Lie algebra

From Wikipedia, the free encyclopedia

In mathematics, a Lie algebra (pronounced /l/) is a vector space over a field together with a non-associative, alternating bilinear map , called the Lie bracket, satisfying the Jacobi identity.

Lie algebras are closely related to Lie groups, which are groups that are also smooth manifolds, with the property that the group operations of multiplication and inversion are smooth maps. Any Lie group gives rise to a Lie algebra. Conversely, to any finite-dimensional Lie algebra over real or complex numbers, there is a corresponding connected Lie group unique up to covering (Lie's third theorem). This correspondence between Lie groups and Lie algebras allows one to study Lie groups in terms of Lie algebras.

Lie algebras were so termed by Hermann Weyl after Sophus Lie in the 1930s. In older texts, the name infinitesimal group is used.

History[edit]

Lie algebras were introduced to study the concept of infinitesimal transformations by Lie in the 1870s[1], and independently discovered by Wilhelm Killing[2] in the 1880s.

Definitions[edit]

A Lie algebra is a vector space over some field F[nb 1] together with a binary operation called the Lie bracket that satisfies the following axioms:

for all scalars a, b in F and all elements x, y, z in .
for all x in .
for all x, y, z in .

Using bilinearity to expand the Lie bracket and using alternativity shows that for all elements x, y in , showing that bilinearity and alternativity together imply

for all elements x, y in . If the field's characteristic is not 2 then anticommutativity implies alternativity.[3]

It is customary to express a Lie algebra in lower-case fraktur, like . If a Lie algebra is associated with a Lie group, then the spelling of the Lie algebra is the same as that Lie group. For example, the Lie algebra of SU(n) is written as .

Generators and dimension[edit]

Elements of a Lie algebra are said to be generators of the Lie algebra if the smallest subalgebra of containing them is itself. The dimension of a Lie algebra is its dimension as a vector space over F. The cardinality of a minimal generating set of a Lie algebra is always less than or equal to its dimension.

Subalgebras, ideals and homomorphisms[edit]

The Lie bracket is not associative in general, meaning that need not equal . (However, it is flexible.) Nonetheless, much of the terminology that was developed in the theory of associative rings or associative algebras is commonly applied to Lie algebras. A subspace that is closed under the Lie bracket is called a Lie subalgebra. If a subspace satisfies a stronger condition that

then is called an ideal in the Lie algebra .[4] A homomorphism between two Lie algebras (over the same base field) is a linear map that is compatible with the respective Lie brackets:

for all elements x and y in . As in the theory of associative rings, ideals are precisely the kernels of homomorphisms; given a Lie algebra and an ideal in it, one constructs the factor algebra or quotient algebra , and the first isomorphism theorem holds for Lie algebras.

Let S be a subset of . The set of elements x such that for all s in S forms a subalgebra called the centralizer of S. The centralizer of itself is called the center of . Similar to centralizers, if S is a subspace,[5] then the set of x such that is in S for all s in S forms a subalgebra called the normalizer of S.

Direct sum and semidirect product[edit]

Given two Lie algebras and , their direct sum is the Lie algebra consisting of the vector space , of the pairs , with the operation

Let be a Lie algebra and an ideal of . If the canonical map splits (i.e., admits a section), then is said to be a semidirect product of and , . See also semidirect sum of Lie algebras.

Levi's theorem says that a finite-dimensional Lie algebra is a semidirect product of its radical and the complementary subalgebra (Levi subalgebra).

Enveloping algebra[edit]

For any associative algebra A with multiplication , one can construct a Lie algebra L(A). As a vector space, L(A) is the same as A. The Lie bracket of two elements of L(A) is defined to be their commutator in A:

The associativity of the multiplication in A implies the Jacobi identity of the commutator in L(A). For example, the associative algebra of n × n matrices over a field F gives rise to the general linear Lie algebra The associative algebra A is called an enveloping algebra of the Lie algebra L(A). Every Lie algebra can be embedded into one that arises from an associative algebra in this fashion; see universal enveloping algebra.

Representations[edit]

Given a vector space V, let denote the Lie algebra enveloped by the associative algebra of all linear endomorphisms of V. A representation of a Lie algebra on V is a Lie algebra homomorphism

A representation is said to be faithful if its kernel is trivial. Every finite-dimensional Lie algebra has a faithful representation on a finite-dimensional vector space (Ado's theorem).[6]

For example,

given by is a representation of on the vector space called the adjoint representation.

Derivations[edit]

A derivation on the Lie algebra (in fact on any non-associative algebra) is a linear map that obeys the Leibniz law, that is,

for all x and y in the algebra. For any x, is a derivation; a consequence of the Jacobi identity. Thus, the image of lies in the subalgebra of consisting of derivations on . A derivation that happens to be in the image of is called an inner derivation. If is semisimple, every derivation on is inner.

Examples[edit]

Vector spaces[edit]

Any vector space endowed with the identically zero Lie bracket becomes a Lie algebra. Such Lie algebras are called abelian, cf. below. Any one-dimensional Lie algebra over a field is abelian, by the antisymmetry of the Lie bracket.

  • The real vector space of all n × n skew-hermitian matrices is closed under the commutator and forms a real Lie algebra denoted . This is the Lie algebra of the unitary group U(n).

Associative algebra[edit]

  • On an associative algebra over a field with multiplication , a Lie bracket may be defined by the commutator . With this bracket, is a Lie algebra.[7]
  • The associative algebra of endomorphisms of a -vector space with the above Lie bracket is denoted . If , the notation is or .[8]

Subspaces[edit]

Every subalgebra (subspace closed under the Lie bracket) of a Lie algebra is a Lie algebra in its own right.

  • The subspace of the general linear Lie algebra consisting of matrices of trace zero is a subalgebra,[9] the special linear Lie algebra, denoted

Real matrix groups[edit]

  • Any Lie group G defines an associated real Lie algebra =Lie(G). The definition in general is somewhat technical, but in the case of real matrix groups, it can be formulated via the exponential map, or the matrix exponent. The Lie algebra consists of those matrices X for which exp(tX) ∈ G, ∀ real numbers t.
The Lie bracket of is given by the commutator of matrices. As a concrete example, consider the special linear group SL(n,R), consisting of all n × n matrices with real entries and determinant 1. This is a matrix Lie group, and its Lie algebra consists of all n × n matrices with real entries and trace 0.

Two Dimensions[edit]

  • On any field there is, up to isomorphism, a single two-dimensional nonabelian Lie algebra with generators and bracket defined as .

Three dimensions[edit]

.
It is explicitly realized as the space of 3×3 strictly upper-triangular matrices, with the Lie bracket given by the matrix commutator,
Any element of the Heisenberg group is thus representable as a product of group generators, i.e., matrix exponentials of these Lie algebra generators,
  • The commutation relations between the x, y, and z components of the angular momentum operator in quantum mechanics are the same as those of and
.

(The physicist convention for Lie algebras is used in the above equations, hence the factor of i.) The Lie algebra formed by these operators have, in fact, representations of all finite dimensions.

Infinite dimensions[edit]

  • An important class of infinite-dimensional real Lie algebras arises in differential topology. The space of smooth vector fields on a differentiable manifold M forms a Lie algebra, where the Lie bracket is defined to be the commutator of vector fields. One way of expressing the Lie bracket is through the formalism of Lie derivatives, which identifies a vector field X with a first order partial differential operator LX acting on smooth functions by letting LX(f) be the directional derivative of the function f in the direction of X. The Lie bracket [X,Y] of two vector fields is the vector field defined through its action on functions by the formula:

Structure theory and classification[edit]

Lie algebras can be classified to some extent. In particular, this has an application to the classification of Lie groups.

Abelian, nilpotent, and solvable[edit]

Analogously to abelian, nilpotent, and solvable groups, defined in terms of the derived subgroups, one can define abelian, nilpotent, and solvable Lie algebras.

A Lie algebra is abelian if the Lie bracket vanishes, i.e. [x,y] = 0, for all x and y in . Abelian Lie algebras correspond to commutative (or abelian) connected Lie groups such as vector spaces or tori and are all of the form meaning an n-dimensional vector space with the trivial Lie bracket.

A more general class of Lie algebras is defined by the vanishing of all commutators of given length. A Lie algebra is nilpotent if the lower central series

becomes zero eventually. By Engel's theorem, a Lie algebra is nilpotent if and only if for every u in the adjoint endomorphism

is nilpotent.

More generally still, a Lie algebra is said to be solvable if the derived series:

becomes zero eventually.

Every finite-dimensional Lie algebra has a unique maximal solvable ideal, called its radical. Under the Lie correspondence, nilpotent (respectively, solvable) connected Lie groups correspond to nilpotent (respectively, solvable) Lie algebras.

Simple and semisimple[edit]

A Lie algebra is "simple" if it has no non-trivial ideals and is not abelian. A Lie algebra is called semisimple if its radical is zero. Equivalently, is semisimple if it does not contain any non-zero abelian ideals. In particular, a simple Lie algebra is semisimple. Conversely, it can be proven that any semisimple Lie algebra is the direct sum of its minimal ideals, which are canonically determined simple Lie algebras.

The concept of semisimplicity for Lie algebras is closely related with the complete reducibility (semisimplicity) of their representations. When the ground field F has characteristic zero, any finite-dimensional representation of a semisimple Lie algebra is semisimple (i.e., direct sum of irreducible representations.) In general, a Lie algebra is called reductive if the adjoint representation is semisimple. Thus, a semisimple Lie algebra is reductive.

Cartan's criterion[edit]

Cartan's criterion gives conditions for a Lie algebra to be nilpotent, solvable, or semisimple. It is based on the notion of the Killing form, a symmetric bilinear form on defined by the formula

where tr denotes the trace of a linear operator. A Lie algebra is semisimple if and only if the Killing form is nondegenerate. A Lie algebra is solvable if and only if

Classification[edit]

The Levi decomposition expresses an arbitrary Lie algebra as a semidirect sum of its solvable radical and a semisimple Lie algebra, almost in a canonical way. Furthermore, semisimple Lie algebras over an algebraically closed field have been completely classified through their root systems. However, the classification of solvable Lie algebras is a 'wild' problem, and cannot[clarification needed] be accomplished in general.

Relation to Lie groups[edit]

Although Lie algebras are often studied in their own right, historically they arose as a means to study Lie groups.

Lie's fundamental theorems describe a relation between Lie groups and Lie algebras. In particular, any Lie group gives rise to a canonically determined Lie algebra (concretely, the tangent space at the identity); and, conversely, for any finite-dimensional Lie algebra there is a corresponding connected Lie group (Lie's third theorem; see the Baker–Campbell–Hausdorff formula). This Lie group is not determined uniquely; however, any two connected Lie groups with the same Lie algebra are locally isomorphic, and in particular, have the same universal cover. For instance, the special orthogonal group SO(3) and the special unitary group SU(2) give rise to the same Lie algebra, which is isomorphic to R3 with the cross-product, while SU(2) is a simply-connected twofold cover of SO(3).

Given a Lie group, a Lie algebra can be associated to it either by endowing the tangent space to the identity with the differential of the adjoint map, or by considering the left-invariant vector fields as mentioned in the examples. In the case of real matrix groups, the Lie algebra consists of those matrices X for which exp(tX) ∈ G for all real numbers t, where exp is the exponential map.

Some examples of Lie algebras corresponding to Lie groups are the following:

  • The Lie algebra for the group is the algebra of complex n×n matrices
  • The Lie algebra for the group is the algebra of complex n×n matrices with trace 0
  • The Lie algebras for the group and for are both the algebra of real anti-symmetric n×n matrices (See Antisymmetric matrix: Infinitesimal rotations for a discussion)
  • The Lie algebra for the group is the algebra of skew-Hermitian complex n×n matrices while the Lie algebra for is the algebra of skew-Hermitian, traceless complex n×n matrices.

In the above examples, the Lie bracket (for and matrices in the Lie algebra) is defined as .

Given a set of generators Ta, the structure constants f abc express the Lie brackets of pairs of generators as linear combinations of generators from the set, i.e., [Ta, Tb] = f abc Tc. The structure constants determine the Lie brackets of elements of the Lie algebra, and consequently nearly completely determine the group structure of the Lie group. The structure of the Lie group near the identity element is displayed explicitly by the Baker–Campbell–Hausdorff formula, an expansion in Lie algebra elements X, Y and their Lie brackets, all nested together within a single exponent, exp(tX) exp(tY) = exp(tX+tYt2[X,Y] + O(t3) ).

The mapping from Lie groups to Lie algebras is functorial, which implies that homomorphisms of Lie groups lift to homomorphisms of Lie algebras, and various properties are satisfied by this lifting: it commutes with composition, it maps Lie subgroups, kernels, quotients and cokernels of Lie groups to subalgebras, kernels, quotients and cokernels of Lie algebras, respectively.

The functor L that takes each Lie group to its Lie algebra and each homomorphism to its differential is faithful and exact. It is however not an equivalence of categories: different Lie groups may have isomorphic Lie algebras (for example SO(3) and SU(2) ), and there are (infinite dimensional) Lie algebras that are not associated to any Lie group.[10]

However, when the Lie algebra is finite-dimensional, one can associate to it a simply connected Lie group having as its Lie algebra. More precisely, the Lie algebra functor L has a left adjoint functor Γ from finite-dimensional (real) Lie algebras to Lie groups, factoring through the full subcategory of simply connected Lie groups.[11] In other words, there is a natural isomorphism of bifunctors

The adjunction (corresponding to the identity on ) is an isomorphism, and the other adjunction is the projection homomorphism from the universal cover group of the identity component of H to H. It follows immediately that if G is simply connected, then the Lie algebra functor establishes a bijective correspondence between Lie group homomorphisms G→H and Lie algebra homomorphisms L(G)→L(H).

The universal cover group above can be constructed as the image of the Lie algebra under the exponential map. More generally, we have that the Lie algebra is homeomorphic to a neighborhood of the identity. But globally, if the Lie group is compact, the exponential will not be injective, and if the Lie group is not connected, simply connected or compact, the exponential map need not be surjective.

If the Lie algebra is infinite-dimensional, the issue is more subtle. In many instances, the exponential map is not even locally a homeomorphism (for example, in Diff(S1), one may find diffeomorphisms arbitrarily close to the identity that are not in the image of exp). Furthermore, some infinite-dimensional Lie algebras are not the Lie algebra of any group.

The correspondence between Lie algebras and Lie groups is used in several ways, including in the classification of Lie groups and the related matter of the representation theory of Lie groups. Every representation of a Lie algebra lifts uniquely to a representation of the corresponding connected, simply connected Lie group, and conversely every representation of any Lie group induces a representation of the group's Lie algebra; the representations are in one-to-one correspondence. Therefore, knowing the representations of a Lie algebra settles the question of representations of the group.

As for classification, it can be shown that any connected Lie group with a given Lie algebra is isomorphic to the universal cover mod a discrete central subgroup. So classifying Lie groups becomes simply a matter of counting the discrete subgroups of the center, once the classification of Lie algebras is known (solved by Cartan et al. in the semisimple case).

Category theoretic definition[edit]

Using the language of category theory, a Lie algebra can be defined as an object A in Veck, the category of vector spaces over a field k of characteristic not 2, together with a morphism [.,.]: AAA, where ⊗ refers to the monoidal product of Veck, such that

where τ(ab) := ba and σ is the cyclic permutation braiding (id ⊗ τA,A) ° (τA,A ⊗ id). In diagrammatic form:

Lie ring[edit]

A Lie ring arises as a generalisation of Lie algebras, or through the study of the lower central series of groups. A Lie ring is defined as a nonassociative ring with multiplication that is anticommutative and satisfies the Jacobi identity. More specifically we can define a Lie ring to be an abelian group with an operation that has the following properties:

  • Bilinearity:
for all x, y, zL.
  • The Jacobi identity:
for all x, y, z in L.
  • For all x in L:

Lie rings need not be Lie groups under addition. Any Lie algebra is an example of a Lie ring. Any associative ring can be made into a Lie ring by defining a bracket operator . Conversely to any Lie algebra there is a corresponding ring, called the universal enveloping algebra.

Lie rings are used in the study of finite p-groups through the Lazard correspondence'. The lower central factors of a p-group are finite abelian p-groups, so modules over Z/pZ. The direct sum of the lower central factors is given the structure of a Lie ring by defining the bracket to be the commutator of two coset representatives. The Lie ring structure is enriched with another module homomorphism, the pth power map, making the associated Lie ring a so-called restricted Lie ring.

Lie rings are also useful in the definition of a p-adic analytic groups and their endomorphisms by studying Lie algebras over rings of integers such as the p-adic integers. The definition of finite groups of Lie type due to Chevalley involves restricting from a Lie algebra over the complex numbers to a Lie algebra over the integers, and the reducing modulo p to get a Lie algebra over a finite field.

Examples[edit]

  • Any Lie algebra over a general ring instead of a field is an example of a Lie ring. Lie rings are not Lie groups under addition, despite the name.
  • Any associative ring can be made into a Lie ring by defining a bracket operator .
  • For an example of a Lie ring arising from the study of groups, let be a group with the commutator operation, and let be a central series in — that is the commutator subgroup is contained in for any . Then
is a Lie ring with addition supplied by the group operation (which will be commutative in each homogeneous part), and the bracket operation given by
extended linearly. Note that the centrality of the series ensures the commutator gives the bracket operation the appropriate Lie theoretic properties.

See also[edit]

Remarks[edit]

  1. ^ Script error: The function "harvard_core" does not exist. allows more generally for a module over a commutative ring with unit element.

Notes[edit]

  1. ^ O'Connor & Robertson 2000
  2. ^ O'Connor & Robertson 2005
  3. ^ Humphreys 1978, p. 1
  4. ^ Due to the anticommutativity of the commutator, the notions of a left and right ideal in a Lie algebra coincide.
  5. ^ Jacobson 1962, pg. 28
  6. ^ Jacobson 1962, Ch. VI
  7. ^ Bourbaki 1989, §1.2. Example 1.
  8. ^ Bourbaki 1989, §1.2. Example 2.
  9. ^ Humphreys p.2
  10. ^ Beltita 2005, pg. 75
  11. ^ Adjoint property is discussed in more general context in Hofman & Morris (2007) (e.g., page 130) but is a straightforward consequence of, e.g., Bourbaki (1989) Theorem 1 of page 305 and Theorem 3 of page 310.

References[edit]

  • Beltita, Daniel. Smooth Homogeneous Structures in Operator Theory, CRC Press, 2005. ISBN 978-1-4200-3480-6
  • Boza, Luis; Fedriani, Eugenio M. & Núñez, Juan. A new method for classifying complex filiform Lie algebras, Applied Mathematics and Computation, 121 (2-3): 169–175, 2001
  • Bourbaki, N. (1989). Lie Groups and Lie Algebras: Chapters 1-3. Berlin·Heidelberg·New York: Springer. ISBN 978-3-540-64242-8. {{cite book}}: Invalid |ref=harv (help)
  • Erdmann, Karin & Wildon, Mark. Introduction to Lie Algebras, 1st edition, Springer, 2006. ISBN 1-84628-040-0
  • Hofman, Karl & Morris, Sidney. "The Lie Theory of Connected Pro-Lie Groups", European Mathematical Society, 2007, ISBN 978-3-03719-032-6
  • Steeb, W.-H. Continuous Symmetries, Lie Algebras, Differential Equations and Computer Algebra, second edition, World Scientific, 2007, ISBN 978-981-270-809-0

External links[edit]