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

\jgccdoi

151211728 \jgccheadingLABEL:LastPageMar. 28, 2022Sept. 15, 2023

Bounding conjugacy depth functions for wreath products of finitely generated abelian groups

Michal Ferov and Mark Pengitore
Abstract.

In this article, we study the asymptotic behaviour of conjugacy separability for wreath products of abelian groups. We fully characterise the asymptotic class in the case of lamplighter groups and give exponential upper and lower bounds for generalised lamplighter groups. In the case where the base group is infinite, we give superexponential lower and upper bounds. We apply our results to obtain lower bounds for conjugacy depth functions of various wreath products of groups where the acting group is not abelian.

The first author is currently support by the Australian Research Council Laureate Fellowship FL170100032 of Professor George Willis. The second author is partially supported by NSF grant DMS-1812028 “Geometry, Topology, and Dynamics of Spaces of Non-positive Curvature”

1. Introduction

Studying infinite, finitely generated groups through their finite quotients is a common method in group theory. Groups in which one can distinguish elements using their finite quotients are called residually finite. Formally speaking, a group GG is said to be residually finite if for every pair of distinct elements f,gGf,g\in G there exists a finite group QQ and a surjective homomorphism π:GQ\pi\colon G\to Q such that π(f)π(g)\pi(f)\neq\pi(g) in QQ. Group properties of this type are called separability properties and are usually defined by what types of subsets we want to distinguish. In this article, we study conjugacy separability, meaning that we will study groups in which one can distinguish conjugacy classes using finite quotients. To be more specific, a group GG is said to be conjugacy separable if for every pair of nonconjugate elements f,gGf,g\in G there exists a finite group QQ and a surjective homomorphism π:GQ\pi\colon G\to Q such that π(f)\pi(f) is not conjugate to π(g)\pi(g) in QQ.

1.1. Motivation


One of the original reasons for studying separability properties in groups is that they provide an algebraic analogue to decision problems in finitely presented groups. To be more specific, if SGS\subseteq G is a separable subset such that SS is recursively enumerable and where one can always effectively construct the image of SS under the canonical projection onto a finite quotient of GG, then one can then decide whether a word in the generators of GG represents an element belonging to SS simply by checking finite quotients. Indeed, it was proved by Mal’tsev [16], adapting the result of McKinsey [18] to the setting of finitely presented groups, that the word problem is solvable for finitely presented, residually finite groups in the following way. Given a finite presentation XR\langle X\mid R\rangle and a word wF(X)w\in F(X) where F(X)F(X) is the free group with the generating set XX, one runs two algorithms in parallel. The first algorithm enumerates all the products of conjugates of the relators and their inverses and checks whether ww appears on the list. On the other hand, the second algorithm enumerates all finite quotients of GG and checks whether the image of the element of GG represented by ww is nontrivial. In other words, the first algorithm is looking for a witness of the triviality of ww whereas the second algorithm is looking for a witness of the nontriviality of ww. Using an analogous approach, Mostowski [20] showed that the conjugacy problem is solvable for finitely presented, conjugacy separable groups. In a similar fashion, finitely presented, LERF groups have solvable generalised word problem meaning that the membership problem is uniformly solvable for every finitely generated subgroup. In general, algorithms that involve enumerating finite quotients of an algebraic structure are sometimes called algorithms of Mal’tsev-Mostowski type or McKinsey’s algorithms.

Given an algorithm, it is natural to ask how much computing power is necessary to produce an answer. In the case of algorithms of Mal’tsev-Mostowski type, one can measure their space complexity by the associated depth functions which we go into more detail. Given a residually finite group GG with a finite generating set SS, its residual finiteness depth function RFG,S:\text{RF}_{G,S}\colon\mathbb{N}\to\mathbb{N} quantifies how deep within the lattice of normal subgroups of finite index of GG one needs to look to be able to decide whether or not a word of length at most nn represents a nontrivial element. In particular, if ww is a word in SS of length at most nn, then either GG has a finite quotient of size at most RFG,S(n)\text{RF}_{G,S}(n) in which the image of the element represented by ww is nontrivial, and if there is no finite quotient of size less than or equal to RFG,S(n)\text{RF}_{G,S}(n) in which the image of ww is nontrivial, then ww must represent the trivial word. In particular, we see that the the residual finiteness depth function of GG with respect to the generating set SS fully determines the size of finite quotients that McKinsey’s algorithm needs to generate in order to give produce an answer. Since every finite group can be fully described by its Cayley table, we see that the space complexity of the word problem of GG with respect to the generating set SS can be bounded from above by (RFG,S(n))2(\text{RF}_{G,S}(n))^{2}. Moreover, the notion of depth function can be generalised to different separability properties. In this note, we study conjugacy separability depth functions which denote as ConjG,S(n)\operatorname{Conj}_{G,S}(n) which is a function that measures how deep within the lattice of normal subgroups of finite index one needs to go in order to be able to distinguish distinct conjugacy classes of elements of word length at most nn with respect to the finite generating subset SS. Just like in computational complexity, we study these functions up to asymptotic equivalence. See subsection 2.1 for the precise definitions of depth functions and the corresponding asymptotic notions.

1.2. Statement of the results


Not much is known about the asymptotic behaviour of the function ConjG,S(n)\operatorname{Conj}_{G,S}(n) for different classes of groups. The first result of this kind was by Lawton, Louder, and McReynolds [14] who showed that if GG is a nonabelian free group or the fundamental group of a closed oriented surface of genus g2g\geq 2, then ConjG(n)nn2\operatorname{Conj}_{G}(n)\preceq n^{n^{2}}. For the class of finitely generated nilpotent groups, the second named author and Deré [7] showed that if GG is a finite extension of a finitely generated abelian group, then ConjG(n)(log(n))d\operatorname{Conj}_{G}(n)\preceq(\log(n))^{d} for some natural number dd, and when GG is a finite extension of a finitely generated nilpotent group that is not virtually abelian, then there exist natural numbers d1d_{1} and d2d_{2} such that nd1ConjG(n)nd2n^{d_{1}}\preceq\operatorname{Conj}_{G}(n)\preceq n^{d_{2}}. Finally, in [10], the authors of this note gave upper bounds for ConjAB(n)\operatorname{Conj}_{A\wr B}(n) of wreath products of conjugacy separable groups AA and BB which generalises Remeslennikov’s classification of conjugacy separable wreath products [21]. However, when applied directly to wreath products of abelian groups, the formulas given in [10] produce rather coarse upper bounds. Applying [10, Theorem C] to the lamplighter group 𝔽2\mathbb{F}_{2}\wr\mathbb{Z}, one can then demonstrate that its conjugacy depth function can be bounded from above by the function 2nn22^{n^{n^{2}}}. Similarly, applying [10, Theorem C] to the group \mathbb{Z}\wr\mathbb{Z} one can show that the conjugacy depth function is bounded from above by nnn2n^{n^{n^{2}}}. In this note, we only focus on conjugacy depth functions of wreath products of finitely generated abelian groups. This restriction allows us to use more effective methods to obtain much better upper bounds than those presented in [10]. Additionally, we are able to use methods from commutative algebra to produce lower bounds and in the case of the lamplighter group, we fully determine the asymptotic equivalence class of its conjugacy depth function.

Before stating our main results, we introduce some notation. Letting f,g:f,g\colon\mathbb{N}\to\mathbb{N} be non-decreasing functions, we write fgf\preceq g if there is a constant CC\in\mathbb{N} such that f(n)Cg(Cn)f(n)\leq Cg(Cn) for all nn\in\mathbb{N}. If fgf\preceq g and gfg\preceq f, we then write fgf\approx g.

This first theorem addresses the asymptotic behavior of conjugacy separability of wreath products of the form ABA\wr B where AA is a finite abelian group and BB is an infinite, finitely generated abelian group. For the statements of our theorem, we say the torsion free rank of a finitely generated abelian group AA is the largest such natural number kk such that AkTor(A)A\cong\mathbb{Z}^{k}\oplus\operatorname{Tor}(A) where Tor(A)\operatorname{Tor}(A) is the subgroup of finite order elements of AA.

Theorem 1.1.

Let AA be a finite abelian group, and suppose that BB is an infinite finitely generated abelian group. If the torsion free rank of BB is 11, then

ConjAB(n)2n.\operatorname{Conj}_{A\wr B}(n)\approx 2^{n}.

If the torsion free rank of BB is k2k\geq 2, then

2nConjAB(n)2n2k.2^{n}\preceq\operatorname{Conj}_{A\wr B}(n)\preceq 2^{n^{2k}}.

As a corollary, we are able to compute the precise asymptotic behaviour of conjugacy separability for lamplighter groups.

Corollary 1.2.

Let 𝔽q\mathbb{F}_{q} be the finite field with qq elements. Then

Conj𝔽q(n)2n.\operatorname{Conj}_{\mathbb{F}_{q}\wr\mathbb{Z}}(n)\approx 2^{n}.

This next theorem addresses the asymptotic behavior of conjugacy separability of ABA\wr B when AA and BB are both infinite, finitely generated abelian groups.

Theorem 1.3.

Let AA be an infinite, finitely generated abelian group, and suppose that BB is an infinite finitely generated abelian group. If BB has torsion free rank 11, then

(log(n))nConjAB(n)(log(n))n2.(\log(n))^{n}\preceq\operatorname{Conj}_{A\wr B}(n)\preceq(\log(n))^{n^{2}}.

If BB has torsion free rank k>1k>1, then

(log(n))nConjAB(n)(log(n))n2k+2.(\log(n))^{n}\preceq\operatorname{Conj}_{A\wr B}(n)\preceq(\log(n))^{n^{2k+2}}.

By combining Theorem 1.1 and Theorem 1.3, the following corollary gives the best known result for asymptotic behaviour of conjugacy separability of wreath products of finitely generated abelian groups with an infinite acting group.

Corollary 1.4.

Let AA be a finitely generated abelian group, and suppose that BB is an infinite, finitely generated abelian group.

Suppose that AA is a finitely generated abelian group. If BB has torsion free rank 11, then

(log(n))nConjAB(n)(log(n))n2.(\log(n))^{n}\preceq\operatorname{Conj}_{A\wr B}(n)\preceq(\log(n))^{n^{2}}.

If BB has torsion free rank k>1k>1, then

(log(n))nConjAB(n)(log(n))n2k+2.(\log(n))^{n}\preceq\operatorname{Conj}_{A\wr B}(n)\preceq(\log(n))^{n^{2k+2}}.

Suppose that AA is finite. If BB has torsion free rank 11, then

ConjAB(n)2n\operatorname{Conj}_{A\wr B}(n)\approx 2^{n}

If BB has torsion free rank k>1k>1, then

2nConjAB(n)2n2k.2^{n}\preceq\operatorname{Conj}_{A\wr B}(n)\preceq 2^{n^{2k}}.

This last theorem applies Theorem 1.1 and Theorem 1.3 to provide exponential lower bounds for conjugacy separable wreath products ABA\wr B where Z(B)\mathbb{Z}\leq Z(B) or where BB has an infinite cyclic subgroup as a retract.

Theorem 1.5.

Let AA be a nontrivial finitely generated abelian group, and suppose that GG is a conjugacy separable finitely generated group with separable cyclic subgroups that contains an infinite cyclic group as a retract or satisfies Z(B)\mathbb{Z}\leq Z(B). If AA is finite, then

2nConjAG(n).2^{n}\preceq\operatorname{Conj}_{A\wr G}(n).

Otherwise,

(log(n))nConjAG(n).(\log(n))^{n}\preceq\operatorname{Conj}_{A\wr G}(n).

1.3. Outline of the paper


In Section 2, we recall standard mathematical notions and concepts that will be used throughout the paper. In particular, in subsection 2.1, we recall the notions of word length, depth functions and associated asymptotic notions. In subsection 2.2, we recall the basic terminology of wreath products of groups. Finally, in subsection 2.3, we recall the notion of Laurent polynomial rings and show that groups of the form RR\wr\mathbb{Z}, were RR is a commutative ring, can be realised as R[x,x1]R[x,x^{-1}]\rtimes\mathbb{Z} where R[x,x1]R[x,x^{-1}] is the ring of Laurent polynomials over the RR and \mathbb{Z} acts on R[x,x1]R[x,x^{-1}] via multiplication by xx. We finish this section by giving a criterion for conjugacy for such groups purely in terms of commutative algebra.

In Section 3 we use methods from commutative algebra to produce lower bounds for the conjugacy depth functions by constructing infinite sequences of pairs of non-conjugate elements that require quotients that are at least exponentially large in the word lengths of the nonconjugate pair of elements in order to remain non-conjugate.

In Section 4, we use combinatorial methods together with the conjugacy criterion for wreath products of abelian groups to construct upper bounds for wreath products of abelian groups.

Finally, in Section 5 we combine the lower bounds obtained in Section 3 together with the upper bounds constructed in Section 4 to prove Theorem 1.5. We then proceed to apply our methods to give lower bounds on the conjugacy depth function for wreath products where the acting group may not necessarily be abelian.

2. Preliminaries

We denote 𝔽p\mathbb{F}_{p} as the finite field of pp elements where pp is prime. We denote Sym(n)\operatorname{Sym}(n) as the symmetric group on nn letters. For x,yGx,y\in G, we write xGyx\sim_{G}y if there exists an element zGz\in G such that zxz1=yzxz^{-1}=y and suppress the subscript when the group GG is clear from context. Whenever the given group is abelian, we will use additive notation.

We say that a subgroup HGH\leq G is conjugacy embedded in GG if for every f,gHf,g\in H we have that fHgf\sim_{H}g if and only if fGgf\sim_{G}g. Following the definition, one can easily check that the relation of being conjugacy embedded is transitive. That means if ABCA\leq B\leq C such that AA is conjugacy embedded in BB and BB is conjugacy embedded in CC, then AA is conjugacy embedded in CC.

Given a group GG, we say that a subgroup RGR\leq G is a retract of GG if there exists a surjective homomorphism ρ:GR\rho\colon G\to R such that ρR=idR\rho\restriction_{R}=\operatorname{id}_{R}. The following remark is a natural consequence of the definition of being a retract.

Remark 2.1.

Let GG be a group, and let RGR\leq G be a subgroup. If RR is a retract of GG, then RR is conjugacy embedded in GG

The next lemma allows us to reduce the study of conjugacy in a semidirect product of abelian groups ABA\rtimes B to conjugacy in A(B/K)A\rtimes(B/K) where KK is the kernel of the action of BB on AA. Since wreath products are a special type of a semidirect product, this lemma will be useful throughout the article. Finally, in this lemma, we will be using additive notation, with BB acting on AA by multiplication, i.e. for b1,b2Bb_{1},b_{2}\in B and aAa\in A we write

b1b2a1\displaystyle b_{1}\cdot b_{2}\cdot a_{1} =(b1+b2)a1,\displaystyle=(b_{1}+b_{2})\cdot a_{1},
0a\displaystyle 0\cdot a =a\displaystyle=a
b1(a)\displaystyle b_{1}\cdot(-a) =(b1a)=b1a.\displaystyle=-(b_{1}\cdot a)=-b_{1}\cdot a.

In particular, for a1,a2Aa_{1},a_{2}\in A and b1,b2Bb_{1},b_{2}\in B, we write

(a1,b1)(a2,b2)\displaystyle(a_{1},b_{1})(a_{2},b_{2}) =(a1+b1a2,b1+b2),\displaystyle=(a_{1}+b_{1}\cdot a_{2},b_{1}+b_{2}),
(a1,b1)(a2,b2)\displaystyle(a_{1},b_{1})(a_{2},b_{2}) =(a1+b1a2,b1+b2),\displaystyle=(a_{1}+b_{1}\cdot a_{2},b_{1}+b_{2}),
(a1,b1)1\displaystyle(a_{1},b_{1})^{-1} =((b1)a1,b1)=((b1)(a1),b1)\displaystyle=(-(-b_{1})\cdot a_{1},-b_{1})=((-b_{1})(-a_{1}),-b_{1})
Lemma 2.2.

Suppose that AA and BB are finitely generated abelian groups, and suppose that a1,a2Aa_{1},a_{2}\in A and bBb\in B. Then

(a1,b)AB(a2,b) if and only if (a1,b¯)A(B/K)(a2,b¯),(a_{1},b)\sim_{A\rtimes B}(a_{2},b)\quad\text{ if and only if }\quad(a_{1},\bar{b})\sim_{A\rtimes(B/K)}(a_{2},\bar{b}),

where KK is the kernel of the action of BB on AA.

Proof.

For this proof, we denote the action of bBb\in B on aAa\in A as bab\cdot a. We note that if (a1,b)AB(a2,b)(a_{1},b)\sim_{A\rtimes B}(a_{2},b), then clearly (a1,b¯)A(B/K)(a2,b¯).(a_{1},\bar{b})\sim_{A\rtimes(B/K)}(a_{2},\bar{b}). Thus, we may assume that (a1,b¯)A(B/K)(a2,b¯)(a_{1},\bar{b})\sim_{A\rtimes(B/K)}(a_{2},\bar{b}). Suppose that there exists (x,y)AB(x,y)\in A\rtimes B and kKk\in K such that

(x,y)(a1,b)(x,y)1=(0,k)(a2,b).(x,y)(a_{1},b)(x,y)^{-1}=(0,k)(a_{2},b).

Thus, we have

(x,y)(a1,b)(x,y)1\displaystyle(x,y)(a_{1},b)(x,y)^{-1} =\displaystyle= (0,k)(a2,b)\displaystyle(0,k)(a_{2},b)
(x+ya1,y+b)((y)(x),y)\displaystyle(x+y\cdot a_{1},y+b)((-y)\cdot(-x),-y) =\displaystyle= (ka2,k+b)\displaystyle(k\cdot a_{2},k+b)
(x+ya1+(y+b)(y)(x),y+by)\displaystyle(x+y\cdot a_{1}+(y+b)\cdot(-y)\cdot(-x),y+b-y) =\displaystyle= (a2,k+b)\displaystyle(a_{2},k+b)
(x+ya1+(y+by)(x),y+by)\displaystyle(x+y\cdot a_{1}+(y+b-y)\cdot(-x),y+b-y) =\displaystyle= (a2,k+b)\displaystyle(a_{2},k+b)
(x+ya1bx,b)\displaystyle(x+y\cdot a_{1}-b\cdot x,b) =\displaystyle= (a2,k+b).\displaystyle(a_{2},k+b).

Hence, we must have that k=0k=0. Therefore, we have (x,y)(a1,b)(x,y)1=(a2,b)(x,y)(a_{1},b)(x,y)^{-1}=(a_{2},b) giving our claim. ∎

Given an abelian group BB, we will use Tor(B)\operatorname{Tor}(B) to denote

Tor(B)={bB:bm=1 for some m}.\operatorname{Tor}(B)=\{b\in B:b^{m}=1\text{ for some }m\in\mathbb{Z}\}.

When BB is a finitely generated, it it easy to see that Tor(B)\operatorname{Tor}(B) is a characteristic subgroup which provides a splitting

B=kTor(B)B=\mathbb{Z}^{k}\oplus\operatorname{Tor}(B)

for some kk\in\mathbb{N}. We say that kk is the torsion-free rank of BB. By fixing a splitting, we define τ:BTor(B)\tau\colon B\to\operatorname{Tor}(B) and ϕ:Bk\phi\colon B\to\mathbb{Z}^{k} as the associated retractions. Then every element bBb\in B can be uniquely expressed as b=τ(b)+ϕ(b)b=\tau(b)+\phi(b). We say that τ(b)\tau(b) is the torsion part of bb and ϕ(b)\phi(b) is the torsion-free part of bb. Whenever we say the torsion part or torsion-free part of an element of BB, we are saying that with respect to some fixed splitting of the above form.

To ease notation, we will view direct sums of groups over some indexing set as finitely supported functions on the indexing set with range in the index groups. More precisely, if

G=iIAG=\bigoplus_{i\in I}A

is a direct sum of copies of a group AA indexed by a set II, then for fGf\in G we will write f(i)f(i) to denote the ii-th coordinate of ff. In particular, elements in GG correspond to functions f:Af\colon\mathcal{I}\to A where f(i)=1f(i)=1 for all but finitely many elements in \mathcal{I}. The support of ff which is the set of elements on which ff is not trivial will be denoted as

supp(f)={iIf(i)1}.\operatorname{supp}(f)=\{i\in I\mid f(i)\neq 1\}.

The range of ff will be denoted as

rng(f)={f(i)iI}.\operatorname{rng}(f)=\{f(i)\mid i\in I\}.

2.1. Asymptotic notions and depth functions


Let GG be a finitely generated group equipped with a finite generating subset SS. We define the word length of an element gGg\in G with respect to SS as

gS=min{|w|wF(S) and w=Gg}.\|g\|_{S}=\min\{|w|\mid w\in F(S)\mbox{ and }w=_{G}g\}.

where |w||w| denotes the word length of ww in F(S).F(S). We use BG,S(n)\operatorname{B}_{G,S}(n) to denote the ball of radius nn centered around the identity with respect to the finite generating subset SS. When the finite generating subset is clear from context, we will instead write BG(n).\operatorname{B}_{G}(n).

The conjugacy separability depth function of GG is defined in the following way. Let f,gGf,g\in G be a pair of elements such that f≁Ggf\not\sim_{G}g. The conjugacy depth of the pair (f,g)(f,g), denoted CDG(f,g)\operatorname{CD}_{G}(f,g), is given by

CDG(f,g)=min{|G/N||Nf.i.G and fN≁G/NgN}\operatorname{CD}_{G}(f,g)=\min\{|G/N|\>|\>N\trianglelefteq_{f.i.}G\mbox{ and }fN\not\sim_{G/N}gN\}

with the understanding that CDG(f,g)=\operatorname{CD}_{G}(f,g)=\infty if no such finite quotient exits. Similar to the definition of residual finiteness, we say that GG is conjugacy separable if CDG(g,h)<\operatorname{CD}_{G}(g,h)<\infty for all f,hGf,h\in G such that fGg.f\nsim_{G}g. Given a finite generating subset SGS\subseteq G for a conjugacy separable group GG, the conjugacy separability depth function ConjG,S:\operatorname{Conj}_{G,S}\colon\mathbb{N}\to\mathbb{N} is defined as

ConjG,S(n)=max{CDG(f,g)f,gBG,S(n) and f≁Gg}.\operatorname{Conj}_{G,S}(n)=\operatorname{max}\{\operatorname{CD}_{G}(f,g)\mid f,g\in\operatorname{B}_{G,S}(n)\mbox{ and }f\not\sim_{G}g\}.

We note that ConjG,S(n)\operatorname{Conj}_{G,S}(n) depends on the choice of the finite generating subset SS. However, one can easily check that the asymptotic behaviour does not. It is well known that a change of a finite generating subset is a quasi-isometry. In particular, if S1,S2GS_{1},S_{2}\subset G are two finite generating subsets of a group GG, then S1S2\|\cdot\|_{S_{1}}\approx\|\cdot\|_{S_{2}}. The same holds for depth functions. For non-decreasing functions f,g:f,g\colon\mathbb{N}\to\mathbb{N}, we write fgf\preceq g if there is a constant CC\in\mathbb{N} such that f(n)Cg(Cn)f(n)\leq Cg(Cn) for all nn\in\mathbb{N}, and if fgf\preceq g and gfg\preceq f, we then write fgf\approx g. When GG is conjugacy separable, we have ConjG,S1(n)ConjG,S2(n)\operatorname{Conj}_{G,S_{1}}(n)\approx\operatorname{Conj}_{G,S_{2}}(n); see [14] for more details. As we are only interested in the asymptotic behaviour of the above defined functions, we will suppress the choice of generating subset whenever we reference the depth functions or the word-length.

Let GG be a conjugacy separable group with a finitely generated conjugacy embedded subgroup RR. We now relate the conjugacy depth of a pair of nonconjugate elements r1,r2Rr_{1},r_{2}\in R as elements of RR with the conjugacy depth of r1,r2r_{1},r_{2} as elements of GG.

Lemma 2.3.

Let GG be a finitely generated conjugacy separable group with a finitely generated conjugacy embedded subgroup RGR\leq G. Then for r1,r2Rr_{1},r_{2}\in R where r1Rr2r_{1}\nsim_{R}r_{2}, we have

CDR(r1,r2)CDG(r1,r2).\operatorname{CD}_{R}(r_{1},r_{2})\leq\operatorname{CD}_{G}(r_{1},r_{2}).
Proof.

Suppose that r1Rr2r_{1}\nsim_{R}r_{2} for r1,r2R.r_{1},r_{2}\in R. Since RR is conjugacy embedded into GG, we have that r1Gr2r_{1}\nsim_{G}r_{2}. hus, we will show that CDR(r1,r2)CDG(r1,r2)\operatorname{CD}_{R}(r_{1},r_{2})\leq\operatorname{CD}_{G}(r_{1},r_{2}). Suppose that NRf.iGN_{R}\trianglelefteq_{f.i}G realises CDG(r1,r2)\operatorname{CD}_{G}(r_{1},r_{2}). Since φ|R:Rφ(R)G/NG\varphi|_{R}\colon R\to\varphi(R)\leq G/N_{G}, we see that φ|R(r1)φ|R(r2).\varphi|_{R}(r_{1})\nsim\varphi|_{R}(r_{2}). Hence, we have

CDR(r1,r2)|φ(R)||G/NG|=CDG(r1,r2).\operatorname{CD}_{R}(r_{1},r_{2})\leq|\varphi(R)|\leq|G/N_{G}|=\operatorname{CD}_{G}(r_{1},r_{2}).\qed

We conclude this subsection with the following lemma which relates ConjG(n)\operatorname{Conj}_{G}(n) with ConjR(n)\operatorname{Conj}_{R}(n) where RR is a retract of a finitely generated conjugacy separable group GG.

Lemma 2.4.

Suppose that GG is a finitely generated conjugacy separable group with a finitely generated subgroup RGR\leq G such that RR is a retract. Then RR is conjugacy separable and conjugacy embedded into GG. Moreover, we have

ConjR(n)ConjG(n).\operatorname{Conj}_{R}(n)\preceq\operatorname{Conj}_{G}(n).
Proof.

Let ρ:GR\rho\colon G\to R be the corresponding retraction. We start by showing there is a finite generating set XGX\subseteq G such that X=XR˙XKX=X_{R}\dot{\cup}X_{K}, R=XR,R=\langle X_{R}\rangle, and XKR={1}\langle X_{K}\rangle\cap R=\{1\}. Suppose that G=XG=\langle X^{\prime}\rangle, where X={x1,,xm}X=\{x_{1},\dots,x_{m}\}. We set XR={ρ(x1),,ρ(xm)}X_{R}=\{\rho(x_{1}),\dots,\rho(x_{m})\} and XK={ρ(x1)1x1,,ρ(xm)1xm}X_{K}=\{\rho(x_{1})^{-1}x_{1},\dots,\rho(x_{m})^{-1}x_{m}\}. It is straightforward to see that R=XRR=\langle X_{R}\rangle and XKker(ρ)\langle X_{K}\rangle\leq\ker(\rho), so XKR={1}\langle X_{K}\rangle\cap R=\{1\}. We also have rXR=rX\|r\|_{X_{R}}=\|r\|_{X} for every rRr\in R.

Now suppose that r1,r2BR(n)r_{1},r_{2}\in\operatorname{B}_{R}(n) satisfy r1≁Rr2r_{1}\not\sim_{R}r_{2}. Following the previous paragraph together with remark 2.1, we see that r1,r2BG(n)r_{1},r_{2}\in\operatorname{B}_{G}(n) and that r1≁Gr2r_{1}\not\sim_{G}r_{2}. By Lemma 2.3, we see that

CDR(r1,r2)CDG(r1,r2).\operatorname{CD}_{R}(r_{1},r_{2})\leq\operatorname{CD}_{G}(r_{1},r_{2}).

We note that this inequality holds for all r1,r2Rr_{1},r_{2}\in R where r1Rr2r_{1}\nsim_{R}r_{2}, and since BR(n)BG(n)B_{R}(n)\subset B_{G}(n), it follows that

ConjR(n)ConjG(n).\operatorname{Conj}_{R}(n)\leq\operatorname{Conj}_{G}(n).\qed

2.2. Wreath products


For groups AA and BB, we denote the restricted wreath product of AA and BB, written as ABA\wr B, by

AB=(bBA)B.A\wr B=\left(\bigoplus_{b\in B}A\right)\rtimes B.

where BB acts on bBA\bigoplus_{b\in B}A via left multiplication on the coordinates. An element fbBAf\in\bigoplus_{b\in B}A is understood as a function f:BAf\colon B\to A such that f(b)1f(b)\neq 1 for only finitely many bBb\in B. With a slight abuse of notation, we will use ABA^{B} to denote bBA\bigoplus_{b\in B}A. The action of BB on ABA^{B} is then realised as bf(x)=f(bx)b\cdot f(x)=f(bx).

Following the given notation, if HAH\leq A and KBK\leq B, we will use HKH^{K} to denote the subset

HK={fABsupp(f)K and rng(f)H}.H^{K}=\{f\in A^{B}\mid\operatorname{supp}(f)\subseteq K\mbox{ and }\operatorname{rng}(f)\subseteq H\}.

Keeping this notation in mind, the wreath product HKH\wr K can then be naturally identified with the subgroup HKKABH^{K}\rtimes K\leq A\wr B.

Lemma 2.5.

Let A,BA,B be finitely generated abelian groups, and suppose that RAAR_{A}\leq A and RBBR_{B}\leq B are retracts. Then the group R=RARBR=R_{A}\wr R_{B} is a retract of G=ABG=A\wr B. In particular, RR is conjugacy embedded in GG and ConjR(n)ConjG(n)\operatorname{Conj}_{R}(n)\preceq\operatorname{Conj}_{G}(n).

Proof.

Let ρA:ARA\rho_{A}\colon A\to R_{A} and ρB:BRB\rho_{B}\colon B\to R_{B} be the associated retraction maps. We define a map ρ:ABRARB\rho\colon A\wr B\to R_{A}\wr R_{B} in a following way: given fAB,bBf\in A^{B},b\in B we set ρ((f,b))=ρ(f)ρB(b)\rho((f,b))=\rho(f)\rho_{B}(b) where ρ(f)\rho(f) is a function in RARBR_{A}^{R_{B}} defined as

ρ(f)(xRB)=ρA(yxRbf(y)).\rho(f)(xR_{B})=\rho_{A}\left(\prod_{y\in xR_{b}}f(y)\right).

We see that ρ\rho is a surjective homomorphism and ρ|R=idR\rho|_{R}=\operatorname{id}_{R}, and thus, it follows that RR is a retract of ABA\wr B. We finish by noting that Remark 2.1 implies RR is conjugacy embedded in ABA\wr B. ∎

Suppose that bBb\in B and fABf\in A^{B} is a function with a finite support. We say that ff is minimal with respect to bb if all elements of supp(f)\operatorname{supp}(f) lie in distinct cosets of b\langle b\rangle in BB. We will say that an element fbABfb\in A\wr B is reduced if ff is minimal with respect to bb.

The following lemma is a special case of [10, Lemma 5.13].

Lemma 2.6.

Let AA, BB be finitely generated groups, and suppose that bBb\in B and f:BAf\colon B\to A are given such that fbBAB(n)fb\in\operatorname{B}_{A\wr B}(n). Then there exists a constant CC independent of bb and ff and fABf^{\prime}\in A^{B} such that the following hold:

  1. (1)

    fbfbf^{\prime}b\sim fb

  2. (2)

    fbf^{\prime}b is reduced

  3. (3)

    fbCfb\|f^{\prime}b\|\leq C\|fb\|.

The following statement and its proof which provides a conjugacy criterion for wreath products of abelian groups follows from [10, Lemma 5.14].

Lemma 2.7.

Let A,BA,B be abelian groups, and let G=ABG=A\wr B be their wreath product. Let f1,f2ABf_{1},f_{2}\in A^{B}, b1,b2Bb_{1},b_{2}\in B be such that the elements f1b1f_{1}b_{1} and f2b2f_{2}b_{2} are reduced. Then

f1b1Gf2b2 if and only if b1=b2 and f1b(f2b)B.f_{1}b_{1}\sim_{G}f_{2}b_{2}\quad\text{ if and only if }\quad b_{1}=b_{2}\quad\text{ and }\quad f_{1}b\in(f_{2}b)^{B}.

In particular, there exists an element cBc\in B such that

csupp(f1)=supp(f2) and f1(cx)=f2(x)c\operatorname{supp}(f_{1})=\operatorname{supp}(f_{2})\quad\text{ and }\quad f_{1}(cx)=f_{2}(x)

for all xBx\in B.

One interpretation of Lemma 2.7 is that by ensuring that we are only working with reduced elements of ABA\wr B, we only need to worry about them being conjugate by an element from BB.

Let AA be a finite abelian group and let BB be a finitely generated abelian group of torsion free rank at least 11. This next lemma allows to reduce the study of asymptotic lower bounds for conjugacy separability of groups of the form ABA\wr B to that of groups of the form 𝔽p.\mathbb{F}_{p}\wr\mathbb{Z}.

Lemma 2.8.

Let AA be a finite abelian group where p|A|p\mid|A|, and let BB be an infinite, finitely generated abelian group. The group 𝔽p\mathbb{F}_{p}\wr\mathbb{Z} is conjugacy embedded in the group ABA\wr B and

Conj𝔽p(n)ConjAB(n).\operatorname{Conj}_{\mathbb{F}_{p}\wr\mathbb{Z}}(n)\preceq\operatorname{Conj}_{A\wr B}(n).
Proof.

Since \mathbb{Z} is a retract of BB, we have that AA\wr\mathbb{Z} is a retract of AB.A\wr B. We then have by Lemma 2.4 that ConjA(n)ConjAB(n)\operatorname{Conj}_{A\wr\mathbb{Z}}(n)\preceq\operatorname{Conj}_{A\wr B}(n). Thus, we may assume that B.B\cong\mathbb{Z}.

We now demonstrate that 𝔽p\mathbb{F}_{p}\wr\mathbb{Z} is conjugacy embedded into A.A\wr\mathbb{Z}. Suppose that =b.\mathbb{Z}=\left<b\right>. Let f1,f2:Af_{1},f_{2}\colon\mathbb{Z}\to A be given such that f1bs1,f2bs2𝔽pf_{1}b^{s_{1}},f_{2}b^{s_{2}}\in\mathbb{F}_{p}\wr\mathbb{Z} satisfy f1bs1Aat2bs2f_{1}b^{s_{1}}\sim_{A\wr\mathbb{Z}}a^{t_{2}}b^{s_{2}} where s1,s2.s_{1},s_{2}\in\mathbb{Z}. Moreover, we may assume are both reduced. We claim that f1bs1𝔽pf2bs2f_{1}b^{s_{1}}\sim_{\mathbb{F}_{p}\wr\mathbb{Z}}f_{2}b^{s_{2}}. Since f1bs1Af2bs2f_{1}b^{s_{1}}\sim_{A\wr\mathbb{Z}}f_{2}b^{s_{2}}, we must have bs1bs2b^{s_{1}}\sim_{\mathbb{Z}}b^{s_{2}}, and given that \mathbb{Z} is abelian, we then have s=s1=s2.s=s_{1}=s_{2}. By Lemma 2.7, we have there exists a btb^{t} such that btsupp(f1)=supp(f2)b^{t}\cdot\operatorname{supp}(f_{1})=\operatorname{supp}(f_{2}) and f1(btx)=f2(x)f_{1}(b^{t}x)=f_{2}(x). However, that is equivalent to f1bsf_{1}b^{s} and f2bsf_{2}b^{s} being conjugate in 𝔽p\mathbb{F}_{p}\wr\mathbb{Z} as desired.

For the second part of the statement, we first show that 𝔽p\mathbb{F}_{p}\wr\mathbb{Z} can be realised as an undistorted subgroup of ABA\wr B. If /pe=a\mathbb{Z}/p^{e}\mathbb{Z}=\langle a\rangle, we then see that /p=ape1\mathbb{Z}/p\mathbb{Z}=\langle a^{p^{e-1}}\rangle. Letting Xpe={a,ape1b}LpeX_{p_{e}}=\{a,a^{p^{e-1}}b\}\subseteq L_{p^{e}} and Xp={ape1,b}LpX_{p}=\{a^{p^{e-1}},b\}\subseteq L_{p}, it then follows that Lpe=XpeL_{p^{e}}=\langle X_{p_{e}}\rangle and Lp=XpL_{p}=\langle X_{p}\rangle. One can easily check that for any xLpx\in L_{p} that xXp=xXpe\|x\|_{X_{p}}=\|x\|_{X_{p^{e}}}, and subsequently, BLp,Xp(n)BLpe,Xpe(n)\operatorname{B}_{L_{p},X_{p}}(n)\subseteq\operatorname{B}_{L_{p^{e}},X_{p^{e}}}(n).

Now suppose that x,yB𝔽p(n)x,y\in\operatorname{B}_{\mathbb{F}_{p}\wr\mathbb{Z}}(n) are not conjugate. We then have that f,gBA(n)f,g\in\operatorname{B}_{A\wr\mathbb{Z}}(n), and since 𝔽p\mathbb{F}_{p}\wr\mathbb{Z} is conjugacy embedded into ABA\wr B, Lemma 2.3 implies

CD𝔽p(x,y)CDAB(x,y).\operatorname{CD}_{\mathbb{F}_{p}\wr\mathbb{Z}}(x,y)\leq\operatorname{CD}_{A\wr B}(x,y).

As a consequence of the above inequality, and the previous paragraph, we see that

Conj𝔽p(n)ConjA(n).\operatorname{Conj}_{\mathbb{F}_{p}\wr\mathbb{Z}}(n)\preceq\operatorname{Conj}_{A\wr\mathbb{Z}}(n).\qed

The next lemma, which is a direct consequence of [6, Theorem 3.4], relates the size of the support of its function part and the size of the elements in the range of the function with the word length of an element. We omit the proof in order to avoid having to introduce more technical notation, we encourage a curious reader to inspect [6, Theorem 3.4] and prove check that the statement indeed holds.

Lemma 2.9.

Let A,BA,B be finitely generated groups and let G=ABG=A\wr B be their wreath product. Then there exists a constant C>0C>0 such that if g=fbg=fb where fABf\in A^{B} and bBb\in B, then

  • (i)

    supp(f)BB(Cg)\operatorname{supp}(f)\subseteq\operatorname{B}_{B}(C\|g\|),

  • (ii)

    rng(f)BA(Cg)\operatorname{rng}(f)\subseteq\operatorname{B}_{A}(C\|g\|),

  • (iii)

    bBB(Cg)b\in\operatorname{B}_{B}(C\|g\|).

Given a wreath product ABA\wr B with a surjective homomorphism π:BB¯\pi\colon B\to\overline{B}, we denote π~:ABAB¯\tilde{\pi}\colon A\wr B\to A\wr\overline{B} as the canonical extension of π\pi to all of ABA\wr B given by

π~(f)(bK)=kKf(b+k),\tilde{\pi}(f)(bK)=\sum_{k\in K}f(b+k),

where K=ker(π)K=\ker(\pi). Note that since the group AA is abelian and the function fABf\in A^{B} is finitely supported, the above sum is well defined. Similarly, if π:AA¯\pi\colon A\to\overline{A} is a surjective homomorphism, we let π~:ABA¯B\tilde{\pi}\colon A\wr B\to\overline{A}\wr B as the natural extension of π\pi to all of ABA\wr B.

2.3. Wreath products and Laurent polynomial rings


Much of the following discussion, which includes undefined notation and terms, can be found in [2, 8, 13]. Given a commutative ring RR, we will write R[x]R[x] to denote the ring of polynomials in the variable xx with coefficients in RR, and we will use R[x,x1]R[x,x^{-1}] to denote the ring of Laurent polynomials over RR.

We first note that R[x,x1]R[x,x^{-1}] is the localisation of the ring R[x]R[x] on the set S={xm|m}S=\{x^{m}\>|\>m\in\mathbb{N}\}. We then have that the ideals of R[x,x1]R[x,x^{-1}] are in one-to-one correspondence with ideals of R[x]R[x] that don’t intersect the set SS. In particular, for any ideal R[x]\mathcal{I}\subset R[x] where S=\mathcal{I}\cap S=\emptyset, we have that R[x,x1]/(S1)=S1(R[x]/)R[x,x^{-1}]/(S^{-1}\mathcal{I})=S^{-1}(R[x]/\mathcal{I}). We finish by observing that the maximal ideals of R[x,x1]R[x,x^{-1}] can be written as =(f)\mathcal{I}=(f) where ff is an irreducible polynomial not in SS. If k=deg(f)k=\deg(f), then |R[x,x1]/|=|R|k|R[x,x^{-1}]/\mathcal{I}|=|R|^{k}.

We now focus on the following representation of RR\wr\mathbb{Z} as a semidirect product of the ring R[x,x1]R[x,x^{-1}] and \mathbb{Z}. First, let us define a function P:RR[x,x1]P\colon R^{\mathbb{Z}}\to R[x,x^{-1}] given by

P(f)=mf(m)xm.P(f)=\sum_{m\in\mathbb{Z}}f(m)x^{m}.

One can easily check in the context of finitely supported functions that PP is a bijection and for any rRr\in R, f,gRf,g\in R^{\mathbb{Z}}, and mm\in\mathbb{Z} that the following holds:

  1. (i)

    P(rf)=rP(f)P(rf)=rP(f),

  2. (ii)

    P(f+g)=P(f)+P(g)P(f+g)=P(f)+P(g),

  3. (iii)

    P(mf)=xmP(f)P(m\cdot f)=x^{m}P(f).

We will use these three equalities without mention.

Lemma 2.10.

Let RR be either the ring \mathbb{Z} or 𝔽p\mathbb{F}_{p} where pp is prime. The group RR\wr\mathbb{Z} is isomorphic to R[x,x1]R[x,x^{-1}]\rtimes\mathbb{Z} where R[x,x1]R[x,x^{-1}] is ring of Laurent polynomials with addition and for tt\in\mathbb{Z}, we have tf(x)=xtf(x).t\cdot f(x)=x^{t}f(x).

Proof.

Let φ:RR[x,x1]\varphi\colon R\wr\mathbb{Z}\to R[x,x^{-1}]\rtimes\mathbb{Z} be the map given by φ(fm)=(P(f),m)\varphi\left(fm\right)=(P(f),m). It is then easy to see that this map is an isomorphism. ∎

The following lemma allows us to understand finite quotients of RR\wr\mathbb{Z} in terms of the cofinite ideals of R[x,x1]R[x,x^{-1}]. For the following lemma, we identify R[x,x1]R[x,x^{-1}] with the normal subgroup of RR\wr\mathbb{Z} given by elements of the form (P,0)(P,0) where PR[x,x1].P\in R[x,x^{-1}].

Lemma 2.11.

Let RR be either the ring \mathbb{Z} or 𝔽p\mathbb{F}_{p} where 𝔽p\mathbb{F}_{p} be the field with pp elements. Let NRN\trianglelefteq R\wr\mathbb{Z}. Then NR[x,x1]N\cap R[x,x^{-1}] is an ideal in R[x,x1]R[x,x^{-1}]. In particular, if Nf.iRN\trianglelefteq_{f.i}R\wr\mathbb{Z}, then NR[x,x1]N\cap R[x,x^{-1}] is a cofinite ideal of R[x,x1]R[x,x^{-1}].

Proof.

Let M=NR[x,x1].M=N\cap R[x,x^{-1}]. We note for (P,0)M(P,0)\in M that

(0,m)(P,0)(0,m)=(xmP,0)M\left(0,m\right)\left(P,0\right)\left(0,-m\right)=\left(x^{m}P,0\right)\in M

since MM is normal. In particular, we have that MM is closed under multiplication by xmx^{m} in R[x,x1]R[x,x^{-1}] for all mm\in\mathbb{Z}. Additionally, for (P1,0),(P2,0)M(P_{1},0),(P_{2},0)\in M we have that

(P1,0)(P2,0)=(P1+P2,0).(P_{1},0)(P_{2},0)=(P_{1}+P_{2},0).

That implies MM is closed under addition. Since multiplying PP by rr is the same as adding rr copies of PP and given that MM is a subgroup of R[x,x1]R[x,x^{-1}] with addition as its group operation, we must have that (rP,0)M(rP,0)\in M. Thus, for (P,0)M(P,0)\in M and a general element mamxm\sum_{m\in\mathbb{Z}}a_{m}x^{m} of R[x,x1]R[x,x^{-1}], we may write

Pmamxm=mamxmPM.P\cdot\sum_{m\in\mathbb{Z}}a_{m}x^{m}=\sum_{m\in\mathbb{Z}}a_{m}x^{m}P\in M.

Thus, MM is an ideal in R[x,x1]R[x,x^{-1}]. Moreover, the second part of the statement immediately follows. ∎

The following lemma gives the explicit expression for the conjugacy class of an arbitrary element of RR\wr\mathbb{Z}.

Lemma 2.12.

Let RR be either the ring \mathbb{Z} or 𝔽p\mathbb{F}_{p} where 𝔽p\mathbb{F}_{p} is the field with pp elements. For (P,m)R(P,m)\in R\wr\mathbb{Z}, its conjugacy class is given by

{((xP+(xm1)Q,m)|,QR[x,x1]}.\left\{\left(\left(x^{\ell}P+\left(x^{m}-1\right)Q,m\right)\>\right|\ell\in\mathbb{Z},Q\in R[x,x^{-1}]\right\}.
Proof.

Let QR[x,x1]Q\in R[x,x^{-1}] and \ell\in\mathbb{Z} be arbitrary. We then write

(Q,)(P,m)(Q,)1\displaystyle(Q,\ell)(P,m)(Q,\ell)^{-1} =(Q+xP,+m)(xQ,)\displaystyle=\left(Q+x^{\ell}P,\ell+m\right)\left(-x^{-\ell}Q,-\ell\right)
=(Q+xPx+mxQ,+m)\displaystyle=\left(Q+x^{\ell}P-x^{\ell+m}x^{-\ell}Q,\ell+m-\ell\right)
=(Q+xPxmQ,m)\displaystyle=\left(Q+x^{\ell}P-x^{m}Q,m\right)
=(xP+(1xm)Q,m).\displaystyle=\left(x^{\ell}P+(1-x^{m})Q,m\right).

Since QQ was arbitrary, we may replace it by Q-Q allowing us to write

(Q,)(P,m)(Q,)1=(xP+(xm1)Q,m).(Q,\ell)(P,m)(Q,\ell)^{-1}=\left(x^{\ell}P+(x^{m}-1)Q,m\right).

From here, our statement is clear. ∎

3. Lower bounds

In this section, we construct asymptotic lower bounds for conjugacy separability for the groups 𝔽p\mathbb{F}_{p}\wr\mathbb{Z} and .\mathbb{Z}\wr\mathbb{Z}. which we divide into two subsections. The first subsection goes over the lower bounds for Conj𝔽p(n)\operatorname{Conj}_{\mathbb{F}_{p}\wr\mathbb{Z}}(n). The second subsection constructs lower bounds for Conj(n)\operatorname{Conj}_{\mathbb{Z}\wr\mathbb{Z}}(n).

3.1. Lower bounds for ConjAB(n)\operatorname{Conj}_{A\wr B}(n) where AA is a finite abelian group


In this section, we provide asymptotic lower bounds for ConjAB(n)\operatorname{Conj}_{A\wr B}(n) when AA is a finite abelian group and BB is an infinite, finitely generated abelian group by finding asymptotic bounds for Conj𝔽p(n)\operatorname{Conj}_{\mathbb{F}_{p}\wr\mathbb{Z}}(n) .

Proposition 3.1.

Let AA be a finite abelian group and BB be an infinite, finitely generated abelian group. Then 2nConjAB(n).2^{n}\preceq\operatorname{Conj}_{A\wr B}(n).

Proof.

By Lemma 2.8, we may assume that A𝔽pA\cong\mathbb{F}_{p} for some prime and that BB\cong\mathbb{Z}. We need to find an infinite sequence of pairs of elements {fi,gi}i=1\{f_{i},g_{i}\}^{\infty}_{i=1} such that

  1. (i)

    limimax{fi,gi}=\lim_{i\to\infty}\text{max}\left\{\|f_{i}\|,\|g_{i}\|\right\}=\infty,

  2. (ii)

    figif_{i}\nsim g_{i},

  3. (iii)

    pCmax{fi,gi}CD𝔽p(fi,gi)p^{C\text{max}\left\{\|f_{i}\|,\|g_{i}\|\right\}}\leq\operatorname{CD}_{\mathbb{F}_{p}\wr\mathbb{Z}}(f_{i},g_{i}),

where C>0C>0 is some constant.

Let {qi}i=1\{q_{i}\}_{i=1}^{\infty} be an enumeration of the set of primes greater than pp such that pp is a primitive root mod qiq_{i}. In this case, it is well known that ψqi(x)=i=1qi1xi\psi_{q_{i}}(x)=\sum_{i=1}^{q_{i}-1}x^{i} is an irreducible polynomial over 𝔽p\mathbb{F}_{p}. Let

fi=(xqi1,qi) and gi=(x1+xqi1,qi).f_{i}=\left(x^{q_{i}}-1,q_{i}\right)\quad\text{ and }\quad g_{i}=\left(x-1+x^{q_{i}}-1,q_{i}\right).

Let us consider the quotient 𝔽p[x,x1]/(ψqi(x))(/qi))\mathbb{F}_{p}[x,x^{-1}]/(\psi_{q_{i}}(x))\rtimes(\mathbb{Z}/q_{i}\mathbb{Z})) with the associated projection map πi\pi_{i}. We then see that

|𝔽p[x,x1]/(ψqi(x))(/qi)|=qipqi1|\mathbb{F}_{p}[x,x^{-1}]/(\psi_{q_{i}}(x))\rtimes(\mathbb{Z}/q_{i}\mathbb{Z})|=q_{i}p^{q_{i}-1}

and that

πi(fi)=(0,0) and πi(gi)=(x1,0)(0,0).\pi_{i}(f_{i})=(0,0)\quad\text{ and }\quad\pi_{i}(g_{i})=(x-1,0)\neq(0,0).

It follows that π(fi)π(gi)\pi(f_{i})\nsim\pi(g_{i}). Subsequently, we see that fif_{i} and gig_{i} are not conjugate in 𝔽p\mathbb{F}_{p}\wr\mathbb{Z} and that CD𝔽p(fi,gi)qipqi1\operatorname{CD}_{\mathbb{F}_{p}\wr\mathbb{Z}}(f_{i},g_{i})\leq q_{i}p^{q_{i}-1}.

To finish, we will demonstrate that pqiCD𝔽p(fi,gi)p^{q_{i}}\leq\operatorname{CD}_{\mathbb{F}_{p}\wr\mathbb{Z}}(f_{i},g_{i}) for all ii. In other words, we need to show that if Nf.i.𝔽pN\trianglelefteq_{f.i.}\mathbb{F}_{p}\wr\mathbb{Z} is given such that |(𝔽p)/N|<pqi|(\mathbb{F}_{p}\wr\mathbb{Z})/N|<p^{q_{i}}, then figi mod Nf_{i}\sim g_{i}\text{ mod }N. Suppose that such a normal finite index subgroup NN is given. We note by Lemma 2.11 that 𝒥N=N𝔽p[x,x1]\mathcal{J}_{N}=N\cap\mathbb{F}_{p}[x,x^{-1}] is an ideal in 𝔽p[x,x1]\mathbb{F}_{p}[x,x^{-1}]. In particular, 𝒥NNN\mathcal{J}_{N}\rtimes N\cap\mathbb{Z}\leq N is a normal subgroup in 𝔽p\mathbb{F}_{p}\wr\mathbb{Z} such that if figi mod 𝒥NN,f_{i}\sim g_{i}\text{ mod }\mathcal{J}_{N}\rtimes N\cap\mathbb{Z}, then figi mod Nf_{i}\sim g_{i}\text{ mod }N. Thus,for the purpose of the proof, we may assume that N𝒥tN\cong\mathcal{J}\rtimes t\mathbb{Z} for some tt\in\mathbb{N}. If 𝒥N=𝔽p[x,x1]\mathcal{J}_{N}=\mathbb{F}_{p}[x,x^{-1}], then (𝔽p)/N(\mathbb{F}_{p}\wr\mathbb{Z})/N is a finite abelian group. In particular, we have that 𝔽p[x,x1]ker(πN),\mathbb{F}_{p}[x,x^{-1}]\leq\ker(\pi_{N}), and thus, πN(fi)=πN(gi)\pi_{N}(f_{i})=\pi_{N}(g_{i}). Hence, we may assume that 𝒥N\mathcal{J}_{N} is a proper ideal in 𝔽p[x,x1]\mathbb{F}_{p}[x,x^{-1}]. Moreover, we have that |𝔽p[x,x1]/𝒥N|<pqi.|\mathbb{F}_{p}[x,x^{-1}]/\mathcal{J}_{N}|<p^{q_{i}}.

Since 𝔽[x,x1]\mathbb{F}[x,x^{-1}] is a localisation of a principal ideal domain, it is also a principal ideal domain. Therefore, there exists a polynomial P𝔽p[x]P\in\mathbb{F}_{p}[x] such that 𝒥N=(P)\mathcal{J}_{N}=(P). Thus, we note that one of the following cases must hold:

gcd(xqi1,P)={xqi1,ψqi,x1,1.\gcd(x^{q_{i}}-1,P)=\begin{cases}x^{q_{i}}-1,\\ \psi_{q_{i}},\\ x-1,\\ 1\end{cases}.

Let us first note that 𝔽p[x,x1]/(P)(𝔽p[x,x1])/N\mathbb{F}_{p}[x,x^{-1}]/(P)\leq\left(\mathbb{F}_{p}[x,x^{-1}]\rtimes\mathbb{Z}\right)/N. We see that we may ignore the first two cases, as in both we have that pqi|𝔽p/N|p^{q_{i}}\leq\left|\mathbb{F}_{p}\wr\mathbb{Z}/N\right|.

For the third case, we have that x1𝒥Nx-1\in\mathcal{J}_{N}. Therefore, we have

πN(fi)\displaystyle\pi_{N}(f_{i}) =\displaystyle= (xqi1 mod 𝒥N,qi mod t)\displaystyle(x^{q_{i}}-1\text{ mod }\mathcal{J}_{N},q_{i}\text{ mod }t)
=\displaystyle= ((x1)ψqi(x) mod 𝒥N,qi mod t)\displaystyle((x-1)\psi_{q_{i}}(x)\text{ mod }\mathcal{J}_{N},q_{i}\text{ mod }t)
=\displaystyle= (0,qi mod t).\displaystyle(0,q_{i}\text{ mod }t).

Similarly, we have

πN(gi)\displaystyle\pi_{N}(g_{i}) =\displaystyle= (x1+(x1)ψqi(x) mod 𝒥,qi mod t)\displaystyle(x-1+(x-1)\psi_{q_{i}}(x)\text{ mod }\mathcal{J},q_{i}\text{ mod }t)
=\displaystyle= (0,qi mod t).\displaystyle(0,q_{i}\text{ mod }t).

Hence, πN(fi)=πN(gi).\pi_{N}(f_{i})=\pi_{N}(g_{i}).

For the last case, we may assume that gcd(xqi1,P)=1\gcd(x^{q_{i}}-1,P)=1. Let us recall that, following Lemma 2.12, we can write the conjugacy class of fif_{i} as

{(xn(xqi1)+(xqi1)λ,qi)|n,λ𝔽p[x,x1]}.\left\{\left.(x^{n}(x^{q_{i}}-1)+(x^{q_{i}}-1)\lambda,q_{i})\>\right|\>n\in\mathbb{Z},\lambda\in\mathbb{F}_{p}[x,x^{-1}]\right\}.

In order for figi mod Nf_{i}\sim g_{i}\text{ mod }N, we need to have

x1+xqi1{xn(xqi1)+(xqi1)λ|n,λ𝔽p[x,x1]} mod (P).x-1+x^{q_{i}}-1\in\{x^{n}(x^{q_{i}}-1)+(x^{q_{i}}-1)\lambda\>|\>n\in\mathbb{Z},\lambda\in\mathbb{F}_{p}[x,x^{-1}]\}\text{ mod }(P).

The above is equivalent to

x1{(xn+λ1)(xqi1)|n,λ𝔽p[x,x1]} mod (P).x-1\in\{(x^{n}+\lambda-1)(x^{q_{i}}-1)\>|\>n\in\mathbb{Z},\lambda\in\mathbb{F}_{p}[x,x^{-1}]\}\text{ mod }(P).

Using basic algebra, we see that the above is equivalent to

x1{λ(xqi1)|λ𝔽p[x,x1]} mod (P).x-1\in\{\lambda(x^{q_{i}}-1)\>|\lambda\in\mathbb{F}_{p}[x,x^{-1}]\}\text{ mod }(P).

Thus, we have that figi mod Nf_{i}\sim g_{i}\text{ mod }N if and only if x1(xqi1) mod (P)x-1\in(x^{q_{i}}-1)\text{ mod }(P).

Since gcd(xqi1,P)=1,\gcd(x^{q_{i}}-1,P)=1, there exist polynomials α,β𝔽p[X]\alpha,\beta\in\mathbb{F}_{p}[X] such that

(xqi1)α+Pβ=1.(x^{q_{i}}-1)\alpha+P\beta=1.

By multiplying through by x1x-1, we may write

x1=(x1)(xqi1)α+(x1)Pβ.x-1=(x-1)(x^{q_{i}}-1)\alpha+(x-1)P\beta.

Reducing mod (P)(P), we have

x1=(x1)α(xqi1) mod (P).x-1=(x-1)\alpha(x^{q_{i}}-1)\text{ mod }(P).

We see that figi mod N,f_{i}\sim g_{i}\text{ mod }N, and therefore,

pqiCD𝔽p(fi,gi)qipqi1.p^{q_{i}}\leq\operatorname{CD}_{\mathbb{F}_{p}\wr\mathbb{Z}}(f_{i},g_{i})\leq q_{i}p^{q_{i}-1}.

From the construction of the elements fi,gif_{i},g_{i}, it can be easily seen that there is a constant CC^{\prime} such that

qifiCqiandqigiCqi.q_{i}\leq\|f_{i}\|\leq C^{\prime}q_{i}\quad\text{and}\quad q_{i}\leq\|g_{i}\|\leq C^{\prime}q_{i}.

There, we have that pnConj𝔽p(Cn)p^{n}\leq\operatorname{Conj}_{\mathbb{F}_{p}\wr\mathbb{Z}}(C^{\prime}n). Hence, we may write

2nConj𝔽p(n)2^{n}\preceq\operatorname{Conj}_{\mathbb{F}_{p}\wr\mathbb{Z}}(n)

since 2npn2^{n}\approx p^{n}. ∎

3.2. Lower bounds for ConjAB(n)\operatorname{Conj}_{A\wr B}(n) when AA and BB are infinite


In this subsection, we provide asymptotic lower bounds for ConjAB(n)\operatorname{Conj}_{A\wr B}(n) where AA and BB are infinite, finitely generated abelian groups. We start with the group \mathbb{Z}\wr\mathbb{Z} as seen in Proposition 3.3. Before we start, we have the following lemma.

Lemma 3.2.

Let m,nm,n\in\mathbb{N} and dd\in\mathbb{Z}. Then

(xm1)(xgcd(m,n)1) mod (xn1,d).(x^{m}-1)\equiv(x^{\gcd(m,n)}-1)\text{ mod }(x^{n}-1,d).
Proof.

We note that m=gcd(m,n)m=\ell\gcd(m,n) for some integer \ell. Therefore,

xmxgcd(m,n)1 mod (xgcd(m,n)1).x^{m}\equiv x^{\ell\gcd(m,n)}\equiv 1\text{ mod }(x^{\gcd(m,n)}-1).

Hence, xgcd(m,n)1xm1x^{\gcd(m,n)}-1\mid x^{m}-1, and thus,

(xm1)(xgcd(m,n)1) mod (xn1,d).\left(x^{m}-1\right)\subset(x^{\gcd(m,n)}-1)\text{ mod }(x^{n}-1,d).

For the other inclusion, we note that there exist integers t,st,s such that gcd(m,n)=tm+sn.\gcd(m,n)=tm+sn. Hence, we may write

(xtm1)(xsn1)=\displaystyle(x^{tm}-1)(x^{sn}-1)= xtm+snxtmxsn+1\displaystyle x^{tm+sn}-x^{tm}-x^{sn}+1 mod (xn1,d)\displaystyle\text{ mod }(x^{n}-1,d)
=\displaystyle= xgcd(m,n)xsn(xtm1)\displaystyle x^{\gcd(m,n)}-x^{sn}-(x^{tm}-1) mod (xn1,d)\displaystyle\text{ mod }(x^{n}-1,d)
\displaystyle\equiv xgcd(m,n)1(xtm1)\displaystyle x^{\gcd(m,n)}-1-(x^{tm}-1) mod (xn1,d).\displaystyle\text{ mod }(x^{n}-1,d).

Since (xtm1)(xsn1)(xn1,d)(x^{tm}-1)(x^{sn}-1)\in(x^{n}-1,d), we have

xgcd(m,n)1xtm1 mod (xn1,d).x^{\gcd(m,n)}-1\equiv x^{tm}-1\text{ mod }(x^{n}-1,d).

Hence,

(xm1)(xgcd(m,n)1) mod (xn1,d).(x^{m}-1)\equiv(x^{\gcd(m,n)}-1)\text{ mod }(x^{n}-1,d).\qed

We now come to the last proposition of this section.

Proposition 3.3.

Let AA and BB be infinite, finitely generated abelian groups. Then

(logn)nConjAB(n).(\log n)^{n}\preceq\operatorname{Conj}_{A\wr B}(n).
Proof.

Let us first note that we may choose splittings of AA and BB as direct sums AkTor(B)A\simeq\mathbb{Z}^{k}\oplus\operatorname{Tor}(B) and BdTor(B)B\simeq\mathbb{Z}^{d}\oplus\operatorname{Tor}(B). Since we assumed that both A,BA,B are infinite, we see that d,k>0d,k>0. In particular, AA contains an element aa of an infinite order such that a\langle a\rangle is an retract of AA and BB contains an element bb of an infinite order such that b\langle b\rangle is an retract of BB. By Lemma 2.5 we see that the subgroup ab\mathbb{Z}\wr\mathbb{Z}\simeq\langle a\rangle\wr\langle b\rangle is a retract of ABA\wr B and Conj(n)ConjAB(n)\operatorname{Conj}_{\mathbb{Z}\wr\mathbb{Z}}(n)\preceq\operatorname{Conj}_{A\wr B}(n). Hence, we may assume that ABA\wr B\cong\mathbb{Z}\wr\mathbb{Z}.

We need to find an infinite sequence of pairs of nonconjugate elements {fi,gi}\{f_{i},g_{i}\} such that log(Cmax{fi,gi})Cmax{fi,gi}<CD(fi,gi)\log(C\text{max}\{\|f_{i}\|,\|g_{i}\|\})^{C\text{max}\{\|f_{i}\|,\|g_{i}\|\}}<\operatorname{CD}_{\mathbb{Z}\wr\mathbb{Z}}(f_{i},g_{i}) for some C>0C>0. For ease of writing, we denote

[x,x1]\mathbb{Z}\wr\mathbb{Z}\simeq\mathbb{Z}[x,x^{-1}]\rtimes\mathbb{Z}

where \mathbb{Z} acts by multiplication on [x,x1]\mathbb{Z}[x,x^{-1}] by xx.

Let {qi}\{q_{i}\} be an enumeration of the primes, and let α(i)=lcm(1,,qi1)\alpha(i)=\operatorname{lcm}(1,\cdots,q_{i}-1). Finally, let kik_{i} be the smallest integer such that α(i)2ki\alpha(i)\leq 2^{k_{i}}. We define the elements fi,gif_{i},g_{i}\in\mathbb{Z}\wr\mathbb{Z} as

fi=(α(i)(x2ki1),2ki) and gi=(α(i)(x2ki1+x2ki11),2ki).f_{i}=(\alpha(i)(x^{2^{k_{i}}}-1),2^{k_{i}})\quad\text{ and }\quad g_{i}=(\alpha(i)(x^{2^{k_{i}}}-1+x^{2^{k_{i}-1}}-1),2^{k_{i}}).

To see that fif_{i} is not conjugate to gig_{i}, we set 𝔨i\mathfrak{k}_{i} be the ideal in [x,x1]\mathbb{Z}[x,x^{-1}] given by (2ki,x2ki1)(2^{k_{i}},x^{2^{k_{i}}}-1), and let H=𝔨iqiH=\mathfrak{k}_{i}\rtimes q_{i}\mathbb{Z}\leq\mathbb{Z}\wr\mathbb{Z}. We see that |()/H|=2kiqi2ki|(\mathbb{Z}\wr\mathbb{Z})/H|=2^{k_{i}}q_{i}^{2^{k_{i}}} and that

πH(fi)=(0,0) and πH(gi)=(α(i)(x2ki11),0)(0,0)\pi_{H}(f_{i})=(0,0)\quad\text{ and }\quad\pi_{H}(g_{i})=\left(\alpha(i)(x^{2^{k_{i}-1}}-1),0\right)\neq(0,0)

where πH:()/H\pi_{H}\colon\mathbb{Z}\wr\mathbb{Z}\to(\mathbb{Z}\wr\mathbb{Z})/H is the natural projection. Therefore, fi≁gif_{i}\not\sim g_{i} in \mathbb{Z}\wr\mathbb{Z}.

Now suppose that NN\trianglelefteq\mathbb{Z}\wr\mathbb{Z} is a finite index subgroup where |()/N|<qi2ki|(\mathbb{Z}\wr\mathbb{Z})/N|<q_{i}^{2^{k_{i}}}. We will show that fiNgiNf_{i}N\sim g_{i}N in ()/N(\mathbb{Z}\wr\mathbb{Z})/N.

We note by Lemma 2.11 that 𝒥N=N[x,x1]\mathcal{J}_{N}=N\cap\mathbb{Z}[x,x^{-1}] is an ideal in [x,x1]\mathbb{Z}[x,x^{-1}]. In particular, 𝒥(N)\mathcal{J}\rtimes(N\cap\mathbb{Z}) is a normal subgroup in \mathbb{Z}\wr\mathbb{Z}. Similarly, N=bN\cap\mathbb{Z}=b\mathbb{Z} for some bb\in\mathbb{Z}. Therefore, we denote N=𝒥NbN^{\prime}=\mathcal{J}_{N}\rtimes b\mathbb{Z}. Thus, it follows that NN^{\prime} is a finite index normal subgroup of \mathbb{Z}\wr\mathbb{Z} where NNN^{\prime}\leq N. In particular, if fiN≁giNf_{i}N\not\sim g_{i}N in ()/N(\mathbb{Z}\wr\mathbb{Z})/N, then fiN≁giNf_{i}N^{\prime}\not\sim g_{i}N^{\prime} in ()/N(\mathbb{Z}\wr\mathbb{Z})/N^{\prime}. Therefore, we may assume that ()/N(\mathbb{Z}\wr\mathbb{Z})/N takes the form ([x,x1]/𝒥)(/b)(\mathbb{Z}[x,x^{-1}]/\mathcal{J})\rtimes(\mathbb{Z}/b\mathbb{Z}) where 𝒥\mathcal{J} is a cofinite ideal and bb is an integer.

Following Lemma 2.2, we see that fiN′′giN′′f_{i}N^{\prime\prime}\nsim g_{i}N^{\prime\prime} for N′′=𝒥(b+K)N^{\prime\prime}=\mathcal{J}\rtimes(b\mathbb{Z}+K) where KK\leq\mathbb{Z} is the preimage under the projection modulo bb of the kernel of the action of /b\mathbb{Z}/b\mathbb{Z} on [x,x1]/𝒥N\mathbb{Z}[x,x^{-1}]/\mathcal{J}_{N}. Letting b00b_{0}\geq 0 be such that b0=b+Kb_{0}\mathbb{Z}=b\mathbb{Z}+K, we note that N′′N^{\prime\prime} is a finite index normal subgroup where

()/N′′=([x,x1])/(𝒥Nb)([x,x1]/𝒥N)(/b).(\mathbb{Z}\wr\mathbb{Z})/N^{\prime\prime}=\left(\mathbb{Z}[x,x^{-1}]\rtimes\mathbb{Z}\right)/(\mathcal{J}_{N}\rtimes b\mathbb{Z})\simeq(\mathbb{Z}[x,x^{-1}]/\mathcal{J}_{N})\rtimes(\mathbb{Z}/b\mathbb{Z}).

Therefore, by the above discussion, we may assume that

()/N([x,x1]/𝒥)(/b)(\mathbb{Z}\wr\mathbb{Z})/N\cong(\mathbb{Z}[x,x^{-1}]/\mathcal{J})\rtimes(\mathbb{Z}/b\mathbb{Z})

where /b\mathbb{Z}/b\mathbb{Z} acts faithfully on [x,x1]/𝒥.\mathbb{Z}[x,x^{-1}]/\mathcal{J}.

We now show we may assume that /b\mathbb{Z}/b\mathbb{Z} acts freely on [x,x1]/𝒥\mathbb{Z}[x,x^{-1}]/\mathcal{J}. Suppose that there are polynomials ρ(x),λ(x)[x,x1]/𝒥\rho(x),\lambda(x)\in\mathbb{Z}[x,x^{-1}]/\mathcal{J} such that

xmρ(x)+𝒥N=xmλ(x)+𝒥Nx^{m}\rho(x)+\mathcal{J}_{N}=x^{m}\lambda(x)+\mathcal{J}_{N}

for some 0m<b0\leq m<b. Since xmx^{m} is a unit, we may cancel and write

ρ(x)+𝒥N=λ(x)+𝒥N\rho(x)+\mathcal{J}_{N}=\lambda(x)+\mathcal{J}_{N}

which gives our claim.

Let \ell be the multiplicative order of x+𝒥x+\mathcal{J} in [x,x1]/𝒥\mathbb{Z}[x,x^{-1}]/\mathcal{J}. We claim that =b\ell=b. By definition, we have that bb is the smallest integer such that

xbρ(x)+𝒥=ρ(x)+𝒥x^{b}\rho(x)+\mathcal{J}=\rho(x)+\mathcal{J}

for all ρ(x)[x,x1]\rho(x)\in\mathbb{Z}[x,x^{-1}]. In particular, we have that xb1=1mod𝒥x^{b}\cdot 1=1\mod\mathcal{J}. Thus, we have that b\ell\mid b. If b\ell\lneq b, we then have that xρ(x)=ρ(x)mod𝒥x^{\ell}\rho(x)=\rho(x)\mod\mathcal{J} for all ρ(x)[x,x1]\rho(x)\in\mathbb{Z}[x,x^{-1}]. However, that implies /b\mathbb{Z}/b\mathbb{Z} doesn’t act faithfully on [x,x1]/𝒥\mathbb{Z}[x,x^{-1}]/\mathcal{J} which is a contradiction. Therefore, we have that =b\ell=b.

Since /b\mathbb{Z}/b\mathbb{Z} acts freely and transitively on the set of powers of x mod 𝒥x\text{ mod }\mathcal{J} in [x,x1]/𝒥\mathbb{Z}[x,x^{-1}]/\mathcal{J}, we have that |[x,x1]/𝒥|=db|\mathbb{Z}[x,x^{-1}]/\mathcal{J}|=d^{b} where dd is the characteristic of the finite ring [x,x1]/𝒥N\mathbb{Z}[x,x^{-1}]/\mathcal{J}_{N}. We note that the ideal (d,xb1)(d,x^{b}-1) is contained in the ideal 𝒥\mathcal{J} and that |[x,x1]/(d,xb1)|=db|\mathbb{Z}[x,x^{-1}]/(d,x^{b}-1)|=d^{b}. It follows that 𝒥=(d,xb1)\mathcal{J}=(d,x^{b}-1).

If d<qid<q_{i}, then dα(i)d\mid\alpha(i), and subsequently,

α(i)(x2ki1),α(i)(x2ki1+x2ki11)(d,xb01).\alpha(i)(x^{2^{k_{i}}}-1),\alpha(i)(x^{2^{k_{i}}}-1+x^{2^{k_{i}-1}}-1)\in(d,x^{b_{0}}-1).

Hence, fi=gi mod Nf_{i}=g_{i}\text{ mod }N. Therefore, we may assume that dqi.d\geq q_{i}.

By Lemma 2.12, we may write the conjugacy class of fif_{i} as

{((x(x2ki1)+(x2ki1)Q,2ki)|,Q[x,x1]}.\{((x^{\ell}(x^{2^{k_{i}}}-1)+(x^{2^{k_{i}}}-1)Q,2^{k_{i}})\>|\>\ell\in\mathbb{Z},Q\in\mathbb{Z}[x,x^{-1}]\}.

Thus, we have that figif_{i}\sim g_{i} if and only if

x2ki1+x2ki11{x(x2ki1)+(x2ki1)Q,|,Q[x,x1]}x^{2^{k_{i}}}-1+x^{2^{k_{i}-1}}-1\in\{x^{\ell}(x^{2^{k_{i}}}-1)+(x^{2^{k_{i}}}-1)Q,\>|\>\ell\in\mathbb{Z},Q\in\mathbb{Z}[x,x^{-1}]\}

which is equivalent to

x2ki11{(x1+Q)(x2ki1)|,Q[x,x1]}.x^{2^{k_{i}-1}}-1\in\{(x^{\ell}-1+Q)(x^{2^{k_{i}}}-1)\>|\ell\in\mathbb{Z},Q\in\mathbb{Z}[x,x^{-1}]\}.

Since x1+Qx^{\ell}-1+Q can be any Laurent polynomial, we have that figif_{i}\sim g_{i} if and only if

x2ki11{Q(x2ki1)|Q[x,x1]}.x^{2^{k_{i}-1}}-1\in\{Q(x^{2^{k_{i}}}-1)\>|\>Q\in\mathbb{Z}[x,x^{-1}]\}.

By Lemma 3.2, we have that

x2ki1xgcd(2ki,b)1 mod (d,xb1).x^{2^{k_{i}}}-1\equiv x^{\gcd(2^{k_{i}},b)}-1\text{ mod }(d,x^{b}-1).

Therefore, we may write the conjugacy class of fif_{i} in ()/N(\mathbb{Z}\wr\mathbb{Z})/N as

{(Q(x2t1)|Q[x,x1]} mod (d,xb1).\{(Q(x^{2^{t}}-1)\>|\>Q\in\mathbb{Z}[x,x^{-1}]\}\text{ mod }(d,x^{b}-1).

where 0t<ki.0\leq t<k_{i}. Therefore, fiNgiNf_{i}N\sim g_{i}N if and only if

x2ki11{(Q(x2t1)|Q[x,x1]} mod (d,xb1).x^{2^{k_{i}-1}}-1\in\{(Q(x^{2^{t}}-1)\>|\>Q\in\mathbb{Z}[x,x^{-1}]\}\text{ mod }(d,x^{b}-1).

Since 2t2ki12^{t}\mid 2^{k_{i}-1}, it is well known that x2t1x2ki11x^{2^{t}}-1\mid x^{2^{k_{i}-1}}-1. Therefore,

x2ki11{(Q(x2t1)|Q[x,x1]} mod (d,xb1)x^{2^{k_{i}-1}}-1\in\{(Q(x^{2^{t}}-1)\>|\>Q\in\mathbb{Z}[x,x^{-1}]\}\text{ mod }(d,x^{b}-1)

when gcd(2ki1,b)2ki1\gcd(2^{k_{i}-1},b)\leq 2^{k_{i}-1}. Hence, if b2ki1,b\leq 2^{k_{i}-1}, then gcd(2ki1,b)2ki1\gcd(2^{k_{i}-1},b)\leq 2^{k_{i}-1}, and subsequently, fiNgiNf_{i}N\sim g_{i}N. We see that

qi2ki<CD(fi,gi)<2kiqi2ki.q_{i}^{2^{k_{i}}}<\operatorname{CD}_{\mathbb{Z}\wr\mathbb{Z}}(f_{i},g_{i})<2^{k_{i}}q_{i}^{2^{k_{i}}}.

Recall that α(i)=exp{υ(qi1)}\alpha(i)=\exp\{\upsilon(q_{i}-1)\}, where υ:\upsilon\colon\mathbb{N}\to\mathbb{N} is the second Chebyshev’s function. The Prime Number Theorem [25, 1.2] then implies that there are constants C0,C0+>0C_{0}^{-},C_{0}^{+}>0 such that 2C0qiα(i)2C0+qi2^{C^{-}_{0}q_{i}}\leq\alpha(i)\leq 2^{C^{+}_{0}q_{i}}. Following the definition of kik_{i}, we see that there are constants C1,C1+>0C_{1}^{-},C_{1}^{+}>0 such that 2C1qi2ki2C1+qi2^{C^{-}_{1}q_{i}}\leq 2^{k_{i}}\leq 2^{C^{+}_{1}q_{i}}. From the construction of the elements fi,gif_{i},g_{i}, it can be easily seen that there is a constant CC^{\prime} such that

α(i)2kiniCα(i)2ki,\alpha(i)2^{k_{i}}\leq n_{i}\leq C^{\prime}\alpha(i)2^{k_{i}},

where ni=max{fi,gi}n_{i}=\max\{\|f_{i}\|,\|g_{i}\|\}. Following the previous discussion, we see that there are constants C2,C2+>0C_{2}^{-},C_{2}^{+}>0 such that 2C3qini2C3+qi2^{C_{3}^{-}q_{i}}\leq n_{i}\leq 2^{C_{3}^{+}q_{i}}. In particular, we see that qilog(Cni)q_{i}\leq\log(Cn_{i}) for some C>0C>0. Therefore, qi2kilog(Cni)Cniq_{i}^{2^{k_{i}}}\geq\log(C^{-}n_{i})^{C^{-}n_{i}}. Thu,s we constructed an infinite sequence of non-conjugate elements fi,gif_{i},g_{i}\in\mathbb{Z}\wr\mathbb{Z} that are conjugate in every finite quotient of \mathbb{Z}\wr\mathbb{Z} of size smaller than log(Cni)Cmax{fi,gi}\log(C^{-}n_{i})^{C^{-}\max\{\|f_{i}\|,\|g_{i}\|\}} where C>0C^{-}>0 is some constant. Subsequently, we see that

log(Cni)Cmax{fi,gi}<CD(fi,gi).\log(C^{-}n_{i})^{C^{-}\max\{\|f_{i}\|,\|g_{i}\|\}}<\operatorname{CD}_{\mathbb{Z}\wr\mathbb{Z}}(f_{i},g_{i}).

Therefore,

(log(n))nConj(n),(\log(n))^{n}\preceq\operatorname{Conj}_{\mathbb{Z}\wr\mathbb{Z}}(n),

which concludes the proof. ∎

4. Upper bounds

The aim of this section is to construct upper bounds for the conjugacy depth function of a wreath product ABA\wr B of finitely generated abelian groups. The idea is to show that we can always find a quotient of the acting group BB such that Lemma 2.7 can be used to demonstrate that the images of the elements are not conjugate and provide asymptotic bounds on the size of this quotient. Recall that one of the assumptions of Lemma 2.7 is that we are working with reduced elements, i.e. the elements of the supports lie in distinct cosets of the acting element. Thus, in order to ensure we are working with reduced elements, 4.1 we show how to construct a finite quotient of the acting group that separates finite subsets and infinite cyclic subgroups. Subsection 4.2 then deals with the conditions that Lemma 2.7 uses to establish non-conjugacy. In particular, we show that if a quotient of a finitely generated abelian group is of sufficient size, then certain finite subsets do not become translates of each other in the quotient. Finally, subsection 4.3 combines these methods to construct a finite quotient preserving non-conjugacy of our given non-conjugate elements and gives an upper bound on its size in terms of their word lengths.

Before we proceed, we recall some notation. If BB is a finitely generated abelian group, we by fixing a splitting may write B=kTor(B)B=\mathbb{Z}^{k}\oplus\operatorname{Tor}(B) where Tor(B)\operatorname{Tor}(B) is the subgroup of finite order elements of BB and kk is the torsion-free rank of BB. Letting ϕ:Bk\phi\colon B\to\mathbb{Z}^{k} and τ:BTor(B)\tau\colon B\to\operatorname{Tor}(B) denote the natural projections associated to the fixed splitting, we may then write every xBx\in B uniquely as x=ϕ(x)+τ(x)x=\phi(x)+\tau(x) where we refer to ϕ(x)\phi(x) as the torsion-free part of xx and τ(x)\tau(x) as the torsion part of xx. When given a vector b=(b1,,bk)kb=(b_{1},\dots,b_{k})\in\mathbb{Z}^{k}, we denote gcd(b)=gcd(b1,,bk).\gcd(b)=\gcd(b_{1},\dots,b_{k}). Given two real numbers a<ba<b, we let [a,b][a,b] denote closed interval from aa to bb. Given two vectors v,wkv,w\in\mathbb{R}^{k}, we denote their dot product as vwv\cdot w. Finally, for a finite group TT, we denote its exponent as exp(T).\exp(T).

4.1. Simultaneous cosets


In this subsection, we study effective separability of cosets of cyclic subgroups in finitely generated abelian groups. Given an infinite, finitely generated abelian group GG, an element bGb\in G, and a finite subset SBG()S\subseteq B_{G}(\ell), we give an upper bound in terms of b\|b\| and \ell on the size of a finite quotient of the group GG such that each pair of cosets of the cyclic subgroup generated by bb corresponding to two distinct elements in SS remain distinct. In the following arguments, we use the observation that s1b=s2bs_{1}\langle b\rangle=s_{2}\langle b\rangle if and only if s11s2bs_{1}^{-1}s_{2}\in\langle b\rangle.

The following lemma is important for the proof of Lemma 4.5.

Lemma 4.1.

Let bb\in\mathbb{Z} satisfy b[n,n]b\in[-n,n] and SS\subseteq\mathbb{Z} be a subset such that S[Cn,Cn]S\subseteq[-Cn,Cn] for some constant C>0C>0. Suppose that cc is a natural number where |b|c>CN|b|c>CN, and let m=2|b|cm=2|b|c. Finally, let π:/m\pi\colon\mathbb{Z}\to\mathbb{Z}/m\mathbb{Z} be the natural projection.

Then for every sSs\in S, we have that

π(s)π(b) in /m if and only if sb in .\pi(s)\in\langle\pi(b)\rangle\text{ in }\mathbb{Z}/m\mathbb{Z}\quad\text{ if and only if }\quad s\in\langle b\rangle\text{ in }\mathbb{Z}.

Furthermore, if π(s)π(b)\pi(s)\in\langle\pi(b)\rangle, then π(s)=tπ(b)\pi(s)=t\pi(b) for the smallest integer tt with respect to the absolute value such that s=tbs=tb. In particular, |t|m|t|\leq m.

Proof.

Observe that the map π:/m\pi\colon\mathbb{Z}\to\mathbb{Z}/m\mathbb{Z} is injective on the interval [c|b|1,c|b|][-c|b|-1,c|b|]. We then note that

π(b)=π(b)=π({(c1)|b|,(c2)|b|,,|b|,0,|b|,,(c1)|b|,c|b|}).\pi(\langle b\rangle)=\pi(b\mathbb{Z})=\pi\left(\{-(c-1)|b|,-(c-2)|b|,\dots,-|b|,0,|b|,\dots,(c-1)|b|,c|b|\}\right).

Thus, if π(s)π(b)\pi(s)\in\pi(\langle b\rangle) for some sSs\in S, then sbs\in\langle b\rangle since the map π\pi is injective on the interval [c|b|1,c|b|][-c|b|-1,c|b|].

Finally, suppose that π(s)π(b)\pi(s)\in\langle\pi(b)\rangle and that π(s)=aπ(b)\pi(s)=a\pi(b) in /m\mathbb{Z}/m\mathbb{Z} where aa\in\mathbb{Z} is the smallest such value with respect to the absolute value. Following the previous argument, it follows that ab[c|b|1,c|b|]ab\in[-c|b|-1,c|b|], and therefore, we have s=abs=ab in \mathbb{Z}. ∎

To deal with the higher-dimensional cases, we first prove two technical lemmas. This first lemma gives bounds of lengths of a free generating basis for the kernel of the linear map given by the dot product with a vector in terms of size of the entries of the vector. For this lemma, when given vectors v1,,vknv_{1},\ldots,v_{k}\in\mathbb{Z}^{n}, we denote v1,,vk\langle v_{1},\ldots,v_{k}\rangle as the subgroup generated by the set {v1,,vk}.\{v_{1},\ldots,v_{k}\}.

Lemma 4.2.

Let b=(b1,,bk)kb=(b_{1},\dots,b_{k})\in\mathbb{Z}^{k} be non-trivial, and let φb:k\varphi_{b}\colon\mathbb{Z}^{k}\to\mathbb{Z} be the homomorphism given by φb(u)=ub\varphi_{b}(u)=u\cdot b. Then there are vectors λ(1),,λ(k1)k\lambda^{(1)},\dots,\lambda^{(k-1)}\in\mathbb{Z}^{k} such that ker(φb)=λ(1),,λ(k1)\ker(\varphi_{b})=\langle\lambda^{(1)},\dots,\lambda^{(k-1)}\rangle and λ(i)2k1b\|\lambda^{(i)}\|\leq 2^{k-1}\|b\| for all ii.

Proof.

We define vectors b(1),,b(k1)b^{(1)},\dots,b^{(k-1)} in the following way:

b(1)\displaystyle b^{(1)} =(b2,b1,0,,0),\displaystyle=(-b_{2},b_{1},0,\dots,0),
\displaystyle\vdots
b(k1)\displaystyle b^{(k-1)} =(0,,0,bk,bk1).\displaystyle=(0,\dots,0,-b_{k},b_{k-1}).

We set

λ(1)=1gcd(b2,b1)b(1),\lambda^{(1)}=\frac{1}{\gcd(-b_{2},b_{1})}b^{(1)},

and we note that if k=2k=2, then ker(φb)=λ(1)\ker(\varphi_{b})=\langle\lambda^{(1)}\rangle. Since λ(1)b\|\lambda^{(1)}\|\leq\|b\|, we are done. For k>2k>2, we will inductively build a generating set for ker(φb)\ker(\varphi_{b}) satisfying the statement of the lemma. We start with some basic observations.

By construction, we have that b(1),,b(k1)ker(φb)b^{(1)},\dots,b^{(k-1)}\in\ker(\varphi_{b}). Let Λi\Lambda_{i} be the maximal subgroup of k\mathbb{Z}^{k} of rank ii that contains b(1),,b(i)b^{(1)},\dots,b^{(i)}. Since the vectors b(1),,b(k1)b^{(1)},\dots,b^{(k-1)} are linearly independent over \mathbb{R}, we immediately see that Λ1Λk1=ker(φb)\Lambda_{1}\leq\dots\leq\Lambda_{k-1}=\ker(\varphi_{b}) and that Λi/Λi1\Lambda_{i}/\Lambda_{i-1}\simeq\mathbb{Z} for every i=2,,k1i=2,\dots,k-1.

Now assume that we already have a set of generators for Λi1\Lambda_{i-1} which we denote as λ(1),,λ(i1)\lambda^{(1)},\dots,\lambda^{(i-1)}. By construction, the elements {λ(1),,λ(i1)}\{\lambda^{(1)},\dots,\lambda^{(i-1)}\} satisfy Λj=λ1,,λj\Lambda_{j}=\langle\lambda_{1},\dots,\lambda_{j}\rangle for all j<ij<i where λj2i1jbj\|\lambda_{j}\|\leq 2^{i-1-j}\|b_{j}\| for 1ji1.1\leq j\leq i-1. Denote Li=Λi1,b(i)L_{i}=\langle\Lambda_{i-1},b^{(i)}\rangle. Since Λi1Λi1,b(i)Λi\Lambda_{i-1}\leq\langle\Lambda_{i-1},b^{(i)}\rangle\leq\Lambda_{i}, we see that Λi/Li\Lambda_{i}/L_{i} is a finite cyclic group. Furthermore, a preimage of some of its generator must be contained within the ii-dimensional parallelogram given by the vectors λ(1),,λ(i1),b(i)\lambda^{(1)},\dots,\lambda^{(i-1)},b^{(i)}. In particular, we see that

λ(i)b(i)+j=1i1λj.\|\lambda^{(i)}\|\leq\|b^{(i)}\|+\sum_{j=1}^{i-1}\|\lambda^{j}\|.

One can then easily check that

λ(i)b(i)+j=1i12i1jb(j).\|\lambda^{(i)}\|\leq\|b^{(i)}\|+\sum_{j=1}^{i-1}2^{i-1-j}\|b^{(j)}\|.

Noting that b(i)=|bi|+|bi+1|\|b^{(i)}\|=|b_{i}|+|b_{i+1}|, we see that λ(i)2i1b\|\lambda^{(i)}\|\leq 2^{i-1}\|b\| as desired. ∎

This next lemma implies any vector in k\mathbb{Z}^{k} whose entries have greatest common denominator as 11 is a part of a free base of k\mathbb{Z}^{k}. This lemma also shows that there exists an matrix TGLk()T\in\operatorname{GL}_{k}(\mathbb{Z}) which sends the vector to an element of the canonical basis and gives a bound on how much the matrix TT stretches the unit cube in k\mathbb{R}^{k} in terms of the size of the entries of the vector.

Lemma 4.3.

Let k>1k>1, and suppose that b=(b1,,bk)kb=(b_{1},\dots,b_{k})\in\mathbb{Z}^{k} is a vector where gcd(b1,,bk)=1\gcd(b_{1},\ldots,b_{k})=1. Then bb belongs to some free base of k\mathbb{Z}^{k}, and moreover, there exists a matrix TGLk()T\in\operatorname{GL}_{k}(\mathbb{Z}) such that

T(b)=(1,0,,0) and T(i=1k[1,1])i=1k[2k1kb,2k1kb].T(b)=(1,0,\dots,0)\quad\text{ and }\quad T\left(\prod_{i=1}^{k}[-1,1]\right)\subseteq\prod_{i=1}^{k}[2^{k-1}k\|b\|,2^{k-1}k\|b\|].
Proof.

[15, Theorem 9] implies there are integers a1,aka_{1},\dots a_{k}\in\mathbb{Z} such that

i=1kaibi=gcd(b1,,bk)=1\sum_{i=1}^{k}a_{i}b_{i}=\gcd(b_{1},\cdots,b_{k})=1

and where max{|ai|}12max{|bi|}\max\{|a_{i}|\}\leq\frac{1}{2}\max\{|b_{i}|\}. Thus, we denote a=(a1,,ak)ka=(a_{1},\dots,a_{k})\in\mathbb{Z}^{k}.

Let φb:k\varphi_{b}\colon\mathbb{Z}^{k}\to\mathbb{Z} be the linear map given by φb(x)=xb\varphi_{b}(x)=x\cdot b. Lemma 4.2 implies there are vectors

λ(1),,λ(k1)Bk(2k1n)i=1k[2k1b,2k1b]\lambda^{(1)},\dots,\lambda^{(k-1)}\in\operatorname{B}_{\mathbb{Z}^{k}}(2^{k-1}n)\subseteq\prod_{i=1}^{k}[2^{k-1}\|b\|,2^{k-1}\|b\|]

that freely generate ker(φb)\ker(\varphi_{b}). We can then form the matrix TT by setting the first row to be equal to the vector aa and the remaining k1k-1 vectors to be equal to the vectors λ(1),,λ(k1)\lambda^{(1)},\dots,\lambda^{(k-1)}, respectively. By construction, we see that T(b)=(1,0,,0)T(b)=(1,0,\dots,0). Since Im(φb)=φb(a)\operatorname{Im}(\varphi_{b})=\langle\varphi_{b}(a)\rangle, we see that k=bker(φb)\mathbb{Z}^{k}=\langle b\rangle\oplus\ker(\varphi_{b}) which implies that the row vectors of TT generate k\mathbb{Z}^{k}. Therefore, TGLk()T\in\operatorname{GL}_{k}(\mathbb{Z}).

To finish the proof, we recall that

ai=1k[12b,12b] and λ(i)2k1ba\in\prod_{i=1}^{k}\left[-\frac{1}{2}\|b\|,\frac{1}{2}\|b\|\right]\quad\text{ and }\quad\|\lambda^{(i)}\|\leq 2^{k-1}\|b\|

for all ii. In particular, this means that for every i,j{1,,k}i,j\in\{1,\dots,k\} we have the (i,j)(i,j)-th entry of TT which we denote as Ti,jT_{i,j} satisfies |Ti,j|2k1b|T_{i,j}|\leq 2^{k-1}\|b\|. Therefore, we have

T(i=1k[1,1])i=1k[2k1kb,2k1kb].T\left(\prod_{i=1}^{k}[-1,1]\right)\subseteq\prod_{i=1}^{k}\left[-2^{k-1}k\|b\|,2^{k-1}k\|b\|\right].\qed

The following corollary is not consequential for this paper, but we feel it an interesting result in its own right.

Corollary 4.4.

Let b=(b1,,bk)b=(b_{1},\dots,b_{k})\in\mathbb{Z} be given such that gcd(b1,,bk)=1\gcd(b_{1},\dots,b_{k})=1. Then there are elements λ(1),,λ(k1)k\lambda^{(1)},\dots,\lambda^{(k-1)}\in\mathbb{Z}^{k} such that the set {b,λ(1),,λ(k1)}\{b,\lambda^{(1)},\dots,\lambda^{(k-1)}\} is a free base of k\mathbb{Z}^{k} and λ(i)2kb\|\lambda^{(i)}\|\leq 2^{k}\|b\|.

For a vector bkb\in\mathbb{Z}^{k}, this next lemma gives bounds on the size of the integer mm we reduce entries in k\mathbb{Z}^{k} mod mm to preserves cosets of the infinite cyclic subgroup generated by bb in terms of the size of the entries in bb.

Lemma 4.5.

Let k>1k>1 and nn\in\mathbb{N} be fixed. Let SkS\subseteq\mathbb{Z}^{k} and b=(b1,,bk)kb=(b_{1},\dots,b_{k})\in\mathbb{Z}^{k} satisfy bBk(n)b\in\operatorname{B}_{\mathbb{Z}^{k}}(n) and where SBk(Cn)S\subseteq\operatorname{B}_{\mathbb{Z}^{k}}(Cn) for some C>0C>0. Suppose mm is an integer satisfying the following:

  1. (1)

    mm is divisible by gcd(b)\gcd(b);

  2. (2)

    m>2kkCn2m>2^{k}kCn^{2};

  3. (3)

    the homomorphism π:k(/m)k\pi\colon\mathbb{Z}^{k}\to\left(\mathbb{Z}/m\mathbb{Z}\right)^{k} given by reducing every mod mm is injective on the set SS.

Then for every sSs\in S we have that π(s)π(b)\pi(s)\in\langle\pi(b)\rangle if and only if sbs\in\langle b\rangle.

Furthermore, if π(s)π(b)\pi(s)\in\langle\pi(b)\rangle, then there is an integer tt\in\mathbb{Z} such that s=tbs=tb and |t|m/c|t|\leq m/c.

Proof.

Set b=1cbb^{\prime}=\frac{1}{c}b and denote b=(b1,,bk)b^{\prime}=(b_{1}^{\prime},\dots,b_{k}^{\prime}). Since gcd(b1,,bk)=1\gcd(b_{1}^{\prime},\dots,b_{k}^{\prime})=1, Lemma 4.3 implies there exists a matrix TGLd()T\in\operatorname{GL}_{d}(\mathbb{Z}) such that T(b)=e1T(b^{\prime})=e_{1} and

T(i=1k[1,1])i=1k[2k1kn,2k1kn]T\left(\prod_{i=1}^{k}[-1,1]\right)\subseteq\prod_{i=1}^{k}[-2^{k-1}kn,2^{k-1}kn]

where {e1,,ek}\{e_{1},\dots,e_{k}\} is the canonical free basis of k\mathbb{Z}^{k}. Since TT is an automorphism of d\mathbb{Z}^{d}, we see that T(s)T(b)=ce1T(s)\in\langle T(b)\rangle=\langle ce_{1}\rangle if and only if sbs\in\langle b\rangle. We note that

T(S)T(Bk(Cn))T(i=1k[Cn,Cn])i=1k[2k1kCn2,2k1kCn2].T(S)\subseteq T\left(B_{\mathbb{Z}^{k}}(Cn)\right)\subseteq T\left(\prod_{i=1}^{k}[-Cn,Cn]\right)\subseteq\prod_{i=1}^{k}\left[-2^{k-1}kCn^{2},2^{k-1}kCn^{2}\right].

Set m=2clm=2cl where ll\in\mathbb{N} is the smallest natural number such that cl>2k1kCn2cl>2^{k-1}kCn^{2}, and denote K=mkkK=m\mathbb{Z}^{k}\leq\mathbb{Z}^{k}. By construction, we have that m2k+1kCn2m\leq 2^{k+1}kCn^{2}. Therefore, we see that the projection πK:kk/mk\pi_{K}\colon\mathbb{Z}^{k}\to\mathbb{Z}^{k}/m\mathbb{Z}^{k} is injective on the hypercube i=1k[m+1,m]\prod_{i=1}^{k}[-m+1,m] where πK\pi_{K} is the reduction of each coordinate mod mm. In particular, since Si=1k[m+1,m]S\subseteq\prod_{i=1}^{k}[-m+1,m], for any sSs\in S we have that T(s)(mk)e1KT(s)\left(m\mathbb{Z}^{k}\right)\subseteq\langle e_{1}\rangle K if and only if T(s)e1T(s)\in\langle e_{1}\rangle. It then follows that π(s)π(b)\pi(s)\notin\langle\pi(b)\rangle whenever sbs\notin\langle b^{\prime}\rangle.

Now suppose that sbs\in\langle b^{\prime}\rangle, i.e. T(s)e1T(s)\in\langle e_{1}\rangle. In this case, we may retract onto the first coordinate and assume that we are working in \mathbb{Z}. The rest of the statement then follows by Lemma 4.1. ∎

This next lemma extends Lemma 4.5 to when the abelian group has torsion.

Lemma 4.6.

Let BB be a finitely generated infinite abelian group of torsion free rank kk. Let bBB(n)b\in\operatorname{B}_{B}(n) and SBB(Cn)S\subset\operatorname{B}_{B}(Cn) be given for some constant C>0C>0.

If k=1k=1, assume that mm\in\mathbb{N} satisfies m2Cnm\geq 2Cn and where both ϕ(b)\|\phi(b)\| and exp(T)\exp(T) divide mm. If k2k\geq 2, assume that mm\in\mathbb{N} is such that mk2kCn2m\geq k2^{k}Cn^{2} and both c=gcd(ϕ(b))c=\gcd(\phi(b)) and exp(Tor(B))\exp(\operatorname{Tor}(B)) divide mm. Then the homomorphism

π:kTor(B)(/m)kTor(B)\pi\colon\mathbb{Z}^{k}\oplus\operatorname{Tor}(B)\to(\mathbb{Z}/m\mathbb{Z})^{k}\oplus\operatorname{Tor}(B)

defined as the identity on Tor(B)\operatorname{Tor}(B) and as the coordinate-wise projection on k\mathbb{Z}^{k} is injective on the set SS and for every sSs\in S we have that π(s)π(b)\pi(s)\in\langle\pi(b)\rangle if and only if sbs\in\langle b\rangle.

Proof.

Lemma 4.5 implies we may assume that Tor(B)0\operatorname{Tor}(B)\neq 0. Therefore, set e=exp(Tor(B))e=\exp(\operatorname{Tor}(B)).

The main argument of the proof when k=1k=1 is analogous to the case when k2k\geq 2, but instead of Lemma 4.5 one would use Lemma 4.1. For this reason, we leave proof in the case when k=1k=1 as an exercise.

Denote e=exp(Tor(B)),e=\exp(\operatorname{Tor}(B)), and suppose that m>0m>0 and π:kTor(B)(/m)kTor(B)\pi\colon\mathbb{Z}^{k}\oplus\operatorname{Tor}(B)\to(\mathbb{Z}/m\mathbb{Z})^{k}\oplus\operatorname{Tor}(B) are as in the statement of the lemma. Assuming that π(s)π(b)\pi(s)\in\langle\pi(b)\rangle for some sSs\in S, there is some tt\in\mathbb{N} such that π(s)=tπ(b)\pi(s)=t\pi(b) which we pick to be as small possible. In particular, we see that tgcd(m,e)=mt\leq\gcd(m,e)=m.

We write:

π(s)\displaystyle\pi(s) =tπ(b)\displaystyle=t\pi(b)
τ(s)+π(ϕ(s))\displaystyle\tau(s)+\pi(\phi(s)) =tτ(s))+tπ(ϕ(b)),\displaystyle=t\tau(s))+t\pi(\phi(b)),

from which immediately see that τ(s)=tτ(b)\tau(s)=t\tau(b) and π(ϕ(s))=tπ(ϕ(b))\pi(\phi(s))=t\pi(\phi(b)). Following Lemma 4.5, we see that ϕ(s)=tϕ(b)\phi(s)=t\phi(b). Therefore, we may write

s=tτ(b)+tϕ(b)=tb.s=t\tau(b)+t\phi(b)=tb.\qed

4.2. Translations


We say that an ordered list X=(x1,xm)GmX=(x_{1},\dots x_{m})\subseteq G^{m} is a translate of an ordered list Y=(y1,,yn)GmY=(y_{1},\dots,y_{n})\subseteq G^{m} if n=mn=m and there exists a permutation σSym(n)\sigma\in\operatorname{Sym}(n) such that

xσ(1)y11==xσ(n)yn1.x_{\sigma(1)}y_{1}^{-1}=\dots=x_{\sigma(n)}y_{n}^{-1}.

In this case, we say that σ\sigma realises a translation of XX onto YY.

We have the following lemma which gives conditions of when two sets in a product of groups are translates of each other in terms of translations of their images in the projection onto the factor groups.

Lemma 4.7.

Let G1,G2G_{1},G_{2} be groups, set G=G1×G2G=G_{1}\times G_{2} and let π1,π2\pi_{1},\pi_{2} denote the canonical projections π1:G1×G2G1\pi_{1}\colon G_{1}\times G_{2}\to G_{1} and π2:G1×G2G2\pi_{2}\colon G_{1}\times G_{2}\to G_{2}. Suppose that X,YGX,Y\subseteq G are two finite subsets where X={x1,,x}X=\{x_{1},\ldots,x_{\ell}\} and Y={y1,,ym}Y=\{y_{1},\ldots,y_{m}\}. Then XX is a translate of YY if and only if the following two properties hold:

  1. (1)

    |X|=|Y|=n|X|=|Y|=n

  2. (2)

    There exists σSym(n)\sigma\in\operatorname{Sym}(n) such that σ\sigma simultaneously realises a translation of the list (π1(x1),,π1(xn))(\pi_{1}(x_{1}),\dots,\pi_{1}(x_{n})) onto the list (π1(y1),,π1(yn))(\pi_{1}(y_{1}),\dots,\pi_{1}(y_{n})) and a translation of the list (π2(x1),,π2(xn))(\pi_{2}(x_{1}),\dots,\pi_{2}(x_{n})) onto the list (π2(y1),,π2(yn))(\pi_{2}(y_{1}),\dots,\pi_{2}(y_{n})).

Proof.

It is straightforward to see that XX is a translate of YY then |X|=|Y||X|=|Y|. Thus, we may assume that |X|,|Y|=n|X|,|Y|=n for some nn\in\mathbb{N}. As mentioned above, we have that XX is a translate of YY if and only if there is σSym(n)\sigma\in\operatorname{Sym}(n) such that

xσ(1)y11==xσ(n)yn1.x_{\sigma(1)}y_{1}^{-1}=\dots=x_{\sigma(n)}y_{n}^{-1}.

If terms of Cartesian coordinates, this means that

π1(xσ(1))π1(y11)==π1(xσ(n))π1(yn1),\displaystyle\pi_{1}(x_{\sigma(1)})\pi_{1}(y_{1}^{-1})=\dots=\pi_{1}(x_{\sigma(n)})\pi_{1}(y_{n}^{-1}),
π2(xσ(1))π2(y11)==π2(xσ(n))π2(yn1).\displaystyle\pi_{2}(x_{\sigma(1)})\pi_{2}(y_{1}^{-1})=\dots=\pi_{2}(x_{\sigma(n)})\pi_{2}(y_{n}^{-1}).

That is equivalent to saying that σ\sigma realises a translation of (π1(x1),,π1(xn))(\pi_{1}(x_{1}),\dots,\pi_{1}(x_{n})) onto (π1(y1),,π1(yn))(\pi_{1}(y_{1}),\dots,\pi_{1}(y_{n})) and that σ\sigma realises a translation of the list (π2(x1),,π2(xn))(\pi_{2}(x_{1}),\dots,\pi_{2}(x_{n})) onto the list (π2(y1),,π2(yn))(\pi_{2}(y_{1}),\dots,\pi_{2}(y_{n})). ∎

This next lemma tells us there exists a constant \ell such that when given two finite subsets X,YX,Y in k\mathbb{Z}^{k} whose coordinates of each element have absolute value at most \ell, then XX and YY are translations of each other if and only if their images are translations in the group (/4)k(\mathbb{Z}/4\ell\mathbb{Z})^{k} where we reduce each coordinate mod 4.4\ell.

Lemma 4.8.

Suppose that X=(x1,,xn)X=(x_{1},\dots,x_{n}) and Y=(y1,,yn)Y=(y_{1},\dots,y_{n}) are two finite ordered lists in k\mathbb{Z}^{k} such that

X,Yi=1d[(1),1]X,Y\subseteq\prod_{i=1}^{d}[-(\ell-1),\ell-1]

for some \ell\in\mathbb{N}, and suppose that c4c\geq 4\ell Let π:k(/c)k\pi\colon\mathbb{Z}^{k}\to\left(\mathbb{Z}/c\mathbb{Z}\right)^{k} be the homomorphism given by reducing each coordinate mod cc. Then the following are equivalent:

  1. (1)

    π(X)=(π(x1),,π(xn))\pi(X)=(\pi(x_{1}),\dots,\pi(x_{n})) is a translate of π(Y)=(π(y1),,π(yn))\pi(Y)=(\pi(y_{1}),\dots,\pi(y_{n})) in (/c)k\left(\mathbb{Z}/c\mathbb{Z}\right)^{k}.

  2. (2)

    XX is a translate of YY in k\mathbb{Z}^{k}.

Furthermore, for all σSym(n)\sigma\in\operatorname{Sym}(n), we have that σ\sigma realises a translation of π(X)\pi(X) onto π(Y)\pi(Y) if and only if σ\sigma realises a translation of XX onto YY.

Proof.

We will only prove the ‘furthermore part of the statement since the first part follows from it.

Since the image of a translate is a translate of an image, the implication from left to right holds trivially. Therefore, we need only consider the other direction. Suppose that c4c\geq 4\ell. For every σSym(n)\sigma\in\operatorname{Sym}(n), define

Dσ(X,Y)={xσ(i)yii=1,,n},D_{\sigma}(X,Y)=\{x_{\sigma(i)}-y_{i}\mid i=1,\dots,n\},

and define Dσ(π(X),π(Y))D_{\sigma}(\pi(X),\pi(Y)) analogously. We observe that σ\sigma realises a translation of XX onto YY if and only if |Dσ(X,Y)|=1|D_{\sigma}(X,Y)|=1. Similarly, σ\sigma realises a translation of π(X)\pi(X) onto π(Y)\pi(Y) if and only if |Dσ(π(X),π(Y))|=1|D_{\sigma}(\pi(X),\pi(Y))|=1. We also have that

Dσ(X,Y)i=1k[2+1,21].D_{\sigma}(X,Y)\subseteq\prod_{i=1}^{k}[-2\ell+1,2\ell-1].

We see that π\pi is injective on i=1k[2+1,21]\prod_{i=1}^{k}[-2\ell+1,2\ell-1]. Hence, π\pi is injective on Dσ(X,Y)D_{\sigma}(X,Y) for every σSym(n)\sigma\in\operatorname{Sym}(n). Since π\pi is a homomorphism, we see that Dσ(π(X),π(Y))=π(Dσ(X,Y))D_{\sigma}(\pi(X),\pi(Y))=\pi\left(D_{\sigma}(X,Y)\right) for all σSym(n)\sigma\in\operatorname{Sym}(n). Hence, we have that |Dσ(π(X),π(Y))|=|Dσ(X,Y)||D_{\sigma}(\pi(X),\pi(Y))|=|D_{\sigma}(X,Y)|. In particular, we see that

|Dσ(π(X),π(Y))|=1 if and only if |Dσ(X,Y)|=1.|D_{\sigma}(\pi(X),\pi(Y))|=1\quad\text{ if and only if }\quad|D_{\sigma}(X,Y)|=1.

Thus, σ\sigma realises a translation of π(X)\pi(X) onto π(Y)\pi(Y) if and only if σ\sigma realises a translation of XX onto YY. ∎

The last lemma of this subsection extends Lemma 4.8 to infinite finitely generated abelian groups with torsion.

Lemma 4.9.

Let BB be a finitely generated abelian group of torsion-free rank kk, and suppose that X,YBB()X,Y\subseteq\operatorname{B}_{B}(\ell) are given. If c4c\geq 4\ell, then the homomorphism π:B(/c)kTor(B)\pi\colon B\to\left(\mathbb{Z}/c\mathbb{Z}\right)^{k}\oplus\operatorname{Tor}(B) given by the identity on Tor(B)\operatorname{Tor}(B) and by the coordinate-wise projection mod cc on the torsion-free part of BB is injective on XYX\cup Y.

Moreover, π(X)\pi(X) is a translate of π(Y)\pi(Y) if and only if XX is a translate of YY in k\mathbb{Z}^{k}. Furthermore, for all permutations σ\sigma, we have that σ\sigma realises a translation of π(X)\pi(X) onto π(Y)\pi(Y) if and only if σ\sigma realises a translation of XX onto YY.

Proof.

By assumption, we have

π(X),π(Y)i=1k[,]\pi(X),\pi(Y)\subseteq\prod_{i=1}^{k}[-\ell,\ell]

which implies that π\pi is injective on XYX\cup Y. In particular, if |X||Y||X|\neq|Y|, then |π(X)||π(Y)||\pi(X)|\neq|\pi(Y)|, and subsequently, π(X)\pi(X) is not a translate of π(Y)\pi(Y). Therefore, we may assume that |X|=|Y||X|=|Y|.

Let {x1,,xm}\{x_{1},\dots,x_{m}\} and {y1,,ym}\{y_{1},\dots,y_{m}\} be enumerations of XX and YY, respectively. Following Lemma 4.7, we see that π(X)\pi(X) is a translate of π(Y)\pi(Y) if and only if the list given by {π(ϕ(x1)),,π(ϕ(xm))}\{\pi(\phi(x_{1})),\dots,\pi(\phi(x_{m}))\} is a translation of the list given by {π(ϕ(y1)),,π(ϕ(ym))}\{\pi(\phi(y_{1})),\dots,\pi(\phi(y_{m}))\} and the list {π(τ(xm)),,π(τ(xm))}\{\pi(\tau(x_{m})),\dots,\pi(\tau(x_{m}))\} is a translation of the list {π(τ(y1)),,π(τ(ym))}\{\pi(\tau(y_{1})),\dots,\pi(\tau(y_{m}))\} where the translation is realised by the same permutation. However, by Lemma 4.8, we see that a permutation σSym(m)\sigma\in\operatorname{Sym}(m) realises a translation of the list given by {π(ϕ(x1)),π(ϕ(xm))}\{\pi(\phi(x_{1})),\dots\pi(\phi(x_{m}))\} onto the list {π(ϕ(y1)),π(ϕ(ym))}\{\pi(\phi(y_{1})),\dots\pi(\phi(y_{m}))\} if and only if it realises a translation of the list given by {ϕ(x1),,ϕ(xm)}\{\phi(x_{1}),\dots,\phi(x_{m})\} onto the list given by {ϕ(y1),,ϕ(ym)}\{\phi(y_{1}),\dots,\phi(y_{m})\}. Since π\pi is defined as the identity on Tor(B)\operatorname{Tor}(B), we see that π(X)\pi(X) is a translate of π(Y)\pi(Y) if and only if XX is a translate of YY, which concludes the proof. ∎

4.3. Finite base groups


Let AA be a finite abelian group, and let BB be an infinite, finitely generated abelian group. Using Lemma 4.6 and Lemma 4.2, the next proposition demonstrates when given non-conjugate elements x,yx,y in a s ABA\wr B that there exists a finite quotient B¯\overline{B} of BB such that the images of xx and yy in AB¯A\wr\overline{B} remain non-conjugate. Moreover, this lemma gives a bound of the size of the quotient of BB in terms of the word lengths of xx and yy.

Proposition 4.10.

Let AA be an abelian group and BB be an infinite, finitely generated abelian group. Let f,g:BAf,g\colon B\to A be finitely supported functions and bBb\in B an element such that fb,gbBAB(n)fb,gb\in\operatorname{B}_{A\wr B}(n) and fb≁ABgbfb\not\sim_{A\wr B}gb. Then there exists a surjective homomorphism π:BB¯\pi\colon B\to\overline{B} to a finite group such that π~(fb)≁π~(gb)\tilde{\pi}(fb)\not\sim\tilde{\pi}(gb) in AB¯A\wr\overline{B}.

Moreover, there exists a constant C>0C>0 such that if BB has torsion free rank 11, then we have |B¯|Cn|\bar{B}|\leq Cn, and if BB is of torsion-free rank k>1k>1, we then have |B¯|Cn2k.|\bar{B}|\leq Cn^{2k}.

Proof.

Fix a splitting of BB into kTor(B)\mathbb{Z}^{k}\oplus\operatorname{Tor}(B) with associated associated free projection ϕ\phi. Following Lemma 2.6, we may assume that both the functions ff and gg are given such that the elements fbfb and gbgb are reduced, i.e. the individual elements of their respective supports lie in distinct cosets of b\langle b\rangle in BB.

Following Lemma 2.7, there are two cases to distinguish:

  1. (i)

    supp(f)\operatorname{supp}(f) is not a translate of supp(g)\operatorname{supp}(g) in BB,

  2. (ii)

    for every aBa\in B such that a+supp(f)=supp(g)a+\operatorname{supp}(f)=\operatorname{supp}(g), there exists some xsupp(g)x\in\operatorname{supp}(g) such that f(x+a)g(x)f(x+a)\neq g(x).

We will construct a finite quotient B¯\overline{B} such that the images of fbfb and gbgb are still reduced in AB¯A\wr\overline{B} whether (i) or (ii) is the case.

Lemma 2.9 implies that there is constant C1>0C_{1}>0 such that

{b}supp(f)supp(g)BB(C1n).\{b\}\cup\operatorname{supp}(f)\cup\operatorname{supp}(g)\subseteq B_{B}(C_{1}n).

In particular, we see that

ϕ(supp(f)),ϕ(supp(g))Bk(C1n)i=1k[C1n,C1n].\phi(\operatorname{supp}(f)),\phi(\operatorname{supp}(g))\subseteq B_{\mathbb{Z}^{k}}(C_{1}n)\subseteq\prod_{i=1}^{k}[-C_{1}n,C_{1}n].

Set =C1n\ell=C_{1}n. It then follows that

ϕ(supp(f)),ϕ(supp(g))i=1k[,].\phi(\operatorname{supp}(f)),\phi(\operatorname{supp}(g))\subseteq\prod_{i=1}^{k}[-\ell,\ell].

We set

S={s2s1s1,s2supp(f)supp(g)}S=\{s_{2}-s_{1}\mid s_{1},s_{2}\in\operatorname{supp}(f)\cup\operatorname{supp}(g)\}

and see that SBB(2)S\subseteq\operatorname{B}_{B}(2\ell) and bBB()\|b\|\in\operatorname{B}_{B}(\ell). Finally, we set e=exp(Tor(B)).e=\exp(\operatorname{Tor}(B)).

If k=1k=1, let mm be the smallest integer such that m>4m>4\ell and where both ee and ϕ(b)\|\phi(b)\| divide mm. It is then straightforward to see that m8m\leq 8\ell. If k2k\geq 2, let mm\in\mathbb{N} be smallest possible such that m>k2k22m>k2^{k}2\ell^{2} and where both ee and gcd(ϕ(b))\gcd(\phi(b)) divide mm. Without loss of generality, we may assume that gcd(ϕ(b))\gcd(\phi(b)) divides ϕ(b)\|\phi(b)\|. In particular, we see that m<k2k+122m<k2^{k+1}2\ell^{2}.

Via Lemma 4.6, we see that the homomorphism π:kTor(B)(/m)kTor(B)\pi\colon\mathbb{Z}^{k}\oplus\operatorname{Tor}(B)\to(\mathbb{Z}/m\mathbb{Z})^{k}\oplus\operatorname{Tor}(B) defined as the identity on Tor(B)\operatorname{Tor}(B) and as the coordinate-wise reduction mod mm on k\mathbb{Z}^{k} is injective on the set SS and for every sSs\in S we have that π(s)π(b)\pi(s)\in\langle\pi(b)\rangle if and only if sbs\in\langle b\rangle. In particular, this means that for every s,ssupp(f)supp(g)s,s^{\prime}\in\operatorname{supp}(f)\cup\operatorname{supp}(g) we have that π(s)π(b)=π(s)π(b)\pi(s)\langle\pi(b)\rangle=\pi(s^{\prime})\langle\pi(b)\rangle if and only if sb=sbs\langle b\rangle=s^{\prime}\langle b\rangle. Set B¯=(/m)kTor(B)\overline{B}=(\mathbb{Z}/m\mathbb{Z})^{k}\oplus\operatorname{Tor}(B), and let π~:ABAB¯\tilde{\pi}\colon A\wr B\to A\wr\overline{B} be the canonical extension of π\pi to the whole of ABA\wr B. From the construction of the map π\pi, we see that π~\tilde{\pi} is injective on supp(f)supp(g)\operatorname{supp}(f)\cup\operatorname{supp}(g). Therefore, it follows that supp(π~(f))=π(supp(f))\operatorname{supp}(\tilde{\pi}(f))=\pi(\operatorname{supp}(f)) and supp(π~(g))=π(supp(g))\operatorname{supp}(\tilde{\pi}(g))=\pi(\operatorname{supp}(g)). Furthermore, we see that for every two s,ssupp(f)supp(g)s,s^{\prime}\in\operatorname{supp}(f)\cup\operatorname{supp}(g) we have that π~(s)π~(b)=π~(s)π~(b)\tilde{\pi}(s)\langle\tilde{\pi}(b)\rangle=\tilde{\pi}(s^{\prime})\langle\tilde{\pi}(b)\rangle in B¯\overline{B} if and only if sb=sbs\langle b\rangle=s^{\prime}\langle b\rangle in BB. In particular, we see that the elements π~(fb)\tilde{\pi}(fb) and π~(gb)\tilde{\pi}(gb) are in reduced form.

Since the elements π~(fb)\tilde{\pi}(fb) and π~(gb)\tilde{\pi}(gb) are in reduced form, we may use Lemma 2.7 to check whether or not they are conjugate in AB¯A\wr\overline{B}. We note that regardless of whether k=1k=1 or k2k\geq 2, we have that m4m\geq 4\ell. Additionally, if |supp(f)||supp(g)||\operatorname{supp}(f)|\neq|\operatorname{supp}(g)|, then |supp(π~(f))||supp(π~(g))||\operatorname{supp}(\tilde{\pi}(f))|\neq|\operatorname{supp}(\tilde{\pi}(g))|. Subsequently, since supp(π~(f))\operatorname{supp}(\tilde{\pi}(f)) is not a translate of supp(π~(g))\operatorname{supp}(\tilde{\pi}(g)), we have that π~(fb)\tilde{\pi}(fb) is not conjugate to π~(fb)\tilde{\pi}(fb) by Lemma 2.7. Therefore, we may assume that |supp(f)|=|supp(g)||\operatorname{supp}(f)|=|\operatorname{supp}(g)|.

Let

supp(f)={x1,,xm} and supp(g)={y1,,ym}.\operatorname{supp}(f)=\{x_{1},\dots,x_{m}\}\quad\text{ and }\quad\operatorname{supp}(g)=\{y_{1},\dots,y_{m}\}.

Via Lemma 4.7, we see that supp(π~(f))\operatorname{supp}(\tilde{\pi}(f)) is a translate of supp(π~(g))\operatorname{supp}(\tilde{\pi}(g)) if and only if {π(ϕ(x1)),,π(ϕ(xm))}\{\pi(\phi(x_{1})),\dots,\pi(\phi(x_{m}))\} is a translation of {π(ϕ(y1)),,π(ϕ(ym))}\{\pi(\phi(y_{1})),\dots,\pi(\phi(y_{m}))\} and the set {π(τ(xm)),,π(τ(xm))}\{\pi(\tau(x_{m})),\dots,\pi(\tau(x_{m}))\} is a translation of {π(τ(y1)),,π(τ(ym))}\{\pi(\tau(y_{1})),\dots,\pi(\tau(y_{m}))\}. Moreover, the translation of both pairs of sets is realised by the same permutation. Lemma 4.8 implies that a permutation σSym(m)\sigma\in\operatorname{Sym}(m) realises a translation of {π(ϕ(x1)),π(ϕ(xm))}\{\pi(\phi(x_{1})),\dots\pi(\phi(x_{m}))\} onto {π(ϕ(y1)),π(ϕ(ym))}\{\pi(\phi(y_{1})),\dots\pi(\phi(y_{m}))\} if and only if it realises a translation of {ϕ(x1),,ϕ(xm)}\{\phi(x_{1}),\dots,\phi(x_{m})\} onto the list {ϕ(y1),,ϕ(ym)}\{\phi(y_{1}),\dots,\phi(y_{m})\}. Since π\pi is defined as the identity on Tor(B)\operatorname{Tor}(B), we see that supp((~π)(f))\operatorname{supp}(\tilde{(}\pi)(f)) is a translate of supp(π~(g))\operatorname{supp}(\tilde{\pi}(g)) if and only if supp(f)\operatorname{supp}(f) is a translate of supp(g)\operatorname{supp}(g). Therefore, if supp(f)\operatorname{supp}(f) is not a translate of supp(g)\operatorname{supp}(g) in BB, we see by Lemma 2.7 that π~(fb)\tilde{\pi}(fb) is not conjugate to π~(gb)\tilde{\pi}(gb). Thus, we may suppose that supp(π~(f))\operatorname{supp}(\tilde{\pi}(f)) is a translate of supp(π~(g))\operatorname{supp}(\tilde{\pi}(g)).

Now suppose that a+supp(f)=supp(g)a+\operatorname{supp}(f)=\operatorname{supp}(g) for some aBa\in B. By assumption, there exists some xsupp(g)x\in\operatorname{supp}(g) such that f(x+a)g(x)f(x+a)\neq g(x). As mentioned before, Lemma 4.8 implies that every translation of supp(π~(f))\operatorname{supp}(\tilde{\pi}(f)) onto supp(π~(g))\operatorname{supp}(\tilde{\pi}(g)) must have already occurred in BB. By the construction of π\pi, we see that for every xi=1k[m,m]×Tor(B)x\in\prod_{i=1}^{k}[-m,m]\times\operatorname{Tor}(B) we have π~(f)(x+K)=f(x)\tilde{\pi}(f)(x+K)=f(x) and π~(g)(x+K)=g(x)\tilde{\pi}(g)(x+K)=g(x). We see that for every aTor(B)×(/m)ka\in\operatorname{Tor}(B)\times(\mathbb{Z}/m\mathbb{Z})^{k} such that π(a)+supp(π~(f))=supp(π~(g))\pi(a)+\operatorname{supp}(\tilde{\pi}(f))=\operatorname{supp}(\tilde{\pi}(g)), there must exist x¯supp(π~(f))\overline{x}\in\operatorname{supp}(\tilde{\pi}(f)) such that π~(f)(π(a)+x¯)π~(g)(x)\tilde{\pi}(f)(\pi(a)+\overline{x})\neq\tilde{\pi}(g)(x). That means that π~(fb)\tilde{\pi}(fb) is not conjugate to π~(gc)\tilde{\pi}(gc).

If k=1k=1, set C=8C1e|Tor(B)|C=8C_{1}e|\operatorname{Tor}(B)|. We then have

|B¯|=m|Tor(B)|(8C1en)|Tor(B)|=Cn.\left|\overline{B}\right|=m|\operatorname{Tor}(B)|\leq\left(8C_{1}en\right)|\operatorname{Tor}(B)|=Cn.

If k2k\geq 2, set C=kk2k(k+1)C12ke2k|Tor(B)|C=k^{k}2^{k(k+1)}C_{1}^{2k}e^{2k}|\operatorname{Tor}(B)|. We then have

|B¯|=mk|Tor(B)|(k2k+12(C1n)2)k|Tor(B)|=Cn2k\left|\overline{B}\right|=m^{k}|\operatorname{Tor}(B)|\leq\left(k2^{k+1}2(C_{1}n)^{2}\right)^{k}|\operatorname{Tor}(B)|=Cn^{2k}

which concludes our proof. ∎

As an immediate consequence of Proposition 4.10, we get the following upper bound for wreath products of abelian groups with finite base group.

Proposition 4.11.

Let AA is a finite abelian group and BB is a finitely generated abelian group of torsion free rank kk. If k=1k=1,

ConjAB(n)2n.\operatorname{Conj}_{A\wr B}(n)\preceq 2^{n}.

Otherwise, for k>1k>1, we have

ConjAB(n)2n2k.\operatorname{Conj}_{A\wr B}(n)\preceq 2^{n^{2k}}.
Proof.

Suppose that f,bABf,b\in A^{B} are finitely supported functions and b,cAb,c\in A are elements such that fb,bcBG(n)fb,bc\in B_{G}(n) and where fb≁Ggcfb\not\sim_{G}gc.

Suppose first that bcb\neq c. Since bcBB(2n)b-c\in\operatorname{B}_{B}(2n), [4, Corollary 2.3] implies there exists a constant C1>0C_{1}>0 and a surjective homomorphism φ:BQ\varphi\colon B\to Q such that φ(b)φ(c)\varphi(b)\neq\varphi(c) and where |Q|C1log(C1n).|Q|\leq C_{1}\>\log(C_{1}n). Since QQ is abelian, we have that φ(b)\varphi(b) and φ(c)\varphi(c) are non-conjugate. By composing φ\varphi with the projection of ABA\wr B onto BB, which we also denote φ\varphi, we have a surjective homomorphism φ:ABQ\varphi\colon A\wr B\to Q such that φ(fb)φ(gc)\varphi(fb)\nsim\varphi(gc) and where |Q|C1log(C1n).|Q|\leq C_{1}\log(C_{1}n). Therefore, we may assume that b=c.b=c.

Following Proposition 4.10, we have two cases. When the torsion free rank is 11, we see that there exists a finite abelian group B¯\overline{B} together with a surjective homomorphism ϕ:ABAB¯\phi\colon A\wr B\to A\wr\overline{B} such that |B¯|C2n|\overline{B}|\leq C_{2}n and where π(fb)\pi(fb) is not conjugate to π(gc)\pi(gc) in AB¯A\wr\overline{B} for some constant C2>0C_{2}>0. We see that

|AB¯|=|A||B¯||B¯||A|C2nC2n.|A\wr\overline{B}|=|A|^{|\overline{B}|}|\overline{B}|\leq|A|^{C_{2}n}C_{2}n.

Interpreting the size of |AB¯||A\wr\overline{B}| as a function of nn, we get that

|AB¯||B¯||A|C2nC2n|A|nn|A|n2n.|A\wr\overline{B}|\leq|\overline{B}||A|^{C_{2}n}C_{2}n\preceq|A|^{n}n\preceq|A|^{n}\preceq 2^{n}.

Subsequently, we see that ConjG(n)2n\operatorname{Conj}_{G}(n)\preceq 2^{n}.

When the torsion free rank is greater than 11, we see that there exists a finite abelian group B¯\overline{B} together with a surjective homomorphism ϕ:ABAB¯\phi\colon A\wr B\to A\wr\overline{B} such that |B¯|(C2n)2k|\overline{B}|\leq(C_{2}n)^{2k} and where π(fb)\pi(fb) is not conjugate to π(gc)\pi(gc) in AB¯A\wr\overline{B} for some constant C2>0C_{2}>0. We see that |AB¯|2n2k.|A\wr\overline{B}|\preceq 2^{n^{2k}}. Consequently, we see that ConjG(n)2n2k\operatorname{Conj}_{G}(n)\preceq 2^{n^{2k}}. ∎

4.4. Infinite base groups


When given a wreath product of finitely generated abelian groups ABA\wr B where AA is infinite, the following lemma will allow us to construct an upper bound for size of the quotient of the base group given two elements of fbfb and gbgb where f,g:BAf,g\colon B\to A are finitely supported functions.

Lemma 4.12.

Let AA and BB be finitely generated abelian groups where AA is infinite and BB is finite. Let f,g:BAf,g\colon B\to A, bBb\in B be such that fb,gbBG(n)fb,gb\in\operatorname{B}_{G}(n), the elements fbfb and gbgb are reduced, and fb≁Ggbfb\not\sim_{G}gb. Then there exists a surjective homomorphism π:AA¯\pi\colon A\to\overline{A} to a finite group A¯\overline{A} such that

|A¯|min{log(Cn)2|B|,log(Cn)Cn2}\left|\overline{A}\right|\leq\min\left\{\log(Cn)^{2|B|},\log(Cn)^{Cn^{2}}\right\}

and where π~(fb)≁π~(gc)\tilde{\pi}(fb)\not\sim\tilde{\pi}(gc) in A¯B\overline{A}\wr B for some constant C>0C>0 independent of f,g,b,nf,g,b,n.

Proof.

Since fb≁Ggbfb\not\sim_{G}gb and the elements fb,gbfb,gb are reduced, Lemma 2.7 implies that one of the following must be true:

  1. (1)

    supp(f)\operatorname{supp}(f) is not a translate of supp(g)\operatorname{supp}(g) in BB,

  2. (2)

    for every aBa\in B such that a+supp(f)=supp(g)a+\operatorname{supp}(f)=\operatorname{supp}(g) there exists some xsupp(g)x\in\operatorname{supp}(g) such that f(x+a)g(x)f(x+a)\neq g(x).

Lemma 2.9 implies there exists a constant C1>0C_{1}>0 such that

|supp(f)|,|supp(g)|C1n|\operatorname{supp}(f)|,|\operatorname{supp}(g)|\leq C_{1}n

and that rng(f),rng(g)BA(C1n)\operatorname{rng}(f),\operatorname{rng}(g)\subseteq B_{A}(C_{1}n). Denote R=rng(f)rng(g)R=\operatorname{rng}(f)\cup\operatorname{rng}(g), Clearly |R|2C1n|R|\leq 2C_{1}n and RBA(C1n)R\subseteq B_{A}(C_{1}n).

First, we will show that there is a finite group A¯\overline{A} satisfying the requirements on conjugacy such that |A¯|log(Cn)|B||\overline{A}|\leq\log(Cn)^{|B|}. Suppose that supp(f)\operatorname{supp}(f) is not a translate of supp(g)\operatorname{supp}(g) in BB. Since AA is a finitely generated abelian group, its residual finiteness depth function is equivalent to log(n)\log(n). It then follows that for every rRBr\in R\subseteq B there is a normal finite index subgroup KrK_{r} of AA such that rKrr\notin K_{r} and |A:Kr|log(C1n)|A:K_{r}|\leq\log(C_{1}n). Set K=rRKrK=\cap_{r\in R}K_{r} with natural projection given by π:AA/K\pi\colon A\to A/K. As none of the elements in RR get mapped to the identity, we see that supp(π~(f))=supp(f)\operatorname{supp}(\tilde{\pi}(f))=\operatorname{supp}(f) and supp(π~(g))=supp(g)\operatorname{supp}(\tilde{\pi}(g))=\operatorname{supp}(g). In particular, we see that the elements π~(fb),π~(gb)\tilde{\pi}(fb),\tilde{\pi}(gb) are reduced. It follows that supp(f~)\operatorname{supp}(\tilde{f}) is not a translate of supp(g~)\operatorname{supp}(\tilde{g}). Therefore, we have that π~(fb)\tilde{\pi}(fb) is not conjugate to π~(gc)\tilde{\pi}(gc) in (A/K)B(A/K)\wr B by Lemma 2.7. We see that

|A/K|log(C1n)|R|log(C1n)|B|.|A/K|\leq\log(C_{1}n)^{|R|}\leq\log(C_{1}n)^{|B|}.

Suppose that supp(f)\operatorname{supp}(f) is a translate of supp(g)\operatorname{supp}(g) and let TBT\subseteq B be the set of all elements of BB that translate supp(f)\operatorname{supp}(f) onto supp(g)\operatorname{supp}(g). By assumption, for every tTt\in T there is xBx\in B such that f(t+x)g(x)f(t+x)\neq g(x). For every such tt, there exists a normal finite index subgroup KtK_{t} of AA such that f(t+xt)Ktg(xt)Ktf(t+x_{t})K_{t}\neq g(x_{t})K_{t} for some xtBx_{t}\in B and |A:Kt|log(C1n)|A:K_{t}|\leq\log(C_{1}n). Denote

K=rRKrtTKt,K=\bigcap_{r\in R}K_{r}\quad\cap\bigcap_{t\in T}K_{t},

where KrK_{r} is defined as in the previous paragraph, and let π:AA/K\pi\colon A\to A/K be the natural projection. Clearly, supp(π~(f))=supp(f)\operatorname{supp}(\tilde{\pi}(f))=\operatorname{supp}(f) and supp(π~(g))=supp(g)\operatorname{supp}(\tilde{\pi}(g))=\operatorname{supp}(g) and, again, we see that the elements π~(fb),π~(gb)\tilde{\pi}(fb),\tilde{\pi}(gb) are reduced. From the construction of the KK we see that for every tTt\in T there is xtBx_{t}\in B such that

π~(f)(xt+t)=π(f(xt+t))π(g(xt))=π~(g)(xt).\tilde{\pi}(f)(x_{t}+t)=\pi(f(x_{t}+t))\neq\pi(g(x_{t}))=\tilde{\pi}(g)(x_{t}).

That means π~(fb)\tilde{\pi}(fb) is not conjugate to π~(gb)\tilde{\pi}(gb) by Lemma 2.7. To bound the index of KK, we can write

|A/K|log(C1n)|R|log(C1n)|T|log(C1n)2|B|.|A/K|\leq\log(C_{1}n)^{|R|}\log(C_{1}n)^{|T|}\leq\log(C_{1}n)^{2|B|}. (1)

Now we show that there is a finite group A¯\overline{A} satisfying the requirements on conjugacy such that |A¯|log(Cn)Cn2.|\overline{A}|\leq\log(Cn)^{Cn^{2}}. Suppose that supp(f)\operatorname{supp}(f) is not a translate of supp(g)\operatorname{supp}(g) in BB. Since AA is a finitely generated abelian group, [4, Corollary 2.3] implies that for every rRr\in R there is Krf.i.AK_{r}\operatorname{\unlhd_{\text{f.i.}}}A such that rKrr\notin K_{r} and |A/Kr|log(C2n)|A/K_{r}|\leq\log(C_{2}n) for some constant C2>0C_{2}>0. Set K=rRKrK=\cap_{r\in R}K_{r} with natural projection given by π:AA/K\pi\colon A\to A/K. As none of the elements in RR get mapped to the identity, we see that supp(π~(f))=supp(f)\operatorname{supp}(\tilde{\pi}(f))=\operatorname{supp}(f) and supp(π~(g))=supp(g)\operatorname{supp}(\tilde{\pi}(g))=\operatorname{supp}(g) and that the elements π~(fb),π~(gb)\tilde{\pi}(fb),\tilde{\pi}(gb) are reduced. It follows that supp(f~)\operatorname{supp}(\tilde{f}) is not a translate of supp(g~)\operatorname{supp}(\tilde{g}). Therefore, we have that π~(fb)\tilde{\pi}(fb) is not conjugate to π~(gc)\tilde{\pi}(gc) in (A/K)B(A/K)\wr B by Lemma 2.7. Clearly,

|A/K|log(C2n)|R|log(C2n)2C1nlog(2C2n)2C1n.|A/K|\leq\log(C_{2}n)^{|R|}\leq\log(C_{2}n)^{2C_{1}n}\leq\log(2C_{2}n)^{2C_{1}n}.

Now suppose for every aBa\in B such that a+supp(f)=supp(g)a+\operatorname{supp}(f)=\operatorname{supp}(g) there exists some xsupp(g)x\in\operatorname{supp}(g) such that f(x+a)g(x)f(x+a)\neq g(x). For every rRr\in R, let KrK_{r} be defined as in the previous paragraph recalling that |A/Kr|log(C0n)|A/K_{r}|\leq\log(C_{0}n). For every {r,s}R\{r,s\}\subset R there is Kr,sf.i.AK_{r,s}\operatorname{\unlhd_{\text{f.i.}}}A such that rsKr,sr-s\notin K_{r,s} and |A/Kr,s|log(2C2n)|A/K_{r,s}|\leq\log(2C_{2}n). Denote

K=rRKr{r,s}RKr,sK=\bigcap_{r\in R}K_{r}\quad\cap\bigcap_{\{r,s\}\in R}K_{r,s}

with associated natural projection given by π:AA/K\pi\colon A\to A/K. Following the same argument as in the previous case, we see that supp(π~(f))=supp(f)\operatorname{supp}(\tilde{\pi}(f))=\operatorname{supp}(f), supp(π~(g))=supp(g)\operatorname{supp}(\tilde{\pi}(g))=\operatorname{supp}(g), and the elements π~(fb),π~(gb)\tilde{\pi}(fb),\tilde{\pi}(gb) are reduced. Now suppose that aBa\in B is given such that a+supp(π~((f))=supp(π~(g))a+\operatorname{supp}(\tilde{\pi}((f))=\operatorname{supp}(\tilde{\pi}(g)). That is equivalent to a+supp(f)=supp(g)a+\operatorname{supp}(f)=\operatorname{supp}(g). Thus, there is some xaBx_{a}\in B such that f(xa+a)g(xa)f(x_{a}+a)\neq g(x_{a}). However, from the construction of KK we see that

π~(f)(xa+a)=π(f(xa+a))π(g(xa))=π~(g)(xa).\tilde{\pi}(f)(x_{a}+a)=\pi(f(x_{a}+a))\neq\pi(g(x_{a}))=\tilde{\pi}(g)(x_{a}).

In particular, Lemma 2.7 implies that π~(fb)\tilde{\pi}(fb) is not conjugate to π~(g)\tilde{\pi}(g) in (A/K)B(A/K)\wr B. Finally, we finish with

|A/K|log(C2n)|R|log(2C2n)|(R2)|log(2C1n)2(2C2n2)log(4C2n)4C2n2.|A/K|\leq\log(C_{2}n)^{|R|}\cdot\log(2C_{2}n)^{\left|\binom{R}{2}\right|}\leq\log(2C_{1}n)^{2\binom{2C_{2}n}{2}}\leq\log(4C_{2}n)^{4C_{2}n^{2}}. (2)

Combining (1) and (2) immediately yields the result. ∎

Combining Proposition 4.10 together with Lemma 4.12 gives the following upper bound for wreath products of infinite, finitely generated abelian groups.

Proposition 4.13.

Suppose that A,BA,B are infinite finitely generated abelian groups. If BB is virtually cyclic then

ConjAB(n)(log(n))n2.\operatorname{Conj}_{A\wr B}(n)\preceq(\log(n))^{n^{2}}.

Otherwise,

ConjAB(n)(log(n))n2k+2\operatorname{Conj}_{A\wr B}(n)\preceq(\log(n))^{n^{2k+2}}

where kk is the torsion-free rank of BB.

Proof.

Let kk denote the torsion-free rank of BB, and suppose that nn\in\mathbb{N}, f,gABf,g\in A^{B}, and b,cBb,c\in B are given such that fb,gcBAB(n)fb,gc\in B_{A\wr B}(n) and fb≁gcfb\not\sim gc.

Suppose first that bcb\neq c. Since bcBB(2n)b-c\in\operatorname{B}_{B}(2n), [4, Corollary 2.3] implies there exists a constant C1>0C_{1}>0 and a surjective homomorphism φ:BQ\varphi\colon B\to Q such that φ(b)φ(c)\varphi(b)\neq\varphi(c) and where |Q|C1log(C1n).|Q|\leq C_{1}\log(C_{1}n). Since AA is abelian, we have that φ(b)\varphi(b) and φ(c)\varphi(c) are non-conjugate. By composing φ\varphi with the projection of ABA\wr B onto BB which we also denote φ\varphi, we have a surjective homomorphism φ:ABQ\varphi\colon A\wr B\to Q such that φ(fb)φ(gc)\varphi(fb)\nsim\varphi(gc) and where |Q|C1log(C1n).|Q|\leq C_{1}\>\log(C_{1}n). Thus, we may assume that b=c.b=c.

Following Lemma 2.6 there exist a constant C0C_{0} and functions f,gABf^{\prime},g^{\prime}\in A^{B} such that fbfbf^{\prime}b\sim fb and gbgbg^{\prime}b\sim gb, the elements fb,gbf^{\prime}b,g^{\prime}b are reduced, and fb,gbBAB(C0n)fb,gb\in B_{A\wr B}(C_{0}n).

Following Proposition 4.10, there is a constant CBC_{B} (independent of n,f,g,b,cn,f,g,b,c) and a finite abelian group B¯\overline{B} together with a surjective homomorphism πB:BB¯\pi_{B}\colon B\to\overline{B} such that the elements πB~(fc)\tilde{\pi_{B}}(f^{\prime}c) and πB~(gc)\tilde{\pi_{B}}(g^{\prime}c) are reduced, πB~(fb)\tilde{\pi_{B}}(f^{\prime}b) is not conjugate to πB~(gc)\tilde{\pi_{B}}(g^{\prime}c) in AB¯A\wr\overline{B}, where |B¯|C0CBn|\overline{B}|\leq C_{0}C_{B}n if BB has torsion free rank 11 and where |B¯|C0CBn2k|\overline{B}|\leq C_{0}C_{B}n^{2k} if BB has torsion free rank k>1k>1.

Set G¯=AB¯\overline{G}=A\wr\overline{B}. One can easily check that π~B(BG(C0n))BG¯(C0n)\tilde{\pi}_{B}(B_{G}(C_{0}n))\subseteq B_{\overline{G}}(C_{0}n). In particular, π~B(fb),π~B(gc)BB¯(C0n)\tilde{\pi}_{B}(f^{\prime}b),\tilde{\pi}_{B}(g^{\prime}c)\subseteq B_{\overline{B}}(C_{0}n). Lemma 4.12 implies that there is a constant CAC_{A} (also independent of n,f,g,b,cn,f,g,b,c) and a finite abelian group A¯\overline{A} together with a surjective homomorphism πA:AA¯\pi_{A}\colon A\to\overline{A} such that

|A¯|min{log(CAn)2|B|,log(CAn)CAn2}|\overline{A}|\leq\min\left\{\log(C_{A}n)^{2|B|},\log(C_{A}n)^{C_{A}n^{2}}\right\}

and where π~A(π~B(fb))\tilde{\pi}_{A}(\tilde{\pi}_{B}(fb)) is not conjugate to π~A(π~B(gc))\tilde{\pi}_{A}(\tilde{\pi}_{B}(gc)) in A¯B¯\overline{A}\wr\overline{B}.

Therefore, A¯B¯\overline{A}\wr\overline{B} is a finite group. If BB has torsion free rank 11, then |B|C0CBn|B|\leq C_{0}C_{B}n. If BB is virtually cyclic, we set C=2C0CB2C=2C_{0}C_{B}^{2} and compute

|A¯B¯|=|B¯||A¯||B¯|\displaystyle\left|\overline{A}\wr\overline{B}\right|=\left|\overline{B}\right|\cdot\left|\overline{A}\right|^{\left|\overline{B}\right|} C0CBn(log(CAn)2C0CBn)C0CBn\displaystyle\leq C_{0}C_{B}n\cdot\left(\log(C_{A}n)^{2C_{0}C_{B}n}\right)^{C_{0}C_{B}n}
=C0CBn(log(CAn))2C02CB2n2\displaystyle=C_{0}C_{B}n\cdot\left(\log(C_{A}n)\right)^{2C_{0}^{2}C_{B}^{2}n^{2}}
Cn(log(Cn))Cn2.\displaystyle\leq Cn\left(\log(Cn)\right)^{Cn^{2}}.

Interpreting |A¯B¯|\left|\overline{A}\wr\overline{B}\right| as a function of nn immediately yields

|A¯B¯|Cn(log(Cn))Cn2nlog(n)n2log(n)n2.\left|\overline{A}\wr\overline{B}\right|\leq Cn\left(\log(Cn)\right)^{Cn^{2}}\approx n\log(n)^{n^{2}}\approx\log(n)^{n^{2}}.

Therefore, we have that

ConjG(n)log(n)n2.\operatorname{Conj}_{G}(n)\preceq\log(n)^{n^{2}}.

Now suppose that k>1k>1. Set C=C0CACBC=C_{0}C_{A}C_{B}. We then compute

|A¯B¯|=|B¯||A¯||B¯|\displaystyle\left|\overline{A}\wr\overline{B}\right|=\left|\overline{B}\right|\cdot\left|\overline{A}\right|^{\left|\overline{B}\right|} C0CBn2k(log(CAn)CAn2)C0CBn2k\displaystyle\leq C_{0}C_{B}n^{2k}\cdot\left(\log(C_{A}n)^{C_{A}n^{2}}\right)^{C_{0}C_{B}n^{2k}}
=C0CBn2k(log(CAn))C0CACBn2k+2\displaystyle=C_{0}C_{B}n^{2k}\cdot\left(\log(C_{A}n)\right)^{C_{0}C_{A}C_{B}n^{2k+2}}
Cn2k(log(Cn))(Cn)2k+2.\displaystyle\leq Cn^{2k}\left(\log(Cn)\right)^{(Cn)^{2k+2}}.

Interpreting |A¯B¯|\left|\overline{A}\wr\overline{B}\right| as a function of nn immediately yields

|A¯B¯|Cn2k(log(Cn))Cn2k+2n2klog(n)n2k+2log(n)n2k+2,\left|\overline{A}\wr\overline{B}\right|\leq Cn^{2k}\left(\log(Cn)\right)^{Cn^{2k+2}}\approx n^{2k}\log(n)^{n^{2k+2}}\approx\log(n)^{n^{2k+2}},

and therefore,

ConjG(n)log(n)n2k+2,\operatorname{Conj}_{G}(n)\preceq\log(n)^{n^{2k+2}},

which concludes our proof. ∎

5. Applications

In this section, we use Corollary 1.4 to derive lower bounds for the conjugacy separability depth function where the acting group is not necessarily abelian. For the statement of the theorem, we denote the center of a group GG as Z(G)Z(G).

Theorem 5.1.

Let AA be a nontrivial finitely generated abelian group, and suppose that GG is a conjugacy separable finitely generated group with separable cyclic subgroups. Suppose that Z(G)\mathbb{Z}\leq Z(G). If AA is finite, then

2nConjAG(n).2^{n}\preceq\operatorname{Conj}_{A\wr G}(n).

Otherwise,

(log(n))nConjAG(n).(\log(n))^{n}\preceq\operatorname{Conj}_{A\wr G}(n).
Proof.

Let B=kZ(G)B=\mathbb{Z}^{k}\leq Z(G). We claim that if (f,b),(g,b)AB(f,b),(g,b)\in A\wr B then xAGyx\sim_{A\wr G}y if and only if there exists zABz\in A\wr B such that zxz1=yz\>x\>z^{-1}=y. Suppose first that xAGyx\sim_{A\wr G}y. [10, Lemma 5.13] implies we may assume that elements of supp(f)\operatorname{supp}(f) (respectively supp(g)\operatorname{supp}(g)) lie in different right cosets of b\left<b\right>. [10, Lemma 5.14] implies that xAGyx\sim_{A\wr G}y if and only if there exists an element cCG(b)c\in C_{G}(b) such that cfc1=g.cfc^{-1}=g. That implies for all for xGx\in G, we have that f(cx)=g(x)f(cx)=g(x). We note that g(x)0g(x)\neq 0 only if xBx\in B. That implies f(cx)0f(cx)\neq 0 only if cxBcx\in B. Hence, cB.c\in B. In particular, if (h,c)AB(h,c)\in A\wr B where cBc\notin B, then (h,c)x(h,c)1y(h,c)\cdot x\cdot(h,c)^{-1}\neq y. Thus, if (h,c)x(h,c)1=y(h,c)\cdot x\cdot(h,c)^{-1}=y, then cBc\in B. Hence, we may write

(h,c)(f,b)(h,c)1\displaystyle(h,c)(f,b)(h,c)^{-1} =\displaystyle= (h,c)(f,b)((c1h,c1)\displaystyle(h,c)(f,b)((c^{-1}\cdot h,c^{-1})
=\displaystyle= (h+cf,cb)(c1h,c1)\displaystyle(h+c\cdot f,cb)(c^{-1}\cdot h,c^{-1})
=\displaystyle= (h+cf+(cb)c1h,g)\displaystyle(h+c\cdot f+(cb)\cdot c^{-1}\cdot h,g)
=\displaystyle= (h+cf+bh,g).\displaystyle(h+c\cdot f+b\cdot h,g).

Thus, if (h,c)x(h,c)=y(h,c)\cdot x\cdot(h,c)=y, we must have that

supp(h+gf+gh)B.\operatorname{supp}(h+g^{\prime}\cdot f+g\cdot h)\subseteq B.

Suppose supp(h)B\operatorname{supp}(h)\not\subseteq B. In this case, we note that supp(h+gh)B\operatorname{supp}(h+g\cdot h)\not\subseteq B and that supp(cf)B\operatorname{supp}(c\cdot f)\subseteq B. Therefore, supp(h+gf+gh)B.\operatorname{supp}(h+g^{\prime}\cdot f+g\cdot h)\not\subseteq B. Hence, we have that (h,c)x(h,c)1y(h,c)\cdot x\cdot(h,c)^{-1}\neq y which is a contradiction. Therefore, supp(f)B\operatorname{supp}(f)\subseteq B which implies (h,c)AB.(h,c)\in A\wr B. Since the other direction is clear, we have our claim.

Subsequently, we have ConjAB(n)ConjAG(n).\operatorname{Conj}_{A\wr B}(n)\preceq\operatorname{Conj}_{A\wr G}(n). Our theorem then follows from Corollary 1.4. ∎

Since the rank of the center of an infinite, finitely generated nilpotent group is always positive, we have the following corollary.

Corollary 5.2.

Let AA be a finitely generated abelian group, and let NN be an infinite, finitely generated nilpotent group. If AA is finite, then

2nConjAN(n).2^{n}\preceq\operatorname{Conj}_{A\wr N}(n).

Otherwise,

(log(n))nConjAN(n).(\log(n))^{n}\preceq\operatorname{Conj}_{A\wr N}(n).

In both cases, we have that ConjAN(n)\operatorname{Conj}_{A\wr N}(n) has at least exponential growth.

Finally, we consider the case when the acting group contains the integers as a retract.

Theorem 5.3.

Let AA be a nontrivial finitely generated abelian group, and suppose that GG is a conjugacy separable finitely generated group with separable cyclic subgroups that contains an infinite cyclic group as a retract. If AA is finite, then

2nConjAG(n).2^{n}\preceq\operatorname{Conj}_{A\wr G}(n).

Otherwise,

(log(n))nConjAG(n).(\log(n))^{n}\preceq\operatorname{Conj}_{A\wr G}(n).
Proof.

Let gGg\in G be an element of infinite order such that g\langle g\rangle is a retract in GG. Then by Lemma 2.5 we see that AgAA\wr\langle g\rangle\simeq A\wr\mathbb{Z} is a retract in AGA\wr G and ConjAg(n)ConjAG(n)\operatorname{Conj}_{A\wr\langle g\rangle}(n)\leq\operatorname{Conj}_{A\wr G}(n). The rest then follows by Corollary 1.4. ∎

Corollary 5.4.

Let AA be a finitely generated abelian group, and suppose that GG belongs to one of the following classes of groups:

  1. (i)

    right-angled Artin groups,

  2. (ii)

    infinite finitely generated nilpotent and polycyclic groups,

  3. (iii)

    limit groups,

  4. (iv)

    fundamental groups of hyperbolic fibered 3-manifolds,

  5. (v)

    graph products of any of the above.

If AA is finite, then

2nConjAG(n).2^{n}\preceq\operatorname{Conj}_{A\wr G}(n).

Otherwise,

(log(n))nConjAG(n).(\log(n))^{n}\preceq\operatorname{Conj}_{A\wr G}(n).
Proof.

Right-angled Artin groups were shown to be conjugacy separable by Minasyan in [19, Theorem 1.1], cyclic subgroup separable by Green in [11, Theorem 2.16], and checking that a right-right angled Artin group admits an infinite cyclic retract is easy - just consider an endomorphism that maps all but one generator to the identity.

For fundamental groups of closed orientable surfaces, they are known to be subgroup separable due to Scott [22, 23] and basic structure theory of abelian groups. The fact that they are conjugacy separable follows from Martino [17]. Lastly, it is clear that they have infinite cyclic retracts due to the fact that they have infinite abelianizations.

Infinite polycyclic and finitely generated nilpotent groups are well known to conjugacy separable and subgroup separable. Proofs of these results can be found in [24, Theorem 3, pg59] and [16]. The fact that these classes of groups admit an infinite cyclic retract follows from the fact that infinite polycyclic groups and infinite finitely generated abelian groups always have infinite abelianization by basic structure results found in [24].

For limit groups, it was shown by Wilton that they are subgroup separable in [5, Theorem A], and Chagas and Zalesskii demonstrated that they are conjugacy separable in [5, Theorem 1.1]. The fact that they have infinite cyclic retracts follows from the fact that limit groups are fully residually free.

The proof that fundamental groups of hyperbolic 33-manifold groups are conjugacy separable follows from Hamilton, Wilton, and Zalesskii [12, Theorem 1.3] and subgroup separability follows from [1, Corollary 4.2.3]. The fact that they have infinite cyclic retracts follows from these fact that these groups have the form π1(Σg)\pi_{1}(\Sigma_{g})\rtimes\mathbb{Z} where Σg\Sigma_{g} is a closed orientable genus g2g\geq 2 surface.

The class of conjugacy separable groups is closed under forming graph products by [9, Theorem 1.1]. The class of cyclic subgroup separable groups is closed under forming graph products by [3, Theorem A]. Finally, to see that a graph product of groups that admits an infinite cyclic retract again admits an infinite cyclic retracts easily follows from the fact that all vertex groups are themselves retracts. ∎

The vast range of examples we were able to construct in this paper either with Theorem 5.1 and Theorem 5.3 lead us to believe that the lower bounds we produced cannot be relaxed and therefore we state the following conjecture.

Conjecture 5.5.

Let AA be a finitely generated abelian group and let GG be a conjugacy separable group with separable cyclic subgroups. If AA is finite, then

2nConjAG(n).2^{n}\preceq\operatorname{Conj}_{A\wr G}(n).

Otherwise,

(log(n))nConjAG(n).(\log(n))^{n}\preceq\operatorname{Conj}_{A\wr G}(n).

6. Open questions

The constructions of lower bounds for 𝔽p\mathbb{F}_{p}\wr\mathbb{Z} and \mathbb{Z}\wr\mathbb{Z} given in Section 3 relies heavily on the fact that both 𝔽p\mathbb{F}_{p}\wr\mathbb{Z} and \mathbb{Z}\wr\mathbb{Z} can be represented as a semidirect product of the additive group of the ring of Laurent polynomials over 𝔽p\mathbb{F}_{p} and \mathbb{Z}, respectively, and that subgroups of finite index correspond to cofinite ideals in the said polynomial rings. This approach can be generalised: let RR denote either 𝔽p\mathbb{F}_{p} or \mathbb{Z}, then

RkR[X1,X11,,Xk,Xk1]k,R\wr\mathbb{Z}^{k}\simeq R[X_{1},X_{1}^{-1},\dots,X_{k},X_{k}^{-1}]\rtimes\mathbb{Z}^{k},

where the action of k\mathbb{Z}^{k} on R[X1,X11,,Xk,Xk1]\simeq R[X_{1},X_{1}^{-1},\dots,X_{k},X_{k}^{-1}] is given by

(b1,,bk)X1e1Xkek=X1b1+e1Xkek+bk.(b_{1},\dots,b_{k})\cdot X_{1}^{e_{1}}\dots X_{k}^{e_{k}}=X_{1}^{b_{1}+e_{1}}\dots X_{k}^{e_{k}+b_{k}}.

One can easily check that finite index subgroups of RkR\wr\mathbb{Z}^{k} correspond to cofinite ideals of R[X1,X11,,Xk,Xk1]R[X_{1},X_{1}^{-1},\dots,X_{k},X_{k}^{-1}] and statement similar to Lemma 2.12 can be proved, but the arguments based on divisibility in R[X,X1]R[X,X^{-1}] do not carry over to R[X1,X11,,Xk,Xk1]R[X_{1},X_{1}^{-1},\dots,X_{k},X_{k}^{-1}].

Question 6.1.

Can the proofs of Proposition 3.1 and Proposition 3.3 be modified to produce lower bounds for the conjugacy depth functions of 𝔽pk\mathbb{F}_{p}\wr\mathbb{Z}^{k} and k\mathbb{Z}\wr\mathbb{Z}^{k}, respectively, that dominate and are not dominated by 2n2^{n} and log(n)n\log(n)^{n}, respectively? In particular, can it be shown that

2nkConj𝔽pk(n)2^{n^{k}}\preceq\operatorname{Conj}_{\mathbb{F}_{p}\wr\mathbb{Z}^{k}}(n)

and

log(n)nkConjk(n)?\log(n)^{n^{k}}\preceq\operatorname{Conj}_{\mathbb{Z}\wr\mathbb{Z}^{k}}(n)?

The upper bounds given by Proposition 4.11 and Proposition 4.13 treat wreath products of abelian groups differently, based on the torsion-free rank of the acting group, which might feel somewhat disappointing. The following question is therefore a naturally arising one.

Question 6.2.

Can the proofs of Proposition Proposition 4.11 and Proposition 4.13 be modified to produce upper bounds for the conjugacy depth functions of 𝔽pk\mathbb{F}_{p}\wr\mathbb{Z}^{k} and k\mathbb{Z}\wr\mathbb{Z}^{k}, respectively, that are given by a single closed-form formula that does not have cases based on the torsion-free rank of the acting group? In particular, can it be shown that

Conj𝔽pk(n)2nk\operatorname{Conj}_{\mathbb{F}_{p}\wr\mathbb{Z}^{k}}(n)\preceq 2^{n^{k}}

and

Conjk(n)log(n)n2k?\operatorname{Conj}_{\mathbb{Z}\wr\mathbb{Z}^{k}}(n)\preceq\log(n)^{n^{2k}}?

References

  • [1] Matthias Aschenbrenner, Stefan Friedl, and Henry Wilton. 3-manifold groups. EMS Series of Lectures in Mathematics. European Mathematical Society (EMS), Zürich, 2015.
  • [2] M. F. Atiyah and I. G. Macdonald. Introduction to commutative algebra. Addison-Wesley Series in Mathematics. Westview Press, Boulder, CO, economy edition, 2016. For the 1969 original see [ MR0242802].
  • [3] Federico Berlai and Michal Ferov. Separating cyclic subgroups in graph products of groups. Journal of Algebra, 531:19–56, 2019.
  • [4] Khalid Bou-Rabee. Quantifying residual finiteness. J. Algebra, 323(3):729–737, 2010.
  • [5] S. C. Chagas and P. A. Zalesskii. Limit groups are conjugacy separable. Internat. J. Algebra Comput., 17(4):851–857, 2007.
  • [6] Tara C. Davis and Alexander Yu. Olshanskii. Subgroup distortion in wreath products of cyclic groups. J. Pure Appl. Algebra, 215(12):2987–3004, 2011.
  • [7] Jonas Deré and Mark Pengitore. Effective twisted conjugacy separability of nilpotent groups. Math. Z., 292(3-4):763–790, 2019.
  • [8] David Eisenbud. Commutative algebra, volume 150 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1995. With a view toward algebraic geometry.
  • [9] M. Ferov. On conjugacy separability of graph products of groups. J. Algebra, 447:135–182, 2016.
  • [10] Michal Ferov and Mark Pengitore. Quantifying conjugacy separability in wreath products of groups. arXiv preprint arXiv:1910.02174, 2019.
  • [11] E. R. Green. Graph products of groups. PhD thesis, University of Leeds, 1990.
  • [12] Emily Hamilton, Henry Wilton, and Pavel A. Zalesskii. Separability of double cosets and conjugacy classes in 3-manifold groups. J. Lond. Math. Soc. (2), 87(1):269–288, 2013.
  • [13] Serge Lang. Algebra, volume 211 of Graduate Texts in Mathematics. Springer-Verlag, New York, third edition, 2002.
  • [14] Sean Lawton, Larsen Louder, and D. B. McReynolds. Decision problems, complexity, traces, and representations. Groups Geom. Dyn., 11(1):165–188, 2017.
  • [15] Bohdan S. Majewski and George Havas. The complexity of greatest common divisor computations. In Leonard M. Adleman and Ming-Deh Huang, editors, Algorithmic Number Theory, pages 184–193, Berlin, Heidelberg, 1994. Springer Berlin Heidelberg.
  • [16] A. I. Mal’cev. On homomorphisms onto finite groups. American Mathematical Society Translations, Series, 2(119):67–79, 1983.
  • [17] Armando Martino. A proof that all Seifert 3-manifold groups and all virtual surface groups are conjugacy separable. J. Algebra, 313(2):773–781, 2007.
  • [18] J. C. C. McKinsey. The decision problem for some classes of sentences without quantifiers. The Journal of Symbolic Logic, 8(2):61–76, 1943.
  • [19] A. Minasyan. Hereditary conjugacy separability of right-angled Artin groups and its applications. Groups Geom. Dyn., 6(2):335–388, 2012.
  • [20] A. Mostowski. On the decidability of some problems in special classes of groups. Fund. Math., 59:123–135, 1966.
  • [21] V. N. Remeslennikov. Finite approximability of groups with respect to conjugacy. Sibirsk. Mat. Ž., 12:1085–1099, 1971.
  • [22] Peter Scott. Subgroups of surface groups are almost geometric. J. London Math. Soc., 17:555–565, 1978.
  • [23] Peter Scott. Corrections to ”subgroups of surface groups are almost geometric”. J. London Math. Soc., 32:217–220, 1985.
  • [24] Daniel Segal. Polycyclic Groups. Cambridge University Press, 1983.
  • [25] Gérald Tenenbaum. Introduction to analytic and probabilistic number theory, volume 46 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 1995. Translated from the second French edition (1995) by C. B. Thomas.