Canonical basis

Summary

In mathematics, a canonical basis is a basis of an algebraic structure that is canonical in a sense that depends on the precise context:

Representation theory edit

The canonical basis for the irreducible representations of a quantized enveloping algebra of type   and also for the plus part of that algebra was introduced by Lusztig [2] by two methods: an algebraic one (using a braid group action and PBW bases) and a topological one (using intersection cohomology). Specializing the parameter   to   yields a canonical basis for the irreducible representations of the corresponding simple Lie algebra, which was not known earlier. Specializing the parameter   to   yields something like a shadow of a basis. This shadow (but not the basis itself) for the case of irreducible representations was considered independently by Kashiwara;[3] it is sometimes called the crystal basis. The definition of the canonical basis was extended to the Kac-Moody setting by Kashiwara [4] (by an algebraic method) and by Lusztig [5] (by a topological method).

There is a general concept underlying these bases:

Consider the ring of integral Laurent polynomials   with its two subrings   and the automorphism   defined by  .

A precanonical structure on a free  -module   consists of

  • A standard basis   of  ,
  • An interval finite partial order on  , that is,   is finite for all  ,
  • A dualization operation, that is, a bijection   of order two that is  -semilinear and will be denoted by   as well.

If a precanonical structure is given, then one can define the   submodule   of  .

A canonical basis of the precanonical structure is then a  -basis   of   that satisfies:

  •   and
  •  

for all  .

One can show that there exists at most one canonical basis for each precanonical structure.[6] A sufficient condition for existence is that the polynomials   defined by   satisfy   and  .

A canonical basis induces an isomorphism from   to  .

Hecke algebras edit

Let   be a Coxeter group. The corresponding Iwahori-Hecke algebra   has the standard basis  , the group is partially ordered by the Bruhat order which is interval finite and has a dualization operation defined by  . This is a precanonical structure on   that satisfies the sufficient condition above and the corresponding canonical basis of   is the Kazhdan–Lusztig basis

 

with   being the Kazhdan–Lusztig polynomials.

Linear algebra edit

If we are given an n × n matrix   and wish to find a matrix   in Jordan normal form, similar to  , we are interested only in sets of linearly independent generalized eigenvectors. A matrix in Jordan normal form is an "almost diagonal matrix," that is, as close to diagonal as possible. A diagonal matrix   is a special case of a matrix in Jordan normal form. An ordinary eigenvector is a special case of a generalized eigenvector.

Every n × n matrix   possesses n linearly independent generalized eigenvectors. Generalized eigenvectors corresponding to distinct eigenvalues are linearly independent. If   is an eigenvalue of   of algebraic multiplicity  , then   will have   linearly independent generalized eigenvectors corresponding to  .

For any given n × n matrix  , there are infinitely many ways to pick the n linearly independent generalized eigenvectors. If they are chosen in a particularly judicious manner, we can use these vectors to show that   is similar to a matrix in Jordan normal form. In particular,

Definition: A set of n linearly independent generalized eigenvectors is a canonical basis if it is composed entirely of Jordan chains.

Thus, once we have determined that a generalized eigenvector of rank m is in a canonical basis, it follows that the m − 1 vectors   that are in the Jordan chain generated by   are also in the canonical basis.[7]

Computation edit

Let   be an eigenvalue of   of algebraic multiplicity  . First, find the ranks (matrix ranks) of the matrices  . The integer   is determined to be the first integer for which   has rank   (n being the number of rows or columns of  , that is,   is n × n).

Now define

 

The variable   designates the number of linearly independent generalized eigenvectors of rank k (generalized eigenvector rank; see generalized eigenvector) corresponding to the eigenvalue   that will appear in a canonical basis for  . Note that

 

Once we have determined the number of generalized eigenvectors of each rank that a canonical basis has, we can obtain the vectors explicitly (see generalized eigenvector).[8]

Example edit

This example illustrates a canonical basis with two Jordan chains. Unfortunately, it is a little difficult to construct an interesting example of low order.[9] The matrix

 

has eigenvalues   and   with algebraic multiplicities   and  , but geometric multiplicities   and  .

For   we have  

  has rank 5,
  has rank 4,
  has rank 3,
  has rank 2.

Therefore  

 
 
 
 

Thus, a canonical basis for   will have, corresponding to   one generalized eigenvector each of ranks 4, 3, 2 and 1.

For   we have  

  has rank 5,
  has rank 4.

Therefore  

 
 

Thus, a canonical basis for   will have, corresponding to   one generalized eigenvector each of ranks 2 and 1.

A canonical basis for   is

 

  is the ordinary eigenvector associated with  .   and   are generalized eigenvectors associated with  .   is the ordinary eigenvector associated with  .   is a generalized eigenvector associated with  .

A matrix   in Jordan normal form, similar to   is obtained as follows:

 
 

where the matrix   is a generalized modal matrix for   and  .[10]

See also edit

Notes edit

  1. ^ Bronson (1970, p. 196)
  2. ^ Lusztig (1990)
  3. ^ Kashiwara (1990)
  4. ^ Kashiwara (1991)
  5. ^ Lusztig (1991)
  6. ^ Lusztig (1993, p. 194)
  7. ^ Bronson (1970, pp. 196, 197)
  8. ^ Bronson (1970, pp. 197, 198)
  9. ^ Nering (1970, pp. 122, 123)
  10. ^ Bronson (1970, p. 203)

References edit

  • Bronson, Richard (1970), Matrix Methods: An Introduction, New York: Academic Press, LCCN 70097490
  • Deng, Bangming; Ju, Jie; Parshall, Brian; Wang, Jianpan (2008), Finite Dimensional Algebras and Quantum Groups, Mathematical surveys and monographs, vol. 150, Providence, R.I.: American Mathematical Society, ISBN 9780821875315
  • Kashiwara, Masaki (1990), "Crystalizing the q-analogue of universal enveloping algebras", Communications in Mathematical Physics, 133 (2): 249–260, Bibcode:1990CMaPh.133..249K, doi:10.1007/bf02097367, ISSN 0010-3616, MR 1090425, S2CID 121695684
  • Kashiwara, Masaki (1991), "On crystal bases of the q-analogue of universal enveloping algebras", Duke Mathematical Journal, 63 (2): 465–516, doi:10.1215/S0012-7094-91-06321-0, ISSN 0012-7094, MR 1115118
  • Lusztig, George (1990), "Canonical bases arising from quantized enveloping algebras", Journal of the American Mathematical Society, 3 (2): 447–498, doi:10.2307/1990961, ISSN 0894-0347, JSTOR 1990961, MR 1035415
  • Lusztig, George (1991), "Quivers, perverse sheaves and quantized enveloping algebras", Journal of the American Mathematical Society, 4 (2): 365–421, doi:10.2307/2939279, ISSN 0894-0347, JSTOR 2939279, MR 1088333
  • Lusztig, George (1993), Introduction to quantum groups, Boston, MA: Birkhauser Boston, ISBN 0-8176-3712-5, MR 1227098
  • Nering, Evar D. (1970), Linear Algebra and Matrix Theory (2nd ed.), New York: Wiley, LCCN 76091646