Signalizer functor

(Learn how and when to remove this message)

In mathematics, in the area of abstract algebra, a signalizer functor is a mapping from a potential finite subgroup to the centralizers of the nontrivial elements of an abelian group. The signalizer functor theorem provides the conditions under which the source of such a functor is in fact a subgroup.

The signalizer functor was first defined by Daniel Gorenstein.[1] George Glauberman proved the Solvable Signalizer Functor Theorem for solvable groups[2] and Patrick McBride proved it for general groups.[3][4] Results concerning signalizer functors play a major role in the classification of finite simple groups.

Definition

Let A be a non-cyclic elementary abelian p-subgroup of the finite group G. An A-signalizer functor on G (or simply a signalizer functor when A and G are clear) is a mapping θ from the set of nonidentity elements of A to the set of A-invariant p′-subgroups of G satisfying the following properties:

The second condition above is called the balance condition. If the subgroups θ ( a ) {\displaystyle \theta (a)} are all solvable, then the signalizer functor θ {\displaystyle \theta } itself is said to be solvable.

Solvable signalizer functor theorem

Given θ , {\displaystyle \theta ,} certain additional, relatively mild, assumptions allow one to prove that the subgroup W = θ ( a ) a A , a 1 {\displaystyle W=\langle \theta (a)\mid a\in A,a\neq 1\rangle } of G {\displaystyle G} generated by the subgroups θ ( a ) {\displaystyle \theta (a)} is in fact a p {\displaystyle p'} -subgroup.

The Solvable Signalizer Functor Theorem proved by Glauberman states that this will be the case if θ {\displaystyle \theta } is solvable and A {\displaystyle A} has at least three generators.[2] The theorem also states that under these assumptions, W {\displaystyle W} itself will be solvable.

Several weaker versions of the theorem were proven before Glauberman's proof was published. Gorenstein proved it under the stronger assumption that A {\displaystyle A} had rank at least 5.[1] David Goldschmidt proved it under the assumption that A {\displaystyle A} had rank at least 4 or was a 2-group of rank at least 3.[5][6] Helmut Bender gave a simple proof for 2-groups using the ZJ theorem,[7] and Paul Flavell gave a proof in a similar spirit for all primes.[8] Glauberman gave the definitive result for solvable signalizer functors.[2] Using the classification of finite simple groups, McBride showed that W {\displaystyle W} is a p {\displaystyle p'} -group without the assumption that θ {\displaystyle \theta } is solvable.[3][4]

Completeness

The terminology of completeness is often used in discussions of signalizer functors. Let θ {\displaystyle \theta } be a signalizer functor as above, and consider the set И of all A {\displaystyle A} -invariant p {\displaystyle p'} -subgroups H {\displaystyle H} of G {\displaystyle G} satisfying the following condition:

For example, the subgroups θ ( a ) {\displaystyle \theta (a)} belong to И as a result of the balance condition of θ.

The signalizer functor θ {\displaystyle \theta } is said to be complete if И has a unique maximal element when ordered by containment. In this case, the unique maximal element can be shown to coincide with W {\displaystyle W} above, and W {\displaystyle W} is called the completion of θ {\displaystyle \theta } . If θ {\displaystyle \theta } is complete, and W {\displaystyle W} turns out to be solvable, then θ {\displaystyle \theta } is said to be solvably complete.

Thus, the Solvable Signalizer Functor Theorem can be rephrased by saying that if A {\displaystyle A} has at least three generators, then every solvable A {\displaystyle A} -signalizer functor on G {\displaystyle G} is solvably complete.

Examples of signalizer functors

The easiest way to obtain a signalizer functor is to start with an A {\displaystyle A} -invariant p {\displaystyle p'} -subgroup M {\displaystyle M} of G , {\displaystyle G,} and define θ ( a ) = M C G ( a ) {\displaystyle \theta (a)=M\cap C_{G}(a)} for all nonidentity a A . {\displaystyle a\in A.} However, it is generally more practical to begin with θ {\displaystyle \theta } and use it to construct the A {\displaystyle A} -invariant p {\displaystyle p'} -group.

The simplest signalizer functor used in practice is θ ( a ) = O p ( C G ( a ) ) . {\displaystyle \theta (a)=O_{p'}(C_{G}(a)).}

As defined above, θ ( a ) {\displaystyle \theta (a)} is indeed an A {\displaystyle A} -invariant p {\displaystyle p'} -subgroup of G {\displaystyle G} , because A {\displaystyle A} is abelian. However, some additional assumptions are needed to show that this θ {\displaystyle \theta } satisfies the balance condition. One sufficient criterion is that for each nonidentity a A , {\displaystyle a\in A,} the group C G ( a ) {\displaystyle C_{G}(a)} is solvable (or p {\displaystyle p} -solvable or even p {\displaystyle p} -constrained).

Verifying the balance condition for this θ {\displaystyle \theta } under this assumption can be done using Thompson's P × Q {\displaystyle P\times Q} -lemma.

Coprime action

To obtain a better understanding of signalizer functors, it is essential to know the following general fact about finite groups:

This fact can be proven using the Schur–Zassenhaus theorem to show that for each prime q {\displaystyle q} dividing the order of X , {\displaystyle X,} the group X {\displaystyle X} has an E {\displaystyle E} -invariant Sylow q {\displaystyle q} -subgroup. This reduces to the case where X {\displaystyle X} is a q {\displaystyle q} -group. Then an argument by induction on the order of X {\displaystyle X} reduces the statement further to the case where X {\displaystyle X} is elementary abelian with E {\displaystyle E} acting irreducibly. This forces the group E / C E ( X ) {\displaystyle E/C_{E}(X)} to be cyclic, and the result follows. [9][10]

This fact is used in both the proof and applications of the Solvable Signalizer Functor Theorem.

For example, one useful result is that it implies that if θ {\displaystyle \theta } is complete, then its completion is the group W {\displaystyle W} defined above.

Normal completion

Another result that follows from the fact above is that the completion of a signalizer functor is often normal in G {\displaystyle G} :

Let θ {\displaystyle \theta } be a complete A {\displaystyle A} -signalizer functor on G {\displaystyle G} .

Let B {\displaystyle B} be a noncyclic subgroup of A . {\displaystyle A.} Then the coprime action fact together with the balance condition imply that W = θ ( a ) a A , a 1 = θ ( b ) b B , b 1 . {\displaystyle W=\langle \theta (a)\mid a\in A,a\neq 1\rangle =\langle \theta (b)\mid b\in B,b\neq 1\rangle .}

To see this, observe that because θ ( a ) {\displaystyle \theta (a)} is B-invariant, θ ( a ) = θ ( a ) C G ( b ) b B , b 1 θ ( b ) b B , b 1 . {\displaystyle \theta (a)=\langle \theta (a)\cap C_{G}(b)\mid b\in B,b\neq 1\rangle \subseteq \langle \theta (b)\mid b\in B,b\neq 1\rangle .}

The equality above uses the coprime action fact, and the containment uses the balance condition. Now, it is often the case that θ {\displaystyle \theta } satisfies an "equivariance" condition, namely that for each g G {\displaystyle g\in G} and nonidentity a A {\displaystyle a\in A} , θ ( a g ) = θ ( a ) g {\displaystyle \theta (a^{g})=\theta (a)^{g}\,} where the superscript denotes conjugation by g . {\displaystyle g.} For example, the mapping a O p ( C G ( a ) ) {\displaystyle a\mapsto O_{p'}(C_{G}(a))} , the example of a signalizer functor given above, satisfies this condition.

If θ {\displaystyle \theta } satisfies equivariance, then the normalizer of B {\displaystyle B} will normalize W . {\displaystyle W.} It follows that if G {\displaystyle G} is generated by the normalizers of the noncyclic subgroups of A , {\displaystyle A,} then the completion of θ {\displaystyle \theta } (i.e., W) is normal in G . {\displaystyle G.}

References

  1. ^ a b Gorenstein, D. (1969), "On the centralizers of involutions in finite groups", Journal of Algebra, 11 (2): 243–277, doi:10.1016/0021-8693(69)90056-8, ISSN 0021-8693, MR 0240188
  2. ^ a b c Glauberman, George (1976), "On solvable signalizer functors in finite groups", Proceedings of the London Mathematical Society, Third Series, 33 (1): 1–27, doi:10.1112/plms/s3-33.1.1, ISSN 0024-6115, MR 0417284
  3. ^ a b McBride, Patrick Paschal (1982a), "Near solvable signalizer functors on finite groups" (PDF), Journal of Algebra, 78 (1): 181–214, doi:10.1016/0021-8693(82)90107-7, hdl:2027.42/23875, ISSN 0021-8693, MR 0677717
  4. ^ a b McBride, Patrick Paschal (1982b), "Nonsolvable signalizer functors on finite groups", Journal of Algebra, 78 (1): 215–238, doi:10.1016/0021-8693(82)90108-9, hdl:2027.42/23876, ISSN 0021-8693
  5. ^ Goldschmidt, David M. (1972a), "Solvable signalizer functors on finite groups", Journal of Algebra, 21: 137–148, doi:10.1016/0021-8693(72)90040-3, ISSN 0021-8693, MR 0297861
  6. ^ Goldschmidt, David M. (1972b), "2-signalizer functors on finite groups", Journal of Algebra, 21 (2): 321–340, doi:10.1016/0021-8693(72)90027-0, ISSN 0021-8693, MR 0323904
  7. ^ Bender, Helmut (1975), "Goldschmidt's 2-signalizer functor theorem", Israel Journal of Mathematics, 22 (3): 208–213, doi:10.1007/BF02761590, ISSN 0021-2172, MR 0390056
  8. ^ Flavell, Paul (2007), A new proof of the Solvable Signalizer Functor Theorem (PDF), archived from the original (PDF) on 2012-04-14
  9. ^ Aschbacher, Michael (2000), Finite Group Theory, Cambridge University Press, ISBN 978-0-521-78675-1
  10. ^ Kurzweil, Hans; Stellmacher, Bernd (2004), The theory of finite groups, Universitext, Berlin, New York: Springer-Verlag, doi:10.1007/b97433, ISBN 978-0-387-40510-0, MR 2014408