††thanks: This work was completed with the support of CONACYT.
Standard automorphisms of semisimple Lie algebras and their relations
David Reynoso-Mercado
Instituto de Matemáticas, Facultad de Ciencias Exactas y Naturales, Universidad de Antioquia UdeA,\brCalle 70 No. 52-21, Medellín, Colombia
[email protected]
(Date: February 28, 2023)
Abstract.
Let be the simple Lie algebra of square matrices with zero trace. There are certain relations concerning standard automorphisms that are considered “folklore”. One can find a complete proof of these in Millson and Toledano Laredo’s work [Transformation Groups, Vol. 10, No. 2, 2005, pp. 217– 254], but said proof is based on topological arguments which are non-trivial. The aim of this work is to verify these relations in an elementary way.
Key words and phrases:
Lie algebras, semisimple Lie algebra, standar automorphisms, Braid group
1991 Mathematics Subject Classification:
Primary 17Bxx; Secondary 17B10
1. Introduction
Let be the simple Lie algebra of square matrices with zero trace. We denote the standard basis of the space of square matrices by and the set by .
As a Lie algebra, is generated by the elements and for .
Note that with the corresponding simple co-roots , with , we have that and is a Cartan subalgebra of .
Let , be a vector subspace of . The dual base of the standard base of is denoted as , and the corresponding elements of are denoted as .
With this notation
is the root system of , that is, the
are the root spaces of with respect to . The Weyl group of , denoted by , is identified with the symmetric group , which acts linearly on such that for
is generated by the simple transpositions with , and the following relations form a complete system of relations
(1.1)
Let us denote the adjoint representation of by , , and define
(1.2)
For all , it is easy to see that for all whit . In [3, 21.2], we can see that the weight spaces of the adjoint representation are permuted by the in the same way that the corresponding weights (roots in this case) are permuted by the , that is .
The aim of this work is to verify the following realations by the ’s in an elementary way.
Theorem 1.1.
Let be a positive integer and ’s be the automorphisms of defined in (1.2). Then the automorphisms , , fulfil to the relations
(1.3)
(1.4)
(1.5)
(1.6)
for any , where is the same for the relation in (1.1).
Note that the relations (1.1) for imply . Under this premise, the remaining relations of (1.1) are equivalent to (1.3), also known as braid relations.
It is evident that directly verifying the relations of Theorem 1.1 is a challenging task. To prove our result, we follow the steps of [5, Proposition 2.14], which is a more general result.
We will study the elements
and verify that these elements of the simple connected Lie group fulfill “the same” relations as the elements of Theorem 1.1 inspired by the results of Millson and Toledano Laredo [5]. It is important to note that the “braid relations” in this context are classical ([7],[1]), while the other relations are considered “folklore”.
The theorem’s result will follow once we observe that
It is important to note that the automorphisms can be defined not only for the adjoint representation, but also for any finite-dimensional representation of , see [3, 21.2]. In this context, they are relevant when one wants to see that the character of is invariant under the action of the Weyl group.
2. Preliminaries
2.1. Braid group
The Artin braid group is the group generated by the elements and the braid relations:
(1)
, for all and in , with , and
(2)
, for in .
Let be the symmetric group and its generators, where is the adjacent transposition that swaps and . It is known that there is a unique group homomorphism such that , see [4, Lemma 1.2], the is called the natural projection. Moreover, since is the generator set we have that is surjective.
The kernel of is called the pure braid group and is denoted by . The elements of are called pure braids on strings.
2.2. Lie algebras and their representations
Let be a complex matrix. Recall that the exponential of , denoted by or , is the usual power series
Proposition 2.1.
For all , the series converges and is a continuous function of .
Let be a finite dimensional vector space over and denote the set of linear transformations . We write for , viewed as Lie algebra and call it the general linear algebra. If , then , and we use to denote . Recall that the special linear Lie algebra denoted by is the set of matrices such that , where is the trace of .
A representation of a Lie algebra is a pair , where is a vector space over and is a homeomorphism of Lie algebras, that is is a linear transformation such that . If is a Lie algebra and , we define the adjoint representation of as where is a homomorphism of Lie algebras.
Remark 2.2.
Let . Then .
Let be matrices of , defined as follows
(2.7)
Then the set is a basis of , such that , . The following theorem is taken from [6, Theorem 1.2.14].
Theorem 2.3.
Let be a field with and the Lie algebra . Then the following holds
(1)
For any an integer positive, there is an unique simple representation (up to isomorphisms) such that .
(2)
Let be a basis of , then each simple representation of dimension have a decomposes into one-dimensional eigenspaces under
for the integer eigenvalues , and additionally from already follows as well as .
2.3. Root space decomposition
Let and be the subset of whose elements are diagonal matrices. Then is subalgebra of . Let be a representation of with semisimple for all . Let and be the standard basis of , where be the matrix whose -th entry is and the remaining ones is , that . We can define a map by , then
Let be the set of weights of , it is easy to see that . For then and , if .
Let a semisimple Lie algebra over and be a subalgebra. We call a Cartan subalgebra if and only if is abelian, it consists of semisimple elements and is maximal with this property. Let be a subalgebra of a semisimple Lie algebra over . We will write to denote for all and and we use to denote the adjoint map . Since is Cartan subalgebra, then its elements are semisimple. Then it follows from [6, Proposición 2.2.2] that is semisimple . Thus, is equal to , where
Now, consider and the subalgebra whose elements are diagonal matrices with trace equal zero. Then, with is a basis of . Moreover, we have that . Thus,
The following theorem is taken from [6, Theorem 2.3.12].
Theorem 2.4.
Let be a semisimple Lie algebra over , a Cartan subalgebra and the corresponding root system. The following statements hold:
(1)
is itself centralizer.
(2)
All root spaces have one dimension. Moreover, for all , there exists an injective Lie algebras homomorphism defined as
In particular, if and only if .
(3)
If , then .
(4)
If then .
For , we can define a co-root as an element in such that and .
Theorem 2.5.
Let , and be as in Theorem 2.4. The following statements hold:
For the Lie algebra we have that its corresponding root system and is the set of simple roots.
Let be the Weyl group of . We want to show that . Due to the definition of Weyl group, it is enough to show that for all . First, from the definition of (see theorem 2.5), we have
(2.12)
for any . On other hand, for any and , we have
(2.18)
Using equation (2.12) and (2.18), it is easy to verify that
Therefore, . Moreover, simple computations reveal that is isomorphic to the symmetric group .
3. Tits extension
In this section, we will recall the arguments about generalised braid groups presented by Millson and Toledano ([5, Section 2]).
Let be a complex, semisimple Lie algebra with Cartan subalgebra and be the Weyl group. Let be a root system, a basis of and be a finite-dimensional -module. We denote by the set of regular elements.
Fix a basepoint . Let be its image in , and let , be the generalized pure and full braid groups of type . The fibration gives rise to the exact sequence
with is obtained by associating to to the unique such that , where is the unique lift of to a path in such that .
Let be the complex, connected, and simply-connected Lie group with Lie algebra , its torus with Lie algebra , and the normalizer of in , so that . We regard as acting on by choosing a homeomorphism compatible with the diagram in Figure 1.
Figure 1. Diagram
By Brieskorn’s theorem (see [1, page 57]), is presented on generators , labelled by the simple reflections , with relations
(3.1)
where is the order of in the Weyl group. One of the first works in which the relations from (3.1) appears is [7, Lemma 56].
Tits has given a simple construction of a canonical class of homomorphisms that differ from each other via conjugation by an element of , see [8].
This section will divide the proposition of Milson and Toledano [5, Proposition 2.14], into three new propositions to verify the relations (1.3)-(1.6) in our particular case, the Lie algebra .
From this point on, we will be working with the Lie group of matrices , with its Lie algebra. For a simple root as in section 2.3, we can define
where is the identity matrix in and the matrix is composed entirely of zeros.
It is easy to see that and that the Lie algebra is generated by
For any , let , for , be an algebraic torus or more precisely:
Let be the set of diagonal matrices with determinant equal to .
Proposition 4.1.
Let be the torus of group as above, then is equal to .
Proof.
Since it is clear that any set is a subset of , then .
On other hand, for an element of , for any there exists a complex number such that . We can define
and since the determinant of is equal to , we have . Then,
Thus, and .
∎
Obviously, the Lie algebra of is equal to , where is the subgroup of diagonal matrices which trace equal to zero.
Proposition 4.2.
Let be the normalizer of in . Then, we have the following equality
(4.1)
where .
Proof.
Let’s take in the right set of Equation (4.1). Is clear that is the inverse matrix of . Now, let be as follows
After some simple calculations, we have that is a diagonal matrix and is equal to . Thus, .
On other hand, let’s take an element of , with inverse being . Then, for all , we have that , including the generators of . Let , for , then
where
Since , then for we have
But for , we have that . Thus,
It follows that . Therefore, for all such that .
For any , we can define the set , whose elements are in the -th row. Since we have that .
Now, let . Then, for all . It follows that . Thus, for any . Therefore is the unique non zero element in the -column of . Thus, for we have that , hence . There exist such that . Therefore
∎
For any , we use to denote , where sgn is the sign of (which is equal to if is even and is if is odd). It is easy to see that . Then, we have that the set is a subset of .
On other hand, for any , it holds true that:
Let , and be such that . Then, is equal to and . Thus, , and , with .
We have that, for any such that , is not a diagonal matrix, since . Therefore, for any , we have if and only if . Thus, we can conclude that
that is, is isomophic to , the simetric group. Given that at the end of section 2.3 we saw that , then .
A straightforward calculation reveals that the normalizer of in , denoted by , has the following form:
from which it is evident that .
The orthogonal reflection corresponding to is denoted by .
The proposition of Millson and Toledano [5, Proposition 2.14] will now be divided into three parts, each of which will be proven for our specific case.
Proposition 4.3.
For any choice of , , the assignment extends uniquely to a homomorphism , that makes the diagram in Figure (1) commute.
Proof.
We must show that the satisfy the braid relations, which can be verified through simple calculations. Let be an element in , and
Recall that
We have the following cases.
Case 1:
When , we have to prove that . Without loss of generality, we can chose . Due to the definition of matrices and , it suffices to observe the behavior of their following submatrices in .
We have
It follows that .
Case 2:
When , we have to prove that . Without loss of generality, we can chose . Consider the submatrices of and below:
Then
Therefore, the braid relations are satisfied.
∎
Proposition 4.4.
If are the homomorphisms corresponding to the choices and , with , respectively, then there exists a such that, for any , .
Proof.
For any , let and be the following elements
We can chose , as follows
From our choice of , it is clear that . Let be a complex number such that . Now, take the co-weights defined by , where is Kronecker’s delta. In our case, with
We choose a diagonal matrix as follow . Since and its inverse matrix are diagonal matrices, we only need to consider the following product of submatrices:
Then, we can conclude that
Thus, we have for all that , with .
Therefore, .
∎
Proposition 4.5.
For any , the homomorphism corresponding to the choice , is subjected to the relations:
(4.14)
(4.15)
(4.16)
(4.17)
for any , where the number of factors on each side of (4.14) is equal to the order of .
Proof.
In proposition 4.3, we have already proved that satisfies equality (4.14). Additionally, we note that for all , if we take
which can also be viewed as , it follows that
It is then evident that the satisfy the relations (4.15) y (4.16).
For the last equation, we can take , and thus have two cases:
Case 1:
If , then . Let’s suppose , this implies that:
The case in which is analogous to the previous one.
Case 2:
If , then .
Thus, . Since , we have that , from where we have that
Therefore, it can be concluded that equality (4.17) has been achieved.
∎
Corollary 4.6.
Let be a positive integer and ’s be the automorphisms of defined in (1.2). Then the automorphisms , are subjected to the relations (4.14)-(4.17).
Proof.
As we saw in Remark (2.2), for any matrices in , we have that is equal to , including the particular case where or .
For all , let be equal to . Then,
From the definition of element and , we have that one of the possible choices of elements of is
Thus, satisfies the relations (4.14)-(4.17). Since , we can conclude that the standard automorphisms satisfy the desired relations.
∎
Acknowledgment
I would like to thank CONACYT for the Support for the Training of Human Resources with registration number 25412 of the Research Project CB-2014/239255 and Christof Geiss for his support, guidance and encouragement. Part of this work was done during my Master’s studies at the Universidad Nacional Autónoma de México.
References
[1]E. Brieskorn, Die Fundamentalgruppe des Raumes der Regulären Orbits einer Endlichen Komplexen Spiegelungsgruppe, Invent. Math. 12 (1971), 57– 61. Available from: http://dx.doi.org/10.1007/bf01389827
[2] Brian C. Hall, Lie groups, Lie Algebras, and Representations. Springer Verlag, 2003.
[3] James E. Humphreys, Introduction to Lie Algebras and Representation Theory. Springer Verlag, 1972.
[4] C. Kasssel, V. Turaev, Braid Groups. Springer Science+Business Media, LLC 2008.
[5] John J. Millson, V. Toledano Laredo, Casimir operators and monodromy representations of generalised braid groups, Transformation Groups, Vol. 10, No. 2, 2005, pp. 217– 254.
[6] W. Soergel, Halbeinfache Lie-algebren. Manuscript available on the page of the author:
http://home.mathematik.uni-freiburg.de/soergel/Skripten/XXHL.pdf
[7] R. Steinberg, Lectures on Chevalley groups. Notes prepared by John Faulkner and Robert Wilson. Yale University, New Haven, Conn., 1968 iii+277pp.
[8]J. Tits, Normalisateurs de tores. I. Groupes de Coxeter etendus, J. Algebra 4 (1966), 96 –116.