[datatype=bibtex] \map \step[fieldsource=pmid, fieldtarget=pubmed]
Deformation theory of nearly manifolds
Abstract
We study the deformation theory of nearly manifolds. These are seven dimensional manifolds admitting real Killing spinors. We show that the infinitesimal deformations of nearly structures are obstructed in general. Explicitly, we prove that the infinitesimal deformations of the homogeneous nearly structure on the Aloff–Wallach space are all obstructed to second order. We also completely describe the cohomology of nearly manifolds.
1. Introduction
Given a -dimensional smooth manifold , a nearly structure on is a non-degenerate (or positive) -form such that for some non-zero real constant ,
(1.1) |
where the metric and the orientation and hence the Hodge star are all induced by . The existence of a nearly structure was shown to be equivalent to the existence of a real Killing spinor in [BFGK91]. A Killing spinor on a Riemannian spin manifold is a section of the spinor bundle such that
(1.2) |
for any vector field on and some . Here is the Clifford multiplication. It was proved by Friedrich [Fri80] that any manifold with a Killing spinor is Einstein with and one of the three cases must hold:
-
•
in which case is a parallel spinor and has holonomy contained in , , or .
-
•
is non-zero and is purely imaginary.
-
•
is non-zero and real, in which case is a real Killing spinor and if is complete then since it is positive Einstein, it is compact with finite.
Given a nearly structure on that satisfies equation (1.1), there exists a real Killing spinor that satisfies equation (1.2) with and vice-versa. See [BFGK91] for more details.
Using the equivalence with real Killing spinors, nearly structures on homogeneous spaces, excluding the case of the round -sphere, were classified in [FKMS97]. Their classification is based on the dimension of the space of Killing spinors . They showed that different types can occur:
-
1.
dim( - nearly structures of type 1.
-
2.
dim( - nearly structures of type 2.
-
3.
dim( - nearly structures of type 3.
A -dimensional manifold with a nearly structure is a nearly manifold (see §2 for more details). Other examples apart from the round include the squashed , Aloff–Wallach spaces , the Berger space and the Stiefel manifold . Another important aspect of nearly manifolds is that the Riemannian cone over has holonomy contained in the Lie group . In that case, the possible holonomies are , or depending on whether the link of the cone is a nearly manifold of type or respectively.
In this paper, we study the deformation theory of nearly manifolds. The infinitesimal deformations of nearly 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 -forms on of type . We address the question of whether nearly 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 manifolds. Another applicability of studying the deformation theory of nearly manifolds can be to develop the deformation theory of conifolds which are asymptotically conical and conically singular manifolds, similar to the theory developed by Karigiannis–Lotay [KL20] for conifolds. Lehmann [Leh20] studies the deformation theory of asymptotically conical 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 -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 -forms. Using this, they proved that the nearly Kähler structures on and are rigid and the flag manifold admits an -dimensional space of infinitesimal deformations. Later, Foscolo proved [Fos17, Theorem 5.3] that the infinitesimal deformations of the flag manifold are all obstructed.
Nearly manifolds are in many ways similar to nearly Kähler -manifolds. Both admit real Killing spinors and hence are positive Einstein. The minimal hypersurfaces in both nearly Kähler 6-manifolds and nealy manifolds behave in a similar way [Dwi19]. It was proved in [AS12] that the nearly structures on the squashed and the Berger space are rigid while the space of infinitesimal nearly deformations of the Aloff–Wallach space is -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 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 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 -manifolds.
We follow a strategy similar to [Fos17] in this paper. After introducing the Dirac operator and a modified Dirac operator on nearly manifolds in §3, we use their properties and the Hodge decomposition theorem to completely describe the cohomology of a complete nearly manifold. We prove our first two main results of the paper which characterize harmonic forms. These are the following.
Theorem 3.8. Let be a complete nearly manifold, not isometric to round . Then every harmonic -form lies in . Equivalently, every harmonic -form lies in .
Theorem 3.9 Let be a complete nearly manifold, not isometric to round . Then every harmonic -form lies in . Equivalently, every harmonic -form lies in .
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 structures on in Proposition 4.2. Using this, in Theorem 4.3 we obtain a new proof of the identification of the space of nearly deformations with an eigenspace of the Laplacian acting on co-closed -forms of type , a result originally due to Alexandrov–Semmelmann [AS12].
To study higher order deformations of nearly manifolds, we use the point view of Hitchin [Hit01] where he interprets nearly structures as constrained critical points of a functional defined on the space . 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 -manifolds. The advantage of this approach is that it allows us to view the nearly equation (2.24) as the vanishing of a smooth map (cf. equation (4.9))
where denotes the space of exact positive -forms on . Thus the obstructions on the first order deformations of a nearly structure to be integrated to higher order deformations can be characterized by which we do in Proposition 4.6.
Finally, we use the general deformation theory of nearly structures developed in the first part of the paper to study the infinitesimal deformations of the Aloff–Wallach space . 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 structure on the Aloff–Wallach space are all obstructed.
The proof of the above theorem is inspired from the ideas in [Fos17]. However, we note that since in the nearly 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 and nearly structures in §2. We discuss the decomposition of space of differential forms on manifolds with a 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 -forms and -forms on manifolds with nearly 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 deformations. We interpret the nearly equation as the vanishing of a smooth map and prove the characterization for a first order deformation of a nearly 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 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 -forms on nearly 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 geometry
We start this section by defining structures and nearly 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 be a smooth manifold. A structure on M is a reduction of the structure group of the frame bundle from to the Lie group . Such a structure exists on 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 structure on M is equivalently defined by a -form on that satisfies a certain pointwise algebraic non-degeneracy condition. Such a -form nonlinearly induces a Riemannian metric and an orientation on and hence a Hodge star operator . We denote the Hodge dual -form by . Pointwise we have , where the norm is taken with respect to the metric induced by .
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 to . Our convention for labelling the Riemann curvature tensor is
in terms of coordinate vector fields. With this convention, the Ricci tensor is , and the Ricci identity is
(2.1) |
We will use the metric to identify the vector fields and -forms by the musical isomorphisms. As such, throughout the paper, we will use them interchangeably without mention.
We have the following contraction identities between and , whose proofs can be found in [Kar09].
(2.2) | ||||
(2.3) |
and
(2.4) | ||||
(2.5) | ||||
(2.6) | ||||
(2.7) |
A structure on induces a splitting of the spaces of differential forms on into irreducible representations. The space of -forms and -forms decompose as
(2.8) | ||||
(2.9) |
where has pointwise dimension . More precisely, we have the following description of the space of forms :
(2.10) | ||||
(2.11) |
In local coordinates, the above conditions can be re-written as
(2.12) | ||||
(2.13) |
Similarly, for -forms
(2.14) | ||||
(2.15) | ||||
(2.16) |
Moreover, the space is isomorphic to the space of sections of , the traceless symmetric -tensors on M, where the isomorphism is given explicitly as
(2.17) | |||
The decompositions of and are obtained by taking the Hodge star of (2.9) and (2.8) respectively.
Given a structure on , we can decompose and according to (2.8) and (2.9). This defines the torsion forms, which are unique differential forms , , and such that (see [Kar09])
(2.18) | ||||
(2.19) |
Let denote the Levi-Civita connection of the metric induced by the structure. The full torsion tensor of a structure is a -tensor satisfying
(2.20) | ||||
(2.21) | ||||
(2.22) |
The full torsion is related to the torsion forms by (see [Kar09])
(2.23) |
Remark 2.1.
The space is isomorphic to the space of vector fields and hence to the space of -forms. Thus in (2.23), we are viewing as an element of which justifies the expression .
A structure is called torsion-free if or equivalently . We can now define nearly structures.
Definition 2.2.
A structure is a nearly structure if is the only nonvanishing component of the torsion, that is
(2.24) |
In this case, we see from (2.23) that .
Remark 2.3.
If is a nearly structure on then since , we can differentiate this to get and hence , as wedge product with is an isomorphism from to . Thus is a constant, if is connected.
Given a structure with torsion , we have the expressions for the Ricci curvature and the scalar curvature of its associated metric which can be found in [Bry06] or [Kar09] as
(2.25) | ||||
(2.26) |
where is the matrix norm in (2.26).
In particular, for a manifold with a nearly structure , we see that
(2.27) | ||||
(2.28) |
Finally, we remark that with the round metric and also the squashed are examples of manifolds with nearly structure (see [FKMS97] for more on nearly structures. The authors in [FKMS97] call such structures nearly parallel structures but we will call them nearly structures.) In particular, with radius has scalar curvature , so comparing with (2.24) we get that .
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 is a -form and is a vector field then
(2.29) | ||||
(2.30) |
If is a -form then we have the following identities
(2.31) | ||||
(2.32) | ||||
(2.33) | ||||
(2.34) |
Suppose is a vector field then we have the following identities
(2.35) | ||||
(2.36) | ||||
(2.37) | ||||
(2.38) |
Let be the non-linear map which associates to any structure , the dual -form with respect to the metric . We note that is defined only when we fix the orientation on . See [Hit01, §8] for more details. We will need the following result from [Joy00, Proposition 10.3.5], later.
Proposition 2.4.
Suppose be a structure on with . Let be a -form which has sufficiently small pointwise norm with respect to so that is still a positive -form and be a -form with small enough pointwise norm so that is a positive -form. Then
-
(1)
the image of under the linearization of at is
(2.39) -
(2)
the image of under the linearization of at is
(2.40)
2.1 First order differential operators
In this section, we discuss various first order differential operators on a manifold with a nearly structure and prove some identities involving them.
For , we have the vector field given by
and for any vector field we have the divergence of which is a function
On a manifold with a structure , for a vector field , we define the curl of , as
(2.41) |
which can also be written as
(2.42) |
and so up to -equivariant isomorphisms, the vector field is the projection of the -form onto the component. In fact, we have the following
Proposition 2.5.
Let be a vector field on . The component of is given by
(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 structure and will later state the results for nearly structures. These formulas are generalizations of the formulas first proved for torsion-free structures by Karigiannis [Kar06, Proposition 4.4].
Proposition 2.6.
Let and be a vector field on with a structure . Then
(2.44) | ||||
(2.45) | ||||
(2.46) |
Remark 2.7.
For fixed , the Riemann curvature tensor is skew-symmetric in and and hence
Explicitly,
Moreover, from [Kar09, eq. (4.17)], we have
(2.47) |
Proof.
We compute
as is skew-symmetric, thus proving (2.44). For (2.45) we use the Ricci identity (2.1) to get
where we used (2.12), (2.13) and (2.47). We have also used the fact that the symmetric part of will vanish when contracted with .
Finally we use the contraction identities (2.2) and (2.4) and the Ricci identity (2.1) to compute
where we used the fact that for the third term in the fourth equality and (2.13) to cancel the components which contract on two indices with for the last two terms in the fourth equality. Thus, we get
∎
For a nearly structure we have and . Moreover from [Kar09, eq. ],
Thus using the Weitzenböck formula for , , we get the following
Corollary 2.8.
Let and be a vector field on with a nearly structure . Then
(2.48) | ||||
(2.49) | ||||
(2.50) | ||||
(2.51) |
2.2 Identities for -forms and -forms
In this subsection, we prove some identities for -forms and -forms on a manifold with a nearly structure. These identities will be used several times in the paper.
Lemma 2.9.
Let be a manifold with a structure. If is a -form then
-
(1)
.
-
(2)
.
Proof.
Lemma 2.10.
Let be a manifold with a structure. Let be a -form on and let for some vector field on . Then
-
(1)
.
-
(2)
.
Proof.
For we have
(2.52) |
where we have used the fact that lies in the kernel of wedge product with and (2.35) in the last equality. For we note that lies in the kernel of wedge product with and . ∎
Next, we explicitly derive the expressions for exterior derivative and the divergence of various components of -forms and -forms on a manifold with a nearly 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 is a manifold with a nearly structure. Let , and . Then
-
(1)
.
-
(2)
.
-
(3)
.
-
(4)
.
-
(5)
.
-
(6)
.
Proof.
We prove part . Since is a -form so
(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
for some . Since lies in the kernel of wedge product with and for , we have
and hence
Suppose for . Using (2.11) and Lemma 2.10 , we have
Thus
which proves
Since is a -form, so we will write
(2.54) |
and will calculate each term on the right hand side. As before, assume
for some . Then
and hence . So we get that
Assume that
for some . Using the fact that lies in the kernel of wedge product with we get
So we get
which gives
and hence
Recall the map from (2.17). To calculate we have
(2.55) |
We calculate the left hand side of (2.55). We have
where we have used (2.20) and (2.24). So
We use the contraction identities (2.2), (2.3) and (2.4) to get
and so from (2.55) we get
and thus
which completes the proof of .
We obtain by
To prove part , we notice that since , which is the image of under the linearization of the map . 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 be a nearly structure on and be a -form so that
where with where is a symmetric traceless -tensor. Then
(2.56) | ||||
(2.57) | ||||
(2.58) |
Proof.
We note that and since is a nearly structure hence
(2.59) |
and
(2.60) |
Now for some . We use Lemma 2.11 to get,
(2.61) |
The first term on the right hand side of (2.61) is as and . The last term is also as from (2.16)
Thus we get that
which gives (2.56).
To derive (2.57) and (2.58), we will need to contract with on two indices and with on three indices. Using (2.17) and the contraction identities (2.2) and (2.5), a short computation gives
(2.62) | ||||
(2.63) |
Suppose for some -form . Note that for an arbitrary -form we have
So from (2.59) we have
(2.64) |
We first use Lemma 2.11 to calculate the second term on the right hand side of (2.64). We have
So in (2.64), we have
(2.65) |
We compute in local coordinates
We now use (2.62), (2.63) and the fact that is traceless to get
Thus from (2.65) we get
and since is arbitrary, we get
which establishes (2.57).
Next, we see from (2.60) and (2.10) that
which on using (2.43) becomes
(2.66) |
Suppose for some -form . For any -form we note that
Thus using (2.66) and the orthogonality of the spaces and , we have
(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
and hence we get
Since is arbitrary we get
which gives (2.58). ∎
Remark 2.13.
The main point of the previous lemma is to exhibit a relation between and . Such a relation is expected because of the form of the linearization of the map . More precisely, from (2.39), applying the linearization of to Lie derivatives, we have , and . The computations in local coordinates was done to relate and to the divergence of the symmetric -tensor .
Remark 2.14.
The previous lemma generalizes Proposition 2.17 from [KL20] where the structure was assumed to be torsion-free .
We have the following corollary of Lemma 2.12.
Corollary 2.15.
Let be a nearly structure and let . Then
-
(1)
If is closed then .
-
(2)
If is co-closed then .
Proof.
We also have a result similar to Lemma 2.12 for -forms which we state below. The proof follows from the proof of Lemma 2.12 by taking and noting that . We expect that both Lemma 2.12 and Lemma 2.16 will be useful in other contexts as well.
Lemma 2.16.
Let be a nearly structure on and be a -form on so that
where and with where is a symmetric traceless -tensor. Then
(2.68) | ||||
(2.69) | ||||
(2.70) |
We get the following corollary.
Corollary 2.17.
Let be a nearly structure on and let . Then
-
1.
If then .
-
2.
If then .
3. Hodge theory of nearly manifolds
3.1 Dirac operators on nearly manifolds
We begin this section by defining the Dirac operator on with a nearly structure. We then define a modified Dirac operator which is more suitable for our purposes. A structure on induces a spin structure, so admits an associated Dirac operator on its spinor bundle . Since is constant, by rescaling the metric induced by the nearly structure, we can change the magnitude of and by changing the orientation, we can change its sign. In the later part of the paper, we study deformations of nearly structures through nearly structures . Since the underlying metric of any nearly structure is positive Einstein, the family of metrics corresponding to will be positive Einstein and so by [Bes87, Corollary 2.12], the scalar curvature is constant in . Thus, by (2.28), will be constant through the deformation. Henceforth, we will assume that . The results of the paper do not depend on the value of chosen. Recall the following definition from §1 with .
Definition 3.1.
A spinor is called a Killing spinor if for any
(3.1) |
where is the Clifford multiplication.
The real spinor bundle , as a representation, is isomorphic to , where the isomorphism is
For comparison with the Dirac-type operator which we define later, let us derive a formula for the Dirac operator on a nearly manifold in terms of this isomorphism.
A unit spinor on a nearly manifold satisfies (3.1). Thus
where we have used the fact that . Also,
which on using and (3.1) becomes
Thus we get
(3.2) |
Now is a -form, hence . Since the Lie group preserves the nearly structure , it preserves the real Killing spinor induced by and , the Lie algebra of , we have . Also, we know from (2.42) that and it follows from the definition of the Clifford multiplication, for instance as in [Kar06, §4.2], that for any , we get that
which we will write as
(3.3) |
Definition 3.2.
The Dirac operator is a first-order differential operator on defined as follows. Let . Then
(3.4) |
The Dirac operator is formally self-adjoint, that is, and is also an elliptic operator.
Consider the Dirac Laplacian . We relate it to the Hodge Laplacian in the following
Proposition 3.3.
Let be a section of the spinor bundle . Then
(3.5) |
Thus is equal to up to lower order terms.
We need a modification of the Dirac operator defined above. The spinor bundle is isomorphic to and hence, via a -equivariant isomorphism, it is also isomorphic to . We define the modified Dirac operator, which we denote by , as follows. Consider the map
Using Lemma 2.11 with , we get
(3.6) |
Remark 3.4.
We note that is defined in the same way as in [KL20] where the authors denote the operator by .
We find the kernel of . Let be in the kernel of . Then
Taking of the second equation and using the first equation and equation (2.49), we get
Since is a non-negative operator, . For , we have
We want to prove that is a Killing vector field. Let . Then
Therefore From Lemma 2.9 , we have
On the other hand, since is compact, using integration by parts we have
Therefore, and . Now using Lemma 2.11 , along with the fact that , i.e., and , we get
and hence is a Killing vector field. Therefore is isomorphic to the set of Killing vector fields such that . We denote by , that is,
(3.7) |
Remark 3.5.
Note that the above can also be proved using the identity , since for .
Remark 3.6.
If we also want the vector field to preserve the structure, then
but since , this implies . Hence the only vector fields in that preserve the structure are trivial. Note that when is a nearly structure of type-1, that is , every Killing vector field preserves the structure and hence .
The motivation for defining the modified Dirac operator can be understood from the following.
Consider the following operator
From previous calculations and Lemma 2.11 we know that
Thus
Doing a similar calculation as we did for , we observe that if , then
and so . Since and are self-adjoint operators, we have the following identification
(3.8) | ||||
This is used in the following important
Proposition 3.7.
Let be a nearly manifold. Then the following holds.
-
1.
.
-
2.
We have an -orthogonal decomposition .
Proof.
The first part of the proposition follows from the decomposition of in equation (3.8).
For the second part we note that the space is -orthogonal to exact -forms. To prove the -orthogonality of the remaining summands we proceed term by term. Let , and , such that for some exact -form . Using the pointwise orthogonality of and , we have
Note that since , Lemma 2.11 implies that , and hence is exact. Thus, . Let . The -orthogonality of and , along with the identity implies
The orthogonality of and follows from the -orthogonality of and . ∎
Thus, from the previous proposition, we know that any -form on a nearly manifold can be written as , for some and . Since for , , one can choose in the previous proposition.
Thus for every -form there exists unique and such that
3.2 Harmonic -forms and -forms on nearly manifolds
The above decomposition of -forms has a very useful application in determining the cohomology of nearly manifolds. We first note that since nearly manifolds are positive Einstein, it follows from Bochner formula and Hodge theory that any harmonic -form is and hence . The next two theorems describe the degree , and degree and cohomology of a nearly manifold.
Theorem 3.8.
Let be a complete nearly manifold. Then every harmonic -form lies in . Equivalently, every harmonic -form lies in .
Proof.
Let be a harmonic -form that is . From Proposition 3.7 there exists and such that
Since and hence , by Lemma 2.11 , and since , we have
Thus and hence .
Since , applying Corollary 2.17 on implies . This identity together with the closedness of gives us
as . Hence or equivalently , thus which implies that which completes the proof of the theorem. ∎
We also describe the degree 2 (and hence degree 5) cohomology on nearly manifolds below. In combination with Theorem 3.8, this completely describes the cohomology of a nearly manifold.
Theorem 3.9.
Let be a complete nearly manifold with . Let be a -form with
If is harmonic then .
Proof.
Suppose is harmonic. Then and since and are linear, we have
which on using Lemma 2.11 , and imply
and
Thus we get
and so
(3.9) |
Now , so taking of both sides and using (2.51) with , we get
Thus is harmonic. Since nearly manifolds are positive Einstein, it follows from Bochner formula and Myers theorem that . Hence . ∎
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 -instantons on a nearly manifold, as already described in [BO19, Corollary 14]. For any , by Theorem 3.9, there is a unique -instanton on a complex line bundle with .
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 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 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 structure (not necessarily nearly ) with suitable modifications. The authors are currently investigating this.
4. Deformations of nearly structures
Let be a nearly manifold with a nearly structure . We are interested in studying the deformation problem of in the space of nearly 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 be the space of structures on , that is, the set of all with . Given a point we define the tangent space .
Lemma 4.1.
The tangent space is the set of all such that
for some and .
4.1 Infinitesimal deformations
We want to study deformations of a given nearly structure on a compact manifold by nearly structures . We will only be interested in deformations of the nearly structures modulo the action of the group where denotes the space of diffeomorphisms of which are isotopic to the identity. We first use Proposition 3.7 to find a slice for the action of diffeomorphism group on which is used to find the space of infinitesimal nearly deformations, a result originally due Alexandrov–Semmelmann [AS12].
For the purposes of doing analysis, we consider the Hölder space of structures on such that and are of class , . Let be a nearly structure such that the induced metric is not isometric to round . Denote the orbit of under the action of – diffeomorphisms isotopic to the identity, by . The tangent space is the space of Lie derivatives for . We are interested in finding a complement of in .
If then using Proposition 3.7 , we can write
for unique and . From Lemma 2.11 we know that
and since
from Lemma 2.11 , we see that
Thus up to an element in we get that
(4.1) |
and hence from Lemma 4.1
(4.2) |
Now, if then from Lemma 2.11 and we see that
and hence
which implies that up to an element in combined with the above observation, we can write
(4.3) |
which implies that
(4.4) |
and hence we get a splitting where which consists of 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 of of the origin such that the “exponentiation” of is a slice for the action of on a sufficiently small neighbourhood of .
With this choice of slice we determine the infinitesimal deformations of the nearly structure which gives a new proof of a result of Alexandrov–Semmelmann [AS12, Theorem 3.5].
Theorem 4.3.
Let be a complete nearly manifold, not isometric to the round . Then the infinitesimal deformations of the nearly structure are in one to one correspondence with with
(4.5) |
Hence is co-closed as well. Moreover, .
Proof.
Let be an infinitesimal nearly deformation of a structure . So must be exact and hence from Proposition 3.7 , we can remove the term, in which case (4.1) and (4.2) become
(4.6) |
Moreover, for infinitesimal nearly deformations we must have
and hence (4.6) implies
Using Lemma 2.11 for the fourth term above and taking inner product with gives
But since and hence we get . Since is not isometric to round , Obata’s theorem then implies that and hence
(4.7) |
which proves the one to one correspondence between the infinitesimal nearly deformations and . Since and is a Killing vector field, we have which is the second part of (4.5). Since is exact, . From (4.7) and the fact that , we get
and hence
Taking of both sides give . Moreover,
which completes the proof of the theorem. ∎
Remark 4.4.
Motivated from the study of deformations of nearly Kähler -manifolds by Foscolo [Fos17, §4] where he used observations of Hitchin [Hit01], we also want to interpret the nearly condition (2.24) as the vanishing of a smooth map on the space of exact positive -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 be an exact positive -form, not necessarily satisfying the nearly condition. Let be the first order deformation of . Hitchin in [Hit01] defined a volume functional for exact -form given by
and a quadratic form
where and are exact -forms. We denote by . When is compact, Hitchin proves [Hit01, Theorem 5] that stable -forms (which is the same as a positive -form in our case) is a critical point of the volume functional subject to the constraint if and only if defines a nearly structure. The linearization of the volume functional at is given by
For the linearization of the quadratic form, suppose is exact with . We use integration by parts to get
Let us define an energy functional on exact -forms by
Then from above calculations
Therefore is a critical point of if and only if for every that is if and only if
Hence the critical points of the functional on are nearly structures. Since the energy functional is diffeomorphism invariant, we can introduce an extra vector field, as will vanish in the direction of Lie derivatives. Thus being a stable exact -form can be given by the formula
for some . We use these observations to write the nearly condition (2.24) as the vanishing of a smooth map. Let us denote by the space of stable and stable, exact -forms, i.e., . We have the following
Proposition 4.5.
Suppose satisfies
(4.8) |
for some vector field and denotes the Hodge star with respect to a fixed background metric. Then is a nearly structure.
Proof.
Suppose we want to describe the local moduli space of nearly structures on a manifold . If denotes the space of nearly structures on then the local moduli space is . A natural way to study this problem is to view the nearly structures on 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 .
Now let be a nearly structure on . Let be a small neighborhood of the -form . Since the condition of being stable is open we can assume the existence of such a neighborhood. Thus for with sufficiently small norm with respect to the metric induced by , is also a stable exact -form. From Proposition 4.5 the pair of stable forms defines a nearly structure if there exists a such that
This condition is equivalent to the vanishing of the map
(4.9) |
Thus, the nearly structures are the zero locus of the map modulo diffeomorphisms.
Let be the dual of under the Hitchin’s duality map as in Proposition 2.4. The linearization of the map at the point is given by
Thus the obstructions on the first order deformations of the nearly structure are given by 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 be a nearly structure and be a first order deformation in . Then lies in the image of if and only if
for all co-closed such that .
Proof.
From Proposition 3.7 , there exists and such that
and from Lemma 4.1, the -form
By Proposition 3.7, for some . Such an lies in the image of if
From Lemma 2.11 (5)
Comparing the last term in the above expression with that of in Lemma 2.11 we get
Using these expressions for and we get
Thus, for finding the , we need to solve the equations
(4.10) | ||||
Let . Then by Implicit Function Theorem, a solution of the first pair of equations always exist if the operator
is invertible in a small neighborhood of its zero locus. Since differs from the modified Dirac operator in (3.6) only by self-adjoint zeroth-order term, it is self-adjoint and hence . A pair is in the kernel of the operator if and only if
Applying the operator on the second equation and using the fact that gives
Thus as is a non-negative operator. The second equation then becomes
and Proposition 2.5 implies that . Using Lemma 2.9 (2) we get that
On the other hand
Combining these two equations we get and hence as well. Thus and is invertible when and we can always solve the first pair of equations in (4.10). Thus there are no restrictions on to be in the image of . Moreover if satisfies the third equation in (4.10) then
which on using the fact that is co-closed implies
Thus is a solution to the equation (4.10) (3) if and only if
for all co-closed such that . To complete the proof of the proposition we now only need to prove the -orthogonality condition for . But observe that since
and so . Since is co-closed, from Corollary 2.15 and
Similarly
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 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 -manifolds, we define the notion of second order deformations of nearly structures.
Definition 4.8.
Given a nearly structure and an infinitesimal deformation , a second order deformation of in the direction of is a pair such that
is a nearly structure up to terms of order . An infinitesimal deformation 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 structures on a manifold .
Remark 4.10.
In a similar way, we can define higher order deformations of a nearly 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 structure , we look for formal power series defining positive exact -form
where and a vector field
which satisfy (4.8), that is
(4.11) |
where is the dual of . Note that the Hodge star is taken with respect to .
Since we are interested in second order deformations, given an infinitesimal nearly deformation , we set and look for such that (4.11) is satisfied upto terms of . Explicitly, we write
where denotes the linearization of Hitchin’s duality map for stable forms in Proposition 2.4 and is the quadratic term of Hitchin’s duality map. Since we want solutions to (4.11) up to second order, we look for such that
(4.12) |
as and is the only second order term in . 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 structure and solutions to (4.12). We do this in the following lemma.
Lemma 4.11.
Proof.
We start with
Since , hence from (2.18) and (2.19) we see that for any vector field , . Thus the terms which are in vanish, that is
Using the facts that , , being an -form on a seven dimensional manifold and we get that
Taking proves that . From (4.12) we get that
which proves that is a second-order deformation of in the direction of in the sense of Definition 4.8. Conversely, suppose that is a second-order deformation of . Then . ∎
From the previous proposition and Proposition 4.6 we have that if is a second order deformation of the nearly structure in the sense of Definition 4.8 then
(4.13) |
for all such that . The above equation simplifies to
Moreover, if is an infinitesimal deformation of , then by Theorem 4.3 satisfies (which of course implies ) and so the above equation is equivalent to
5. Deformations on the Aloff-Wallach space
In [AS12, Prop. 8.3] Alexandrov–Semmelmann established that the space of infinitesimal deformations of the nearly structure on the Aloff–Wallach space is an eight dimensional space isomorphic to , the Lie algebra of . The rest of the paper is devoted to prove that these deformations are obstructed to second order.
The embedding of and in , which we denote by and , following [AS12], is given by
The Lie algebra splits as
where is the -dimensional orthogonal complement of with respect to , the Killing form of . The normal nearly metric on is then given by where the constant comes from our choice of . If we denote by the standard -dimensional complex irreducible representation of and by the -dimensional complex irreducible representation of with highest weight , then as an -representation
Let be the basis of . If we define , we have
This basis is orthonormal with respect to the metric . We use the shorthand to denote the -form . The nearly structure is given by
As an representation, where
By Theorem 4.3, the space of first order deformations is given by . In this example, it was found to be isomorphic to . As an representation, is isomorphic to the span of . The -invariant homomorphism from to is given by where
Let us fix an . The adjoint action of is given by
where are functions on .
The infinitesimal deformation associated to such that is given by
We can now compute the -form by using the relation . In order to show that the infinitesimal deformation associated to is obstructed to second order, we need to compute the quadratic term as discussed in equation (4.13) and find an element for which the -inner product is non-zero.
To compute , one can use the algorithm for stable -forms on manifolds with structures as discussed in [Hit01]. Using the fact that , one can easily show that for some non-zero constant , where is the quadratic term associated to . Thus, we will instead compute and show that the inner product to prove obstructedness.
Consider to be a positive -form for small . We will denote the metric and the volume form induced by by and respectively. We have a Taylor series expansion
Then one can define the symmetric bi-linear form by
The zero order term of , denoted by is given by . Similarly, one can compute the linear term and the quadratic term . The metric is then defined using the relation (see for example, [Kar09])
The linear term in is proportional to and thus vanishes since . Using the above formula we get that
where is a quadratic polynomial in and . Using the Taylor series expansion of and , we can compute the Taylor series expansion of the Hodge star associated to , . The Hodge star operator can be computed using the formula
The quadratic term is then given by
In the present case, for a general element , the quadratic term turns out to be very complicated and is not very enlightening. We define the cubic polynomial on by
Note that is cubic in since and are quadratic and linear in respectively. This cubic polynomial can be lifted to a polynomial on the Lie group by
This lift enables us to calculate the average of on by using the Peter–Weyl theorem. To express the polynomial in a compact form, we will set . Then the cubic polynomial is given by
(5.1) |
The next step in proving obstructedness is to show that the average value of on is non-zero. For this, we appeal to the Peter–Weyl theorem. The Peter–Weyl theorem states that for any compact Lie group , we have
where denotes the set of all non-isomorphic irreducible representations of .
The cubic polynomial lies in the representation . The average value of the function on is the same as the average value of where is the projection of to the invariant polynomials. This is because lies in the non-trivial part of the Peter–Weyl decomposition and has an average value of zero. The unique trivial sub-representation of is generated by the determinant polynomial on which is given by
The average value of the polynomial can be computed by computing the inner product of with . On , since the Killing form is non-degenarate, defines an inner product on . The inner product induces an inner product on in the natural way. All the computations that follow are done using .
If denotes the matrix with as the -th entry and zero elsewhere, then the subspace of generated by is orthogonal to . Moreover are also orthogonal to each other. Thus the only non-trivial terms occurring in the inner product of and are,
From (5.1) and the above computations we have that
Thus we get the following theorem.
Theorem 5.1.
The infinitesimal deformations of the homogeneous nearly structure on the Aloff–Wallach space are all obstructed.
References
- [AS12] Bogdan Alexandrov and Uwe Semmelmann “Deformations of nearly parallel -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 -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 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 -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 and structures” In Canad. J. Math. 57.5, 2005, pp. 1012–1055 DOI: 10.4153/CJM-2005-039-x
- [Kar06] Spiro Karigiannis “Some Notes on and Geometry”, 2006 arXiv:math/0608618 [math.DG]
- [Kar09] Spiro Karigiannis “Flows of -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 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 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 -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]