{\infty}_{+}(M)_{N}) of f,f, a smooth one {ϕ(t)}t(ϵ,ϵ)(DiffC0)\{\phi(t)\}_{t\in(-\epsilon,\epsilon)}(\subset\mathrm{Diff}_{C_{0}}) of idMid_{M} and a smooth one {g¯(t)}t(ϵ,ϵ)(𝔖ˇC01)\{\bar{g}(t)\}_{t\in(-\epsilon,\epsilon)}(\subset\check{\mathfrak{S}}_{C^{1}_{0}}) of g¯\bar{g} with δ(f(0)g¯+fg¯(0))=0\delta\bigl{(}f^{{}^{\prime}}(0)\bar{g}+f\bar{g}^{{}^{\prime}}(0)\bigr{)}=0 such that

g(t)=(f(t)ϕ(t))ϕ(t)g¯(t).g(t)=\bigl{(}f(t)\circ\phi(t)\bigr{)}\phi(t)^{*}\bar{g}(t).
Proof.

Reverse the order of applying Theorem 3.1 and Theorem 3.2 in the proof of Main Theorem. ∎

4 Applications

We use the same notations as those in the above sections. We give two applications of the following.

4.1 Some rigidity theorems for relative constant scalar curvature metrics

In the case of M=,\partial M=\emptyset, a metric gg in a given conformal class CC is called a Yamabe metric if gg is a minimizer of the restriction |C\mathcal{E}|_{C} of the normalized Einstein-Hilbert functional .\mathcal{E}. The infimum of |C\mathcal{E}|_{C} is called the Yamabe constant Y(M,C)Y(M,C) of C.C. By combining the Koiso’s decomposition theorem with the existence of a Yamabe metric in each conformal class, Böhm-Wang-Ziller proved the following (see the proof of [6, Theorem 5.1]):

Theorem 4.1.

Let (Mn,g)(M^{n},g_{\infty}) be a closed Riemannian manifold of dimension n3.n\geq 3. Assume that gg_{\infty} is a unique constant scalar curvature (csc metric for brevity) in its conformal class up to rescaling with λ1(Δg)>Rgn1.\lambda_{1}(-\Delta_{g_{\infty}})>\frac{R_{g_{\infty}}}{n-1}. Here, λ1(Δg)\lambda_{1}(-\Delta_{g_{\infty}}) denotes the first non-zero eigenvalue of Δg.-\Delta_{g_{\infty}}. Then each csc metric sufficiently close to gg_{\infty} with respect to the CC^{\infty}-topology is a Yamabe metric in its conformal class.

Remark 4.1.

In the above, the condition λ1(Δg)>Rgn1\lambda_{1}(-\Delta_{g_{\infty}})>\frac{R_{g_{\infty}}}{n-1} implies that (Mn,[g])(M^{n},[g_{\infty}]) is not conformally equivalent to the standard nn-sphere (Sn,[gstd]).(S^{n},[g_{{\rm std}}]).

On the other hand, using a compactness theorem of the space of all csc metrics in a fixed conformal class (proved by Khuri-Maqrues-Schoen [14, Theorem 1.1]), one can also get the following:

Theorem 4.2.

Let (Mn,g)(M^{n},g_{\infty}) be a closed Riemannian manifold of dimension either 3n7,3\leq n\leq 7, or both 8n248\leq n\leq 24 and that MM is spin. Assume that gg_{\infty} is a unique csc metric in its conformal class up to rescaling with λ1(Δg)>Rgn1\lambda_{1}(-\Delta_{g_{\infty}})>\frac{R_{g_{\infty}}}{n-1}. Then each csc metric sufficiently close to gg_{\infty} with respect to the CC^{\infty}-topology is also a unique csc metric up to rescaling in its conformal class.

We can prove a similar statement below on a manifold with boundary. When M,\partial M\neq\emptyset, for gg\in\mathscr{M} with Hg=0H_{g}=0 along M,\partial M,

Y(M,[g]0):=infh[g]0(h)Y(M,[g]_{0}):=\inf_{h\in[g]_{0}}\mathcal{E}(h)

is called the relative Yamabe constant of [g]0.[g]_{0}. A metric hh with Hh=0H_{h}=0 along M\partial M is called a relative Yamabe metric if Y(M,[g]0)=(h)Y(M,[g]_{0})=\mathcal{E}(h) (see [3] for more details). Here, [g]0[g]_{0} denotes the relative conformal class of gg, that is, [g]0:={h[g]|Hh=0alongM}={ug|uC+(M),νg(u)|M=0}.[g]_{0}:=\bigl{\{}h\in[g]\bigm{|}H_{h}=0~{}\mathrm{along}~{}\partial M\bigr{\}}=\bigl{\{}u\cdot g\bigm{|}u\in C^{\infty}_{+}(M),~{}\nu_{g}(u)|_{\partial M}=0\bigr{\}}. Then we can prove the following:

Theorem 4.3.

Let (Mn,g)(M^{n},g_{\infty}) be a compact connected Riemannian manifold of dimension n3n\geq 3 with smooth non-empty minimal boundary M\partial M (i.e., Hg=0H_{g_{\infty}}=0 along M\partial M). Assume that gg_{\infty} is a unique relative csc metric in [g]0[g_{\infty}]_{0} up to rescaling with λ1(Δg;Neumann)>Rgn1,\lambda_{1}(-\Delta_{g_{\infty}};\mathrm{Neumann})>\frac{R_{g_{\infty}}}{n-1}, where λ1(Δg;Neumann)\lambda_{1}(-\Delta_{g_{\infty}};\mathrm{Neumann}) denotes the first non-zero eigenvalue of Δg-\Delta_{g_{\infty}} with the Neumann boundary condition (see [1, Proposition 2.6] and [24, Section 1]) Moreover, we assume the following: either

(a) (Mn,g)(M^{n},g_{\infty}) has a nonumbilic point on M,\partial M,

or

(b) (Mn,g)(M^{n},g_{\infty}) is umbilic boundary (therefore, M\partial M is totally geodesic) satisfying that one of the followings (b1)-(b3) holds:

(b1) the Weyl tensor does not vanish identically on M\partial M and n6,n\geq 6,

(b2) MM is locally conformally flat,

(b3) n=3,4,n=3,4, or 5.5.

Then each relative csc metric sufficiently close to gg_{\infty} in C01\mathscr{M}_{C^{1}_{0}} with respect to he CC^{\infty}-topology is a relative Yamabe metric.

Remark 4.2.

Escobar [11] proved that, for any (M,g)(M,g_{\infty}) satisfying either (a) or (b) in the above, then there exists a relative Yamabe metric in [g]0.[g_{\infty}]_{0}. Note also that, the condition λ1(Δg;Neumann)>Rgn1\lambda_{1}(-\Delta_{g_{\infty}};\mathrm{Neumann})>\frac{R_{g_{\infty}}}{n-1} implies that (M,[g]0)(M,[g_{\infty}]_{0}) is not conformally equivalent to the standard hemisphere (S+n,[gstd]0).(S^{n}_{+},[g_{{\rm std}}]_{0}).

The following follows directly from Main Theorem.

Corollary 4.1.

Assume that g𝔖C01g_{\infty}\in\mathfrak{S}_{C^{1}_{0}} satisfies λ1(Δg;Neumann)>Rgn1.\lambda_{1}(-\Delta_{g_{\infty}};\mathrm{Neumann})>\frac{R_{g_{\infty}}}{n-1}. Let {gi}\{g_{i}\} and {g~i:=ui4n2gi}𝔖C01\{\tilde{g}_{i}:=u_{i}^{\frac{4}{n-2}}g_{i}\}\subset\mathfrak{S}_{C^{1}_{0}} be sequences each of which converges to gg_{\infty} with respect to the CC^{\infty}-topology. Then, except for a finite number of ii, gi=g~i.g_{i}=\tilde{g}_{i}.

Proof.

Since CC^{\infty}-topology is stronger than the ILH-topology, this claim is directly derived from Main Theorem. ∎

Outline of the proof of Theorem 4.3.

We can assume that (M,g)(M,g_{\infty}) has unit volume. Let {gi}𝔖C01\{g_{i}\}\subset\mathfrak{S}_{C^{1}_{0}} be a sequence which converges to gg_{\infty} with respect to the CC^{\infty}-topology. For each i,i, let uiC+(M)Nu_{i}\in C^{\infty}_{+}(M)_{N} be a solution of relative Yamabe problem in [gi]0,[g_{i}]_{0}, that is, g~i:=ui4n2gi\tilde{g}_{i}:=u_{i}^{\frac{4}{n-2}}g_{i} is a relative Yamabe metric of [gi]0[g_{i}]_{0} with unit volume. Since g~i𝔖C01\tilde{g}_{i}\in\mathfrak{S}_{C^{1}_{0}} is a relative Yamabe metric, then the following hold:

uiL2nn2(gi)=1,||u_{i}||_{L^{\frac{2n}{n-2}}(g_{i})}=1, (8)
4n1n2Δgiui+Rgiui=Y(M,[gi]0)uin+2n2onM,-4\frac{n-1}{n-2}\Delta_{g_{i}}u_{i}+R_{g_{i}}u_{i}=Y(M,[g_{i}]_{0})u_{i}^{\frac{n+2}{n-2}}~{}\mathrm{on}~{}M, (9)
νgi(ui)=0onM.\nu_{g_{i}}(u_{i})=0~{}\mathrm{on}~{}\partial M. (10)

By [11], the assumptions in Theorem 4.3 implies that

Y(M,[g]0)<Y(S+n,[gstd]0).Y(M,[g_{\infty}]_{0})<Y(S^{n}_{+},[g_{\mathrm{std}}]_{0}).

Since gY(M,[g]0)g\mapsto Y(M,[g]_{0}) is continuous with respect to the C2C^{2}-topology,

Y(M,[gi]0)<Y(S+n,[gstd]0)Y(M,[g_{i}]_{0})<Y(S^{n}_{+},[g_{{\rm std}}]_{0})

for sufficiently large i.i. Hence, we can apply the similar argument in the proof of Theorem 5.1 in [6] to that on a manifold with boundary after slight modifications. Then, there exists a subsequaence {uik}{ui}\{u_{i_{k}}\}\subset\{u_{i}\} and uC+(M)N,R~u_{\infty}\in C^{\infty}_{+}(M)_{N},~{}\tilde{R}\in\mathbb{R} such that

g~i=uik4n2gikg~:=u4n2gask\tilde{g}_{i}=u_{i_{k}}^{\frac{4}{n-2}}g_{i_{k}}\rightarrow\tilde{g}_{\infty}:=u_{\infty}^{\frac{4}{n-2}}g_{\infty}~{}\mathrm{as}~{}k\rightarrow\infty

and

uL2nn2(g)=1,||u_{\infty}||_{L^{\frac{2n}{n-2}}(g_{\infty})}=1, (11)
4n1n2Δgu+Rgu=R~un+2n2onM,-4\frac{n-1}{n-2}\Delta_{g_{\infty}}u_{\infty}+R_{g_{\infty}}u_{\infty}=\tilde{R}u_{\infty}^{\frac{n+2}{n-2}}~{}\mathrm{on}~{}M, (12)
νg(u)=0onM.\nu_{g_{\infty}}(u_{\infty})=0~{}\mathrm{on}~{}\partial M. (13)

Here, the above convergence g~ig~(k)\tilde{g}_{i}\rightarrow\tilde{g}_{\infty}(k\rightarrow\infty) is the CC^{\infty}-convergence with respect to g.g_{\infty}. Then, from the regularity theorem([7]) and the maximum principle, uC+(M)Nu_{\infty}\in C^{\infty}_{+}(M)_{N}. From (11) and (12), we have g~𝔖C01\tilde{g}_{\infty}\in\mathfrak{S}_{C^{1}_{0}}. Hence, from the uniqueness assumption for gg_{\infty}, g~=g.\tilde{g}_{\infty}=g_{\infty}. Therefore g~ik=gik\tilde{g}_{i_{k}}=g_{i_{k}} from Corollary 4.1, except for finite number of k.k.

By using a compactness result proved by Discozi-Khuri [9, Theorem 1.1], we can also prove the following:

Theorem 4.4.

Let (Mn,g)(M^{n},g_{\infty}) be a compact connected Riemannian manifold with smooth non-empty totally geodesic boundary M.\partial M. Assume that either 3n7,3\leq n\leq 7, or both 8n248\leq n\leq 24 and that MM is spin. Let g𝔖C01g_{\infty}\in\mathfrak{S}_{C^{1}_{0}} be a unique relative csc metric in [g]0[g_{\infty}]_{0} up to rescaling with λ1(Δg;Neumann)>Rgn1\lambda_{1}(-\Delta_{g_{\infty}};\mathrm{Neumann})>\frac{R_{g_{\infty}}}{n-1}. Then each relative csc metric sufficiently close to gg_{\infty} in C01\mathscr{M}_{C^{1}_{0}} with respect to he CC^{\infty}-topology is also a relative unique csc metric up to rescaling in its relative conformal class.

Proof.

In the proof of Theorem 4.3, we will take g~i:=ui4n2gi\tilde{g}_{i}:=u_{i}^{\frac{4}{n-2}}g_{i} as another relative csc metric in [gi]0[g_{i}]_{0} with unit volume. Then, from the compactness result [9, Theorem 1.1], there exist a subsequence {uik}{ui}\{u_{i_{k}}\}\subset\{u_{i}\} and uC+(M)N,R~u_{\infty}\in C^{\infty}_{+}(M)_{N},~{}\tilde{R}\in\mathbb{R} such that

g~i=uik4n2gikg~:=u4n2gask\tilde{g}_{i}=u_{i_{k}}^{\frac{4}{n-2}}g_{i_{k}}\rightarrow\tilde{g}_{\infty}:=u_{\infty}^{\frac{4}{n-2}}g_{\infty}~{}\mathrm{as}~{}k\rightarrow\infty

and

uL2nn2(g)=1,||u_{\infty}||_{L^{\frac{2n}{n-2}}(g_{\infty})}=1,
4n1n2Δgu+Rgu=R~un+2n2onM,-4\frac{n-1}{n-2}\Delta_{g_{\infty}}u_{\infty}+R_{g_{\infty}}u_{\infty}=\tilde{R}u_{\infty}^{\frac{n+2}{n-2}}~{}\mathrm{on}~{}M,
νg(u)=0onM.\nu_{g_{\infty}}(u_{\infty})=0~{}\mathrm{on}~{}\partial M.

Then, the same argument as that in the proof of Theorem 4.3 implies that g~ik=gik\tilde{g}_{i_{k}}=g_{i_{k}} except for a finite number of kk. This completes the proof. ∎

4.2 A characterization of relative Einstein metrics

In the case of M=,\partial M=\emptyset, we recall the Yamabe invariant Y(M)Y(M) of MM (cf. [15], [24]) defined by

Y(M):=supC𝒞(M)Y(M,C)=supC𝒞(M)(infgCY(M,C)),Y(M):=\sup_{C\in\mathcal{C}(M)}Y(M,C)=\sup_{C\in\mathcal{C}(M)}\bigl{(}\inf_{g\in C}Y(M,C)\bigr{)},

where 𝒞(M)\mathcal{C}(M) denotes the set of all conformal classes on M.M. By the Koiso’s decomposition theorem, one can get the following:

Theorem 4.5 (cf. [4, Proposition 5.89]).

Let MnM^{n} be a closed manifold of dimension 3\geq 3 and gg a unique csc metric (up to rescaling) in its conformal class [g].[g]. Assume that Y(M)Y(M) is attained by gg and that λ1(Δg)>Rgn1.\lambda_{1}(-\Delta_{g})>\frac{R_{g}}{n-1}. Then, gg is an Einstein metric.

We can prove a similar statement below on a manifold with boundary. Let MnM^{n} be a compact connected Riemannian manifold of dimension n3n\geq 3 with non-empty smooth boundary M\partial M. For each conformal class CC on M,M, we define an invariant 𝒴(M;C)\mathcal{Y}(M;C):

𝒴(M;C):=supC¯𝒞Cinfg¯C¯0(g¯)=supgC01Y(M,[g]0),\mathcal{Y}(M;C):=\sup_{\bar{C}\in\mathcal{C}_{C}}\inf_{\bar{g}\in\bar{C}_{0}}\mathcal{E}(\bar{g})=\sup_{g\in\mathscr{M}_{C^{1}_{0}}}Y(M,[g]_{0}),

where 𝒞C:={C¯;conformalclassonM|C¯||M1=C||M1}\mathcal{C}_{C}:=\bigl{\{}\bar{C};~{}\mathrm{conformal~{}class~{}on}~{}M\bigm{|}\bar{C}||^{1}_{\partial M}=C||^{1}_{\partial M}\bigr{\}}. From the Aubin-type inequality, it holds that 𝒴(M;C)𝒴(S+n,[gstd])\mathcal{Y}(M;C)\leq\mathcal{Y}(S^{n}_{+},[g_{\mathrm{{\rm std}}}]) (see [1] for instance). Then, we can prove the following:

Theorem 4.6.

Fix a conformal class CC on M.M. Let g~\tilde{g} be a unique relative csc metric (up to rescaling) in [g~]0[\tilde{g}]_{0} with g~||M1C0||M1.\tilde{g}||^{1}_{\partial M}\in C_{0}||^{1}_{\partial M}. Assume that 𝒴(M;C)\mathcal{Y}(M;C) is attained by g~\tilde{g} and that λ1(Δg~;Neumann)>Rg~n1.\lambda_{1}(-\Delta_{\tilde{g}};\mathrm{Neumann})>\frac{R_{\tilde{g}}}{n-1}. Moreover, assume that 𝒴(M;C)<Y(S+n,[gstd]0).\mathcal{Y}(M;C)<Y(S^{n}_{+},[g_{{\rm std}}]_{0}). Then, g~\tilde{g} is a relative Einstein metric.

Proof.

From Main Theorem, every critical point of |𝔖C01\mathcal{E}|_{\mathfrak{S}_{C^{1}_{0}}} is Einstein (cf. [5, Proposition 4.47]). Hence, it is enough to prove that ddt(g(t))|t=0=0\frac{d}{dt}\mathcal{E}(g(t))|_{t=0}=0 for any smooth deformation g(t)g(t) of g~\tilde{g} in 𝔖C01\mathfrak{S}_{C^{1}_{0}}.

We now remark that, under the condition Y(M,[g]0)<Y(S+n,[gstd]0),Y(M,[g]_{0})<Y(S^{n}_{+},[g_{{\rm std}}]_{0}), there exists always a relative Yamabe metric in [g]0[g]_{0} ([7]). Hence, by Theorem 4.3, Y(M,[g(t)]0)=(g(t))Y(M,[g(t)]_{0})=\mathcal{E}(g(t)) for sufficiently small |t|<<1.|t|<<1. On the other hand, since Y(M,[g(t)])𝒴(M;C)Y(M,[g(t)])\leq\mathcal{Y}(M;C) and Y(M,[g(0)]0)=𝒴(M;C)Y(M,[g(0)]_{0})=\mathcal{Y}(M;C), we have

0=ddtY(M,[g(t)])|t=0=ddt(g(t))|t=0.0=\frac{d}{dt}Y(M,[g(t)])|_{t=0}=\frac{d}{dt}\mathcal{E}(g(t))|_{t=0}.

Therefore, g~\tilde{g} is a relative Einstein. ∎

Remark 4.3.

From the characterization Theorem 4.6 for relative Einstein metrics, we would like to suggest that the relative Yamabe invariant

Y(M,M,C|M)Y(M,\partial M,C|_{\partial M}) defined in [3, Section 1]:

Y(M,M,C|M):=supgC0|Y(M,[g]0)Y(M,\partial M,C|_{\partial M}):=\sup_{g\in\mathscr{M}_{C_{0}|_{\partial}}}Y(M,[g]_{0})

should be replaced by the above 𝒴(M;C)=supgC01Y(M,[g]0).\mathcal{Y}(M;C)=\sup_{g\in\mathscr{M}_{C^{1}_{0}}}Y(M,[g]_{0}).

References

  • [1] Akutagawa, K.: Notes on the relative Yamabe invariant. Differential Geometry (ed by Q.-M. Cheng, Josai Univ) 3, 105–113 (2001)
  • [2] Akutagawa, K.: An Obata type theorem on compact Einstein manifolds with boundary, preprint (2020)
  • [3] Akutagawa, K., Botvinnik, B.: The relative Yamabe invariant. Comm. Anal. Geom 10, 925–954 (2002)
  • [4] Aubin, T.: Some nonlinear problems in Riemannian geometry. Springer Science & Business Media (2013)
  • [5] Besse, A.: Einstein manifolds. Spring-Verlag, Berlin (1987)
  • [6] Böhm, C., Wang, M., Ziller, W.: A variational approach for compact homogeneous Einstein manifolds. Geometric & Functional Analysis GAFA 14(4), 681–733 (2004)
  • [7] Cherrier, P.: Problemes de Neumann non linéaires sur les variétés Riemanniennes. Journal of Functional Analysis 57(2), 154–206 (1984)
  • [8] Cruz, T., Vitório, F.: Prescribing the curvature of Riemannian manifolds with boundary. Calculus of Variations and Partial Differential Equations 58(4), 124 (2019)
  • [9] Disconzi, M.M., Khuri, M.A.: Compactness and non-compactness for the Yamabe problem on manifolds with boundary. Journal für die reine und angewandte Mathematik 2017(724), 145–201 (2017)
  • [10] Ebin, D.G.: The manifold of Riemannian metrics, in: Global Analysis, Berkeley, Calif., 1968. In: Proc. Sympos. Pure Math., vol. 15, pp. 11–40 (1970)
  • [11] Escobar, J.F., et al.: The Yamabe problem on manifolds with boundary. Journal of Differential Geometry 35(1), 21–84 (1992)
  • [12] Gilbarg, D., Trudinger, N.: Elliptic partial differential equations of second order, reprint of the 2nd ed. Berlin Heidelberg New York 1983. corr. 3rd printing 1998 ed (2013)
  • [13] Hörmander, L.: The analysis of linear partial differential operators. iii, volume 274 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences] (1985)
  • [14] Khuri, M.A., Marques, F.C., Schoen, R.M., et al.: A compactness theorem for the Yamabe problem. Journal of Differential Geometry 81(1), 143–196 (2009)
  • [15] Kobayashi, O.: Scalar curvature of a metric with unit volume. Mathematische Annalen 279(2), 253–265 (1987)
  • [16] Kobayashi, S., Nomizu, K.: Foundations of differential geometry, vol. 1. New York, London (1963)
  • [17] Koiso, N.: Nondeformability of Einstein metrics. Osaka Journal of Mathematics 15(2), 419–433 (1978)
  • [18] Koiso, N.: A decomposition of the space \mathscr{M} of Riemannian metrics on a manifold. Osaka Journal of Mathematics 16(2), 423–429 (1979)
  • [19] Lang, S.: Introduction to differentiable manifolds, Second edition. Springer (2002)
  • [20] Myers, S.B., Steenrod, N.E.: The group of isometries of a Riemannian manifold. Annals of Mathematics 40, 400–416 (1939)
  • [21] Omori, H.: On the group of diffeomorphisms on a compact manifold. In: Proc. Symp. Pure Appl. Math., XV, Amer. Math. Soc, pp. 167–183 (1970)
  • [22] Omori, H.: Infinite-dimensional Lie groups, vol. 158. American Mathematical Soc. (2017)
  • [23] Palais, R.S.: Foundations of global non-linear analysis. Benjamin (1968)
  • [24] Schoen, R.M.: Variational theory for the total scalar curvature functional for Riemannian metrics and related topics. In: Topics in calculus of variations, pp. 120–154. Springer (1989)

E-mail adress[email protected]

Department Of Mathematics, Chuo University, Tokyo 112-8551, Japan