3-transposition group

Last updated

In mathematical group theory, a 3-transposition group is a group generated by a conjugacy class of involutions, called the 3-transpositions, such that the product of any two involutions from the conjugacy class has order at most 3.

Contents

They were first studied by BerndFischer  ( 1964 , 1970 , 1971 ) who discovered the three Fischer groups as examples of 3-transposition groups.

History

Fischer (1964) first studied 3-transposition groups in the special case when the product of any two distinct 3-transpositions has order 3. He showed that a finite group with this property is solvable, and has a (nilpotent) 3-group of index 2. Manin (1986) used these groups to construct examples of non-abelian CH-quasigroups and to describe the structure of commutative Moufang loops of exponent 3.

Fischer's theorem

Suppose that G is a group that is generated by a conjugacy class D of 3-transpositions and such that the 2 and 3 cores O2(G) and O3(G) are both contained in the center Z(G) of G. Then Fischer (1971) proved that up to isomorphism G/Z(G) is one of the following groups and D is the image of the given conjugacy class:

The missing cases with n small above either do not satisfy the condition about 2 and 3 cores or have exceptional isomorphisms to other groups on the list.

Important examples

The group Sn has order n! and for n>1 has a subgroup An of index 2 that is simple if n>4.

The symmetric group Sn is a 3-transposition group for all n>1. The 3-transpositions are the elements that exchange two points, and leaving each of the remaining points fixed. These elements are the transpositions (in the usual sense) of Sn. (For n=6 there is a second class of 3-transpositions, namely the class of the elements of S6 which are products of 3 disjoint transpositions.)

The symplectic group Sp2n(2) has order

It is a 3-transposition group for all n≥1. It is simple if n>2, while for n=1 it is S3, and for n=2 it is S6 with a simple subgroup of index 2, namely A6. The 3-transpositions are of the form xx+(x,v)v for non-zero v.

The special unitary group SUn(2) has order

The projective special unitary group PSUn(2) is the quotient of the special unitary group SUn(2) by the subgroup M of all the scalar linear transformations in SUn(2). The subgroup M is the center of SUn(2). Also, M has order gcd(3,n).

The group PSUn(2) is simple if n>3, while for n=2 it is S3 and for n=3 it has the structure 32:Q8 (Q8 = quaternion group).

Both SUn(2) and PSUn(2) are 3-transposition groups for n=2 and for all n≥4. The 3-transpositions of SUn(2) for n=2 or n≥4 are of the form xx+(x,v)v for non-zero vectors v of zero norm. The 3-transpositions of PSUn(2) for n=2 or n≥4 are the images of the 3-transpositions of SUn(2) under the natural quotient map from SUn(2) to PSUn(2)=SUn(2)/M.

The orthogonal group O2n±(2) has order

(Over fields of characteristic 2, orthogonal group in odd dimensions are isomorphic to symplectic groups.) It has an index 2 subgroup (sometimes denoted by Ω2n±(2)), which is simple if n>2.

The group O2nμ(2) is a 3-transposition group for all n>2 and μ=±1. The 3-transpositions are of the form xx+(x,v)v for vectors v such that Q(v)=1, where Q is the underlying quadratic form for the orthogonal group.

The orthogonal groups On±(3) are the automorphism groups of quadratic forms Q over the field of 3 elements such that the discriminant of the bilinear form (a,b)=Q(a+b)−Q(a)−Q(b) is ±1. The group Onμ,σ(3), where μ and σ are signs, is the subgroup of Onμ(3) generated by reflections with respect to vectors v with Q(v)=+1 if σ is +, and is the subgroup of Onμ(3) generated by reflections with respect to vectors v with Q(v)=-1 if σ is −.

For μ=±1 and σ=±1, let POnμ,σ(3)=Onμ,σ(3)/Z, where Z is the group of all scalar linear transformations in Onμ,σ(3). If n>3, then Z is the center of Onμ,σ(3).

For μ=±1, let Ωnμ(3) be the derived subgroup of Onμ(3). Let PΩnμ(3)= Ωnμ(3)/X, where X is the group of all scalar linear transformations in Ωnμ(3). If n>2, then X is the center of Ωnμ(3).

If n=2m+1 is odd the two orthogonal groups On±(3) are isomorphic and have order

and On+,+(3) ≅ On−,−(3) (center order 1 for n>3), and On−,+(3) ≅ On+,−(3) (center order 2 for n>3), because the two quadratic forms are scalar multiples of each other, up to linear equivalence.

If n=2m is even the two orthogonal groups On±(3) have orders

and On+,+(3) ≅ On+,−(3), and On−,+(3) ≅ On−,−(3), because the two classes of transpositions are exchanged by an element of the general orthogonal group that multiplies the quadratic form by a scalar. If n=2m, m>1 and m is even, then the centre of On+,+(3) ≅ On+,−(3) has order 2, and the centre of On−,+(3) ≅ On−,−(3) has order 1. If n=2m, m>2 and m is odd, then the centre of On+,+(3) ≅ On+,−(3) has order 1, and the centre of On−,+(3) ≅ On−,−(3) has order 2.

If n>3, and μ=±1 and σ=±1, the group Onμ,σ(3) is a 3-transposition group. The 3-transpositions of the group Onμ,σ(3) are of the form xx−(x,v)v/Q(v)=x+(x, v)/(v,v) for vectors v with Q(v)=σ, where Q is the underlying quadratic form of Onμ(3).

If n>4, and μ=±1 and σ=±1, then Onμ,σ(3) has index 2 in the orthogonal group Onμ(3). The group Onμ,σ(3) has a subgroup of index 2, namely Ωnμ(3), which is simple modulo their centers (which have orders 1 or 2). In other words, PΩnμ(3) is simple.

If n>4 is odd, and (μ,σ)=(+,+) or (−,−), then Onμ,+(3) and POnμ,+(3) are both isomorphic to SOnμ(3)=Ωnμ(3):2, where SOnμ(3) is the special orthogonal group of the underlying quadratic form Q. Also, Ωnμ(3) is isomorphic to PΩnμ(3), and is also non-abelian and simple.

If n>4 is odd, and (μ,σ)=(+,−) or (−,+), then Onμ,+(3) is isomorphic to Ωnμ(3)×2, and Onμ,+(3) is isomorphic to Ωnμ(3). Also, Ωnμ(3) is isomorphic to PΩnμ(3), and is also non-abelian and simple.

If n>5 is even, and μ=±1 and σ=±1, then Onμ,+(3) has the form Ωnμ(3):2, and POnμ,+(3) has the form PΩnμ(3):2. Also, PΩnμ(3) is non-abelian and simple.

Fi22 has order 217.39.52.7.11.13 = 64561751654400 and is simple.

Fi23 has order 218.313.52.7.11.13.17.23 = 4089470473293004800 and is simple.

Fi24 has order 222.316.52.73.11.13.17.23.29 and has a simple subgroup of index 2, namely Fi24'.

Isomorphisms and solvable cases

There are numerous degenerate (solvable) cases and isomorphisms between 3-transposition groups of small degree as follows ( Aschbacher 1997 , p.46):

Solvable groups

The following groups do not appear in the conclusion of Fisher's theorem as they are solvable (with order a power of 2 times a power of 3).

has order 1.
has order 2, and it is a 3-transposition group.
is elementary abelian of order 4, and it is not a 3-transposition group.
has order 6, and it is a 3-transposition group.
is elementary abelian of order 8, and it is not a 3-transposition group.
has order 24, and it is a 3-transposition group.
has order 72, and it is not a 3-transposition group, where Q8 denotes the quaternion group.
has order 72, and it is not a 3-transposition group.
has order 216, and it is not a 3-transposition group, where 31+2 denotes the extraspecial group of order 27 and exponent 3, and Q8 denotes the quaternion group.
has order 288, and it is not a 3-transposition group.
has order 576, where * denotes the non-direct central product, and it is not a 3-transposition group.

Isomorphisms

There are several further isomorphisms involving groups in the conclusion of Fischer's theorem as follows. This list also identifies the Weyl groups of ADE Dynkin diagrams, which are all 3-transposition groups except W(D2)=22, with groups on Fischer's list (W stands for Weyl group).

has order 120, and the group is a 3-transposition group.
has order 720 (and 2 classes of 3-transpositions), and the group is a 3-transposition group.
has order 40320, and the group is a 3-transposition group.
has order 51840, and the group is a 3-transposition group.
has order 25920, and the group is a 3-transposition group.
has order 2903040, and the group is a 3-transposition group.
has order 69672960, and the group is a 3-transposition group.
for all s≥1, and the group is a 3-transposition group if s≥2.
for all s≥1, and the group is a 3-transposition group for all s≥1.
for all s≥0, and the group is a 3-transposition group for all s≥0.
for all s≥0, and the group is a 3-transposition group if s≥1.
for all m≥0, and the group is a 3-transposition group if m≥1.
for all m≥0, and the group is a 3-transposition group if m=0 or m≥2.
for all n≥1, and the group is a 3-transposition group for all n≥1.
for all n≥2, and the group is a 3-transposition group if n≥3.

Proof

The idea of the proof is as follows. Suppose that D is the class of 3-transpositions in G, and dD, and let H be the subgroup generated by the set Dd of elements of D commuting with d. Then Dd is a set of 3-transpositions of H, so the 3-transposition groups can be classified by induction on the order by finding all possibilities for G given any 3-transposition group H. For simplicity assume that the derived group of G is perfect (this condition is satisfied by all but the two groups involving triality automorphisms.)

3-transpositions and graph theory

It is fruitful to treat 3-transpositions as vertices of a graph. Join the pairs that do not commute, i. e. have a product of order 3. The graph is connected unless the group has a direct product decomposition. The graphs corresponding to the smallest symmetric groups are familiar graphs. The 3 transpositions of S3 form a triangle. The 6 transpositions of S4 form an octahedron. The 10 transpositions of S5 form the complement of the Petersen graph.

The symmetric group Sn can be generated by n–1 transpositions: (1 2), (2 3), ..., (n−1 n) and the graph of this generating set is a straight line. It embodies sufficient relations to define the group Sn. [1]

Related Research Articles

<span class="mw-page-title-main">Pauli matrices</span> Matrices important in quantum mechanics and the study of spin

In mathematical physics and mathematics, the Pauli matrices are a set of three 2 × 2 complex matrices which are Hermitian, involutory and unitary. Usually indicated by the Greek letter sigma, they are occasionally denoted by tau when used in connection with isospin symmetries.

<span class="mw-page-title-main">Symmetric group</span> Type of group in abstract algebra

In abstract algebra, the symmetric group defined over any set is the group whose elements are all the bijections from the set to itself, and whose group operation is the composition of functions. In particular, the finite symmetric group defined over a finite set of symbols consists of the permutations that can be performed on the symbols. Since there are such permutation operations, the order of the symmetric group is .

<span class="mw-page-title-main">Clifford algebra</span> Algebra based on a vector space with a quadratic form

In mathematics, a Clifford algebra is an algebra generated by a vector space with a quadratic form, and is a unital associative algebra. As K-algebras, they generalize the real numbers, complex numbers, quaternions and several other hypercomplex number systems. The theory of Clifford algebras is intimately connected with the theory of quadratic forms and orthogonal transformations. Clifford algebras have important applications in a variety of fields including geometry, theoretical physics and digital image processing. They are named after the English mathematician William Kingdon Clifford.

<span class="mw-page-title-main">Semidirect product</span> Operation in group theory

In mathematics, specifically in group theory, the concept of a semidirect product is a generalization of a direct product. There are two closely related concepts of semidirect product:

In linear algebra, an orthogonal matrix, or orthonormal matrix, is a real square matrix whose columns and rows are orthonormal vectors.

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

In group theory, the quaternion group Q8 (sometimes just denoted by Q) is a non-abelian group of order eight, isomorphic to the eight-element subset of the quaternions under multiplication. It is given by the group presentation

<span class="mw-page-title-main">Orthogonal group</span> Type of group in mathematics

In mathematics, the orthogonal group in dimension n, denoted O(n), is the group of distance-preserving transformations of a Euclidean space of dimension n that preserve a fixed point, where the group operation is given by composing transformations. The orthogonal group is sometimes called the general orthogonal group, by analogy with the general linear group. Equivalently, it is the group of n×n orthogonal matrices, where the group operation is given by matrix multiplication. The orthogonal group is an algebraic group and a Lie group. It is compact.

<span class="mw-page-title-main">Unitary group</span> Group of unitary matrices

In mathematics, the unitary group of degree n, denoted U(n), is the group of n × n unitary matrices, with the group operation of matrix multiplication. The unitary group is a subgroup of the general linear group GL(n, C). Hyperorthogonal group is an archaic name for the unitary group, especially over finite fields. For the group of unitary matrices with determinant 1, see Special unitary group.

<span class="mw-page-title-main">Lorentz group</span> Lie group of Lorentz transformations

In physics and mathematics, the Lorentz group is the group of all Lorentz transformations of Minkowski spacetime, the classical and quantum setting for all (non-gravitational) physical phenomena. The Lorentz group is named for the Dutch physicist Hendrik Lorentz.

In mathematics, particularly in matrix theory, a permutation matrix is a square binary matrix that has exactly one entry of 1 in each row and each column and 0s elsewhere. Each such matrix, say P, represents a permutation of m elements and, when used to multiply another matrix, say A, results in permuting the rows or columns of the matrix A.

In abstract algebra and number theory, Kummer theory provides a description of certain types of field extensions involving the adjunction of nth roots of elements of the base field. The theory was originally developed by Ernst Eduard Kummer around the 1840s in his pioneering work on Fermat's Last Theorem. The main statements do not depend on the nature of the field – apart from its characteristic, which should not divide the integer n – and therefore belong to abstract algebra. The theory of cyclic extensions of the field K when the characteristic of K does divide n is called Artin–Schreier theory.

In mathematics, a maximal compact subgroupK of a topological group G is a subgroup K that is a compact space, in the subspace topology, and maximal amongst such subgroups.

<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 projective geometry and linear algebra, the projective orthogonal group PO is the induced action of the orthogonal group of a quadratic space V = (V,Q) on the associated projective space P(V). Explicitly, the projective orthogonal group is the quotient group

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

In mathematics, the classical groups are defined as the special linear groups over the reals R, the complex numbers C and the quaternions H together with special automorphism groups of symmetric or skew-symmetric bilinear forms and Hermitian or skew-Hermitian sesquilinear forms defined on real, complex and quaternionic finite-dimensional vector spaces. Of these, the complex classical Lie groups are four infinite families of Lie groups that together with the exceptional groups exhaust the classification of simple Lie groups. The compact classical groups are compact real forms of the complex classical groups. The finite analogues of the classical groups are the classical groups of Lie type. The term "classical group" was coined by Hermann Weyl, it being the title of his 1939 monograph The Classical Groups.

In mathematics, Macdonald polynomialsPλ(x; t,q) are a family of orthogonal symmetric polynomials in several variables, introduced by Macdonald in 1987. He later introduced a non-symmetric generalization in 1995. Macdonald originally associated his polynomials with weights λ of finite root systems and used just one variable t, but later realized that it is more natural to associate them with affine root systems rather than finite root systems, in which case the variable t can be replaced by several different variables t=(t1,...,tk), one for each of the k orbits of roots in the affine root system. The Macdonald polynomials are polynomials in n variables x=(x1,...,xn), where n is the rank of the affine root system. They generalize many other families of orthogonal polynomials, such as Jack polynomials and Hall–Littlewood polynomials and Askey–Wilson polynomials, which in turn include most of the named 1-variable orthogonal polynomials as special cases. Koornwinder polynomials are Macdonald polynomials of certain non-reduced root systems. They have deep relationships with affine Hecke algebras and Hilbert schemes, which were used to prove several conjectures made by Macdonald about them.

In group theory, a branch of mathematics, the automorphisms and outer automorphisms of the symmetric groups and alternating groups are both standard examples of these automorphisms, and objects of study in their own right, particularly the exceptional outer automorphism of S6, the symmetric group on 6 elements.

<span class="mw-page-title-main">Hyperoctahedral group</span> Group of symmetries of an n-dimensional hypercube

In mathematics, a hyperoctahedral group is an important type of group that can be realized as the group of symmetries of a hypercube or of a cross-polytope. It was named by Alfred Young in 1930. Groups of this type are identified by a parameter n, the dimension of the hypercube.

In mathematical physics, higher-dimensional gamma matrices generalize to arbitrary dimension the four-dimensional Gamma matrices of Dirac, which are a mainstay of relativistic quantum mechanics. They are utilized in relativistically invariant wave equations for fermions in arbitrary space-time dimensions, notably in string theory and supergravity. The Weyl–Brauer matrices provide an explicit construction of higher-dimensional gamma matrices for Weyl spinors. Gamma matrices also appear in generic settings in Riemannian geometry, particularly when a spin structure can be defined.

In mathematics, symmetric cones, sometimes called domains of positivity, are open convex self-dual cones in Euclidean space which have a transitive group of symmetries, i.e. invertible operators that take the cone onto itself. By the Koecher–Vinberg theorem these correspond to the cone of squares in finite-dimensional real Euclidean Jordan algebras, originally studied and classified by Jordan, von Neumann & Wigner (1934). The tube domain associated with a symmetric cone is a noncompact Hermitian symmetric space of tube type. All the algebraic and geometric structures associated with the symmetric space can be expressed naturally in terms of the Jordan algebra. The other irreducible Hermitian symmetric spaces of noncompact type correspond to Siegel domains of the second kind. These can be described in terms of more complicated structures called Jordan triple systems, which generalize Jordan algebras without identity.

References

  1. Dickson, L. E. (2003) [1900], Linear Groups: With an Exposition of the Galois Field Theory, p. 287, ISBN   978-0-486-49548-4