+ All Categories
Home > Documents > LINEAR GROUPS AND GROUP RINGSpassman/lingroup.pdf · 2005. 10. 4. · LINEAR GROUPS AND GROUP RINGS...

LINEAR GROUPS AND GROUP RINGSpassman/lingroup.pdf · 2005. 10. 4. · LINEAR GROUPS AND GROUP RINGS...

Date post: 06-Sep-2021
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
21
LINEAR GROUPS AND GROUP RINGS J. Z. GONC ¸ALVES AND D. S. PASSMAN Abstract. This paper consists of two parts. The first is concerned with free products in linear groups and uses the usual “ping pong” lemma and attractors to prove the results. What is new here is that we allow certain subspaces of V associated with the semisimple and generalized transvection operators to have dimensions larger than 1. The second part is concerned with applications of this machinery to integral groups rings Z[G] of finite groups. We show, for example, that if G is nonabelian of order prime to 6, then Z[G] contains two Bass cyclic units that generate a nonabelian free group. 1. Linear operators and attractors Let F be a field and let || : F R + = {r R : r 0} be an absolute value defined on F . Here R is the field of real numbers and, by definition, we have |ab| = |a|·|b| |a + b|≤|a| + |b| |a| =0 ⇐⇒ a =0 for all a, b F . In particular, |1| =1= |- 1|. Indeed, |a| = 1 if a is any root of unity in F . See [J, Chapter 9] or [B, Chapter VI] for additional basic properties. If we define δ : F × F R + by δ(a, b)= |a - b|, then F clearly becomes a metric space using δ as a metric. We assume throughout that F is locally compact in this topology, so that each a F has a neighborhood with compact closure. As a consequence of the product formula |ab| = |a|·|b|, it follows that every closed, bounded subset of F is compact. Furthermore, || : F R + is a continuous function and F is complete, in that every Cauchy sequence has a limit. If || is archimedean, then Ostrowski’s Theorem [J, page 552] implies that F = R or C, the field of complex numbers, and that || is the ordinary absolute value defined on C. In particular, || is the identity function on |F | = R + F , where |F | is the set of absolute values taken on by elements of F . Thus, if 0 = a F , then a/|a| is an element of F having absolute value 1. On the other hand, if || is non-archimedean, then we know that |a + b|≤ max{|a|, |b|} for all a, b F . Indeed, if |a| = |b|, then |a + b| = max{|a|, |b|}. In this situation, O = {a F : |a|≤ 1} is a subring of F , the valuation ring associated with ||, and O has the unique maximal ideal m = {a F : |a| < 1}. Since F is locally compact, O is compact and hence the residue class field O/m is finite, since the 2000 Mathematics Subject Classification. 16S34, 16U60, 20E06, 20H20. The first author’s reasearch was supported in part by CNPq grant 303.756/82-5 and Fapesp- Brazil, Proj. Tematico 00/07.291-0. The second author’s research was supported in part by NSA grant 144-LQ65. 1
Transcript
Page 1: LINEAR GROUPS AND GROUP RINGSpassman/lingroup.pdf · 2005. 10. 4. · LINEAR GROUPS AND GROUP RINGS J. Z. GONC¸ALVES AND D. S. PASSMAN Abstract. This paper consists of two parts.

LINEAR GROUPS AND GROUP RINGS

J. Z. GONCALVES AND D. S. PASSMAN

Abstract. This paper consists of two parts. The first is concerned with free

products in linear groups and uses the usual “ping pong” lemma and attractors

to prove the results. What is new here is that we allow certain subspaces of Vassociated with the semisimple and generalized transvection operators to have

dimensions larger than 1. The second part is concerned with applications ofthis machinery to integral groups rings Z[G] of finite groups. We show, forexample, that if G is nonabelian of order prime to 6, then Z[G] contains twoBass cyclic units that generate a nonabelian free group.

1. Linear operators and attractors

Let F be a field and let | | : F → R+ = {r ∈ R : r ≥ 0} be an absolute valuedefined on F . Here R is the field of real numbers and, by definition, we have

|ab| = |a|·|b||a + b| ≤ |a|+ |b||a| = 0 ⇐⇒ a = 0

for all a, b ∈ F . In particular, |1| = 1 = | − 1|. Indeed, |a| = 1 if a is any root ofunity in F . See [J, Chapter 9] or [B, Chapter VI] for additional basic properties.

If we define δ : F × F → R+ by δ(a, b) = |a − b|, then F clearly becomes ametric space using δ as a metric. We assume throughout that F is locally compactin this topology, so that each a ∈ F has a neighborhood with compact closure.As a consequence of the product formula |ab| = |a|·|b|, it follows that every closed,bounded subset of F is compact. Furthermore, | | : F → R+ is a continuous functionand F is complete, in that every Cauchy sequence has a limit.

If | | is archimedean, then Ostrowski’s Theorem [J, page 552] implies that F = Ror C, the field of complex numbers, and that | | is the ordinary absolute valuedefined on C. In particular, | | is the identity function on |F | = R+ ⊆ F , where |F |is the set of absolute values taken on by elements of F . Thus, if 0 6= a ∈ F , thena/|a| is an element of F having absolute value 1.

On the other hand, if | | is non-archimedean, then we know that

|a + b| ≤ max{|a|, |b|}for all a, b ∈ F . Indeed, if |a| 6= |b|, then |a + b| = max{|a|, |b|}. In this situation,O = {a ∈ F : |a| ≤ 1} is a subring of F , the valuation ring associated with | |,and O has the unique maximal ideal m = {a ∈ F : |a| < 1}. Since F is locallycompact, O is compact and hence the residue class field O/m is finite, since the

2000 Mathematics Subject Classification. 16S34, 16U60, 20E06, 20H20.The first author’s reasearch was supported in part by CNPq grant 303.756/82-5 and Fapesp-

Brazil, Proj. Tematico 00/07.291-0. The second author’s research was supported in part by NSA

grant 144-LQ65.

1

Page 2: LINEAR GROUPS AND GROUP RINGSpassman/lingroup.pdf · 2005. 10. 4. · LINEAR GROUPS AND GROUP RINGS J. Z. GONC¸ALVES AND D. S. PASSMAN Abstract. This paper consists of two parts.

2 J. Z. GONCALVES AND D. S. PASSMAN

cosets of m yield an open cover of O. Furthermore, for any real number 0 < r < 1,mr = {a ∈ F : |a| < r} is an open ideal of O and we see again that O/mr is finite.It follows that there are only finitely many different values for |F | between r and1. In particular, |F | takes on a unique largest value < 1, and with this, we see thatthe valuation is discrete.

Conversely, if the residue class field is finite and if | | is a complete discrete valu-ation, then [B, Proposition VI.5.2] implies that F is locally compact. Furthermore,by [B, Theorem VI.9.1], this occurs if and only if F is either a finite algebraicextension of a p-adic field or the field of formal Laurent series over a finite field.

Now let V = Fn be the vector space of n-tuples of F . If v = (a1, a2, . . . , an) ∈ V ,we define the norm of v by

‖v‖ = max {|ai| : i = 1, 2, . . . , n}.Then, it is easy to see that

‖av‖ = |a|·‖v‖‖v + w‖ ≤ ‖v‖+ ‖w‖‖v‖ = 0 ⇐⇒ v = 0

for all a ∈ F and v, w ∈ V . Thus, V becomes a metric space using the metricδ(v, w) = ‖v − w‖ and, since F is locally compact, so is V . Again, as a consequenceof the product formula ‖av‖ = |a|·‖v‖, it follows that every closed, bounded subsetof V is compact. In particular, the unit sphere in V given by S = {v ∈ V : ‖v‖ = 1}is compact since ‖ ‖ : V → R+ is continuous. Note that ‖V ‖ = |F | so that for each0 6= v ∈ V there exists a ∈ F with ‖v‖ = |a|. In particular, v/a has norm 1 andhence Fv ∩ S 6= ∅.

We say that X ⊆ V is a projective subset of V if X contains a nonzero vectorand if it is closed under multiplication by F • = F \ 0, that is F •X ⊆ X. Thus,except for the possible presence or absence of the zero vector, these projectivesubsets correspond in a one-to-one manner to the nonempty subsets of P(V ), theprojective space of V . Because of this, we say that projective subsets X and Y aredisjoint if X ∩ Y ⊆ {0}. Furthermore, we define the distance between them by

d(X, Y ) = inf{‖x− y‖ : x ∈ X ∩ S, y ∈ Y ∩ S}.If 0 6= v, w ∈ V , we set

d(v, w) = d(Fv, Fw) = inf{‖av − bw‖ : a, b ∈ F, ‖av‖ = 1 = ‖bw‖}.As we observed, field elements a and b do indeed exist with ‖av‖ = 1 = ‖bw‖.Finally, we define

d(Y, v) = d(v, Y ) = d(Fv, Y ) = inf{d(v, y) : y ∈ Y }.

Lemma 1.1. With the above notation, we havei. If X and Y are nonzero subspaces of V , then there exist x0 ∈ X ∩ S and

y0 ∈ Y ∩S with d(X, Y ) = ‖x0 − y0‖. In particular, if X and Y are disjoint,then d(X, Y ) > 0.

ii. The distance function d defines a metric on the projective space P(V ). Withrespect to this metric, P(V ) has diameter ≤ 2. It has diameter ≤ 1 if | | isa non-archimedean absolute value.

iii. If 0 6= v, w ∈ V , then d(v, w) ≤ 2·‖v − w‖/‖v‖. If | | is non-archimedean,then d(v, w) ≤ ‖v − w‖/‖v‖.

Page 3: LINEAR GROUPS AND GROUP RINGSpassman/lingroup.pdf · 2005. 10. 4. · LINEAR GROUPS AND GROUP RINGS J. Z. GONC¸ALVES AND D. S. PASSMAN Abstract. This paper consists of two parts.

LINEAR GROUPS AND GROUP RINGS 3

Proof. (i) This is clear since X ∩ S and Y ∩ S are compact, and since the map(X ∩ S)× (Y ∩ S) → R+ given by (x, y) 7→ ‖x− y‖ is continuous.

(ii) Let u, v, w be nonzero vectors in V . By (i), there exist a, b, b′, c′ ∈ F such that‖au‖ = ‖bv‖ = ‖b′v‖ = ‖c′w‖ = 1, d(u, v) = ‖au− bv‖ and d(v, w) = ‖b′v − c′w‖.Then |b| = |b′| and therefore (b/b′)(b′v − c′w) has the same norm as b′v − c′w. Inother words, if we set c = (b/b′)c′ ∈ F , then ‖cw‖ = 1, d(v, w) = ‖bv − cw‖ and

d(u, v) + d(v, w) = ‖au− bv‖+ ‖bv − cw‖ ≥ ‖au− cw‖ ≥ d(u, w).

Furthermore, if u and v correspond to distinct points in P(V ), then Fu 6= Fvand hence Fu ∩ Fv = 0. It now follows from (i) that d(u, v) > 0. Finally, forany 0 6= u, v ∈ V , we have, as above, d(u, v) = ‖au− bv‖ ≤ ‖au‖ + ‖bv‖ = 2 soP(V ) has diameter at most 2. In the non-archimedean case, d(u, v) = ‖au− bv‖ ≤max{‖au‖, ‖bv‖} = 1.

(iii) We first assume that | | is archimedean. Then |F | ⊆ F , so that v/‖v‖ andw/‖w‖ both have norm 1. Now note that

u =v

‖v‖− w

‖w‖=

v − w

‖v‖− w

‖w‖·‖v‖ − ‖w‖

‖v‖= u′ − u′′.

Since ‖u′‖ = ‖v − w‖/‖v‖ and ‖u′′‖ =∣∣‖v‖ − ‖w‖

∣∣/‖v‖ ≤ ‖v − w‖/‖v‖, it followsthat

d(v, w) ≤ ‖u‖ ≤ ‖u′‖+ ‖u′′‖ ≤ 2·‖v − w‖‖v‖

,

as required. Finally, suppose that | | is non-archimedean. If ‖v‖ 6= ‖w‖, then (ii)yields

d(v, w) ≤ 1 ≤ max{‖v‖, ‖w‖}/‖v‖ = ‖v − w‖/‖v‖.On the other hand, if ‖v‖ = ‖w‖, choose a ∈ F with |a| equal to this commonnorm. Then

d(v, w) ≤ ‖v/a− w/a‖ = ‖v − w‖/|a| = ‖v − w‖/‖v‖and the lemma is proved. �

It is clear from the above that if X and Z are projective subsets of V and if0 6= y ∈ V , then

d(X, Z) = inf{d(x, z) : 0 6= x ∈ X, 0 6= z ∈ Z}d(X, Z) ≤ d(X, y) + d(y, Z).

As usual, if σ : V → W is a linear transformation to another normed vector spaceW , then we can define

‖σ‖ = sup{‖σ(v)‖ : v ∈ S}.Since the unit sphere S is compact, ‖σ‖ = ‖σ(v0)‖ for some v0 ∈ S. Furthermore,‖σ(v)‖ ≤ ‖σ‖·‖v‖ for all v ∈ V . Of course, if σ 6= 0, then ‖σ‖ 6= 0.

Now, let T : V → V be a linear operator and let I be a subspace of V determinedby T . Then I is an attractor for T if, for certain subspaces Y of V , T maps vectorsclose to Y to vectors that are close to I. More precise versions are given below. Forconvenience, if X is any projective subset of V and if ε > 0, we define the (closed)ε-neighborhood of X by Nε(X) = {0 6= v ∈ V : d(v,X) ≤ ε}.

We first consider operators T that can be viewed as generalized transvections.Specifically, let τ : V → V be a nonzero operator of square 0 and set I = im τ =τ(V ). If T = 1+aτ with a ∈ F and |a| large, then the aτ term should dominate T ,

Page 4: LINEAR GROUPS AND GROUP RINGSpassman/lingroup.pdf · 2005. 10. 4. · LINEAR GROUPS AND GROUP RINGS J. Z. GONC¸ALVES AND D. S. PASSMAN Abstract. This paper consists of two parts.

4 J. Z. GONCALVES AND D. S. PASSMAN

and hence I should be an attractor for T . Indeed, the following is a slight variantof an argument given in [P].

Proposition 1.2. Let T = 1 + aτ be an operator on the normed F -vector spaceV , where a ∈ F and τ : V → V is nonzero and has square 0. Set I = im τ = τ(V ),and let K = ker τ . Suppose X is a subspace of V with V = X ⊕ K, let κ be apositive real number ≤ d(X, K)/2, and let ε > 0. If X = Nκ(X) and I = Nε(I),then T (X) ⊆ I for all suitably large |a|.

Proof. It is clear that both X and I are projective subsets of V . Furthermore, if0 6= u ∈ X is arbitrary, then by definition of κ and X, we have

κ + d(u, K) ≥ d(X, u) + d(u, K) ≥ d(X, K) ≥ 2κ,

so d(u, K) ≥ κ. In particular, d(X,K) ≥ κ and X is disjoint from K. Since P(V )has diameter ≤ 2, we also have κ ≤ 1.

Let v ∈ X and write v = x + y ∈ V = X ⊕ K, with x ∈ X and y ∈ K. SinceX ∩ K = ∅, we have x 6= 0. Indeed, we claim that ‖x‖ ≥ (κ/2)·‖v‖. For this, ify = 0, then x = v so ‖x‖ ≥ (κ/2)·‖v‖, since κ ≤ 1. On the other hand, if y 6= 0,then Lemma 1.1(iii) yields

κ ≤ d(X,K) ≤ d(v,K) ≤ d(v, y) ≤ 2·‖v − y‖/‖v‖ = 2·‖x‖/‖v‖.

So again we obtain ‖x‖ ≥ (κ/2)·‖v‖, as required.Now I ∼= V/K ∼= X and X ∩ K = 0, so the restriction of τ to X yields an

isomorphism σ : X → I. Let σ−1 : I → X be the inverse of σ and set s = ‖σ−1‖. Ifz = τ(x) = σ(x), then x = σ−1(z) and hence ‖x‖ ≤ ‖σ−1‖·‖z‖ = s·‖τ(x)‖. Thus,we conclude that ‖τ(x)‖ ≥ s−1·‖x‖ ≥ (κ/2s)·‖v‖.

Finally, note that T = 1 + aτ , so T (v) = v + aτ(x + y) = v + aτ(x). Since, T (v)and aτ(x) are nonzero, and since aτ(x) ∈ I, Lemma 1.1(iii) yields

d(T (v), I) ≤ d(T (v), aτ(x)) ≤ ‖T (v)− aτ(x)‖/‖aτ(x)‖ = ‖v‖/‖aτ(x)‖.

But ‖aτ(x)‖ = |a|·‖τ(x)‖ ≥ |a|·(k/2s)·‖v‖, so

d(T (v), I) ≤ ‖v‖/‖aτ(x)‖ ≤ 2s/(κ·|a|).

In particular, if |a| ≥ 2s/(κε), then d(T (v), I) ≤ ε and we conclude that T (v) ∈ I.Note that this lower bound 2s/(κε) depends upon τ,X and κ, but not upon thechoice of v ∈ X. �

Next, suppose that T : V → V is diagonalizable, that is, T is semisimple withall eigenvalues in F . If I is the subspace of V spanned by the eigenvectors corre-sponding to those eigenvalues of maximal absolute value, then it is reasonable thatI should be an attractor for all Tn, with n a sufficiently large positive integer. Forthis, it is convenient to first isolate the following facts.

Lemma 1.3. Let T : V → V be diagonalizable, and let n be a positive integer.i. If all eigenvalues of T have absolute value ≤ r, then there exists a real

constant k > 0 such that ‖Tn(v)‖ ≤ krn·‖v‖ for all v ∈ V .ii. If all eigenvalues of T have absolute value ≥ s, then there exists a real

constant k′ > 0 with ‖Tn(v)‖ ≥ k′sn·‖v‖ for all v ∈ V .

Proof. Let {v1, v2, . . . , vm} be a basis for V consisting of eigenvectors of T , with vi

corresponding to the eigenvalue λi. Then each v ∈ V can be written uniquely as

Page 5: LINEAR GROUPS AND GROUP RINGSpassman/lingroup.pdf · 2005. 10. 4. · LINEAR GROUPS AND GROUP RINGS J. Z. GONC¸ALVES AND D. S. PASSMAN Abstract. This paper consists of two parts.

LINEAR GROUPS AND GROUP RINGS 5

v =∑m

i=1 πi(v)vi, where πi : V → F is a nonzero linear functional. Now define anew norm ‖ ‖′ on V by ‖v‖′ = max{|πi(v)| : 1 ≤ i ≤ m}.

Assume that each λi satisfies r ≥ |λi| ≥ s. Since Tn(v) =∑m

i=1 λni πi(v)·vi, we

see that

‖Tn(v)‖′ = max{|λni πi(v)| : 1 ≤ i ≤ m}

≤ rn·max{|πi(v)| : 1 ≤ i ≤ m} = rn·‖v‖′.

Similarly, ‖Tn(v)‖′ ≥ sn·‖v‖′.Finally, note that the two norms ‖ ‖ and ‖ ‖′ are equivalent. In other words,

there are positive constants a and b with a·‖v‖′ ≥ ‖v‖ ≥ b·‖v‖′ for all v ∈ V . Thus

‖Tn(v)‖ ≤ a·‖Tn(v)‖′ ≤ arn·‖v‖′ ≤ (a/b)rn·‖v‖,

and similarly

‖Tn(v)‖ ≥ b·‖Tn(v)‖′ ≥ bsn·‖v‖′ ≥ (b/a)sn·‖v‖.

The result follows with k = a/b and k′ = b/a. �

With this, we can now prove the following generalization of [T, Lemma 3.9].

Proposition 1.4. Let T : V → V be a diagonalizable, nonsingular operator on thenormed F -vector space V . Let I be the subspace of V spanned by the eigenspaces ofT corresponding to the eigenvalues of absolute value ≥ r > 0, and let 0 6= K ⊆ Vbe spanned by the remaining eigenspaces. Suppose X is a subspace of V disjointfrom K, let κ be a positive real number ≤ d(X, K)/2, and let ε > 0. If we setX = Nκ(X) and I = Nε(I), then we have Tn(X) ⊆ I for all suitably large positiveintegers n.

Proof. Since T is diagonalizable, we have V = I ⊕K, and it is clear that both Xand I are projective subsets of V . Furthermore, if 0 6= u ∈ X is arbitrary, then bydefinition of κ and X, we have

κ + d(u, K) ≥ d(X, u) + d(u, K) ≥ d(X, K) ≥ 2κ,

so d(u, K) ≥ κ. Hence d(X,K) ≥ κ.Let v ∈ X and write v = z + y ∈ V = I ⊕K, with z ∈ I and y ∈ K. If y = 0,

then v ∈ I, so Tn(v) ∈ I, for all n, and there is nothing to prove. Thus, we cansuppose that y 6= 0. By Lemma 1.1(iii), we have

κ ≤ d(X,K) ≤ d(v,K) ≤ d(v, y) ≤ 2·‖v − y‖/‖v‖ = 2·‖z‖/‖v‖,

so ‖z‖ ≥ (κ/2)·‖v‖ and z 6= 0. On the other hand, if π : V → K is the naturalprojection with kernel I, then y = π(v), so ‖y‖ ≤ h·‖v‖, where we set h = ‖π‖.

By the definition of r, part (ii) of the previous lemma implies that ‖Tn(z)‖ ≥k′rn‖z‖ for some positive constant k′. Also, if s is the maximum absolute valueof all the eigenvalues of the restriction of T to K, then s < r and part (i) of theprevious lemma implies that ‖Tn(y)‖ ≤ ksn‖y‖ for some positive constant k.

Finally, since Tn(v), Tn(z) 6= 0 and Tn(z) ∈ I, Lemma 1.1(iii) yields

d(Tn(v), I) ≤ d(Tn(v), Tn(z)) ≤ 2·‖Tn(v − z)‖/‖Tn(z)‖= 2·‖Tn(y)‖/‖Tn(z)‖.

Page 6: LINEAR GROUPS AND GROUP RINGSpassman/lingroup.pdf · 2005. 10. 4. · LINEAR GROUPS AND GROUP RINGS J. Z. GONC¸ALVES AND D. S. PASSMAN Abstract. This paper consists of two parts.

6 J. Z. GONCALVES AND D. S. PASSMAN

Furthermore, by the above, we have ‖Tn(y)‖ ≤ ksn·‖y‖ ≤ khsn·‖v‖ and ‖Tn(z)‖ ≥k′rn·‖z‖ ≥ k′rn(κ/2)·‖v‖. Thus

d(Tn(v), I) ≤ 2·‖Tn(y)‖‖Tn(z)‖

≤ 4kh

k′κ·(s/r)n.

But 0 < (s/r) < 1 and (4kh)/(k′κ) is a positive constant, so it is clear that if n issufficiently large, then this upper bound for d(Tn(v), I) can be made to be ≤ ε. Inother words, Tn(v) ∈ I for all sufficiently large n, where the bound on n dependson T , X and κ, but not on the particular choice of v ∈ X. �

As a consequence of the above, essentially replacing T by T−1, we obtain

Proposition 1.5. Let T : V → V be a diagonalizable, nonsingular operator on thenormed F -vector space V . Let I be the subspace of V spanned by the eigenspacesof T corresponding to the eigenvalues of absolute value ≤ r, and let 0 6= K ⊆ V bespanned by the remaining eigenspaces. Suppose X is a subspace of V disjoint fromK, let κ be a positive real number ≤ d(X, K)/2, and let ε > 0. If we set X = Nκ(X)and I = Nε(I), then we have T−n(X) ⊆ I for all suitably large positive integers n.

2. Free products in linear groups

Our goal now is to obtain applications of the attractor results to the existenceof free products as is done, for example, in [T, §3]. Since, most of the argumentshere tend to be similiar, our proofs are somewhat skimpy. We first need

Lemma 2.1. Let T : V → V be a nonsingular linear transformation, and let Xand Y be projective subsets of V . Then

d(T (X), T (Y )) ≤ 2 d(X, Y )·‖T‖·‖T−1‖.

In particular, if 0 6= x ∈ V , then

d(T (x), T (Y )) ≤ 2 d(x, Y )·‖T‖·‖T−1‖.

Proof. Let x ∈ X ∩ S and y ∈ Y ∩ S, where S denotes the unit sphere of V . Then

d(T (X), T (Y )) ≤ d(T (x), T (y)) ≤ 2 ‖T (x− y)‖/‖T (x)‖,

by Lemma 1.1(iii). Now ‖T (x− y)‖ ≤ ‖T‖·‖x− y‖, and x = T−1(T (x)) impliesthat 1 = ‖x‖ ≤ ‖T−1‖·‖T (x)‖. Thus

d(T (X), T (Y )) ≤ 2 ‖x− y‖·‖T‖·‖T−1‖,

and the result follows since d(X, Y ) = inf{‖x− y‖ : x ∈ X ∩ S, y ∈ Y ∩ S}. �

As is to be expected, the proof of the existence of free products ultimatelydepends upon the “ping-pong” lemma of F. Klein (see [H, Lemma II.24]). Forconvenience, we state and quickly prove this elementary, but powerful, result. Here,we use G# to denote the nonidentity elements of a group G.

Lemma 2.2. Let Γ be a group generated by the nonidentity subgroups G and H,and suppose that Γ acts on a set X having nonempty subsets P and Q with Q 6= P .If G#Q ⊆ P , H#P ⊆ Q, and |H| > 2, then Γ is naturally isomorphic to the freeproduct G ∗H.

Page 7: LINEAR GROUPS AND GROUP RINGSpassman/lingroup.pdf · 2005. 10. 4. · LINEAR GROUPS AND GROUP RINGS J. Z. GONC¸ALVES AND D. S. PASSMAN Abstract. This paper consists of two parts.

LINEAR GROUPS AND GROUP RINGS 7

Proof. It suffices to show that 1 ∈ Γ cannot be written as a nonempty alternatingproduct of elements coming from G# and H#. Suppose by way of contradictionthat such a product 1 = γ1γ2 · · · γn exists with n ≥ 1. If the product starts and endsin G#, that is if γ1, γn ∈ G#, then by conjugating this expression by a nonidentityelement of H, we obtain a similar expression, but this time starting and ending inH#. Next, if γ1 ∈ G# and γn ∈ H#, then since |H| > 2, we can conjugate theexpression by an element of H#, different from γ−1

n , to obtain a similar productbut starting and ending in H#. Since the same argument handles the γ1 ∈ H#,γn ∈ G# situation, we can therefore replace any such expression by one withγ1, γn ∈ H#. But then, the alternating nature of the action of G# and H# on Pand Q yields 1P = P and γ1γ2 · · · γnP ⊆ Q, and hence P ⊆ Q. Furthermore, byconjugating the expression for 1 by a nonidentity element of G, we obtain a similarexpression but now starting and ending in G#. This time, the alternating nature ofthe action yields 1Q = Q and γ1γ2 · · · γnQ ⊆ P , so we obtain the reverse inclusionQ ⊆ P . Hence P = Q, contradiction. �

The following is essentially [P, Theorem 1.1]. Suppose T : V → V is given byT = 1 + aτ , where 0 6= a ∈ F and τ : V → V is a nonzero operator of square0. Since Tn = 1 + naτ , we see that T has infinite order if charF = 0 and it hasprime order p if charF = p > 0. Thus the condition below that either |G| ≥ 3or char F 6= 2 guarantees that one of the two generating subgroups in 〈G, T 〉 hasorder at least 3. Furthermore, note that |na| = |n| |a|. Hence, in order to applyProposition 1.2 to Tn, we need to know that |na| is at least as large as |a|, whenn 6= 0 in F . Specifically, we use 1Z to denote the set of integer multiples of 1 in F ,so that 1Z = Z if charF = 0 and 1Z = GF(p) if charF = p > 0. In the latter case,it is clear that |1Z \ 0| = 1, but in the former situation, a number of possibilitiesexist. Thus, the hypothesis below that |1Z \ 0| ≥ 1 comes into play only in thecharacteristic 0 situation. This hypothesis is needed, since the condition fails, forexample, in p-adic fields. Indeed, in view of [B, Theorem VI.9.1] it fails preciselywhen F is a finite algebraic extension of a p-adic field.

Theorem 2.3. Let F be a locally compact field, let V be a finite-dimensional F -vector space, and let G be a nonidentity finite subgroup of the general linear groupGL(V ). Assume, in fact, that |G| ≥ 3 when charF = 2. Furthermore, let τ : V → Vbe a nonzero linear transformation of square 0, and write K = ker τ and I = im τ =τ(V ). If gI ∩ K = 0 for all g ∈ G# and if |1Z \ 0| ≥ 1, then for all a ∈ F ofsufficiently large absolute value, we have 〈G, T 〉 ∼= G ∗ 〈T 〉 where T = 1 + aτ .

Proof. Let 2κ be the minimum of the finitely many distances d(gI,K) for all g ∈G#. Then κ > 0, by assumption and Lemma 1.1(i), and we set P =

⋃g∈G# Nκ(gI).

Next, let r = max{2·‖g‖·‖g−1‖ : g ∈ G#}, set ε = κ/r, and define Q = Nε(I).We claim that G#Q ⊆ P . To this end, let g ∈ G# and v ∈ Q. Then v ∈ Nε(I), so

d(v, I) ≤ ε. Therefore, by Lemma 2.1, we have d(gv, gI) ≤ 2·‖g‖·‖g−1‖·ε ≤ rε = κ,so gv ∈ Nκ(gI) ⊆ P , as required. Note also that I ⊆ Q, but that I ∩ P = ∅ by thedefinition of κ and the fact that I ⊆ K. Thus P 6= Q.

Finally, by Propositions 1.2, we know that I is an attractor for T . Specifically,by applying this result to each of the finitely many subspaces gI, with g ∈ G#, wesee that there exists a positive real number s so that if |a| ≥ s, then T ·Nκ(gI) ⊆Nε(I) = Q. Thus since Tn = (1 + aτ)n = 1 + naτ and |na| ≥ |a| for all n ∈ Z with

Page 8: LINEAR GROUPS AND GROUP RINGSpassman/lingroup.pdf · 2005. 10. 4. · LINEAR GROUPS AND GROUP RINGS J. Z. GONC¸ALVES AND D. S. PASSMAN Abstract. This paper consists of two parts.

8 J. Z. GONCALVES AND D. S. PASSMAN

1·n 6= 0, we have 〈T 〉#P ⊆ Q. Since at least one of the groups G or 〈T 〉 has order≥ 3, we conclude from Lemma 2.2 that 〈G, T 〉 = G ∗ 〈T 〉. �

The next result is the semisimple analog of the above. To avoid repetition, ifT : V → V is any nonsingular, diagonalizable operator, then we say that V =X+ ⊕ X0 ⊕ X− is a T -decomposition of V if there exist real numbers r > s > 0with X+ 6= 0 spanned by the eigenspaces of T corresponding to the eigenvaluesof absolute value ≥ r, X− 6= 0 spanned by the eigenspaces of T correspondingto the eigenvalues of absolute value ≤ s, and with X0 the span of the remainingeigenspaces. Note that the hypothesis below implies that X+ 6= X− and thatdimF X+ = dimF X−. In particular, T must have infinite multiplicative ordersince all its eigenvalues cannot be roots of unity, which all have absolute value 1.

Theorem 2.4. Let F be a locally compact field, let V be a finite-dimensional F -vector space, and let G 6= 1 be a finite subgroup of GL(V ). Furthermore, supposeT : V → V is a nonsingular, diagonalizable linear transformation and let V =X+ ⊕X0 ⊕X− be a T -decomposition of V . Assume that, for all g ∈ G#, gX+ andgX− are disjoint from both X0 ⊕X− and X+ ⊕X0. Then, for all sufficiently largeintegers n, 〈G, Tn〉 ∼= G ∗ 〈Tn〉.

Proof. Let 2κ be the minimum of the finitely many distances d(gX+, X0 ⊕ X−),d(gX−, X0 ⊕X−), d(gX+, X+ ⊕X0) and d(gX−, X+ ⊕X0) for all g ∈ G#. Thenκ > 0, by assumption and Lemma 1.1(i), and we set

P =⋃

g∈G#

Nκ(gX+) ∪⋃

g∈G#

Nκ(gX−).

Next, let t = max{2·‖g‖·‖g−1‖ : g ∈ G#}, set ε = κ/t, and define

Q = Nε(X+) ∪Nε(X−).

We claim that G#Q ⊆ P . To this end, let g ∈ G# and v ∈ Q. Then v ∈Nε(X±) for some choice of ±, so d(v,X±) ≤ ε. Therefore, by Lemma 2.1, we haved(gv, gX±) ≤ 2·‖g‖·‖g−1‖·ε ≤ tε = κ, so gv ∈ Nκ(gX±) ⊆ P , as required. Notealso that X+ ⊆ Q, but that X+ ∩ P = ∅ by the definition of κ, so P 6= Q.

Finally, by Propositions 1.4 and 1.5, we know that X+ is an attractor for T andthat X− is an attractor for T−1. Specifically, by applying those results to each ofthe finitely many subspaces gX+ and gX−, with g ∈ G#, we see that there existsa positive integer n0 so that if n ≥ n0, then Tn·Nκ(gX±) ⊆ Nε(X+) ⊆ Q andT−n·Nκ(gX±) ⊆ Nε(X−) ⊆ Q, for all g ∈ G#. It follows that if n ≥ n0, then〈Tn〉#P ⊆ Q. Thus, since T has infinite multiplicative order, Lemma 2.2 impliesthat 〈G, Tn〉 = G ∗ 〈Tn〉. �

The remaining theorems in this section are concerned with groups generatedby two operators S and T that are either both diagonalizable, both generalizedtransvections, or one of each. As above, in the case of generalized transvections,we need to assume that |1Z \ 0| ≥ 1. Furthermore, when both S and T are gener-alized transvections, we suppose that char F 6= 2 to avoid the possibility that bothoperators have order 2. It is easy to see that the hypotheses below imply that thevarious subspaces I, J , X± and Y± must all have the same dimension.

Theorem 2.5. Let V be a finite-dimensional F -vector space and let S, T : V →V be two nonsingular operators. Suppose S and T are both diagonalizable with

Page 9: LINEAR GROUPS AND GROUP RINGSpassman/lingroup.pdf · 2005. 10. 4. · LINEAR GROUPS AND GROUP RINGS J. Z. GONC¸ALVES AND D. S. PASSMAN Abstract. This paper consists of two parts.

LINEAR GROUPS AND GROUP RINGS 9

V = X+ ⊕X0 ⊕X− and V = Y+ ⊕ Y0 ⊕ Y− being S- and T -decompositions of V ,respectively. If the eight intersections X±∩(Y0⊕Y±) and Y±∩(X0⊕X±) are trivial,then for all sufficiently large positive integers m,n, we have 〈Sm, Tn〉 = 〈Sm〉∗〈Tn〉.

Proof. Let 2κ be the smallest value of the eight distances d(X±, Y0 ⊕ Y±) andd(Y±, X0 ⊕ X±), so that κ > 0 by assumption and Lemma 1.1(i). Define P =Nκ(X+) ∪ Nκ(X−) and Q = Nκ(Y+) ∪ Nκ(Y−). Since X+ is an attractor forS and X− is an attractor for S−1, it follows from Propositions 1.4 and 1.5 that〈Sm〉#Q ⊆ P for all sufficiently large positive integers m. Similarly, 〈Tn〉#P ⊆ Qfor all sufficiently large n. Since X+ ⊆ P and X+ ∩Q = ∅, we see that P 6= Q, andhence the result follows from Lemma 2.2. �

The above generalizes [T, Proposition 3.2], but also follows from it. Namely, ifk is the common dimension of X± and Y±, then we can let G = 〈S, T 〉 act on theexterior power W = ∧kV . Since S and T are diagonalizable on W and since ∧kX±and ∧kY± reduce to points in the projective space P(W ), we can conclude from[T, Proposition 3.2] that the image of G is a free product and hence so is G. Onthe other hand, if either S or T is a generalized transvection, then its structure asan operator on ∧kV becomes quite different in nature, and therefore this exteriorpower trick no longer applies.

Theorem 2.6. Let F be a locally compact field, let V be a finite-dimensional F -vector space, and let S, T : V → V be two nonsingular operators. Specifically, S =1 + aσ and T = 1 + bτ are both generalized transvections, where σ, τ : V → V arenonzero operators of square 0. Assume that |1Z \ 0| ≥ 1 and char F 6= 2. WriteI = σ(V ) = im σ, K = kerσ, J = τ(V ) = im τ , and L = ker τ . If the intersectionsI ∩ L and J ∩ K are both trivial, then for all a, b ∈ F with |a| and |b| sufficientlylarge, we have 〈S, T 〉 = 〈S〉 ∗ 〈T 〉.

Proof. Let 2κ be the smaller of the distances d(I, L) and d(J,K), so that κ > 0 byassumption and Lemma 1.1(i). Define P = Nκ(I) and Q = Nκ(J). Since I is anattractor for S and |1Z \ 0| ≥ 1, it follows from Proposition 1.2 that 〈S〉#Q ⊆ Pfor all a ∈ F with sufficiently large absolute value. Similarly, 〈T 〉#P ⊆ Q for allb ∈ F with sufficiently large absolute value. Since I ⊆ K, we have I ⊆ P andI ∩Q = ∅. Thus P 6= Q and, since S and T have order at least 3, the result followsfrom Lemma 2.2. �

Finally, we consider the mixed case.

Theorem 2.7. Let V be a finite-dimensional F -vector space and let S, T : V → Vbe two nonsingular operators. Suppose S is diagonalizable with an S-decompositiongiven by V = X+ ⊕ X0 ⊕ X−. Furthermore, suppose T = 1 + aτ is a generalizedtransvection, where τ : V → V is a nonzero operator of square 0 with I = τ(V ) =im τ and K = ker τ . Assume also that |1Z\0| ≥ 1. If the four intersections X±∩Kand I ∩ (X0⊕X±) are trivial, then for all sufficiently large integers n and all a ∈ Fof sufficiently large absolute value, we have 〈Sn, T 〉 = 〈Sn〉 ∗ 〈T 〉.

Proof. Let 2κ be the smallest of the four distances d(I,X0 ⊕ X±) and d(X±,K),so that κ > 0 by assumption and Lemma 1.1(i). Define P = Nκ(X+) ∪ Nκ(X−)and Q = Nκ(I). Since X+ is an attractor for S and X− is an attractor for S−1,it follows from Propositions 1.4 and 1.5 that 〈Sn〉#Q ⊆ P for all sufficiently largepositive integers n. Similarly, since I is an attractor for T , Proposition 1.2 and

Page 10: LINEAR GROUPS AND GROUP RINGSpassman/lingroup.pdf · 2005. 10. 4. · LINEAR GROUPS AND GROUP RINGS J. Z. GONC¸ALVES AND D. S. PASSMAN Abstract. This paper consists of two parts.

10 J. Z. GONCALVES AND D. S. PASSMAN

the hypothesis imply that 〈T 〉#P ⊆ Q for all a ∈ F with sufficiently large absolutevalue. Finally, I ⊆ Q and I∩P = ∅, so P 6= Q, and Lemma 2.2 yields the result. �

3. Bass cyclic units

The goal of the remainder of this paper is to apply linear group results, andin particular Theorem 2.5, to the unit group of an integral group ring. We firstintroduce some notation.

Let G be a group and let x be an elements of G of finite order d. We work inthe integral group ring Z[X] ⊆ Z[G], where X is the cyclic group generated by x.To start with, let x = X =

∑d−1i=0 xi denote the sum of the elements of X. As is

well known, xjX = Xxj = X for all j. Now define

uk,m(x) = (1 + x + · · ·+ xk−1)m +1− km

dx,

where 1 ≤ k, gcd(k, d) = 1, and where m is a multiple of the Euler functionϕ(d). The latter two conditions imply that km ≡ 1 mod d and hence uk,m(x) ∈Z[X]. Recall that the augmentation map of Z[X] is the homomorphism Z[X] →Z determined by x 7→ 1. Then each uk,m(x) has augmentation 1, and indeed,uk,m(x) = (1 + x + · · ·+ xk−1)m + cx where c is the unique integer such that thiselement has the augmentation 1. We can, of course, view uk,m(x) as a polynomialfunction on x subject to xd = 1. In particular, we can evaluate this function on anyy satisfying yd = 1. For example, we can take y = xj for any integer j, or y = εwhere ε is any complex dth root of unity.

Lemma 3.1. With the above notation, we havei. u(k+d),m(x) = uk,m(x).ii. uk,m(x)·uk,n(x) = uk,(m+n)(x).iii. uk,m(x)·u`,m(xk) = ukl,m(x).iv. u1,m(x) = 1 and uk,m(x)−1 = u`,m(xk) where k` ≡ 1 mod d.

Proof. For parts (i), (ii) and (iii), we use the identities xj x = xxj = x and xd = 1to easily show that the right and left sides of each equation differ by an integermultiple of x, say cx. Furthermore, since both sides have augmentation 1, theirdifference has augmentation 0. But the augmentation of cx is equal to cd, so itfollows that c = 0 and hence both sides of each equation are equal. For part (iv), itis clear that u1,m(x) = 1 and hence, by part (i), ur,m(x) = 1 for any positive integerr ≡ 1 mod d. Finally, if k` ≡ 1 mod d, then (iii) implies that uk,m(x)·u`,m(xk) = 1,as required. �

In view of (iv) above, each uk,m(x) is a unit in Z[G] and, as in [S, Chapter 2],these elements are called Bass cyclic units. Furthermore, in view of (i), uk,m(x) isdetermined by k modulo d and hence we can assume that 1 ≤ k ≤ d − 1. Whenthe first parameter is equal to 1, then u1,m(x) = 1, and when this parameter isequal to d − 1, then it is easy to see that ud−1,m(x) = x(d−1)m. Because of this,we usually take 2 ≤ k ≤ d − 2 and hence d ≥ 5. Finally, it follows from (ii) thatuk,m(x)a = uk,ma(x) for all integers a ≥ 1.

Lemma 3.2. Let θ : Z[G] → Z[H] be the group ring homomorphism determined bythe group epimorphism θ : G → H, and let y be an element of H of order d. Ifuk,m(y) is a Bass cyclic unit of Z[H], then there exists an element x ∈ G, whose

Page 11: LINEAR GROUPS AND GROUP RINGSpassman/lingroup.pdf · 2005. 10. 4. · LINEAR GROUPS AND GROUP RINGS J. Z. GONC¸ALVES AND D. S. PASSMAN Abstract. This paper consists of two parts.

LINEAR GROUPS AND GROUP RINGS 11

order has the same prime factors as those of y, and a Bass cyclic unit uk,m′(x) ofZ[G] such that uk,m′(x) maps to a positive integer power of uk,m(y).

Proof. If y is a π-element, then there exists a π-element x ∈ G with θ(x) = y. Inparticular, we now know that k is prime to d′, the order of x. On the other hand,we may have d′ > d so it is not necessarily true that m divides ϕ(d′). Nevertheless,there certainly exists a positive integer a so that ϕ(d′) divides m′ = ma. Thenuk,m′(x) is a Bass cyclic unit and, since θ(x) is an integer multiple of y, it followsfrom part (ii) of the preceding lemma that θ(uk,m′(x)) and uk,m(y)a agree up toan integer multiple of y, say cy. But both of these terms have augmentation 1, sowe conclude that c = 0, as required. �

We will apply Theorem 2.5 to matrix images of these units, so it is necessaryto understand their eigenvalues. For this, we first isolate the following simple factusing a bit of calculus.

Lemma 3.3. Let r and k be real numbers with 2 ≤ k and 0 < r ≤ 1/(2k). Thenthe real-valued function

f(z) =∣∣∣∣ sin kπz

sinπz

∣∣∣∣defined on the interval [r, 1/2] takes on its maximum at z = r. Moreover, f(r) > 1.

Proof. We first study the function f(z) in the open interval (0, 1/k). Note that bothnumerator and denominator are positive here, so the absolute value is unnecessary.Furthermore, if z ∈ [1/(2k), 1/k) then the numerator of f(z) is decreasing andthe denominator is increasing, so f(z) is certainly strictly decreasing. On theother hand, if z ∈ (0, 1/(2k)), then all the trigonometric functions involved arepositive and it is easy to see that the derivative ∂f(z)/∂z is a positive multiple ofk tanπz− tan kπz < 0. Thus f(z) is also decreasing in (0, 1/(2k)), and we concludethat f(z) is strictly decreasing to 0 in (0, 1/k). Since r is in this interval, we seethat f(r) > f(z) for all z ∈ (r, 1/k).

It remains to compare f(r) with the values f(z) for z ∈ [1/k, 1/2]. To do this,we use the inequalities y ≥ sin y ≥ y− y3/6 = y(1− y2/6) which hold for all y ≥ 0.To start with

sin kπr ≥ kπr·[1− (kπr)2/6] ≥ kπr·[1− π2/24]

since kr ≤ 1/2. Thus, sin πr ≤ πr yields

f(r) =sin kπr

sinπr≥ kπr·[1− π2/24]

πr= k·[1− π2/24].

Indeed, using k ≥ 2, we have f(r) ≥ 2[1− π2/24] > 1.177 > 1.On the other hand, if z ∈ [1/k, 1/2], then | sin kπz| ≤ 1 and

sinπz ≥ sinπ/k ≥ (π/k)·[1− (π/k)2/6] ≥ (π/k)·[1− π2/24],

using k ≥ 2. Thus, since π·[1− π2/24]2 > 1.089 > 1, we have

f(z) =∣∣∣∣ sin kπz

sinπz

∣∣∣∣ ≤ k

π·[1− π2/24]< k·[1− π2/24] ≤ f(r),

and the lemma is proved. �

Page 12: LINEAR GROUPS AND GROUP RINGSpassman/lingroup.pdf · 2005. 10. 4. · LINEAR GROUPS AND GROUP RINGS J. Z. GONC¸ALVES AND D. S. PASSMAN Abstract. This paper consists of two parts.

12 J. Z. GONCALVES AND D. S. PASSMAN

It is instructive to look at a computer plot of the function f(z) in the interval[0, 1/2] for large values of k. One sees that the decreasing aspect of f(z) fromf(0) = k to f(1/k) = 0 is quite precipitous. On the other hand, the values of f(z)in the interval [1/k, 1/2] stay relatively small.

Lemma 3.4. Let ε = e2πi/d be a primitive complex dth root of unity and let a bean integer. Assume that 2 ≤ k ≤ d− 2 and that gcd(k, d) = 1.

i. uk,m(1) = 1 and if εa 6= 1 then

|uk,m(εa)| =∣∣∣∣εak/2 − ε−ak/2

εa/2 − ε−a/2

∣∣∣∣m =∣∣∣∣ sin(kπa/d)

sin(πa/d)

∣∣∣∣m .

ii. The largest absolute value |uk,m(εa)| occurs when a ≡ ±1 mod d.iii. The smallest absolute value |uk,m(εa)| occurs when ak ≡ ±1 mod d.

Proof. (i) Since x evaluated at 1 is equal to d, we have uk,m(1) = km+(1−km) = 1.On the other hand, if εa 6= 1, then x evaluated at εa is 0 and

uk,m(εa) = [1 + (εa) + (εa)2 + · · ·+ (εa)k−1]m =(

εak − 1εa − 1

)m

.

Hence, since |εa/2| = 1, this yields

|uk,m(εa)| =∣∣∣∣εak/2 − ε−ak/2

εa/2 − ε−a/2

∣∣∣∣m .

Note that the numerator and denominator here are twice the imaginary parts ofεak/2 and εa/2 respectively, so

|uk,m(εa)| =∣∣∣∣ sin(kπa/d)

sin(πa/d)

∣∣∣∣m .

(ii) Let us first assume that εa 6= 1. Then from the above formula, it is clear that|uk,m(εa)| = |u(d−k),m(εa)|. In particular, by replacing k by d − k if necessary, itsuffices to assume that 2 ≤ k ≤ d/2. Furthermore, since |uk,m(εa)| = |uk,m(ε−a)|,it suffices to restrict our attention to the possibilities a = 1, 2, . . . , bd/2c. For this,consider the real-valued function

f(z) =∣∣∣∣ sin kπz

sinπz

∣∣∣∣and observe that |uk,m(εa)| = f(a/d)m with m > 0. Furthermore, each a/d iscontained in the closed interval [r, 1/2] with r = 1/d. Since rk = k/d ≤ 1/2,the preceding lemma now implies that the maximum value of f(z) on this intervaloccurs at z = r = 1/d and that this largest value is > 1. Thus a = 1 and f(1/d) > 1.Taking into account the ± symmetry, we see that the maximum value of |uk,m(εa)|with εa 6= 1 occurs precisely when a ≡ ±1 mod d. Indeed, since this value is largerthan 1 and since uk,m(1) = 1, we see that |uk,m(ε±1)| is the maximum value of|uk,m(εa)| over all complex dth roots of unity.

(iii) The smallest value of |uk,m(εa)| occurs precisely when |uk,m(εa)−1| takes onits largest value. Thus, since uk,m(x)−1 = u`,m(xk) with k` ≡ 1 mod d, we see that|uk,m(εa)| is minimal when |u`,m(εka)| is maximal. Since 2 ≤ ` ≤ d−2, we concludefrom the above that this occurs precisely when ak ≡ ±1 mod d, as required. �

Finally, we discuss all the values of |uk,m(εa)| at least when d is a prime power.

Page 13: LINEAR GROUPS AND GROUP RINGSpassman/lingroup.pdf · 2005. 10. 4. · LINEAR GROUPS AND GROUP RINGS J. Z. GONC¸ALVES AND D. S. PASSMAN Abstract. This paper consists of two parts.

LINEAR GROUPS AND GROUP RINGS 13

Lemma 3.5. Let p be a prime.

i. Suppose ε1, ε2, . . . , εr and δ1, δ2, . . . , δr are complex pnth roots of unity thatsatisfy ε1 + ε2 + · · ·+ εr = δ1 + δ2 + · · ·+ δr. If r ≤ p− 1 then, by relabelingthe δ’s if necessary, we have εi = δi for all i.

ii. Set d = pn, suppose 2 ≤ k ≤ d−2, and let ε be a primitive complex dth rootof unity. If p ≥ 5 and |uk,m(εa)| = |uk,m(εb)|, then a ≡ ±b mod d.

Proof. (i) Let tr denote the Galois trace in the field of pnth roots of unity dividedby pn−1. Then tr 1 = p − 1, tr ε = −1 if ε is a primitive pth root of unity, andtr ε = 0 if ε is a primitive path root of unity with 2 ≤ a ≤ n. We first show thatsome δi must equal ε1. For this, by multiplying through by ε−1

1 if necessary, itsuffices to assume that ε1 = 1. Since r ≤ p−1, the trace of the left hand side of theequation is ≥ (p− 1)− (p− 2) > 0 and thus there must exist some δi with tr δi > 0.In other words, δi = 1 = ε1, and the result follows by induction on r.

(ii) Since d is odd, each dth root of unity is a square. Thus, to avoid fractionalexponents, we replace a by 2a and b by 2b. Suppose first that ε2a and ε2b are not1. Then, by Lemma 3.4(i) and the fact that (εak − ε−ak)/(εa − ε−a) is real, theequality |uk,m(ε2a)| = |uk,m(ε2b)| implies that

εak − ε−ak

εa − ε−a= κ·ε

bk − ε−bk

εb − ε−b

where κ = ±1.For convenience, let <(ϑ) = ϑ + ϑ denote twice the real part of ϑ. Then cross

multiplying the above displayed equation yields

<(εak+b − εak−b

)= κ·<

(εbk+a − εbk−a

)= <

(εbk+κa − εbk−κa

),

where the last equality of course holds only for κ = ±1. Thus

<(εak+b + εbk−κa

)= <

(εbk+κa + εak−b

)and, since both sides are sums of four pnth roots of unity and since p ≥ 5, weconclude from part (i) that the right-hand and left-hand exponents must matchmodulo d. But certainly ak + b 6≡ ±(ak − b) mod d, so we obtain

ak + b ≡ ±(bk + κa) mod d

ak − b ≡ ±(bk − κa) mod d.

If the two ± signs above disagree, then adding the equations yields 2ak ≡ ±2κaand hence k ≡ ±1 mod d, a contradiction. Thus the signs must agree and this timeadding yields 2ak ≡ ±2bk so 2a ≡ ±2b mod d, as required.

When ε2b = 1, the argument is of course simpler. Suppose, by way of contradic-tion that ε2a 6= 1. Then |uk,m(ε2a)| = |uk,m(ε2b)| = 1 yields

εak − ε−ak

εa − ε−a= κ = ±1,

so εka − ε−ka = κ(εa − ε−a) = εκa − ε−κa. Thus εka + ε−κa = ε−ka + εκa and thisis a contradiction since ka 6≡ −ka mod d and ka 6≡ κa mod d. �

Page 14: LINEAR GROUPS AND GROUP RINGSpassman/lingroup.pdf · 2005. 10. 4. · LINEAR GROUPS AND GROUP RINGS J. Z. GONC¸ALVES AND D. S. PASSMAN Abstract. This paper consists of two parts.

14 J. Z. GONCALVES AND D. S. PASSMAN

4. Free subgroups of the unit group

In this section, we prove our main group ring result which asserts that if Gis a nonabelian group of order prime to 6, then Z[G] has two Bass cyclic unitsthat generate a nonabelian free subgroup of the unit group. For this, it is firstconvenient to restate Theorem 2.5 in a more usable form, based on the projectionmaps to the plus and minus components of the S- and T -decompositions of V .Here, all the assumptions of the first two sections apply. In particular, F is anabsolute-valued field that is locally compact in the metric topology. As will beapparent, the reformulation below includes a slight change in notation.

Corollary 4.1. Let V be a finite-dimensional F -vector space and let S and T betwo nonsingular operators on V . Suppose S and T are both diagonalizable withV = S+ ⊕ S0 ⊕ S− and V = T+ ⊕ T0 ⊕ T− being S- and T -decompositions of V ,respectively. Assume that dim S+ = dim S− = r = dim T+ = dim T− and considerthe four projections σ+ : V → S+, σ− : V → S−, τ+ : V → T+, and τ− : V → T−.If the idempotent conditions rank σiτj = r = rank τjσi hold for all i, j ∈ {+,−},then 〈Sm, Tn〉 = 〈Sm〉 ∗ 〈Tn〉 for all sufficiently large positive integers m and n.

Proof. Since rank σ+τ+ = r = rank τ+, we have V σ+τ+ = V τ+. Thus the mapτ+ : V σ+ → V τ+ is onto and consequently also one-to-one. In other words, V σ+ isdisjoint from ker τ+ = V (1 − τ+), and we see that S+ ∩ (T0 ⊕ T−) = 0. Similarly,the seven remaining idempotent conditions yield the seven remaining intersectionconditions of Theorem 2.5, and hence the result follows. �

Obviously, Theorems 2.6 and 2.7 have similar interpretations based on suitableprojections, but these maps are not canonically defined. For integral group ringapplications, we will of course apply Corollary 4.1 with F = C, the field of complexnumbers. We now begin the proof of our main result by considering a number ofspecial cases. The following lemma is well-known. We use it to fix notation formuch of the remainder of this paper.

Lemma 4.2. Let G = AoX, where A is a normal abelian subgroup of G and whereX = 〈x〉 is cyclic of prime order p. Let X : C[G] → Mn(C) be a complex irreduciblerepresentation of G of degree n > 1, with associated character χ : G → C, and letµ be an irreducible constituent of the restriction of χ to A. Then we have

i. n = p and we can assume that

X(a) = diag(µ(a), µx(a), . . . , µxp−1(a))

for all a ∈ A. Here µxi

(a) = µ(xiax−i) and these p linear characters areall distinct. Furthermore, X(x) is then the permutation matrix

X(x) =

1

1. . .

11

.

Page 15: LINEAR GROUPS AND GROUP RINGSpassman/lingroup.pdf · 2005. 10. 4. · LINEAR GROUPS AND GROUP RINGS J. Z. GONC¸ALVES AND D. S. PASSMAN Abstract. This paper consists of two parts.

LINEAR GROUPS AND GROUP RINGS 15

ii. Let ε1, ε2, . . . , εp be all the complex pth roots of unity, define

P =1√

p

1 1 · · · 1ε1 ε2 · · · εp

......

...εp−11 εp−1

2 · · · εp−1p

and let Q = P ∗, where ∗ denotes transpose conjugate. Then P is a unitarymatrix, Q = P−1 and QX(x)P = D = diag(ε1, ε2, . . . , εp).

iii. If A = 〈a〉 is cyclic, then µ(a), µx(a), . . . , µxp−1(a) are all distinct and not

equal to 1. In addition, if p is odd, then no two of these elements can becomplex conjugates of each other.

Proof. Part (i) follows from Clifford’s Theorem [I, Theorem 6.5] and a result of Ito[I, Theorem 6.15]. Note that if two of the characters µxi

and µxj

are equal, then allof these conjugate characters are identical and hence X(A) is central in the matrixring. But then X(G) is abelian, contradicting the assumption that n > 1. Thus thevarious µxi

must be distinct. Part (ii) is an immediate consequence of the simplecomputation X(x)P = PD along with the fact that P is clearly unitary. Finally, for(iii), since A = 〈a〉, we see that the linear characters of A are determined by theirvalue on a. In particular, by (i), it follows that µ(a), µx(a), . . . , µxp−1

(a) are alldistinct. Furthermore, if some µxi

(a) = 1, then µxi

= 1A and hence µxk

= 1A forall k, certainly a contradiction. Finally, if µxi

(a) and µxj

(a) are complex conjugatesfor some i 6≡ j mod p, then xj−i sends µxi

to its complex conjugate character, andhence x2(j−i) fixes µxi

. By (i), this can only occur when p = 2. �

Our first special cases, considered below, are fairly easy to handle.

Lemma 4.3. Let G = A o X be a nonabelian group, where A / G and X = 〈x〉 iscyclic of prime order p ≥ 5. Assume that either A is cyclic of prime power orderor A is abelian of type (p, p). Then, for some a ∈ A, there exist Bass cyclic unitsuk,t(a) and ur,s(x) that generate a nonabelian free subgroup of the unit group of theintegral group ring Z[G].

Proof. If A is cyclic, we take a ∈ A to be a generator of the group. By assumption,a has order qk for some prime q. Since |Aut(A)| = (q − 1)qk−1 and since G isnonabelian, it follows that either q = p or p divides q − 1. In either case, we haveq ≥ 5. On the other hand, if A is abelian of type (p, p), then we can take a ∈ Ato be any element that is not central in G. Now choose any two Bass cyclic unitsuk,t(a) and ur,s(x) with k 6≡ ±1 modulo the order of a and with r 6≡ ±1 mod p.Since a and x both have odd order ≥ 5, one possibility here is that k = r = 2.

Since G is nonabelian, there exists a nonlinear irreducible representation X ofC[G] ⊇ Z[G] with associated character χ : G → C. Say X : C[G] → Mu(C) for someu > 1. By the preceding lemma, we have u = p, and we can assume that X(a) =diag(α1, α2, . . . , αp) for suitable αi ∈ C. Indeed, if A is cyclic, then α1, α2, . . . , αp

are distinct qkth roots of unity and no two of these are complex conjugates of eachother. On the other hand, if A is abelian of type (p, p), then we know that χ(a) = 0and therefore {α1, α2, . . . , αp} are all the complex pth roots of unity.

Note that

S = X(uk,t(a)) = diag(uk,t(α1), uk,t(α2), . . . , uk,t(αp)).

Page 16: LINEAR GROUPS AND GROUP RINGSpassman/lingroup.pdf · 2005. 10. 4. · LINEAR GROUPS AND GROUP RINGS J. Z. GONC¸ALVES AND D. S. PASSMAN Abstract. This paper consists of two parts.

16 J. Z. GONCALVES AND D. S. PASSMAN

In particular, if A is cyclic, then by Lemma 3.5(ii) the eigenvalues of S have distinctabsolute values, and hence we can choose an S-decomposition of the space V = Cp

with dim S+ = dim S− = 2. Similarly, by Lemma 3.4(ii)(iii), this holds for anynoncentral a ∈ A if A is abelian of type (p, p). Thus we have V = S+ ⊕ S0 ⊕ S−and we denote by σ+ and σ− the projections of V into S+ and S−, respectively.Certainly, there exist distinct subscripts i, j, i′, j′ satisfying σ+ = ei,i + ej,j andσ− = ei′,i′ + ej′,j′ .

Next, by Lemma 4.2(ii), there exists a suitably described unitary matrix P withP−1X(x)P = D = diag(ε1, ε2, . . . , εp), where ε1, ε2, . . . , εp are the p distinct com-plex pth roots of unity, written in any order we choose. Thus, if T = X(ur,s(x)),then P−1TP = diag(ur,s(ε1), ur,s(ε2), . . . , ur,s(εp)). In other words, we know theeigenvalues of T and, by Lemma 3.4(ii)(iii), precisely two of these have the largestabsolute value and precisely two have the smallest absolute value. These thereforegive rise to a T -decomposition V = T+ ⊕ T0 ⊕ T− of V with dimT+ = dim T− = 2.Indeed, by ordering the eigenvalues of X(x) appropriately, we can assume that ε1

and ε2 yield the largest absolute value, while ε3 and ε4 yield the smallest. It thenfollows that the corresponding projections τ+ and τ− satisfy τ+ = P (e1,1+e2,2)P−1

and τ− = P (e3,3 + e4,4)P−1.It remains to verify the eight idempotent conditions of Corollary 4.1. For this,

we will just consider σ+τ− and τ−σ+, since the six additional products follow inthe same manner. To start with, note that σ+τ− = (ei,i + ej,j)·P (e3,3 + e4,4)P−1,so σ+τ− and (ei,i + ej,j)P (e3,3 + e4,4) have the same rank. Furthermore, the lattermatrix has only four nonzero entries and these form the 2× 2 submatrix

1√

εi−13 εi−1

4

εj−13 εj−1

4

which has a nonzero determinant. Thus we conclude that rankσ+τ− = 2. On theother hand, since P is a unitary matrix, it is clear that both σ+ and τ− are Herme-tian. It follows that τ−σ+ = (σ+τ−)∗, and therefore rank τ−σ+ = rankσ+τ− = 2.

It follows from Corollary 4.1 that there exist positive integers m and n suchthat 〈Sm, Tn〉 = 〈Sm〉∗〈Tn〉 is a free group of rank 2. Thus, since S = X(uk,t(a))and T = X(ur,s(x)), we conclude that uk,t(a)m and ur,s(x)n generate a nonabelianfree subgroup of the unit group of Z[G]. But uk,t(a)m = uk,tm(a) and ur,s(x)n =ur,sn(x) by Lemma 3.1(ii), so the result follows. �

When G = A o X with A an elementary abelian q-group of order > q, then theabove proof cannot apply since the eigenvalues of X(a) for any a ∈ A are difficultto control. Thus we are forced to take a different approach, and for this we need

Lemma 4.4. Let 〈x〉 be a group of prime order p acting faithfully and irreduciblyon an elementary abelian q-group A. Here p ≥ 5, q ≥ 3 is a prime different from p,and |A| > q. If 1 6= a ∈ A, then the p− 1 elements a1+x, a1+x2

, . . . , a1+xp−1cannot

all be 〈x〉-conjugate.

Proof. Note that CA(x) = 1 and hence 1+x+x2+ · · ·+xp−1 = 0 in its action on A.Let 1 6= a ∈ A and suppose, by way of contradiction, that a1+x, a1+x2

, . . . , a1+xp−1

are all 〈x〉-conjugate. Since a〈x〉 = A, it follows that these p − 1 elements are alldistinct and therefore if b ∈ A is the pth element of this 〈x〉-conjugacy class, thenb∏p−1

i=1 a1+xi ∈ CA(x) = 1. Thus, since x + x2 + · · · + xp−1 = −1, we conclude

Page 17: LINEAR GROUPS AND GROUP RINGSpassman/lingroup.pdf · 2005. 10. 4. · LINEAR GROUPS AND GROUP RINGS J. Z. GONC¸ALVES AND D. S. PASSMAN Abstract. This paper consists of two parts.

LINEAR GROUPS AND GROUP RINGS 17

that bap−1a−1 = 1 and hence b = a2−p. It follows that there exists a one-to-onefunction f from {1, 2, . . . , p− 1} to itself such that

a(2−p)xi

= bxi

= a1+xf(i)

for all 1 ≤ i ≤ p− 1. In particular, since a〈x〉 = A, we see that (2− p)xi = 1 + xf(i)

as operators on A.Since 〈x〉 acts irreducibly on A, we can now view A as the additive group of

GF(qn). Furthermore, x can be taken to be an element of order p in this fieldacting by right multiplication on A, and x generates GF(qn) over the prime subfieldGF(q). The operator equations (2− p)xi = 1+xf(i) are now equations in the field.In particular, if i = f(i) for some 1 ≤ i ≤ p − 1, then xi ∈ GF(q), so x ∈ GF(q)and n = 1, contrary to our assumption.

Next, we multiply the equation (2 − p)xi = 1 + xf(i) by x−i, and then settingj ≡ −i mod p, we get

(2− p) = xj + xg(j),

where g(j) = f(i) − i 6≡ 0 mod p. Thus g is also a one-to-one function from{1, 2, . . . , p − 1} to itself, and by summing the above displayed equation over allsuch j, we obtain (2− p)(p− 1) ≡ (−1) + (−1) = −2 mod q. Thus p2 ≡ 3p mod qand hence p ≡ 3 mod q since p 6= q. In particular, q 6= 3 and 2− p ≡ −1 mod q, sowe have 1 + xj = −xg(j) for all j = 1, 2, . . . , p− 1. Since p is odd and xp = 1, thisyields (1 + xj)p = −1.

In other words, x, x2, . . . , xp−1 ∈ GF(qn) are all roots of the polynomial equa-tions (1+ζ)p = −1 and ζp = 1 in GF(q)[ζ]. Hence, they are roots of 2+(1+ζ)p−ζp,a polynomial of degree p− 1. It follows that this polynomial must be a scalar mul-tiple of 1 + ζ + ζ2 + · · ·+ ζp−1 and, by considering the constant term, we see thatthis scalar is 3. We conclude that

(pk

)≡ 3 mod q for k = 1, 2, . . . , p − 1. However,

since p ≡ 3 mod q and p, q ≥ 5, we see that(p3

)≡ 1 6≡ 3 mod q, and this is the

required contradiction. �

We briefly comment on the above situation for the missing small primes. Ifp = 2, then |A| = q, so this case does not occur. If p = 3, then 1 + x + x2 = 0,so a1+x = a−x2

and a1+x2= a−x. Hence, these elements are always 〈x〉-conjugate.

Finally, if q = 2, then the preceding proof yields (1 + ζ)p = 1 + ζ + · · ·+ ζp−1 + ζp

in the polynomial ring GF(2)[ζ]. Thus (1 + ζ)p+1 = 1 + ζp+1 and it follows easilythat p must be a Mersenne prime. Conversely, if p = 2n − 1 is such a prime, then|A| = 2n and all nonidentity elements of A are 〈x〉-conjugate.

At this point, it is convenient to isolate certain matrix computations. We assumein that following that G = A o X, X, χ, P and Q = P ∗ are as in Lemma 4.2.

Lemma 4.5. Let i 6= i′ be subscripts with εi′ = εi, and let j 6= j′ be subscripts withεj′ = εj Furthermore, let a ∈ A be an element of order prime to p, and assume thatthe matrix

M = (ei,i + ei′,i′)·QX(a)P ·(ej,j + ej′,j′)

does not have rank 2.

i. If i = j or j′, then χ(aaxd)

= χ(aax

)for all d = 1, 2, . . . , p− 1.

ii. If i 6= j and i 6= j′, then χ(aaxd)

= χ(aaxtd)

for all d = 1, 2, . . . , p − 1,where t 6≡ ±1 mod p satisfies εj/εi = (εj/εi)t.

Page 18: LINEAR GROUPS AND GROUP RINGSpassman/lingroup.pdf · 2005. 10. 4. · LINEAR GROUPS AND GROUP RINGS J. Z. GONC¸ALVES AND D. S. PASSMAN Abstract. This paper consists of two parts.

18 J. Z. GONCALVES AND D. S. PASSMAN

Proof. Let a have order q so that, by assumption, q is prime to p, and write X(a) =diag(α0, α1, . . . , αp−1), where each αd is a qth root of unity and where we view thesubscripts modulo p. Observe that X(x)−1X(a)X(x) = diag(αp−1, α0, . . . , αp−2)and thus X(x)−dX(a)X(x)d = diag(α0−d, α1−d, . . . , αp−1−d). It follows that

X(aaxd

) = diag(α0α0−d, α1α1−d, . . . , αp−1αp−1−d)

and therefore χ(aaxd)

=∑p−1

r=0 αrαr−d.Next, if ρ is any pth root of unity, we set Tr(ρ, a) =

∑p−1r=0 ρrαr. Then

Tr(ρ, a)Tr(ρ−1, a) =p−1∑r=0

ρrαr·p−1∑s=0

ρ−sαs

=p−1∑d=0

ρd·p−1∑r=0

αrαr−d =p−1∑d=0

ρd·χ(aaxd)

.

Now note that the (i, j)th entry of QX(a)P is equal to

1p

p−1∑r=0

εri αrε

rj =

1pTr(εj/εi, a),

since Q = P ∗. In particular, since εi′ = εi and εj′ = εj , we see that the 2 × 2submatrix of M corresponding to rows i and i′ and columns j and j′ is given by

M2×2 =1p·

Tr(εj/εi, a) Tr(εj/εi, a)

Tr(εj/εi, a) Tr(εj/εi, a)

.

Setting σ = εj/εi and τ = εj/εi, we see that σ 6= τ, τ and

M2×2 =1p·

Tr(σ, a) Tr(τ , a)

Tr(τ, a) Tr(σ, a)

.

Since rankM 6= 2, by assumption, it follows that detM2×2 = 0 and thereforeTr(σ, a)·Tr(σ, a) = Tr(τ, a)·Tr(τ , a). In other words, we have

(∗)p−1∑d=0

σd·χ(aaxd)

=p−1∑d=0

τd·χ(aaxd)

.

Now a is an element of order q, so each χ(aaxd)

is contained in Q[δ], where δ isa primitive complex qth root of unity. In particular, (∗) is a polynomial equationsatisfied by a primitive pth root of unity over Q[δ] and, since gcd(p, q) = 1, we knowthat any such equation is a multiple of 1 + ζ + · · ·+ ζp−1.

Suppose first that i = j. Then εi = εj , so σ = 1 and τ is a primitive pth rootof unity. Since (∗) is a polynomial equation in τ of degree ≤ p − 1, we concludefrom the above remarks that all coefficients of the powers of τ must be equal. Inparticular, χ

(aaxd)

= χ(aax

)for all d = 1, 2, . . . , p − 1. Similarly, if i = j′, then

τ = 1 and the same argument applies.On the other hand, if i 6= j and i 6= j′, then both σ and τ are primitive pth roots

of unity, and we can write τ = σt for some t ∈ {1, 2, . . . , p−1}. Indeed, since τ 6= σ

Page 19: LINEAR GROUPS AND GROUP RINGSpassman/lingroup.pdf · 2005. 10. 4. · LINEAR GROUPS AND GROUP RINGS J. Z. GONC¸ALVES AND D. S. PASSMAN Abstract. This paper consists of two parts.

LINEAR GROUPS AND GROUP RINGS 19

and τ 6= σ, we have t 6= 1 or p− 1. Viewing the exponents modulo p, equation (∗)now becomes

p−1∑d=0

σd·χ(aaxd)

=p−1∑d=0

σdt·χ(aaxd)

,

and since the d = 0 coefficients match, we see that χ(aaxd)

= χ(aaxdt)

for all d.Of course, εj/εi = τ = σt = (εj/εi)t, so the result follows. �

With Lemmas 4.4 and 4.5 in hand, we can now deal with the Frobenius case.Unlike the proof of Lemma 4.3, where the first parameters k and r are essentiallyarbitrary, here our proof requires that k be selected in a rather careful manner.

Lemma 4.6. Let X = 〈x〉 be a cyclic group of prime order p that acts faithfullyand irreducibly on an elementary abelian q-group A, with q a prime different fromp and with |A| > q. If p ≥ 5 and q ≥ 3, then for any 1 6= a ∈ A, there existsuitable Bass cyclic units uk,m(x) and uk,m(a−1xa) that generate a nonabelian freesubgroup of the unit group of Z[A o X].

Proof. Write G = A o X and let 1 6= a ∈ A. Since |A| > q, Lemma 4.4, impliesthat the elements a1+x, a1+x2

, . . . , a1+xp−1cannot be all X-conjugate to a1+x. In

other words, there exists t ∈ {1, 2, . . . , p− 1} with a1+x not X-conjugate to a1+xt

.Since (a1+x)x−1

= a1+x−1, it is clear that t 6= 1 or p− 1. Thus 2 ≤ t ≤ p− 2, and

we setk ≡ (t− 1)/(t + 1) mod p.

Certainly, k 6≡ 0 mod p and, since t ≡ (1 − k)/(1 + k) mod p, it is clear thatk 6≡ ±1 mod p. Thus we can assume that 2 ≤ k ≤ p − 2. For any suitable integerm, for example m = ϕ(p) = p−1, we can now consider the Bass cyclic units uk,m(x)and uk,m(a−1xa) = a−1uk,m(x)a.

Since G acts on A as X does, it follows that a1+x and a1+xt

are not G-conjugate.Thus there exists an irreducible representation X of C[G], with corresponding char-acter χ, such that χ(a1+x) 6= χ(a1+xt

). Taking complex conjugates, we see thatχ(a−(1+x)) 6= χ(a−(1+xt)). Now X acts faithfully and irreducibly on A, so G isnonabelian and clearly A = G′, the commutator subgroup of G. It follows thatχ(1) 6= 1 and therefore Lemma 4.2 applies. In particular, χ(1) = p and we canassume that X(x) and X(a) are as described in that lemma. Furthermore, let ε bethe primitive complex pth root of unity whose angle with the real axis is 2π/p, andthen use ε` = ε` in the matrix P , for all ` = 1, 2, . . . , p.

Now Lemma 4.2(ii) implies that P−1X(x)P = D = diag(ε1, ε2, . . . , εp) and henceT = X(uk,m(x)) satisfies P−1TP = diag(uk,m(ε1), uk,m(ε2), . . . , uk,m(εp)). In otherwords, we know the eigenvalues of T and, since 2 ≤ k ≤ p − 2, Lemma 3.4(ii)(iii)implies that there are precisely two of these that have largest absolute value, namelyuk,m(εi) and uk,m(εi′) with i = 1, i′ = p − 1, and εi′ = εi. Furthermore, thereare precisely two of these that have smallest absolute value, namely uk,m(εj) anduk,m(εj′) with j ≡ k−1 mod p, j′ ≡ −k−1 mod p, and εj′ = εj . These thereforegive rise to a T -decomposition Cp = V = T+⊕T0⊕T− with dim T+ = dim T− = 2.Since Q = P−1, the corresponding projections τ+ : V → T+ and τ− : V → T−satisfy τ+ = P (ei,i + ei′,i′)Q and τ− = P (ej,j + ej′,j′)Q.

Next, let S = X(uk,m(a−1xa)) = X(a)−1X(uk,m(x))X(a). Since S is conjugateto T , it is clear that there is an S-decomposition V = S+⊕S0⊕S− with dim S+ =

Page 20: LINEAR GROUPS AND GROUP RINGSpassman/lingroup.pdf · 2005. 10. 4. · LINEAR GROUPS AND GROUP RINGS J. Z. GONC¸ALVES AND D. S. PASSMAN Abstract. This paper consists of two parts.

20 J. Z. GONCALVES AND D. S. PASSMAN

dim S− = 2. Furthermore, the corresponding projections σ+ and σ− satisfy

σ+ = X(a)−1τ+X(a) = X(a)−1P (ei,i + ei′,i′)QX(a), and

σ− = X(a)−1τ−X(a) = X(a)−1P (ej,j + ej′,j′)QX(a).

Our goal now is to apply Corollary 4.1 to S and T , and for this we need to verifythe eight idempotent conditions, namely that the eight possible products all haverank 2. However, since P and X(a) are unitary matrices and since ei,i + ei′,i′ andej,j +ej′.j′ are Hermetian, we see that σ± and τ± are also Hermetian. In particular,(τ±σ±)∗ = σ∗±τ∗± = σ±τ±, so rank τ±σ± = rankσ±τ±. It follows that we need onlycheck that the four products σ±τ± each have rank 2.

Next, since P , Q, X(a) and X(a)−1 are nonsingular matrices, we see that

rank σ±τ± = rank (er,r + er′,r′)·QX(a)P ·(es,s + es′,s′),

where r = i or j and s = i or j. Two of these cases are quite simple. Namely, ifr = s, then Lemma 4.5(i) asserts that if these matrices do not have rank 2, thenχ(a1+x) = χ(a1+xd

) for all d = 1, 2, . . . , p − 1, and this is a contradiction since weknow that χ(a1+x) 6= χ(a1+xt

). Thus, only two cases remain. But observe that

[(ej,j + ej′,j′)·QX(a)P ·(ei,i + ei′,i′)]∗

= (ei,i + ei′,i′)·QX(a−1)P ·(ej,j + ej′,j′),

and we know that both χ(a1+x) 6= χ(a1+xt

) and χ(a−(1+x)) 6= χ(a−(1+xt)).Thus, we need only consider the rank of

M = (ei,i + ei′,i′)·QX(a)P ·(ej,j + ej′,j′),

since the other argument will be similar. Now if rank M 6= 2, and if u is definedby εj/εi = (εj/εi)u, then Lemma 4.5(ii) asserts that χ(a1+x) = χ(a1+xu

). But,working with exponents modulo p, we have εj/εi = εk−1+1 and εj/εi = εk−1−1, sou(k−1 − 1) ≡ k−1 + 1 mod p and therefore u ≡ (1 + k)/(1− k) ≡ t mod p. This is,of course, a contradiction by our choice of the parameters t and k.

It follows that all the idempotent conditions are verified and we conclude fromCorollary 4.1 that, for some positive integer n, we have 〈Sn, Tn〉 = 〈Sn〉∗〈Tn〉is a free group of rank 2. Since X(uk,m(a−1xa)n) = Sn and X(uk,m(x)n) = Tn,Lemma 3.1(ii) implies that uk,mn(xa) = uk,m(xa)n and uk,mn(x) = uk,m(x)n gen-erate a nonabelian free subgroup of the unit group of Z[G]. �

We can now quickly prove our main result.

Theorem 4.7. If G is a finite nonabelian group of order prime to 6, then thereexist two elements x, y ∈ G of prime power order and two Bass cyclic units uk,m(x)and ur,s(y) such that 〈uk,m(x), ur,s(y)〉 is a nonabelian free subgroup of the unitgroup of the integral group ring Z[G].

Proof. We proceed by induction on |G|. Certainly, we can assume that all propersubgroups of G are abelian. Furthermore, suppose G is a proper homomorphicimage of G that is nonabelian and, by induction, let x, y ∈ G be elements of primepower order such that uk,m(x) and ur,s(y) generate a nonabelian free group. ByLemma 3.2, there exist elements x, y ∈ G of prime power order such that uk,m′(x)and ur,s′(y) map to powers of uk,m(x) and ur,s(y), respectively, under the naturalhomomorphism Z[G] → Z[G]. Thus 〈uk,m′(x), ur,s′(y)〉 is the required free group ofunits in Z[G]. Thus, we can now also assume that all proper homomorphic images

Page 21: LINEAR GROUPS AND GROUP RINGSpassman/lingroup.pdf · 2005. 10. 4. · LINEAR GROUPS AND GROUP RINGS J. Z. GONC¸ALVES AND D. S. PASSMAN Abstract. This paper consists of two parts.

LINEAR GROUPS AND GROUP RINGS 21

of G are abelian. It therefore follows from [M] that G = A o X, where X is cyclicof prime order p. Furthermore, either A is cyclic, A is abelian of type (p, p), or Ais an elementary abelian q-group for some prime q 6= p. In the latter situation, Xacts faithfully and irreducibly on A. The result now follows from the special casesalready obtained in Lemmas 4.3 and 4.6. �

Note that some assumption on the primes dividing |G| is required in the abovebecause, as we have seen, there are no Bass cyclic units for group elements of order2 or 3. For example, suppose G = A o X, where X is cyclic of order p = 2 or 3and where X acts in a fixed-point-free manner on the normal abelian subgroup A.Then G is a Frobenius group, so all elements of G \A have order p = 2 or 3. Thusthe only Bass cyclic units of the integral group ring Z[G] come from elements ofA and these all commute. In particular, we cannot find two Bass cyclic units thatgenerate a nonabelian free group.

References

[B] N. Bourbaki, Elements of Mathematics, Commutative Algebra, Hermann, Paris, 1972.

[H] P. de la Harpe, Topics in Geometric Group Theory, Chicago Lectures in Mathematics, Univ.of Chicago Press, Chicago, 2000.

[I] I. M. Isaacs, Character Theory of Finite Groups, Academic Press, New York, 1976.

[J] N. Jacobson, Basic Algebra II, Freeman, San Francisco, 1980.[M] G. Miller and H. Moreno Non-abelian groups in which every subgroup is abelian, Trans. AMS

4 (1903), 398–404.

[P] D. S. Passman, Free products in linear groups, Proc. AMS 132 (2004), 37–46.[S] S. K. Sehgal, Units in Integral Group Rings, Longman Scientific, Harlow, 1993.

[T] J. Tits, Free subgroups in linear groups, J. Algebra 20 (1972), 250–270.

Department of Mathematics, University of Sao Paulo, Sao Paulo 05389-970, Brazil

E-mail address: [email protected]

Department of Mathematics, University of Wisconsin, Madison, Wisconsin 53706

E-mail address: [email protected]


Recommended