Jump to content

Lie algebra: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
Last sentence of the first paragraph was false as written.
BTotaro (talk | contribs)
Revised and expanded
Tag: harv-error
Line 4: Line 4:
{{Ring theory sidebar}}
{{Ring theory sidebar}}


In [[mathematics]], a '''Lie algebra''' (pronounced {{IPAc-en|l|iː}} {{respell|LEE}}) is a [[vector space]] <math>\mathfrak g</math> together with an [[Binary operation|operation]] called the '''Lie bracket''', an [[Alternating multilinear map|alternating bilinear map]] <math>\mathfrak g \times \mathfrak g \rightarrow \mathfrak g</math>, that satisfies the [[Jacobi identity]]. In other words, a Lie algebra is an [[algebra over a field]] for which the multiplication operation (called the Lie bracket) is alternating and satisfies the Jacobi identity. The Lie bracket of two vectors <math>x</math> and <math>y</math> is denoted <math>[x,y]</math>. A Lie algebra is typically a [[non-associative algebra]]. However, every [[associative algebra]] gives rise to a Lie algebra, with the Lie bracket defined as the [[commutator]] <math>[x,y] = x y - yx </math>.
In [[mathematics]], a '''Lie algebra''' (pronounced {{IPAc-en|l|iː}} {{respell|LEE}}) is a [[vector space]] <math>\mathfrak g</math> together with an operation called the '''Lie bracket''', an [[Alternating multilinear map|alternating bilinear map]] <math>\mathfrak g \times \mathfrak g \rightarrow \mathfrak g</math>, that satisfies the [[Jacobi identity]]. In other words, a Lie algebra is an [[algebra over a field]] for which the multiplication operation (called the Lie bracket) is alternating and satisfies the Jacobi identity. The Lie bracket of two vectors <math>x</math> and <math>y</math> is denoted <math>[x,y]</math>. A Lie algebra is typically a [[non-associative algebra]]. However, every [[associative algebra]] gives rise to a Lie algebra, consisting of the same vector space with the [[commutator]] Lie bracket, <math>[x,y] = xy - yx </math>.


Lie algebras are closely related to [[Lie group]]s, which are [[group (mathematics)|group]]s that are also [[smooth manifolds]]: any Lie group gives rise to a Lie algebra, which is its [[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 space|connected]] Lie group unique up to coverings ([[Lie's third theorem]]). This [[Lie group–Lie algebra correspondence|correspondence]] allows one to study the structure and [[List of simple Lie groups|classification]] of Lie groups in terms of Lie algebras.
Lie algebras are closely related to [[Lie group]]s, which are [[group (mathematics)|group]]s 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 number|real]] or [[complex number]]s, there is a corresponding [[connected space|connected]] Lie group, unique up to [[covering space]]s ([[Lie's third theorem]]). This [[Lie group–Lie algebra correspondence|correspondence]] allows one to study the structure and [[List of simple Lie groups|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 <math>\mathfrak{g}</math> to ''G'' at the identity. To second order, the group operation may be non-commutative, and the second-order terms describing the non-commutativity of ''G'' near the identity give <math>\mathfrak{g}</math> the structure of a Lie algebra. It is a remarkable fact that these second-order terms (the Lie algebra) completely determine the group structure of ''G'' near the identity. They even determine ''G'' globally, up to covering spaces.


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.
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 space of three dimensional vectors <math>\mathfrak{g}=\mathbb{R}^3</math> with the Lie bracket operation defined by the [[cross product]] <math>[x,y]=x\times y.</math> This is skew-symmetric since <math>x\times y = -y\times x</math>, and instead of associativity it satisfies the Jacobi identity:
An elementary example (not directly coming from an associative algebra) is the 3-dimensional space <math>\mathfrak{g}=\mathbb{R}^3</math> with Lie bracket defined by the [[cross product]] <math>[x,y]=x\times y.</math> This is skew-symmetric since <math>x\times y = -y\times x</math>, and instead of associativity it satisfies the Jacobi identity:
:<math> x\times(y\times z) \ =\ (x\times y)\times z \ +\ y\times(x\times z). </math>
:<math> x\times(y\times z)+\ y\times(z\times x)+\ z\times(x\times y)\ =\ 0. </math>
This is the Lie algebra of the Lie group of [[3D rotation group|rotations of space]], and each vector <math>v\in\R^3</math> may be pictured as an infinitesimal rotation around the axis <math>v</math>, with velocity equal to the magnitude of <math>v</math>. The Lie bracket is a measure of the non-commutativity between two rotations: since a rotation commutes with itself, one has the alternating property <math>[x,x]=x\times x = 0</math>.
This is the Lie algebra of the Lie group of [[3D rotation group|rotations of space]], and each vector <math>v\in\R^3</math> may be pictured as an infinitesimal rotation around the axis <math>v</math>, with angular speed equal to the magnitude
of <math>v</math>. The Lie bracket is a measure of the non-commutativity between two rotations. Since a rotation commutes with itself, one has the alternating property <math>[x,x]=x\times x = 0</math>.


== History ==
== History ==
Lie algebras were introduced to study the concept of [[infinitesimal transformation]]s by [[Sophus Lie|Marius Sophus Lie]] in the 1870s,<ref>{{harvnb|O'Connor|Robertson|2000}}</ref> and independently discovered by [[Wilhelm Killing]]<ref>{{harvnb|O'Connor|Robertson|2005}}</ref> in the 1880s. The name ''Lie algebra'' was given by [[Hermann Weyl]] in the 1930s; in older texts, the term ''infinitesimal group'' is used.
Lie algebras were introduced to study the concept of [[infinitesimal transformation]]s by [[Sophus Lie]] in the 1870s,<ref>{{harvnb|O'Connor|Robertson|2000}}.</ref> and independently discovered by [[Wilhelm Killing]]<ref>{{harvnb|O'Connor|Robertson|2005}}.</ref> in the 1880s. The name ''Lie algebra'' was given by [[Hermann Weyl]] in the 1930s; in older texts, the term ''infinitesimal group'' was used.


==Definition of a Lie algebra==
== Definitions ==
A Lie algebra is a vector space <math>\,\mathfrak{g}</math> over a [[field (mathematics)|field]] <math>F</math> together with a [[binary operation]] <math>[\,\cdot\,,\cdot\,]: \mathfrak{g}\times\mathfrak{g}\to\mathfrak{g}</math> called the Lie bracket, satisfying the following axioms:{{efn|More generally, one has the notion of a Lie algebra over any [[commutative ring]] ''R'': an ''R''-module with an alternating ''R''-bilinear map that satisfies the Jacobi identity ({{harvtxt|Bourbaki|1989|loc=Section 2}}).}}
===Definition of a Lie algebra===
A Lie algebra is a [[vector space]] <math>\,\mathfrak{g}</math> over some [[field (mathematics)|field]] <math>F</math> together with a [[binary operation]] <math>[\,\cdot\,,\cdot\,]: \mathfrak{g}\times\mathfrak{g}\to\mathfrak{g}</math> called the Lie bracket satisfying the following axioms:{{efn|{{harvtxt|Bourbaki|1989|loc=Section 2.}} allows more generally for a [[Module (mathematics)|module]] over a [[commutative ring]]; in this article, this is called a [[#Lie ring|Lie ring]].}}


* [[Bilinear operator|Bilinearity]],
* ''Bilinearity'',
::<math> [a x + b y, z] = a [x, z] + b [y, z], </math>
::<math> [a x + b y, z] = a [x, z] + b [y, z], </math>
::<math> [z, a x + b y] = a[z, x] + b [z, y] </math>
::<math> [z, a x + b y] = a[z, x] + b [z, y] </math>
:for all scalars <math>a</math>, <math>b</math> in <math>F</math> and all elements <math>x</math>, ''<math>y</math>'', ''<math>z</math>'' in <math>\mathfrak{g}</math>.
:for all scalars <math>a,b</math> in <math>F</math> and all elements <math>x,y,z</math> in <math>\mathfrak{g}</math>.


* The ''Alternating'' property,
* [[Alternatization|Alternativity]],
::<math> [x,x]=0\ </math>
::<math> [x,x]=0\ </math>
:for all <math>x</math> in <math>\mathfrak{g}</math>.
:for all <math>x</math> in <math>\mathfrak{g}</math>.


* The [[Jacobi identity]],
* The ''Jacobi identity'',
:: <math> [x,[y,z]] + [y,[z,x]] + [z,[x,y]] = 0 \ </math>
:: <math> [x,[y,z]] + [y,[z,x]] + [z,[x,y]] = 0 \ </math>
:for all <math>x</math>, ''<math>y</math>'', ''<math>z</math>'' in <math>\mathfrak{g}</math>.
:for all <math>x,y,z</math> in <math>\mathfrak{g}</math>.


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 <math> [x+y,x+y] </math> and using alternativity shows that <math> [x,y] + [y,x]=0\ </math> for all elements <math>x</math>, ''<math>y</math>'' in <math>\mathfrak{g}</math>, showing that bilinearity and alternativity together imply

Using bilinearity to expand the Lie bracket <math> [x+y,x+y] </math> and using the alternating property shows that <math> [x,y] + [y,x]=0 </math> for all <math>x,y</math> in <math>\mathfrak{g}</math>. Thus bilinearity and the alternating property together imply
* [[Anticommutativity]],
* [[Anticommutativity]],
:: <math> [x,y] = -[y,x],\ </math>
:: <math> [x,y] = -[y,x],\ </math>
:for all elements <math>x</math>, ''<math>y</math>'' in <math>\mathfrak{g}</math>. If the field's [[Characteristic (algebra)|characteristic]] is not 2 then anticommutativity implies alternativity, since it implies <math>[x,x]=-[x,x].</math><ref>{{harvnb|Humphreys|1978|p=1}}</ref>
:for all <math>x,y</math> in <math>\mathfrak{g}</math>. If the field does not have [[Characteristic (algebra)|characteristic]] 2, then anticommutativity implies the alternating property, since it implies <math>[x,x]=-[x,x].</math><ref>{{harvnb|Humphreys|1978|p=1.}}</ref>


It is customary to denote a Lie algebra by a lower-case [[fraktur]] letter such as <math>\mathfrak{g, h, b, n}</math>. If a Lie algebra is associated with a [[Lie group]], then the algebra is denoted by the fraktur version of the group: for example the Lie algebra of [[special unitary group|SU(''n'')]] is <math>\mathfrak{su}(n)</math>.
It is customary to denote a Lie algebra by a lower-case [[fraktur]] letter such as <math>\mathfrak{g, h, b, n}</math>. If a Lie algebra is associated with a Lie group, then the algebra is denoted by the fraktur version of the group's name: for example, the Lie algebra of [[special unitary group|SU(''n'')]] is <math>\mathfrak{su}(n)</math>.


===Generators and dimension===
===Generators and dimension===
The ''dimension'' of a Lie algebra over a field means its [[dimension (vector space)|dimension as a vector space]]. In physics, a vector space [[basis (linear algebra)|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 <math>\mathfrak{g}</math> means a subset of <math>\mathfrak{g}</math> such that any Lie subalgebra (as defined below) that contains ''S'' must be all of <math>\mathfrak{g}</math>. Equivalently, <math>\mathfrak{g}</math> is spanned (as a vector space) by all iterated brackets of elements of ''S''.
Elements of a Lie algebra <math>\mathfrak{g}</math> are said to [[Generator (mathematics)|generate]] it if the smallest subalgebra containing these elements is <math>\mathfrak{g}</math> itself. The ''dimension'' of a Lie algebra is its [[dimension (vector space)|dimension as a vector space]] over ''<math>F</math>''. The cardinality of a minimal generating set of a Lie algebra is always less than or equal to its dimension.


== Basic examples ==
See the [[classification of low-dimensional real Lie algebras]] for other small examples.
===Abelian Lie algebras===
Any vector space <math>V</math> endowed with the identically zero Lie bracket becomes a Lie algebra. Such a Lie algebra is called '''abelian'''. Every one-dimensional Lie algebra is abelian, by the alternating property of the Lie bracket.


=== Subalgebras, ideals and homomorphisms ===
=== The Lie algebra of matrices ===
* On an associative algebra <math>A</math> over a field <math>F</math> with multiplication written as <math>xy</math>, a Lie bracket may be defined by the commutator <math>[x,y] = xy - yx</math>. With this bracket, <math>A</math> is a Lie algebra. (The Jacobi identity follows from the associativity of the multiplication on <math>A</math>.) <ref>{{harvnb|Bourbaki|1989|loc=§1.2. Example 1.}}</ref>
The Lie bracket is not required to be [[associative]], meaning that <math>[[x,y],z]</math> need not equal <math>[x,[y,z]]</math>. Nonetheless, much of the terminology of associative [[ring (mathematics)|rings]] and [[associative algebra|algebra]]s has analogs for Lie algebras. A ''Lie subalgebra'' is a linear subspace <math>\mathfrak{h} \subseteq \mathfrak{g}</math> which is closed under the Lie bracket. An ''ideal'' <math>\mathfrak i\subseteq\mathfrak{g}</math> is a subalgebra satisfying the stronger condition:<ref>Due to the anticommutativity of the commutator, the notions of a left and right ideal in a Lie algebra coincide.</ref>
* The [[endomorphism ring]] of an <math>F</math>-vector space <math>V</math> with the above Lie bracket is denoted <math>\mathfrak{gl}(V)</math>.
*For a field ''F'' and a positive integer ''n'', the space of ''n'' × ''n'' [[matrix (mathematics)|matrices]] over ''F'', denoted <math>\mathfrak{gl}(n, F)</math> or <math>\mathfrak{gl}_n(F)</math>, is a Lie algebra with bracket given by the commutator of matrices: <math>[X,Y]=XY-YX</math>.<ref>{{harvnb|Bourbaki|1989|loc=§1.2. Example 2.}}</ref> This is a special case of the previous example; it is probably the most important example of a Lie algebra. It is called the '''general linear''' Lie algebra.


:When ''F'' is the real numbers, <math>\mathfrak{gl}(n,\mathbb{R})</math> is the Lie algebra of the [[general linear group]] <math>\mathrm{GL}(n,\mathbb{R})</math>, the group of [[invertible matrix|invertible]] ''n'' x ''n'' real matrices (or equivalently, matrices with nonzero [[determinant]]), where the group operation is matrix multiplication. Likewise, <math>\mathfrak{gl}(n,\mathbb{C})</math> is the Lie algebra of the complex Lie group <math>\mathrm{GL}(n,\mathbb{C})</math>. The Lie bracket on <math>\mathfrak{gl}(n,\R)</math> describes the failure of commutativity for matrix multiplication, or equivalently for the composition of [[linear map]]s. For any field ''F'', <math>\mathfrak{gl}(n,F)</math> can be viewed as the Lie algebra of the [[algebraic group]] <math>\mathrm{GL}(n)_F</math>.

==Definitions==
=== Subalgebras, ideals and homomorphisms ===
The Lie bracket is not required to be [[associative]], meaning that <math>[[x,y],z]</math> need not be equal to <math>[x,[y,z]]</math>. Nonetheless, much of the terminology for associative [[ring (mathematics)|rings]] and algebras (and also for groups) has analogs for Lie algebras. A '''Lie subalgebra''' is a linear subspace <math>\mathfrak{h} \subseteq \mathfrak{g}</math> which is closed under the Lie bracket. An '''ideal''' <math>\mathfrak i\subseteq\mathfrak{g}</math> is a linear subspace that satisfies the stronger condition:<ref>By the anticommutativity of the commutator, the notions of a left and right ideal in a Lie algebra coincide.</ref>
:<math>[\mathfrak{g},\mathfrak i]\subseteq \mathfrak i.</math>
:<math>[\mathfrak{g},\mathfrak i]\subseteq \mathfrak i.</math>


In the correspondence between Lie groups and Lie algebras, subgroups correspond to Lie subalgebras, and [[normal subgroup]]s correspond to ideals.
A Lie algebra ''homomorphism'' is a linear map compatible with the respective Lie brackets:


A Lie algebra '''homomorphism''' is a linear map compatible with the respective Lie brackets:
:<math> \phi: \mathfrak{g}\to\mathfrak{g'}, \quad \phi([x,y])=[\phi(x),\phi(y)] \ \text{for all}\
x,y \in \mathfrak g. </math>
:<math> \phi\colon \mathfrak{g}\to\mathfrak{g'}, \quad \phi([x,y])=[\phi(x),\phi(y)]\ \text{for all}\ x,y \in \mathfrak g. </math>
An '''isomorphism''' of Lie algebras is a [[bijective]] homomorphism.


As for associative rings, ideals are precisely the [[kernel (algebra)|kernels]] of homomorphisms; given a Lie algebra <math>\mathfrak{g}</math> and an ideal <math>\mathfrak i</math> in it, one constructs the ''factor algebra'' or ''quotient algebra'' <math>\mathfrak{g}/\mathfrak i</math>, and the [[first isomorphism theorem]] holds for Lie algebras.
As with normal subgroups in groups, ideals in Lie algebras are precisely the [[kernel (algebra)|kernels]] of homomorphisms. Given a Lie algebra <math>\mathfrak{g}</math> and an ideal <math>\mathfrak i</math> in it, the ''quotient Lie algebra'' <math>\mathfrak{g}/\mathfrak i</math> is defined, with a surjective homomorphism <math>\mathfrak{g}\to\mathfrak{g}/\mathfrak{i}</math> of Lie algebras. The [[first isomorphism theorem]] holds for Lie algebras.


Since the Lie bracket is a kind of infinitesimal [[commutator]] of the corresponding Lie group, two elements <math>x,y\in\mathfrak g</math> are said to ''commute'' if their bracket vanishes: <math>[x,y]=0</math>.
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 <math>x,y\in\mathfrak g</math> are said to ''commute'' if their bracket vanishes: <math>[x,y]=0</math>.


The [[centralizer]] subalgebra of a subset <math>S\subset \mathfrak{g}</math> is the set of elements commuting with ''<math>S</math>'': that is, <math>\mathfrak{z}_{\mathfrak g}(S) = \{x\in\mathfrak g\ \mid\ [x, s] = 0 \ \text{ for all } s\in S\}</math>. The centralizer of <math>\mathfrak{g}</math> itself is the ''center'' <math>\mathfrak{z}(\mathfrak{g})</math>. Similarly, for a subspace ''S'', the [[normalizer]] subalgebra of ''<math>S</math>'' is <math>\mathfrak{n}_{\mathfrak g}(S) = \{x\in\mathfrak g\ \mid\ [x,s]\in S \ \text{ for all}\ s\in S\}</math>.<ref>{{harvnb|Jacobson|1962|p=28}}</ref> Equivalently, if <math>S</math> is a Lie subalgebra, <math>\mathfrak{n}_{\mathfrak g}(S)</math> is the largest subalgebra such that <math>S</math> is an ideal of <math>\mathfrak{n}_{\mathfrak g}(S)</math>.
The [[centralizer]] subalgebra of a subset <math>S\subset \mathfrak{g}</math> is the set of elements commuting with ''<math>S</math>'': that is, <math>\mathfrak{z}_{\mathfrak g}(S) = \{x\in\mathfrak g\ \mid\ [x, s] = 0 \ \text{ for all } s\in S\}</math>. The centralizer of <math>\mathfrak{g}</math> itself is the ''center'' <math>\mathfrak{z}(\mathfrak{g})</math>. Similarly, for a subspace ''S'', the [[normalizer]] subalgebra of ''<math>S</math>'' is <math>\mathfrak{n}_{\mathfrak g}(S) = \{x\in\mathfrak g\ \mid\ [x,s]\in S \ \text{ for all}\ s\in S\}</math>.<ref>{{harvnb|Jacobson|1979|p=28.}}</ref> If <math>S</math> is a Lie subalgebra, <math>\mathfrak{n}_{\mathfrak g}(S)</math> is the largest subalgebra such that <math>S</math> is an ideal of <math>\mathfrak{n}_{\mathfrak g}(S)</math>.


==== Examples ====
==== Example ====
For <math>\mathfrak{d}(2) \subset \mathfrak{gl}(2)</math>, the commutator of two elements <math>g \in \mathfrak{gl}(2)</math> and <math>d \in \mathfrak{d}(2)</math>:<blockquote><math>\begin{align}
The subspace <math>\mathfrak{t}_n</math> of diagonal matrices in <math>\mathfrak{gl}(n,F)</math> is an abelian Lie subalgebra. (It is a [[Cartan subalgebra]] of <math>\mathfrak{gl}(n)</math>, analogous to a [[maximal torus]] in the theory of [[compact Lie group]]s.) Here <math>\mathfrak{t}_n</math> is not an ideal in <math>\mathfrak{gl}(n)</math> for <math>n\geq 2</math>. For example, when <math>n=2</math>, this follows from the calculation:
<blockquote><math>\begin{align}
\left[
\left[
\begin{bmatrix}
\begin{bmatrix}
a & b \\
a & b \\
c & d
c & d
\end{bmatrix},
\end{bmatrix},
\begin{bmatrix}
\begin{bmatrix}
x & 0 \\
x & 0 \\
Line 84: Line 99:
c(x-y) & 0
c(x-y) & 0
\end{bmatrix}
\end{bmatrix}
\end{align}</math></blockquote>
(which is not always in <math>\mathfrak{t}_2</math>).


\end{align}</math></blockquote>shows <math>\mathfrak{d}(2)</math> is a subalgebra, but not an ideal. In fact, every one-dimensional linear subspace of a Lie algebra has an induced abelian Lie algebra structure, which is generally not an ideal. For any simple Lie algebra, all abelian Lie algebras can never be ideals.
Every one-dimensional linear subspace of a Lie algebra <math>\mathfrak{g}</math> is an abelian Lie subalgebra, but it need not be an ideal.


=== Direct sum and semidirect product ===
=== Product and semidirect product ===
For two Lie algebras <math>\mathfrak{g^{}}</math> and <math>\mathfrak{g'}</math>, their [[Direct sum of modules|direct sum]] Lie algebra is the vector space <math>\mathfrak{g}\oplus\mathfrak{g'}</math>consisting of all pairs <math>\mathfrak{}(x,x'), \,x\in\mathfrak{g}, \ x'\in\mathfrak{g'}</math>, with the operation
For two Lie algebras <math>\mathfrak{g}</math> and <math>\mathfrak{g'}</math>, the ''[[direct product|product]]'' Lie algebra is the vector space <math>\mathfrak{g}\times \mathfrak{g'}</math> consisting of all ordered pairs <math>(x,x'), \,x\in\mathfrak{g}, \ x'\in\mathfrak{g'}</math>, with Lie bracket<ref>{{harvnb|Bourbaki|1989|loc=section I.1.1.}}</ref>
:<math> [(x,x'),(y,y')]=([x,y],[x',y']).</math>
This is the product in the [[product (category theory)|category]] of Lie algebras. Note that the copies of <math>\mathfrak g</math> and <math>\mathfrak g'</math> in <math>\mathfrak{g}\times \mathfrak{g'}</math> commute with each other: <math>[(x,0), (0,x')] = 0.</math>


Let <math>\mathfrak{g}</math> be a Lie algebra and <math>\mathfrak{i}</math> an ideal of <math>\mathfrak{g}</math>. If the canonical map <math>\mathfrak{g} \to \mathfrak{g}/\mathfrak{i}</math> splits (i.e., admits a section <math>\mathfrak{g}/\mathfrak{i}\to \mathfrak{g}</math>, as a homomorphism of Lie algebras), then <math>\mathfrak{g}</math> is said to be a [[semidirect product]] of <math>\mathfrak{i}</math> and <math>\mathfrak{g}/\mathfrak{i}</math>, <math>\mathfrak{g}=\mathfrak{g}/\mathfrak{i}\ltimes\mathfrak{i}</math>. See also [[Lie algebra extension#By semidirect sum|semidirect sum of Lie algebras]].
:<math> [(x,x'),(y,y')]=([x,y],[x',y']),</math>

so that the copies of <math>\mathfrak g, \mathfrak g'</math> commute with each other: <math>[(x,0), (0,x')] = 0.</math>

Let <math>\mathfrak{g}</math> be a Lie algebra and <math>\mathfrak{i}</math> an ideal of <math>\mathfrak{g}</math>. If the canonical map <math>\mathfrak{g} \to \mathfrak{g}/\mathfrak{i}</math> splits (i.e., admits a section), then <math>\mathfrak{g}</math> is said to be a [[semidirect product]] of <math>\mathfrak{i}</math> and <math>\mathfrak{g}/\mathfrak{i}</math>, <math>\mathfrak{g}=\mathfrak{g}/\mathfrak{i}\ltimes\mathfrak{i}</math>. See also [[Lie algebra extension#By semidirect sum|semidirect sum of Lie algebras]].

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


=== Derivations ===
=== Derivations ===
For an [[algebra over a field|algebra]] ''A'' over a field ''F'', a [[derivation (abstract algebra)|''derivation'']] of ''A'' is a [[linear map]] <math>D\colon A\to A</math> that satisfies the [[General Leibniz rule|Leibniz rule]]
For an [[algebra over a field|algebra]] ''A'' over a field ''F'', a [[derivation (abstract algebra)|''derivation'']] of ''A'' over ''F'' is a linear map <math>D\colon A\to A</math> that satisfies the [[product rule|Leibniz rule]]
:<math>D(xy) = D(x)y + xD(y)</math>
:<math>D(xy) = D(x)y + xD(y)</math>
for all <math>x,y\in A</math>. (The definition makes sense for a possibly [[non-associative algebra]].) Given two derivations <math>D_1</math> and <math>D_2</math>, their commutator <math>[D_1,D_2]:=D_1D_2-D_2D_1</math> is again a derivation. This operation makes the space <math>\text{Der}_k(A)</math> of all derivations of ''A'' over ''F'' into a Lie algebra.<ref>{{harvnb|Humphreys|1978|p=4}}</ref>
for all <math>x,y\in A</math>. (The definition makes sense for a possibly [[non-associative algebra]].) Given two derivations <math>D_1</math> and <math>D_2</math>, their commutator <math>[D_1,D_2]:=D_1D_2-D_2D_1</math> is again a derivation. This operation makes the space <math>\text{Der}_k(A)</math> of all derivations of ''A'' over ''F'' into a Lie algebra.<ref>{{harvnb|Humphreys|1978|p=4.}}</ref>


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 very natural way to construct Lie algebras: they are the "infinitesimal automorphisms" of ''A''. Indeed, writing out the condition that
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
:<math>(1+\epsilon D)(xy)=(1+\epsilon D)(x)\cdot (1+\epsilon D)(y) \pmod{\epsilon^2}</math>
:<math>(1+\epsilon D)(xy) \equiv (1+\epsilon D)(x)\cdot (1+\epsilon D)(y) \pmod{\epsilon^2}</math>
(where 1 denotes the identity map on ''A'') gives exactly the definition of ''D'' being a derivation.
(where 1 denotes the identity map on ''A'') gives exactly the definition of ''D'' being a derivation.


'''Example: the Lie algebra of vector fields.''' Let ''A'' be the ring <math>C^{\infty}(X)</math> of [[smooth function]]s on a smooth manifold ''X''. Then a derivation of ''A'' over '''R''' is equivalent to a [[vector field]] on ''X''. (A vector field ''v'' gives a derivation of the space of smooth functions by differentiating functions in the direction of ''v''.) This makes the space <math>\text{Vect}(X)</math> of vector fields into a Lie algebra (see [[Lie bracket of vector fields]]).<ref>{{harvnb|Varadarajan|2004|p=49}}</ref> Informally speaking, <math>\text{Vect}(X)</math> is the Lie algebra of the [[diffeomorphism group]] of ''X''. So the Lie bracket of vector fields describes the non-commutativity of the diffeomorphism group. An [[group action|action]] of a Lie group ''G'' on a manifold ''X'' determines a homomorphism of Lie algebras <math>\mathfrak{g}\to \text{Vect}(X)</math>.
'''Example: the Lie algebra of vector fields.''' Let ''A'' be the ring <math>C^{\infty}(X)</math> of [[smooth function]]s on a smooth manifold ''X''. Then a derivation of ''A'' over <math>\mathbb{R}</math> is equivalent to a [[vector field]] on ''X''. (A vector field ''v'' gives a derivation of the space of smooth functions by differentiating functions in the direction of ''v''.) This makes the space <math>\text{Vect}(X)</math> of vector fields into a Lie algebra (see [[Lie bracket of vector fields]]).<ref>{{harvnb|Varadarajan|1984|p=49.}}</ref> Informally speaking, <math>\text{Vect}(X)</math> is the Lie algebra of the [[diffeomorphism group]] of ''X''. So the Lie bracket of vector fields describes the non-commutativity of the diffeomorphism group. An [[group action|action]] of a Lie group ''G'' on a manifold ''X'' determines a homomorphism of Lie algebras <math>\mathfrak{g}\to \text{Vect}(X)</math>.


A Lie algebra can be viewed as a non-associative algebra, and so each Lie algebra <math>\mathfrak{g}</math> over a field ''F'' determines its Lie algebra of derivations, <math>\text{Der}_F(\mathfrak{g})</math>. That is, a derivation of <math>\mathfrak{g}</math> is a linear map <math>D\colon \mathfrak{g}\to \mathfrak{g}</math> such that
A Lie algebra can be viewed as a non-associative algebra, and so each Lie algebra <math>\mathfrak{g}</math> over a field ''F'' determines its Lie algebra of derivations, <math>\text{Der}_F(\mathfrak{g})</math>. That is, a derivation of <math>\mathfrak{g}</math> is a linear map <math>D\colon \mathfrak{g}\to \mathfrak{g}</math> such that
:<math>D([x,y])=[D(x),y]+[x,D(y)]</math>.
:<math>D([x,y])=[D(x),y]+[x,D(y)]</math>.
The ''inner derivation'' associated to any <math>x\in\mathfrak g</math> is the adjoint mapping <math>\mathrm{ad}_x</math> defined by <math>\mathrm{ad}_x(y):=[x,y]</math>. (This is a derivation as a consequence of the Jacobi identity.) That gives a homomorphism of Lie algebras, <math>\operatorname{ad}\colon\mathfrak{g}\to \text{Der}_F(\mathfrak{g})</math>. The image <math>\text{Inn}_F(\mathfrak{g})</math> is an ideal in <math>\text{Der}_F(\mathfrak{g})</math>, and the Lie algebra of '''outer derivations''' is defined as the quotient Lie algebra, <math>\text{Out}_F(\mathfrak{g})=\text{Der}_F(\mathfrak{g})/\text{Inn}_F(\mathfrak{g})</math>. (This is exactly analogous to the [[outer automorphism group]] of a group.) For a [[semisimple Lie algebra]] over a field of characteristic zero, every derivation is inner.
The ''inner derivation'' associated to any <math>x\in\mathfrak g</math> is the adjoint mapping <math>\mathrm{ad}_x</math> defined by <math>\mathrm{ad}_x(y):=[x,y]</math>. (This is a derivation as a consequence of the Jacobi identity.) That gives a homomorphism of Lie algebras, <math>\operatorname{ad}\colon\mathfrak{g}\to \text{Der}_F(\mathfrak{g})</math>. The image <math>\text{Inn}_F(\mathfrak{g})</math> is an ideal in <math>\text{Der}_F(\mathfrak{g})</math>, and the Lie algebra of ''outer derivations'' is defined as the quotient Lie algebra, <math>\text{Out}_F(\mathfrak{g})=\text{Der}_F(\mathfrak{g})/\text{Inn}_F(\mathfrak{g})</math>. (This is exactly analogous to the [[outer automorphism group]] of a group.) For a [[semisimple Lie algebra]] (defined below) over a field of characteristic zero, every derivation is inner.<ref>{{harvnb|Serre|2006|loc=Part I, section VI.3.}}</ref> This is related to the theorem that the outer automorphism group of a semisimple Lie group is finite.<ref>{{harvnb|Fulton|Harris|1991|loc=Proposition D.40.}}</ref>


In contrast, an abelian Lie algebra has many outer derivations. Namely, for a vector space <math>V</math> with Lie bracket zero, the Lie algebra <math>\text{Out}_F(V)</math> can be identified with <math>\mathfrak{gl}(V)</math>.
==== Examples ====
For example, given a Lie algebra ideal <math>\mathfrak{i} \subset \mathfrak{g}</math> the adjoint representation <math>\mathfrak{ad}_\mathfrak {g}</math> of <math>\mathfrak{g}</math> acts as outer derivations on <math>\mathfrak{i}</math> since <math>[x,i] \subset \mathfrak{i}</math> for any <math>x \in \mathfrak{g}</math> and <math>i \in \mathfrak{i}</math>. For the Lie algebra <math>\mathfrak{b}_n</math> of upper triangular matrices in <math>\mathfrak{gl}(n)</math>, it has an ideal <math>\mathfrak{n}_n</math> of strictly upper triangular matrices (where the only non-zero elements are above the diagonal of the matrix). For instance, the commutator of elements in <math>\mathfrak{b}_3</math> and <math>\mathfrak{n}_3</math> gives<blockquote><math>\begin{align}
\left[
\begin{bmatrix}
a & b & c \\
0 & d & e \\
0 & 0 & f
\end{bmatrix},
\begin{bmatrix}
0 & x & y \\
0 & 0 & z \\
0 & 0 & 0
\end{bmatrix}
\right] &= \begin{bmatrix}
0 & ax & ay+bz \\
0 & 0 & dz \\
0 & 0 & 0
\end{bmatrix} - \begin{bmatrix}
0 & dx & ex+yf \\
0 & 0 & fz \\
0 & 0 & 0
\end{bmatrix} \\
&= \begin{bmatrix}
0 & (a-d)x & (a-f)y-ex+bz \\
0 & 0 & (d-f)z \\
0 & 0 & 0
\end{bmatrix}
\end{align}</math></blockquote>shows there exist outer derivations from <math>\mathfrak{b}_3</math> in <math>\text{Der}(\mathfrak{n}_3)</math>.

=== Split Lie algebra ===
Let ''V'' be a finite-dimensional vector space over a field ''F'', <math>\mathfrak{gl}(V)</math> the Lie algebra of linear transformations and <math>\mathfrak{g} \subseteq \mathfrak{gl}(V)</math> a Lie subalgebra. Then <math>\mathfrak{g}</math> is said to be '''split''' if the roots of the characteristic polynomials of all linear transformations in <math>\mathfrak{g}</math> are in the base field ''F''.<ref>{{harvnb|Jacobson|1962|p=42}}</ref> More generally, a finite-dimensional Lie algebra <math>\mathfrak{g}</math> is said to be split if it has a Cartan subalgebra whose image under the [[adjoint representation]] <math>\operatorname{ad}: \mathfrak{g} \to \mathfrak{gl}(\mathfrak g)</math> is a split Lie algebra. A [[split real form]] of a complex semisimple Lie algebra (cf. [[#Real form and complexification]]) is an example of a split real Lie algebra. See also [[split Lie algebra]] for further information.

=== Vector space basis ===
For practical calculations, it is often convenient to choose an explicit [[vector space basis]] for the algebra. A common construction for this basis is sketched in the article [[structure constant]]s.

===Definition using category-theoretic notation===
Although the definitions above are sufficient for a conventional understanding of Lie algebras, once this is understood, additional insight can be gained by using notation common to [[category theory]], that is, by defining a Lie algebra in terms of [[linear map]]s—that is, [[morphism]]s of the [[category of vector spaces]]—without considering individual elements. (In this section, the [[field (mathematics)|field]] over which the algebra is defined is supposed to be of [[characteristic (algebra)|characteristic]] different from two.)

For the category-theoretic definition of Lie algebras, two [[tensor product#Tensor powers and braiding|braiding isomorphisms]] are needed. If {{mvar|A}} is a vector space, the ''interchange isomorphism'' <math>\tau: A\otimes A \to A\otimes A</math> is defined by
:<math>\tau(x\otimes y)= y\otimes x.</math>
The ''cyclic-permutation braiding'' <math>\sigma:A\otimes A\otimes A \to A\otimes A\otimes A </math> is defined as
:<math>\sigma=(\mathrm{id}\otimes \tau)\circ(\tau\otimes \mathrm{id}),</math>
where <math>\mathrm{id}</math> is the identity morphism.
Equivalently, <math>\sigma</math> is defined by
:<math>\sigma(x\otimes y\otimes z)= y\otimes z\otimes x.</math>

With this notation, a Lie algebra can be defined as an [[object (category theory)|object]] <math>A</math> in the category of vector spaces together with a [[morphism]]
:<math>[\cdot,\cdot]:A\otimes A\rightarrow A</math>
that satisfies the two morphism equalities
:<math>[\cdot,\cdot]\circ(\mathrm{id}+\tau)=0,</math>
and
:<math>[\cdot,\cdot]\circ ([\cdot,\cdot]\otimes \mathrm{id}) \circ (\mathrm{id} +\sigma+\sigma^2)=0.</math>


== Examples ==
== Examples ==
=== Matrix Lie algebras ===
A [[Linear group|matrix group]] is a Lie group consisting of invertible matrices, <math>G\subset \mathrm{GL}(n,\mathbb{R})</math>, where the group operation of ''G'' is matrix multiplication. The corresponding Lie algebra <math>\mathfrak g</math> is the space of matrices which are tangent vectors to ''G'' inside the linear space <math>M_n(\mathbb{C})</math>: this consists of derivatives of smooth curves in ''G'' at the [[identity matrix]] <math>I</math>:
:<math>\mathfrak{g} = \{ X = c'(0) \in M_n(\mathbb{C}) \ \mid\ \text{ smooth } c: \mathbb{R}\to G, \ c(0) = I \}.</math>
The Lie bracket of <math>\mathfrak{g}</math> is given by the commutator of matrices, <math>[X,Y]=XY-YX</math>. Given a Lie algebra <math>\mathfrak{g}\subset \mathfrak{gl}(n,\mathbb{R})</math>, one can recover the Lie group as the subgroup generated by the [[matrix exponential]] of elements of <math>\mathfrak{g}</math>.<ref>{{harvnb|Varadarajan|1984|loc=section 2.10, Remark 2.}}</ref> (To be precise, this gives the [[identity component]] of ''G'', if ''G'' is not connected.) Here the exponential mapping <math>\exp: M_n(\mathbb{R})\to M_n(\mathbb{R})</math> is defined by <math>\exp(X) = I + X + \tfrac{1}{2!}X^2 + \tfrac{1}{3!}X^3 + \cdots</math>, which converges for every matrix <math>X</math>.


The same comments apply to complex Lie subgroups of <math>GL(n,\mathbb{C})</math> and the complex matrix exponential, <math>\exp: M_n(\mathbb{C})\to M_n(\mathbb{C})</math> (defined by the same formula).
=== Vector spaces ===
Any vector space <math>V</math> endowed with the identically zero Lie bracket becomes a Lie algebra. Such Lie algebras are called [[Abelian Lie algebra|abelian]], cf. below. Any one-dimensional Lie algebra over a field of characteristic different from 2 is abelian, by the alternating property of the Lie bracket.


Here are some matrix Lie groups and their Lie algebras.<ref>{{harvnb|Hall|2015|loc=§3.4.}}</ref>
=== Associative algebra with commutator bracket ===
* On an [[associative algebra]] <math>A</math> over a field <math>F</math> with multiplication <math>(x, y) \mapsto xy</math>, a Lie bracket may be defined by the [[Commutator#Ring theory|commutator]] <math>[x,y] = xy - yx</math>. With this bracket, <math>A</math> is a Lie algebra.<ref>{{harvnb|Bourbaki|1989|loc=§1.2. Example 1.}}</ref> The associative algebra ''A'' is called an ''enveloping algebra'' of the Lie algebra <math>(A, [\,\cdot\, , \cdot \,])</math>. Every Lie algebra can be embedded into one that arises from an associative algebra in this fashion; see [[universal enveloping algebra]].
* The [[endomorphism ring|associative algebra of the endomorphisms]] of an ''F''-vector space <math>V</math> with the above Lie bracket is denoted <math>\mathfrak{gl}(V)</math>.
*For a finite dimensional vector space <math>V = F^n</math>, the previous example is exactly the Lie algebra of ''n'' × ''n'' matrices, denoted <math>\mathfrak{gl}(n, F)</math> or <math>\mathfrak{gl}_n(F)</math>,<ref>{{harvnb|Bourbaki|1989|loc=§1.2. Example 2.}}</ref> and with bracket <math>[X,Y]=XY-YX</math> where adjacency indicates matrix multiplication. This is the Lie algebra of the [[general linear group]], consisting of invertible matrices.


* For a positive integer ''n'', the [[special linear group]] <math>\mathrm{SL}(n,\mathbb{R})</math> consists of all real {{math|''n''&nbsp;×&nbsp;''n''}} matrices with determinant 1. This is the group of linear maps from <math>\mathbb{R}^n</math> to itself that preserve volume and [[orientability|orientation]]. More abstractly, <math>\mathrm{SL}(n,\mathbb{R})</math> is the [[commutator subgroup]] of the general linear group <math>\mathrm{GL}(n,\R)</math>. Its Lie algebra <math>\mathfrak{sl}(n,\mathbb{R})</math> consists of all real {{math|''n''&nbsp;×&nbsp;''n''}} matrices with [[trace (mathematics)|trace]] 0. Similarly, one can define the analogous complex Lie group <math>{\rm SL}(n,\mathbb{C})</math> and its Lie algebra <math>\mathfrak{sl}(n,\mathbb{C})</math>.
=== Special matrices ===
* The [[orthogonal group]] <math>\mathrm{O}(n)</math> plays a basic role in geometry: it is the group of linear maps from <math>\mathbb{R}^n</math> to itself that preserve the length of vectors. For example, rotations and reflections belong to <math>\mathrm{O}(n)</math>. Equivalently, this is the group of ''n'' x ''n'' orthogonal matrices, meaning that <math>A^{\mathrm{T}}=A^{-1}</math>. The orthogonal group has two connected components; the identity component is called the ''special orthogonal group'' <math>\mathrm{SO}(n)</math>, consisting of the orthogonal matrices with determinant 1. Both groups have the same Lie algebra <math>\mathfrak{so}(n)</math>, the subspace of skew-symmetric matrices in <math>\mathfrak{gl}(n,\mathbb{R})</math> (<math>X^{\rm T}=-X</math>). See also [[Skew-symmetric matrix#Infinitesimal rotations|infinitesimal rotations with skew-symmetric matrices]].
Two important subalgebras of <math>\mathfrak{gl}_n(F)</math> are:
:The complex orthogonal group <math>\mathrm{O}(n,\mathbb{C})</math>, its identity component <math>\mathrm{SO}(n,\mathbb{C})</math>, and the Lie algebra <math>\mathfrak{so}(n,\mathbb{C})</math> are given by the same formulas applied to ''n'' x ''n'' complex matrices. Equivalently, <math>\mathrm{O}(n,\mathbb{C})</math> is the subgroup of <math>\mathrm{GL}(n,\mathbb{C})</math> that preserves the standard [[symmetric bilinear form]] on <math>\mathbb{C}^n</math>.
* The [[unitary group]] <math>\mathrm{U}(n)</math> is the subgroup of <math>\mathrm{GL}(n,\mathbb{C})</math> that preserves the length of vectors in <math>\mathbb{C}^n</math> (with respect to the standard [[Hermitian inner product]]). Equivalently, this is the group of ''n''&nbsp;×&nbsp;''n'' unitary matrices (satisfying <math>A^*=A^{-1}</math>, where <math>A^*</math> denotes the [[conjugate transpose]] of a matrix). Its Lie algebra <math>\mathfrak{u}(n)</math> consists of the skew-hermitian matrices in <math>\mathfrak{gl}(n,\mathbb{C})</math> (<math>X^*=-X</math>). This is a Lie algebra over <math>\mathbb{R}</math>, not over <math>\mathbb{C}</math>. (Indeed, ''i'' times a skew-hermitian matrix is hermitian, rather than skew-hermitian.) Likewise, the unitary group <math>\mathrm{U}(n)</math> is a real Lie subgroup of the complex Lie group <math>\mathrm{GL}(n,\mathbb{C})</math>. For example, <math>\mathrm{U}(1)</math> is the [[circle group]], and its Lie algebra (from this point of view) is <math>i\mathbb{R}\subset \mathbb{C}=\mathfrak{gl}(1,\mathbb{C})</math>.
* The [[special unitary group]] <math>\mathrm{SU}(n)</math> is the subgroup of matrices with determinant 1 in <math>\mathrm{U}(n)</math>. Its Lie algebra <math>\mathrm{su}(n)</math> consists of the skew-hermitian matrices with trace zero.
*The [[symplectic group]] <math>\mathrm{Sp}(2n,\R)</math> is the subgroup of <math>\mathrm{GL}(2n,\mathbb{R})</math> that preserves the standard [[symplectic vector space|alternating bilinear form]] on <math>\mathbb{R}^{2n}</math>. Its Lie algebra is the [[symplectic Lie algebra]] <math>\mathfrak{sp}(2n,\mathbb{R})</math>.
*The [[classical Lie algebra]]s are those listed above, along with variants over any field.


=== Two dimensions ===
* The matrices of [[Trace (linear algebra)|trace]] zero form the [[special linear Lie algebra]] <math>\mathfrak{sl}_n(F)</math>, the Lie algebra of the [[special linear group]] <math>\mathrm{SL}_n(F)</math>.<ref>{{harvnb|Humphreys|1978|p=2}}</ref>
Some Lie algebras of low dimension are described here. See the [[classification of low-dimensional real Lie algebras]] for further examples.
*The [[skew-hermitian]] matrices form the unitary Lie algebra <math>\mathfrak u(n)</math>, the Lie algebra of the [[unitary group]]&nbsp;''U''(''n'').


* There is a unique nonabelian Lie algebra <math>\mathfrak{g}</math> of dimension 2 over any field ''F'', up to isomorphism.<ref>{{harvnb|Erdmann|Wildon|2006|loc=Theorem 3.1.}}</ref> Here <math>\mathfrak{g}</math> has a basis <math>X,Y</math> for which the bracket is given by <math> \left [X, Y\right ] = Y</math>. (This determines the Lie bracket completely, because the axioms imply that <math>[X,X]=0</math> and <math>[Y,Y]=0</math>.) Over the real numbers, <math>\mathfrak{g}</math> can be viewed as the Lie algebra of the Lie group <math>G=\mathrm{Aff}(1,\mathbb{R})</math> of [[Affine group|affine transformations]] of the real line, <math>x\mapsto ax+b</math>.
=== Matrix Lie algebras ===
A complex [[Linear group|matrix group]] is a Lie group consisting of matrices, <math>G\subset M_n(\mathbb{C})</math>, where the multiplication of ''G'' is matrix multiplication. The corresponding Lie algebra <math>\mathfrak g</math> is the space of matrices which are tangent vectors to ''G'' inside the linear space <math>M_n(\mathbb{C})</math>: this consists of derivatives of smooth curves in ''G'' at the identity: <blockquote><math>\mathfrak{g} = \{ X = c'(0) \in M_n(\mathbb{C}) \ \mid\ \text{ smooth } c : \mathbb{R}\to G, \ c(0) = I \}.</math></blockquote>The Lie bracket of <math>\mathfrak{g}</math> is given by the commutator of matrices, <math>[X,Y]=XY-YX</math>. Given the Lie algebra, one can recover the Lie group as the image of the [[matrix exponential]] mapping <math>\exp: M_n(\mathbb{C})\to M_n(\mathbb{C})</math> defined by <math>\exp(X) = I + X + \tfrac{1}{2!}X^2+\cdots</math>, which converges for every matrix <math>X</math>: that is, <math>G=\exp(\mathfrak g)</math>.
The following are examples of Lie algebras of matrix Lie groups:<ref>{{harvnb|Hall|2015|loc=§3.4}}</ref>


:The affine group ''G'' can be identified with the group of matrices
* The [[special linear group]] <math>{\rm SL}_n(\mathbb{C})</math>, consisting of all {{math|''n''&nbsp;×&nbsp;''n''}} matrices with determinant 1. Its Lie algebra <math>\mathfrak{sl}_n(\mathbb{C})</math> consists of all {{math|''n''&nbsp;×&nbsp;''n''}} matrices with complex entries and trace 0. Similarly, one can define the corresponding real Lie group <math>{\rm SL}_n(\mathbb{R})</math> and its Lie algebra <math>\mathfrak{sl}_n(\mathbb{R})</math>.
::<math> \left( \begin{array}{cc} a & b\\ 1 & 0 \end{array} \right) </math>
* The [[unitary group]] <math>U(n)</math> consists of ''n''&nbsp;×&nbsp;''n'' unitary matrices (satisfying <math>U^*=U^{-1}</math>). Its Lie algebra <math>\mathfrak{u}(n)</math> consists of skew-self-adjoint matrices (<math>X^*=-X</math>).
:under matrix multiplication, with <math>a,b \in \mathbb{R} </math>, <math>a \neq 0</math>. Its Lie algebra is the Lie subalgebra <math>\mathfrak{g}</math> of <math>\mathfrak{gl}(2,\mathbb{R})</math> consisting of all matrices
* The special [[orthogonal group]] <math>\mathrm{SO}(n)</math>, consisting of real determinant-one orthogonal matrices (<math>A^{\mathrm{T}}=A^{-1}</math>). Its Lie algebra <math>\mathfrak{so}(n)</math> consists of real skew-symmetric matrices (<math>X^{\rm T}=-X</math>). The full orthogonal group <math>\mathrm{O}(n)</math>, without the determinant-one condition, consists of <math>\mathrm{SO}(n)</math> and a separate connected component, so it has the ''same'' Lie algebra as <math>\mathrm{SO}(n)</math>. See also [[Skew-symmetric matrix#Infinitesimal rotations|infinitesimal rotations with skew-symmetric matrices]]. Similarly, one can define a complex version of this group and algebra, simply by allowing complex matrix entries.
::<math> \left( \begin{array}{cc} c & d\\ 0 & 0 \end{array}\right). </math>
:In these terms, the basis above for <math>\mathfrak{g}</math> is given by the matrices
::<math> X= \left( \begin{array}{cc} 1 & 0\\ 0 & 0 \end{array}\right), \qquad Y= \left( \begin{array}{cc} 0 & 1\\ 0 & 0 \end{array}\right). </math>


:For any field <math>F</math>, the 1-dimensional subspace <math>F\cdot Y</math> is an ideal in the 2-dimensional Lie algebra <math>\mathfrak{g}</math>, by the formula <math>[X,Y]=Y\in F\cdot Y</math>. Both of the Lie algebras <math>F\cdot Y</math> and <math>\mathfrak{g}/(F\cdot Y)</math> are abelian (because 1-dimensional). In this sense, <math>\mathfrak{g}</math> can be broken into abelian "pieces", meaning that it is solvable (though not nilpotent), in the terminology below.
=== Two dimensions ===

* On any field <math>F</math> there is, up to isomorphism, a single two-dimensional nonabelian Lie algebra. With generators ''x, y,'' its bracket is defined as <math> \left [x, y\right ] = y</math>. It generates the [[Affine group#Matrix representation|affine group in one dimension]].

:This can be realized by the matrices:
::<math> x= \left( \begin{array}{cc} 1 & 0\\ 0 & 0 \end{array}\right), \qquad y= \left( \begin{array}{cc} 0 & 1\\ 0 & 0 \end{array}\right). </math>
Since
:<math> \left( \begin{array}{cc} 1 & c\\ 0 & 0 \end{array}\right)^{n+1} = \left( \begin{array}{cc} 1 & c\\ 0 & 0 \end{array}\right)</math>
for any natural number <math>n</math> and any <math>c</math>, one sees that the resulting Lie group elements are upper triangular 2×2 matrices with unit lower diagonal:
::<math> \exp(a\cdot{}x+b\cdot{}y)= \left( \begin{array}{cc} e^a & \tfrac{b}{a}(e^a-1)\\ 0 & 1 \end{array}\right) = 1 + \tfrac{e^a-1}{a}\left(a\cdot{}x+b\cdot{}y\right). </math>


=== Three dimensions ===
=== Three dimensions ===
* The [[Heisenberg algebra]] <math>{\rm H}_3(\mathbb{R})</math> is a three-dimensional Lie algebra generated by elements {{mvar|x}}, {{mvar|y}}, and {{mvar|z}} with Lie brackets
* The [[Heisenberg algebra]] <math>\mathfrak{h}_3(F)</math> over a field ''F'' is the three-dimensional Lie algebra with a basis <math>X,Y,Z</math> such that<ref>{{harvnb|Erdmann|Wildon|2006|loc=section 3.2.1.}}</ref>
::<math>[X,Y] = Z,\quad [X,Z] = 0, \quad [Y,Z] = 0</math>.

:It can be viewed as the Lie algebra of 3&times;3 strictly [[upper-triangular]] matrices, with the commutator Lie bracket and the basis
::<math>[x,y] = z,\quad [x,z] = 0, \quad [y,z] = 0</math>.
:It is usually realized as the space of 3&times;3 strictly upper-triangular matrices, with the commutator Lie bracket and the basis
::<math>
::<math>
x = \left( \begin{array}{ccc}
X = \left( \begin{array}{ccc}
0&1&0\\
0&1&0\\
0&0&0\\
0&0&0\\
0&0&0
0&0&0
\end{array}\right),\quad
\end{array}\right),\quad
y = \left( \begin{array}{ccc}
Y = \left( \begin{array}{ccc}
0&0&0\\
0&0&0\\
0&0&1\\
0&0&1\\
0&0&0
0&0&0
\end{array}\right),\quad
\end{array}\right),\quad
z = \left( \begin{array}{ccc}
Z = \left( \begin{array}{ccc}
0&0&1\\
0&0&1\\
0&0&0\\
0&0&0\\
Line 224: Line 182:
</math>
</math>


:Over the real numbers, <math>\mathfrak{h}_3(\mathbb{R})</math> is the Lie algebra of the [[Heisenberg group]] <math>\mathrm{H}_3(\mathbb{R})</math>, that is, the group of matrices
:Any element of the [[Heisenberg group]] has a representation as a product of group generators, i.e., [[matrix exponential]]s of these Lie algebra generators,
::<math>\left( \begin{array}{ccc}
::<math>\left( \begin{array}{ccc}
1&a&c\\
1&a&c\\
0&1&b\\
0&1&b\\
0&0&1
0&0&1
\end{array}\right)= e^{by} e^{cz} e^{ax}~.
\end{array}\right)
</math>
</math>
:under matrix multiplication.


:For any field ''F'', the center of <math>\mathfrak{h}_3(F)</math> is the 1-dimensional ideal <math>F\cdot Z</math>, and the quotient <math>\mathfrak{h}_3(F)/(F\cdot Z)</math> is abelian, isomorphic to <math>F^2</math>. In the terminology below, it follows that <math>\mathfrak{h}_3(F)</math> is nilpotent (though not abelian).
* The Lie algebra <math>\mathfrak{so}(3)</math> of the group SO(3) is spanned by the three matrices<ref>{{harvnb|Hall|2015|loc=Example 3.27}}</ref>

* The Lie algebra <math>\mathfrak{so}(3)</math> of the rotation group <math>\mathrm{SO}(3)</math> is the space of skew-symmetric 3 x 3 matrices over <math>\mathbb{R}</math>. A basis is given by the three matrices<ref>{{harvnb|Hall|2015|loc=Example 3.27.}}</ref>
::<math>
::<math>
F_1 = \left( \begin{array}{ccc}
F_1 = \left( \begin{array}{ccc}
Line 252: Line 213:
:The commutation relations among these generators are
:The commutation relations among these generators are
::<math>[F_1, F_2] = F_3,</math>
::<math>[F_1, F_2] = F_3,</math>
:: <math>[F_2, F_3] = F_1,</math>
::<math>[F_2, F_3] = F_1,</math>
:: <math>[F_3, F_1] = F_2.</math>
::<math>[F_3, F_1] = F_2.</math>


:The three-dimensional [[Euclidean space]] <math>\mathbb{R}^3</math> with the Lie bracket given by the [[cross product]] of [[Vector (geometric)|vectors]] has the same commutation relations as above: thus, it is isomorphic to <math>\mathfrak{so}(3)</math>. This Lie algebra is unitarily equivalent to the usual [[Spin (physics)]] angular-momentum component operators for spin-1 particles in [[quantum mechanics]].
:The cross product of vectors in <math>\mathbb{R}^3</math> is given by the same formula in terms of the standard basis; so that Lie algebra is isomorphic to <math>\mathfrak{so}(3)</math>. Also, <math>\mathfrak{so}(3)</math> is equivalent to the [[Spin (physics)]] angular-momentum component operators for spin-1 particles in [[quantum mechanics]].<ref name="quantum">{{harvnb|Wigner|1959|loc=Chapters 17 and 20.}}</ref>

:The Lie algebra <math>\mathfrak{so}(3)</math> cannot be broken into pieces in the way that the previous examples can: it is ''simple'', meaning that it is not abelian and its only ideals are 0 and all of <math>\mathfrak{so}(3)</math>.

* Another simple Lie algebra of dimension 3, in this case over <math>\mathbb{C}</math>, is the space <math>\mathfrak{sl}(2,\mathbb{C})</math> of 2 x 2 matrices of trace zero. A basis is given by the three matrices
:<math>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).</math>
:The Lie bracket is given by:
::<math>[H, E] = 2E,</math>
::<math>[H, F] = -2F,</math>
::<math>[E, F] = H.</math>

:Using these formulas, one can show that the Lie algebra <math>\mathfrak{sl}(2,\mathbb{C})</math> is simple, and classify its finite-dimensional representations (defined below).<ref>{{harvnb|Erdmann|Wildon|2006|loc=Chapter 8.}}</ref> In the terminology of quantum mechanics, one can think of ''E'' and ''F'' as [[ladder operator|raising and lowering operators]]. Indeed, for any representation of <math>\mathfrak{sl}(2,\mathbb{C})</math>, the relations above imply that ''E'' maps the ''c''-[[eigenspace]] of ''H'' (for a complex number ''c'') into the <math>(c+2)</math>-eigenspace, while ''F'' maps the ''c''-eigenspace into the <math>(c-2)</math>-eigenspace.

:The Lie algebra <math>\mathfrak{sl}(2,\mathbb{C})</math> is isomorphic to the [[complexification]] of <math>\mathfrak{so}(3)</math>, meaning the [[tensor product]] <math>\mathfrak{so}(3)\otimes_{\mathbb{R}}\mathbb{C}</math>. The formulas for the Lie bracket are easier to analyze in the case of <math>\mathfrak{sl}(2,\mathbb{C})</math>. As a result, it is common to analyze complex representations of the group <math>\mathrm{SO}(3)</math> by relating them to representations of the Lie algebra <math>\mathfrak{sl}(2,\mathbb{C})</math>.


=== Infinite dimensions ===
=== Infinite dimensions ===
* The Lie algebra of vector fields on a smooth manifold of positive dimension is an infinite-dimensional Lie algebra over <math>\mathbb{R}</math>.
*[[Kac–Moody algebra]]s are a large class of infinite-dimensional Lie algebras whose structure is very similar to the finite-dimensional cases above.
* The [[Moyal bracket|Moyal algebra]] is an infinite-dimensional Lie algebra that contains all [[Classical Lie groups#Relationship with bilinear forms|classical Lie algebra]]s as subalgebras.
* The [[Kac–Moody algebra]]s are a large class of infinite-dimensional Lie algebras, say over <math>\mathbb{C}</math>, with structure much like that of the finite-dimensional simple Lie algebras (such as <math>\mathfrak{sl}(n,\C)</math>).
* The [[Moyal bracket|Moyal algebra]] is an infinite-dimensional Lie algebra that contains all the [[Classical Lie groups#Relationship with bilinear forms|classical Lie algebra]]s as subalgebras.
* The [[Virasoro algebra]] is of paramount importance in [[string theory]].
* The [[Virasoro algebra]] is important in [[string theory]].
* The functor that takes a Lie algebra over a field ''F'' to the underlying vector space has a [[left adjoint]] <math>V\mapsto L(V)</math>, called the ''[[free Lie algebra]]'' on a vector space ''V''. It is spanned by all iterated Lie brackets of elements of ''V'', modulo only the relations coming from the definition of a Lie algebra. The free Lie algebra <math>L(V)</math> is infinite-dimensional for ''V'' of dimension at least 2.<ref>{{harvnb|Serre|2006|loc=Part I, Chapter IV.}}</ref>


== Representations ==
== Representations ==
Line 266: Line 243:


===Definitions===
===Definitions===
Given a vector space ''V'', let <math>\mathfrak{gl}(V)</math> denote the Lie algebra consisting of all linear [[endomorphism]]s of ''V'', with bracket given by <math>[X,Y]=XY-YX</math>. A ''representation'' of a Lie algebra <math>\mathfrak{g}</math> on ''V'' is a Lie algebra homomorphism
Given a vector space ''V'', let <math>\mathfrak{gl}(V)</math> denote the Lie algebra consisting of all linear maps from ''V'' to itself, with bracket given by <math>[X,Y]=XY-YX</math>. A ''representation'' of a Lie algebra <math>\mathfrak{g}</math> on ''V'' is a Lie algebra homomorphism
:<math>\pi: \mathfrak g \to \mathfrak{gl}(V).</math>
:<math>\pi\colon \mathfrak g \to \mathfrak{gl}(V).</math>
That is, <math>\pi</math> sends each element of <math>\mathfrak{g}</math> to a linear map from ''V'' to itself, in such a way that the Lie bracket on <math>\mathfrak{g}</math> corresponds to the commutator of linear maps.


A representation is said to be ''faithful'' if its kernel is zero. [[Ado's theorem]]<ref>{{harvnb|Jacobson|1962|loc=Ch. VI}}</ref> states that every finite-dimensional Lie algebra over a field of characteristic zero has a faithful representation on a finite-dimensional vector space.
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.<ref>{{harvnb|Jacobson|1979|loc=Ch. VI.}}</ref> Equivalently, every finite-dimensional Lie algebra over a field ''F'' is isomorphic to a Lie subalgebra of <math>\mathfrak{gl}(n,F)</math> for some positive integer ''n''.


===Adjoint representation===
===Adjoint representation===
Line 277: Line 255:


===Goals of representation theory===
===Goals of representation theory===
One important aspect of the study of Lie algebras (especially semisimple Lie algebras) is the study of their representations. (Indeed, most of the books listed in the references section devote a substantial fraction of their pages to representation theory.) 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 <math>\mathfrak{g}</math>. Indeed, in the semisimple case, the adjoint representation is already faithful. Rather the goal is to understand ''all'' possible representations of <math>\mathfrak{g}</math>, up to the natural notion of equivalence. In the semisimple case over a field of characteristic zero, [[Weyl's theorem on complete reducibility|Weyl's theorem]]<ref>{{harvnb|Hall|2015|loc=Theorem 10.9}}</ref> says that every finite-dimensional representation is a direct sum of irreducible representations (those with no nontrivial invariant subspaces). The irreducible representations, in turn, are classified by a [[Lie algebra representation#Classifying finite-dimensional representations of Lie algebras|theorem of the highest weight]].
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 <math>\mathfrak{g}</math>. Indeed, in the semisimple case, the adjoint representation is already faithful. Rather, the goal is to understand all possible representations of <math>\mathfrak{g}</math>. For a semisimple Lie algebra over a field of characteristic zero, [[Weyl's theorem on complete reducibility|Weyl's theorem]]<ref name="reducibility">{{harvnb|Hall|2015|loc=Theorem 10.9.}}</ref> says that every finite-dimensional representation is a direct sum of irreducible representations (those with no nontrivial invariant subspaces). The finite-dimensional irreducible representations are well understood from several points of view; see the [[representation theory of semisimple Lie algebras]] and the [[Weyl character formula]].

===Universal enveloping algebra===
{{main|Universal enveloping algebra}}
The functor that takes an associative algebra ''A'' over a field ''F'' to ''A'' as a Lie algebra (by <math>[X,Y]:=XY-YX</math>) has a [[left adjoint]] <math>\mathfrak{g}\mapsto U(\mathfrak{g})</math>, called the '''universal enveloping algebra'''. To construct this: given a Lie algebra <math>\mathfrak{g}</math>, let
:<math>T(\mathfrak{g})=F\oplus \mathfrak{g} \oplus (\mathfrak{g}\otimes\mathfrak{g}) \oplus (\mathfrak{g}\otimes\mathfrak{g}\otimes\mathfrak{g})\oplus \cdots</math>
be the [[tensor algebra]] on <math>\mathfrak{g}</math>, also called the free associative algebra on the vector space <math>\mathfrak{g}</math>. Here <math>\otimes</math> denotes the [[tensor product]] of ''F''-vector spaces. Let ''I'' be the two-sided [[ideal (ring theory)|ideal]] in <math>T(\mathfrak{g})</math> generated by the elements <math>XY-YX-[X,Y]</math> for <math>X,Y\in\mathfrak{g}</math>; then the universal enveloping algebra is the quotient ring <math>U(\mathfrak{g}) = T(\mathfrak{g}) / I</math>. It satisfies the [[Poincaré–Birkhoff–Witt theorem]]: if <math>e_1,\ldots,e_n</math> is a basis for <math>\mathfrak{g}</math> as a ''k''-vector space, then a basis for <math>U(\mathfrak{g})</math> is given by all ordered products <math>e_1^{i_1}\cdots e_n^{i_n}</math> with <math>i_1,\ldots,i_n</math> natural numbers. In particular, the map <math>\mathfrak{g}\to U(\mathfrak{g})</math> is injective.<ref>{{harvnb|Humphreys|1978|loc=section 17.3.}}</ref>

Representations of <math>\mathfrak{g}</math> are equivalent to [[module (mathematics)|modules]] over the universal enveloping algebra. The fact that <math>\mathfrak{g}\to U(\mathfrak{g})</math> is injective implies that every Lie algebra (possibly of infinite dimension) has a faithful representation (of infinite dimension), namely its representation on <math>U(\mathfrak{g})</math>. This also shows that every Lie algebra is contained in the Lie algebra associated to some associative algebra.


===Representation theory in physics===
===Representation theory in physics===
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 would be the [[angular momentum operator]]s, whose commutation relations are those of the Lie algebra <math>\mathfrak{so}(3)</math> of the [[rotation group SO(3)]]. Typically, the space of states is very far from being irreducible under the pertinent operators, but one can attempt to decompose it into irreducible pieces. In doing so, one needs to know the irreducible representations of the given Lie algebra. In the study of the quantum [[Hydrogen-like atom|hydrogen atom]], for example, quantum mechanics textbooks give (without calling it that) a classification of the irreducible representations of the Lie algebra <math>\mathfrak{so}(3)</math>.
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 operator]]s, whose commutation relations are those of the Lie algebra <math>\mathfrak{so}(3)</math> of the [[rotation group SO(3)]]. Typically, the space of states is far from being irreducible under the pertinent operators, but
one can attempt to decompose it into irreducible pieces. In doing so, one needs to know the irreducible representations of the given Lie algebra. In the study of the quantum [[Hydrogen-like atom|hydrogen atom]], for example, quantum mechanics textbooks classify (more or less explicitly) the finite-dimensional irreducible representations of the Lie algebra <math>\mathfrak{so}(3)</math>.<ref name="quantum" />


== Structure theory and classification ==
== Structure theory and classification ==
Lie algebras can be classified to some extent. This is a powerful approach to the classification of Lie groups.

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


=== Abelian, nilpotent, and solvable ===
=== Abelian, nilpotent, and solvable ===
Analogously to abelian, nilpotent, and solvable groups, defined in terms of the derived subgroups, one can define abelian, nilpotent, and solvable Lie algebras.
Analogously to [[abelian group|abelian]], [[nilpotent group|nilpotent]], and [[solvable group]]s, one can define abelian, nilpotent, and solvable Lie algebras.


A Lie algebra <math>\mathfrak{g}</math> is ''abelian{{anchor|abelian}}'' if the Lie bracket vanishes, i.e. [''x'',''y''] = 0, for all ''x'' and ''y'' in <math>\mathfrak{g}</math>. Abelian Lie algebras correspond to commutative (or [[abelian group|abelian]]) connected Lie groups such as vector spaces <math>\mathbb{K}^n</math> or [[torus|tori]] <math>\mathbb{T}^n</math>, and are all of the form <math>\mathfrak{k}^n,</math> meaning an ''n''-dimensional vector space with the trivial Lie bracket.
A Lie algebra <math>\mathfrak{g}</math> is ''abelian{{anchor|abelian}}'' if the Lie bracket vanishes; that is, [''x'',''y''] = 0 for all ''x'' and ''y'' in <math>\mathfrak{g}</math>. In particular, the Lie algebra of an abelian Lie group (such as the group <math>\mathbb{R}^n</math> under addition or the [[torus|torus group]] <math>\mathbb{T}^n</math>) is abelian. Every finite-dimensional abelian Lie algebra over a field <math>F</math> is isomorphic to <math>F^n</math> for some <math>n\geq 0</math>, meaning an ''n''-dimensional vector space with Lie bracket zero.


A more general class of Lie algebras is defined by the vanishing of all commutators of given length. A Lie algebra <math>\mathfrak{g}</math> is ''[[nilpotent Lie algebra|nilpotent]]'' if the [[lower central series]]
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 <math>\mathfrak{g}</math> is <math>[\mathfrak{g},\mathfrak{g}]</math>, meaning the linear subspace spanned by all brackets <math>[x,y]</math> with <math>x,y\in\mathfrak{g}</math>. The commutator subalgebra is an ideal in <math>\mathfrak{g}</math>, in fact the smallest ideal such that the quotient Lie algebra is abelian. It is analogous to the [[commutator subgroup]] of a group.

:<math> \mathfrak{g} > [\mathfrak{g},\mathfrak{g}] > [[\mathfrak{g},\mathfrak{g}],\mathfrak{g}] > [[[\mathfrak{g},\mathfrak{g}],\mathfrak{g}],\mathfrak{g}] > \cdots</math>

becomes zero eventually. By [[Engel's theorem]], a Lie algebra is nilpotent if and only if for every ''u'' in <math>\mathfrak{g}</math> the [[adjoint endomorphism]]


A Lie algebra <math>\mathfrak{g}</math> is ''[[nilpotent Lie algebra|nilpotent]]'' if the [[lower central series]]
:<math> \mathfrak{g} \supseteq [\mathfrak{g},\mathfrak{g}] \supseteq [[\mathfrak{g},\mathfrak{g}],\mathfrak{g}] \supseteq [[[\mathfrak{g},\mathfrak{g}],\mathfrak{g}],\mathfrak{g}] \supseteq \cdots</math>
becomes zero after finitely many steps. Equivalently, <math>\mathfrak{g}</math> is nilpotent if there is a finite sequence of ideals in <math>\mathfrak{g}</math>,
:<math>0=\mathfrak{a}_0 \subseteq \mathfrak{a}_1 \subseteq \cdots \subseteq \mathfrak{a}_r = \mathfrak{g},</math>
such that <math>\mathfrak{a}_j/\mathfrak{a}_{j-1}</math> is central in <math>\mathfrak{g}/\mathfrak{a}_{j-1}</math> for each ''j''. By [[Engel's theorem]], a Lie algebra over any field is nilpotent if and only if for every ''u'' in <math>\mathfrak{g}</math> the adjoint endomorphism
:<math>\operatorname{ad}(u):\mathfrak{g} \to \mathfrak{g}, \quad \operatorname{ad}(u)v=[u,v]</math>
:<math>\operatorname{ad}(u):\mathfrak{g} \to \mathfrak{g}, \quad \operatorname{ad}(u)v=[u,v]</math>
is [[nilpotent endomorphism|nilpotent]].<ref>{{harvnb|Jacobson|1979|loc=section II.3.}}</ref>


More generally, a Lie algebra <math>\mathfrak{g}</math> is said to be ''[[solvable Lie algebra|solvable]]'' if the [[derived series]]:
is [[nilpotent endomorphism|nilpotent]].
:<math> \mathfrak{g} \supseteq [\mathfrak{g},\mathfrak{g}] \supseteq [[\mathfrak{g},\mathfrak{g}],[\mathfrak{g},\mathfrak{g}]] \supseteq [[[\mathfrak{g},\mathfrak{g}],[\mathfrak{g},\mathfrak{g}]],[[\mathfrak{g},\mathfrak{g}],[\mathfrak{g},\mathfrak{g}]]] \supseteq \cdots</math>
becomes zero after finitely many steps. Equivalently, <math>\mathfrak{g}</math> is solvable if there is a finite sequence of Lie subalgebras,
:<math>0=\mathfrak{m}_0 \subseteq \mathfrak{m}_1 \subseteq \cdots \subseteq \mathfrak{m}_r = \mathfrak{g},</math>
such that <math>\mathfrak{m}_{j-1}</math> is an ideal in <math>\mathfrak{m}_{j}</math> with <math>\mathfrak{m}_{j}/\mathfrak{m}_{j-1}</math> abelian for each ''j''.<ref>{{harvnb|Jacobson|1979|loc=section I.7.}}</ref>


Every finite-dimensional Lie algebra over a field has a unique maximal solvable ideal, called its [[radical of a Lie algebra|radical]].<ref>{{harvnb|Jacobson|1979|p=24.}}</ref> Under the Lie correspondence, nilpotent (respectively, solvable) Lie groups correspond to nilpotent (respectively, solvable) Lie algebras over <math>\mathbb{R}</math>.
More generally still, a Lie algebra <math>\mathfrak{g}</math> is said to be ''[[solvable Lie algebra|solvable]]'' if the [[derived series]]:


For example, for a positive integer ''n'', the radical of <math>\mathfrak{gl}(n,F)</math> is its center, the 1-dimensional subspace spanned by the identity matrix. An example of a solvable Lie algebra is the space <math>\mathfrak{b}_{n}</math> of upper-triangular matrices in <math>\mathfrak{gl}(n)</math>; this is not nilpotent when <math>n\geq 2</math>. An example of a nilpotent Lie algebra is the space <math>\mathfrak{u}_{n}</math> of strictly upper-triangular matrices in <math>\mathfrak{gl}(n)</math>;
:<math> \mathfrak{g} > [\mathfrak{g},\mathfrak{g}] > [[\mathfrak{g},\mathfrak{g}],[\mathfrak{g},\mathfrak{g}]] > [[[\mathfrak{g},\mathfrak{g}],[\mathfrak{g},\mathfrak{g}]],[[\mathfrak{g},\mathfrak{g}],[\mathfrak{g},\mathfrak{g}]]] > \cdots</math>
this is not abelian when <math>n\geq 3</math>.


=== Simple and semisimple ===
becomes zero eventually.
{{main|Semisimple Lie algebra}}
A Lie algebra <math>\mathfrak{g}</math> is called ''[[Simple Lie algebra|simple]]'' if it is not abelian and the only ideals in <math>\mathfrak{g}</math> are 0 and <math>\mathfrak{g}</math>. (In particular, a one-dimensional—necessarily abelian—Lie algebra <math>\mathfrak{g}</math> is by definition not simple, even though its only ideals are 0 and <math>\mathfrak{g}</math>.) A finite-dimensional Lie algebra <math>\mathfrak{g}</math> is called ''[[semisimple Lie algebra|semisimple]]'' if the only solvable ideal in <math>\mathfrak{g}</math> is 0. In characteristic zero, a Lie algebra <math>\mathfrak{g}</math> is semisimple if and only if it is isomorphic to a product of simple Lie algebras, <math>\mathfrak{g} \cong \mathfrak{g}_1 \times \cdots \times \mathfrak{g}_r</math>.<ref>{{harvnb|Jacobson|1979|loc=Ch. III, § 5.}}</ref>


For example, the Lie algebra <math>\mathfrak{sl}(n,F)</math> is simple for every <math>n\geq 2</math> and every field ''F'' of characteristic zero (or just of characteristic not dividing ''n''). The Lie algebra <math>\mathfrak{su}(n)</math> over <math>\mathbb{R}</math> is simple for every <math>n\geq 2</math>. The Lie algebra <math>\mathfrak{so}(n)</math> over <math>\mathbb{R}</math> is simple if <math>n=3</math> or <math>n\geq 5</math>.<ref>{{harvnb|Erdmann|Wildon|2006|loc=Theorem 12.1.}}</ref> (There are "exceptional isomorphisms" <math>\mathfrak{so}(3)\cong\mathfrak{su}(2)</math> and <math>\mathfrak{so}(4)\cong\mathfrak{su}(2) \times \mathfrak{su}(2)</math>.)
The second term <math> [\mathfrak{g},\mathfrak{g}]</math> in the derived series (or, equivalently, the second term in the lower central series) is called the commutator subalgebra or commutator ideal of <math>\mathfrak{g}</math>.


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 representation|semisimple]] (that is, a direct sum of irreducible representations).<ref name="reducibility" />
Every finite-dimensional Lie algebra has a unique maximal solvable ideal, called its [[radical of a Lie algebra|radical]]. Under the Lie correspondence, nilpotent (respectively, solvable) connected Lie groups correspond to nilpotent (respectively, solvable) Lie algebras.


A finite-dimensional Lie algebra over a field of characteristic zero is called [[reductive Lie algebra|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.<ref>{{harvnb|Varadarajan|1984|loc=Theorem 3.16.3.}}</ref>
=== Simple and semisimple ===
{{main|Semisimple Lie algebra}}
A Lie algebra is "[[Simple Lie algebra|simple]]" if it has no non-trivial ideals and is not abelian. (This implies that a one-dimensional—necessarily abelian—Lie algebra is by definition not simple, even though it has no nontrivial ideals.) A Lie algebra <math>\mathfrak{g}</math> is called ''[[semisimple Lie algebra|semisimple]]'' if it is isomorphic to a direct sum of simple algebras. There are several equivalent characterizations of semisimple algebras, such as having no nonzero solvable ideals.


For example, <math>\mathfrak{gl}(n,F)</math> is reductive for ''F'' of characteristic zero: for <math>n\geq 2</math>, it is isomorphic to the product
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 (field)|characteristic]] zero, any finite-dimensional representation of a semisimple Lie algebra is [[semisimple representation|semisimple]] (i.e., direct sum of irreducible representations). In general, a Lie algebra is called [[reductive Lie algebra|reductive]] if the adjoint representation is semisimple. Thus, a semisimple Lie algebra is reductive.<!--
:<math>\mathfrak{gl}(n,F) \cong F\times \mathfrak{sl}(n,F),</math>
of a Lie algebra <math>\mathfrak{g}</math> over ''F'' is equivalent to the complete reducibility of all finite-dimensional [[Lie algebra representation|representations]] of <math>\mathfrak{g}.</math> An early proof of this statement proceeded via connection with compact groups ([[Weyl's unitary trick]]), but later entirely algebraic proofs were found.-->
where ''F'' denotes the center of <math>\mathfrak{gl}(n,F)</math>, the 1-dimensional subspace spanned by the identity matrix. Since the special linear Lie algebra <math>\mathfrak{sl}(n,F)</math> is simple, <math>\mathfrak{gl}(n,F)</math> contains few ideals: only 0, the center ''F'', <math>\mathfrak{sl}(n,F)</math>, and all of <math>\mathfrak{gl}(n,F)</math>.


=== Cartan's criterion ===
=== Cartan's criterion ===
[[Cartan's criterion]] gives conditions for a Lie algebra of characteristic zero to be nilpotent, solvable, or semisimple. It is based on the notion of the [[Killing form]], a [[symmetric bilinear form]] on <math>\mathfrak{g}</math> defined by the formula
[[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 <math>\mathfrak{g}</math> defined by
: <math>K(u,v)=\operatorname{tr}(\operatorname{ad}(u)\operatorname{ad}(v)),</math>
:<math>K(u,v)=\operatorname{tr}(\operatorname{ad}(u)\operatorname{ad}(v)),</math>
where tr denotes the [[Trace (linear algebra)|trace of a linear operator]]. A Lie algebra <math>\mathfrak{g}</math> is semisimple if and only if the Killing form is [[nondegenerate form|nondegenerate]]. A Lie algebra <math>\mathfrak{g}</math> is solvable if and only if <math>K(\mathfrak{g},[\mathfrak{g},\mathfrak{g}])=0.</math>
where tr denotes the trace of a linear operator. Namely: a Lie algebra <math>\mathfrak{g}</math> is semisimple if and only if the Killing form is [[nondegenerate form|nondegenerate]]. A Lie algebra <math>\mathfrak{g}</math> is solvable if and only if <math>K(\mathfrak{g},[\mathfrak{g},\mathfrak{g}])=0.</math><ref>{{harvnb|Varadarajan|1984|loc=section 3.9.}}</ref>


=== Classification ===
=== Classification ===
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. (Such a decomposition exists for a finite-dimensional Lie algebra over a field of characteristic zero.<ref>{{harvnb|Jacobson|1962|loc=Ch. III, § 9.}}</ref>) Furthermore, semisimple Lie algebras over an algebraically closed field of characteristic zero have been completely classified through their [[root system]]s.
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.<ref>{{harvnb|Jacobson|1979|loc=Ch. III, § 9.}}</ref> 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 system]]s. Namely, every simple Lie algebra is of type A<sub>''n''</sub>, B<sub>''n''</sub>, C<sub>''n''</sub>, D<sub>''n''</sub>, E<sub>6</sub>, E<sub>7</sub>, E<sub>8</sub>, F<sub>4</sub>, or G<sub>2</sub>.<ref>{{harvnb|Jacobson|1979|loc=section IV.6.}}</ref> Here the simple Lie algebra of type A<sub>''n''</sub> is <math>\mathfrak{sl}(n+1,F)</math>, B<sub>''n''</sub> is <math>\mathfrak{so}(2n+1,F)</math>, C<sub>''n''</sub> is <math>\mathfrak{sp}(2n,F)</math>, and D<sub>''n''</sub> is <math>\mathfrak{so}(2n,F)</math>. The other five are known as the [[exceptional Lie algebra]]s.

The classification of finite-dimensional simple Lie algebras over <math>\mathbb{R}</math> is more complicated, but it was also solved by Cartan (see [[simple Lie group]] for an equivalent classification). One can analyze a Lie algebra <math>\mathfrak{g}</math> over <math>\mathbb{R}</math> by considering its complexification <math>\mathfrak{g}\otimes_{\mathbb{R}}\mathbb{C}</math>.

In the years leading up to 2004, the finite-dimensional simple Lie algebras over an algebraically closed field of characteristic <math>p>3</math> were classified by [[Richard Earl Block]], Robert Lee Wilson, Alexander Premet, and Helmut Strade. (See [[restricted Lie algebra#Classification of simple Lie algebras]].) It turns out that there are many more simple Lie algebras in positive characteristic than in characteristic zero.


== Relation to Lie groups ==
== Relation to Lie groups ==
{{main|Lie group–Lie algebra correspondence}}
{{main|Lie group–Lie algebra correspondence}}
[[Image:Image Tangent-plane.svg|thumb| The tangent space of a [[sphere]] at a point <math>x</math>. If <math>x</math> is the identity element, then the tangent space is also a Lie algebra.]]
[[Image:Image Tangent-plane.svg|thumb| The tangent space of a [[sphere]] at a point <math>x</math>. If <math>x</math> were the identity element of a Lie group, the tangent space would be a Lie algebra.]]
Although Lie algebras are often studied in their own right, historically they arose as a means to study [[Lie group]]s.
Although Lie algebras can be studied in their own right, historically they arose as a means to study [[Lie group]]s.


We now briefly outline the relationship between Lie groups and Lie algebras. Any Lie group gives rise to a canonically determined Lie algebra over '''R''' (concretely, ''the tangent space at the identity''). Conversely, for any finite-dimensional Lie algebra <math>\mathfrak g</math>, there exists a corresponding connected Lie group <math>G</math> with Lie algebra <math>\mathfrak g</math>. This is [[Lie's third theorem]]; see the [[Baker–Campbell–Hausdorff formula]]. This Lie group is not determined uniquely; however, any two 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 <math>\mathbb{R}^3</math> with the cross-product, but SU(2) is a simply-connected twofold cover of SO(3).
The relationship between Lie groups and Lie algebras can be summarized as follows. Each Lie group determines a Lie algebra over <math>\mathbb{R}</math> (concretely, the tangent space at the identity). Conversely, for every finite-dimensional Lie algebra <math>\mathfrak g</math>, there is a connected Lie group <math>G</math> with Lie algebra <math>\mathfrak g</math>. This is [[Lie's third theorem]]; see the [[Baker–Campbell–Hausdorff formula]]. This Lie group is not determined uniquely; however, any two Lie groups with the same Lie algebra are ''locally isomorphic'', and more strongly, they 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 <math>\mathbb{R}^3</math> with the cross product, but SU(2) is a [[simply connected]] double cover of SO(3).


If we consider ''simply connected'' Lie groups, however, we have a one-to-one correspondence: For each (finite-dimensional real) Lie algebra <math>\mathfrak g</math>, there is a unique simply connected Lie group <math>G</math> with Lie algebra <math>\mathfrak g</math>.
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 <math>\mathbb{R}</math>.<ref>{{harvnb|Varadarajan|1984|loc=Theorems 2.7.5 and 3.15.1.}}</ref>


The correspondence between Lie algebras and Lie groups is used in several ways, including in the [[list of simple Lie groups|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 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.
The correspondence between Lie algebras and Lie groups is used in several ways, including in the [[list of simple Lie groups|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.


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 (group theory)|center]], once the classification of Lie algebras is known (solved by [[Élie Cartan|Cartan]] et al. in the [[Semisimple Lie algebra|semisimple]] case).
Every connected Lie group is isomorphic to its universal cover modulo a [[discrete group|discrete]] central subgroup.<ref>{{harvnb|Varadarjan|1984|loc=section 2.6.}}</ref> So classifying Lie groups becomes simply a matter of counting the discrete subgroups of the [[Center (group theory)|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.


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('''S'''<sup>1</sup>), 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.
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.<ref>{{harvnb|Milnor|2010|loc=Warnings 1.6 and 8.5.}}</ref>

Lie theory also does not work so neatly for infinite-dimensional representations of a finite-dimensional group. Even for the additive group <math>G=\mathbb{R}</math>, an infinite-dimensional representation of <math>G</math> can usually not be differentiated to produce a representation of its Lie algebra on the same space, or vice versa.<ref>{{harvnb|Knapp|2001|loc=section III.3, Problem III.5.}}</ref> The theory of [[Harish-Chandra module]]s is a more subtle relation between infinite-dimensional representations for groups and Lie algebras.


== Real form and complexification ==
== Real form and complexification ==
Given a [[complex Lie algebra]] <math>\mathfrak g</math>, a real Lie algebra <math>\mathfrak{g}_0</math> is said to be a ''[[real form]]'' of <math>\mathfrak g</math> if the [[complexification]] <math>\mathfrak{g}_0 \otimes_{\mathbb R} \mathbb{C} \simeq \mathfrak{g}</math> is isomorphic to <math>\mathfrak{g}</math>.<ref name="Fulton 26">{{harvnb|Fulton|Harris|1991|loc=§26.1.}}</ref> A real form need not be unique; for example, <math>\mathfrak{sl}_2 \mathbb{C}</math> has two real forms, <math>\mathfrak{sl}_2 \mathbb{R}</math> and <math>\mathfrak{su}_2</math>.<ref name="Fulton 26" />
Given a [[complex Lie algebra]] <math>\mathfrak g</math>, a real Lie algebra <math>\mathfrak{g}_0</math> is said to be a ''[[real form]]'' of <math>\mathfrak g</math> if the complexification <math>\mathfrak{g}_0 \otimes_{\mathbb R} \mathbb{C}</math> is isomorphic to <math>\mathfrak{g}</math>. A real form need not be unique; for example, <math>\mathfrak{sl}(2,\mathbb{C})</math> has two real forms up to isomorphism, <math>\mathfrak{sl}(2,\mathbb{R})</math> and <math>\mathfrak{su}(2)</math>.<ref name="Fulton 26">{{harvnb|Fulton|Harris|1991|loc=§26.1.}}</ref>


Given a semisimple finite-dimensional complex Lie algebra <math>\mathfrak g</math>, a ''[[split form]]'' of it is a real form that splits; i.e., it has a Cartan subalgebra which acts via an adjoint representation with real eigenvalues. A split form exists and is unique (up to isomorphisms).<ref name="Fulton 26" /> A ''[[compact form]]'' is a real form that is the Lie algebra of a compact Lie group. A compact form exists and is also unique.<ref name="Fulton 26" />
Given a semisimple complex Lie algebra <math>\mathfrak g</math>, a ''[[split form]]'' of it is a real form that splits; i.e., it has a Cartan subalgebra which acts via an adjoint representation with real eigenvalues. A split form exists and is unique (up to isomorphism). A ''[[compact form]]'' is a real form that is the Lie algebra of a compact Lie group. A compact form exists and is also unique up to isomorphism.<ref name="Fulton 26" />


== Lie algebra with additional structures ==
== Lie algebra with additional structures ==
A Lie algebra can be equipped with some additional structures that are assumed to be compatible with the bracket. For example, a [[graded Lie algebra]] is a Lie algebra with a graded vector space structure. If it also comes with a differential (so that the underlying graded vector space is a [[chain complex]]), then it is called a [[differential graded Lie algebra]].
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 group]]s 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 number]]s <math>\mathbb{Q}</math> to describe [[rational homotopy theory]] in algebraic terms.<ref>{{harvnb|Quillen|1969|loc=Corollary II.6.2.}}</ref>
A [[simplicial Lie algebra]] is a [[simplicial object]] in the category of Lie algebras; in other words, it is obtained by replacing the underlying set with a [[simplicial set]] (so it might be better thought of as a family of Lie algebras).


== Lie ring ==
== Lie ring ==
The definition of a Lie algebra over a field extends to define a Lie algebra over any [[commutative ring]] ''R''. Namely, a Lie algebra <math>\mathfrak{g}</math> over ''R'' is an ''R''-[[module (mathematics)|module]] with an alternating ''R''-bilinear map <math>[\ , \ ]\colon \mathfrak{g} \times \mathfrak{g} \to \mathfrak{g}</math> that satisfies the Jacobi identity. A Lie algebra over the ring <math>\mathbb{Z}</math> of [[integer]]s is sometimes called a '''Lie ring'''. (This is not directly related to the notion of a Lie group.)
A ''Lie ring'' arises as a generalisation of Lie algebras, or through the study of the [[lower central series]] of [[Group (mathematics)|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 <math>L</math> to be an [[abelian group]] with an operation <math>[\cdot,\cdot]</math> that has the following properties:


Lie rings are used in the study of finite [[p-group]]s (for a prime number ''p'') through the ''Lazard correspondence''.<ref>{{harvnb|Khukhro|1998|loc=Ch. 6.}}</ref> 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.
* Bilinearity:


[[p-adic analytic group|p-adic Lie groups]] are related to Lie algebras over the field <math>\mathbb{Q}_p</math> of [[p-adic number]]s as well as over the ring <math>\mathbb{Z}_p</math> of [[p-adic integer]]s.<ref>{{harvnb|Serre|2006|loc=Part II, section V.1.}}</ref> Part of [[Claude Chevalley]]'s construction of the finite [[groups of Lie type]] involves showing that a simple Lie algebra over the complex numbers comes from a Lie algebra over the integers, and then (with more care) a [[group scheme]] over the integers.<ref>{{harvnb|Humphreys|1978|loc=section 25.}}</ref>
::<math> [x + y, z] = [x, z] + [y, z], \quad [z, x + y] = [z, x] + [z, y] </math>

:for all ''x'', ''y'', ''z'' &isin; ''L''.

* The ''Jacobi identity'':

:: <math> [x,[y,z]] + [y,[z,x]] + [z,[x,y]] = 0 \quad </math>

:for all ''x'', ''y'', ''z'' in ''L''.

* For all ''x'' in ''L'':

::<math> [x,x]=0.</math>

Lie rings need not be [[Lie group]]s 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 <math>[x,y] = xy - yx</math>. 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-group]]s through the ''Lazard correspondence''. The lower central factors of a ''p''-group are finite abelian ''p''-groups, so modules over '''Z'''/''p'''''Z'''. 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 ''p''th 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 group]]s 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 then reducing modulo ''p'' to get a Lie algebra over a finite field.


=== Examples ===
=== Examples ===
* Here is a construction of Lie rings arising from the study of abstract groups. For elements <math>x,y</math> of a group, define the commutator <math>[x,y]= x^{-1}y^{-1}xy</math>. Let <math>G = G_0 \supseteq G_1 \supseteq G_2 \supseteq \cdots \supseteq G_n \supseteq \cdots</math> be a ''filtration'' of a group <math>G</math>, that is, a chain of subgroups such that <math>[G_i,G_j]</math> is contained in <math>G_{i+j}</math> for all <math>i,j</math>. (For the Lazard correspondence, one takes the filtration to be the lower central series of ''G''.) Then
* Any Lie algebra over a general [[Ring (mathematics)|ring]] instead of a [[Field (mathematics)|field]] is an example of a Lie ring. Lie rings are ''not'' [[Lie group]]s under addition, despite the name.
:: <math>L = \bigoplus_{i\geq 0} G_i/G_{i+1}</math>
* Any associative ring can be made into a Lie ring by defining a bracket operator
:is a Lie ring, with addition given by the group multiplication (which is abelian on each quotient group <math>G_i/G_{i+1}</math>), and with Lie bracket <math>G_i/G_{i+1} \times G_j/G_{j+1} \to G_{i+j}/G_{i+j+1}</math> given by commutators in the group:<ref>{{harvnb|Serre|2006|loc=Part I, Chapter II.}}</ref>
:: <math>[x,y] = xy - yx.</math>
:: <math>[xG_{i+1}, yG_{j+1}] := [x,y]G_{i+j+1}. </math>
* For an example of a Lie ring arising from the study of [[Group (mathematics)|groups]], let <math>G</math> be a group with <math>[x,y]= x^{-1}y^{-1}xy</math> the commutator operation, and let <math>G = G_0 \supseteq G_1 \supseteq G_2 \supseteq \cdots \supseteq G_n \supseteq \cdots</math> be a [[central series]] in <math>G</math> &mdash; that is the commutator subgroup <math>[G_i,G_j]</math> is contained in <math>G_{i+j}</math> for any <math>i,j</math>. Then


: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 <math>\mathbb{Z}/2\mathbb{Z}</math>.
:: <math>L = \bigoplus G_i/G_{i+1}</math>


==Definition using category-theoretic notation==
:is a Lie ring with addition supplied by the group operation (which is abelian in each homogeneous part), and the bracket operation given by
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, [[morphism]]s 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 [[tensor product#Tensor powers and braiding|braiding isomorphisms]] are needed. If {{mvar|A}} is a vector space, the ''interchange isomorphism'' <math>\tau: A\otimes A \to A\otimes A</math> is defined by
:: <math>[xG_i, yG_j] = [x,y]G_{i+j}\ </math>
:<math>\tau(x\otimes y)= y\otimes x.</math>
The ''cyclic-permutation braiding'' <math>\sigma:A\otimes A\otimes A \to A\otimes A\otimes A </math> is defined as
:<math>\sigma=(\mathrm{id}\otimes \tau)\circ(\tau\otimes \mathrm{id}),</math>
where <math>\mathrm{id}</math> is the identity morphism. Equivalently, <math>\sigma</math> is defined by
:<math>\sigma(x\otimes y\otimes z)= y\otimes z\otimes x.</math>


With this notation, a Lie algebra can be defined as an object <math>A</math> in the category of vector spaces together with a morphism
:extended linearly. The centrality of the series ensures that the commutator <math>[x,y]</math> gives the bracket operation the appropriate Lie theoretic properties.
:<math>[\cdot,\cdot]\colon A\otimes A\rightarrow A</math>
that satisfies the two morphism equalities
:<math>[\cdot,\cdot]\circ(\mathrm{id}+\tau)=0,</math>
and
:<math>[\cdot,\cdot]\circ ([\cdot,\cdot]\otimes \mathrm{id}) \circ (\mathrm{id}+\sigma+\sigma^2)=0.</math>


==See also==
==See also==
{{div col|colwidth=28em}}
{{div col|colwidth=28em}}
* [[Adjoint representation of a Lie algebra]]
* [[Affine Lie algebra]]
* [[Affine Lie algebra]]
* [[Anyonic Lie algebra]]
* [[Automorphism of a Lie algebra]]
* [[Automorphism of a Lie algebra]]
* [[Chiral Lie algebra]]
* [[Gelfand–Fuks cohomology]]
* [[Free Lie algebra]]<!-- Need to be discussed somewhere in this article but not clear where, probably not too early. -->
* [[Index of a Lie algebra]]
* [[Index of a Lie algebra]]
* [[Lie algebra cohomology]]
* [[Lie algebra cohomology]]
Line 407: Line 399:
* [[Particle physics and representation theory]]
* [[Particle physics and representation theory]]
* [[Lie superalgebra]]
* [[Lie superalgebra]]
* [[Orthogonal symmetric Lie algebra]]
* [[Poisson algebra]]
* [[Poisson algebra]]
* [[Pre-Lie algebra]]
* [[Pre-Lie algebra]]
Line 415: Line 408:
* [[Restricted Lie algebra]]
* [[Restricted Lie algebra]]
* [[Serre relations]]
* [[Serre relations]]
* [[Symmetric Lie algebra]] <!-- missing article -->
* [[Gelfand–Fuks cohomology]]
{{div col end}}
{{div col end}}


Line 426: Line 417:


== Sources ==
== Sources ==
* {{cite book |last=Beltiţă |first= Daniel |title=Smooth Homogeneous Structures in Operator Theory |series=CRC Monographs and Surveys in Pure and Applied Mathematics |volume=137 |publisher=CRC Press |year= 2006 |isbn=978-1-4200-3480-6 |mr=2188389 |url=https://books.google.com/books?id=x8NQc-pWbLQC&q=%22Lie+algebra%22}}
* {{cite book |last=Bourbaki |first=Nicolas |author-link=Nicolas Bourbaki |title=Lie Groups and Lie Algebras: Chapters 1-3 |year=1989 |isbn=978-3-540-64242-8 |publisher=Springer |url=https://books.google.com/books?id=brSYF_rB2ZcC|mr=1728312}}
* {{Cite book |last1=Erdmann |first1=Karin |author-link=Karin Erdmann |last2=Wildon |first2=Mark |title=Introduction to Lie Algebras |publisher=Springer |year=2006 |isbn=1-84628-040-0|mr=2218355}}
* {{Cite journal |last1=Boza |first1=Luis |last2=Fedriani |first2=Eugenio M. |last3=Núñez |first3=Juan |date=2001-06-01 |title=A new method for classifying complex filiform Lie algebras |journal=Applied Mathematics and Computation |volume=121 |issue=2–3 |pages=169–175 |doi=10.1016/s0096-3003(99)00270-2 |issn=0096-3003}}
* {{cite book |last=Bourbaki |first=Nicolas |author-link=Nicolas Bourbaki |title=Lie Groups and Lie Algebras: Chapters 1-3 |year=1989 |isbn=978-3-540-64242-8 |publisher=Springer |url=https://books.google.com/books?id=brSYF_rB2ZcC}}
* {{Cite book |last1=Erdmann |first1=Karin |author-link=Karin Erdmann |last2=Wildon |first2=Mark |title=Introduction to Lie Algebras |publisher=Springer |year=2006 |isbn=1-84628-040-0}}
* {{Fulton-Harris}}
* {{Fulton-Harris}}
* {{cite book |last=Hall |first=Brian C. |title=Lie groups, Lie algebras, and Representations: An Elementary Introduction |edition=2nd |series=Graduate Texts in Mathematics |volume=222 |publisher=Springer |year=2015 |isbn=978-3319134666 |doi=10.1007/978-3-319-13467-3 |issn=0072-5285}}
* {{cite book |last=Hall |first=Brian C. |title=Lie groups, Lie Algebras, and Representations: An Elementary Introduction |edition=2nd |series=Graduate Texts in Mathematics |volume=222 |publisher=Springer |year=2015 |isbn=978-3319134666 |doi=10.1007/978-3-319-13467-3 |issn=0072-5285|mr=3331229}}
* {{cite book |last=Humphreys |first=James E. |author-link=James E. Humphreys |title=Introduction to Lie Algebras and Representation Theory |edition=2nd |series=Graduate Texts in Mathematics |volume=9 |publisher=Springer-Verlag |year=1978 |isbn=978-0-387-90053-7 |url-access=registration |url=https://archive.org/details/introductiontoli00jame|mr=0499562}}
* {{cite book |last1=Hofmann |first1=Karl H. |last2=Morris |first2=Sidney A |title=The Lie Theory of Connected Pro-Lie Groups |publisher=European Mathematical Society |year=2007 |isbn=978-3-03719-032-6}}
* {{cite book |last=Humphreys |first=James E. |author-link=James E. Humphreys |title=Introduction to Lie Algebras and Representation Theory |edition=2nd |series=Graduate Texts in Mathematics |volume=9 |publisher=Springer-Verlag |year=1978 |isbn=978-0-387-90053-7 |url-access=registration |url=https://archive.org/details/introductiontoli00jame}}
* {{cite book |last=Jacobson |first=Nathan |author-link=Nathan Jacobson |title=Lie Algebras |orig-year=1962 |publisher=Dover |year=1979 |isbn=978-0-486-63832-4|mr=0559927}}
* {{cite book |last=Jacobson |first=Nathan |author-link=Nathan Jacobson |title=Lie algebras |orig-year=1962 |publisher=Dover |year=1979 |isbn=978-0-486-63832-4 |ref={{harvid |Jacobson |1962}}}}
* {{citation |mr=1615819 | last1=Khukhro | first1=E. I. | title=p-Automorphisms of Finite p-Groups | publisher=[[Cambridge University Press]] | year=1998 | isbn=0-521-59717-X | doi=10.1017/CBO9780511526008}}
* {{citation | last=Knapp | first=Anthony W. | author-link=Anthony W. Knapp|title=Representation Theory of Semisimple Groups: an Overview Based on Examples | publisher=[[Princeton University Press]] | year=2001 | origyear=1986 |isbn=0-691-09089-0 | mr=1880691}}
* {{cite book |last1=Kac |first1=Victor G. |author-link1=Victor Kac |display-authors=etal |title=Course notes for MIT 18.745: Introduction to Lie Algebras |url=http://math.mit.edu/~lesha/745lec/ |url-status=bot: unknown |archive-url=https://web.archive.org/web/20100420004313/http://math.mit.edu/~lesha/745lec/ |archive-date=2010-04-20}}
* {{citation|last=Milnor|first=John|author-link=John Milnor|chapter=Remarks on infinite-dimensional Lie groups|title=Collected Papers of John Milnor|volume=5|year=2010|origyear=1986|pages=91–141|mr=0830252|isbn=978-0-8218-4876-0}}
* {{cite journal |last1=Mubarakzyanov |first1=G.M. |title=On solvable Lie algebras |journal=Izv. Vys. Ucheb. Zaved. Matematika |volume=1 |issue=32 |year=1963 |pages=114–123 |mr=153714 |zbl=0166.04104 |url=http://mi.mathnet.ru/eng/ivm2141 |language=ru}}
* {{citation|author1-first=Daniel|author1-last=Quillen|author1-link=Daniel Quillen|title=Rational homotopy theory|journal=[[Annals of Mathematics]]|volume=90|year=1969|pages=205–295|doi=10.2307/1970725|issue=2|mr=0258031|jstor=1970725}}
* {{cite web |last1=O'Connor |first1=J.J |author-link1=John J. O'Connor (mathematician) |last2=Robertson |first2=E.F. |author-link2=Edmund F. Robertson |title=Biography of Sophus Lie |year=2000 |publisher=MacTutor History of Mathematics Archive |url=http://www-history.mcs.st-andrews.ac.uk/Mathematicians/Lie.html}}
* {{cite book |last=Serre |first= Jean-Pierre |author-link=Jean-Pierre Serre |title=Lie Algebras and Lie Groups |edition=2nd |publisher=Springer |year=2006 |isbn=978-3-540-55008-2|mr=2179691}}
* {{cite web |last1=O'Connor |first1=J.J |last2=Robertson |first2=E.F. |title=Biography of Wilhelm Killing |year=2005 |publisher=MacTutor History of Mathematics Archive |url=http://www-history.mcs.st-andrews.ac.uk/Mathematicians/Killing.html}}
* {{cite book |last=Varadarajan |first=Veeravalli S. |author-link=Veeravalli S. Varadarajan |title=Lie Groups, Lie Algebras, and Their Representations |publisher=Springer |year=1984 | origyear=1974 |isbn=978-0-387-90969-1|mr=0746308}}
* {{cite journal |last1=Popovych |first1=R.O. |last2=Boyko |first2=V.M. |last3=Nesterenko |first3=M.O. |last4=Lutfullin |first4=M.W. |display-authors=etal |title=Realizations of real low-dimensional Lie algebras |journal=J. Phys. A: Math. Gen. |volume=36 |issue=26 |year=2003 |pages=7337–60 |doi=10.1088/0305-4470/36/26/309 |arxiv=math-ph/0301029 |bibcode=2003JPhA...36.7337P |s2cid=9800361}}
* {{cite book |last=Serre |first= Jean-Pierre |author-link=Jean-Pierre Serre |title=Lie Algebras and Lie Groups |edition=2nd |publisher=Springer |year=2006 |isbn=978-3-540-55008-2}}
* {{cite book |last=Wigner |first=Eugene |author-link=Eugene Wigner |title=Group Theory and its Application to the Quantum Mechanics of Atomic Spectra |translator=J. J. Griffin | publisher=[[Academic Press]] |year=1959 |isbn=978-0127505503|mr=0106711}}
* {{cite book |last=Steeb |first= Willi-Hans |title=Continuous Symmetries, Lie Algebras, Differential Equations and Computer Algebra |edition=2nd |publisher= World Scientific |year= 2007 |isbn=978-981-270-809-0 |mr=2382250 |doi=10.1142/6515}}
* {{cite book |last=Varadarajan |first=Veeravalli S. |author-link=Veeravalli S. Varadarajan |title=Lie Groups, Lie Algebras, and Their Representations |edition=1st |publisher=Springer |year=2004 |isbn=978-0-387-90969-1}}


==External links==
==External links==
* {{cite book |last1=Kac |first1=Victor G. |author-link1=Victor Kac |display-authors=etal |title=Course notes for MIT 18.745: Introduction to Lie Algebras |url=http://math.mit.edu/~lesha/745lec/ |url-status=dead |archive-url=https://web.archive.org/web/20100420004313/http://math.mit.edu/~lesha/745lec/ |archive-date=2010-04-20}}
* {{springer|title=Lie algebra|id=p/l058370}}
* {{springer|title=Lie algebra|id=p/l058370}}
* {{cite web |last=McKenzie |first=Douglas |year=2015 |url=http://www.liealgebrasintro.com |title=An Elementary Introduction to Lie Algebras for Physicists}}
* {{cite web |last=McKenzie |first=Douglas |year=2015 |url=http://www.liealgebrasintro.com |title=An Elementary Introduction to Lie Algebras for Physicists}}
* {{cite web |last1=O'Connor |first1=J.J |author-link1=John J. O'Connor (mathematician) |last2=Robertson |first2=E.F. |author-link2=Edmund F. Robertson |title=Marius Sophus Lie |year=2000 |publisher=MacTutor History of Mathematics Archive |url=https://mathshistory.st-andrews.ac.uk/Biographies/Lie/}}
* {{cite web |last1=O'Connor |first1=J.J |author-link1=John J. O'Connor (mathematician) |last2=Robertson |first2=E.F. | author-link2=Edmund F. Robertson |title=Wilhelm Karl Joseph Killing |year=2005 |publisher=MacTutor History of Mathematics Archive |url=https://mathshistory.st-andrews.ac.uk/Biographies/Killing/}}


{{Authority control}}
{{Authority control}}

Revision as of 22:36, 8 February 2024

In mathematics, a Lie algebra (pronounced /l/ LEE) is a vector space together with an operation called the Lie bracket, an alternating bilinear map , that satisfies the Jacobi identity. In other words, a Lie algebra is an algebra over a field for which the multiplication operation (called the Lie bracket) is alternating and satisfies the Jacobi identity. The Lie bracket of two vectors and is denoted . A Lie algebra is typically a non-associative algebra. However, every associative algebra gives rise to a Lie algebra, consisting of the same vector space with the commutator Lie bracket, .

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 to G at the identity. To second order, the group operation may be non-commutative, and the second-order terms describing the non-commutativity of G near the identity give the structure of a Lie algebra. It is a remarkable fact that these second-order terms (the Lie algebra) completely determine the group structure of G near the identity. They even determine G globally, up to covering spaces.

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 with Lie bracket defined by the cross product This is skew-symmetric since , and instead of associativity it satisfies the Jacobi identity:

This is the Lie algebra of the Lie group of rotations of space, and each vector may be pictured as an infinitesimal rotation around the axis , with angular speed equal to the magnitude of . The Lie bracket is a measure of the non-commutativity between two rotations. Since a rotation commutes with itself, one has the alternating property .

History

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.

Definition of a Lie algebra

A Lie algebra is a vector space over a field together with a binary operation called the Lie bracket, satisfying the following axioms:[a]

  • Bilinearity,
for all scalars in and all elements in .
  • The Alternating property,
for all in .
  • The Jacobi identity,
for all in .

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 and using the alternating property shows that for all in . Thus bilinearity and the alternating property together imply

for all in . If the field does not have characteristic 2, then anticommutativity implies the alternating property, since it implies [3]

It is customary to denote a Lie algebra by a lower-case fraktur letter such as . If a Lie algebra is associated with a Lie group, then the algebra is denoted by the fraktur version of the group's name: for example, the Lie algebra of SU(n) is .

Generators and dimension

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 means a subset of such that any Lie subalgebra (as defined below) that contains S must be all of . Equivalently, is spanned (as a vector space) by all iterated brackets of elements of S.

Basic examples

Abelian Lie algebras

Any vector space endowed with the identically zero Lie bracket becomes a Lie algebra. Such a Lie algebra is called abelian. Every one-dimensional Lie algebra is abelian, by the alternating property of the Lie bracket.

The Lie algebra of matrices

  • On an associative algebra over a field with multiplication written as , a Lie bracket may be defined by the commutator . With this bracket, is a Lie algebra. (The Jacobi identity follows from the associativity of the multiplication on .) [4]
  • The endomorphism ring of an -vector space with the above Lie bracket is denoted .
  • For a field F and a positive integer n, the space of n × n matrices over F, denoted or , is a Lie algebra with bracket given by the commutator of matrices: .[5] This is a special case of the previous example; it is probably the most important example of a Lie algebra. It is called the general linear Lie algebra.
When F is the real numbers, is the Lie algebra of the general linear group , the group of invertible n x n real matrices (or equivalently, matrices with nonzero determinant), where the group operation is matrix multiplication. Likewise, is the Lie algebra of the complex Lie group . The Lie bracket on describes the failure of commutativity for matrix multiplication, or equivalently for the composition of linear maps. For any field F, can be viewed as the Lie algebra of the algebraic group .

Definitions

Subalgebras, ideals and homomorphisms

The Lie bracket is not required to be associative, meaning that need not be equal to . Nonetheless, much of the terminology for associative rings and algebras (and also for groups) has analogs for Lie algebras. A Lie subalgebra is a linear subspace which is closed under the Lie bracket. An ideal is a linear subspace that satisfies the stronger condition:[6]

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:

An isomorphism of Lie algebras is a bijective homomorphism.

As with normal subgroups in groups, ideals in Lie algebras are precisely the kernels of homomorphisms. Given a Lie algebra and an ideal in it, the quotient Lie algebra is defined, with a surjective homomorphism of Lie algebras. The first isomorphism theorem holds for Lie algebras.

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 are said to commute if their bracket vanishes: .

The centralizer subalgebra of a subset is the set of elements commuting with : that is, . The centralizer of itself is the center . Similarly, for a subspace S, the normalizer subalgebra of is .[7] If is a Lie subalgebra, is the largest subalgebra such that is an ideal of .

Example

The subspace of diagonal matrices in is an abelian Lie subalgebra. (It is a Cartan subalgebra of , analogous to a maximal torus in the theory of compact Lie groups.) Here is not an ideal in for . For example, when , this follows from the calculation:

(which is not always in ).

Every one-dimensional linear subspace of a Lie algebra is an abelian Lie subalgebra, but it need not be an ideal.

Product and semidirect product

For two Lie algebras and , the product Lie algebra is the vector space consisting of all ordered pairs , with Lie bracket[8]

This is the product in the category of Lie algebras. Note that the copies of and in commute with each other:

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

Derivations

For an algebra A over a field F, a derivation of A over F is a linear map that satisfies the Leibniz rule

for all . (The definition makes sense for a possibly non-associative algebra.) Given two derivations and , their commutator is again a derivation. This operation makes the space of all derivations of A over F into a Lie algebra.[9]

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

(where 1 denotes the identity map on A) gives exactly the definition of D being a derivation.

Example: the Lie algebra of vector fields. Let A be the ring of smooth functions on a smooth manifold X. Then a derivation of A over is equivalent to a vector field on X. (A vector field v gives a derivation of the space of smooth functions by differentiating functions in the direction of v.) This makes the space of vector fields into a Lie algebra (see Lie bracket of vector fields).[10] Informally speaking, is the Lie algebra of the diffeomorphism group of X. So the Lie bracket of vector fields describes the non-commutativity of the diffeomorphism group. An action of a Lie group G on a manifold X determines a homomorphism of Lie algebras .

A Lie algebra can be viewed as a non-associative algebra, and so each Lie algebra over a field F determines its Lie algebra of derivations, . That is, a derivation of is a linear map such that

.

The inner derivation associated to any is the adjoint mapping defined by . (This is a derivation as a consequence of the Jacobi identity.) That gives a homomorphism of Lie algebras, . The image is an ideal in , and the Lie algebra of outer derivations is defined as the quotient Lie algebra, . (This is exactly analogous to the outer automorphism group of a group.) For a semisimple Lie algebra (defined below) over a field of characteristic zero, every derivation is inner.[11] This is related to the theorem that the outer automorphism group of a semisimple Lie group is finite.[12]

In contrast, an abelian Lie algebra has many outer derivations. Namely, for a vector space with Lie bracket zero, the Lie algebra can be identified with .

Examples

Matrix Lie algebras

A matrix group is a Lie group consisting of invertible matrices, , where the group operation of G is matrix multiplication. The corresponding Lie algebra is the space of matrices which are tangent vectors to G inside the linear space : this consists of derivatives of smooth curves in G at the identity matrix :

The Lie bracket of is given by the commutator of matrices, . Given a Lie algebra , one can recover the Lie group as the subgroup generated by the matrix exponential of elements of .[13] (To be precise, this gives the identity component of G, if G is not connected.) Here the exponential mapping is defined by , which converges for every matrix .

The same comments apply to complex Lie subgroups of and the complex matrix exponential, (defined by the same formula).

Here are some matrix Lie groups and their Lie algebras.[14]

  • For a positive integer n, the special linear group consists of all real n × n matrices with determinant 1. This is the group of linear maps from to itself that preserve volume and orientation. More abstractly, is the commutator subgroup of the general linear group . Its Lie algebra consists of all real n × n matrices with trace 0. Similarly, one can define the analogous complex Lie group and its Lie algebra .
  • The orthogonal group plays a basic role in geometry: it is the group of linear maps from to itself that preserve the length of vectors. For example, rotations and reflections belong to . Equivalently, this is the group of n x n orthogonal matrices, meaning that . The orthogonal group has two connected components; the identity component is called the special orthogonal group , consisting of the orthogonal matrices with determinant 1. Both groups have the same Lie algebra , the subspace of skew-symmetric matrices in (). See also infinitesimal rotations with skew-symmetric matrices.
The complex orthogonal group , its identity component , and the Lie algebra are given by the same formulas applied to n x n complex matrices. Equivalently, is the subgroup of that preserves the standard symmetric bilinear form on .
  • The unitary group is the subgroup of that preserves the length of vectors in (with respect to the standard Hermitian inner product). Equivalently, this is the group of n × n unitary matrices (satisfying , where denotes the conjugate transpose of a matrix). Its Lie algebra consists of the skew-hermitian matrices in (). This is a Lie algebra over , not over . (Indeed, i times a skew-hermitian matrix is hermitian, rather than skew-hermitian.) Likewise, the unitary group is a real Lie subgroup of the complex Lie group . For example, is the circle group, and its Lie algebra (from this point of view) is .
  • The special unitary group is the subgroup of matrices with determinant 1 in . Its Lie algebra consists of the skew-hermitian matrices with trace zero.
  • The symplectic group is the subgroup of that preserves the standard alternating bilinear form on . Its Lie algebra is the symplectic Lie algebra .
  • The classical Lie algebras are those listed above, along with variants over any field.

Two dimensions

Some Lie algebras of low dimension are described here. See the classification of low-dimensional real Lie algebras for further examples.

  • There is a unique nonabelian Lie algebra of dimension 2 over any field F, up to isomorphism.[15] Here has a basis for which the bracket is given by . (This determines the Lie bracket completely, because the axioms imply that and .) Over the real numbers, can be viewed as the Lie algebra of the Lie group of affine transformations of the real line, .
The affine group G can be identified with the group of matrices
under matrix multiplication, with , . Its Lie algebra is the Lie subalgebra of consisting of all matrices
In these terms, the basis above for is given by the matrices
For any field , the 1-dimensional subspace is an ideal in the 2-dimensional Lie algebra , by the formula . Both of the Lie algebras and are abelian (because 1-dimensional). In this sense, can be broken into abelian "pieces", meaning that it is solvable (though not nilpotent), in the terminology below.

Three dimensions

  • The Heisenberg algebra over a field F is the three-dimensional Lie algebra with a basis such that[16]
.
It can be viewed as the Lie algebra of 3×3 strictly upper-triangular matrices, with the commutator Lie bracket and the basis
Over the real numbers, is the Lie algebra of the Heisenberg group , that is, the group of matrices
under matrix multiplication.
For any field F, the center of is the 1-dimensional ideal , and the quotient is abelian, isomorphic to . In the terminology below, it follows that is nilpotent (though not abelian).
  • The Lie algebra of the rotation group is the space of skew-symmetric 3 x 3 matrices over . A basis is given by the three matrices[17]
The commutation relations among these generators are
The cross product of vectors in is given by the same formula in terms of the standard basis; so that Lie algebra is isomorphic to . Also, is equivalent to the Spin (physics) angular-momentum component operators for spin-1 particles in quantum mechanics.[18]
The Lie algebra cannot be broken into pieces in the way that the previous examples can: it is simple, meaning that it is not abelian and its only ideals are 0 and all of .
  • Another simple Lie algebra of dimension 3, in this case over , is the space of 2 x 2 matrices of trace zero. A basis is given by the three matrices
The Lie bracket is given by:
Using these formulas, one can show that the Lie algebra is simple, and classify its finite-dimensional representations (defined below).[19] In the terminology of quantum mechanics, one can think of E and F as raising and lowering operators. Indeed, for any representation of , the relations above imply that E maps the c-eigenspace of H (for a complex number c) into the -eigenspace, while F maps the c-eigenspace into the -eigenspace.
The Lie algebra is isomorphic to the complexification of , meaning the tensor product . The formulas for the Lie bracket are easier to analyze in the case of . As a result, it is common to analyze complex representations of the group by relating them to representations of the Lie algebra .

Infinite dimensions

  • The Lie algebra of vector fields on a smooth manifold of positive dimension is an infinite-dimensional Lie algebra over .
  • The Kac–Moody algebras are a large class of infinite-dimensional Lie algebras, say over , with structure much like that of the finite-dimensional simple Lie algebras (such as ).
  • The Moyal algebra is an infinite-dimensional Lie algebra that contains all the classical Lie algebras as subalgebras.
  • The Virasoro algebra is important in string theory.
  • The functor that takes a Lie algebra over a field F to the underlying vector space has a left adjoint , called the free Lie algebra on a vector space V. It is spanned by all iterated Lie brackets of elements of V, modulo only the relations coming from the definition of a Lie algebra. The free Lie algebra is infinite-dimensional for V of dimension at least 2.[20]

Representations

Definitions

Given a vector space V, let denote the Lie algebra consisting of all linear maps from V to itself, with bracket given by . A representation of a Lie algebra on V is a Lie algebra homomorphism

That is, sends each element of to a linear map from V to itself, in such a way that the Lie bracket on corresponds to the commutator of linear maps.

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.[21] Equivalently, every finite-dimensional Lie algebra over a field F is isomorphic to a Lie subalgebra of for some positive integer n.

Adjoint representation

For any Lie algebra , the adjoint representation is the representation

given by .

Goals of representation theory

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 . Indeed, in the semisimple case, the adjoint representation is already faithful. Rather, the goal is to understand all possible representations of . For a semisimple Lie algebra over a field of characteristic zero, Weyl's theorem[22] says that every finite-dimensional representation is a direct sum of irreducible representations (those with no nontrivial invariant subspaces). The finite-dimensional irreducible representations are well understood from several points of view; see the representation theory of semisimple Lie algebras and the Weyl character formula.

Universal enveloping algebra

The functor that takes an associative algebra A over a field F to A as a Lie algebra (by ) has a left adjoint , called the universal enveloping algebra. To construct this: given a Lie algebra , let

be the tensor algebra on , also called the free associative algebra on the vector space . Here denotes the tensor product of F-vector spaces. Let I be the two-sided ideal in generated by the elements for ; then the universal enveloping algebra is the quotient ring . It satisfies the Poincaré–Birkhoff–Witt theorem: if is a basis for as a k-vector space, then a basis for is given by all ordered products with natural numbers. In particular, the map is injective.[23]

Representations of are equivalent to modules over the universal enveloping algebra. The fact that is injective implies that every Lie algebra (possibly of infinite dimension) has a faithful representation (of infinite dimension), namely its representation on . This also shows that every Lie algebra is contained in the Lie algebra associated to some associative algebra.

Representation theory in physics

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 of the rotation group SO(3). Typically, the space of states is far from being irreducible under the pertinent operators, but one can attempt to decompose it into irreducible pieces. In doing so, one needs to know the irreducible representations of the given Lie algebra. In the study of the quantum hydrogen atom, for example, quantum mechanics textbooks classify (more or less explicitly) the finite-dimensional irreducible representations of the Lie algebra .[18]

Structure theory and classification

Lie algebras can be classified to some extent. This is a powerful approach to the classification of Lie groups.

Abelian, nilpotent, and solvable

Analogously to abelian, nilpotent, and solvable groups, one can define abelian, nilpotent, and solvable Lie algebras.

A Lie algebra is abelian if the Lie bracket vanishes; that is, [x,y] = 0 for all x and y in . In particular, the Lie algebra of an abelian Lie group (such as the group under addition or the torus group ) is abelian. Every finite-dimensional abelian Lie algebra over a field is isomorphic to for some , meaning an n-dimensional vector space with Lie bracket zero.

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 is , meaning the linear subspace spanned by all brackets with . The commutator subalgebra is an ideal in , in fact the smallest ideal such that the quotient Lie algebra is abelian. It is analogous to the commutator subgroup of a group.

A Lie algebra is nilpotent if the lower central series

becomes zero after finitely many steps. Equivalently, is nilpotent if there is a finite sequence of ideals in ,

such that is central in for each j. By Engel's theorem, a Lie algebra over any field is nilpotent if and only if for every u in the adjoint endomorphism

is nilpotent.[24]

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

becomes zero after finitely many steps. Equivalently, is solvable if there is a finite sequence of Lie subalgebras,

such that is an ideal in with abelian for each j.[25]

Every finite-dimensional Lie algebra over a field has a unique maximal solvable ideal, called its radical.[26] Under the Lie correspondence, nilpotent (respectively, solvable) Lie groups correspond to nilpotent (respectively, solvable) Lie algebras over .

For example, for a positive integer n, the radical of is its center, the 1-dimensional subspace spanned by the identity matrix. An example of a solvable Lie algebra is the space of upper-triangular matrices in ; this is not nilpotent when . An example of a nilpotent Lie algebra is the space of strictly upper-triangular matrices in ; this is not abelian when .

Simple and semisimple

A Lie algebra is called simple if it is not abelian and the only ideals in are 0 and . (In particular, a one-dimensional—necessarily abelian—Lie algebra is by definition not simple, even though its only ideals are 0 and .) A finite-dimensional Lie algebra is called semisimple if the only solvable ideal in is 0. In characteristic zero, a Lie algebra is semisimple if and only if it is isomorphic to a product of simple Lie algebras, .[27]

For example, the Lie algebra is simple for every and every field F of characteristic zero (or just of characteristic not dividing n). The Lie algebra over is simple for every . The Lie algebra over is simple if or .[28] (There are "exceptional isomorphisms" and .)

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).[22]

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.[29]

For example, is reductive for F of characteristic zero: for , it is isomorphic to the product

where F denotes the center of , the 1-dimensional subspace spanned by the identity matrix. Since the special linear Lie algebra is simple, contains few ideals: only 0, the center F, , and all of .

Cartan's criterion

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 defined by

where tr denotes the trace of a linear operator. Namely: a Lie algebra is semisimple if and only if the Killing form is nondegenerate. A Lie algebra is solvable if and only if [30]

Classification

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.[31] 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.[32] Here the simple Lie algebra of type An is , Bn is , Cn is , and Dn is . The other five are known as the exceptional Lie algebras.

The classification of finite-dimensional simple Lie algebras over is more complicated, but it was also solved by Cartan (see simple Lie group for an equivalent classification). One can analyze a Lie algebra over by considering its complexification .

In the years leading up to 2004, the finite-dimensional simple Lie algebras over an algebraically closed field of characteristic were classified by Richard Earl Block, Robert Lee Wilson, Alexander Premet, and Helmut Strade. (See restricted Lie algebra#Classification of simple Lie algebras.) It turns out that there are many more simple Lie algebras in positive characteristic than in characteristic zero.

Relation to Lie groups

The tangent space of a sphere at a point . If were the identity element of a Lie group, the tangent space would be a Lie algebra.

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 (concretely, the tangent space at the identity). Conversely, for every finite-dimensional Lie algebra , there is a connected Lie group with Lie algebra . This is Lie's third theorem; see the Baker–Campbell–Hausdorff formula. This Lie group is not determined uniquely; however, any two Lie groups with the same Lie algebra are locally isomorphic, and more strongly, they 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 with the cross product, but SU(2) is a simply connected double cover of SO(3).

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 .[33]

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.[34] 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.[35]

Lie theory also does not work so neatly for infinite-dimensional representations of a finite-dimensional group. Even for the additive group , an infinite-dimensional representation of can usually not be differentiated to produce a representation of its Lie algebra on the same space, or vice versa.[36] The theory of Harish-Chandra modules is a more subtle relation between infinite-dimensional representations for groups and Lie algebras.

Real form and complexification

Given a complex Lie algebra , a real Lie algebra is said to be a real form of if the complexification is isomorphic to . A real form need not be unique; for example, has two real forms up to isomorphism, and .[37]

Given a semisimple complex Lie algebra , a split form of it is a real form that splits; i.e., it has a Cartan subalgebra which acts via an adjoint representation with real eigenvalues. A split form exists and is unique (up to isomorphism). A compact form is a real form that is the Lie algebra of a compact Lie group. A compact form exists and is also unique up to isomorphism.[37]

Lie algebra with additional structures

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 to describe rational homotopy theory in algebraic terms.[38]

Lie ring

The definition of a Lie algebra over a field extends to define a Lie algebra over any commutative ring R. Namely, a Lie algebra over R is an R-module with an alternating R-bilinear map that satisfies the Jacobi identity. A Lie algebra over the ring of integers is sometimes called a Lie ring. (This is not directly related to the notion of a Lie group.)

Lie rings are used in the study of finite p-groups (for a prime number p) through the Lazard correspondence.[39] 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 of p-adic numbers as well as over the ring of p-adic integers.[40] Part of Claude Chevalley's construction of the finite groups of Lie type involves showing that a simple Lie algebra over the complex numbers comes from a Lie algebra over the integers, and then (with more care) a group scheme over the integers.[41]

Examples

  • Here is a construction of Lie rings arising from the study of abstract groups. For elements of a group, define the commutator . Let be a filtration of a group , that is, a chain of subgroups such that is contained in for all . (For the Lazard correspondence, one takes the filtration to be the lower central series of G.) Then
is a Lie ring, with addition given by the group multiplication (which is abelian on each quotient group ), and with Lie bracket given by commutators in the group:[42]
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 .

Definition using category-theoretic notation

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 A is a vector space, the interchange isomorphism is defined by

The cyclic-permutation braiding is defined as

where is the identity morphism. Equivalently, is defined by

With this notation, a Lie algebra can be defined as an object in the category of vector spaces together with a morphism

that satisfies the two morphism equalities

and

See also

Remarks

  1. ^ More generally, one has the notion of a Lie algebra over any commutative ring R: an R-module with an alternating R-bilinear map that satisfies the Jacobi identity (Bourbaki (1989, Section 2)).

References

  1. ^ O'Connor & Robertson 2000.
  2. ^ O'Connor & Robertson 2005.
  3. ^ Humphreys 1978, p. 1.
  4. ^ Bourbaki 1989, §1.2. Example 1.
  5. ^ Bourbaki 1989, §1.2. Example 2.
  6. ^ By the anticommutativity of the commutator, the notions of a left and right ideal in a Lie algebra coincide.
  7. ^ Jacobson 1979, p. 28.
  8. ^ Bourbaki 1989, section I.1.1.
  9. ^ Humphreys 1978, p. 4.
  10. ^ Varadarajan 1984, p. 49.
  11. ^ Serre 2006, Part I, section VI.3.
  12. ^ Fulton & Harris 1991, Proposition D.40.
  13. ^ Varadarajan 1984, section 2.10, Remark 2.
  14. ^ Hall 2015, §3.4.
  15. ^ Erdmann & Wildon 2006, Theorem 3.1.
  16. ^ Erdmann & Wildon 2006, section 3.2.1.
  17. ^ Hall 2015, Example 3.27.
  18. ^ a b Wigner 1959, Chapters 17 and 20.
  19. ^ Erdmann & Wildon 2006, Chapter 8.
  20. ^ Serre 2006, Part I, Chapter IV.
  21. ^ Jacobson 1979, Ch. VI.
  22. ^ a b Hall 2015, Theorem 10.9.
  23. ^ Humphreys 1978, section 17.3.
  24. ^ Jacobson 1979, section II.3.
  25. ^ Jacobson 1979, section I.7.
  26. ^ Jacobson 1979, p. 24.
  27. ^ Jacobson 1979, Ch. III, § 5.
  28. ^ Erdmann & Wildon 2006, Theorem 12.1.
  29. ^ Varadarajan 1984, Theorem 3.16.3.
  30. ^ Varadarajan 1984, section 3.9.
  31. ^ Jacobson 1979, Ch. III, § 9.
  32. ^ Jacobson 1979, section IV.6.
  33. ^ Varadarajan 1984, Theorems 2.7.5 and 3.15.1.
  34. ^ Varadarjan 1984, section 2.6.
  35. ^ Milnor 2010, Warnings 1.6 and 8.5.
  36. ^ Knapp 2001, section III.3, Problem III.5.
  37. ^ a b Fulton & Harris 1991, §26.1.
  38. ^ Quillen 1969, Corollary II.6.2.
  39. ^ Khukhro 1998, Ch. 6.
  40. ^ Serre 2006, Part II, section V.1.
  41. ^ Humphreys 1978, section 25.
  42. ^ Serre 2006, Part I, Chapter II.

Sources