Root system

Last updated

In mathematics, a root system is a configuration of vectors in a Euclidean space satisfying certain geometrical properties. The concept is fundamental in the theory of Lie groups and Lie algebras, especially the classification and representation theory of semisimple Lie algebras. Since Lie groups (and some analogues such as algebraic groups) and Lie algebras have become important in many parts of mathematics during the twentieth century, the apparently special nature of root systems belies the number of areas in which they are applied. Further, the classification scheme for root systems, by Dynkin diagrams, occurs in parts of mathematics with no overt connection to Lie theory (such as singularity theory). Finally, root systems are important for their own sake, as in spectral graph theory. [1]

Contents

Definitions and examples

The six vectors of the root system A2 Root system A2 with labels.png
The six vectors of the root system A2

As a first example, consider the six vectors in 2-dimensional Euclidean space, R2, as shown in the image at the right; call them roots. These vectors span the whole space. If you consider the line perpendicular to any root, say β, then the reflection of R2 in that line sends any other root, say α, to another root. Moreover, the root to which it is sent equals α + , where n is an integer (in this case, n equals 1). These six vectors satisfy the following definition, and therefore they form a root system; this one is known as A2.

Definition

Let E be a finite-dimensional Euclidean vector space, with the standard Euclidean inner product denoted by . A root system in E is a finite set of non-zero vectors (called roots) that satisfy the following conditions: [2] [3]

  1. The roots span E.
  2. The only scalar multiples of a root that belong to are itself and .
  3. For every root , the set is closed under reflection through the hyperplane perpendicular to .
  4. (Integrality) If and are roots in , then the projection of onto the line through is an integer or half-integer multiple of .

An equivalent way of writing conditions 3 and 4 is as follows:

  1. For any two roots , the set contains the element
  2. For any two roots , the number is an integer.

Some authors only include conditions 13 in the definition of a root system. [4] In this context, a root system that also satisfies the integrality condition is known as a crystallographic root system. [5] Other authors omit condition 2; then they call root systems satisfying condition 2 reduced. [6] In this article, all root systems are assumed to be reduced and crystallographic.

In view of property 3, the integrality condition is equivalent to stating that β and its reflection σα(β) differ by an integer multiple of α. Note that the operator

defined by property 4 is not an inner product. It is not necessarily symmetric and is linear only in the first argument.

Rank-2 root systems
Root system A1xA1.svg Root system D2.svg
Root system
Dyn-node n1.pngDyn-2.pngDyn-node n2.png
Root system
Dyn2-nodes.png
Root system A2.svg Root system G2.svg
Root system
Dyn2-node n1.pngDyn2-3.pngDyn2-node n2.png
Root system
Dyn2-nodeg n1.pngDyn2-6a.pngDyn2-node n2.png
Root system B2.svg Root system C2 (fixed).svg
Root system
Dyn2-nodeg n1.pngDyn2-4a.pngDyn2-node n2.png
Root system
Dyn2-node n1.pngDyn2-4b.pngDyn2-nodeg n2.png

The rank of a root system Φ is the dimension of E. Two root systems may be combined by regarding the Euclidean spaces they span as mutually orthogonal subspaces of a common Euclidean space. A root system which does not arise from such a combination, such as the systems A2, B2, and G2 pictured to the right, is said to be irreducible.

Two root systems (E1, Φ1) and (E2, Φ2) are called isomorphic if there is an invertible linear transformation E1  E2 which sends Φ1 to Φ2 such that for each pair of roots, the number is preserved. [7]

The root lattice of a root system Φ is the Z-submodule of E generated by Φ. It is a lattice in E.

Weyl group

The Weyl group of the
A
2
{\displaystyle A_{2}}
root system is the symmetry group of an equilateral triangle A2 Weyl group (revised).png
The Weyl group of the root system is the symmetry group of an equilateral triangle

The group of isometries of E generated by reflections through hyperplanes associated to the roots of Φ is called the Weyl group of Φ. As it acts faithfully on the finite set Φ, the Weyl group is always finite. The reflection planes are the hyperplanes perpendicular to the roots, indicated for by dashed lines in the figure below. The Weyl group is the symmetry group of an equilateral triangle, which has six elements. In this case, the Weyl group is not the full symmetry group of the root system (e.g., a 60-degree rotation is a symmetry of the root system but not an element of the Weyl group).

Rank one example

There is only one root system of rank 1, consisting of two nonzero vectors . This root system is called .

Rank two examples

In rank 2 there are four possibilities, corresponding to , where . [8] The figure at right shows these possibilities, but with some redundancies: is isomorphic to and is isomorphic to .

Note that a root system is not determined by the lattice that it generates: and both generate a square lattice while and both generate a hexagonal lattice.

Whenever Φ is a root system in E, and S is a subspace of E spanned by Ψ = Φ  S, then Ψ is a root system in S. Thus, the exhaustive list of four root systems of rank 2 shows the geometric possibilities for any two roots chosen from a root system of arbitrary rank. In particular, two such roots must meet at an angle of 0, 30, 45, 60, 90, 120, 135, 150, or 180 degrees.

Root systems arising from semisimple Lie algebras

If is a complex semisimple Lie algebra and is a Cartan subalgebra, we can construct a root system as follows. We say that is a root of relative to if and there exists some such that

for all . One can show [9] that there is an inner product for which the set of roots forms a root system. The root system of is a fundamental tool for analyzing the structure of and classifying its representations. (See the section below on Root systems and Lie theory.)

History

The concept of a root system was originally introduced by Wilhelm Killing around 1889 (in German, Wurzelsystem [10] ). [11] He used them in his attempt to classify all simple Lie algebras over the field of complex numbers. Killing originally made a mistake in the classification, listing two exceptional rank 4 root systems, when in fact there is only one, now known as F4. Cartan later corrected this mistake, by showing Killing's two root systems were isomorphic. [12]

Killing investigated the structure of a Lie algebra , by considering what is now called a Cartan subalgebra . Then he studied the roots of the characteristic polynomial , where . Here a root is considered as a function of , or indeed as an element of the dual vector space . This set of roots form a root system inside , as defined above, where the inner product is the Killing form. [11]

Elementary consequences of the root system axioms

The integrality condition for
<
b
,
a
> 
{\displaystyle \langle \beta ,\alpha \rangle }
is fulfilled only for b on one of the vertical lines, while the integrality condition for
<
a
,
b
> 
{\displaystyle \langle \alpha ,\beta \rangle }
is fulfilled only for b on one of the red circles. Any b perpendicular to a (on the Y axis) trivially fulfills both with 0, but does not define an irreducible root system.
Modulo reflection, for a given a there are only 5 nontrivial possibilities for b, and 3 possible angles between a and b in a set of simple roots. Subscript letters correspond to the series of root systems for which the given b can serve as the first root and a as the second root (or in F4 as the middle 2 roots). Integrality of root systems.svg
The integrality condition for is fulfilled only for β on one of the vertical lines, while the integrality condition for is fulfilled only for β on one of the red circles. Any β perpendicular to α (on the Y axis) trivially fulfills both with 0, but does not define an irreducible root system.
Modulo reflection, for a given α there are only 5 nontrivial possibilities for β, and 3 possible angles between α and β in a set of simple roots. Subscript letters correspond to the series of root systems for which the given β can serve as the first root and α as the second root (or in F4 as the middle 2 roots).


The cosine of the angle between two roots is constrained to be one-half of the square root of a positive integer. This is because and are both integers, by assumption, and

Since , the only possible values for are and , corresponding to angles of 90°, 60° or 120°, 45° or 135°, 30° or 150°, and 0° or 180°. Condition 2 says that no scalar multiples of α other than 1 and −1 can be roots, so 0 or 180°, which would correspond to 2α or −2α, are out. The diagram at right shows that an angle of 60° or 120° corresponds to roots of equal length, while an angle of 45° or 135° corresponds to a length ratio of and an angle of 30° or 150° corresponds to a length ratio of .

In summary, here are the only possibilities for each pair of roots. [13]

Positive roots and simple roots

The labeled roots are a set of positive roots for the
G
2
{\displaystyle G_{2}}
root system, with
a
1
{\displaystyle \alpha _{1}}
and
a
2
{\displaystyle \alpha _{2}}
being the simple roots Base for the G2 root system.png
The labeled roots are a set of positive roots for the root system, with and being the simple roots

Given a root system we can always choose (in many ways) a set of positive roots. This is a subset of such that

If a set of positive roots is chosen, elements of are called negative roots. A set of positive roots may be constructed by choosing a hyperplane not containing any root and setting to be all the roots lying on a fixed side of . Furthermore, every set of positive roots arises in this way. [14]

An element of is called a simple root (also fundamental root) if it cannot be written as the sum of two elements of . (The set of simple roots is also referred to as a base for .) The set of simple roots is a basis of with the following additional special properties: [15]

For each root system there are many different choices of the set of positive roots—or, equivalently, of the simple roots—but any two sets of positive roots differ by the action of the Weyl group. [16]

Dual root system, coroots, and integral elements

The dual root system

If Φ is a root system in E, the coroot α of a root α is defined by

The set of coroots also forms a root system Φ in E, called the dual root system (or sometimes inverse root system). By definition, α∨ ∨ = α, so that Φ is the dual root system of Φ. The lattice in E spanned by Φ is called the coroot lattice. Both Φ and Φ have the same Weyl group W and, for s in W,

If Δ is a set of simple roots for Φ, then Δ is a set of simple roots for Φ. [17]

In the classification described below, the root systems of type and along with the exceptional root systems are all self-dual, meaning that the dual root system is isomorphic to the original root system. By contrast, the and root systems are dual to one another, but not isomorphic (except when ).

Integral elements

A vector in E is called integral [18] if its inner product with each coroot is an integer:

Since the set of with forms a base for the dual root system, to verify that is integral, it suffices to check the above condition for .

The set of integral elements is called the weight lattice associated to the given root system. This term comes from the representation theory of semisimple Lie algebras, where the integral elements form the possible weights of finite-dimensional representations.

The definition of a root system guarantees that the roots themselves are integral elements. Thus, every integer linear combination of roots is also integral. In most cases, however, there will be integral elements that are not integer combinations of roots. That is to say, in general the weight lattice does not coincide with the root lattice.

Classification of root systems by Dynkin diagrams

Pictures of all the connected Dynkin diagrams Finite Dynkin diagrams.svg
Pictures of all the connected Dynkin diagrams

A root system is irreducible if it cannot be partitioned into the union of two proper subsets , such that for all and .

Irreducible root systems correspond to certain graphs, the Dynkin diagrams named after Eugene Dynkin. The classification of these graphs is a simple matter of combinatorics, and induces a classification of irreducible root systems.

Constructing the Dynkin diagram

Given a root system, select a set Δ of simple roots as in the preceding section. The vertices of the associated Dynkin diagram correspond to the roots in Δ. Edges are drawn between vertices as follows, according to the angles. (Note that the angle between simple roots is always at least 90 degrees.)

The term "directed edge" means that double and triple edges are marked with an arrow pointing toward the shorter vector. (Thinking of the arrow as a "greater than" sign makes it clear which way the arrow is supposed to point.)

Note that by the elementary properties of roots noted above, the rules for creating the Dynkin diagram can also be described as follows. No edge if the roots are orthogonal; for nonorthogonal roots, a single, double, or triple edge according to whether the length ratio of the longer to shorter is 1, , . In the case of the root system for example, there are two simple roots at an angle of 150 degrees (with a length ratio of ). Thus, the Dynkin diagram has two vertices joined by a triple edge, with an arrow pointing from the vertex associated to the longer root to the other vertex. (In this case, the arrow is a bit redundant, since the diagram is equivalent whichever way the arrow goes.)

Classifying root systems

Although a given root system has more than one possible set of simple roots, the Weyl group acts transitively on such choices. [19] Consequently, the Dynkin diagram is independent of the choice of simple roots; it is determined by the root system itself. Conversely, given two root systems with the same Dynkin diagram, one can match up roots, starting with the roots in the base, and show that the systems are in fact the same. [20]

Thus the problem of classifying root systems reduces to the problem of classifying possible Dynkin diagrams. A root systems is irreducible if and only if its Dynkin diagrams is connected. [21] The possible connected diagrams are as indicated in the figure. The subscripts indicate the number of vertices in the diagram (and hence the rank of the corresponding irreducible root system).

If is a root system, the Dynkin diagram for the dual root system is obtained from the Dynkin diagram of by keeping all the same vertices and edges, but reversing the directions of all arrows. Thus, we can see from their Dynkin diagrams that and are dual to each other.

Weyl chambers and the Weyl group

The shaded region is the fundamental Weyl chamber for the base
{
a
1
,
a
2
}
{\displaystyle \{\alpha _{1},\alpha _{2}\}} Weyl chambers for A2.png
The shaded region is the fundamental Weyl chamber for the base

If is a root system, we may consider the hyperplane perpendicular to each root . Recall that denotes the reflection about the hyperplane and that the Weyl group is the group of transformations of generated by all the 's. The complement of the set of hyperplanes is disconnected, and each connected component is called a Weyl chamber. If we have fixed a particular set Δ of simple roots, we may define the fundamental Weyl chamber associated to Δ as the set of points such that for all .

Since the reflections preserve , they also preserve the set of hyperplanes perpendicular to the roots. Thus, each Weyl group element permutes the Weyl chambers.

The figure illustrates the case of the root system. The "hyperplanes" (in this case, one dimensional) orthogonal to the roots are indicated by dashed lines. The six 60-degree sectors are the Weyl chambers and the shaded region is the fundamental Weyl chamber associated to the indicated base.

A basic general theorem about Weyl chambers is this: [22]

Theorem: The Weyl group acts freely and transitively on the Weyl chambers. Thus, the order of the Weyl group is equal to the number of Weyl chambers.

In the case, for example, the Weyl group has six elements and there are six Weyl chambers.

A related result is this one: [23]

Theorem: Fix a Weyl chamber . Then for all , the Weyl-orbit of contains exactly one point in the closure of .

Root systems and Lie theory

Irreducible root systems classify a number of related objects in Lie theory, notably the following:

In each case, the roots are non-zero weights of the adjoint representation.

We now give a brief indication of how irreducible root systems classify simple Lie algebras over , following the arguments in Humphreys. [24] A preliminary result says that a semisimple Lie algebra is simple if and only if the associated root system is irreducible. [25] We thus restrict attention to irreducible root systems and simple Lie algebras.

For connections between the exceptional root systems and their Lie groups and Lie algebras see E8, E7, E6, F4, and G2.

Properties of the irreducible root systems

Φ|Φ||Φ<|ID|W|
An (n ≥ 1)n(n + 1)n + 1(n + 1)!
Bn (n ≥ 2)2n22n222nn!
Cn (n ≥ 3)2n22n(n − 1)2n−122nn!
Dn (n ≥ 4)2n(n − 1)42n−1n!
E6 72351840
E7 12622903040
E8 2401696729600
F4 4824411152
G2 1263112

Irreducible root systems are named according to their corresponding connected Dynkin diagrams. There are four infinite families (An, Bn, Cn, and Dn, called the classical root systems) and five exceptional cases (the exceptional root systems). The subscript indicates the rank of the root system.

In an irreducible root system there can be at most two values for the length (α, α)1/2, corresponding to short and long roots. If all roots have the same length they are taken to be long by definition and the root system is said to be simply laced; this occurs in the cases A, D and E. Any two roots of the same length lie in the same orbit of the Weyl group. In the non-simply laced cases B, C, G and F, the root lattice is spanned by the short roots and the long roots span a sublattice, invariant under the Weyl group, equal to r2/2 times the coroot lattice, where r is the length of a long root.

In the adjacent table, |Φ<| denotes the number of short roots, I denotes the index in the root lattice of the sublattice generated by long roots, D denotes the determinant of the Cartan matrix, and |W| denotes the order of the Weyl group.

Explicit construction of the irreducible root systems

An

Model of the
A
3
{\displaystyle A_{3}}
root system in the Zometool system A3vzome.jpg
Model of the root system in the Zometool system
Simple roots in A3
e1e2e3e4
α11−100
α201−10
α3001−1
Dyn2-node n1.pngDyn2-3.pngDyn2-node n2.pngDyn2-3.pngDyn2-node n3.png

Let E be the subspace of Rn+1 for which the coordinates sum to 0, and let Φ be the set of vectors in E of length 2 and which are integer vectors, i.e. have integer coordinates in Rn+1. Such a vector must have all but two coordinates equal to 0, one coordinate equal to 1, and one equal to −1, so there are n2 + n roots in all. One choice of simple roots expressed in the standard basis is αi = eiei+1 for 1 ≤ in.

The reflection σi through the hyperplane perpendicular to αi is the same as permutation of the adjacent ith and (i + 1)th coordinates. Such transpositions generate the full permutation group. For adjacent simple roots, σi(αi+1) = αi+1 + αi = σi+1(αi) = αi + αi+1, that is, reflection is equivalent to adding a multiple of 1; but reflection of a simple root perpendicular to a nonadjacent simple root leaves it unchanged, differing by a multiple of 0.

The An root lattice – that is, the lattice generated by the An roots – is most easily described as the set of integer vectors in Rn+1 whose components sum to zero.

The A2 root lattice is the vertex arrangement of the triangular tiling.

The A3 root lattice is known to crystallographers as the face-centered cubic (or cubic close packed) lattice. [29] It is the vertex arrangement of the tetrahedral-octahedral honeycomb.

The A3 root system (as well as the other rank-three root systems) may be modeled in the Zometool Construction set. [30]

In general, the An root lattice is the vertex arrangement of the n-dimensional simplectic honeycomb.

Bn

Simple roots in B4
e1e2e3e4
α1 1−100
α20  1−10
α300 1−1
α4000 1
Dyn2-node n1.pngDyn2-3.pngDyn2-node n2.pngDyn2-3.pngDyn2-node n3.pngDyn2-4b.pngDyn2-nodeg n4.png

Let E = Rn, and let Φ consist of all integer vectors in E of length 1 or 2. The total number of roots is 2n2. One choice of simple roots is αi = eiei+1 for 1 ≤ in – 1 (the above choice of simple roots for An−1), and the shorter root αn = en.

The reflection σn through the hyperplane perpendicular to the short root αn is of course simply negation of the nth coordinate. For the long simple root αn−1, σn−1(αn) = αn + αn−1, but for reflection perpendicular to the short root, σn(αn−1) = αn−1 + 2αn, a difference by a multiple of 2 instead of 1.

The Bn root lattice—that is, the lattice generated by the Bn roots—consists of all integer vectors.

B1 is isomorphic to A1 via scaling by 2, and is therefore not a distinct root system.

Cn

Root system B3, C3, and A3 = D3 as points within a cube and octahedron Root vectors b3 c3-d3.png
Root system B3, C3, and A3 = D3 as points within a cube and octahedron
Simple roots in C4
e1e2e3e4
α1 1−100
α20 1−10
α300 1−1
α4000 2
Dyn2-nodeg n1.pngDyn2-3.pngDyn2-nodeg n2.pngDyn2-3.pngDyn2-nodeg n3.pngDyn2-4a.pngDyn2-node n4.png

Let E = Rn, and let Φ consist of all integer vectors in E of length 2 together with all vectors of the form 2λ, where λ is an integer vector of length 1. The total number of roots is 2n2. One choice of simple roots is: αi = eiei+1, for 1 ≤ in − 1 (the above choice of simple roots for An−1), and the longer root αn = 2en. The reflection σn(αn−1) = αn−1 + αn, but σn−1(αn) = αn + 2αn−1.

The Cn root lattice—that is, the lattice generated by the Cn roots—consists of all integer vectors whose components sum to an even integer.

C2 is isomorphic to B2 via scaling by 2 and a 45 degree rotation, and is therefore not a distinct root system.

Dn

Simple roots in D4
e1e2e3e4
α1 1−100
α20 1−10
α300 1−1
α400 1 1
DynkinD4 labeled.png

Let E = Rn, and let Φ consist of all integer vectors in E of length 2. The total number of roots is 2n(n − 1). One choice of simple roots is αi = eiei+1 for 1 ≤ in − 1 (the above choice of simple roots for An−1) together with αn = en−1 + en.

Reflection through the hyperplane perpendicular to αn is the same as transposing and negating the adjacent n-th and (n − 1)-th coordinates. Any simple root and its reflection perpendicular to another simple root differ by a multiple of 0 or 1 of the second root, not by any greater multiple.

The Dn root lattice – that is, the lattice generated by the Dn roots – consists of all integer vectors whose components sum to an even integer. This is the same as the Cn root lattice.

The Dn roots are expressed as the vertices of a rectified n-orthoplex, Coxeter–Dynkin diagram: CDel node.pngCDel 3.pngCDel node 1.pngCDel 3.png...CDel 3.pngCDel node.pngCDel split1.pngCDel nodes.png. The 2n(n − 1) vertices exist in the middle of the edges of the n-orthoplex.

D3 coincides with A3, and is therefore not a distinct root system. The twelve D3 root vectors are expressed as the vertices of CDel node.pngCDel split1.pngCDel nodes 11.png, a lower symmetry construction of the cuboctahedron.

D4 has additional symmetry called triality. The twenty-four D4 root vectors are expressed as the vertices of CDel node.pngCDel 3.pngCDel node 1.pngCDel split1.pngCDel nodes.png, a lower symmetry construction of the 24-cell.

E6, E7, E8

E6Coxeter.svg
72 vertices of 122 represent the root vectors of E6
(Green nodes are doubled in this E6 Coxeter plane projection)
E7Petrie.svg
126 vertices of 231 represent the root vectors of E7
E8 graph.svg
240 vertices of 421 represent the root vectors of E8
DynkinE6AltOrder.svg DynkinE7AltOrder.svg DynkinE8AltOrder.svg

The root system has 240 roots. The set just listed is the set of vectors of length 2 in the E8 root lattice, also known simply as the E8 lattice or Γ8. This is the set of points in R8 such that:

  1. all the coordinates are integers or all the coordinates are half-integers (a mixture of integers and half-integers is not allowed), and
  2. the sum of the eight coordinates is an even integer.

Thus,

Simple roots in E8: even coordinates
1−1000000
01−100000
001−10000
0001−1000
00001−100
000001−10
00000110
1/21/21/21/21/21/21/21/2

An alternative description of the E8 lattice which is sometimes convenient is as the set Γ'8 of all points in R8 such that

The lattices Γ8 and Γ'8 are isomorphic; one may pass from one to the other by changing the signs of any odd number of coordinates. The lattice Γ8 is sometimes called the even coordinate system for E8 while the lattice Γ'8 is called the odd coordinate system.

One choice of simple roots for E8 in the even coordinate system with rows ordered by node order in the alternate (non-canonical) Dynkin diagrams (above) is:

αi = eiei+1, for 1 ≤ i ≤ 6, and
α7 = e7 + e6

(the above choice of simple roots for D7) along with

Simple roots in E8: odd coordinates
1−1000000
01−100000
001−10000
0001−1000
00001−100
000001−10
0000001−1
1/21/21/21/21/2 1/2 1/2 1/2

One choice of simple roots for E8 in the odd coordinate system with rows ordered by node order in alternate (non-canonical) Dynkin diagrams (above) is

αi = eiei+1, for 1 ≤ i ≤ 7

(the above choice of simple roots for A7) along with

α8 = β5, where

(Using β3 would give an isomorphic result. Using β1,7 or β2,6 would simply give A8 or D8. As for β4, its coordinates sum to 0, and the same is true for α1...7, so they span only the 7-dimensional subspace for which the coordinates sum to 0; in fact −2β4 has coordinates (1,2,3,4,3,2,1) in the basis (αi).)

Since perpendicularity to α1 means that the first two coordinates are equal, E7 is then the subset of E8 where the first two coordinates are equal, and similarly E6 is the subset of E8 where the first three coordinates are equal. This facilitates explicit definitions of E7 and E6 as

E7 = {αZ7 ∪ (Z+1/2)7 : Σαi2 + α12 = 2, Σαi + α1 ∈ 2Z},
E6 = {αZ6 ∪ (Z+1/2)6 : Σαi2 + 2α12 = 2, Σαi + 2α1 ∈ 2Z}

Note that deleting α1 and then α2 gives sets of simple roots for E7 and E6. However, these sets of simple roots are in different E7 and E6 subspaces of E8 than the ones written above, since they are not orthogonal to α1 or α2.

F4

Simple roots in F4
e1e2e3e4
α11−100
α201−10
α30010
α41/21/21/21/2
Dyn2-node n1.pngDyn2-3.pngDyn2-node n2.pngDyn2-4b.pngDyn2-nodeg n3.pngDyn2-3.pngDyn2-nodeg n4.png
48-root vectors of F4, defined by vertices of the 24-cell and its dual, viewed in the Coxeter plane F4 roots by 24-cell duals.svg
48-root vectors of F4, defined by vertices of the 24-cell and its dual, viewed in the Coxeter plane

For F4, let E = R4, and let Φ denote the set of vectors α of length 1 or 2 such that the coordinates of 2α are all integers and are either all even or all odd. There are 48 roots in this system. One choice of simple roots is: the choice of simple roots given above for B3, plus .

The F4 root lattice—that is, the lattice generated by the F4 root system—is the set of points in R4 such that either all the coordinates are integers or all the coordinates are half-integers (a mixture of integers and half-integers is not allowed). This lattice is isomorphic to the lattice of Hurwitz quaternions.

G2

Simple roots in G2
e1e2e3
α11 −1  0
β−12−1
Dyn2-nodeg n1.pngDyn2-6a.pngDyn2-node n2.png

The root system G2 has 12 roots, which form the vertices of a hexagram. See the picture above.

One choice of simple roots is (α1, β = α2α1) where αi = eiei+1 for i = 1, 2 is the above choice of simple roots for A2.

The G2 root lattice—that is, the lattice generated by the G2 roots—is the same as the A2 root lattice.

The root poset

Hasse diagram of E6 root poset with edge labels identifying added simple root position E6HassePoset.svg
Hasse diagram of E6 root poset with edge labels identifying added simple root position

The set of positive roots is naturally ordered by saying that if and only if is a nonnegative linear combination of simple roots. This poset is graded by , and has many remarkable combinatorial properties, one of them being that one can determine the degrees of the fundamental invariants of the corresponding Weyl group from this poset. [31] The Hasse graph is a visualization of the ordering of the root poset.

See also

Notes

  1. Cvetković, Dragoš (2002). "Graphs with least eigenvalue −2; a historical survey and recent developments in maximal exceptional graphs". Linear Algebra and Its Applications. 356 (1–3): 189–210. doi: 10.1016/S0024-3795(02)00377-4 .
  2. Bourbaki, Ch.VI, Section 1
  3. Humphreys 1972 , p. 42
  4. Humphreys 1992 , p. 6
  5. Humphreys 1992 , p. 39
  6. Humphreys 1992 , p. 41
  7. Humphreys 1972 , p. 43
  8. Hall 2015 Proposition 8.8
  9. Hall 2015 , Section 7.5
  10. Killing 1889
  11. 1 2 Bourbaki 1998 , p. 270
  12. Coleman 1989 , p. 34
  13. Hall 2015 Proposition 8.6
  14. Hall 2015 , Theorems 8.16 and 8.17
  15. Hall 2015 , Theorem 8.16
  16. Hall 2015 , Proposition 8.28
  17. Hall 2015 , Proposition 8.18
  18. Hall 2015 , Section 8.7
  19. This follows from Hall 2015 , Proposition 8.23
  20. Hall 2015 , Proposition 8.32
  21. Hall 2015 , Proposition 8.23
  22. Hall 2015 , Propositions 8.23 and 8.27
  23. Hall 2015 , Proposition 8.29
  24. See various parts of Chapters III, IV, and V of Humphreys 1972, culminating in Section 19 in Chapter V
  25. Hall 2015, Theorem 7.35
  26. Humphreys 1972 , Section 16
  27. Humphreys 1972 , Part (b) of Theorem 18.4
  28. Humphreys 1972 Section 18.3 and Theorem 18.4
  29. Conway, John; Sloane, Neil J.A. (1998). "Section 6.3". Sphere Packings, Lattices and Groups. Springer. ISBN   978-0-387-98585-5.
  30. Hall 2015 Section 8.9
  31. Humphreys 1992 , Theorem 3.20

Related Research Articles

<span class="mw-page-title-main">Dynkin diagram</span> Pictorial representation of symmetry

In the mathematical field of Lie theory, a Dynkin diagram, named for Eugene Dynkin, is a type of graph with some edges doubled or tripled. Dynkin diagrams arise in the classification of semisimple Lie algebras over algebraically closed fields, in the classification of Weyl groups and other finite reflection groups, and in other contexts. Various properties of the Dynkin diagram correspond to important features of the associated Lie algebra.

In the mathematical field of representation theory, a weight of an algebra A over a field F is an algebra homomorphism from A to F, or equivalently, a one-dimensional representation of A over F. It is the algebra analogue of a multiplicative character of a group. The importance of the concept, however, stems from its application to representations of Lie algebras and hence also to representations of algebraic and Lie groups. In this context, a weight of a representation is a generalization of the notion of an eigenvalue, and the corresponding eigenspace is called a weight space.

<span class="mw-page-title-main">Weyl group</span> Subgroup of a root systems isometry group

In mathematics, in particular the theory of Lie algebras, the Weyl group of a root system Φ is a subgroup of the isometry group of that root system. Specifically, it is the subgroup which is generated by reflections through the hyperplanes orthogonal to the roots, and as such is a finite reflection group. In fact it turns out that most finite reflection groups are Weyl groups. Abstractly, Weyl groups are finite Coxeter groups, and are important examples of these.

<span class="mw-page-title-main">Quantum group</span> Algebraic construct of interest in theoretical physics

In mathematics and theoretical physics, the term quantum group denotes one of a few different kinds of noncommutative algebras with additional structure. These include Drinfeld–Jimbo type quantum groups, compact matrix quantum groups, and bicrossproduct quantum groups. Despite their name, they do not themselves have a natural group structure, though they are in some sense 'close' to a group.

E<sub>8</sub> (mathematics) 248-dimensional exceptional simple Lie group

In mathematics, E8 is any of several closely related exceptional simple Lie groups, linear algebraic groups or Lie algebras of dimension 248; the same notation is used for the corresponding root lattice, which has rank 8. The designation E8 comes from the Cartan–Killing classification of the complex simple Lie algebras, which fall into four infinite series labeled An, Bn, Cn, Dn, and five exceptional cases labeled G2, F4, E6, E7, and E8. The E8 algebra is the largest and most complicated of these exceptional cases.

<span class="mw-page-title-main">Compact group</span> Topological group with compact topology

In mathematics, a compact (topological) group is a topological group whose topology realizes it as a compact topological space. Compact groups are a natural generalization of finite groups with the discrete topology and have properties that carry over in significant fashion. Compact groups have a well-understood theory, in relation to group actions and representation theory.

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

In mathematics, a reductive group is a type of linear algebraic group over a field. One definition is that a connected linear algebraic group G over a perfect field is reductive if it has a representation that has a finite kernel and is a direct sum of irreducible representations. Reductive groups include some of the most important groups in mathematics, such as the general linear group GL(n) of invertible matrices, the special orthogonal group SO(n), and the symplectic group Sp(2n). Simple algebraic groups and (more generally) semisimple algebraic groups are reductive.

In mathematics, a Kac–Moody algebra is a Lie algebra, usually infinite-dimensional, that can be defined by generators and relations through a generalized Cartan matrix. These algebras form a generalization of finite-dimensional semisimple Lie algebras, and many properties related to the structure of a Lie algebra such as its root system, irreducible representations, and connection to flag manifolds have natural analogues in the Kac–Moody setting.

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

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

In mathematics and physics, specifically the study of field theory and partial differential equations, a Toda field theory, named after Morikazu Toda, is specified by a choice of Lie algebra and a specific Lagrangian.

In mathematics, an affine Lie algebra is an infinite-dimensional Lie algebra that is constructed in a canonical fashion out of a finite-dimensional simple Lie algebra. Given an affine Lie algebra, one can also form the associated affine Kac-Moody algebra, as described below. From a purely mathematical point of view, affine Lie algebras are interesting because their representation theory, like representation theory of finite-dimensional semisimple Lie algebras, is much better understood than that of general Kac–Moody algebras. As observed by Victor Kac, the character formula for representations of affine Lie algebras implies certain combinatorial identities, the Macdonald identities.

Verma modules, named after Daya-Nand Verma, are objects in the representation theory of Lie algebras, a branch of mathematics.

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

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

In mathematical group theory, the root datum of a connected split reductive algebraic group over a field is a generalization of a root system that determines the group up to isomorphism. They were introduced by Michel Demazure in SGA III, published in 1970.

In mathematics, the spin representations are particular projective representations of the orthogonal or special orthogonal groups in arbitrary dimension and signature. More precisely, they are two equivalent representations of the spin groups, which are double covers of the special orthogonal groups. They are usually studied over the real or complex numbers, but they can be defined over other fields.

In mathematics, the Dynkin index of a finite-dimensional highest-weight representation of a compact simple Lie algebra with highest weight is defined by

In algebra, the Nichols algebra of a braided vector space is a braided Hopf algebra which is denoted by and named after the mathematician Warren Nichols. It takes the role of quantum Borel part of a pointed Hopf algebra such as a quantum groups and their well known finite-dimensional truncations. Nichols algebras can immediately be used to write down new such quantum groups by using the Radford biproduct.

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

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

<span class="mw-page-title-main">Glossary of Lie groups and Lie algebras</span>

This is a glossary for the terminology applied in the mathematical theories of Lie groups and Lie algebras. For the topics in the representation theory of Lie groups and Lie algebras, see Glossary of representation theory. Because of the lack of other options, the glossary also includes some generalizations such as quantum group.

In abstract algebra, specifically the theory of Lie algebras, Serre's theorem states: given a root system , there exists a finite-dimensional semisimple Lie algebra whose root system is the given .

References

Further reading