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

\DeclareSourcemap\maps

[datatype=bibtex] \map \step[fieldsource=pmid, fieldtarget=pubmed]

A gradient flow of Spin(7)-structures

Shubham Dwivedi
(March 15, 2025)
Abstract

We formulate and study the negative gradient flow of an energy functional of Spin(7)-structures on compact 88-manifolds. The energy functional is the L2L^{2}-norm of the torsion of the Spin(7)-structure. Our main result is the short-time existence and uniqueness of solutions to the flow. We also explain how this negative gradient flow is the most general flow of Spin(7)-structures. We also study solitons of the flow and prove a non-existence result for compact expanding solitons.

1.   Introduction

Given an 88-dimensional smooth manifold MM, a Spin(7)-structure on MM is a reduction of the structure group of the frame bundle Fr(M)\text{Fr}(M) to the Lie group Spin(7) SO(8)\subset\mathrm{SO}(8). While many 88-manifolds admit Spin(7)-structures, the existence of torsion-free Spin(7)-structures is a difficult problem. Given the success of geometric flows in proving the existence of special geometric structures on manifolds it is natural to attempt to tackle such questions using a suitable geometric flow of Spin(7)-structures.

Unlike the case of geometric flows of G2\mathrm{G}_{2}-structures, there have been fewer studies of flows of Spin(7)-structures. A flow of isometric/harmonic Spin(7)-structures was introduced and studied in detail in [DLE24]. The authors proved various analytic and geometric results for the flow. This flow can be seen as an instance of the more general theory of harmonic HH-structures which has been studied in a fairly detailed manner in [LS23] and [FLME22]. One disadvantage of the isometric flow is that it runs in an isometry class of a given Spin(7)-structure and hence the metric remains fixed throughout the flow. We remedy this here by studying the most general flow of Spin(7)-structures (in particular, the metric is also changing) which include the isometric flow as a special case. On a fixed oriented spin Riemannian 88-manifold, a Spin(7)-structure is equivalent (up to sign) to a unit spinor field. Using a spinorial approach, Ammann–Weiss–Witt [AWW16] also studied the negative gradient flow of a Dirichlet energy, thought of as a function of a unit spinor field. They proved general short-time existence and uniqueness. In contrast to their work, our approach is more direct and uses the Riemannian geometric properties of a Spin(7)-structure with the flow equation and related quantities all given in terms of the curvature of the underlying metric and the torsion of the Spin(7)-structure. We also describe all possible second order quasilinear flows of Spin(7)-structures. A study of some natural functionals of Spin(7)-structures and the corresponding Euler-Lagrange equations has also been done recently by Krasnov [Kra24].

Our point of view of studying general flows of Spin(7)-structures is that of [DGK23] where the authors study the most general second order flows of G2\mathrm{G}_{2}-structures by classifying all linearly independent second order (in the G2\mathrm{G}_{2}-structure) tensors coming from the Riemann curvature tensor and the covariant derivative of the torsion which can be made into a 33-form.

We now describe the energy functional and its negative gradient flow. More details are given in § 3.

Definition 1.1.

Let M8M^{8} be a compact manifold which admits Spin(7)-structures. The energy functional EE on the set of Spin(7)-structures on MM is

E(Φ)=12M|TΦ|2volΦ\displaystyle E(\Phi)=\frac{1}{2}\int_{M}|T_{\Phi}|^{2}\operatorname{vol}_{\Phi} (1.1)

where TΦT_{\Phi} is the torsion of the Spin(7)-structure Φ\Phi and the norm and the volume form are with respect to the metric induced by Φ\Phi. The critical points of the functional EE are torsion-free Spin(7)-structures which are also absolute minimizers.

We compute the first variation of EE in Lemma 3.3 which also gives the negative gradient of the energy functional. We describe the negative gradient flow below. See § 2 and § 3 for the definition of the terms and more details.

Definition 1.2 (Negative gradient flow).

Let (M8,Φ0)(M^{8},\Phi_{0}) be a compact manifold with a Spin(7)-structure. Let (TT)ij=8Tb;ilTj;lb8Tj;ilTb;lb+2Ti;lbTj;lb(T*T)_{ij}=8T_{b;il}T_{j;lb}-8T_{j;il}T_{b;lb}+2T_{i;lb}T_{j;lb}. The negative gradient flow of the energy functional EE is the flow of Spin(7)-structures which is the following initial value problem

{Φt=(Ric+2(T8g)+(TT)|T|2g+2divT)Φ,Φ(0)=Φ0.\displaystyle\left\{\begin{array}[]{rl}&\dfrac{\partial\Phi}{\partial t}=\left(-\mathrm{Ric}+2(\mathcal{L}_{T_{8}}g)+(T*T)-|T|^{2}g+2\operatorname{div}T\right)\diamond\Phi,\\ &\Phi(0)=\Phi_{0}.\end{array}\right. (GF)

where Ric\mathrm{Ric} is the Ricci curvature of the underlying metric, T8T_{8} is the 88-dimensional component of the torsion tensor TT and \diamond is the operator described in (2.13).

Given any geometric flow, one of the main questions is whether there exist a solution to the flow for a short time and if any two solutions starting with the same initial conditions are unique. The main theorem of the paper is as follows.

Theorem 4.11. Let (M8,Φ0)(M^{8},\Phi_{0}) be a compact 88-manifold with a Spin(7)-structure Φ0\Phi_{0} and consider the negative gradient flow of the natural energy functional EE in (1.1). This is the flow (GF). Then there exists ε>0\varepsilon>0 and a unique smooth solution Φ(t)\Phi(t) of (GF) for t[0,ε)t\in[0,\varepsilon) with ε=ε(Φ0)\varepsilon=\varepsilon(\Phi_{0}).

We prove the main theorem by explicitly calculating the principal symbols of the operators involved in the flow and then using a modified DeTurck’s trick.

As discussed at the end of § 2, the most general flow of Spin(7)-structures which is second order can be written as

tΦ(t)=(Ric+aT8g+bdivT+l.o.t)Φ\displaystyle\frac{\partial}{\partial t}\Phi(t)=(-\mathrm{Ric}+a\mathcal{L}_{T_{8}}g+b\operatorname{div}T+\text{l.o.t})\diamond\Phi (1.3)

where a,ba,b\in\mathbb{R} and l.o.t are the terms which are lower order in Φ\Phi. The situation in the Spin(7)-case is a bit simpler than the corresponding case of the G2\mathrm{G}_{2}-structures where there are more terms which could appear in a flow of G2\mathrm{G}_{2}-structures [DGK23, §1]. This is because of the fact that for pMp\in M, the subbundle ApMA_{p}M of admissible 4-forms in Λ4TpM\Lambda^{4}T^{*}_{p}M has codimension 2727 and thus any variation of a Spin(7)-structure is given only by 44-forms in Ω1+7+354\Omega^{4}_{1+7+35}. Consequently, many naturally occurring tensors, for example the 2727-dimensional Weyl curvature tensor cannot be a candidate for a flow of Spin(7)-structures (see § 2 for details) even though they can be made into a 44-form.

The methods in § 4 can be used to establish short-time existence and uniqueness results for the general flow in (1.3) subjected to conditions on the coefficients a,ba,b which will be sufficient conditions to apply the modified DeTurck’s trick described in § 4.4. The advantage of looking at a specific flow in the family (1.3), which for us is (GF) which corresponds to a,b=2a,b=2 is that it also provides us with the lower order terms which are important when one wants to study further analytic properties of the flow. An example of the importance of lower order terms is our result on the non-existence of compact expanding solitons for the flow in § 5.

Solitons of a geometric flow are special solutions which move only by scalings and diffeomorphisms. These are also called self-similar solutions. A soliton can be viewed as a generalized fixed point of the flow modulo scalings and diffeomorphisms. These special solutions are important because they serve as singularity models for the flow. We study solitons of the flow (GF). We derive equations for the underlying metric and the divergence of the torsion of a soliton solution. We expect that if we have a Cheeger–Gromov–Hamilton type compactness theorem for the solutions of the flow then the solitons will play a role in understanding the behaviour of the flow near a singularity.

The paper is organized as follows. We start by reviewing the notion of Spin(7)-structures, the decomposition of space of forms and the torsion tensor in § 2. The first variation of the energy functional is computed in § 3. We give a brief review of elliptic and parabolic differential operators in § 4.1 which is followed by a modified DeTurck’s trick in § 4.4. The main theorem, Theorem 4.11, is proved in § 4.5. Finally, in § 5 we study solitons of the flow and prove Proposition 5.2 which is a non-existence result for compact expanding solitons of the flow.

Throughout the paper, we compute in a local orthonormal frame, so all indices are subscripts and any repeated indices are summed over all values from 11 to 88. Our convention for labelling the Riemann curvature tensor is

Rijklxl=(ijji)xk\displaystyle R_{ijkl}\frac{\partial}{\partial x^{l}}=(\nabla_{i}\nabla_{j}-\nabla_{j}\nabla_{i})\frac{\partial}{\partial x^{k}}

in terms of coordinate vector fields and as a result, the Ricci tensor is Rij=RlijlR_{ij}=R_{lijl}. We also have the Ricci identity which, for instance, for a 22-tensor SS reads as

ijSkljiSkl=RijkmSmlRijlmSkm.\displaystyle\nabla_{i}\nabla_{j}S_{kl}-\nabla_{j}\nabla_{i}S_{kl}=-R_{ijkm}S_{ml}-R_{ijlm}S_{km}. (1.4)

Acknowledgements. We are grateful to Panagiotis Gianniotis and Spiro Karigiannis for various discussions on this project in particular and on flows of special geometric structures in general. We would like to thank Thomas Walpuski for discussions on the torsion decomposition of Spin(7)-structures and Ragini Singhal for explaining the representation theoretic aspects of Spin(7)-structures. A part of this work was done when the author is a Simons-CRM Scholar-in-Residence at the CRM, Montreal. We would like to thank the Simons Foundation and the CRM for funding and CRM for providing a wonderful environment to work in.

2.   Preliminaries on Spin(7)-structures

In this section, we briefly review the notion of a Spin(7)-structure on an 88-dimensional manifold MM and the associated decomposition of differential forms. We also discuss the torsion tensor of a Spin(7)-structure. More details can be found in [Joy00, Chapter 10] and [Kar08].

A Spin(7)-structure on MM is a 44-form Φ\Phi on MM. The existence of such a structure is a topological condition. The space of 44-forms which determine a Spin(7)-structure on MM is a subbundle AA of Ω4(M)\Omega^{4}(M), called the bundle of admissible 44-forms. This is not a vector subbundle and it is not even an open subbundle, unlike the case for G2\mathrm{G}_{2}-structures. For pMp\in M, the subbundle Ap(M)A_{p}(M) is of codimension 2727 in Λ4(TpM)\Lambda^{4}(T^{*}_{p}M).

A Spin(7)-structure Φ\Phi determines a Riemannian metric and an orientation on MM in a nonlinear way which can be described in local coordinates as follows. Let pMp\in M and extend a non-zero tangent vector vTpMv\in T_{p}M to a local frame {v,e1,,e7}\{v,e_{1},\cdots,e_{7}\} near pp. Define

Bij(v)\displaystyle B_{ij}(v) =((eivΦ)(ejvΦ)(vΦ))(e1,,e7),\displaystyle=((e_{i}\lrcorner v\lrcorner\Phi)\wedge(e_{j}\lrcorner v\lrcorner\Phi)\wedge(v\lrcorner\Phi))(e_{1},\cdots,e_{7}),
A(v)\displaystyle A(v) =((vΦ)Φ)(e1,,e7).\displaystyle=((v\lrcorner\Phi)\wedge\Phi)(e_{1},\cdots,e_{7}).

Then the metric induced by Φ\Phi is given by

(gΦ(v,v))2=73673(detBij(v))13A(v)3.\displaystyle(g_{\Phi}(v,v))^{2}=-\frac{7^{3}}{6^{\frac{7}{3}}}\frac{(\textup{det}\ B_{ij}(v))^{\frac{1}{3}}}{A(v)^{3}}. (2.1)

The metric and the orientation determine a Hodge star operator \star, and the 44-form is self-dual, i.e., Φ=Φ\star\Phi=\Phi.

Definition 2.1.

Let \nabla be the Levi-Civita connection of the metric gΦg_{\Phi}. The pair (M8,Φ)(M^{8},\Phi) is a Spin(7)-manifold if Φ=0\nabla\Phi=0. This is a non-linear partial differential equation for Φ\Phi, since \nabla depends on gg, which in turn depends non-linearly on Φ\Phi. A Spin(7)-manifold has Riemannian holonomy contained in the subgroup Spin(7)SO(8)\mathrm{Spin}(7)\subset\mathrm{SO}(8). Such a parallel Spin(7)-structure is also called torsion free.

2.1 Decomposition of the space of forms

The existence of a Spin(7)-structure Φ\Phi induces a decomposition of the space of differential forms on MM into irreducible Spin(7)-representations. We have the following orthogonal decomposition, with respect to gΦg_{\Phi}:

Ω2=Ω72Ω212,Ω3=Ω83Ω483,Ω4=Ω14Ω74Ω274Ω354,\displaystyle\Omega^{2}=\Omega^{2}_{7}\oplus\Omega^{2}_{21},\ \ \ \ \ \ \ \Omega^{3}=\Omega^{3}_{8}\oplus\Omega^{3}_{48},\ \ \ \ \ \ \ \ \ \Omega^{4}=\Omega^{4}_{1}\oplus\Omega^{4}_{7}\oplus\Omega^{4}_{27}\oplus\Omega^{4}_{35},

where Ωlk\Omega^{k}_{l} has pointwise dimension ll. Explicitly, Ω2\Omega^{2} and Ω3\Omega^{3} are described as follows:

Ω72\displaystyle\Omega^{2}_{7} ={βΩ2(Φβ)=3β},\displaystyle=\{\beta\in\Omega^{2}\mid\star(\Phi\wedge\beta)=-3\beta\}, (2.2)
Ω212\displaystyle\Omega^{2}_{21} ={βΩ2(Φβ)=β},\displaystyle=\{\beta\in\Omega^{2}\mid\star(\Phi\wedge\beta)=\beta\}, (2.3)

and

Ω83\displaystyle\Omega^{3}_{8} ={XΦXΓ(TM)},\displaystyle=\{X\lrcorner\Phi\mid X\in\Gamma(TM)\}, (2.4)
Ω483\displaystyle\Omega^{3}_{48} ={γΩ3γΦ=0}.\displaystyle=\{\gamma\in\Omega^{3}\mid\gamma\wedge\Phi=0\}. (2.5)

For our computations, it is useful to describe the spaces of forms in local coordinates. For βΩ2(M)\beta\in\Omega^{2}(M),

βijΩ72βabΦabij\displaystyle\beta_{ij}\in\Omega^{2}_{7}\iff\beta_{ab}\Phi_{abij} =6βij,\displaystyle=-6\beta_{ij}, (2.6)
βijΩ212βabΦabij\displaystyle\beta_{ij}\in\Omega^{2}_{21}\iff\beta_{ab}\Phi_{abij} =2βij.\displaystyle=2\beta_{ij}. (2.7)
Remark 2.2.

A convention different than ours is also prevalent in the literature. In this convention, Ω72\Omega^{2}_{7} and Ω212\Omega^{2}_{21} are the +3+3 and 1-1 eigenspaces of the map β(Φβ),\beta\mapsto*(\Phi\wedge\beta), respectively. As a result, the constants on the right hand sides of (2.6) and (2.7) are +6+6 and 2-2 respectively.

For γΩ3(M)\gamma\in\Omega^{3}(M),

γijkΩ83\displaystyle\gamma_{ijk}\in\Omega^{3}_{8} γijk=XlΦijklfor someXΓ(TM),\displaystyle\iff\gamma_{ijk}=X_{l}\Phi_{ijkl}\ \ \ \ \textup{for\ some}\ X\in\Gamma(TM), (2.8)
γijkΩ483\displaystyle\gamma_{ijk}\in\Omega^{3}_{48} γijkΦijkl=0.\displaystyle\iff\gamma_{ijk}\Phi_{ijkl}=0. (2.9)

If π7\pi_{7} and π21\pi_{21} are the projection operators on Ω2\Omega^{2}, it follows from (2.6) and (2.7) that

π7(β)ij\displaystyle\pi_{7}(\beta)_{ij} =14βij18βabΦabij,\displaystyle=\frac{1}{4}\beta_{ij}-\frac{1}{8}\beta_{ab}\Phi_{abij}, (2.10)
π21(β)ij\displaystyle\pi_{21}(\beta)_{ij} =34βij+18βabΦabij.\displaystyle=\frac{3}{4}\beta_{ij}+\frac{1}{8}\beta_{ab}\Phi_{abij}. (2.11)

We will be using these equations throughout the paper. Finally, for βijΩ212\beta_{ij}\in\Omega^{2}_{21},

βabΦbpqr\displaystyle\beta_{ab}\Phi_{bpqr} =βpiΦiqra+βqiΦirpa+βriΦipqa,\displaystyle=\beta_{pi}\Phi_{iqra}+\beta_{qi}\Phi_{irpa}+\beta_{ri}\Phi_{ipqa}, (2.12)

which can be used to show that Ω212𝔰𝔬(7)\Omega^{2}_{21}\equiv\mathfrak{so}(7) is the Lie algebra of Spin(7), with the commutator of matrices

[α,β]ij=αilβljαjlβli.\displaystyle[\alpha,\beta]_{ij}=\alpha_{il}\beta_{lj}-\alpha_{jl}\beta_{li}.

To describe Ω4\Omega^{4} in local coordinates, we use the \diamond operator which was first described in [DGK21] for the G2\mathrm{G}_{2} case and in [DLE24] for the Spin(7) (see also [FLME22, eq. (1.14)] for the general case of HH-structures). Given AΓ(TMTM)A\in\Gamma(T^{*}M\otimes TM), define

AΦ=124(AipΦpjkl+AjpΦipkl+AkpΦijpl+AlpΦijkp)dxidxjdxkdxl,\displaystyle A\diamond\Phi=\frac{1}{24}(A_{ip}\Phi_{pjkl}+A_{jp}\Phi_{ipkl}+A_{kp}\Phi_{ijpl}+A_{lp}\Phi_{ijkp})dx^{i}\wedge dx^{j}\wedge dx^{k}\wedge dx^{l}, (2.13)

and hence

(AΦ)ijkl=AipΦpjkl+AjpΦipkl+AkpΦijpl+AlpΦijkp.\displaystyle(A\diamond\Phi)_{ijkl}=A_{ip}\Phi_{pjkl}+A_{jp}\Phi_{ipkl}+A_{kp}\Phi_{ijpl}+A_{lp}\Phi_{ijkp}. (2.14)

Recall that Γ(TMTM)=Ω0S0Ω2\Gamma(T^{*}M\otimes TM)=\Omega^{0}\oplus S_{0}\oplus\Omega^{2}, and Ω2\Omega^{2} further splits orthogonally into (2.2) and (2.3), so

Γ(TMTM)=Ω0S0Ω72Ω212.\displaystyle\Gamma(T^{*}M\otimes TM)=\Omega^{0}\oplus S_{0}\oplus\Omega^{2}_{7}\oplus\Omega^{2}_{21}. (2.15)

With respect to this splitting, we can write A=18(trA)g+A35+A7+A21A=\frac{1}{8}(\operatorname{tr}A)g+A_{35}+A_{7}+A_{21} where A35A_{35} is a symmetric traceless 22-tensor. The diamond contraction (2.14) defines a linear map AAΦA\mapsto A\diamond\Phi, from Ω0S0Ω72Ω212\Omega^{0}\oplus S_{0}\oplus\Omega^{2}_{7}\oplus\Omega^{2}_{21} to Ω4(M)\Omega^{4}(M). We record the following properties of the \diamond operator whose proof can be found in [Kar08] or [DLE24].

Proposition 2.3.

Let (M,Φ)(M,\Phi) be a manifold with a Spin(7)-structure. Then

  1. 1.

    The Hodge star of AΦA\diamond\Phi is

    (AΦ)\displaystyle\star(A\diamond\Phi) =(A¯Φ)whereA¯=14(trA)gijAji.\displaystyle=(\bar{A}\diamond\Phi)\ \ \textup{where}\ \ \bar{A}=\frac{1}{4}(\operatorname{tr}A)g_{ij}-A_{ji}. (2.16)
  2. 2.

    If BΓ(TMTM)B\in\Gamma(T^{*}M\otimes TM) and A35,B35A_{35},\ B_{35} denote the traceless symmetric parts of AA and BB respectively and A7,B7A_{7},\ B_{7} denote their Ω72\Omega^{2}_{7} component. Then

    AΦ,BΦ\displaystyle\langle A\diamond\Phi,B\diamond\Phi\rangle =24(trA)(trB)+96A35,B35+384A7,B7.\displaystyle=24(\operatorname{tr}A)(\operatorname{tr}B)+96\langle A_{35},B_{35}\rangle+384\langle A_{7},B_{7}\rangle. (2.17)

More properties of the \diamond operator for general HH-structures have been proved in [FLME22, Lemma 1.4].

The following proposition was originally proved in [Kar08, Prop. 2.3].

Proposition 2.4.

The kernel of the map AAΦA\mapsto A\diamond\Phi is isomorphic to the subspace Ω212\Omega^{2}_{21}. The remaining three summands Ω0,S0\Omega^{0},\ S_{0} and Ω72\Omega^{2}_{7} are mapped isomorphically onto the subspaces Ω14,Ω354\Omega^{4}_{1},\ \Omega^{4}_{35} and Ω74\Omega^{4}_{7} respectively.

Proof.

Using (2.17) with A=BA=B gives

|AΦ|2\displaystyle|A\diamond\Phi|^{2} =24(trA)2+96|A35|2+384|A7|2\displaystyle=24(\operatorname{tr}A)^{2}+96|A_{35}|^{2}+384|A_{7}|^{2}

and hence AΦ=0A=A21A\diamond\Phi=0\iff A=A_{21}. If A21=0A_{21}=0 then since all the coefficients in |AΦ|2|A\diamond\Phi|^{2} are positive hence the map AAΦA\mapsto A\diamond\Phi is injective on (Ω212)(\Omega^{2}_{21})^{\perp} and by dimension count, the map is a linear isomorphism. ∎

To understand Ω274\Omega^{4}_{27}, we need another characterization of the space of 44-forms using the Spin(7)-structure. Following [Kar08], we adopt the following:

Definition 2.5.

On (M,Φ)(M,\Phi), define a Spin(7)-equivariant linear operator ΛΦ\Lambda_{\Phi} on Ω4\Omega^{4} as follows. Let σΩ4(M)\sigma\in\Omega^{4}(M) and let (σΦ)ijkl=σijmnΦmnkl(\sigma\cdot\Phi)_{ijkl}=\sigma_{ijmn}\Phi_{mnkl}. Then

(ΛΦ(σ))ijkl=(σΦ)ijkl+(σΦ)iklj+(σΦ)iljk+(σΦ)jkil+(σΦ)jlki+(σΦ)klij.\displaystyle(\Lambda_{\Phi}(\sigma))_{ijkl}=(\sigma\cdot\Phi)_{ijkl}+(\sigma\cdot\Phi)_{iklj}+(\sigma\cdot\Phi)_{iljk}+(\sigma\cdot\Phi)_{jkil}+(\sigma\cdot\Phi)_{jlki}+(\sigma\cdot\Phi)_{klij}. (2.18)
Proposition 2.6.

The spaces Ω14\Omega^{4}_{1}, Ω74,Ω274\Omega^{4}_{7},\ \Omega^{4}_{27} and Ω354\Omega^{4}_{35} are all eigenspaces of ΛΦ\Lambda_{\Phi} with distinct eigenvalues:

Ω14={σΩ4ΛΦ(σ)=24σ},Ω274={σΩ4ΛΦ(σ)=4σ},\displaystyle\begin{aligned} \Omega^{4}_{1}&=\{\sigma\in\Omega^{4}\mid\Lambda_{\Phi}(\sigma)=-24\sigma\},\\ \Omega^{4}_{27}&=\{\sigma\in\Omega^{4}\mid\Lambda_{\Phi}(\sigma)=4\sigma\},\end{aligned} Ω74={σΩ4ΛΦ(σ)=12σ},Ω354={σΩ4ΛΦ(σ)=0}.\displaystyle\begin{aligned} \Omega^{4}_{7}&=\{\sigma\in\Omega^{4}\mid\Lambda_{\Phi}(\sigma)=-12\sigma\},\\ \Omega^{4}_{35}&=\{\sigma\in\Omega^{4}\mid\Lambda_{\Phi}(\sigma)=0\}.\end{aligned} (2.19)

Moreover, the decomposition of Ω4(M)\Omega^{4}(M) into self-dual and anti-self-dual parts is

Ω+4={σΩ4σ=σ}=Ω14Ω74Ω274,Ω4={σΩ4σ=σ}=Ω354.\displaystyle\Omega^{4}_{+}=\{\sigma\in\Omega^{4}\mid\star\sigma=\sigma\}=\Omega^{4}_{1}\oplus\Omega^{4}_{7}\oplus\Omega^{4}_{27},\ \ \ \ \ \Omega^{4}_{-}=\{\sigma\in\Omega^{4}\mid\star\sigma=-\sigma\}=\Omega^{4}_{35}. (2.20)

Before we discuss the torsion of a Spin(7)-structure, we note some contraction identities involving the 44-form Φ\Phi. In local coordinates {x1,,x8}\{x^{1},\cdots,x^{8}\}, the 44-form Φ\Phi is

Φ=124Φijkldxidxjdxkdxl\displaystyle\Phi=\frac{1}{24}\Phi_{ijkl}\ dx^{i}\wedge dx^{j}\wedge dx^{k}\wedge dx^{l}

where Φijkl\Phi_{ijkl} is totally skew-symmetric. We have the following identities

ΦijklΦabcl\displaystyle\Phi_{ijkl}\Phi_{abcl} =giagjbgkc+gibgjcgka+gicgjagkb\displaystyle=g_{ia}g_{jb}g_{kc}+g_{ib}g_{jc}g_{ka}+g_{ic}g_{ja}g_{kb}
giagjcgkbgibgjagkcgicgjbgka\displaystyle\quad-g_{ia}g_{jc}g_{kb}-g_{ib}g_{ja}g_{kc}-g_{ic}g_{jb}g_{ka}
giaΦjkbcgibΦjkcagicΦjkab\displaystyle\quad-g_{ia}\Phi_{jkbc}-g_{ib}\Phi_{jkca}-g_{ic}\Phi_{jkab}
gjaΦkibcgjbΦkicagjcΦkiab\displaystyle\quad-g_{ja}\Phi_{kibc}-g_{jb}\Phi_{kica}-g_{jc}\Phi_{kiab}
gkaΦijbcgkbΦijcagkcΦijab\displaystyle\quad-g_{ka}\Phi_{ijbc}-g_{kb}\Phi_{ijca}-g_{kc}\Phi_{ijab} (2.21)
ΦijklΦabkl\displaystyle\Phi_{ijkl}\Phi_{abkl} =6giagjb6gibgja4Φijab\displaystyle=6g_{ia}g_{jb}-6g_{ib}g_{ja}-4\Phi_{ijab} (2.22)
ΦijklΦajkl\displaystyle\Phi_{ijkl}\Phi_{ajkl} =42gia\displaystyle=42g_{ia} (2.23)
ΦijklΦijkl\displaystyle\Phi_{ijkl}\Phi_{ijkl} =336.\displaystyle=336. (2.24)

We also have contraction identities involving Φ\nabla\Phi and Φ\Phi

(mΦijkl)Φabkl\displaystyle(\nabla_{m}\Phi_{ijkl})\Phi_{abkl} =Φijkl(mΦabkl)4mΦijab\displaystyle=-\Phi_{ijkl}(\nabla_{m}\Phi_{abkl})-4\nabla_{m}\Phi_{ijab} (2.25)
(mΦijkl)Φajkl\displaystyle(\nabla_{m}\Phi_{ijkl})\Phi_{ajkl} =Φijkl(mΦajkl)\displaystyle=-\Phi_{ijkl}(\nabla_{m}\Phi_{ajkl}) (2.26)
(mΦijkl)Φijkl\displaystyle(\nabla_{m}\Phi_{ijkl})\Phi_{ijkl} =0.\displaystyle=0. (2.27)

We now describe the torsion of a Spin(7)-structure. Given XΓ(TM)X\in\Gamma(TM), we know from [Kar08, Lemma 2.10] that XΦ\nabla_{X}\Phi lies in the subbundle Ω74Ω4\Omega^{4}_{7}\subset\Omega^{4}.

Definition 2.7.

The torsion tensor of a Spin(7)-structure Φ\Phi is the element of Ω81Ω72\Omega^{1}_{8}\otimes\Omega^{2}_{7} defined by expressing Φ\nabla\Phi Since XΦΩ74\nabla_{X}\Phi\in\Omega^{4}_{7}, by Proposition 2.4, Φ\nabla\Phi can be written as

mΦijkl=(TmΦ)ijkl=Tm;ipΦpjkl+Tm;jpΦipkl+Tm;kpΦijpl+Tm;lpΦijkp\displaystyle\nabla_{m}\Phi_{ijkl}=(T_{m}\diamond\Phi)_{ijkl}=T_{m;ip}\Phi_{pjkl}+T_{m;jp}\Phi_{ipkl}+T_{m;kp}\Phi_{ijpl}+T_{m;lp}\Phi_{ijkp} (2.28)

where Tm;abΩ72T_{m;ab}\in\Omega^{2}_{7}, for each fixed mm. This defines the torsion tensor TT of a Spin(7)-structure, which is an element of Ω81Ω72\Omega^{1}_{8}\otimes\Omega^{2}_{7}.

In terms of Φ\nabla\Phi, the torsion TT is given by

Tm;ab=196(mΦajkl)Φbjkl\displaystyle T_{m;ab}=\frac{1}{96}(\nabla_{m}\Phi_{ajkl})\Phi_{bjkl} (2.29)

since TT is an element of Ω81Ω72\Omega^{1}_{8}\otimes\Omega^{2}_{7}. Thus we have the following result, originally due to Fernández [Fer86].

Theorem 2.8.

[Fer86] The Spin(7)-structure Φ\Phi is torsion free if and only if dΦ=0d\Phi=0. Since Φ=Φ\star\Phi=\Phi, this is equivalent to dΦ=0d^{*}\Phi=0.

Remark 2.9.

The notation Tm;abT_{m;ab} should not be confused with taking two covariant derivatives of TmT_{m}. The torsion tensor T is an element of Ω81Ω72\Omega^{1}_{8}\otimes\Omega^{2}_{7} and thus, for each fixed index mm, Tm;abΩ72T_{m;ab}\in\Omega^{2}_{7}.

Finally, an important bit of information which we need about the torsion is that it satisfies a “Bianchi-type identity”. This was first proved by Karigiannis [Kar08, Theorem 4.2] using the diffeomorphism invariance of the torsion tensor and a different proof using the Ricci identity (1.4) was given in [DLE24, Theorem 3.9].

Theorem 2.10.

The torsion tensor TT satisfies the following “Bianchi-type” identity

iTj;abjTi;ab=2Ti;amTj;mb2Tj;amTi;mb+14Rjiab18RjimnΦmnab.\displaystyle\nabla_{i}T_{j;ab}-\nabla_{j}T_{i;ab}=2T_{i;am}T_{j;mb}-2T_{j;am}T_{i;mb}+\frac{1}{4}R_{jiab}-\frac{1}{8}R_{jimn}\Phi_{mnab}. (2.30)

Using the Riemannian Bianchi identity, we see that

RijklΦajkl=(Rjkil+Rkijl)Φajkl=RiljkΦaljkRikjlΦakjl\displaystyle R_{ijkl}\Phi_{ajkl}=-(R_{jkil}+R_{kijl})\Phi_{ajkl}=-R_{iljk}\Phi_{aljk}-R_{ikjl}\Phi_{akjl}

and hence we have the fact that

RijklΦajkl=0.\displaystyle R_{ijkl}\Phi_{ajkl}=0. (2.31)

Using this and contracting (2.30) on jj and bb gives the expression for the Ricci curvature of a metric induced by a Spin(7)-structure. Precisely,

Rij=4iTa;ja4aTi;ja8Ti;jbTa;ba+8Ta;jbTi;ba\displaystyle R_{ij}=4\nabla_{i}T_{a;ja}-4\nabla_{a}T_{i;ja}-8T_{i;jb}T_{a;ba}+8T_{a;jb}T_{i;ba} (2.32)

which also proves that the metric of a torsion free Spin(7)-structure is Ricci-flat, a result originally due to Bonan. Taking the trace of (2.32) gives the expression of the scalar curvature RR

R\displaystyle R =4iTa;ia4aTi;ia+8|T|2+8Ta;jbTj;ba,\displaystyle=4\nabla_{i}T_{a;ia}-4\nabla_{a}T_{i;ia}+8|T|^{2}+8T_{a;jb}T_{j;ba}, (2.33)
=(2.39)8divT8+8|T8|2+8Ta;jbTj;ba.\displaystyle\stackrel{{\scriptstyle\eqref{T8des}}}{{=}}8\operatorname{div}T_{8}+8|T_{8}|^{2}+8T_{a;jb}T_{j;ba}. (2.34)

Since T=T8+T48T=T_{8}+T_{48}, the T8T_{8} component of torsion can be viewed as a vector field on MM and hence we want to look at alternate expression for the symmetric 22-tensor T8g\mathcal{L}_{T_{8}}g which will be used in § 3.1. We have the following linear algebra result whose proof can be found in [DGK23, Lemma 4.1].

Lemma 2.11.

Let VV and WW be finite-dimensional real vector spaces equipped with positive definite inner products, and suppose that

V=V1VmV=V_{1}\oplus_{\perp}\cdots\oplus_{\perp}V_{m}

is an orthogonal direct sum of subspaces. Let ι:VW\iota\colon V\to W and ρ:WV\rho\colon W\to V be linear maps. Suppose that for every 1km1\leq k\leq m, there exist bk,ckb_{k},c_{k} both nonzero, such that for all vkVkv_{k}\in V_{k} and wWw\in W, we have

(i)ριvk=bkvk and (ii)ρw,vk=ckw,ιvk.\mathrm{(i)}\,\,\rho\iota v_{k}=b_{k}v_{k}\quad\text{ and }\quad\mathrm{(ii)}\,\,\langle\rho w,v_{k}\rangle=c_{k}\langle w,\iota v_{k}\rangle. (2.35)

Then in fact we have an isomorphism of WW with an orthogonal direct sum

W(kerρ)V=(kerρ)V1Vk.W\cong(\ker\rho)\oplus_{\perp}V=(\ker\rho)\oplus_{\perp}V_{1}\oplus_{\perp}\cdots\oplus_{\perp}V_{k}. (2.36)

Let V=Λ81V=\Lambda^{1}_{8} and W=Λ81Λ72W=\Lambda^{1}_{8}\otimes\Lambda^{2}_{7} in the previous lemma. Define ι:Λ81Λ81Λ72\iota:\Lambda^{1}_{8}\rightarrow\Lambda^{1}_{8}\otimes\Lambda^{2}_{7} by

ι(X)ijk=XkgijXjgik+XpΦpijk,\displaystyle\iota(X)_{ijk}=X_{k}g_{ij}-X_{j}g_{ik}+X_{p}\Phi_{pijk},

and ρ:Λ81Λ72Λ81\rho:\Lambda^{1}_{8}\otimes\Lambda^{2}_{7}\rightarrow\Lambda^{1}_{8} by

ρ(γ)j=γi;ji.\displaystyle\rho(\gamma)_{j}=\gamma_{i;ji}.

Clearly ι(X)ijk=ι(X)ikj\iota(X)_{ijk}=-\iota(X)_{ikj}. Consider

ι(X)ijkΦabjk\displaystyle\iota(X)_{ijk}\Phi_{abjk} =(XkgijXjgik+XpΦpijk)Φabjk\displaystyle=\left(X_{k}g_{ij}-X_{j}g_{ik}+X_{p}\Phi_{pijk}\right)\Phi_{abjk}
=XkΦabikXjΦabji+Xp(6gpagib6gpbgia4Φpiab)\displaystyle=X_{k}\Phi_{abik}-X_{j}\Phi_{abji}+X_{p}(6g_{pa}g_{ib}-6g_{pb}g_{ia}-4\Phi_{piab})
=2XkΦabik+6Xagib6Xbgia+4XpΦabip\displaystyle=2X_{k}\Phi_{abik}+6X_{a}g_{ib}-6X_{b}g_{ia}+4X_{p}\Phi_{abip}
=6Xagib6Xbgia+6XkΦabik=6ι(X)iab\displaystyle=6X_{a}g_{ib}-6X_{b}g_{ia}+6X_{k}\Phi_{abik}=-6\iota(X)_{iab} (2.37)

where we have used (2.23) in the second equality. Hence (2.6) implies ι(X)Λ81Λ72\iota(X)\in\Lambda^{1}_{8}\otimes\Lambda^{2}_{7}. We compute

(ρι(X))j\displaystyle(\rho\iota(X))_{j} =ρ(XkgijXjgik+XpΦpijk)j=7Xj.\displaystyle=\rho(X_{k}g_{ij}-X_{j}g_{ik}+X_{p}\Phi_{pijk})_{j}=-7X_{j}.

Also,

ργ,X\displaystyle\langle\rho\gamma,X\rangle =γi;jiXj,\displaystyle=\gamma_{i;ji}X_{j},

and, using (2.6),

γ,ιX\displaystyle\langle\gamma,\iota X\rangle =γi;jk(XkgijXjgik+XpΦpijk)\displaystyle=\gamma_{i;jk}(X_{k}g_{ij}-X_{j}g_{ik}+X_{p}\Phi_{pijk})
=γi;ikXkXjγi;ji+Xpγi;jkΦpijk\displaystyle=\gamma_{i;ik}X_{k}-X_{j}\gamma_{i;ji}+X_{p}\gamma_{i;jk}\Phi_{pijk}
=8Xjγi;ji.\displaystyle=-8X_{j}\gamma_{i;ji}.

Thus, (i) and (ii) in (2.35) are satisfied with bk=7b_{k}=-7 and ck=8c_{k}=-8 respectively. We deduce from Lemma 2.11 that

Λ81Λ72(kerρ)Λ81\displaystyle\Lambda^{1}_{8}\otimes\Lambda^{2}_{7}\cong(\ker\rho)\oplus_{\perp}\Lambda^{1}_{8} (2.38)

with kerρ={γi;jkΛ81Λ72γi;ji=0}\ker\rho=\{\gamma_{i;jk}\in\Lambda^{1}_{8}\otimes\Lambda^{2}_{7}\mid\gamma_{i;ji}=0\} being 4848-dimensional. In particular, since TΦΛ81Λ72T_{\Phi}\in\Lambda^{1}_{8}\otimes\Lambda^{2}_{7}, we see that

(T8)j=Ti;ji(T_{8})_{j}=T_{i;ji} (2.39)

and thus

(T8g)ij=i(T8)j+j(T8)i=i(Tk;jk)+j(Tk:ik).\displaystyle\left(\mathcal{L}_{T_{8}}g\right)_{ij}=\nabla_{i}(T_{8})_{j}+\nabla_{j}(T_{8})_{i}=\nabla_{i}(T_{k;jk})+\nabla_{j}(T_{k:ik}). (2.40)

We will also need the expression for the Lie derivative of the Spin(7)-structure in § 4. Let WΓ(TM)W\in\Gamma(TM). Using (2.28) and (2.14), we have

(WΦ)ijkl\displaystyle(\mathcal{L}_{W}\Phi)_{ijkl} =WppΦijkl+iWpΦpjkl+jWpΦipkl+kWpΦijpl+lWpΦijkp\displaystyle=W_{p}\nabla_{p}\Phi_{ijkl}+\nabla_{i}W_{p}\Phi_{pjkl}+\nabla_{j}W_{p}\Phi_{ipkl}+\nabla_{k}W_{p}\Phi_{ijpl}+\nabla_{l}W_{p}\Phi_{ijkp}
=Wp(TpΦ)ijkl+(WΦ)ijkl\displaystyle=W_{p}(T_{p}\diamond\Phi)_{ijkl}+(\nabla W\diamond\Phi)_{ijkl}

and since W=(W)sym+(W)skew=12Wg+(W)7+(W)21\nabla W=(\nabla W)_{\text{sym}}+(\nabla W)_{\text{skew}}=\frac{1}{2}\mathcal{L}_{W}g+(\nabla W)_{7}+(\nabla W)_{21}, where (W)7(\nabla W)_{7} and (W)21(\nabla W)_{21} denote the projection of the 22-form (W)skew(\nabla W)_{\text{skew}} onto the Ω72\Omega^{2}_{7} and Ω212\Omega^{2}_{21} components respectively. Using the fact that Ω212Φ=0\Omega^{2}_{21}\diamond\Phi=0, we get

(WΦ)ijkl\displaystyle(\mathcal{L}_{W}\Phi)_{ijkl} =(12Wg+T(W)+(W)7Φ)ijkl.\displaystyle=\left(\frac{1}{2}\mathcal{L}_{W}g+T(W)+(\nabla W)_{7}\diamond\Phi\right)_{ijkl}. (2.41)

We record this observation in the following

Proposition 2.12.

A vector field WW is an infinitesimal symmetry of Φ\Phi, i.e., WΦ=0\mathcal{L}_{W}\Phi=0 if and only if

Wg=0andT(W)=(W)7.\displaystyle\mathcal{L}_{W}g=0\ \ \ \ \ \text{and}\ \ \ T(W)=-(\nabla W)_{7}. (2.42)

Remark 2.13.

Proposition 2.12 and (2.41) is essentially [DLE24, Lemma 2.6] where it was proved for arbitrary HH-structures.

Comparing with the case of manifolds with G2\mathrm{G}_{2}-structures, the most general flow of G2\mathrm{G}_{2}-structures which are second order quasilinear are described in [DGK23] and the authors prove the short-time existence and uniqueness theorem for a large class of flows. Such a family of flows was obtained by explicitly computing all independent second order differential invariants of G2\mathrm{G}_{2}-structures which are 33-forms. Since the Riemann curvature tensor Rm\mathrm{Rm} and T\nabla T are the only two second order invariants of a G2\mathrm{G}_{2}-structure, decomposing them into irreducible G2\mathrm{G}_{2}-representations and then picking the linearly independent components (since Rm\mathrm{Rm} and T\nabla T are related by the so called G2\mathrm{G}_{2}-Bianchi identity) which can be made into a 33-form provides a way to write the most general flow of G2\mathrm{G}_{2}-structures. See [DGK23, §4] for more details. The existence of a Spin(7)-structure induces a decomposition of tensor bundles into irreducible Spin(7)-representations and one can similarly decompose Rm\mathrm{Rm} and T\nabla T into irreducible Spin(7)-representations and look at the linearly independent components which can be made into an admissible 44-form. We describe this briefly below without any proofs.

Let us denote the kk-dimensional irreducible Spin(7)-representation by 𝐤\mathbf{k}. Recall that the space 𝒦\mathcal{K} of curvature tensors on (M,g)(M,g) is the subspace of S2(Λ2)=Γ(S2(Λ2TM))S^{2}(\Lambda^{2})=\Gamma(\mathrm{S}^{2}(\Lambda^{2}T^{*}M)) of elements satisfying the first Bianchi identity and hence we have the orthogonal decomposition

S2(Λ2)\displaystyle S^{2}(\Lambda^{2}) =Ω4𝒦.\displaystyle=\Omega^{4}\oplus_{\perp}\mathcal{K}.

Since for a manifold with a Spin(7)-structure Φ\Phi, we have Λ2=𝟕𝟐𝟏\Lambda^{2}=\mathbf{7}\oplus\mathbf{21} (see § 2.1), we have

S2(Λ2)=S2(𝟕𝟐𝟏)=S2(𝟕)𝟕𝟐𝟏S2(𝟐𝟏).\displaystyle S^{2}(\Lambda^{2})=S^{2}(\mathbf{7}\oplus\mathbf{21})=S^{2}(\mathbf{7})\oplus\mathbf{7}\otimes\mathbf{21}\oplus S^{2}(\mathbf{21}).

One can check the following decomposition of tensor bundles into irreducible Spin(7)-representations.

𝟕𝟕49\displaystyle\underbrace{\mathbf{7}\otimes\mathbf{7}}_{49} =𝟏𝟐𝟏𝟐𝟕\displaystyle=\mathbf{1}\oplus\mathbf{21}\oplus\mathbf{27} (2.43)
𝟕𝟐𝟏147\displaystyle\underbrace{\mathbf{7}\otimes\mathbf{21}}_{147} =𝟑𝟓𝟕𝟏𝟎𝟓\displaystyle=\mathbf{35}\oplus\mathbf{7}\oplus\mathbf{105} (2.44)
𝟐𝟏𝟐𝟏441\displaystyle\underbrace{\mathbf{21}\otimes\mathbf{21}}_{441} =𝟐𝟏𝟏𝟖𝟗𝟏𝟔𝟖𝟐𝟕𝟏𝟑𝟓\displaystyle=\mathbf{21}\oplus\mathbf{189}\oplus\mathbf{168}\oplus\mathbf{27}\oplus\mathbf{1}\oplus\mathbf{35} (2.45)
S2(21)231\displaystyle\underbrace{S^{2}(21)}_{231} =𝟐𝟕𝟏𝟔𝟖𝟑𝟓𝟏\displaystyle=\mathbf{27}\oplus\mathbf{168}\oplus\mathbf{35}\oplus\mathbf{1} (2.46)

and hence we have

S2(Λ2)\displaystyle S^{2}(\Lambda^{2}) =(𝟏𝟐𝟕)(𝟑𝟓𝟕𝟏𝟎𝟓)(𝟐𝟕𝟏𝟔𝟖𝟑𝟓𝟏).\displaystyle=(\mathbf{1}\oplus\mathbf{27})\oplus(\mathbf{35}\oplus\mathbf{7}\oplus\mathbf{105})\oplus(\mathbf{27}\oplus\mathbf{168}\oplus\mathbf{35}\oplus\mathbf{1}). (2.47)

Since the space of curvature type tensors 𝒦\mathcal{K} is orthogonal to the space of 44-forms Ω4=𝟏𝟕𝟐𝟕𝟑𝟓\Omega^{4}=\mathbf{1}\oplus\mathbf{7}\oplus\mathbf{27}\oplus\mathbf{35}, we get the following decomposition

𝒦\displaystyle\mathcal{K} =𝟏𝟑𝟓Ricci𝟐𝟕𝟏𝟎𝟓𝟏𝟔𝟖Weyl.\displaystyle=\underbrace{\mathbf{1}\oplus\mathbf{35}}_{\text{Ricci}}\oplus\underbrace{\mathbf{27}\oplus\mathbf{105}\oplus\mathbf{168}}_{\text{Weyl}}.

Thus, in the presence of a Spin(7)-structure, the Weyl tensor further decomposes as W=W27+W105+W168W=W_{27}+W_{105}+W_{168}. Like the G2\mathrm{G}_{2}-case there is again a W27W_{27}, however unlike the G2\mathrm{G}_{2}-case, this W27W_{27} cannot be used for a flow of Spin(7)-structures as the space of admissible 44-forms is (pointwise) 6363-dimensional and is 𝟏𝟕𝟑𝟓\mathbf{1}\oplus\mathbf{7}\oplus\mathbf{35}. Consequently, the only contribution from Rm\mathrm{Rm} for a flow of Spin(7)-structures are the traceless Ricci curvature Ric\mathrm{Ric}_{\circ} and the scalar curvature RgRg.

In a similar way we look at the possible contributions from T\nabla T for a flow of Spin(7)-structures. Since T=T8+T48T=T_{8}+T_{48}, hence at every point it lies in the representation 𝟖𝟒𝟖\mathbf{8}\oplus\mathbf{48}. Thus,

T\displaystyle\nabla T =T8+T48\displaystyle=\nabla T_{8}+\nabla T_{48}
(𝟖𝟖)(𝟖𝟒𝟖)\displaystyle\in(\mathbf{8}\otimes\mathbf{8})\oplus(\mathbf{8}\otimes\mathbf{48})
=(𝟏𝟕𝟐𝟏𝟑𝟓)(𝟑𝟓𝟐𝟕𝟏𝟖𝟗𝟐𝟏𝟏𝟎𝟓𝟕)\displaystyle=(\mathbf{1}\oplus\mathbf{7}\oplus\mathbf{21}\oplus\mathbf{35})\oplus(\mathbf{35}\oplus\mathbf{27}\oplus\mathbf{189}\oplus\mathbf{21}\oplus\mathbf{105}\oplus\mathbf{7})

and hence the components which can possibly contribute to a flow of Spin(7)-structures are 𝟏,𝟕\mathbf{1},\mathbf{7} and 𝟑𝟓\mathbf{35}. The 𝟏\mathbf{1} component from TT is the divT8\operatorname{div}T_{8} term which can be written in terms of the scalar curvature RR and lower order terms as demonstrated in (2.34). Similarly, the two 𝟑𝟓\mathbf{35} components are a combination from T8g\mathcal{L}_{T_{8}}g and Ric\mathrm{Ric} because T\nabla T and Rm\mathrm{Rm} are related to each other by the Spin(7)-Bianchi identity (2.30). The contribution for the 𝟕\mathbf{7} component from T\nabla T is the divT\operatorname{div}T term (since in the Spin(7)-case divT=divTt\operatorname{div}T=-\operatorname{div}T^{t}). One can do an explicit description of all the components above from Rm\mathrm{Rm} and T\nabla T as in [DGK23]. However, since we are only interested in flows of Spin(7)-structures for which we have all the relevant components, namely, Ric,T8g\mathrm{Ric},\mathcal{L}_{T_{8}}g and divT\operatorname{div}T, we do not pursue the former here and we are currently doing this in collaboration with R. Singhal.

3.   A gradient flow of Spin(7)-structures

3.1 Derivation of the flow

In this section we define the flow of Spin(7)-structures studied in this paper and prove that it is a negative gradient flow.

Definition 3.1.

Let M8M^{8} be a compact manifold. The energy functional EE on the set of Spin(7)-structures on MM is

E(Φ)=12M|TΦ|2volΦ\displaystyle E(\Phi)=\frac{1}{2}\int_{M}|T_{\Phi}|^{2}\operatorname{vol}_{\Phi} (3.1)

where TΦT_{\Phi} is the torsion of Φ\Phi and the norm and the volume form are with respect to the metric induced by Φ\Phi.

The natural question here is whether, given a Spin(7)-structure Φ0\Phi_{0}, there is a “best” Spin(7)-structure. An obvious way to study this question is to consider the negative gradient flow of EE.

The most general flow of Spin(7)-structures [Kar08] is given by

Φt=AΦ=(h+X)Φ\displaystyle\frac{\partial\Phi}{\partial t}=A\diamond\Phi=(h+X)\diamond\Phi (3.2)

where Aij=hij+XijA_{ij}=h_{ij}+X_{ij} with hijS2h_{ij}\in S^{2} and XijΩ72X_{ij}\in\Omega^{2}_{7}. Thus, the most general flow of Spin(7)-structures is given by a time-dependent family of symmetric 22-tensors and a time-dependent family of 22-forms lying in the subspace Ω72\Omega^{2}_{7}. In this case the evolution of the metric, the inverse of the metric and the volume form are given by

gt=2hij,gij1t=2hij,volt=trhvol.\displaystyle\frac{\partial g}{\partial t}=2h_{ij},\ \ \ \ \frac{\partial g^{-1}_{ij}}{\partial t}=-2h_{ij},\ \ \ \ \frac{\partial\operatorname{vol}}{\partial t}=\operatorname{tr}h\operatorname{vol}. (3.3)

We start by considering the variation of the torsion TT with respect to variation of the Spin(7)-structures.

Proposition 3.2.

[Kar08, Thm. 3.4] Let (Φt)t(ε,ε)(\Phi_{t})_{t\in(-\varepsilon,\varepsilon)} be a smooth family of Spin(7)-structures. By (3.2), we can write t|t=0Φt=AΦ\left.\frac{\partial}{\partial t}\right|_{t=0}\Phi_{t}=A\diamond\Phi for some hΓ(S2TM)h\in\Gamma(S^{2}T^{*}M), XΩ72X\in\Omega^{2}_{7}. If TtT_{t} is the torsion of Φt\Phi_{t} then

t|t=0(Tt)m;ab=(hapTm;pbhbpTm;pa)+π7(bhamahbm)+XapTm;pbXbpTm;pa+π7(mXab).\displaystyle\left.\frac{\partial}{\partial t}\right|_{t=0}(T_{t})_{m;ab}=(h_{ap}T_{m;pb}-h_{bp}T_{m;pa})+\pi_{7}(\nabla_{b}h_{am}-\nabla_{a}h_{bm})+X_{ap}T_{m;pb}-X_{bp}T_{m;pa}+\pi_{7}(\nabla_{m}X_{ab}). (3.4)

Let EE be the energy functional from (3.1). The following result motivates our formulation of the flow, cf. Definition 3.2.

Lemma 3.3.

Let (Φt)t(ε,ε)(\Phi_{t})_{t\in(-\varepsilon,\varepsilon)} be a smooth family of Spin(7)-structures and t|t=0Φt=(h+X)Φ\left.\frac{\partial}{\partial t}\right|_{t=0}\Phi_{t}=(h+X)\diamond\Phi for some hΓ(S2TM)h\in\Gamma(S^{2}T^{*}M), XΩ72X\in\Omega^{2}_{7}. The gradient of the energy functional E(Φt)E(\Phi_{t}) from (3.1) is given by

ddt|t=0E(Φt)\displaystyle\left.\frac{d}{dt}\right|_{t=0}E(\Phi_{t}) =ham(12Ram(T8g)am4Tb;alTm;lb+4Tm;alTb;lbTa;lbTm;lb\displaystyle=\int h_{am}\left(\frac{1}{2}R_{am}-(\mathcal{L}_{T_{8}}g)_{am}-4T_{b;al}T_{m;lb}+4T_{m;al}T_{b;lb}-T_{a;lb}T_{m;lb}\right.
+12|T|2gam)volXabdivTabvol.\displaystyle\qquad\qquad\quad\left.+\frac{1}{2}|T|^{2}g_{am}\right)\operatorname{vol}-\int X_{ab}\operatorname{div}T_{ab}\operatorname{vol}. (3.5)
Proof.

We use Proposition 3.2 and (3.3) to compute

ddt|t=0E(Φt)\displaystyle\left.\frac{d}{dt}\right|_{t=0}E(\Phi_{t}) =ddt|t=012M(Tt)m;ab(Tt)n;cdgmngacgbdvol\displaystyle=\left.\frac{d}{dt}\right|_{t=0}\frac{1}{2}\int_{M}(T_{t})_{m;ab}(T_{t})_{n;cd}\ g^{mn}g^{ac}g^{bd}\operatorname{vol}
=MTm;abtTm;abvol\displaystyle=\int_{M}T_{m;ab}\partial_{t}T_{m;ab}\operatorname{vol}
M(hmnTm;abTn;ab+hacTm;abTm;cb+hbdTm;abTm;ad)vol\displaystyle\quad-\int_{M}(h_{mn}T_{m;ab}T_{n;ab}+h_{ac}T_{m;ab}T_{m;cb}+h_{bd}T_{m;ab}T_{m;ad})\operatorname{vol}
+12M|T|2trhvol.\displaystyle\quad+\frac{1}{2}\int_{M}|T|^{2}\operatorname{tr}h\operatorname{vol}. (3.6)

We calculate the first term on the right hand side of (3.6). Using (2.10) for the π7\pi_{7} component and integration by parts, we have

Tm;abtTm;abvol\displaystyle\int T_{m;ab}\partial_{t}T_{m;ab}\operatorname{vol} =Tm;ab[(hapTm;pbhbpTm;pa)+π7(bhamahbm)]vol\displaystyle=\int T_{m;ab}[(h_{ap}T_{m;pb}-h_{bp}T_{m;pa})+\pi_{7}(\nabla_{b}h_{am}-\nabla_{a}h_{bm})]\operatorname{vol}
+Tm;ab[XapTm;pbXbpTm;pa+π7(mXab)]vol\displaystyle\quad+\int T_{m;ab}[X_{ap}T_{m;pb}-X_{bp}T_{m;pa}+\pi_{7}(\nabla_{m}X_{ab})]\operatorname{vol}
=[2hapTm;pbTm;ab+2π7(bham)Tm;ab]vol\displaystyle=\int[2h_{ap}T_{m;pb}T_{m;ab}+2\pi_{7}(\nabla_{b}h_{am})T_{m;ab}]\operatorname{vol}
+[2XapTm;pbTm;ab+π7(mXab)Tm;ab]vol\displaystyle\quad+\int[2X_{ap}T_{m;pb}T_{m;ab}+\pi_{7}(\nabla_{m}X_{ab})T_{m;ab}]\operatorname{vol}
=[2hapTm;pbTm;ab+(12bham14ihjmΦijba)Tm;ab]vol\displaystyle=\int\left[2h_{ap}T_{m;pb}T_{m;ab}+\left(\frac{1}{2}\nabla_{b}h_{am}-\frac{1}{4}\nabla_{i}h_{jm}\Phi_{ijba}\right)T_{m;ab}\right]\operatorname{vol}
+(14mXabTm;ab18mXijΦijabTm;ab)vol\displaystyle\quad+\int\left(\frac{1}{4}\nabla_{m}X_{ab}T_{m;ab}-\frac{1}{8}\nabla_{m}X_{ij}\Phi_{ijab}T_{m;ab}\right)\operatorname{vol}
=(2hapTm;pbTm;ab+12bhamTm;ab32ihjmTm;ij)vol\displaystyle=\int\left(2h_{ap}T_{m;pb}T_{m;ab}+\frac{1}{2}\nabla_{b}h_{am}T_{m;ab}-\frac{3}{2}\nabla_{i}h_{jm}T_{m;ij}\right)\operatorname{vol}
+mXabTm;abvol\displaystyle\quad+\int\nabla_{m}X_{ab}T_{m;ab}\operatorname{vol}
=(2hapTm;pbTm;ab+2bhamTm;ab)vol\displaystyle=\int\left(2h_{ap}T_{m;pb}T_{m;ab}+2\nabla_{b}h_{am}T_{m;ab}\right)\operatorname{vol}
XabdivTabvol,\displaystyle\quad-\int X_{ab}\operatorname{div}T_{ab}\operatorname{vol},

where (divT)ab=mTm;ab(\operatorname{div}T)_{ab}=\nabla_{m}T_{m;ab} and we used the fact that Tm;pbTm;abT_{m;pb}T_{m;ab} is symmetric in a,ba,b while XabX_{ab} is skew-symmetric in the second equality, the fact that Tm;abΩ1Ω72T_{m;ab}\in\Omega^{1}\otimes\Omega^{2}_{7} and (2.6) in the second last equality and integration by parts in the last equality. Integrating by parts and then using the Spin(7)-Bianchi identity (2.30) gives

2bhamTm;abvol\displaystyle 2\int\nabla_{b}h_{am}T_{m;ab}\operatorname{vol} =2hambTm;abvol\displaystyle=-2\int h_{am}\nabla_{b}T_{m;ab}\operatorname{vol}
=2ham(mTb;ab+2Tb;alTm;lb2Tm;alTb;lb+14Rmbab18RmbklΦklab)vol\displaystyle=-2\int h_{am}\left(\nabla_{m}T_{b;ab}+2T_{b;al}T_{m;lb}-2T_{m;al}T_{b;lb}+\frac{1}{4}R_{mbab}-\frac{1}{8}R_{mbkl}\Phi_{klab}\right)\operatorname{vol}
=2ham(14Ram+mTb;ab+2Tb;alTm;lb2Tm;alTb;lb)vol.\displaystyle=-2\int h_{am}\left(-\frac{1}{4}R_{am}+\nabla_{m}T_{b;ab}+2T_{b;al}T_{m;lb}-2T_{m;al}T_{b;lb}\right)\operatorname{vol}.

from which we infer

Tm;abtTm;abvol\displaystyle\int T_{m;ab}\partial_{t}T_{m;ab}\operatorname{vol} =ham(12Ram2mTb;ab4Tb;alTm;lb+4Tm;alTb;lb2Tl;mbTl;ab)vol\displaystyle=\int h_{am}\left(\frac{1}{2}R_{am}-2\nabla_{m}T_{b;ab}-4T_{b;al}T_{m;lb}+4T_{m;al}T_{b;lb}-2T_{l;mb}T_{l;ab}\right)\operatorname{vol}
XabdivTabvol.\displaystyle\quad-\int X_{ab}\operatorname{div}T_{ab}\operatorname{vol}. (3.7)

Using (3.1) in (3.6) and the expression of T8g\mathcal{L}_{T_{8}}g from (2.40), we get

ddt|t=0E(Φt)\displaystyle\left.\frac{d}{dt}\right|_{t=0}E(\Phi_{t}) =ham(12Ram2mTb;ab4Tb;alTm;lb+4Tm;alTb;lb2Tl;mbTl;ab\displaystyle=\int h_{am}\left(\frac{1}{2}R_{am}-2\nabla_{m}T_{b;ab}-4T_{b;al}T_{m;lb}+4T_{m;al}T_{b;lb}-2T_{l;mb}T_{l;ab}\right.
Tm;lbTa;lbTl;abTl;mbTl;baTl;bm+12|T|2gam)vol\displaystyle\qquad\qquad\quad\left.-T_{m;lb}T_{a;lb}-T_{l;ab}T_{l;mb}-T_{l;ba}T_{l;bm}+\frac{1}{2}|T|^{2}g_{am}\right)\operatorname{vol}
XabdivTabvol\displaystyle\quad-\int X_{ab}\operatorname{div}T_{ab}\operatorname{vol}
=ham(12Ram(T8g)am4Tb;alTm;lb+4Tm;alTb;lbTa;lbTm;lb\displaystyle=\int h_{am}\left(\frac{1}{2}R_{am}-(\mathcal{L}_{T_{8}}g)_{am}-4T_{b;al}T_{m;lb}+4T_{m;al}T_{b;lb}-T_{a;lb}T_{m;lb}\right.
+12|T|2gam)volXabdivTabvol,\displaystyle\qquad\qquad\quad\left.+\frac{1}{2}|T|^{2}g_{am}\right)\operatorname{vol}-\int X_{ab}\operatorname{div}T_{ab}\operatorname{vol},

which completes the proof. ∎

Remark 3.4.

A similar calculation for HH-structures is done in [FLME22, Prop. 1.44]. The terms coming from Φ\Phi and gg are explicit for us as we are looking at a particular HH-structure.

We are interested in the negative gradient flow of E(Φt)E(\Phi_{t}) and based on the above computations, we propose the following flow of Spin(7)-structures.

Definition 3.5.

Let (M8,Φ0)(M^{8},\Phi_{0}) be a compact manifold with a Spin(7)-structure. Let (TT)ij=8Tb;ilTj;lb8Tj;ilTb;lb+2Ti;lbTj;lb(T*T)_{ij}=8T_{b;il}T_{j;lb}-8T_{j;il}T_{b;lb}+2T_{i;lb}T_{j;lb}. A gradient flow of Spin(7)-structures is the following initial value problem

{Φt=(Ric+2(T8g)+TT|T|2g+2divT)Φ,Φ(0)=Φ0.\displaystyle\left\{\begin{array}[]{rl}&\dfrac{\partial\Phi}{\partial t}=\left(-\mathrm{Ric}+2(\mathcal{L}_{T_{8}}g)+T*T-|T|^{2}g+2\operatorname{div}T\right)\diamond\Phi,\\ &\Phi(0)=\Phi_{0}.\end{array}\right. (GF)

It follows from (3.3) and (2.33) that along (GF), the underlying metric and volume form evolve as

tgij\displaystyle\partial_{t}g_{ij} =2Rij+4(T8g)ij+16Tb;ilTj;lb16Tj;ilTb;lb+4Ti;lbTj;lb2|T|2gij,\displaystyle=-2R_{ij}+4(\mathcal{L}_{T_{8}}g)_{ij}+16T_{b;il}T_{j;lb}-16T_{j;il}T_{b;lb}+4T_{i;lb}T_{j;lb}-2|T|^{2}g_{ij}, (3.10)
t(g1)ij\displaystyle\partial_{t}(g^{-1})_{ij} =2Rij4(T8g)ij16Tb;ilTj;lb+16Tj;ilTb;lb4Ti;lbTj;lb+2|T|2gij,\displaystyle=2R_{ij}-4(\mathcal{L}_{T_{8}}g)_{ij}-16T_{b;il}T_{j;lb}+16T_{j;il}T_{b;lb}-4T_{i;lb}T_{j;lb}+2|T|^{2}g_{ij}, (3.11)
tvol\displaystyle\partial_{t}\operatorname{vol} =(4divT8+6|T|2)vol.\displaystyle=-(4\operatorname{div}T_{8}+6|T|^{2})\operatorname{vol}. (3.12)

We discuss the effect of scaling on tensors induced from a Spin(7)-structure. It follows from [DLE24, §4.3.1] that if Φ\Phi is a Spin(7)-structure then so is Φ~=c4Φ\widetilde{\Phi}=c^{4}\Phi, c>0c>0 constant. In this case, g~=c2g,g1~=c2g1\widetilde{g}=c^{2}g,\ \widetilde{g^{-1}}=c^{-2}g^{-1} and vol~=c8vol\widetilde{\operatorname{vol}}=c^{8}\operatorname{vol}. Moreover,

~=,T~=c2T,andRic~=Ric.\displaystyle\widetilde{\nabla}=\nabla,\ \ \widetilde{T}=c^{2}T,\ \text{and}\ \widetilde{\mathrm{Ric}}=\mathrm{Ric}.

The natural parabolic rescaling for a geometric evolution equation involves scaling the time variable t by c2tc^{2}t, when the space variable scales by cc and hence, keeping in mind that taking \diamond with Φ\Phi involves contraction on one index, we see that each term on the right hand side of (GF) indeed has the correct scaling. We record for future reference that for Φ~=c4Φ\widetilde{\Phi}=c^{4}\Phi

|~jRm~|g~=c(2+j)|jRm|g,|~jT~|g~=c(1+j)|jT|g.\displaystyle|\widetilde{\nabla}^{j}\widetilde{\mathrm{Rm}}|_{\widetilde{g}}=c^{-(2+j)}|\nabla^{j}\mathrm{Rm}|_{g},\ \ \ |\widetilde{\nabla}^{j}\widetilde{T}|_{\widetilde{g}}=c^{-(1+j)}|\nabla^{j}T|_{g}. (3.13)

In particular, E(c4Φ)=c6E(Φ)E(c^{4}\Phi)=c^{6}E(\Phi) is a positively homogeneous functional and hence an application of Euler’s theorem for positively homogeneous functional shows that the critical points of EE are torsion-free Spin(7)-structures, which in fact, are absolute minimizers and thus the flow (GF) is capable of detecting torsion-free Spin(7) structures.

4.   Short-time existence and uniqueness

In this section we establish short-time existence and uniqueness of the flow (GF) of Spin(7)-structures, using a modification of the DeTurck’s trick and the explicit computation of the principal symbols of the second order linear differential operators defining the flow. We first calculate the principal symbols of the highest order terms on the right hand side of (GF), i.e., Ric,T8g\mathrm{Ric},\mathcal{L}_{T_{8}}g and divT\operatorname{div}T in § 4.2. We use these to show that (GF) is a weakly parabolic PDE and the failure of parabolicity is only due to the diffeomorphism invariance of the tensors involved. We use a modification of the DeTurck’s trick from the Ricci flow to prove the short-time existence and uniqueness of solution in § 4.3.

4.1 Differential operators, ellipticity, and parabolicity

We give a brief review of parabolic PDEs and the existence and uniqueness of solutions of such equations. Other sources for the discussion below are [CK04, §3.2][AH11, §5.1],  [Top06, §4] and [DGK23, §6.1] (for the G2\mathrm{G_{2}} case).

Let E,FE,F be vector bundles over a Riemannian manifold (M,g)(M,g) and let L:Γ(E)Γ(F)L\colon\Gamma(E)\to\Gamma(F) be a linear differential operator of order mm. We write LL, for every xMx\in M, in terms of local frames for EE and MM, as

L(σ)b(x)=l=0m[L^l(x)]ab,i1,,il[i1,,illσ(x)]a=l=0m[L^l(x)]b(lσ(x))L(\sigma)^{b}(x)=\sum_{l=0}^{m}[\hat{L}_{l}(x)]_{a}^{b,i_{1},\ldots,i_{l}}[\nabla^{l}_{i_{1},\ldots,i_{l}}\sigma(x)]^{a}=\sum_{l=0}^{m}[\hat{L}_{l}(x)]^{b}(\nabla^{l}\sigma(x)) (4.1)

where for each l=0,1,,ml=0,1,\ldots,m, we write lσΓ((TM)lE)\nabla^{l}\sigma\in\Gamma((T^{*}M)^{\otimes l}\otimes E) to denote the ll-th covariant derivative of σ\sigma, and L^lΓ((TM)lHom(E,F))\hat{L}_{l}\in\Gamma((TM)^{\otimes l}\otimes\mathrm{Hom}(E,F)). Here the index aa corresponds to a local frame for EE and the index bb corresponds to a local frame for FF.

For any such linear differential operator, we define its principal symbol so that for each xMx\in M and ξTxM\xi\in T^{*}_{x}M, the map

σξ(L):ExFx\sigma_{\xi}(L)\colon E_{x}\to F_{x}

is the linear homomorphism

b =[L^m(x)]b(ξ,,ξ,σ),\displaystyle=[\hat{L}_{m}(x)]^{b}(\xi,\ldots,\xi,\sigma), (4.2)
=[L^m(x)]ab,i1,,imξi1ξimσa.\displaystyle=[\hat{L}_{m}(x)]^{b,i_{1},\ldots,i_{m}}_{a}\xi_{i_{1}}\cdots\xi_{i_{m}}\sigma^{a}.

The principal symbol satisfies the fundamental properties

σξ(P+Q)=σξ(P)+σξ(Q),σξ(PQ)=σξ(P)σξ(Q),\sigma_{\xi}(P+Q)=\sigma_{\xi}(P)+\sigma_{\xi}(Q),\qquad\sigma_{\xi}(P\circ Q)=\sigma_{\xi}(P)\circ\sigma_{\xi}(Q),

whenever PP, QQ are linear differential operators so that either P+QP+Q or PQP\circ Q is well defined. We have the following

Definition 4.1.

A linear differential operator L:Γ(E)Γ(F)L\colon\Gamma(E)\to\Gamma(F) is called elliptic if for any xMx\in M, ξTxM\xi\in T^{*}_{x}M, ξ0\xi\neq 0, the principal symbol σξ(L):ExFx\sigma_{\xi}(L)\colon E_{x}\to F_{x} is a linear isomorphism.

Let EE be a vector bundle over MM with a fibre metric ,\langle\cdot,\cdot\rangle. Consider a second order linear differential operator L:Γ(E)Γ(E)L\colon\Gamma(E)\to\Gamma(E). If there is a constant c>0c>0 such that for any ξTxM\xi\in T^{*}_{x}M, ξ0\xi\neq 0 and vExv\in E_{x}, we have

σξ(L)(v),vc|ξ|2|v|2,\langle\sigma_{\xi}(L)(v),v\rangle\geq c|\xi|^{2}|v|^{2},

then LL is called strongly elliptic.

Definition 4.2.

Let EE, FF be vector bundles over MM, let 𝒰Γ(E)\mathcal{U}\subseteq\Gamma(E) be open, and let P:𝒰Γ(F)P\colon\mathcal{U}\to\Gamma(F) be a nonlinear differential operator. The operator PP is called elliptic at v𝒰v\in\mathcal{U} if the linearization

DvP\displaystyle D_{v}P :Γ(E)Γ(F),\displaystyle\colon\Gamma(E)\to\Gamma(F),
(DvP)(w)\displaystyle(D_{v}P)(w) :=dds|s=0P(v+sw),\displaystyle:={\left.{\frac{d}{ds}}\right|}_{{s=0}}P(v+sw),

is an elliptic linear differential operator.

Similarly, if P:𝒰Γ(E)P\colon\mathcal{U}\to\Gamma(E) is a second order differential operator and EE is endowed with a bundle metric ,\langle\cdot,\cdot\rangle, we say that PP is strongly elliptic at σ𝒰\sigma\in\mathcal{U} if its linearization DσP:Γ(E)Γ(E)D_{\sigma}P\colon\Gamma(E)\to\Gamma(E) is a strongly elliptic linear differential operator.

A nonlinear evolution equation of the form tσ=P(σ)\frac{\partial}{\partial t}\sigma=P(\sigma), where σ𝒰\sigma\in\mathcal{U}, is called parabolic at σ\sigma if PP is strongly elliptic at σ\sigma.

The importance of the above definition is due to the following standard result.

Theorem 4.3.

Let MM be a Riemannian manifold, let EE be a vector bundle over MM endowed with a fibre metric ,\langle\cdot,\cdot\rangle, and let 𝒰Γ(E)\mathcal{U}\subseteq\Gamma(E) be open. Let P:𝒰Γ(E)P\colon\mathcal{U}\to\Gamma(E) be a second order quasilinear differential operator, which is strongly elliptic at σ0𝒰\sigma_{0}\in\mathcal{U}. Then there exists ϵ>0\epsilon>0 and for any t[0,ϵ)t\in[0,\epsilon) a unique σ(t)𝒰\sigma(t)\in\mathcal{U}, such that

σ(t)t=P(σ(t)),σ(0)=σ0.\frac{\partial\sigma(t)}{\partial t}=P(\sigma(t)),\qquad\sigma(0)=\sigma_{0}. (4.3)

That is, a nonlinear evolution equation tσ=P(σ)\frac{\partial}{\partial t}\sigma=P(\sigma) which is parabolic at σ0\sigma_{0} has a unique short time smooth solution with initial condition σ(0)=σ0\sigma(0)=\sigma_{0}. ∎

4.2 Principal Symbols

Let Φ\Phi be a Spin(7)-structure and consider a variation (Φt)t(ε,ε)(\Phi_{t})_{t\in(\varepsilon,\varepsilon)} with Φ0=Φ\Phi_{0}=\Phi and

Φ˙ijkl=t|t=0Φijkl=AipΦpjkl+AjpΦipkl+AkpΦijpl+AlpΦijkp,\displaystyle\dot{\Phi}_{ijkl}=\left.\frac{\partial}{\partial t}\right|_{t=0}\Phi_{ijkl}=A_{ip}\Phi_{pjkl}+A_{jp}\Phi_{ipkl}+A_{kp}\Phi_{ijpl}+A_{lp}\Phi_{ijkp}, (4.4)

where Aij=hij+XijA_{ij}=h_{ij}+X_{ij}. Since the variation of the associated Riemannian metric is t|t=0gt=2h\left.\frac{\partial}{\partial t}\right|_{t=0}g_{t}=2h hence the linearization DΓiarD\Gamma^{r}_{ia} of the Christoffel symbols satisfies

δqrDΓiar=ihaq+ahiqqhia,\delta_{qr}D\Gamma^{r}_{ia}=\nabla_{i}h_{aq}+\nabla_{a}h_{iq}-\nabla_{q}h_{ia}, (4.5)

where \nabla is the Levi-Civita connection of the metric gg associated to the Spin(7)-structure Φ\Phi.

It follows that the principal symbol of DΓiarD\Gamma^{r}_{ia}, for any non-zero ξTxM\xi\in T_{x}^{*}M, is

σ(δrqDΓiar)(x,ξ)(Φ˙)=(ξihaq+ξahiqξqhia).\sigma(\delta_{rq}D\Gamma^{r}_{ia})(x,\xi)(\dot{\Phi})=(\xi_{i}h_{aq}+\xi_{a}h_{iq}-\xi_{q}h_{ia}). (4.6)

Now we compute the linearization D(Φ)mijklD(\nabla\Phi)_{mijkl} of mΦijkl\nabla_{m}\Phi_{ijkl}:

D(Φ)(Φ˙)mijkl=DΓmir(Φ˙)ΦrjklDΓmjr(Φ˙)ΦirklDΓmkr(Φ˙)ΦijrlDΓmlr(Φ˙)Φijkr+mΦ˙ijkl.\displaystyle D(\nabla\Phi)(\dot{\Phi})_{mijkl}=-D\Gamma^{r}_{mi}(\dot{\Phi})\Phi_{rjkl}-D\Gamma^{r}_{mj}(\dot{\Phi})\Phi_{irkl}-D\Gamma^{r}_{mk}(\dot{\Phi})\Phi_{ijrl}-D\Gamma^{r}_{ml}(\dot{\Phi})\Phi_{ijkr}+\nabla_{m}\dot{\Phi}_{ijkl}.

The principal symbol of the differential operator D(Φ)D(\nabla\Phi) is:

σ(D(Φ))(x,ξ)(Φ˙)mijkl\displaystyle\sigma(D(\nabla\Phi))(x,\xi)(\dot{\Phi})_{mijkl} =(ξmhiq+ξihmqξqhmi)Φqjkl(ξmhjq+ξjhmqξqhmj)Φiqkl\displaystyle=-(\xi_{m}h_{iq}+\xi_{i}h_{mq}-\xi_{q}h_{mi})\Phi_{qjkl}-(\xi_{m}h_{jq}+\xi_{j}h_{mq}-\xi_{q}h_{mj})\Phi_{iqkl}
(ξmhkq+ξkhmqξqhmk)Φijql(ξmhlq+ξlhmqξqhml)Φijkq\displaystyle\quad-(\xi_{m}h_{kq}+\xi_{k}h_{mq}-\xi_{q}h_{mk})\Phi_{ijql}-(\xi_{m}h_{lq}+\xi_{l}h_{mq}-\xi_{q}h_{ml})\Phi_{ijkq}
+ξm(hipΦpjkl+hjpΦipkl+hkpΦijpl+hlpΦijkp\displaystyle\quad+\xi_{m}(h_{ip}\Phi_{pjkl}+h_{jp}\Phi_{ipkl}+h_{kp}\Phi_{ijpl}+h_{lp}\Phi_{ijkp}
+XipΦpjkl+XjpΦipkl+XkpΦijpl+XlpΦijkp)\displaystyle\qquad\qquad+X_{ip}\Phi_{pjkl}+X_{jp}\Phi_{ipkl}+X_{kp}\Phi_{ijpl}+X_{lp}\Phi_{ijkp})
=(ξihmqξqhmi)Φqjkl(ξjhmqξqhmj)Φiqkl\displaystyle=-(\xi_{i}h_{mq}-\xi_{q}h_{mi})\Phi_{qjkl}-(\xi_{j}h_{mq}-\xi_{q}h_{mj})\Phi_{iqkl}
(ξkhmqξqhmk)Φijql(ξlhmqξqhml)Φijkq\displaystyle\quad-(\xi_{k}h_{mq}-\xi_{q}h_{mk})\Phi_{ijql}-(\xi_{l}h_{mq}-\xi_{q}h_{ml})\Phi_{ijkq}
+ξm(XipΦpjkl+XjpΦipkl+XkpΦijpl+XlpΦijkp).\displaystyle\quad+\xi_{m}(X_{ip}\Phi_{pjkl}+X_{jp}\Phi_{ipkl}+X_{kp}\Phi_{ijpl}+X_{lp}\Phi_{ijkp}).

Namely

σ(D(Φ))(x,ξ)(Φ˙)mijkl\displaystyle\sigma(D(\nabla\Phi))(x,\xi)(\dot{\Phi})_{mijkl} =(ξihmqξqhmi)Φqjkl(ξjhmqξqhmj)Φiqkl\displaystyle=-(\xi_{i}h_{mq}-\xi_{q}h_{mi})\Phi_{qjkl}-(\xi_{j}h_{mq}-\xi_{q}h_{mj})\Phi_{iqkl} (4.7)
(ξkhmqξqhmk)Φijql(ξlhmqξqhml)Φijkq\displaystyle\quad-(\xi_{k}h_{mq}-\xi_{q}h_{mk})\Phi_{ijql}-(\xi_{l}h_{mq}-\xi_{q}h_{ml})\Phi_{ijkq}
+ξm(XipΦpjkl+XjpΦipkl+XkpΦijpl+XlpΦijkp).\displaystyle\quad+\xi_{m}(X_{ip}\Phi_{pjkl}+X_{jp}\Phi_{ipkl}+X_{kp}\Phi_{ijpl}+X_{lp}\Phi_{ijkp}).

The linearization DTm;ibDT_{m;ib} of the torsion Tm;ib=196mΦijklΦbjklT_{m;ib}=\frac{1}{96}\nabla_{m}\Phi_{ijkl}\Phi_{bjkl} is given by

DT(Φ˙)m;ib=196D(Φ)mijklΦbjkl+l.o.t..DT(\dot{\Phi})_{m;ib}=\frac{1}{96}D(\nabla\Phi)_{mijkl}\Phi_{bjkl}+\textrm{l.o.t.}. (4.8)

Hence, its principal symbol is given by

σ(DT)(x,ξ)(Φ˙)m;ib\displaystyle\sigma(DT)(x,\xi)(\dot{\Phi})_{m;ib} =196[(ξihmqξqhmi)Φqjkl(ξjhmqξqhmj)Φiqkl\displaystyle=\frac{1}{96}\big{[}-(\xi_{i}h_{mq}-\xi_{q}h_{mi})\Phi_{qjkl}-(\xi_{j}h_{mq}-\xi_{q}h_{mj})\Phi_{iqkl}
(ξkhmqξqhmk)Φijql(ξlhmqξqhml)Φijkq\displaystyle\qquad\qquad-(\xi_{k}h_{mq}-\xi_{q}h_{mk})\Phi_{ijql}-(\xi_{l}h_{mq}-\xi_{q}h_{ml})\Phi_{ijkq}
+ξm(XiqΦqjkl+XjqΦiqkl+XkqΦijql+XlqΦijkq)]Φbjkl\displaystyle\qquad\qquad+\xi_{m}(X_{iq}\Phi_{qjkl}+X_{jq}\Phi_{iqkl}+X_{kq}\Phi_{ijql}+X_{lq}\Phi_{ijkq})\big{]}\Phi_{bjkl} (4.9)

We use the contraction identities in (2.21)-(2.24) and (2.10) to compute the principal symbol of DTDT,

σ(DT)(x,ξ)(Φ˙)m;ib=(14(ξbhimξihmb+ξjhmqΦibjq)+ξmXib).\displaystyle\sigma(DT)(x,\xi)(\dot{\Phi})_{m;ib}=\left(\frac{1}{4}(\xi_{b}h_{im}-\xi_{i}h_{mb}+\xi_{j}h_{mq}\Phi_{ibjq})+\xi_{m}X_{ib}\right). (4.10)

Using (4.10) and (2.39), the principal symbols of divTib\operatorname{div}T_{ib} and (T8g)il(\mathcal{L}_{T_{8}}g)_{il} are

σ(DdivT)(x,ξ)(Φ˙)ib\displaystyle\sigma(D\operatorname{div}T)(x,\xi)(\dot{\Phi})_{ib} =14(ξmξbhimξmξihmb+ξmξjhmqΦibjq)+|ξ|2Xib,\displaystyle=\frac{1}{4}(\xi_{m}\xi_{b}h_{im}-\xi_{m}\xi_{i}h_{mb}+\xi_{m}\xi_{j}h_{mq}\Phi_{ibjq})+|\xi|^{2}X_{ib}, (4.11)
σ(DT8g)(x,ξ)(Φ˙)li\displaystyle\sigma(D\mathcal{L}_{T_{8}}g)(x,\xi)(\dot{\Phi})_{li} =14(ξlξmhim+ξiξmhlm2ξiξltrh+ξlξjhmqΦimjq+ξiξjhmqΦlmjq)\displaystyle=\frac{1}{4}\left(\xi_{l}\xi_{m}h_{im}+\xi_{i}\xi_{m}h_{lm}-2\xi_{i}\xi_{l}\operatorname{tr}h+\xi_{l}\xi_{j}h_{mq}\Phi_{imjq}+\xi_{i}\xi_{j}h_{mq}\Phi_{lmjq}\right)
+ξlξmXim+ξiξmXlm\displaystyle\quad+\xi_{l}\xi_{m}X_{im}+\xi_{i}\xi_{m}X_{lm}
=14(ξlξmhim+ξiξmhlm2ξiξltrh+)+ξlξmXim+ξiξmXlm.\displaystyle=\frac{1}{4}\left(\xi_{l}\xi_{m}h_{im}+\xi_{i}\xi_{m}h_{lm}-2\xi_{i}\xi_{l}\operatorname{tr}h+\right)+\xi_{l}\xi_{m}X_{im}+\xi_{i}\xi_{m}X_{lm}. (4.12)

Recall that the Ricci curvature in terms of the torsion tensor is given by

Rij=4iTa;ja4aTi;ja8Ti;jbTa;ba+8Ta;jbTi;ba.\displaystyle R_{ij}=4\nabla_{i}T_{a;ja}-4\nabla_{a}T_{i;ja}-8T_{i;jb}T_{a;ba}+8T_{a;jb}T_{i;ba}.

From the computations in the previous section we obtain that the symbol of DRicD\mathrm{Ric} is given by

σ(DRic)(x,ξ)(Φ˙)ij\displaystyle\sigma(D\mathrm{Ric})(x,\xi)(\dot{\Phi})_{ij} =4σ(DT)ia;ja4σ(DT)ai;ja\displaystyle=4\sigma(D\nabla T)_{ia;ja}-4\sigma(D\nabla T)_{ai;ja}
=(ξiξahajξiξjtrh+4ξiξaXja)(ξaξjhia+|ξ|2hij+4ξaξiXja)\displaystyle=(\xi_{i}\xi_{a}h_{aj}-\xi_{i}\xi_{j}\operatorname{tr}h+4\xi_{i}\xi_{a}X_{ja})-(-\xi_{a}\xi_{j}h_{ia}+|\xi|^{2}h_{ij}+4\xi_{a}\xi_{i}X_{ja})

which simplifies to

σ(DRic)(x,ξ)(Φ˙)jk\displaystyle\sigma(D\mathrm{Ric})(x,\xi)(\dot{\Phi})_{jk} =|ξ|2hijξiξjtrh+(ξiξahaj+ξaξjhia).\displaystyle=-|\xi|^{2}h_{ij}-\xi_{i}\xi_{j}\operatorname{tr}h+(\xi_{i}\xi_{a}h_{aj}+\xi_{a}\xi_{j}h_{ia}). (4.13)

Similarly, the principal symbol of the scalar curvature is

σ(DR)(x,ξ)(Φ˙)\displaystyle\sigma(DR)(x,\xi)(\dot{\Phi}) =2|ξ|2trh+2ξjξahja.\displaystyle=-2|\xi|^{2}\operatorname{tr}h+2\xi_{j}\xi_{a}h_{ja}. (4.14)

4.3 Failure of parabolicity of the flow

Consider the differential operator L:Γ(S2TM)Ω72(M)Ω4(M)L:\Gamma(S^{2}T^{*}M)\oplus\Omega^{2}_{7}(M)\rightarrow\Omega^{4}(M)

L(h,X)\displaystyle L(h,X) =hΦ+XΦ.\displaystyle=h\diamond\Phi+X\diamond\Phi.

Clearly, (GF) is tΦ(t)=L(Ric+2(T8g)+(TT)|T|2g,2divT)\partial_{t}\Phi(t)=L(-\mathrm{Ric}+2(\mathcal{L}_{T_{8}}g)+(T*T)-|T|^{2}g,2\operatorname{div}T). Since Ric\mathrm{Ric} and TT are diffeomorphism invariant tensors, φL(Φ)=L(φΦ)\varphi^{*}L(\Phi)=L(\varphi^{*}\Phi) for the operator in (GF). For the purposes of short time existence of the flow we are only interested in the highest order term so we instead consider the operator (which we still LL)

L(Φ)=L(Ric+2T8g,2divT)=(Ric+2T8g+2divT)Φ.\displaystyle L(\Phi)=L(-\mathrm{Ric}+2\mathcal{L}_{T_{8}}g,2\operatorname{div}T)=(-\mathrm{Ric}+2\mathcal{L}_{T_{8}}g+2\operatorname{div}T)\diamond\Phi. (4.15)

Moreover, using Proposition 2.4, we will view L:Γ(S2TM)Ω72Γ(S2TM)Ω72L:\Gamma(S^{2}T^{*}M)\oplus\Omega^{2}_{7}\rightarrow\Gamma(S^{2}T^{*}M)\oplus\Omega^{2}_{7}. Since the bundle metric (h1,X1),(h2,X2)=h1,h2+X1,X2\langle(h_{1},X_{1}),(h_{2},X_{2})\rangle=\langle h_{1},h_{2}\rangle+\langle X_{1},X_{2}\rangle on Γ(S2TM)Ω72(M)\Gamma(S^{2}T^{*}M)\oplus\Omega^{2}_{7}(M) is uniformly equivalent to the natural inner product on Ω17354(M)\Omega^{4}_{1\oplus 7\oplus 35}(M), the operator LL is strongly elliptic if and only if there is a constant c>0c>0 such that for any xMx\in M, ξTxM\xi\in T^{*}_{x}M, ξ0\xi\neq 0, and any (h,X)Γ(S2TM)Ω72(M)(h,X)\in\Gamma(S^{2}T^{*}M)\oplus\Omega^{2}_{7}(M), we have

σξ(L)(h,X),(h,X)c|(h,X)|2=c(|h|2+|X|2).\displaystyle\langle\sigma_{\xi}(L)(h,X),(h,X)\rangle\geq c|(h,X)|^{2}=c(|h|^{2}+|X|^{2}). (4.16)

We will see below that the operator LL is, in fact, not elliptic (and hence (GF) not parabolic) but the failure of ellipticity is only due to the diffeomorphism invariance of the tensors in the definition of LL. As a result, we use a modified DeTurck’s trick to prove short-time existence in Theorem 4.11.

We first note, using (4.13), (4.12) and (4.11), that

σξ(L)(h,X)ij\displaystyle\sigma_{\xi}(L)(h,X)_{ij} =|ξ|2hij12(ξiξahaj+ξaξjhia)+2ξiξmXjm+2ξjξmXim\displaystyle=|\xi|^{2}h_{ij}-\frac{1}{2}(\xi_{i}\xi_{a}h_{aj}+\xi_{a}\xi_{j}h_{ia})+2\xi_{i}\xi_{m}X_{jm}+2\xi_{j}\xi_{m}X_{im}
+12(ξmξjhimξmξihmj+ξmξkhmqΦijkq)+2|ξ|2Xij.\displaystyle\quad+\frac{1}{2}(\xi_{m}\xi_{j}h_{im}-\xi_{m}\xi_{i}h_{mj}+\xi_{m}\xi_{k}h_{mq}\Phi_{ijkq})+2|\xi|^{2}X_{ij}. (4.17)

In order to analyze the kernel of σξ(L)\sigma_{\xi}(L), we define the map

δ:Γ(TM)Ω4,δW=WΦ.\delta^{*}\colon\Gamma(TM)\to\Omega^{4},\qquad\delta^{*}W=\mathcal{L}_{W}\Phi.

From (2.41) we have

δW=WΦ=(12(Wg)+T(W)+(W)7)Φ.\delta^{*}W=\mathcal{L}_{W}\Phi=\left(\tfrac{1}{2}(\mathcal{L}_{W}g)+T(W)+(\nabla W)_{7}\right)\diamond\Phi. (4.18)
Proposition 4.4.

Let δ:Γ(TM)Ω4\delta^{*}\colon\Gamma(TM)\to\Omega^{4} be as in (4.18). For any nonzero ξTxM\xi\in T^{*}_{x}M, we have

jk =12(ξjWk+ξkWj),\displaystyle=\tfrac{1}{2}(\xi_{j}W_{k}+\xi_{k}W_{j}), (4.19)
[π7σξ(δ)(W)]jk\displaystyle[\pi_{7}\circ\sigma_{\xi}(\delta^{*})(W)]_{jk} =18ξjWk18ξkWj18ξaWbΦabjk,\displaystyle=\frac{1}{8}\xi_{j}W_{k}-\frac{1}{8}\xi_{k}W_{j}-\frac{1}{8}\xi_{a}W_{b}\Phi_{abjk},

and σξ(δ):TxMΛ4(TxM)\sigma_{\xi}(\delta^{*})\colon T^{*}_{x}M\to\Lambda^{4}(T^{*}_{x}M) is injective. Here π7\pi_{7} denotes the identification of Ω74\Omega^{4}_{7} component of σξ(δ)\sigma_{\xi}(\delta^{*}) with an element in Ω72\Omega^{2}_{7}.

Proof.

The expressions in (4.19) follow from (4.18), because (Wg)jk=jWk+kWj(\mathcal{L}_{W}g)_{jk}=\nabla{j}W_{k}+\nabla{k}W_{j} and ((W)7)jk=14((W)skew)jk18((W)skew)abΦabjk=18jWk18kWj18aWbΦabjk((\nabla W)_{7})_{jk}=\frac{1}{4}((\nabla W)_{\text{skew}})_{jk}-\frac{1}{8}((\nabla W)_{\text{skew}})_{ab}\Phi_{abjk}=\frac{1}{8}\nabla_{j}W_{k}-\frac{1}{8}\nabla_{k}W_{j}-\frac{1}{8}\nabla_{a}W_{b}\Phi_{abjk}. Suppose Wkerσξ(δ)W\in\ker\sigma_{\xi}(\delta^{*}). In particular we get ξjWk+ξkWj=0\xi_{j}W_{k}+\xi_{k}W_{j}=0. Multiplying by ξjWk\xi_{j}W_{k} and summing, we obtain

0=|ξ|2|W|2+W,ξ2,0=|\xi|^{2}|W|^{2}+\langle W,\xi\rangle^{2},

which implies that W=0W=0 as ξ0\xi\neq 0, so σξ(δ)\sigma_{\xi}(\delta^{*}) is injective. ∎

Recall that L(Φ)L(\Phi) is invariant under diffeomorphisms, that is,

L(φΦ)=φ(L(Φ)),L(\varphi^{*}\Phi)=\varphi^{*}(L(\Phi)),

for any diffeomorphism φ:MM\varphi\colon M\to M. It follows that for any vector field WΓ(TM)W\in\Gamma(TM), we have

W(L(Φ))=DΦL(WΦ).\mathcal{L}_{W}(L(\Phi))=D_{\Phi}L(\mathcal{L}_{W}\Phi). (4.20)

Since WW(L(Φ))W\mapsto\mathcal{L}_{W}(L(\Phi)) is a first order linear differential operator on WW, whereas

WDΦL(WΦ)=(DΦLδ)(W)W\mapsto D_{\Phi}L(\mathcal{L}_{W}\Phi)=(D_{\Phi}L\circ\delta^{*})(W)

is a priori a third order differential operator, it follows that

σξ(DΦLδ)=σξ(DΦL)σξ(δ)=0.\sigma_{\xi}(D_{\Phi}L\circ\delta^{*})=\sigma_{\xi}(D_{\Phi}L)\circ\sigma_{\xi}(\delta^{*})=0.

and hence

im(σξ(δ))={12ξiVj+12ξjVi+18ξiVj18ξjVi18ξaVbΦabij:VTxM}kerσξ(DΦL).\text{im}(\sigma_{\xi}(\delta^{*}))=\left\{\frac{1}{2}\xi_{i}V_{j}+\frac{1}{2}\xi_{j}V_{i}+\frac{1}{8}\xi_{i}V_{j}-\frac{1}{8}\xi_{j}V_{i}-\frac{1}{8}\xi_{a}V_{b}\Phi_{abij}:V\in T^{*}_{x}M\right\}\subseteq\ker\sigma_{\xi}(D_{\Phi}L). (4.21)

Hence, by the injectivity of σξ(δ)\sigma_{\xi}(\delta^{*}), the principal symbol of LL has

dimker(σ(L))8=dimM\displaystyle\text{dim}\ker(\sigma(L))\geq 8=\text{dim}\ M (4.22)

that is due to diffeomorphism invariance, so LL is never an elliptic differential operator.

Remark 4.5.

The relation in (4.21) can be checked directly by putting hij=12(ξiVj+ξjVi)h_{ij}=\frac{1}{2}(\xi_{i}V_{j}+\xi_{j}V_{i}) and Xij=18(ξiVjξjViξaVbΦabij)X_{ij}=\frac{1}{8}(\xi_{i}V_{j}-\xi_{j}V_{i}-\xi_{a}V_{b}\Phi_{abij}) in (4.3) (which is, of course, expected).

We now show that the dimension of kernel of σ(L)\sigma(L) is at most 88. To do so, we introduce the following two operators. Consider

B1\displaystyle B_{1} :S2(TxM)TxM,\displaystyle\colon S^{2}(T^{*}_{x}M)\to T^{*}_{x}M, B1(h)k\displaystyle B_{1}(h)_{k} =ξahak12ξktrh,\displaystyle=\xi_{a}h_{ak}-\tfrac{1}{2}\xi_{k}\operatorname{tr}h, (4.23)
B2\displaystyle B_{2} :TxMΛ72(TxM),\displaystyle\colon T^{*}_{x}M\to\Lambda^{2}_{7}(T^{*}_{x}M), B2(W)ij\displaystyle B_{2}(W)_{ij} =18(ξiWjξjWiξaWbΦabij).\displaystyle=\frac{1}{8}(\xi_{i}W_{j}-\xi_{j}W_{i}-\xi_{a}W_{b}\Phi_{abij}).

We recall that the map B1B_{1} is the symbol of the Bianchi map 𝒮2Ω1\mathcal{S}^{2}\to\Omega^{1} given by hihik12k(trh)h\mapsto\nabla_{i}h_{ik}-\frac{1}{2}\nabla_{k}(\operatorname{tr}h). The Ricci curvature Ric\mathrm{Ric} lies in the kernel of the Bianchi map by the twice contracted Riemannian second Bianchi identity. The map B2B_{2} is the π7σξ(δ)\pi_{7}\circ\sigma_{\xi}(\delta^{*}) map in (4.19) and is precisely the symbol of the operator W(W)7W\mapsto(\nabla W)_{7}. The reason we need the map B2B_{2} is because while doing a modified version of the DeTurck’s trick, we need to add a term of the form WΦ\mathcal{L}_{W}\Phi to the right hand side of (GF), and the operator (W)7,WΓ(TM)(\nabla W)_{7},W\in\Gamma(TM) shows up in the Ω74\Omega^{4}_{7} part of WΦ\mathcal{L}_{W}\Phi, by equation (2.41).

It is convenient to define the operator B~:S2Ω72Ω1\widetilde{B}\colon S^{2}\oplus\Omega^{2}_{7}\to\Omega^{1} by

B~(h,X)k=B1(h)k2σξ(DΦT8)(h,X)k.\tilde{B}(h,X)_{k}=B_{1}(h)_{k}-2\sigma_{\xi}(D_{\Phi}T_{8})(h,X)_{k}. (4.24)

We can rewrite the components of σξ(DL)\sigma_{\xi}(DL) in terms of the map B~\widetilde{B}.

Proposition 4.6.

Let LL be the operator as in (4.15). In terms of the maps B1,B2B_{1},B_{2} of (4.23) and B~\widetilde{B} of (4.24), the 1+351+35 and 77 parts of the principal symbol of linearization DΦLD_{\Phi}L can be expressed as

π1+35σξ(DL)(h,X)ij\displaystyle\pi_{1+35}\circ\sigma_{\xi}(DL)(h,X)_{ij} =|ξ|2hijξiB~(h,X)jξjB~(h,X)i,\displaystyle=|\xi|^{2}h_{ij}-\xi_{i}\widetilde{B}(h,X)_{j}-\xi_{j}\widetilde{B}(h,X)_{i}, (4.25)
π7σξ(DL)(h,X)ij\displaystyle\pi_{7}\circ\sigma_{\xi}(DL)(h,X)_{ij} =2|ξ|2Xij4(B2(B~(h,X))ij+14(ξiξmhmjξjξmhimξaξmhmbΦabij)\displaystyle=2|\xi|^{2}X_{ij}-4(B_{2}(\widetilde{B}(h,X))_{ij}+\frac{1}{4}\left(\xi_{i}\xi_{m}h_{mj}-\xi_{j}\xi_{m}h_{im}-\xi_{a}\xi_{m}h_{mb}\Phi_{abij}\right)
+ξiξmXjmξjξmXimξaξmXbmΦabij.\displaystyle\quad+\xi_{i}\xi_{m}X_{jm}-\xi_{j}\xi_{m}X_{im}-\xi_{a}\xi_{m}X_{bm}\Phi_{abij}.
=2|ξ2|Xij8(B2(B~(h,X)))ij+2ξjξmXim2ξiξmXjm+2ξkξmXqmΦkqij\displaystyle=2|\xi^{2}|X_{ij}-8(B_{2}(\widetilde{B}(h,X)))_{ij}+2\xi_{j}\xi_{m}X_{im}-2\xi_{i}\xi_{m}X_{jm}+2\xi_{k}\xi_{m}X_{qm}\Phi_{kqij}
Proof.

We note that (4.3) implies

π1+35σξ(DL)(h,X)ij\displaystyle\pi_{1+35}\circ\sigma_{\xi}(DL)(h,X)_{ij} =|ξ|2hij12(ξiξahaj+ξaξjhia)+2ξiξmXjm+2ξjξmXim.\displaystyle=|\xi|^{2}h_{ij}-\frac{1}{2}(\xi_{i}\xi_{a}h_{aj}+\xi_{a}\xi_{j}h_{ia})+2\xi_{i}\xi_{m}X_{jm}+2\xi_{j}\xi_{m}X_{im}. (4.26)

Further, the definitions of the maps B1B_{1}, B~\widetilde{B} and equations (4.10), (2.39) give

B~(h,X)k\displaystyle\widetilde{B}(h,X)_{k} =ξihik12ξktrh2(14(ξmhkmξktrh+ξjhmqΦkmjq)+ξmXkm)\displaystyle=\xi_{i}h_{ik}-\frac{1}{2}\xi_{k}\operatorname{tr}h-2\left(\frac{1}{4}(\xi_{m}h_{km}-\xi_{k}\operatorname{tr}h+\xi_{j}h_{mq}\Phi_{kmjq})+\xi_{m}X_{km}\right)
=12ξihik2ξmXkm\displaystyle=\frac{1}{2}\xi_{i}h_{ik}-2\xi_{m}X_{km} (4.27)

and hence ξihik=2B~(h,X)k+4ξmXkm\xi_{i}h_{ik}=2\widetilde{B}(h,X)_{k}+4\xi_{m}X_{km}. Substituting this in (4.26) gives the first equation in (4.25). We see from (4.3) that

π7σξ(DL)(h,X)ij\displaystyle\pi_{7}\circ\sigma_{\xi}(DL)(h,X)_{ij} =2|ξ|2Xij+12(ξjξmhmiξiξmhmj+ξkξmhmqΦkqij).\displaystyle=2|\xi|^{2}X_{ij}+\frac{1}{2}(\xi_{j}\xi_{m}h_{mi}-\xi_{i}\xi_{m}h_{mj}+\xi_{k}\xi_{m}h_{mq}\Phi_{kqij}). (4.28)
Using the expression for ξihik=2B~(h,X)k+4ξmXim\xi_{i}h_{ik}=2\widetilde{B}(h,X)_{k}+4\xi_{m}X_{im} in (4.28) gives
=2|ξ|2Xij+12(ξj(2B~(h,X)i+4ξmXim)ξi(2B~(h,X)j+4ξmXjm)\displaystyle=2|\xi|^{2}X_{ij}+\frac{1}{2}\left(\xi_{j}(2\widetilde{B}(h,X)_{i}+4\xi_{m}X_{im})-\xi_{i}(2\widetilde{B}(h,X)_{j}+4\xi_{m}X_{jm})\right.
+ξk(2B~(h,X)q+4ξmXqm)Φkqij),\displaystyle\qquad\qquad\qquad\qquad\left.+\xi_{k}(2\widetilde{B}(h,X)_{q}+4\xi_{m}X_{qm})\Phi_{kqij}\right), (4.29)
which on using the definition of the map B2B_{2} from (4.23) give
=2|ξ2|Xij8(B2(B~(h,X)))ij+2ξjξmXim2ξiξmXjm+2ξkξmXqmΦkqij\displaystyle=2|\xi^{2}|X_{ij}-8(B_{2}(\widetilde{B}(h,X)))_{ij}+2\xi_{j}\xi_{m}X_{im}-2\xi_{i}\xi_{m}X_{jm}+2\xi_{k}\xi_{m}X_{qm}\Phi_{kqij}

which is the second equation in (4.25). ∎

Remark 4.7.

The operator B~\widetilde{B} plays a role similar to the role of the Bianchi operator B1B_{1} in the analysis of the principal symbol of the Ricci tensor, for instance in the Ricci flow and for the analysis of the symbol of large class of flows of G2\mathrm{G_{2}}-structures as in [DGK23].

Our main goal is to prove that the dimension of kernel of the symbol of the operator LL is 88 which will prove that the diffeomorphism invariance of Ric\mathrm{Ric} and TT are the only reason for the failure of parabolicity of (GF) and hence we can use a modified version of the DeTurck’s trick. To calculate an upper bound on dimker(σ(DL))\text{dim}\ \ker(\sigma(DL)), we compute the adjoint of the map B~\widetilde{B}.

Let YΓ(TM)Y\in\Gamma(TM). Then

B~((h,X)),Y\displaystyle\langle\widetilde{B}((h,X)),Y\rangle =(12ξihik2ξmXkm)Yk\displaystyle=\left(\frac{1}{2}\xi_{i}h_{ik}-2\xi_{m}X_{km}\right)Y_{k}
=14hik(ξiYk+ξkYi)2Xkm(18ξmYkξkYm18ξaYbΦabmk),\displaystyle=\frac{1}{4}h_{ik}(\xi_{i}Y_{k}+\xi_{k}Y_{i})-2X_{km}\left(\frac{1}{8}\xi_{m}Y_{k}-\xi_{k}Y_{m}-\frac{1}{8}\xi_{a}Y_{b}\Phi_{abmk}\right),

where we used (2.6) for XΩ72(M)X\in\Omega^{2}_{7}(M). Thus, B~:Ω1S2Ω72\widetilde{B}^{*}:\Omega^{1}\to S^{2}\oplus\Omega^{2}_{7}, B~(Y)=(B~1(Y),B~2(Y))\widetilde{B}^{*}(Y)=(\widetilde{B}_{1}^{*}(Y),\widetilde{B}^{*}_{2}(Y)) with

B~1(Y)ij\displaystyle\widetilde{B}^{*}_{1}(Y)_{ij} =14(ξiYk+ξkYi)\displaystyle=\frac{1}{4}(\xi_{i}Y_{k}+\xi_{k}Y_{i}) (4.30)
B~2(Y)ij\displaystyle\widetilde{B}^{*}_{2}(Y)_{ij} =14(ξiYjξjYiξaYbΦabij).\displaystyle=-\frac{1}{4}(\xi_{i}Y_{j}-\xi_{j}Y_{i}-\xi_{a}Y_{b}\Phi_{abij}).
Lemma 4.8.

The map B~:Ω1S2Ω72\widetilde{B}^{*}\colon\Omega^{1}\to S^{2}\oplus\Omega^{2}_{7} is injective. Consequently, dim(kerB~)=35\dim(\ker\widetilde{B})=35.

Proof.

Let YkerB~Y\in\ker\widetilde{B}^{*}. Then B~1(Y)=0\widetilde{B}^{*}_{1}(Y)=0 and hence

0\displaystyle 0 =|B~1(Y)|2\displaystyle=|\widetilde{B}^{*}_{1}(Y)|^{2}
=116(ξiYk+ξkYi)(ξiYk+ξkYi)\displaystyle=\frac{1}{16}(\xi_{i}Y_{k}+\xi_{k}Y_{i})(\xi_{i}Y_{k}+\xi_{k}Y_{i})
=18(|ξ|2|Y|2+ξ,Y2)\displaystyle=\frac{1}{8}(|\xi|^{2}|Y|^{2}+\langle\xi,Y\rangle^{2})

and since ξ0\xi\neq 0, Y=0Y=0. This proves that B~\widetilde{B}^{*} is injective and hence dimim(B~)=8\text{dim}\ \text{im}(\widetilde{B}^{*})=8. Since S2Ω72=kerB~imB~S^{2}\oplus\Omega^{2}_{7}=\ker\widetilde{B}^{*}\oplus\text{im}\ \widetilde{B}^{*}, we get that dim(kerB~)=36+78=35\text{dim}(\ker\widetilde{B})=36+7-8=35. ∎

We prove our main result in this section on the dim ker(σ(DL))\ker(\sigma(DL)).

Proposition 4.9.

Consider the operator LL from (4.15). Then ker(σξ(DL))=im(σξ(δ))\ker(\sigma_{\xi}(DL))=\text{im}(\sigma_{\xi}(\delta^{*})) and hence dimker(σ(DL))=8=dimM\text{dim}\ker(\sigma(DL))=8=\text{dim}\ M and the failure of parabolicity of (GF) is only due to diffeomorphism invariance of the tensors involved.

Proof.

We observe from Proposition 4.6, equation (4.25) that

σξ(DL)kerB~(h,X)ij\displaystyle\sigma_{\xi}(DL)\mid_{\ker\widetilde{B}}(h,X)_{ij} =|ξ|2hij+2|ξ|2Xij+2ξjξmXim2ξiξmXjm+2ξkξmXqmΦkqij\displaystyle=|\xi|^{2}h_{ij}+2|\xi|^{2}X_{ij}+2\xi_{j}\xi_{m}X_{im}-2\xi_{i}\xi_{m}X_{jm}+2\xi_{k}\xi_{m}X_{qm}\Phi_{kqij} (4.31)

and hence if (h,X)ker(σξ(DL))kerB~(h,X)\in\ker(\sigma_{\xi}(DL))\cap\ker\widetilde{B} then h=0h=0 and

|ξ|2Xij+ξjξmXimξiξmXjm+ξkξmXqmΦkqij\displaystyle|\xi|^{2}X_{ij}+\xi_{j}\xi_{m}X_{im}-\xi_{i}\xi_{m}X_{jm}+\xi_{k}\xi_{m}X_{qm}\Phi_{kqij} =0.\displaystyle=0. (4.32)

Multiplying both sides by ξi\xi_{i} gives

|ξ|2ξiXij+ξiξjξmXim|ξ|2ξmXjm+ξiξkξmXqmΦkqij=0\displaystyle|\xi|^{2}\xi_{i}X_{ij}+\xi_{i}\xi_{j}\xi_{m}X_{im}-|\xi|^{2}\xi_{m}X_{jm}+\xi_{i}\xi_{k}\xi_{m}X_{qm}\Phi_{kqij}=0
which on using the fact that ξiξj\xi_{i}\xi_{j} is symmetric in ii and jj whereas XijX_{ij} and Φijkl\Phi_{ijkl} are skew-symmetric in i,ji,j and ξ0\xi\neq 0 gives
ξiXij=0,\displaystyle\xi_{i}X_{ij}=0,

and so X(ξ)=0X(\xi)=0. Therefore, from (4.32), we get |ξ|2Xij=0|\xi|^{2}X_{ij}=0 and hence X=0X=0. Thus,

ker(σξ(DL))kerB~={0}.\displaystyle\ker(\sigma_{\xi}(DL))\cap\ker\widetilde{B}=\{0\}. (4.33)

Since dimker(B~)=35\text{dim}\ker(\widetilde{B})=35 by Lemma 4.8, we get

dimker(B~)36+735=8,\displaystyle\text{dim}\ker(\widetilde{B})\leq 36+7-35=8, (4.34)

which in combination with (4.22) gives dimker(σξ(DL))=8\text{dim}\ker(\sigma_{\xi}(DL))=8. ∎

4.4 A modified DeTurck’s trick

In this section we prove that, given a background Spin(7)-structure Φ~\widetilde{\Phi}, it is always possible to modify the operator LL from (4.15) to an operator which is strongly elliptic at Φ~\widetilde{\Phi}, that is an operator L~\widetilde{L} whose symbol satisfies

[σξ(DΦ~L~)](h,X),(h,X)c|ξ|2|(h,X)|2=c|ξ|2(|h|2+|X|2)\langle[\sigma_{\xi}(D_{\tilde{\Phi}}\widetilde{L})](h,X),(h,X)\rangle\geq c|\xi|^{2}|(h,X)|^{2}=c|\xi|^{2}(|h|^{2}+|X|^{2})

for some constant c>0c>0. We do this by doing a modification of the DeTurck’s which was originally formulated to give an alternative proof of the short-time existence and uniqueness of solutions to the Ricci flow by DeTurck [DeT83].

Let Φ~\widetilde{\Phi} be a fixed Spin(7)-structure on MM, for instance, one can take the initial Spin(7)-structure when running the flow. Motivated by the definition of the map B~\widetilde{B} in (4.24), we define the vector field W(Φ,Φ~)W(\Phi,\widetilde{\Phi}) on MM by

Wk\displaystyle W^{k} =gij(ΓijkΓ~ijk)4(T8)k=W~k4(T8)k,\displaystyle=g^{ij}\left(\Gamma^{k}_{ij}-\widetilde{\Gamma}^{k}_{ij}\right)-4(T_{8})^{k}=\widetilde{W}^{k}-4(T_{8})^{k}, (4.35)

where g~\widetilde{g} is the Riemannian metric induced by Φ~\widetilde{\Phi} and Γ~\widetilde{\Gamma} is its Christoffel symbols. The vector field W~\widetilde{W} is the same vector field which is used in the DeTurck’s trick for the Ricci flow and the vector field 4T8-4T_{8} is the extra term which we need for the DeTurck’s trick in the Spin(7)-case. We define the modified operator L~\widetilde{L} as

L~(Φ)\displaystyle\widetilde{L}(\Phi) =L(Φ)+W(Φ,Φ~)Φ\displaystyle=L(\Phi)+\mathcal{L}_{W(\Phi,\widetilde{\Phi})}\Phi (4.36)

where LL is the operator in (4.15). Using (2.41) we have

WΦ=W~4T8Φ=(12W~g2T8g+T(W~4T8)+(W~4T8)7)Φ.\displaystyle\mathcal{L}_{W}\Phi=\mathcal{L}_{\widetilde{W}-4T_{8}}\Phi=\left(\frac{1}{2}\mathcal{L}_{\widetilde{W}}g-2\mathcal{L}_{T_{8}}g+T(\widetilde{W}-4T_{8})+(\nabla\widetilde{W}-4\nabla T_{8})_{7}\right)\diamond\Phi.

Since we only need the highest order terms for the purpose of short-time existence and uniqueness, we neglect the T(W~4T8)T(\widetilde{W}-4T_{8}) term above and consequently, the operator L~\widetilde{L} is given by

L~(Φ)=(Ric+12W~g+2divT+(W~4T8)7)Φ.\displaystyle\widetilde{L}(\Phi)=\left(-\mathrm{Ric}+\frac{1}{2}\mathcal{L}_{\widetilde{W}}g+2\operatorname{div}T+(\nabla\widetilde{W}-4\nabla T_{8})_{7}\right)\diamond\Phi. (4.37)

We claim that the operator L~\widetilde{L} is elliptic. We need to calculate the principal symbol of the linearization of L~\widetilde{L}. It is well known (for example, see [CK04, §3.2]) that the linearization of W~\widetilde{W}, upto lower order terms, is

(DΦ~W~)(h,X)=2(divg~h12trg~h).\displaystyle(D_{\widetilde{\Phi}}\widetilde{W})(h,X)=2\left(\operatorname{div}_{\widetilde{g}}h-\frac{1}{2}\nabla\operatorname{tr}_{\widetilde{g}}h\right).

The factor of 2 is here because the variation of metric in the Spin(7) case is given by 2h2h. Thus,

σξ(12DW~g)(h,X)ij\displaystyle\sigma_{\xi}\left(\frac{1}{2}D\mathcal{L}_{\widetilde{W}}g\right)(h,X)_{ij} =σξ(12(i(2divhtrh)j+j(2divhtrh)i))\displaystyle=\sigma_{\xi}\left(\frac{1}{2}\left(\nabla_{i}(2\operatorname{div}h-\nabla\operatorname{tr}h)_{j}+\nabla_{j}(2\operatorname{div}h-\nabla\operatorname{tr}h)_{i}\right)\right)
=ξiξmhmj+ξjξmhmjξiξjtrh,\displaystyle=\xi_{i}\xi_{m}h_{mj}+\xi_{j}\xi_{m}h_{mj}-\xi_{i}\xi_{j}\operatorname{tr}h,

which on using (4.13) gives

σξ(D(Ric+12W~g))\displaystyle\sigma_{\xi}\left(D\left(-\mathrm{Ric}+\frac{1}{2}\mathcal{L}_{\widetilde{W}}g\right)\right) =|ξ|2hij.\displaystyle=|\xi|^{2}h_{ij}. (4.38)

We now calculate the symbol of the remaining terms in (4.37). Since

(W~7)ij\displaystyle(\nabla\widetilde{W}_{7})_{ij} =18iW~j18jW~i18aW~bΦabij\displaystyle=\frac{1}{8}\nabla_{i}\widetilde{W}_{j}-\frac{1}{8}\nabla_{j}\widetilde{W}_{i}-\frac{1}{8}\nabla_{a}\widetilde{W}_{b}\Phi_{abij}

so

(DΦ~W~7)ij\displaystyle(D_{\widetilde{\Phi}}\nabla\widetilde{W}_{7})_{ij} =18i(2divhtrh)j18j(2divhtrh)i18a(2divhtrh)bΦabij\displaystyle=\frac{1}{8}\nabla_{i}(2\operatorname{div}h-\operatorname{tr}h)_{j}-\frac{1}{8}\nabla_{j}(2\operatorname{div}h-\operatorname{tr}h)_{i}-\frac{1}{8}\nabla_{a}(2\operatorname{div}h-\operatorname{tr}h)_{b}\Phi_{abij}

and hence

σξ(DΦ~W~7)ij\displaystyle\sigma_{\xi}(D_{\widetilde{\Phi}}\nabla\widetilde{W}_{7})_{ij} =18ξi(2ξmhmjξjtrh)18ξj(2ξmhmiξitrh)18ξa(2ξmhmbξbtrh)Φabij\displaystyle=\frac{1}{8}\xi_{i}(2\xi_{m}h_{mj}-\xi_{j}\operatorname{tr}h)-\frac{1}{8}\xi_{j}(2\xi_{m}h_{mi}-\xi_{i}\operatorname{tr}h)-\frac{1}{8}\xi_{a}(2\xi_{m}h_{mb}-\xi_{b}\operatorname{tr}h)\Phi_{abij}
=14ξiξmhmj14ξjξmhmi14ξaξmhmbΦabij.\displaystyle=\frac{1}{4}\xi_{i}\xi_{m}h_{mj}-\frac{1}{4}\xi_{j}\xi_{m}h_{mi}-\frac{1}{4}\xi_{a}\xi_{m}h_{mb}\Phi_{abij}. (4.39)

Similarly,

4((T8)7)ij\displaystyle-4((\nabla T_{8})_{7})_{ij} =12i(Tm;jm)+12j(Tm;im)+12a(Tm;bm)Φabij\displaystyle=-\frac{1}{2}\nabla_{i}(T_{m;jm})+\frac{1}{2}\nabla_{j}(T_{m;im})+\frac{1}{2}\nabla_{a}(T_{m;bm})\Phi_{abij}

which on using (4.10) gives

4(σξ(DΦ~(T8)7))ij\displaystyle-4(\sigma_{\xi}(D_{\widetilde{\Phi}}(\nabla T_{8})_{7}))_{ij} =12ξi(14(ξmhjmξjtrh+ξkhmqΦjmkq)+ξmXjm)\displaystyle=-\frac{1}{2}\xi_{i}\left(\frac{1}{4}(\xi_{m}h_{jm}-\xi_{j}\operatorname{tr}h+\xi_{k}h_{mq}\Phi_{jmkq})+\xi_{m}X_{jm}\right)
+12ξj(14(ξmhimξitrh+ξkhmqΦimkq)+ξmXim)\displaystyle\quad+\frac{1}{2}\xi_{j}\left(\frac{1}{4}(\xi_{m}h_{im}-\xi_{i}\operatorname{tr}h+\xi_{k}h_{mq}\Phi_{imkq})+\xi_{m}X_{im}\right)
+12ξa(14(ξmhbmξbtrh+ξkhmqΦbmkq)+ξmXbm)Φabij\displaystyle\quad+\frac{1}{2}\xi_{a}\left(\frac{1}{4}(\xi_{m}h_{bm}-\xi_{b}\operatorname{tr}h+\xi_{k}h_{mq}\Phi_{bmkq})+\xi_{m}X_{bm}\right)\Phi_{abij}
=18ξiξmhjm+18ξjξmhim+18ξaξmhbmΦabij\displaystyle=-\frac{1}{8}\xi_{i}\xi_{m}h_{jm}+\frac{1}{8}\xi_{j}\xi_{m}h_{im}+\frac{1}{8}\xi_{a}\xi_{m}h_{bm}\Phi_{abij}
12ξiξmXjm+12ξjξmXim+12ξaξmXbmΦabij.\displaystyle\quad-\frac{1}{2}\xi_{i}\xi_{m}X_{jm}+\frac{1}{2}\xi_{j}\xi_{m}X_{im}+\frac{1}{2}\xi_{a}\xi_{m}X_{bm}\Phi_{abij}. (4.40)

Using (4.38), (4.11), (4.4) and (4.40) we compute the symbol of L~\widetilde{L} as

σξ(DL~(Φ)(h,X))ij\displaystyle\sigma_{\xi}(D\widetilde{L}(\Phi)(h,X))_{ij} =|ξ|2hij+12(ξmξjhimξmξihmj+ξmξahmbΦijab)+2|ξ|2Xij\displaystyle=|\xi|^{2}h_{ij}+\frac{1}{2}(\xi_{m}\xi_{j}h_{im}-\xi_{m}\xi_{i}h_{mj}+\xi_{m}\xi_{a}h_{mb}\Phi_{ijab})+2|\xi|^{2}X_{ij}
+14ξiξmhmj14ξjξmhmi14ξaξmhmbΦabij18ξiξmhjm+18ξjξmhim+18ξaξmhbmΦabij\displaystyle\quad+\frac{1}{4}\xi_{i}\xi_{m}h_{mj}-\frac{1}{4}\xi_{j}\xi_{m}h_{mi}-\frac{1}{4}\xi_{a}\xi_{m}h_{mb}\Phi_{abij}-\frac{1}{8}\xi_{i}\xi_{m}h_{jm}+\frac{1}{8}\xi_{j}\xi_{m}h_{im}+\frac{1}{8}\xi_{a}\xi_{m}h_{bm}\Phi_{abij}
12ξiξmXjm+12ξjξmXim+12ξaξmXbmΦabij\displaystyle\quad-\frac{1}{2}\xi_{i}\xi_{m}X_{jm}+\frac{1}{2}\xi_{j}\xi_{m}X_{im}+\frac{1}{2}\xi_{a}\xi_{m}X_{bm}\Phi_{abij}
=|ξ|2hij+2|ξ|2Xij38ξiξmhmj+38ξjξmhmi+38ξaξmhmbΦabij\displaystyle=|\xi|^{2}h_{ij}+2|\xi|^{2}X_{ij}-\frac{3}{8}\xi_{i}\xi_{m}h_{mj}+\frac{3}{8}\xi_{j}\xi_{m}h_{mi}+\frac{3}{8}\xi_{a}\xi_{m}h_{mb}\Phi_{abij}
12ξiξmXjm+12ξjξmXim+12ξaξmXbmΦabij\displaystyle\quad-\frac{1}{2}\xi_{i}\xi_{m}X_{jm}+\frac{1}{2}\xi_{j}\xi_{m}X_{im}+\frac{1}{2}\xi_{a}\xi_{m}X_{bm}\Phi_{abij} (4.41)

Using (4.41) we calculate

σξ(DL~(Φ)(h,X)),(h,X)\displaystyle\left\langle\sigma_{\xi}(D\widetilde{L}(\Phi)(h,X)),(h,X)\right\rangle =|ξ|2|h|2+2|ξ|2|X|23ξiξmhmjXij4ξiξmXjmXij\displaystyle=|\xi|^{2}|h|^{2}+2|\xi|^{2}|X|^{2}-3\xi_{i}\xi_{m}h_{mj}X_{ij}-4\xi_{i}\xi_{m}X_{jm}X_{ij}
=|ξ|2|h|2+2|ξ|2|X|23h(ξ),X(ξ)+4|X(ξ)|2\displaystyle=|\xi|^{2}|h|^{2}+2|\xi|^{2}|X|^{2}-3\langle h(\xi),X(\xi)\rangle+4|X(\xi)|^{2}
|ξ|2|h|2+2|ξ|2|X|23|h(ξ)||X(ξ)|+4|X(ξ)|2\displaystyle\geq|\xi|^{2}|h|^{2}+2|\xi|^{2}|X|^{2}-3|h(\xi)||X(\xi)|+4|X(\xi)|^{2}
which on using Young’s inequality on the 3rd term gives
|ξ|2|h|2+2|ξ|2|X|212|h(ξ)|292|X(ξ)|2+4|X(ξ)|2\displaystyle\geq|\xi|^{2}|h|^{2}+2|\xi|^{2}|X|^{2}-\frac{1}{2}|h(\xi)|^{2}-\frac{9}{2}|X(\xi)|^{2}+4|X(\xi)|^{2}
and we use |h(ξ)|2|h|2|ξ|2|h(\xi)|^{2}\leq|h|^{2}|\xi|^{2} and |X(ξ)|2|X|2|ξ|2|X(\xi)|^{2}\leq|X|^{2}|\xi|^{2} to get
|ξ|2|h|2+2|ξ|2|X|212|h|2|ξ|212|X|2|ξ|2\displaystyle\geq|\xi|^{2}|h|^{2}+2|\xi|^{2}|X|^{2}-\frac{1}{2}|h|^{2}|\xi|^{2}-\frac{1}{2}|X|^{2}|\xi|^{2}
12|ξ|2(|h|2+|X|2)=12|(h,X)|2.\displaystyle\geq\frac{1}{2}|\xi|^{2}(|h|^{2}+|X|^{2})=\frac{1}{2}|(h,X)|^{2}. (4.42)

The above computations leading to (4.42) and (4.16) prove the following

Proposition 4.10.

Let (M8,Φ~)(M^{8},\widetilde{\Phi}) be an 88-manifold with a Spin(7)-structure Φ~\widetilde{\Phi} and let L~\widetilde{L} be the operator

L~(Φ)=L(Φ)+W(Φ,Φ~)Φ=(Ric+2T8g+12Wg+2divT+(W~4T8)7)Φ,\displaystyle\widetilde{L}(\Phi)=L(\Phi)+\mathcal{L}_{W(\Phi,\widetilde{\Phi})}\Phi=\left(-\mathrm{Ric}+2\mathcal{L}_{T_{8}}g+\frac{1}{2}\mathcal{L}_{W}g+2\operatorname{div}T+(\nabla\widetilde{W}-4\nabla T_{8})_{7}\right)\diamond\Phi, (4.43)

where W(Φ,Φ~)k=gij(ΓijkΓ~ijk4(T8)k)W(\Phi,\widetilde{\Phi})^{k}=g^{ij}\left(\Gamma^{k}_{ij}-\widetilde{\Gamma}^{k}_{ij}-4(T_{8})^{k}\right). Then L~(Φ)\widetilde{L}(\Phi) is strongly elliptic at Φ~\widetilde{\Phi}. ∎

4.5 Short-time existence and uniqueness

In this section, we prove the main theorem of the paper by using the modified DeTurck’s trick whose details were given in § 4.4.

Theorem 4.11.

Let (M8,Φ0)(M^{8},\Phi_{0}) be a compact 88-manifold with a Spin(7)-structure Φ0\Phi_{0} and consider the negative gradient flow (GF) of the natural energy functional EE in (3.1). Then there exists ε>0\varepsilon>0 and a unique smooth solution Φ(t)\Phi(t) of (GF) for t[0,ε)t\in[0,\varepsilon) with ε=ε(Φ0)\varepsilon=\varepsilon(\Phi_{0}).

Proof.

Let Φ~=Φ0\widetilde{\Phi}=\Phi_{0} and define W(Φ,Φ~)W(\Phi,\widetilde{\Phi}) as in (4.35). Since the terms 8Tb;alTm;lb8Tm;alTb;lb+2Ta;lbTm;lb|T|2g8T_{b;al}T_{m;lb}-8T_{m;al}T_{b;lb}+2T_{a;lb}T_{m;lb}-|T|^{2}g are at most first order in Φ\Phi, we deduce from Proposition 4.10 that the linearization of the operator

L~(Φ)=(Ric+2(T8g)+8Tb;alTm;lb8Tm;alTb;lb+2Ta;lbTm;lb|T|2g+2divT)Φ+WΦ\displaystyle\widetilde{L}(\Phi)=\left(-\mathrm{Ric}+2(\mathcal{L}_{T_{8}}g)+8T_{b;al}T_{m;lb}-8T_{m;al}T_{b;lb}+2T_{a;lb}T_{m;lb}-|T|^{2}g+2\operatorname{div}T\right)\diamond\Phi+\mathcal{L}_{W}\Phi

is strongly elliptic. We obtain from standard parabolic theory (Theorem 4.3) that there is a unique smooth solution Φ~(t)\widetilde{\Phi}(t) for t[0,ε)t\in[0,\varepsilon) of the initial value problem

tΦ¯(t)\displaystyle\frac{\partial}{\partial t}\bar{\Phi}(t) =L~(Φ¯(t)),\displaystyle=\widetilde{L}(\bar{\Phi}(t)),
Φ¯(0)\displaystyle\bar{\Phi}(0) =Φ0.\displaystyle=\Phi_{0}.

Let φt:MM,t[0,ε)\varphi_{t}:M\to M,\ t\in[0,\varepsilon) be the one-parameter family of diffeomorphisms defined by

tφt\displaystyle\frac{\partial}{\partial t}\varphi_{t} =W(Φ¯(t),Φ~)φt,\displaystyle=-W(\bar{\Phi}(t),\widetilde{\Phi})\circ\varphi_{t},
φ0\displaystyle\varphi_{0} =IdM.\displaystyle=\operatorname{Id}_{M}.

Since MM is compact, the family of diffeomorphisms φt\varphi_{t} exists by [CK04, Lemma 3.15] as long as the solutions Φ¯(t)\bar{\Phi}(t) exists. The family Φ(t)=φt(Φ¯(t))\Phi(t)=\varphi^{*}_{t}(\bar{\Phi}(t)) then satisfies

Φ(t)t=t(φt(Φ¯(t)))\displaystyle\dfrac{\partial\Phi(t)}{\partial t}=\partial_{t}(\varphi^{*}_{t}(\bar{\Phi}(t))) =φt(W(t)Φ¯(t)+tΦ¯(t))\displaystyle=\varphi^{*}_{t}\left(\mathcal{L}_{-W(t)}\bar{\Phi}(t)+\partial_{t}\bar{\Phi}(t)\right)
=φt((W(t)Φ¯(t)Ric¯+2T¯8g¯+l.o.t.¯+2div¯T¯)Φ¯(t)Φ¯(t)+W(t)Φ¯(t))\displaystyle=\varphi^{*}_{t}\left(\left(\mathcal{L}_{-W(t)}\bar{\Phi}(t)-\overline{\mathrm{Ric}}+2\mathcal{L}_{\bar{T}_{8}}\bar{g}+\overline{\text{l.o.t.}}+2\overline{\operatorname{div}}\bar{T}\right)\diamond_{\bar{\Phi}(t)}\bar{\Phi}(t)+\mathcal{L}_{W(t)}\bar{\Phi}(t)\right)
=(Ric¯+2T¯8g¯+l.o.t.¯+2div¯T¯)(φt(Φ¯(t)))(φt(Φ¯(t)))\displaystyle=\left(-\overline{\mathrm{Ric}}+2\mathcal{L}_{\bar{T}_{8}}\bar{g}+\overline{\text{l.o.t.}}+2\overline{\operatorname{div}}\bar{T}\right)\diamond_{(\varphi^{*}_{t}(\bar{\Phi}(t)))}(\varphi^{*}_{t}(\bar{\Phi}(t)))
=(Ric+2(T8g)+8Tb;alTm;lb8Tm;alTb;lb+2Ta;lbTm;lb|T|2g+2divT)Φ\displaystyle=\left(-\mathrm{Ric}+2(\mathcal{L}_{T_{8}}g)+8T_{b;al}T_{m;lb}-8T_{m;al}T_{b;lb}+2T_{a;lb}T_{m;lb}-|T|^{2}g+2\operatorname{div}T\right)\diamond\Phi

with Φ(0)=Id(Φ¯(0))=Φ0\Phi(0)=\operatorname{Id}^{*}(\bar{\Phi}(0))=\Phi_{0} and l.o.t¯=8T¯b;alT¯m;lb8T¯m;alT¯b;lb+2T¯a;lbT¯m;lb|T¯|g¯2g¯\overline{\text{l.o.t}}=8\bar{T}_{b;al}\bar{T}_{m;lb}-8\bar{T}_{m;al}\bar{T}_{b;lb}+2\bar{T}_{a;lb}\bar{T}_{m;lb}-|\bar{T}|^{2}_{\bar{g}}\bar{g}. This proves the short-time existence of solutions to (GF).

We now prove uniqueness, by using the uniqueness of solutions to the harmonic map heat flow. Suppose Φi(t),i=1,2\Phi_{i}(t),i=1,2 are solutions to (GF) with the same initial condition. Let (Fi)t:MM(F_{i})_{t}:M\to M be a one-parameter family of diffeomorphisms given as

t(Fi)t\displaystyle\frac{\partial}{\partial t}(F_{i})_{t} =4(T8)Φi(t)(Fi)t,\displaystyle=-4(T_{8})_{\Phi_{i}(t)}\circ(F_{i})_{t}, (4.44)
(Fi)0\displaystyle(F_{i})_{0} =IdM.\displaystyle=\operatorname{Id}_{M}.

Using (2.41), we see that the family Φ¯i(t)=(Fi)tΦi(t)\bar{\Phi}_{i}(t)=(F_{i})^{*}_{t}\Phi_{i}(t) solves

tΦ¯i(t)\displaystyle\frac{\partial}{\partial t}\bar{\Phi}_{i}(t) =(Ricg¯i(t)4(¯T¯8)7+2divT¯8+l.o.t¯)Φ¯i(t)\displaystyle=\left(-\mathrm{Ric}_{\bar{g}_{i}(t)}-4(\bar{\nabla}\bar{T}_{8})_{7}+2\operatorname{div}\bar{T}_{8}+\overline{\text{l.o.t}}\right)\diamond\bar{\Phi}_{i}(t) (4.45)
Φ¯i(0)=Φ0.\displaystyle\bar{\Phi}_{i}(0)=\Phi_{0}.

If we set Φ^i(t)=((φi)t1)Φ¯i(t)\hat{\Phi}_{i}(t)=((\varphi_{i})^{-1}_{t})^{*}\bar{\Phi}_{i}(t) by defining the diffeomorphisms φi(t)\varphi_{i}(t) as t(φi)t=W~(g^i(t),g~)\partial_{t}(\varphi_{i})_{t}=-\widetilde{W}(\hat{g}_{i}(t),\widetilde{g}) then it is known [CK04, §4.3] that φi(t):MM,t[0,εi)\varphi_{i}(t):M\to M,\ t\in[0,\varepsilon_{i}^{{}^{\prime}}) is a solution to

tφi(t)\displaystyle\frac{\partial}{\partial t}\varphi_{i}(t) =Δgi(t),g0(φi)(t),\displaystyle=\Delta_{g_{i}(t),g_{0}}(\varphi_{i})(t), (4.46)
φi(0)\displaystyle\varphi_{i}(0) =IdM\displaystyle=\operatorname{Id}_{M}

which is the harmonic map heat flow from (M8,g0)(M^{8},g_{0}) to (M8,gi(t))(M^{8},g_{i}(t)) with the initial value being the identity map. Again using (2.41), we see that the maps Φ^i(t)\hat{\Phi}_{i}(t) satisfy

tΦ^i(t)\displaystyle\frac{\partial}{\partial t}\hat{\Phi}_{i}(t) =(Ricg^i(t)+12W~(g^i(t),g~)g^i(t)+(^W~4^T^8)7+2divT^8+l.o.t.^)Φ^i(t)\displaystyle=\left(-\mathrm{Ric}_{\hat{g}_{i}(t)}+\frac{1}{2}\mathcal{L}_{\widetilde{W}(\hat{g}_{i}(t),\widetilde{g})}\hat{g}_{i}(t)+(\hat{\nabla}\widetilde{W}-4\hat{\nabla}\hat{T}_{8})_{7}+2\operatorname{div}\hat{T}_{8}+\widehat{\text{l.o.t.}}\right)\diamond\hat{\Phi}_{i}(t) (4.47)
Φ^i(0)\displaystyle\hat{\Phi}_{i}(0) =Φ0.\displaystyle=\Phi_{0}.

for t[0,εi).t\in[0,\varepsilon_{i}^{{}^{\prime}}). Since we have proved above that the operator in (4.47) is parabolic, we know from the standard theory of uniqueness of parabolic equations that Φ^1(t)=Φ^2(t)\hat{\Phi}_{1}(t)=\hat{\Phi}_{2}(t) for all t[0,ε)t\in[0,\varepsilon^{{}^{\prime}}) with ε=min{ε1,ε2}\varepsilon^{{}^{\prime}}=\text{min}\{\varepsilon_{1}^{{}^{\prime}},\varepsilon_{2}^{{}^{\prime}}\} and as a result g^1(t)=g^2(t)\hat{g}_{1}(t)=\hat{g}_{2}(t) which gives φ1(t)=φ2(t)\varphi_{1}(t)=\varphi_{2}(t). Consequently

Φ¯1(t)=(φ1)tΦ^1(t)=(φ2)tΦ^2(t)=Φ¯2(t),for allt[0,ε).\displaystyle\bar{\Phi}_{1}(t)=(\varphi_{1})^{*}_{t}\hat{\Phi}_{1}(t)=(\varphi_{2})^{*}_{t}\hat{\Phi}_{2}(t)=\bar{\Phi}_{2}(t),\ \ \ \ \ \ \ \ \text{for\ all}\ \ \ t\in[0,\varepsilon^{{}^{\prime}}).

Finally, we have

0=IdMt\displaystyle 0=\frac{\partial\operatorname{Id}_{M}}{\partial t} =t((Fi)t(Fi)t1)\displaystyle=\frac{\partial}{\partial t}\left((F_{i})_{t}\circ(F_{i})^{-1}_{t}\right)
=(Fi)tt(Fi)t1+((Fi)t)((Fi)t1t)\displaystyle=\frac{\partial(F_{i})_{t}}{\partial t}\circ(F_{i})^{-1}_{t}+((F_{i})_{t})_{*}\left(\frac{\partial(F_{i})^{-1}_{t}}{\partial t}\right)
=4(T8)Φi(t)+((Fi)t)((Fi)t1t)\displaystyle=-4(T_{8})_{\Phi_{i}(t)}+((F_{i})_{t})_{*}\left(\frac{\partial(F_{i})^{-1}_{t}}{\partial t}\right)

which gives

((Fi)t1t)=4((Fi)t1)(T8)Φi(t)=4(T8)(Fi)tΦi(t)(Fi)t1=4(T8)Φ¯i(t)(Fi)t1.\displaystyle\left(\frac{\partial(F_{i})^{-1}_{t}}{\partial t}\right)=4((F_{i})^{-1}_{t})_{*}(T_{8})_{\Phi_{i}(t)}=4(T_{8})_{(F_{i})^{*}_{t}\Phi_{i}(t)}\circ(F_{i})^{-1}_{t}=4(T_{8})_{\bar{\Phi}_{i}(t)}\circ(F_{i})^{-1}_{t}.

But since we proved above that Φ¯1(t)=Φ¯2(t),\bar{\Phi}_{1}(t)=\bar{\Phi}_{2}(t), we get (F1)t1=(F2)t1(F_{1})^{-1}_{t}=(F_{2})^{-1}_{t} and hence Φ1(t)=Φ2(t)\Phi_{1}(t)=\Phi_{2}(t) for all t[0,ε)t\in[0,\varepsilon^{{}^{\prime}}) which proves the uniqueness of solutions to (GF) and completes the proof of the theorem. ∎

Remark 4.12.

A particularly interesting flow of Spin(7)-structures is the "coupling" of the Ricci flow of the metric and the isometric/harmonic flow of Spin(7)-structures

tΦ(t)=(Ric+divT)Φ.\frac{\partial}{\partial t}\Phi(t)=(-\mathrm{Ric}+\operatorname{div}T)\diamond\Phi.

This flow induces precisely the Ricci flow tg=2Ric\frac{\partial}{\partial t}g=-2\mathrm{Ric} on the metric, and the only other thing it does to the Spin(7)-structure Φ\Phi is to deform it by the isometric flow which was studied in detail in [DLE24]. This “coupling” of Ricci flow with the isometric flow has good short-time existence and uniqueness. This can be seen by going through the modified DeTurck’s trick in § 4.4 and choosing W=W~W=\widetilde{W} and then checking that the resulting modified flow is strictly parabolic. The isometric flow of Spin(7)-structures has many good properties, in particular there is an almost monotonicity formula for the solutions to the flow (see [DLE24]). It would be interesting to study whether the Ricci flow coupled with the harmonic flow has similar analytic properties.

5.   Solitons

Let M8M^{8} be an 88-manifold. A soliton for (GF) is a triple (Φ,Y,λ)(\Phi,Y,\lambda) with YΓ(TM)Y\in\Gamma(TM) and λ\lambda\in\mathbb{R} such that

(Ric+2(T8g)+TT|T|2g+2divT)Φ=λΦ+YΦ\displaystyle\left(-\mathrm{Ric}+2(\mathcal{L}_{T_{8}}g)+T*T-|T|^{2}g+2\operatorname{div}T\right)\diamond\Phi=\lambda\Phi+\mathcal{L}_{Y}\Phi (5.1)

where (TT)ij=8Tb;ilTj;lb8Tj;ilTb;lb+2Ti;lbTj;lb(T*T)_{ij}=8T_{b;il}T_{j;lb}-8T_{j;il}T_{b;lb}+2T_{i;lb}T_{j;lb}. Those Spin(7)-structures which satisfy (5.1) are special as they give self-similar solutions to (GF). Recall that a self-similar solution to (GF) is a solution of the form

Φ(t)=λ(t)4φ(t)Φ(0)\displaystyle\Phi(t)=\lambda(t)^{4}\varphi(t)^{*}\Phi(0)

where λ(t)\lambda(t) are time-dependent scalings and φ(t):MM\varphi(t):M\to M are a one-parameter family of diffeomorphisms. Note that the power of 44 on λ(t)\lambda(t) is just for convenience in calculations. A straightforward calculation (for instance see [DGK21, Lemma 2.17], [DLE24, Prop. 2.11] or [FLME22, Prop. 1.55]) shows that the solitons for (GF) and self-similar solutions are in one-to-one correspondence.

We say a soliton (Φ,Y,λ)(\Phi,Y,\lambda) is expanding if λ>0\lambda>0; steady if λ=0\lambda=0; and shrinking if λ<0\lambda<0.

We now derive the condition satisfied by the metric gg induced by Φ\Phi and divT\operatorname{div}T when (Φ,Y,λ)(\Phi,Y,\lambda) is a soliton, which we expect to have further use.

Proposition 5.1.

Let (Φ,Y,λ)(\Phi,Y,\lambda) be a solitons as defined in (5.1). Then the induced metric gg satisfies

Rij+2(T8g)ij+8Tb;ilTj;lb8Tj;ilTb;lb+2Ti;lbTj;lb|T|2gij=λ4gij+12(Yg)ij,\displaystyle-R_{ij}+2(\mathcal{L}_{T_{8}}g)_{ij}+8T_{b;il}T_{j;lb}-8T_{j;il}T_{b;lb}+2T_{i;lb}T_{j;lb}-|T|^{2}g_{ij}=\frac{\lambda}{4}g_{ij}+\frac{1}{2}(\mathcal{L}_{Y}g)_{ij}, (5.2)

and divT\operatorname{div}T satisfies

(divT)ij=12(T(Y)+(Y)7)ij.\displaystyle(\operatorname{div}T)_{ij}=\frac{1}{2}\left(T(Y)+(\nabla Y)_{7}\right)_{ij}. (5.3)
Proof.

The definition of the \diamond operator in (2.13) implies Φ=(g4Φ)\Phi=\left(\dfrac{g}{4}\diamond\Phi\right). Moreover, (2.41) gives YΦ=(12Yg+T(Y)+(Y)7)Φ\mathcal{L}_{Y}\Phi=\left(\frac{1}{2}\mathcal{L}_{Y}g+T(Y)+(\nabla Y)_{7}\right)\diamond\Phi. Putting these expressions in (5.1) and using the fact that Ω1+354\Omega^{4}_{1+35} is orthogonal to Ω74\Omega^{4}_{7} gives (5.2) and (5.3). ∎

We can prove the non-existence theorem for compact expanding solitons of (GF).

Proposition 5.2.
  1. 1.

    There are no compact expanding solitons of (GF).

  2. 2.

    The only compact steady solitons of (GF) are given by torsion-free Spin(7)-structures.

Proof.

Taking the trace of (5.2) gives

R+4divT8+8Tb;ilTi;lb+8|T8|2+2|T|28|T|2=2λ+divY\displaystyle-R+4\operatorname{div}T_{8}+8T_{b;il}T_{i;lb}+8|T_{8}|^{2}+2|T|^{2}-8|T|^{2}=2\lambda+\operatorname{div}Y

which on using the expression for the scalar curvature (2.34) simplifies to

4divT86|T|2=2λ+divY.\displaystyle-4\operatorname{div}T_{8}-6|T|^{2}=2\lambda+\operatorname{div}Y. (5.4)

Integrating (5.4) on compact MM gives

3M|T|2vol=λVol(M).\displaystyle-3\int_{M}|T|^{2}\operatorname{vol}=\lambda\text{Vol}(M).

So λ0\lambda\leq 0 and λ=0\lambda=0 if and only if T0T\equiv 0. ∎

References

  • [AH11] Ben Andrews and Christopher Hopper “The Ricci flow in Riemannian geometry. A complete proof of the differentiable 1/4-pinching sphere theorem” 2011, Lect. Notes Math. Berlin: Springer, 2011 DOI: 10.1007/978-3-642-16286-2
  • [AWW16] Bernd Ammann, Hartmut Weiss and Frederik Witt “A spinorial energy functional: critical points and gradient flow” In Math. Ann. 365.3-4, 2016, pp. 1559–1602 DOI: 10.1007/s00208-015-1315-8
  • [CK04] Bennett Chow and Dan Knopf “The Ricci flow: an introduction” 110, Math. Surv. Monogr. Providence, RI: American Mathematical Society (AMS), 2004
  • [DeT83] Dennis M. DeTurck “Deforming metrics in the direction of their Ricci tensors” In J. Differ. Geom. 18, 1983, pp. 157–162 DOI: 10.4310/jdg/1214509286
  • [DGK21] Shubham Dwivedi, Panagiotis Gianniotis and Spiro Karigiannis “A gradient flow of isometric G2\mathrm{G}_{2}-structures” In J. Geom. Anal. 31.2, 2021, pp. 1855–1933 DOI: 10.1007/s12220-019-00327-8
  • [DGK23] Shubham Dwivedi, Panagiotis Gianniotis and Spiro Karigiannis “Flows of G2G_{2}-structures, II: Curvature, torsion, symbols, and functionals”, 2023 DOI: 10.48550/arXiv.2311.05516
  • [DLE24] Shubham Dwivedi, Eric Loubeau and Henrique N. Earp “Harmonic flow of Spin(7)\mathrm{Spin}(7)-structures” In Ann. Sc. Norm. Super. Pisa Cl. Sci. (5) Vol. XXV (2024), 151-215, 2024 arXiv:2109.06340 [math.DG]
  • [Fer86] Marisa Fernández “A classification of Riemannian manifolds with structure group Spin(7){\rm Spin}(7) In Ann. Mat. Pura Appl. (4) 143, 1986, pp. 101–122 DOI: 10.1007/BF01769211
  • [FLME22] Daniel Fadel, Eric Loubeau, Andrés J. Moreno and Henrique N. Earp “Flows of geometric structures” In arXiv e-prints, 2022, pp. arXiv:2211.05197 DOI: 10.48550/arXiv.2211.05197
  • [Joy00] Dominic D. Joyce “Compact manifolds with special holonomy”, Oxford Mathematical Monographs Oxford University Press, Oxford, 2000, pp. xii+436
  • [Kar08] Spiro Karigiannis “Flows of Spin(7)-structures” In Differential geometry and its applications World Sci. Publ., Hackensack, NJ, 2008, pp. 263–277 DOI: 10.1142/9789812790613_0023
  • [Kra24] Kirill Krasnov “Dynamics of Cayley Forms”, 2024 arXiv:2403.16661 [math.DG]
  • [LS23] Eric Loubeau and Henrique N. Sá Earp “Harmonic flow of geometric structures” Id/No 23 In Ann. Global Anal. Geom. 64.4, 2023, pp. 42 DOI: 10.1007/s10455-023-09928-7
  • [Top06] Peter Topping “Lectures on the Ricci flow” 325, Lond. Math. Soc. Lect. Note Ser. Cambridge: Cambridge University Press, 2006

Institut für Mathematik, Humboldt-Universität zu Berlin, Rudower Chaussee 25, 12489 Berlin.
dwivedis@hu-berlin.de