The Density Hales-Jewett Theorem in a Measure Preserving Framework Daniel Glasscock, April 2013

Similar documents
RAMSEY THEORY. 1 Ramsey Numbers

Idempotents in Compact Semigroups and Ramsey Theory H. Furstenberg and Y. Katznelson ( )

Notes on the Dual Ramsey Theorem

March 3, The large and small in model theory: What are the amalgamation spectra of. infinitary classes? John T. Baldwin

Löwenheim-Skolem Theorems, Countable Approximations, and L ω. David W. Kueker (Lecture Notes, Fall 2007)

Large subsets of semigroups

Ramsey theory and the geometry of Banach spaces

FUNCTIONAL ANALYSIS LECTURE NOTES: COMPACT SETS AND FINITE-DIMENSIONAL SPACES. 1. Compact Sets

5 Measure theory II. (or. lim. Prove the proposition. 5. For fixed F A and φ M define the restriction of φ on F by writing.

Combinatorial proofs of Gowers FIN k and FIN ± k

(U) =, if 0 U, 1 U, (U) = X, if 0 U, and 1 U. (U) = E, if 0 U, but 1 U. (U) = X \ E if 0 U, but 1 U. n=1 A n, then A M.

MAT1000 ASSIGNMENT 1. a k 3 k. x =

(1) Consider the space S consisting of all continuous real-valued functions on the closed interval [0, 1]. For f, g S, define

Indeed, if we want m to be compatible with taking limits, it should be countably additive, meaning that ( )

Introductory Analysis I Fall 2014 Homework #5 Solutions

Ergodic theorem involving additive and multiplicative groups of a field and

Problem set 1, Real Analysis I, Spring, 2015.

A generalization of modal definability

Measure-theoretic unfriendly colorings

RAMSEY ALGEBRAS AND RAMSEY SPACES

Part V. 17 Introduction: What are measures and why measurable sets. Lebesgue Integration Theory

Math 541 Fall 2008 Connectivity Transition from Math 453/503 to Math 541 Ross E. Staffeldt-August 2008

ULTRAFILTER AND HINDMAN S THEOREM

Measures. 1 Introduction. These preliminary lecture notes are partly based on textbooks by Athreya and Lahiri, Capinski and Kopp, and Folland.

V (v i + W i ) (v i + W i ) is path-connected and hence is connected.

Math 426 Homework 4 Due 3 November 2017

Problem Set 2: Solutions Math 201A: Fall 2016

Notions such as convergent sequence and Cauchy sequence make sense for any metric space. Convergent Sequences are Cauchy

An Introduction to Non-Standard Analysis and its Applications

Problems - Section 17-2, 4, 6c, 9, 10, 13, 14; Section 18-1, 3, 4, 6, 8, 10; Section 19-1, 3, 5, 7, 8, 9;

RAMSEY THEORY: VAN DER WAERDEN S THEOREM AND THE HALES-JEWETT THEOREM

HILBERT SPACES AND THE RADON-NIKODYM THEOREM. where the bar in the first equation denotes complex conjugation. In either case, for any x V define

l(y j ) = 0 for all y j (1)

Filters in Analysis and Topology

Part III. 10 Topological Space Basics. Topological Spaces

MULTIPLICATIVE RICHNESS OF ADDITIVELY LARGE SETS IN Z d

HINDMAN S THEOREM AND IDEMPOTENT TYPES. 1. Introduction

MATH41011/MATH61011: FOURIER SERIES AND LEBESGUE INTEGRATION. Extra Reading Material for Level 4 and Level 6

Final. due May 8, 2012

W if p = 0; ; W ) if p 1. p times

Measurable functions are approximately nice, even if look terrible.

arxiv: v1 [math.fa] 14 Jul 2018

Disjoint Hamiltonian Cycles in Bipartite Graphs

A disjoint union theorem for trees

Math 117: Topology of the Real Numbers

UNIVERSALITY OF THE LATTICE OF TRANSFORMATION MONOIDS

1 The Local-to-Global Lemma

ON THE EQUIVALENCE OF CONGLOMERABILITY AND DISINTEGRABILITY FOR UNBOUNDED RANDOM VARIABLES

SYNCHRONOUS RECURRENCE. Contents

NAME: Mathematics 205A, Fall 2008, Final Examination. Answer Key

Spanning and Independence Properties of Finite Frames

A NOTE ON TENSOR CATEGORIES OF LIE TYPE E 9

Banach Spaces V: A Closer Look at the w- and the w -Topologies

Sequences of height 1 primes in Z[X]

MATH 8253 ALGEBRAIC GEOMETRY WEEK 12

Banach Spaces II: Elementary Banach Space Theory

Rudiments of Ergodic Theory

Locally convex spaces, the hyperplane separation theorem, and the Krein-Milman theorem

The geometry of secants in embedded polar spaces

Free Subgroups of the Fundamental Group of the Hawaiian Earring

Chromatic Ramsey number of acyclic hypergraphs

Monochromatic Forests of Finite Subsets of N

Automorphism groups of wreath product digraphs

Introduction to Bases in Banach Spaces

LEBESGUE INTEGRATION. Introduction

arxiv:math/ v1 [math.lo] 5 Mar 2007

Remarks on a Ramsey theory for trees

5 Birkhoff s Ergodic Theorem

MATH5011 Real Analysis I. Exercise 1 Suggested Solution

SEPARABLE MODELS OF RANDOMIZATIONS

THE SEMIGROUP βs APPLICATIONS TO RAMSEY THEORY

Math 209B Homework 2

4 CONNECTED PROJECTIVE-PLANAR GRAPHS ARE HAMILTONIAN. Robin Thomas* Xingxing Yu**

16 1 Basic Facts from Functional Analysis and Banach Lattices

II - REAL ANALYSIS. This property gives us a way to extend the notion of content to finite unions of rectangles: we define

Analysis II Lecture notes

THE RADON NIKODYM PROPERTY AND LIPSCHITZ EMBEDDINGS

Consequences of Continuity

Axioms of separation

Real Analysis Chapter 1 Solutions Jonathan Conder

Real Variables: Solutions to Homework 3

Section 2: Classes of Sets

Topological dynamics: basic notions and examples

Ergodic Theory. Constantine Caramanis. May 6, 1999

be a path in L G ; we can associated to P the following alternating sequence of vertices and edges in G:

Topological properties

RAMSEY THEORY. Contents 1. Introduction Arithmetic Ramsey s Theorem

CHEVALLEY S THEOREM AND COMPLETE VARIETIES

Lebesgue Measure on R n

A dyadic endomorphism which is Bernoulli but not standard

Corrections to Introduction to Topological Manifolds (First edition) by John M. Lee December 7, 2015

Formal power series rings, inverse limits, and I-adic completions of rings

INVARIANT PROBABILITIES ON PROJECTIVE SPACES. 1. Introduction

Lecture 2: Connecting the Three Models

INVERSE LIMITS AND PROFINITE GROUPS

BALANCING GAUSSIAN VECTORS. 1. Introduction

Graph coloring, perfect graphs

AN ALGEBRAIC APPROACH TO GENERALIZED MEASURES OF INFORMATION

Multi-coloring and Mycielski s construction

THE FUNDAMENTAL GROUP AND CW COMPLEXES

Transcription:

The Density Hales-Jewett Theorem in a Measure Preserving Framework Daniel Glasscock, April 2013 The purpose of this note is to explain in detail the reduction of the combinatorial density Hales-Jewett theorem (DHJ) to an equivalent measure-theoretic dynamical theorem. This approach was used by Furstenberg and Katznelson in the first proof [3] of DJH in 1991. The approach taken here more closely resembles their earlier work [2] on DHJ 3. 1. Introduction Let k N, and write [k] for {1, 2,..., k}. The set of all words with letters from [k] of length n is then [k] n. Let [k] <ω denote the free semigroup of all words with letters from [k] of positive, finite length with concatenation. A variable word is a word with letters from [k] {t} in which t appears. Variable words should be considered non-constant functions [k] [k] <ω defined by substituting letters from [k] in place of t. A combinatorial line in [k] <ω is the image of a variable word. Theorem (DHJc): For all k N and ɛ > 0, there exists an N N such that for all n N and A [k] n with A > ɛk n, the set A contains a combinatorial line. A [k] <ω -system is a probability space (X, µ) together with k infinite sequences of invertible, measure preserving transformations {T (i) j } i [k] j N and a map T : [k]<ω {T (i) j } i [k] j N defined by where w = w 1 w 2 w n, w i [k]. T (w) = T (w1) 1 T (w2) 2 T n (wn) Theorem (DHJm): For all k N, [k] <ω -systems (X, µ, T ), and A X of positive measure, there exists a combinatorial line L such that ( ) µ T (w) 1 A > 0. w L It is in this form that Furstenberg and Katznelson [3] proved the density Hales-Jewett theorem. The goal here is to prove the equivalence between these two forms of the theorem. 2. Equivalent forms The first step is to move DHJc into a static measure-theoretic form. A [k] <ω -process in a probability space (X, µ) is a collection of measureable sets {B w } w [k] <ω indexed by [k] <ω. A [k] n -process is defined similarly. We will show by combinatorial arguments that DHJc is equivalent to the following static form of the density Hales-Jewett theorem. Theorem (DHJf): Let (X, µ) be a probability space. For all k N and ɛ > 0, there exists an N N such that for all n N and [k] n -processes {B w } in X for which µ(b w ) > ɛ for all w [k] n, there exists a combinatorial line L in [k] n such that µ( w L B w ) > 0. The following equivalent infinitary version will be more useful in establishing the dynamical connection. The equivalence is left as an easy exercise. 1

Theorem (DHJi): Let (X, µ) be a probability space. For all k N, ɛ > 0, and [k] <ω -processes {B w } in X for which µ(b w ) > ɛ for all w [k] <ω, there exists a combinatorial line L such that µ( w L B w ) > 0. The probability space (X, µ) plays an unimportant role here. What really matters is the relative position of the sets amongst themselves. In order to highlight this, we will focus on the following object. The (joint) distribution of a [k] <ω -process {B w } in X is the function d which for all n N is defined on subsets W [k] n by ( ) d(w ) = µ B w. Since many different [k] <ω -processes may have the same distribution, it is worthwhile to consider distributions without making reference to the processes from which they come. If Z k = n P([k] n ) is the disjoint union of the power sets of each [k] n, then distributions of [k] <ω -processes (as functions on Z k with values in [0, 1]) are points in the compact metric space [0, 1] Z k. (An example of a metric: n n 2 kn W [k] d(w ).) A point ϕ [0, n 1]Z k is called a (joint) distribution if it is the distribution of a [k] <ω -process in some probability space (such a process is called a parent process for ϕ). The equivalence between the previous theorem and the next is left as an easy exercise. w W Theorem (DHJd): For all k N, ɛ > 0, and distributions d [0, 1] Z k w [k] <ω, there exists a combinatorial line L such that d(l) > 0. for which d({w}) > ɛ for all To motivate the next definition, consider the [k] <ω -process {T (w) 1 A} associated with a [k] <ω -system (X, µ, T ) and a set A X of positive measure. This process is stationary in the following sense: if w 1 and w 2 are of the same length, then for all v [k] <ω, µ ( T (w 1 v) 1 A T (w 2 v) 1 A ) = µ ( T (w 1 ) 1 A T (w 2 ) 1 A ). A [k] <ω -process {B w } in (X, µ) is stationary if for all n N, W [k] n, and v [k] <ω, ( ) ( ) µ B wv = µ B w. w W A distribution d is called stationary if for all n N, W [k] n, and v [k] <ω, d(w v) = d(w ). It follows from the definitions that a distribution is stationary if and only its parent processes are stationary. It will be shown later that an arbitrary distribution has subdistributions which approximate stationary ones. This fact will allow us to show the equivalence of DHJd and the following stationary form. w W Theorem (DHJs): For all k N, ɛ > 0, and stationary distributions d [0, 1] Z k for all w [k] <ω, there exists a combinatorial line L such that d(l) > 0. for which d({w}) > ɛ The final step will be to show the equivalence of DHJs and DHJm. This is accomplished by realizing any stationary distribution as a distribution of a [k] <ω -orbit of some set of positive measure. In summary, we will prove DHJc DHJf DHJi DHJd DHJs DHJm. 3. Equivalence of DHJc and DHJf Both of the following proofs are directly from [3]. First, DHJc implies DHJf. 2

Proof. Let k N, ɛ > 0. Let N be as given in DHJc. Let n N and {B w } be a [k] n -process in X for which µ(b w ) > ɛ for all w [k] n. For all x X, let A(x) = {w [k] n x B w }. Then A(x) = w [k] χ n B w, so A(x) dµ > ɛk n. X Thus there exists a C X of positive measure such that for all x C, A(x) > ɛk n. By DHJc, for all x C, the set A(x) [k] n contains a combinatorial line L(x). Since there are only finitely many combinatorial lines in [k] n, there exists a C C of positive measure such that L(x) = L is constant on C. This means that C w L B w, and we are done since C has positive measure. Conversely, DHJf implies DHJc. Proof. Let k N. First, we claim that that DHJc holds with N = 2 if ɛ > 1 1/k. Let n N and A [k] n with A > ɛk n. For i [k], let A i = {w A w 1 = i} and A i [k]n 1 be the projection of A i onto the last n 1 letters. There exists a v i [k] A i since otherwise A kn k n 1, a contradiction. Then {iv i [k]} is a combinatorial line in A. Let ɛ 0 be the infimum over all ɛ for which the conclusion of DHJc is valid. If ɛ 0 = 0, we are done. Otherwise, 0 < ɛ 0 1 1/k. Let m be large enough to satisfy the conclusions of DHJf for processes of measure greater than ɛ 0 /2. Let ɛ 1 = ɛ 0 (1 k m 2 ) so that ɛ 2 = ɛ 1 + ɛ 0 2 k m > ɛ 0. Since ɛ 2 > ɛ 0, there exists an M such that the conclusion of DHJc is valid for n M and sets of density greater than ɛ 2. We claim that the conclusion of DHJc valid for sets of density greater than ɛ 1 and n m + M, which, since ɛ 1 < ɛ 0, contradicts the definition of ɛ 0. Let n m + M and A [k] n have density greater than ɛ 1. Consider the [k] m -process {A w} in [k] n m defined by A w = {u [k] n m wu A}. There are two cases to consider. Case 1: each A w has density great than ɛ 0 /2. By DHJf, there exists a combinatorial line L [k] m such that w L A w has positive measure (density). In particular, the intersection is non-empty, so there exists a v w L A w. Now Lv is a combinatorial line in A. Case 2: there exists a w [k] m such that A w has density at most ɛ 0 /2. By double counting A, the average of the densities of the A w s is the density of A. Since A has density greater than ɛ 1, by the definition of ɛ 2, there exists a v [k] m such that A v [k] n m has density great than ɛ 2. Since n m M, there exists a combinatorial line L in A v. Then vl is a combinatorial line in A. 4. Subspaces, subprocesses, and subdistributions Notions of subspaces of [k] <ω, subprocesses, and subdistributions are essential to proving these equivalences. A variable sentence Σ is a concatenation w 1 (t 1 )w 2 (t 2 ) ([k] {t i } i N ) N of infinitely many variable words, each with a distinct variable. Variable sentences may be thought of as functions Σ : [k] <ω [k] <ω by defining Σv = w 1 (v 1 ) w n (v n ) where v = v 1 v n, v i [k]. A (infinite dimensional) subspace of [k] <ω is a subset of [k] <ω of the form Σ[k] <ω for some variable sentence Σ. Thus, elements of subspaces of [k] <ω are indexed by [k] <ω. Note that subspaces of [k] <ω and variable sentences are in 1-to-1 correspondence. Let Σ 1, Σ 2 be variable sentences. It is left to the reader to check that Σ 1 [k] <ω Σ 2 [k] <ω if and only if there exists a variable sentence Σ 3 for which Σ 1 = Σ 2 Σ 3. In case either (both) hold, Σ 1 [k] <ω is called a subspace of Σ 2 [k] <ω. The first characterization is more apparent, while the second means that 3

subspaces of subspaces of [k] <ω are subspaces of [k] <ω. Subsets of [k] <ω of the form Σ[k] n for a variable sentence Σ are called n-dimensional subspaces of [k] <ω. Note that a combinatorial line is exactly a 1-dimensional subspace, and so a combinatorial line in a subspace of [k] <ω is a combinatorial line in [k] <ω. Let {B w } be a [k] <ω -process in (X, µ). A subprocess of {B w } is a [k] <ω -process of the form {B Σw } for some variable sentence Σ. Subprocesses of {B w } are also Σ[k] <ω -processes, but it will be advantageous to consider all processes and subprocesses as indexed by [k] <ω. Similarly, a subdistribution of a distribution d is a distribution of the form d Σ for some variable sentence Σ. (By a slight abuse of notation, Σ acts element-wise to take subsets of [k] <ω to subsets of [k] <ω.) 5. The regular process and space of distributions The next lemma shows that the probability space (X, µ) plays a secondary role. Let m be the Lebesgue measure on [0, 1], and for each n N, fix an ordering of P([k] n ). Lemma 1: For all k N and distributions d [0, 1] Z k, there exists a parent process in ([0, 1], m). Proof. Let k N, d [0, 1] Z k be a distribution, and {B w } be a parent [k] <ω -process in (X, µ). We will describe a [k] <ω -process in ([0, 1], m) with distribution d. Let n N, and for each W [k] n, let p(w ) = µ( w W B w \ w W B w ). That is, p measures the cells of the partition of X created by {B w } w [k] n. For each W [k] n, in order, let I W be a closed interval of length p(w ) in [0, 1] disjoint from the previous ones with left endpoint situated as close to 0 as possible. This is possible since W [k] p(i n W ) = 1. For each w [k] n, let C w = W w I W. It is straightforward to check that the resulting [k] <ω -process {C w } in [0, 1] has distribution d. Given a distribution d, the parent process {C w } created above will be called the regular process for d. Lemma 2: For all k N, the set of distributions in [0, 1] Z k is closed. Proof. Let (d i ) i N be a sequence of distributions converging to ϕ in [0, 1] Z k. For each i N, let {C i,w } be the regular [k] <ω -process for d i. For each w [k] <ω, the sequence of endpoints of the intervals composing C i,w converge to the endpoints of a set of intervals which compose a C w. It is easily verified that the resulting [k] <ω -process {C w } in [0, 1] has distribution ϕ. 6. Equivalence of DHJd and DHJs That DHJd implies DHJs is clear. The reverse implication will follow by the compactness of [0, 1] Z k and Lemma 2. More specifically, we will show that any distribution has subdistributions limiting to a stationary distribution. After applying DHJs to the corresponding stationary distribution, we may approximate it well enough to carry the conclusion back to a subdistribution of the original distribution. In order to accomplish this, we need to formalize the notion of being close to stationary. If ɛ > 0, a distribution d is ɛ-stationary if for all n ɛ 1, W [k] n, and v [k] <ω, d (W v) d (W ) < ɛ. For a distribution d [0, 1] Z k, let S(d) be the closure in [0, 1] Z k of the set of subdistributions of d. To prove the next lemma, we will make use of the following theorem due to Carlson and Simpson [1]. It can be thought of as having the same relationship to the Hales-Jewett theorem as Hindman s theorem has to Schur s. 4

Theorem (CS): For all k N and finite colorings of [k] <ω, there exists a monochromatic subspace. Lemma 3: For all k N and distributions d [0, 1] Z k, S(d) contains stationary distributions. Proof. Suppose first that S(d) contains ɛ-stationary distributions for arbitrary small ɛ. For each i N, let d i be an ɛ i -stationary distribution in S(d) where ɛ i 0. Since [0, 1] Z k is compact, there exists a subsequence (d ni ) i N which converges to a point ϕ [0, 1] Z k. Since S(d) and the set of distributions are both closed, ϕ is a distribution in S(d). To see that ϕ is stationary, let n N, W [k] n, and v [k] <ω. Let δ > 0. Choose i N such that ɛ ni < min(1/n, δ/3) and such that the distance between d ni and ϕ in the metric on [0, 1] Z k is so small that ϕ(w ) d ni (W ) < δ/3 and ϕ(w v) d ni (W v) < δ/3. Then ϕ(w v) ϕ(w ) ϕ(w v) d ni (W v) + d ni (W v) d ni (W ) + d ni (W ) ϕ(w ) < δ. Since δ was arbitrary, ϕ(w v) = ϕ(w ), so ϕ is stationary. Thus it suffices to show that for all ɛ > 0, S(d) contains an ɛ-stationary distribution. Let ɛ > 0. We will prove the following by induction: for all N N, there exists a variable word Σ such that for all n N, W [k] n, and v [k] <ω, d(σw ) d(σw v) < ɛ. Base case: N = 1. Partition [0, 1] into length ɛ subintervals α 1, α 2,, α P, and color [k] <ω with P 2k colors in the following way. For each v [k] <ω, let c(v) = (c W ) W [k] where d(w v) α cw. By CS, there exists a variable sentence Σ = w 1 (t 1 )w 2 (t 2 ) with monochromatic range. Let Σ = w 0 (t 0 )w 2 (t 2 ) where w 0 (t 0 ) = t 0 w 1 (1). Now suppose W [k] and v [k] <ω. Then for w W, Σw = wσ 1 and Σwv = wσ 1v. Since Σ 1 and Σ 1v have the same color, d(σw ) d(σw v) = d(w Σ 1) d(w Σ 1v) < ɛ. Assume now that the induction hypothesis holds for some N N. Since sub-subdistributions are subdistributions, we may without loss of generality assume that d is such that for all n N, W [k] n, and v [k] <ω, d(w ) d(w v) < ɛ. Partition [0, 1] into length ɛ subintervals α 1, α 2,, α P, and color [k] <ω with P 2kn colors in the following way. For each v [k] <ω, let c(v) = (c W ) W [k] n where d(w v) α cw. By CS, there exists a variable sentence Σ = w 1 (t 1 )w 2 (t 2 ) with monochromatic range. Let Σ = t N t 1 w 0 (t 0 )w 2 (t 2 ) where w 0 (t 0 ) = t 0 w 1 (1). Suppose n N, W [k] n, and v [k] <ω. Then for w W, Σw = w and Σwv = wṽ where ṽ is a word independent of w. Since W [k] n, by the induction hypothesis, d(σw ) d(σw v) = d(w ) d(w ṽ) < ɛ. Suppose n = N + 1, W [k] n, and v [k] <ω. Then for w W, Σw = wσ 1 and Σwv = wσ 1v. Since Σ 1 and Σ 1v have the same color, d(σw ) d(σw v) = d(w Σ 1) d(w Σ 1v) < ɛ. The induction yields an ɛ-stationary distribution in S(d). We now have the tools to prove that DHJs implies DHJd. Proof. Let k N, ɛ > 0, and d [0, 1] Z k be a distribution for which d({w}) > ɛ for all w [k] <ω. By Lemma 3, there exists a stationary distribution d in S(d). Let (Σ i ) i N be a sequence of variable sentences such that d Σ i d in [0, 1] Z k. 5

For all w [k] <ω, d Σ i ({w}) d({w}), whereby d({w}) > ɛ/2. Invoking DHJs for d, there exists a combinatorial line L such that d(l) > 0. Since d Σ i (L) d(l), there exists an i N for which d Σ i (L) > 0. Now L = Σ i L is a combinatorial line for which d(l ) > 0. 7. Equivalence of DHJs and DHJm It remains to show that DHJs is equivalent to DHJm. Since the [k] <ω -process {T (w) 1 A} associated with a [k] <ω -system (X, µ, T ) and a set A X of positive measure has stationary distribution, it is clear that DHJs implies DHJm. The reverse implication will follow from the fact that any stationary distribution looks like the distribution of a [k] <ω -orbit of a set of positive measure. We will rely on the following lemma in order to craft an appropriate [k] <ω -system on which to apply DHJm. Lemma 4: Let B 0 and B 1 be finite algebras of measureable sets in ([0, 1], m). Assume that there is a measure preserving isomorphism ψ : B 0 B 1. Then there exists an invertible measure preserving transformation ψ of ([0, 1], m) which induces ψ, i.e. ψ(b) = ψ 1 (B). We conclude this note with the proof that DHJm implies DHJs. Proof. Let k N, ɛ > 0, and d [0, 1] Z k be a stationary distribution for which d({w}) > ɛ for all w [k] <ω. Using the subspace Σ = 1t 1 t 2 and passing to the subdistribution d Σ, we may assume without loss of generality that d is constant on the set {{1},..., {k}}. Let {C w } be the regular [k] <ω - process for d. For a finite set Y of measurable sets, denote by Y the finite algebra generated by Y. For all v [k], {C v } is isomorphic to {C 1 } because m(c v ) = m(c 1 ). By Lemma 4, for all v [k], there exists an invertible, measure preserving transformation T (v) 1 of [0, 1] such that T (v) 1 C v = C 1. Let n 2. Since d is stationary, for all v [k], the obvious correspondence between {C w w [k] n 1 } and {C wv w [k] n 1 } is an isomorphism. By Lemma 4, for all v [k], there exists an invertible, measure preserving transformation T n (v) such that for all w [k] n 1, T n (v) C wv = C w. Now ([0, 1], m), the maps {T (i) j } i [k] j N (w1) defined above, and T (w) = T 1 T (w2) 2 T n (wn) yield a [k] <ω - system. By the choice of the T (i) j s, for all w [k] <ω, T (w)c w = C 1. Applying DHJm with A = C 1, there exists a combinatorial line L such that m( w L T (w) 1 C 1 ) > 0. But T (w) 1 C 1 = C w, and so d( w L C w ) > 0. References [1] Timothy J. Carlson and Stephen G. Simpson. A dual form of Ramsey s theorem. Adv. in Math., 53(3):265 290, 1984. [2] H. Furstenberg and Y. Katznelson. A density version of the Hales-Jewett theorem for k = 3. Discrete Math., 75(1-3):227 241, 1989. Graph theory and combinatorics (Cambridge, 1988). [3] H. Furstenberg and Y. Katznelson. A density version of the Hales-Jewett theorem. J. Anal. Math., 57:64 119, 1991. 6