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

Rigidity of spherical product Ricci solitons

Ao Sun Department of Mathematics, University of Chicago, 5734 S. University Avenue, Chicago, IL 60637, USA [email protected]  and  Jonathan J. Zhu Mathematical Sciences Institute, Australian National University, Hanna Neumann Building, Science Road, Canberra, ACT 2601, Australia [email protected]
Abstract.

We show that S2×S2S^{2}\times S^{2} is isolated as a shrinking Ricci soliton in the space of metrics, up to scaling and diffeomorphism. We also prove the same rigidity for S2×NS^{2}\times N, where NN belongs to a certain class of closed Einstein manifolds. These results are the Ricci flow analogues of our results for Clifford-type shrinking solitons for the mean curvature flow.

1. Introduction

Ricci solitons are metrics that flow by rescaling (up to diffeomorphism) under the Ricci flow. They were first introduced by Hamilton [Ham82] to study the singular behaviour of the Ricci flow. Shrinking solitons are those that contract under Ricci flow, and relate to the formation of singularities. A gradient shrinking Ricci soliton is a Riemannian manifold (M,g)(M,g) that satisfies the equation

(1.1) Ric(g)+Hessgf=12τg\operatorname{Ric}(g)+\operatorname{Hess}^{g}f=\frac{1}{2\tau}g

for some smooth function ff on MM and some τ>0\tau>0. Note that equation (1.1) has two natural symmetries: diffeomorphisms of MM, and rescaling (which results in a rescaled τ\tau). In the compact setting, every closed shrinking Ricci soliton is a gradient shrinking soliton. Shrinking solitons often arise as Type I singularity models of Ricci flow, and also as tangent flows, which were recently introduced by Bamler in [Bam20].

In this paper, we study the rigidity of certain gradient shrinking Ricci solitons (M,g)(M,g), that is, whether (M,g)(M,g) is isolated in the space of shrinking Ricci solitons (modulo diffeomorphism and rescaling). Our first main theorem is the following local rigidity theorem for S2×S2S^{2}\times S^{2}.

Theorem 1.1.

Let (M,g)(M,g) be S2×S2S^{2}\times S^{2} with the round unit metric on each factor. There is a C2,αC^{2,\alpha}-neighbourhood 𝒰\mathcal{U} of gg such that any other gradient shrinking Ricci soliton in 𝒰\mathcal{U}, is equivalent to gg up to diffeomorphism and rescaling.

The question of local rigidity of S2×S2S^{2}\times S^{2} was previously posed by Kröncke [Krö16, Example 6.4]. Our approach here is the Ricci flow analogue of our methods in [SZ20] for Sm×SnS^{m}\times S^{n} as mean curvature flow solitons (see also [ELS20], [Zhu20]). These methods are needed to handle non-integrable infinitesimal deformations. In contrast to the MCF setting, as Ricci solitons such deformations only arise when at least one spherical factor has dimension 22.

Our methods also can be used to show local rigidity for a class of product manifolds S2×NS^{2}\times N, where NN is a 1-Einstein manifold satisfying

  • (\dagger)

    λ1(ΔN)>2\lambda_{1}(\Delta^{N})>2, and kerTT(ΔLN+2)=0\ker_{TT}(\Delta_{L}^{N}+2)=0.

Here we use the sign convention Δ=div\Delta=\operatorname{div}\nabla for the Laplacian on functions, λ1\lambda_{1} is the first nonzero eigenvalue and ΔL\Delta_{L} is the Lichnerowicz Laplacian; kerTT\ker_{TT} refers to the kernel when acting on transverse traceless 2-tensors. We prove the following:

Theorem 1.2.

Let (M,g)(M,g) be S2×NS^{2}\times N, where S2S^{2} has the round unit metric and (N,gN)(N,g_{N}) is a closed 11-Einstein manifold satisfying ()(\dagger). There is a C2,αC^{2,\alpha}-neighbourhood 𝒰\mathcal{U} of gg such that any other gradient shrinking Ricci soliton in 𝒰\mathcal{U}, is equivalent to gg up to diffeomorphism and rescaling.

There are many NN which satisfy the assumptions in Theorem 1.2. For example, any NN listed in [CH15, Table 2], that is not listed as “i.d.” and has (in their notation) λ1μfns>2-\lambda^{-1}\mu_{fns}>2.

Both theorems above are direct consequences of the following quantitative rigidity theorem. For this we introduce the shrinker quantity

(1.2) Φ(g):=τg(Ric(g)+Hessgfg)g2,\Phi(g):=\tau^{g}(\operatorname{Ric}(g)+\operatorname{Hess}^{g}f_{g})-\frac{g}{2},

where fg,τgf^{g},\tau^{g} are the pair realising the Perelman shrinker entropy ν\nu; an important property is that Φ\Phi is equivariant under pullback by diffeomorphisms, and under rescalings (see Section 2). Any compact gradient shrinking Ricci soliton must satisfy (1.1) with (f,τ)=(fg,τg)(f,\tau)=(f_{g},\tau_{g}), that is, Φ(g)=0\Phi(g)=0 (see [Krö15c, Remark 3.4]).

Theorem 1.3.

Let (M,g)(M,g) be either S2×S2S^{2}\times S^{2} or S2×NS^{2}\times N, where S2S^{2} has the round unit metric and (N,gN)(N,g_{N}) is a closed 11-Einstein manifold satisfying ()(\dagger). There exist C,ϵ0>0C,\epsilon_{0}>0 so that if g~\tilde{g} is a metric on MM with g~gC2,αε0\|\tilde{g}-g\|_{C^{2,\alpha}}\leq\varepsilon_{0}, then there exists c(11C,1+1C)c\in(1-\frac{1}{C},1+\frac{1}{C}) and a diffeomorphism ψ\psi of MM so that,

(1.3) cψg~gC2,α3CΦ(g~)C0,α,\|c\psi^{*}\tilde{g}-g\|_{C^{2,\alpha}}^{3}\leq C\|\Phi(\tilde{g})\|_{C^{0,\alpha}},

As Φ\Phi may be regarded as the L2L^{2}-gradient of the shrinker entropy ν\nu, Theorem 1.3 may be regarded as the Hölder version of a Łojasiewicz inequality for ν\nu about the critical points S2×S2S^{2}\times S^{2}, S2×NS^{2}\times N. The L2L^{2} version is somewhat more natural and would yield a gradient Łojasiewicz inequality as well, but the theory is significantly more technical so we leave it for future work; the Hölder version is more than sufficient for rigidity.

Background

The question of rigidity - whether a geometric object is isolated in its class, modulo the symmetries of that class - has been studied for various geometric structures in many contexts. For Ricci solitons, low dimensional solitons have been essentially classified, so their rigidity or non-rigidity is clear (the dimension 22 case is due to Hamilton [Ham88] and the dimension 33 case is due to Ivey [Ive93]). In higher dimensions, the classification is far from well-understood. From a geometric PDE perspective, Ricci solitons are natural generalizations of Einstein manifolds. Kröncke studied the local rigidity of Ricci solitons in [Krö16], using similar techniques as for Einstein manifolds in [Krö15b].

For non-rigidity results, similar to the Einstein case, Inoue [Ino19] studied the moduli space of Fano manifolds with Kähler-Ricci solitons. As a consequence, Inoue proved that some Kähler-Ricci solitons are non-rigid. It seems that the local rigidity of a Ricci soliton actually has deep connections to the topological/algebraic structure of the manifold, and we look forward to such interpretations for the class of Ricci solitons we study in this paper.

The study of rigidity of Einstein manifolds has a more developed history. The first rigidity result for Einstein manifolds was proved by Berger in [Ber66], where Berger proved that all Einstein metrics on SnS^{n} whose sectional curvature is (n2)(n1)(n-2)(n-1)-pinched are isometric to the standard sphere metric. We consider this to be a ‘global’ result in that it gives a large explicit neighbourhood in which the special structure is unique. Later, Berger-Ebin [BE69] proposed a general strategy to prove local rigidity of Einstein metrics. In particular, they introduced the notion of “infinitesimal deformations”. The local rigidity of symmetric spaces as Einstein manifolds was first studied by [Koi80], and Kröncke extended Koiso’s result in [Krö15b].

Meanwhile, it is known that certain moduli spaces of Einstein manifolds, even with an extra structure (e.g. Kähler), can be nontrivial. Then any metric in such a moduli space is an example of non-rigid Einstein manifold. For example, even in dimension 2, the moduli space of surfaces with genus gg of constant curvature 1-1 has dimension 6g66g-6. Hence when g>1g>1, such hyperbolic metrics are not rigid. We refer the readers to [Bes87, Chapter 12.J] for further discussion.

The Ricci soliton equation is both more general and more complicated than the Einstein equation, making it more challenging to obtain a rigidity result. It is also interesting to compare Ricci flow with other geometric flows, especially the most natural extrinsic flow - the mean curvature flow. There are a number of results on the rigidity of shrinking solitons to the mean curvature flow - see [CIM15], [ELS20], [SZ20], [Zhu20]. All these results are about product self-shrinkers. During the completion of this work, we learnt that Colding-Minicozzi are developing techniques which could be applied to study certain noncompact Ricci solitons [CMa, CMb].

Deformation theory for solitons

The starting point in the study of local rigidity is to consider infinitesimal deformations which satisfy the equation Φ=0\Phi=0 to first order. That is, one looks for 22-tensors in the kernel of the linearized operator 𝒟Φ\mathcal{D}\Phi. By analogy with the mean curvature flow setting, we call elements of this kernel Jacobi deformations; in the Ricci flow literature these are sometimes referred to as infinitesimal solitonic deformations [Krö16]. A Jacobi deformation is called integrable if it induces a continuous family of gradient shrinking Ricci solitons. Any diffeomorphism or rescaling generates an integrable Jacobi deformation.

In [Krö16], Kröncke found conditions under which certain Einstein manifolds are rigid as gradient shrinking Ricci solitons. His idea was to expand Φ=0\Phi=0 to higher and higher order and examine whether there are solutions to each order. If there are no solutions at a given order, we call this an obstruction of that order. Kröncke calculates only up to order 2: 2n\mathbb{CP}^{2n} has a second order obstruction, whilst for S2×S2S^{2}\times S^{2} all Jacobi deformations are integrable to second order. (For 2n+1\mathbb{CP}^{2n+1}, almost all Jacobi deformations are non-integrable to second order, but there is a set of positive codimension which is indeed integrable.)

The manifolds we study in this paper, also have Jacobi deformations which are integrable to order 2, but we will show they have an obstruction at order 3. In particular, we resolve the case of S2×S2S^{2}\times S^{2}. We also hope that our method may be adapted to determine the rigidity of 2n+1\mathbb{CP}^{2n+1}. As described below, expanding at orders higher than 2 result in inhomogenous elliptic PDEs, which makes the analysis significantly more complicated. For instance, we use the special structure of eigenfunctions on S2S^{2} to solve (1.6) explicitly.

In [SZ20], we studied the rigidity of products of spheres Sm×SnS^{m}\times S^{n} as mean curvature flow self-shrinkers, and found that despite the existence of nontrivial Jacobi deformations, there is an obstruction at third order. (Note that [ELS20], by a similar method, had previously covered the m,n=1m,n=1 case.) As mentioned above, when m,n3m,n\geq 3, as a Ricci shrinker Sm×SnS^{m}\times S^{n} does not have Jacobi deformations other than those from symmetry (a first order obstruction). However, when at least one of the factors is S2S^{2}, nontrivial Jacobi deformations appear once more. We remark that when considered in the class of Einstein manifolds, there are again no nontrivial deformations, even for S2S^{2} products.

Proof strategy

Our proof strategy is a deformation-obstruction theory for the shrinker quantity Φ\Phi, mirroring that we used in [SZ20] (also developed by the second named author in [Zhu20]). One may consult those papers for an overview of the method, including in toy cases, but we reproduce a sketch below for the readers’ convenience. The basic idea is similar to that of Kröncke [Krö16], but we study the expansion of Φ\Phi more directly to find a ‘uniformly’ obstructed term, which gives the quantitative rigidity.

Consider an elliptic functional \mathcal{F} defined on a Banach space \mathcal{E}. Let 𝒢:\mathcal{G}:\mathcal{E}\to\mathcal{E}^{\prime} be the Euler-Lagrange operator of \mathcal{F} and suppose 0 is a critical point of \mathcal{F}, so that 𝒢(0)=0\mathcal{G}(0)=0. In the Ricci soliton setting, \mathcal{F} will be the shrinker entropy ν\nu and 𝒢\mathcal{G} will be the Ricci shrinker quantity Φ(g+)\Phi(g+\cdot).

For hh\in\mathcal{E}, we have the formal Taylor expansion

(1.4) 𝒢(h)=𝒟𝒢(h)+12𝒟2𝒢(h,h)+16𝒟3𝒢(h,h,h)+.\mathcal{G}(h)=\mathcal{D}\mathcal{G}(h)+\frac{1}{2}\mathcal{D}^{2}\mathcal{G}(h,h)+\frac{1}{6}\mathcal{D}^{3}\mathcal{G}(h,h,h)+\cdots.

Here 𝒟𝒢(h)=Lh\mathcal{D}\mathcal{G}(h)=Lh, where LL is the linearised operator at 0. Its kernel 𝒦\mathcal{K} is the space of Jacobi deformations.

If 𝒦=0\mathcal{K}=0, ellipticity gives that LL is invertible, hence 𝒢\mathcal{G} is obstructed at order 1. (As mentioned above, modulo symmetries, this is the case for Sm×SnS^{m}\times S^{n}, m,n3m,n\geq 3.) Otherwise 𝒦0\mathcal{K}\neq 0, and the first order term suggests the decomposition h=h1+k1h=h_{1}+k_{1}, where k1𝒦k_{1}\in\mathcal{K} and h1h_{1} is in a complement 1\mathcal{E}_{1} of 𝒦\mathcal{K}. This gives the further expansion

(1.5) 𝒢(h)=Lh1+12𝒟2𝒢(k1,k1)+𝒟2𝒢(k1,h1)+16𝒟3𝒢(h1,h1,h1)+.\mathcal{G}(h)=Lh_{1}+\frac{1}{2}\mathcal{D}^{2}\mathcal{G}(k_{1},k_{1})+\mathcal{D}^{2}\mathcal{G}(k_{1},h_{1})+\frac{1}{6}\mathcal{D}^{3}\mathcal{G}(h_{1},h_{1},h_{1})+\cdots.

Invertibility of LL on 1\mathcal{E}_{1} implies that h1h_{1} is (at least) second order compared to hh and k1k_{1}. The second order term is then Lh1+12𝒟2𝒢(k1,k1)Lh_{1}+\frac{1}{2}\mathcal{D}^{2}\mathcal{G}(k_{1},k_{1}). By elliptic theory, there exists h1h_{1} for which this term is 0, if and only if π𝒦(𝒟2𝒢(k1,k1))=0\pi_{\mathcal{K}}(\mathcal{D}^{2}\mathcal{G}(k_{1},k_{1}))=0. In [Krö16], Kröncke essentially finds Einstein manifolds for which π𝒦(𝒟2𝒢(k1,k1))0\pi_{\mathcal{K}}(\mathcal{D}^{2}\mathcal{G}(k_{1},k_{1}))\neq 0 for all nonzero k1𝒦k_{1}\in\mathcal{K}. This would imply π𝒦(𝒟2𝒢(k1,k1))δk12\|\pi_{\mathcal{K}}(\mathcal{D}^{2}\mathcal{G}(k_{1},k_{1}))\|\geq\delta\|k_{1}\|^{2}, which we refer to as an order 2 obstruction, and in turn implies hk1𝒢12\|h\|\simeq\|k_{1}\|\lesssim\|\mathcal{G}\|^{\frac{1}{2}}.

For the cases considered in this paper, it turns out that π𝒦(𝒟2𝒢(k1,k1))0\pi_{\mathcal{K}}(\mathcal{D}^{2}\mathcal{G}(k_{1},k_{1}))\equiv 0. That is, there is always a solution of

(1.6) Lk2+12𝒟2𝒢(k1,k1)=0Lk_{2}+\frac{1}{2}\mathcal{D}^{2}\mathcal{G}(k_{1},k_{1})=0

for some k21k_{2}\in\mathcal{E}_{1}, which suggests the further decomposition h1=k2+h2h_{1}=k_{2}+h_{2}. The expansion then becomes

(1.7) 𝒢(h)=Lh2+𝒟2𝒢(k1,k2)+16𝒟3𝒢(k1,k1,k1)+,\mathcal{G}(h)=Lh_{2}+\mathcal{D}^{2}\mathcal{G}(k_{1},k_{2})+\frac{1}{6}\mathcal{D}^{3}\mathcal{G}(k_{1},k_{1},k_{1})+\cdots,

where again invertibility of LL shows that the residual h2h_{2} is of third order compared to hh. Repeating the process, we may attempt to solve the third order condition

(1.8) Lh2+𝒟2𝒢(k1,h2)+16𝒟3𝒢(k1,k1,k1)=0,Lh_{2}+\mathcal{D}^{2}\mathcal{G}(k_{1},h_{2})+\frac{1}{6}\mathcal{D}^{3}\mathcal{G}(k_{1},k_{1},k_{1})=0,

which depends on the projection π𝒦(𝒟2𝒢(k1,h2)+16𝒟3𝒢(k1,k1,k1)).\pi_{\mathcal{K}}\left(\mathcal{D}^{2}\mathcal{G}(k_{1},h_{2})+\frac{1}{6}\mathcal{D}^{3}\mathcal{G}(k_{1},k_{1},k_{1})\right). Note that the third order condition involves some lower order cross terms.

If the projection vanishes identically, we proceed to the next higher order. If at the mmth order, the corresponding projection is nonzero for any nonzero k1𝒦k_{1}\in\mathcal{K}, then we have an mmth order obstruction, which would imply the quantitative rigidity h𝒢1m\|h\|\lesssim\|\mathcal{G}\|^{\frac{1}{m}}. Again, for the cases considered here, we find a third order obstruction

(1.9) π𝒦(𝒟2𝒢(k1,k2)+16𝒟3𝒢(k1,k1,k1))δk13.\left\|\pi_{\mathcal{K}}\left(\mathcal{D}^{2}\mathcal{G}(k_{1},k_{2})+\frac{1}{6}\mathcal{D}^{3}\mathcal{G}(k_{1},k_{1},k_{1})\right)\right\|\geq\delta\|k_{1}\|^{3}.

Organisation of the paper

We collect some preliminary results in Section 2 and record descriptions of the Jacobi deformations on our S2S^{2} products in Section 3. In Section 4 we compute the variation of the shrinker quantities ν,τ,f,Φ\nu,\tau,f,\Phi, with the variation formulae for more basic geometric quantities recorded in Appendix A. Using the explicit description of Jacobi deformations, we specialise these variations to our S2S^{2} products in Section 5 and establish a formal obstruction at order 3. Finally, in Section 6, we use Taylor expansion and the formal obstruction to prove the main quantitative rigidity Theorem 1.3. The a priori bounds on 𝒟kΦ\mathcal{D}^{k}\Phi are deferred to Appendix B.

Acknowledgements.

The authors want to thank Professors Bill Minicozzi and Toby Colding for sharing in enlightening discussions about their work and ours. JZ would also like to thank Prof. Richard Bamler for helpful conversations. We thank Prof. Klaus Kröncke for rectifying our understanding of his work as well as Jiangtao Yu for pointing out an inaccuracy in the prior computation.

JZ was supported in part by the Australian Research Council under grant FL150100126.

2. Preliminaries

Throughout, we consider a closed Riemannian manifold (M,g)(M,g).

Our sign convention for the (rough) Laplacian is Δ=div\Delta=\operatorname{div}\nabla. Here, and henceforth, we suppress dependence on the metric gg when clear from context. When we wish to emphasise the domain, we will use Δ0\Delta_{0} for the Laplacian on functions, Δ1\Delta_{1} for the rough Laplacian on 1-forms and Δ2\Delta_{2} for the rough Laplacian on symmetric 2-tensors. When not otherwise specified or clear from context, Δ\Delta should be understood as the Laplacian on functions. With our sign convention, we say u0u\neq 0 is an eigenfunction of Δ\Delta with eigenvalue λ\lambda if Δu=λu\Delta u=-\lambda u, and similarly for other operators. We write λ1\lambda_{1} for the first nonzero eigenvalue.

A metric is said to be (μ(\mu-) Einstein if Ric(g)=μg\operatorname{Ric}(g)=\mu g for some μ\mu\in\mathbb{R}. Let 𝒮2(M)\mathcal{S}^{2}(M) denote the space of (smooth) symmetric 22-tensors on MM, and Ω1(M)\Omega^{1}(M) the space of (smooth) 1-forms on MM. Throughout the paper, we will use several different norms like the Sobolev norm and the Hölder norm. We can also study the tensors in the completion under these norms, but approximation by smooth sections easily yields the same estimates.

An important subspace is the space of transverse traceless (TT) 22-tensors, which we denote by 𝒮̊g2={h𝒮2(M)|trgh=0,divgh=0}\mathring{\mathcal{S}}^{2}_{g}=\{h\in\mathcal{S}^{2}(M)|\operatorname{tr}_{g}h=0,\operatorname{div}_{g}h=0\}. Then we have the well-known (Lg2L^{2}_{g}-) orthogonal decomposition

(2.1) 𝒮2(M)=(C(M)g+im)𝒮̊g2.\mathcal{S}^{2}(M)=(C^{\infty}(M)g+\operatorname*{im}\mathcal{L})\oplus\mathring{\mathcal{S}}^{2}_{g}.

Here :Ω1(M)𝒮2(M)\mathcal{L}:\Omega^{1}(M)\to\mathcal{S}^{2}(M) is the Lie derivative, (ω)ij=iωj+jωi(\mathcal{L}\omega)_{ij}=\nabla_{i}\omega_{j}+\nabla_{j}\omega_{i}. Note that \mathcal{L} is actually independent of the metric, and 12-\frac{1}{2}\mathcal{L} is the Lg2L^{2}_{g}-adjoint of divg\operatorname{div}_{g} for any gg.

We define the Lichnerowicz Laplacian ΔL:𝒮̊g2𝒮̊g2\Delta_{L}:\mathring{\mathcal{S}}^{2}_{g}\to\mathring{\mathcal{S}}^{2}_{g} as acting by

ΔLhij=Δhij+2RkijlhlkRikhjkRjkhik.\Delta_{L}h_{ij}=\Delta h_{ij}+2R^{l}_{kij}h^{k}_{l}-R^{k}_{i}h_{jk}-R^{k}_{j}h_{ik}.

Note that this definition also makes sense on all of 𝒮2(M)\mathcal{S}^{2}(M) but we choose to restrict the domain so that ker(ΔL)=kerTT(ΔT)\ker(\Delta_{L})=\ker_{TT}(\Delta_{T}) and so forth will refer to the TT kernel.

Define

𝒲(g,f,τ):=[τ(|f|2+R)+fn](4πτ)n/2efdVg.\mathcal{W}(g,f,\tau):=\int[\tau(|\nabla f|^{2}+R)+f-n](4\pi\tau)^{-n/2}e^{-f}\mathop{}\!\mathrm{d}V_{g}.

The shrinker entropy is defined by

(2.2) ν(g)=inf{𝒲(g,f,τ)|τ>0,(4πτ)n/2efdVg=1}.\nu(g)=\inf\left\{\mathcal{W}(g,f,\tau)|\tau>0,(4\pi\tau)^{-n/2}\int e^{-f}\mathop{}\!\mathrm{d}V_{g}=1\right\}.

As in [Krö15c, Lemma 4.1], given any gradient shrinking Ricci soliton, there is a C2,αC^{2,\alpha} neighbourhood in the space of metrics on which the minimisers fg,τgf^{g},\tau^{g} are unique and depend analytically on the metric; consequently ν\nu also depends analytically on the metric on this neighbourhood. Moreover, the first variation of ν\nu is given by

(2.3) 𝒟νg(h)=(4πτg)n/2Φ(g),hefgdVg,\mathcal{D}\nu_{g}(h)=-(4\pi\tau_{g})^{-n/2}\int\langle\Phi(g),h\rangle e^{-f_{g}}\mathop{}\!\mathrm{d}V_{g},

where the Ricci shrinker quantity Φ:𝒮2(M)𝒮2(M)\Phi:\mathcal{S}^{2}(M)\to\mathcal{S}^{2}(M) is defined by

(2.4) Φ(g)=τg(Ric(g)+Hessgfg)g2.\Phi(g)=\tau^{g}(\operatorname{Ric}(g)+\operatorname{Hess}^{g}f^{g})-\frac{g}{2}.

(Compact) gradient shrinking Ricci solitons are those metrics which satisfy Φ(g)=0\Phi(g)=0. For convenience we denote wg=efg(4πτg)n/2w^{g}=\frac{e^{-{f^{g}}}}{(4\pi{\tau^{g}})^{n/2}}. By definition, we have the normalisation wgdVg=1\int w^{g}\mathop{}\!\mathrm{d}V_{g}=1. Henceforth, for readability, we will suppress dependence on the metric where clear from context.

Any Einstein manifold (M,g)(M,g) is a gradient shrinking Ricci soliton with fgf_{g} constant. Henceforth, L2L^{2} will always refer to the Lebesgue space weighted by wgw^{g}; if gg is Einstein this is equivalent to Lg2L^{2}_{g} up to the constant weight ww.

It is known that (see [CCG+07]) the Euler-Lagrange equations for f,τf,\tau are

(2.5) τ(2Δf+|df|2R)f+n+ν=0.\tau(-2\Delta f+|df|^{2}-R)-f+n+\nu=0.
(2.6) fwdVg=n2+ν.\int fw\mathop{}\!\mathrm{d}V_{g}=\frac{n}{2}+\nu.

Through this paper, C,C,C′′C,C^{\prime},C^{\prime\prime} will denote constants which may change from line to line but retain their stated dependencies.

We record the following observation:

Lemma 2.1.
(2.7) divgΦ(g)=Φ(g)gf.\operatorname{div}_{g}\Phi(g)=\Phi(g)\cdot\nabla^{g}f.
Proof.

Using the Bianchi identity gives divRic=12R\operatorname{div}\operatorname{Ric}=\frac{1}{2}\nabla R. By commuting derivatives we have divHessf=Δf+Ricf=Δf+1τΦf+12τfHessff\operatorname{div}\operatorname{Hess}f=\nabla\Delta f+\operatorname{Ric}\cdot\nabla f=\nabla\Delta f+\frac{1}{\tau}\Phi\cdot\nabla f+\frac{1}{2\tau}\nabla f-\operatorname{Hess}f\cdot\nabla f. On the other hand, differentiating the defining equation for ff in space gives

τ(2Δf+2HessffR)f=0.\tau(-2\nabla\Delta f+2\operatorname{Hess}f\cdot\nabla f-\nabla R)-\nabla f=0.

This implies the result. ∎

2.1. Variations of shrinker quantities

Consider a soliton metric Φ(g)=0\Phi(g)=0. We consider the formal expansion of Φ\Phi given by

Φ(g+h)=𝒟Φg(h)+12𝒟2Φg(h,h)+16𝒟3Φg(h,h,h)+\Phi(g+h)=\mathcal{D}\Phi_{g}(h)+\frac{1}{2}\mathcal{D}^{2}\Phi_{g}(h,h)+\frac{1}{6}\mathcal{D}^{3}\Phi_{g}(h,h,h)+\cdots

where 𝒟kΦg\mathcal{D}^{k}\Phi_{g} is a symmetric kk-linear map 𝒮2(M)k𝒮2(M)\mathcal{S}^{2}(M)^{k}\to\mathcal{S}^{2}(M). For instance, L=𝒟ΦgL=\mathcal{D}\Phi_{g} is the linearisation of Φ\Phi at gg. We will henceforth suppress dependence on the initial metric gg. We will use the corresponding notation for variations of other quantities such as ν(g),τg,fg\nu(g),\tau_{g},f_{g}.

2.2. Deformations and Jacobi deformations

Given a 1-parameter family of metrics g(s)g(s) with g(0)=gg(0)=g, the corresponding infinitesimal deformation is g(0)𝒮2(M)g^{\prime}(0)\in\mathcal{S}^{2}(M). At a shrinking soliton Φ(g)=0\Phi(g)=0, we define 𝒦:=kerL\mathcal{K}:=\ker L to be the kernel of L=𝒟ΦgL=\mathcal{D}\Phi_{g}. We call elements of 𝒦\mathcal{K} Jacobi deformations; they correspond to deformations which preserve Φ=0\Phi=0 to first order and have previously been referred to as infinitesimal solitonic deformations.

The Ricci shrinker quantity has two natural symmetries: under scaling by c>0c>0 we have Φ(cg)=cΦ(g)\Phi(cg)=c\Phi(g), and under pullback by a diffeomorphism ψDiff(M)\psi\in\operatorname*{Diff}(M) we have Φ(ψg)=ψΦ(g)\Phi(\psi^{*}g)=\psi^{*}\Phi(g). The corresponding spaces of infinitesimal deformations are g\mathbb{R}g and im\operatorname*{im}\mathcal{L}, and if Φ(g)=0\Phi(g)=0 these are integrable subspaces of Jacobi deformations. We define the subspace of Jacobi deformations from symmetry to be 𝒦0:=im+g𝒦\mathcal{K}_{0}:=\operatorname*{im}\mathcal{L}+\mathbb{R}g\subset\mathcal{K}.

If gg is a μ\mu-Einstein metric, Kröncke [Krö16, Section 6] checked that

kerdivg={(Δv+μv)g+Hessgv}.\ker\operatorname{div}_{g}=\{(\Delta v+\mu v)g+\operatorname{Hess}^{g}v\}.

Note that Hessgv=12(v)\operatorname{Hess}^{g}v=\frac{1}{2}\mathcal{L}(\nabla v)^{\sharp}. A theorem of Lichnerowicz ([Lic58], see also [IV15, Theorem 2.1]) implies that λ1(Δ)>μ\lambda_{1}(\Delta)>\mu, and in particular that Δ+μ\Delta+\mu is invertible.

Let C̊g={uC(M)|vdVg=0}\mathring{C}^{\infty}_{g}=\{u\in C^{\infty}(M)|\int v\mathop{}\!\mathrm{d}V_{g}=0\}, so that the conformal deformation space satisfies

(2.8) C(M)g+im=(Δ+μ)C̊gg+𝒦0.C^{\infty}(M)g+\operatorname*{im}\mathcal{L}=(\Delta+\mu)\mathring{C}^{\infty}_{g}\cdot g+\mathcal{K}_{0}.

Without loss of generality, we will assume μ=1\mu=1. The action of LL is well-known and also computed in Section 4.1: It acts on the conformal part vgvg, where v=(Δ+1)v~v=(\Delta+1)\tilde{v}, v~C̊g\tilde{v}\in\mathring{C}^{\infty}_{g}, by

L(vg)=14(Δ+2)vg+14Hess(Δ+2)v~,L(vg)=-\frac{1}{4}(\Delta+2)vg+\frac{1}{4}\operatorname{Hess}(\Delta+2)\tilde{v},

and on the TT part h𝒮̊g2h\in\mathring{\mathcal{S}}^{2}_{g} by

L(h)=14(ΔL+2)h.L(h)=-\frac{1}{4}(\Delta_{L}+2)h.

Recall that we defined ΔL\Delta_{L} as an operator on 𝒮̊g2\mathring{\mathcal{S}}^{2}_{g}. The above implies the Lg2L^{2}_{g}-orthogonal decomposition of the kernel

(2.9) 𝒦=(𝒦0+𝒦1)ker(ΔL+2),\mathcal{K}=(\mathcal{K}_{0}+\mathcal{K}_{1})\oplus\ker(\Delta_{L}+2),

where

(2.10) 𝒦1:={vg|vker(Δ+2)}.\mathcal{K}_{1}:=\{vg|v\in\ker(\Delta+2)\}.

We denote by π𝒦\pi_{\mathcal{K}} the L2L^{2}-projection to 𝒦\mathcal{K}, and similarly for other subspaces.

To study deformations of a gradient shrinking Ricci soliton, we need to account for the symmetry action at an integrated level, not just infinitesimal. More precisely, given a gradient shrinking Ricci soliton (M,g)(M,g) and a metric g~\tilde{g} close to gg, we want to compare g~\tilde{g} to the symmetry orbit of gg (or vice versa).

Lemma 2.2.

Suppose (M,g)(M,g) is a μ\mu-Einstein manifold. For any δ>0\delta>0 there exists ϵ0\epsilon_{0} such that if g~gC2,αε0\|\tilde{g}-g\|_{C^{2,\alpha}}\leq\varepsilon_{0}, then there exists c(1δ,1+δ)c\in(1-\delta,1+\delta) and ψDiff(M)\psi\in\operatorname*{Diff}(M) such that

h:=cψg~g(Δ+μ)C̊gg𝒮̊g2,h:=c\psi^{*}\tilde{g}-g\in(\Delta+\mu)\mathring{C}^{\infty}_{g}\cdot g\oplus\mathring{\mathcal{S}}^{2}_{g},

and moreover hC2,αδ\|h\|_{C^{2,\alpha}}\leq\delta.

The proof is standard, using implicit function theorem for Banach spaces. We refer the readers to the proof of Theorem 3.6 in [Via14].

3. Jacobi deformations on Einstein manifolds

In this section we discuss Jacobi deformations on products of Einstein manifolds, with the goal of giving an explicit description of the Jacobi deformations on S2×S2S^{2}\times S^{2} and certain products S2×NS^{2}\times N. We also demonstrate that there are no Jacobi deformations on Sm×SnS^{m}\times S^{n}, m,n3m,n\geq 3, other than those from scaling and diffeomorphism.

3.1. Spectrum on product Einstein manifolds

By a standard separation of variables, spectra on a product manifold can be derived from the spectra of each component. For the Laplacian on functions, we have

(3.1) spec(Δ0M×N)=spec(Δ0M)+spec(Δ0N).\text{spec}(\Delta^{M\times N}_{0})=\text{spec}(\Delta^{M}_{0})+\text{spec}(\Delta^{N}_{0}).

For the spectrum of ΔL\Delta_{L}, it is convenient to first work on all of 𝒮2(M)\mathcal{S}^{2}(M). To match with previous literature, we define the Einstein operator ΔE:𝒮2(M)𝒮2(M)\Delta_{E}:\mathcal{S}^{2}(M)\to\mathcal{S}^{2}(M) by

ΔEhij=Δhij+2Rkijlhlk.\Delta_{E}h_{ij}=\Delta h_{ij}+2R^{l}_{kij}h^{k}_{l}.

Note that this differs from the expression for ΔL\Delta_{L} by exactly the Ricci terms; in the case of a μ\mu-Einstein manifold, for h𝒮̊g2h\in\mathring{\mathcal{S}}^{2}_{g} we have ΔEh=(ΔL+2)h\Delta_{E}h=(\Delta_{L}+2)h.

We have the following theorem (see [AM11] and [Krö15a]; note that our sign convention is opposite to Kröncke’s).

Theorem 3.1.

Suppose M,NM,N are μ\mu-Einstein manifolds. Then

(3.2) spec(ΔEM×N)=(spec(ΔEM)+spec(Δ0N))(spec(ΔEN)+spec(Δ0M))(spec(Δ1M)+spec(Δ1N)).\text{spec}(\Delta_{E}^{M\times N})=(\text{spec}(\Delta_{E}^{M})+\text{spec}(\Delta_{0}^{N}))\cup(\text{spec}(\Delta_{E}^{N})+\text{spec}(\Delta_{0}^{M}))\cup(\text{spec}(\Delta_{1}^{M})+\text{spec}(\Delta_{1}^{N})).

Moreover, the eigentensors on the product manifold may be given by (symmetric tensor) products of the corresponding eigensections.

3.2. Spectra on round spheres

We will always consider SnS^{n} with the metric induced as the sphere of radius n1\sqrt{n-1} in n+1\mathbb{R}^{n+1}, which is 11-Einstein. The spectra of the Laplace operators has been computed (see for instance [Bou99, Bou09]):

For k,nk,n\in\mathbb{N} define λk,n0=k(k+n1)n1\lambda^{0}_{k,n}=\frac{k(k+n-1)}{n-1}, λk,n1=k(k+n1)+n2n1\lambda^{1}_{k,n}=\frac{k(k+n-1)+n-2}{n-1} and λk,n2=k(k+n1)+2(n1)n1\lambda^{2}_{k,n}=\frac{k(k+n-1)+2(n-1)}{n-1}.

  • The spectrum of the Laplacian on functions Δ0\Delta_{0} consists of eigenfunctions φk,n\varphi_{k,n} with eigenvalue λk,n0\lambda^{0}_{k,n}, for k0k\geq 0.

  • The Hodge Laplacian acts on Ω1(Sn)\Omega^{1}(S^{n}) by Δ11\Delta_{1}-1 and has spectrum consisting of dφk,nd\varphi_{k,n}, k1k\geq 1, together with divergence-free eigenforms ωk,n\omega_{k,n} with eigenvalue λk,n1\lambda^{1}_{k,n}, for k1k\geq 1.

  • The (generally defined) Lichnerowicz Laplacian acts on 𝒮2(Sn)\mathcal{S}^{2}(S^{n}) by ΔE2\Delta_{E}-2, and its spectrum consists of: φk,ng\varphi_{k,n}g, k0k\geq 0; Hessφk,n\operatorname{Hess}\varphi_{k,n}, k2k\geq 2; ωk,n\mathcal{L}\omega_{k,n}, for k2k\geq 2; and TT eigenforms hk,nh_{k,n} with eigenvalue λk,n2\lambda^{2}_{k,n}, for k2k\geq 2.

That is,

spec(Δ0Sn)={λk,n0}k0,\text{spec}(\Delta^{S^{n}}_{0})=\{\lambda^{0}_{k,n}\}_{k\geq 0},
spec(Δ1Sn)={λk,n01}k1{λk,n11}k1,\text{spec}(\Delta^{S^{n}}_{1})=\{\lambda^{0}_{k,n}-1\}_{k\geq 1}\cup\{\lambda^{1}_{k,n}-1\}_{k\geq 1},
spec(ΔESn)={λk,n02}k0{λk,n12}k2{λk,n22}k2,\text{spec}(\Delta^{S^{n}}_{E})=\{\lambda^{0}_{k,n}-2\}_{k\geq 0}\cup\{\lambda^{1}_{k,n}-2\}_{k\geq 2}\cup\{\lambda^{2}_{k,n}-2\}_{k\geq 2},

In particular, since λ2,n2=4nn1>2\lambda^{2}_{2,n}=\frac{4n}{n-1}>2 for any n2n\geq 2, there are no TT eigentensors of eigenvalue 0, that is, ker(ΔL+2)=0\ker(\Delta_{L}+2)=0.

Later we focus on the n=2n=2 case (which is in fact the unit sphere), so we record separately

{λk,20}k0={0,2,6},{λk,21}k1={2,6,},{λk,22}k2={8,}.\{\lambda^{0}_{k,2}\}_{k\geq 0}=\{0,2,6\cdots\},\qquad\{\lambda^{1}_{k,2}\}_{k\geq 1}=\{2,6,\cdots\},\qquad\{\lambda^{2}_{k,2}\}_{k\geq 2}=\{8,\cdots\}.

On S2S^{2}, the 22-eigenspace ker(Δ0+2)\ker(\Delta_{0}+2) consists of the coordinate functions θi=x,ei\theta_{i}=\langle x,e_{i}\rangle, which satisfy θi=eiT\nabla\theta_{i}=e_{i}^{T}, θi,θj=δijθiθj\langle\nabla\theta_{i},\nabla\theta_{j}\rangle=\delta_{ij}-\theta_{i}\theta_{j}, Hessθi=θig\operatorname{Hess}\theta_{i}=-\theta_{i}g and Δ0(θiθj)=2δij6θiθj\Delta_{0}(\theta_{i}\theta_{j})=2\delta_{ij}-6\theta_{i}\theta_{j}.

3.3. Jacobi deformations on product manifolds

If the product manifold is Sm×SnS^{m}\times S^{n} for m,n3m,n\geq 3, there is no Jacobi deformation other than those induced by diffeomorphisms from the spectrum on product manifolds. This is a consequence of the following lemma.

Lemma 3.2.

Let M=Sm×SnM=S^{m}\times S^{n}, with the round 1-Einstein metric gg. If m,n3m,n\geq 3 then

ker(Δ0+2)=0,ker(ΔL+2)=0.\ker(\Delta_{0}+2)=0,\qquad\ker(\Delta_{L}+2)=0.
Proof.

For the Laplacian on functions, note that for n3n\geq 3 we have λ1,n0>1\lambda^{0}_{1,n}>1, so the sum of any two nonzero eigenvalues is strictly greater than 2. But λ2,n0>2\lambda^{0}_{2,n}>2, so λ1(Δ0M)>2\lambda_{1}(\Delta^{M}_{0})>2.

For the Lichnerowicz Laplacian, note that spec(Δ1Sn)\text{spec}(\Delta^{S^{n}}_{1}) is strictly positive, so clearly 0spec(Δ1Sm)+spec(Δ1Sn)0\notin\text{spec}(\Delta^{S^{m}}_{1})+\text{spec}(\Delta^{S^{n}}_{1}). Also note that the least eigenvalue of ΔESm\Delta_{E}^{S^{m}} is 2-2, and the next eigenvalue is >1>-1. Since λ1,n0>1\lambda^{0}_{1,n}>1 and λ2,n0>2\lambda^{0}_{2,n}>2 we conclude using Theorem 3.1 that 0spec(ΔEM)0\notin\text{spec}(\Delta^{M}_{E}). ∎

In the following we only focus on the product manifolds with S2S^{2} components.

We give precise descriptions of Jacobi deformations on S2×S2S^{2}\times S^{2} and S2×NS^{2}\times N. Recall that we decomposed the kernel as

(3.3) 𝒦=(𝒦0+𝒦1)ker(ΔL+2),\mathcal{K}=(\mathcal{K}_{0}+\mathcal{K}_{1})\oplus\ker(\Delta_{L}+2),

where 𝒦0\mathcal{K}_{0} is the space of Jacobi deformations from symmetries, and

(3.4) 𝒦1:={vg|vker(Δ+2)}.\mathcal{K}_{1}:=\{vg|v\in\ker(\Delta+2)\}.

Let π1,π2\pi_{1},\pi_{2} be the projections onto each factor.

Lemma 3.3.

Let M=S2×S2M=S^{2}\times S^{2} with the round 1-Einstein metric gg, we have

ker(Δ+2)={π1u1+π2u2|u1,u2ker(ΔS2+2)},ker(ΔL+2)=0.\ker(\Delta+2)=\{\pi_{1}^{*}u_{1}+\pi_{2}^{*}u_{2}|u_{1},u_{2}\in\ker(\Delta^{S^{2}}+2)\},\qquad\ker(\Delta_{L}+2)=0.
Proof.

For the Laplacian on functions, since λ1,20=2\lambda^{0}_{1,2}=2, the only product eigenfunctions with eigenvalue 2 are those which are constant on one factor and have eigenvalue 2 on the other.

For the Lichnerowicz Laplacian, spec(Δ1S2)\text{spec}(\Delta^{S^{2}}_{1}) is strictly positive, so clearly 0spec(Δ1S2)+spec(Δ1S2)0\notin\text{spec}(\Delta^{S^{2}}_{1})+\text{spec}(\Delta^{S^{2}}_{1}). Also the least eigenvalue of ΔES2\Delta_{E}^{S^{2}} is 2-2, and the next eigenvalue is 0. Moreover the eigenspaces are ker(ΔES2)={vgS2|vker(Δ0S2+2)}\ker(\Delta^{S^{2}}_{E})=\{vg_{S^{2}}|v\in\ker(\Delta^{S^{2}}_{0}+2)\} and ker(ΔES22)=gS2\ker(\Delta^{S^{2}}_{E}-2)=\mathbb{R}g_{S^{2}}. So the only product eigentensors in ker(ΔEM)\ker(\Delta^{M}_{E}) are of the forms 1v2g21\cdot v_{2}g^{2}, v1g2v_{1}\cdot g^{2}, v1g11v_{1}g^{1}\cdot 1 or g1v2g^{1}\cdot v_{2}. Here gbg^{b} is the (pulled-back) metric on the bbth S2S^{2} factor, b=1,2b=1,2, and vb=πbubv_{b}=\pi_{b}^{*}u_{b}, ubker(ΔS2+2)u_{b}\in\ker(\Delta^{S^{2}}+2). But vb(g1+g2)=vbgC(M)gv_{b}(g^{1}+g^{2})=v_{b}g\in C^{\infty}(M)g, and vbgb=Hessvbimv_{b}g^{b}=-\operatorname{Hess}v_{b}\in\operatorname*{im}\mathcal{L}. The TT kernel is therefore ker(ΔLM+2)=0\ker(\Delta^{M}_{L}+2)=0. ∎

We will consider 1-Einstein manifolds NN, which additionally satisfy:

  • (\dagger)

    λ1(Δ0N)>2\lambda_{1}(\Delta^{N}_{0})>2, and ker(ΔLN+2)=0\ker(\Delta^{N}_{L}+2)=0.

Lemma 3.4.

Let M=S2×NM=S^{2}\times N, where S2S^{2} has the round 1-Einstein metric and NN is a 11-Einstein manifold satisfying (\dagger), we have

ker(Δ+2)={π1u1|u1ker(ΔS2+2)},ker(ΔL+2)=0.\ker(\Delta+2)=\{\pi_{1}^{*}u_{1}|u_{1}\in\ker(\Delta^{S^{2}}+2)\},\qquad\ker(\Delta_{L}+2)=0.
Proof.

For the Laplacian on functions, since λ1,20=2\lambda^{0}_{1,2}=2, the only product eigenfunctions with eigenvalue 2 are those which are constant on NN and have eigenvalue 2 on the S2S^{2} factor.

For the Lichnerowicz Laplacian, spec(Δ1S2)\text{spec}(\Delta^{S^{2}}_{1}) is strictly positive and spec(Δ1N)\text{spec}(\Delta^{N}_{1}) is nonnegative for any Einstein manifold (cf. [Krö15a, Lemma 4.4]), so 0spec(Δ1S2)+spec(Δ1N)0\notin\text{spec}(\Delta^{S^{2}}_{1})+\text{spec}(\Delta^{N}_{1}). Then by ()(\dagger), the only product eigentensors in ker(ΔEM)\ker(\Delta^{M}_{E}) are of the forms 1π2(Nω)1\cdot\pi_{2}^{*}(\mathcal{L}_{N}\omega) or vgS21vg_{S^{2}}\cdot 1, where ωker(Δ1N+1)\omega\in\ker(\Delta^{N}_{1}+1) and v=π1uv=\pi_{1}^{*}u, uker(Δ0S2+2)u\in\ker(\Delta^{S^{2}}_{0}+2). But π2(Nω)=M(π2ω)\pi_{2}^{*}(\mathcal{L}_{N}\omega)=\mathcal{L}_{M}(\pi_{2}^{*}\omega), and again vgS2=HessvimMvg_{S^{2}}=-\operatorname{Hess}v\in\operatorname*{im}\mathcal{L}_{M}. The TT kernel is therefore ker(ΔLM+2)=0\ker(\Delta^{M}_{L}+2)=0. ∎

Later, it will be convenient to collect the two previous cases as M=b=1BSb2×N~M=\prod_{b=1}^{B}S^{2}_{b}\times\tilde{N}. We will write any vker(Δ+2)v\in\ker(\Delta+2) as v=vbv=\sum v_{b}, where vbker(Δb+2)v_{b}\in\ker(\Delta^{b}+2) and Δb\Delta^{b} is the Laplacian on the bbth S2S^{2} factor. It will be understood that each vbv_{b} is independent of any other factors; explicitly, vb=πbubv_{b}=\pi_{b}^{*}u_{b} in the above.

4. Variations

In this section we compute the variation of the quantities ν,f,τ,Φ\nu,f,\tau,\Phi, by considering certain families of metrics g(s,t)g(s,t) with g(0,0)=gg(0,0)=g. The initial soliton will be an Einstein manifold (M,g)(M,g), with Einstein constant μ=1\mu=1, hence τ=τ0=12\tau=\tau_{0}=\frac{1}{2} and f=f0=n2+νf=f_{0}=\frac{n}{2}+\nu. In particular the initial weight w=w0w=w_{0} is also a constant.

Note that the variations of more familiar differential geometric quantities are given in Appendix A. For the more complicated quantities here, the procedure will be as follows:

  1. (1)

    Use lower order variations to calculate variation of ν\nu.

  2. (2)

    Differentiate (2.6) and wdVg=1\int w\mathop{}\!\mathrm{d}V_{g}=1 to calculate variation of τ\tau.

  3. (3)

    Differentiate (2.5) to calculate variation of ff.

  4. (4)

    Substitute in to find variation of Φ\Phi.

We always denote u=(Δ+1)u~u=(\Delta+1)\tilde{u} where udVg=u~dVg=0\int u\mathop{}\!\mathrm{d}V_{g}=\int\tilde{u}\mathop{}\!\mathrm{d}V_{g}=0, and similarly for vv.

Note that any variation of Hess\operatorname{Hess} or Δ\Delta will act as 0 on any constant function on MM. In particular, at the initial metric f=f0f=f_{0} is constant so 𝒟kHessf=0,𝒟kΔf=0\mathcal{D}^{k}\operatorname{Hess}\cdot f=0,\mathcal{D}^{k}\Delta\cdot f=0, k1k\geq 1.

For convenience, we define

η=log(wdVg),\eta=\log(w\mathop{}\!\mathrm{d}V_{g}),

where dVg\mathop{}\!\mathrm{d}V_{g} understood as the volume ratio using the initial metric as the reference metric.

Step (2) is a general procedure, but here as our initial soliton is Einstein, it will be expedient for us to actually consider ϕwdVg\int\phi w\mathop{}\!\mathrm{d}V_{g}, where ϕ=(1f0)+f\phi=(1-f_{0})+f. Note that ϕwdVg=(1f0)+n2+ν\int\phi w\mathop{}\!\mathrm{d}V_{g}=(1-f_{0})+\frac{n}{2}+\nu, so variations of ϕwdVg\int\phi w\mathop{}\!\mathrm{d}V_{g} match those of ν\nu. Moreover, ϕ0=1\phi_{0}=1, while all variations of ϕ\phi match those of ff.

4.1. First variation

We begin with the first variation as it is distinguished in our method. So we consider the 1-parameter family of metrics g(t)g(t), where to start gg^{\prime} is a general element of 𝒮2(M)\mathcal{S}^{2}(M) satisfying (trgg)dVg=0\int(\operatorname{tr}_{g}g^{\prime})\mathop{}\!\mathrm{d}V_{g}=0. Here we use primes to denote the derivative in tt, evaluated at 0.

As Φ=0\Phi=0 at t=0t=0, we immediately have

𝒟ν(g)=ν=Φ,gwdVg=0.\mathcal{D}\nu(g^{\prime})=\nu^{\prime}=-\int\langle\Phi,g^{\prime}\rangle w\mathop{}\!\mathrm{d}V_{g}=0.

Note that η=(log(wdVg))=nτ2τf+12trgg\eta^{\prime}=(\log(w\mathop{}\!\mathrm{d}V_{g}))^{\prime}=-\frac{n\tau^{\prime}}{2\tau}-f^{\prime}+\frac{1}{2}\operatorname{tr}_{g}g^{\prime}. Then (ϕwdVg)=wdVg(ϕη+ϕ).(\int\phi w\mathop{}\!\mathrm{d}V_{g})^{\prime}=\int w\mathop{}\!\mathrm{d}V_{g}(\phi\eta^{\prime}+\phi^{\prime}).

Taking ϕ=(1f0)+f\phi=(1-f_{0})+f, since ν=0\nu^{\prime}=0 we have 0=wdVg(η+f)=wdVg(nτ2τ+12trgg).0=\int w\mathop{}\!\mathrm{d}V_{g}(\eta^{\prime}+f^{\prime})=\int w\mathop{}\!\mathrm{d}V_{g}(-\frac{n\tau^{\prime}}{2\tau}+\frac{1}{2}\operatorname{tr}_{g}g^{\prime}). Therefore

𝒟τ(g)=τ=0.\mathcal{D}\tau(g^{\prime})=\tau^{\prime}=0.

Differentiating (2.5) and using ν=τ=0\nu^{\prime}=\tau^{\prime}=0 gives

(4.1) 0=fΔf12R.0=-f^{\prime}-\Delta f^{\prime}-\frac{1}{2}R^{\prime}.

4.1.1. TT direction

Here we compute 𝒟Φ(h)\mathcal{D}\Phi(h), where h𝒮̊g2h\in\mathring{\mathcal{S}}^{2}_{g} is transverse traceless. To match the notation used for higher variations, we use subscripts to denote derivatives in tt.

As trh=divh=0\operatorname{tr}h=\operatorname{div}h=0, the variation formulae in Appendix A.1 give Rt=0R_{t}=0 and therefore

𝒟f(h)=ft=0.\mathcal{D}f(h)=f_{t}=0.

They also give

𝒟Ric(h)=Rict=Rikjlhkl+hij12Δhij=12ΔLh,\mathcal{D}\operatorname{Ric}(h)=\operatorname{Ric}_{t}=-R_{ikjl}h_{kl}+h_{ij}-\frac{1}{2}\Delta h_{ij}=-\frac{1}{2}\Delta_{L}h,

where ΔL\Delta_{L} is the Lichnerowicz Laplacian.

Then

𝒟Φ(h)=Φt=τRictgt2=14(ΔLh+2h).\mathcal{D}\Phi(h)=\Phi_{t}=\tau\operatorname{Ric}_{t}-\frac{g_{t}}{2}=-\frac{1}{4}(\Delta_{L}h+2h).

4.1.2. Conformal direction

Here we compute 𝒟Φ(vg)\mathcal{D}\Phi(vg), where v=(Δ+1)v~v=(\Delta+1)\tilde{v}, v~dVg=0\int\tilde{v}\mathop{}\!\mathrm{d}V_{g}=0. For use in higher variations, it will be convenient to change the parameter to ss. That is, here we consider the family (1+sv)g(1+sv)g.

Since Δ\Delta commutes with itself, (4.1) gives that

𝒟f(vg)=fs=(Δ+1)1Rs2=(n2+n12Δ)v~=12v~+n12v.\mathcal{D}f(vg)=f_{s}=-(\Delta+1)^{-1}\frac{R_{s}}{2}=\left(\frac{n}{2}+\frac{n-1}{2}\Delta\right)\tilde{v}=\frac{1}{2}\tilde{v}+\frac{n-1}{2}v.

Here we have used the variation formulae in Appendix A.1 (or Appendix A.2). It will be useful for later to record

𝒟Ric(vg)=Rics=n22Hessv12(Δv)g.\mathcal{D}\operatorname{Ric}(vg)=\operatorname{Ric}_{s}=-\frac{n-2}{2}\operatorname{Hess}v-\frac{1}{2}(\Delta v)g.

Then we have

𝒟Φ(vg)=Φs\displaystyle\mathcal{D}\Phi(vg)=\Phi_{s} =\displaystyle= τ(Rics+Hessfs)gs2\displaystyle\tau(\operatorname{Ric}_{s}+\operatorname{Hess}f_{s})-\frac{g_{s}}{2}
=\displaystyle= 12μ(n22Hessv12(Δv)g+0+12Hess(v~+(n1)v))vg2\displaystyle\frac{1}{2\mu}\left(-\frac{n-2}{2}\operatorname{Hess}v-\frac{1}{2}(\Delta v)g+0+\frac{1}{2}\operatorname{Hess}(\tilde{v}+(n-1)v)\right)-\frac{vg}{2}
=\displaystyle= 14(Δ+2)vg+14Hess(Δ+2)v~.\displaystyle-\frac{1}{4}(\Delta+2)vg+\frac{1}{4}\operatorname{Hess}(\Delta+2)\tilde{v}.

This identifies the conformal kernel as {vg|vker(Δ+2)}\{vg|v\in\ker(\Delta+2)\}.

4.2. Higher variations in the conformal kernel direction

We now proceed to find variations in the directions vgvg, where vker(Δ+2)v\in\ker(\Delta+2), by consider the 1-parameter family of metrics (1+sv)g(1+sv)g.

4.2.1. Second variation

First, we have νss=wdVg(Φs,gs+Φ,gss+Φ,gsηs)\nu_{ss}=-\int w\mathop{}\!\mathrm{d}V_{g}(\langle\Phi_{s},g_{s}\rangle+\langle\Phi,g_{ss}\rangle+\langle\Phi,g_{s}\rangle\eta_{s}). Since at s=0s=0 we have Φ=Φs=0\Phi=\Phi_{s}=0, this immediately implies

𝒟2ν(vg,vg)=νss=0.\mathcal{D}^{2}\nu(vg,vg)=\nu_{ss}=0.

We calculate

(ϕw)ss=wdVg(ϕηs2+ϕηss+2ϕsηs+ϕss).\left(\int\phi w\right)_{ss}=\int w\mathop{}\!\mathrm{d}V_{g}(\phi\eta_{s}^{2}+\phi\eta_{ss}+2\phi_{s}\eta_{s}+\phi_{ss}).

By the above, since Δv=2v\Delta v=-2v, we have fs=n22vf_{s}=\frac{n-2}{2}v and

ηs=(log(wdVg))s=nτs2τfs+12trggs=v.\eta_{s}=(\log(w\mathop{}\!\mathrm{d}V_{g}))_{s}=-\frac{n\tau_{s}}{2\tau}-f_{s}+\frac{1}{2}\operatorname{tr}_{g}g_{s}=v.

We also compute

ηss=nτss2τ+n(τs)22τ2fss+12trggss12|gs|2=nτss2τfssn2v2.\eta_{ss}=-\frac{n\tau_{ss}}{2\tau}+\frac{n(\tau_{s})^{2}}{2\tau^{2}}-f_{ss}+\frac{1}{2}\operatorname{tr}_{g}g_{ss}-\frac{1}{2}|g_{s}|^{2}=-\frac{n\tau_{ss}}{2\tau}-f_{ss}-\frac{n}{2}v^{2}.

Then once again taking ϕ=(1f0)+f\phi=(1-f_{0})+f, we have

0=wdVg(ηs2+ηss+2fsηs+fss)=wdVg(v2+(n2)v2nτss2τn2v2).0=\int w\mathop{}\!\mathrm{d}V_{g}(\eta_{s}^{2}+\eta_{ss}+2f_{s}\eta_{s}+f_{ss})=\int w\mathop{}\!\mathrm{d}V_{g}\left(v^{2}+(n-2)v^{2}-\frac{n\tau_{ss}}{2\tau}-\frac{n}{2}v^{2}\right).

This gives

𝒟2τ(vg,vg)=τss=n22nv2wdVg,\mathcal{D}^{2}\tau(vg,vg)=\tau_{ss}=\frac{n-2}{2n}\int v^{2}w\mathop{}\!\mathrm{d}V_{g},

Now differentiating (2.5) twice, we have

0=τssRfss+12(2Δfss4Δsfs+2|dfs|2Rss).0=-\tau_{ss}R-f_{ss}+\frac{1}{2}(-2\Delta f_{ss}-4\Delta_{s}f_{s}+2|df_{s}|^{2}-R_{ss}).

Using the conformal variation formulae in Appendix A.2 gives

(4.2) 𝒟2f(vg,vg)=fss=nτss+(1+Δ)1(nv23n24|dv|2).\mathcal{D}^{2}f(vg,vg)=f_{ss}=-n\tau_{ss}+(1+\Delta)^{-1}\left(nv^{2}-\frac{3n-2}{4}|dv|^{2}\right).

At s=0s=0 we then calculate Φss=τss(Ric+Hessf)+τ(Ricss+Hessfss+2Hesssfs)\Phi_{ss}=\tau_{ss}(\operatorname{Ric}+\operatorname{Hess}f)+\tau(\operatorname{Ric}_{ss}+\operatorname{Hess}f_{ss}+2\operatorname{Hess}_{s}f_{s}), so using Φ=0\Phi=0 we have

(4.3) 𝒟2Φ(vg,vg)=Φss=τss2g+12Hessfss+n22vHessv+n24dvdv+12(2v2+|dv|2)g.\mathcal{D}^{2}\Phi(vg,vg)=\Phi_{ss}=\frac{\tau_{ss}}{2}g+\frac{1}{2}\operatorname{Hess}f_{ss}+\frac{n-2}{2}v\operatorname{Hess}v+\frac{n-2}{4}dv\otimes dv+\frac{1}{2}(-2v^{2}+|dv|^{2})g.

Note that, by differentiating (2.7) twice, at s=0s=0 (using Φ=Φs=0\Phi=\Phi_{s}=0) we have

(4.4) div𝒟2Φ(vg,vg)=divΦss=0.\operatorname{div}\mathcal{D}^{2}\Phi(vg,vg)=\operatorname{div}\Phi_{ss}=0.

4.2.2. Third variation

Here we compute 𝒟3Φ(vg,vg,vg),vgL2\langle\mathcal{D}^{3}\Phi(vg,vg,vg),vg\rangle_{L^{2}}, where vker(Δ+2)v\in\ker(\Delta+2).

For brevity, we will omit the solution of 𝒟3τ(vg,vg,vg)=τsss\mathcal{D}^{3}\tau(vg,vg,vg)=\tau_{sss} (and νsss\nu_{sss}) - it is constant on MM, and therefore will not contribute after integration against vv anyway.

The third derivative of (2.5) gives

0=τsssRfsss+3τss(2ΔfsRs)+τ(2Δfsss6Δsfss6ΔssfsRsss+6(gij)s(fs),i(fs),j+6gij(fs),i(fss)j).\begin{split}0=&-\tau_{sss}R-f_{sss}+3\tau_{ss}(-2\Delta f_{s}-R_{s})\\ &+\tau\left(-2\Delta f_{sss}-6\Delta_{s}f_{ss}-6\Delta_{ss}f_{s}-R_{sss}+6(g^{ij})_{s}(f_{s})_{,i}(f_{s})_{,j}+6g^{ij}(f_{s})_{,i}(f_{ss})_{j}\right).\end{split}

Using the variation formulae in Appendix A.2 then gives the following defining equation for 𝒟3f(vg,vg,vg)=fsss\mathcal{D}^{3}f(vg,vg,vg)=f_{sss},

(4.5) (1+Δ)fsss+nτsss=3(n2)τssv+3vΔfss3v3(3n2)12n275n+664v|dv|2.\begin{split}(1+\Delta)f_{sss}+n\tau_{sss}=&-3(n-2)\tau_{ss}v+3v\Delta f_{ss}\\ &-3v^{3}(3n-2)-\frac{12n^{2}-75n+66}{4}v|dv|^{2}.\end{split}

We then have

Φsss=τsss(Ric+Hessf)+3τss(Rics+Hessfs)+τ(Ricsss+3Hessssfs+3Hesssfss+Hessfsss).\Phi_{sss}=\tau_{sss}(\operatorname{Ric}+\operatorname{Hess}f)+3\tau_{ss}(\operatorname{Ric}_{s}+\operatorname{Hess}f_{s})+\tau(\operatorname{Ric}_{sss}+3\operatorname{Hess}_{ss}f_{s}+3\operatorname{Hess}_{s}f_{ss}+\operatorname{Hess}f_{sss}).

Since Φ=Φs=0\Phi=\Phi_{s}=0 at s=0s=0, we then have

(4.6) 𝒟3Φ(vg,vg,vg)=12Hessfsss+τsssg+3τssvg3(n2)2v2Hessv3(n2)vdvdv+3v3g3(3n10)4v|dv|2g+34(dvdfssdfssdv+dv,dfssg).\begin{split}\mathcal{D}^{3}\Phi(vg,vg,vg)=&\frac{1}{2}\operatorname{Hess}f_{sss}+\tau_{sss}g+3\tau_{ss}vg-\frac{3(n-2)}{2}v^{2}\operatorname{Hess}v\\ &-3(n-2)vdv\otimes dv+3v^{3}g-\frac{3(3n-10)}{4}v|dv|^{2}g\\ &+\frac{3}{4}(-dv\otimes df_{ss}-df_{ss}\otimes dv+\langle dv,df_{ss}\rangle g).\end{split}

Note that the τsssg\tau_{sss}g term will vanish upon integration against vgvg.

Taking the inner product we have

(4.7) 𝒟3Φ(vg,vg,vg),vgL2=(6(n1)v49n218n244v2|dv|2+3(n2)4vdv,dfss)wdVg+12vΔfssswdVg+3(n2)2(v2wdVg)2.\begin{split}\langle\mathcal{D}^{3}\Phi(vg,vg,vg)&,vg\rangle_{L^{2}}\\ &=\int\left(6(n-1)v^{4}-\frac{9n^{2}-18n-24}{4}v^{2}|dv|^{2}+\frac{3(n-2)}{4}v\langle dv,df_{ss}\rangle\right)w\mathop{}\!\mathrm{d}V_{g}\\ &\quad+\frac{1}{2}\int v\Delta f_{sss}w\mathop{}\!\mathrm{d}V_{g}+\frac{3(n-2)}{2}\left(\int v^{2}w\mathop{}\!\mathrm{d}V_{g}\right)^{2}.\end{split}

Recall fssf_{ss} and fsssf_{sss} are defined by (4.2) and (4.5) respectively.

4.3. Second variation cross terms

Finally, we calculate certain cross terms for 𝒟2Φ\mathcal{D}^{2}\Phi.

4.3.1. Kernel-TT direction

Here we compute 𝒟2Φ(vg,h),ugL2\langle\mathcal{D}^{2}\Phi(vg,h),ug\rangle_{L^{2}}, where vker(Δ+2)v\in\ker(\Delta+2), h𝒮̊g2h\in\mathcal{\mathring{S}}^{2}_{g} and uu is any smooth function on MM, by considering the 2-parameter variation g+svg+thg+svg+th.

By the calculations in Section 4.1 we have τs=τt=0\tau_{s}=\tau_{t}=0 and fs=n22vf_{s}=\frac{n-2}{2}v, ft=0f_{t}=0.

First, using Φ=0\Phi=0 at (s,t)=(0,0)(s,t)=(0,0), we have

𝒟2ν(vg,h)=νst=Φs,gtwdVg=vtrg(Φt)wdVg=0,\mathcal{D}^{2}\nu(vg,h)=\nu_{st}=-\int\langle\Phi_{s},g_{t}\rangle w\mathop{}\!\mathrm{d}V_{g}=-\int v\operatorname{tr}_{g}(\Phi_{t})w\mathop{}\!\mathrm{d}V_{g}=0,

as the Lichnerowicz Laplacian preserves the TT decomposition.

Calculate

(ϕw)st=wdVg(ϕηsηt+ϕsηt+ϕtηs+ϕηst+ϕst).\left(\int\phi w\right)_{st}=\int w\mathop{}\!\mathrm{d}V_{g}(\phi\eta_{s}\eta_{t}+\phi_{s}\eta_{t}+\phi_{t}\eta_{s}+\phi\eta_{st}+\phi_{st}).

Now as gs,gt=vg,h=0\langle g_{s},g_{t}\rangle=v\langle g,h\rangle=0 we have ηst=n2τstτfst\eta_{st}=-\frac{n}{2}\frac{\tau_{st}}{\tau}-f_{st}, and from the calculations in Section 4.1 we have ηs=v\eta_{s}=v, ηt=0\eta_{t}=0. Taking ϕ=1+ff0\phi=1+f-f_{0} again, we have ϕt=ft=0\phi_{t}=f_{t}=0 and ϕst=0\phi_{st}=0. So since νst=0\nu_{st}=0 we have

0=wdVg(ηsηt+fsηt+ftηs+ηst+fst)0=\int w\mathop{}\!\mathrm{d}V_{g}(\eta_{s}\eta_{t}+f_{s}\eta_{t}+f_{t}\eta_{s}+\eta_{st}+f_{st})

and hence

𝒟2τ(vg,h)=τst=0.\mathcal{D}^{2}\tau(vg,h)=\tau_{st}=0.

Differentiating (2.5) in ss then tt gives τ(2Δtfs2ΔfstRst)fst=0.\tau(-2\Delta_{t}f_{s}-2\Delta f_{st}-R_{st})-f_{st}=0.

Note that Δt\Delta_{t} acts as Δtϕ=h,Hessϕ\Delta_{t}\phi=-\langle h,\operatorname{Hess}\phi\rangle. Using the formulae in Section A.3 then gives (Δ+1)fst=n22h,Hessv12Rst=0,(\Delta+1)f_{st}=\frac{n-2}{2}\langle h,\operatorname{Hess}v\rangle-\frac{1}{2}R_{st}=0, hence

𝒟2f(vg,h)=fst=0.\mathcal{D}^{2}f(vg,h)=f_{st}=0.

As gst=0g_{st}=0 we then have Φst=τ(Ricst+Hesstfs).\Phi_{st}=\tau(\operatorname{Ric}_{st}+\operatorname{Hess}_{t}f_{s}). Since hh is TT, the first variation Lemma A.1 gives Hesst=0\operatorname{Hess}_{t}=0. Therefore Φst,ugL2=12Ricst,ugL2\langle\Phi_{st},ug\rangle_{L^{2}}=\frac{1}{2}\langle\operatorname{Ric}_{st},ug\rangle_{L^{2}}.

Again using Appendix A.3 we conclude that

(4.8) 𝒟2Φ(vg,h),ugL2=12𝒟2Ric(vg,h),ugL2=n24uh,HessvwdVg.\langle\mathcal{D}^{2}\Phi(vg,h),ug\rangle_{L^{2}}=\frac{1}{2}\langle\mathcal{D}^{2}\operatorname{Ric}(vg,h),ug\rangle_{L^{2}}=\frac{n-2}{4}\int u\langle h,\operatorname{Hess}v\rangle w\mathop{}\!\mathrm{d}V_{g}.

4.3.2. Kernel-conformal direction

Here we compute 𝒟2Φ(vg,ug),vgL2\langle\mathcal{D}^{2}\Phi(vg,ug),vg\rangle_{L^{2}}, where vker(Δ+2)v\in\ker(\Delta+2), u=(Δ+1)u~u=(\Delta+1)\tilde{u} and udVg=0\int u\mathop{}\!\mathrm{d}V_{g}=0, by considering the metric variation (1+sv+tu)g(1+sv+tu)g. We furthermore assume u~\tilde{u} (hence uu) is orthogonal to ker(Δ+2)\ker(\Delta+2). By calculations in Section 4.1 we have τs=τt=0\tau_{s}=\tau_{t}=0 and fs=n22vf_{s}=\frac{n-2}{2}v, ft=12u~+n12uf_{t}=\frac{1}{2}\tilde{u}+\frac{n-1}{2}u.

Again since Φs=0\Phi_{s}=0 at s=0s=0, we have

𝒟2ν(vg,ug)=νst=τΦs,gtwdVg=0.\mathcal{D}^{2}\nu(vg,ug)=\nu_{st}=-\int\tau\langle\Phi_{s},g_{t}\rangle w\mathop{}\!\mathrm{d}V_{g}=0.

As before we have

(ϕw)st=wdVg(ϕηsηt+ϕsηt+ϕtηs+ϕηst+ϕst).\left(\int\phi w\right)_{st}=\int w\mathop{}\!\mathrm{d}V_{g}(\phi\eta_{s}\eta_{t}+\phi_{s}\eta_{t}+\phi_{t}\eta_{s}+\phi\eta_{st}+\phi_{st}).

Now as gs,gt=nuv\langle g_{s},g_{t}\rangle=nuv we have ηst=n2τstτfst12nuv\eta_{st}=-\frac{n}{2}\frac{\tau_{st}}{\tau}-f_{st}-\frac{1}{2}nuv, and as in Section 4.1 we have ηs=v\eta_{s}=v, ηt=12(uu~)\eta_{t}=\frac{1}{2}(u-\tilde{u}). Taking ϕ=1+ff0\phi=1+f-f_{0} again, since νst=0\nu_{st}=0 we again have

0=wdVg(ηsηt+fsηt+ftηs+ηst+fst).0=\int w\mathop{}\!\mathrm{d}V_{g}(\eta_{s}\eta_{t}+f_{s}\eta_{t}+f_{t}\eta_{s}+\eta_{st}+f_{st}).

It then follows from u,u~u,\tilde{u} being orthogonal to vker(Δ+2)v\in\ker(\Delta+2) that

𝒟2τ(vg,ug)=τst=0.\mathcal{D}^{2}\tau(vg,ug)=\tau_{st}=0.

Differentiating (2.5) in ss then tt gives τ(2Δsft2Δtfs2ΔfstRst)fst=0\tau(-2\Delta_{s}f_{t}-2\Delta_{t}f_{s}-2\Delta f_{st}-R_{st})-f_{st}=0. This yields the defining equation for 𝒟2f(vg,ug)=fst\mathcal{D}^{2}f(vg,ug)=f_{st},

(4.9) (Δ+1)fst=12vΔu~n12vΔun24du~,dvn24du,dv.(\Delta+1)f_{st}=\frac{1}{2}v\Delta\tilde{u}-\frac{n-1}{2}v\Delta u-\frac{n-2}{4}\langle d\tilde{u},dv\rangle-\frac{n^{2}}{4}\langle du,dv\rangle.

Then, using the conformal cross-variations in Appendix A.2.1, we have

𝒟2Φ(vg,ug)=Φst\displaystyle\mathcal{D}^{2}\Phi(vg,ug)=\Phi_{st} =\displaystyle= τ(Ricst+Hesssft+Hesstfs+Hessfst)\displaystyle\tau(\operatorname{Ric}_{st}+\operatorname{Hess}_{s}f_{t}+\operatorname{Hess}_{t}f_{s}+\operatorname{Hess}f_{st})
=\displaystyle= 3(n2)8(dudv+dvdu)+n24(uHessv+vHessu)\displaystyle\frac{3(n-2)}{8}(du\otimes dv+dv\otimes du)+\frac{n-2}{4}(u\operatorname{Hess}v+v\operatorname{Hess}u)
+14(2uv+vΔu)gn44du,dvg\displaystyle+\frac{1}{4}(-2uv+v\Delta u)g-\frac{n-4}{4}\langle du,dv\rangle g
+14(dvft+dftdv+dv,dftg)\displaystyle+\frac{1}{4}(-dv\otimes f_{t}+df_{t}\otimes dv+\langle dv,df_{t}\rangle g)
+n28(dudvdvdu+du,dvg)+12Hessfst.\displaystyle+\frac{n-2}{8}(-du\otimes dv-dv\otimes du+\langle du,dv\rangle g)+\frac{1}{2}\operatorname{Hess}f_{st}.

Taking the inner product with vgvg and collecting terms, it follows that

(4.10) 𝒟2Φ(vg,ug),vgL2=(n24vdu,dv(n1)uv2+n24vdu~,dv+n12v2Δu)wdVg+12vΔfstwdVg,\begin{split}\langle\mathcal{D}^{2}\Phi(vg,ug),vg\rangle_{L^{2}}=&\int\left(\frac{n^{2}}{4}v\langle du,dv\rangle-(n-1)uv^{2}+\frac{n-2}{4}v\langle d\tilde{u},dv\rangle+\frac{n-1}{2}v^{2}\Delta u\right)w\mathop{}\!\mathrm{d}V_{g}\\ &+\frac{1}{2}\int v\Delta f_{st}w\mathop{}\!\mathrm{d}V_{g},\end{split}

where fstf_{st} is defined as in (4.9).

5. Specialising to S2S^{2} products

In this section we consider a product of 1-Einstein manifolds M=S2×NM=S^{2}\times N, where S2S^{2} has the (unit) round metric. The goal is to derive a formal obstruction for Φ\Phi at order 3. To do so, we will use explicit knowledge of the Jacobi deformations on MM together with the formulae in the previous section.

5.0.1. Explicit form

We assume that NN either is S2S^{2}, or satisfies assumption (\dagger), so that the results of Section 3.3 apply. It will actually be convenient to unify the cases by collecting S2S^{2} factors: We write M=b=1BSb2×N~M=\prod_{b=1}^{B}S^{2}_{b}\times\tilde{N}, where B=2,N~=B=2,\tilde{N}=* if N=S2N=S^{2}, and otherwise B=1,N~=NB=1,\tilde{N}=N. Note here the subscript bb serves only to distinguish the isometric S2S^{2} factors.

Throughout this section, we consider vker(Δ+2)v\in\ker(\Delta+2). As in Section 3.3, we have v=bvbv=\sum_{b}v_{b}, where we may write vb=αbθbv_{b}=\alpha_{b}\theta^{b} and θb=x,yb\theta^{b}=\langle x,y_{b}\rangle for some αb\alpha_{b}\in\mathbb{R} and some vector |yb|=1|y_{b}|=1. Then as in Section 3.2, we have |dvb|2=αb2vb2|dv_{b}|^{2}=\alpha_{b}^{2}-v_{b}^{2} and Δvb2=2αb26vb2\Delta v_{b}^{2}=2\alpha_{b}^{2}-6v_{b}^{2}.

We introduce the notation σk=bαbk\sigma_{k}=\sum_{b}\alpha_{b}^{k} and Sv=vb2S_{v}=\sum v_{b}^{2}. Then v2=αaαbθaθbv^{2}=\sum\alpha_{a}\alpha_{b}\theta^{a}\theta^{b}, |dv|2=σ2Sv|dv|^{2}=\sigma_{2}-S_{v}, and Δ(v2)=2σ24v22Sv\Delta(v^{2})=2\sigma_{2}-4v^{2}-2S_{v}. Finally note Hessv=vbgb\operatorname{Hess}v=-\sum v_{b}g^{b}.

5.1. Spherical integrals

Lemma 5.1.

Let M=b=1BSb2×N~M=\prod_{b=1}^{B}S^{2}_{b}\times\tilde{N} and v=vbv=\sum v_{b} be as in Section 5.0.1. Then

wdVg=1,,wvb2dVg=αb2/3,w\int\mathop{}\!\mathrm{d}V_{g}=1,,\qquad w\int v_{b}^{2}\mathop{}\!\mathrm{d}V_{g}=\alpha_{b}^{2}/3,
wvb4dVg=αb4/5,wva2vb2dVg=αa2αb2/9 if ab.w\int v_{b}^{4}\mathop{}\!\mathrm{d}V_{g}=\alpha_{b}^{4}/5,\qquad w\int v_{a}^{2}v_{b}^{2}\mathop{}\!\mathrm{d}V_{g}=\alpha_{a}^{2}\alpha_{b}^{2}/9\quad\text{ if }a\neq b.

Moreover,

wv2dVg=wSvdVg=13σ2,wv2vb2dVg=19σ2αb2+445αb4,w\int v^{2}\mathop{}\!\mathrm{d}V_{g}=w\int S_{v}\mathop{}\!\mathrm{d}V_{g}=\frac{1}{3}\sigma_{2},\qquad w\int v^{2}v_{b}^{2}\mathop{}\!\mathrm{d}V_{g}=\frac{1}{9}\sigma_{2}\alpha_{b}^{2}+\frac{4}{45}\alpha_{b}^{4},
wv4dVg=13σ22215σ4,wSv2dVg=19σ22+445σ4.w\int v^{4}\mathop{}\!\mathrm{d}V_{g}=\frac{1}{3}\sigma_{2}^{2}-\frac{2}{15}\sigma_{4},\qquad w\int S_{v}^{2}\mathop{}\!\mathrm{d}V_{g}=\frac{1}{9}\sigma_{2}^{2}+\frac{4}{45}\sigma_{4}.
Proof.

Note that S2θj2=4π3\int_{S^{2}}\theta_{j}^{2}=\frac{4\pi}{3}, S2θj4=4π5\int_{S^{2}}\theta_{j}^{4}=\frac{4\pi}{5} and |S2|=4π|S^{2}|=4\pi. Also note that w1=|M|=(4π)B|N~|w^{-1}=|M|=(4\pi)^{B}|\tilde{N}|. The first four equations are then routine computations. For the last four, note that spherical polynomials of odd degree integrate to 0. ∎

5.2. Explicit solutions

In this subsection, we use the explicit form of vv to explicitly solve for certain quantities in Section 4.

We will frequently use the ansatz that a function ϕ\phi is a linear combination

(5.1) ϕ=A1v2+A2Sv+A3σ2+B1vvb+B2vb2+B3αb2,\phi=A_{1}v^{2}+A_{2}S_{v}+A_{3}\sigma_{2}+B_{1}vv_{b}+B_{2}v_{b}^{2}+B_{3}\alpha_{b}^{2},

and formally solve the resulting equations as linear systems. Under this ansatz, we represent such a function by the column vector (A1A2A3B1B2B3)\begin{pmatrix}A_{1}\\ A_{2}\\ A_{3}\\ B_{1}\\ B_{2}\\ B_{3}\end{pmatrix}, and the Laplacian Δ\Delta acts as

M=(400000260000220000000400000260000220).M=\begin{pmatrix}-4&0&0&0&0&0\\ -2&-6&0&0&0&0\\ 2&2&0&0&0&0\\ 0&0&0&-4&0&0\\ 0&0&0&-2&-6&0\\ 0&0&0&2&2&0\end{pmatrix}.

The Laplacian Δb\Delta^{b} on a particular Sb2S^{2}_{b} factor acts as

Mb=(000000000000000000400200260460220220).M_{b}=\begin{pmatrix}0&0&0&0&0&0\\ 0&0&0&0&0&0\\ 0&0&0&0&0&0\\ -4&0&0&-2&0&0\\ -2&-6&0&-4&-6&0\\ 2&2&0&2&2&0\end{pmatrix}.

5.2.1. Second variation of ff

Here we find an explicit solution for the function fss=𝒟2f(vg,vg)f_{ss}=\mathcal{D}^{2}f(vg,vg) in Section 4.2.1. Recall that the defining equation (4.2) is

(1+Δ)fss=nv23n24|dv|2nτss=nv2+3n24Sv11n1012σ2,(1+\Delta)f_{ss}=nv^{2}-\frac{3n-2}{4}|dv|^{2}-n\tau_{ss}=nv^{2}+\frac{3n-2}{4}S_{v}-\frac{11n-10}{12}\sigma_{2},

where we have used

τss=𝒟2τ(vg,vg)=n22nv2wdVg=n26nσ2.\tau_{ss}=\mathcal{D}^{2}\tau(vg,vg)=\frac{n-2}{2n}\int v^{2}w\mathop{}\!\mathrm{d}V_{g}=\frac{n-2}{6n}\sigma_{2}.

We use the ansatz fss=A1v2+A2Sv+A3σ2f_{ss}=A_{1}v^{2}+A_{2}S_{v}+A_{3}\sigma_{2}, and solve the linear system as

(A1A2A3000)=(1+M)1(n3n2411n1012000)=(n3n66013n3860000).\begin{pmatrix}A_{1}\\ A_{2}\\ A_{3}\\ 0\\ 0\\ 0\end{pmatrix}=(1+M)^{-1}\begin{pmatrix}n\\ \frac{3n-2}{4}\\ -\frac{11n-10}{12}\\ 0\\ 0\\ 0\end{pmatrix}=\begin{pmatrix}-\frac{n}{3}\\ -\frac{n-6}{60}\\ -\frac{13n-38}{60}\\ 0\\ 0\\ 0\end{pmatrix}.

One may verify that this solution indeed satisfies the defining equation, which proves

Lemma 5.2.

Suppose M=b=1BSb2×N~M=\prod_{b=1}^{B}S^{2}_{b}\times\tilde{N} and vv is as in Section 5.0.1. Then

(5.2) 𝒟2f(vg,vg)=fss=n3v2n660Sv13n3860σ2\mathcal{D}^{2}f(vg,vg)=f_{ss}=-\frac{n}{3}v^{2}-\frac{n-6}{60}S_{v}-\frac{13n-38}{60}\sigma_{2}

is the unique solution of (4.2).

5.2.2. Second variation of Φ\Phi

Proposition 5.3.

Let M=b=1BSb2×N~M=\prod_{b=1}^{B}S^{2}_{b}\times\tilde{N} and v=vbv=\sum v_{b} be as in Section 5.0.1. Then

π𝒦𝒟2Φ(vg,vg)=0.\pi_{\mathcal{K}}\mathcal{D}^{2}\Phi(vg,vg)=0.
Proof.

By the results of Section 3.3, we have 𝒦=𝒦0+𝒦1\mathcal{K}=\mathcal{K}_{0}+\mathcal{K}_{1}, and 𝒦1={vg|vker(Δ+2)}\mathcal{K}_{1}=\{vg|v\in\ker(\Delta+2)\}. Since div𝒟2Φ(vg,vg)=0\operatorname{div}\mathcal{D}^{2}\Phi(vg,vg)=0, we immediately have π𝒦0𝒟2Φ(vg,vg)=0\pi_{\mathcal{K}_{0}}\mathcal{D}^{2}\Phi(vg,vg)=0.

Now the second variation formula (4.3) and Lemma 5.2, we see that trg𝒟2Φ(vg,vg)\operatorname{tr}_{g}\mathcal{D}^{2}\Phi(vg,vg) is a spherical polynomial (on b=1BSb2\prod_{b=1}^{B}S^{2}_{b}) of even (total) degree. Then for any vker(Δ+2)v^{\prime}\in\ker(\Delta+2), we see that 𝒟2Φ(vg,vg),vgL2\langle\mathcal{D}^{2}\Phi(vg,vg),v^{\prime}g\rangle_{L^{2}} is the integral of a spherical polynomial of odd degree, and therefore vanishes. This implies π𝒦1𝒟2Φ(vg,vg)=0\pi_{\mathcal{K}_{1}}\mathcal{D}^{2}\Phi(vg,vg)=0 and hence the result. ∎

5.2.3. Second variation of the metric

Since π𝒦𝒟2Φ(vg,vg)=0\pi_{\mathcal{K}}\mathcal{D}^{2}\Phi(vg,vg)=0, there is now a unique u~C̊g\tilde{u}\in\mathring{C}^{\infty}_{g}, u~ker(Δ+2)\tilde{u}\perp\ker(\Delta+2) and h𝒮̊g2h\in\mathring{\mathcal{S}}^{2}_{g} such that

(5.3) 𝒟Φ(ug+h)=12𝒟2Φ(vg,vg),\mathcal{D}\Phi(ug+h)=-\frac{1}{2}\mathcal{D}^{2}\Phi(vg,vg),

where u=(Δ+1)u~u=(\Delta+1)\tilde{u}. We will give an explicit solution for uu. For hh, it will later suffice to give an explicit solution for h,gb\langle h,g^{b}\rangle modulo ker(Δ+2)\ker(\Delta+2).

First, tracing (5.3) using the first variation formulae in Section 4.1 and the second variation formulae (4.3) gives the defining equation for u~\tilde{u}:

(5.4) (Δ+2)(n+(n1)Δ)u~=4tr𝒟Φ(ug)=2tr𝒟2Φ(vg,vg)=n𝒟2τ(vg,vg)+Δ𝒟2f(vg,vg)4(n1)v2+3n22|dv|2=Δ𝒟2f(vg,vg)4(n1)v23n22Sv+5n43σ2.\begin{split}(\Delta+2)(n+(n-1)\Delta)\tilde{u}&=-4\operatorname{tr}\mathcal{D}\Phi(ug)=2\operatorname{tr}\mathcal{D}^{2}\Phi(vg,vg)\\ &=n\mathcal{D}^{2}\tau(vg,vg)+\Delta\mathcal{D}^{2}f(vg,vg)-4(n-1)v^{2}+\frac{3n-2}{2}|dv|^{2}\\ &=\Delta\mathcal{D}^{2}f(vg,vg)-4(n-1)v^{2}-\frac{3n-2}{2}S_{v}+\frac{5n-4}{3}\sigma_{2}.\end{split}

Using the same ansatz u~=A1v2+A2Sv+A3σ2\tilde{u}=A^{\prime}_{1}v^{2}+A^{\prime}_{2}S_{v}+A^{\prime}_{3}\sigma_{2}, and the explicit solution of 𝒟2f(vg,vg)\mathcal{D}^{2}f(vg,vg), we can solve the linear system as

(A1A2A3000)\displaystyle\begin{pmatrix}A^{\prime}_{1}\\ A^{\prime}_{2}\\ A^{\prime}_{3}\\ 0\\ 0\\ 0\end{pmatrix} =\displaystyle= (n+(n1)M)1(2+M)1M(n3n66013n3860000)\displaystyle(n+(n-1)M)^{-1}(2+M)^{-1}M\begin{pmatrix}-\frac{n}{3}\\ -\frac{n-6}{60}\\ -\frac{13n-38}{60}\\ 0\\ 0\\ 0\end{pmatrix}
+(n+(n1)M)1(2+M)1(4(n1)3n225n43000)\displaystyle+(n+(n-1)M)^{-1}(2+M)^{-1}\begin{pmatrix}-4(n-1)\\ -\frac{3n-2}{2}\\ \frac{5n-4}{3}\\ 0\\ 0\\ 0\end{pmatrix}
=\displaystyle= (2(2n3)3(3n4)247n2678n+45660(3n4)(5n6)(n6)(4n5)30n(5n6)000).\displaystyle\begin{pmatrix}-\frac{2(2n-3)}{3(3n-4)}\\ \frac{247n^{2}-678n+456}{60(3n-4)(5n-6)}\\ -\frac{(n-6)(4n-5)}{30n(5n-6)}\\ 0\\ 0\\ 0\end{pmatrix}.

One may verify that the corresponding solution

u~=2(2n3)3(3n4)v2+247n2678n+45660(3n4)(5n6)Sv(n6)(4n5)30n(5n6)σ2\tilde{u}=-\frac{2(2n-3)}{3(3n-4)}v^{2}+\frac{247n^{2}-678n+456}{60(3n-4)(5n-6)}S_{v}-\frac{(n-6)(4n-5)}{30n(5n-6)}\sigma_{2}

indeed satisfies equation (5.4). For convenience we also record

(5.5) u=(Δ+1)u~=2(2n3)3n4v229n282n+564(3n4)(5n6)Sv11n219n+66n(5n6)σ2.u=(\Delta+1)\tilde{u}=\frac{2(2n-3)}{3n-4}v^{2}-\frac{29n^{2}-82n+56}{4(3n-4)(5n-6)}S_{v}-\frac{11n^{2}-19n+6}{6n(5n-6)}\sigma_{2}.

With u~\tilde{u} in hand, we consider the pointwise inner products hb:=h,gbh_{b}:=\langle h,g^{b}\rangle. We have

𝒟Φ(ug+h),gb=14Δb(Δ+2)u~12(Δ+2)u14(Δ+2)hb,\langle\mathcal{D}\Phi(ug+h),g^{b}\rangle=\frac{1}{4}\Delta_{b}(\Delta+2)\tilde{u}-\frac{1}{2}(\Delta+2)u-\frac{1}{4}(\Delta+2)h_{b},

so pairing (5.3) with gbg^{b} we find the defining equation

(5.6) (Δ+2)hb\displaystyle(\Delta+2)h_{b} =\displaystyle= (Δ+2)(Δbu~2u)+2𝒟2Φ(vg,vg),gb\displaystyle(\Delta+2)(\Delta^{b}\tilde{u}-2u)+2\langle\mathcal{D}^{2}\Phi(vg,vg),g^{b}\rangle
=\displaystyle= (Δ+2)(Δbu~2u)+n23nσ2+Δb𝒟2f(vg,vg)\displaystyle(\Delta+2)(\Delta^{b}\tilde{u}-2u)+\frac{n-2}{3n}\sigma_{2}+\Delta^{b}\mathcal{D}^{2}f(vg,vg)
2(n2)vvb+n22|dvb|24v2+2|dv|2.\displaystyle-2(n-2)vv_{b}+\frac{n-2}{2}|dv_{b}|^{2}-4v^{2}+2|dv|^{2}.

We will find h~b\tilde{h}_{b} so that (Δ+2)h~b(\Delta+2)\tilde{h}_{b} equals the right hand side of (5.6); then hbh~bker(Δ+2)h_{b}-\tilde{h}_{b}\in\ker(\Delta+2). Here we use the ansatz h~b=A1′′v2+A2′′Sv+A3′′σ2+B1vvb+B2vb2+B3αb2\tilde{h}_{b}=A^{\prime\prime}_{1}v^{2}+A^{\prime\prime}_{2}S_{v}+A^{\prime\prime}_{3}\sigma_{2}+B_{1}vv_{b}+B_{2}v_{b}^{2}+B_{3}\alpha_{b}^{2}. Using the solution of u~\tilde{u} (hence uu), we may solve the linear system as

(A1′′A2′′A3′′B1B2B3)\displaystyle\begin{pmatrix}A^{\prime\prime}_{1}\\ A^{\prime\prime}_{2}\\ A^{\prime\prime}_{3}\\ B_{1}\\ B_{2}\\ B_{3}\end{pmatrix} =\displaystyle= Mb(2(2n3)3(3n4)247n2678n+45660(3n4)(5n6)(n6)(4n5)30n(5n6)000)2(2(2n3)3n429n282n+564(3n4)(5n6)11n219n+66n(5n6)000)\displaystyle M_{b}\begin{pmatrix}-\frac{2(2n-3)}{3(3n-4)}\\ \frac{247n^{2}-678n+456}{60(3n-4)(5n-6)}\\ -\frac{(n-6)(4n-5)}{30n(5n-6)}\\ 0\\ 0\\ 0\end{pmatrix}-2\begin{pmatrix}\frac{2(2n-3)}{3n-4}\\ -\frac{29n^{2}-82n+56}{4(3n-4)(5n-6)}\\ -\frac{11n^{2}-19n+6}{6n(5n-6)}\\ 0\\ 0\\ 0\end{pmatrix}
+(M+2)1Mb(n3n66013n3860000)+(M+2)1(427n23n2(n2)n22n22)\displaystyle+(M+2)^{-1}M_{b}\begin{pmatrix}-\frac{n}{3}\\ -\frac{n-6}{60}\\ -\frac{13n-38}{60}\\ 0\\ 0\\ 0\end{pmatrix}+(M+2)^{-1}\begin{pmatrix}-4\\ -2\\ \frac{7n-2}{3n}\\ -2(n-2)\\ -\frac{n-2}{2}\\ \frac{n-2}{2}\end{pmatrix}
=\displaystyle= (2(n2)3n4(n2)(7n8)(3n4)(5n6)(n2)(17n18)6n(5n6)n(n2)3n4n(n2)(7n8)2(3n4)(5n6)(n1)(n2)5n6.).\displaystyle\begin{pmatrix}-\frac{2(n-2)}{3n-4}\\ \frac{(n-2)(7n-8)}{(3n-4)(5n-6)}\\ \frac{(n-2)(17n-18)}{6n(5n-6)}\\ \frac{n(n-2)}{3n-4}\\ -\frac{n(n-2)(7n-8)}{2(3n-4)(5n-6)}\\ -\frac{(n-1)(n-2)}{5n-6}.\end{pmatrix}.

One may verify that the corresponding solution indeed solves (5.6). In conclusion, we have proven

Lemma 5.4.

Let M=b=1BSb2×N~M=\prod_{b=1}^{B}S^{2}_{b}\times\tilde{N} and v=vbv=\sum v_{b} be as in Section 5.0.1. There is a unique solution (u~,h)(\tilde{u},h) of 𝒟Φ(ug+h)=12𝒟2Φ(vg,vg)\mathcal{D}\Phi(ug+h)=-\frac{1}{2}\mathcal{D}^{2}\Phi(vg,vg), where u=(Δ+1)u~u=(\Delta+1)\tilde{u}, u~C̊g\tilde{u}\in\mathring{C}^{\infty}_{g}, u~ker(Δ+2)\tilde{u}\perp\ker(\Delta+2), h𝒮̊g2h\in\mathring{\mathcal{S}}^{2}_{g}. Moreover they satisfy

(5.7) u~=2(2n3)3(3n4)v2+247n2678n+45660(3n4)(5n6)Sv(n6)(4n5)30n(5n6)σ2,\tilde{u}=-\frac{2(2n-3)}{3(3n-4)}v^{2}+\frac{247n^{2}-678n+456}{60(3n-4)(5n-6)}S_{v}-\frac{(n-6)(4n-5)}{30n(5n-6)}\sigma_{2},
(5.8) hb=h,gb=2(n2)3n4v2+(n2)(7n8)(3n4)(5n6)Sv+(n2)(17n18)6n(5n6)σ2+n(n2)3n4vvbn(n2)(7n8)2(3n4)(5n6)vb2(n1)(n2)5n6αb2+φb,\begin{split}h_{b}=\langle h,g^{b}\rangle=&-\frac{2(n-2)}{3n-4}v^{2}+\frac{(n-2)(7n-8)}{(3n-4)(5n-6)}S_{v}+\frac{(n-2)(17n-18)}{6n(5n-6)}\sigma_{2}\\ &+\frac{n(n-2)}{3n-4}vv_{b}-\frac{n(n-2)(7n-8)}{2(3n-4)(5n-6)}v_{b}^{2}-\frac{(n-1)(n-2)}{5n-6}\alpha_{b}^{2}\\ &+\varphi_{b},\end{split}

for some φbker(Δ+2)\varphi_{b}\in\ker(\Delta+2).

5.3. Integration by parts

We will use the following integration by parts tricks. The first equation below, in particular, is used for terms involving derivatives of fgf_{g}.

Lemma 5.5.

Let M=b=1BSb2×N~M=\prod_{b=1}^{B}S^{2}_{b}\times\tilde{N} and vv is as in Section 5.0.1. Then for any smooth function ϕ\phi, we have

vΔϕdVg=2v(1+Δ)ϕdVg.\int v\Delta\phi\mathop{}\!\mathrm{d}V_{g}=2\int v(1+\Delta)\phi\mathop{}\!\mathrm{d}V_{g}.
v2ΔϕdVg=ϕ(2σ24v22Sv)dVg,\int v^{2}\Delta\phi\mathop{}\!\mathrm{d}V_{g}=\int\phi(2\sigma_{2}-4v^{2}-2S_{v})\mathop{}\!\mathrm{d}V_{g},
vdϕ,dvdVg=ϕ(2v2σ2+Sv)dVg.\int v\langle d\phi,dv\rangle\mathop{}\!\mathrm{d}V_{g}=\int\phi(2v^{2}-\sigma_{2}+S_{v})\mathop{}\!\mathrm{d}V_{g}.

5.4. Second variation cross terms

Proposition 5.6.

Let M=b=1BSb2×N~M=\prod_{b=1}^{B}S^{2}_{b}\times\tilde{N} and vker(Δ+2)v\in\ker(\Delta+2). Let u=(Δ+1)u~,hu=(\Delta+1)\tilde{u},h be the unique solution of 𝒟Φ(ug+h)=12𝒟2Φ(vg,vg)\mathcal{D}\Phi(ug+h)=-\frac{1}{2}\mathcal{D}^{2}\Phi(vg,vg) as in Lemma 5.4. Then we have

(5.9) 𝒟2Φ(vg,ug),vgL2=2235n39866n2+14364n6888675(3n4)(5n6)σ487n3265n2+186n+24108n(3n4)σ22,\begin{split}\langle\mathcal{D}^{2}\Phi(vg,ug),vg\rangle_{L^{2}}=&\frac{2235n^{3}-9866n^{2}+14364n-6888}{675(3n-4)(5n-6)}\sigma_{4}\\ &-\frac{87n^{3}-265n^{2}+186n+24}{108n(3n-4)}\sigma_{2}^{2},\end{split}
(5.10) 𝒟2Φ(vg,h),vgL2=(n2)2(41n2+40n104)180(3n4)(5n6)σ4+(n2)2(29n359n+90n72)72n(3n4)(5n6)σ22.\begin{split}\langle\mathcal{D}^{2}\Phi(vg,h),vg\rangle_{L^{2}}=&-\frac{(n-2)^{2}(41n^{2}+40n-104)}{180(3n-4)(5n-6)}\sigma_{4}\\ &+\frac{(n-2)^{2}(29n^{3}-59n+90n-72)}{72n(3n-4)(5n-6)}\sigma_{2}^{2}.\end{split}
Proof.

Pair the second variation formula (4.10) with vgvg. Using Lemma 5.5 and the explicit forms (5.2, 5.7) of fssf_{ss} and u~\tilde{u}, every term in 𝒟2Φ(vg,ug),vgL2\langle\mathcal{D}^{2}\Phi(vg,ug),vg\rangle_{L^{2}} may be expanded as a product of the functions v2,Sv,σ2v^{2},S_{v},\sigma_{2}. All such products have integral computed in Lemma 5.1, and collecting terms gives the first equation.

By formula (4.8), we have 𝒟2Φ(vg,h),vgL2=n24(vvbhb)wdVg\langle\mathcal{D}^{2}\Phi(vg,h),vg\rangle_{L^{2}}=-\frac{n-2}{4}\int(\sum vv_{b}h_{b})w\mathop{}\!\mathrm{d}V_{g}. Now note that for any vker(Δ+2)v^{\prime}\in\ker(\Delta+2), the product vvbvvv_{b}v^{\prime} is a spherical polynomial of total degree 3, and hence vvbvdVg=0\int vv_{b}v^{\prime}\mathop{}\!\mathrm{d}V_{g}=0. Therefore we need only integrate against the explicit portion of the solution (5.8) for hbh_{b} - that is, against h~b\tilde{h}_{b} in the proof of Lemma 5.4.

Consider the basis functions v2,Sv,σ2,vvb,vb2,αb2v^{2},S_{v},\sigma_{2},vv_{b},v_{b}^{2},\alpha_{b}^{2}. Multiplying by vvbvv_{b} and summing over bb respectively give v4,v2Sv2,σ2v2,v2Sv,vvb3,σ2vvbv^{4},v^{2}S_{v}^{2},\sigma_{2}v^{2},v^{2}S_{v},v\sum v_{b}^{3},\sigma_{2}vv_{b}. Only the last two do not have integral listed in Lemma 5.1. For these we again note that spherical polynomials of odd degree integrate to 0, which implies v(vb3)dVg=vb4dVg\int v(\sum v_{b}^{3})\mathop{}\!\mathrm{d}V_{g}=\sum\int v_{b}^{4}\mathop{}\!\mathrm{d}V_{g} and vvbdVg=vb2dVg\int vv_{b}\mathop{}\!\mathrm{d}V_{g}=\int v_{b}^{2}\mathop{}\!\mathrm{d}V_{g}. Collecting terms then gives the second equation. ∎

5.5. Third variation

Proposition 5.7.

Let M=b=1BSb2×N~M=\prod_{b=1}^{B}S^{2}_{b}\times\tilde{N} and vker(Δ+2)v\in\ker(\Delta+2). Then

(5.11) 𝒟3Φ(vg,vg,vg),vgL2=457n21852n+660900σ423n395n2+36n+1218nσ22.\begin{split}\langle\mathcal{D}^{3}\Phi(vg,vg,vg),vg\rangle_{L^{2}}=&\frac{457n^{2}-1852n+660}{900}\sigma_{4}\\ &-\frac{23n^{3}-95n^{2}+36n+12}{18n}\sigma_{2}^{2}.\end{split}
Proof.

Again, using Lemma 5.5, every term in formula (4.7) for 𝒟3Φ(vg,vg,vg),vgL2\langle\mathcal{D}^{3}\Phi(vg,vg,vg),vg\rangle_{L^{2}} may be expanded as a product of the functions v2,Sv,σ2v^{2},S_{v},\sigma_{2}, which have integral computed in Lemma 5.1. Collecting terms gives the result. ∎

5.6. Formal obstruction

Combining the above formulae, we have

Theorem 5.8.

Let M=b=1BSb2×N~M=\prod_{b=1}^{B}S^{2}_{b}\times\tilde{N} and vker(Δ+2)v\in\ker(\Delta+2). Let u=(Δ+1)u~,hu=(\Delta+1)\tilde{u},h be the unique solution of 𝒟Φ(ug+h)=12𝒟2Φ(vg,vg)\mathcal{D}\Phi(ug+h)=-\frac{1}{2}\mathcal{D}^{2}\Phi(vg,vg). Then we have

(5.12) 𝒟3Φ(vg,vg,vg)+6𝒟2Φ(ug+h,vg),vgL2=Q4σ4+Q2σ22,\langle\mathcal{D}^{3}\Phi(vg,vg,vg)+6\mathcal{D}^{2}\Phi(ug+h,vg),vg\rangle_{L^{2}}=Q_{4}\sigma_{4}+Q_{2}\sigma_{2}^{2},

where

Q4=5625n423546n3+15316n2+28104n26784900(3n4)(5n6),Q_{4}=\frac{5625n^{4}-23546n^{3}+15316n^{2}+28104n-26784}{900(3n-4)(5n-6)},
Q2=603n53203n4+4384n3+108n23120n+115236n(3n4)(5n6).Q_{2}=-\frac{603n^{5}-3203n^{4}+4384n^{3}+108n^{2}-3120n+1152}{36n(3n-4)(5n-6)}.
Corollary 5.9.

Let M=b=1BSb2×N~M=\prod_{b=1}^{B}S^{2}_{b}\times\tilde{N} and suppose vker(Δ+2)v\in\ker(\Delta+2). Let u=(Δ+1)u~,hu=(\Delta+1)\tilde{u},h be the unique solution of 𝒟Φ(ug+h)=12𝒟2Φ(vg,vg)\mathcal{D}\Phi(ug+h)=-\frac{1}{2}\mathcal{D}^{2}\Phi(vg,vg). If n=4n=4, or if B=1B=1, then there exists δ>0\delta>0 such that

(5.13) π𝒦1(𝒟3Φ(vg,vg,vg)+6𝒟2Φ(ug+h,vg))L2δvL23.\|\pi_{\mathcal{K}_{1}}(\mathcal{D}^{3}\Phi(vg,vg,vg)+6\mathcal{D}^{2}\Phi(ug+h,vg))\|_{L^{2}}\geq\delta\|v\|_{L^{2}}^{3}.
Proof.

Recall that vL22=v2wdVg=σ23\|v\|_{L^{2}}^{2}=\int v^{2}w\mathop{}\!\mathrm{d}V_{g}=\frac{\sigma_{2}}{3}.

First assume n=4n=4. Then have explicitly Q4=41211575Q_{4}=\frac{4121}{1575} and Q2=535126Q_{2}=-\frac{535}{126}. By the trivial inequality σ4σ22\sigma_{4}\leq\sigma_{2}^{2} we then have

Q4σ4+Q2σ22(Q4+Q2)σ22=17111050σ22<0.Q_{4}\sigma_{4}+Q_{2}\sigma_{2}^{2}\leq(Q_{4}+Q_{2})\sigma_{2}^{2}=-\frac{1711}{1050}\sigma_{2}^{2}<0.

Now assume B=1B=1. Then σ4=σ22\sigma_{4}=\sigma_{2}^{2}, and

Q4+Q2=1050n44881n3+3968n2+2468n2400300n(5n6).Q_{4}+Q_{2}=-\frac{1050n^{4}-4881n^{3}+3968n^{2}+2468n-2400}{300n(5n-6)}.

Note that the numerator has no integer roots.

Therefore, in either case,

|𝒟3Φ(vg,vg,vg)+6𝒟2Φ(ug+h,vg),vgL2|δvL24,|\langle\mathcal{D}^{3}\Phi(vg,vg,vg)+6\mathcal{D}^{2}\Phi(ug+h,vg),vg\rangle_{L^{2}}|\geq\delta\|v\|_{L^{2}}^{4},

which implies the result. ∎

Note that one may replace L2L^{2} in (5.13) with any norm since 𝒦1\mathcal{K}_{1} is finite-dimensional.

6. Hölder theory

In this section, we use Taylor expansion to turn the formal obstruction into the quantitative rigidity Theorem 1.3. Throughout this section (M,g)(M,g) will denote the 1-Einstein manifold S2×NS^{2}\times N, where either N=S2N=S^{2} or NN satisfies (\dagger).

Recall that 𝒦\mathcal{K} is the space of Jacobi deformations, and 𝒦1={vg|(Δ+2μ)v=0}\mathcal{K}_{1}=\{vg|(\Delta+2\mu)v=0\} is the subspace of Jacobi deformations modulo symmetries. We use T1T2T_{1}*T_{2} to denote an unspecified contraction of the tensors T1,T2T_{1},T_{2}. We will freely use that Hölder norms behave well under contractions, i.e. T1T2Ck,αT1Ck,αT2Ck,α\|T_{1}*T_{2}\|_{C^{k,\alpha}}\leq\|T_{1}\|_{C^{k,\alpha}}\|T_{2}\|_{C^{k,\alpha}}.

6.1. Taylor expansion

In the sequel, we would like to apply Taylor expansion to Φ(s):=Φ(g+sh)\Phi(s):=\Phi(g+sh). To do so, we need a priori bounds for 𝒟kΦ\mathcal{D}^{k}\Phi in terms of a finite number of derivatives of gg. However, this is not as simple as in [SZ20], since the explicit form of Φ\Phi depends not only on the 2-jet of gg, but also on the implicitly defined τg\tau_{g} and (the 2-jet of) fgf_{g}.

The goal is to show that 𝒟kν\mathcal{D}^{k}\nu, 𝒟kτ\mathcal{D}^{k}\tau, 𝒟kf\mathcal{D}^{k}f and 𝒟kΦ\mathcal{D}^{k}\Phi are uniformly bounded as kk-linear operators on C2,α(𝒮2(M))C^{2,\alpha}(\mathcal{S}^{2}(M)) with values in \mathbb{R}, \mathbb{R}, C2,α(M)C^{2,\alpha}(M) and C0,α(𝒮2(M))C^{0,\alpha}(\mathcal{S}^{2}(M)) respectively. Explicitly, let

𝒟kνg:=suphiC2,α1|𝒟kνg(h1,,hk)|,\|\mathcal{D}^{k}\nu_{g}\|:=\sup_{\|h_{i}\|_{C^{2,\alpha}}\leq 1}|\mathcal{D}^{k}\nu_{g}(h_{1},\cdots,h_{k})|,
𝒟kτg:=suphiC2,α1|𝒟kτg(h1,,hk)|,\|\mathcal{D}^{k}\tau_{g}\|:=\sup_{\|h_{i}\|_{C^{2,\alpha}}\leq 1}|\mathcal{D}^{k}\tau_{g}(h_{1},\cdots,h_{k})|,
𝒟kfgC2,α:=suphiC2,α1𝒟kfg(h1,,hk)C2,α,\|\mathcal{D}^{k}f_{g}\|_{C^{2,\alpha}}:=\sup_{\|h_{i}\|_{C^{2,\alpha}}\leq 1}\|\mathcal{D}^{k}f_{g}(h_{1},\cdots,h_{k})\|_{C^{2,\alpha}},
𝒟kΦgC0,α:=suphiC2,α1𝒟kΦg(h1,,hk)C0,α.\|\mathcal{D}^{k}\Phi_{g}\|_{C^{0,\alpha}}:=\sup_{\|h_{i}\|_{C^{2,\alpha}}\leq 1}\|\mathcal{D}^{k}\Phi_{g}(h_{1},\cdots,h_{k})\|_{C^{0,\alpha}}.

Then we have

Lemma 6.1.

Fix a reference soliton metric Φ(g0)=0\Phi(g_{0})=0. There exists Ck,ϵ>0C_{k},\epsilon>0 such that if gg0C2,α<ϵ\|g-g_{0}\|_{C^{2,\alpha}}<\epsilon, then for any kk the variations at gg satisfy

(6.1) 𝒟kνg+𝒟kτg+𝒟kfgC2,α+𝒟kΦgC0,αCk,\|\mathcal{D}^{k}\nu_{g}\|+\|\mathcal{D}^{k}\tau_{g}\|+\|\mathcal{D}^{k}f_{g}\|_{C^{2,\alpha}}+\|\mathcal{D}^{k}\Phi_{g}\|_{C^{0,\alpha}}\leq C_{k},

The proof of Lemma 6.1 is deferred to Appendix B, and uses an inductive process similar to Section 4. We now proceed with our Taylor expansion:

Lemma 6.2.

Fix a metric gg. There exist C,ϵ>0C,\epsilon>0 so that if hC2,α<ϵ\|h\|_{C^{2,\alpha}}<\epsilon, then

(6.2) Φ(g+h)Φ(g)j=1k1j!𝒟jΦ(h,,h)C0,αChC2,αk+1.\left\|\Phi(g+h)-\Phi(g)-\sum_{j=1}^{k}\frac{1}{j!}\mathcal{D}^{j}\Phi(h,\cdots,h)\right\|_{C^{0,\alpha}}\leq C\|h\|_{C^{2,\alpha}}^{k+1}.
Proof.

Set Φ(s)=Φ(g+sh)\Phi(s)=\Phi(g+sh), so that dkdskΦ(s)=(𝒟kΦ)g+sh(h,,h)\frac{d^{k}}{ds^{k}}\Phi(s)=(\mathcal{D}^{k}\Phi)_{g+sh}(h,\cdots,h). We then apply Taylor’s theorem with remainder (in integral form) pointwise to Φ(s)\Phi(s) and Φ(s)|xΦ(s)|y\Phi(s)|_{x}-\Phi(s)|_{y} to get the result. Note that Lemma 6.1 gives uniform bounds on 𝒟k+1ΦC0,α\|\mathcal{D}^{k+1}\Phi\|_{C^{0,\alpha}} in a neighbourhood of gg, which controls the remainder. ∎

6.2. Hölder estimate

Here we first prove a quantitative rigidity for deformations modulo symmetries as follows:

Theorem 6.3.

Let M=S2×NM=S^{2}\times N, where N=S2N=S^{2} or NN satisfies (\dagger). There exists C,ε1>0C,\varepsilon_{1}>0 such that if h𝒮̊g2(Δ+1)C̊ggh\in\mathring{\mathcal{S}}^{2}_{g}\oplus(\Delta+1)\mathring{C}^{\infty}_{g}\cdot g and hC2,αε1\|h\|_{C^{2,\alpha}}\leq\varepsilon_{1}, then

(6.3) hC2,α3CΦ(g+h)C0,α.\|h\|_{C^{2,\alpha}}^{3}\leq C\|\Phi(g+h)\|_{C^{0,\alpha}}.

The proof of Theorem 6.3 is somewhat formal and proceeds as in [SZ20]. We reproduce the strategy here in detail for those unfamiliar with the setting there.

We will repeatedly use the following estimate, which takes advantage of the fact that 𝒦1\mathcal{K}_{1} is finite-dimensional.

Lemma 6.4.

For any h𝒮2(M)h\in\mathcal{S}^{2}(M), we have

(6.4) π𝒦1(h)C2,αCπ𝒦1(h)L2ChL2C′′hC0,α.\|\pi_{\mathcal{K}_{1}}(h)\|_{C^{2,\alpha}}\leq C\|\pi_{\mathcal{K}_{1}}(h)\|_{L^{2}}\leq C\|h\|_{L^{2}}\leq C^{\prime\prime}\|h\|_{C^{0,\alpha}}.
Proof.

The first two inequalities are the equivalence of norms on a finite-dimensional space. Pythagoras theorem in L2L^{2} space implies that π𝒦1(h)L2hL2.\|\pi_{\mathcal{K}_{1}}(h)\|_{L^{2}}\leq\|h\|_{L^{2}}. The last inequality is just integration as MM has finite volume.

We also need the following elliptic Schauder estimate:

Lemma 6.5.

There exists CC so that if h𝒮2(M)h\in\mathcal{S}^{2}(M), then

hπ𝒦(h)C2,αCLhC0,α.\|h-\pi_{\mathcal{K}}(h)\|_{C^{2,\alpha}}\leq C\|Lh\|_{C^{0,\alpha}}.

In what follows we consider a soliton metric Φ(g)=0\Phi(g)=0, and recall that L=𝒟ΦL=\mathcal{D}\Phi. Also, we assume hC2,α<ϵ\|h\|_{C^{2,\alpha}}<\epsilon for small enough ϵ1\epsilon\ll 1 so that Φ(g+h)C0,α<1\|\Phi(g+h)\|_{C^{0,\alpha}}<1. We will frequently use this to absorb terms of higher degree, and also freely use Young’s inequality.

6.2.1. 1st order analysis

By Lemma 6.2 with k=1k=1 we have

Φ(g+h)Φ(g)𝒟Φ(h)C0,αChC2,α2.\|\Phi(g+h)-\Phi(g)-\mathcal{D}\Phi(h)\|_{C^{0,\alpha}}\leq C\|h\|^{2}_{C^{2,\alpha}}.

We consider h𝒮̊g2(Δ+1)C̊ggh\in\mathring{\mathcal{S}}^{2}_{g}\oplus(\Delta+1)\mathring{C}^{\infty}_{g}\cdot g, and decompose h=h1+k1h=h_{1}+k_{1}, where k1𝒦1k_{1}\in\mathcal{K}_{1} and so h1𝒮̊g2h_{1}\in\mathring{\mathcal{S}}^{2}_{g}. Note that indeed k1=vg=π𝒦(h)=π𝒦1(h)k_{1}=vg=\pi_{\mathcal{K}}(h)=\pi_{\mathcal{K}_{1}}(h) by the assumption on hh. The above then becomes

(6.5) Φ(g+h)LhC0,αChC2,α2.\|\Phi(g+h)-Lh\|_{C^{0,\alpha}}\leq C\|h\|^{2}_{C^{2,\alpha}}.

This allows us to show that the residual h1h_{1} is of higher order again:

Lemma 6.6.

There exist C,ϵ>0C,\epsilon>0 so that if hC2,α<ϵ\|h\|_{C^{2,\alpha}}<\epsilon and h1h_{1} is as above, then

(6.6) hC2,αC(hC0,α+Φ(g+h)C0,α),\|h\|_{C^{2,\alpha}}\leq C(\|h\|_{C^{0,\alpha}}+\|\Phi(g+h)\|_{C^{0,\alpha}}),
(6.7) h1C2,αC(hC0,α2+Φ(g+h)C0,α).\|h_{1}\|_{C^{2,\alpha}}\leq C(\|h\|^{2}_{C^{0,\alpha}}+\|\Phi(g+h)\|_{C^{0,\alpha}}).
Proof.

By Lemma 6.5 and (6.5), we have

(6.8) h1C2,αC𝒟Φ(h1)C0,αC(hC2,α2+Φ(g+h)C0,α).\|h_{1}\|_{C^{2,\alpha}}\leq C\|\mathcal{D}\Phi(h_{1})\|_{C^{0,\alpha}}\leq C(\|h\|^{2}_{C^{2,\alpha}}+\|\Phi(g+h)\|_{C^{0,\alpha}}).

Now by the triangle inequality

(6.9) hC2,αh1C2,α+vC2,αC(hC0,α+hC2,α2+Φ(g+h)C0,α).\|h\|_{C^{2,\alpha}}\leq\|h_{1}\|_{C^{2,\alpha}}+\|v\|_{C^{2,\alpha}}\leq C(\|h\|_{C^{0,\alpha}}+\|h\|_{C^{2,\alpha}}^{2}+\|\Phi(g+h)\|_{C^{0,\alpha}}).

Absorbing the higher power of hC2,α<ϵ1\|h\|_{C^{2,\alpha}}<\epsilon\ll 1 into the left hand side gives the first inequality. Substituting back in to (6.8) and absorbing higher order terms again gives the second inequality. ∎

6.2.2. 2nd order analysis

By Lemma 6.2 with k=2k=2 we have

Φ(g+h)Φ(g)𝒟Φ(h)12𝒟2Φ(h,h)C0,αChC2,α3.\|\Phi(g+h)-\Phi(g)-\mathcal{D}\Phi(h)-\frac{1}{2}\mathcal{D}^{2}\Phi(h,h)\|_{C^{0,\alpha}}\leq C\|h\|^{3}_{C^{2,\alpha}}.

Expanding 𝒟2Φ\mathcal{D}^{2}\Phi by our decomposition of hh and using the uniform estimates on 𝒟2Φ\mathcal{D}^{2}\Phi,

Φ(g+h)Lh112𝒟2Φ(vg,vg)C0,αC(hC2,α3+h1C2,α2+h1C2,αvC2,α).\|\Phi(g+h)-Lh_{1}-\frac{1}{2}\mathcal{D}^{2}\Phi(vg,vg)\|_{C^{0,\alpha}}\leq C(\|h\|^{3}_{C^{2,\alpha}}+\|h_{1}\|_{C^{2,\alpha}}^{2}+\|h_{1}\|_{C^{2,\alpha}}\|v\|_{C^{2,\alpha}}).

As discussed in Section 5.2.3, there is a unique solution L(ug+k2)=12𝒟2Φ(vg,vg)L(ug+k_{2})=-\frac{1}{2}\mathcal{D}^{2}\Phi(vg,vg), where u=(Δ+1)u~u=(\Delta+1)\tilde{u}, u~C̊g\tilde{u}\in\mathring{C}^{\infty}_{g}, u~ker(Δ+2)\tilde{u}\perp\ker(\Delta+2) and k2𝒮̊g2k_{2}\in\mathring{\mathcal{S}}^{2}_{g}. Then we have the further decomposition h1=ug+k2+h2h_{1}=ug+k_{2}+h_{2}, and the above becomes

(6.10) Φ(g+h)Lh2C0,αC(hC2,α3+h1C2,α2+h1C2,αvC2,α).\|\Phi(g+h)-Lh_{2}\|_{C^{0,\alpha}}\leq C(\|h\|^{3}_{C^{2,\alpha}}+\|h_{1}\|_{C^{2,\alpha}}^{2}+\|h_{1}\|_{C^{2,\alpha}}\|v\|_{C^{2,\alpha}}).

This allows us to show that the residual h2h_{2} is of higher order:

Lemma 6.7.

There exist C,ϵ>0C,\epsilon>0 so that if hC2,α<ϵ\|h\|_{C^{2,\alpha}}<\epsilon and h2h_{2} is as above, then

(6.11) h2C2,αC(hC0,α3+Φ(g+h)C0,α).\|h_{2}\|_{C^{2,\alpha}}\leq C(\|h\|^{3}_{C^{0,\alpha}}+\|\Phi(g+h)\|_{C^{0,\alpha}}).
Proof.

Using Lemma 6.5, and (6.10), we have

h2C2,α\displaystyle\|h_{2}\|_{C^{2,\alpha}} \displaystyle\leq CLh2C0,αC(Φ(g+h)C0,α+hC2,α3+h1C2,α2+h1C2,αvC2,α)\displaystyle C\|Lh_{2}\|_{C^{0,\alpha}}\leq C^{\prime}(\|\Phi(g+h)\|_{C^{0,\alpha}}+\|h\|^{3}_{C^{2,\alpha}}+\|h_{1}\|_{C^{2,\alpha}}^{2}+\|h_{1}\|_{C^{2,\alpha}}\|v\|_{C^{2,\alpha}})
\displaystyle\leq C′′(Φ(g+h)C0,α+Φ(g+h)C0,α3+hC0,α3)\displaystyle C^{\prime\prime}(\|\Phi(g+h)\|_{C^{0,\alpha}}+\|\Phi(g+h)\|_{C^{0,\alpha}}^{3}+\|h\|^{3}_{C^{0,\alpha}})
+C′′(Φ(g+h)C0,α2+hC0,α4+Φ(g+h)C0,αhC0,α+hC0,α3)\displaystyle+C^{\prime\prime}(\|\Phi(g+h)\|_{C^{0,\alpha}}^{2}+\|h\|^{4}_{C^{0,\alpha}}+\|\Phi(g+h)\|_{C^{0,\alpha}}\|h\|_{C^{0,\alpha}}+\|h\|^{3}_{C^{0,\alpha}}) .

In the last line we have used Lemma 6.6 for the h,h1h,h_{1} terms and Lemma 6.4 to estimate vC2,αChC0,α\|v\|_{C^{2,\alpha}}\leq C\|h\|_{C^{0,\alpha}}. Absorbing higher order terms then yields the result. ∎

We also record estimates on u,k2u,k_{2}:

Lemma 6.8.

There exist C,ϵ>0C,\epsilon>0 so that if hC2,α<ϵ\|h\|_{C^{2,\alpha}}<\epsilon and u,k2u,k_{2} are as above, then

(6.12) uC2,α+k2C2,αChC0,α2,\|u\|_{C^{2,\alpha}}+\|k_{2}\|_{C^{2,\alpha}}\leq C\|h\|_{C^{0,\alpha}}^{2},
Proof.

By the elliptic estimate Lemma 6.5, we have

uC2,α+k2C2,αCL(ug+k2)C2,αC𝒟2Φ(vg,vg)C0,α.\|u\|_{C^{2,\alpha}}+\|k_{2}\|_{C^{2,\alpha}}\leq C\|L(ug+k_{2})\|_{C^{2,\alpha}}\leq C^{\prime}\|\mathcal{D}^{2}\Phi(vg,vg)\|_{C^{0,\alpha}}.

The uniform bounds on 𝒟2Φ\mathcal{D}^{2}\Phi and Lemma 6.4 then give

𝒟2Φ(vg,vg)C0,αCvC2,α2ChC0,α2.\|\mathcal{D}^{2}\Phi(vg,vg)\|_{C^{0,\alpha}}\leq C\|v\|_{C^{2,\alpha}}^{2}\leq C^{\prime}\|h\|^{2}_{C^{0,\alpha}}.

6.2.3. 3rd order analysis

By Lemma 6.2 with k=3k=3, we have

Φ(g+h)Φ(g)𝒟Φ(h)12𝒟2Φ(h,h)16𝒟3Φ(h,h,h)C0,αChC2,α4.\|\Phi(g+h)-\Phi(g)-\mathcal{D}\Phi(h)-\frac{1}{2}\mathcal{D}^{2}\Phi(h,h)-\frac{1}{6}\mathcal{D}^{3}\Phi(h,h,h)\|_{C^{0,\alpha}}\leq C\|h\|^{4}_{C^{2,\alpha}}.

Using our decomposition h=vg+h1h=vg+h_{1}, h1=ug+k2+h2h_{1}=ug+k_{2}+h_{2} this becomes

Φ(g+h)Lh212𝒟2Φ(h1,h1)𝒟2Φ(h1,vg)16𝒟3Φ(h,h,h)C0,αChC2,α4.\|\Phi(g+h)-Lh_{2}-\frac{1}{2}\mathcal{D}^{2}\Phi(h_{1},h_{1})-\mathcal{D}^{2}\Phi(h_{1},vg)-\frac{1}{6}\mathcal{D}^{3}\Phi(h,h,h)\|_{C^{0,\alpha}}\leq C\|h\|^{4}_{C^{2,\alpha}}.

By the uniform estimates on 𝒟kΦ\mathcal{D}^{k}\Phi, after expanding further we get

(6.13) Φ(g+h)Lh2𝒟2Φ(ug+k2,vg)16𝒟3Φ(vg,vg,vg)C0,αC(hC2,α4+h1C2,α2+h2C2,αvC2,α)+C(vC2,α2h1C2,α+vC2,αh1C2,α2+h1C2,α3).\begin{split}\|\Phi(g+h)-Lh_{2}&-\mathcal{D}^{2}\Phi(ug+k_{2},vg)-\frac{1}{6}\mathcal{D}^{3}\Phi(vg,vg,vg)\|_{C^{0,\alpha}}\\ &\leq C(\|h\|^{4}_{C^{2,\alpha}}+\|h_{1}\|_{C^{2,\alpha}}^{2}+\|h_{2}\|_{C^{2,\alpha}}\|v\|_{C^{2,\alpha}})\\ &\quad+C(\|v\|_{C^{2,\alpha}}^{2}\|h_{1}\|_{C^{2,\alpha}}+\|v\|_{C^{2,\alpha}}\|h_{1}\|_{C^{2,\alpha}}^{2}+\|h_{1}\|_{C^{2,\alpha}}^{3}).\end{split}
Proof of Theorem 6.3.

Decompose h=vg+ug+k2+h2h=vg+ug+k_{2}+h_{2} as above. By the formal obstruction Corollary 5.9 and equivalence of norms on the finite-dimensional space 𝒦1𝒦\mathcal{K}_{1}\subset\mathcal{K}, we have

(6.14) vC2,α3Cπ𝒦(𝒟2Φ(ug+h,vg)+16𝒟3Φ(vg,vg,vg))C2,αLh2+𝒟2Φ(ug+k2,vg)+16𝒟3Φ(vg,vg,vg)C0,α.\begin{split}\|v\|_{C^{2,\alpha}}^{3}&\leq C\|\pi_{\mathcal{K}}(\mathcal{D}^{2}\Phi(ug+h,vg)+\frac{1}{6}\mathcal{D}^{3}\Phi(vg,vg,vg))\|_{C^{2,\alpha}}\\ &\leq\|Lh_{2}+\mathcal{D}^{2}\Phi(ug+k_{2},vg)+\frac{1}{6}\mathcal{D}^{3}\Phi(vg,vg,vg)\|_{C^{0,\alpha}}.\end{split}

In the second line we have used Lemma 6.4 and that π𝒦(Lh2)=0\pi_{\mathcal{K}}(Lh_{2})=0 (by self-adjointness). Then by the triangle inequality and (6.13), we have

(6.15) vC2,α3C(Φ(g+h)C0,α+hC2,α4+h1C2,α2+h2C2,αvC2,α)+C(vC2,α2h1C2,α+vC2,αh1C2,α2+h1C2,α3).\begin{split}\|v\|_{C^{2,\alpha}}^{3}\leq&C(\|\Phi(g+h)\|_{C^{0,\alpha}}+\|h\|^{4}_{C^{2,\alpha}}+\|h_{1}\|_{C^{2,\alpha}}^{2}+\|h_{2}\|_{C^{2,\alpha}}\|v\|_{C^{2,\alpha}})\\ &\quad+C(\|v\|_{C^{2,\alpha}}^{2}\|h_{1}\|_{C^{2,\alpha}}+\|v\|_{C^{2,\alpha}}\|h_{1}\|_{C^{2,\alpha}}^{2}+\|h_{1}\|_{C^{2,\alpha}}^{3}).\end{split}

Using Lemmas 6.4, 6.6 and 6.7 this gives that

1CvC2,α3Φ(g+h)C0,α+Φ(g+h)C0,α4+hC0,α4+Φ(g+h)C0,α2+hC0,α4+hC0,α(Φ(g+h)C0,α+hC0,α3)+hC0,α2(Φ(g+h)C0,α+hC0,α2)+hC0,α(Φ(g+h)C0,α2+hC0,α4)+Φ(g+h)C0,α3+hC0,α6.\begin{split}\frac{1}{C}\|v\|_{C^{2,\alpha}}^{3}\leq&\|\Phi(g+h)\|_{C^{0,\alpha}}+\|\Phi(g+h)\|^{4}_{C^{0,\alpha}}+\|h\|^{4}_{C^{0,\alpha}}+\|\Phi(g+h)\|^{2}_{C^{0,\alpha}}+\|h\|^{4}_{C^{0,\alpha}}\\ &+\|h\|_{C^{0,\alpha}}(\|\Phi(g+h)\|_{C^{0,\alpha}}+\|h\|^{3}_{C^{0,\alpha}})+\|h\|_{C^{0,\alpha}}^{2}(\|\Phi(g+h)\|_{C^{0,\alpha}}+\|h\|^{2}_{C^{0,\alpha}})\\ &+\|h\|_{C^{0,\alpha}}(\|\Phi(g+h)\|^{2}_{C^{0,\alpha}}+\|h\|^{4}_{C^{0,\alpha}})+\|\Phi(g+h)\|^{3}_{C^{0,\alpha}}+\|h\|^{6}_{C^{0,\alpha}}.\end{split}

Absorbing higher order terms yields

(6.16) vC2,α3Φ(g+h)C0,α+hC0,α4.\|v\|_{C^{2,\alpha}}^{3}\leq\|\Phi(g+h)\|_{C^{0,\alpha}}+\|h\|^{4}_{C^{0,\alpha}}.

Finally, to estimate h=vg+ug+k2+h2h=vg+ug+k_{2}+h_{2}, we combine this with Lemmas 6.7 and 6.8 to find

(6.17) hC2,αvC2,α+ug+k2C2,α+h2C2,αC(Φ(g+h)C0,α13+hC0,α43+hC0,α2+hC0,α3+Φ(g+h)C0,α).\begin{split}\|h\|_{C^{2,\alpha}}&\leq\|v\|_{C^{2,\alpha}}+\|ug+k_{2}\|_{C^{2,\alpha}}+\|h_{2}\|_{C^{2,\alpha}}\\ &\leq C(\|\Phi(g+h)\|_{C^{0,\alpha}}^{\frac{1}{3}}+\|h\|^{\frac{4}{3}}_{C^{0,\alpha}}+\|h\|_{C^{0,\alpha}}^{2}+\|h\|_{C^{0,\alpha}}^{3}+\|\Phi(g+h)\|_{C^{0,\alpha}}).\end{split}

Absorbing higher order terms again gives the desired estimate

(6.18) hC2,αCΦ(g+h)C0,α13.\|h\|_{C^{2,\alpha}}\leq C\|\Phi(g+h)\|_{C^{0,\alpha}}^{\frac{1}{3}}.

The proof of the quantitative rigidity theorem follows by combining Theorem 6.3 and the slicing Lemma 2.2.

Proof of Theorem 1.3.

Let δ<ϵ1\delta<\epsilon_{1}, where ϵ1\epsilon_{1} is as in Theorem 6.3. Then by Lemma 2.2, there exist c(1δ,1+δ)c\in(1-\delta,1+\delta), ψDiff(M)\psi\in\operatorname*{Diff}(M) and h𝒮̊g2(Δ+1)C̊ggh\in\mathring{\mathcal{S}}^{2}_{g}\oplus(\Delta+1)\mathring{C}^{\infty}_{g}\cdot g so that cψg~=g+hc\psi^{*}\tilde{g}=g+h and hC2,α<ϵ1\|h\|_{C^{2,\alpha}}<\epsilon_{1}. By the symmetries of Φ\Phi, we have Φ(cψg~)C0,α(1+δ)Φ(g~)C0,α\|\Phi(c\psi^{*}\tilde{g})\|_{C^{0,\alpha}}\leq(1+\delta)\|\Phi(\tilde{g})\|_{C^{0,\alpha}}. The result then follows from Theorem 6.3.

Appendix A Variation of basic geometric quantities

In this section we compute various variation formulae for the geometric quantities Ricg,Hessg\operatorname{Ric}_{g},\operatorname{Hess}^{g} and their traces Rg,ΔgR_{g},\Delta^{g}.

A.1. First variation

Here we list the well-known first variations of the basic geometric quantities.

Lemma A.1.

Consider a manifold MM with a 1-parameter family of metrics g(s)g(s). Then at s=0s=0, if eie_{i} is an orthonormal frame, we have

(A.1) Ricij=Rijklgkl+12(ikgjk+jkgik+Rikgjk+RjkgikΔgijij(trgg).\operatorname{Ric}^{\prime}_{ij}=-R_{ijkl}g^{\prime}_{kl}+\frac{1}{2}(\nabla_{i}\nabla_{k}g^{\prime}_{jk}+\nabla_{j}\nabla_{k}g^{\prime}_{ik}+R_{ik}g^{\prime}_{jk}+R_{jk}g^{\prime}_{ik}-\Delta g^{\prime}_{ij}-\nabla_{i}\nabla_{j}(\operatorname{tr}_{g}g^{\prime}).
(A.2) R=2gijRij+divgdivgg+Ric,gΔ(trgg).R^{\prime}=-2g^{\prime}_{ij}R_{ij}+\operatorname{div}_{g}\operatorname{div}_{g}g^{\prime}+\langle\operatorname{Ric},g^{\prime}\rangle-\Delta(\operatorname{tr}_{g}g^{\prime}).
(A.3) (Hessϕ)ij=12(igjk+jgikkgij)kϕ.({\operatorname{Hess}^{\prime}}\phi)_{ij}=-\frac{1}{2}(\nabla_{i}g^{\prime}_{jk}+\nabla_{j}g^{\prime}_{ik}-\nabla_{k}g^{\prime}_{ij})\nabla_{k}\phi.
(A.4) Δϕ=gij(Hessϕ)ijdivgg,dϕ+12trgg,ϕ.\Delta^{\prime}\phi=-g^{\prime}_{ij}(\operatorname{Hess}\phi)_{ij}-\langle\operatorname{div}_{g}g^{\prime},d\phi\rangle+\frac{1}{2}\langle\nabla\operatorname{tr}_{g}g^{\prime},\nabla\phi\rangle.

A.2. Conformal direction

Here we consider higher order variations under conformal change.

Lemma A.2.

Let MM be a manifold with the metric family (1+sv)g(1+sv)g, where vv (and ϕ\phi) is a smooth function on MM. Then we have the variation formulae

Rics=n22Hessv12(Δv)g,Ricss=(n2)vHessv+3n22dvdv+(vΔvn42|dv|2)g,Ricsss=3(n2)v2Hessv9(n2)vdvdv(3v2Δv+3(n4)v|dv|2)g,\begin{split}\operatorname{Ric}_{s}&=-\frac{n-2}{2}\operatorname{Hess}v-\frac{1}{2}(\Delta v)g,\\ \operatorname{Ric}_{ss}&=(n-2)v\operatorname{Hess}v+3\frac{n-2}{2}dv\otimes dv+(v\Delta v-\frac{n-4}{2}|dv|^{2})g,\\ \operatorname{Ric}_{sss}&=-3(n-2)v^{2}\operatorname{Hess}v-9(n-2)vdv\otimes dv-(3v^{2}\Delta v+3(n-4)v|dv|^{2})g,\end{split}
Rs=vR(n1)Δv,Rss=2v2R+4(n1)vΔv(n1)(n6)2|dv|2,Rsss=6v3R18(n1)v2Δv+92(n1)(n6)v|dv|2,\begin{split}R_{s}&=-vR-(n-1)\Delta v,\\ R_{ss}&=2v^{2}R+4(n-1)v\Delta v-\frac{(n-1)(n-6)}{2}|dv|^{2},\\ R_{sss}&=-6v^{3}R-18(n-1)v^{2}\Delta v+\frac{9}{2}(n-1)(n-6)v|dv|^{2},\end{split}
Hesssϕ=12(dvdϕdϕdv+dϕ,dvg),Hessssϕ=v(dvdϕdϕdv+dϕ,dvg),Hesssssϕ=3v2(dvdϕdϕdv+dϕ,dvg),\begin{split}\operatorname{Hess}_{s}\phi&=\frac{1}{2}(-dv\otimes d\phi-d\phi\otimes dv+\langle d\phi,dv\rangle g),\\ \operatorname{Hess}_{ss}\phi&=-v(-dv\otimes d\phi-d\phi\otimes dv+\langle d\phi,dv\rangle g),\\ \operatorname{Hess}_{sss}\phi&=3v^{2}(-dv\otimes d\phi-d\phi\otimes dv+\langle d\phi,dv\rangle g),\end{split}
Δsϕ=vΔϕ+n22dv,dϕ,Δssϕ=2v2Δϕ2(n2)dϕ,dv,Δsssϕ=6v3Δϕ+9(n2)v2dϕ,dv.\begin{split}\Delta_{s}\phi&=-v\Delta\phi+\frac{n-2}{2}\langle dv,d\phi\rangle,\\ \Delta_{ss}\phi&=2v^{2}\Delta\phi-2(n-2)\langle d\phi,dv\rangle,\\ \Delta_{sss}\phi&=-6v^{3}\Delta\phi+9(n-2)v^{2}\langle d\phi,dv\rangle.\end{split}
Proof.

The relevant quantities are given explicitly under conformal variation g~=e2ψg\widetilde{g}=e^{2\psi}g by

Ric~=Ric(n2)(Hessψdψdψ)(Δψ+(n2)|dψ|2)g,\widetilde{\operatorname{Ric}}=\operatorname{Ric}-(n-2)(\operatorname{Hess}\psi-d\psi\otimes d\psi)-(\Delta\psi+(n-2)|d\psi|^{2})g,
R~=e2ψ(R2(n1)Δψ(n1)(n2)|dψ|2),\widetilde{R}=e^{-2\psi}(R-2(n-1)\Delta\psi-(n-1)(n-2)|d\psi|^{2}),
Hess~f=Hessfdψdfdfdψ+f,ψg,\widetilde{\operatorname{Hess}}f=\operatorname{Hess}f-d\psi\otimes df-df\otimes d\psi+\langle\nabla f,\nabla\psi\rangle g,
Δ~f=e2ψ(Δf+(n2)ψ,f).\widetilde{\Delta}f=e^{-2\psi}(\Delta f+(n-2)\langle\nabla\psi,\nabla f\rangle).

The above variation formulae follow by setting e2ψ=(1+sv)e^{2\psi}=(1+sv) and expanding in ss to third order. Note that ϕ=12log(1+sv)=s2vs24v2+s36v3\phi=\frac{1}{2}\log(1+sv)=\frac{s}{2}v-\frac{s^{2}}{4}v^{2}+\frac{s^{3}}{6}v^{3}-\cdots.

A.2.1. Second variation cross term

Lemma A.3.

Let MM be a manifold with the 2-parameter family of metrics (1+sv+tu)g(1+sv+tu)g, where u,vu,v are smooth functions on MM. Then we have

Ricst=3(n2)4(dudv+dvdu)+n22(uHessv+vHessu)+12(uΔv+vΔu)gn42du,dvg,\operatorname{Ric}_{st}=\frac{3(n-2)}{4}(du\otimes dv+dv\otimes du)+\frac{n-2}{2}(u\operatorname{Hess}v+v\operatorname{Hess}u)+\frac{1}{2}(u\Delta v+v\Delta u)g-\frac{n-4}{2}\langle du,dv\rangle g,
Hessstϕ=12(d(uv)dϕ+dϕd(uv)12dϕ,d(uv)g,\operatorname{Hess}_{st}\phi=\frac{1}{2}(d(uv)\otimes d\phi+d\phi\otimes d(uv)-\frac{1}{2}\langle d\phi,d(uv)\rangle g,
Rst=2uvR+2(n1)(uΔv+vΔu)(n1)(n6)2du,dv.R_{st}=2uvR+2(n-1)(u\Delta v+v\Delta u)-\frac{(n-1)(n-6)}{2}\langle du,dv\rangle.
Proof.

Polarise the second variation formulae in the previous lemma. ∎

A.3. Second variation in kernel-TT direction

Here we compute the basic variations needed in Section 4.3.1, by considering the 2-parameter variation (1+sv)(g+th)(1+sv)(g+th). This will allow us to use the variation formulae under conformal change, although now gst=vhg_{st}=vh so for any quantity QQ we will have Qst=𝒟2Q(vg,h)+𝒟Q(vh)Q_{st}=\mathcal{D}^{2}Q(vg,h)+\mathcal{D}Q(vh).

Lemma A.4.

Let (M,g)(M,g) be a closed 1-Einstein manifold, and vker(Δ+2)v\in\ker(\Delta+2) and h𝒮̊g2h\in\mathring{\mathcal{S}}^{2}_{g}. Then for any smooth function uu on MM, we have

𝒟2Ric(vg,h),ugL2=n22uh,HessvwdVg,\langle\mathcal{D}^{2}\operatorname{Ric}(vg,h),ug\rangle_{L^{2}}=\frac{n-2}{2}\int u\langle h,\operatorname{Hess}v\rangle w\mathop{}\!\mathrm{d}V_{g},
𝒟2R(vg,h)=(n2)h,Hessv.\mathcal{D}^{2}R(vg,h)=(n-2)\langle h,\operatorname{Hess}v\rangle.
Proof.

First, by the conformal variation Lemma A.2, at s=0s=0 we have

Rics=n22Hessg+thv12(Δg+thv)(g+th).\operatorname{Ric}_{s}=-\frac{n-2}{2}\operatorname{Hess}^{g+th}v-\frac{1}{2}(\Delta^{g+th}v)(g+th).

Since trh=0,divh=0\operatorname{tr}h=0,\operatorname{div}h=0, by the general first variation Lemma A.1 we have Hesst=0\operatorname{Hess}_{t}=0 and hence Δtϕ=h,Hessϕ\Delta_{t}\phi=-\langle h,\operatorname{Hess}\phi\rangle. So differentiating in tt now, at (s,t)=(0,0)(s,t)=(0,0) we have

Ricst=12h,Hessvg12(Δv)h.\operatorname{Ric}_{st}=\frac{1}{2}\langle h,\operatorname{Hess}v\rangle g-\frac{1}{2}(\Delta v)h.

This implies

Ricst,ugL2=u(n2h,Hessv12(Δv)trh)wdVg=n2uh,HessvwdVg.\langle\operatorname{Ric}_{st},ug\rangle_{L^{2}}=\int u\left(\frac{n}{2}\langle h,\operatorname{Hess}v\rangle-\frac{1}{2}(\Delta v)\operatorname{tr}h\right)w\mathop{}\!\mathrm{d}V_{g}=\frac{n}{2}\int u\langle h,\operatorname{Hess}v\rangle w\mathop{}\!\mathrm{d}V_{g}.

By Lemma A.1 again we have

𝒟Ric(vh),ugL2=vuRic,hwdVg+u(divdiv(vh)+vRic,hΔ(trvh))wdVg=12uh,HessvwdVg,\begin{split}\langle\mathcal{D}\operatorname{Ric}(vh),ug\rangle_{L^{2}}&=-\int vu\langle\operatorname{Ric},h\rangle w\mathop{}\!\mathrm{d}V_{g}+\int u(\operatorname{div}\operatorname{div}(vh)+v\langle\operatorname{Ric},h\rangle-\Delta(\operatorname{tr}vh))w\mathop{}\!\mathrm{d}V_{g}\\ &=\frac{1}{2}\int u\langle h,\operatorname{Hess}v\rangle w\mathop{}\!\mathrm{d}V_{g},\end{split}

where we used that the initial metric satisfies Ric=g\operatorname{Ric}=g. Therefore

𝒟2Ric(vg,h),ugL2=Ricst𝒟Ric(vh),ugL2=n22uh,HessvwdVg.\langle\mathcal{D}^{2}\operatorname{Ric}(vg,h),ug\rangle_{L^{2}}=\langle\operatorname{Ric}_{st}-\mathcal{D}\operatorname{Ric}(vh),ug\rangle_{L^{2}}=\frac{n-2}{2}\int u\langle h,\operatorname{Hess}v\rangle w\mathop{}\!\mathrm{d}V_{g}.

Now for any 2-tensor TT, in an orthonormal frame at (s,t)=(0,0)(s,t)=(0,0) we have

(trT)st=trg(Tst)gt,Tsgs,Ttgst,T+2(gs)ij(gt)ikTkj.(\operatorname{tr}T)_{st}=\operatorname{tr}_{g}(T_{st})-\langle g_{t},T_{s}\rangle-\langle g_{s},T_{t}\rangle-\langle g_{st},T\rangle+2(g_{s})_{ij}(g_{t})_{ik}T_{kj}.

In particular this gives

Rst=n2h,Hessv12(Δv)trhh,Δv2gn22HessvvtrRictvh,Ric+2vh,Ric.R_{st}=\frac{n}{2}\langle h,\operatorname{Hess}v\rangle-\frac{1}{2}(\Delta v)\operatorname{tr}h-\langle h,-\frac{\Delta v}{2}g-\frac{n-2}{2}\operatorname{Hess}v\rangle-v\operatorname{tr}\operatorname{Ric}_{t}-\langle vh,\operatorname{Ric}\rangle+2v\langle h,\operatorname{Ric}\rangle.

Again using that Ric=g\operatorname{Ric}=g and trh=0,divg=0\operatorname{tr}h=0,\operatorname{div}g=0, we get Rst=2n22h,HessvR_{st}=\frac{2n-2}{2}\langle h,\operatorname{Hess}v\rangle. By Lemma A.1 we have 𝒟R(vh)=divdiv(vh)=h,Hessv\mathcal{D}R(vh)=\operatorname{div}\operatorname{div}(vh)=\langle h,\operatorname{Hess}v\rangle, therefore

𝒟2R(vg,h)=Rst𝒟R(vh)=(n2)h,Hessv.\mathcal{D}^{2}R(vg,h)=R_{st}-\mathcal{D}R(vh)=(n-2)\langle h,\operatorname{Hess}v\rangle.

Appendix B Uniform bounds for shrinker quantities

Here we prove Lemma 6.1, which is reproduced below. Recall that T1T2T_{1}*T_{2} denotes an unspecified contraction of the tensors T1,T2T_{1},T_{2}, and T1T2Ck,αT1Ck,αT2Ck,α\|T_{1}*T_{2}\|_{C^{k,\alpha}}\leq\|T_{1}\|_{C^{k,\alpha}}\|T_{2}\|_{C^{k,\alpha}}.

Lemma B.1 (= Lemma 6.1).

Fix a reference soliton metric Φ(g0)=0\Phi(g_{0})=0. There exists Ck,ϵ>0C_{k},\epsilon>0 such that if gg0C2,α<ϵ\|g-g_{0}\|_{C^{2,\alpha}}<\epsilon, then for any kk the variations at gg satisfy

(B.1) 𝒟kνg+𝒟kτg+𝒟kfgC2,α+𝒟kΦgC0,αCk,\|\mathcal{D}^{k}\nu_{g}\|+\|\mathcal{D}^{k}\tau_{g}\|+\|\mathcal{D}^{k}f_{g}\|_{C^{2,\alpha}}+\|\mathcal{D}^{k}\Phi_{g}\|_{C^{0,\alpha}}\leq C_{k},
Proof.

We proceed by induction on kk. For brevity we abbreviate 𝐡=(h1,,hk)\mathbf{h}=(h_{1},\cdots,h_{k}) and 𝐡k=i=1khi\|\mathbf{h}\|^{k}=\prod_{i=1}^{k}\|h_{i}\| . The main idea is that we can estimate derivatives of τ,f\tau,f by differentiating their defining equations in the 𝐡\mathbf{h} direction and focussing on the highest order terms; all other terms will be estimated by the inductive hypothesis.

The case k=0k=0 is clear as each quantity depends smoothly on gg. So suppose that

𝒟jνg+𝒟jτg+𝒟jfgC2,α+𝒟jΦgC0,αCj\|\mathcal{D}^{j}\nu_{g}\|+\|\mathcal{D}^{j}\tau_{g}\|+\|\mathcal{D}^{j}f_{g}\|_{C^{2,\alpha}}+\|\mathcal{D}^{j}\Phi_{g}\|_{C^{0,\alpha}}\leq C_{j}

for all j<kj<k. For a tensor quantity TT we denote an unspecified linear combination of contractions of copies of 𝒟T\mathcal{D}^{\cdot}T with total order jj by

(𝒟)jT=j1++jα=j𝒟j1T𝒟jlT.(*\mathcal{D})^{j}T=\sum_{j_{1}+\cdots+j_{\alpha}=j}\mathcal{D}^{j_{1}}T*\cdots*\mathcal{D}^{j_{l}}T.

In particular, if j1+j2+j3=j<kj_{1}+j_{2}+j_{3}=j<k, 𝐡𝐡\mathbf{h}^{\prime}\subset\mathbf{h} then by the inductive hypothesis

((𝒟)j1τ(𝒟)j2f(𝒟)j3Φ)(𝐡)C0,αCj𝐡C2,αj.\left\|\left((*\mathcal{D})^{j_{1}}\tau*(*\mathcal{D})^{j_{2}}f*(*\mathcal{D})^{j_{3}}\Phi\right)(\mathbf{h}^{\prime})\right\|_{C^{0,\alpha}}\leq C_{j}\|\mathbf{h}^{\prime}\|^{j}_{C^{2,\alpha}}.

Recall that 𝒟νg(h)=Φ(g),hwgdVg\mathcal{D}\nu_{g}(h)=-\int\langle\Phi(g),h\rangle w^{g}\mathop{}\!\mathrm{d}V_{g}, where w=(4πτ)n/2efw=(4\pi\tau)^{-n/2}e^{-f}. Differentiating this further, we find that

𝒟kνg(𝐡)=wdVg(h1(k1+k2+k3=k1(𝒟)k1τ(𝒟)k2f(𝒟)k3Φ)(𝐡)),\mathcal{D}^{k}\nu_{g}(\mathbf{h})=-\int w\mathop{}\!\mathrm{d}V_{g}\left(h_{1}*\left(\sum_{k_{1}+k_{2}+k_{3}=k-1}(*\mathcal{D})^{k_{1}}\tau*(*\mathcal{D})^{k_{2}}f*(*\mathcal{D})^{k_{3}}\Phi\right)(\mathbf{h}^{\prime})\right),

where 𝐡=(h2,,hk)\mathbf{h}^{\prime}=(h_{2},\cdots,h_{k}). It follows immediately that |𝒟kνg(𝐡)|C𝐡C2,αk|\mathcal{D}^{k}\nu_{g}(\mathbf{h})|\leq C\|\mathbf{h}\|_{C^{2,\alpha}}^{k}.

By differentiating the normalisation wgdVg=1\int w^{g}\mathop{}\!\mathrm{d}V_{g}=1, we have

n2τ𝒟kτ(𝐡)𝒟kf(𝐡)wdVg=wdVg(k1+k2=kk1,k2<k(𝒟)k1τ(𝒟)k2f)(𝐡).-\frac{n}{2\tau}\mathcal{D}^{k}\tau(\mathbf{h})-\int\mathcal{D}^{k}f(\mathbf{h})w\mathop{}\!\mathrm{d}V_{g}=\int w\mathop{}\!\mathrm{d}V_{g}\left(\sum_{\begin{subarray}{c}k_{1}+k_{2}=k\\ k_{1},k_{2}<k\end{subarray}}(*\mathcal{D})^{k_{1}}\tau*(*\mathcal{D})^{k_{2}}f\right)(\mathbf{h}).

Similarly by differentiating fgwgdVg=n2+ν(g)\int f^{g}w^{g}\mathop{}\!\mathrm{d}V_{g}=\frac{n}{2}+\nu(g), we have

n2τ𝒟kτ(𝐡)+(1νn2)𝒟kf(𝐡)wdVg=𝒟kν(𝐡)+wdVg(k1+k2=kk1,k2<k(𝒟)k1τ(𝒟)k2f)(𝐡).\begin{split}-\frac{n}{2\tau}\mathcal{D}^{k}\tau(\mathbf{h})+&\left(1-\nu-\frac{n}{2}\right)\int\mathcal{D}^{k}f(\mathbf{h})w\mathop{}\!\mathrm{d}V_{g}\\ &=\mathcal{D}^{k}\nu(\mathbf{h})+\int w\mathop{}\!\mathrm{d}V_{g}\left(\sum_{\begin{subarray}{c}k_{1}+k_{2}=k\\ k_{1},k_{2}<k\end{subarray}}(*\mathcal{D})^{k_{1}}\tau*(*\mathcal{D})^{k_{2}}f\right)(\mathbf{h}).\end{split}

Eliminating 𝒟kf(𝐡)wdVg\int\mathcal{D}^{k}f(\mathbf{h})w\mathop{}\!\mathrm{d}V_{g}, we then have

n2+ν22τn𝒟kτ(𝐡)=𝒟kν(𝐡)+wdVg(k1+k2=kk1,k2<k(𝒟)k1τ(𝒟)k2f)(𝐡).\frac{\frac{n}{2}+\nu-2}{\frac{2\tau}{n}}\mathcal{D}^{k}\tau(\mathbf{h})=\mathcal{D}^{k}\nu(\mathbf{h})+\int w\mathop{}\!\mathrm{d}V_{g}\left(\sum_{\begin{subarray}{c}k_{1}+k_{2}=k\\ k_{1},k_{2}<k\end{subarray}}(*\mathcal{D})^{k_{1}}\tau*(*\mathcal{D})^{k_{2}}f\right)(\mathbf{h}).

It follows that |𝒟kτg(𝐡)|C𝐡C2,αk|\mathcal{D}^{k}\tau_{g}(\mathbf{h})|\leq C\|\mathbf{h}\|_{C^{2,\alpha}}^{k}.

Recall the defining equation (2.5) for ff, which states

τg(2Δgfg+|dfg|g2Rg)fg+n+ν(g)=0.\tau^{g}(-2\Delta^{g}f^{g}+|df^{g}|_{g}^{2}-R_{g})-f^{g}+n+\nu(g)=0.

All the basic geometric quantities involve at most the 2-jet of gg. Let Δf:=Δdf,d()\Delta_{f}:=\Delta-\langle df,d(\cdot)\rangle denote the drift Laplacian. By differentiating the defining equation, we will have

(Δf+12τ)(𝒟kf(𝐡))=12τ𝒟kν(𝐡)+12τ𝒟kτ(𝐡)(2Δf+|df|2R)+𝐡𝐡′′=𝐡#𝐡=k1#𝐡′′=k2(𝒥2)(𝐡)(j1+j2=k2j2<k(𝒟)j1τ(𝒟)j2𝒥2(f))(𝐡′′).\begin{split}(\Delta_{f}+\frac{1}{2\tau})\left(\mathcal{D}^{k}f(\mathbf{h})\right)=&\frac{1}{2\tau}\mathcal{D}^{k}\nu(\mathbf{h})+\frac{1}{2\tau}\mathcal{D}^{k}\tau(\mathbf{h})(-2\Delta f+|df|^{2}-R)\\ &+\sum_{\begin{subarray}{c}\mathbf{h}^{\prime}\cup\mathbf{h}^{\prime\prime}=\mathbf{h}\\ \#\mathbf{h}^{\prime}=k_{1}\\ \#\mathbf{h}^{\prime\prime}=k_{2}\end{subarray}}(*\mathcal{J}_{2})(\mathbf{h}^{\prime})*\left(\sum_{\begin{subarray}{c}j_{1}+j_{2}=k_{2}\\ j_{2}<k\end{subarray}}(*\mathcal{D})^{j_{1}}\tau*(*\mathcal{D})^{j_{2}}\mathcal{J}_{2}(f)\right)(\mathbf{h}^{\prime\prime}).\end{split}

Here 𝒥2(f)\mathcal{J}_{2}(f), 𝒥2(h)\mathcal{J}_{2}(h) denote the 2-jets of ff and hh (with respect to the reference metric g0g_{0} on MM), (𝒥2)(h1,,hj)=𝒥2(h1)𝒥2(hj)(*\mathcal{J}_{2})(h_{1},\cdots,h_{j})=\mathcal{J}_{2}(h_{1})\cdots*\mathcal{J}_{2}(h_{j}) and the sum is over all partitions of 𝐡\mathbf{h}.

By the previous estimates, we then have

(Δf+12τ)(𝒟kf(𝐡))C0,αC(|𝒟kν(𝐡)|+|𝒟kτ(𝐡)|)+C𝐡𝐡′′=𝐡#𝐡=k1#𝐡′′=k2<k𝐡C2,αk1(j1+j2=k2(𝒟)j1τ(𝒟)j2𝒥2(f))(𝐡′′)C0,αC𝐡C2,αk.\begin{split}\|(\Delta_{f}+\frac{1}{2\tau})&\left(\mathcal{D}^{k}f(\mathbf{h})\right)\|_{C^{0,\alpha}}\\ &\leq C(|\mathcal{D}^{k}\nu(\mathbf{h})|+|\mathcal{D}^{k}\tau(\mathbf{h})|)\\ &\quad+C\sum_{\begin{subarray}{c}\mathbf{h}^{\prime}\cup\mathbf{h}^{\prime\prime}=\mathbf{h}\\ \#\mathbf{h}^{\prime}=k_{1}\\ \#\mathbf{h}^{\prime\prime}=k_{2}<k\end{subarray}}\|\mathbf{h}^{\prime}\|^{k_{1}}_{C^{2,\alpha}}\left\|\left(\sum_{j_{1}+j_{2}=k_{2}}(*\mathcal{D})^{j_{1}}\tau*(*\mathcal{D})^{j_{2}}\mathcal{J}_{2}(f)\right)(\mathbf{h}^{\prime\prime})\right\|_{C^{0,\alpha}}\\ &\leq C^{\prime}\|\mathbf{h}\|^{k}_{C^{2,\alpha}}.\end{split}

For a shrinking soliton, it is known that 12τg0specΔfg0\frac{1}{2\tau^{g_{0}}}\notin\text{spec}\Delta^{g_{0}}_{f} (see for instance [CZ12]). Therefore, by continuity, for ϵ\epsilon small enough we will have 12τgspecΔfg\frac{1}{2\tau^{g}}\notin\text{spec}\Delta^{g}_{f}. In particular, (Δf+12τ)(\Delta_{f}+\frac{1}{2\tau}) is coercive (uniformly in gg), so by elliptic Schauder theory

𝒟kfg(𝐡)C2,αC(Δfg+12τg)(𝒟kfg(𝐡))C0,αC𝐡C2,αk.\|\mathcal{D}^{k}f_{g}(\mathbf{h})\|_{C^{2,\alpha}}\leq C\|(\Delta^{g}_{f}+\frac{1}{2\tau^{g}})\left(\mathcal{D}^{k}f_{g}(\mathbf{h})\right)\|_{C^{0,\alpha}}\leq C^{\prime}\|\mathbf{h}\|^{k}_{C^{2,\alpha}}.

Finally, note that again all the basic geometric quantities in the definition

Φ(g)=τg(Ric(g)+Hessgfg)g2\Phi(g)=\tau^{g}(\operatorname{Ric}(g)+\operatorname{Hess}^{g}f^{g})-\frac{g}{2}

depend on at most the 2-jet of gg. Then

𝒟kΦ(𝐡)=𝐡𝐡′′=𝐡#𝐡=k1#𝐡′′=k2(𝒥2)(𝐡)(j1+j2=k2(𝒟)j1τ(𝒟)j2f)(𝐡′′)+h1hk.\mathcal{D}^{k}\Phi(\mathbf{h})=\sum_{\begin{subarray}{c}\mathbf{h}^{\prime}\cup\mathbf{h}^{\prime\prime}=\mathbf{h}\\ \#\mathbf{h}^{\prime}=k_{1}\\ \#\mathbf{h}^{\prime\prime}=k_{2}\end{subarray}}(*\mathcal{J}_{2})(\mathbf{h}^{\prime})*\left(\sum_{j_{1}+j_{2}=k_{2}}(*\mathcal{D})^{j_{1}}\tau*(*\mathcal{D})^{j_{2}}f\right)(\mathbf{h}^{\prime\prime})+h_{1}*\cdots*h_{k}.

By the previous estimates we conclude that indeed 𝒟kΦg(𝐡)C0,αC𝐡C2,αk\|\mathcal{D}^{k}\Phi_{g}(\mathbf{h})\|_{C^{0,\alpha}}\leq C\|\mathbf{h}\|^{k}_{C^{2,\alpha}}, which completes the induction. ∎

References

  • [AM11] Lars Andersson and Vincent Moncrief, Einstein spaces as attractors for the Einstein flow, J. Differential Geom. 89 (2011), no. 1, 1–47. MR 2863911
  • [Bam20] Richard Bamler, Compactness theory of the space of super ricci flows, arXiv:2008.09298 (2020).
  • [BE69] M. Berger and D. Ebin, Some decompositions of the space of symmetric tensors on a Riemannian manifold, J. Differential Geometry 3 (1969), 379–392. MR 266084
  • [Ber66] M. Berger, Sur les variétés d’Einstein compactes, Comptes Rendus de la IIIe Réunion du Groupement des Mathématiciens d’Expression Latine (Namur, 1965), Librairie Universitaire, Louvain, 1966, pp. 35–55. MR 0238226
  • [Bes87] Arthur L. Besse, Einstein manifolds, Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)], vol. 10, Springer-Verlag, Berlin, 1987. MR 867684
  • [Bou99] Mohamed Boucetta, Spectre des laplaciens de Lichnerowicz sur les sphères et les projectifs réels, Publ. Mat. 43 (1999), no. 2, 451–483. MR 1744618
  • [Bou09] by same author, Spectra and symmetric eigentensors of the Lichnerowicz Laplacian on SnS^{n}, Osaka J. Math. 46 (2009), no. 1, 235–254. MR 2531148
  • [CCG+07] Bennett Chow, Sun-Chin Chu, David Glickenstein, Christine Guenther, James Isenberg, Tom Ivey, Dan Knopf, Peng Lu, Feng Luo, and Lei Ni, The Ricci flow: techniques and applications. Part I, Mathematical Surveys and Monographs, vol. 135, American Mathematical Society, Providence, RI, 2007, Geometric aspects. MR 2302600
  • [CH15] Huai-Dong Cao and Chenxu He, Linear stability of Perelman’s ν\nu-entropy on symmetric spaces of compact type, J. Reine Angew. Math. 709 (2015), 229–246. MR 3430881
  • [CIM15] Tobias Holck Colding, Tom Ilmanen, and William P. Minicozzi, II, Rigidity of generic singularities of mean curvature flow, Publ. Math. Inst. Hautes Études Sci. 121 (2015), 363–382. MR 3349836
  • [CMa] Tobias Holck Colding and William P. Minicozzi, II, Optimal growth bounds for eigenfunctions, in preparation.
  • [CMb] by same author, Singularities of ricci flow and diffeomorphisms, in preparation.
  • [CZ12] Huai-Dong Cao and Meng Zhu, On second variation of Perelman’s Ricci shrinker entropy, Math. Ann. 353 (2012), no. 3, 747–763. MR 2923948
  • [ELS20] Christopher G. Evans, Jason D. Lotay, and Felix Schulze, Remarks on the self-shrinking Clifford torus, J. Reine Angew. Math. 765 (2020), 139–170. MR 4129358
  • [Ham82] Richard S. Hamilton, Three-manifolds with positive Ricci curvature, J. Differential Geometry 17 (1982), no. 2, 255–306. MR 664497
  • [Ham88] by same author, The Ricci flow on surfaces, Mathematics and general relativity (Santa Cruz, CA, 1986), Contemp. Math., vol. 71, Amer. Math. Soc., Providence, RI, 1988, pp. 237–262. MR 954419
  • [Ino19] Eiji Inoue, The moduli space of Fano manifolds with Kähler-Ricci solitons, Adv. Math. 357 (2019), 106841, 65. MR 4017922
  • [IV15] Stefan Ivanov and Dimiter Vassilev, The lichnerowicz and obata first eigenvalue theorems and the obata uniqueness result in the yamabe problem on cr and quaternionic contact manifolds, arXiv:1504.03259 (2015).
  • [Ive93] Thomas Ivey, Ricci solitons on compact three-manifolds, Differential Geom. Appl. 3 (1993), no. 4, 301–307. MR 1249376
  • [Koi80] Norihito Koiso, Rigidity and stability of Einstein metrics—the case of compact symmetric spaces, Osaka Math. J. 17 (1980), no. 1, 51–73. MR 558319
  • [Krö15a] Klaus Kröncke, On infinitesimal Einstein deformations, Differential Geom. Appl. 38 (2015), 41–57. MR 3304669
  • [Krö15b] by same author, On the stability of Einstein manifolds, Ann. Global Anal. Geom. 47 (2015), no. 1, 81–98. MR 3302177
  • [Krö15c] by same author, Stability and instability of Ricci solitons, Calc. Var. Partial Differential Equations 53 (2015), no. 1-2, 265–287. MR 3336320
  • [Krö16] by same author, Rigidity and infinitesimal deformability of Ricci solitons, J. Geom. Anal. 26 (2016), no. 3, 1795–1807. MR 3511458
  • [Lic58] André Lichnerowicz, Géométrie des groupes de transformations, Travaux et Recherches Mathématiques, III. Dunod, Paris, 1958. MR 0124009
  • [SZ20] Ao Sun and Jonathan J. Zhu, Rigidity and łojasiewicz inequalities for clifford self-shrinkers, arXiv:2011.01636 (2020).
  • [Via14] Jeff A. Viaclovsky, Critical metrics for riemannian curvature functionals, arXiv:1405.6080 (2014).
  • [Zhu20] Jonathan J. Zhu, Łojasiewicz inequalities for cylindrical self-shrinkers by perturbation, In preparation (2020).