[datatype=bibtex] \map \step[fieldsource=pmid, fieldtarget=pubmed]
A gradient flow of Spin(7)-structures
Abstract
We formulate and study the negative gradient flow of an energy functional of Spin(7)-structures on compact -manifolds. The energy functional is the -norm of the torsion of the Spin(7)-structure. Our main result is the short-time existence and uniqueness of solutions to the flow. We also explain how this negative gradient flow is the most general flow of Spin(7)-structures. We also study solitons of the flow and prove a non-existence result for compact expanding solitons.
1. Introduction
Given an -dimensional smooth manifold , a Spin(7)-structure on is a reduction of the structure group of the frame bundle to the Lie group Spin(7) . While many -manifolds admit Spin(7)-structures, the existence of torsion-free Spin(7)-structures is a difficult problem. Given the success of geometric flows in proving the existence of special geometric structures on manifolds it is natural to attempt to tackle such questions using a suitable geometric flow of Spin(7)-structures.
Unlike the case of geometric flows of -structures, there have been fewer studies of flows of Spin(7)-structures. A flow of isometric/harmonic Spin(7)-structures was introduced and studied in detail in [DLE24]. The authors proved various analytic and geometric results for the flow. This flow can be seen as an instance of the more general theory of harmonic -structures which has been studied in a fairly detailed manner in [LS23] and [FLME22]. One disadvantage of the isometric flow is that it runs in an isometry class of a given Spin(7)-structure and hence the metric remains fixed throughout the flow. We remedy this here by studying the most general flow of Spin(7)-structures (in particular, the metric is also changing) which include the isometric flow as a special case. On a fixed oriented spin Riemannian -manifold, a Spin(7)-structure is equivalent (up to sign) to a unit spinor field. Using a spinorial approach, Ammann–Weiss–Witt [AWW16] also studied the negative gradient flow of a Dirichlet energy, thought of as a function of a unit spinor field. They proved general short-time existence and uniqueness. In contrast to their work, our approach is more direct and uses the Riemannian geometric properties of a Spin(7)-structure with the flow equation and related quantities all given in terms of the curvature of the underlying metric and the torsion of the Spin(7)-structure. We also describe all possible second order quasilinear flows of Spin(7)-structures. A study of some natural functionals of Spin(7)-structures and the corresponding Euler-Lagrange equations has also been done recently by Krasnov [Kra24].
Our point of view of studying general flows of Spin(7)-structures is that of [DGK23] where the authors study the most general second order flows of -structures by classifying all linearly independent second order (in the -structure) tensors coming from the Riemann curvature tensor and the covariant derivative of the torsion which can be made into a -form.
We now describe the energy functional and its negative gradient flow. More details are given in § 3.
Definition 1.1.
Let be a compact manifold which admits Spin(7)-structures. The energy functional on the set of Spin(7)-structures on is
(1.1) |
where is the torsion of the Spin(7)-structure and the norm and the volume form are with respect to the metric induced by . The critical points of the functional are torsion-free Spin(7)-structures which are also absolute minimizers.
We compute the first variation of in Lemma 3.3 which also gives the negative gradient of the energy functional. We describe the negative gradient flow below. See § 2 and § 3 for the definition of the terms and more details.
Definition 1.2 (Negative gradient flow).
Let be a compact manifold with a Spin(7)-structure. Let . The negative gradient flow of the energy functional is the flow of Spin(7)-structures which is the following initial value problem
(GF) |
where is the Ricci curvature of the underlying metric, is the -dimensional component of the torsion tensor and is the operator described in (2.13).
Given any geometric flow, one of the main questions is whether there exist a solution to the flow for a short time and if any two solutions starting with the same initial conditions are unique. The main theorem of the paper is as follows.
Theorem 4.11. Let be a compact -manifold with a Spin(7)-structure and consider the negative gradient flow of the natural energy functional in (1.1). This is the flow (GF). Then there exists and a unique smooth solution of (GF) for with .
We prove the main theorem by explicitly calculating the principal symbols of the operators involved in the flow and then using a modified DeTurck’s trick.
As discussed at the end of § 2, the most general flow of Spin(7)-structures which is second order can be written as
(1.3) |
where and l.o.t are the terms which are lower order in . The situation in the Spin(7)-case is a bit simpler than the corresponding case of the -structures where there are more terms which could appear in a flow of -structures [DGK23, §1]. This is because of the fact that for , the subbundle of admissible 4-forms in has codimension and thus any variation of a Spin(7)-structure is given only by -forms in . Consequently, many naturally occurring tensors, for example the -dimensional Weyl curvature tensor cannot be a candidate for a flow of Spin(7)-structures (see § 2 for details) even though they can be made into a -form.
The methods in § 4 can be used to establish short-time existence and uniqueness results for the general flow in (1.3) subjected to conditions on the coefficients which will be sufficient conditions to apply the modified DeTurck’s trick described in § 4.4. The advantage of looking at a specific flow in the family (1.3), which for us is (GF) which corresponds to is that it also provides us with the lower order terms which are important when one wants to study further analytic properties of the flow. An example of the importance of lower order terms is our result on the non-existence of compact expanding solitons for the flow in § 5.
Solitons of a geometric flow are special solutions which move only by scalings and diffeomorphisms. These are also called self-similar solutions. A soliton can be viewed as a generalized fixed point of the flow modulo scalings and diffeomorphisms. These special solutions are important because they serve as singularity models for the flow. We study solitons of the flow (GF). We derive equations for the underlying metric and the divergence of the torsion of a soliton solution. We expect that if we have a Cheeger–Gromov–Hamilton type compactness theorem for the solutions of the flow then the solitons will play a role in understanding the behaviour of the flow near a singularity.
The paper is organized as follows. We start by reviewing the notion of Spin(7)-structures, the decomposition of space of forms and the torsion tensor in § 2. The first variation of the energy functional is computed in § 3. We give a brief review of elliptic and parabolic differential operators in § 4.1 which is followed by a modified DeTurck’s trick in § 4.4. The main theorem, Theorem 4.11, is proved in § 4.5. Finally, in § 5 we study solitons of the flow and prove Proposition 5.2 which is a non-existence result for compact expanding solitons of the flow.
Throughout the paper, we compute in a local orthonormal frame, so all indices are subscripts and any repeated indices are summed over all values from to . Our convention for labelling the Riemann curvature tensor is
in terms of coordinate vector fields and as a result, the Ricci tensor is . We also have the Ricci identity which, for instance, for a -tensor reads as
(1.4) |
Acknowledgements. We are grateful to Panagiotis Gianniotis and Spiro Karigiannis for various discussions on this project in particular and on flows of special geometric structures in general. We would like to thank Thomas Walpuski for discussions on the torsion decomposition of Spin(7)-structures and Ragini Singhal for explaining the representation theoretic aspects of Spin(7)-structures. A part of this work was done when the author is a Simons-CRM Scholar-in-Residence at the CRM, Montreal. We would like to thank the Simons Foundation and the CRM for funding and CRM for providing a wonderful environment to work in.
2. Preliminaries on Spin(7)-structures
In this section, we briefly review the notion of a Spin(7)-structure on an -dimensional manifold and the associated decomposition of differential forms. We also discuss the torsion tensor of a Spin(7)-structure. More details can be found in [Joy00, Chapter 10] and [Kar08].
A Spin(7)-structure on is a -form on . The existence of such a structure is a topological condition. The space of -forms which determine a Spin(7)-structure on is a subbundle of , called the bundle of admissible -forms. This is not a vector subbundle and it is not even an open subbundle, unlike the case for -structures. For , the subbundle is of codimension in .
A Spin(7)-structure determines a Riemannian metric and an orientation on in a nonlinear way which can be described in local coordinates as follows. Let and extend a non-zero tangent vector to a local frame near . Define
Then the metric induced by is given by
(2.1) |
The metric and the orientation determine a Hodge star operator , and the -form is self-dual, i.e., .
Definition 2.1.
Let be the Levi-Civita connection of the metric . The pair is a Spin(7)-manifold if . This is a non-linear partial differential equation for , since depends on , which in turn depends non-linearly on . A Spin(7)-manifold has Riemannian holonomy contained in the subgroup . Such a parallel Spin(7)-structure is also called torsion free.
2.1 Decomposition of the space of forms
The existence of a Spin(7)-structure induces a decomposition of the space of differential forms on into irreducible Spin(7)-representations. We have the following orthogonal decomposition, with respect to :
where has pointwise dimension . Explicitly, and are described as follows:
(2.2) | ||||
(2.3) |
and
(2.4) | ||||
(2.5) |
For our computations, it is useful to describe the spaces of forms in local coordinates. For ,
(2.6) | ||||
(2.7) |
Remark 2.2.
For ,
(2.8) | ||||
(2.9) |
If and are the projection operators on , it follows from (2.6) and (2.7) that
(2.10) | ||||
(2.11) |
We will be using these equations throughout the paper. Finally, for ,
(2.12) |
which can be used to show that is the Lie algebra of Spin(7), with the commutator of matrices
To describe in local coordinates, we use the operator which was first described in [DGK21] for the case and in [DLE24] for the Spin(7) (see also [FLME22, eq. (1.14)] for the general case of -structures). Given , define
(2.13) |
and hence
(2.14) |
Recall that , and further splits orthogonally into (2.2) and (2.3), so
(2.15) |
With respect to this splitting, we can write where is a symmetric traceless -tensor. The diamond contraction (2.14) defines a linear map , from to . We record the following properties of the operator whose proof can be found in [Kar08] or [DLE24].
Proposition 2.3.
Let be a manifold with a Spin(7)-structure. Then
-
1.
The Hodge star of is
(2.16) -
2.
If and denote the traceless symmetric parts of and respectively and denote their component. Then
(2.17) ∎
More properties of the operator for general -structures have been proved in [FLME22, Lemma 1.4].
The following proposition was originally proved in [Kar08, Prop. 2.3].
Proposition 2.4.
The kernel of the map is isomorphic to the subspace . The remaining three summands and are mapped isomorphically onto the subspaces and respectively.
Proof.
Using (2.17) with gives
and hence . If then since all the coefficients in are positive hence the map is injective on and by dimension count, the map is a linear isomorphism. ∎
To understand , we need another characterization of the space of -forms using the Spin(7)-structure. Following [Kar08], we adopt the following:
Definition 2.5.
On , define a Spin(7)-equivariant linear operator on as follows. Let and let . Then
(2.18) |
Proposition 2.6.
The spaces , and are all eigenspaces of with distinct eigenvalues:
(2.19) |
Moreover, the decomposition of into self-dual and anti-self-dual parts is
(2.20) |
Before we discuss the torsion of a Spin(7)-structure, we note some contraction identities involving the -form . In local coordinates , the -form is
where is totally skew-symmetric. We have the following identities
(2.21) | ||||
(2.22) | ||||
(2.23) | ||||
(2.24) |
We also have contraction identities involving and
(2.25) | ||||
(2.26) | ||||
(2.27) |
We now describe the torsion of a Spin(7)-structure. Given , we know from [Kar08, Lemma 2.10] that lies in the subbundle .
Definition 2.7.
The torsion tensor of a Spin(7)-structure is the element of defined by expressing Since , by Proposition 2.4, can be written as
(2.28) |
where , for each fixed . This defines the torsion tensor of a Spin(7)-structure, which is an element of .
In terms of , the torsion is given by
(2.29) |
since is an element of . Thus we have the following result, originally due to Fernández [Fer86].
Theorem 2.8.
[Fer86] The Spin(7)-structure is torsion free if and only if . Since , this is equivalent to .
Remark 2.9.
The notation should not be confused with taking two covariant derivatives of . The torsion tensor T is an element of and thus, for each fixed index , .
Finally, an important bit of information which we need about the torsion is that it satisfies a “Bianchi-type identity”. This was first proved by Karigiannis [Kar08, Theorem 4.2] using the diffeomorphism invariance of the torsion tensor and a different proof using the Ricci identity (1.4) was given in [DLE24, Theorem 3.9].
Theorem 2.10.
The torsion tensor satisfies the following “Bianchi-type” identity
(2.30) |
∎
Using the Riemannian Bianchi identity, we see that
and hence we have the fact that
(2.31) |
Using this and contracting (2.30) on and gives the expression for the Ricci curvature of a metric induced by a Spin(7)-structure. Precisely,
(2.32) |
which also proves that the metric of a torsion free Spin(7)-structure is Ricci-flat, a result originally due to Bonan. Taking the trace of (2.32) gives the expression of the scalar curvature
(2.33) | ||||
(2.34) |
Since , the component of torsion can be viewed as a vector field on and hence we want to look at alternate expression for the symmetric -tensor which will be used in § 3.1. We have the following linear algebra result whose proof can be found in [DGK23, Lemma 4.1].
Lemma 2.11.
Let and be finite-dimensional real vector spaces equipped with positive definite inner products, and suppose that
is an orthogonal direct sum of subspaces. Let and be linear maps. Suppose that for every , there exist both nonzero, such that for all and , we have
(2.35) |
Then in fact we have an isomorphism of with an orthogonal direct sum
(2.36) |
∎
Let and in the previous lemma. Define by
and by
Clearly . Consider
(2.37) |
where we have used (2.23) in the second equality. Hence (2.6) implies . We compute
Also,
and, using (2.6),
Thus, (i) and (ii) in (2.35) are satisfied with and respectively. We deduce from Lemma 2.11 that
(2.38) |
with being -dimensional. In particular, since , we see that
(2.39) |
and thus
(2.40) |
We will also need the expression for the Lie derivative of the Spin(7)-structure in § 4. Let . Using (2.28) and (2.14), we have
and since , where and denote the projection of the -form onto the and components respectively. Using the fact that , we get
(2.41) |
We record this observation in the following
Proposition 2.12.
A vector field is an infinitesimal symmetry of , i.e., if and only if
(2.42) |
∎
Remark 2.13.
Comparing with the case of manifolds with -structures, the most general flow of -structures which are second order quasilinear are described in [DGK23] and the authors prove the short-time existence and uniqueness theorem for a large class of flows. Such a family of flows was obtained by explicitly computing all independent second order differential invariants of -structures which are -forms. Since the Riemann curvature tensor and are the only two second order invariants of a -structure, decomposing them into irreducible -representations and then picking the linearly independent components (since and are related by the so called -Bianchi identity) which can be made into a -form provides a way to write the most general flow of -structures. See [DGK23, §4] for more details. The existence of a Spin(7)-structure induces a decomposition of tensor bundles into irreducible Spin(7)-representations and one can similarly decompose and into irreducible Spin(7)-representations and look at the linearly independent components which can be made into an admissible -form. We describe this briefly below without any proofs.
Let us denote the -dimensional irreducible Spin(7)-representation by . Recall that the space of curvature tensors on is the subspace of of elements satisfying the first Bianchi identity and hence we have the orthogonal decomposition
Since for a manifold with a Spin(7)-structure , we have (see § 2.1), we have
One can check the following decomposition of tensor bundles into irreducible Spin(7)-representations.
(2.43) | ||||
(2.44) | ||||
(2.45) | ||||
(2.46) |
and hence we have
(2.47) |
Since the space of curvature type tensors is orthogonal to the space of -forms , we get the following decomposition
Thus, in the presence of a Spin(7)-structure, the Weyl tensor further decomposes as . Like the -case there is again a , however unlike the -case, this cannot be used for a flow of Spin(7)-structures as the space of admissible -forms is (pointwise) -dimensional and is . Consequently, the only contribution from for a flow of Spin(7)-structures are the traceless Ricci curvature and the scalar curvature .
In a similar way we look at the possible contributions from for a flow of Spin(7)-structures. Since , hence at every point it lies in the representation . Thus,
and hence the components which can possibly contribute to a flow of Spin(7)-structures are and . The component from is the term which can be written in terms of the scalar curvature and lower order terms as demonstrated in (2.34). Similarly, the two components are a combination from and because and are related to each other by the Spin(7)-Bianchi identity (2.30). The contribution for the component from is the term (since in the Spin(7)-case ). One can do an explicit description of all the components above from and as in [DGK23]. However, since we are only interested in flows of Spin(7)-structures for which we have all the relevant components, namely, and , we do not pursue the former here and we are currently doing this in collaboration with R. Singhal.
3. A gradient flow of Spin(7)-structures
3.1 Derivation of the flow
In this section we define the flow of Spin(7)-structures studied in this paper and prove that it is a negative gradient flow.
Definition 3.1.
Let be a compact manifold. The energy functional on the set of Spin(7)-structures on is
(3.1) |
where is the torsion of and the norm and the volume form are with respect to the metric induced by .
The natural question here is whether, given a Spin(7)-structure , there is a “best” Spin(7)-structure. An obvious way to study this question is to consider the negative gradient flow of .
The most general flow of Spin(7)-structures [Kar08] is given by
(3.2) |
where with and . Thus, the most general flow of Spin(7)-structures is given by a time-dependent family of symmetric -tensors and a time-dependent family of -forms lying in the subspace . In this case the evolution of the metric, the inverse of the metric and the volume form are given by
(3.3) |
We start by considering the variation of the torsion with respect to variation of the Spin(7)-structures.
Proposition 3.2.
Let be the energy functional from (3.1). The following result motivates our formulation of the flow, cf. Definition 3.2.
Lemma 3.3.
Let be a smooth family of Spin(7)-structures and for some , . The gradient of the energy functional from (3.1) is given by
(3.5) |
Proof.
We calculate the first term on the right hand side of (3.6). Using (2.10) for the component and integration by parts, we have
where and we used the fact that is symmetric in while is skew-symmetric in the second equality, the fact that and (2.6) in the second last equality and integration by parts in the last equality. Integrating by parts and then using the Spin(7)-Bianchi identity (2.30) gives
from which we infer
(3.7) |
Remark 3.4.
A similar calculation for -structures is done in [FLME22, Prop. 1.44]. The terms coming from and are explicit for us as we are looking at a particular -structure.
We are interested in the negative gradient flow of and based on the above computations, we propose the following flow of Spin(7)-structures.
Definition 3.5.
Let be a compact manifold with a Spin(7)-structure. Let . A gradient flow of Spin(7)-structures is the following initial value problem
(GF) |
It follows from (3.3) and (2.33) that along (GF), the underlying metric and volume form evolve as
(3.10) | ||||
(3.11) | ||||
(3.12) |
We discuss the effect of scaling on tensors induced from a Spin(7)-structure. It follows from [DLE24, §4.3.1] that if is a Spin(7)-structure then so is , constant. In this case, and . Moreover,
The natural parabolic rescaling for a geometric evolution equation involves scaling the time variable t by , when the space variable scales by and hence, keeping in mind that taking with involves contraction on one index, we see that each term on the right hand side of (GF) indeed has the correct scaling. We record for future reference that for
(3.13) |
In particular, is a positively homogeneous functional and hence an application of Euler’s theorem for positively homogeneous functional shows that the critical points of are torsion-free Spin(7)-structures, which in fact, are absolute minimizers and thus the flow (GF) is capable of detecting torsion-free Spin(7) structures.
4. Short-time existence and uniqueness
In this section we establish short-time existence and uniqueness of the flow (GF) of Spin(7)-structures, using a modification of the DeTurck’s trick and the explicit computation of the principal symbols of the second order linear differential operators defining the flow. We first calculate the principal symbols of the highest order terms on the right hand side of (GF), i.e., and in § 4.2. We use these to show that (GF) is a weakly parabolic PDE and the failure of parabolicity is only due to the diffeomorphism invariance of the tensors involved. We use a modification of the DeTurck’s trick from the Ricci flow to prove the short-time existence and uniqueness of solution in § 4.3.
4.1 Differential operators, ellipticity, and parabolicity
We give a brief review of parabolic PDEs and the existence and uniqueness of solutions of such equations. Other sources for the discussion below are [CK04, §3.2], [AH11, §5.1], [Top06, §4] and [DGK23, §6.1] (for the case).
Let be vector bundles over a Riemannian manifold and let be a linear differential operator of order . We write , for every , in terms of local frames for and , as
(4.1) |
where for each , we write to denote the -th covariant derivative of , and . Here the index corresponds to a local frame for and the index corresponds to a local frame for .
For any such linear differential operator, we define its principal symbol so that for each and , the map
is the linear homomorphism
b | (4.2) | |||
The principal symbol satisfies the fundamental properties
whenever , are linear differential operators so that either or is well defined. We have the following
Definition 4.1.
A linear differential operator is called elliptic if for any , , , the principal symbol is a linear isomorphism.
Let be a vector bundle over with a fibre metric . Consider a second order linear differential operator . If there is a constant such that for any , and , we have
then is called strongly elliptic.
Definition 4.2.
Let , be vector bundles over , let be open, and let be a nonlinear differential operator. The operator is called elliptic at if the linearization
is an elliptic linear differential operator.
Similarly, if is a second order differential operator and is endowed with a bundle metric , we say that is strongly elliptic at if its linearization is a strongly elliptic linear differential operator.
A nonlinear evolution equation of the form , where , is called parabolic at if is strongly elliptic at .
The importance of the above definition is due to the following standard result.
Theorem 4.3.
Let be a Riemannian manifold, let be a vector bundle over endowed with a fibre metric , and let be open. Let be a second order quasilinear differential operator, which is strongly elliptic at . Then there exists and for any a unique , such that
(4.3) |
That is, a nonlinear evolution equation which is parabolic at has a unique short time smooth solution with initial condition . ∎
4.2 Principal Symbols
Let be a Spin(7)-structure and consider a variation with and
(4.4) |
where . Since the variation of the associated Riemannian metric is hence the linearization of the Christoffel symbols satisfies
(4.5) |
where is the Levi-Civita connection of the metric associated to the Spin(7)-structure .
It follows that the principal symbol of , for any non-zero , is
(4.6) |
Now we compute the linearization of :
The principal symbol of the differential operator is:
Namely
(4.7) | ||||
The linearization of the torsion is given by
(4.8) |
Hence, its principal symbol is given by
(4.9) |
We use the contraction identities in (2.21)-(2.24) and (2.10) to compute the principal symbol of ,
(4.10) |
(4.11) | ||||
(4.12) |
Recall that the Ricci curvature in terms of the torsion tensor is given by
From the computations in the previous section we obtain that the symbol of is given by
which simplifies to
(4.13) |
Similarly, the principal symbol of the scalar curvature is
(4.14) |
4.3 Failure of parabolicity of the flow
Consider the differential operator
Clearly, (GF) is . Since and are diffeomorphism invariant tensors, for the operator in (GF). For the purposes of short time existence of the flow we are only interested in the highest order term so we instead consider the operator (which we still )
(4.15) |
Moreover, using Proposition 2.4, we will view . Since the bundle metric on is uniformly equivalent to the natural inner product on , the operator is strongly elliptic if and only if there is a constant such that for any , , , and any , we have
(4.16) |
We will see below that the operator is, in fact, not elliptic (and hence (GF) not parabolic) but the failure of ellipticity is only due to the diffeomorphism invariance of the tensors in the definition of . As a result, we use a modified DeTurck’s trick to prove short-time existence in Theorem 4.11.
Proposition 4.4.
Let be as in (4.18). For any nonzero , we have
jk | (4.19) | |||
and is injective. Here denotes the identification of component of with an element in .
Proof.
Recall that is invariant under diffeomorphisms, that is,
for any diffeomorphism . It follows that for any vector field , we have
(4.20) |
Since is a first order linear differential operator on , whereas
is a priori a third order differential operator, it follows that
and hence
(4.21) |
Hence, by the injectivity of , the principal symbol of has
(4.22) |
that is due to diffeomorphism invariance, so is never an elliptic differential operator.
Remark 4.5.
We now show that the dimension of kernel of is at most . To do so, we introduce the following two operators. Consider
(4.23) | ||||||
We recall that the map is the symbol of the Bianchi map given by . The Ricci curvature lies in the kernel of the Bianchi map by the twice contracted Riemannian second Bianchi identity. The map is the map in (4.19) and is precisely the symbol of the operator . The reason we need the map is because while doing a modified version of the DeTurck’s trick, we need to add a term of the form to the right hand side of (GF), and the operator shows up in the part of , by equation (2.41).
It is convenient to define the operator by
(4.24) |
We can rewrite the components of in terms of the map .
Proposition 4.6.
Proof.
We note that (4.3) implies
(4.26) |
Further, the definitions of the maps , and equations (4.10), (2.39) give
(4.27) |
and hence . Substituting this in (4.26) gives the first equation in (4.25). We see from (4.3) that
(4.28) | ||||
Using the expression for in (4.28) gives | ||||
(4.29) | ||||
which on using the definition of the map from (4.23) give | ||||
which is the second equation in (4.25). ∎
Remark 4.7.
The operator plays a role similar to the role of the Bianchi operator in the analysis of the principal symbol of the Ricci tensor, for instance in the Ricci flow and for the analysis of the symbol of large class of flows of -structures as in [DGK23].
Our main goal is to prove that the dimension of kernel of the symbol of the operator is which will prove that the diffeomorphism invariance of and are the only reason for the failure of parabolicity of (GF) and hence we can use a modified version of the DeTurck’s trick. To calculate an upper bound on , we compute the adjoint of the map .
Lemma 4.8.
The map is injective. Consequently, .
Proof.
Let . Then and hence
and since , . This proves that is injective and hence . Since , we get that . ∎
We prove our main result in this section on the dim .
Proposition 4.9.
Proof.
We observe from Proposition 4.6, equation (4.25) that
(4.31) |
and hence if then and
(4.32) |
Multiplying both sides by gives
which on using the fact that is symmetric in and whereas and are skew-symmetric in and gives | ||||
and so . Therefore, from (4.32), we get and hence . Thus,
(4.33) |
Since by Lemma 4.8, we get
(4.34) |
which in combination with (4.22) gives . ∎
4.4 A modified DeTurck’s trick
In this section we prove that, given a background Spin(7)-structure , it is always possible to modify the operator from (4.15) to an operator which is strongly elliptic at , that is an operator whose symbol satisfies
for some constant . We do this by doing a modification of the DeTurck’s which was originally formulated to give an alternative proof of the short-time existence and uniqueness of solutions to the Ricci flow by DeTurck [DeT83].
Let be a fixed Spin(7)-structure on , for instance, one can take the initial Spin(7)-structure when running the flow. Motivated by the definition of the map in (4.24), we define the vector field on by
(4.35) |
where is the Riemannian metric induced by and is its Christoffel symbols. The vector field is the same vector field which is used in the DeTurck’s trick for the Ricci flow and the vector field is the extra term which we need for the DeTurck’s trick in the Spin(7)-case. We define the modified operator as
(4.36) |
where is the operator in (4.15). Using (2.41) we have
Since we only need the highest order terms for the purpose of short-time existence and uniqueness, we neglect the term above and consequently, the operator is given by
(4.37) |
We claim that the operator is elliptic. We need to calculate the principal symbol of the linearization of . It is well known (for example, see [CK04, §3.2]) that the linearization of , upto lower order terms, is
The factor of 2 is here because the variation of metric in the Spin(7) case is given by . Thus,
which on using (4.13) gives
(4.38) |
We now calculate the symbol of the remaining terms in (4.37). Since
so
and hence
(4.39) |
Using (4.41) we calculate
which on using Young’s inequality on the 3rd term gives | ||||
and we use and to get | ||||
(4.42) |
Proposition 4.10.
Let be an -manifold with a Spin(7)-structure and let be the operator
(4.43) |
where . Then is strongly elliptic at . ∎
4.5 Short-time existence and uniqueness
In this section, we prove the main theorem of the paper by using the modified DeTurck’s trick whose details were given in § 4.4.
Theorem 4.11.
Proof.
Let and define as in (4.35). Since the terms are at most first order in , we deduce from Proposition 4.10 that the linearization of the operator
is strongly elliptic. We obtain from standard parabolic theory (Theorem 4.3) that there is a unique smooth solution for of the initial value problem
Let be the one-parameter family of diffeomorphisms defined by
Since is compact, the family of diffeomorphisms exists by [CK04, Lemma 3.15] as long as the solutions exists. The family then satisfies
with and . This proves the short-time existence of solutions to (GF).
We now prove uniqueness, by using the uniqueness of solutions to the harmonic map heat flow. Suppose are solutions to (GF) with the same initial condition. Let be a one-parameter family of diffeomorphisms given as
(4.44) | ||||
Using (2.41), we see that the family solves
(4.45) | ||||
If we set by defining the diffeomorphisms as then it is known [CK04, §4.3] that is a solution to
(4.46) | ||||
which is the harmonic map heat flow from to with the initial value being the identity map. Again using (2.41), we see that the maps satisfy
(4.47) | ||||
for Since we have proved above that the operator in (4.47) is parabolic, we know from the standard theory of uniqueness of parabolic equations that for all with and as a result which gives . Consequently
Finally, we have
which gives
But since we proved above that we get and hence for all which proves the uniqueness of solutions to (GF) and completes the proof of the theorem. ∎
Remark 4.12.
A particularly interesting flow of Spin(7)-structures is the "coupling" of the Ricci flow of the metric and the isometric/harmonic flow of Spin(7)-structures
This flow induces precisely the Ricci flow on the metric, and the only other thing it does to the Spin(7)-structure is to deform it by the isometric flow which was studied in detail in [DLE24]. This “coupling” of Ricci flow with the isometric flow has good short-time existence and uniqueness. This can be seen by going through the modified DeTurck’s trick in § 4.4 and choosing and then checking that the resulting modified flow is strictly parabolic. The isometric flow of Spin(7)-structures has many good properties, in particular there is an almost monotonicity formula for the solutions to the flow (see [DLE24]). It would be interesting to study whether the Ricci flow coupled with the harmonic flow has similar analytic properties.
5. Solitons
Let be an -manifold. A soliton for (GF) is a triple with and such that
(5.1) |
where . Those Spin(7)-structures which satisfy (5.1) are special as they give self-similar solutions to (GF). Recall that a self-similar solution to (GF) is a solution of the form
where are time-dependent scalings and are a one-parameter family of diffeomorphisms. Note that the power of on is just for convenience in calculations. A straightforward calculation (for instance see [DGK21, Lemma 2.17], [DLE24, Prop. 2.11] or [FLME22, Prop. 1.55]) shows that the solitons for (GF) and self-similar solutions are in one-to-one correspondence.
We say a soliton is expanding if ; steady if ; and shrinking if .
We now derive the condition satisfied by the metric induced by and when is a soliton, which we expect to have further use.
Proposition 5.1.
Proof.
We can prove the non-existence theorem for compact expanding solitons of (GF).
Proposition 5.2.
References
- [AH11] Ben Andrews and Christopher Hopper “The Ricci flow in Riemannian geometry. A complete proof of the differentiable 1/4-pinching sphere theorem” 2011, Lect. Notes Math. Berlin: Springer, 2011 DOI: 10.1007/978-3-642-16286-2
- [AWW16] Bernd Ammann, Hartmut Weiss and Frederik Witt “A spinorial energy functional: critical points and gradient flow” In Math. Ann. 365.3-4, 2016, pp. 1559–1602 DOI: 10.1007/s00208-015-1315-8
- [CK04] Bennett Chow and Dan Knopf “The Ricci flow: an introduction” 110, Math. Surv. Monogr. Providence, RI: American Mathematical Society (AMS), 2004
- [DeT83] Dennis M. DeTurck “Deforming metrics in the direction of their Ricci tensors” In J. Differ. Geom. 18, 1983, pp. 157–162 DOI: 10.4310/jdg/1214509286
- [DGK21] Shubham Dwivedi, Panagiotis Gianniotis and Spiro Karigiannis “A gradient flow of isometric -structures” In J. Geom. Anal. 31.2, 2021, pp. 1855–1933 DOI: 10.1007/s12220-019-00327-8
- [DGK23] Shubham Dwivedi, Panagiotis Gianniotis and Spiro Karigiannis “Flows of -structures, II: Curvature, torsion, symbols, and functionals”, 2023 DOI: 10.48550/arXiv.2311.05516
- [DLE24] Shubham Dwivedi, Eric Loubeau and Henrique N. Earp “Harmonic flow of -structures” In Ann. Sc. Norm. Super. Pisa Cl. Sci. (5) Vol. XXV (2024), 151-215, 2024 arXiv:2109.06340 [math.DG]
- [Fer86] Marisa Fernández “A classification of Riemannian manifolds with structure group ” In Ann. Mat. Pura Appl. (4) 143, 1986, pp. 101–122 DOI: 10.1007/BF01769211
- [FLME22] Daniel Fadel, Eric Loubeau, Andrés J. Moreno and Henrique N. Earp “Flows of geometric structures” In arXiv e-prints, 2022, pp. arXiv:2211.05197 DOI: 10.48550/arXiv.2211.05197
- [Joy00] Dominic D. Joyce “Compact manifolds with special holonomy”, Oxford Mathematical Monographs Oxford University Press, Oxford, 2000, pp. xii+436
- [Kar08] Spiro Karigiannis “Flows of Spin(7)-structures” In Differential geometry and its applications World Sci. Publ., Hackensack, NJ, 2008, pp. 263–277 DOI: 10.1142/9789812790613_0023
- [Kra24] Kirill Krasnov “Dynamics of Cayley Forms”, 2024 arXiv:2403.16661 [math.DG]
- [LS23] Eric Loubeau and Henrique N. Sá Earp “Harmonic flow of geometric structures” Id/No 23 In Ann. Global Anal. Geom. 64.4, 2023, pp. 42 DOI: 10.1007/s10455-023-09928-7
- [Top06] Peter Topping “Lectures on the Ricci flow” 325, Lond. Math. Soc. Lect. Note Ser. Cambridge: Cambridge University Press, 2006
Institut für Mathematik, Humboldt-Universität zu Berlin, Rudower Chaussee 25, 12489 Berlin.
dwivedis@hu-berlin.de