Chapter 9

Series of Subgroups

In the future, as in the past, the great ideas must be simplifying ideas.

—André Weil

If G is a finite abelian group, it can be shown that G is isomorphic to a direct product of cyclic groups (see Chapter 7). This result is an example of a structure theorem, that is, a theorem showing that every group in a suitable defined class may be constructed in a systematic way from well understood groups in the class. Such theorems are hard to come by, and the result for finite abelian groups is a stunning example. The structure of nonabelian finite groups is much more complicated.

Suppose that groups K and H are given. It is a very difficult problem to describe all groups G that have a normal subgroup K1 isomorphic to K such that G/K1 is isomorphic to H. If we could solve this extension problem, the solution would give an inductive method for constructing all finite groups. Direct and semidirect products solve this problem in very special cases. Although the general problem is far from being solved, the classes of groups that can be built up this way are of interest.

To illustrate, suppose that we use only abelian groups as building blocks. Starting with an abelian group G0, we construct G1G0 such that G0 img G1 and G1/G0 is abelian. Next, we extend G1 to obtain G2G1 such that G1 img G2 and G2/G1 is abelian. After n steps, we have a chain

G = 389GnGn−1imgG1G0,

where Gi img Gi+1 and Gi+1/Gi is abelian for each i. Such a group G is called solvable, and the theory of these groups is successful in the following sense: The class of solvable groups is large (it contains all finite groups of odd order), but at the same time, many theorems are true for all solvable groups but do not hold in general. We investigate solvable groups in Section 9.2.

If we use simple groups as building blocks in this way, the resulting groups are those studied in Section 9.1. In this case, the above chain of subgroups is called a composition series for G, and the famous Jordan–Hö lder theorem asserts that G uniquely determines the series of groups Gn/Gn−1, Gn−1/Gn−2, . . ., G1/G0, G0. This leads to the useful notion of the composition length of a group.

Section 9.3 deals with somewhat more specialized central series and begins the study of finite nilpotent groups. These groups are characterized as the groups that are the direct product of their Sylow subgroups, equivalently if every Sylow subgroup is normal. In addition, the Frattini subgroup is defined for every finite group and shown to be nilpotent.

9.1 The Jordan–Hölder Theorem

Much of what we do in this chapter is concerned with groups G that admit a chain of subgroups with certain nice properties. A subnormal series for G is a chain

G = G0G1G2imgGn = {1}

of subgroups of G such that Gi+1 img Gi for each i. The factor groups Gi/Gi+1 are called the factors of the subnormal series. Note that we do not insist that the subgroups Gi are normal in G. Moreover, by possibly deleting some of the groups Gi, we clearly may assume that GiGi+1 for each i.

Example 1. G ⊇ {1} is a subnormal series for any group G. The only factor is G.

Example 2. A4KHε is a subnormal series for A4, we have K = {ε, (12)(34), (13)(24), (14)(23)} and H = {ε, (12)(34)}. The factors are C3, C2, and C2 in order. Note that H is not normal in A4.

The most interesting cases are the groups that admit a subnormal series in which every factor is abelian or every factor is simple. We investigate the first case in Section 9.2; the second calls for another definition.

If G is a group, a subnormal series G = G0G1imgGn = {1} is called a composition series for G if each factor Gi/Gi+1 is simple. In this case, the factors Gi/Gi+1 are called composition factors of G, and the integer n is called the length of the composition series. If G = {1}, we say that G has a composition series of length 0.

Example 3. The simple groups are those with a composition series of length 1.

Example 4. Every finite group G has a composition series. This holds by definition if G = {1}. If G ≠ {1}, write G = G0 and choose a maximal normal subgroup G1 of G0 (it exists because G is finite). Then G0/G1 is simple by Theorem 6 §8.1. If G1 ≠ {1}, choose a maximal normal subgroup G2 of G1 and continue in this way. The series G = G0G1G2img must reach {1} eventually because G is finite, so it is a composition series.

The converse of Example 4 is false: Any infinite simple group has a composition series of length 1. However, the converse does hold for abelian groups.

Example 5. An abelian group G has a composition series if and only if G is finite. Indeed, if G = G0G1imgGn = {1} is a composition series, each composition factor Gi/Gi+1 is a simple abelian group and so is finite. Hence (Exercise 11), |G| = |G/G1||G1/G2| img |Gn−1/Gn| is also finite.

The finite abelian groups are not the only ones having all composition factors abelian. Theorem 8 §8.2 shows that every finite p-group has this property:

Example 6. If p is a prime, each finite p -group has a composition series in which every composition factor is isomorphic to Cp.

A group may have several different composition series. For example, let G be a cyclic group of order 12 and, for each divisor d of 12, let Hd be the unique subgroup of G of order d. Then G has three composition series. These series, along with their composition factors, are

img

Note that the length is 3 in each case and the factors are also the same except for the order in which they appear. Hence, these series are all equivalent in the following sense: Two composition series for a group are said to be equivalent if they have the same length and the composition factors can be paired in such a way that corresponding factors are isomorphic. This is clearly an equivalence relation on the set of all composition series of a group G (assuming that there is one). The remarkable thing is that any two composition series for G are equivalent. This is the most important theorem in this section.

Theorem 1. Jordan–Hölder Theorem. If a group has a composition series, any two composition series are equivalent.

We will give the proof at the end of this section.

If a group G has a composition series, it uniquely determines the length of the series and the composition factors (including multiplicities). Hence, we can speak of the composition length of the group G, denoted lengthG, and of the composition factors of G.

Composition series were first discussed by Camille Jordan. In 1869, he showed (for groups of permutations) that the orders of the composition factors are the same for every composition series of the group. However, it was not until 20 years later, after the abstract definition of a group had been given, that Otto H ölder observed that the group uniquely determines the composition factors themselves and that they are the same in any composition series.

Example 7. If n ≥ 5, then SnA5 ⊃ {ε} is a composition series because An is simple (Theorem 8 §2.8). Hence, Sn has length 2 and the composition factors are C2Sn/An and An. If n = 4, we get a composition series

S4A4KH ⊃ {ε},

where K = {ε, (12)(34), (13)(24), (14)(23)} and H = {ε, (12)(34)}. Hence, S4 has length 4 and the composition factors are C2, C3, C2, and C2.

The Jordan–Hölder theorem is a type of unique factorization theorem. In the following corollary, we use it to give another proof that the factorization of an integer n into primes is unique. Here, the composition factors play the role of primes.

Corollary. The factorization of an integer n ≥ 2 into primes is unique.

Proof. Let n = p1p2 img pr, where pi are (not necessarily distinct) primes. If img is a cyclic group of order n, then

img

is a composition series for G because the factors are img (Indeed, img for each k by Theorem 5 §2.4.) Since any other factorization of n into primes must yield the same composition factors, it follows that n uniquely determines the number r = lengthG and the primes pi.

If K img G and G/KH, the group G is called an extension of K by H. So if

G = G0G1imgGr = {1}

is a composition series for G, then each Gi is an extension of Gi+1 by a simple group. Thus, each finite group G is the result of a finite number of extensions by finite simple groups, and the Jordan–Hö lder theorem shows that G uniquely determines the simple groups used (up to order). Moreover, we know all the finite simple groups (see the discussion at the end of Section 2.8), so the complete description of all finite groups comes down to the extension problem: For a simple group H, describe all extensions of a given group G by H. This is a very difficult task.

We are going to prove that subgroups and homomorphic images of groups with a composition series again have composition series. The proof requires the follow-ing technical lemma that gives important information about subnormal series in general and will be referred to again later. For composition series, we use it to deduce some important properties of the length of a group and prove the Jordan–Hölder theorem itself.

Lemma 1. Let G = G0G1imgGn = {1} be a subnormal series for the group G and let K img G.

1. K = KG0KG1imgKGn = {1} is a subnormal series for K, and the factor (KGi)/(KGi+1) is isomorphic to a normal subgroup of Gi/Gi+1 for each i.

2. G/K = (KG0)/K ⊇ (KG1)/Kimg ⊇ (KGn)/K = {K} is a subnormal series for G/K, and the factor [(KGi)/K]/[(KGi+1)/K] is a homomorphic image of Gi/Gi+1 for each i.

Proof. We leave to the reader the verification that the series are subnormal.

(1) Define α: KGiGi/Gi+1 by α(x) = xGi+1. This is clearly a group homomorphism and ker α = {x img KGi img x img Gi+1} = KGi+1. Hence, it remains to prove that α(KGi) img (Gi/Gi+1). But if x img KGi and y img Gi, then

(yGi+1)α(x)(yGi+1)−1 = (yxy−1)Gi+1 = α(yxy−1) img α(KGi)

because yxy−1 img (yKy−1) ∩ Gi = KGi.

(2) Since KGi+1 img KGi (Exercise 15 §8.1), the third isomorphism theorem (Theorem 7 §8.1) shows that

img

To show that (KGi)/(KGi+1) is an image of Gi/Gi+1, define

img by  α(xGi+1) = xKGi+1 for all x img Gi.

This is well defined because xGi+1 = yGi+1 implies that y−1x img Gi+1KGi+1. It is clearly a group homomorphism, and (as K img G) it is onto because

kxKGi+1 = x(x−1kx)KGi+1 = xKGi+1 = α(xGi+1)

holds for all k img K and x img Gi.

Now suppose that G = G0G1imgGn = {1} is a composition series for G. If K img G, the subnormal series for K and G/K in Lemma 1 are also composition series. Indeed, in both cases, the factors are isomorphic to either normal subgroups or images of the simple groups Gi/Gi+1 and so are all either simple or {1}. Hence, after we eliminate equalities, these series become composition series for K and G/K, respectively, each with factors from G. This proves part of Theorem 2.

Theorem 2. Let G be a group and let K img G. Then G has a composition series if and only if both K and G/K have composition series. Moreover, in this case,

1. length G = length K + length G/K.

2. The composition factors of G are exactly those of K and G/K.

3. G has a composition series containing K.

Proof. If G has a composition series, we have already seen that this is true of K and G/K. Conversely, suppose

K = K0K1imgKm = {1}  and img

are composition series for K and G/K. Because GiGi+1Gi/KGi+1/K, the series

G = G0G1imgGr = K = K0K1imgKm = {1}

is a composition series for G. Now (1)–(3) are apparent.

Corollary 1. If θ: GH is a homomorphism, then G has a composition series if and only if both ker θ and θ(G) have composition series. In this case,

lengthG = length ker θ + lengthθ(G).

Proof. Since G/ker θθ(G), Theorem 2 applies with K = ker θ.

Corollary 2. If G1, G2, img, Gn are groups, then G1 × G2 × img × Gn has a composition series if and only if the same is true of each Gi. In this case,

length(G1 × G2 × img × Gn) = lengthG1 + lengthG2 + img + lengthGn.

Solution. We prove it for n = 2, and then the general case follows by induction. Define θ: G1 × G2G2 by θ(g1, g2) = g2 for all g1 img G1 and g2 img G2. Then θ is a homomorphism, ker θ = G1 × {1} ≅ G1, and θ(G1 × G2) = G2. Use Corollary 1.

Example 8. Let G be an abelian group of order img where pi are distinct primes. Show that lengthG = n1 + n2 + img + nr.

Solution. Proceed by induction on |G|. If |G| = 1, it is clear. In general, let K be a maximal (normal) subgroup of G. Then |G/K| is a prime divisor of |G|, say |G/K| = p1. Hence, img so induction and Theorem 2 give

img

We conclude this section with a proof of the Jordan–Hölder theorem. The proof requires the following lemma.

Lemma 2. Let G be a group and let H and K be distinct maximal normal subgroups of G. Then HK is maximal normal in both H and K. Moreover,

img and   img

Proof. We first claim that KH = G. Indeed, HKH img G and KKH img G, so if KHG, the fact that H and K are maximal normal in G implies that H = KH = K, contrary to assumption. Hence, KH = G, so the second isomorphism theorem (Theorem 1 §8.1) gives GK = KHKHHK. Because G/K is simple, this shows that KH is maximal normal in H. The rest is proved in the same way.

Proof of the Jordan–Hölder Theorem. Suppose the group G has a composition series

(1) img

of length n. We show by induction on n that every composition series

(2) img

for G is equivalent to series (1). If n = 1, then G is simple, so G1 = {1} = H1 and the theorem holds. So assume that n ≥ 2 and that the theorem holds for all groups with a composition series of length less than n. In particular, it holds for G1 because G1G2imgGn = {1} has length n − 1. If it happens that H1 = G1, then G1H2imgHm = {1} is another composition series for G1 and so is equivalent to G1G2imgGn = {1} by induction and the theorem follows.

So assume that H1G1 and let H1G1 = L0L1imgLs = {1} be a composition series for H1G1 by Theorem 2 as H1G1 img G. Now consider the following series for G:

(3) img

(4) img

As H1G1, Lemma 2 asserts that H1G1 is maximal normal in each of H1 and G1, so both (3) and (4) are composition series for G. Moreover, G/G1H1/(H1G1) and G/H1G1/(H1G1) by Lemma 2, so (3) and (4) are equivalent; denote this as (3) img (4). Note that this equivalence holds even if s = 0, that is, if H1G1 = {1}.

Now G1G2imgGn = {1} and G1 ⊃ (H1G1) ⊃ L1imgLs = {1} are composition series for G1 and so are equivalent by induction. This implies that (1) img (3) and also that n − 1 = s + 1. But then the composition series H1 ⊃ (H1G1) ⊃ L1imgLs = {1} has length n − 1 and so (again by induction) is equivalent to H1H2imgHm = {1}. This in turn implies that (2) img (4). Piecing these equivalent series together gives (1) img (3) img (4) img (2), which proves the Jordan–Hölder theorem.

Exercises 9.1

1. In each case, find the length of the group and exhibit the composition factors.

(a) C8  (b) C12  (c) D4   (d) A4  (e) Q (see Section 2.8)

2. If n ≥ 1 and p is prime, show that img has exactly one composition series.

3. Find all composition series for

a. C24

b. C30

4. Find two nonisomorphic finite groups with identical composition factors.

5. Find all composition series for C4 × C2.

6. If img find the length of Dn. [Hint: Example 8.]

7. Find a composition series for D16 that contains the center Z(D16) and find one that does not contain Z(D16).

8.

a. For each m ≥ 2, find a group of length m.

b. For each m ≥ 2, find a group of length 1 with a subgroup of length m.

9. Let G = K0 × K1 × img × Kr, where each Ki is simple. Show that the groups Ki are the composition factors of G.

10. Describe the groups of length 2 by using Exercise 6 §8.1.

11. Let G = G0G1imgGn= {1} be any subnormal series. If each factor Gi/Gi+1 is finite, show that G is finite and that |G| = |G0/G1| · |G1/G2| img |Gn−1/Gn|.

12. If p is a prime, show that a finite group is a p-group if and only if all its composition factors are isomorphic to Cp.

13. Suppose that G is a group with a composition series. Show that any subnormal series G = G0G1imgGn ={1} can be refined (by inserting groups if necessary) to a composition series for G.

14. Suppose that G has a composition series with no two factors isomorphic.

a. Show that no two proper normal subgroups of G are isomorphic. [Hint: If H img G and K img G, find a composition series through HK, H, and HK and one through HK, K, and HK. Use Exercise 13.]

b. Show that every normal subgroup of G is characteristic in G.

15. Let img where the pi are distinct primes and each ni ≥ 1.

a. Show that Cn has exactly r maximal normal subgroups.

b. If m = n1 + n2 + img + nr, show that Cn has m ! n1 ! n2 ! img nr ! composition series. [Hint: induct on m.]

16. Prove the Zassenhaus lemma:100 Let H1 img H and K1 img Kbe subgroups of a group G. Then H1(HK1) img H1(HK), K1(H1K) img K1(HK), and we have H1(HK)H1(HK1) ≅ K1(HK)K1(H1K). [Hint: Each group is isomorphic to HK(H1K)(HK1) by Theorem 4 §2.10.]

17. Prove the Schreier refinement theorem:101 Two subnormal series G = G0G1imgGn ={1} and H = H0H1imgHm = 1 can be refined (by inserting groups) in such a way that the resulting series are equivalent. [Hint:

Gi = Gi+1(GiH0) ⊇ Gi+1(GiH1) ⊇ imgGi+1(GiHm) = Gi+1.

Do a similar construction with the Hj and use the Zassenhaus lemma.]

18. Use the Schreier refinement theorem to prove the Jordan–Hö lder theorem.

9.2 Solvable Groups

In Section 9.1, we were concerned with groups that admit a composition series, that is, a subnormal series in which all the factors are simple. Although those groups are of interest, we obtain an even more important class of groups when the factors are required to be abelian rather than simple.

A group G is called a solvable group102 if there exists a subnormal series

G = G0G1imgGn = {1}

such that each factor Gi/Gi+1 is abelian. Such a series is called a solvable series for G. Note that {1} is solvable.

Example 1. Every abelian group G is solvable because G ⊇ {1} is a solvable series.

Example 2. If p is a prime, every finite p -group is solvable. In fact, it has a composition series in which each factor is isomorphic to Cp (Theorem 8 §8.2).

Example 3. Dn is solvable for each n. Indeed, Dn has a cyclic subgroup H of index 2, so H img Dn and DnH ⊇ {1} is a solvable series.

Example 4. S4 is solvable because S4A4K ⊇ {ε} is a solvable series, where K = {ε, (12)(34), (13)(24), (14)(23)}.

Example 5. If p and q are primes, show that any group of order pq is solvable.

Solution. Because G is not simple by the Sylow theorems (Example 5 §8.4), let K ≠ {1} be a proper normal subgroup. Then |K| = p or |K| = q; either way GK ⊃ {1} is a solvable series with factors Cp and Cq.

Suppose that G1 is an abelian group. It is difficult to describe how to construct all groups G2 such that G1 img G2 and G2/G1 is abelian. Nonetheless, suppose we carry out this construction and then repeat it to construct a group G3 such that G2 img G3 and G3/G2 is abelian. If we continue this procedure, each group constructed is clearly solvable, and we can obtain every solvable group in this way—constructed from the bottom up, as it were. Viewing solvable groups in this way is useful, but an analogous top–down construction is actually more important. It is based on the derived subgroup introduced in Section 2.9.

Recall that an element of the form aba−1b−1 in a group G is called a commutator, denoted [a, b]. The set G′ of all products of commutators is a subgroup, called the derived subgroup of G, and has the following properties (Theorem 3 §2.9):

1. Gimg G and G/Gis abelian.

2. If K img G and G/K is abelian, then G′ ⊆ K.

For a group G, repeatedly taking the derived subgroup leads to a subnormal series of subgroups GG′ ⊇ G′′G′′′img in which each factor is abelian. There is a standard notation for these subgroups. Given a group G, construct subgroups G(0), G(1), G(2), img of G as follows:

1. Define G(0) = G.

2. If G(i) has been constructed for i ≥ 0, define G(i+1) = [G(i)]′.

Thus, G(1) = G′, G(2) = G′′, G(3) = G′′′, and so on. Furthermore, G(i+1) img G(i) for each i and the subnormal series

G = G(0)G(1)G(2)img

is called the derived series for G. Note that G(i)/G(i+1) is abelian for each i by Theorem 3 §2.9. The groups G(i) are called the higher derived subgroups of G, and they are actually normal in G as the next theorem shows. The proof requires the following lemma; we leave the proof as Exercise 8.

Lemma 1. Let G denote a group and let H be a subgroup.

1. If α: GG is a homomorphism, then α(G′) ⊆ G′.

2. G′ ⊆ H if and only if H img G and G/H is abelian.

3. If HG is a subgroup, then H′ ⊆ G′.

Theorem 1. If G is a group, we have G(i) img G for all i ≥ 0.

Proof. We use induction on i ≥ 0. It is clear if i = 0, so assume inductively that G(i) img G for some i. If a img G then σa(G(i)) = G(i), where σa is the inner automorphism of G determined by a. But then Lemma 1 gives

σa[G(i+1)] = σa[(G(i))′] ⊆ (G(i))′ = G(i+1).

This shows that G(i+1) img G and so completes the induction.

The solvable groups G are just those for which the derived series reaches {1}.

Theorem 2. A group G is solvable if and only if G(n) = {1} for some n ≥ 1.

Proof. If G(n) = {1}, then

G = G(0)G(1)G(2)imgG(n) = {1}

is a solvable series for G because G(i)/G(i+1) = G(i)/[G(i)]′ is abelian for each i. Conversely, let G = G0G1imgGn = {1} be a solvable series for G. It suffices to show that G(i)Gi holds for each i. This is clear if i = 0, so assume that G(i)Gi for some i ≥ 0. As Gi/Gi+1 is abelian, we have img Hence,

img

by Lemma 1, which completes the induction.

If G is solvable, then G(i)G(i+1) is strictly for all i by Corollary 1 of Theorem 3.

Theorem 2 provides a quick method of establishing several basic properties of solvable groups. We begin with the following result.

Theorem 3. Every subgroup and image of a solvable group is again solvable.

Proof. Suppose that G is solvable and let G(n) = {1}. If H is a subgroup of G, it suffices to show that H(i)G(i) for each i. This follows by induction: It is clear when i = 0, and if H(i)G(i), then Lemma 1 gives

H(i+1) = [H(i)]′ ⊆ [G(i)]′ = G(i+1).

Now let α: GK be an onto group homomorphism. It suffices to show that K(i)α[G(i)] for each i. This is clear if i = 0 because α is onto, so assume that K(i)α(G(i)). Then, given x and y in K(i), write x = α(a) and y = α(b), where a, b img G(i). Hence,

[x, y] = [α(a), α(b)] = α([a, b]) img α(G(i+1)),

so K(i+1)α(G(i+1)) because K(i+1) = (K(i))′.

Corollary 1. If H ≠ {1} is a subgroup of a solvable group G, then H′ ≠ H.

Proof. If H′ = H, then H(2) = [H′]′ = H′ = H and an induction shows that H(i) = H ≠ {1} holds for each i. As H is solvable by Theorem 3, this result contradicts Theorem 2.

Corollary 2. A simple group is solvable if and only if it is abelian (of prime order).

Proof. Let G be solvable. If G is simple, then either G′ = {1} (so G is abelian) or G′ = G, (contradicting Corollary 1). The converse is clear.

Example 6. If n ≥ 5, the symmetric group Sn is not solvable. For if Sn were solvable, An would be solvable by Theorem 3 so, as An is simple, Corollary 2 would imply that An is abelian, which is not the case. Hence, Sn is not solvable.

Example 6 explains the origin of the term solvable. A classical problem in the theory of equations was to find a formula for the roots of a real polynomial xn + an−1xn−1 + img + a1x + a0 in terms of the coefficients ai. If n = 2, the solution is the famous quadratic formula: img In general, such a formula should give the roots in terms of the coefficients ai using only arithmetic operations and the extraction of roots. Such formulas were found for n = 3 and n = 4, but the case n = 5 proved to be difficult. Call a polynomial f solvable if such a formula exists. It will be shown in Chapter 10 that f is solvable if and only if a certain group (called the Galois group of f) is a solvable group. For example, the polynomial x5 − 6x + 2 has Galois group S5 (Example 1 §10.3) and so cannot be solvable. Incidentally, the first proof that a nonsolvable polynomial exists was given in 1824 by the young Norwegian mathematician Niels Henrik Abel, building on the work of Paolo Ruffini.

Theorem 4 gives a useful way to show that a group is solvable.

Theorem 4. If K img G, then G is solvable if and only if both K and G/K are solvable.

Proof. Assume that K and G/K are solvable and let

img and img

be solvable series. Then

G = G0G1imgGr = K = K0K1imgKm = {1}

is a subnormal series for G and the factors are abelian because GiGi+1Gi/KGi+1/K for each i. Hence, G is solvable. The converse follows by Theorem 3.

Example 7. If G1, G2, img, Gn are groups, then G1 × G2 × img × Gn is solvable if and only if the same is true of each Gi.

Solution. By induction, it suffices to prove it for n = 2. Let θ: G1 × G2G2 be the projection given by θ(g1, g2) = g2. Then (G1 × G2)/ker θθ(G1 × G2) = G2 and ker θ = G1 × {1} ≅ G1 are both solvable, so Theorem 4 applies.

The above theorems are valid for arbitrary groups. We now give some conditions equivalent to solvability in a finite group.

Theorem 5. The following conditions are equivalent for a finite group G.

1. G is solvable.

2. The composition factors of G are all abelian.

3. H′ ≠ H for every subgroup H ≠ {1} of G.

Proof. Note first that G has a composition series because it is finite.

(1) ⇒ (2). Each composition factor is simple and solvable (Theorem 3), and so is abelian by Corollary 2 of Theorem 3.

(2) ⇒ (3). Any composition series for G is a solvable series by (2).

(3) ⇒ (1). The derived series G = G(0)G(1)G(2)img reaches {1} because G is finite and G(i) ⊃ (G(i))′ = G(i+1) for each i by (3).

Example 8. Let R be any ring. If n ≥ 3, show that the group G of all invertible n × n matrices over R is not solvable.

Solution. Let Eij denote the n × n matrix with (i, j)-entry 1 and zeros elsewhere. Then EijEjk = Eik, whereas EijElk = 0 if jl. If I is the n × n identity matrix, this shows that I + Eij is in G whenever ij and that (I + Eij)−1 = IEij. Now let H be the subgroup of G generated by the matrices I + Eij, ij. If i, j, and k are distinct indices (they exist because n ≥ 3), compute

img

This shows that every generator of H is a commutator from H and hence H′ = H. Thus, G is not solvable by Corollary 1 of Theorem 3.

If F is a field, Example 8 shows that the general linear group GLn(F) of all n × n invertible matrices over F is not solvable if n ≥ 3. If F is finite, Theorem 5 shows that a nonabelian simple group is lurking among the composition factors of GLn(F). In fact, such a group exists even if F is infinite. The mapping A img det A is an onto homomorphism GLn(F) → F* and the kernel is the special linear group SLn(F) of all matrices with determinant 1. It is not difficult to verify that the center of SLn(F) consists of all scalar matrices aI, where a img F satisfies an = 1. The factor group

img

is called the projective special linear group (of degree n) over F. These groups comprise another infinite family of finite, simple, nonabelian groups (in addition to the alternating groups An, n ≥ 5). The theorem was proved in 1870 by Camille Jordan for img a prime and, in early 1900, Leonard Eugene Dickson proved it for all finite fields.

Theorem 6. Jordan–Dickson Theorem. If F is a finite field, then PSLn(F) is a finite nonabelian simple group for all n ≥ 2, except for img and img

The proof is beyond the scope of this book.103

The class of solvable groups is large. Of course, it contains all abelian groups, and a celebrated theorem of William Burnside asserts that every group of order pnqm is solvable, where p and q are primes. In a different direction, Georg Frobenius showed that every group of square-free order is solvable. In 1911, Burnside conjectured that every nonabelian finite simple group has even order, equivalently (Exercise 13) that every group of odd order is solvable. This conjecture remained an open question until 1963 when two American algebraists Walter Feit and John Thompson proved that it is true. The proof is 254 pages long and fills an entire issue of the Pacific Journal of Mathematics,104 and it is widely regarded as the best single paper in finite group theory. Thompson went on to classify all minimal finite simple groups, that is, those in which every proper subgroup is solvable, and played an important role in the classification of all finite simple groups. He was awarded the Fields Medal in 1970, the highest honor a mathematician can attain.

Even though the class of solvable groups is very large, many theorems are true for solvable groups that are not true of groups in general. One such theorem, a fundamental strengthening of the Sylow theorems in any solvable group, was first proved in 1928 by the British mathematician Philip Hall.

Theorem 7. Hall's Theorem. Let G be a group of order nm, where n and m are relatively prime. If G is solvable, then

1. G has a subgroup of order n and any two are conjugate.

2. If k|n, each subgroup of order k is contained in a subgroup of order n.

We omit the proof.105 Hall went on to develop the theory of finite solvable groups and influenced an entire generation of group theorists.

Exercises 9.2

1. Is Z(G)≠{1} for every solvable group G ≠ {1}? Support your answer.

2. If G is solvable, is N(H) ≠ H for each subgroup HG? Support your answer.

3. Is G′ abelian for every solvable group G? Support your answer.

4. Does every solvable group of order n have a subgroup of order m for each divisor m of n? Support your answer.

5. Give an example of a nonsolvable group in which every Sylow subgroup is abelian.

6. Show that a nonsolvable group of minimal order must be simple.

7. Suppose G has a solvable, maximal normal subgroup. Either prove G is solvable or give a counterexample.

8. Prove Lemma 1.

9. If |G| = p2q, p, q primes, show that G is solvable. [Hint: Exercise 14 §8.4.]

10. If |G| = p2q2, p, q primes, show that G is solvable. [Hint: Example 9 §8.4 and the preceding exercise.]

11.

a. Show that

img

is a solvable group for any field F.

b. Show that

img

is a solvable group for any field F.

12. If p and q are primes, show that every group of order pmqn is solvable if and only if the only simple groups of this type are the cyclic groups of order p or q. (Burnside proved that these statements are true.)

13. Show that every group of odd order is solvable if and only if every finite nonabelian simple group has even order. (These statements are true by Feit and Thompson.)

14. Find the composition length of a solvable group of order img where pi are distinct primes. [Hint: Example 8 §9.1.]

15. Show that a solvable group is finite if and only if it has a composition series.

16. Show that the following are equivalent for a group G.

(a) G is solvable.  (b) G′ is solvable.  (c) G/Z(G) is solvable.

17. If H img G and K img G show that G/(HK) is solvable if and only if the product (G/H) × (G/K) is solvable.

18. If Ki img G for i = 1, 2, img, n, put K = K1K2imgKn. If G/Ki is solvable for each i, show that G/K is solvable.

19. If H and K are solvable subgroups of G and K img G, show that HK is solvable.

20.

a. If G is a finite group and Z(G/K) is nontrivial for all K img G, KG, show that G is solvable.

b. Show that the converse of (a) is false for a finite group G.

21. If G ≠ 1 is solvable, show that

a. G has a nontrivial abelian factor group.

b. G has a nontrivial abelian normal subgroup.

22. Show that the following are equivalent for a nontrivial finite group G.

1. G is solvable.

2. Every nontrivial normal subgroup of G has a nontrivial abelian factor group.

3. Every nontrivial factor group of G has a nontrivial abelian normal subgroup.

23. If G is a finite group, define R = R(G) = img {K img G img G/K is solvable}.

a. Show that R = R(G) is the smallest normal subgroup of G such that G/R is solvable. [Hint: Exercise 18.]

b. Show that G is solvable if and only if R(G) = {1}.

c. If α: GH is a group homomorphism, show that α[R(G)] ⊆ R(H). Hint: Consider {k img α(k) img R(H)}.

d. If HG is a subgroup, show that R(H) ⊆ HR(G), with equality if H img G.

24. If G is a finite group, define S(G) = ∏ {K img G img K is solvable}.

a. Prove S(G) is the largest solvable, normal subgroup of G. [Hint: Exercise 19.]

b. Prove G is solvable if and only if S(G) = G.

c. If α: GH is an onto group homomorphism, show that α[S(G)] ⊆ S(H).

d. If HG is a subgroup, show that HS(G) ⊆ S(H), with equality if H img G.

e. Show that S(G/S) = {1}.

25. A group G is called polycyclic if it has a solvable series with every factor cyclic.

a. Show that every finite solvable group is polycyclic.

b. Show that every polycyclic group is finitely generated.

c. Show that every subgroup and homomorphic image of a polycyclic group is polycyclic. [Hint: Lemma 1 §9.1.]

d. If K img G, show that G is polycyclic if and only if K and G/K are polycyclic.

e. Show that the following are equivalent for a group G. [Hint: Theorem 3 §7.2.]   

i. G is polycyclic   

ii. Every subgroup of G is solvable and finitely generated.   

iii. Every normal subgroup of G is solvable and finitely generated.

26. A class img of groups is called a subvariety if img and each subgroup and ho-momorphic image of a group in img is again in img. Examples: abelian groups, p-groups for a fixed prime p and torsion groups (each element has finite order). If img is a subvariety, a group G is called img-solvable if there is a subnormal series G = G0G1imgGn= {1} with Gi/Gi+1 in img for each i. If G is a group and K img G, show that G is img-solvable if and only if both K and G/K are img-solvable. [Hint: Lemma 1 §9.1.]

27. A subvariety img of groups (Exercise 26) is called a variety if, in addition, G × H is in img whenever G and H are in img. Examples: abelian groups, p -groups for a fixed prime p, torsion groups, and img-solvable groups, where img is any subvariety (by Exercise 26). If img is a variety and G is a finite group, the img-derived subgroup of G is defined to be img is in img Let G denote a finite group. (a) Show that img and img is in img (b) If K img G, show that G/K is in img if and only if img (c) If H is a subgroup of G, show that img [Hint: Lemma 2 §9.1.] (d) If α: GH is a homomorphism of groups, show that img

28. If img is a variety of finite groups, define img and img for each k ≥ 0. Let G denote a finite group. Show that (a) If α: GH is a group homomorphism, then img for all k. (b) img for each k. (c) G is img-solvable if and only if img{1} for some k. (d) Every subgroup of a img-solvable group is img -solvable. (e) G is img-solvable if and only if img for all subgroups H≠{1} of G.

9.3 Nilpotent Groups

If G is a group, the definition of the derived subgroup G′ guarantees that G is abelian if and only if G′ = {1}. If the process of taking the derived subgroup is iterated, the derived series G = G(0)G(1)G(2)img is obtained, and G is solvable if and only if this series (of normal subgroups of G) reaches {1} in a finite number of steps (Theorem 2 §9.2). Note that G(1) = G′. Now the center Z(G) plays an analogous role to G′ in the sense that G is abelian if and only if Z(G) = G. In view of this, an irresistible question arises: Is there a way to iterate the formation of the center so as to create a series {1} = Z0Z1Z2img of normal subgroups of G (with Z(G) = Z1) such that G is solvable if and only if this series reaches G in a finite number of steps? The answer is yes and no. Yes, there is a natural way to define such a series. No, it does not characterize the solvable groups in this way. Rather, it characterizes a smaller class of groups called the nilpotent groups. In this section, we define these groups and show that the finite ones are precisely the finite groups that are isomorphic to the direct product of their Sylow subgroups.

Central Series

If G is a group, define a series Z0(G), Z1(G), Z2(G), . . . of normal subgroups of G inductively as follows:

1. Take Z0(G) = {1}.

2. If Zi(G) img G has been constructed, define Zi+1(G) the unique normal subgroup of G that contains Zi(G) and satisfies img

Then Zi+1(G) ⊇ Zi(G) and Zi(G) img Zi+1(G) for each i ≥ 0, and the series

{1} = Z0(G)⊆ Z1(G) ⊆ Z2(G) ⊆ img

is called the ascending central series of G. Note that

img

The ascending central series may never reach G, even if G is solvable:

Example 1. Suppose Z(G) = {1} where G ≠ {1}—for example the solvable group S3. Then Z1(G) = Z(G) = {1}, so we have img Hence, Z2(G) = {1}, and this process continues inductively to show that Zk(G) = {1} for all k.

To characterize the groups G for which the ascending central series reaches G, it is useful to define a related descending central series. This requires a new notion. Recall that the derived subgroup G′ is generated by the commutators [a, b] = aba−1b−1 in G. We extend this idea as follows. If H and K are subgroups of a group G, define

img

to be the subgroup generated by the commutators [h, k], with h img H and k img K. Note that [h, k]−1 = [k, h] for all h img H and k img K. Hence, [H, K] = [K, H], and this group consists of all products of commutators of the form [h, k] or [k, h], where h img H and k img K. In particular, [G, G] = G′. Lemma 1 collects several other useful facts about these subgroups.

Lemma 1. Let H, K, H1, and K1 be subgroups of a group G.

1. [H, K] = [K, H].

2. If HH1 and KK1, then [H, K] ⊆ [H1, K1].

3. If H img G and K img G, then [H, K] img G.

4. H img G if and only if [H, G] ⊆ H.

5. Suppose that KHG and K img G. Then H/KZ(G/K)  if and only if   [H, G] ⊆ K.

Proof. We prove (5) and leave the rest as Exercise 2. If img, then we have hKgK = gKhK for all g img G and h img H; that is, [h, g] img K. Hence, [H, G] ⊆ K. Since this argument works in reverse, we have proved (5).

Now, given a group G, define a series Γ0(G), Γ1(G), Γ2(G), . . . of subgroups of G inductively as follows:

1. Take Γ0(G) = G.

2. If Γi(G) has been constructed, define Γi+1(G) = [Γi(G), G].

Then Γi(G) img G for all i ≥ 0 by induction on i (using (3) of Lemma 1), whence Γi+1(G) = [Γi(G), G] ⊆ Γi(G) for each i by (4) of Lemma 1. Thus, we obtain a series of normal subgroups

G = Γ0(G) ⊇ Γ1(G) ⊇ Γ2(G) ⊇ img.

This is called the descending central series of G. The name comes from the fact that, using (5) of Lemma 1,

img

holds for each i ≥ 0. Note that

Γ1(G) = [G, G] = G′ is the derived subgroup of G.

If G is an abelian group, then Z1(G) = G and Γ1(G) = {1}. On the other hand, there are groups (even solvable ones by Example 1) for which the ascending central series does not reach G and the descending central series does not reach 1. However, if either possibility occurs, so does the other.

Lemma 2. The following are equivalent for a group G and an integer n ≥ 0:

1. Γn(G) = {1}.

2. Zn(G) = G.

3. A series G = G0G1imgGn = {1} exists with Gi img G for each i and Gi/Gi+1Z(G/Gi+1).

Proof. Write Γi(G) = Γi and Zi(G) = Zi for each i.

(1) ⇒ (2). If Γn = {1}, we show that ΓniZi for each i = 0, 1, 2, . . . (so Zn = G). This is clear if i = 0 by (1), so assume ΓniZi, where i > 0. If a img Γni−1, then, for all g img G, [a, g] imgni−1, G] = ΓniZi. Thus, aZi is in the center of G/Zi and so a img Zi+1. Hence, Γni−1Zi+1 as required.

(2) ⇒ (3). Given (2), use G = ZnZn−1imgZ0 = {1} in (3).

(3) ⇒ (1). Given (3), we show that ΓiGi for each i = 0, 1, 2, . . . (so Γn = {1}). This is clear if i = 0, so assume that ΓiGi for some i > 0. We must show that [Γi, G] = Γi+1Gi+1, so we show that [a, g] img Gi+1 for all a img Γi, g img G. But img by (3), so a img ΓiGi implies that aGi+1 commutes with gGi+1 for all g img G. This in turn implies that [a, g] img Gi+1, as required.

A group G is called a nilpotent group if the conditions in Lemma 2 are satisfied for some n ≥ 0. The smallest integer n for which Γn(G) = {1}, equivalently Zn(G) = G, is called the nilpotency class of G. Thus, if G is nilpotent, then G has class 1 if and only if it is abelian, and (Exercise 11) G has class 2 if and only if it is nonabelian and G′ ⊆ Z(G).

A series as in (3) of Lemma 2 is called a central series for G. Suppose that G = G0G1imgGn = {1} is any central series for a nilpotent group G. Then the proof that (3) ⇒ (1) in Lemma 2 derives the first of the following inclusions:

img

We leave the second inclusion for the reader (Exercise 7). Hence, we often call the series G0(G) ⊇ Γ1(G) ⊇ Γ2(G) ⊇ img and 1 = Z0(G)⊆ Z1(G) ⊆ Z2(G) ⊆ img the lower and upper central series, respectively.

Example 2. Every abelian group is nilpotent.

Example 3. If p is a prime, every finite p -group is nilpotent. In fact, if we write Zi(G) = Zi for each i, Theorem 6 §8.2 shows that Zi+1/Zi = Z(G/Zi) is not trivial if ZiG because G/Zi is a p-group. Hence, 1 ⊂ Z1Z2img, which eventually reaches G because G is finite.

Example 4. Show that every nilpotent group G is solvable, but not conversely.

Solution. If G is nilpotent, the series {1} = Z0(G) ⊆ imgZn(G) = G is a solvable series. However, S3 is solvable but not nilpotent by Example 1.

Theorem 1. Every subgroup and image of a nilpotent group is again nilpotent.

Proof. Let G be nilpotent. To show that a subgroup KG is nilpotent, it suffices to show that Γi(K) ⊆ Γi(G) for each i. This is clear if i = 0. If Γi(K) ⊆ Γi(G) for some i, then the induction goes through because

Γi+1(K) = [Γi(K), K] ⊆ [Γi(G), G] = Γi+1(G).

Now let α: GH be an onto homomorphism; we show that Γi(H) ⊆ αi(G)] for each i. If i = 0, it is because α is onto. In general, let Γi(H) ⊆ αi(G)), and let y img Γi(H) and h img H. Write y = α(x), where x img Γi(G), and (as α is onto) write h = α(g), g img G. Then

[y, h] = [α(x), α(g)] = α[x, g] img αi(G), G] = αi+1(G)).

Hence, Γi+1(H) = [Γi(H), H] ⊆ αi+1(G)), as required.

Corollary. A group G is nilpotent if and only if G′ is nilpotent.

Proof. Write D = G′. If D(n) = {1}, then G(n+1) = D(n) = {1}.

By Theorem 1, if K img G and G is nilpotent, then both K and G/K are nilpotent. The converse is false (S3 is again a counterexample) in contrast to the situation for solvable groups. However, the converse does hold when KZ(G).

Theorem 2. If G is a group, KZ(G), and G/K is nilpotent, then G is nilpotent.

Proof. Assume that G/K is nilpotent, and let θ: GG/K be the coset map. By hypothesis, let θ(G) = X0X1imgXn = {K} be a central series for θ(G), where img for 0 ≤ in − 1. Write Xi = Gi/K = θ(Gi) for 0 ≤ in. Then we obtain the series

G = G0G1imgGn = KGn+1 = {1}.

We show this is a central series for G. First, img because KZ(G). To see that img for 0 ≤ i < n, let a img Gi. Then θ(a) img θ(Gi) = Xi, so θ(a) Xi+1 commutes with θ(g) Xi+1 for all g img G. Thus, θ[a, g] = [θ(a), θ(g)] img Xi+1 = θ(Gi+1), say θ[a, g] = θ(b), b img Gi+1. Thus, [a, g]b−1 img ker θ = KGi+1, so [a, g] img Gi+1b = Gi+1. This means aGi+1 commutes with gGi+1, that is, img, as required.

The next result will be needed in Theorem 4.

Theorem 3. If G1, G2, img, Gn are nilpotent, so also is G1 × G2 × img × Gn.

Proof. This follows because Γi(G1 × G2 × img × Gn) ⊆ Γi(G1) × img × Γi(Gn) for each i, a fact that we leave as Exercise 6.

Theorem 3 and Example 3 combine to show that any finite direct product of finite p-groups (for various primes p) is nilpotent. In fact, every finite nilpotent group is isomorphic to such a direct product. We need the following notion.

A subgroup M of a group G is said to be maximal in G if MG and the only subgroups H such that MHG are H = M and H = G. Clearly, every proper subgroup K of a finite group is contained in a maximal subgroup—one of maximal order containing K. If G is finite, every subgroup of prime index is maximal by Example 6 §2.6.106 The converse is not necessarily true (any subgroup of index 4 in A4 is maximal), but it does hold in a finite p-group. Moreover, in this case, the maximal subgroups (of index p) are necessarily normal (see the corollary to Theorem 1 §8.3). This property characterizes the finite nilpotent groups.

Theorem 4. Burnside–Wielandt Theorem. 107 The following conditions are equivalent for a finite group G ≠ {1}:

1. G is nilpotent.

2. N(H) ≠ H for all subgroups HG of G.

3. Every maximal subgroup of G is normal in G.

4. Every Sylow subgroup of G is normal in G.

5. G is isomorphic to the direct product of its Sylow subgroups.

Proof. (1) ⇒ (2). Write Zi = Zi(G) for each i and assume that Zn = G. If HG is a subgroup of G, then Z0H but Zn img H, so an integer k ≥ 0 exists such that ZkH but Zk+1 img H. Choose a img Zk+1, aH. Then aZk is in the center of G/Zk, so if h img H, aZk and hZk commute. Hence, hah−1a−1 img ZkH, from which aHa−1H. Thus, a img N(H), and so N(H) ≠ H.

(2) ⇒ (3). Let M be a maximal subgroup of G. Since MN(M) ⊆ G, (2) implies that N(M) = G. Hence, M img G.

(3) ⇒ (4). Suppose P is a nonnormal Sylow p-subgroup of G. Then N(P) ≠ G, so let N(P) ⊆ M, where M is a maximal subgroup of G. Because PM, (3) gives aPa−1aMa−1 = M for all a img G. Hence, both P and aPa−1 are Sylow p-subgroups of M and so are conjugate in M, say P = m(aPa−1)m−1 for some m img M. But then ma img N(P), so a img M. Because a img G was arbitrary, this means GM, a contradiction. This proves (4).

(4) ⇒ (5). Let P1, P2, img, Pr denote the distinct Sylow subgroups of G.

Claim 1. P1P2 img PkP1 × P2 × img × Pk for each k = 2, 3, img, r.

Proof. It is clear if k = 1. Assume inductively that P1P2 img PkP1 × P2 × img × Pk for some k > 1. Then (P1P2 img Pk) ∩ Pk+1 = {1} because elements in the two subgroups have relatively prime orders. By Theorem 6 §2.8,

(P1P2 img Pk)Pk+1 ≅ (P1 × P2 × img × Pk) × Pk+1P1 × P2 × img × Pk+1

because P1P2 img Pk img G and Pk+1 img G. This proves the claim.

The claim gives |P1P2 img Pr| ≅ |P1||P2| img |Pr| = |G|. Hence G = P1P2 img Pr and (5) follows, again by the Claim.

(5) ⇒ (1). This follows from Theorem 3 and Example 3.

It is interesting to compare (2) in Theorem 4 with the result (Theorem 5 §9.2) that a finite group G is solvable if and only if H′ ≠ H for every subgroup H ≠ {1}.

Since every finite abelian group is nilpotent, the implication (1) ⇒ (5) in Theorem 4 gives another proof of the primary decomposition theorem for finite abelian groups (Corollary 2 of Theorem 3 §7.2). We reformulate (1)img(5) as

Corollary 1. A finite group G is nilpotent if and only if G is isomorphic to a finite direct product of p-groups for various primes p.

Frattini and Fitting Subgroups

One of the most important aspects of the study of nilpotent groups is that every finite group G contains a nilpotent subgroup Φ, which is characteristic in G (that is, σ(Φ) = Φ for every automorphism σ of G—these subgroups are discussed in Corollary 3 of Theorem 3 §2.8). We now turn to a discussion of this.

If G ≠ {1} is a finite group, define the Frattini subgroup

Φ(G) = img {MG img M is a maximal subgroup of G}.

Define Φ{1} = {1}. This was introduced in 1885 by Giovanni Frattini.

Example 5. Φ(A4) = {ε}. Indeed, K = {ε, (12)(34), (13)(24), (14)(23)} is maximal, being of index 3, and M = {ε, (123), (132)} is maximal (it has index 4, but A4 has no subgroup of index 2). Hence, we have Φ(A4) ⊆ KM = {ε}.

Example 6. If Q = { ± 1, ± i, ± j, ± k} is the quaternion group, then Φ(Q) = {1, − 1} because img and img are the only maximal subgroups.

Example 7. If img and o(a) = pn, where p is a prime, img because img is the unique maximal subgroup (of index p).

Theorem 5. Let G be a group and write Φ = Φ(G). Then the following hold:

1. If α: GH is an onto group homomorphism, then α(Φ) ⊆ Φ(H).

2. In particular, Φ is a characteristic subgroup of G.

3. G is nilpotent if and only if G′ ⊆ Φ.

Proof. (1) If UH is a maximal subgroup, we must show that α(Φ) ⊆ U. If we define M = {m img G img α(m) img U}, then it suffices to show that M is a maximal subgroup of G (since then Φ ⊆ M). But if MKG are subgroups of G then α(M) ⊆ α(K) ⊆ α(G). Since α is onto, this is Uα(K) ⊆ H, so α(K) = U or σ(K) = H. These imply that K = M or K = G.

(2) This follows from (1) if H = G and α is an automorphism of G.

(3) G′ ⊆ Φ if and only if G′ ⊆ M for each maximal subgroup M of G, if and only if M img G for each M, and if and only if G is nilpotent by Theorem 4.

Corollary 1. The following are equivalent for a finite group G:

1. G is nilpotent.

2. G/Φ(G) is abelian.

3. G/Φ(G) is nilpotent.

Proof. (1)img(2) restates (3) of Theorem 5 and (2)⇒(3) is obvious.

(3)⇒(1). Given (3), we show that every maximal subgroup M of G is normal. Write Φ = Φ(G). Since Φ ⊆ M, the subgroup M/Φ is maximal in G/Φ by the correspondence theorem, so Mimg G/Φ by (3) and Theorem 4. But then M img G, again by the correspondence theorem.

To see that Φ(G) is a nilpotent group, we first characterize it in terms of the following concept. An element t img G is called a nongenerator in G if it can be omitted from any generating set X of G; that is, if img, then img

Theorem 6. Let G denote any finite group. Then the following hold:

1. Φ(G) = {t img t is a nongenerator of G}.

2. Φ(G)is a nilpotent group.

Proof. For convenience, write Φ = Φ(G).

(1) Write N = {t img t is a nongenerator of G} and let a img Φ. If aN, then XG exists such that img but img So let img where M is a maximal subgroup of G. Then a img M because Φ ⊆ M, whence img a contradiction. Hence, Φ ⊆ N. Conversely, if t img N and M is a maximal subgroup of G, then g img M (otherwise img). Hence, N ⊆ Φ.

(2) By Theorem 4 we show that every Sylow p-subgroup P of Φ is normal in G. If g img G, then gPg−1gΦg−1 = Φ by (2) of Theorem 5. Hence, both gPg−1 and P are Sylow p-subgroups of Φ and so are conjugate in Φ, say t(gPg−1)t−1 = P, where t img Φ. Thus, tg img N(P), which yields G = ΦN(P). But then img so, as Φ is finite, img by (1). Hence, P img G as required.

Note the proof of (2) shows that Sylow p-subgroups of Φ(G) are normal in G.

Corollary 1. Assume that Φ(G) is finitely generated (for example, if G is finite). If HΦ(G) = G, where H is a subgroup, then H = G.

Proof. Write Φ(G) = Φ. If img then img and the nongenerators ti can be removed one by one.

The next result extends a useful theorem about finite p-groups.

Theorem 7. If G is a nilpotent group and {1} ≠ H img G, then HZ(G) ≠ {1}.

Proof. Write Γi = Γi(G) for each i. We have G = Γ0 ⊇ Γ1img ⊇ Γn = {1} for some n, so H ∩ Γn = {1} while H ∩ Γ0 ≠ {1}. So there exists k such that H ∩ Γk ≠ {1} while H ∩ Γk+1 = {1}. Choose 1 ≠ h img H ∩ Γk. If g img G, then [h, g] imgk, G] = Γk+1. Also, [h, g] = h−1g−1hg img H(g−1Hg) = H because H img G. So [h, g] img H ∩ Γk+1 = {1}, whence hg = gh. Thus, h img Z(G) ∩ H.

In 1938, Hans Fitting identified a largest nilpotent normal subgroup in every finite group. His key result was

Theorem 8. Fitting's Theorem. If H and K are nilpotent, normal subgroups of a finite group G, so also is HK.

Proof. We proceed by induction on We have HK img G, so we may assume (by induction) that G = HK and that H ≠ {1} ≠ K. Write W = Z(K). Then W ≠ {1} by Theorem 7. Also, W img G being characteristic in K img G. If we write N = [W, H], the proof falls into two cases.

Case 1. N = {1}. Then W centralizes H (and K), so WZ(HK) = Z(G). But img img Moreover, img and img are both nilpotent by Theorem 1, so img is nilpotent by induction. Hence, G is nilpotent by Theorem 2.

Case 2. N ≠ {1}. We have NWH because W img G and H img G. In particular, V = NZ(H) ≠ {1} again by Theorem 7. As before, img img and both img and img are nilpotent, so img is nilpotent by induction. But V centralizes H, and it also centralizes K because VNW = Z(K). Hence, VZ(HK) = Z(G), so G is nilpotent by Theorem 2 and we are done in this case too.

Now let G be any finite group. If N1 = {1}, N2, . . ., Nk denote all the nilpotent, normal subgroups of G, define the Fitting subgroup F(G) of G by

img

Then F(G) img G, and it is nilpotent by Theorem 8 and induction on k. This proves

Theorem 9. If G is a finite group, then F(G) is the largest nilpotent, normal subgroup of G in the sense that it contains every such subgroup.

Corollary 1. If α: GH is an onto homomorphism, then α[F(G)] ⊆ F(H).

Proof. α[F(G)] img H because α is onto, and it is nilpotent by Theorem 1.

Hence, a finite group G is nilpotent if and only if F(G) = G. Clearly, Z(G) ⊆ F(G), and Φ(G) ⊆ F(G) because it is nilpotent (Theorem 6) and normal in G (Theorem 5). Moreover, F(G) is a characteristic subgroup of G by the above corollary.

Lemma 3. If G is finite and N img G, then N is nilpotent if and only if N′ ⊆ Φ(G).

Proof. Write Φ(G) = Φ. If N′ ⊆ Φ, then N′ is nilpotent by Theorem 6, and so N is nilpotent by the corollary to Theorem 1. Conversely, if N is nilpotent, then N′ ⊆ Φ(N) by Theorem 5, and it remains to show that Φ(N) ⊆ Φ. Write Φ(N) = H and suppose H6 ⊆ Φ. Then H6 ⊆ M for some maximal subgroup M of G, so HM = G. Since HN, the modular law (Lemma 1 §8.1) gives

N = GN = HMN = H(MN).

Thus, img and H = Φ(N) consist of (a finite number of) non-generators of N. Hence, N = MN, whence NM, a contradiction.

We can now describe the relationship between the Frattini and Fitting subgroups in a finite group.

Theorem 10. Let G be a finite group and write Φ = Φ(G) and F = F(G).

1. F′ ⊆ Φ ⊆ F.

2. F(G/Φ) = F/Φ.

Proof. (1) We have Fimg G because it is characteristic in F img G, so F′ ⊆ Φ by Lemma 3 because F is nilpotent. On the other hand, Φ ⊆ F by Theorem 9 because Φ is nilpotent. This proves (1).

(2) This follows from a more general result (Theorem 11).108

Theorem 11. If G is a finite group and Φ(G) ⊆ N img G, then N is nilpotent if and only if N/Φ(G) is nilpotent.

Proof. Write Φ(G) = Φ. If N is nilpotent, so is its image img For the converse, assume that img is nilpotent. To show that N is nilpotent, we show that every Sylow p -subgroup P of N is normal in N (and invoke Theorem 4). First, img is a Sylow p-subgroup of img (by Example 4 §8.4), whence img by hypothesis. But then img is characteristic in img and it follows that img Thus, ΦP img G. But P is a Sylow p-subgroup of ΦP (because ΦPN), so G = (ΦP)NG(P) by Lemma 1 §8.4. Since G is finite and Φ consists of nongenerators, it follows that G = P NG(P) = NG(P). Hence, P img G, so certainly P img N as required.

There is much more information available on nilpotent groups in books on group theory.109

Exercises 9.3

1.

a. Show that An is not nilpotent if n ≥ 3.

b. Show that every nilpotent group is solvable, but not conversely.

2. Prove (1)–(4) in Lemma 1.

3. If H and K are subgroups of G and α: GG1 is a homomorphism, show that α[H, K] = [α(H), α(K)]. Conclude that [H, K] is normal in G (characteristic in G) if the same is true of H and K.

4. If α: GH is any homomorphism, show that αi(G)] = Γi[α(G)].

5. If G(k) is the kth derived subgroup of G (see Theorem 1, Section 9.2), show that G(k+1) = [G(k), G(k)] for each k ≥ 0.

6.

a. Show that Γi(G1 × img × Gn) ⊆ Γi(G1) × img × Γi(Gn) for all i ≥ 0.

b. Show that equality holds in (a).

7. If G = G0G1imgGn= {1} is any central series for a group G, show that GniZi(G) for each i.

8. Let

img

where F is a field. Is G nilpotent?

9. Show that Dn is nilpotent if and only if n is a power of 2.

10. Show that a finite group is nilpotent if and only if any two elements of relatively prime orders commute.

11. Show that a group G is nilpotent of class 2 if and only if G is nonabelian and G′ ⊆ Z(G).

12. Show that a finite group G is nilpotent if and only if Z(G/K) is nontrivial for all K img G, KG. [Hint: Theorem 7.]

13. If G is a finite nilpotent group, let K be of minimal order in {K img {1} ≠ K img G}. Show that KZ(G) and that |K| is a prime. [Hint: Theorem 7.]

14. If H img G and K img G, show that G/(HK) is nilpotent if and only if the product group (G/H) × (G/K) is nilpotent.

15. Show that a finite group G is nilpotent if and only if G has a normal subgroup of order m for every divisor m of |G|.

16. A subgroup H of a group G is called subnormal in G if a chain of subgroups H = H0H1imgHn = G exists such that Hi img Hi+1 for each i. Show that a finite group G is nilpotent if and only if every subgroup is subnormal.

17. If KZ(G) and G/K is nilpotent, show that G is nilpotent using only (3) of Theorem 4.

18.

a. If G is nilpotent, show that Z(H)≠{1} for all subgroups H ≠ {1}.

b. Show that the converse is false by considering Q6.

19. If G is nilpotent and G/G′ is cyclic, show that G is abelian. [Hint: Apply Theorem 2 §2.9 to G/[Γ2(G)] and conclude that Γ2(G) = Γ1(G).]

20. If G is a finite group show that (2) img (3) img (4) and (2) ⇒ (1), but (1) ⇒/(2).

(1) G′ is abelian. (2) G/Z(G) is abelian. (3) Γ2(G) = {1}. (4) Z2(G) = G.

21. Let img where o(a) = n, o(b) = 2, and aba = b. Show that (a) Φ(D4) = {1, a2},  (b) Φ(D12) ⊆ {1, a6},  and (c) Φ(Dpq) ={1}, where pq are primes. [Hint: If img has index m, show that img is a subgroup of index m.]

22. If o(a)img where pi are distinct primes, show that img, where m = p1p2 img pr.

23. Let |G| = p3, where p is a prime. If G is nonabelian, show that Φ(G) = G′ = Z(G) and that this subgroup has order p. [Hint: Exercise 26 §8.2.]

24. If G is a finite group, write Φ = Φ(G). Show that the following are equivalent:

(1) G is nilpotent (2) G/Φ is abelian (3) G/Φ is nilpotent

25. If K img G, where G is finite, and Φ(G/K) = {K}, show that Φ(G) ⊆ K.

26.

a. If G is finite, K img G, and K ⊆ Φ(G), show that Φ(G/K) = Φ(G)/K.

b. Show that Φ(G/Φ(G)) = {1}.

27. Show that Φ(G × H) = Φ(G) × Φ(H) for finite groups G and H.

28. If G is a finite group and H img G, show that Φ(H) ⊆ H ∩ Φ(G). [Hint: If Φ(H) img M, where M is maximal in G, show that G = Φ(H)M and apply Lemma 1 §8.1.]

29.

a. If G is a finite group and MG is a maximal subgroup, show that either Z(G) ⊆ M or G′ ⊆ M. [Hint: MZ(G) = G implies that M img G.]

b. Show that Z(G) ∩ G′ ⊆ Φ(G) for all finite groups G.

30. Show that a finite group G can be generated by n elements if and only if the same is true of G/Φ(G).

31. If G is a finite p-group, p a prime, show that img

Notes

100. Due to Hans Zassenhaus.

101. Due to Otto Schreier.

102. These groups are also called soluble groups.

103. See Kargapolov, M.I. and Merzljakov, J.I., Fundamentals of the Theory of Groups, Springer-Verlag, 1979; Rotman, J.J., The Theory of Groups: An Introduction, 2nd ed., Boston: Allyn & Bacon, 1973; Artin, E., Geometric Algebra, New York: Interscience, 1957. For an elementary proof when n = 2, see Lang, S., Undergraduate Algebra, Berlin: Springer-Verlag, 1987.

104. Feit, W. and Thomson, J.G., Solvability of groups of odd order, Pacific Journal of Mathematics, 13 (1963), 775–1029.

105. See Kargapolov, M.I. and Merzljakov, J.I., Fundamentals of the Theory of Groups, Springer-Verlag, 1979; MacDonald, I.D., The Theory of Groups, London: Oxford University Press, 1968; Rotman, J.J., The Theory of Groups: An Introduction, 2nd ed., Boston: Allyn & Bacon, 1973.

106. This is also true if G is infinite by Exercise 31, §2.6.

107. The name honors William Burnside and Helmut Wielandt.

108. Because of (2) and the corollary to Theorem 9, the Fitting subgroup of G is also called the nilpotent radical of G.

109. The following books contain excellent introductions to the theory of nilpotent groups: Mac-Donald, I.D., The Theory of Groups, London: Oxford University Press, 1968; Rose, J.S., A Course on Group Theory, Cambridge, England: Cambridge University Press, 1978; Kargapolov, M.I. and Merzljakov, J.I., Fundamentals of the Theory of Groups, New York: Springer-Verlag, 1979; Gorenstsein, D., Finite Groups, 2nd ed., New York: Chelsea, 1980.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
3.135.213.214