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

Free structures and limiting density

Johanna N. Y. Franklin, Meng-Che “Turbo” Ho , and Julia Knight The first author was supported in part by Simons Foundation Collaboration Grant #420806.The second author acknowledges support from the National Science Foundation under Grant No. DMS-2054558.The first and third authors acknowledge support from the National Science Foundation under Grant #DMS-1800692.This material is based upon work supported by the National Science Foundation under Grant No. DMS-1928930 while the second and third authors participated in a program hosted by the Mathematical Sciences Research Institute in Berkeley, California, during Fall 2020 and Summer 2022 programs.
Abstract

Gromov asked what a typical (finitely presented) group looks like, and he suggested a way to make the question precise in terms of limiting density. The typical finitely generated group is known to share some important properties with the non-abelian free groups. We ask Gromov’s question more generally, for structures in an arbitrary algebraic variety (in the sense of universal algebra), with presentations of a specific form. We focus on elementary properties. We give examples illustrating different behaviors of the limiting density. Based on the examples, we identify sufficient conditions for the elementary first-order theory of the free structure to match that of the typical structure; i.e., a sentence is true in the free structure iff it has limiting density 11.

1 Introduction

In the paper where he introduced the notion of hyperbolic group, Gromov [8] asked what a typical group looks like. He was thinking of finitely presented groups. He described, in terms of limiting density, what it might mean for a typical group to have some property QQ, and he stated that the typical group is hyperbolic. Gromov’s notion has been made precise in different ways; see, for instance, the survey [17]. Ol’shanskii [18] cleaned up the statement and the proof that the typical group is hyperbolic. The third author conjectured that for the typical group obtained from a presentation consisting of m2m\geq 2 generators and a single relator, the elementary first order theory matches that of the non-abelian free groups. In this paper, we generalize Gromov’s question to arbitrary equational classes, or algebraic varieties (in the sense of universal algebra). Here, as for groups, the notions of finite presentation and free structure make sense. We find examples exhibiting different behavior. Our main results are for varieties with only unary functions. With some rather strong conditions on the variety, and on the presentations, we obtain the analogue of the conjecture for groups: the sentences true in the free structure are exactly those with limiting density 11.

1.1 Organization

We begin with Gromov’s original question, which concerned finitely presented groups. We then generalize this question to finitely presented members of a variety VV. Section 2 has some background on algebraic varieties. Section 3 has four examples illustrating different possible behaviors and also introduces key ideas that will appear in later proofs. The first example shows that in a bijective structure with a single identity, the sentences true in the free structures have limiting density 11. In the second example, the set of sentences with limiting density 1 is the theory of a structure in the variety, but this structure is not finitely generated, nor is it free. In the third example, we look at bijective structures as in the first example but with two identities, and we give a sentence for which the limiting density does not exist. In the fourth example, we look at abelian groups and give sentences where the limiting density is strictly between 0 and 11.

In Section 4, we consider more general bijective varieties. We show that for these varieties and presentations with a single generator and a single identity, a sentence has limiting density 11 if and only if it is true in the free structure. For a language with unary function symbols f1,,fnf_{1},\ldots,f_{n}, injective and commuting, we say how to find elementary invariants.111We are grateful to Sergei Starchenko, who, having seen our result for bijective structures, suggested that we look further at varieties for languages with unary functions. We use a version of Gaifman’s Locality Theorem, which we prove using saturation.222We are grateful to Phokion Kolaitis for alerting us to Gaifman’s Theorem and providing references. In Section 5, we consider sentences with constants. For an arbitrary variety and presentations with a fixed tuple a¯\bar{a} of generators, we give conditions guaranteeing that, for sentences φ\varphi in the language with added constants naming the generators, φ\varphi has limiting density 11 if and only if it is true in the free structure on a¯\bar{a}. In Section 6, we give further examples illustrating the results from Sections 4 and 5.

1.2 Gromov’s question about groups

Here, we recall Gromov’s original question and mention some prior work on typical or random groups. The usual language for groups has a binary operation symbol (for the group operation), a unary operation symbol (for the inverse), and a constant (for the identity). Let TT be the theory of groups. Recall that a group presentation consists of a tuple a¯\bar{a} of generators and a set RR of words wi(a¯)w_{i}(\bar{a}) on these generators, called relators. In the group GG with presentation a¯|R\langle\bar{a}|R\rangle, Gt(a¯)=eG\models t(\bar{a})=e if and only if T{w(a¯)=e:w(a¯)R}t(a¯)=eT\cup\{w(\bar{a})=e:w(\bar{a})\in R\}\models t(\bar{a})=e. Suppose FF is the free group on a¯\bar{a} and NN is the subgroup of FF consisting of the elements t(a¯)t(\bar{a}) such that T{w(a¯)=e}t(a¯)=eT\cup\{w(\bar{a})=e\}\models t(\bar{a})=e. Then GF/NG\cong F/N.

The notion of limiting density depends not just on the variety, but also on the allowable group presentations, Ol’shanskii [18] considered presentations with mm generators and kk relators, all reduced. Kapovich and Schupp [11] considered the case where k=1k=1. In the Gromov “density” model, the number of relators may vary but is bounded in terms of the length of the relators and a parameter dd. It is important to bound the number of relators in some way; otherwise, the typical group will almost surely be trivial [18].

Definition 1.1.

Let QQ be a property of interest. Let PsP_{s} be the number of presentations in which the relators have length at most ss, and let Ps(Q)P_{s}(Q) be the number of these presentations for which the resulting group has property QQ. The limiting density for QQ is limsPs(Q)Ps\lim_{s\rightarrow\infty}\frac{P_{s}(Q)}{P_{s}}, if this limit exists.

We consider the typical group to have property QQ if QQ has density 11. We are particularly interested in elementary properties. For a sentence φ\varphi, the density of the property of satisfying φ\varphi will be called simply the density of φ\varphi. The typical group, in the sense of limiting density, is also called the random group.333Harrison-Trainor, Khoussainov, and Turetsky [9] took a different approach and considered random structures more along the lines of the Rado graph. The typical group has some properties of free groups. Gromov introduced the property of hyperbolicity and stated that the typical group is hyperbolic. Ol’shanskii [18] showed that for presentations with mm generators and kk reduced relators, the property of being hyperbolic has limiting density 11. Kapovich and Schupp [11] showed that for presentations with mm generators and 11 reduced relator, the property that all minimal generating tuples are Nielsen equivalent has limiting density 11. Nielsen equivalence means that one tuple can be transformed into the other by a finite sequence of simple, obviously reversible, kinds of steps.

Benjamin Fine, in conversation with the third author at the JMM in January of 2013, made an off-hand comment to the effect that in the limiting density sense, all groups look free. Fine’s comment gave rise to the conjecture below, saying that for presentations with m2m\geq 2 generators and 11 relator, the typical group has the same elementary first order theory as the the free group. By a result of Sela [20] (see also Kharlampovich-Miasnikov [12]), the elementary first-order theory of all non-abelian free groups is the same. The conjecture is given in [10, Conjecture 2.2].

Conjecture 1.2 (Knight).

Take groups given by presentations with a fixed mm-tuple of generators, for m2m\geq 2, and 11 relator (the Kapovich-Schupp model). For all elementary first order sentences φ\varphi,

  1. 1.

    the limiting density exists, and

  2. 2.

    the density has value 11 if φ\varphi is true in the non-abelian free groups, and 0 otherwise.

There is some evidence for the conjecture. By a result of Arzhantseva and Ol’shanskii [2, 1], a random group obtained from a presentation with mm generators and kk relators has many free subgroups. Thus, an existential sentence true in the free group is also true in a random group. Kharlampovich and Sklinos [13] used Gromov’s density model, with parameter dd. In this setting, they showed the following.

Theorem 1.3 (Kharlampovich and Sklinos [13]).

A random group, in Gromov’s density model with d1/16d\leq 1/16, satisfies a universal sentence if and only if the sentence is true in the non-abelian free groups.

The Kharlampovich-Sklinos result implies the conjecture (for universal sentences), but we will not give a proof here.

2 Generalizing Gromov’s question

The question that Gromov asked about groups makes sense for other algebraic varieties as well. We begin by presenting our definition of an algebraic variety; then we discuss the types of presentations we will allow and give some basic lemmas.

2.1 Algebraic varieties

Definition 2.1.

A language is algebraic if it consists only of function symbols and constants.

The term “algebraic variety” is used to mean different things in algebraic geometry and in universal algebra. The definition that we give below is the one from universal algebra.

Definition 2.2 (Algebraic variety).

For a fixed algebraic language LL, a class VV of LL-structures is an algebraic variety, or simply variety, if it is closed under substructures, homomorphic images, and direct products.

For our purposes, it is convenient to use the following equivalent definition, of “equational class.”

Definition 2.3 (Equational class).

For a fixed algebraic language LL, a class VV of LL-structures is an equational class if it is axiomatized by sentences of the form (x¯)t(x¯)=t(x¯)(\forall\bar{x})t(\bar{x})=t^{\prime}(\bar{x})—universal quantifiers in front of an equation.

Birkhoff showed that these two definitions are equivalent. Mal’tsev defined a broader class of theories whose models have well-defined presentations. See [3] for a general overview of universal algebra, where the result below appears as Theorem 11.9.

Theorem 2.4.

For a fixed algebraic language LL, a class of LL-structures is an equational class if and only if it is a variety.

In the usual language for groups, namely {,1,e}\{\cdot,^{-1},e\}, the group axioms have the required form. Thus, groups form a variety.


Now we consider an arbitrary algebraic variety VV. For a fixed generating tuple a¯\bar{a}, there is a well-defined free structure FF generated by a¯\bar{a}. If 𝒜\mathcal{A} is a structure in VV generated by a¯\bar{a}, then 𝒜\mathcal{A} is a quotient of FF under an appropriate equivalence relation on terms t(a¯)t(\bar{a}). This equivalence relation becomes equality in 𝒜\mathcal{A}.

Definition 2.5.

For a variety VV, a presentation has the form a¯|R\bar{a}|R, where a¯\bar{a} is a generating tuple and RR is a set of identities on a¯\bar{a}. We write a¯|R\langle\bar{a}|R\rangle for the structure 𝒜\mathcal{A} such that the identities t(a¯)=t(a¯)t(\bar{a})=t^{\prime}(\bar{a}) true in 𝒜\mathcal{A} are just the ones that follow logically from RR and the axioms for VV.

We ask what the typical behavior is for members of a variety given by presentations of a specific form.

2.2 Allowable presentations

In this paper, almost all of the languages of our varieties will be either the group language or a language with just unary function symbols. We consider presentations with a fixed generating tuple a¯\bar{a}, say of length mm. For the analogue of the Ol’shanskii setting, we consider presentations with kk identities for some fixed kk. For the analogue of the Kapovich-Schupp setting, we set k=1k=1. This is the primary case we will consider. Where we do consider k>1k>1, our presentations have the form a¯|R\bar{a}|R, where RR is an unordered set of identities.

We may restrict the identities in certain natural ways. For groups, we do what the group theorists do; that is, we suppose that the identities have the form w(a¯)=ew(\bar{a})=e, where w(a¯)w(\bar{a}) is a word representing a product of various aia_{i} and ai1a_{i}^{-1}. For the variety in the language consisting of two unary function symbols S,S1S,S^{-1} with axioms saying that the two functions are inverses, we may restrict in a similar way, allowing only identities of the form t(ai)=ajt(a_{i})=a_{j}; that is, with function symbols only on the left. For the language with finitely many unary function symbols f1,,fnf_{1},\ldots,f_{n} and varieties that do not have axioms explicitly saying that one fjf_{j} is the inverse of another fif_{i}, our identities have the form t(ai)=t(aj)t(a_{i})=t^{\prime}(a_{j}), where t(x)t(x) and t(x)t^{\prime}(x) are terms built up from the function symbols.

2.2.1 Length

We will need to measure length of identities in our presentations. We will use the following conventions, based on the restrictions described above.

Definition 2.6.
  • In the setting of groups, the length of an identity of the form w(a¯)=ew(\bar{a})=e is the number of occurrences of the various aia_{i} and ai1a_{i}^{-1} in the word w(a¯)w(\bar{a}). This is the usual length of the relator.

  • For varieties in the language LL with just the unary function symbols f1,,fnf_{1},\ldots,f_{n}, the length of an identity of the form t(ai)=t(aj)t(a_{i})=t^{\prime}(a_{j}) is the total number of occurrences of the function symbols in the terms tt and tt^{\prime}.

2.2.2 Limiting density

As for groups, we consider limiting density. Here is the formal definition of limiting density.

Definition 2.7.

Fix a language, a variety, and a set of allowable presentations with an mm-tuple a¯\overline{a} of generators and kk identities. We write PsP_{s} for the number of presentations in which all of the identities have length at most ss. For a property QQ, let Ps(Q)P_{s}(Q) be the number of these presentations for which the resulting structure has property QQ. Then the limiting density of QQ is limsPs(Q)Ps\lim_{s\rightarrow\infty}\frac{P_{s}(Q)}{P_{s}}, provided that this limit exists.

We are particularly interested in the case where QQ is the property of satisfying an elementary first-order sentence φ\varphi in the language of the variety, possibly with added constants for the generators. We write Ps(φ)P_{s}(\varphi) for the number of presentations in which the identities have length at most ss, and the resulting structure satisfies φ\varphi. We say that φ\varphi has limiting density dd if limsPs(φ)Ps=d\lim\limits_{s\to\infty}\frac{P_{s}(\varphi)}{P_{s}}=d.

Definition 2.8.

We say the variety VV, with a specified set of allowable presentations, satisfies the zero–one law if for every elementary first-order sentence φ\varphi in LL, φ\varphi has limiting density 11 or 0.

2.2.3 Sets versus tuples of identities

We have said that our presentations consist of a tuple of generators and an unordered set of distinct identities. Other possibilities would be to consider ordered tuples of identities, with or without repetition. In practice, most of the time, we will consider a single identity. When we do consider more than one identity, we show that the results would be the same for ordered tuples of identities allowing repetition, ordered tuples not allowing repetition, and unordered sets of identities.

As above, we write PsP_{s} for the number of unordered sets of kk identities of length at most ss. In the result below, we write PsP^{*}_{s} for the number of ordered kk-tuples allowing repetition, and PsP^{**}_{s} for the number of ordered kk-tuples not allowing repetition.

Proposition 2.9.

Let N(s)N(s) be the number of identities in LL of length at most ss and suppose that limsN(s)=\lim\limits_{s\to\infty}N(s)=\infty. Then for any sentence φ\varphi,

limsPs(φ)Ps=limsPs(φ)Ps=limsPs(φ)Ps.\lim\limits_{s\to\infty}\frac{P_{s}(\varphi)}{P_{s}}=\lim\limits_{s\to\infty}\frac{P^{*}_{s}(\varphi)}{P^{*}_{s}}=\lim\limits_{s\to\infty}\frac{P^{**}_{s}(\varphi)}{P^{**}_{s}}\ .
Proof.

By definition, Ps=(N(s)k)P_{s}={N(s)\choose k}, Ps=N(s)kP^{*}_{s}=N(s)^{k}, and Ps=k!PsP^{**}_{s}=k!\cdot P_{s}. Each unordered set of kk identities yields k!k! ordered kk-tuples of identities. Thus, it is clear that

limsPs(φ)Ps=limsPs(φ)Ps.\lim\limits_{s\to\infty}\frac{P_{s}(\varphi)}{P_{s}}=\lim\limits_{s\to\infty}\frac{P^{**}_{s}(\varphi)}{P^{**}_{s}}\ .

To compare Ps(φ)Ps\frac{P_{s}(\varphi)}{P_{s}} and Ps(φ)Ps\frac{P^{*}_{s}(\varphi)}{P^{*}_{s}}, we need the following:


Claim:

  1. 1.

    k!PsPs1\frac{k!P_{s}}{P^{*}_{s}}\rightarrow 1,

  2. 2.

    N(s)kk!PsPs0\frac{N(s)^{k}-k!P_{s}}{P^{*}_{s}}\rightarrow 0.

Proof of Claim.

For (1), the denominator is N(s)kN(s)^{k} and the numerator is a polynomial in N(s)N(s) with leading term N(s)kN(s)^{k}. For (2), the numerator is a polynomial in N(s)N(s) of degree less than kk, and the denominator is N(s)kN(s)^{k}. ∎

Now, we note that

k!Ps(φ)Ps(φ)k!Ps(φ)+N(s)kk!(N(s)k).k!P_{s}(\varphi)\leq P^{*}_{s}(\varphi)\leq k!P_{s}(\varphi)+N(s)^{k}-k!{N(s)\choose k}.

Dividing by Ps=N(s)kP^{*}_{s}=N(s)^{k} and letting ss\to\infty, we get

limsk!Ps(φ)PslimsPs(φ)Pslimsk!Ps(φ)+N(s)kk!(N(s)k)Ps.\lim\limits_{s\to\infty}\frac{k!P_{s}(\varphi)}{P^{*}_{s}}\leq\lim\limits_{s\to\infty}\frac{P^{*}_{s}(\varphi)}{P^{*}_{s}}\leq\lim\limits_{s\to\infty}\frac{k!P_{s}(\varphi)+N(s)^{k}-k!{N(s)\choose k}}{P^{*}_{s}}.

Using the claim, we see that the right-hand side is

limsk!Ps(φ)k!Ps=limsPs(φ)Ps.\lim\limits_{s\to\infty}\frac{k!P^{*}_{s}(\varphi)}{k!P^{*}_{s}}=\lim\limits_{s\to\infty}\frac{P^{*}_{s}(\varphi)}{P^{*}_{s}}.

We can now phrase the questions we are interested in more formally.

Question 2.10.
  1. 1.

    Which varieties (with allowable presentations involving a fixed mm-tuple a¯\bar{a} of generators) satisfy the zero–one law?

  2. 2.

    Given that the zero–one law holds, when do the sentences with limiting density 11 match those true in the free structure?

2.3 Basic lemmas

Before we begin, we state three lemmas that hold very generally.

Lemma 2.11.

(φψ)(\varphi\vee\psi) has limiting density 0 if and only if φ\varphi and ψ\psi both have limiting density 0; in fact, this holds for any finite disjunction.

Proof.

We have Ps(φ)Ps,Ps(ψ)PsPs(φψ)PsPs(φ)Ps+Ps(ψ)Ps\frac{P_{s}(\varphi)}{P_{s}},\frac{P_{s}(\psi)}{P_{s}}\leq\frac{P_{s}(\varphi\vee\psi)}{P_{s}}\leq\frac{P_{s}(\varphi)}{P_{s}}+\frac{P_{s}(\psi)}{P_{s}}. From this, the lemma is clear. ∎

Lemma 2.12.

φ\varphi has limiting density 0 just in case ¬φ\neg{\varphi} has limiting density 11.

Proof.

We have 1=Ps(φ)Ps+Ps(¬φ)Ps1=\frac{P_{s}(\varphi)}{P_{s}}+\frac{P_{s}(\neg{\varphi})}{P_{s}}. Again, the lemma is clear. ∎

Lemma 2.13.

Let SS be the set of LL-sentences with limiting density 1. Then SS is consistent and is closed under logical implication—if φ1,,φnS\varphi_{1},\ldots,\varphi_{n}\in S and φ1,,φnψ\varphi_{1},\ldots,\varphi_{n}\vdash\psi, then ψS\psi\in S.

Proof.

Suppose SS is not consistent. By the Compactness Theorem, some finite subset is inconsistent. As every sentence in this set has density 1, there is a model of TT that realizes all these (finitely many) sentences, a contradiction. By the previous two lemmas, we have that each ¬φi\neg\varphi_{i} has limiting density 0, so ¬φi\bigvee\neg\varphi_{i} also has limiting density 0, and so φi\bigwedge\varphi_{i} has limiting density 1. But, φiψ\bigwedge\varphi_{i}\vdash\psi, so Ps(ψ)PsPs(φi)Ps=1\frac{P_{s}({\psi})}{P_{s}}\geq\frac{P_{s}(\bigwedge\varphi_{i})}{P_{s}}=1, and the lemma follows. ∎

3 Illustrative examples

In this section, we consider some of varieties and classes of presentations that illustrate different possibilities. First, we consider the variety of bijective structures, and presentations with a single generator and a single identity in which the function symbols occur only on the left. We show that the sentences true in the free structure are exactly those with limiting density 11. Second, we consider a variety of a single unary function and presentations with a single generator and a single identity. Here, we show that a specific sentence that is true in the free structure has limiting density 0. Next, we again consider bijective structures, and presentations with a single generator but with two identities. Here, we give sample sentences for which the limiting density does not exist. Finally, we consider the variety of abelian groups, and presentations with a single generator and a single relator. We give sentences for which the limiting density does not exist, and sentences for which the limiting density exists but is neither 0 nor 11.

3.1 Bijective structures

We start with the variety of bijective structures. Recall that the language consists of two unary function symbols S,S1S,S^{-1}. The axioms are

(x)SS1(x)=x and (x)S1S(x)=x.(\forall x)SS^{-1}(x)=x\mbox{\ \ and\ \ }(\forall x)S^{-1}S(x)=x\ .

These guarantee that the function SS is 111-1 and onto and that S1S^{-1} is the inverse of SS. Let TT be the theory generated by these axioms. The models consist of infinite \mathbb{Z}-chains and finite cycles m\mathbb{Z}_{m}. While these structures lack the mathematical interest and importance of groups, it is instructive to consider them because there are relatively simple elementary invariants, and for presentations with a single generator and a single identity, we can calculate the limiting densities for these sentences. It turns out that the analogue of Conjecture 1.2 holds.

Lemma 3.1.

Over the theory TT, every sentence is equivalent to a Boolean combination of sentences of the following basic types:

  1. 1.

    α(n,k)\alpha(n,k), saying that there are at least kk cycles of size nn,

  2. 2.

    β(n)\beta(n), saying that there is a chain of length at least nn.

Proof.

For any model 𝒜\mathcal{A} of TT, we have a natural equivalence relation \sim on the universe, where aba\sim b if Sm(a)=bS^{m}(a)=b for some integer mm. Each \sim-class is a copy of \mathbb{Z} or a finite cycle. The isomorphism type of 𝒜\mathcal{A} is determined by the number of \sim-classes of different types. Each model 𝒜\mathcal{A} of TT is elementarily equivalent to a saturated model 𝒜\mathcal{A}^{*}, where 𝒜\mathcal{A}^{*} has infinitely many copies of \mathbb{Z} if there is no finite bound on the sizes of the \sim-classes. From this, we see that the isomorphism type of 𝒜\mathcal{A}^{*} and the elementary first order theory of 𝒜\mathcal{A} are determined by the sentences α(n,k)\alpha(n,k) and β(n)\beta(n). ∎

We consider bijective structures with a single generator aa. There is a single \sim-class, which has the form \mathbb{Z}, an infinite chain, or m\mathbb{Z}_{m}, a cycle of size mm. We note that in either \mathbb{Z} or m\mathbb{Z}_{m}, all elements are automorphic. The following lemma is clear from the meanings of the sentences α(n,k)\alpha(n,k) and β(n)\beta(n).

Lemma 3.2.
  1. 1.

    For k>1k>1, α(n,k)\alpha(n,k) is false in both \mathbb{Z} and m\mathbb{Z}_{m},

  2. 2.

    α(n,1)\alpha(n,1) is true only in n\mathbb{Z}_{n},

  3. 3.

    β(n)\beta(n) is true in \mathbb{Z}; it is true in m\mathbb{Z}_{m} if and only if m>nm>n.

For models with a single generator, α(n,k)\alpha(n,k) is false for k>1k>1, and β(n)\beta(n) is equivalent to mn¬α(n,1)\bigwedge_{m\leq n}\neg{\alpha(n,1)}. Thus, it is enough to consider the elementary invariants of the form α(n,1)\alpha(n,1).


Here our presentations have a single identity, of the form t(a)=at(a)=a (function symbols occur only on the left). We may refer to the term t(a)t(a) as a relator. For a single relator t(a)t(a), we get \mathbb{Z} if for some kk, t(a)t(a) has kk occurrences of SS and kk occurrences of S1S^{-1}. We get m\mathbb{Z}_{m} if for some kk, t(a)t(a) has either m+km+k occurrences of SS and kk of S1S^{-1} or m+km+k occurrences of S1S^{-1} and kk of SS.


We will show that for all n1n\geq 1, α(n,1)\alpha(n,1) has limiting density 0. This implies that ¬α(n,1)\neg\alpha(n,1), which is true in the free structure, has limiting density 11. We will use two combinatorial lemmas. The first is an approximation for (2kk)\left(\begin{array}[]{c}2k\\ k\end{array}\right), good for large kk. The proof requires the use of Stirling’s formula on all three factorials (see the website of Das [5]).

Lemma 3.3.

(2kk)=(1+o(1))22kπk\left(\begin{array}[]{c}2k\\ k\end{array}\right)=(1+o(1))\frac{2^{2k}}{\sqrt{\pi k}}.

The second combinatorial lemma is an inequality.

Lemma 3.4.

For all n1n\geq 1 and all kk, 2(n+2kk)<(n+2(k+1)k+1)2\left(\begin{array}[]{c}n+2k\\ k\end{array}\right)<\left(\begin{array}[]{c}n+2(k+1)\\ k+1\end{array}\right).

Proof.

Recall Pascal’s Identity

(nk)+(nk+1)=(n+1k+1).\left(\begin{array}[]{c}n\\ k\end{array}\right)+\left(\begin{array}[]{c}n\\ k+1\end{array}\right)=\left(\begin{array}[]{c}n+1\\ k+1\end{array}\right).

We prove the inequality by induction on kk. First, for k=0k=0, the inequality just says that 2<n+22<n+2. Now, suppose k>0k>0. Applying Pascal’s Identity to the right side of the inequality, we get

(n+2(k+1)k+1)=(n+2k+1k)+(n+2k+1k+1)\left(\begin{array}[]{c}n+2(k+1)\\ k+1\end{array}\right)=\left(\begin{array}[]{c}n+2k+1\\ k\end{array}\right)+\left(\begin{array}[]{c}n+2k+1\\ k+1\end{array}\right)

and then

(n+2kk1)+(n+2kk)+(n+2kk)+(n+2kk+1).\left(\begin{array}[]{c}n+2k\\ k-1\end{array}\right)+\left(\begin{array}[]{c}n+2k\\ k\end{array}\right)+\left(\begin{array}[]{c}n+2k\\ k\end{array}\right)+\left(\begin{array}[]{c}n+2k\\ k+1\end{array}\right)\ .

This is clearly greater than 2(n+2kk)2\left(\begin{array}[]{c}n+2k\\ k\end{array}\right). ∎

To show that α(n,1)\alpha(n,1) has limiting density 0, we use several further lemmas.

Lemma 3.5.

Ps=2s+11P_{s}=2^{s+1}-1.

Proof.

The number of terms of length mm is 2m2^{m}, so the number of terms of length at most ss is 1+2++2s=2s+111+2+\ldots+2^{s}=2^{s+1}-1. ∎

Lemma 3.6.

Ps(α(m,1))=m+2ks2(m+2kk)P_{s}(\alpha(m,1))=\sum_{m+2k\leq s}2\left(\begin{array}[]{c}m+2k\\ k\end{array}\right).

Proof.

For each m1m\geq 1, and each kk, we have (m+2kk)\left(\begin{array}[]{c}m+2k\\ k\end{array}\right) terms with m+km+k occurrences of SS and kk of S1S^{-1}, and the same number with the symbols switched. ∎

The next lemma bounds the sum Pn+2k(α(n,1))P_{n+2k}(\alpha(n,1)) by a single term.

Lemma 3.7.

For all n1n\geq 1 and all k0k\geq 0, Pn+2k(α(n,1))<(n+2(k+1)k+1)P_{n+2k}(\alpha(n,1))<\left(\begin{array}[]{c}n+2(k+1)\\ k+1\end{array}\right).

Proof.

We fix nn and proceed by induction on kk. For k=0k=0, the left side is Pn(α(n,1))=(n0)=1P_{n}(\alpha(n,1))=\left(\begin{array}[]{c}n\\ 0\end{array}\right)=1, and the right side is (n+21)=n+2>1\left(\begin{array}[]{c}n+2\\ 1\end{array}\right)=n+2>1. Supposing that the statement holds for kk, we prove it for k+1k+1. By Lemma 3.4,

(n+2(k+2)k+2)>2(n+2(k+1)k+1)=(n+2(k+1)k+1)+(n+2(k+1)k+1).\left(\begin{array}[]{c}n+2(k+2)\\ k+2\end{array}\right)>2\left(\begin{array}[]{c}n+2(k+1)\\ k+1\end{array}\right)=\left(\begin{array}[]{c}n+2(k+1)\\ k+1\end{array}\right)+\left(\begin{array}[]{c}n+2(k+1)\\ k+1\end{array}\right)\ .

By the Induction Hypothesis, this is greater than

(n+2(k+1)k+1)+Pn+2k(φn)=Pn+2(k+1)(φn).\left(\begin{array}[]{c}n+2(k+1)\\ k+1\end{array}\right)+P_{n+2k}(\varphi_{n})=P_{n+2(k+1)}(\varphi_{n})\ .

Now we can show that the limiting density of α(n,1)\alpha(n,1) is 0 for any n1n\geq 1. To do so, we must make an odd/even case distinction because the only way to get presentations of different lengths of the same structure is for these lengths to differ by a multiple of two, so Pn+2k+1(α(n,1))P_{n+2k+1}(\alpha(n,1)) will equal Pn+2k(α(n,1))P_{n+2k}(\alpha(n,1)). However, if s=n+2ks=n+2k for some kk, then Ps(α(n,1))P_{s}(\alpha(n,1)) has a new last term (n+2kk)\left(\begin{array}[]{c}n+2k\\ k\end{array}\right), and Ps+1(α(n,1))=Ps(α(n,1))P_{s+1}(\alpha(n,1))=P_{s}(\alpha(n,1)). Therefore, it is enough to show that

Pn+2k(α(n,1))Pn+2k0.\frac{P_{n+2k}(\alpha(n,1))}{P_{n+2k}}\rightarrow 0\ .

By Lemma 3.4, the last term of Pn+2k(α(n,1))P_{n+2k}(\alpha(n,1)) is greater than the sum of the earlier terms. Thus, Pn+2k(α(n,1))<2(n+2kk)P_{n+2k}(\alpha(n,1))<2\left(\begin{array}[]{c}n+2k\\ k\end{array}\right). Recall that Pn+2k=2n+2k+11P_{n+2k}=2^{n+2k+1}-1, which is strictly greater than 2n+2k2^{n+2k}, so Pn+2k(α(n,1))Pn+2k<22n+2k(n+2kk)\frac{P_{n+2k}(\alpha(n,1))}{P_{n+2k}}<\frac{2}{2^{n+2k}}\left(\begin{array}[]{c}n+2k\\ k\end{array}\right). To prove that the limiting density of α(n,1)\alpha(n,1) is 0, it is enough to prove that

22n+2k(n+2kk)0.\frac{2}{2^{n+2k}}\left(\begin{array}[]{c}n+2k\\ k\end{array}\right)\rightarrow 0.

We can express 22n+2k(n+2kk)\frac{2}{2^{n+2k}}\left(\begin{array}[]{c}n+2k\\ k\end{array}\right) as a product of two factors, one involving the fixed nn and the other not. The first factor is

22n(2k+nk+n)(2k+(n1)k+(n1))(2k+1k+1).\frac{2}{2^{n}}\left(\frac{2k+n}{k+n}\right)\left(\frac{2k+(n-1)}{k+(n-1)}\right)\ldots\left(\frac{2k+1}{k+1}\right)\ .

This is an (n+1)(n+1)-fold product with limit 22 as kk\rightarrow\infty. The second factor is 122k(2kk)\frac{1}{2^{2k}}\left(\begin{array}[]{c}2k\\ k\end{array}\right). By Lemma 3.3 above, this is (1+o(1))1πk(1+o(1))\frac{1}{\sqrt{\pi k}}, which has limit 0. All together, we have shown the following.

Proposition 3.8.

For n1n\geq 1, limsPs(α(n,1))Ps=0\lim_{s\rightarrow\infty}\frac{P_{s}(\alpha(n,1))}{P_{s}}=0.

From this, we get the following.

Theorem 3.9.

For bijective structures with a single generator and a single relator, each sentence φ\varphi has limiting density equal to 11 if φ\ \mathbb{Z}\models\varphi and 0\ 0 otherwise.

For later use, we state below another immediate consequence of Proposition 3.8. For a term t(a)t(a), let XX be the difference between the number of occurrences of SS and the number of occurrences of S1S^{-1} in t(a)t(a). The relators that make α(n,1)\alpha(n,1) true are exactly those for which |X|=n|X|=n.

Lemma 3.10.

For each kk\in\mathbb{Z}, the set of presentations a,t(a)=aa,t(a)=a for which X=kX=k has density 0.

3.2 Structures with a single unary function

Next, take the language with a single unary function symbol ff, and the variety with no axioms. We consider presentations with a single generator aa and a single identity. At first, we focus on the sentence φ\varphi saying that ff is not 111-1. We will see that this sentence is false in the free structure, but it has limiting density 11. After that, we will consider arbitrary sentences, and prove a zero-one law.

For an identity f(r)(a)=f(r)(a)f^{(r)}(a)=f^{(r^{\prime})}(a), where r,r0r,r^{\prime}\geq 0, the length is r+rr+r^{\prime}. If r=rr=r^{\prime}, then we get an ω\omega-chain. If 0<r<r0<r<r^{\prime}, then we get a chain of length rr leading to a cycle of length rrr^{\prime}-r. If 0=r<r0=r<r^{\prime}, then we get a cycle of length rr^{\prime}. Similarly, if 0<r<r0<r^{\prime}<r, we get a chain of length rr^{\prime} leading to a cycle of length rrr-r^{\prime}, and if 0=r<r0=r^{\prime}<r, we get a cycle of length rr.

We write m+nm+\mathbb{Z}_{n} for a chain of length mm leading to a cycle of length nn, allowing the possibility that m=0m=0. Any structure in our variety obtained from a single generator and a single identity has one of the following forms:

  1. 1.

    an ω\omega-chain—this is the free structure,

  2. 2.

    a finite chain leading to a finite cycle, or

  3. 3.

    a finite cycle.

Lemma 3.11.

Ps=12(s2+3s+2)P_{s}=\frac{1}{2}(s^{2}+3s+2).

Proof.

The number of identities of length mm is m+1m+1. Then
Ps=1+2++(s+1)=12(s+2)(s+1)=12(s2+3s+2)P_{s}=1+2+\ldots+(s+1)=\frac{1}{2}(s+2)(s+1)=\frac{1}{2}(s^{2}+3s+2). ∎

Recall that φ\varphi holds if the function is not 111-1. We can see that φ\varphi is false in ω\omega and the finite cycles n\mathbb{Z}_{n}, and true in the structures m+nm+\mathbb{Z}_{n} for m1m\geq 1, so an identity fn(a)=fn(a)f^{n}(a)=f^{n^{\prime}}(a) yields a structure in which φ\varphi is false if n=nn=n^{\prime} or if one of n,nn,n^{\prime} is 0, and true otherwise. We can easily count the identities of length mm that make φ\varphi false. For m=0m=0, there is just one identity, and it makes φ\varphi false. For m=1m=1, there are two identities, and both make φ\varphi false. For m2m\geq 2, if mm is odd, there are just two identities that make φ\varphi false, and if mm is even, there are three identities that make φ\varphi false. Thus, Ps(¬φ)=1+2+3+2++(52+12(1)s)P_{s}(\neg{\varphi})=1+2+3+2+\ldots+(\frac{5}{2}+\frac{1}{2}(-1)^{s}).

Proposition 3.12.

The limiting density for φ\varphi is 11.

Proof.

For s1s\geq 1, Ps(¬φ)3sP_{s}(\neg{\varphi})\leq 3s, and lims3sPs=0\lim_{s\rightarrow\infty}\frac{3s}{P_{s}}=0. Therefore, ¬φ\neg{\varphi} has limiting density 0, and φ\varphi has limiting density 11. ∎

m=3m=3n=4n=4
Figure 1: m+nm+\mathbb{Z}_{n} where m=3m=3 and n=4n=4

In fact, we will see that each sentence has limiting density 11 or 0. The set of sentences with limiting density 11 is not that of the free structure. It is the theory of a structure 𝒜\mathcal{A} that we may think of as a limit of the structures m+nm+\mathbb{Z}_{n} (see Figure 1). The limit structure 𝒜\mathcal{A} consists of an ω\omega-chain together with two ω\omega^{*}-chains and a single ω\omega-chain that come together at a special point—this point is the end of the two ω\omega^{*}-chains and the beginning of the ω\omega-chain (see Figure 2). The chain of length mm is replaced, in the limit, by an ω\omega-chain plus one of the ω\omega^{*}-chains, and the nn-cycle is replaced, in the limit, by the other ω\omega-chain and ω\omega^{*}-chain. We note that 𝒜\mathcal{A} is not finitely generated.

ω\omegaω\omega^{*}ω\omega^{*}ω\omega
Figure 2: The limit structure
Lemma 3.13.

The theory of 𝒜\mathcal{A} is generated by the following sentences; note that the elements aa and cc are defined and not named by constants.

  1. 1.

    ψa\psi_{a}, saying that there is a unique element aa with no ff-pre-image,

  2. 2.

    ψc\psi_{c}, saying that there is a unique element cc with two ff-pre-images,

  3. 3.

    ψ\psi, saying that there is no element with more than two ff-pre-images,

  4. 4.

    for every nωn\in\omega, αn\alpha_{n} saying that there is no cycle of length nn, and

  5. 5.

    for every nωn\in\omega, βn\beta_{n} saying that aa and cc are not connected by a chain of length nn.

Proof.

In any structure in our variety, there is an equivalence relation \sim, where xyx\sim y if there is a finite sequence x0,,xnx_{0},\ldots,x_{n} such that x=x0x=x_{0}, y=xny=x_{n}, and for each i<ni<n, either f(xi)=xi+1f(x_{i})=x_{i+1} or f(xi+1)=xif(x_{i+1})=x_{i}. If \mathcal{B} and 𝒞\mathcal{C} are two saturated models of the sentences above, of the same cardinality, then 𝒞\mathcal{B}\cong\mathcal{C}. We map the special elements aa and cc in \mathcal{B} to the corresponding elements of 𝒞\mathcal{C}. The \sim-class of aa is an ω\omega-chain that does not include cc. The \sim-class of cc has two ω\omega^{*}-chains ending in cc and one ω\omega-chain starting with cc. The other \sim-classes, not containing aa or cc, are \mathbb{Z}-chains. There are the same number of these in \mathcal{B} and 𝒞\mathcal{C}. ∎

Lemma 3.14.

The sentences above that generate the theory of 𝒜\mathcal{A} all have limiting density 11.

Proof.

The sentence ψ\psi, saying that there is no element with more than two ff-pre-images, is true in all of the models that we get from a single generator aa and a single identity, so this has limiting density 11. We have seen that the set of presentations that give models in which ff is not 111-1 has limiting density 11. The models have the form m+nm+\mathbb{Z}_{n} for m,n>0m,n>0. The sentences ψa\psi_{a}, saying that there is a unique element aa with no ff-pre-image, ψc\psi_{c}, saying that there is a unique element cc with two ff-pre-images, and ψ\psi are true in all of these models, so these sentences have limiting density 11.

Consider αn\alpha_{n}, saying that there is no cycle of length nn. Let BB be the set of presentations that make αn\alpha_{n} false—the resulting model has a cycle of length nn. These identities have the form fn+k(a)=fk(a)f^{n+k}(a)=f^{k}(a) or fk(a)=fn+k(a)f^{k}(a)=f^{n+k}(a). The number of such identities of length mm is 0 if mnm-n is odd and 22 if mnm-n is even. Then Ps(αn)2sP_{s}(\alpha_{n})\leq 2s. Since Ps=O(s2)P_{s}=O(s^{2}), the limiting density of BB is 0, so the density of αn\alpha_{n} is 11. Finally, consider βn\beta_{n}, saying that aa and cc are not connected by a chain of length nn. Let CC be the set of presentations that make this false. These identities have the form fn(a)=fn+k(a)f^{n}(a)=f^{n+k}(a) or fn+k(a)=fn(a)f^{n+k}(a)=f^{n}(a). The number of such identities of length mm is 22 for all m2nm\geq 2n. Then Ps(C)2sP_{s}(C)\leq 2s. Again the limiting density of CC is 0, so the density of βn=1\beta_{n}=1. ∎

We conclude the following.

Proposition 3.15.

For all sentences φ\varphi true in 𝒜\mathcal{A}, the limiting density is 11.

3.3 Bijective structures with two identities

In the next example, we return to the variety of bijective structures as in Section 3.1. As before, our presentations have a single generator aa, but there are two identities instead of just one. We will show that the limiting density need not exist. Recall that the language consists of unary function symbols S,S1S,S^{-1} and that the axioms say that SS and S1S^{-1} are inverses. The identities have the form t(a)=at(a)=a, with function symbols only on the left. Each identity is thus equivalent to one of the form Sm(a)=aS^{m}(a)=a, where mm\in\mathbb{Z}.

Proposition 3.16.

For bijective structures, and presentations with a single generator aa and two identities, the sentence φ\varphi saying that the structure is a 11-cycle does not have a limiting density.

The proof is somewhat involved. We begin with some elementary lemmas, but eventually we will consider a random walk on a group and appeal to results from random group theory that depend on the Central Limit Theorem. The lemma below tells us when the sentence φ\varphi is true.

Lemma 3.17.

Let 𝒜\mathcal{A} be the structure given by an unordered set consisting of two identities, equivalent to Sm(a)=aS^{m}(a)=a and Sm(a)=aS^{m^{\prime}}(a)=a. Then 𝒜\mathcal{A} is a 11-cycle if and only if GCD(m,m)=1GCD(m,m^{\prime})=1.

Proof.

Note that Sk(a)=aS^{k}(a)=a if and only if a=Sk(a)a=S^{-k}(a). Thus, we may suppose that both m,m0m,m^{\prime}\geq 0. First, suppose that GCD(m,m)=1GCD(m,m^{\prime})=1. In this case, there are r,sr,s\in\mathbb{Z} such that mr+ms=1mr+m^{\prime}s=1. Then we have

S(a)=Smr+ms(a)=Smr(Sms(a))=SmSmSmSm(a)=a,S(a)=S^{mr+m^{\prime}s}(a)=S^{mr}(S^{m^{\prime}s}(a))=S^{m}\circ\cdots\circ S^{m}\circ S^{m^{\prime}}\circ\cdots\circ S^{m^{\prime}}(a)=a,

so 𝒜\mathcal{A} is a 11-cycle. Now, suppose that 𝒜\mathcal{A} is a 11-cycle, and let GCD(m,m)=dGCD(m,m^{\prime})=d. The axioms of TT and the identities Sm(a)=aS^{m}(a)=a and Sm(a)=aS^{m^{\prime}}(a)=a are both satisfied in a dd-cycle, and 𝒜\mathcal{A} can only be a 11-cycle if d=1d=1. ∎

Our presentations have two identities, but we also need some facts about single identities. We indicate with that we are considering single identities, writing P=mP^{\prime}_{=m} for the number of identities of length mm and PsP^{\prime}_{s} for the number of length at most ss, and writing P=m(B)P^{\prime}_{=m}(B), Ps(B)P^{\prime}_{s}(B) for the number of these identities in a set BB. We reserve PsP_{s} for the number of unordered pairs of identities of length at most ss, and we write Ps(B2)P_{s}(B^{2}) for the number such that both identities are in BB.

For a single identity of the form t(a)=at(a)=a, let XX be the difference between the number of occurrences of SS and the number of occurrences of S1S^{-1} in tt. Intuition may suggest that the statement n|Xn|X should have limiting density 1n\frac{1}{n}. This turns out to be true for odd nn. However, for n=2n=2, we find that the limiting density for the statement 2|X2|X does not exist. Essentially, the reason is that the last term of Ps(2|X)=msP=m(2|X)P^{\prime}_{s}(2|X)=\sum_{m\leq s}P^{\prime}_{=m}(2|X) may be greater than the sum of all earlier terms, and this term depends on the parity of ss. The lemma below says what happens to P=s(2|X)P_{=s}(2|X) as the parity of ss changes.

Lemma 3.18.
  • For even mm, all identities of length mm satisfy 2X2\mid X; none satisfies 2X2\nmid X.

  • For odd mm, all identities of length mm satisfy 2X2\nmid X and none satisfies 2X2\mid X.

Proof.

For m=0m=0, there is just one identity, and for this identity, X=0X=0. Supposing that the statements hold for mm, if tt has length mm, then tt has two extensions of length m+1m+1, and the parity of XX changes. ∎

The next lemma gives the proportion of single identities of length at most ss for which 2|X2|X holds. The value depends on the parity of ss.

Lemma 3.19.
  1. 1.

    limsP2s(2X)P2s=23\displaystyle\lim\limits_{s\rightarrow\infty}\frac{P^{\prime}_{2s}(2\mid X)}{P^{\prime}_{2s}}=\frac{2}{3} and limsP2s(2X)P2s=13\displaystyle\lim\limits_{s\rightarrow\infty}\frac{P^{\prime}_{2s}(2\nmid X)}{P^{\prime}_{2s}}=\frac{1}{3}.

  2. 2.

    limsP2s+1(2X)P2s+1=13\displaystyle\lim_{s\rightarrow\infty}\frac{P^{\prime}_{2s+1}(2\mid X)}{P^{\prime}_{2s+1}}=\frac{1}{3} and limsP2s+1(2X)P2s+1=23\displaystyle\lim_{s\rightarrow\infty}\frac{P^{\prime}_{2s+1}(2\nmid X)}{P^{\prime}_{2s+1}}=\frac{2}{3}.

Proof.

The calculation is based on Lemma 3.18. For (1), the even case, we have P2s=22s+11=4s21P^{\prime}_{2s}=2^{2s+1}-1=4^{s}\cdot 2-1, and

P2s(2X)=ms22m=ms4m=4s+113.P^{\prime}_{2s}(2\mid X)=\sum_{m\leq s}2^{2m}=\sum_{m\leq s}4^{m}=\frac{4^{s+1}-1}{3}.

Then P2s(2X)P2s=4s+113(4s21)23\displaystyle\frac{P^{\prime}_{2s}(2\mid X)}{P^{\prime}_{2s}}=\frac{4^{s+1}-1}{3(4^{s}\cdot 2-1)}\rightarrow\frac{2}{3}.


For (2), the odd case, we have P2s+1=22s+21=4s+11P^{\prime}_{2s+1}=2^{2s+2}-1=4^{s+1}-1 and

P2s+1(2X)=P2s(2X)=4s+113.P^{\prime}_{2s+1}(2\mid X)=P^{\prime}_{2s}(2\mid X)=\frac{4^{s+1}-1}{3}.

Then P2s+1(2X)P2s+1=4s+113(4s+11)=13\displaystyle\frac{P^{\prime}_{2s+1}(2\mid X)}{P^{\prime}_{2s+1}}=\frac{4^{s+1}-1}{3\cdot(4^{s+1}-1)}=\frac{1}{3}. ∎

So far, the lemmas have involved only elementary calculations. The next result is from random group theory [4, 19], concerning a random walk on a group. The elements of the group G=nG=\mathbb{Z}_{n} represent the possible remainders after division of an integer zz by nn. In general, for a random walk, there are finitely many states, and given just the current state ss, with no more prior history, we have fixed probabilities of passing next to state ss^{\prime}. We allow s=ss^{\prime}=s.

Our states are group elements. We write μ(g)\mu(g) for the probability of going in one step from the identity to gg, and we write μ(k)(g)\mu^{(k)}(g) for the probability of going in kk steps from the identity to gg. For the result below, the probability measure μ\mu, defined on GG, is supported on a special generating set Σ\Sigma. For μ\mu to be supported on Σ\Sigma means that μ\mu assigns non-zero probability to the elements of Σ\Sigma, i.e., the set Σ\Sigma will consist of the group elements reachable from the identity in one step. For any nn, Σ\Sigma is also the set of differences ggg^{\prime}-g, where gg^{\prime} is a successor of gg reachable in one step. The values of μk(g)\mu^{k}(g), for k>0k>0 are obtained by considering the tree of 11-step extensions of length kk starting from the identity. We multiply probabilities along the paths, and then sum over the paths leading to gg.

The result below tells us that the probability of each remainder gg in n\mathbb{Z}_{n} is approximately 1n\frac{1}{n}, and that the convergence (as kk\rightarrow\infty) has a great deal of uniformity.

Theorem 3.20 ([19, Theorem 7.3]).

There exist α,β>0\alpha,\beta>0 such that for any group GG of the form n\mathbb{Z}_{n}, any generating set Σ\Sigma containing the group identity element, and a probability measure μ\mu supported on Σ\Sigma, we have that for all gGg\in G and all kωk\in\omega,

|μ(k)(g)1n|<αe(βkn2)\left|\mu^{(k)}(g)-\frac{1}{n}\right|<\alpha e^{-(\frac{\beta k}{n^{2}})}

To adapt this theorem to our setting, we will consider an odd nn and identities of even length mm. We break the identity into pieces of length 2, so each piece has 1/41/4 chance of being each of SSSS, S1S1S^{-1}S^{-1}, SS1SS^{-1}, or S1SS^{-1}S. These correspond to 2,22,-2, and 0 in the random walk on n\mathbb{Z}_{n}, and when nn is odd, Σ={2,0,2}\Sigma=\{-2,0,2\} generates n\mathbb{Z}_{n}. Just as the identities of even length approach a uniform distribution, so do the identities of odd length, and, consequently, so do the identities of length at most ss. We have the following:

Corollary 3.21.

For any odd number nn and any ss,

limsPs(nX)Ps=1n and limsPs(nX)Ps=n1n.\displaystyle\lim_{s\to\infty}\frac{P^{\prime}_{s}(n\mid X)}{P^{\prime}_{s}}=\frac{1}{n}\text{ and }\displaystyle\lim_{s\to\infty}\frac{P^{\prime}_{s}(n\nmid X)}{P^{\prime}_{s}}=\frac{n-1}{n}.

Our presentations have an unordered set of two identities. However, it is easier to count ordered pairs, allowing repetition. By Proposition 2.9, we get the same limiting densities, so we will count ordered pairs allowing repetition of elements instead. Let CC be the set of presentations in which the difference functions X1,X2X_{1},X_{2} are both divisible by some prime pp. It follows from Lemma 3.10 that X1=0X_{1}=0, and X2=0X_{2}=0 both have limiting density 0. The important part of CC consists of the presentations such that X10X_{1}\not=0 and X20X_{2}\not=0, and in what follows, we write CC for this important part. For the presentations in CC, there is some prime that divides both X1X_{1} and X2X_{2}, and both X1,X2X_{1},X_{2} are non-zero.

Definition 3.22.

For each ss, let psp_{s} be the greatest prime pp such that
23pln(s)2\cdot 3\cdots p\leq ln(s).

For every ss, we split CC into two parts, C1=Cs,1C_{1}=C_{s,1} and C2=Cs,2C_{2}=C_{s,2}. Note that this splitting depends on ss. For a presentation in CC, let dd be the least prime that divides both X1,X2X_{1},X_{2}. Then the presentation is in Cs,1C_{s,1} if dpsd\leq p_{s} and in Cs,2C_{s,2} if d>psd>p_{s}. We will show that Cs,2C_{s,2} has limiting density limsPs(Cs,2)Ps=0\lim\limits_{s\to\infty}\frac{P_{s}(C_{s,2})}{P_{s}}=0 and that the limiting density of C1C_{1} does not exist—it toggles between two values, one for even ss and the other for odd ss. Among the primes, 22 behaves differently from the odd primes. We have shown that, for single identities with difference XX, the limiting density of 2|X2|X does not exist. We will see later that this explains why the limiting density of C1C_{1} does not exist.


Our first goal is to show that C2C_{2} has limiting density 0. Toward this, we consider a single identity with difference XX.

Lemma 3.23.

For each odd prime pp and all ss, Ps(p|X&X0)Ps2p+1\displaystyle\frac{P^{\prime}_{s}(p|X\ \&\ X\not=0)}{P^{\prime}_{s}}\leq\frac{2}{p+1}.

Proof.

We will first prove that P=m(p|X&X0)P=m2p+1\displaystyle\frac{P^{\prime}_{=m}(p|X\ \&\ X\not=0)}{P^{\prime}_{=m}}\leq\frac{2}{p+1} for all mm. Note that the numbers P=m(X=n)P^{\prime}_{=m}(X=n), for mnm-m\leq n\leq m, form a Pascal triangle. At the top, for m=0m=0, we have 11, corresponding to X=0X=0. For m=1m=1, we have 11’s corresponding to X=±1X=\pm 1. In general, for even mm, XX takes the even values nn in the interval [m,m][-m,m], and for odd mm, XX takes the odd values in the interval [m,m][-m,m]. In both cases, P=m+1(X=n)=P=m(X=n1)+P=m(X=n+1)P^{\prime}_{=m+1}(X=n)=P^{\prime}_{=m}(X=n-1)+P^{\prime}_{=m}(X=n+1). We can see that P=m(X=n)P^{\prime}_{=m}(X=n) decreases as |n||n| increases, and that P=m(X=n)=P=m(X=n)P^{\prime}_{=m}(X=n)=P^{\prime}_{=m}(X=-n).


For odd mm (so that XX is odd), we have

P=m(X=±1)P=m(X=±3)P=m(X=±p)P^{\prime}_{=m}(X=\pm 1)\geq P^{\prime}_{=m}(X=\pm 3)\geq\cdots\geq P^{\prime}_{=m}(X=\pm p)\ \,
P=m(X=±(p+2))P=m(X=±(p+4))P=m(X=±3p)P^{\prime}_{=m}(X=\pm(p+2))\geq P^{\prime}_{=m}(X=\pm(p+4))\geq\cdots\geq P^{\prime}_{=m}(X=\pm 3p)
\cdots

Note that there are pp terms in each of the lines, except the first line, which has only p+12\frac{p+1}{2} terms. Therefore, we have

1=n:oddP=m(X=n)P=m\displaystyle 1=\sum\limits_{n:\text{odd}}\frac{P^{\prime}_{=m}(X=n)}{P^{\prime}_{=m}} =P=m(X=±1)P=m+P=m(X=±3)P=m+\displaystyle=\frac{P^{\prime}_{=m}(X=\pm 1)}{P^{\prime}_{=m}}+\frac{P^{\prime}_{=m}(X=\pm 3)}{P^{\prime}_{=m}}+\cdots
(p+12)P=m(X=±p)P=m+pP=m(X=±3p)P=m+\displaystyle\geq\left(\frac{p+1}{2}\right)\frac{P^{\prime}_{=m}(X=\pm p)}{P^{\prime}_{=m}}+p\cdot\frac{P^{\prime}_{=m}(X=\pm 3p)}{P^{\prime}_{=m}}+\cdots
(p+12)(P=m(X=±p)P=m+P=m(X=±3p)P=m+)\displaystyle\geq\left(\frac{p+1}{2}\right)\left(\frac{P^{\prime}_{=m}(X=\pm p)}{P^{\prime}_{=m}}+\frac{P^{\prime}_{=m}(X=\pm 3p)}{P^{\prime}_{=m}}+\cdots\right)
=(p+12)P=m(pX&X0)P=m.\displaystyle=\left(\frac{p+1}{2}\right)\frac{P^{\prime}_{=m}(p\mid X\ \&\ X\neq 0)}{P^{\prime}_{=m}}.

If mm is even (so that XX is even), then we have

P=m(X=±2)P=m(X=±4)P=m(X=±2p)P^{\prime}_{=m}(X=\pm 2)\geq P^{\prime}_{=m}(X=\pm 4)\geq\cdots\geq P^{\prime}_{=m}(X=\pm 2p)
P=m(X=±(2p+2))P=m(X=±(2p+4))P=m(X=±4p)P^{\prime}_{=m}(X=\pm(2p+2))\geq P^{\prime}_{=m}(X=\pm(2p+4))\geq\cdots\geq P^{\prime}_{=m}(X=\pm 4p)
\cdots

In this case, each line has pp terms, and we get the following slightly stronger inequality:

1=n:evenP=m(X=n)P=m\displaystyle 1=\sum\limits_{n:\text{even}}\frac{P^{\prime}_{=m}(X=n)}{P^{\prime}_{=m}} =P=m(X=0)P=m+P=m(X=±2)P=m+P=m(X=±4)P=m+\displaystyle=\frac{P^{\prime}_{=m}(X=0)}{P^{\prime}_{=m}}+\frac{P^{\prime}_{=m}(X=\pm 2)}{P^{\prime}_{=m}}+\frac{P^{\prime}_{=m}(X=\pm 4)}{P^{\prime}_{=m}}+\cdots
P=m(X=±2)P=m+P=m(X=±4)P=m+\displaystyle\geq\frac{P^{\prime}_{=m}(X=\pm 2)}{P^{\prime}_{=m}}+\frac{P^{\prime}_{=m}(X=\pm 4)}{P^{\prime}_{=m}}+\cdots
pP=m(X=±2p)P=m+pP=m(X=±4p)P=m+\displaystyle\geq p\cdot\frac{P^{\prime}_{=m}(X=\pm 2p)}{P^{\prime}_{=m}}+p\cdot\frac{P^{\prime}_{=m}(X=\pm 4p)}{P^{\prime}_{=m}}+\cdots
=pP=m(pX&X0)P=m.\displaystyle=p\cdot\frac{P^{\prime}_{=m}(p\mid X\ \&\ X\neq 0)}{P^{\prime}_{=m}}.

Combining the even and odd case, we get the desired P=m(p|X&X0)P=m2p+1\frac{P^{\prime}_{=m}(p|X\ \&\ X\not=0)}{P^{\prime}_{=m}}\leq\frac{2}{p+1}.

Now, we turn our attention back to the inequality in the lemma, which concerns identities up to a certain length. The quotient Ps(pX&X0)Ps\frac{P^{\prime}_{s}(p\mid X\ \&\ X\neq 0)}{P^{\prime}_{s}} is a weighted average (weighted by the proportion of identities of each length) of the probabilities P=m(pX&X0)P=m\frac{P^{\prime}_{=m}(p\mid X\ \&\ X\neq 0)}{P^{\prime}_{=m}}, where msm\leq s. Thus, the lemma follows from the inequality on identities of a fixed length P=m(p|X&X0)P=m2p+1\frac{P^{\prime}_{=m}(p|X\ \&\ X\not=0)}{P^{\prime}_{=m}}\leq\frac{2}{p+1}. ∎

We are now ready to consider both identities.

Lemma 3.24.

limsPs(C2)Ps=0\displaystyle\lim\limits_{s\to\infty}\frac{P_{s}(C_{2})}{P_{s}}=0.

Proof.

Below, we will appeal to Proposition 2.9 and consider, for each ss, the probability space consisting of the ordered pairs of identities, each of length at most ss. Then the random variables X1,X2X_{1},X_{2} are independent. Counting ordered pairs of identities and allowing repetition, we see that for each ss,

Ps(C2)\displaystyle P_{s}(C_{2}) p>psPs(X1,X20&p|X1&p|X2)\displaystyle\leq\sum\limits_{p>p_{s}}P_{s}(X_{1},X_{2}\not=0\ \&\ p|X_{1}\ \&\ p|X_{2})
=p>psPs(X10&p|X1)Ps(X20&p|X2).\displaystyle=\sum\limits_{p>p_{s}}P^{\prime}_{s}(X_{1}\neq 0\ \&\ p|X_{1})P^{\prime}_{s}(X_{2}\neq 0\ \&\ p|X_{2}).

So, it follows from the previous lemma that

Ps(C2)Ps\displaystyle\frac{P_{s}(C_{2})}{P_{s}} p>psPs(X10&p|X1)PsPs(X20&p|X2)Ps\displaystyle\leq\sum\limits_{p>p_{s}}\frac{P^{\prime}_{s}(X_{1}\neq 0\ \&\ p|X_{1})}{P^{\prime}_{s}}\frac{P^{\prime}_{s}(X_{2}\neq 0\ \&\ p|X_{2})}{P^{\prime}_{s}}
p>ps(2p+1)2\displaystyle\leq\sum_{p>p_{s}}\left(\frac{2}{p+1}\right)^{2}

By a well-known fact from number theory, the sum of the squares of the reciprocals of primes (or of all natural numbers) converges. Since limsps=\lim\limits_{s\to\infty}p_{s}=\infty, we have that limsp>ps(2p+1)2=0\lim\limits_{s\to\infty}\sum\limits_{p>p_{s}}(\frac{2}{p+1})^{2}=0. Thus, C2C_{2} has limiting density 0. ∎

We turn to C1C_{1}. Again, we consider first a single identity.

Lemma 3.25.

We write DsD_{s} for the set of identities of length at most ss but greater than s\sqrt{s}. Then Ps(Ds)Ps1\frac{P^{\prime}_{s}(D_{s})}{P^{\prime}_{s}}\rightarrow 1.

Proof.

We have Ps(Ds)=PsPsP^{\prime}_{s}(D_{s})=P^{\prime}_{s}-P^{\prime}_{\sqrt{s}}, and Ps(Ds)Ps=12s+112s+111\frac{P^{\prime}_{s}(D_{s})}{P^{\prime}_{s}}=1-\frac{2^{\sqrt{s}+1}-1}{2^{s+1}-1}\rightarrow 1. ∎

Lemma 3.25 may be interpreted as saying that most identities of length at most ss have length at least s\sqrt{s}. We write Ps(Ds2)P_{s}(D_{s}^{2}) for the number of pairs of identities of length at most ss such that both have length at least s\sqrt{s}. The next lemma says that for most pairs of identities of length at most ss, the length of both is at least s\sqrt{s}.

Lemma 3.26.

limsPs(Ds2)Ps1\lim_{s\rightarrow\infty}\frac{P_{s}(D_{s}^{2})}{P_{s}}\rightarrow 1.

Now, Ps(C1)Ps\frac{P_{s}(C_{1})}{P_{s}} is the probability that, among pairs of identities of length at most ss, with difference functions X1X_{1} and X2X_{2}, there is some prime ppsp\leq p_{s} such that p|X1&p|X2p|X_{1}\ \&\ p|X_{2}. We may suppose that both identities have length greater than s\sqrt{s}. We have seen that the limiting probability that 22 divides both X1,X2X_{1},X_{2} does not exist—for even ss, it approaches 49\frac{4}{9}, while for odd ss, it approaches 19\frac{1}{9} 3.21. Here, we consider odd primes. We have justified thinking of the random variables X1,X2X_{1},X_{2} (for identities of length at most ss) as independent.

We would like to assume that for i=1,2i=1,2, the events p|Xip|X_{i} for different primes pp are independent. This turns out to be “approximately” true. The probability that X1,X2X_{1},X_{2} are not both divisible by 33 is approximately 11321-\frac{1}{3^{2}}. The probability that X1,X2X_{1},X_{2} are not both divisible by 33 and not both divisible by 55 is approximately (1132)(1152)(1-\frac{1}{3^{2}})(1-\frac{1}{5^{2}}). The probability that X1,X2X_{1},X_{2} are not both divisible by any odd prime ppsp\leq p_{s} is approximately 3pps(11p2)\prod\limits_{3\leq p\leq p_{s}}(1-\frac{1}{p^{2}}). This formula matches what we would get by laborious inclusion-exclusion counting.

In fact, the divisibilities of XiX_{i} by different primes may not be independent. However, we can apply the Chinese Remainder Theorem and consider the residue of XiX_{i} modulo Ns=ppspN_{s}=\prod\limits_{p\leq p_{s}}p, which is 23ppsp2\prod\limits_{3\leq p\leq p_{s}}p. It follows from the Definition of psp_{s} (Definition 3.22) that Nsln(s)N_{s}\leq\ln(s). This is where the random walk on the group comes in. We will use Theorem 3.20.

Theorem 3.27.

Below, we let pp range over all primes:

  1. 1.

    P2s(Cs,1c)P2s593p(11p2)\displaystyle\frac{P_{2s}(C_{s,1}^{c})}{P_{2s}}\rightarrow\frac{5}{9}\cdot\prod_{3\leq p}\left(1-\frac{1}{p^{2}}\right).

  2. 2.

    P2s+1(Cs,1c)P2s+1893p(11p2)\displaystyle\frac{P_{2s+1}(C_{s,1}^{c})}{P_{2s+1}}\rightarrow\frac{8}{9}\cdot\prod_{3\leq p}\left(1-\frac{1}{p^{2}}\right).

Proof.

We prove (1). Take α,β>0\alpha,\beta>0 as in Theorem 3.20. Recall that X1X_{1} and X2X_{2} are the difference functions associated with the first and second identities, where in each identity, the function symbols are all on the left. Fixing ss, we consider the residue of X1X_{1} and X2X_{2} modulo Ns=23ppspN_{s}=2\prod\limits_{3\leq p\leq p_{s}}p, where NsN_{s} is at most ln(s)\ln(s). For a single identity, we consider a string tt of length kk. For the fixed ss, let GsG_{s} be the group of possible remainders after division of X1X_{1} by NsN_{s}. Theorem 3.20 tells us that for every 0a<Ns0\leq a<N_{s} and every kk,

|P=k(X1=a(modNs))P=k2Ns|<αeβkNs2 if a and k have the same parity, and\left|\frac{P^{\prime}_{=k}(X_{1}=a\pmod{N_{s}})}{P^{\prime}_{=k}}-\frac{2}{N_{s}}\right|<\alpha e^{\frac{-\beta k}{N_{s}^{2}}}\text{ if $a$ and $k$ have the same parity, and}
P=k(X1=a(modNs))P=k=0 if a and k have different parities.\frac{P^{\prime}_{=k}(X_{1}=a\pmod{N_{s}})}{P^{\prime}_{=k}}=0\text{ if $a$ and $k$ have different parities}.

When we sum up the identities of length at most ss, the previous lemma says that most of them will have length some ksk\geq\sqrt{s}. Thus, we may assume that ksk\geq\sqrt{s}. The previous inequality yields

|P=k(X1=a(modNs))P=k2Ns|<αeβkNs2<αeβsln(s)2.\left|\frac{P^{\prime}_{=k}(X_{1}=a\pmod{N_{s}})}{P^{\prime}_{=k}}-\frac{2}{N_{s}}\right|<\alpha e^{\frac{-\beta k}{N_{s}^{2}}}<\alpha e^{\frac{-\beta\sqrt{s}}{\ln(s)^{2}}}.

By Lemma 3.19, among the identities of length up to 2s2s, 2/32/3 of them are even, and 1/31/3 of them are odd. The probability P2s(X1=a(modNs))P2s\frac{P^{\prime}_{2s}(X_{1}=a\pmod{N_{s}})}{P^{\prime}_{2s}} is a weighted sum of P=k(X1=a(modNs))P=k\frac{P^{\prime}_{=k}(X_{1}=a\pmod{N_{s}})}{P^{\prime}_{=k}}. Noticing that the rest of the previous inequality does not depend on kk, doing a weighted sum gives

|P2s(X1=a(modNs))P2s232Ns|<αeβsln(s)2 if a is even, and\left|\frac{P^{\prime}_{2s}(X_{1}=a\pmod{N_{s}})}{P^{\prime}_{2s}}-\frac{2}{3}\cdot\frac{2}{N_{s}}\right|<\alpha e^{\frac{-\beta\sqrt{s}}{\ln(s)^{2}}}\text{ if }a\text{ is even, and}
|P2s(X1=a(modNs))P2s132Ns|<αeβsln(s)2 if a is odd.\left|\frac{P^{\prime}_{2s}(X_{1}=a\pmod{N_{s}})}{P^{\prime}_{2s}}-\frac{1}{3}\cdot\frac{2}{N_{s}}\right|<\alpha e^{\frac{-\beta\sqrt{s}}{\ln(s)^{2}}}\text{ if }a\text{ is odd.}

By the independence of X1,X2X_{1},X_{2}, we see that for all sufficiently large ss,

|P2s(X1=a(modNs)&X2=b(modNs))P2sc4Ns2|<αeβsln(s)2\left|\frac{P_{2s}(X_{1}=a\pmod{N_{s}}\ \&\ X_{2}=b\pmod{N_{s}})}{P_{2s}}-c\frac{4}{N_{s}^{2}}\right|<\alpha e^{\frac{-\beta\sqrt{s}}{\ln(s)^{2}}}

where c={19if a,b are both odd,29if one of a,b is odd and the other is even,49if a,b are both even.c=\begin{cases}\frac{1}{9}&\mbox{if $a,b$ are both odd,}\\ \frac{2}{9}&\mbox{if one of $a,b$ is odd and the other is even,}\\ \frac{4}{9}&\mbox{if $a,b$ are both even.}\\ \end{cases}


Now, we consider pairs (a,b)(a,b) modulo NsN_{s} such that no ppsp\leq p_{s} divides both aa and bb. Note that (a,b)(a,b) cannot both be even. As we have seen, up to the parities of aa and bb, for large ss, the distribution of XiX_{i} is approximately uniform, and the distribution of ordered pairs (X1,X2)(X_{1},X_{2}) is also approximately uniform. For each odd prime pp, the fraction of the pairs (a,b)(a,b) such that a,ba,b are both divisible by pp is approximately 1p2\frac{1}{p^{2}}. Thus, considering all primes, approximately 3pps(11p2)\prod\limits_{3\leq p\leq p_{s}}(1-\frac{1}{p^{2}}) of the possible pairs do not have a common odd prime factor ps\leq p_{s}. More precisely,

|P2s(no odd prime pps divides both X1,X2)P2s3pps(11p2)|<Ns2αeβsln(s)2.\left|\frac{P_{2s}(\text{no odd prime }p\leq p_{s}\text{ divides both }X_{1},X_{2})}{P_{2s}}-\prod\limits_{3\leq p\leq p_{s}}\left(1-\frac{1}{p^{2}}\right)\right|<N_{s}^{2}\alpha e^{\frac{-\beta\sqrt{s}}{\ln(s)^{2}}}.

Finally, considering p=2p=2, the probability that both a,ba,b are even is approximately 49\frac{4}{9}. Thus, we have that P2s(Cs,1c)P2s\frac{P_{2s}(C_{s,1}^{c})}{P_{2s}}, the probability that no prime ppsp\leq p_{s} divides both X1X_{1} and X2X_{2}, satisfies

|P2s(Cs,1c)P2s593pps(11p2)|<Ns2αeβsln(s)2(lns)2αeβsln(s)2.\left|\frac{P_{2s}(C_{s,1}^{c})}{P_{2s}}-\frac{5}{9}\cdot\prod\limits_{3\leq p\leq p_{s}}\left(1-\frac{1}{p^{2}}\right)\right|<N_{s}^{2}\alpha e^{\frac{-\beta\sqrt{s}}{\ln(s)^{2}}}\leq(\ln s)^{2}\alpha e^{\frac{-\beta\sqrt{s}}{\ln(s)^{2}}}.

Note that the right hand side of the inequality goes to 0 as ss\to\infty. Thus,

limsP2s(Cs,1c)P2s=593p(11p2).\lim\limits_{s\to\infty}\frac{P_{2s}(C_{s,1}^{c})}{P_{2s}}=\frac{5}{9}\cdot\prod_{3\leq p}\left(1-\frac{1}{p^{2}}\right).

The proof of (2) is similar. ∎

3.4 Abelian groups

Let VV be the variety of abelian groups. To axiomatize VV, we add to the group axioms the sentence (x)(y)x+y=y+x(\forall x)(\forall y)x+y=y+x.

3.4.1 Elementary invariants

Szmielew [21] carried out an elimination of quantifiers for abelian groups, and she gave elementary invariants. Later, Eklof and Fisher [6] used saturation to give elementary invariants for modules. Their methods also yield the Szmielew invariants for abelian groups. We give invariants for abelian groups below. For a prime pp, we write G[p]G[p] for the set {xG:px=0}\{x\in G:px=0\}, which consists of the identity and the elements of order pp.

  1. 1.

    α(p,n,k)\alpha(p,n,k), saying |pnG|k|p^{n}G|\geq k,

  2. 2.

    β(p,n,k)\beta(p,n,k), saying dim(pnG/pn+1G)kdim(p^{n}G/_{p^{n+1}G})\geq k,

  3. 3.

    γ(p,n,k)\gamma(p,n,k), saying dim(pnG[p])kdim(p^{n}G[p])\geq k,

  4. 4.

    δ(p,n,k)\delta(p,n,k), saying dim(pnG[p]/pn+1G[p])kdim(p^{n}G[p]/_{p^{n+1}G[p]})\geq k.

We consider presentations with a single generator aa and a single relator. The free abelian group on one generator is \mathbb{Z}, and the other abelian groups on one generator are the finite cyclic groups CmC_{m}. We focus on the sentences β(p,n,1)\beta(p,n,1), which say that there is an element divisible by pnp^{n} and not by pn+1p^{n+1}. We will see that these sentences are true in \mathbb{Z} and do not have limiting density 0 or 11. For p=2p=2, the limiting density does not exist, while for odd primes pp, the limiting density exists and has a value strictly between 0 and 11.

Lemma 3.28.
  1. 1.

    β(p,n,1)\beta(p,n,1) is true in \mathbb{Z}.

  2. 2.

    β(p,n,1)\beta(p,n,1) is true in CmC_{m} if and only if pn+1|mp^{n+1}|m.

Proof.

For (1), we note that in \mathbb{Z}, the element pnp^{n} witnesses the truth of β(p,n,1)\beta(p,n,1). For (2), consider CmC_{m}. For some rr (possibly 0) and some mm^{\prime} relatively prime to pp, we have m=prmm=p^{r}\cdot m^{\prime}, and then CmCprCmC_{m}\cong C_{p^{r}}\oplus C_{m^{\prime}}. If r>nr>n, then CprC_{p^{r}} has an element divisible by pnp^{n} and not by pn+1p^{n+1}, and otherwise, there is no such element. Furthermore, all elements of CmC_{m^{\prime}} are divisible by all powers of pp. So, CmC_{m} has elements divisible by pnp^{n} if and only if r>nr>n. ∎

A relator of length mm has the form w(a)=imdiaw(a)=\sum_{i\leq m}d_{i}a, where di=±1d_{i}=\pm 1. We consider the relator of length 0, representing the empty sum, to be 0.

Lemma 3.29.

Ps=1+2++2s=2s+11P_{s}=1+2+\ldots+2^{s}=2^{s+1}-1.

Proof.

There is one relator of length 0. For m1m\geq 1, there are 2m2^{m} possible relators t(a)=d1a++dmat(a)=d_{1}a+\ldots+d_{m}a, di=±1d_{i}=\pm 1. Then we have
Ps=1+2+22++2s=12s+112=2s+11P_{s}=1+2+2^{2}+\ldots+2^{s}=\frac{1-2^{s+1}}{1-2}=2^{s+1}-1. ∎

We consider the limiting density for β(p,n,1)\beta(p,n,1) for various combinations of pp and nn. Recall that the sentence β(p,n,1)\beta(p,n,1) is true in \mathbb{Z}, and it is true in CmC_{m} provided that pn+1|mp^{n+1}|m. Let AA be the set of relators that give \mathbb{Z}, and for r<pn+1r<p^{n+1}, let BrB_{r} be the set of relators that give m\mathbb{Z}_{m} for the various mpn+1rm\equiv_{p^{n+1}}r. We write Ps(A)P_{s}(A) and Ps(Br)P_{s}(B_{r}) for the number of relators of length at most ss in the sets AA and BrB_{r}.

Lemma 3.30.
  1. 1.

    Ps(A)=1+0<2ms(2mm)P_{s}(A)=1+\sum\limits_{0<2m\leq s}\left(\begin{array}[]{cc}2m\\ m\end{array}\right),

  2. 2.

    AA has limiting density 0.

Proof.

For (1), note that the relator t(a)t(a) gives \mathbb{Z} just in case t(a)t(a) has even length 2m2m for some mm, and aa and a-a each occur mm times. For (2), look back at Section 3.1 where we encountered the same quantities as the current PsP_{s} and Ps(A)P_{s}(A). There, we saw that limsPs(A)Ps=0\lim_{s\rightarrow\infty}\frac{P_{s}(A)}{P_{s}}=0. ∎

Recall that for a relator t(a)t(a), XX is the difference between the number of occurrences of aa and the number of occurrences of a1a^{-1} in the term t(a)t(a). For a relator of even length 2m2m, XX takes only even values 0,±2,,±2m0,\pm 2,\ldots,\pm 2m. For an identity of odd length 2m+12m+1, XX takes only odd values ±1\pm 1, ±3\pm 3, \ldots, ±(2m+1)\pm(2m+1). Using arguments similar to those in Section 3.3, we will show that the limiting density of 2n+1|X2^{n+1}|X does not exist, and for odd primes pp, the limiting density of pn+1|Xp^{n+1}|X is 1pn+1\frac{1}{p^{n+1}}.

Lemma 3.31.

For p=2p=2 and n0n\geq 0, the limiting density of 2n+1|X2^{n+1}|X does not exist. In particular, P2s(2n+1|X)P2s(23)(12n)\frac{P_{2s}(2^{n+1}|X)}{P_{2s}}\rightarrow(\frac{2}{3})(\frac{1}{2^{n}}), and P2s+1(2n+1|X)p2s+1(13)(12n)\frac{P_{2s+1}(2^{n+1}|X)}{p_{2s+1}}\rightarrow(\frac{1}{3})(\frac{1}{2^{n}}).

Proof.

We begin with the case where n=0n=0. Here the calculations are straightforward. We have P=2m(2|X)=22mP_{=2m}(2|X)=2^{2m}, and P=2m+1(2|X)=0P_{=2m+1}(2|X)=0. Then P2s(2|X)=1+22+24++22s=1+4+42++4s=4s+113P_{2s}(2|X)=1+2^{2}+2^{4}+\ldots+2^{2s}=1+4+4^{2}+\ldots+4^{s}=\frac{4^{s+1}-1}{3}. Therefore, P2s(2|X)P2s+1=(4s+113)22s+2123\frac{P_{2s}(2|X)}{P_{2s+1}}=\frac{(\frac{4^{s+1}-1}{3})}{2^{2s+2}-1}\rightarrow\frac{2}{3}. Since P2s+1(2|X)=P2sP_{2s+1}(2|X)=P_{2s}, we have P2s+1(2|X)P2s+1=(4s+113)22s+2113\frac{P_{2s+1}(2|X)}{P_{2s+1}}=\frac{(\frac{4^{s+1}-1}{3})}{2^{2s+2}-1}\rightarrow\frac{1}{3}. What we have shown is that if EE is the set of relators of even length, then P2s(E)P2s23\frac{P_{2s}(E)}{P_{2s}}\rightarrow\frac{2}{3} and P2s+1(E)P2s+113\frac{P_{2s+1}(E)}{P_{2s+1}}\rightarrow\frac{1}{3}.

For n1n\geq 1, we again use Theorem 3.20. For every presentation in EcE^{c}, we have 2n+1X2^{n+1}\nmid X, so we may condition on EE when we consider even length 2m2m. In the relation w(a)=i2mdiaw(a)=\sum_{i\leq 2m}d_{i}a, we can consider the sums d1+d2,d3+d4,,d2m1+d2md_{1}+d_{2},d_{3}+d_{4},\dots,d_{2m-1}+d_{2m}. This gives us an mm-step random walk on 2n+1\mathbb{Z}_{2^{n+1}} with each step being 22 with probability 14\frac{1}{4}, 2-2 with probability 14\frac{1}{4}, and 0 with probability 12\frac{1}{2}. Dividing everything by 2, we get a random walk with support {1,0,1}\{1,0,-1\} on 2n\mathbb{Z}_{2^{n}}. Thus, Theorem 3.20 applies. We have 2n+1X2^{n+1}\mid X exactly when the random walk ends at 02n0\in\mathbb{Z}_{2^{n}}. The probability of this is P=2m(2n+1|X)P=2m1/2n\frac{P_{=2m}(2^{n+1}|X)}{P_{=2m}}\to 1/2^{n}.

Now, as in the proof of Theorem 3.27, we have that most identities of length at most ss will have length at least s\geq\sqrt{s}. Since the rate of convergence in Theorem 3.20 is exponential and all identities in EE have even length, we can pass from the probability for identities of a fixed length to the probability for identities with length s\leq s, and we get P2s(2n+1|XE)P2s(E)1/2n\frac{P_{2s}(2^{n+1}|X\mid E)}{P_{2s}(E)}\to 1/2^{n}.

Since 2n+1X2^{n+1}\mid X only when XX is even, i.e., the identity is in EE, and the above probability 1/2n1/2^{n} was conditioned to EE, we have the desired

P2s(2n+1|X)P2s=P2s(2n+1|XE)P2s=P2s(2n+1|XE)P2s(E)P2s(E)P2s(12n)(23).\frac{P_{2s}(2^{n+1}|X)}{P_{2s}}=\frac{P_{2s}(2^{n+1}|X\mid E)}{P_{2s}}=\frac{P_{2s}(2^{n+1}|X\mid E)}{P_{2s}(E)}\cdot\frac{P_{2s}(E)}{P_{2s}}\to(\frac{1}{2^{n}})\cdot(\frac{2}{3}).

The odd case can be proved similarly. ∎

Lemma 3.32.

For odd primes pp, pn+1|Xp^{n+1}|X has limiting density 1pn+1\frac{1}{p^{n+1}}.

Proof.

For a fixed even length 2m2m, we get a random walk on pn+1\mathbb{Z}_{p^{n+1}} supported on {2,0,2}\{2,0,-2\}—a single step increases the length by 22. By Theorem 3.20, we have P2m(pn+1X)P2n1pn+1\frac{P_{2m}(p^{n+1}\mid X)}{P_{2n}}\to\frac{1}{p^{n+1}}. This random walk converges to the uniform distribution for even lengths, and the same is true for the odd lengths.

As before, we split the set of relators of length at most ss into two parts, those of length less than s\sqrt{s}, and those of length at least s\sqrt{s}. Let SsS_{s} be the number of relators of length at most ss for which the length is less than s\sqrt{s}, and let LsL_{s} be the number for which the length is at least s\sqrt{s}. Then SsPs0\frac{S_{s}}{P_{s}}\rightarrow 0, so LsPs1\frac{L_{s}}{P_{s}}\rightarrow 1. Then the exponential rate of convergence of Theorem 3.20 gives

Ps(pn+1|X)Ps1pn+1.\frac{P_{s}(p^{n+1}|X)}{P_{s}}\rightarrow\frac{1}{p^{n+1}}.

4 Generalizing

In this section, we give general conditions that imply some of the behaviors that we saw in Section 3. Our languages will have finitely many unary function symbols, and we may also allow finitely many constants.

4.1 Generalized bijective varieties

In Section 3.1, we considered the variety with axioms saying of a pair of unary function symbols S,S1S,S^{-1} that they are inverses, and we showed that for presentations with a single generator aa and a single identity of the form t(a)=at(a)=a, the sentences true in the free structure are exactly those with limiting density 11. In this subsection and the next, we turn our attention to varieties of structures with multiple bijective unary functions, possibly with additional axioms. We might suppose that the language of has unary function symbols g1,g11,,gn,gn1g_{1},g_{1}^{-1},\cdots,g_{n},g_{n}^{-1}, and that our varieties have axioms saying that for each ii, gig_{i} and gi1g_{i}^{-1} are inverses. However, the assumption that the functions have inverses named by function symbols turns out to be unnecessary once we know that the functions are 111-1 and onto.

Definition 4.1.

Let LL be a language with unary function symbols f1,,fnf_{1},\ldots,f_{n}, and let VV be an algebraic variety with theory TT. The variety is generalized bijective if for all ii, T(x,y)(fi(x)=fi(y)x=y)T\vdash(\forall x,y)(f_{i}(x)=f_{i}(y)\rightarrow x=y) and
T(y)(x)fi(x)=yT\vdash(\forall y)(\exists x)\ f_{i}(x)=y.

The result below says that for a generalized bijective variety, the basic functions have inverses named by terms.

Proposition 4.2.

Let TT be the theory of a generalized bijective variety in the language {f1,,fn}\{f_{1},\ldots,f_{n}\}. Then for each fif_{i}, there is some word uiu_{i} such that
T(x)fiui(x)=uifi(x)=xT\vdash(\forall x)f_{i}\circ u_{i}(x)=u_{i}\circ f_{i}(x)=x.

Proof.

Fix ii, and let FF be the free structure on one generator aa. There is some bFb\in F with fi(b)=af_{i}(b)=a. We can express bb as ui(a)u_{i}(a) for some word uiu_{i}. Then Ffiui(a)=aF\models f_{i}\circ u_{i}(a)=a. Recall that in a variety, if an atomic formula is true of the generating tuplea in the free structure, then it holds on all tuples in all structures [3, Theorem 11.4]. Thus, T(x)fiui(x)=xT\vdash(\forall x)\ f_{i}\circ u_{i}(x)=x. In FF, let x=fi(a)x=f_{i}(a). We have fiuifi(a)=fi(a)f_{i}\circ u_{i}\circ f_{i}(a)=f_{i}(a). Since fif_{i} is injective, this means that Fuifi(a)=aF\models u_{i}\circ f_{i}(a)=a. Hence, T(x)uifi(x)=xT\models(\forall x)\ u_{i}\circ f_{i}(x)=x. This completes the proof. ∎

Definition 4.3.

Let VV be a variety in the language {f1,,fn}\{f_{1},\ldots,f_{n}\}, where each fif_{i} is unary. The variety is commutative if the axioms imply (x)fi(fj(x))=fj(fi(x))(\forall x)f_{i}(f_{j}(x))=f_{j}(f_{i}(x)) for all i,ji,j.

Our main general result, Theorem 4.26, says that for a commutative generalized bijective variety VV and presentations with a single generator and a single identity, the zero–one law holds. Moreover, the sentences with density 11 are those true in the free structure. To prove Theorem 4.26, we will use a version of Gaifman’s Locality Theorem, which we discuss below.

4.2 Gaifman’s Locality Theorem

We state a special version of Gaifman’s Locality Theorem for generalized bijective varieties, and we sketch a proof using saturation. Fix a language LL consisting of unary function symbols f1,,fmf_{1},\ldots,f_{m}. Below, we define the Gaifman graph of an LL-structure. Gaifman defined the graph for structures in a finite relational language. When convenient, we treat the unary functions as binary relations.

Definition 4.4.

Let 𝒜\mathcal{A} be an LL-structure. The Gaifman graph of 𝒜\mathcal{A} is the undirected graph with universe equal to that of 𝒜\mathcal{A}, and with an edge between xx and yy if and only if fi(x)=yf_{i}(x)=y or fi(y)=xf_{i}(y)=x for some ii.

We define an equivalence relation \sim on 𝒜\mathcal{A} such that xyx\sim y if xx and yy belong to the same connected component in the Gaifman graph; i.e., there is a finite path leading from xx to yy.

Definition 4.5 (distance, d(x,y)d(x,y)).

For x,y𝒜x,y\in\mathcal{A}, the distance between xx and yy is the least rr such that there is a path of length rr from xx to yy. We write d(x,y)rd(x,y)\geq r, d(x,y)>rd(x,y)>r to indicate that the distance is, respectively, at least rr, or greater than rr.

Remark: Elements x,yx,y lie in different connected components just in case d(x,y)>rd(x,y)>r for all rr.


We consider substructures of 𝒜\mathcal{A}. Note that two connected components, thought of as substructures, are isomorphic if there is a map from one onto the other that preserves the unary functions fif_{i}, which we think of as binary relations. The structure 𝒜\mathcal{A} is determined, up to isomorphism, by the number of connected components of different isomorphism types.

Definition 4.6 (rr-ball, Br(a)B_{r}(a), Br(a¯)B_{r}(\bar{a})).

Let 𝒜\mathcal{A} be a structure and let rωr\in\omega.

  1. 1.

    For a𝒜a\in\mathcal{A}, the rr-ball around aa is Br(a)={x𝒜:d(a,x)r}B_{r}(a)=\{x\in\mathcal{A}:d(a,x)\leq r\}.

  2. 2.

    For a¯𝒜n\bar{a}\in\mathcal{A}^{n}, we write Br(a¯)B_{r}(\bar{a}) for the set i<nBr(ai)\cup_{i<n}B_{r}(a_{i}).

  3. 3.

    We write B(a)B_{\infty}(a) for the connected component of aa, or rBr(a)\cup_{r}B_{r}(a).

  4. 4.

    We write B(a¯)B_{\infty}(\bar{a}) for the union of the connected components of elements of a¯\bar{a}, or iB(ai)\cup_{i}B_{\infty}(a_{i}).

Let VV be a generalized bijective variety for the language LL. For 𝒜V\mathcal{A}\in V, each element has a unique image and a unique pre-image under each fif_{i}. We show that for each rr and nn, there is a finite set of formulas α(x¯)\alpha(\bar{x}) that describe, for all 𝒜V\mathcal{A}\in V, the possible substructures Br(a¯)B_{r}(\bar{a}) for nn-tuples a¯\bar{a}.

Lemma 4.7.

Let VV be a generalized bijective variety for the language LL. For each rr and nn, there is a finite set Cr,nC_{r,n} of formulas α(x¯)\alpha(\bar{x}), such that

  1. 1.

    for each 𝒜V\mathcal{A}\in V, each nn-tuple a¯\bar{a} in 𝒜\mathcal{A} satisfies a unique formula α(x¯)Cr,n\alpha(\bar{x})\in C_{r,n},

  2. 2.

    for all 𝒜,𝒜V\mathcal{A},\mathcal{A}^{\prime}\in V, if nn-tuples a¯\bar{a} in 𝒜\mathcal{A} and a¯\bar{a}^{\prime} in 𝒜\mathcal{A}^{\prime} satisfy the same formula α(x¯)Cr,n\alpha(\bar{x})\in C_{r,n}, then there is an isomorphism from Br(a¯)B_{r}(\bar{a}) onto Br(a¯)B_{r}(\bar{a}^{\prime}) that takes a¯\bar{a} to a¯\bar{a}^{\prime}.

Moreover, we may take the formulas α(x¯)\alpha(\bar{x}) in Cr,nC_{r,n} to be existential. We may equally well take them to be universal.

Proof.

We describe the possible elements of Br(x¯)B_{r}(\bar{x}) inductively as follows. The set B0(x¯)B_{0}(\bar{x}) has just the members of the nn-tuple x¯\bar{x} as possible elements. Now, suppose we have the possible elements of Br(x¯)B_{r}(\bar{x}) for some r0r\geq 0. We will set the possible elements of Br+1(x¯)B_{r+1}(\bar{x}) to be the elements of Br(x¯)B_{r}(\bar{x}) together with additional possible elements zz obtained as follows: Take some yBr(x¯)y\in B_{r}(\bar{x}) corresponding to a node at a distance rr from some xx¯x\in\bar{x} and follow an arrow labeled fif_{i} or fi1f_{i}^{-1} from yy to zz; please note that fi1f_{i}^{-1} is shorthand for the term that acts as an inverse to fif_{i} from Proposition 4.2.

We may think of the possible elements of Br(x¯)B_{r}(\bar{x}) as terms u(x)u(x), where uu is a string of fif_{i}, fi1f_{i}^{-1} of length at most rr. For an actual structure in our generalized bijective variety, with an actual tuple a¯\bar{a} corresponding to x¯\bar{x}, we may have equalities—different paths may lead to the same point. For 𝒜V\mathcal{A}\in V generated by a¯\bar{a}, the elements of Br(a¯)B_{r}(\bar{a}) are equivalence classes of terms u(ai)u(a_{i}), where uu is a string of fi,fi1f_{i},f_{i}^{-1} of length at most rr. We have an existential formula saying that there exist yy’s corresponding to the possible elements of Br(x¯)B_{r}(\bar{x}) such that the structure has a specific atomic diagram. We also have a universal formula saying that for all yy’s corresponding to the possible elements of Br(x¯)B_{r}(\bar{x}), the structure has a specific atomic diagram. ∎

We fix sets of formulas Cr,nC_{r,n} as in the lemma. Gaifman’s Locality Theorem says that any formula φ(x¯)\varphi(\bar{x}) (in a relational language) can be expressed as a finite Boolean combination of “local” formulas and “local” sentences (see the references [15], [7], [14]). For our setting, we take the local formulas and local sentences to be as follows.

Definition 4.8.
  1. 1.

    The rr-local formulas x¯\bar{x} are those in Cr,nC_{r,n} for various nn.

  2. 2.

    The rr-local sentences have one of the following forms:

    1. (a)

      (v1,,vs)(iαi(vi)&i<jd>2r(vi,vj))(\exists v_{1},\cdots,v_{s})\left(\bigwedge\limits_{i}\alpha_{i}(v_{i})\ \&\bigwedge\limits_{i<j}d^{>2r}(v_{i},v_{j})\right),
      for some ss and αi(x)Cr,1\alpha_{i}(x)\in C_{r,1},

    2. (b)

      (v)α(v)(\exists v)\alpha(v), for some αCr,1\alpha\in C_{r,1}.

Remark.

This definition is similar to Gaifman’s, except that we allow only special formulas in Cr,nC_{r,n}. Note that the formulas in Cr,nC_{r,n} already give information on whether the distance between xix_{i} and xjx_{j} is greater than 2r2r. Indeed, if d(xi,xj)2rd(x_{i},x_{j})\leq 2r, the formula will contain a conjunct that says (in the rational language) t(xi)=t(xj)t(x_{i})=t^{\prime}(x_{j}) for some t,tt,t^{\prime} of length at most rr. Thus, we may equivalently replace 2(a) by (v1,,vs)α(v1,,vs)(\exists v_{1},\cdots,v_{s})\alpha(v_{1},\cdots,v_{s}) for some αCr,s\alpha\in C_{r,s}. We chose the form above to stay closer to Gaifman’s definition.

Definition 4.9.

A formula or sentence is local if it is rr-local for some rr.

Here is our special version of Gaifman’s Locality Theorem, where the local formulas and sentences are as defined above.

Theorem 4.10.

Let VV be a generalized bijective variety with theory TT.

  1. 1.

    Any elementary first order sentence φ\varphi is equivalent over TT to a sentence φ\varphi^{*} that is a finite Boolean combination of local sentences.

  2. 2.

    Any elementary first order formula φ(x¯)\varphi(\bar{x}) with free variables x¯\bar{x} is equivalent over TT to a formula φ(x¯)\varphi^{*}(\bar{x}) that is finite Boolean combination of local sentences and local formulas. In fact, we may take φ(x¯)\varphi^{*}(\bar{x}) to be a finite disjunction of formulas αi(x¯)&β\alpha_{i}(\bar{x})\ \&\ \beta, where for each ii, αi(x¯)\alpha_{i}(\bar{x}) is a single local formula, and β\beta is a finite conjunction of local sentences and negations of local sentences.

We sketch a proof using saturation. We begin with some definitions and lemmas.

Definition 4.11.
  • For 𝒜V\mathcal{A}\in V, the local theory of 𝒜\mathcal{A} is the set of all local sentences and negations of local sentences that are true in 𝒜\mathcal{A}.

  • For a¯\bar{a} in 𝒜\mathcal{A}, the local type of a¯\bar{a} is the set of formulas generated by the local theory and the set of local formulas true of a¯\bar{a} in 𝒜\mathcal{A}.

Note that for a¯\bar{a} in 𝒜\mathcal{A} and a¯\bar{a}^{\prime} in 𝒜\mathcal{A}^{\prime} of the same length, if the local type of a¯\bar{a} in 𝒜\mathcal{A} is contained in the local type of a¯\bar{a}^{\prime} in 𝒜\mathcal{A}^{\prime}, then the local types are the same.

Lemma 4.12.

Let 𝒜,𝒜V\mathcal{A},\mathcal{A}^{\prime}\in V. If nn-tuples a¯\bar{a} in 𝒜\mathcal{A} and a¯\bar{a^{\prime}} in 𝒜\mathcal{A}^{\prime} satisfy the same local type, then there is a partial isomorphism ff from B(a¯)B_{\infty}(\bar{a}) onto B(a¯)B_{\infty}(\bar{a}^{\prime}) such that f(a¯)=a¯f(\bar{a})=\bar{a}^{\prime}.

Proof.

The fact that the tuples a¯\bar{a} and a¯\bar{a}^{\prime} satisfy the same local type means that the structures 𝒜\mathcal{A} and 𝒜\mathcal{A}^{\prime} satisfy the same local theory, and for each rr, the tuples a¯,a¯\bar{a},\bar{a}^{\prime} satisfy the same unique formula α(x¯)Cr,n\alpha(\bar{x})\in C_{r,n}. By Lemma 4.7, for each rr, there is an isomorphism pp from Br(a¯)B_{r}(\bar{a}) onto Br(a¯)B_{r}(\bar{a}^{\prime}) taking a¯\bar{a} to a¯\bar{a}^{\prime}. We have a tree of these finite partial isomorphisms pp between B(a¯)𝒜B_{\infty}(\bar{a})\subseteq\mathcal{A} and B(a¯)𝒜B_{\infty}(\bar{a}^{\prime})\subseteq\mathcal{A}^{\prime}, where at level rr, we put the isomorphisms from Br(a¯)B_{r}(\bar{a}) onto Br(a¯)B_{r}(\bar{a}^{\prime}) that take a¯\bar{a} to a¯\bar{a}^{\prime}, and at level r+1r+1, the successors of a given partial isomorphism pp from level nn are the extensions of pp taking Br+1(a¯)B_{r+1}(\bar{a}) isomorphically onto Br+1(a¯)B_{r+1}(\bar{a}^{\prime}). If B(a¯)B_{\infty}(\bar{a}) is infinite, then the tree is infinite, and it is finitely branching, so by König’s Lemma, there is a path (pr)rω(p_{r})_{r\in\omega}. The desired isomorphism is rpr\cup_{r}p_{r}. If the substructure B(a¯)B_{\infty}(\bar{a}) is finite, then it is contained in Br(a¯)B_{r}(\bar{a}) for some rr, and prp_{r} is the desired isomorphism. ∎

For any 𝒜V\mathcal{A}\in V, the isomorphism type of 𝒜\mathcal{A} is determined by the number of connected components of each isomorphism type. Suppose 𝒜\mathcal{A} is saturated, of infinite cardinality κ\kappa. In 𝒜\mathcal{A}, a local type Γ(x¯)\Gamma(\bar{x}) is satisfied if it is finitely satisfied. For a local type Γ(x)={αr(x):rω}\Gamma(x)=\{\alpha_{r}(x):r\in\omega\}, there are at least nn realizations of Γ(x)\Gamma(x) on different connected components if and only if for all rr, 𝒜\mathcal{A} satisfies the rr-local sentence saying that there are at least nn elements satisfying αr(x)\alpha_{r}(x) and at a distance greater than 2r2r. The number of connected components with an element satisfying Γ(x)\Gamma(x) is either finite or κ\kappa. This yields the following.

Lemma 4.13.

Suppose 𝒜,𝒜V\mathcal{A},\mathcal{A}^{\prime}\in V are saturated and of the same cardinality κ\kappa. If 𝒜,𝒜\mathcal{A},\mathcal{A}^{\prime} satisfy the same local sentences, then 𝒜𝒜\mathcal{A}\cong\mathcal{A}^{\prime}.

Proof.

Since 𝒜,𝒜\mathcal{A},\mathcal{A}^{\prime} are saturated, of the same cardinality, and satisfy the same local sentences, they realize the same local types, and they have the same number of connected components of each isomorphism type. Hence, they are isomorphic. ∎

Knowing what the saturated structures in the variety VV look like, we see that for any countable 𝒜V\mathcal{A}\in V, there exists a saturated structure 𝒜\mathcal{A}^{*} of cardinality 202^{\aleph_{0}} such that 𝒜,𝒜\mathcal{A},\mathcal{A}^{*} satisfy the same local sentences.

Lemma 4.14.

If 𝒜,𝒜V\mathcal{A},\mathcal{A}^{\prime}\in V have the same local theory, then they are elementarily equivalent.

Proof.

Let 𝒜\mathcal{A}^{*} and (𝒜)(\mathcal{A}^{\prime})^{*} be saturated models of the common local theory of 𝒜,𝒜\mathcal{A},\mathcal{A}^{\prime} such that 𝒜,(𝒜)\mathcal{A}^{*},(\mathcal{A}^{\prime})^{*} both have cardinality 202^{\aleph_{0}}. Applying Lemma 4.13, we see that 𝒜(𝒜)\mathcal{A}^{*}\cong(\mathcal{A}^{\prime})^{*}. Hence, 𝒜,𝒜\mathcal{A},\mathcal{A}^{\prime} are elementarily equivalent. ∎

Lemma 4.15.

Take nn-tuples a¯,a¯\bar{a},\bar{a}^{\prime} in 𝒜\mathcal{A}. If a¯,a¯\bar{a},\bar{a}^{\prime} satisfy the same local type, then there is an automorphism of 𝒜\mathcal{A} that takes a¯\bar{a} to a¯\bar{a}^{\prime}.

Proof.

We have a partial isomorphism ff from B(a¯)B_{\infty}(\bar{a}) onto B(a¯)B_{\infty}(\bar{a}^{\prime}) such that f(a¯)=a¯f(\bar{a})=\bar{a}^{\prime}. This extends to an automorphism that agrees with ff on B(a¯)B_{\infty}(\bar{a}), with f1f^{-1} on B(a¯)B_{\infty}(\bar{a}^{\prime}), and with the identity on the rest of 𝒜\mathcal{A}. ∎

Lemma 4.16.

If 𝒜φ\mathcal{A}\models\varphi, then there is a sentence ψ\psi true in 𝒜\mathcal{A} such that ψ\psi is a finite conjunction of local sentences and negations of local sentences and T(ψφ)T\vdash(\psi\rightarrow\varphi).

Proof.

If SS is the local theory of 𝒜\mathcal{A}, then TSφT\cup S\vdash\varphi. Then there is some ψ\psi, the conjunction of a finite subset of SS, such that T(ψφ)T\vdash(\psi\rightarrow\varphi). ∎

For a formula φ(x¯)\varphi(\bar{x}) with an nn-tuple x¯\bar{x} of variables, we have the following.

Lemma 4.17.

If 𝒜φ(a¯)\mathcal{A}\models\varphi(\bar{a}), then there is a formula ψ(x¯)=(α(x¯)&β)\psi(\bar{x})=(\alpha(\bar{x})\ \&\ \beta) such that β\beta is a finite conjunction of sentences in the local theory of 𝒜\mathcal{A}, α(x¯)\alpha(\bar{x}) is a local formula satisfied by a¯\bar{a} in 𝒜\mathcal{A}, and T(x¯)(ψ(x¯)φ(x¯))T\vdash(\forall\bar{x})(\psi(\bar{x})\rightarrow\varphi(\bar{x})).

Proof.

We have a saturated model \mathcal{B} of cardinality 202^{\aleph_{0}} with a tuple b¯\bar{b} satisfying the type of a¯\bar{a}. If \mathcal{B}^{\prime} is saturated and satisfies the local theory of 𝒜\mathcal{A} and \mathcal{B}, there is an isomorphism ff from \mathcal{B} onto \mathcal{B}^{\prime}. If b¯\bar{b}^{\prime} is an nn-tuple in \mathcal{B}^{\prime} satisfying the local type of a¯\bar{a} and b¯\bar{b}, we may suppose that f(b¯)=b¯f(\bar{b})=\bar{b}^{\prime}. Hence, b¯\bar{b}^{\prime} realizes the complete type of a¯\bar{a}. This shows that the local theory of 𝒜\mathcal{A} and the local type of a¯\bar{a} generate the full theory and type. If χ(x¯)\chi(\bar{x}) is a finite conjunction of local formulas and negations of local formulas in the local type of a¯\bar{a}, then there is a single formula α(x¯)\alpha(\bar{x}) in the local type of a¯\bar{a} that implies χ(x¯)\chi(\bar{x})—take α(x¯)Cr,n\alpha(\bar{x})\in C_{r,n} for sufficiently large rr. ∎

A standard model-theoretic argument gives the following.

Proposition 4.18.

Any elementary first order sentence φ\varphi is equivalent over TT to a finite disjunction of local sentences and negations of such sentences.

Proof.

For each 𝒜V\mathcal{A}\in V satisfying φ\varphi, choose ψ\psi as in Lemma 4.16, a finite conjunction of local sentences and negations, true in 𝒜\mathcal{A}, such that T(ψφ)T\vdash(\psi\rightarrow\varphi). Let SS be the set of chosen sentences. Now, T{¬ψ:ψS}{φ}T\cup\{\neg{\psi}:\psi\in S\}\cup\{\varphi\} is inconsistent, so there is a finite set SSS^{\prime}\subseteq S such that T(φψSψ)T\vdash(\varphi\rightarrow\bigvee_{\psi\in S^{\prime}}\psi). Then φ\varphi is equivalent over TT to the disjunction of the sentences in SS^{\prime}. ∎

Here is the companion result for formulas with free variables.

Proposition 4.19.

For any formula φ(x¯)\varphi(\bar{x}) with free variables among x¯\bar{x}, there is a formula φ(x¯)\varphi^{*}(\bar{x}) equivalent over TT to φ(x¯)\varphi(\bar{x}) such that φ(x¯)\varphi^{*}(\bar{x}) is a finite disjunction of formulas (α(x¯)&β)(\alpha(\bar{x})\ \&\ \beta), where β\beta is a conjunction of local sentences and negations and α(x¯)\alpha(\bar{x}) is a local formula.

Proof.

We replace x¯\bar{x} with a tuple of constants c¯\bar{c}. For each model 𝒜\mathcal{A} of TT and each tuple a¯\bar{a} satisfying φ(x¯)\varphi(\bar{x}), choose a formula ψ(x¯)\psi(\bar{x}) in the local type of a¯\bar{a} such that T(ψ(c¯)φ(c¯))T\vdash(\psi(\bar{c})\rightarrow\varphi(\bar{c})). Let SS be the set of chosen formulas. Now, T{¬ψ(c¯):ψ(c¯)S}{φ(c¯)}T\cup\{\neg{\psi(\bar{c})}:\psi(\bar{c})\in S\}\cup\{\varphi(\bar{c})\} is inconsistent, so for some finite SSS^{\prime}\subseteq S, T(φ(c¯)ψ(c¯)Sψ(c¯))T\vdash(\varphi(\bar{c})\rightarrow\bigvee_{\psi(\bar{c})\in S^{\prime}}\psi(\bar{c})). We may take ψ\psi of the form α(x¯)&β\alpha(\bar{x})\ \&\ \beta, where β\beta is the conjunction of the local sentences in SS^{\prime} and α(x¯)\alpha(\bar{x}) is the local formula in Cr,nC_{r,n} that is true of a¯\bar{a}, where rr is greatest such that SS^{\prime} contains a formula in Cr,nC_{r,n}. ∎

Remark.

For our special version of Gaifman’s Locality Theorem, the local formulas may be taken to be either existential or universal. Thus, over a completion of TT (or over the set of local sentences in the complete theory), each formula is equivalent to an existential formula, and to a universal formula.

4.3 The group associated to a generalized bijective variety

Let VV be a generalized bijective variety with theory TT. There is an equivalence relation on strings of function symbols such that strings t,tt,t^{\prime} are equivalent if T(x)t(x)=t(x)T\vdash(\forall x)t(x)=t^{\prime}(x). For a string of symbols tt, we may write len(t)\operatorname{len}(t) for the length of tt. We will associate to the variety VV a group G(V)G(V), whose elements are the equivalence classes of strings.

Definition 4.20 (Gaifman group, G(V)G(V)).

For a generalized bijective variety VV, the Gaifman group is the group G(V)G(V) consisting of equivalence classes of strings of symbols under the operation induced by concatenation of strings.

The identity in G(V)G(V) is the equivalence class of the empty string. For each function symbol fif_{i}, we fix a term uiu_{i} that names the inverse, as in Proposition 4.2. We may write fi1f_{i}^{-1} for uiu_{i}. The inverse function extends in a natural way to any word vv in f1,,fnf_{1},\cdots,f_{n}. Let FF be the element of VV obtained as the free structure generated by the finite tuple a¯\bar{a}. The group G(V)G(V) has a natural action on FF, taking tG(V)t\in G(V) and bFb\in F to t(b)t(b). Since b=t(a)b=t^{\prime}(a) for some tt^{\prime}, the action takes t(a)t^{\prime}(a) to tt(a)t\circ t^{\prime}(a).

Definition 4.21 (orbit under action of G(V)G(V)).

For 𝒜V\mathcal{A}\in V and b𝒜b\in\mathcal{A}, the orbit of bb under the action of G(V)G(V) is the set of all xx such that for some tG(V)t\in G(V), t(b)=xt(b)=x.

Note.

For 𝒜V\mathcal{A}\in V and b𝒜b\in\mathcal{A}, the orbit of bb under the action of G(V)G(V) is just the set of elements of 𝒜\mathcal{A} generated by bb. The automorphism orbit of bb results from the action of the group of automorphisms.

Lemma 4.22.

Let VV be a generalized bijective variety, and let FF be the free structure in VV generated by the tuple a¯\bar{a}. The action of G(V)G(V) on FF is well defined and simply transitive on the orbits.

Proof.

We first prove that the action is well defined. Suppose t1=t2t_{1}=t_{2} in G(V)G(V). Without loss of generality, we assume that t1t_{1} is obtained from t2t_{2} by applying an identity of G(V)G(V), say w=ww=w^{\prime}. This means that t1=uw(w)1vt_{1}=uw(w^{\prime})^{-1}v and t2=uvt_{2}=uv for some words u,vu,v. Then (t1,t(a))t1t(a)=uw(w)1vt(a)(t_{1},t^{\prime}(a))\mapsto t_{1}\circ t^{\prime}(a)=uw(w^{\prime})^{-1}vt^{\prime}(a). Since w=ww=w^{\prime} is an identity in G(V)G(V), we have T(x)w(x)=(w)(x)T\vdash(\forall x)w(x)=(w^{\prime})(x) and for an element aa of FF, Fuw(w)1vt(a)=uw(w)1vt(a)=uvt(a)=t2t(a)F\models uw(w^{\prime})^{-1}vt^{\prime}(a)=uw^{\prime}(w^{\prime})^{-1}vt^{\prime}(a)=uvt^{\prime}(a)=t_{2}t^{\prime}(a). Thus, the action is well defined.

Recall that every element xx of FF has the form t(ai)t(a_{i}) for some generator aia_{i}, and every such xx is in the orbit of aia_{i}. Thus, every orbit in FF has the form {t(ai):tG(V)}\{t(a_{i}):t\in G(V)\} for some generator aia_{i}. Now, take x=t(ai)x=t(a_{i}) in FF and suppose that Fut(ai)=vt(ai)F\models u\circ t(a_{i})=v\circ t(a_{i}). Since FF is free, we have that T(x)ut(x)=vt(x)T\models(\forall x)ut(x)=vt(x). Therefore, ut=vtut=vt holds in the group G(V)G(V), so by cancellation, we have u=vu=v. Thus, the action is simply transitive on its orbits. ∎

For our commutative generalized bijective variety with theory TT, we have the following.

Lemma 4.23.
  1. 1.

    For u,v,wG(V)u,v,w\in G(V), T(x)(u(w(x))=v(w(x))u(x)=v(x))T\vdash(\forall x)(u(w(x))=v(w(x))\leftrightarrow u(x)=v(x)).

  2. 2.

    For αCr,1\alpha\in C_{r,1}, T(x)(α(w(x))α(x))T\vdash(\forall x)(\alpha(w(x))\leftrightarrow\alpha(x)).

For structures 𝒜V\mathcal{A}\in V with a single generator aa, all elements have the same local type. In fact, they are in the same automorphism orbit as well as the same orbit under the action of G(V)G(V).

Lemma 4.24.

Suppose 𝒜V\mathcal{A}\in V is generated by aa. For αCr,1\alpha\in C_{r,1},
𝒜(x)α(x)(x)α(x)\mathcal{A}\models(\exists x)\alpha(x)\leftrightarrow(\forall x)\alpha(x).

Consider a local sentence ρ\rho saying that there exists x¯\bar{x} with xix_{i} satisfying αiCr,1\alpha_{i}\in C_{r,1} and with d(xi,xj)>2rd(x_{i},x_{j})>2r for i<ji<j. For 𝒜\mathcal{A} generated by a single element aa, ρ\rho cannot be true unless the αi\alpha_{i}’s are all the same and 𝒜\mathcal{A} has a tuple of elements x¯\bar{x} such that d(xi,xj)>2rd(x_{i},x_{j})>2r for i<ji<j. Thus, the important local invariants are the sentences (x)α(x)(\exists x)\alpha(x) for αCr,1\alpha\in C_{r,1} and the sentences saying that there are at least nn elements at a distance at least 2r2r. We will show that for these important sentences, the ones true in FF have density 11.

For a string tt of function symbols, we write tnt^{n} for the nn-fold concatenation of tt. We write t\langle t\rangle for the subgroup of G(V)G(V) generated by the equivalence class of tt—the elements are the equivalence classes of the strings tnt^{n}, tn=(t1)nt^{-n}=(t^{-1})^{n}. We need to understand truth in the structure 𝒜\mathcal{A} with presentation a|Ra|R, where RR is a single identity. Any identity is equivalent over TT to a canonically chosen identity of the form t(a)=at^{*}(a)=a, where the length of tt^{*} is bounded by a constant multiple of the length of RR. The next lemma will tell us a great deal about truth in 𝒜\mathcal{A}.

Lemma 4.25.

Let VV be a generalized bijective variety, and consider presentations a|Ra|R, where RR is an identity equivalent to one of the form t(a)=at^{*}(a)=a. Then for u,vG(V)u,v\in G(V), a|Ru(a)=v(a)\langle a|R\rangle\models u(a)=v(a) iff u,vu,v are in the same left coset of t\langle t^{*}\rangle.

Proof.

Let 𝒜=a|R\mathcal{A}=\langle a|R\rangle.


\Leftarrow: Without loss of generality, suppose v=u(t)nv=u(t^{*})^{n}. In 𝒜\mathcal{A}, we have

v(a)=u(t)n(a)=u(t)n1(a)==u(a).v(a)=u(t^{*})^{n}(a)=u(t^{*})^{n-1}(a)=\cdots=u(a)\ .

\Rightarrow: Now, suppose u(a)=v(a)u(a)=v(a) in 𝒜\mathcal{A}. Then T{t(a)=a}T\cup\{t^{*}(a)=a\} must prove

u(a)=x0(a)=x1(a)==x(a)=v(a),u(a)=x_{0}(a)=x_{1}(a)=\cdots=x_{\ell}(a)=v(a)\ ,

where for each i<i<\ell, we have one of the following:

  1. (i)

    xi+1=xitx_{i+1}=x_{i}t^{*},

  2. (ii)

    xi=xi+1tx_{i}=x_{i+1}t^{*}, or

  3. (iii)

    xi(a)=xi+1(a)x_{i}(a)=x_{i+1}(a).

In the first two cases, xix_{i} and xi+1x_{i+1} are clearly in the same left coset of t\langle t^{*}\rangle. In the third case, xi=xi+1x_{i}=x_{i+1} in G(V)G(V), so again xix_{i} and xi+1x_{i+1} are in the same left coset of t\langle t^{*}\rangle. ∎

For a given identity u(a)=v(a)u(a)=v(a), we are interested in the identities RR such that a|Ru(a)=v(a)\langle a|R\rangle\models u(a)=v(a). The lemma above lets us recognize these identities. We come to the theorem that gives conditions under which the sentences true in the free structure have limiting density 11.

Theorem 4.26.

Let VV be a commutative generalized bijective variety in the language {f1,,fn}\{f_{1},\cdots,f_{n}\}, and consider presentations with a single generator aa and a single identity. Let FF be the free structure on aa. If FF is infinite, then the sentences true in FF have limiting density 11.

Proof.

We show that for the important sentences α\alpha, if α\alpha is true in FF, then it has density 11, and if α\alpha is false in FF, then it has density 0. For structures in VV with generator aa, the important sentences say one of the following:

  1. 1.

    (x)α(x)(\exists x)\alpha(x) for αCr,1\alpha\in C_{r,1}—this is equivalent to a finite conjunction of formulas of the form u(a)=v(a)u(a)=v(a) or u(a)v(a)u(a)\not=v(a).

  2. 2.

    (x1,xn)i<jd(xi,xj)>2r(\exists x_{1},\ldots x_{n})\bigwedge_{i<j}d(x_{i},x_{j})>2r.

If FF is infinite, then we can show that any sentence of the second form true in FF is implied over TT by a sentence of the first form true in FF. A saturated model of the theory of FF has infinitely many connected components, and the sentence (x1,xn)i<jd(xi,xj)>2r(\exists x_{1},\ldots x_{n})\bigwedge_{i<j}d(x_{i},x_{j})>2r is clearly true in this model. Therefore, it is true in FF. Take witnesses x1,,xnx_{1},\ldots,x_{n}, where xi=wi(a)x_{i}=w_{i}(a). Choose kk such that all xix_{i} are in Bk(a)B_{k}(a), and take αC1,k\alpha\in C_{1,k} true of aa in FF. Then over TT, (x)α(x)(\exists x)\alpha(x) implies (x1,xn)i<jd(xi,xj)>2r(\exists x_{1},\ldots x_{n})\bigwedge_{i<j}d(x_{i},x_{j})>2r.

The group G(V)G(V) is abelian and finitely generated, so it is a finite direct product of cyclic groups generated by some elements b1,,bkb_{1},\cdots,b_{k}. We write Πi(x)\Pi_{i}(x) for the projection of an element xx on the subgroup generated by bib_{i}. Since G(V)G(V) is infinite, some bib_{i} must have infinite order. Without loss of generality, we suppose b1b_{1} has infinite order and generates a copy of \mathbb{Z}. We focus on Π1(x)\Pi_{1}(x), and we suppose that the values are integers.

Each identity RR has the form t(a)=t(a)t(a)=t^{\prime}(a), but this is equivalent to an identity of the form t(a)=at^{*}(a)=a. Let e0=maxi|Π1(fi)|e_{0}=\max_{i}|\Pi_{1}(f_{i})|. If len(t)r\operatorname{len}(t)\leq r, then the projection Π1(t)\Pi_{1}(t) is an integer bounded by re0r\cdot e_{0}. If len(t),len(t)r\operatorname{len}(t),\operatorname{len}(t^{\prime})\leq r, then d(t,t)2rd(t,t^{\prime})\leq 2r. Then |Π1(t)Π1(t)|2re0|\Pi_{1}(t)-\Pi_{1}(t^{\prime})|\leq 2r\cdot e_{0}. To prove Theorem 4.26, it is enough to show that all statements of the form t(a)=t(a)t(a)=t^{\prime}(a) or t(a)t(a)t(a)\not=t^{\prime}(a) true in FF have limiting density 11. The proof consists of two steps.

  1. 1.

    The first step is to show that for a fixed kk, the set of presentations
    a|t(a)=t(a)a|t(a)=t^{\prime}(a) such that |Π1(t)Π1(t)|<k|\Pi_{1}(t)-\Pi_{1}(t^{\prime})|<k has limiting density 0.

  2. 2.

    The second step is to show that for a fixed kk and a fixed identity RR of the form t(a)=t(a)t(a)=t^{\prime}(a), if |Π1(t)Π1(t))|>e0k|\Pi_{1}(t)-\Pi_{1}(t^{\prime}))|>e_{0}k, then for any uu, vv such that d(u,v)kd(u,v)\leq k in the Gaifman graph G(F)G(F), we have Fu(a)=v(a)F\models u(a)=v(a) if and only if aRu(a)=v(a)\langle a\mid R\rangle\models u(a)=v(a).

Toward the first step, we prove some lemmas.

Lemma 4.27.
  1. 1.

    The number of identities of length mm is nm(m+1)n^{m}(m+1). Furthermore, for every 0km+10\leq k\leq m+1, there are exactly nmn^{m} identities of length mm in which tt (the string of function symbols on the left side) has length kk.

  2. 2.

    Ps=ns+1(s+2)(n1)+1(n1)2P_{s}=\frac{n^{s+1}(s+2)(n-1)+1}{(n-1)^{2}}.

Proof.

For (1), the number of strings of function symbols of length mm is nmn^{m}. To determine an identity t(a)=t(a)t(a)=t^{\prime}(a), we choose one of the m+1m+1 initial segments to serve as the left-hand side. For (2), we simply note that

Ps\displaystyle P_{s} =\displaystyle= 0ms(m+1)nm=(1+2n++(s+1)ns)\displaystyle\sum_{0\leq m\leq s}(m+1)n^{m}=(1+2n+\ldots+(s+1)n^{s})
=\displaystyle= (s+2)ns+2(s+2)ns+1+1(n1)2=ns+1(s+2)(n1)+1(n1)2.\displaystyle\displaystyle\frac{(s+2)n^{s+2}-(s+2)n^{s+1}+1}{(n-1)^{2}}=\frac{n^{s+1}(s+2)(n-1)+1}{(n-1)^{2}}.

The next lemma may by interpreted as saying that a random identity of length s\leq s has length >s>\sqrt{s}.

Lemma 4.28.

limsPs2PsPs2=1\lim_{s\rightarrow\infty}\frac{P_{s^{2}}-P_{s}}{P_{s^{2}}}=1.

Proof.

Using Lemma 4.27, we get PsPs2=ns+1(s+2)(n1)+1ns2+1(s2(n1)+2)+1\frac{P_{s}}{P_{s^{2}}}=\frac{n^{s+1}(s+2)(n-1)+1}{n^{s^{2}+1}(s^{2}(n-1)+2)+1}. This clearly has limit 0, so Ps2PsPs2=1PsPs2\frac{P_{s^{2}}-P_{s}}{P_{s^{2}}}=1-\frac{P_{s}}{P_{s^{2}}} has limit 11. ∎

Let P=mP_{=m} be the number of identities of length exactly mm, and let P=m(A)P_{=m}(A) be the number of identities in AA of length equal to mm. Calculating the limit of P=s(A)P=s\frac{P_{=s}(A)}{P_{=s}} is often easier than calculating the limit of Ps(A)Ps\frac{P_{s}(A)}{P_{s}}. The lemma below gives us permission to do that.

Lemma 4.29.

For any set AA of identities of arbitrary length, if P=s(A)P=s\frac{P_{=s}(A)}{P_{=s}} has limit 0, then so does Ps(A)Ps\frac{P_{s}(A)}{P_{s}}.

Proof.

We show that for ϵ>0\epsilon>0, there is some mm such that for sms\geq m, Ps(A)Ps<ϵ\frac{P_{s}(A)}{P_{s}}<\epsilon. Take m1m_{1} such that for all sm1s\geq m_{1}, we have P=s(A)P=s<ϵ2\frac{P_{=s}(A)}{P_{=s}}<\frac{\epsilon}{2}, and take m2m_{2} such that for all ss such that sm2\sqrt{s}\geq m_{2}, we have PsPs<ϵ2\frac{P_{\sqrt{s}}}{P_{s}}<\frac{\epsilon}{2}. Let sm1,m2s\geq m_{1},m_{2}. Then

Ps(A)Ps(A)=s<msP=m(A)<ϵ2s<msP=m=ϵ2(PsPs).P_{s}(A)-P_{\sqrt{s}}(A)=\sum_{\sqrt{s}<m\leq s}P_{=m}(A)<\frac{\epsilon}{2}\sum_{\sqrt{s}<m\leq s}P_{=m}=\frac{\epsilon}{2}(P_{s}-P_{\sqrt{s}}).

This gives us

Ps(A)Ps=PsPs+Ps(A)Ps(A)Ps<ϵ2+ϵ2PsPsPs<ϵ2+ϵ2=ϵ.\frac{P_{s}(A)}{P_{s}}=\frac{P_{\sqrt{s}}}{P_{s}}+\frac{P_{s}(A)-P_{\sqrt{s}}(A)}{P_{s}}<\frac{\epsilon}{2}+\frac{\epsilon}{2}\cdot\frac{P_{s}-P_{\sqrt{s}}}{P_{s}}<\frac{\epsilon}{2}+\frac{\epsilon}{2}=\epsilon.\

The next lemma will complete the first step of the proof of Theorem 4.26. We write tt and tt^{\prime} for both strings of function symbols and elements of G(V)G(V).

Lemma 4.30.

For every kk\in\mathbb{N}, we have

limsPs(|Π1(t)Π1(t)|<k)Ps=0.\lim\limits_{s\to\infty}\frac{P_{s}(|\Pi_{1}(t)-\Pi_{1}(t^{\prime})|<k)}{P_{s}}=0\ .
Proof.

By Lemma 4.29, it suffices to prove that

limsP=s(|Π1(t)Π1(t)|<k)P=s=0.\lim\limits_{s\to\infty}\frac{P_{=s}(|\Pi_{1}(t)-\Pi_{1}(t^{\prime})|<k)}{P_{=s}}=0.

Furthermore, since kk is fixed, it is enough to prove that for every kk\in\mathbb{Z},

limsP=s(Π1(t)Π1(t)=k)P=s=0.\lim\limits_{s\to\infty}\frac{P_{=s}(\Pi_{1}(t)-\Pi_{1}(t^{\prime})=k)}{P_{=s}}=0.

Fix ss. The identities of length ss form a finite probability space, and the random variables Π1(t)\Pi_{1}(t) and Π1(t)\Pi_{1}(t^{\prime}) are not independent. By Lemma 4.27, we may consider Π1(t)Π1(t)\Pi_{1}(t)-\Pi_{1}(t^{\prime}) conditioned on tt having length \ell. Then len(t)=s\operatorname{len}(t^{\prime})=s-\ell. For each s\ell\leq s, the number of identities with len(t)=\operatorname{len}(t)=\ell and len(t)=s\operatorname{len}(t^{\prime})=s-\ell is equal to the number of strings of length ss, so the probability that len(t)=\operatorname{len}(t)=\ell is 1s+1\frac{1}{s+1}. The probability that Π1(t)Π1(t)=k\Pi_{1}(t)-\Pi_{1}(t^{\prime})=k is the sum over s\ell\leq s of the probability that len(t)=\operatorname{len}(t)=\ell times the conditional probability that Π1(t)Π1(t)=k\Pi_{1}(t)-\Pi_{1}(t^{\prime})=k given len(t)=\operatorname{len}(t)=\ell. We have

P=s(Π1(t)Π1(t)=k)P=s\frac{P_{=s}(\Pi_{1}(t)-\Pi_{1}(t^{\prime})=k)}{P_{=s}}
=1s+1=0sP=s(Π1(t)Π1(t)=k&len(t)=&len(t)=s)P=s(len(t)=&len(t)=s).=\frac{1}{s+1}\sum\limits_{\ell=0}^{s}\frac{P_{=s}(\Pi_{1}(t)-\Pi_{1}(t^{\prime})=k\ \&\ \operatorname{len}(t)=\ell\ \&\ \operatorname{len}(t^{\prime})=s-\ell)}{P_{=s}(\operatorname{len}(t)=\ell\ \&\ \operatorname{len}(t^{\prime})=s-\ell)}.

We write XX_{\ell} for Π1(t)\Pi_{1}(t) conditioned on tt having length \ell. Then, as a random variable, XX_{\ell} is a sum of \ell i.i.d. random variables YkY_{\ell_{k}} whose value is equal to the projection of the kthk^{th} symbol. All function symbols are equally likely. Thus, with probability 1n\frac{1}{n}, YY will be Π1(fi)\Pi_{1}(f_{i}) for 1in1\leq i\leq n. As ss\to\infty, we have \ell\to\infty. By the Central Limit Theorem, we have that X/X_{\ell}/\ell converges to a normal distribution. This means that, in particular, for every ϵ\epsilon, there is some ϵ\ell_{\epsilon} such that for every >ϵ\ell>\ell_{\epsilon}, the probability that X=iX_{\ell}=i is less than ϵ\epsilon for all ii; i.e.,

P=s(Π1(t)=i&len(t)=)P=s(len(t)=)<ϵ.\frac{P_{=s}(\Pi_{1}(t)=i\ \&\ \operatorname{len}(t)=\ell)}{P_{=s}(\operatorname{len}(t)=\ell)}<\epsilon.

Without loss of generality, we will assume that len(t)len(t)\operatorname{len}(t)\geq\operatorname{len}(t^{\prime}), so s/2\ell\geq s/2. Thus, >ϵ\ell>\ell_{\epsilon} whenever s>2ϵs>2\ell_{\epsilon}.

Now, we have that

P=s(Π1(t)Π1(t)=k&len(t)=&len(t)=s)P=s(len(t)=&len(t)=s)\displaystyle\frac{P_{=s}(\Pi_{1}(t)-\Pi_{1}(t^{\prime})=k\ \&\ \operatorname{len}(t)=\ell\ \&\ \operatorname{len}(t^{\prime})=s-\ell)}{P_{=s}(\operatorname{len}(t)=\ell\ \&\ \operatorname{len}(t^{\prime})=s-\ell)}
=\displaystyle= iP=s(Π1(t)=i&len(t)=)P=s(len(t)=)P=s(Π1(t)=ik&len(t)=s)P=s(len(t)=s)\displaystyle\sum\limits_{i}\frac{P_{=s}(\Pi_{1}(t)=i\ \&\ \operatorname{len}(t)=\ell)}{P_{=s}(\operatorname{len}(t)=\ell)}\cdot\frac{P_{=s}(\Pi_{1}(t^{\prime})=i-k\ \&\ \operatorname{len}(t^{\prime})=s-\ell)}{P_{=s}(\operatorname{len}(t^{\prime})=s-\ell)}
<\displaystyle< iϵP=s(Π1(t)=ik&len(t)=s)P=s(len(t)=s)\displaystyle\sum\limits_{i}\epsilon\cdot\frac{P_{=s}(\Pi_{1}(t^{\prime})=i-k\ \&\ \operatorname{len}(t^{\prime})=s-\ell)}{P_{=s}(\operatorname{len}(t^{\prime})=s-\ell)}
<\displaystyle< ϵ.\displaystyle\ \ \epsilon.

Combining these, we get

limsP=s(Π1(t)Π1(t)=k)P=s\displaystyle\lim\limits_{s\to\infty}\frac{P_{=s}(\Pi_{1}(t)-\Pi_{1}(t^{\prime})=k)}{P_{=s}}
=\displaystyle= lims1s+1=0sP=s(Π1(t)Π1(t)=k&len(t)=&len(t)=s)P=s(len(t)=&len(t)=s)\displaystyle\lim\limits_{s\to\infty}\frac{1}{s+1}\sum\limits_{\ell=0}^{s}\frac{P_{=s}(\Pi_{1}(t)-\Pi_{1}(t^{\prime})=k\ \&\ \operatorname{len}(t)=\ell\ \&\ \operatorname{len}(t^{\prime})=s-\ell)}{P_{=s}(\operatorname{len}(t)=\ell\ \&\ \operatorname{len}(t^{\prime})=s-\ell)}
=\displaystyle= 0.\displaystyle\ \ 0.

We proceed to the second step of the proof. Recall that e0=maxi|Π1(fi)|e_{0}=\max_{i}|\Pi_{1}(f_{i})|.

Lemma 4.31.

Fix RR of the form t(a)=t(a)t(a)=t^{\prime}(a), and fix kk such that
|Π1(t)Π1(t))|>e0k|\Pi_{1}(t)-\Pi_{1}(t^{\prime}))|>e_{0}k. For any uu, vv at a distance k\leq k in the Gaifman graph of FF, we have Fu(a)=v(a)F\models u(a)=v(a) if and only if aRu(a)=v(a)\langle a\mid R\rangle\models u(a)=v(a), where aR\langle a\mid R\rangle is the structure given by the presentation a|Ra|R.

Proof.

Based on the discussion before Lemma 4.27, we can see that if u(a)u(a) and v(a)v(a) are adjacent in the Gaifman graph, then |Π1(u)Π1(v)|=|Π1(fi)|e0|\Pi_{1}(u)-\Pi_{1}(v)|=|\Pi_{1}(f_{i})|\leq e_{0} for some fif_{i}. Thus, if d(u,v)kd(u,v)\leq k in the Gaifman graph, then |Π1(u)Π1(v)|e0k|\Pi_{1}(u)-\Pi_{1}(v)|\leq e_{0}k. We will also write t=t1tt^{*}=t^{-1}\circ t^{\prime}, where t1t^{-1} is the term that is the inverse of tt in the theory of the commutative generalized bijective variety. Note that t1t^{-1} may be longer than tt, but this does not affect the argument below.

It is easy to see that if u(a)=v(a)u(a)=v(a) holds in FF, then it holds in the structure a|R\langle a|R\rangle, where RR is t(a)=t(a)t(a)=t^{\prime}(a), which is equivalent to t(a)=at^{*}(a)=a. Suppose aRu(a)=v(a)\langle a\mid R\rangle\models u(a)=v(a). By Lemma 4.25, this implies that u,vu,v are in the same left coset of t\langle t^{*}\rangle; i.e., u1vtu^{-1}v\in\langle t^{*}\rangle. Taking the projection Π1\Pi_{1}, we see that Π1(u1v)Π1(t)\Pi_{1}(u^{-1}v)\in\Pi_{1}(\langle t^{*}\rangle). For some integer kk, we have u1v=(t)ktu^{-1}v=(t^{*})^{k}\in\langle t^{*}\rangle, and Π1((t)k)=kΠ1(t)\Pi_{1}((t^{*})^{k})=k\cdot\Pi_{1}(t^{*}). However, by assumption, |Π1(t)|=|Π1(t1t)|>e0|k||\Pi_{1}(t^{*})|=|\Pi_{1}(t^{-1}t^{\prime})|>e_{0}|k|, and we have |Π1(u1v)|e0|k||\Pi_{1}(u^{-1}v)|\leq e_{0}|k|. Therefore, we must have k=0k=0. It follows that Π1(u1v)=0Π1(t)=0\Pi_{1}(u^{-1}v)=0\cdot\Pi_{1}(t^{*})=0. Moreover, u1v=(t)0u^{-1}v=(t^{*})^{0}. It follows that u=vu=v in G(V)G(V), and Fu(a)=v(a)F\models u(a)=v(a). ∎

We are ready to complete the proof of the theorem. We just need to show that the sentences of the form u(a)=v(a)u(a)=v(a) or u(a)v(a)u(a)\not=v(a) true in FF have limiting density 11. By Lemma 4.30, for any integer kk, the set of identities t(a)=t(a)t(a)=t^{\prime}(a) such that |Π1(t)Π1(t)|>ϵ0|k||\Pi_{1}(t)-\Pi_{1}(t^{\prime})|>\epsilon_{0}|k| has density 11. For a fixed sentence u(a)=v(a)u(a)=v(a), take kk such that u,vu,v both have length at most k2\frac{k}{2}, so that u(a),v(a)u(a),v(a) are at distance at most kk. Then by Lemma 4.31, the sentence u(a)=v(a)u(a)=v(a) holds in FF iff it holds in the structures given by identities t(a)=t(a)t(a)=t^{\prime}(a) such that |Π1(t)=Π1(t)|>ϵ0k|\Pi_{1}(t)=\Pi_{1}(t^{\prime})|>\epsilon_{0}k, where this set has density 11. ∎

This theorem can be generalized to presentations with multiple generators.

Proposition 4.32.

Let VV be a commutative generalized bijective variety in the language {f1,,fn}\{f_{1},\cdots,f_{n}\} and suppose that the free structure on aa is infinite. Then for the structures in VV with an mm-tuple a¯\overline{a} of generators and a single identity, the sentences true in the free structure on a¯\overline{a} have limiting density 11.

To do so, we need the following lemma.

Lemma 4.33.

Let VV be a commutative generalized bijective variety, with theory TT. Let FmF_{m} be the free structure on mm generators. Suppose that F1F_{1} is infinite. Then for all m1m\geq 1, FmF_{m} and F1F_{1} satisfy the same theory.

Proof.

All elements of F1F_{1} have the same local type. Now, F1F_{1} has a saturated elementary extension FF^{*} whose Gaifman graph has infinitely many connected components. Let AA be the substructure of FF^{*} extending F1F_{1} and generated by an mm-tuple a1,,ama_{1},\ldots,a_{m} from different connected components. Clearly, F1F_{1} and FF^{*} satisfy the same special local sentences. Since the sentences are existential, any special local sentence true in F1F_{1} is true in AA, and any special local sentence true in AA is true in FF^{*}.

We may suppose that FmF_{m} has generators a1,,ana_{1},\ldots,a_{n}. The connected component of aia_{i} in FmF_{m} and in AA is generated by aia_{i}—the elements are named by terms t(ai)t(a_{i}). The special rr-local formula α(x)Cr,1\alpha(x)\in C_{r,1} true of the elements of F1F_{1} is true of each aia_{i} in FmF_{m} and in AA. We have an isomorphism from FmF_{m} onto AA that takes aia_{i} to aia_{i} and takes Br(ai)B_{r}(a_{i}) in FmF_{m} to Br(ai)B_{r}(a_{i}) in AA. Then F1F_{1} and FmF_{m} have the same theory. ∎

Proof of Proposition 4.32.

For presentations with mm generators and a single identity, we consider separately the set MM of presentations in which the identity involves a single generator and the complementary set ¬M\neg{M} in which the identity involves two distinct generators. For a presentation a¯|t1(ai)=t2(ai)\bar{a}|t_{1}(a_{i})=t_{2}(a_{i}) in MM, the resulting structure is the disjoint union of the structure ai|t1(ai)=t2(ai)\langle a_{i}|t_{1}(a_{i})=t_{2}(a_{i})\rangle (with generator aia_{i}) and (m1)(m-1) copies of F1F_{1} (one for each of the other aja_{j}’s). The identities in ¬M\neg{M} have the form t1(ai)=t2(aj)t_{1}(a_{i})=t_{2}(a_{j}) for iji\not=j. In the structure a¯|t1(ai)=t2(aj)\langle\overline{a}|t_{1}(a_{i})=t_{2}(a_{j})\rangle, the connected component of aia_{i} and the connected component of aja_{j} are collapsed via the relation t(ai)=t11t2t(aj)t(a_{i})=t_{1}^{-1}t_{2}t(a_{j}). Thus, the structure is a disjoint union of (m1)(m-1) copies of the free structure on one generator.

For fixed ss, we have a finite probability space. For a sentence φ\varphi, the probability that φ\varphi is true is Ps(φ)Ps=Ps(M&φ)Ps+Ps(¬M&φ)Ps\frac{P_{s}(\varphi)}{P_{s}}=\frac{P_{s}(M\ \&\ \varphi)}{P_{s}}+\frac{P_{s}(\neg{M}\ \&\ \varphi)}{P_{s}}. For presentations with a single generator, we write PsP^{\prime}_{s} and Ps(φ)P_{s}^{\prime}(\varphi). By Theorem 4.26,

Ps(φ)Ps{1if F1φ,0otherwise.\frac{P^{\prime}_{s}(\varphi)}{P^{\prime}_{s}}\rightarrow\left\{\begin{array}[]{cc}1&\mbox{if $F_{1}\models\varphi$,}\\ 0&\mbox{otherwise}.\end{array}\right.

Now, Ps(M&φ)Ps\frac{P_{s}(M\ \&\ \varphi)}{P_{s}} is the probability of (M&φ)(M\ \&\ \varphi). This is equal to the probability of MM times the conditional probability of φ\varphi given MM. The probability of MM is 1n\frac{1}{n}. The conditional probability of φ\varphi given MM is the same as the probability of φ\varphi for presentations with a single generator; namely, Ps(φ)Ps\frac{P^{\prime}_{s}(\varphi)}{P^{\prime}_{s}}. Thus, Ps(M&φ)Ps=(1n)(Ps(φ)Ps)\frac{P_{s}(M\ \&\ \varphi)}{P_{s}}=(\frac{1}{n})(\frac{P^{\prime}_{s}(\varphi)}{P^{\prime}_{s}}). As ss\rightarrow\infty, this approaches 1n\frac{1}{n} if φ\varphi is true in the free structures and 0 otherwise.

Similarly, the probability of (¬M&φ)(\neg{M}\ \&\ \varphi) is the probability of ¬M\neg{M} times the conditional probability of φ\varphi given ¬M\neg{M}. The probability of ¬M\neg{M} is (n1)n\frac{(n-1)}{n}. The conditional probability of φ\varphi given ¬M\neg{M} is 11 if φ\varphi is true in Fm1F_{m-1} and 0 otherwise. Thus,

Ps(¬M&φ)Ps={(n1)nif Fm1φ0otherwise.\frac{P_{s}(\neg{M}\ \&\ \varphi)}{P_{s}}=\left\{\begin{array}[]{cc}\frac{(n-1)}{n}&\mbox{if $F_{m-1}\models\varphi$}\\ 0&\mbox{otherwise.}\end{array}\right.

In total, Ps(φ)Ps\frac{P_{s}(\varphi)}{P_{s}} has limit 1n+n1n=1\frac{1}{n}+\frac{n-1}{n}=1 if φ\varphi is true in the free structures and 0 otherwise. ∎

Remark.

Using the multidimensional Central Limit Theorem [22], we can generalize the theorem and corollary above to any commutative generalized bijective variety VV where k\mathbb{Z}^{k} embeds into G(V)G(V). In this case, the random structures in VV with a single generator and kk identities satisfy the zero–one conjecture, and the limiting theory agrees with the theory of the free structure. However, without the condition that k\mathbb{Z}^{k} embeds in G(V)G(V), the statement is false, as witnessed by the bijective structures with two identities considered in Section 3.3.

4.3.1 Superstability

We make a brief comment on the superstability of completions of the theory of generalized bijective varieties. Recall that for an infinite cardinal κ\kappa, a (complete) theory TT is κ\kappa-stable if for every set AA in a model of TT, if AA has cardinality κ\kappa, then the set of complete types over AA has cardinality κ\kappa as well. A theory is stable if it is κ\kappa-stable for some κ\kappa, and it is superstable if it is κ\kappa-stable for all sufficiently large κ\kappa. If the language of TT is countable, then κ20\kappa\geq 2^{\aleph_{0}} will suffice. For more on stable theories, see Chapter 4 of [16].

Proposition 4.34.

All completions of the theory of a generalized bijective variety are superstable.

Proof.

Let TT be a completion of this theory, and let XX be a subset of some model of TT of cardinality κ20\kappa\geq 2^{\aleph_{0}}. We show that the number of 11-types over XX is at most κ\kappa.

A type in a variable xx over XX will say one of the following:

  1. 1.

    x=t(a)x=t(a) for some aXa\in X.

  2. 2.

    For every term tt and every aXa\in X, xt(a)x\not=t(a), and xx satisfies a certain quantifier-free 11-type p(x)p(x).

From this, it follows that if κ\kappa is the cardinality of XX, then the number of 11-types over XX is at most κ+20\kappa+2^{\aleph_{0}}. Thus, for κ20\kappa\geq 2^{\aleph_{0}}, TT is κ\kappa-stable. ∎

Remark.

If we drop the condition that TT is a completion of the theory of a generalized bijective variety, then there are theories in a language with finitely many unary function symbols that are unstable. We will not give an example here, although one is easily obtained taking the theory of a structure in the variety which we will study in Section 6.3.

4.4 Failure of the zero–one law

The next result gives conditions under which the zero–one law fails.

Theorem 4.35.

Let LL be a language with nn unary functions, including ff, where n2n\geq 2. Let VV be a variety such that for some term tt involving a symbol apart from ff, the theory TT of VV contains the sentence (x)(y)t(x)=t(y)(\forall x)(\forall y)t(x)=t(y). Consider presentations with an mm-tuple a¯\bar{a} of generators and a single identity, and suppose that in the free structure, f(t(a))t(a)f(t(a))\not=t(a). Then there is a sentence with limiting density neither 0 nor 11.

Remark.

The sentence (x)(y)t(x)=t(y)(\forall x)(\forall y)t(x)=t(y) says that tt has a constant value. If tt involved just the symbol ff, then the free structure would satisfy the sentence f(t(a))=t(f(a))=t(a)f(t(a))=t(f(a))=t(a).

Proof of Theorem 4.35.

Let φ\varphi be a sentence saying that ff fixes the constant given by tt. For instance, we may take φ=(x)f(t(x))=t(x)\varphi=(\forall x)f(t(x))=t(x). We show that φ\varphi does not have limiting density 0 or 1. We consider presentations with a tuple a¯\bar{a} of mm generators and a single identity. Let AA be the set of identities of the form u(ai)=v(aj)u(a_{i})=v(a_{j}), where u(ai)=t(u(ai))u(a_{i})=t(u^{\prime}(a_{i})) and v(aj)=f(t(v(aj)))v(a_{j})=f(t(v^{\prime}(a_{j}))). In the resulting structures, ff fixes the constant, so φ\varphi is true. Let BB be the set of identities of the form u(ai)=v(aj)u(a_{i})=v(a_{j}), where u(ai)=t(u(ai))u(a_{i})=t(u^{\prime}(a_{i})) and v(aj)=t(v(aj))v(a_{j})=t(v^{\prime}(a_{j})). The resulting structure is free and ff does not fix the constant, so φ\varphi is false. We show that neither AA nor BB has density 0. It follows that neither φ\varphi nor ¬φ\neg{\varphi} has density 0.

The number of identities of length \ell is m2n(+1)m^{2}n^{\ell}(\ell+1). Then

Ps\displaystyle P_{s} =m20s(+1)n=m2(1+2n++(s+1)ns)\displaystyle=m^{2}\sum_{0\leq\ell\leq s}(\ell+1)n^{\ell}=m^{2}(1+2n+\ldots+(s+1)n^{s})
=m2((s+2)ns+2(s+2)ns+1+1(n1)2)=m2(ns+1(s+2)(n1)+1(n1)2).\displaystyle=m^{2}\left(\frac{(s+2)n^{s+2}-(s+2)n^{s+1}+1}{(n-1)^{2}}\right)=m^{2}\left(\frac{n^{s+1}(s+2)(n-1)+1}{(n-1)^{2}}\right).

Say that tt has length rr. Then the identities in AA have length at least 2r+12r+1, and for =2r+1+\ell=2r+1+\ell^{\prime}, the number of identities in AA of length \ell is
m2n(+1)m^{2}n^{\ell^{\prime}}(\ell^{\prime}+1). Then

Ps(A)\displaystyle P_{s}(A) =m22r+1+s(+1)n=m2(1+2n++(s2r)ns2r1)\displaystyle=m^{2}\sum_{2r+1+\ell^{\prime}\leq s}(\ell^{\prime}+1)n^{\ell^{\prime}}=m^{2}(1+2n+\ldots+(s-2r)n^{s-2r-1})
=m2(ns2r(s2r+1)(n1)+1(n1)2),\displaystyle=m^{2}\left(\frac{n^{s-2r}(s-2r+1)(n-1)+1}{(n-1)^{2}}\right),

and

Ps(A)Ps=1n2r+1(s2r+1)(n1)+1(s+2)(n1)+11n2r+1.\frac{P_{s}(A)}{P_{s}}=\frac{1}{n^{2r+1}}\frac{(s-2r+1)(n-1)+1}{(s+2)(n-1)+1}\rightarrow\frac{1}{n^{2r+1}}\ .

The identities in BB have length at least 2r2r. For =2r+\ell=2r+\ell^{\prime}, the number of identities in BB of length \ell is m2n(+1)m^{2}n^{\ell^{\prime}}(\ell^{\prime}+1). Then

Ps(B)\displaystyle P_{s}(B) =m22r+s(+1)n=m2(1+2n+(s2r+1)ns2r)\displaystyle=m^{2}\sum_{2r+\ell^{\prime}\leq s}(\ell^{\prime}+1)n^{\ell^{\prime}}=m^{2}(1+2n+\ldots(s-2r+1)n^{s-2r})
=m2(ns2r+1(s2r+2)(n1)+1(n1)2),\displaystyle=m^{2}\left(\frac{n^{s-2r+1}(s-2r+2)(n-1)+1}{(n-1)^{2}}\right),

and

Ps(B)Ps=1n2r(s2r+2)(n1)+1(s+2)(n1)+11n2r.\frac{P_{s}(B)}{P_{s}}=\frac{1}{n^{2r}}\frac{(s-2r+2)(n-1)+1}{(s+2)(n-1)+1}\rightarrow\frac{1}{n^{2r}}.

Since n2n\geq 2, both of these limits are strictly between 0 and 11. ∎

5 Naming the generators

5.1 A general result

Let VV be a variety in a language LL with axioms generating a theory TT. We consider presentations with a fixed generating tuple a¯\bar{a}, and kk identities. Let LL^{\prime} be the result of adding to LL constants for the generators. We ask when the LL^{\prime}-sentences true in the free structure have limiting density 11.

Proposition 5.1.

Let TFT_{F} be the set of LL^{\prime}-sentences true in the free structure FF generated by a¯\bar{a}, and let SS be the set of LL^{\prime}-sentences with limiting density 11. Then the following are equivalent:

  1. 1.

    TFST_{F}\subseteq S,

  2. 2.

    TF=ST_{F}=S,

  3. 3.

    SS has the following two properties:

    1. (a)

      SS includes the sentences from TFT_{F} of the form t(a¯)t(a¯)t(\bar{a})\not=t^{\prime}(\bar{a}),

    2. (b)

      for any LL^{\prime}-formula φ(x)\varphi(x) with just xx free, if φ(t(a¯))S\varphi(t(\bar{a}))\in S for all closed terms t(a¯)t(\bar{a}), then (x)φ(x)S(\forall x)\varphi(x)\in S.

Proof.

We will prove (1)(2)(3)(1)(1)\Rightarrow(2)\Rightarrow(3)\Rightarrow(1). First, we assume (1) and prove (2). We must show that STFS\subseteq T_{F}. Take φS\varphi\in S. If φTF\varphi\notin T_{F}, then ¬φ\neg{\varphi} must be in TFT_{F}, so it is in SS. Then φ\varphi has limiting density 0, and we have a contradiction. Next, we assume (2) and prove (3). We can see that TFT_{F} has properties (a) and (b), so SS does as well. Finally, we assume (3) and prove (1). The set SS has properties (a) and (b). Sentences that are logically equivalent have the same limiting density as well as the same truth value in the free structure FF. We show by induction on φ(a¯)\varphi(\bar{a}) that if φ(a¯)TF\varphi(\bar{a})\in T_{F}, then φ(a¯)S\varphi(\bar{a})\in S. We suppose that the negations in our formulas are brought inside, next to the atomic formulas.

  1. 1.

    Suppose φ\varphi has the form t(a¯)=t(a¯)t(\bar{a})=t^{\prime}(\bar{a}). If FφF\models\varphi, then TFφT_{F}\vdash\varphi, so the limiting density is 11.

  2. 2.

    Suppose φ\varphi has the form t(a¯)t(a¯)t(\bar{a})\not=t^{\prime}(\bar{a}). By (a), if FφF\models\varphi, then φ\varphi has limiting density 11.

  3. 3.

    Suppose φ=(φ1&φ2)\varphi=(\varphi_{1}\ \&\ \varphi_{2}). If FφF\models\varphi, then both conjuncts are true, so both have limiting density 11. Then φ\varphi also has limiting density 11.

  4. 4.

    Suppose φ=(φ1φ2)\varphi=(\varphi_{1}\ \vee\ \varphi_{2}). If FφF\models\varphi, then at least one disjunct is true, so it has limiting density 11. Then φ\varphi also has limiting density 11.

  5. 5.

    Suppose φ=(x)ψ(x)\varphi=(\exists x)\psi(x). If FφF\models\varphi, then Fψ(t(a¯))F\models\psi(t(\bar{a})) for some t(x¯)t(\bar{x}). Then this sentence has limiting density 11, so φ\varphi also has limiting density 11.

  6. 6.

    Suppose φ=(x)ψ(x)\varphi=(\forall x)\psi(x). If FφF\models\varphi, then Fψ(t(a¯))F\models\psi(t(\bar{a})) for all closed terms t(a¯)t(\bar{a}). Then the sentence ψ(t(a¯))\psi(t(\bar{a})) has limiting density 11 for all t(a¯)t(\bar{a}), and by (b), (x)ψ(x)S(\forall x)\psi(x)\in S.

Consider the following further property.


Property (c): If (x)ψ(x)S(\exists x)\psi(x)\in S, then ψ(t(a¯))S\psi(t(\bar{a}))\in S for some t(a¯)t(\bar{a}).

Lemma 5.2.

If SS is complete (i.e., we have the zero–one law), then (b) and (c) are equivalent.

Proof.

First, suppose that (b) holds and that (x)ψ(x)S(\exists x)\psi(x)\in S. If there is no t(a¯)t(\bar{a}) such that ψ(t(a¯))S\psi(t(\bar{a}))\in S, then ¬ψ(t(a¯))S\neg{\psi(t(\bar{a}))}\in S for all t(a¯)t(\bar{a}), and (x)¬ψ(x)S(\forall x)\neg{\psi(x)}\in S for a contradiction. Now, suppose (c) holds and that ψ(t(a¯))S\psi(t(\bar{a}))\in S for all t(a¯)t(\bar{a}). If ¬(x)ψ(x)S\neg{(\forall x)\psi(x)}\in S, then (x)¬ψ(x)S(\exists x)\neg{\psi(x)}\in S. By (c), ¬ψ(t(a¯))S\neg{\psi(t(\bar{a}))}\in S for some t(a¯)t(\bar{a}) for a contradiction, so (x)ψ(x)S(\forall x)\psi(x)\in S. ∎

Lemma 5.3.

Suppose SS satisfies (a) and (b). Then for all formulas φ(x,y)\varphi(x,y) with free variables x,yx,y, if φ(t(a¯),t(a¯))S\varphi(t(\bar{a}),t^{\prime}(\bar{a}))\in S for all terms t(a¯),t(a¯)t(\bar{a}),t^{\prime}(\bar{a}), then
(x)(y)φ(x,y)S(\forall x)(\forall y)\varphi(x,y)\in S.

Proof.

For a fixed term t(a¯)t(\bar{a}), suppose φ(t(a¯),t(a¯))S\varphi(t(\bar{a}),t^{\prime}(\bar{a}))\in S for all t(a¯)t^{\prime}(\bar{a}). By (b),
(y)φ(t(a¯),y)S(\forall y)\varphi(t(\bar{a}),y)\in S for all t(a¯)t(\bar{a}). So, by (b), (x)(y)φ(a¯,x,y)S(\forall x)(\forall y)\varphi(\bar{a},x,y)\in S. ∎

If the orbit of a¯\bar{a} in 𝒜\mathcal{A} is defined by an LL-formula ψ(x¯)\psi(\bar{x}), then for each LL^{\prime}-sentence φ\varphi, we have 𝒜φ(a¯)\mathcal{A}\models\varphi(\bar{a}) iff 𝒜\mathcal{A} satisfies the LL-sentences (x¯)(ψ(x¯)&φ(x¯))(\exists\bar{x})(\psi(\bar{x})\ \&\ \varphi(\bar{x})) and (x¯)(ψ(x¯)φ(x¯))(\forall\bar{x})(\psi(\bar{x})\rightarrow\varphi(\bar{x})).

Proposition 5.4.

Let FF be the free structure generated by a¯\bar{a}. Suppose that the orbit of a¯\bar{a} is defined by the LL-formula ψ(x¯)\psi(\bar{x}) and the LL^{\prime}-sentence ψ(a¯)\psi(\bar{a}) has limiting density 11. Suppose also that for all LL-sentences φ\varphi, φ\varphi is true in FF if and only if it has limiting density 11. Then the same is true for all LL^{\prime}-sentences.

Proof.

Take an LL^{\prime}-sentence φ(a¯)\varphi(\bar{a}) that is true in FF. In FF, this is equivalent to the LL-sentence φ=(x¯)(ψ(x¯)φ(x¯))\varphi^{*}=(\forall\bar{x})(\psi(\bar{x})\rightarrow\varphi(\bar{x})). The sentence φ\varphi^{*} is true in FF, so it has limiting density 11. The set of sentences with limiting density 11 is closed under logical consequence, so since ψ(a¯)\psi(\bar{a}) has limiting density 11, it follows that φ(a¯)\varphi(\bar{a}) has limiting density 11. ∎

5.2 Generalized bijective structures and sentences with constants

We have seen that for the basic bijective variety and for the broader class of commutative generalized bijective varieties, when we consider presentations with a single generator and a single identity, the sentences (in the language of the variety) true in the free structure have density 11. We can apply Proposition 5.1 to extend this to sentences with a constant naming the generator.

Example 5.5.

Let VV be a commutative generalized bijective variety in the language LL. Consider presentations with a single generator aa and a single identity, and let LL^{\prime} be the extension of LL with a constant for the generator. Suppose that the free structure FF generated by aa is infinite. In FF, all elements are automorphic. In particular, aa and t(a)t(a) are automorphic via the automorphism xt(x)x\mapsto t(x). Preparing to apply Proposition 5.1, we take ψ(x)\psi(x) to be x=xx=x. Clearly, ψ(a)\psi(a) has limiting density 11. By Theorem 4.26, the LL-sentences true in the free structure have limiting density 11. Then Proposition 5.1 says that this holds also for the LL^{\prime}-sentences (involving aa).

For a generating tuple a¯\bar{a}, the sentences φ(a¯)\varphi(\bar{a}) true in the free structure on a¯\bar{a} have density 11. To establish this, we need to take a closer look at the formulas Cr,nC_{r,n} from Section 4.2 in the expanded language LL^{\prime} with constants naming the generators. We consider the unary functions fif_{i} as binary relations and the constants aia_{i} as unary relations. Thus, we have atomic formulas with the meanings x=aix=a_{i} and fi(x)=yf_{i}(x)=y.

Lemma 5.6.

Let VV be a commutative generalized bijective variety, and consider presentations with a generating mm-tuple a¯\bar{a} and a single identity. Let FF be the free structure. If FF is infinite, then for every rωr\in\omega, there is a set SS of presentations such that

  1. 1.

    SS has limiting density 11, and

  2. 2.

    for α(x¯)Cr,m\alpha(\bar{x})\in C_{r,m}, the following are equivalent:

    1. (a)

      α(a¯)\alpha(\bar{a}) holds in FF,

    2. (b)

      α(a¯)\alpha(\bar{a}) holds in some structure given by a presentation in SS,

    3. (c)

      α(a¯)\alpha(\bar{a}) holds in all structures given by a presentation in SS.

If 𝒜\mathcal{A} is the structure given by the identity t1(ai)=t2(aj)t_{1}(a_{i})=t_{2}(a_{j}), then for each rr, we get an isomorphism pp from B2r(a¯)B_{2r}(\bar{a}) in FF to B2r(a¯)B_{2r}(\bar{a}) in 𝒜\mathcal{A}, given by p(u(ak))=u(a)p(u(a_{k}))=u(a).

Proof.

We use notation from the proof of Theorem 4.26. Recall that Π1\Pi_{1} is the projection onto the copy of \mathbb{Z} generated by b1b_{1}, where b1b_{1} is an element of infinite order in the abelian group G(V)G(V) associated with the variety VV. Let SS be the set of presentations in which the identity t1(ai)=t2(aj)t_{1}(a_{i})=t_{2}(a_{j}) satisfies |Π1(t1)Π1(t2)|>e0r|\Pi_{1}(t_{1})-\Pi_{1}(t_{2})|>e_{0}r, where e0=maxi|Π1(fi)|+4e_{0}=\max_{i}|\Pi_{1}(f_{i})|+4. The fact that SS has limiting density 11 follows from the proof of Theorem 4.26.

Since the formulas in Cr,mC_{r,m} uniquely describe the isomorphism type of Br(x¯)B_{r}(\bar{x}), it suffices to show that pp is an isomorphism. We know that pp is surjective since 𝒜\mathcal{A} is a quotient of FF. As in the proof of Theorem 4.26, it is also injective. Indeed, if i=ji=j, then the projection from FF to 𝒜\mathcal{A} is injective on the substructure generated by a1,,ai1,ai+1,,ama_{1},\cdots,a_{i-1},a_{i+1},\cdots,a_{m}. On the substructure generated by aia_{i}, if u(ai)=u(ai)u(a_{i})=u^{\prime}(a_{i}), then we can apply Lemma 4.25, and we see that (the equivalence class of) u1uu^{-1}u^{\prime} is in t1t21\langle t_{1}t_{2}^{-1}\rangle as an element of G(V)G(V). Since |Π1(t1)Π1(t2)|>e0r|\Pi_{1}(t_{1})-\Pi_{1}(t_{2})|>e_{0}r and the length of u1uu^{-1}u^{\prime} is at most 4r4r, this is not possible. If iji\neq j, then, as in Corollary 4.32, the substructure generated by aia_{i} and that generated by aja_{j} are identified via aj=t1t21(ai)a_{j}=t_{1}t_{2}^{-1}(a_{i}), while the projection map is injective on the substructure generated by the further generators aka_{k}. Thus, if u(ai)=u(aj)u(a_{i})=u^{\prime}(a_{j}), then we must have uu1=t1t21uu^{\prime-1}=t_{1}t_{2}^{-1}, but since |Π1(t1)Π1(t2)|>e0r|\Pi_{1}(t_{1})-\Pi_{1}(t_{2})|>e_{0}r and the length of u1uu^{-1}u^{\prime} is at most 4r4r, this is again impossible.

Recall that we are thinking of the language as relational and we have atomic formulas with the meanings x=aix=a_{i} and fi(x)=yf_{i}(x)=y. The formula saying x=aix=a_{i} holds exactly on aia_{i}, in either B2r(a¯;F)B_{2r}(\bar{a};F) or B2r(a¯;𝒜)B_{2r}(\bar{a};\mathcal{A}). If the formula saying fi(x)=yf_{i}(x)=y holds in B2r(a¯;F)B_{2r}(\bar{a};F), then it holds of p(x)p(x) and p(y)p(y) in B2r(a¯;𝒜)B_{2r}(\bar{a};\mathcal{A}) because 𝒜\mathcal{A} is a quotient of FF. Thus, we have fi(p(x))=p(y)f_{i}(p(x))=p(y). Finally, suppose that in B2r(a¯;𝒜)B_{2r}(\bar{a};\mathcal{A}), fi(p(x))=p(y)f_{i}(p(x))=p(y). Then p(x)=u(aj)p(x)=u(a_{j}) for some jj and p(y)=fi(u(aj))p(y)=f_{i}(u(a_{j})). However, the map pp is bijective, so Fx=u(aj)F\models x=u(a_{j}) and Fy=fi(u(aj))F\models y=f_{i}(u(a_{j})) as well. Then fi(x)=yf_{i}(x)=y holds in B2r(a¯;F)B_{2r}(\bar{a};F). This shows that pp is an isomorphism, completing the proof. ∎

Theorem 5.7.

Let VV be a commutative generalized bijective variety in the language LL, and suppose that the free structure on one generator is infinite. Consider presentations with a fixed generating mm-tuple a¯\bar{a} and a single identity. Let FF be the free structure on a¯\bar{a}. Let LL^{\prime} be the result of adding constants for the elements of a¯\bar{a} to LL. Then an LL^{\prime}-sentence is true in FF iff it has limiting density 11.

Proof.

Let φ\varphi^{\prime} be an LL^{\prime}-sentence that is true in FF, so φ=φ(a¯)\varphi^{\prime}=\varphi(\bar{a}) for some LL-formula φ\varphi. By Theorem 4.10, φ(x¯)\varphi(\bar{x}) can be expressed as a finite disjunction iφi(x¯)\bigvee_{i}\varphi_{i}(\bar{x}) for φi(x¯)\varphi_{i}(\bar{x}) of the form ρi(x¯)&χi\rho_{i}(\bar{x})\ \&\ \chi_{i}, where ρi(x¯)Cr,m\rho_{i}(\bar{x})\in C_{r,m} and χi\chi_{i} is a conjunction of special sentences and negations of special sentences. Recall that the special sentences have the form (v1,,vs)(iαi(vi)&i<jd>2r(vi,vj))(\exists v_{1},\cdots,v_{s})\left(\bigwedge\limits_{i}\alpha_{i}(v_{i})\ \&\bigwedge\limits_{i<j}d^{>2r}(v_{i},v_{j})\right), where αi(vi)Cr,1\alpha_{i}(v_{i})\in C_{r,1} and d>2r(vi,vj)d^{>2r}(v_{i},v_{j}) is the formula saying that the distance between viv_{i} and vjv_{j} in the Gaifman graph is greater than 2r2r. Since Fφ(a¯)F\models\varphi(\bar{a}), we have Fφi(a¯)F\models\varphi_{i}(\bar{a}) for all ii, i.e., Fρi(a¯)&χiF\models\rho_{i}(\bar{a})\ \&\ \chi_{i}. Since χi\chi_{i} is an LL-sentence true in FF, it has limiting density 11 by Theorem 4.26. On the other hand, ρi(x¯)Cr,m\rho_{i}(\bar{x})\in C_{r,m}, and by Lemma 5.6, ρi(a¯)\rho_{i}(\bar{a}) also has limiting density 11. Thus, φ\varphi^{\prime} has limiting density 11. ∎

6 More examples

In Section 3, we gave some examples illustrating different possible behaviors of limiting density. We considered sentences with no constants. In Section 4, we gave conditions guaranteeing that the sentences with limiting density 11 are those true in the free structure. In Section 5, we gave some results for sentences with constants naming the generators. In the current section, we look again at some of the examples from Section 3 in light of the results from Sections 4 and 5. We we also give some further examples, illustrating more subtle points suggested by these results.

6.1 Examples of Proposition 5.1

Let VV be a variety in the language LL. Consider presentations with a fixed tuple a¯\bar{a} of generators, and some number of identities, and let LL^{\prime} be the result of adding to LL constants for the generators. Here, for reference, is the statement of Proposition 5.1.


Proposition 5.1. Let TFT_{F} be the set of LL^{\prime}-sentences true in the free structure FF generated by a¯\bar{a}, and let SS be the set of LL^{\prime}-sentences with limiting density 11. Then the following are equivalent:

  1. 1.

    TFST_{F}\subseteq S,

  2. 2.

    TF=ST_{F}=S,

  3. 3.

    SS has the following two properties:

    1. (a)

      SS includes the sentences from TFT_{F} of the form t(a¯)t(a¯)t(\bar{a})\not=t^{\prime}(\bar{a}),

    2. (b)

      for any LL^{\prime}-formula φ(x)\varphi(x) with just xx free, if φ(t(a¯))S\varphi(t(\bar{a}))\in S for all closed terms t(a¯)t(\bar{a}), then (x)φ(x)S(\forall x)\varphi(x)\in S.

The proposition says that conditions (a) and (b) are necessary and sufficient for the LL^{\prime}-sentences true in FF to have density 11. We revisit some examples and see what the result says about them.

6.1.1 Generalized bijective structures

In Theorem 5.7, we saw that for the variety of generalized bijective structures and presentations with a single generator aa and a single identity, any sentence, possibly involving the constant aa, has limiting density 11 iff it is true in the free structure on aa. Hence, we must have both properties (a) and (b) from Proposition 5.1.

6.1.2 Abelian groups

In Section 3.4, we saw that for abelian groups and presentations with a single generator and a single relator, the zero–one law fails.

Proposition 6.1.

For abelian groups and presentations with a single generator and a single relator, Property (a) holds and Property (b) fails, witnessed by the formulas φ(a,x)\varphi(a,x) saying that pn+1xpnap^{n+1}x\not=p^{n}a, where pp is an odd prime.

Proof.

The free structure is \mathbb{Z}. Take a sentence of the form ma0ma\not=0. This is true in \mathbb{Z}, and the sentence is in SS since all relators longer than |m||m| make it true. Thus, Property (a) holds. Now fix nn and an odd prime pp. For all closed terms t(a)t(a), the sentence pn+1t(a)pn(a)p^{n+1}t(a)\not=p^{n}(a) is in SS. We have pn+1t(a)pnap^{n+1}t(a)\not=p^{n}a for all terms t(a)=mat(a)=ma. By Property (a), the sentences pn+1t(a)pnap^{n+1}t(a)\not=p^{n}a are all in SS. If we had Property (b), then the sentence (x)pn+1xpna(\forall x)p^{n+1}x\not=p^{n}a would be in SS. However, recall from Section 3.4 that the sentence β(p,n,1)\beta(p,n,1) says that there is an element divisible by pnp^{n} and not by pn+1p^{n+1}. This is true in \mathbb{Z} but not in SS—by Lemmas 3.28 and 3.32, the limiting density is 1pn+1\frac{1}{p^{n+1}}. Since pna=pnap^{n}a=p^{n}a is logically valid, pnap^{n}a is divisible by pnp^{n} in all models. Thus, if the sentence (x)pn+1xpna(\forall x)p^{n+1}x\not=p^{n}a is in SS, then β(p,n,1)\beta(p,n,1) is in SS—in the models satisfying (x)pn+1xpna(\forall x)p^{n+1}x\not=p^{n}a, pnap^{n}a is not divisible by pn+1p^{n+1}. This is a contradiction. ∎

6.1.3 Structures with a single unary function, one generator, and one identity

The next example is from Section 3.2. The variety of unary functions has a single unary function symbol ff and no axioms.

Proposition 6.2.

For the variety of unary functions and presentations of the form a|fm(a)=fn(a)a|f^{m}(a)=f^{n}(a), Property (a) holds and Property (b) fails.

Proof.

To show that Property (a) holds, consider a sentence of the form fi(a)fj(a)f^{i}(a)\neq f^{j}(a). Note that the set of presentations with identity fn(a)=fm(a)f^{n}(a)=f^{m}(a) for n,m>i+j+1n,m>i+j+1 has limiting density 11. Moreover, for any such presentation, we recall from Section 3.2 that the resulting structure is a finite chain leading to a finite cycle where the chain is longer than both i+1i+1 and j+1j+1. Then we have fi(a)fj(a)f^{i}(a)\neq f^{j}(a) in the structure. Thus, fi(a)fj(a)f^{i}(a)\neq f^{j}(a) is in SS.

To show that Property (b) fails, let φ(x)=(y)(xyf(x)f(y))\varphi(x)=(\forall y)(x\not=y\rightarrow f(x)\not=f(y)). We will show that this witnesses the failure of (b). For any fixed x=fi(a)x=f^{i}(a), note that the set of presentations with identity fn(a)=fm(a)f^{n}(a)=f^{m}(a) for n,m>i+1n,m>i+1 has limiting density 11. In any such presentation, we have fj+1(a)f(y)f^{j+1}(a)\neq f(y) unless y=fj(a)y=f^{j}(a). Thus, the sentence saying that the formula

φ(x)=(y)(xyf(x)f(y))\varphi(x)=(\forall y)(x\not=y\rightarrow f(x)\not=f(y))

holds for x=fi(a)x=f^{i}(a) is in SS for any closed term fi(a)f^{i}(a). On the other hand, the sentence (x)(y)(xyf(x)f(y))(\forall x)(\forall y)(x\not=y\rightarrow f(x)\not=f(y)) saying that ff is injective has limiting density 0, as shown in Section 3.2. Thus Property (b) fails in this variety. ∎

6.1.4 A new example

In the next example, we modify the variety of bijective structures to obtain an example in which Property (b) holds but Property (a) fails.

Example 6.3.

Let LcL_{c} be the language that consists of unary function symbols SS, S1S^{-1} and a constant cc, and let VV be the variety with axioms saying that SS and S1S^{-1} are inverse functions and S3(c)=cS^{3}(c)=c. Consider presentations with a single generator aa and one identity. For the resulting structure AA, let AaA_{a} be the cycle generated by aa and let AcA_{c} be the cycle generated by cc. We describe the structures obtained from all possible identities, and we give some limiting densities.

  1. 1.

    Let S1S_{1} be the set of identities of the form Sn(a)=Sm(a)S^{n}(a)=S^{m}(a). This has density 14\frac{1}{4}. If k=|mn|k=|m-n|, then AaA_{a} is a kk-cycle if k>0k>0 and a \mathbb{Z}-chain if k=0k=0. AcA_{c} is always a 33-cycle in this case.

  2. 2.

    Let S2S_{2} be the set of identities of the form Sn(c)=Sm(c)S^{n}(c)=S^{m}(c). This has density 14\frac{1}{4}. Then AaA_{a} is a \mathbb{Z}-chain (always the same), and AcA_{c} is a 33-cycle or a 11-cycle.

  3. 3.

    Let S3S_{3} be the set of identities of the form Sn(a)=Sm(c)S^{n}(a)=S^{m}(c) or Sn(c)=Sm(a)S^{n}(c)=S^{m}(a). This set has density 12\frac{1}{2}. In the resulting structure, Aa=AcA_{a}=A_{c} is a 33-cycle.

To see that Property (a) fails, consider the sentence cS(c)c\neq S(c). This is true in the free structure but fails exactly in the subset of S2S_{2} where AcA_{c} is a 1-cycle, which has limiting density 16\frac{1}{6}.

To show that Property (b) holds, assume for some φ(x)\varphi(x), φ(t)\varphi(t) has limiting density 1 for all closed terms t=t(c,a)t=t(c,a). We will show that for i=1,2,3i=1,2,3, the set of identities in SiS_{i} for which the resulting structure satisfies (x)φ(x)(\forall x)\varphi(x) has the same density as SiS_{i}. For any finite set σ\sigma of closed terms t=t(c)t=t(c) or t=t(a)t=t(a), the sentence ψ(c,a)=tσφ(t)\psi(c,a)=\bigwedge_{t\in\sigma}\varphi(t) has density 11. This makes the case S3S_{3} easy. For the structures given by identities in S3S_{3}, all xx are named by terms cc, S(c)S(c), or S2(c)S^{2}(c).

For the remaining cases, we use Gaifman’s Locality Theorem. Consider a formula φ(u,v,x)\varphi^{\prime}(u,v,x) in the language of bijective structures such that φ(x)=φ(c,a,x)\varphi(x)=\varphi^{\prime}(c,a,x). By Gaifman, φ(u,v,x)\varphi^{\prime}(u,v,x) is equivalent in bijective structures to a formula i(αi(u,v,x)&βi)\bigvee_{i}(\alpha_{i}(u,v,x)\ \&\ \beta_{i}), where for some rr, βi\beta_{i} is a conjunction of local sentences and negations, each rr^{\prime}-local for rrr^{\prime}\leq r, and αi(u,v,x)\alpha_{i}(u,v,x) is an rr-local formula that describes the union of the rr-balls around u,v,xu,v,x.

For identities in S2S_{2}, AcA_{c} may have one element or three, and AaA_{a} is fixed. Let σ\sigma be a finite set with closed terms naming the elements of AcA_{c} and the elements of AaA_{a} that are not far from aa, with d(x,a)2rd(x,a)\leq 2r, plus one more element x=t(a)x=t^{*}(a) where d(x,a)>2rd(x,a)>2r. The sentence ψ(a,c)\psi(a,c) saying that φ(t)\varphi(t) holds for all of these terms has density 11. For the other xAax\in A_{a}, the ones far from aa, the balls Br(x)B_{r}(x) are isomorphic. If t(a)t^{*}(a) satisfies αi(c,a,x)\alpha_{i}(c,a,x), then all elements do, so φ(x)\varphi(x) holds. Then (x)φ(x)(\forall x)\varphi(x) has density 11.

Finally, for an identity in S1S_{1}, AcA_{c} is a fixed 33-cycle, while AaA_{a} varies with the identity. Consider the disjuncts αi(c,a,x)&βi\alpha_{i}(c,a,x)\ \&\ \beta_{i} that might be satisfied by some xAax\in A_{a}. The same identities also yield plain bijective structures AaA_{a}. Let αi(a,x)\alpha^{\prime}_{i}(a,x) be the part of αi(c,a,x)\alpha_{i}(c,a,x) describing the rr-balls around aa and xx. For x=t(a)x=t(a), xx satisfies αi(c,a,x)&βi\alpha_{i}(c,a,x)\ \&\ \beta_{i}, iff βi\beta_{i} holds in AaZ3A_{a}\cup Z_{3} and αi(a,x)\alpha^{\prime}_{i}(a,x) holds in AaA_{a}. For each βi\beta_{i}, there is a sentence βi\beta^{\prime}_{i} such that for the structures AA given by an identity in S1S_{1}, AβiA\models\beta_{i} iff AaβiA_{a}\models\beta_{i}^{\prime}. We may take βi\beta^{\prime}_{i} to be a finite disjunction of conjunctions of sentences that, in the setting of bijective structures, are rr^{\prime}-local sentences or negations. We can see this by using the Feferman-Vaught Theorem or, less formally, just by thinking about what βi\beta_{i} says. Let φ(a,x)=iαi(a,x)&βi\varphi^{\prime}(a,x)=\bigvee_{i}\alpha^{\prime}_{i}(a,x)\ \&\ \beta^{\prime}_{i}. For x=t(a)x=t(a), the formula φ(a,x)\varphi^{\prime}(a,x) has density 11 in the bijective structure AaA_{a}. Then by our earlier result, (x)φ(x)(\forall x)\varphi^{\prime}(x) has density 11. For each bijective structure generated by aa in which (x)φ(x)(\forall x)\varphi^{\prime}(x) is true, we consider the structure to be AaA_{a}, and in the variety we are currently considering, A=AaZ3A=A_{a}\cup Z_{3} satisfies (x)φ(x)(\forall x)\varphi(x), so this has density 11.

This example also shows that the zero–one law may fail if we allow constants in the language, giving an obstacle for generalizing Theorem 4.26 (on the zero–one law for generalized bijective structures) to varieties in a language with constants. Note that in Section 5, we did add constants naming a tuple of generators. However, these constants were not part of the language of the variety—they did not appear in the axioms.

6.2 Structures with a single unary function and more generators and identities

For the language with a single unary function symbol ff and the variety with no axioms, we saw in Section 3.2 that for presentations with a single generator and a single identity, the zero–one law holds, but the limiting theory is not that of the free structure. In particular, the sentence φ\varphi saying that ff is not injective has density 11, but it is false in the free structure. We now consider presentations with multiple generators and identities.

Proposition 6.4.

Let LL be the language with a single unary function symbol ff, and let VV be the variety with no axioms. For presentations with mm generators and kk identities, the sentence φ=(x,y)(f(x)=f(y)&xy)\varphi=(\exists x,y)(f(x)=f(y)\ \&\ x\neq y) has density 11.

Let the generators be a1,,ama_{1},\ldots,a_{m}. The identities have the form fp(ai)=fq(aj)f^{p}(a_{i})=f^{q}(a_{j}). As before, the sentence φ\varphi is true if the chosen identities all satisfy that p,qp,q are both non-zero and pqp\neq q. Indeed, in this case, without loss of generality, we may take n,in,i such that fn(ai)f^{n}(a_{i}) appears as one side of some identity and there is no m<nm<n such that fm(ai)f^{m}(a_{i}) appears as one side of some identity. Suppose fn(ai)=fq(aj)f^{n}(a_{i})=f^{q}(a_{j}) is one of the identities. Then x=fn1(ai)x=f^{n-1}(a_{i}) and y=fq1(aj)y=f^{q-1}(a_{j}) witness φ\varphi.

We can show that φ\varphi has density 11. The number of identities of length rr is m2(r+1)m^{2}(r+1). The number of length at most ss is m2(1+2++(s+1))=m2(s+2)(s+1)2m^{2}(1+2+\cdots+(s+1))=m^{2}\frac{(s+2)(s+1)}{2}. The number of unordered sets of kk identities of length at most ss is Ps=(m2(s+2)(s+1)2k)P_{s}=\left(\begin{array}[]{c}m^{2}\frac{(s+2)(s+1)}{2}\\ k\end{array}\right). We count the identities of length rr such that p,q0p,q\neq 0 and pqp\neq q. If rr is even, then there are at most 3m23m^{2} identities of length rr for which the condition fails; namely, fr(ai)=ajf^{r}(a_{i})=a_{j}, ai=fr(aj)a_{i}=f^{r}(a_{j}), and fr/2(ai)=fr/2(aj)f^{r/2}(a_{i})=f^{r/2}(a_{j}). (If rr is odd, then the number is at most 2m22m^{2}.)

Thus, there are at least m2(r2)m^{2}(r-2) identities of length rr satisfying the condition, and there are at least m2(s1)(s2)2m^{2}\frac{(s-1)(s-2)}{2} identities of length at most ss satisfying the condition. Let AA be the set of presentations with all identities satisfying the condition. Then Ps(A)(m2(s1)(s2)2k)P_{s}(A)\geq\left(\begin{array}[]{c}m^{2}\frac{(s-1)(s-2)}{2}\\ k\end{array}\right).

It is now a calculus exercise to show that Ps(A)PS1\frac{P_{s}(A)}{P_{S}}\to 1, and the proposition follows.

f

Remark.

We saw that when m=k=1m=k=1, the zero–one law holds. However, it does not hold in the case where m=1m=1 and k=2k=2. Suppose the two identities are fp(a)=fq(a)f^{p}(a)=f^{q}(a), fp(a)=fq(a)f^{p^{\prime}}(a)=f^{q^{\prime}}(a), and consider the sentence ψ=(x)f(x)=x\psi=(\exists x)f(x)~=~x. This case is similar to the case of bijective structures with two identities in Section 3.3. The sentence ψ\psi is true if and only if GCD(pq,pq)=1GCD(p-q,p^{\prime}-q^{\prime})=1. An argument like that in Section 3.3 shows that ψ\psi has density strictly between 0 and 11. We omit the proof here.

6.3 Structures with multiple unary functions

We turn our attention to a more complicated case. Take the language with nn function symbols f1,,fnf_{1},\ldots,f_{n} and the variety with no axioms, and consider presentations with mm generators and kk identities. We begin with the case where k=1k=1.

Proposition 6.5.

Let φ\varphi be the sentence

(x)(y)(xy&1i,jnfi(x)=fj(y)).(\exists x)(\exists y)\left(x\not=y\ \&\ \bigvee_{1\leq i,j\leq n}f_{i}(x)=f_{j}(y)\right)\ .

This sentence is false in the free structure, but it has limiting density 11 among structures given by presentations with generators a1,,ama_{1},\ldots,a_{m} and a single identity of the form t(ai)=t(aj)t(a_{i})=t^{\prime}(a_{j}).

Proof.

For mm generators a1,,ama_{1},\ldots,a_{m}, the free structure FF is the join of disjoint substructures generated by the separate aja_{j}. In FF, each element is uniquely expressed as t(ai)t(a_{i}), where the term tt is built up out of the functions fjf_{j}. The terms are all distinct, and the sentence φ\varphi is false. For an identity t(ai)=t(aj)t(a_{i})=t^{\prime}(a_{j}), the length is the sum of the lengths of t,tt,t^{\prime}. The number of identities of length \ell is m2n(+1)m^{2}n^{\ell}(\ell+1), so the number of identities of length at most ss, or PsP_{s}, is m20sn(+1)m^{2}\sum_{0\leq\ell\leq s}n^{\ell}(\ell+1), which is equal to m2(n1)(s+2)ns+1(ns+21)(n1)2m^{2}\frac{(n-1)(s+2)n^{s+1}-(n^{s+2}-1)}{(n-1)^{2}}.

Let AA be the set of identities t(ai)=t(aj)t(a_{i})=t^{\prime}(a_{j}) such that tt has length 0. We show that AA has limiting density 0. The number of identities in AA of length \ell is m2nm^{2}n^{\ell}, so the number of length at most ss is m2(0sn)m^{2}(\sum_{0\leq\ell\leq s}n^{\ell}), or m2ns+11n1m^{2}\frac{n^{s+1}-1}{n-1}. This is Ps(A)P_{s}(A). It is not difficult to verify that limsPs(A)Ps=0\lim_{s\rightarrow\infty}\frac{P_{s}(A)}{P_{s}}=0. Similarly, let BB be the set of identities t(ai)=t(aj)t(a_{i})=t^{\prime}(a_{j}) such that tt^{\prime} has length 0. Then BB also has limiting density 0. Therefore, the limiting density of ABA\cup B is 0. Let CC be the set of identities not in ABA\cup B. This will have limiting density 11. The identities in CC have the form t(ai)=t(aj)t(a_{i})=t^{\prime}(a_{j}) where t,tt,t^{\prime} both have length at least 11. Say that t(ai)=fi(t(ai))t(a_{i})=f_{i^{\prime}}(t^{*}(a_{i})) and t(aj)=fj(t(aj))t^{\prime}(a_{j})=f_{j^{\prime}}(t^{\prime*}(a_{j})) for terms tt^{*} and tt^{\prime*}. In the model given by the identity t(ai)=t(aj)t(a_{i})=t^{\prime}(a_{j}), we have t(ai)t(aj)t^{*}(a_{i})\not=t^{\prime*}(a_{j}). The elements x=t(ai)x=t^{*}(a_{i}) and y=t(aj)y=t^{\prime*}(a_{j}) witness that the sentence φ\varphi is true. ∎

Now, we consider presentations with more than one identity. We let φ\varphi be as in Proposition 6.5.

Proposition 6.6.

For the language with nn unary function symbol f1,,fnf_{1},\dots,f_{n}, let VV be the variety with no axioms. For presentations with a fixed mm-tuple of generators and kk identities, where k2k\geq 2, the sentence φ\varphi has limiting density 11.

Proof.

Let IsI_{s} be the number of identities of length at most ss. Then the number of presentations in which all identities have length at most ss is Ps=(Isk)P_{s}=\left(\begin{array}[]{c}I_{s}\\ k\end{array}\right). Consider the identities t(ai)=t(aj)t(a_{i})=t^{\prime}(a_{j}) in which neither side has length 0. The number of these identities of length \ell, where 2\ell\geq 2, is m2n(1)m^{2}n^{\ell}(\ell-1), so the number of length at most ss is m22sn(1)=m2(n1)sns1(ns1))(n1)2m^{2}\sum_{2\leq\ell\leq s}n^{\ell}(\ell-1)=m^{2}\frac{(n-1)sn^{s-1}-(n^{s}-1))}{(n-1)^{2}}. For convenience, we call this JsJ_{s}. Let CC be the set of presentations with kk identities in which neither side has length 0. Then Ps(C)=(Jsk)P_{s}(C)=\left(\begin{array}[]{c}J_{s}\\ k\end{array}\right).

We show by induction on kk that limsPs(C)Ps=1\lim_{s\rightarrow\infty}\frac{P_{s}(C)}{P_{s}}=1. We write PskP_{s}^{k} and Psk(C)P_{s}^{k}(C) to indicate the value of kk under consideration. For k=2k=2, Ps2(C)Ps2=Js(Js1)Is(Is1)\frac{P_{s}^{2}(C)}{P_{s}^{2}}=\frac{J_{s}(J_{s}-1)}{I_{s}(I_{s}-1)}. We know that JsIs1\frac{J_{s}}{I_{s}}\rightarrow 1. For the expression Ps2(C)Ps2\frac{P_{s}^{2}(C)}{P_{s}^{2}}, we divide top and bottom both by IsI_{s} and get a new numerator JsIs1Is\frac{J_{s}}{I_{s}}-\frac{1}{I_{s}} that goes to 11 and a new denominator IsIs1Is\frac{I_{s}}{I_{s}}-\frac{1}{I_{s}} that also goes to 11. Now, supposing that the statement holds for kk, we show that it holds for k+1k+1. We have Psk+1(C)Psk=(Psk(C)Psk)(JskIsk)\frac{P_{s}^{k+1}(C)}{P_{s}^{k}}=\left(\frac{P_{s}^{k}(C)}{P_{s}^{k}}\right)\left(\frac{J_{s}-k}{I_{s}-k}\right). By the Induction Hypothesis, the first factor goes to 11. For the second factor, we again divide top and bottom by IsI_{s}. The new numerator is JsIskIs\frac{J_{s}}{I_{s}}-\frac{k}{I_{s}}, which has limit 11. The new denominator is IsIskIs\frac{I_{s}}{I_{s}}-\frac{k}{I_{s}}, which also has limit 11.

We claim that the sentence φ\varphi is true in all structures obtained from presentations in CC. Take any presentation in CC and consider the resulting model 𝒜\mathcal{A}. No aia_{i} is in the range of any function in any model of this sort. The given identities all take us from a non-trivial term in some aia_{i} to a non-trivial term in some aja_{j} and do not force us to assign values aia_{i}, so we can fill out the rest of the function values without ever using these values aia_{i}. Thus, all nontrivial identities true in 𝒜\mathcal{A} are in all structures with presentations in CC. Take an identity of shortest length, say t(ai)=t(aj)t(a_{i})=t^{\prime}(a_{j}), and proceed as for a single identity. Say that t(ai)=fi(t(ai))t(a_{i})=f_{i^{\prime}}(t^{*}(a_{i})) and t(aj)=fj(t(aj))t^{\prime}(a_{j})=f_{j^{\prime}}(t^{\prime*}(a_{j})) for terms tt^{*} and tt^{\prime*}. By the minimality of the length of t(ai)=t(aj)t(a_{i})=t^{\prime}(a_{j}), we have t(ai)t(aj)t^{*}(a_{i})\not=t^{\prime*}(a_{j}). This witnesses the truth of φ\varphi. ∎

References

  • [1] G. N. Arzhantseva. On groups in which subgroups with a fixed number of generators are free. Fundam. Prikl. Mat., 3(3):675–683, 1997.
  • [2] G. N. Arzhantseva and A. Yu. Ol’shanskiĭ. Generality of the class of groups in which subgroups with a lesser number of generators are free. Mat. Zametki, 59(4):489–496, 638, 1996.
  • [3] Stanley Burris and H. P. Sankappanavar. A course in universal algebra, volume 78 of Graduate Texts in Mathematics. Springer-Verlag, New York-Berlin, 1981.
  • [4] Matthew Cordes, Moon Duchin, Yen Duong, Meng-Che Ho, and Andrew P. Sánchez. Random nilpotent groups I. Int. Math. Res. Not. IMRN, (7):1921–1953, 2018.
  • [5] Shagnik Das. A brief note on estimates of binomial coefficients. http://page.mi.fu-berlin.de/shagnik/notes/binomials.pdf. Accessed March 1, 2022.
  • [6] Paul C. Eklof and Edward R. Fischer. The elementary theory of abelian groups. Ann. Math. Logic, 4:115–171, 1972.
  • [7] Haim Gaifman. On local and nonlocal properties. In Proceedings of the Herbrand symposium (Marseilles, 1981), volume 107 of Stud. Logic Found. Math., pages 105–135. North-Holland, Amsterdam, 1982.
  • [8] M. Gromov. Hyperbolic groups. In Essays in group theory, volume 8 of Math. Sci. Res. Inst. Publ., pages 75–263. Springer, New York, 1987.
  • [9] Matthew Harrison-Trainor, Bakh Khoussainov, and Daniel Turetsky. Effective aspects of algorithmically random structures. Computability, 8(3-4):359–375, 2019.
  • [10] Meng-Che Ho. Randomizing and Describing Groups. ProQuest LLC, Ann Arbor, MI, 2017. Thesis (Ph.D.)–The University of Wisconsin - Madison.
  • [11] Ilya Kapovich and Paul Schupp. Genericity, the Arzhantseva-Olshanskii method and the isomorphism problem for one-relator groups. Math. Ann., 331(1):1–19, 2005.
  • [12] Olga Kharlampovich and Alexei Myasnikov. Elementary theory of free non-abelian groups. J. Algebra, 302(2):451–552, 2006.
  • [13] Olga Kharlampovich and Rizos Sklinos. First-order sentences in random groups. arXiv preprint arXiv:2106.05461, 2022.
  • [14] Phokion G. Kolaitis. On the expressive power of logics on finite models. In Finite model theory and its applications, Texts Theoret. Comput. Sci. EATCS Ser., pages 27–123. Springer, Berlin, 2007.
  • [15] Leonid Libkin. Elements of finite model theory. Texts in Theoretical Computer Science. An EATCS Series. Springer-Verlag, Berlin, 2004.
  • [16] David Marker. Model theory, volume 217 of Graduate Texts in Mathematics. Springer-Verlag, New York, 2002. An introduction.
  • [17] Yann Ollivier. A January 2005 invitation to random groups, volume 10 of Ensaios Matemáticos [Mathematical Surveys]. Sociedade Brasileira de Matemática, Rio de Janeiro, 2005.
  • [18] A. Yu. Ol’shanskiĭ. Almost every group is hyperbolic. Internat. J. Algebra Comput., 2(1):1–17, 1992.
  • [19] Laurent Saloff-Coste. Random walks on finite groups. In Probability on discrete structures, volume 110 of Encyclopaedia Math. Sci., pages 263–346. Springer, Berlin, 2004.
  • [20] Z. Sela. Diophantine geometry over groups. VI. The elementary theory of a free group. Geom. Funct. Anal., 16(3):707–730, 2006.
  • [21] W. Szmielew. Elementary properties of Abelian groups. Fund. Math., 41:203–271, 1955.
  • [22] A. W. van der Vaart. Asymptotic statistics, volume 3 of Cambridge Series in Statistical and Probabilistic Mathematics. Cambridge University Press, Cambridge, 1998.