Special Lagrangians in nearly Kähler
Abstract
This article explores special Lagrangian submanifolds in , viewed as a nearly Kähler manifold, from two different perspectives. Intrinsically, using a moving frame set-up, and extrinsically, using moment-type maps. We describe new homogeneous examples, from both perspectives, and classify totally geodesic special Lagrangian submanifolds. We show that every special Lagrangian in , or the flag manifold admitting a symmetry of an subgroup of nearly Kähler automorphisms is automatically homogeneous.
Nearly Kähler manifolds were first introduced in the 1970s and in the last two decades fundamental questions about the structure and existence of these manifolds were settled [Nag02, But10, FH17], making them a trending topic in differential geometry. In dimension 6, they form a special class of -structures and are important for Riemannian geometry as they provide examples of Einstein manifolds. In addition, they are of interest in exceptional holonomy as their cones are torsion free manifolds, which makes nearly Kähler manifolds crucial for understanding manifolds with singularities.
One peculiarity of Lagrangian submanifolds of nearly Kähler manifolds is that they are automatically special Lagrangian. Because of their simple definition they are natural objects to study in nearly Kähler geometry, following the strategy to understand an ambient space by studying its distinguished submanifolds. Just as for -holomorphic curves in nearly Kähler manifolds there are two additional lines of motivation to study Lagrangian submanifolds. The first one comes from Riemannian geometry, for any special Lagrangian in a nearly Kähler manifold is minimal. The second comes from special holonomy, for the cone of a special Lagrangian is coassociative in the -cone of .
In the last few decades, many constructions for special Lagrangian submanifolds of have been found and various subclasses of special Lagrangians have been classified, see for example [Vra03, Lot11]. More recently, the ambient spaces and have received attention, for example in [Bek+19, Sto20a]. This article is dedicated to the ambient nearly Kähler space . The main results of this article are the classification of totally geodesic special Lagrangians in 3.18 and of special Lagrangians admitting a symmetry of a 3-dimensional group of nearly Kähler automorphisms in 4.11. These results are obtained through two different approaches to describe special Lagrangians in .
The first approach is intrinsic as it uses the structure equations describing a special Lagrangian. We derive them in the general nearly Kähler setting in section 2 and show in 2.1 that for a homogeneous ambient space and a simply connected domain there is a unique Lagrangian immersion for every solution to the structure equations.
In section 3, we adapt these equations to the twistor fibration . We introduce an angle function parametrising the Lagrangian at a tangent level. Generically, intersects every twistor fibre transversally. The points where are those where this is not the case, and the intersection is then diffeomorphic to a circle. We identify Lagrangians with as circle bundles over superminimal surfaces in , a construction discovered in [Sto20] and in [Kon17].
Finally, we classify all Lagrangians where takes the boundary value . In fact, there are just two such examples and they are both homogeneous. We describe another somewhat surprising homogeneous example with arising from the irreducible representation of on . We also show that the standard in is the only totally geodesic Lagrangian submanifold of .
The second approach to exploring special Lagrangians in is extrinsic. In section 4, we introduce moment maps in nearly Kähler geometry. They encode the symmetry of the nearly Kähler manifold in a set of equivariant functions . We use these moment-type maps to show a general existence result of special Lagrangians with symmetry, in 4.3 and to classify special Lagrangians admitting an action of a group of automorphisms, in 4.11. We show that they are, in fact, all homogeneous and describe the examples found in section 3 extrinsically. We show an analogous result for the flag manifold by studying the action of three-dimensional subgroups of on .
Most of the material presented in this article originates from author’s PhD thesis, [Asl22].
Acknowledgement. The author is grateful to Jason Lotay, Simon Salamon and Thomas Madsen for their advice and support. The review from Lorenzo Foscolo and Luc Vrancken of the author’s PhD thesis has led to a significant improvement of the material.
This work was supported by the London Mathematical Society grant ECF-2021-01.
1 Background
Let be a 6-dimensional almost Hermitian manifold. Then is called nearly Kähler provided there is a complex-valued three-form defining an -structure satisfying
Remark 1.1.
Often, the Calabi-Yau case is also included in the definition of a nearly Kähler manifold. We choose to exclude this case, so there is no need to introduce the subclass of strictly nearly Kähler manifolds.
Every nearly Kähler manifold admits a unique connection with totally skew-symmetric torsion and holonomy contained in , i.e. , cf. [Gra70]. This connection is called the characteristic connection and related to the Levi-Civita connection by
(1.1) |
see [MNS08]. Examples of (compact) nearly Kähler manifolds are very scarce. In fact, there are only six known examples of compact simply-connected nearly Kähler manifolds.
Proposition 1.2.
[But10, Theorem 1] If is a homogeneous strictly nearly Kähler manifold of dimension six, then is one of the following: , , or .
In each case, the identity component of the group of nearly Kähler automorphisms is equal to , see [DC12]. There are infinitely many freely-acting finite subgroups of the automorphism group of the homogeneous nearly Kähler , cf. [CV15].
In addition, there are two known examples of compact, simply-connected nearly Kähler manifolds which are not homogeneous. They were constructed by Foscolo and Haskins via cohomogeneity one actions on and [FH17].
A 3-dimensional submanifold of a nearly Kähler manifold is called Lagrangian if . For the general set-up, we will also work with the more flexible notion of an immersed Lagrangian submanifold, i.e. we a smooth immersion such that . However, all examples we encounter are embedded submanifolds.
Because of the nearly Kähler identity , Lagrangian submanifolds are automatically special Lagrangian. Special Lagrangians in nearly Kähler geometry share some important general properties with special Lagrangians in Calabi-Yau manifolds. Every special Lagrangian in is minimal and orientable, see for example [VS+19].
Let be second fundamental form of in . Then the cubic form is fully-symmetric, i.e. an element of , and traceless when contracted in any two components, see [SS10]. The cubic form is also called the fundamental cubic of and takes values in the intrinsic bundle . This means that in order to study special Lagrangians where satisfies special properties one does not need to specify the normal bundle of . One such special property would be that , or equivalently the second fundamental form, is a parallel section. However, it turns out that this assumption is rather restrictive. Any such Lagrangian is automatically totally geodesic [Zha+16, Theorem 1.1].
Another special property of is that it admits symmetries. This approach has been developed in [Bry06a] for special Lagrangian submanifolds of . By picking a frame in a point one regards as a harmonic polynomial of degree three in three variables, i.e. an element of , which is a seven-dimensional vector space. The space as an module and a generic element in does not have any symmetries in . The possible symmetry groups are classified in [Bry06a, Proposition 1]. The classification gives a natural ansatz for finding special Lagrangian submanifolds. Impose one of the pointwise symmetries above to every point in . This ansatz has led to the construction of new special Lagrangians in the Calabi-Yau in [Bry06a] and in the nearly Kähler [Vra03, Lot11]. For however, this ansatz is less fruitful since the curvature tensor is more complicated so we do not have -freedom to change frames as we will see later. However, this framework gives us a way to categorise examples of special Lagrangian that are constructed in different ways.
The following result is known for Calabi-Yau manifolds and the nearly Kähler but it holds for any nearly Kähler manifold.
Proposition 1.3.
Every real analytic surface on which vanishes can locally be uniquely thickened to a special Lagrangian submanifold in . Special Lagrangian submanifolds in a nearly Kähler manifold locally depend on two functions of two variables.
Proof.
See [Lot11], the proof is based on the fact that the Cartan test holds and thus holds for any structure. ∎
Infinitesimal deformations of nearly Kähler manifolds correspond to eigensections of a rotation operator on [Kaw17]. It is shown in [VS+19], that the moduli space of smooth Lagrangian deformations of special Lagrangians is a finite dimensional analytic variety. All formally unobstructed infinitesimal deformations are smoothly unobstructed.
1.1 The nearly Kähler Structure on
The nearly Kähler structure on can be defined through the twistor fibration . The fibres are projective lines and totally geodesic for the Kähler structure on . Since is a sphere bundle inside the twistor fibration has a natural connection . The nearly Kähler structure on is defined via the Kähler structure by squashing the metric and reversing the almost complex structure on the vertical fibres.
For explicit computations it is convenient to define the nearly Kähler structure from the homogeneous space structure . Identify with via . This identification gives an action of on which descends to and acts transitively on that space. The stabiliser of the element is
which shows . Following [Xu10], consider the Maurer-Cartan form on which can be written in components as
(1.2) |
Since has values in , the one-forms and are complex-valued and are real-valued. The equation implies the torsion identity
(1.3) |
and the curvature formula
The nearly Kähler structure on is defined by declaring the forms and to be unitary forms for any local section of the bundle . The resulting almost complex structure and metric do not depend on the choice of . The nearly Kähler forms are pullbacks of
respectively. In general, we will treat the nearly Kähler forms as basic forms on . However, Killing vector fields typically have a simple expression in local coordinates. To contract the nearly Kähler forms on with Killing vector fields we pull back the local unitary forms on the chart with the local section
Here,
This gives the following expressions for the pull-backs
(1.4) |
To show these formulae, note that the pullback of the Maurer-Cartan form via is
Combining this with eq. 1.2 yields
where is a real term. Equations 1.4 follow by splitting the quaternionic-valued differential forms on the right-hand side into their and part.
2 Structure Equations for Special Lagrangians
The structure equations for a special Lagrangian manifold in Calabi-Yau were established in [Bry06a] and for nearly Kähler in [Lot11]. We generalise the equations to the setting of a general nearly Kähler manifold. The main difference is the appearance of an extra curvature term. We characterise nearly Kähler manifolds by differential identities on the frame bundle, as done in [Bry06]. If an index appears on the right-hand side but not on the left-hand side of an equation, summation over the index set is implicit.
Let be a nearly Kähler manifold and consider the -frame bundle . Let be the tautological one-forms on and let be the nearly Kähler connection one-form on , giving the torsion relation
(2.1) |
and the curvature identity
(2.2) |
In particular, the curvature of is always of type . In [Bry06] it is remarked that the tensor can be written as sum
where has the following symmetries
The tensor vanishes exactly when is the round six-sphere. The nearly Kähler forms are expressed in terms of by
(2.3) |
Note the difference from [Bry06] in the convention for in order to satisfy the standard nearly Kähler integrability equations.
The torsion-relation eq. 2.1 and curvature-relation eq. 2.2 yield differential identities for the connection one-form and tautological one-form on the frame bundle . If is a special Lagrangian submanifold in then one obtains more differential identities because the frame bundle admits a natural reduction to an bundle over . The reason for this is that, at the tangent level, a Lagrangian subspace looks like in , which defines the restriction
If are the standard complex-valued one-forms on then is characterised as the 3-dimensional subspace of on which the imaginary parts of vanish. Similarly, our aim is to describe the reduction as the vanishing set of one forms on . To that end, split the forms and into real and imaginary part. The bundle is now defined by imposing the condition .
This characterisation implies more differential identities. From the torsion-relation we get
where is an cyclic permutation of . The condition implies . By Cartan’s lemma, we have or where is a fully symmetric three-tensor. In fact, this tensor corresponds to the fundamental cubic up to a factor, just as in the case of special Lagrangians in or in .
On the reduced bundle, we split into real and imaginary part,
This also allows us to split the curvature identity into real imaginary part
To write these equations more invariantly, let
We can summarise the equations on the reduced bundle over in tensor notation
(2.4) | ||||
(2.5) | ||||
(2.6) | ||||
(2.7) |
where . The matrix of one forms is completely defined by the symmetric tensor . The advantage to work with is that its components are not one-forms but functions, allowing us to rewrite eq. 2.4, eq. 2.6 and eq. 2.7
The Levi-Civita connection one-form of the induced metric on is . Note that the forms differ by a factor 2 from the orthonormal one forms considered in [Lot11].
If is one of the homogeneous nearly Kähler manifolds then a special Lagrangian submanifold can locally be recovered from a solution to eq. 2.4-eq. 2.6, which we will make precise now. There is a splitting such that . The nearly Kähler structure then yields an invariant special unitary basis on . Up to a cover, embeds into the -frame bundle via the adjoint action . Under this identification
is the Maurer-Cartan form on . In other words, the nearly Kähler connection is equal to the canonical homogeneous connection on , see [But10]. The following proposition guarantees that for the homogeneous nearly Kähler manifolds we can locally recover the special Lagrangian from a solution of the structure equations. Since and determine the first and second fundamental form, this can be viewed a Bonnet-type theorem.
Proposition 2.1.
Let be a homogeneous nearly Kähler manifold, be a simply-connected three manifold and , defining a linearly independent co-frame at each point, and satisfying the equations 2.4-2.7. Then there is a special Lagrangian immersion , unique up to isometries, with determining the metric and second fundamental form of in .
Proof.
Remark 2.2.
Note that the tautological one form can also be regarded as an element in . With this identification, a local section of gives a section . Then vanishes on while vanishes on the normal bundle.
3 An Angle Function for Special Lagrangians
Since twistor fibres are -holomorphic they can never be contained in a special Lagrangian submanifold. Generically, a special Lagrangian intersects every twistor fibre transversally. However, there is a special class of special Lagrangians which are circle bundles over superminimal surfaces in . We review this construction and define an angle function which has value if intersects a twistor fibre non-transversally. We use a gauge transformation, which depends on , to use the moving frame setup from the previous section for special Lagrangians in . We identify special solutions to the resulting structure equations, all of which turn out to be homogeneous.
3.1 The Linear Model
We start with the study of Lagrangian subspaces in a twistor space on the tangent level. The space of special Lagrangian subspaces of is identified with the homogeneous space . Twistor nearly Kähler spaces have the property that the holonomy of the nearly Kähler connection reduces to . The two-form splits into a horizontal and vertical part . So, in order to understand how frames can be adapted further to a special Lagrangian of a twistor space, we study the linear problem first.
Let denote the standard basis of with dual basis and let as well as . Let be the stabiliser of inside . Let also be the complex-valued three form on . We have abused notation slightly here, since are forms on the nearly Kähler manifold but also denote their linear models on .
For a complex subspace denote by the set of all special Lagrangian subspaces of . By we refer to the subspace spanned by and . Note that and that acts from the left on this space. The quotient is an interval and the following lemma gives a description of each representative.
Lemma 3.1.
Under the action of any element in has a unique representative of the form , for .
Proof.
Special Lagrangian planes in are parametrised by . Thus, we have to find a unique representative of the action of on , which is the action of a maximal torus in acting on . The standard torus acting on admits unique representatives of the form for . The statement follows by conjugating the action of to the standard torus action. ∎
For any subspace denote by the kernel of the projection onto and by its dimension. Let
(3.4) |
and be the image of when applied to the standard in , i.e. .
Proposition 3.2.
Any special Lagrangian subspace admits a unique representative for , under the action of . Furthermore, if and only if .
Proof.
Since is Lagrangian, . If then is represented by the standard in and if and only if . So from now on we assume that . Consider the map . Note that and that descends to a map To show surjectivity observe that for we have . So by acting with we can achieve that is spanned by . Furthermore, observe that
which means that for . We have shown that any element in is represented by a for . The uniqueness follows from the observation that has norm when restricted to the vector space . ∎
If is a basis of such that and then can be computed by the formula
(3.5) |
3.2 gives a geometric interpretation of the boundary value . In 3.11 we relate the case to CR-manifolds in the Kähler , so we study this case on the linear level first. Motivated by the existence of the two almost complex structures and on the twistor space, consider the almost complex structure
(3.6) |
on . Any special Lagrangian subspace in splits as .
Lemma 3.3.
If then is orthogonal to . The subspace is invariant under if and only if .
Proof.
The endomorphism commutes with the action of on , so it suffices to prove the statement for . If then is one-dimensional so it cannot be invariant under . Otherwise, is spanned by and equals . Clearly is orthogonal to . The statement follows by observing that is invariant under the endomorphism if and only if . ∎
The following lemma can be proven by standard computations in and is important for adapting frames on special Lagrangians in twistor spaces.
Lemma 3.4.
Let be the stabiliser group of in with Lie algebra . Then
Then is generated by and the element . In particular, is isomorphic to if , to if and to otherwise.
The action of on is a smooth cohomogeneity one action. The orbit at is diffeomorphic to and is singular for and of principal type otherwise. The principal orbits are diffeomorphic to . The orbit of is diffeomorphic to . Observe that is conjugated to the subgroup generated by
This subgroup is equal to the preimage of of the map
Here acts as the antipodal map on both and on . Hence, the orbit of is diffeomorphic to .
The following lemma summarises these observations.
Lemma 3.5.
The action of on is of cohomogeneity one. The principal orbit is diffeomorphic to , two singular orbits occur at and . The orbit is diffeomorphic to and that of to .
3.2 Adapting Frames
We now assume that is a nearly Kähler twistor space over a Riemannian four manifold . In other words, is either or the flag manifold . Lagrangian submanifolds of the latter have been studied in [Sto20a] so our interest is in in this chapter. Before using the explicit description of we give a few general statements that could be useful for generalisations to other spaces, such as non-nearly Kähler twistor spaces.
Given a special Lagrangian submanifold , we clearly have for . Since the frame bundle reduces to there is a map . Hence, can be understood as a map from to the interval and from to . We now apply our knowledge of the action of on to obtain a further frame reduction for special Lagrangian submanifolds in nearly Kähler twistor spaces. In that case, the holonomy of the nearly Kähler connection on reduces to , so reduces to an -bundle and we can assume . This means that there are two different reductions of : The first is to an -bundle , simply because itself reduces to an bundle. The second reduction is to an -bundle or equivalently by imposing as in section 2.
If is a rank one bundle, or equivalently , then the intersection is a bundle. We will derive its structure equations in section 3.3. If avoids the value then the intersection is trivial and which precludes the existence of a distinguished frame. However, by 3.4 we can apply a gauge transformation to guarantee a non-empty intersection.
For there is a frame in which maps to . Such a frame is unique up to the action of the stabiliser of in , which is computed in 3.4. This means that
(3.7) |
This is a principal bundle over with structure group given as in 3.4 if is either equal to or everywhere or if avoids these values altogether. In the latter case, the structure group is discrete. We first describe all special Lagrangians where is constant and equal to one of the boundary values everywhere. If then intersects every twistor fibre in a circle and maps to a surface in .
3.3 Lagrangians with
There is a general construction for Lagrangian submanifolds in the twistor space of an arbitrary Riemannian four-manifold due to Storm [Sto20] and Konstantinov [Kon17]. To make sense of how a Lagrangian submanifold in is defined, recall that carries two almost complex structures and metrics for . For a surface define the circle bundle with fibre over equal to . Geometrically, the fibre of at is the equator in each twistor fibre, which is diffeomorphic to , relative to the twistor lift of at . It turns out that this construction gives a lot of examples of Lagrangians in twistor spaces.
Proposition 3.6.
[Sto20] The submanifold is Lagrangian in for both and and every if is superminimal. Conversely, if is Lagrangian for any and , then is superminimal.
Assume is Lagrangian with so and are a rank two and a rank one bundle and So is also Lagrangian for and arises via the construction above. In this case the intersection is an bundle which is defined by imposing for on . Since the equation implies that lies in the span of and lie in the span of and . Since this implies that , i.e. takes values in when restricted to .
We can view locally as an orthonormal co-frame on , see 2.2. The forms vanish on the normal bundle while vanish on . The form is dual to the unit vector field tangent along the fibres of . Since this means that the fibres of are in fact geodesics. Since twistor fibres are totally geodesic these geodesics are great circles in the twistor fibres.
Since for this implies so the fundamental cubic is of the form
We have therefore shown.
Proposition 3.7.
The fundamental cubic of in a nearly Kähler twistor space either vanishes or has stabiliser .
We can also recover the result that is superminimal by showing that the second fundamental form in is complex-linear and using [MU97, Proposition 1c].
Remark 3.8.
Bryant considers special Lagrangians of the form . The cubic form of such submanifolds is always stabilised by . These examples are somewhat analogous to horizontal Lagrangians whose fundamental cubic also admits an symmetry.
From now on, we will work specifically with . We have seen that is either totally geodesic or its fundamental cubic has stabiliser . If is homogeneous then is a homogeneous superminimal surface in . Such a surface is equal to a totally geodesic or the Veronese curve in . Hence, there are only two different examples of homogeneous special Lagrangian submanifolds with . Both of them are known as Lagrangians for the Kähler structure on .
Example 3.9 (The standard ).
The standard is a totally geodesic special Lagrangian submanifold. It fibres over a totally geodesic in under the twistor fibration. It is the orbit of on .
The second example was discovered in [Chi04] and is described in [Kon17] in terms of the twistor fibration.
Example 3.10 (Chiang Lagrangian).
The subgroup of which comes from the irreducible representation of on has a special Lagrangian orbit at . This example is known as the Chiang Lagrangian and fibres over the Veronese surface in . The subgroup acts with stabiliser on . The stabiliser subgroup induces the full symmetry group of the fundamental cubic since the Chiang Lagrangian is not totally geodesic.
Since superminimal curves in have an explicit Weierstraß parametrisation one can produce many (explicit) examples of special Lagrangians in . However, our focus is on exploring special Lagrangians which do not arise from superminimal surfaces.
3.4 Changing the Gauge
If one expresses the nearly Kähler structure on in terms of local coordinates one can work out a system of PDE’s which, at least locally, describes special Lagrangian submanifolds. However, this approach is not very likely to succeed since local coordinates on are not an elegant way to define its nearly Kähler structure. Of more geometric importance are the first and second fundamental form and 2.1 shows that locally they contain all information about the submanifold. We use a gauge transformation, which depends on the function to describe the structure equations for a special Lagrangian in .
The bundle embeds into the frame bundle of via the adjoint action of on which factors through the double cover . So, on the level of structure equations we identify with . We apply the gauge transformation to , which defines the bundle as in eq. 3.7. This bundle has a reduced structure group, depending on the behaviour of the function , which is made precise in 3.4. For example, if avoids the values and then the structure group of is .
Recall from section 1.1 that is an principal bundle over . A local unitary frame for the nearly Kähler structure on is obtained by pulling back the forms , which are components of the Maurer-Cartan form on . We can realise the bundle by setting , where is defined in eq. 3.4, and imposing the equations
(3.8) |
Our aim is to compute the differentials of the one forms and also of and on the reduced bundle . We will achieve this by first computing the connection and curvature form in the transformed frame and then applying eq. 2.4-eq. 2.7. We begin by applying the transformation formula for a connection-one form under the gauge transformation
(3.9) |
Here is the connection form defined on , see eq. 1.3. Since lies in the torsion transforms trivially and we have by eq. 2.1. So in order to compute the differentials of we compute the transformed connection one-form from eq. 3.9. We split into real and imaginary part to get
(3.10) |
and
(3.11) |
To obtain expressions for the differentials of and we use the curvature equation 2.7. The curvature tensor transforms tensorially under gauge transformations yielding the explicit expressions
(3.15) | ||||
(3.19) |
Finally, combining the explicit expressions of with eq. 2.4-2.7 results in the following differential identities
(3.20) |
with and . Recall, that the equation implies more algebraic identities.
These differential identities are satisfied on any special Lagrangian in . Conversely, a special Lagrangian submanifold can locally be reconstructed from such a solution. There is little hope of working out all solutions of eq. 3.20. Instead, one typically imposes additional conditions and then tries to classify all special Lagrangians satisfying the extra condition. For example, one can already see that for the equations simplify considerably.
If everywhere there is a splitting where is the kernel of the projection . Recall that the standard complex structure on agrees with the nearly Kähler structure on and differs by a sign on .
Proposition 3.11.
The distribution is invariant under the standard complex structure on if and only if . In that case, is a CR submanifold for the Kähler structure on with being the -invariant distribution on .
Proof.
CR immersion from to the Kähler have been studied in [HYL18]. The splitting gives an ansatz for Lagrangians arising as a product such that . Indeed, we will give such an example for later. However, we first show that this ansatz fails when and is compact. Note that
and that is a multiple of , which vanishes on . This implies the following.
Lemma 3.12.
If is constant then defines a calibration on . The fibres of are the calibrated subspaces of .
Since is closed on it defines a cohomology class in . We have that when pulled back to , this class vanishes since . If takes values in the structure group of is just , generated by in which corresponds to . This element leaves and invariant so in particular reduces to a form on and we have shown.
Proposition 3.13.
If takes values in then . In particular, in this case does not have any compact two-dimensional submanifold which is tangent to .
3.5 Lagrangians with
The structure equations simplify significantly under the assumption . Indeed, plugging into eq. 3.20 yields and implies . Since we have Cartan’s lemma implies that there is a single function such that
The fundamental cubic is equal to which has stabiliser . Let , then
The structure equations are then equivalent to
(3.21) |
Hence, there are two examples of special Lagrangian submanifolds with , both of which are homogeneous and in particular compact. Neither of them is totally geodesic. Note that, as a subset of , the adapted frame bundle is defined by the equations
(3.22) |
They are both orbits of a Lie group with Lie algebra equal to the span of
Eq. 3.21 shows that is in fact constant and must be equal to either or . So, there are two distinct examples of special Lagrangians with . We describe the geometry of each of them.
Example 3.14.
There is a unique special Lagrangian with and . The reduced frame bundle over this submanifold is described by
These are the structure equations of where carries a metric of constant curvature .
Example 3.15.
There is a unique special Lagrangian with and . The reduced frame bundle over this submanifold is described by
which are the structure equations of a Berger sphere.
3.6 Lagrangians avoiding Boundary Values
From now on we assume that is not identically zero or and denote by the open set where avoids these values. On the frame bundle reduces to a discrete bundle. For -holomorphic curves in the second fundamental form is determined by two angle functions [Asl21]. In contrast, alone does not determine the second fundamental form of . We let and such that is entirely determined by and these quantities. Clearly all of these functions are constant on orbits of Lie subgroups of , the converse is also true.
Proposition 3.16.
Any solution with and constant is an orbit of a Lie group.
Proof.
Let be the special Lagrangian corresponding to this solution with adapted frame bundle . Then is an integral submanifold of the EDS generated by . By assumption, has constant coefficients which means that the equations describe a linear subspace of and hence is a Lie group and is a double cover of Lie groups. ∎
In principle, we could derive a set of PDE’s for and from the structure equations but this is not practical in full generality. However, there is a somewhat surprising homogeneous example.
Example 3.17.
Setting
is a solution to eq. 3.20 and hence corresponds to a unique special Lagrangian in . The fundamental cubic of this example is given by whose orientation preserving symmetry group is coming from . The Ricci curvature is diagonal in the (dual) frame in which
By 3.16, this example is homogeneous. We have found new examples by imposing conditions on . In each case, the fundamental cubic has non-trivial symmetries. The structure equations eq. 3.20 only hold in a fixed gauge. This makes it difficult to classify special Lagrangians where the fundamental cubic has a symmetry everywhere. We do not have the gauge freedom to bring them into the standard form as in [Bry06a, Proposition 1]. However, this poses no problem for the totally geodesic case.
Proposition 3.18.
Up to isometries, the standard is the unique totally geodesic special Lagrangian in .
Proof.
There is no totally geodesic Lagrangian which lies in . This is because in that case forces but this is a contradiction to the first equation of 3.20.
If is a totally geodesic Lagrangian with then the adapted frame bundle is a four-dimensional submanifold of on which and vanish. If then vanishes on the adapted bundle and by eq. 2.7 the ideal generated by and is closed under differentials. By Frobenius’ theorem, there is a unique maximal submanifold on which these forms vanish that passes through the identity . Hence, up to isometries, there is a unique totally geodesic special Lagrangian in with . We have already found this example, it is the standard . ∎
4 Classifying invariant Special Lagrangians
Instead of imposing symmetries on the fundamental cubic, we shall now impose them on the special Lagrangian itself. We have already encountered examples of homogoneous special Lagrangians.
There are examples of special Lagrangians admitting a cohomogeneity one action of in both and . In , there is a unique example of this type, the squashed three-sphere [Lot11, Example 6.4]. In , the Harvey-Lawson examples [Bry06a, HL82]
admit a cohomogeneity one action of for .
The situation in is different. We show in this section that all special Lagrangians that admit an action of an group of automorphism are in fact homogeneous and have already been described in the previous section. We introduce moment-type maps to prove this classification.
4.1 Moment Maps
Assume that acts effectively on with three-dimensional principal orbits and by nearly Kähler automorphisms. Let be a basis of such that . Denote the corresponding fundamental vector fields by . The map is an anti Lie algebra homomorphism. Hence, the vector fields obey the standard Pauli commutator relationships . Consider the map
Then is an equivariant map with respect to the action of on coming from the double cover . In addition, define the invariant scalar function
The map is not a multi-moment-type map in the sense of [MS13, Definition 3.5]. The Lie-kernel of is trivial, so there is no non-trivial multi-moment map for the three form . On the other hand, the map is trivial and is a multi-moment map with values in for . We will refer to and as multi-moment-type maps.
The general strategy to obtain moment-type maps is to contract Killing vector fields with the nearly Kähler forms. Using a standard argument, the following lemma shows that all such combinations are exhausted by and .
Lemma 4.1.
The form vanishes on orbits, i.e. .
Proof.
Let be a three-dimensional orbit of . Since acts by isometries on we have that is a invariant form on . The same holds for . So there is such that . Since is exact i.e. . ∎
Since is non-degenerate this means that vanishes if and only if are linearly dependent over . By Cartan’s formula and the nearly Kähler structure equations we get
(4.1) |
where is a cylic permutation of . The following proposition is somewhat similar to the toric situation [Dix19] as we can identify invariant special Lagrangians orbits by the values of the maps and .
Proposition 4.2.
The orbit of a point is special Lagrangian if and only if and . The set is a union of fixed points of the action and two-spheres on which vanishes. If has non-vanishing Euler characteristic then lies in the image of . The function is not constant and the set of points in which and consists of special Lagrangian orbits.
Proof.
By the definition of , the two-form vanishes on the orbit of if and only if . If then are linearly independent at and the orbit at is 3-dimensional, which implies the first statement. If then the orbit has dimension less than three and the second statement follows from the fact that lower-dimensional orbits must be points or two-spheres.
Eq. 4.1 implies that if is constant then are linearly dependent everywhere which contradicts the principal orbit type being three-dimensional. If then any vector field must have a zero, which forces to vanish. Finally, consider a point in which and . We want to show that . Using the action of we can assume that . Then . But and hence . ∎
Since either the maximum or minimum of is not zero this implies an existence result for special Lagrangians.
Corollary 4.3.
If is compact then the action has a special Lagrangian orbit.
If is a special Lagrangian submanifold on which a subgroup acts then will lie in the vanishing set of . So we can classify all invariant special Lagrangian submanifolds of by computing the vanishing set of for every subgroup of .
Definition 4.4.
Define the three subgroups of as , , arising from the inclusion , and which comes from the irreducible representation of on .
Any three-dimensional subgroup of is conjugate to one of .
Remark 4.5.
Note that contains two subgroups that do not stabilise a vector in . They are not conjugated to each other and, on the Lie algebra level, correspond to the splitting of into self-dual and anti-self-dual two forms. However, in , these two Lie algebras are conjugated to each other, for example via the element . Since the same holds true for the corresponding subgroups in .
The groups naturally act on through the double cover . The group acts via , the group via the double cover leaving a plane in invariant and acts irreducibly on and factors through . To relate the group invariant examples to those found in the previous section we compute the function for group orbits, for which we use eq. 3.5. To this end, it makes sense to define The Killing vector fields corresponding to the subgroups admit quite simple expressions in local coordinates. So, to express for in homogeneous coordinates we need to do so for the nearly Kähler form . This is the essence of eq. 1.4, where the forms are pulled back to a chart in by a local section.
It will be challenging to compute for and , so we first establish representation theoretic results to simplify the computations. In [GDV02], it is shown that given an irreducible finite-dimensional continuous real representation of a compact Lie group , the intersection of any hyperplane and any group orbit is non-empty. The authors of [GDV02] pose the question whether the same statement holds for complex representations, in particular irreducible representations of . There is a general framework to relate this question to the existence of nowhere vanishing sections in bundles over the flag manifold [AÐ10]. The following result follows a similar strategy and gives a direct proof for .
Lemma 4.6.
Let be a finite dimensional unitary representation of with all weights non-zero and be a hyperplane which is invariant under the maximal torus . Then intersects every orbit.
Proof.
Since is invariant there is a linear equivariant map such that . Assume that there is an such that . Then is a non-vanishing equivariant map . Restricting this map to gives a representation of on of weight .
Note that the principal bundle is the Hopf fibration and that gives rise to a nowhere vanishing section of the associated bundle over . Since the Hopf fibration has non-trivial Chern class, the complex line bundle is trivial which forces . This is a contradiction because restricts to an equivariant isomorphism from to , so is a zero-weight subspace. ∎
Note that, in the situation above, is invariant under and the action of on splits into one-dimensional components. Then every orbit also intersects the set where one of the components is restricted to the set .
All the actions of on factor through an action of on . The irreducible action of on has weights . The action of on factors through on and through on . In particular neither has a zero weight, so 4.6 applies to these cases.
4.1.1
Recall , we compute the Killing vector fields on the chart
We contract these vector fields with the nearly Kähler forms and in homogeneous coordinates from eq. 1.4, which gives
Hence, vanishes on the line of fixed points or when . Note that is the centraliser of in , acts with cohomogeneity one on and the orbits of that action are the level sets of . In particular acts transitively on which means that up to isometries there is a unique special Lagrangian on which acts. Hence, for simplification we consider the orbit at the point . At this point, annihilates which means that evaluating eq. 3.5 at yields
Hence and is diffeomorphic to . It is also the orbit of the larger group .
Lemma 4.7.
The unique special Lagrangian invariant under is which is identified with example 3.15.
Remark 4.8.
The multi-moment map for torus symmetry is an eigenfunction of the Laplace operator on , cf. [RS19, Lemma 3.1]. However, this is not the case for the multi-moment map , as it is non-positive everywhere, so , this integral vanishes for eigenfunctions of Laplace operator.
4.1.2
The group lies inside . Let , which commutes with all elements in the Lie algebra of . Again, we compute
For , the map is equal to
by eq. 1.4. We apply 4.6 and compute on the set and , and w.l.o.g we assume . Then we have
The set is a -holomorphic quadric and hence diffeomorphic to . The action of on this quadric is of cohomogeneity one. The principal orbit is and the singular orbit .
If then vanishes if and . Denote this point by , the orbit is special Lagrangian and is horizontal on . We compute via eq. 3.5
i.e. .
If then and only occurs for and . Denote this point by and note vanishes on . Hence, and the orbit is diffeomorphic to .
Lemma 4.9.
All special Lagrangians that admit a action are which corresponds to example 3.14 and example 3.9 respectively.
4.1.3
To compute the Killing vector fields for we need the explicit description of
see for example [Kaw18]. Now we can compute the Killing vector fields for on
Again, we apply 4.6 and restrict ourselves to compute and for and . Let furthermore , then by eq. 1.4
Hence, the only solutions of are . The solutions with are in the orbit of the point . The point is also in the same orbit as . So, it suffices to consider the points and .
Note that and which must hence be the minimum and maximum of respectively. The map vanishes at and hence the orbit satisfies and is in fact the Chiang Lagrangian.
Lemma 4.10.
All invariant special Lagrangians are given by the orbits and , which correspond to example 3.10 and example 3.17 respectively.
As remarked after 1.2, the identity component of nearly Kähler automorphisms of is . So, combining all results of this section results in the following theorem.
Theorem 4.11.
Every Special Lagrangian in that admits a non-trivial action of a three-dimensional group of nearly Kähler automorphisms is homogeneous and one of the following orbits.
Example
Properties
Group orbit
Stabiliser group of
3.15
Berger Sphere
0
3.14
0
3.9
standard
(tot. geodesic)
3.10
Chiang Lagrangian
3.17
distinct e’values
For the definition of the subgroups see 4.4.
4.2 The Flag Manifold
4.11 classifies homogeneous special Lagrangians and also rules out the existence of special Lagrangians admitting a cohomogeneity one action of a three-dimensional group of nearly Kähler automorphisms. The aim of this section is to prove the analogous statement for the nearly Kähler flag manifold . The homogeneous special Lagrangians in the flag manifold are classified in [Sto20a], so we restrict ourselves to the cohomogeneity-one case.
We could achieve this by computing the moment-type maps and determine the zero sets, as we did for . However, the statement can also be shown by analysing the group actions of 3-dimensional subgroups of , which is the identity component of nearly Kähler automorphisms as remarked after 1.2. We will show the set of elements with one-dimensional stabilisers are two-dimensional, so they cannot be special Lagrangian. To understand the action of three-dimensional subgroups of on the flag manifold we exploit the fact that the flag manifold is an adjoint orbit for .
Up to conjugation there are two three-dimensional subgroups of the standard and the subgroup fixing the element . Consider the adjoint action of on its lie algebra . Every element is then conjugate to a diagonal matrix, the orbits are distinguished by the set of purely imaginary eigenvalues.
Hence, the orbits are the level sets of the functions
There are three orbit types. The principal stabiliser type is a maximal torus in . Every element with distinct eigenvalues is of principal type. If has a repeated eigenvalue the stabiliser type is , unless all eigenvalues are zero.
So we fix an element with distinct eigenvalues and identify the flag manifold with the adjoint orbit of . Our aim is to determine the set of elements in with one-dimensional stabiliser under the action of and .
Proposition 4.12.
Every element in with non-principal stabiliser for either the action of or has a representative in the set
Proof.
With respect to the adjoint action on splits as . It is known that the action of on is irreducible and has trivial stabiliser unless the element in has repeated eigenvalues, in which case the stabiliser is . Every such element is conjugate to . Let , the stabiliser of in intersects the stabiliser of in a one-dimensional set if and only if is of the form
which implies the statement for .
With respect to the subgroup , the representation splits as where the action on the first summand is the adjoint action, is irreducible on the second summand and trivial on the third summand. For the stabiliser to be non-trivial the component in has to vanish. The trivial component is spanned by the element . Finally, every element in is conjugate to under the action. ∎
Theorem 4.13.
There are no special Lagrangians in which admits a cohomogeneity one action of nearly Kähler automorphisms.
Proof.
We show that for a three-dimensional group acting by nearly Kähler automorphisms on the flag manifold, the set of elements in with one-dimensional stabiliser is two-dimensional.
The identity component of nearly Kähler automorphisms of the flag manifold is . Since and are the only three-dimensional subgroups of it suffices to show that the intersection is finite. Since is two-dimensional it only remains to check that the function has full rank on . A direct computation shows that the determinant of the Jacobian is
which vanishes if and only if . In each case, the resulting element in has repeated eigenvalues, so it does not lie in since has distinct eigenvalues. ∎
References
- [AÐ10] Jinpeng An and Dragomir Ž Ðoković “Universal subspaces for compact Lie groups” In Journal für die reine und angewandte Mathematik 2010.647 De Gruyter, 2010, pp. 151–173
- [Asl21] Benjamin Aslan “Transverse -holomorphic curves in nearly Kaehler ” In Annals of Global Analysis and Geometry, https://doi.org/10.1007/s10455-021-09806-0 Springer, 2021
- [Asl22] Benjamin Aslan “Special submanifolds in nearly Kähler 6-manifolds”, 2022
- [Bek+19] Burcu Bektaş, Marilena Moruz, Joeri Van der Veken and Luc Vrancken “Lagrangian submanifolds of the nearly Kähler from minimal surfaces in ” In Proceedings of the Royal Society of Edinburgh Section A: Mathematics 149.3 Royal Society of Edinburgh Scotland Foundation, 2019, pp. 655–689
- [Bry06] Robert Bryant “On the geometry of almost complex 6-manifolds” In Asian Journal of Mathematics 10.3 International Press of Boston, 2006, pp. 561–605
- [Bry06a] Robert Bryant “Second order families of special Lagrangian 3-folds” In Perspectives in Riemannian Geometry, CRM Proc. Lecture Notes 40, 2006, pp. 63–98
- [But10] Jean-Baptiste Butruille “Homogeneous nearly Kähler manifolds” In Handbook of pseudo-Riemannian geometry and supersymmetry, 2010, pp. 399–423
- [Chi04] River Chiang “New Lagrangian submanifolds of ” In International Mathematics Research Notices 2004.45 Hindawi Publishing Corporation, 2004, pp. 2437–2441
- [CV15] Vincente Cortés and Jose Vásquez “Locally homogeneous nearly Kähler manifolds” In Annals of Global Analysis and Geometry 48.3 Springer, 2015, pp. 269–294
- [DC12] JC Dávila and F Martı́n Cabrera “Homogeneous nearly Kähler manifolds” In Annals of Global Analysis and Geometry 42.2 Springer, 2012, pp. 147–170
- [Dix19] Kael Dixon “The multi-moment map of the nearly Kähler ” In Geom Dedicata 200 Springer, 2019, pp. 351–362
- [FH17] Lorenzo Foscolo and Mark Haskins “New -holonomy cones and exotic nearly Kähler structures on and ” In Annals of Mathematics JSTOR, 2017, pp. 59–130
- [GDV02] Jorge Galindo, Pierre De La Harpe and Thierry Vust “Two observations on irreducible representations of groups.” In Journal of Lie Theory 12.2 Heldermann Verlag, 2002, pp. 535–538
- [Gra70] Alfred Gray “Nearly Kähler manifolds” In Journal of Differential Geometry 4 Lehigh University, 1970, pp. 283–309
- [HL82] Reese Harvey and Blaine Lawson “Calibrated geometries” In Acta Mathematica 148 Institut Mittag-Leffler, 1982, pp. 47–157
- [HYL18] Zejun Hu, Jiabin Yin and Zhenqi Li “Equivariant CR minimal immersions from into ” In Annals of Global Analysis and Geometry 54.1 Springer, 2018, pp. 1–24
- [Kaw17] Kotaro Kawai “Deformations of homogeneous associative submanifolds in nearly parallel -manifolds” In Asian Journal of Mathematics 21, 2017, pp. 429–462
- [Kaw18] Kotaro Kawai “Second-Order deformations of associative submanifolds in nearly parallel -manifolds” In The Quarterly Journal of Mathematics 69.1 Oxford University Press, 2018, pp. 241–270
- [Kon17] Momchil Konstantinov “Higher rank local systems in Lagrangian Floer theory” In arXiv preprint arXiv:1701.03624, 2017
- [Lot11] Jason Lotay “Ruled Lagrangian submanifolds of the 6-sphere” In Transactions of the American Mathematical Society 363.5, 2011, pp. 2305–2339
- [MNS08] Andrei Moroianu, Paul-Andi Nagy and Uwe Semmelmann “Deformations of nearly Kähler structures” In Pacific Journal of Mathematics 235.1 Mathematical Sciences Publishers, 2008, pp. 57–72
- [MS13] Thomas Bruun Madsen and Andrew Swann “Closed forms and multi-moment maps” In Geometriae Dedicata 165.1 Springer, 2013, pp. 25–52
- [MU97] Sebastián Montiel and Francisco Urbano “Second variation of superminimal surfaces into self-dual Einstein four-manifolds” In Transactions of the American Mathematical Society 349.6, 1997, pp. 2253–2269
- [Nag02] Paul-Andi Nagy “Nearly Kähler geometry and Riemannian foliations” In Asian J. Math. 6, 2002
- [RS19] Giovanni Russo and Andrew Swann “Nearly Kähler six-manifolds with two-torus symmetry” In Journal of Geometry and Physics 138 Elsevier, 2019, pp. 144–153
- [SS10] Lars Schäfer and Knut Smoczyk “Decomposition and minimality of Lagrangian submanifolds in nearly Kähler manifolds” In Annals of Global Analysis and Geometry 37.3 Springer, 2010, pp. 221–240
- [Sto20] Reinier Storm “A note on Lagrangian submanifolds of twistor spaces and their relation to superminimal surfaces” In Differential Geometry and its Applications 73 Elsevier, 2020, pp. 101669
- [Sto20a] Reinier Storm “Lagrangian submanifolds of the nearly Kähler full flag manifold ” In Journal of Geometry and Physics 158 Elsevier, 2020, pp. 103844
- [Vra03] Luc Vrancken “Special Lagrangian submanifolds of the nearly Kaehler 6-sphere” In Glasgow Mathematical Journal 45.3 Cambridge University Press, 2003, pp. 415–426
- [VS+19] Hông Vân Lê and Lorenz Schwachhöfer “Lagrangian submanifolds in strict nearly Kähler 6-manifolds” In Osaka Journal of Mathematics 56.3 Osaka UniversityOsaka City University, Departments of Mathematics, 2019, pp. 601–629
- [Xu10] Feng Xu “Pseudo-holomorphic curves in nearly Kähler ” In Differential Geometry and its Applications 28.1 Elsevier, 2010, pp. 107–120
- [Zha+16] Yinshan Zhang, Bart Dioos, Z Hu, Luc Vrancken and Xiafeng Wang “Lagrangian submanifolds in the 6-dimensional nearly Kähler manifolds with parallel second fundamental form” In Journal of Geometry and Physics 108 Elsevier, 2016, pp. 21–37
Department of Mathematics, University of West London, UK
E-mail address: benjamin.aslan.17@ucl.ac.uk