Canonical basis

Basis of a type of algebraic structure

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

  • In a coordinate space, and more generally in a free module, it refers to the standard basis defined by the Kronecker delta.
  • In a polynomial ring, it refers to its standard basis given by the monomials, ( X i ) i {\displaystyle (X^{i})_{i}} .
  • For finite extension fields, it means the polynomial basis.
  • In linear algebra, it refers to a set of n linearly independent generalized eigenvectors of an n×n matrix A {\displaystyle A} , if the set is composed entirely of Jordan chains.[1]
  • In representation theory, it refers to the basis of the quantum groups introduced by Lusztig.

Representation theory

The canonical basis for the irreducible representations of a quantized enveloping algebra of type A D E {\displaystyle ADE} 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 q {\displaystyle q} to q = 1 {\displaystyle q=1} yields a canonical basis for the irreducible representations of the corresponding simple Lie algebra, which was not known earlier. Specializing the parameter q {\displaystyle q} to q = 0 {\displaystyle q=0} 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 Z := Z [ v , v 1 ] {\displaystyle {\mathcal {Z}}:=\mathbb {Z} \left[v,v^{-1}\right]} with its two subrings Z ± := Z [ v ± 1 ] {\displaystyle {\mathcal {Z}}^{\pm }:=\mathbb {Z} \left[v^{\pm 1}\right]} and the automorphism ¯ {\displaystyle {\overline {\cdot }}} defined by v ¯ := v 1 {\displaystyle {\overline {v}}:=v^{-1}} .

A precanonical structure on a free Z {\displaystyle {\mathcal {Z}}} -module F {\displaystyle F} consists of

  • A standard basis ( t i ) i I {\displaystyle (t_{i})_{i\in I}} of F {\displaystyle F} ,
  • An interval finite partial order on I {\displaystyle I} , that is, ( , i ] := { j I j i } {\displaystyle (-\infty ,i]:=\{j\in I\mid j\leq i\}} is finite for all i I {\displaystyle i\in I} ,
  • A dualization operation, that is, a bijection F F {\displaystyle F\to F} of order two that is ¯ {\displaystyle {\overline {\cdot }}} -semilinear and will be denoted by ¯ {\displaystyle {\overline {\cdot }}} as well.

If a precanonical structure is given, then one can define the Z ± {\displaystyle {\mathcal {Z}}^{\pm }} submodule F ± := Z ± t j {\textstyle F^{\pm }:=\sum {\mathcal {Z}}^{\pm }t_{j}} of F {\displaystyle F} .

A canonical basis of the precanonical structure is then a Z {\displaystyle {\mathcal {Z}}} -basis ( c i ) i I {\displaystyle (c_{i})_{i\in I}} of F {\displaystyle F} that satisfies:

  • c i ¯ = c i {\displaystyle {\overline {c_{i}}}=c_{i}} and
  • c i j i Z + t j  and  c i t i mod v F + {\displaystyle c_{i}\in \sum _{j\leq i}{\mathcal {Z}}^{+}t_{j}{\text{ and }}c_{i}\equiv t_{i}\mod vF^{+}}

for all i I {\displaystyle i\in I} .

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 r i j Z {\displaystyle r_{ij}\in {\mathcal {Z}}} defined by t j ¯ = i r i j t i {\textstyle {\overline {t_{j}}}=\sum _{i}r_{ij}t_{i}} satisfy r i i = 1 {\displaystyle r_{ii}=1} and r i j 0 i j {\displaystyle r_{ij}\neq 0\implies i\leq j} .

A canonical basis induces an isomorphism from F + F + ¯ = i Z c i {\displaystyle \textstyle F^{+}\cap {\overline {F^{+}}}=\sum _{i}\mathbb {Z} c_{i}} to F + / v F + {\displaystyle F^{+}/vF^{+}} .

Hecke algebras

Let ( W , S ) {\displaystyle (W,S)} be a Coxeter group. The corresponding Iwahori-Hecke algebra H {\displaystyle H} has the standard basis ( T w ) w W {\displaystyle (T_{w})_{w\in W}} , the group is partially ordered by the Bruhat order which is interval finite and has a dualization operation defined by T w ¯ := T w 1 1 {\displaystyle {\overline {T_{w}}}:=T_{w^{-1}}^{-1}} . This is a precanonical structure on H {\displaystyle H} that satisfies the sufficient condition above and the corresponding canonical basis of H {\displaystyle H} is the Kazhdan–Lusztig basis

C w = y w P y , w ( v 2 ) T w {\displaystyle C_{w}'=\sum _{y\leq w}P_{y,w}(v^{2})T_{w}}

with P y , w {\displaystyle P_{y,w}} being the Kazhdan–Lusztig polynomials.

Linear algebra

If we are given an n × n matrix A {\displaystyle A} and wish to find a matrix J {\displaystyle J} in Jordan normal form, similar to A {\displaystyle A} , 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 D {\displaystyle D} 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 A {\displaystyle A} possesses n linearly independent generalized eigenvectors. Generalized eigenvectors corresponding to distinct eigenvalues are linearly independent. If λ {\displaystyle \lambda } is an eigenvalue of A {\displaystyle A} of algebraic multiplicity μ {\displaystyle \mu } , then A {\displaystyle A} will have μ {\displaystyle \mu } linearly independent generalized eigenvectors corresponding to λ {\displaystyle \lambda } .

For any given n × n matrix A {\displaystyle A} , 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 A {\displaystyle A} 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 x m 1 , x m 2 , , x 1 {\displaystyle \mathbf {x} _{m-1},\mathbf {x} _{m-2},\ldots ,\mathbf {x} _{1}} that are in the Jordan chain generated by x m {\displaystyle \mathbf {x} _{m}} are also in the canonical basis.[7]

Computation

Let λ i {\displaystyle \lambda _{i}} be an eigenvalue of A {\displaystyle A} of algebraic multiplicity μ i {\displaystyle \mu _{i}} . First, find the ranks (matrix ranks) of the matrices ( A λ i I ) , ( A λ i I ) 2 , , ( A λ i I ) m i {\displaystyle (A-\lambda _{i}I),(A-\lambda _{i}I)^{2},\ldots ,(A-\lambda _{i}I)^{m_{i}}} . The integer m i {\displaystyle m_{i}} is determined to be the first integer for which ( A λ i I ) m i {\displaystyle (A-\lambda _{i}I)^{m_{i}}} has rank n μ i {\displaystyle n-\mu _{i}} (n being the number of rows or columns of A {\displaystyle A} , that is, A {\displaystyle A} is n × n).

Now define

ρ k = rank ( A λ i I ) k 1 rank ( A λ i I ) k ( k = 1 , 2 , , m i ) . {\displaystyle \rho _{k}=\operatorname {rank} (A-\lambda _{i}I)^{k-1}-\operatorname {rank} (A-\lambda _{i}I)^{k}\qquad (k=1,2,\ldots ,m_{i}).}

The variable ρ k {\displaystyle \rho _{k}} designates the number of linearly independent generalized eigenvectors of rank k (generalized eigenvector rank; see generalized eigenvector) corresponding to the eigenvalue λ i {\displaystyle \lambda _{i}} that will appear in a canonical basis for A {\displaystyle A} . Note that

rank ( A λ i I ) 0 = rank ( I ) = n . {\displaystyle \operatorname {rank} (A-\lambda _{i}I)^{0}=\operatorname {rank} (I)=n.}

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

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

A = ( 4 1 1 0 0 1 0 4 2 0 0 1 0 0 4 1 0 0 0 0 0 5 1 0 0 0 0 0 5 2 0 0 0 0 0 4 ) {\displaystyle A={\begin{pmatrix}4&1&1&0&0&-1\\0&4&2&0&0&1\\0&0&4&1&0&0\\0&0&0&5&1&0\\0&0&0&0&5&2\\0&0&0&0&0&4\end{pmatrix}}}

has eigenvalues λ 1 = 4 {\displaystyle \lambda _{1}=4} and λ 2 = 5 {\displaystyle \lambda _{2}=5} with algebraic multiplicities μ 1 = 4 {\displaystyle \mu _{1}=4} and μ 2 = 2 {\displaystyle \mu _{2}=2} , but geometric multiplicities γ 1 = 1 {\displaystyle \gamma _{1}=1} and γ 2 = 1 {\displaystyle \gamma _{2}=1} .

For λ 1 = 4 , {\displaystyle \lambda _{1}=4,} we have n μ 1 = 6 4 = 2 , {\displaystyle n-\mu _{1}=6-4=2,}

( A 4 I ) {\displaystyle (A-4I)} has rank 5,
( A 4 I ) 2 {\displaystyle (A-4I)^{2}} has rank 4,
( A 4 I ) 3 {\displaystyle (A-4I)^{3}} has rank 3,
( A 4 I ) 4 {\displaystyle (A-4I)^{4}} has rank 2.

Therefore m 1 = 4. {\displaystyle m_{1}=4.}

ρ 4 = rank ( A 4 I ) 3 rank ( A 4 I ) 4 = 3 2 = 1 , {\displaystyle \rho _{4}=\operatorname {rank} (A-4I)^{3}-\operatorname {rank} (A-4I)^{4}=3-2=1,}
ρ 3 = rank ( A 4 I ) 2 rank ( A 4 I ) 3 = 4 3 = 1 , {\displaystyle \rho _{3}=\operatorname {rank} (A-4I)^{2}-\operatorname {rank} (A-4I)^{3}=4-3=1,}
ρ 2 = rank ( A 4 I ) 1 rank ( A 4 I ) 2 = 5 4 = 1 , {\displaystyle \rho _{2}=\operatorname {rank} (A-4I)^{1}-\operatorname {rank} (A-4I)^{2}=5-4=1,}
ρ 1 = rank ( A 4 I ) 0 rank ( A 4 I ) 1 = 6 5 = 1. {\displaystyle \rho _{1}=\operatorname {rank} (A-4I)^{0}-\operatorname {rank} (A-4I)^{1}=6-5=1.}

Thus, a canonical basis for A {\displaystyle A} will have, corresponding to λ 1 = 4 , {\displaystyle \lambda _{1}=4,} one generalized eigenvector each of ranks 4, 3, 2 and 1.

For λ 2 = 5 , {\displaystyle \lambda _{2}=5,} we have n μ 2 = 6 2 = 4 , {\displaystyle n-\mu _{2}=6-2=4,}

( A 5 I ) {\displaystyle (A-5I)} has rank 5,
( A 5 I ) 2 {\displaystyle (A-5I)^{2}} has rank 4.

Therefore m 2 = 2. {\displaystyle m_{2}=2.}

ρ 2 = rank ( A 5 I ) 1 rank ( A 5 I ) 2 = 5 4 = 1 , {\displaystyle \rho _{2}=\operatorname {rank} (A-5I)^{1}-\operatorname {rank} (A-5I)^{2}=5-4=1,}
ρ 1 = rank ( A 5 I ) 0 rank ( A 5 I ) 1 = 6 5 = 1. {\displaystyle \rho _{1}=\operatorname {rank} (A-5I)^{0}-\operatorname {rank} (A-5I)^{1}=6-5=1.}

Thus, a canonical basis for A {\displaystyle A} will have, corresponding to λ 2 = 5 , {\displaystyle \lambda _{2}=5,} one generalized eigenvector each of ranks 2 and 1.

A canonical basis for A {\displaystyle A} is

{ x 1 , x 2 , x 3 , x 4 , y 1 , y 2 } = { ( 4 0 0 0 0 0 ) , ( 27 4 0 0 0 0 ) , ( 25 25 2 0 0 0 ) , ( 0 36 12 2 2 1 ) , ( 3 2 1 1 0 0 ) , ( 8 4 1 0 1 0 ) } . {\displaystyle \left\{\mathbf {x} _{1},\mathbf {x} _{2},\mathbf {x} _{3},\mathbf {x} _{4},\mathbf {y} _{1},\mathbf {y} _{2}\right\}=\left\{{\begin{pmatrix}-4\\0\\0\\0\\0\\0\end{pmatrix}},{\begin{pmatrix}-27\\-4\\0\\0\\0\\0\end{pmatrix}},{\begin{pmatrix}25\\-25\\-2\\0\\0\\0\end{pmatrix}},{\begin{pmatrix}0\\36\\-12\\-2\\2\\-1\end{pmatrix}},{\begin{pmatrix}3\\2\\1\\1\\0\\0\end{pmatrix}},{\begin{pmatrix}-8\\-4\\-1\\0\\1\\0\end{pmatrix}}\right\}.}

x 1 {\displaystyle \mathbf {x} _{1}} is the ordinary eigenvector associated with λ 1 {\displaystyle \lambda _{1}} . x 2 , x 3 {\displaystyle \mathbf {x} _{2},\mathbf {x} _{3}} and x 4 {\displaystyle \mathbf {x} _{4}} are generalized eigenvectors associated with λ 1 {\displaystyle \lambda _{1}} . y 1 {\displaystyle \mathbf {y} _{1}} is the ordinary eigenvector associated with λ 2 {\displaystyle \lambda _{2}} . y 2 {\displaystyle \mathbf {y} _{2}} is a generalized eigenvector associated with λ 2 {\displaystyle \lambda _{2}} .

A matrix J {\displaystyle J} in Jordan normal form, similar to A {\displaystyle A} is obtained as follows:

M = ( x 1 x 2 x 3 x 4 y 1 y 2 ) = ( 4 27 25 0 3 8 0 4 25 36 2 4 0 0 2 12 1 1 0 0 0 2 1 0 0 0 0 2 0 1 0 0 0 1 0 0 ) , {\displaystyle M={\begin{pmatrix}\mathbf {x} _{1}&\mathbf {x} _{2}&\mathbf {x} _{3}&\mathbf {x} _{4}&\mathbf {y} _{1}&\mathbf {y} _{2}\end{pmatrix}}={\begin{pmatrix}-4&-27&25&0&3&-8\\0&-4&-25&36&2&-4\\0&0&-2&-12&1&-1\\0&0&0&-2&1&0\\0&0&0&2&0&1\\0&0&0&-1&0&0\end{pmatrix}},}
J = ( 4 1 0 0 0 0 0 4 1 0 0 0 0 0 4 1 0 0 0 0 0 4 0 0 0 0 0 0 5 1 0 0 0 0 0 5 ) , {\displaystyle J={\begin{pmatrix}4&1&0&0&0&0\\0&4&1&0&0&0\\0&0&4&1&0&0\\0&0&0&4&0&0\\0&0&0&0&5&1\\0&0&0&0&0&5\end{pmatrix}},}

where the matrix M {\displaystyle M} is a generalized modal matrix for A {\displaystyle A} and A M = M J {\displaystyle AM=MJ} .[10]

See also

Notes

  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

  • 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