• Aucun résultat trouvé

Pseudo-Riemannian metrics on bicovariant bimodules

N/A
N/A
Protected

Academic year: 2022

Partager "Pseudo-Riemannian metrics on bicovariant bimodules"

Copied!
23
0
0

Texte intégral

(1)

BLAISE PASCAL

Jyotishman Bhowmick & Sugato Mukhopadhyay Pseudo-Riemannian metrics on bicovariant bimodules Volume 27, no2 (2020), p. 159-180.

<http://ambp.centre-mersenne.org/item?id=AMBP_2020__27_2_159_0>

Cet article est mis à disposition selon les termes de la licence Creative Commons attribution 4.0.

https://creativecommons.org/licenses/4.0/

L’accès aux articles de la revue « Annales mathématiques Blaise Pascal » (http://ambp.centre-mersenne.org/), implique l’accord avec les conditions générales d’utilisation (http://ambp.centre-mersenne.org/legal/).

Publication éditée par le laboratoire de mathématiques Blaise Pascal de l’université Clermont Auvergne, UMR 6620 du CNRS

Clermont-Ferrand — France

Publication membre du

(2)

Pseudo-Riemannian metrics on bicovariant bimodules

Jyotishman Bhowmick Sugato Mukhopadhyay

Abstract

We study pseudo-Riemannian invariant metrics on bicovariant bimodules over Hopf algebras. We clarify some properties of such metrics and prove that pseudo-Riemannian invariant metrics on a bicovariant bimodule and its cocycle deformations are in one to one correspondence.

1. Introduction

The notion of metrics on covariant bimodules on Hopf algebras have been studied by a number of authors including Heckenberger and Schmüdgen ([6, 7, 8]) as well as Beggs, Majid and their collaborators ([11] and references therein). The goal of this article is to characterize bicovariant pseudo-Riemannian metrics on a cocycle-twisted bicovariant bimodule. As in [8], the symmetry of the metric comes from Woronowicz’s braiding map𝜎on the bicovariant bimodule. However, since our notion of non-degeneracy of the metric is slightly weaker than that in [8], we consider a slightly larger class of metrics than those in [8]. The positive-definiteness of the metric does not play any role in what we do.

We refer to the later sections for the definitions of pseudo-Riemannian metrics and cocycle deformations. Our strategy is to exploit the covariance of the various maps between bicovariant bimodules to view them as maps between the finite-dimensional vector spaces of left-invariant elements of the respective bimodules. This was already observed and used crucially by Heckenberger and Schmüdgen in the paper [8]. We prove that bi- invariant pseudo-Riemannian metrics are automatically bicovariant maps and compare our definition of pseudo-Riemannian metric with some of the other definitions available in the literature. Finally, we prove that the pseudo-Riemannian bi-invariant metrics on a bicovariant bimodule and its cocycle deformation are in one to one correspondence.

These results will be used in the companion article [2].

In Section 2, we discuss some generalities on bicovariant bimodules. In Section 3, we define and study pseudo-Riemannian left metrics on a bicovariant differential calculus.

Finally, in Section 4, we prove our main result on bi-invariant metrics on cocycle- deformations.

Let us set up some notations and conventions that we are going to follow. All vector spaces will be assumed to be over the complex field. For vector spaces 𝑉1 and 𝑉2,

Keywords:bicovariant bimodules, pseudo-Riemannian metric, cocycle deformations.

(3)

𝜎can :𝑉1C𝑉2 →𝑉2C𝑉1will denote the canonical flip map, i.e,𝜎can(𝑣1C𝑣2) = 𝑣2C𝑣1.For the rest of the article,(A,Δ)will denote a Hopf algebra. We will use the Sweedler notation for the coproductΔ. Thus, we will write

Δ(𝑎)=𝑎(1)C𝑎(2). (1.1) For a rightA-module𝑉 ,the notation𝑉will stand for the set HomA(𝑉 ,A).

Following [16], the comodule coaction on a leftA-comodule𝑉will be denoted by the symbolΔ𝑉.Thus,Δ𝑉 is aC-linear mapΔ𝑉 :𝑉→ A ⊗C𝑉such that

(Δ⊗Cid)Δ𝑉 =(id⊗CΔ𝑉𝑉, (𝜖⊗Cid)Δ𝑉(𝑣)=𝑣 for all𝑣in𝑉(here𝜖is the counit ofA). We will use the notation

Δ𝑉(𝑣)=𝑣(−1)C𝑣(0). (1.2) Similarly, the comodule coaction on a rightA-comodule will be denoted by𝑉Δand we will write

𝑉Δ(𝑣)=𝑣(0)C𝑣(1). (1.3) Finally, for a Hopf algebraA,AAMA,AMAA,A

AMAA will denote the categories of various types of mixed Hopf-bimodules as in [13, Subsection 1.9].

2. Covariant bimodules on quantum groups

In this section we recall and prove some basic facts on covariant bimodules. These objects were studied by many Hopf-algebraists (as Hopf-bimodules) including Abe ([1]) and Sweedler ([15]). During the 1980’s, they were re-introduced by Woronowicz ([16]) for studying differential calculi over Hopf algebras. Schauenburg ([14]) proved a categorical equivalence between bicovariant bimodules and Yetter–Drinfeld modules over a Hopf algebra, the latter being introduced by Yetter in [17].

We start by recalling the notions on covariant bimodules from Section 2 of [16].

Suppose 𝑀 is a bimodule over A such that (𝑀 ,Δ𝑀) is a left A-comodule. Then (𝑀 ,Δ𝑀)is called a left-covariant bimodule if this tuplet is an object of the category

A

AMA, i.e, for all𝑎inAand𝑚in𝑀, the following equation holds.

Δ𝑀(𝑎𝑚)= Δ(𝑎)Δ𝑀(𝑚), Δ𝑀(𝑚 𝑎)= Δ𝑀(𝑚)Δ(𝑎).

Similarly, if 𝑀Δis a right comodule coaction on𝑀 , then (𝑀 ,𝑀Δ) is called a right covariant bimodule if it is an object of the categoryAMAA, i.e, for any𝑎inAand𝑚 in𝑀,

𝑀Δ(𝑎𝑚)= Δ(𝑎)𝑀Δ(𝑚), 𝑀Δ(𝑚 𝑎)=𝑀Δ(𝑚)Δ(𝑎).

(4)

Finally, let𝑀be a bimodule overAandΔ𝑀 :𝑀→ A ⊗C𝑀 and𝑀Δ:𝑀 →𝑀 ⊗CA beC-linear maps. Then we say that(𝑀 ,Δ𝑀,𝑀Δ)is a bicovariant bimodule if this triplet is an object ofAAMAA.Thus,

(i) (𝑀 ,Δ𝑀)is left-covariant bimodule, (ii) (𝑀 ,𝑀Δ)is a right-covariant bimodule, (iii) (id⊗C𝑀Δ)Δ𝑀 =(Δ𝑀Cid)𝑀Δ.

The vector space of left (respectively, right) invariant elements of a left (respectively, right) covariant bimodules will play a crucial role in the sequel and we introduce notations for them here.

Definition 2.1. Let (𝑀 ,Δ𝑀) be a left-covariant bimodule overA. The subspace of left-invariant elements of𝑀is defined to be the vector space

0𝑀 :={𝑚∈𝑀 :Δ𝑀(𝑚)=1⊗C𝑚}.

Similarly, if(𝑀 ,𝑀Δ)is a right-covariant bimodule overA, the subspace of right-invariant elements of𝑀 is the vector space

𝑀0:={𝑚∈𝑀 :𝑀Δ(𝑚)=𝑚⊗C1}.

Remark 2.2. We will say that a bicovariant bimodule(𝑀 ,Δ𝑀,𝑀Δ)is finite if0𝑀 is a finite dimensional vector space. Throughout this article, we will only work with bicovariant bimodules which are finite.

Let us note the immediate consequences of the above definitions.

Lemma 2.3([16, Theorem 2.4]). Suppose𝑀is a bicovariantA-A-bimodule. Then

𝑀Δ(0𝑀) ⊆0𝑀⊗CA, Δ𝑀(𝑀0) ⊆ A ⊗C𝑀0.

More precisely, if{𝑚𝑖}𝑖is a basis of0𝑀 ,then there exist elements{𝑎𝑗 𝑖}𝑖 , 𝑗inAsuch that

𝑀Δ(𝑚𝑖)=∑︁

𝑗

𝑚𝑗C𝑎𝑗 𝑖. (2.1)

Proof. This is a simple consequence of the fact that𝑀Δcommutes withΔ𝑀. The categoryAAMAhas a natural monoidal structure. Indeed, if(𝑀 ,Δ𝑀)and(𝑁 ,Δ𝑁) are left-covariant bimodules overA,then we have a left coactionΔ𝑀A𝑁 ofAon𝑀⊗A𝑁 defined by the following formula:

Δ𝑀A𝑁(𝑚⊗A𝑛)=𝑚(−1)𝑛(−1)C𝑚(0)A𝑛(0).

(5)

Here, we have made use of the Sweedler notation introduced in (1.2). This makes𝑀⊗A𝑁 into a left covariant A-A-bimodule. Similarly, there is a right coaction 𝑀A𝑁Δon 𝑀 ⊗A 𝑁if(𝑀 ,𝑀Δ)and(𝑁 ,𝑁Δ)are right covariant bimodules.

The fundamental theorem of Hopf modules (Theorem 1.9.4 of [13]) states that if𝑉 is a left-covariant bimodule overA,then𝑉is a free as a left (as well as a right)A-module.

This was reproved by Woronowicz in [16]. In fact, one has the following result:

Proposition 2.4([16, Theorem 2.1 and Theorem 2.3]). Let(𝑀 ,Δ𝑀)be a bicovariant bimodule over A. Then the multiplication maps 0𝑀 ⊗CA → 𝑀, A ⊗C0𝑀 → 𝑀, 𝑀0CA →𝑀andA ⊗C𝑀0→ 𝑀are isomorphisms.

Corollary 2.5. Let(𝑀 ,Δ𝑀)and(𝑁 ,Δ𝑁)be left-covariant bimodules overAand{𝑚𝑖}𝑖

and{𝑛𝑗}𝑗be vector space bases of0𝑀and0𝑁respectively. Then each element of𝑀⊗A𝑁 can be written asÍ

𝑖 𝑗𝑎𝑖 𝑗𝑚𝑖A𝑛𝑗andÍ

𝑖 𝑗𝑚𝑖A𝑛𝑗𝑏𝑖 𝑗, where𝑎𝑖 𝑗and𝑏𝑖 𝑗are uniquely determined.

A similar result holds for right-covariant bimodules(𝑀 ,𝑀Δ)and(𝑁 ,𝑁Δ)overA.

Finally, if(𝑀 ,Δ𝑀)is a left-covariant bimodule overAwith basis{𝑚𝑖}𝑖 of0𝑀, and (𝑁 ,𝑁Δ)is a right-covariant bimodule overAwith basis{𝑛𝑗}𝑗 of𝑁0, then any element of𝑀⊗A 𝑁can be written uniquely asÍ

𝑖 𝑗𝑎𝑖 𝑗𝑚𝑖A𝑛𝑗 as well asÍ

𝑖 𝑗𝑚𝑖A𝑛𝑗𝑏𝑖 𝑗. Proof. The proof of this result is an adaptation of [16, Lemma 3.2] and we omit it.

The next proposition will require the definition of right Yetter–Drinfeld modules for which we refer to [17] and [14, Definition 4.1].

Proposition 2.6 ([14, Theorem 5.7]). The functor 𝑀 ↦→ 0𝑀 induces a monoidal equivalence of categories AAMAA and the category of right Yetter–Drinfeld modules.

Therefore, if(𝑀 ,Δ𝑀)and(𝑁 ,Δ𝑀)be left-covariant bimodules overA,then

0(𝑀⊗A 𝑁)=spanC{𝑚⊗A𝑛:𝑚∈0𝑀 , 𝑛∈0𝑁}. (2.2) Similarly, if(𝑀 ,𝑀Δ)and(𝑁 ,𝑁Δ)are right-covariant bimodules overA, then we have that

(𝑀⊗A 𝑁)0=spanC{𝑚⊗A𝑛:𝑚∈𝑀0, 𝑛∈𝑁0}. Thus,0(𝑀 ⊗A 𝑁)=0𝑀⊗C0𝑁and(𝑀⊗A𝑁)0=𝑀0C𝑁0.

Remark 2.7. In the light of Proposition 2.6, we are allowed to use the notations0𝑀⊗C0𝑁 and0(𝑀⊗A𝑁)interchangeably.

We recall now the definition of covariant maps between bimodules.

(6)

Definition 2.8. Let(𝑀 ,Δ𝑀,𝑀Δ)and(𝑁 ,Δ𝑁,𝑁Δ)be bicovariantA-bimodules and𝑇 be aC-linear map from𝑀 to𝑁 .

𝑇is called left-covariant if𝑇 is a morphism in the categoryAM, i.e, for all𝑚∈𝑀 , (id⊗C𝑇) (Δ𝑀(𝑚))= Δ𝑁(𝑇(𝑚)).

𝑇is called right-covariant if𝑇 is a morphism in the categoryMA.Thus, for all𝑚∈𝑀 , (𝑇 ⊗Cid)𝑀Δ(𝑚)=𝑁Δ(𝑇(𝑚)).

Finally, a map which is both left and right covariant will be called a bicovariant map. In other words, a bicovariant map is a morphism in the categoryAMA.

We end this section by recalling the following fundamental result of Woronowicz.

Proposition 2.9([16, Proposition 3.1]). Given a bicovariant bimoduleEthere exists a unique bimodule homomorphism

𝜎:E ⊗A E → E ⊗A E such that𝜎(𝜔⊗A𝜂)=𝜂⊗A𝜔 (2.3) for any left-invariant element𝜔and right-invariant element𝜂inE.𝜎is invertible and is a bicovariantA-bimodule map fromE ⊗AEto itself. Moreover,𝜎satisfies the following braid equation onE ⊗A E ⊗A E:

(id⊗A𝜎) (𝜎⊗Aid) (id⊗A𝜎)=(𝜎⊗Aid) (id⊗A𝜎) (𝜎⊗Aid). 3. Pseudo-Riemannian metrics on bicovariant bimodules

In this section, we study pseudo-Riemannian metrics on bicovariant differential calculus on Hopf algebras. After defining pseudo-Riemannian metrics, we recall the definitions of left and right invariance of a pseudo-Riemannian metrics from [8]. We prove that a pseudo-Riemannian metric is left (respectively, right) invariant if and only if it is left (respectively, right) covariant. The coefficients of a left-invariant pseudo-Riemannian metric with respect to a left-invariant basis ofEare scalars. We use this fact to clarify some properties of pseudo-Riemannian invariant metrics. We end the section by comparing our definition with those by Heckenberger and Schmüdgen ([8]) as well as by Beggs and Majid.

Definition 3.1([8]).SupposeEis a bicovariantA-bimoduleEand𝜎:E ⊗AE → E ⊗AE be the map as in Proposition 2.9. A pseudo-Riemannian metric for the pair(E, 𝜎)is a rightA-linear map𝑔:E ⊗A E → Asuch that the following conditions hold:

(i) 𝑔◦𝜎=𝑔.

(ii) If𝑔(𝜌⊗A𝜈)=0 for all𝜈inE,then𝜌=0.

(7)

For other notions of metrics on covariant differential calculus, we refer to [11] and references therein.

Definition 3.2([8]). A pseudo-Riemannian metric𝑔on a bicovariantA-bimoduleEis said to be left-invariant if for all𝜌, 𝜈inE,

(id⊗C𝜖 𝑔) (Δ( E ⊗AE)(𝜌⊗A 𝜈))=𝑔(𝜌⊗A𝜈).

Similarly, a pseudo-Riemannian metric𝑔on a bicovariantA-bimoduleEis said to be right-invariant if for all𝜌, 𝜈inE,

(𝜖 𝑔⊗Cid) (( E ⊗AE)Δ(𝜌⊗A 𝜈))=𝑔(𝜌⊗A 𝜈).

Finally, a pseudo-Riemannian metric𝑔on a bicovariantA-AbimoduleEis said to be bi-invariant if it is both left-invariant as well as right-invariant.

We observe that a pseudo-Riemannian metric is invariant if and only if it is covariant.

Proposition 3.3. Let𝑔be a pseudo-Riemannian metric on the bicovariant bimoduleE.

Then𝑔is left-invariant if and only if𝑔:E ⊗AE → Ais a left-covariant map. Similarly, 𝑔is right-invariant if and only if𝑔:E ⊗AE → Ais a right-covariant map.

Proof. Let𝑔be a left-invariant metric onE, and𝜌,𝜈be elements ofE. Then the following computation shows that𝑔is a left-covariant map.

Δ𝑔(𝜌⊗A 𝜈)= Δ( (id⊗C𝜖 𝑔) (Δ( E ⊗AE)(𝜌⊗A𝜈)))

= Δ(id⊗C𝜖 𝑔) (𝜌(−1)𝜈(−1)C𝜌(0)A𝜈(0))

= Δ(𝜌(−1)𝜈(−1))𝜖 𝑔(𝜌(0)A 𝜈(0))

=(𝜌(−1))(1)(𝜈(−1))(1)C (𝜌(−1))(2)(𝜈(−1))(2)𝜖 𝑔(𝜌(0)A 𝜈(0))

=(𝜌(−1))(1)(𝜈(−1))(1)C( (id⊗C𝜖 𝑔) ( (𝜌(−1))(2)(𝜈(−1))(2)C𝜌(0)A𝜈(0)))

=𝜌(−1)𝜈(−1)C( (id⊗C𝜖 𝑔) (Δ( E ⊗AE)(𝜌(0)A𝜈(0))))

(where we have used co associativity of comodule coactions)

=𝜌(−1)𝜈(−1)C𝑔(𝜌(0)A𝜈(0))

=(id⊗C𝑔) (Δ( E ⊗AE)(𝜌⊗A𝜈)).

On the other hand, suppose𝑔 :E ⊗A E → Ais a left-covariant map. Then we have (id⊗C𝜖 𝑔)Δ( E ⊗AE)(𝜌⊗A𝜈)=(id⊗C𝜖) (id⊗C𝑔)Δ( E ⊗AE)(𝜌⊗A 𝜈)

=(id⊗C𝜖)Δ𝑔(𝜌⊗A𝜈)=𝑔(𝜌⊗A𝜈).

The proof of the right-covariant case is similar.

The following key result will be used throughout the article.

(8)

Lemma 3.4 ([8]). If 𝑔 is a pseudo-Riemannian metric which is left-invariant on a left-covariantA-bimoduleE, then𝑔(𝜔1A𝜔2) ∈C.1for all𝜔1, 𝜔2in0E.Similarly, if 𝑔is a right-invariant pseudo-Riemannian metric on a right-covariantA-bimodule, then 𝑔(𝜂1A𝜂2) ∈C.1for all𝜂1, 𝜂2inE0.

Let us clarify some of the properties of a left-invariant and right-invariant pseudo- Riemannian metrics. To that end, we make the next definition which makes sense as we always work with finite bicovariant bimodules (see Remark 2.2). The notations used in the next definition will be used throughout the article.

Definition 3.5. LetEand𝑔be as above. For a fixed basis{𝜔1, . . . , 𝜔𝑛}of0E,we define 𝑔𝑖 𝑗 =𝑔(𝜔𝑖A𝜔𝑗).Moreover, we define𝑉𝑔 :E → E =HomA(E,A)to be the map defined by the formula

𝑉𝑔(𝑒) (𝑓)=𝑔(𝑒⊗A 𝑓).

Proposition 3.6. Let 𝑔 be a left-invariant pseudo-Riemannian metric for E as in Definition 3.1. Then the following statements hold:

(i) The map𝑉𝑔is a one-one rightA-linear map fromEtoE.

(ii) If𝑒∈ Eis such that𝑔(𝑒⊗A 𝑓)=0for all 𝑓 ∈0E,then𝑒=0.In particular, the map𝑉𝑔|0Eis one-one and hence an isomorphism from0Eto(0E).

(iii) The matrix( (𝑔𝑖 𝑗))𝑖 𝑗is invertible.

(iv) Let𝑔𝑖 𝑗denote the(𝑖, 𝑗)-th entry of the inverse of the matrix( (𝑔𝑖 𝑗))𝑖 𝑗. Then𝑔𝑖 𝑗 is an element ofC.1for all𝑖, 𝑗.

(v) If𝑔(𝑒⊗A 𝑓)=0for all𝑒in0E, then 𝑓 =0.

Proof. The rightA-linearity of𝑉𝑔 follows from the fact that𝑔is a well-defined map fromE ⊗A EtoA.The condition (ii) of Definition 3.1 forces𝑉𝑔 to be one-one. This proves (i).

For proving (ii), note that𝑉𝑔|0E is the restriction of a one-one map to a subspace.

Hence, it is a one-oneC-linear map. Since𝑔is left-invariant, by Lemma 3.4, for any𝑒in

0E,𝑉𝑔(𝑒) (0E)is contained inC. Therefore,𝑉𝑔maps0Einto(0E). Since,0Eand(0E) have the same finite dimension as vector spaces,𝑉𝑔|0E :0E → (0E)is an isomorphism.

This proves (ii).

Now we prove (iii). Let {𝜔𝑖}𝑖 be a basis of 0E and {𝜔

𝑖}𝑖 be a dual basis, i.e, 𝜔

𝑖(𝜔𝑗)=𝛿𝑖 𝑗.Since𝑉𝑔|0Eis a vector space isomorphism from0Eto(0E)by part (ii),

(9)

there exist complex numbers𝑎𝑖 𝑗such that (𝑉𝑔)−1(𝜔

𝑖)=∑︁

𝑗

𝑎𝑖 𝑗𝜔𝑗.

This yields

𝛿𝑖 𝑘=𝜔

𝑖(𝜔𝑘)=𝑔

∑︁

𝑗

𝑎𝑖 𝑗𝜔𝑗A 𝜔𝑘

!

=∑︁

𝑗

𝑎𝑖 𝑗𝑔𝑗 𝑘.

Therefore,( (𝑎𝑖 𝑗))𝑖 𝑗is the left-inverse and hence the inverse of the matrix( (𝑔𝑖 𝑗))𝑖 𝑗. This proves (iii).

For proving (iv), we use the fact that𝑔𝑖 𝑗is an element ofC.1 for all𝑖, 𝑗. Since

∑︁

𝑘

𝑔(𝜔𝑖A𝜔𝑘)𝑔𝑘 𝑗 =𝛿𝑖 𝑗.1=∑︁

𝑘

𝑔𝑖 𝑘𝑔(𝜔𝑘A𝜔𝑗)=𝛿𝑖 𝑗,

we have

∑︁

𝑘

𝑔(𝜔𝑖A 𝜔𝑘)𝜖(𝑔𝑘 𝑗)=𝛿𝑖 𝑗 =∑︁

𝑘

𝜖(𝑔𝑖 𝑘)𝑔(𝜔𝑘A𝜔𝑗).

So, the matrix( (𝜖(𝑔𝑖 𝑗)))𝑖 𝑗is also an inverse to the matrix( (𝑔(𝜔𝑖A𝜔𝑗)))𝑖 𝑗and hence 𝑔𝑖 𝑗=𝜖(𝑔𝑖 𝑗)and𝑔𝑖 𝑗is inC.1.

Finally, we prove (v) using (iv). Suppose 𝑓 be an element inEsuch that𝑔(𝑒⊗A 𝑓)=0 for all𝑒in0E. Let 𝑓 =Í

𝑘𝜔𝑘𝑎𝑘for some elements𝑎𝑘inA. Then for any fixed index𝑖0, we obtain

0=𝑔

∑︁

𝑗

𝑔𝑖0𝑗𝜔𝑗A

∑︁

𝑘

𝜔𝑘𝑎𝑘

!

=∑︁

𝑘

∑︁

𝑗

𝑔𝑖0𝑗𝑔𝑗 𝑘𝑎𝑘 =∑︁

𝑘

𝛿𝑖

0𝑘𝑎𝑘 =𝑎𝑖

0.

Hence, we have that 𝑓 =0. This finishes the proof.

We apply the results in Proposition 3.6 to exhibit a basis of the free rightA-module 𝑉𝑔(E).This will be used in making Definition 4.11 which is needed to prove our main Theorem 4.15.

Lemma 3.7. Suppose{𝜔𝑖}𝑖 is a basis of0Eand{𝜔

𝑖}𝑖be the dual basis as in the proof of Proposition 3.6. If𝑔is a pseudo-Riemannian left-invariant metric onE,then𝑉𝑔(E)is a free rightA-module with basis{𝜔

𝑖}𝑖.

Proof. We will use the notations(𝑔𝑖 𝑗)𝑖 𝑗and𝑔𝑖 𝑗from of Proposition 3.6. Since𝑉𝑔is a rightA-linear map,𝑉𝑔(E)is a rightA-module. Since

𝑉𝑔(𝜔𝑖)=∑︁

𝑗

𝑔𝑖 𝑗𝜔𝑗 (3.1)

(10)

and the inverse matrix(𝑔𝑖 𝑗)𝑖 𝑗has scalar entries (Proposition 3.6), we get 𝜔

𝑘 =∑︁

𝑖

𝑔𝑘 𝑖𝑉𝑔(𝜔𝑖)

and so𝜔

𝑘belongs to𝑉𝑔(E)for all𝑘 .By the rightA-linearity of the map𝑉𝑔,we conclude that the set{𝜔

𝑖}𝑖 is rightA-total in𝑉𝑔(E). Finally, if𝑎𝑖are elements inAsuch thatÍ

𝑘𝜔

𝑘𝑎𝑘 =0,then by (3.1), we have 0=∑︁

𝑖 , 𝑘

𝑔𝑘 𝑖𝑉𝑔(𝜔𝑖)𝑎𝑘 =𝑉𝑔

∑︁

𝑖

𝜔𝑖

∑︁

𝑘

𝑔𝑘 𝑖𝑎𝑘

! ! .

As𝑉𝑔is one-one and{𝜔𝑖}𝑖 is a basis of the free moduleE,we get

∑︁

𝑘

𝑔𝑘 𝑖𝑎𝑘 =0 ∀ 𝑖 .

Multiplying by𝑔𝑖 𝑗and summing over𝑖yields𝑎𝑗 =0.This proves that{𝜔

𝑖}𝑖 is a basis of

𝑉𝑔(E)and finishes the proof.

Remark 3.8. Let us note that for all𝑒∈ E,the following equation holds:

𝑒=∑︁

𝑖

𝜔𝑖𝜔𝑖(𝑒). (3.2)

The following proposition was kindly pointed out to us by the referee for which we will need the notion of a left dual of an object in a monoidal category. We refer to Definition 2.10.1 of [5] or Definition XIV.2.1 of [9] for the definition.

Proposition 3.9. Suppose𝑔is a pseudo-RiemannianA-bilinear pseudo-Riemannian metric on a finite bicovariantA-bimodule. LeteEdenote the left dual of the objectEin the categoryAAMAA.ThenEeis isomorphic toEas objects in the categoryAAMAA via the morphism𝑉𝑔.

Proof. It is well-known thatEeandEare isomorphic objects in the categoryAAMAA. This follows by using the bicovariantA-bilinear maps

ev :E ⊗e AE −→ A; coev :A −→ E ⊗A E;e 𝜙⊗A𝑒↦−→𝜙(𝑒), 1↦−→∑︁

𝑖

𝜔𝑖A 𝜔

𝑖

We define ev𝑔 :E ⊗AE → Aand coev𝑔:A → E ⊗AEby the following formulas:

ev𝑔(𝑒⊗A 𝑓)=𝑔(𝑒⊗A 𝑓), coev𝑔(1)=∑︁

𝑖

𝜔𝑖A𝑉𝑔1(𝜔

𝑖).

(11)

Then since𝑔is both left and rightA-linear, ev𝑔and coev𝑔areA-A-bilinear maps. The bicovariance of𝑔 implies the bicovariance of ev𝑔 while the bicovariance of coev𝑔 = (id⊗A𝑉𝑔−1) ◦coev follows from the bicovariance of𝑉𝑔and coev.

Since the left dual of an object is unique upto isomorphism, we need to check the following identities for all𝑒inE:

(ev𝑔Aid) (id⊗Acoev𝑔) (𝑒)=𝑒, (id⊗Aev𝑔) (coev𝑔Aid) (𝑒)=𝑒 . But these follow by a simple computation using the fact that0Eis rightA-total inEand the identity (3.2).

From the above discussion, we have thatE andE are two left duals of the object E in the category AAMAA.Then by the proof of Proposition 2.10.5 of [5], we know that(ev𝑔AidE) (idEAcoev)is an isomorphism fromEtoE.But it can be easily checked that(ev𝑔AidE) (idEAcoev)=𝑉𝑔.This completes the proof.

Now we state a result on bi-invariant (i.e both left and right-invariant) pseudo- Riemannian metric.

Proposition 3.10. Let𝑔be a pseudo-Riemannian metric onEand the symbols{𝜔𝑖}𝑖, {𝑔𝑖 𝑗}𝑖 𝑗be as above. If

EΔ(𝜔𝑖)=∑︁

𝑗

𝜔𝑗C𝑅𝑗 𝑖 (3.3)

(see(2.1)), then𝑔is bi-invariant if and only if the elements𝑔𝑖 𝑗are scalar and 𝑔𝑖 𝑗=∑︁

𝑘 𝑙

𝑔𝑘 𝑙𝑅𝑘 𝑖𝑅𝑙 𝑗. (3.4)

Proof. Since𝑔is left-invariant, the elements𝑔𝑖 𝑗are inC.1. Moreover, the right-invariance of𝑔implies that𝑔is right-covariant (Proposition 3.3), i.e.

1⊗C𝑔𝑖 𝑗= Δ(𝑔𝑖 𝑗)=(𝑔⊗Aid)E ⊗AEΔ(𝜔𝑖C𝜔𝑗)

=(𝑔⊗Aid) ∑︁

𝑘 𝑙

𝜔𝑘A𝜔𝑙C𝑅𝑘 𝑖𝑅𝑙 𝑗

!

=1⊗C∑︁

𝑘 𝑙

𝑔𝑘 𝑙𝑅𝑘 𝑖𝑅𝑙 𝑗,

so that

𝑔𝑖 𝑗=∑︁

𝑘 𝑙

𝑔𝑘 𝑙𝑅𝑘 𝑖𝑅𝑙 𝑗. (3.5)

Conversely, if𝑔𝑖 𝑗 =𝑔(𝜔𝑖A𝜔𝑗)are scalars and (3.4) is satisfied, then𝑔is left-invariant and right-covariant. By Proposition 3.3,𝑔is right-invariant.

We end this section by comparing our definition of pseudo-Riemannian metrics with some of the other definitions available in the literature.

(12)

Proposition 3.6 shows that our notion of pseudo-Riemannian metric coincides with the rightA-linear version of a “symmetric metric” introduced in Definition 2.1 of [8] if we impose the condition of left-invariance.

Next, we compare our definition with the one used by Beggs and Majid in Proposition 4.2 of [10] (also see [11] and references therein). To that end, we need to recall the construction of the two forms by Woronowicz ([16]).

IfEis a bicovariantA-bimodule and𝜎be the map as in Proposition 2.9, Woronowicz defined the space of two forms as:

Ω2(A):=(E ⊗A E)

Ker(𝜎−1). The symbol∧will denote the quotient map

∧:E ⊗A E →Ω2(A). Thus,

Ker(∧)=Ker(𝜎−1). (3.6)

In Proposition 4.2 of [10], the authors define a metric on a bimoduleEover a (possibly) noncommutative algebraAas an elementℎofE ⊗AEsuch that∧(ℎ)=0.We claim that metrics in the sense of Beggs and Majid are in one to one correspondence with elements 𝑔∈HomA(E ⊗AE,A)(not necessarily left-invariant) such that𝑔◦𝜎=𝑔 .Thus, modulo the nondegeneracy condition (ii) of Definition 3.1, our notion of pseudo-Riemannian metric matches with the definition of metric by Beggs and Majid.

Indeed, if𝑔 ∈ HomA(E ⊗A E,A) as above and{𝜔𝑖}𝑖,is a basis of0E,then the equation𝑔◦𝜎=𝑔implies that

𝑔◦𝜎(𝜔𝑖A𝜔𝑗)=𝑔(𝜔𝑖A𝜔𝑗). However, by equation (3.15) of [16], we know that

𝜎(𝜔𝑖A𝜔𝑗)=∑︁

𝑘 ,𝑙

𝜎𝑘 𝑙

𝑖 𝑗𝜔𝑘A𝜔𝑙

for some scalars𝜎𝑘 𝑙

𝑖 𝑗.Therefore, we have

∑︁

𝑘 ,𝑙

𝜎𝑘 𝑙

𝑖 𝑗𝑔(𝜔𝑘A𝜔𝑙)=𝑔(𝜔𝑖A𝜔𝑗). (3.7) We claim that the elementℎ=Í

𝑖 , 𝑗𝑔(𝜔𝑖A𝜔𝑗)𝜔𝑖A𝜔𝑗 satisfies∧(ℎ)=0.Indeed, by virtue of (3.6), it is enough to prove that (𝜎−1) (ℎ) =0.But this directly follows from (3.7) using the leftA-linearity of𝜎 .

This argument is reversible and hence starting fromℎ∈ E ⊗A Esatisfying∧(ℎ)=0, we get an element𝑔 ∈HomA(E ⊗AE,A)such that for all𝑖, 𝑗 ,

𝑔◦𝜎(𝜔𝑖A𝜔𝑗)=𝑔(𝜔𝑖A𝜔𝑗).

(13)

Since{𝜔𝑖A𝜔𝑗 :𝑖, 𝑗}is rightA-total inE ⊗A E(Corollary 2.5) and the maps𝑔, 𝜎 are rightA-linear, we get that𝑔◦𝜎=𝑔 .This proves our claim. Let us note that since we did not assume𝑔to be left invariant, the quantities𝑔(𝜔𝑖A𝜔𝑗)need not be scalars.

However, the proof goes through since the elements𝜎

𝑖 𝑗

𝑘 𝑙are scalars.

4. Pseudo-Riemannian metrics for cocycle deformations

This section concerns the braiding map and pseudo-Riemannian metrics of bicovariant bimodules under cocycle deformations of Hopf algebras. This section contains two main results. We start by recalling that a bicovariant bimoduleEover a Hopf algebraAcan be deformed in the presence of a 2-cocycle𝛾onAto a bicovariantA𝛾-bimoduleE𝛾.We prove that the canonical braiding map of the bicovariant bimoduleE𝛾(Proposition 2.9) is a cocycle deformation of the canonical braiding map of E.Finally, we prove that pseudo-Riemannian bi-invariant metrics onEandE𝛾are in one to one correspondence.

Throughout this section, we will make heavy use of the Sweedler notations as spelled out in (1.1), (1.2) and (1.3). The coassociativity ofΔwill be expressed by the following equation:

(Δ⊗Cid)Δ(𝑎)=(id⊗CΔ)Δ(𝑎)=𝑎(1)C𝑎(2)C𝑎(3). Also, when𝑚is an element of a bicovariant bimodule, we will use the notation

(id⊗C𝑀Δ)Δ𝑀(𝑚)=(Δ𝑀Cid)𝑀Δ(𝑚)=𝑚(−1)C𝑚(0)C𝑚(1). (4.1) Definition 4.1. A cocycle𝛾on a Hopf algebra(A,Δ)is aC-linear map𝛾:A ⊗CA →C such that it is convolution invertible, unital, i.e,

𝛾(𝑎⊗C1)=𝜖(𝑎)=𝛾(1⊗C𝑎) and for all𝑎, 𝑏, 𝑐inA,

𝛾(𝑎(1)C𝑏(1))𝛾(𝑎(2)𝑏(2)C𝑐)=𝛾(𝑏(1)C𝑐(1))𝛾(𝑎⊗C𝑏(2)𝑐(2)). (4.2) Given a Hopf algebra (A,Δ)and such a cocycle𝛾as above, we have a new Hopf algebra(A𝛾𝛾)which is equal toAas a vector space, the coproductΔ𝛾is equal toΔ while the algebra structure∗𝛾onA𝛾is defined by the following equation:

𝑎∗𝛾𝑏=𝛾(𝑎(1)C𝑏(1))𝑎(2)𝑏(2)𝛾(𝑎(3)C𝑏(3)). (4.3) Here,𝛾is the convolution inverse to𝛾which is unital and satisfies the following equation:

𝛾(𝑎(1)𝑏(1)C𝑐)𝛾(𝑎(2)C𝑏(2))=𝛾(𝑎⊗C𝑏(1)𝑐(1))𝛾(𝑏(2)C𝑐(2)). (4.4) We refer to [4, Theorem 1.6] for more details.

Suppose𝑀 is a bicovariantA-A-bimodule. Then𝑀 can also be deformed in the presence of a cocycle. This is the content of the next proposition.

(14)

Proposition 4.2 ([12, Theorem 2.5]). Suppose 𝑀 is a bicovariant A-bimodule and 𝛾is a2-cocycle onA.Then we have a bicovariantA𝛾-bimodule𝑀𝛾 which is equal to𝑀 as a vector space but the left and rightA𝛾-module structures are defined by the following formulas:

𝑎∗𝛾𝑚=𝛾(𝑎(1)C𝑚(−1))𝑎(2).𝑚(0)𝛾(𝑎(3)C𝑚(1)) (4.5) 𝑚∗𝛾𝑎 =𝛾(𝑚(−1)C𝑎(1))𝑚(0).𝑎(2)𝛾(𝑚(1)C𝑎(3)), (4.6) for all elements𝑚of𝑀and for all elements𝑎ofA. Here,∗𝛾 denotes the right and left A𝛾-module actions, and.denotes the right and leftA-module actions.

TheA𝛾-bicovariant structures are given by

Δ𝑀𝛾 := Δ𝑀 :𝑀𝛾→ A𝛾C𝑀𝛾and𝑀𝛾Δ:=𝑀Δ:𝑀𝛾→ 𝑀𝛾CA𝛾. (4.7) Remark 4.3. From Proposition 4.2, it is clear that if𝑀is a finite bicovariant bimodule (see Remark 2.2), then𝑀𝛾is also a finite bicovariant bimodule.

We end this subsection by recalling the following result on the deformation of bicovariant maps.

Proposition 4.4([12, Theorem 2.5]). Let(𝑀 ,Δ𝑀,𝑀Δ)and(𝑁 ,Δ𝑁,𝑁Δ)be bicovariant A-bimodules,𝑇 :𝑀 →𝑁be aC-linear bicovariant map and𝛾be a cocycle as above.

Then there exists a map𝑇𝛾 :𝑀𝛾→𝑁𝛾defined by𝑇𝛾(𝑚)=𝑇(𝑚)for all𝑚in𝑀. Thus, 𝑇𝛾 =𝑇 asC-linear maps. Moreover, we have the following:

(i) the deformed map𝑇𝛾:𝑀𝛾→ 𝑁𝛾is anA𝛾bicovariant map,

(ii) if𝑇is a bicovariant right (respectively left)A-linear map, then𝑇𝛾is a bicovariant right (respectively left)A𝛾-linear map,

(iii) if(𝑃,Δ𝑃,𝑃Δ)is another bicovariantA-bimodule, and𝑆:𝑁→𝑃is a bicovari- ant map, then(𝑆◦𝑇)𝛾:𝑀𝛾 →𝑃𝛾is a bicovariant map and𝑆𝛾◦𝑇𝛾=(𝑆◦𝑇)𝛾. 4.1. Deformation of the braiding map

SupposeEis a bicovariantA-bimodule,𝜎be the bicovariant braiding map of Propo- sition 2.9 and 𝑔 be a bi-invariant metric. Then Proposition 4.4 implies that we have deformed maps𝜎𝛾and𝑔𝛾.In this subsection, we study the map𝜎𝛾.The map𝑔𝛾will be discussed in the next subsection. We will need the following result:

(15)

Proposition 4.5([12, Theorem 2.5]). Let(𝑀 ,Δ𝑀,𝑀Δ)and(𝑁 ,Δ𝑁,𝑁Δ)be bicovariant bimodules over a Hopf algebraA and 𝛾be a cocycle as above. Then there exists a bicovariantA𝛾- bimodule isomorphism

𝜉:𝑀𝛾A𝛾 𝑁𝛾→ (𝑀⊗A𝑁)𝛾. The isomorphism𝜉and its inverse are respectively given by

𝜉(𝑚⊗A𝛾𝑛)=𝛾(𝑚(−1)C𝑛(−1))𝑚(0)A𝑛(0)𝛾(𝑚(1)C𝑛(1)) 𝜉−1(𝑚⊗A𝑛)=𝛾(𝑚(−1)C𝑛(−1))𝑚(0)A𝛾𝑛(0)𝛾(𝑚(1)C𝑛(1))

As an illustration, we make the following computation which will be needed later in this subsection:

Lemma 4.6. Suppose𝜔∈0E, 𝜂∈ E0.Then the following equation holds:

𝜉−1(𝛾(𝜂(−1)C1)𝜂(0)A𝜔(0)𝛾(1⊗C𝜔(1)))=𝜂⊗A𝛾𝜔.

Proof. Let us first clarify that we view𝛾(𝜂(−1)C1)𝜂(0)A 𝜔(0)𝛾(1⊗C𝜔(1))as an element in(E ⊗AE)𝛾.Then the equation holds because of the following computation:

𝜉−1(𝛾(𝜂(−1)C1)𝜂(0)A𝜔(0)𝛾(1⊗C𝜔(1)))

=𝛾(𝜂(−1)C1)𝜉−1(𝜂(0)A𝜔(0))𝛾(1⊗C𝜔(1))

=𝛾(𝜂(−2)C1)𝛾(𝜂(−1)C1)𝜂(0)A𝛾𝜔(0)𝛾(1⊗C𝜔(1))𝛾(1⊗C𝜔(2)) (since𝜔∈0E, 𝜂∈ E0)

=𝜖(𝜂(−2))𝜖(𝜂(−1))𝜂(0)A𝛾𝜔(0)𝜖(𝜔(1))𝜖(𝜔(2)) (since𝛾and𝛾are normalised)

=𝜂⊗A𝛾𝜔.

Now, we are in a position to study the map𝜎𝛾.By Proposition 4.2,E𝛾is a bicovariant A𝛾-bimodule. Then Proposition 2.9 guarantees the existence of a canonical braiding fromE𝛾A𝛾E𝛾to itself. We show that this map is nothing but the deformation𝜎𝛾of the map𝜎associated with the bicovariantA-bimoduleE. By the definition of𝜎𝛾,it is a map from(E ⊗AE)𝛾to(E ⊗A E)𝛾.However, by virtue of Proposition 4.5, the map𝜉 defines an isomorphism fromE𝛾A𝛾 E𝛾 to(E ⊗AE)𝛾.By an abuse of notation, we will denote the map

𝜉−1𝜎𝛾𝜉:E𝛾A𝛾 E𝛾→ E𝛾A𝛾E𝛾

by the symbol𝜎𝛾again.

Theorem 4.7([12, Theorem 2.5]). LetEbe a bicovariantA-bimodule and𝛾be a cocycle as above. Then the deformation𝜎𝛾of𝜎is the unique bicovariantA𝛾-bimodule braiding map onE𝛾given by Proposition 2.9.

(16)

Proof. Since𝜎 is a bicovariantA-bimodule map fromE ⊗A E to itself, part (ii) of Proposition 4.4 implies that𝜎𝛾is a bicovariantA𝛾-bimodule map from(E ⊗A E)𝛾 E𝛾A𝛾E𝛾to itself. By Proposition 2.9, there exists a uniqueA𝛾-bimodule map𝜎0from E𝛾A𝛾 E𝛾to itself such that𝜎0(𝜔⊗A𝛾 𝜂)=𝜂⊗A𝛾𝜔for all𝜔in0(E𝛾),𝜂in(E𝛾)0.

Since0(E𝛾)=0Eand(E𝛾)0 =E0, it is enough to prove that𝜎𝛾(𝜔⊗A𝛾𝜂)=𝜂⊗A𝛾𝜔 for all𝜔in0E,𝜂inE0.

We will need the concrete isomorphism betweenE𝛾A𝛾E𝛾 and(E ⊗AE)𝛾defined in Proposition 4.5. Since𝜔is in0Eand𝜂is inE0, this isomorphism maps the element 𝜔⊗A𝛾 𝜂to𝛾(1⊗C𝜂(−1))𝜔(0)A𝜂(0)𝛾(𝜔(1)C1). Then, by the definition of𝜎𝛾, we compute the following:

𝜎𝛾(𝜔⊗A𝛾𝜂)=𝜎(𝛾(1⊗C𝜂(−1))𝜔(0)A𝜂(0)𝛾(𝜔(1)C1))

=𝜎(𝜖(𝜂(−1))𝜔(0)A 𝜂(0)𝜖(𝜔(1)))=𝜖(𝜂(−1))𝜂(0)A𝜔(0)𝜖(𝜔(1))

=𝛾(𝜂(−1)C1)𝜂(0)A𝜔(0)𝛾(1⊗C𝜔(1))=𝜂⊗A𝛾𝜔,

where, in the last step we have used Lemma 4.6.

Remark 4.8. Proposition 4.2, Proposition 4.4, Proposition 4.5 and Theorem 4.7 together imply that the categories AAMAA and AA𝛾

𝛾

MAA𝛾

𝛾 are isomorphic as braided monoidal categories. This was the content of Theorem 2.5 of [12]. The referee has pointed out that this is a special case of a much more generalized result of Bichon ([3, Theorem 6.1]) which says that if two Hopf algebras are monoidally equivalent, then the corresponding categories of right-right Yetter Drinfeld modules are also monoidally equivalent.

However, in Theorem 4.7, we have proved in addition that the braiding onAA𝛾

𝛾

MAA𝛾

𝛾 is precisely the Woronowicz braiding of Proposition 2.9.

Corollary 4.9. If the unique bicovariantA-bimodule braiding map𝜎for a bicovariant A-bimoduleEsatisfies the equation𝜎2=1, then the bicovariantA𝛾-bimodule braiding map𝜎𝛾for the bicovariantA𝛾-bimoduleE𝛾 also satisfies𝜎2

𝛾 =1.

In particular, ifAis the commutative Hopf algebra of regular functions on a compact semisimple Lie group𝐺andEis its canonical space of one-forms, then the braiding map 𝜎𝛾forE𝛾 satisfies𝜎𝛾2=1.

Proof. By Theorem 4.7,𝜎𝛾is the unique braiding map for the bicovariantA𝛾-bimodule E𝛾. Since, by our hypothesis,𝜎2 =1, the deformed map𝜎𝛾 also satisfies𝜎2

𝛾 =1 by part (ii) of Proposition 4.4.

Next, ifAis a commutative Hopf algebra as in the statement of the corollary andEis its canonical space of one-forms, then we know that the braiding map𝜎is just the flip map, i.e. for all𝑒1, 𝑒2inE,

𝜎(𝑒1A 𝑒2)=𝑒2A𝑒1,

(17)

and hence it satisfies𝜎2 =1. Therefore, for every cocycle deformationE𝛾 ofE, the corresponding braiding map satisfies𝜎2

𝛾 =1.

4.2. Pseudo-Riemannian bi-invariant metrics onE𝛾

SupposeE is a bicovariantA-bimodule and E𝛾 be its cocycle deformation as above.

The goal of this subsection is to prove that a pseudo-Riemannian bi-invariant metric onEnaturally deforms to a pseudo-Riemannian bi-invariant metric onE𝛾.Since𝑔is a bicovariant (i.e, both left and right covariant) map from the bicovariant bimoduleE ⊗AE to itself, then by Proposition 4.4, we have a rightA𝛾-linear bicovariant map𝑔𝛾 from E𝛾A𝛾 E𝛾to itself. We need to check the conditions (i) and (ii) of Definition 3.1 for the map𝑔𝛾.

The proof of the equality 𝑔𝛾 = 𝑔𝛾 ◦𝜎𝛾 is straightforward. However, checking condition (ii), i.e, verifying that the map𝑉𝑔

𝛾 is an isomorphism onto its image needs some work. The root of the problem is that we do not yet know whetherE=𝑉𝑔(E).Our strategy to verify condition (ii) is the following: we show that the rightA-module𝑉𝑔(E) is a bicovariant rightA-module (see Definition 4.10) in a natural way. Let us remark that since the map𝑔(hence𝑉𝑔) is not leftA-linear,𝑉𝑔(E)need not be a leftA-module.

Since bicovariant rightA-modules and bicovariant maps between them can be deformed (Proposition 4.13), the map𝑉𝑔deforms to a rightA𝛾-linear isomorphism fromE𝛾 to (𝑉𝑔(E))𝛾.Then in Theorem 4.15, we show that(𝑉𝑔)𝛾 coincides with the map𝑉𝑔

𝛾and the latter is an isomorphism onto its image. This is the only subsection where we use the theory of bicovariant right modules (as opposed to bicovariant bimodules).

Definition 4.10. Let𝑀 be a rightA-module, andΔ𝑀 :𝑀 → A ⊗C𝑀 and𝑀Δ:𝑀 → 𝑀⊗CAbeC-linear maps. We say that(𝑀 ,Δ𝑀,𝑀Δ)is a bicovariant rightA-module if the triplet is an object of the category AMAA,i.e,

(i) (𝑀 ,Δ𝑀)is a leftA-comodule, (ii) (𝑀 ,𝑀Δ)is a rightA-comodule, (iii) (id⊗C𝑀Δ)Δ𝑀 =(Δ𝑀Cid)𝑀Δ, (iv) For any𝑎inAand𝑚in𝑀,

Δ𝑀(𝑚 𝑎)= Δ𝑀(𝑚)Δ(𝑎), 𝑀Δ(𝑚 𝑎)=𝑀Δ(𝑚)Δ(𝑎).

For the rest of the subsection,E will denote a bicovariantA-bimodule. Moreover, {𝜔𝑖}𝑖will denote a basis of0Eand{𝜔

𝑖}𝑖the dual basis, i.e,𝜔

𝑖(𝜔𝑗)=𝛿𝑖 𝑗.

(18)

Let us recall that (2.1) implies the existence of elements𝑅𝑖 𝑗inAsuch that

EΔ(𝜔𝑖)=∑︁

𝑖 𝑗

𝜔𝑗C𝑅𝑗 𝑖. (4.8)

We want to show that𝑉𝑔(E)is a bicovariant rightA-module. To this end, we recall that (Lemma 3.7)𝑉𝑔(E)is a free rightA-module with basis{𝜔

𝑖}𝑖.This allows us to make the following definition.

Definition 4.11. Let{𝜔𝑖}𝑖and{𝜔

𝑖}𝑖be as above and𝑔a bi-invariant pseudo-Riemannian metric on E. Then we can endow𝑉𝑔(E) with a left-coaction Δ𝑉𝑔( E) : 𝑉𝑔(E) → A ⊗C𝑉𝑔(E) and a right-coaction 𝑉𝑔( E)Δ : 𝑉𝑔(E) → 𝑉𝑔(E) ⊗C A, defined by the formulas

Δ𝑉𝑔( E)

∑︁

𝑖

𝜔𝑖𝑎𝑖

!

=∑︁

𝑖

(1⊗C𝜔𝑖)Δ(𝑎𝑖), 𝑉

𝑔( E)Δ ∑︁

𝑖

𝜔𝑖𝑎𝑖

!

=∑︁

𝑖 𝑗

(𝜔

𝑗C𝑆(𝑅𝑖 𝑗))Δ(𝑎𝑖),

(4.9)

where the elements𝑅𝑖 𝑗are as in (4.8).

Then we have the following result.

Proposition 4.12. The triplet(𝑉𝑔(E),Δ𝑉𝑔( E),𝑉

𝑔( E)Δ)is a bicovariant rightA-module.

Moreover, the map𝑉𝑔:E →𝑉𝑔(E)is bicovariant, i.e, we have Δ𝑉𝑔( E)(𝑉𝑔(𝑒))=(id⊗C𝑉𝑔E(𝑒), 𝑉

𝑔( E)Δ(𝑉𝑔(𝑒))=(𝑉𝑔Cid)EΔ(𝑒). (4.10) Proof. The fact that(𝑉𝑔(E),Δ𝑉𝑔( E),𝑉

𝑔( E)Δ)is a bicovariant rightA-module follows immediately from the definition of the mapsΔ𝑉𝑔( E) and𝑉𝑔( E)Δ.So we are left with proving (4.10). Let𝑒 ∈ E.Then there exist elements 𝑎𝑖 inAsuch that 𝑒= Í

𝑖𝜔𝑖𝑎𝑖. Hence, by (3.1), we obtain

Δ𝑉𝑔( E)(𝑉𝑔(𝑒))= Δ𝑉𝑔( E) 𝑉𝑔

∑︁

𝑖

𝜔𝑖𝑎𝑖

! !

= Δ𝑉𝑔( E)

∑︁

𝑖 𝑗

𝑔𝑖 𝑗𝜔

𝑗𝑎𝑖

!

=∑︁

𝑖 𝑗

(1⊗C𝑔𝑖 𝑗𝜔

𝑗)Δ(𝑎𝑖)=∑︁

𝑖

( (id⊗C𝑉𝑔) (1⊗C𝜔𝑖))Δ(𝑎𝑖)

=∑︁

𝑖

(id⊗C𝑉𝑔) (ΔE(𝜔𝑖))Δ(𝑎𝑖)

=∑︁

𝑖

(id⊗C𝑉𝑔E(𝜔𝑖𝑎𝑖)=(id⊗C𝑉𝑔E(𝑒).

This proves the first equation of (4.10).

Références

Documents relatifs

First, it is to introduce in Continuum Mechanics the manifold Met( B ) of Riemannian metrics on the body B and to show that each covariant derivative on Met( B ) induces an

Theorem A is a true generalization of Ledrappier and Wang’s result: when Γ\X is assumed to be compact, X has bounded curvature and the Poincar´e exponent and the volume growth

other hand we prove a uniqueness theorem for harmonic maps which takes values in a neighbourhood of a

The solution of the geodesic equation (Theorem 4.1) gives a precise sense to how geodesics in ( M , g N ) generalize those discovered by Calabi [4,5] for the subspace of Kähler

The remainder of the article is arranged as follows: in Section 1.2 conditions required on the background metric are introduced so that an initial statement of the theorem can be

Toute utilisa- tion commerciale ou impression systématique est constitutive d’une in- fraction pénale.. Toute copie ou impression de ce fichier doit conte- nir la présente mention

We consider the isospectral deformations confined to Lax’s sense which are called L-isospectral deformations, and set up the following problem.. PROBLEM B: Is there a

In 1999-2000, Aprodu, Aprodu and Brˆınz˘ anescu [1] introduced and studied the notion of pseudo-horizontally homothetic map which, if it is also a harmonic map, is an