Skip to main content
Log in

A canonical structure on the tangent bundle of a pseudo- or para-Kähler manifold

  • Published:
Monatshefte für Mathematik Aims and scope Submit manuscript

Abstract

It is a classical fact that the cotangent bundle \(T^* {\mathcal {M}}\) of a differentiable manifold \({\mathcal {M}}\) enjoys a canonical symplectic form \(\Omega ^*\). If \(({\mathcal {M}},\mathrm{J} ,g,\omega )\) is a pseudo-Kähler or para-Kähler \(2n\)-dimensional manifold, we prove that the tangent bundle \(T{\mathcal {M}}\) also enjoys a natural pseudo-Kähler or para-Kähler structure \(({\tilde{\hbox {J}}},\tilde{g},\Omega )\), where \(\Omega \) is the pull-back by \(g\) of \(\Omega ^*\) and \(\tilde{g}\) is a pseudo-Riemannian metric with neutral signature \((2n,2n)\). We investigate the curvature properties of the pair \(({\tilde{\hbox {J}}},\tilde{g})\) and prove that: \(\tilde{g}\) is scalar-flat, is not Einstein unless \(g\) is flat, has nonpositive (resp. nonnegative) Ricci curvature if and only if \(g\) has nonpositive (resp. nonnegative) Ricci curvature as well, and is locally conformally flat if and only if \(n=1\) and \(g\) has constant curvature, or \(n>2\) and \(g\) is flat. We also check that (i) the holomorphic sectional curvature of \(({\tilde{\hbox {J}}},\tilde{g})\) is not constant unless \(g\) is flat, and (ii) in \(n=1\) case, that \(\tilde{g}\) is never anti-self-dual, unless conformally flat.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. We might also define an “almost para-Kähler” structure on \(T{\mathcal {M}}\) by introducing the para-Sasaki metric

    $$\begin{aligned} G'( (X_h ,X_v),(Y_h ,Y_v)) := g(X_h,Y_h) - g(X_v,Y_v). \end{aligned}$$

    This metric has neutral signature \((m,m)\) (\(m\) being the dimension of \({\mathcal {M}}\)), is compatible with \(J'\) and verifies \(\Omega :=-G'(J'.,.)\).

  2. The terminology split-holomorphic is sometimes used.

  3. The Reader should be aware of the conflicting notation: the splitting of \(T T {\mathcal {M}}\simeq {\mathbb {R}} ^{4n}\) as \({\mathbb {R}}^{2n} \oplus {\mathbb {R}}^{2n}\) induced by the coordinate charts (e.g. \(\bar{X}\simeq ((x,\xi ),(X,\Xi ))\)) differs a priori from the connection-induced splitting \(\bar{X} \simeq (\Pi \bar{X}, \mathrm{K} \bar{X})\).

  4. Note that computations in [11] are done for the Sasaki metric, hence direct results do not apply.

  5. On the contrary, some authors seem to imply that scalar flatness is equivalent to anti-self-duality, see [7]). However this contradiction could possibly come from a different choice of orientation, which would exchange self-dual with anti-self-dual.

References

  1. Anciaux, H.: Minimal Submanifolds in Pseudo-Riemannian Geometry. World Scientific, Singapore (2010)

    Book  Google Scholar 

  2. Anciaux, H., Guilfoyle, B., Romon, P.: Minimal Lagrangian surfaces in the tangent bundle of a Riemannian surface. J. Geom. Phys. 61, 237–247 (2011)

    Article  MATH  MathSciNet  Google Scholar 

  3. Besse, A.L.: Einstein Manifolds. Classics in Mathematics. Springer, Berlin (1987)

  4. Boeckx, E., Vanhecke, L.: Geometry of Riemannian manifolds and their unit tangent sphere bundles. Publ. Math. Debr. 57(3–4), 509–533 (2000)

    MATH  MathSciNet  Google Scholar 

  5. Derdziński, A.: Self-dual Kähler manifolds and Einstein manifolds of dimension four. Compositio Math. 49(3), 405–433 (1983)

    MATH  MathSciNet  Google Scholar 

  6. Dombrowski, P.: On the geometry of the tangent bundle. J. Reine Angew. Math. 210, 73–88 (1962)

    MATH  MathSciNet  Google Scholar 

  7. Dunajski, M., West, S.: Anti-self-dual conformal structures in neutral signature. In: Alekseevsky, D., Baum, H. (eds.) Recent Developments in Pseudo-Riemannian Geometry. ESI-Series on Mathematics and Physics. European Mathematical Society, Zürich (2008)

  8. Gudmundsson, S., Kappos, E.: On the geometry of tangent bundles. Expositiones Math. 20(1), 1–41 (2002)

    Article  MATH  MathSciNet  Google Scholar 

  9. Guilfoyle, B., Klingenberg, W.: An indefinite Kähler metric on the space of oriented lines. J. Lond. Math. Soc. 72, 497–509 (2005)

    Article  MATH  MathSciNet  Google Scholar 

  10. Guilfoyle, B., Klingenberg, W.: On area-stationary surfaces in certain neutral Kaehler 4-manifolds. Beiträge Algebra Geom. 49, 481–490 (2008)

    MATH  MathSciNet  Google Scholar 

  11. Kowalski, O.: Curvature of the induced Riemannian metric on the tangent bundle. J. Reine Angew. Math. 250, 124–129 (1971)

    MATH  MathSciNet  Google Scholar 

  12. Lafontaine, J.: Some relevant Riemannian geometry. In: Lafontaine, J., Audin, M. (eds.) Holomorphic Curves in Symplectic Geometry. Birkhäuser, Basel (1994)

    Google Scholar 

  13. Lempert, L., Szöke, R.: The tangent bundle of an almost complex manifold. Can. Math. Bull. 44(1), 70–79 (2001)

    Article  MATH  Google Scholar 

  14. Musso, E., Tricerri, F.: Riemannian metrics on tangent bundles. Ann. Math. Pura Appl. 150, 1–20 (1988)

    Article  MATH  MathSciNet  Google Scholar 

  15. Yano, K., Ishihara, S.: Tangent and Cotangent Bundles: Differential Geometry. Marcel Dekker Inc, New York (1973)

    MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Henri Anciaux.

Additional information

Communicated by D. V. Alekseevsky.

H. Anciaux was supported by CNPq (PQ 306154/2011-0) and Fapesp (2011/21362-2).

Appendix: the Weyl tensor in the pseudo-Kähler or para-Kähler cases

Appendix: the Weyl tensor in the pseudo-Kähler or para-Kähler cases

The Riemann curvature tensor \({\hbox {Rm}}\) of a pseudo-Riemannian manifold \({\mathcal {N}}\) may be seen as a symmetric form \(\text {R}\) on bivectors of \(\Lambda ^2 T{\mathcal {N}}\) (see [3] for references). Splitting \(\text {R}\) along the eigenspaces \(\Lambda ^+ \oplus \Lambda ^-\) of the Hodge operator \(*\) on \(\Lambda ^2 T{\mathcal {N}}\), yields the following block decomposition

$$\begin{aligned} \text {R}= \left( \begin{array}{cc} \text {W}^+ + \frac{\text {Scal}}{12} I &{} \text {Z}\\ \text {Z}^* &{} \text {W}^- + \frac{\text {Scal}}{12} I \end{array}\right) \end{aligned}$$

where \(Z^*\) denotes the adjoint w.r.t. the induced metric on \(\Lambda ^2 T{\mathcal {N}}\), so that \(\text {W}= \text {W}^+ \oplus \text {W}^-\), the Weyl tensor seen as a 2-form on \(\Lambda ^2 T{\mathcal {N}}\), is the traceless, Hodge-commuting part of the Riemann curvature operator \(\text {R}\). Hence the following formula

$$\begin{aligned} \text {W}= {\hbox {Rm}}- \frac{1}{2} \text {Ric}\bigcirc \!\!\!\!\!\!\wedge \,g + \frac{\text {Scal}}{12} g \bigcirc \!\!\!\!\!\!\wedge \,g \, . \end{aligned}$$

If, additionally, \({\mathcal {N}}\) is a four dimensional Kähler manifold, then

Theorem 7.1

W\(^+\) can be written as a multiple of the scalar curvature by a parallel non-trivial 2-form on \(\Lambda ^2 T{\mathcal {N}}\).

See Prop. 2 in [5] for a proof and the explicit formula for the tensor involved. We do not need it explicitly since we are only interested in the following

Corollary 6

\(({\mathcal {N}}, g, \mathrm{J} )\) is anti-self-dual (\(\text {W}^+ = 0\)) if and only if the scalar curvature vanishes.

The result extends to the two cases considered in this article: (1) neutral pseudo-Kähler manifolds and (2) para-Kähler manifolds, with a slight twist: \(\text {W}^+\) is replaced by \(\text {W}^-\). Precisely:

Theorem 7.2

Let \(({\mathcal {N}}, g, \mathrm{J} )\) be a four dimensional manifold endowed with a pseudo-Kähler neutral metric (respectively a para-Kähler metric, necessarily neutral). Then the Weyl tensor W commutes with the Hodge operator and \({\mathcal {N}}\) is self-dual (W\(^- = 0\)) if and only if the scalar curvature vanishes.

The result for neutral pseudo-Kähler manifolds is probably known and relates to representation theory (see [3] for introduction and references), but since we could not find an explicit proof in the literatureFootnote 5, we will give a simple one below. To our knowledge, the proof for the para-Kähler case is new (albeit similar).

1.1 A.1 The pseudo-Kähler case

We will write explicitly the Weyl tensor in a given positively oriented orthonormal frame, denoted by \(( e_1, e_{1'}, e_2, e_{2'})\), where \(e_{1'} = \mathrm{J} e_1, e_{2'} = \mathrm{J} e_2, g ( e_1) = g ( e_{1'}) = - 1\) and \(g ( e_2) = g ( e_{2'}) = + 1\). (For brevity, \(g(X)\) denotes the norm \(g(X,X)\).) The pseudo-metric \(g\) extends to bivectors, has signature \(( 2, 4)\), and will be again denoted by \(g\): \(g ( e_a \wedge e_b) = g ( e_a) g ( e_b) - g ( e_a, e_b)^2 = g ( e_a) g ( e_b)\), so that \(\mathcal {B}= ( e_1 \wedge e_{1'}, e_1 \wedge e_2, e_1 \wedge e_{2'}, e_{1'} \wedge e_2, e_{1'} \wedge e_{2'}, e_2 \wedge e_{2'})\) is an orthonormal frame of \(\Lambda ^2\), with \(g ( e_a \wedge e_b) = - 1\), except for \(g ( e_1 \wedge e_{1'}) = g ( e_2 \wedge e_{2'}) = + 1\). (Note that the other convention, taking \(- g\) does not change the induced metric on \(\Lambda ^2\).)

Since the volume \(e_1 \wedge e_{1'} \wedge e_2 \wedge e_{2'}\) is positively oriented, we construct an orthonormal eigenbasis for the Hodge star on \(\Lambda ^2 T{\mathcal {N}}\):

$$\begin{aligned} \left\{ \begin{array}{l} E_1^{\pm } = \frac{\sqrt{2}}{2} ( e_1 \wedge e_{1'} \pm e_2 \wedge e_{2'})\\ E_2^{\pm } = \frac{\sqrt{2}}{2} ( e_1 \wedge e_2 \pm e_{1'} \wedge e_{2'})\\ E_3^{\pm } = \frac{\sqrt{2}}{2} ( e_1 \wedge e_{2'} \mp e_{1'} \wedge e_2) \end{array} \right. \end{aligned}$$

so that \(\Lambda ^{\pm }\) is generated by \(E_1^{\pm }, E_2^{\pm }, E_3^{\pm }\).

The Kähler condition implies

$$\begin{aligned} {\hbox {Rm}}( \mathrm{J} X, \mathrm{J} Y, Z, T) = {\hbox {Rm}}( X, Y, Z, T) = {\hbox {Rm}}( X, Y, \mathrm{J} Z, \mathrm{J} T), \end{aligned}$$

because \(\mathrm{J} \) is isometric and parallel. The matrix of the symmetric 2-form \(\text {R}\) in the orthonormal frame \(\mathcal {B}\) is

 

\(e_{11'}\)

\(e_{12}\)

\(e_{12'}\)

\(e_{1'2}\)

\(e_{1'2'}\)

\(e_{22'}\)

\(e_{11'}\)

\(\text {R}_{11' 11'}\)

\(\text {R}_{11' 12}\)

\(\text {R}_{11' 12'}\)

\(\text {R}_{11' 1' 2} = - \text {R}_{11' 12'}\)

\(\text {R}_{11' 1' 2'} = \text {R}_{11' 12}\)

\(\text {R}_{11' 22'}\)

\(e_{12}\)

 

\(\text {R}_{1212}\)

\(\text {R}_{1212'} \)

\(\text {R}_{121' 2}= - \text {R}_{1212'}\)

\(\text {R}_{131' 2'} = \text {R}_{1212} \)

\(\text {R}_{1222'}\)

\(e_{12'}\)

  

\(\text {R}_{12' 12'}\)

\(\text {R}_{12' 1' 2} = - \text {R}_{12' 12'}\)

\(\text {R}_{12' 1' 2'} = \text {R}_{1212'}\)

\(\text {R}_{12' 22'}\)

\(e_{1'2}\)

   

\(\text {R}_{1' 21' 2} = \text {R}_{12' 12'}\)

\(\text {R}_{1' 21' 2'} = - \text {R}_{1212'}\)

\(\text {R}_{1' 222'} = - \text {R}_{12' 22'}\)

\(e_{1'2'}\)

    

\(\text {R}_{1' 2' 1' 2'} = \text {R}_{1212}\)

\(\text {R}_{1' 2'22'} = \text {R}_{1222'}\)

\(e_{22'}\)

     

\(\text {R}_{22' 22'}\)

where \(e_{ab}\) stands for \(e_a \wedge e_b\), for greater legibility. We have written the matrix as a table for clarity and to make symmetries more obvious, and because \(\text {R}\) is symmetric we need only write half the matrix. We have used the internal symmetries of \(\text {R}\), to choose among equivalent coefficients the ones lowest in the lexicographic order of the indices.

The Weyl tensor satisfies some of the \(\mathrm{J} \)-symmetries of \(\text {R}\): indeed

$$\begin{aligned} \text {Ric}( \mathrm{J} X, \mathrm{J} Y)&= \sum _{i=1}^4 g ( e_i) {\hbox {Rm}}( \mathrm{J} X, e_i, \mathrm{J} Y, e_i) =\sum _{i=1}^4 g ( e_i) {\hbox {Rm}}( X, \mathrm{J} e_i, Y, \mathrm{J} e_i)\\&= \sum _{i=1}^4 g ( \mathrm{J} e_i) {\hbox {Rm}}( X, \mathrm{J} e_i, Y, \mathrm{J} e_i) = \text {Ric}( X, Y) \end{aligned}$$

because \(( \mathrm{J} e_i)\) is again an orthonormal basis. In particular, this invariance implies \(r_{11'} = \text {Ric}( e_1, e_{1'}) = r_{1' 1} = - r_{11'}\), so \(r_{11'}\) vanish (and so does \(r_{22'}\)). For the Kulkarni–Nomizu product,

$$\begin{aligned} \text {Ric}\bigcirc \!\!\!\!\!\!\wedge \,g ( \mathrm{J} X, Y, Z, T)&= \text {Ric}(\mathrm{J} X, Z) g (Y, T) + \text {Ric}(Y, T) g (\mathrm{J} X, Z)\\&- \text {Ric}(\mathrm{J} X, T) g (Y, Z) - \text {Ric}(Y,Z) g (\mathrm{J} X, T)\\&= - \text {Ric}(X, \mathrm{J} Z) g (\mathrm{J} Y, \mathrm{J} T) - \text {Ric}(\mathrm{J} Y, \mathrm{J} T) g (X, \mathrm{J} Z)\\&+ \text {Ric}(X, \mathrm{J} T) g (\mathrm{J} Y, \mathrm{J} Z) + \text {Ric}(\mathrm{J} Y, \mathrm{J} Z) g (X, \mathrm{J} T)\\&= - \text {Ric}\bigcirc \!\!\!\!\!\!\wedge \,g ( X, \mathrm{J} Y, \mathrm{J} Z, \mathrm{J} T) \end{aligned}$$

so

$$\begin{aligned} \text {Ric}\bigcirc \!\!\!\!\!\!\wedge \,g ( \mathrm{J} X, \mathrm{J} Y, Z, T) = - \text {Ric}\bigcirc \!\!\!\!\!\!\wedge \,g ( X, \mathrm{J} ^2 Y, \mathrm{J} Z, \mathrm{J} T) = \text {Ric}\bigcirc \!\!\!\!\!\!\wedge \,g ( X, Y, \mathrm{J} Z, \mathrm{J} T) \, . \end{aligned}$$

Hence the following symmetries (fewer than for \({\hbox {Rm}}\)) in the coefficients of \(\text {Ric}\bigcirc \!\!\!\!\!\!\wedge \,g\), \(g \bigcirc \!\!\!\!\!\!\wedge \,g\) and \({\hbox {Rm}}\), and therefore \(\text {W}\):

 

\(e_{1 1'}\)

\(e_{1} \wedge e_{2}\)

\(e_{1 2'}\)

\(e_{1'} \wedge e_{2}\)

\(e_{1' 2'}\)

\(e_{2 2'}\)

\(e_{1 1'}\)

\(\text {W}_{11' 11'}\)

\(\text {W}_{11' 12}\)

\(\text {W}_{11' 12'}\)

\(\text {W}_{11' 1' 2} = - \text {W}_{11' 12'}\)

\(\text {W}_{11' 1' 2'} = \text {W}_{11' 12}\)

\(\text {W}_{11' 22'}\)

\(e_{1 2}\)

 

\(\text {W}_{1212}\)

\(\text {W}_{1212'}\)

\(\text {W}_{121' 2}\)

\(\text {W}_{121' 2'}\)

\(\text {W}_{1222'}\)

\(e_{1 2'}\)

  

\(\text {W}_{12' 12'}\)

\(\text {W}_{12' 1' 2}\)

\(\text {W}_{12' 1' 2'} = - \text {W}_{121' 2}\)

\(\text {W}_{12' 22'}\)

\(e_{1' 2}\)

   

\(\text {W}_{1' 21' 2} = \text {W}_{12' 12'}\)

\(\text {W}_{1' 21' 2'} = - \text {W}_{1212'}\)

\(\text {W}_{1' 222'} = - \text {W}_{12' 22'}\)

\(e_{1' 2'}\)

    

\(\text {W}_{1' 2' 1' 2'} = \text {W}_{1212}\)

\(\text {W}_{1' 2' 22'} = \text {W}_{1222'}\)

\(e_{2 2'}\)

     

\(\text {W}_{22' 22'}\)

Expanding on the above eigenbasis of \(\Lambda ^+ \oplus \Lambda ^-\) (which differs from the one in the positive definite case) yields the following Weyl tensor coefficients, which we have simplified using the symmetries above (up to a factor \(1 / 2\) due to normalization):

 

\(E_1^+\)

\(E_2^+\)

\(E_3^+\)

\(E_1^+\)

\(\text {W}_{11' 11'}\) + \(\text {W}_{22' 22'}\) + \(2 \text {W}_{11' 22'}\)

\(2 ( \text {W}_{11' 12} + \text {W}_{1222'})\)

\(2 ( \text {W}_{11' 12'} + \text {W}_{12' 22'})\)

\(E_2^+\)

 

\(2 ( \text {W}_{1212} + \text {W}_{121' 2'})\)

\(2 ( \text {W}_{1212'} - \text {W}_{121' 2})\)

\(E_3^+\)

  

\(2 ( \text {W}_{12' 12'} - \text {W}_{12' 1' 2})\)

\(E_1^-\)

   

\(E_2^-\)

   

\(E_3^-\)

   
 

\(E_1^-\)

\(E_2^-\)

\(E_3^-\)

\(E_1^+\)

\(\text {W}_{11' 11'} - \text {W}_{22' 22'}\)

0

0

\(E_2^+\)

\(2 ( \text {W}_{11' 12} - \text {W}_{1222'})\)

0

0

\(E_3^+\)

\(2 ( \text {W}_{11' 12'} - \text {W}_{12' 22'})\)

0

0

\(E_1^-\)

\(\text {W}_{11' 11'} + \text {W}_{22' 22'} - 2 \text {W}_{11' 22'}\)

0

0

\(E_2^-\)

 

\(2 ( \text {W}_{1212} - \text {W}_{121' 2'})\)

\(2 ( \text {W}_{1212'} + \text {W}_{121' 2})\)

\(E_3^-\)

  

\(2 ( \text {W}_{12' 12'} + \text {W}_{12' 1' 2})\)

(Again only half the coefficients are written down.) Further simplifications come from computing \(\text {W}\), and using

$$\begin{aligned} \text {Scal}&= - r_{11} - r_{1' 1'} + r_{22} + r_{2' 2'} = 2 ( r_{22} -r_{11})\\&= 2 ( - ( - \text {R}_{11' 11'} + \text {R}_{1212} + \text {R}_{12' 12'}) + ( - \text {R}_{1212} - \text {R}_{1' 21' 2} + \text {R}_{22' 22'}))\\&= 2 ( \text {R}_{11' 11'} - 2 ( \text {R}_{1212} + \text {R}_{12' 12'}) + \text {R}_{22' 22'}) \, . \end{aligned}$$

First prove that the Hodge star commutes with \(\text {W}\) by considering \(\text {W}( \Lambda ^+, \Lambda ^-)\):

$$\begin{aligned} \text {W}_{11' 11'}&= \text {R}_{11' 11'} + \frac{1}{2} ( r_{11} + r_{1' 1'}) + \frac{\text {Scal}}{6} = \text {R}_{11' 11'} + r_{11} + \frac{\text {Scal}}{6}\\&= \text {R}_{1212} + \text {R}_{12' 12'} + \frac{\text {Scal}}{6}\\ \text {W}_{22' 22'}&= \text {R}_{22' 22'} - \frac{1}{2} ( r_{22} + r_{2' 2'}) + \frac{\text {Scal}}{6} = \text {R}_{22' 22'} - r_{22} + \frac{\text {Scal}}{6}\\&= \text {R}_{1212} + \text {R}_{12' 12'} + \frac{\text {Scal}}{6} \end{aligned}$$

so that \(\text {W}_{11' 11'} - \text {W}_{22' 22'} = 0\). Similarly

$$\begin{aligned} \text {W}_{11' 12} = \text {R}_{11' 12} + \frac{r_{1' 2}}{2}, \text {W}_{1222'} = \text {R}_{1222'} + \frac{r_{12'}}{2} = \text {R}_{1222'} - \frac{r_{1' 2}}{2} \end{aligned}$$

so

$$\begin{aligned}&\text {W}_{11' 12} - \text {W}_{1222'} = \text {R}_{11' 12} - \text {R}_{1222'} + r_{1' 2} = 0\\&\text {W}_{11' 12'} = \text {R}_{11' 12'} + \frac{r_{1' 2'}}{2} = \text {R}_{11' 12'} + \frac{r_{12}}{2}, \text {W}_{12' 22'} = \text {R}_{12' 22'} - \frac{r_{12}}{2},\\&\text {W}_{11' 12'} - \text {W}_{12' 22'} = \text {R}_{11' 12'} - \text {R}_{12' 22'} + r_{12} = 0 . \end{aligned}$$

That proves that \(\text {W}\) is block-diagonal.

The \(\text {W}^-\) term satisfies

$$\begin{aligned} \text {W}_{11' 11'} + \text {W}_{22' 22'} - 2 \text {W}_{11' 22'}&= \text {R}_{11' 11'} + r_{11} + \text {R}_{22' 22'} - r_{22} + \frac{\text {Scal}}{3}\\&- 2 \text {R}_{11' 22'}\\&= \text {R}_{11' 11'} + \text {R}_{22' 22'} - 2 \text {R}_{11' 22'} - \frac{\text {Scal}}{6}\\&= \text {R}_{11' 11'} + \text {R}_{22' 22'} - 2 ( \text {R}_{1212} + \text {R}_{12' 12'})\\&- \frac{\text {Scal}}{6}\\&= \frac{\text {Scal}}{2} - \frac{\text {Scal}}{6} = \frac{\text {Scal}}{3} \end{aligned}$$

using the first Bianchi identity (and the invariance of \({\hbox {Rm}}\)):

$$\begin{aligned} \text {R}_{11' 22'}&= - \text {R}_{1' 212'} - \text {R}_{211' 2'} = \text {R}_{12' 12'} + \text {R}_{1212} .\\ \text {W}_{1212} - \text {W}_{121' 2'}&= \text {R}_{1212} + \frac{r_{22} - r_{11}}{2} - \frac{\text {Scal}}{6} - \text {R}_{121' 2'} = \frac{\text {Scal}}{4} - \frac{\text {Scal}}{6}\\&= \frac{\text {Scal}}{12}\\ \text {W}_{12' 12'} + \text {W}_{12' 1' 2}&= \text {R}_{12' 12'} + \frac{\text {Scal}}{4} - \frac{\text {Scal}}{6} + \text {R}_{12' 1' 2} = \frac{\text {Scal}}{12}\\ \text {W}_{1212'} + \text {W}_{121' 2}&= \text {R}_{1212'} + \frac{r_{22'}}{2} + \text {R}_{121' 2} - \frac{r_{11'}}{2} = \frac{1}{2} ( r_{22'} - r_{11'}) = 0 . \end{aligned}$$

Finally,

$$\begin{aligned} \text {W}^- = \text {Scal} \left( \begin{array}{ccc} 1 / 3 &{} &{} \\ &{} 1 / 6 &{} \\ &{} &{} 1 / 6 \end{array}\right) = \frac{\text {Scal}}{6} \text {Id} + \frac{\text {Scal}}{6} E^-_1 \otimes E^-_1 \end{aligned}$$

(and indeed this matrix is traceless w.r.t. the pseudo-metric \(g\)). One should note that the above expression differs from the Riemannian case, where

$$\begin{aligned} \text {W}^+ = \text {Scal} \left( \begin{array}{ccc} 1 / 3 &{} &{} \\ &{} - 1 / 6 &{} \\ &{} &{} - 1 / 6 \end{array}\right) = - \frac{\text {Scal}}{6} \text {Id} + \frac{\text {Scal}}{3} E_1^+ \otimes E_1^+ . \end{aligned}$$

We let the Reader check that in the neutral case, the \(\text {W}^+\) part is not a multiple of the scalar curvature, which completes the proof of Theorem 7.2.

1.2 A.2 The para-Kähler case

The computations are almost identical, but the results differ from the pseudo-Kähler setup, because the para-complex structure \(\mathrm{J} \) is now an anti-isometry: \(\text {R}( \mathrm{J} X, \mathrm{J} Y)Z = - \text {R}( X, Y) Z\). We pick an orthonormal basis \(( e_1, e_{1'}, e_2, e_{2'})\) with \(e_{1'} = \mathrm{J} e_1, e_{2'} = \mathrm{J} e_2\), and \(g(e_1) = g ( e_2) = + 1\), \(g ( e_{1'}) = g ( e_{2'}) = - 1\). The frame \(\mathcal {B}= ( e_1 \wedge e_{1'}, e_1 \wedge e_2, e_1 \wedge e_{2'}, e_{1'} \wedge e_2, e_{1'} \wedge e_{2'}, e_2 \wedge e_{2'})\) of \(\Lambda ^2 T{\mathcal {N}}\) is also orthonormal w.r.t. the induced metric on \(\Lambda ^2\), again denoted by \(g\), which has signature \(( 2, 4)\): \(g ( e_a \wedge e_b) = g ( e_a) g ( e_b) = - 1\), except for \(g ( e_1 \wedge e_2) = g ( e_{1'} \wedge e_{2'}) = + 1\).

An orthonormal eigenbasis for the Hodge operator is the following:

$$\begin{aligned} \left\{ \begin{array}{l} E_1^{\pm } = \frac{\sqrt{2}}{2} ( e_1 \wedge e_{1'} \mp e_2 \wedge e_{2'})\\ E_2^{\pm } = \frac{\sqrt{2}}{2} ( e_1 \wedge e_2 \mp e_{1'} \wedge e_{2'})\\ E_3^{\pm } = \frac{\sqrt{2}}{2} ( e_1 \wedge e_{2'} \mp e_{1'} \wedge e_2) \end{array} \right. \end{aligned}$$

where the \(E_a^+\) (resp. \(E_a^-\)) span \(\Lambda ^+\) (resp. \(\Lambda ^-\)). (Note the sign differences w.r.t. the pseudo-Kähler case.)

Since \(\mathrm{J} \) is anti-isometric and parallel,

$$\begin{aligned} {\hbox {Rm}}( \mathrm{J} X, \mathrm{J} Y, Z, T) = - {\hbox {Rm}}( X,Y, Z, T) = {\hbox {Rm}}( X, Y, \mathrm{J} Z, \mathrm{J} T) \, . \end{aligned}$$

Hence the following symmetries of the Riemannian curvature operator \(\text {R}\), expressed in the frame \(\mathcal {B}\) (for symmetry reasons and greater legibility, lower left coefficients are not written in this and the subsequent matrices):

 

\(e_{1 1'}\)

\(e_{12}\)

\(e_{1 2'}\)

\(e_{1'} \wedge e_{2}\)

\(e_{1' 2'}\)

\(e_{2 2'}\)

\(e_{1 1'}\)

\(\text {R}_{11' 11'}\)

\(\text {R}_{11' 12}\)

\(\text {R}_{11' 12'}\)

\(\text {R}_{11' 1' 2} = - \text {R}_{11' 12'}\)

\(\text {R}_{11' 1' 2'} = - \text {R}_{11' 12}\)

\(\text {R}_{11' 22'}\)

\(e_{1 2}\)

 

\(\text {R}_{1212}\)

\(\text {R}_{1212'}\)

\(\text {R}_{121' 2} = - \text {R}_{1212'}\)

\(\text {R}_{121' 2'} = - \text {R}_{1212}\)

\(\text {R}_{1222'}\)

\(e_{1 2'}\)

  

\(\text {R}_{12' 12'}\)

\(\text {R}_{12' 1' 2} = - \text {R}_{12' 12'}\)

\(\text {R}_{12' 1' 2'} = - \text {R}_{1212'}\)

\(\text {R}_{12' 22'}\)

\(e_{1' 2}\)

   

\(\text {R}_{1' 21' 2} = \text {R}_{12' 12'}\)

\(\text {R}_{1' 21' 2'} = \text {R}_{1212'}\)

\(\text {R}_{1' 222'} = - \text {R}_{12' 22'}\)

\(e_{1' 2'}\)

    

\(\text {R}_{1' 2' 1' 2'} = \text {R}_{1212}\)

\(\text {R}_{1' 2'22'} = - \text {R}_{1222'}\)

\(e_{2 2'}\)

     

\(\text {R}_{22' 22'}\)

(Note again the similarity with the pseudo-Kähler case: only a few signs change.)

The Weyl tensor satisfies some of the \(\mathrm{J} \)-symmetries of \({\hbox {Rm}}\) since

$$\begin{aligned} \text {Ric}( \mathrm{J} X, \mathrm{J} Y)&= \sum _{i=1}^4 g(e_i) {\hbox {Rm}}( \mathrm{J} X, e_i, \mathrm{J} Y, e_i) = \sum _{i=1}^4 g ( e_i) {\hbox {Rm}}( X, \mathrm{J} e_i, Y, \mathrm{J} e_i)\\&= - \sum _{i=1}^4 g ( \mathrm{J} e_i) {\hbox {Rm}}( X, \mathrm{J} e_i, Y, \mathrm{J} e_i) = - \text {Ric}( X, Y) \end{aligned}$$

since \(( \mathrm{J} e_i)\) is also an orthonormal basis. In particular this invariance implies \(r_{1' 1} = r_{11'} = - r_{1' 1}\), so \(r_{11'}\) vanishes (and so does \(r_{22'}\)). Finally,

$$\begin{aligned} \frac{\text {Scal}}{2} = r_{11} + r_{22} = - \text {R}_{11' 11'} + 2 ( \text {R}_{1212} - \text {R}_{12' 12'}) - \text {R}_{22' 22'} . \end{aligned}$$

The Kulkarni–Nomizu product \(\text {Ric}\bigcirc \!\!\!\!\!\!\wedge \,g\) satisfies

$$\begin{aligned} \text {Ric}\bigcirc \!\!\!\!\!\!\wedge \,g ( \mathrm{J} X, Y, Z, T)&= \text {Ric}(\mathrm{J} X, Z) g (Y, T) + \text {Ric}(Y, T) g (\mathrm{J} X, Z) \\&- \text {Ric}(\mathrm{J} X, T) g (Y, Z) - \text {Ric}(Y,Z) g (\mathrm{J} X, T) \\&= \text {Ric}(X, \mathrm{J} Z) g (\mathrm{J} Y, \mathrm{J} T) + \text {Ric}(\mathrm{J} Y, \mathrm{J} T) g (X, \mathrm{J} Z)\\&- \text {Ric}(X, \mathrm{J} T) g (\mathrm{J} Y, \mathrm{J} Z) - \text {Ric}(\mathrm{J} Y, \mathrm{J} Z) g (X, \mathrm{J} T)\\&= \text {Ric}\bigcirc \!\!\!\!\!\!\wedge \,g ( X, \mathrm{J} Y, \mathrm{J} Z, \mathrm{J} T) \end{aligned}$$

so

$$\begin{aligned} \text {Ric}\bigcirc \!\!\!\!\!\!\wedge \,g ( \mathrm{J} X, \mathrm{J} Y, Z, T) = \text {Ric}\bigcirc \!\!\!\!\!\!\wedge \,g ( X, \mathrm{J} ^2 Y, \mathrm{J} Z, \mathrm{J} T) = \text {Ric}\bigcirc \!\!\!\!\!\!\wedge \,g ( X, Y, \mathrm{J} Z, \mathrm{J} T) \end{aligned}$$

and the same property holds for \(g \bigcirc \!\!\!\!\!\!\wedge \,g\). Hence the following symmetries (fewer than for \({\hbox {Rm}}\)) in the coefficients of \(\text {Ric}\bigcirc \!\!\!\!\!\!\wedge \,g\), \(g \bigcirc \!\!\!\!\!\!\wedge \,g\) and \({\hbox {Rm}}\), and therefore \(\text {W}\):

 

\(e_{1 1'}\)

\(e_{1 2}\)

\(e_{1 2'}\)

\(e_{1' 2}\)

\(e_{1' 2'}\)

\(e_{2 2'}\)

\(e_{1 1'}\)

\(\text {W}_{11' 11'}\)

\(\text {W}_{11' 12}\)

\(\text {W}_{11' 12'}\)

\(\text {W}_{11' 1' 2} = - \text {W}_{11' 12'}\)

\(\text {W}_{11' 1' 2'} = - \text {W}_{11' 12}\)

\(\text {W}_{11' 22'}\)

\(e_{1 2}\)

 

\(\text {W}_{1212}\)

\(\text {W}_{1212'}\)

\(\text {W}_{121' 2}\)

\(\text {W}_{121' 2'}\)

\(\text {W}_{1222'}\)

\(e_{1 2'}\)

  

\(\text {W}_{12' 12'}\)

\(\text {W}_{12' 1' 2}\)

\(\text {W}_{12' 1' 2'} = \text {W}_{121' 2}\)

\(\text {W}_{12' 22'}\)

\(e_{1' 2}\)

   

\(\text {W}_{1' 21' 2} = \text {W}_{12' 12'}\)

\(\text {W}_{1' 21' 2'} = \text {W}_{1212'}\)

\(\text {W}_{1' 222'} = - \text {W}_{12' 22'}\)

\(e_{1' 2'}\)

    

\(\text {W}_{1' 2' 1' 2'} = \text {W}_{1212}\)

\(\text {W}_{1' 2'22'} = - \text {W}_{1222'}\)

\(e_{2 2'}\)

     

\(\text {W}_{22' 22'}\)

Let us now express \(\text {W}\) in the Hodge basis defined earlier, using the above symmetries (up to a factor \(1 / 2\) due to normalization).

 

\(E_1^+\)

\(E_2^+\)

\(E_3^+\)

\(E_1^+\)

\(\text {W}_{11' 11'} + \text {W}_{22' 22'} - 2 \text {W}_{11' 22'}\)

\(2 ( \text {W}_{11' 12} - \text {W}_{1222'})\)

\(2 ( \text {W}_{11' 12'} - \text {W}_{12' 22'})\)

\(E_2^+\)

 

\(2 ( \text {W}_{1212} - \text {W}_{121' 2'})\)

\(2 ( \text {W}_{1212'} - \text {W}_{121' 2})\)

\(E_3^+\)

  

\(2 ( \text {W}_{12' 12'} - \text {W}_{12' 1' 2})\)

\(E_1^-\)

   

\(E_2^-\)

   

\(E_3^-\)

   
 

\(E_1^-\)

\(E_2^-\)

\(E_3^-\)

\(E_1^+\)

\(\text {W}_{11' 11'} - \text {W}_{22' 22'}\)

0

0

\(E_2^+\)

\(2 ( \text {W}_{11' 12} + \text {W}_{1222'})\)

0

0

\(E_3^+\)

\(2 ( \text {W}_{11' 12'} + \text {W}_{12' 22'})\)

0

0

\(E_1^-\)

\(\text {W}_{11' 11'} + \text {W}_{22' 22'} + 2 \text {W}_{11' 22'}\)

0

0

\(E_2^-\)

 

\(2 ( \text {W}_{1212} + \text {W}_{121' 2'})\)

\(2 ( \text {W}_{1212'} + \text {W}_{121' 2})\)

\(E_3^-\)

  

\(2 ( \text {W}_{12' 12'} + \text {W}_{12' 1' 2})\)

Only three terms in the off-block-diagonal part are not obviously zero.

$$\begin{aligned} \text {W}_{11' 11'}&= \text {R}_{11' 11'} - \frac{1}{2} ( - r_{11} + r_{1' 1'}) - \frac{\text {Scal}}{6} = \text {R}_{11' 11'} + r_{11} - \frac{\text {Scal}}{6}\\ \text {W}_{22' 22'}&= \text {R}_{22' 22'} - \frac{1}{2} ( - r_{22} + r_{2' 2'}) - \frac{\text {Scal}}{6} = \text {R}_{22' 22'} + r_{22} - \frac{\text {Scal}}{6} \end{aligned}$$

but \(r_{11} = - \text {R}_{11' 11'} + \text {R}_{1212} - \text {R}_{12' 12'}\) and \(r_{22} = \text {R}_{2121} - \text {R}_{21' 21'} - \text {R}_{22' 22'} = \text {R}_{1212} - \text {R}_{12' 12'} - \text {R}_{22' 22'}\) so that

$$\begin{aligned} \text {W}_{11' 11'} - \text {W}_{22' 22'} = \text {R}_{11' 11'} - \text {R}_{22' 22'} + r_{11} - r_{22} = 0 . \end{aligned}$$

Similarly

$$\begin{aligned}&\text {W}_{11' 12} + \text {W}_{1222'} = \text {R}_{11' 12} - \frac{r_{1' 2}}{2} + \text {R}_{1222'} + \frac{r_{12'}}{2} = \text {R}_{11' 12} + \text {R}_{1222'} - r_{1' 2} = 0\\&\text {W}_{11' 12'} + \text {W}_{12' 22'} = \text {R}_{11' 12'} - \frac{r_{1' 2'}}{2} + \text {R}_{12' 22'} + \frac{r_{12}}{2} = \text {R}_{11' 12'} + \text {R}_{12' 22'} + r_{12} = 0 \end{aligned}$$

which proves that \(\text {W}\) is block-diagonal, i.e. commutes with the Hodge operator.

Let us now look more closely at the \(\text {W}^-\) term

$$\begin{aligned}&\left( \begin{array}{ccc} \text {W}_{11' 11'} + \text {W}_{22' 22'} + 2 \text {W}_{11' 22'} &{} 0 &{} 0\\ &{} 2 ( \text {W}_{1212} + \text {W}_{121' 2'}) &{} 2 ( \text {W}_{1212'} + \text {W}_{121' 2})\\ &{} &{} 2 ( \text {W}_{12' 12'} + \text {W}_{12' 1' 2}) \end{array}\right) \\&\text {W}_{11' 11'} + \text {W}_{22' 22'} + 2 \text {W}_{11' 22'}\\&\qquad = \text {R}_{11' 11'} + r_{11} - \frac{\text {Scal}}{6} + \text {R}_{22' 22'} + r_{22} - \frac{\text {Scal}}{6} + 2 \text {R}_{11' 22'}\\&\qquad = \text {R}_{11' 11'} + \text {R}_{22' 22'} + 2 \text {R}_{11' 22'} + \frac{\text {Scal}}{2} - \frac{\text {Scal}}{3}\\&\qquad = \text {R}_{11' 11'} + \text {R}_{22' 22'} + 2 ( - \text {R}_{1212} + \text {R}_{12' 12'}) + \frac{\text {Scal}}{6} = - \frac{\text {Scal}}{3} \end{aligned}$$

where we have used the first Bianchi identity (and the invariance of \({\hbox {Rm}}\))

$$\begin{aligned} \text {R}_{11' 22'}&= - \text {R}_{1' 212'} - \text {R}_{211' 2'} = \text {R}_{12' 12'} - \text {R}_{1212} .\\ \text {W}_{1212} + \text {W}_{121' 2'}&= \text {R}_{1212} - \frac{r_{22} + r_{11}}{2} + \frac{\text {Scal}}{6} + \text {R}_{121' 2'}\\&= \text {R}_{1212} - \frac{\text {Scal}}{4} + \frac{\text {Scal}}{6} + \text {R}_{121' 2'} = - \frac{\text {Scal}}{12}\\ \text {W}_{12' 12'} + \text {W}_{12' 1' 2}&= \text {R}_{12' 12'} + \frac{\text {Scal}}{4} - \frac{\text {Scal}}{6} + \text {R}_{12' 1' 2} = \frac{\text {Scal}}{12}\\ \text {W}_{1212'} + \text {W}_{121' 2}&= \text {R}_{1212'} - \frac{r_{22'}}{2} + \text {R}_{121' 2} - \frac{r_{11'}}{2} = 0 . \end{aligned}$$

Finally,

$$\begin{aligned} \text {W}^- = \text {Scal} \left( \begin{array}{ccc} - 1 / 3 &{} &{} \\ &{} - 1 / 6 &{} \\ &{} &{} 1 / 6 \end{array}\right) \end{aligned}$$

vanishes if and only if \(\text {Scal} = 0\). (The Reader will check that this matrix is indeed traceless w.r.t. the pseudo-metric \(g\).)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Anciaux, H., Romon, P. A canonical structure on the tangent bundle of a pseudo- or para-Kähler manifold. Monatsh Math 174, 329–355 (2014). https://doi.org/10.1007/s00605-014-0630-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00605-014-0630-6

Keywords

Mathematics Subject Classification (2010)

Navigation