Skip to content
Publicly Available Published by De Gruyter October 30, 2018

Maximal deviation of large powers in the Nottingham group

  • Tuǧba Aslan EMAIL logo and Pál Hegedüs
From the journal Journal of Group Theory

Abstract

The group of formal power series under substitution over 𝔽p, the so-called Nottingham group, is a pro-p group so it is a metric space. The intertwining of these two structures is an important object of study. This paper is concerned with how the power map influences the distance of elements. Suppose that d(f,g)=d1 while d(f,1)=d2d1. In this paper we provide a sharp bound for d(fj,gj) in terms of d1,d2 and the exponent j. This bound confirms a conjecture by K. Keating.

1 Introduction

Let 𝔽p be the field of order p, for a fixed prime p. Let 𝒩 be the Nottingham group over 𝔽p which is the group of all formal power series f(x)𝔽p[[x]] with leading term x, under formal substitution: given g,f𝒩 set (fg)(x)=f(g(x)). For brevity, f is used instead of f(x) if it is not confusing. Consider the following chain of open subgroups of 𝒩:

𝒩k={g𝒩:g=x(1+i=kαixi),αi𝔽p}for each k.

The depth of f𝒩{1}, D(f), is the integer k1 such that f𝒩k𝒩k+1, while D(1)=. The distance of f and g is usually defined by

𝒩d(g,f)=p-D(fg-1)

which makes 𝒩 an ultrametric space. It is clear that d(f,g)=d(fj,gj) if pj. So in understanding the distance of powers the crucial case is when the exponent j is a power of p.

The following definition is by Keating [3].

Definition 1.

Let nk1 and let k0 be the least nonnegative residue of k modulo p. Set

e(k,n)={0if pk and n=k,1if pk,pn, and n>k,0if pk and pn,iif pk and n2k-i(modp) for some 0ik0,k0if pk and n2k-i(modp) for all 0ik0.

He defined this constant to prove the following theorem [3, Corollary 2] and to show that for m=1 the bound is sharp [3, Theorem 1 (b)].

Theorem 1.1 (Keating).

Let p be a prime and nk1. Suppose f,gN are such that D(f)k and D(gf-1)n. Then for all m1 we have

D(gpmf-pm)n+(pm-1)k+pm-pp-1k0+e(k,n).

He conjectured that the bound was sharp for every m>1 as well. In this paper we confirm his conjecture except for the case p=2, k=1, when it is not sharp for m>2.

Theorem 1.2.

Let p be a prime and nk1. If p=2, then assume k>1. Then there exist f,gN such that D(f)=k and D(gf-1)=n and for all m1 we have

D(gpmf-pm)=n+(pm-1)k+pm-pp-1k0+e(k,n).

For p=2, k=1 we have D(g2mf-2m)n+2m+2-11+e(k,n) which is strictly stronger than the bound of Theorem 1.1 if m>2. This and the exact bounds for the various values of n are treated fully in [1].

Theorems 1.1 and 1.2 have the following metric version.

Theorem 1.3.

Let d1=p-nd2=p-k. Suppose that f,gN are such that d(f,g)d1 and d(f,1)d2. Let j=pmj be an integer such that pj. Then

d(fj,gj)d1d2pm-1p-(pm-pp-1k0+e(k,n))

and for every choice of d1,d2,j (except for p=2, d2=p-1) there exist f,g as above such that equality holds.

Crucial in the proof is the understanding of the Engel words. The general formulas (see Lemma 2.3 and Lemma 2.4), albeit aesthetic, are not sufficient for direct constructions. But reinterpreting them in linear algebraic terms allows direct computation of the leading terms of [a,s(p-1)b] for arbitrary s0, see the formulation in Corollary 3.4.

The choice for f,g in Theorem 1.2 provides controllably increasing sequences D(f)<D(fp)<D(fp2)< and D(g)<D(gp)<D(gp2)<. In fact, these are the slowest possible, see Lemma 2.2.

Corollary 1.4.

Let p>2 be prime, and nk1 integers. Suppose that f,gN are such that Theorem 1.2 is satisfied with them. Then for all m1 we have

(1.1)D(fpm)=pmk+pm-1p-1k0.

If n>k, then we also have

D(gpm)=pmk+pm-1p-1k0.

2 Preliminaries

First we take a closer look at the commutators in the Nottingham group. Let a,b𝒩. The commutator of a and b is [a,b]=[a,1b]=a-1b-1ab. The Engel word [a,mb] is defined inductively as [a,mb]=[[a,m-1b],b] for m2.

Lemma 2.1.

Let g,hN with D(g)=n, D(h)=k. Write g=x+γ(x) and h=δ(x), where deg(γ(x))=n+1 and deg(δ(x)-x)=k+1. Then

[g,h]x-γ(x)+γ(δ(x))δ(x)(modx2n+k+1),

where δ(x) is the formal derivative of δ(x).

Proof.

We know that hg=δ(x+γ(x)) and gh=δ(x)+γ(δ(x)). Let us denote

ε(x)=-γ(x)+γ(δ(x))δ(x).

To prove the claim, we have to verify

(2.1)δ(x+ε(x)+γ(x+ε(x)))δ(x)+γ(δ(x))(modx2n+k+1).

Note first that

γ(δ(x))γ(x)+γ(x)(δ(x)-x)(modxn+2k+1),

so

ε(x)=γ(δ(x))-γ(x)δ(x)δ(x)
γ(x)(δ(x)-x)-γ(x)(δ(x)-1)δ(x)(modxn+2k+1).

Here the numerator has degree at least n+k+1 (with equality if and only if nk(modp)), so ε(x) has degree at least n+k+1. Consequently,

γ(x+ε(x))γ(x)(modx2n+k+1).

The required approximation (2.1) follows:

δ(x+ε(x)+γ(x+ε(x)))δ(x+ε(x)+γ(x))
=δ(x+γ(δ(x))δ(x))
δ(x)+γ(δ(x))(modx2n+k+1).

For later use we state the following two simple corollaries of Lemma 2.1, which are well known and fundamental for the Nottingham group. The first was noted for ε(x) in the proof, the second follows from the first:

(2.2)D([g,h])D(g)+D(h),with “=” if and only if D(g)D(h)(modp),D(gp)D(g)(modp),if gp1.

By [2, Lemma 1], D(gp)pD(g) if gp1. Combined with (2.2) we get the form we will use.

Lemma 2.2.

Let k0 be the remainder of D(g) divided by p. Then

D(gp)pD(g)+k0.

By induction, for every m1,

D(gpm)pmD(g)+pm-1p-1k0.

Theorem 1.1 is, in fact, a generalisation of this lemma.

The next two results will be used in the proof of the main theorem but they are also of independent interest. The first is a generalisation of Lemma 2.1. Note that it also holds when b=1, k=.

Lemma 2.3.

Let a,bN with D(a)=n and D(b)=k. Suppose that a(x)=x+α(x) and b(x)=β(x). Set β0(x)=x, β1(x)=β(x), βi(x)=βi-1(β(x)) for i1. Then

(2.3)[a,mb]x+i=0m(mi)(-1)m+iα(βi(x))βi(x)(modx2n+mk+1).

Moreover, if pk and either mp+1 or m=p,p2n-k, then (2.3) is valid modulo x2n+mk+2.

Proof.

The proof is by induction on m. The m=1 case is exactly Lemma 2.1. Suppose that we have the result for m1, that is we can write [a,mb]=cd, where

c=x+i=0m(mi)(-1)m+iα(βi(x))βi(x)

and D(d)2n+mk. Now

[a,m+1b]=[cd,b]=[c,b][c,b,d][d,b],

and here D([c,b,d])D(c)+D(d)+D(b)>2n+(m+1)k. On the other hand we have

D([d,b])D(b)+D(d)

with equality if and only if D(b)D(d)(modp), see (2.2). So

[a,m+1b][c,b](modx2n+(m+1)k+1)

holds unconditionally. But even

[a,m+1b][c,b](modx2n+(m+1)k+2)

holds if

(2.4)D(d)>2n+mk,ork=D(b)2n+mk(modp).

If pk, then there is a unique 1m0p which is a solution of

k2n+m0k(modp).

This m0=p if only if p2n-k. For m=m0 the second condition of (2.4) holds, while for m>m0 the first does.

For [c,b] we apply Lemma 2.1. As D(c)D(a)+mD(b)=n+mk by (2.2), the congruence is valid modulo x2n+(2m+1)k+1

[c,b]x-i=0m(mi)(-1)m+iα(βi(x))βi(x)+i=0m(mi)(-1)m+iα(βi(β(x)))βi(β(x))β(x)
x+i=0m(mi)(-1)m+i+1α(βi(x))βi(x)+i=0m(mi)(-1)m+iα(βi+1(x))βi+1(x)
x+(-1)m+1α(x)+α(βm+1(x))βm+1(x)
+i=1m((mi)+(mi-1))(-1)m+i-1α(βi(x))βi(x)
x+i=0m+1(m+1i)(-1)m+1+iα(βi(x))βi(x)(modx2n+(2m+1)k+1)

Thus we are done by induction. ∎

Lemma 2.4.

Let a,bN with D(a)=n and D(b)=k. Then

(2.5)[a,mbpl][a,mplb](modx2n+mplk+1)for m,l1.

Moreover, if pk, then (2.5) is valid modulo x2n+mplk+2, unless p>2 and m=l=1.

Proof.

Note that bpl is of depth at least plD(b) by Lemma 2.2. Note also that

(mplj)=0if plj

and

(mplipl)=(mi)

in 𝔽p, a field of characteristic p. We also clearly have (-1)m+j=(-1)pl(m+j). So over 𝔽p

x+i=0m(mi)(-1)m+iα(βipl(x))βipl(x)=x+j=0mpl(mplj)(-1)mpl+jα(βj(x))βj(x).

By Lemma 2.3 the first expression is [a,mbpl] modulo x2n+mD(bpl)+1, while the second is [a,mplb] modulo x2n+mplk+1, as required.

If pk then D(bpl)>plk, by Lemma 2.2, so we get the refinement from Lemma 2.3. (Note that for p=2, k2n(modp) implies that k is even.) ∎

Now we will describe Keating’s constructions and what we will need from his results. Let b,u𝒩 such that

b=x+xk+1β(x),

where β(x)=rk+rk+1x+rk+2x2+, and

D(u)=nk+k0.

Assume pk. For sk, h1 we construct matrices Ah:=Ah,n,sMs×s(𝔽p). This matrix Ah is a linear operator that takes s-many coefficients of [u,h-1b] starting from x(h-1)k+n+j+1 and gives s-many coefficients of [u,hb] starting from xhk+n+j+1. This matrix is constructed formally using formula (2.9), below. The product Πp-1:=A1A2Ap-1 will be our main concern. It is an operator that takes s-many “leading” coefficients of u and gives s-many “leading” coefficients of [u,p-1b].

Define w1=u and wh+1=[wh,b] for h0. Following [3], we determine the coefficients of wh+1 from those of

wh=x+x(h-1)k+n+1θ(x),

where θ(x)=t0+t1x+t2x2+:

(2.6)

[wh,b]x+((h-2)k+n)xhk+n+1β(x)θ(x)
+xhk+n+2(β(x)θ(x)-β(x)θ(x))(modx(h+1)k+n+1).

It follows from the expansion above that for 0jk-1 the coefficient of xhk+n+j+1 in wh+1 is

(2.7)i=0j((h-2)k+n+2i-j)rk+j-iti=i=0j(nh+2i-j)rk+j-iti,

using the abbreviation

(2.8)nh=(h-2)k+n.

For sk and h1 we define the matrix Ah=Ah,n,sMs×s(𝔽p) as

(2.9)Ah=(nhrk(nh-1)rk+1(nh-2)rk+2(nh-s+1)rk+s-10(nh+1)rknhrk+1(nh-s+3)rk+s-2000(nh+s-1)rk).

Clearly Ah depends only on s, b and the remainder of n modulo p.

Let now νh𝔽ps be the row vector whose entries are the coefficients of x(h-1)k+n+j+1 in wh for 0js-1. It follows from (2.7) that

νh+1=νhAh,n,s.

For h1 define a matrix Πh=Πh,n,sMs×s(𝔽p) by setting Πh=A1A2Ah. Then we have νp=ν1Πp-1.

Recall now the definition of k0 and e(k,n) from Definition 1. Keating introduced the matrices above for s=e(k,n)+1, and we follow his notation. He went on to show that in his e(k,n)+1×e(k,n)+1 matrix Πp-1 the first e(k,n) columns are 0 (see [3, Cases 3–4]). (NB: He claimed this only for n>(p-1)k+p, because he used it only in that case, however it works also for arbitrary nk+k0.) Using our extended notation, we may restate his result as follows:

Lemma 2.5.

For pk and e=e(k,n) we have Πp-1,n,e=0.

From this we instantly get:

Corollary 2.6.

Let pk>k0 so 2k0<k and write e=e(k,n). For a fixed s<2k0+1 denote by T the top right-hand corner s-e×s-e block of Πp-1,n,s and let ν be the first s-e coefficients of ν1. The first e entries of νp are 0. The remaining s-e entries are νT.

Proof.

The top left-hand corner e×e block of Πp-1,n,s is zero by Lemma 2.5. Similarly, the bottom right-hand corner e×e block of Πp-1,n,s, which is exactly Πp-1,n+e,e, is zero by Lemma 2.5. As the matrix Πp-1,n,s is a triangular matrix, nonzero entries of Πp-1,n,s only occur in the top right-hand corner s-e×s-e block of Πp-1,n,s, which is called T. So, the first e entries of νp are 0 and the remaining s-e entries are νT. ∎

In the rest of the paper our concern will be this top right-hand corner block, T.

3 Constructions

Lemma 4 of [3] can be generalized as the following lemma, the proof being almost identical.

Lemma 3.1.

Suppose that n>nk are such that Theorem 1.1 holds for (k,n), and Theorem 1.2 holds for (k,n). If

n+e(k,n)=n+e(k,n),

then Theorem 1.2 also holds for (k,n).

Consider a fixed k. By the definition of e(k,n), the sum n+e(k,n) is constant in the intervals k+tpnk+tp+k0 (with t0). Lemma 3.1 allows us to assume that k+k0+tpn<k+(t+1)p. If n=k+k0+tp, then e(k,n)=0, otherwise e(k,n)=k0. We are going to assume this in the proof of Theorem 1.2, below.

Let pk and k+k0+tpnk+(t+1)p for some nonnegative integer t. We define

e(k,n)={0if k+k0+tp<n<k+(t+1)p,k0if n=k+k0+tp or n=k+(t+1)p.

This describes the number of significant digits for the following definition. For u𝒩n let μn(u)𝔽pe(k,n)+1 denote the vector consisting of the first e(k,n)+1 coefficients of u.

For given n,k as above and λ𝔽p we define an e(k,n)+1×e(k,n)+1 matrix Tn[λ] as

(-λ002λ20λ00000λ0-1002λ)if nk(modp),

and

(-1002λ00000000)if n2k(modp),

and Tn[λ]=(λ) if k+k0+pt<n<k+p(t+1).

Lemma 3.2.

Let k>p, pk and k+k0+ptnk+(t+1)p for some nonnegative integer t. Define b=x+xk+1+xk+k0+1. For every uNn we have [u,p-1b]Nn* and

μn*([u,p-1b])=μn(u)Tn[-1],

where n*=n+(p-1)k+e(k,n).

Proof.

We have to show that the top right-hand e(k,n)+1×e(k,n)+1 block of Πp-1,n,e(k,n)+e(k,n)+1 is Tn[-1], by Corollary 2.6. From the definition in (2.9) and the choice of our b we see that in the matrices Ah the entries (i,j) are 0 unless j-i{0,k0}. Consequently, in the products Πh=A1Ah the entries (i,j) are 0 unless j-i is a nonnegative integer multiple of k0.

First we deal with the case n2k(modp). Pick some mk(modp). Now e(k,n)=0 and e(k,m)=k0, while e(k,n)=e(k,m)=k0. From the definition in (2.9) we also see the following block decomposition of the matrices Ah,m,2k0+1,Πh,m,2k0+1M(2k0+1)×(2k0+1)(𝔽p) for h1:

(3.1)Ah,m,2k0+1=(Ah,m,k0*0Ah,n,k0+1),Πh,m,2k0+1=(Πh,m,k0*0Πh,n,k0+1).

It is enough to compute the larger matrices Ah,m,2k0+1M(2k0+1)×(2k0+1)(𝔽p), and their product Πp-1,m,2k0+1M(2k0+1)×(2k0+1)(𝔽p).

For convenience we index the rows and columns of Ah from 0 to 2k0. By (2.9), the (i,j) entry of Ah for 0ij2k0 is

ahij={mh+iif i=j,mh-k0+iif j=k0+i,0otherwise,

where we used again the convention (2.8). Note that

mh+k0=(h-2)k+m+k0mh+1(modp).

Thus Ah has the form

0k02k00mh00mh-10000mh+100mh-1+1000000000k00000mh+100mh002k000000000mh+2.

As remarked above, the (i,j) entry of Πh is zero if j-i{0,k0,2k0}. Now we determine the rest.

By induction on h1, we get that the (i,i+k0) entry of Πh is

(t=1h-1(m1+i)(m2+i)(mt+i)2(mt+3+i)(mt+4+i)(mh+1+i))
(3.2)+(m0+i)(m3+i)(mh+1+i).

Finally, evaluating for h=p-1, we get that the (i,i+k0) entry of Πp-1 is

(3.3)

πi,i+k0=t=0p-2(mt+i)j=3p(mt+j+i).

Note that each summand in (3.3) is a product of p-2 consecutive numbers from {mt+i}t=1p, the last with multiplicity 2. Pick the unique l{1,,p} such that ml+i=0. So jl(mj+i)=-1 and hence

(ml-1+i)j=3p(ml-1+j+i)=--k0k0=1,
(ml-2+i)j=3p(ml-2+j+i)=--2k0-k0=-2,

and

(3.4)πi,i+k0={-2+1=-1if l1,p,-2if l=p,1if l=1.

By Lemma 2.5 we know that the first k0 columns and the last k0 rows of Πp-1 are zero. The remaining diagonal entry is

πk0,k0=m2m3m4mp-1=(2k0)(2k0)(3k0)((p-2)k0)=-1.

To determine π0,2k0 note that ap-1,2k0,2k0=mp+1=0 and ap-1,k0,2k0=mp-1 in Ap-1. So π0,2k0 is mp-1 times the (0,k0) entry of Πp-2. Using (3.2), we get

π0,2k0=m0m3m4mp-1mp-1=(-k0)(2k0)(3k0)((p-2)k0)2=2.

Thus, we have determined the top right-hand corner k0+1×k0+1 matrix of Πp-1,m,2k0+1 corresponding to mk(modp), it is

(3.5)Tm[-1]=(10020-100000-10-100-2).

By formulas (3.1) and (3.5), the top right-hand corner k0+1×k0+1 matrix of Πp-1,n,k0+1M(k0+1)×(k0+1)(𝔽p) corresponding to n2k(modp), is

Tn[-1]=(-100-200000000)

We now turn to the case of k+k0+pt<n<k+p(t+1), that is, we have n2k-i(modp) for any 0ik0. So e(k,n)=k0 and e(k,n)=0. Then the corresponding matrix Ah,n,k0+1 is

0k00nh00nh-10nh+1000000k00000nh+1.

As above, we determine the (i,j) entries of Πp-1=Πp-1,n,k0+1:

π0,k0=(t=1p-2n1n2nt2nt+3nt+4np)+n0n3np=t=0p-2ntj=3pnt+j.

Note that we get n1=n-k0, np=(p-2)k+n0 by our assumptions, so π0,k0=-1 as in (3.4). For the diagonal entries we again get

πi,i=t=1p-1(nt+i){0,-1},

with πi,i=-1 if and only if 0n0+i=n-2k+i(modp). In our case we have n2k-i(modp) for any 0ik0 so all πi,i=0. We cut the first k0 columns and last k0 rows and we get the matrix Tn[-1]=(-1). ∎

Lemma 3.3.

Let p>2. Suppose k=k0<p and k+k0+ptnk+(t+1)p for some nonnegative integer t. If k<p-1, then let b=x+xk+1, otherwise let b=x+xk+1+x2k+1. For every uNn we have [u,p-1b]Nn* and

μn*([u,p-1b])=μn(u)Tn[λ],

where n*=n+(p-1)k+e(k,n) and the nonzero λFp is

λ={k-12=-1if k=p-1,k+12if k<p-1.

Proof.

The proof follows the argument of Lemma 3.2. However, as k is small, we have to be concerned with more terms than in (2.6) and in (2.7). We extend the dimension of the matrices to e(k,n)+e(k,n)+12k+1. Then we can use Lemma 2.5 and it remains to show that the top right-hand e(k,n)+1×e(k,n)+1 block of Πp-1,n,e(k,n)+e(k,n)+1 is Tn[λ].

The refinement of (2.6) (using n2k) for uh+1 reads as follows:

(3.6)

[uh,b]x+((h-2)k+n)β(x)θ(x)xhk+n+1
+(β(x)θ(x)-β(x)θ(x))xhk+n+2
+(((h-1)k+n+12)-(k+1)((h-2)k+n))
×β(x)2θ(x)x(h+1)k+n+1
+(((h-2)k+n)β(x)2θ(x)
+((3-h)k-n+1)β(x)β(x)θ(x))x(h+1)k+n+2
+(12β(x)2θ′′(x)+β(x)2θ(x)
-β(x)β(x)θ(x))x(h+1)k+n+3
+Eh,n(modx(h+2)k+n+2).

Here, for some qh,n,qh,n𝔽p,

Eh,n={qh,nx(h+2)k+n+1β(x)3θ(x)if (h-1)k+n>2k,qh,nx(h+2)k+n+1β(x)3θ(x)+qh,nβ(x)θ(x)2if (h-1)k+n=2k.

The coefficient of xhk+n+j+1 in uh+1 for 0j<k is the same as in (2.7), and for kj<2k it is

(3.7)

i=1k+a((h-2)k+n+2i-j)rk+j-iti
+((nh+a2)+(k+12))rk2ta+i=01c1irk+irk+1-i)ta-1
+(i=02c2irk+irk+2-i)ta-2++(i=0acairk+irk+a-i)t0,

where a=j-k and csw𝔽p for 1sa, 0wa, and ws. For j=2k there will be an extra term in the coefficient of t0 and the rest will satisfy the above formula. But its value is irrelevant for our proof.

We again get that the (i,j) entry of Πh is 0 unless j-i{0,k,2k} as in Lemma 3.2.

As before, we assume first that n2k(modp) and pick some mk(modp) and hence e(k,n)=0 and e(k,m)=k. As a subcase assume that k<p-1 so b=x+xk+1. Now from (3.7) we compute the entries of the matrix Ah,m,2k+1:

ahij={mh+iif i=j,(mh+i2)+(k+12)if j=i+k,Ch,mif (i,j)=(0,2k),0otherwise.

The value of Ch,m𝔽p will turn out to be irrelevant. By induction on h1, we get that the (i,i+k) entry of Πh is

(3.8)

(t=1h(m1+i)(m2+i)(mt+i)(mt+2+i)(mt+3+i)
(mh+1+i)mt+i-12)
+(k+12)t=1h(m1+i)(m2+i)
(mt-1+i)(mt+2+i)(mh+1+i).

Consider h=p-1. For i=0 the first sum is 0, because m1=0 is a factor in each product. For i>0 pick the unique l{2,,p} such that ml+i=0 and observe that the first sum has a single nonzero term, that for t=l-1. Its value is k+12. For i=0,k the second sum has a single nonzero summand, otherwise two. For i=0 the value is -k-12, for i=k it is k+12 and for i0,k the two terms add up to 0. So finally

(3.9)πi,i+k={-k-12if i=0,k+1if i=k,k+12otherwise.

For the diagonal entries we again have

πi,i=h=1p-1(mh+i){0,-1},

with πi,i=-1 if and only if 0m0+i=m-2k+i-k+i(modp). Note that Πp-2 has (0,0) entry 0 since m1=0 and (0,k) entry k+12k by (3.8). Also Ap-1 has (k,2k) entry k(k+1) and (2k,2k) entry 0. So

π0,2k=(k+1)22.

Put λ=k+120 as k<p-1. Now we have the claimed matrix

Tm[λ]=(-λ002λ20λ00000λ0-1002λ).

As (3.1) is still valid, we also have

Tn[λ]=(-1002λ00000000).

Now suppose that k=p-1 so take b=x+xp+x2p-1. We continue to assume that n2k(modp) and mk(modp) and hence e(k,n)=0,e(k,m)=k. By using (3.6), (3.7) we have the corresponding matrix Ah,m,2k+1 whose (i,j) entries for 0i,j2k are

ahij={mh+iif i=j,mh+i-k+(mh+i2)+(k+12)if j=k+i,Dh,mif (i,j)=(0,2k),0otherwise.

The value of Dh,m𝔽p is uninteresting, as before. Again the (i,j) entry of Πh,m,2k+1 is zero if j-i{0,k,2k}. By the value of ahii+k we get that the (i,i+k) entry of Πh is the sum of (3.8) and (3.2), so

πi,i+k={-k-12+1=-k-12if i=0,k+1-2=k-1if i=k,k+12-1=k-12otherwise.

For the diagonal entries we again have

πi,i=h=1p-1(mh+i){0,-1},

with πi,i=-1 if and only if 0m0+i=m-2k+i-k+i(modp). Note that Πp-2 has (0,0) entry 0 and (0,k) entry k+12k+-1k=k-12k. On the other hand, Ap-1 has (k,2k) entry -2k+k(k+1)=k(k-1) and (2k,2k) entry 0. So

π0,2k=(k-1)22.

Put λ=k-12. (We use the assumption p>2 to claim λ0.) We obtained the claimed matrix

Tm[λ]=(-λ002λ20λ00000λ0-1002λ).

We get Tn[λ] from this exactly as above.

Assume finally that k+k0+pt<n<k+p(t+1) that is n2k-i(modp) for any 0ik0, so e(k,n)=k0. Of course k<p-1. So we take the element b=x+xk+1 and obtain the corresponding matrix Tn[λ]=(k+12) (from (3.9)) exactly as in Lemma 3.2. ∎

We collect in the following corollary what we need for the proof of Theorem 1.2.

Corollary 3.4.

Let pk and k>1 if p=2. In addition, let k+k0+ptnk+p(t+1) for some nonnegative integer t. Define

(3.10)

b=x+xk+1+xk+k0+1,λ=-1if kp-1,
b=x+xk+1,λ=k+12if k<p-1.

Then for every uNn and for s1 we have [u,s(p-1)b]Nn* and

μn*([u,s(p-1)b])=λs-1μn(u)Tn[λ],

where n*=n+s(p-1)k+(s-1)k0+e(k,n).

Proof.

Suppose that k+k0+pt<n<k+(t+1), for some integer t. Now e(k,n)=0. Recall the matrix Tn[λ]=(λ), where λ is as above, exactly as in Lemmas 3.2 and 3.3. If u𝒩n is such that μn(u)=(α), then by these lemmas,

μn*([u,s(p-1)b])=(λsα)=λs-1μn(u)Tn[λ],

as claimed.

Suppose now that n=k+k0+pt or n=k+p(t+1) for some nonnegative integer t. Hence e(k,n)=k0. Note that n+(p-1)k+e(k,n)k(modp) in both cases, so from the second step there will be no difference between the two cases. Recall the matrices X=Tk[λ],Y=T2k[λ]Mk0+1×k0+1(𝔽p), where λ is as above, exactly as in Lemmas 3.2 and 3.3. Note that YX=λY and X2=λX.

If n=k+p(t+1), then these lemmas imply by induction on s that

μn*([u,s(p-1)b])=μn(u)Xs=λs-1μn(u)Tn[λ],

as claimed.

If n=k+k0+pt, then these lemmas imply first that

μn+(p-1)+e(k,n)([u,p-1b])=μn(u)Y.

Then (using e(k,k)=k0) we obtain by induction on s that

μn*([u,s(p-1)b])=μn(u)YXs-1=λs-1μn(u)Tn[λ],

as claimed. ∎

Proof of Theorem 1.2.

Suppose that pk. By the remark before Lemma 3.1 we may assume that k+k0+ptn<k+(t+1)p, for some integer t0. If n=k+k0+pt, then e(k,n)=0, otherwise e(k,n)=k0.

Let g-1=b=x+xk+1+xk+k0+1, or x+xk+1 as in Lemmas 3.2 and 3.3. We pick some u=u0𝒩n of depth n, its leading coefficient, α0 will be chosen appropriately later. Set f-1=g-1u and then um=gpmf-pm.

Let nm=n+(pm-1)k+pm-pp-1k0+e(k,n). We prove that D(um)=nm by induction on m. For that, assume m1 and that for every smaller index the theorem holds. Note that

nm=nm-1+(p-1)pm-1k+pm-1k0.

If n2k(modp), then for the inductive proof to work we will need information on the subsequent k0=e(k,nm)=e(k,n) coefficients of um.

To this end, we recall some facts about formal basic commutators for two generators, a,b. These are defined inductively and are totally ordered. First, a and b are basic commutators of weight 1 and a>b. Second, if x>y are basic commutators of weight p and q, then [x,y] is called a basic commutator of weight p+q provided that if x=[z,w], then yw. The total ordering is arbitrary among basic commutators of equal weights but if x has weight larger than y, then x>y.

Every basic commutator has naturally defined weights, Wa and Wb in the variables: the number of times that variable occurs in the commutator. More about the formal basic commutators and relevant descriptions of weight can be found in [4, pp. 9–12]. The following finer approximation is from there [4, Proposition 1.1.32 (i)]:

umum-1p[um-1,g-pm-1](p2)[um-1,2g-pm-1](p3)
[um-1,p-1g-pm-1](modK(g-pm-1,um-1)),

where K(g-pm-1,um-1) is the normal closure in 𝒩 of the set of all formal basic commutators in {g-pm-1,um-1} of weight at least p and of weight at least 2 in um-1 and also the p-th powers of all basic commutators of weight at least 3 and at most p-1 and of weight at least 2 in um-1. The only weight p commutator with weight 2 in um-1 is w=[[um-1,p-2g-pm-1],um-1]. We denote its multiplicity by δ0. (The exact value of δ0 is p2-p-1, but it is irrelevant for our proof.) Now our goal is to show we can approximate um by the commutator [um-1,p-1g-pm-1].

If m=1 and nm-1=n=k+k0, then e(k,n)=0 and e(k,n)=k0 so

D([u,p-1g-1])=n+(p-1)k=n+(p-1)k+e(k,n),

by (2.2). The element w in K(g-1,u) has depth exactly

2n+(p-2)k=n+(p-1)k+e(k,n)+e(k,n),

every other element in K(g-1,u) has larger depth. The leading coefficient of w is

(n-k)αn(n+k)(n+(p-4)k)α(p-2)k=-2α2,

where α is the leading coefficient of u. If m1 is arbitrary with nm-1>k+k0, then every element in K(g-pm-1,um-1) has depth at least

2nm-1+(p-2)k>nm-1+(p-1)k+e(k,nm-1)+e(k,nm-1).

From now let m1 arbitrary. Also note that

D([um-1,ig-pm-1](pi+1))p(nm-1+k)>nm-1+(p-1)k+e(k,nm-1)+e(k,nm-1)

for 1ip-2 for every nm-1k+k0. Let nm-1¯ denote the remainder of nm-1 modulo p. Then

D(um-1p)pnm-1+nm-1¯=nm-1+(p-1)k+k0+(p-2)k0+(p-1)(nm-1-k-k0)+nm-1¯>nm-1+(p-1)k+e(k,nm-1)+e(k,nm-1).

We obtain a first version for the coefficient vector

(3.11)

μnm(um)
={μnm([um-1,p-1g-pm-1])if nm-1>k+k0,μn1([u,p-1g-1])+(0,,0,-2δ0α2)if m=1,n=k+k0.

Here δ0 is the exponent of w in u1[u,p-1g-1]wδ0(modxn1+e(k,n)+1) in the second case. Consider m>1. By the last line of Lemma 2.4 (which holds in our case, as p-1=1 would imply p=2),

[um-1,p-1g-pm-1][um-1,(p-1)pm-1g-1](modx2nm-1+(p-1)pm-1k+2).

Now

2nm-1+(p-1)pm-1k+2=nm+nm-1-pm-1k0+2=nm+n+(pm-1-1)k+(-pm-1+pm-1-pp-1)k0+e(k,n)+2>nm+e(k,nm)+1.

So we obtain μnm(um)=μnm([um-1,(p-1)pm-1g-1]) for m>1. By induction and by (3.11), we get a second version for the coefficient vector

(3.12)μnm(um)={μnm([u1,pm-pg-1])for m>1,μnm([u,pm-1g-1])for m1 if n>k+k0.

After these preliminary observations we turn to the proof.

First let k+k0+pt<n<k+(t+1)p. Then e(k,n)=0 and e(k,n)=k0, and so nmnk,2k(modp) for every m0. By Corollary 3.4 and (3.12),

D(um)=D([um-1,p-1g-pm-1])=D([u,pm-1g-1])=nm,

as required.

Now suppose that n=k+k0+tp. Then e(k,n)=k0 and e(k,n)=0, and so nmk(modp) for every m1. Recall that α0 denotes the leading coefficient of u𝒩n. By the definition of g-1=b and λ in (3.10), Corollary 3.4 shows that for the first k0 coefficients of v=[u,p-1g-1] we have

μn1(v)=μn(u)Tn[λ]=(-α,0,,0,2λα).

By (3.11),

μn1(gpf-p)=(-α,0,,0,2λα-2δα2),

where δ=δ0 if n=k+k0 and δ=0 otherwise. The leading coefficient is -α0, so

D(u1)=D(gpf-p)=n+(p-1)k+e(k,n)=n1

holds. Now n1k(modp) so e(k,n1)=e(k,n1)=k0.

We apply Corollary 3.4 again to get

D(u2)=D([u1,p-1g-p])=D([u1,(p-1)pg-1])=n2.

We also obtain

(3.13)μn2(gp2f-p2)=(-λα+2δα2,0,,0,2λ2α-4λδα2).

The leading coefficient is nonzero if α0 and 2αδλ. Such an α exists because λ0. Fix this α. The vector at (3.13) is an eigenvector of Tnm[λ] with eigenvalue λ0 for all m1.

By Corollary 3.4 and (3.12), we obtain

D(um)=D([um-1,p-1g-pm-1])=D([u1,pm-pg-1]=nm,

as required.

Now it remains to consider the case of pk. Our arguments are similar to those of [3]. Suppose that pn. Let g,u be arbitrary and set f-1:=g-1u. Then

D(gpf-p)=D([u,p-1g-1])=n+(p-1)kn(modp)

since 2n+(p-2)k>n+(p-1)k. We can proceed with induction to get

D(gpmf-pm)=D([u,pm-1(p-1)g-1])=n+(pm-1)k

for m1. Suppose finally that pn. If n=k, then let g be arbitrary of depth k and f=gp. Then

gpmfp-m=gpm(1-p)

and so

D(gpmfp-m)=pmk

by [2, Lemma 1]. If n>k, then pick g and f to work for the pair k,n+1. Then e(k,n)=1=e(k,n+1)+1 so we have the result by Lemma 3.1. ∎

Proof of Corollary 1.4.

Assume the contrary: for a particular m1 we have

(p-1)D(fm)>(pm+1-pm)k+(pm-1)k0.

By Theorem 1.2, we have

D(gpmf-pm)=n1=n+(pm-1)k+pm-pp-1k0+e(k,n),

where n1n(modp) if n2k-i(modp) for any 0ik0, and otherwise n1k0(modp). Therefore, in any case e(k,n1)=k0. Now apply Theorem 1.1 for gpm,fpm to get

D((gpm)p(fpm)-p)n1+(p-1)D(fpm)+k0
>n+(pm-1)k+pm-pp-1k0+e(k,n)
+(pm+1-pm)k+(pm-1)k0+k0
=n+(pm+1-1)k+pm+1-pp-1k0+e(k,n)
=D(gpm+1f-pm+1).

Here the last equation is the statement of Theorem 1.2 for m+1. The strict inequality is a contradiction, so (1.1) holds.

Now suppose that n>k and hence n-k+e(k,n)-k0>0. Since Theorem 1.2 holds for f,g, we have

D(g-pmfpm)=n+(pm-1)k+pm-pp-1k0+e(k,n)
=n-k+e(k,n)-k0+pmk+pm-1p-1k0
>pmk+pm-1p-1k0
=D(fpm),

by the first part. Therefore,

D(gpm)=D(fpm)=pmk+pm-1p-1k0,

as claimed. ∎


Communicated by Evgenii I. Khukhro


Funding statement: The research of the second author was partly supported by National Research, Development and Innovation Office – NKFIH (K 115799).

References

[1] T. Aslan, The distance of large powers in the Nottingham group for p=2, preprint. Search in Google Scholar

[2] R. D. Camina, The Nottingham group, New Horizons in Pro-p Groups, Progr. Math. 184, Birkhäuser, Boston (2000), 205–221. 10.1007/978-1-4612-1380-2_6Search in Google Scholar

[3] K. Keating, How close are p-th powers in the Nottingham group?, J. Algebra 287 (2005), no. 2, 294–309. 10.1016/j.jalgebra.2005.02.011Search in Google Scholar

[4] C. R. Leadham-Green and S. McKay, The Structure of Groups of Prime Power Order, Oxford University Press, Oxford, 2002. Search in Google Scholar

Received: 2018-06-21
Revised: 2018-09-18
Published Online: 2018-10-30
Published in Print: 2019-05-01

© 2019 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 1.5.2024 from https://www.degruyter.com/document/doi/10.1515/jgth-2018-0133/html
Scroll to top button