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

Special Lagrangians in nearly Kähler 3\mathbb{CP}^{3}

Benjamin Aslan
Abstract

This article explores special Lagrangian submanifolds in 3\mathbb{CP}^{3}, viewed as a nearly Kähler manifold, from two different perspectives. Intrinsically, using a moving frame set-up, and extrinsically, using SU(2)\mathrm{SU}(2) 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 3\mathbb{CP}^{3}, or the flag manifold 𝔽1,2(3)\mathbb{F}_{1,2}(\mathbb{C}^{3}) admitting a symmetry of an SU(2)\mathrm{SU}(2) 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 SU(3)\mathrm{SU}(3)-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 G2G_{2} manifolds, which makes nearly Kähler manifolds crucial for understanding G2G_{2} 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 JJ-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 G2G_{2}-cone of MM.

In the last few decades, many constructions for special Lagrangian submanifolds of S6S^{6} have been found and various subclasses of special Lagrangians have been classified, see for example [Vra03, Lot11]. More recently, the ambient spaces 𝔽\mathbb{F} and S3×S3S^{3}\times S^{3} have received attention, for example in [Bek+19, Sto20a]. This article is dedicated to the ambient nearly Kähler space M=3M=\mathbb{CP}^{3}. 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 3\mathbb{CP}^{3}.

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 3S4\mathbb{CP}^{3}\to S^{4}. We introduce an angle function θ:L[0,π4]\theta\colon L\to[0,\frac{\pi}{4}] parametrising the Lagrangian at a tangent level. Generically, LL intersects every twistor fibre transversally. The points where θ=π4\theta=\frac{\pi}{4} are those where this is not the case, and the intersection is then diffeomorphic to a circle. We identify Lagrangians with θπ4\theta\equiv\frac{\pi}{4} as circle bundles over superminimal surfaces in S4S^{4}, a construction discovered in [Sto20] and in [Kon17].

Finally, we classify all Lagrangians where θ\theta takes the boundary value 0. In fact, there are just two such examples and they are both homogeneous. We describe another somewhat surprising homogeneous example with θ12arccos(7255)\theta\equiv\frac{1}{2}\arccos(\frac{7\sqrt{2}}{5\sqrt{5}}) arising from the irreducible representation of SU(2)\mathrm{SU}(2) on S3(2)S^{3}(\mathbb{C}^{2}). We also show that the standard 3\mathbb{RP}^{3} in 3\mathbb{CP}^{3} is the only totally geodesic Lagrangian submanifold of 3\mathbb{CP}^{3}.

The second approach to exploring special Lagrangians in 3\mathbb{CP}^{3} is extrinsic. In section 4, we introduce SU(2)\mathrm{SU}(2) moment maps in nearly Kähler geometry. They encode the symmetry of the nearly Kähler manifold in a set of SU(2)\mathrm{SU}(2) equivariant functions M3M\to\mathbb{R}^{3}\oplus\mathbb{R}. We use these moment-type maps to show a general existence result of special Lagrangians with SU(2)\mathrm{SU}(2) symmetry, in 4.3 and to classify special Lagrangians admitting an action of a SU(2)\mathrm{SU}(2) 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 𝔽\mathbb{F} by studying the action of three-dimensional subgroups of SU(3)\mathrm{SU}(3) on 𝔽\mathbb{F}.

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 (M,g,J,ω)(M,g,J,\omega) be a 6-dimensional almost Hermitian manifold. Then MM is called nearly Kähler provided there is a complex-valued three-form ψ=Reψ+iImψΛ3,0(M)\psi=\mathop{\rm Re}\nolimits\psi+i\mathop{\rm Im}\nolimits\psi\in\Lambda^{3,0}(M) defining an SU(3)\mathrm{SU}(3)-structure satisfying

dω\displaystyle\mathrm{d}\omega =3Reψ\displaystyle=3\mathop{\rm Re}\nolimits\psi
dImψ\displaystyle\mathrm{d}\mathop{\rm Im}\nolimits\psi =2ωω.\displaystyle=-2\omega\wedge\omega.
Remark 1.1.

Often, the Calabi-Yau case dω=0,dImψ=0\mathrm{d}\omega=0,\mathrm{d}\mathop{\rm Im}\nolimits\psi=0 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 ¯\overline{\nabla} with totally skew-symmetric torsion and holonomy contained in SU(3)\mathrm{SU}(3), i.e. ¯g=¯J=¯ψ=0\overline{\nabla}g=\overline{\nabla}J=\overline{\nabla}\psi=0, cf. [Gra70]. This connection is called the characteristic connection and related to the Levi-Civita connection by

g(¯XY,Z)=g(XY,Z)+12Reψ(X,Y,JZ),\displaystyle g(\overline{\nabla}_{X}Y,Z)=g(\nabla_{X}Y,Z)+\frac{1}{2}\mathop{\rm Re}\nolimits\psi(X,Y,JZ), (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 M=G/HM=G/H is a homogeneous strictly nearly Kähler manifold of dimension six, then MM is one of the following: S6=G2/SU(3)S^{6}=G_{2}/{\mathrm{SU}(3)}, S3×S3=SU(2)3/ΔSU(2)S^{3}\times S^{3}=\mathrm{SU}(2)^{3}/\Delta{\mathrm{SU}(2)}, 3=Sp(2)/U(1)×Sp(1)\mathbb{CP}^{3}=\mathrm{Sp}(2)/{\mathrm{U}(1)\times\mathrm{Sp}(1)} or 𝔽=SU(3)/𝕋2\mathbb{F}=\mathrm{SU}(3)/{\mathbb{T}^{2}}.

In each case, the identity component of the group of nearly Kähler automorphisms is equal to GG, see [DC12]. There are infinitely many freely-acting finite subgroups of the automorphism group of the homogeneous nearly Kähler S3×S3S^{3}\times S^{3}, 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 S3×S3S^{3}\times S^{3} and S6S^{6} [FH17].

A 3-dimensional submanifold LL of a nearly Kähler manifold is called Lagrangian if ω|L=0\omega|_{L}=0. 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 ι:LM\iota\colon L\to M such that ιω=0\iota^{*}\omega=0. However, all examples we encounter are embedded submanifolds.

Because of the nearly Kähler identity dω=3Reψ\mathrm{d}\omega=3\mathop{\rm Re}\nolimits\psi, 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 LL in MM is minimal and orientable, see for example [VS+19].

Let IIM\mathrm{I\!I}_{M} be second fundamental form of LL in MM. Then the cubic form C(X,Y,Z)=ω(IIM(X,Y),Z)C(X,Y,Z)=\omega(\mathrm{I\!I}_{M}(X,Y),Z) is fully-symmetric, i.e. an element of Γ(S3(TL))\Gamma(S^{3}(T^{*}L)), and traceless when contracted in any two components, see [SS10]. The cubic form CC is also called the fundamental cubic of LL and takes values in the intrinsic bundle Γ(S3(TL))\Gamma(S^{3}(T^{*}L)). This means that in order to study special Lagrangians where CC satisfies special properties one does not need to specify the normal bundle of LL. One such special property would be that CC, 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 CC is that it admits symmetries. This approach has been developed in [Bry06a] for special Lagrangian submanifolds of 3\mathbb{C}^{3}. By picking a frame in a point xLx\in L one regards CC as a harmonic polynomial of degree three in three variables, i.e. an element of 3(3)\mathcal{H}_{3}(\mathbb{R}^{3}), which is a seven-dimensional vector space. The space 3(3)\mathcal{H}_{3}(\mathbb{R}^{3}) as an SO(3)\mathrm{SO}(3) module and a generic element in 3(3)\mathcal{H}_{3}(\mathbb{R}^{3}) does not have any symmetries in SO(3)\mathrm{SO}(3). 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 LL. This ansatz has led to the construction of new special Lagrangians in the Calabi-Yau 3\mathbb{C}^{3} in [Bry06a] and in the nearly Kähler S6S^{6} [Vra03, Lot11]. For 3\mathbb{CP}^{3} however, this ansatz is less fruitful since the curvature tensor is more complicated so we do not have SO(3)\mathrm{SO}(3)-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 S6S^{6} but it holds for any nearly Kähler manifold.

Proposition 1.3.

Every real analytic surface on which ω\omega vanishes can locally be uniquely thickened to a special Lagrangian submanifold in MM. 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 SU(3)\mathrm{SU}(3) structure. ∎

Infinitesimal deformations of nearly Kähler manifolds correspond to eigensections of a rotation operator on LL [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 3\mathbb{CP}^{3}

The nearly Kähler structure on 3\mathbb{CP}^{3} can be defined through the twistor fibration 3S4\mathbb{CP}^{3}\to S^{4}. The fibres are projective lines and totally geodesic for the Kähler structure on 3\mathbb{CP}^{3}. Since 3\mathbb{CP}^{3} is a sphere bundle inside Λ2(S4)\Lambda^{2}_{-}(S^{4}) the twistor fibration has a natural connection T3=𝒱T\mathbb{CP}^{3}=\mathcal{H}\oplus\mathcal{V}. The nearly Kähler structure on 3\mathbb{CP}^{3} 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 3=Sp(2)/S1×S3\mathbb{CP}^{3}=\mathrm{Sp}(2)/{S^{1}\times S^{3}}. Identify 2\mathbb{H}^{2} with 4\mathbb{C}^{4} via =j\mathbb{H}=\mathbb{C}\oplus j\mathbb{C}. This identification gives an action of Sp(2)\mathrm{Sp}(2) on 4\mathbb{C}^{4} which descends to 3\mathbb{CP}^{3} and acts transitively on that space. The stabiliser of the element (1,0,0,0)4(1,0,0,0)\in\mathbb{C}^{4} is

{(z00q)zS1,qS3}\left\{\begin{pmatrix}z&0\\ 0&q\end{pmatrix}\mid z\in S^{1}\subset\mathbb{C},\quad q\in S^{3}\subset\mathbb{H}\right\}

which shows 3=Sp(2)/S1×S3\mathbb{CP}^{3}=\mathrm{Sp}(2)/{S^{1}\times S^{3}}. Following [Xu10], consider the Maurer-Cartan form on Sp(2)\mathrm{Sp}(2) which can be written in components as

ΩMC=(iρ1+jω3¯ω1¯2+jω22ω12+jω22iρ2+jτ).\displaystyle\Omega_{MC}=\begin{pmatrix}i\rho_{1}+j\overline{\omega_{3}}&-\frac{\overline{\omega_{1}}}{\sqrt{2}}+j\frac{\omega_{2}}{\sqrt{2}}\\ \frac{\omega_{1}}{\sqrt{2}}+j\frac{\omega_{2}}{\sqrt{2}}&i\rho_{2}+j\tau\end{pmatrix}. (1.2)

Since ΩMC\Omega_{MC} has values in 𝔰𝔭(2)\mathfrak{sp}(2), the one-forms ω1,ω2,ω3\omega_{1},\omega_{2},\omega_{3} and τ\tau are complex-valued and ρ1,ρ2\rho_{1},\rho_{2} are real-valued. The equation dΩMC+[ΩMC,ΩMC]=0\mathrm{d}\Omega_{MC}+[\Omega_{MC},\Omega_{MC}]=0 implies the torsion identity

d(ω1ω2ω3)=(i(ρ2ρ1)τ¯0τi(ρ1+ρ2)0002iρ1)Aω:=(ω1ω2ω3)+(ω2¯ω3¯ω3¯ω1¯ω1¯ω2¯).\displaystyle\mathrm{d}\begin{pmatrix}\omega_{1}\\ \omega_{2}\\ \omega_{3}\end{pmatrix}=-\underbrace{\begin{pmatrix}i(\rho_{2}-\rho_{1})&-\overline{\tau}&0\\ \tau&-i(\rho_{1}+\rho_{2})&0\\ 0&0&2i\rho_{1}\end{pmatrix}}_{A_{\omega}:=}\wedge\begin{pmatrix}\omega_{1}\\ \omega_{2}\\ \omega_{3}\end{pmatrix}+\begin{pmatrix}\overline{\omega_{2}}\wedge\overline{\omega_{3}}\\ \overline{\omega_{3}}\wedge\overline{\omega_{1}}\\ \overline{\omega_{1}}\wedge\overline{\omega_{2}}\end{pmatrix}. (1.3)

and the curvature formula

dAω=AωAω+(ω1ω1¯ω3ω3¯ω1ω2¯0ω2ω1¯ω2ω2¯ω3ω3¯000ω1ω1¯ω2ω2¯+2ω3ω3¯).\mathrm{d}A_{\omega}=-A_{\omega}\wedge A_{\omega}+\begin{pmatrix}\omega_{1}\wedge\overline{\omega_{1}}-\omega_{3}\wedge\overline{\omega_{3}}&\omega_{1}\wedge\overline{\omega_{2}}&0\\ \omega_{2}\wedge\overline{\omega_{1}}&\omega_{2}\wedge\overline{\omega_{2}}-\omega_{3}\wedge\overline{\omega_{3}}&0\\ 0&0&-\omega_{1}\wedge\overline{\omega_{1}}-\omega_{2}\wedge\overline{\omega_{2}}+2\omega_{3}\wedge\overline{\omega_{3}}\end{pmatrix}.

The nearly Kähler structure on 3\mathbb{CP}^{3} is defined by declaring the forms sω1,sω2s^{*}\omega_{1},s^{*}\omega_{2} and sω3s^{*}\omega_{3} to be unitary (1,0)(1,0) forms for any local section ss of the bundle Sp(2)3\mathrm{Sp}(2)\to\mathbb{CP}^{3}. The resulting almost complex structure and metric do not depend on the choice of ss. The nearly Kähler forms ω,ψ\omega,\psi are pullbacks of

i2i=13ωiω¯i, and iω1ω2ω3,\displaystyle\frac{i}{2}\sum_{i=1}^{3}\omega_{i}\wedge\overline{\omega}_{i},\text{ and }-i\omega_{1}\wedge\omega_{2}\wedge\omega_{3},

respectively. In general, we will treat the nearly Kähler forms as basic forms on Sp(2)\mathrm{Sp}(2). However, Killing vector fields typically have a simple expression in local coordinates. To contract the nearly Kähler forms on 3\mathbb{CP}^{3} with Killing vector fields we pull back the local unitary (1,0)(1,0) forms ω1,ω2,ω3\omega_{1},\omega_{2},\omega_{3} on the chart 𝔸0={Z00}\mathbb{A}_{0}=\{Z_{0}\neq 0\} with the local section

s:𝔸0Sp(2),(1,Z1,Z2,Z3)(h1|Z|1h¯11h¯2ah2|Z|1a).s\colon\mathbb{A}_{0}\to\mathrm{Sp}(2),\quad(1,Z_{1},Z_{2},Z_{3})\mapsto\begin{pmatrix}h_{1}|Z|^{-1}&-\overline{h}_{1}^{-1}\overline{h}_{2}a\\ h_{2}|Z|^{-1}&a\end{pmatrix}.

Here,

|Z|2=1+|Z1|2+|Z2|2+|Z3|2,h1=1+jZ1,h2=Z2+jZ3,a=(1+|h2|2|h1|2)1/2.|Z|^{2}=1+|Z_{1}|^{2}+|Z_{2}|^{2}+|Z_{3}|^{2},\quad h_{1}=1+jZ_{1},\quad h_{2}=Z_{2}+jZ_{3},\quad a=(1+\frac{|h_{2}|^{2}}{|h_{1}|^{2}})^{-1/2}.

This gives the following expressions for the pull-backs

sω1=2|Z|2((Z3¯Z1¯Z2)dZ1+(1+|Z1|2)dZ2)sω2=2|Z|2((Z2¯Z1¯Z3)dZ1+(1+|Z1|2)dZ3)sω3=|Z|2(dZ1¯Z3¯dZ2¯+Z2¯dZ3¯).\displaystyle\begin{split}s^{*}\omega_{1}&=\sqrt{2}|Z|^{-2}((\overline{Z_{3}}-\overline{Z_{1}}Z_{2})\mathrm{d}Z_{1}+(1+|Z_{1}|^{2})\mathrm{d}Z_{2})\\ s^{*}\omega_{2}&=\sqrt{2}|Z|^{-2}((-\overline{Z_{2}}-\overline{Z_{1}}Z_{3})\mathrm{d}Z_{1}+(1+|Z_{1}|^{2})\mathrm{d}Z_{3})\\ s^{*}\omega_{3}&=|Z|^{-2}(\mathrm{d}\overline{Z_{1}}-\overline{Z_{3}}\mathrm{d}\overline{Z_{2}}+\overline{Z_{2}}\mathrm{d}\overline{Z_{3}}).\end{split} (1.4)

To show these formulae, note that the pullback of the Maurer-Cartan form via ss is

s(ΩMC)=(h¯1|Z|1h2¯|Z|1h2h11aa)(d(|Z|1h1)d(h¯11h¯2a)d(|Z|1h2)da).s^{*}(\Omega_{MC})=\begin{pmatrix}\overline{h}_{1}|Z|^{-1}&\overline{h_{2}}|Z|^{-1}\\ -h_{2}h_{1}^{-1}a&a\end{pmatrix}\begin{pmatrix}\mathrm{d}(|Z|^{-1}h_{1})&d(-\overline{h}_{1}^{-1}\overline{h}_{2}a)\\ \mathrm{d}(|Z|^{-1}h_{2})&\mathrm{d}a\end{pmatrix}.

Combining this with eq. 1.2 yields

(isρ1+jsω3¯)\displaystyle(is^{*}\rho_{1}+js^{*}\overline{\omega_{3}}) =|Z|2(h1¯dh1+h2¯dh2)+R\displaystyle=|Z|^{-2}(\overline{h_{1}}\mathrm{d}h_{1}+\overline{h_{2}}\mathrm{d}h_{2})+R
12a|Z|(sω1+jsω2)\displaystyle\frac{1}{\sqrt{2}a|Z|}(s^{*}\omega_{1}+js^{*}\omega_{2}) =h2h11dh1+dh2,\displaystyle=-h_{2}h_{1}^{-1}\mathrm{d}h_{1}+\mathrm{d}h_{2},

where RR is a real term. Equations 1.4 follow by splitting the quaternionic-valued differential forms on the right-hand side into their \mathbb{C} and jj\mathbb{C} part.

2 Structure Equations for Special Lagrangians

The structure equations for a special Lagrangian manifold in Calabi-Yau 3\mathbb{C}^{3} were established in [Bry06a] and for nearly Kähler S6S^{6} 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 {1,2,3}\{1,2,3\} is implicit.

Let M6M^{6} be a nearly Kähler manifold and consider the SU(3)\mathrm{SU}(3)-frame bundle PSU(3)P_{\mathrm{SU}(3)}. Let (ζ1,ζ2,ζ3)Ω1(PSU(3),3)(\zeta_{1},\zeta_{2},\zeta_{3})\in\Omega^{1}(P_{\mathrm{SU}(3)},\mathbb{C}^{3}) be the tautological one-forms on PSU(3)P_{\mathrm{SU}(3)} and let ϕΩ1(P,𝔰𝔲(3))\phi\in\Omega^{1}(P,\mathfrak{su}(3)) be the nearly Kähler connection one-form on PSU(3)P_{\mathrm{SU}(3)}, giving the torsion relation

d(ζ1ζ2ζ3)=ϕ(ζ1ζ2ζ3)+(ζ¯2ζ¯3ζ¯3ζ¯1ζ¯1ζ¯2)\displaystyle\mathrm{d}\begin{pmatrix}\zeta_{1}\\ \zeta_{2}\\ \zeta_{3}\end{pmatrix}=-\phi\wedge\begin{pmatrix}\zeta_{1}\\ \zeta_{2}\\ \zeta_{3}\end{pmatrix}+\begin{pmatrix}\overline{\zeta}_{2}\wedge\overline{\zeta}_{3}\\ \overline{\zeta}_{3}\wedge\overline{\zeta}_{1}\\ \overline{\zeta}_{1}\wedge\overline{\zeta}_{2}\end{pmatrix} (2.1)

and the curvature identity

dϕij=ϕikϕkj+Kijpqζqζp¯.\displaystyle\mathrm{d}\phi_{ij}=-\phi_{ik}\wedge\phi_{kj}+K_{ijpq}\zeta_{q}\wedge\overline{\zeta_{p}}. (2.2)

In particular, the curvature of ¯\overline{\nabla} is always of type (1,1)(1,1). In [Bry06] it is remarked that the tensor KK can be written as sum

Kijpq=Kijpq+34δpiδqj14δijδpqK_{ijpq}=K^{\prime}_{ijpq}+\frac{3}{4}\delta_{pi}\delta_{qj}-\frac{1}{4}\delta_{ij}\delta_{pq}

where KK^{\prime} has the following symmetries

Kijpq=Kpjiq=Kiqpj=Kjiqp¯, and iKiipq=0.K^{\prime}_{ijpq}=K^{\prime}_{pjiq}=K^{\prime}_{iqpj}=\overline{K^{\prime}_{jiqp}},\text{ and }\sum_{i}K^{\prime}_{iipq}=0.

The tensor KK^{\prime} vanishes exactly when MM is the round six-sphere. The nearly Kähler forms are expressed in terms of ζi\zeta_{i} by

ω=i2iζiζi¯,ψ=iζ1ζ2ζ3.\displaystyle\omega=\frac{i}{2}\sum_{i}\zeta_{i}\wedge\overline{\zeta_{i}},\quad\psi=-i\zeta_{1}\wedge\zeta_{2}\wedge\zeta_{3}. (2.3)

Note the difference from [Bry06] in the convention for ψ\psi 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 PSU(3)P_{\mathrm{SU}(3)}. If LL is a special Lagrangian submanifold in MM then one obtains more differential identities because the frame bundle PSU(3)P_{\mathrm{SU}(3)} admits a natural reduction to an SO(3)\mathrm{SO}(3) bundle over LL. The reason for this is that, at the tangent level, a Lagrangian subspace looks like 3\mathbb{R}^{3} in 3\mathbb{C}^{3}, which defines the restriction

PSO(3)={p:3TM,pPSU(3)|Lp(3)=TL}.P_{\mathrm{SO}(3)}=\{p\colon\mathbb{C}^{3}\to TM,\quad p\in P_{\mathrm{SU}(3)}|_{L}\mid p(\mathbb{R}^{3})=TL\}.

If dz1,dz2,dz3\mathrm{d}z_{1},\mathrm{d}z_{2},\mathrm{d}z_{3} are the standard complex-valued one-forms on 3\mathbb{C}^{3} then 33\mathbb{R}^{3}\subset\mathbb{C}^{3} is characterised as the 3-dimensional subspace of 3\mathbb{C}^{3} on which the imaginary parts of dzi\mathrm{d}z_{i} vanish. Similarly, our aim is to describe the reduction PSO(3)P_{\mathrm{SO}(3)} as the vanishing set of one forms on PSU(3)P_{\mathrm{SU}(3)}. To that end, split the forms ζi=σi+iηi\zeta_{i}=\sigma_{i}+i\eta_{i} and ϕ=α+iβ\phi=\alpha+i\beta into real and imaginary part. The bundle PSO(3)P_{\mathrm{SO}(3)} is now defined by imposing the condition ηi=0\eta_{i}=0.

This characterisation implies more differential identities. From the torsion-relation we get

dσi\displaystyle\mathrm{d}\sigma_{i} =αijσj+βijηj+σkσlηkηl\displaystyle=-\alpha_{ij}\wedge\sigma_{j}+\beta_{ij}\wedge\eta_{j}+\sigma_{k}\wedge\sigma_{l}-\eta_{k}\wedge\eta_{l}
dηi\displaystyle\mathrm{d}\eta_{i} =βijσjαijηjσkηlηkσl\displaystyle=-\beta_{ij}\wedge\sigma_{j}-\alpha_{ij}\wedge\eta_{j}-\sigma_{k}\wedge\eta_{l}-\eta_{k}\wedge\sigma_{l}

where (i,k,l)(i,k,l) is an cyclic permutation of (1,2,3)(1,2,3). The condition ηi=0\eta_{i}=0 implies βijσj=0\beta_{ij}\wedge\sigma_{j}=0. By Cartan’s lemma, we have βij=hijkσk\beta_{ij}=h_{ijk}\sigma_{k} or β=hσ\beta=h\sigma where hh 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 3\mathbb{C}^{3} or in S6S^{6}.

On the reduced bundle, we split KK into real and imaginary part,

Kijpqζqζp¯=Kijpqσqσp=(Rijpq+iSijpq)σqσp=(RijpqiSijpq)σpσq.K_{ijpq}\zeta_{q}\wedge\overline{\zeta_{p}}=K_{ijpq}\sigma_{q}\wedge\sigma_{p}=(R_{ijpq}+iS_{ijpq})\sigma_{q}\wedge\sigma_{p}=(-R_{ijpq}-iS_{ijpq})\sigma_{p}\wedge\sigma_{q}.

This also allows us to split the curvature identity into real imaginary part

dαij\displaystyle\mathrm{d}\alpha_{ij} =αikαkj+βikβkjRijpqσpσq\displaystyle=-\alpha_{ik}\wedge\alpha_{kj}+\beta_{ik}\wedge\beta_{kj}-R_{ijpq}\sigma_{p}\wedge\sigma_{q}
dβij\displaystyle\mathrm{d}\beta_{ij} =βikαkjαikβkjSijpqσpσq.\displaystyle=-\beta_{ik}\wedge\alpha_{kj}-\alpha_{ik}\wedge\beta_{kj}-S_{ijpq}\sigma_{p}\wedge\sigma_{q}.

To write these equations more invariantly, let

[σ]=(0σ3σ2σ30σ1σ2σ10).[\sigma]=\begin{pmatrix}0&\sigma_{3}&-\sigma_{2}\\ -\sigma_{3}&0&\sigma_{1}\\ \sigma_{2}&-\sigma_{1}&0\end{pmatrix}.

We can summarise the equations on the reduced bundle over LL in tensor notation

βσ\displaystyle\beta\wedge\sigma =0\displaystyle=0 (2.4)
dσ\displaystyle\mathrm{d}\sigma =ασ12[σ]σ\displaystyle=-\alpha\wedge\sigma-\frac{1}{2}[\sigma]\wedge\sigma (2.5)
dα\displaystyle\mathrm{d}\alpha =αα+ββRσσ\displaystyle=-\alpha\wedge\alpha+\beta\wedge\beta-R\sigma\wedge\sigma (2.6)
dβ\displaystyle\mathrm{d}\beta =βααβSσσ\displaystyle=-\beta\wedge\alpha-\alpha\wedge\beta-S\sigma\wedge\sigma (2.7)

where (σσ)pq=σpσq(\sigma\wedge\sigma)_{pq}=\sigma_{p}\wedge\sigma_{q}. The matrix of one forms β\beta is completely defined by the symmetric tensor hh. The advantage to work with hh is that its components are not one-forms but functions, allowing us to rewrite eq. 2.4, eq. 2.6 and eq. 2.7

β=hσdα=αα+hσhσ+34σσRσσ,0=(dh+((hα+12h[σ]))+Sσ)σ.\displaystyle\beta=h\sigma\quad\mathrm{d}\alpha=-\alpha\wedge\alpha+h\sigma\wedge h\sigma+\frac{3}{4}\sigma\wedge\sigma-R\sigma\wedge\sigma,\quad 0=(\mathrm{d}h+((h\alpha+\frac{1}{2}h[\sigma]))+S\sigma)\wedge\sigma.

The Levi-Civita connection one-form of the induced metric on LL is α+12[σ]\alpha+\frac{1}{2}[\sigma]. Note that the forms σ\sigma differ by a factor 2 from the orthonormal one forms considered in [Lot11].

If M=G/HM=G/H 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 𝔤=𝔥𝔪\mathfrak{g}=\mathfrak{h}\oplus\mathfrak{m} such that AdH(𝔪)𝔪\mathrm{Ad}_{H}(\mathfrak{m})\subset\mathfrak{m}. The nearly Kähler structure then yields an Ad(H)\mathrm{Ad}(H) invariant special unitary basis ω1,ω2,ω3\omega_{1},\omega_{2},\omega_{3} on 𝔪3\mathfrak{m}\cong\mathbb{C}^{3}. Up to a cover, GG embeds into the SU(3)\mathrm{SU}(3)-frame bundle PSU(3)P_{\mathrm{SU}(3)} via the adjoint action HSU(𝔪)H\to\mathrm{SU}(\mathfrak{m}). Under this identification

ψ+(ζ1,ζ2,ζ3)𝔥3𝔥𝔪\psi+(\zeta_{1},\zeta_{2},\zeta_{3})\in\mathfrak{h}\oplus\mathbb{C}^{3}\cong\mathfrak{h}\oplus\mathfrak{m}

is the Maurer-Cartan form ωG\omega_{G} on GG. In other words, the nearly Kähler connection is equal to the canonical homogeneous connection on GMG\to M, 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 α\alpha and β\beta determine the first and second fundamental form, this can be viewed a Bonnet-type theorem.

Proposition 2.1.

Let M=G/HM=G/H be a homogeneous nearly Kähler manifold, L3L^{3} be a simply-connected three manifold and σΩ1(L,3)\sigma\in\Omega^{1}(L,\mathbb{R}^{3}), defining a linearly independent co-frame at each point, αΩ1(L,𝔰𝔬(3))\alpha\in\Omega^{1}(L,\mathfrak{so}(3)) and βΩ1(L,S2(3))\beta\in\Omega^{1}(L,S^{2}(\mathbb{R}^{3})) satisfying the equations 2.4-2.7. Then there is a special Lagrangian immersion LML\to M, unique up to isometries, with α,β\alpha,\beta determining the metric and second fundamental form of LL in MM.

Proof.

Define the form γ=α+iβ+(σ1,σ2,σ3)𝔥𝔪𝔤\gamma=\alpha+i\beta+(\sigma_{1},\sigma_{2},\sigma_{3})\in\mathfrak{h}\oplus\mathfrak{m}\cong\mathfrak{g}. Since σ,α,β\sigma,\alpha,\beta satisfy the equations 2.4-2.7 we have dγ+[γ,γ]=0\mathrm{d}\gamma+[\gamma,\gamma]=0. The statement now follows from Cartan’s theorem, just as the classical Bonnet theorem for surfaces in 3\mathbb{R}^{3}. ∎

Remark 2.2.

Note that the tautological one form (ζ1,ζ2,ζ3)(\zeta_{1},\zeta_{2},\zeta_{3}) can also be regarded as an element in Γ(P,End(3))\Gamma(P,\mathrm{End}(\mathbb{C}^{3})). With this identification, a local section ss of LUPSO(3)L\supset U\to P_{\mathrm{SO}(3)} gives a section Γ(U,(TM)|L3)Ω1(U,3)\Gamma(U,(T^{\vee}M)|_{L}\otimes\mathbb{C}^{3})\cong\Omega^{1}(U,\mathbb{C}^{3}). Then sηis^{*}\eta_{i} vanishes on TLTL while sσs^{*}\sigma vanishes on the normal bundle.

3 An Angle Function for Special Lagrangians

Since twistor fibres are JJ-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 S4S^{4}. We review this construction and define an angle function L[0,π4]L\to[0,\frac{\pi}{4}] which has value π4\frac{\pi}{4} if LL intersects a twistor fibre non-transversally. We use a gauge transformation, which depends on θ\theta, to use the moving frame setup from the previous section for special Lagrangians in 3\mathbb{CP}^{3}. 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 n\mathbb{C}^{n} is identified with the homogeneous space SU(n)/SO(n)\mathrm{SU}(n)/{\mathrm{SO}(n)}. Twistor nearly Kähler spaces have the property that the holonomy of the nearly Kähler connection reduces to U(2)S(U(2)×U(1))\mathrm{U}(2)\cong\mathrm{S}(\mathrm{U}(2)\times\mathrm{U}(1)). The two-form splits into a horizontal and vertical part ω=ω+ω𝒱\omega=\omega_{\mathcal{H}}+\omega_{\mathcal{V}}. 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 (b1,b2,b3)(b_{1},b_{2},b_{3}) denote the standard basis of 3\mathbb{C}^{3} with dual basis (ω1,ω2,ω3)(\omega_{1},\omega_{2},\omega_{3}) and let ω=i2(ω1ω¯1+ω2ω¯2)\omega_{\mathcal{H}}=\frac{i}{2}(\omega_{1}\wedge\bar{\omega}_{1}+\omega_{2}\wedge\bar{\omega}_{2}) as well as ω𝒱=i2(ω3ω¯3)\omega_{\mathcal{V}}=\frac{i}{2}(\omega_{3}\wedge\bar{\omega}_{3}). Let HS(U(2)×U(1))H\cong\mathrm{S}(\mathrm{U}(2)\times\mathrm{U}(1)) be the stabiliser of ω𝒱\omega_{\mathcal{V}} inside SU(3)\mathrm{SU}(3). Let also ψ=Reψ+iImψ\psi=\mathop{\rm Re}\nolimits\psi+i\mathop{\rm Im}\nolimits\psi be the complex-valued three form iω1ω2ω3-i\omega_{1}\wedge\omega_{2}\wedge\omega_{3} on 3\mathbb{C}^{3}. We have abused notation slightly here, since ω,ψ\omega,\psi are forms on the nearly Kähler manifold but also denote their linear models on 3\mathbb{C}^{3}.

For a complex subspace W3W\subset\mathbb{C}^{3} denote by SLag(W)\mathrm{SLag}(W) the set of all special Lagrangian subspaces of WW. By 23\mathbb{C}^{2}\subset\mathbb{C}^{3} we refer to the subspace spanned by b2b_{2} and b3b_{3}. Note that SLag(2)S2\mathrm{SLag}(\mathbb{C}^{2})\cong S^{2} and that U(1)SU(2)\mathrm{U}(1)\subset\mathrm{SU}(2) 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 U(1){diag(eiφ,eiφ)}SU(2)\mathrm{U}(1)\cong\{\mathrm{diag}(e^{i\varphi},e^{-i\varphi})\}\subset\mathrm{SU}(2) any element in SLag(2)\mathrm{SLag}(\mathbb{C}^{2}) has a unique representative of the form Vθ=span(ieiθb2eiθb3,eiθb2+ieiθb3)V_{\theta}=\mathrm{span}(-ie^{-i\theta}b_{2}-e^{-i\theta}b_{3},e^{i\theta}b_{2}+ie^{i\theta}b_{3}), for 0θπ/20\leq\theta\leq\pi/2.

Proof.

Special Lagrangian planes in 2\mathbb{C}^{2} are parametrised by SU(2)/SO(2)\mathrm{SU}(2)/{\mathrm{SO}(2)}. Thus, we have to find a unique representative of the action (A,B)X=AXB1(A,B)X=AXB^{-1} of K=U(1)×SO(2)K=\mathrm{U}(1)\times\mathrm{SO}(2) on SU(2)\mathrm{SU}(2), which is the action of a maximal torus in SO(4)\mathrm{SO}(4) acting on S3S^{3}. The standard torus U(1)×U(1)U(2)SO(4)\mathrm{U}(1)\times\mathrm{U}(1)\subset\mathrm{U}(2)\subset\mathrm{SO}(4) acting on S3S^{3} admits unique representatives of the form (cos(θ),0,sin(θ),0)(\cos(\theta),0,\sin(\theta),0) for 0θπ/20\leq\theta\leq\pi/2. The statement follows by conjugating the action of KK to the standard torus action. ∎

For any subspace W3W\subset\mathbb{C}^{3} denote by KWK_{W} the kernel of the projection onto span(b3)\mathrm{span}(b_{3}) and by nWn_{W} its dimension. Let

Tθ=(1000ieiθ2eiθ20eiθ2ieiθ2)\displaystyle T_{\theta}=\left(\begin{array}[]{ccc}1&0&0\\ 0&-\frac{ie^{-i\theta}}{\sqrt{2}}&\frac{e^{i\theta}}{\sqrt{2}}\\ 0&-\frac{e^{-i\theta}}{\sqrt{2}}&\frac{ie^{i\theta}}{\sqrt{2}}\\ \end{array}\right) (3.4)

and WθW_{\theta} be the image of TθT_{\theta} when applied to the standard 3\mathbb{R}^{3} in 3\mathbb{C}^{3}, i.e. Wθ=span(b1,ieiθb2eiθb3,eiθb2+ieiθb3)W_{\theta}=\mathrm{span}(b_{1},-ie^{-i\theta}b_{2}-e^{-i\theta}b_{3},e^{i\theta}b_{2}+ie^{i\theta}b_{3}).

Proposition 3.2.

Any special Lagrangian subspace W3W\subset\mathbb{C}^{3} admits a unique representative WθW_{\theta} for 0θπ/40\leq\theta\leq\pi/4, under the action of HH. Furthermore, nW=2n_{W}=2 if and only if θ=π/4\theta=\pi/4.

Proof.

Since WW is Lagrangian, nW1n_{W}\geq 1. If nW=2n_{W}=2 then WW is represented by the standard 3\mathbb{R}^{3} in 3\mathbb{C}^{3} and nWθ=2n_{W_{\theta}}=2 if and only if θ=π/4\theta=\pi/4. So from now on we assume that nW=1n_{W}=1. Consider the map l:Gr2(2)Gr3(3),Vspan(b1,V)l\colon\mathrm{Gr}_{2}(\mathbb{C}^{2})\to\mathrm{Gr}_{3}(\mathbb{C}^{3}),\quad V\mapsto\mathrm{span}(b_{1},V). Note that Wθ=l(Vθ)W_{\theta}=l(V_{\theta}) and that ll descends to a map l^:Lag(2)/U(1)Lag(3)/H.\hat{l}\colon\mathrm{Lag}(\mathbb{C}^{2})/{\mathrm{U}(1)}\to\mathrm{Lag}(\mathbb{C}^{3})/H. To show surjectivity observe that for WLag(3)W\in\mathrm{Lag}(\mathbb{C}^{3}) we have KWspan(b1,b2)K_{W}\subset\mathrm{span}(b_{1},b_{2}). So by acting with HH we can achieve that KWK_{W} is spanned by b1b_{1}. Furthermore, observe that

(100010001)Tθ(100001010)=Tπ/2θ\begin{pmatrix}-1&0&0\\ 0&-1&0\\ 0&0&1\end{pmatrix}T_{\theta}\begin{pmatrix}-1&0&0\\ 0&0&1\\ 0&1&0\end{pmatrix}=T_{\pi/2-\theta}

which means that Wθ=WθW_{\theta}=W_{\theta^{\prime}} for θ+θ=π/2\theta+\theta^{\prime}=\pi/2. We have shown that any element in Lag(3)\mathrm{Lag}(\mathbb{C}^{3}) is represented by a WθW_{\theta} for 0θπ/40\leq\theta\leq\pi/4. The uniqueness follows from the observation that ω𝒱\omega_{\mathcal{V}} has norm 12|cos(2θ)|\frac{1}{2}|\mathrm{cos}(2\theta)| when restricted to the vector space WθW_{\theta}. ∎

If w1,w2,w3w_{1},w_{2},w_{3} is a basis of WW such that w1KWw_{1}\in K_{W} and w2,w2KWw_{2},w_{2}\in K_{W}^{\perp} then θ\theta can be computed by the formula

12w1|cos(2θ)|ψ(w1,w2,w3)=ω𝒱(w2,w3).\displaystyle\frac{1}{2\|w_{1}\|}|\mathrm{cos}(2\theta)|\psi^{-}(w_{1},w_{2},w_{3})=\omega_{\mathcal{V}}(w_{2},w_{3}). (3.5)

3.2 gives a geometric interpretation of the boundary value π/4\pi/4. In 3.11 we relate the case θ=0\theta=0 to CR-manifolds in the Kähler 3\mathbb{CP}^{3}, so we study this case on the linear level first. Motivated by the existence of the two almost complex structures J1J_{1} and J2J_{2} on the twistor space, consider the almost complex structure

J:(b1,b2,b3)(ib1,ib2,ib3)\displaystyle J^{\prime}\colon(b_{1},b_{2},b_{3})\mapsto(ib_{1},ib_{2},-ib_{3}) (3.6)

on 3\mathbb{C}^{3}. Any special Lagrangian subspace WW in 3\mathbb{C}^{3} splits as KWKWK_{W}\oplus K_{W}^{\perp}.

Lemma 3.3.

If θπ4\theta\neq\frac{\pi}{4} then J(KW)J(K_{W}) is orthogonal to WW. The subspace KWK_{W}^{\perp} is invariant under JJ^{\prime} if and only if θ=0\theta=0.

Proof.

The endomorphism JJ^{\prime} commutes with the action of HH on 3\mathbb{C}^{3}, so it suffices to prove the statement for WθW_{\theta}. If θ=π4\theta=\frac{\pi}{4} then KWK_{W}^{\perp} is one-dimensional so it cannot be invariant under JJ^{\prime}. Otherwise, KWK_{W} is spanned by b1b_{1} and KWK_{W}^{\perp} equals VθV_{\theta}. Clearly Jb1=ib1Jb_{1}=ib_{1} is orthogonal to WW. The statement follows by observing that VθV_{\theta} is invariant under the endomorphism (b2,b3)(ib2,ib3)(b_{2},b_{3})\mapsto(ib_{2},-ib_{3}) if and only if θ=0\theta=0. ∎

The following lemma can be proven by standard computations in SU(3)\mathrm{SU}(3) and is important for adapting frames on special Lagrangians in twistor spaces.

Lemma 3.4.

Let Hθ=Tθ1HTθSO(3)H_{\theta}=T_{\theta}^{-1}HT_{\theta}\cap\mathrm{SO}(3) be the stabiliser group of WθW_{\theta} in HH with Lie algebra 𝔥θ\mathfrak{h}_{\theta}. Then

𝔥θ={(011100100)θ=π/4{0}𝔰𝔬(2)θ=0{0}otherwiseTθ𝔥θTθ1={(01+i01+i00000)θ=π/4{0}𝔰(𝔲(1)𝔲(1))θ=0{0}otherwise\displaystyle\mathfrak{h}_{\theta}=\begin{cases}\mathbb{R}\cdot\begin{pmatrix}0&1&-1\\ -1&0&0\\ 1&0&0\end{pmatrix}&\theta=\pi/4\\ \{0\}\oplus\mathfrak{so}(2)&\theta=0\\ \{0\}&\text{otherwise}\end{cases}\qquad T_{\theta}\mathfrak{h}_{\theta}T_{\theta}^{-1}=\begin{cases}\mathbb{R}\cdot\begin{pmatrix}0&-1+i&0\\ 1+i&0&0\\ 0&0&0\end{pmatrix}&\theta=\pi/4\\ \{0\}\oplus\mathfrak{s(u(1)\oplus u(1))}&\theta=0\\ \{0\}&\text{otherwise}\end{cases}

Then HθH_{\theta} is generated by exp(𝔥θ)\exp(\mathfrak{h}_{\theta}) and the element diag(1,1,1)\mathrm{diag}(1,-1,-1). In particular, HθH_{\theta} is isomorphic to O(2)\mathrm{O}(2) if θ=π/4\theta=\pi/4, to SO(2)\mathrm{SO}(2) if θ=0\theta=0 and to 2\mathbb{Z}_{2} otherwise.

The action of HH on Lag(3)\mathrm{Lag}(\mathbb{C}^{3}) is a smooth cohomogeneity one action. The orbit at WθW_{\theta} is diffeomorphic to H/(TθHθTθ1)H/({T_{\theta}H_{\theta}T_{\theta}^{-1}}) and is singular for θ=0,π/4\theta=0,\pi/4 and of principal type otherwise. The principal orbits are diffeomorphic to H/diag(1,1,1)U(2)/diag(1,1)H/{\langle\mathrm{diag}(1,-1,-1)\rangle}\cong\mathrm{U}(2)/{\langle\mathrm{diag}(1,-1)\rangle}. The orbit of W0W_{0} is diffeomorphic to H/({1}×S(U(1)×U(1)))U(2)/({1}×U(1))S3H/{(\{1\}\times S(\mathrm{U}(1)\times\mathrm{U}(1)))}\cong\mathrm{U}(2)/{(\{1\}\times\mathrm{U}(1))}\cong S^{3}. Observe that Tπ/4Hπ/4Tπ/41{T_{\pi/4}H_{\pi/4}T_{\pi/4}^{-1}} is conjugated to the O(2)\mathrm{O}(2) subgroup generated by

S(U(1)×U(1)) and (0110).S(\mathrm{U}(1)\times\mathrm{U}(1))\text{ and }\begin{pmatrix}0&-1\\ -1&0\end{pmatrix}.

This subgroup is equal to the preimage of [([1,0],1)[([1,0],1) of the map

U(2)(1×S1)/2,A[[A(1,0)T],det(A)].\mathrm{U}(2)\to(\mathbb{CP}^{1}\times S^{1})/{\mathbb{Z}_{2}},\quad A\mapsto[[A(1,0)^{T}],\mathrm{det}(A)].

Here 2\mathbb{Z}_{2} acts as the antipodal map on both 1S2\mathbb{CP}^{1}\cong S^{2} and on S1S^{1}. Hence, the orbit of Wπ/4W_{\pi/4} is diffeomorphic to (S2×S1)/2(S^{2}\times S^{1})/\mathbb{Z}_{2}.

The following lemma summarises these observations.

Lemma 3.5.

The action of HH on Lag(3)\mathrm{Lag}(\mathbb{C}^{3}) is of cohomogeneity one. The principal orbit is diffeomorphic to U(2)/2\mathrm{U}(2)/{\mathbb{Z}_{2}}, two singular orbits occur at θ=0\theta=0 and θ=π4\theta=\frac{\pi}{4}. The orbit W0W_{0} is diffeomorphic to S3S^{3} and that of Wπ/4W_{\pi/4} to (S2×S1)/2(S^{2}\times S^{1})/\mathbb{Z}_{2}.

3.2 Adapting Frames

We now assume that MM is a nearly Kähler twistor space over a Riemannian four manifold NN. In other words, MM is either 3\mathbb{CP}^{3} or the flag manifold 𝔽\mathbb{F}. Lagrangian submanifolds of the latter have been studied in [Sto20a] so our interest is in 3\mathbb{CP}^{3} in this chapter. Before using the explicit description of 3\mathbb{CP}^{3} 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 LML\subset M, we clearly have TxLLag(TxM)T_{x}L\in\mathrm{Lag}(T_{x}M) for xLx\in L. Since the frame bundle reduces to HH there is a map Lag(TM|L)Lag(3)/H\mathrm{Lag}(TM|_{L})\to\mathrm{Lag}(\mathbb{C}^{3})/H. Hence, θ\theta can be understood as a map from LL to the interval [0,π4][0,\frac{\pi}{4}] and TθT_{\theta} from LL to SU(3)\mathrm{SU}(3). We now apply our knowledge of the action of HH on Lag(3)\mathrm{Lag}(\mathbb{C}^{3}) 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 MM reduces to HH, so PSU(3)P_{\mathrm{SU}(3)} reduces to an HH-bundle and we can assume ϕ13=ϕ23=ϕ31=ϕ32=0\phi_{13}=\phi_{23}=\phi_{31}=\phi_{32}=0. This means that there are two different reductions of P|LP|_{L}: The first is to an HH-bundle PH={p:3TM,pPSU(3)|Lp(b3)𝒱}P_{H}=\{p\colon\mathbb{C}^{3}\to TM,p\in P_{\mathrm{SU}(3)}|_{L}\mid p(b_{3})\in\mathcal{V}\}, simply because PSU(3)P_{\mathrm{SU}(3)} itself reduces to an HH bundle. The second reduction is to an SO(3)\mathrm{SO}(3)-bundle PSO(3)={p:3TM,pPSU(3)|Lp(3)=TL}P_{\mathrm{SO}(3)}=\{p\colon\mathbb{C}^{3}\to TM,p\in P_{\mathrm{SU}(3)}|_{L}\mid p(\mathbb{R}^{3})=TL\} or equivalently by imposing ηi=0\eta_{i}=0 as in section 2.

If TL𝒱TL\cap\mathcal{V} is a rank one bundle, or equivalently θπ4\theta\equiv\frac{\pi}{4}, then the intersection PSO(3)PHP_{\mathrm{SO}(3)}\cap P_{H} is a HSO(3)H\cap\mathrm{SO}(3) bundle. We will derive its structure equations in section 3.3. If θ\theta avoids the value π4\frac{\pi}{4} then the intersection TL𝒱TL\cap\mathcal{V} is trivial and PSO(3)PH=P_{\mathrm{SO}(3)}\cap P_{H}=\emptyset 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 xLx\in L there is a frame in PHP_{H} which maps WθW_{\theta} to TLTL. Such a frame is unique up to the action of the stabiliser of WθW_{\theta} in HH, which is computed in 3.4. This means that

Q=PHTθPSO(3).\displaystyle Q=P_{H}T_{\theta}\cap P_{\mathrm{SO}(3)}\neq\emptyset. (3.7)

This is a principal bundle over LL with structure group given as in 3.4 if θ\theta is either equal to 0 or π4\frac{\pi}{4} everywhere or if θ\theta avoids these values altogether. In the latter case, the structure group is discrete. We first describe all special Lagrangians where θ\theta is constant and equal to one of the boundary values everywhere. If θπ4\theta\equiv\frac{\pi}{4} then LL intersects every twistor fibre in a circle and maps to a surface in NN.

3.3 Lagrangians with θπ4\theta\equiv\frac{\pi}{4}

There is a general construction for Lagrangian submanifolds in the twistor space ZZ of an arbitrary Riemannian four-manifold NN due to Storm [Sto20] and Konstantinov [Kon17]. To make sense of how a Lagrangian submanifold in ZZ is defined, recall that ZZ carries two almost complex structures J1,J2J_{1},J_{2} and metrics gλg_{\lambda} for λ0\lambda\in\mathbb{R}_{\geq 0}. For a surface XNX\subset N define the circle bundle LXZ(N)L_{X}\subset Z(N) with fibre over xXx\in X equal to {JZx(N)J(TxX)=νx}\{J\in Z_{x}(N)\mid J(T_{x}X)=\nu_{x}\}. Geometrically, the fibre of LXL_{X} at xXx\in X is the equator in each twistor fibre, which is diffeomorphic to S2S^{2}, relative to the twistor lift of XX at xx. It turns out that this construction gives a lot of examples of Lagrangians in twistor spaces.

Proposition 3.6.

[Sto20] The submanifold LXL_{X} is Lagrangian in ZZ for both J1J_{1} and J2J_{2} and every gλg_{\lambda} if XX is superminimal. Conversely, if LXL_{X} is Lagrangian for any JaJ_{a} and gλg_{\lambda}, then XX is superminimal.

Assume LL is Lagrangian with θπ4\theta\equiv\frac{\pi}{4} so TLTL\cap\mathcal{H} and TL𝒱TL\cap\mathcal{V} are a rank two and a rank one bundle and TL=TLTL𝒱.TL=TL\cap\mathcal{H}\oplus TL\cap\mathcal{V}. So LL is also Lagrangian for J1J_{1} and LL arises via the construction above. In this case the intersection PO(2)=PHPSO(3)P_{\mathrm{O}(2)}=P_{H}\cap P_{\mathrm{SO}(3)} is an S(O(2)×O(1))S(\mathrm{O}(2)\times O(1)) bundle which is defined by imposing ηi=0\eta_{i}=0 for i=1,,3i=1,\dots,3 on PHP_{H}. Since β32=β23=β31=β13=0\beta_{32}=\beta_{23}=\beta_{31}=\beta_{13}=0 the equation βσ=0\beta\wedge\sigma=0 implies that β33\beta_{33} lies in the span of σ3\sigma_{3} and β11,β22\beta_{11},\beta_{22} lie in the span of σ1\sigma_{1} and σ2\sigma_{2}. Since Tr(ϕ)=0\mathrm{Tr}(\phi)=0 this implies that β33=0=β11+β22\beta_{33}=0=\beta_{11}+\beta_{22}, i.e. ϕ\phi takes values in 𝔰𝔲(2)\mathfrak{su}(2) when restricted to PO(2)P_{\mathrm{O}(2)}.

We can view (σ1,σ2,σ3,η1,η2,η3)(\sigma_{1},\sigma_{2},\sigma_{3},\eta_{1},\eta_{2},\eta_{3}) locally as an orthonormal co-frame on TM|LTM|_{L}, see 2.2. The forms σi\sigma_{i} vanish on the normal bundle while ηi\eta_{i} vanish on TLTL. The form σ3\sigma_{3} is dual to the unit vector field tangent along the fibres of LXL\to X. Since β33=0\beta_{33}=0 this means that the fibres of LXL\to X are in fact geodesics. Since twistor fibres are totally geodesic 1M\mathbb{CP}^{1}\subset M these geodesics are great circles in the twistor fibres.

Since β3i=0\beta_{3i}=0 for i=1,2,3i=1,2,3 this implies h3ij=0h_{3ij}=0 so the fundamental cubic is of the form

a(x133x1x22)+b(x233x2x12).a(x_{1}^{3}-3x_{1}x_{2}^{2})+b(x_{2}^{3}-3x_{2}x_{1}^{2}).

We have therefore shown.

Proposition 3.7.

The fundamental cubic of LXL_{X} in a nearly Kähler twistor space either vanishes or has stabiliser S3S_{3}.

We can also recover the result that XX is superminimal by showing that the second fundamental form XX in NN is complex-linear and using [MU97, Proposition 1c].

Remark 3.8.

Bryant considers special Lagrangians of the form C1×Σ22=3C^{1}\times\Sigma^{2}\subset\mathbb{C}\oplus\mathbb{C}^{2}=\mathbb{C}^{3}. The cubic form of such submanifolds is always stabilised by S3S_{3}. These examples are somewhat analogous to horizontal Lagrangians whose fundamental cubic also admits an S3S_{3} symmetry.

From now on, we will work specifically with M=3M=\mathbb{CP}^{3}. We have seen that LXL_{X} is either totally geodesic or its fundamental cubic has stabiliser S3S_{3}. If LXL_{X} is homogeneous then XX is a homogeneous superminimal surface in S4S^{4}. Such a surface is equal to a totally geodesic S2S4S^{2}\subset S^{4} or the Veronese curve in S4S^{4}. Hence, there are only two different examples of homogeneous special Lagrangian submanifolds with θπ4\theta\equiv\frac{\pi}{4}. Both of them are known as Lagrangians for the Kähler structure on 3\mathbb{CP}^{3}.

Example 3.9 (The standard 3\mathbb{RP}^{3}).

The standard 33\mathbb{RP}^{3}\subset\mathbb{CP}^{3} is a totally geodesic special Lagrangian submanifold. It fibres over a totally geodesic S2S^{2} in S4S^{4} under the twistor fibration. It is the orbit of {(ab¯ba¯)a,bj,|a|2+|b|2=1}SU(2)\left\{\begin{pmatrix}a&-\bar{b}\\ b&\bar{a}\end{pmatrix}\mid a,b\in\mathbb{R}\oplus j\mathbb{R},\quad|a|^{2}+|b|^{2}=1\right\}\cong\mathrm{SU}(2) on [1,0,0,0][1,0,0,0].

The second example was discovered in [Chi04] and is described in [Kon17] in terms of the twistor fibration.

Example 3.10 (Chiang Lagrangian).

The SU(2)\mathrm{SU}(2) subgroup of Sp(2)\mathrm{Sp}(2) which comes from the irreducible representation of SU(2)\mathrm{SU}(2) on 4=S3(2)\mathbb{C}^{4}=S^{3}(\mathbb{C}^{2}) has a special Lagrangian orbit at [1,0,0,1]3[1,0,0,1]\in\mathbb{CP}^{3}. This example is known as the Chiang Lagrangian and fibres over the Veronese surface in S4S^{4}. The SU(2)\mathrm{SU}(2) subgroup acts with stabiliser S3S_{3} on [1,0,0,1][1,0,0,1]. The stabiliser subgroup induces the full symmetry group of the fundamental cubic since the Chiang Lagrangian is not totally geodesic.

Since superminimal curves in S4S^{4} have an explicit Weierstraß parametrisation one can produce many (explicit) examples of special Lagrangians in 3\mathbb{CP}^{3}. 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 3\mathbb{CP}^{3} 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 3\mathbb{CP}^{3} 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 θ\theta to describe the structure equations for a special Lagrangian in 3\mathbb{CP}^{3}.

The bundle Sp(2)\mathrm{Sp}(2) embeds into the frame bundle of 3\mathbb{CP}^{3} via the adjoint action of S1×S3S^{1}\times S^{3} on 𝔪\mathfrak{m} which factors through the double cover S1×S3U(2)S^{1}\times S^{3}\to\mathrm{U}(2). So, on the level of structure equations we identify PHP_{H} with Sp(2)\mathrm{Sp}(2). We apply the gauge transformation TθT_{\theta} to Sp(2)\mathrm{Sp}(2), which defines the bundle QQ as in eq. 3.7. This bundle has a reduced structure group, depending on the behaviour of the function θ\theta, which is made precise in 3.4. For example, if θ\theta avoids the values 0 and π4\frac{\pi}{4} then the structure group of QQ is 2\mathbb{Z}_{2}.

Recall from section 1.1 that Sp(2)\mathrm{Sp}(2) is an S1×S3S^{1}\times S^{3} principal bundle over 3\mathbb{CP}^{3}. A local unitary frame for the nearly Kähler structure on 3\mathbb{CP}^{3} is obtained by pulling back the forms (ω1,ω2,ω3)(\omega_{1},\omega_{2},\omega_{3}), which are components of the Maurer-Cartan form on Sp(2)\mathrm{Sp}(2). We can realise the bundle QQ by setting Tθ1(ω1,ω2,ω3)=(ζ1,ζ2,ζ3)T^{-1}_{\theta}(\omega_{1},\omega_{2},\omega_{3})=(\zeta_{1},\zeta_{2},\zeta_{3}), where TθT_{\theta} is defined in eq. 3.4, and imposing the equations

η1=0,η2=0,η3=0.\displaystyle\eta_{1}=0,\quad\eta_{2}=0,\quad\eta_{3}=0. (3.8)

Our aim is to compute the differentials of the one forms ζi\zeta_{i} and also of ρi\rho_{i} and τ\tau on the reduced bundle QQ. 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 TθT_{\theta}

ϕ=Tθ1AωTθ+Tθ1dTθ.\displaystyle\phi=T_{\theta}^{-1}A_{\omega}T_{\theta}+T_{\theta}^{-1}dT_{\theta}. (3.9)

Here AωA_{\omega} is the connection form defined on Sp(2)\mathrm{Sp}(2), see eq. 1.3. Since TθT_{\theta} lies in SU(3)\mathrm{SU}(3) the torsion transforms trivially and we have dζ=ϕζ[ζ]ζ\mathrm{d}\zeta=-\phi\wedge\zeta-[\zeta]\wedge\zeta by eq. 2.1. So in order to compute the differentials of ζ\zeta we compute the transformed connection one-form ϕ\phi from eq. 3.9. We split ϕ\phi into real and imaginary part ϕ=α+iβ\phi=\alpha+i\beta to get

α=(012Re(iexp(iθ)τ¯)12Re(exp(iθ)τ¯)12Re(iexp(iθ)τ)012(3ρ1+ρ2)cos(2θ)12Re(exp(iθ)τ)12(3ρ1+ρ2)cos(2θ)0)\displaystyle\alpha=\begin{pmatrix}0&\frac{1}{\sqrt{2}}\mathop{\rm Re}\nolimits(i\exp(i\theta)\bar{\tau})&\frac{-1}{\sqrt{2}}\mathop{\rm Re}\nolimits(\exp(-i\theta)\bar{\tau})\\ \frac{1}{\sqrt{2}}\mathop{\rm Re}\nolimits(i\exp(i\theta)\tau)&0&\frac{1}{2}(3\rho_{1}+\rho_{2})\cos(2\theta)\\ \frac{1}{\sqrt{2}}\mathop{\rm Re}\nolimits(\exp(-i\theta)\tau)&-\frac{1}{2}(3\rho_{1}+\rho_{2})\cos(2\theta)&0\end{pmatrix} (3.10)

and

β=(ρ1+ρ212Im(iexp(iθ)τ¯)12Im(exp(iθ)τ¯)12Im(iexp(iθ)τ)12(ρ1ρ22dθ)12(3ρ1+ρ2)sin(2θ)12Im(exp(iθ)τ)12(3ρ1+ρ2)sin(2θ)12(ρ1ρ2+2dθ)).\displaystyle\beta=\begin{pmatrix}-\rho_{1}+\rho_{2}&\frac{1}{\sqrt{2}}\mathop{\rm Im}\nolimits(i\exp(i\theta)\bar{\tau})&\frac{-1}{\sqrt{2}}\mathop{\rm Im}\nolimits(\exp(-i\theta)\bar{\tau})\\ \frac{1}{\sqrt{2}}\mathop{\rm Im}\nolimits(i\exp(i\theta)\tau)&\frac{1}{2}(\rho_{1}-\rho_{2}-2\mathrm{d}\theta)&\frac{1}{2}(3\rho_{1}+\rho_{2})\sin(2\theta)\\ \frac{1}{\sqrt{2}}\mathop{\rm Im}\nolimits(\exp(-i\theta)\tau)&\frac{1}{2}(3\rho_{1}+\rho_{2})\sin(2\theta)&\frac{1}{2}(\rho_{1}-\rho_{2}+2\mathrm{d}\theta)\end{pmatrix}. (3.11)

To obtain expressions for the differentials of ρi\rho_{i} and τ\tau we use the curvature equation 2.7. The curvature tensor (R+iS)σσ(R+iS)\sigma\wedge\sigma transforms tensorially under gauge transformations yielding the explicit expressions

Sσσ\displaystyle S\sigma\wedge\sigma =(cos(2θ)σ2σ312cos(2θ)σ1σ312cos(2θ)σ1σ212cos(2θ)σ1σ312cos(2θ)σ2σ354sin(4θ)σ2σ312cos(2θ)σ1σ254sin(4θ)σ2σ312cos(2θ)σ2σ3)\displaystyle=\left(\begin{array}[]{ccc}-\cos(2\theta)\sigma_{2}\wedge\sigma_{3}&-\frac{1}{2}\cos(2\theta)\sigma_{1}\wedge\sigma_{3}&\frac{1}{2}\cos(2\theta)\sigma_{1}\wedge\sigma_{2}\\ -\frac{1}{2}\cos(2\theta)\sigma_{1}\wedge\sigma_{3}&\frac{1}{2}\cos(2\theta)\sigma_{2}\wedge\sigma_{3}&\frac{5}{4}\sin(4\theta)\sigma_{2}\wedge\sigma_{3}\\ \frac{1}{2}\cos(2\theta)\sigma_{1}\wedge\sigma_{2}&\frac{5}{4}\sin(4\theta)\sigma_{2}\wedge\sigma_{3}&\frac{1}{2}\cos(2\theta)\sigma_{2}\wedge\sigma_{3}\\ \end{array}\right) (3.15)
Rσσ\displaystyle R\sigma\wedge\sigma =12(0σ1(σ2sin(2θ)σ3)σ1(σ3sin(2θ)σ2)σ1(sin(2θ)σ3σ2)05cos2(2θ)σ2σ3σ1(sin(2θ)σ2σ3)5cos2(2θ)σ2σ30).\displaystyle=\frac{1}{2}\left(\begin{array}[]{ccc}0&\sigma_{1}\wedge(\sigma_{2}-\sin(2\theta)\sigma_{3})&\sigma_{1}\wedge(\sigma_{3}-\sin(2\theta)\sigma_{2})\\ \sigma_{1}\wedge(\sin(2\theta)\sigma_{3}-\sigma_{2})&0&5\cos^{2}(2\theta)\sigma_{2}\wedge\sigma_{3}\\ \sigma_{1}\wedge(\sin(2\theta)\sigma_{2}-\sigma_{3})&-5\cos^{2}(2\theta)\sigma_{2}\wedge\sigma_{3}&0\\ \end{array}\right). (3.19)

Finally, combining the explicit expressions of ϕ,R,S\phi,R,S with eq. 2.4-2.7 results in the following differential identities

dρ1=32cos(2θ)σ2σ3,dρ2=12cos(2θ)σ2σ3+iττ¯dτ=2iτρ2+12σ1(iσ2exp(iθ)σ3exp(iθ))dσ1=(ϵ1σ2+ϵ2σ3)+σ2σ3dσ2=12cos(2θ)(3ρ1+ρ2)σ3ϵ1σ1σ1σ3dσ3=12cos(2θ)(3ρ1+ρ2)σ2ϵ2σ1+σ1σ2\displaystyle\begin{split}\mathrm{d}\rho_{1}&=\frac{3}{2}\cos(2\theta)\sigma_{2}\wedge\sigma_{3},\quad\mathrm{d}\rho_{2}=\frac{1}{2}\cos(2\theta)\sigma_{2}\wedge\sigma_{3}+i\tau\wedge\bar{\tau}\\ \mathrm{d}\tau&=-2i\tau\wedge\rho_{2}+\frac{1}{\sqrt{2}}\sigma_{1}\wedge(i\sigma_{2}\exp(-i\theta)-\sigma_{3}\exp(i\theta))\\ \mathrm{d}\sigma_{1}&=(\epsilon_{1}\wedge\sigma_{2}+\epsilon_{2}\wedge\sigma_{3})+\sigma_{2}\wedge\sigma_{3}\\ \mathrm{d}\sigma_{2}&=-\frac{1}{2}\cos(2\theta)(3\rho_{1}+\rho_{2})\wedge\sigma_{3}-\epsilon_{1}\wedge\sigma_{1}-\sigma_{1}\wedge\sigma_{3}\\ \mathrm{d}\sigma_{3}&=\frac{1}{2}\cos(2\theta)(3\rho_{1}+\rho_{2})\wedge\sigma_{2}-\epsilon_{2}\wedge\sigma_{1}+\sigma_{1}\wedge\sigma_{2}\end{split} (3.20)

with ϵ1=i22(exp(iθ)τexp(iθ)τ¯)\epsilon_{1}=\frac{i}{2\sqrt{2}}(\exp(i\theta)\tau-\exp(-i\theta)\bar{\tau}) and ϵ2=122(exp(iθ)τ+exp(iθ)τ¯)\epsilon_{2}=\frac{1}{2\sqrt{2}}(\exp(-i\theta)\tau+\exp(i\theta)\bar{\tau}). Recall, that the equation βσ\beta\wedge\sigma implies more algebraic identities.

These differential identities are satisfied on any special Lagrangian in 3\mathbb{CP}^{3}. 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 θ=0\theta=0 the equations simplify considerably.

If θπ/4\theta\neq\pi/4 everywhere there is a splitting TL=EETL=E\oplus E^{\perp} where EE^{\perp} is the kernel of the projection TL𝒱TL\to\mathcal{V}. Recall that the standard complex structure J1J_{1} on 3\mathbb{CP}^{3} agrees with the nearly Kähler structure J2J_{2} on \mathcal{H} and differs by a sign on 𝒱\mathcal{V}.

Proposition 3.11.

The distribution EE is invariant under the standard complex structure J1J_{1} on 3\mathbb{CP}^{3} if and only if θ0\theta\equiv 0. In that case, LL is a CR submanifold for the Kähler structure on 3\mathbb{CP}^{3} with EE being the J1J_{1}-invariant distribution on LL.

Proof.

In each point xLx\in L we can pick a frame p:TxM3p:T_{x}M\to\mathbb{C}^{3} such that TLTL is identified with Wθ(x)W_{\theta(x)}, 𝒱\mathcal{V} with span(b3)\mathrm{span}(b_{3}) and J1J_{1} with JJ^{\prime} from eq. 3.6. The statement follows from 3.3. If θ=0\theta=0 then EE is invariant under J1J_{1} and J1(E)J_{1}(E^{\perp}) is orthogonal to TLTL, as required. ∎

CR immersion from S3S^{3} to the Kähler n\mathbb{CP}^{n} have been studied in [HYL18]. The splitting TL=EETL=E\oplus E^{\perp} gives an ansatz for Lagrangians arising as a product X2×S1X^{2}\times S^{1} such that TX=ETX=E. Indeed, we will give such an example for θ0\theta\equiv 0 later. However, we first show that this ansatz fails when θ0\theta\neq 0 and XX is compact. Note that

ω𝒱=i2(ω3ω3¯)=12cos(2θ)σ2σ3\omega_{\mathcal{V}}=\frac{i}{2}(\omega_{3}\wedge\bar{\omega_{3}})=\frac{1}{2}\cos(2\theta)\sigma_{2}\wedge\sigma_{3}

and that dω𝒱\mathrm{d}\omega_{\mathcal{V}} is a multiple of Reψ\mathop{\rm Re}\nolimits\psi, which vanishes on LL. This implies the following.

Lemma 3.12.

If θπ/4\theta\neq\pi/4 is constant then 2cos(2θ)ω𝒱\frac{2}{\cos(2\theta)}\omega_{\mathcal{V}} defines a calibration on LL. The fibres of EE are the calibrated subspaces of 2cos(2θ)ω𝒱\frac{2}{\cos(2\theta)}\omega_{\mathcal{V}}.

Since ω𝒱\omega_{\mathcal{V}} is closed on LL it defines a cohomology class in H2(L,)H^{2}(L,\mathbb{R}). We have that when pulled back to Sp(2)\mathrm{Sp}(2), this class vanishes since dρ1=2ω𝒱ω=3ω𝒱\mathrm{d}\rho_{1}=2\omega_{\mathcal{V}}-\omega_{\mathcal{H}}=3\omega_{\mathcal{V}}. If θ\theta takes values in (0,π4)(0,\frac{\pi}{4}) the structure group of QQ is just 2\mathbb{Z}_{2}, generated by diag(1,1,1)\mathrm{diag}(1,-1,-1) in HH which corresponds to diag(i,i)S1×S3Sp(2)\mathrm{diag}(i,i)\in S^{1}\times S^{3}\subset\mathrm{Sp}(2). This element leaves ρ1\rho_{1} and ρ2\rho_{2} invariant so in particular ρ1\rho_{1} reduces to a form on LL and we have shown.

Proposition 3.13.

If θ\theta takes values in (0,π/4)(0,\pi/4) then [ω𝒱]=0[\omega_{\mathcal{V}}]=0. In particular, in this case LL does not have any compact two-dimensional submanifold which is tangent to EE.

3.5 Lagrangians with θ0\theta\equiv 0

The structure equations simplify significantly under the assumption θ0\theta\equiv 0. Indeed, plugging θ=0\theta=0 into eq. 3.20 yields β32=0\beta_{32}=0 and dθ=0\mathrm{d}\theta=0 implies β22=β33\beta_{22}=\beta_{33}. Since we have βσ=0\beta\wedge\sigma=0 Cartan’s lemma implies that there is a single function f:Lf\colon L\to\mathbb{R} such that

β22=β33=1/2β11=1/2fσ1,β21=1/2fσ2,β31=1/2σ3.\beta_{22}=\beta_{33}=-1/2\beta_{11}=1/2f\sigma_{1},\quad\beta_{21}=1/2f\sigma_{2},\quad\beta_{31}=1/2\sigma_{3}.

The fundamental cubic is equal to f(x13+3/2x22x1+3/2x32x1)f(-x_{1}^{3}+3/2x_{2}^{2}x_{1}+3/2x_{3}^{2}x_{1}) which has stabiliser SO(2)\mathrm{SO}(2). Let γ=ρ1+ρ2\gamma=\rho_{1}+\rho_{2}, then

α21=12fσ3,α31=12fσ2,α32=γ12fσ1.\alpha_{21}=-\frac{1}{2}f\sigma_{3},\quad\alpha_{31}=\frac{1}{2}f\sigma_{2},\quad\alpha_{32}=-\gamma-\frac{1}{2}f\sigma_{1}.

The structure equations are then equivalent to

1+f+2f2=0,dγ1=12(5f)σ2σ3.\displaystyle-1+f+2f^{2}=0,\quad d\gamma_{1}=\frac{1}{2}(5-f)\sigma_{2}\wedge\sigma_{3}. (3.21)

Hence, there are two examples of special Lagrangian submanifolds LfL_{f} with θ=0\theta=0, both of which are homogeneous and in particular compact. Neither of them is totally geodesic. Note that, as a subset of Sp(2)\mathrm{Sp}(2), the adapted frame bundle is defined by the equations

fσ1=ρ1ρ2,τ=fσ2+iσ32.\displaystyle f\sigma_{1}=\rho_{1}-\rho_{2},\quad\tau=f\frac{\sigma_{2}+i\sigma_{3}}{\sqrt{2}}. (3.22)

They are both orbits of a Lie group with Lie algebra equal to the span of

𝔪1=(i00i),𝔪2=(if/211if/2)𝔪3=(jji2ji2jf),𝔪4=(jij12j12jif).\displaystyle\mathfrak{m}_{1}=\begin{pmatrix}i&0\\ 0&i\end{pmatrix},\quad\mathfrak{m}_{2}=\begin{pmatrix}if/{\sqrt{2}}&-1\\ 1&-if/{\sqrt{2}}\end{pmatrix}\quad\mathfrak{m}_{3}=\begin{pmatrix}-j&-j\frac{i}{\sqrt{2}}\\ -j\frac{i}{\sqrt{2}}&jf\end{pmatrix},\quad\mathfrak{m}_{4}=\begin{pmatrix}-ji&j\frac{1}{\sqrt{2}}\\ j\frac{1}{\sqrt{2}}&jif\end{pmatrix}.

Eq. 3.21 shows that ff is in fact constant and must be equal to either 1-1 or 12\frac{1}{2}. So, there are two distinct examples of special Lagrangians with θ=0\theta=0. We describe the geometry of each of them.

Example 3.14.

There is a unique special Lagrangian with θ=0\theta=0 and f=1f=-1. The reduced frame bundle over this submanifold is described by

dσ1=0,dσ2=γσ3,dσ3=γσ2,dγ=3σ2σ3.\mathrm{d}\sigma_{1}=0,\quad\mathrm{d}\sigma_{2}=-\gamma\wedge\sigma_{3},\quad\mathrm{d}\sigma_{3}=\gamma\wedge\sigma_{2},\quad\mathrm{d}\gamma=3\sigma_{2}\wedge\sigma_{3}.

These are the structure equations of S1×S2S^{1}\times S^{2} where S2S^{2} carries a metric of constant curvature 33.

Example 3.15.

There is a unique special Lagrangian with θ=0\theta=0 and f=12f=\frac{1}{2}. The reduced frame bundle over this submanifold is described by

dσ1=32σ2σ3,dσ2=(γ+32σ1)σ3,dσ3=(γ+32σ1)σ2,dγ=94σ2σ3.\mathrm{d}\sigma_{1}=\frac{3}{2}\sigma_{2}\wedge\sigma_{3},\quad\mathrm{d}\sigma_{2}=-(\gamma+\frac{3}{2}\sigma_{1})\wedge\sigma_{3},\quad\mathrm{d}\sigma_{3}=(\gamma+\frac{3}{2}\sigma_{1})\wedge\sigma_{2},\quad\mathrm{d}\gamma=\frac{9}{4}\sigma_{2}\wedge\sigma_{3}.

which are the structure equations of a Berger sphere.

3.6 Lagrangians avoiding Boundary Values

From now on we assume that θ\theta is not identically zero or π/4\pi/4 and denote by LL^{*} the open set where θ\theta avoids these values. On LL^{*} the frame bundle reduces to a discrete bundle. For JJ-holomorphic curves in 3\mathbb{CP}^{3} the second fundamental form is determined by two angle functions [Asl21]. In contrast, θ\theta alone does not determine the second fundamental form of LL^{*}. We let dθ=t1σ1+t2σ2+t3σ3\mathrm{d}\theta=t_{1}\sigma_{1}+t_{2}\sigma_{2}+t_{3}\sigma_{3} and x=h221,y=h222,z=h322,w=h321x=h_{221},y=h_{222},z=h_{322},w=h_{321} such that β\beta is entirely determined by θ\theta and these quantities. Clearly all of these functions are constant on orbits of Lie subgroups of Sp(2)\mathrm{Sp}(2), the converse is also true.

Proposition 3.16.

Any solution with x,y,z,wx,y,z,w and θ(0,π/4)\theta\in(0,\pi/4) constant is an orbit of a Lie group.

Proof.

Let LL be the special Lagrangian corresponding to this solution with adapted frame bundle L^\hat{L}. Then L^\hat{L} is an integral submanifold of the EDS generated by ηi,βhσ\eta_{i},\beta-h\sigma. By assumption, hh has constant coefficients which means that the equations ηi=0,β=hσ\eta_{i}=0,\beta=h\sigma describe a linear subspace of 𝔰𝔭(2)\mathfrak{sp}(2) and hence L^\hat{L} is a Lie group and L^L\hat{L}\to L is a double cover of Lie groups. ∎

In principle, we could derive a set of PDE’s for θ,x,y,z\theta,x,y,z and tit_{i} from the structure equations but this is not practical in full generality. However, there is a somewhat surprising homogeneous example.

Example 3.17.

Setting

x=2/5,y=0,z=0,w=353/2,θ=12arccos(7255)x=-\sqrt{2/5},\quad y=0,\quad z=0,\quad w=-\frac{3}{5}\sqrt{3/2},\quad\theta=\frac{1}{2}\arccos(\frac{7\sqrt{2}}{5\sqrt{5}})

is a solution to eq. 3.20 and hence corresponds to a unique special Lagrangian in 3\mathbb{CP}^{3}. The fundamental cubic of this example is given by 25(2x133x1x223x1x32)956x1x2x3,\sqrt{\frac{2}{5}}(2x_{1}^{3}-3x_{1}x_{2}^{2}-3x_{1}x_{3}^{2})-\frac{9}{5}\sqrt{6}x_{1}x_{2}x_{3}, whose orientation preserving symmetry group is 2\mathbb{Z}_{2} coming from (x2,x3)(x2,x3)(x_{2},x_{3})\mapsto(-x_{2},-x_{3}). The Ricci curvature is diagonal in the (dual) frame σ1,σ2+σ3,σ2σ3\sigma_{1},\sigma_{2}+\sigma_{3},\sigma_{2}-\sigma_{3} in which Ric=diag(99/50,27/50(2+15),27/50(2+15)).\mathrm{Ric}=\mathrm{diag}(-99/50,-27/50(-2+\sqrt{15}),27/50(2+\sqrt{15})).

By 3.16, this example is homogeneous. We have found new examples by imposing conditions on θ\theta. 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 3\mathbb{RP}^{3} is the unique totally geodesic special Lagrangian in 3\mathbb{CP}^{3}.

Proof.

There is no totally geodesic Lagrangian which lies in θ[0,π4)\theta\in[0,\frac{\pi}{4}). This is because in that case β=0\beta=0 forces ρ1=0\rho_{1}=0 but this is a contradiction to the first equation of 3.20.

If LL is a totally geodesic Lagrangian with θπ4\theta\equiv\frac{\pi}{4} then the adapted frame bundle QQ is a four-dimensional submanifold of Sp(2)\mathrm{Sp}(2) on which ηi\eta_{i} and β\beta vanish. If θπ4\theta\equiv\frac{\pi}{4} then SS vanishes on the adapted bundle QQ and by eq. 2.7 the ideal generated by ηi\eta_{i} and βij\beta_{ij} is closed under differentials. By Frobenius’ theorem, there is a unique maximal submanifold on which these forms vanish that passes through the identity eSp(2)e\in\mathrm{Sp}(2). Hence, up to isometries, there is a unique totally geodesic special Lagrangian in 3\mathbb{CP}^{3} with θ=π4\theta=\frac{\pi}{4}. We have already found this example, it is the standard 33\mathbb{RP}^{3}\subset\mathbb{CP}^{3}. ∎

4 Classifying SU(2)\mathrm{SU}(2) 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 SU(2)\mathrm{SU}(2) in both S6S^{6} and 3\mathbb{C}^{3}. In S6S^{6}, there is a unique example of this type, the squashed three-sphere [Lot11, Example 6.4]. In 3\mathbb{C}^{3}, the Harvey-Lawson examples [Bry06a, HL82]

Lc={(s+it)uuS23,t33s2t=c3}L_{c}=\{(s+it)u\mid u\in S^{2}\subset\mathbb{R}^{3},t^{3}-3s^{2}t=c^{3}\}

admit a cohomogeneity one action of SO(3)\mathrm{SO}(3) for c0c\neq 0.

The situation in 3\mathbb{CP}^{3} is different. We show in this section that all special Lagrangians that admit an action of an SU(2)\mathrm{SU}(2) group of automorphism are in fact homogeneous and have already been described in the previous section. We introduce SU(2)\mathrm{SU}(2) moment-type maps to prove this classification.

4.1 SU(2)\mathrm{SU}(2) Moment Maps

Assume that SU(2)\mathrm{SU}(2) acts effectively on MM with three-dimensional principal orbits and by nearly Kähler automorphisms. Let {ξ1,ξ2,ξ3}\{\xi_{1},\xi_{2},\xi_{3}\} be a basis of 𝔰𝔲(2)\mathfrak{su}(2) such that [ξi,ξj]=ϵijkξk[\xi_{i},\xi_{j}]=-\epsilon_{ijk}\xi_{k}. Denote the corresponding fundamental vector fields by KξiK^{\xi_{i}}. The map ξKξ\xi\to K^{\xi} is an anti Lie algebra homomorphism. Hence, the vector fields KξiK^{\xi_{i}} obey the standard Pauli commutator relationships [Kξi,Kξj]=ϵijkKξk[K^{\xi_{i}},K^{\xi_{j}}]=\epsilon_{ijk}K^{\xi_{k}}. Consider the map

μ=(μ1,μ2,μ3)=(ω(Kξ2,Kξ3),ω(Kξ3,Kξ1),ω(Kξ1,Kξ2)).\mu=(\mu_{1},\mu_{2},\mu_{3})=(\omega(K^{\xi_{2}},K^{\xi_{3}}),\omega(K^{\xi_{3}},K^{\xi_{1}}),\omega(K^{\xi_{1}},K^{\xi_{2}})).

Then μ:M3\mu\colon M\to\mathbb{R}^{3} is an SU(2)\mathrm{SU}(2) equivariant map with respect to the action of SU(2)\mathrm{SU}(2) on 3\mathbb{R}^{3} coming from the double cover SU(2)SO(3)\mathrm{SU}(2)\to\mathrm{SO}(3). In addition, define the invariant scalar function

ν=Imψ(Kξ1,Kξ2,Kξ3).\nu=\mathop{\rm Im}\nolimits\psi(K^{\xi_{1}},K^{\xi_{2}},K^{\xi_{3}}).

The map μ\mu is not a multi-moment-type map in the sense of [MS13, Definition 3.5]. The Lie-kernel of Λ2𝔰𝔲(2)𝔰𝔲(2)\Lambda^{2}\mathfrak{su}(2)\to\mathfrak{su}(2) is trivial, so there is no non-trivial multi-moment map for the three form Reψ\mathop{\rm Re}\nolimits\psi. On the other hand, the map Λ3𝔰𝔲(2)Λ2𝔰𝔲(2)\Lambda^{3}\mathfrak{su}(2)\to\Lambda^{2}\mathfrak{su}(2) is trivial and ν\nu is a multi-moment map with values in Λ3𝔰𝔲(2)\mathbb{R}\cong\Lambda^{3}\mathfrak{su}(2) for 12ωω-\frac{1}{2}\omega\wedge\omega. We will refer to μ\mu and ν\nu 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 μ\mu and ν\nu.

Lemma 4.1.

The form Reψ\mathop{\rm Re}\nolimits\psi vanishes on SU(2)SU(2) orbits, i.e. Reψ(Kξ1,Kξ2,Kξ3)=0\mathop{\rm Re}\nolimits\psi(K^{\xi_{1}},K^{\xi_{2}},K^{\xi_{3}})=0.

Proof.

Let 𝒪\mathcal{O} be a three-dimensional orbit of SU(2)SU(2). Since SU(2)SU(2) acts by isometries on MM we have that vol𝒪\mathrm{vol}_{\mathcal{O}} is a SU(2)SU(2) invariant form on 𝒪\mathcal{O}. The same holds for Reψ|𝒪\mathop{\rm Re}\nolimits\psi|_{\mathcal{O}}. So there is λ\lambda\in\mathbb{R} such that Reψ|𝒪=λvol𝒪\mathop{\rm Re}\nolimits\psi|_{\mathcal{O}}=\lambda\mathrm{vol}_{\mathcal{O}}. Since Reψ\mathop{\rm Re}\nolimits\psi is exact λvol(𝒪)=𝒪Reψ=0\lambda\mathrm{vol}(\mathcal{O})=\int_{\mathcal{O}}\mathop{\rm Re}\nolimits\psi=0 i.e. λ=0\lambda=0. ∎

Since ψ=Reψ+iImψ\psi=\mathop{\rm Re}\nolimits\psi+i\mathop{\rm Im}\nolimits\psi is non-degenerate this means that ν\nu vanishes if and only if Kξ1,Kξ2,Kξ3K^{\xi_{1}},K^{\xi_{2}},K^{\xi_{3}} are linearly dependent over \mathbb{C}. By Cartan’s formula and the nearly Kähler structure equations we get

dν=2lμlω(Kξl,), and dμk\displaystyle\mathrm{d}\nu=2\sum_{l}\mu_{l}\omega(K^{\xi_{l}},\cdot),\text{ and }\mathrm{d}\mu_{k} =ω(Kξk,)+3Reψ(Kξi,Kξj,)\displaystyle=-\omega(K^{\xi_{k}},\cdot)+3\mathop{\rm Re}\nolimits\psi(K^{\xi_{i}},K^{\xi_{j}},\cdot) (4.1)

where (i,j,k)(i,j,k) is a cylic permutation of (1,2,3)(1,2,3). The following proposition is somewhat similar to the toric situation [Dix19] as we can identify SU(2)\mathrm{SU}(2) invariant special Lagrangians orbits by the values of the maps μ\mu and ν\nu.

Proposition 4.2.

The orbit of a point xMx\in M is special Lagrangian if and only if ν(x)0\nu(x)\neq 0 and μ(x)=0\mu(x)=0. The set ν1(0)μ1(0)\nu^{-1}(0)\cap\mu^{-1}(0) is a union of fixed points of the SU(2)SU(2) action and two-spheres on which ω\omega vanishes. If MM has non-vanishing Euler characteristic then 0 lies in the image of ν\nu. The function ν\nu is not constant and the set of points in which dν=0\mathrm{d}\nu=0 and ν0\nu\neq 0 consists of special Lagrangian orbits.

Proof.

By the definition of μ\mu, the two-form ω\omega vanishes on the SU(2)\mathrm{SU}(2) orbit of xx if and only if μ(x)=0\mu(x)=0. If ν(x)0\nu(x)\neq 0 then KξiK^{\xi_{i}} are linearly independent at xx and the orbit at xx is 3-dimensional, which implies the first statement. If ν(x)=0\nu(x)=0 then the orbit has dimension less than three and the second statement follows from the fact that lower-dimensional SU(2)\mathrm{SU}(2) orbits must be points or two-spheres.

Eq. 4.1 implies that if ν\nu is constant then (Kξ1,Kξ2,Kξ3)(K^{\xi_{1}},K^{\xi_{2}},K^{\xi_{3}}) are linearly dependent everywhere which contradicts the principal orbit type being three-dimensional. If χ(M)0\chi(M)\neq 0 then any vector field KξiK^{\xi_{i}} must have a zero, which forces ν\nu to vanish. Finally, consider a point xx in which dν=0\mathrm{d}\nu=0 and ν0\nu\neq 0. We want to show that μ(x)=(0,0,0)\mu(x)=(0,0,0). Using the action of SU(2)SU(2) we can assume that μ2(x),μ3(x)=0\mu_{2}(x),\mu_{3}(x)=0. Then 0=JKξ1ν=2Kξ12μ10=JK^{\xi_{1}}\nu=-2\|K^{\xi_{1}}\|^{2}\mu_{1}. But ν0\nu\neq 0 and hence μ1(x)=0\mu_{1}(x)=0. ∎

Since either the maximum or minimum of ν\nu is not zero this implies an existence result for special Lagrangians.

Corollary 4.3.

If MM is compact then the SU(2)SU(2) action has a special Lagrangian orbit.

If LL is a special Lagrangian submanifold on which a SU(2)\mathrm{SU}(2) subgroup acts then LL will lie in the vanishing set of μ\mu. So we can classify all SU(2)\mathrm{SU}(2) invariant special Lagrangian submanifolds of 3\mathbb{CP}^{3} by computing the vanishing set of μ\mu for every SU(2)\mathrm{SU}(2) subgroup of Sp(2)\mathrm{Sp}(2).

Definition 4.4.

Define the three SU(2)\mathrm{SU}(2) subgroups of Sp(2)\mathrm{Sp}(2) as K1={1}×Sp(1)K_{1}=\{1\}\times\mathrm{Sp}(1), K2=SU(2)K_{2}=\mathrm{SU}(2), arising from the inclusion 22\mathbb{C}^{2}\subset\mathbb{H}^{2}, and K3K_{3} which comes from the irreducible representation of SU(2)\mathrm{SU}(2) on S3(2)4S^{3}(\mathbb{C}^{2})\cong\mathbb{C}^{4}.

Any three-dimensional subgroup of Sp(2)\mathrm{Sp}(2) is conjugate to one of K1,K2,K3K_{1},K_{2},K_{3}.

Remark 4.5.

Note that SO(4)\mathrm{SO}(4) contains two SU(2)\mathrm{SU}(2) subgroups that do not stabilise a vector in 4\mathbb{R}^{4}. They are not conjugated to each other and, on the Lie algebra level, correspond to the splitting of Λ2(4)\Lambda^{2}(\mathbb{R}^{4}) into self-dual and anti-self-dual two forms. However, in SO(5)\mathrm{SO}(5), these two Lie algebras are conjugated to each other, for example via the element (x4,x5)(x4,x5)(x_{4},x_{5})\mapsto(-x_{4},-x_{5}). Since Sp(2)=Spin(5)\mathrm{Sp}(2)=\mathrm{Spin}(5) the same holds true for the corresponding SU(2)\mathrm{SU}(2) subgroups in Sp(2)\mathrm{Sp}(2).

The groups KiK_{i} naturally act on S4S^{4} through the double cover Sp(2)SO(5)\mathrm{Sp}(2)\to\mathrm{SO}(5). The group K1K_{1} acts via SU(2)SO(4)\mathrm{SU}(2)\subset\mathrm{SO}(4), the group K2K_{2} via the double cover SU(2)SO(3)\mathrm{SU}(2)\to\mathrm{SO}(3) leaving a plane in 5\mathbb{R}^{5} invariant and K3K_{3} acts irreducibly on 5\mathbb{R}^{5} and factors through SO(3)\mathrm{SO}(3). To relate the group invariant examples to those found in the previous section we compute the function θ\theta for group orbits, for which we use eq. 3.5. To this end, it makes sense to define μ𝒱=(ω𝒱(Kξ2,Kξ3),ω𝒱(Kξ3,Kξ1),ω𝒱(Kξ1,Kξ2)).\mu_{\mathcal{V}}=(\omega_{\mathcal{V}}(K^{\xi_{2}},K^{\xi_{3}}),\omega_{\mathcal{V}}(K^{\xi_{3}},K^{\xi_{1}}),\omega_{\mathcal{V}}(K^{\xi_{1}},K^{\xi_{2}})). The Killing vector fields corresponding to the subgroups KiK_{i} admit quite simple expressions in local coordinates. So, to express ν\nu for KiK_{i} in homogeneous coordinates we need to do so for the nearly Kähler form ω=i2iωiω¯i\omega=\frac{i}{2}\sum_{i}\omega_{i}\wedge\bar{\omega}_{i}. This is the essence of eq. 1.4, where the forms ωi\omega_{i} are pulled back to a chart in 3\mathbb{CP}^{3} by a local section.

It will be challenging to compute ν\nu for K2K_{2} and K3K_{3}, 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 GG, 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 SU(2)\mathrm{SU}(2). There is a general framework to relate this question to the existence of nowhere vanishing sections in bundles over the flag manifold G/TG/T [AÐ10]. The following result follows a similar strategy and gives a direct proof for G=SU(2)G=\mathrm{SU}(2).

Lemma 4.6.

Let (V,ρ)(V,\rho) be a finite dimensional unitary representation of G=SU(2)G=\mathrm{SU}(2) with all weights non-zero and HH be a hyperplane which is invariant under the maximal torus U(1)SU(2)\mathrm{U}(1)\subset\mathrm{SU}(2). Then HH intersects every GG orbit.

Proof.

Since HH is U(1)\mathrm{U}(1) invariant there is a linear U(1)\mathrm{U}(1) equivariant map f:Vf\colon V\to\mathbb{C} such that ker(f)=H\mathrm{ker}(f)=H. Assume that there is an xVx\in V such that G.xH=G.x\cap H=\emptyset. Then s:gf(gx)s\colon g\mapsto f(gx) is a non-vanishing U(1)\mathrm{U}(1) equivariant map SU(2)\mathrm{SU}(2)\to\mathbb{C}. Restricting this map to U(1)SU(2)\mathrm{U}(1)\subset\mathrm{SU}(2) gives a representation τ\tau of U(1)\mathrm{U}(1) on \mathbb{C} of weight kk\in\mathbb{Z}.

Note that the principal bundle SU(2)SU(2)/U(1)=S2\mathrm{SU}(2)\to\mathrm{SU}(2)/{\mathrm{U}(1)}=S^{2} is the Hopf fibration and that ss gives rise to a nowhere vanishing section of the associated bundle E=SU(2)×τE=\mathrm{SU}(2)\times_{\tau}\mathbb{C} over S2S^{2}. Since the Hopf fibration has non-trivial Chern class, the complex line bundle EE is trivial which forces k=0k=0. This is a contradiction because ff restricts to an equivariant isomorphism from HH^{\perp} to \mathbb{C}, so HH^{\perp} is a zero-weight subspace. ∎

Note that, in the situation above, HH is invariant under U(1)\mathrm{U}(1) and the action of U(1)\mathrm{U}(1) on HH splits into one-dimensional components. Then every GG orbit also intersects the set HHH^{\prime}\subset H where one of the \mathbb{C} components is restricted to the set 0\mathbb{R}_{\geq 0}.

All the actions of KiK_{i} on 3\mathbb{CP}^{3} factor through an action of SU(4)\mathrm{SU}(4) on 4\mathbb{C}^{4}. The irreducible action ρk\rho_{k} of SU(2)\mathrm{SU}(2) on Sk(2)S^{k}(\mathbb{C}^{2}) has weights (k,k2,,k+2,k)(k,k-2,\dots,-k+2,-k). The action of K2K_{2} on 3\mathbb{CP}^{3} factors through ρ1ρ1\rho_{1}\oplus\rho_{1} on 4\mathbb{C}^{4} and K3K_{3} through ρ3\rho_{3} on 4\mathbb{C}^{4}. In particular neither has a zero weight, so 4.6 applies to these cases.

4.1.1 K1K_{1}

Recall K1={1}×Sp(1)K_{1}=\{1\}\times\mathrm{Sp}(1), we compute the Killing vector fields on the chart 𝔸0={Z00}\mathbb{A}_{0}=\{Z_{0}\neq 0\}

Kξ1=Im(Z2Z2Z3Z3),Kξ2=Re(Z3Z2Z2Z3),Kξ3=Im(Z3Z2+Z2Z3).K^{\xi_{1}}=-\mathop{\rm Im}\nolimits(Z_{2}\frac{\partial}{\partial Z_{2}}-Z_{3}\frac{\partial}{\partial Z_{3}}),\quad K^{\xi_{2}}=\mathop{\rm Re}\nolimits(Z_{3}\frac{\partial}{\partial Z_{2}}-Z_{2}\frac{\partial}{\partial Z_{3}}),\quad K^{\xi_{3}}=\mathop{\rm Im}\nolimits(Z_{3}\frac{\partial}{\partial Z_{2}}+Z_{2}\frac{\partial}{\partial Z_{3}}).

We contract these vector fields with the nearly Kähler forms ω\omega and ψ\psi in homogeneous coordinates from eq. 1.4, which gives

ν\displaystyle\nu =|Z|612(|Z0|2+|Z1|2)(|Z2|2+|Z3|2)2,μ1=|Z|2(|Z3|2|Z2|2)f\displaystyle=-|Z|^{-6}\frac{1}{2}(|Z_{0}|^{2}+|Z_{1}|^{2})(|Z_{2}|^{2}+|Z_{3}|^{2})^{2},\quad\mu_{1}=|Z|^{-2}(|Z_{3}|^{2}-|Z_{2}|^{2})f
μ2+iμ3\displaystyle\mu_{2}+i\mu_{3} =2i|Z|2Z2Z3¯f,f=14|Z|2(2(|Z0|2+|Z1|2)+(|Z2|2+|Z3|2)).\displaystyle=2i|Z|^{-2}Z_{2}\overline{Z_{3}}f,\quad f=\frac{1}{4}|Z|^{-2}(-2(|Z_{0}|^{2}+|Z_{1}|^{2})+(|Z_{2}|^{2}+|Z_{3}|^{2})).

Hence, ν\nu vanishes on the line of fixed points {Z2=Z3=0}\{Z_{2}=Z_{3}=0\} or when f=0f=0. Note that Sp(1)×Sp(1)\mathrm{Sp}(1)\times\mathrm{Sp}(1) is the centraliser of K1K_{1} in Sp(2)\mathrm{Sp}(2), acts with cohomogeneity one on 3\mathbb{CP}^{3} and the orbits of that action are the level sets of ff. In particular Sp(1)×Sp(1)\mathrm{Sp}(1)\times\mathrm{Sp}(1) acts transitively on f=0f=0 which means that up to isometries there is a unique special Lagrangian on which K1K_{1} acts. Hence, for simplification we consider the orbit 𝒪11\mathcal{O}_{11} at the point P11=[1,0,2,0]P_{11}=[1,0,\sqrt{2},0]. At this point, Kξ1K^{\xi_{1}} annihilates ω𝒱\omega_{\mathcal{V}} which means that evaluating eq. 3.5 at P11P_{11} yields

12|cos(2θ)|=Kξ1|ω𝒱(Kξ2,Kξ3)ν|=12.\frac{1}{2}|\cos(2\theta)|=\|K^{\xi_{1}}\||\frac{\omega_{\mathcal{V}}(K^{\xi_{2}},K^{\xi_{3}})}{\nu}|=\frac{1}{2}.

Hence θ=0\theta=0 and 𝒪11\mathcal{O}_{11} is diffeomorphic to S3S^{3}. It is also the orbit of the larger group S1×Sp(1)S^{1}\times\mathrm{Sp}(1).

Lemma 4.7.

The unique special Lagrangian invariant under K1K_{1} is 𝒪11\mathcal{O}_{11} which is identified with example 3.15.

Remark 4.8.

The multi-moment map for 𝕋2\mathbb{T}^{2} torus symmetry is an eigenfunction of the Laplace operator on MM, cf. [RS19, Lemma 3.1]. However, this is not the case for the SU(2)\mathrm{SU}(2) multi-moment map ν\nu, as it is non-positive everywhere, so MνvolM<0\int_{M}\nu\mathrm{vol}_{M}<0, this integral vanishes for eigenfunctions of Laplace operator.

4.1.2 K2K_{2}

The group K2K_{2} lies inside U(2)Sp(2)\mathrm{U}(2)\subset\mathrm{Sp}(2). Let ξ0=diag(i,i)𝔰𝔭(2)\xi_{0}=\mathrm{diag}(i,i)\in\mathfrak{sp}(2), which commutes with all elements in the Lie algebra of K2K_{2}. Again, we compute

Kξ0\displaystyle K^{\xi_{0}} =2Re(iZ1Z1)Kξ1=2Re(2i(Z1Z1+Z2Z2)),\displaystyle=2\mathop{\rm Re}\nolimits(iZ_{1}\frac{\partial}{\partial Z_{1}})\qquad K^{\xi_{1}}=2\mathop{\rm Re}\nolimits(-2i(Z_{1}\frac{\partial}{\partial Z_{1}}+Z_{2}\frac{\partial}{\partial Z_{2}})),
Kξ2+iKξ3\displaystyle K^{\xi_{2}}+iK^{\xi_{3}} =2(Z3Z1Z2Z1(1+Z22)Z2(Z1+Z2Z3)Z3).\displaystyle=2(Z_{3}-Z_{1}Z_{2}\frac{\partial}{\partial Z_{1}}-(1+Z_{2}^{2})\frac{\partial}{\partial Z_{2}}-(Z_{1}+Z_{2}Z_{3})\frac{\partial}{\partial Z_{3}}).

For K2K_{2}, the map ν\nu is equal to

8|Z0Z1+Z2Z3|2|Z|2-8\frac{|Z_{0}Z_{1}+Z_{2}Z_{3}|^{2}}{|Z|^{2}}

by eq. 1.4. We apply 4.6 and compute μ\mu on the set Z2=0Z_{2}=0 and Z1=r0Z_{1}=r\geq 0, and w.l.o.g we assume Z0=1Z_{0}=1. Then we have

μ1=2|Z|4(1+r4+4|Z3|2|Z3|4),μ2iμ3=4i|Z|4rZ3(2+r2+|Z3|2).\displaystyle\mu_{1}=-2|Z|^{-4}(-1+r^{4}+4|Z_{3}|^{2}-|Z_{3}|^{4}),\quad\mu_{2}-i\mu_{3}=-4i|Z|^{-4}rZ_{3}(-2+r^{2}+|Z_{3}|^{2}).

The set ν=0\nu=0 is a J1J_{1}-holomorphic quadric and hence diffeomorphic to S2×S2S^{2}\times S^{2}. The action of U(2)\mathrm{U}(2) on this quadric is of cohomogeneity one. The principal orbit is S2×S1S^{2}\times S^{1} and the singular orbit S2S^{2}.

If ν=0\nu=0 then ν\nu vanishes if r=0r=0 and |Z3|=2+3|Z_{3}|=\sqrt{2+\sqrt{3}}. Denote this point by P21P_{21}, the U(2)\mathrm{U}(2) orbit 𝒪21\mathcal{O}_{21} is special Lagrangian and Kξ0K^{\xi_{0}} is horizontal on ν1(0)\nu^{-1}(0). We compute via eq. 3.5

12|cos(2θ)|=Kξ0|ω𝒱(Kξ2,Kξ3)||Imψ(Kξ0,Kξ2,Kξ3)|=12,\frac{1}{2}|\cos(2\theta)|=\|K^{\xi_{0}}\|\frac{|\omega_{\mathcal{V}}(K^{\xi_{2}},K^{\xi_{3}})|}{|\mathop{\rm Im}\nolimits\psi(K^{\xi_{0}},K^{\xi_{2}},K^{\xi_{3}})|}=\frac{1}{2},

i.e. θ=0\theta=0.

If ν0\nu\neq 0 then r0r\neq 0 and ν=0\nu=0 only occurs for Z3=0Z_{3}=0 and r=1r=1. Denote this point by P22P_{22} and note ν𝒱\nu_{\mathcal{V}} vanishes on P22P_{22}. Hence, θ=π/4\theta=\pi/4 and the orbit 𝒪22\mathcal{O}_{22} is diffeomorphic to 3\mathbb{RP}^{3}.

Lemma 4.9.

All special Lagrangians that admit a K2K_{2} action are 𝒪21,𝒪22\mathcal{O}_{21},\mathcal{O}_{22} which corresponds to example 3.14 and example 3.9 respectively.

4.1.3 K3K_{3}

To compute the Killing vector fields for K3K_{3} we need the explicit description of K3SU(4)K_{3}\subset\mathrm{SU}(4)

K3={(a33a2b¯3ab¯2b¯33a2ba(|a|22|b|2)b¯(2|a|2|b|2)3a¯b¯23ab2b(2|a|2|b|2)a¯(|a|22|b|2)3a¯2b¯b33a¯b23a¯2ba¯3)(a,b)S32},\displaystyle K_{3}=\left\{\left(\begin{array}[]{cccc}a^{3}&-\sqrt{3}a^{2}\overline{b}&\sqrt{3}a\overline{b}^{2}&-\overline{b}^{3}\\ \sqrt{3}a^{2}b&a(|a|^{2}-2|b|^{2})&-\overline{b}(2|a|^{2}-|b|^{2})&\sqrt{3}\overline{a}\overline{b}^{2}\\ \sqrt{3}ab^{2}&b(2|a|^{2}-|b|^{2})&\overline{a}(|a|^{2}-2|b|^{2})&-\sqrt{3}\overline{a}^{2}\overline{b}\\ b^{3}&\sqrt{3}\overline{a}b^{2}&\sqrt{3}\overline{a}^{2}b&\overline{a}^{3}\end{array}\right)\mid(a,b)\in S^{3}\subset\mathbb{C}^{2}\right\},

see for example [Kaw18]. Now we can compute the Killing vector fields for K3K_{3} on 𝔸0\mathbb{A}_{0}

Kξ1=\displaystyle K^{\xi_{1}}= 2Im(3Z1Z1+2Z2Z2+Z3Z3)\displaystyle 2\mathop{\rm Im}\nolimits(3Z_{1}\frac{\partial}{\partial Z_{1}}+2Z_{2}\frac{\partial}{\partial Z_{2}}+Z_{3}\frac{\partial}{\partial Z_{3}})
Kξ2=\displaystyle K^{\xi_{2}}= Re(3(Z1Z3+Z2)Z1+(3Z1(2+3Z2)Z3)Z2\displaystyle\mathop{\rm Re}\nolimits(-\sqrt{3}(Z_{1}Z_{3}+Z_{2})\frac{\partial}{\partial Z_{1}}+(\sqrt{3}Z_{1}-(2+\sqrt{3}Z_{2})Z_{3})\frac{\partial}{\partial Z_{2}}
+(2Z23(1+Z32))Z3)\displaystyle+(2Z_{2}-\sqrt{3}(1+Z_{3}^{2}))\frac{\partial}{\partial Z_{3}})
Kξ3=\displaystyle K^{\xi_{3}}= Im(3(Z1Z3+Z2)Z1+(3Z1+(23Z2)Z3)Z2\displaystyle\mathop{\rm Im}\nolimits(\sqrt{3}(-Z_{1}Z_{3}+Z_{2})\frac{\partial}{\partial Z_{1}}+(\sqrt{3}Z_{1}+(2-\sqrt{3}Z_{2})Z_{3})\frac{\partial}{\partial Z_{2}}
+(2Z23(1+Z32))Z3).\displaystyle+(2Z_{2}-\sqrt{3}(-1+Z_{3}^{2}))\frac{\partial}{\partial Z_{3}}).

Again, we apply 4.6 and restrict ourselves to compute ν\nu and ν\nu for Z0=1,Z2=r>0Z_{0}=1,Z_{2}=r>0 and Z3=0Z_{3}=0. Let furthermore Z1=exp(iϕ)sZ_{1}=\exp(i\phi)s, then by eq. 1.4

μ1\displaystyle\mu_{1} =2|Z|4(5r44r2s216r23s4+3)\displaystyle=2|Z|^{-4}\left(5r^{4}-4r^{2}s^{2}-16r^{2}-3s^{4}+3\right)
μ2\displaystyle\mu_{2} =|Z|44rssin(ϕ)(r(3r9)+3(s28))\displaystyle=|Z|^{-4}4rs\sin(\phi)\left(r(\sqrt{3}r-9)+\sqrt{3}\left(s^{2}-8\right)\right)
μ3\displaystyle\mu_{3} =|Z|44rscos(ϕ)(r(3r9)3(s28))\displaystyle=|Z|^{-4}4rs\cos(\phi)\left(r(-\sqrt{3}r-9)-\sqrt{3}\left(s^{2}-8\right)\right)
ν\displaystyle\nu =|Z|38(4r4(s25)123r3s2cos(2ϕ)+3r2(s2+4)9(s4+s2)).\displaystyle=|Z|^{-3}8\left(4r^{4}\left(s^{2}-5\right)-12\sqrt{3}r^{3}s^{2}\cos(2\phi)+3r^{2}\left(s^{2}+4\right)-9\left(s^{4}+s^{2}\right)\right).

Hence, the only solutions of μ=(0,0,0)\mu=(0,0,0) are (r,s){(0,1),(3,0),(1/5,0)}(r,s)\in\{(0,1),(\sqrt{3},0),(1/{\sqrt{5}},0)\}. The solutions with r=0r=0 are in the U(1)\mathrm{U}(1) orbit of the point P31=[1,0,1,0]P_{31}=[1,0,1,0]. The point [1,3,0,0][1,\sqrt{3},0,0] is also in the same K3K_{3} orbit as P31P_{31}. So, it suffices to consider the points P31P_{31} and P32=[1,1/5,0,0]P_{32}=[1,1/{\sqrt{5}},0,0].

Note that ν(P31)=18\nu(P_{31})=-18 and ν(P32)=200/27\nu(P_{32})=200/27 which must hence be the minimum and maximum of ν\nu respectively. The map μ𝒱\mu_{\mathcal{V}} vanishes at P31P_{31} and hence the orbit 𝒪31\mathcal{O}_{31} satisfies θ=0\theta=0 and is in fact the Chiang Lagrangian.

Furthermore, μ𝒱(P32)=(149,0,0)\mu_{\mathcal{V}}(P_{32})=(-\frac{14}{9},0,0) which means that Kξ1K^{\xi_{1}} is horizontal at P32P_{32}. By eq. 3.5 we have

12|cos(2θ)|=Kξ1|ν𝒱||ν|=7510,θ=12arccos(7255)0.24\frac{1}{2}|\cos(2\theta)|=\|K^{\xi_{1}}\|\frac{|\nu_{\mathcal{V}}|}{|\nu|}=\frac{7}{5\sqrt{10}},\quad\theta=\frac{1}{2}\arccos(\frac{7\sqrt{2}}{5\sqrt{5}})\approx 0.24

on 𝒪32\mathcal{O}_{32}.

Lemma 4.10.

All K3K_{3} invariant special Lagrangians are given by the orbits 𝒪31\mathcal{O}_{31} and 𝒪32\mathcal{O}_{32}, which correspond to example 3.10 and example 3.17 respectively.

As remarked after 1.2, the identity component of nearly Kähler automorphisms of 3\mathbb{CP}^{3} is Sp(2)\mathrm{Sp}(2). So, combining all results of this section results in the following theorem.

Theorem 4.11.

Every Special Lagrangian in 3\mathbb{CP}^{3} 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 θ\theta Group orbit Stabiliser group of CC 3.15 Berger Sphere 0 K1K_{1} SO(2)\mathrm{SO}(2) 3.14 S1×S2S^{1}\times S^{2} 0 U(2)K2\mathrm{U}(2)\supset K_{2} SO(2)\mathrm{SO}(2) 3.9 standard 3\mathbb{RP}^{3} π/4\pi/4 K2K_{2} SO(3)\mathrm{SO}(3) (tot. geodesic) 3.10 Chiang Lagrangian π/4\pi/4 K3K_{3} S3S_{3} 3.17 distinct Ric\mathrm{Ric} e’values 0.24\approx 0.24 K3K_{3} 2\mathbb{Z}_{2}

For the definition of the SU(2)\mathrm{SU}(2) subgroups KiK_{i} 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 𝔽=SU(3)/𝕋2\mathbb{F}=\mathrm{SU}(3)/{\mathbb{T}^{2}}. 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 3\mathbb{CP}^{3}. However, the statement can also be shown by analysing the group actions of 3-dimensional subgroups of SU(3)\mathrm{SU}(3), 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 SU(3)\mathrm{SU}(3) on the flag manifold we exploit the fact that the flag manifold is an adjoint orbit for SU(3)\mathrm{SU}(3).

Up to conjugation there are two three-dimensional subgroups of SU(3)\mathrm{SU}(3) the standard SO(3)SU(3)\mathrm{SO}(3)\subset\mathrm{SU}(3) and the SU(2)\mathrm{SU}(2) subgroup fixing the element (0,0,1)(0,0,1). Consider the adjoint action of SU(3)\mathrm{SU}(3) on its lie algebra 𝔰𝔲(3)\mathfrak{su}(3). Every element A𝔰𝔲(3)A\in\mathfrak{su}(3) is then conjugate to a diagonal matrix, the SU(3)\mathrm{SU}(3) orbits are distinguished by the set of purely imaginary eigenvalues.

Hence, the SU(3)\mathrm{SU}(3) orbits are the level sets of the functions

ρ1(A)=Tr(A2)ρ2(A)=Tr(A3).\displaystyle\rho_{1}(A)=\mathrm{Tr}(A^{2})\quad\rho_{2}(A)=\mathrm{Tr}(A^{3}).

There are three orbit types. The principal stabiliser type is a maximal torus in SU(3)\mathrm{SU}(3). Every element with distinct eigenvalues is of principal type. If AA has a repeated eigenvalue the stabiliser type is SU(2)×U(1)\mathrm{SU}(2)\times\mathrm{U}(1), unless all eigenvalues are zero.

So we fix an element A𝔰𝔲(3)A\in\mathfrak{su}(3) with distinct eigenvalues and identify the flag manifold SU(3)/𝕋2\mathrm{SU}(3)/\mathbb{T}^{2} with the adjoint orbit of AA. Our aim is to determine the set of elements in 𝔽\mathbb{F} with one-dimensional stabiliser under the action of SO(3)SU(3)\mathrm{SO}(3)\subset\mathrm{SU}(3) and SU(2)\mathrm{SU}(2).

Proposition 4.12.

Every element in 𝔰𝔲(3)\mathfrak{su}(3) with non-principal stabiliser for either the action of SO(3)\mathrm{SO}(3) or SU(2)\mathrm{SU}(2) has a representative in the set

S={(iμλ0λiμ0002iμ)λ,μ}.S=\left\{\left(\begin{array}[]{ccc}i\mu&-\lambda&0\\ \lambda&i\mu&0\\ 0&0&-2i\mu\\ \end{array}\right)\mid\lambda,\mu\in\mathbb{R}\right\}.
Proof.

With respect to SO(3)\mathrm{SO}(3) the adjoint action on 𝔰𝔲(3)\mathfrak{su}(3) splits as Λ2(3)iS02(3)\Lambda^{2}(\mathbb{R}^{3})\oplus iS^{2}_{0}(\mathbb{R}^{3}). It is known that the action of SO(3)\mathrm{SO}(3) on S02(3)S^{2}_{0}(\mathbb{R}^{3}) is irreducible and has trivial stabiliser unless the element in S02(3)S^{2}_{0}(\mathbb{R}^{3}) has repeated eigenvalues, in which case the stabiliser is O(2)\mathrm{O}(2). Every such element is conjugate to diag(μ,μ,2μ)\mathrm{diag}(\mu,\mu,-2\mu). Let AΛ2(3)A\in\Lambda^{2}(\mathbb{R}^{3}), the stabiliser of AA in SO(3)\mathrm{SO}(3) intersects the stabiliser of diag(μ,μ,2μ)\mathrm{diag}(\mu,\mu,-2\mu) in a one-dimensional set if and only if AA is of the form

Aλ=(0λ0λ00000),A_{\lambda}=\left(\begin{array}[]{ccc}0&-\lambda&0\\ \lambda&0&0\\ 0&0&0\\ \end{array}\right),

which implies the statement for SO(3)\mathrm{SO}(3).

With respect to the subgroup SU(2)\mathrm{SU}(2), the representation splits as 𝔰𝔲(2)2\mathfrak{su}(2)\oplus\mathbb{C}^{2}\oplus\mathbb{R} 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 2\mathbb{C}^{2} has to vanish. The trivial component is spanned by the element diag(μ,μ,2μ)\mathrm{diag}(\mu,\mu,-2\mu). Finally, every element in 𝔰𝔲(2)\mathfrak{su}(2) is conjugate to AλA_{\lambda} under the SU(2)\mathrm{SU}(2) action. ∎

Theorem 4.13.

There are no special Lagrangians in 𝔽\mathbb{F} 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 𝔽\mathbb{F} with one-dimensional stabiliser is two-dimensional.

The identity component of nearly Kähler automorphisms of the flag manifold is SU(3)\mathrm{SU}(3). Since SU(2)\mathrm{SU}(2) and SO(3)\mathrm{SO}(3) are the only three-dimensional subgroups of SU(3)\mathrm{SU}(3) it suffices to show that the intersection 𝔽S\mathbb{F}\cap S is finite. Since SS is two-dimensional it only remains to check that the function ρ=(ρ1,ρ2)\rho=(\rho_{1},\rho_{2}) has full rank on SS. A direct computation shows that the determinant of the Jacobian is

det(Jρ)=24iλ3216iλμ2\det(J_{\rho})=24i\lambda^{3}-216i\lambda\mu^{2}

which vanishes if and only if λ{0,3μ,3μ}\lambda\in\{0,3\mu,-3\mu\}. In each case, the resulting element in SS has repeated eigenvalues, so it does not lie in 𝔽\mathbb{F} since AA 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 JJ-holomorphic curves in nearly Kaehler 3\mathbb{CP}^{3} 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 S3×S3S^{3}\times S^{3} from minimal surfaces in S3S^{3} 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 n\mathbb{CP}^{n} 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 S3×S3S^{3}\times S^{3} In Geom Dedicata 200 Springer, 2019, pp. 351–362
  • [FH17] Lorenzo Foscolo and Mark Haskins “New G2G_{2}-holonomy cones and exotic nearly Kähler structures on S6S^{6} and S3×S3S^{3}\times S^{3} 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 S3S^{3} into n\mathbb{CP}^{n} 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 G2G_{2}-manifolds” In Asian Journal of Mathematics 21, 2017, pp. 429–462
  • [Kaw18] Kotaro Kawai “Second-Order deformations of associative submanifolds in nearly parallel G2G_{2}-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 𝔽1,2(3)\mathbb{F}_{1,2}(\mathbb{C}^{3}) 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 CP3\textbf{CP}^{3} 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