Loading [MathJax]/jax/output/HTML-CSS/jax.js

Thursday, 28 February 2013

Simplicity of the projective special linear groups

Theorem If n2 then PSLn(q) is simple provided (n,q) is not (2,2) or (2,3).
proof: Consider PSLn(q) acting on Pn1(q) for n2 and the exceptions do not occur, then it is primitive since it's 2-transitive and by previous results SLn(q) is perfect, so since PSLn(q) is a quotient of that it's perfect too. Take dV# so that [d]Pn1(q) let A be the image of T(d) in PSLn(q) applying [???] and taking quotients we see that A is a normal abelian subgroup of the stabilizer PSLn(q)[d] and it's conjugates generate PSLn(q) thus the conditions of Iwasawa's lemma are satisfied.

Wednesday, 27 February 2013

M11

Let Δ={δ1,δ2,δ3,δ4,δ5} and let Ω be the unordered pairs of these. We will use the names:
  • 0: {δ1,δ2},  1: {δ1,δ3},  2: {δ1,δ4}
  • 3: {δ1,δ5},  4: {δ2,δ3},  5: {δ2,δ4}
  • 6: {δ2,δ5},  7: {δ3,δ4},  8: {δ3,δ5}
  • 9: {δ4,δ5}
Let G=AΔ (basically A5) this acts transitively on Ω!
  • a=(δ1δ2)(δ3δ5)=(16)(25)(34)(79)
  • b=(δ1δ2)(δ4δ5)=(14)(26)(35)(78)
  • g2=(δ2δ3)(δ4δ5)=(01)(23)(58)(67)
  • g3=(δ1δ4)(δ2δ3)=(07)(15)(39)(68)
you can clearly combine these to take 1 anywhere, which proves transitivity.

The stabilizer of 0 has size |G0|=|G|/|Ω|=60/10=6 (orbit stabilizer) so G0=a,b={1,a,b,ab,ba,aba=bab} and the G0 orbits are:
  • {0}
  • {1,2,3,4,5,6}
  • {7,8,9}
the double cosets corresponding to these suborbits are represented by 1,g2 and g3.

gap> a:=(1,2)(3,5);;b:=(1,2)(4,5);;
gap> DoubleCosets(Group((1,2,3,4,5),(1,2,3)),Group(a,b),Group(a,b));
[ DoubleCoset(Group( [ (1,2)(3,5), (1,2)(4,5) ] ),(),Group( 
    [ (1,2)(3,5), (1,2)(4,5) ] )), 
  DoubleCoset(Group( [ (1,2)(3,5), (1,2)(4,5) ] ),(2,3)(4,5),Group( 
    [ (1,2)(3,5), (1,2)(4,5) ] )), 
  DoubleCoset(Group( [ (1,2)(3,5), (1,2)(4,5) ] ),(1,3)(2,4),Group( 
    [ (1,2)(3,5), (1,2)(4,5) ] )) ]

so take Ω and set Ω+=Ω{} then choose x=(0)(23)(46)(89) then
  • x2=1G0
  • ax=b and (since x has order 2) bx=a so Gx0=G0.
  • gx2=(1)(23)(59)(47)=xg2GxG.
  • gx3=(7)(15)(28)(49)=xg3abGxG.
Thus L=G,x is a one point extension. It has order 660=11106 and is 2-transitive (hence a primitive permutation group). We saw that G (A5) - the stabilizer of in L - is simple so if L had a regular normal subgroup (with a transitive action on the 10+1 points, with all its stabilizers trivial so..) of order 11 which must (by Lagrange) be C11. |NS11(C11)|=110 (by counting cycles) which is strictly less than 660 so it can't be normal. Therefore by the proposition we have:

Theorem L is simple.

1 and x are (G,G)-double coset reps in L. Take ωΩ+ and form Ω=Ω+{ω}. y=(ω)(14)(25)(36) then
  • y2=1G=L
  • Gy=G since G=a,b,g2,g3 and ay=a, by=b, gy2=gb2, gy3=g3ab.
  • xy=yxabLyL
Therefore we have another one point extension M11=L,y, which is 3-transitive of order 7920 acting on 12 points: so it can't have any regular normal subgroup from the corollary as none of these groups have order 12) but since L is simple by the proposition we have:

Theorem M11 is simple.

From Jordan's theorem we see that we cannot perform another one point extension.

A_n is simple!

Lemma A5 is perfect.
proof: A5 is generated by (123) and (12345) both are commutators:
gap> a := (1,5,2);; b := (4,2,3);; a^(-1)*b^(-1)*a*b;
(1,2,3)
gap> a := (1,2,3)*(3,4,5);; b := (1,4,2)*(3,5,2);; a^(-1)*b^(-1)*a*b;
(1,2,3,4,5)

Theorem A5 is simple.
proof: The most basic proof using cycles directly can be found in Goodman.
proof: The conjugacy classes have sizes: 1, 15, 20, 12, 12. No sum of these that includes 1 is a divisor of 60 so there are no normal subgroups (which would necessarily be a union of conjugacy classes).
proof: A perfect group is not solvable, and every smaller group whose order divides |A5|=60 is solvable so A5 has no normal subgroups (else it would be solvable too!)

gap> List(AllSmallGroups(2), StructureDescription);    
[ "C2" ]
gap> List(AllSmallGroups(2^2), StructureDescription);
[ "C4", "C2 x C2" ]
gap> List(AllSmallGroups(2*3), StructureDescription);
[ "S3", "C6" ]
gap> List(AllSmallGroups(2*5), StructureDescription);
[ "D10", "C10" ]
gap> List(AllSmallGroups(2*3*5), StructureDescription);
[ "C5 x S3", "C3 x D10", "D30", "C30" ]
gap> List(AllSmallGroups(2^2*3), StructureDescription);  
[ "C3 : C4", "C12", "A4", "D12", "C6 x C2" ]
gap> List(AllSmallGroups(2^2*5), StructureDescription);
[ "C5 : C4", "C20", "C5 : C4", "D20", "C10 x C2" ] 

Theorem An is simple.
proof: Induction on n with base case 5. An is n2 transitive in the natural action (by the multiple-transitivity section) so for n>5 this action is (at least) 2-transitive so primitive (by primitivity section) which by the powerful corollary about transitivity with regular normal subgroups tells us that a regular normal subgroup would have to be C22 in the case of A6 and there isn't one otherwise but C22 doesn't have enough elements to be transitive on 6 points so it can't be regular -  so An has no regular normal subgroups: therefore by the proposition in that section it's simple.

Tuesday, 26 February 2013

Iwasawa's lemma

Lemma (Iwasawa) If (G,Ω) is a primitive permutation group with G perfect and for some αΩ, Gα has a normal abelian subgroup A whose conjugates generate G, then G is simple.
proof: Suppose 1NG, we will gradually show that N must be the whole group. By primitivity N is transitive on Ω and Gα is a maximal subgroup, so NGα and NGα=G. Any g may be written nx for some nN, xGα so Ag=Anx=AxAN and these conjugates cover the whole group so AN=G. Now G/NA/(AN) is abelian, but G is perfect so N=G.

Projective Spaces and Groups

We aim to "fix" the following two issues: GL and SL are not 2-transitive because they can't linearly dependent vectors to linearly independent ones. SL is not simple (even though it's perfect) because it has a center. Let V=Vn(q) and n2 throughout.

We define an equivalence relation R on V# by vRw iff v=λw for some nonzero λFq.

Definition We then have the projective space P(V) of projective vectors. Write Pn1(q).

For a subspace UV the set of equivalence classes (projective vectors) [U] (the image of U#) is a subspace of P(V) so it inherits geometric structure. The dimension of [U] is the dimension of U minus 1. A point is a class [v] for some vector v, and a line is the projective class of a 2D subspace.

Given gGL(V), vV# and λF#q we have (λv)g=λ(vg)[vg] so we can (well) define an action by [v]g=[vg]. In this way GL(V) and SL(V) act on P(V), but not faithfully.

Lemma Let G be GL(V) or SL(V), the kernel of the action of G on P(V) is Z(G).
proof: From the previous post we have seen what Z(G) is: scalar multiples of the identity. If gs=1 then clearly g acts trivially on P(V). Conversely if gGL(V) is in the kernel of the action then [vg]=[v] for all [v]P(V), then for every vector V we have vg=λvv for some λvF# and the proof concludes in the same way as before.

Definition The projective general linear group and projective special linear group are defined by PGL(V)=GL(V)/Z(GL(V)) and PSL(V)=SL(V)/Z(SL(V)). They have faithful action on P(V). (Note: PSL might not be a subgroup of PGL anymore).

We call [g]PGL or PSL the image (???) if gGL or SL if [v][g]=[vg] for vP(V). PGLn(q) and PSLn(q) act on Pn1(q) by the order calculations in the previous post we find |PGLn(q)|=qn(n1)/2ni=2(qi1) and |PSLn(q)|=qn(n1)/2(n,q1)ni=2(qi1). Thus |PGL2(q)|=(q+1)q(q1).

Proposition The permutation group PGL2(q) is sharply 3-transitive on P1(q).
proof: For g=(abcd)GL2(q) and vV2(q)#, vg=(aλ1+cλ2,bλ1+dλ2). These work out exactly as the mobius transformations when regarding P1(q) as Fq{}. We have already shown this one point extension is genrated by the mobius transforms!

Proposition Both PGL(V) and PSL(V) act 2-transitively on P(V).
proof: Given ([e1],[e2]), ([e1],[e2]) distinct then extend to a basis and find a map between them (why doesn't this work for all n-transitivity?)

Saturday, 23 February 2013

Transvections

Assume n2 throughout,

Definition A linear functional on V is a linear map from V to Fq. The set of these is the dual space V. Given fV write Vf for the kernel of f, if f0 then Vf is a subspace of dimension n1 (any such space of codimension 1 we call a hyperplane).

Lemma If you have f,fV with the same hyperplane then there exists λ such that f=λf.
proof: The result is clear if Vf=V assume not so f,f0. Take vVVf (i.e. not in the common hyperplane) so vf,vf0 and put λ=(vf)(vf)1 then fλfV and its kernel is contained in Vf,v=V so it equals zero.

Definition A linear automorphism τGL(V) is called a transvection with direction dV# if τ fixes d and vτv is a scalar multiple of v for all v.

Clearly 1 is a transvection of any direction.

Lemma If τ is a transvection with direction d then the vectorspace fix(τ) is a hyperplane containing d.
proof: Define f:VFq by (vf)d=vτv i.e. f is that scalar multiple which a transvection defines. Since τ is linear fV. The result is clear for τ=1 and if not f0 so dfix(τ)=Vf.

Corollary Any transvection can be written as τf,d mapping v to v+(vf)d for some fV, d(Vf)#.

Lemma For f,fV, dV#, gGL(V) we have
  • τgf,g=τg1f,dg
  • τf,dτf,d=τf+f,d
proof: first we need to check (dg)(g1f)=0 and d(f+f)=0 for the notation to be meaningful. Now the results follow by computation.

Definition For a direction dV# set T(d)={τf,d|fV,dVf} and T the union over all directions (all transvections).

Proposition T# is a single conjugacy class in GL(V) and lies in SL(V). If n3 then T# is a single conjugacy class in SL(V).
proof: By the calculation lemma previous, we know that T# is closed under conjugation. Let τf,d,τf,dT# and write e1=d,e1=d then take bases e1,,en1 and e1,,en1 of the hyperplanes Vf,Vf, choose en,en such that enf=enf=1. Now we have bases for V so there is a GL map g from one to the other.
For i<n we have ei(g1f)=eif=0 so Vg1f is the space spanned by e1,,en1=Vf so they are scalar multiples of each other, let λ be such that f=λ(g1f) and 1=enf=λen(g1f)=λenf=λ so since dg=d it follows that τgf,d=τf,d!
If they are all conjugate they all have the same determinant δ, now det(τf,dτf,d)=det(τf+f,d) so δ2=δ proves they lie in SL.
For n3 we can use the μ trick as before to get them in SL.

Proposition If dV# then T(d) is an abelian normal subgroup of the stabilizer SL(V)d; T(d) are all conjugates in SL(V).
proof: Certainly elements of T(d) stabilize d, by the computational lemma before we see that it is an abelian group (commutative and closed, therefore has identity and inverses). If gSL(V) with dg=d then T(d)g=T(d) (by the previous) so T(d)SL(V)d. Given d,dV# by the transitivity lemma (that requires dimension > 1) exists gSL(V) which takes d to d then T(d)g=T(d).

Proposition The set T generates SL(V).
proof: Elementary matrices.

Definition A group if perfect if G=G. This is equivalent to there being no nontrivial abelian quotients: Clearly if G/[G,G]1 then G[G,G]. Conversely if G/NA={Ng} then we always have Ngg=Ngg i.e. there is some n such that gg=ngg. So every commutator [g,g] is an element of N.

Proposition If n2 the group SLn(q) is perfect provided (n,q) isn't (2,2) or (2,3).
proof: We just need to show that each τT# is a commutator since that set generates our group. If τ has direction d take some σT(d)# not equal to τ1, then στT(d)# so there is a gSLn(q) that conjugates στ=σg, whence τ=σ1g1σg=[σ,g].
For n=2 we will use 2×2 matrices directly, in some basis τ=(1γ01) (nonzero γ), now for any nonzero λ and μFq we have (λ00λ1)(1μ01)(λ100λ)(1μ01)=(1μ(λ21)01) so for q>3 take λ0,1,1 then λ210 so let μ=γ(λ21)1.

Proposition
  • Z(GL(V))={λ1|λF#q}
  • Z(SL(V))={λ1|λF#q,λn=1}
proof: Certainly these are contained in the center, suppose gZ(GL(V)) or Z(SL(V)) then for each nonzero vector dV# take a transvection τf,dT(d), this should be unaffected by conjugation but we know τgf,d=τg1f,dg so dg=λdd. Then λdd+λvd=(d+v)g=λd+v(d+v)=λd+vd+λd+vv so λd=λv=λ and g=λI. For SL we also require det(g)=λn=1. This tells us that Z(GLn(q))F#q and Z(SLn(q))C(n,q1).

Finite Fields and Finite Vector Spaces

Definition An Affine transformation of Fq is a map fa,b taking λ to aλ+b where aF#q is nonzero. The group of such maps is called A(Fq).

Proposition A(Fq) is sharply 2-transitive of order q(q1).
proof: Let α,β distinct, the system of equations α=a0+b, β=a1+b has a unique solution.

Corollary By a general lemma about sharply 2-transitive groups this group must have a regular characteristic subgroup, this group is {f1,b|bFq}.

Proposition A(Fq) has a one point extension which is sharply 3-transitive of degree q+1.
proof: This is a straightforward application of the one point extension theorem, adjoin to Fq and define x on Fq{} to swap 0 and and invert all other elements λx=λ1. Clearly x2=1. Let G0={fa,0aF#q} and note fxa,0 fixes 0 and while for λF#q we fxa,0=fa1,0 so Gx0=G0. Finally a system for the double cosets is given by just 1 and any other element e.g. f=f1,1 will do (so λf=1λ and f2=1). We see that xf acts on the ,1,0 by cycling them and for the remaining elements λ(xf)3=1111λ=λ and (apparently...) xf=fxGxG so we have a one point extension.

Definition Vn(q) is the n-dimensional vector space over Fq, clearly |Vn(q)|=qn.

Definition If V  is a vector space then a linear automorphism of V is a bijective linear map VV. The group of these is called the general linear group GL(V) or GLn(q) when V=Vn(q).

Definition The special linear group SL(V) of linear automorphisms of determinant 1.

Lemma If V is a vector space over Fq then SL(V) is a normal subgroup of GL(V) and the index is q1: |GL(V):SL(V)|=q1.
proof: SL is just the kernel of the surjective determinant map from GL to F×q. As a consequence GL/SLF#q so |GL:SL|=q1.

Lemma The group GL(V) acts transitively on V# and if the dimension of V is >1 the same is true of SL(V).
proof: Take two nonzero vectors e1,f1 then to get a map between them choose bases e1,,en and f1,,fn this gives gGL(V) mapping between them. If n>1, since we don't necessarily have det(g)=1 let det(g)=μ and replace en by μ1en. Now g mapping between these bases has determinant 1. (did I get this right?)

Proposition |GLn(q)|=qn(n1)/2ni=1(qi1) and |SLn(q)|=qn(n1)/2ni=2(qi1).
proof: GLn(q) acts regularly on ordered bases of Vn(q), so the size of GLn(q) is equal to the number of ordered bases: qn1 choices for the first element, and having chosen e1,,ei (which spans a qi sized space) already there are qnqi choices for the next. The size of SL comes from the lemma before the previous.

Saturday, 16 February 2013

Sharply t-transitive groups

Sharply transitive groups are the smallest possible transitive groups, given tN a t-transitive group G is called sharply t-transitive if Gα1α2αt=1 for distinct αs. Equivalently, there is exactly one group element that takes (α1,,αt) to any other triple of distinct symbols.

Theorem If G is a t-transitive group of degree n (i.e. it acts on n symbols) then it is sharply t-transitive iff |G|=n(n1)(nt1).
proof: For t=1 this is the regular action. For t+1 any Gα will be t-transitive and |G|=n|Gα| since the orbit of α is the whole of Ω, conversely every stabilizer Gα will have the same order - and there are n of them so each |Gα|=|G|/n is t1-transitive so G is t-transitive.

Example Sn is sharply n-transitive in the natural actions, it is also n1-sharply transitive. Note this isn't a contradiction from the group order result because |G|=|G|1. Intuitively what's happening is if we choose exactly where n1 symbols go, then it's already decided where the last one must go.

Proposition If G is sharply 2-transitive of degree n then G has a regular characteristic subgroup which is an elementary abelian p-group for some prime p (so n is power of p).
proof: Given G, let K be the union of the fixed point free elements and 1. We will show it is a group. Since for any distinct α,βΩ, GαGβ=Gαβ=1 so G is the disjoint union KαΩG#α, |G#α|=|G|n1 (using n=|G|/|Gα| from the orbit-stab. theorem) so |K|=n. Take α,βΩ distinct and choose kK# such that αkα. As G is 2-transitive there exists gGα with (αh)g=β then hg has the same order and fixedpoints as h i.e. none, therefore its in K, and αkg=β so K is a transitive "set". For any αΩ the map KΩ given by kαk is surjective and hence bijective. Now take take distinct x,yK we know αxαy so αxy1α for all α, so xy1K. This proves K a subgroup!
Given kK# all its cycles on Ω must all have the same length since otherwise some non-identity power of k would have fixed points. So the order of k divides n. Similarly for elements of G#α the same argument tells us all its cycles on Ω{α} have the same length, so its order divides n1. Hence K consists of the elements of G of order dividing n - a property preserved by conjugation - therefore it's a characteristic subgroup. By the corollary from the section on regular normal subgroups it is an elementary abelian p-group.

Thursday, 14 February 2013

One-point extensions

Reversing the idea of a point stabilizer Let (G,Ω) be a transitive permutation group, take a point ωΩ and form Ω+=Ω{ω}, extend the action by ωg=ω for all gG:

Definition A one-point extensionof (G,Ω) is a transitive permutation group (G+,Ω+) with (G+)ω=G

By the stabilizer-orbit theorem |G+|=|G|(|Ω|+1). If G+ is t-transitive then G is (t1)-transitive.

Example Sn and An have one point extensions Sn+1 and An+1.
Non-example D8 doesn't have one by Sylow theory.

Take αΩ, we know the rank r of G equal to the number of double cosets in G. For g1,,grG we have a complete representation of the double coset system iff αg1,,αgr is a complete set of representatives of Gα-orbits. wlog take g1=1, if (G+,Ω+) is a one point extension of (G,Ω) then it is 2-transitive so it has a primitive action and hence the point stabilizer G is a maximal subgroup of G+: for any xG+G we must have x,G=G+. We can wlog choose x to interchange α and ω.

Theorem Let (G,Ω) be a transitive permutation group of rank r for αΩ, let g1=1,g2,gr be a complete set of representatives of the double coset system. Take ωΩ and form Ω+=Ω{ω}. Take xS(Ω+) (so some permutation from the symmetric group) with αx=ω, ωx=α and set G+=x,G then (G+,Ω+) is a one point extension iff:
  1. x2Gα
  2. (Gα)x=Gα
  3. gxiGxG for all i>1.
proof: () (1) definition of x (2) easy (3) As xG the 2-transitivity of G+ means that G+=GGxG (any permutation that can't be done with g can be done with x in the middle), for 2ir we have giGGα so ωgxi=αgixαx=ω thus gxi(G+)ω=G so gxiGxG. () Set H=GGxGas a set, we will show its a group HSΩ+, to show it's a group we need HH=H=H1. Since G1=G and (gxg)1=g1x(x2)1g1GxG . By (2) xGαx=Gα so for i2 we have xGαgiGαx=Gα(xgix)Gα=Gαx2gxiGαGxG so xGx=xGαxri=2xGαgiGαxGαGxGH. Now HH=GGxGGxGxGHGHG=H, we have shown it's a group! Now G+=G,x=H and for all g,g take ω(gxg)=ωxg=αgω so (G+)ω=G.

Tuesday, 12 February 2013

Multiple-transitivity

Consider actions of rank 2 (meaning that there are two suborbits of Gα), Ω{α} forms a single Gα orbit (because one its other orbit is αGα={α}). G acts on Ω2 non-transitively because g(β,β)=(γ,γ) but if we define the diagonal Delta={(β,β)Ω} this is a single orbit due to transitivity and so Ω2Δ is the interesting part:

Lemma The action of G on Ω is of rank 2 iff Ω2Δ is a single orbit.
proof: If the action has rank 2 then let (β1,β2),(γ1,γ2) lie off the diagonal and pick x,yG such that αx=β1, αy=γ1 then neither of β2x1 and γy1 are equal to α (otherwise β1=β2 or γ1=γ2) so there exists hGα which maps one to the other β2x1h=γ2y1. Let g=x1hy and compute (β1,β2)g=(γ1,γ2).
In the other direction if  Ω2Δ is a single orbit the action is at least 2 (since we can fix any one element β in the first component and map any other element γ not equal to beta to any other element not equal to beta). So suppose the rank were larger than 2, then pick β,γ in different Gα orbits of Ω{α} and take (α,β),(α,γ)Ω2Δ there's clearly no way to map from one to the other.

We can generalize this to Ωt for any natural t|Ω|. Again Δ={(α,α,)} is a single orbit, but the interesting part is Ω(t)={(α1,α2,)|αiαj}.

Definition The action of G on Ω is t-transitive if the induced action on Ω(t) is transitive. This is equivalent to saying it can simultaneously map any t distinct points to any other t distinct points.

Lemma In terms of cosets, a group action is 2-transitive iff G=GαGαgGα for any gGα.
proof:  We saw previous that double cosets correspond to suborbits, write Gα=Gα1Gα to see this is the same as rank 2.

Lemma The action of G on Ω is t-transitive iff the action of Gα on Ω{α} is (t1)-transitive.
proof: write this out.

Corollary If G acts t-transitively on Ω and |Ω|=n then |G| is divisible by n(n1)(nt+1). WHICH THEOREM DOES THIS DEPEND ON? CHECK OTHER BOOK

Theorem In the natural action Sn is n-transitive while An is n2 transitive and not n1 transitive.
proof: This is obvious from the fact Sn contains every permutation. For An we use induction: it clearly holds for A3. For n3 the stabilizer of any point of An is An1 so by the lemma we complete the induction.

Regular Normal Subgroups and Semidirect Products

Let G act transitively on Ω throughout. We shall be concerned with the case where G has a regular normal subgroup K, meaning that KG acts transitively on Ω and the stabilizer Kα is trivial for every α (this is the regular action, it's equivalent to the action on the cosets 1g).

Lemma If G acts transitively on Ω with regular normal subgroup K choose αΩ the action of Gα on K by conjugation is equivalent to its action on Ω with 1K corresponding to αΩ.
proof: Define θ:KΩ by kαk. This is a bijection since K is regular. We just need to show that the two actions are equivalent: (kθ)g=αkg=αgkg=αkg=kgθ.

If G is multiply transitive and has a regular normal subgroup K the subgroup Gα is transitive on Ω{α} and hence K#=K{1} as conjugation is an automorphism of a group, this means in particular that the group K ???? automorphisms ?????? ff

Lemma An automorphism preserves the order of a group element.
proof: Let θ:GG be an automorphism and g have order n, then (θg)i1 for any 0<i<n since if it did θ(gi)=1 implies that gi=1.

Proposition Let K be a group then
  1. If AutK acts transitively on K# then K is elementary abelian
  2. If AutK is 2-transitive on K# then either KCs2 or C3
  3. If AutK is 3-transitive on K# then KC22
  4. AutK is not 4-transitive.
 proof: (1) Since automorphisms preserve the order of elements all elements must have the same order, and that order must be a prime otherwise there would be elements with smaller order, by Cauchy's theorem we can say this is a p-group and so Z(K) is non-trivial. In fact we know Z(K)charG (because the center is preserved by conjugation) now any automorphism of K fixes a characteristic subgroup like Z(K) so for transitivity to be possible Z(K)=K. This implies K is abelian and therefore a direct product of cyclic groups, which we know must all be Cp!
(2) Since 2-transitivity implies primitivity of the action we define a relation on K# that is an AutK-congruence: kRk if k=k or k1=k. Clearly this is an equivalence relation preserved by the action hence by primitivity it must be the equality relation (implying k1=k so the group is Cs2) or the entire relation (implying every two non-identity elements are equal or inverse so the group will be C3).
(3) If the group is further 3-transitive, then it's impossible that it's C3: That's too small! Take some kK# and consider the stabilizer (AutK)k which is 2-transitive hence primitive so again let's define a congruence on it k1Rk2 if k2=k1 or k2=kk1 - meaning that k1,k2 "differ by k" - this can't be the equality relation since that would imply k=1 so it's the universal relation implying KC22.
(3)  This is impossible again because the group is too small.

Corollary Suppose G acts t-transitively on Ω with a regular normal subgroup K so |Ω|=|K| then
  1. If t=2 then KCsp
  2. If t=3 then KCs2 or C3
  3. If t=4 then KC22
  4. t<5
Proposition If (G,Ω) is a primitive permutation group such that some Gα is simple (therefore all Gα are simple) then either G is simple or has a regular normal subgroup.
proof: Suppose G is not simple, then there's a K such that 1KG. In that case KGαGα is simple so KGα=1 or KGα=Gα but K is transitive by a corollary on primitive permutation groups and Gα is a maximal subgroup by another corollary so we cannot have GαK (certainly Gα<K can't happen, and if G=Gα then \alpha gGα=αg so Gα=Gβ==G contradiction). Then KGα=1 so K is regular because it's a normal subgroup which we showed acts transitively and also it has the regular action because we saw that Kα=1.

Definition Let H,K be groups with a homomorphism θ:HAutK. For kK, hH write kh short for khθ. Then the semidirect product KH is defined as the group K×H with binary operation:
(k1,h1)(k2,h2)=(k1kh112,h1h2).
proof: prove this is a group

Define subgroups of the semidirect product ˜K={(k,1)} and ˜H={(1,h)} then ˜KKH, ˜HKH, ˜K˜H=1,,˜K˜H=KH and ˜KKH but we don't know that ˜HKH. This is an "external" construction of the semidirect product. We also have an internal one. An extension G of K by H is such that KG, G/KH i.e. the short exact sequence 1KGH1 in which case we write G=K.H, the . is "neutral" meaning that it doesn't really tell you how the group has been extended.

Definition if we have a s.e.s. like above with KG, KH=1 and KH=G we call G a split extension of K by H and write G=K:H.

Theorem If G is a split ext then define θ:HAutK by khθ=kh. Then GKH.

Corollary If G is transitive on Ω with a regular normal subgroup then G=K:Gα.

Saturday, 9 February 2013

Primitivity

The section is about decomposing group actions, assume Ω transitive.

Definition A non-empty set ΓΩ is called a block if for all gG either Γg=Γ or ΓgΓ={}. If Γ is a block then the set Σ=Σ(Γ)={ΓggG} of all translates of Γ is a block system.

Lemma A block system partitions Ω. proof: Let αΓ and βΩ then β=αg for some g so βΓg and the blocks cover Ω. If ΓgΓh is non-empty then Γgh1=Γ so Γg=Γh.

Definition A G-congruence is an equivalence relation R on Ω such that αRβ implies αgRβg.

Definition If R is a G-congruence then the R-equiv classes form a block system and conversely if Γ is a block we define R by αRβ iff α,βΓg.
proof: easy

The trivial G-congruences are the equality relation and the one induced by the block Ω.

Definition The (transitive) action on Ω is called imprimitive if there is a non-trivial G-congruence. If there are no non-trivial G-congruences an action is primitive.

Proposition Let αΩ, write B(α) for the set of blocks containing α and S(α) for the set of subgroups containing Gα.
  • There are mutually inverse bijections Ψ:B(α)S(α) and Φ:S(α)B(α) defined by ΓΨ=GΓ, HΦ=αH.
  • For Γ,ΓB(α) we have ΓΓ iff ΓΨΓΨ.
proof: long.

Corollary The action of G on Ω is primitive iff each Gα is a maximal subgroup.

Proposition If the action of G on Ω is 2-transitive then it is primitive.
proof: If the action is 2-trans take αΩ. Suppose Γ is a block containing α with |Γ|>1. Take βΓ{α} then for any βΩ{α} there exists gGα with βg=β. As αΓgΓ we must have Γg=Γ so βΓ and hence Γ=Ω. So α lies in no nontrivial block.

Note: The converse is not true, you can have imprimitive actions that aren't 2-transitive.

Proposition If NG the set of N-orbits in Ω is a block system.
proof:  Let Γ be an N-orbit, If gG with ΓgΓ{} let αΓgΓ so α=βg with βΓ then Γ=αN=βN. So Γg=βNg=βgN=αN=Γ.

Corollary If G is a primitive permutation group (no non-identity elements fix all elements of Ω) and 1NG then N acts transitively.
proof: Since 1N there is an N orbit of size > 1 so by the previous theorem it gives a block system, by primitivity it's the whole of Ω, so N must act transitively.

Thursday, 7 February 2013

Suborbits and double-cosets

Throughout we will assume that G acts transitively on Ω. The idea motivating this section is that since each orbit is conjugate, what freedom remains when we pick some fixed α and consider Gα (The stabilizer of α)?

Since Gα is not just a subset of G but in fact a subgroup, the action of G on Ω induces an action of Gα on Ω.

Definition The orbits of Gα on Ω (by this induced action) are called suborbits, their sizes are called subdegrees and the rank is how many there are.


Recall the orbit/stabilizer theorem:
  • αG={αgΩ|gG}
  • Gα={gG|αg=α}
  • Since αg=αg iff gg1Gα iff Gαg=Gαg the bijection αgGαg is well defined, it also respects the group action.
By transitivity every element βΩ can be written as αg for some g, in fact {β}=αGαg so elements of Ω are in bijection with the right cosets {Gαg|gG}.

Suborbits of Gα are the orbits βGα which are thus in bijection with the double cosets GαgGα.

We call these (Gα,Gα)-double cosets and they partition the group. The size of a double coset divided by |Gα| gives the subdegree (similar to Lagrange's theorem).

Lemma The rank of the action of G is 1|G|gG|fix(g)|2.
proof: Apply Burnside's lemma to the action of Gα on Ω to get |Ω||G|gGα|fix(g)| since |Gα|=|G||Ω| now sum over all α to get 1|G|αΩgGα|fix(g)|=1|G|gGαfix(g)|fix(g)|.

Definition Let α,βΩ. The 2-point stabilizer Gα,β is GαGβ.
Definition The pointwise stabilizer of a set of points ΓΩ, G(Γ) is bigcapγΓGγ.
Definition The setwise stabilizer GΓ={gG|Γg=Γ}.

Lemma Given βΩ the subdegree corresponding to β is |Gα:Gα,β|.
proof: In the action of Gα on the suborbit βGα the stab. of β is GαGβ. The result follows from orb/stab theorem.

Definition If Gα=1 the action is regular and has rank |Ω|

Tuesday, 5 February 2013

Permutation groups: Burnsides Lemma and the Fixed point free theorem

Definition If G acts on Ω and gG then the fixed point set fix(g) is {αΩ|αg=α}. If fix(g)={} then g is FPF (fixed point free).

Lemma ("Burnsides lemma" except it was actually proved by Frobenius) If G acts on Ω then the number of orbits is 1|G|gG|fix(g)|.
proof: Consider the set S={(α,g)Ω×G|αg=α}. We shall count it in two different ways:
First given α, consider the number of ways it can occur as the first component of the pair: that's just |Gα| the size of its stabilizer, from the fact |Gαg|=|Ggα|=|Gα| proved in the previous post we get that the contribution to S from the orbit is αG is |Gα||αG|=|G| therefore |S|=|G|number of orbits.
Secondly given g, consider the number of ways it can occur as the second component of the pair: that's just fix(g).
Putting these together gives the formula.

Corollary If G acts transitively on Ω (and |Ω|>1) some element of g is fixed point free. proof: There is only 1 orbit, so the average number (over G) of elements fixed is 1.. but the identity has |G| fixed points - every element!

Permutation groups: actions, orbit and stabilizer

Notation backwards notation: (123)(1324)=(14)

Notation (24)(12453)=(25), in general σπ sends iπiσπ so you can compute these by hand.

 Definition A permutation group is a finite set Ω and a group of permutations (that is, bijections ΩΩ). We'll write SΩ for the group of all permutations on a set. The degree of a permutation group is the cardinality |Ω|.

Notation Let αΩ then αg for the image of α through g. Group homomorphisms are written after elements too.

Definition If G acts on two sets Ω and Ω then the actions are equivalent if there is a bijection θ between them such that gG,αΩ,(αθ)g=(αg)θ.

Definition We write αG={αg|gG} for the orbit of G containing α.

Definition We write Gα={gG|αg=α} for the stabilizer of α.

Definition A group action is transitive if Ω is a single orbit.

Definition The kernel of the action is G(Ω)={gG|αΩ,αg=α}. Clearly GS(Ω) defined by g(ααg) is a homomorphism with kernel G(Ω), this G/G(Ω) can be identified with its image in SΩ giving a permutation group (G/G(Ω),Ω).

Definition An action is said to be faithful if the kernel is trivial.

Definition The core of H in G is HG=xGHx, this is the largest normal subgroup of G contained in H.

Examples
(i) If (G,Ω) is a permutation gp and G acts on Ω faithfully this is called the natural action.
(ii) If HG we have an action of G on the set (G:H) of right cosets of H in G by (Hx)g=H(xg). This is called a coset action and if H=1 it is the regular action. The kernel of the regular action is the core: We have gG(Ω) iff x,Hxg=Hx iff x,xgx1H iff x,gHx.
(iii) The action of G on (G:H) given by HxΩ is clearly transitive. The stabilizer of H is H while that of Hx is {gG|Hxg=Hx}=Hx.

Lemma If G acts on Ω, given gG and αΩ we have Gαg=Ggα.
proof: xGαg iff αgx=αg iff αgxg1=α iff  xGgα.

Theorem (orbit stabilizer) If G acts on Ω and αΩ then the actions of G on αG and (G:Gα) are equivalent.
proof: given g,hG, αg=αh iff gh1Gα iff Gαg=Gαh. Thus θ=αgGαg is a well defined bijection, and it respects the action since ((αg)θ)x=Gαgx=((αg)x)θ.

Corollary Any transitive action is equivalent to a coset action.

Corollary |G|=|Gα||αG|. (because |αG|=|(G:Gα)| by the theorem and |G|=|Gα||(G:Gα)| is a triviality just write it down).

Saturday, 2 February 2013

Sylow bases - Structural character of Solvable groups

Notation Let p denote the set of all primes other than p.

Let G=pam with p|m, a Sylow p-subgroup P has order pa whereas a Hall p-subgroup H of G has order m. HP=1 and so |G|=|H||P| and G=HP.

Lemma If H,KG and |G:H|,|G:K| are coprime then |G:HK|=|G:H||G:K| (note, this doesn't assume normality).
proof: Put a=|G:H|, b=|G:K|, c=|G:HK|. Since |G:HK|=|G:H||H:HK| we have a|c and similarly b|c thus ab|c. In the other direction we may define a map (HK)x(Hx,Kx):{cosets of HK}{cosets of H}×{cosets of K} this is well defined since (HK)x=(HK)y iff xy1H and K iff Hx=Hy and Kx=Ky. Since this map is injective cab.

Definition Let G be a group whose order factors into powers of primes p1,,pk. A Sylow basis for G is a collection of Sylow subgroups P1,,Pk such that for all i, Pi is a Sylow pi-subgroup and for all i,j PiPj=PjPi.

Any product of a subset of the Sylow basis will give a Hall π-subgroup and you can get a Hall π-subgroup for any π this way.

Definition Two Sylow bases P and B are said to be conjugate when there exists a single g such that forall i, Pi=Bgi.

Theorem (Hall - 1937) If G is solvable then it has a Sylow basis and any two such bases are conjugate.
proof: (Existence) Let G be a solvable group and write |G|=pa11pakk, let S={1,,k}. For each iS let Qi be a Hall pi-subgroup so |G:Qi|=paii by the previous theorem of Hall. Given any TS the intersection tTQt is a Hall π-subgroup (for the appropriate π={pj|jST}). In particular Pi=tiQt is a Hall {pi}-subgroup (A Sylow pi-subgroup) of G. To see that this gives a basis take i,jS not equal and PiPj=1 by coprimality, therefore we have (considered as sets) |PiPj|=|Pi||Pj|=|PjPi|. Write T=S{pi,pj} then tTQt is a group that contains Pi and Pj hence PiPj and PjPi but |tTQt|=paiipajj as it's a Hall π-group!
(Uniqueness up to conjugacy) Let B1,,Bk be any other Sylow basis for G with (by renumbering) |Bi|=|Pi|. For tinS form the Hall pt-subgroup Ct=ΠitBi. Instead of showing each Bi is conjugate to Pi, we show that each Qi is conjugate to Ci then deduce that. Let d be the number of t such that CtQt, and we prove by induction on d that there exists some g such that for every t, Ct=Qgt: the base case d=0 is trivial. Assume (by renumbering if necessary) that Ct=Qt for all t>d. Write H=t>dQt by the uniqueness up to conjugacy of Hall π-subgroups for solvable groups there exists x such that Cd=Qxd, since Qd contains each Pi except Pd - which lies in H - G=QdH and so x=gh for some gQd, hH. Now Cd=Qghd=Qhd and for t<d Ct=Qt=Qht so there are at most d1 values of t (the t<d) such that CtQt therefore we have by induction zG such that t, Ct=(Qht)z=Qhzt.
Finally for all i, Bi=tiCi=tiQgi=Pgi.

Hall's Theorem

Definition For π a set of primes, a π-group is a group whose orders is a product of prime powers taken from π.

Definition A Hall π-subgroup H is a π-subgroup of G where the index |G:H| is a product of primes not from π.

Theorem (Hall - 1928) Let G be a solvable group and π a set of primes then
  1. G has a Hall π-subgroup.
  2. Any two Hall π-subgroups are conjugate.
  3. Any π-subgroup of G is contained in a Hall π-subgroup.
proof: We will show by induction on |G| that:
  • (a) G has a π-subgroup H
  • (b) Any π-subgroup L of G is contained in a conjugate of H.
 (b) implies (2) and (a,b) together imply (1) since any Sylow p-subgroup is a π-subgroup (when pπ) we can ensure that the order |G:H| has no powers of p in it.

The base case is trivial. Let |G|=mn where m is a product of powers of primes from π and n contains no primes from π. The case m=1 is trivial. We split the proof into two cases:

(Case A.a) There is a minimal normal subgroup M with order dividing m. By earlier results M is elementary abelian and a p-group for some p. Write |M|=pa so that |G/M|=m1n where m1=mpa. By induction G/M has a subgroup - which by isomorphism theorems, is of the form H/M - with order m1; then H is a subgroup of G with order m this proves part (a).
(Case A.b) Given a π-subgroup L we have LMG (since L is a π-group and M is normal in G todo: make sure this is right) and by isomorphism theorems LM/ML/(LM) so LM/M is a π-subgroup of G/M. By induction it lies in some conjugate of H/M, say LM/M(H/M)Mg=Hg/M and so we have LLMHg gving (b).

(Case B) In this case there does not exist a minimal normal subgroup of order dividing m. Let M be an minimal normal subgroup then it's elementary abelian and it must be a q-group for some prime q|n (i.e. q is a prime not in π). Write |M|=qb so |G/M|=mnqb=mn1 with n1=nqb and we split this into two more cases:

(Case Bi.a) Assume n1>1, then by induction G/M has a subgroup K/M of order m (because the index |G/M:K/M| is n1), |K|=mqb<mn so by induction again K has a π-subgroup H of order m as required for (a)
(Case Bi.b) Given L as before LM/M is a π-subgroup of G/M so by induction it lies in some conjugate of K/M, say LM/M(K/M)Mg=Kg/M then Lg1 is a pi-subgroup of K so by induction it lies in some conjugate of H, say Lg1Hk whence LHkg1 as required for (b).

(Case Bii.) The final case is when G has no minimal normal subgroups of order dividing m and n1=1 (refer back to Case B for definition of n1). Let N/M be a minimal normal subgroup of G/M then we know it is an elementary abelian p-group for some p|m, say |N/M|=pa then NG and |N|=paqb.
Let P be a Sylow p-subgroup of N and write H=NG(P) by the Frattini argument G=HN since N=PM and PH so G=HPM=HM. Let J=HM then JHM=G by minimality of M we must have J=1 or J=M:
(Case Bii.J=M) This case is easily disposed of. HM=M so MH so G=HM=H=NG(P) so PG (by normalizer facts) and some subgroup of P is then a minimal normal subgroup of of G whose order does divide m, contradiction.
(Case Bii.J=1.a) In this case |HM|=1 and thus mqb=|G|=|H||M|=qb|H| so |H|=m giving (a).
(Case Bii.J=1.b) Given L we have (and we can form this because M is a q group and L is a p-group) LM=LMG=LMHM=(LMH)M (by the ABC lemma). Now LMH is a π-subgroup of LM and |LM:LMH|=|(LMH)M|/|LMH|=|M|/|LMHM|=|M| therefore it is in fact a Hall π-subgroup so if LM<G we are done by induction. It only remains to deal with the case where LM=G: in this case since LM=1 by coprime orders, |G|=|L||M| implying |L|=m. Since MN (since MN from the fact the quotient makes sense) LN=G and thus |LN|=|L||N|/|G|=pa so by Sylow's theorem LN is conjugate to P (L=Pn for some nN). LNL since NG so LNG(LN)=NG(Pn)=NG(P)n=Hn giving (b).

Thus we have (a) and (b) for all groups, this implies (i) and (ii). For (iii) let K be any Hall π-subgroup of G by (b) we know that KHg for some gG, since |H|=|K|, K=Hg.