[go: up one dir, main page]

Jump to content

Coxeter group: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
Citation bot (talk | contribs)
m Citations: [184] added: postscript, doi, issue. Tweaked: journal. Unified citation types. Rjwilmsi
Stwitzel (talk | contribs)
m Davies -> Davis
Line 3: Line 3:
Coxeter groups find applications in many areas of mathematics. Examples of finite Coxeter groups include the [[symmetry group]]s of [[regular polytope]]s, and the [[Weyl group]]s of [[simple Lie algebra]]s. Examples of infinite Coxeter groups include the [[triangle group]]s corresponding to [[Tessellation#Regular and irregular tessellations|regular tessellation]]s of the [[Euclidean plane]] and the [[hyperbolic plane]], and the Weyl groups of infinite-dimensional [[Kac–Moody algebra]]s.
Coxeter groups find applications in many areas of mathematics. Examples of finite Coxeter groups include the [[symmetry group]]s of [[regular polytope]]s, and the [[Weyl group]]s of [[simple Lie algebra]]s. Examples of infinite Coxeter groups include the [[triangle group]]s corresponding to [[Tessellation#Regular and irregular tessellations|regular tessellation]]s of the [[Euclidean plane]] and the [[hyperbolic plane]], and the Weyl groups of infinite-dimensional [[Kac–Moody algebra]]s.


Standard references include {{Harv|Humphreys|1990}} and {{Harv|Davies|2007}}.
Standard references include {{Harv|Humphreys|1990}} and {{Harv|Davis|2007}}.


==Definition==
==Definition==

Revision as of 20:00, 20 September 2010

In mathematics, a Coxeter group, named after H.S.M. Coxeter, is an abstract group that admits a formal description in terms of mirror symmetries. Indeed, the finite Coxeter groups are precisely the finite Euclidean reflection groups; the symmetry groups of regular polyhedra are an example. However, not all Coxeter groups are finite, and not all can be described in terms of symmetries and Euclidean reflections. Coxeter groups were introduced (Coxeter 1934) as abstractions of reflection groups, and finite Coxeter groups were classified in (Coxeter 1935).

Coxeter groups find applications in many areas of mathematics. Examples of finite Coxeter groups include the symmetry groups of regular polytopes, and the Weyl groups of simple Lie algebras. Examples of infinite Coxeter groups include the triangle groups corresponding to regular tessellations of the Euclidean plane and the hyperbolic plane, and the Weyl groups of infinite-dimensional Kac–Moody algebras.

Standard references include (Humphreys 1990) and (Davis 2007).

Definition

Formally, a Coxeter group can be defined as a group with the presentation

where and for . The condition means no relation of the form should be imposed.

A number of conclusions can be drawn immediately from the above definition.

  • The relation mi i = 1 means that (riri )1 = (ri )2 = 1 for all i ; the generators are involutions.
  • If mi j = 2, then the generators ri and rj commute. This follows by observing that
xx = yy = 1,
together with
xyxy = 1
implies that
xy = x(xyxy)y = (xx)yx(yy) = yx.
Alternatively, since the generators are involutions, , so , and thus is equal to the commutator.
  • In order to avoid redundancy among the relations, it is necessary to assume that mi j=mj i. This follows by observing that
yy = 1,
together with
(xy)m = 1
implies that
(yx)m = (yx)myy = y(xy)my = yy = 1.
Alternatively, and are conjugate elements, as .

The Coxeter matrix is the n×n, symmetric matrix with entries mi j. Indeed, every symmetric matrix with positive integer and ∞ entries and with 1's on the diagonal serves to define a Coxeter group. The Coxeter matrix can be conveniently encoded by a Coxeter graph, as per the following rules.

  • The vertices of the graph are labelled by generator subscripts.
  • Vertices i and j are connected if and only if mi j ≥ 3.
  • An edge is labelled with the value of mi j whenever it is 4 or greater.

In particular, two generators commute if and only if they are not connected by an edge. Furthermore, if a Coxeter graph has two or more connected components, the associated group is the direct product of the groups associated to the individual components. Thus the disjoint union of Coxeter graphs yields a direct sum of Coxeter matrices and a direct product of Coxeter groups.

An example

The graph in which vertices 1 through n are placed in a row with each vertex connected by an unlabelled edge to its immediate neighbors gives rise to the symmetric group Sn+1; the generators correspond to the transpositions (1 2), (2 3), ... (n n+1). Two non-consecutive transpositions always commute, while (k k+1) (k+1 k+2) gives the 3-cycle (k k+2 k+1). Of course this only shows that Sn+1 is a quotient group of the Coxeter group described by the graph, but it is not too difficult to check that equality holds.

Connection with reflection groups

Coxeter groups are deeply connected with reflection groups. Simply put, Coxeter groups are abstract groups (given via a presentation), while reflection groups are concrete groups (given as subgroups of linear groups or various generalizations). Coxeter groups grew out of the study of reflection groups — they are an abstraction: a reflection group is a subgroup of a linear group generated by reflections (which have order 2), while a Coxeter group is an abstract group generated by involutions (elements of order 2, abstracting from reflections), and whose relations have a certain form (, corresponding to hyperplanes meeting at an angle of , with being of order k abstracting from a rotation by ).

The abstract group of a reflection group is a Coxeter group, while conversely a reflection group can be seen as a linear representation of a Coxeter group. For finite reflection groups, this yields an exact correspondence: every finite Coxeter group admits a faithful representation as a finite reflection group of some Euclidean space. For infinite Coxeter groups, however, a Coxeter group may not admit a representation as a reflection group.

Historically, (Coxeter 1934) proved that every reflection group is a Coxeter group (i.e., has a presentation where all relations are of the form or ), and indeed this paper introduced the notion of a Coxeter group, while (Coxeter 1935) proved that every finite Coxeter group had a representation as a reflection group, and classified finite Coxeter groups.

Finite Coxeter groups

Coxeter graphs of the finite Coxeter groups.

Classification

The finite Coxeter groups were classified in (Coxeter 1935), in terms of Coxeter–Dynkin diagrams; they are all represented by reflection groups of finite-dimensional Euclidean spaces.

The finite Coxeter groups consist of three one-parameter families of increasing rank one one-parameter family of dimension two, and six exceptional groups: and

Weyl groups

Many, but not all of these, are Weyl groups, and every Weyl group can be realized as a Coxeter group. The Weyl groups are the families and and the exceptions and denoted in Weyl group notation as The non-Weyl groups are the exceptions and and the family except where this coincides with one of the Weyl groups (namely and ).

This can be proven by comparing the restrictions on (undirected) Dynkin diagrams with the restrictions on Coxeter diagrams of finite groups:[note 1] Dynkin diagrams have the additional restriction that the only permitted edge labels are 2, 3, 4, and 6, which yields the above. Geometrically, this corresponds to the crystallographic restriction theorem, and the fact that excluded polytopes do not fill space or tile the plane – for the dodecahedron (dually, icosahedron) does not fill space; for the 120-cell (dually, 600-cell) does not fill space; for a p-gon does not tile the plane except for or 6 (the triangular, square, and hexagonal tilings, respectively).

Note further that the (directed) Dynkin diagrams Bn and Cn give rise to the same Weyl group (hence Coxeter group), because they differ as directed graphs, but agree as undirected graphs – direction matters for root systems but not for the Weyl group; this corresponds to the hypercube and cross-polytope being different regular polytopes but having the same symmetry group.

Properties

Some properties of the finite Coxeter groups are given in the following table:

Group
symbol
Alternate
symbol
Bracket notation Rank Order Related polytopes Coxeter-Dynkin diagram
An An [3n] n (n + 1)! n-simplex ...
Bn = Cn Cn [4,3n-1] n 2n n! n-hypercube / n-cross-polytope ...
Dn Bn [3n-3,1,1] n 2n−1 n! n-demihypercube ...
E6 E6 [32,2,1] 6 51840 221 polytope
E7 E7 [33,2,1] 7 2903040 321 polytope
E8 E8 [34,2,1] 8 696729600 421 polytope
F4 F4 [3,4,3] 4 1152 24-cell
H3 G3 [3,5] 3 120 icosahedron / dodecahedron
H4 G4 [3,3,5] 4 14400 120-cell / 600-cell
I2(p) D2p [p] 2 2p p-gon

Symmetry groups of regular polytopes

All symmetry groups of regular polytopes are finite Coxeter groups, and note that dual polytopes have the same symmetry group.

There are three series of regular polytopes in all dimensions. The symmetry group of a regular n-simplex is the symmetric group Sn+1, also known as the Coxeter group of type An. The symmetry group of the n-cube and its dual, the n-cross-polytope is BCn, and is known as the hyperoctahedral group.

The exceptional regular polytopes in dimensions two, three, and four, correspond to other Coxeter groups. In two dimensions, the dihedral groups, which are the symmetry groups of regular polygons, form the series I2(p). In three dimensions, the symmetry group of the regular dodecahedron and its dual, the regular icosahedron, is H3, known as the full icosahedral group. In four dimensions, there are three special regular polytopes, the 24-cell, the 120-cell, and the 600-cell. The first has symmetry group F4, while the other two are dual and have symmetry group H4.

The Coxeter groups of type Dn, E6, E7, and E8 are the symmetry groups of certain semiregular polytopes.

Affine Weyl groups

The affine Weyl groups form a second important series of Coxeter groups. These are not finite themselves, but each contains a normal abelian subgroup such that the corresponding quotient group is finite. In each case, the quotient group is itself a Weyl group, and the Coxeter graph is obtained from the Coxeter graph of the Weyl group by adding an additional vertex and one or two additional edges. For example, for n ≥ 2, the graph consisting of n+1 vertices in a circle is obtained from An in this way, and the corresponding Coxeter group is the affine Weyl group of An. For n = 2, this can be pictured as the symmetry group of the standard tiling of the plane by equilateral triangles.

A list of the affine Coxeter groups follows:

Group
symbol
Alternate
symbol
Bracket notation Related uniform tessellation(s) Coxeter-Dynkin diagram
A~n-1 Pn [3[n]] Simplex-rectified-simplex honeycomb
A~2:Triangular tiling
A~3:Tetrahedral-octahedral honeycomb
...
B~n-1 Sn [4,3n-3,31,1] Demihypercubic honeycomb ...
C~n-1 Rn [4,3n-2,4] Hypercubic honeycomb ...
D~n-1 Qn [ 31,1,3n-4,31,1] Demihypercubic honeycomb ...
E~6 T7 [32,2,2] 2 22 honeycomb
E~7 T8 [33,3,1] 3 31 honeycomb
1 33 honeycomb
E~8 T9 [35,2,1] 5 21 honeycomb
2 51 honeycomb
1 52 honeycomb
F~4 U5 [3,4,3,3] Hexadecachoric honeycomb and
Icositetrachoric honeycomb
H~2 V3 [6,3] Hexagonal tiling and
Triangular tiling
I~1 W2 [∞] apeirogon

Note the subscript is one less than the number of nodes in each case, since each of these groups was obtained by adding a node to a finite group's graph.

Hyperbolic Coxeter groups

There are infinitely many hyperbolic Coxeter groups describing reflection groups in hyperbolic space, notably including the hyperbolic triangle groups.

Partial orders

A choice of reflection generators gives rise to a length function l on a Coxeter group, namely the minimum number of uses of generators required to express a group element; this is precisely the length in the word metric in the Cayley graph. An expression for v using l(v) generators is a reduced word. For example, the permutation (13) in S3 has two reduced words, (12)(23)(12) and (23)(12)(23). The function defines a map generalizing the sign map for the symmetric group.

Using reduced words one may define three partial orders on the Coxeter group, the weak order, the absolute order and the Bruhat order (named for François Bruhat). An element v exceeds an element u in the Bruhat order if some (or equivalently, any) reduced word for v contains a reduced word for u as a substring, where some letters (in any position) are dropped. In the weak order, v ≥ u if some reduced word for v contains a reduced word for u as an initial segment. Indeed, the word length makes this into a graded poset. The Hasse diagrams corresponding to these orders are objects of study, and are related to the Cayley graph determined by the generators. The absolute order is defined analogously to the weak order, but with generating set/alphabet consisting of all conjugates of the Coxeter generators.

For example, the permutation (1 2 3) in S3 has only one reduced word, (12)(23), so covers (12) and (23) in the Bruhat order but only covers (12) in the weak order.

Homology

Since a Coxeter group W is generated by finitely many elements of order 2, its abelianization is an elementary abelian 2-group, i.e. it is isomorphic to the direct sum of several copies of the cyclic group Z2. This may be restated in terms of the first homology group of W.

The Schur multiplier M(W) (related to the second homology) was computed in (Ihara & Yokonuma 1965) for finite reflection groups and in (Yokonuma 1965) for affine reflection groups, with a more unified account given in (Howlett 1988). In all cases, the Schur multiplier is also an elementary abelian 2-group. For each infinite family {Wn} of finite or affine Weyl groups, the rank of M(W) stabilizes as n goes to infinity.

See also

Notes

  1. ^ Formally, the Coxeter graph can be obtained from the Dynkin diagram by discarding the direction of the edges, and replacing every double edge with an edge labelled 4 and every triple edge by an edge labelled 6.

References

  1. ^ Brink and Howlett (1993), "A finiteness property and an automatic structure for Coxeter groups", Mathematische Annalen, Springer Berlin / Heidelberg, ISSN 0025-5831.
  • Coxeter, H.S.M. (1934), "Discrete groups generated by reflections", Ann. Of Math., 35: 588–621, doi:10.2307/1968753
  • Coxeter, H.S.M. (1935), "The complete enumeration of finite groups of the form ", J. London Math. Soc., 10: 21–25
  • Davis, Michael W. (2007), The Geometry and Topology of Coxeter Groups (PDF), ISBN 978-0-691-13138-2
  • Larry C Grove and Clark T. Benson, Finite Reflection Groups, Graduate texts in mathematics, vol. 99, Springer, (1985)
  • James E. Humphreys, Reflection Groups and Coxeter Groups, Cambridge studies in advanced mathematics, 29 (1990)
  • Richard Kane, Reflection Groups and Invariant Theory, CMS Books in Mathematics, Springer (2001)
  • Anders Björner and Francesco Brenti, Combinatorics of Coxeter Groups, Graduate Texts in Mathematics, vol. 231, Springer, (2005)
  • Howard Hiller, Geometry of Coxeter groups. Research Notes in Mathematics, 54. Pitman (Advanced Publishing Program), Boston, Mass.-London, 1982. iv+213 pp. ISBN 0-273-08517-4
  • Nicolas Bourbaki, Lie Groups and Lie Algebras: Chapter 4-6, Elements of Mathematics, Springer (2002). ISBN 978-3540426509
  • Howlett, Robert B. (1988), "On the Schur Multipliers of Coxeter Groups", Journal of the London Mathematical Society, 2, 38 (2): 263–276, doi:10.1112/jlms/s2-38.2.263
  • E. B. Vinberg, Absence of crystallographic groups of reflections in Lobachevski spaces of large dimension, Trudy Moskov. Mat. Obshch. 47 (1984)