Next Article in Journal
Model for Predicting Maize Crop Yield on Small Farms Using Clusterwise Linear Regression and GRASP
Previous Article in Journal
Enhancing Efficacy in Breast Cancer Screening with Nesterov Momentum Optimization Techniques
Previous Article in Special Issue
Rigidity and Triviality of Gradient r-Almost Newton-Ricci-Yamabe Solitons
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Twistor and Reflector Spaces for Paraquaternionic Contact Manifolds

1
Faculty of Mathematics and Informatics, University of Sofia, Blvd. James Bourchier 5, 1164 Sofia, Bulgaria
2
Institute of Mathematics and Informatics, Bulgarian Academy of Sciences, 1000 Sofia, Bulgaria
*
Author to whom correspondence should be addressed.
Mathematics 2024, 12(21), 3355; https://doi.org/10.3390/math12213355
Submission received: 8 October 2024 / Revised: 20 October 2024 / Accepted: 22 October 2024 / Published: 25 October 2024
(This article belongs to the Special Issue Differentiable Manifolds and Geometric Structures)

Abstract

:
We consider certain fiber bundles over paraquaternionic contact manifolds, called twistor and reflector spaces. We show that the twistor space carries an integrable CR structure (Cauchy–Riemann structure) and the reflector space is an integrable para-CR structure, both with neutral signatures.
MSC:
58G30; 53C17

1. Introduction

The geometry of paraquaternionic contact structures is essentially a tool to study a special type of co-dimension three distribution on ( 4 n + 3 ) manifolds with properties closely related to the algebra of paraquaternions, known also as split-quaternions [1], quaternions of the second kind [2], and complex product structures [3]. The paraquaternionic contact structure, introduced in [4], can be considered a generalization of the para three-Sasakian geometry developed in [1,5]. In many ways, paraquaternionic contact structures resemble the geometry of quaternionic contact manifolds, introduced by O. Biquard [6], which has been very useful in relation to the quaternionic contact Yamabe problem and the determination of extremals and the best constant in the L 2 Folland–Stein inequality on the quaternionic Heisenberg group [7,8,9,10,11]. Despite the similarities between these two types of geometry, there are also some major differences determined mainly by the fact that in the paraquaternionic contact setting, one is often forced to consider sub-hyperbolic PDEs instead of sub-elliptic PDEs.
As shown in [6], the study of quaternionic contact structures leads back in a natural way to the study of a particular class of integrable CR manifolds (which are never pseudo-convex), called twistor spaces, which appear as certain sphere bundles over the base quaternionic contact manifold (see also [12]). This is a generalization of the concept of a twistor space of a quaternionic Kähler manifold [13]. In the paraquaternionic contact case, we have two different types of bundles: the twistor space Z and the reflector space R . The situation is very similar to the discussion in [14]. The fibers of Z are diffeomorphic to the two-sheeted hyperboloid x 2 + y 2 z 2 = 1 in R 3 , whereas the fibers of R are diffeomorphic to the one-sheeted hyperboloid x 2 + y 2 z 2 = 1 (see Section 3 below for the details). The purpose of this paper is to demonstrate the following:
Theorem 1. 
If ( M , H ) is any paraquaternionic contact manifold with twistor space Z and reflector space R , then we have a natural integrable CR structure on Z and a natural integrable para-CR structure on R . The Levi form for each of these structures is of signature ( 2 n + 2 , 2 n + 2 ) .
The proof of this theorem is divided into several steps throughout the paper and follows the results obtained in Propositions 1, 2, 3, and 5.
Conventions. 
In this paper we use the following general conventions:
(a) 
Indices s and t usually run from 1 to 3 (when nothing else specified).
(b) 
Indices i , j , k always represent a positive (cyclic) permutation of 1 , 2 , 3 .
(c) 
The summation symbol ( i j k ) indicates summation over all positive permutations ( i j k ) of 1 , 2 , 3 ; that is,
( i j k ) X i j k = X 123 + X 231 + X 312 .
(d) 
We fix the following signs: ϵ 1 = 1 , ϵ 2 = 1 , and ϵ 3 = 1 .

2. Preliminaries

2.1. CR and Para-CR Structures on Manifolds

A CR structure (or a Cauchy–Riemann structure) on a differentiable manifold is a type of geometric structure that models the geometry of a real hypersurface in a complex manifold. Formally, a CR manifold is a differentiable manifold N of odd dimension, say 2 n + 1 , endowed with a complex subbundle K of the complexified tangent bundle C T N = T N R C , so that the fibers of K are of complex dimension n; [ K , K ] K (i.e., K is formally integrable), and K K ¯ = { 0 } .
If we set D to be the real component of K K ¯ , then D is a 2n-dimensional (real) distribution on N. There is a natural field J of endomorphisms of the distribution D with the following properties: J 2 = Id D ; the fibers of K and K ¯ are eigenspaces of J with eigenvalues of 1 and 1 , respectively. The Levi form of the CR structure ( D , J ) is a vector-valued hermitian 2-form L, defined on D, where the values in the line bundle T N / D . L is given by the following formula:
L ( x , y ) = [ x , J y ] m o d D , x , y D .
For a more detailed discussion on this topic, see [15].
Similarly, a para-CR structure on a 2 n + 1 -dimensional differentiable manifold N can be defined as a pair ( D , J ) of a co-dimension distribution D on N and a field of endomorphisms J of D with the following properties: J 2 = Id D and J ± Id D ; [ K , K ] K and [ K ˜ , K ˜ ] K ˜ , where K and K ˜ are now the 1 and 1 eigenspaces of J. The Levi form, in this case, is a vector-valued symmetric 2-form L, defined on D, with values in the line bundle T N / D , which are given again by Formula (1). See, for example, ref. [16] or [17] and the references contained therein for a more detailed discussion on para-CR manifolds and their applications.

2.2. The Algebra of Split-Quaternions

Both the quaternions and the split-quaternions are real Clifford algebras generated by a two-dimensional non-degenerate quadratic form. In the negative-definite case, we obtain the algebra of quaternions, whereas in the other two cases, i.e., of a positive-definite or indefinite quadratic form, we get the same (up to an isomorphism) Clifford algebra, which is denoted here by B and is called the algebra of the split-quaternions (or paraquaternions) (see, e.g., [1]). The elements of B are generally represented in the following form:
a = a 0 + a 1 j 1 + a 2 j 2 + a 3 j 3 ,
where a s are real numbers, and j s are basic split-quaternions; that is, some fixed elements of B satisfy the following identities:
j 1 2 = j 2 2 = 1 , j 1 j 2 = j 2 j 1 = j 3 .
The remaining multiplication rules for B are easily derived from the following:
j 3 2 = 1 , j 2 j 3 = j 3 j 2 = j 1 , j 3 j 1 = j 1 j 3 = j 2 .
The conjugate to a is defined by a ¯ = a 0 a 1 j 1 a 2 j 2 a 3 j 3 . We obtain the typical identity a b ¯ = b ¯ a ¯ . The real and imaginary parts of a split-quaternion are given by R e ( a ) = a 0 and I m ( a ) = a 1 j 1 + a 2 j 2 + a 3 j 3 . There is a natural inner product on I m ( B ) = R 3 ,
a , b = R e ( a b ) = a 1 b 1 a 2 b 2 + a 3 b 3 ,
and a cross product “×”,
a × b = s , t = 1 s t 3 ( a s b t ) j s j t ,
so that
a × b , c = det a 1 b 1 c 1 a 2 b 2 c 2 a 3 b 3 c 3 , a , b , c I m ( B ) .
We observe that B is isomorphic to the algebra M 2 ( R ) of all 2 × 2 matrices with real entries under the identification
j 1 = 1 0 0 1 , j 2 = 0 1 1 0 , j 3 = 0 1 1 0 .
Let S O ( 1 , 2 ) be the group of all 3 × 3 real matrices of determinant 1 that preserve the inner product (2). We need the following basic lemma, which is easily derived from the multiplication rules of B .
Lemma 1. 
Three split-quaternions, γ 1 , γ 2 , and γ 3 , satisfy the identities
γ 1 2 = γ 2 2 = 1 , γ 1 γ 2 = γ 2 γ 1 = γ 3 ,
if and only if there exists a matrix A = ( a s t ) S O ( 1 , 2 ) so that γ s = t a s t j t , s = 1 , 2 , 3 .
If we regard the vector space B n (the elements of B n are thought of as column vectors) as a right B module, the multiplication from the left with n × n matrices with entries in B represents the space of all B -linear endomorphisms of B n . We define S p ( n , B ) to be the group of all B -linear transformations that preserve the inner product x , y = R e ( x ¯ T y ) , x , y B n ,
S p ( n , B ) = A M n ( B ) : A ¯ T A = 1 .
In particular, S p ( 1 , B ) is the group of the unite split-quaternions,
S p ( 1 , B ) = z = z 0 + z 1 j 1 + z 2 j 2 + z 3 j 3 : z 0 2 z 1 2 z 2 2 + z 3 2 = 1 .
Consider the action of the direct product S p ( n , B ) × S p ( 1 , B ) on the vector space B n , defined by
( A , z ) · x = A x z ¯ ,
and let us fix (once and for all) identification B n = R 4 n . Since the induced inner product is of signature ( 2 n , 2 n ) , we obtain an embedding of the quotient group
S p ( n , B ) × S p ( 1 , B ) ± ( 1 , 1 )
into the matrix group S O ( 2 n , 2 n ) . The image of this embedding is denoted by S p ( n , B ) S p ( 1 , B ) and consists of all elements of S O ( 2 n , 2 n ) that preserve the three-dimensional subspace Q E n d ( R 4 n ) generated by the right action of I m ( B ) on B n .

2.3. Paraquaternionic Contact Structures

Consider a 4 n -dimensional smooth distribution H on a ( 4 n + 3 ) -dimensional manifold M. Suppose that at each point p in an open subset U M , we are given a triple ( η 1 , η 2 , η 3 ) of 1-forms on T p M , a triple ( I 1 , I 2 , I 3 ) of endomorphisms of H p T p M , and a non-degenerate quadratic form g on H p , all depending smoothly on point p. The list ( η s , I s , g ) is called a (local) paraquaternionic contact (shortly: pqc) structure for H on U if the following three conditions are satisfied at each p U :
(i)
H p = A T p M : η 1 ( A ) = η 2 ( A ) = η 3 ( A ) = 0 ;
(ii)
d η s ( X , Y ) = 2 g ( I s X , Y ) , X , Y H p ,  s = 1, 2, 3;
(iii)
I 1 2 = I 2 2 = id , I 1 I 2 = I 2 I 1 = I 3 .
Clearly, for every local pqc-structure ( η s , I s , g ) for H, the quadratic form g must be of signature (2n,2n). The pair ( M , H ) is called a paraquaternionic contact manifold if, around each point of M, there exists at least one local pqc-structure for H. Here arises the natural question: to what extent are the different local pqc-structures determined by distribution H? The answer is given by the following.
Lemma 2. 
Suppose that ( M , H ) is a pqc manifold. If ( η s , I s , g ) and ( η s , I s , g ) are two pqc-structures for H on an open set U M , then
( η 1 , η 2 , η 3 ) = f ( η 1 , η 2 , η 3 ) S , ( I 1 , I 2 , I 3 ) = ( I 1 , I 2 , I 3 ) S , g = f g ,
for some non-vanishing real valued smooth function f on U and some matrix-valued smooth function S = ( a i j ) : U S O ( 1 , 2 ) .
Proof. 
By assumption H = s = 1 3 K e r ( η s ) = s = 1 3 K e r ( η s ) , there exists a matrix-valued function A = ( a s t ) : U G L ( 3 ) so that η s = t = 1 3 a s t η t , s = 1 , 2 , 3 . Applying the exterior derivative to both sides of this equation and taking the restriction of the resulting two forms to distribution H, we obtain
d η s | H = t a s t ( d η t | H ) .
If G is a field of endomorphisms of H defined by the equation g ( X , Y ) = g ( G X , Y ) , X , Y H , then d η s ( X , Y ) = g ( G I s X , Y ) , and using (6),
G I s = t a s t I t .
This yields
I 1 = ( I 2 ) 1 I 3 = ( G I 2 ) 1 ( G I 3 ) = ( s a 2 s I s ) 1 ( t a 3 t I t ) span R { i d H , I 1 , I 2 , I 3 } ,
similarly to I 2 and I 3 . Let us observe that span R { i d H , I 1 , I 2 , I 3 } E n d ( H ) is an algebra with respect to the usual composition of endomorphisms, which is isomorphic to the algebra of split-quaternions. Therefore, using Lemma 1, we have
span R { I 1 , I 2 , I 3 } = span R { I 1 , I 2 , I 3 } .
In particular, this yields that I 1 , I 2 , I 3 are skew-symmetric with respect to both g and g . Furthermore, we calculate the following:
g G I 1 I 2 + I 2 G I 1 X , Y = g ( G I 3 X , Y ) g ( G I 1 X , I 2 Y ) = g ( I 3 X , Y ) g ( I 1 X , I 2 Y ) = 0 ,
i.e., G I 1 anti-commutes with I 2 , similarly to I 3 . Therefore, G I 1 must be proportional to I 1 , i.e., G is proportional to the identity. This means
g = f g
for some appropriate non-vanishing real-valued function f. The rest follows from Lemma 1. □
An important consequence of the above lemma is that for each pqc manifold ( M , H ) , we can associate a canonical line bundle G ( M ) M so that if ( η s , I s , g ) is a local pqc structure for H, then g is a local section of G ( M ) . Furthermore, the vector bundle π : Q ( M ) M with fiber (over p)
Q p = span { I 1 , I 2 , I 3 } ,
is also globally defined. It has a canonical inner product,
I s , I t = ϵ s , if s = t 0 , otherwise , ϵ 1 = ϵ 2 = ϵ 3 = 1
of signature ( , , + ) and an orientation defined by the ordering of I 1 , I 2 and I 3 .

2.4. Invariant Tensor Decomposition

Let ( M , H ) be a pqc manifold and consider some local pqc-structure ( η s , I s , g ) for H, defined around a fixed p M . Each endomorphism Ψ E n d ( H p ) can be decomposed uniquely into a sum of four components, Ψ = Ψ + + + + Ψ + + Ψ + + Ψ + , where Ψ + + + commutes with I 1 , I 2 , and I 3 and Ψ + commutes with I 1 and anti-commutes with I 2 and I 3 , etc. Explicitly,
4 Ψ + + + = Ψ + I 1 Ψ I 1 + I 2 Ψ I 2 I 3 Ψ I 3 ; 4 Ψ + = Ψ + I 1 Ψ I 1 I 2 Ψ I 2 + I 3 Ψ I 3 ; 4 Ψ + = Ψ I 1 Ψ I 1 + I 2 Ψ I 2 + I 3 Ψ I 3 ; 4 Ψ + = Ψ I 1 Ψ I 1 I 2 Ψ J 2 I 3 Ψ I 3 .
Clearly, this decomposition depends on the particular choice of a pqc structure. To obtain invariant decomposition, we shall consider the action of the Casimir operator † on E n d ( H p ) , given by
( Ψ ) = I 1 Ψ I 1 + I 2 Ψ I 2 I 3 Ψ I 3 .
The leading signs ( + , + , ) in the above summation are opposite to the signature of the invariant inner product on Q p (cf. (8)); therefore, † must be invariant too. It is easily seen that this Casimir operator has eigenvalues 3 and 1 , and that, if Ψ = Ψ [ 3 ] + Ψ [ 1 ] is the induced decomposition of Ψ E n d ( H p ) into a sum of eigenvectors, then
Ψ [ 3 ] = Ψ + + + and Ψ [ 1 ] = Ψ + + Ψ + + Ψ + .

2.5. The Canonical Connection

In general, a pqc manifold ( M , H ) is a parabolic type of geometry that cannot be characterized by a linear connection on the tangent bundle of M; it requires more complicated construction involving a certain Cartan connection, which we shall not deal with here. Instead, we shall use an auxiliary assumption. We require that the naturally induced line bundle G ( M ) M (cf. Section 2.3) admits a global non-vanishing section g; that is, there is a globally defined g on M so that around each point, one can find at least one local pqc structure for H of the form ( η s , I s , g ) (with last entry the same g).
The triple ( M , H , g ) is already a much simpler type of geometry that can be characterized by a unique linear connection ∇ on the tangent bundle of M (as shown in [4]) called the canonical connection of the triple. We shall summarize all the relevant properties of this connection below. Let us first observe that the differential invariants produced by ∇ depend strongly on the choice of g. If we are interested only in the geometry defined by ( M , H ) , we need to consider those differential invariants that remain unchanged after an arbitrary multiplication of g by a non-vanishing function (cf. Lemma 2). The relationship between ( M , H , g ) and ( M , H ) is similar to that between the Riemannian and conformal Riemannian geometry.
In [4], it is shown (with a slightly different notation) that if the dimension of M is at least 11, to each choice of (a global) g, there exists a unique complementary (vertical) distribution V T M on M,
T M = H V .
If we pick any local pqc structure ( η s , I s , g ) for H, then V is the real span of local vector fields ξ 1 , ξ 2 and ξ 3 on M, called Reeb vector fields, which are defined by the following equations:
( i ) η s ( ξ t ) = ϵ s , if s = t 0 , otherwise , ϵ 1 = ϵ 2 = ϵ 3 = 1 ; ( i i ) d η s ( ξ t , X ) + d η t ( ξ s , X ) = 0 , X H , s , t = 1 , 2 , 3 .
Remark 1. 
In (the lowest) dimension 7, the existence of Reeb vector fields is an additional condition on the structure, which we shall assume is always satisfied.
At each p M , the vector space H p is isomorphic as a S p ( n , B ) S p ( 1 , B ) module to B n (with the action (5)), and the set of all isomorphisms from H p to B n constitutes a fiber over p of a certain principle bundle P ( M ) M with a structure group S p ( n , B ) S p ( 1 , B ) . The Reeb vector fields (10) allow us to extend the action of S p ( n , B ) S p ( 1 , B ) on H p to an action on the whole tangent space at p, T p M = H p V p by declaring that S p ( n , B ) S p ( 1 , B ) acts on the Reeb vector fields ξ s in the same way as it acts on the endomorphisms I s E n d ( H p ) . It is easily verified (using Lemma 2) that this action remains unchanged if we replace the initial pqc structure ( η s , I s , g ) with another (of course, the Rieb vector fields must undergo a respective transformation as well) as long as the g-entry remains the same; that is, the choice of g allows us to consider T p M as a S p ( n , B ) S p ( 1 , B ) module isomorphic to B n I m ( B ) , and the set of all isomorphisms is a principle fiber bundle P ( M ) M with a structure group S p ( n , B ) S p ( 1 , B ) .
The canonical connection ∇ is a principle S p ( n , B ) S p ( 1 , B ) -connection on P , whose torsion tensor
T ( A , B ) = A B B A [ A , B ] , A , B T M ,
can be described as follows.
We define three (local) two-forms, ω 1 , ω 2 , and ω 3 , on M by setting
ω s ( A , B ) = g I s ( A H ) , B H , s = 1 , 2 , 3 ,
where by subscript H we mean projection onto H w.r.t. the decomposition (9). There exists a (unique) triple ( S c a l , τ , μ ) , where S c a l is a (global) function on M; τ and μ are globally defined as traceless symmetric sections of the endomorphism bundle E n d ( H ) M , satisfying τ = τ [ 1 ] , μ = μ [ 3 ] (cf. Section 2.4), so that the following is true:
( i ) T ( X , Y ) = 2 ω 1 ( X , Y ) ξ 1 2 ω 2 ( X , Y ) ξ 2 + 2 ω 3 ( X , Y ) ξ 3 , ( i i ) T ( ξ s , X ) = 1 4 I s τ τ I s + I s μ X , ( i i i ) T ( ξ s , ξ t ) = S c a l 8 n ( n + 1 ) ξ s × ξ t ξ s , ξ t H c f , ( 15 )
for any X , Y H , s , t = 1 , 2 , 3 . Notice also that the vertical distribution V has an induced inner product , of signature ( , , + ) , so that
ξ s , ξ t = ϵ s , if s = t 0 , otherwise .
On V, we also have a natural orientation and a cross product “×” (cf. (3)):
ξ 1 × ξ 2 = ξ 3 , ξ 2 × ξ 3 = ξ 1 , ξ 3 × ξ 1 = ξ 2 .
Clearly, the two vector bundles V M and π : Q ( M ) M over M (cf. (7)) are isomorphic, and the cross product on V p corresponds to the half-commutator on Q p :
1 2 I , J = 1 2 I J J I , I , J Q p .
Since both , and “×” are ∇-parallel, locally, on the same domain where the considered local pqc-structure ( η s , I s , g ) is defined, we can find certain one-forms, i.e., α 1 , α 2 and α 3 (called connection 1-forms), so that
A I s = 1 2 t α t ( A ) I t , I s ,
or equivalently,
A ξ s = t α t ( A ) ξ t × ξ s ,
for all A T p M and s = 1 , 2 , 3 . As shown in [4], the connection one-forms are completely determined by the exterior derivatives of the three one-forms η s and the function S c a l ,
α i ( X ) = d η k ( ξ j , X ) = d η j ( ξ k , X ) , α i ( ξ s ) = d η s ( ξ j , ξ k ) δ i s ( S c a l 16 n ( n + 2 ) , + 1 2 d η 1 ( ξ 2 , ξ 3 ) + d η 2 ( ξ 3 , ξ 1 ) + d η 3 ( ξ 1 , ξ 2 ) ) ,
for all X H and s = 1 , 2 , 3 , where δ i s is the Kronecker delta, and ( i j k ) is any positive permutation of 1 , 2 , 3 .

2.6. Curvature

It turns out that not only the torsion (cf. (13)) but also many of the contractions of the curvature tensor,
R ( A , B ) = A , B [ A , B ] , A , B T M ,
are completely determined by the triple ( S c a l , τ , μ ) . Consider a local frame e a H , 1 a 4 n for H, and let e a * H be its dual; that is, the frame defined by the following equations:
g ( e a , e b * ) = 1 , if a = b 0 , otherwise , a , b = 1 , , 4 n .
The Ricci curvature, R i c , is defined by
R i c ( A , B ) = a g ( R ( e a , A ) B , e a * ) , A , B T M .
According to [4], we have
R i c ( X , Y ) = g S c a l 4 n X + ( 2 n + 2 ) τ ( X ) + ( 4 n + 10 ) μ ( X ) , Y ,
for all X , Y H . In particular, S c a l = a R i c ( e a , e a * ) , i.e., S c a l is indeed the scalar curvature of ∇.
Since, by design, ∇ is a principle S p ( n , B ) S p ( 1 , B ) connection, its curvature splits into a sum of two components, s p ( n , B ) s p ( 1 , B ) . We shall use the Ricci two-forms ρ s to represent the s p ( 1 , B ) component of the curvature:
R ( A , B ) , I s = t = 1 3 ρ t ( A , B ) I t , I s , A , B T M , s = 1 , 2 , 3 ,
or equivalently,
ρ s ( A , B ) = ϵ s 4 n a g R ( A , B ) e a , I s e a * .
Using the result in [4],
ρ s ( X , Y ) = ϵ s g 1 2 τ I s + I s τ X + 2 μ + S c a l 16 n ( n + 2 ) I s X , Y , ρ i ( X , ξ i ) = d S c a l ( X ) 32 n ( n + 2 ) + 1 2 g I i ξ j , ξ k H + I j ξ k , ξ i H + I k ξ i , ξ j H , X , ρ i ( X , ξ s ) = g I s ξ j , ξ k H , X , i s ,
for all X , Y H and s = 1 , 2 , 3 , where ( i j k ) is any positive permutation of 1,2,3 ( d S c a l is the differential of S c a l ). For the values of the three Ricci two-forms on a pair of vertical vector fields, we have the identity
ρ i ( ξ i , ξ j ) + ρ k ( ξ k , ξ j ) = d S c a l ( ξ j ) 16 n ( n + 2 ) .

3. Twistor and Reflector Spaces

The twistor space Z and the reflector space R of a pqc manifold ( M , H ) are defined as subbundles of the canonical vector bundle π : Q ( M ) M (cf. (7)). The corresponding fibers over a point p M are
Z p = I Q p ( M ) : I 2 = id and R p = I Q p ( M ) : I 2 = id .
The purpose of this section is to prove the two following propositions.
Proposition 1. 
On the twistor space Z , there exists a natural co-dimension one distribution K T Z and a smooth field J of endomorphisms of K , which satisfies J 2 = i d (such a pair ( K , J ) is called an almost CR structure).
Furthermore, if η is any local one-form on Z with K = ker ( η ) , then at each I Z , d η ( J . , . ) there is a non-degenerate symmetric two-tensor on K I of signature ( 2 n + 2 , 2 n + 2 ) , dim ( M ) = 4 n + 3 ; that is, the Levi form of the almost CR structure on Z is of signature ( 2 n + 2 , 2 n + 2 ) .
Proposition 2. 
On the reflector space R , there exists a natural co-dimension one distribution K T R and a smooth field J of endomorphisms of K , which satisfies J 2 = i d (such a pair ( K , J ) is called an almost para-CR structure).
Furthermore, if η is any local one-form on R with K = ker ( η ) , then at each I R , d η ( J . , . ) is a non-degenerate symmetric two-tensor on K I of signature ( 2 n + 2 , 2 n + 2 ) ; that is, the Levi form of the almost para-CR structure is of signature ( 2 n + 2 , 2 n + 2 ) .
Later in this paper (Section 4), we will show that both the almost CR structure on Z and the almost para-CR structure on R are in fact integrable.

3.1. The Induced Structure on Q

To begin with, let us fix an arbitrary non-vanishing section g of the line bundle G ( M ) M (cf. Section 2.3) and consider the corresponding canonical connection ∇ on T M . We shall use ∇ to induce a certain structure on the tangent space of the vector bundle Q = Q ( M ) . Indeed, since ∇ preserves the vector bundle Q E n d ( T M ) , it defines a horizontal distribution D T Q so that the horizontal lift A h (w.r.t. ∇) of any vector field A on M is a vector field on Q tangent to D . On the other hand, there is a distribution F = ker ( π * ) T Q that consists of all vectors that are tangent to the fibers of the bundle π : Q M . We have the following direct sum decomposition:
T Q = D F .
The differential π * of the projection map π : Q M at any I Q is an isomorphism between D I and T p M , where p = π ( I ) . There is also a natural isomorphism F I Q p that identifies the tangent vector to a curve t I ( t ) Q p at I ( 0 ) = I (that is, any element of F I ) with the respective derivative d I ( t ) d t | t = 0 (which is as an element of the fiber Q p ).
Let us consider a (small enough) domain U of local coordinates u α , 1 α 4 n + 3 on M. For each I π 1 ( U ) Q , we know that I = x 1 I 1 + x 2 I 2 + x 3 I 3 ; thus, we may consider the functions
u α π , x 1 , x 2 , x 3 , 1 α 4 n + 3 ,
as local coordinates on Q (we shall abbreviate u α π to u α ). In this coordinate chart, the isomorphism between F I and Q p identifies x s with I s for s = 1 , 2 , 3 .
Lemma 3. 
Within the coordinate chart (23), the horizontal lift A h of a vector field
A = a = 1 4 n + 3 A s u a
on M, at I = s x s I s Q , is given by
A I h = α = 1 4 n + 3 A α u α s , t = 1 3 x s A I s , ϵ t I t x t = α = 1 4 n + 3 A α u α + ( i j k ) ϵ i x j α k ( A ) x k α j ( A ) x i ,
where A I h denotes the value of A h at I, and α s are the connection one-forms of ∇ (cf. (16)).
Proof. 
Consider a curve t u α ( t ) , x s ( t ) within the coordinate chart (23), passing through a fixed I Q at a time t = 0 . Suppose that the tangent vector to this curve at t = 0 is A I h . Then,
0 = A s x s ( t ) I s = s x ˙ s ( 0 ) I s + x s ( 0 ) A I s ;
therefore, since x s ( 0 ) I s = I ,
x ˙ s ( 0 ) = t x t A I t , ϵ s I s ;
that is, for the horizontal lift A I h we have
A I h = α = 1 4 n + 3 A α u α + s = 1 3 x ˙ s ( 0 ) x s ,
where x ˙ s ( 0 ) are given by (25). Applying (16) to the latter yields the result. □
Lemma 4. 
For any two vector fields A and B on M, within a coordinate chart like (23), the commutator of their respective horizontal lifts A h and B h at any I = s x s I s Q is given by
A h , B h I = A , B I h + ( i j k ) 2 ϵ i x j ρ k ( A , B ) x k ρ j ( A , B ) x i ,
where ρ s are the corresponding Ricci two-forms (cf. (20)).
Proof. 
Using (24), we calculate
A h , B h I = α = 1 4 n + 3 A , B α u α s , t = 1 3 x s A B I s B A I s , ϵ t I t x t = A , B I h s , t = 1 3 x s R ( A , B ) , I s , ϵ t I t x t .
The result follows from (20). □
Next, we consider two naturally defined (global) vector fields, χ and N , on Q . At any I = s x s I s Q , we set, with respect to the coordinate chart (23),
χ = s x s ξ s h and N = s x s x s .
Clearly, N is a section of the vertical distribution F T Q . On the other hand, the splitting of T M = H V (cf. (9)) defines the splitting of the horizontal distribution, D = H V , and the vector field χ is tangent everywhere to V .
Suppose that I Q , considered as an endomorphism of the vector space H p T p M , does not square to 0, I 2 0 . Letting W I be the orthogonal complement of N in F I , and U I the orthogonal complement of χ in V I (the orthogonality is with respect to (14) and (8)), we obtain the splitting
T I Q = H I U I R · χ I V I D I W I R · N I F I .
We now consider a canonical one-form η on Q , defined at any I = x s I s Q , by
η = s x s π * ( η s ) ,
where π * ( η s ) is the pullback of η s via π : Q M . In order to calculate the exterior derivative of η , we introduce three local one-forms ϕ 1 , ϕ 2 and ϕ 3 on Q using the following formula
ϕ i = ϵ i d x i x j π * ( α k ) x k π * ( α j )
for any positive permutation ( i j k ) of 1 , 2 , 3 . Clearly, the forms ϕ s are only defined within the coordinate chart (23). According to Lemma 3, each ϕ s vanishes on the horizontal distribution D , and we have
ϕ s x t = ϵ s , if s = t 0 , otherwise .
For any A T I Q , we have
A = π * A H h + s ϵ s η s ( A ) ξ s h + ϕ s ( A ) x s .
By subscript H we mean projection onto H w.r.t. the decomposition (9).
Lemma 5. 
The exterior derivative of the canonical one-form η on Q is given (within the coordinate chart (23)) by
d η = ( i j k ) 2 x i π * ( ω i ) + ϵ i ϕ i π * ( η i ) S c a l 8 n ( n + 2 ) ϵ i x i π * ( η j η k ) .
The two-forms ω s are as in (12); for the wedge product, we use the formula ϕ i π * ( η i ) ( A , B ) = ϕ i ( A ) η i ( π * B ) ϕ i ( B ) η i ( π * A ) .
Proof. 
Differentiating (28) yields
d η = s d x s π * ( η s ) + x s π * ( d η s ) .
We calculate:
d η ( X h , X ˜ h ) = s x s d η s ( X , X ˜ ) = 2 s x s g ( I s X , X ˜ ) ;
d η X h , ξ i h = ϵ i d x i ( X h ) + s x s d η s ( X , ξ i ) = c f . ( 24 ) x j α k ( X ) x k α j ( X ) + s x s d η s ( X , ξ i ) = c f . ( 18 ) 0 ;
d η ξ i h , ξ j h = ϵ j d x j ( ξ i h ) ϵ i d x i ( ξ j h ) + s x s d η s ( ξ i , ξ j ) = c f . ( 24 ) x k α i ( ξ i ) x i α k ( ξ i ) x j α k ( ξ j ) + x k α j ( ξ j ) + s x s d η s ( ξ i , ξ j ) = c f . ( 18 ) S c a l 8 n ( n + 2 ) x k ;
d η ξ s h , x t = η t ( ξ s ) .
As a consequence of the previous lemma, we obtain the following.
Corollary 1. 
At any I = s x s I s Q , the canonical one-form η and the vector field χ (cf. (26)) satisfy
η ( χ ) = s ϵ s x s 2 a n d χ d η = s ϵ s x s d x s ,
where χ d η ( A ) = d η ( χ , A ) .
Let Q o Q be the open subset consisting of all I Q with I 2 0 . Clearly, the twistor and the reflector spaces Z and R are submanifolds in Q o . On the manifold Q o , we have the distribution
K = H U W T Q o .
Using local coordinates (23) and one-forms ϕ s (cf. (30)), K can be described with the equations
s x s π * ( η s ) = 0 and s x s ϕ s = 0 .
We introduce a natural field J of endomorphisms of the distribution K that satisfies J 2 = I , I id by setting
J ( X + U + W ) = I π * X I h + χ I × U + N I × W ,
where X H I , U U I and W W I . For any A K I within the coordinate chart (23), we have (cf. (31))
J ( A ) = s x s I s π * ( A ) H h + ( i j k ) ϵ j x j η k ( π * A ) ϵ k x k η j ( π * A ) ξ i h + ( i j k ) ϵ j x j ϕ k ( A ) ϵ k x k ϕ j ( A ) x i .
Let us denote by G the bilinear form
G ( A , B ) = 1 2 I , I d η ( J A , B ) , A , B K I .
Since J 2 = I , I id , we have
d η ( A , B ) = 2 G ( J A , B ) , A , B K I .
Lemma 6. 
At any I Q o , G is a symmetric two-form on K I T I Q o (cf. (32)) of signature ( 2 n + 2 , 2 n + 2 ) , which satisfies the relation
G ( J A , B ) = G ( A , J B )
for A , B K I . Explicitly, within the coordinate chart (23), we have that
G ( A , B ) = g π * A H , π * B H , S c a l 16 n ( n + 2 ) s ϵ s η s ( π * A ) η s ( π * B ) 1 2 I , I ( i j k ) ϵ i x i ( ϕ j ( A ) η k ( π * B ) + η k ( π * A ) ϕ j ( B ) , ϕ k ( A ) η j ( π * B ) η j ( π * A ) ϕ k ( B ) ) .
Proof. 
Formula (35) is a straightforward application of Lemma 5. To calculate the signature of G on K I , we first observe that the two subspaces H I and U I + W I are G-orthogonal, and the restriction of G to H I has the same signature as g. Therefore, we only need to show that the restriction of G to U I + W I is of signature ( 2 , 2 ) .
For any fixed I Q o , we can pick a local pqc structure ( η s , I s , g ) in such a way so that either I = λ I 3 or I = λ I 1 , λ R . In the first case, the restriction of G to U I + W I is given, w.r.t. the frame { ξ 1 h , ξ 2 h } of U I and the frame x 1 , x 2 of W I , by the matrix
h 0 0 l 0 h l 0 0 l 0 0 l 0 0 0 ,
where h = S c a l 16 n ( n + 2 ) and l = 1 2 λ . This matrix has two eigenvalues, each with multiplications of two: 1 2 h ± h 2 + 4 l 2 . Therefore, the restriction of G to U I + W I has signature (2,2).
Similarly, in the second case (when I = λ I 1 ), the restriction of G to U I + W I is given, w.r.t. the frame { ξ 2 h , ξ 3 h } of U I and the frame x 2 , x 3 of W I , by
h 0 0 l 0 h l 0 0 l 0 0 l 0 0 0 .
Given matrix has two positive and two negative eigenvalues:
1 2 ± h + h 2 + 4 l 2 and 1 2 ± h h 2 + 4 l 2 ;
thus, the signature is again ( 2 , 2 ) . □

3.2. Invariance

For the definition of the distribution K T Q o and the respective field J (cf. (33)), we have used, as an essential tool, the concept of a horizontal lift of a vector fields w.r.t. ∇. Since ∇ is the canonical connection determined by a choice of a section g of the canonical line bundle G M (cf. Section 2.3), the whole construction depends on that choice as well. Our purpose here is to show that this dependence is only formal and, in fact, if we replace g with
g ¯ = 1 2 f g ,
where f is any smooth and non-vanishing function on M, then both K and J remain unchanged.
If A is a vector field on M with a horizontal lift A h to Q w.r.t. g and ∇, we shall denote A h ¯ as the respective horizontal lift of A to Q w.r.t. g ¯ and its canonical connection ¯ . Clearly, if ( η s , I s , g ) is any local pqc structure for H, then so is ( η ¯ s , I s , g ¯ ) , where η ¯ s = 1 2 f η s . More generally, we shall use the bar on objects related to the pqc structure ( η s , I s , g ) to indicate the respective objects related to ( η ¯ s , I s , g ¯ ) , e.g., ξ ¯ s will denote the Reeb vector fields (cf. (10)), defined by
( i ) η ¯ s ( ξ ¯ t ) = ϵ s , if s = t 0 , otherwise ( i i ) d η ¯ s ( ξ ¯ t , X ) + d η ¯ t ( ξ ¯ s , X ) = 0 , X H .
One can easily derive from the above that
ξ ¯ s = 2 f ξ s + I s f ,
where f is the horizontal gradient of the function f; that is, the unique section of the distribution H, which satisfies g ( f , X ) = d f ( X ) for all X H . According to [18] (here we are using slightly different sign conventions), we have the following formulas concerning the connection one-forms α ¯ s (cf. (16)) of ¯ :
α ¯ s ( X ) = α s ( X ) + ϵ s f d f ( I s X ) , s = 1 , 2 , 3 , X H , α ¯ i ( ξ ¯ i ) = 2 f α i ( ξ i ) Δ f 2 n + g ( f , f ) n f , α ¯ j ( ξ ¯ i ) = 2 f α j ( ξ i ) + α j ( I i f ) 2 ϵ i d f ( ξ k ) , α ¯ k ( ξ ¯ i ) = 2 f α k ( ξ i ) + α k ( I i f ) + 2 ϵ i d f ( ξ j ) ,
where ( i j k ) is any positive permutation of 1 , 2 , 3 , and Δ f = a d f ( e a , e a * ) (cf. (19)).
Lemma 7. 
Within the coordinate chart (23) on Q , we have the following formulas for the horizontal lift of a vector fields from M to Q :
X h ¯ = X h + N × t ϵ t d f ( I t X ) x t , ξ ¯ s h ¯ = 2 f ξ s h + I s f h + 2 d f π * χ x s 2 ϵ s x s t ϵ t d f ( ξ t ) x t + Δ f 2 n + g ( f , f ) n f N × x s ,
where X is any section of H, and s = 1 , 2 , 3 .
Proof. 
The proof is obtained by a straightforward calculation using (38) and (24). □
Let us observe that the vector field N on Q (cf. (26)) does not depend on the choice of g and ∇, whereas the field χ is changed as follows:
χ ¯ = s x s ξ ¯ s h ¯ = χ + 2 d f ( π * χ ) N + J f h 2 I , I t ϵ t d f ( ξ t ) x t .
Proposition 3. 
The distribution K on Q o (defined by (32)) and the field J of endomorphisms of K (defined by (33)) do not depend on the choice of g and.
Proof. 
Let us begin by constructing a distribution K ¯ on Q o as in (32) using (37) in place of g and ¯ in place of ∇. Within the coordinate chart (23), we have an orthogonal decomposition
span ξ ¯ 1 h ¯ , ξ ¯ 2 h ¯ , ξ ¯ 3 h ¯ = U ¯ R . χ ¯
that defines a distribution U ¯ . Then,
K ¯ = H ¯ U ¯ W ,
where the distribution H ¯ is defined by the requirement that its sections are precisely the horizontal lifts, w.r.t. the connection ¯ , of vector fields on M tangent to the distribution H T M , and W is as in (27).
If A = s a s ξ ¯ s h ¯ is any element of U ¯ I , I = s x s I s , then s ϵ s a s x s = 0 . Using (39), we calculate
A = s a s ξ ¯ s h ¯ = s ( 2 f a s ξ s h + a s I s f h + 2 d f π * χ a s x s + Δ f 2 n + g ( f , f ) n f N × a s x s ) ,
which yields that A H U W and U ¯ K . Similarly, if X is any section of H, then, by (39), X h ¯ H W ; thus, H ¯ K . Therefore, we get that K ¯ = K , and K does not depend on the choice of g and ∇. The invariance of J is shown similarly. □

3.3. Proof of Propositions 1 and 2

The restriction of the onr-form η to the twistor space Z Q o and the reflector space R Q o , respectively, satisfies
η d η 2 n 0 ;
therefore, it defines a contact structure on both Z and R . The tangent bundles T Z and T R , considered as subbundles in T Q o , are described by the equation
s x s ϕ s = 0 cf . ( 30 ) .
The vector field χ (cf. (26)) is tangent to Z ( R ), and if we restrict to the tangent space of Z (resp. R ), we obtain that (cf. Corollary 1)
η ( χ ) = I , I and χ d η = 0 ;
that is, χ is a Reeb vector field for the contact form η on Z ( R ).
At each I Z ( I R ), the kernel of η (cf. (32)) is given by the subspace K I T I Z ( K I T I R ) and the endomorphism J (cf. (34)) of K I satisfies J 2 = i d ( J 2 = i d ). The pair ( K , J ) defines an almost CR structure on the twistor space Z and an almost para-CR structure on the reflector space R . The signature of d η ( J . , . ) is given by Lemma 6. By Proposition 3, the pair ( K , J ) is uniquely determined by the pqc distribution H T M , which does not depend on the particular choice of the local pqc structure ( η s , I s , g ) for H.

4. Integrability

In this section, we consider the integrability question for the previously introduced (Section 3) almost CR structure ( K , J ) on the twistor space Z , and for the respective almost para-CR structure on the reflector space R .
Observe that using Lemma 6, if A and B are any two sections of K , then
J A , B + A , J B
is also a section of K . Therefore, the integrability of the almost CR structure ( K , J ) on Z is equivalent to the equation N Z ( A , B ) = 0 , where N Z is the so called Nijenhuis tensor, defined by
N Z ( A , B ) = A , B + J A , J B J J A , B + A , J B ,
for any two vector fields A and B on Z that are tangent to the distribution K T Z .
The complexified distribution K c = K R C on Z splits as
K c = K 1 K 1 ,
where K 1 and K 1 are the eigenspaces of J with eigenvalues 1 and 1 . The vanishing of the Nijenhuis tensor N Z is equivalent to the formal integrability of the complex distributions K 1 and K 1 ; that is, to any of the following two conditions
K 1 , K 1 K 1 and K 1 , K 1 K 1 .
Similarly, the almost para-CR structure ( K , J ) on the reflector space R is integrable if N R ( A , B ) = 0 for any two sections A and B of the distribution K T R , where
N R ( A , B ) = A , B + J A , J B J J A , B + A , J B .
Here, the complexified distribution splits as K c = K + 1 K 1 , where K + 1 and K 1 are the ± 1 eigenspaces of J. The vanishing of N R is equivalent to the formal integrability of K + 1 and K 1 , i.e., to the following conditions:
K + 1 , K + 1 K + 1 and K 1 , K 1 K 1 .
The following result is obtained as a straightforward application of Proposition 5 below.
Proposition 4. 
The almost CR structure ( K , J ) on the twistor space Z and the respective almost para-CR structure on the reflector space R are integrable.

Integrability on the Ambient Space Q o

The distribution K (cf. (32)) can be considered as a vector bundle over the manifold Q o . We introduce a Nijenhuis-like tensor field N defined for any two vector fields A and B on Q o that are tangent to the distribution K by the following formula:
N ( A , B ) = I , I A , B + J A , J B J J A , B + A , J B ,
where N is indeed a tensor field, meaning that the value of N ( A , B ) at any given I Q o depends only on the values of A and B at I, due to the obvious property N ( f A , h B ) = f h N ( A , B ) for any functions f and h on Q o . Notice that the expression on the right hand side in (44) also makes sense, since, by Lemma 5, the vector field J A , B + A , J B is tangent to the distribution K , and the action of J is well defined there (by definition J is a field of endomorphisms of K , cf. (33)). Furthermore, applying Lemma 5 one more time, we observe that N ( A , B ) is always a section of K ; thus, N : K × K K . Clearly, if restricted to the twistor space Z Q o , N coincides with the Nijenhuis tensor N Z (cf. (42)), and, similarly, on R Q o , it coincides with N R (cf. (43)).
Proposition 5. 
On Q o , we have that
N ( A , B ) = 0 ,
for any two sections A and B of K .
Proof. 
To begin with, we fix an arbitrary non-vanishing section g of the line bundle G ( M ) M (cf. Section 2.3) and consider the corresponding canonical connection ∇ on T M . Using Lemma 2, for any fixed I Q o , we can pick a local pqc structure ( η s , I s , g ) in such a way so that either I = λ I 1 or I = λ I 3 , λ R . Let us assume that I = λ I 1 (in the other case the proof is similar). Using the corresponding Reeb vector fields ξ s , we construct a coordinate chart as in (23) around the fixed point I = λ I 1 Q o .
Following the structure (32) of K and observing that N ( A , B ) = N ( B , A ) , we see that there are six different cases to consider in the proof: (I) A , B H ; (II) A H , B U ; (III) A H , B W ; (IV) A , B U ; (V) A U , B W ; (VI) A , B W .
Case (I) A , B H :
Without loss of generality, in this case, we may assume that A = X h and B = Y h for some vector fields X and Y on M that are tangent to the distribution H. We calculate the following:
N ( X h , Y h ) | I = λ I 1 = = λ 2 X h , Y h + s , t x s ( I s X ) h , x t ( I t Y ) h | I = λ I 1 λ I 1 s x s ( I s X ) h , Y h + X h , x s ( I s Y ) h | I = λ I 1 ,
= λ 2 ( X , Y + [ I 1 X , I 1 Y ] I 1 I 1 X , Y + X , I 1 Y ) | I = λ I 1 h + s ( d x s J X h ( I s Y ) h d x s J Y h ( I s X ) h d x s ( X h ) J ( I s Y ) h + d x s ( Y h ) J ( I s X ) h ) | I = λ I 1 + 2 λ 3 ρ 3 ( X , Y ) + ρ 3 ( I 1 X , I 1 Y ) ρ 2 ( I 1 X , Y ) ρ 2 ( X , I 1 Y ) x 2 + 2 λ 3 ρ 2 ( X , Y ) + ρ 2 ( I 1 X , I 1 Y ) ρ 3 ( I 1 X , Y ) ρ 3 ( X , I 1 Y ) x 2 .
We observe that the last two lines in the above expression vanish as a consequence of (21). We may represent the remaining part of the expression as
λ 2 Σ 1 | I = λ I 1 h + λ 2 Σ 2 | I = λ I 1 h ,
where
Σ 1 = X , Y + [ I 1 X , I 1 Y ] I 1 I 1 X , Y + X , I 1 Y , Σ 2 = 1 λ s ( d x s ( I 1 X ) h ( I s Y ) d x s ( I 1 Y ) h ( I s X ) d x s ( X h ) I 1 ( I s Y ) + d x s ( Y h ) I 1 ( I s X ) ) .
Using the canonical connection ∇ on M and its torsion T (cf. (11)), we calculate
Σ 1 = X Y Y X T ( X , Y ) + I 1 X ( I 1 Y ) I 1 Y ( I 1 X ) T ( I 1 X , I 1 Y ) I 1 ( I 1 X Y Y ( I 1 X ) T ( I 1 X , Y ) + X ( I 1 Y ) I 1 Y X T ( X , I 1 Y ) ,
= I 1 X I 1 Y + I 1 Y I 1 X + I 1 X I 1 Y I 1 Y I 1 X T ( X , Y ) T ( I 1 X , I 1 Y ) + I 1 T ( I 1 X , Y ) + T ( X , I 1 Y ) ,
= α 2 ( X ) + α 3 ( I 1 X ) I 2 Y α 2 ( Y ) + α 3 ( I 1 Y ) I 2 X + α 3 ( X ) α 2 ( I 1 X ) I 3 Y α 3 ( Y ) α 2 ( I 1 Y ) I 3 X , ,
where the last equality follows from (16) and (13).
Applying (24) to the expression Σ 2 gives
Σ 2 = α 2 ( X ) + α 3 ( I 1 X ) I 2 Y + α 2 ( Y ) + α 3 ( I 1 Y ) I 2 X α 3 ( X ) α 2 ( I 1 X ) I 3 Y + α 3 ( Y ) α 2 ( I 1 Y ) I 3 X .
Therefore, using (45), we get N ( X h , Y h ) = 0 .
Case (II) A H , B U :
Here, we may assume that A = X h and B = μ 2 ξ 2 h + μ 3 ξ 3 h , where μ 2 and μ 3 are any real numbers, and X is a section of H T M . We obtain that
N ( A , B ) | I = λ I 1 = μ 2 N ( X h , ξ 2 h ) | I = λ I 1 + μ 3 N ( X h , ξ 3 h ) | I = λ I 1 .
In order to calculate the quantity N ( X h , ξ 2 h ) | I = λ I 1 , consider the vector field
ξ 2 h + x 2 I , I χ .
Clearly, (47) is a vector filed tangent to the distribution U K T Q (cf. (32)) that is defined in a neighborhood of the fixed point I = λ I 1 , so that its value at this point coincides with the value of ξ 2 h . Therefore, we obtain that (cf. (44) and (33))
N ( X h , ξ 2 h ) | I = λ I 1 = N X h , ξ 2 h + x 2 I , I χ | I = λ I 1 = λ 2 X h , ξ 2 h + x 2 I , I χ | I = λ I 1 + s , t x s ( I s X ) h , x t ξ t × ξ 2 h | I = λ I 1 λ s J x s ( I s X ) h , ξ 2 h + x 2 I , I χ + X h , x s ( ξ s × ξ 2 ) h | I = λ I 1 ,
= λ 2 X , ξ 2 h + λ 2 I 1 X , ξ 3 h λ J I 1 X , ξ 2 + X , ξ 3 h λ 2 ( α 2 ( ξ 2 ) + α 3 ( ξ 3 ) I 2 X + α 2 ( ξ 3 ) α 3 ( ξ 2 ) I 3 X + α 3 ( X ) α 2 ( I 1 X ) ξ 1 ) h + 2 λ 3 ρ 3 ( X , ξ 2 ) + ρ 3 ( I 1 X , ξ 3 ) ρ 2 ( I 1 X , ξ 2 ) ρ 2 ( X , ξ 3 ) x 2 + 2 λ 3 ρ 2 ( X , ξ 2 ) + ρ 2 ( I 1 X , ξ 3 ) ρ 3 ( I 1 X , ξ 2 ) ρ 3 ( X , ξ 3 ) x 3 .
By the properties (21) of ρ ( X , ξ s ) , the last four lines in the above expression vanish. Using the canonical connection ∇ on M and its torsion T (cf. (11)), we calculate that
λ 2 [ X , ξ 2 ] h + λ 2 I 1 X , ξ 3 h λ J I 1 X , ξ 2 + X , ξ 3 h = λ 2 ( I 1 ξ 2 I 1 X ξ 3 I 1 X T ( X , ξ 2 ) + I 1 T ( I 1 X , ξ 2 ) T ( I 1 X , ξ 3 ) + I 1 T ( X , ξ 3 ) + X ξ 2 + I 1 X ξ 3 ξ 1 × ( I 1 X ) ξ 2 + X ξ 3 ) h ,
= λ 2 ( α 2 ( ξ 2 ) + α 3 ( ξ 3 ) I 2 X + α 2 ( ξ 3 ) α 3 ( ξ 2 ) I 3 X + α 3 ( X ) α 2 ( I 1 X ) ξ 1 ) h , ,
where for the last identity, we have use Formulas (16), (17), and (13). Substituting (49) into (48), we get
N ( X h , ξ 2 h ) | I = λ I 1 = 0 .
Similarly, one can also show that
N ( X h , ξ 3 h ) | I = λ I 1 = 0 ;
therefore, we obtain that, in this case, N ( A , B ) = 0 .
Case (III) A H , B W :
We may assume here that A = X h and B = μ 2 x 2 + μ 3 x 3 , where μ 2 and μ 3 are any real numbers, and X is a section of H T M . Then,
N ( A , B ) | I = λ I 1 = μ 2 N X h , x 2 | I = λ I 1 + μ 3 N X h , x 3 | I = λ I 1 .
In order to show that N X h , x 3 | I = λ I 1 vanishes (the vanishing of the other summand is shown similarly), we consider the vector field
x 3 x 3 I , I N .
Clearly, this is a vector filed tangent to the distribution W K T Q (cf. (32)), which is defined in a neighborhood of the fixed point I = λ I 1 , so that its value at this point coincides with the value of x 3 . Therefore, using (44) and (33) we get
N ( X h , x 3 ) | I = λ I 1 = N X h , x 3 x 3 I , I N | I = λ I 1 = λ 2 X h , x 3 x 3 I , I N | I = λ I 1 + s , t x s ( I s X ) h , x t x t × x 2 | I = λ I 1 λ s J x s ( I s X ) h , x 3 x 3 I , I N + X h , x t x t × x 2 | I = λ I 1 ,
= λ 2 X h , x 3 + λ 2 ( I 1 X ) h , x 2 λ J ( I 1 X ) h , x 3 + X h , x 2 + λ 2 α 2 ( X ) α 3 ( I 1 X ) x 1 = 0 . .
Case (IV) A , B U :
It suffices to assume A = ξ 2 h , B = ξ 3 h . Using (44) and (24), we calculate
N ( ξ 2 h , ξ 3 h ) | I = λ I 1 = N ξ 2 h + x 2 I , I , ξ 3 h x 3 I , I | I = λ I 1 = λ 2 ξ 2 h + x 2 I , I , ξ 3 h x 3 I , I | I = λ I 1 + s , t x s ( ξ s × ξ 2 ) h , x t ( ξ t × ξ 3 ) h | I = λ I 1 λ s J ( x s ( ξ s × ξ 2 ) h , ξ 3 h x 3 I , I + ξ 2 h + x 2 I , I , x s ( ξ s × ξ 3 ) h ) | I = λ I 1 = 0 .
Case (V) A U , B W :
Here, we need to consider the following assumptions: A = ξ s h and B = x t for s , t = 2 , 3 . We shall consider only the case where s = 2 and t = 3 ; the remaining three possibilities are entirely analogous.
N ( ξ 2 h , x 3 ) | I = λ I 1 = N ξ 2 h + x 2 I , I , x 3 x 3 I , I N | I = λ I 1 = λ 2 ξ 2 h + x 2 I , I , x 3 x 3 I , I N | I = λ I 1 + s , t x s ( ξ s × ξ 2 ) h , x t x t × x 2 | I = λ I 1 λ s J ( x s ( ξ s × ξ 2 ) h , x 3 x 3 I , I N + ξ 2 h + x 2 I , I , x t x t × x 3 ) | I = λ I 1 = 0 .
Case (VI) A , B W :
It suffices to consider only the case A = x 2 , B = x 3 .
N ( x 2 , x 3 ) | I = λ I 1 = N x 2 + x 2 I , I N , x 3 x 3 I , I N | I = λ I 1 = λ 2 x 2 + x 2 I , I N , x 3 x 3 I , I N | I = λ I 1 + s , t x s x s × x 2 , x t x t × x 3 | I = λ I 1 λ s J ( x s x s × x 2 , x 3 x 3 I , I N + x 2 + x 2 I , I N , x s x s × x 3 ) | I = λ I 1 = 0 .

Author Contributions

Writing—original draft, S.I., I.M. and M.T. All authors have read and agreed to the published version ot he manuscript.

Funding

The research of S.I. is partially supported by Contract KP-06-H72-1/05.12.2023 with the National Science Fund of Bulgaria, by Contract 80-10-181/22.4.2024 with Sofia University “St.Kl.Ohridski”, and the National Science Fund of Bulgaria, National Scientific Program “VIHREN”, Project KP-06-DV-7. The research of I.M. is partially financed by the European Union–Next Generation EU through the National Recovery and Resilience Plan of the Republic of Bulgaria, project N: BG-RRP-2.004-0008-C01. The research of M. Tch. is partially supported by Contract KP-06-H72-1/05.12.2023 with the National Science Fund of Bulgaria and by Contract 80-10-181/22.4.2024 with the Sofia University “St.Kl.Ohridski”.

Data Availability Statement

Required data can be found in section “References” of our paper.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Dancer, A.; Jorgensen, H.; Swann, A. Metric geometries over the split quaternions. Rend. Sem. Mat. Univ. Politec. Torino 2005, 63, 119–139. [Google Scholar]
  2. Libermann, P. Sur le probleme d’equivalence de certains structures infinitesimales. Ann. Mat. Pura Appl. 1954, 36, 27–120. [Google Scholar] [CrossRef]
  3. Andrada, A.; Salamon, S. Complex product structures on Lie algebras. Forum Math. 2005, 17, 261–295. [Google Scholar] [CrossRef]
  4. Ivanov, S.; Tchomakova, M.; Zamkovoy, S. Geometry of paraquaternionic contact structures. arXiv 2024, arXiv:2404.16713. [Google Scholar]
  5. Alekseevsky, D.; Kamishima, Y. Quaternionic and para-quaternionic CR structure on (4n+3)-dimensional manifolds. Centr. Eur. J. Math. 2004, 2, 732–753. [Google Scholar] [CrossRef]
  6. Biquard, O. Métriques d’Einstein asymptotiquement symétriques. Astérisque 2000, 265. [Google Scholar] [CrossRef]
  7. Ivanov, S.; Minchev, I.; Vassilev, D. Quaternionic Contact Einstein Structures and the Quaternionic Contact Yamabe Problem. Mem. AMS 2014, 231, 1086. [Google Scholar]
  8. Ivanov, S.; Minchev, I.; Vassilev, D. Extremals for the Sobolev inequality on the seven dimensional quaternionic Heisenberg group and the quaternionic contact Yamabe problem. J. Eur. Math. Soc. 2010, 12, 1041–1067. [Google Scholar] [CrossRef]
  9. Ivanov, S.; Minchev, I.; Vassilev, D. Solution of the qc Yamabe equation on a 3-Sasakian manifold and the quaternionic Heisenberg group. Anal. PDE 2023, 16, 839–860. [Google Scholar] [CrossRef]
  10. Ivanov, S.; Petkov, A. The qc Yamabe problem on non-spherical quaternionic contact manifolds. J. Math. Pures Appl. 2018, 118, 44–81. [Google Scholar] [CrossRef]
  11. Ivanov, S.; Vassilev, D. Extremals for the Sobolev Inequality and the Quaternionic Contact Yamabe Problem; World Scientific Publishing Co. Pvt. Ltd.: Hackensack, NJ, USA, 2011. [Google Scholar]
  12. Davidov, J.; Ivanov, S.; Minchev, I. The twistor space of a quaternionic contact manifold. J. Differ. Geom. 2012, 63, 873–890. [Google Scholar] [CrossRef]
  13. Salamon, S. Quaternionic Kähler manifolds. Invent. Math. 1982, 67, 143–171. [Google Scholar] [CrossRef]
  14. Ivanov, S.; Minchev, I.; Zamkovoy, S. Twistor and reflector spaces of almost para-quaternionic manifolds. In Handbook of Pseudo-Riemannian Geometry and Supersymmetry; IRMA Lectures in Mathematics and Theoretical Physics; EMS Press: Berlin, Germany, 2010; Volume 16, pp. 477–496. [Google Scholar]
  15. Lempert, L. Spaces of Cauchy-Riemann Manifolds. Adv. Stud. Pure Math. Geom. Overdeterm. Syst. 1997, 25, 221–236. [Google Scholar]
  16. Alekseevsky, D.; Medori, C.; Tomassini, A. Maximally homogeneous para-CR manifolds. Ann. Glob. Anal Geom. 2006, 30, 1–27. [Google Scholar] [CrossRef]
  17. Hill, C.; Nurowski, P. Differential equations and para-CR structures. Boll. Dell’Unione Mat. Ital. 2010, 3, 25–91. [Google Scholar]
  18. Ivanov, S.; Tchomakova, M.; Zamkovoy, S. Conformal paraquaternionic contact curvature and the local flatness theorem. arXiv 2024, arXiv:2404.16703. [Google Scholar]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ivanov, S.; Minchev, I.; Tchomakova, M. Twistor and Reflector Spaces for Paraquaternionic Contact Manifolds. Mathematics 2024, 12, 3355. https://doi.org/10.3390/math12213355

AMA Style

Ivanov S, Minchev I, Tchomakova M. Twistor and Reflector Spaces for Paraquaternionic Contact Manifolds. Mathematics. 2024; 12(21):3355. https://doi.org/10.3390/math12213355

Chicago/Turabian Style

Ivanov, Stefan, Ivan Minchev, and Marina Tchomakova. 2024. "Twistor and Reflector Spaces for Paraquaternionic Contact Manifolds" Mathematics 12, no. 21: 3355. https://doi.org/10.3390/math12213355

APA Style

Ivanov, S., Minchev, I., & Tchomakova, M. (2024). Twistor and Reflector Spaces for Paraquaternionic Contact Manifolds. Mathematics, 12(21), 3355. https://doi.org/10.3390/math12213355

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop