• Nie Znaleziono Wyników

Some framed f-structures on transversally Finsler foliations

N/A
N/A
Protected

Academic year: 2021

Share "Some framed f-structures on transversally Finsler foliations"

Copied!
10
0
0

Pełen tekst

(1)

U N I V E R S I T A T I S M A R I A E C U R I E - S K Ł O D O W S K A L U B L I N – P O L O N I A

VOL. LXV, NO. 1, 2011 SECTIO A 87–96

CRISTIAN IDA

Some framed f -structures on transversally Finsler foliations

Abstract. Some problems concerning to Liouville distribution and framed f -structures are studied on the normal bundle of the lifted Finsler foliation to its normal bundle. It is shown that the Liouville distribution of transver- sally Finsler foliations is an integrable one and some natural framed f (3, ε)- structures of corank 2 exist on the normal bundle of the lifted Finsler foliation.

1. Introduction and preliminaries. The study of structures on mani- folds defined by a tensor field satisfying f3± f = 0 has the origin in a paper by K. Yano [15]. Later on, these structures have been generically called f - structures. On the tangent manifold of a Finsler space, the notion of framed f (3, 1)-structure was defined and studied by M. Anastasiei in [2]. Further developments concerning framed f (3, −1)-structure on such manifold was studied in [5, 6]. In a paper by A. Miernowski and W. Mozgawa [9] was defined the notion of transversally Finsler foliation and there it is proved that the normal bundle of the lifted Finsler foliation to its normal bundle has a local model of tangent manifold and it is the Riemannian one. Thus, some problems specific for tangent manifolds can be extended and studied on the normal bundle of the lifted Finsler foliation. Firstly, following [4], we define a Liouville distribution in the vertical bundle and we prove that it is integrable. Next, by analogy with [2], some framed f (3, ε)-structures

2000 Mathematics Subject Classification. 53C12, 53B40.

Key words and phrases. Transversally Finsler foliation, Liouville distribution, framed f -structures.

(2)

on the normal bundle of the lifted Finsler foliation are defined and studied in the paper.

In this section we briefly recall some basic facts about transversally Finsler foliations (see [9]).

Let (M, F ) be a n-dimensional Finsler manifold, where F : T M → R is a Finsler metric (for necessary definitions, see [3, 11]).

Definition 1.1 ([9]). A diffeomorphism f : M → M is said to be a Finsler isometry if F (f (x), f(y)) = F (x, y) for each x ∈ M and y ∈ TxM .

Definition 1.2 ([9]). A foliated cocycle {Ui, fi, γij} on a (n + m)-dimen- sional manifold M is said to be a Finslerian foliation F if

(i) {Ui}, i ∈ I is an open covering of M;

(ii) fi : Ui → M is a submersion, where (M, F ) is a Finsler manifold;

(iii) γij is a local Finslerian isometry of (M, F ) such that for each u ∈ Ui∩ Uj

fi(u) = (γij◦ fj)(u).

The Finsler manifold (M, F ) will be called the transversal manifold of foliation F .

The local submersions {fi} define by pull-back a Finsler metric in the normal bundle Q = Q(M) of the foliation F , denoted by FQ and given by (1.1) FQ(u, p(Xu)) = F (fi(u), (fi)(Xu))

for any u ∈ M, Xu ∈ TuM, where p : T M → Q is the natural projection (for Lagrangian case, see [12]).

We denote by u = (xi, xα), α = 1, . . . , n, i = n + 1, . . . , n + m = dim M the adapted coordinates in a local foliated chart on M and let {∂xi,∂xα} be a local frame of T M. If we denote by {∂xα} the corresponding local frame of Q, then we can induce a chart (xi, xα, yα) on Q, where yα ∂∂xα is a transversal vector at a point (xi, xα). Note that in this coordinate system the metric FQ in Q does not depend on (xi).

According to [9], the distribution spanned by {∂xi}, i = n + 1, . . . , n + m defines a foliation FQ on Q called the natural lift of F to Q.

Let us consider Q(Q) to be the bundle over Q transversal to the foliation FQ. The canonical projection π : Q → M, π(xi, xα, yα) = (xi, xα) induces another projection π : T Q → T M which maps the tangent vectors to FQ

in the tangent vectors to F . Thus, π induces a mapping π : Q(Q) → Q defined byπ = p ◦ π◦ p−1, where p : T Q → Q(Q) is the natural projection.

If we denote by V := V (Q) = ker π, it is a vertical bundle spanned by the vectors {∂yα}, α = 1, . . . , n.

(3)

Let us put GQ = FQ2 and Gα = ∂G∂yαQ, Gαβ =

2GQ

∂yα∂yβ, etc. Then, by the same arguments as in [3, 11] we have

(1.2) Gαβyα= Gβ; Gαβyαyβ = 2GQ; Gαβγyγ= 0.

According to [9], for any vertical vectors X = ˙Xα ∂∂yα and Y = ˙Yβ ∂∂yβ the formula

(1.3) Gv(X, Y ) = Gαβαβ defines a Riemannian metric in the vertical bundle V .

An important global vector field is defined by (1.4) Γ : Q → V ; Γ(xi, xα, yα) = yα

∂yα

and it is called the Liouville vector field (or radial vertical vector field).

Moreover, by the second equality of (1.2), we have

(1.5) GQ= 1

2Gv(Γ, Γ) > 0.

Also, by using the technique of good vertical connection often used in Finsler geometry (see [1]), in [9] it is proved that the normal bundle Q(Q) has a local model of tangent manifold. Thus we have the spliting Q(Q) = H(Q) ⊕ V (Q), where the horizontal bundle H := H(Q) is spanned by the vectors {δxδα = ∂xα − Nαβ

∂yβ}, where the coefficients Nαβ are related only in terms of Finsler metric GQ.

In the sequel we will use the adapted basis {δα := δxδα,∂.α:= ∂yα} as well as its dual {dxα, δyα:= dyα+ Nβαdxβ}.

We notice that Gv induces a Riemannian metric on the horizontal bundle denoted by Gh and we consider G = Gh + Gv the Sasaki type lift of the fundamental tensor Gαβ, locally given by

(1.6) G = Gαβdxα⊗ dxβ+ Gαβδyα⊗ δyβ. 2. A Liouville distribution. Let us consider ξ = Γ

FQ

2 to be the unit Liouville vector field with respect to Gv, i.e.

(2.1) Gv(ξ, ξ) = 1.

Using Gv and ξ, we define a vertical 1-form η ∈ Γ(V) by (2.2) η(X) = Gv(X, ξ), ∀ X ∈ Γ(V ).

Denote by {ξ} the line vector bundle over Q spanned by ξ and we define the Liouville distribution as the complementary orthogonal distribution SQ to {ξ} in V with respect to Gv, namely V = SQ ⊕ {ξ}. Hence, SQ is defined by η, that is

(2.3) Γ(SQ) = {X ∈ Γ(V ) ; η(X) = 0}.

(4)

Thus, any vertical vector field X ∈ Γ(V ) can be expressed as

(2.4) X = ΦX + η(X)ξ,

where Φ is the projection morphism of V on SQ. By direct calculations, one gets

Proposition 2.1. For any vertical vector fields X, Y ∈ Γ(V ), we have (2.5) Gv(X, ΦY ) = Gv(ΦX, ΦY ) = Gv(X, Y ) − η(X)η(Y ).

Then the local components of η and Φ with respect to the basis {δyα} and {δyα⊗∂.α}, respectively, are given by

(2.6) ηα = Gα

FQ√ 2 and

(2.7) Φβα= δβα− ηαyβ FQ

√ 2, where δβα denotes the Kronecker symbol.

Theorem 2.1. The Liouville distribution SQ is integrable.

Proof. Let X, Y ∈ Γ(SQ). As V is an integrable distribution on Q, it is sufficient to prove that [X, Y ] has no component with respect to ξ. By using (1.3) and (2.2), we obtain that X ∈ Γ(SQ) if and only if

(2.8) Gαβyαβ = 0,

where ˙Xβ are the components of X. Differentiating (2.8) with respect to yγ, we get

(2.9) Gαβγyαβ+ Gγββ+ Gαβyα.γβ = 0 , ∀ γ = 1, . . . , n and taking into account the last equality of (1.2), we get

(2.10) Gγββ+ Gαβyα.γβ = 0 , ∀ γ = 1, . . . , n.

Then, by direct calculations using (1.3), (1.4) and (2.10), we have Gv([X, Y ], ξ) = 1

FQ

2Gαβyα h.

γ( ˙Yβ) ˙Xγ−∂.γ ( ˙Xβ) ˙Yγ i

= − 1 FQ

√ 2



Gγββγ− Gγββγ



= 0

which completes the proof. 

(5)

Let us consider ∇ : X (V ) → X (TQ ⊗ V ) to be the unique good vertical Bott connection introduced in [9]. We notice that for vertical vector fields, it is locally given by

(2.11) ∇.

α

.

β= Cαβγ.γ, where the local vertical coefficients are given by

(2.12) Cαβγ = 1

2Gδγ.α (Gβδ), where (Gδγ) denotes the inverse of (Gγδ).

Now, contracting (2.12) by yα, we deduce

(2.13) Cαβγ yα= 0.

By straightforward calculations using (2.4), (2.5), (2.6), (2.7) and (2.13) we obtain

Proposition 2.2. The vertical covariant derivatives, with respect to ∇ of ξ, η and Φ, are

(2.14) ∇Xξ = 1

FQ

√ 2ΦX

(2.15) (∇Xη)Y = 1

FQ

2Gv(ΦX, ΦY ) (2.16) (∇XΦ)Y = − 1

FQ

2[Gv(ΦX, ΦY )ξ + η(Y )ΦX]

for any X, Y ∈ Γ(V ).

3. Some framed f (3, ε)-structures on Q(Q). A framed f (3, 1)-struc- ture of corank s on a (2n+s)-dimensional manifold N is a natural generaliza- tion of an almost contact structure on N and it is a triplet (f, (ξa), (ωa)), a = 1, . . . , s, where f is a tensor field of type (1, 1), (ξa) are vector fields and (ωa) are 1-forms on N such that

(3.1) ωab) = δba; f (ξa) = 0 ; ωa◦ f = 0 ; f2= −I +X

a

ωa⊗ ξa, where I denotes the Kronecker tensor field on N . The name of f (3, 1)- structure was suggested by the identity f3+ f = 0. For an account of this kind of structures we refer to [10].

The linear operator φ given in the local adapted basis by (3.2) φ(δα) =∂.α; φ(∂.α) = −δα

defines an almost complex structure on Q(Q) and it is easy to see that (3.3) G(φ(X), φ(Y )) = G(X, Y ), ∀ X, Y ∈ Γ(Q(Q)).

(6)

Let us put ξ1 = yα

FQ

2δα and ξ2 = yα

FQ

2 .

α. From the definition of φ, it follows:

Proposition 3.1. We have φ(ξ1) = ξ2 and φ(ξ2) = −ξ1. Now, let us consider the 1-forms ω1 = yα

FQ

2dxα and ω2 = yα

FQ

2δyα, where yα = Gαβyβ. By the second equality of (1.2) we have yαyα = 2GQ

and so ωab) = δba. By a direct calculation, we obtain Proposition 3.2. We have ω1◦ φ = −ω2 and ω2◦ φ = ω1.

Proposition 3.3. We have ω1(X) = G(X, ξ1) and ω2(X) = G(X, ξ2), for any X ∈ Γ(Q(Q)).

Now, we define a tensor field f of type (1, 1) on Q(Q) by (3.4) f (X) = φ(X) − ω1(X)ξ2+ ω2(X)ξ1

for any X ∈ Γ(Q(Q)).

Theorem 3.1. The triplet (f, (ξa), (ωa)), a = 1, 2 provides a framed f (3, 1)- structure on Q(Q), namely

(i) ωab) = δab ; f (ξa) = 0 ; ωa◦ f = 0;

(ii) f2(X) = −X + ω1(X)ξ1+ ω2(X)ξ2, for any X ∈ Γ(Q(Q));

(iii) f is of rank 2n − 2 and f3+ f = 0.

Proof. Using (3.4) and Propositions 3.1 and 3.2, by direct calculations we get (i) and (ii). Applying f on the equality (ii) and taking into account the equality (i), we obtain f3+ f = 0. Now, from the second equations in (i), we see that span{ξ1, ξ2} ⊆ ker f . If X = Xαδα+ ˙Xα .α belongs to ker f and it is not in span{ξ1, ξ2}, by using (3.4), we have

f (X) =



Xβ−Xαyα

2GQ

yβ

 .

β − X˙β−X˙αyα

2GQ

yβ

!

δβ = 0.

Thus, X = Xαyα

FQ

2ξ1+ X˙

αyα

FQ

2ξ2 ∈ span{ξ1, ξ2}, contradiction. Hence ker f =

span{ξ1, ξ2} and rankf = 2n − 2. 

Theorem 3.2. The Riemannian metric G verifies

(3.5) G(f (X), f (Y )) = G(X, Y ) − ω1(X)ω1(Y ) − ω2(X)ω2(Y ) for any X, Y ∈ Γ(Q(Q)).

Proof. Since G(ξ1, ξ2) = 0 and G(ξ1, ξ1) = G(ξ2, ξ2) = 1, by using (3.4) and

Propositions 3.2 and 3.3 we get (3.5). 

Remark 3.1. In the local basis {δα,∂.α}, we have (3.6) f (δα) =



δαβ−yαyβ 2GQ

 .

β; f (∂.α) =



−δαβ+yαyβ 2GQ

 δβ

(7)

and using (3.6) one finds

(3.7)

G(f (δα), f (δβ)) = Gαβ−yαyβ 2GQ

G(f (δα), f (∂.β)) = 0

G(f (∂.α), f (∂.β)) = Gαβ−yαyβ

2GQ. Now, from (3.7) easily follows (3.5).

Theorem 3.2 says that (f, G) is a Riemannian framed f (3, 1)-structure on Q(Q).

Let us put ϕ(X, Y ) = G(f (X), Y ) for any X, Y ∈ Γ(Q(Q)). We have Theorem 3.3. ϕ is a 2-form on Q(Q) and the annihilator of ϕ is spanned by {ξ1, ξ2}.

Proof. ϕ is bilinear since G is bilinear. Now, using Proposition 3.3 and Theorems 3.1 and 3.2, by direct calculations we have ϕ(Y, X) = −ϕ(X, Y ) which says that ϕ is a 2-form on Q(Q).

By the second equality (i) from Theorem 3.1 it follows that ϕ(ξ1, ξ1) = ϕ(ξ1, ξ2) = ϕ(ξ2, ξ2) = 0, hence span{ξ1, ξ2} is contained in the null space of ϕ. Conversely, if X = Xαδα+ ˙Xα .α∈ Γ(Q(Q)) such that ϕ(X, X) = 0, by direct calculations, we get X = Xαyα

FQ

2ξ1+X˙

αyα

FQ

2ξ2 ∈ span{ξ1, ξ2}.  Remark 3.2. Locally, we have

(3.8) ϕ =



Gαβ −yαyβ 2GQ



dxα∧ δyβ



Gαβ−yαyβ 2GQ



δyα∧ dxβ and it appears as a deformation of the symplectic structure ϕ0(X, Y ) = G(φX, Y ) for any X, Y ∈ Γ(Q(Q)).

Finally, we make a similar study concerning framed f (3, −1)-structures on Q(Q).

A framed f (3, −1)-structure of corank s on a (2n + s)-dimensional mani- fold N is a natural generalization of an almost paracontact structure on N and it consists of a triplet ( ef , (ξa), (ωa)), a = 1, . . . , s, where ef is a tensor field of type (1, 1), (ξa) are vector fields and (ωa) are 1-forms on N such that

(3.9) ωab) = δba; ef (ξa) = 0 ; ωa◦ ef = 0 ; ef2 = I −X

a

ωa⊗ ξa. The name of f (3, −1)-structure was suggested by the identity ef3− ef = 0.

This is in some sense dual to the framed f (3, 1)-structure on N .

Let us consider the linear operator eφ given in the local adapted basis by (3.10) φ(δe α) = δα; eφ(∂.α) = −∂.α.

(8)

It is easy to check that eφ2 = I and

(3.11) G( eφ(X), eφ(Y )) = G(X, Y )

for any X, Y ∈ Γ(Q(Q)). From the definition of eφ it follows Proposition 3.4. We have eφ(ξ1) = ξ1 and eφ(ξ2) = −ξ2. Proposition 3.5. We have ω1◦ eφ = ω1 and ω2◦ eφ = −ω2.

Now, we define a tensor field ef of type (1, 1) on Q(Q) by (3.12) f (X) = ee φ(X) − ω1(X)ξ1+ ω2(X)ξ2

for any X ∈ Γ(Q(Q)).

Theorem 3.4. The triplet ( ef , (ξa), (ωa)), a = 1, 2 provides a framed f (3, −1)-structure on Q(Q), namely

(i) ωab) = δab; ef (ξa) = 0; ωa◦ ef = 0;

(ii) ef2(X) = X − ω1(X)ξ1− ω2(X)ξ2, for any X ∈ Γ(Q(Q));

(iii) ef is of rank 2n − 2 and ef3− ef = 0.

Proof. Using (3.12) and Propositions 3.4 and 3.5, by direct calculations we get (i) and (ii). Applying ef on the equality (ii) and taking into account the equality (i), we obtain ef3− ef = 0. The equality (iii) follows by the same

argument as in the proof of Theorem 3.1. 

Theorem 3.5. The Riemannian metric G verifies

(3.13) G( ef (X), ef (Y )) = G(X, Y ) − ω1(X)ω1(Y ) − ω2(X)ω2(Y ) for any X, Y ∈ Γ(Q(Q)).

Proof. It follows by direct calculations by using (3.12) and Propositions

3.3 and 3.5. 

Remark 3.3. In the local basis {δα,∂.α}, we have (3.14) f (δe α) =



δαβ−yαyβ 2GQ



δβ; ef (∂.α) =



−δβα+ yαyβ 2GQ

 .

β

and using (3.14) one finds

(3.15)

G( ef (δα), ef (δβ)) = Gαβ−yαyβ

2GQ G( ef (δα), ef (∂.β)) = 0

G( ef (∂.α), ef (∂.β)) = Gαβ−yαyβ

2GQ. Now, (3.13) easily follows from (3.15).

(9)

Theorem 3.5. says that ( ef , G) is a Riemannian framed f (3, −1)-structure on Q(Q).

Let us putϕ(X, Y ) = G( ee f (X), Y ) for any X, Y ∈ Γ(Q(Q)). We have Theorem 3.6. ϕ is a symmetric bilinear form on Q(Q) and the annihilatore of ϕ is spanned by {ξe 1, ξ2}.

Proof. It follows in a similar manner with the proof of Theorem 3.3, by using Proposition 3.3 and Theorems 3.4 and 3.5. 

Locally, we have (3.16) ϕ =e



Gαβ−yαyβ 2GQ



dxα⊗ dxβ



Gαβ −yαyβ 2GQ



δyα⊗ δyβ with det(Gαβy2Gαyβ

Q) = 0, since (Gαβy2Gαyβ

Q)yβ = yα− yα= 0.

Remark 3.4. The mapϕ is a singular pseudo-Riemannian metric on Q(Q).e Acknowledgement. The author is thankful to the anonymous referee for valuable comments which improved this paper.

References

[1] Abate, M., Patrizio, G., Finsler Metrics – A Global Approach, Lecture Notes in Math., 1591, Springer-Verlag, Berlin, 1994.

[2] Anastasiei, M., A framed f -structure on tangent bundle of a Finsler space, An. Univ.

Bucure¸sti, Mat.-Inf., 49 (2000), 3–9.

[3] Bao, D., Chern, S. S. and Shen, Z., An Introduction to Riemannian Finsler Geometry, Graduate Texts in Math., 200, Springer-Verlag, New York, 2000.

[4] Bejancu, A., Farran, H. R., On the vertical bundle of a pseudo-Finsler manifold, Int.

J. Math. Math. Sci. 22 (3) (1997), 637–642.

[5] Gˆırt¸u, M., An almost paracontact structure on the indicatrix bundle of a Finsler space, Balkan J. Geom. Appl. 7(2) (2002), 43–48.

[6] Gˆırt¸u, M., A framed f (3, −1)-structure on the tangent bundle of a Lagrange space, Demonstratio Math. 37(4) (2004), 955–961.

[7] Hasegawa, I., Yamaguchi, K. and Shimada, H., Sasakian structures on Finsler man- ifolds, Antonelli, P. L., Miron R. (eds.), Lagrange and Finsler Geometry, Kluwer Acad. Publ., Dordrecht, 1996, 75–80.

[8] Miernowski, A., A note on transversally Finsler foliations, Ann. Univ. Mariae Curie- Skłodowska Sect. A 60 (2006), 57–64.

[9] Miernowski, A., Mozgawa, W., Lift of the Finsler foliations to its normal bundle, Differential Geom. Appl. 24 (2006), 209–214.

[10] Mihai, I., Ro¸sca, R. and Verstraelen, L., Some aspects of the differential geometry of vector fields, PADGE, Katholieke Univ. Leuven, vol. 2 (1996).

[11] Miron, R., Anastasiei, M., The Geometry of Lagrange Spaces: Theory and Applica- tions, Kluwer Acad. Publ., Dordrecht, 1994.

[12] Popescu, P., Popescu, M., Lagrangians adapted to submersions and foliations, Dif- ferential Geom. Appl. 27 (2009), 171–178.

[13] Singh, K. D., Singh, R., Some f (3, ε)-structure manifold, Demonstratio Math. 10 (3–4) (1977), 637–645.

(10)

[14] Vaisman, I., Lagrange geometry on tangent manifolds, Int. J. Math. Math. Sci. 51 (2003), 3241–3266.

[15] Yano, K., On a structure defined by a tensor field of type (1, 1) satisfying f3+ f = 0, Tensor (N.S.) 14 (1963), 99–109.

Cristian Ida

Department of Algebra, Geometry and Differential Equations Transilvania University of Bra¸sov Bra¸sov 500091, Str. Iuliu Maniu 50 Romˆania

e-mail: cristian.ida@unitbv.ro Received April 23, 2010

Cytaty

Powiązane dokumenty

Application of a linear Padé approximation In a similar way as for standard linear systems Kaczorek, 2013, it can be easily shown that if sampling is applied to the

The larger segment W is also a twisted square based prism, but the sections W t are obtained by rotating the base with angle velocity φ/2 over the t-interval [0, 2π/φ]... By

Therefore, Theorem 4.3 may be generalized to all line graphs of multigraphs which possess maximal matchable subsets of vertices – for example, the line graphs of multigraphs

However, as was shown by Mioduszewski (1961), this involution, when restricted to any arc of S, has at most one discontinuity point and becomes continuous if we change the value φ(x)

In the proof of this theorem, the key role is played by an effective interpretation of the well-known fact that an irreducible polynomial which is reducible over the algebraic

We prove that, for every γ ∈ ]1, ∞[, there is an element of the Gevrey class Γ γ which is analytic on Ω, has F as its set of defect points and has G as its set of

This descriptive definition of our integral is then used to show that our process of integration extends the one of Lebesgue, and to establish a quite general divergence theorem..

Another class of Dirac structures related to Poisson reduction and dual pairs of Poisson manifolds, called null Dirac structures, is characterized in the following way..