1 Introduction

We study the quaternionic contact structure [3] (qc-structure for short) on \(4n+3\)-manifolds X to construct quaternionic Hermitian 4n-manifolds as their quotients. In the previous paper [2], we studied a qc-structure \(({{\mathsf {D}}},Q)\) whose \(\mathrm{Im}\, {\mathbb {H}}\)-valued (globally defined) 1-form \(\omega\) representing \({{\mathsf {D}}}\) satisfies that each distribution defined by \(\displaystyle {\text {d}}\omega _\alpha +2\omega _\beta \wedge \omega _\gamma =0\) \(((\alpha ,\beta ,\gamma )\sim (1,2,3))\) has the three-dimensional common kernel on X. \(({{\mathsf {D}}},\omega , Q)\) is called a quaternionic CR-structure (cf. [2, Definition 2.1]). It has shown in [1, 2] that every positive definite quaternionic CR-structure \((X,(\omega ,Q))\) induces a 3-Sasaki manifold. Then, X admits a (local) principal \(\mathrm{Sp}(1)\)-bundle : \(\mathrm{Sp}(1){\rightarrow }X{\longrightarrow }X/\mathrm{Sp}(1)\) over a quaternionic Kähler orbifold \(X/\mathrm{Sp}(1)\). In particular, according to the results [8, 9] of Biquard’s connection [3], X is a qc-Einstein manifold with nonzero qc-scalar curvature. For the remaining case of vanishing qc-scalar curvature, there is no nondegenerate quaternionic CR-structure on X since the integrability of quaternionic CR-structure does not hold. Taking into account these results, we shall interpret a qc-Einstein manifold with vanishing qc-scalar curvature in terms of the differentiable equations of the contact forms \(\omega _\alpha\) \((\alpha =1,2,3)\). Given a quaternionic contact manifold X, let

$$\begin{aligned} {{\mathsf {E}}}=\{\xi \in TX\ \mid {\text {d}}\omega _1(\xi ,\varvec{v})={\text {d}}\omega _2(\xi ,\varvec{v})= {\text {d}}\omega _3(\xi ,\varvec{v})=0,\, {}^\forall \, \varvec{v}\in TX\} \end{aligned}$$
(1.1)

be the distribution on X. If \({{\mathsf {E}}}\) has the three-dimensional kernel, then we call \(({{\mathsf {D}}},\omega ,Q)\) a strict qc-structure on X.

When X is a qc-Einstein manifold with vanishing qc-scalar curvature, it follows from Lemma 6.4 [8] (also (1) of Proposition 6.3) that the Reeb fields \(\{\xi _\alpha \}_{\alpha =1,2,3}\) of \(\omega\) are Killing and generate a (local) abelian Lie group (that is, \([\xi _\alpha ,\xi _\beta ]=0\)), it is easy to see that \({{\mathsf {E}}}=\{\xi _\alpha \}_{\alpha =1,2,3}\). Thus, a qc-Einstein manifold with vanishing qc-scalar curvature is a strict qc-manifold. Conversely, if X is a strict qc-manifold, then we prove in Proposition 2.5 of Sect. 2 that \({{\mathsf {E}}}\) generates a three-dimensional local abelian Lie group \({{\mathcal {R}}}\) and if \({{\mathcal {R}}}\) extends to a global \({\mathbb {R}}^3\)-action on X, then there is a principal bundle : \(\displaystyle {\mathbb {R}}^3{\rightarrow }X{\mathop {{\longrightarrow }}\limits ^{{{\mathsf {p}}}}}X/{\mathbb {R}}^3\) over the hyperKähler manifold \(X/{\mathbb {R}}^3\). (This holds always locally over an appropriate neighborhood of X in case \({{\mathcal {R}}}\) is a local qc-action.) Since \(X/{\mathbb {R}}^3\) is hyperKähler (locally in general), using the pullback by \({{\mathsf {p}}}\), both qc-Ricci tensor and qc-scalar curvature of X vanish by the definition (cf. [8]), so X is a qc-Einstein manifold with vanishing qc-scalar curvature. Thus, a strict quaternionic contact manifold is the same as a qc-Einstein manifold with vanishing qc-scalar curvature. Indeed, we owe a lot to the referee who pointed out this equivalence in our earlier draft.

If a Lie group G admits a left invariant strict qc-structure, then G is called a strict qc-group. An example is the quaternionic Heisenberg nilpotent Lie group \({{\mathcal {M}}}\) with the standard qc-structure admitting a nontrivial central extension \(\displaystyle 1{\rightarrow }{\mathbb {R}}^3{\rightarrow }\, {{\mathcal {M}}}{\longrightarrow }\, {\mathbb {H}}^n{\rightarrow }1\) (cf. Sect. 3.2). We construct a family of simply connected strict qc solvable Lie subgroups \({{\mathcal {M}}}(k,\ell )\) of \({{\mathcal {M}}}\rtimes T^n\) where \(k+\ell =n\), \(T^n\le \mathop {\mathrm{Sp}}\nolimits (n)\), (see Sect. 3.3, cf. [4]).

Theorem A

If G is a contractible unimodular strict qc-group, then G is isomorphic to \({{\mathcal {M}}}(k,\ell )\).

A \(4n+3\)-dimensional qc-manifold X is uniformizable (or spherical) if X is locally modeled over \((\mathop {\mathrm{PSp}}\nolimits (n+1,1), S^{4n+3})\). (This is the case \(W^{qc}=0\), see [10] also.) The pair \((\mathop {\mathrm{PSp}}\nolimits (n+1,1), S^{4n+3})\) is obtained from projective compactification of the complete simply connected quaternionic hyperbolic space \({\mathbb {H}}^{n+1}_{\mathbb {H}}\) with \(\mathop {\mathrm{Isom}}\nolimits ({\mathbb {H}}^{n+1}_{\mathbb {H}})=\mathop {\mathrm{PSp}}\nolimits (n+1,1)\).

Denote by \(\mathop {\mathrm{Aut}}\nolimits _{qc}(X)\) the group of qc-transformations of X. If there exists a discrete subgroup \(\Gamma \le \mathop {\mathrm{Aut}}\nolimits _{qc}(X)\) acting properly with compact quotient \(X/\Gamma\), then X is said to be divisible (cf. Definition 4.3). The following result [13, Theorem 1.1] was proved for the compact case.

Theorem B

Let M be a \((4n+3)\)-dimensional compact uniformizable strict qc-manifold. Then, M is qc-conformal to the quaternionic infranilmanifold \({\mathcal {M}}/\Gamma\) (some finite cover of which is a principal \(T^3\)-bundle over the quaternionic flat torus \(T^n_{\mathbb {H}}\).)

The following uniqueness theorem characterizes especially the noncompact case (cf. Theorem 4.4).

Theorem C

Let \((X,{{\mathsf {D}}},\omega ,\{J_\alpha \}_{\alpha =1}^3)\) be a noncompact simply connected uniformizable strict qc-manifold. Put \({{\mathsf {E}}}=\{\xi _1,\xi _2,\xi _3\}\). Suppose X is divisible by \(\Gamma\).

(1) If \(\mathop {\mathrm{Aut}}\nolimits _{qc}(X)\) leaves \({{\mathsf {E}}}\) invariant, then the developing pair reduces to the equivariant immersion:

$$\begin{aligned} (\rho ,\mathop {\mathop {\mathrm{dev}}\nolimits }):(\mathop {\mathrm{Aut}}\nolimits _{qc}(X),X){\rightarrow }(\mathop {\mathrm{Aut}}\nolimits _{qc}({{\mathcal {M}}}),{{\mathcal {M}}}). \end{aligned}$$

In addition,

(2) For any \(\gamma \in \Gamma\) and \(\alpha =1,2,3\), suppose \(\displaystyle \gamma _*\xi _\alpha =\sum \nolimits _{\beta =1}^3a_{\alpha \beta }\xi _\beta\) for some function \(a_{\alpha \beta }:X{\rightarrow }\mathrm{SO}(3)\). Then,

  1. (i)

    \(\mathop {\mathrm{dev}}\nolimits : X{\rightarrow }{{\mathcal {M}}}\) is a qc-diffeomorphism so that \({{\mathcal {R}}}={\mathbb {R}}^3\).

  2. (ii)

    There exists a strict qc-structure \(({{\mathsf {D}}}, \eta ,\{J_\alpha '\}_{\alpha =1}^3)\) qc-conformal to \((\omega ,\{J_\alpha \}_{\alpha =1}^3)\). The quotient \((X/{\mathbb {R}}^3, \{\Theta _\alpha ,{\hat{J}}_\alpha '\}_{\alpha =1}^3)\) is a hyperKähler manifold isometric to \({\mathbb {H}}^n\).

For the difference between Theorem B and Theorem C, we remark that in Theorem B there is a \(T^3\)-action on \(X/\Gamma\) which lifts to X an \({\mathbb {R}}^3\)-action centralizing \(\Gamma\), while in Theorem C X is divisible by \(\Gamma\), but the intersection \({\mathbb {R}}^3\cap \Gamma\) is not necessarily uniform in \({\mathbb {R}}^3\), which does not imply to induce a \(T^3\)-action on \(X/\Gamma\).

The second part of this paper treats the quaternionic Hermitian quotient in place of the hyperKähler quotient. We construct a noncompact qc-manifold to obtain a quaternionic Hermitian manifold \((Y,\{\hat{\Omega }_\alpha ,{\hat{J}}_\alpha \}_{\alpha =1}^3)\) such that one of \((\hat{\Omega }_\alpha ,{\hat{J}}_\alpha )\)’s is Kähler. (See Theorem 6.5, Theorem 7.2, Corollary 7.3.)

Theorem D

There exists a uniformizable noncompact qc-manifold X whose quotient by the \({\mathbb {R}}^3\)-action gives a 4n-dimensional quaternionic Hermitian manifold \((Y, {\hat{g}}, \{\hat{\Omega }_\alpha , {\hat{J}}_\alpha \}_{\alpha =1}^3)\). Moreover,

  1. 1.

    \((Y, {\hat{g}} , \{\hat{\Omega }_1, {\hat{J}}_1\})\) is a Bochner flat complex Kähler manifold.

  2. 2.

    \((Y,{\hat{g}})\) is not Einstein. In particular, Y is not isometric to any domain of the quaternionic euclidean space \({\mathbb {H}}^n\).

  3. 3.

    The quaternionic Hermitian isometry group \(\displaystyle \mathop {\mathrm{Isom}}\nolimits _{qH}(Y,{\hat{g}}, \{{\hat{J}}_\alpha \}_{\alpha =1}^3)\) is isomorphic to a k-torus \(T^k\) for some k where \(n+1\le k\le 2n\).

The paper is organized as follows In Sect. 2, we give some basic facts on strict qc-structure. The fundamental property of strict qc-manifolds is proved in Proposition 2.5 which produces hyperKähler structures on their \({\mathbb {R}}^3\)-quotients as mentioned. In Sect. 3, we review quaternionic Heisenberg nilpotent Lie group \({{\mathcal {M}}}\) where the group structure and qc-structure are explained explicitly. We give a nontrivial strict qc-group as a qc manifold in Theorem 3.3. From another viewpoint, we discuss strict qc manifolds in connection with spherical (uniformizable) qc geometry \((\mathop {\mathrm{PSp}}\nolimits (n+1,1), S^{4n+3})\) in Sect. 4. Theorem 4.4 gives a sufficient condition for a divisible group \(\Gamma\) of the qc-automorphism group \(\mathop {\mathrm{Aut}}\nolimits _{qc}(X)\) characterizing that the quotient \(X/{\mathbb {R}}^3\) may be isometric to \({\mathbb {H}}^n\) as the standard hyperKähler manifold. In Sect. 5, we relax the condition strict on \({{\mathsf {E}}}\) in order to get a quaternionic Hermitian structure \(({\hat{g}},\{\hat{\Omega }_\alpha , {\hat{J}}_\alpha \}_{\alpha =1}^3)\) on the quotient domain \(Y=X/{\mathbb {R}}^3\) of \({\mathbb {H}}^n\). This can be achieved by the conformal change of the \(\mathrm{Im}\, {\mathbb {H}}\)-valued one-form \(\omega _0\) which represents the standard qc-structure on \({{\mathcal {M}}}\). We can show that one of them, say \(({\hat{g}},{\hat{\Omega }}_1, {\hat{J}}_1)\) is a Kähler metric on Y. Moreover, in Sect. 6 a prominent property of this construction is that \((Y, {\hat{\Omega }}_1, {\hat{J}}_1)\) admits a Bochner flat Kähler structure. In particular, Y is not locally isometric to any domain of the flat space \({\mathbb {H}}^n\). In Sect 7, we discuss the quaternionic isometry group \(\mathop {\mathrm{Isom}}\nolimits _{\!\, qH}(Y, {\hat{g}},\{\hat{\Omega }_\alpha , {\hat{J}}_\alpha \}_{\alpha =1}^3)\). In course of discussion, we obtain a strictly pseudoconvex spherical pseudo-Hermitian structure \(\{{\hat{\omega }}_1, J_1\}\) on the \((4n+1)\)-quotient \(X/{\mathbb {R}}^2\) such that the pseudo-Hermitian transformation group \(\mathop {\mathrm{Psh}}\nolimits (X/{\mathbb {R}}^2)\) is isomorphic to \({\mathbb {R}}\times T^{2n}\). Theorem D is a consequence of the results of Sects. 6 and 7.1.

2 Strict quaternionic contact manifolds

The hypercomplex structure \(\{J_\alpha ,J_\beta ,J_\gamma \}\) on \({{\mathsf {D}}}\) is defined by the following equation \(((\alpha , \beta ,\gamma ) \sim (1,2,3))\):

$$\begin{aligned} {\text {d}}\omega _\alpha (\varvec{u},\varvec{v})={\text {d}}\omega _\beta (J_\gamma \varvec{u},\varvec{v}) \ \, ({}^\forall \,\varvec{u},\varvec{v}\in {{\mathsf {D}}}). \end{aligned}$$
(2.1)

There is the reciprocity on \({{\mathsf {D}}}\):

$$\begin{aligned} \begin{aligned} {\text {d}}\omega _\alpha (J_\alpha \varvec{u},\varvec{v})={\text {d}}\omega _\beta (J_\beta \varvec{u},\varvec{v}) ={\text {d}}\omega _\gamma (J_\gamma \varvec{u},\varvec{v})\ \, ((\alpha , \beta ,\gamma ) \sim (1,2,3)). \end{aligned} \end{aligned}$$
(2.2)

It is easy to see from (2.2)

$$\begin{aligned} {\text {d}}\omega _\alpha (J_\alpha \varvec{u},J_\alpha \varvec{v})={\text {d}}\omega _\alpha (\varvec{u},\varvec{v})\ \, (\alpha =1,2,3). \end{aligned}$$
(2.3)

2.1 Strict qc-manifolds

Let \((X,D,\omega ,\{J_\alpha \}_{\alpha =1}^3)\) be a strict qc-manifold with distribution \({{\mathsf {E}}}\) (cf. (1.1)).

Lemma 2.1

\({{\mathsf {E}}}\) generates a three-dimensional local abelian Lie group \({{\mathcal {R}}}\).

Proof

Since \({{\mathsf {E}}}\) is of dimension 3 and transverse to \({{\mathsf {D}}}\), it follows \(\omega ({{\mathsf {E}}})=\mathrm{Im}\,{\mathbb {H}}\). There exist vector fields \(\{\xi _\alpha \}_{\alpha =1,2,3}\subset {{\mathsf {E}}}\) such that

$$\begin{aligned} \omega _\alpha (\xi _\beta )=\delta _{\alpha \beta }. \end{aligned}$$
(2.4)

(Equivalently \(\displaystyle \omega (\xi _\alpha )= \omega _1(\xi _\alpha )i+\omega _2(\xi _\alpha )j+ \omega _3(\xi _\alpha )k=\delta _{1\alpha }i+\delta _{2\alpha }j+\delta _{3\alpha }k\).) By (1.1), \(\displaystyle 2{\text {d}}\omega _\alpha (\xi _\beta ,\xi _\gamma )=-\omega _\alpha ( [\xi _\beta ,\xi _\gamma ])=0\) and so \([\xi _\beta ,\xi _\gamma ]\in {{\mathsf {D}}}\). For any \(\varvec{v}\in {{\mathsf {D}}}\), \(0=2{\text {d}}\omega _\alpha (\xi _\beta ,\varvec{v})=-\omega _\alpha ([\xi _\beta ,\varvec{v}])\) \((\alpha =1,2,3)\) so \([\xi _\beta ,\varvec{v}]\in {{\mathsf {D}}}\) \(((\alpha ,\beta ,\gamma )\sim (1,2,3))\). Using the Jacobi identity,

$$\begin{aligned} \begin{aligned} 2{\text {d}}\omega _\alpha ([\xi _\beta ,\xi _\gamma ],\varvec{v})&=-\omega _\alpha ([\,[\xi _\beta ,\xi _\gamma ],\varvec{v}]) =\omega _\alpha ([\,[\xi _\gamma ,\varvec{v}],\xi _\beta ])+\omega _\alpha ([\,[\varvec{v},\xi _\beta ],\xi _\gamma ])\\&=-2{\text {d}}\omega _\alpha ([\xi _\gamma ,\varvec{v}],\xi _\beta )-2{\text {d}}\omega _\alpha ([\varvec{v},\xi _\beta ],\xi _\gamma )=0.\\ \end{aligned} \end{aligned}$$

By the non-degeneracy of \({\text {d}}\omega _\alpha\) on \({{\mathsf {D}}}\), it follows \([\xi _\beta ,\xi _\gamma ]=0\) for any \(\beta ,\gamma\). Thus, \({{\mathsf {E}}}=\{\xi _\alpha , \alpha =1,2,3\}\) generates a local abelian Lie group. \(\square\)

Proposition 2.2

Denote by \({{\mathcal {L}}}_\xi\) the Lie derivative of a vector field \(\xi\) on X.

  1. (1)

    \({{\mathcal {L}}}_{\xi _\alpha }\omega _\beta =0\), \({{\mathcal {L}}}_{\xi _\alpha }{\text {d}}\omega _\beta =0\)    \((\alpha ,\beta =1,2,3)\). In particular, \({{\mathcal {L}}}_{\xi _\alpha }{{\mathsf {D}}}={{\mathsf {D}}}\).

  2. (2)

    \({{\mathcal {L}}}_{\xi _\alpha }J_\beta =0\)  \((\alpha ,\beta =1,2,3)\).

Proof

First note that \(\displaystyle {\mathcal {L}}_{\xi _{\beta }}\omega _\alpha =({\text {d}}\iota _{\xi _{\alpha }}+\iota _{\xi _{\alpha }}d)\omega _\beta = \iota _{\xi _{\alpha }}{\text {d}}\omega _\alpha =0,\ \ {\mathcal {L}}_{\xi _{\alpha }}{\text {d}}\omega _\beta ={\text {d}}{\mathcal {L}}_{\xi _{\alpha }}\omega _\beta =0\) from (2.4), (1.1). For any \(\varvec{v}\in {{\mathsf {D}}}\), \(0=({\mathcal {L}}_{\xi _{\alpha }}\omega )(\varvec{v}) =-\omega ({\mathcal {L}}_{\xi _{\alpha }}\varvec{v})\) so \({\mathcal {L}}_{\xi _{\alpha }}\varvec{v}\in {{\mathsf {D}}}\). Thus, \(({\mathcal {L}}_{\xi _{\alpha }}J_\beta )\varvec{v}={\mathcal {L}}_{\xi _{\alpha }}(J_\beta \varvec{v})- J_\beta ({\mathcal {L}}_{\xi _{\alpha }}\varvec{v})\in {{\mathsf {D}}}\). For \({}^\forall \,\varvec{u},\varvec{v}\in {{\mathsf {D}}}\), \(\displaystyle ({\mathcal {L}}_{\xi _{\alpha }}{\text {d}}\omega _\gamma )(J_\alpha \varvec{u},\varvec{v})=0\), which equals

$$\begin{aligned}\begin{aligned} {\mathcal {L}}_{\xi _{\alpha }}({\text {d}}\omega _\gamma (J_\alpha \varvec{u},\varvec{v})) -{\text {d}}\omega _\gamma ({\mathcal {L}}_{\xi _{\alpha }}(J_\alpha \varvec{u}),\varvec{v})-{\text {d}}\omega _\gamma (J_\alpha \varvec{u},{\mathcal {L}}_{\xi _{\alpha }}\varvec{v})=0. \end{aligned}\end{aligned}$$
$$\begin{aligned} \begin{aligned} {\text {d}}\omega _\gamma (({\mathcal {L}}_{\xi _{\alpha }}J_\alpha )\varvec{u},\varvec{v})&={\text {d}}\omega _\gamma ({\mathcal {L}}_{\xi _{\alpha }}(J_\alpha \varvec{u}),\varvec{v})- {\text {d}}\omega _\gamma (J_\alpha {\mathcal {L}}_{\xi _{\alpha }}\varvec{u},\varvec{v})\\&={\mathcal {L}}_{\xi _{\alpha }}({\text {d}}\omega _\gamma (J_\alpha \varvec{u},\varvec{v}))-{\text {d}}\omega _\gamma (J_\alpha \varvec{u},{\mathcal {L}}_{\xi _{\alpha }}\varvec{v}) -{\text {d}}\omega _\gamma (J_\alpha {\mathcal {L}}_{\xi _{\alpha }}\varvec{u},\varvec{v})\\&={\mathcal {L}}_{\xi _{\alpha }}({\text {d}}\omega _\beta (\varvec{u},\varvec{v}))- {\text {d}}\omega _\beta (\varvec{u},{\mathcal {L}}_{\xi _{\alpha }}\varvec{v})-{\text {d}}\omega _\beta ({\mathcal {L}}_{\xi _{\alpha }}\varvec{u},\varvec{v})\ \, (2.1)\\&=({\mathcal {L}}_{\xi _{\alpha }}{\text {d}}\omega _\beta )(\varvec{u},\varvec{v})=0.\\ \end{aligned} \end{aligned}$$

Similarly, it follows \(\displaystyle {\text {d}}\omega _\alpha (({\mathcal {L}}_{\xi _{\alpha }}J_\beta )\varvec{u},\varvec{v})=0\). By the non-degeneracy of \({\text {d}}\omega _\alpha\) on \({{\mathsf {D}}}\), it follows \({\mathcal {L}}_{\xi _{\alpha }}J_\alpha =0,\ {\mathcal {L}}_{\xi _{\alpha }}J_\beta =0\).

Let \({{\mathcal {R}}}\) denote the local abelian group obtained from Lemma 2.1.

Proposition 2.3

Suppose that \({{\mathcal {R}}}\) generates a global abelian group of a strict qc-manifold X. Then, \({{\mathcal {R}}}\) acts properly on X as qc-transformations, that is a closed subgroup \({{\mathcal {R}}}\le \mathop {\mathrm{Aut}}\nolimits _{qc}(X)\).

Proof

By (1), (2) of Proposition 2.2, it follows \(\displaystyle t^*\omega _\alpha =\omega _\alpha ,\ t_*\circ J_\alpha =J_\alpha \circ t_*\) for any \(t\in {{\mathcal {R}}}\) \((\alpha =1,2,3)\). Define a Riemannian metric on X by

$$\begin{aligned} g(A,B)=\sum _{i=1}^{3}\omega _i(A)\cdot \omega _i(B)+{\text {d}}\omega _1(J_1 A,B)\ \ ({}^\forall \, A,B\in TX). \end{aligned}$$
(2.5)

(We may choose whichever \({\text {d}}\omega _\alpha \circ J_\alpha\) from the reciprocity \(\displaystyle {\text {d}}\omega _1\circ J_1= {\text {d}}\omega _2\circ J_2={\text {d}}\omega _3\circ J_3\).) Then, note that \({{\mathcal {R}}}\le \mathop {\mathrm{Isom}}\nolimits (X,g)\le \mathop {\mathrm{Aut}}\nolimits _{qc}(X)\). If \(\displaystyle {\overline{{{\mathcal {R}}}}}\) is the closure of \({{\mathcal {R}}}\) in \(\mathop {\mathrm{Isom}}\nolimits (X,g)\), then it acts properly on X. Let \(\varvec{\tau }\) be a vector field induced by a one-parameter subgroup of \({\overline{{{\mathcal {R}}}}}\). Then, there is a sequence of vector fields \(\{\xi ^{(n)}\}\subset {{\mathsf {E}}}\) such that \(\displaystyle {\text {d}}\omega _1(\varvec{\tau },A)=\lim _{n{\rightarrow }\infty } {\text {d}}\omega _1(\xi ^{(n)},A)=0\) \(({}^\forall \, A\in TX)\) by (1.1). And so \(\varvec{\tau }\in {{\mathsf {E}}}\). This implies \({\overline{{{\mathcal {R}}}}}={{\mathcal {R}}}\).

For example, if X is complete with respect to g of (2.5), then \({{\mathcal {R}}}\) extends to a global action of X. If a strict qc-manifold \((X,\omega ,{{\mathsf {D}}},\{J_\alpha \}_{\alpha =1}^3)\) admits a global \({\mathbb {R}}^3\)-action induced by \({{\mathsf {E}}}\), then \({\mathbb {R}}^3\) acts properly by Proposition 2.3 and hence freely on X. There is a principal bundle over a 4n-dimensional manifold \(Y=X/{\mathbb {R}}^3\): \(\displaystyle {\mathbb {R}}^3{\rightarrow }X{\mathop {{\longrightarrow }}\limits ^{\pi }}Y\). We will show that Y admits a hyperKähler metric. Since each \(t\in {\mathbb {R}}^3\) satisfies \(J_\alpha \cdot t_*=t_*\cdot J_\alpha\) on \({{\mathsf {D}}}\) by (2) of Proposition 2.2, \({\mathbb {R}}^3\) induces a well-defined almost complex structure \({\hat{J}}_\alpha\) on Y such that \(\displaystyle \pi _*\cdot J_\alpha ={\hat{J}}_\alpha \cdot \pi _*: {{\mathsf {D}}}{\rightarrow }TY\) at each point of X. \(\{{\hat{J}}_\alpha \}_{\alpha =1}^3\) constitutes a quaternionic structure on Y. Define a 2-form \(\Omega _\alpha\) \((\alpha =1,2,3)\) on Y to be

$$\begin{aligned} \pi ^*\Omega _\alpha ={\text {d}}\omega _\alpha . \end{aligned}$$
(2.6)

Proposition 2.4

The 2-form \(\Omega _\alpha\) is a well-defined closed 2-form \((\alpha =1,2,3)\) satisfying the following equality:

$$\begin{aligned} \Omega _1({\hat{J}}_1 {\hat{A}}, {\hat{B}})=\Omega _2({\hat{J}}_2{\hat{A}},{\hat{B}}) =\Omega _3({\hat{J}}_3{\hat{A}},{\hat{B}}) \, \ ({\hat{A}},{\hat{B}}\in TY). \end{aligned}$$

Moreover,

$$\begin{aligned} g({\hat{A}},{\hat{B}})=\Omega _\alpha ({\hat{J}}_\alpha {\hat{A}},{\hat{B}})\ \, (\alpha =1,2,3) \end{aligned}$$

is a hyperKähler metric on \((Y,\{{\hat{J}}_\alpha \}_{\alpha =1,2,3})\).

Proof

Let \(A=V+\varvec{u},\, B=W+\varvec{v}\in TX\) \(({}^\exists \,V,W\in {{\mathsf {E}}}={\mathsf {R}}^3\), \({}^\exists \,\varvec{u},\varvec{v}\in {{\mathsf {D}}})\). (Similarly, \(A'=V'+\varvec{u}',\, B'=W'+\varvec{v}'\).) Suppose \(\pi _*A_p=\pi _*A'_q\), \(\pi _*B_p=\pi _*B'_q\). Then, \(q=tp\in X\) for some \(t\in {\mathbb {R}}^3\) in which \(A'=t_*T+t_*A\)\(B'=t_*T'+t_*B\) \(({}^\exists \, T, T'\in {{\mathsf {R}}}^3_p={{\mathsf {E}}}_p)\). Since each element t leaves \(\omega _\alpha\) invariant by (1) of Proposition 2.2, (1.1) shows \({\text {d}}\omega _\alpha (A',B')={\text {d}}\omega _\alpha (A,B)\), thus (2.6) is well-defined. In particular,

$$\begin{aligned} {\text {d}}\omega _\alpha =0 \text{ on } \, Y. \end{aligned}$$

Furthermore, as \(V,W\in {{\mathsf {E}}}\), \(\displaystyle {\text {d}}\omega _\alpha (A,B)={\text {d}}\omega _\alpha (\varvec{u},\varvec{v})\). Since \({\hat{A}}= \pi _*A=\pi _*\varvec{u},\ {\hat{B}}= \pi _*B=\pi _* \varvec{v}\), it follows \(\displaystyle {\text {d}}\omega _\alpha (J_\alpha \varvec{u},\varvec{v})=\pi ^*\Omega _\alpha (J_\alpha \varvec{u},\varvec{v}) =\Omega _\alpha ({\hat{J}}_\alpha {\hat{A}},{\hat{B}})\). Since \({\text {d}}\omega _1(J_1 \varvec{u},\varvec{v})={\text {d}}\omega _2(J_2 \varvec{u},\varvec{v})={\text {d}}\omega _3(J_3 \varvec{u},\varvec{v})\) is positive definite on \({{\mathsf {D}}}\) [cf. (2.2)], we have a positive definite 2-form on Y:

$$\begin{aligned} g({\hat{A}},{\hat{B}})=\Omega _1({\hat{J}}_1 {\hat{A}},{\hat{B}})= \Omega _2({\hat{J}}_2 {\hat{A}},{\hat{B}})=\Omega _3({\hat{J}}_3 {\hat{A}},{\hat{B}})\ \ ({\hat{A}},{\hat{B}}\in TY). \end{aligned}$$

By (2.6), \(\displaystyle \Omega _\alpha ({\hat{J}}_\alpha {\hat{A}},{\hat{J}}_\alpha {\hat{B}})= \Omega _\alpha ({\hat{A}},{\hat{B}})\). It follows

$$\begin{aligned} g({\hat{J}}_\alpha {\hat{A}},{\hat{J}}_\alpha {\hat{B}})=g({\hat{A}},{\hat{B}}) \, \text{ on } \, Y\ \, (\alpha =1,2,3). \end{aligned}$$

By the definition, g is a hyperKähler metric on Y. \(\square\)

In summary, we obtain the result implied in Introduction.

Proposition 2.5

Let \((X,{{\mathsf {D}}}, (\omega ,\{J_\alpha \}_{\alpha =1}^3))\) be a strict qc-manifold. Let \({{\mathcal {R}}}\) be a local abelian group generated by the distribution \({{\mathsf {E}}}\). If \({{\mathcal {R}}}\) extends to a global action of \({\mathbb {R}}^3\) on X, then the quotient manifold \(Y=X/{\mathbb {R}}^3\) supports a hyperKähler structure \((g, \{\Omega _\alpha ,{\hat{J}}_{\alpha }\}_{\alpha =1}^3)\).

3 Quaternionic Heisenberg Lie group \({{\mathcal {M}}}\)

3.1 Quick review of quaternionic parabolic geometry

We recall parabolic quaternionic group derived from the quaternionic hyperbolic group. The quaternionic hyperbolic space \({\mathbb {H}}^{n+1}_{\mathbb {H}}\) has a (projective) compactification whose boundary is diffeomorphic to \(S^{4n+3}\). The isometric action of the quaternionic hyperbolic group \(\mathrm{Isom}\,({\mathbb {H}}^{n+1}_{\mathbb {H}})=\mathrm{PSp}(n+1,1)\) extends to an analytic action on \(S^{4n+3}\), which we may call a quaternionic contact action on \(S^{4n+3}\). Let \(\infty\) be the point at infinity of \(S^{4n+3}\). The standard sphere \(S^{4n+3}\) with \(\infty\) removed admits a qc-structure isomorphic to the quaternionic Heisenberg Lie group \({{\mathcal {M}}}\) with \(\mathop {\mathrm{Aut}}\nolimits _{qc}({{\mathcal {M}}})={{\mathcal {M}}}\rtimes (\mathop {\mathrm{Sp}}\nolimits (n)\cdot \mathop {\mathrm{Sp}}\nolimits (1)\times {\mathbb {R}}^+)\). Recall the definition of \({\mathcal {M}}\) from [2]. Put \(t=(t_1,t_2,t_3), s=(s_1,s_2,s_3)\in {\mathbb {R}}^3=\text{ Im }\,{\mathbb {H}}\), and \(z={}^t(z_1,\ldots ,z_n), w={}^t(w_1,\ldots ,w_n)\in {\mathbb {H}}^n\) and so on. Then, \({{\mathcal {M}}}\) is the product \({\mathbb {R}}^3\times {\mathbb {H}}^n\) with group law:

$$\begin{aligned}(t,z)\cdot (s,w)=(t+s-\text{ Im }\langle z,w\rangle , z+w)\end{aligned}$$

where \(\langle z,w\rangle ={}^t{\bar{z}}w\) is the Hermitian inner product. \({{\mathcal {M}}}\) is a nilpotent Lie group such that the center is the commutator subgroup \([{\mathcal {M}},{\mathcal {M}}]={\mathbb {R}}^3\) consisting of elements (t, 0).

Each element \(h=\big ((t,v), \sqrt{u} A\cdot a\big )\in \mathop {\mathrm{Aut}}\nolimits _{qc}({{\mathcal {M}}})\) acts on \({{\mathcal {M}}}\) as

$$\begin{aligned} h(s,z)=(t+ u a s{\bar{a}} -\text{ Im }\langle v,\sqrt{u} Az{\bar{a}}\rangle , v+\sqrt{u} Az{{\bar{a}}})\ \, ({}^\forall \,(s,z)\in {{\mathcal {M}}}). \end{aligned}$$
(3.1)

In particular, \(\displaystyle \mathrm{E}({{\mathcal {M}}})={\mathcal {M}}\rtimes \bigl (\mathrm{Sp}(n)\cdot \mathrm{Sp}(1)\bigr )\) is a normal subgroup of \(\mathop {\mathrm{Aut}}\nolimits _{qc}({{\mathcal {M}}})\) acting properly and transitively on \({{\mathcal {M}}}\) in the manner of (3.1).

3.2 The qc-structure of \({{\mathcal {M}}}\)

The \(\mathrm{Im}\,{\mathbb {H}}\)-valued 1-form on \({\mathcal {M}}\) is defined by

$$\begin{aligned} \begin{aligned} \omega _0={\text {d}}t_1i+{\text {d}}t_2j+{\text {d}}t_3k+\mathrm{Im}\langle z,{\text {d}}z\rangle . \end{aligned} \end{aligned}$$
(3.2)

Put \(\displaystyle \omega _0=\omega _1i+\omega _2j+\omega _3k\). Let \(\displaystyle {{{\mathsf {D}}}}_0= \mathop {\cap }_{i=1}^{3}\mathrm{ker}\, \omega _i=\mathrm{ker}\, \omega _0\) which denotes the codimension 3-subbundle on \({{\mathcal {M}}}\) satisfying \(\displaystyle T{{\mathcal {M}}}={{\mathsf {D}}}_0\oplus \langle \frac{\text {d}}{{\text {d}}t_1},\frac{\text {d}}{{\text {d}}t_2}, \frac{\text {d}}{\text {d}t_3}\rangle\). Let \(x_\alpha\) \((\alpha =1,\dots ,4n)\) be a real number so that \({\mathbb {R}}^{4n}\) is identified with \(\displaystyle {\mathbb {H}}^n=(x_1+ix_2+jx_3+kx_4;\ \ldots \ldots ;x_{4n-3}+ix_{4n-2}+jx_{4n-3}+kx_{4n})\). A direct calculation shows

$$\begin{aligned} \begin{aligned} \omega _1&={\text {d}}t_1+(x_1{\text {d}}x_2-x_2{\text {d}}x_1)+(x_4{\text {d}}x_3-x_3{\text {d}}x_4)+ \cdots \\&\quad \cdots +(x_{4n-3}{\text {d}}x_{4n-2}-x_{4n-2}{\text {d}}x_{4n-3})+(x_{4n}{\text {d}}x_{4n-1}-x_{4n-1}{\text {d}}x_{4n}).\\ \omega _2&=dt_2+(x_1dx_3-x_3dx_1)+(x_2dx_4-x_4{\text {d}}x_2)+ \cdots \\&\quad \cdots +(x_{4n-3}{\text {d}}x_{4n-1}-x_{4n-1}{\text {d}}x_{4n-3})+(x_{4n-2}{\text {d}}x_{4n}-x_{4n}{\text {d}}x_{4n-2}).\\ \omega _3&={\text {d}}t_3+(x_1{\text {d}}x_4-x_4{\text {d}}x_1)+(x_3{\text {d}}x_2-x_2{\text {d}}x_3)+ \cdots \\&\quad \cdots +(x_{4n-3}{\text {d}}x_{4n}-x_{4n}{\text {d}}x_{4n-3})+(x_{4n-1}{\text {d}}x_{4n-2}-x_{4n-2}{\text {d}}x_{4n-1}). \end{aligned} \end{aligned}$$
(3.3)

The hypercomplex structure \(\{J_1,J_2,J_3\}\) on \(\displaystyle {{{\mathsf {D}}}}_0\) is given as in (2.1). Alternatively if \(\pi :{{\mathcal {M}}}{\rightarrow }{\mathbb {H}}^n\) is the canonical projection (homomorphism), then \(\pi _*:{{\mathsf {D}}}_0{\rightarrow }T{\mathbb {H}}^n\) is an isomorphism at each point of \({{\mathcal {M}}}\) for which each \(J_\alpha\) on \({{\mathsf {D}}}_0\) is defined by the commutative rule:

$$\begin{aligned} \pi _*\circ J_\alpha = I_\alpha \circ \pi _* \end{aligned}$$
(3.4)

where \(\{I_\alpha \}_{\alpha =1}^3\) of the right hand side is the standard quaternionic structure \(\{i,j,k\}\) on \({\mathbb {H}}^n\), respectively.

Proposition 3.1

\(({{\mathcal {M}}}, ({{\mathsf {D}}}_0, \omega _0, \{J_\alpha \}_{\alpha =1}^3))\) is a strict qc-manifold for which

  1. (1)

    \(\displaystyle {{\mathsf {E}}}_0= \left\langle \frac{\text {d}}{{\text {d}t}_1},\frac{\text {d}}{{\text {d}}t_2}, \frac{\text {d}}{{\text {d}}t_3}\right\rangle\) generates the center \({\mathbb {R}}^3\) of \({{\mathcal {M}}}\), transverse to \({{\mathsf {D}}}_0\).

  2. (2)

    There is a principal bundle: \(\displaystyle {\mathbb {R}}^3{\rightarrow }{{\mathcal {M}}}{\mathop {{\longrightarrow }}\limits ^{\pi }}{\mathbb {H}}^n\) whose qc-structure \((\omega _0,\{J_\alpha \}_{\alpha =1}^3)\) induces the standard hyperKähler structure on \({\mathbb {H}}^n\).

Proof

It follows \(\displaystyle {\text {d}}\omega _\alpha \left( \frac{d}{dt_\beta }\right) =\delta _{\alpha \beta }\), \(\displaystyle {\text {d}}\omega _\alpha \left( \frac{d}{dt_\beta },X\right) =0\) \(({}^\forall \, X\in T{{\mathcal {M}}})\) by (3.3). And so \(\displaystyle {{\mathsf {E}}}_0=\left\langle \frac{\text {d}}{\text {d}t_1},\frac{\text {d}}{\text {d}t_2}, \frac{\text {d}}{\text {d}t_3}\right\rangle\). The remaining follows from Proposition 2.5. Explicitly, if \(g_{\mathbb {H}}\) is the standard quaternionic euclidean metric on \({\mathbb {H}}^n\), then (3.4), (3.2) show \({\text {d}}\omega _\alpha (J_\alpha X,Y)=g_{\mathbb {H}}(\pi _*X,\pi _*Y)\) \(({}^\forall \, X,Y\in {{\mathsf {D}}}_0)\). \(\square\)

Remark 3.2

There is a canonical equivariant Riemannian submersion :

(3.5)

where \(g_\omega =\mathop {\sum }_{\alpha =1}^{3}\omega _\alpha \cdot \omega _\alpha +{\text {d}}\omega _1\circ J_1\) is a Riemannian metric (cf. (2.5)). Note that this metric is not a 3-Sasaki metric globally defined on \({{\mathcal {M}}}\). Here, \(\mathrm{E}({\mathbb {H}}^n)={\mathbb {H}}^n\rtimes \mathrm{Sp}(n)\cdot \mathrm{Sp}(1)\) is the quaternionic isometry group \(\mathop {\mathrm{Isom}}\nolimits ({\mathbb {H}}^n,g_{\mathbb {H}})\).

From (3.1) (cf. [2]), we see that any element \(h=\big ((t,v),\sqrt{u} A\cdot a\big )\in \mathop {\mathrm{Aut}}\nolimits _{qc}({{\mathcal {M}}})\) satisfies

$$\begin{aligned} h^*\omega _0=u\cdot a\omega _0 {{\bar{a}}} \end{aligned}$$
(3.6)

which thus preserves \({{\mathsf {D}}}_0\). Thus, for \(h\in {{\mathcal {M}}}\rtimes \mathrm{Sp}(n)\), it follows

$$\begin{aligned} h^*\omega _0=\omega _0 \text{ on } {{\mathcal {M}}}, \ \, h_*J_\alpha =J_\alpha h_* \text{ on } {{\mathsf {D}}}_0. \end{aligned}$$
(3.7)

In particular, note that \(t_*J_\alpha =J_\alpha t_*\) \(({}^\forall \, t\in {\mathbb {R}}^3=C({{\mathcal {M}}}))\).

3.3 Strict qc-group

Let \(({{\mathsf {D}}},\omega ,\{J_\alpha \}_{\alpha =1}^3\)) be a qc-structure on X. Put

$$\begin{aligned} \mathop {\mathrm{Psh}}\nolimits _{qc}(X)=\{h\in \mathop {\mathrm{Diff}}\nolimits (X)\mid h^*\omega =\omega ,\, h_*J_\alpha =J_\alpha h_*|_{{{\mathsf {D}}}},\, \alpha =1,2,3\}. \end{aligned}$$
(3.8)

It is a subgroup of \(\mathop {\mathrm{Aut}}\nolimits _{qc}(X)\). We apply the similar constriction for Sasaki groups (cf. [4]). Let \(\rho :{\mathbb {H}}^\ell \rightarrow \mathrm{Sp}(k)\) be a non-trivial homomorphism \((k+\ell =n)\). Define \({\mathbb {H}}(k,\ell )\) to be the semidirect product \(\displaystyle {\mathbb {H}}^k{\rtimes }_{\rho }^{}{\mathbb {H}}^\ell\) which is canonically embedded to the group of hyperkähler isometries \(\displaystyle {\mathbb {H}}^n\rtimes \mathrm{Sp}(n)\) of flat quaternionic space \({\mathbb {H}}^{n}\). Since \({\mathbb {H}}(k,\ell )\) acts simply transitively on \({\mathbb {H}}^n\), it is a flat hyperKähler group. (In fact, in view of [7, Theorem II], every flat hyperKähler Lie group contained in \({\mathbb {H}}^n\rtimes \mathrm{Sp}(n)\) may be conjugate to some \({\mathbb {H}}(k,\ell )\).) Let \(\mathop {\mathrm{Psh}}\nolimits _{qc}({{\mathcal {M}}})= {{\mathcal {M}}}\rtimes \mathrm{Sp}(n)\) be the normal subgroup of \(E({{\mathcal {M}}})\). Take the pull-back \({{\mathcal {M}}}(k,\ell )\) of \({\mathbb {H}}(k,\ell )\) in the following central extension :

(3.9)

Here, \({{\mathcal {M}}}(n,0)={{\mathcal {M}}}\). Then, \({{\mathcal {M}}}(k,\ell )\) is a simply connected solvable Lie group acting simply transitively by qc-transformations on the strict qc-manifold \({{\mathcal {M}}}\). Thus, \({{\mathcal {M}}}(k,\ell )\) admits a strict qc structure as a Lie group.

Theorem 3.3

Let G be a contractible unimodular strict qc Lie group. Then, G is isomorphic to \({{\mathcal {M}}}(k,\ell )\).

Proof

G is viewed as a strict qc-manifold endowed with a left invariant strict qc-structure \((\omega ,\{J_\alpha \}_{\alpha =1}^3)\). Then, \(G\le \mathop {\mathrm{Psh}}\nolimits _{qc}(G)\) by (3.8). If \({\mathbb {R}}^3\) is the abelian group generated by \({{\mathsf {E}}}=\langle \xi _\alpha , \alpha =1,2,3\rangle\), then \({\mathbb {R}}^3\le \mathop {\mathrm{Psh}}\nolimits _{qc}(G)\) by Proposition 2.2. Let \(\mathop {\mathrm{Isom}}\nolimits _{hk}(G/{\mathbb {R}}^3)=\{h\in \mathop {\mathrm{Diff}}\nolimits (G/{\mathbb {R}}^3)\mid h^*\Omega _\alpha =\Omega _\alpha , h_*{\hat{J}}_\alpha = {\hat{J}}_\alpha h_*, \alpha =1,2,3\}\) be a subgroup of isometries of the hyperKähler manifold \(G/{\mathbb {R}}^3\) as in Proposition 2.5. Denote by \(\mathop {\mathrm{Isom}}\nolimits _{h}(G/{\mathbb {R}}^3)=\{h\in \mathop {\mathrm{Diff}}\nolimits (G/{\mathbb {R}}^3)\mid h^*\Omega _1=\Omega _1, h_*{\hat{J}}_1={\hat{J}}_1 h_* \}\) the holomorphic isometry group of \(G/{\mathbb {R}}^3\) as a Kähler manifold. Recall that \(\mathop {\mathrm{Psh}}\nolimits (G/{\mathbb {R}}^2)=\{h\in \mathop {\mathrm{Diff}}\nolimits (G/{\mathbb {R}}^2)\mid h^*\omega _1=\omega _1, h_*J_1=J_1 h_* \}\) is the group of strictly pseudoconvex pseudo-Hermitian transformations of \((G/{\mathbb {R}}^2,(\omega _1,J_1))\). There is the commutative diagram of central group extensions (cf. [4, Proposition 3.4]):

(3.10)

It follows from [4, Theorem 2] that \(\mathop {\mathrm{Isom}}\nolimits _{h}(G/{\mathbb {R}}^3)^0=({\mathbb {C}}^k\rtimes \mathrm{U(k)})\times S_0\) for which \(S_0\) is a semisimple Lie group of noncompact type. It acts transitively on the Kähler manifold \(G/{\mathbb {R}}^3\) holomorphically isometric to the product \({\mathbb {C}}^k\times D\) where D is the bounded symmetric domain.

Consider (1)  \({\mathbb {R}}^3\) is normal in G. Then, \(G/{\mathbb {R}}^3\) is a flat hyperKähler group by Hano’s theorem [7] and so \(G/{\mathbb {R}}^3={\mathbb {H}}(k,\ell )\) as above. Then, the pull back of this in (3.10) gives \(G={{\mathcal {M}}}(k,\ell )\). Otherwise, (2)  \({\mathbb {R}}^k\) is normal in G where \(k=0,1,2\). Case (i)  If \({\mathbb {R}}^2\) is normal in G, then the quotient group \({\hat{G}} =G/{\mathbb {R}}^2\) is a Sasaki group for which \(\displaystyle S^1={\mathbb {R}}/{\mathbb {Z}}{\rightarrow }{\bar{G}}={\hat{G}}/{\mathbb {Z}}{\longrightarrow }G/{\mathbb {R}}^3\) is a pseudo-Hermitian (Sasaki) bundle. As in the proof of [4, Theorem 2], \(G/{\mathbb {R}}^3={\bar{G}}/S^1\) is a bounded symmetric domain so that \(G/{\mathbb {R}}^3={\mathbb {H}}^2_{\mathbb {R}}\) with \({\bar{G}}=\mathop {\mathrm{PSL}}\nolimits (2,{\mathbb {R}})\). It is impossible for \(G/{\mathbb {R}}^3\) to admit a quaternionic structure. Case (ii)  If \({\mathbb {R}}\) is normal in G, then put \({\hat{G}}=G/{\mathbb {R}}\). The principal bundle \(\displaystyle {\mathbb {R}}^2/{\mathbb {Z}}^2 {\rightarrow }{\bar{G}}={\hat{G}}/{\mathbb {Z}}^2{\longrightarrow }G/{\mathbb {R}}^3\) becomes a principal bundle of homogeneous space \(\displaystyle T^2{\rightarrow }{\bar{G}} {\longrightarrow }{\bar{G}}/T^2=G/{\mathbb {R}}^3\). As \(\bar{G}/T^2\) is a bounded symmetric domain, \({\bar{G}}\) is locally isomorphic to \(\mathop {\mathrm{PSL}}\nolimits (2,{\mathbb {R}})\times \mathop {\mathrm{PSL}}\nolimits (2,{\mathbb {R}})\), which is impossible since \(\bar{G}/T^2=\mathop {\mathrm{PSL}}\nolimits (2,{\mathbb {R}})/S^1\times \mathop {\mathrm{PSL}}\nolimits (2,{\mathbb {R}})/S^1 ={\mathbb {H}}^2_{\mathbb {R}}\times {\mathbb {H}}^2_{\mathbb {R}}\) has a positive scalar curvature which is not hyperKähler. Finally, Case (iii) \({\bar{G}}=G/{\mathbb {Z}}^3\) and \(T^3={\mathbb {R}}^3/{\mathbb {Z}}^3\) such that \(G/{\mathbb {R}}^3={\bar{G}}/T^3\) is a bounded symmetric domain. Thus, \({\bar{G}}\) is locally isomorphic to \(\mathop {\mathrm{PSL}}\nolimits (2,{\mathbb {R}})\times \mathop {\mathrm{PSL}}\nolimits (2,{\mathbb {R}})\times \mathop {\mathrm{PSL}}\nolimits (2,{\mathbb {R}})\) and \(G/{\mathbb {R}}^3={\mathbb {H}}^2_{\mathbb {R}}\times {\mathbb {H}}^2_{\mathbb {R}}\times {\mathbb {H}}^2_{\mathbb {R}}\), which cannot be a quaternionic manifold. \(\square\)

4 Spherical qc-manifolds

The following theorem is a supporting example to Proposition 2.5 which is implied by Schoen’s result [14]. (Compare [6, 12] for the proofs of the quaternionic case.)

Theorem 4.1

Let \((X,{{\mathsf {D}}},\{J_\alpha \}_{\alpha =1}^3)\) be a noncompact quaternionic contact manifold. If \(\mathop {\mathrm{Aut}}\nolimits _{qc}(X)\) does not act properly on X, then X admits the spherical qc-structure qc-conformal to the quaternionic Heisenberg Lie group \({{\mathcal {M}}}\).

Let \((\tilde{\omega },{{\mathsf {D}}}_0, \{I_\alpha \}_{\alpha =1}^3)\) be the standard qCR-structure on \(S^{4n+3}\) such that \(\ker \,\tilde{\omega }={{\mathsf {D}}}_0\) (cf. [2]).

Definition 4.2

A qc-manifold \((M,{{\mathsf {D}}}, (\omega ,\{J_\alpha \}_{\alpha =1}^3))\) is spherical (or uniformizable) with respect to \((\mathop {\mathrm{PSp}}\nolimits (n+1,1), S^{4n+3})\) if there exists a \(\rho\)-equivariant developing immersion from the universal covering manifold X of M:

$$\begin{aligned} (\rho ,\mathop {\mathop {\mathrm{dev}}\nolimits }):(\mathop {\mathrm{Aut}}\nolimits _{qc}(X),X){\rightarrow }(\mathop {\mathrm{PSp}}\nolimits (n+1,1), S^{4n+3}) \end{aligned}$$

such that

  1. (i)

      \(\displaystyle \mathop {\mathrm{dev}}\nolimits ^*\tilde{\omega }=\lambda \cdot \omega \cdot {\bar{\lambda }}=u\cdot a\omega {\bar{a}}\) for some map \(\lambda =\sqrt{u}\cdot a:X{\rightarrow }{\mathbb {R}}^+\times {\mathop {\mathrm{Sp}}\nolimits }(1)\) \((u\in {\mathbb {R}}^+, a\in {\mathop {\mathrm{Sp}}\nolimits }(1))\).

  2. (ii)

      If the conjugate by the map \(a:X{\rightarrow }{\mathop {\mathrm{Sp}}\nolimits }(1)\) represents the matrix \((a_{\alpha \beta }): X{\rightarrow }\mathrm{SO}(3)\), then \(\mathop {\mathrm{dev}}\nolimits _*\circ J_\alpha =\sum _{\beta }a_{\alpha \beta }I_\beta \circ \mathop {\mathrm{dev}}\nolimits _*: {{\mathsf {D}}}{\rightarrow }{{\mathsf {D}}}_0\).

  3. (iii)

      \(\rho :\mathop {\mathop {\mathrm{Aut}}\nolimits }_{qc}(X){\rightarrow }\mathop {\mathrm{PSp}}\nolimits (n+1,1)\) is the holonomy homomorphism such that \(\displaystyle \mathop {\mathop {\mathrm{dev}}\nolimits }\circ h=\rho (h)\circ \mathop {\mathop {\mathrm{dev}}\nolimits }\)   \(({}^\forall \, h\in {\mathop {\mathrm{Aut}}\nolimits }_{qc}(X))\).

Definition 4.3

A qc-manifold X is divisible if there exists a discrete subgroup \(\Gamma \le \mathop {\mathrm{Aut}}\nolimits _{qc}(X)\) which acts properly discontinuously with compact quotient.

Theorem 4.4

Let \((X,{{\mathsf {D}}},\omega ,\{J_\alpha \}_{\alpha =1}^3)\) be a simply connected noncompact uniformizable strict qc-manifold. Put \({{\mathsf {E}}}=\{\xi _\alpha ,\alpha =1,2,3\}\). Suppose X is divisible by \(\Gamma\).

(1) If \(\mathop {\mathrm{Aut}}\nolimits _{qc}(X)\) leaves \({{\mathsf {E}}}\) invariant, then the developing pair reduces to the equivariant immersion:

$$\begin{aligned} (\rho ,\mathop {\mathop {\mathrm{dev}}\nolimits }):(\mathop {\mathrm{Aut}}\nolimits _{qc}(X),X){\rightarrow }(\mathop {\mathrm{Aut}}\nolimits _{qc}({{\mathcal {M}}}),{{\mathcal {M}}}). \end{aligned}$$

(2) For any \(\gamma \in \Gamma\) and \(\alpha =1,2,3\), suppose \(\gamma _*\xi _\alpha =\sum _{\beta =1}^3a_{\alpha \beta }\xi _\beta\) for some function \(a_{\alpha \beta }:X{\rightarrow }\mathrm{SO}(3)\). Then,

  1. (i)

    \(\mathop {\mathrm{dev}}\nolimits : X{\rightarrow }{{\mathcal {M}}}\) is a qc-diffeomorphism so that \({{\mathcal {R}}}={\mathbb {R}}^3\).

  2. (ii)

    There exists a strict qc-structure \(\displaystyle (\eta ,\{J_\alpha '\}_{\alpha =1}^3,{\mathbb {R}}^3)\) qc-conformal to \((\omega ,\{J_\alpha \}_{\alpha =1}^3)\). The quotient \((X/{\mathbb {R}}^3, \{\Theta _\alpha ,{\hat{J}}_\alpha '\}_{\alpha =1}^3)\) is a hyperKähler manifold isometric to \({\mathbb {H}}^n\).

The method of proof is based on that of [13] by taking into account the results of [14] (cf. [6, 12]).

Proof

Put \(\mathop {\mathrm{Aut}}\nolimits _{qc}(X)=\mathop {\mathrm{Aut}}\nolimits (X)\). Let \(\displaystyle G={\overline{\rho (\mathop {\mathrm{Aut}}\nolimits (X))}}^0\) be the identity component of the closure of the holonomy image \(\displaystyle \rho (\mathop {\mathrm{Aut}}\nolimits (X))\) in \(\mathop {\mathrm{PSp}}\nolimits (n+1,1)\). We first show that (i) G is not compact.

Case 1. Suppose G is compact. If G has no fixed point on \(S^{4n+3}\), then G has the unique fixed point at the origin \(\varvec{0}\) in \({\mathbb {H}}^{n+1}_{\mathbb {H}}\) where \(S^{4n+3}=\partial {\mathbb {H}}^{n+1}_{\mathbb {H}}\). As in the proof of [13], \(\mathop {\mathrm{dev}}\nolimits :X{\rightarrow }S^{4n+3}\) is shown to be an isometry, which is excluded by the non-compactness of X. So G has the fixed point set F in \(S^{4n+3}\). We may assume that \(\mathop {\mathrm{Aut}}\nolimits (X)\) acts properly on X by Theorem 4.1, so \(\mathop {\mathrm{dev}}\nolimits\) misses F. It reduces to an immersion \(\mathop {\mathrm{dev}}\nolimits :X{\rightarrow }S^{4n+3}-F\). As \(\mathop {\mathrm{Aut}}\nolimits (S^{4n+3}-F)\) acts properly on \(S^{4n+3}-F\) by the result of [14], there is a Riemannian metric on \(S^{4n+3}-F\) invariant under \(\mathop {\mathrm{Aut}}\nolimits (S^{4n+3}-F)\). Since X is divisible, X is complete with respect to the pullback metric, \(\mathop {\mathrm{dev}}\nolimits : X{\rightarrow }S^{4n+3}-F\) is a covering map. On the other hand, if we note that the action of G is linear on \(S^{4n+3}\), F must be a subsphere \(S^{k}\) \((0\le k<4n+3)\) such that the complement \(S^{4n+3}-F\) is unknotted, that is, homeomorphic to \({\mathbb {R}}^{k+1}\times S^{4n+2-k}\). Moreover, it is shown in [13, Lemma 3.1] (also [5, p.77]) that \(S^{4n+3}-F\) is either one of the following:

  1. (1)

    \(S^{4n+3}-S^{m-1}\) where \(F=S^{m-1}=\partial {\mathbb {H}}^{m}_{{\mathbb {R}}}\) \((1\le m\le n+1)\). \({\mathop {\mathrm{Aut}}\nolimits }(S^{4n+3}-S^{m-1})=P(\mathrm{O}(m,1)\cdot \mathrm{Sp}(1)\times \mathrm{Sp}(n-m+1))\).

  2. (2)

    \(S^{4n+3}-S^{2m-1}\) where \(F=S^{2m-1}=\partial {\mathbb {H}}^{m}_{{\mathbb {C}}}\) \((1\le m\le n+1)\). \({\mathop {\mathrm{Aut}}\nolimits }(S^{4n+3}-S^{2m-1})= P(\mathrm{U}(m,1)\cdot \mathrm{U}(1)\times \mathrm{Sp}(n-m+1))\).

  3. (3)

    \(S^{4n+3}-S^{4m-1}\) where \(F=S^{4m-1}=\partial {\mathbb {H}}^{m}_{{\mathbb {H}}}\) \((1\le m\le n)\). \({\mathop {\mathrm{Aut}}\nolimits }(S^{4n+3}-S^{4m-1})=\mathrm{Sp}(m,1)\cdot \mathrm{Sp}(n-m+1)\).

  4. (4)

    \(S^{4n+3}-S^{2}\) where \(S^2=\partial {\mathbb {H}}_{\mathrm{Im}}\). \({\mathop {\mathrm{Aut}}\nolimits }(S^{4n+3}-S^{2})=\mathrm{SL}(2,{\mathbb {C}})\cdot \mathrm{Sp}(n)\). (This case reduces to (3).)

In particular, \(S^{4n+3}-F\) is simply connected in each case. Hence, \(\mathop {\mathrm{dev}}\nolimits : X{\rightarrow }S^{4n+3}-F\) is diffeomorphic, so \(\rho\) is an isomorphism. However, this case does not occur since \(\mathop {\mathrm{Aut}}\nolimits (X)^0\cong G=\rho (\mathop {\mathrm{Aut}}\nolimits (X))=\mathop {\mathrm{Aut}}\nolimits (S^{4n+3}-F)^0\) which is noncompact (in fact \(\mathop {\mathrm{Aut}}\nolimits (S^{4n+3}-F)^0\) contains a noncompact subgroup \(\mathrm{O}(m,1),\,\mathrm{U}(m,1)\) \((1\le m\le n+1)\), \(\mathop {\mathrm{SL}}\nolimits (2,{\mathbb {C}})\) or \(\mathop {\mathrm{Sp}}\nolimits (m,1)\) \((1\le m\le n)\), respectively.) As a consequence, the case G is compact does not occur.

Case 2. Suppose G is noncompact. Then, either G has a common fixed point \(\{\infty \}\) in \(S^{4n+3}\) or G leaves invariant a totally geodesic subspace of \({\mathbb {H}}^{n+1}_{\mathbb {H}}\). (See [5, Theorem 4.4.1].) In the latter case, G (possibly \(\mathop {\mathrm{PSp}}\nolimits (n+1,1)\)) is either one of the identity component of the above groups (1), (2), (3), (4). On the other hand, let \({{\mathcal {R}}}\) be the local abelian group generated by \({{\mathsf {E}}}\) as before. The holonomy homomorphism \(\rho\) maps \({{\mathcal {R}}}\) into \(\mathop {\mathrm{PSp}}\nolimits (n+1,1)\). By Liouville’s theorem, \(\rho ({{\mathcal {R}}})\) extends globally to a subgroup of \(\mathop {\mathrm{PSp}}\nolimits (n+1,1)\) on \(S^{4n+3}\). As \(\mathop {\mathrm{Aut}}\nolimits _{qc}(X)\) leaves \({{\mathsf {E}}}\) invariant, \(\mathop {\mathrm{Aut}}\nolimits (X)^0\) normalizes \({{\mathcal {R}}}\). Thus, G normalizes \(\overline{\rho ({{\mathcal {R}}})}\) also. In particular, the radical of G is nontrivial, so G is not semisimple. On the other hand, G is semisimple except for the case \(m=1\) of (1) such that \({\mathop {\mathrm{Aut}}\nolimits }(S^{4n+3}-S^{0})^0=\mathrm{SO}(1,1)^0\times \mathrm{Sp}(1)\cdot \mathrm{Sp}(n) \cong {\mathbb {R}}^+\times \mathrm{Sp}(n)\cdot \mathrm{Sp}(1)\). For this case, G has exactly two fixed points \(\{\varvec{0},\infty \}\). Noting from Theorem 4.1, as in the argument of (i), \(\displaystyle \mathop {\mathrm{dev}}\nolimits : X{\rightarrow }S^{4n+3}-\{\varvec{0},\infty \}= S^{4n+2}\times {\mathbb {R}}^+\) is a diffeomorphism so that \(\rho\) maps the radical of \(\mathop {\mathrm{Aut}}\nolimits (X)^0\) isomorphically onto the radical \({\mathbb {R}}^+\). Since the radical contains \(\rho ({{\mathcal {R}}})\) of dimension three, it is impossible. As a consequence, by the non-ellipticity of elements in \(\mathop {\mathrm{PSp}}\nolimits (n+1,1)\), G has a unique common fixed point \(\{\infty \}\). Noting \(\mathop {\mathrm{Aut}}\nolimits (X)\) acts properly on X, \(\mathop {\mathrm{dev}}\nolimits\) misses \(\{\infty \}\). This proves (1).

(2). Suppose some \(\rho (\gamma )\in \rho (\Gamma )\) has a nontrivial summand in \({\mathbb {R}}^+\) of \(\mathop {\mathrm{Aut}}\nolimits ({{\mathcal {M}}})\). It follows from (3.6) that \(\displaystyle \rho (\gamma )^*\omega _0= v\cdot b\omega _0\bar{b}\) where \(\lambda =\sqrt{v}\cdot b\in \mathop {\mathrm{Sp}}\nolimits (1)\times {\mathbb {R}}^+\). On the other hand, by our hypothesis (2), \(\gamma _*\xi _\alpha =\sum _{\beta =1}^3a_{\alpha \beta }\xi _\beta\) for some function \(a_{\alpha \beta }:X{\rightarrow }\mathrm{SO}(3)\). Put \(\mathop {\mathrm{dev}}\nolimits _*\xi _\alpha =\tilde{\xi }_\alpha\) on \({{\mathcal {M}}}\). As \(\mathop {\mathrm{dev}}\nolimits \circ \gamma =\rho (\gamma )\circ \mathop {\mathrm{dev}}\nolimits\), letting \((i,j,k)\sim (i_\alpha ,i_\beta ,i_\gamma )\), it follows

$$\begin{aligned}\rho (\gamma )_*(\tilde{\xi }_\alpha i_\alpha )=a(\tilde{\xi }_\alpha i_\alpha ){\bar{a}} \end{aligned}$$

where the conjugate of a represents the matrix \((a_{\alpha \beta })\in \mathrm{SO}(3)\). Calculate

$$\begin{aligned} \begin{aligned} \omega _0(\rho (\gamma )_*\tilde{\xi }_\alpha i_\alpha )&=a\omega _0(\tilde{\xi }_\alpha i_\alpha ){\bar{a}}\\ \rho (\gamma )^*\omega _0(\tilde{\xi }_\alpha i_\alpha )&=u\cdot b\omega _0(\tilde{\xi }_\alpha i_\alpha ){\bar{b}}. \end{aligned}\end{aligned}$$
(4.1)

Taking the norm in \({\mathbb {H}}\), it follows

$$\begin{aligned} \big |\omega _0(\tilde{\xi }_\alpha i_\alpha )\big |= \big |a\omega _0(\tilde{\xi }_\alpha i_\alpha ){\bar{a}}\big | =\big |u\cdot b\omega _0(\tilde{\xi }_\alpha i_\alpha )\bar{b}\big |=u\big |\omega _0(\tilde{\xi }_\alpha i_\alpha )\big |. \end{aligned}$$

Hence, \(u=1\) on X. This implies \(\rho (\gamma )\in \mathrm{E}({{\mathcal {M}}})\) so that \(\rho (\Gamma )\le \mathrm{E}({{\mathcal {M}}})\). As usual, there is the \(\mathrm{E}({{\mathcal {M}}})\)-invariant Riemannian metric on \({{\mathcal {M}}}\). Since X is divisible, X is complete with respect to the pullback metric. Thus, \((\rho ,\mathop {\mathrm{dev}}\nolimits ):(\Gamma ,X){\rightarrow }(\mathrm{E}({{\mathcal {M}}}),{{\mathcal {M}}})\) is an equivariant isometry. As \(\rho :\mathop {\mathrm{Aut}}\nolimits _{qc}(X){\rightarrow }\mathop {\mathrm{Aut}}\nolimits _{qc}({{\mathcal {M}}})\) is an isomorphism, and \({{\mathcal {R}}}\) is normalized by \(\mathop {\mathrm{Aut}}\nolimits _{qc}(X)\), so does \(\rho ({{\mathcal {R}}})\) in \(\mathop {\mathrm{Aut}}\nolimits _{qc}({{\mathcal {M}}})\). By the action of (3.1) and the group structure of \(\mathop {\mathrm{Aut}}\nolimits _{qc}({{\mathcal {M}}})\) we note \(\rho ({{\mathcal {R}}})={\mathbb {R}}^3\) which is the center of \({{\mathcal {M}}}\). This proves (i). In particular, \(\displaystyle \mathop {\mathrm{dev}}\nolimits _*{{\mathsf {E}}}={{\mathsf {E}}}_0\).

(ii). Let \((\omega _0,\{J_\alpha \}_{\alpha =1}^3)\) be the standard (spherical) qc-structure on \({{\mathcal {M}}}\) where \(\omega _0=\omega _1i+\omega _2j+\omega _3k\). By the definition, it satisfies \(\displaystyle \mathop {\mathrm{dev}}\nolimits ^*\omega _0=u\cdot a\omega {\bar{a}}\) for some \(u\in {\mathbb {R}}^+, a\in {\mathop {\mathrm{Sp}}\nolimits }(1)\). When a represents \((a_{\alpha \beta })\in \mathrm{SO}(3)\) as before, it follows \(\mathop {\mathrm{dev}}\nolimits _*J_\alpha =\sum _{\beta }a_{\alpha \beta }J_\beta \mathop {\mathrm{dev}}\nolimits _*\) (cf. Definition 4.2). Put

$$\begin{aligned} \begin{aligned} \eta&=\mathop {\mathrm{dev}}\nolimits ^*\omega _0 \, (\eta _\alpha =\mathop {\mathrm{dev}}\nolimits ^*\omega _\alpha ), \ J_\alpha '=\mathop {\sum }_{\beta =1}^3a_{\beta \alpha }J_\beta \,(\alpha =1,2,3). \end{aligned} \end{aligned}$$
(4.2)

We check \(({{\mathsf {D}}}, \eta ,\{J_\alpha '\}_{\alpha =1}^3)\) is a qc-structure qc-conformal to \(({{\mathsf {D}}}_0,\omega , \{J_\alpha \}_{\alpha =1}^3)\). For this, let \(X,Y\in {{\mathsf {D}}}\) so that \(\mathop {\mathrm{dev}}\nolimits _*X,\mathop {\mathrm{dev}}\nolimits _*Y\in {{\mathsf {D}}}_0\). Note that \(J_\gamma \mathop {\mathrm{dev}}\nolimits _*=\mathop {\mathrm{dev}}\nolimits _*\sum _{\mu }a_{\mu \gamma }J_\mu\) as above. Let \(\displaystyle {{\tilde{J}}}_\gamma =(d\eta _\beta )^{-1}\circ d\eta _\alpha\) as well as \(\displaystyle J_\alpha =({\text {d}}\omega _\beta )^{-1}\circ {\text {d}}\omega _\alpha\). By calculation,

$$\begin{aligned} \begin{aligned} {\text {d}}\omega _\beta (J_\gamma (\mathop {\mathrm{dev}}\nolimits _*\varvec{u}), \mathop {\mathrm{dev}}\nolimits _*\varvec{v})&={\text {d}}\omega _\alpha (\mathop {\mathrm{dev}}\nolimits _*\varvec{u}, \mathop {\mathrm{dev}}\nolimits _*\varvec{v})= d\eta _\alpha (\varvec{u},\varvec{v})\\ ={\text {d}}\omega _\beta \bigl (\mathop {\mathrm{dev}}\nolimits _*({\sum }_{\mu }a_{\mu \gamma }J_\mu \varvec{u}&), \mathop {\mathrm{dev}}\nolimits _*\varvec{v}\bigr ) =d\eta _\beta (J'_\gamma \varvec{u}, \varvec{v}). \end{aligned} \end{aligned}$$

Noting \(d\eta _\beta ({{\tilde{J}}}_\gamma \varvec{u},\varvec{v})=d\eta _\alpha (\varvec{u},\varvec{v})\), the non-degeneracy of \(d\eta _\beta\) implies \({{\tilde{J}}}_\gamma =J'_\gamma\). The equations \(\mathop {\mathrm{dev}}\nolimits ^*\omega _0=\eta\), \(\mathop {\mathrm{dev}}\nolimits _*{{\mathsf {E}}}={{\mathsf {E}}}_0\) imply \(d\eta ({{\mathsf {E}}},A)=0\) \(({}^\forall \, A\in TX)\). Thus, \(({{\mathsf {D}}},\eta ,\{J_\alpha '\}_{\alpha =1}^3)\) is a strict qc-structure. Noting that \({{\mathsf {R}}}\) acts properly and freely on X, we have a smooth manifold \(Y=X/{{\mathsf {R}}}\). As \(\mathop {\mathrm{dev}}\nolimits :X{\rightarrow }{{\mathcal {M}}}\) is \(\rho\)-equivariant, \(\mathop {\mathrm{dev}}\nolimits\) induces a diffeomorphism \(\hat{\mathop {\mathrm{dev}}\nolimits }: Y{\rightarrow }{\mathbb {H}}^n\) with the commutative diagram:

Applying Proposition 2.5 (cf. (2.6)), the 2-form \(\Theta _\alpha\) on Y is defined by

$$\begin{aligned} \pi ^*\Theta _\alpha =d\eta _\alpha \ \, (\alpha =1,2,3). \end{aligned}$$
(4.3)

A quaternionic structure \(\{{\hat{J}}'_\alpha \}_{\alpha =1}^3\) on Y is also induced by

$$\begin{aligned} {\hat{J}}'_\alpha \pi _*=\pi _*J'_\alpha . \end{aligned}$$
(4.4)

Using the reciprocity \(\displaystyle J'_\gamma =(d\eta _\beta )^{-1}\circ d\eta _\alpha\), it follows for any \({\hat{A}},{\hat{B}}\in TY\)

$$\begin{aligned} \Theta _\alpha ({\hat{J}}'_\alpha {\hat{A}}, {\hat{J}}'_\alpha {\hat{B}})=\Theta _\alpha ({\hat{A}},{\hat{B}})\ \, (\alpha =1,2,3). \end{aligned}$$

Thus, \(\Theta _\alpha\) is a Kähler form on \((Y,\{J_\alpha '\}_{\alpha =1}^3)\). The quaternionic Hermitian metric

$$\begin{aligned} g({\hat{A}},{\hat{B}})=\Theta _\alpha ({\hat{J}}'_\alpha {\hat{A}},{\hat{B}})= \Theta _\beta ({\hat{J}}'_\beta {\hat{A}},{\hat{B}})= \Theta _\gamma ({\hat{J}}'_\gamma {\hat{A}},{\hat{B}}) \end{aligned}$$
(4.5)

is a hyperKähler metric on \((Y,\{{\hat{J}}_\alpha '\}_{\alpha =1}^3)\). Let \(({\mathbb {H}}^n, g_{\mathbb {H}}, \{I_\alpha \}_{\alpha =1}^3)\) be the standard euclidean metric as in Remark 3.2. Noting \({\text {d}}\omega _\alpha \circ J_\alpha =\pi ^*g_{\mathbb {H}}\) and (4.2), a calculation shows

$$\begin{aligned} \begin{aligned} \hat{\mathop {\mathrm{dev}}\nolimits ^*}g_0=g, \hat{\mathop {\mathrm{dev}}\nolimits }_*\circ {\hat{J}}_\alpha '=I_\alpha \circ {\hat{\mathop {\mathrm{dev}}\nolimits }}_*. \end{aligned} \end{aligned}$$
(4.6)

This gives an isometry of \((Y, g, \{{\hat{J}}_\alpha '\}_{\alpha =1}^3)\) onto \(({\mathbb {H}}^n, g_{\mathbb {H}}, \{I_\alpha \}_{\alpha =1}^3)\). \(\square\)

Remark 4.5

The new Kähler form \(\Theta _\alpha\) and \(\Theta = \Theta _1 i+\Theta _2 j+\Theta _3 k\) are related to the original forms \(\Omega\) and \(\Omega _\alpha\) as follows. For some constant \(c>0\),

$$\begin{aligned} \Theta =c\cdot a\Omega {\bar{a}},\, \Theta _\alpha =c\cdot \sum _{\beta =1}^3a_{\alpha \beta }\Omega _\beta . \end{aligned}$$

In fact, as we put \(\displaystyle \eta =\mathop {\mathrm{dev}}\nolimits ^*\omega _0=u\cdot a\omega \bar{a}\), it follows \(\displaystyle {\text {d}}\eta =u\cdot a\cdot {\text {d}}\omega \cdot \bar{a}\,|_{{\mathsf {D}}}\) so that \(\displaystyle \Theta ={\hat{u}}\cdot \alpha \Omega \bar{\alpha }\) where \({\hat{u}},\alpha\) are induced functions on Y. As usual, \(\displaystyle \Theta ^2= {\hat{u}}^2 \Omega ^2\) which shows that \({\hat{u}}\) is a constant \(c>0\). We have \(\Theta _\alpha =c \sum _{\beta =1}^3\alpha _{\alpha \beta }\Omega _\beta\). \((\pi ^*\alpha =a,\ \pi ^*\alpha _{\alpha \beta }=a_{\alpha \beta })\).

5 Quotient quaternionic Hermitian manifolds

For the strict qc-structure \(({{\mathsf {D}}}_0, \omega _0, \{J_\alpha \}_{\alpha =1}^3)\) on the quaternionic Heisenberg Lie group \({{\mathcal {M}}}\), we consider a qc-structure \(\eta =\eta _1i+\eta _2 j+\eta _3 k\) which is qc-conformal to \(\omega\). Take a one-form, say \(\eta _1\) to define a distribution:

$$\begin{aligned} {{\mathsf {E}}}_1=\{\xi \mid d\eta _1(\xi ,A)=0,\ \, {}^\forall \, A\in TX\}. \end{aligned}$$
(5.1)

\({{\mathsf {E}}}_1\) does not induce a distribution such as \({{\mathsf {E}}}\). When \({{\mathsf {E}}}_1\) generates a three-dimensional abelian Lie group \({{\mathcal {R}}}\), we shall show that there is an invariant domain X such that the quotient \(X/{{\mathcal {R}}}\) admits a special kind of quaternionic Hermitian structure.

Choose numbers \(a_1,\ldots ,a_n\) such that

$$\begin{aligned} 0<a_1<a_2<\cdots <a_n. \end{aligned}$$
(5.2)

Let \(A_t\) be the diagonal matrix

$$\begin{aligned} \mathrm{diag}(\,e^{ia_1t},\,e^{ia_2t},\,\ldots \, e^{ia_nt}\,) \in T^n\le \mathop {\mathrm{Sp}}\nolimits (n). \end{aligned}$$
(5.3)

Define a homomorphism \(\displaystyle \rho _1:{\mathbb {R}}^3{\rightarrow }({\mathbb {R}}^3,0)\times \mathrm{Sp}(n)\le {{\mathcal {M}}}\rtimes \mathop {\mathrm{Sp}}\nolimits (n)\) to be

$$\begin{aligned} \begin{aligned} \rho _1(t_1)&=\bigl (((t_1,0,0),0),\, A_{t_1}\bigr ),\\ \rho _1(t_2)&=\bigl (((0,t_2,0),0),\, I\bigr ),\ \ \rho _1(t_3)=\bigl (((0,0,t_3),0),\, I\bigr ). \\ \end{aligned} \end{aligned}$$
(5.4)

More precisely, this action is defined on \({{\mathcal {M}}}={\mathbb {R}}^3\times {\mathbb {H}}^n\) as

$$\begin{aligned} \begin{aligned} \rho _1(t_1)\bigl ((s_1,s_2,s_3), z_1,\dots ,z_n\bigr )&=\bigl (((s_1+t_1), s_2,s_3 ),\\&\qquad e^{ia_1t_1}z_1,\dots , e^{ia_nt_1}z_n \bigr ),\\ \rho _1(t_2)\bigl ((s_1,s_2,s_3), z_1,\dots ,z_n\bigr )&=\bigl ((s_1,(s_2+t_2),s_3), z_1,\dots ,z_n\bigr ),\\ \rho _1(t_3)\bigl ((s_1,s_2,s_3),z_1,\dots ,z_n\bigr )&=\bigl ((s_1,s_2,(s_3+t_3)),z_1,\dots ,z_n\bigr ). \end{aligned} \end{aligned}$$
(5.5)

In view of (5.4), the group \(\langle \rho _1(t_1),\rho _1(t_2),\rho _1(t_3)\rangle\) forms a 3-dimensional abelian Lie group \(\rho _1({\mathbb {R}}^3)\). If \(\xi _1\) is the vector field induced by \(\{\rho _1(t_1)\}_{t_1\in {\mathbb {R}}}\), then it follows

$$\begin{aligned} \begin{aligned} \xi _1&=\frac{\rho _1(t_1)}{{\text {d}}t_1}|_{t_1=0}=\frac{\text {d}}{{\text {d}}t_1}+ (a_1iz_1,\dots , a_n i z_n) =\ \frac{\text {d}}{{\text {d}}t_1}+\\&\sum _{k=1}^n a_k(-x_{4k-2}\frac{{\text {d}}}{{\text {d}}x_{4k-3}}+x_{4k-3}\frac{{\text {d}}}{{\text {d}}x_{4k-2}} - x_{4k} \frac{{\text {d}}}{{\text {d}}x_{4k-1}}+x_{4k-1}\frac{{\text {d}}}{{\text {d}}x_{4k}}).\\ \end{aligned}\end{aligned}$$
(5.6)

Letting \(\displaystyle z_k=u_k+v_k j= (x_{4k-3}+ix_{4k-2})+(x_{4k-1}+ix_{4k})j\in {\mathbb {H}}\)  \((k=1,\cdots ,n)\), a calculation using (3.3) shows that

$$\begin{aligned} \begin{aligned} \omega _1((\xi _1)_p)&=1+\sum _{k=1}^n a_k(x_{4k-3}^2+x_{4k-2}^2-x_{4k}^2-x_{4k-1}^2)\\&=1+\sum _{k=1}^n a_k(|u_k|^2-|v_k|^2)\ \, \text{ at }\ p=((s_1,s_2,s_3), z_1,\dots ,z_n). \end{aligned}\end{aligned}$$
(5.7)

The singular set \(S=\{p\in {{\mathcal {M}}}\mid \omega _1((\xi _1)_p)=0\}\) is not entirely equal to \({{\mathcal {M}}}\) from (5.7). Denote the domain of \({{\mathcal {M}}}\) by

$$\begin{aligned} X=\{p\in {{\mathcal {M}}}\mid \omega _1((\xi _1)_p)\ne 0\} \end{aligned}$$
(5.8)

(if necessary taking the component containing the origin \((0,0)\in {{\mathcal {M}}}\)). Since \(\rho _1({\mathbb {R}}^3)\) leaves S invariant, so does X. Put \(\displaystyle Y=X/\rho _1({\mathbb {R}}^3)\). Then, there is a commutative diagram of principal bundles.

(5.9)

The image \(\pi _1S=\{\pi _1(p)\in {\mathbb {H}}^n\mid \sum _{k=1}^n a_k(|u_k|^2-|v_k|^2)=-1\}\) is a real hypersurface in \({\mathbb {R}}^{4n}={\mathbb {H}}^n\). Noting \(Y={\mathbb {H}}^n-\pi _1S\), it follows each component of Y is simply connected in \({\mathbb {H}}^n\).

5.1 Conformal change of \(\omega _0\)

Let \(\omega _0=\omega _1i+\omega _2j+\omega _3k\) be the qc-form on \({{\mathcal {M}}}\) (cf. Sect. 3.2). We introduce new 1-forms on X:

$$\begin{aligned} \eta _\alpha =\frac{1}{\omega _1(\xi _1)}\omega _\alpha \, (\alpha =1,2,3). \end{aligned}$$
(5.10)

Put \(\displaystyle \eta =\eta _1i+\eta _2j+\eta _3k\) on X. Since \(\eta\) is conformal to \(\omega _0\), it follows \(\ker \, \eta =\ker \, \omega _0={{\mathsf {D}}}_0\) on X. As \(\rho _1({\mathbb {R}}^3)\) leaves \(\omega _\alpha\) invariant, so does \(\eta _\alpha\). The hypercomplex structure of \(\eta\), \(\displaystyle \{{{\tilde{J}}}_\gamma =(d\eta _\beta |\mathsf{D_0})^{-1}\circ (d\eta _{\alpha }|\mathsf{D_0})\}_{\gamma =1}^3\) coincides with \(\{J_\gamma \}_{\gamma =1}^3\) of \(\omega _0\) on \({{\mathsf {D}}}_0\). Noting \(\displaystyle \rho _1(t)_*J_\alpha =J_\alpha \rho _1(t)_*\) \((t\in {\mathbb {R}}^3)\) from (3.7), \(\{J_\alpha \}\) induces a quaternionic structure \(\{{\hat{J}}_\alpha \}_{\alpha =1}^3\) on Y such that

$$\begin{aligned} \pi _{1*} J_\alpha (\varvec{v})={\hat{J}}_\alpha \pi _{1*} (\varvec{v})\ \, ({}^\forall \, \varvec{v}\in {{\mathsf {D}}}_0). \end{aligned}$$
(5.11)

Proposition 5.1

\((Y,\{{\hat{J}}_\alpha \}_{\alpha =1}^3)\) is a quaternionic Hermitian manifold.

Proof

Define \(\hat{\Omega }_\alpha\) \((\alpha =1,2,3)\) to be

$$\begin{aligned} \hat{\Omega }_\alpha (\pi _{1*}\varvec{u},\pi _{1*}\varvec{v})=d\eta _\alpha (\varvec{u},\varvec{v}) ({}^\forall \, \varvec{u}, \varvec{v}\in {{\mathsf {D}}}_0). \end{aligned}$$
(5.12)

Since \(\eta _\alpha\) is \(\rho _1({\mathbb {R}}^3)\)-invariant and the distribution by \(\rho _1({\mathbb {R}}^3)\) is transverse to \({{\mathsf {D}}}_0\), \(\hat{\Omega }_\alpha\) is well-defined on Y (cf. Lemma 5.2). Put

$$\begin{aligned} {\hat{g}}(\hat{\varvec{u}},\hat{\varvec{v}})=\hat{\Omega }_\alpha ({\hat{J}}_\alpha \hat{\varvec{u}}, \hat{\varvec{v}}) \, ({}^\forall \, \hat{\varvec{u}}, \hat{\varvec{v}}\in TY). \end{aligned}$$
(5.13)

As \(\hat{\Omega }_\alpha\) is invariant under \({\hat{J}}_\alpha\), it follows \(\displaystyle {\hat{g}}({\hat{J}}_\alpha \hat{\varvec{u} }, {\hat{J}}_\alpha \hat{\varvec{v}})= {\hat{g}}(\hat{\varvec{u}}, \hat{\varvec{v}})\). Thus, \({\hat{g}}\) is a quaternionic Hermitian metric on \((Y,\{{\hat{J}}_\alpha \}_{\alpha =1}^3)\).

As in (5.4), the distribution \(\displaystyle \left\{ \xi _1,\frac{\text {d}}{{\text {d}}t_2},\frac{\text {d}}{{\text {d}}t_3}\right\}\) generates \(\rho _1({\mathbb {R}}^3)\le {{\mathcal {M}}}\rtimes \mathop {\mathrm{Sp}}\nolimits (n)\). Note from (3.3) that

$$\begin{aligned} \eta _1(\xi _1)=1, \eta _1(\frac{\text {d}}{{\text {d}}t_2})=0, \, \eta _1(\frac{\text {d}}{{\text {d}}t_3})=0. \end{aligned}$$
(5.14)

Lemma 5.2

\(\displaystyle {{\mathsf {E}}}_1=\langle \xi _1, \frac{\text {d}}{{\text {d}}t_2},\frac{\text {d}}{{\text {d}}t_3}\rangle\).

Proof

For any \(A\in TX\), we prove \(\displaystyle {\text {d}}\eta _1(\xi _1, A)=0,\, \ {\text {d}}\eta _1(\frac{\text {d}}{{\text {d}}t_\beta }, A)=0\) \((\beta =2, 3)\). If \(A\in {{\mathsf {D}}}_0\), then \([\xi _1,A]\in {{\mathsf {D}}}_0\). Since the distribution \(\displaystyle \langle \xi _1, \frac{\text {d}}{{\text {d}}t_2},\frac{\text {d}}{{\text {d}}t_3}\rangle\) generates \({\mathbb {R}}^3\), \(\displaystyle [\xi _1,\frac{\text {d}}{{\text {d}}t_\beta }]=0\) \((\beta =2,3)\). Then, it is easy to see that \(\displaystyle d\eta _1(\xi _1,A)=d\eta _1(\xi _1, \frac{\text {d}}{{\text {d}}t_\beta })=0\). As \(\displaystyle \frac{\text {d}}{{\text {d}}t_\beta }\) \((\beta =2,3)\) are induced from the central subgroup \((0,{\mathbb {R}}^2)\) of \({\mathbb {R}}^3\), (5.14) shows \(\displaystyle {\text {d}}\eta _1(\frac{\text {d}}{{\text {d}}t_\beta }, B)=0\), \(\displaystyle d\eta _1(\frac{\text {d}}{{\text {d}}t_\beta },\frac{\text {d}}{{\text {d}}t_\gamma })=0\) \(({}^\forall \, B\in {{\mathsf {D}}}_0)\). \(\square\)

Lemma 5.3

Let \(\hat{\Omega }_1\) be the 2-form on Y as in (5.12). Then, \(d\hat{\Omega }_1=0\).

Proof

It suffices to show

$$\begin{aligned} \pi _1^*\hat{\Omega }_1={\text {d}}\eta _1\quad \text{ on } Y. \end{aligned}$$
(5.15)

\(\pi _1^*\hat{\Omega }_1={\text {d}}\eta _1\) on \({{\mathsf {D}}}_0\) from (5.12). For any \(\xi \in {{\mathsf {E}}}_1\) and \(A\in TX\), \(\displaystyle {\text {d}}\eta _1(\xi , A)=0\) by Lemma 5.2. As \({{\mathsf {E}}}_1\oplus {{\mathsf {D}}}_0=TX\), it follows \(\pi _1^*\hat{\Omega }_1={\text {d}}\eta _1\) on X. \(\square\)

6 Pseudo-Hermitian structure

6.1 Heisenberg Lie group \({{\mathcal {N}}}\)

Let \({{\mathcal {N}}}\) be the \(4n+1\)-dimensional Heisenberg Lie group which has a central group extension \(\displaystyle 1{\rightarrow }{\mathbb {R}}{\rightarrow }\, {{\mathcal {N}}}{\mathop {{\longrightarrow }}\limits ^{p}}{\mathbb {C}}^{2n}{\rightarrow }1\). A pseudo-Hermitian structure \((\omega _{{\mathcal {N}}},J_{{\mathcal {N}}})\) consists of a contact form

$$\begin{aligned} \omega _{{\mathcal {N}}}={\text {d}}t+\mathfrak {I}\langle (z,w),({\text {d}}z,{\text {d}}w)\rangle =dt+\mathfrak {I}({}^t\bar{z}dz+{}^t{\bar{w}}dw) \ \ \, ((z,w)\in {\mathbb {C}}^{2n})\end{aligned}$$

together with a complex structure \(J_{{\mathcal {N}}}\) on \(\mathrm{ker}\, \omega _{{\mathcal {N}}}\) which is isomorphic to the standard complex structure on \({\mathbb {C}}^{2n}\) at each point of \({{\mathcal {N}}}\) (cf. [11]). As \({\mathbb {R}}^2=\{(0,t_2,t_3)\}\) is a central subgroup of \({\mathbb {R}}^3=C({{\mathcal {M}}})\), there is a quotient nilpotent Lie group \({{\mathcal {M}}}/{\mathbb {R}}^2\) with central group extension \(\displaystyle 1{\rightarrow }{\mathbb {R}}{\rightarrow }\, {{\mathcal {M}}}/{\mathbb {R}}^2{\mathop {{\longrightarrow }}\limits ^{\hat{\pi }}}{\mathbb {H}}^n{\rightarrow }1\). We shall find an explicit isomorphism to identify \({{\mathcal {M}}}/{\mathbb {R}}^2\) with \({{\mathcal {N}}}\). For our use, let \(z+wj\in {\mathbb {H}}^n\) such that \(z,w\in {\mathbb {C}}^n\). Then, \({{\mathcal {M}}}/{\mathbb {R}}^2\) is the product \(\displaystyle {\mathbb {R}}\times {\mathbb {H}}^n=\mathrm{Im}\, {\mathbb {H}}\,( \mathrm{mod}\, {\mathbb {R}}j+{\mathbb {R}}k)\times {\mathbb {H}}^n\) with group law:

$$\begin{aligned} \begin{aligned} \bigl (a,z+wj\bigr )\cdot \bigl (b,z'+w'j\bigr )&=\bigl (a+b-\mathfrak {I}({}^t\bar{z}z'+{}^tw{\bar{w}}'),z+z'+(w+w')j\bigr ). \end{aligned} \end{aligned}$$

Define a diffeomorphism \(\varphi : {{\mathcal {M}}}/{\mathbb {R}}^2={\mathbb {R}}\times {\mathbb {H}}^n{\rightarrow }\, {{\mathcal {N}}}={\mathbb {R}}\times {\mathbb {C}}^{2n}\) to be

$$\begin{aligned} \varphi (a,(z+wj))=(a,(z,{\bar{w}})). \end{aligned}$$
(6.1)

As we see that \(\displaystyle \varphi ((a,z+wj)\cdot (b,z'+w'j)) =\bigl (a,(z,{\bar{w}})\bigr )\cdot \bigl (b,(z',{\bar{w}}')\bigr )\),

Lemma 6.1

\(\varphi\) is a Lie group isomorphism of \({{\mathcal {M}}}/{\mathbb {R}}^2\) onto \({{\mathcal {N}}}\).

Consider the projection

(6.2)

for which the subbundle \(\displaystyle \langle \frac{\text {d}}{{\text {d}}t_2}, \frac{\text {d}}{{\text {d}}t_3}\rangle\) is tangent to the fiber \({\mathbb {R}}^2\). For \(\omega _0=\omega _1i+\omega _2j+\omega _3k\), noting (3.3), \(\omega _1\) induces a 1-form \(\hat{\omega }_1\) on \({{\mathcal {M}}}/{\mathbb {R}}^2\) such that

$$\begin{aligned} p_1^*{\hat{\omega }}_1=\omega _1\ \, \text{ on }\ {{\mathcal {M}}}. \end{aligned}$$
(6.3)

As \(\displaystyle {\hat{\omega }}_1={\text {d}}t_1+\mathfrak {I}\langle (z,{\bar{w}}),({\text {d}}z,{\text {d}}\bar{w})\rangle\) from (3.2) \((z+wj\in {\mathbb {H}}^n)\), (6.1) shows

$$\begin{aligned} \varphi ^*\omega _{{\mathcal {N}}}=\hat{\omega }_1\ \text{ on }\ {{\mathcal {M}}}/{\mathbb {R}}^2. \end{aligned}$$
(6.4)

Let \({\hat{\varphi }}:{\mathbb {H}}^n{\rightarrow }{\mathbb {C}}^{2n}\) be a diffeomorphism defined by

$$\begin{aligned} {\hat{\varphi }}(z+wj)=(z,{\bar{w}}), \end{aligned}$$
(6.5)

with the following commutative diagram from (6.1):

(6.6)

Take the standard complex structure \(J_1\) on \({\mathbb {H}}^n\) such that \(\displaystyle J_1 u=u{\bar{i}}\). As \(\displaystyle u{\bar{i}}=(z+wj)\bar{i}=({\bar{i}}z+iwj)\), it follows \(\displaystyle \hat{\varphi }(u\bar{i})={\bar{i}}\hat{\varphi }(u)\). If we take the anti-complex structure \(J_{\mathbb {C}}'\) on \({\mathbb {C}}^{2n}\) such as \(\displaystyle J_{\mathbb {C}}'(v)={\bar{i}} v\), it follows

$$\begin{aligned} \hat{\varphi }_*\circ J_1=J_{\mathbb {C}}'\circ \hat{\varphi }_*: T{\mathbb {H}}^n{\rightarrow }T{\mathbb {C}}^{2n}. \end{aligned}$$
(6.7)

As in (3.4) of Sect. 3.2, the almost complex structure \(J_1\) on \({{\mathsf {D}}}_0\) induces an almost complex structure \(J_1\) on \(p_{1*}{{\mathsf {D}}}_0={\ker }\,\hat{\omega }_1\) such that

$$\begin{aligned} \hat{\pi }_{*}\circ J_1=J_1\circ \hat{\pi }_{*}\, : \ p_{1*}{{\mathsf {D}}}_0\,{\rightarrow }T{\mathbb {H}}^n. \end{aligned}$$
(6.8)

Similarly, if \(J'_{{\mathcal {N}}}\) is the anti-complex structure on \({\ker }\,\omega _{{\mathcal {N}}}\) of \({{\mathcal {N}}}\), it follows

$$\begin{aligned} p_{*}\circ J'_{{\mathcal {N}}}=J'_{\mathbb {C}}\circ p_{*} : \mathrm{ker}\, \omega _{{\mathcal {N}}}{\rightarrow }T{\mathbb {C}}^{2n}. \end{aligned}$$
(6.9)

Then, (6.7), (6.8) and (6.9) imply

$$\begin{aligned} \varphi _*\circ J_1=J'_{{\mathcal {N}}}\circ \varphi _* \text{ on } {\ker }\,\hat{\omega }_1\, (=p_{1*}{{\mathsf {D}}}_0). \end{aligned}$$
(6.10)

In particular, \(J_1\) is a complex structure on \(p_{1*}{{\mathsf {D}}}_0={\ker }\,\hat{\omega }_1\). We have the principal bundle induced from (5.9):

(6.11)

There is a CR-structure \(({\ker }\,\hat{\omega }_1,J_1)\) on \(X/{\mathbb {R}}^2\subset {{\mathcal {M}}}/{\mathbb {R}}^2\) as above. Let \({\hat{J}}_1\) be the almost complex structure on Y as in (5.11), that is \(\displaystyle {\hat{\pi }}_{1*}J_1={\hat{J}}_1{\hat{\pi }}_{1*} : {\ker }\,\hat{\omega }_1{\rightarrow }TY\).

Lemma 6.2

\({\hat{J}}_1\) is integrable on Y.

Proof

Let \(\displaystyle {\ker }\,\hat{\omega }_1\otimes {\mathbb {C}}=P^{1,0}\oplus P^{0,1}\) be the eigenspace decomposition for \(J_1\). We have an isomorphism \(\displaystyle \hat{\pi }_{1*}:{\ker }\,\hat{\omega }_1\otimes {\mathbb {C}}{\longrightarrow }TY\otimes {\mathbb {C}}=Q^{1,0}\oplus Q^{0,1}\) (the eigenspace decomposition for \({\hat{J}}_1\), respectively). Since \(J_1\) is integrable, \(\displaystyle [\varvec{u},\varvec{v}]\in P^{1,0}\) for \(\varvec{u},\varvec{v}\in P^{1,0}\). Then, \(\hat{\pi }_{1*}([\varvec{u},\varvec{v}])=[\hat{\pi }_{1*}(\varvec{u}),\hat{\pi }_{1*}(\varvec{v})]\). It follows \(\displaystyle [\hat{\pi }_{1*}(\varvec{u}),\hat{\pi }_{1*}(\varvec{v})]\in Q^{1,0}\). Thus, \({\hat{J}}_1\) is integrable. \(\square\)

Combining this lemma with Lemma 5.3, we obtain

Proposition 6.3

\((Y,\{\hat{\Omega }_1,{\hat{J}}_1\})\) is a Kähler manifold.

6.2 Bochner flat structure on \((Y,{\hat{J}}_1)\)

Put \(\displaystyle e^{iat}z=(e^{ia_1t}z_1,\dots ,e^{ia_nt}z_n)\) for short, similarly for \(e^{-iat}w\) for \(a=(a_1,\ldots ,a_n)\) satisfying (5.2). Let \({\mathbb {R}}\) act on \({{\mathcal {N}}}\) by

$$\begin{aligned} \rho (t)(s,(z,w))=(t+s, (e^{iat}z, e^{-iat}w)) \end{aligned}$$
(6.12)

such that \(\rho ({\mathbb {R}})\le {\mathbb {R}}\times \mathrm{U}(2n)\le {{\mathcal {N}}}\rtimes \mathrm{U}(2n)\le \mathop {\mathrm{Aut}}\nolimits _{CR}({{\mathcal {N}}})\). There induces another principal bundle \(\displaystyle \rho ({\mathbb {R}}){\rightarrow }\, {{\mathcal {N}}}{\mathop {{\longrightarrow }}\limits ^{q}}{\mathbb {C}}^{2n}\). By (6.1) and (5.5),

$$\begin{aligned}\begin{aligned} \varphi \bigl (\rho _1(t)(s,(z+wj))\bigr )&= \varphi \bigl (s+t, (e^{iat}z+e^{iat}wj)\bigr )\\&=\bigl (s+t, (e^{iat}z, e^{-iat}{\bar{w}})\bigr ) =\rho (t)\varphi \bigl (s,(z+wj)\bigr ), \end{aligned}\end{aligned}$$

there is the bundle isomorphism (cf. (5.9)):

(6.13)

Let \(\xi\) be the vector field induced by \(\rho ({\mathbb {R}})\) on \({{\mathcal {N}}}\). Put \(\displaystyle p_{1*}\xi _1={\hat{\xi }}_1\) from (5.6), (6.2), which is the Reeb field of \(({{\mathcal {M}}}/{\mathbb {R}}^2,(\hat{\omega }_1,J_1))\). We have \(\displaystyle \varphi _*{\hat{\xi }}_1=\xi\). At \(p=((s_1,s_2,s_3), z_1,\dots ,z_n)\in {{\mathcal {M}}}\) with \(z_k=u_k+v_k j\), (6.4), (6.3) and (5.7) imply

$$\begin{aligned} \omega _{{\mathcal {N}}}(\xi ) =\hat{\omega }_1({\hat{\xi }}_1)= \omega _1(\xi _1)=1+\sum _{k=1}^n a_k(|u_k|^2-|v_k|^2). \end{aligned}$$
(6.14)

Corresponding to \(X/{\mathbb {R}}^2=\{x\in p_1(X)\mid \hat{\omega }_1((\hat{\xi }_1)_x)\ne 0\}\), the bundle isomorphism \(\varphi\) maps \(X/{\mathbb {R}}^2\) onto the domain \(\displaystyle {{\mathcal {N}}}_1=\{z\in {{\mathcal {N}}}\mid \omega _{{\mathcal {N}}}(\xi _z)\ne 0\}\) of \({{\mathcal {N}}}\). As in (5.10), define the contact forms to be

$$\begin{aligned} \begin{aligned} \hat{\eta }_1&=\frac{1}{{\hat{\omega }}_1(\hat{\xi }_1)}\hat{\omega }_1 \, \text{ on }\ X/{\mathbb {R}}^2, \quad \eta _{{{\mathcal {N}}}_1}=\frac{1}{\omega _{{\mathcal {N}}}(\xi )}\omega _{{\mathcal {N}}}\ \, \text{ on } {{\mathcal {N}}}_1. \end{aligned} \end{aligned}$$
(6.15)

In particular, \(p_{1*}{{\mathsf {D}}}_0={\ker }\,\hat{\eta }_1\). Noting (6.14), (6.4) shows that

$$\begin{aligned} \varphi ^*{\eta _{{{\mathcal {N}}}_1}}=\hat{\eta }_1. \end{aligned}$$

Put \(J_{{{\mathcal {N}}}_1}=J_{{{\mathcal {N}}}}|_{{{\mathcal {N}}}_1}\). Since \((\eta _{{{\mathcal {N}}}_1},J_{{{\mathcal {N}}}_1})\) represents a spherical pseudo-Hermitian structure on \({{\mathcal {N}}}_1\), this equation together with (6.10) implies (cf. [11])

Proposition 6.4

The pseudo-Hermitian structure \((X/{\mathbb {R}}^2,\hat{\eta }_1, J_1)\) is anti-holomorphically isomorphic to the spherical CR-structure \(({{\mathcal {N}}}_1,{\eta _{{{\mathcal {N}}}_1}}, J_{{{\mathcal {N}}}_1})\).

The projection \(\displaystyle \hat{\pi }_1:X/{\mathbb {R}}^2{\rightarrow }Y\) of (6.11) induces the following from (5.15), (5.11):

$$\begin{aligned} \begin{aligned} \hat{\pi }_1^*\hat{\Omega }_1&=d\hat{\eta }_1 \, \text{ on } \,X/{\mathbb {R}}^2,\\ {\hat{J}}_\alpha \circ \hat{\pi }_{1*}&=\hat{\pi }_{1*}\circ J_\alpha \, \text{ on } p_{1*}{{\mathsf {D}}}_0\ \, (\alpha =1,2,3). \end{aligned} \end{aligned}$$
(6.16)

Theorem 6.5

\((Y,{\hat{g}}, \{\hat{\Omega }_1, {\hat{J}}_1\})\) is an anti-holomorphic Bochner flat Kähler manifold for the quaternionic Hermitian manifold \((Y, {\hat{g}},\{\hat{\Omega }_\alpha , {\hat{J}}_\alpha \}_{\alpha =1}^3)\).

Proof

Since \(d\hat{\eta }_1=\hat{\pi }_1^*\hat\Omega _1\) from \(X/{\mathbb {R}}^2\) by (6.16), \(\displaystyle \rho _1({\mathbb {R}}){\rightarrow }X/{\mathbb {R}}^2{\mathop {{\longrightarrow }}\limits ^{\hat{\pi }_1}}Y\) is a pseudo-Hermitian bundle with the Reeb field \(\hat{\xi }_1\), the Bochner curvature tensor of the Kähler manifold \((Y, \{\hat{\Omega }_1, {\hat{J}}_1\})\) coincides with the Chern–Moser curvature tensor on \((X/{\mathbb {R}}^2, \hat{\eta }_1,J_1)\) by the pull-back of \(\hat{\pi }_1\) in (6.16) (cf. [16]). As \(X/{\mathbb {R}}^2\) is spherical CR by Proposition 6.4, the Chern–Moser curvature tensor is zero and thus \((Y,\{\hat{\Omega }_1,{\hat{J}}_1\})\) is a Bochner flat manifold. \(\square\)

7 Pseudo-Hermitian group \(\mathop {\mathrm{Psh}}\nolimits (X/{\mathbb {R}}^2)\)

We determine the holomorphic isometry group \(\displaystyle \mathop {\mathrm{Isom}}\nolimits _{\!\, h}(Y)=\{f\in \mathop {\mathrm{Diff}}\nolimits (Y)\mid f^*\hat{\Omega }_1=\hat{\Omega }_1, \, f_*\circ {\hat{J}}_1={\hat{J}}_1\circ f_*\}\) of the Kähler manifold \((Y,(\hat{\Omega }_1,{\hat{J}}_1))\). In order to do so, consider the pseudo-Hermitian group of the pseudo-Hermitian manifold \((X/{\mathbb {R}}^2,\hat{\eta }_1, J_1)\) (cf. [11]):

$$\begin{aligned} \mathop {\mathrm{Psh}}\nolimits (X/{\mathbb {R}}^2)=\{{{\tilde{f}}}\in \mathop {\mathrm{Diff}}\nolimits (X/{\mathbb {R}}^2) \mid {{\tilde{f}}}^*\hat{\eta }_1={\hat{\eta }}_1,\ {{\tilde{f}}}_*\circ J_1=J_1\circ {{\tilde{f}}}_*\}. \end{aligned}$$

As \(H^1(Y;{\mathbb {R}})=0\) (see the remark below (5.9)), it associates the exact sequence from [4, Proposition 3.4]:

(7.1)

Since \(\rho _1({\mathbb {R}})\) induces the Reeb field \({\hat{\xi }}_1\) of \(\hat{\eta }_1\), \(\mathop {\mathrm{Psh}}\nolimits (X/{\mathbb {R}}^2)\) itself is the centralizer of \(\rho _1({\mathbb {R}})\) in \(\mathop {\mathrm{Psh}}\nolimits (X/{\mathbb {R}}^2)\) (cf. [4, 11]). If we recall the representation of \(\rho ({\mathbb {R}})\) from (6.12):

$$\begin{aligned} \rho (t)=\bigl (t,(e^{ia_1t},\dots ,e^{ia_nt}, e^{-ia_1t},\dots ,e^{-ia_nt})\bigr )\in {\mathbb {R}}\times \mathrm{U}(2n)\le {{\mathcal {N}}}\rtimes \mathrm{U}(2n), \end{aligned}$$

under the equivariant diffeomorphism \(\varphi\) of (6.13), the condition (5.2) implies that

$$\begin{aligned} \mathop {\mathrm{Psh}}\nolimits (X/{\mathbb {R}}^2)={\mathbb {R}}\times T^n\times T^n \le {\mathbb {R}}\times \mathrm{U}(2n). \end{aligned}$$
(7.2)

We have from (7.1) that

Proposition 7.1

\(\mathop {\mathrm{Isom}}\nolimits _{\!\, h}(Y)=T^{2n}\).

Theorem 7.2

The quaternionic Hermitian manifold \((Y,{\hat{g}},\{{\hat{J}}_\alpha \}_{\alpha =1}^3)\) is not Einstein. Moreover, it is never holomorphically isometric to any domain of the quaternionic euclidean space \({\mathbb {H}}^n\).

Proof

When the Bochner flat manifold \((Y,{\hat{g}},\{\Omega _1,{\hat{J}}_1\})\) is Einstein, it is of constant holomorphic curvature by Tachibana’s result [15]. Thus, Y is locally holomorphically isometric to the flat space \({\mathbb {C}}^{2n}\). We may assume that the origin \(\varvec{0}\) belongs to Y (cf. (5.7)). Then, the stabilizer at \(\varvec{0}\) is the maximal compact subgroup isomorphic to \(\mathop {\mathrm{U}}\nolimits (2n)\). Since \(T^{2n}\) is the full holomorphic isometry group of \((Y,{\hat{g}},{\hat{J}}_1)\) by Proposition 7.1, it is impossible. \((Y,{\hat{g}},{\hat{J}}_1)\) is not holomorphically isometric to any domain of \(({\mathbb {H}}^n,g_{\mathbb {H}})\) with the standard euclidean metric \(g_{\mathbb {H}}\). \(\square\)

7.1 Quaternionic Hermitian isometry group of \((Y,{\hat{g}})\)

The quaternionic Hermitian isometry group of the quaternionic Hermitian manifold \((Y, {\hat{g}},\{\hat{\Omega }_\alpha ,{\hat{J}}_\alpha \}_{\alpha =1}^3)\) may be denoted naturally by the following:

$$\begin{aligned} \mathrm{Isom}_{\,\!qH}(Y)=\bigl \{ {\hat{h}}\in \mathop {\mathrm{Diff}}\nolimits (Y) \ \big |\ {{\hat{h}}}^*{\hat{\Omega }}_\alpha = \sum _{\beta =1}^{3}\hat{\Omega }_\beta {\hat{a}}_{\beta \alpha },\ {\hat{h}}_* \circ {\hat{J}}_\alpha =\sum _{\beta =1}^{3}{\hat{a}}_{\alpha \beta }{\hat{J}}_\beta \circ {\hat{h}}_*\bigr \}, \end{aligned}$$

where \(\displaystyle ({\hat{a}}_{\alpha \beta })_{\alpha ,\beta =1,2,3}:Y{\longrightarrow }\mathrm{SO}(3)\) are smooth maps.

For the abelian group \(\displaystyle \rho _1({\mathbb {R}}^3)\) defined by (5.4), let \(N_{\mathop {\mathrm{Aut}}\nolimits ({{\mathcal {M}}})}(\rho _1({\mathbb {R}}^3))\) be the normalizer of \(\rho _1({\mathbb {R}}^3)\) in \(\mathop {\mathrm{Aut}}\nolimits ({{\mathcal {M}}})\). By the formula (5.3) of \(A_t\), the normalizer of \(\{A_{t}\}\) in \(\mathop {\mathrm{Sp}}\nolimits (n)\) is isomorphic to \(T^n\). Since the only subgroup \(S^1=\langle e^{i\theta }\rangle\) of \(\mathop {\mathrm{Sp}}\nolimits (1)\) normalizes \(\rho _1({\mathbb {R}}^3)\) in view of the actions (3.1) and (5.4), it follows that

$$\begin{aligned} \begin{aligned} N_{\mathop {\mathrm{Aut}}\nolimits ({{\mathcal {M}}})}(\rho _1({\mathbb {R}}^3))&=({\mathbb {R}}^3,0)\times T^n\cdot S^1 \le {{\mathcal {M}}}\rtimes \mathop {\mathrm{Sp}}\nolimits (n)\cdot \mathop {\mathrm{Sp}}\nolimits (1) \end{aligned}\end{aligned}$$
(7.3)

where \(\displaystyle T^n\cdot S^1=T^n\times _{\{\pm 1\}} S^1\). Recall every element of \(\mathop {\mathrm{Aut}}\nolimits (X)\) extends to an element of \(\mathop {\mathrm{Aut}}\nolimits ({{\mathcal {M}}})\). Since each element of \(N_{\mathop {\mathrm{Aut}}\nolimits ({{\mathcal {M}}})}(\rho _1({\mathbb {R}}^3))\) preserves both \(\omega _1\) and \(\xi _1\) from (5.8), it follows \(\displaystyle N_{\mathop {\mathrm{Aut}}\nolimits (X)}(\rho _1({\mathbb {R}}^3))=N_{\mathop {\mathrm{Aut}}\nolimits ({{\mathcal {M}}})}(\rho _1({\mathbb {R}}^3))\).

Let \(\eta =\eta _1i+\eta _2j+\eta _3k\) be as before (cf. (5.10)). Note \(h^*\eta =a\cdot \eta \cdot {\bar{a}}\) for \(h=((t,0),A\cdot a)\in N_{\mathop {\mathrm{Aut}}\nolimits (X)}(\rho _1({\mathbb {R}}^3))\). By (7.3), the projection \(\pi _1:X{\rightarrow }Y\) of (5.9) induces an element \({\hat{h}}:Y{\rightarrow }Y\). Since \(\pi _1^*\hat{\Omega }=d\eta\) for \(\hat{\Omega }=\hat{\Omega }_1i+\hat{\Omega }_2j+\hat{\Omega }_3k\), we have \({\hat{h}}^*\hat{\Omega }= a\cdot \hat{\Omega }\cdot {\bar{a}}\) for \(a\in S^1\). Thus, it assigns an element \({\hat{h}}\in \mathop {\mathrm{Isom}}\nolimits _{\!\,qH}(Y, {\hat{g}},\{\hat{\Omega }_\alpha ,{\hat{J}}_\alpha \}_{\alpha =1}^3) =\mathop {\mathrm{Isom}}\nolimits _{\!\, qH}(Y)\). Letting \(a=(a_{\alpha \beta })_{\alpha ,\beta =2,3}\in \mathrm{SO}(2)\), \(\mathop {\mathrm{Isom}}\nolimits _{\!\, qH}(Y)\) can be described as

$$\begin{aligned} \begin{aligned} \bigl \{{\hat{h}}\in \mathop {\mathrm{Diff}}\nolimits (Y)\mid&\ {\hat{h}}^*{\hat{\Omega }}_1=\hat{\Omega }_1, {\hat{h}}_*\circ {\hat{J}}_1 = {\hat{J}}_1\circ {\hat{h}}_*, \\&\quad {\hat{h}}^*\hat{\Omega }_\alpha = \sum _{\beta =2,3}\hat{\Omega }_\beta a_{\beta \alpha }, \quad {\hat{h}}_* \circ {\hat{J}}_\alpha \, =\, \sum _{\beta =2,3} a_{\alpha \beta }{\hat{J}}_\beta \circ {\hat{h}}_*\ \bigr \}. \end{aligned} \end{aligned}$$
(7.4)

Setting \(\displaystyle \tilde{\phi }(h)={\hat{h}}\), (7.3) gives an exact sequence:

(7.5)

If \(\iota : \mathop {\mathrm{Isom}}\nolimits _{\!\,qH}(Y){\rightarrow }\mathop {\mathrm{Isom}}\nolimits _{h}(Y)=T^{2n}\) (cf. Proposition 7.1) is the natural inclusion (that is, forgetting the quaternionic structure but leaving the holomorphic structure as it is), then under the equivariant diffeomorphism \(\hat{\varphi }\) of (6.5), it follows \(\displaystyle \iota (T^{n}\cdot S^1)= \{(z_1,\ldots ,z_n,{\bar{z}}_1,\ldots , \bar{z}_n)\times (e^{-i\theta },\ldots ,e^{-i\theta })\}\le T^{2n}\) where \(z_i\in S^1\) \((i=1,\ldots ,n)\). We obtain

Corollary 7.3

The quaternionic Hermitian isometry group \(\mathop {\mathrm{Isom}}\nolimits _{\!\,qH}(Y)^0\) is isomorphic to the torus \(T^k\) for some k \((n+1\le k\le 2n)\).

Remark 7.4

Since the forms \(\hat{\Omega }_2, \hat{\Omega }_3\) are not Kähler, the equation \(\displaystyle \pi _{1}^*\hat{\Omega }_\alpha =d\eta _\alpha\) does not hold on X (only on \({{\mathsf {D}}}_0\)), the method of [4, Propositions 3.4, 3.1, 3.2] cannot be applied to show the surjectivity of \(\tilde{\phi }\) in (7.5).