Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Lie Algebras
Lie Algebras
Lie Algebras
Ebook516 pages6 hours

Lie Algebras

Rating: 2.5 out of 5 stars

2.5/5

()

Read preview

About this ebook

Lie group theory, developed by M. Sophus Lie in the nineteenth century, ranks among the more important developments in modern mathematics. Lie algebras comprise a significant part of Lie group theory and are being actively studied today. This book, by Professor Nathan Jacobson of Yale, is the definitive treatment of the subject and can be used as a text for graduate courses.
Chapter 1 introduces basic concepts that are necessary for an understanding of structure theory, while the following three chapters present the theory itself: solvable and nilpotent Lie algebras, Cartan’s criterion and its consequences, and split semi-simple Lie algebras. Chapter 5, on universal enveloping algebras, provides the abstract concepts underlying representation theory. The basic results on representation theory are given in three succeeding chapters: the theorem of Ado-Iwasawa, classification of irreducible modules, and characters of the irreducible modules. In Chapter 9 the automorphisms of semi-simple Lie algebras over an algebraically closed field of characteristic zero are determined. These results are applied in Chapter 10 to the problems of sorting out the simple Lie algebras over an arbitrary field. The reader, to fully benefit from this tenth chapter, should have some knowledge about the notions of Galois theory and some of the results of the Wedderburn structure theory of associative algebras.
Nathan Jacobson, presently Henry Ford II Professor of Mathematics at Yale University, is a well-known authority in the field of abstract algebra. His book, Lie Algebras, is a classic handbook both for researchers and students. Though it presupposes knowledge of linear algebra, it is not overly theoretical and can be readily used for self-study.
LanguageEnglish
Release dateSep 16, 2013
ISBN9780486136790
Lie Algebras

Related to Lie Algebras

Related ebooks

Mathematics For You

View More

Related articles

Reviews for Lie Algebras

Rating: 2.5 out of 5 stars
2.5/5

3 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Lie Algebras - Nathan Jacobson

    ALGEBRAS

    CHAPTER I

    Basic Concepts

    The theory of Lie algebras is an outgrowth of the Lie theory of continuous groups. The main result of the latter is the reduction of local problems concerning Lie groups to corresponding problems on Lie algebras, thus to problems in linear algebra. One associates with every Lie group a Lie algebra over the reals or complexes and one establishes a correspondence between the analytic subgroups of the Lie group and the subalgebras of its Lie algebra, in which invariant subgroups correspond to ideals, abelian subgroups to abelian subalgebras, etc. Isomorphism of the Lie algebras is equivalent to local isomorphism of the corresponding Lie groups. We shall not discuss these matters in detail since excellent modern accounts of the Lie theory are available. The reader may consult one of the following books: Chevalley’s Theory of Lie Groups, Cohn’s Lie Groups, Pontrjagin’s Topological Groups.

    More recently, two other types of group theory have been aided by the introduction of appropriate Lie algebras in their study. The first of these is the theory of free groups which can be studied by means of free Lie algebras using a method which was originated by Magnus. Although the connection here is not so close as in the Lie theory, significant results on free groups and other types of discrete groups have been obtained using Lie algebras. Particularly noteworthy are the results on the so-called restricted Burnside problem: Is there a bound for the orders of the finite groups with a fixed number r of generators and satisfying the relation xm = 1, m a fixed positive integer? It is worth mentioning that Lie algebras of prime characteristic play an important role in these applications to discrete group theory. Again we shall not enter into the details but refer the interested reader to two articles which give a good account of this method in group theory. These are: Lazard [2] and Higman [1].

    The type of correspondence between subgroups of a Lie group and subalgebras of its Lie algebra which obtains in the Lie theory has a counterpart in Chevalley’s theory of linear algebraic groups. Roughly speaking, a linear algebraic group is a subgroup of the group of non-singular n × n matrices which is specified by a set of polynomial equations in the entries of the matrices. An example is the orthogonal group which is defined by the set of equations

    on the entries αij of the matrix (αij). With each linear algebraic group Chevalley has defined a corresponding Lie algebra (see Chevalley [2]) which gives useful information on the group and is decisive in the theory of linear algebraic groups of characteristic zero.

    In view of all this group theoretic background it is not surprising that the basic concepts in the theory of Lie algebras have a group-theoretic flavor. This should be kept in mind throughout the study of Lie algebras and particularly in this chapter, which gives the foundations that are adequate for the main structure theory to be developed in Chapters II to IV. Questions on foundations are taken up again in Chapter V. These concern some concepts that are necessary for the representation theory, which will be treated in Chapters VI and VII.

    1. Definition and construction of Lie and associative algebras

    over a field Φ. over Φ in which a bilinear composition is defined. Thus for every pair (x, y), x, y , we can associate a product x y and this satisfies the bilinearity conditions∈

    A similar definition can be given for a non-associative algebra over a commutative ring Φ having an identity element (unit) 1. This is a left Φ-module with a product x y satisfying (1) and (2). We shall be interested mainly in the case of algebras over fields and, in fact, in such algebras which are finite-dimensional as vector spaces. For such an algebra we have a basis (e1, e2, …, ewhere the γ’s are in Φ. The n³ γijk are called the constants of multiplication of the algebra (relative to the chosen basis). They give the values of every product eiej, i, j = 1, 2, …, n. Thus let x and y . Then, by (1) and (2),

    and this is determined by the eiej.

    and a basis (ei. For every pair (i, j) we define in any way we please eiej we define

    One checks immediately that this is bilinear in the sense that (1) and (2) are valid. The choice of eiej is equivalent to the choice of the elements γ ijk in Φ

    The notion of a non-associative algebra is too general to lead to interesting structural results. In order to obtain such results one must impose some further conditions on the multiplication. The most important ones—and the ones which will concern us here, are the associative laws and the Lie conditions.

    is said to be associative if its multiplication satisfies the associative law

    is said to be a Lie algebra if its multiplication satisfies the Lie conditions

    The second of these is called the Jacobi identity.

    is a Lie algebra and x, y then 0 = (x + y)² = x² + xy + yx + y² = xy + yx so that

    holds in any Lie algebra. Conversely, if this condition holds then 2x² = 0, so that, if the characteristic is not two, then x² = 0. Hence for algebras of characteristic ≠ 2 the condition (6) can be used for the first of (5) in the definition of a Lie algebra.

    PROPOSITION 1. A Non-Associative Algebra with basis (e1, e2, …, en) over Φ is associative if and only if (eiej)ek = ei(ejek) for i, j, k = 1, 2, …, n. If these conditions are equivalent to

    The algebra is Lie if and only if

    for i, j, k = 1, 2, …, n. These conditions are equivalent to

    Proof: is associative, then (eiej)ek = ei(ejek). Conversely, assume these conditions hold for the ei. If

    . Hence (xy)z = x(yz. Hence the linear independence of the ei implies that the conditions (eiej)ek = ei(ejek) are equivalent to (7). The proof in the Lie case is similar to the foregoing and will be omitted.

    In actual practice the general procedure we have indicated is not often used in constructing examples of associative and of Lie algebras except for algebras of low dimensionalities. We shall employ this in determining the Lie algebras of one, two, and three dimensions in and eiej = – ejei in an algebra, then the validity of (eiej)ek + (ejek)ei + (ekei)ej = 0 for a particular triple i, j, k implies (ejei)ek + (eiek)ej + (ekej)ei = 0. Since cyclic permutations of i, j, k are clearly allowed it follows that the Jacobi identity for (ei, ej,)ek is valid for i′, j′, k′, a permutation of i, j, k. Next let i = j. Then

    or, what is the same thing, ximplies that the Jacobi identities are satisfied for ei, ei, ej. In particular, the Jacobi identities are consequences of x= 3, then the only identity we have to check is (e1e2)e3 + (e2e3)e1 + (e3e1)e2 = 0.

    2. Algebras of linear transformations. Derivations

    be a vector space over a field Φ into itself. We recall that if A, B and α Φ, then A + B, αA and AB are defined by x(A + B) = xA + xB, x(α A) = α(xA), x(AB) = (xA)B for x is a vector space relative to + and the scalar multiplication and that multiplication is associative and satisfies is m-dimensional, m is m²-dimensional over Φ. If (e1, e2, …, emover Φ, then the linear transformations Eij such that eiEij = ej, erEij = 0 if r i, i, j = 1, …, mover Φ. If A , then we can write ei A = Σjαijej, i = 1, …, m, and (α) = (αij) is the matrix of A relative to the basis (ei). The correspondence A → (αonto the algebra Φm of m × m matrices with entries αij in Φ.

    is called the (associative) algebra of linear transformations over Φwhich is closed under multiplication, is called an algebra of linear transformations.

    is an arbitrary non-associative algebra and a , then the mapping aR which sends any x into xa is a linear transformation. It is well known and easy to check that (a + b)R = aR + bR, (αa)R = αaR is associative, (ab)R = aRbRis an associative algebra, the mapping a aR has an identity (or unit) 1, then a aR + 1 (cf. Jacobson [2], vol. I, is isomorphic to an algebra of linear transformations in a finite-dimensional vector space.

    be an associative algebra. If x, y , then we define the Lie product or (additive) commutator of x and y as

    One checks immediately that

    Moreover,

    Thus the product [xy] satisfies all the conditions on the product in a Lie algebra. The Lie algebra obtained in this way is called the Lie algebra LL L is called a Lie algebra of linear transformationsLassociative. In view of the result just proved on associative algebras this is equivalent to showing that every Lie algebra is isomorphic to a Lie algebra of linear transformations.

    Lover a field Φ.

    Orthogonal Lie algebrabe equipped with a non-degenerate symmetric bilinear form (x,yfinite-dimensional. Then any linear transformation A has an adjoint A* relative to (x,y); that is, A* is linear and satisfies: (x A, y) = (x, y A*). The mapping A A* : (A + B)* = A* + B*, (αA)* = αA*, (AB)* = B* Adenote the set of A which are skew in the sense that A* = − Aand if A* = − A, B* = − B, then [AB]* = (AB − BA)* = B*A* − A*B* − BA AB = [BA] = − [AB]. Hence [ABand is a subalgebra of L.

    If Φ relative to (x,y). This is the group of linear transformations O which are orthogonal in the sense that (xO, yO) = (x, y), x, y the orthogonal Lie algebra relative to (x, y).

    Symplectic Lie algebra. Here we suppose (x, y) is a non-degenerate alternate form: (x, x= 2l is even. Again let A* be the adjoint of A) relative to (x,yof skew (A* = – AL. This is related to the symplectic group and so we shall call it the symplectic Lie algebra of the alternate form (x,y).

    Triangular linear transformationsi = i be the set of linear transformations T L L. We can choose a basis (x1, x2, …, xmso that (x1, x2, …, xiiimplies that the matrix of T relative to (x1, x2, …, xm) is of the form

    Such a matrix is called triangular and correspondingly we shall call any T a triangular linear transformation.

    Derivation algebrasbe an arbitrary non-associative algebra. A derivation D satisfying

    , then

    if α Φ. We have

    Interchange of 1, 2 and subtraction gives

    (xy)[D1D2] = (x[D1D2])y + x(y[D1D2]).

    Hence [D1DL. We shall call this the Lie algebra of derivations or derivation algebra of .

    is a finite-dimensional algebra over the field of real numbers. We shall not prove any of our assertions on the relation between Lie groups and Lie algebras but refer the reader to the literature on Lie groups for this. However, in the present instance we shall indicate the link between the group of automorphisms and the Lie algebra of derivations.

    Let D be a derivation. Then induction on n gives the Leibniz rule:

    If the characteristic of Φ is 0 we can divide by n! and obtain

    is finite-dimensional over the field of reals, then it is easy to prove (cf. Jacobson [2], vol. II, p. 197) that the series

    converges for every linear mapping D , and the linear mapping exp D defined by (13) is 1:1. Also it is easy to see, using (12′), that if D is a derivation, then G = exp D satisfies (xy)G = (xG)(yG). Hence G .

    is arbitrary of characteristic 0. Let D be a nilpotent derivation, say, DN = 0. Consider the mapping

    We write this as G = 1 + Z, Z = D + (D²/2!) + … + (DN–1/(N – 1)!) and note that ZN = 0. Hence G = 1 + Z has the inverse 1 – Z + Z² + … ± Zn–1 and so G . We have

    Hence G .

    3. Inner derivations of associative and Lie algebras

    If a , then a determines two mappings aL: x → ax and aR: x xa into itself. These are called the left multiplication and right multiplication determined by a. The defining conditions (1) and (2) for an algebra show that aL and aR are linear mappings and the mappings a aL, a aR be associative and set Da = aR aL. Hence Da is the linear mapping x xa ax. We have

    hence Da . We shall call this the inner derivation determined by a.

    by [xy] and we shall do this from now on. Also, it is usual to denote the right multiplication aR (= − aL since [xa] = – [ax]) by ad a and to call this the adjoint mapping determined by a. We have

    hence ad a: x → [xa] is a derivation. We call this also the inner derivation determined by a .

    is called an ideal if , (2) ab, ba for any a , b . Consider the set of elements of the form Σaibi, ai, bi ² and to call this the derived algebra is an ideal if and only if [ab] (or [bafor every a , b of elements c such that [ac] = 0 for all a is an ideal. This is called the center is called abelian ′ = 0.

    .

    Proof. In any non-associative algebra we have (a + b)L = aL + bL, (αa)L = αaL, (a + b)R = aR + bR, (αa)R = αaR. Hence if Da = aR aL then Da + b = Da + Db, Dαa = αDa . Let D . Then (ax)D = (aD)x + a(xD), or (ax)D a(xD) = (aD)x. In operator form this reads (xaL) – (xD)aL = x(aD)L, or [aLD] = aLD DaL = (aD)L. Similarly, [aRD] = (aD)R and consequently also [DaD] = DaDis associative or Lie and I is an inner derivation and D any derivation, then [ID] .

    Examplebe the algebra with basis (e, f) such that [ef] = e = – [fe] and all other products of base elements are 0. Then [aa′ = Φe. If D Hence if D then eD = δe. Also ad(δf) has the property e(ad δf) = [e, δf] = δe. Hence if E = D – ad δf, then E is a derivation and eE = 0. Then e = [ef] gives 0 = [e, fE]. It follows that fE = γe. Now ad(– γe) satisfies e ad(– γe) = 0, f ad(– γe) = [f, – γe] = γ[ef] = γe. Hence E = ad(– γe) is inner and D = E + ad δf = Φe + Φf is inner.

    In group theory one defines a group to be complete if all of its automorphisms are inner and its center is the identity. If H is complete and invariant in G then H is a direct factor of G. By analogy we shall call a Lie algebra complete if its derivations are all inner and its center is 0.

    PROPOSITION 3. If is complete and an ideal in , then = where is an ideal.

    Proof : , that is, the set of elements b such that [kb] = 0 for all k is evidently a subspace and if b and a , then [k[ba]] = − [a[kb]] – [b[ak]] = 0 − [b, k′], k′ = [ak; hence [k[ba]] = 0 for all k and [babe complete. If c , then c and so c = 0. Next let a into itself and hence it induces a derivation D . This is inner and so we have a k such that xD = [xa] = [xk] for every x . Then b = a – k and a = b + k, b , k as required.

    Example. The algebra Φe + Φf of the last example is complete.

    4. Determination of the Lie algebras of low dimensionalities

    3. If (e1, e2, …, en, then [eiej] = 0 and [eiej] = – [eiej]. Hence in giving the multiplication table for the basis, it suffices to give the products [eiej] for i < j. We shall use these abbreviated multiplication tables in our discussion.

    = Φe [ee] = 0.

    = 2.

    ′ = 0, 2 is abelian.

    = Φe + Φf′ = Φ[ef] is one-dimensional. We may choose e ′ = Φe. Then [ef] = αe ≠ 0 and replacement of f by α–1f permits us to take [ef] = eis the algebra of the example of § 3. This can now be characterized as the non-abelian two-dimensional Lie algebra.

    = 3.

    abelian.

    ′ = Φe= Φe + Φf + Φg′ = Φ[fg]. Hence we may suppose [fg] = ehas basis (e, f, g), with multiplication table

    We have only one Lie algebra satisfying our conditions. (If we have (16), then the Jacobi condition is satisfied.)

    ′ = Φe, then there is an f such that [ef] ≠ 0. Then [ef] = βe ≠ 0 and we may suppose [ef] = e. Hence Φe + Φf = Φghas basis (e, f, g) with multiplication table

    ′ = Φe + Φf = Φe + Φf + Φg′ = Φ[eg] + Φ[fg] and so ad g ′. Hence we have basis (e, f, g) with

    with basis (e, f, g) we can define a product [ab] so that [aa] = 0 and (18) holds. Then [[ef]g] + [[fg]e] + [[ge]fis a Lie algebra. What changes can be made in the multiplication table (18)? Our choice of basis amounts to this: We have chosen a basis (e, f′ and supplemented this with a g ′ will change A to a similar matrix M–1AM. The type of change allowable for g is to replace it by ρg + x, ρ ≠ 0 in Φ, x ′. Then [e, ρg + x] = ρ[eg], [f, ρg + x] = ρ[fg] so this changes A to ρA. Hence the different matrices A which can be used in (18) are the non-zero multiples of the matrices similar to A′ = 2 and the conjugacy classes in the two dimensional collineation group.

    If the field is algebraically closed we can choose A in one of the following forms:

    These give the multiplication tables

    Different choices of α give different algebras unless αα′ = 1. Hence we get an infinite number of non-isomorphic algebras.

    ′ = 3. Let (e1, e2, e3) be a basis and set [e2e3] = f1, [e3e1] = f2, [e1e2] = f3. Then (fi, f2, f, A = (αij) non-singular. The only Jacobi condition which has to be imposed is that [f1e1] + [f2e2] + [f3e3] = 0. This gives

    Hence αij = αji and so A non-singular. Set

    . We have for (i, j, k) any cyclic permutation of (1, 2, 3)

    The matrix N = (vij) = adj M′ = (M′)–1 det M′. The matrix relating the f’s to the e’s is A and that relating the e′s is Mis the matrix (ᾱijthen

    Two matrices A, B are called multiplicatively cogredient if B = ρNAN where N is non-singular and p ≠ 0 in Φ. In this case we may write B = ρσ²(σ–1N)′ A(σ–1N), σ = ρ det N and if the matrices are of three rows and columns, then we take M = σN–1 and B = μ(M–1)′ AM–1, μ = ρσ² = det M. Thus we have the relation (19). Thus the conditions on A satisfying our conditions we can associate a unique class of non-singular multiplicatively cogredient symmetric matrices. We have as many algebras as there are classes of such matrices. For the remainder of this section we assume the characteristic is not two. Then each co-gredience class contains a diagonal matrix of the form diag {α, β, 1}, αβ ≠ 0. This implies that the basis can be chosen so that

    If the base field is the field of reals, then we have two different algebras obtained by taking α = β = 1 and α = − 1, β = 1. If the field is algebraically closed we can take α = β = 1.

    contains an element h such that ad h has a characteristic root α ≠ 0 belonging to Φ. Then we have a vector e ≠ 0 such that [eh] = e ad h = αe ≠ 0 and since [hh] = 0, e and h are linearly independent and are part of a basis (e1, e2, e3) = (e, h, f). If (f1, f2, f3) are defined as before, the symmetric matrix (αij) is now

    Then we have [eh] = αe, [hh] = 0, [fh] = –αf α11h α12e, which implies that the characteristic roots of ad h are 0, α and – α. We may replace f by a characteristic vector belonging to the root – α of ad h. This is linearly independent of (e, h) and may be used for f. Hence we may suppose that [eh] = αe, [fh] = – αf. If we replace h by 2α–1h we obtain [eh] – 2e, [fh] = – 2f. The form of (21) now gives [ef] = βh ≠ 0. If we replace f by β–1f, then we obtain the basis (e, f, h) such that

    ′ ≠ 0. The particular algebra we have singled out by the condition that it contains h with ad h having a non-zero characteristic root in Φ is called the split three-dimensional simple Lie algebra. It will play an important role in the sequel.

    5. Representations and modules

    is an associative algebra over a field Φ, then a representation over Φ. If a A, b B in the representation, then, by definition, a + b A + B, αa αA, α ∈ Φ, and ab AB. A right -module over Φ mapping (x, a), x , a , into an element xa such that

    If a A -module by defining xa = xA. Thus we will have

    -module, then for any a we let A denote the mapping x xa. Then the first part of 1 and the first part of 2 show that A over Φ. The rest of the conditions in 1, 2, and 3 imply that a A is a representation.

    In the theory of representations and in other parts of the theory of associative algebras, algebras with an identity play a preponderant role. In fact, for most considerations it is convenient to confine attention to these algebras and to consider only those homomorphisms which map the identity into the identity of the image algebra. In the sequel we shall find it useful at times to deal with associative algebras which need not have identities. We shall therefore adopt the following conventions on terminology: Algebra without any modifier will mean associative algebra with an identity For these subalgebra will mean subalgebra in the usual sense containing the identity, and homomorphism will mean homomorphism in the usual sense mapping 1 into 1. In particular, this will be understood for representations. The corresponding notion of a module is defined by 1 through 3 above, together with the condition

    4. x1 = x, x .

    ."

    -module by taking xa . Then 1, 2, and 3 hold as a consequence of the axioms for an algebra and 4. holds since 1 is the identity. The representation a A where A is the linear transformation x xa is called the regular representation. We have seen (§ 2) that the regular representation is faithful, that is, an isomorphism.

    be a Lie algebra. Then we define a representation to be a homomorphism l L Lover Φ. The conditions here are that if l1 → L1, l2 → L2, then

    We now define xl for x , l by xl = xL. Then (23) and the linearity of L gives the following conditions:

    over Φ such that the result xl satisfies 1, 2, and 3 above.

    As in the associative case, the concepts of module and representation are equivalent. Thus we have indicated that if l L is any module, then for any l we let L denote the mapping x xl. Then L over Φ and l L over Φ.

    -module by taking xl to be the product [xl. Then 1 and 2 are consequences of the axioms for an algebra and 3. follows from the Jacobi identity and skew symmetry. We have denoted the representing transformation of l: x → [xl] by ad l. The representation l → ad l determined by this module is called the adjoint representation . We recall that the mappings ad l .

    as an abelian Lie algebra with product [xy] = 0. Then the mappings x → xl -module which is at the same time a Lie algebra, and we assume that the module mappings x xl and 1, 2, and 3 above, we have also

    4. [x1x2]l = [x1l, x2] + [x1l, x2l].

    a multiplication

    Enjoying the preview?
    Page 1 of 1