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

Dp-minimal profinite groups and
valuations on the integers

Tim Clausen111The author is partially supported by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany’s Excellence Strategy EXC 2044–390685587, Mathematics Münster: Dynamics-Geometry-Structure.
Abstract

We study dp-minimal infinite profinite groups that are equipped with a uniformly definable fundamental system of open subgroups. We show that these groups have an open subgroup AA such that either AA is a direct product of countably many copies of 𝔽p\mathbb{F}_{p} for some prime pp, or AA is of the form Appαp×ApA\cong\prod_{p}\mathbb{Z}_{p}^{\alpha_{p}}\times A_{p} where αp<ω\alpha_{p}<\omega and ApA_{p} is a finite abelian pp-group for each prime pp. Moreover, we show that if AA is of this form, then there is a fundamental system of open subgroups such that the expansion of AA by this family of subgroups is dp-minimal. Our main ingredient is a quantifier elimination result for a class of valued abelian groups. We also apply it to (,+)(\mathbb{Z},+) and we show that if we expand (,+)(\mathbb{Z},+) by any chain of subgroups (Bi)i<ω(B_{i})_{i<\omega}, we obtain a dp-minimal structure. This structure is distal if and only if the size of the quotients Bi/Bi+1B_{i}/B_{i+1} is bounded.

1 Introduction

A profinite group GG together with a fundamental system {Ki:iI}\{K_{i}:i\in I\} of open subgroups can be viewed as a two-sorted structure (G,I)(G,I) in the two-sorted language prof\mathcal{L}_{\text{prof}}. In these structures the fundamental system of open subgroups is definable. Since a fundamental system of open subgroups is a neighborhood basis at the identity, this implies that the topology on GG is definable.

These structures have been studied by Macpherson and Tent in [19]. They mainly considered full profinite groups, i.e. profinite groups GG where the family {Ki:iI}\{K_{i}:i\in I\} consists of all open subgroups. Their main result states that a full profinite group (G,I)(G,I) is NIP if and only if it is NTP2 if and only if it is virtually a finite direct product of analytic pro-pp groups.

Since analytic pro-pp groups can be described as products of copies of p\mathbb{Z}_{p} with a twisted multiplication, profinite NIP groups are composed of “one-dimensional” profinite NIP groups. In the setting of full profinite groups the combinatorial structure of the lattice of open subgroups is visible in the model theoretic structure. This plays an important role in the classification.

Without the fullness assumption, only a portion of this lattice is visible. In general the family {Ki:iI}\{K_{i}:i\in I\} could simply consist of a chain of open subgroups. In this more general setting, we will restrict ourselves to the “one-dimensional”, i.e. dp-minimal case. A profinite group (G,I)(G,I) is dp-minimal if it has NIP and is dp-minimal in the group sort. We prove the following classification result:

Theorem.

Let (G,I)(G,I) be a dp-minimal profinite group. Then GG has an open abelian subgroup AA such that either AA is a direct product of countably many copies of 𝔽p\mathbb{F}_{p} for some prime pp, or AA is isomorphic to ppαp×Ap\prod_{p}\mathbb{Z}_{p}^{\alpha_{p}}\times A_{p} where αp<ω\alpha_{p}<\omega and ApA_{p} is a finite abelian pp-group for each prime pp. Moreover, every abelian profinite group AA of the above form admits a fundamental system of open subgroups such that the corresponding prof\mathcal{L}_{\text{prof}}-structure is dp-minimal.

The main ingredient of this theorem is a quantifier elimination result which is also applicable in other settings. We apply it to this situation: Consider the structure (,+)(\mathbb{Z},+). If we expand it by the full lattice of subgroups, then the expanded structure interprets Peano Arithmetic and hence is not tame in any sense. However, if we only name a chain in this lattice, we obtain a tame structure. A chain of subgroups =B0>B1>\mathbb{Z}=B_{0}>B_{1}>\dots is the same as a valuation v:ω{}v:\mathbb{Z}\rightarrow\omega\cup\{\infty\} defined by

v(a)=max{i:aBi}.v(a)=\max\{i:a\in B_{i}\}.
Theorem.

Let (Bi)i<ω(B_{i})_{i<\omega} be a strictly descending chain of subgroups of \mathbb{Z}, B0=B_{0}=\mathbb{Z}, and let v:ω{}v:\mathbb{Z}\rightarrow\omega\cup\{\infty\} be the valuation defined by

v(x)=max{i:xBi}.v(x)=\max\{i:x\in B_{i}\}.

Then (,0,1,+,v)(\mathbb{Z},0,1,+,v) is dp-minimal. Moreover, (,0,1,+,v)(\mathbb{Z},0,1,+,v) is distal if and only if the size of the quotients Bi/Bi+1B_{i}/B_{i+1} is bounded.

This stays true if we expand the value sort by unary predicates and monotone binary relations. There has been recent interest in dp-minimal expansions of (,+)(\mathbb{Z},+) (e.g. [2], [21], [1], and [22]). Alouf and d’Elbée showed in [2] that if pp is a prime and vpv_{p} denotes the pp-adic valuation, then (,0,1,+,vp)(\mathbb{Z},0,1,+,v_{p}) is a minimal expansion of (,0,1,+)(\mathbb{Z},0,1,+) in the sense that there are no proper intermediate expansions. We show that this does not hold true for all valuations and we conjecture that the pp-adic valuations are essentially the only examples with this property among valuations vv such that (,0,1,+,v)(\mathbb{Z},0,1,+,v) is distal.

The proof of the classification theorem for dp-minimal profinite groups consists of three parts: We analyze the algebraic structure of dp-minimal profinite groups in Section 3. This will imply the first part of the theorem. It then remains to show that these groups appear as dp-minimal profinite groups. This is done in Section 4. The case where the group is given by an 𝔽p\mathbb{F}_{p}-vector space has already been done by Maalouf in [10]. We explain this result in Section 4.1. The remaining case is handled by a quantifier elimination result (see Section 4.2). This quantifier elimination result allows us to show that a certain class of profinite groups as prof\mathcal{L}_{\text{prof}}-structures is dp-minimal (Theorem 4.29) and we are able to characterize distality in this class (Theorem 4.32).

We will also apply the quantifier elimination result to valuations on (,+)(\mathbb{Z},+). This will be done in Section 5 where we discuss the second theorem and its consequences for the study of dp-minimal expansions of (,+,0,1)(\mathbb{Z},+,0,1). We also show that the pp-adic valuations have a limit theory (Proposition 5.9) and we consider expansions of (,+,0,1)(\mathbb{Z},+,0,1) given by multiple valuations.

Section 6 contains a few results which are related to dp-minimal profinite groups. We show that our main result implies some structural consequences for uniformly definable families of finite index subgroups in dp-minimal groups (Proposition 6.4). Jarden and Lubotzky [8] showed that two elementarily equivalent profinite groups are isomorphic if one of them is finitely generated. This was generalized to strongly complete profinite groups by Helbig [7]. We will give an alternative proof for these results in Section 6.2. Finally, we prove a result about uniformly definable families of normal subgroups in NTP2 groups (Proposition 6.13)222Thanks to Pierre Simon for bringing this question to my attention.: If such a family is closed under finite intersections, then it must be defined by an NIP formula.

Acknowledgments

I would like to thank my supervisor, Katrin Tent, for her help and support during the last years and for giving me the opportunity to work on these exciting topics. I would also like to thank Pierre Simon for interesting and useful discussions and valuable suggestions while I visited UC Berkeley.

2 Preliminaries

We assume that the reader is familiar with both profinite groups and model theory. We will give a quick overview about the notions and tools that are used to prove the main result.

2.1 Profinite groups

A topological group is profinite if it is the inverse limit of an inverse system of (discrete) finite groups. This condition is equivalent to the group being Hausdorff, compact, and totally disconnected. If GG is a profinite group, then

GlimG/NG\cong\varprojlim G/N

where NN ranges over all open normal subgroups.

The open subgroups generate the topology on GG, i.e. every open set is a union of cosets of open subgroups. A fundamental system of open subgroups is a family \mathcal{F} consisting of open subgroups which generate the topology on GG. Equivalently, every open subgroup of GG contains a subgroup in \mathcal{F}. If 𝒫\mathcal{P} is a property of groups, we will say that GG is virtually 𝒫\mathcal{P} if GG has an open subgroup HH which satisfies 𝒫\mathcal{P}.

We will use a number of results about the structure of abelian profinite groups. Recall that a profinite group is pro-pp if it is the inverse limit of finite pp-groups. A free abelian pro-pp group is a direct product of copies of p\mathbb{Z}_{p}.

Proposition 2.1 (Theorem 4.3.4 of [13]).

Let pp be a prime.

  1. (a)

    If GG is a torsion free pro-pp abelian group, then GG is a free abelian pro-pp group.

  2. (b)

    Let GG be a finitely generated pro-pp abelian group. Then the torsion subgroup tor(G)\mathrm{tor}(G) is finite and

    GFtor(G)G\cong F\oplus\mathrm{tor}(G)

    where FF is a free pro-pp abelian group of finite rank.

Proposition 2.2 (Corollary 4.3.9 of [13]).

Let GG be a torsion profinite abelian group. Then there is a finite set of primes π\pi and a natural number ee such that

Gpπ(i=1e(m(i,p)Cpi))G\cong\prod_{p\in\pi}(\prod_{i=1}^{e}(\prod_{m(i,p)}C_{p^{i}}))

where each m(i,p)m(i,p) is a cardinal and each CpiC_{p^{i}} is the cyclic group of order pip^{i}. In particular, GG is of finite exponent.

Proposition 2.3 (Proposition 1.13 and Proposition 1.14 of [6]).

Let GG be a pro-pp group. Then GG is (topologically) finitely generated if and only if the Frattini subgroup Φ(G)=Gp[G,G]¯\Phi(G)=\overline{G^{p}[G,G]} is open in GG.

Proposition 2.4.

Let AA be an abelian profinite group. Then nAAnA\leq A is an open subgroup for all n1n\geq 1 if and only if

Appαp×ApA\cong\prod_{p}\mathbb{Z}_{p}^{\alpha_{p}}\times A_{p}

where αp<ω\alpha_{p}<\omega and ApA_{p} is a finite abelian pp-group for each prime pp.

Proof.

An abelian profinite group is the direct product of its pp-Sylow subgroups. Let PP be a pp-Sylow subgroup of AA. If pPPpP\leq P has finite index, then PP is finitely generated by Proposition 2.3. Then by Proposition 2.1 the pp-Sylow subgroup PP has the desired form. ∎

We will also need the following result by Zelmanov:

Theorem 2.5 (Theorem 2 of [24]).

Every infinite compact group has an infinite abelian subgroup.

We will view profinite groups as two-sorted structures in the following language which was introduced in [19]:

Definition 2.6.

prof\mathcal{L}_{\text{prof}} is a two-sorted language containing the group sort 𝒢\mathcal{G} and the index sort \mathcal{I}. The language prof\mathcal{L}_{\text{prof}} then consists of:

  • the usual language of groups on 𝒢\mathcal{G},

  • a binary relation \leq on \mathcal{I}, and

  • a binary relation K𝒢×K\subseteq\mathcal{G}\times\mathcal{I}.

Remark 2.7.

A profinite group GG together with a fundamental system of open subgroups {Ki:iI}\{K_{i}:i\in I\} can be viewed as an prof\mathcal{L}_{\text{prof}} structure (G,I)(G,I) as follows:

  • we set iji\leq j if and only if KiKjK_{i}\supseteq K_{j}, and

  • the relation KK is defined by K(G,i)=KiK(G,i)=K_{i} for all iIi\in I.

2.2 Model theoretic notions of complexity

We will mostly work in the context of an NIP theory. We use [16] as our main reference for this section.

2.2.1 The independence property

An important class of model theoretic theories is the class of NIP (or dependent) theories, i.e. the class of theories which cannot code the \in-relation on an infinite set. This notion was introduced by Shelah.

Definition 2.8.

A formula φ(x,y)\varphi(x,y) has the independence property (IP) if there are sequences (ai)i<ω(a_{i})_{i<\omega} and (bJ)Jω(b_{J})_{J\subseteq\omega} such that

φ(ai,bJ)iJ.\models\varphi(a_{i},b_{J})\iff i\in J.

We say that φ(x,y)\varphi(x,y) has NIP if it does not have IP. This notion is symmetric in the sense that a formula φ(x,y)\varphi(x,y) has NIP if and only if the formula ψ(y,x)φ(x,y)\psi(y,x)\equiv\varphi(x,y) has NIP (see Lemma 2.5 of [16]).

We will make use of the following characterization of IP:

Lemma 2.9 (Lemma 2.7 of [16]).

A formula φ(x,y)\varphi(x,y) has IP if and only if there exists an indiscernible sequence (ai)i<ω(a_{i})_{i<\omega} and a tuple bb such that

φ(ai,b)i is odd.\models\varphi(a_{i},b)\iff i\text{ is odd}.

We call a theory NIP if all formulas have NIP.

Definition 2.10.

A subset XMTX\subseteq M\models T is externally definable if there is a formula φ(x,y)\varphi(x,y), an elementary expansion MM^{*} of MM, and an element bMb\in M^{*} such that X=φ(M,b)X=\varphi(M,b).

By a result of Shelah, naming all externally definable sets in an NIP structure preserves NIP:

Theorem 2.11 (Proposition 3.23 and Corollary 3.24 of [16]).

Let MM be a model of an NIP theory and let MShM^{\text{Sh}} be the Shelah expansion, i.e. the expansion of MM by all externally definable sets. Then MShM^{\text{Sh}} has quantifier elimination and is NIP.

Theorem 2.12 (Baldwin-Saxl, Theorem 2.13 of [16]).

Let GG be an NIP group and let {Hi:iI}\{H_{i}:i\in I\} be a family of uniformly definable subgroups of GG. Then there is a constant KK such that for any finite subset JIJ\subseteq I there is J0JJ_{0}\subseteq J of size |J0|K|J_{0}|\leq K such that

{Hi:iJ}={Hi:iJ0}.\bigcap\{H_{i}:i\in J\}=\bigcap\{H_{i}:i\in J_{0}\}.

As an easy consequence we obtain:

Corollary 2.13.

If (G,I)(G,I) is an NIP profinite group, then {Ki:iI}\{K_{i}:i\in I\} can only contain finitely many subgroups of any given finite index.

By a result of Shelah, abelian subgroups of NIP groups have definable envelopes given by centralizers of definable sets:

Theorem 2.14 (Proposition 2.27 of [16]).

Let GG be an NIP group and let XX be a set of commuting elements. Then there is a formula φ(x,y)\varphi(x,y) and a parameter bb (in some elementary extension GG^{*}) such that Cen(Cen(φ(G,b)))\mathrm{Cen}(\mathrm{Cen}(\varphi(G^{*},b))) is an abelian (definable) subgroup of GG^{*} and contains XX.

2.2.2 Dp-minimality

NIP theories admit a notion of dimension given by dp-rank:

Definition 2.15 (Definition 4.2 of [16]).

Let pp be a partial type over a set AA, and let κ\kappa be a cardinal. We define

dp-rk(p,A)<κ\textrm{dp-rk}(p,A)<\kappa

if and only if for every family (It)t<κ(I_{t})_{t<\kappa} of mutually indiscernible sequences over AA and bpb\models p, one of these sequences is indiscernible over AbAb.

A theory is called dp-minimal if dp-rk(x=x,)=1\textrm{dp-rk}(x=x,\emptyset)=1 where xx is a singleton. We call a multi-sorted theory with distinguished home-sort dp-minimal if it is NIP and it is dp-minimal in the home-sort, i.e. dp-rk(x=x,)=1\textrm{dp-rk}(x=x,\emptyset)=1 where xx is a singleton in the home-sort.

Remark 2.16.

As a consequence of the quantifier elimination in Theorem 2.11 the Shelah expansion of a dp-minimal structure is dp-minimal.

We will use the fact that definable subgroups in a dp-minimal group are always comparable in the following sense:

Lemma 2.17 (Claim in Lemma 4.31 of [16]).

Suppose GG is dp-minimal and H1H_{1} and H2H_{2} are definable subgroups. Then |H1:H1H2||H_{1}:H_{1}\cap H_{2}| or |H2:H1H2||H_{2}:H_{1}\cap H_{2}| is finite.

2.2.3 Distality

Distality is a notion introduced by Simon to describe the unstable part of an NIP theory. The general definition of distality is slightly more complicated than the definitions of NIP and dp-minimality (see Definition 2.1 in [15] or Chapter 9 in [16]). In case of a dp-minimal theory distality can be described as follows:

Proposition 2.18.

A dp-minimal theory TT is distal if and only if there is no infinite non-constant totally indiscernible set of singletons.

Proof.

This characterization follows from Example 2.4 and Lemma 2.10 in [15]. ∎

By Exercise 9.12 of [16] distality is preserved under going to TeqT^{\text{eq}}:

Proposition 2.19.

If TT is distal, then so is TeqT^{\mathrm{eq}}.

2.3 Quantifier elimination

Recall that a theory TT has quantifier elimination if every formula is equivalent to a quantifier free formula modulo TT. The proof of Theorem 3.2.5 in [20] gives the following useful criterion for quantifier elimination:

Proposition 2.20.

Let TT be a theory and let φ(x)\varphi(x) be a formula. Then φ(x)\varphi(x) is equivalent to a quantifier free formula modulo TT if and only if for all 1,2T\mathcal{M}_{1},\mathcal{M}_{2}\models T with common substructure 𝒜\mathcal{A} and all a𝒜a\in\mathcal{A} we have

1φ(a)2φ(a).\mathcal{M}_{1}\models\varphi(a)\implies\mathcal{M}_{2}\models\varphi(a).

If TT is a two-sorted theory and the only symbols that connect the two sorts are functions from one sort to the other, then it suffices to check quantifier elimination for very specific formulas:

Lemma 2.21.

Let TT be a theory in a two-sorted language =01{fj:jJ}\mathcal{L}=\mathcal{L}_{0}\cup\mathcal{L}_{1}\cup\{f_{j}:j\in J\} with sorts 𝒮0\mathcal{S}_{0} and 𝒮1\mathcal{S}_{1} where 0\mathcal{L}_{0} is purely in the sort 𝒮0\mathcal{S}_{0}, 1\mathcal{L}_{1} is purely in the sort 𝒮1\mathcal{S}_{1}, and each fjf_{j} is a function from sort 𝒮0\mathcal{S}_{0} to sort 𝒮1\mathcal{S}_{1}. Suppose

  • (a)

    every 1\mathcal{L}_{1}-formula is equivalent to a quantifier free formula modulo TT and

  • (b)

    every formula of the form

    x𝒮0rRφr(x,y¯r,z¯r)\exists x\in\mathcal{S}_{0}\bigwedge_{r\in R}\varphi_{r}(x,\bar{y}_{r},\bar{z}_{r})

    is equivalent to a quantifier free formula modulo TT where xx is a singleton, y¯r𝒮0\bar{y}_{r}\subseteq\mathcal{S}_{0}, z¯r𝒮1\bar{z}_{r}\subseteq\mathcal{S}_{1}, and each φr\varphi_{r} is either a basic 0\mathcal{L}_{0}-formula or is of the form fj(t(x,y¯r))=zf_{j}(t(x,\bar{y}_{r}))=z where tt is an 0\mathcal{L}_{0} term and zz is one of the variables in the tuple z¯r\bar{z}_{r}.

Then TT eliminates quantifiers.

Proof.

To show quantifier elimination it suffices to consider simple existential formulas. Consider a formula of the form

γ𝒮1rRφr(γ,yr¯,zr¯)\exists\gamma\in\mathcal{S}_{1}\bigwedge_{r\in R}\varphi_{r}(\gamma,\bar{y_{r}},\bar{z_{r}})

where γ\gamma is a singleton, yr¯𝒮0\bar{y_{r}}\subseteq\mathcal{S}_{0}, zr¯𝒮1\bar{z_{r}}\subseteq\mathcal{S}_{1}, and each φr\varphi_{r} is a basic formula. We may assume that γ\gamma appears non-trivially in each formula φr\varphi_{r}. Then each φr\varphi_{r} is a basic 1\mathcal{L}_{1}-formula where the variables y¯r\bar{y}_{r} only appear as terms of the form

f(t(yr¯))f(t(\bar{y_{r}}))

where ff is a function symbol and tt is an 0\mathcal{L}_{0}-term. Now the 𝒮1\mathcal{S}_{1}-quantifier can be eliminated by (a).

Now consider a formula of the form

x𝒮0rRφr(x,yr¯,zr¯)\exists x\in\mathcal{S}_{0}\bigwedge_{r\in R}\varphi_{r}(x,\bar{y_{r}},\bar{z_{r}})

where xx is a singleton, yr¯𝒮0\bar{y_{r}}\subseteq\mathcal{S}_{0}, zr¯𝒮1\bar{z_{r}}\subseteq\mathcal{S}_{1}, and each φr\varphi_{r} is a basic formula.

Let R~R\tilde{R}\subseteq R be the set of all rRr\in R such that φr\varphi_{r} is a basic 1\mathcal{L}_{1}-formula. If rR~r\in\tilde{R}, then we may write φr\varphi_{r} as

φrψr(fr¯(tr¯(x,yr¯)),zr¯)\varphi_{r}\equiv\psi_{r}(\bar{f_{r}}(\bar{t_{r}}(x,\bar{y_{r}})),\bar{z_{r}})

where ψr\psi_{r} is a basic 1\mathcal{L}_{1}-formula such that all variables of ψr\psi_{r} are in 𝒮1\mathcal{S}_{1}. Then φr\varphi_{r} is equivalent to

ξ¯𝒮1:(ξ¯=fr¯(tr¯(x,yr¯))ψr(ξ¯,zr¯)).\exists\bar{\xi}\in\mathcal{S}_{1}:(\bar{\xi}=\bar{f_{r}}(\bar{t_{r}}(x,\bar{y_{r}}))\land\psi_{r}(\bar{\xi},\bar{z_{r}})).

Now we may rewrite

x𝒮0rRφr(x,yr¯,zr¯)\exists x\in\mathcal{S}_{0}\bigwedge_{r\in R}\varphi_{r}(x,\bar{y_{r}},\bar{z_{r}})

as a formula of the form

(ξr¯)rR~𝒮1:((rR~ψr(ξ¯,zr¯))(x𝒮0rR~ξr¯=fr¯(tr¯(x,yr¯))rRR~φr(x,yr¯))).\exists(\bar{\xi_{r}})_{r\in\tilde{R}}\in\mathcal{S}_{1}:((\bigwedge_{r\in\tilde{R}}\psi_{r}(\bar{\xi},\bar{z_{r}}))\land(\exists x\in\mathcal{S}_{0}\bigwedge_{r\in\tilde{R}}\bar{\xi_{r}}=\bar{f_{r}}(\bar{t_{r}}(x,\bar{y_{r}}))\land\bigwedge_{r\in R\setminus\tilde{R}}\varphi_{r}(x,\bar{y_{r}}))).

We can now eliminate the 𝒮0\mathcal{S}_{0}-quantifier by (b) and then eliminate the 𝒮1\mathcal{S}_{1}-quantifiers as in the first step. ∎

3 Algebraic properties of dp-minimal profinite groups

We view a profinite group GG together with a fundamental system of open subgroups {Ki:iI}\{K_{i}:i\in I\} as an prof\mathcal{L}_{\text{prof}}-structure (G,I)(G,I) (as in Remark 2.7). The aim of this chapter is to prove the first part of the main theorem: If (G,I)(G,I) is a dp-minimal profinite group, then GG has an open abelian subgroup AA such that either AA is a vector space over 𝔽p\mathbb{F}_{p} for some prime pp, or Appαp×ApA\cong\prod_{p}\mathbb{Z}_{p}^{\alpha_{p}}\times A_{p} where αp<ω\alpha_{p}<\omega and ApA_{p} is a finite abelian pp-group for each prime pp.

Simon showed in [14] that all dp-minimal groups are abelian-by-finite-exponent. An example of a dp-minimal group that is not abelian-by-finite was given by Simonetta in [17].

We will show that all dp-minimal profinite groups have an open abelian subgroup. We will then analyze the structure of this abelian profinite group. For dp-minimal profinite groups the fundamental system of open subgroups can always be replaced by a chain of open subgroups:

Lemma 3.1.

Let (G,I)(G,I) be a dp-minimal profinite group. Then the subgroups

Hi:={Kj:|G:Kj||G:Ki|}H_{i}:=\bigcap\{K_{j}:|G:K_{j}|\leq|G:K_{i}|\}

are uniformly definable open subgroups and hence the topology on GG is generated by a definable chain of open subgroups.

Proof.

The HiH_{i} are open subgroups by Corollary 2.13. By Lemma 2.17 and compactness we can find a constant KK such that for all i,ji,j |Ki:KiKj|<K|K_{i}:K_{i}\cap K_{j}|<K or |Kj:KiKj|<K|K_{j}:K_{i}\cap K_{j}|<K. Given i,jIi,j\in I we have

|G:Ki||G:Kj||Ki:KiKj||Kj:KiKj|.|G:K_{i}|\leq|G:K_{j}|\iff|K_{i}:K_{i}\cap K_{j}|\geq|K_{j}:K_{i}\cap K_{j}|.

Moreover, we have |Ki:KiKj|<K|K_{i}:K_{i}\cap K_{j}|<K or |Kj:KiKj|<K|K_{j}:K_{i}\cap K_{j}|<K. Therefore this is a definable condition and hence the subgroups HiH_{i} are uniformly definable. ∎

In a dp-minimal profinite group we cannot find infinite definable subgroups of infinite index:

Lemma 3.2.

Let (G,I)(G,I) be a dp-minimal profinite group. Let (G,I)(G^{*},I^{*}) be an elementary extension and let H<GH<G^{*} be a definable subgroup. If GHG\cap H is infinite, then |G:H||G^{*}:H| is finite.

Proof.

If |Ki:KiH||K_{i}^{*}:K_{i}^{*}\cap H| is finite for some iIi\in I, then clearly |G:H|<|G^{*}:H|<\infty.

Now assume |Ki:KiH||K_{i}^{*}:K_{i}^{*}\cap H| is infinite for all iIi\in I. We aim to show that |H:KiH||H:K_{i}^{*}\cap H| must be unbounded: Since GHG\cap H is infinite and iIKi=1\bigcap_{i\in I}K_{i}=1, |GH:KiH||G\cap H:K_{i}\cap H| must be unbounded. Therefore |H:KiH||H:K_{i}^{*}\cap H| must be unbounded. This contradicts Lemma 2.17. ∎

As a consequence of Zelmanov’s theorem (Theorem 2.5) and the existence of definable envelopes for abelian subgroups (Theorem 2.14) we get that a dp-minimal profinite group must be virtually abelian:

Proposition 3.3.

Let (G,I)(G,I) be a dp-minimal profinite group. Then GG is virtually abelian.

Proof.

By Theorem 2.5 GG has an infinite abelian subgroup AA. By Theorem 2.14, we can find an elementary extension (G,I)(G^{*},I^{*}), a formula φ(x,y)\varphi(x,y), and a parameter b(G,I)b\in(G^{*},I^{*}) such that Cen(Cen(φ(G,b)))\mathrm{Cen}(\mathrm{Cen}(\varphi(G^{*},b))) is an abelian subgroup of GG^{*} and contains AA. Therefore Cen(Cen(φ(G,b)))\mathrm{Cen}(\mathrm{Cen}(\varphi(G^{*},b))) has finite index in GG^{*} by Lemma 3.2. By elementarity there is some b(G,I)b^{\prime}\in(G,I) such that Cen(Cen(φ(G,b)))\mathrm{Cen}(\mathrm{Cen}(\varphi(G,b^{\prime}))) is an abelian group and has finite index in GG. Moreover, Cen(Cen(φ(G,b)))\mathrm{Cen}(\mathrm{Cen}(\varphi(G,b^{\prime}))) is closed since it is a centralizer. Closed subgroups of finite index are open and therefore Cen(Cen(φ(G,b)))\mathrm{Cen}(\mathrm{Cen}(\varphi(G,b^{\prime}))) is an open abelian subgroup of GG. ∎

We are now able to prove the first part of the main theorem:

Theorem 3.4.

Let (A,I)(A,I) be an abelian dp-minimal profinite group. Then either AA is virtually a direct product of countably many copies of 𝔽p\mathbb{F}_{p} for some prime pp, or Appαp×ApA\cong\prod_{p}\mathbb{Z}_{p}^{\alpha_{p}}\times A_{p} where αp<ω\alpha_{p}<\omega and ApA_{p} is a finite abelian pp-group for each prime pp.

Proof.

Consider the closed subgroup A[n]:={xA:nx=0}A[n]:=\{x\in A:nx=0\}. Suppose there is a minimal nn such that A[n]A[n] is infinite. Then A[n]A[n] has finite index in AA (by Lemma 3.2) and hence is an open subgroup of AA. Therefore we may assume A=A[n]A=A[n]. Now the minimality of nn and Proposition 2.2 imply that nn must be prime and therefore AA is a direct product of copies of 𝔽p\mathbb{F}_{p} (again by Proposition 2.2). Since AA admits a countable fundamental system of open subgroups, this direct product must be a direct product of countably many copies of 𝔽p\mathbb{F}_{p}.

Now assume A[n]A[n] is finite for all nn. Then the closed subgroup nAnA must be open in AA for all nn (by Lemma 3.2). Now Proposition 2.4 implies the theorem. ∎

4 Valued abelian profinite groups

If AA is an abelian group and (Ai)i<ω(A_{i})_{i<\omega} is a strictly descending chain of subgroups such that A0=AA_{0}=A and i<ωAi={0}\bigcap_{i<\omega}A_{i}=\{0\}, then we can define a valuation map v:Aω{}v:A\rightarrow\omega\cup\{\infty\} by setting

v(x)=max{i:xAi}.v(x)=\max\{i:x\in A_{i}\}.

We have v(x)=v(x)=\infty if and only if x=0x=0, and this valuation satisfies the inequality

v(xy)min{v(x),v(y)}v(x-y)\geq\min\{v(x),v(y)\}

where we have equality in case v(x)v(y)v(x)\neq v(y).

The valued group (A,v)(A,v) can be seen as a two-sorted structure consisting of the group AA, the linear order (ω{},)(\omega\cup\{\infty\},\leq), and the valuation v:Aω{}v:A\rightarrow\omega\cup\{\infty\}.

Our goal is to classify dp-minimal profinite groups up to finite index. We know by Lemma 3.1 that the fundamental system of open subgroups can be assumed to be a chain. Moreover, by Theorem 3.4 we only need to consider groups of the form

i<ω𝔽porppαp×Ap\prod_{i<\omega}\mathbb{F}_{p}\quad\text{or}\quad\prod_{p}\mathbb{Z}_{p}^{\alpha_{p}}\times A_{p}

where αp<ω\alpha_{p}<\omega and ApA_{p} is a finite abelian pp-group for each prime pp.

If AA is such a group and {Bi:i<ω}\{B_{i}:i<\omega\} is a fundamental system of open subgroups which is given by a strictly descending chain, then the above construction yields a definable valuation v:Aω{}v:A\rightarrow\omega\cup\{\infty\}. Conversely, given such a valuation vv, we can recover the fundamental system of open subgroups by setting

Bi={aA:v(a)i}.B_{i}=\{a\in A:v(a)\geq i\}.

Hence the valuation and the fundamental system are interdefinable.

We will show that if AA is of the above form, then AA admits a fundamental system given by a chain of open subgroups such that the expansion of AA by the corresponding valuation (and hence the corresponding prof\mathcal{L}_{\text{prof}}-structure) is dp-minimal. If A=i<ω𝔽pA=\prod_{i<\omega}\mathbb{F}_{p}, this follows from results by Maalouf in [10] and will be explained in Section 4.1.

Definition 4.1.
  1. (a)

    The subgroups Bi={aA:v(a)i}B_{i}=\{a\in A:v(a)\geq i\} are called the vv-balls of radius ii. We will also denote them by BivB_{i}^{v} to emphasize that they correspond to the valuation vv.

  2. (b)

    A valuation v:Aω{}v:A\rightarrow\omega\cup\{\infty\} is good if for all i<ωi<\omega the subgroup BiB_{i} is of the form Bi=nAB_{i}=nA for some positive integer nn.

In case Appαp×ApA\cong\prod_{p}\mathbb{Z}_{p}^{\alpha_{p}}\times A_{p}, we will prove a quantifier elimination result for good valuations. Note that by Proposition 2.4 each such group can be equipped with a good valuation such that {Bi:i<ω}\{B_{i}:i<\omega\} is a fundamental system of open subgroups. We will show the following theorem:

Theorem 4.2.

Let Appαp×ApA\cong\prod_{p}\mathbb{Z}_{p}^{\alpha_{p}}\times A_{p} as above and let vv be a good valuation. Then the structure (A,+,v)(A,+,v) is dp-minimal. Moreover, it is distal if and only if the size of the quotients Bi/Bi+1B_{i}/B_{i+1} is bounded.

This theorem will be proven in this chapter. If π\pi is a set of primes, a natural number n1n\geq 1 is called a π\pi-number if the prime decomposition of nn only contains primes in π\pi. An immediate consequence of the above theorem is the following:

Corollary 4.3.

Let (πi)i<ω(\pi_{i})_{i<\omega} be a sequence of finite non-empty disjoint sets of primes. For each i<ωi<\omega fix a finite non-trivial abelian group AiA_{i} such that |Ai||A_{i}| is a πi\pi_{i}-number. Set

A=i<ωAiA=\prod_{i<\omega}A_{i}

and let vv be the valuation defined by

v((ai)i<ω)=min{i:ai0}.v((a_{i})_{i<\omega})=\min\{i:a_{i}\neq 0\}.

Then (A,+,v)(A,+,v) is dp-minimal but not distal.

Proof.

We have Bkv=(i<k|Ai|)AB_{k}^{v}=(\prod_{i<k}|A_{i}|)A. Hence vv is a good valuation and the theorem applies. ∎

4.1 Valued vector spaces

Valued vector spaces have been studies by S. Kuhlmann and F.-V. Kuhlmann in [9] and by Maalouf in [10]. Set A=i<ω𝔽pA=\prod_{i<\omega}\mathbb{F}_{p} and let v:Aω{}v:A\rightarrow\omega\cup\{\infty\} be the valuation given by

v((xi)i<ω)=min{i:xi0}.v((x_{i})_{i<\omega})=\min\{i:x_{i}\neq 0\}.

It follows from results by Maalouf in [10] that this valued abelian profinite group is dp-minimal:

Proposition 4.4.

The valued abelian profinite group (A,v)(A,v) is dp-minimal.

Proof.

Set B=i<ω𝔽pB=\bigoplus_{i<\omega}\mathbb{F}_{p} and let w:Bω{}w:B\rightarrow\omega\cup\{\infty\} be the valuation given by

w((xi)i<ω)=min{i:xi0}.w((x_{i})_{i<\omega})=\min\{i:x_{i}\neq 0\}.

By Proposition 4 of [10] the valued vector space (B,w)(B,w) is C-minimal and hence dp-minimal (by Theorem A.7 of [16]).

Théorème 1 of [10] implies that (A,v)(A,v) and (B,w)(B,w) are elementarily equivalent. Hence (A,v)(A,v) is dp-minimal. ∎

Remark 4.5.

The last step of the previous proof also follows from results in Section 6.1. Let (B,w)(B,w) be as in the proof of Proposition 4.4 and set

Bi={xB:w(x)i}.B_{i}=\{x\in B:w(x)\geq i\}.

Then Alimi<ωB/BiA\cong\varprojlim_{i<\omega}B/B_{i} and hence (A,v)(A,v) is dp-minimal by Lemma 6.2.

4.2 A quantifier elimination result

We denote the set of primes by \mathbb{P}. For each prime pp\in\mathbb{P} we fix an integer αp0\alpha_{p}\geq 0 and a finite pp-group ApA_{p}. Let

Zppαp×ApZ\prec\prod_{p\in\mathbb{P}}\mathbb{Z}_{p}^{\alpha_{p}}\times A_{p}

be an abelian group that is (as a pure group) an elementary substructure. We will always assume ZZ to be infinite. We fix a set of constants {cj:j<ω}Z\{c_{j}:j<\omega\}\subseteq Z containing 0 such that the set is dense with respect to the profinite topology on ZZ and contains every torsion element. It follows from Proposition 2.4 that the set of constants is also dense with respect to the profinite topology on ppαp×Ap\prod_{p\in\mathbb{P}}\mathbb{Z}_{p}^{\alpha_{p}}\times A_{p}.

Definition 4.6.

If π\pi is a set of primes, we set

Zπ=Z(ppαp×pπAp)andAπ=ZpπAp.Z_{\pi}=Z\cap(\prod_{p\in\mathbb{P}}\mathbb{Z}_{p}^{\alpha_{p}}\times\prod_{p\in\mathbb{P}\setminus\pi}A_{p})\quad\text{and}\quad A_{\pi}=Z\cap\prod_{p\in\pi}A_{p}.

Note that we have Z=Zπ×AπZ=Z_{\pi}\times A_{\pi} for any set π\pi\subseteq\mathbb{P}. The group AπA_{\pi} is the π\pi-torsion part of ZZ and the group ZπZ_{\pi} has no π\pi-torsion.

Let vv be a good valuation and set I:=ω{+,}I:=\omega\cup\{+\infty,-\infty\}. For each l1l\geq 1 we define a function vl:ZIv^{l}:Z\rightarrow I by

vl(a)={+iff a=0iiff alBilBi+1iff alZ.v^{l}(a)=\begin{cases}+\infty&\text{iff }a=0\\ i&\text{iff }a\in lB_{i}\setminus lB_{i+1}\\ -\infty&\text{iff }a\not\in lZ.\end{cases}

Note that if alZa\in lZ, then vl(a)=v(a/l)v^{l}(a)=v(a/l). Now ZZ together with the valuation vv may be viewed as a two-sorted structure with group sort 𝒵\mathcal{Z} and value sort \mathcal{I} in the language =𝒵v\mathcal{L}^{-}=\mathcal{L}_{\mathcal{Z}}\cup\mathcal{L}_{v}\cup\mathcal{L}_{\mathcal{I}}^{-} where

  • 𝒵={+,,cj:j<ω}\mathcal{L}_{\mathcal{Z}}=\{+,-,c_{j}:j<\omega\} is the obvious language on ZZ,

  • v={vl:l1}\mathcal{L}_{v}=\{v^{l}:l\geq 1\} consists of symbols for the functions vlv^{l}, and

  • ={,0,+,}\mathcal{L}_{\mathcal{I}}^{-}=\{\leq,0,+\infty,-\infty\} is the obvious language on II.

Since we consider the group ZZ and the constants cjc_{j} to be fixed, this structure only depends on the valuation and we denote it by (Z,v)(Z,v).

We define the following binary relations on II:

  • Indkπ,l(i,j)ij\mathrm{Ind}_{k}^{\pi,l}(i,j)\iff i\leq j and |ZπlBi:ZπlBj|k|Z_{\pi}\cap lB_{i}:Z_{\pi}\cap lB_{j}|\geq k,

  • Divqkπ,l(i,j)ij\mathrm{Div}_{q^{k}}^{\pi,l}(i,j)\iff i\leq j and qkαqq^{k\alpha_{q}} divides |ZπlBi:ZπlBj||Z_{\pi}\cap lB_{i}:Z_{\pi}\cap lB_{j}|,

where π\pi is a finite set of primes, qπq\in\pi is a prime, and k0k\geq 0. We set Indkπ,l(i,+)\mathrm{Ind}_{k}^{\pi,l}(i,+\infty) and Divqkπ,l(i,+)\mathrm{Div}_{q^{k}}^{\pi,l}(i,+\infty) to be always true.

Observation 4.7.
  • (a)

    If qπq\in\pi then qkαqq^{k\alpha_{q}} divides |ZπlBi:ZπlBj||Z_{\pi}\cap lB_{i}:Z_{\pi}\cap lB_{j}| if and only if (ZπlBi)/(ZπlBj)(Z_{\pi}\cap lB_{i})/(Z_{\pi}\cap lB_{j}) has an element of order qkq^{k}. In particular, the predicate Divqkπ,l\mathrm{Div}_{q^{k}}^{\pi,l} is definable.

  • (b)

    In the standard model Divqkπ,l(i,j)\mathrm{Div}_{q^{k}}^{\pi,l}(i,j) is equivalent to the statement that qkq^{k} divides |qlBi:qlBj||\mathbb{Z}_{q}\cap lB_{i}:\mathbb{Z}_{q}\cap lB_{j}|. In that sense the expression |qlBi:qlBj||\mathbb{Z}_{q}\cap lB_{i}:\mathbb{Z}_{q}\cap lB_{j}| makes sense even in non-standard models.

  • (c)

    We have xnZx\in nZ if and only if vn(x)0v^{n}(x)\geq 0. Hence the subgroups nZnZ are quantifier free 0-definable. Since the subgroups nZnZ generate the profinite topology on ZZ, this implies that the open subgroups ZπZ_{\pi} are quantifier free 0-definable for finite subsets π\pi\subseteq\mathbb{P}. Moreover, in that case AπA_{\pi} is also quantifier free 0-definable since it is a finite set of constants.

Let VV be the set of good valuations on ZZ. We set TZ:=vVTh((Z,v))T_{Z}:=\bigcap_{v\in V}\mathrm{Th}((Z,v)) to be the common \mathcal{L}^{-}-Theory of structures (Z,v)(Z,v), vVv\in V. The following quantifier elimination result will be shown in the next sections:

Theorem 4.8.

Let \mathcal{L}_{\mathcal{I}}\supseteq\mathcal{L}_{\mathcal{I}}^{-} be an expansion on the sort \mathcal{I} and let TTZT\supseteq T_{Z} be an expansion of TZT_{Z} to the language =𝒵v\mathcal{L}=\mathcal{L}_{\mathcal{Z}}\cup\mathcal{L}_{v}\cup\mathcal{L}_{\mathcal{I}}. Suppose that:

  1. 1.

    The relations Divqkπ,l\mathrm{Div}_{q^{k}}^{\pi,l} and Indkπ,l\mathrm{Ind}_{k}^{\pi,l} are quantifier free 0-definable modulo TT.

  2. 2.

    The successor function on \mathcal{I} is contained in \mathcal{L}_{\mathcal{I}}.

  3. 3.

    Every \mathcal{L}_{\mathcal{I}}-formula is equivalent to a quantifier free \mathcal{L}-formula modulo TT.

Then TT eliminates quantifiers.

To prove the quantifier elimination result we will need to understand formulas that describe systems of linear congruences. Therefore we will need to understand linear congruences in models of the theory TT.

4.2.1 Linear congruences in \mathbb{Z}

We will need generalizations of the following well-known fact:

Fact 4.9.

A linear congruence nxamodmnx\equiv a\mod m in \mathbb{Z} has a solution if and only if d=gcd(n,m)d=\gcd(n,m) divides aa. In that case it has exactly dd many solutions modulo mm. If ss is a solution, then a complete system of solutions modulo mm is given by

s+tm/d,t=0,d1.s+tm/d,\quad t=0,\dots d-1.
Observation 4.10.

4.9 has two important consequences:

  1. (a)

    If nxamodmnx\equiv a\mod m has a solution and n=gcd(n,m)n=\gcd(n,m), then nn divides aa and hence a/na/n is a solution.

  2. (b)

    If nxamodmnx\equiv a\mod m has a solution, then all solutions agree modulo m/dm/d.

Part (a) will be important since in that case a solution will be determined by the constant aa. Part (b) tells us that solutions of linear congruences can “collapse”. We will need to understand this collapsing of solutions.

We now fix a group 𝔸\mathbb{A} of the form 𝔸=ppαp\mathbb{A}=\prod_{p\in\mathbb{P}}\mathbb{Z}_{p}^{\alpha_{p}}, αp<ω\alpha_{p}<\omega. If nn is a positive integer, let n(p)n(p) be the unique integer such that n=pn(p)n=\prod p^{n(p)}. Note that 4.9 can be applied to p\mathbb{Z}_{p} because p/kp=/pk(p)\mathbb{Z}_{p}/k\mathbb{Z}_{p}=\mathbb{Z}/p^{k(p)}\mathbb{Z}.

We consider linear congruences

nxamodmnx\equiv a\mod m

in 𝔸\mathbb{A} where nn and mm are positive integers and a𝔸a\in\mathbb{A}. Note that solving the above linear congruence is equivalent to solving it in each copy of p\mathbb{Z}_{p} in the product 𝔸=ppαp\mathbb{A}=\prod_{p\in\mathbb{P}}\mathbb{Z}_{p}^{\alpha_{p}}:

Lemma 4.11.
  1. (a)

    Let nxamodmnx\equiv a\mod m be a linear congruence in 𝔸\mathbb{A}. Write

    a=(ap,i)p,i<αp𝔸=ppαp.a=(a_{p,i})_{p\in\mathbb{P},i<\alpha_{p}}\in\mathbb{A}=\prod_{p\in\mathbb{P}}\mathbb{Z}_{p}^{\alpha_{p}}.

    The solutions for nxamodmnx\equiv a\mod m in 𝔸\mathbb{A} are exactly the tuples s=(sp,i)p,i<αps=(s_{p,i})_{p\in\mathbb{P},i<\alpha_{p}} where each sp,is_{p,i} is a solution for nxap,imodmnx\equiv a_{p,i}\mod m in p\mathbb{Z}_{p}.

  2. (b)

    Set d=gcd(n,m)d=\gcd(n,m). Then the linear congruence nxamodmnx\equiv a\mod m has a solution if and only if dd divides aa in 𝔸\mathbb{A} (i.e. ad𝔸a\in d\mathbb{A}). In that case it has exactly p|dpαpd(p)\prod_{p|d}p^{\alpha_{p}d(p)} many solutions modulo mm in 𝔸\mathbb{A}.

We call a finite family of linear congruences (and negations of linear congruences) a system of linear congruences. Recall Bézout’s identity:

Fact 4.12 (Bézout’s identity).

If a1,ana_{1},\dots a_{n} are integers, then gcd(a1,an)\gcd(a_{1},\dots a_{n}) is a \mathbb{Z}-linear combination of a1,ana_{1},\dots a_{n}.

We will look at systems of linear congruences where the modulus is fixed:

Proposition 4.13.

Let nrxarmodm,rR,n_{r}x\equiv a_{r}\mod m,\,r\in R, be a system 𝒮\mathcal{S} of linear congruences in 𝔸\mathbb{A} (where RR is a finite set). Set n=gcd(nr:rR)n=\gcd(n_{r}:r\in R) and d=gcd(n,m)d=\gcd(n,m). By Bézout’s identity we can find integers zrz_{r} such that n=rRzrnrn=\sum_{r\in R}z_{r}n_{r}. Put a=rRzrara=\sum_{r\in R}z_{r}a_{r}.

  1. (a)

    If the system 𝒮\mathcal{S} has a common solution, the solutions of 𝒮\mathcal{S} are exactly the solutions of nxamodmnx\equiv a\mod m.

  2. (b)

    Set k=n/dk=n/d and dr=nr/kd_{r}=n_{r}/k. Then the system 𝒮\mathcal{S} has a solution if and only if the system 𝒯\mathcal{T}:

    drxarmodm,rR,d_{r}x\equiv a_{r}\mod m,\quad r\in R,

    has a solution. Moreover, the systems 𝒮\mathcal{S} and 𝒯\mathcal{T} have the same number of solutions modulo mm.

Proof.

(a) It suffices to show this for each factor in the product 𝔸=ppαp\mathbb{A}=\prod_{p\in\mathbb{P}}\mathbb{Z}_{p}^{\alpha_{p}}. Hence we may assume 𝔸=p\mathbb{A}=\mathbb{Z}_{p} and m=pm(p)m=p^{m(p)}. Clearly any common solution of the system 𝒮\mathcal{S} solves nxamodmnx\equiv a\mod m.

Now suppose ss is a solution of 𝒮\mathcal{S} (and hence a solution of nxamodmnx\equiv a\mod m). Then by 4.9 all solutions of nxamodmnx\equiv a\mod m are of the form s+tm/ds+tm/d where d=gcd(n,m)d=\gcd(n,m). Fix rRr\in R. Now dd divides dr=gcd(nr,m)d_{r}=\gcd(n_{r},m), say dr=krdd_{r}=k_{r}d. Therefore

s+tm/d=s+tkrm/drs+tm/d=s+tk_{r}m/d_{r}

solves nrxarmodmn_{r}x\equiv a_{r}\mod m for all t=0,d1t=0,\dots d-1 (by 4.9). Hence every solution of nxamodmnx\equiv a\mod m solves 𝒮\mathcal{S}.

(b) We have n=gcd(nr:rR)=rRzrnrn=\gcd(n_{r}:r\in R)=\sum_{r\in R}z_{r}n_{r}. If we divide by kk, we get that d=gcd(dr:rR)=rRzrdrd=\gcd(d_{r}:r\in R)=\sum_{r\in R}z_{r}d_{r}. We aim to show that 𝒮\mathcal{S} has a solution if and only if 𝒯\mathcal{T} has a solution. If ss is a solution for 𝒮\mathcal{S}, then ksks solves 𝒯\mathcal{T}. Now assume that 𝒯\mathcal{T} has a solution. Then by (a) the system 𝒯\mathcal{T} has the same solutions as the linear congruence

dxamodm.dx\equiv a\mod m.

Since we assume that 𝒯\mathcal{T} has a solution, this implies that d=gcd(d,m)d=\gcd(d,m) divides aa (by part (b) of Lemma 4.11). Then the linear congruence

nxamodmnx\equiv a\mod m

also has solutions by part (b) of Lemma 4.11 since d=gcd(n,m)d=\gcd(n,m) divides aa. If ss solves nxamodmnx\equiv a\mod m, then ksks solves dxamodmdx\equiv a\mod m and hence is a solution for 𝒯\mathcal{T}. This implies that ss solves 𝒮\mathcal{S}. Hence 𝒮\mathcal{S} has a solution if and only if 𝒯\mathcal{T} has a solution. Moreover, if 𝒮\mathcal{S} and 𝒯\mathcal{T} have solutions, then by (a) the solutions of 𝒮\mathcal{S} are exactly the solutions of nxamodmnx\equiv a\mod m and the solutions of 𝒯\mathcal{T} are exactly the solutions of dxamodmdx\equiv a\mod m. Hence they have the same number of solutions modulo mm by part (b) of Lemma 4.11. ∎

We will now consider systems of linear congruences where we vary the modulus:

Lemma 4.14.

Let nxamodpmnx\equiv a\mod p^{m} be a linear congruence in p\mathbb{Z}_{p}. Set d=gcd(n,pm)d=\gcd(n,p^{m}) and suppose l>0l>0 divides pmp^{m} such that dd divides pm/lp^{m}/l. Then

nxamodpmandnxamoddlnx\equiv a\mod p^{m}\quad\text{and}\quad nx\equiv a\mod dl

have exactly dd solutions modulo pmp^{m} respectively dldl and all these solutions agree modulo ll.

Proof.

The assumption implies that dldl divides pmp^{m}. Therefore

d=gcd(n,pm)=gcd(n,dl).d=\gcd(n,p^{m})=\gcd(n,dl).

By 4.9 the congruences have a solution if and only if dd divides aa. In that case nxamodpmnx\equiv a\mod p^{m} has exactly dd solutions modulo pmp^{m} and the congruence nxamoddlnx\equiv a\mod dl has exactly dd solutions modulo dldl. Moreover, by part (b) of 4.10 all these solutions agree modulo ll. ∎

Proposition 4.15.

Let nxamodmnx\equiv a\mod m be a linear congruence in 𝔸\mathbb{A}. Set d=gcd(n,m)d=\gcd(n,m). Suppose l>0l>0 divides mm and is such that for all p|dp|d we have pd(p)|(m/l)p^{d(p)}|(m/l) or pp does not divide (m/l)(m/l). Set

k=p|d,p|(m/l)pd(p).k=\prod_{p|d,p|(m/l)}p^{d(p)}.

Then the linear congruences

nxamodmandnxamodklnx\equiv a\mod m\quad\text{and}\quad nx\equiv a\mod kl

have the same number of solutions modulo mm respectively klkl. Moreover, if XX is the set of solutions modulo mm of nxamodmnx\equiv a\mod m, YY is the set of solutions modulo klkl of nxamodklnx\equiv a\mod kl, and X/lX/l and Y/lY/l are the images of XX and YY in 𝔸/l𝔸\mathbb{A}/l\mathbb{A}, then X/l=Y/lX/l=Y/l and each element in X/lX/l (resp. Y/lY/l) has exactly p|d,p|(m/l)pαpd(p)\prod_{p|d,p|(m/l)}p^{\alpha_{p}d(p)} many preimages in XX (resp. YY).

Proof.

By an application of Lemma 4.11 it suffices to show this in case 𝔸=p\mathbb{A}=\mathbb{Z}_{p}. Hence we will assume 𝔸=p\mathbb{A}=\mathbb{Z}_{p}.

If pp does not divide dd, then dd is a unit in p\mathbb{Z}_{p} and hence each of the congruences has a unique solution in p\mathbb{Z}_{p}.

Hence we may assume pp divides dd. If pp does not divide m/lm/l, then m(p)=l(p)m(p)=l(p) and k(p)=0k(p)=0. Then mp=klp=lpm\mathbb{Z}_{p}=kl\mathbb{Z}_{p}=l\mathbb{Z}_{p} and therefore the linear congruences

nxamodm,nxamodkl,andnxamodlnx\equiv a\mod m,\quad\quad nx\equiv a\mod kl,\quad\text{and}\quad nx\equiv a\mod l

have the same solutions (in p\mathbb{Z}_{p}). Since solutions modulo mm (resp. modulo klkl) are the same as solutions modulo ll, each element of X/lX/l (resp. Y/lY/l) has a unique preimage in XX (resp. YY).

Now assume pp divides m/lm/l. Then by assumption pd(p)p^{d(p)} divides m/lm/l. In that case the result follows by Lemma 4.14. Note that each element in X/lX/l (resp. Y/lY/l) has exactly dd preimages in XX (resp. YY). ∎

4.2.2 Linear congruences in 𝒵\mathcal{Z}

Fix TTZT\supset T_{Z} as in Theorem 4.8 (for a group Zppαp×ApZ\prec\prod_{p\in\mathbb{P}}\mathbb{Z}_{p}^{\alpha_{p}}\times A_{p} as in the beginning of Section 4.2). We have vl(a)iv^{l}(a)\geq i if and only if aBivl=lBia\in B_{i}^{v^{l}}=lB_{i}. Therefore we will consider certain formulas as linear congruences:

vl(nxa)i\displaystyle v^{l}(nx-a)\geq i nxamodlBi,\displaystyle\iff nx\equiv a\mod{lB_{i}},
vl(nxa)<i\displaystyle v^{l}(nx-a)<i nxamodlBi.\displaystyle\iff nx\not\equiv a\mod{lB_{i}}.

Here xx will be a variable and aa will be a constant. The integer nn will be part of the formula. In particular, it will always be a standard integer. Recall that for a subset π\pi\subseteq\mathbb{P} a natural number n1n\geq 1 is called a π\pi-number if the prime decomposition of nn only contains primes in π\pi.

We will often work in the π\pi-torsion free group ZπZ_{\pi} defined in Definition 4.6. If we assume that π\pi is finite, then by part (c) of 4.7 the subgroup ZπZ_{\pi} is quantifier free 0-definable. If MTM\models T is any model, then we set Zπ(M)Z_{\pi}(M) to be the subgroup defined by the formula which defines ZπZ_{\pi} in ZZ. The subgroup Aπ(M)A_{\pi}(M) is defined analogously.

If we use the notation in part (b) of 4.7, then

gcd(n,lBi):=gcd(n,{p:αp>0}|p:plBi|)\gcd(n,lB_{i}):=\gcd(n,\prod_{\{p:\alpha_{p}>0\}}|\mathbb{Z}_{p}:\mathbb{Z}_{p}\cap lB_{i}|)

is well-defined even if lBilB_{i} has infinite index because nn is always a standard integer. Therefore the results in Section 4.2.1 can be formulated using the divisibility predicates and they will hold true for models of TT.

Proposition 4.16.

Let MM be a model of TT and let nxamodlBinx\equiv a\mod lB_{i} be a linear congruence in Z(M)Z(M). Let π\pi be a finite set of primes such that nn is a π\pi-number and aZπ(M)a\in Z_{\pi}(M). Then the linear congruence has a solution in Zπ(M)Z_{\pi}(M) if and only if d=gcd(n,lBi)d=\gcd(n,lB_{i}) divides aa (i.e. adZπ(M)a\in dZ_{\pi}(M)). In that case there are exactly pαpd(p)\prod p^{\alpha_{p}d(p)} many solutions modulo lBilB_{i} in Zπ(M)Z_{\pi}(M).

Proof.

This is essentially part (b) of Lemma 4.11. Since this is a first-order statement, it suffices to consider good valuations vv on ZZ. Since the statement only affects the quotients Z/lBiZ/lB_{i}, we may assume that ZZ is of the form

Z=ppαp×Ap.Z=\prod_{p\in\mathbb{P}}\mathbb{Z}_{p}^{\alpha_{p}}\times A_{p}.

Put 𝔸=ppαp\mathbb{A}=\prod_{p\in\mathbb{P}}\mathbb{Z}_{p}^{\alpha_{p}} and H=pπApH=\prod_{p\not\in\pi}A_{p}. Then

Zπ=𝔸×HZ_{\pi}=\mathbb{A}\times H

and HH has no π\pi-torsion. Write a=a0ha=a_{0}h for a0𝔸a_{0}\in\mathbb{A} and hHh\in H. Then we can apply Lemma 4.11 to the linear congruence

nxa0modlBinx\equiv a_{0}\mod lB_{i}

in 𝔸\mathbb{A}. Note that the linear congruence

nxhmodlBinx\equiv h\mod lB_{i}

has a unique solution modulo lBilB_{i} in HH (namely h/nh/n). This shows the proposition. ∎

Proposition 4.17.

Let MM be a model of TT and let nrxarmodlBi,rR,n_{r}x\equiv a_{r}\mod lB_{i},\,r\in R, be a system of linear congruences. Let π\pi be a finite set of primes such that all nrn_{r} are π\pi-numbers and all ara_{r} are contained in Zπ(M)Z_{\pi}(M). Set n=gcd(nr:rR)n=\gcd(n_{r}:r\in R) and d=gcd(n,lBi)d=\gcd(n,lB_{i}). By Bézout’s identity we can find integers zrz_{r} such that n=rRzrnrn=\sum_{r\in R}z_{r}n_{r}. Put a=rRzrara=\sum_{r\in R}z_{r}a_{r}.

  1. (a)

    If the system has a common solution in Zπ(M)Z_{\pi}(M), the solutions of the system in Zπ(M)Z_{\pi}(M) are exactly the solutions of nxamodlBinx\equiv a\mod lB_{i} in Zπ(M)Z_{\pi}(M).

  2. (b)

    Set k=n/dk=n/d and dr=nr/kd_{r}=n_{r}/k. Then the system has a solution in Zπ(M)Z_{\pi}(M) if and only if the system

    drxarmodlBi,rR,d_{r}x\equiv a_{r}\mod lB_{i},\quad r\in R,

    has a solution in Zπ(M)Z_{\pi}(M). Moreover, these systems have the same number of solutions modulo lBilB_{i} in Zπ(M)Z_{\pi}(M).

Proof.

This follows from Proposition 4.13 by the same arguments that are used in Proposition 4.16. ∎

Proposition 4.18.

Let MM be a model of TT and let nxamodlBinx\equiv a\mod lB_{i} be a linear congruence. Let π\pi be a finite set of primes such that nn is a π\pi-number and aZπ(M)a\in Z_{\pi}(M). Set d=gcd(n,lBi)d=\gcd(n,lB_{i}). Fix u>0u>0 and j>ij>i such that ulBj<lBiulB_{j}<lB_{i} is a subgroup and is such that for all p|dp|d we have that pd(p)p^{d(p)} divides |plBi:pulBi||\mathbb{Z}_{p}\cap lB_{i}:\mathbb{Z}_{p}\cap ulB_{i}| or pp does not divide |plBi:pulBi||\mathbb{Z}_{p}\cap lB_{i}:\mathbb{Z}_{p}\cap ulB_{i}|. Set

k={pd(p):p divides d and |plBi:pulBi|}.k=\prod\{p^{d(p)}\,:\,p\text{ divides }d\text{ and }|\mathbb{Z}_{p}\cap lB_{i}:\mathbb{Z}_{p}\cap ulB_{i}|\}.

Then the linear congruences

nxamodlBiandnxamodkulBjnx\equiv a\mod lB_{i}\quad\text{and}\quad nx\equiv a\mod kulB_{j}

have the same number of solutions modulo lBilB_{i} respectively kulBjkulB_{j} in Zπ(M)Z_{\pi}(M). Moreover, if XX is the set of solutions modulo lBilB_{i} of nxamodlBinx\equiv a\mod lB_{i}, YY is the set of solutions modulo kulBjkulB_{j} of nxamodkulBjnx\equiv a\mod kulB_{j}, and X/ulBjX/ulB_{j} and Y/ulBjY/ulB_{j} are the images of XX and YY modulo ulBjulB_{j}, then X/ulBj=Y/ulBjX/ulB_{j}=Y/ulB_{j} and each element in X/ulBjX/ulB_{j} (resp. Y/ulBjY/ulB_{j}) has exactly p|d,p|(m/l)pαpd(p)\prod_{p|d,p|(m/l)}p^{\alpha_{p}d(p)} many preimages in XX (resp. YY).

Proof.

This follows from Proposition 4.15 by the same arguments that are used in Proposition 4.16. ∎

The following lemma will often be useful:

Lemma 4.19.

Fix a model MTM\models T, let π\pi be a finite set of primes, let aZπ(M)a\in Z_{\pi}(M), and let tt be a π\pi-number. Then the linear congruences

nxamodlBiandtnxtamodtlBinx\equiv a\mod{lB_{i}}\quad\text{and}\quad tnx\equiv ta\mod{tlB_{i}}

have the same solutions in Zπ(M)Z_{\pi}(M).

Proof.

Multiplying by tt is injective since Zπ(M)Z_{\pi}(M) does not have tt-torsion. ∎

4.2.3 Systems of linear congruences in 𝒵\mathcal{Z}

Fix TT as in Theorem 4.8. Note that we assume that the successor function SS (on \mathcal{I}) is contained in \mathcal{L}_{\mathcal{I}}.

Lemma 4.20.

Let M1M_{1} and M2M_{2} be models of TT and let (A,J)(A,J) be a common substructure. Let π\pi be a finite set of primes and let 𝒮:\mathcal{S}:

nrxarmodlBi,rR,n_{r}x\equiv a_{r}\mod lB_{i},\quad r\in R,

be a system of linear congruences where each nrn_{r} is a π\pi-number, arZπ(A)a_{r}\in Z_{\pi}(A), and iJi\in J. Suppose there is a π\pi-number tt and a constant bAb\in A such that tt divides bb and b/tb/t solves 𝒮\mathcal{S} in Zπ(M1)Z_{\pi}(M_{1}). Then b/tb/t solves 𝒮\mathcal{S} in Zπ(M2)Z_{\pi}(M_{2}).

Proof.

We have tt divides bb if and only if vt(b)0v^{t}(b)\geq 0. This does not depend on the model. Moreover, b/tb/t solves nrxarmodlBin_{r}x\equiv a_{r}\mod lB_{i} if and only if vl(nr(b/t)ar)iv^{l}(n_{r}(b/t)-a_{r})\geq i. By Lemma 4.19 we have

vl(nr(b/t)ar)=vtl(nrbtar).v^{l}(n_{r}(b/t)-a_{r})=v^{tl}(n_{r}b-ta_{r}).

Therefore this value does not depend on the model. ∎

Lemma 4.21.

Let M1M_{1} and M2M_{2} be models of TT and let (A,J)(A,J) be a common substructure. Let π\pi be a finite set of primes and let

nrxarmodlBi,rR,n_{r}x\equiv a_{r}\mod lB_{i},\quad r\in R,

be a system of linear congruences where each nrn_{r} is a π\pi-number, arZπ(A)a_{r}\in Z_{\pi}(A), and iJi\in J. Then the system has the same number of solutions modulo lBilB_{i} in Zπ(M1)Z_{\pi}(M_{1}) and Zπ(M2)Z_{\pi}(M_{2}).

Proof.

Set n:=gcd(nr:rR)n:=\gcd(n_{r}:r\in R), say n=rRzrnrn=\sum_{r\in R}z_{r}n_{r} (by Bézout’s identity), and put a:=rRzrara:=\sum_{r\in R}z_{r}a_{r}. Set d:=gcd(n,{p:αp>0}|p:plBi|)d:=\gcd(n,\prod_{\{p:\alpha_{p}>0\}}|\mathbb{Z}_{p}:\mathbb{Z}_{p}\cap lB_{i}|), k=n/dk=n/d, and dr=nr/kd_{r}=n_{r}/k. Then by Proposition 4.17 (b) the system

nrxarmodlBi,rR,n_{r}x\equiv a_{r}\mod lB_{i},\,r\in R,

has the same number of solutions modulo lBilB_{i} in Zπ(M1)Z_{\pi}(M_{1}) (resp. Zπ(M2)Z_{\pi}(M_{2})) as the system

drxarmodlBi,rR.d_{r}x\equiv a_{r}\mod lB_{i},\,r\in R.

We have d=gcd(dr:rR)=rRzrdr.d=\gcd(d_{r}:r\in R)=\sum_{r\in R}z_{r}d_{r}. By Proposition 4.17 (a) any solution of the system

drxarmodlBi,rR,d_{r}x\equiv a_{r}\mod lB_{i},\,r\in R,

is a solution of dxamodlBidx\equiv a\mod lB_{i}. Now by Proposition 4.16 the linear congruence dxamodlBidx\equiv a\mod lB_{i} has a solution if and only if dd divides aa. In that case a/da/d must be a solution and we can apply Lemma 4.20 to see that this must hold true in both models.

Hence Zπ(M1)Z_{\pi}(M_{1}) contains a solution if and only if Zπ(M2)Z_{\pi}(M_{2}) contains a solution. In that case the solutions are exactly the solutions of nxamodlBinx\equiv a\mod lB_{i} and by Proposition 4.16 the number of solutions modulo lBilB_{i} does not depend on the model. ∎

Lemma 4.22.

Let M1M_{1} and M2M_{2} be models of TT and let (A,J)(A,J) be a common substructure. Let π\pi be a finite set of primes and let 𝒮\mathcal{S} be a system

nrxarmodlBir,\displaystyle n_{r}x\equiv a_{r}\mod lB_{i_{r}},\quad rR,\displaystyle r\in R,

of linear congruences where ll and each nrn_{r} is a π\pi-number, arZπ(A)a_{r}\in Z_{\pi}(A), irJi_{r}\in J. Suppose moreover, that the index |Bir:Bir||B_{i_{r}}:B_{i_{r^{\prime}}}| is a π\pi-number whenever it is finite. Fix rmaxRr_{\text{max}}\in R such that irmaxi_{r_{\text{max}}} is maximal. Then 𝒮\mathcal{S} has the same number of solutions modulo lBirmaxlB_{i_{r_{\text{max}}}} in Zπ(M1)Z_{\pi}(M_{1}) and Zπ(M2)Z_{\pi}(M_{2}).

Proof.

If |lBir:lBir||lB_{i_{r}}:lB_{i_{r^{\prime}}}| is finite, then there is a π\pi-number tt such that lBir=tlBirlB_{i_{r}}=tlB_{i_{r^{\prime}}}. Lemma 4.19 allows us to replace the linear congruence

nrxarmodlBirn_{r^{\prime}}x\equiv a_{r^{\prime}}\mod lB_{i_{r^{\prime}}}

by the linear congruence

tnrxtarmodlBir.tn_{r^{\prime}}x\equiv ta_{r^{\prime}}\mod lB_{i_{r}}.

Hence we may assume that the index |lBir:lBir||lB_{i_{r}}:lB_{i_{r^{\prime}}}| is infinite whenever ir<iri_{r}<i_{r^{\prime}}.

For r0Rr_{0}\in R set R[r0]={rR:ir=ir0}R[r_{0}]=\{r\in R:i_{r}=i_{r_{0}}\} and consider the system 𝒮r0\mathcal{S}_{r_{0}}:

nrxarmodlBir,rR[r0].n_{r}x\equiv a_{r}\mod lB_{i_{r}},\quad r\in R[r_{0}].

By Lemma 4.21 the system 𝒮r0\mathcal{S}_{r_{0}} has the same number of solutions modulo lBir0lB_{i_{r_{0}}} in Zπ(M1)Z_{\pi}(M_{1}) and Zπ(M2)Z_{\pi}(M_{2}). If 𝒮r0\mathcal{S}_{r_{0}} has no solution, then 𝒮\mathcal{S} has no solution and we are done. Hence assume that 𝒮r0\mathcal{S}_{r_{0}} has a solution (and hence has the same number of solutions in both models by Lemma 4.21).

Then by Proposition 4.17 we can replace the system 𝒮r0\mathcal{S}_{r_{0}} by a single linear congruence without changing the solutions.

Hence we may assume

ir=irr=ri_{r}=i_{r^{\prime}}\iff r=r^{\prime}

for all r,rRr,r^{\prime}\in R. Now we may write R={r0,rm}R=\{r_{0},\dots r_{m}\} such that ir0>>irmi_{r_{0}}>\dots>i_{r_{m}}. We prove the lemma by induction on mm. The case m=0m=0 is done by Lemma 4.21. Hence we assume m>0m>0.

The system 𝒮\mathcal{S} has the form

nr0x\displaystyle n_{r_{0}}x ar0modlBir0,\displaystyle\equiv a_{r_{0}}\mod lB_{i_{r_{0}}},
\displaystyle\vdots
nrmx\displaystyle n_{r_{m}}x ar0modlBirm.\displaystyle\equiv a_{r_{0}}\mod lB_{i_{r_{m}}}.

Now set d:=gcd(nr0,p,αp>0|p:plBir0|)d:=\gcd(n_{r_{0}},\prod_{p\in\mathbb{P},\alpha_{p}>0}|\mathbb{Z}_{p}:\mathbb{Z}_{p}\cap lB_{i_{r_{0}}}|) and put

u=p|d,αp>0{|plBir1:plBir0|:|plBir1:plBir0| is finite}.u=\prod_{p|d,\alpha_{p}>0}\{|\mathbb{Z}_{p}\cap lB_{i_{r_{1}}}:\mathbb{Z}_{p}\cap lB_{i_{r_{0}}}|:|\mathbb{Z}_{p}\cap lB_{i_{r_{1}}}:\mathbb{Z}_{p}\cap lB_{i_{r_{0}}}|\text{ is finite}\}.

Set k={pd(p):αp>0 and k=\prod\{p^{d(p)}:\alpha_{p}>0\text{ and }p does not divide u}\text{ does not divide }u\} and consider the system 𝒮:\mathcal{S}^{\prime}:

nr0x\displaystyle n_{r_{0}}x ar0modkulBir1,\displaystyle\equiv a_{r_{0}}\mod kulB_{i_{r_{1}}},
unr1x\displaystyle un_{r_{1}}x uar1modulBir1,\displaystyle\equiv ua_{r_{1}}\mod ulB_{i_{r_{1}}},
unr2x\displaystyle un_{r_{2}}x uar2modulBir2,\displaystyle\equiv ua_{r_{2}}\mod ulB_{i_{r_{2}}},
\displaystyle\vdots
unrmx\displaystyle un_{r_{m}}x uarmmodulBirm.\displaystyle\equiv ua_{r_{m}}\mod ulB_{i_{r_{m}}}.

By Proposition 4.18 the linear congruences nr0xar0modlBir0n_{r_{0}}x\equiv a_{r_{0}}\mod lB_{i_{r_{0}}} and nr0xar0modkulBir1n_{r_{0}}x\equiv a_{r_{0}}\mod kulB_{i_{r_{1}}} have the same number of solutions modulo lBir0lB_{i_{r_{0}}} respectively kulBir1kulB_{i_{r_{1}}} and the sets of solutions agree modulo ulBir1ulB_{i_{r_{1}}}. The statement about the number of preimages in Proposition 4.18 implies that 𝒮\mathcal{S} and 𝒮\mathcal{S}^{\prime} have the same number of solutions modulo lBir0lB_{i_{r_{0}}} respectively kulBir1kulB_{i_{r_{1}}}. By Lemma 4.19 we can rewrite 𝒮\mathcal{S}^{\prime} as follows:

nr0x\displaystyle n_{r_{0}}x ar0modkulBir1,\displaystyle\equiv a_{r_{0}}\mod kulB_{i_{r_{1}}},
kunr1x\displaystyle kun_{r_{1}}x kuar1modkulBir1,\displaystyle\equiv kua_{r_{1}}\mod kulB_{i_{r_{1}}},
kunr2x\displaystyle kun_{r_{2}}x kuar2modkulBir2,\displaystyle\equiv kua_{r_{2}}\mod kulB_{i_{r_{2}}},
\displaystyle\vdots
kunrmx\displaystyle kun_{r_{m}}x kuarmmodkulBirm.\displaystyle\equiv kua_{r_{m}}\mod kulB_{i_{r_{m}}}.

By induction the system 𝒮\mathcal{S}^{\prime} has the same number of solutions modulo kulBir1kulB_{i_{r_{1}}} in Zπ(M1)Z_{\pi}(M_{1}) and Zπ(M2)Z_{\pi}(M_{2}). Hence 𝒮\mathcal{S} has the same number of solutions modulo lBir0lB_{i_{r_{0}}} in Zπ(M1)Z_{\pi}(M_{1}) and Zπ(M2)Z_{\pi}(M_{2}). ∎

To deal with the general case we will make use of the following:

Fact 4.23 (Inclusion-exclusion priciple).

Let A1,AnA_{1},\dots A_{n} be finite sets. Then

|i=1nAi|=J{1,n}(1)|J|+1|jJAj|.|\bigcup_{i=1}^{n}A_{i}|=\sum_{\emptyset\neq J\subseteq\{1,\dots n\}}(-1)^{|J|+1}|\bigcap_{j\in J}A_{j}|.
Proposition 4.24.

Let M1M_{1} and M2M_{2} be models of TT and let (A,J)(A,J) be a common substructure. Let π\pi be a finite set of primes and let 𝒮\mathcal{S} be a system

nrxarmodlBir,\displaystyle n_{r}x\equiv a_{r}\mod lB_{i_{r}},\quad rR,\displaystyle r\in R,
nsxasmodlBis,\displaystyle n_{s}x\not\equiv a_{s}\mod lB_{i_{s}},\quad sS,\displaystyle s\in S,

of linear congruences where each ntn_{t} is a π\pi-number, atZπ(A)a_{t}\in Z_{\pi}(A), itJi_{t}\in J for all tRSt\in R\cup S. Assume there is r0Rr_{0}\in R such that ir0i_{r_{0}} is maximal in {it:tRS}\{i_{t}:t\in R\cup S\}. Suppose moreover, that the index |Bit:Bit||B_{i_{t}}:B_{i_{t^{\prime}}}| is a π\pi-number whenever it is finite. Then 𝒮\mathcal{S} has the same number of solutions modulo lBir0lB_{i_{r_{0}}} in Zπ(M1)Z_{\pi}(M_{1}) and Zπ(M2)Z_{\pi}(M_{2}).

Proof.

For pairwise distinct s1,snSs_{1},\dots s_{n}\in S let As1,snA_{s_{1},\dots s_{n}} be the set of solutions modulo lBir0lB_{i_{r_{0}}} of the system 𝒮s1,sn\mathcal{S}_{s_{1},\dots s_{n}}:

nrx\displaystyle n_{r}x armodlBir,rR,\displaystyle\equiv a_{r}\mod lB_{i_{r}},\quad r\in R,
ns1x\displaystyle n_{s_{1}}x as1modlBis1,\displaystyle\equiv a_{s_{1}}\mod lB_{i_{s_{1}}},
\displaystyle\vdots
nsnx\displaystyle n_{s_{n}}x asnmodlBisn.\displaystyle\equiv a_{s_{n}}\mod lB_{i_{s_{n}}}.

By Lemma 4.22 the system 𝒮s1,sn\mathcal{S}_{s_{1},\dots s_{n}} has the same number of solutions in Zπ(M1)Z_{\pi}(M_{1}) and Zπ(M2)Z_{\pi}(M_{2}). In particular, this holds true for the system 𝒮\mathcal{S}_{\emptyset}:

nrxarmodlBir,rR.n_{r}x\equiv a_{r}\mod lB_{i_{r}},\quad r\in R.

Moreover, sSAs\bigcup_{s\in S}A_{s} is exactly the set of solutions modulo lBir0lB_{i_{r_{0}}} for 𝒮\mathcal{S}_{\emptyset} that do not solve 𝒮\mathcal{S}.

Note that As1Asn=As1,snA_{s_{1}}\cap\dots\cap A_{s_{n}}=A_{s_{1},\dots s_{n}} and hence by an application of the inclusion-exclusion principle the number |sSAs||\bigcup_{s\in S}A_{s}| (which is finite since we only count solutions modulo lBir0lB_{i_{r_{0}}}) does not depend on the model.

Now the system 𝒮\mathcal{S} is solved by exactly

|A||sSAs||A_{\emptyset}|-|\bigcup_{s\in S}A_{s}|

many solutions modulo lBir0lB_{i_{r_{0}}} and this number does not depend on the model. ∎

4.2.4 Proof of quantifier elimination

Proof of Theorem 4.8.

By Lemma 2.21 it suffices to show that every formula of the form

ψ(z¯,i¯)x𝒵rRφr(x,z¯,i¯)\psi(\bar{z},\bar{i})\equiv\exists x\in\mathcal{Z}\bigwedge_{r\in R}\varphi_{r}(x,\bar{z},\bar{i})

is equivalent to a quantifier free formula modulo TT where z¯𝒵\bar{z}\subseteq\mathcal{Z}, i¯\bar{i}\subseteq\mathcal{I} and each φr\varphi_{r} is either a basic Z\mathcal{L}_{Z}-formula or is of the form

vlr(tr(x,z¯))=irv^{l_{r}}(t_{r}(x,\bar{z}))=i_{r}

where trt_{r} is an Z\mathcal{L}_{Z}-term and iri_{r} is one variable in the tuple i¯\bar{i}.

Write R=R0R1R2R=R_{0}\cup R_{1}\cup R_{2} such that

φr(x,z¯,i¯)\displaystyle\varphi_{r}(x,\bar{z},\bar{i}) nrxtr(z¯)=0,for rR0,\displaystyle\equiv n_{r}x-t_{r}(\bar{z})=0,\quad\text{for }r\in R_{0},
φr(x,z¯,i¯)\displaystyle\varphi_{r}(x,\bar{z},\bar{i}) nrxtr(z¯)0,for rR1,and\displaystyle\equiv n_{r}x-t_{r}(\bar{z})\neq 0,\quad\text{for }r\in R_{1},\quad\text{and}
φr(x,z¯,i¯)\displaystyle\varphi_{r}(x,\bar{z},\bar{i}) vlr(nrxtr(z¯))=ir,for rR2.\displaystyle\equiv v^{l_{r}}(n_{r}x-t_{r}(\bar{z}))=i_{r},\quad\text{for }r\in R_{2}.

Now let π\pi be a finite set of primes such that nrn_{r}, lrl_{r}, and the cardinalities of all finite quotients |lBir:lBir||lB_{i_{r}}:lB_{i_{r^{\prime}}}| are π\pi-numbers. Fix two models M1,M2M_{1},M_{2} of TT and let (A,J)(A,J) be a common substructure, a¯A,η¯J\bar{a}\subseteq A,\bar{\eta}\subseteq J. Set ar:=tr(a¯)a_{r}:=t_{r}(\bar{a}). We have A=Zπ(A)×Aπ(A)A=Z_{\pi}(A)\times A_{\pi}(A) (since AπA_{\pi} is a finite set of constants) and hence each ara_{r} can be written as ar=arπbrπa_{r}=a_{r}^{\pi}b_{r}^{\pi} with arπZπ(A)a_{r}^{\pi}\in Z_{\pi}(A) and brπAπ(A)b_{r}^{\pi}\in A_{\pi}(A). Now suppose the formula ψ(a¯,η¯)\psi(\bar{a},\bar{\eta}) has a solution in M1M_{1}. By Proposition 2.20 it suffices to show that it has a solution in M2M_{2}.

If rR0r\in R_{0}, then φr\varphi_{r} must be satisfied in Zπ(M1)Z_{\pi}(M_{1}) and Aπ(M1)A_{\pi}(M_{1}). If rR1r\in R_{1}, then it suffices if φr\varphi_{r} is satisfied in Zπ(M1)Z_{\pi}(M_{1}) or Aπ(M1)A_{\pi}(M_{1}). If rR2r\in R_{2}, then we have

φr(x,a¯,η¯)vlr(nrxar)=ir.\varphi_{r}(x,\bar{a},\bar{\eta})\equiv v^{l_{r}}(n_{r}x-a_{r})=i_{r}.

This is satisfied if we have “=” in Zπ(M1)Z_{\pi}(M_{1}) or Aπ(M1)A_{\pi}(M_{1}) and “\geq” in the other subgroup. Hence there are subsets R1πR1R_{1}^{\pi}\subseteq R_{1} and R2πR2R_{2}^{\pi}\subseteq R_{2} such that the formulas

ψπxZπ\displaystyle\psi^{\pi}\equiv\exists x\in Z_{\pi} rR0nrxarπ=0\displaystyle\bigwedge_{r\in R_{0}}n_{r}x-a_{r}^{\pi}=0
rR1πnrxarπ0\displaystyle\land\bigwedge_{r\in R_{1}^{\pi}}n_{r}x-a_{r}^{\pi}\neq 0
rR2πvlr(nrxarπ)=ir\displaystyle\land\bigwedge_{r\in R_{2}^{\pi}}v^{l_{r}}(n_{r}x-a_{r}^{\pi})=i_{r}
rR2R2πvlr(nrxarπ)ir\displaystyle\land\bigwedge_{r\in R_{2}\setminus R_{2}^{\pi}}v^{l_{r}}(n_{r}x-a_{r}^{\pi})\geq i_{r}

and

ψπ¯xAπ\displaystyle\overline{\psi^{\pi}}\equiv\exists x\in A_{\pi} rR0nrxbrπ=0\displaystyle\bigwedge_{r\in R_{0}}n_{r}x-b_{r}^{\pi}=0
rR1R1πnrxbrπ0\displaystyle\land\bigwedge_{r\in R_{1}\setminus R_{1}^{\pi}}n_{r}x-b_{r}^{\pi}\neq 0
rR2R2πvlr(nrxbrπ)=ir\displaystyle\land\bigwedge_{r\in R_{2}\setminus R_{2}^{\pi}}v^{l_{r}}(n_{r}x-b_{r}^{\pi})=i_{r}
rR2πvlr(nrxbrπ)ir\displaystyle\land\bigwedge_{r\in R_{2}^{\pi}}v^{l_{r}}(n_{r}x-b_{r}^{\pi})\geq i_{r}

have a solution in Zπ(M1)Z_{\pi}(M_{1}) respectively Aπ(M1)A_{\pi}(M_{1}). Since AπA_{\pi} is a finite set of constants, this implies that ψπ¯\overline{\psi^{\pi}} has a solution in Aπ(M2)A_{\pi}(M_{2}). It remains to show that ψπ\psi^{\pi} has a solution in Zπ(M2)Z_{\pi}(M_{2}).

If R2=R_{2}=\emptyset, then we are done since the formulas xnZx\in nZ are quantifier free 0-definable and hence the result follows from the usual quantifier elimination for abelian groups. Therefore we assume R2R_{2}\neq\emptyset.

If R0R_{0}\neq\emptyset, say r0R0r_{0}\in R_{0}, then ar0π/nr0a_{r_{0}}^{\pi}/n_{r_{0}} is the solution of ψπ\psi^{\pi} in Zπ(M1)Z_{\pi}(M_{1}). Lemma 4.20 implies that ar0π/nr0a_{r_{0}}^{\pi}/n_{r_{0}} also solves ψπ\psi^{\pi} in Zπ(M2)Z_{\pi}(M_{2}). Hence we may assume R0=R_{0}=\emptyset.

If ir=+i_{r}=+\infty for some rr, then we have

vlr(nrxarπ)irnrxarπ=0.v^{l_{r}}(n_{r}x-a_{r}^{\pi})\geq i_{r}\iff n_{r}x-a_{r}^{\pi}=0.

Hence we may assume ir<+i_{r}<+\infty for all rR2r\in R_{2}.

Given l1l^{\prime}\geq 1 there is a finite set of constants ClC_{l^{\prime}} in the language such that the formula vl(t(x,z¯))=v^{l^{\prime}}(t(x,\bar{z}))=-\infty is equivalent to

cClvl(t(x,z¯)c)0.\bigvee_{c\in C_{l^{\prime}}}v^{l^{\prime}}(t(x,\bar{z})-c)\geq 0.

Thus we may also assume ir>i_{r}>-\infty for all rR2r\in R_{2}.

Note that each formula of the form nrxarπ0n_{r}x-a_{r}^{\pi}\neq 0 excludes only a single solution. Since we assume R2R_{2}\neq\emptyset and all formulas of the form

vlr(nrxarπ)=ir or vlr(nrxarπ)irv^{l_{r}}(n_{r}x-a_{r}^{\pi})=i_{r}\text{ or }v^{l_{r}}(n_{r}x-a_{r}^{\pi})\geq i_{r}

are solved by cosets of lrBir+1l_{r}B_{i_{r}+1}, we may moreover assume R1=R_{1}=\emptyset.

By Lemma 4.19 we have vlr(nrxarπ)=vmlr(mnrxmarπ)v^{l_{r}}(n_{r}x-a_{r}^{\pi})=v^{ml_{r}}(mn_{r}x-ma_{r}^{\pi}) for all π\pi-numbers mm. Thus we may use Lemma 4.19 to replace each lrl_{r^{\prime}} by l:=lcm(lr:rR2)l:=\mathrm{lcm}(l_{r}:r\in R_{2}).

We consider formulas as linear congruences:

vl(nrxarπ)=ir\displaystyle v^{l}(n_{r}x-a_{r}^{\pi})=i_{r} (nrxarπ0modlBirnrxarπ0modlBir+1),\displaystyle\iff(n_{r}x-a_{r}^{\pi}\equiv 0\mod lB_{i_{r}}\quad\land\quad n_{r}x-a_{r}^{\pi}\not\equiv 0\mod lB_{i_{r}+1}),
vl(nrxarπ)ir\displaystyle v^{l}(n_{r}x-a_{r}^{\pi})\geq i_{r} nrxarπ0modlBir.\displaystyle\iff n_{r}x-a_{r}^{\pi}\equiv 0\mod lB_{i_{r}}.

Hence it suffices to show that the system of linear congruences

nrxarπ\displaystyle n_{r}x-a_{r}^{\pi} 0modlBir,rR2,\displaystyle\equiv 0\mod lB_{i_{r}},\quad r\in R_{2},
nrxarπ\displaystyle n_{r}x-a_{r}^{\pi} 0modlBir+1,rR2π,\displaystyle\not\equiv 0\mod lB_{i_{r}+1},\quad r\in R_{2}^{\pi},

has a solution in Zπ(M2)Z_{\pi}(M_{2}). After slightly adjusting the system (by using Lemma 4.19) and renaming, we get a system

nsxbs\displaystyle n_{s}x-b_{s} 0modlBis,sS,\displaystyle\equiv 0\mod lB_{i_{s}},\quad s\in S,
ntxbt\displaystyle n_{t}x-b_{t} 0modlBit,tT,\displaystyle\not\equiv 0\mod lB_{i_{t}},\quad t\in T,

where SS\neq\emptyset and every index |Bir:Bir|,r,rST|B_{i_{r}}:B_{i_{r^{\prime}}}|,r,r^{\prime}\in S\cup T is infinite or trivial. If there is an element sSs\in S such that isi_{s} is maximal in {ir:rST}\{i_{r}:r\in S\cup T\}, then we are done by Proposition 4.24. Hence suppose there is t0Tt_{0}\in T such that it0>isi_{t_{0}}>i_{s} for all sSs\in S. Then |Bis:Bit0||B_{i_{s}}:B_{i_{t_{0}}}| is infinite for all sSs\in S. In particular, the congruence

nt0xbt00modlBit0n_{t_{0}}x-b_{t_{0}}\not\equiv 0\mod lB_{i_{t_{0}}}

can be ignored, since each lBislB_{i_{s}}-class consists of infinitely many lBt0lB_{t_{0}} classes. Hence we removed one linear congruence from the system. After iterating this, we can find sSs\in S such that isi_{s} is maximal. ∎

4.3 The monotone hull

Theorem 4.8 gives quantifier elimination up to a suitable language on \mathcal{I}. The following gives a tame expansion of \mathcal{L}_{\mathcal{I}}^{-} which allows us to analyze the definable sets.

A binary relation RR on a linear ordering is called monotone if and only if it satisfies

xxRyyimpliesxRy.x^{\prime}\leq xRy\leq y^{\prime}\quad\text{implies}\quad x^{\prime}Ry^{\prime}.

The following result by Simon states that expanding a linear ordering by monotone binary relations is tame:

Proposition 4.25 (Proposition 4.1 and Proposition 4.2 of [14]).

Let (I,,Rα,Cβ)α,β(I,\leq,R_{\alpha},C_{\beta})_{\alpha,\beta} be a linear order equipped with monotone binary relations and unary predicates such that every \emptyset-definable monotone binary relation is given by one of the RαR_{\alpha} and every \emptyset-definable unary predicate is given by one of the CβC_{\beta}. Then (I,,Rα,Cβ)α,β(I,\leq,R_{\alpha},C_{\beta})_{\alpha,\beta} has quantifier elimination and is dp-minimal.

Fix a theory TZT_{Z} as in the quantifier elimination statement and let MTZM\models T_{Z} be a model. Note that the definable relations \leq, Divqkπ,l\mathrm{Div}_{q^{k}}^{\pi,l}, and Indkπ,l\mathrm{Ind}_{k}^{\pi,l} are monotone.

Definition 4.26.

Let SS be a set of unary predicates and monotone binary relations on the value set of MM.

  1. (a)

    We define ,monS\mathcal{L}_{\mathcal{I},\text{mon}}^{S} to be the monotone hull of

    S:={Divqkπ,l,Indkπ,l}q,π,l,kS,\mathcal{L}_{\mathcal{I}}^{S}:=\mathcal{L}_{\mathcal{I}}^{-}\cup\{\mathrm{Div}_{q^{k}}^{\pi,l},\mathrm{Ind}_{k}^{\pi,l}\}_{q,\pi,l,k}\cup S,

    i.e. the expansion of S\mathcal{L}_{\mathcal{I}}^{S} by all 0-definable (in S\mathcal{L}_{\mathcal{I}}^{S}) unary relations and all 0-definable monotone binary relations on the value sort.

  2. (b)

    Set monS=𝒵v,monS\mathcal{L}_{\mathrm{mon}}^{S}=\mathcal{L}_{\mathcal{Z}}\cup\mathcal{L}_{v}\cup\mathcal{L}_{\mathcal{I},\mathrm{mon}}^{S} and define mon=mon\mathcal{L}_{\mathrm{mon}}=\mathcal{L}_{\mathrm{mon}}^{\emptyset}.

Note that mon\mathcal{L}_{\mathrm{mon}}\supseteq\mathcal{L}^{-} is an expansion by definitions.

Proposition 4.27.

Let SS be as in Definition 4.26. Then Th(M)\mathrm{Th}(M) admits quantifier elimination in the language monS\mathcal{L}_{\mathrm{mon}}^{S}.

Proof.

The successor function and its inverse are 0-definable. If R,monSR\in\mathcal{L}_{\mathcal{I},\text{mon}}^{S} is a monotone binary relation, then so is Rm,n(x,y)R(x+m,y+n)R_{m,n}(x,y)\iff R(x+m,y+n) for all m,nm,n\in\mathbb{Z}. The same holds true for 0-definable unary predicates. Therefore adding the successor function to the language does not add any new definable sets in \mathcal{I}. Hence Theorem 4.8 and Proposition 4.25 imply quantifier elimination in monS\mathcal{L}_{\mathrm{mon}}^{S}. ∎

4.4 Dp-minimality and distality

Let TT be a complete monS\mathcal{L}_{\mathrm{mon}}^{S}-theory as in Proposition 4.27.

Lemma 4.28.

Let (Z,I,v)T(Z,I,v)\models T be a sufficiently saturated model and let (aj)jJ1(a_{j})_{j\in J_{1}} and (bj)jJ2(b_{j})_{j\in J_{2}} be mutually indiscernible sequences in the group sort. Let γI\gamma\in I be a singleton. Then one of the sequences is indiscernible over γ\gamma.

Proof.

We may assume that both sequences are indexed by a dense linear order. Suppose (aj)jJ1(a_{j})_{j\in J_{1}} is not indiscernible over γ\gamma. By the quantifier elimination result this must be witnessed by a formula of the form

R(vl(t(x¯)),γ)R(v^{l}(t(\bar{x})),\gamma)

where tt is an 𝒵\mathcal{L}_{\mathcal{Z}}-term, RR is a monotone binary relation on II, and l1l\geq 1. Hence we can find tuples j0¯,j1¯J1\bar{j_{0}},\bar{j_{1}}\subseteq J_{1} of the same order type such that

R(vl(t(a¯j0¯)),c)and⊧̸R(vl(t(a¯j1¯)),c)\models R(v^{l}(t(\overline{a}_{\overline{j_{0}}})),c)\quad\text{and}\quad\not\models R(v^{l}(t(\overline{a}_{\overline{j_{1}}})),c)

where a¯ji¯=(aj)jji¯\overline{a}_{\overline{j_{i}}}=(a_{j})_{j\in\overline{j_{i}}} is the tuple corresponding to ji¯J1\overline{j_{i}}\subseteq J_{1}.

After replacing j0¯\bar{j_{0}} or j1¯\bar{j_{1}} if necessary, we may assume that j0¯\bar{j_{0}} and j1¯\bar{j_{1}} have disjoint convex hulls in J1J_{1}. We can extend to a sequence (ji¯)i<ω(\bar{j_{i}})_{i<\omega} such that (a¯ji¯)i<ω(\overline{a}_{\overline{j_{i}}})_{i<\omega} is an indiscernable sequence. Then

(vl(t(a¯ji¯)))i<ω(v^{l}(t(\overline{a}_{\overline{j_{i}}})))_{i<\omega}

is a non-constant indiscernible sequence in the value sort that is not indiscernible over γ\gamma.

By Proposition 4.25 the value sort is dp-minimal. Therefore (bj)jJ2(b_{j})_{j\in J_{2}} must be indiscernible over γ\gamma: Otherwise we could apply the above argument to the sequence (bj)jJ2(b_{j})_{j\in J_{2}} to get a second non-constant indiscernible sequence in the value sort which is not indiscernible over γ\gamma. Since these two sequences would be mutually indiscernible, this would contradict dp-minimality of the value sort. ∎

Theorem 4.29.

TT is dp-minimal.

Proof.

Let M=(Z,I,v)TM=(Z,I,v)\models T be a sufficiently saturated model and let J1J_{1} and J2J_{2} be mutually indiscernible sequences. We will assume that both of them are indexed by a dense linear order. Let zZz\in Z be a singleton. We aim to show that one of the sequences is indiscernible over zz.

Since II is essentially an imaginary sort, we may assume that the sequences J1J_{1} and J2J_{2} live in the sort ZZ. Note that equality on the value sort can be expressed using the monotone binary relation \leq. By the quantifier elimination result, the failure of indiscernibility must be witnessed by formulas of the following form:

  1. 1.

    t(x¯)nz=0t(\bar{x})-nz=0,

  2. 2.

    C(vl(t(x¯)nz))C(v^{l}(t(\bar{x})-nz)),

  3. 3.

    R(vl1(t1(x¯)n1z),vl2(t2(x¯)n2z))R(v^{l_{1}}(t_{1}(\bar{x})-n_{1}z),v^{l_{2}}(t_{2}(\bar{x})-n_{2}z)),

where tt is an 𝒵\mathcal{L}_{\mathcal{Z}}-term, CC is a coloring on \mathcal{I}, RR is a monotone binary relation on \mathcal{I}, l1l\geq 1, and nn\in\mathbb{Z}. One of the terms in the third case could also be a quantifier free 0-definable constant in the value sort. This case is analogous to case (b) below and therefore we will not consider it separately.

Note that a formula of the first type would imply that zz is algebraic over the parameters plugged in for x¯\bar{x}. Hence it suffices to consider the other two types of formulas. If an indiscernible sequence JJ is not indiscernible over zz, then this must be witnessed by a¯,a¯J\bar{a},\bar{a}^{\prime}\subseteq J of the same order type such that we are in one of the following cases:

  1. (a)

    We have

    vl1(t(a¯)nz)vl1(t(a¯)nz)andvl2(t(a¯)nz)vl2(t(a¯)nz)v^{l_{1}}(t(\bar{a})-nz)\neq v^{l_{1}}(t(\bar{a}^{\prime})-nz)\quad\text{and}\quad v^{l_{2}}(t^{\prime}(\bar{a})-n^{\prime}z)\neq v^{l_{2}}(t^{\prime}(\bar{a}^{\prime})-n^{\prime}z)

    and

    R(vl1(t(a¯)nz),vl2(t(a¯)nz))and⊧̸R(vl1(t(a¯)nz),vl2(t(a¯)nz))\models R(v^{l_{1}}(t(\bar{a})-nz),v^{l_{2}}(t^{\prime}(\bar{a})-n^{\prime}z))\quad\text{and}\quad\not\models R(v^{l_{1}}(t(\bar{a}^{\prime})-nz),v^{l_{2}}(t^{\prime}(\bar{a}^{\prime})-n^{\prime}z))

    for some choice of t,t,n0,n0,t,t^{\prime},n\neq 0,n^{\prime}\neq 0, and a relation RR.

  2. (b)

    We have

    vl1(t(a¯)nz)vl1(t(a¯)nz)v^{l_{1}}(t(\bar{a})-nz)\neq v^{l_{1}}(t(\bar{a}^{\prime})-nz)

    and

    R(vl1(t(a¯)nz),vl2(t(a¯)))and⊧̸R(vl1(t(a¯)nz),vl2(t(a¯)))\models R(v^{l_{1}}(t(\bar{a})-nz),v^{l_{2}}(t^{\prime}(\bar{a})))\quad\text{and}\quad\not\models R(v^{l_{1}}(t(\bar{a}^{\prime})-nz),v^{l_{2}}(t^{\prime}(\bar{a}^{\prime})))

    for some choice of t,t,n0,t,t^{\prime},n\neq 0, and a relation RR.

  3. (c)

    We have

    vl1(t(a¯)nz)=vl1(t(a¯)nz)andvl2(t(a¯))<vl2(t(a¯))v^{l_{1}}(t(\bar{a})-nz)=v^{l_{1}}(t(\bar{a}^{\prime})-nz)\quad\text{and}\quad v^{l_{2}}(t^{\prime}(\bar{a}))<v^{l_{2}}(t^{\prime}(\bar{a}^{\prime}))

    and

    ⊧̸R(vl1(t(a¯)nz),vl2(t(a¯)))andR(vl1(t(a¯)nz),vl2(t(a¯)))\not\models R(v^{l_{1}}(t(\bar{a})-nz),v^{l_{2}}(t^{\prime}(\bar{a})))\quad\text{and}\quad\models R(v^{l_{1}}(t(\bar{a}^{\prime})-nz),v^{l_{2}}(t^{\prime}(\bar{a}^{\prime})))

    or

    R(vl2(t(a¯),vl1(t(a¯)nz)))and⊧̸R(vl2(t(a¯)),vl1(t(a¯)nz))\models R(v^{l_{2}}(t^{\prime}(\bar{a}),v^{l_{1}}(t(\bar{a})-nz)))\quad\text{and}\quad\not\models R(v^{l_{2}}(t^{\prime}(\bar{a}^{\prime})),v^{l_{1}}(t(\bar{a}^{\prime})-nz))

    for some choice of t,t,n0,t,t^{\prime},n\neq 0, and a monotone binary relation RR.

  4. (d)

    We have

    vl1(t(a¯)nz)=vl1(t(a¯)nz)andvl2(t(a¯)nz)<vl2(t(a¯)nz)v^{l_{1}}(t(\bar{a})-nz)=v^{l_{1}}(t(\bar{a}^{\prime})-nz)\quad\text{and}\quad v^{l_{2}}(t^{\prime}(\bar{a})-n^{\prime}z)<v^{l_{2}}(t^{\prime}(\bar{a}^{\prime})-n^{\prime}z)

    and

    ⊧̸R(vl1(t(a¯)nz),vl2(t(a¯)nz))andR(vl1(t(a¯)nz),vl2(t(a¯)nz))\not\models R(v^{l_{1}}(t(\bar{a})-nz),v^{l_{2}}(t^{\prime}(\bar{a})-n^{\prime}z))\quad\text{and}\quad\models R(v^{l_{1}}(t(\bar{a}^{\prime})-nz),v^{l_{2}}(t^{\prime}(\bar{a}^{\prime})-n^{\prime}z))

    or

    R(vl2(t(a¯)nz),vl1(t(a¯)nz)))and⊧̸R(vl2(t(a¯)nz),vl1(t(a¯)nz))\models R(v^{l_{2}}(t^{\prime}(\bar{a})-n^{\prime}z),v^{l_{1}}(t(\bar{a})-nz)))\quad\text{and}\quad\not\models R(v^{l_{2}}(t^{\prime}(\bar{a}^{\prime})-n^{\prime}z),v^{l_{1}}(t(\bar{a}^{\prime})-nz))

    for some choice of t,t,n0,n0,t,t^{\prime},n\neq 0,n^{\prime}\neq 0, and a monotone binary relation RR.

The case corresponding to a coloring is essentially the same as (b) so we will not do it explicitly.

We will use Lemma 4.19 to assume that all the lil_{i} coincide: Let π\pi be a finite set of primes. We want to be able to work in Zπ(M)Z_{\pi}(M). Fix a term

vl(t(a¯)nz)v^{l}(t(\bar{a})-nz)

and write t(a¯)=b0(a¯)+b1(a¯)t(\bar{a})=b_{0}(\bar{a})+b_{1}(\bar{a}), z=c0+c1z=c_{0}+c_{1} for b0(a¯),c0Zπ(M)b_{0}(\bar{a}),c_{0}\in Z_{\pi}(M), b1(a¯),c1Aπ(M)b_{1}(\bar{a}),c_{1}\in A_{\pi}(M). Since Aπ(M)A_{\pi}(M) is a finite set of constants, the value of b1(a¯)b_{1}(\bar{a}) only depends on the order type of a¯\bar{a}. Therefore

γ=vl(b1(a¯)+nc1)Aπ(M)\gamma=v^{l}(b_{1}(\bar{a})+nc_{1})\in A_{\pi}(M)

also only depends on the order type of a¯\bar{a}. We have

vl(t(x¯)nz)=min{vl(b0(x¯)nc0),γ}v^{l}(t(\bar{x})-nz)=\min\{v^{l}(b_{0}(\bar{x})-nc_{0}),\gamma\}

because Z=Zπ(M)×Aπ(M)Z=Z_{\pi}(M)\times A_{\pi}(M). If vl(t(a¯)nz)=γv^{l}(t(\bar{a}^{\prime})-nz)=\gamma for all a¯\bar{a}^{\prime} of the same order type as a¯\bar{a}, then this value is a constant. If vl(t(a¯)nz)=vl(b0(a¯)nc0)v^{l}(t(\bar{a}^{\prime})-nz)=v^{l}(b_{0}(\bar{a}^{\prime})-nc_{0}) for all a¯\bar{a}^{\prime} of the same order type as a¯\bar{a}, then this value can always be calculated in Zπ(M)Z_{\pi}(M). If we are not in one of these two cases, then the quantifier free 0-definable coloring

C<γ(i)i<γC_{<\gamma}(i)\iff i<\gamma

witnesses (in Zπ(M)Z_{\pi}(M)) that JJ is not indiscernible over zz. Hence we can work in Zπ(M)Z_{\pi}(M) and therefore we can assume that all the lil_{i} coincide (by Lemma 4.19). Moreover, to simplify the notation we will assume that all the lil_{i} are equal to 11.

We say that an indiscernible sequence JJ has an approximation for zz over αI\alpha\in I if there is a set DD such that JJ is indiscernible over DD, α\alpha is definable over DD, and the residue class of zz modulo BαB_{\alpha} is algebraic (in Z/BαZ/B_{\alpha}) over parameters in DD.

We now assume that the mutually indiscernible sequences J1J_{1} and J2J_{2} both fail to be indiscernible over zz. Then this must be witnessed as in (a) to (d). Such a witness for J1J_{1} (resp. J2J_{2}) is good if J2J_{2} (resp. J1J_{1}) has an approximation for zz for a suitable α\alpha defined as follows:

  • If the witness is given as in (a), then we set

    α=max{v(t(a¯)t(a¯)),v(t(a¯)t(a¯))}+1.\alpha=\max\{v(t(\bar{a})-t(\bar{a}^{\prime})),v(t^{\prime}(\bar{a})-t^{\prime}(\bar{a}^{\prime}))\}+1.

    If (for example) α=v(t(a¯)t(a¯))+1\alpha=v(t(\bar{a})-t(\bar{a}^{\prime}))+1 (=v(t(a¯)nz)+1<v(t(a¯)nz)+1=v(t(\bar{a})-nz)+1<v(t(\bar{a}^{\prime})-nz)+1), then t(a¯)nzmodBαt(\bar{a^{\prime}})\equiv n^{\prime}z\mod B_{\alpha}. Therefore the residue class of zz modulo BαB_{\alpha} is algebraic over t(a¯)t(\bar{a}^{\prime}).

  • If the witness is given as in (b), then we set

    α=v(t(a¯)t(a¯))+1=min{v(t(a¯)nz),v(t(a¯)nz)}+1.\alpha=v(t(\bar{a})-t(\bar{a}^{\prime}))+1=\min\{v(t(\bar{a})-nz),v(t(\bar{a}^{\prime})-nz)\}+1.

    If v(t(a¯)nz)<v(t(a¯)nz)v(t(\bar{a})-nz)<v(t(\bar{a}^{\prime})-nz), then t(a¯)nzmodBαt(\bar{a^{\prime}})\equiv n^{\prime}z\mod B_{\alpha} and therefore the residue class of zz modulo BαB_{\alpha} is algebraic over t(a¯)t(\bar{a}^{\prime}).

  • If the witness is given as in (c), we set

    α=v(t(a¯)nz).\alpha=v(t(\bar{a})-nz).
  • Now assume the witness is given as in (d). We set

    α1=v(t(a¯)nz)andα2=v(t(a¯)nz)+1.\alpha_{1}=v(t(\bar{a})-nz)\quad\text{and}\quad\alpha_{2}=v(t^{\prime}(\bar{a})-n^{\prime}z)+1.

    Now put α=max{α1,α2}\alpha=\max\{\alpha_{1},\alpha_{2}\}.

In particular, every witness of type (a) or (b) is good because J1J_{1} and J2J_{2} are mutually indiscernible. Recall that if v(x)<v(y)v(x)<v(y), then v(xy)=v(x)v(x-y)=v(x). We aim to show that we can always find a good witness:

Suppose the witness is given as in (c). Choose a¯0J1\bar{a}_{0}\subseteq J_{1} of the same order type as a¯\bar{a} and a¯\bar{a}^{\prime} such that all indices involved in a¯0\bar{a}_{0} are smaller than the indices in a¯\bar{a} and a¯\bar{a}^{\prime} (from now on, we will write a0¯a¯,a¯\bar{a_{0}}\ll\bar{a},\bar{a}^{\prime} in that case). If v(t(a¯0)nz)v(t(a¯)nz)v(t(\bar{a}_{0})-nz)\neq v(t(\bar{a})-nz), then either the pair (a0¯,a¯)(\bar{a_{0}},\bar{a}) or the pair (a¯0,a¯)(\bar{a}_{0},\bar{a}^{\prime}) gives a good witness as in case (b).

Hence we will assume v(t(a¯0)nz)=v(t(a¯)nz)v(t(\bar{a}_{0})-nz)=v(t(\bar{a})-nz). Let J1>a¯0J_{1}^{>\bar{a}_{0}} be the sequence consisting of all elements of J1J_{1} with index larger than all indices in a¯0\bar{a}_{0} and set J2a¯0J_{2}\cup\bar{a}_{0} to be the sequence J2J_{2} where each tuple is expanded by a¯0\bar{a}_{0}. Then J1>a¯0J_{1}^{>\bar{a}_{0}} and J2a¯0J_{2}\cup\bar{a}_{0} are mutually indiscernible. Moreover, J1>a¯0J_{1}^{>\bar{a}_{0}} is not indiscernible over α=v(t(a¯)nz)\alpha=v(t(\bar{a})-nz) (as witnessed by a¯\bar{a} and a¯\bar{a}^{\prime}). Hence J2a¯0J_{2}\cup\bar{a}_{0} is indiscernible over α\alpha by Lemma 4.28. Now J2J_{2} is indiscernible over the set {a¯0,α}\{\bar{a}_{0},\alpha\} and we have a¯0nzmodBα\bar{a}_{0}\equiv nz\mod B_{\alpha}. Therefore the residue class of zz modulo BαB_{\alpha} is algebraic over a¯0\bar{a}_{0} and hence the witness is good.

Now suppose the witness is given as in (d). We set

α1=v(t(a¯)nz)andα2=v(t(a¯)nz)+1.\alpha_{1}=v(t(\bar{a})-nz)\quad\text{and}\quad\alpha_{2}=v(t^{\prime}(\bar{a})-n^{\prime}z)+1.

If α2α1\alpha_{2}\geq\alpha_{1}, then we have a good witness by the same arguments as in (a) and (b). Hence assume α:=α1>α2\alpha:=\alpha_{1}>\alpha_{2}. Suppose for all a¯0a¯,a¯\bar{a}_{0}\ll\bar{a},\bar{a}^{\prime} we have v(t(a¯0)nz)=αv(t(\bar{a}_{0})-nz)=\alpha. Fix

a¯0a¯1a¯,a¯.\bar{a}_{0}\ll\bar{a}_{1}\ll\bar{a},\bar{a}^{\prime}.

Consider the mutually indiscernible sequences J1>a¯0J_{1}^{>\bar{a}_{0}} and J2a0¯J_{2}\cup\bar{a_{0}}.

Assume that J1>a¯0J_{1}^{>\bar{a}_{0}} is indiscernible over α\alpha. Then the residue class of zz modulo BαB_{\alpha} is algebraic over t(a¯1)t(\bar{a}_{1}). Since t(a¯)nzmodBαt^{\prime}(\bar{a})\not\equiv n^{\prime}z\mod B_{\alpha}, we get t(a¯)nzmodBαt^{\prime}(\bar{a}^{\prime})\not\equiv n^{\prime}z\mod B_{\alpha} by indiscernibility (applied to α\alpha and a¯1\bar{a}_{1}). Therefore v(t(a¯)nz)v(t^{\prime}(\bar{a})-n^{\prime}z) and v(t(a¯)nz)v(t^{\prime}(\bar{a}^{\prime})-n^{\prime}z) only depend on the residue class of zz modulo BαB_{\alpha} (and can be calculated in Z/BαZ/B_{\alpha}) and hence cannot witness the failure of indiscernibility over zz.

Hence J1>a¯0J_{1}^{>\bar{a}_{0}} is not indiscernible over α\alpha. Then J2a¯0J_{2}\cup\bar{a}_{0} is indiscernible over α\alpha by Lemma 4.28. Therefore J2J_{2} is indiscernible over {a¯0,α}\{\bar{a}_{0},\alpha\} and the residue class of zz modulo BαB_{\alpha} is algebraic over a¯0\bar{a}_{0}. Hence we have a good witness.

Hence we assume that there is a¯0a¯,a¯\bar{a}_{0}\ll\bar{a},\bar{a}^{\prime} such that

v(t(a¯0)nz)α.v(t(\bar{a}_{0})-nz)\neq\alpha.

If v(t(a¯0)nz)>αv(t(\bar{a}_{0})-nz)>\alpha, then α=v(t(a¯0)t(a¯))\alpha=v(t(\bar{a}_{0})-t(\bar{a})) and we have a good witness as in cases (a) and (b). Hence we assume v(t(a¯0)nz)<αv(t(\bar{a}_{0})-nz)<\alpha.

If v(t(a¯0)nz){v(t(a¯)nz),v(t(a¯)nz)}v(t^{\prime}(\bar{a}_{0})-n^{\prime}z)\not\in\{v(t^{\prime}(\bar{a})-n^{\prime}z),v(t^{\prime}(\bar{a}^{\prime})-n^{\prime}z)\}, then (a¯0,a¯)(\bar{a}_{0},\bar{a}) or (a¯0,a¯)(\bar{a}_{0},\bar{a}^{\prime}) gives a good witness as in case (a). If v(t(a¯0)nz)=v(t(a¯)nz)v(t^{\prime}(\bar{a}_{0})-n^{\prime}z)=v(t^{\prime}(\bar{a})-n^{\prime}z), then either (a¯0,a¯)(\bar{a}_{0},\bar{a}^{\prime}) gives a witness as in case (a) or the new witness is given by (a¯0,a¯)(\bar{a}_{0},\bar{a}) and we have

v(t(a¯)nz)>v(t(a¯0)nz)andv(ta¯)nz)=v(t(a¯0)nz).v(t(\bar{a})-nz)>v(t(\bar{a}_{0})-nz)\quad\text{and}\quad v(t^{\prime}\bar{a})-n^{\prime}z)=v(t^{\prime}(\bar{a}_{0})-n^{\prime}z).

Hence we are again in case (d) but J2J_{2} is indiscernible over

v(t(a¯)nz)=v(t(a¯)t(a¯))v(t^{\prime}(\bar{a})-n^{\prime}z)=v(t^{\prime}(\bar{a})-t^{\prime}(\bar{a}^{\prime}))

and hence this witness given by (a¯0,a¯)(\bar{a}_{0},\bar{a}) must be good.

Now only the case v(t(a¯0)nz)=v(t(a¯)nz)v(t^{\prime}(\bar{a}_{0})-n^{\prime}z)=v(t^{\prime}(\bar{a}^{\prime})-n^{\prime}z) is left. We then have

v(t(a¯)nz)=v(t(a¯)nz)>v(t(a¯0)nz),v(t(\bar{a})-nz)=v(t(\bar{a}^{\prime})-nz)>v(t(\bar{a}_{0})-nz),
v(t(a¯)nz)<v(t(a¯)nz)=v(t(a0¯)nz).v(t^{\prime}(\bar{a})-n^{\prime}z)<v(t^{\prime}(\bar{a}^{\prime})-n^{\prime}z)=v(t^{\prime}(\bar{a_{0}})-n^{\prime}z).

Assume the witnessing formula was of the form

R(v(t(x¯)nz),v(t(x¯)nz))R(v(t(\bar{x})-nz),v(t^{\prime}(\bar{x})-n^{\prime}z))

for a monotone binary relation RR (the other case is done analogously).

We then have the following implications by monotonicity:

R(v(t(a¯)nz),v(t(a¯)nz))\displaystyle\models R(v(t(\bar{a})-nz),v(t^{\prime}(\bar{a})-n^{\prime}z))
\displaystyle\implies R(v(t(a¯)nz),v(t(a¯)nz))\displaystyle\models R(v(t(\bar{a}^{\prime})-nz),v(t^{\prime}(\bar{a}^{\prime})-n^{\prime}z))
\displaystyle\implies R(v(t(a¯0)nz),v(t(a¯0)nz)).\displaystyle\models R(v(t(\bar{a}_{0})-nz),v(t^{\prime}(\bar{a}_{0})-n^{\prime}z)).

Hence R(v(t(a¯)nz),v(t(a¯)nz))R(v(t(\bar{a})-nz),v(t^{\prime}(\bar{a})-n^{\prime}z)) must be false and R(v(t(a¯)nz),v(t(a¯)nz))R(v(t(\bar{a}^{\prime})-nz),v(t^{\prime}(\bar{a}^{\prime})-n^{\prime}z)) must be true (since this was a witness for the failure of indiscernibility over zz). Then R(v(t(a¯0)nz),v(t(a¯0)nz))R(v(t(\bar{a}_{0})-nz),v(t^{\prime}(\bar{a}_{0})-n^{\prime}z)) must be true. But then a¯\bar{a} and a¯0\bar{a}_{0} give a witness as in (a). Hence we can always find a good witness.

Since we assume that both J1J_{1} and J2J_{2} fail to be indiscernible over zz, we can find a good witness for each of them. Let α\alpha be the constant for the witness in J1J_{1} and let β\beta be the constant for the witness in J2J_{2}. We assume αβ\alpha\leq\beta. Then J1J_{1} is indiscernible over β\beta and over the residue class of zz in Z/BβZ/B_{\beta}.

Suppose the witness for J1J_{1} is given as in (a) or (b). If we have

v(t(a¯)nz)<v(t(a¯)nz),v(t(\bar{a})-nz)<v(t(\bar{a}^{\prime})-nz),

then v(t(a¯)nz)<βv(t(\bar{a})-nz)<\beta and indiscernibility (and algebraicity of zz modulo BβB_{\beta} over a suitable parameter) imply that v(t(a¯)nz)<βv(t(\bar{a}^{\prime})-nz)<\beta. Hence those values only depend on the residue class of zz modulo BβB_{\beta} (and can be calculated in Z/BβZ/B_{\beta} with the restricted valuation). Therefore they cannot witness the failure of indiscernibility over zz.

Now suppose the witness for J1J_{1} is given as in case (c). If α=β\alpha=\beta, then this cannot be a witness for the failure for indiscernibility. Hence we must have α<β\alpha<\beta. But then

v(t(a¯)nz)=v(t(a¯)nz)v(t(\bar{a})-nz)=v(t(\bar{a}^{\prime})-nz)

only depends on the residue class of zz in BβB_{\beta} and we can argue as before. The same arguments work if the witness for J1J_{1} is given as in case (d).

Hence J1J_{1} or J2J_{2} must be indiscernible over zz. ∎

To characterize distality we will show that the quotients Bi/Bi+1B_{i}/B_{i+1} are stable. We will make use of the following lemma:

Lemma 4.30 (Lemma 5.13 of [2]).

Let 0\mathcal{L}_{0} be any language and let T0T_{0} be an unstable 0\mathcal{L}_{0}-theory. Let 00\mathcal{L}_{0}^{-}\subseteq\mathcal{L}_{0} be such that T0|0T_{0}|_{\mathcal{L}_{0}^{-}} is stable. Then there exists an 0\mathcal{L}_{0}-formula φ(x,y)\varphi(x,y), |x|=1|x|=1, over \emptyset and a parameter bb such that φ(x,b)\varphi(x,b) is not 0\mathcal{L}_{0}^{-}-definable.

Proposition 4.31.

Suppose Bi/Bi+1B_{i}/B_{i+1} is infinite. Then the induced structure on Bi/Bi+1B_{i}/B_{i+1} is stable.

Proof.

Suppose Bi/Bi+1B_{i}/B_{i+1} is infinite. By Lemma 4.30 it suffices to show that for every formula φ~(x~,y~)\tilde{\varphi}(\tilde{x},\tilde{y}) (in Bi/Bi+1B_{i}/B_{i+1}) and every constant b~Bi/Bi+1\tilde{b}\in B_{i}/B_{i+1} the formula φ~(x~,b~)\tilde{\varphi}(\tilde{x},\tilde{b}) is definable in the pure group (Bi/Bi+1,+)(B_{i}/B_{i+1},+).

Given such a formula there is an monS\mathcal{L}_{\text{mon}}^{S}-formula φ(x,y)\varphi(x,y) such that φ\varphi is the preimage of φ~\tilde{\varphi} under the natural projection

πi:BiBi/Bi+1.\pi_{i}:B_{i}\rightarrow B_{i}/B_{i+1}.

Now fix a preimage bb of b~\tilde{b}. Note that φ(Bi,b)\varphi(B_{i},b) is a union of cosets of Bi+1B_{i+1}.

By the quantifier elimination result φ\varphi is equivalent to a boolean combination of atomic monS\mathcal{L}_{\text{mon}}^{S}-formulas. We aim to show the following:

Claim.

There is a formula ψ(x,y)\psi(x,y) which is defined in the pure abelian group (Bi,+)(B_{i},+) such that φ(Bi,b)\varphi(B_{i},b) and ψ(Bi,b)\psi(B_{i},b) coincide on all but finitely many cosets of Bi+1B_{i+1}.

It suffices to prove the claim for atomic formulas. Therefore we may assume that φ\varphi is atomic. Then we are in one of the following cases:

  1. (a)

    φ(x,b)nxt(b)=0\varphi(x,b)\equiv nx-t(b)=0,

  2. (b)

    φ(x,b)R(vl1(n1xt1(b)),vl2(n2xt2(b)))\varphi(x,b)\equiv R(v^{l_{1}}(n_{1}x-t_{1}(b)),v^{l_{2}}(n_{2}x-t_{2}(b))),

  3. (c)

    φ(x,b)C(vl(nxt(b)))\varphi(x,b)\equiv C(v^{l}(nx-t(b))).

In case (a) there is nothing to show. Therefore we consider the cases (b) and (c) which include valuations. We show that sets of the form

(a+lBj)Bi(a+lB_{j})\cap B_{i}

are definable in the pure abelian group (Bi,+)(B_{i},+) up to a unique coset of Bi+1B_{i+1}:

If j<ij<i, then Bj>BiB_{j}>B_{i} and lBjlB_{j} has finite index in BjB_{j}. Hence lBjBilB_{j}\cap B_{i} has finite index in BiB_{i}. Moreover, lBjBilB_{j}\cap B_{i} is of the form

lBjBi=lBilB_{j}\cap B_{i}=l^{\prime}B_{i}

for a positive integer ll^{\prime} because this holds true for the standard models. The cosets of lBil^{\prime}B_{i} are definable in the pure group language. If j=ij=i, then lBiBi=lBilB_{i}\cap B_{i}=lB_{i} is definable in the pure group language. Now assume j>ij>i. Then lBj<BjBi+1lB_{j}<B_{j}\leq B_{i+1} and therefore Bi(a+lBj)B_{i}\cap(a+lB_{j}) is trivial outside of a single coset of Bi+1B_{i+1}.

This also shows that there are only finitely many intersections of the form

(a+lBj)(Bi(a+Bi+1))(a+lB_{j})\cap(B_{i}\setminus(a+B_{i+1}))

where everything except jj is fixed. Therefore the restriction of vl(xa)v^{l}(x-a) to Bi(a+Bi+1)B_{i}\setminus(a+B_{i+1}) is given by a finite chain of definable subgroups (in the pure abelian group (Bi,+)(B_{i},+)).

Since nx0modBi+1nx\equiv 0\mod B_{i+1} has only finitely many solutions modulo Bi+1B_{i+1} the same holds true for the valuation vl(nxa)v^{l}(nx-a) restricted to BiB_{i}: Outside of finitely many cosets of Bi+1B_{i+1} it is given by a finite chain of (Bi,+)(B_{i},+)-definable subgroups. In that sense vl(nxa)v^{l}(nx-a) is (Bi,+)(B_{i},+)-definable outside of finitely many cosets of Bi+1B_{i+1}.

Therefore the formula φ\varphi in (a) or (b) is definable in (Bi,+)(B_{i},+) outside of finitely many cosets of Bi+1B_{i+1}. This shows the claim.

Hence we can find such a formula ψ(x,y)\psi(x,y) defined in the pure abelian group (Bi,+)(B_{i},+) such that φ(Bi,b)\varphi(B_{i},b) and ψ(Bi,b)\psi(B_{i},b) coincide on all but finitely many cosets of Bi+1B_{i+1}. The usual quantifier elimination result for abelian groups shows that ψ(B1,b)\psi(B_{1},b) is a boolean combination of cosets of the trivial subgroup and groups of the form lBilB_{i} for l1l\geq 1. Each subgroup lBilB_{i} has finite index in BiB_{i} and the family {lBi:l1}\{lB_{i}:l\geq 1\} is closed under finite intersections. Hence a boolean combination of such groups is a union of finitely many cosets of lBilB_{i} for a suitable ll.

Since φ(Bi,b)\varphi(B_{i},b) and ψ(Bi,b)\psi(B_{i},b) agree on all but finitely many cosets of Bi+1B_{i+1} and φ(Bi,b)\varphi(B_{i},b) is a union of cosets of Bi+1B_{i+1}, the same must be true for φ~(Bi/Bi+1,b~)\tilde{\varphi}(B_{i}/B_{i+1},\tilde{b}), i.e. φ~(Bi/Bi+1,b~)\tilde{\varphi}(B_{i}/B_{i+1},\tilde{b}) is a boolean combination of cosets of (Bi/Bi+1,+)(B_{i}/B_{i+1},+)-definable subgroups. Therefore φ~(Bi/Bi+1,b~)\tilde{\varphi}(B_{i}/B_{i+1},\tilde{b}) is definable in (Bi/Bi+1,+)(B_{i}/B_{i+1},+). ∎

Theorem 4.32.

TT is distal if and only if there is a constant k<ωk<\omega such that

|Bi/Bi+1|k|B_{i}/B_{i+1}|\leq k

holds for all i<i<\infty.

Proof.

Suppose |Bi/Bi+1||B_{i}/B_{i+1}| is unbounded. Then there is some i0i_{0} such that |Bi0/Bi0+1||B_{i_{0}}/B_{i_{0}+1}| is infinite. By Proposition 4.31 the induced structure on Bi0/Bi0+1B_{i_{0}}/B_{i_{0}+1} is stable. Hence it follows from Proposition 2.19 that TT is not distal.

Now let XX be a non-constant totally indiscernible set of singletons and fix x,yXx,y\in X. Put i0=v(x1y)i_{0}=v(x^{-1}y). If xyx\neq y, then i0<i_{0}<\infty and hence

xBi0=yBi0 xBi0+1xBi0+1.xB_{i_{0}}=yB_{i_{0}}\quad\text{\quad}xB_{i_{0}+1}\neq xB_{i_{0}+1}.

It follows easily from total indiscernibility that i0i_{0} does not depend on the choice of xyx\neq y. Hence |X||Bi0/Bi0+1||X|\leq|B_{i_{0}}/B_{i_{0}+1}|. ∎

5 Valuations on the integers

The most well-known example of a dp-minimal expansion of (,+)(\mathbb{Z},+) is (,+,)(\mathbb{Z},+,\leq). Based on work by Palacín and Sklinos [12], Conant and Pillay [5] proved the remarkable result that (,+,0)(\mathbb{Z},+,0) has no proper stable expansions of finite dp-rank. Hence any proper dp-minimal expansion must be unstable. The other known examples of dp-minimal expansions are:

  • (,+,vp)(\mathbb{Z},+,v_{p}) where vpv_{p} is the pp-adic valuation on \mathbb{Z}. This was shown by Alouf and d’Elbée in [2].

  • (,+,C)(\mathbb{Z},+,C) where CC is cyclic order. These were found by Tran and Walsberg in [21].

  • Proper dp-minimal expansions of (,+,S)(\mathbb{Z},+,S), where SS is a dense cyclic order, and (,+,vp)(\mathbb{Z},+,v_{p}) were very recently found by Walsberg in [22].

An overview about the current research on dp-minimal expansions of (,+)(\mathbb{Z},+) is given by Walsberg in Section 6 of [22].

5.1 A single valuation

We add the following family of examples which generalize the pp-adic examples by Alouf and d’Elbée:

Theorem 5.1.

Let (Bi)i<ω(B_{i})_{i<\omega} be a strictly descending chain of subgroups of \mathbb{Z}, B0=B_{0}=\mathbb{Z}, let v:ω{}v:\mathbb{Z}\rightarrow\omega\cup\{\infty\} be the valuation defined by

v(x)=max{i:xBi},v(x)=\max\{i:x\in B_{i}\},

and let SS be a set of unary predicates and monotone binary relations on the value set. Then (,0,1,+,v,S)(\mathbb{Z},0,1,+,v,S) admits quantifier elimination in the language monS\mathcal{L}_{\mathrm{mon}}^{S} (with 0 and 11 as constants) and is dp-minimal. Moreover, (,0,1,+,v,S)(\mathbb{Z},0,1,+,v,S) is distal if and only if the size of the quotients Bi/Bi+1B_{i}/B_{i+1} is bounded.

Proof.

Note that any infinite strictly descending chain of subgroups of \mathbb{Z} must have trivial intersection. Moreover, every non-trivial subgroup of \mathbb{Z} is of the form nn\mathbb{Z} for some n1n\geq 1 and hence vv is a good valuation in the sense of Definition 4.1.

Moreover, ^=pp\mathbb{Z}\prec\hat{\mathbb{Z}}=\prod_{p}\mathbb{Z}_{p} and 1=\langle 1\rangle=\mathbb{Z} is dense. Hence Proposition 4.27 implies the quantifier elimination result. Dp-minimality follows by Theorem 4.29 and the claim about distality follows by Theorem 4.32. ∎

In case of the pp-adic valuation Alouf and d’Elbée proved in Theorem 1.1 of [2] that (,+,vp)(\mathbb{Z},+,v_{p}) has quantifier elimination in the language pE={+,,0,1,|p,Dn)n1\mathcal{L}_{p}^{E}=\{+,-,0,1,|_{p},D_{n})_{n\geq 1} where

x|pyvp(x)vp(y)andDn=n.x|_{p}y\iff v_{p}(x)\leq v_{p}(y)\quad\text{and}\quad D_{n}=n\mathbb{Z}.

Conant [4] showed that the structure (,+,0,1,)(\mathbb{Z},+,0,1,\leq) is a minimal proper expansion of (,+,0,1)(\mathbb{Z},+,0,1), i.e. there is no proper intermediate expansion. Alouf and d’Elbée proved the same for (,+,0,1,vp)(\mathbb{Z},+,0,1,v_{p}). We will show that this does not hold true for arbitrary valuations.

Proposition 5.2.

Fix distinct primes p0,p1,qp_{0},p_{1},q\in\mathbb{P} and put s=p0p1qs=p_{0}p_{1}q. For i<ωi<\omega fix σiSym({0,1})\sigma_{i}\in\mathrm{Sym}(\{0,1\}), set n0=1n_{0}=1, and recursively define

n3l+m={n3l1q iff m=0,n3lpσl(0) iff m=1,n3l+1pσl(1) iff m=2.n_{3l+m}=\begin{cases}n_{3l-1}q&\text{ iff }m=0,\\ n_{3l}p_{\sigma_{l}(0)}&\text{ iff }m=1,\\ n_{3l+1}p_{\sigma_{l}(1)}&\text{ iff }m=2.\end{cases}

Set vσv_{\sigma} to be the valuation corresponding to (ni)i<ω(n_{i}\mathbb{Z})_{i<\omega} and let ww be the valuation corresponding to (si)i<ω(s^{i}\mathbb{Z})_{i<\omega}. Then ww is definable in (,+,vσ)(\mathbb{Z},+,v_{\sigma}).

Proof.

If a{0}a\in\mathbb{Z}\setminus\{0\}, then there is a unique ta{p0,p1,q}t_{a}\in\{p_{0},p_{1},q\} such that vσ(a)<vσ(taa)v_{\sigma}(a)<v_{\sigma}(t_{a}a). Let a,b{0}a,b\in\mathbb{Z}\setminus\{0\}. If |vσ(a)vσ(b)|3|v_{\sigma}(a)-v_{\sigma}(b)|\geq 3, then

w(a)<w(b)vσ(a)<vσ(b).w(a)<w(b)\iff v_{\sigma}(a)<v_{\sigma}(b).

If |vσ(a)vσ(b)|<3|v_{\sigma}(a)-v_{\sigma}(b)|<3, then w(a)w(b)w(a)\leq w(b) can be determined using tat_{a} and tbt_{b}. ∎

Corollary 5.3.

Let ww be as in Proposition 5.2. Then there are 202^{\aleph_{0}} many valuations vv such that ww is definable in (,+,v)(\mathbb{Z},+,v). Only countably many of those can be definable in (,+,w)(\mathbb{Z},+,w).

Proof.

There are 202^{\aleph_{0}} many valuations vσv_{\sigma} as in Proposition 5.2 and ww is definable in each (,+,vσ)(\mathbb{Z},+,v_{\sigma}). On the other hand, (,+,w)(\mathbb{Z},+,w) has only countably many definable sets. ∎

Remark 5.4.

Note that by Theorem 4.32 all these structures are distal. Hence not even all dp-minimal distal expansions by valuations are minimal expansions.

The fact that expansions by arbitrary valuations are dp-minimal allows us to construct other non-trivial examples: For k2k\geq 2 let vkv_{k} denote the valuation corresponding to the sequence (ki)i<ω(k^{i}\mathbb{Z})_{i<\omega}.

Proposition 5.5.

Let rr and ss be coprime positive integers. Then the expansion

(,+,vr(x)<vs(x))(\mathbb{Z},+,v_{r}(x)<v_{s}(x))

is dp-minimal.

Proof.

We have

vr(x)<vs(x)vrs(x)<vrs(rx).v_{r}(x)<v_{s}(x)\iff v_{rs}(x)<v_{rs}(rx).

Hence the relation vr(x)<vs(x)v_{r}(x)<v_{s}(x) is definable in the dp-minimal structure (,+,vrs)(\mathbb{Z},+,v_{rs}). ∎

It seems unlikely that vrsv_{rs} is definable from vr(x)vs(x)v_{r}(x)\leq v_{s}(x).

The induced structure on the index set ω{}\omega\cup\{\infty\} seems to be important. If it is not o-minimal and Xω{}X\subseteq\omega\cup\{\infty\} is a definable infinite and co-infinite subset, then the set

A={a:w(a)X}A=\{a\in\mathbb{Z}:w(a)\in X\}\subseteq\mathbb{Z}

is definable. It is not clear if ww is definable in (,+,0,1,A)(\mathbb{Z},+,0,1,A).

If the induced structure on ω{}\omega\cup\{\infty\} is o-minimal, then k=|Bi/Bi+1|{}k=|B_{i}/B_{i+1}|\in\mathbb{N}\cup\{\infty\} must be constant for all sufficiently large ii (in some elementary extension). If kk is finite, then the size of the quotients Bi/Bi+1B_{i}/B_{i+1} is bounded and hence we are in the distal case.

Conjecture 5.6.

Let (,+,0,1,v)(\mathbb{Z},+,0,1,v) be distal. Then the following are equivalent:

  1. (a)

    (,+,0,1,v)(\mathbb{Z},+,0,1,v) is a minimal expansion of (,+,0,1)(\mathbb{Z},+,0,1),

  2. (b)

    there is a prime pp such that |Bi/Bi+1|=p|B_{i}/B_{i+1}|=p for almost all i<ωi<\omega,

  3. (c)

    vv is interdefinable with a pp-adic valuation for some prime pp,

  4. (d)

    the (,+,v)(\mathbb{Z},+,v)-induced structure on the value set of vv^{\prime} is o-minimal for all (,+,v)(\mathbb{Z},+,v)-definable valuations vv^{\prime}.

Proposition 5.7.

If (a) implies (d), then 5.6 holds.

Proof.

We already know (b) \implies (c) \implies (a) and by assumption (a) \implies (d) holds. Hence (d) \implies (b) remains to be shown.

Let (,+,0,1,v)(\mathbb{Z},+,0,1,v) be distal and assume (d). Then there is k>1k>1 such that

|Bi:Bi+1|=k|B_{i}:B_{i+1}|=k

for almost all i<ωi<\omega. Therefore vv and vkv_{k} are interdefinable and we may assume v=vkv=v_{k}.

If k=stk=st where ss and tt are coprime, then vv is interdefinable with the valuation ww such that |Biw/Bi+1w||B_{i}^{w}/B_{i+1}^{w}| alternates between ss and tt. Then the induced structure on the value set of ww is not o-minimal. This contradicts (d).

Hence we may assume k=pnk=p^{n} for some prime pp and n1n\geq 1. If n>1n>1, then the pp-adic valuation vpv_{p} is definable by

vp(x)vp(y)r=0n1vpn(prx)vpn(pry).v_{p}(x)\leq v_{p}(y)\iff\bigwedge_{r=0}^{n-1}v_{p^{n}}(p^{r}x)\leq v_{p^{n}}(p^{r}y).

Now the set vp({a:vpn(pa)>vpn(a)})v_{p}(\{a\in\mathbb{Z}:v_{p^{n}}(pa)>v_{p^{n}}(a)\}) contradicts (d).

Hence k=pk=p must be a prime. This shows (b). ∎

There are non-distal candidates for minimal expansions:

Question 5.8.

Let (pi)0<i<ω(p_{i})_{0<i<\omega} be an enumeration of the primes such that each prime appears exactly once and let vv be the valuation corresponding to (p1pi)i<ω(p_{1}\cdots p_{i}\mathbb{Z})_{i<\omega}. Then the induced structure on the value set is o-minimal. Is (,+,0,1,v)(\mathbb{Z},+,0,1,v) a minimal expansion of (,+,0,1)(\mathbb{Z},+,0,1)?

We end this section with the observation that the pp-adic valuations have a limit theory:

Proposition 5.9.

For each prime pp let vpv_{p} denote the pp-adic valuation on \mathbb{Z}. Then the corresponding limit theory exists, i.e.

Th(p(,+,vp)/𝒰)\mathrm{Th}(\prod_{p}(\mathbb{Z},+,v_{p})/\mathcal{U})

does not depend on the choice of the non-principal ultrafilter 𝒰𝒫()\mathcal{U}\subseteq\mathcal{P}(\mathbb{P}).

Proof.

We fix the common \mathcal{L}^{-}-theory

T=𝒰Th(p(,+,vp)/𝒰)T=\bigcap_{\mathcal{U}}\mathrm{Th}_{\mathcal{L}^{-}}(\prod_{p}(\mathbb{Z},+,v_{p})/\mathcal{U})

of these ultraproducts. Note that the predicate Divqkl(i,j)\mathrm{Div}_{q^{k}}^{l}(i,j) fails for all 0<i<j0<i<j and the predicate Indkl\mathrm{Ind}_{k}^{l} holds true for all 0<i<j0<i<j. Thus they are quantifier free 0-definable after naming the successor function SS on \mathcal{I}. Therefore TT has quantifier elimination after naming SS by Theorem 4.8 (because (ω,0,,S)(\omega,0,\leq,S) has quantifier elimination). The constants in Z\mathcal{L}_{Z} generate a subgroup that is isomorphic to \mathbb{Z}. An element a{0}a\in\mathbb{Z}\setminus\{0\} must have valuation 0 in all models of TT. Therefore all models of TT have isomorphic substructures and hence TT is complete. ∎

5.2 Multiple valuations

If PP\subseteq\mathbb{P} is a non-empty set of primes, then Alouf and d’Elbée proved that the structure (,+,vp)pP(\mathbb{Z},+,v_{p})_{p\in P} has dp-rank exactly |P||P|. We will generalize this result to expansions of (,+)(\mathbb{Z},+) by arbitrary valuations which involve disjoint sets of primes.

Let VV be a non-empty family of non-trivial valuations v:ω{}v:\mathbb{Z}\rightarrow\omega\cup\{\infty\}. For each vVv\in V set

πv={p:p divides |Biv/Bi+1v| for some i<ω}.\pi_{v}=\{p\in\mathbb{P}:p\text{ divides }|B_{i}^{v}/B_{i+1}^{v}|\text{ for some }i<\omega\}.

We view (,+)(\mathbb{Z},+) together with these valuations as a multi-sorted structure with group sort 𝒵\mathcal{Z} and with a distinct value sort v\mathcal{I}_{v} for each valuation vVv\in V. Now put

  • 𝒵={0,1,+,}\mathcal{L}_{\mathcal{Z}}=\{0,1,+,-\},

  • v={vl:l1}\mathcal{L}_{v}=\{v^{l}:l\geq 1\}, and

  • v={,0,+,,Divqkl,Indkl}q,k,l\mathcal{L}_{\mathcal{I}_{v}}^{-}=\{-\infty,0,+\infty,\leq,\mathrm{Div}_{q^{k}}^{l},\mathrm{Ind}_{k}^{l}\}_{q,k,l}

for each valuation vVv\in V. Let monv\mathcal{L}_{\text{mon}}^{v} be the monotone hull of v\mathcal{L}_{\mathcal{I}_{v}}^{-} as in Section 4.3 and set

mon=𝒵(vVv)(vVmonv)\mathcal{L}_{\text{mon}}=\mathcal{L}_{\mathcal{Z}}\cup(\bigcup_{v\in V}\mathcal{L}_{v})\cup(\bigcup_{v\in V}\mathcal{L}_{\text{mon}}^{v})

to be the disjoint union of these languages.

Proposition 5.10.

Suppose the sets πv\pi_{v} are pairwise disjoint. Then (,+,v)vV(\mathbb{Z},+,v)_{v\in V} has quantifier elimination in the language mon\mathcal{L}_{\text{mon}}.

Proof.

This is very similar to the proof of Theorem 4.8. Note that a multi-sorted version of Lemma 2.21 holds true in this setting. As in the proof of Theorem 4.8 it suffices to show the back-and-forth property for systems of linear congruences. Let 𝒮\mathcal{S} be the system

nsxbs\displaystyle n_{s}x-b_{s} 0modlsBisv,sSv,\displaystyle\equiv 0\mod l_{s}B_{i_{s}}^{v},\quad s\in S^{v},
ntxbt\displaystyle n_{t}x-b_{t} 0modltBitv,tTv,\displaystyle\not\equiv 0\mod l_{t}B_{i_{t}}^{v},\quad t\in T^{v},

where SvS^{v}\neq\emptyset and TvT^{v} are finite index sets for each vV0v\in V_{0} for a finite subset V0VV_{0}\subseteq V. By an application of Lemma 4.19 we may assume that all the lsl_{s} and ltl_{t} have the same value which we denote by ll.

Let aa be a solution. We will show that we can assume that aa and all constants bsb_{s} and btb_{t} are contained in lB0lB_{0}: If aa is in lB0lB_{0}, then all bsb_{s} must be contained in lB0lB_{0} since otherwise the congruences can not be satisfied. If btb_{t} is not contained in lB0lB_{0}, then the congruence

ntxbt0modlBitvn_{t}x-b_{t}\not\equiv 0\mod lB_{i_{t}}^{v}

does not impose any restrictions on lB0lB_{0} and we can ignore it without changing the solutions in lB0lB_{0}.

If aa is not contained in lB0lB_{0}, then there is a constant cc\in\mathbb{Z} (and hence in the language) such that aclB0a-c\in lB_{0}. In that case the shifted system 𝒮c\mathcal{S}^{c}:

ns(x+c)bs\displaystyle n_{s}(x+c)-b_{s} 0modlBisv,sSv,\displaystyle\equiv 0\mod lB_{i_{s}}^{v},\quad s\in S^{v},
nt(x+c)bt\displaystyle n_{t}(x+c)-b_{t} 0modlBitv,tTv,\displaystyle\not\equiv 0\mod lB_{i_{t}}^{v},\quad t\in T^{v},

is solved by aclB0a-c\in lB_{0} and all the constants nscbsn_{s}c-b_{s} and ntcbtn_{t}c-b_{t} can be assumed to lie in lB0lB_{0}. Thus we can replace 𝒮\mathcal{S} by 𝒮c\mathcal{S}^{c}.

Hence we may assume that 𝒮\mathcal{S} is a system of linear congruences in the subgroup lB0lB_{0}. We have

lB0^=pplB_{0}\equiv\hat{\mathbb{Z}}=\prod_{p}\mathbb{Z}_{p}

and the valuations vlv^{l} involve disjoint sets of primes. Therefore the system 𝒮\mathcal{S} can be solved independently for each valuation vVv\in V. This is done as in the proof of Theorem 4.8. ∎

Theorem 5.11.

Suppose the sets πv\pi_{v} are pairwise disjoint. Then

dp-rk((,+,v)vV)=|V|.\textrm{dp-rk}((\mathbb{Z},+,v)_{v\in V})=|V|.
Proof.

\geq is shown exactly as in the case of the pp-adic valuations which was done by Alouf and d’Elbée (Theorem 1.2 of [2]).

Now assume κ:=dp-rk((,+,v)vV)>|V|\kappa:=\textrm{dp-rk}((\mathbb{Z},+,v)_{v\in V})>|V|. As in the proof of Theorem 4.29 this is witnessed by mutually indiscernible sequences (Ii)i<κ(I_{i})_{i<\kappa} (in the group sort) and a singleton cc in the group sort such that no sequence is indiscernible over cc. As argued in Theorem 4.29, the fact that a sequence II is not indiscernible over cc must be witnessed by an atomic mon\mathcal{L}_{\text{mon}}-formula which involves a valuation.

Since κ>|V|\kappa>|V|, there must be two sequences I1I_{1} and I2I_{2} for which this witnessing formula involves the same valuation vv. This is a contradiction because (,+,v)(\mathbb{Z},+,v) is dp-minimal by Theorem 4.29. ∎

6 Further results

6.1 Uniformly definable families of finite-index subgroups of dp-minimal groups

The classification of NIP profinite groups by Macpherson and Tent in [19] yields information about uniformly definable families of finite index subgroups in arbitrary NIP groups (see Theorem 8.7 in [3]). We will do the same in the dp-minimal case. The arguments are almost identical to those in Section 8 of [3] (see also Remark 5.5 in [19]), we only need to make sure that the construction presented there preserves dp-minimality.

Let HH be a group and let (Ni:iI)(N_{i}:i\in I) be a family of normal subgroups of finite index such that

i,jk:NkNiNj.\forall i,j\exists k:N_{k}\leq N_{i}\cap N_{j}.

We view HH as an prof\mathcal{L}_{\text{prof}}-structure =(H,I)\mathcal{H}=(H,I). Let fj:limH/NiH/Njf_{j}:\varprojlim H/N_{i}\rightarrow H/N_{j} be the projection maps. Then {kerfj:jI}\{\ker f_{j}:j\in I\} is a neighborhood basis at the identity. Therefore we may view limH/Ni\varprojlim H/N_{i} as an prof\mathcal{L}_{\text{prof}}-structure (limH/Ni,I)(\varprojlim H/N_{i},I).

Lemma 6.1.

Let =(H,I)\mathcal{H}^{*}=(H^{*},I^{*}) be an |I|+|I|^{+}-saturated elementary extension of \mathcal{H}. Then

/iINilimiIH/Ni\mathcal{H}^{*}/\bigcap_{i\in I}N_{i}^{*}\cong\varprojlim_{i\in I}H/N_{i}

and (Nj/iINi:jI)(N_{j}^{*}/\bigcap_{i\in I}N_{i}^{*}:j\in I) is a neighborhood basis for the identity consisting of open normal subgroups.

Proof.

By elementarity we have |H:Ni|=|H:Ni||H^{*}:N_{i}^{*}|=|H:N_{i}| for all iIi\in I. Using this and elementarity it is easy to see that

limiIH/Ni=limiIH/Ni.\varprojlim_{i\in I}H^{*}/N_{i}^{*}=\varprojlim_{i\in I}H/N_{i}.

Now write

limiIH/Ni={(giNi)i:ij:giNj=gjNj}\varprojlim_{i\in I}H^{*}/N_{i}^{*}=\{(g_{i}N_{i}^{*})_{i}:\forall i\geq j:g_{i}N_{j}^{*}=g_{j}N_{j}^{*}\}

and let f:HlimiIH/Ni,g(gNi)if:H^{*}\rightarrow\varprojlim_{i\in I}H^{*}/N_{i}^{*},g\mapsto(gN_{i}^{*})_{i} be the natural homomorphism. Clearly kerf=iINi\ker f=\bigcap_{i\in I}N_{i}^{*}. It remains to show that ff is surjective.

Fix (giNi)ilimiIH/Ni(g_{i}N_{i})_{i}\in\varprojlim_{i\in I}H^{*}/N_{i}^{*} and consider the partial type

Σ(x)={xgiNi:iI}.\Sigma(x)=\{x\in g_{i}N_{i}^{*}:i\in I\}.

Given IoII_{o}\subseteq I finite, there is jIj\in I such that NjiI0NiN_{j}\leq\bigcap_{i\in I_{0}}N_{i}. Then giNj=giNjg_{i}N_{j}=g_{i^{\prime}}N_{j} for all i,iI0i,i^{\prime}\in I_{0}. Hence Σ(x)\Sigma(x) is finitely satisfiable and as \mathcal{H}^{*} is |I|+|I|^{+}-saturated, there exists gHg\in H^{*} such that ggiNig\in g_{i}N_{i}^{*} for all iIi\in I and hence f(g)=(giNi)f(g)=(g_{i}N_{i}^{*}).

The family (Nj/iINi:jI)(N_{j}^{*}/\bigcap_{i\in I}N_{i}^{*}:j\in I) is a neighborhood basis for the identity consisting of open normal subgroups. ∎

Lemma 6.2.

If =(H,I)\mathcal{H}=(H,I) has NIP, then (limiIH/Ni,I)(\varprojlim_{i\in I}H/N_{i},I) has NIP. If moreover (H,I)(H,I) is dp-minimal, then (limiIH/Ni,I)(\varprojlim_{i\in I}H/N_{i},I) is dp-minimal.

Proof.

Since (H,I)(H,I) has NIP, every uniformly definable family of subgroups contains only finitely many subgroups of each finite index. Let (H,I)(H^{*},I^{*}) be an |I|+|I|^{+}-saturated elementary extension. Then II is externally definable (since I={iI:|H:Ki|<}I=\{i\in I^{*}:|H^{*}:K_{i}^{*}|<\infty\}). If (H,I)(H,I) is dp-minimal, then (H,I,I)(H^{*},I^{*},I) is dp-minimal by Remark 2.16. By the above lemma the structure (limiIH/Ni,I)(\varprojlim_{i\in I}H/N_{i},I) is interpretable as a quotient in (H,I,I)(H^{*},I^{*},I) and hence is NIP (resp. dp-minimal). ∎

Let GG be a group and let φ(x,y)\varphi(x,y) be a formula. Set 𝒩φ={Ni:iI}\mathcal{N}_{\varphi}=\{N_{i}:i\in I\} to be the family of all normal subgroups which are finite intersections of conjugates of φ\varphi-definable subgroups of finite index. Note that every φ\varphi-definable subgroup of finite index contains some N𝒩φN\in\mathcal{N}_{\varphi}. The profinite group limiIG/Ni\varprojlim_{i\in I}G/N_{i} naturally becomes an LprofL_{\text{prof}}-structure 𝒢φ=(limiIG/Ni,I)\mathcal{G}_{\varphi}=(\varprojlim_{i\in I}G/N_{i},I).

Proposition 6.3.

Let GG and φ\varphi be as above. If GG is NIP, then 𝒢φ\mathcal{G}_{\varphi} is NIP. If moreover GG is dp-minimal, then 𝒢φ\mathcal{G}_{\varphi} is dp-minimal in the group sort.

Proof.

By Baldwin-Saxl finite intersections of conjugates of φ\varphi-definable subgroups are uniformly definable by some formula ψ(x,z)\psi(x,z). The set J={b:ψ(G,b)G}J=\{b:\psi(G,b)\trianglelefteq G\} is definable. Put J0={b:|G:ψ(G,b)|<}J_{0}=\{b:|G:\psi(G,b)|<\infty\}. Then 𝒩φ={ψ(G,b):bJ0}\mathcal{N}_{\varphi}=\{\psi(G,b):b\in J_{0}\}. Since 𝒩φ\mathcal{N}_{\varphi} is closed under intersections, it follows that J0J_{0} is externally definable. Let EE be the equivalence relation defined by aEbψ(G,a)=ψ(G,b)aEb\iff\psi(G,a)=\psi(G,b). Now apply the previous lemma to the structure (G,J0/E)(G,J_{0}/E). ∎

By Proposition 3.3 every dp-minimal profinite group (G,I)(G,I) has an open abelian subgroup. Now Proposition 6.3 implies the following:

Proposition 6.4.

Let GG be a dp-minimal group and let φ(x,y)\varphi(x,y) be a formula. Let 𝒩φ\mathcal{N}_{\varphi} be the family of all normal subgroups which are finite intersections of conjugates of φ\varphi-definable subgroups of finite index. If 𝒩φ\mathcal{N}_{\varphi} is infinite, then there is N𝒩φN\in\mathcal{N}_{\varphi} such that for all M𝒩φM\in\mathcal{N}_{\varphi} the quotient N/(NM)N/(N\cap M) is abelian.

Proof.

The profinite group (limN𝒩φG/N,𝒩φ)(\varprojlim_{N\in\mathcal{N}_{\varphi}}G/N,\mathcal{N}_{\varphi}) is dp-minimal and therefore is virtually abelian by Proposition 3.3. Since the quotients G/NG/N are preserved, this implies the proposition. ∎

Remark 6.5.

By Theorem 3.4 there are essentially two types of dp-minimal profinite groups. This will also be seen in the abelian quotients in the statement of Proposition 6.4.

Remark 6.6.

By Proposition 5.1 of [19] every profinite NIP group (G,I)(G,I) has an open prosolvable subgroup. Hence if we only assume NIP in the previous theorem, the quotients will be solvable instead of abelian (see Theorem 8.7 of [3]).

6.2 Strong homogeneity of profinite groups

Jarden and Lubotzky [8] showed that two elementarily equivalent profinite groups are isomorphic if one of them is finitely generated. This was generalized to strongly complete profinite groups by Helbig [7]. A profinite group is strongly complete if all subgroups of finite index are open. The tools used by Helbig and the construction in Section 6.1 give a proof for strong homogeneity.

Let GG be a profinite group and suppose (Ni:iI)(N_{i}:i\in I) is a neighborhood basis at the identity consisting of open normal subgroups. Let P\mathcal{L}_{P} be the group language expanded by a family of unary predicates (Pi:iI)(P_{i}:i\in I). We consider GG as an p\mathcal{L}_{p} structure by setting Pi(G)=NiP_{i}(G)=N_{i}. Note that if GG^{*} is an elementary expansion, then there is a natural P\mathcal{L}_{P}-structure on the quotient G/(iIPi(G))G^{*}/(\bigcap_{i\in I}P_{i}(G^{*})).

Lemma 6.7.

Let GG a profinite group equipped with an P\mathcal{L}_{P} structure as above. Let GG^{*} be an elementary extension of GG in the language P\mathcal{L}_{P}. Then the composition

GGG/(iIPi(G))G\rightarrow G^{*}\rightarrow G^{*}/(\bigcap_{i\in I}P_{i}(G^{*}))

is an P\mathcal{L}_{P}-isomorphism.

Proof.

The lemma follows from the same arguments as Lemma 6.1. ∎

Proposition 6.8.

Let GG and HH be profinite groups as P\mathcal{L}_{P} structures such that the predicates (Pi:iI)(P_{i}:i\in I) encode neighborhood bases at the identity consisting of open normal subgroups in both groups. Suppose AGA\subseteq G is a subset and f:AHf:A\rightarrow H is an elementary map with respect to the language P\mathcal{L}_{P}. Then ff extends to an P\mathcal{L}_{P}-isomorphism between GG and HH.

Proof.

Let GG^{*} be a common strongly |A|+|A|^{+}-homogeneous elementary extension of GG and HH. We can find f~Aut(G)\tilde{f}\in\mathrm{Aut}(G^{*}) such that f~|A=f\tilde{f}|_{A}=f. Since f~\tilde{f} is an LPL_{P}-automorphism, it induces an automorphism of G/(iIPi(G))G^{*}/(\bigcap_{i\in I}P_{i}(G^{*})). Now use Lemma 6.7 to get the desired isomorphism between GG and HH. ∎

The following observation in Remark 3.12 in [7] is a consequence of Theorem 2 in [18] and Corollary 52.12 in [11]:

Theorem 6.9.

Let GG be a profinite group. Then the following are equivalent:

  • (a)

    GG is strongly complete.

  • (b)

    For each finite group AA there exists a group word ww such that w(A)=1w(A)=1 and w(G)w(G) is open in GG.

Recall that a group word has finite width if w(G)=w(G)n\langle w(G)\rangle=w(G)^{n} for some n>ωn>\omega. We will make use of the following result:

Proposition 6.10 (Proposition 5.2(b) of [23]).

Let GG be a profinite group. If ww is a group word, then w(G)w(G) is closed in GG if and only if ww has finite width in GG.

Proposition 6.11.

Let GG and HH be profinite groups. Let AGA\subseteq G be a subset of GG and let f:AHf:A\rightarrow H be an elementary map. If one of the groups is strongly complete, then ff extends to an isomorphism.

Proof.

By Theorem 6.9 and Proposition 6.10 strong completeness is a first-order property among profinite groups. For each finite group AA there is a group word wAw_{A} such that wA(A)=1w_{A}(A)=1, wA(G)w_{A}(G) is open in GG, and wA(H)w_{A}(H) is open in HH. Note that by Proposition 6.10 and elementary equivalence of GG and HH, wA(G)w_{A}(G) and wA(H)w_{A}(H) are definable by the same formula without parameters.

If NN is an open normal subgroup of GG then wG/N(G/N)=1w_{G/N}(G/N)=1 and hence wG/N(G)Nw_{G/N}(G)\subseteq N. Therefore the family (wB(G):B a finite group)(w_{B}(G):B\text{ a finite group}) is a neighborhood basis at the identity.

Hence we may consider GG and HH as P\mathcal{L}_{P}-structures where the predicates are given by PB(G)=wB(G)P_{B}(G)=w_{B}(G). By Proposition 6.8 ff extends to an isomorphism. ∎

6.3 A result on families of subgroups of NTP2 groups

By Theorem 1.1 of [19] a full profinite group is NIP if and only if it is NTP2. Since the structure of these groups is determined by the lattice of subgroups, this only depends on a single formula. We will show a version for formulas in NTP2 groups. We will use the following lemma by Macpherson and Tent on groups in NTP2:

Lemma 6.12 (Lemma 4.3 in [19]).

Let GG be an \emptyset-definable group in a structure with NTP2 theory, and φ(x,y¯)\varphi(x,\bar{y}) a formula implying xGx\in G. Then there is k=kφk=k_{\varphi}\in{\mathbb{N}} such that the following holds. Suppose that HH is a subgroup of GG, π:HΠiJTi\pi:H\longrightarrow\Pi_{i\in J}T_{i} is an epimorphism to the Cartesian product of the groups TiT_{i}, and πj:HTj\pi_{j}:H\longrightarrow T_{j} is for each jJj\in J the composition of π\pi with the canonical projection ΠiJTiTj\Pi_{i\in J}T_{i}\to T_{j}. Suppose also that for each jJj\in J, there is a subgroup R¯jG\bar{R}_{j}\leq G and group Rj<TjR_{j}<T_{j} with R¯jH=πj1(Rj)\bar{R}_{j}\cap H=\pi_{j}^{-1}(R_{j}), such that finite intersections of the groups R¯j\bar{R}_{j} are uniformly definable by instances of φ(x,y¯)\varphi(x,\bar{y}). Then |J|k|J|\leq k.

Proposition 6.13.

Let GG be an NTP2 group and let φ(x,y)\varphi(x,y) be a formula such that |x|=1|x|=1. Suppose that the family {φ(G,b):bG}\{\varphi(G,b):b\in G\} consists of normal subgroups of GG and is closed under finite intersections. Then φ(x,y)\varphi(x,y) has NIP.

Proof.

We aim to show that φ\varphi satisfies the Baldwin-Saxl condition. Let N1,NnN_{1},\dots N_{n} be instances of φ\varphi and fix kφk_{\varphi} as in Lemma 6.12. Now set Ck={Ni:1in,ik}C_{k}=\bigcap\{N_{i}:1\leq i\leq n,i\neq k\} and C={Ni:1in}C=\bigcap\{N_{i}:1\leq i\leq n\}. Note that CiNi=CC_{i}\cap N_{i}=C and we have CiNjC_{i}\subseteq N_{j} if iji\neq j. Now set

H=Ci:1in.H=\langle C_{i}:1\leq i\leq n\rangle.

We then have

H/Ci=1nCi/C.H/C\cong\prod_{i=1}^{n}C_{i}/C.

Now set Ti=Ci/CT_{i}=C_{i}/C and assume that all TiT_{i} are non-trivial. Let

π:HH/C=i=1nTi\pi:H\rightarrow H/C=\prod_{i=1}^{n}T_{i}

be the natural projection and let πj:HTj\pi_{j}:H\rightarrow T_{j} be the projection on TjT_{j}. Set Rj=1<TjR_{j}=1<T_{j} and put R¯j=Nj\bar{R}_{j}=N_{j}. Then

R¯jH=NjH=πj1(Rj).\bar{R}_{j}\cap H=N_{j}\cap H=\pi_{j}^{-1}(R_{j}).

Hence Lemma 6.12 implies nkφn\leq k_{\varphi}.

If n>kφn>k_{\varphi}, then there must be 1in1\leq i\leq n such that Ti=1T_{i}=1 and hence C=CiC=C_{i} can be written as in intersection of n1n-1 instances of φ\varphi. Inductively this shows that any intersection of instances of φ\varphi is an intersection of at most kφk_{\varphi} instances. Hence φ\varphi satisfies the Baldwin-Saxl condition.

This implies that φ\varphi has NIP: Otherwise we can find a constant aa and an indiscernible sequence (bi)i<ω(b_{i})_{i<\omega} such that

φ(a,bi)i is odd.\models\varphi(a,b_{i})\iff i\text{ is odd.}

Set n=kφn=k_{\varphi} and take i0,in<ωi_{0},\dots i_{n}<\omega, all of them odd. By the Baldwin-Saxl lemma we may assume

Hi0Hin=Hi1Hin.H_{i_{0}}\cap\dots H_{i_{n}}=H_{i_{1}}\cap\dots H_{i_{n}}.

By indiscernibility this implies

Hi0+1Hin=Hi1Hin.H_{i_{0}+1}\cap\dots H_{i_{n}}=H_{i_{1}}\cap\dots H_{i_{n}}.

But this is a contradiction since aHi0+1a\not\in H_{i_{0}+1}. ∎

While it is clearly sufficient to assume that the φ\varphi-definable subgroups normalize each other, the above proof requires some normality assumption.

Question 6.14.

Does Proposition 6.13 hold even without any normality assumption?

References

  • [1] Eran Alouf. On dp-minimal expansions of the integers, 2020. arXiv:2001.11480 [math.LO].
  • [2] Eran Alouf and Christian D’Elbée. A new dp-minimal expansion of the integers. The Journal of Symbolic Logic, 84(2):632–663, 2019.
  • [3] Tim Clausen and Katrin Tent. Some model theory of profinite groups. Lectures in Model Theory, Münster Lectures in Mathematics, EMS Publishing, 2016.
  • [4] Gabriel Conant. There are no intermediate structures between the group of integers and presburger arithmetic. The Journal of Symbolic Logic, 83(1):187–207, 2018.
  • [5] Gabriel Conant and Anand Pillay. Stable groups and expansions of (,+,0)(\mathbb{Z},+,0). Fundamenta Mathematicae, 01 2016.
  • [6] J. D. Dixon, M. P. F. Du Sautoy, A. Mann, and D. Segal. Analytic Pro-P Groups. Cambridge Studies in Advanced Mathematics. Cambridge University Press, 2 edition, 1999.
  • [7] Patrick Helbig. On small profinite groups. Journal of Group Theory, 20(5), 2017.
  • [8] Moshe Jarden and Alexander Lubotzky. Elementary equivalence of profinite groups. Bulletin of The London Mathematical Society, 40:887–896, 07 2008.
  • [9] Franz-Viktor Kuhlmann and Salma Kuhlmann. Ax-Kochen-Ershov principles for valued and ordered vector spaces. In W. Charles Holland, editor, Ordered Algebraic Structures, pages 237–259, Dordrecht, 1997. Kluwer.
  • [10] Fares Maalouf. Espaces vectoriels c-minimaux. J. Symbolic Logic, 75(2):741–758, 06 2010.
  • [11] Hanna Neumann. Varieties of Groups, volume 37 of Ergebnisse der Mathematik und ihrer Grenzgebiete. 2. Folge. Springer, first edition, 1967.
  • [12] Daniel Palacín and Rizos Sklinos. On superstable expansions of free abelian groups. Notre Dame J. Formal Log., 59:157–169, 2018.
  • [13] Luis Ribes and Pavel Zalesskii. Profinite Groups, volume 40 of A Series of Modern Surveys in Mathematics. Springer, second edition, 2000.
  • [14] Pierre Simon. On dp-minimal ordered structures. J. Symb. Log., 76:448–460, 2011.
  • [15] Pierre Simon. Distal and non-distal nip theories. Ann. Pure Appl. Log., 164:294–318, 2013.
  • [16] Pierre Simon. A Guide to NIP Theories. Cambridge, 2015.
  • [17] Patrick Simonetta. An example of a c-minimal group which is not abelian-by-finite. Proceedings of the American Mathematical Society, 131, 12 2003.
  • [18] M.G. Smith and John Wilson. On subgroups of finite index in compact hausdorff groups. Archiv der Mathematik, 80:123–129, 04 2003.
  • [19] Katrin Tent and Dugald Macpherson. Profinite groups with NIP theory and p-adic analytic groups. Bull. Lond. Math. Soc., 48(6):1037–1049, 2016.
  • [20] Katrin Tent and Martin Ziegler. A Course in Model Theory. Lecture Notes in Logic. Cambridge University Press, 2012.
  • [21] Minh Tran and Erik Walsberg. A family of dp-minimal expansions of (;+)(\mathbb{Z};+), 2017. arXiv:1711.04390 [math.LO].
  • [22] Erik Walsberg. Dp-minimal expansions of (,+)(\mathbb{Z},+) via dense pairs via Mordell-Lang, 2020. arXiv:2004.06847 [math.LO].
  • [23] J.S. Wilson. Profinite Groups. Clarendon Press, 1998.
  • [24] E. I. Zelmanov. On periodic compact groups. Israel Journal of Mathematics, 77:83–95, 1992.