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

A Simple Combinatorial Proof of Szemerédi’s Theorem via Three Levels of InfinitiesMathematics Subject Classification 2020: Primary 11B25, Secondary 03H05

Renling Jin This work was partially supported by a collaboration grant (ID: 513023) from Simons Foundation.
Abstract

We present a nonstandard simple elementary proof of Szemerédi’s theorem by a straightforward induction with the help of three levels of infinities and four different bounded elementary embeddings in a nonstandard universe.

\dajAUTHORdetails

title = A Simple Combinatorial Proof of Szemerédi’s Theorem via Three Levels of Infinities, author = Renling Jin, plaintextauthor = Renling Jin, plaintexttitle = A Simple Combinatorial Proof of Szemerédi’s Theorem via Three Levels of Infinities, runningtitle = A Simple Proof, runningauthor = Renling Jin, copyrightauthor = Renling Jin, keywords = arithmetic progression, Szemerédi’s theorem, nonstandard analysis, iterated nonstandard extensions, \dajEDITORdetailsyear=2023, number=15, received=11 April 2022, revised=5 July 2023, published=25 September 2023, doi=10.19086/da.87772,

[classification=text]

1 Introduction

This article is strongly influenced by Terence Tao’s notes [9].

Theorem 1.1 (van der Waerden, 1927)

Given any k,nk,n\in{\mathbb{N}}, there exists Γ(k,n)\Gamma(k,n)\in{\mathbb{N}} called van der Waerden number, such that if {U1,U2,,Un}\{U_{1},U_{2},\ldots,U_{n}\} is a partition of {1,2,,Γ(k,n)}\{1,2,\ldots,\Gamma(k,n)\}, then there is a unu\leq n such that UuU_{u} contains a kk–term arithmetic progression.

Theorem 1.2 (E. Szemerédi, 1975 [8])

If DD\subseteq{\mathbb{N}} has a positive upper density, then DD contains a kk–term arithmetic progression for every kk\in{\mathbb{N}}.

Szemerédi’s theorem confirms a conjecture of P. Erdős and P. Turán made in 1936, which implies van der Waerden’s theorem.

Nonstandard versions of Furstenberg’s ergodic proof and Gowers’s harmonic proof of Szemerédi’s theorem have been tried by T. Tao (see Tao’s blog post [10]). In the workshop Nonstandard methods in combinatorial number theory sponsored by American Institute of Mathematics in San Jose, CA, August 2017, Tao gave a series of lectures to explain Szemerédi’s original combinatorial proof and hope to simplify it so that the proof can be better understood. He believed that Szemerédi’s combinatorial method should have a greater impact on combinatorics.

During these lectures Tao challenged the audience to produce a nonstandard proof of Szemerédi’s theorem which is noticeably simpler and more transparent than Szemerédi’s original proof. The current article is the result of Tao’s challenge and inspiration. However, in his later blog post [11], Tao commented that “in fact there are now signs that perhaps nonstandard analysis is not the optimal framework in which to place this argument.” We disagree. The current article is our effort to show that with the help of a nonstandard universe with three levels of infinities, Szemerédi’s original argument can be made simpler and more transparent.

The main simplification in our proof of Szemerédi’s theorem compared to the standard proof in [8, 9] is that a Tower of Hanoi type induction in [9, Theorem 6.6] and in [8, Lemma 5, Lemma 6, and Fact 12] is replaced by a straightforward induction (see Lemma 5.1 below), which makes Szemerédi’s idea more transparent. To achieve this, we work within a chain of nonstandard extensions 𝕍0𝕍1𝕍2𝕍3{\mathbb{V}}_{0}\prec{\mathbb{V}}_{1}\prec{\mathbb{V}}_{2}\prec{\mathbb{V}}_{3} which supply three levels of infinities, plus various bounded elementary embeddings from 𝕍j{\mathbb{V}}_{j} to 𝕍j{\mathbb{V}}_{j^{\prime}} for some 0j<j30\leq j<j^{\prime}\leq 3.

The paper is organized in the following sections. §2 is a brief introduction of logic foundation for constructing the nonstandard extensions 𝕍0𝕍1𝕍2𝕍3{\mathbb{V}}_{0}\prec{\mathbb{V}}_{1}\prec{\mathbb{V}}_{2}\prec{\mathbb{V}}_{3} and bounded elementary embeddings including i0i_{0}, ii_{*}, i1i_{1}, and i2i_{2} to the reader who does not have logic background. The reader who is only interested in applications can familiarize with the notation, Property 2.7, Proposition 2.14, and Proposition 2.15, safely skip the proof, and return to it at a later time. In §3 we translate the density along arithmetic progressions in standard setting to strong upper Banach density in nonstandard setting as well as translate some consequences of the double counting argument in standard setting to nonstandard setting. The reader who is only interested in applications can familiarize with the notation, Lemma 3.3, Lemma 3.4, and Lemma 3.5, safely skip the proof, and return to it at a later time. In §4 we re-write, in a nonstandard setting, the proof of a so–called mixing lemma in [9] based on a weak regularity lemma. This section does not offer new idea but is included only for self-containment. §5 is the main part of the paper where we present the proof of Szemerédi’s theorem. In §6 we pose a question whether the presented proof of Szmerédi’s theorem can be carried out without the axiom of choice.

2 Construction of Nonstandard Extensions

The notation we use here should be consistent with some standard textbooks. Consult, for example, [1, 4, 7] for more details. If f:ABf:A\to B is a function, then f(a)f(a) denotes the image of aa as an element in BB and f[C]:={f(a)aC}f[C]:=\{f(a)\mid a\in C\} for some CAC\subseteq A.


§2.1 Superstructure Let ω\omega be the set of all standard non-negative integers used at the meta-level and XX be an infinite set of urelements, i.e., elements without members. The superstructure on XX (cf. [1, page 263]), denoted by 𝕍(X){\mathbb{V}}(X), is composed of the base set V(X)V(X) and the membership relation \in on V(X)V(X) where

V(X):=nωV(X,n)V(X):=\bigcup_{n\in\omega}V(X,n)

and V(X,n)V(X,n) is defined recursively by letting

V(X,0):=X and V(X,n+1):=V(X,n)𝒫(V(X,n))V(X,0):=X\,\mbox{ and }\,V(X,n+1):=V(X,n)\cup\mathscr{P}(V(X,n))

for every nωn\in\omega where 𝒫\mathscr{P} is the powerset operator. For notational convenience we often write 𝕍(X){\mathbb{V}}(X) also for its base set V(X)V(X). Hopefully, this will not cause confusion. One can define a rank function for every a𝕍(X)a\in{\mathbb{V}}(X) recursively. Set rank(a)=0\mbox{rank}(a)=0 iff (the abbreviation of “if and only if”) aXa\in X and

rank(a)=1+max{rank(b)ba}\mbox{rank}(a)=1+\max\{\mbox{rank}(b)\mid b\in a\} (1)

for any a𝕍(X)Xa\in{\mathbb{V}}(X)\setminus X. Notice that rank(a)=n\mbox{rank}(a)=n iff a𝕍(X,n)𝕍(X,n1)a\in{\mathbb{V}}(X,n)\setminus{\mathbb{V}}(X,n-1).

Let 0{\mathbb{N}}_{0} be the set of all standard positive integers and 0{\mathbb{R}}_{0} be the set of all standard real numbers. By the standard universe we mean the superstructure 𝕍0:=𝕍(0){\mathbb{V}}_{0}:={\mathbb{V}}({\mathbb{R}}_{0}) on X=0X={\mathbb{R}}_{0}. Notice that all standard mathematical objects mentioned in this paper have ranks below, say, 100100. For example, an ordered pair (a,b)(a,b) of standard real numbers can be viewed as the set {{a},{a,b}}𝕍(0,2)\{\{a\},\{a,b\}\}\in{\mathbb{V}}({\mathbb{R}}_{0},2) and a function f:00f:{\mathbb{N}}_{0}\to{\mathbb{R}}_{0} can be viewed as a set of ordered pairs in 𝕍(0,2){\mathbb{V}}({\mathbb{R}}_{0},2). Hence f𝕍(0,3)f\in{\mathbb{V}}({\mathbb{R}}_{0},3).


§2.2 Logic and model theory Before introducing nonstandard universe we should mention briefly, without rigor, some concepts in model theory. For simplicity we consider only model theory on finite relational languages. We call a set \mathscr{L} of finitely many symbols P1,P2,,PnP_{1},P_{2},\ldots,P_{n} with arity miωm_{i}\in\omega for each PiP_{i} a (relational) language. An \mathscr{L}–model \mathcal{M} is a structure composed of a nonempty base set MM and an mim_{i}-nary relation PiMmiP^{\mathcal{M}}_{i}\subseteq M^{m_{i}}, called the interpretation of PiP_{i} in \mathcal{M}, for each symbol PiP_{i}\in\mathscr{L}. For notational convenience we sometimes write \mathcal{M} for the base set of \mathcal{M}.

We can define (first-order) \mathscr{L}–formulas recursively starting from atomic formulas. If xx and xx^{\prime} are variables, then x=xx=x^{\prime} is an atomic formula. If PiP_{i}\in\mathscr{L} and x¯:=(x1,x2,,xmi)\overline{x}:=(x_{1},x_{2},\ldots,x_{m_{i}}) is an mim_{i}-tuple of variables, then Pi(x¯)P_{i}(\overline{x}) is an atomic formula. The word “first-order” means that these variables are intended to take only elements of some \mathscr{L}–model as their values. All formulas mentioned in this paper are first-order. We will use the symbol a¯\overline{a} to represent an mm-tuple of elements with some suitable generic number mωm\in\omega. When x¯\overline{x} is intended to be substituted by a¯\overline{a}, we assume implicitly that they have the same length.

If φ\varphi and ψ\psi are \mathscr{L}–formulas and xx is a variable, then ¬φ\neg\varphi, φψ\varphi\wedge\psi, φψ\varphi\vee\psi, φψ\varphi\to\psi, φψ\varphi\leftrightarrow\psi, xφ\forall x\varphi, and xφ\exists x\varphi are also \mathscr{L}–formulas. The symbols ¬\neg, \wedge, \vee, \to, and \leftrightarrow are called logic connectives, and \forall and \exists are called universal and existential quantifiers. An occurrence of xx in φ\varphi is called bounded if it occurs in a sub-formula of the form xψ\forall x\psi or xψ\exists x\psi in φ\varphi. An occurrence of xx in φ\varphi is called free if it is not bounded. A formula without free variable is called a sentence.

Let \mathcal{M} be an \mathscr{L}–model. For any \mathscr{L}–formula φ(x¯)\varphi(\overline{x}), where x¯\overline{x} is an mm-tuple of variables containing all free variables in φ\varphi, and any a¯m\overline{a}\in\mathcal{M}^{m}, called parameters, we can define φ(a¯)\mathcal{M}\models\varphi(\overline{a}), meaning φ(a¯)\varphi(\overline{a}) is true in \mathcal{M}, recursively by (a) a=a\mathcal{M}\models a=a^{\prime} iff aa and aa^{\prime} are identical elements in \mathcal{M} and Pi(a¯)\mathcal{M}\models P_{i}(\overline{a}) iff a¯Pi\overline{a}\in P_{i}^{\mathcal{M}}; (b) for any \mathscr{L}–formulas φ\varphi and ψ\psi with all free variables x¯\overline{x} being substituted by a¯\overline{a} in \mathcal{M}, ¬φ\mathcal{M}\models\neg\varphi iff ⊧̸φ\mathcal{M}\not\models\varphi (¬\neg means “not”), φψ\mathcal{M}\models\varphi\vee\psi iff φ\mathcal{M}\models\varphi or ψ\mathcal{M}\models\psi (\vee means “or”), φψ\mathcal{M}\models\varphi\wedge\psi iff φ\mathcal{M}\models\varphi and ψ\mathcal{M}\models\psi, (\wedge means “and”), φψ\mathcal{M}\models\varphi\to\psi iff ¬φψ\mathcal{M}\models\neg\varphi\vee\psi (\to means “imply”), and φψ\mathcal{M}\models\varphi\leftrightarrow\psi iff (φψ)(ψφ)\mathcal{M}\models(\varphi\to\psi)\wedge(\psi\to\varphi) (\leftrightarrow means “if and only if”); (c) for any \mathscr{L}–formulas φ(y,x¯)\varphi(y,\overline{x}), yφ(y,a¯)\mathcal{M}\models\forall y\,\varphi(y,\overline{a}) iff φ(b,a¯)\mathcal{M}\models\varphi(b,\overline{a}) for every bb\in\mathcal{M} (\forall means “for all”) and yφ(y,a¯)\mathcal{M}\models\exists y\varphi(y,\overline{a}) iff φ(b,a¯)\mathcal{M}\models\varphi(b,\overline{a}) for some bb\in\mathcal{M} (\exists means “there exist”). From the truth definition above, for every \mathscr{L}–formula φ(x¯)\varphi(\overline{x}) there is another \mathscr{L}–formula ψ(x¯)\psi(\overline{x}) using only logic connectives ¬\neg and \vee, and only quantifier \exists, such that φ(a¯)\mathcal{M}\models\varphi(\overline{a}) iff ψ(a¯)\mathcal{M}\models\psi(\overline{a}) for any \mathscr{L}–model \mathcal{M} and any a¯m\overline{a}\in\mathcal{M}^{m}.

We sometimes call a formula with all free variables being substituted by parameters a sentence. Clearly, the truth value of a sentence with parameters from \mathcal{M} is determined in \mathcal{M}. When we write φ(x¯,a¯)\varphi(\overline{x},\overline{a}), we mean implicitly that all free variables in the formula are among x¯\overline{x} and all parameters from model \mathcal{M} are among a¯\overline{a}.

Suppose \mathcal{M} and 𝒩\mathcal{N} are two \mathscr{L}–models. A function i:𝒩i:\mathcal{M}\to\mathcal{N} is called an elementary embedding from \mathcal{M} to 𝒩\mathcal{N} if for any \mathscr{L}–formula φ(x¯)\varphi(\overline{x}) and any mm-tuple a¯=(a1,a2,,am)m\overline{a}=(a_{1},a_{2},\ldots,a_{m})\in\mathcal{M}^{m} we have

φ(a¯) iff 𝒩φ(i(a¯))\mathcal{M}\models\varphi(\overline{a})\,\mbox{ iff }\,\mathcal{N}\models\varphi(i(\overline{a})) (2)

where i(a¯):=(i(a1),i(a2),,i(am))𝒩mi(\overline{a}):=(i(a_{1}),i(a_{2}),\ldots,i(a_{m}))\in\mathcal{N}^{m}. An elementary embedding is necessarily injective. If there exists an elementary embedding i:𝒩i:\mathcal{M}\to\mathcal{N}, we can view \mathcal{M} as an elementary submodel of 𝒩\mathcal{N} and call 𝒩\mathcal{N} an elementary extension of \mathcal{M}, denoted by 𝒩\mathcal{M}\preceq\mathcal{N}. We sometimes write 𝒩\mathcal{M}\prec\mathcal{N} to emphasize that ii is not surjective.

If \mathcal{M} is an \mathscr{L}–model, :={Pn+1,,Pk}\mathscr{L}^{\prime}:=\mathscr{L}\cup\{P_{n+1},\ldots,P_{k}\}, and PimiP_{i}^{\mathcal{M}}\subseteq\mathcal{M}^{m_{i}} for n<ikn<i\leq k, the \mathscr{L}^{\prime}–model \mathcal{M^{\prime}} by adding the relations PiP_{i}^{\mathcal{M}} to \mathcal{M} is denoted by (;Pn+1,,Pk)(\mathcal{M};P^{\mathcal{M}}_{n+1},\ldots,P^{\mathcal{M}}_{k}). We call \mathcal{M}^{\prime} a model expansion of \mathcal{M}.


§2.3 Ultrapower construction Next we construct an elementary extension of a model using ultrapower construction (cf. [1, §4]).

Definition 2.1

Let XX be an infinite set. A set 𝒫(X){\mathcal{F}}\subseteq\mathscr{P}(X) is a non-principal ultrafilter on XX if it satisfies the following:

  1. 1.

    XX\in{\mathcal{F}} and FF\not\in{\mathcal{F}} for any F𝒫<ω(X)F\in\mathscr{P}_{<\omega}(X) where 𝒫<ω(X)\mathscr{P}_{<\omega}(X) is the collection of all finite subsets of XX,

  2. 2.

    A,B𝒫(X)(A,B implies AB)\forall A,B\in\mathscr{P}(X)\,(A,B\in{\mathcal{F}}\,\mbox{ implies }\,A\cap B\in{\mathcal{F}}),

  3. 3.

    A,B𝒫(X)(A and AB imply B)\forall A,B\in\mathscr{P}(X)\,(A\in{\mathcal{F}}\,\mbox{ and }\,A\subseteq B\,\mbox{ imply }\,B\in{\mathcal{F}}),

  4. 4.

    A𝒫(X)(A or (XA))\forall A\in\mathscr{P}(X)\,(A\in{\mathcal{F}}\,\mbox{ or }\,(X\setminus A)\in{\mathcal{F}}).

It is well known that the existence of non-principal ultrafilters on XX follows from 𝖹𝖥𝖢\mathsf{ZFC}.

Definition 2.2

Let \mathcal{M} be an \mathscr{L}–model and {\mathcal{F}} be a non-principal ultrafilter on an infinite set XX. Let X\mathcal{M}^{X} be the set of all functions from XX to \mathcal{M}. For any f,gXf,g\in\mathcal{M}^{X} define

fg iff {nXf(n)=g(n)}.f\sim_{{\mathcal{F}}}g\,\mbox{ iff }\,\{n\in X\mid f(n)=g(n)\}\in{\mathcal{F}}.

It is easy to check that \sim_{{\mathcal{F}}} is an equivalence relation. Denote [f][f]_{{\mathcal{F}}} for the equivalence class containing ff. The ultrapower of \mathcal{M} modulo {\mathcal{F}}, denoted by X/\mathcal{M}^{X}/{\mathcal{F}}, is an \mathscr{L}–model which is composed of the base set

{[f]fX}\{[f]_{{\mathcal{F}}}\mid f\in\mathcal{M}^{X}\}

and the interpretation of PiP_{i} by

PiX/:={([f1],[f2],,[fmi]){nX(f1(n),f2(n),,fmi(n))Pi}}P_{i}^{\mathcal{M}^{X}/{\mathcal{F}}}:=\left\{([f_{1}]_{{\mathcal{F}}},[f_{2}]_{{\mathcal{F}}},\ldots,[f_{m_{i}}]_{{\mathcal{F}}})\mid\{n\in X\mid(f_{1}(n),f_{2}(n),\ldots,f_{m_{i}}(n))\in P_{i}^{\mathcal{M}}\}\in{\mathcal{F}}\right\}

for every PiP_{i}\in\mathscr{L}. Notice that

PiX/={[f¯]f¯ is a function from X to Pi}.P_{i}^{\mathcal{M}^{X}/{\mathcal{F}}}=\{[\overline{f}]_{{\mathcal{F}}}\mid\overline{f}\,\mbox{ is a function from }\,X\,\mbox{ to }\,P_{i}^{\mathcal{M}}\}.

The right side above is the ultrapower of PiP_{i}^{\mathcal{M}} modulo {\mathcal{F}}.

For an element cc\in\mathcal{M} denote ϕc:X\phi_{c}:X\to\mathcal{M} for the constant function with value cc. Let i:X/i:\mathcal{M}\to\mathcal{M}^{X}/{\mathcal{F}} be the natural embedding, i.e., i(c)=[ϕc]i(c)=[\phi_{c}]_{{\mathcal{F}}}. The following is often called Łoś’s theorem. (cf. [1, Theorem 4.1.9, Corollary 4.1.13].)

Proposition 2.3

For any \mathscr{L}–formula φ(x¯)\varphi(\overline{x}) and any [a]¯X/\overline{[a]}_{{\mathcal{F}}}\in\mathcal{M}^{X}/{\mathcal{F}} it is true that

X/φ([a]¯) iff {nXφ(a(n)¯)}.\mathcal{M}^{X}/{\mathcal{F}}\models\varphi(\overline{[a]}_{{\mathcal{F}}})\,\mbox{ iff }\,\left\{n\in X\mid\mathcal{M}\models\varphi(\overline{a(n)})\right\}\in{\mathcal{F}}.

The proof of Proposition 2.3 is done by induction on the complexity of φ\varphi.

Corollary 2.4

The natural embedding i:X/i:\mathcal{M}\to\mathcal{M}^{X}/{\mathcal{F}} with i(c)=[ϕc]i(c)=[\phi_{c}]_{{\mathcal{F}}} is an elementary embedding. Furthermore, if =(;R)\mathcal{M}^{\prime}=(\mathcal{M};R) is a model expansion of \mathcal{M}, then ii is also an elementary embedding from \mathcal{M}^{\prime} to X/=(X/;RX/)\mathcal{M}^{\prime X}/{\mathcal{F}}=(\mathcal{M}^{X}/{\mathcal{F}};R^{X}/{\mathcal{F}}).

So, the model \mathcal{M} can be viewed as an elementary submodel of X/\mathcal{M}^{X}/{\mathcal{F}} via the natural embedding ii and X/\mathcal{M}^{X}/{\mathcal{F}} is an elementary extension of \mathcal{M}.


§2.4 Construction of 𝕍1{\mathbb{V}}_{1} Fix a non-principal ultrafilter 0{\mathcal{F}}_{0} on 0{\mathbb{N}}_{0}.

\blacklozenge: From now on let ={}\mathscr{L}=\{\in\}. The ultrafilters which will be used are 0{\mathcal{F}}_{0}, the tensor product of 0{\mathcal{F}}_{0}’s, and nonstandard versions of them.

Recall that 𝕍0{\mathbb{V}}_{0} represents the standard universe, which is an \mathscr{L}–model. Let 𝕍00/0{\mathbb{V}}_{0}^{{\mathbb{N}}_{0}}/{\mathcal{F}}_{0} be the ultrapower of 𝕍0{\mathbb{V}}_{0} modulo 0{\mathcal{F}}_{0} and i0:𝕍0𝕍00/0i^{0}:{\mathbb{V}}_{0}\to{\mathbb{V}}_{0}^{{\mathbb{N}}_{0}}/{\mathcal{F}}_{0} be the natural embedding. Denote {}^{*}\!\!\in for the interpretation of \in in 𝕍00/0{\mathbb{V}}_{0}^{{\mathbb{N}}_{0}}/{\mathcal{F}}_{0}. Notice that aba\in b iff i0(a)i0(b)i^{0}(a)\,^{*}\!\!\in i^{0}(b) for any a,b𝕍0a,b\in{\mathbb{V}}_{0}. One can define the rank function rank(b)\mbox{rank}(b) for every b𝕍00/0b\in{\mathbb{V}}_{0}^{{\mathbb{N}}_{0}}/{\mathcal{F}}_{0} according to (1) with \in being replaced by {}^{*}\!\!\in. Notice that some elements [f][f]_{{\mathcal{F}}} in 𝕍00/0{\mathbb{V}}_{0}^{{\mathbb{N}}_{0}}/{\mathcal{F}}_{0} may not have a finite {}^{*}\!\!\in-rank. For example, if f(1)=0f(1)=0 and f(n+1)={f(n)}f(n+1)=\{f(n)\} for every n0n\in{\mathbb{N}}_{0}, then [f]0𝕍00/0[f]_{{\mathcal{F}}_{0}}\in{\mathbb{V}}_{0}^{{\mathbb{N}}_{0}}/{\mathcal{F}}_{0} does not have a finite {}^{*}\!\!\in-rank. Let

V1:={b𝕍00/0-rank(b)ω}.V_{1}:=\left\{b\in{\mathbb{V}}_{0}^{{\mathbb{N}}_{0}}/{\mathcal{F}}_{0}\mid\,^{*}\!\!\in\!\mbox{-rank}(b)\in\omega\right\}.

The set V1V_{1} is just the ultrapower of 𝕍0{\mathbb{V}}_{0} modulo 0{\mathcal{F}}_{0} truncated at {}^{*}\!\!\in-rank ω\omega. Notice that -rank(a)=-rank(i0(a))\in\!\mbox{-rank}(a)=\,^{*}\!\!\in\!\mbox{-rank}(i^{0}(a)) for every a𝕍0a\in{\mathbb{V}}_{0} and V1=nωi0(V(0,n))V_{1}=\bigcup_{n\in\omega}i^{0}(V({\mathbb{R}}_{0},n)).

Let 1:=i0(0){\mathbb{R}}_{1}:=i^{0}({\mathbb{R}}_{0}) and 1:=i0(0){\mathbb{N}}_{1}:=i^{0}({\mathbb{N}}_{0}). Assume that every element in 1{\mathbb{R}}_{1} is an urelement and identify i0(r)i^{0}(r) by rr for every r0r\in{\mathbb{R}}_{0}. Then 01{\mathbb{R}}_{0}\subseteq{\mathbb{R}}_{1}. Since the natural order \leq on 0{\mathbb{R}}_{0}, addition ++ and multiplication ×\times on 0{\mathbb{R}}_{0} can be viewed as elements in 𝕍0{\mathbb{V}}_{0}, we have that i0()i^{0}(\leq) is a linear order on 1{\mathbb{R}}_{1} extending \leq, i0(+)i^{0}(+) is the addition on 1{\mathbb{R}}_{1} extending ++, and i0(×)i^{0}(\times) is the multiplication on 1{\mathbb{R}}_{1} extending ×\times. For notational convenience we write \leq, ++, and ×\times for i0()i^{0}(\leq), i0(+)i^{0}(+), and i0(×)i^{0}(\times), respectively. By the elementality of i0i^{0} the structure (1;+,×,,0,1)({\mathbb{R}}_{1};+,\times,\leq,0,1) is an ordered field containing the standard real field (0;+,×,,0,1)({\mathbb{R}}_{0};+,\times,\leq,0,1) as its subfield. Notice that if Id(n)=n\mbox{Id}(n)=n for every n0n\in{\mathbb{N}}_{0}, then [Id]0i0(0)=1[\mbox{Id}]_{{\mathcal{F}}_{0}}\in i^{0}({\mathbb{N}}_{0})={\mathbb{N}}_{1} and [Id]0r[\mbox{Id}]_{{\mathcal{F}}_{0}}\geq r for every r0r\in{\mathbb{R}}_{0}, i.e., 1{\mathbb{N}}_{1} contains natural numbers such as [Id]0[\mbox{Id}]_{{\mathcal{F}}_{0}} which are infinitely large relative to real numbers in 0{\mathbb{R}}_{0}.

Let \mathscr{M} be the Mostowski collapsing map on V1V_{1}, i.e., (a)=a\mathscr{M}(a)=a for every a1a\in{\mathbb{R}}_{1} and

(b):={(a)ab}\mathscr{M}(b):=\{\mathscr{M}(a)\mid a\,^{*}\!\!\in b\}

for every bV11b\in V_{1}\setminus{\mathbb{R}}_{1}. Then \mathscr{M} is an injection and aba\,^{*}\!\!\in b iff (a)(b)\mathscr{M}(a)\in\mathscr{M}(b). If one identifies V1V_{1} with the image of V1V_{1} under \mathscr{M}, one can pretend that {}^{*}\!\!\in is the true membership relation and consider V1V_{1} as a subset of the superstructure 𝕍(1){\mathbb{V}}({\mathbb{R}}_{1}). Hence, we can pretend that {}^{*}\!\!\in is the true membership relation \in and drop for notational convenience. The purpose of the truncation of 𝕍00/0{\mathbb{V}}_{0}^{{\mathbb{N}}_{0}}/{\mathcal{F}}_{0} at {}^{*}\!\!\in-rank ω\omega is to make sure \mathscr{M} is well defined in the standard sense.

Let 𝕍1:=(V1;){\mathbb{V}}_{1}:=(V_{1};\in). We call 𝕍1{\mathbb{V}}_{1} a nonstandard universe extending 𝕍0{\mathbb{V}}_{0}. We will extend 𝕍1{\mathbb{V}}_{1} further later. Notice that due to the truncation, 𝕍1{\mathbb{V}}_{1} is no longer an elementary extension of 𝕍0{\mathbb{V}}_{0} from the model theoretic point of view. However, 𝕍1{\mathbb{V}}_{1} is a so-called bounded elementary extension of 𝕍0{\mathbb{V}}_{0}.

An \mathscr{L}–formula θ\theta has bounded quantifiers if every occurrence of quantifiers \forall and \exists in θ\theta has the form xy\forall x\!\in\!y and xy\exists x\!\in\!y. Notice that xyφ\forall x\!\in\!y\,\varphi is the abbreviation of x(xyφ)\forall x(x\!\in\!y\to\varphi) and xyφ\exists x\!\in\!y\,\varphi is the abbreviation of x(xyφ)\exists x(x\!\in\!y\wedge\varphi). Similar to (2), it is easy to show that for any \mathscr{L}–formula φ(x¯)\varphi(\overline{x}) with bounded quantifiers and any a¯𝕍0\overline{a}\in{\mathbb{V}}_{0} we have

𝕍0φ(a¯) iff 𝕍1φ(i0(a¯)).{\mathbb{V}}_{0}\models\varphi(\overline{a})\,\mbox{ iff }\,{\mathbb{V}}_{1}\models\varphi(i^{0}(\overline{a})). (3)

So, the map i0i^{0} is called a bounded elementary embedding from 𝕍0{\mathbb{V}}_{0} to 𝕍1{\mathbb{V}}_{1}. It is a common abuse of notation to write 𝕍0𝕍1{\mathbb{V}}_{0}\preceq{\mathbb{V}}_{1} to indicate the existence of the bounded elementary embedding (instead of just elementary embedding) i0i^{0} and 𝕍0𝕍1{\mathbb{V}}_{0}\prec{\mathbb{V}}_{1} to emphasize that i0i^{0} is not surjective. The property (3) is sometimes called the transfer principle in nonstandard analysis. Notice that if A𝕍0A\in{\mathbb{V}}_{0}, then i0(A)i^{0}(A) can be viewed as the ultrapower of AA modulo 0{\mathcal{F}}_{0}, i.e., i0(A)={[f]0fA0}i^{0}(A)=\{[f]_{{\mathcal{F}}_{0}}\mid f\in A^{{\mathbb{N}}_{0}}\}.

Each i0(a)i^{0}(a) for a𝕍0a\in{\mathbb{V}}_{0} is called 𝕍0{\mathbb{V}}_{0}–internal (or “standard” in some literature) and each b𝕍1b\in{\mathbb{V}}_{1} is called 𝕍1{\mathbb{V}}_{1}–internal. Hence 𝕍0{\mathbb{V}}_{0}–internal set is also a 𝕍1{\mathbb{V}}_{1}–internal set. For example, 1=i0(0){\mathbb{R}}_{1}=i^{0}({\mathbb{R}}_{0}) and 1=i0(0){\mathbb{N}}_{1}=i^{0}({\mathbb{N}}_{0}) are 𝕍0{\mathbb{V}}_{0}–internal sets. Some 𝕍1{\mathbb{V}}_{1}–internal sets are not 𝕍0{\mathbb{V}}_{0}–internal. For example, the set {1,2,,[Id]0}\{1,2,\ldots,[\mbox{Id}]_{{\mathcal{F}}_{0}}\} is 𝕍1{\mathbb{V}}_{1}–internal subset of 1{\mathbb{N}}_{1} but not 𝕍0{\mathbb{V}}_{0}–internal. Some subsets of a 𝕍1{\mathbb{V}}_{1}–internal set are not 𝕍1{\mathbb{V}}_{1}–internal. For example, 0{\mathbb{N}}_{0} as a subset of 1{\mathbb{N}}_{1} is not 𝕍1{\mathbb{V}}_{1}–internal because it is bounded above in 1{\mathbb{N}}_{1} and has no largest element. Notice that i0[0]=0i^{0}[{\mathbb{N}}_{0}]={\mathbb{N}}_{0} and i0(0)=1i^{0}({\mathbb{N}}_{0})={\mathbb{N}}_{1}. The following proposition says that a subset of B𝕍1B\in{\mathbb{V}}_{1} defined by an \mathscr{L}–formula with parameters from 𝕍1{\mathbb{V}}_{1} is 𝕍1{\mathbb{V}}_{1}–internal. The proposition is an easy consequence of Proposition 2.3.

Proposition 2.5

Let φ(a¯,x¯)\varphi(\overline{a},\overline{x}) be an \mathscr{L}–formula with bounded quantifiers, and parameters a¯\overline{a} and BmB^{m} being 𝕍1{\mathbb{V}}_{1}–internal. Then {b¯Bm𝕍1φ(a¯,b¯)}\left\{\overline{b}\in B^{m}\mid{\mathbb{V}}_{1}\models\varphi(\overline{a},\overline{b})\right\} is a 𝕍1{\mathbb{V}}_{1}–internal set. (cf. [1, Theorem 4.4.14].)

The following is called the overspill principle in nonstandard analysis. Notice that 0{\mathbb{N}}_{0} is an infinite initial segment of 1{\mathbb{N}}_{1}.

Proposition 2.6

Let U1U\subseteq{\mathbb{N}}_{1} be an infinite proper initial segment of 1{\mathbb{N}}_{1} and not 𝕍1{\mathbb{V}}_{1}–internal. Let AA be an 𝕍1{\mathbb{V}}_{1}–internal subset of 1{\mathbb{N}}_{1}. If AUA\cap U is upper unbounded in UU, then A(1U)A\cap({\mathbb{N}}_{1}\setminus U)\not=\emptyset.

The proof of Proposition 2.6 is easy. If A(1U)=A\cap({\mathbb{N}}_{1}\setminus U)=\emptyset, then UU can be defined by a formula with bounded quantifiers and parameter AA. Hence UU is 𝕍1{\mathbb{V}}_{1}–internal.


§2.5 Construction of 𝕍2{\mathbb{V}}_{2} and 𝕍3{\mathbb{V}}_{3} We now extend 𝕍1{\mathbb{V}}_{1} further to 𝕍2{\mathbb{V}}_{2} and 𝕍3{\mathbb{V}}_{3} to form a nonstandard extension chain

𝕍0𝕍1𝕍2𝕍3{\mathbb{V}}_{0}\prec{\mathbb{V}}_{1}\prec{\mathbb{V}}_{2}\prec{\mathbb{V}}_{3}

using ultrapower construction and show the existence of bounded elementary embeddings i0i_{0}, ii_{*}, i1i_{1}, and i2i_{2} besides the natural embeddings i1:𝕍1𝕍2i^{1}:{\mathbb{V}}_{1}\to{\mathbb{V}}_{2} and i2:𝕍2𝕍3i^{2}:{\mathbb{V}}_{2}\to{\mathbb{V}}_{3}. The chain and embeddings will satisfy the following properties. Let j+1=ij(j){\mathbb{N}}_{j+1}=i^{j}({\mathbb{N}}_{j}) and j+1=ij(j){\mathbb{R}}_{j+1}=i^{j}({\mathbb{R}}_{j}) be the set of all positive integers and the set of all real numbers, respectively, in 𝕍j+1{\mathbb{V}}_{j+1} for j=1,2j=1,2.

Property 2.7
  1. 1.

    For j=0,1,2j=0,1,2, j+1{\mathbb{N}}_{j+1} is an end–extension of j{\mathbb{N}}_{j}, i.e., every number in j+1j{\mathbb{N}}_{j+1}\setminus{\mathbb{N}}_{j} is greater than each number in j{\mathbb{N}}_{j}.

  2. 2.

    There is a bounded elementary embedding ii_{*} from the \mathscr{L}^{\prime}–model
    (𝕍2;0,1)({\mathbb{V}}_{2};{\mathbb{R}}_{0},{\mathbb{R}}_{1}) to the \mathscr{L}^{\prime}–model (𝕍3;1,2)({\mathbb{V}}_{3};{\mathbb{R}}_{1},{\mathbb{R}}_{2}), where :={P1,P2}\mathscr{L}^{\prime}:=\mathscr{L}\cup\{P_{1},P_{2}\} for two new unary predicate symbols P1P_{1} and P2P_{2} not in \mathscr{L}. Furthermore, the map i1:=i𝕍1i_{1}:=i_{*}\!\upharpoonright\!{\mathbb{V}}_{1} is a bounded elementary embedding from (𝕍1;0)({\mathbb{V}}_{1};{\mathbb{R}}_{0}) to (𝕍2;1)({\mathbb{V}}_{2};{\mathbb{R}}_{1}). Notice that i1(a)21i_{1}(a)\in{\mathbb{N}}_{2}\setminus{\mathbb{N}}_{1} for each a10a\in{\mathbb{N}}_{1}\setminus{\mathbb{N}}_{0}.

  3. 3.

    There is a bounded elementary embedding i2i_{2} from 𝕍2{\mathbb{V}}_{2} to 𝕍3{\mathbb{V}}_{3} such that i21i_{2}\!\upharpoonright\!{\mathbb{N}}_{1} is an identity map and i2(a)32i_{2}(a)\in{\mathbb{N}}_{3}\setminus{\mathbb{N}}_{2} for each a21a\in{\mathbb{N}}_{2}\setminus{\mathbb{N}}_{1}.

We are now going to work towards establishing this property in this section.

\blacklozenge: From now on, let 𝕍jj{\mathbb{V}}_{j}^{{\mathbb{N}}_{j}} always represent, for notational convenience, the set of all functions ff from j{\mathbb{N}}_{j} to 𝕍j{\mathbb{V}}_{j} such that {rank(f(n))nj}\{\mbox{rank}(f(n))\mid n\in{\mathbb{N}}_{j}\} is a bounded set in ω\omega.

Notice that 1:=i0(0)𝕍1{\mathcal{F}}_{1}:=i^{0}({\mathcal{F}}_{0})\in{\mathbb{V}}_{1} satisfies Part 1–4 of Definition 2.1 with XX, 𝒫<ω(X)\mathscr{P}_{<\omega}(X), and 𝒫(X)\mathscr{P}(X) being replaced by 1:=i0(0){\mathbb{N}}_{1}:=i^{0}({\mathbb{N}}_{0}), i0(𝒫<0(0))={A1A𝕍1 and |A|1}i^{0}(\mathscr{P}_{<{\mathbb{N}}_{0}}({\mathbb{N}}_{0}))=\left\{A\subseteq{\mathbb{N}}_{1}\mid A\in{\mathbb{V}}_{1}\,\mbox{ and }\,|A|\in{\mathbb{N}}_{1}\right\}, and i0(𝒫(0))=𝕍1𝒫(1)i^{0}(\mathscr{P}({\mathbb{N}}_{0}))={\mathbb{V}}_{1}\cap\mathscr{P}({\mathbb{N}}_{1}), respectively. We call 1{\mathcal{F}}_{1} a 𝕍1{\mathbb{V}}_{1}–internal non-principal ultrafilter on 1{\mathbb{N}}_{1}.

Definition 2.8

Let 1:=i0(0){\mathcal{F}}_{1}:=i^{0}({\mathcal{F}}_{0}). Denote 𝕍2{\mathbb{V}}_{2} for the model (V2;2)(V_{2};\,\in_{2}) such that

V2:=(𝕍11𝕍1)/1 andV_{2}:=({\mathbb{V}}_{1}^{{\mathbb{N}}_{1}}\cap{\mathbb{V}}_{1})/{\mathcal{F}}_{1}\,\mbox{ and}
[f]12[g]1 iff {n1f(n)g(n)}1[f]_{{\mathcal{F}}_{1}}\in_{2}[g]_{{\mathcal{F}}_{1}}\,\mbox{ iff }\,\{n\in{\mathbb{N}}_{1}\mid f(n)\in g(n)\}\in{\mathcal{F}}_{1}

for all f,g𝕍11𝕍1f,g\in{\mathbb{V}}_{1}^{{\mathbb{N}}_{1}}\cap{\mathbb{V}}_{1}. Let i1:𝕍1𝕍2i^{1}:{\mathbb{V}}_{1}\to{\mathbb{V}}_{2} be the natural embedding that i1(c)=[ϕc]1i^{1}(c)=[\phi_{c}]_{{\mathcal{F}}_{1}} for every c𝕍1c\in{\mathbb{V}}_{1}.

We call 𝕍2{\mathbb{V}}_{2} a 𝕍1{\mathbb{V}}_{1}–internal ultrapower of 𝕍1{\mathbb{V}}_{1} modulo the 𝕍1{\mathbb{V}}_{1}–internal ultrafilter 1{\mathcal{F}}_{1}.

Proposition 2.9

If φ(x¯)\varphi(\overline{x}) is an \mathscr{L}–formula with bounded quantifier and
[a]¯1𝕍2\overline{[a]}_{{\mathcal{F}}_{1}}\in{\mathbb{V}}_{2}, then

𝕍2φ([a]¯1) iff {n1𝕍1φ(a(n)¯)}1.{\mathbb{V}}_{2}\models\varphi(\overline{[a]}_{{\mathcal{F}}_{1}})\,\mbox{ iff }\,\left\{n\in{\mathbb{N}}_{1}\mid{\mathbb{V}}_{1}\models\varphi(\overline{a(n)})\right\}\in{\mathcal{F}}_{1}.
Corollary 2.10

The natural embedding i1:𝕍1𝕍2i^{1}:{\mathbb{V}}_{1}\to{\mathbb{V}}_{2} is a bounded elementary embedding from 𝕍1{\mathbb{V}}_{1} to 𝕍2{\mathbb{V}}_{2}.

The proof of Proposition 2.9 is almost the same as the proof of Proposition 2.3 except one step that shows {n1𝕍1xB(n)φ(a(n)¯,x)}1\left\{n\in{\mathbb{N}}_{1}\mid{\mathbb{V}}_{1}\models\exists x\!\in\!B(n)\,\varphi(\overline{a(n)},x)\right\}\in{\mathcal{F}}_{1} implies 𝕍2x[B]1φ([a]¯1,x){\mathbb{V}}_{2}\models\exists x\!\in\![B]_{{\mathcal{F}}_{1}}\,\varphi(\overline{[a]}_{{\mathcal{F}}_{1}},x). Let Mz=i0(𝕍(0,z))𝕍1M_{z}=i^{0}({\mathbb{V}}({\mathbb{R}}_{0},z))\in{\mathbb{V}}_{1} with zωz\in\omega such that B,a¯MzB,\overline{a}\in M_{z}. By the axiom of choice there is a well-order \lhd on 𝕍(0,z){\mathbb{V}}({\mathbb{R}}_{0},z). So, every nonempty set AMzA\in M_{z} has a i0()i^{0}(\lhd)-least element by the transfer principle. Suppose that {n1𝕍1xB(n)φ(a(n)¯,x)}=X1\left\{n\in{\mathbb{N}}_{1}\mid{\mathbb{V}}_{1}\models\exists x\!\in\!B(n)\,\varphi(\overline{a(n)},x)\right\}=X\in{\mathcal{F}}_{1}. For each n1n\in{\mathbb{N}}_{1}, if nXn\not\in X, let f(n)=0f(n)=0 and if nXn\in X, let f(n)f(n) be the i0()i^{0}(\lhd)-least element in the nonempty set {xB(n)𝕍1φ(a(n),x)}\left\{x\in B(n)\mid{\mathbb{V}}_{1}\models\varphi(a(n),x)\right\}. Since B,a¯,i0()B,\overline{a},i^{0}(\lhd) are in 𝕍1{\mathbb{V}}_{1}, so does the function ff by Proposition 2.5. Hence [f]1𝕍2[B]1[f]_{{\mathcal{F}}_{1}}\in{\mathbb{V}}_{2}\cap[B]_{{\mathcal{F}}_{1}}. By the induction hypothesis we have that

{n1𝕍1φ(a(n),f(n))}X1 implies 𝕍2φ([a]¯1,[f]1),\left\{n\in{\mathbb{N}}_{1}\mid{\mathbb{V}}_{1}\models\varphi(a(n),f(n))\right\}\supseteq X\in{\mathcal{F}}_{1}\,\mbox{ implies }\,{\mathbb{V}}_{2}\models\varphi(\overline{[a]}_{{\mathcal{F}}_{1}},[f]_{{\mathcal{F}}_{1}}),

which implies 𝕍2x[B]1φ([a]¯1,x){\mathbb{V}}_{2}\models\exists x\!\in\![B]_{{\mathcal{F}}_{1}}\,\varphi(\overline{[a]}_{{\mathcal{F}}_{1}},x).

Let

i0=i1i0.i_{0}=i^{1}\!\circ\!i^{0}. (4)

Then i0i_{0} is a bounded elementary embedding from 𝕍0{\mathbb{V}}_{0} to 𝕍2{\mathbb{V}}_{2}, which will be used in Theorem 5.4.

Same as for 𝕍1{\mathbb{V}}_{1} we can assume by Mostowski collapsing that 𝕍2{\mathbb{V}}_{2} is a subset of the superstructure 𝕍(2){\mathbb{V}}({\mathbb{R}}_{2}) and 2\in_{2} is the true membership relation \in. Notice that i11i^{1}\!\upharpoonright\!{\mathbb{R}}_{1} is an identity map. The element a𝕍2a\in{\mathbb{V}}_{2} is called 𝕍2{\mathbb{V}}_{2}–internal, i1(b)i^{1}(b) is called 𝕍1{\mathbb{V}}_{1}–internal for any b𝕍1b\in{\mathbb{V}}_{1}, and i0(c)i_{0}(c) is called 𝕍0{\mathbb{V}}_{0}–internal for every c𝕍0c\in{\mathbb{V}}_{0}. Notice that 0{\mathbb{N}}_{0} and 1{\mathbb{N}}_{1} as subsets of 2{\mathbb{N}}_{2} are not 𝕍2{\mathbb{V}}_{2}–internal.

If f𝕍1f\in{\mathbb{V}}_{1} is a function from 1{\mathbb{N}}_{1} to [n][n] for some n1n\in{\mathbb{N}}_{1}, then [f]1=m[f]_{{\mathcal{F}}_{1}}=m for some m1m\in{\mathbb{N}}_{1} by (3) with 𝕍0,𝕍1,i0{\mathbb{V}}_{0},{\mathbb{V}}_{1},i^{0} being replaced by 𝕍1,𝕍2,i1{\mathbb{V}}_{1},{\mathbb{V}}_{2},i^{1}. Hence 2{\mathbb{N}}_{2} is a proper end-extension of 1{\mathbb{N}}_{1}.

By the same way, we can define 𝕍3{\mathbb{V}}_{3} as a 𝕍2{\mathbb{V}}_{2}–internal ultrapower of 𝕍2{\mathbb{V}}_{2} modulo i1(1)i^{1}({\mathcal{F}}_{1}).

Definition 2.11

Let 2:=i1(1){\mathcal{F}}_{2}:=i^{1}({\mathcal{F}}_{1}) be the 𝕍2{\mathbb{V}}_{2}–internal ultrafilter on 2{\mathbb{N}}_{2}. Let 𝕍3{\mathbb{V}}_{3} be the model (V3;3)(V_{3};\,\in_{3}) such that

V3:=(𝕍22𝕍2)/2 andV_{3}:=({\mathbb{V}}_{2}^{{\mathbb{N}}_{2}}\cap{\mathbb{V}}_{2})/{\mathcal{F}}_{2}\,\mbox{ and}
[f]23[g]2 iff {n2f(n)g(n)}2[f]_{{\mathcal{F}}_{2}}\in_{3}[g]_{{\mathcal{F}}_{2}}\,\mbox{ iff }\,\{n\in{\mathbb{N}}_{2}\mid f(n)\in g(n)\}\in{\mathcal{F}}_{2}

for all f,g𝕍22𝕍2f,g\in{\mathbb{V}}_{2}^{{\mathbb{N}}_{2}}\cap{\mathbb{V}}_{2}, and define the natural mebedding i2:𝕍2𝕍3i^{2}:{\mathbb{V}}_{2}\to{\mathbb{V}}_{3} by i2(c)=[ϕc]2i^{2}(c)=[\phi_{c}]_{{\mathcal{F}}_{2}} for every c𝕍2c\in{\mathbb{V}}_{2}.

Generalizing the arguments above, we have that the map i2i^{2} is a bounded elementary embedding from 𝕍2{\mathbb{V}}_{2} to 𝕍3{\mathbb{V}}_{3}. We say that 𝕍3{\mathbb{V}}_{3} is a nonstandard extension of 𝕍2{\mathbb{V}}_{2}. It is also easy to see that 3:=i2(2){\mathbb{N}}_{3}:=i^{2}({\mathbb{N}}_{2}) is an end-extension of 2{\mathbb{N}}_{2}.

We have completed the construction of 𝕍0𝕍1𝕍2𝕍3{\mathbb{V}}_{0}\prec{\mathbb{V}}_{1}\prec{\mathbb{V}}_{2}\prec{\mathbb{V}}_{3} and verified that Part 1 of Property 2.7 is true.


§2.6 Bounded elementary embeddings ii_{*}, i1i_{1}, and i2i_{2} To verify Part 2–3 of Property 2.7 we have to view the construction of 𝕍j{\mathbb{V}}_{j} for j=2,3j=2,3 from a different angle. For a set A0×0A\subseteq{\mathbb{N}}_{0}\times{\mathbb{N}}_{0} and n0n\in{\mathbb{N}}_{0}, let An:={m0(m,n)A}A_{n}:=\left\{m\in{\mathbb{N}}_{0}\mid(m,n)\in A\right\}. Let 0{\mathcal{F}}_{0} and 0{\mathcal{F}}_{0}^{\prime} be two non-principal ultrafilters on 0{\mathbb{N}}_{0}. The tensor product of 0{\mathcal{F}}_{0} and 0{\mathcal{F}}^{\prime}_{0} is defined by

00:={A0×0{m0Am0}0}.{\mathcal{F}}_{0}\otimes{\mathcal{F}}_{0}^{\prime}:=\left\{A\subseteq{\mathbb{N}}_{0}\times{\mathbb{N}}_{0}\mid\{m\in{\mathbb{N}}_{0}\mid A_{m}\in{\mathcal{F}}_{0}\}\in{\mathcal{F}}_{0}^{\prime}\right\}.

It is easy to check that 00{\mathcal{F}}_{0}\otimes{\mathcal{F}}_{0}^{\prime} is a non-principal ultrafilter on 0×0{\mathbb{N}}_{0}\times{\mathbb{N}}_{0}. For simplicity we assume that 0{\mathcal{F}}_{0} and 0{\mathcal{F}}^{\prime}_{0} are the same ultrafilter.

Lemma 2.12

Let 𝕍2:=(V00×0/00;){\mathbb{V}}^{\prime}_{2}:=(V_{0}^{{\mathbb{N}}_{0}\times{\mathbb{N}}_{0}}/{\mathcal{F}}_{0}\otimes{\mathcal{F}}_{0}^{\prime};\in^{\prime}) where [f]00[g]00[f]_{{\mathcal{F}}_{0}\otimes{\mathcal{F}}^{\prime}_{0}}\in^{\prime}[g]_{{\mathcal{F}}_{0}\otimes{\mathcal{F}}^{\prime}_{0}} iff
{(m,n)0×0f(m,n)g(m,n)}00\left\{(m,n)\in{\mathbb{N}}_{0}\times{\mathbb{N}}_{0}\mid f(m,n)\in g(m,n)\right\}\in{\mathcal{F}}_{0}\otimes{\mathcal{F}}_{0}^{\prime}. Then

𝕍2𝕍2.{\mathbb{V}}_{2}\cong{\mathbb{V}}_{2}^{\prime}.

Proof Let mm range over the first copy of 0{\mathbb{N}}_{0} and nn over the second copy of 0{\mathbb{N}}_{0} in 0×0{\mathbb{N}}_{0}\times{\mathbb{N}}_{0}. Given a𝕍2a\in{\mathbb{V}}_{2}, there is an fa𝕍11𝕍1f_{a}\in{\mathbb{V}}_{1}^{{\mathbb{N}}_{1}}\cap{\mathbb{V}}_{1} such that a=[fa]1a=[f_{a}]_{{\mathcal{F}}_{1}}. Since the ranks of all image of faf_{a} is bounded, the range of faf_{a} is in a 𝕍0{\mathbb{V}}_{0}–internal set i0(B)i^{0}(B). So, by identifying faf_{a} with its graph, we have fa1×i0(B)f_{a}\subseteq{\mathbb{N}}_{1}\times i^{0}(B). Since fa𝕍1f_{a}\in{\mathbb{V}}_{1}, there is a ga𝕍00g_{a}\in{\mathbb{V}}_{0}^{{\mathbb{N}}_{0}} such that fa=[ga]0f_{a}=[g_{a}]_{{\mathcal{F}}_{0}} where ga(n)0×Bg_{a}(n)\subseteq{\mathbb{N}}_{0}\times B is the graph of a function ga(n):0Bg_{a}(n):{\mathbb{N}}_{0}\to B for all n0n\in{\mathbb{N}}_{0}. Now let Fa:0×0BF_{a}:{\mathbb{N}}_{0}\times{\mathbb{N}}_{0}\to B be such that Fa(m,n)=ga(n)(m)BF_{a}(m,n)=g_{a}(n)(m)\in B. Then Fa𝕍00×0F_{a}\in{\mathbb{V}}_{0}^{{\mathbb{N}}_{0}\times{\mathbb{N}}_{0}}.

We view a[Fa]00a\mapsto[F_{a}]_{{\mathcal{F}}_{0}\otimes{\mathcal{F}}^{\prime}_{0}} as a relation between 𝕍2{\mathbb{V}}_{2} and 𝕍2{\mathbb{V}}^{\prime}_{2}. Notice that [X]01[X]_{{\mathcal{F}}^{\prime}_{0}}\in{\mathcal{F}}_{1} iff {n0X(n)0}0\{n\in{\mathbb{N}}_{0}\mid X(n)\in{\mathcal{F}}_{0}\}\in{\mathcal{F}}^{\prime}_{0}. Notice also that for any a,b𝕍2a,b\in{\mathbb{V}}_{2}, we have

a=b iff\displaystyle a=b\,\mbox{ iff }
{[x]01fa([x]0)=fb([x]0)}1 iff\displaystyle\left\{[x]_{{\mathcal{F}}^{\prime}_{0}}\in{\mathbb{N}}_{1}\mid f_{a}([x]_{{\mathcal{F}}^{\prime}_{0}})=f_{b}([x]_{{\mathcal{F}}^{\prime}_{0}})\right\}\in{\mathcal{F}}_{1}\,\mbox{ iff}
{[x]01[ga]0([x]0)=[gb]0([x]0)}1 iff\displaystyle\left\{[x]_{{\mathcal{F}}^{\prime}_{0}}\in{\mathbb{N}}_{1}\mid[g_{a}]_{{\mathcal{F}}^{\prime}_{0}}([x]_{{\mathcal{F}}^{\prime}_{0}})=[g_{b}]_{{\mathcal{F}}^{\prime}_{0}}([x]_{{\mathcal{F}}^{\prime}_{0}})\right\}\in{\mathcal{F}}_{1}\,\mbox{ iff}
{[x]01[ga(n)(x(n))]0=[gb(n)(x(n))]0}1 iff\displaystyle\left\{[x]_{{\mathcal{F}}^{\prime}_{0}}\in{\mathbb{N}}_{1}\mid[g_{a}(n)(x(n))]_{{\mathcal{F}}^{\prime}_{0}}=[g_{b}(n)(x(n))]_{{\mathcal{F}}^{\prime}_{0}}\right\}\in{\mathcal{F}}_{1}\,\mbox{ iff}
{n0{m=x(n)0ga(n)(x(n))=gb(n)(x(n))}0}0 iff\displaystyle\left\{n\in{\mathbb{N}}_{0}\mid\{m=x(n)\in{\mathbb{N}}_{0}\mid g_{a}(n)(x(n))=g_{b}(n)(x(n))\}\in{\mathcal{F}}_{0}\right\}\in{\mathcal{F}}^{\prime}_{0}\,\mbox{ iff}
{n0{m0ga(n)(m)=gb(n)(m)}0}0 iff\displaystyle\left\{n\in{\mathbb{N}}_{0}\mid\{m\in{\mathbb{N}}_{0}\mid g_{a}(n)(m)=g_{b}(n)(m)\}\in{\mathcal{F}}_{0}\right\}\in{\mathcal{F}}^{\prime}_{0}\,\mbox{ iff}
{n0{m0Fa(m,n)=Fb(m,n))}0}0 iff\displaystyle\left\{n\in{\mathbb{N}}_{0}\mid\{m\in{\mathbb{N}}_{0}\mid F_{a}(m,n)=F_{b}(m,n))\}\in{\mathcal{F}}_{0}\right\}\in{\mathcal{F}}^{\prime}_{0}\,\mbox{ iff}
{(m,n)0×0Fa(m,n)=Fb(m,n))}00 iff\displaystyle\left\{(m,n)\in{\mathbb{N}}_{0}\times{\mathbb{N}}_{0}\mid F_{a}(m,n)=F_{b}(m,n))\right\}\in{\mathcal{F}}_{0}\otimes{\mathcal{F}}^{\prime}_{0}\,\mbox{ iff}
[Fa]00=[Fb]00.\displaystyle[F_{a}]_{{\mathcal{F}}_{0}\otimes{\mathcal{F}}^{\prime}_{0}}=[F_{b}]_{{\mathcal{F}}_{0}\otimes{\mathcal{F}}^{\prime}_{0}}.

Hence, the relation a[Fa]00a\mapsto[F_{a}]_{{\mathcal{F}}_{0}\otimes{\mathcal{F}}^{\prime}_{0}} is an injective function from 𝕍2{\mathbb{V}}_{2} to 𝕍2{\mathbb{V}}^{\prime}_{2}.

On the other hand, given [F]00𝕍00×0/00[F]_{{\mathcal{F}}_{0}\otimes{\mathcal{F}}^{\prime}_{0}}\in{\mathbb{V}}_{0}^{{\mathbb{N}}_{0}\times{\mathbb{N}}_{0}}/{\mathcal{F}}_{0}\otimes{\mathcal{F}}^{\prime}_{0}, let g(n)𝕍00g(n)\in{\mathbb{V}}_{0}^{{\mathbb{N}}_{0}} be such that g(n)(m)=F(m,n)g(n)(m)=F(m,n). Let f:1𝕍1f:{\mathbb{N}}_{1}\to{\mathbb{V}}_{1} be such that f([x]0):=[g]0([x]0)=[g(n)(x(n))]0f([x]_{{\mathcal{F}}_{0}}):=[g]_{{\mathcal{F}}^{\prime}_{0}}([x]_{{\mathcal{F}}_{0}})=[g(n)(x(n))]_{{\mathcal{F}}^{\prime}_{0}}. Then f=[g]0𝕍1f=[g]_{{\mathcal{F}}^{\prime}_{0}}\in{\mathbb{V}}_{1}. So, a=[f]1𝕍2a=[f]_{{\mathcal{F}}_{1}}\in{\mathbb{V}}_{2} is such that [Fa]00=[F]00[F_{a}]_{{\mathcal{F}}_{0}\otimes{\mathcal{F}}^{\prime}_{0}}=[F]_{{\mathcal{F}}_{0}\otimes{\mathcal{F}}^{\prime}_{0}}. This shows that a[Fa]00a\mapsto[F_{a}]_{{\mathcal{F}}_{0}\otimes{\mathcal{F}}^{\prime}_{0}} is surjective.

Notice also that for any a,b𝕍2a,b\in{\mathbb{V}}_{2},

ab iff\displaystyle a\in b\,\mbox{ iff}
{[x]01fa([x]0)fb([x]0)}1 iff\displaystyle\left\{[x]_{{\mathcal{F}}^{\prime}_{0}}\in{\mathbb{N}}_{1}\mid f_{a}([x]_{{\mathcal{F}}^{\prime}_{0}})\in f_{b}([x]_{{\mathcal{F}}^{\prime}_{0}})\right\}\in{\mathcal{F}}_{1}\,\mbox{ iff}
{[x]01[ga]0([x]0)[gb]0([x]0)}1 iff\displaystyle\left\{[x]_{{\mathcal{F}}^{\prime}_{0}}\in{\mathbb{N}}_{1}\mid[g_{a}]_{{\mathcal{F}}^{\prime}_{0}}([x]_{{\mathcal{F}}^{\prime}_{0}})\in[g_{b}]_{{\mathcal{F}}^{\prime}_{0}}([x]_{{\mathcal{F}}^{\prime}_{0}})\right\}\in{\mathcal{F}}_{1}\,\mbox{ iff}
{[x]01[ga(n)(x(n))]0[gb(n)(x(n))]0}1 iff\displaystyle\left\{[x]_{{\mathcal{F}}^{\prime}_{0}}\in{\mathbb{N}}_{1}\mid[g_{a}(n)(x(n))]_{{\mathcal{F}}^{\prime}_{0}}\in[g_{b}(n)(x(n))]_{{\mathcal{F}}^{\prime}_{0}}\right\}\in{\mathcal{F}}_{1}\,\mbox{ iff}
{n0{m=x(n)0ga(n)(x(n))gb(n)(x(n))}0}0 iff\displaystyle\left\{n\in{\mathbb{N}}_{0}\mid\{m=x(n)\in{\mathbb{N}}_{0}\mid g_{a}(n)(x(n))\in g_{b}(n)(x(n))\}\in{\mathcal{F}}_{0}\right\}\in{\mathcal{F}}^{\prime}_{0}\,\mbox{ iff}
{n0{m0ga(n)(m)gb(n)(m)}0}0 iff\displaystyle\left\{n\in{\mathbb{N}}_{0}\mid\{m\in{\mathbb{N}}_{0}\mid g_{a}(n)(m)\in g_{b}(n)(m)\}\in{\mathcal{F}}_{0}\right\}\in{\mathcal{F}}^{\prime}_{0}\,\mbox{ iff}
{n0{m0Fa(m,n)Fb(m,n))}0}0 iff\displaystyle\left\{n\in{\mathbb{N}}_{0}\mid\{m\in{\mathbb{N}}_{0}\mid F_{a}(m,n)\in F_{b}(m,n))\}\in{\mathcal{F}}_{0}\right\}\in{\mathcal{F}}^{\prime}_{0}\,\mbox{ iff}
{(m,n)0×0Fa(m,n)Fb(m,n))}00 iff\displaystyle\left\{(m,n)\in{\mathbb{N}}_{0}\times{\mathbb{N}}_{0}\mid F_{a}(m,n)\in F_{b}(m,n))\right\}\in{\mathcal{F}}_{0}\otimes{\mathcal{F}}^{\prime}_{0}\,\mbox{ iff}
[Fa]00[Fb]00.\displaystyle[F_{a}]_{{\mathcal{F}}_{0}\otimes{\mathcal{F}}^{\prime}_{0}}\in^{\prime}[F_{b}]_{{\mathcal{F}}_{0}\otimes{\mathcal{F}}^{\prime}_{0}}.

This completes the proof of 𝕍2𝕍2{\mathbb{V}}_{2}\cong{\mathbb{V}}^{\prime}_{2}. \blacksquare

If we identify each of 𝕍2{\mathbb{V}}_{2} and 𝕍2{\mathbb{V}}^{\prime}_{2} with its image of Mostowski collapsing, then 𝕍2{\mathbb{V}}_{2} and 𝕍2{\mathbb{V}}^{\prime}_{2} can be viewed as the same model.

Lemma 2.13

Let 𝕍2′′:=((V00/0)0/0;′′){\mathbb{V}}^{\prime\prime}_{2}:=\left((V_{0}^{{\mathbb{N}}_{0}}/{\mathcal{F}}_{0})^{{\mathbb{N}}_{0}}/{\mathcal{F}}_{0}^{\prime};\in^{\prime\prime}\right) where [[f]0]0′′[[g]0]0[[f]_{{\mathcal{F}}_{0}}]_{{\mathcal{F}}^{\prime}_{0}}\in^{\prime\prime}[[g]_{{\mathcal{F}}_{0}}]_{{\mathcal{F}}^{\prime}_{0}} iff
{n0{m0f(m,n)g(m,n)}0}0\left\{n\in{\mathbb{N}}_{0}\mid\left\{m\in{\mathbb{N}}_{0}\mid f(m,n)\in g(m,n)\right\}\in{\mathcal{F}}_{0}\right\}\in{\mathcal{F}}_{0}^{\prime}. Then

𝕍2𝕍2′′{\mathbb{V}}^{\prime}_{2}\cong{\mathbb{V}}_{2}^{\prime\prime}

for any f,g:0×0𝕍0f,g:{\mathbb{N}}_{0}\times{\mathbb{N}}_{0}\to{\mathbb{V}}_{0}.

The proof of Lemma 2.13 can be found in [1, Proposition 6.5.2]. We call 𝕍2′′{\mathbb{V}}_{2}^{\prime\prime} the external ultrapower of 𝕍1{\mathbb{V}}_{1} modulo 0{\mathcal{F}}^{\prime}_{0}. By Lemma 2.12 and Lemma 2.13 we can view 𝕍2{\mathbb{V}}_{2} and 𝕍2′′{\mathbb{V}}_{2}^{\prime\prime} as the same model and write \in for ′′\in^{\prime\prime}. To summarize, we have that

𝕍2′′=(𝕍00/0)0/0=𝕍00×0/00=(𝕍11𝕍1)/1=𝕍2.{\mathbb{V}}^{\prime\prime}_{2}=({\mathbb{V}}_{0}^{{\mathbb{N}}_{0}}/{\mathcal{F}}_{0})^{{\mathbb{N}}_{0}}/{\mathcal{F}}_{0}^{\prime}={\mathbb{V}}_{0}^{{\mathbb{N}}_{0}\times{\mathbb{N}}_{0}}/{\mathcal{F}}_{0}\otimes{\mathcal{F}}_{0}^{\prime}=({\mathbb{V}}_{1}^{{\mathbb{N}}_{1}}\cap{\mathbb{V}}_{1})/{\mathcal{F}}_{1}={\mathbb{V}}_{2}. (5)

The term on the left side is the two-step iteration when we take the ultrapowers by using 0{\mathcal{F}}_{0} first and 0{\mathcal{F}}^{\prime}_{0} second. The term on the right side is when we use 0{\mathcal{F}}^{\prime}_{0} first so that V00V_{0}^{{\mathbb{N}}_{0}} becomes 𝕍1{\mathbb{V}}_{1}, 0{\mathbb{N}}_{0} becomes 1{\mathbb{N}}_{1}, and 0{\mathcal{F}}_{0} becomes 1{\mathcal{F}}_{1}.

However, the bounded elementary embedding from 𝕍1{\mathbb{V}}_{1} to 𝕍2{\mathbb{V}}_{2} induced by the 𝕍1{\mathbb{V}}_{1}–internal ultrapower modulo 1{\mathcal{F}}_{1} and the bounded elementary embedding from 𝕍1{\mathbb{V}}_{1} to 𝕍2′′{\mathbb{V}}_{2}^{\prime\prime} induced by the external ultrapower of 𝕍1{\mathbb{V}}_{1} modulo 0{\mathcal{F}}_{0}^{\prime} are different. Let i1:𝕍1𝕍2′′=𝕍10/0i_{1}:{\mathbb{V}}_{1}\to{\mathbb{V}}_{2}^{\prime\prime}={\mathbb{V}}_{1}^{{\mathbb{N}}_{0}}/{\mathcal{F}}^{\prime}_{0} be such that i1(c)=[ϕc]0i_{1}(c)=[\phi_{c}]_{{\mathcal{F}}^{\prime}_{0}} for each c𝕍1c\in{\mathbb{V}}_{1}. If c01c\in{\mathbb{N}}_{0}\subseteq{\mathbb{N}}_{1}, then clearly we have [ϕc]0=c[\phi_{c}]_{{\mathcal{F}}_{0}^{\prime}}=c. If c,c10c,c^{\prime}\in{\mathbb{N}}_{1}\setminus{\mathbb{N}}_{0} with c=[f]0c=[f]_{{\mathcal{F}}^{\prime}_{0}} and c=[g]0c^{\prime}=[g]_{{\mathcal{F}}^{\prime}_{0}} for some f,g00f,g\in{\mathbb{N}}_{0}^{{\mathbb{N}}_{0}}, then c>g(n)c>g(n) for every n0n\in{\mathbb{N}}_{0}. Hence i1(c)=[ϕc]0>[g]0=ci_{1}(c)=[\phi_{c}]_{{\mathcal{F}}_{0}^{\prime}}>[g]_{{\mathcal{F}}_{0}^{\prime}}=c^{\prime}. This shows that i1(c)21i_{1}(c)\in{\mathbb{N}}_{2}\setminus{\mathbb{N}}_{1}. The map i1i_{1} will be i𝕍1i_{*}\!\upharpoonright\!{\mathbb{V}}_{1} for a bounded elementary embedding ii_{*} defined below. Notice that above arguments still work if the ultrafilters 0{\mathcal{F}}_{0} and 0{\mathcal{F}}_{0}^{\prime} are different.

Generalizing the construction further and with the help of Mostowski collapsing, we have that

𝕍3′′:=((𝕍00/0)0/0)0/0′′=𝕍20/0′′\displaystyle{\mathbb{V}}^{\prime\prime}_{3}:=(({\mathbb{V}}_{0}^{{\mathbb{N}}_{0}}/{\mathcal{F}}_{0})^{{\mathbb{N}}_{0}}/{\mathcal{F}}_{0}^{\prime})^{{\mathbb{N}}_{0}}/{\mathcal{F}}_{0}^{\prime\prime}={\mathbb{V}}_{2}^{{\mathbb{N}}_{0}}/{\mathcal{F}}^{\prime\prime}_{0}
=𝕍00×0×0/000′′\displaystyle={\mathbb{V}}_{0}^{{\mathbb{N}}_{0}\times{\mathbb{N}}_{0}\times{\mathbb{N}}_{0}}/{\mathcal{F}}_{0}\otimes{\mathcal{F}}_{0}^{\prime}\otimes{\mathcal{F}}_{0}^{\prime\prime}
=(𝕍11×1𝕍1)/11=(𝕍21𝕍1)/1=𝕍3\displaystyle=({\mathbb{V}}_{1}^{{\mathbb{N}}_{1}\times{\mathbb{N}}_{1}}\cap{\mathbb{V}}_{1})/{\mathcal{F}}_{1}\otimes{\mathcal{F}}^{\prime}_{1}=({\mathbb{V}}_{2}^{{\mathbb{N}}_{1}}\cap{\mathbb{V}}_{1})/{\mathcal{F}}^{\prime}_{1}={\mathbb{V}}_{3}^{\prime}
=(𝕍11×1𝕍1)/11=(𝕍22𝕍2)/2=𝕍3.\displaystyle=({\mathbb{V}}_{1}^{{\mathbb{N}}_{1}\times{\mathbb{N}}_{1}}\cap{\mathbb{V}}_{1})/{\mathcal{F}}_{1}\otimes{\mathcal{F}}^{\prime}_{1}=({\mathbb{V}}_{2}^{{\mathbb{N}}_{2}}\cap{\mathbb{V}}_{2})/{\mathcal{F}}_{2}={\mathbb{V}}_{3}.

Let i:𝕍2𝕍3′′=𝕍20/0′′i_{*}:{\mathbb{V}}_{2}\to{\mathbb{V}}_{3}^{\prime\prime}={\mathbb{V}}_{2}^{{\mathbb{N}}_{0}}/{\mathcal{F}}_{0}^{\prime\prime} be the bounded elementary embedding determined by (2) with i(c)=[ϕc]0′′i_{*}(c)=[\phi_{c}]_{{\mathcal{F}}_{0}^{\prime\prime}} for every c𝕍2c\in{\mathbb{V}}_{2}. Since 00/0′′=1{\mathbb{R}}_{0}^{{\mathbb{N}}_{0}}/{\mathcal{F}}^{\prime\prime}_{0}={\mathbb{R}}_{1} and 10/0′′=(11𝕍1)/1=2{\mathbb{R}}_{1}^{{\mathbb{N}}_{0}}/{\mathcal{F}}^{\prime\prime}_{0}=({\mathbb{R}}_{1}^{{\mathbb{N}}_{1}}\cap{\mathbb{V}}_{1})/{\mathcal{F}}_{1}={\mathbb{R}}_{2} we conclude that the map ii_{*} is also a bounded elementary embedding from model expansion (𝕍2;0,1)({\mathbb{V}}_{2};{\mathbb{R}}_{0},{\mathbb{R}}_{1}) to (𝕍3;1,2)({\mathbb{V}}_{3};{\mathbb{R}}_{1},{\mathbb{R}}_{2}). Since 𝕍2=𝕍10/0′′{\mathbb{V}}_{2}={\mathbb{V}}_{1}^{{\mathbb{N}}_{0}}/{\mathcal{F}}_{0}^{\prime\prime}, we have that i𝕍1i_{*}\!\upharpoonright\!{\mathbb{V}}_{1} coincides with i1i_{1} mentioned above when 𝕍2′′{\mathbb{V}}_{2}^{\prime\prime} is constructed, and is a bounded elementary embedding from (𝕍1;0)({\mathbb{V}}_{1};{\mathbb{R}}_{0}) to (𝕍2;1)({\mathbb{V}}_{2};{\mathbb{R}}_{1}). Hence Part 2 of Property 2.7 is verified.

Let i2:𝕍2𝕍3=(𝕍21𝕍1)/1i_{2}:{\mathbb{V}}_{2}\to{\mathbb{V}}_{3}^{\prime}=({\mathbb{V}}_{2}^{{\mathbb{N}}_{1}}\cap{\mathbb{V}}_{1})/{\mathcal{F}}_{1}^{\prime} be the bounded elementary embedding determined by (2) with i2(c)=[ϕc]1i_{2}(c)=[\phi_{c}]_{{\mathcal{F}}_{1}^{\prime}} for every c𝕍2c\in{\mathbb{V}}_{2}. Since every 𝕍1{\mathbb{V}}_{1}–internal function from 1{\mathbb{N}}_{1} to [n][n] for some n1n\in{\mathbb{N}}_{1} is equivalent to a constant function modulo 1{\mathcal{F}}^{\prime}_{1} we conclude that i21i_{2}\!\upharpoonright\!{\mathbb{N}}_{1} is an identity map. Similar to the argument for i1i_{1}, we have i2(c)32i_{2}(c)\in{\mathbb{N}}_{3}\setminus{\mathbb{N}}_{2} for every c21c\in{\mathbb{N}}_{2}\setminus{\mathbb{N}}_{1}. This verifies Part 3 of Property 2.7.

The following propositions are 𝕍j{\mathbb{V}}_{j}–versions of Proposition 2.5 and Proposition 2.6. Let 0<j30<j\leq 3.

Proposition 2.14

Let φ(a¯,x¯)\varphi(\overline{a},\overline{x}) be an \mathscr{L}–formula with bounded quantifiers, and parameters a¯\overline{a} and BmB^{m} be 𝕍j{\mathbb{V}}_{j}–internal. Then {b¯Bm𝕍jφ(a¯,b¯)}\left\{\overline{b}\in B^{m}\mid{\mathbb{V}}_{j}\models\varphi(\overline{a},\overline{b})\right\} is a 𝕍j{\mathbb{V}}_{j}–internal set.

Proposition 2.15

Let UU be an infinite proper initial segment of j{\mathbb{N}}_{j} and not 𝕍j{\mathbb{V}}_{j}–internal. Let AjA\subseteq{\mathbb{N}}_{j} be 𝕍j{\mathbb{V}}_{j}–internal. If AUA\cap U is upper unbounded in UU, then A(jU)A\cap({\mathbb{N}}_{j}\setminus U)\not=\emptyset.

We would like to mention that 𝕍1{\mathbb{V}}_{1}, 𝕍2{\mathbb{V}}_{2}, and 𝕍3{\mathbb{V}}_{3} are ultrapowers of 𝕍0{\mathbb{V}}_{0} modulo the ultrafilters 0{\mathcal{F}}_{0}, 00{\mathcal{F}}_{0}\otimes{\mathcal{F}}_{0}^{\prime}, and 000′′{\mathcal{F}}_{0}\otimes{\mathcal{F}}^{\prime}_{0}\otimes{\mathcal{F}}^{\prime\prime}_{0}, respectively. Hence, they are all countably saturated (cf. [1, Corollary 4.4.24]) although the countable saturation is not used in this paper. The proofs of Proposition 2.14 and Proposition 2.15 are the same as Proposition 2.5 and Proposition 2.6. Notice also that 𝕍2{\mathbb{V}}_{2} and 𝕍3{\mathbb{V}}_{3} are 𝕍1{\mathbb{V}}_{1}–internal ultrapowers of 𝕍1{\mathbb{V}}_{1} modulo the 𝕍1{\mathbb{V}}_{1}–internal ultrafilters 1{\mathcal{F}}_{1} and 11{\mathcal{F}}_{1}\otimes{\mathcal{F}}^{\prime}_{1}, respectively.

The ultrafilters 0{\mathcal{F}}_{0}, 0{\mathcal{F}}^{\prime}_{0}, and 0′′{\mathcal{F}}^{\prime\prime}_{0} do not have to be on the countable set 0{\mathbb{N}}_{0} and do not have to be the same. The only restriction is that they have to be in 𝕍0{\mathbb{V}}_{0}.

Iterated nonstandard extensions were used in combinatorial number theory before, e.g. in [2, 3, 6]. But we will use them in a new way by exploring the advantages of various bounded elementary embeddings between 𝕍j{\mathbb{V}}_{j} and 𝕍j{\mathbb{V}}_{j^{\prime}} (see Property 2.7). For most of the time we work within 𝕍2{\mathbb{V}}_{2}. The nonstandard universe 𝕍3{\mathbb{V}}_{3} is only used for one step in the proof of Lemma 5.1.

3 Nonstandard Versions of Some Facts

The Greek letters α,β,η,ϵ\alpha,\beta,\eta,\epsilon will represent “standard” reals unless specified otherwise. Let 0j<j30\leq j<j^{\prime}\leq 3 in this paragraph. An r3r\in{\mathbb{R}}_{3} is a 𝕍j{\mathbb{V}}_{j}–infinitesimal, denoted by rj0r\approx_{j}0, if |r|<1/n|r|<1/n for every njn\in{\mathbb{N}}_{j}. By an infinitesimal we mean a 𝕍0{\mathbb{V}}_{0}–infinitesimal. Denote stjst_{j} for the 𝕍j{\mathbb{V}}_{j}–standard part map, i.e., stj(r)st_{j}(r) is the unique real number rjr^{\prime}\in{\mathbb{R}}_{j} such that rrj0r-r^{\prime}\approx_{j}0 when r3(m,m)r\in{\mathbb{R}}_{3}\cap(-m,m) for some mjm\in{\mathbb{N}}_{j}. Notice that stjst_{j} and j{\mathbb{N}}_{j} are definable by a formula with bounded quantifiers and parameters in (𝕍j;j)({\mathbb{V}}_{j^{\prime}};{\mathbb{R}}_{j}). Sometimes, the subscript 0 will be dropped. For example, \approx means 0\approx_{0} and stst means st0st_{0}. For any two positive integers m,n3m,n\in{\mathbb{N}}_{3} we denote mnm\ll n for mjm\in{\mathbb{N}}_{j} and njjn\in{\mathbb{N}}_{j^{\prime}}\setminus{\mathbb{N}}_{j}. Hence m1m\gg 1 means m30m\in{\mathbb{N}}_{3}\setminus{\mathbb{N}}_{0}. Notice that if r1(m,m)r\in{\mathbb{R}}_{1}\cap(-m,m) for some m0m\in{\mathbb{N}}_{0} and st(r)=αst(r)=\alpha, then st(i1(r))=αst(i_{1}(r))=\alpha where i1i_{1} is in Part 2 of Property 2.7. This is true because i1(α)=αi_{1}(\alpha)=\alpha, i1(1/n)=1/ni_{1}(1/n)=1/n, and |rα|<1/n|r-\alpha|<1/n iff |i1(r)i1(α)|<i1(1/n)|i_{1}(r)-i_{1}(\alpha)|<i_{1}(1/n) for each n0n\in{\mathbb{N}}_{0}. Similarly, we have st1(i2(r))=st1(r)st_{1}(i_{2}(r))=st_{1}(r) for r2r\in{\mathbb{R}}_{2} in the domain of st1st_{1}.

Capital letters AA, BB, CC, \ldots represent sets of integers except HH, JJ, KK, NN which are reserved for integers in 30{\mathbb{N}}_{3}\setminus{\mathbb{N}}_{0}. The letter k3k\geq 3 represents exclusively the length of the arithmetic progression in Szemerédi’s theorem and ll represents an integer between 11 and kk. All unspecified sets mentioned in this paper will be either standard subsets of 0{\mathbb{N}}_{0} or 𝕍j{\mathbb{V}}_{j}–internal sets for some j=1,2,3j=1,2,3. For any n3n\in{\mathbb{N}}_{3} let [n]:={1,2,,n}[n]:=\{1,2,\ldots,n\}.

For any bounded set AjA\subseteq{\mathbb{N}}_{j^{\prime}} and njn\in{\mathbb{N}}_{j^{\prime}} denote δn(A)\delta_{n}(A) for the quantity |A|/n|A|/n in 𝕍j{\mathbb{V}}_{j^{\prime}} where |A||A| means the internal cardinality of AA in 𝕍j{\mathbb{V}}_{j^{\prime}}. Denote μnj(A):=stj(δn(A))\mu^{j}_{n}(A):=st_{j}(\delta_{n}(A)) for j<jj<j^{\prime}. Notice that δn\delta_{n} is an internal function while μnj\mu^{j}_{n} are often external functions but definable in (𝕍j;j)({\mathbb{V}}_{j^{\prime}};{\mathbb{R}}_{j}) for j>jj^{\prime}>j, i.e.,

μnj(A)=α iff njj(|δH(A)α|<1n).\mu^{j}_{n}(A)=\alpha\,\mbox{ iff }\,\forall n\in{\mathbb{N}}_{j^{\prime}}\cap{\mathbb{R}}_{j}\,\left(|\delta_{H}(A)-\alpha|<\frac{1}{n}\right).

We often write μn\mu_{n} for μn0\mu^{0}_{n}. If AΩA\subseteq\Omega and |Ω|=H|\Omega|=H, then μH(A)\mu_{H}(A) coincides with the Loeb measure of AA in Ω\Omega. The term δH\delta_{H} is often used for an internal argument.

\blacklozenge: The abbreviation a.p. stands for “arithmetic progression” and nn–a.p. stands for “nn-term arithmetic progression.”

The length of an a.p. pp is the number of terms in pp which can be written as |p||p|. The letters P,Q,RP,Q,R are reserved exclusively for a.p.’s of length 1\gg 1, and p,q,rp,q,r for a.p.’s of length kk or other standard length. When we run out of letters, we may also use x,y\vec{x},\vec{y} for kk–a.p.’s. If 1l|p|1\leq l\leq|p|, then p(l)p(l) represents the ll-th term of pp. We denote pAp\subseteq A for p(l)Ap(l)\in A for all 1l|p|1\leq l\leq|p|. We allow the common difference dd of an a.p. to be any integer including, occasionally, the trivial case for d=0d=0. If pp and qq are two a.p.’s of the same length, then pqp\oplus q represents the |p||p|–a.p. {p(l)+q(l)l=1,2,,|p|}\{p(l)+q(l)\mid l=1,2,\ldots,|p|\}. If pp is an a.p. and XX is an element or a set, then pXp\oplus X represents the sequence {p(l)+X1l|p|}\{p(l)+X\mid 1\leq l\leq|p|\}. By pqXp\sqsubseteq q\oplus X we mean p(l)q(l)+Xp(l)\in q(l)+X for every 1l|p|=|q|1\leq l\leq|p|=|q|.

If XjX\subseteq{\mathbb{R}}_{j} and X𝕍jX\in{\mathbb{V}}_{j} let supj(X){\sup}_{j}(X) be the supremum of XX in the sense of 𝕍j{\mathbb{V}}_{j}, i.e., the unique least upper bound of XX in j{\mathbb{R}}_{j}, or \infty if XX is unbounded above in j{\mathbb{R}}_{j}.

Let 1j<j31\leq j<j^{\prime}\leq 3. For any AjjA\subseteq{\mathbb{N}}_{j^{\prime}-j} and any collection of a.p.’s 𝒫𝕍jj{\mathcal{P}}\in{\mathbb{V}}_{j^{\prime}-j}, there exists X𝕍0X\in{\mathbb{V}}_{0} with X0X\subseteq{\mathbb{R}}_{0} (because every subset of 0{\mathbb{R}}_{0} is in 𝕍0{\mathbb{V}}_{0}) such that xXx\in X iff there exists a P𝒫P\in{\mathcal{P}} with |P|jj0|P|\in{\mathbb{N}}_{j^{\prime}-j}\setminus{\mathbb{N}}_{0} and μ|P|(AP)=x\mu_{|P|}(A\cap P)=x. By the elementality of ii_{*} in Part 2 of Property 2.7 we have that for any AjA\subseteq{\mathbb{N}}_{j^{\prime}} and any collection of a.p.’s 𝒫𝕍j{\mathcal{P}}\in{\mathbb{V}}_{j^{\prime}}, there exists X𝕍jX\in{\mathbb{V}}_{j} with XjX\subseteq{\mathbb{R}}_{j} such that xXx\in X iff there exists a P𝒫P\in{\mathcal{P}} with |P|jj|P|\in{\mathbb{N}}_{j^{\prime}}\setminus{\mathbb{N}}_{j} and μ|P|j(AP)=x\mu^{j}_{|P|}(A\cap P)=x. Therefore, the operator supj{\sup}_{j} and hence SDjS\!D^{j} below are well defined. Similarly, SDSjS\!D_{S}^{j} below is also well defined.

Definition 3.1

For 0j<j30\leq j<j^{\prime}\leq 3 and AjA\subseteq{\mathbb{N}}_{j^{\prime}} with |A|jj|A|\in{\mathbb{N}}_{j^{\prime}}\setminus{\mathbb{N}}_{j} the strong upper Banach density SDj(A)S\!D^{j}(A) of AA in 𝕍j{\mathbb{V}}_{j} is defined by

SDj(A):=supj{μ|P|j(AP)|P|jj}.S\!D^{j}(A):={\sup}_{j}\left\{\mu^{j}_{|P|}(A\cap P)\mid|P|\in{\mathbb{N}}_{j^{\prime}}\setminus{\mathbb{N}}_{j}\right\}. (8)

The letter PP above always represents an a.p. If SjS\subseteq{\mathbb{N}}_{j^{\prime}} has SDj(S)=ηjS\!D^{j}(S)=\eta\in{\mathbb{R}}_{j} and AjA\subseteq{\mathbb{N}}_{j^{\prime}}, the strong upper Banach density SDSjS\!D_{S}^{j} of AA relative to SS is defined by

SDSj(A):=supj{μ|P|j(AP)|P|jj, and μ|P|j(SP)=η}.S\!D^{j}_{S}(A):={\sup}_{j}\left\{\mu^{j}_{|P|}(A\cap P)\mid|P|\in{\mathbb{N}}_{j^{\prime}}\setminus{\mathbb{N}}_{j},\mbox{ and }\mu^{j}_{|P|}(S\cap P)=\eta\right\}. (9)

When SDSj(A)S\!D^{j}_{S}(A), defined in (9), is used in this paper, the set AA is often a subset of SS although there is no such restriction in the definition.

Definition 3.2

If A0A\subseteq{\mathbb{N}}_{0}, then the strong upper Banach density of AA is defined by SD(A):=SD0(i0(A))S\!D(A):=S\!D^{0}(i_{0}(A)) where SD0S\!D^{0} is defined by (8) and i0i_{0} is defined by (4).

We would like to point out that for standard sets A0A\subseteq{\mathbb{N}}_{0},

SD(A)=limn0,nsup{δ|p|(Ap)|p|n}.S\!D(A)=\lim_{n\in{\mathbb{N}}_{0},\,n\to\infty}\sup\{\delta_{|p|}(A\cap p)\mid|p|\geq n\}.

This equality will not be used. The purpose here is to give the reader some intuition because the right side is a standard expression.

The superscript 0 in SD0S\!D^{0} will be omitted. Notice that the upper density of a set A0A\subseteq{\mathbb{N}}_{0} is less than or equal to the upper Banach density of AA, which is less than or equal to the strong upper Banach density of AA. The strong upper Banach density of AA is the nonstandard version of the density of AA along a collection of arbitrarily long arithmetic progressions satisfying the double counting property in [9]. Of course, if we know that Szemerédi’s theorem is true, then SD(S)=η>0S\!D(S)=\eta>0 implies that η=1\eta=1. Also SD(S)=1S\!D(S)=1 and SDS(A)=α>0S\!D_{S}(A)=\alpha>0 imply that α=1\alpha=1.

Suppose that the strong upper Banach density of A0A\subseteq{\mathbb{N}}_{0} is a positive real number α\alpha. Instead of looking for kk–a.p.’s in AA we will look for kk–a.p.’s in i0(A)Pi_{0}(A)\cap P for some infinitely long a.p. PP such that the distribution of i0(A)Pi_{0}(A)\cap P in PP is very uniform, i.e., the measure and strong upper Banach density of i0(A)Pi_{0}(A)\cap P in PP are the same value α\alpha. The uniformity allows the use of an argument similar to the so called density increment argument in the standard literature. The next lemma is the beginning of this effort.

Lemma 3.3

For 0j<j30\leq j<j^{\prime}\leq 3 let ASjA\subseteq S\subseteq{\mathbb{N}}_{j^{\prime}} with |A|jj|A|\in{\mathbb{N}}_{j^{\prime}}\setminus{\mathbb{N}}_{j} and α,ηj\alpha,\eta\in{\mathbb{R}}_{j} with 0αη10\leq\alpha\leq\eta\leq 1. Then the following are true:

  1. 1.

    SDj(S)ηS\!D^{j}(S)\geq\eta iff there exists a PP with |P|jj|P|\in{\mathbb{N}}_{j^{\prime}}\setminus{\mathbb{N}}_{j} and μ|P|j(SP)η\mu^{j}_{|P|}(S\cap P)\geq\eta;

  2. 2.

    If SDj(S)=ηS\!D^{j}(S)=\eta, then there exists a PP with |P|jj|P|\in{\mathbb{N}}_{j^{\prime}}\setminus{\mathbb{N}}_{j} such that μ|P|j(SP)=SDj(SP)=η\mu^{j}_{|P|}(S\cap P)=S\!D^{j}(S\cap P)=\eta;

  3. 3.

    Suppose SDj(S)=ηS\!D^{j}(S)=\eta. Then SDSj(A)αS\!D^{j}_{S}(A)\geq\alpha iff there exists a PP with |P|jj|P|\in{\mathbb{N}}_{j^{\prime}}\setminus{\mathbb{N}}_{j}, μ|P|j(SP)=η\mu^{j}_{|P|}(S\cap P)=\eta, and μ|P|j(AP)α\mu^{j}_{|P|}(A\cap P)\geq\alpha;

  4. 4.

    Suppose SDj(S)=ηS\!D^{j}(S)=\eta. If SDSj(A)=αS\!D^{j}_{S}(A)=\alpha, then there exists a PP with |P|jj|P|\in{\mathbb{N}}_{j^{\prime}}\setminus{\mathbb{N}}_{j} such that μ|P|j(SP)=η\mu^{j}_{|P|}(S\cap P)=\eta and μ|P|j(AP)=SDSPj(AP)=α\mu^{j}_{|P|}(A\cap P)=S\!D^{j}_{S\cap P}(A\cap P)=\alpha.

Proof Part 1: If SDj(S)ηS\!D^{j}(S)\geq\eta, then there is a PnP_{n} with |Pn|jj|P_{n}|\in{\mathbb{N}}_{j^{\prime}}\setminus{\mathbb{N}}_{j} such that δ|Pn|(SPn)>η1/n\delta_{|P_{n}|}(S\cap P_{n})>\eta-1/n for every njn\in{\mathbb{N}}_{j}. Let

A:={njPj(|P|nδ|P|(SP)>η1/n)}.A:=\left\{n\in{\mathbb{N}}_{j^{\prime}}\mid\exists P\subseteq{\mathbb{N}}_{j^{\prime}}\,(|P|\geq n\wedge\delta_{|P|}(S\cap P)>\eta-1/n)\right\}.

Then AA is 𝕍j{\mathbb{V}}_{j^{\prime}}–internal and AjA\cap{\mathbb{N}}_{j} is unbounded above in j{\mathbb{N}}_{j}. By Proposition 2.15, there is a JAjJ\in A\setminus{\mathbb{N}}_{j}. Hence there is an a.p. PJjP_{J}\subseteq{\mathbb{N}}_{j^{\prime}} such that |PJ|Jjj|P_{J}|\geq J\in{\mathbb{N}}_{j^{\prime}}\setminus{\mathbb{N}}_{j} and δ|PJ|(SPJ)>η1/Jjη\delta_{|P_{J}|}(S\cap P_{J})>\eta-1/J\approx_{j}\eta. Therefore, μ|PJ|j(SPJ)η\mu^{j}_{|P_{J}|}(S\cap P_{J})\geq\eta. On the other hand, if μ|P|j(SP)η\mu^{j}_{|P|}(S\cap P)\geq\eta, then SDj(S)ηS\!D^{j}(S)\geq\eta by the definition of SDjS\!D^{j} in (8).

Part 2: If SDj(S)=ηS\!D^{j}(S)=\eta, we can find PP with |P|jj|P|\in{\mathbb{N}}_{j^{\prime}}\setminus{\mathbb{N}}_{j} such that μ|P|j(SP)=ηη\mu^{j}_{|P|}(S\cap P)=\eta^{\prime}\geq\eta by Part 1. Clearly, η=SDj(S)SDj(SP)μ|P|j(SP)=η\eta=S\!D^{j}(S)\geq S\!D^{j}(S\cap P)\geq\mu^{j}_{|P|}(S\cap P)=\eta^{\prime} by the definition of SDjS\!D^{j}. Hence η=η\eta=\eta^{\prime}.

Part 3: If SDSj(A)αS\!D^{j}_{S}(A)\geq\alpha, then there is a PP with |P|>n|P|>n such that |δ|P|(SP)η|<1/n|\delta_{|P|}(S\cap P)-\eta|<1/n and δ|P|(AP)>α1/n\delta_{|P|}(A\cap P)>\alpha-1/n for every njn\in{\mathbb{N}}_{j}. By Proposition 2.15 as in the proof of Part 1 there is a PJP_{J} for some JjjJ\in{\mathbb{N}}_{j^{\prime}}\setminus{\mathbb{N}}_{j} with |PJ|J|P_{J}|\geq J such that |δ|PJ|(SPJ)η|<1/J|\delta_{|P_{J}|}(S\cap P_{J})-\eta|<1/J and δ|PJ|(APJ)>α1/J\delta_{|P_{J}|}(A\cap P_{J})>\alpha-1/J, which implies μ|PJ|j(SP)=η\mu^{j}_{|P_{J}|}(S\cap P)=\eta and μ|PJ|j(APJ)α\mu^{j}_{|P_{J}|}(A\cap P_{J})\geq\alpha. On the other hand, if μ|P|(SP)=η\mu_{|P|}(S\cap P)=\eta and μ|P|j(AP)α\mu^{j}_{|P|}(A\cap P)\geq\alpha, then SDSj(A)αS\!D_{S}^{j}(A)\geq\alpha by the definition of SDSjS\!D_{S}^{j} in (9).

Part 4: If SDSj(A)=αS\!D^{j}_{S}(A)=\alpha, then μ|P|j(SP)=η\mu^{j}_{|P|}(S\cap P)=\eta and μ|P|j(AP)=αα\mu^{j}_{|P|}(A\cap P)=\alpha^{\prime}\geq\alpha for some PP with |P|jj|P|\in{\mathbb{N}}_{j^{\prime}}\setminus{\mathbb{N}}_{j} by Part 3. Clearly, α=SDSj(A)SDSPj(AP)μ|P|j(AP)=α\alpha=S\!D^{j}_{S}(A)\geq S\!D^{j}_{S\cap P}(A\cap P)\geq\mu^{j}_{|P|}(A\cap P)=\alpha^{\prime} by the definition of SDSjS\!D_{S}^{j}. Hence α=α\alpha=\alpha^{\prime}. \blacksquare

The following lemma is the internal version of an argument similar to so-called the double counting property in the standard literature. Let 1HN/21\ll H\leq N/2. Roughly speaking, if CC is very uniformly distributed in [N][N] with measure α\alpha, then for almost all x[NH]x\in[N-H] the measure of C(x+[H])C\cap(x+[H]) inside x+[H]x+[H] is α\alpha. Since the measure μH\mu_{H} is not an internal function we use δH\delta_{H} instead and require |δH(C(x+[H])α|<1/J|\delta_{H}(C\cap(x+[H])-\alpha|<1/J for some infinite JJ instead of μH(C(x+[H])=α\mu_{H}(C\cap(x+[H])=\alpha. The lemma is stated in a more general case with 𝕍j{\mathbb{V}}_{j} being viewed as the “standard” universe in 𝕍j{\mathbb{V}}_{j^{\prime}}.

Lemma 3.4

Let N,HjjN,H\in{\mathbb{N}}_{j^{\prime}}\setminus{\mathbb{N}}_{j}, HN/2H\leq N/2, and C[N]C\subseteq[N] with μNj(C)=SDj(C)=αj\mu^{j}_{N}(C)=S\!D^{j}(C)=\alpha\in{\mathbb{R}}_{j} for 0j<j30\leq j<j^{\prime}\leq 3. For each njn\in{\mathbb{N}}_{j^{\prime}} let

Dn,H,C:={x[NH]|δH(C(x+[H]))α|<1n}.D_{n,H,C}:=\left\{x\in[N-H]\mid\left|\delta_{H}(C\cap(x+[H]))-\alpha\right|<\frac{1}{n}\right\}. (10)

Then there exists a JjjJ\in{\mathbb{N}}_{j^{\prime}}\setminus{\mathbb{N}}_{j} such that μNHj(DJ,H,C)=1\mu^{j}_{N-H}(D_{J,H,C})=1.

Notice that Dn,H,CDn,H,CD_{n,H,C}\subseteq D_{n^{\prime},H,C} if nnn\geq n^{\prime}.

Proof Fix NN, HH, and CC. The subscripts HH and CC in Dn,H,CD_{n,H,C} will be omitted in the proof. If stj(H/N)>0st_{j}(H/N)>0, then for every x[NH]x\in[N-H] we have μHj(x+[H])=α\mu^{j}_{H}(x+[H])=\alpha by the supremality of α\alpha. Hence the maximal JJ with JHJ\leq H such that |δH(A(x+[H]))α|<1/J|\delta_{H}(A\cap(x+[H]))-\alpha|<1/J for every x[NH]x\in[N-H] is in jj{\mathbb{N}}_{j^{\prime}}\setminus{\mathbb{N}}_{j}. Now DJ=[NH]D_{J}=[N-H] works.

Assume that stj(H/N)=0st_{j}(H/N)=0. So, μNHj\mu^{j}_{N-H} and μNj\mu^{j}_{N} coincide. If δN(Dn)j1\delta_{N}(D_{n})\approx_{j}1 for every njn\in{\mathbb{N}}_{j}, then the maximal JJ satisfying |δN(DJ)1|<1/J|\delta_{N}(D_{J})-1|<1/J must be in jj{\mathbb{N}}_{j^{\prime}}\setminus{\mathbb{N}}_{j} by Proposition 2.15. Hence μNj(DJ)=1\mu^{j}_{N}(D_{J})=1. So we can assume that μNj(Dn)<1\mu^{j}_{N}(D_{n})<1 for some njn\in{\mathbb{N}}_{j} and derive a contradiction.

Notice that for each x[NH]x\in[N-H], it is impossible to have μHj(C(x+[H]))>α=SDj(C)\mu^{j}_{H}(C\cap(x+[H]))>\alpha=S\!D^{j}(C) by the definition of SDjS\!D^{j}. Let D¯n:=[NH]Dn\overline{D}_{n}:=[N-H]\setminus D_{n}. Then μNj(D¯n)=1μNj(Dn)>0\mu^{j}_{N}(\overline{D}_{n})=1-\mu^{j}_{N}(D_{n})>0. Notice that xD¯nx\in\overline{D}_{n} implies δH(C(x+[H]))α1/n\delta_{H}(C\cap(x+[H]))\leq\alpha-1/n. By the following double counting argument, by ignoring some 𝕍j{\mathbb{V}}_{j}–infinitesimal amount inside stjst_{j}, we have

α=stj(1Hy=1HδN(Cy))=stj(1HNy=1Hx=1NχC(x+y))\displaystyle\alpha=st_{j}\left(\frac{1}{H}\sum_{y=1}^{H}\delta_{N}(C-y)\right)=st_{j}\left(\frac{1}{HN}\sum_{y=1}^{H}\sum_{x=1}^{N}\chi_{C}(x+y)\right)
=stj(1NHx=1Ny=1HχC(x+y))=stj(1Nx=1NδH(C(x+[H])))\displaystyle=st_{j}\left(\frac{1}{NH}\sum_{x=1}^{N}\sum_{y=1}^{H}\chi_{C}(x+y)\right)=st_{j}\left(\frac{1}{N}\sum_{x=1}^{N}\delta_{H}(C\cap(x+[H]))\right)
=stj(1NxDnδH(C(x+[H]))+1NxD¯nδH(C(x+[H])))\displaystyle=st_{j}\left(\frac{1}{N}\sum_{x\in D_{n}}\delta_{H}(C\cap(x+[H]))+\frac{1}{N}\sum_{x\in\overline{D}_{n}}\delta_{H}(C\cap(x+[H]))\right)
αμNj(Dn)+(α1n)μNj(D¯n)<α\displaystyle\leq\alpha\mu^{j}_{N}(D_{n})+\left(\alpha-\frac{1}{n}\right)\mu^{j}_{N}(\overline{D}_{n})<\alpha

which is absurd. This completes the proof. \blacksquare

Suppose 0j<j30\leq j<j^{\prime}\leq 3, NH1N\geq H\gg 1 in j{\mathbb{N}}_{j^{\prime}}, U[N]U\subseteq[N], AS[N]A\subseteq S\subseteq[N], 0αη10\leq\alpha\leq\eta\leq 1, and x[N]x\in[N]. For each njn\in{\mathbb{N}}_{j} let ξ(x,α,η,A,S,U,H,n)\xi(x,\alpha,\eta,A,S,U,H,n) be the following internal statement:

|δH(x+[H])U)1|<1/n,|δH((x+[H])S)η|<1/n, and|δH((x+[H])A)α|<1/n.\begin{array}[]{rcl}\vspace{0.1in}|\delta_{H}(x+[H])\cap U)-1|&<&1/n,\\ \vspace{0.1in}|\delta_{H}((x+[H])\cap S)-\eta|&<&1/n,\mbox{ and}\\ |\delta_{H}((x+[H])\cap A)-\alpha|&<&1/n.\end{array} (11)

The statement ξ(x,α,η,A,S,U,H,n)\xi(x,\alpha,\eta,A,S,U,H,n) infers that the densities of A,S,UA,S,U in the interval x+[H]x+[H] go to α,η,1\alpha,\eta,1, respectively, as nn\to\infty in j{\mathbb{N}}_{j}. The statement ξ\xi will be referred a few times in Lemma 5.1 and its proof.

The following lemma is the application of Lemma 3.4 to the sets U,S,AU,S,A simultaneously.

Lemma 3.5

Let NjjN\in{\mathbb{N}}_{j^{\prime}}\setminus{\mathbb{N}}_{j}, U[N]U\subseteq[N], and AS[N]A\subseteq S\subseteq[N] be such that μNj(U)=1\mu^{j}_{N}(U)=1, μNj(S)=SD(S)=η\mu^{j}_{N}(S)=S\!D(S)=\eta, and μNj(A)=SDSj(A)=α\mu^{j}_{N}(A)=S\!D^{j}_{S}(A)=\alpha for some η,αj\eta,\alpha\in{\mathbb{R}}_{j} and 0j<j30\leq j<j^{\prime}\leq 3. For any n,hjn,h\in{\mathbb{N}}_{j^{\prime}} let

Gn,h:={x[Nh]𝕍jξ(x,α,η,A,S,U,h,n)}.G_{n,h}:=\{x\in[N-h]\mid{\mathbb{V}}_{j^{\prime}}\models\xi(x,\alpha,\eta,A,S,U,h,n)\}. (12)
  1. (a)

    For each HjjH\in{\mathbb{N}}_{j^{\prime}}\setminus{\mathbb{N}}_{j} with HN/2H\leq N/2 there exists a JjjJ\in{\mathbb{N}}_{j^{\prime}}\setminus{\mathbb{N}}_{j} such that μNHj(GJ,H)=1\mu^{j}_{N-H}(G_{J,H})=1;

  2. (b)

    For each njn\in{\mathbb{N}}_{j}, there is an hnjh_{n}\in{\mathbb{N}}_{j} with hn>nh_{n}>n such that δN(Gn,hn)>11/n\delta_{N}(G_{n,h_{n}})>1-1/n.

Proof Part (a): Applying Lemma 3.4 for UU and SS we can find J1,J2jjJ_{1},J_{2}\in{\mathbb{N}}_{j^{\prime}}\setminus{\mathbb{N}}_{j} such that μNHj(DJ1,H,U)=1\mu^{j}_{N-H}(D_{J_{1},H,U})=1 and μNHj(DJ2,H,S)=1\mu^{j}_{N-H}(D_{J_{2},H,S})=1 where Dn,h,CD_{n,h,C} is defined in (10) and α\alpha is replaced by 11 for UU and η\eta for SS. Let G:=DJ1,H,UDJ2,H,SG^{\prime}:=D_{J_{1},H,U}\cap D_{J_{2},H,S}. For each nmin{J1,J2}n\leq\min\{J_{1},J_{2}\} let

G¯n′′:={x[NH]δH(A(x+[H]))>α+1n}, and\overline{G}^{\prime\prime}_{n}:=\left\{x\in[N-H]\mid\delta_{H}(A\cap(x+[H]))>\alpha+\frac{1}{n}\right\},\mbox{ and}
G¯n′′:={x[NH]δH(A(x+[H]))<α1n}.\underline{G}^{\prime\prime}_{n}:=\left\{x\in[N-H]\mid\delta_{H}(A\cap(x+[H]))<\alpha-\frac{1}{n}\right\}.

Notice that both G¯n′′\overline{G}^{\prime\prime}_{n} and G¯n′′\underline{G}^{\prime\prime}_{n} are 𝕍j{\mathbb{V}}_{j^{\prime}}–internal. If μNHj(G¯n′′)>0\mu^{j}_{N-H}(\overline{G}^{\prime\prime}_{n})>0 for some njn\in{\mathbb{N}}_{j}, then G¯n′′G\overline{G}^{\prime\prime}_{n}\cap G^{\prime}\not=\emptyset. Let x0G¯n′′Gx_{0}\in\overline{G}^{\prime\prime}_{n}\cap G^{\prime}. Then we have μHj(S(x0+[H]))=η\mu^{j}_{H}(S\cap(x_{0}+[H]))=\eta and μHj(A(x0+[H]))>α+1/n\mu^{j}_{H}(A\cap(x_{0}+[H]))>\alpha+1/n, which contradicts SDSj(A)=αS\!D^{j}_{S}(A)=\alpha. Hence δNH(G¯n′′)j0\delta_{N-H}(\overline{G}^{\prime\prime}_{n})\approx_{j}0 for every njn\in{\mathbb{N}}_{j}. By Proposition 2.15 we can find J+jjJ_{+}\in{\mathbb{N}}_{j^{\prime}}\setminus{\mathbb{N}}_{j} such that μNHj(G¯n′′)=0\mu^{j}_{N-H}(\overline{G}^{\prime\prime}_{n})=0 for any nJ+n\leq J_{+}. If μNHj(G¯n′′)>0\mu^{j}_{N-H}(\underline{G}^{\prime\prime}_{n})>0 for some njn\in{\mathbb{N}}_{j}, then μNHj(G¯m′′)>0\mu^{j}_{N-H}(\overline{G}_{m}^{\prime\prime})>0 for some mjm\in{\mathbb{N}}_{j} by the fact that μNHj(A)=α\mu^{j}_{N-H}(A)=\alpha. Hence δNH(G¯n′′)j0\delta_{N-H}(\underline{G}_{n}^{\prime\prime})\approx_{j}0 for every njn\in{\mathbb{N}}_{j}. By Proposition 2.15 again we can find JjjJ_{-}\in{\mathbb{N}}_{j^{\prime}}\setminus{\mathbb{N}}_{j} such that μNHj(G¯n′′)=0\mu^{j}_{N-H}(\underline{G}^{\prime\prime}_{n})=0 for any nJn\leq J_{-}. The proof is complete by setting J:=min{J1,J2,J+,J}J:=\min\{J_{1},J_{2},J_{+},J_{-}\} and

GJ,H:=(DJ,H,UDJ.H,S)(G¯J′′G¯J′′).G_{J,H}:=(D_{J,H,U}\cap D_{J.H,S})\setminus(\overline{G}^{\prime\prime}_{J}\cup\underline{G}^{\prime\prime}_{J}).

Part (b): Suppose Part (b) is not true. Then there exists an njn\in{\mathbb{N}}_{j} such that δNh(Gn,h)11/n\delta_{N-h}(G_{n,h})\leq 1-1/n for any h>nh>n in j{\mathbb{N}}_{j}. By Proposition 2.15 there is an HjjH\in{\mathbb{N}}_{j^{\prime}}\setminus{\mathbb{N}}_{j} such that δNH(Gn,H)11/n\delta_{N-H}(G_{n,H})\leq 1-1/n. By Part (a) there is a JnJ\gg n such that μNHj(GJ,H)=1\mu^{j}_{N-H}(G_{J,H})=1. We have a contradiction because n<Jn<J and hence GJ,HGn,HG_{J,H}\subseteq G_{n,H}. \blacksquare

Notice that for a given nn one can choose hnh_{n} to be the least such that δN(Gn,hn)>11/n\delta_{N}(G_{n,h_{n}})>1-1/n in Lemma 3.5 (b). So we can assume that hnh_{n} is an internal function of nn. Hence we can assume that Gn,hnG_{n,h_{n}} is also an internal function of nn.

4 Mixing Lemma

We work within 𝕍j{\mathbb{V}}_{j^{\prime}} for 0<j30<j^{\prime}\leq 3 in this section. Any unspecified sets are 𝕍j{\mathbb{V}}_{j^{\prime}}–internal. The letter VV will sometimes be used for a set other than the standard/nonstandard universes in §2. Hopefully, no confusion will arise. The following standard lemma is a consequence of Szemerédi’s Regularity Lemma in [8]. The proof of the lemma can be found in the appendix of [9].

Lemma 4.1

Let U,WU,W be finite sets, let ϵ>0\epsilon>0, and for each wWw\in W, let EwE_{w} be a subset of  UU. Then there exists a partition U=U1U2UnϵU=U_{1}\cup U_{2}\cup\cdots\cup U_{n_{\epsilon}} for some nϵ0n_{\epsilon}\in{\mathbb{N}}_{0}, and real numbers 0cu,w10\leq c_{u,w}\leq 1 in 0{\mathbb{R}}_{0} for u[nϵ]u\in[n_{\epsilon}] and wWw\in W such that for any set FUF\subseteq U, one has

FEw|u=1nϵcu,w|FUuϵ|U|\left||F\cap E_{w}|-\sum_{u=1}^{n_{\epsilon}}c_{u,w}|F\cap U_{u}|\right|\leq\epsilon|U|

for all but ϵ|W|\epsilon|W| values of wWw\in W.

The following lemma, the nonstandard version of so–called mixing lemma in [9], can be derived from Lemma 4.1. We present a proof similar to the proof in [9] in a nonstandard setting. Part (i) and Part (ii) of the lemma are used to prove part (iii) and only Part (iii) will be referred in the proof of Lemma 5.1.

Lemma 4.2 (Mixing Lemma)

Let Nj0N\in{\mathbb{N}}_{j^{\prime}}\setminus{\mathbb{N}}_{0}, AS[N]A\subseteq S\subseteq[N], 1HN/21\ll H\leq N/2, and R[NH]R\subseteq[N-H] be an a.p. with |R|1|R|\gg 1 such that

μN(S)=SD(S)=η>0,μN(A)=SDS(A)=α>0,\mu_{N}(S)=S\!D(S)=\eta>0,\,\mu_{N}(A)=S\!D_{S}(A)=\alpha>0, (13)
μH((x+[H])S)=η, and μH((x+[H])A)=α\mu_{H}((x+[H])\cap S)=\eta,\,\mbox{ and }\,\mu_{H}((x+[H])\cap A)=\alpha (14)

for every xRx\in R. Then the following are true.

  1. (i)

    For any set E[H]E\subseteq[H] with μH(E)>0\mu_{H}(E)>0, there is an xRx\in R such that

    μH(A(x+E))αμH(E);\mu_{H}(A\cap(x+E))\geq\alpha\mu_{H}(E);
  2. (ii)

    Let m1m\gg 1 be such that the van der Waerden number Γ(3m,m)|R|\Gamma\left(3^{m},m\right)\leq|R|. For any internal partition {Unn[m]}\{U_{n}\mid n\in[m]\} of [H][H] there exists an mm–a.p. PRP\subseteq R, a set I[m]I\subseteq[m] with μH(UI)=1\mu_{H}(U_{I})=1 where UI={UnnI}U_{I}=\bigcup\{U_{n}\mid n\in I\}, and an infinitesimal ϵ>0\epsilon>0 such that

    |δH(A(x+Un))αδH(Un)|ϵδH(Un)|\delta_{H}(A\cap(x+U_{n}))-\alpha\delta_{H}(U_{n})|\leq\epsilon\delta_{H}(U_{n})

    for all nIn\in I and all xPx\in P;

  3. (iii)

    Given an internal collection of sets {Ew[H]wW}\{E_{w}\subseteq[H]\mid w\in W\} with |W|1|W|\gg 1 and μH(Ew)>0\mu_{H}(E_{w})>0 for every wWw\in W, there exists an xRx\in R and TWT\subseteq W such that μ|W|(T)=1\mu_{|W|}(T)=1 and

    μH(A(x+Ew))=αμH(Ew)\mu_{H}(A\cap(x+E_{w}))=\alpha\mu_{H}(E_{w})

    for every wTw\in T.

Proof Part (i): Assume that (i) is not true. For each xRx\in R let rxr_{x} be such that δH(A(E+x))=(αrx)δH(E)\delta_{H}(A\cap(E+x))=(\alpha-r_{x})\delta_{H}(E). Then rxr_{x} must be positive non-infinitesimal. We can set r:=min{rxxR}r:=\min\{r_{x}\mid x\in R\} since the function xrxx\mapsto r_{x} is internal. Clearly, the number rr is positive non-infinitesimal. Hence δH(A(E+x))(αr)δH(E)\delta_{H}(A\cap(E+x))\leq(\alpha-r)\delta_{H}(E) for all xRx\in R. Notice that by (13) and (14), for μH\mu_{H}–almost all y[H]y\in[H] we have μ|R|(S(y+R))=η\mu_{|R|}(S\cap(y+R))=\eta which implies that for μH\mu_{H}–almost all y[H]y\in[H] we have μ|R|(A(y+R))=α\mu_{|R|}(A\cap(y+R))=\alpha. So

αμH(E)1HyE1|R|xRχA(x+y)=1|R|xR1Hy=1HχA(E+x)(x+y)\displaystyle\alpha\mu_{H}(E)\approx\frac{1}{H}\sum_{y\in E}\frac{1}{|R|}\sum_{x\in R}\chi_{A}(x+y)=\frac{1}{|R|}\sum_{x\in R}\frac{1}{H}\sum_{y=1}^{H}\chi_{A\cap(E+x)}(x+y)
1|R|xR(αr)δH(E)=(αr)δH(E)(αst(r))μH(E)<αμH(E),\displaystyle\leq\frac{1}{|R|}\sum_{x\in R}(\alpha-r)\delta_{H}(E)=(\alpha-r)\delta_{H}(E)\approx(\alpha-st(r))\mu_{H}(E)<\alpha\mu_{H}(E),

which is absurd.

Part (ii): To make the argument explicitly internal we use δH\delta_{H} instead of μH\mu_{H}. For each tjt\in{\mathbb{N}}_{j^{\prime}}, xRx\in R, and n[m]n\in[m] let

cnt(x)={1 if δH((x+Un)A)(α+1t)δH(Un),0 if (α1t)δH(Un)<δH((x+Un)A)<(α+1t)δH(Un),1 if δH((x+Un)A)(α1t)δH(Un).c^{t}_{n}(x)=\left\{\begin{array}[]{cl}\vskip 6.0pt plus 2.0pt minus 2.0pt1&\,\mbox{ if }\,\delta_{H}((x+U_{n})\cap A)\geq\left(\alpha+\frac{1}{t}\right)\delta_{H}(U_{n}),\\ \vskip 6.0pt plus 2.0pt minus 2.0pt0&\,\mbox{ if }\,\left(\alpha-\frac{1}{t}\right)\delta_{H}(U_{n})<\delta_{H}((x+U_{n})\cap A)<\left(\alpha+\frac{1}{t}\right)\delta_{H}(U_{n}),\\ -1&\,\mbox{ if }\,\delta_{H}((x+U_{n})\cap A)\leq\left(\alpha-\frac{1}{t}\right)\delta_{H}(U_{n}).\end{array}\right.

and let ct:P{1,0,1}[m]c^{t}:P\to\{-1,0,1\}^{[m]} be such that ct(x)(n)=cnt(x)c^{t}(x)(n)=c^{t}_{n}(x). For each t0t\in{\mathbb{N}}_{0}, since the van der Waerden number Γ(3m,m)|R|\Gamma(3^{m},m)\leq|R|, there exists an mm–a.p. PtRP_{t}\subseteq R such that ct(x)=ct(x)c^{t}(x)=c^{t}(x^{\prime}) for any x,xPtx,x^{\prime}\in P_{t}. For each xPtx\in P_{t} let

It+={n[m]ct(x)(n)=1},It={n[m]ct(x)(n)=1}, andIt=[m](It+It), and\begin{array}[]{rcl}\vskip 6.0pt plus 2.0pt minus 2.0ptI^{+}_{t}&=&\{n\in[m]\mid c^{t}(x)(n)=1\},\\ \vskip 6.0pt plus 2.0pt minus 2.0ptI^{-}_{t}&=&\{n\in[m]\mid c^{t}(x)(n)=-1\},\mbox{ and}\\ I_{t}&=&[m]\setminus(I^{+}_{t}\cup I^{-}_{t}),\mbox{ and}\end{array}
Ut+={UnnIt+},Ut={UnnIt}, andUt=[H](Ut+Ut).\begin{array}[]{rcl}\vskip 6.0pt plus 2.0pt minus 2.0ptU^{+}_{t}&=&\bigcup\{U_{n}\mid n\in I^{+}_{t}\},\\ \vskip 6.0pt plus 2.0pt minus 2.0ptU^{-}_{t}&=&\bigcup\{U_{n}\mid n\in I^{-}_{t}\},\,\mbox{ and}\\ U_{t}&=&[H]\setminus(U^{+}_{t}\cup U^{-}_{t}).\end{array}

Clearly, δH((x+Ut)A)(α1/t)δH(Ut)\delta_{H}((x+U^{-}_{t})\cap A)\leq(\alpha-1/t)\delta_{H}(U^{-}_{t}) because UtU^{-}_{t} is a disjoint union of the UnU_{n}’s for nItn\in I^{-}_{t}. Since t0t\in{\mathbb{N}}_{0} we have that μH(Ut)=0\mu_{H}(U^{-}_{t})=0 by (i) with PtP_{t} in the place of RR and UtU^{-}_{t} in the place of EE. Notice that δH(A(x+Ut+))(α+1/t)δH(Ut+)\delta_{H}(A\cap(x+U^{+}_{t}))\geq(\alpha+1/t)\delta_{H}(U^{+}_{t}). Since αμH(A(x+Ut+))(α+1/t)μH(Ut+)\alpha\geq\mu_{H}(A\cap(x+U^{+}_{t}))\geq(\alpha+1/t)\mu_{H}(U^{+}_{t}), we have that μH(Ut+)<1\mu_{H}(U^{+}_{t})<1, which implies μH(Ut)>0\mu_{H}(U_{t})>0. If μH(Ut+)>0\mu_{H}(U^{+}_{t})>0, then δH(A(x+Ut+))(α+1/t)δH(Ut+)\delta_{H}(A\cap(x+U^{+}_{t}))\geq(\alpha+1/t)\delta_{H}(U^{+}_{t}) implies μH(A(x+Ut))<αμH(Ut)\mu_{H}(A\cap(x+U_{t}))<\alpha\mu_{H}(U_{t}) for all xPtx\in P_{t}, which again contradicts (i). Hence μH(Ut+)=0\mu_{H}(U^{+}_{t})=0 and therefore, δH(Ut)>11/t\delta_{H}(U_{t})>1-1/t is true for every t0t\in{\mathbb{N}}_{0}.

Since the set of all tjt\in{\mathbb{N}}_{j^{\prime}} with δH(Ut)>11/t\delta_{H}(U_{t})>1-1/t is 𝕍j{\mathbb{V}}_{j^{\prime}}–internal, by Proposition 2.15 there is a J1J\gg 1 such that δH(UJ)>11/J1\delta_{H}(U_{J})>1-1/J\approx 1. The proof of (ii) is completed by letting P:=PJP:=P_{J}, I:=IJI:=I_{J}, and UI:=UJU_{I}:=U_{J}.

Part (iii): Choose a sufficiently large positive infinitesimal ϵ\epsilon satisfying that there is an internal partition of [H]=U0U1Um[H]=U_{0}\cup U_{1}\cup\cdots\cup U_{m} and real numbers 0cn,w10\leq c_{n,w}\leq 1 for each n[m]n\in[m] and wWw\in W such that the van der Waerden number Γ(3m,m)|R|\Gamma(3^{m},m)\leq|R|, and for any internal set F[H]F\subseteq[H] there is a TFWT_{F}\subseteq W with |WTF|ϵ|W||W\setminus T_{F}|\leq\epsilon|W| such that

FEw|n=1mcn,w|FUnϵH\left||F\cap E_{w}|-\sum_{n=1}^{m}c_{n,w}|F\cap U_{n}|\right|\leq\epsilon H (15)

for all wTFw\in T_{F}. Notice that such ϵ\epsilon exists because if ϵ\epsilon is a standard positive real, then m=nϵm=n_{\epsilon} is in 0{\mathbb{N}}_{0}. From (15) with FF being replaced by [H][H] we have

Ew|n=1mcn,w|UnϵH\left||E_{w}|-\sum_{n=1}^{m}c_{n,w}|U_{n}|\right|\leq\epsilon H (16)

for all wT[H]w\in T_{[H]}. By (ii) we can find a PRP\subseteq R of length mm, a positive infinitesimal ϵ1\epsilon_{1}, and I[m]I\subseteq[m] where, for some xPx\in P,

I:={n[m]|δH((x+Un)A)αδH(Un)|<ϵ1δH(Un)}I:=\left\{n\in[m]\mid|\delta_{H}((x+U_{n})\cap A)-\alpha\delta_{H}(U_{n})|<\epsilon_{1}\delta_{H}(U_{n})\right\}

(II is independent of the choice of xx), and V:={UnnI}V:=\bigcup\{U_{n}\mid n\in I\} with μH(V)=1\mu_{H}(V)=1. Let I=[m]II^{\prime}=[m]\setminus I and V=[H]VV^{\prime}=[H]\setminus V. Then for each wT:=T[H]T(Ax)[H]w\in T:=T_{[H]}\cap T_{(A-x)\cap[H]} we have

|δH(A(x+Ew))αδH(Ew)|\displaystyle\left|\delta_{H}(A\cap(x+E_{w}))-\alpha\delta_{H}(E_{w})\right|
1H(||A(x+Ew)|n[m]cn,w|A(x+Un)||\displaystyle\leq\frac{1}{H}\left(\left||A\cap(x+E_{w})|-\sum_{n\in[m]}c_{n,w}|A\cap(x+U_{n})|\right|\right.
+|n[m]cn,w|A(x+Un)|n[m]cn,wα|Un||\displaystyle\quad+\left|\sum_{n\in[m]}c_{n,w}|A\cap(x+U_{n})|-\sum_{n\in[m]}c_{n,w}\alpha|U_{n}|\right|
+|αn[m]cn,w|Un|α|Ew||)\displaystyle\quad\left.+\left|\alpha\sum_{n\in[m]}c_{n,w}|U_{n}|-\alpha|E_{w}|\right|\right)
ϵ+1HnIcn,wϵ1|Un|+2δH(V)+αϵ\displaystyle\leq\epsilon+\frac{1}{H}\sum_{n\in I}c_{n,w}\epsilon_{1}|U_{n}|+2\delta_{H}(V^{\prime})+\alpha\epsilon
ϵ+ϵ1δH(V)+2δH(V)+αϵ0.\displaystyle\leq\epsilon+\epsilon_{1}\delta_{H}(V)+2\delta_{H}(V^{\prime})+\alpha\epsilon\approx 0.

Hence μH(A(x+Ew))=αμH(Ew)\mu_{H}(A\cap(x+E_{w}))=\alpha\mu_{H}(E_{w}) for all wTw\in T. Notice that μ|W|(T)=1\mu_{|W|}(T)=1 because ϵ0\epsilon\approx 0 and μ|W|(T[H])=μ|W|(T[H](Ax))=1\mu_{|W|}(T_{[H]})=\mu_{|W|}(T_{[H]\cap(A-x)})=1. \blacksquare

The set SS in Lemma 4.2, although seems unnecessary, is needed in the proof of Lemma 5.1.

5 Proof of Szemerédi’s Theorem

We work within 𝕍2{\mathbb{V}}_{2} in this section except in the proof of Claim 1 in Lemma 5.1 where 𝕍3{\mathbb{V}}_{3} is needed.

Szemerédi’s theorem is an easy consequence of Lemma 5.1, denoted by 𝐋(m){\bf L}(m) for all m[k]m\in[k]. For an integer n2k+1n\geq 2k+1 define an interval Cn[n]C_{n}\subseteq[n] by

Cn:=[kn2k+1,(k+1)n2k+1].C_{n}:=\left[\left\lceil\frac{kn}{2k+1}\right\rceil,\,\left\lfloor\frac{(k+1)n}{2k+1}\right\rfloor\right]. (17)

The set CnC_{n} is the subinterval of [n][n] in the middle of [n][n] with the length n/(2k+1)±ι\lfloor n/(2k+1)\rfloor\pm\iota for ι=0\iota=0 or 11. If n1n\gg 1, then μn(Cn)=1/(2k+1)\mu_{n}(C_{n})=1/(2k+1). For notational convenience we denote

D:=3k3 and η0:=11D.D:=3k^{3}\,\mbox{ and }\,\eta_{0}:=1-\frac{1}{D}. (18)

\blacklozenge: Fix a K10K\in{\mathbb{N}}_{1}\setminus{\mathbb{N}}_{0}. The number KK is the length of an interval which will play an important role in Lemma 5.1. Keeping KK unchanged is one of the advantages from nonstandard analysis, which is unavailable in the standard setting.

There is a summary of ideas used in the proof of Lemma 5.1 right after the proof. It explains some motivation of the steps taken in the proof.

Lemma 5.1 (𝐋(m){\bf L}(m))

Given any α>0\alpha>0, η>η0\eta>\eta_{0}, any N21N\in{\mathbb{N}}_{2}\setminus\!{\mathbb{N}}_{1}, and any AS[N]A\subseteq S\subseteq[N] and U[N]U\subseteq[N] with

μN(U)=1,μN(S)=SD(S)=η, and μN(A)=SDS(A)=α,\mu_{N}(U)=1,\mu_{N}(S)=S\!D(S)=\eta,\mbox{ and }\,\mu_{N}(A)=S\!D_{S}(A)=\alpha, (19)

the following are true:

  1. L

    (m)1(α,η,N,A,S,U,K){}_{1}(m)(\alpha,\eta,N,A,S,U,K): There exists a kk–a.p. xU\vec{x}\subseteq U with x[K][N]\vec{x}\oplus[K]\subseteq[N] satisfying the statement (n0)ξ(x(l),α,η,A,S,U,K,n)(\forall n\in{\mathbb{N}}_{0})\,\xi(\vec{x}(l),\alpha,\eta,A,S,U,K,n) for l[k]l\in[k], where ξ\xi is defined in (11), and there exist TlCKT_{l}\subseteq C_{K} with μ|CK|(Tl)=1\mu_{|C_{K}|}(T_{l})=1 where CKC_{K} is defined in (17) and Vl[K]V_{l}\subseteq[K] with μK(Vl)=1\mu_{K}(V_{l})=1 for every lml\geq m, and collections of kk–a.p.’s

    𝒫:={𝒫l,ttTl and lm} and𝒬:={𝒬l,vvVl and lm} such that\begin{array}[]{rcl}\vskip 6.0pt plus 2.0pt minus 2.0pt{\mathcal{P}}&:=&\bigcup\{{\mathcal{P}}_{l,t}\mid t\in T_{l}\,\mbox{ and }\,l\geq m\}\,\mbox{ and}\\ {\mathcal{Q}}&:=&\bigcup\{{\mathcal{Q}}_{l,v}\mid v\in V_{l}\,\mbox{ and }\,l\geq m\}\,\mbox{ such that}\end{array}
    𝒫l,t{p(x[K])Ul<m(p(l)A) and p(l)=x(l)+t}{\mathcal{P}}_{l,t}\subseteq\{p\sqsubseteq(\vec{x}\oplus[K])\cap U\mid\forall l^{\prime}<m\,(p(l^{\prime})\in A)\mbox{ and }p(l)=\vec{x}(l)+t\} (20)

    satisfying μK(𝒫l,t)=αm1/k\mu_{K}({\mathcal{P}}_{l,t})=\alpha^{m-1}/k for all lml\geq m and tTlt\in T_{l}, and

    𝒬l,v={qx[K]l<m(q(l)A) and q(l)=x(l)+v}{\mathcal{Q}}_{l,v}=\{q\sqsubseteq\vec{x}\oplus[K]\mid\forall l^{\prime}<m\,(q(l^{\prime})\in A)\,\mbox{ and }\,q(l)=\vec{x}(l)+v\} (21)

    satisfying μK(𝒬l,v)αm1\mu_{K}({\mathcal{Q}}_{l,v})\leq\alpha^{m-1} for all lml\geq m and vVlv\in V_{l}.

  2. L

    (m)2(α,η,N,A,S,K{}_{2}(m)(\alpha,\eta,N,A,S,K): There exist a set W0SW_{0}\subseteq S of min{K,1/D(1η)}\min\{K,\lfloor 1/D(1-\eta)\rfloor\}–consecutive integers where DD is defined in (18) and a collection of kk–a.p.’s ={rwwW0}{\mathcal{R}}=\{r_{w}\mid w\in W_{0}\} such that for each wW0w\in W_{0} we have rw(l)Ar_{w}(l)\in A for l<ml<m, rw(l)Sr_{w}(l)\in S for l>ml>m, and rw(m)=wr_{w}(m)=w.

Remark 5.2
  1. (a)

    𝐋2(m){\bf L}_{2}(m) is an internal statement in 𝕍2{\mathbb{V}}_{2}. Both 𝐋1(m){\bf L}_{1}(m) and 𝐋2(m){\bf L}_{2}(m) depend on KK. Since KK is fixed throughout whole proof, it, as a parameter, may be omitted in some expressions.

  2. (b)

    If H1H\gg 1 and T[H]T\subseteq[H] with μH(T)>1ϵ\mu_{H}(T)>1-\epsilon, then  TT contains 1/ϵ\lfloor 1/\epsilon\rfloor consecutive integers because otherwise we have μH(T)(1/ϵ1)/1/ϵ\mu_{H}(T)\leq(\lfloor 1/\epsilon\rfloor-1)/\lfloor 1/\epsilon\rfloor =11/1/ϵ11/(1/ϵ)=1ϵ=1-1/\lfloor 1/\epsilon\rfloor\leq 1-1/(1/\epsilon)=1-\epsilon.

  3. (c)

    The purpose of defining CKC_{K} is that if tCKt\in C_{K}, then the number of kk–a.p.’s px[K]p\sqsubseteq\vec{x}\oplus[K] with p(l)=x(l)+tp(l)=\vec{x}(l)+t is guaranteed to be at least K/(k1)K/(k-1).

  4. (d)

    It is not essential to require specific constant c=1/kc=1/k for μK(𝒫l,t)=cαm1\mu_{K}({\mathcal{P}}_{l,t})=c\alpha^{m-1} in 𝐋1(m){\bf L}_{1}(m). Just requiring that μK(𝒫l,t)cαm1\mu_{K}({\mathcal{P}}_{l,t})\geq c\alpha^{m-1} for some positive standard real cc is sufficient. We use more specific expression “μK(𝒫l,t)=αm1/k\mu_{K}({\mathcal{P}}_{l,t})=\alpha^{m-1}/k” for notational simplicity.

  5. (e)

    Some “bad” kk–a.p.’s in 𝒫{\mathcal{P}} in 𝐋1(m){\bf L}_{1}(m) will be thinned out so that {\mathcal{R}} in 𝐋2(m){\bf L}_{2}(m) can be constructed from 𝒫{\mathcal{P}}. The collection 𝒬{\mathcal{Q}} is only used to prevent 𝒫{\mathcal{P}} from being thinned out too much. See the proof of Lemma 5.3.

  6. (f)

    It is important to notice that in 𝐋1(m){\bf L}_{1}(m) the collection 𝒫l,t{\mathcal{P}}_{l,t} is a part of the collection at the right side of (20) while the collection 𝒬l,v{\mathcal{Q}}_{l,v} is equal to the collection at the right side of (21).

Lemma 5.3

𝐋1(m)(α,η,N,A,S,U){\bf L}_{1}(m)(\alpha,\eta,N,A,S,U) implies 𝐋2(m)(α,η,N,A,S){\bf L}_{2}(m)(\alpha,\eta,N,A,S) for any α,η,N,A,S,U\alpha,\eta,N,A,S,U satisfying the conditions of Lemma 5.1.

Proof Assume we have obtained the kk–a.p. xU\vec{x}\subseteq U with x[K][N]\vec{x}\oplus[K]\subseteq[N], sets TlCKT_{l}\subseteq C_{K} and Vl[K]V_{l}\subseteq[K] with μ|CK|(Tl)=1\mu_{|C_{K}|}(T_{l})=1 and μK(Vl)=1\mu_{K}(V_{l})=1, and collections of kk–a.p.’s 𝒫{\mathcal{P}} and 𝒬{\mathcal{Q}} as in 𝐋1(m){\bf L}_{1}(m).

Call a kk–a.p. p𝒫m:={𝒫m,ttTm}p\in{\mathcal{P}}_{m}:=\bigcup\{{\mathcal{P}}_{m,t}\mid t\in T_{m}\} good if p(l)S(x(l)+[K])p(l)\in S\cap(\vec{x}(l)+[K]) for lml\geq m and bad otherwise. Let 𝒫mg𝒫m{\mathcal{P}}_{m}^{g}\subseteq{\mathcal{P}}_{m} be the collection of all good kk–a.p.’s and 𝒫mb:=𝒫m𝒫mg{\mathcal{P}}^{b}_{m}:={\mathcal{P}}_{m}\setminus{\mathcal{P}}_{m}^{g} be the collection of all bad kk–a.p.’s. Let Tmg:={p(m)x(m)p𝒫mg}T_{m}^{g}:=\{p(m)-\vec{x}(m)\mid p\in{\mathcal{P}}_{m}^{g}\}. Then TmgTm(Sx(m))CKT_{m}^{g}\subseteq T_{m}\cap(S-\vec{x}(m))\cap C_{K}. We show that μ|CK|(Tmg)>1D(1η)\mu_{|C_{K}|}(T_{m}^{g})>1-D(1-\eta).

Let Q:={qx[K]q(l)A for l<m}Q:=\{q\sqsubseteq\vec{x}\oplus[K]\mid q(l^{\prime})\in A\,\mbox{ for }\,l^{\prime}<m\}. Notice that

𝒫mblm{qQq(l)S}{\mathcal{P}}_{m}^{b}\subseteq\bigcup_{l\geq m}\{q\in Q\mid q(l)\not\in S\}

and for each vVlv\in V_{l}, q𝒬l,vq\in{\mathcal{Q}}_{l,v} iff qQq\in Q and q(l)=x(l)+vq(l)=\vec{x}(l)+v.

Hence, |𝒫mb|l=mkw[K](Sx(l))|{qQq(l)=x(l)+w}|\displaystyle\mbox{Hence, }\,|{\mathcal{P}}_{m}^{b}|\leq\sum_{l=m}^{k}\sum_{w\in[K]\setminus(S-\vec{x}(l))}|\{q\in Q\mid q(l)=\vec{x}(l)+w\}|
l=mk(w[K]Vl|{qQq(l)=x(l)+w}|+vVl(Sx(l))|𝒬l,v|)\displaystyle\leq\sum_{l=m}^{k}\left(\sum_{w\in[K]\setminus V_{l}}|\{q\in Q\mid q(l)=\vec{x}(l)+w\}|+\sum_{v\in V_{l}\setminus(S-\vec{x}(l))}|{\mathcal{Q}}_{l,v}|\right)
Kl=mk(|[K]Vl|+|Vl(Sx(l))|αm1).\displaystyle\leq K\sum_{l=m}^{k}(|[K]\setminus V_{l}|+|V_{l}\setminus(S-\vec{x}(l))|\alpha^{m-1}).
So |𝒫mg|=|𝒫m||𝒫mb|tTm|𝒫m,t|Kl=mk(|[K]Vl|+|Vl(Sx(l))|αm1).\mbox{So }\,\,|{\mathcal{P}}_{m}^{g}|=|{\mathcal{P}}_{m}|-|{\mathcal{P}}_{m}^{b}|\geq\sum_{t\in T_{m}}|{\mathcal{P}}_{m,t}|-K\sum_{l=m}^{k}(|[K]\setminus V_{l}|+|V_{l}\setminus(S-\vec{x}(l))|\alpha^{m-1}).

Notice that μK([K]Vl)=0\mu_{K}([K]\setminus V_{l})=0. Hence we have |[K]Vl|/|CK|0|[K]\setminus V_{l}|/|C_{K}|\approx 0 and

μ|CK|(Tmg)αm1k=st(1|CK|tTmg1K|𝒫m,t|)st(1|CK|K|𝒫mg|)\displaystyle\mu_{|C_{K}|}(T_{m}^{g})\cdot\frac{\alpha^{m-1}}{k}=st\left(\frac{1}{|C_{K}|}\sum_{t\in T_{m}^{g}}\frac{1}{K}|{\mathcal{P}}_{m,t}|\right)\geq st\left(\frac{1}{|C_{K}|K}|{\mathcal{P}}_{m}^{g}|\right)
st(1|CK|tTm1K|𝒫m,t|1|CK|l=mk(|Vl(Sx(l))|αm1))\displaystyle\geq st\left(\frac{1}{|C_{K}|}\sum_{t\in T_{m}}\frac{1}{K}|{\mathcal{P}}_{m,t}|-\frac{1}{|C_{K}|}\sum_{l=m}^{k}(|V_{l}\setminus(S-\vec{x}(l))|\alpha^{m-1})\right)
μ|CK|(Tm)αm1k(2k+1)k(1η)αm1\displaystyle\geq\mu_{|C_{K}|}(T_{m})\cdot\frac{\alpha^{m-1}}{k}-(2k+1)k(1-\eta)\cdot\alpha^{m-1}
=(1k(2k+1)k(1η))αm1,\displaystyle=\left(\frac{1}{k}-(2k+1)k(1-\eta)\right)\cdot\alpha^{m-1},

which implies μ|CK|(Tmg)1(2k+1)k2(1η)>1D(1η)\displaystyle\mu_{|C_{K}|}(T_{m}^{g})\geq 1-(2k+1)k^{2}(1-\eta)>1-D(1-\eta). Recall that TmgCKT^{g}_{m}\subseteq C_{K}. Hence x(m)+Tmg\vec{x}(m)+T^{g}_{m} contains a set W0W_{0} of 1/D(1η)\lfloor 1/D(1-\eta)\rfloor consecutive integers. So, 𝐋2(m){\bf L}_{2}(m) is proven if we let :={rwwW0}{\mathcal{R}}:=\{r_{w}\mid w\in W_{0}\} where rwr_{w} is one of the kk–a.p.’s in 𝒫mg{\mathcal{P}}_{m}^{g} such that rw(m)=x(m)+wr_{w}(m)=\vec{x}(m)+w. \blacksquare


The idea of the proof of Lemma 5.3 is due to Szemerédi. See [9].


Proof of Lemma 5.1 We prove 𝐋(m){\bf L}(m) by induction on mm. By Lemmas 5.3 it suffices to prove 𝐋1(m){\bf L}_{1}(m).

For 𝐋(1){\bf L}(1), given any α>0\alpha>0, η>η0\eta>\eta_{0}, N21N\in{\mathbb{N}}_{2}\setminus{\mathbb{N}}_{1}, AA, SS, and UU satisfying (19), by Lemma 3.5 (b) we can find a kk–a.p. x[N]\vec{x}\subseteq[N] such that (n0)ξ(x(l),α,η,A,S,U,K,n)(\forall n\in{\mathbb{N}}_{0})\,\xi(\vec{x}(l),\alpha,\eta,A,S,U,K,n) is true for l[k]l\in[k], where ξ\xi is defined in (11). For each l[k]l\in[k] let Tl=CK(Ux(l))T_{l}=C_{K}\cap(U-\vec{x}(l)) and Vl=[K]V_{l}=[K]. For each l[k]l\in[k], tTlt\in T_{l}, and vVlv\in V_{l} let

𝒫l,t:={p(x[K])Up(l)=x(l)+t}𝒬l,v:={q(x[K])q(l)=x(l)+v}.\begin{array}[]{rcl}\vskip 6.0pt plus 2.0pt minus 2.0pt{\mathcal{P}}_{l,t}&:=&\{p\sqsubseteq(\vec{x}\oplus[K])\cap U\mid p(l)=\vec{x}(l)+t\}\\ {\mathcal{Q}}_{l,v}&:=&\{q\sqsubseteq(\vec{x}\oplus[K])\mid q(l)=\vec{x}(l)+v\}.\end{array}

Clearly, we have μK(𝒫l,t)1/(k1)>1/k\mu_{K}({\mathcal{P}}_{l,t})\geq 1/(k-1)>1/k. By some pruning we can assume that μK(𝒫l,t)=1/k\mu_{K}({\mathcal{P}}_{l,t})=1/k. It is trivial that μK(𝒬l,v)1\mu_{K}({\mathcal{Q}}_{l,v})\leq 1 and q𝒬l,vq\in{\mathcal{Q}}_{l,v} iff q(l)=x(l)+vq(l)=\vec{x}(l)+v for each qx[K]q\sqsubseteq\vec{x}\oplus[K]. This completes the proof of 𝐋1(1)(α,η,N,A,S,U){\bf L}_{1}(1)(\alpha,\eta,N,A,S,U). 𝐋2(1)(α,η,N,A,S){\bf L}_{2}(1)(\alpha,\eta,N,A,S) follows from Lemma 5.3.

Assume 𝐋(m1){\bf L}(m-1) is true for some 2mk2\leq m\leq k.

We now prove 𝐋(m){\bf L}(m). Given any α>0\alpha>0 and η>η0\eta>\eta_{0}, fix N21N\in{\mathbb{N}}_{2}\setminus{\mathbb{N}}_{1}, U[N]U\subseteq[N], and AS[N]A\subseteq S\subseteq[N] satisfying (19). For each n10n\in{\mathbb{N}}_{1}\setminus{\mathbb{N}}_{0}, by Lemma 3.5 (b), there is an hn>nh_{n}>n in 1{\mathbb{N}}_{1} and Gn,hn[N]G_{n,h_{n}}\subseteq[N] defined in (12) such that dn:=δNhn(Gn,hn)>11/nd_{n}:=\delta_{N-h_{n}}(G_{n,h_{n}})>1-1/n. Notice that dn1μNhn1(Gn,hn)>η0d_{n}\approx_{1}\mu^{1}_{N-h_{n}}(G_{n,h_{n}})>\eta_{0} because n1n\gg 1 and μNhn(Gn,hn))=1\mu_{N-h_{n}}(G_{n,h_{n}}))=1. Let ηn1:=μNhn1(Gn,hn)\eta^{1}_{n}:=\mu^{1}_{N-h_{n}}(G_{n,h_{n}}) and fix an n10n\in{\mathbb{N}}_{1}\setminus{\mathbb{N}}_{0}.

Claim 1 The following internal statement θ(n,A,N)\theta(n,A,N) is true:

W[N](W is an a.p.|W|min{K,1/2D(1dn)}={rwwW}\exists W\subseteq[N]\,\exists{\mathcal{R}}\,(W\,\mbox{ is an a.p.}\,\wedge|W|\geq\min\{K,\lfloor 1/2D(1-d_{n})\rfloor\}\,\wedge\,{\mathcal{R}}=\{r_{w}\mid w\in W\} is a collection of kk–a.p.’s such that

wW((lm)(rw(l)Gn,hn)rw(m1)=w(l,lm2)\forall w\in W\,((\forall l\geq m)\,(r_{w}(l)\in\,G_{n,h_{n}})\,\wedge\,r_{w}(m-1)=w\,\wedge\,(\forall l,l^{\prime}\leq m-2)
((A(rw(l)+[hn]))rw(l)=(A(rw(l)+[hn]))rw(l))).((A\cap(r_{w}(l)+[h_{n}]))-r_{w}(l)=(A\cap(r_{w}(l^{\prime})+[h_{n}]))-r_{w}(l^{\prime}))).

Proof of Claim 1 Working in 𝕍2{\mathbb{V}}_{2} by considering 𝕍1{\mathbb{V}}_{1} as the standard universe, we can find P[N]P\subseteq[N] with |P|21|P|\in{\mathbb{N}}_{2}\setminus{\mathbb{N}}_{1} by Lemma 3.3 and Part 2 of Property 2.7 such that

SD1(Gn,hn)=μ|P|1(PGn,hn)=SD1(Gn,hnP)=ηn1.S\!D^{1}(G_{n,h_{n}})=\mu^{1}_{|P|}(P\cap G_{n,h_{n}})=S\!D^{1}(G_{n,h_{n}}\cap P)=\eta^{1}_{n}.

For each xPGn,hnx\in P\cap G_{n,h_{n}} let τx=((x+[hn])A)x\tau_{x}=((x+[h_{n}])\cap A)-x. Since there are at most 2hn12^{h_{n}}\in{\mathbb{N}}_{1} different τx\tau_{x}’s and |P|2hn|P|\gg 2^{h_{n}}, we can find one, say, τn[hn]\tau_{n}\subseteq[h_{n}] such that the set

Bn:={xPGn,hnτx=τn}B_{n}:=\{x\in P\cap G_{n,h_{n}}\mid\tau_{x}=\tau_{n}\}

satisfies μ|P|1(Bn)ηn1/2hn>0\mu^{1}_{|P|}(B_{n})\geq\eta^{1}_{n}/2^{h_{n}}>0. Notice that μ|P|0(Bn)\mu^{0}_{|P|}(B_{n}) could be 0.

Let PPP^{\prime}\subseteq P with |P|=N21|P^{\prime}|=N^{\prime}\in{\mathbb{N}}_{2}\setminus{\mathbb{N}}_{1} be such that μN1(Gn,hnP)=ηn1\mu^{1}_{N^{\prime}}(G_{n,h_{n}}\cap P^{\prime})=\eta^{1}_{n} and

βn1:=μN1(BnP)=SDGn,hnP1(BnP)=SDGn,hnP1(BnP)μ|P|1(Bn)>0\begin{array}[]{rcl}\vspace{0.1in}\beta^{1}_{n}&:=&\mu^{1}_{N^{\prime}}(B_{n}\cap P^{\prime})=S\!D^{1}_{G_{n,h_{n}}\cap P}(B_{n}\cap P)\\ &=&S\!D^{1}_{G_{n,h_{n}}\cap P^{\prime}}(B_{n}\cap P^{\prime})\geq\mu^{1}_{|P|}(B_{n})>0\end{array}

by Part 4 of Lemma 3.3 and Part 2 of Property 2.7. Let dd be the common difference of the a.p. PP^{\prime} and φ:P[N]\varphi:P^{\prime}\to[N^{\prime}] be the order-preserving bijection, i.e.,

φ(x):=1+(xminP)/d.\varphi(x):=1+(x-\min P^{\prime})/d.

Let B:=φ[BnP]B^{\prime}:=\varphi[B_{n}\cap P^{\prime}] and S:=φ[Gn,hnP]S^{\prime}:=\varphi[G_{n,h_{n}}\cap P^{\prime}]. We have that B,S,NB^{\prime},S^{\prime},N^{\prime} and βn1,ηn1\beta^{1}_{n},\eta^{1}_{n} in the place of A,S,NA,S,N and α,η\alpha,\eta satisfy the 𝕍1{\mathbb{V}}_{1}–version of (19) with μ\mu, SDS\!D  and SDSS\!D_{S} being replaced by μ1\mu^{1}, SD1S\!D^{1},  and SDS1S\!D^{1}_{S^{\prime}}.

Let N′′=i2(N)N^{\prime\prime}=i_{2}(N^{\prime}), B′′=i2(B)B^{\prime\prime}=i_{2}(B^{\prime}), and S′′=i2(S)S^{\prime\prime}=i_{2}(S^{\prime}) where i2i_{2} is in Part 3 of Property 2.7. Recall that i2𝕍1i_{2}\!\upharpoonright\!{\mathbb{V}}_{1} is an identity map. Since N21N^{\prime}\in{\mathbb{N}}_{2}\setminus{\mathbb{N}}_{1}, we have N′′32N^{\prime\prime}\in{\mathbb{N}}_{3}\setminus{\mathbb{N}}_{2}. Notice also that μN′′1(S′′)=SD1(S′′)=ηn1\mu^{1}_{N^{\prime\prime}}(S^{\prime\prime})=S\!D^{1}(S^{\prime\prime})=\eta^{1}_{n} and μN′′1(B′′)=SDS′′1(B′′)=βn1\mu^{1}_{N^{\prime\prime}}(B^{\prime\prime})=S\!D^{1}_{S^{\prime\prime}}(B^{\prime\prime})=\beta^{1}_{n}. By the induction hypothesis that 𝐋(m1){\bf L}(m-1) is true we have

(𝕍2;0,1)α,η0N21A,S[N]\displaystyle({\mathbb{V}}_{2};{\mathbb{R}}_{0},{\mathbb{R}}_{1})\models\forall\alpha,\eta\in{\mathbb{R}}_{0}\,\forall N\in{\mathbb{N}}_{2}\setminus{\mathbb{N}}_{1}\,\forall A,S\subseteq[N]
(α>0η>η0ASμN(S)=SD(S)=ημN(A)=SDS(A)\displaystyle(\alpha>0\wedge\eta>\eta_{0}\wedge A\subseteq S\wedge\mu_{N}(S)=S\!D(S)=\eta\wedge\mu_{N}(A)=S\!D_{S}(A)
𝐋2(m1)(α,η,N,A,S)).\displaystyle\to{\bf L}_{2}(m-1)(\alpha,\eta,N,A,S)).

Since (𝕍2;0,1)({\mathbb{V}}_{2};{\mathbb{R}}_{0},{\mathbb{R}}_{1}) and (𝕍3;1,2)({\mathbb{V}}_{3};{\mathbb{R}}_{1},{\mathbb{R}}_{2}) are elementarily equivalent by Part 2 of Property 2.7 via ii_{*}, we have, by universal instantiation, that

(𝕍3;1,2)𝐋2(m1)(βn1,ηn1,N′′,B′′,S′′).({\mathbb{V}}_{3};{\mathbb{R}}_{1},{\mathbb{R}}_{2})\models{\bf L}_{2}(m-1)(\beta^{1}_{n},\eta^{1}_{n},N^{\prime\prime},B^{\prime\prime},S^{\prime\prime}). (23)

Notice that the right side above no longer depends on 1{\mathbb{R}}_{1} or 2{\mathbb{R}}_{2}. So, we have

𝕍3𝐋2(m1)(i2(βn1),i2(ηn1),i2(N),i2(B),i2(S)){\mathbb{V}}_{3}\models{\bf L}_{2}(m-1)(i_{2}(\beta^{1}_{n}),i_{2}(\eta^{1}_{n}),i_{2}(N^{\prime}),i_{2}(B^{\prime}),i_{2}(S^{\prime})) (24)

because i2(βn1)=βn1i_{2}(\beta^{1}_{n})=\beta^{1}_{n} and i2(ηn1)=ηn1i_{2}(\eta^{1}_{n})=\eta^{1}_{n}. Since i2i_{2} is a bounded elementary embedding, we have

𝕍2𝐋2(m1)(βn1,ηn1,N,B,S),{\mathbb{V}}_{2}\models{\bf L}_{2}(m-1)(\beta^{1}_{n},\eta^{1}_{n},N^{\prime},B^{\prime},S^{\prime}),

which means that there is a set W[N]W^{\prime}\subseteq[N^{\prime}] of min{K,1/D(1ηn1)}\min\{K,\lfloor 1/D(1-\eta^{1}_{n})\rfloor\}–consecutive integers and a collection of kk–a.p.’s ={rwwW}{\mathcal{R}}^{\prime}=\{r^{\prime}_{w}\mid w\in W^{\prime}\} such that for every wWw\in W^{\prime} we have rw(l)Br^{\prime}_{w}(l)\in B^{\prime} for l<m1l<m-1, rw(m1)=wr^{\prime}_{w}(m-1)=w, and rw(l)Sr^{\prime}_{w}(l)\in S^{\prime} for lml\geq m. Notice that φ1[[N]][N]\varphi^{-1}[[N^{\prime}]]\subseteq[N]. Let W=φ1[W]W=\varphi^{-1}[W^{\prime}] and ={rwwW}{\mathcal{R}}=\{r_{w}\mid w\in W\}, where rw=φ1[rφ(w)]r_{w}=\varphi^{-1}[r^{\prime}_{\varphi(w)}], such that for each wWw\in W we have rw(l)φ1[B]Bnr_{w}(l)\in\varphi^{-1}[B^{\prime}]\subseteq B_{n} for l<m1l<m-1, rw(m1)=wr_{w}(m-1)=w, and rw(l)φ1[S]Gn,hnr_{w}(l)\in\varphi^{-1}[S^{\prime}]\subseteq G_{n,h_{n}} for lml\geq m. If ηn1=1\eta^{1}_{n}=1, then |W|K|W|\geq K. If ηn1<1\eta^{1}_{n}<1, then 2(1dn)>1ηn12(1-d_{n})>1-\eta^{1}_{n}. Hence |W|min{K,1/2D(1dn)}|W|\geq\min\{K,\lfloor 1/2D(1-d_{n})\rfloor\}. \blacksquare (Claim 1)

The following claim follows from Claim 1 by Proposition 2.15.

Claim 2 There exists a J21J\in{\mathbb{N}}_{2}\setminus{\mathbb{N}}_{1} such that the θ(J,A,N)\theta(J,A,N) is true, i.e., W[N](W is an a.p.|W|min{K,1/2D(1dJ)}={rwwW}\exists W\subseteq[N]\,\exists{\mathcal{R}}\,(W\,\mbox{ is an a.p.}\,\wedge|W|\geq\min\{K,\lfloor 1/2D(1-d_{J})\rfloor\}\,\wedge\,{\mathcal{R}}=\{r_{w}\mid w\in W\} is a collection of kk–a.p.’s such that wW((lm)(rw(l)GJ,hJ)\forall w\in W\,((\forall l\geq m)\,(r_{w}(l)\in\,G_{J,h_{J}}), rw(m1)=wr_{w}(m-1)=w, and (l,lm2)((A(rw(l)+[hJ]))rw(l)=(A(rw(l)+[hJ]))rw(l))))(\forall l,l^{\prime}\leq m-2)\,((A\cap(r_{w}(l)+[h_{J}]))-r_{w}(l)=(A\cap(r_{w}(l^{\prime})+[h_{J}]))-r_{w}(l^{\prime})))).

For notational convenience let WH:=WW_{H}:=W and H:={\mathcal{R}}_{H}:={\mathcal{R}} be obtained in Claim 2 and rename H:=hJH:=h_{J}, SH:=GJ,hJS_{H}:=G_{J,h_{J}}, τH:=(A(rw(l)+[hJ]))rw(l)\tau_{H}:=(A\cap(r_{w}(l)+[h_{J}]))-r_{w}(l) for some (or any) wWHw\in W_{H} and l<m1l<m-1. Let {ws1s|WH|}\{w_{s}\mid 1\leq s\leq|W_{H}|\} be the increasing enumeration of WHW_{H}. Notice that H21H\in{\mathbb{N}}_{2}\setminus{\mathbb{N}}_{1}. We now go back to consider 𝕍0{\mathbb{V}}_{0} as our standard universe. Notice that μNH(SH)=1\mu_{N-H}(S_{H})=1, |WH|1|W_{H}|\gg 1, and (n0)ξ(x,α,η,A,S,U,H,n)(\forall n\in{\mathbb{N}}_{0})\,\xi(x,\alpha,\eta,A,S,U,H,n) is true for every xSHx\in S_{H} where ξ\xi is defined in (11).

Claim 3 For each s0s\in{\mathbb{N}}_{0} we can find an internal Us[H]U_{s}\subseteq[H] with μH(Us)=1\mu_{H}(U_{s})=1 such that for each yUsy\in U_{s} and each l[k]l\in[k], rws(l)+yUr_{w_{s}}(l)+y\in U and (n0)ξ(rws(l)+y,α,η,A,S,U,K,n)(\forall n\in{\mathbb{N}}_{0})\,\xi(r_{w_{s}}(l)+y,\alpha,\eta,A,S,U,K,n) is true.

Proof of Claim 3 For each l[k]l\in[k] we have ξ(rws(l),α,η,A,S,U,H,n)\xi(r_{w_{s}}(l),\alpha,\eta,A,S,U,H,n) is true because rws(l)SHr_{w_{s}}(l)\in S_{H}. By Lemma 3.5 (a), we can find a set Glrws(l)+[H]G_{l}\subseteq r_{w_{s}}(l)+[H] with μH(Gl)=1\mu_{H}(G_{l})=1 such that
ξ(rws(l)+y,α,η,A,S,U,K,n)\xi(r_{w_{s}}(l)+y,\alpha,\eta,A,S,U,K,n) is true for every rws(l)+yGlr_{w_{s}}(l)+y\in G_{l}. Set

Us:=l=1k((UGl)rws(l)).U_{s}:=\bigcap_{l=1}^{k}((U\cap G_{l})-r_{w_{s}}(l)).

Then we have Us[H]U_{s}\subseteq[H] and μH(Us)=1\mu_{H}(U_{s})=1. \blacksquare (Claim 3)

Notice that δH(i=1sUi)>11/s\delta_{H}(\bigcap_{i=1}^{s}U_{i})>1-1/s. By Proposition 2.15 we can find 1I|WH|1\ll I\leq|W_{H}| and

U:={Us1sI}U^{\prime}:=\bigcap\{U_{s}\mid 1\leq s\leq I\}

such that δH(U)>11/I\delta_{H}(U^{\prime})>1-1/I. Hence μH(U)=1\mu_{H}(U^{\prime})=1. Applying the induction hypothesis for 𝐋1(m1)(α,1,H,τH,[H],U){\bf L}_{1}(m-1)(\alpha,1,H,\tau_{H},[H],U^{\prime}), we obtain a kk–a.p. yU\vec{y}\subseteq U^{\prime} with y[K][H]\vec{y}\oplus[K]\subseteq[H], TlCKUT^{\prime}_{l}\subseteq C_{K}\cap U^{\prime} with μ|CK|(Tl)=1\mu_{|C_{K}|}(T^{\prime}_{l})=1 and Vl[K]V^{\prime}_{l}\subseteq[K] with μK(Vl)=1\mu_{K}(V^{\prime}_{l})=1 for each lm1l\geq m-1, and collections of kk–a.p.’s

𝒫={𝒫l,ttTl and lm1} and𝒬={𝒬l,vvVl and lm1}\begin{array}[]{rcl}\vspace{0.1in}{\mathcal{P}}^{\prime}&=&\bigcup\{{\mathcal{P}}^{\prime}_{l,t}\mid t\in T^{\prime}_{l}\,\mbox{ and }\,l\geq m-1\}\,\mbox{ and}\\ {\mathcal{Q}}^{\prime}&=&\bigcup\{{\mathcal{Q}}^{\prime}_{l,v}\mid v\in V^{\prime}_{l}\,\mbox{ and }\,l\geq m-1\}\end{array}

such that (i) for each lm1l\geq m-1 and tTlt\in T^{\prime}_{l} we have μK(𝒫l,t)=αm2/k\mu_{K}({\mathcal{P}}^{\prime}_{l,t})=\alpha^{m-2}/k and for each p𝒫l,tp\in{\mathcal{P}}^{\prime}_{l,t} we have p(y[K])Up\sqsubseteq(\vec{y}\oplus[K])\cap U^{\prime}, p(l)τHp(l^{\prime})\in\tau_{H} for l<m1l^{\prime}<m-1, p(l)=y(l)+tp(l)=\vec{y}(l)+t, and (ii) for each lm1l\geq m-1 and vVlv\in V^{\prime}_{l} we have μK(𝒬l,v)αm2\mu_{K}({\mathcal{Q}}^{\prime}_{l,v})\leq\alpha^{m-2}, and for each qy[K]q\sqsubseteq\vec{y}\oplus[K] we have q𝒬l,vq\in{\mathcal{Q}}^{\prime}_{l,v} iff q(l)τHq(l^{\prime})\in\tau_{H} for every l<m1l^{\prime}<m-1 and q(l)=y(l)+vq(l)=\vec{y}(l)+v. For each lml\geq m, tTlt\in T_{l}, and vVlv\in V_{l} let

El,t:={p(m1)p𝒫l,t} and Fl,v:={q(m1)q𝒬l,v}.E_{l,t}:=\{p(m-1)\mid p\in{\mathcal{P}}^{\prime}_{l,t}\}\,\mbox{ and }\,F_{l,v}:=\{q(m-1)\mid q\in{\mathcal{Q}}^{\prime}_{l,v}\}.

Then El,t,Fl,vy(m1)+[K]E_{l,t},F_{l,v}\subseteq\vec{y}(m-1)+[K], μK(El,t)=μK(𝒫l,t)=αm2/k\mu_{K}(E_{l,t})=\mu_{K}({\mathcal{P}}^{\prime}_{l,t})=\alpha^{m-2}/k, and μK(Fl,v)=μK(𝒬l,v)αm2\mu_{K}(F_{l,v})=\mu_{K}({\mathcal{Q}}^{\prime}_{l,v})\leq\alpha^{m-2}. Since yU\vec{y}\subseteq U^{\prime} we have that for each l[k]l\in[k], (n0)ξ(rws(l)+y(l),α,η,A,S,U,K,n)(\forall n\in{\mathbb{N}}_{0})\,\xi(r_{w_{s}}(l)+\vec{y}(l),\alpha,\eta,A,S,U,K,n) is true.

Applying Part (iii) of Lemma 4.2 with R:={ws+y(m1)1sI}R:=\{w_{s}+\vec{y}(m-1)\mid 1\leq s\leq I\} and HH being replaced by KK we can find s0[I]s_{0}\in[I], TlTlT_{l}\subseteq T^{\prime}_{l} with μ|CK|(Tl)=1\mu_{|C_{K}|}(T_{l})=1 and VlVlV_{l}\subseteq V^{\prime}_{l} with μK(Vl)=1\mu_{K}(V_{l})=1 for each lml\geq m such that for each tTlt\in T_{l} and vVlv\in V_{l} we have

μK((ws0+El,t)((ws0+y(m1)+[K])A))=αμK(El,t)=α(αm2/k)=αm1/k and\begin{array}[]{l}\vspace{0.1in}\mu_{K}((w_{s_{0}}+E_{l,t})\cap((w_{s_{0}}+\vec{y}(m-1)+[K])\cap A))\\ \qquad=\alpha\mu_{K}(E_{l,t})=\alpha(\alpha^{m-2}/k)=\alpha^{m-1}/k\,\mbox{ and}\end{array} (25)
μK((ws0+Fl,v)((ws0+y(m1)+[K])A))=αμK(Fl,t)ααm2=αm1.\begin{array}[]{l}\vspace{0.1in}\mu_{K}((w_{s_{0}}+F_{l,v})\cap((w_{s_{0}}+\vec{y}(m-1)+[K])\cap A))\\ \qquad\qquad=\alpha\mu_{K}(F_{l,t})\leq\alpha\!\cdot\!\alpha^{m-2}=\alpha^{m-1}.\end{array} (26)

Let x:=rws0y\vec{x}:=r_{w_{s_{0}}}\oplus\vec{y}. Clearly, we have x[K][N]\vec{x}\oplus[K]\subseteq[N]. We also have that xU\vec{x}\subseteq U, μK((x(l)+[K])S)=η\mu_{K}((\vec{x}(l)+[K])\cap S)=\eta, and μK((x(l)+[K])A)=α\mu_{K}((\vec{x}(l)+[K])\cap A)=\alpha because rws0SHr_{w_{s_{0}}}\subseteq S_{H} and yUUs0\vec{y}\subseteq U^{\prime}\subseteq U_{s_{0}}. For each lml\geq m, tTlt\in T_{l}, and vVlv\in V_{l} let

𝒫l,t:={rws0pp𝒫l,t and\displaystyle{\mathcal{P}}_{l,t}:=\{r_{w_{s_{0}}}\oplus p\mid p\in{\mathcal{P}}^{\prime}_{l,t}\,\mbox{ and}
p(m1)El,t(((ws0+y(m1)+[K])A)ws0)},\displaystyle\,p(m-1)\in E_{l,t}\cap(((w_{s_{0}}+\vec{y}(m-1)+[K])\cap A)-w_{s_{0}})\},
𝒬l,v:={rws0qq𝒬l,t and\displaystyle{\mathcal{Q}}_{l,v}:=\{r_{w_{s_{0}}}\oplus q\mid q\in{\mathcal{Q}}^{\prime}_{l,t}\,\mbox{ and}
q(m1)Fl,v(((ws0+y(m1)+[K])A)ws0)}.\displaystyle\,q(m-1)\in F_{l,v}\cap(((w_{s_{0}}+\vec{y}(m-1)+[K])\cap A)-w_{s_{0}})\}.

Then μK(𝒫l,t)=αm1/k\mu_{K}({\mathcal{P}}_{l,t})=\alpha^{m-1}/k by (25). If q¯x[K]\bar{q}\sqsubseteq\vec{x}\oplus[K], then there is a qy[K]q\sqsubseteq\vec{y}\oplus[K] such that q¯=rws0q\bar{q}=r_{w_{s_{0}}}\oplus q. If q¯(l)A\bar{q}(l^{\prime})\in A for l<ml^{\prime}<m and vVlv\in V_{l} for some lml\geq m such that q¯(l)=x(l)+v\bar{q}(l)=\vec{x}(l)+v, then q(l)τHq(l^{\prime})\in\tau_{H} for l<m1l^{\prime}<m-1, vVlv\in V^{\prime}_{l}, and q(l)=y(l)+vq(l)=\vec{y}(l)+v, which imply q𝒬l,vq\in{\mathcal{Q}}^{\prime}_{l,v} by induction hypothesis. Hence we have q(m1)Fl,vq(m-1)\in F_{l,v}. Clearly, q¯(m1)=ws0+q(m1)A\bar{q}(m-1)=w_{s_{0}}+q(m-1)\in A implies q(m1)Fl,v(((ws0+y(m1)+[K])A)ws0)q(m-1)\in F_{l,v}\cap(((w_{s_{0}}+\vec{y}(m-1)+[K])\cap A)-w_{s_{0}}). Thus we have q¯𝒬l,v\bar{q}\in{\mathcal{Q}}_{l,v}. Clearly, μK(𝒬l,v)αm1\mu_{K}({\mathcal{Q}}_{l,v})\leq\alpha^{m-1} by (26).

Summarizing the argument above we have that for each rws0p𝒫l,tr_{w_{s_{0}}}\oplus p\in{\mathcal{P}}_{l,t}

  • rws0(l)+p(l)rws0(l)+τHAr_{w_{s_{0}}}(l^{\prime})+p(l^{\prime})\in r_{w_{s_{0}}}(l^{\prime})+\tau_{H}\subseteq A for l<m1l^{\prime}<m-1 because rws0(l)BHr_{w_{s_{0}}}(l^{\prime})\in B_{H},

  • rws0(m1)+p(m1)=ws0+p(m1)r_{w_{s_{0}}}(m-1)+p(m-1)=w_{s_{0}}+p(m-1)

    (ws0+El,t)(ws0+y(m1)+[K])AA\in(w_{s_{0}}+E_{l,t})\cap(w_{s_{0}}+\vec{y}(m-1)+[K])\cap A\subseteq A,

  • rws0(l)+p(l)(x(l)+[K])UUr_{w_{s_{0}}}(l^{\prime})+p(l^{\prime})\in(\vec{x}(l^{\prime})+[K])\cap U\subseteq U for lml^{\prime}\geq m because of pUp\subseteq U^{\prime},

  • rws0(l)+p(l)=rws0(l)+y(l)+t=x(l)+tr_{w_{s_{0}}}(l)+p(l)=r_{w_{s_{0}}}(l)+\vec{y}(l)+t=\vec{x}(l)+t.

For each q¯x[K]\bar{q}\sqsubseteq\vec{x}\oplus[K], q¯𝒬l,v\bar{q}\in{\mathcal{Q}}_{l,v} iff there is a qy[K]q\sqsubseteq\vec{y}\oplus[K] with q¯=rws0q\bar{q}=r_{w_{s_{0}}}\oplus q such that

  • rws0(l)+q(l)rws0(l)+τHAr_{w_{s_{0}}}(l^{\prime})+q(l^{\prime})\in r_{w_{s_{0}}}(l^{\prime})+\tau_{H}\subseteq A for l<m1l^{\prime}<m-1 because rws0(l)BHr_{w_{s_{0}}}(l^{\prime})\in B_{H},

  • rws0(m1)+q(m1)=ws0+q(m1)Ar_{w_{s_{0}}}(m-1)+q(m-1)=w_{s_{0}}+q(m-1)\in A which is equivalent to

    ws0+q(m1)(ws0+Fl,v)(ws0+y(m1)+[K])AAw_{s_{0}}+q(m-1)\in(w_{s_{0}}+F_{l,v})\cap(w_{s_{0}}+\vec{y}(m-1)+[K])\cap A\subseteq A,

  • rws0(l)+q(l)=rws0(l)+y(l)+v=x(l)+vr_{w_{s_{0}}}(l)+q(l)=r_{w_{s_{0}}}(l)+\vec{y}(l)+v=\vec{x}(l)+v.

This completes the proof of 𝐋1(m)(α,η,N,A,S,U){\bf L}_{1}(m)(\alpha,\eta,N,A,S,U) as well as 𝐋(m){\bf L}(m) by Lemma 5.3. \blacksquare


Summary of the ideas used in the proof of Lemma 5.1:  We want to use 𝐋2(m1){\bf L}_{2}(m-1) to create a sequence H{\mathcal{R}}_{H} of kk blocks rw[H]r_{w}\oplus[H] of size HH that are in arithmetic progression such that (a) the set AA in the first m2m-2 blocks rw(l)+[H]r_{w}(l)+[H] for lm2l\leq m-2 are identical copies of τH\tau_{H} in [H][H], (b) the initial points {rw(m1)wWH}\{r_{w}(m-1)\mid w\in W_{H}\} of all (m1)(m-1)-st blocks form an a.p. of infinite length which should be used when applying Part (iii) of Lemma 4.2, (c) the initial points of the rest of the blocks satisfy an appropriate version of (11) for some infinite n=Jn=J, i.e., rwSHr_{w}\subseteq S_{H}. Then we work inside [H][H], using L1(m1)L_{1}(m-1) to create collections of kk–a.p.’s 𝒫{\mathcal{P}}^{\prime} and 𝒬{\mathcal{Q}}^{\prime} in [H][H] instead of [N][N]. For applying Part (iii) of Lemma 4.2 we want to make sure, if we can, that the (m1)(m-1)-st terms of all p𝒫l,tp\in{\mathcal{P}}^{\prime}_{l,t} form a set of positive measure and the (m1)(m-1)-st terms of all q𝒬l,vq\in{\mathcal{Q}}^{\prime}_{l,v} form a set of positive measure. Then mixing one rwr_{w} for some wWHw\in W_{H} with 𝒫{\mathcal{P}}^{\prime} and 𝒬{\mathcal{Q}}^{\prime} at (m1)(m-1)-st terms yields 𝒫{\mathcal{P}} and 𝒬{\mathcal{Q}} validating 𝐋1(m){\bf L}_{1}(m).

Unfortunately, using 𝐋1(m1){\bf L}_{1}(m-1) to create collections 𝒫{\mathcal{P}}^{\prime} and 𝒬{\mathcal{Q}}^{\prime} of kk–a.p.’s in [H][H] cannot guarantee that the set El,tE_{l,t} of the (m1)(m-1)-st terms of all p𝒫l,tp\in{\mathcal{P}}^{\prime}_{l,t} and the set Fl,vF_{l,v} of the (m1)(m-1)-st terms of all q𝒬l,vq\in{\mathcal{Q}}^{\prime}_{l,v} have positive measures in [H][H] (the positive measures can be guaranteed when k4k\leq 4 but not for k>4k>4). So, instead of getting the sets El,tE_{l,t} and Fl,vF_{l,v} to have positive measures in [H][H] we make sure that the set El,tE_{l,t} and Fl,vF_{l,v} have positive measures in some subinterval y(m1)+[K]\vec{y}(m-1)+[K] of length KK in [H][H], where KK is fixed and could be much smaller than HH. This is achieved by requiring yU\vec{y}\subseteq U^{\prime}. When we use the mixing lemma we want to make sure that AA in all of these relevant intervals (rwsy)(m1)+[K](r_{w_{s}}\oplus\vec{y})(m-1)+[K] of length KK has measure α\alpha. This requirement is again achieved by requiring rwsSHr_{w_{s}}\subseteq S_{H} and yU\vec{y}\subseteq U^{\prime} after shrinking WHW_{H} to its initial segment {wss[I]}\{w_{s}\mid s\in[I]\}.

Since we want KK to be significantly smaller than NN we assume that NN and KK is at least one universe apart. Hence NN must be at least in 21{\mathbb{N}}_{2}\setminus{\mathbb{N}}_{1}. If we want to apply 𝐋1(m1){\bf L}_{1}(m-1) for HH instead of NN, then HH must also be at least in 21{\mathbb{N}}_{2}\setminus{\mathbb{N}}_{1}. If we want BHB_{H} to have a 𝕍2{\mathbb{V}}_{2}–standard positive measure, NN^{\prime} must be in 4{\mathbb{N}}_{4} in order to use the 𝕍4{\mathbb{V}}_{4}–version of 𝐋(m1){\bf L}(m-1) with 𝕍2{\mathbb{V}}_{2} being considered as the “standard” universe. But NN^{\prime} can only be guaranteed one universe apart from HH by Definition 3.1 even though NN is assumed to be at least two universes apart from HH. Therefore, we use hn1h_{n}\in{\mathbb{N}}_{1} (instead of HH) which leads to N21N^{\prime}\in{\mathbb{N}}_{2}\setminus{\mathbb{N}}_{1} and use i2i_{2} to lift NN^{\prime} to N′′32N^{\prime\prime}\in{\mathbb{N}}_{3}\setminus{\mathbb{N}}_{2} while keeping hn1h_{n}\in{\mathbb{N}}_{1}. This allows the use of 𝕍3{\mathbb{V}}_{3}–version of 𝐋2(m1){\bf L}_{2}(m-1) with 𝕍1{\mathbb{V}}_{1} being considered as the “standard” universe. Then spill hnh_{n} over to H21H\in{\mathbb{N}}_{2}\setminus{\mathbb{N}}_{1} and apply 𝐋1(m1){\bf L}_{1}(m-1) with N,A,S,UN,A,S,U being replaced by H,τH,[H],UH,\tau_{H},[H],U^{\prime} to obtain desired collections 𝒫{\mathcal{P}}^{\prime} and 𝒬{\mathcal{Q}}^{\prime} of kk–a.p.’s. All these steps rely on the fact that 𝐋2(m1){\bf L}_{2}(m-1) is an internal statement with internal parameters.

Theorem 5.4 (E. Szemerédi, 1975)

Let k0k\in{\mathbb{N}}_{0}. If D0D\subseteq{\mathbb{N}}_{0} has positive upper density, then DD contains nontrivial kk-term arithmetic progressions.

Proof It suffices to find a nontrivial kk–a.p. in i0(D)i_{0}(D). Let PP be an a.p. such that |P|1|P|\gg 1 and μ|P|(i0(D)P)=SD(D)=α\mu_{|P|}(i_{0}(D)\cap P)=S\!D(D)=\alpha. Then α>0\alpha>0 because α\alpha is greater than or equal to the upper density of DD. Let A=i0(D)PA=i_{0}(D)\cap P. Without loss of generality, we can assume P=[N]P=[N] for some N1N\gg 1. We can also assume that N21N\in{\mathbb{N}}_{2}\setminus{\mathbb{N}}_{1} because otherwise replace NN by i1(N)i_{1}(N) and AA by i1(A)i_{1}(A). Then we have μN(A)=SD(A)=α\mu_{N}(A)=S\!D(A)=\alpha. Set U=S=[N]U=S=[N]. Trivially, μN(S)=SD(S)=η=1\mu_{N}(S)=S\!D(S)=\eta=1, ASA\subseteq S, and SDS(A)=SD(A)=αS\!D_{S}(A)=S\!D(A)=\alpha. To start with k=k+1k^{\prime}=k+1 instead of kk, we have many nontrivial kk^{\prime}–a.p.’s p𝒫p\in{\mathcal{P}} such that p(l)Ap(l)\in A for lk1=kl\leq k^{\prime}-1=k in 𝐋1(k){\bf L}_{1}(k^{\prime}). So there must be many nontrivial kk–a.p.’s in Ai0(D)A\subseteq i_{0}(D). By 𝕍0𝕍2{\mathbb{V}}_{0}\prec{\mathbb{V}}_{2}, there must be nontrivial kk–a.p.’s in DD. \blacksquare

6 A Question

The construction of the nonstandard universes above requires the existence of a non-principal ultrafilter 0{\mathcal{F}}_{0} on 0{\mathbb{N}}_{0} and a well-order \lhd on some 𝕍(0,z){\mathbb{V}}({\mathbb{R}}_{0},z), which is a consequence of 𝖹𝖥𝖢\mathsf{ZFC}. Notice that 𝖹𝖥\mathsf{ZF} cannot guarantee the existence of 0{\mathcal{F}}_{0} although 𝖹𝖥\mathsf{ZF} plus the existence of 0{\mathcal{F}}_{0} and \lhd is strictly weaker than 𝖹𝖥𝖢\mathsf{ZFC}. However, assuming the existence of 0{\mathcal{F}}_{0} and \lhd may be avoided by the axiomatic approach of nonstandard analysis developed in [5]. In [5] two systems of axioms 𝖲𝖯𝖮𝖳\mathsf{SPOT} and 𝖲𝖢𝖮𝖳\mathsf{SCOT} are introduced. Roughly speaking, 𝖲𝖯𝖮𝖳\mathsf{SPOT} contains 𝖹𝖥\mathsf{ZF} plus some primitive tools for nonstandard analysis and 𝖲𝖢𝖮𝖳\mathsf{SCOT} contains 𝖹𝖥\mathsf{ZF} plus the axiom of dependent choice and some primitive tools for nonstandard analysis. It is shown in [5] that 𝖲𝖯𝖮𝖳\mathsf{SPOT} is a conservative extension of 𝖹𝖥\mathsf{ZF} and sufficient for developing basic calculus while 𝖲𝖢𝖮𝖳\mathsf{SCOT} is a conservative extension of 𝖹𝖥\mathsf{ZF} plus the axiom of dependent choice and sufficient for developing basic calculus and Lebesgue integration.

Question 6.1

Can the nonstandard proof of Szemerédi’s theorem in §5 be carried out in 𝖲𝖢𝖮𝖳\mathsf{SCOT} or even in 𝖲𝖯𝖮𝖳\mathsf{SPOT}?

Acknowledgments

The author would like to thank the American Institute of Mathematics which sponsored the workshop Nonstandard Methods in Additive Combinatorics, where he had an opportunity to attend Terence Tao’s lectures and learn from Tao’s interpretation of Szemerédi’s original proof of Szemerédi’s theorem [9]. The author would also like to thank Steven Leth, Isaac Goldbring, Mikhail Katz, Michael Benedikt, Karel Hrbáček, and the anonymous referee for comments, suggestions, and correcting some mistakes and typos in earlier versions of the paper.

References

  • [1] Chen Chung Chang and H. Jerome Keisler. Model Theory. (3rd edition), North-Holland, 1990.
  • [2] Mauro Di Nasso. Hypernatural numbers as ultrafilters. in Nonstandard Analysis for the Working Mathematician (Springer, Dordrecht, 2015), 443–474.
  • [3] Mauro Di Nasso, Isaac Goldbring, and Martino Lupini. Nonstandard Methods in Ramsey Theory and Combinatorial Number Theory. Lecture Notes in Mathematics book series, volume 2239, Springer, 2019.
  • [4] Robert Goldblatt. Lectures on the hyperreals–an introduction to nonstandard analysis. Springer, 1998.
  • [5] Karel Hrbáček and Mikhail G. Katz. Infinitesimal analysis without the Axiom of Choice. Annals of Pure and Applied Logic, 172 (6), June 2021
  • [6] Lorenzo Luperi Baglini. Hyperintegers and Nonstandard Techniques in Combinatorics of Numbers. PhD Dissertation (2012), University of Siena, arXiv: 1212.2049.
  • [7] Peter A. Loeb and Manfred P. H. Wolff, editors. Nonstandard Analysis for the Working Mathematician–Second edition. Springer, Dordrecht, 2015.
  • [8] Endre Szemerédi. On sets of integers containing no kk elements in arithmetic progression. Collection of articles in memory of Juriǐ Vladimirovič Linnik. Acta Arithmatica. 27 (1975): 199–245.
  • [9] Terence Tao. Szemerédi’s proof of Szemerédi’s theorem. Acta Mathematica Hungarica. 161 (2020): 443–487. https://terrytao.files.wordpress.com/2017/09/szemeredi-proof1.pdf
  • [10] Terence Tao. A nonstandard analysis proof of Szemerédi’s theorem.
    https://terrytao.wordpress.com/2015/07/20/a-nonstandard-analysis-proof-of-szemeredis-theorem/
  • [11] Terence Tao. Szemeredi’s proof of Szemeredi’s theorem.
    https://terrytao.wordpress.com/2017/09/12/szemeredis-proof-of-szemeredis-theorem/.
{dajauthors}{authorinfo}

[pgom] Renling Jin
College of Charleston
Charleston, South Carolina, USA

jinr\imageatcofc\imagedotedu
\urlhttp://jinr.people.cofc.edu