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]

Deformation theory of nearly G2\mathrm{G}_{2} manifolds

Shubham Dwivedi                   Ragini Singhal
(January 5, 2025)
Abstract

We study the deformation theory of nearly G2\mathrm{G}_{2} manifolds. These are seven dimensional manifolds admitting real Killing spinors. We show that the infinitesimal deformations of nearly G2\mathrm{G}_{2} structures are obstructed in general. Explicitly, we prove that the infinitesimal deformations of the homogeneous nearly G2\mathrm{G}_{2} structure on the Aloff–Wallach space are all obstructed to second order. We also completely describe the cohomology of nearly G2\mathrm{G}_{2} manifolds.

2000 Mathematics Subject Classification: 53C15, 53C25, 53C29.

1.   Introduction

Given a 77-dimensional smooth manifold MM, a nearly G2\mathrm{G}_{2} structure on MM is a non-degenerate (or positive) 33-form φ\varphi such that for some non-zero real constant τ0\tau_{0},

dφ=τ0φφ\displaystyle d\varphi=\tau_{0}*_{\varphi}\varphi (1.1)

where the metric and the orientation and hence the Hodge star * are all induced by φ\varphi. The existence of a nearly G2\mathrm{G}_{2} structure was shown to be equivalent to the existence of a real Killing spinor in [BFGK91]. A Killing spinor on a Riemannian spin manifold (Mn,g)(M^{n},g) is a section of the spinor bundle μΓ((M))\mu\in\Gamma(\not{S}(M)) such that

Xμ=αXμ\displaystyle\nabla_{X}\mu=\alpha X\cdot\mu (1.2)

for any vector field XX on MM and some α\alpha\in\mathbb{C}. Here \cdot is the Clifford multiplication. It was proved by Friedrich [Fri80] that any manifold with a Killing spinor is Einstein with Ric(g)=4(n1)α2g\mathrm{Ric}(g)=4(n-1)\alpha^{2}g and one of the three cases must hold:

  • α=0\alpha=0 in which case μ\mu is a parallel spinor and MM has holonomy contained in SU(n2)\mathrm{SU}(\frac{n}{2}), Sp(n4)\mathrm{Sp}(\frac{n}{4}), G2\mathrm{G}_{2} or Spin(7)\mathrm{Spin}(7).

  • α\alpha is non-zero and is purely imaginary.

  • α\alpha is non-zero and real, in which case μ\mu is a real Killing spinor and if MM is complete then since it is positive Einstein, it is compact with π1(M)\pi_{1}(M) finite.

Given a nearly G2\mathrm{G}_{2} structure φ\varphi on MM that satisfies equation (1.1), there exists a real Killing spinor μ\mu that satisfies equation (1.2) with α=18τ0\alpha=-\frac{1}{8}\tau_{0} and vice-versa. See [BFGK91] for more details.

Using the equivalence with real Killing spinors, nearly G2\mathrm{G}_{2} structures on homogeneous spaces, excluding the case of the round 77-sphere, were classified in [FKMS97]. Their classification is based on the dimension of the space of Killing spinors KK\not{S}. They showed that 33 different types can occur:

  1. 1.

    dim(K)=1K\not{S})=1 - nearly G2\mathrm{G}_{2} structures of type 1.

  2. 2.

    dim(K)=2K\not{S})=2 - nearly G2\mathrm{G}_{2} structures of type 2.

  3. 3.

    dim(K)=3K\not{S})=3 - nearly G2\mathrm{G}_{2} structures of type 3.

A 77-dimensional manifold (M,φ)(M,\varphi) with a nearly G2\mathrm{G}_{2} structure φ\varphi is a nearly G2\mathrm{G}_{2} manifold (see §2 for more details). Other examples apart from the round S7S^{7} include the squashed S7S^{7}, Aloff–Wallach spaces N(k,l)N(k,l), the Berger space SO(5)/SO(3)\mathrm{SO}(5)/\mathrm{SO}(3) and the Stiefel manifold V5,2V_{5,2}. Another important aspect of nearly G2\mathrm{G}_{2} manifolds is that the Riemannian cone C(M)C(M) over MM has holonomy contained in the Lie group Spin(7)\mathrm{Spin}(7). In that case, the possible holonomies are Spin(7)\mathrm{Spin}(7), SU(4)\mathrm{SU}(4) or Sp(2)\mathrm{Sp}(2) depending on whether the link of the cone is a nearly G2\mathrm{G}_{2} manifold of type 1, 21,\ 2 or 33 respectively.

In this paper, we study the deformation theory of nearly G2\mathrm{G}_{2} manifolds. The infinitesimal deformations of nearly G2\mathrm{G}_{2} manifolds were studied by Alexandrov–Semmelmann in [AS12] where they identified the space of infinitesimal deformations with an eigenspace of the Laplacian acting on co-closed 33-forms on MM of type Ω273\Omega^{3}_{27}. We address the question of whether nearly G2\mathrm{G}_{2} manifolds have smooth obstructed or unobstructed deformations, i.e., whether infinitesimal deformations can be integrated to genuine deformations. This could potentially give new examples of nearly G2\mathrm{G}_{2} manifolds. Another applicability of studying the deformation theory of nearly G2\mathrm{G}_{2} manifolds can be to develop the deformation theory of Spin(7)\mathrm{Spin}(7) conifolds which are asymptotically conical and conically singular Spin(7)\mathrm{Spin}(7) manifolds, similar to the theory developed by Karigiannis–Lotay [KL20] for G2\mathrm{G}_{2} conifolds. Lehmann [Leh20] studies the deformation theory of asymptotically conical Spin(7)\mathrm{Spin}(7)-manifolds.

The study of deformation theory of special algebraic structures is not new. Deformations of Einstein metrics were studied by Koiso where he showed [Koi82, Theorem 6.12] that the infinitesimal deformations of Einstein metrics is in general obstructed, by exhibiting certain Einstein symmetric spaces which admit non-trivial infinitesimal Einstein deformations which cannot be integrated to second order. The deformation theory of nearly Kähler structures on homogeneous 66-manifolds was studied by Moroianu–Nagy–Semmelmann in [MNS08]. They identified the space of infinitesimal deformations with an eigenspace of the Laplacian acting on co-closed primitive (1,1)(1,1)-forms. Using this, they proved that the nearly Kähler structures on 3\mathbb{CP}^{3} and S3×S3S^{3}\times S^{3} are rigid and the flag manifold 𝔽3\mathbb{F}_{3} admits an 88-dimensional space of infinitesimal deformations. Later, Foscolo proved [Fos17, Theorem 5.3] that the infinitesimal deformations of the flag manifold 𝔽3\mathbb{F}_{3} are all obstructed.

Nearly G2\mathrm{G}_{2} manifolds are in many ways similar to nearly Kähler 66-manifolds. Both admit real Killing spinors and hence are positive Einstein. The minimal hypersurfaces in both nearly Kähler 6-manifolds and nealy G2\mathrm{G}_{2} manifolds behave in a similar way [Dwi19]. It was proved in [AS12] that the nearly G2\mathrm{G}_{2} structures on the squashed S7S^{7} and the Berger space SO(5)/SO(3)\mathrm{SO}(5)/\mathrm{SO}(3) are rigid while the space of infinitesimal nearly G2\mathrm{G}_{2} deformations of the Aloff–Wallach space X1,1X_{1,1} is 88-dimensional. It is therefore natural to ask whether these infinitesimal deformations are obstructed to second order.

To address this question, we use a Dirac-type operator on nearly G2\mathrm{G}_{2} manifolds (cf. equation (3.6)). The use of Dirac operators to study deformation theory has been very useful. Nordström in [Nor08] used Dirac operators to study the deformation theory of compact manifolds with special holonomy from a different point of view than Joyce [Joy00]. In particular, the mapping properties of the Dirac type operators can be used to prove slice theorems for the action of the diffeomorphism group. This approach has also been very effective in studying the deformation theory of non-compact manifolds with special holonomy, most notably by Nordström [Nor08] for asymptotically cylindrical manifolds with exceptional holonomy and by Karigiannis–Lotay [KL20] for G2\mathrm{G}_{2} conifolds. Dirac-type operators, in a way very close to the use made by the authors in this paper, were also used by Foscolo [Fos17] to study the deformation theory of nearly Kähler 66-manifolds.

We follow a strategy similar to [Fos17] in this paper. After introducing the Dirac operator and a modified Dirac operator on nearly G2\mathrm{G}_{2} manifolds in §3, we use their properties and the Hodge decomposition theorem to completely describe the cohomology of a complete nearly G2\mathrm{G}_{2} manifold. We prove our first two main results of the paper which characterize harmonic forms. These are the following.

Theorem 3.8. Let (M,φ,ψ)(M,\varphi,\psi) be a complete nearly G2\mathrm{G}_{2} manifold, not isometric to round S7S^{7}. Then every harmonic 44-form lies in Ω274\Omega^{4}_{27}. Equivalently, every harmonic 33-form lies in Ω273\Omega^{3}_{27}.

Theorem 3.9 Let (M,φ,ψ)(M,\varphi,\psi) be a complete nearly G2\mathrm{G}_{2} manifold, not isometric to round S7S^{7}. Then every harmonic 22-form lies in Ω142\Omega^{2}_{14}. Equivalently, every harmonic 55-form lies in Ω145\Omega^{5}_{14}.

We note that Theorem 3.9 was originally proved by Ball–Oliveira [BO19, Remark 15]. We give a different proof in this paper.

We use the properties of the modified Dirac operator, explicitly we use Proposition 3.7, to prove a slice theorem for the action of the diffeomorphism group on the space of nearly G2\mathrm{G}_{2} structures on MM in Proposition 4.2. Using this, in Theorem 4.3 we obtain a new proof of the identification of the space of nearly G2\mathrm{G}_{2} deformations with an eigenspace of the Laplacian acting on co-closed 33-forms of type Ω273\Omega^{3}_{27}, a result originally due to Alexandrov–Semmelmann [AS12].

To study higher order deformations of nearly G2\mathrm{G}_{2} manifolds, we use the point view of Hitchin [Hit01] where he interprets nearly G2\mathrm{G}_{2} structures as constrained critical points of a functional defined on the space Ω3×Ωexact4\Omega^{3}\times\Omega^{4}_{\operatorname{\textup{exact}}}. This approach is inspired from the work of Foscolo [Fos17] where he used similar ideas to study second order deformations of nearly Kähler structures on 66-manifolds. The advantage of this approach is that it allows us to view the nearly G2\mathrm{G}_{2} equation (2.24) as the vanishing of a smooth map (cf. equation (4.9))

Φ:Ω+,exact4×Γ(TM)Ωexact4\displaystyle\Phi:\Omega^{4}_{+,\operatorname{\textup{exact}}}\times\Gamma(TM)\longrightarrow\Omega^{4}_{\operatorname{\textup{exact}}}

where Ω+,exact4\Omega^{4}_{+,\operatorname{\textup{exact}}} denotes the space of exact positive 44-forms on MM. Thus the obstructions on the first order deformations of a nearly G2\mathrm{G}_{2} structure to be integrated to higher order deformations can be characterized by Im(DΦ)\operatorname{Im}(D\Phi) which we do in Proposition 4.6.

Finally, we use the general deformation theory of nearly G2\mathrm{G}_{2} structures developed in the first part of the paper to study the infinitesimal deformations of the Aloff–Wallach space SU(3)×SU(2)SU(2)×U(1)\frac{\mathrm{SU}(3)\times\mathrm{SU}(2)}{\mathrm{SU}(2)\times\mathrm{U}(1)}. It was expected in [Fos17] that the infinitesimal deformations of the Aloff–Wallach space might be obstructed to higher orders. In §5 we confirm this expectation. More precisely, we prove the following.

Theorem 5.1. The infinitesimal deformations of the homogeneous nearly G2\mathrm{G}_{2} structure on the Aloff–Wallach space X1,1SU(3)×SU(2)SU(2)×U(1)X_{1,1}\cong\frac{\mathrm{SU}(3)\times\mathrm{SU}(2)}{\mathrm{SU}(2)\times\mathrm{U}(1)} are all obstructed.

The proof of the above theorem is inspired from the ideas in [Fos17]. However, we note that since in the nearly G2\mathrm{G}_{2} case we only have one stable form and the other is the dual of it, unlike the nearly Kähler case, the expressions and computations involved are more complicated and the proof of the theorem is computationally much more involved.

The paper is organized as follows. We discuss some preliminaries on G2\mathrm{G}_{2} and nearly G2\mathrm{G}_{2} structures in §2. We discuss the decomposition of space of differential forms on manifolds with a G2\mathrm{G}_{2} structure. We describe some first order differential operators in §2.1 which appear throughout the paper. In §2.2, we prove many important identities for 22-forms and 33-forms on manifolds with nearly G2\mathrm{G}_{2} structures. Some of these appear to be new, at least in the present form and we believe that they will be useful in other contexts as well. We introduce the Dirac and the modified Dirac operator in §3 and use the mapping properties of the latter to prove Theorem 3.8 and Theorem 3.9. We begin the discussion on infinitesimal deformations in §4.1. We prove a slice theorem and use that to obtain a new proof of the result of Alexandrov–Semmmelmann on infinitesimal nearly G2\mathrm{G}_{2} deformations. We interpret the nearly G2\mathrm{G}_{2} equation as the vanishing of a smooth map and prove the characterization for a first order deformation of a nearly G2\mathrm{G}_{2} structure to be integrated to second order in Proposition 4.6. Finally, in §5, we prove Theorem 5.1.

Note. The almost simultaneous preprint [NS20] by Semmelmann–Nagy has some overlap with the present paper and some of the ideas involved are the same. We also characterize the cohomology of nearly G2\mathrm{G}_{2} manifolds. The second version of their paper also contains a discussion of the deformations of the Aloff–Wallach spaces.

Acknowledgments.

We are indebted to Spiro Karigiannis and Benoit Charbonneau for various discussions related to the paper and for constant encouragement and advice. We are grateful to Ben Webster for an important discussion on representation theory. We thank Gavin Ball and Gonçalo Oliveira for pointing us out to their result about harmonic 22-forms on nearly G2\mathrm{G}_{2} manifolds in their paper [BO19]. We are grateful to Gonçalo Oliveira for discussions on the material in §5. Finally, we would like to thank the referee for a very careful reading of the paper and for many useful remarks and suggestions which have improved the quality of the paper.

2.   Preliminaries on G2\mathrm{G}_{2} geometry

We start this section by defining G2\mathrm{G}_{2} structures and nearly G2\mathrm{G}_{2} structures on a seven dimensional manifold and also discuss the decomposition of space of differential forms on such a manifold. We also collect together various identities which will be used throughout the paper.

Let M7M^{7} be a smooth manifold. A G2\mathrm{G}_{2} structure on M is a reduction of the structure group of the frame bundle from GL(7,)\mathrm{GL}(7,\mathbb{R}) to the Lie group G2SO(7)\mathrm{G}_{2}\subset\mathrm{SO}(7). Such a structure exists on MM if and only if the manifold is orientable and spinnable, conditions which are respectively equivalent to the vanishing of the first and second Stiefel–Whitney classes. From the point of view of differential geometry, a G2\mathrm{G}_{2} structure on M is equivalently defined by a 33-form φ\varphi on MM that satisfies a certain pointwise algebraic non-degeneracy condition. Such a 33-form nonlinearly induces a Riemannian metric gφg_{\varphi} and an orientation volφ\operatorname{vol}_{\varphi} on MM and hence a Hodge star operator φ*_{\varphi}. We denote the Hodge dual 44-form φφ*_{\varphi}\varphi by ψ\psi. Pointwise we have |φ|=|ψ|=7|\varphi|=|\psi|=7, where the norm is taken with respect to the metric induced by φ\varphi.

Notations and conventions. 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 77. Our convention for labelling the Riemann curvature tensor is

Rijkmxm=(ijji)xk,R_{ijkm}\frac{\partial}{\partial x^{m}}=(\nabla_{i}\nabla_{j}-\nabla_{j}\nabla_{i})\frac{\partial}{\partial x^{k}},

in terms of coordinate vector fields. With this convention, the Ricci tensor is Rjk=RljklR_{jk}=R_{ljkl}, and the Ricci identity is

ijXkjiXk=RijklXl.\displaystyle\nabla_{i}\nabla_{j}X_{k}-\nabla_{j}\nabla_{i}X_{k}=-R_{ijkl}X_{l}. (2.1)

We will use the metric to identify the vector fields and 11-forms by the musical isomorphisms. As such, throughout the paper, we will use them interchangeably without mention.

We have the following contraction identities between φ\varphi and ψ\psi, whose proofs can be found in [Kar09].

φijkφabk\displaystyle\varphi_{ijk}\varphi_{abk} =giagjbgibgja+ψijab,\displaystyle=g_{ia}g_{jb}-g_{ib}g_{ja}+\psi_{ijab}, (2.2)
φijkφajk\displaystyle\varphi_{ijk}\varphi_{ajk} =6gia\displaystyle=6g_{ia} (2.3)

and

φijkψabck\displaystyle\varphi_{ijk}\psi_{abck} =gjaφibc+gjbφaic+gjcφabigiaφjbcgibφajcgicφabj,\displaystyle=g_{ja}\varphi_{ibc}+g_{jb}\varphi_{aic}+g_{jc}\varphi_{abi}-g_{ia}\varphi_{jbc}-g_{ib}\varphi_{ajc}-g_{ic}\varphi_{abj}, (2.4)
φijkψabjk\displaystyle\varphi_{ijk}\psi_{abjk} =4φiab,\displaystyle=4\varphi_{iab}, (2.5)
ψijklψabkl\displaystyle\psi_{ijkl}\psi_{abkl} =4giagjb4gibgja+2ψijab\displaystyle=4g_{ia}g_{jb}-4g_{ib}g_{ja}+2\psi_{ijab} (2.6)
ψijklψajkl\displaystyle\psi_{ijkl}\psi_{ajkl} =24gia.\displaystyle=24g_{ia}. (2.7)

A G2\mathrm{G}_{2} structure on MM induces a splitting of the spaces of differential forms on MM into irreducible G2\mathrm{G}_{2} representations. The space of 22-forms Ω2(M)\Omega^{2}(M) and 33-forms Ω3(M)\Omega^{3}(M) decompose as

Ω2(M)\displaystyle\Omega^{2}(M) =Ω72(M)Ω142(M),\displaystyle=\Omega^{2}_{7}(M)\oplus\Omega^{2}_{14}(M), (2.8)
Ω3(M)\displaystyle\Omega^{3}(M) =Ω13(M)Ω73(M)Ω273(M)\displaystyle=\Omega^{3}_{1}(M)\oplus\Omega^{3}_{7}(M)\oplus\Omega^{3}_{27}(M) (2.9)

where Ωlk\Omega^{k}_{l} has pointwise dimension ll. More precisely, we have the following description of the space of forms :

Ω72(M)\displaystyle\Omega^{2}_{7}(M) ={XφXΓ(TM)}={βΩ2(M)(φβ)=2β},\displaystyle=\{X\lrcorner\varphi\mid X\in\Gamma(TM)\}=\{\beta\in\Omega^{2}(M)\mid*(\varphi\wedge\beta)=2\beta\}, (2.10)
Ω142(M)\displaystyle\Omega^{2}_{14}(M) ={βΩ2(M)βψ=0}={βΩ2(M)(φβ)=β}.\displaystyle=\{\beta\in\Omega^{2}(M)\mid\beta\wedge\psi=0\}=\{\beta\in\Omega^{2}(M)\mid*(\varphi\wedge\beta)=-\beta\}. (2.11)

In local coordinates, the above conditions can be re-written as

βΩ72\displaystyle\beta\in\Omega^{2}_{7}\ \ \ βijψabij=4βab,\displaystyle\iff\ \ \ \beta_{ij}\psi_{abij}=4\beta_{ab}, (2.12)
βΩ142\displaystyle\beta\in\Omega^{2}_{14}\ \ βijψabij=2βabβijφijk=0.\displaystyle\ \iff\ \ \ \beta_{ij}\psi_{abij}=-2\beta_{ab}\ \ \ \iff\ \ \ \beta_{ij}\varphi_{ijk}=0. (2.13)

Similarly, for 33-forms

Ω13\displaystyle\Omega^{3}_{1} ={fφfC(M)},\displaystyle=\{f\varphi\mid f\in C^{\infty}(M)\}, (2.14)
Ω73\displaystyle\Omega^{3}_{7} ={XψXΓ(TM)}={(αφ)αΩ1},\displaystyle=\{X\lrcorner\psi\mid X\in\Gamma(TM)\}=\{*(\alpha\wedge\varphi)\mid\alpha\in\Omega^{1}\}, (2.15)
Ω273\displaystyle\Omega^{3}_{27} ={ηΩ3ηφ=0=ηψ}.\displaystyle=\{\eta\in\Omega^{3}\ \mid\ \eta\wedge\varphi=0=\eta\wedge\psi\}. (2.16)

Moreover, the space Ω273\Omega^{3}_{27} is isomorphic to the space of sections of S02(TM)S^{2}_{0}(T^{*}M), the traceless symmetric 22-tensors on M, where the isomorphism iφi_{\varphi} is given explicitly as

η=16ηijkdxidxjdxkΩ273iφhabdxadxbC(S02(TM))\displaystyle\eta=\frac{1}{6}\eta_{ijk}dx^{i}\wedge dx^{j}\wedge dx^{k}\in\Omega^{3}_{27}\ \ \ \ \ \ \ \overset{i_{\varphi}}{\longleftrightarrow}\ \ \ \ \ \ \ h_{ab}dx^{a}dx^{b}\in C^{\infty}(S^{2}_{0}(T^{*}M)) (2.17)
whereηijk=hipφpjk+hjpφipk+hkpφijp.\displaystyle\textup{where}\ \ \ \ \ \eta_{ijk}=h_{ip}\varphi_{pjk}+h_{jp}\varphi_{ipk}+h_{kp}\varphi_{ijp}.

The decompositions of Ω4(M)=Ω14(M)Ω74(M)Ω274(M)\Omega^{4}(M)=\Omega^{4}_{1}(M)\oplus\Omega^{4}_{7}(M)\oplus\Omega^{4}_{27}(M) and Ω5(M)=Ω75(M)Ω145(M)\Omega^{5}(M)=\Omega^{5}_{7}(M)\oplus\Omega^{5}_{14}(M) are obtained by taking the Hodge star of (2.9) and (2.8) respectively.

Given a G2\mathrm{G}_{2} structure φ\varphi on MM, we can decompose dφd\varphi and dψd\psi according to (2.8) and (2.9). This defines the torsion forms, which are unique differential forms τ0Ω0(M)\tau_{0}\in\Omega^{0}(M), τ1Ω1(M)\tau_{1}\in\Omega^{1}(M), τ2Ω142(M)\tau_{2}\in\Omega^{2}_{14}(M) and τ3Ω273(M)\tau_{3}\in\Omega^{3}_{27}(M) such that (see [Kar09])

dφ\displaystyle d\varphi =τ0ψ+3τ1φ+φτ3,\displaystyle=\tau_{0}\psi+3\tau_{1}\wedge\varphi+*_{\varphi}\tau_{3}, (2.18)
dψ\displaystyle d\psi =4τ1ψ+φτ2.\displaystyle=4\tau_{1}\wedge\psi+*_{\varphi}\tau_{2}. (2.19)

Let \nabla denote the Levi-Civita connection of the metric induced by the G2\mathrm{G}_{2} structure. The full torsion tensor TT of a G2\mathrm{G}_{2} structure is a 22-tensor satisfying

iφjkl\displaystyle\nabla_{i}\varphi_{jkl} =Timψmjkl,\displaystyle=T_{im}\psi_{mjkl}, (2.20)
Tlm\displaystyle T_{lm} =124(lφabc)ψmabc,\displaystyle=\frac{1}{24}(\nabla_{l}\varphi_{abc})\psi_{mabc}, (2.21)
mψijkl\displaystyle\nabla_{m}\psi_{ijkl} =Tmiφjkl+TmjφiklTmkφijl+Tmlφijk.\displaystyle=-T_{mi}\varphi_{jkl}+T_{mj}\varphi_{ikl}-T_{mk}\varphi_{ijl}+T_{ml}\varphi_{ijk}. (2.22)

The full torsion TT is related to the torsion forms by (see [Kar09])

Tlm=τ04glm(τ3)lm(τ1)lm12(τ2)lm.T_{lm}=\frac{\tau_{0}}{4}g_{lm}-(\tau_{3})_{lm}-(\tau_{1})_{lm}-\frac{1}{2}(\tau_{2})_{lm}. (2.23)
Remark 2.1.

The space Ω72\Omega^{2}_{7} is isomorphic to the space of vector fields and hence to the space of 11-forms. Thus in (2.23), we are viewing τ1\tau_{1} as an element of Ω72\Omega^{2}_{7} which justifies the expression (τ1)lm(\tau_{1})_{lm}.

A G2\mathrm{G}_{2} structure φ\varphi is called torsion-free if φ=0\nabla\varphi=0 or equivalently T=0T=0. We can now define nearly G2\mathrm{G}_{2} structures.

Definition 2.2.

A G2\mathrm{G}_{2} structure φ\varphi is a nearly G2\mathrm{G}_{2} structure if τ0\tau_{0} is the only nonvanishing component of the torsion, that is

dφ=τ0ψanddψ=0.\displaystyle d\varphi=\tau_{0}\psi\ \ \ \ \textup{and}\ \ \ \ d\psi=0. (2.24)

In this case, we see from (2.23) that Tij=τ04gijT_{ij}=\dfrac{\tau_{0}}{4}g_{ij}.

Remark 2.3.

If φ\varphi is a nearly G2\mathrm{G}_{2} structure on MM then since dφ=τ0ψd\varphi=\tau_{0}\psi, we can differentiate this to get dτ0ψ=0d\tau_{0}\wedge\psi=0 and hence dτ0=0d\tau_{0}=0, as wedge product with ψ\psi is an isomorphism from Ω71(M)\Omega^{1}_{7}(M) to Ω75(M)\Omega^{5}_{7}(M). Thus τ0\tau_{0} is a constant, if MM is connected.

Given a G2\mathrm{G}_{2} structure φ\varphi with torsion TlmT_{lm}, we have the expressions for the Ricci curvature RijR_{ij} and the scalar curvature RR of its associated metric gg which can be found in [Bry06] or [Kar09] as

Rjk\displaystyle R_{jk} =(iTjmjTim)φmkiTjlTlk+tr(T)TjkTjbTlpψlpbk,\displaystyle=(\nabla_{i}T_{jm}-\nabla_{j}T_{im})\varphi_{mki}-T_{jl}T_{lk}+\operatorname{tr}(T)T_{jk}-T_{jb}T_{lp}\psi_{lpbk}, (2.25)
R\displaystyle R =12i(τ1)i+218τ02|τ3|2+5|τ1|214|τ2|2.\displaystyle=-12\nabla_{i}(\tau_{1})_{i}+\frac{21}{8}{\tau_{0}}^{2}-|\tau_{3}|^{2}+5|\tau_{1}|^{2}-\frac{1}{4}|\tau_{2}|^{2}. (2.26)

where |C|2=CijCklgikgjl|C|^{2}=C_{ij}C_{kl}g^{ik}g^{jl} is the matrix norm in (2.26).

In particular, for a manifold MM with a nearly G2\mathrm{G}_{2} structure φ\varphi, we see that

Rij\displaystyle R_{ij} =38τ02gij,\displaystyle=\frac{3}{8}{\tau_{0}}^{2}g_{ij}, (2.27)
R\displaystyle R =218τ02.\displaystyle=\frac{21}{8}{\tau_{0}}^{2}. (2.28)

Finally, we remark that S7S^{7} with the round metric and also the squashed S7S^{7} are examples of manifolds with nearly G2\mathrm{G}_{2} structure (see [FKMS97] for more on nearly G2\mathrm{G}_{2} structures. The authors in [FKMS97] call such structures nearly parallel G2\mathrm{G}_{2} structures but we will call them nearly G2\mathrm{G}_{2} structures.) In particular, S7S^{7} with radius 11 has scalar curvature 4242, so comparing with (2.24) we get that τ0=4\tau_{0}=4.

We use the following identities throughout the paper. They are all proved in [Kar05, Lemma 2.2.1 and Lemma 2.2.3] and we collect them here for the convenience of the reader. First, we note that if α\alpha is a kk-form and ww is a vector field then

(wα)\displaystyle*(w\lrcorner\alpha) =(1)k+1(wα),\displaystyle=(-1)^{k+1}(w\wedge*\alpha), (2.29)
(wα)\displaystyle*(w\wedge\alpha) =(1)k(wα).\displaystyle=(-1)^{k}(w\lrcorner*\alpha). (2.30)

If α\alpha is a 11-form then we have the following identities

(φ(φα))\displaystyle*(\varphi\wedge*(\varphi\wedge\alpha)) =4α,\displaystyle=-4\alpha, (2.31)
ψ(φα)\displaystyle\psi\wedge*(\varphi\wedge\alpha) =0,\displaystyle=0, (2.32)
(ψ(ψα))\displaystyle*(\psi\wedge*(\psi\wedge\alpha)) =3α,\displaystyle=3\alpha, (2.33)
φ(ψα)\displaystyle\varphi\wedge*(\psi\wedge\alpha) =2(ψα).\displaystyle=2(\psi\wedge\alpha). (2.34)

Suppose ww is a vector field then we have the following identities

φ(wψ)\displaystyle\varphi\wedge(w\lrcorner\psi) =4w,\displaystyle=-4*w, (2.35)
ψ(wψ)\displaystyle\psi\wedge(w\lrcorner\psi) =0,\displaystyle=0, (2.36)
ψ(wφ)\displaystyle\psi\wedge(w\lrcorner\varphi) =3w,\displaystyle=3*w, (2.37)
φ(wφ)\displaystyle\varphi\wedge(w\lrcorner\varphi) =2(wφ).\displaystyle=2*(w\lrcorner\varphi). (2.38)

Let Θ:Ω+3Ω+4\Theta:\Omega^{3}_{+}\rightarrow\Omega^{4}_{+} be the non-linear map which associates to any G2\mathrm{G}_{2} structure φ\varphi, the dual 44-form ψ=Θ(φ)=φ\psi=\Theta(\varphi)=*\varphi with respect to the metric gφg_{\varphi}. We note that Θ1:Ω+4Ω+3\Theta^{-1}:\Omega^{4}_{+}\rightarrow\Omega^{3}_{+} is defined only when we fix the orientation on MM. See [Hit01, §8] for more details. We will need the following result from [Joy00, Proposition 10.3.5], later.

Proposition 2.4.

Suppose φ\varphi be a G2\mathrm{G}_{2} structure on MM with ψ=φ\psi=*\varphi. Let ξ\xi be a 33-form which has sufficiently small pointwise norm with respect to gφg_{\varphi} so that φ+ξ\varphi+\xi is still a positive 33-form and η\eta be a 44-form with small enough pointwise norm so that ψ+η\psi+\eta is a positive 44-form. Then

  1. (1)

    the image of ξ\xi under the linearization of Θ\Theta at φ\varphi is

    Θ(ξ)=φ(43π1(ξ)+π7(ξ)π27(ξ)).\displaystyle\Theta(\xi)=*_{\varphi}\Big{(}\frac{4}{3}\pi_{1}(\xi)+\pi_{7}(\xi)-\pi_{27}(\xi)\Big{)}. (2.39)
  2. (2)

    the image of η\eta under the linearization of Θ1\Theta^{-1} at ψ\psi is

    Θ1(η)=φ(34π1(η)+π7(η)π27(η)).\displaystyle\Theta^{-1}(\eta)=*_{\varphi}\Big{(}\frac{3}{4}\pi_{1}(\eta)+\pi_{7}(\eta)-\pi_{27}(\eta)\Big{)}. (2.40)

2.1 First order differential operators

In this section, we discuss various first order differential operators on a manifold with a nearly G2\mathrm{G}_{2} structure and prove some identities involving them.

For fC(M)f\in C^{\infty}(M), we have the vector field gradf\operatorname{grad}f given by

(gradf)k=kf(\operatorname{grad}f)_{k}=\nabla_{k}f

and for any vector field XX we have the divergence of XX which is a function

divX=kXk.\operatorname{div}X=\nabla_{k}X_{k}.

On a manifold with a G2\mathrm{G}_{2} structure φ\varphi, for a vector field XΓ(TM)X\in\Gamma(TM), we define the curl of XX, as

(curlX)k=iXjφijk\displaystyle(\operatorname{curl}X)_{k}=\nabla_{i}X_{j}\varphi_{ijk} (2.41)

which can also be written as

(curlX)=(dXψ)\displaystyle(\operatorname{curl}X)=*(dX\wedge\psi) (2.42)

and so up to G2\mathrm{G}_{2}-equivariant isomorphisms, the vector field curlX\operatorname{curl}X is the projection of the 22-form dXdX onto the Ω72\Omega^{2}_{7} component. In fact, we have the following

Proposition 2.5.

Let XX be a vector field on MM. The Ω72\Omega^{2}_{7} component of dXdX is given by

π7(dX)=13(curlX)φ=13(curlXψ).\displaystyle\pi_{7}(dX)=\frac{1}{3}(\operatorname{curl}X)\lrcorner\varphi=\frac{1}{3}*(\operatorname{curl}X\wedge\psi). (2.43)
Proof.

We know that π7(dX)=Wφ\pi_{7}(dX)=W\lrcorner\varphi for some vector field WW. Using (2.37) we compute

curlX\displaystyle\operatorname{curl}X =(dXψ)=(π7(dX)ψ)=((Wφ)ψ)=3W\displaystyle=*(dX\wedge\psi)=*(\pi_{7}(dX)\wedge\psi)=*((W\lrcorner\varphi)\wedge\psi)=3W

which gives (2.43). ∎

In the next proposition we state and prove various relations among the first order differential operators described above. We prove the results for any G2\mathrm{G}_{2} structure and will later state the results for nearly G2\mathrm{G}_{2} structures. These formulas are generalizations of the formulas first proved for torsion-free G2\mathrm{G}_{2} structures by Karigiannis [Kar06, Proposition 4.4].

Proposition 2.6.

Let fC(M)f\in C^{\infty}(M) and XX be a vector field on MM with a G2\mathrm{G}_{2} structure φ\varphi. Then

curl(gradf)\displaystyle\operatorname{curl}(\operatorname{grad}f) =0,\displaystyle=0, (2.44)
div(curlX)\displaystyle\operatorname{div}(\operatorname{curl}X) =iXj(4(τ1)ij(τ2)ij)+(π7(Rm))jljXl,\displaystyle=\nabla_{i}X_{j}(4(\tau_{1})_{ij}-(\tau_{2})_{ij})+(\pi_{7}(\mathrm{Rm}))_{jl}^{j}X_{l}, (2.45)
curl(curlX)l\displaystyle\operatorname{curl}(\operatorname{curl}X)_{l} =l(divX)+RlmXmΔXl(curlX)mTml(lXiiXl)(τ1)msφmsi\displaystyle=\nabla_{l}(\operatorname{div}X)+R_{lm}X_{m}-\Delta X_{l}-(\operatorname{curl}X)_{m}T_{ml}-(\nabla_{l}X_{i}-\nabla_{i}X_{l})(\tau_{1})_{ms}\varphi_{msi}
+trT(curlX)l+iXjTisφjsl+iXjTjsφsil.\displaystyle\quad+\operatorname{tr}T(\operatorname{curl}X)_{l}+\nabla_{i}X_{j}T_{is}\varphi_{jsl}+\nabla_{i}X_{j}T_{js}\varphi_{sil}. (2.46)
Remark 2.7.

For fixed i,ji,\ j, the Riemann curvature tensor RijklR_{ijkl} is skew-symmetric in kk and ll and hence

Rijkl=(π7(Rm))ijkl+(π14(Rm))ijkl.\displaystyle R_{ijkl}=(\pi_{7}(\mathrm{Rm}))_{ijkl}+(\pi_{14}(\mathrm{Rm}))_{ijkl}.

Explicitly,

(π7(Rm))ijkl=13Rijkl+16Rabklψabij,(π14(Rm))ijkl=23Rijkl16Rabklψabij.\displaystyle(\pi_{7}(\mathrm{Rm}))_{ijkl}=\frac{1}{3}R_{ijkl}+\frac{1}{6}R_{abkl}\psi_{abij},\ \ \ (\pi_{14}(\mathrm{Rm}))_{ijkl}=\frac{2}{3}R_{ijkl}-\frac{1}{6}R_{abkl}\psi_{abij}.

Moreover, from [Kar09, eq. (4.17)], we have

(π7(Rm))ijkl=(π7(Rm))ijmφmklwhereπ7(Rm)ijm=16Rijklφklm.\displaystyle(\pi_{7}(\mathrm{Rm}))_{ijkl}=(\pi_{7}(\mathrm{Rm}))^{m}_{ij}\varphi_{mkl}\ \ \ \ \ \textup{where}\ \ \ \ \ \pi_{7}(\mathrm{Rm})^{m}_{ij}=\frac{1}{6}R_{ijkl}\varphi_{klm}. (2.47)
Proof.

We compute

curl(gradf)\displaystyle\operatorname{curl}(\operatorname{grad}f) =i(jf)φijk=0\displaystyle=\nabla_{i}(\nabla_{j}f)\varphi_{ijk}=0

as φ\varphi is skew-symmetric, thus proving (2.44). For (2.45) we use the Ricci identity (2.1) to get

div(curlX)\displaystyle\operatorname{div}(\operatorname{curl}X) =k(iXjφijk)\displaystyle=\nabla_{k}(\nabla_{i}X_{j}\varphi_{ijk})
=kiXjφijk+iXjkφijk\displaystyle=\nabla_{k}\nabla_{i}X_{j}\varphi_{ijk}+\nabla_{i}X_{j}\nabla_{k}\varphi_{ijk}
=12(kiXjikXj)φijk+iXjTkmψmijk\displaystyle=\frac{1}{2}(\nabla_{k}\nabla_{i}X_{j}-\nabla_{i}\nabla_{k}X_{j})\varphi_{ijk}+\nabla_{i}X_{j}T_{km}\psi_{mijk}
=12RkijlXlφijk+iXj(4(τ1)ij(τ2)ij)\displaystyle=-\frac{1}{2}R_{kijl}X_{l}\varphi_{ijk}+\nabla_{i}X_{j}(4(\tau_{1})_{ij}-(\tau_{2})_{ij})
=3(π7(Rm))ljjXl+iXj(4(τ1)ij(τ2)ij)\displaystyle=3(\pi_{7}(\mathrm{Rm}))_{lj}^{j}X_{l}+\nabla_{i}X_{j}(4(\tau_{1})_{ij}-(\tau_{2})_{ij})

where we used (2.12), (2.13) and (2.47). We have also used the fact that the symmetric part of TT will vanish when contracted with ψ\psi.

Finally we use the contraction identities (2.2) and (2.4) and the Ricci identity (2.1) to compute

(curl(curlX))l\displaystyle(\operatorname{curl}(\operatorname{curl}X))_{l} =m(iXjφijk)φmkl\displaystyle=\nabla_{m}(\nabla_{i}X_{j}\varphi_{ijk})\varphi_{mkl}
=(miXjφijk+iXjTmsψsijk)φlmk\displaystyle=(\nabla_{m}\nabla_{i}X_{j}\varphi_{ijk}+\nabla_{i}X_{j}T_{ms}\psi_{sijk})\varphi_{lmk}
=miXj(gilgjmgimgjl+ψijlm)\displaystyle=\nabla_{m}\nabla_{i}X_{j}(g_{il}g_{jm}-g_{im}g_{jl}+\psi_{ijlm})
+iXjTms(gmsφlij+gmiφslj+gmjφsilglsφmijgliφsmjgljφsim)\displaystyle\quad+\nabla_{i}X_{j}T_{ms}(g_{ms}\varphi_{lij}+g_{mi}\varphi_{slj}+g_{mj}\varphi_{sil}-g_{ls}\varphi_{mij}-g_{li}\varphi_{smj}-g_{lj}\varphi_{sim})
=jlXjΔXl+12(miXjimXj)ψijlm+trTiXjφijl+iXjTisφslj\displaystyle=\nabla_{j}\nabla_{l}X_{j}-\Delta X_{l}+\frac{1}{2}(\nabla_{m}\nabla_{i}X_{j}-\nabla_{i}\nabla_{m}X_{j})\psi_{ijlm}+\operatorname{tr}T\nabla_{i}X_{j}\varphi_{ijl}+\nabla_{i}X_{j}T_{is}\varphi_{slj}
+iXmTmsφsiliXjTmlφmijlXjTmsφsmjiXlTmsφmsi\displaystyle\quad+\nabla_{i}X_{m}T_{ms}\varphi_{sil}-\nabla_{i}X_{j}T_{ml}\varphi_{mij}-\nabla_{l}X_{j}T_{ms}\varphi_{smj}-\nabla_{i}X_{l}T_{ms}\varphi_{msi}
=l(divX)+RlmXmΔXl+trT(curlX)l+iXjTisφjsl+iXmTmsφsil\displaystyle=\nabla_{l}(\operatorname{div}X)+R_{lm}X_{m}-\Delta X_{l}+\operatorname{tr}T(\operatorname{curl}X)_{l}+\nabla_{i}X_{j}T_{is}\varphi_{jsl}+\nabla_{i}X_{m}T_{ms}\varphi_{sil}
(curlX)mTmllXj(τ1)msφmsj+iXl(τ1)msφmsi\displaystyle\quad-(\operatorname{curl}X)_{m}T_{ml}-\nabla_{l}X_{j}(\tau_{1})_{ms}\varphi_{msj}+\nabla_{i}X_{l}(\tau_{1})_{ms}\varphi_{msi}

where we used the fact that Rabcdψabck=0R_{abcd}\psi_{abck}=0 for the third term in the fourth equality and (2.13) to cancel the τ2\tau_{2} components which contract on two indices with φ\varphi for the last two terms in the fourth equality. Thus, we get

(curl(curlX))l\displaystyle(\operatorname{curl}(\operatorname{curl}X))_{l} =l(divX)+RlmXmΔXl(curlX)mTml(iXllXi)(τ1)msφmsi\displaystyle=\nabla_{l}(\operatorname{div}X)+R_{lm}X_{m}-\Delta X_{l}-(\operatorname{curl}X)_{m}T_{ml}-(\nabla_{i}X_{l}-\nabla_{l}X_{i})(\tau_{1})_{ms}\varphi_{msi}
+trT(curlX)l+iXjTisφjsl+iXjTjsφsil.\displaystyle\quad+\operatorname{tr}T(\operatorname{curl}X)_{l}+\nabla_{i}X_{j}T_{is}\varphi_{jsl}+\nabla_{i}X_{j}T_{js}\varphi_{sil}.

For a nearly G2\mathrm{G}_{2} structure we have Tij=τ04gijT_{ij}=\dfrac{\tau_{0}}{4}g_{ij} and Rij=3τ028gijR_{ij}=\dfrac{3{\tau_{0}}^{2}}{8}g_{ij}. Moreover from [Kar09, eq. (4.18)(4.18)],

(π7(Rm))jlj=l(trT)+j(Tlj)+TlaTjbφabj=0.\displaystyle(\pi_{7}(\mathrm{Rm}))_{jl}^{j}=-\nabla_{l}(\operatorname{tr}T)+\nabla_{j}(T_{lj})+T_{la}T_{jb}\varphi_{abj}=0.

Thus using the Weitzenböck formula for XX, Xl=jjXl=(ΔdX)l+RilXi\nabla^{*}\nabla X_{l}=-\nabla_{j}\nabla_{j}X_{l}=(\Delta_{d}X)_{l}+R_{il}X_{i}, we get the following

Corollary 2.8.

Let fC(M)f\in C^{\infty}(M) and XX be a vector field on MM with a nearly G2\mathrm{G}_{2} structure φ\varphi. Then

curl(gradf)\displaystyle\operatorname{curl}(\operatorname{grad}f) =0,\displaystyle=0, (2.48)
div(curlX)\displaystyle\operatorname{div}(\operatorname{curl}X) =0,\displaystyle=0, (2.49)
curl(curlX)\displaystyle\operatorname{curl}(\operatorname{curl}X) =grad(divX)ΔX+3τ028X+τ0(curlX),\displaystyle=\operatorname{grad}(\operatorname{div}X)-\Delta X+\dfrac{3{\tau_{0}}^{2}}{8}X+\tau_{0}(\operatorname{curl}X), (2.50)
=ΔdX+grad(divX)+τ0(curlX).\displaystyle=\Delta_{d}X+\operatorname{grad}(\operatorname{div}X)+\tau_{0}(\operatorname{curl}X). (2.51)

2.2 Identities for 22-forms and 33-forms

In this subsection, we prove some identities for 22-forms and 33-forms on a manifold with a nearly G2\mathrm{G}_{2} structure. These identities will be used several times in the paper.

Lemma 2.9.

Let (M,φ)(M,\varphi) be a manifold with a G2\mathrm{G}_{2} structure. If β=β7+β14\beta=\beta_{7}+\beta_{14} is a 22-form then

  1. (1)

    (βφ)=2β7β14*(\beta\wedge\varphi)=2\beta_{7}-\beta_{14}.

  2. (2)

    (ββφ)=2|β7|2|β14|2*(\beta\wedge\beta\wedge\varphi)=2|\beta_{7}|^{2}-|\beta_{14}|^{2}.

Proof.

The identity in (1)(1) follows from (2.10) and (2.11). For (2)(2) we note that for 77-dimensional manifolds 2(α)=α*^{2}(\alpha)=\alpha for a kk-form α\alpha, so

ββφ=β2(βφ)=β(2β7β14)\displaystyle\beta\wedge\beta\wedge\varphi=\beta\wedge*^{2}(\beta\wedge\varphi)=\beta\wedge*(2\beta_{7}-\beta_{14})

and the decomposition of 22-forms is orthogonal. ∎

Lemma 2.10.

Let (M,φ)(M,\varphi) be a manifold with a G2\mathrm{G}_{2} structure. Let σ=fφ+σ7+σ27\sigma=f\varphi+\sigma_{7}+\sigma_{27} be a 33-form on MM and let σ7=Xψ\sigma_{7}=X\lrcorner\psi for some vector field XX on MM. Then

  1. (1)

    (σφ)=4X*(\sigma\wedge\varphi)=4X.

  2. (2)

    (σψ)=7f*(\sigma\wedge\psi)=7f.

Proof.

For (1)(1) we have

(σφ)\displaystyle*(\sigma\wedge\varphi) =((fφ+σ7+σ27)φ)=(σ7φ)=((Xφ)φ)\displaystyle=*((f\varphi+\sigma_{7}+\sigma_{27})\wedge\varphi)=*(\sigma_{7}\wedge\varphi)=*((X\lrcorner*\varphi)\wedge\varphi)
=4X\displaystyle=4X (2.52)

where we have used the fact that Ω13Ω273\Omega^{3}_{1}\oplus\Omega^{3}_{27} lies in the kernel of wedge product with φ\varphi and (2.35) in the last equality. For (2)(2) we note that Ω73Ω273\Omega^{3}_{7}\oplus\Omega^{3}_{27} lies in the kernel of wedge product with ψ\psi and φψ=7vol\varphi\wedge\psi=7\operatorname{vol}. ∎

Next, we explicitly derive the expressions for exterior derivative and the divergence of various components of 22-forms and 33-forms on a manifold with a nearly G2\mathrm{G}_{2} structure. Some of these identities are new, at least in the present form and we believe that they will be useful in other contexts as well.

Lemma 2.11.

Suppose (M,φ)(M,\varphi) is a manifold with a nearly G2\mathrm{G}_{2} structure. Let fC(M)f\in C^{\infty}(M), βΩ142\beta\in\Omega^{2}_{14} and XΓ(TM)X\in\Gamma(TM). Then

  1. (1)

    d(fφ)=dfφ+τ0fψd(f\varphi)=df\wedge\varphi+\tau_{0}f\psi.

  2. (2)

    d(fφ)=(df)φd^{*}(f\varphi)=-(df)\lrcorner\varphi.

  3. (3)

    dβ=14(dβφ)+π27(dβ)d\beta=\dfrac{1}{4}*(d^{*}\beta\wedge\varphi)+\pi_{27}(d\beta).

  4. (4)

    d(Xφ)=37(dX)φ+12((3τ02XcurlX)φ)+iφ(12(iXj+jXi)+17(dX)gij)d(X\lrcorner\varphi)=-\dfrac{3}{7}(d^{*}X)\varphi+\dfrac{1}{2}*\Big{(}\Big{(}\dfrac{3\tau_{0}}{2}X-\operatorname{curl}X\Big{)}\wedge\varphi\Big{)}+i_{\varphi}\Big{(}\dfrac{1}{2}(\nabla_{i}X_{j}+\nabla_{j}X_{i})+\dfrac{1}{7}(d^{*}X)g_{ij}\Big{)}.

  5. (5)

    d(Xφ)=curlXd^{*}(X\lrcorner\varphi)=\operatorname{curl}X.

  6. (6)

    d(Xψ)=47dXψ(12curlX+τ04X)φiφ(12(iXj+jXj)+17(dX)gij)d(X\lrcorner\psi)=-\dfrac{4}{7}d^{*}X\psi-\Big{(}\dfrac{1}{2}\operatorname{curl}X+\dfrac{\tau_{0}}{4}X\Big{)}\wedge\varphi-*i_{\varphi}\Big{(}\dfrac{1}{2}(\nabla_{i}X_{j}+\nabla_{j}X_{j})+\dfrac{1}{7}(d^{*}X)g_{ij}\Big{)}.

Proof.

We have

d(fφ)\displaystyle d(f\varphi) =dfφ+fdφ\displaystyle=df\wedge\varphi+fd\varphi
=dfφ+τ0fψ\displaystyle=df\wedge\varphi+\tau_{0}f\psi

where we have used (2.24) which proves (1)(1). For part (2)(2) we compute

d(fφ)=d(fφ)=d(fφ)=(dfφ)=dfφ\displaystyle d^{*}(f\varphi)=-*d*(f\varphi)=-*d(f*\varphi)=-*(df\wedge*\varphi)=-df\lrcorner\varphi

as dψ=0d\psi=0.

We prove part (3)(3). Since dβd\beta is a 33-form so

dβ=π1(dβ)+π7(dβ)+π27(dβ).\displaystyle d\beta=\pi_{1}(d\beta)+\pi_{7}(d\beta)+\pi_{27}(d\beta). (2.53)

We compute each term on the right hand side of (2.53). We will repeatedly use the identities (2.29)–(2.38). Suppose

π1(dβ)=aφ\displaystyle\pi_{1}(d\beta)=a\varphi

for some aC(M)a\in C^{\infty}(M). Since Ω73Ω273\Omega^{3}_{7}\oplus\Omega^{3}_{27} lies in the kernel of wedge product with ψ\psi and βψ=0\beta\wedge\psi=0 for βΩ142\beta\in\Omega^{2}_{14}, we have

0=d(βψ)=dβψ=π1(dβ)ψ=7avol\displaystyle 0=d(\beta\wedge\psi)=d\beta\wedge\psi=\pi_{1}(d\beta)\wedge\psi=7a\operatorname{vol}

and hence

π1(dβ)=0.\displaystyle\pi_{1}(d\beta)=0.

Suppose π7(dβ)=Xψ\pi_{7}(d\beta)=X\lrcorner\psi for XΓ(TM)X\in\Gamma(TM). Using (2.11) and Lemma 2.10 (1)(1), we have

dβ=d(β)=d(βφ)=(dβφ)τ0(βψ)=4X.\displaystyle d^{*}\beta=*d*(\beta)=-*d(\beta\wedge\varphi)=-*(d\beta\wedge\varphi)-\tau_{0}*(\beta\wedge\psi)=-4X.

Thus

π7(dβ)=14dβψ=14(dβφ),\displaystyle\pi_{7}(d\beta)=-\frac{1}{4}d^{*}\beta\lrcorner\psi=\frac{1}{4}*(d^{*}\beta\wedge\varphi),

which proves (3).(3).

Since d(Xφ)d(X\lrcorner\varphi) is a 33-form, so we will write

d(Xφ)=π1(d(Xφ))+π7(d(Xφ))+π27(d(Xφ))\displaystyle d(X\lrcorner\varphi)=\pi_{1}(d(X\lrcorner\varphi))+\pi_{7}(d(X\lrcorner\varphi))+\pi_{27}(d(X\lrcorner\varphi)) (2.54)

and will calculate each term on the right hand side. As before, assume

π1(d(Xφ))=aφ\displaystyle\pi_{1}(d(X\lrcorner\varphi))=a\varphi

for some aC(M)a\in C^{\infty}(M). Then

d((Xφ)ψ)=π1(d(Xφ))ψ=7avol\displaystyle d((X\lrcorner\varphi)\wedge\psi)=\pi_{1}(d(X\lrcorner\varphi))\wedge\psi=7a\operatorname{vol}

and hence 7a=d((Xφ)ψ)=d(3X)7a=*d((X\lrcorner\varphi)\wedge\psi)=*d(3*X). So we get that

a=37dX=37dX.\displaystyle a=\frac{3}{7}*d*X=-\frac{3}{7}d^{*}X.

Assume that

π7(d(Xφ))=Yψ\displaystyle\pi_{7}(d(X\lrcorner\varphi))=Y\lrcorner\psi

for some YΓ(TM)Y\in\Gamma(TM). Using the fact that Ω13Ω273\Omega^{3}_{1}\oplus\Omega^{3}_{27} lies in the kernel of wedge product with φ\varphi we get

d((Xφ)φ)=d(Xφ)φ+(Xφ)dφ=π7(d(Xφ))φ+τ0(Xφ)ψ=(Yψ)φ+3τ0X.\displaystyle d((X\lrcorner\varphi)\wedge\varphi)=d(X\lrcorner\varphi)\wedge\varphi+(X\lrcorner\varphi)\wedge d\varphi=\pi_{7}(d(X\lrcorner\varphi))\wedge\varphi+\tau_{0}(X\lrcorner\varphi)\wedge\psi=(Y\lrcorner\psi)\wedge\varphi+3\tau_{0}*X.

So we get

4Y+3τ0X=d((Xφ)φ)=d(2(Xφ))=2d(Xψ)=2(dX)ψ\displaystyle 4*Y+3\tau_{0}*X=d((X\lrcorner\varphi)\wedge\varphi)=d(2*(X\lrcorner\varphi))=2d(X\wedge\psi)=2(dX)\wedge\psi

which gives

Y=12(((dX)ψ)3τ02X)=12(curlX3τ02X)\displaystyle Y=\frac{1}{2}\Big{(}*((dX)\wedge\psi)-\frac{3\tau_{0}}{2}X\Big{)}=\frac{1}{2}\Big{(}\operatorname{curl}X-\frac{3\tau_{0}}{2}X\Big{)}

and hence

π7(d(Xφ))=12((curlX3τ02X)φ).\displaystyle\pi_{7}(d(X\lrcorner\varphi))=-\frac{1}{2}*\Big{(}\Big{(}\operatorname{curl}X-\frac{3\tau_{0}}{2}X\Big{)}\wedge\varphi\Big{)}.

Recall the map iφi_{\varphi} from (2.17). To calculate π27(d(Xφ))\pi_{27}(d(X\lrcorner\varphi)) we have

d(Xφ)imnφjmn+d(Xφ)jmnφimn\displaystyle d(X\lrcorner\varphi)_{imn}\varphi_{jmn}+d(X\lrcorner\varphi)_{jmn}\varphi_{imn} =[37(dX)φimn+12((curlX3τ02X)ψ)imn+i(h0)imn]φjmn\displaystyle=\Big{[}\frac{-3}{7}(d^{*}X)\varphi_{imn}+\frac{1}{2}\Big{(}\Big{(}\operatorname{curl}X-\frac{3\tau_{0}}{2}X\Big{)}\lrcorner\psi\Big{)}_{imn}+i(h_{0})_{imn}\Big{]}\varphi_{jmn}
+[37(dX)φjmn+12((curlX3τ02X)ψ)jmn+i(h0)jmn]φimn\displaystyle\quad+\Big{[}\frac{-3}{7}(d^{*}X)\varphi_{jmn}+\frac{1}{2}\Big{(}\Big{(}\operatorname{curl}X-\frac{3\tau_{0}}{2}X\Big{)}\lrcorner\psi\Big{)}_{jmn}+i(h_{0})_{jmn}\Big{]}\varphi_{imn}
=367(dX)gij+8(h0)ij+12(curlX3τ02X)sψsimnφjmn\displaystyle=-\frac{36}{7}(d^{*}X)g_{ij}+8(h_{0})_{ij}+\frac{1}{2}\Big{(}\operatorname{curl}X-\frac{3\tau_{0}}{2}X\Big{)}_{s}\psi_{simn}\varphi_{jmn}
+(curlX3τ02X)sψsjmnφimn\displaystyle\quad+\Big{(}\operatorname{curl}X-\frac{3\tau_{0}}{2}X\Big{)}_{s}\psi_{sjmn}\varphi_{imn}
=367(dX)gij+8(h0)ij.\displaystyle=-\frac{36}{7}(d^{*}X)g_{ij}+8(h_{0})_{ij}. (2.55)

We calculate the left hand side of (2.55). We have

d(Xφ)imnφjmn+d(Xφ)jmnφimn\displaystyle d(X\lrcorner\varphi)_{imn}\varphi_{jmn}+d(X\lrcorner\varphi)_{jmn}\varphi_{imn} =(i(Xlφlmn)m(Xlφlin)+n(Xlφlim))φjmn\displaystyle=(\nabla_{i}(X_{l}\varphi_{lmn})-\nabla_{m}(X_{l}\varphi_{lin})+\nabla_{n}(X_{l}\varphi_{lim}))\varphi_{jmn}
+(j(Xlφlmn)m(Xlφljn)+n(Xlφljm))φimn\displaystyle\quad+(\nabla_{j}(X_{l}\varphi_{lmn})-\nabla_{m}(X_{l}\varphi_{ljn})+\nabla_{n}(X_{l}\varphi_{ljm}))\varphi_{imn}
=(iXlφlmnmXlφlin+nXlφlim)φjmn\displaystyle=(\nabla_{i}X_{l}\varphi_{lmn}-\nabla_{m}X_{l}\varphi_{lin}+\nabla_{n}X_{l}\varphi_{lim})\varphi_{jmn}
+τ04(XlψilmnXlψmlin+Xlψnlim)φjmn\displaystyle\quad+\dfrac{\tau_{0}}{4}(X_{l}\psi_{ilmn}-X_{l}\psi_{mlin}+X_{l}\psi_{nlim})\varphi_{jmn}
+(jXlφlmnmXlφljn+nXlφljm)φimn\displaystyle\quad+(\nabla_{j}X_{l}\varphi_{lmn}-\nabla_{m}X_{l}\varphi_{ljn}+\nabla_{n}X_{l}\varphi_{ljm})\varphi_{imn}
+τ04(XlψjlmnXlψmljn+Xlψnljm)φimn\displaystyle\quad+\dfrac{\tau_{0}}{4}(X_{l}\psi_{jlmn}-X_{l}\psi_{mljn}+X_{l}\psi_{nljm})\varphi_{imn}

where we have used (2.20) and (2.24). So

d(Xφ)imnφjmn+d(Xφ)jmnφimn\displaystyle d(X\lrcorner\varphi)_{imn}\varphi_{jmn}+d(X\lrcorner\varphi)_{jmn}\varphi_{imn} =(iXlφlmnφjmn2mXlφlinφjmn)\displaystyle=(\nabla_{i}X_{l}\varphi_{lmn}\varphi_{jmn}-2\nabla_{m}X_{l}\varphi_{lin}\varphi_{jmn})
+τ04(XlψilmnXlψmlin+Xlψnlim)φjmn\displaystyle\quad+\dfrac{\tau_{0}}{4}(X_{l}\psi_{ilmn}-X_{l}\psi_{mlin}+X_{l}\psi_{nlim})\varphi_{jmn}
(jXlφlmnφimn2mXlφljnφimn)\displaystyle\quad(\nabla_{j}X_{l}\varphi_{lmn}\varphi_{imn}-2\nabla_{m}X_{l}\varphi_{ljn}\varphi_{imn})
+τ04(XlψjlmnXlψmljn+Xlψnljm)φimn.\displaystyle\quad+\dfrac{\tau_{0}}{4}(X_{l}\psi_{jlmn}-X_{l}\psi_{mljn}+X_{l}\psi_{nljm})\varphi_{imn}.

We use the contraction identities (2.2), (2.3) and (2.4) to get

d(Xφ)imnφjmn+d(Xφ)jmnφimn\displaystyle d(X\lrcorner\varphi)_{imn}\varphi_{jmn}+d(X\lrcorner\varphi)_{jmn}\varphi_{imn} =4iXj+4jXi+4(divX)gij\displaystyle=4\nabla_{i}X_{j}+4\nabla_{j}X_{i}+4(\operatorname{div}X)g_{ij}
+τ04(4Xlφilj+4Xlφlij+4Xlφlij)\displaystyle\quad+\dfrac{\tau_{0}}{4}(-4X_{l}\varphi_{ilj}+4X_{l}\varphi_{lij}+4X_{l}\varphi_{lij})
+τ04(4Xlφjli+4Xlφlji+4Xlφlji)\displaystyle\quad+\dfrac{\tau_{0}}{4}(-4X_{l}\varphi_{jli}+4X_{l}\varphi_{lji}+4X_{l}\varphi_{lji})
=4iXj+4jXi4(dX)gij\displaystyle=4\nabla_{i}X_{j}+4\nabla_{j}X_{i}-4(d^{*}X)g_{ij}

and so from (2.55) we get

367(dX)gij+8(h0)ij\displaystyle-\dfrac{36}{7}(d^{*}X)g_{ij}+8(h_{0})_{ij} =4iXj+4jXi4(dX)gij\displaystyle=4\nabla_{i}X_{j}+4\nabla_{j}X_{i}-4(d^{*}X)g_{ij}

and thus

(h0)ij=12(iXj+jXi)+17(dX)gij\displaystyle(h_{0})_{ij}=\dfrac{1}{2}(\nabla_{i}X_{j}+\nabla_{j}X_{i})+\dfrac{1}{7}(d^{*}X)g_{ij}

which completes the proof of (4)(4).

We obtain (5)(5) by

d(Xφ)=d(Xφ)=d(Xψ)=(dXψ)=curlX.\displaystyle d^{*}(X\lrcorner\varphi)=*d*(X\lrcorner\varphi)=*d(X\wedge\psi)=*(dX\wedge\psi)=\operatorname{curl}X.

To prove part (6)(6), we notice that since dψ=0d\psi=0, d(Xψ)=Xψd(X\lrcorner\psi)=\mathcal{L}_{X}\psi which is the image of Xφ=d(Xφ)+τ0Xψ\mathcal{L}_{X}\varphi=d(X\lrcorner\varphi)+\tau_{0}X\lrcorner\psi under the linearization of the map Θ\Theta. We then use part (4) of the lemma and (2.39) to get part (6).

We use the following important lemma on several occasions.

Lemma 2.12.

Let φ\varphi be a nearly G2\mathrm{G}_{2} structure on MM and σ\sigma be a 33-form so that

σ=fφ+(Xφ)+η\displaystyle\sigma=f\varphi+*(X\wedge\varphi)+\eta

where ηΩ273\eta\in\Omega^{3}_{27} with η=iφ(h)\eta=i_{\varphi}(h) where hh is a symmetric traceless 22-tensor. Then

π1(dσ)\displaystyle\pi_{1}(d\sigma) =(τ0f+47dX)ψ,\displaystyle=\Big{(}\tau_{0}f+\frac{4}{7}d^{*}X\Big{)}\psi, (2.56)
π7(dσ)\displaystyle\pi_{7}(d\sigma) =(df+τ04X+12curlX12divh)φ,\displaystyle=\Big{(}df+\frac{\tau_{0}}{4}X+\frac{1}{2}\operatorname{curl}X-\frac{1}{2}\operatorname{div}h\Big{)}\wedge\varphi, (2.57)
π7(dσ)\displaystyle\pi_{7}(d^{*}\sigma) =((df+τ0X23curlX23divh)ψ).\displaystyle=*\Big{(}(-df+\tau_{0}X-\frac{2}{3}\operatorname{curl}X-\frac{2}{3}\operatorname{div}h)\wedge\psi\Big{)}. (2.58)
Proof.

We note that σ=fψ+(Xφ)+η*\sigma=f\psi+(X\wedge\varphi)+*\eta and since φ\varphi is a nearly G2\mathrm{G}_{2} structure hence

dσ=dfφ+τ0fψ+d(Xφ)+dη\displaystyle d\sigma=df\wedge\varphi+\tau_{0}f\psi+d*(X\wedge\varphi)+d\eta (2.59)

and

dσ=dσ=(dfψ)d(Xφ)+dη.\displaystyle d^{*}\sigma=-*d*\sigma=-*(df\wedge\psi)-*d(X\wedge\varphi)+d^{*}\eta. (2.60)

Now π1(dσ)=λψ\pi_{1}(d\sigma)=\lambda\psi for some λC(M)\lambda\in C^{\infty}(M). We use Lemma 2.11 (6)(6) to get,

7λ\displaystyle 7\lambda =λψ,ψ=π1(dσ),ψ=dσ,ψ\displaystyle=\langle\lambda\psi,\psi\rangle=\langle\pi_{1}(d\sigma),\psi\rangle=\langle d\sigma,\psi\rangle
=dfφ+τ0fψ+d(Xφ)+dη,ψ\displaystyle=\langle df\wedge\varphi+\tau_{0}f\psi+d*(X\wedge\varphi)+d\eta,\psi\rangle
=dfφ,ψ+7τ0f+4dX+dη,ψ.\displaystyle=\langle df\wedge\varphi,\psi\rangle+7\tau_{0}f+4d^{*}X+\langle d\eta,\psi\rangle. (2.61)

The first term on the right hand side of (2.61) is 0 as dfφΩ74df\wedge\varphi\in\Omega^{4}_{7} and ψΩ14\psi\in\Omega^{4}_{1}. The last term is also 0 as from (2.16)

dη,ψvol\displaystyle\langle d\eta,\psi\rangle\operatorname{vol} =dηφ=d(ηφ)+τ0ηψ=0.\displaystyle=d\eta\wedge\varphi=d(\eta\wedge\varphi)+\tau_{0}\eta\wedge\psi=0.

Thus we get that

7λ=7τ0f+4dXλ=τ0f+47dX\displaystyle 7\lambda=7\tau_{0}f+4d^{*}X\ \ \ \ \ \ \ \ \implies\ \ \ \ \ \ \ \lambda=\tau_{0}f+\frac{4}{7}d^{*}X

which gives (2.56).

To derive (2.57) and (2.58), we will need to contract ηΩ273\eta\in\Omega^{3}_{27} with φ\varphi on two indices and with ψ\psi on three indices. Using (2.17) and the contraction identities (2.2) and (2.5), a short computation gives

ηijkφajk\displaystyle\eta_{ijk}\varphi_{ajk} =4hia,\displaystyle=4h_{ia}, (2.62)
ηijkψaijk\displaystyle\eta_{ijk}\psi_{aijk} =0.\displaystyle=0. (2.63)

Suppose π7(dσ)=Yφ\pi_{7}(d\sigma)=Y\wedge\varphi for some 11-form YY. Note that for an arbitrary 11-form ZZ we have

Yφ,Zφvol\displaystyle\langle Y\wedge\varphi,Z\wedge\varphi\rangle\operatorname{vol} =Yφ(Zφ)\displaystyle=Y\wedge\varphi\wedge*(Z\wedge\varphi)
=Yφ(Zψ)=4YZ\displaystyle=-Y\wedge\varphi\wedge(Z\lrcorner\psi)=4Y\wedge*Z
=4Y,Zvol.\displaystyle=4\langle Y,Z\rangle\operatorname{vol}.

So from (2.59) we have

4Y,Z\displaystyle 4\langle Y,Z\rangle =Yφ,Zφ=π7(dσ),Zφ=dσ,Zφ\displaystyle=\langle Y\wedge\varphi,Z\wedge\varphi\rangle=\langle\pi_{7}(d\sigma),Z\wedge\varphi\rangle=\langle d\sigma,Z\wedge\varphi\rangle
=dfφ+τ0fψ+d(Xφ)+dη,Zφ\displaystyle=\langle df\wedge\varphi+\tau_{0}f\psi+d*(X\wedge\varphi)+d\eta,Z\wedge\varphi\rangle
=4df,Z+d(Xφ),Zφ+dη,Zφ.\displaystyle=4\langle df,Z\rangle+\langle d*(X\wedge\varphi),Z\wedge\varphi\rangle+\langle d\eta,Z\wedge\varphi\rangle. (2.64)

We first use Lemma 2.11 (6)(6) to calculate the second term on the right hand side of (2.64). We have

d(Xφ),Zφ\displaystyle\langle d*(X\wedge\varphi),Z\wedge\varphi\rangle =(12curlX+τ04X)φ,Zφ=2curlX+τ0X,Z\displaystyle=\left\langle(\frac{1}{2}\operatorname{curl}X+\frac{\tau_{0}}{4}X)\wedge\varphi,Z\wedge\varphi\right\rangle=\langle 2\operatorname{curl}X+\tau_{0}X,Z\rangle

So in (2.64), we have

4Y,Z\displaystyle 4\langle Y,Z\rangle =4df+τ0X+2curlX,Z+dη,Zφ.\displaystyle=\langle 4df+\tau_{0}X+2\operatorname{curl}X,Z\rangle+\langle d\eta,Z\wedge\varphi\rangle. (2.65)

We compute in local coordinates

dη,Zφ\displaystyle\langle d\eta,Z\wedge\varphi\rangle =124(dη)ijkl(Zφ)ijkl\displaystyle=\frac{1}{24}(d\eta)_{ijkl}(Z\wedge\varphi)_{ijkl}
=124(iηjkljηikl+kηijllηijk)(Zφ)ijkl\displaystyle=\frac{1}{24}(\nabla_{i}\eta_{jkl}-\nabla_{j}\eta_{ikl}+\nabla_{k}\eta_{ijl}-\nabla_{l}\eta_{ijk})(Z\wedge\varphi)_{ijkl}
=16(iηjkl)(ZiφjklZjφiklZkφjilZlφjki)\displaystyle=\frac{1}{6}(\nabla_{i}\eta_{jkl})(Z_{i}\varphi_{jkl}-Z_{j}\varphi_{ikl}-Z_{k}\varphi_{jil}-Z_{l}\varphi_{jki})
=16(Ziiηjklφjkl3Zjiηjklφikl)\displaystyle=\frac{1}{6}(Z_{i}\nabla_{i}\eta_{jkl}\varphi_{jkl}-3Z_{j}\nabla_{i}\eta_{jkl}\varphi_{ikl})
=16(Zii(ηjklφjkl)τ04Ziηjklψijkl3Zji(ηjklφikl)+3τ04Zjηjklψiikl).\displaystyle=\frac{1}{6}(Z_{i}\nabla_{i}(\eta_{jkl}\varphi_{jkl})-\frac{\tau_{0}}{4}Z_{i}\eta_{jkl}\psi_{ijkl}-3Z_{j}\nabla_{i}(\eta_{jkl}\varphi_{ikl})+\frac{3\tau_{0}}{4}Z_{j}\eta_{jkl}\psi_{iikl}).

We now use (2.62), (2.63) and the fact that hh is traceless to get

dη,Zφ\displaystyle\langle d\eta,Z\wedge\varphi\rangle =16(Zii(4trh)03Zji(4hji))\displaystyle=\frac{1}{6}(Z_{i}\nabla_{i}(4\operatorname{tr}h)-0-3Z_{j}\nabla_{i}(4h_{ji}))
=2divh,Z.\displaystyle=-2\langle\operatorname{div}h,Z\rangle.

Thus from (2.65) we get

Y,Z\displaystyle\langle Y,Z\rangle =df+τ04X+12curlX12divh,Z\displaystyle=\Big{\langle}df+\frac{\tau_{0}}{4}X+\frac{1}{2}\operatorname{curl}X-\frac{1}{2}\operatorname{div}h,Z\Big{\rangle}

and since ZZ is arbitrary, we get

Y=df+τ04X+12curlX12divh\displaystyle Y=df+\frac{\tau_{0}}{4}X+\frac{1}{2}\operatorname{curl}X-\frac{1}{2}\operatorname{div}h

which establishes (2.57).

Next, we see from (2.60) and (2.10) that

dσ\displaystyle d^{*}\sigma =(dfψ)(dXφ)+τ0(Xψ)+dη\displaystyle=-*(df\wedge\psi)-*(dX\wedge\varphi)+*\tau_{0}(X\wedge\psi)+d^{*}\eta
=(dfτ0Xψ)2π7(dX)+π14(dX)+dη\displaystyle=-*(df-\tau_{0}X\wedge\psi)-2\pi_{7}(dX)+\pi_{14}(dX)+d^{*}\eta

which on using (2.43) becomes

dσ\displaystyle d^{*}\sigma =((dfτ0X+23curlX)ψ)+π14(dX)+dη.\displaystyle=-*\Big{(}\Big{(}df-\tau_{0}X+\frac{2}{3}\operatorname{curl}X\Big{)}\wedge\psi\Big{)}+\pi_{14}(dX)+d^{*}\eta. (2.66)

Suppose π7(dσ)=(Wψ)\pi_{7}(d^{*}\sigma)=*(W\wedge\psi) for some 11-form WW. For any 11-form ZZ we note that

(Wψ),(Zψ)vol\displaystyle\langle*(W\wedge\psi),*(Z\wedge\psi)\rangle\operatorname{vol} =(Wψ)Zψ=(Wψ)ψZ=3WZ=3W,Zvol.\displaystyle=*(W\wedge\psi)\wedge Z\wedge\psi=*(W\wedge\psi)\wedge\psi\wedge Z=3*W\wedge Z=3\langle W,Z\rangle\operatorname{vol}.

Thus using (2.66) and the orthogonality of the spaces Ω72\Omega^{2}_{7} and Ω142\Omega^{2}_{14}, we have

3W,Z\displaystyle 3\langle W,Z\rangle =(Wψ),(Zψ)=π7(dσ),(Zψ)=dσ,(Zψ)\displaystyle=\langle*(W\wedge\psi),*(Z\wedge\psi)\rangle=\langle\pi_{7}(d^{*}\sigma),*(Z\wedge\psi)\rangle=\langle d^{*}\sigma,*(Z\wedge\psi)\rangle
=((dfτ0X+23curlX)ψ)+π14(dX)+dη,(Zψ)\displaystyle=\langle-*((df-\tau_{0}X+\frac{2}{3}\operatorname{curl}X)\wedge\psi)+\pi_{14}(dX)+d^{*}\eta,*(Z\wedge\psi)\rangle
=3df+3τ0X2curlX,Z+dη,(Zψ).\displaystyle=\langle-3df+3\tau_{0}X-2\operatorname{curl}X,Z\rangle+\langle d^{*}\eta,*(Z\wedge\psi)\rangle. (2.67)

Using (2.62) and (2.63), we compute the last term on the right hand side of (2.67), in local coordinates. We have

dη,(Zψ)\displaystyle\langle d^{*}\eta,*(Z\wedge\psi)\rangle =dη,Zφ=12(dη)ijZmφmij=12p(ηpij)Zmφmij\displaystyle=\langle d^{*}\eta,Z\lrcorner\varphi\rangle=\frac{1}{2}(d^{*}\eta)_{ij}Z_{m}\varphi_{mij}=-\frac{1}{2}\nabla_{p}(\eta_{pij})Z_{m}\varphi_{mij}
=12Zm(p(ηpijφmij)τ04ηpijψpmij)\displaystyle=-\frac{1}{2}Z_{m}(\nabla_{p}(\eta_{pij}\varphi_{mij})-\frac{\tau_{0}}{4}\eta_{pij}\psi_{pmij})
=12Zm(4phpm0)=2divh,Z\displaystyle=-\frac{1}{2}Z_{m}(4\nabla_{p}h_{pm}-0)=-2\langle\operatorname{div}h,Z\rangle

and hence we get

W,Z\displaystyle\langle W,Z\rangle =df+τ0X23curlX23divh,Z.\displaystyle=\Big{\langle}-df+\tau_{0}X-\frac{2}{3}\operatorname{curl}X-\frac{2}{3}\operatorname{div}h,Z\Big{\rangle}.

Since ZZ is arbitrary we get

W=df+τ0X23curlX23divh\displaystyle W=-df+\tau_{0}X-\frac{2}{3}\operatorname{curl}X-\frac{2}{3}\operatorname{div}h

which gives (2.58). ∎

Remark 2.13.

The main point of the previous lemma is to exhibit a relation between π7(dη)\pi_{7}(d\eta) and π7(dη)\pi_{7}(d^{*}\eta). Such a relation is expected because of the form of the linearization of the map Θ\Theta. More precisely, from (2.39), applying the linearization of Θ\Theta to Lie derivatives, we have π27(Xψ)=π27(Xφ)\pi_{27}(\mathcal{L}_{X}\psi)=-*\pi_{27}(\mathcal{L}_{X}\varphi), dη,ZφL2=η,XψL2\langle d\eta,Z\wedge\varphi\rangle_{L^{2}}=-\langle\eta,*\mathcal{L}_{X}\psi\rangle_{L^{2}} and dη,ZφL2=η,XφL2\langle d^{*}\eta,Z\lrcorner\varphi\rangle_{L^{2}}=\langle\eta,\mathcal{L}_{X}\varphi\rangle_{L^{2}}. The computations in local coordinates was done to relate π7(dη)\pi_{7}(d\eta) and π7(dη)\pi_{7}(d^{*}\eta) to the divergence of the symmetric 22-tensor hh.

Remark 2.14.

The previous lemma generalizes Proposition 2.17 from [KL20] where the G2\mathrm{G}_{2} structure was assumed to be torsion-free (τ0=0)(\tau_{0}=0).

We have the following corollary of Lemma 2.12.

Corollary 2.15.

Let φ\varphi be a nearly G2\mathrm{G}_{2} structure and let ηΩ273\eta\in\Omega^{3}_{27}. Then

  1. (1)

    If η\eta is closed then dηΩ142d^{*}\eta\in\Omega^{2}_{14}.

  2. (2)

    If η\eta is co-closed then dηΩ274d\eta\in\Omega^{4}_{27}.

Proof.

In the notation of Lemma 2.12 we get that f=X=0f=X=0 and σ=η\sigma=\eta. Thus we get that

π7(dη)=0π7(dη)=0\pi_{7}(d\eta)=0\ \ \ \ \ \ \iff\ \ \ \ \ \ \pi_{7}(d^{*}\eta)=0

as from Lemma 2.12, both conditions are equivalent to divh=0\operatorname{div}h=0. Now if dη=0d\eta=0 then π7(dη)=0\pi_{7}(d^{*}\eta)=0 and hence dηΩ142d^{*}\eta\in\Omega^{2}_{14}. If dη=0d^{*}\eta=0 then π7(dη)=0\pi_{7}(d\eta)=0. Also, since f=X=0f=X=0, we know from (2.56) that π1(dη)=0\pi_{1}(d\eta)=0. So dηΩ274d\eta\in\Omega^{4}_{27}. ∎

We also have a result similar to Lemma 2.12 for 44-forms which we state below. The proof follows from the proof of Lemma 2.12 by taking ζ=σ\zeta=*\sigma and noting that iφ(h)=iφ(h)*i_{\varphi}(h)=-i_{\varphi}(h). We expect that both Lemma 2.12 and Lemma 2.16 will be useful in other contexts as well.

Lemma 2.16.

Let φ\varphi be a nearly G2\mathrm{G}_{2} structure on MM and ζ\zeta be a 44-form on MM so that

ζ=fψ+Xφ+ζ0\displaystyle\zeta=f\psi+X\wedge\varphi+\zeta_{0}

where XΩ1(M)X\in\Omega^{1}(M) and ζ0Ω274\zeta_{0}\in\Omega^{4}_{27} with ζ0=iφ(h)\zeta_{0}=*i_{\varphi}(h) where hh is a symmetric traceless 22-tensor. Then

π7(dζ)\displaystyle\pi_{7}(d\zeta) =WψwhereW=dfτ0X+23curlX23divh,\displaystyle=W\wedge\psi\ \ \ \textup{where}\ \ \ W=df-\tau_{0}X+\frac{2}{3}\operatorname{curl}X-\frac{2}{3}\operatorname{div}h, (2.68)
π1(dζ)\displaystyle\pi_{1}(d^{*}\zeta) =(τ0f+47dX)φ,\displaystyle=\Big{(}\tau_{0}f+\frac{4}{7}d^{*}X\Big{)}\varphi, (2.69)
π7(dζ)\displaystyle\pi_{7}(d^{*}\zeta) =YψwhereY=df12curlXτ04X12divh.\displaystyle=Y\lrcorner\psi\ \ \ \textup{where}\ \ \ Y=-df-\frac{1}{2}\operatorname{curl}X-\frac{\tau_{0}}{4}X-\frac{1}{2}\operatorname{div}h. (2.70)

We get the following corollary.

Corollary 2.17.

Let φ\varphi be a nearly G2\mathrm{G}_{2} structure on MM and let ζ0Ω274\zeta_{0}\in\Omega^{4}_{27}. Then

  1. 1.

    If dζ0=0d\zeta_{0}=0 then dζ0Ω273d^{*}\zeta_{0}\in\Omega^{3}_{27}.

  2. 2.

    If dζ0=0d^{*}\zeta_{0}=0 then dζ0Ω145d\zeta_{0}\in\Omega^{5}_{14}.

3.   Hodge theory of nearly G2\mathrm{G}_{2} manifolds

3.1 Dirac operators on nearly G2\mathrm{G}_{2} manifolds

We begin this section by defining the Dirac operator on (M,φ)(M,\varphi) with a nearly G2\mathrm{G}_{2} structure. We then define a modified Dirac operator which is more suitable for our purposes. A G2\mathrm{G}_{2} structure on MM induces a spin structure, so MM admits an associated Dirac operator \not{D} on its spinor bundle (M)\not{S}(M). Since τ0\tau_{0} is constant, by rescaling the metric induced by the nearly G2\mathrm{G}_{2} structure, we can change the magnitude of τ0\tau_{0} and by changing the orientation, we can change its sign. In the later part of the paper, we study deformations of nearly G2\mathrm{G}_{2} structures through nearly G2\mathrm{G}_{2} structures φt{\varphi}_{t}. Since the underlying metric of any nearly G2\mathrm{G}_{2} structure is positive Einstein, the family of metrics gtg_{t} corresponding to φt{\varphi}_{t} will be positive Einstein and so by [Bes87, Corollary 2.12], the scalar curvature RtR_{t} is constant in tt. Thus, by (2.28), τ0\tau_{0} will be constant through the deformation. Henceforth, we will assume that τ0=4\tau_{0}=4. The results of the paper do not depend on the value of τ0\tau_{0} chosen. Recall the following definition from §1 with τ0=4\tau_{0}=4.

Definition 3.1.

A spinor ηΓ((M))\eta\in\Gamma(\not{S}(M)) is called a Killing spinor if for any XΓ(TM)X\in\Gamma(TM)

Xη=12Xη\nabla_{X}\eta=-\frac{1}{2}X\cdot\eta (3.1)

where ``"``\cdot" is the Clifford multiplication.

The real spinor bundle (M)\not{S}(M), as a G2\mathrm{G}_{2} representation, is isomorphic to Ω0Ω1\Omega^{0}\oplus\Omega^{1}, where the isomorphism is

(f,X)fη+Xη.\displaystyle(f,X)\longrightarrow f\cdot\eta+X\cdot\eta.

For comparison with the Dirac-type operator which we define later, let us derive a formula for the Dirac operator \not{D} on a nearly G2\mathrm{G}_{2} manifold in terms of this isomorphism.

A unit spinor η\eta on a nearly G2\mathrm{G}_{2} manifold MM satisfies (3.1). Thus

(fη)=i=17eiei(fη)=fη+72fη,\displaystyle\not{D}(f\eta)=\sum_{i=1}^{7}e_{i}\cdot\nabla_{e_{i}}(f\eta)=\nabla f\cdot\eta+\frac{7}{2}f\eta,

where we have used the fact that eiei=1e_{i}\cdot e_{i}=-1. Also,

(Xη)\displaystyle\not{D}(X\cdot\eta) =i=17eiei(Xη)=i=17(eieiXη+eiXeiη)\displaystyle=\sum_{i=1}^{7}e_{i}\cdot\nabla_{e_{i}}(X\cdot\eta)=\sum_{i=1}^{7}(e_{i}\cdot\nabla_{e_{i}}X\cdot\eta+e_{i}\cdot X\cdot\nabla_{e_{i}}\eta)
=(dX)η+(dX)η+i=17eiXeiη\displaystyle=(dX)\cdot\eta+(d^{*}X)\eta+\sum_{i=1}^{7}e_{i}\cdot X\cdot\nabla_{e_{i}}\eta

which on using Xei+eiX=2X,eiX\cdot e_{i}+e_{i}\cdot X=-2\langle X,e_{i}\rangle and (3.1) becomes

(Xη)\displaystyle\not{D}(X\cdot\eta) =(dX)η+(dX)ηi=17(Xeieiη+2X,eieiη)\displaystyle=(dX)\cdot\eta+(d^{*}X)\eta-\sum_{i=1}^{7}(X\cdot e_{i}\cdot\nabla_{e_{i}}\eta+2\langle X,e_{i}\rangle\nabla_{e_{i}}\eta)
=(dX)η+(dX)η72Xη+Xη=(dX)η+(dX)η52Xη.\displaystyle=(dX)\cdot\eta+(d^{*}X)\eta-\frac{7}{2}X\cdot\eta+X\cdot\eta=(dX)\cdot\eta+(d^{*}X)\eta-\frac{5}{2}X\cdot\eta.

Thus we get

(fη+Xη)=(72f+dX)η+(f+dX52X)η.\displaystyle\not{D}(f\eta+X\cdot\eta)=\Big{(}\frac{7}{2}f+d^{*}X\Big{)}\eta+\Big{(}\nabla f+dX-\frac{5}{2}X\Big{)}\cdot\eta. (3.2)

Now dXdX is a 22-form, hence dX=π7(dX)+π14(dX)dX=\pi_{7}(dX)+\pi_{14}(dX). Since the Lie group G2\mathrm{G}_{2} preserves the nearly G2\mathrm{G}_{2} structure φ\varphi, it preserves the real Killing spinor η\eta induced by φ\varphi and Ω142(M)𝔤2\Omega^{2}_{14}(M)\cong\mathfrak{g}_{2}, the Lie algebra of G2\mathrm{G}_{2}, we have π14(dX)η=0\pi_{14}(dX)\cdot\eta=0. Also, we know from (2.42) that π7(dX)=13(curlX)φ\pi_{7}(dX)=\dfrac{1}{3}(\operatorname{curl}X)\lrcorner\varphi and it follows from the definition of the Clifford multiplication, for instance as in [Kar06, §4.2], that (Yφ)η=3Yη(Y\lrcorner\varphi)\cdot\eta=3Y\cdot\eta for any YΓ(TM)Y\in\Gamma(TM), we get that

(f,X)=(72f+dX)η+(f+curlX52X)η\displaystyle\not{D}(f,X)=\Big{(}\frac{7}{2}f+d^{*}X\Big{)}\eta+\Big{(}\nabla f+\operatorname{curl}X-\frac{5}{2}X\Big{)}\cdot\eta

which we will write as

(f,X)=(72f+dX,f+curlX52X).\displaystyle\not{D}(f,X)=\Big{(}\frac{7}{2}f+d^{*}X,\nabla f+\operatorname{curl}X-\frac{5}{2}X\Big{)}. (3.3)
Definition 3.2.

The Dirac operator \not{D} is a first-order differential operator on (M)\not{S}(M) defined as follows. Let s=(f,X)Γ((M))s=(f,X)\in\Gamma(\not{S}(M)). Then

(f,X)=(72f+dX,f+curlX52X).\not{D}(f,X)=\Big{(}\frac{7}{2}f+d^{*}X,\nabla f+\operatorname{curl}X-\frac{5}{2}X\Big{)}. (3.4)

The Dirac operator is formally self-adjoint, that is, =\not{D}^{*}=\not{D} and is also an elliptic operator.

Consider the Dirac Laplacian 2=\not{D}^{2}=\not{D}^{*}\not{D}. We relate it to the Hodge Laplacian in the following

Proposition 3.3.

Let s=(f,X)s=(f,X) be a section of the spinor bundle (M)\not{S}(M). Then

2(f,X)\displaystyle\not{D}^{2}(f,X) =(Δf+494f+dX,ΔdX+curlX+254X+f).\displaystyle=\Big{(}\Delta f+\frac{49}{4}f+d^{*}X,\ \Delta_{d}X+\operatorname{curl}X+\frac{25}{4}X+\nabla f\Big{)}. (3.5)

Thus 2\not{D}^{2} is equal to Δd\Delta_{d} up to lower order terms.

Proof.

Using Corollary 2.8, we calculate

2(f,X)\displaystyle\not{D}^{2}(f,X) =(72(72f+dX)+d(f+curlX52X),d(72f+dX)+curl(f+curlX52X)\displaystyle=\Big{(}\frac{7}{2}\Big{(}\frac{7}{2}f+d^{*}X\Big{)}+d^{*}\Big{(}\nabla f+\operatorname{curl}X-\frac{5}{2}X\Big{)},d\Big{(}\frac{7}{2}f+d^{*}X\Big{)}+\operatorname{curl}\Big{(}\nabla f+\operatorname{curl}X-\frac{5}{2}X\Big{)}
52(f+curlX52X))\displaystyle\qquad\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \qquad\qquad\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ -\frac{5}{2}\Big{(}\nabla f+\operatorname{curl}X-\frac{5}{2}X\Big{)}\Big{)}
=(Δf+494f+dX,ΔdX+curlX+254X+f)\displaystyle=\Big{(}\Delta f+\frac{49}{4}f+d^{*}X,\ \Delta_{d}X+\operatorname{curl}X+\frac{25}{4}X+\nabla f\Big{)}

which proves (3.5). ∎

We need a modification of the Dirac operator defined above. The spinor bundle (M)\not{S}(M) is isomorphic to Ω10Ω71\Omega^{0}_{1}\oplus\Omega^{1}_{7} and hence, via a G2\mathrm{G}_{2}-equivariant isomorphism, it is also isomorphic to Ω13Ω73\Omega^{3}_{1}\oplus\Omega^{3}_{7}. We define the modified Dirac operator, which we denote by DD, as follows. Consider the map

D:\displaystyle D: Ω10Ω71Ω13Ω73\displaystyle\ \Omega^{0}_{1}\oplus\Omega^{1}_{7}\longrightarrow\Omega^{3}_{1}\oplus\Omega^{3}_{7}
(f,X)12d(fφ)+π17(d(Xφ)).\displaystyle(f,X)\mapsto\frac{1}{2}*d(f\varphi)+\pi_{1\oplus 7}(d(X\lrcorner\varphi)).

Using Lemma 2.11 (4)(4) with τ0=4\tau_{0}=4, we get

D(f,X)=(2f37dX,12df+6XcurlX).\displaystyle D(f,X)=\Big{(}2f-\frac{3}{7}d^{*}X,\frac{1}{2}df+6X-\operatorname{curl}X\Big{)}. (3.6)
Remark 3.4.

We note that DD is defined in the same way as in [KL20] where the authors denote the operator by ˇ\check{\not{D}}.

We find the kernel of DD. Let (f,X)Ω0Ω1(f,X)\in\Omega^{0}\oplus\Omega^{1} be in the kernel of DD. Then

2f37dX\displaystyle 2f-\frac{3}{7}d^{*}X =0,\displaystyle=0,
12df+6XcurlX\displaystyle\frac{1}{2}df+6X-\operatorname{curl}X =0.\displaystyle=0.

Taking dd^{*} of the second equation and using the first equation and equation (2.49), we get

Δf=ddf\displaystyle\Delta f=d^{*}df =2dcurlX12dX=56f.\displaystyle=2d^{*}\operatorname{curl}X-12d^{*}X=-56f.

Since Δ\Delta is a non-negative operator, f=0f=0. For XX, we have

dX=0andcurlX=6X.\displaystyle d^{*}X=0\ \ \ \ \ \ \ \textup{and}\ \ \ \ \ \ \ \operatorname{curl}X=6X.

We want to prove that XX is a Killing vector field. Let dX=Yφ+π14(dX)dX=Y\lrcorner\varphi+\pi_{14}(dX). Then

dXψ\displaystyle dX\wedge\psi =(Yφ)ψ\displaystyle=(Y\lrcorner\varphi)\wedge\psi
=3Y.\displaystyle=3*Y.

Therefore π7(dX)=13(dXψ)φ=13(curlX)φ=2Xφ.\pi_{7}(dX)=\dfrac{1}{3}*(dX\wedge\psi)\lrcorner\varphi=\dfrac{1}{3}(\operatorname{curl}X)\lrcorner\varphi=2X\lrcorner\varphi. From Lemma 2.9 (2)(2), we have

M𝑑XdXφ\displaystyle\int_{M}dX\wedge dX\wedge\varphi =22Xφ2π14(dX)2=8Xφ,Xφπ14(dX)2\displaystyle=2\|2X\lrcorner\varphi\|^{2}-\|\pi_{14}(dX)\|^{2}=8\langle X\lrcorner\varphi,X\lrcorner\varphi\rangle-\|\pi_{14}(dX)\|^{2}
=8X,((Xφ)ψ)π14(dX)2=24X2π14(dX)2.\displaystyle=8\langle X,*((X\lrcorner\varphi)\wedge\psi)\rangle-\|\pi_{14}(dX)\|^{2}=24\|X\|^{2}-\|\pi_{14}(dX)\|^{2}.

On the other hand, since MM is compact, using integration by parts we have

M𝑑XdXφ\displaystyle\int_{M}dX\wedge dX\wedge\varphi =MXdXdφ=4MXdXψ=4MX(6X)=24X2.\displaystyle=\int_{M}X\wedge dX\wedge d\varphi=4\int_{M}X\wedge dX\wedge\psi=4\int_{M}X\wedge(6*X)=24\|X\|^{2}.

Therefore, π14(dX)=0\pi_{14}(dX)=0 and dX=π7(dX)=2XφdX=\pi_{7}(dX)=2X\lrcorner\varphi. Now using Lemma 2.11 (4)(4), along with the fact that XkerDX\in\ker D, i.e., dX=0d^{*}X=0 and curlX=6X\operatorname{curl}X=6X, we get

0\displaystyle 0 =d(dX)=d(Xφ)=12iφ(Xg),\displaystyle=d(dX)=d(X\lrcorner\varphi)=\frac{1}{2}i_{\varphi}(\mathcal{L}_{X}g),

and hence XX is a Killing vector field. Therefore kerD\ker D is isomorphic to the set of Killing vector fields XX such that curlX=6X\operatorname{curl}X=6X. We denote kerD\ker D by 𝒦\mathcal{K}, that is,

kerD=𝒦={XΓ(TM)Xg=0andcurlX=6X}.\displaystyle\ker D=\mathcal{K}=\{X\in\Gamma(TM)\mid\mathcal{L}_{X}g=0\ \textup{and}\ \operatorname{curl}X=6X\}. (3.7)
Remark 3.5.

Note that the above can also be proved using the identity ΔX=ddX=2d(Xφ)=12X\Delta X=d^{*}dX=-2d^{*}(X\lrcorner\varphi)=12X, since Ricg=6g\mathrm{Ric}_{g}=6g for τ0=4\tau_{0}=4.

Remark 3.6.

If we also want the vector field X𝒦X\in\mathcal{K} to preserve the G2\mathrm{G}_{2} structure, then

Xφ=d(Xφ)+Xdφ=4Xψ=0,\displaystyle\mathcal{L}_{X}\varphi=d(X\lrcorner\varphi)+X\lrcorner\ d\varphi=4X\lrcorner\psi=0,

but since Ω1Ω74\Omega^{1}\cong\Omega^{4}_{7}, this implies X=0X=0. Hence the only vector fields in 𝒦\mathcal{K} that preserve the G2\mathrm{G}_{2} structure are trivial. Note that when φ\varphi is a nearly G2\mathrm{G}_{2} structure of type-1, that is dim(K)=1\mathrm{dim}(K\not{S})=1, every Killing vector field preserves the G2\mathrm{G}_{2} structure and hence 𝒦={0}\mathcal{K}=\{0\}.

The motivation for defining the modified Dirac operator can be understood from the following.

Consider the following operator

D+:Ω13Ω75\displaystyle D^{+}\ :\ \Omega^{3}_{1}\oplus\Omega^{5}_{7} Ω174\displaystyle\to\Omega^{4}_{1\oplus 7}
(fφ,Xψ)\displaystyle(f\varphi,X\wedge\psi) π17(d(fφ)+d(Xψ)).\displaystyle\mapsto\pi_{1\oplus 7}(d(f\varphi)+d^{*}(X\wedge\psi)).

From previous calculations and Lemma 2.11 we know that

d(fφ)\displaystyle d(f\varphi) =dfφ+4fψΩ174,\displaystyle=df\wedge\varphi+4f\psi\in\Omega^{4}_{1\oplus 7},
π17(d(Xψ))\displaystyle\pi_{1\oplus 7}(d^{*}(X\wedge\psi)) =37(dX)ψ+12(curlX6X)φ.\displaystyle=\frac{3}{7}(d^{*}X)\psi+\frac{1}{2}\Big{(}\operatorname{curl}X-6X\Big{)}\wedge\varphi.

Thus

D+(fφ,Xψ)\displaystyle D^{+}(f\varphi,X\wedge\psi) =(4f+37(dX),df+12(curlX6X)).\displaystyle=\Big{(}4f+\frac{3}{7}(d^{*}X),df+\frac{1}{2}\Big{(}\operatorname{curl}X-6X\Big{)}\Big{)}.

Doing a similar calculation as we did for kerD\ker D, we observe that if (f,X)kerD+(f,X)\in\ker D^{+}, then

Δf=28f,curlX=6Xf=0=dXhenceX𝒦\displaystyle\Delta f=-28f,\ \ \operatorname{curl}X=6X\ \ \ \implies\ \ f=0=d^{*}X\ \ \textup{hence}\ \ X\in\mathcal{K}

and so kerD+=kerD\ker D^{+}=\ker D. Since Ω13Ω75Ω174\Omega^{3}_{1}\oplus\Omega^{5}_{7}\cong\Omega^{4}_{1\oplus 7} and D,D+D,D^{+} are self-adjoint operators, we have the following identification

Ω174\displaystyle\Omega^{4}_{1\oplus 7} =ImD+kerD+=ImD+kerD\displaystyle=\operatorname{Im}D^{+}\oplus\ker D^{+}=\operatorname{Im}D^{+}\oplus\ker D (3.8)
=dΩ13π17(dΩ75){Xφ|X𝒦}.\displaystyle=d\Omega^{3}_{1}\oplus\pi_{1\oplus 7}(d^{*}\Omega^{5}_{7})\oplus\{X\wedge\varphi|X\in\mathcal{K}\}.

This is used in the following important

Proposition 3.7.

Let (M,φ,ψ)(M,\varphi,\psi) be a nearly G2\mathrm{G}_{2} manifold. Then the following holds.

  1. 1.

    Ω4={Xφ|X𝒦}dΩ13dΩ75Ω274\Omega^{4}=\{X\wedge\varphi|X\in\mathcal{K}\}\oplus d\Omega^{3}_{1}\oplus d^{*}\Omega^{5}_{7}\oplus\Omega^{4}_{27}.

  2. 2.

    We have an L2L^{2}-orthogonal decomposition Ωexact4={XφX𝒦}dΩ13Ω27,exact4\Omega^{4}_{\textup{exact}}=\{X\wedge\varphi\mid X\in\mathcal{K}\}\oplus d\Omega^{3}_{1}\oplus\Omega^{4}_{27,\textup{exact}}.

Proof.

The first part of the proposition follows from the decomposition of Ω174\Omega^{4}_{1\oplus 7} in equation (3.8).

For the second part we note that the space dΩ75d^{*}\Omega^{5}_{7} is L2L^{2}-orthogonal to exact 44-forms. To prove the L2L^{2}-orthogonality of the remaining summands we proceed term by term. Let X𝒦X\in\mathcal{K}, d(fφ)dΩ13d(f\varphi)\in d\Omega^{3}_{1} and γΩ274\gamma\in\Omega^{4}_{27}, such that dα=Xφ+d(fφ)+βd\alpha=X\wedge\varphi+d(f\varphi)+\beta for some exact 44-form dαd\alpha. Using the pointwise orthogonality of Ω14\Omega^{4}_{1} and Ω74\Omega^{4}_{7}, we have

Xφ,d(fφ)L2\displaystyle\langle X\wedge\varphi,d(f\varphi)\rangle_{L^{2}} =Xφ,dfφ+4fψL2\displaystyle=\langle X\wedge\varphi,df\wedge\varphi+4f\psi\rangle_{L^{2}}
=Xφ,dfφL2\displaystyle=\langle X\wedge\varphi,df\wedge\varphi\rangle_{L^{2}}
=4X,dfL2=4dX,fL2=0.\displaystyle=4\langle X,df\rangle_{L^{2}}=4\langle d^{*}X,f\rangle_{L^{2}}=0.

Note that since X𝒦X\in\mathcal{K}, Lemma 2.11 (6)(6) implies that Xφ=d(14Xψ)X\wedge\varphi=d\left(-\frac{1}{4}X\lrcorner\psi\right), and hence is exact. Thus, βΩ27,exact4\beta\in\Omega^{4}_{27,\textup{exact}}. Let β=dα0\beta=d\alpha_{0}. The L2L^{2}-orthogonality of Ω274\Omega^{4}_{27} and Ω14\Omega^{4}_{1}, along with the identity φdα=0\varphi\wedge*d\alpha=0 implies

dα0,d(fφ)L2\displaystyle\langle d\alpha_{0},d(f\varphi)\rangle_{L^{2}} =dα0,dfφ+4fψL2\displaystyle=\langle d\alpha_{0},df\wedge\varphi+4f\psi\rangle_{L^{2}}
=dα0,dfφL2+dα0,4fψL2=0.\displaystyle=\langle d\alpha_{0},df\wedge\varphi\rangle_{L^{2}}+\langle d\alpha_{0},4f\psi\rangle_{L^{2}}=0.

The orthogonality of XφX\wedge\varphi and dα0d\alpha_{0} follows from the L2L^{2}-orthogonality of Ω74\Omega^{4}_{7} and Ω274\Omega^{4}_{27}. ∎

Thus, from the previous proposition, we know that any 44-form α\alpha on a nearly G2\mathrm{G}_{2} manifold can be written as α=Xφ+d(fφ)+d(Yψ)+α0\alpha=X\wedge\varphi+d(f\varphi)+d^{*}(Y\wedge\psi)+\alpha_{0}, for some X𝒦,fC(M),YΓ(TM)X\in\mathcal{K},\ f\in C^{\infty}(M),Y\in\Gamma(TM) and α0Ω274\alpha_{0}\in\Omega^{4}_{27}. Since for Y𝒦Y\in\mathcal{K}, d(Yψ)=0d^{*}(Y\wedge\psi)=0, one can choose Y𝒦L2Y\in\mathcal{K}^{\perp_{L^{2}}} in the previous proposition.

Thus for every 44-form α\alpha there exists unique X𝒦,Y𝒦L2,fC(M)X\in\mathcal{K},\ Y\in\mathcal{K}^{\perp_{L^{2}}},\ f\in C^{\infty}(M) and α0Ω274\alpha_{0}\in\Omega^{4}_{27} such that

α=Xφ+d(fφ)+d(Yψ)+α0.\alpha=X\wedge\varphi+d(f\varphi)+d^{*}(Y\wedge\psi)+\alpha_{0}.

3.2 Harmonic 22-forms and 33-forms on nearly G2\mathrm{G}_{2} manifolds

The above decomposition of 44-forms has a very useful application in determining the cohomology of nearly G2\mathrm{G}_{2} manifolds. We first note that since nearly G2\mathrm{G}_{2} manifolds are positive Einstein, it follows from Bochner formula and Hodge theory that any harmonic 11-form is 0 and hence 1(M)=6(M)=0\mathcal{H}^{1}(M)=\mathcal{H}^{6}(M)=0. The next two theorems describe the degree 33, 44 and degree 22 and 55 cohomology of a nearly G2\mathrm{G}_{2} manifold.

Theorem 3.8.

Let (M,φ,ψ)(M,\varphi,\psi) be a complete nearly G2\mathrm{G}_{2} manifold. Then every harmonic 44-form lies in Ω274\Omega^{4}_{27}. Equivalently, every harmonic 33-form lies in Ω273\Omega^{3}_{27}.

Proof.

Let α\alpha be a harmonic 44-form that is dα=dα=0d\alpha=d^{*}\alpha=0. From Proposition 3.7 there exists X𝒦,fC(M),Y𝒦L2X\in\mathcal{K},\ f\in C^{\infty}(M),\ Y\in\mathcal{K}^{\perp_{L^{2}}} and α0Ω274\alpha_{0}\in\Omega^{4}_{27} such that

α\displaystyle\alpha =Xφ+d(fφ)+d(Yψ)+α0.\displaystyle=X\wedge\varphi+d(f\varphi)+d^{*}(Y\wedge\psi)+\alpha_{0}.

Since X𝒦X\in\mathcal{K} and hence 6X=curlX6X=\operatorname{curl}X, by Lemma 2.11 (6)(6), d(Xφ)=4XψΩ73d^{*}(X\wedge\varphi)=4X\lrcorner\psi\in\Omega^{3}_{7} and since d(fφ)=dfφ+4fψΩ174d(f\varphi)=df\wedge\varphi+4f\psi\in\Omega^{4}_{1\oplus 7}, we have

0=α,d(fφ)L2\displaystyle 0=\langle\alpha,d(f\varphi)\rangle_{L^{2}} =Xφ,d(fφ)L2+d(fφ)L22+d(Yψ),d(fφ)L2+α0,d(fφ)L2\displaystyle=\langle X\wedge\varphi,d(f\varphi)\rangle_{L^{2}}+\|d(f\varphi)\|^{2}_{L^{2}}+\langle d^{*}(Y\wedge\psi),d(f\varphi)\rangle_{L^{2}}+\langle\alpha_{0},d(f\varphi)\rangle_{L^{2}}
=d(Xφ),fφL2+d(fφ)L22\displaystyle=\langle d^{*}(X\wedge\varphi),f\varphi\rangle_{L^{2}}+\|d(f\varphi)\|^{2}_{L^{2}}
=d(fφ)L22.\displaystyle=\|d(f\varphi)\|^{2}_{L^{2}}.

Thus d(fφ)=0d(f\varphi)=0 and hence f=0f=0.

Now, 0=dα=d(Xφ)+dα0=4Xψ+dα00=d^{*}\alpha=d^{*}(X\wedge\varphi)+d^{*}\alpha_{0}=4X\lrcorner\psi+d^{*}\alpha_{0}. Using the identity, (Xψ)φ=4X(X\lrcorner\psi)\wedge\varphi=4*X we have

dα0L22\displaystyle\|d^{*}\alpha_{0}\|^{2}_{L^{2}} =16Xψ,XψL2\displaystyle=16\langle X\lrcorner\psi,X\lrcorner\psi\rangle_{L^{2}}
=16X,((Xψ)φ)L2=64XL22.\displaystyle=16\langle X,*((X\lrcorner\psi)\wedge\varphi)\rangle_{L^{2}}=64\|X\|^{2}_{L^{2}}.

On the other hand, again by Lemma 2.11 (6)(6)

dα0L22\displaystyle\|d^{*}\alpha_{0}\|^{2}_{L^{2}} =dα0,dα0L2\displaystyle=\langle d^{*}\alpha_{0},d^{*}\alpha_{0}\rangle_{L^{2}}
=4dα0,XψL2\displaystyle=-4\langle d^{*}\alpha_{0},X\lrcorner\psi\rangle_{L^{2}}
=4α0,d(Xψ)L2=16α0,XφL2=0,\displaystyle=-4\langle\alpha_{0},d(X\lrcorner\psi)\rangle_{L^{2}}=16\langle\alpha_{0},X\wedge\varphi\rangle_{L^{2}}=0,

which implies X=0X=0. So α=d(Yψ)+α0\alpha=d^{*}(Y\wedge\psi)+\alpha_{0}.

Since dα0=0d^{*}\alpha_{0}=0, applying Corollary 2.17 on α0\alpha_{0} implies dα0Ω145d\alpha_{0}\in\Omega^{5}_{14}. This identity together with the closedness of α\alpha gives us

0=α,d(Yψ)L2\displaystyle 0=\langle\alpha,d^{*}(Y\wedge\psi)\rangle_{L^{2}} =d(Yψ)L22+α0,d(Yψ)L2\displaystyle=\|d^{*}(Y\wedge\psi)\|^{2}_{L^{2}}+\langle\alpha_{0},d^{*}(Y\wedge\psi)\rangle_{L^{2}}
=d(Yψ)L22+dα0,YψL2=d(Yψ)L22.\displaystyle=\|d^{*}(Y\wedge\psi)\|^{2}_{L^{2}}+\langle d\alpha_{0},Y\wedge\psi\rangle_{L^{2}}=\|d^{*}(Y\wedge\psi)\|^{2}_{L^{2}}.

as YψΩ75Y\wedge\psi\in\Omega^{5}_{7}. Hence d(Yψ)=0d^{*}(Y\wedge\psi)=0 or equivalently Y𝒦Y\in\mathcal{K}, thus Y=0Y=0 which implies that α=α0\alpha=\alpha_{0} which completes the proof of the theorem. ∎

We also describe the degree 2 (and hence degree 5) cohomology on nearly G2\mathrm{G}_{2} manifolds below. In combination with Theorem 3.8, this completely describes the cohomology of a nearly G2\mathrm{G}_{2} manifold.

Theorem 3.9.

Let (M,φ,ψ)(M,\varphi,\psi) be a complete nearly G2\mathrm{G}_{2} manifold with τ0=4\tau_{0}=4. Let β\beta be a 22-form with

β=β7+β14=(Xφ)+β14for someXΓ(TM).\displaystyle\beta=\beta_{7}+\beta_{14}=(X\lrcorner\varphi)+\beta_{14}\ \ \textup{for\ some}\ X\in\Gamma(TM).

If β\beta is harmonic then βΩ142\beta\in\Omega^{2}_{14}.

Proof.

Suppose βΩ2(M)\beta\in\Omega^{2}(M) is harmonic. Then dβ=dβ=0d\beta=d^{*}\beta=0 and since dd and dd^{*} are linear, we have

dβ7+dβ14=0,dβ7+dβ14=0\displaystyle d\beta_{7}+d\beta_{14}=0,\ \ \ \ \ \ d^{*}\beta_{7}+d^{*}\beta_{14}=0

which on using Lemma 2.11 (3)(3), (4)(4) and (5)(5) imply

37(dX)φ+12((6XcurlX)φ)+iφ(12(Xg)+17(dX)g)+14(dβ14φ)+π27(dβ14)=0\displaystyle-\frac{3}{7}(d^{*}X)\varphi+\frac{1}{2}*((6X-\operatorname{curl}X)\wedge\varphi)+i_{\varphi}\Big{(}\frac{1}{2}(\mathcal{L}_{X}g)+\frac{1}{7}(d^{*}X)g\Big{)}+\frac{1}{4}*(d^{*}\beta_{14}\wedge\varphi)+\pi_{27}(d\beta_{14})=0

and

dβ14=curlX.\displaystyle d^{*}\beta_{14}=-\operatorname{curl}X.

Thus we get

37(dX)φ+12((6XcurlX12curlX)φ)+iφ(12(Xg)+17(dX)g)+π27(dβ14)=0\displaystyle-\frac{3}{7}(d^{*}X)\varphi+\frac{1}{2}*((6X-\operatorname{curl}X-\frac{1}{2}\operatorname{curl}X)\wedge\varphi)+i_{\varphi}\Big{(}\frac{1}{2}(\mathcal{L}_{X}g)+\frac{1}{7}(d^{*}X)g\Big{)}+\pi_{27}(d\beta_{14})=0

and so

dX=0,curlX=4Xand12(Xg)+π27(dβ14)=0.\displaystyle d^{*}X=0,\ \ \ \operatorname{curl}X=4X\ \ \ \textup{and}\ \ \ \frac{1}{2}(\mathcal{L}_{X}g)+\pi_{27}(d\beta_{14})=0. (3.9)

Now curlX=4X\operatorname{curl}X=4X, so taking curl\operatorname{curl} of both sides and using (2.51) with dX=0d^{*}X=0, we get

ΔdX+4curlX=4curlXΔdX=0.\displaystyle\Delta_{d}X+4\operatorname{curl}X=4\operatorname{curl}X\ \ \ \implies\ \ \ \Delta_{d}X=0.

Thus XX is harmonic. Since nearly G2\mathrm{G}_{2} manifolds are positive Einstein, it follows from Bochner formula and Myers theorem that X=0X=0. Hence β=β14Ω142\beta=\beta_{14}\in\Omega^{2}_{14}. ∎

Remark 3.10.

Theorem 3.9 was also proved in a very different way in [BO19, Remark 15]. The theorem has the following interesting interpretation in the context of G2\mathrm{G}_{2}-instantons on a nearly G2\mathrm{G}_{2} manifold, as already described in [BO19, Corollary 14]. For any αH2(M,)\alpha\in H^{2}(M,\mathbb{Z}), by Theorem 3.9, there is a unique G2\mathrm{G}_{2}-instanton on a complex line bundle LL with c1(L)=αc_{1}(L)=\alpha.

Remark 3.11.

It was brought to the attention of the authors by Uwe Semmelmann and Paul-Andi Nagy that Theorem 3.8 also follows from the description of nearly G2\mathrm{G}_{2} manifolds using Killing spinors which is based on an old result of Hijazi saying that the Clifford product of a harmonic form and a Killing spinor vanishes. We also describe degree 22 cohomology by our methods. We believe that the methods and the identities described here, apart from being useful in other contexts, also have the potential to be extended to manifolds with any G2\mathrm{G}_{2} structure (not necessarily nearly G2\mathrm{G}_{2}) with suitable modifications. The authors are currently investigating this.

4.   Deformations of nearly G2\mathrm{G}_{2} structures

Let (M,φ,ψ)(M,\varphi,\psi) be a nearly G2\mathrm{G}_{2} manifold with a nearly G2\mathrm{G}_{2} structure (φ,ψ)(\varphi,\psi). We are interested in studying the deformation problem of (φ,ψ)(\varphi,\psi) in the space of nearly G2\mathrm{G}_{2} structures. The infinitesimal version of this problem was settled by Alexandrov and Semmelmann in [AS12]. We will obtain new proofs of some of their results using the results proved in the previous sections.

Let 𝒫\mathcal{P} be the space of G2\mathrm{G}_{2} structures on MM, that is, the set of all (φ,ψ)Ω+3×Ω+4(\varphi,\psi)\in\Omega^{3}_{+}\times\Omega^{4}_{+} with Θ(φ)=ψ\Theta(\varphi)=\psi. Given a point 𝔭=(φ,ψ)𝒫\mathfrak{p}=(\varphi,\psi)\in\mathcal{P} we define the tangent space T𝔭𝒫T_{\mathfrak{p}}\mathcal{P}.

Lemma 4.1.

The tangent space T𝔭𝒫T_{\mathfrak{p}}\mathcal{P} is the set of all (ξ,η)Ω3(M)×Ω4(M)(\xi,\eta)\in\Omega^{3}(M)\times\Omega^{4}(M) such that

ξ\displaystyle\xi =3fφXψ+γ\displaystyle=3f\varphi-X\lrcorner\psi+\gamma
η\displaystyle\eta =4fψ+Xφγ\displaystyle=4f\psi+X\wedge\varphi-*\gamma

for some fΩ0(M),XΓ(TM)f\in\Omega^{0}(M),\ X\in\Gamma(TM) and γΩ273\gamma\in\Omega^{3}_{27}.

Proof.

The proof immediately follows from equations (2.39) and (2.40) from Proposition 2.4. ∎

4.1 Infinitesimal deformations

We want to study deformations of a given nearly G2\mathrm{G}_{2} structure φ\varphi on a compact manifold MM by nearly G2\mathrm{G}_{2} structures φt\varphi_{t}. We will only be interested in deformations of the nearly G2\mathrm{G}_{2} structures modulo the action of the group ×Diff0(M){\mathbb{R}}^{*}\times\textup{Diff}_{0}(M) where Diff0(M)\textup{Diff}_{0}(M) denotes the space of diffeomorphisms of MM which are isotopic to the identity. We first use Proposition 3.7 to find a slice for the action of diffeomorphism group on 𝒫\mathcal{P} which is used to find the space of infinitesimal nearly G2\mathrm{G}_{2} deformations, a result originally due Alexandrov–Semmelmann [AS12].

For the purposes of doing analysis, we consider the Hölder space 𝒫k,α\mathcal{P}^{k,\alpha} of G2\mathrm{G}_{2} structures on MM such that φ\varphi and ψ\psi are of class Ck,αC^{k,\alpha}, k1,α(0,1)k\geq 1,\ \alpha\in(0,1). Let 𝔭=(φ,ψ)𝒫k,α\mathfrak{p}=(\varphi,\psi)\in\mathcal{P}^{k,\alpha} be a nearly G2\mathrm{G}_{2} structure such that the induced metric is not isometric to round S7S^{7}. Denote the orbit of 𝔭\mathfrak{p} under the action of Diff0k+1,α(M)\textup{Diff}_{0}^{k+1,\alpha}(M)Ck+1,αC^{k+1,\alpha} diffeomorphisms isotopic to the identity, by 𝒪𝔭\mathcal{O}_{\mathfrak{p}}. The tangent space T𝔭𝒪𝔭T_{\mathfrak{p}}\mathcal{O}_{\mathfrak{p}} is the space of Lie derivatives X(φ,ψ)\mathcal{L}_{X}(\varphi,\psi) for XΓ(TM)X\in\Gamma(TM). We are interested in finding a complement 𝒞\mathcal{C} of T𝔭𝒪𝔭T_{\mathfrak{p}}\mathcal{O}_{\mathfrak{p}} in T𝔭𝒫T_{\mathfrak{p}}\mathcal{P}.

If (ξ,η)T𝔭𝒫(\xi,\eta)\in T_{\mathfrak{p}}\mathcal{P} then using Proposition 3.7 (1)(1), we can write

η=Xφ+dfφ+4fψ+d(Yψ)+η0\displaystyle\eta=X\wedge\varphi+df\wedge\varphi+4f\psi+d^{*}(Y\wedge\psi)+\eta_{0}

for unique X𝒦fΩ0(M),Y𝒦L2X\in\mathcal{K}\ f\in\Omega^{0}(M),\ Y\in\mathcal{K}^{\perp_{L^{2}}} and η0Ω274\eta_{0}\in\Omega^{4}_{27}. From Lemma 2.11 (4)(4) we know that

d(Yψ)\displaystyle*d*(Y\wedge\psi) =d(Yφ)=37(dY)ψ(3Y12curlY)φiφ(12(iYj+jYi)+17(dY)gij)\displaystyle=-*d(Y\lrcorner\varphi)=\frac{3}{7}(d^{*}Y)\psi-(3Y-\frac{1}{2}\operatorname{curl}Y)\wedge\varphi-*i_{\varphi}\Big{(}\frac{1}{2}(\nabla_{i}Y_{j}+\nabla_{j}Y_{i})+\frac{1}{7}(d^{*}Y)g_{ij}\Big{)}

and since

Yψ=d(Yψ)=47dYψ(12curlY+Y)φiφ(12(iYj+jYj)+17(dY)gij)\mathcal{L}_{Y}\psi=d(Y\lrcorner\psi)=-\dfrac{4}{7}d^{*}Y\psi-\Big{(}\dfrac{1}{2}\operatorname{curl}Y+Y\Big{)}\wedge\varphi-*i_{\varphi}\Big{(}\dfrac{1}{2}(\nabla_{i}Y_{j}+\nabla_{j}Y_{j})+\dfrac{1}{7}(d^{*}Y)g_{ij}\Big{)}

from Lemma 2.11 (6)(6), we see that

d(Yψ)\displaystyle d^{*}(Y\wedge\psi) =17(dY)ψ+(curlY2Y)φYψ.\displaystyle=-\frac{1}{7}(d^{*}Y)\psi+(\operatorname{curl}Y-2Y)\wedge\varphi-\mathcal{L}_{Y}\psi.

Thus up to an element in T𝔭𝒪𝔭T_{\mathfrak{p}}\mathcal{O}_{\mathfrak{p}} we get that

η=(4f17dY)ψ+(X+df+curlY2Y)φ+η0\displaystyle\eta=\Big{(}4f-\frac{1}{7}d^{*}Y\Big{)}\psi+(X+df+\operatorname{curl}Y-2Y)\wedge\varphi+\eta_{0} (4.1)

and hence from Lemma 4.1

ξ=(3f328dY)φ(X+df+curlY2Y)ψη0.\displaystyle\xi=(3f-\frac{3}{28}d^{*}Y)\varphi-(X+df+\operatorname{curl}Y-2Y)\lrcorner\psi-*\eta_{0}. (4.2)

Now, if X𝒦X\in\mathcal{K} then from Lemma 2.11 (6)(6) and curlX=6X\operatorname{curl}X=6X we see that

X4ψ=d(X4ψ)=Xφ\displaystyle\mathcal{L}_{-\frac{X}{4}}\psi=d\left(-\frac{X}{4}\lrcorner\psi\right)=X\wedge\varphi

and hence

η=X4ψ+d(fφ)+d(Yψ)+η0\eta=\mathcal{L}_{-\frac{X}{4}}\psi+d(f\varphi)+d^{*}(Y\wedge\psi)+\eta_{0}

which implies that up to an element in T𝔭𝒪𝔭T_{\mathfrak{p}}\mathcal{O}_{\mathfrak{p}} combined with the above observation, we can write

η=(4f17dY)ψ+(df+curlY2Y)φ+η0\displaystyle\eta=\Big{(}4f-\frac{1}{7}d^{*}Y\Big{)}\psi+(df+\operatorname{curl}Y-2Y)\wedge\varphi+\eta_{0} (4.3)

which implies that

ξ=(3f328dY)φ(df+curlY2Y)ψη0\displaystyle\xi=(3f-\frac{3}{28}d^{*}Y)\varphi-(df+\operatorname{curl}Y-2Y)\lrcorner\psi-*\eta_{0} (4.4)

and hence we get a splitting T𝔭𝒫=T𝔭𝒪𝔭𝒞T_{\mathfrak{p}}\mathcal{P}=T_{\mathfrak{p}}\mathcal{O}_{\mathfrak{p}}\oplus\mathcal{C} where 𝒞Ω0(M)×𝒦L2×Ω274\mathcal{C}\cong\Omega^{0}(M)\times\mathcal{K}^{\perp_{L^{2}}}\times\Omega^{4}_{27} which consists of (ξ,η)T𝔭𝒫(\xi,\eta)\in T_{\mathfrak{p}}\mathcal{P} of the form (4.4) and (4.3) respectively. This gives a choice of slice. In fact, as discussed in [Nor08, pg. 49 & Theorem 3.1.4] we have

Proposition 4.2.

There exists an open neighbourhood UU of 𝒞\mathcal{C} of the origin such that the “exponentiation” of UU is a slice for the action of Diff0k+1,α(M)\textup{Diff}_{0}^{k+1,\alpha}(M) on a sufficiently small neighbourhood of 𝔭𝒫k,α\mathfrak{p}\in\mathcal{P}^{k,\alpha}.

With this choice of slice we determine the infinitesimal deformations of the nearly G2\mathrm{G}_{2} structure 𝔭\mathfrak{p} which gives a new proof of a result of Alexandrov–Semmelmann [AS12, Theorem 3.5].

Theorem 4.3.

Let (M,φ,ψ)(M,\varphi,\psi) be a complete nearly G2\mathrm{G}_{2} manifold, not isometric to the round S7S^{7}. Then the infinitesimal deformations of the nearly G2\mathrm{G}_{2} structure are in one to one correspondence with (X,ξ0)𝒦×Ω273(X,\xi_{0})\in\mathcal{K}\times\Omega^{3}_{27} with

dξ0=4ξ0andΔX=12X.\displaystyle*d\xi_{0}=-4\xi_{0}\ \ \ \ \ \ \ \ \ \ \textup{and}\ \ \ \ \ \ \ \ \ \ \Delta X=12X. (4.5)

Hence ξ0\xi_{0} is co-closed as well. Moreover, Δdξ0=16ξ0\Delta_{d}\xi_{0}=16\xi_{0}.

Proof.

Let (ξ,η)T𝔭𝒫(\xi,\eta)\in T_{\mathfrak{p}}\mathcal{P} be an infinitesimal nearly G2\mathrm{G}_{2} deformation of a G2\mathrm{G}_{2} structure 𝔭𝒫\mathfrak{p}\in\mathcal{P}. So η\eta must be exact and hence from Proposition 3.7 (2)(2), we can remove the d(Yψ)d^{*}(Y\wedge\psi) term, in which case (4.1) and (4.2) become

η=4fψ+(X+df)φ+η0andξ=3fφ(X+df)ψη0.\displaystyle\eta=4f\psi+(X+df)\wedge\varphi+\eta_{0}\ \ \ \ \ \textup{and}\ \ \ \ \ \xi=3f\varphi-(X+df)\lrcorner\psi-*\eta_{0}. (4.6)

Moreover, for infinitesimal nearly G2\mathrm{G}_{2} deformations we must have

dξ=4ηd\xi=4\eta

and hence (4.6) implies

4fψ+(4X+df)φ+4η0+d((X+df)ψ)+dη0=0.\displaystyle 4f\psi+(4X+df)\wedge\varphi+4\eta_{0}+d((X+df)\lrcorner\psi)+d*\eta_{0}=0.

Using Lemma 2.11 (6)(6) for the fourth term above and taking inner product with ψ\psi gives

28f4d(X+df)=0.\displaystyle 28f-4d^{*}(X+df)=0.

But since X𝒦dX=0X\in\mathcal{K}\implies d^{*}X=0 and hence we get Δf=7f\Delta f=7f. Since MM is not isometric to round S7S^{7}, Obata’s theorem then implies that f=0f=0 and hence

η=Xφ+η0andξ=Xψη0\displaystyle\eta=X\wedge\varphi+\eta_{0}\ \ \ \ \ \textup{and}\ \ \ \ \ \xi=-X\lrcorner\psi-*\eta_{0} (4.7)

which proves the one to one correspondence between the infinitesimal nearly G2\mathrm{G}_{2} deformations and 𝒦×Ω273\mathcal{K}\times\Omega^{3}_{27}. Since Ric=6g\mathrm{Ric}=6g and XX is a Killing vector field, we have ΔX=12X\Delta X=12X which is the second part of (4.5). Since η0\eta_{0} is exact, dη0=0d\eta_{0}=0. From (4.7) and the fact that dξ=4ηd\xi=4\eta, we get

dη0=4η0\displaystyle d*\eta_{0}=-4\eta_{0}

and hence

dξ0=4ξ0.\displaystyle*d\xi_{0}=-4\xi_{0}.

Taking dd^{*} of both sides give dξ0=0d^{*}\xi_{0}=0. Moreover,

Δdξ0=ddξ0=4dξ0=4(dξ0)=16ξ0\displaystyle\Delta_{d}\xi_{0}=d^{*}d\xi_{0}=-4d^{*}*\xi_{0}=-4*(d\xi_{0})=16\xi_{0}

which completes the proof of the theorem. ∎

Remark 4.4.

From the computations for the proof of Proposition 4.2 we know that for X𝒦X\in\mathcal{K},

4Xφ=Xψ.\displaystyle-4X\wedge\varphi=\mathcal{L}_{X}\psi.

Thus, from Theorem 4.3 we see that the infinitesimal deformations of a nearly G2\mathrm{G}_{2} structure modulo diffeomorphisms are in one-to-one correspondence with ξ0Ω273\xi_{0}\in\Omega^{3}_{27} such that dξ0=4ξ0*d\xi_{0}=-4\xi_{0}.

Motivated from the study of deformations of nearly Kähler 66-manifolds by Foscolo [Fos17, §4] where he used observations of Hitchin [Hit01], we also want to interpret the nearly G2\mathrm{G}_{2} condition (2.24) as the vanishing of a smooth map on the space of exact positive 44-forms. Moreover, in order to study the second order deformations, it will be convenient to enlarge the space by introducing a vector field as an additional parameter which is natural when one considers the action of the diffeomorphism group. We elaborate on this below.

Let ψ=dα\psi=d\alpha be an exact positive 44-form, not necessarily satisfying the nearly G2\mathrm{G}_{2} condition. Let ηΩexact4\eta\in\Omega^{4}_{\operatorname{\textup{exact}}} be the first order deformation of ψ\psi. Hitchin in [Hit01] defined a volume functional for exact 44-form ρ=dγ\rho=d\gamma given by

V(ρ)\displaystyle V(\rho) =Mρρ,\displaystyle=\int_{M}*\rho\wedge\rho,

and a quadratic form

W(ρ,ρ)\displaystyle W(\rho,\rho^{\prime}) =Mγρ=Mργ,\displaystyle=\int_{M}\gamma\wedge\rho^{\prime}=\int_{M}\rho\wedge\gamma^{\prime},

where ρ=dγ\rho=d\gamma and ρ=dγ\rho^{\prime}=d\gamma^{\prime} are exact 44-forms. We denote W(ρ,ρ)W(\rho,\rho) by W(ρ)W(\rho). When MM is compact, Hitchin proves [Hit01, Theorem 5] that stable 44-forms (which is the same as a positive 44-form in our case) ηΩexact4(M)\eta\in\Omega^{4}_{\operatorname{\textup{exact}}}(M) is a critical point of the volume functional VV subject to the constraint W(η)=constantW(\eta)=\textup{constant} if and only if η\eta defines a nearly G2\mathrm{G}_{2} structure. The linearization of the volume functional at ψ\psi is given by

dV(η)=ddt|t=0V(ψ+tη)\displaystyle dV(\eta)=\left.\frac{d}{dt}\right|_{t=0}V(\psi+t\eta) =Mφη+Mηψ\displaystyle=\int_{M}\varphi\wedge\eta+\int_{M}*\eta\wedge\psi
=2Mφη.\displaystyle=2\int_{M}\varphi\wedge\eta.

For the linearization of the quadratic form, suppose ψ\psi is exact with ψ=dα\psi=d\alpha. We use integration by parts to get

dW(η)=ddt|t=0W(ψ+tη)\displaystyle dW(\eta)=\left.\frac{d}{dt}\right|_{t=0}W(\psi+t\eta) =Mαη+Mγψ\displaystyle=\int_{M}\alpha\wedge\eta+\int_{M}\gamma\wedge\psi
=2Mαη.\displaystyle=2\int_{M}\alpha\wedge\eta.

Let us define an energy functional EE on exact 44-forms by

E(ρ)\displaystyle E(\rho) :=V(ρ)4W(ρ).\displaystyle:=V(\rho)-4W(\rho).

Then from above calculations

dE(η)\displaystyle dE(\eta) =M(φ4α)η=Md((φ4α)γ.\displaystyle=\int_{M}(\varphi-4\alpha)\wedge\eta=\int_{M}d((\varphi-4\alpha)\wedge\gamma.

Therefore ψ=dα\psi=d\alpha is a critical point of EE if and only if dE(η)=0dE(\eta)=0 for every ηΩexact4\eta\in\Omega^{4}_{\operatorname{\textup{exact}}} that is if and only if

dφ4dα\displaystyle d\varphi-4d\alpha =dφ4ψ=0.\displaystyle=d\varphi-4\psi=0.

Hence the critical points of the functional EE on Ω+,exact4\Omega^{4}_{+,\operatorname{\textup{exact}}} are nearly G2\mathrm{G}_{2} structures. Since the energy functional EE is diffeomorphism invariant, we can introduce an extra vector field, as dEdE will vanish in the direction of Lie derivatives. Thus ψ\psi being a stable exact 44-form can be given by the formula

ψ\displaystyle\psi =14d(φd(Zψ))\displaystyle=\frac{1}{4}d(\varphi-*d(Z\lrcorner\psi))

for some ZΓ(TM)Z\in\Gamma(TM). We use these observations to write the nearly G2\mathrm{G}_{2} condition (2.24) as the vanishing of a smooth map. Let us denote by P^\widehat{P} the space of stable 33 and stable, exact 44-forms, i.e., (φ,ψ)Ω+3×Ω+,exact4(\varphi,\psi)\in\Omega^{3}_{+}\times\Omega^{4}_{+,\operatorname{\textup{exact}}}. We have the following

Proposition 4.5.

Suppose (φ,ψ)𝒫^(\varphi,\psi)\in\widehat{\mathcal{P}} satisfies

dφ4ψ=dd(Zψ)\displaystyle d\varphi-4\psi=d*d(Z\lrcorner\psi) (4.8)

for some vector field ZZ and * denotes the Hodge star with respect to a fixed background metric. Then (φ,ψ)(\varphi,\psi) is a nearly G2\mathrm{G}_{2} structure.

Proof.

We will prove that d(Zψ)=0d(Z\lrcorner\psi)=0. We note from (2.32) that

(Zψ)ψ=0(Z\lrcorner\psi)\wedge\psi=0

So from (4.8) we get that

d(Zψ)L22\displaystyle\|d(Z\lrcorner\psi)\|^{2}_{L^{2}} =d(Zψ),d(Zψ)L2\displaystyle=\langle d(Z\lrcorner\psi),d(Z\lrcorner\psi)\rangle_{L^{2}}
=(Zψ),dd(Zψ)L2\displaystyle=\langle(Z\lrcorner\psi),*d*d(Z\lrcorner\psi)\rangle_{L^{2}}
=(Zψ),(dφ4ψ)L2\displaystyle=\langle(Z\lrcorner\psi),*(d\varphi-4\psi)\rangle_{L^{2}}
=M(Zψ)(dφ4ψ)=M(Zψ)dφ\displaystyle=\int_{M}(Z\lrcorner\psi)\wedge(d\varphi-4\psi)=\int_{M}(Z\lrcorner\psi)\wedge d\varphi

Since φ\varphi is a G2\mathrm{G}_{2} structure and dψ=0d\psi=0 from (4.8), we know from (2.19) that τ1=0\tau_{1}=0 and hence dφd\varphi has no component in Ω74\Omega^{4}_{7}. Thus

(Zψ),dφ=0\langle(Z\lrcorner\psi),*d\varphi\rangle=0

which implies that

d(Zψ)L22=M(Zψ)dφ=0\displaystyle\|d(Z\lrcorner\psi)\|^{2}_{L^{2}}=\int_{M}(Z\lrcorner\psi)\wedge d\varphi=0

which proves the proposition. ∎

Suppose we want to describe the local moduli space of nearly G2\mathrm{G}_{2} structures on a manifold MM. If 𝒩𝒫\mathcal{NP} denotes the space of nearly G2\mathrm{G}_{2} structures on MM then the local moduli space is =𝒩𝒫/Diff0(M)\mathcal{M}=\mathcal{NP}/\textup{Diff}_{0}(M). A natural way to study this problem is to view the nearly G2\mathrm{G}_{2} structures on MM as the zero locus of an appropriate function, find the linearization of the function and prove its surjectivity, so that an Implicit Function Theorem argument describes \mathcal{M}.

Now let (φ,ψ)(\varphi,\psi) be a nearly G2\mathrm{G}_{2} structure on MM. Let UΩ+,exact4U\subset\Omega^{4}_{+,\operatorname{\textup{exact}}} be a small neighborhood of the 44-form ψ\psi. Since the condition of being stable is open we can assume the existence of such a neighborhood. Thus for ηΩexact4\eta\in\Omega^{4}_{\operatorname{\textup{exact}}} with sufficiently small norm with respect to the metric induced by φ\varphi, ψ~=ψ+η\tilde{\psi}=\psi+\eta is also a stable exact 44-form. From Proposition 4.5 the pair of stable forms (φ~,ψ~)(\tilde{\varphi},\tilde{\psi}) defines a nearly G2\mathrm{G}_{2} structure if there exists a ZΓ(TM)Z\in\Gamma(TM) such that

dφ~4ψ~=dd(Zψ~).\displaystyle d\tilde{\varphi}-4\tilde{\psi}=d*d(Z\lrcorner\tilde{\psi}).

This condition is equivalent to the vanishing of the map

Φ\displaystyle\Phi :U×Γ(TM)Ωexact4\displaystyle:U\times\Gamma(TM)\to\Omega^{4}_{\operatorname{\textup{exact}}}
(ψ~,Z)\displaystyle(\tilde{\psi},Z) dψ~4ψ~dd(Zψ~).\displaystyle\mapsto d*\tilde{\psi}-4\tilde{\psi}-d*d(Z\lrcorner\tilde{\psi}). (4.9)

Thus, the nearly G2\mathrm{G}_{2} structures are the zero locus of the map Φ\Phi modulo diffeomorphisms.

Let ξ\xi be the dual of η\eta under the Hitchin’s duality map Θ\Theta as in Proposition 2.4. The linearization of the map Φ\Phi at the point (ψ,0)(\psi,0) is given by

dξ4η=dd(Zψ).\displaystyle d\xi-4\eta=d*d(Z\lrcorner\psi).

Thus the obstructions on the first order deformations of the nearly G2\mathrm{G}_{2} structure (φ,ψ)(\varphi,\psi) are given by Im(DΦ)\operatorname{Im}(D\Phi) which is characterized in the following proposition, whose proof is inspired from a similar theorem in the nearly Kähler case by Foscolo [Fos17, Proposition 4.5].

Proposition 4.6.

Let (φ,ψ)(\varphi,\psi) be a nearly G2\mathrm{G}_{2} structure and (ξ,η)Ω3×Ωexact4(\xi,\eta)\in\Omega^{3}\times\Omega^{4}_{\operatorname{\textup{exact}}} be a first order deformation in 𝒫\mathcal{P}. Then αΩexact4\alpha\in\Omega^{4}_{\operatorname{\textup{exact}}} lies in the image of DΦD\Phi if and only if

dα4α,χL2=0\displaystyle\langle d^{*}\alpha-4*\alpha,\chi\rangle_{L^{2}}=0

for all co-closed χΩ273\chi\in\Omega^{3}_{27} such that Δχ=16χ\Delta\chi=16\chi.

Proof.

From Proposition 3.7 (2)(2), there exists X𝒦,fC(M)X\in\mathcal{K},f\in C^{\infty}(M) and η0Ω27,exact4\eta_{0}\in\Omega^{4}_{27,\operatorname{\textup{exact}}} such that

η\displaystyle\eta =Xφ+d(fφ)+η0\displaystyle=X\wedge\varphi+d(f\varphi)+\eta_{0}
=d(14Xψ+fφ)+η0\displaystyle=d\left(-\frac{1}{4}X\lrcorner\psi+f\varphi\right)+\eta_{0}

and from Lemma 4.1, the 33-form

ξ\displaystyle\xi =3fφ(df+X)ψη0.\displaystyle=3f\varphi-(df+X)\lrcorner\psi-*\eta_{0}.

By Proposition 3.7, α=Yφ+d(hφ)+α0\alpha=Y\wedge\varphi+d(h\varphi)+\alpha_{0} for some Y𝒦,hC(M),α0Ω27,exact4Y\in\mathcal{K},h\in C^{\infty}(M),\alpha_{0}\in\Omega^{4}_{27,\operatorname{\textup{exact}}}. Such an α\alpha lies in the image of DΦD\Phi if

dξ4ηdd(Zψ)\displaystyle d\xi-4\eta-d*d(Z\lrcorner\psi) =α=d(14Yψ+hφ)+α0.\displaystyle=\alpha=d\Big{(}-\frac{1}{4}Y\lrcorner\psi+h\varphi\Big{)}+\alpha_{0}.

From Lemma 2.11 (5)

d(Zψ)\displaystyle d^{*}(Z\wedge\psi) =d(Zφ)\displaystyle=-*d(Z\lrcorner\varphi)
=37(dZ)ψ12(6ZcurlZ)φiφ(12(iZj+jZi)+17(dZ)gij)\displaystyle=\dfrac{3}{7}(d^{*}Z)\psi-\dfrac{1}{2}\Big{(}6Z-\operatorname{curl}Z\Big{)}\wedge\varphi-*i_{\varphi}\Big{(}\dfrac{1}{2}(\nabla_{i}Z_{j}+\nabla_{j}Z_{i})+\dfrac{1}{7}(d^{*}Z)g_{ij}\Big{)}

Comparing the last term in the above expression with that of d(Zψ)d(Z\lrcorner\psi) in Lemma 2.11 we get

d(Zψ)\displaystyle d(Z\lrcorner\psi) =17dZψ+(2ZcurlZ)φ+d(Zψ).\displaystyle=\dfrac{1}{7}d^{*}Z\psi+(2Z-\operatorname{curl}Z)\wedge\varphi+d^{*}(Z\wedge\psi).

Using these expressions for ξ,η\xi,\eta and d(Zψ)d(Z\lrcorner\psi) we get

dξ4ηdd(Zψ)\displaystyle d\xi-4\eta-d*d(Z\lrcorner\psi) =d((f17dZ)φ(df2Z+curlZ)ψ)dη04η0.\displaystyle=d((-f-\dfrac{1}{7}d^{*}Z)\varphi-(df-2Z+\operatorname{curl}Z)\lrcorner\psi)-d*\eta_{0}-4\eta_{0}.

Thus, for finding the Im(DΦ)\operatorname{Im}(D\Phi), we need to solve the equations

f+17dZ\displaystyle f+\dfrac{1}{7}d^{*}Z =h\displaystyle=-h (4.10)
df2Z+curlZ\displaystyle df-2Z+\operatorname{curl}Z =14Y\displaystyle=\dfrac{1}{4}Y
dη04η0\displaystyle-d*\eta_{0}-4\eta_{0} =α0.\displaystyle=\alpha_{0}.

Let α0=0\alpha_{0}=0. Then by Implicit Function Theorem, a solution of the first pair of equations always exist if the operator

D~:Ω0×Ω1\displaystyle\tilde{D}\ \colon\ \Omega^{0}\times\Omega^{1} Ω0×Ω1\displaystyle\to\Omega^{0}\times\Omega^{1}
(f,Z)\displaystyle(f,Z) (f+17dZ,df2Z+curlZ)\displaystyle\mapsto\Big{(}f+\dfrac{1}{7}d^{*}Z,df-2Z+\operatorname{curl}Z\Big{)}

is invertible in a small neighborhood of its zero locus. Since D~\tilde{D} differs from the modified Dirac operator DD in (3.6) only by self-adjoint zeroth-order term, it is self-adjoint and hence ker(D~)=coker(D~)\ker(\tilde{D})=\text{coker}(\tilde{D}). A pair (f,Z)(f,Z) is in the kernel of the operator DD if and only if

f+17dZ\displaystyle f+\dfrac{1}{7}d^{*}Z =0\displaystyle=0
df2Z+curlZ\displaystyle df-2Z+\operatorname{curl}Z =0.\displaystyle=0.

Applying the operator dd^{*} on the second equation and using the fact that d(curlZ)=0d^{*}(\operatorname{curl}Z)=0 gives

0\displaystyle 0 =ddf2dZ=ddf+14f.\displaystyle=d^{*}df-2d^{*}Z=d^{*}df+14f.

Thus f=0f=0 as Δ\Delta is a non-negative operator. The second equation then becomes

curlZ\displaystyle\operatorname{curl}Z =d(Zφ)=(dZψ)=2Z\displaystyle=d^{*}(Z\lrcorner\varphi)=*(dZ\wedge\psi)=2Z

and Proposition 2.5 implies that dZ=23Zφ+π14(dZ)dZ=\dfrac{2}{3}Z\lrcorner\varphi+\pi_{14}(dZ). Using Lemma 2.9 (2) we get that

M𝑑ZdZφ\displaystyle\int_{M}dZ\wedge dZ\wedge\varphi =89Zφ2π14(dZ)2\displaystyle=\dfrac{8}{9}\|Z\lrcorner\varphi\|^{2}-\|\pi_{14}(dZ)\|^{2}
=83Z2π14(dZ)2.\displaystyle=\dfrac{8}{3}\|Z\|^{2}-\|\pi_{14}(dZ)\|^{2}.

On the other hand

M𝑑ZdZφ\displaystyle\int_{M}dZ\wedge dZ\wedge\varphi =4MZdZψ=8Z2.\displaystyle=4\int_{M}Z\wedge dZ\wedge\psi=8\|Z\|^{2}.

Combining these two equations we get 163Z2=π14(dZ)2\dfrac{16}{3}\|Z\|^{2}=-\|\pi_{14}(dZ)\|^{2} and hence Z=0Z=0 as well. Thus ker(D~)=coker(D~)=0\ker(\tilde{D})=\text{coker}(\tilde{D})=0 and D~\tilde{D} is invertible when α0=0\alpha_{0}=0 and we can always solve the first pair of equations in (4.10). Thus there are no restrictions on Y,hY,h to be in the image of DΦD\Phi. Moreover if α00\alpha_{0}\neq 0 satisfies the third equation in (4.10) then

dα0\displaystyle d^{*}\alpha_{0} =ddη04dη0,\displaystyle=-d^{*}d*\eta_{0}-4d^{*}\eta_{0},
α0\displaystyle*\alpha_{0} =dη04η0\displaystyle=-d^{*}\eta_{0}-4*\eta_{0}

which on using the fact that η0*\eta_{0} is co-closed implies

dα04α0\displaystyle d^{*}\alpha_{0}-4*\alpha_{0} =16η0ddη0=16η0Δdη0.\displaystyle=16*\eta_{0}-d^{*}d*\eta_{0}=16*\eta_{0}-\Delta_{d}*\eta_{0}.

Thus α0Ω27,exact4\alpha_{0}\in\Omega^{4}_{27,\operatorname{\textup{exact}}} is a solution to the equation (4.10) (3) if and only if

dα04α0,ξ0L2=0\displaystyle\langle d^{*}\alpha_{0}-4*\alpha_{0},\xi_{0}\rangle_{L^{2}}=0

for all co-closed ξ0Ω273\xi_{0}\in\Omega^{3}_{27} such that Δξ=16ξ\Delta\xi=16\xi. To complete the proof of the proposition we now only need to prove the L2L^{2}-orthogonality condition for α\alpha. But observe that since Y𝒦Y\in\mathcal{K}

dα\displaystyle d^{*}\alpha =d(Yφ)+dd(hφ)+dα0=4Yψ+dd(hφ)+dα0,\displaystyle=d^{*}(Y\wedge\varphi)+d^{*}d(h\varphi)+d^{*}\alpha_{0}=-4Y\lrcorner\psi+d^{*}d(h\varphi)+d^{*}\alpha_{0},

and so dα4α=dd(hφ)4d(hφ)+dα04α0d^{*}\alpha-4*\alpha=d^{*}d(h\varphi)-4*d(h\varphi)+d^{*}\alpha_{0}-4*\alpha_{0}. Since ξ\xi is co-closed, from Corollary 2.15 dξΩ274d\xi\in\Omega^{4}_{27} and

dd(hφ),ξL2\displaystyle\langle d^{*}d(h\varphi),\xi\rangle_{L^{2}} =d(hφ),dξL2=0.\displaystyle=\langle d(h\varphi),d\xi\rangle_{L^{2}}=0.

Similarly

d(hφ),ξL2\displaystyle\langle*d(h\varphi),\xi\rangle_{L^{2}} =d(hψ),ξL2=hψ,dξL2=0\displaystyle=\langle d^{*}(h\psi),\xi\rangle_{L^{2}}=\langle h\psi,d\xi\rangle_{L^{2}}=0

which completes the proof of the proposition. ∎

Remark 4.7.

Proposition 4.6 puts a very strong restriction on the first order deformations of a nearly G2\mathrm{G}_{2} structure to be unobstructed.

4.2 Second-order deformations

Following the work of Koiso [Koi82] on deformations of Einstein metrics and the work of Foscolo [Fos17] on the second order deformations of nearly Kähler structures on 66-manifolds, we define the notion of second order deformations of nearly G2\mathrm{G}_{2} structures.

Definition 4.8.

Given a nearly G2\mathrm{G}_{2} structure (φ0,ψ0)(\varphi_{0},\psi_{0}) and an infinitesimal deformation (ξ1,η1)(\xi_{1},\eta_{1}), a second order deformation of (φ0,ψ0)(\varphi_{0},\psi_{0}) in the direction of (ξ1,η1)(\xi_{1},\eta_{1}) is a pair (ξ2,η2)Ω3×Ω4(\xi_{2},\eta_{2})\in\Omega^{3}\times\Omega^{4} such that

φ=φ0+ϵξ1+ϵ22ξ2,ψ=ψ0+ϵη1+ϵ22η2\displaystyle\varphi=\varphi_{0}+\epsilon\xi_{1}+\frac{\epsilon^{2}}{2}\xi_{2},\ \ \ \ \ \ \psi=\psi_{0}+\epsilon\eta_{1}+\frac{\epsilon^{2}}{2}\eta_{2}

is a nearly G2\mathrm{G}_{2} structure up to terms of order O(ϵ2)O(\epsilon^{2}). An infinitesimal deformation (ξ1,η1)(\xi_{1},\eta_{1}) is said to be obstructed to second order if there exists no second-order deformation in its direction.

Remark 4.9.

Second order deformations are the same as the second derivative of a curve of nearly G2\mathrm{G}_{2} structures on a manifold MM.

Remark 4.10.

In a similar way, we can define higher order deformations of a nearly G2\mathrm{G}_{2} structure.

Following the discussion in the previous section and in particular Proposition 4.5, in order to find second order deformations of a given nearly G2\mathrm{G}_{2} structure (φ0,ψ0)(\varphi_{0},\psi_{0}), we look for formal power series defining positive exact 44-form

ψϵ=ψ0+ϵη1+ϵ22η2+\displaystyle\psi_{\epsilon}=\psi_{0}+\epsilon\eta_{1}+\frac{\epsilon^{2}}{2}\eta_{2}+\cdots

where ηiΩexact4\eta_{i}\in\Omega^{4}_{\operatorname{\textup{exact}}} and a vector field

Zϵ=ϵZ1+ϵ22Z2+\displaystyle Z_{\epsilon}=\epsilon Z_{1}+\frac{\epsilon^{2}}{2}Z_{2}+\cdots

which satisfy (4.8), that is

dφϵ4ψϵ=dd(Zϵψϵ)\displaystyle d\varphi_{\epsilon}-4\psi_{\epsilon}=d*d(Z_{\epsilon}\lrcorner\psi_{\epsilon}) (4.11)

where φϵ\varphi_{\epsilon} is the dual of ψϵ\psi_{\epsilon}. Note that the Hodge star * is taken with respect to φϵ\varphi_{\epsilon}.

Since we are interested in second order deformations, given an infinitesimal nearly G2\mathrm{G}_{2} deformation (ξ1,η1)(\xi_{1},\eta_{1}), we set Z1=0Z_{1}=0 and look for η2Ωexact4\eta_{2}\in\Omega^{4}_{\operatorname{\textup{exact}}} such that (4.11) is satisfied upto terms of O(ϵ3)O(\epsilon^{3}). Explicitly, we write

φϵ=φ0+ϵξ1+ϵ22(η2^Q3(η1))\displaystyle\varphi_{\epsilon}=\varphi_{0}+\epsilon\xi_{1}+\frac{\epsilon^{2}}{2}(\widehat{\eta_{2}}-Q_{3}(\eta_{1}))

where η2^\widehat{\eta_{2}} denotes the linearization of Hitchin’s duality map Θ\Theta for stable forms in Proposition 2.4 and Q3(η1)Q_{3}(\eta_{1}) is the quadratic term of Hitchin’s duality map. Since we want solutions to (4.11) up to second order, we look for η2\eta_{2} such that

dη2^4η2=d(Q3(η1))+dd(Z2ψ0)\displaystyle d\widehat{\eta_{2}}-4\eta_{2}=d(Q_{3}(\eta_{1}))+d*d(Z_{2}\lrcorner\psi_{0}) (4.12)

as Z1=0Z_{1}=0 and Z2ψ0Z_{2}\lrcorner\psi_{0} is the only second order term in ZϵψϵZ_{\epsilon}\lrcorner\psi_{\epsilon}. We know from Proposition 4.6 that there are obstructions to finding second order deformations and hence in solving the above equation. We want to establish a one-to-one correspondence between second order deformations of a nearly G2\mathrm{G}_{2} structure and solutions to (4.12). We do this in the following lemma.

Lemma 4.11.

Suppose η2\eta_{2} is a solution of (4.12). Then d(Z2ψ0)=0d(Z_{2}\lrcorner\psi_{0})=0 and (η2^Q3(η1),η2)(\widehat{\eta_{2}}-Q_{3}(\eta_{1}),\eta_{2}) defines a second-order deformation of (φ0,ψ0)(\varphi_{0},\psi_{0}) in the direction of (ξ1,η1)(\xi_{1},\eta_{1}) in the sense of Definition 4.8. Conversely, every second order deformation (ξ2,η2)(\xi_{2},\eta_{2}) is a solution to (4.12).

Proof.

We start with

d(Z2ψ0)L22\displaystyle\|d(Z_{2}\lrcorner\psi_{0})\|^{2}_{L^{2}} =Z2ψ0,dd(Z2ψ0)L2\displaystyle=\langle Z_{2}\lrcorner\psi_{0},d^{*}d(Z_{2}\lrcorner\psi_{0})\rangle_{L^{2}}
=Z2ψ0,dd(Z2ψ0)L2\displaystyle=\langle Z_{2}\lrcorner\psi_{0},*d*d(Z_{2}\lrcorner\psi_{0})\rangle_{L^{2}}
=Z2ψ0,(dη2^4η2dQ3(η1))L2\displaystyle=\langle Z_{2}\lrcorner\psi_{0},*(d\widehat{\eta_{2}}-4\eta_{2}-dQ_{3}(\eta_{1}))\rangle_{L^{2}}

Since dψϵ=O(ϵ3)d\psi_{\epsilon}=O(\epsilon^{3}), hence from (2.18) and (2.19) we see that for any vector field YY, 𝑑φϵ(Yψϵ)=O(ϵ3)\int d\varphi_{\epsilon}\wedge(Y\lrcorner\psi_{\epsilon})=O(\epsilon^{3}). Thus the terms which are O(ϵ2)O(\epsilon^{2}) in 𝑑φϵ(Yψϵ)\int d\varphi_{\epsilon}\wedge(Y\lrcorner\psi_{\epsilon}) vanish, that is

𝑑φ0(Yη2)+dξ1η1+d(η2^Q3(η1))(Yψ0)=0.\displaystyle\int d\varphi_{0}\wedge(Y\lrcorner\eta_{2})+d\xi_{1}\wedge\eta_{1}+d(\widehat{\eta_{2}}-Q_{3}(\eta_{1}))\wedge(Y\lrcorner\psi_{0})=0.

Using the facts that dφ0=4ψ0d\varphi_{0}=4\psi_{0}, dξ1η1=0d\xi_{1}\wedge\eta_{1}=0, being an 88-form on a seven dimensional manifold and (Yη2)ψ0=(Yψ0)η2(Y\lrcorner\eta_{2})\wedge\psi_{0}=-(Y\lrcorner\psi_{0})\wedge\eta_{2} we get that

d(η2^Q3(η1))(Yψ0)4η2(Yψ0)=0\displaystyle\int d(\widehat{\eta_{2}}-Q_{3}(\eta_{1}))\wedge(Y\lrcorner\psi_{0})-4\eta_{2}\wedge(Y\lrcorner\psi_{0})=0

Taking Y=Z2Y=Z_{2} proves that d(Z2ψ0)=0d(Z_{2}\lrcorner\psi_{0})=0. From (4.12) we get that

d(η2^Q3(η1))=4η2d(\widehat{\eta_{2}}-Q_{3}(\eta_{1}))=4\eta_{2}

which proves that ((η2^Q3(η1),η2))((\widehat{\eta_{2}}-Q_{3}(\eta_{1}),\eta_{2})) is a second-order deformation of (φ0,ψ0)(\varphi_{0},\psi_{0}) in the direction of (ξ1,η1)(\xi_{1},\eta_{1}) in the sense of Definition 4.8. Conversely, suppose that (ξ2,η2)(\xi_{2},\eta_{2}) is a second-order deformation of (φ0,ψ0)(\varphi_{0},\psi_{0}). Then dξ2=4η2d\xi_{2}=4\eta_{2}. ∎

From the previous proposition and Proposition 4.6 we have that if (ξ2,η2)(\xi_{2},\eta_{2}) is a second order deformation of the nearly G2\mathrm{G}_{2} structure (φ0,ψ0)(\varphi_{0},\psi_{0}) in the sense of Definition 4.8 then

ddQ3(η1)4dQ3(η1),χL2=0\displaystyle\langle d^{*}dQ_{3}(\eta_{1})-4*dQ_{3}(\eta_{1}),\chi\rangle_{L^{2}}=0 (4.13)

for all χΩ273\chi\in\Omega^{3}_{27} such that dχ=0,Δχ=16χd^{*}\chi=0,\Delta\chi=16\chi. The above equation simplifies to

Q3(η1),dχ4χL2\displaystyle\langle*Q_{3}(\eta_{1}),d\chi-4*\chi\rangle_{L^{2}} =0.\displaystyle=0.

Moreover, if χ\chi is an infinitesimal deformation of (φ0,ψ0)(\varphi_{0},\psi_{0}), then by Theorem 4.3 χ\chi satisfies dχ=4χd\chi=-4*\chi (which of course implies dχ=0andΔχ=16χd^{*}\chi=0\ \textup{and}\ \Delta\chi=16\chi) and so the above equation is equivalent to

Q3(η1),χL2\displaystyle\langle Q_{3}(\eta_{1}),\chi\rangle_{L^{2}} =0.\displaystyle=0.

5.   Deformations on the Aloff-Wallach space

In [AS12, Prop. 8.3] Alexandrov–Semmelmann established that the space of infinitesimal deformations of the nearly G2\mathrm{G}_{2} structure on the Aloff–Wallach space X1,1SU(3)×SU(2)SU(2)×U(1)X_{1,1}\cong\frac{\mathrm{SU}(3)\times\mathrm{SU(2)}}{\mathrm{SU}(2)\times\mathrm{U}(1)} is an eight dimensional space isomorphic to 𝔰𝔲(3)\mathfrak{su}(3), the Lie algebra of SU(3)\mathrm{SU}(3). The rest of the paper is devoted to prove that these deformations are obstructed to second order.

The embedding of 𝔰𝔲(2)\mathfrak{su}(2) and 𝔲(1)\mathfrak{u}(1) in 𝔰𝔲(3)𝔰𝔲(2)\mathfrak{su}(3)\oplus\mathfrak{su}(2), which we denote by 𝔰𝔲(2)d\mathfrak{su}(2)_{d} and 𝔲(1)\mathfrak{u}(1), following [AS12], is given by

𝔰𝔲(2)d\displaystyle\mathfrak{su}(2)_{d} ={((a000),a)a𝔰𝔲(2)},\displaystyle=\left\{\Big{(}\begin{pmatrix}a&0\\ 0&0\end{pmatrix},a\Big{)}\mid a\in\mathfrak{su}(2)\right\},
𝔲(1)\displaystyle\mathfrak{u}(1) =span{C}=span{((i000i0002i),0)}.\displaystyle=\text{span}\{C\}=\text{span}\left\{(\begin{pmatrix}i&0&0\\ 0&i&0\\ 0&0&-2i\end{pmatrix},0)\right\}.

The Lie algebra 𝔰𝔲(3)𝔰𝔲(2)\mathfrak{su}(3)\oplus\mathfrak{su}(2) splits as

𝔰𝔲(3)𝔰𝔲(2)\displaystyle\mathfrak{su}(3)\oplus\mathfrak{su}(2) =𝔰𝔲(2)𝔲(1)𝔪\displaystyle=\mathfrak{su}(2)\oplus\mathfrak{u}(1)\oplus\mathfrak{m}

where 𝔪\mathfrak{m} is the 77-dimensional orthogonal complement of 𝔰𝔲(2)𝔲(1)\mathfrak{su}(2)\oplus\mathfrak{u}(1) with respect to BB, the Killing form of 𝔰𝔲(3)𝔰𝔲(2)\mathfrak{su}(3)\oplus\mathfrak{su}(2). The normal nearly G2\mathrm{G}_{2} metric on X1,1X_{1,1} is then given by 340B-\frac{3}{40}B where the constant 340-\frac{3}{40} comes from our choice of τ0=4\tau_{0}=4. If we denote by WW the standard 22-dimensional complex irreducible representation of SU(2)\mathrm{SU}(2) and by F(k)F(k) the 11-dimensional complex irreducible representation of U(1)\mathrm{U}(1) with highest weight kk, then as an SU(2)×U(1)\mathrm{SU}(2)\times\mathrm{U}(1)-representation

𝔰𝔲(3)\displaystyle\mathfrak{su}(3)_{\mathbb{C}} S2WWF(3)WF(3).\displaystyle\cong S^{2}W\oplus WF(3)\oplus WF(-3)\oplus\mathbb{C}.

Let {ei}i=17\{e_{i}\}_{i=1}^{7} be the basis of 𝔪\mathfrak{m}. If we define I=(i00i),J=(0110)andK=(0ii0)I=\begin{pmatrix}i&0\\ 0&-i\end{pmatrix},J=\begin{pmatrix}0&-1\\ 1&0\end{pmatrix}\ \textup{and}\ K=\begin{pmatrix}0&i\\ i&0\end{pmatrix}, we have

e1\displaystyle e_{1} :=13((2I000),3I),e2:=13((2J000),3J),e3:=13((2K000),3K),\displaystyle:=\frac{1}{3}\left(\begin{pmatrix}2I&0\\ 0&0\end{pmatrix},-3I\right),\ \ \ e_{2}:=\frac{1}{3}\left(\begin{pmatrix}2J&0\\ 0&0\end{pmatrix},-3J\right),\ \ \ e_{3}:=\frac{1}{3}\left(\begin{pmatrix}2K&0\\ 0&0\end{pmatrix},-3K\right),
e4\displaystyle e_{4} :=53((002000200),0),e5:=53((002i0002i00),0),\displaystyle:=\frac{\sqrt{5}}{3}\left(\begin{pmatrix}0&0&\sqrt{2}\\ 0&0&0\\ -\sqrt{2}&0&0\end{pmatrix},0\right),\ \ \ e_{5}:=\frac{\sqrt{5}}{3}\left(\begin{pmatrix}0&0&\sqrt{2}i\\ 0&0&0\\ \sqrt{2}i&0&0\end{pmatrix},0\right),
e6\displaystyle e_{6} :=53((000002020),0),e7:=53((000002i02i0),0).\displaystyle:=\frac{\sqrt{5}}{3}\left(\begin{pmatrix}0&0&0\\ 0&0&\sqrt{2}\\ 0&-\sqrt{2}&0\end{pmatrix},0\right),\ \ \ e_{7}:=\frac{\sqrt{5}}{3}\left(\begin{pmatrix}0&0&0\\ 0&0&\sqrt{2}i\\ 0&\sqrt{2}i&0\end{pmatrix},0\right).

This basis is orthonormal with respect to the metric g=340Bg=-\frac{3}{40}B. We use the shorthand ei1i2ine^{i_{1}i_{2}\dots i_{n}} to denote the nn-form ei1ei2eine^{i_{1}}\wedge e^{i_{2}}\wedge\dots\wedge e^{i_{n}}. The nearly G2\mathrm{G}_{2} structure φ\varphi is given by

φ\displaystyle\varphi =e123+e145e167+e246+e257+e347e356.\displaystyle=e^{123}+e^{145}-e^{167}+e^{246}+e^{257}+e^{347}-e^{356}.

As an SU(2)×U(1)\textup{SU(2)}\times\textup{U(1)} representation, 𝔪S2WWF(3)WF(3)\mathfrak{m}_{\mathbb{C}}\cong S^{2}W\oplus WF(3)\oplus WF(-3) where

S2W\displaystyle S^{2}W =Span{e1,e2,e3},WF(3)=Span{e4ie5,e6ie7},WF(3)=Span{e4+ie5,e6+ie7}.\displaystyle=\textup{Span}\{e^{1},e^{2},e^{3}\},\ \ \ WF(3)=\textup{Span}\{e^{4}-ie^{5},e^{6}-ie^{7}\},\ \ \ WF(-3)=\textup{Span}\{e^{4}+ie^{5},e^{6}+ie^{7}\}.

By Theorem 4.3, the space of first order deformations is given by {ξΩ273|dξ=4ξ}\{\xi\in\Omega^{3}_{27}\ |\ d\xi=-4*\xi\}. In this example, it was found to be isomorphic to 𝔰𝔲(3)\mathfrak{su}(3). As an SU(2)×U(1)\mathrm{SU}(2)\times\mathrm{U}(1) representation, 𝔰𝔲(3)\mathfrak{su}(3) is isomorphic to the span of {C,e1,,e7}\{C,e_{1},\dots,e_{7}\}. The SU(2)×U(1)\mathrm{SU}(2)\times\mathrm{U}(1)-invariant homomorphism from 𝔰𝔲(3)\mathfrak{su}(3) to Ω273(X1,1)\Omega^{3}_{27}(X_{1,1}) is given by Span{A}\text{Span}\{A\} where

A(C)\displaystyle A(C) =φ7e123,A(e1)=53(e145+e167),\displaystyle=\varphi-7e^{123},\ \ \ A(e_{1})=\frac{5}{3}(e^{145}+e^{167}),
A(e2)\displaystyle A(e_{2}) =53(e245+e267),A(e3)=53(e345+e367),\displaystyle=\frac{5}{3}(e^{245}+e^{267}),\ \ \ A(e_{3})=\frac{5}{3}(e^{345}+e^{367}),
A(e4)\displaystyle A(e_{4}) =59(3e467+e137+e126+e234),A(e5)=59(3e567+e235e136+e127),\displaystyle=\frac{5}{9}(3e^{467}+e^{137}+e^{126}+e^{234}),\ \ \ A(e_{5})=\frac{5}{9}(3e^{567}+e^{235}-e^{136}+e^{127}),
A(e6)\displaystyle A(e_{6}) =59(3e456e236e135+e124),A(e7)=59(3e457e237+e125+e134).\displaystyle=\frac{5}{9}(3e^{456}-e^{236}-e^{135}+e^{124}),\ \ \ A(e_{7})=\frac{5}{9}(3e^{457}-e^{237}+e^{125}+e^{134}).

Let us fix an α𝔰𝔲(3)\alpha\in\mathfrak{su}(3). The adjoint action of h=(h1,h2)SU(3)×SU(2)h=(h_{1},h_{2})\in\mathrm{SU}(3)\times\mathrm{SU}(2) is given by

h1αh\displaystyle h^{-1}\alpha h =h11αh1=(iv1x1+ix2x3+ix4x1+ix2iv2x5+ix6x3+ix4x5+ix6i(v1+v2))\displaystyle=h_{1}^{-1}\alpha h_{1}=\begin{pmatrix}iv_{1}&x_{1}+ix_{2}&x_{3}+ix_{4}\\ -x_{1}+ix_{2}&iv_{2}&x_{5}+ix_{6}\\ -x_{3}+ix_{4}&-x_{5}+ix_{6}&-i(v_{1}+v_{2})\end{pmatrix}

where v1,v2,x1,x2,x3,x4,x5,x6v_{1},v_{2},x_{1},x_{2},x_{3},x_{4},x_{5},x_{6} are functions on X1,1X_{1,1}.

The infinitesimal deformation ξα\xi_{\alpha} associated to α\alpha such that dξα=4ξαd\xi_{\alpha}=-4*\xi_{\alpha} is given by

ξα=v1+v22A(C)+v1v22A(e1)+i=16xiA(ei+1).\displaystyle\xi_{\alpha}=\frac{v_{1}+v_{2}}{2}A(C)+\frac{v_{1}-v_{2}}{2}A(e_{1})+\sum_{i=1}^{6}x_{i}A(e_{i+1}).

We can now compute the 44-form ηα\eta_{\alpha} by using the relation dξα=4ηα=4ξαd\xi_{\alpha}=4\eta_{\alpha}=-4*\xi_{\alpha}. In order to show that the infinitesimal deformation (ξα,ηα)(\xi_{\alpha},\eta_{\alpha}) associated to α\alpha is obstructed to second order, we need to compute the quadratic term Q3(ηα)Q_{3}(\eta_{\alpha}) as discussed in equation (4.13) and find an element β𝔰𝔲(3)\beta\in\mathfrak{su}(3) for which the L2L^{2}-inner product is non-zero.

To compute Q3(ηα)Q_{3}(\eta_{\alpha}), one can use the algorithm for stable 44-forms on manifolds with G2\mathrm{G}_{2} structures as discussed in [Hit01]. Using the fact that ξα=ηα\xi_{\alpha}=-*\eta_{\alpha}, one can easily show that for some non-zero constant c1c_{1}, Q3(ηα)=c1Q4(ξα)Q_{3}(\eta_{\alpha})=c_{1}*Q_{4}(\xi_{\alpha}) where Q4(ξα)Q_{4}(\xi_{\alpha}) is the quadratic term associated to ξα\xi_{\alpha}. Thus, we will instead compute Q4(ξα)Q_{4}(\xi_{\alpha}) and show that the inner product Q4(ξα),ξαL20\langle*Q_{4}(\xi_{\alpha}),\xi_{\alpha}\rangle_{L^{2}}\neq 0 to prove obstructedness.

Consider φt=φ+tξα\varphi_{t}=\varphi+t\xi_{\alpha} to be a positive 33-form for small tt. We will denote the metric and the volume form induced by φt\varphi_{t} by gtg_{t} and volt\operatorname{vol}_{t} respectively. We have a Taylor series expansion

gt\displaystyle g_{t} =g0+tg1+t2g2+O(t3)).\displaystyle=g_{0}+tg_{1}+t^{2}g_{2}+O(t^{3})).

Then one can define the symmetric bi-linear form BtB_{t} by

(Bt)ij\displaystyle(B_{t})_{ij} =((eiφt)(ejφt)φt)(e1,,e7).\displaystyle=((e_{i}\lrcorner\varphi_{t})\wedge(e_{j}\lrcorner\varphi_{t})\wedge\varphi_{t})(e_{1},\dots,e_{7}).

The zero order term of BtB_{t}, denoted by B0B_{0} is given by (B0)ij=((eiφ)(ejφ)φ)(e1,,e7)=δij(B_{0})_{ij}=((e_{i}\lrcorner\varphi)\wedge(e_{j}\lrcorner\varphi)\wedge\varphi)(e_{1},\dots,e_{7})=\delta_{ij}. Similarly, one can compute the linear term (B1)ij=3((eiφ)(ejφ)ξα)(e1,,e7)(B_{1})_{ij}=3((e_{i}\lrcorner\varphi)\wedge(e_{j}\lrcorner\varphi)\wedge\xi_{\alpha})(e_{1},\dots,e_{7}) and the quadratic term (B2)ij=3((eiξα)(ejξα)φ)(e1,,e7)(B_{2})_{ij}=3((e_{i}\lrcorner\xi_{\alpha})\wedge(e_{j}\lrcorner\xi_{\alpha})\wedge\varphi)(e_{1},\dots,e_{7}). The metric is then defined using the relation (see for example, [Kar09])

(Bt)ij=6(gt)ijdetgt.\displaystyle(B_{t})_{ij}=6(g_{t})_{ij}\sqrt{\det g_{t}}.

The linear term in volt\operatorname{vol}_{t} is proportional to φηα+ψξα\varphi\wedge\eta_{\alpha}+\psi\wedge\xi_{\alpha} and thus vanishes since (ξα,ηα)Ω273×Ω274(\xi_{\alpha},\eta_{\alpha})\in\Omega^{3}_{27}\times\Omega^{4}_{27}. Using the above formula we get that

volt\displaystyle\operatorname{vol}_{t} =detgt=1+At2+O(t3),\displaystyle=\sqrt{\det g_{t}}=1+At^{2}+O(t^{3}),

where AA is a quadratic polynomial in v1,v2v_{1},v_{2} and xi,i=1..6x_{i},i=1..6. Using the Taylor series expansion of gtg_{t} and detgt\sqrt{\det g_{t}}, we can compute the Taylor series expansion of the Hodge star associated to φt\varphi_{t}, t=0+t1+t22+O(t3)*_{t}=*_{0}+t*_{1}+t^{2}*_{2}+O(t^{3}). The Hodge star operator t*_{t} can be computed using the formula

t(ei1i2ik)\displaystyle*_{t}(e^{i_{1}i_{2}\dots i_{k}}) =volt(7k)!gti1j1gtikjkϵj1j7ejk+1j7.\displaystyle=\frac{\operatorname{vol}_{t}}{(7-k)!}g_{t}^{i_{1}j_{1}}\dots g_{t}^{i_{k}j_{k}}\epsilon_{j_{1}\dots j_{7}}e^{j_{k+1}\dots j_{7}}.

The quadratic term Q4(ξα)Q_{4}(\xi_{\alpha}) is then given by

Q4(ξα)\displaystyle Q_{4}(\xi_{\alpha}) =2φ+1ξα.\displaystyle=*_{2}\varphi+*_{1}\xi_{\alpha}.

In the present case, for a general element α𝔰𝔲(3)\alpha\in\mathfrak{su}(3), the quadratic term turns out to be very complicated and is not very enlightening. We define the cubic polynomial on X1,1X_{1,1} by

fα([h])\displaystyle f_{\alpha}([h]) =Q4(ξα),ξαL2.\displaystyle=\langle*Q_{4}(\xi_{\alpha}),\xi_{\alpha}\rangle_{L^{2}}.

Note that fαf_{\alpha} is cubic in α\alpha since Q4(ξα)Q_{4}(\xi_{\alpha}) and ξα\xi_{\alpha} are quadratic and linear in α\alpha respectively. This cubic polynomial can be lifted to a polynomial PP on the Lie group SU(3)×SU(2)\mathrm{SU}(3)\times\mathrm{SU}(2) by

fα([h])\displaystyle f_{\alpha}([h]) =P(h1αh).\displaystyle=P(h^{-1}\alpha h).

This lift enables us to calculate the average of PP on SU(3)×SU(2)\mathrm{SU}(3)\times\mathrm{SU}(2) by using the Peter–Weyl theorem. To express the polynomial PP in a compact form, we will set z1=x2+ix1,z2=x4ix3,z3=x6+ix5z_{1}=x_{2}+ix_{1},z_{2}=x_{4}-ix_{3},z_{3}=x_{6}+ix_{5}. Then the cubic polynomial PP is given by

P(h1αh)=976(v12v2+v22v1)+259Re(z1z2z3)296(v13+v23)+53(v1+v2)|z1|2+3718(v1|z3|3+v2|z2|2)+319(v1|z2|3+v2|z3|2)\displaystyle\begin{split}P(h^{-1}\alpha h)=&-\frac{97}{6}(v_{1}^{2}v_{2}+v_{2}^{2}v_{1})+\frac{25}{9}\mathrm{Re}(z_{1}z_{2}z_{3})-\frac{29}{6}(v_{1}^{3}+v_{2}^{3})+\frac{5}{3}(v_{1}+v_{2})|z_{1}|^{2}\\ &+\frac{37}{18}(v_{1}|z_{3}|^{3}+v_{2}|z_{2}|^{2})+\frac{31}{9}(v_{1}|z_{2}|^{3}+v_{2}|z_{3}|^{2})\end{split} (5.1)

The next step in proving obstructedness is to show that the average value of PP on SU(3)×SU(2)\mathrm{SU}(3)\times\mathrm{SU}(2) is non-zero. For this, we appeal to the Peter–Weyl theorem. The Peter–Weyl theorem states that for any compact Lie group GG, we have

L2(G)=VγGirrHom(Vγ,G)Vγ\displaystyle L^{2}(G)=\underset{V_{\gamma}\in G_{irr}}{\bigoplus}\mathrm{Hom}(V_{\gamma},G)\otimes V_{\gamma}

where GirrG_{irr} denotes the set of all non-isomorphic irreducible representations of GG.

The cubic polynomial PP lies in the SU(3)×SU(2)\mathrm{SU}(3)\times\mathrm{SU}(2) representation Sym3𝔰𝔲(3)\textup{Sym}^{3}\mathfrak{su}(3). The average value of the function P(g1ξg)P(g^{-1}\xi g) on SU(3)×SU(2)\textup{SU}(3)\times\mathrm{SU}(2) is the same as the average value of R(h1αh)R(h^{-1}\alpha h) where RR is the projection of PP to the invariant polynomials. This is because (PR)(h1αh)(P-R)(h^{-1}\alpha h) lies in the non-trivial part of the Peter–Weyl decomposition and has an average value of zero. The unique trivial sub-representation of Sym3𝔰𝔲(3)\textup{Sym}^{3}\mathfrak{su}(3) is generated by the determinant polynomial ideti\det on 𝔰𝔲(3)\mathfrak{su}(3) which is given by

idet(g1αg)=\displaystyle i\det(g^{-1}\alpha g)= (v1v22+v2v12)+(v1+v2)|z1|2(v1|z3|2+v2|z2|2)+2Re(z1z2z3).\displaystyle-(v_{1}v_{2}^{2}+v_{2}v_{1}^{2})+(v_{1}+v_{2})|z_{1}|^{2}-(v_{1}|z_{3}|^{2}+v_{2}|z_{2}|^{2})+2\mathrm{Re}(z_{1}z_{2}z_{3}).

The average value of the polynomial PP can be computed by computing the inner product of PP with ideti\det. On 𝔰𝔲(3)\mathfrak{su}(3), since the Killing form BB is non-degenarate, g=112Bg=-\frac{1}{12}B defines an inner product on 𝔰𝔲(3)\mathfrak{su}(3). The inner product gg induces an inner product on Sym3𝔰𝔲(3)\textup{Sym}^{3}\mathfrak{su}(3) in the natural way. All the computations that follow are done using gg.

If EijE_{ij} denotes the matrix with 11 as the (i,j)(i,j)-th entry and zero elsewhere, then the subspace of 𝔰𝔲(3)\mathfrak{su}(3) generated by {EijEji+i(Eij+Eji)i,j=1,2,3,ij}\{E_{ij}-E_{ji}+i(E_{ij}+E_{ji})\mid i,j=1,2,3,i\neq j\} is orthogonal to Span{E11iE33,E22iE33}\mathrm{Span}\{E_{11}-iE_{33},E_{22}-iE_{33}\}. Moreover EijEji+i(Eij+Eji),i,j=1,2,3,ijE_{ij}-E_{ji}+i(E_{ij}+E_{ji}),\ i,j=1,2,3,i\neq j are also orthogonal to each other. Thus the only non-trivial terms occurring in the inner product of PP and ideti\det are,

v12v2+v22v12=13,Re(z1z2z3)2=23,v13+v23,v12v2+v22v1=14,\displaystyle\|v_{1}^{2}v_{2}+v_{2}^{2}v_{1}\|^{2}=\frac{1}{3},\ \ \ \ \ \|\mathrm{Re}(z_{1}z_{2}z_{3})\|^{2}=\frac{2}{3},\ \ \ \ \ \langle v_{1}^{3}+v_{2}^{3},v_{1}^{2}v_{2}+v_{2}^{2}v_{1}\rangle=-\frac{1}{4},
(v1+v2)|z1|22=1,v1|z3|2+v2|z2|22=43,v1|z2|2+v2|z3|2,v1|z3|2+v2|z2|2=13.\displaystyle\|(v_{1}+v_{2})|z_{1}|^{2}\|^{2}=1,\ \ \ \|v_{1}|z_{3}|^{2}+v_{2}|z_{2}|^{2}\|^{2}=\frac{4}{3},\ \ \ \langle v_{1}|z_{2}|^{2}+v_{2}|z_{3}|^{2},v_{1}|z_{3}|^{2}+v_{2}|z_{2}|^{2}\rangle=-\frac{1}{3}.

From (5.1) and the above computations we have that

P,idet\displaystyle\langle P,i\det\rangle =976(13)+509(23)+296(14)+53(1)3718(43)319(13)=191240.\displaystyle=\frac{97}{6}\left(\frac{1}{3}\right)+\frac{50}{9}\left(\frac{2}{3}\right)+\frac{29}{6}\left(-\frac{1}{4}\right)+\frac{5}{3}(1)-\frac{37}{18}\left(\frac{4}{3}\right)-\frac{31}{9}\left(-\frac{1}{3}\right)=\frac{191}{24}\neq 0.

Thus we get the following theorem.

Theorem 5.1.

The infinitesimal deformations of the homogeneous nearly G2\mathrm{G}_{2} structure on the Aloff–Wallach space X1,1SU(3)×SU(2)SU(2)×U(1)X_{1,1}\cong\frac{\mathrm{SU}(3)\times\mathrm{SU}(2)}{\mathrm{SU}(2)\times\mathrm{U}(1)} are all obstructed.

References

  • [AS12] Bogdan Alexandrov and Uwe Semmelmann “Deformations of nearly parallel G2{\rm G}_{2}-structures” In Asian J. Math. 16.4, 2012, pp. 713–744 DOI: 10.4310/AJM.2012.v16.n4.a6
  • [Bes87] Arthur L. Besse “Einstein manifolds” 10, Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)] Springer-Verlag, Berlin, 1987, pp. xii+510 DOI: 10.1007/978-3-540-74311-8
  • [BFGK91] Helga Baum, Thomas Friedrich, Ralf Grunewald and Ines Kath “Twistors and Killing spinors on Riemannian manifolds” With German, French and Russian summaries 124, Teubner-Texte zur Mathematik [Teubner Texts in Mathematics] B. G. Teubner Verlagsgesellschaft mbH, Stuttgart, 1991, pp. 180
  • [BO19] Gavin Ball and Goncalo Oliveira “Gauge theory on Aloff-Wallach spaces” In Geom. Topol. 23.2, 2019, pp. 685–743 DOI: 10.2140/gt.2019.23.685
  • [Bry06] Robert L. Bryant “Some remarks on G2G_{2}-structures” In Proceedings of Gökova Geometry-Topology Conference 2005 Gökova Geometry/Topology Conference (GGT), Gökova, 2006, pp. 75–109
  • [Dwi19] Shubham Dwivedi “Minimal hypersurfaces in nearly G2\rm G_{2} manifolds” In J. Geom. Phys. 135, 2019, pp. 253–264 DOI: 10.1016/j.geomphys.2018.10.007
  • [FKMS97] Thomas Friedrich, Ines Kath, Andrei Moroianu and Uwe Semmelmann “On nearly parallel G2\mathrm{G}_{2}-structures” In Journal of Geometry and Physics 23.3, 1997, pp. 259–286 DOI: https://doi.org/10.1016/S0393-0440(97)80004-6
  • [Fos17] Lorenzo Foscolo “Deformation theory of nearly Kähler manifolds” In J. Lond. Math. Soc. (2) 95.2, 2017, pp. 586–612 DOI: 10.1112/jlms.12033
  • [Fri80] Thomas Friedrich “Der erste Eigenwert des Dirac-Operators einer kompakten, Riemannschen Mannigfaltigkeit nichtnegativer Skalarkrümmung” In Math. Nachr. 97, 1980, pp. 117–146 DOI: 10.1002/mana.19800970111
  • [Hit01] Nigel Hitchin “Stable forms and special metrics” In Global differential geometry: the mathematical legacy of Alfred Gray (Bilbao, 2000) 288, Contemp. Math. Amer. Math. Soc., Providence, RI, 2001, pp. 70–89 DOI: 10.1090/conm/288/04818
  • [Joy00] Dominic D. Joyce “Compact manifolds with special holonomy”, Oxford Mathematical Monographs Oxford University Press, Oxford, 2000, pp. xii+436
  • [Kar05] Spiro Karigiannis “Deformations of G2G_{2} and Spin(7){\rm Spin}(7) structures” In Canad. J. Math. 57.5, 2005, pp. 1012–1055 DOI: 10.4153/CJM-2005-039-x
  • [Kar06] Spiro Karigiannis “Some Notes on G2\mathrm{G}_{2} and Spin(7)\mathrm{Spin}(7) Geometry”, 2006 arXiv:math/0608618 [math.DG]
  • [Kar09] Spiro Karigiannis “Flows of G2G_{2}-structures. I” In Q. J. Math. 60.4, 2009, pp. 487–522 DOI: 10.1093/qmath/han020
  • [KL20] Spiro Karigiannis and Jason Lotay “Deformation theory of G2\mathrm{G}_{2} conifolds” In to appear, Communications in Analysis and Geometry, 2020 arXiv:1212.6457 [math.DG]
  • [Koi82] Norihito Koiso “Rigidity and infinitesimal deformability of Einstein metrics” In Osaka Math. J. 19.3, 1982, pp. 643–668
  • [Leh20] Fabian Lehmann “Deformations of asymptotically conical Spin(7)\mathrm{Spin}(7)-manifolds” In in preparation, 2020
  • [MNS08] Andrei Moroianu, Paul-Andi Nagy and Uwe Semmelmann “Deformations of nearly Kähler structures” In Pacific J. Math. 235.1, 2008, pp. 57–72 DOI: 10.2140/pjm.2008.235.57
  • [Nor08] Johannes Nordström “Deformations and gluing of asymptotically cylindrical manifolds with exceptional holonomy”, 2008 URL: https://people.bath.ac.uk/jlpn20/thesis_final_twoside.pdf
  • [NS20] Paul-Andi Nagy and Uwe Semmelmann “Deformations of nearly G2G_{2}-structures” In arXiv e-prints, 2020 arXiv:2007.01657 [math.DG]

Department of Pure Mathematics, University of Waterloo, Waterloo, ON, N2L 3G1, Canada.
e-mail address (SD): [email protected]
e-mail address (RS): [email protected]