akg
akg x akg → akg
x
y
[x,y]
[x,y]=xy-yx
Lie algebras are closely related to Lie groups, which are groups that are also smooth manifolds: every Lie group gives rise to a Lie algebra, which is the tangent space at the identity. (In this case, the Lie bracket measures the failure of commutativity for the Lie group.) Conversely, to any finite-dimensional Lie algebra over the real or complex numbers, there is a corresponding connected Lie group, unique up to covering spaces (Lie's third theorem). This correspondence allows one to study the structure and classification of Lie groups in terms of Lie algebras, which are simpler objects of linear algebra.
In more detail: for any Lie group, the multiplication operation near the identity element 1 is commutative to first order. In other words, every Lie group G is (to first order) approximately a real vector space, namely the tangent space
ak{g}
ak{g}
In physics, Lie groups appear as symmetry groups of physical systems, and their Lie algebras (tangent vectors near the identity) may be thought of as infinitesimal symmetry motions. Thus Lie algebras and their representations are used extensively in physics, notably in quantum mechanics and particle physics.
An elementary example (not directly coming from an associative algebra) is the 3-dimensional space
ak{g}=R3
[x,y]=x x y.
x x y=-y x x
x x (y x z)+ y x (z x x)+ z x (x x y) = 0.
v\in\R3
v
v
[x,x]=x x x=0
Lie algebras were introduced to study the concept of infinitesimal transformations by Sophus Lie in the 1870s,[1] and independently discovered by Wilhelm Killing[2] in the 1880s. The name Lie algebra was given by Hermann Weyl in the 1930s; in older texts, the term infinitesimal group was used.
A Lie algebra is a vector space
ak{g}
F
[ ⋅ , ⋅ ]:ak{g} x ak{g}\toak{g}
[ax+by,z]=a[x,z]+b[y,z],
[z,ax+by]=a[z,x]+b[z,y]
for all scalars
a,b
F
x,y,z
ak{g}
[x,x]=0
for all
x
ak{g}
[x,[y,z]]+[y,[z,x]]+[z,[x,y]]=0
for all
x,y,z
ak{g}
Given a Lie group, the Jacobi identity for its Lie algebra follows from the associativity of the group operation.
Using bilinearity to expand the Lie bracket
[x+y,x+y]
[x,y]+[y,x]=0
x,y
ak{g}
[x,y]=-[y,x],
for all
x,y
ak{g}
[x,x]=-[x,x].
It is customary to denote a Lie algebra by a lower-case fraktur letter such as
ak{g,h,b,n}
ak{su}(n)
The dimension of a Lie algebra over a field means its dimension as a vector space. In physics, a vector space basis of the Lie algebra of a Lie group G may be called a set of generators for G. (They are "infinitesimal generators" for G, so to speak.) In mathematics, a set S of generators for a Lie algebra
ak{g}
ak{g}
ak{g}
ak{g}
Any vector space
V
A
F
xy
[x,y]=xy-yx
A
A
F
V
ak{gl}(V)
ak{gl}(n,F)
ak{gl}n(F)
[X,Y]=XY-YX
When F is the real numbers,
ak{gl}(n,R)
GL(n,R)
ak{gl}(n,C)
GL(n,C)
ak{gl}(n,\R)
ak{gl}(n,F)
GL(n)
The Lie bracket is not required to be associative, meaning that
[[x,y],z]
[x,[y,z]]
ak{h}\subseteqak{g}
aki\subseteqak{g}
[ak{g},aki]\subseteqaki.
In the correspondence between Lie groups and Lie algebras, subgroups correspond to Lie subalgebras, and normal subgroups correspond to ideals.
A Lie algebra homomorphism is a linear map compatible with the respective Lie brackets:
\phi\colonak{g}\toak{h}, \phi([x,y])=[\phi(x),\phi(y)] forall x,y\inakg.
As with normal subgroups in groups, ideals in Lie algebras are precisely the kernels of homomorphisms. Given a Lie algebra
ak{g}
aki
ak{g}/aki
ak{g}\toak{g}/ak{i}
\phi\colonak{g}\toak{h}
\phi
ak{h}
ak{g}/ker(\phi)
For the Lie algebra of a Lie group, the Lie bracket is a kind of infinitesimal commutator. As a result, for any Lie algebra, two elements
x,y\inakg
[x,y]=0
The centralizer subalgebra of a subset
S\subsetak{g}
S
ak{z}akg(S)=\{x\inakg:[x,s]=0 foralls\inS\}
ak{g}
ak{z}(ak{g})
S
ak{n}akg(S)=\{x\inakg:[x,s]\inS forall s\inS\}
S
ak{n}akg(S)
S
ak{n}akg(S)
The subspace
ak{t}n
ak{gl}(n,F)
ak{gl}(n)
ak{t}n
ak{gl}(n)
n\geq2
n=2
(which is not always in\begin{align} \left[ \begin{bmatrix} a&b\\ c&d \end{bmatrix}, \begin{bmatrix} x&0\\ 0&y \end{bmatrix} \right]&=\begin{bmatrix} ax&by\\ cx&dy\\ \end{bmatrix}-\begin{bmatrix} ax&bx\\ cy&dy\\ \end{bmatrix}\\ &=\begin{bmatrix} 0&b(y-x)\\ c(x-y)&0 \end{bmatrix} \end{align}
ak{t}2
Every one-dimensional linear subspace of a Lie algebra
ak{g}
For two Lie algebras
ak{g}
ak{g'}
ak{g} x ak{g'}
(x,x'),x\inak{g}, x'\inak{g'}
[(x,x'),(y,y')]=([x,y],[x',y']).
akg
akg'
ak{g} x ak{g'}
[(x,0),(0,x')]=0.
Let
ak{g}
ak{i}
ak{g}
ak{g}\toak{g}/ak{i}
ak{g}/ak{i}\toak{g}
ak{g}
ak{i}
ak{g}/ak{i}
ak{g}=ak{g}/ak{i}\ltimesak{i}
For an algebra A over a field F, a derivation of A over F is a linear map
D\colonA\toA
D(xy)=D(x)y+xD(y)
x,y\inA
D1
D2
[D1,D2]:=D1D2-D2D1
Derk(A)
Informally speaking, the space of derivations of A is the Lie algebra of the automorphism group of A. (This is literally true when the automorphism group is a Lie group, for example when F is the real numbers and A has finite dimension as a vector space.) For this reason, spaces of derivations are a natural way to construct Lie algebras: they are the "infinitesimal automorphisms" of A. Indeed, writing out the condition that
(1+\epsilonD)(xy)\equiv(1+\epsilonD)(x) ⋅ (1+\epsilonD)(y)\pmod{\epsilon2}
Example: the Lie algebra of vector fields. Let A be the ring
Cinfty(X)
R
Vect(X)
Vect(X)
ak{g}\toVect(X)
A Lie algebra can be viewed as a non-associative algebra, and so each Lie algebra
ak{g}
DerF(ak{g})
ak{g}
D\colonak{g}\toak{g}
D([x,y])=[D(x),y]+[x,D(y)]
x\inakg
adx
adx(y):=[x,y]
\operatorname{ad}\colonak{g}\toDerF(ak{g})
InnF(ak{g})
DerF(ak{g})
OutF(ak{g})=DerF(ak{g})/InnF(ak{g})
In contrast, an abelian Lie algebra has many outer derivations. Namely, for a vector space
V
OutF(V)
ak{gl}(V)
A matrix group is a Lie group consisting of invertible matrices,
G\subsetGL(n,R)
akg
Mn(R)
I
ak{g}=\{X=c'(0)\inMn(R):smoothc:R\toG, c(0)=I\}.
ak{g}
[X,Y]=XY-YX
ak{g}\subsetak{gl}(n,R)
ak{g}
\exp:Mn(R)\toMn(R)
\exp(X)=I+X+\tfrac{1}{2!}X2+\tfrac{1}{3!}X3+ …
X
The same comments apply to complex Lie subgroups of
GL(n,C)
\exp:Mn(C)\toMn(C)
Here are some matrix Lie groups and their Lie algebras.
SL(n,R)
Rn
SL(n,R)
GL(n,\R)
ak{sl}(n,R)
{\rmSL}(n,C)
ak{sl}(n,C)
O(n)
Rn
O(n)
AT=A-1
AT
SO(n)
ak{so}(n)
ak{gl}(n,R)
X\rm=-X
The complex orthogonal group
O(n,C)
SO(n,C)
ak{so}(n,C)
O(n,C)
GL(n,C)
Cn
U(n)
GL(n,C)
Cn
A*=A-1
A*
ak{u}(n)
ak{gl}(n,C)
X*=-X
R
C
U(n)
GL(n,C)
U(1)
iR\subsetC=ak{gl}(1,C)
SU(n)
U(n)
su(n)
Sp(2n,\R)
GL(2n,R)
R2n
ak{sp}(2n,R)
Some Lie algebras of low dimension are described here. See the classification of low-dimensional real Lie algebras for further examples.
ak{g}
ak{g}
X,Y
\left[X,Y\right]=Y
[X,X]=0
[Y,Y]=0
ak{g}
G=Aff(1,R)
x\mapstoax+b
The affine group G can be identified with the group of matrices
\left(\begin{array}{cc}a&b\ 0&1\end{array}\right)
under matrix multiplication, with
a,b\inR
a ≠ 0
ak{g}
ak{gl}(2,R)
\left(\begin{array}{cc}c&d\ 0&0\end{array}\right).
In these terms, the basis above for
ak{g}
X=\left(\begin{array}{cc}1&0\ 0&0\end{array}\right), Y=\left(\begin{array}{cc}0&1\ 0&0\end{array}\right).
For any field
F
F ⋅ Y
ak{g}
[X,Y]=Y\inF ⋅ Y
F ⋅ Y
ak{g}/(F ⋅ Y)
ak{g}
ak{h}3(F)
X,Y,Z
[X,Y]=Z, [X,Z]=0, [Y,Z]=0
It can be viewed as the Lie algebra of 3×3 strictly upper-triangular matrices, with the commutator Lie bracket and the basis
X=\left(\begin{array}{ccc} 0&1&0\\ 0&0&0\\ 0&0&0 \end{array}\right), Y=\left(\begin{array}{ccc} 0&0&0\\ 0&0&1\\ 0&0&0 \end{array}\right), Z=\left(\begin{array}{ccc} 0&0&1\\ 0&0&0\\ 0&0&0 \end{array}\right)~.
Over the real numbers,
ak{h}3(R)
H3(R)
\left(\begin{array}{ccc} 1&a&c\\ 0&1&b\\ 0&0&1 \end{array}\right)
under matrix multiplication.
For any field F, the center of
ak{h}3(F)
F ⋅ Z
ak{h}3(F)/(F ⋅ Z)
F2
ak{h}3(F)
ak{so}(3)
R
F1=\left(\begin{array}{ccc} 0&0&0\\ 0&0&-1\\ 0&1&0 \end{array}\right), F2=\left(\begin{array}{ccc} 0&0&1\\ 0&0&0\\ -1&0&0 \end{array}\right), F3=\left(\begin{array}{ccc} 0&-1&0\\ 1&0&0\\ 0&0&0 \end{array}\right)~.
The commutation relations among these generators are
[F1,F2]=F3,
[F2,F3]=F1,
[F3,F1]=F2.
The cross product of vectors in
R3
ak{so}(3)
ak{so}(3)
The Lie algebra
ak{so}(3)
ak{so}(3)
C
ak{sl}(2,C)
H=\left(\begin{array}{cc}1&0\ 0&-1\end{array}\right), E=\left (\begin{array}{cc}0&1\ 0&0\end{array}\right), F=\left(\begin{array}{cc}0&0\ 1&0\end{array}\right).
The Lie bracket is given by:
[H,E]=2E,
[H,F]=-2F,
[E,F]=H.
Using these formulas, one can show that the Lie algebra
ak{sl}(2,C)
ak{sl}(2,C)
(c+2)
(c-2)
The Lie algebra
ak{sl}(2,C)
ak{so}(3)
ak{so}(3) ⊗ RC
ak{sl}(2,C)
SO(3)
ak{sl}(2,C)
R
C
ak{sl}(n,\C)
V\mapstoL(V)
L(V)
See main article: Lie algebra representation.
Given a vector space V, let
ak{gl}(V)
[X,Y]=XY-YX
ak{g}
\pi\colonakg\toak{gl}(V).
\pi
ak{g}
ak{g}
A representation is said to be faithful if its kernel is zero. Ado's theorem states that every finite-dimensional Lie algebra over a field of characteristic zero has a faithful representation on a finite-dimensional vector space. Kenkichi Iwasawa extended this result to finite-dimensional Lie algebras over a field of any characteristic. Equivalently, every finite-dimensional Lie algebra over a field F is isomorphic to a Lie subalgebra of
ak{gl}(n,F)
For any Lie algebra
ak{g}
\operatorname{ad}\colonak{g}\toak{gl}(ak{g})
\operatorname{ad}(x)(y)=[x,y]
ak{g}
One important aspect of the study of Lie algebras (especially semisimple Lie algebras, as defined below) is the study of their representations. Although Ado's theorem is an important result, the primary goal of representation theory is not to find a faithful representation of a given Lie algebra
ak{g}
ak{g}
See main article: Universal enveloping algebra. The functor that takes an associative algebra A over a field F to A as a Lie algebra (by
[X,Y]:=XY-YX
ak{g}\mapstoU(ak{g})
ak{g}
T(ak{g})=F ⊕ ak{g} ⊕ (ak{g} ⊗ ak{g}) ⊕ (ak{g} ⊗ ak{g} ⊗ ak{g}) ⊕ …
ak{g}
ak{g}
⊗
T(ak{g})
XY-YX-[X,Y]
X,Y\inak{g}
U(ak{g})=T(ak{g})/I
e1,\ldots,en
ak{g}
U(ak{g})
i1 | |
e | |
1 |
…
in | |
e | |
n |
i1,\ldots,in
ak{g}\toU(ak{g})
Representations of
ak{g}
ak{g}\toU(ak{g})
U(ak{g})
The representation theory of Lie algebras plays an important role in various parts of theoretical physics. There, one considers operators on the space of states that satisfy certain natural commutation relations. These commutation relations typically come from a symmetry of the problem—specifically, they are the relations of the Lie algebra of the relevant symmetry group. An example is the angular momentum operators, whose commutation relations are those of the Lie algebra
ak{so}(3)
SO(3)
ak{so}(3)
Lie algebras can be classified to some extent. This is a powerful approach to the classification of Lie groups.
Analogously to abelian, nilpotent, and solvable groups, one can define abelian, nilpotent, and solvable Lie algebras.
A Lie algebra
ak{g}
ak{g}
Rn
Tn
F
Fn
n\geq0
A more general class of Lie algebras is defined by the vanishing of all commutators of given length. First, the commutator subalgebra (or derived subalgebra) of a Lie algebra
ak{g}
[ak{g},ak{g}]
[x,y]
x,y\inak{g}
ak{g}
A Lie algebra
ak{g}
ak{g}\supseteq[ak{g},ak{g}]\supseteq[[ak{g},ak{g}],ak{g}]\supseteq[[[ak{g},ak{g}],ak{g}],ak{g}]\supseteq …
ak{g}
ak{g}
0=ak{a}0\subseteqak{a}1\subseteq … \subseteqak{a}r=ak{g},
ak{a}j/ak{a}j-1
ak{g}/ak{a}j-1
ak{g}
\operatorname{ad}(u):ak{g}\toak{g}, \operatorname{ad}(u)v=[u,v]
More generally, a Lie algebra
ak{g}
ak{g}\supseteq[ak{g},ak{g}]\supseteq[[ak{g},ak{g}],[ak{g},ak{g}]]\supseteq[[[ak{g},ak{g}],[ak{g},ak{g}]],[[ak{g},ak{g}],[ak{g},ak{g}]]]\supseteq …
ak{g}
0=ak{m}0\subseteqak{m}1\subseteq … \subseteqak{m}r=ak{g},
ak{m}j-1
ak{m}j
ak{m}j/ak{m}j-1
Every finite-dimensional Lie algebra over a field has a unique maximal solvable ideal, called its radical. Under the Lie correspondence, nilpotent (respectively, solvable) Lie groups correspond to nilpotent (respectively, solvable) Lie algebras over
R
For example, for a positive integer n and a field F of characteristic zero, the radical of
ak{gl}(n,F)
ak{b}n
ak{gl}(n)
n\geq2
ak{u}n
ak{gl}(n)
n\geq3
See main article: Semisimple Lie algebra. A Lie algebra
ak{g}
ak{g}
ak{g}
ak{g}
ak{g}
ak{g}
ak{g}
ak{g}
ak{g}\congak{g}1 x … x ak{g}r
For example, the Lie algebra
ak{sl}(n,F)
n\geq2
ak{su}(n)
R
n\geq2
ak{so}(n)
R
n=3
n\geq5
ak{so}(3)\congak{su}(2)
ak{so}(4)\congak{su}(2) x ak{su}(2)
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, every finite-dimensional representation of a semisimple Lie algebra is semisimple (that is, a direct sum of irreducible representations).
A finite-dimensional Lie algebra over a field of characteristic zero is called reductive if its adjoint representation is semisimple. Every reductive Lie algebra is isomorphic to the product of an abelian Lie algebra and a semisimple Lie algebra.
For example,
ak{gl}(n,F)
n\geq2
ak{gl}(n,F)\congF x ak{sl}(n,F),
ak{gl}(n,F)
ak{sl}(n,F)
ak{gl}(n,F)
ak{sl}(n,F)
ak{gl}(n,F)
Cartan's criterion (by Élie Cartan) gives conditions for a finite-dimensional Lie algebra of characteristic zero to be solvable or semisimple. It is expressed in terms of the Killing form, the symmetric bilinear form on
ak{g}
K(u,v)=\operatorname{tr}(\operatorname{ad}(u)\operatorname{ad}(v)),
ak{g}
ak{g}
K(ak{g},[ak{g},ak{g}])=0.
The Levi decomposition asserts that every finite-dimensional Lie algebra over a field of characteristic zero is a semidirect product of its solvable radical and a semisimple Lie algebra. Moreover, a semisimple Lie algebra in characteristic zero is a product of simple Lie algebras, as mentioned above. This focuses attention on the problem of classifying the simple Lie algebras.
The simple Lie algebras of finite dimension over an algebraically closed field F of characteristic zero were classified by Killing and Cartan in the 1880s and 1890s, using root systems. Namely, every simple Lie algebra is of type An, Bn, Cn, Dn, E6, E7, E8, F4, or G2. Here the simple Lie algebra of type An is
ak{sl}(n+1,F)
ak{so}(2n+1,F)
ak{sp}(2n,F)
ak{so}(2n,F)
The classification of finite-dimensional simple Lie algebras over
R
ak{g}
R
ak{g} ⊗ RC
In the years leading up to 2004, the finite-dimensional simple Lie algebras over an algebraically closed field of characteristic
p>3
See main article: Lie group–Lie algebra correspondence. Although Lie algebras can be studied in their own right, historically they arose as a means to study Lie groups.
The relationship between Lie groups and Lie algebras can be summarized as follows. Each Lie group determines a Lie algebra over
R
akg
G
akg
For simply connected Lie groups, there is a complete correspondence: taking the Lie algebra gives an equivalence of categories from simply connected Lie groups to Lie algebras of finite dimension over
R
The correspondence between Lie algebras and Lie groups is used in several ways, including in the classification of Lie groups and the representation theory of Lie groups. For finite-dimensional representations, there is an equivalence of categories between representations of a real Lie algebra and representations of the corresponding simply connected Lie group. This simplifies the representation theory of Lie groups: it is often easier to classify the representations of a Lie algebra, using linear algebra.
Every connected Lie group is isomorphic to its universal cover modulo a discrete central subgroup. So classifying Lie groups becomes simply a matter of counting the discrete subgroups of the center, once the Lie algebra is known. For example, the real semisimple Lie algebras were classified by Cartan, and so the classification of semisimple Lie groups is well understood.
For infinite-dimensional Lie algebras, Lie theory works less well. The exponential map need not be a local homeomorphism (for example, in the diffeomorphism group of the circle, there are diffeomorphisms arbitrarily close to the identity that are not in the image of the exponential map). Moreover, in terms of the existing notions of infinite-dimensional Lie groups, some infinite-dimensional Lie algebras do not come from any group.
Lie theory also does not work so neatly for infinite-dimensional representations of a finite-dimensional group. Even for the additive group
G=R
G
akg
ak{g}0
akg
ak{g}0 ⊗ RC
ak{g}
ak{sl}(2,C)
ak{sl}(2,R)
ak{su}(2)
Given a semisimple complex Lie algebra
akg
A Lie algebra may be equipped with additional structures that are compatible with the Lie bracket. For example, a graded Lie algebra is a Lie algebra (or more generally a Lie superalgebra) with a compatible grading. A differential graded Lie algebra also comes with a differential, making the underlying vector space a chain complex.
For example, the homotopy groups of a simply connected topological space form a graded Lie algebra, using the Whitehead product. In a related construction, Daniel Quillen used differential graded Lie algebras over the rational numbers
Q
The definition of a Lie algebra over a field extends to define a Lie algebra over any commutative ring R. Namely, a Lie algebra
ak{g}
[ , ]\colonak{g} x ak{g}\toak{g}
Z
Lie rings are used in the study of finite p-groups (for a prime number p) through the Lazard correspondence. The lower central factors of a finite p-group are finite abelian p-groups. 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; see the example below.
p-adic Lie groups are related to Lie algebras over the field
Qp
Zp
x,y
[x,y]=x-1y-1xy
G=G1\supseteqG2\supseteqG3\supseteq … \supseteqGn\supseteq …
G
[Gi,Gj]
Gi+j
i,j
L=oplusi\geqGi/Gi+1
is a Lie ring, with addition given by the group multiplication (which is abelian on each quotient group
Gi/Gi+1
Gi/Gi+1 x Gj/Gj+1\toGi+j/Gi+j+1
[xGi+1,yGj+1]:=[x,y]Gi+j+1.
For example, the Lie ring associated to the lower central series on the dihedral group of order 8 is the Heisenberg Lie algebra of dimension 3 over the field
Z/2Z
The definition of a Lie algebra can be reformulated more abstractly in the language of category theory. Namely, one can define a Lie algebra in terms of linear maps—that is, morphisms in the category of vector spaces—without considering individual elements. (In this section, the field over which the algebra is defined is assumed to be of characteristic different from 2.)
For the category-theoretic definition of Lie algebras, two braiding isomorphisms are needed. If is a vector space, the interchange isomorphism
\tau:A ⊗ A\toA ⊗ A
\tau(x ⊗ y)=y ⊗ x.
\sigma:A ⊗ A ⊗ A\toA ⊗ A ⊗ A
\sigma=(id ⊗ \tau)\circ(\tau ⊗ id),
id
\sigma
\sigma(x ⊗ y ⊗ z)=y ⊗ z ⊗ x.
With this notation, a Lie algebra can be defined as an object
A
[ ⋅ , ⋅ ]\colonA ⊗ A → A
[ ⋅ , ⋅ ]\circ(id+\tau)=0,
[ ⋅ , ⋅ ]\circ([ ⋅ , ⋅ ] ⊗ id)\circ(id+\sigma+\sigma2)=0.