Roche Wef
Roche Wef
Bulletin
Number 88, Winter 2021, 57–68
ISSN 0791-5578
ALAN ROCHE
Abstract. Wedderburn proved in 1905 that a finite division ring is always a field.
His result has intrigued generations of mathematicians, spurring generalizations and
alternative proofs. The shortest, most elegant proof is surely Witt’s from 1931, now
the standard textbook treatment. Following a strategy of Zassenhaus, we present four
overlapping group-theoretic proofs. The first uses a counting argument; the others
hinge on properties of special classes of finite groups.
1. Introduction
Wedderburn’s Little Theorem says that a finite division ring is a field. That is, if each
nonzero element of a finite ring with identity has a multiplicative inverse then the ring is
commutative. In the words of I. N. Herstein, the theorem “has caught the imagination
of mathematicians because it is so unexpected, interrelating two seemingly unrelated
things, the number of elements in a certain algebraic system and the multiplication of
that system” [9]. In the words of Emil Artin, “this result of Wedderburn has fascinated
most algebraists to a very high degree” [2]. As often in mathematics, the first proofs
were somewhat clumsy. Quoting Artin again, Wedderburn [17] makes use of “divisibility
properties which are hard to establish” and “several attempts were made to simplify the
proofs.” In 1931, Witt gave a short, elegant proof. It uses only elementary group theory
in the form of the class equation (which Wedderburn also used) and easy divisibility
properties. It is indeed a proof from THE BOOK (see [1, Chap. 6]).
Witt’s argument, however, was far from the last word. There have been many other
proofs—ones like Witt’s that use essentially elementary methods and ones that stem
from a broader perspective. As an example of the latter, note that Wedderburn’s Big
Theorem [18] puts the Little Theorem in a wider setting—the theory of central simple
algebras and the Brauer group. This is a profound theory concerned, in its classical
incarnation, with describing division algebras that are finite dimensional over their
centres. Various results in the overall theory yield proofs of the Little Theorem. For
instance, by properties of what are called cyclic algebras the theorem reduces to an easy
calculation—checking surjectivity of the norm map for a finite extension of finite fields
(see, for example, [10, Section 8.4]).
Our starting point is a technical lemma due to Zassenhaus [20]. It says that if E is
a subfield of a finite division ring D then the normalizer and centralizer of E× in D×
coincide. Zassenhaus goes on to deduce the Little Theorem by a purely group-theoretic
argument (see Remark 1 below). We follow the same path: with Zassenhaus’s lemma
in hand, we apply tools from finite group theory to obtain the Little Theorem—four
times. Out of a perverse sense of symmetry, we also include four proofs of Zassenhaus’s
lemma.
The first proof of the Little Theorem is direct. Beyond the technical lemma, it
uses only elementary group theory and simple counting. The other proofs rely on
more advanced material: Burnside’s Normal Complement Theorem, Carter subgroups,
a property of Frobenius groups. For anyone new to these topics, we’ve tried to fill in
enough background so that the bulk of the paper still makes sense. More precisely, a
reader on friendly terms with (a) finite groups up to the Sylow theorems plus the notion
of a nilpotent group and (b) fields and basic Galois theory should be able to follow the
main thread of argument.
Three of our proofs share a common strategy: they proceed by showing that the
maximal fields in a noncommutative finite division ring D (with D sometimes of minimal
order) form a single conjugacy class (under the action of D× ). Since each element of D
is contained in a maximal field, it follows that if E is a fixed maximal field in D then the
group D× is a union of conjugates of E× . A finite group, however, is never a union of
conjugates of a proper subgroup (see Lemma 2 below), and so a noncommutative finite
division ring cannot exist. The same strategy underlies the classic proof of the Little
Theorem in van der Waerden’s Moderne Algebra, the influential text based on lectures
by Artin and Noether, first published in 1930–31. Van der Waerden’s argument, which
is due to Noether, uses early results in the theory of central simple algebras. It is
probably the most reproduced proof after Witt’s of the Little Theorem. 1
Except for the first, our proofs are more artificial than Zassenhaus’s. We could be
accused (we have been) of using a series of sledgehammers to crack the Little Theorem.
But they are such beautiful sledgehammers. The actual arguments, once the necessary
background is in place, are short, requiring just a tap from any sledgehammers we
wield. The proofs are meant as mathematical entertainments—an amusing application,
we hope, of parts of finite group theory.
Acknowledgements. The treatment of Zassenhaus’s lemma in Section 7 owes much to com-
ments provided by a reviewer of a previous (pre-Bulletin) version of the paper. In addition, I’ve
adopted several suggestions made by the referee of the current (Bulletin) version.
I would also like to note an older, more personal debt to two master expositors of mathematics,
T. J. Laffey and D. L. McQuillan, under whose stimulating guidance I first learned of the world
containing Wedderburn’s theorem.
Notation
Throughout D denotes a finite divison ring. The centre of D is a finite field which we
write always as F. For any ring R (with identity), R× denotes the group of units of R.
We write q for the cardinality of F and q N for the cardinality of D, so that N is the
dimension of D as an F-vector space.
For H a subgroup of a group G, the normalizer and centralizer of H in G are written
as N (H) and C(H). The ambient group G should be clear from context. In fact,
we work mostly with a subfield E of D. Then N (E× ) and C(E× ) always mean the
normalizer and centralizer in D× (that is, the ambient group is invariably D× ).
1Van der Waerden, Witt and Zassenhaus were part of the remarkable flowering of abstract algebra
in Germany in the 1920s and early 1930s in which Artin at Hamburg and Noether at Göttingen were
leading figures. Witt was Noether’s doctoral student and Zassenhaus was Artin’s. Van der Waerden
studied at Göttingen and was in the group (or ring) of young mathematicians centred around Noether;
he also spent a year at Hamburg where he worked closely with Artin [16]. And then the cataclysm.
In March 1933 the Nazis took power. In April a decree dismissed “non-Aryans” from government
institutions including universities. Within weeks, a vibrant mathematical culture was destroyed [12].
Hilbert, when asked in 1934 by the new minister of education about mathematics at Göttingen “now
that it was freed of the Jewish influence,” replied: “Mathematics at Göttingen? There is really none
anymore” [13, p. 205].
Wedderburn’s Little Theorem 59
For S a finite set, |S| denotes the number of elements in S. Given a subset T of S,
we write S \ T for the difference of S and T , that is, the set of elements of S that do
not belong to T .
2. Zassenhaus’s Lemma
The following technical result, due to Zassenhaus ([20, p. 59]), underpins each of our
approaches to Wedderburn’s theorem.
Lemma 1. If E is a subfield of a finite division ring D, then the normalizer and cen-
tralizer of E× in D× coincide.
We only need the statement for now and so defer the task of proving the lemma to
Section 7.
Remark 1. The formulation in [20] is superficially different: it says that if A is an
abelian subgroup of D× then N (A) = C(A). To obtain this version for a given abelian
subgroup A of D× , simply apply Lemma 1 to F(A), the subfield of D generated over F
by A. Zassenhaus goes on to prove the Little Theorem by establishing a pleasant result:
if a finite group G has the property that N (A) = C(A) for every abelian subgroup A
then G is abelian ([20, Theorem 7]).
Remark 2. Let E be a subfield of D that strictly contains the centre F, so that the
automorphism group Aut(E/F) is nontrivial. A foundational result in the theory of
central simple algebras—the Skolem–Noether Theorem—implies that each element of
Aut(E/F) is implemented by conjugation by some element of D× (see, for example,
[5, Theorem 3.14]). The action of N (E× ) on E by conjugation therefore induces an
isomorphism N (E× )/C(E× ) ≃ Aut(E/F), and so N (E× ) 6= C(E× ). Thus, if we were
willing to make use of the theory of algebras, Lemma 1 yields a one-line (or one-
paragraph) proof of the Little Theorem.
Rearranging, we obtain
qN − q q m1 − q q mr − q
= + · · · + . (2)
qN − 1 q m1 − 1 q mr − 1
1
The left side is less than 1. Further, each term on the right side is greater than .
2
Indeed, for m ≥ 2,
qm − q 1
m
> ⇐⇒ 2(q m − q) > q m − 1
q −1 2
⇐⇒ q m − 2q + 1 > 0,
Expanding each side, cancelling common terms and rearranging then gives
q N +1 − q N = q m+1 − q m .
For use in this section and the next, we list some properties of finite groups in which
all Sylow subgroups are cyclic. More can be said: in fact, a finite group has cyclic
Sylow subgroups if and only if it’s a semidirect product of cyclic groups whose orders
are relatively prime ([8, Theorem 5.16]).
Proposition 2. Let G be a finite group in which every Sylow subgroup is cyclic and
write l for the largest prime divisor of |G|. Then:
(a) a Sylow l-subgroup of G is normal;
(b) the group G is solvable;
(c) if G is nonabelian, it contains a normal abelian subgroup that is strictly larger
than the centre.
The proposition is a consequence of a famous result of Burnside ([8, Theorem 5.13]).
Burnside’s Normal Complement Theorem. Suppose a finite group G admits a
Sylow subgroup S such that N (S) = C(S). Then S has a normal complement in G.
That is, there is a normal subgroup N of G such that G = N S and N ∩ S = {1}.
2The author of the argument is better known today for activities outside mathematics.
62 ALAN ROCHE
The result is also known as Burnside’s Transfer Theorem as it follows from properties
of the transfer map—a natural homomorphism from a group G to the commutator group
H/H ′ of a subgroup H of finite index . Isaacs’ book [8] contains a cogent account of
this map and several of its applications including Burnside’s result.
Proof of Proposition 2. If G has prime-power order then (a) is obvious and (b) and
(c) are well known. For the remainder of the proof, we assume that the order of G is
divisible by at least two primes.
To establish part (a), we argue by induction on |G|. Write l1 for the least prime divisor
of G and let S1 be a Sylow l1 -subgroup of G. First, we note that Burnside’s Theorem
implies that S1 has a normal complement in G. To this end, observe that S1 ⊂ C(S1 ),
so that [N (S1 ) : C(S1 )] is not divisible by l1 . The action of N (S1 ) on S1 by conjugation
induces an embedding from N (S1 )/C(S1 ) into Aut(S1 ). We have |S1 | = l1e for some
positive integer e. Since S1 is cyclic, Aut(S1 ) has order φ(l1e ) = l1e−1 (l1 − 1). Thus
[N (S1 ) : C(S1 )] divides l1 − 1, and so N (S1 ) = C(S1 ) (as l1 is the least prime divisor
of |G|). Hence, by Burnside’s Theorem, G = N S1 for a normal subgroup N of G such
that N ∩ S1 = {1}.
Every Sylow subgroup of N is again cyclic. Our inductive hypothesis therefore implies
that a Sylow l-subgroup S of N is normal in N , and thus is the unique Sylow l-subgroup
of N . Now S is also a Sylow l-subgroup of G. Further, for g ∈ G, the group gSg −1 is
a Sylow l-subgroup of N (by normality of N ). Using uniqueness of S, we see that S is
normal in G.
For part (b), we also use induction on |G|. With notation as in the proof of part
(a), the quotient G/S inherits from G the property that each of its Sylow subgroups is
cyclic. Thus, by our inductive hypothesis, G/S is solvable. As S is certainly solvable,
it follows that G is solvable.
For part (c), suppose G is nonabelian and write Z for the centre of G. Then the
nontrivial group G/Z has cyclic Sylow subgroups. Using part (a), we see that G/Z
admits a nontrivial cyclic normal subgroup, say H/Z. Note that H is abelian (since
H/Z is cyclic), normal in G and strictly contains Z. We’ve proved part (c).
We now use Proposition 1 and Proposition 2 (c) to prove Wedderburn’s theorem.
Proof. We assume that D is noncommutative and derive a contradiction.
By Proposition 1 and Proposition 2 (c), the group D× contains a normal abelian
subgroup S that is strictly larger than F× . We set E = F(S), the subfield of D generated
over F by S. Since S is normal in D× , we see that E× is also normal in D× .
As in the first proof, we write down two ways to complete the argument (see the last
two paragraphs of [11] for yet another way).
Method 1. Since N (E× ) = D× , Lemma 1 says that C(E× ) = D× , and so E is contained
in the centre F of D—a contradiction.
Method 2. Recall that |F| = q and |D| = q N . As E strictly contains F, we have
C(E) 6= D. Thus
|E| = q m and |C(E)| = q n
for integers m and n with m ≤ n < N . Note that n divides N : in fact, N = ne where
e is the dimension of D as a C(E)-vector space.
The action of D× on E by conjugation induces an embedding of groups
D× /C(E)× ֒→ Aut(E/F).
In particular, [D× : C(E× )] is at most |Aut(E/F)| = m, that is,
q ne − 1
≤ m.
qn − 1
Wedderburn’s Little Theorem 63
is direct, and hence has dimension l2 over K1 . The containment is therefore an equality.
That is, each element of End K1 (K) admits a unique expression as l−1 i
P
i=0 λi σ with each
λi ∈ K. Using the same notation, we see that there is a well-defined map of K-vector
spaces
l−1
X Xl−1
i
λi σ 7−→ λi β i : End K1 (K) −→ D1 . (16)
i=0 i=0
We claim that the map is a ring homomorphism. Note that the lemma follows. In-
deed, given the claim, the kernel of (16) is a two-sided ideal in the simple ring End K1 (K)
and hence is trivial. This means that (16) is injective and so the division ring D1 con-
tains a copy of the matrix ring End K1 (K)—a contradiction. Alternatively, the map is
surjective (its image, for example, contains K and β). Using simplicity of End K1 (K)
again, the rings End K1 (K) and D1 must be isomorphic which is absurd.
To establish the claim, we appeal to the following principle whose proof is immediate.
(α) Let R and S be rings and let φ : R → S be a homomorphism of abelian groups.
Suppose R is generated as an abelian group by a subset Γ and that
φ(γ1 γ2 ) = φ(γ1 )φ(γ2 )
for all γ1 , γ2 ∈ Γ. Then φ is a homomorphism of rings.
From our discussion, End K1 (K) is generated as an abelian group by the elements λσ i
for λ ∈ K and i ∈ Z. Writing φ for the map in (16) and using (α), we see that we only
have to check one family of relations:
φ(λσ i µσ j ) = φ(λσ i )φ(µσ j ), λ, µ ∈ K, i, j ∈ Z.
By (15), λσ i µσ j = λσ i (µ)σ i+j , and so the left side is λσ i (µ)β i+j . By (5), the right side
is
λβ i µβ j = λσ i (µ)β i+j .
Thus (16) is a ring homomorphism and the (fourth) proof is complete.
68 ALAN ROCHE
References
[1] M. Aigner and G. M. Ziegler, Proofs from THE BOOK, 6th ed. Springer, 2018.
[2] E. Artin, The influence of J. H. M. Wedderburn on the development of modern algebra. Bull. Amer.
Math. Soc. 56 (1950), 65-72.
[3] R. P. Burn and D. M. Madaram, Frobenius groups and Wedderburn’s theorem. Amer. Math.
Monthly. 77(9) (1970), 983-984.
[4] R. W. Carter, Nilpotent self-normalizing subgroups of soluble groups. Math. Zeit. 75 (1961), 136-
139.
[5] R. K. Dennis and B. Farb, Noncommutative Algebra. Graduate Texts in Math., 144, Springer-
Verlag, 1993.
[6] S. Ebey and K. Sitaram, Frobenius groups and Wedderburn’s theorem. Amer. Math. Monthly. 76(5)
(1969), 526-528.
[7] W. Feit, Characters of Finite Groups. W. A. Benjamin, 1967.
[8] I. M. Isaacs, Finite Group Theory. Graduate Studies in Math., 92, Amer. Math. Soc., 2008.
[9] I. N. Herstein, Topics in Algebra, 2nd ed. John Wiley & Sons, 1975.
[10] N. Jacobson, Basic Algebra II, 2nd ed. W. H. Freeman and Co., 1989.
[11] T. J. Kaczynski, Another proof of Wedderburn’s theorem. Amer. Math. Monthly. 71(6) (1964),
652-653.
[12] S. Mac Lane, Mathematics at Göttingen under the Nazis. Notices Amer. Math. Soc. 42(10) (1995),
1134-1138.
[13] C. Reid, Hilbert. Reprint of 1970 original. Copernicus, 1996.
[14] J.-P. Serre, Finite Groups: An Introduction. International Press, 2016.
[15] M. Chiara Tamburini, The conjugacy conjecture for Carter subgroups. Milan J. Math. 75 (2007),
357-377.
[16] B. L. van der Waerden, On the sources of my book Moderne Algebra. Hist. Math. 2 (1975), 31-40.
[17] J. H. M. Wedderburn, A theorem on finite algebras. Trans. Amer. Math. Soc. 6 (1905), no. 3,
349-352.
[18] J. H. M. Wedderburn, On hypercomplex numbers. Proc. London Math. Soc. (2) 6 (1908), 77-118.
[19] E. P. Vdovin, Carter subgroups of finite groups. Siberian Adv. Math. 19 (2009), no. 1, 27-74.
[20] H. J. Zassenhaus, A group-theoretic proof of a theorem of Maclagan-Wedderburn. Proc. Glascow
Math. Assoc. 1 (1952), 53-63.
Alan Roche studied at University College Dublin and the University of Chicago, and has
worked at the University of Oklahoma since 2001. His principal mathematical interests are in
number theory and representation theory.
Department of Mathematics, University of Oklahoma, Norman, OK 73019, USA.
E-mail address: [email protected]