This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Asymptotic and Assouad-Nagata dimension of finitely generated groups and their subgroups

Levi Sledd
Abstract

We prove that for all k,m,n{}k,m,n\in\mathbb{N}\cup\{\infty\} with 4kmn4\leq k\leq m\leq n, there exists a finitely generated group GG with a finitely generated subgroup HH such that asdim(G)=k\operatorname{asdim}(G)=k, asdimAN(G)=m\operatorname{asdim_{\textnormal{AN}}}(G)=m, and asdimAN(H)=n\operatorname{asdim_{\textnormal{AN}}}(H)=n. This simultaneously answers two open questions in asymptotic dimension theory.

0 Introduction

This is the second in a series of two papers on asymptotic dimension and Assouad-Nagata dimension of finitely generated groups: the first is [Sledd]. Asymptotic dimension (asdim\operatorname{asdim}) and asymptotic Assouad-Nagata dimension (asdimAN\operatorname{asdim_{\textnormal{AN}}}) are two distinct but related ways of defining the large-scale dimension of a metric space. Each is invariant under quasi-isometry, and thus can be considered as an invariant of finitely generated groups. For countable groups with proper left-invariant metrics, asymptotic Assouad-Nagata dimension is equivalent to Assouad-Nagata dimension (dimAN\operatorname{dim_{\textnormal{AN}}}). So from now on we use this shorter term when talking about groups, although we continue to denote it by asdimAN\operatorname{asdim_{\textnormal{AN}}}.

Given a way of defining ‘dimension’ for an algebraic structure, it is natural to ask whether it is monotonic with respect to substructures: that is, whether ABA\leqslant B implies that the dimension of AA is no greater than the dimension of BB. Is our dimension like that of a vector space, where this natural monotonicity holds, or is it like the rank of a free group, where it fails spectacularly? Since asdim\operatorname{asdim} is actually a coarse invariant, it follows that asdim\operatorname{asdim} is well defined for all countable groups, and if GG is a countable group and HGH\leqslant G, then asdim(H)asdim(G)\operatorname{asdim}(H)\leq\operatorname{asdim}(G). In this paper we show that Assouad-Nagata dimension behaves quite differently. Namely, we prove the following theorem.

Theorem 1.

For any k,m,n{}k,m,n\in\mathbb{N}\cup\{\infty\} with 4kmn4\leq k\leq m\leq n, there exist finitely generated, recursively presented groups GG and HH with HGH\leqslant G, such that

asdim(G)\displaystyle\operatorname{asdim}(G) =k\displaystyle=k
asdimAN(G)\displaystyle\operatorname{asdim_{\textnormal{AN}}}(G) =m\displaystyle=m
asdimAN(H)\displaystyle\operatorname{asdim_{\textnormal{AN}}}(H) =n.\displaystyle=n\,.

In [Higes2], Higes constructs an infinitely generated, locally finite abelian group, and a proper left-invariant metric with respect to which the group has asymptotic dimension 0 but infinite Assouad-Nagata dimension. In [Brodskiy_Dydak_Lang], Brodskiy, Dydak, and Lang construct finitely generated groups with a similar gap, showing that (for example) 22\mathbb{Z}_{2}\wr\mathbb{Z}^{2} has asymptotic dimension 2 but infinite Assouad-Nagata dimension. Previously it was not known whether a finitely generated group GG could satisfy asdim(G)<asdimAN(G)<\operatorname{asdim}(G)<\operatorname{asdim}_{AN}(G)<\infty (Question (2) of [Higes2]), nor was it known whether a finitely generated group could contain a finitely generated subgroup of greater Assouad-Nagata dimension (Questions 8.6 and 8.7 of [Brodskiy_etal]). With 1, we show that both these things are possible.

If HGH\leqslant G but asdimAN(H)>asdimAN(G)\operatorname{asdim_{\textnormal{AN}}}(H)>\operatorname{asdim_{\textnormal{AN}}}(G), it must be that HH is distorted in GG, and that this distortion collapses HH to a space of lesser Assouad-Nagata dimension in GG. However, distortion does not always affect the Assouad-Nagata dimension of the distorted subgroup. For example, in BS(1,2)=a,bb1aba2BS(1,2)=\langle a,b\mid b^{-1}aba^{-2}\rangle, the subgroup a\langle a\rangle is distorted, but still has Assouad-Nagata dimension 11. We call distortion which affects Assouad-Nagata dimension Assouad-Nagata dimension distortion. The author hopes that Assouad-Nagata dimension distortion will be an interesting phenomenon to study in its own right, and that more examples can be found in nature.

The paper is organized as follows. In Section 1, we fix countable group KK, constructed as a direct sum of cyclic groups of increasing order. We then show that for each m,n{}m,n\in\mathbb{N}\cup\{\infty\} with m<nm<n, there are two different proper left-invariant metrics on KK such that asdimAN(K)=m\operatorname{asdim_{\textnormal{AN}}}(K)=m with respect to one, and asdimAN(K)=n\operatorname{asdim_{\textnormal{AN}}}(K)=n with respect to the other.

In Section 2, we use techniques from small cancellation theory to establish a highly technical lemma. This lemma allows us to quasi-isometrically embed KK, with respect to each proper left-invariant metric, into a finitely generated group.

In Section 3, we embed KK into finitely generated groups AA and BB. This is done is such a way that, calling KAK_{A} the copy of KK in AA and KBK_{B} the copy of KK in BB, we have that asdimAN(KA)=m\operatorname{asdim_{\textnormal{AN}}}(K_{A})=m and asdimAN(KB)=n\operatorname{asdim_{\textnormal{AN}}}(K_{B})=n. We then identify the two with an isomorphism ϕ:KAKB\phi:K_{A}\to K_{B}, and let G=AϕBG=A*_{\phi}B. Our technical small cancellation lemma comes back to help us a second time by showing that ϕ\phi ‘crushes’ the image of KBK_{B} in GG to the size of KAK_{A}. With a few calculations using well-known extension theorems for Assouad-Nagata dimension, we are able to prove the following.

Proposition 1.

For any m,n{}m,n\in\mathbb{N}\cup\{\infty\} with m<nm<n, there exists a group G=AϕBG=A*_{\phi}B where GG, AA, and BB are finitely generated and recursively presented, such that

1\displaystyle 1 asdim(G)2\displaystyle\leq\operatorname{asdim}(G)\leq 2
m+1\displaystyle m+1 asdimAN(G)m+2\displaystyle\leq\operatorname{asdim_{\textnormal{AN}}}(G)\leq m+2
n+1\displaystyle n+1 asdimAN(B)n+2.\displaystyle\leq\operatorname{asdim_{\textnormal{AN}}}(B)\leq n+2\,.

Using the free product formulas for asymptotic and Assouad-Nagata dimension and the Morita theorem for Assouad-Nagata dimension, it is then easy to derive 1 from 1.

There are many technical restrictions placed on the presentations of AA and BB from 1. In Section 4 we give explicit presentations where these conditions are satisfied. Curiously, although we are able to give an explicit presentation of a group GG satisfying the conditions of 1, we are not quite able to do the same for 1. However, we can explicitly give presentations of two groups, one of which must be a group satisfying the conclusion of 1.

1 Adapting a construction of Higes

We refer the reader to [Sledd], Section 1 for basic conventions and notation regarding metric spaces, as well as definitions of the terms asymptotic dimension (asdim\operatorname{asdim}), asymptotic Assouad-Nagata dimension (asdimAN\operatorname{asdim_{\textnormal{AN}}}), and control function. We assume that the reader is familiar with the notions of quasi-isometry and bi-Lipschitz equivalence. Occasionally we will mention terms such as ‘coarse’ map/embedding/equivalence. Since we will not need to work with these directly, we do not give a definition here, but one may be found in any text on coarse geometry, for example [Roe, pp. 9]. What matters to us is that, if XX and YY are metric spaces which are coarsely equivalent, then asdim(X)=asdim(Y)\operatorname{asdim}(X)=\operatorname{asdim}(Y), and that a quasi-isometry or bi-Lipschitz map is a special case of a coarse equivalence.

In this paper, we adopt the convention that the Cartesian product of two metric spaces is always endowed with the 1\ell^{1} product metric. That is, if XX and YY are metric spaces, then X×YX\times Y is equipped with the metric defined by

dX×Y((x,y),(x,y))=dX(x,x)+dY(y,y)d_{X\times Y}((x,y),(x^{\prime},y^{\prime}))=d_{X}(x,x^{\prime})+d_{Y}(y,y^{\prime})

for all x,xXx,x^{\prime}\in X and y,yYy,y^{\prime}\in Y. With this convention in mind, if \sim stands for either “is coarsely equivalent to,” “is quasi-isometric to,” or “is bi-Lipschitz equivalent to,” then we have that XXX\sim X^{\prime} and YYY\sim Y^{\prime} implies X×YX×YX\times Y\sim X^{\prime}\times Y^{\prime}. In addition, asdim\operatorname{asdim} and asdimAN\operatorname{asdim_{\textnormal{AN}}} are subadditive with respect to taking direct products, in a sense that is made precise by the following two theorems. We will use them often throughout this paper.

Lemma 1.1.

[Bell_Dranishnikov, Brodskiy_etal] Let X,YX,Y be metric spaces. Then

asdim(X×Y)\displaystyle\operatorname{asdim}(X\times Y) asdim(X)+asdim(Y)\displaystyle\leq\operatorname{asdim}(X)+\operatorname{asdim}(Y)
asdimAN(X×Y)\displaystyle\operatorname{asdim_{\textnormal{AN}}}(X\times Y) asdimAN(X)+asdimAN(Y).\displaystyle\leq\operatorname{asdim_{\textnormal{AN}}}(X)+\operatorname{asdim_{\textnormal{AN}}}(Y).

1.1 Normed groups

We denote the identity element of an arbitrary group by 1, and of an abelian group by 0. Let GG be a group. A norm on GG is a function :G0+\|\cdot\|:G\to\mathbb{R}_{0}^{+} such that, for all g,hGg,h\in G,

  • g=0\|g\|=0 if and only if g=1g=1.

  • g=g1\|g\|=\|g^{-1}\|.

  • ghg+h\|gh\|\leq\|g\|+\|h\|.

Some authors call this a length function or weight function on GG.

A norm is proper if {gGgN}\{g\in G\mid\|g\|\leq N\} is finite for all N0N\geq 0. There is a natural one-to-one correspondence between norms and left-invariant metrics, given by d(g,h)=g1hd(g,h)=\|g^{-1}h\| and g=d(1,g)\|g\|=d(1,g), and a left-invariant metric on a group is proper if and only if the corresponding norm is proper. Every countable group admits a proper norm, and any two proper norms on the same countable group are coarsely equivalent [Dranishnikov_Smith2, Proposition 1.1]. Thus asdim\operatorname{asdim} is an invariant of countable groups: in particular, if GG is a countable group and HGH\leqslant G, then asdim(H)asdim(G)\operatorname{asdim}(H)\leq\operatorname{asdim}(G). It is easy to show that a countable group has asymptotic dimension zero if and only if it is locally finite, a fact which we will use many times.

Formally, a normed group should be an ordered pair (G,G)(G,\|\cdot\|_{G}). But from now on, whenever we say that GG is a normed group, it is understood that GG is equipped with a norm, which is always called G\|\cdot\|_{G}. With this convention in mind we eliminate the norm from the notation wherever possible.

If GG is a normed group and ss is a positive real number, then the function sG:G0+,gsgGs\|\cdot\|_{G}:G\to\mathbb{R}_{0}^{+},g\mapsto s\|g\|_{G} is also a norm on GG. We call the normed group (G,sG)(G,s\|\cdot\|_{G}) a scaled normed group, and denote it briefly by sGsG.

Given two normed groups G0G_{0} and G1G_{1} and scaling constants s0,s1s_{0},s_{1}, we define their scaled direct product s0G0×s1G1s_{0}G_{0}\times s_{1}G_{1} to be the group G0×G1G_{0}\times G_{1} endowed with the norm (s0,s1)\|\cdot\|_{(s_{0},s_{1})} defined by

(g0,g1)(s0,s1)=s0g0G0+s1g1G1\|(g_{0},g_{1})\|_{(s_{0},s_{1})}=s_{0}\|g_{0}\|_{G_{0}}+s_{1}\|g_{1}\|_{G_{1}}

for all g0G0g_{0}\in G_{0} and g1G1g_{1}\in G_{1}. This is just the 1\ell^{1} product norm on s0G0×s1G1s_{0}G_{0}\times s_{1}G_{1}. For any kk\in\mathbb{N}, we define the scaled direct product of finitely many scaled normed groups i=0ksiGi\prod_{i=0}^{k}s_{i}G_{i} by iterating this construction. Note that for finite direct products we have that i=0ksiGi\prod_{i=0}^{k}s_{i}G_{i} is bi-Lipschitz equivalent to i=0kGi\prod_{i=0}^{k}G_{i} without scaling.

To avoid frequently having to state that certain sets are nonempty, we declare iGi\prod_{i\in\emptyset}G_{i} to be the trivial group. Let II be a set and (Gi)iI(G_{i})_{i\in I} an II-tuple of groups. For g=(gi)iIiIGig=(g_{i})_{i\in I}\in\prod_{i\in I}G_{i}, we denote the support of gg by supp(g)\operatorname{supp}(g); that is, supp(g)={iIgi1}\operatorname{supp}(g)=\{i\in I\mid g_{i}\neq 1\}. By definition iIGi\bigoplus_{i\in I}G_{i} is the subgroup of iIGi\prod_{i\in I}G_{i} consisting of all giIGig\in\prod_{i\in I}G_{i} such that supp(g)\operatorname{supp}(g) is finite. The notion of scaled direct products can then be extended to general direct sums in the following natural way.

Definition 1.2.

Let II be a set, let (Gi)iI(G_{i})_{i\in I} be an II-tuple of normed groups, and let s=(si)iIs=(s_{i})_{i\in I} an II-tuple of scaling constants. Let G=iIGiG=\bigoplus_{i\in I}G_{i}. Then the scaled direct sum iIsiGi\bigoplus_{i\in I}s_{i}G_{i} is defined to be the normed group (G,s)(G,\|\cdot\|_{s}), where s\|\cdot\|_{s} is given by

gs=iIsigiGi\|g\|_{s}=\sum_{i\in I}s_{i}\|g_{i}\|_{G_{i}}

for all gGg\in G. We call s\|\cdot\|_{s} the norm induced by ss.

Lemma 1.3.

Let II be a set, s=(si)iIs=(s_{i})_{i\in I} an II-tuple of scaling constants bounded away from zero. Then iIsiGi\bigoplus_{i\in I}s_{i}G_{i} is bi-Lipschitz equivalent to iIsiGi\bigoplus_{i\in I}s_{i}^{\prime}G_{i}, where sis_{i}^{\prime} is a positive integer for all iIi\in I.

Proof.

Suppose that ε>0\varepsilon>0 is such that siεs_{i}\geq\varepsilon for all iIi\in I. Let s=(si)iI=(si)iIs^{\prime}=(s_{i}^{\prime})_{i\in I}=(\lceil s_{i}\rceil)_{i\in I}, and let g=(gi)iIiIGig=(g_{i})_{i\in I}\in\bigoplus_{i\in I}G_{i}. Then clearly gsgs\|g\|_{s}\leq\|g\|_{s^{\prime}}, and

gs=iIsigiGiiI(si+1si)sigiGi(1+1ε)iIsigiGi=(1+1ε)gs.\|g\|_{s^{\prime}}=\sum_{i\in I}\lceil s_{i}\rceil\|g_{i}\|_{G_{i}}\leq\sum_{i\in I}\left(\tfrac{s_{i}+1}{s_{i}}\right)s_{i}\|g_{i}\|_{G_{i}}\leq\left(1+\tfrac{1}{\varepsilon}\right)\sum_{i\in I}s_{i}\|g_{i}\|_{G_{i}}=\left(1+\tfrac{1}{\varepsilon}\right)\|g\|_{s}\,.

1.2 A fixed group with varying norms

The next set of lemmas deal specifically with direct sums of cyclic groups. Here we assume that a cyclic group comes equipped with the natural norm, that is x=min(x,x)\|x\|_{\mathbb{Z}_{\ell}}=\min(x,\ell-x) for all xx\in\mathbb{Z}_{\ell}, and x=|x|\|x\|_{\mathbb{Z}}=|x| for all xx\in\mathbb{Z}. Unless otherwise noted, tuples are sequences indexed by \mathbb{N}, e.g. (si)(s_{i}) stands for (si)i(s_{i})_{i\in\mathbb{N}}.

Definition 1.4.

Let (xi)ii(x_{i})\in\bigoplus_{i\in\mathbb{N}}\mathbb{Z}_{\ell_{i}}. The geodesic form of (xi)(x_{i}) is the unique sequence of integers (yi)(y_{i}) such that for all ii\in\mathbb{N},

  • yiximodi, andy_{i}\equiv x_{i}\mod\ell_{i}\,\text{, and}

  • yi{i12,,1,0,1,,i2}.y_{i}\in\left\{-\left\lfloor\tfrac{\ell_{i}-1}{2}\right\rfloor,\ldots,-1,0,1,\ldots,\left\lfloor\tfrac{\ell_{i}}{2}\right\rfloor\right\}_{\,.}

Note that if s=(si)s=(s_{i}) is a sequence of scaling constants, x=(xi)isiix=(x_{i})\in\bigoplus_{i\in\mathbb{N}}s_{i}\mathbb{Z}_{\ell_{i}}, and (yi)(y_{i}) is the geodesic form of xx, then we have

xs=isi|yi|.\|x\|_{s}=\sum_{i\in\mathbb{N}}s_{i}|y_{i}|\,. (1)
Definition 1.5.

For s+s\in\mathbb{R}^{+} and k,nk,n\in\mathbb{N}, assume that the set

s{0,,k}n={0,s,2s,,ks}nns\{0,\ldots,k\}^{n}=\{0,s,2s,\ldots,ks\}^{n}\subset\mathbb{R}^{n}

is equipped with the 1\ell^{1} metric. Then an expanded nn-dimensional cube is a space isometric to s{0,,k}ns\{0,\ldots,k\}^{n} for some s1s\geq 1 and kk\in\mathbb{N}.

In accordance with Definition 1.5, whenever ss is a scaling constant and s1s\geq 1, we call ss an expansion constant. Sequences of expanded cubes are useful for establishing lower bounds on the asymptotic Assouad-Nagata dimension of a metric space.

Lemma 1.6.

[Higes, Corollary 2.7] Let XX be a metric space, nn\in\mathbb{N}. If XX contains a sequence of expanded nn-dimensional cubes sj{0,,kj}ns_{j}\{0,\ldots,k_{j}\}^{n} where limjkj=\lim_{j\to\infty}k_{j}=\infty, then asdimAN(X)n\operatorname{asdim_{\textnormal{AN}}}(X)\geq n.

Suppose that PP is a set with |P|n|P|\geq n, (i)iP(\ell_{i})_{i\in P} is a PP-tuple of natural numbers, and sPs_{P} is an expansion constant. Let kPk_{P} be a natural number with kPmin{i/2iP}k_{P}\leq\min\{\ell_{i}/2\mid i\in P\}. Then by (1), sPiPis_{P}\bigoplus_{i\in P}\mathbb{Z}_{\ell_{i}} contains an expanded nn-dimensional cube sP{0,,kP}ns_{P}\{0,\ldots,k_{P}\}^{n}. With this observation and Lemma 1.6, one can construct a group which can achieve any positive Assouad-Nagata dimension. The idea is to take a direct sum of cyclic groups, block every nn of them together, and scale the blocks appropriately. In [Higes] Higes uses this idea to construct, for any n+{}n\in\mathbb{Z}^{+}\cup\{\infty\}, a normed group GnG_{n} with asymptotic dimension zero but Assouad-Nagata dimension nn. However, in Higes’ examples, if mnm\neq n, then GmG_{m} and GnG_{n} are not isomorphic. For our purposes, it is important that the group be fixed, with only the norm varying. The rest of this section is devoted to working out the details of this construction. To smooth the process, we introduce the following ad hoc notation.

Definition 1.7.

For each m+{}m\in\mathbb{Z}^{+}\cup\{\infty\}, let 𝒫m={P(m,j)j}\mathcal{P}_{m}=\{P_{(m,j)}\mid j\in\mathbb{N}\} be the partition of \mathbb{N} given by

P(m,j)={{jm,jm+1,,(j+1)m1} if m+{j2,j2+1,,(j+1)21} if m=.P_{(m,j)}=\begin{cases}\{jm,jm+1,\ldots,(j+1)m-1\}&\text{ if }m\in\mathbb{Z}^{+}\\ \{j^{2},j^{2}+1,\ldots,(j+1)^{2}-1\}&\text{ if }m=\infty\,.\end{cases}
Definition 1.8.

Let s=(si)s=(s_{i}) be a sequence, m+{}m\in\mathbb{Z}^{+}\cup\{\infty\}. Let the mm-inflation of ss, denoted m×sm\times s, be the sequence defined by

(m×s)i=sjiP(m,j).(m\times s)_{i}=s_{j}\Leftrightarrow i\in P_{(m,j)}\,.

For example, if s=(1,2,3,)s=(1,2,3,\ldots), then 2×s=(1,1,2,2,3,3,)2\times s=(1,1,2,2,3,3,\ldots) and ×s=(1,2,2,2,3,3,3,3,)\infty\times s=(1,2,2,2,3,3,3,3,\ldots). By definition,

si={(m×s)im if m+(m×s)i2 if m=and(m×s)i={si/m if m+si if m=.s_{i}=\begin{cases}(m\times s)_{im}&\text{ if }m\in\mathbb{Z}^{+}\\ (m\times s)_{i^{2}}&\text{ if }m=\infty\end{cases}\quad\text{and}\quad(m\times s)_{i}=\begin{cases}s_{\lfloor i/m\rfloor}&\text{ if }m\in\mathbb{Z}^{+}\\ s_{\lfloor\sqrt{i}\rfloor}&\text{ if }m=\infty\,.\end{cases} (2)
Lemma 1.9.

Let dd\in\mathbb{N}, and let (c0,,cd1)(c_{0},\ldots,c_{d-1}) be a finite sequence of scaling constants. Let m+{}m\in\mathbb{Z}^{+}\cup\{\infty\} be fixed, let (si)(s_{i}) be an increasing sequence of expansion constants, and let (i)(\ell_{i}) be an increasing sequence of positive integers. Let

Zd\displaystyle Z_{d} =i=0d1ci,\displaystyle=\bigoplus_{i=0}^{d-1}c_{i}\mathbb{Z}\,, Km=i(m×s)ii.\displaystyle K_{m}=\bigoplus_{i\in\mathbb{N}}(m\times s)_{i}\mathbb{Z}_{\ell_{i}}\,.

Then asdimAN(Zd×Km)d+m\operatorname{asdim_{\textnormal{AN}}}(Z_{d}\times K_{m})\geq d+m.

Proof.

By Lemma 1.3, we may assume without loss of generality that all sis_{i} are positive integers. Since finite direct products preserve bi-Lipschitz equivalence, we may also assume that all cic_{i} are equal to 1, so that Zd=dZ_{d}=\mathbb{Z}^{d}.

Now note that

Zd×Km=d×j(sjiP(m,j)i),Z_{d}\times K_{m}=\mathbb{Z}^{d}\times\bigoplus_{j\in\mathbb{N}}\left(s_{j}\bigoplus_{i\in P_{(m,j)}}\mathbb{Z}_{\ell_{i}}\right)_{\,,}

where d×sjiP(m,j)i\mathbb{Z}^{d}\times s_{j}\bigoplus_{i\in P_{(m,j)}}\mathbb{Z}_{\ell_{i}} is an isometrically embedded subgroup for each jj\in\mathbb{N}. Let

kj=min{i/2iP(m,j)}={jm/2 if m+j2/2 if m=.k_{j}=\min\{\lfloor\ell_{i}/2\rfloor\mid i\in P_{(m,j)}\}=\begin{cases}\lfloor\ell_{jm}/2\rfloor&\text{ if }m\in\mathbb{Z}^{+}\\ \lfloor\ell_{j^{2}}/2\rfloor&\text{ if }m=\infty\,.\end{cases}

Then limjkj=\lim_{j\to\infty}k_{j}=\infty.

If m+m\in\mathbb{Z}^{+}, then |P(m,j)|=m|P_{(m,j)}|=m for all jj\in\mathbb{N}. Then since sjs_{j} is an integer, d×sjiP(m,j)i\mathbb{Z}^{d}\times s_{j}\bigoplus_{i\in P_{(m,j)}}\mathbb{Z}_{\ell_{i}} contains the expanded (d+m)(d+m)-dimensional cube sj{0,,kj}d+ms_{j}\{0,\ldots,k_{j}\}^{d+m} for all jj\in\mathbb{N}. Since limjkj=\lim_{j\to\infty}k_{j}=\infty, by Lemma 1.6 we have asdimAN(d×Km)d+m\operatorname{asdim_{\textnormal{AN}}}(\mathbb{Z}^{d}\times K_{m})\geq d+m.

If m=m=\infty, let n+n\in\mathbb{Z}^{+}. Then |P(m,j)|=(j+1)2j2=2j+1n|P_{(m,j)}|=(j+1)^{2}-j^{2}=2j+1\geq n for all jnj\geq n. Therefore sjiP(m,j)is_{j}\bigoplus_{i\in P_{(m,j)}}\mathbb{Z}_{\ell_{i}} contains the expanded nn-dimensional cube sj{0,,kj}ns_{j}\{0,\ldots,k_{j}\}^{n} for all jnj\geq n. Since limjkj=\lim_{j\to\infty}k_{j}=\infty, by Lemma 1.6 we have asdimAN(K)n\operatorname{asdim_{\textnormal{AN}}}(K_{\infty})\geq n. Since n+n\in\mathbb{Z}^{+} was chosen arbitrarily, asdimAN(K)=\operatorname{asdim_{\textnormal{AN}}}(K_{\infty})=\infty, thus asdimAN(Zd×K)=\operatorname{asdim_{\textnormal{AN}}}(Z_{d}\times K_{\infty})=\infty. ∎

Now, in the notation of Lemma 1.9, we wish to impose certain conditions on the sequence (si)(s_{i}) of expansion constants to guarantee that asdimAN(Zd×Km)=d+m\operatorname{asdim_{\textnormal{AN}}}(Z_{d}\times K_{m})=d+m exactly. We will use a lemma of Higes; in order to do so we need to introduce a little notation, and consider a different norm on countable direct sums of scaled normed groups.

Definition 1.10.

Let (Gi)(G_{i}) be a sequence of normed groups and s=(si)s=(s_{i}) a sequence of scaling constants. Let G=iGiG=\bigoplus_{i\in\mathbb{N}}G_{i}. For convenience, let us define the height function h:Gh:G\to\mathbb{N} by

h(g)={0if g=1max(supp(g)) otherwise.h(g)=\begin{cases}0&\text{if }g=1\\ \max(\operatorname{supp}(g))&\text{ otherwise.}\end{cases}

Now define the quasi-ultranorm on GG induced by ss, denoted squ\|\cdot\|_{s}^{\textnormal{qu}}, by

gsqu=shghGh\|g\|_{s}^{\textnormal{qu}}=s_{h}\|g_{h}\|_{G_{h}} (3)

for all g=(gi)Gg=(g_{i})\in G, where h=h(g)h=h(g).

In [Higes], Higes calls the metric associated to this norm the quasi-ultrametric generated by the sequence of metrics (dGi)(d_{G_{i}}), where dGid_{G_{i}} is the metric associated to the scaled norm siGis_{i}\|\cdot\|_{G_{i}} for each ii\in\mathbb{N}. For this reason we call the norm in (3) the quasi-ultranorm on GG induced by ss, and put ‘qu’ in the superscript. The next lemma says that if all GiG_{i} are finite then, under mild assumptions about the growth of the sequence ss, the norms s\|\cdot\|_{s} and squ\|\cdot\|_{s}^{\textnormal{qu}} are, for our purposes, interchangeable.

Lemma 1.11.

Let (Gi)(G_{i}) be a sequence of normed groups and s=(si)s=(s_{i}) a sequence scaling constants. Let G=iGiG=\bigoplus_{i\in\mathbb{N}}G_{i}. Suppose that Gi,Gi,siG_{i},\|\cdot\|_{G_{i}},s_{i} satisfy the following conditions for all ii\in\mathbb{N}:

  • giGi1\|g_{i}\|_{G_{i}}\geq 1 for all giGi{1}g_{i}\in G_{i}\smallsetminus\{1\}.

  • diam(Gi+1)diam(Gi)\operatorname{diam}(G_{i+1})\geq\operatorname{diam}(G_{i}).

  • si+12sidiam(Gi)s_{i+1}\geq 2s_{i}\operatorname{diam}(G_{i}).

Then the norm s\|\cdot\|_{s} and quasi-ultranorm squ\|\cdot\|_{s}^{\textnormal{qu}} induced by ss are bi-Lipschitz equivalent.

Proof.

Clearly gsqugs\|g\|_{s}^{\textnormal{qu}}\leq\|g\|_{s} for all gGg\in G.

We now prove by induction on h(g)h(g) that gs2gsqu\|g\|_{s}\leq 2\|g\|_{s}^{\textnormal{qu}}. This is clear when h(g)=0h(g)=0. Now suppose that h(g)=k1h(g)=k\geq 1. Write gg as gg′′g^{\prime}g^{\prime\prime}, where gj=gjg_{j}^{\prime}=g_{j} exactly when j=kj=k and is equal to 1 otherwise, and h(g′′)=i<kh(g^{\prime\prime})=i<k. Then we have

gs\displaystyle\|g\|_{s} gs+g′′s=gsqu+g′′sgsqu+2g′′squ\displaystyle\leq\|g^{\prime}\|_{s}+\|g^{\prime\prime}\|_{s}=\|g^{\prime}\|_{s}^{\textnormal{qu}}+\|g^{\prime\prime}\|_{s}\leq\|g^{\prime}\|_{s}^{\textnormal{qu}}+2\|g^{\prime\prime}\|_{s}^{\textnormal{qu}}
gsqu+2sidiam(Gi)gsqu+2sk1diam(Gk1)\displaystyle\leq\|g^{\prime}\|_{s}^{\textnormal{qu}}+2s_{i}\operatorname{diam}(G_{i})\leq\|g^{\prime}\|_{s}^{\text{qu}}+2s_{k-1}\operatorname{diam}(G_{k-1})
gsqu+sk2gsqu=2gsqu.\displaystyle\leq\|g^{\prime}\|_{s}^{\textnormal{qu}}+s_{k}\leq 2\|g^{\prime}\|_{s}^{\textnormal{qu}}=2\|g\|_{s}^{\textnormal{qu}}\,.

Lemma 1.12.

[Higes, Proof of Corollary 4.11] Let (i)(\ell_{i}) be an increasing sequence of positive integers with 02\ell_{0}\geq 2. Let mm be a fixed positive integer. Let s=(si)s=(s_{i}) be a sequence of expansion constants such that

si+11+sidiam(im)=1+(mi/2)si.s_{i+1}\geq 1+s_{i}\operatorname{diam}(\mathbb{Z}_{\ell_{i}}^{m})=1+(m\lfloor\ell_{i}/2\rfloor)s_{i}\,.

Let Kmqu=(iim,squ)K_{m}^{\textnormal{qu}}=(\bigoplus_{i\in\mathbb{N}}\mathbb{Z}_{\ell_{i}}^{m},\|\cdot\|_{s}^{\textnormal{qu}}). Then for any kk\in\mathbb{N} we have asdimAN(k×Kmqu)=k+m\operatorname{asdim_{\textnormal{AN}}}(\mathbb{Z}^{k}\times K_{m}^{\textnormal{qu}})=k+m.

We use this lemma in the case k=0,m=1k=0,m=1 to obtain the slightly generalized lemma that we need.

Lemma 1.13.

Let dd\in\mathbb{N}, and let (c0,,cd1)(c_{0},\ldots,c_{d-1}) be a finite sequence of scaling constants. Let (i)(\ell_{i}) be a sequence of positive integers, and let m+{}m\in\mathbb{Z}^{+}\cup\{\infty\} be fixed. Let (sj)(s_{j}) be an increasing sequence of expansion constants such that, if m+m\in\mathbb{Z}^{+}, we have

sj+1((j+1)m)sjs_{j+1}\geq(\ell_{(j+1)m})s_{j}

for all jj\in\mathbb{N}. Now let

Zd=i=0d1ci\displaystyle Z_{d}=\bigoplus_{i=0}^{d-1}c_{i}\mathbb{Z} Km=i(m×s)ii.\displaystyle K_{m}=\bigoplus_{i\in\mathbb{N}}(m\times s)_{i}\mathbb{Z}_{\ell_{i}}\,.

Then asdimAN(Zd×Km)=d+m\operatorname{asdim_{\textnormal{AN}}}(Z_{d}\times K_{m})=d+m.

Proof.

The lower bound is established in Lemma 1.9. For the upper bound, suppose that m+m\in\mathbb{Z}^{+}. Then

Km=r=0m1(jsjjm+r).K_{m}=\bigoplus_{r=0}^{m-1}\left(\bigoplus_{j\in\mathbb{N}}s_{j}\mathbb{Z}_{\ell_{jm+r}}\right)_{\,.}

Since (i)(\ell_{i}) is increasing, for all jj\in\mathbb{N} and r{0,,m1}r\in\{0,\ldots,m-1\} we have that

sj+1((j+1)m)sj(jm+r)sj(2jm+r/2)sj=(2diam(jm+r))sj1+sjdiam(jm+r).s_{j+1}\geq(\ell_{(j+1)m})s_{j}\geq(\ell_{jm+r})s_{j}\geq(2\lfloor\ell_{jm+r}/2\rfloor)s_{j}=(2\operatorname{diam}(\mathbb{Z}_{\ell_{jm+r}}))s_{j}\geq 1+s_{j}\operatorname{diam}(\mathbb{Z}_{\ell_{jm+r}})\,.

Therefore for any fixed r{0,,m1}r\in\{0,\ldots,m-1\}, the sequences (jm+r),(jm+r),(\ell_{jm+r}),(\mathbb{Z}_{\ell_{jm+r}}), and (sj)(s_{j}) together satisfy the assumptions of Lemmas 1.11 and 1.12. Hence for all r{0,,m1}r\in\{0,\ldots,m-1\},

asdimAN(jsjjm+r)=asdimAN(jjm+r,s)=asdimAN(jjm+r,squ)=1.\operatorname{asdim_{\textnormal{AN}}}\left(\bigoplus_{j\in\mathbb{N}}s_{j}\mathbb{Z}_{jm+r}\right)=\operatorname{asdim_{\textnormal{AN}}}\left(\bigoplus_{j\in\mathbb{N}}\mathbb{Z}_{jm+r},\|\cdot\|_{s}\right)=\operatorname{asdim_{\textnormal{AN}}}\left(\bigoplus_{j\in\mathbb{N}}\mathbb{Z}_{jm+r},\|\cdot\|_{s}^{\textnormal{qu}}\right)=1\,.

Thus by Lemma 1.1,

asdimAN(Km)r=0m1asdimAN(jsjjm+r)m,\operatorname{asdim_{\textnormal{AN}}}(K_{m})\leq\sum_{r=0}^{m-1}\operatorname{asdim_{\textnormal{AN}}}\left(\bigoplus_{j\in\mathbb{N}}s_{j}\mathbb{Z}_{jm+r}\right)\leq m\,,

and asdimAN(Zd)=asdimAN(d)=d\operatorname{asdim_{\textnormal{AN}}}(Z_{d})=\operatorname{asdim_{\textnormal{AN}}}(\mathbb{Z}^{d})=d. Therefore asdimAN(Zd×Km)d+m\operatorname{asdim_{\textnormal{AN}}}(Z_{d}\times K_{m})\leq d+m. ∎

The importance of Lemma 1.13 lies in the fact that if (i)(\ell_{i}) is fixed and m,n+{}m,n\in\mathbb{Z}^{+}\cup\{\infty\} are distinct, then KmK_{m} and KnK_{n} are merely the same group with different norms. Later, we will construct two finitely generated groups AA and BB with subgroups that are isomorphic and bi-Lipschitz equivalent to KmK_{m} and KnK_{n}, respectively. Since KmK_{m} and KnK_{n} are isomorphic, we construct a finitely generated group GG which is the amalgamated product of AA and BB along an isomorphism between KmK_{m} and KnK_{n}. The isomorphism ‘collapses’ KnK_{n}, so that the Assouad-Nagata dimension of GG is not much more than mm, while the Assouad-Nagata dimension of BB is at least nn. To construct A,BA,B, and GG such that all of the aforementioned geometric properties hold, we use some small cancellation theory. This is the topic of the next section.

2 van Kampen diagrams and the C(λ)C^{\prime}(\lambda) condition

The goal of this section is to prove Lemma 2.22, which states that words of a certain form are quasigeodesic in certain central extensions of C(λ)C^{\prime}(\lambda) groups, where 0<λ<1/120<\lambda<\nicefrac{{1}}{{12}}. This is a generalization [Olshanskii_Osin_Sapir, Lemma 5.10], originally used to construct finitely generated groups with circle-tree asymptotic cones. The proof of Lemma 2.22 is a technical argument that involves performing surgery on van Kampen diagrams.

We assume that the reader is familiar with the C(λ)C^{\prime}(\lambda) condition and the notion of a van Kampen diagram. However, there are myriad definitions of van Kampen diagram in the literature, and for our purposes it is necessary to define the C(λ)C^{\prime}(\lambda) condition in a way which, though clearly equivalent to the usual definition, is slightly non-standard. Therefore in Sections 2.1 and 2.2 we fix terminology and notation, and provide all necessary definitions for the following sections.

In Section 2.3, we define signed and unsigned rr-face counts, where rr is a relation of a presentation. We also introduce various operations on van Kampen diagrams, and examine how each of these operations affects the signed and unsigned rr-face counts. Our approach is to treat van Kampen diagrams as graphs embedded in the plane, so that the 2-cells are simply the bounded faces enclosed by the graph. In this way we manipulate van Kampen diagrams directly in the plane and keep topological considerations to a minimum. In Section 2.4, we collect some facts about van Kampen diagrams over C(1/6)C^{\prime}(\nicefrac{{1}}{{6}}) presentations that are used in the proof of Lemma 2.22. Finally, in Section 2.5, we prove Lemma 2.22.

2.1 The C(λ)C^{\prime}(\lambda) condition

Let SS be a set. Let S1S^{-1} be the set of formal inverses of SS, let 11 be a new symbol not in SS, and declare 11=11^{-1}=1. Let

S1=S{1}S=SS1{1}.\begin{split}S_{1}&=S\cup\{1\}\\ S_{\circ}&=S\cup S^{-1}\cup\{1\}.\end{split} (4)

The length of a word ww in the free monoid SS_{\circ}^{*} is denoted |w||w|. There is a unique word of length 0 called the empty word and denoted ε\varepsilon. We define w0w^{0} to be ε\varepsilon for any wSw\in S_{\circ}^{*}. A word wSw\in S_{\circ}^{*} is reduced if ww does not contain a subword of the form 1,ss1,1,ss^{-1}, or s1ss^{-1}s for any sSs\in S, and cyclically reduced if every cylcic shift of ww (including ww itself) is reduced.

Let RR be a language over the alphabet SS_{\circ}, that is, RSR\subseteq S_{\circ}^{*}. Then RR_{*} denotes the closure of RR under taking cyclic shifts and formal inverses of its elements. We say that RR is reduced if every element of RR is reduced, and cyclically reduced if RR_{*} is reduced. We say that RR is cyclically minimal if it does not contain two distinct words, one of which is a cyclic shift of the other word or its inverse. That is, RR is cyclically minimal if R{r}={r}R\cap\{r\}_{*}=\{r\} for each rRr\in R.

A presentation is a pair SR\langle S\mid R\rangle, where SS is a set and RSR\subseteq S_{\circ}^{*}. The notation G=SRG=\langle S\mid R\rangle means that SR\langle S\mid R\rangle is a presentation and GF(S)/RG\cong F(S)/\langle\langle R\rangle\rangle, where F(S)F(S) is the free group with basis SS, and R\langle\langle R\rangle\rangle is the normal closure of RR as a subset of F(S)F(S).

If GG is a group generated by a set SS, there is a natural monoid epimorphism from SS_{\circ}^{*} to GG that evaluates a word in SS_{\circ}^{*} as a product of generators and their inverses, and sends 11 to the identity element. If both GG and SS are understood, then for a word wSw\in S_{\circ}^{*} we denote the image of ww under this homomorphism by w¯\bar{w}. For a group element gGg\in G, we write w=Ggw=_{G}g to abbreviate that w¯=g\bar{w}=g. The word norm on GG with respect to SS is defined by

gG=min{|w|wS,w=Gg}.\|g\|_{G}=\min\{|w|\mid w\in S_{\circ}^{*},w=_{G}g\}\,.

We omit the generating set from the notation since any other choice of finite generating set yields a norm which is bi-Lipschitz equivalent. What matters is that the generating set is fixed throughout any proof in which the word norm plays a role.

A word wSw\in S_{\circ}^{*} is called geodesic in GG if |w|=w¯G|w|=\|\bar{w}\|_{G}. If u,wSu,w\in S_{\circ}^{*}, gGg\in G, ww is geodesic, and w=Gu=Ggw=_{G}u=_{G}g, then ww is called a geodesic representative of uu or of gg in GG. If K,C0K,C\geq 0 are fixed constants, then we say that a word wSw\in S_{\circ}^{*} is (K,C)(K,C)-quasigeodesic in GG if |w|Kw¯G+C|w|\leq K\|\bar{w}\|_{G}+C.

Given two words u,vSu,v\in S_{\circ}^{*}, we say that pp is a piece (of uu and of vv) if there exists u{u},v{v}u^{\prime}\in\{u\}_{*},v^{\prime}\in\{v\}_{*} such that pp is a common prefix of uu^{\prime} and vv^{\prime}.

Definition 2.1.

Let SS be a set, RSR\subseteq S_{\circ}^{*} a language, and λ\lambda a real number with 0<λ<10<\lambda<1. Then RR satisfies C(λ)C^{\prime}(\lambda) if, whenever u,vRu,v\in R and u{u},v{v}u^{\prime}\in\{u\}_{*},v^{\prime}\in\{v\}_{*} witness that pp is a piece of uu and vv, then either u=vu^{\prime}=v^{\prime} or |p|<λmin(|u|,|v|)|p|<\lambda\min(|u|,|v|).

In this case we say that RR is a C(λ)C^{\prime}(\lambda) language. If GG is a group and G=SRG=\langle S\mid R\rangle for some C(λ)C^{\prime}(\lambda) language RR, then SR\langle S\mid R\rangle is called a C(λ)C^{\prime}(\lambda) presentation and GG is called a C(λ)C^{\prime}(\lambda) group.

In most treatments of the C(λ)C^{\prime}(\lambda) condition, it is assumed that R=RR=R_{*}, and a piece is defined to be a common prefix of two distinct words in RR. In our case, however, it is important to assume that RR is cyclically minimal (in particular RRR\neq R_{*}), in order to ensure that the signed rr-face count (Definition 2.5 below) is well defined. For this reason we give the definition above, which, though not the usual definition of the C(λ)C^{\prime}(\lambda) condition, is clearly equivalent.

2.2 van Kampen Diagrams

Let Γ\Gamma be a connected graph. By a path in Γ\Gamma we mean a combinatorial path, which may have repeated edges or vertices: in graph-theoretic terms, our ‘path’ is really a walk. Since points in the interiors of edges generally don’t matter to us, we write xΓx\in\Gamma to mean that xV(Γ)x\in V(\Gamma). Likewise, if α\alpha is a path in Γ\Gamma, then xαx\in\alpha means that xx is a vertex visited by α\alpha.

Let Γ\Gamma be any directed graph, and suppose that Lab:E(Γ)S1\operatorname{Lab}:E(\Gamma)\to S_{1} (see (1) above) is a function which assigns labels from S1S_{1} to the edges of Γ\Gamma. Then we extend Lab\operatorname{Lab} to a map from the set of all paths in Γ\Gamma to SS_{\circ}^{*} in the following natural way.

  • If e=(x,y)e=(x,y) is a directed edge labeled ss, then Lab(x,e,y)=s\operatorname{Lab}(x,e,y)=s and Lab(y,e,x)=s1\operatorname{Lab}(y,e,x)=s^{-1}.

  • If α=(x0,e1,x1,,xn1,en,xn)\alpha=(x_{0},e_{1},x_{1},\ldots,x_{n-1},e_{n},x_{n}) is a path, then

    Lab(α)=Lab(x0,e1,x1)Lab(x1,e2,x2)Lab(xn1,en,xn).\operatorname{Lab}(\alpha)=\operatorname{Lab}(x_{0},e_{1},x_{1})\operatorname{Lab}(x_{1},e_{2},x_{2})\cdots\operatorname{Lab}(x_{n-1},e_{n},x_{n}).

For a path α\alpha we define (α)\ell(\alpha), the length of α\alpha, to be the number of edges traversed by α\alpha, counting multiplicity. Equivalently, (α)=|Lab(α)|\ell(\alpha)=|\operatorname{Lab}(\alpha)|.

A plane graph is a graph which is topologically embedded in 2\mathbb{R}^{2}. A face of a plane graph MM is the closure of a connected component of 2M\mathbb{R}^{2}\smallsetminus M. Let FF be a face of a finite directed plane graph MM with edges labeled by elements of S1S_{1}. Choosing a base point xFx\in\partial F and an orientation counterclockwise (+)(+) or clockwise ()(-), there is a unique circuit which traverses F\partial F exactly once, called the boundary path and denoted (F,x,±)(\partial F,x,\pm). If all properties of (F,x,±)(\partial F,x,\pm) that we care about are preserved after changing its base point and orientation, then we leave these choices out of the notation and write F\partial F. The boundary label of FF is Lab(F,x,±)\operatorname{Lab}(\partial F,x,\pm), sometimes denoted by just Lab(F)\operatorname{Lab}(\partial F). We write M\partial M instead of F\partial F if FF is the unbounded face; from now on, ‘face’ will mean ‘bounded face’ unless otherwise stated.

Definition 2.2.

A van Kampen diagram over a presentation SR\langle S\mid R\rangle is a finite, connected, directed plane graph MM with edges labeled by elements of S1S_{1}, such that if FF is a face of MM, then either Lab(F)R\operatorname{Lab}(\partial F)\in R_{*} or Lab(F)=F(S)1\operatorname{Lab}(\partial F)=_{F(S)}1.

An edge is essential if it is labeled by an element of SS, and inessential if it is labeled by 1. A face FF is called essential if Lab(F)R\operatorname{Lab}(\partial F)\in R_{*} and inessential if Lab(F)=F(S)1\operatorname{Lab}(\partial F)=_{F(S)}1. If RR is cyclically reduced then these cases are mutually exclusive. A face with boundary label rRr\in R is called an rr-face. We call a van Kampen diagram bare if it contains no inessential faces, and padded otherwise.

A subdiagram of a van Kampen diagram MM is a simply connected union of faces of MM. If MM is a van Kampen diagram and DD is a subdiagram of MM, then we call DD simple if D\partial D is a simple closed curve in the plane. Likewise, a face FF of MM is called simple of F\partial F is a simple closed curve.

Let α\alpha and β\beta be two paths in a van Kampen diagram. Then we say that αβ\alpha\cap\beta is trivial if it contains at most one vertex, and nontrivial otherwise. We say that α\alpha and β\beta intersect simply if αβ\alpha\cap\beta a single subpath of both α\alpha and (β\beta or the reverse path of β\beta). Note that this is not the same as saying that αβ\alpha\cap\beta is connected. We apply this terminology to faces as well. For example, if we say that FF and α\alpha intersect simply, it means that there is a choice of base point xFx\in\partial F such that (F,x,+)(\partial F,x,+) and α\alpha intersect simply. If we say that two faces FF and FF^{\prime} intersect simply, it means that (F,x,+)(\partial F,x,+) and (F,x,)(\partial F^{\prime},x,-) intersect simply for some xFFx\in\partial F\cap\partial F^{\prime}.

Let MM be a van Kampen diagram, and suppose FF and FF^{\prime} are distinct faces of MM. Then we say that FF and FF^{\prime} cancel if there exists an edge e=(x,y)e=(x,y) in FF\partial F\cap\partial F^{\prime} such that Lab(F,x,+)=Lab(F,x,)\operatorname{Lab}(\partial F,x,+)=\operatorname{Lab}(\partial F^{\prime},x,-). A van Kampen diagram is called reduced if no two of its faces cancel. We have the following geometric interpretation of the C(λ)C^{\prime}(\lambda) condition, which follows immediately from the definition.

Lemma 2.3.

Let SR\langle S\mid R\rangle be a presentation where RR satisfies C(λ)C^{\prime}(\lambda), and let MM be a van Kampen diagram over SR\langle S\mid R\rangle. Suppose that F,FF,F^{\prime} are essential faces of MM and α\alpha is a common subpath of F\partial F and F\partial F^{\prime}. Then either FF and FF^{\prime} cancel, or (α)<λmin((F),(F))\ell(\alpha)<\lambda\min(\ell(\partial F),\ell(\partial F^{\prime})).

Whenever GG is a group generated by SS, the Cayley graph of GG with respect to SS is denoted Γ(G,S)\Gamma(G,S).

Lemma 2.4 (van Kampen Lemma).

[Lyndon_Schupp, Chapter V, Section 1] Let G=SRG=\langle S\mid R\rangle and wSw\in S_{\circ}^{*}. Then w=G1w=_{G}1 if and only if there exists a van Kampen diagram MM over SR\langle S\mid R\rangle and xMx\in\partial M such that Lab(M,x,+)=w\operatorname{Lab}(\partial M,x,+)=w. Furthermore, given gGg\in G, there exists a combinatorial map f:MΓ(G,S)f:M\to\Gamma(G,S) preserving labels and orientations of edges, such that f(x)=gf(x)=g. In particular, ff does not increase distances, i.e. is 11-Lipschitz.

2.3 Operations on van Kampen diagrams

Given a van Kampen diagram MM over a presentation SR\langle S\mid R\rangle, there are various ways to deform MM within the plane to get another van Kampen diagram MM^{\prime}. To check that the resulting graph MM^{\prime} is really a van Kampen diagram, it suffices to show that the operation preserves connectedness and produces a planar embedding of MM^{\prime}. If one also requires that MM^{\prime} is a van Kampen diagram over the same presentation, one needs to check that any new faces enclosed by the operation have a boundary label which is either in RR_{*} or equal to the identity in F(S)F(S). In this section we list a few operations on van Kampen diagrams that are needed for the proof of Lemma 2.22. In our case it will be necessary to keep track of how each operation affects the boundary label Lab(M)\operatorname{Lab}(\partial M), as well as two quantities that we call the signed and unsigned rr-face counts.

Definition 2.5.

Let MM be a van Kampen diagram over a presentation SR\langle S\mid R\rangle, where RR is cyclically minimal. Let rRr\in R. Then the (unsigned) rr-face count κ(M,r)\kappa(M,r) is the total number of rr-faces in MM.

Definition 2.6.

Let MM be a van Kampen diagram over a presentation SR\langle S\mid R\rangle where RR is cyclically minimal, and let rRr\in R. Then the signed rr-face count σ(M,r)\sigma(M,r) is defined as follows.

  • If FF is a face of MM, then

    σ(F,r)={1 if Lab(F,x,+)=r for some xF1 if Lab(F,x,)=r for some xF0 otherwise.\sigma(F,r)=\begin{cases}1&\text{ if }\operatorname{Lab}(\partial F,x,+)=r\text{ for some }x\in\partial F\\ -1&\text{ if }\operatorname{Lab}(\partial F,x,-)=r\text{ for some }x\in\partial F\\ 0&\text{ otherwise.}\end{cases}
  • σ(M,r)={σ(F,r)F is a face of M}\sigma(M,r)=\sum\{\sigma(F,r)\mid F\text{ is a face of }M\}.

The assumption that RR is cyclically minimal ensures that each face contributes to the signed or unsigned rr-face count of at most one rRr\in R. Note that if FF and FF^{\prime} are two faces of MM that cancel with each other, then σ(F,r)=σ(F,r)\sigma(F,r)=-\sigma(F^{\prime},r) for all rRr\in R.

Operation 2.7 (Removing an inessential edge).

Suppose that e=(x,y)e=(x,y) is an inessential edge of a van Kampen diagram MM over a presentation SR\langle S\mid R\rangle, where RR is cyclically reduced and cyclically minimal, and Lab(M)\operatorname{Lab}(\partial M) is cyclically reduced. Then ee is on the boundary of exactly two inessential bounded faces. There are two possibilities.

  1. (a)

    If xyx\neq y, contract ee to remove it. This will produce a connected, planar embedding of the new graph. This changes two inessential faces with labels 1u1u and 1v1v to two inessential faces with labels uu and vv. Since RR is cyclically reduced, this does not affect the rr-face count for any rRr\in R.

  2. (b)

    If x=yx=y, delete ee to remove it. Since ee is a loop, this will leave the graph connected. This replaces two inessential faces on either side of ee with labels u1u1 and 1v1v with a single inessential face labeled uvuv. Again since RR is cyclically reduced, this operation does not affect σ(M,r)\sigma(M,r) for any rRr\in R.

Note that neither (a) nor (b) can introduce new self-intersections in the boundary path of any face of MM. Also, since Lab(M)\operatorname{Lab}(\partial M) is cyclically reduced, neither operation affects Lab(M)\operatorname{Lab}(\partial M).

Operation 2.8 (Removing a simple subdiagram with trivial boundary label).

Let MM be a van Kampen diagram over SR\langle S\mid R\rangle, where RR and Lab(M)\operatorname{Lab}(\partial M) are both cyclically reduced. Suppose that MM contains a simple subdiagram DD such that D\partial D contains no inessential edges and Lab(D)=F(S)1\operatorname{Lab}(\partial D)=_{F(S)}1. Then D=α+α\partial D=\alpha_{+}\alpha_{-}, where Lab(α)=Lab(α+)1\operatorname{Lab}(\alpha_{-})=\operatorname{Lab}(\alpha_{+})^{-1}. We may then remove DD by replacing DD with a simple inessential face FF and deforming α+\alpha_{+} onto α\alpha_{-} through the interior of FF. This does not affect the boundary label of MM.

Refer to caption
α+\alpha_{+}
α\alpha_{-}
DD
α+\alpha_{+}
α\alpha_{-}
FF
Refer to caption
α+=α\alpha_{+}=\alpha_{-}
Refer to caption
Figure 1:

Note that if FF and FF^{\prime} are simple faces that intersect simply, and FF cancels with FF^{\prime}, then FFF\cup F^{\prime} is a simple subdiagram of MM with trivial boundary label, which may be removed by applying 2.8. Perhaps surprisingly, 2.8 does not always preserve the signed rr-face count, as the following example shows.

Example 2.9.

Figure 2 depicts a van Kampen diagram MM over the presentation
a,ba2,aba1b\langle a,b\mid a^{2},aba^{-1}b\rangle with boundary label bb1bb^{-1}, such that σ(M,aba1b)=2\sigma(M,aba^{-1}b)=2.

Refer to caption
aa
aa
aa
aa
bb
bb
bb
Refer to caption
Figure 2:

However, 2.8 does preserve the signed rr-face count of van Kampen diagrams over C(1/6)C^{\prime}(\nicefrac{{1}}{{6}}) presentations. This is because C(1/6)C^{\prime}(\nicefrac{{1}}{{6}}) presentations are aspherical. The definition of a spherical van Kampen diagram is the same as that of a van Kampen diagram with 2\mathbb{R}^{2} replaced by S2S^{2}: in particular, every face is bounded. A presentation SR\langle S\mid R\rangle is aspherical if every bare spherical van Kampen diagram over SR\langle S\mid R\rangle contains a pair of faces that cancel. The following is a special case of a lemma of Olshanskii.

Lemma 2.10.

[Olshanskii, Lemma 31.1 part 2)] Let SR\langle S\mid R\rangle be an aspherical presentation, and suppose that MM is a van Kampen diagram over SR\langle S\mid R\rangle with boundary label ww, where w=F(S)1w=_{F(S)}1. Then σ(M,r)=0\sigma(M,r)=0 for all rRr\in R.

Operation 2.11 (Padding a vertex).

Suppose that xx is a vertex of MM which appears twice in the boundary path of some face (bounded or unbounded) of MM. Choose ε>0\varepsilon>0 small enough so that B(x,ε)2B(x,\varepsilon)\subset\mathbb{R}^{2} contains only the ends of edges incident to xx. Now B(x,ε)MB(x,\varepsilon)\smallsetminus M consists of finitely many connected components: let these be denoted C0,C1,,CkC_{0},C_{1},\ldots,C_{k}. For each i{0,,k}i\in\{0,\ldots,k\}, insert a clone xix_{i} of xx into CiC_{i}, and connect it to xx with an inessential edge. Then duplicate the edges on either side of xix_{i}, attaching the endpoint meant for xx to xix_{i} instead: see Figure 3. The resulting graph has the same essential faces and boundary path as MM, and one fewer vertex that is a point of self-intersection of the boundary path of a face. Each new inessential face has boundary label 1ss11ss^{-1} for some sSs\in S.

Refer to caption
xx
bb
aa
Refer to caption
xx
aa
Refer to caption
bb
bb
aa
11
CiC_{i}
xix_{i}
Refer to caption
Figure 3:
Operation 2.12 (Quotienting simple faces).

Suppose that G=SRGG=\langle S\mid R_{G}\rangle and H=SRHH=\langle S\mid R_{H}\rangle is a quotient of GG, so every word in RGR_{G} represents the identity element of HH. Suppose that MGM_{G} is a van Kampen diagram over SRG\langle S\mid R_{G}\rangle. Let FF be a simple face of MGM_{G}, and let MFM_{F} be a chosen van Kampen diagram over SRH\langle S\mid R_{H}\rangle with boundary label Lab(F)\operatorname{Lab}(\partial F). Then we may quotient FF to a copy of MFM_{F} without affecting the boundary label of MGM_{G}: see Figure 4. Applying this operation once produces a van Kampen diagram over SRGRH\langle S\mid R_{G}\cup R_{H}\rangle. If FF is the last face of MGM_{G} with label in RGRHR_{G}\smallsetminus R_{H}, then this results in a van Kampen diagram over SRH\langle S\mid R_{H}\rangle. Thus, if this operation can be applied to every essential face of MGM_{G} in sequence, then we obtain a “quotient van Kampen diagram” MHM_{H} over SRH\langle S\mid R_{H}\rangle with the same boundary label as MGM_{G}.

Refer to caption
FF
MGM_{G}
Refer to caption
MFM_{F}
Refer to caption
Figure 4:
Operation 2.13 (Excising a subpath of M\partial M).

Let MM be a van Kampen diagram over a presentation SR\langle S\mid R\rangle, where RR is cyclically minimal and cyclically reduced. Let zMz\in\partial M, and suppose we can write (M,z,+)(\partial M,z,+) as αβ\alpha*\beta, where α\alpha and β\beta are paths of positive length. Suppose that α=α0ρα1\alpha=\alpha_{0}*\rho*\alpha_{1}, where Lab(ρ)\operatorname{Lab}(\rho) is a cyclic shift of r±1r^{\pm 1} for some rRr\in R. Let xx be the initial and yy the terminal vertex of ρ\rho, and suppose xyx\neq y. Then we may contract xx to yy through the unbounded face, identifying the two vertices to obtain a new van Kampen diagram MM^{\prime}. Now MM^{\prime} has exactly one new face FF^{\prime}, where (F,x,)=ρ(\partial F^{\prime},x,-)=\rho, so MM^{\prime} is a van Kampen diagram over the same presentation SR\langle S\mid R\rangle. Also, (M,z,+)=αβ(\partial M^{\prime},z,+)=\alpha^{\prime}*\beta, where α=α0α1\alpha^{\prime}=\alpha_{0}*\alpha_{1}: see Figure 5. Note that ρ\rho may intersect itself: in that case F\partial F^{\prime} will have self-intersections in MM^{\prime}, but this is fine. The only topological feature of MM which is essential to this operation is that xx and yy are distinct.

Refer to caption
ρ\rho
xx
yy
Refer to caption
ρ\rho
Refer to caption
x=yx=y
Refer to caption
β\beta
α\alpha
β\beta
α\alpha
Refer to caption
Figure 5:

Now (α)=(α)|r|\ell(\alpha^{\prime})=\ell(\alpha)-|r|, and β\beta is unaffected by the operation. Also, for all rRr^{\prime}\in R,

κ(M,r)\displaystyle\kappa(M^{\prime},r^{\prime}) ={κ(M,r)+1 if r=rκ(M,r) otherwise.\displaystyle=\begin{cases}\kappa(M,r^{\prime})+1&\text{ if }r^{\prime}=r\\ \kappa(M,r^{\prime})&\text{ otherwise.}\end{cases}
σ(M,r)\displaystyle\sigma(M^{\prime},r^{\prime}) ={σ(M,r)1 if r=r and Lab(ρ) is a cyclic shift of rσ(M,r)+1 if r=r and Lab(ρ) is a cyclic shift of r1σ(M,r) otherwise.\displaystyle=\begin{cases}\sigma(M,r^{\prime})-1&\text{ if }r^{\prime}=r\text{ and }\operatorname{Lab}(\rho)\text{ is a cyclic shift of }r\\ \sigma(M,r^{\prime})+1&\text{ if }r^{\prime}=r\text{ and }\operatorname{Lab}(\rho)\text{ is a cyclic shift of }r^{-1}\\ \sigma(M,r^{\prime})&\text{ otherwise.}\end{cases}

Note that, since RR is cyclically reduced, these last three cases are all distinct. Indeed, it is an easy exercise to show that if a word rRr\in R is a cyclic shift of r1r^{-1}, then rr is not reduced.

2.4 Reductions that preserve signed rr-face counts

Later we will need to use Lemma 2.27, a result that applies only to bare, reduced van Kampen diagrams over C(1/6)C^{\prime}(\nicefrac{{1}}{{6}}) presentations. At the same time, we would like to apply this result to van Kampen diagrams with signed rr-face counts that are carefully controlled. Thus, we need to establish a method of taking a van Kampen diagram over a C(1/6)C^{\prime}(\nicefrac{{1}}{{6}}) presentation, and making it bare and reduced without affecting the signed rr-face counts. In this subsection, we develop such a process, which is encapsulated in Lemma 2.20. We then prove Lemma 2.21, which allows us to construct certain ‘quotient’ van Kampen diagrams with controlled rr-face counts.

Lemma 2.14.

Let MM be a van Kampen diagram over SR\langle S\mid R\rangle such that RR is cyclically minimal and cyclically reduced, and Lab(M)\operatorname{Lab}(\partial M) is cyclically reduced. Then there exists a van Kampen diagram MM^{\prime} such that all of the following conditions hold.

  1. (a)

    Lab(M)=Lab(M)\operatorname{Lab}(\partial M^{\prime})=\operatorname{Lab}(\partial M).

  2. (b)

    σ(M,r)=σ(M,r)\sigma(M^{\prime},r)=\sigma(M,r) for all rRr\in R.

  3. (c)

    Every inessential face of MM^{\prime} has boundary label ss1ss^{-1} or 1ss11ss^{-1} for some sSs\in S.

  4. (d)

    All inessential edges of MM^{\prime} are loops.

Proof.

Let II be the set of all inessential faces in MM whose boundary labels are not equal to 1ss11ss^{-1} or ss1ss^{-1} for some sSs\in S. Let FIF\in I. If F\partial F consists of a single inessential edge loop, we simply contract this loop to remove FF; it is easy to see that this preserves the boundary label as well as the signed and unsigned face counts of MM, and whether MM satisfies (c) or (d). Therefore we assume that F\partial F is at least two edges long. Now we repeatedly pad vertices of F\partial F (2.11) until FF is simple. Since each inessential face added in the process has boundary label 1ss11ss^{-1} for some sSs\in S, this does not increase |I||I|.

We claim that, without loss of generality, we may assume that F\partial F contains no inessential edges. Suppose that F\partial F contains an inessential edge ee. Since M\partial M and RR are both cyclically reduced, ee lies on the boundary of exactly two inessential, bounded faces, one of which is FF: call the other one FF^{\prime}. Then for some u,uSu,u^{\prime}\in S_{\circ}^{*} we have Lab(F)=1u\operatorname{Lab}(F)=1u and Lab(F)=1u\operatorname{Lab}(F^{\prime})=1u^{\prime}. We know that ee must have distinct endpoints since F\partial F is a simple closed curve and (F)2\ell(\partial F)\geq 2. Therefore we may remove ee using 2.7 (a). This changes the boundary label of FF from 1u1u to uu, and the boundary label of FF^{\prime} from 1u1u^{\prime} to uu^{\prime}. Thus it does not change whether or not FF or FF^{\prime} is a member of II. Therefore removing ee does not change |I||I|, and without loss of generality we may assume that F\partial F contains no inessential edges.

Since FF is simple, Lab(F)=F(S)1\operatorname{Lab}(\partial F)=_{F(S)}1, and F\partial F contains no inessential edges, we may remove FF using 2.8. This reduces |I||I| by 1. Since Lab(M)\operatorname{Lab}(\partial M) is cyclically reduced, none of the previous operations affect Lab(M)\operatorname{Lab}(\partial M). Since only inessential faces were removed, and RR is cyclically reduced, σ(M,r)\sigma(M,r) is also preserved for all rRr\in R. Repeating this process, we obtain a diagram MM^{\prime} for which (a) and (b) hold and |I|=0|I|=0, i.e. such that (a)-(c) hold. At this point we may repeatedly apply 2.7 (a) to remove all inessential edges of MM^{\prime} with distinct endpoints, so that (d) holds in MM^{\prime}. Reasoning as in the previous paragraph, one can see that this does not interfere with conditions (a)-(c). Thus (a)-(d) hold in MM^{\prime}, finishing the construction. ∎

Lemma 2.15.

Let GG be a group given by presentation SR\langle S\mid R\rangle, where RR is cyclically minimal and cyclically reduced, and sG1s\neq_{G}1 for any sSs\in S. Let MM be a van Kampen diagram over SR\langle S\mid R\rangle such that Lab(M)\operatorname{Lab}(\partial M) is cyclically reduced. Then there exists a van Kampen diagram MM^{\prime} over SR\langle S\mid R\rangle such that all of the following conditions hold.

  1. (a)

    Lab(M)=Lab(M)\operatorname{Lab}(\partial M^{\prime})=\operatorname{Lab}(\partial M).

  2. (b)

    σ(M,r)=σ(M,r)\sigma(M^{\prime},r)=\sigma(M,r) for all rRr\in R.

  3. (c)

    Every inessential face of MM^{\prime} is contained in a simple subdiagram whose boundary label is equal to ss1ss^{-1} for some sSs\in S.

Proof.

We may assume that we have a van Kampen diagram MM^{\prime} that satisfies (a)-(d) of Lemma 2.14. We prove here that, in the presence of the assumption that sG1s\neq_{G}1 for all sSs\in S, it follows that MM^{\prime} also satisfies conclusion (c) of the current lemma.

Let FF be an inessential face of MM^{\prime}. There are two cases: either Lab(F)=ss1\operatorname{Lab}(\partial F)=ss^{-1} or Lab(F)=1ss1\operatorname{Lab}(\partial F)=1ss^{-1} for some sSs\in S.

If Lab(F)=ss1\operatorname{Lab}(\partial F)=ss^{-1}, then since sG1s\neq_{G}1, we have that FF is simple. Thus FF itself is a simple subdiagram of MM^{\prime} which contains FF and has boundary label ss1ss^{-1}.

Suppose on the other hand that Lab(F)=1aa1\operatorname{Lab}(\partial F)=1aa^{-1}, where aSa\in S. Let ee be the inessential edge of F\partial F. Since M\partial M^{\prime} and RR are both cyclically reduced, ee lies on the boundary paths of exactly two inessential, bounded faces, one of which is FF: call the other one FF^{\prime}. Since FF^{\prime} is an inessential face of MM^{\prime} with an inessential edge on its boundary path, we have that Lab(F)=1bb1\operatorname{Lab}(F^{\prime})=1bb^{-1} for some bSb\in S. Now ee is a loop since MM^{\prime} satisfies (d) of Lemma 2.14. Since sG1s\neq_{G}1, the endpoints of each of the aa-labeled edges of F\partial F are distinct: similarly for the bb-labeled edges of F\partial F^{\prime}. Therefore FFF\cup F^{\prime} takes the form depicted in Figure 6, allowing that the roles of FF and FF^{\prime} may be switched.

Refer to caption
bb
aa
bb
Refer to caption
11
Refer to caption
aa
FF^{\prime}
FF
Figure 6:

Notice that in Figure 6, FFF\cup F^{\prime} is enclosed in a simple subdiagram with boundary label aa1aa^{-1} (or bb1bb^{-1}, if the roles of FF and FF^{\prime} are switched). This finishes the second case, thus (c) holds for MM^{\prime} and we are done. ∎

Corollary 2.16.

Let GG be a group given by an aspherical presentation SR\langle S\mid R\rangle, where RR is cyclically minimal and cyclically reduced, and sG1s\neq_{G}1 for all sSs\in S. Let MM be a van Kampen diagram over SR\langle S\mid R\rangle such that Lab(M)\operatorname{Lab}(\partial M) is cyclically reduced. Then there exists a van Kampen diagram MM^{\prime} such that all of the following conditions hold.

  1. (a)

    Lab(M)=Lab(M)\operatorname{Lab}(\partial M^{\prime})=\operatorname{Lab}(\partial M).

  2. (b)

    σ(M,r)=σ(M,r)\sigma(M^{\prime},r)=\sigma(M,r) for all rRr\in R.

  3. (c)

    MM^{\prime} is bare.

Proof.

We may assume that MM^{\prime} satisfies (a)-(c) of Lemma 2.15. Now all inessential faces of MM^{\prime} are contained in simple subdiagrams of MM^{\prime} with boundary label ss1ss^{-1} for some sSs\in S. Thus we may make MM^{\prime} bare by repeatedly applying 2.8. 2.8 always preserves the boundary label of a van Kampen diagram, so (a) holds. Since SR\langle S\mid R\rangle is aspherical, it follows from Lemma 2.10 that each application of 2.8 preserves σ(M,r)\sigma(M^{\prime},r) for all rRr\in R. Thus (a)-(c) hold for MM^{\prime}, and we are done. ∎

Often one would like to take a van Kampen diagram MM which is not reduced, and reduce it using 2.8. However, the canceling faces may not be simple, or may not intersect each other simply. A common solution is to pad the van Kampen diagram with inessential faces. However, if MM is a van Kampen diagram over a C(1/6)C^{\prime}(\nicefrac{{1}}{{6}}) presentation, then MM is topologically well-behaved enough to perform this operation without the use of inessential faces. We make this statement precise in the following two lemmas. The second is a consequence of the first, which is the famous Greendlinger Lemma.

Lemma 2.17 (Greendlinger Lemma).

[Lyndon_Schupp, Chapter V, Theorem 4.5] Let MM be a bare and reduced van Kampen diagram over a cyclically reduced C(λ)C^{\prime}(\lambda) presentation, where λ1/6\lambda\leq\nicefrac{{1}}{{6}}, such that MM has at least one bounded face and Lab(M)\operatorname{Lab}(\partial M) is cyclically reduced. Then there exists a face FF of MM such that F\partial F and M\partial M share a common subpath of length more than 12(F)\frac{1}{2}\ell(\partial F).

Lemma 2.18.

Let MM be a bare van Kampen diagram over a cyclically reduced C(1/6)C^{\prime}(\nicefrac{{1}}{{6}}) presentation. Then

  1. (a)

    If MM is reduced, then every face of MM is simple.

  2. (b)

    If MM is reduced, then every two faces of MM that intersect nontrivially also intersect simply.

  3. (c)

    If MM is not reduced, then there exists a pair of faces that cancel and intersect simply.

Proof.

We refer the reader to [Lyndon_Schupp, Chapter V, Lemma 4.1] for the proof of part (a). For part (b), suppose that MM is a counterexample with the minimum number of faces, and that FF and FF^{\prime} are two faces of MM that do not intersect simply, i.e. such that F\partial F intersects F\partial F^{\prime} in more than one maximal common subpath. Then F\partial F and F\partial F^{\prime} together enclose at least one simple subdiagram of MM, call it DD. Since MM is reduced, so is DD. By the Greendlinger Lemma, there exists a face EE of DD such that E\partial E intersects D\partial D in a subpath of length at least 12(E)\frac{1}{2}\ell(\partial E). Therefore E\partial E intersects one of FF or FF^{\prime}, say FF, in a common subpath of length at least 14(E)\frac{1}{4}\ell(\partial E). But then EE and FF cancel, contradicting the assumption that MM is reduced.

For part (c), suppose that MM is a counterexample with the minimum number of faces. Then MM is not reduced, and there are two faces FF and FF^{\prime} that cancel but do not intersect simply. Therefore F\partial F and F\partial F^{\prime} together enclose a simple subdiagram DD. Again DD must be reduced, this time by minimality of MM. By an argument similar to the one in the preceding paragraph, there is a face EE of DD that cancels with FF. By assumption, EE and FF cannot intersect simply. But then DFD\cup F is a subdiagram of MM that is a counterexample with strictly fewer faces than MM, since it does not include FF^{\prime}. This contradicts minimality of MM, finishing the proof. ∎

One can then use Lemma 2.18 to prove the following corollary.

Corollary 2.19.

Let GG be a group given by presentation SR\langle S\mid R\rangle, where RR is cyclically reduced and satisfies C(1/6)C^{\prime}(\nicefrac{{1}}{{6}}). Then for every rRr\in R, if uu is a subword of an element of {r}\{r\}_{*}, then uG1u\neq_{G}1. In particular,

  1. (a)

    For all generators sSs\in S, if sRs\not\in R then sG1s\neq_{G}1.

  2. (b)

    If MM is a bare van Kampen diagram over SR\langle S\mid R\rangle, then every face of MM is simple.

  3. (c)

    If MM is a bare van Kampen diagram over SR\langle S\mid R\rangle which is not reduced, then there exists a pair of cancelling faces that are simple and intersect simply.

Proof.

Suppose otherwise, and choose uSu\in S_{\circ}^{*} to be a word of minimum length which is a proper subword of {r}\{r\}_{*} for some rRr\in R. Without loss of generality, suppose that uu is a prefix of rr, so that r=uvr=uv for some vSv\in S_{\circ}^{*}. Clearly v=G1v=_{G}1, so by minimality of uu we have that |v||u||v|\geq|u|, hence |u|12|r||u|\leq\frac{1}{2}|r|.

Let MM be a reduced van Kampen diagram over SR\langle S\mid R\rangle such that Lab(M)=u\operatorname{Lab}(\partial M)=u. Since RR is cyclically reduced, so is uu: in particular, MM has at least one bounded face. By the Greendlinger Lemma, there exists a face FF of MM such that FF shares a common subpath of length more than 12(F)\frac{1}{2}\ell(\partial F) with M\partial M. Let the label of this common subpath be ww. Then ww is a piece of rr and Lab(F)\operatorname{Lab}(\partial F), of length more than 12(F)\frac{1}{2}\ell(\partial F). By the C(1/6)C^{\prime}(\nicefrac{{1}}{{6}}) condition, Lab(F)=r\operatorname{Lab}(\partial F)=r. But then |w|>12|r||u||w|>\frac{1}{2}|r|\geq|u|. Since ww is the label of a subpath of M\partial M, this is a contradiction.

Conclusions (a) and (b) follow directly. Part (c) follows from part (b) of the current lemma and Lemma 2.18 (c). ∎

Lemma 2.20.

Let MM be a van Kampen diagram over a C(1/6)C^{\prime}(\nicefrac{{1}}{{6}}) presentation SR\langle S\mid R\rangle, where RR is cyclically minimal and cyclically reduced, and |r|2|r|\geq 2 for all rRr\in R. Then there exists a van Kampen diagram MM^{\prime} over SR\langle S\mid R\rangle such that

  1. (a)

    Lab(M)=Lab(M)\operatorname{Lab}(\partial M^{\prime})=\operatorname{Lab}(\partial M).

  2. (b)

    σ(M,r)=σ(M,r)\sigma(M^{\prime},r)=\sigma(M,r) for all rRr\in R.

  3. (c)

    MM^{\prime} is bare and reduced.

Proof.

We may assume that MM^{\prime} satisfies (a)-(c) of Corollary 2.16. Thus we only have to show that it is possible to transform MM^{\prime} so that it is reduced, while preserving the boundary label and signed rr-face count for each rRr\in R, and without adding any inessential faces.

Suppose that MM^{\prime} is not reduced. Since MM^{\prime} is bare, by Corollary 2.19 there exist two simple faces FF and FF^{\prime} that cancel and intersect simply. Thus FFF\cup F^{\prime} is a simple subdiagram of MM with trivial boundary label. Now remove FFF\cup F^{\prime} with 2.8. Since SR\langle S\mid R\rangle is C(1/6)C^{\prime}(\nicefrac{{1}}{{6}}), and therefore aspherical, this operation preserves σ(M,r)\sigma(M^{\prime},r) for all rRr\in R. Repeating, we end up with a reduced van Kampen diagram. ∎

Lemma 2.21.

Let G,HG,H be groups given by presentations

G\displaystyle G =SRG\displaystyle=\langle S\mid R_{G}\rangle
H\displaystyle H =SRH\displaystyle=\langle S\mid R_{H}\rangle

where SRH\langle S\mid R_{H}\rangle is a cyclically reduced C(1/6)C^{\prime}(\nicefrac{{1}}{{6}}) presentation, and |rH|2|r_{H}|\geq 2 for all rHRHr_{H}\in R_{H}. Suppose that rG=H1r_{G}=_{H}1 for all rGRGr_{G}\in R_{G}, so HH is a quotient of GG. Let MGM_{G} be a van Kampen diagram over SRG\langle S\mid R_{G}\rangle, and for each face FF of MGM_{G}, let MFM_{F} be a van Kampen diagram over SR\langle S\mid R\rangle with boundary label Lab(F)\operatorname{Lab}(\partial F). Then there exists a “quotient van Kampen diagram” MHM_{H} over SRH\langle S\mid R_{H}\rangle such that

  1. (a)

    Lab(MG)=Lab(MH)\operatorname{Lab}(\partial M_{G})=\operatorname{Lab}(\partial M_{H}).

  2. (b)

    σ(MH,r)={σ(MF,r)F is an essential face of MG}\sigma(M_{H},r)=\sum\{\sigma(M_{F},r)\mid F\text{ is an essential face of }M_{G}\}.

  3. (c)

    MHM_{H} is bare and reduced.

Proof.

Start with MGM_{G}. By repeatedly padding vertices, we may assume that all essential faces of MGM_{G} are simple.

Now take an essential disk face FF of MGM_{G}, and quotient it to MFM_{F}. This may introduce self-intersections among essential faces in MGM_{G}. Pad vertices again until all essential faces of MGM_{G} are simple, and repeat as many times as necessary to quotient all essential faces that were originally in MGM_{G}. Since padding vertices and quotienting simple faces preserve the boundary label, we obtain a van Kampen diagram MHM_{H} over SRH\langle S\mid R_{H}\rangle, possibly with many inessential faces, such that Lab(MH)=Lab(MG)\operatorname{Lab}(\partial M_{H})=\operatorname{Lab}(\partial M_{G}). Thus (a) holds. In addition, for all rRHr\in R_{H},

σ(MH,r)={σ(MF,r)F is a face of MG},\sigma(M_{H},r)=\sum\{\sigma(M_{F},r)\mid F\text{ is a face of }M_{G}\}\,,

so (b) holds as well.

Note that we do not require RGR_{G} to by cyclically reduced for any of the previous steps to work. However, RHR_{H} is cyclically reduced, thus by Lemma 2.20 we may ensure that (c) holds, without interfering with conditions (a) or (b). ∎

2.5 A technical lemma

This section is devoted to proving the following lemma. In essence it is similar to [Olshanskii_Osin_Sapir, Lemma 5.10], but for our purposes we need the more general version stated here. In order to avoid constantly reiterating the assumptions, the notation used in this lemma will be ‘globally fixed’ for this section. Thus until the next section, GG will always refer to the group with presentation given in Lemma 2.22, etc. Any new notation introduced in the body of this section will also remain fixed until the beginning of the next section.

Lemma 2.22.

Let λ\lambda be a real number, where 0<λ<1/120<\lambda<\nicefrac{{1}}{{12}}. Let {ii}\{\ell_{i}\mid i\in\mathbb{N}\} be a set of positive integers, where each i2\ell_{i}\geq 2. Let SS be a finite set. Let

U={uii}S\displaystyle U=\{u_{i}\mid i\in\mathbb{N}\}\subset S_{\circ}^{*} V={vii}S\displaystyle V=\{v_{i}\mid i\in\mathbb{N}\}\subset S_{\circ}^{*}

be languages, and let u~S\tilde{u}\in S_{\circ}^{*} be a word, such that the following conditions are satisfied for all i,ii,i^{\prime}\in\mathbb{N}.

  1. (a)

    UVU\cup V is cyclically minimal and cyclically reduced, and satisfies C(λ)C^{\prime}(\lambda).

  2. (b)

    2|ui||vi|2\leq|u_{i}|\leq|v_{i}|.

  3. (c)

    If pp is a piece of u~\tilde{u} and uiu_{i}, then |p|<λ|ui||p|<\lambda|u_{i}|, and the same statement holds if uiu_{i} is replaced with viv_{i}.

  4. (d)

    If ui=uiu_{i}=u_{i^{\prime}}, vi=viv_{i}=v_{i^{\prime}}, or ui=viu_{i}=v_{i^{\prime}}, then i=ii=i^{\prime}.

Now let

G\displaystyle G =SRG:=S[s,ui],uii,uivi1:sS,i\displaystyle=\langle S\mid R_{G}\rangle:=\langle S\mid[s,u_{i}],u_{i}^{\ell_{i}},u_{i}v_{i}^{-1}:s\in S,i\in\mathbb{N}\rangle
H\displaystyle H =SRH:=SUV=Sui,vi:i.\displaystyle=\langle S\mid R_{H}\rangle:=\langle S\mid U\cup V\rangle=\langle S\mid u_{i},v_{i}:i\in\mathbb{N}\rangle\,.

Let (ki)(k_{i}) be a sequence of integers where |ki|i/2|k_{i}|\leq\ell_{i}/2 for all ii\in\mathbb{N}, and ki=0k_{i}=0 for all but finitely many ii\in\mathbb{N}. Let uSu\in S_{\circ}^{*} be a word of the form

u=u~i=0uiki.u=\tilde{u}\prod_{i=0}^{\infty}{u_{i}^{k_{i}}}_{\,.}

Then uu is (3112λ,0)\left({\frac{3}{1-12\lambda}}_{\,,}0\right)-quasigeodesic in GG.

Note that if UVU\cup V is C(λ)C^{\prime}(\lambda), then so is (UV{ui1}){ui}(U\cup V\cup\{u_{i}^{-1}\})\smallsetminus\{u_{i}\}. Therefore assume without loss of generality that all kik_{i} are nonnegative.

Let ww be a geodesic representative of uu in GG. Then uw1=G1uw^{-1}=_{G}1, so by the van Kampen Lemma, there exists a van Kampen diagram MGM_{G} with Lab(MG)=uw1\operatorname{Lab}(\partial M_{G})=uw^{-1}.

Lemma 2.23.

There exists a van Kampen diagram MHM_{H} over SUV\langle S\mid U\cup V\rangle such that

  1. (a)

    MHM_{H} is bare and reduced.

  2. (b)

    MH=αβ\partial M_{H}=\alpha*\beta, where Lab(α)=u\operatorname{Lab}(\alpha)=u and Lab(β)=w1\operatorname{Lab}(\beta)=w^{-1}.

  3. (c)

    σ(MH,ui)+σ(MH,vi)0modi\sigma(M_{H},u_{i})+\sigma(M_{H},v_{i})\equiv 0\mod\ell_{i} for all ii\in\mathbb{N}.

  4. (d)

    σ(MH,ui)0modi\sigma(M_{H},u_{i})\equiv 0\mod\ell_{i} for all ii\in\mathbb{N} such that ui=viu_{i}=v_{i}.

Proof.

Each face FF of MGM_{G} has boundary label equal to either [s,ui][s,u_{i}], uiiu_{i}^{\ell_{i}}, or uivi1u_{i}v_{i}^{-1}. Each of these words represents the trivial elment of HH. For each face FF of MGM_{G}, choose a van Kampen diagram MFM_{F} over SRH\langle S\mid R_{H}\rangle, of one of forms depicted in Figure 7.

Refer to caption
uiu_{i}
uiu_{i}
i\ell_{i}
uiu_{i}
uiu_{i}
uiu_{i}
Refer to caption
ss
Refer to caption
uiu_{i}
viv_{i}
Figure 7:

Applying Lemma 2.21, there exists a bare, reduced van Kampen diagram MHM_{H} with Lab(MH)=Lab(MG)=uw1\operatorname{Lab}(M_{H})=\operatorname{Lab}(M_{G})=uw^{-1}, such that for all ii\in\mathbb{N},

σ(M,ui)+σ(M,vi)={σ(MF,ui)+σ(MF,vi)F is a face of MG}.\sigma(M,u_{i})+\sigma(M,v_{i})=\sum\{\sigma(M_{F},u_{i})+\sigma(M_{F},v_{i})\mid F\text{ is a face of }M_{G}\}\,.

Thus (a) and (b) hold. Now notice that for all MFM_{F} depicted in Figure 7, σ(MF,ui)+σ(MF,vi)0modi\sigma(M_{F},u_{i})+\sigma(M_{F},v_{i})\equiv 0\mod\ell_{i}. Also, if ui=viu_{i}=v_{i}, then σ(MF,ui)0modi\sigma(M_{F},u_{i})\equiv 0\mod\ell_{i}. Thus (c) and (d) hold as well. ∎

Let MHM_{H} be the van Kampen diagram from Lemma 2.23. Let k=ikik=\sum_{i\in\mathbb{N}}k_{i}. Then we may write

α=α~α0αkβ\alpha=\tilde{\alpha}*\alpha_{0}*\dots*\alpha_{k}*\beta

where Lab(α~)=u~\operatorname{Lab}(\tilde{\alpha})=\tilde{u}, and for all j{0,,k}j\in\{0,\ldots,k\} we have Lab(αj)=ui\operatorname{Lab}(\alpha_{j})=u_{i} for some ii\in\mathbb{N}.

Lemma 2.24.

There exists a van Kampen diagram MHM_{H}^{\prime} over SUV\langle S\mid U\cup V\rangle and natural numbers {hii}\{h_{i}\mid i\in\mathbb{N}\} satisfying all of the following conditions for all ii\in\mathbb{N}.

  1. (a)

    MHM_{H}^{\prime} is bare and reduced.

  2. (b)

    κ(MH,ui)=κ(MH,ui)hi\kappa(M_{H}^{\prime},u_{i})=\kappa(M_{H},u_{i})-h_{i}.

  3. (c)

    MH=αβ\partial M_{H}^{\prime}=\alpha^{\prime}*\beta^{\prime}, where Lab(α)=u~i=0uikihi\operatorname{Lab}(\alpha^{\prime})=\tilde{u}\prod_{i=0}^{\infty}u_{i}^{k_{i}-h_{i}} and Lab(β)=w\operatorname{Lab}(\beta^{\prime})=w.

  4. (d)

    No face FF of MHM_{H}^{\prime} intersects α\alpha^{\prime} in a common subpath of length at least 2λ(F)2\lambda\ell(\partial F).

  5. (e)

    0hikii/20\leq h_{i}\leq k_{i}\leq\ell_{i}/2.

Proof.

If MHM_{H} already satisfies (d), then all conditions are satisfied by setting MH=MHM_{H}^{\prime}=M_{H} and hi=0h_{i}=0 for all ii\in\mathbb{N}. Therefore suppose that MHM_{H} does not satisfy (d), i.e. there exists a face FF of MHM_{H} such that F\partial F intersects α\alpha in a common subpath of length at least 2λ(F)2\lambda\ell(\partial F). Then there must be a common subpath of F\partial F and α~\tilde{\alpha} or αj\alpha_{j} for some j{0,,k}j\in\{0,\ldots,k\}, of length at least λ(F)\lambda\ell(\partial F). The former possibility is excluded by condition (c) of Lemma 2.22. Thus F\partial F intersects αj\alpha_{j} in a common subpath of length at least λ(F)\lambda\ell(\partial F) for some j{0,,k}j\in\{0,\ldots,k\}. Call this common subpath γ\gamma. Since Lab(αj)=ui\operatorname{Lab}(\alpha_{j})=u_{i} for some ii\in\mathbb{N}, by the C(λ)C^{\prime}(\lambda) condition we have that Lab(F)=ui\operatorname{Lab}(\partial F)=u_{i} as well.

Now apply 2.13 to excise αj\alpha_{j} from α\alpha. Let FF^{\prime} be the new uiu_{i}-face created by this operation. Then γ\gamma is a common subpath of FF and FF^{\prime} of length at least λ(F)=λ(F)\lambda\ell(\partial F)=\lambda\ell(\partial F^{\prime}), so FF and FF^{\prime} cancel. Since MHM_{H} was reduced, FF and FF^{\prime} are the only pair of faces that cancel at this stage. Therefore by Corollary 2.19 (c), FF and FF^{\prime} are simple and intersect simply. Thus FFF\cup F^{\prime} is a simple subdiagram of MM with trivial boundary label, which we may remove with 2.8.

Let M^H\hat{M}_{H} be the van Kampen diagram obtained in this way. Then clearly M^H\hat{M}_{H} satisfies (a). We added one uiu_{i}-face and removed two, so κ(M^H,ui)=κ(MH,ui)1\kappa(\hat{M}_{H},u_{i})=\kappa(M_{H},u_{i})-1. Since uiuju_{i}\neq u_{j} whenever iji\neq j, no uju_{j}-face counts were affected for any jij\neq i. Therefore (b) is satisfied with hi=1h_{i}=1. Now after excising αj\alpha_{j} the boundary path becomes α^β:=α~α0αj1αj+1αkβ\hat{\alpha}*\beta:=\tilde{\alpha}*\alpha_{0}*\dots*\alpha_{j-1}*\alpha_{j+1}*\dots*\alpha_{k}*\beta. Removing FFF\cup F^{\prime} does not change the boundary label, so (c) is satisfied with hi=1h_{i}=1. Because of this, we may iterate the process. By construction, M^H\hat{M}_{H} has one fewer face than MHM_{H} which fails to satisfy (d). Therefore repeat as many times as there are faces in MHM_{H} failing to satisfy (d) to get MHM_{H}^{\prime}. Each such face must be a uiu_{i}-face for some ii\in\mathbb{N}, so for each ii\in\mathbb{N}, let hih_{i} be the number of uiu_{i}-faces in MHM_{H} failing to satisfy (d). Since the boundary label becomes shorter at each step, by (c) it follows that hikih_{i}\leq k_{i} for all ii\in\mathbb{N}. Therefore MHM_{H}^{\prime} satisfies (e), and we are done. ∎

For the next step in the proof, the following ad hoc lemma is useful.

Lemma 2.25.

Let MM be a bare, reduced van Kampen diagram over a cyclically reduced C(1/6)C^{\prime}(\nicefrac{{1}}{{6}}) presentation. Let α\alpha be a subpath of M\partial M such that no face FF of MM intersects α\alpha in a common subpath of length at least 14(F)\frac{1}{4}\ell(\partial F). Then every face of MM that intersects α\alpha nontrivially, intersects α\alpha simply.

Proof.

Suppose that FF is a face of MM such that F\partial F shares more than one vertex with α\alpha, but F\partial F does not intersect α\alpha simply. Then there exist subpaths of α\alpha and F\partial F that together enclose a simple subdiagram DD of MM. Since MM, and therefore DD, is reduced, by the Greendlinger Lemma there exists a face FF^{\prime} of DD such that F\partial F^{\prime} shares a common subpath of length at least 12(F)\frac{1}{2}\ell(\partial F^{\prime}) with D\partial D. Thus F\partial F^{\prime} intersects either F\partial F or α\alpha in a common subpath of length at least 14(F)\frac{1}{4}\ell(\partial F^{\prime}). The latter possibility is ruled out by assumption, so FF cancels with FF^{\prime} by the C(1/6)C^{\prime}(\nicefrac{{1}}{{6}}) condition. But this contradicts our assumption that MM is reduced. ∎

Definition 2.26.

Let MM be a van Kampen diagram over a presentation SR\langle S\mid R\rangle. Then the perimeter sum of MM, denoted PS(M)\operatorname{PS}(M), is defined by

PS(M)={(F)F is a face of M}.\operatorname{PS}(M)=\sum\{\ell(\partial F)\mid F\text{ is a face of }M\}\,.

Note that if MM is bare, then

PS(M)=rR|r|κ(M,r).\operatorname{PS}(M)=\sum_{r\in R}|r|\kappa(M,r)\,.

To obtain bounds on (α)\ell(\alpha^{\prime}) in terms of (β)\ell(\beta), and on (α)\ell(\alpha) in terms of (α)\ell(\alpha^{\prime}), we use the following fact about van Kampen diagrams over C(1/6)C^{\prime}(\nicefrac{{1}}{{6}}) presentations. It is the final puzzle piece in the proof.

Lemma 2.27.

[Olshanskii_Osin_Sapir, Lemma 3.8] Let MM be a bare and reduced van Kampen diagram over a cyclically reduced C(λ)C^{\prime}(\lambda) presentation, where λ1/6\lambda\leq\nicefrac{{1}}{{6}}. Then (16λ)PS(M)(M)(1-6\lambda)\operatorname{PS}(M)\leq\ell(\partial M).

With this in mind, we resume our proof.

Lemma 2.28.

(α)<2(β)\ell(\alpha^{\prime})<2\ell(\beta).

Proof.

Note that an edge of α\alpha^{\prime} is shared by the boundary path of some face of MHM_{H}^{\prime} if and only if it is not also an edge of β\beta^{\prime}. We have by Lemma 2.24 (d) that no face FF intersects α\alpha^{\prime} in a common subpath of length at least 2λ(F)<16(F)<14(F)2\lambda\ell(\partial F)<\frac{1}{6}\ell(\partial F)<\frac{1}{4}\ell(\partial F). Therefore by Lemma 2.25, every face whose boundary path shares an edge with α\alpha^{\prime} intersects α\alpha^{\prime} in a single common subpath. Thus

PS(MH)>12λ(αβ)12λ((α)(β))=12λ((α)(β))\operatorname{PS}(M_{H})>\tfrac{1}{2\lambda}\ell(\alpha^{\prime}\smallsetminus\beta^{\prime})\geq\tfrac{1}{2\lambda}\left(\ell(\alpha^{\prime})-\ell(\beta^{\prime})\right)=\tfrac{1}{2\lambda}\left(\ell(\alpha^{\prime})-\ell(\beta^{\prime})\right)

On the other hand, (MH)=(α)+(β)\ell(\partial M_{H}^{\prime})=\ell(\alpha^{\prime})+\ell(\beta^{\prime}). Thus by Lemma 2.27,

12λ((α)(β))<PS(MH)\displaystyle\tfrac{1}{2\lambda}\left(\ell(\alpha^{\prime})-\ell(\beta^{\prime})\right)<\operatorname{PS}(M_{H}^{\prime}) 116λ(MH)=116λ((α)+(β))\displaystyle\leq\tfrac{1}{1-6\lambda}\ell(\partial M_{H}^{\prime})=\tfrac{1}{1-6\lambda}\left(\ell(\alpha^{\prime})+\ell(\beta^{\prime})\right)
(16λ)((α)(β))\displaystyle(1-6\lambda)(\ell(\alpha^{\prime})-\ell(\beta^{\prime})) <2λ((α)+(β))\displaystyle<2\lambda(\ell(\alpha^{\prime})+\ell(\beta^{\prime}))
(18λ)(α)\displaystyle(1-8\lambda)\ell(\alpha^{\prime}) <(14λ)(β)\displaystyle<(1-4\lambda)\ell(\beta^{\prime})
(α)\displaystyle\ell(\alpha^{\prime}) <14λ18λ(β)<2(β)=2(β),\displaystyle<\tfrac{1-4\lambda}{1-8\lambda}\ell(\beta^{\prime})<2\ell(\beta^{\prime})=2\ell(\beta)\,,

since 0<λ<1/120<\lambda<\nicefrac{{1}}{{12}}. ∎

Lemma 2.29.

PS(MH)2((α)(α))\operatorname{PS}(M_{H})\geq 2(\ell(\alpha)-\ell(\alpha^{\prime})).

Proof.

Let I={iui=vi}I=\{i\in\mathbb{N}\mid u_{i}=v_{i}\}. By Lemma 2.23, if iIi\in I then σ(MH,ui)0modi\sigma(M_{H},u_{i})\equiv 0\mod\ell_{i}. If iIi\not\in I, then σ(MH,ui)+σ(MH,vi)0modi\sigma(M_{H},u_{i})+\sigma(M_{H},v_{i})\equiv 0\mod\ell_{i}. Note that κ(MH,ui)+κ(MH,vi)κ(MH,ui)hi\kappa(M_{H},u_{i})+\kappa(M_{H},v_{i})\geq\kappa(M_{H},u_{i})\geq h_{i} by Lemma 2.24 (b). Since hii/2h_{i}\leq\ell_{i}/2, it follows that there are at least 2hi2h_{i} faces in MHM_{H} with boundary label either ui±1u_{i}^{\pm 1} or vi±1v_{i}^{\pm 1}. If ui=viu_{i}=v_{i}, this says that κ(MH,ui)2hi\kappa(M_{H},u_{i})\geq 2h_{i}. If uiviu_{i}\neq v_{i}, this means κ(MH,ui)+κ(MH,vi)2hi\kappa(M_{H},u_{i})+\kappa(M_{H},v_{i})\geq 2h_{i}. Therefore

PS(MH)\displaystyle\operatorname{PS}(M_{H}) =rRH|r|κ(MH,r)\displaystyle=\sum_{r\in R_{H}}|r|\kappa(M_{H},r)
=iI|ui|κ(MH,ui)+iI(|ui|κ(MH,ui)+|vi|κ(MH,vi))\displaystyle=\sum_{i\in I}|u_{i}|\kappa(M_{H},u_{i})+\sum_{i\not\in I}(|u_{i}|\kappa(M_{H},u_{i})+|v_{i}|\kappa(M_{H},v_{i}))
iI|ui|κ(MH,ui)+iI|ui|(κ(MH,ui)+κ(MH,vi))\displaystyle\geq\sum_{i\in I}|u_{i}|\kappa(M_{H},u_{i})+\sum_{i\not\in I}|u_{i}|(\kappa(M_{H},u_{i})+\kappa(M_{H},v_{i}))
iI2hi|ui|+iI2hi|ui|=i2hi|ui|=2((α)(α)),\displaystyle\geq\sum_{i\in I}2h_{i}|u_{i}|+\sum_{i\not\in I}2h_{i}|u_{i}|=\sum_{i\in\mathbb{N}}2h_{i}|u_{i}|=2(\ell(\alpha)-\ell(\alpha^{\prime}))\,,

where the last equality follows from Lemma 2.24 (c). ∎

Now we are ready to prove Lemma 2.22.

Proof of Lemma 2.22.

Continuing to use the terminology and notation built up in this section, since ww is a geodesic representative of uu, |u|=(α)|u|=\ell(\alpha), and |w|=(β)|w|=\ell(\beta), it suffices to prove that (α)<3112λ(β)\ell(\alpha)<\frac{3}{1-12\lambda}\ell(\beta). By Lemmas 2.27, 2.28, and 2.29,

2((α)(α))PS(MH)\displaystyle 2(\ell(\alpha)-\ell(\alpha^{\prime}))\leq\operatorname{PS}(M_{H}) 116λ(MH)=116λ((α)+(β))\displaystyle\leq\tfrac{1}{1-6\lambda}\ell(\partial M_{H})=\tfrac{1}{1-6\lambda}\left(\ell(\alpha)+\ell(\beta)\right)
(112λ)(α)\displaystyle(1-12\lambda)\ell(\alpha) (212λ)(α)+(β)<(α)+(β)<3(β)\displaystyle\leq(2-12\lambda)\ell(\alpha^{\prime})+\ell(\beta)<\ell(\alpha^{\prime})+\ell(\beta)<3\ell(\beta)
(α)\displaystyle\ell(\alpha) <3112λ(β).\displaystyle<\tfrac{3}{1-12\lambda}\ell(\beta).

Therefore uu is (3112λ,0)\left(\tfrac{3}{1-12\lambda},0\right)-quasigeodesic, as desired. ∎

For this to be a meaningful bound we must have 0<λ<1/120<\lambda<\nicefrac{{1}}{{12}}, explaining our initial choice of λ\lambda.

3 Proof of the main result

In this section we prove the following proposition.

Proposition 3.1.

Let m,n+{}m,n\in\mathbb{Z}^{+}\cup\{\infty\} with m<nm<n. Then there exist finitely generated, recursively presented groups GG and BB such that BGB\leqslant G and

1\displaystyle 1 asdim(G)2\displaystyle\leq\operatorname{asdim}(G)\leq 2
m+1\displaystyle m+1 asdimAN(G)m+2\displaystyle\leq\operatorname{asdim_{\textnormal{AN}}}(G)\leq m+2
n+1\displaystyle n+1 asdimAN(B)n+2.\displaystyle\leq\operatorname{asdim_{\textnormal{AN}}}(B)\leq n+2\,.

Since the proof requires many auxiliary lemmas, we again ‘globally fix’ all notation in this section.

Let mm be a fixed positive integer, and let n+{}n\in\mathbb{Z}^{+}\cup\{\infty\} with m<nm<n. Let (i)(\ell_{i}) be an increasing sequence of positive integers with 02\ell_{0}\geq 2. Let SA,SBS_{A},S_{B} be disjoint finite sets, and let 0<λ<1/120<\lambda<\nicefrac{{1}}{{12}}. Suppose we have two languages

UA={uii}(SA)\displaystyle U_{A}=\{u_{i}\mid i\in\mathbb{N}\}\subset(S_{A})_{\circ}^{*} VB={vii}(SB)\displaystyle V_{B}=\{v_{i}\mid i\in\mathbb{N}\}\subset(S_{B})_{\circ}^{*}

satisfying all of the following conditions for all i,i,ji,i^{\prime},j\in\mathbb{N}.

  1. (a)

    UA,VBU_{A},V_{B} are cyclically minimal and cyclically reduced, and satisfy C(λ)C^{\prime}(\lambda).

  2. (b)

    There exists a nonempty word y(SB)y\in(S_{B})_{\circ}^{*} such that, for all hh\in\mathbb{Z}, if pp is a piece of yhy^{h} and viv_{i}, then |p|<λ|vi||p|<\lambda|v_{i}|.

  3. (c)

    2|ui||vi|2\leq|u_{i}|\leq|v_{i}|.

  4. (d)

    If ui=uiu_{i}=u_{i^{\prime}} or vi=viv_{i}=v_{i^{\prime}}, then i=ii=i^{\prime}.

  5. (e)

    The sequence of word lengths (|ui|)(|u_{i}|) is constant on blocks of the partition 𝒫m\mathcal{P}_{m} and (|vi|)(|v_{i}|) is constant on blocks of 𝒫n\mathcal{P}_{n} (see Definition 1.7).

  6. (f)

    |u(j+1)m|(j+1)m|ujm||u_{(j+1)m}|\geq\ell_{(j+1)m}|u_{jm}|. If n+n\in\mathbb{Z}^{+} then |v(j+1)n|(j+1)n|vjn||v_{(j+1)n}|\geq\ell_{(j+1)n}|v_{jn}|, and if n=n=\infty then
    |v(j+1)2|(j+1)2|vj2||v_{(j+1)^{2}}|\geq\ell_{(j+1)^{2}}|v_{j^{2}}|.

  7. (g)

    UA,VBU_{A},V_{B} are recursive.

We construct an example of languages UA,VBU_{A},V_{B} satisfying (a)-(f) in the next section, and show that they can be recursive in the process. Assuming we already have UA,VBU_{A},V_{B} satisfying (a)-(g), let HA,HBH_{A},H_{B} be given by the presentations

HA=SAUA\displaystyle H_{A}=\langle S_{A}\mid U_{A}\rangle HB=SBVB\displaystyle H_{B}=\langle S_{B}\mid V_{B}\rangle

and let A,BA,B be central extensions of HA,HBH_{A},H_{B}, respectively, defined by

A\displaystyle A =SARA:=SA[a,ui],uii:aSA,i\displaystyle=\langle S_{A}\mid R_{A}\rangle:=\langle S_{A}\mid[a,u_{i}],u_{i}^{\ell_{i}}:a\in S_{A},i\in\mathbb{N}\rangle
B\displaystyle B =SBRB:=SB[b,vi],vii:bSB,i.\displaystyle=\langle S_{B}\mid R_{B}\rangle:=\langle S_{B}\mid[b,v_{i}],v_{i}^{\ell_{i}}:b\in S_{B},i\in\mathbb{N}\rangle\,.

Since all elements in RA,RBR_{A},R_{B} represent the trivial element in HA,HBH_{A},H_{B}, respectively, there are natural epimorphisms πA:AHA\pi_{A}:A\to H_{A} and πB:BHB\pi_{B}:B\to H_{B}. Recall that for a word ww in (SA)(S_{A})_{\circ}^{*} or (SB)(S_{B})_{\circ}^{*}, we denote by w¯\bar{w} the element of AA or BB, respectively, that ww represents. Let

KA\displaystyle K_{A} =Ker(πA)=u¯i:iZ(A)\displaystyle=\operatorname{Ker}(\pi_{A})=\langle\bar{u}_{i}:i\in\mathbb{N}\rangle\leqslant Z(A)
KB\displaystyle K_{B} =Ker(πB)=v¯i:iZ(B)\displaystyle=\operatorname{Ker}(\pi_{B})=\langle\bar{v}_{i}:i\in\mathbb{N}\rangle\leqslant Z(B)

where we consider KAK_{A} as a normed group, equipped with the restriction to KAK_{A} of the word norm on AA with respect to the generating set SAS_{A}, which we will denote A\|\cdot\|_{A}: similarly for KBK_{B}.

By condition (c), there exist sequences s=(sj),t=(tj)s=(s_{j}),t=(t_{j}) such that |ui|=(m×s)i|u_{i}|=(m\times s)_{i} and |vi|=(n×t)i|v_{i}|=(n\times t)_{i} for all ii\in\mathbb{N}. Define normed groups KmK_{m}, KnK_{n} similar to the normed group defined in Lemma 1.13, as follows:

Km=i|ui|i=i(m×s)ii\displaystyle K_{m}=\bigoplus_{i\in\mathbb{N}}|u_{i}|\mathbb{Z}_{\ell_{i}}=\bigoplus_{i\in\mathbb{N}}(m\times s)_{i}\mathbb{Z}_{\ell_{i}} Kn=i|vi|i=i(n×t)ii.\displaystyle K_{n}=\bigoplus_{i\in\mathbb{N}}|v_{i}|\mathbb{Z}_{\ell_{i}}=\bigoplus_{i\in\mathbb{N}}(n\times t)_{i}\mathbb{Z}_{\ell_{i}}\,.

Suppose that xx is a word over SAS_{A} satisfying (b) with respect to UAU_{A}, except possibly the condition that xx not be the empty word. Now condition (d) guarantees that ss and tt are increasing sequences of positive integers, such that for all jj\in\mathbb{N}, sj+1sj(j+1)ms_{j+1}\geq s_{j}\ell_{(j+1)m} and tj+1tj(j+1)nt_{j+1}\geq t_{j}\ell_{(j+1)n} if n+n\in\mathbb{Z}^{+}. Condition (e) guarantees that s02s_{0}\geq 2 and t02t_{0}\geq 2. Therefore all hypotheses of Lemmas 1.9 and 1.13 are satisfied, and we have

asdimAN(|x|×Km)={m if x=εm+1 otherwise\displaystyle\operatorname{asdim_{\textnormal{AN}}}(|x|\mathbb{Z}\times K_{m})=\begin{cases}m&\text{ if }x=\varepsilon\\ m+1&\text{ otherwise}\end{cases} asdimAN(|y|×Kn)=n+1.\displaystyle\operatorname{asdim_{\textnormal{AN}}}(|y|\mathbb{Z}\times K_{n})=n+1\,.

Now KAK_{A} is abelian, KAK_{A} satisfies u¯ii=1\bar{u}_{i}^{\ell_{i}}=1 for all ii\in\mathbb{N}, and, since KAK_{A} is central in AA, we have x¯,KAx¯×KA\langle\bar{x},K_{A}\rangle\cong\langle\bar{x}\rangle\times K_{A}. All the corresponding statements hold for yy and KBK_{B}. Therefore there exist natural epimorphisms ϕA\phi_{A} and ϕB\phi_{B} defined by

ϕA:|x|×Kmx¯,KA\displaystyle\phi_{A}:|x|\mathbb{Z}\times K_{m}\to\langle\bar{x},K_{A}\rangle ϕB:|y|×Kny¯,KB\displaystyle\phi_{B}:|y|\mathbb{Z}\times K_{n}\to\langle\bar{y},K_{B}\rangle
(h,z)x¯hiu¯izi\displaystyle(h,z)\mapsto\bar{x}^{h}\prod_{i\in\mathbb{N}}\bar{u}_{i}^{z_{i}} (h,z)y¯hiv¯izi\displaystyle(h,z)\mapsto\bar{y}^{h}\prod_{i\in\mathbb{N}}\bar{v}_{i}^{z_{i}}

for all hh\in\mathbb{Z} and z=(zi)Kmz=(z_{i})\in K_{m} or KnK_{n}. In the case that x=εx=\varepsilon we have that |x|=0|x|=0 and 0={0}0\mathbb{Z}=\{0\}, so ϕA:KmKA\phi_{A}:K_{m}\to K_{A}.

Lemma 3.2.

Each of the epimorphisms ϕA,ϕB\phi_{A},\phi_{B} is bi-Lipschitz, hence each is a quasi-isometry and an isomorphism.

Proof.

We prove the statement for ϕA\phi_{A}. Let \|\cdot\| be the norm on KmK_{m}. Let hh\in\mathbb{Z} and z=(zi)Kmz=(z_{i})\in K_{m}. Let (ki)(k_{i}) be the geodesic form of zz (see Definition 1.4). Then

ϕA(h,z)A=x¯hiu¯ikiA|xhiuiki|=h|x|+i|ki||ui|=(h,z).\|\phi_{A}(h,z)\|_{A}=\left\|\bar{x}^{h}\prod_{i\in\mathbb{N}}\bar{u}_{i}^{k_{i}}\right\|_{A}\leq\left|x^{h}\prod_{i\in\mathbb{N}}u_{i}^{k_{i}}\right|=h|x|+\sum_{i\in\mathbb{N}}|k_{i}||u_{i}|=\|(h,z)\|.

Now kii/2k_{i}\leq\ell_{i}/2 for all ii\in\mathbb{N}, and xhx^{h} satisfies condition (c) of Lemma 2.22. Furthermore,

A=SA[a,ui],uii:aSA,i=SA[a,ui],uii,ui(ui)1:aSA,iA=\langle S_{A}\mid[a,u_{i}],u_{i}^{\ell_{i}}:a\in S_{A},i\in\mathbb{N}\rangle=\langle S_{A}\mid[a,u_{i}],u_{i}^{\ell_{i}},u_{i}(u_{i})^{-1}:a\in S_{A},i\in\mathbb{N}\rangle

and UAUA=UA={uii}U_{A}\cup U_{A}=U_{A}=\{u_{i}\mid i\in\mathbb{N}\} is a cyclically reduced, cyclically minimal C(λ)C^{\prime}(\lambda) language, where 2|ui||ui|2\leq|u_{i}|\leq|u_{i}| and ui=uiu_{i}=u_{i^{\prime}} implies that i=ii=i^{\prime} for all i,ii,i^{\prime}\in\mathbb{N}. Thus we may apply Lemma 2.22 with G=A,U=UA,V=UA,G=A,U=U_{A},V=U_{A}, and u~=xh\tilde{u}=x^{h}. This yields

(h,z)=h|x|+i|ui||ki|=|xhiuiki|(3112λ)x¯hiu¯ikiA=(3112λ)ϕA(h,z)A,\|(h,z)\|=h|x|+\sum_{i\in\mathbb{N}}|u_{i}||k_{i}|=\left|x^{h}\prod_{i\in\mathbb{N}}u_{i}^{k_{i}}\right|\leq\left(\frac{3}{1-12\lambda}\right)\left\|\bar{x}^{h}\prod_{i\in\mathbb{N}}\bar{u}_{i}^{k_{i}}\right\|_{A}=\left(\tfrac{3}{1-12\lambda}\right)\|\phi_{A}(h,z)\|_{A}\,,

hence (112λ3)(h,z)ϕA(k,z)A(h,z)\left(\frac{1-12\lambda}{3}\right)\|(h,z)\|\leq\|\phi_{A}(k,z)\|_{A}\leq\|(h,z)\| and ϕA\phi_{A} is bi-Lipschitz. ∎

By replacing xx or yy with ε\varepsilon, we obtain the following.

Corollary 3.3.

Both ϕA|Km:KmKA\phi_{A}|_{K_{m}}:K_{m}\to K_{A} and ϕB|Kn:KnKB\phi_{B}|_{K_{n}}:K_{n}\to K_{B} are bi-Lipschitz maps. Therefore asdimAN(KA)=asdimAN(Km)=m\operatorname{asdim_{\textnormal{AN}}}(K_{A})=\operatorname{asdim_{\textnormal{AN}}}(K_{m})=m and asdimAN(KB)=asdimAN(Kn)=n\operatorname{asdim_{\textnormal{AN}}}(K_{B})=\operatorname{asdim_{\textnormal{AN}}}(K_{n})=n.

In order to get our bounds on asdimAN(G)\operatorname{asdim_{\textnormal{AN}}}(G) and asdimAN(B)\operatorname{asdim_{\textnormal{AN}}}(B), we use the extension theorems for asymptotic and Assouad-Nagata dimension.

Lemma 3.4 (Extension Theorems).

[Bell_Dranishnikov2, Brodskiy_etal] Let

1KGH11\to K\to G\to H\to 1

be a short exact sequence, where GG and HH are finitely generated groups equipped with the word norm with respect to some finite generating set, and the norm on KK is the restriction to KK of the norm on GG. Then

asdim(G)\displaystyle\operatorname{asdim}(G) asdim(K)+asdim(H)\displaystyle\leq\operatorname{asdim}(K)+\operatorname{asdim}(H)
asdimAN(G)\displaystyle\operatorname{asdim_{\textnormal{AN}}}(G) asdimAN(K)+asdimAN(H).\displaystyle\leq\operatorname{asdim_{\textnormal{AN}}}(K)+\operatorname{asdim_{\textnormal{AN}}}(H)\,.
Lemma 3.5.

[Sledd] Let HH be a finitely generated C(1/6)C^{\prime}(\nicefrac{{1}}{{6}}) group. Then asdimAN(H)2\operatorname{asdim_{\textnormal{AN}}}(H)\leq 2.

Corollary 3.6.

We have

1asdim(A)\displaystyle 1\leq\operatorname{asdim}(A) 2\displaystyle\leq 2 1asdim(B)2\displaystyle 1\leq\operatorname{asdim}(B)\leq 2
masdimAN(A)\displaystyle m\leq\operatorname{asdim_{\textnormal{AN}}}(A) m+2\displaystyle\leq m+2 n+1asdimAN(B)n+2.\displaystyle n+1\leq\operatorname{asdim_{\textnormal{AN}}}(B)\leq n+2\,.

Also, if xεx\neq\varepsilon, then asdimAN(A)m+1\operatorname{asdim_{\textnormal{AN}}}(A)\geq m+1.

Proof.

We establish the bounds for AA: the argument for BB is similar. Since AA is finitely generated and infinite, asdimAN(A)1\operatorname{asdim_{\textnormal{AN}}}(A)\geq 1. By Corollary 3.3, asdimAN(A)asdimAN(KA)=m\operatorname{asdim_{\textnormal{AN}}}(A)\geq\operatorname{asdim_{\textnormal{AN}}}(K_{A})=m. If xεx\neq\varepsilon,

asdimAN(A)asdimAN(x¯,KA)=asdimAN(|x|×Km)=m+1\operatorname{asdim_{\textnormal{AN}}}(A)\geq\operatorname{asdim_{\textnormal{AN}}}(\langle\bar{x},K_{A}\rangle)=\operatorname{asdim_{\textnormal{AN}}}(|x|\mathbb{Z}\times K_{m})=m+1

since |x|>0|x|>0. This gives the lower bounds on the asymptotic and Assouad-Nagata dimension of AA. For the upper bounds, note that AA is constructed so that there is a short exact sequence

1KAAHA11\to K_{A}\to A\to H_{A}\to 1

where HAH_{A} is a finitely generated C(1/6)C^{\prime}(\nicefrac{{1}}{{6}}) group and hence asdim(HA)asdimAN(HA)2\operatorname{asdim}(H_{A})\leq\operatorname{asdim_{\textnormal{AN}}}(H_{A})\leq 2. Since KAK_{A} is locally finite, asdim(KA)=0\operatorname{asdim}(K_{A})=0. Now by Lemma 3.4,

asdim(A)\displaystyle\operatorname{asdim}(A) asdim(KA)+asdim(HA)2\displaystyle\leq\operatorname{asdim}(K_{A})+\operatorname{asdim}(H_{A})\leq 2
asdimAN(A)\displaystyle\operatorname{asdim_{\textnormal{AN}}}(A) asdimAN(KA)+asdimAN(HA)m+2.\displaystyle\leq\operatorname{asdim_{\textnormal{AN}}}(K_{A})+\operatorname{asdim_{\textnormal{AN}}}(H_{A})\leq m+2\,.

By Corollary 3.3, the maps ϕA|Km:KmKA\phi_{A}|_{K_{m}}:K_{m}\to K_{A} and ϕB|Kn:KnKB\phi_{B}|_{K_{n}}:K_{n}\to K_{B} are isomorphisms. Therefore the map defined by u¯iv¯i\bar{u}_{i}\mapsto\bar{v}_{i} for all ii\in\mathbb{N} extends to an isomorphism from KAK_{A} to KBK_{B}. Let ϕ:KAKB\phi:K_{A}\to K_{B} be this isomorphism. Let

G=AϕB:=ABaϕ(a)1:aA.G=A*_{\phi}B:=\langle A\sqcup B\mid a\phi(a)^{-1}:a\in A\rangle\,.

Let S=SASBS=S_{A}\sqcup S_{B}. Then GG admits the presentation

G=SRG:=S[s,ui],uii,uivi1:sS,i,G=\langle S\mid R_{G}\rangle:=\langle S\mid[s,u_{i}],u_{i}^{\ell_{i}},u_{i}v_{i}^{-1}:s\in S,i\in\mathbb{N}\rangle\,,

which is recursive if UAU_{A} and VBV_{B} are recursive. Let

H=SRH:=SASBUAVB=SASBui,vi:i.H=\langle S\mid R_{H}\rangle:=\langle S_{A}\sqcup S_{B}\mid U_{A}\sqcup V_{B}\rangle=\langle S_{A}\sqcup S_{B}\mid u_{i},v_{i}:i\in\mathbb{N}\rangle\,.

Since UAU_{A} and VBV_{B} are C(λ)C^{\prime}(\lambda) languages over disjoint alphabets, HH is a C(λ)C^{\prime}(\lambda) group. Furthermore, all words in RGR_{G} represent the trivial element of HH, so there is a natural epimorphism π:GH\pi:G\to H. Let K=Ker(π)K=\operatorname{Ker}(\pi). Then

K=u¯i:iZ(G).K=\langle\bar{u}_{i}:i\in\mathbb{N}\rangle\leqslant Z(G)\,.

We consider KK as a normed group, where the norm on KK is the restriction to KK of the word norm on GG with respect to SS. Thus we have a short exact sequence

1KGH1.1\to K\to G\to H\to 1\,.

Let bSBb\in S_{B}. Considering the relations of KK and the fact that KK is central in GG, there exists a natural epimorphism ϕK:×Kmb¯,K\phi_{K}:\mathbb{Z}\times K_{m}\to\langle\bar{b},K\rangle given by

ϕK(h,z)=b¯hiu¯izi\displaystyle\phi_{K}(h,z)=\bar{b}^{h}\prod_{i\in\mathbb{N}}\bar{u}_{i}^{z_{i}}

for all hh\in\mathbb{Z} and z=(zi)Kmz=(z_{i})\in K_{m}. Now we have the following.

Lemma 3.7.

The epimorphism ϕK\phi_{K} is bi-Lipschitz, in particular ϕK\phi_{K} is a quasi-isometry and an isomorphism.

Proof.

The proof is similar to that of Lemma 3.2. The only difference is that now we apply Lemma 2.22 with U=UA,V=VB,U=U_{A},V=V_{B}, and u~=bh\tilde{u}=b^{h}. Since bb is a word over an alphabet disjoint from SAS_{A}, clearly condition (c) of Lemma 2.22 is satisfied with u~=bh\tilde{u}=b^{h} for any hh\in\mathbb{N}. Since 2|ui||vi|2\leq|u_{i}|\leq|v_{i}| and uiviu_{i}\neq v_{i^{\prime}} for all i,ii,i^{\prime}\in\mathbb{N}, all hypotheses of Lemma 2.22 are satisfied. ∎

We are now ready to prove Proposition 3.1.

Proof of Proposition 3.1.

Let B,GB,G be defined as in this section. The bounds on asdimAN(B)\operatorname{asdim_{\textnormal{AN}}}(B) are established in Corollary 3.6. Since GG is finitely generated and infinite, asdim(G)1\operatorname{asdim}(G)\geq 1. For the lower bound on the Assouad-Nagata dimension of GG, note that

asdimAN(G)asdimAN(b¯,K)=asdimAN(×Km)=m+1.\operatorname{asdim_{\textnormal{AN}}}(G)\geq\operatorname{asdim_{\textnormal{AN}}}(\langle\bar{b},K\rangle)=\operatorname{asdim_{\textnormal{AN}}}(\mathbb{Z}\times K_{m})=m+1\,.

By Lemma 3.5, we have asdim(H)asdimAN(H)2\operatorname{asdim}(H)\leq\operatorname{asdim_{\textnormal{AN}}}(H)\leq 2. Applying the extension theorems to the short exact sequence 1KGH11\to K\to G\to H\to 1 yields that asdim(G)2\operatorname{asdim}(G)\leq 2 and asdimAN(G)m+2\operatorname{asdim_{\textnormal{AN}}}(G)\leq m+2. ∎

We give a presentation of a group GG satisfying the conditions of Proposition 3.1 in the next section. For now, we derive the main result of this paper as a corollary. To do this, we will need to recall two theorems of asymptotic dimension theory. The first a theorem of Dranishnikov and Smith, known as the Morita theorem for asymptotic Assouad-Nagata dimension. We state a special case of it here.

Theorem 3.8 (Morita theorem for asdimAN\operatorname{asdim_{\textnormal{AN}}}).

[Dranishnikov_Smith] Let GG be a finitely generated group. Then asdimAN(G×)=asdimAN(G)+1\operatorname{asdim_{\textnormal{AN}}}(G\times\mathbb{Z})=\operatorname{asdim_{\textnormal{AN}}}(G)+1.

The second is the free product formulas for asymptotic and Assouad-Nagata dimension. The theorem for asdim\operatorname{asdim} is due to Dranishnikov, and its counterpart for asdimAN\operatorname{asdim_{\textnormal{AN}}} is due to Brodskiy and Higes.

Theorem 3.9.

[Dranishnikov, Brodskiy_Higes] Let AA and BB be finitely generated groups. Then

asdim(AB)\displaystyle\operatorname{asdim}(A*B) =max{asdim(A),asdim(B),1}\displaystyle=\max\{\operatorname{asdim}(A),\operatorname{asdim}(B),1\}
asdimAN(AB)\displaystyle\operatorname{asdim_{\textnormal{AN}}}(A*B) =max{asdimAN(A),asdimAN(B),1}\displaystyle=\max\{\operatorname{asdim_{\textnormal{AN}}}(A),\operatorname{asdim_{\textnormal{AN}}}(B),1\}

We are now ready to prove the main theorem.

Theorem 3.10.

For all k,m,n{}k,m,n\in\mathbb{N}\cup\{\infty\} with 4kmn4\leq k\leq m\leq n, there exist finitely generated, recursively presented groups GG and HH with HGH\leqslant G, such that

asdim(G)\displaystyle\operatorname{asdim}(G) =k\displaystyle=k
asdimAN(G)\displaystyle\operatorname{asdim_{\textnormal{AN}}}(G) =m\displaystyle=m
asdimAN(H)\displaystyle\operatorname{asdim_{\textnormal{AN}}}(H) =n.\displaystyle=n\,.
Proof.

Applying Proposition 3.1 with m3m-3 and n2n-2, there exist finitely generated, recursively presented groups G0G_{0} and B0B_{0} with B0G0B_{0}\leqslant G_{0}, such that

1\displaystyle 1 asdim(G0)2\displaystyle\leq\operatorname{asdim}(G_{0})\leq 2
m2\displaystyle m-2 asdimAN(G0)m1\displaystyle\leq\operatorname{asdim_{\textnormal{AN}}}(G_{0})\leq m-1
n1\displaystyle n-1 asdimAN(B0)n.\displaystyle\leq\operatorname{asdim_{\textnormal{AN}}}(B_{0})\leq n\,.

Let

G1={G0×2 if asdimAN(G0)=m2G0× if asdimAN(G0)=m1.G_{1}=\begin{cases}G_{0}\times\mathbb{Z}^{2}&\text{ if }\operatorname{asdim_{\textnormal{AN}}}(G_{0})=m-2\\ G_{0}\times\mathbb{Z}&\text{ if }\operatorname{asdim_{\textnormal{AN}}}(G_{0})=m-1\,.\end{cases}

Then by the Morita theorem for Assouad-Nagata dimension, we have asdim(G1)=m\operatorname{asdim}(G_{1})=m. By the extension theorem for asymptotic dimension, we have that asdim(G1)asdim(G0)+24\operatorname{asdim}(G_{1})\leq\operatorname{asdim}(G_{0})+2\leq 4. Now let G=G1kG=G_{1}*\mathbb{Z}^{k}. Then since 4km4\leq k\leq m, by the free product formulas for asymptotic and Assouad-Nagata dimension it follows that asdim(G)=k\operatorname{asdim}(G)=k and asdimAN(G)=m\operatorname{asdim_{\textnormal{AN}}}(G)=m. Note that B0B_{0} and B0×B_{0}\times\mathbb{Z} are both subgroups of GG. Therefore, let

H={B0× if asdimAN(B0)=n1B0 if asdimAN(B0)=n.H=\begin{cases}B_{0}\times\mathbb{Z}&\text{ if }\operatorname{asdim_{\textnormal{AN}}}(B_{0})=n-1\\ B_{0}&\text{ if }\operatorname{asdim_{\textnormal{AN}}}(B_{0})=n\,.\end{cases}

Again by the Morita theorem, we have that asdimAN(H)=n\operatorname{asdim_{\textnormal{AN}}}(H)=n, and HGH\leqslant G. This completes the proof. ∎

4 A concrete example

In this section we construct an example of a group of the sort described in Proposition 3.1. In doing so, we show that such a group can be given by an explicit presentation, i.e. is recursively presented. The following lemma shows one way of constructing C(λ)C^{\prime}(\lambda) languages, which was used by Bowditch in [Bowditch2] to construct 202^{\aleph_{0}} small cancellation groups in distinct quasi-isometry classes.

Lemma 4.1.

Let U={uii}{a,x}U=\{u_{i}\mid i\in\mathbb{N}\}\subset\{a,x\}_{\circ}^{*} be a language where we define

ui=(amixmi)niu_{i}=(a^{m_{i}}x^{m_{i}})^{n_{i}}

for some positive integers mi,nim_{i},n_{i}, for each ii\in\mathbb{N}. Let k2k\geq 2 be an integer, and suppose that all of the following conditions hold.

  1. (a)

    nikn_{i}\geq k for all ii\in\mathbb{N}.

  2. (b)

    mimim_{i}\neq m_{i^{\prime}} for all distinct i,ii,i^{\prime}\in\mathbb{N}.

Then all of the following conclusions hold for all ii\in\mathbb{N}.

  1. (i)

    UU is cyclically minimal and cyclically reduced, and satisfies C(1k1)C^{\prime}\left(\frac{1}{k-1}\right).

  2. (ii)

    For all hh\in\mathbb{Z}, if pp is a piece of xhx^{h} and uiu_{i}, then |p|<1k1|ui||p|<\frac{1}{k-1}|u_{i}|.

  3. (iii)

    2|ui|2\leq|u_{i}|.

  4. (iv)

    If ui=uiu_{i}=u_{i^{\prime}} then i=ii=i^{\prime}.

Proof.

If uiUu_{i}\in U, then no cyclic shift of ui1u_{i}^{-1} is in UU: if u~i\tilde{u}_{i} is a cyclic shift of uiu_{i} that belongs to UU, then |u~i|=|ui||\tilde{u}_{i}|=|u_{i}| and u~i\tilde{u}_{i} must begin with aa and end with xx, in which case u~i=ui\tilde{u}_{i}=u_{i}. Therefore UU is cyclically minimal. Since all uiu_{i} are positive words (that is, do not contain letters a1a^{-1} or x1x^{-1}), it is clear that UU is cyclically reduced. For the same reason, when talking about pieces of some uiu_{i} and another positive word ww, it suffices to consider only cyclic shifts of uiu_{i} and ww, and we may ignore cyclic shifts of ui1u_{i}^{-1} or w1w^{-1}. To show that UU satisfies C(1k1)C^{\prime}(\frac{1}{k-1}), suppose i,ii,i^{\prime}\in\mathbb{N} are distinct. Let pp be a maximal piece of uiu_{i} and uiu_{i^{\prime}}. Since mimim_{i}\neq m_{i^{\prime}}, suppose without loss of generality that mi<mim_{i}<m_{i^{\prime}}. Then pp must have the form amixmia^{m_{i}}x^{m_{i}}. But then ni|p||ui|n_{i}|p|\leq|u_{i}| and ni|p||ui|n_{i^{\prime}}|p|\leq|u_{i^{\prime}}|. Since ni,nikn_{i},n_{i^{\prime}}\geq k, we have |p|1kmin(|ui|,|ui|)<1k1min(|ui|,|ui|)|p|\leq\frac{1}{k}\min(|u_{i}|,|u_{i^{\prime}}|)<\frac{1}{k-1}\min(|u_{i}|,|u_{i^{\prime}}|). Therefore UU satisfies C(1k1)C^{\prime}\left(\frac{1}{k-1}\right). Conclusion (ii) says only that any power of xx makes up less than 1k1\frac{1}{k-1} of a cyclic shift of some uiu_{i}. But a maximal subword of a cyclic shift of uiu_{i} of the form xhx^{h} must be xmix^{m_{i}}, which has length at most 12ni|ui|12k|ui|<1k1|ui|\frac{1}{2n_{i}}|u_{i}|\leq\frac{1}{2k}|u_{i}|<\frac{1}{k-1}|u_{i}|, so this is clear. Parts (iii) and (iv) are obvious. ∎

Lemma 4.2.

Let m+{}m\in\mathbb{Z}^{+}\cup\{\infty\}, and let 𝒫m={P(m,j)j}\mathcal{P}_{m}=\{P_{(m,j)}\mid j\in\mathbb{N}\} be the partition of \mathbb{N} given in Definition 1.7. Let k2k\geq 2 be an integer. For each ii\in\mathbb{N}, let ri=imin(P(m,j))r_{i}=i-\min(P_{(m,j)}) whenever iP(m,j)i\in P_{(m,j)}. Let (pj)(p_{j}) be an increasing sequence of positive integers. Let U={uii}{a,x}U=\{u_{i}\mid i\in\mathbb{N}\}\subset\{a,x\}_{\circ}^{*} be given by

ui=(ak(pjri)xk(pjri))k(ri+1)u_{i}=\left(a^{k^{(p_{j}-r_{i})}}x^{k^{(p_{j}-r_{i})}}\right)^{k^{(r_{i}+1)}}

whenever iP(m,j)i\in P_{(m,j)}. Let (i)(\ell_{i}) be an increasing sequence of positive integers. Suppose that the sequence (pj)(p_{j}) satisfies

pj+1pj+logk((j+1)m)+|P(m,j+1)| if m+pj+1pj+logk((j+1)2)+|P(m,j+1)| if m=.\begin{split}p_{j+1}\geq p_{j}+\log_{k}(\ell_{(j+1)m})+|P_{(m,j+1)}|\text{ if }m\in\mathbb{Z}^{+}\\ p_{j+1}\geq p_{j}+\log_{k}(\ell_{(j+1)^{2}})+|P_{(m,j+1)}|\text{ if }m=\infty\,.\end{split} (5)

Then all of the following conclusions hold for all ii\in\mathbb{N}.

  1. (i)

    UU is cyclically minimal and cyclically reduced, and satisfies C(1k1)C^{\prime}\left(\frac{1}{k-1}\right).

  2. (ii)

    For all hh\in\mathbb{N}, if p{a,x}p\in\{a,x\}_{\circ}^{*} is a piece of xhx^{h} and wiw_{i}, then |p|<1k1|ui||p|<\frac{1}{k-1}|u_{i}|.

  3. (iii)

    2|ui|2\leq|u_{i}|.

  4. (iv)

    If ui=uiu_{i}=u_{i^{\prime}}, then i=ii=i^{\prime}.

  5. (v)

    The sequence of word lengths (|ui|)(|u_{i}|) is constant on blocks of 𝒫m\mathcal{P}_{m}.

  6. (vi)

    If m+m\in\mathbb{Z}^{+} then |u(j+1)m|(j+1)m|ujm||u_{(j+1)m}|\geq\ell_{(j+1)m}|u_{jm}|, and if m=m=\infty then |u(j+1)2|(j+1)2|uj2||u_{(j+1)^{2}}|\geq\ell_{(j+1)^{2}}|u_{j^{2}}|.

Proof.

Note that, if iP(m,j)i\in P_{(m,j)}, then |ui|=2kpjrikri+1=2kpj+12|u_{i}|=2k^{p_{j}-r_{i}}k^{r_{i}+1}=2k^{p_{j}+1}\geq 2, which depends only on jj. This establishes (iv) and (v). Define the sequence s=(sj)s=(s_{j}) by

sj=2kpj+1s_{j}=2k^{p_{j}+1}

for all jj\in\mathbb{N}. Then |ui|=(m×s)i|u_{i}|=(m\times s)_{i}.

For (vi), note that logk(s(j+1)m)=logk(2)+pj+1+1\log_{k}(s_{(j+1)m})=\log_{k}(2)+p_{j+1}+1. If m+m\in\mathbb{Z}^{+}, then we have pj+1pj+logk((j+1)m)p_{j+1}\geq p_{j}+\log_{k}(\ell_{(j+1)m}), implying that sj+1(j+1)msjs_{j+1}\geq\ell_{(j+1)m}s_{j} for all jj\in\mathbb{N}. If m=m=\infty, then logk(sj+1)pj+1pj+logk((j+1)2)\log_{k}(s_{j+1})\geq p_{j+1}\geq p_{j}+\log_{k}(\ell_{(j+1)^{2}}), so sj+1(j+1)2sjs_{j+1}\geq\ell_{(j+1)^{2}}s_{j} for all jj\in\mathbb{N}. This establishes (vi).

For parts (i)-(iv), we use Lemma 4.1. Obviously part (a) of Lemma 4.1 is satisfied, so we only need to check part (b). For this is suffices to show that if iP(m,j),iP(m,j)i\in P_{(m,j)},i^{\prime}\in P_{(m,j^{\prime})}, and iii\neq i^{\prime}, then pjripjrip_{j}-r_{i}\neq p_{j^{\prime}}-r_{i^{\prime}}. If j=jj^{\prime}=j then this is immediate. If j=j+1j^{\prime}=j+1 then we have

pjri=pj+1ripj+1|Pj+1|+1>pjpjri.p_{j^{\prime}}-r_{i^{\prime}}=p_{j+1}-r_{i^{\prime}}\geq p_{j+1}-|P_{j+1}|+1>p_{j}\geq p_{j}-r_{i}\,.

This shows that pjrip_{j}-r_{i} increases with jj no matter the choice of iP(m,j)i\in P_{(m,j)}, so we are done. ∎

We are ready to construct our example. Let m,n+{}m,n\in\mathbb{Z}^{+}\cup\{\infty\} with m<nm<n. Let SA={a,x}S_{A}=\{a,x\}, SB={b,y}S_{B}=\{b,y\} be disjoint two-element alphabets. Let k=14k=14 and let i=14i\ell_{i}=14^{i} for all ii\in\mathbb{N}. Let (pj),(qj)(p_{j}),(q_{j}) be increasing sequences of positive integers. Let UA={uii}(SA)U_{A}=\{u_{i}\mid i\in\mathbb{N}\}\subset(S_{A})_{\circ}^{*} be the language constructed with respect to m,k,(i)m,k,(\ell_{i}) and (pj)(p_{j}) as in Lemma 4.2. Similarly define VB={vii}(SB)V_{B}=\{v_{i}\mid i\in\mathbb{N}\}\subset(S_{B})_{\circ}^{*} with respect to n,k,(i)n,k,(\ell_{i}), and (qj)(q_{j}).

Lemma 4.3.

Suppose that for all i,ji,j\in\mathbb{N} we have

  1. (a)

    pj+1pj+(j+2)mp_{j+1}\geq p_{j}+(j+2)m.

  2. (b)

    qj+1qj+(j+2)nq_{j+1}\geq q_{j}+(j+2)n if n+n\in\mathbb{Z}^{+}, and qj+1qj+(j+2)2q_{j+1}\geq q_{j}+(j+2)^{2} if n=n=\infty.

  3. (c)

    pi/mqi/np_{\lfloor i/m\rfloor}\leq q_{\lfloor i/n\rfloor} if n+n\in\mathbb{Z}^{+}, and pi/mqip_{\lfloor i/m\rfloor}\leq q_{\lfloor\sqrt{i}\rfloor} if n=n=\infty.

Then UA,VBU_{A},V_{B} satisfy conditions (a)-(f) listed in the proof of Proposition 3.1.

Proof.

Note that

logk((j+1)n)+|P(n,j+1)|\displaystyle\log_{k}(\ell_{(j+1)n})+|P_{(n,j+1)}| =log14(14(j+1)n)+n=(j+2)n\displaystyle=\log_{14}(14^{(j+1)n})+n=(j+2)n if n+\displaystyle\text{ if }n\in\mathbb{Z}^{+}
logk((j+1)2)+|P(n,j+1)|\displaystyle\log_{k}(\ell_{(j+1)^{2}})+|P_{(n,j+1)}| =log14(14(j+1)2)+(2j+1)(j+2)2\displaystyle=\log_{14}(14^{(j+1)^{2}})+(2j+1)\leq(j+2)^{2} if n=.\displaystyle\text{ if }n=\infty\,.

Therefore assumptions (a) and (b) guarantee that (pj)(p_{j}) and (qj)(q_{j}) satisfy (5) with respect to (i)(\ell_{i}) and m,nm,n, respectively, and so UA,VBU_{A},V_{B} satisfy all conditions listed in the proof of Proposition 3.1, except possibly that |ui||vi||u_{i}|\leq|v_{i}| for all ii\in\mathbb{N}. Now, if iP(n,j)𝒫ni\in P_{(n,j)}\in\mathcal{P}_{n}, then j=i/nj=\lfloor i/n\rfloor if n+n\in\mathbb{Z}^{+}, and j=ij=\lfloor\sqrt{i}\rfloor if n=n=\infty. It follows that assumption (c) is necessary and sufficient to guarantee that |ui||vi||u_{i}|\leq|v_{i}| for all ii\in\mathbb{N}. ∎

Example 4.4.

Let

pj\displaystyle p_{j} =m(j+2)2\displaystyle=m(j+2)^{2} qj={n2(j+3)2 if n+m(j+3)4 if n=.\displaystyle q_{j}=\begin{cases}n^{2}(j+3)^{2}&\text{ if }n\in\mathbb{Z}^{+}\\ m(j+3)^{4}&\text{ if }n=\infty.\end{cases}

Then (pj),(qj)(p_{j}),(q_{j}) satisfy the hypotheses of Lemma 4.3. The verification of this is no more than a tedious calculation, so we omit it. Note that, in the notation of Lemma 4.2,

ri={imodn if n+i2i2 if n=.r_{i}=\begin{cases}i\mod n&\text{ if }n\in\mathbb{Z}^{+}\\ i^{2}-\lfloor\sqrt{i}\rfloor^{2}&\text{ if }n=\infty\,.\end{cases}

Also, if iP(n,j)i\in P_{(n,j)}, then j=i/nj=\lfloor i/n\rfloor if n+n\in\mathbb{Z}^{+}, and j=ij=\lfloor\sqrt{i}\rfloor if n=n=\infty. So, expanding the forms of uiu_{i} and viv_{i} according to Lemma 4.2 with respect to the sequences (pj)(p_{j}) and (qj)(q_{j}) given above yields

ui\displaystyle u_{i} =(a14m(i/m+2)2(i mod m)x14m(i/m+2)2(i mod m))14(i mod m)+1\displaystyle=\left(a^{14^{m(\lfloor i/m\rfloor+2)^{2}-(i\text{ mod }m)}}x^{14^{m(\lfloor i/m\rfloor+2)^{2}-(i\text{ mod }m)}}\right)^{14^{(i\text{ mod }m)+1}}
vi\displaystyle v_{i} ={(b14n2(i/n+3)2(i mod n)y14n2(i/n+3)2(i mod n))14(i mod n)+1 if n+(b14m(i+3)4(ii2)y14m(i+3)4(ii2))14(ii2)+1 if n=.\displaystyle=\begin{cases}\left(b^{14^{n^{2}(\lfloor i/n\rfloor+3)^{2}-(i\text{ mod }n)}}y^{14^{n^{2}(\lfloor i/n\rfloor+3)^{2}-(i\text{ mod }n)}}\right)^{14^{(i\text{ mod }n)+1}}&\text{ if }n\in\mathbb{Z}^{+}\\ \left(b^{14^{m(\lfloor\sqrt{i}\rfloor+3)^{4}-(i-\lfloor\sqrt{i}\rfloor^{2})}}y^{14^{m(\lfloor\sqrt{i}\rfloor+3)^{4}-(i-\lfloor\sqrt{i}\rfloor^{2})}}\right)^{14^{(i-\lfloor\sqrt{i}\rfloor^{2})+1}}&\text{ if }n=\infty\,.\end{cases}

Then the languages {uii}\{u_{i}\mid i\in\mathbb{N}\} and {vii}\{v_{i}\mid i\in\mathbb{N}\} satisfy conditions (a)-(f) listed in the proof of Proposition 3.1, and are clearly recursive. Thus the group GG with presentation

G=a,b,x,y[a,ui],[x,ui],[b,ui],[y,ui],ui14i,uivi1:iG=\langle a,b,x,y\mid[a,u_{i}],[x,u_{i}],[b,u_{i}],[y,u_{i}],u_{i}^{14^{i}},u_{i}v_{i}^{-1}:i\in\mathbb{N}\rangle

is a finitely generated, recursively presented group of Assouad-Nagata dimension at most m+2m+2, containing a finitely generated subgroup of Assouad-Nagata dimension at least n+1n+1.

References