Complexification (Lie group)

Last updated

In mathematics, the complexification or universal complexification of a real Lie group is given by a continuous homomorphism of the group into a complex Lie group with the universal property that every continuous homomorphism of the original group into another complex Lie group extends compatibly to a complex analytic homomorphism between the complex Lie groups. The complexification, which always exists, is unique up to unique isomorphism. Its Lie algebra is a quotient of the complexification of the Lie algebra of the original group. They are isomorphic if the original group has a quotient by a discrete normal subgroup which is linear.

Contents

For compact Lie groups, the complexification, sometimes called the Chevalley complexification after Claude Chevalley, can be defined as the group of complex characters of the Hopf algebra of representative functions, i.e. the matrix coefficients of finite-dimensional representations of the group. In any finite-dimensional faithful unitary representation of the compact group it can be realized concretely as a closed subgroup of the complex general linear group. It consists of operators with polar decomposition g = u • exp iX, where u is a unitary operator in the compact group and X is a skew-adjoint operator in its Lie algebra. In this case the complexification is a complex algebraic group and its Lie algebra is the complexification of the Lie algebra of the compact Lie group.

Universal complexification

Definition

If G is a Lie group, a universal complexification is given by a complex Lie group GC and a continuous homomorphism φ: GGC with the universal property that, if f: GH is an arbitrary continuous homomorphism into a complex Lie group H, then there is a unique complex analytic homomorphism F: GCH such that f = Fφ.

Universal complexifications always exist and are unique up to a unique complex analytic isomorphism (preserving inclusion of the original group).

Existence

If G is connected with Lie algebra 𝖌, then its universal covering group G is simply connected. Let GC be the simply connected complex Lie group with Lie algebra 𝖌C = 𝖌 ⊗ C, let Φ: GGC be the natural homomorphism (the unique morphism such that Φ*: 𝖌 ↪ 𝖌 ⊗ C is the canonical inclusion) and suppose π: GG is the universal covering map, so that ker π is the fundamental group of G. We have the inclusion Φ(ker π) ⊂ Z(GC), which follows from the fact that the kernel of the adjoint representation of GC equals its centre, combined with the equality

which holds for any k ∈ ker π. Denoting by Φ(ker π)* the smallest closed normal Lie subgroup of GC that contains Φ(ker π), we must now also have the inclusion Φ(ker π)* ⊂ Z(GC). We define the universal complexification of G as

In particular, if G is simply connected, its universal complexification is just GC. [1]

The map φ: GGC is obtained by passing to the quotient. Since π is a surjective submersion, smoothness of the map πC ∘ Φ implies smoothness of φ.

Construction of the complexification map Complexification-quotient-map.svg
Construction of the complexification map

For non-connected Lie groups G with identity component Go and component group Γ = G / Go, the extension

induces an extension

and the complex Lie group GC is a complexification of G. [2]

Proof of the universal property

The map φ: GGC indeed possesses the universal property which appears in the above definition of complexification. The proof of this statement naturally follows from considering the following instructive diagram.

Universal property of complexification.svg

Here, is an arbitrary smooth homomorphism of Lie groups with a complex Lie group as the codomain.

Existence of the map F

For simplicity, we assume is connected. To establish the existence of , we first naturally extend the morphism of Lie algebras to the unique morphism of complex Lie algebras. Since is simply connected, Lie's second fundamental theorem now provides us with a unique complex analytic morphism between complex Lie groups, such that . We define as the map induced by , that is: for any . To show well-definedness of this map (i.e. ), consider the derivative of the map . For any , we have

,

which (by simple connectedness of ) implies . This equality finally implies , and since is a closed normal Lie subgroup of , we also have . Since is a complex analytic surjective submersion, the map is complex analytic since is. The desired equality is imminent.

Uniqueness of the map F

To show uniqueness of , suppose that are two maps with . Composing with from the right and differentiating, we get , and since is the inclusion , we get . But is a submersion, so , thus connectedness of implies .

Uniqueness

The universal property implies that the universal complexification is unique up to complex analytic isomorphism.

Injectivity

If the original group is linear, so too is the universal complexification and the homomorphism between the two is an inclusion. [3] Onishchik & Vinberg (1994) give an example of a connected real Lie group for which the homomorphism is not injective even at the Lie algebra level: they take the product of T by the universal covering group of SL(2,R) and quotient out by the discrete cyclic subgroup generated by an irrational rotation in the first factor and a generator of the center in the second.

Basic examples

The following isomorphisms of complexifications of Lie groups with known Lie groups can be constructed directly from the general construction of the complexification.

.
This follows from the isomorphism of Lie algebras
,
together with the fact that is simply connected.
.
This follows from the isomorphism of Lie algebras
,
together with the fact that is simply connected.
,
where denotes the proper orthochronous Lorentz group. This follows from the fact that is the universal (double) cover of , hence:
.
We also use the fact that is the universal (double) cover of .
.
This follows from the same isomorphism of Lie algebras as in the second example, again using the universal (double) cover of the proper orthochronous Lorentz group.
.
This follows from the fact that is the universal (double) cover of , hence and so .

The last two examples show that Lie groups with isomorphic complexifications may not be isomorphic. Furthermore, the complexifications of Lie groups and show that complexification is not an idempotent operation, i.e. (this is also shown by complexifications of and ).

Chevalley complexification

Hopf algebra of matrix coefficients

If Gis a compact Lie group, the *-algebra A of matrix coefficients of finite-dimensional unitary representations is a uniformly dense *-subalgebra of C(G), the *-algebra of complex-valued continuous functions on G. It is naturally a Hopf algebra with comultiplication given by

The characters of A are the *-homomorphisms of A into C. They can be identified with the point evaluations ff(g) for g in G and the comultiplication allows the group structure on G to be recovered. The homomorphisms of A into C also form a group. It is a complex Lie group and can be identified with the complexification GC of G. The *-algebra A is generated by the matrix coefficients of any faithful representation σ of G. It follows that σ defines a faithful complex analytic representation of GC. [4]

Invariant theory

The original approach of Chevalley (1946) to the complexification of a compact Lie group can be concisely stated within the language of classical invariant theory, described in Weyl (1946). Let G be a closed subgroup of the unitary group U(V) where V is a finite-dimensional complex inner product space. Its Lie algebra consists of all skew-adjoint operators X such that exp tX lies in G for all real t. Set W = VC with the trivial action of G on the second summand. The group G acts on WN, with an element u acting as uN. The commutant (or centralizer algebra) is denoted by AN = EndGWN. It is generated as a *-algebra by its unitary operators and its commutant is the *-algebra spanned by the operators uN. The complexification GC of G consists of all operators g in GL(V) such that gN commutes with AN and g acts trivially on the second summand in C. By definition it is a closed subgroup of GL(V). The defining relations (as a commutant) show that G is an algebraic subgroup. Its intersection with U(V) coincides with G, since it is a priori a larger compact group for which the irreducible representations stay irreducible and inequivalent when restricted to G. Since AN is generated by unitaries, an invertible operator g lies in GC if the unitary operator u and positive operator p in its polar decomposition g = up both lie in GC. Thus u lies in G and the operator p can be written uniquely as p = exp T with T a self-adjoint operator. By the functional calculus for polynomial functions it follows that hN lies in the commutant of AN if h = exp zT with z in C. In particular taking z purely imaginary, T must have the form iX with X in the Lie algebra of G. Since every finite-dimensional representation of G occurs as a direct summand of WN, it is left invariant by GC and thus every finite-dimensional representation of G extends uniquely to GC. The extension is compatible with the polar decomposition. Finally the polar decomposition implies that G is a maximal compact subgroup of GC, since a strictly larger compact subgroup would contain all integer powers of a positive operator p, a closed infinite discrete subgroup. [5]

Decompositions in the Chevalley complexification

Cartan decomposition

The decomposition derived from the polar decomposition

where 𝖌 is the Lie algebra of G, is called the Cartan decomposition of GC. The exponential factor P is invariant under conjugation by G but is not a subgroup. The complexification is invariant under taking adjoints, since G consists of unitary operators and P of positive operators.

Gauss decomposition

The Gauss decomposition is a generalization of the LU decomposition for the general linear group and a specialization of the Bruhat decomposition. For GL(V) it states that with respect to a given orthonormal basis e1, ..., en an element g of GL(V) can be factorized in the form

with X lower unitriangular, Y upper unitriangular and D diagonal if and only if all the principal minors of g are non-vanishing. In this case X, Y and D are uniquely determined.

In fact Gaussian elimination shows there is a unique X such that X−1g is upper triangular. [6]

The upper and lower unitriangular matrices, N+ and N, are closed unipotent subgroups of GL(V). Their Lie algebras consist of upper and lower strictly triangular matrices. The exponential mapping is a polynomial mapping from the Lie algebra to the corresponding subgroup by nilpotence. The inverse is given by the logarithm mapping which by unipotence is also a polynomial mapping. In particular there is a correspondence between closed connected subgroups of N± and subalgebras of their Lie algebras. The exponential map is onto in each case, since the polynomial function log ( eAeB ) lies in a given Lie subalgebra if A and B do and are sufficiently small. [7]

The Gauss decomposition can be extended to complexifications of other closed connected subgroups G of U(V) by using the root decomposition to write the complexified Lie algebra as [8]

where 𝖙 is the Lie algebra of a maximal torus T of G and 𝖓± are the direct sum of the corresponding positive and negative root spaces. In the weight space decomposition of V as eigenspaces of T, 𝖙 acts as diagonally, 𝖓+ acts as lowering operators and 𝖓 as raising operators. 𝖓± are nilpotent Lie algebras acting as nilpotent operators; they are each other's adjoints on V. In particular T acts by conjugation of 𝖓+, so that 𝖙C ⊕ 𝖓+ is a semidirect product of a nilpotent Lie algebra by an abelian Lie algebra.

By Engel's theorem, if 𝖆 ⊕ 𝖓 is a semidirect product, with 𝖆 abelian and 𝖓 nilpotent, acting on a finite-dimensional vector space W with operators in 𝖆 diagonalizable and operators in 𝖓 nilpotent, there is a vector w that is an eigenvector for 𝖆 and is annihilated by 𝖓. In fact it is enough to show there is a vector annihilated by 𝖓, which follows by induction on dim 𝖓, since the derived algebra 𝖓' annihilates a non-zero subspace of vectors on which 𝖓 / 𝖓' and 𝖆 act with the same hypotheses.

Applying this argument repeatedly to 𝖙C ⊕ 𝖓+ shows that there is an orthonormal basis e1, ..., en of V consisting of eigenvectors of 𝖙C with 𝖓+ acting as upper triangular matrices with zeros on the diagonal.

If N± and TC are the complex Lie groups corresponding to 𝖓+ and 𝖙C, then the Gauss decomposition states that the subset

is a direct product and consists of the elements in GC for which the principal minors are non-vanishing. It is open and dense. Moreover, if T denotes the maximal torus in U(V),

These results are an immediate consequence of the corresponding results for GL(V). [9]

Bruhat decomposition

If W = NG(T) / T denotes the Weyl group of T and B denotes the Borel subgroup TCN+, the Gauss decomposition is also a consequence of the more precise Bruhat decomposition

decomposing GC into a disjoint union of double cosets of B. The complex dimension of a double coset BσB is determined by the length of σ as an element of W. The dimension is maximized at the Coxeter element and gives the unique open dense double coset. Its inverse conjugates B into the Borel subgroup of lower triangular matrices in GC. [10]

The Bruhat decomposition is easy to prove for SL(n,C). [11] Let B be the Borel subgroup of upper triangular matrices and TC the subgroup of diagonal matrices. So N(TC) / TC = Sn. For g in SL(n,C), take b in B so that bg maximizes the number of zeros appearing at the beginning of its rows. Because a multiple of one row can be added to another, each row has a different number of zeros in it. Multiplying by a matrix w in N(TC), it follows that wbg lies in B. For uniqueness, if w1bw2 = b0, then the entries of w1w2 vanish below the diagonal. So the product lies in TC, proving uniqueness.

Chevalley (1955) showed that the expression of an element g as g = b1σb2 becomes unique if b1 is restricted to lie in the upper unitriangular subgroup Nσ = N+σ Nσ−1. In fact, if Mσ = N+σ N+σ−1, this follows from the identity

The group N+ has a natural filtration by normal subgroups N+(k) with zeros in the first k − 1 superdiagonals and the successive quotients are Abelian. Defining Nσ(k) and Mσ(k) to be the intersections with N+(k), it follows by decreasing induction on k that N+(k) = Nσ(k) ⋅ Mσ(k). Indeed, Nσ(k)N+(k + 1) and Mσ(k)N+(k + 1) are specified in N+(k) by the vanishing of complementary entries (i, j) on the kth superdiagonal according to whether σ preserves the order i < j or not. [12]

The Bruhat decomposition for the other classical simple groups can be deduced from the above decomposition using the fact that they are fixed point subgroups of folding automorphisms of SL(n,C). [13] For Sp(n,C), let J be the n × n matrix with 1's on the antidiagonal and 0's elsewhere and set

Then Sp(n,C) is the fixed point subgroup of the involution θ(g) = A (gt)−1A−1 of SL(2n,C). It leaves the subgroups N±, TC and B invariant. If the basis elements are indexed by n, n−1, ..., 1, −1, ..., −n, then the Weyl group of Sp(n,C) consists of σ satisfying σ(j) = −j, i.e. commuting with θ. Analogues of B, TC and N± are defined by intersection with Sp(n,C), i.e. as fixed points of θ. The uniqueness of the decomposition g = nσb = θ(n) θ(σ) θ(b) implies the Bruhat decomposition for Sp(n,C).

The same argument works for SO(n,C). It can be realised as the fixed points of ψ(g) = B (gt)−1B−1 in SL(n,C) where B = J.

Iwasawa decomposition

The Iwasawa decomposition

gives a decomposition for GC for which, unlike the Cartan decomposition, the direct factor AN is a closed subgroup, but it is no longer invariant under conjugation by G. It is the semidirect product of the nilpotent subgroup N by the Abelian subgroup A.

For U(V) and its complexification GL(V), this decomposition can be derived as a restatement of the Gram–Schmidt orthonormalization process. [14]

In fact let e1, ..., en be an orthonormal basis of V and let g be an element in GL(V). Applying the Gram–Schmidt process to ge1, ..., gen, there is a unique orthonormal basis f1, ..., fn and positive constants ai such that

If k is the unitary taking (ei) to (fi), it follows that g−1k lies in the subgroup AN, where A is the subgroup of positive diagonal matrices with respect to (ei) and N is the subgroup of upper unitriangular matrices. [15]

Using the notation for the Gauss decomposition, the subgroups in the Iwasawa decomposition for GC are defined by [16]

Since the decomposition is direct for GL(V), it is enough to check that GC = GAN. From the properties of the Iwasawa decomposition for GL(V), the map G × A × N is a diffeomorphism onto its image in GC, which is closed. On the other hand, the dimension of the image is the same as the dimension of GC, so it is also open. So GC = GAN because GC is connected. [17]

Zhelobenko (1973) gives a method for explicitly computing the elements in the decomposition. [18] For g in GC set h = g*g. This is a positive self-adjoint operator so its principal minors do not vanish. By the Gauss decomposition, it can therefore be written uniquely in the form h = XDY with X in N, D in TC and Y in N+. Since h is self-adjoint, uniqueness forces Y = X*. Since it is also positive D must lie in A and have the form D = exp iT for some unique T in 𝖙. Let a = exp iT/2 be its unique square root in A. Set n = Y and k = gn−1a−1. Then k is unitary, so is in G, and g = kan.

Complex structures on homogeneous spaces

The Iwasawa decomposition can be used to describe complex structures on the G-orbit s in complex projective space of highest weight vectors of finite-dimensional irreducible representations of G. In particular the identification between G / T and GC / B can be used to formulate the Borel–Weil theorem. It states that each irreducible representation of G can be obtained by holomorphic induction from a character of T, or equivalently that it is realized in the space of sections of a holomorphic line bundle on G / T.

The closed connected subgroups of G containing T are described by Borel–de Siebenthal theory. They are exactly the centralizers of tori ST. Since every torus is generated topologically by a single element x, these are the same as centralizers CG(X) of elements X in 𝖙. By a result of Hopf CG(x) is always connected: indeed any element y is along with S contained in some maximal torus, necessarily contained in CG(x).

Given an irreducible finite-dimensional representation Vλ with highest weight vector v of weight λ, the stabilizer of Cv in G is a closed subgroup H. Since v is an eigenvector of T, H contains T. The complexification GC also acts on V and the stabilizer is a closed complex subgroup P containing TC. Since v is annihilated by every raising operator corresponding to a positive root α, P contains the Borel subgroup B. The vector v is also a highest weight vector for the copy of sl2 corresponding to α, so it is annihilated by the lowering operator generating 𝖌α if (λ, α) = 0. The Lie algebra p of P is the direct sum of 𝖙C and root space vectors annihilating v, so that

The Lie algebra of H = PG is given by p ∩ 𝖌. By the Iwasawa decomposition GC = GAN. Since AN fixes Cv, the G-orbit of v in the complex projective space of Vλ coincides with the GC orbit and

In particular

Using the identification of the Lie algebra of T with its dual, H equals the centralizer of λ in G, and hence is connected. The group P is also connected. In fact the space G / H is simply connected, since it can be written as the quotient of the (compact) universal covering group of the compact semisimple group G / Z by a connected subgroup, where Z is the center of G. [19] If Po is the identity component of P, GC / P has GC / Po as a covering space, so that P = Po. The homogeneous space GC / P has a complex structure, because P is a complex subgroup. The orbit in complex projective space is closed in the Zariski topology by Chow's theorem, so is a smooth projective variety. The Borel–Weil theorem and its generalizations are discussed in this context in Serre (1954), Helgason (1994), Duistermaat & Kolk (2000) and Sepanski (2007).

The parabolic subgroup P can also be written as a union of double cosets of B

where Wλ is the stabilizer of λ in the Weyl group W. It is generated by the reflections corresponding to the simple roots orthogonal to λ. [20]

Noncompact real forms

There are other closed subgroups of the complexification of a compact connected Lie group G which have the same complexified Lie algebra. These are the other real forms of GC. [21]

Involutions of simply connected compact Lie groups

If G is a simply connected compact Lie group and σ is an automorphism of order 2, then the fixed point subgroup K = Gσ is automatically connected. (In fact this is true for any automorphism of G, as shown for inner automorphisms by Steinberg and in general by Borel.) [22]

This can be seen most directly when the involution σ corresponds to a Hermitian symmetric space. In that case σ is inner and implemented by an element in a one-parameter subgroup exp tT contained in the center of Gσ. The innerness of σ implies that K contains a maximal torus of G, so has maximal rank. On the other hand, the centralizer of the subgroup generated by the torus S of elements exp tT is connected, since if x is any element in K there is a maximal torus containing x and S, which lies in the centralizer. On the other hand, it contains K since S is central in K and is contained in K since z lies in S. So K is the centralizer of S and hence connected. In particular K contains the center of G. [23]

For a general involution σ, the connectedness of Gσ can be seen as follows. [24]

The starting point is the Abelian version of the result: if T is a maximal torus of a simply connected group G and σ is an involution leaving invariant T and a choice of positive roots (or equivalently a Weyl chamber), then the fixed point subgroup Tσ is connected. In fact the kernel of the exponential map from onto T is a lattice Λ with a Z-basis indexed by simple roots, which σ permutes. Splitting up according to orbits, T can be written as a product of terms T on which σ acts trivially or terms T2 where σ interchanges the factors. The fixed point subgroup just corresponds to taking the diagonal subgroups in the second case, so is connected.

Now let x be any element fixed by σ, let S be a maximal torus in CG(x)σ and let T be the identity component of CG(x, S). Then T is a maximal torus in G containing x and S. It is invariant under σ and the identity component of Tσ is S. In fact since x and S commute, they are contained in a maximal torus which, because it is connected, must lie in T. By construction T is invariant under σ. The identity component of Tσ contains S, lies in CG(x)σ and centralizes S, so it equals S. But S is central in T, to T must be Abelian and hence a maximal torus. For σ acts as multiplication by −1 on the Lie algebra , so it and therefore also are Abelian.

The proof is completed by showing that σ preserves a Weyl chamber associated with T. For then Tσ is connected so must equal S. Hence x lies in S. Since x was arbitrary, Gσ must therefore be connected.

To produce a Weyl chamber invariant under σ, note that there is no root space on which both x and S acted trivially, for this would contradict the fact that CG(x, S) has the same Lie algebra as T. Hence there must be an element s in S such that t = xs acts non-trivially on each root space. In this case t is a regular element of T—the identity component of its centralizer in G equals T. There is a unique Weyl alcove A in such that t lies in exp A and 0 lies in the closure of A. Since t is fixed by σ, the alcove is left invariant by σ and hence so also is the Weyl chamber C containing it.

Conjugations on the complexification

Let G be a simply connected compact Lie group with complexification GC. The map c(g) = (g*)−1 defines an automorphism of GC as a real Lie group with G as fixed point subgroup. It is conjugate-linear on and satisfies c2 = id. Such automorphisms of either GC or are called conjugations. Since GC is also simply connected any conjugation c1 on corresponds to a unique automorphism c1 of GC.

The classification of conjugations c0 reduces to that of involutions σ of G because given a c1 there is an automorphism φ of the complex group GC such that

commutes with c. The conjugation c0 then leaves G invariant and restricts to an involutive automorphism σ. By simple connectivity the same is true at the level of Lie algebras. At the Lie algebra level c0 can be recovered from σ by the formula

for X, Y in .

To prove the existence of φ let ψ = c1c an automorphism of the complex group GC. On the Lie algebra level it defines a self-adjoint operator for the complex inner product

where B is the Killing form on . Thus ψ2 is a positive operator and an automorphism along with all its real powers. In particular take

It satisfies

Cartan decomposition in a real form

For the complexification GC, the Cartan decomposition is described above. Derived from the polar decomposition in the complex general linear group, it gives a diffeomorphism

On GC there is a conjugation operator c corresponding to G as well as an involution σ commuting with c. Let c0 = c σ and let G0 be the fixed point subgroup of c. It is closed in the matrix group GC and therefore a Lie group. The involution σ acts on both G and G0. For the Lie algebra of G there is a decomposition

into the +1 and −1 eigenspaces of σ. The fixed point subgroup K of σ in G is connected since G is simply connected. Its Lie algebra is the +1 eigenspace . The Lie algebra of G0 is given by

and the fixed point subgroup of σ is again K, so that GG0 = K. In G0, there is a Cartan decomposition

which is again a diffeomorphism onto the direct and corresponds to the polar decomposition of matrices. It is the restriction of the decomposition on GC. The product gives a diffeomorphism onto a closed subset of G0. To check that it is surjective, for g in G0 write g = up with u in G and p in P. Since c0g = g, uniqueness implies that σu = u and σp = p−1. Hence u lies in K and p in P0.

The Cartan decomposition in G0 shows that G0 is connected, simply connected and noncompact, because of the direct factor P0. Thus G0 is a noncompact real semisimple Lie group. [25]

Moreover, given a maximal Abelian subalgebra in , A = exp is a toral subgroup such that σ(a) = a−1 on A; and any two such 's are conjugate by an element of K. The properties of A can be shown directly. A is closed because the closure of A is a toral subgroup satisfying σ(a) = a−1, so its Lie algebra lies in and hence equals by maximality. A can be generated topologically by a single element exp X, so is the centralizer of X in . In the K-orbit of any element of there is an element Y such that (X,Ad k Y) is minimized at k = 1. Setting k = exp tT with T in , it follows that (X,[T,Y]) = 0 and hence [X,Y] = 0, so that Y must lie in . Thus is the union of the conjugates of . In particular some conjugate of X lies in any other choice of , which centralizes that conjugate; so by maximality the only possibilities are conjugates of . [26]

A similar statements hold for the action of K on in . Morevoer, from the Cartan decomposition for G0, if A0 = exp , then

Iwasawa decomposition in a real form

See also

Notes

  1. See:
  2. Bourbaki 1981 , pp. 210–214
  3. Hochschild 1966
  4. See:
  5. See:
  6. Zhelobenko 1973 , p. 28
  7. Bump 2004 , pp. 202–203
  8. See:
  9. Zhelobenko 1973
  10. See:
  11. Steinberg 1974 , p. 73
  12. Chevalley 1955 , p. 41
  13. See:
  14. Sepanski 2007 , p. 8
  15. Knapp 2001 , p. 117
  16. See:
  17. Bump 2004 , pp. 203–204
  18. Zhelobenko 1973 , p. 289
  19. Helgason 1978
  20. See:
  21. Dieudonné 1977 , pp. 164–173
  22. See:
  23. Wolf 2010
  24. See: Bourbaki 1982 , pp. 46–48
  25. Dieudonné 1977 , pp. 166–168
  26. Helgason 1978 , p. 248

Related Research Articles

<span class="mw-page-title-main">Lie algebra</span> Vector space with a binary operation satisfying the Jacobi identity

In mathematics, a Lie algebra is a vector space together with an operation called the Lie bracket, an alternating bilinear map , that satisfies the Jacobi identity. The Lie bracket of two vectors and is denoted . The vector space together with this operation is a non-associative algebra, meaning that the Lie bracket is not necessarily associative.

<span class="mw-page-title-main">Lie group</span> Group that is also a differentiable manifold with group operations that are smooth

In mathematics, a Lie group is a group that is also a differentiable manifold. A manifold is a space that locally resembles Euclidean space, whereas groups define the abstract concept of a binary operation along with the additional properties it must have to be a group, for instance multiplication and the taking of inverses (division), or equivalently, the concept of addition and the taking of inverses (subtraction). Combining these two ideas, one obtains a continuous group where multiplying points and their inverses are continuous. If the multiplication and taking of inverses are smooth (differentiable) as well, one obtains a Lie group.

<span class="mw-page-title-main">Adjoint representation</span>

In mathematics, the adjoint representation of a Lie group G is a way of representing the elements of the group as linear transformations of the group's Lie algebra, considered as a vector space. For example, if G is , the Lie group of real n-by-n invertible matrices, then the adjoint representation is the group homomorphism that sends an invertible n-by-n matrix to an endomorphism of the vector space of all linear transformations of defined by: .

In the field of representation theory in mathematics, a projective representation of a group G on a vector space V over a field F is a group homomorphism from G to the projective linear group

<span class="mw-page-title-main">Spin group</span> Double cover Lie group of the special orthogonal group

In mathematics the spin group Spin(n) is the double cover of the special orthogonal group SO(n) = SO(n, R), such that there exists a short exact sequence of Lie groups

In mathematics, the Iwasawa decomposition of a semisimple Lie group generalises the way a square real matrix can be written as a product of an orthogonal matrix and an upper triangular matrix. It is named after Kenkichi Iwasawa, the Japanese mathematician who developed this method.

<span class="mw-page-title-main">Cartan subalgebra</span> Nilpotent subalgebra of a Lie algebra

In mathematics, a Cartan subalgebra, often abbreviated as CSA, is a nilpotent subalgebra of a Lie algebra that is self-normalising. They were introduced by Élie Cartan in his doctoral thesis. It controls the representation theory of a semi-simple Lie algebra over a field of characteristic .

<span class="mw-page-title-main">Semisimple Lie algebra</span> Direct sum of simple Lie algebras

In mathematics, a Lie algebra is semisimple if it is a direct sum of simple Lie algebras..

In mathematics, a maximal compact subgroupK of a topological group G is a subgroup K that is a compact space, in the subspace topology, and maximal amongst such subgroups.

<span class="mw-page-title-main">Symmetric space</span> A (pseudo-)Riemannian manifold whose geodesics are reversible.

In mathematics, a symmetric space is a Riemannian manifold whose group of symmetries contains an inversion symmetry about every point. This can be studied with the tools of Riemannian geometry, leading to consequences in the theory of holonomy; or algebraically through Lie theory, which allowed Cartan to give a complete classification. Symmetric spaces commonly occur in differential geometry, representation theory and harmonic analysis.

<span class="mw-page-title-main">Representation theory of the Lorentz group</span> Representation of the symmetry group of spacetime in special relativity

The Lorentz group is a Lie group of symmetries of the spacetime of special relativity. This group can be realized as a collection of matrices, linear transformations, or unitary operators on some Hilbert space; it has a variety of representations. This group is significant because special relativity together with quantum mechanics are the two physical theories that are most thoroughly established, and the conjunction of these two theories is the study of the infinite-dimensional unitary representations of the Lorentz group. These have both historical importance in mainstream physics, as well as connections to more speculative present-day theories.

<span class="mw-page-title-main">Hermitian symmetric space</span> Manifold with inversion symmetry

In mathematics, a Hermitian symmetric space is a Hermitian manifold which at every point has an inversion symmetry preserving the Hermitian structure. First studied by Élie Cartan, they form a natural generalization of the notion of Riemannian symmetric space from real manifolds to complex manifolds.

<span class="mw-page-title-main">Classical group</span>

In mathematics, the classical groups are defined as the special linear groups over the reals R, the complex numbers C and the quaternions H together with special automorphism groups of symmetric or skew-symmetric bilinear forms and Hermitian or skew-Hermitian sesquilinear forms defined on real, complex and quaternionic finite-dimensional vector spaces. Of these, the complex classical Lie groups are four infinite families of Lie groups that together with the exceptional groups exhaust the classification of simple Lie groups. The compact classical groups are compact real forms of the complex classical groups. The finite analogues of the classical groups are the classical groups of Lie type. The term "classical group" was coined by Hermann Weyl, it being the title of his 1939 monograph The Classical Groups.

In mathematics, a zonal spherical function or often just spherical function is a function on a locally compact group G with compact subgroup K that arises as the matrix coefficient of a K-invariant vector in an irreducible representation of G. The key examples are the matrix coefficients of the spherical principal series, the irreducible representations appearing in the decomposition of the unitary representation of G on L2(G/K). In this case the commutant of G is generated by the algebra of biinvariant functions on G with respect to K acting by right convolution. It is commutative if in addition G/K is a symmetric space, for example when G is a connected semisimple Lie group with finite centre and K is a maximal compact subgroup. The matrix coefficients of the spherical principal series describe precisely the spectrum of the corresponding C* algebra generated by the biinvariant functions of compact support, often called a Hecke algebra. The spectrum of the commutative Banach *-algebra of biinvariant L1 functions is larger; when G is a semisimple Lie group with maximal compact subgroup K, additional characters come from matrix coefficients of the complementary series, obtained by analytic continuation of the spherical principal series.

In mathematics, the Plancherel theorem for spherical functions is an important result in the representation theory of semisimple Lie groups, due in its final form to Harish-Chandra. It is a natural generalisation in non-commutative harmonic analysis of the Plancherel formula and Fourier inversion formula in the representation theory of the group of real numbers in classical harmonic analysis and has a similarly close interconnection with the theory of differential equations. It is the special case for zonal spherical functions of the general Plancherel theorem for semisimple Lie groups, also proved by Harish-Chandra. The Plancherel theorem gives the eigenfunction expansion of radial functions for the Laplacian operator on the associated symmetric space X; it also gives the direct integral decomposition into irreducible representations of the regular representation on L2(X). In the case of hyperbolic space, these expansions were known from prior results of Mehler, Weyl and Fock.

<span class="mw-page-title-main">Borel–de Siebenthal theory</span>

In mathematics, Borel–de Siebenthal theory describes the closed connected subgroups of a compact Lie group that have maximal rank, i.e. contain a maximal torus. It is named after the Swiss mathematicians Armand Borel and Jean de Siebenthal who developed the theory in 1949. Each such subgroup is the identity component of the centralizer of its center. They can be described recursively in terms of the associated root system of the group. The subgroups for which the corresponding homogeneous space has an invariant complex structure correspond to parabolic subgroups in the complexification of the compact Lie group, a reductive algebraic group.

In mathematics, symmetric cones, sometimes called domains of positivity, are open convex self-dual cones in Euclidean space which have a transitive group of symmetries, i.e. invertible operators that take the cone onto itself. By the Koecher–Vinberg theorem these correspond to the cone of squares in finite-dimensional real Euclidean Jordan algebras, originally studied and classified by Jordan, von Neumann & Wigner (1934). The tube domain associated with a symmetric cone is a noncompact Hermitian symmetric space of tube type. All the algebraic and geometric structures associated with the symmetric space can be expressed naturally in terms of the Jordan algebra. The other irreducible Hermitian symmetric spaces of noncompact type correspond to Siegel domains of the second kind. These can be described in terms of more complicated structures called Jordan triple systems, which generalize Jordan algebras without identity.

In mathematics, Lie group–Lie algebra correspondence allows one to correspond a Lie group to a Lie algebra or vice versa, and study the conditions for such a relationship. Lie groups that are isomorphic to each other have Lie algebras that are isomorphic to each other, but the converse is not necessarily true. One obvious counterexample is and which are non-isomorphic to each other as Lie groups but their Lie algebras are isomorphic to each other. However, by restricting our attention to the simply connected Lie groups, the Lie group-Lie algebra correspondence will be one-to-one.

<span class="mw-page-title-main">Exponential map (Lie theory)</span>

In the theory of Lie groups, the exponential map is a map from the Lie algebra of a Lie group to the group, which allows one to recapture the local group structure from the Lie algebra. The existence of the exponential map is one of the primary reasons that Lie algebras are a useful tool for studying Lie groups.

<span class="mw-page-title-main">Lie algebra extension</span> Creating a "larger" Lie algebra from a smaller one, in one of several ways

In the theory of Lie groups, Lie algebras and their representation theory, a Lie algebra extensione is an enlargement of a given Lie algebra g by another Lie algebra h. Extensions arise in several ways. There is the trivial extension obtained by taking a direct sum of two Lie algebras. Other types are the split extension and the central extension. Extensions may arise naturally, for instance, when forming a Lie algebra from projective group representations. Such a Lie algebra will contain central charges.

References