Home  

Random  

Nearby  



Log in  



Settings  



Donate  



About Wikipedia  

Disclaimers  



Wikipedia





Structure constants





Article  

Talk  



Language  

Watch  

Edit  





Inmathematics, the structure constantsorstructure coefficients of an algebra over a field are the coefficients of the basis expansion (into linear combination of basis vectors) of the products of basis vectors. Because the product operation in the algebra is bilinear, by linearity knowing the product of basis vectors allows to compute the product of any elements (just like a matrix allows to compute the action of the linear operator on any vector by providing the action of the operator on basis vectors). Therefore, the structure constants can be used to specify the product operation of the algebra (just like a matrix defines a linear operator). Given the structure constants, the resulting product is obtained by bilinearity and can be uniquely extended to all vectors in the vector space, thus uniquely determining the product for the algebra.

Using the cross product as a Lie bracket, the algebra of 3-dimensional real vectors is a Lie algebra isomorphic to the Lie algebras of SU(2) and SO(3). The structure constants are , where is the antisymmetric Levi-Civita symbol.

Structure constants are used whenever an explicit form for the algebra must be given. Thus, they are frequently used when discussing Lie algebrasinphysics, as the basis vectors indicate specific directions in physical space, or correspond to specific particles (recall that Lie algebras are algebras over a field, with the bilinear product being given by the Lie bracket, usually defined via the commutator).

Definition

edit

Given a set of basis vectors   for the underlying vector space of the algebra, the product operation is uniquely defined by the products of basis vectors:

 .

The structure constantsorstructure coefficients   are just the coefficients of   in the same basis:

 .

Otherwise said they are the coefficients that express  aslinear combination of the basis vectors  .

The upper and lower indices are frequently not distinguished, unless the algebra is endowed with some other structure that would require this (for example, a pseudo-Riemannian metric, on the algebra of the indefinite orthogonal group so(p,q)). That is, structure constants are often written with all-upper, or all-lower indexes. The distinction between upper and lower is then a convention, reminding the reader that lower indices behave like the components of a dual vector, i.e. are covariant under a change of basis, while upper indices are contravariant.

The structure constants obviously depend on the chosen basis. For Lie algebras, one frequently used convention for the basis is in terms of the ladder operators defined by the Cartan subalgebra; this is presented further down in the article, after some preliminary examples.

Example: Lie algebras

edit

For a Lie algebra, the basis vectors are termed the generators of the algebra, and the product rather called the Lie bracket (often the Lie bracket is an additional product operation beyond the already existing product, thus necessitating a separate name). For two vectors   and   in the algebra, the Lie bracket is denoted  .

Again, there is no particular need to distinguish the upper and lower indices; they can be written all up or all down. In physics, it is common to use the notation   for the generators, and  or  (ignoring the upper-lower distinction) for the structure constants. The linear expansion of the Lie bracket of pairs of generators then looks like

 .

Again, by linear extension, the structure constants completely determine the Lie brackets of all elements of the Lie algebra.

All Lie algebras satisfy the Jacobi identity. For the basis vectors, it can be written as

 

and this leads directly to a corresponding identity in terms of the structure constants:

 

The above, and the remainder of this article, make use of the Einstein summation convention for repeated indexes.

The structure constants play a role in Lie algebra representations, and in fact, give exactly the matrix elements of the adjoint representation. The Killing form and the Casimir invariant also have a particularly simple form, when written in terms of the structure constants.

The structure constants often make an appearance in the approximation to the Baker–Campbell–Hausdorff formula for the product of two elements of a Lie group. For small elements   of the Lie algebra, the structure of the Lie group near the identity element is given by

 

Note the factor of 1/2. They also appear in explicit expressions for differentials, such as  ; see Baker–Campbell–Hausdorff formula#Infinitesimal case for details.

Lie algebra examples

edit

𝔰𝔲(2) and 𝔰𝔬(3)

edit

The algebra   of the special unitary group SU(2) is three-dimensional, with generators given by the Pauli matrices  . The generators of the group SU(2) satisfy the commutation relations (where   is the Levi-Civita symbol):   where  

In this case, the structure constants are  . Note that the constant 2i can be absorbed into the definition of the basis vectors; thus, defining  , one can equally well write  

Doing so emphasizes that the Lie algebra   of the Lie group SU(2) is isomorphic to the Lie algebra  ofSO(3). This brings the structure constants into line with those of the rotation group SO(3). That is, the commutator for the angular momentum operators are then commonly written as   where   are written so as to obey the right hand rule for rotations in 3-dimensional space.

The difference of the factor of 2i between these two sets of structure constants can be infuriating, as it involves some subtlety. Thus, for example, the two-dimensional complex vector space can be given a real structure. This leads to two inequivalent two-dimensional fundamental representationsof , which are isomorphic, but are complex conjugate representations; both, however, are considered to be real representations, precisely because they act on a space with a real structure.[1] In the case of three dimensions, there is only one three-dimensional representation, the adjoint representation, which is a real representation; more precisely, it is the same as its dual representation, shown above. That is, one has that the transpose is minus itself:  

In any case, the Lie groups are considered to be real, precisely because it is possible to write the structure constants so that they are purely real.

𝔰𝔲(3)

edit

A less trivial example is given by SU(3):[2]

Its generators, T, in the defining representation, are:

 

where  , the Gell-Mann matrices, are the SU(3) analog of the Pauli matrices for SU(2):

     
     
   

These obey the relations

 
 

The structure constants are totally antisymmetric. They are given by:

 
 
 

and all other   not related to these by permuting indices are zero.

The d take the values:

 
 
 

𝔰𝔲(N)

edit

For the general case of 𝔰𝔲(N), there exists closed formula to obtain the structure constant, without having to compute commutation and anti-commutation relations between the generators. We first define the   generators of 𝔰𝔲(N), based on a generalisation of the Pauli matrices and of the Gell-Mann matrices (using the bra-ket notation). There are   symmetric matrices,

 ,

  anti-symmetric matrices,

 ,

and   diagonal matrices,

 .

To differenciate those matrices we define the following indices:

 ,
 ,
 ,

with the condition  .

All the non-zero totally anti-symmetric structure constants are

 ,
 ,
 ,
 .

All the non-zero totally symmetric structure constants are

 ,
 ,
 ,
 ,
 ,
 ,
 ,
 .

For more details on the derivation see [3] and.[4]

Examples from other algebras

edit

Hall polynomials

edit

The Hall polynomials are the structure constants of the Hall algebra.

Hopf algebras

edit

In addition to the product, the coproduct and the antipode of a Hopf algebra can be expressed in terms of structure constants. The connecting axiom, which defines a consistency condition on the Hopf algebra, can be expressed as a relation between these various structure constants.

Applications

edit

Choosing a basis for a Lie algebra

edit

One conventional approach to providing a basis for a Lie algebra is by means of the so-called "ladder operators" appearing as eigenvectors of the Cartan subalgebra. The construction of this basis, using conventional notation, is quickly sketched here. An alternative construction (the Serre construction) can be found in the article semisimple Lie algebra.

Given a Lie algebra  , the Cartan subalgebra   is the maximal Abelian subalgebra. By definition, it consists of those elements that commute with one-another. An orthonormal basis can be freely chosen on  ; write this basis as   with

 

where   is the inner product on the vector space. The dimension   of this subalgebra is called the rank of the algebra. In the adjoint representation, the matrices   are mutually commuting, and can be simultaneously diagonalized. The matrices   have (simultaneous) eigenvectors; those with a non-zero eigenvalue   are conventionally denoted by  . Together with the   these span the entire vector space  . The commutation relations are then

 

The eigenvectors   are determined only up to overall scale; one conventional normalization is to set

 

This allows the remaining commutation relations to be written as

 

and

 

with this last subject to the condition that the roots (defined below)   sum to a non-zero value:  . The   are sometimes called ladder operators, as they have this property of raising/lowering the value of  .

For a given  , there are as many   as there are   and so one may define the vector  , this vector is termed a root of the algebra. The roots of Lie algebras appear in regular structures (for example, in simple Lie algebras, the roots can have only two different lengths); see root system for details.

The structure constants   have the property that they are non-zero only when   are a root. In addition, they are antisymmetric:

 

and can always be chosen such that

 

They also obey cocycle conditions:[7]

 

whenever  , and also that

 

whenever  .

References

edit
  1. ^ Fulton, William; Harris, Joe (1991). Representation theory. A first course. Graduate Texts in Mathematics, Readings in Mathematics. Vol. 129. New York: Springer-Verlag. doi:10.1007/978-1-4612-0979-9. ISBN 978-0-387-97495-8. MR 1153249. OCLC 246650103.
  • ^ Weinberg, Steven (1995). The Quantum Theory of Fields. Vol. 1 Foundations. Cambridge University Press. ISBN 0-521-55001-7.
  • ^ Bossion, D.; Huo, P. (2021). "General Formulas of the Structure Constants in the 𝔰𝔲(N) Lie Algebra". arXiv:2108.07219 [math-ph].
  • ^ Bossion, D.; Ying, W.; Chowdhury, S. N.; Huo, P. (2022). "Non-adiabatic mapping dynamics in the phase space of the SU(N) Lie group". J. Chem. Phys. 157 (8): 084105. Bibcode:2022JChPh.157h4105B. doi:10.1063/5.0094893. PMID 36049982. S2CID 251187368.
  • ^ Raghunathan, Madabusi S. (2012) [1972]. "2. Lattices in Nilpotent Lie Groups". Discrete Subgroups of Lie Groups. Springer. ISBN 978-3-642-86428-5.
  • ^ Eidemüller, M.; Dosch, H.G.; Jamin, M. (2000) [1999]. "The field strength correlator from QCD sum rules". Nucl. Phys. B Proc. Suppl. 86 (1–3): 421–5. arXiv:hep-ph/9908318. Bibcode:2000NuPhS..86..421E. doi:10.1016/S0920-5632(00)00598-3. S2CID 18237543.
  • ^ Cornwell, J.F. (1984). Group Theory In Physics. Vol. 2 Lie Groups and their applications. Academic Press. ISBN 0121898040. OCLC 969857292.

  • Retrieved from "https://en.wikipedia.org/w/index.php?title=Structure_constants&oldid=1224150438"
     



    Last edited on 16 May 2024, at 15:04  





    Languages

     


    Deutsch
    Italiano
    Nederlands

    Português
    Русский

     

    Wikipedia


    This page was last edited on 16 May 2024, at 15:04 (UTC).

    Content is available under CC BY-SA 4.0 unless otherwise noted.



    Privacy policy

    About Wikipedia

    Disclaimers

    Contact Wikipedia

    Code of Conduct

    Developers

    Statistics

    Cookie statement

    Terms of Use

    Desktop