13.07.2015 Views

Dynamics of sets of zero density - IMPA

Dynamics of sets of zero density - IMPA

Dynamics of sets of zero density - IMPA

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Yuri Gomes Lima<strong>Dynamics</strong> <strong>of</strong> <strong>sets</strong> <strong>of</strong> <strong>zero</strong> <strong>density</strong>Tese submetida ao Instituto Nacional deMatemática Pura e Aplicada - <strong>IMPA</strong> - paraobtenção do título de Doutor em Ciências.Orientador: Pr<strong>of</strong>. Carlos Gustavo Moreira.Orientador: Pr<strong>of</strong>. Enrique Ramiro Pujals.Rio de JaneiroMarço de 2011


To my beloved mom, Glaucia.


AcknowledgementsMy Ph.D. degree is part <strong>of</strong> at least ten years <strong>of</strong> my life, so I will acknowledge people from thevery ancient times.My academic career began in high school through mathematical olympiads, whose first encouragementscame from my 9th grade teacher Max and developments by Antonio Caminhaand On<strong>of</strong>re Campos. Its an honor having them, nowadays, as true friends.Later on, I was admitted to the undergraduate studies at Universidade Federal do Ceará. Ithank Pr<strong>of</strong>essor Abdênago Barros, who provided all the administrative tools necessary for myexchange from Engineering to Mathematics. During this time, I had the help <strong>of</strong> many pr<strong>of</strong>essors,<strong>of</strong> which I highlight Plácido Andrade, João Lucas Barbosa, Abdênago Barros, AntonioCaminha, LuquésioJorge, Levi Lima, FábioMontenegro, Francisco Pimentel andRomildoSilva.In 2007, I began my graduate studies at <strong>IMPA</strong>, supported by one <strong>of</strong> my two advisors, CarlosGustavo Moreira - Gugu - and Marcelo Viana. Some months later I met my other advisor,Enrique Pujals. A huge acknowledgement is devoted to them, for their valuable mathematicalsupport, covering <strong>of</strong> conference expenses and advices in life. Yet more special thanks are givento my advisors, who (Lebesgue almost-every occasion) patiently heard my doubts and explainedin the simplest way; who gave me insights to the development <strong>of</strong> this thesis. I have learned alot from them, from the happiness <strong>of</strong> Gugu to the optimism <strong>of</strong> Enrique; from the social ideas <strong>of</strong>Gugu to the political ones <strong>of</strong> Enrique. Their flaws are that Enrique does not play football andGugu does terribly bad.In the middle <strong>of</strong> my stay at <strong>IMPA</strong> I had the opportunity <strong>of</strong> doing a semester internship atThe Ohio State University, under the supervision <strong>of</strong> Vitaly Bergelson. I would like to thankhim for such guidance, enormous encouragement and also to the suggestion <strong>of</strong> research topic,which result is Chapter 5 <strong>of</strong> the present work. I made many friends in Columbus, <strong>of</strong> which Ihighlight Naomi Adaniya, Fabian Benitez, Darcey Hull, Kathrin Levine, Daniel Munoz, SanjnaShah, Sylvia Silveira and Laura Tompkins.Carlos Matheus has been an alma mater to me. He teaches me a lot, specially in topics Idon’t have any knowledge. In addition, he shared his blog for posts about my research. He havehad many discussions, together with Aline Cerqueira - hearing everything, telling stories about“the dark side <strong>of</strong> Rio” and supporting us with her special care. I thank Alejandro Kocsard, forii


his patience and availability in my latest project and Jacob Palis for his influence and globalperspectives.I have many considerations to <strong>IMPA</strong> employees. As they are too many, I represent each departmentby one person: Guilherme Brondi, Nelly Carvajal, Fernando Codá, Alexandre Maria,Sonia Melo, Samantha Nunes, Ana Paula Rodrigues, Fatima Russo and Fábio Santos. Mathematicalfriends also contributed (a lot!) to the fulfillment <strong>of</strong> my Ph.D.: Alan Prata, Martin deBorbon, Fernando Carneiro, Thiago Catalan, Pablo Dávalos, Jorge Erick, Tertuliano Franco,Pablo Guarino, Marcelo Hilário, Cristina Lizana, Ives Macedo, José Manuel, Fernando Marek,Jyrko Morris, Adriana Neumann, Ivaldo Nunes, Carlo Pietro, Artem Raibekas, José Régis,Patrícia Romano, Waliston Silva, Wanderson Silva, Vanessa Simões, Ricardo Turolla and FranciscoValenzuela. I thank the financial support provided by Faperj-Brazil, <strong>IMPA</strong> and Pronex.To those who shared the apartment with me: Alex Abreu, Samuel Barbosa, Thiago Barros,Renan Finder and Helio Macedo. Yes, it demands a great effort to stand me. Thanks to themfor this and the allowance <strong>of</strong> transforming the apartment into “home, sweet home”.Outside the mathematical scenario, I would like to acknowledge my friends Bruno Cals,Emanuel Carneiro, Filipe Carlos, Eliza Dias, S<strong>of</strong>ia Galvão, Lívia Leal, Caio Leão, ClarissaLima, Renan Lima, João Marcelo, Davi Maximo, Mayara Mota, Rodrigo Nobre, José Ricardo,André Silveira, Cícero Thiago and Chico Vieira.The golden key <strong>of</strong> my acknowledgements go to the basis <strong>of</strong> everything, my family: parentsGlaucia, Raul and Gilberto; siblings Gisele and Patrick; aunts Fatinha, Gardênia; cousinsGabriela Gomes, Diego Lacerda, Igor Marinho, Harley Teixeira, Mardey Teixeira and RebecaTeixeira. They give me the strength and support that allow me to face the everyday life difficultiesand good moments.iii


“Ora (direis) ouvir estrelas! CertoPerdeste o senso!” E eu vos direi, no entanto,Que, para ouvi-las, muita vez despertoE abro as janelas, pálido de espanto...E conversamos toda a noite, enquantoA via láctea, como um pálio aberto,Cintila. E, ao vir do sol, saudoso e em pranto,Inda as procuro pelo céu deserto.Direis agora: “Tresloucado amigo!Que conversas com elas? Que sentidoTem o que dizem, quando estão contigo?”E eu vos direi: “Amai para entendê-las!Pois só quem ama pode ter ouvidoCapaz de ouvir e de entender estrelas”.Via Láctea, de Olavo Bilac.


AbstractThis thesis is devoted to the study <strong>of</strong> <strong>sets</strong> <strong>of</strong> <strong>zero</strong> <strong>density</strong>, both continuous - inside R - anddiscrete - inside Z. Most <strong>of</strong> the arguments are combinatorial in nature.The interests are tw<strong>of</strong>old. The first investigates arithmetic sums: given two sub<strong>sets</strong> E,F <strong>of</strong>R or Z, what can be said about E+λF for most <strong>of</strong> the parameters λ ∈ R? The continuous casedates back to Marstrand. We give an alternative combinatorial pro<strong>of</strong> <strong>of</strong> his theorem for theparticular case <strong>of</strong> products <strong>of</strong> regular Cantor <strong>sets</strong> and, in the sequel, extend these techniques togive the pro<strong>of</strong> <strong>of</strong> the general case.We also discuss Marstrand’s theorem in discrete spaces. More specifically, we proposeafractaldimension for sub<strong>sets</strong> <strong>of</strong> the integers and establish a Marstrand type theorem in this context.The second interest is ergodic theoretical. It contains constructions <strong>of</strong> Z d -actions withprescribed topological and ergodic properties, such as total minimality, total ergodicity andtotal strict ergodicity. These examples prove that Bourgain’s polynomial pointwise ergodictheorem has not a topological version.v


ResumoO presente trabalho estuda conjuntos de densidade <strong>zero</strong>, tanto contínuos - subconjuntos deR - quanto discretos - subconjuntos de Z. Os argumentos são, em sua maioria, de naturezacombinatória.Os interesses são divididos em duas partes. O primeira investiga somas aritméticas: dadossubconjuntos E,F de R ou Z, o que se podeinferir a respeito doconjunto E+λF paraamaioriados parâmetros λ ∈ R? O caso contínuo tem origem nos trabalhos de Marstrand. Aqui, é dadauma prova alternativa desse teorema no caso em que o conjunto é produto de dois conjuntosde Cantor regulares e, posteriormente, as técnicas aplicadas são estendidas para a obtenção deuma prova do caso geral.Ainda na primeira parte, discutimos o Teorema de Marstrand em espaços discretos. Maisespecificamente, propomos uma dimensão fractal para subconjuntos de inteiros e provamos umresultado do tipo Marstrand nesse contexto. Uma grande quantidade de exemplos é discutida,além de contraexemplos para possíveis extensões do resultado.Osegundointeresseéergódico. ConstruímosZ d -ações compropriedadesergódicasetopológicasprescritas,comototalmenteminimal, totalmenteergódicoetotalmenteestritamenteergódico.Tais exemplos provam que não existe uma versão topológica para o teorema de Bourgain sobrea convergência pontual de médias ergódicas polinomiais.vi


List <strong>of</strong> Publications Z d -actions with prescribed topological and ergodic properties. Toappear in ErgodicTheory& Dynamical Systems. A combinatorial pro<strong>of</strong> <strong>of</strong> Marstrand’s Theorem for products <strong>of</strong> regular Cantor <strong>sets</strong>, withCarlos Gustavo Moreira. To appear in Expositiones Mathematicae. Yet another pro<strong>of</strong> <strong>of</strong> Marstrand’s theorem, with Carlos Gustavo Moreira. To appear inBulletin <strong>of</strong> the Brazilian Mathematical Society. A Marstrand theorem for sub<strong>sets</strong> <strong>of</strong> integers, with Carlos Gustavo Moreira. Available atwww.impa.br/∼yurilima.vii


viii


Contents1 Brief description 11.1 Goals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.1.1 Largeness in R and Z . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.1.2 Zero <strong>density</strong> in R and Z . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31.2 Chapter 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41.3 Chapter 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51.4 Chapter 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61.5 Chapter 5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72 A combinatorial pro<strong>of</strong> <strong>of</strong> Marstrand’s Theorem for products <strong>of</strong> regular Cantor<strong>sets</strong> 92.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92.2 Regular Cantor <strong>sets</strong> <strong>of</strong> class C 1+α . . . . . . . . . . . . . . . . . . . . . . . . . . . 112.3 Pro<strong>of</strong> <strong>of</strong> Theorem 2.1.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142.4 Pro<strong>of</strong> <strong>of</strong> Theorem 2.1.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172.5 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193 Yet another pro<strong>of</strong> <strong>of</strong> Marstrand’s Theorem 213.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213.2 Preliminaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243.2.1 Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243.2.2 Dyadic squares . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243.3 Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253.3.1 Rewriting the integral I i . . . . . . . . . . . . . . . . . . . . . . . . . . . . 273.3.2 Transversality condition . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283.3.3 Good covers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 293.4 Pro<strong>of</strong> <strong>of</strong> Theorems 3.1.1 and 3.1.2 . . . . . . . . . . . . . . . . . . . . . . . . . . 303.5 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32ix


4 A Marstrand theorem for sub<strong>sets</strong> <strong>of</strong> integers 334.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 334.2 Preliminaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 364.2.1 General notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 364.2.2 Counting dimension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 364.3 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 384.3.1 Polynomial sub<strong>sets</strong> <strong>of</strong> Z . . . . . . . . . . . . . . . . . . . . . . . . . . . . 384.3.2 Cantor <strong>sets</strong> in Z . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 394.3.3 Generalized IP-<strong>sets</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 424.4 Regularity and compatibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 444.4.1 Regular <strong>sets</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 444.4.2 Compatible <strong>sets</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 474.4.3 A counterexample <strong>of</strong> Theorem 4.1.1 for regular non-compatible <strong>sets</strong> . . . 474.4.4 Another counterexample <strong>of</strong> Theorem 4.1.1 . . . . . . . . . . . . . . . . . . 494.5 Pro<strong>of</strong>s <strong>of</strong> the Theorems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 504.6 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 545 Z d -actions with prescribed topological and ergodic properties 575.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 575.2 Preliminaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 605.2.1 Group actions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 605.2.2 Subgroups <strong>of</strong> Z d . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 635.2.3 Symbolic spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 635.2.4 Topological entropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 655.2.5 Frequencies and unique ergodicity . . . . . . . . . . . . . . . . . . . . . . 655.2.6 Law <strong>of</strong> Large Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 675.3 Main Constructions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 685.3.1 Minimality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 695.3.2 Total minimality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 695.3.3 Total strict ergodicity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 715.3.4 Pro<strong>of</strong> <strong>of</strong> Theorem 5.1.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 755.4 Pro<strong>of</strong> <strong>of</strong> Theorems 5.1.2 and 5.1.3 . . . . . . . . . . . . . . . . . . . . . . . . . . 765.4.1 Pro<strong>of</strong> <strong>of</strong> Theorem 5.1.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 775.4.2 Pro<strong>of</strong> <strong>of</strong> Theorem 5.1.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78


Chapter 1Brief descriptionThis thesis is devoted to the study <strong>of</strong> <strong>sets</strong> <strong>of</strong> <strong>zero</strong> <strong>density</strong>, both continuous - inside R - anddiscrete - inside Z. Most <strong>of</strong> the arguments are combinatorial in nature.The interests are tw<strong>of</strong>old. The first investigates arithmetic sums: given two sub<strong>sets</strong> E,F <strong>of</strong>R or Z, what can be said about E+λF for most <strong>of</strong> the parameters λ ∈ R? The continuous casedates back to Marstrand, as we will describe in the next section. The discrete case is new [24].The second interest is ergodic theoretical. Specifically, it is shown that Bourgain’s polynomialpointwise ergodic theorem has not a topological version.Each chapter originated an article and for this reason is self-contained. Chapters 2, 3, 4 werewritten together with one <strong>of</strong> my Ph.D. advisors, Carlos Gustavo Moreira (Gugu), and Chapter5 by myself as I was a visiting scholar at The Ohio State University under the supervision <strong>of</strong>Vitaly Bergelson.The present chapter describes the main goals <strong>of</strong> this work and gives a brief description <strong>of</strong>each <strong>of</strong> the next chapters.1.1 GoalsConsider the following table.Continuous DiscretePositive <strong>density</strong> ̌ ̌Zero <strong>density</strong> ̌ ?For each box with the checkmark ̌, we represent a situation for which there is an activeresearch and a solid theory has been developed, as we will describe in the sequel. The only1


2unchecked box concerns discrete <strong>sets</strong> <strong>of</strong> <strong>zero</strong> <strong>density</strong>. These objects are <strong>of</strong> great importance, notonly for intrisic reasons, but also because many <strong>of</strong> them are natural arithmetic <strong>sets</strong> that appearin diverse areas <strong>of</strong> Mathematics, from which we mention• the prime numbers: even having <strong>zero</strong> <strong>density</strong>, B. Green and T. Tao [16] proved that the primenumbers are combinatorially rich in the sense that they contain infinitely long arithmeticprogressions;• J. Bourgain [6] has proved almost sure pointwise convergence <strong>of</strong> ergodic averages along polynomials.The above results strongly rely on the particular combinatorial and analytical structure <strong>of</strong>each set. In this thesis, we obtain results in a more general point <strong>of</strong> view. Yet concerning thetable above, let us contrast the continuous and discrete situations.1.1.1 Largeness in R and ZLet m denote the Lebesgue measure <strong>of</strong> R. We ask the following question: if K ⊂ R is large inthe sense <strong>of</strong> m, does it inherit some structure from R? A satisfactory answer is given byTheorem 1.1.1 (Lebesgue differentiation). If K ⊂ R is a Borel set with m(K) > 0, thenfor m-almost every x ∈ K.m(K ∩(x−ε,x+ε))lim= 1ε→0 + 2εIn other words, copies <strong>of</strong> intervals <strong>of</strong> the real line “almost” appear in large <strong>sets</strong> <strong>of</strong> R. Surprisingly,the same phenomenon holds for sub<strong>sets</strong> <strong>of</strong> Z. Given a subset E ⊂ Z, let d ∗ (E) denotethe upper-Banach <strong>density</strong> <strong>of</strong> E, defined asd ∗ |E ∩{M +1,...,N}|(E) = limsup·N−M→∞ N −MWe thus have a discrete version <strong>of</strong> Theorem 1.1.1.Theorem 1.1.2 (Szemerédi [45]). If E ⊂ Z has positive upper Banach <strong>density</strong>, then E containsarbitrarily long arithmetic progressions.One can interpret this result by saying that <strong>density</strong> represents the correct notion <strong>of</strong> largenessneeded to preserve finite configurations <strong>of</strong> Z. These two situations just described are examples<strong>of</strong> Ramsey’s principle which, roughly speaking, asserts thatIf a substructure occupies a positive portion <strong>of</strong> a structure, then it must contain “copies” <strong>of</strong>the structure.


31.1.2 Zero <strong>density</strong> in R and ZHowever, theorems 1.1.1 and 1.1.2 can not infer anything if the <strong>density</strong> is <strong>zero</strong>, but these <strong>sets</strong>might posses rich combinatorial properties as well. For example, the ternary Cantor set⎧⎫⎨∑⎬K = a⎩ i ·3 −i ; a i = 0 or 2⎭n≥1satisfies K + K = [0,2] and naturally appears in the theory <strong>of</strong> homoclonic bifurcations onsurfaces. Regarding the continuous case, one can investigate largeness <strong>of</strong> a set <strong>of</strong> <strong>zero</strong> <strong>density</strong>via its Hausdorff dimension. It is a classical result in Geometric Measure Theory the followingTheorem 1.1.3 (Marstrand [27]). If K ⊂ R 2 is a Borel set with Hausdorff dimension greaterthan one, then the projection <strong>of</strong> K into R in the direction <strong>of</strong> λ has positive Lebesgue measurefor m-almost every λ ∈ R.Informally speaking, this means that <strong>sets</strong> <strong>of</strong> Hausdorff dimension greater than one are “fat”in almost every direction. In particular, if K = K 1 × K 2 is a cartesian product <strong>of</strong> two onedimensionalsub<strong>sets</strong> <strong>of</strong> R, Marstrand’s theorem translates to “m(K 1 +λK 2 ) > 0 for m-almostevery λ ∈ R”.Up to now, a theory <strong>of</strong> fractal <strong>sets</strong> in Z has not been developed. This constitutes the maingoal <strong>of</strong> this thesis. We will investigate these <strong>sets</strong> from geometrical and ergodic points <strong>of</strong> view,according to the diagram below.PART 1: GEOMETRYMarstrand’s theorem for sub<strong>sets</strong> <strong>of</strong> Z↑Marstrand’s theorem in R: new pro<strong>of</strong>PART 2: ERGODICErgodic theorems along <strong>sets</strong> <strong>of</strong> <strong>zero</strong> <strong>density</strong>↑Z d -actions with prescribed topological and ergodic properties


4Part 1 is the content <strong>of</strong> chapters 2, 3 and 4 and Part 2 <strong>of</strong> chapter 5. In order to obtain aMarstrand’s theorem for sub<strong>sets</strong> <strong>of</strong> Z, we first establish a new pro<strong>of</strong> <strong>of</strong> the classical Marstrand’stheorem that can be adapted to the discrete case. Regarding Part 2, we prove that Bourgain’spolynomial pointwise ergodic theorem has not a topological version by constructing Z d -actionswith prescribed tolopogical and ergodic properties. In the next sections <strong>of</strong> this chapter, we willgive a brief description <strong>of</strong> the ideas involved.1.2 Chapter 2Let m denote the Lebesgue measure <strong>of</strong> R. It is a classical result in Geometric Measure Theorythe followingTheorem 1.2.1 (Marstrand [27]). If K ⊂ R 2 is a Borel set with Hausdorff dimension greaterthan one, then the projection <strong>of</strong> K into R in the direction <strong>of</strong> λ has positive Lebesgue measurefor m-almost every λ ∈ R.In particular, if K = K 1 × K 2 is a cartesian product <strong>of</strong> two one-dimensional sub<strong>sets</strong> <strong>of</strong>R, Marstrand’s theorem translates to “m(K 1 + λK 2 ) > 0 for m-almost every λ ∈ R”. Theinvestigation <strong>of</strong> such arithmetic sums K 1 + λK 2 has been an active area <strong>of</strong> Mathematics, inspecial when K 1 and K 2 are dynamically defined Cantor <strong>sets</strong>. Remarkably, M. Hall [18] proved,in 1947, that the Lagrange spectrum 1 contains a whole half line, by showing that the arithmeticsum K(4)+K(4) <strong>of</strong> a certain Cantor set K(4) ⊂ R with itself contains [6,∞).Marstrand’s Theorem for product <strong>of</strong> Cantor <strong>sets</strong> is also fundamental in certain results <strong>of</strong>dynamical bifurcations, namely homoclinic bifurcations in surfaces. For instance, in [38] it isused to prove that hyperbolicity holds for most small positive values <strong>of</strong> parameter in homoclinicbifurcations associated to horseshoes with Hausdorff dimension smaller than one; in [39] itis used to show that hyperbolicity is not prevalent in homoclinic bifurcations associated tohorseshoes with Hausdorff dimension larger than one.Looking for some kind <strong>of</strong> converse to Marstrand’s theorem, J. Palis conjectured in [36] thatfor generic pairs <strong>of</strong> regular Cantor <strong>sets</strong> (K 1 ,K 2 ) <strong>of</strong> the real line either K 1 +K 2 has <strong>zero</strong> measureor else it contains an interval. Observe that the first part is a consequence <strong>of</strong> Marstrand’stheorem. In [33], it is proved that stable intersections <strong>of</strong> regular Cantor <strong>sets</strong> are dense in theregion where the sum <strong>of</strong> their Hausdorff dimensions is larger than one, thus establishing theremaining <strong>of</strong> Palis conjecture. It is remarkable to point out that Marstrand’s theorem is usedin the pro<strong>of</strong>.1 The Lagrange spectrum is the set <strong>of</strong> best constants <strong>of</strong> rational approximations <strong>of</strong> irrational numbers. See [9]for the specific description.


5In the connection <strong>of</strong> these two applications, I would like to point out that a formula for theHausdorff dimension <strong>of</strong> K 1 +K 2 , under mild assumptions <strong>of</strong> non-linear Cantor <strong>sets</strong> K 1 and K 2 ,has been obtained by Gugu [31] and applied in [32] to prove that the Hausdorff dimension <strong>of</strong>the Lagrange spectrum increases continuously. In parallel to this non-linear setup, Y. Peres andP. Shmerkin proved the same phenomena happen to self-similar Cantor <strong>sets</strong> without algebraicresonance [41]. Finally, M. Hochman and P. Shmerkin extended and unified many resultsconcerning projections <strong>of</strong> products <strong>of</strong> self-similar measures on regular Cantor <strong>sets</strong> [19].Thetechniques used in[31] werebased onestimates obtained in [33]. With these tools, Guguand myself were able to develop in [25] a purely combinatorial pro<strong>of</strong> <strong>of</strong> Marstrand’s theorem forproducts <strong>of</strong> regular dynamically defined Cantor <strong>sets</strong>. Denoting by HD the Hausdorff dimension,the main theorem isTheorem 1.2.2 (Lima-Moreira [25]). If K 1 ,K 2 are regular dynamically defined Cantor <strong>sets</strong> <strong>of</strong>class C 1+α , α > 0, such that d = HD(K 1 )+HD(K 2 ) > 1, then m(K 1 +λK 2 ) > 0 for m-almostevery λ ∈ R.Let m d denote the Hausdorff d-measure on R. The idea <strong>of</strong> the pro<strong>of</strong> relies on the fact thata regular Cantor set <strong>of</strong> Hausdorff dimension d is regular in the sense that the m d -measure <strong>of</strong>arbitrarily small portions <strong>of</strong> it can be uniformly controlled. Once this is true, a double countingargument proves that the arithmetic sum has Lebesgue measure bounded away from <strong>zero</strong> atadvanced stages <strong>of</strong> the construction <strong>of</strong> the Cantor <strong>sets</strong>, except for a small set <strong>of</strong> parameters.The good feature <strong>of</strong> the pro<strong>of</strong> is that this discretization idea may be applied to the discretecontext, as we shall see in Chapter 4.The same idea <strong>of</strong> homogeneity is applied in [26] to prove the general case <strong>of</strong> Marstrand’stheorem. Namely, once some upper regularity on the m d -measure <strong>of</strong> K is assumed, it is possibleto apply a weighted version <strong>of</strong> [25] (the covers <strong>of</strong> K may be composed <strong>of</strong> sub<strong>sets</strong> with differentdiameter scales) to conclude the result.1.3 Chapter 3Chapter 3 extends the techniques developed in Chapter 2 and encloses a combinatorial pro<strong>of</strong><strong>of</strong> Marstrand’s Theorem without any restrictions on K. The pro<strong>of</strong> makes a study on the onedimensionalfibers <strong>of</strong> K in every direction and relies on two facts:(I) Transversality condition: given two squares on the plane, the Lebesgue measure <strong>of</strong> the set<strong>of</strong> angles θ ∈ R for which their projections in the direction <strong>of</strong> θ have nonempty intersection hasan upper bound.(II) After a regularization <strong>of</strong> K, (I) enables us to conclude that, except for a small set <strong>of</strong> anglesθ ∈ R, the fibers in the direction <strong>of</strong> θ are not concentrated in a thin region. As a consequence,


6K projects into a set <strong>of</strong> positive Lebesgue measure.The idea <strong>of</strong> (II) is based on [31] and was employed in [26] to develop a combinatorial pro<strong>of</strong><strong>of</strong> Marstrand’s Theorem when K is the product <strong>of</strong> two regular Cantor <strong>sets</strong> (which is the content<strong>of</strong> Chapter 2).Compared to other pro<strong>of</strong>s <strong>of</strong> Marstrand’s Theorem, the new ideas here are the discretization<strong>of</strong> the argument and the use <strong>of</strong> dyadic covers, which allow the simplification <strong>of</strong> the methodemployed.1.4 Chapter 4Given a subset E ⊂ Z, let d ∗ (E) denote the upper-Banach <strong>density</strong> <strong>of</strong> E, defined asd ∗ |E ∩{M +1,...,N}|(E) = limsup·N−M→∞ N −MSzemerédi’s theorem asserts that if d ∗ (E) > 0, then E contains arbitrarily long arithmeticprogressions. Onecan interpretthisresultbysayingthat <strong>density</strong>representsthecorrect notion <strong>of</strong>largeness needed to preserve finite configurations <strong>of</strong> Z. On the other hand, Szemerédi’s theoremcannot infer any property about sub<strong>sets</strong> <strong>of</strong> <strong>zero</strong> <strong>density</strong>, and many <strong>of</strong> these are important.A class <strong>of</strong> examples are the integer polynomial sequences. The prime numbers are anotherexample. These <strong>sets</strong> may, as well, contain combinatorially rich patterns.A set E ⊂ Z <strong>of</strong> <strong>zero</strong> upper-Banach <strong>density</strong> is characterized as occupying portions in intervals<strong>of</strong> Z that grow in a sublinear fashion as the length <strong>of</strong> the intervals grow. On the other hand,there still may exist some kind <strong>of</strong> growth speed. For example, the number <strong>of</strong> perfect squareson an interval <strong>of</strong> length n is about n 0.5 . This exponent means, in some sense, a dimension <strong>of</strong>{n 2 ; n ∈ Z} inside Z. In general, define the counting dimension <strong>of</strong> E ⊂ Z asD(E) = limsup|I|→∞log|E ∩I|log|I|,where I runs over the intervals <strong>of</strong> Z.In [24], Gugu and myself investigated this dimension when one considers arithmetic sums<strong>of</strong> the form E +⌊λF⌋, λ ∈ R, for a class <strong>of</strong> sub<strong>sets</strong> E,F ⊂ Z. At first sight, by an ingenuousapplication <strong>of</strong> the multiplicative principle, one should expect that E +⌊λF⌋ has dimension atleast D(E)+D(F). This is not true at all, due to two reasons:(1) E and F may be very irregular as sub<strong>sets</strong> <strong>of</strong> Z.(2) E and F may exhibit growth speed inside intervals <strong>of</strong> very different lengths.


7This led to the definition <strong>of</strong> regular compatible <strong>sets</strong>: E is regular if|E ∩I| 1,|I|D(E)where I runs over all intervals <strong>of</strong> Z, and there exists a sequence <strong>of</strong> intervals (I n ) n≥1 such that|I n | → ∞ and|E ∩I n | 1. (1.4.1)|I n | D(E)Two regular sub<strong>sets</strong> E,F ⊂ Z are compatible if there exist two sequences (I n ) n≥1 , (J n ) n≥1<strong>of</strong> intervals with increasing lengths satisfying (1.4.1) and such that I n ∼ J n (above, we usedasymptotic Vinogradov notations and ∼). The main theorem <strong>of</strong> [24] isTheorem 1.4.1 (Lima-Moreira [24]). Let E,F ⊂ Z be two regular compatible <strong>sets</strong>. ThenD(E +⌊λF⌋) ≥ min{1,D(E) +D(F)}for Lebesgue almost every λ ∈ R. If in addition D(E)+D(F) > 1, then E +⌊λF⌋ has positiveupper-Banach <strong>density</strong> for Lebesgue almost every λ ∈ R.The above theorem has direct consequences when applied to integer polynomial sequences.For any p(x) ∈ Z[x] <strong>of</strong> degree d, the set E = {p(n); n ∈ Z} is regular and has dimension 1/d.Also, it is compatible with any other regular set. These observations imply the second result <strong>of</strong>the chapter.Theorem 1.4.2 (Lima-Moreira [24]). Let p i ∈ Z[x] with degree d i and let E i = {p i (n); n ∈ Z},i = 0,...,k. If1d 0+ 1 d 1+···+ 1 d k> 1,then the arithmetic sum E 0 + ⌊λ 1 E 1 ⌋ + ··· + ⌊λ k E k ⌋ has positive upper-Banach <strong>density</strong> forLebesgue almost every (λ 1 ,...,λ k ) ∈ R k .1.5 Chapter 5LetT beamesure-preservingtransformationon theprobability space(X,B,µ) andf : X → Rameasurable function. A successful area in ergodic theory deals with the convergence <strong>of</strong> averagesn −1 ·∑nk=1 f ( T k x ) , x ∈ X, when n converges to infinity. The well known Birkh<strong>of</strong>f’s Theoremstates that such limit exists for almost every x ∈ X whenever f is an L 1 -function. Severalresults have been (and still are being) proved when, instead <strong>of</strong> {1,2,...,n}, average is madealong other sequences <strong>of</strong> natural numbers. A remarkable result on this direction was given by J.Bourgain [6], where he proved that if p(x) is a polynomial with integer coefficients and f is an


L p -function, for some p > 1, then the averages n −1·∑nk=1 f ( T p(k) x ) converge for almost everyx ∈ X. In other words, convergence fails to hold for a negligible set with respect to the measureµ. We mention this result is not true for L 1 -functions, as recently proved by Z. Buczolich andD. Mauldin [7].In [3], V. Bergelson asked if this set is also negligible from the topological point <strong>of</strong> view.It turned out, by a result <strong>of</strong> R. Pavlov [40], that this is not true. He proved that, for everysequence (p n ) n≥1 ⊂ Z <strong>of</strong> <strong>zero</strong> upper-Banach <strong>density</strong>, there exist a totally minimal, totallyuniquely ergodic and topologically mixing transformation (X,T) and a continuous functionf : X → R such that n −1 · ∑nk=1 f (Tp kx) fails to converge for a residual set <strong>of</strong> x ∈ X. Theconditions <strong>of</strong> minimality, unique ergodicity, although they seem artificial, are natural for thequestion posed by Bergelson, in order to avoid pathological counterexamples. The construction<strong>of</strong> Pavlov was inspired in results <strong>of</strong> Hahn and Katznelson [17].In [23], I extended the results <strong>of</strong> Pavlov [40] to the setting <strong>of</strong> Z d -actions. The methodconsisted <strong>of</strong> constructing (totally) minimal and (totally) uniquely ergodic Z d -actions. Theprogram was carried out by constructing closed shift invariant sub<strong>sets</strong> <strong>of</strong> a sequence space.More specifically, a sequence <strong>of</strong> finite configurations (C k ) k≥1 <strong>of</strong> {0,1} Zd was built, where C k+1is essentially formed by the concatenation <strong>of</strong> elements in C k such that each <strong>of</strong> them occursstatistically well-behaved in each element <strong>of</strong> C k+1 . Finally, the set <strong>of</strong> limits <strong>of</strong> shifted C k -configurations as k → ∞ constitutes the shift invariant subset. The main results on [23] isTheorem 1.5.1 (Lima [23]). Given a set P ⊂ Z d <strong>of</strong> <strong>zero</strong> upper-Banach <strong>density</strong>, there exist atotally strictly ergodic Z d -action (X,A,µ,T) and a continuous function f : X → R such thatthe ergodic averages1|P ∩(−n,n) d |fail to converge for a residual set <strong>of</strong> x ∈ X.∑g∈P∩(−n,n) d f (T g x)Yet in the arithmetic setup, let p 1 ,...,p d ∈ Z[x]. Bergelson and Leibman proved in[4] that if T is a minimal Z d -action, then there is a residual set Y ⊂ X for which x ∈{T (p 1(n),...,p d (n)) x; n ∈ Z}, for every x ∈ Y. The same method employed in Theorem 1.5.1also gives the best topological result one can expect in this context.Theorem 1.5.2 (Lima [23]). Given a set P ⊂ Z d <strong>of</strong> <strong>zero</strong> upper-Banach <strong>density</strong>, there existsa totally strictly ergodic Z d -action (X,A,µ,T) and an uncountable set Y ⊂ X for which x ∉{T p x; p ∈ P}, for every x ∈ Y.8


Chapter 2A combinatorial pro<strong>of</strong> <strong>of</strong>Marstrand’s Theorem for products<strong>of</strong> regular Cantor <strong>sets</strong>In a paper from 1954 Marstrand proved that if K ⊂ R 2 has Hausdorff dimension greater than1, then its one-dimensional projection has positive Lebesgue measure for almost-all directions.In this chapter, we give a combinatorial pro<strong>of</strong> <strong>of</strong> this theorem when K is the product <strong>of</strong> regularCantor <strong>sets</strong> <strong>of</strong> class C 1+α , α > 0, for which the sum <strong>of</strong> their Hausdorff dimension is greaterthan 1.2.1 IntroductionIf U is a subset <strong>of</strong> R n , the diameter <strong>of</strong> U is |U| = sup{|x−y|;x,y ∈ U} and, if U is a family <strong>of</strong>sub<strong>sets</strong> <strong>of</strong> R n , the diameter <strong>of</strong> U is defined as‖U‖ = sup |U|.U∈UGiven d > 0, the Hausdorff d-measure <strong>of</strong> a set K ⊆ R n ism d (K) = lim(∑infε→0 U covers K‖U‖ d 0 . We define the Hausdorff dimension <strong>of</strong> K as HD(K) = d 0 .Also, for each θ ∈ R, let v θ = (cosθ,sinθ), L θ the line in R 2 through the origin containing v θ9.


Chapter 2. Marstrand’s Theorem for products <strong>of</strong> Cantor <strong>sets</strong> 10and proj θ : R 2 → L θ the orthogonal projection. From now on, we’ll restrict θ to the interval[−π/2,π/2], because L θ = L θ+π .In 1954, J. M. Marstrand [27] proved the following result on the fractal dimension <strong>of</strong> plane<strong>sets</strong>.Theorem. If K ⊆ R 2 is a Borel set such that HD(K) > 1, then m(proj θ (K)) > 0 for m-almostevery θ ∈ R.The pro<strong>of</strong> is based on a qualitative characterization <strong>of</strong> the “bad” angles θ for which theresult is not true. Specifically, Marstrand exhibits a Borel measurable function f(x,θ), (x,θ) ∈R 2 ×[−π/2,π/2], such that f(x,θ) = ∞ for m d -almost every x ∈ K, for every “bad” angle. Inparticular,∫f(x,θ)dm d (x) = ∞. (2.1.1)On the other hand, using a version <strong>of</strong> Fubini’s Theorem, he proves that∫ π/2 ∫dθ f(x,θ)dm d (x) = 0K−π/2Kwhich, in view <strong>of</strong> (2.1.1), implies thatm({θ ∈ [−π/2,π/2]; m(proj θ (K)) = 0}) = 0.These results are based on the analysis <strong>of</strong> rectangular densities <strong>of</strong> points.Many generalizations and simpler pro<strong>of</strong>s have appeared since. One <strong>of</strong> them appeared in1968 by R. Kaufman who gave a very short pro<strong>of</strong> <strong>of</strong> Marstrand’s theorem using methods <strong>of</strong>potential theory. See [21] for his original pro<strong>of</strong> and [29], [37] for further discussion.In this chapter, we prove a particular case <strong>of</strong> Marstrand’s Theorem.Theorem 2.1.1. If K 1 ,K 2 are regular Cantor <strong>sets</strong> <strong>of</strong> class C 1+α , α > 0, such that d =HD(K 1 )+HD(K 2 ) > 1, then m(proj θ (K 1 ×K 2 )) > 0 for m-almost every θ ∈ R.The argument also works to show that the push-forward measure <strong>of</strong> the restriction <strong>of</strong> m d toK 1 ×K 2 , defined as µ θ = (proj θ ) ∗ (m d | K1 ×K 2), is absolutely continuous with respect to m, form-almost every θ ∈ R. Denoting its Radon-Nykodim derivative by χ θ = dµ θ /dm, we also giveyet another pro<strong>of</strong> <strong>of</strong> the following result.Theorem 2.1.2. χ θ is an L 2 function for m-almost every θ ∈ R.Our pro<strong>of</strong> makes a study on the fibers proj −1 θ (v)∩(K 1 ×K 2 ), (θ,v) ∈ R×L θ , and relieson two facts:(I) A regular Cantor set <strong>of</strong> Hausdorff dimension d is regular in the sense that the m d -measure<strong>of</strong> small portions <strong>of</strong> it has the same exponential behavior.


Chapter 2. Marstrand’s Theorem for products <strong>of</strong> Cantor <strong>sets</strong> 11(II) This enables us to conclude that, except for a small set <strong>of</strong> angles θ ∈ R, the fibersproj θ −1 (v) ∩ (K 1 × K 2 ) are not concentrated in a thin region. As a consequence, K 1 × K 2projects into a set <strong>of</strong> positive Lebesgue measure.The idea <strong>of</strong> (II) is based on [31]. It proves that, if K 1 and K 2 are regular Cantor <strong>sets</strong> <strong>of</strong> classC 1+α , α > 0, and at least one <strong>of</strong> them is non-essentially affine (a technical condition), then thearithmetic sum K 1 +K 2 = {x 1 +x 2 ;x 1 ∈ K 1 ,x 2 ∈ K 2 } has the expected Hausdorff dimension:HD(K 1 +K 2 ) = min{1,HD(K 1 )+HD(K 2 )}.Marstrand’s Theorem for products <strong>of</strong> Cantor <strong>sets</strong> is a source <strong>of</strong> very relevant applicationsin dynamical systems, as pointed out in Chapter 1. It is <strong>of</strong> fundamental importance in certainresults <strong>of</strong> dynamical bifurcations, namely homoclinic bifurcations. For instance, in [38] it isused to prove that hyperbolicity holds for most small positive values <strong>of</strong> parameter in homoclinicbifurcationsassociatedtohorseshoeswithHausdorffdimensionsmallerthanone; in[39]itisusedto show that hyperbolicity is not prevalent in homoclinic bifurcations associated to horseshoeswith Hausdorff dimension larger than one; in [33] it is used to prove that stable intersections<strong>of</strong> regular Cantor <strong>sets</strong> are dense in the region where the sum <strong>of</strong> their Hausdorff dimensions islarger than one; in [34] to show that, for homoclinic bifurcations associated to horseshoes withHausdorff dimension larger than one, typically there are open <strong>sets</strong> <strong>of</strong> parameters with positiveLebesgue <strong>density</strong> at the initial bifurcation parameter corresponding to persistent homoclinictangencies.2.2 Regular Cantor <strong>sets</strong> <strong>of</strong> class C 1+αWe say that K ⊂ R is a regular Cantor set <strong>of</strong> class C 1+α , α > 0, if:(i) there are disjoint compact intervals I 1 ,I 2 ,...,I r ⊆ [0,1] such that K ⊂ I 1 ∪···∪I r andthe boundary <strong>of</strong> each I i is contained in K;(ii) there is a C 1+α expanding map ψ defined in a neighbourhood <strong>of</strong> I 1 ∪ I 2 ∪ ··· ∪ I r suchthat ψ(I i ) is the convex hull <strong>of</strong> a finite union <strong>of</strong> some intervals I j , satisfying:by(ii.1) for each i ∈ {1,2,...,r} and n sufficiently big, ψ n (K ∩I i ) = K;(ii.2) K = ⋂ ψ −n (I 1 ∪I 2 ∪···∪I r ).n∈NThe set {I 1 ,...,I r } is called a Markov partition <strong>of</strong> K. It defines an r ×r matrix B = (b ij )b ij = 1, if ψ(I i ) ⊇ I j= 0, if ψ(I i )∩I j = ∅,


Chapter 2. Marstrand’s Theorem for products <strong>of</strong> Cantor <strong>sets</strong> 12which encodes the combinatorial properties <strong>of</strong> K. Given such matrix, consider the set Σ B ={θ = (θ1 ,θ 2 ,...) ∈ {1,...,r} N ; b θi θ i+1= 1,∀i ≥ 1 } and the shift transformation σ : Σ B → Σ Bgiven by σ(θ 1 ,θ 2 ,...) = (θ 2 ,θ 3 ,...).There is a natural homeomorphism between the pairs (K,ψ) and (Σ B ,σ). For each finiteword a = (a 1 ,...,a n ) such that b ai a i+1= 1, i = 1,...,n−1, the intersectionI a = I a1 ∩ψ −1 (I a2 )∩···∩ψ −(n−1) (I an )is a non-empty interval with diameter |I a | = |I an |/|(ψ n−1 ) ′ (x)| for some x ∈ I a , which isexponentially small if n is large. Then, {h(θ)} = ⋂ n≥1 I (θ 1 ,...,θ n) defines a homeomorphismh : Σ B → K that commutes the diagramΣ Bσ Σ Bh h K ψKIf λ = sup{|ψ ′ (x)|;x ∈ I 1 ∪··· ∪I r } ∈ (1,∞), then ∣ ∣I (θ1∣,...,θ n+1 ) ≥ λ ·∣∣I −1 (θ1∣,...,θ n) and so, forρ > 0 small and θ ∈ Σ B , there is a positive integer n = n(ρ,θ) such thatρ ≤ ∣ I(θ1 ∣,...,θ n) ≤ λρ.Definition 2.2.1. A ρ-decomposition <strong>of</strong> K is any finite set (K) ρ = {I 1 ,I 2 ,...,I r } <strong>of</strong> disjointclosed intervals <strong>of</strong> R, each one <strong>of</strong> them intersecting K, whose union covers K and such thatρ ≤ |I i | ≤ λρ, i = 1,2,...,r.Remark 2.2.2. Although ρ-decompositions are not unique, we use, for simplicity, the notation(K) ρ todenoteany<strong>of</strong>them. Wealsousethesamenotation(K) ρ todenotetheset ⋃ I∈(K) ρI ⊂ Rand the distinction between these two situations will be clear throughout the text.Every regular Cantor set{<strong>of</strong> class C 1+α } has a ρ-decomposition for ρ > 0 small: by thecompactness <strong>of</strong> K, the family I (θ1 ,...,θ n(ρ,θ))has a finite cover (in fact, it is only necessaryθ∈Σ Bfor ψ to be <strong>of</strong> class C 1 ). Also, one can define ρ-decomposition for the product <strong>of</strong> two Cantor<strong>sets</strong> K 1 and K 2 , denoted by (K 1 × K 2 ) ρ . Given ρ ≠ ρ ′ and two decompositions (K 1 × K 2 ) ρ ′and (K 1 ×K 2 ) ρ , consider the partial order(K 1 ×K 2 ) ρ ′ ≺ (K 1 ×K 2 ) ρ ⇐⇒ ρ ′ < ρ and⋃Q ′ ∈(K 1 ×K 2 ) ρ ′In this case, proj θ ((K 1 ×K 2 ) ρ ′) ⊆ proj θ ((K 1 ×K 2 ) ρ ) for any θ.Q ′ ⊆⋃Q∈(K 1 ×K 2 ) ρQ.A remarkable property <strong>of</strong> regular Cantor <strong>sets</strong> <strong>of</strong> class C 1+α , α > 0, is bounded distortion.


Chapter 2. Marstrand’s Theorem for products <strong>of</strong> Cantor <strong>sets</strong> 13Lemma 2.2.3. Let (K,ψ) be a regular Cantor set <strong>of</strong> class C 1+α , α > 0, and {I 1 ,...,I r } aMarkov partition. Given δ > 0, there exists a constant C(δ) > 0, decreasing on δ, with thefollowing property: if x,y ∈ K satisfy(i) |ψ n (x)−ψ n (y)| < δ;(ii) The interval [ψ i (x),ψ i (y)] is contained in I 1 ∪···∪I r , for i = 0,...,n−1,thenIn addition, C(δ) → 0 as δ → 0.e −C(δ) ≤ |(ψn ) ′ (x)||(ψ n ) ′ (y)| ≤ eC(δ) .A direct consequence <strong>of</strong> bounded distortion is the required regularity <strong>of</strong> K, contained in thenext result.Lemma 2.2.4. Let K be a regular Cantor set <strong>of</strong> class C 1+α , α > 0, and let d = HD(K). Then0 < m d (K) < ∞. Moreover, there is c > 0 such that, for any x ∈ K and 0 ≤ r ≤ 1,c −1 ·r d ≤ m d (K ∩B r (x)) ≤ c·r d .ThesamehappensforproductsK 1 ×K 2 <strong>of</strong> Cantor <strong>sets</strong> (without loss<strong>of</strong> generality, consideredwith the box norm).Lemma 2.2.5. Let K 1 ,K 2 be regular Cantor <strong>sets</strong> <strong>of</strong> class C 1+α , α > 0, and let d = HD(K 1 )+HD(K 2 ). Then 0 < m d (K 1 ×K 2 ) < ∞. Moreover, there is c 1 > 0 such that, for any x ∈ K 1 ×K 2and 0 ≤ r ≤ 1,c 1−1 ·r d ≤ m d ((K 1 ×K 2 )∩B r (x)) ≤ c 1 ·r d .See chapter 4 <strong>of</strong> [37] for the pro<strong>of</strong>s <strong>of</strong> these lemmas. In particular, if Q ∈ (K 1 ×K 2 ) ρ , thereis x ∈ (K 1 ∪K 2 )∩Q such that B λ −1 ρ(x) ⊆ Q ⊆ B λρ (x) and so(c 1 λ d) −1·ρ d ≤ m d ((K 1 ×K 2 )∩Q) ≤ c 1 λ d ·ρ d .Changing c 1 by c 1 λ d , we may also assume thatc 1−1 ·ρ d ≤ m d ((K 1 ×K 2 )∩Q) ≤ c 1 ·ρ d ,which allows us to obtain estimates on the cardinality <strong>of</strong> ρ-decompositions.Lemma 2.2.6. Let K 1 ,K 2 be regular Cantor <strong>sets</strong> <strong>of</strong> class C 1+α , α > 0, and let d = HD(K 1 )+HD(K 2 ). Then there is c 2 > 0 such that, for any ρ-decomposition (K 1 × K 2 ) ρ , x ∈ K 1 × K 2and 0 ≤ r ≤ 1,( ) r#{Q ∈ (K 1 ×K 2 ) ρ ;Q ⊆ B r (x)} ≤ c 2 ·d·ρIn addition, c 2−1 ·ρ −d ≤ #(K 1 ×K 2 ) ρ ≤ c 2 ·ρ −d .


Chapter 2. Marstrand’s Theorem for products <strong>of</strong> Cantor <strong>sets</strong> 14Pro<strong>of</strong>. We havec 1 ·r d ≥ m d ((K 1 ×K 2 )∩B r (x))∑≥ m d ((K 1 ×K 2 )∩Q)≥Q⊆B r(x)∑Q⊆B r(x)c 1−1 ·ρ d= #{Q ∈ (K 1 ×K 2 ) ρ ;Q ⊆ B r (x)}·c 1−1 ·ρ dand then( ) r#{Q ∈ (K 1 ×K 2 ) ρ ;Q ⊆ B r (x)} ≤ c 12 ·d·ρOn the other hand,m d (K 1 ×K 2 ) =∑∑m d ((K 1 ×K 2 )∩Q) ≤ c 1 ·ρ d ,Q∈(K 1 ×K 2 ) ρ Q∈(K 1 ×K 2 ) ρimplying that#(K 1 ×K 2 ) ρ ≥ c 1−1 ·m d (K 1 ×K 2 )·ρ −d .Taking c 2 = max{c 1 2 , c 1 /m d (K 1 ×K 2 )}, we conclude the pro<strong>of</strong>.2.3 Pro<strong>of</strong> <strong>of</strong> Theorem 2.1.1Given rectangles Q and ˜Q, let}Θ = Q,˜Q{θ ∈ [−π/2,π/2];proj θ (Q)∩proj θ (˜Q) ≠ ∅ .Lemma 2.3.1. If Q, ˜Q ∈ (K 1 ×K 2 ) ρ and x ∈ (K 1 ×K 2 )∩Q,˜x ∈ (K 1 ×K 2 )∩ ˜Q, then( )ρm Θ Q,˜Q≤ 2πλ·d(x,˜x) ·Pro<strong>of</strong>. Consider the figure.xθ |θ −ϕ 0|˜xL θproj θ (˜x)θproj θ (x)


Chapter 2. Marstrand’s Theorem for products <strong>of</strong> Cantor <strong>sets</strong> 15Since proj θ (Q) has diameter at most λρ, d(proj θ (x),proj θ (˜x)) ≤ 2λρ and then, by elementarygeometry,sin(|θ −ϕ 0 |) = d(proj θ(x),proj θ (˜x))d(x,˜x)ρ≤ 2λ·d(x,˜x)ρ=⇒ |θ −ϕ 0 | ≤ πλ·d(x,˜x) ,because sin −1 y ≤ πy/2. As ϕ 0 is fixed, the lemma is proved.We point out that, although simple, Lemma 2.3.1 expresses the crucial property <strong>of</strong> transversalitythat makes the pro<strong>of</strong> work, and all results related to Marstrand’s theorem use a similaridea in one way or another. See [43] where this tranversality condition is also exploited.Fixed a ρ-decomposition (K 1 ×K 2 ) ρ , let{N (K1 ×K 2 ) ρ(θ) = # (Q, ˜Q)}∈ (K 1 ×K 2 ) ρ ×(K 1 ×K 2 ) ρ ;proj θ (Q)∩proj θ (˜Q) ≠ ∅for each θ ∈ [−π/2,π/2] andE((K 1 ×K 2 ) ρ ) =∫ π/2−π/2N (K1 ×K 2 ) ρ(θ)dθ.Proposition 2.3.2. Let K 1 ,K 2 be regular Cantor <strong>sets</strong> <strong>of</strong> class C 1+α , α > 0, and let d =HD(K 1 )+HD(K 2 ). Then there is c 3 > 0 such that, for any ρ-decomposition (K 1 ×K 2 ) ρ ,E((K 1 ×K 2 ) ρ ) ≤ c 3 ·ρ 1−2d .Pro<strong>of</strong>. Let s 0 = ⌈ log 2 ρ −1⌉ and choose, for each Q ∈ (K 1 ×K 2 ) ρ , a point x ∈ (K 1 ×K 2 )∩Q.By a double counting and using Lemmas 2.2.6 and 2.3.1, we have∑ ( )E((K 1 ×K 2 ) ρ ) = Θ Q,˜Q=Q,˜Q∈(K 1 ×K 2 ) ρms 0 ∑s=1∑mQ, ˜Q∈(K 1 ×K 2 )ρ2 −s 1, c 3 = 2 d+1 πλc 22 ·∑s≥1 2s(1−d) < +∞ satisfies the required inequality.


Chapter 2. Marstrand’s Theorem for products <strong>of</strong> Cantor <strong>sets</strong> 16This implies that, for each ε > 0, the upper boundN (K1 ×K 2 ) ρ(θ) ≤ c 3 ·ρ 1−2d(2.3.1)εholds for every θ except for a set <strong>of</strong> measure at most ε. Letting c 4 = c−2 2 ·c −1 3 , we will showthatm(proj θ ((K 1 ×K 2 ) ρ )) ≥ c 4 ·ε (2.3.2)for every θ satisfying (2.3.1). For this, divide [−2,2] ⊆ L θ in ⌊4/ρ⌋ intervals J ρ 1 ,..., Jρ ⌊4/ρ⌋ <strong>of</strong>equal lenght (at least ρ) and defines ρ,i = #{Q ∈ (K 1 ×K 2 ) ρ ; proj θ (x) ∈ J ρ i}, i = 1,...,⌊4/ρ⌋.Then ∑ ⌊4/ρ⌋i=1s ρ,i = #(K 1 ×K 2 ) ρ and⌊4/ρ⌋∑i=1s ρ,i 2 ≤ N (K1 ×K 2 ) ρ(θ) ≤ c 3 ·ρ 1−2d ·ε −1 .Let S ρ = {1 ≤ i ≤ ⌊4/ρ⌋;s ρ,i > 0}. By Cauchy-Schwarz inequality,⎛ ⎞⎝ ∑ 2s ρ,i⎠i∈S ρ#S ρ ≥∑i∈S ρs ρ,i2≥ c 2 −2 ·ρ −2dc 3 ·ρ 1−2d ·ε −1 = c 4 ·ε·ρFor each i ∈ S ρ , the interval J ρ iis contained in proj θ ((K 1 ×K 2 ) ρ ) and thenwhich proves (2.3.2).m(proj θ ((K 1 ×K 2 ) ρ )) ≥ c 4 ·ε,Pro<strong>of</strong> <strong>of</strong> Theorem 2.1.1. Fix a decreasing sequence(K 1 ×K 2 ) ρ1 ≻ (K 1 ×K 2 ) ρ2 ≻ ··· (2.3.3)<strong>of</strong> decompositions such that ρ n → 0 and, for each ε > 0, consider the <strong>sets</strong>{G n 1−2d ε = θ ∈ [−π/2,π/2]; N (K1 ×K 2 ) ρn(θ) ≤ c 3 ·ρ n ·ε −1} , n ≥ 1.Then m([−π/2,π/2]\G n ε) ≤ ε, and the same holds for the setG ε = ⋂ ∞⋃G l ε.n≥1If θ ∈ G ε , thenl=nm(proj θ ((K 1 ×K 2 ) ρn )) ≥ c 4 ·ε, for infinitely many n,which implies that m(proj θ (K 1 ×K 2 )) ≥ c 4 · ε. Finally, the set G = ⋃ n≥1 G 1/n satisfiesm([−π/2,π/2]\G) = 0 and m(proj θ (K 1 ×K 2 )) > 0, for any θ ∈ G.


Chapter 2. Marstrand’s Theorem for products <strong>of</strong> Cantor <strong>sets</strong> 172.4 Pro<strong>of</strong> <strong>of</strong> Theorem 2.1.2Given any X ⊂ K 1 ×K 2 , let (X) ρ be the restriction <strong>of</strong> the ρ-decomposition (K 1 ×K 2 ) ρ to thoserectangles which intersect X. As done in Section 2.3, we’ll obtain estimates on the cardinality<strong>of</strong> (X) ρ . Being a subset <strong>of</strong> K 1 × K 2 , the upper estimates from Lemma 2.2.6 also hold for X.The lower estimate is given byLemma 2.4.1. Let X be a subset <strong>of</strong> K 1 ×K 2 suchthat m d (X) > 0. Then there is c 6 = c 6 (X) > 0such that, for any ρ-decomposition (K 1 ×K 2 ) ρ and 0 ≤ r ≤ 1,c 6 ·ρ −d ≤ #(X) ρ ≤ c 2 ·ρ −d .Pro<strong>of</strong>. As m d (X) < ∞, there exists c 5 = c 5 (X) > 0 (see Theorem 5.6 <strong>of</strong> [11]) such thatm d (X ∩B r (x)) ≤ c 5 ·r d , for all x ∈ X and 0 ≤ r ≤ 1,and thenm d (X) = ∑Q∈(X) ρm d (X ∩Q) ≤ ∑Q∈(X) ρc 5 ·(λρ) d =(c 5 ·λ d)·ρ d ·#(X) ρ .Just take c 6 = c−1 5 ·λ −d ·m d (X).Proposition 2.4.2. The measure µ θ = (proj θ ) ∗ (m d | K1 ×K 2) is absolutely continuous with respectto m, for m-almost every θ ∈ R.Pro<strong>of</strong>. Note that the implicationX ⊂ K 1 ×K 2 , m d (X) > 0 =⇒ m(proj θ (X)) > 0 (2.4.1)is sufficient for the required absolute continuity. In fact, if Y ⊂ L θ satisfies m(Y) = 0, thenµ θ (Y) = m d (X) = 0,where X = proj −1 θ (Y). Otherwise, by (2.4.1) we would have m(Y) = m(proj θ (X)) > 0,contradicting the assumption.We prove that (2.4.1) holds for every θ ∈ G, where G is the set defined in the pro<strong>of</strong> <strong>of</strong>Theorem 2.1.1. The argument is the same made after Proposition 2.3.2: as, by the previouslemma, #(X) ρ has lower and upper estimates depending only on X and ρ, we obtain that−1 m(proj θ ((X) ρn )) ≥ c 3 ·c 62 ·ε, for infinitely many n,and then m(proj θ (X)) > 0.


Chapter 2. Marstrand’s Theorem for products <strong>of</strong> Cantor <strong>sets</strong> 18Let χ θ = dµ θ /dm. In principle, this is a L 1 function. We prove that it is a L 2 function, forevery θ ∈ G.Pro<strong>of</strong> <strong>of</strong> Theorem 2. Let θ ∈ G 1/m , for some m ∈ N. ThenN (K1 ×K 2 ) ρn(θ) ≤ c 3 ·ρ n1−2d ·m, for infinitely many n. (2.4.2)For each <strong>of</strong> these n, consider the partition P n = {J ρn1 ,...,Jρn ⌊4/ρ n⌋ } <strong>of</strong> [−2,2] ⊂ L θ into intervals<strong>of</strong> equal length and let χ θ,n be the expectation <strong>of</strong> χ θ with respect to P n . As ρ n → 0, thesequence <strong>of</strong> functions (χ θ,n ) n∈N converges pointwise to χ θ . By Fatou’s Lemma, we’re done if weprove that each χ θ,n is L 2 and its L 2 -norm ‖χ θ,n ‖ 2is bounded above by a constant independent<strong>of</strong> n.By definition,µ θ (J ρni) = m d((projθ ) −1 (J ρni) ) ≤ s ρn,i ·c 1 ·ρ n d , i = 1,2,...,⌊4/ρ n ⌋,and thenimplying thati)|J ρni|χ θ,n (x) = µ θ(J ρn‖χ θ,n ‖ 2 2==≤∫≤ c 1 ·s ρn,i ·ρ nd|J ρni|L θ|χ θ,n | 2 dm⌊4/ρ n⌋∑i=1⌊4/ρ n⌋∑i=1∫J ρni|J ρni|·|χ θ,n | 2 dm⌊4/ρ∑ n⌋2d−1 ≤ c 12 ·ρ n ·, ∀x ∈ J ρni,(c1 ·s ρn,i ·ρ ndi=1|J ρni|s ρn,i 2≤ c 12 ·ρ n2d−1 ·N (K1 ×K 2 ) ρn(θ).) 2In view <strong>of</strong> (2.4.2), this last expression is bounded above by( )·( )2d−1 1−2d c 12 ·ρ n c 3 ·ρ n ·m = c 12 ·c 3 ·m,which is a constant independent <strong>of</strong> n.


Chapter 2. Marstrand’s Theorem for products <strong>of</strong> Cantor <strong>sets</strong> 192.5 Concluding remarksThe pro<strong>of</strong>s <strong>of</strong> Theorems 2.1.1 and 2.1.2 work not just for the case <strong>of</strong> products <strong>of</strong> regular Cantor<strong>sets</strong>, but in greater generality, whenever K ⊂ R 2 is a Borel set for which there is a constantc > 0 such that, for any x ∈ K and 0 ≤ r ≤ 1,c −1 ·r d ≤ m d (K ∩B r (x)) ≤ c·r d ,since this alone implies the existence <strong>of</strong> ρ-decompositions for K.


Chapter 2. Marstrand’s Theorem for products <strong>of</strong> Cantor <strong>sets</strong> 20


Chapter 3. Yet another pro<strong>of</strong> <strong>of</strong> Marstrand’s Theorem 22<strong>sets</strong>.In 1954, J. M. Marstrand [27] proved the following result on the fractal dimension <strong>of</strong> planeTheorem 3.1.1. If K ⊂ R 2 is a Borel set such that HD(K) > 1, then m(proj θ (K)) > 0 form-almost every θ ∈ R.The pro<strong>of</strong> is based on a qualitative characterization <strong>of</strong> the “bad” angles θ for which theresult is not true. Specifically, Marstrand exhibits a Borel measurable function f(x,θ), (x,θ) ∈R 2 ×[−π/2,π/2], such that f(x,θ) = ∞ for m s -almost every x ∈ K, for every “bad” angle. Inparticular,∫Kf(x,θ)dm s (x) = ∞. (3.1.1)On the other hand, using a version <strong>of</strong> Fubini’s Theorem, he proves that∫ π/2−π/2which, in view <strong>of</strong> (3.1.1), implies that∫dθ f(x,θ)dm s (x) = 0Km({θ ∈ [−π/2,π/2]; m(proj θ (K)) = 0}) = 0.These results are based on the analysis <strong>of</strong> rectangular densities <strong>of</strong> points.Many generalizations and simpler pro<strong>of</strong>s have appeared since. One <strong>of</strong> them came in 1968 byR. Kaufman who gave a very short pro<strong>of</strong> <strong>of</strong> Marstrand’s Theorem using methods <strong>of</strong> potentialtheory. See [21] for his original pro<strong>of</strong> and [29], [37] for further discussion.In this chapter, we give a new pro<strong>of</strong> <strong>of</strong> Theorem 3.1.1. Our pro<strong>of</strong> makes a study on thefibers K ∩proj −1 θ (v), (θ,v) ∈ R×L θ , and relies on two facts:(I) Transversality condition: given two squares on the plane, the Lebesgue measure <strong>of</strong> the set<strong>of</strong> angles for which their projections have nonempty intersection has an upper bound. SeeSubsection 3.3.2.(II) After a regularization <strong>of</strong> K, (I) enables us to conclude that, except for a small set <strong>of</strong> anglesθ ∈ R, the fibers K ∩proj −1 θ (v) are not concentrated in a thin region. As a consequence, Kprojects into a set <strong>of</strong> positive Lebesgue measure.The idea <strong>of</strong> (II) is based on [31] used in [25] to develop a combinatorial pro<strong>of</strong> <strong>of</strong> Theorem3.1.1 when K is the product <strong>of</strong> two regular Cantor <strong>sets</strong>. In the present chapter, we give acombinatorial pro<strong>of</strong> <strong>of</strong> Theorem 3.1.1 without any restrictions on K. Compared to other pro<strong>of</strong>s<strong>of</strong> Marstrand’s Theorem, the new ideas here are the discretization <strong>of</strong> the argument and the use<strong>of</strong> dyadic covers, which allow the simplification <strong>of</strong> the method employed. These covers may becomposed <strong>of</strong> <strong>sets</strong> with rather different scales and so a weighted sum is necessary to capture theHausdorff s-measure <strong>of</strong> K.


Chapter 3. Yet another pro<strong>of</strong> <strong>of</strong> Marstrand’s Theorem 23The theory developed in [26] works whenever K is an Ahlfors-David regular set, namelywhen there are constants a,b > 0 such thata·r d ≤ m s (K ∩B r (x)) ≤ b·r d , for any x ∈ K and 0 < r ≤ 1.Unfortunately, the general situation can not be reduced to this one, as proved by P. Mattilaand P. Saaranen: in [30], they constructed a compact set <strong>of</strong> R with positive Lebesgue measuresuch that it contains no nonempty Ahlfors-David subset.We also give yet another pro<strong>of</strong> that the push-forward measure <strong>of</strong> the restriction <strong>of</strong> m s to K,defined as µ θ = (proj θ ) ∗ (m s | K ), is absolutely continuous with respect to m, for m-almost everyθ ∈ R, and its Radon-Nykodim derivative is square-integrable.Theorem 3.1.2. The measure µ θ is absolutely continuous with respect to m and its Radon-Nykodim derivative is an L 2 function, for m-almost every θ ∈ R.Marstrand’s Theorem is a classical result in Geometric Measure Theory. In particular,if K = K 1 × K 2 is a cartesian product <strong>of</strong> two one-dimensional sub<strong>sets</strong> <strong>of</strong> R, Marstrand’stheorem translates to “m(K 1 +λK 2 ) > 0 for m-almost every λ ∈ R”. The investigation <strong>of</strong> sucharithmetic sums K 1 + λK 2 has been an active area <strong>of</strong> Mathematics, in special when K 1 andK 2 are dynamically defined Cantor <strong>sets</strong>. Remarkably, M. Hall [18] proved, in 1947, that theLagrange spectrum 1 contains a whole half line, by showing that the arithmetic sum K(4)+K(4)<strong>of</strong> a certain Cantor set K(4) ⊂ R with itself contains [6,∞).In the connection <strong>of</strong> these two applications, we point out that a formula for the Hausdorffdimension <strong>of</strong> K 1 +K 2 , under mild assumptions <strong>of</strong> non-linear Cantor <strong>sets</strong> K 1 and K 2 , has beenobtainedbythesecondauthorin[31]andappliedin[32]toprovethattheHausdorffdimension<strong>of</strong>the Lagrange spectrum increases continuously. In parallel to this non-linear setup, Y. Peres andP. Shmerkin proved the same phenomena happen to self-similar Cantor <strong>sets</strong> without algebraicresonance [41]. Finally, M. Hochman and P. Shmerkin extended and unified many resultsconcerning projections <strong>of</strong> products <strong>of</strong> self-similar measures on regular Cantor <strong>sets</strong> [19].The content <strong>of</strong> this chapter is organized as follows. In Section 3.2 we introduce the basicnotations and definitions. Section 3.3 is devoted to the main calculations, including thetransversality condition in Subsection 3.3.2 and the pro<strong>of</strong> <strong>of</strong> existence <strong>of</strong> good dyadic covers inSubsection 3.3.3. Finally, in Section 3.4 we prove Theorems 3.1.1 and 3.1.2. We also collectfinal remarks in Section 3.5.1 The Lagrange spectrum is the set <strong>of</strong> best constants <strong>of</strong> rational approximations <strong>of</strong> irrational numbers. See [9]for the specific description.


Chapter 3. Yet another pro<strong>of</strong> <strong>of</strong> Marstrand’s Theorem 243.2 Preliminaries3.2.1 NotationThe distance in R 2 will be denoted by |·|. Let B r (x) denote the open ball <strong>of</strong> R 2 centered in xwith radius r. As in Section 3.1, the diameter <strong>of</strong> U ⊂ R 2 is |U| = sup{|x−y|;x,y ∈ U} and, ifU is a family <strong>of</strong> sub<strong>sets</strong> <strong>of</strong> R 2 , the diameter <strong>of</strong> U is defined as‖U‖ = sup |U|.U∈UGiven s > 0, the Hausdorff s-measure <strong>of</strong> a set K ⊂ R 2 is m s (K) and its Hausdorff dimension isHD(K). In this work, we assume K is contained in [0,1) 2 .Definition 3.2.1. A Borel set K ⊂ R 2 is an s-set if HD(K) = s and 0 < m s (K) < ∞.Let m be the Lebesgue measure <strong>of</strong> Lebesgue measurable <strong>sets</strong> on R. For each θ ∈ R, letv θ = (cosθ,sinθ), L θ the line in R 2 through the origin containing v θ and proj θ : R 2 → L θ theorthogonal projection onto L θ .A square [a,a+l)×[b,b+l) ⊂ R 2 will be denoted by Q and its center, the point (a+l/2,b+l/2), by x.Let X be a set. We use the following notation to compare the asymptotic <strong>of</strong> functions.Definition 3.2.2. Let f,g : X → R be two real-valued functions. We say f g if there is aconstant C > 0 such thatIf f g and g f, we write f ∼ g.|f(x)| ≤ C ·|g(x)|, ∀x ∈ X.3.2.2 Dyadic squaresLet D 0 be the family <strong>of</strong> unity squares <strong>of</strong> R 2 congruent to [0,1) 2 and with vertices in the latticeZ 2 . Dilating this family by a factor <strong>of</strong> 2 −i , we obtain the family D i , i ∈ Z.Definition 3.2.3. Let D denote the union <strong>of</strong> D i , i ∈ Z. A dyadic square is any element Q ∈ D.The dyadic squares possess the following properties:(1) Every x ∈ R 2 belongs to exactly one element <strong>of</strong> each family D i .(2) Two dyadic squares are either disjoint or one is contained in the other.(3) A dyadic square<strong>of</strong> D i is contained in exactly one dyadic square<strong>of</strong> D i−1 and contains exactlyfour dyadic squares <strong>of</strong> D i+1 .


Chapter 3. Yet another pro<strong>of</strong> <strong>of</strong> Marstrand’s Theorem 25(4) Given any subset U ⊂ R 2 , there are four dyadic squares, each with side length at most2·|U|, whose union contains U.(1) to (3) are direct. To prove (4), let R be smallest rectangle <strong>of</strong> R 2 with sides parallelto the axis that contains U. The sides <strong>of</strong> R have length at most |U|. Let i ∈ Z such that2 −i−1 < |U| ≤ 2 −i and choose a dyadic square Q ∈ D i that intersects R. If Q contains U, we’redone. If not, Q and three <strong>of</strong> its neighbors cover U.RQUDefinition 3.2.4. A dyadic cover <strong>of</strong> K is a finite subset C ⊂ D <strong>of</strong> disjoint dyadic squares suchthatK ⊂ ⋃ Q∈CQ.First used by A.S. Besicovitch in his demonstration that closed <strong>sets</strong> <strong>of</strong> infinite m s -measurecontainsub<strong>sets</strong><strong>of</strong>positivebutfinitemeasure[5], dyadiccoverswerelateremployedbyMarstrandto investigate the Hausdorff measure <strong>of</strong> cartesian products <strong>of</strong> <strong>sets</strong> [28].Due to (4), for any family U <strong>of</strong> sub<strong>sets</strong> <strong>of</strong> R 2 , there is a dyadic family C such that⋃Q ⊃ ⋃ ∑U and |Q| s < 64· ∑|U| sQ∈C U∈U Q∈C U∈Uand so, if K is an s-set, there exists a sequence (C i ) i≥1 <strong>of</strong> dyadic covers <strong>of</strong> K with diametersconverging to <strong>zero</strong> such that∑|Q| s ∼ 1. (3.2.1)Q∈C i3.3 CalculationsLet K ⊂ R 2 be a Borel set with Hausdorff dimension greater than one. From now on, weassume every cover <strong>of</strong> K is composed <strong>of</strong> dyadic squares <strong>of</strong> sides at most one. Before going intothe calculations, we make the following reduction.


Chapter 3. Yet another pro<strong>of</strong> <strong>of</strong> Marstrand’s Theorem 26Lemma 3.3.1. Let K be a Borel subset <strong>of</strong> R 2 . Given s < HD(K), there exists a compact s-setK ′ ⊂ K such thatm s (K ′ ∩B r (x)) r d , x ∈ R 2 and 0 ≤ r ≤ 1.In other words, there exists a constant b > 0 such thatm s (K ′ ∩B r (x)) ≤ b·r d , for any x ∈ R 2 and 0 ≤ r ≤ 1. (3.3.1)See Theorem 5.4 <strong>of</strong> [11] for a pro<strong>of</strong> <strong>of</strong> the above lemma when K is closed and [10] for thegeneral case. From now on, we assume K is a compact s-set, with s > 1, that satisfies (3.3.1).Given a dyadic cover C <strong>of</strong> K, let, for each θ ∈ [−π/2,π/2], fθ C : L θ → R be the functiondefined byfθ C (x) = Q∈Cχ ∑ projθ (Q)(x)·|Q| s−1 ,where χ projθ (Q) denotes the characteristic function <strong>of</strong> the set proj θ (Q). The reason we considerthis function is that it captures the Hausdorff s-measure <strong>of</strong> K in the sense that∫fθ C (x)dm(x) = ∑ ∫|Q| s−1 · χ projθ (Q)(x)dm(x)L θ Q∈CL θ= ∑ |Q| s−1 ·m(proj θ (Q))Q∈Cwhich, as |Q|/2 ≤ m(proj θ (Q)) ≤ |Q|, satisfies∫L θf C θ (x)dm(x) ∼ ∑ Q∈CIf in addition C satisfies (3.2.1), then∫(x)dm(x) ∼ 1 , ∀θ ∈ [−π/2,π/2]. (3.3.2)L θf C θDenoting the union ⋃ Q∈CQ by C as well, an application <strong>of</strong> the Cauchy-Schwarz inequality givesthatm(proj θ (C))·( ∫proj θ (C)(fCθ) 2dm)≥|Q| s .( ∫ ) 2fθ C dm ∼ 1.proj θ (C)The above inequality implies that if (C i ) i≥1 is a sequence <strong>of</strong> dyadic covers <strong>of</strong> K satisfying (3.2.1)with diameters converging to <strong>zero</strong> and the L 2 -norm <strong>of</strong> f C iθis uniformly bounded, that is∫ ( ) 2dm 1, (3.3.3)proj θ (C i )f C iθ


Chapter 3. Yet another pro<strong>of</strong> <strong>of</strong> Marstrand’s Theorem 27thenm(proj θ (K)) = limi→∞m(proj θ (C i )) 1and so proj θ (K) has positive Lebesgue measure, as wished. This conclusion will be obtainedfor m-almost every θ ∈ [−π/2,π/2] by showing thatI i. =∫ π/2−π/2(dθ∫L θf C iθ) 2dm 1. (3.3.4)3.3.1 Rewriting the integral I iFor simplicity, let f denote f C iθ. Then the interior integral <strong>of</strong> (3.3.4) becomes∫L θf 2 dm =and, using the inequalitiesit follows that∫⎞ ⎛ ⎞⎝ ∑ χ projθ (Q) ·|Q| s−1 ⎠· ⎝ ∑˜Q∈Ci χ projθ (˜Q) ·|˜Q| s−1 ⎠dmQ∈C iL θ⎛= ∑ ∫|Q| s−1 ·|˜Q| s−1 · χ projθ (Q)∩proj θ (˜Q) dmLQ,˜Q∈C θi= ∑|Q| s−1 ·|˜Q| s−1 ·m(proj θ (Q)∩proj θ (˜Q))Q,˜Q∈C im(proj θ (Q)∩proj θ (˜Q)) ≤ min{m(proj θ (Q)),m(proj θ (˜Q)} ≤ min{|Q|,| ˜Q|},∫L θf 2 dm ∑Q,˜Q∈C i|Q| s−1 ·|˜Q| s−1 ·min{|Q|,| ˜Q|}. (3.3.5)We now proceed to prove (3.3.4) by a double-counting argument. To this matter, consider, foreach pair <strong>of</strong> squares (Q, ˜Q) ∈ C i ×C i , the setThenI i Θ Q,˜Q=∑{}θ ∈ [−π/2,π/2];proj θ (Q)∩proj θ (˜Q) ≠ ∅ .|Q| s−1 ·|˜Q| s−1 ·min{|Q|,|˜Q|}·Q,˜Q∈C i= ∑∫ π/2−π/2χ ΘQ, ˜Q(θ)dθQ,˜Q∈C i|Q| s−1 ·|˜Q| s−1 ·min{|Q|,|˜Q|}·m(Θ Q,˜Q). (3.3.6)


Chapter 3. Yet another pro<strong>of</strong> <strong>of</strong> Marstrand’s Theorem 283.3.2 Transversality conditionThis subsection estimates the Lebesgue measure <strong>of</strong> Θ Q,˜Q.Lemma 3.3.2. If Q, ˜Q are squares <strong>of</strong> R 2 and x,˜x ∈ R 2 are its centers, respectively, thenPro<strong>of</strong>. Let θ ∈ Θ Q,˜Qand consider the figure.( )m Θ Q,˜Q≤ 2π · max{|Q|,|˜Q|} ·|x− ˜x|xθ |θ −ϕ 0|˜xL θproj θ (˜x)θproj θ (x)Since proj θ (Q) has diameter at most |Q| (and the same happens to ˜Q), we have |proj θ (x) −proj θ (˜x)| ≤ 2·max{|Q|,|˜Q|} and then, by elementary geometry,sin(|θ −ϕ 0 |) = |proj θ(x)−proj θ (˜x)||x− ˜x|≤ 2· max{|Q|,|˜Q|}|x− ˜x|=⇒ |θ −ϕ 0 | ≤ π · max{|Q|,|˜Q|}|x− ˜x|because sin −1 y ≤ πy/2. As ϕ 0 is fixed, the lemma is proved.,We point out that, although ingenuous, Lemma 3.3.2 expresses the crucial property <strong>of</strong>transversality that makes the pro<strong>of</strong> work, and all results related to Marstrand’s Theorem use asimilar idea in one way or another. See [43] where this tranversality condition is also exploited.By Lemma 3.3.2 and (3.3.6), we obtainI i ∑|x− ˜x| −1 ·|Q| s ·|˜Q| s . (3.3.7)Q,˜Q∈C i


Chapter 3. Yet another pro<strong>of</strong> <strong>of</strong> Marstrand’s Theorem 293.3.3 Good coversThe last summand will be estimated by choosing appropriate dyadic covers C i . Let C be anarbitrary dyadic cover <strong>of</strong> K. Remember K is an s-set satisfying (3.3.1).Definition 3.3.3. The dyadic cover C is good if∑|˜Q| s < max{128b,1}·|Q| s , ∀Q ∈ D. (3.3.8)˜Q∈C˜Q⊂QAny other constant dependingonly on K would work for the definition. Thereason we chosethis specific constant will become clear below, where we provide the existence <strong>of</strong> good dyadiccovers.Proposition 3.3.4. Let K ⊂ R 2 be a compact s-set satisfying (3.3.1). Then, for any δ > 0,there exists a good dyadic cover <strong>of</strong> K with diameter less than δ.Pro<strong>of</strong>. Let i 0 ≥ 1 such that 2 −i 0−1 < δ ≤ 2 −i 0. Begin with a finite cover U <strong>of</strong> K with diameterless than δ/4 such that∑|U| s < 2·m s (K ∩Q) , ∀Q ∈ D i0 .U∈UU⊂QNow, change U by adyadic cover C according to property(4) <strong>of</strong> Subsection 3.2.1. C hasdiameterat most δ and satisfies∑|˜Q| s < 64· ∑|U| s < 128·m s (K ∩Q) , ∀Q ∈ D i0 .˜Q∈C˜Q⊂QU∈UU⊂QBy additivity, the same inequality happens for any Q ∈ ⋃ 0≤i≤i 0D i and so, as m s (K ∩ Q) ≤b·|Q| s , it follows that∑|˜Q| s < 128b·|Q| s , ∀Q ∈ ⋃˜Q∈C˜Q⊂Q0≤i≤i 0D i , (3.3.9)that is, (3.3.8) holds for large scales. To control the small ones, apply the following operation:whenever Q ∈ ⋃ i>i 0D i is such that∑|˜Q| s > |Q| s ,˜Q∈C˜Q⊂Qwechange C by C∪{Q}\{˜Q ∈ C; ˜Q ⊂ Q}. It is clear that suchoperation preserves theinequality(3.3.9) and so, after a finite number <strong>of</strong> steps, we end up with a good dyadic cover.


Chapter 3. Yet another pro<strong>of</strong> <strong>of</strong> Marstrand’s Theorem 30As the constant in (3.3.8) does not depend on δ, there is a sequence (C i ) i≥1 <strong>of</strong> good dyadiccovers <strong>of</strong> K with diameters converging to <strong>zero</strong> such that∑|˜Q| s |Q| s , Q ∈ D and i ≥ 1. (3.3.10)˜Q∈C i˜Q⊂Q3.4 Pro<strong>of</strong> <strong>of</strong> Theorems 3.1.1 and 3.1.2Pro<strong>of</strong> <strong>of</strong> Theorem 3.1.1. Let (C i ) i≥1 beasequence <strong>of</strong> good dyadic covers satisfying (3.3.10) suchthat ‖C i ‖ → 0. By (3.3.7),I i ∑Q,˜Q∈C i|x− ˜x| −1 ·|Q| s ·|˜Q| s= ∑ Q∈C i ∞ ∑j=0≤ ∑ Q∈C i ∞ ∑j=0∑|x− ˜x| −1 ·|Q| s ·|˜Q| s˜Q∈C i2 −j−1 0, the <strong>sets</strong>G i ε = {θ ∈ [−π/2,π/2];(f∫L C iθθ) 2dm< ε−1}Then m ( [−π/2,π/2]\G i ε) ε, and the same holds for the setIf θ ∈ G ε , thenG ε = ⋂ i≥1∞⋃G j ε.j=im(proj θ (C i )) ε, for infinitely many n,, i ≥ 1.whichimpliesthatm(proj θ (K)) > 0. Finally, thesetG = ⋃ i≥1 G 1/i satisfiesm([−π/2,π/2]\G) =0 and m(proj θ (K)) > 0, for any θ ∈ G.


Chapter 3. Yet another pro<strong>of</strong> <strong>of</strong> Marstrand’s Theorem 31A direct consequence is theCorollary 3.4.1. The measure µ θ = (proj θ ) ∗ (m s | K ) is absolutely continuous with respect tom, for m-almost every θ ∈ R.Pro<strong>of</strong>. By Theorem 3.1.1, we have the implicationX ⊂ K, m s (X) > 0 =⇒ m(proj θ (X)) > 0, m-almost every θ ∈ R, (3.4.1)which is sufficient for the required absolute continuity. Indeed, if Y ⊂ L θ satisfies m(Y) = 0,thenµ θ (Y) = m s (X) = 0,where X = proj −1 θ (Y). Otherwise, by (3.4.1) we would have m(Y) = m(proj θ (X)) > 0,contradicting the assumption.Let f θ = dµ θ /dm. By the pro<strong>of</strong> <strong>of</strong> Theorem 3.1.1, we have∥∥f C i∥θ∥L 2 1, m-a.e. θ ∈ R. (3.4.2)Pro<strong>of</strong> <strong>of</strong> Theorem 3.1.2. Define, for each ε > 0, the function f θ,ε : L θ → R byf θ,ε (x) = 1 2ε∫ x+εx−εf θ (y)dm(y), x ∈ L θ .As f θ is an L 1 -function, the Lebesgue differentiation theorem gives that f θ (x) = lim ε→0 f θ,ε (x)for m-almost every x ∈ L θ . If we manage to show that 2‖f θ,ε ‖ L 2 1, m-a.e. θ ∈ R, (3.4.3)then Fatou’s lemma establishes the theorem. To this matter, first observe thatf θ,ε (x) = 1 2ε∫ x+εx−εf θ (y)dm(y)= 1 2ε ·µ θ([x−ε,x+ε])= 1 2ε ·m s((projθ ) −1 ([x−ε,x+ε])∩K ) .2 We consider ‖f θ,ε ‖ L 2 as a function <strong>of</strong> ε > 0.


Chapter 3. Yet another pro<strong>of</strong> <strong>of</strong> Marstrand’s Theorem 32In order to estimate this last term, fix ε > 0 and let i > 0 such that C = C i has diameter lessthan ε. Thenm s((projθ ) −1 ([x−ε,x+ε])∩K ) ≤∼≤∑Q∈Cproj θ (Q)⊂[x−2ε,x+2ε]∑Q∈Cproj θ (Q)⊂[x−2ε,x+2ε]∑Q∈Cproj θ (Q)⊂[x−2ε,x+2ε]∫ x+2εx−2εm s (Q∩K)|Q| sf C θ (y)dm(y),|Q| s−1 ·m(proj θ (Q))where in the second inequality we used (3.3.1). By the Cauchy-Schwarz inequality,|f θ,ε (x)| 2 1 2ε∫ x+2εx−2ε∣ fCθ (y) ∣ 2 dm(y)and so‖f θ,ε ‖ 2 L 2∼∫∫2εL θ1L θ∣ ∣ f C θ= ∥ ∥ fCθ∥ ∥2L 2∫ x+2εx−2ε∣ 2 dm∣ fCθ (y) ∣ 2 dm(y)dm(x)which, by (3.4.2), establishes (3.4.3).3.5 Concluding remarksThe good feature <strong>of</strong> the pro<strong>of</strong> is that the discretization idea may be applied to other contexts.For example, we prove in [24] a Marstrand type theorem in an arithmetical context.


Chapter 4A Marstrand theorem for sub<strong>sets</strong> <strong>of</strong>integersWe prove a Marstrand type theorem for a class <strong>of</strong> sub<strong>sets</strong> <strong>of</strong> the integers. More specifically,after defining the counting dimension D(E) <strong>of</strong> E ⊂ Z and the concepts <strong>of</strong> regularity andcompatibility, we show that if E,F ⊂ Z are two regular compatible <strong>sets</strong>, then D(E +⌊λF⌋) ≥min{1,D(E)+D(F)} for Lebesgue almost every λ ∈ R. If in addition D(E)+D(F) > 1, it isalso shown that E+⌊λF⌋ has positive upper Banach <strong>density</strong> for Lebesgue almost every λ ∈ R.The result has direct consequences when applied to arithmetic <strong>sets</strong>, such as the integer values<strong>of</strong> a polynomial with integer coefficients.4.1 IntroductionThe purpose <strong>of</strong> this chapter is to prove a Marstrand type theorem for a class <strong>of</strong> sub<strong>sets</strong> <strong>of</strong> theintegers.Thewell-known theorem <strong>of</strong> Marstrand [27] states the following: if K ⊂ R 2 is a Borel set suchthat its Hausdorff dimension is greater than one then, for almost every direction, its projectionto R in the respective direction has positive Lebesgue measure. In other words, this means K is“fat” in almost every direction. When K is the product <strong>of</strong> two real sub<strong>sets</strong> K 1 ,K 2 , Marstrand’stheorem can be stated in a more analytical form as: for Lebesgue almost every λ ∈ R, thearithmetic sum K 1 +λK 2 has positive Lebesgue measure. Much research has been done aroundthis topic, specially because the analysis <strong>of</strong> such arithmetic sums has applications in the theory<strong>of</strong> homoclinic birfurcations and also in diophantine approximations.Given a subset E ⊂ Z, let d ∗ (E) denote its upper Banach <strong>density</strong> 1 . A remarkable result in1 See Subsection 4.2.2 for the proper definitions.33


Chapter 4. Marstrand theorem for sub<strong>sets</strong> <strong>of</strong> integers 34additive combinatorics is Szemerédi’s theorem [45]. It asserts that if d ∗ (E) > 0, then E containsarbitrarily long arithmetic progressions. One can interpret this result by saying that <strong>density</strong>represents the correct notion <strong>of</strong> largeness needed to preserve finite configurations <strong>of</strong> Z. On theother hand, Szemerédi’s theorem cannot infer any property about sub<strong>sets</strong> <strong>of</strong> <strong>zero</strong> upper Banach<strong>density</strong>, and many <strong>of</strong> these <strong>sets</strong> are <strong>of</strong> interest. A class <strong>of</strong> examples are the integer values <strong>of</strong> apolynomial with integer coefficients and the prime numbers. These <strong>sets</strong> may, as well, containcombinatorially rich patterns. See, for example, [4] and [16].A set E ⊂ Z <strong>of</strong> <strong>zero</strong> upper Banach <strong>density</strong> is characterized as occupying portions in intervals<strong>of</strong> Z that grow in a sublinear fashion as the length <strong>of</strong> the intervals grow. On the other hand,there still may exist some kind <strong>of</strong> growth speed. For example, the number <strong>of</strong> perfect squares onan initial interval <strong>of</strong> length n is about n 0.5 . This exponent means, in some sense, a dimension<strong>of</strong> {n 2 ; n ∈ Z} inside Z. In the present chapter, we suggest a counting dimension D(E) for Eand establish the following Marstrand type result on the counting dimension <strong>of</strong> most arithmeticsums E +⌊λF⌋ for a class <strong>of</strong> sub<strong>sets</strong> E,F ⊂ Z.Theorem 4.1.1. Let E,F ⊂ Z be two regular compatible <strong>sets</strong>. ThenD(E +⌊λF⌋) ≥ min{1,D(E) +D(F)}for Lebesgue almost every λ ∈ R. If in addition D(E)+D(F) > 1, then E +⌊λF⌋ has positiveupper Banach <strong>density</strong> for Lebesgue almost every λ ∈ R.An immediate consequence <strong>of</strong> Theorem 4.1.1 is that the same result holds for sub<strong>sets</strong> <strong>of</strong> thenatural numbers.Corollary 4.1.2. Let E,F ⊂ N be two regular compatible <strong>sets</strong>. ThenD(E +⌊λF⌋) ≥ min{1,D(E) +D(F)}for Lebesgue almost every λ > 0. If in addition D(E)+D(F) > 1, then E +⌊λF⌋ has positiveupper Banach <strong>density</strong> for Lebesgue almost every λ > 0.The reader should make a parallel between the quantities d ∗ (E) and D(E) <strong>of</strong> sub<strong>sets</strong> E ⊂ Zand Lebesgue measure and Hausdorff dimension <strong>of</strong> sub<strong>sets</strong> <strong>of</strong> R. It is exactly this associationthat allows us to call Theorem 4.1.1 a Marstrand theorem for sub<strong>sets</strong> <strong>of</strong> integers.The notions <strong>of</strong> regularity and compatibility are defined in Subsections 4.4.1 and 4.4.2, respectively.Both are fulfilled for many arithmetic sub<strong>sets</strong> <strong>of</strong> Z, such as the integer values <strong>of</strong> apolynomial with integer coefficients and, more generally, for the universal <strong>sets</strong> (see Definition4.4.5). These are the <strong>sets</strong> that exhibit the expected growth rate along intervals <strong>of</strong> arbitrarylength. The second result <strong>of</strong> this work is


Chapter 4. Marstrand theorem for sub<strong>sets</strong> <strong>of</strong> integers 35Theorem 4.1.3. Let E 0 ,...,E k be universal sub<strong>sets</strong> <strong>of</strong> Z. ThenD(E 0 +⌊λ 1 E 1 ⌋+···+⌊λ k E k ⌋) ≥ min{1,D(E 0 )+D(E 1 )+···+D(E k )}for Lebesgue almost every λ = (λ 1 ,...,λ k ) ∈ R k . If in addition ∑ ki=0 D(E i) > 1, then thearithmetic sum E 0 + ⌊λ 1 E 1 ⌋ + ··· + ⌊λ k E k ⌋ has positive upper Banach <strong>density</strong> for Lebesguealmost every λ = (λ 1 ,...,λ k ) ∈ R k .The integer values <strong>of</strong> a polynomial with integer coefficients have special interest in ergodictheory and its connections with combinatorics, due to ergodic theorems along these sub<strong>sets</strong> [6],as well as its combinatorial implications. See [4] for the remarkable work <strong>of</strong> V. Bergelson andA. Leibman on the polynomial extension <strong>of</strong> Szemerédi’s theorem. Theorem 4.1.3 has a directconsequence when applied to this setting.Corollary 4.1.4. Let p i ∈ Z[x] with degree d i > 0 and let E i = (p i (n)) n∈Z , i = 0,...,k. ThenD(E 0 +⌊λ 1 E 1 ⌋+···+⌊λ k E k ⌋) ≥ min{1, 1 + 1 +···+ 1 }d 0 d 1 d kfor Lebesgue almost every λ = (λ 1 ,...,λ k ) ∈ R k . If in addition ∑ ki=0 d i −1 > 1, then thearithmetic sum E 0 + ⌊λ 1 E 1 ⌋ + ··· + ⌊λ k E k ⌋ has positive upper Banach <strong>density</strong> for Lebesguealmost every λ = (λ 1 ,...,λ k ) ∈ R k .The pro<strong>of</strong> <strong>of</strong> Theorem 4.1.1 is based on the ideas developed in [24] and [26]. It relies onthe fact that the cardinality <strong>of</strong> a regular subset <strong>of</strong> Z along an increasing sequence <strong>of</strong> intervalsexhibits an exponential behavior ruled out by its counting dimension. As this holds for tworegular sub<strong>sets</strong> E,F ⊂ Z, the compatibility assumption allows to estimate the cardinality <strong>of</strong>the arithmetic sum E + ⌊λF⌋ along the respective arithmetic sums <strong>of</strong> intervals and, finally, adouble-counting argument estimates the size <strong>of</strong> the “bad” parameters for which such cardinalityis small. Theorem 4.1.3 follows from Theorem 4.1.1 by a fairly simple induction.This chapter is organized as follows. In Section 4.2 we introduce the basic notations anddefinitions. Section 4.3 is devoted to the discussion <strong>of</strong> examples. In particular, the <strong>sets</strong> given byinteger values <strong>of</strong> a polynomial with integer coefficients are investigated in Subsection 4.3.1. InSection 4.4 we introduce the notions <strong>of</strong> regularity and compatibility. Subsection 4.4.3 provides acounterexample to Theorem 4.1.1 when the <strong>sets</strong> are no longer compatible and Subsection 4.4.4a counterexample to the same theorem when the space <strong>of</strong> parameters is Z. Finally, in Section4.5 we prove Theorems 4.1.1 and 4.1.3. We also collect final remarks and further questions inSection 4.6.


Chapter 4. Marstrand theorem for sub<strong>sets</strong> <strong>of</strong> integers 364.2 Preliminaries4.2.1 General notationGiven a set X, |X| denotes the cardinality <strong>of</strong> X. Z denotes the set <strong>of</strong> integers and N the set <strong>of</strong>positive integers. We use the following notation to compare the asymptotic <strong>of</strong> functions.Definition 4.2.1. Let f,g : Z or N → R be two real-valued functions. We say f g if there isa constant C > 0 such that|f(n)| ≤ C ·|g(n)|, ∀n ∈ Z or N.If f g and g f, we write f ∼ g. We say f ≈ g iff(n)lim|n|→∞ g(n) = 1.For each x ∈ R, ⌊x⌋ denotes the integer part <strong>of</strong> x. For each k ≥ 1, m k denotes the Lebesguemeasure <strong>of</strong> R k . For k = 1, let m = m 1 . The letter I will always denote an interval <strong>of</strong> Z:I = (M,N] = {M +1,...,N}.The length <strong>of</strong> I is equal to its cardinality, |I| = N −M.For E ⊂ Z and λ ∈ R, λE denotes the set {λn; n ∈ E} ⊂ R and ⌊λE⌋ the set {⌊λn⌋; n ∈E} ⊂ Z.4.2.2 Counting dimensionDefinition 4.2.2. The upper Banach <strong>density</strong> <strong>of</strong> E ⊂ Z is equal towhere I runs over all intervals <strong>of</strong> Z.d ∗ (E) = limsup|I|→∞|E ∩I||I|Definition 4.2.3. The counting dimension or simply dimension <strong>of</strong> a set E ⊂ Z is equal towhere I runs over all intervals <strong>of</strong> Z.D(E) = limsup|I|→∞,log|E ∩I|log|I|Obviously, D(E) ∈ [0,1] and D(E) = 1 whenever d ∗ (E) > 0. The counting dimension allowsus to distinguish largeness between two <strong>sets</strong> <strong>of</strong> <strong>zero</strong> upper Banach <strong>density</strong>. Similar definitions toD(E) have appeared in [2] and [13]. We now give another characterization to it that is similarin spirit to the Hausdorff dimension <strong>of</strong> sub<strong>sets</strong> <strong>of</strong> R. Let α be a nonnegative real number.,


Chapter 4. Marstrand theorem for sub<strong>sets</strong> <strong>of</strong> integers 37Definition 4.2.4. The counting α-measure <strong>of</strong> E ⊂ Z is equal towhere I runs over all intervals <strong>of</strong> Z.H α (E) = limsup|I|→∞|E ∩I||I| α ,Clearly, H α (E) ∈ [0,∞]. For a fixed E ⊂ Z, the numbers H α (E) are decreasing in α andone easily checks thatα < D(E) =⇒ H α (E) = ∞ and α > D(E) =⇒ H α (E) = 0,which in turn implies the existence and uniqueness <strong>of</strong> α ≥ 0 such thatH β (E) = ∞ , if 0 ≤ β < α,= 0 , if β > α.The above equalities imply that D(E) = α, that is, the counting dimension is exactly theparameter α where H α (E) decreases from infinity to <strong>zero</strong>. Also, if β > D(E), then|E ∩I| |I| β , (4.2.1)where I runs over all intervals <strong>of</strong> Z and, conversely, if (4.2.1) holds, then D(E) ≤ β.Below, we collect basic properties <strong>of</strong> the counting dimension and α-measure.(i) E ⊂ F =⇒ D(E) ≤ D(F).(ii) D(E ∪F) = max{D(E),D(F)}.(iii) For any λ > 0,H α (⌊λE⌋) = λ −α ·H α (E). (4.2.2)The first two are direct. Let’s prove (iii). For any interval I ⊂ Z, we have |E ∩ I| ≈|⌊λE⌋∩⌊λI⌋| and so|E ∩I||I| α≈ λ α · |⌊λE⌋∩⌊λI⌋||⌊λI⌋| α=⇒ H α (E) = λ α ·H α (⌊λE⌋).Remark 4.2.5. As ⌊−x⌋ = −⌊x⌋ or ⌊−x⌋ = −⌊x⌋ − 1, the <strong>sets</strong> ⌊−λE⌋ and ⌊λE⌋ have thesame counting dimension. Also, 0 < H α (⌊−λE⌋) < ∞ if and only if 0 < H α (⌊λE⌋) < ∞. Forthese reasons, we assume from now on that λ > 0.


Chapter 4. Marstrand theorem for sub<strong>sets</strong> <strong>of</strong> integers 384.3 ExamplesExample 1. Let α ∈ (0,1] andE α ={⌊n 1/α⌋ }; n ∈ N . (4.3.1)We infer that H α (E α ) = 1. To prove this, we make use <strong>of</strong> the inequality 2(x+y) α ≤ x α +y α , x,y ≥ 0,to conclude that|E α ∩(M,N]|(N −M) α ≤ (N +1)α −(M +1) α(N −M) α ≤ 1.This proves that H α (E α ) ≤ 1. On the other hand,|E α ∩(0,N]|N α≥ Nα −1N αand so H α (E α ) ≥ 1. Then H α (E α ) = 1 and, in particular, D(E α ) = α.Example 2. The setE = ⋃ n∈N(n n ,(n+1) n ]∩E 1−1/nhas <strong>zero</strong> upper Banach <strong>density</strong> and D(E) = 1.Example 3. The prime numbers have dimension one. This follows from the prime numbertheorem:log|{1 ≤ p ≤ n; p is prime}| logn−loglognlim= lim = 1.n→∞ logn n→∞ lognExample 4. Sets <strong>of</strong> <strong>zero</strong> upper Banach <strong>density</strong> appear naturally in infinite ergodic theory. Let(X,A,µ,T) be a sigma-finite measure-preserving system, with µ(X) = ∞, and let A ∈ A havefinite measure. Fixed x ∈ A, let E = {n ≥ 1; T n x ∈ A}. By Hopf’s Ratio Ergodic Theorem,E has <strong>zero</strong> upper Banach <strong>density</strong> almost surely. In many specific cases, its dimension can becalculated or at least estimated. See [1] for further details.4.3.1 Polynomial sub<strong>sets</strong> <strong>of</strong> ZDefinition 4.3.1. A polynomial subset <strong>of</strong> Z is a set E = {p(n); n ∈ Z}, where p(x) is anon-constant polynomial with integer coefficients.2 For each t ≥ 0, the function x ↦→ x α +(t−x) α , 0 ≤ x ≤ t, is concave and so attains its minimum in x = 0and x = t. Then x α +(t−x) α ≥ t α for any x ∈ [0,t].


Chapter 4. Marstrand theorem for sub<strong>sets</strong> <strong>of</strong> integers 39These are the <strong>sets</strong> we consider in Theorem 4.1.3. Their counting dimension is easily calculatedasfollows.GivenE,F ⊂ Z,letE andF beasymptoticifE = {··· < a −1 < a 0 < a 1 < ···},F = {··· < b −1 < b 0 < b 1 < ···} and there is i ≥ 0 such thata n−i ≤ b n ≤ a n+i , for every n ∈ Z. (4.3.2)Denote this by E ≈ F.Lemma 4.3.2. If E,F ⊂ Z are asymptotic then H α (E) = H α (F), for any α ≥ 0. In particular,D(E) = D(F).Pro<strong>of</strong>. Let I = (M,N] be an interval and assume E ∩I = {a m+1 ,a m+2 ,...,a n }. By relation(4.3.2),b m−i ≤ a m ≤ M < a m+1 ≤ b m+i+1 and b n−i ≤ a n ≤ N < a n+1 ≤ b n+i+1 ,which imply the inclusions{b m+i+1 ,...,b n−i } ⊂ F ∩I ⊂ {b m−i ,...,b n+i+1 }.Then |E ∩I| ≈ |F ∩I| and soH α (E) = limsup|I|→∞|E ∩I||I| α= limsup|I|→∞|F ∩I||I| α= H α (F).Let E = {p(n); n ∈ Z}, where p(x) ∈ Z[x] has degree d. Assuming p has leading coefficienta > 0, there is i ≥ 0 such that a · (n − i) d < p(n) < a · (n + i) d for every n ∈ Z and thenE ≈ aE 1/d , where E 1/d is defined as in (4.3.1). By Lemma 4.3.2, we getD(E) = 1 d and H 1/d(E) = 1.4.3.2 Cantor <strong>sets</strong> in ZThe famous ternary Cantor set <strong>of</strong> R is formed by the real numbers <strong>of</strong> [0,1] with only 0’s and2’s on their base 3 expansion. In analogy to this, let E ⊂ Z be defined as{ n∑}E = a i ·3 i ; n ∈ N and a i = 0 or 2 . (4.3.3)i=0The set E has been slightly investigated in [11]. There, A. Fisher proved thatH log2/log3 (E) > 0.


Chapter 4. Marstrand theorem for sub<strong>sets</strong> <strong>of</strong> integers 40We prove below that H log2/log3 (E) ≤ 2, which in particular gives that D(E) = log2/log3,as expected. Let I = (M,N] be an interval <strong>of</strong> Z. We can assume M + 1,N ∈ E. Indeed, ifĨ = ( ˜M,Ñ], where ˜M +1 and Ñ are the smallest and largest elements <strong>of</strong> E ∩I, respectively,thenLet M +1,N ∈ E, say|E ∩I||I| α≤|E ∩Ĩ||Ĩ|α ·M +1 = a 0 ·3 0 +a 1 ·3 1 +···+a m−1 ·3 m +2·3 mN = b 0 ·3 0 +b 1 ·3 1 +···+b n−1 ·3 n +2·3 n .We can also assume that m < n. If this is not the case, the quotient |E ∩ I|/|I| α does notdecrease if we change I by (M −2·3 n ,N −2·3 n ]. With these assumptions,and thenN −M ≥ 2·3 n −(3 m+1 −2) ≥ 3 n|E ∩I| |E ∩(0,N]|≤|I| log2/log3 |I| log2/log3 ≤ 2 n+1(3 n = 2.) log2/log3Because I is arbitrary, this gives that H log2/log3 (E) ≤ 2.Observe that the renormalization <strong>of</strong> E ∩ (0,3 n ) via the linear map x ↦→ x/3 n generatesa subset <strong>of</strong> the unit interval (0,1) that is equal to the set <strong>of</strong> left endpoints <strong>of</strong> the remainingintervals <strong>of</strong> the n-th step <strong>of</strong> the construction <strong>of</strong> the ternary Cantor set <strong>of</strong> R. In other words,if K = ⋃ n∈E [n,n + 1], then K/3n is exactly the n-th step <strong>of</strong> the construction <strong>of</strong> the ternaryCantor set <strong>of</strong> R.More generally, let us define a class <strong>of</strong> Cantor <strong>sets</strong> in Z. Fix a basis a ∈ N and a binarymatrix A = (a ij ) 0≤i,j≤a−1 . LetΣ n (A) = { (d 0 d 1···d n−1 d n ); a di−1 d i= 1, 1 ≤ i ≤ n } , n ≥ 0,denote the set <strong>of</strong> admissible words <strong>of</strong> length n+1 and Σ ∗ (A) = ⋃ n≥0 Σ n(A) the set <strong>of</strong> all finiteadmissible words.Definition 4.3.3. The integer Cantor set E A ⊂ Z associated to the matrix A is the setE A = {d 0 ·a 0 +···+d n ·a n ; (d 0 d 1···d n ) ∈ Σ ∗ (A)}.Ourdefinition was inspiredon thefact that dynamically defined topologically mixingCantor<strong>sets</strong> <strong>of</strong> the real line are homeomorphic to subshifts <strong>of</strong> finite type, which is exactly what we didabove, after truncating the numbers. See [24] for more details. The dimension <strong>of</strong> E A , as in the


Chapter 4. Marstrand theorem for sub<strong>sets</strong> <strong>of</strong> integers 41inspiring case, depends on the Perron-Frobenius eigenvalue <strong>of</strong> A. Remember that the Perron-Frobenius eigenvalue is the largest eigenvalue λ + (A) <strong>of</strong> A. It has multiplicity one and maximizesthe absolute value <strong>of</strong> the eigenvalues <strong>of</strong> A. Also, there is a constant c = c(A) > 0 such thatwhose pro<strong>of</strong> may be found in [22].c −1 ·λ + (A) n ≤ |Σ n (A)| ≤ c·λ + (A) n , for every n ≥ 0, (4.3.4)Lemma 4.3.4. If A is a binary a×a matrix, thenD(E A ) = logλ +(A)logaand 0 < Hlogλ + (A)(E A ) < ∞.logaPro<strong>of</strong>. Let I = (M,N]. Again, we may assume M +1,N ∈ E A , sayM +1 = x 0 ·a 0 +···+x n ·a nN = y 0 ·a 0 +···+y n ·a n ,with y n > x n . If y n ≥ x n +2, then{M +1 ≤ (x n +1)·a nand, as I ⊂ (0,a n+1 ), we have=⇒ |I| ≥ anN ≥ (x n +2)·an |E A ∩I||I| logλ + (A)loga≤ |Σ n(A)|a nlogλ + (A)loga≤ c·λ +(A) nλ + (A) n = c. (4.3.5)If y n = x n +1, let i,j ∈ {0,1,...,n−1} be the indices for which(i) x i < a−1 and x i+1 = ··· = x n−1 = a−1,(ii) y j > 0 and y j+1 = ··· = y n−1 = 0.Then {M +1 ≤ (x n +1)·a n −a iN ≥ (x n +1)·a n +a j =⇒ |I| ≥ a i +a j ≥ a max{i,j} .In order to ∑ nk=0 z k ·a k belong to I, one must have z n = x n or z n = x n +1. In the first casez i+1 = ··· = z n−1 = a−1 and in the second z j+1 = ··· = z n−1 = 0. Then|E A ∩I| ≤ |Σ i (A)|+|Σ j (A)| ≤ 2c·λ + (A) max{i,j}and so|E A ∩I||I| logλ + (A)loga≤ 2c·λ +(A) max{i,j}= 2c. (4.3.6)a max{i,j}logλ + (A)loga


Chapter 4. Marstrand theorem for sub<strong>sets</strong> <strong>of</strong> integers 42Estimates (4.3.5) and (4.3.6) give H logλ+ (A)/loga < ∞. On the other hand,|E A ∩(0,a n ]|a nlogλ + (A)logawhich concludes the pro<strong>of</strong>.≥ c−1 ·λ + (A) n−1λ + (A) n = c −1 ·λ + (A) −1 =⇒ H logλ+ (A)/loga > 0,By the above lemma, if X ⊂ {0,...,a−1} and A = (a ij ) is defined by a ij = 1 iff i,j ∈ X,then D(E A ) = log|X|/loga, which extends the results about the ternary Cantor set defined in(4.3.3).Ingeneral, ifE,F aresub<strong>sets</strong><strong>of</strong>ZsuchthatD(E)+D(F) > 1, itisnottruethatd ∗ (E+F) >0, because the elements <strong>of</strong> E +F may have many representations as the sum <strong>of</strong> one element <strong>of</strong>E and other <strong>of</strong> F. This resonance phenomenon decreases the dimension <strong>of</strong> E+F. Lemma 4.3.4provides a simple example to this situation: if E = E A and F = E B , where A = (a ij ) 0≤i,j≤11 ,B = (b ij ) 0≤i,j≤11 are defined bya ij = 1 ⇐⇒ 0 ≤ i,j ≤ 3 and b ij = 1 ⇐⇒ 4 ≤ i,j ≤ 7,then D(E)+D(F) = 2log4/log12, while E +F = E C for C = (c ij ) 0≤i,j≤11 given byc ij = 1 ⇐⇒ 4 ≤ i,j ≤ 10.E+F has counting dimension equal to log7/log12 and so d ∗ (E+F) = 0. What Theorem 4.1.1proves is that resonance is avoided if one is allowed to change the scales <strong>of</strong> the <strong>sets</strong>, multiplyingone <strong>of</strong> them by a factor λ ∈ R.4.3.3 Generalized IP-<strong>sets</strong>This class <strong>of</strong> <strong>sets</strong> was suggested to us by Simon Griffiths and Rob Morris.Definition 4.3.5. Thegeneralized IP-setassociatedtothesequences<strong>of</strong>positiveintegers(k n ) n≥1 ,(d n ) n≥1 is the set { n∑}x i ·d i ; n ∈ N and 0 ≤ x i < k ii=1E =.We assume d n > ∑ n−1i=1 k i·d i , in which case the map ∑ ni=1 x i·d i ↦→ (x 1 ,...,x n ) is a bijectionbetween E and the set <strong>of</strong> finite sequencesΣ ∗ = {(x 1 ,...,x n ); n ∈ N,x n > 0 and 0 ≤ x i < k i }.Also, if Σ ∗ is lexicographically ordered 3 , then the map is order-preserving.3 The sequence (x 1,...,x n) is smaller than (y 1,...,y m) if n < m or if there is i ∈ {1,...,n} such that x i < y iand x j = y j for j = i+1,...,n.


Chapter 4. Marstrand theorem for sub<strong>sets</strong> <strong>of</strong> integers 43Lemma 4.3.6. Let E be the generalized IP-set associated to (k n ) n≥1 , (d n ) n≥1 and let p n =k 1···k n . Thenlimsupn→∞logp n logp n≤ D(E) ≤ limsup · (4.3.7)logk n d n n→∞ logd nIn particular, if logk n /logp n → 0, then D(E) = limsup n→∞ logp n /logd n .Pro<strong>of</strong>. The calculations are similar to those <strong>of</strong> Lemma 4.3.4. Let I = (M,N], sayM +1 = x 1 ·d 1 +···+x n ·d nN = y 1 ·d 1 +···+y n ·d n ,with y n > x n . If y n ≥ x n +2, then{M +1 ≤ (x n +1)·d nN ≥ (x n +2)·d n=⇒ |I| ≥ d nand |E ∩I| ≤ p n , thus givinglog|E ∩I|log|I|If y n = x n +1, let i,j ∈ {1,...,n−1} be the indices for which(i) x i < k i −1 and x j = k j −1, j = i+1,...,n−1,(ii) y j > 0 and y j+1 = ··· = y n−1 = 0.Then≤ logp nlogd n· (4.3.8)n−1∑n−1∑M +1 ≤ x n ·d n +x i ·d i + k l ·d l < x n ·d n −d i + k l ·d l < (x n +1)·d n −d il=1l≠il=1andN ≥ y j ·d j +y n ·d n ≥ (x n +1)·d n +d j ,implying that |I| ≥ d i +d j ≥ d max{i,j} . Now, in order to ∑ nl=1 z l·d l belong to I, one must havez n = x n or z n = x n +1. In the first case z i+1 = k i+1 −1,...,z n−1 = k n−1 −1 and in the secondz j+1 = ··· = z n−1 = 0. Then |E ∩I| ≤ p i +p j ≤ 2p max{i,j} and solog|E ∩I|log|I|≤ log2p max{i,j}logd max{i,j}· (4.3.9)Relations (4.3.8) and (4.3.9) prove the right hand inequality <strong>of</strong> (4.3.7). The other follows byconsidering the intervals (0, ∑ ni=1 k i ·d i ].


Chapter 4. Marstrand theorem for sub<strong>sets</strong> <strong>of</strong> integers 444.4 Regularity and compatibility4.4.1 Regular <strong>sets</strong>Definition 4.4.1. A subset E ⊂ Z is called regular or D(E)-set if 0 < H D(E) (E) < ∞.By Lemmas 4.3.2 and 4.3.4, polynomial <strong>sets</strong> and Cantor <strong>sets</strong> are regular.Definition 4.4.2. Given two sub<strong>sets</strong> E = {··· < x −1 < x 0 < x 1 < ···} and F <strong>of</strong> Z, let E ∗Fdenote the setE ∗F = {x n ; n ∈ F}.This is a subset <strong>of</strong> E whose counting dimension is at most D(E)D(F). To see this, consideran arbitrary interval I ⊂ Z. If E ∩I = {x i+1 ,...,x j }, then(E ∗F)∩I = {x n ;n ∈ F ∩(i,j]}.Given α > D(E) and β > D(F), relation (4.2.1) guarantees that|(E ∗F)∩I| = |F ∩(i,j]| (j −i) β = |E ∩I| β |I| αβand so D(E ∗F) ≤ αβ. Choosing α, β arbitrarily close to D(E), D(F), respectively, it followsthat D(E ∗F) ≤ D(E)D(F).If E is regular, it is possible to choose F with arbitrary dimension in such a way that E∗Fis also regular and has dimension equal to D(E)D(F). To this matter, choose disjoint intervalsI n = (a n ,b n ], n ≥ 1, with lengths going to infinity such that|E ∩I n ||I n | D(E) 1and let E ∩I n = {x in+1 < x in+2 < ··· < x jn }, where i n < j n . Given α ∈ [0,1], letF = ⋃ n≥1(E α +i n )∩(i n ,j n ],where E α is defined as in (4.3.1). Then D(F) = D(E α ) = α and|(E ∗F)∩I n | ≥ |F ∩(i n ,j n ]|= |E α ∩(0,j n −i n ]| (j n −i n ) D(F)= |E ∩I n | D(F) |I n | D(E)D(F) ,


Chapter 4. Marstrand theorem for sub<strong>sets</strong> <strong>of</strong> integers 45implying that|(E ∗F)∩I n | |I n | D(E)D(F) . (4.4.1)This proves the reverse inequality D(E∗F) ≥ D(E)D(F). We thus obtain that, given a regularsubset E ⊂ Z and 0 ≤ α ≤ D(E), there exists a regular subset E ′ ⊂ E such that D(E ′ ) = α. Itis a harder task to prove that this holds even when E is not regular.Proposition 4.4.3. Let E ⊂ Z and 0 ≤ α ≤ 1. If H α (E) > 0, then there exists a regular subsetE ′ ⊂ E such that D(E ′ ) = α. In particular, for any 0 ≤ α < D(E), there is E ′ ⊂ E regularsuch that D(E ′ ) = α.Pro<strong>of</strong>. The idea is to apply a dyadic argument in E to decrease H α (E) in a controlled way.Given an interval I ⊂ Z and a subset F ⊂ Z, defines F (I) = . |F ∩J|supJ⊂I |J| α ·J intervalIf F = {a 1 ,a 2 ,...,a k } ⊂ I, the dyadic operation <strong>of</strong> alternately discard the elements a 2 ,a 4 ,...,a 2⌊(k−1)/2⌋ <strong>of</strong> the set {a 2 ,a 3 ,...,a k−1 },F = {a 1 ,a 2 ,...,a k } F ′ = {a 1 ,a 3 ,a 5 ,...,a 2⌈(k−1)/2⌉−1 ,a k },decreases s F (I) to approximately s F (I)/2. More specifically, if s F (I) > 2, then s F ′(I) > 1/2.Indeed, for every interval J ⊂ I|F ′ ∩J||J| αand, for J maximizing s F (I),≤ 1 2· |F ∩J|+1|J| α≤ s F(I)2+ 1 2 < s F(I)− 1 2|F ′ ∩J||J| α≥ 1 2· |F ∩J|−1|J| α> 1−12·|J| α ≥ 1 2 ·After a finite number <strong>of</strong> these dyadic operations, one obtains a subset F ′ ⊂ F such that12 < s F ′(I) ≤ 2.If H α (E) < ∞, there is nothing to do. Assume that H α (E) = ∞. We proceed inductivelyby constructing a sequence F 1 ⊂ F 2 ⊂ ··· <strong>of</strong> finite sub<strong>sets</strong> <strong>of</strong> E contained in a sequence <strong>of</strong>intervals I n = (a n ,b n ] with increasing lengths such that(i) 1/2 < s Fn (I n ) ≤ 3;(ii) there is an interval J n ⊂ I n such that |J n | ≥ n and|F n ∩J n ||J n | α > 1 2 ·


Chapter 4. Marstrand theorem for sub<strong>sets</strong> <strong>of</strong> integers 46Once these properties are fulfilled, the set E ′ = ⋃ n≥1 F n will satisfy the required conditions.Take any a ∈ E and I 1 = {a}. Assume I n , F n and J n have been defined satisfying (i) and(ii). As H α (E) = ∞, there exists an interval J n+1 disjoint from (a n −|I n | 1/α ,b n +|I n | 1/α ] forwhich|E ∩J n+1 ||J n+1 | α ≥ (n+1) 1−α .This inequality allows to restrict J n+1 to a smaller interval <strong>of</strong> size at least n+1, also denotedJ n+1 , such thats E (J n+1 ) = |E ∩J n+1||J n+1 | α · (4.4.2)Consider F ′ n+1 = E ∩J n+1 and apply the dyadic operation to F ′ n+1 until12 < s F n+1 ′ (J n+1) = |F′ n+1 ∩J n+1||J n+1 | α ≤ 2. (4.4.3)Let I n+1 = I n ∪K n ∪J n+1 be the convex hull 4 <strong>of</strong> I n and J n+1 and F n+1 = F n ∪F ′ n+1 . Condition(ii) is satisfied because <strong>of</strong> (4.4.3). To prove (i), let I be a subinterval <strong>of</strong> I n+1 . We have threecases to consider.• I ⊂ I n ∪K n : by condition (i) <strong>of</strong> the inductive hypothesis,• I ⊂ K n ∪J n+1 : by (4.4.3),|F n+1 ∩I||I| α ≤ |F n ∩(I ∩I n )||I ∩I n | α ≤ 3.|F n+1 ∩I||I| α≤ |F′ n+1 ∩(I ∩J n+1)||I ∩J n+1 | α ≤ 2.• I ⊃ K n : as |K n | ≥ |I n | 1/α ,|F n+1 ∩I||I| α= |F n ∩(I ∩I n )|+|F ′ n+1 ∩(I ∩J n+1)|(|I ∩I n |+|K n |+|I ∩J n+1 |) α≤|F n ∩(I ∩I n )|(|I ∩I n |+|K n |) α + |F′ n+1 ∩(I ∩J n+1)|(|I ∩J n+1 |) α≤ |I n||K n | α +s F ′ n+1 (J n+1)≤ 3.This proves condition (i) and completes the inductive step.By the above proposition and equation (4.2.2), for any α ∈ [0,1] and h ∈ (0,∞), there existsan α-set E ⊂ Z such that H α (E) = h. We’ll use this fact in Subsection 4.4.3.4 Observe that |K n| ≥ |I n| 1/α .


Chapter 4. Marstrand theorem for sub<strong>sets</strong> <strong>of</strong> integers 474.4.2 Compatible <strong>sets</strong>Definition 4.4.4. Two regular sub<strong>sets</strong> E,F ⊂ Z are compatible if there exist two sequences(I n ) n≥1 , (J n ) n≥1 <strong>of</strong> intervals with increasing lengths such that1. |I n | ∼ |J n |,2. |E ∩I n | |I n | D(E) and |F ∩J n | |J n | D(F) .The notion <strong>of</strong> compatibility means that E and F have comparable intervals on which therespective intersections obey the correct growth speed <strong>of</strong> cardinality. Some <strong>sets</strong> have theseintervals in all scales.Definition 4.4.5. A regular subset E ⊂ Z is universal if there exists a sequence (I n ) n≥1 <strong>of</strong>intervals such that |I n | ∼ n and |E ∩I n | |I n | D(E) .It is clear that E and F are compatible whenever one <strong>of</strong> them is universal and the otheris regular. Every E α is universal and the same happens to polynomial sub<strong>sets</strong> E, due to theasymptotic relation E ≈ aE 1/d , where d is the degree <strong>of</strong> the polynomial that defines E (seeSubsection 4.3.1). In particular, any two polynomial sub<strong>sets</strong> are compatible.4.4.3 A counterexample <strong>of</strong> Theorem 4.1.1 for regular non-compatible <strong>sets</strong>In this subsection, we construct regular <strong>sets</strong> E,F ⊂ Z such that D(E)+D(F) > 1 and E+⌊λF⌋has <strong>zero</strong> upper Banach <strong>density</strong> for every λ ∈ R. The idea is, in the contrast to compatibility,|E ∩I| |F ∩J|construct E and F such that the intervals I,J ⊂ Z for which and are bounded|I| D(E) |J| D(F)away from <strong>zero</strong> have totally different sizes.Let α ∈ (1/2,1). We will defineE = ⋃i=1,3,...such that the following conditions hold:(E i ∩I i ) and F = ⋃i=2,4,...(E i ∩I i )(i) E i = ⌊µ i ⌊µ i −1 E 0 ⌋⌋, i ≥ 1, where E 0 ⊂ N is an α-set with H α (E 0 ) = 1/2.(ii) I i = (a i ,b i ], i ≥ 1, is a disjoint sequence <strong>of</strong> intervals <strong>of</strong> increasing length such that|E i ∩I i |limi→∞ |I i | α = 1 2 ·(iii) µ i > b i−121−α for every i ≥ 1.


Chapter 4. Marstrand theorem for sub<strong>sets</strong> <strong>of</strong> integers 48It is clear that we can inductively construct the sequences (µ i ) i≥1 ,(E i ) i≥1 and (I i ) i≥1 in sucha way that 0 < b i < a i+1 for every i ≥ 1. Before going into the calculations, let us explain theabove conditions. Thedilations in(i) guarantee that consecutive elements <strong>of</strong> E i differ at least by(the order <strong>of</strong>) µ i ; (ii) ensures that E,F are α-<strong>sets</strong>; (iii) ensures that the referred incompatibilitybetween E and F holds and, also, that the left endpoint <strong>of</strong> I i is much bigger than the rightendpoint <strong>of</strong> I i−1 .Lemma 4.4.6. Let E,F ⊂ Z with D(E),D(F) < 1, A,B ⊂ Z finite and E ′ = E ∪ A,F ′ = F ∪B. ThenPro<strong>of</strong>. We haved ∗ (E ′ +⌊λF ′ ⌋) = d ∗ (E +⌊λF⌋), ∀λ ∈ R.E ′ +⌊λF ′ ⌋ = (E +⌊λF⌋)∪(A+⌊λF⌋)∪(E +⌊λB⌋)∪(A+⌊λB⌋)and each <strong>of</strong> the <strong>sets</strong> A+⌊λF⌋, E +⌊λB⌋, A+⌊λB⌋ has dimension smaller than one.Fix λ > 0. By removing, if necessary, finitely many intervals I i , we may also assume that(iv) ⌊λI i ⌋∩⌊λI j ⌋ = ∅ for all i ≠ j.(v) b i > max{4λ −1 ,4λ} 1−α1+α for all i ≥ 1.By the previous lemma, this does not affect the value <strong>of</strong> d ∗ (E+⌊λF⌋). With these assumptions,(I i +⌊λI j ⌋)∩(I k +⌊λI l ⌋) ≠ ∅ ⇐⇒ i = k and j,l < i or j = l and i,k < jand then(E +⌊λF⌋)∩(a i ,a i+1 ] =⊔(a+⌊λI j ⌋) , if i is odda∈I ij=2,4,...,i−1=⊔(I j +⌊λb⌋) , if i is evenb∈I ij=1,3,...,i−1In addition, (v) implies that, if i is odd, a,a ′ ∈ I i are distinct and j,k < i are even, then thegap between a+⌊λI j ⌋ and a ′ +⌊λI k ⌋ is at least |a−a ′ |/2; if i is even, b,b ′ ∈ I i are distinct andj,k < i are odd, then the gap between I j +⌊λb⌋ and I k +⌊λb ′ ⌋ is at least |b−b ′ |/4λ.By the above fractal-like structure, it is expected that E + ⌊λF⌋ has <strong>zero</strong> upper Banach<strong>density</strong>. Let us formally prove this. Consider an interval I = (M,N] ⊂ Z and, as usual, assumethat M +1,N ∈ E +⌊λF⌋, sayM +1 = a ′ +⌊λb ′ ⌋ ∈ I k +⌊λI l ⌋N = a+⌊λb⌋ ∈ I i +⌊λI j ⌋.


Chapter 4. Marstrand theorem for sub<strong>sets</strong> <strong>of</strong> integers 49The largest index among i,j,k,l is either i or j. We may assume that i > j. In fact, thereverse case is symmetric, because E+⌊λF⌋ is basically ⌊λ(F +⌊λ −1 E⌋)⌋ and, with this interpretation,theroles<strong>of</strong>I i andI j areinterchanged. Inthissituation, wehavetwocases toconsider:• a = a ′ : an element r +⌊λs⌋ belongs to (E +⌊λF⌋)∩I iff r = a and s ∈ F ∩[b ′ ,b] and then|(E +⌊λF⌋)∩I||I|≤ |b−b′ +1| α|⌊λb⌋−⌊λb ′ ⌋| ∼ |b−b′ | αλ·|b−b ′ | = 1λ·|b−b ′ · (4.4.4)| 1−α• a > a ′ : for r + ⌊λs⌋ belong to (E + ⌊λF⌋) ∩ I, we necessarily have r ∈ E ∩ [a ′ ,a] ands ∈ I 2 ∪I 4 ∪···∪I i−1 and so|(E +⌊λF⌋)∩I||I|≤ |E ∩[a′ ,a]|·b i−1|a−a ′ |/2∼ |a−a′ | α ·b i−1|a−a ′ |≤ b i−1µ i1−α ≤ 1b i−1, (4.4.5)where in the last inequality we used (ii).Estimates (4.4.4) and (4.4.5) establish that E +⌊λF⌋ has <strong>zero</strong> upper Banach <strong>density</strong>.4.4.4 Another counterexample <strong>of</strong> Theorem 4.1.1Wenowprovethatonecannotexpectthattheset<strong>of</strong>parametersgivenbyTheorem4.1.1containsintegers. More specifically, we construct a regular set E ⊂ Z such that D(E) = D(E +λE) forevery λ ∈ Z. In particular, if D(E) ∈ (1/2,1), E +λE has <strong>zero</strong> upper Banach <strong>density</strong> and soevery parameter λ ∈ R for which d ∗ (E +⌊λE⌋) > 0 is not an integer.Fixed α ∈ (1/2,1) and c ∈ Z, consider the generalized IP-set associated to the sequencesObserve thatn−1∑i=1k n = c·2 n and d n =n−1∑k i ·d i ≤ c·i=12 i22α +i ≤ c·⌊2 n2 /2α ⌋ , n ≥ 1.(n−1) 22α +n−1∑j=12 j < c·2 (n−1)22α +n < d nfor large n. Also, by Lemma 4.3.6, this set has counting dimension α. Then, if E is thegeneralized IP-set associated tok n = 2 n and d n =⌊2 n2 /2α ⌋ ,thearithmeticsumE+λE isalsoageneralizedIP-set, associatedtothesequences ˜k n = (λ+1)·k nandd n , whichalsohascountingdimensionequaltoα. ByProposition4.4.3, E contains aregularsubset with dimension greater than 1/2. This subset gives the required counterexample.


Chapter 4. Marstrand theorem for sub<strong>sets</strong> <strong>of</strong> integers 504.5 Pro<strong>of</strong>s <strong>of</strong> the TheoremsLetE,F betwo regularcompatible sub<strong>sets</strong><strong>of</strong> Z. Throughouttherest<strong>of</strong>thepro<strong>of</strong>, fixacompactinterval Λ <strong>of</strong> positive real numbers. Given distinct points z = (a,b) and z ′ = (a ′ ,b ′ ) <strong>of</strong> E ×F,letΛ z,z ′ = { λ ∈ Λ; a+⌊λb⌋ = a ′ +⌊λb ′ ⌋ } .Clearly, Λ z,z ′ is empty if b = b ′ . For b ≠ b ′ , it is possible to estimate its Lebesgue measure.Lemma 4.5.1. Let z = (a,b) and z ′ = (a ′ ,b ′ ) be distinct points <strong>of</strong> Z 2 . If Λ z,z ′ ≠ ∅, then(a) m(Λ z,z ′) |b−b ′ | −1 .(b) minΛ·|b−b ′ |−1 ≤ |a−a ′ | ≤ maxΛ·|b−b ′ |+1.Pro<strong>of</strong>. Assume b > b ′ and let λ ∈ Λ z,z ′, say a+⌊λb⌋ = n = a ′ +⌊λb ′ ⌋. Then{n−a ≤ λb < n−a+1n−a ′ ≤ λb ′ < n−a ′ =⇒ a ′ −a−1 < λ(b−b ′ ) < a ′ −a+1+1and so( a ′ −a−1Λ z,z ′ ⊂b−b ′), a′ −a+1b−b ′ ,which proves (a). The second part also follows from the above inclusion, asanda ′ −a−1b−b ′ ≤ minΛ z,z ′ ≤ maxΛ =⇒ a ′ −a ≤ maxΛ·(b−b ′ )+1a ′ −a+1b−b ′≥ maxΛ z,z ′ ≥ minΛ =⇒ a ′ −a ≥ minΛ·(b−b ′ )−1.By item (b) <strong>of</strong> the above lemma, |a−a ′ | ∼ |b−b ′ | whenever Λ z,z ′ ≠ ∅. We point out that,although naive, Lemma 4.5.1 expresses the crucial property <strong>of</strong> transversality that makes thepro<strong>of</strong> work, and all results related to Marstrand’s theorem use a similar idea in one way oranother.Let (I n ) n≥1 and (J n ) n≥1 be sequences <strong>of</strong> intervals satisfying the compatibility conditions <strong>of</strong>Definition 4.4.4. Associated to these intervals, consider, for each pair (n,λ) ∈ N×Λ, the setN n (λ) = { ((a,b),(a ′ ,b ′ )) ∈ ((E ∩I n )×(F ∩J n )) 2 ; a+⌊λb⌋ = a ′ +⌊λb ′ ⌋ }and, for each n ≥ 1, the integral∫∆ n = |N n (λ)|dm(λ).Λ


Chapter 4. Marstrand theorem for sub<strong>sets</strong> <strong>of</strong> integers 51By a double counting, one has the equality∑∆ n = m(Λ z,z ′). (4.5.1)z,z ′ ∈(E∩I n)×(F×J n)Lemma 4.5.2. Let D(E) = α and D(F) = β as above.(a) If α+β < 1, then ∆ n |I n | α+β .(b) If α+β > 1, then ∆ n |I n | 2α+2β−1 .Pro<strong>of</strong>. Using (4.5.1),and then∆ n ===∑z,z ′ ∈(E∩I n)×(F×J n)∑a∈E∩Inb∈F∩Jn∑a∈E∩Inb∈F∩Jn∑a∈E∩Inb∈F∩Jn |I n | α+β ·log|I n|∑s=1log|I n|∑s=1log|I n|∑s=1log|I n|∑s=1∑a ′ ∈E∩In|a−a ′ |∼e sm(Λ z,z ′)∑m(Λ z,z ′)b ′ ∈F∩Jn|b−b ′ |∼e se −s ·(e s ) α ·(e s ) β(e s ) α+β−1(e α+β−1) s∆ n |I n | α+β ·|I n | α+β−1 = |I n | 2α+2β−1 , if α+β > 1, |I n | α+β ·1 = |I n | α+β , if α+β < 1.Pro<strong>of</strong> <strong>of</strong> Theorem 4.1.1. The pro<strong>of</strong> is divided in three parts.Part 1. α+β < 1: fix ε > 0. By Lemma 4.5.2, the set <strong>of</strong> parameters λ ∈ Λ for which|N n (λ)| |I n| α+βhas Lebesgue measure at least m(Λ)−ε. We will prove thatε(4.5.2)|(E +⌊λF⌋)∩(I n +⌊λJ n ⌋)||I n +⌊λJ n ⌋| α+β ε (4.5.3)


Chapter 4. Marstrand theorem for sub<strong>sets</strong> <strong>of</strong> integers 52whenever λ ∈ Λ satisfies (4.5.2). For each (m,n,λ) ∈ Z×Z×Λ, letThenands(m,n,λ) = |{(a,b) ∈ (E ∩I n )×(F ∩J n ); a+⌊λb⌋ = m}|.∑s(m,n,λ) = |E ∩I n |·|F ∩J n | ∼ |I n | α+β (4.5.4)m∈Z∑s(m,n,λ) 2 = |N n (λ)| |I n| α+β· (4.5.5)εm∈ZThe numerator in (4.5.3) is at least the cardinality <strong>of</strong> the set S(n,λ) = {m ∈ Z; s(m,n,λ) > 0},because (E + ⌊λF⌋) ∩ (I n + ⌊λJ n ⌋) contains S(n,λ). By the Cauchy-Schwarz inequality andrelations (4.5.4) and (4.5.5),and so, as |I n +⌊λJ n ⌋| ∼ |I n |,establishing (4.5.3).|S(n,λ)| ≥( ∑s(m,n,λ)m∈Z∑s(m,n,λ) 2m∈Z(|In | α+β) 2|I n | α+βε= ε·|I n | α+β) 2|(E +⌊λF⌋)∩(I n +⌊λJ n ⌋)||I n +⌊λJ n ⌋| α+β |S(n,λ)| ε,|I n | α+βFor each n ≥ 1, let G n ε = {λ ∈ Λ; (4.5.3) holds}. Then m(Λ\G n ε) ≤ ε, and the same holdsfor the setFor each λ ∈ G ε ,G ε = ⋂ n≥1∞⋃G l ε.l=nD α+β (E +⌊λF⌋) ≥ ε =⇒ D(E +⌊λF⌋) ≥ α+βand then, as the set G = ⋃ n≥1 G 1/n ⊂ Λ has Lebesgue measure m(Λ), Part 1 is completed.Part 2. α+β > 1: for a fixed ε > 0, Lemma 4.5.2 implies that the set <strong>of</strong> parameters λ ∈ Λ forwhich|N n (λ)| |I n| 2α+2β−1(4.5.6)ε


Chapter 4. Marstrand theorem for sub<strong>sets</strong> <strong>of</strong> integers 53has Lebesgue measure at least m(Λ)−ε. In this case,|S(n,λ)| ≥( ∑s(m,n,λ)m∈Z∑s(m,n,λ) 2m∈Z(|In | α+β) 2|I n | 2α+2β−1ε= ε·|I n |) 2and then|(E +⌊λF⌋)∩(I n +⌊λJ n ⌋)||I n +⌊λJ n ⌋|The Borel-Cantelli argument is analogous to Part 1. |S(n,λ)||I n | ε.Part 3. α+β = 1: let n ≥ 1. Being regular, E has a regular subset E n ⊂ E, also compatible 5with F, such that D(E)−1/n < D(E n ) < D(E). Then1− 1 n < D(E n)+D(F) < 1and so, by Part 1, there is a full Lebesgue measure set Λ n such thatD(E n +⌊λF⌋) ≥ 1− 1 n , ∀λ ∈ Λ n.The set Λ = ⋂ n≥1 Λ n has full Lebesgue measure as well andD(E +⌊λF⌋) ≥ 1, ∀λ ∈ Λ.Pro<strong>of</strong> <strong>of</strong> Theorem 4.1.3. We also divide it in parts.Part 1. ∑ ki=0 D(E i) ≤ 1: by Theorem 4.1.1,D(E 0 +⌊λ 1 E 1 ⌋) ≥ D(E 0 )+D(E 1 ) , m−a.e. λ 1 ∈ R.To each <strong>of</strong> these parameters, apply Proposition 4.4.3 to obtain a regular subset F λ1 ⊂ E 0 +⌊λ 1 E 1 ⌋ such that5 This may be assumed because <strong>of</strong> relation (4.4.1).D(F λ1 ) = D(E 0 )+D(E 1 ).


Chapter 4. Marstrand theorem for sub<strong>sets</strong> <strong>of</strong> integers 54As E 2 is universal, another application <strong>of</strong> Theorem 4.1.1 guarantees thatD(F λ1 +⌊λ 2 E 2 ⌋) ≥ D(E 0 )+D(E 1 )+D(E 2 ), m−a.e. λ 2 ∈ Rand them, by Fubini’s theorem,D(E 0 +⌊λ 1 E 1 ⌋+⌊λ 2 E 2 ⌋) ≥ D(E 0 )+D(E 1 )+D(E 2 ), m 2 −a.e. (λ 1 ,λ 2 ) ∈ R 2 .Iterating the above arguments, it follows thatD(E 0 +⌊λ 1 E 1 ⌋+···+⌊λ k E k ⌋) ≥ D(E 0 )+···+D(E k ), m k −a.e. (λ 1 ,...,λ k ) ∈ R k .Part 2. ∑ ki=0 D(E i) > 1: without loss <strong>of</strong> generality, we can assumeD(E 0 )+···+D(E k−1 ) ≤ 1 < D(E 0 )+···+D(E k−1 )+D(E k ).By Part 1,D(E 0 +⌊λ 1 E 1 ⌋+···+⌊λ k−1 E k−1 ⌋) ≥ D(E 0 )+···+D(E k−1 )for m k−1 −a.e. (λ 1 ,...,λ k−1 ) ∈ R k−1 . To each <strong>of</strong> these (k−1)-tuples, consider a regular subsetF (λ1 ,...,λ k−1 ) <strong>of</strong> E 0 +···+⌊λ k−1 E k−1 ⌋ such thatD ( F (λ1 ,...,λ k−1 ))= D(E0 )+···+D(E k−1 ).Finally, because D(F (λ1 ,...,λ k−1 ))+D(E k ) > 1, Theorem 4.1.1 guarantees thatd ∗( F (λ1 ,...,λ k−1 ) +⌊λ k E k ⌋ ) > 0 , m−a.e. λ k ∈ R,which, after another application <strong>of</strong> Fubini’s theorem, concludes the pro<strong>of</strong>.4.6 Concluding remarksWe think there is a more specific way <strong>of</strong> defining the counting dimension that encodes theconditions <strong>of</strong> regularity and compatibility. A natural candidate would be a prototype <strong>of</strong> aHausdorff dimension, where one looks to all covers, properly renormalized in the unit interval,and takes a liminf. An alternative definition has appeared in [35]. It would be a naturalprogram to prove Marstrand type results in this context.Another interesting question is to consider arithmetic sums E+λF, where λ ∈ Z. These aregenuine arithmetic sums and, as we saw in Subsection 4.4.4, their dimension may not increase.We think very strong conditions on the <strong>sets</strong> E,F are needed to prove analogous results aboutE +λF for λ ∈ Z.


Chapter 4. Marstrand theorem for sub<strong>sets</strong> <strong>of</strong> integers 55We also think the results obtained in this chapter work to sub<strong>sets</strong> <strong>of</strong> Z k . Given E ⊂ Z k , theupper Banach <strong>density</strong> <strong>of</strong> E is equal tod ∗ |E ∩(I 1 ×···×I k )|(E) = limsup,|I 1 |,...,|I k |→∞ |I 1 ×···×I k |where I 1 ,...,I k run over all intervals <strong>of</strong> Z, the counting dimension <strong>of</strong> E islog|E ∩(I 1 ×···×I k )|D(E) = limsup,|I 1 |,...,|I k |→∞ log|I 1 ×···×I k |where I 1 ,...,I k run over all intervals <strong>of</strong> Z and, for α ≥ 0, the counting α-measure <strong>of</strong> E is|E ∩(I 1 ×···×I k )|H α (E) = limsup|I 1 |,...,|I k |→∞ |I 1 ×···×I k | α ,whereI 1 ,...,I k runover all intervals <strong>of</strong> Z. These quantities satisfy similar properties to those inSubsection 4.2.2. The notion <strong>of</strong> regularity is defined in an analogous manner. For compatibility,we have to take into account the geometry <strong>of</strong> Z k . Two regular sub<strong>sets</strong> E,F ⊂ Z k are compatibleif there exist sequences <strong>of</strong> rectangles R n = I1 n ×··· ×In k and S n = J1 n ×··· ×Jn k, n ≥ 1, suchthat(i) |Ii n| ∼ |Jn i | for every i = 1,2,...,k, and(ii) |E ∩R n | |R n | D(E) and |F ∩S n | |S n | D(F) .We think the theory <strong>of</strong> this chapter may be extended to prove that if E,F ⊂ Z k are two regularcompatible sub<strong>sets</strong>, thenD(E +⌊λF⌋) ≥ min{1,D(E) +D(F)}for Lebesgue almost every λ ∈ R and, if in addition D(E) + D(F) > 1, then E + ⌊λF⌋ haspositive upper Banach <strong>density</strong> for Lebesgue almost every λ ∈ R.


Chapter 4. Marstrand theorem for sub<strong>sets</strong> <strong>of</strong> integers 56


Chapter 5Z d -actions with prescribedtopological and ergodic propertiesWe extend constructions <strong>of</strong> Hahn-Katznelson [17] and Pavlov [40] to Z d -actions on symbolicdynamical spaces with prescribed topological and ergodic properties. More specifically, wedescribe a method to build Z d -actions which are (totally) minimal, (totally) strictly ergodicand have positive topological entropy.5.1 IntroductionErgodic theory studies statistical and recurrence properties <strong>of</strong> measurable transformations Tacting in a probability space (X,B,µ), where µ is a measure invariant by T, that is, µ(T −1 A) =µ(A), forallA ∈ B. Itinvestigates awideclass<strong>of</strong>notions, suchasergodicity, mixingandentropy.These properties, in some way, give qualitative and quantitative aspects <strong>of</strong> the randomness <strong>of</strong> T.For example, ergodicity means that T is indecomposable in the metric sense with respect to µand entropy is a concept that counts the exponential growth rate for the number <strong>of</strong> statisticallysignificant distinguishable orbit segments.In most cases, the object <strong>of</strong> study has topological structures: X is a compact metric space,B is the Borel σ-algebra <strong>of</strong> X, µ is a Borel measure probability and T is a homeomorphism <strong>of</strong> X.In this case, concepts such as minimality and topological mixing give topological aspects <strong>of</strong> therandomness <strong>of</strong> T. For example, minimality means that T is indecomposable in the topologicalsense, that is, the orbit <strong>of</strong> every point is dense in X.A natural question arises: how do ergodic and topological concepts relate to each other?How do ergodic properties forbid topological phenomena and vice-versa? Are metric and topologicalindecomposability equivalent? This last question was answered negatively in [15] via57


Chapter 5. Z d -actions with prescribed topological and ergodic properties 58the construction <strong>of</strong> a minimal diffeomorphism <strong>of</strong> the torus T 2 which preserves area but is notergodic.AnotherquestionwasraisedbyW.Parry: supposeT hasauniqueBorelprobabilityinvariantmeasure and that (X,T) is a minimal transformation. Can (X,T) have positive entropy? Thedifficulty in answering this at the time was the scarcity <strong>of</strong> a wide class <strong>of</strong> minimal and uniquelyergodictransformations. Thiswas solved affirmatively in [17], whereF. HahnandY. Katznelsondeveloped an inductive method <strong>of</strong> constructing symbolic dynamical systems with the requiredproperties. The principal idea <strong>of</strong> the paper was the weak law <strong>of</strong> large numbers.Later, works <strong>of</strong> Jewett and Krieger (see [42]) proved that every ergodic measure-preservingsystem (X,B,µ,T) is metrically isomorphic to a minimal and uniquely ergodic homeomorphismon a Cantor set and this gives many examples to Parry’s question: if an ergodic system(X,B,µ,T) has positive metric entropy and Φ : (X,B,µ,T) → (Y,C,ν,S) is the metric isomorphismobtained by Jewett-Krieger’s theorem, then (Y,S) has positive topological entropy, bythe variational principle.It is worth mentioning that the situation is quite different in smooth ergodic theory, oncesome regularity on the transformation is assumed. A. Katok showed in [21] that every C 1+αdiffeomorphism <strong>of</strong> a compact surface can not be minimal and simultaneously have positivetopological entropy. More specifically, he proved that the topological entropy can be written interms <strong>of</strong> the exponential growth <strong>of</strong> periodic points <strong>of</strong> a fixed order.Suppose that T is a mesure-preserving transformation on the probability space (X,B,µ)and f : X → R is a measurable function. A successful area in ergodic theory deals withthe convergence <strong>of</strong> averages n −1·∑nk=1 f ( T k x ) , x ∈ X, when n converges to infinity. The wellknown Birkh<strong>of</strong>f’s Theorem states that such limit exists for almost every x ∈ X whenever f is anL 1 -function. Several resultshave been(and still are being) proved when, instead <strong>of</strong> {1,2,...,n},average is made along other sequences <strong>of</strong> natural numbers. A remarkable result on this directionwasgivenbyJ.Bourgain[6], whereheprovedthatifp(x)isapolynomialwithintegercoefficientsand f is an L p -function, for some p > 1, then the averages n −1 ·∑nk=1 f ( T p(k) x ) converge foralmost every x ∈ X. In other words, convergence fails to hold for a negligible set with respectto the measure µ. We mention this result is not true for L 1 -functions, as recently proved by Z.Buczolich and D. Mauldin [7].In [3], V. Bergelson asked if this set is also negligible from the topological point <strong>of</strong> view. Itturnedout, by aresult <strong>of</strong> R. Pavlov [40], that this is not true. Heproved that, for every sequence(p n ) n≥1 ⊂ Z<strong>of</strong><strong>zero</strong>upper-Banach<strong>density</strong>, thereexistatotally minimal, totally uniquelyergodicand topologically mixing transformation (X,T) and a continuous function f : X → R such thatn −1 ·∑nk=1 f (Tp kx) fails to converge for a residual set <strong>of</strong> x ∈ X.Suppose now that (X,T) is totally minimal, that is, (X,T n ) is minimal for every positive


Chapter 5. Z d -actions with prescribed topological and ergodic properties 59integer n. Pavlov alsoprovedthat, foreverysequence(p n ) n≥1 ⊂ Z<strong>of</strong><strong>zero</strong>upper-Banach<strong>density</strong>,there exists a totally minimal, totally uniquely ergodic and topologically mixing continuoustransformation (X,T) such that x ∉ {T pn x; n ≥ 1} for an uncountable number <strong>of</strong> x ∈ X.In this work, we extend the results <strong>of</strong> Hahn-Katznelson and Pavlov, giving a method <strong>of</strong> constructing(totally) minimal and (totally) uniquely ergodic Z d -actions with positive topologicalentropy. We carry out our program by constructing closed shift invariant sub<strong>sets</strong> <strong>of</strong> a sequencespace. More specifically, we build a sequence <strong>of</strong> finite configurations (C k ) k≥1 <strong>of</strong> {0,1} Zd , C k+1being essentially formed by the concatenation <strong>of</strong> elements in C k such that each <strong>of</strong> them occursstatistically well-behaved in each element <strong>of</strong> C k+1 , and consider the set <strong>of</strong> limits <strong>of</strong> shiftedC k -configurations as k → ∞. The main results areTheorem 5.1.1. There exist totally strictly ergodic Z d -actions (X,B,µ,T) with arbitrarily largepositive topological entropy.We should mention that this result is not new, because Jewett-Krieger’s Theorem is true forZ d -actions [46]. This formulation emphasizes to the reader that the constructions, which maybe used in other settings, have the additional advantage <strong>of</strong> controlling the topological entropy.Theorem 5.1.2. Given a set P ⊂ Z d <strong>of</strong> <strong>zero</strong> upper-Banach <strong>density</strong>, there exist a totally strictlyergodic Z d -action (X,B,µ,T) and a continuous function f : X → R such that the ergodicaverages1|P ∩(−n,n) d |∑g∈P∩(−n,n) d f (T g x)fail to converge for a residual set <strong>of</strong> x ∈ X. In addition, (X,B,µ,T) can have arbitrarily largetopological entropy.The above theorem has a special interest when P is an arithmetic set for which classicalergodic theory and Fourier analysis have established almost-sure convergence. This is the case(also proved in [6]) whenP = {(p 1 (n),...,p d (n)); n ∈ Z},where p 1 ,...,p d are polynomials with integer coefficients: for any f ∈ L p , p > 1, the limit1n∑ ( )lim f T (p 1(k),...,p d (k)) xn→∞ nk=1exists almost-surely. Note that P has <strong>zero</strong> upper-Banach <strong>density</strong> whenever one <strong>of</strong> the polynomialshas degree greater than 1.Theorem 5.1.3. Given a set P ⊂ Z d <strong>of</strong> <strong>zero</strong> upper-Banach <strong>density</strong>, there exists a totally strictlyergodic Z d -action (X,B,µ,T) and an uncountable set X 0 ⊂ X for which x ∉ {T p x; p ∈ P}, forevery x ∈ X 0 . In addition, (X,B,µ,T) can have arbitrarily large topological entropy.


Chapter 5. Z d -actions with prescribed topological and ergodic properties 60Yet in the arithmetic setup, Theorem 5.1.3 is the best topological result one can expect.Indeed, Bergelson and Leibman proved in [4] that if T is a minimal Z d -action, then there is aresidual set Y ⊂ X for which x ∈ {T (p 1(n),...,p d (n)) x; n ∈ Z}, for every x ∈ Y.5.2 PreliminariesWe begin with some notation. Consider a metric space X, B its Borel σ-algebra and a G groupwith identity e. Throughout this work, G will denote Z d , d > 1, or one <strong>of</strong> its subgroups.5.2.1 Group actionsDefinition 5.2.1. A G-action on X is a measurable transformation T : G×X → X, denotedby (X,T), such that(i) T(g 1 ,T(g 2 ,x)) = T(g 1 g 2 ,x), for every g 1 ,g 2 ∈ G and x ∈ X.(ii) T(e,x) = x, for every x ∈ X.In other words, for each g ∈ G, the restrictionT g : X −→ Xx ↦−→ T(g,x)is a bimeasurable transformation on X such that T g 1g 2= T g 1T g 2, for every g 1 ,g 2 ∈ G. WhenG is abelian, (T g ) g∈G forms a commutative group <strong>of</strong> bimeasurable transformations on X. Foreach x ∈ X, the orbit <strong>of</strong> X with respect to T is the setO T (x) = . {T g x; g ∈ G}.If F is a subgroup <strong>of</strong> G, the restriction T| F : F ×X → X is clearly a F-action on X.Definition 5.2.2. We say that (X,T) is minimal if O T (x) is dense in X, for every x ∈ X, andtotally minimal if O T|F (x) is dense in X, for every x ∈ X and every subgroup F < G <strong>of</strong> finiteindex.Remind that the index <strong>of</strong> a subgroup F, denoted by (G : F), is the number <strong>of</strong> co<strong>sets</strong> <strong>of</strong> F inG. The above definition extends the notion <strong>of</strong> total minimality <strong>of</strong> Z-actions. In fact, a Z-action(X,T) is totally minimal if and only if T n : X → X is a minimal transformation, for everyn ∈ Z.Consider the set M(X) <strong>of</strong> all Borel probability measures in X. A probability µ ∈ M(X) isinvariant under T or simply T-invariant ifµ(T g A) = µ(A), ∀g ∈ G, ∀A ∈ B.


Chapter 5. Z d -actions with prescribed topological and ergodic properties 61Let M T (X) ⊂ M(X) denote the set <strong>of</strong> all T-invariant probability measures. Such set isnon-empty whenever G is amenable, by a Krylov-Bogolubov argument applied to any Fφlnersequence 1 <strong>of</strong> G.Definition 5.2.3. A G measure-preserving system or simply G-mps is a quadruple (X,B,µ,T),where T is a G-action on X and µ ∈ M T (X).We say that A ∈ B is T-invariant if T g A = A, for all g ∈ G.Definition 5.2.4. The G-mps (X,B,µ,T) is ergodic if it has only trivial invariant <strong>sets</strong>, that is,if µ(A) = 0 or 1 whenever A is a measurable set invariant under T.Definition 5.2.5. The G-action (X,T) is uniquely ergodic if M T (X) is unitary, and totallyuniquely ergodic if, for every subgroup F < G <strong>of</strong> finite index, the restricted F-action (X,T| F )is uniquely ergodic.Definition 5.2.6. We say that (X,T) is strictly ergodic if it is minimal and uniquely ergodic,and totally strictly ergodic if, for every subgroup F < G <strong>of</strong> finite index, the restricted F-action(X,T| F ) is strictly ergodic.The result below was proved in [47] and states the pointwise ergodic theorem for Z d -actions.Theorem 5.2.7. Let (X,B,µ,T) be a Z d -mps. Then, for every f ∈ L 1 (µ), there is a T-invariant function ˜f ∈ L 1 (µ) such thatlimn→∞1 ∑n d f (T g x) = ˜f(x)g∈[0,n) dfor µ-almost every x ∈ X. In particular, if the action is ergodic, ˜f is constant and equal to∫fdµ.Above, [0,n) denotestheset{0,1,...,n−1}, [0,n) d thed-dimensional cube[0,n)×···×[0,n)<strong>of</strong> Z d and by a T-invariant function we mean that f ◦ T g = f, for every g ∈ G. These averagesallow the characterization <strong>of</strong> unique ergodicity. Let C(X) denote the space <strong>of</strong> continuousfunctions from X to R.Proposition 5.2.8. Let (X,T) be a Z d -action on the compact metric space X. The followingitems are equivalent.(a) (X,T) is uniquely ergodic.1 (A n) n≥1 is a Fφlner sequence <strong>of</strong> G if lim n→∞|A n∆gA n|/|A n| = 0 for every g ∈ G.


Chapter 5. Z d -actions with prescribed topological and ergodic properties 62(b) For every f ∈ C(X) and x ∈ X, the limitexists and is independent <strong>of</strong> x.limn→∞(c) For every f ∈ C(X), the sequence <strong>of</strong> functions1 ∑n d f (T g x)g∈[0,n) df n = 1 n d ∑converges uniformly in X to a constant function.g∈[0,n) d f ◦T gPro<strong>of</strong>. The implications (c)⇒(b)⇒(a) are obvious. It remains to prove (a)⇒(c). Let M T (X) ={µ}. We’ll show that f n converges uniformly to ˜f = ∫ fdµ. By contradiction, suppose this isnot the case for some f ∈ C(X). This means that there exist ε > 0, n i → ∞ and x i ∈ X suchthat ∣ ∣∣∣f ni (x i )−∫fdµ∣ ≥ ε.For each i, let ν i ∈ M(X) be the probability measure associated to the linear functionalΘ i : C(X) → R defined byΘ i (ϕ) = 1n id∑g∈[0,n i ) d ϕ(T g x i ), ϕ ∈ C(X).Restricting to a subsequence, if necessary, we assume that ν i → ν in the weak-star topology.Because the cubes A i = [0,n i ) d form a Fφlner sequence in Z d , ν ∈ M T (X). In fact, for eachh ∈ Z d ,(∣∣∫ϕ◦T h) ∫dν −ϕdν∣ = lim1i→∞ n d i∣∑g∈A i +hϕ(T g x i )− ∑ ϕ(T g x i )g∈A i∣#A i ∆(A i +h)≤ max|ϕ(x)|· limx∈X i→∞ #A i= 0.But∣ ∣∣∣∫∫fdν −∫fdµ∣ = lim∣i→∞∫fdν i −∫fdµ∣ = lim∣ f n i(x i )−i→∞fdµ∣ ≥ εand so ν ≠ µ, contradicting the unique ergodicity <strong>of</strong> (X,T).


Chapter 5. Z d -actions with prescribed topological and ergodic properties 635.2.2 Subgroups <strong>of</strong> Z dLet F be the set <strong>of</strong> all subgroups <strong>of</strong> Z d <strong>of</strong> finite index. This set is countable, because eachelement <strong>of</strong> F is generated by d linearly independent vectors <strong>of</strong> Z d . Consider, then, a subset(F k ) k∈N <strong>of</strong> F such that, for each F ∈ F, there exists k 0 > 0 such that F k < F, for everyk ≥ k 0 . For this, just consider an enumeration <strong>of</strong> F and define F k as the intersection <strong>of</strong> thefirst k elements. Such intersections belong to F because(Z d : F ∩F ′ ) ≤ (Z d : F)·(Z d : F ′ ), ∀F,F ′ < Z d .Restricting them, if necessary, we assume that F k = m k · Z d , where (m k ) k≥1 is an increasingsequence <strong>of</strong> positive integers. Observe that (Z d : F k ) = m d k . Such sequence will be fixedthroughout the rest <strong>of</strong> the chapter.Definition 5.2.9. Given asubgroupF < Z d , wesaythattwoelements g 1 ,g 2 ∈ Z d arecongruentmodulo F if g 1 −g 2 ∈ F and denote it by g 1 ≡ F g 2 . The set ¯F ⊂ Z d is a complete residue setmodulo F if, for every g ∈ Z d , there exists a unique h ∈ ¯F such that g ≡ F h.Every complete residue set modulo F is canonically identified to the quocient Z d /F and hasexactly (Z d : F) elements.5.2.3 Symbolic spacesLetC beafinitealphabet andconsidertheset Ω(C) = C Zd <strong>of</strong> all functionsx : Z d → C. We endowC withthediscrete topology andΩ(C) withtheproducttopology. By Tychon<strong>of</strong>f’s theorem, Ω(C)is a compact metric space. We are not interested in a particular metric in Ω(C). Instead, weconsider a basis <strong>of</strong> topology B 0 to be defined below.Consider the family R <strong>of</strong> all finite d-dimensional cubes A = [r 1 ,r 1 +n)×··· ×[r d ,r d +n)<strong>of</strong> Z d , n ≥ 0. We say that A has length n and is centered at g = (r 1 ,...,r d ) ∈ Z d .Definition 5.2.10. A configuration or pattern is a pair b A = (A,b), where A ∈ R and b is afunction from A to C. We say that b A is supported in A with encoding function b.Let Ω A (C) denote the space <strong>of</strong> configurations supported in A and Ω ∗ (C) the space <strong>of</strong> allconfigurations in Z d :Ω ∗ (C) = . {b A ; b A is a configuration}.Given A ∈ R, consider the map Π A : Ω(C) → Ω A (C) defined by the restrictionΠ A (x) : A −→ Cg ↦−→ x(g)In particular, Π {g} (x) = x(g). We use the simpler notation x| A to denote Π A (x).


Chapter 5. Z d -actions with prescribed topological and ergodic properties 64Definition 5.2.11. If A ∈ R is centered at g, we say that x| A is a configuration <strong>of</strong> x centeredat g or that x| A occurs in x centered at g.For A 1 ,A 2 ∈ R such that A 1 ⊂ A 2 , let π A 2A 1: Ω A2 → Ω A1 be the restrictionπ A 2A 1(b) : A 1 −→ Cg ↦−→ b(g)As above, when there is no ambiguity, we denote π A 2A 1(b) simply by b| A1 . It is clear that thediagram below commutes.Π A2Ω(C) Ω A2 (C)Π A1π A 2A 1Ω A1 (C)These maps will help us to control the patterns to appear in the constructions <strong>of</strong> Section 5.3.By a cylinder in Ω(C) we mean the set <strong>of</strong> elements <strong>of</strong> Ω(C) with some fixed configuration.More specifically, given b A ∈ Ω ∗ (C), the cylinder generated by b A is the setCyl(b A ) . = {x ∈ Ω(C); x| A = b A }.The family B 0 := {Cyl(b A )|b A ∈ Ω ∗ (C)} forms a clopen set <strong>of</strong> cylinders generating B. Hencethe set C 0 = {χ B ; B ∈ B 0 } <strong>of</strong> cylinder characteristic functions generates a dense subspace inC(Ω(C)). Let µ be the probability measure defined byµ(Cyl(b A )) = |C| −|A| , ∀b A ∈ Ω ∗ (C),and extended to B by Caratheódory’s Theorem. Above, |·| denotes the number <strong>of</strong> elements <strong>of</strong>a set.Consider the Z d -action T : Z d ×Ω(C) → Ω(C) defined byT g (x) = (x(g +h)) h∈Z d ,also called the shift action. Given B = Cyl(b A ) and g ∈ Z d , let B + g denote the cylinderassociated to b A+g = (˜b,A + g), where ˜b : A + g → {0,1} is defined by ˜b(h) = b(h − g),∀h ∈ A+g. With this notation,χ B ◦T g = χ B+g . (5.2.1)


Chapter 5. Z d -actions with prescribed topological and ergodic properties 65In fact,⇐⇒⇐⇒⇐⇒⇐⇒χ B (T g x) = 1T g x ∈ Bx(g +h) = b(h), ∀h ∈ Ax(h) = ˜b(h), ∀h ∈ A+gx ∈ B +g.Definition 5.2.12. A subshift <strong>of</strong> (Ω(C),T) is a Z d -action (X,T), where X is a closed subset <strong>of</strong>Ω(C) invariant under T.5.2.4 Topological entropyFor each subset X <strong>of</strong> Ω(C) and A ∈ R, letΩ A (C,X) = {x| A ; x ∈ X}denote the set <strong>of</strong> configurations supported in A which occur in elements <strong>of</strong> X and Ω ∗ (C,X) thespace <strong>of</strong> all configurations in Z d occuring in elements <strong>of</strong> X,Ω ∗ (C,X) = ⋃Ω A (C,X).A∈RDefinition 5.2.13. The topological entropy <strong>of</strong> the subshift (X,T) is the limitlog|Ω [0,n) d(C,X)|h(X,T) = limn→∞ n d , (5.2.2)which always exists and is equal to inf n∈N1n d ·log|Ω [0,n) d(C,X)|.5.2.5 Frequencies and unique ergodicityDefinition 5.2.14. Given configurations b A1 ∈ Ω A1 (C) and b A2 ∈ Ω A2 (C), the set <strong>of</strong> ocurrences<strong>of</strong> b A1 in b A2 isS(b A1 ,b A2 ) . = {g ∈ Z d ; A 1 +g ⊂ A 2 and π A 2A 1 +g (b A 2) = b A1 +g}.The frequency <strong>of</strong> b A1 in b A2 is defined asfr(b A1 ,b A2 ) . = |S(b A 1,b A2 )||A 2 |Given F ∈ F and h ∈ Z d , the set <strong>of</strong> ocurrences <strong>of</strong> b A1 in b A2 centered at h modulo F isS(b A1 ,b A2 ,h,F) . = {g ∈ S(b A1 ,b A2 ); A 1 +g is centered at a vertex ≡ F h}·


Chapter 5. Z d -actions with prescribed topological and ergodic properties 66and the frequency <strong>of</strong> b A1 in b A2 centered at h modulo F is the quocientfr(b A1 ,b A2 ,h,F) . = |S(b A 1,b A2 ,h,F)||A 2 |Observe that if ¯F ⊂ Z d is a complete residue set modulo F, thenfr(b A1 ,b A2 ) = ∑ g∈¯Ffr(b A1 ,b A2 ,g,F).·To our purposes, we rewrite Proposition 5.2.8 in a different manner.Proposition 5.2.15. A subshift (X,T) is uniquely ergodic if and only if, for every b A ∈ Ω ∗ (C)and x ∈ X,exists and is independent <strong>of</strong> x.fr(b A ,x) = . ( )lim fr b A ,x|n→∞ [0,n) dPro<strong>of</strong>. By approximation, condition (b) <strong>of</strong> Proposition 5.2.8 holds for C(X) if and only if itholds for C 0 = {χ B ; B ∈ B 0 }. If f = χ Cyl(bA ), (5.2.1) implies thatlim f 1 ∑n(x) = limn→∞ n→∞ n d f (T g x)g∈[0,n) d1 ∑= limn→∞ n d χ Cyl(bA+g) (x)g∈[0,n) d1 ∑= limn→∞ n d χ Cyl(bA+g) (x)g∈[0,n) dA+g⊂[0,n)(d )= lim fr b A ,x|n→∞ [0,n) d= fr(b A ,x),where in the third equality we used that, for a fixed A ∈ R,|{g ∈ [0,n) d ; A+g ⊄ [0,n) d }|limn→∞ n d = 0.Corollary 5.2.16. A subshift (X,T) is totally uniquely ergodic if and only if, for every b A ∈Ω ∗ (C), x ∈ X and F ∈ F,exists and is independent <strong>of</strong> x.fr(b A ,x,F) = . ( )lim fr b A ,x|n→∞ [0,n) d,0,FSo, uniqueergodicity is all aboutconstant frequencies. We’ll obtain thisviatheLaw <strong>of</strong>LargeNumbers, equidistributing ocurrences <strong>of</strong> configurations along residue classes <strong>of</strong> subgroups.


Chapter 5. Z d -actions with prescribed topological and ergodic properties 675.2.6 Law <strong>of</strong> Large NumbersIntuitively, if A is a subset <strong>of</strong> Z d , each letter <strong>of</strong> C appears in x| A with frequency approximately1/|C|, for almost every x ∈ Ω(C). This is what the Law <strong>of</strong> Large Number says. For our purposes,we state this result in a slightly different way. Let (X,B,µ) be a probability space and A ⊂ Z dinfinite. For each g ∈ A, let X g : X → R be a random variable.Theorem 5.2.17. (Law <strong>of</strong> Large Numbers) If (X g ) g∈A is a family <strong>of</strong> independent and identicallydistributed random variables such that E[X g ] = m, for every g ∈ A, then the sequence ( X n)n≥1defined byX n =∑g∈A∩[0,n) d X g|A∩[0,n) d |converges in probability to m, that is, for any ε > 0,lim µ(∣ ∣ Xn −m ∣ ) < ε = 1.n→∞Consider the probability measure space (X,B,µ) defined in Subsection 5.2.3. Fixed w ∈ C,let X g : Ω(C) → R be defined asX g (x) = 1 , if x(g) = w= 0 , if x(g) ≠ w.(5.2.3)It is clear that (X g ) g∈Z d are independent, identically distributed and satisfy∫E[X g ] = X g (x)dµ(x) = 1|C| , ∀g ∈ Zd .In addition,which implies theX n (x) =X∑g∈[0,n) d X g(x)n d =(∣ ∣∣∣S w,x| [0,n) d)∣Corollary 5.2.18. Let w ∈ C, g ∈ Z d , F ∈ F and ε > 0.(a) The number <strong>of</strong> elements b ∈ Ω [0,n) d(C) such that∣ fr(w,b)− 1|C| ∣ < εis asymptotic to |C| nd as n → ∞.(b) The number <strong>of</strong> elements b ∈ Ω [0,n) d(C) such that∣ fr(w,b,g,F)− 1|C|·(Z d : F) ∣ < εis asymptotic to |C| nd as n → ∞.n d( )= fr w,x| [0,n) d ,


Chapter 5. Z d -actions with prescribed topological and ergodic properties 68(c) The number <strong>of</strong> elements b ∈ Ω [0,n) d(C) such that∣ fr(w,b,g,F)− 1|C|·(Z d : F) ∣ < εfor every w ∈ C and g ∈ Z d is asymptotic to |C| nd as n → ∞.Pro<strong>of</strong>. (a) The required number is equal to(|C| nd ·µ {x ∈ Ω(C);∣∣fr( ∣ )w,x| [0,n) d)−|C| −1 ∣∣ < ε}and is asymptotic to |C| nd , as the above µ-measure converges to 1.(b) Take A = F +g and (X h ) h∈A as in (5.2.3). For any x ∈ Ω(C),(∣ ∣∣∣S w,x| [0,n) d,g,F)∣X n (x) =|A∩[0,n) d |( )= fr w,x| [0,n) d,g,F ·(Z d : F)+o(1),because ∣ ∣ A∩[0,n)d is asymptotic to n d /(Z d : F). This implies that for n large)∣ ∣(w,x| fr 1∣∣∣[0,n) d,g,F −|C|·(Z d : F) ∣ < ε ⇐⇒ X n (x)− 1|C| ∣ < ε·(Zd : F)and then Theorem 5.2.17 guarantees the conclusion.(c) As the events are independent, this follows from (b).5.3 Main ConstructionsLet C = {0,1}. In this section, we construct subshifts (X,T) with topological and ergodic prescribedproperties.To this matter, we buildasequence<strong>of</strong> finitenon-empty <strong>sets</strong> <strong>of</strong> configurationsC k ⊂ Ω Ak (C), k ≥ 1, such that:(i) A k = [0,n k ) d , where (n k ) k≥1 is an increasing sequence <strong>of</strong> positive integers.(ii) n 1 = 1 and C 1 = Ω A1 (C) ∼ = {0,1}.(iii) C k is the concatenation <strong>of</strong> elements <strong>of</strong> C k−1 , possibly with the insertion <strong>of</strong> few additionalblocks <strong>of</strong> <strong>zero</strong>es and ones.Given such sequence (C k ) k≥1 , we consider X ⊂ Ω(C) as the set <strong>of</strong> limits <strong>of</strong> shifted C k -patternsas k → ∞, that is, x ∈ X if there exist sequences (w k ) k≥1 , w k ∈ C k , and (g k ) k≥1 ⊂ Z d such thatx = limk→∞ Tg kw k .


Chapter 5. Z d -actions with prescribed topological and ergodic properties 69The above limit has an abuse <strong>of</strong> notation, because T acts in Ω(C) and w k ∉ Ω(C). Formallyspeaking, this means that, for each g ∈ Z d , there exists k 0 ≥ 1 such thatx(g) = w k (g +g k ), ∀k ≥ k 0 .By definition, X is invariant under T and, for any k, every x ∈ X is an infinite concatenation<strong>of</strong> elements <strong>of</strong> C k and additional blocks <strong>of</strong> <strong>zero</strong>es and ones.If C k ⊂ Ω Ak ({0,1}) and A ∈ R, Ω A (C k ) is identified in a natural way to a subset <strong>of</strong>Ω nk A({0,1}). In some situations, to distinguish this association, we use small letters for Ω A (C k )and capital letters for Ω nk A({0,1}) 2 . In this situation, if w ∈ Ω A (C k ) and g ∈ A, the patternw(g) ∈ C k occurs in W ∈ Ω nk A({0,1}) centered at n k g. In other words, if w k ∈ C k , thenS(w k ,W,n k g,F) = n k ·S(w k ,w,g,F). (5.3.1)In each <strong>of</strong> the next subsections, (C k ) k≥1 is constructed with specific combinatorial andstatistical properties.5.3.1 MinimalityThe action (X,T) is minimal if and only if, for each x,y ∈ X, every configuration <strong>of</strong> x is also aconfiguration <strong>of</strong> y. For this, suppose C k ⊂ Ω Ak ({0,1}) is defined and non-empty.By the Law <strong>of</strong> Large Numbers, if l k is large, every element <strong>of</strong> C k occurs in almost everyelement <strong>of</strong> Ω [0,lk ) d(C k) (in fact, by Corollary 5.2.18, each <strong>of</strong> them occurs approximately withfrequency 1/|C k | > 0). Take any subset C k+1 <strong>of</strong> Ω [0,lk ) d(C k) with this property and consider itas a subset <strong>of</strong> Ω [0,nk+1 ) d({0,1}), where n k+1 = l k n k .Let us prove that (X,T) is minimal. Consider x,y ∈ X and x| A a finite configuration <strong>of</strong> x.For large k, x| A is a subconfiguration <strong>of</strong> some w k ∈ C k . As y is formed by the concatenation <strong>of</strong>elements <strong>of</strong> C k+1 , every element <strong>of</strong> C k is a configuration <strong>of</strong> y. In particular, w k (and then x| A )is a configuration <strong>of</strong> y.5.3.2 Total minimalityThe action (X,T) is totally minimal if and only if, for each x,y ∈ X and F ∈ F, everyconfiguration x| A <strong>of</strong> x centered 3 at 0 also occurs in y centered at some g ∈ F. To guarantee thisfor every F ∈ F, we inductively control the ocurrence <strong>of</strong> subconfigurations centered in finitelymany subgroups <strong>of</strong> Z d .2 For example, w ∈ Ω A(C k ) and W ∈ Ω nk A({0,1}) denote the “same” element.3 Because <strong>of</strong> the T-invariance <strong>of</strong> X, we can suppose that x| A is centered in 0 ∈ Z d . In fact, instead <strong>of</strong> x,y, weconsider T g x,T g y.


Chapter 5. Z d -actions with prescribed topological and ergodic properties 70Consider the sequence (F k ) ⊂ F defined in Subsection 5.2.2. By induction, suppose C k ⊂Ω Ak ({0,1}) is non-empty satisfying (i), (ii), (iii) and the additional assumption(iv) gcd(n k ,m k ) = 1 (observe that this holds for k = 1).Take l k large and ˜C k+1 ⊂ Ω [0,lk m k+1 ) d(C k) non-empty such that(v) S(w k ,w| [0,lk m k+1 −1) d,g,F k) ≠ ∅, for every triple (w k ,w,g) ∈ C k × ˜C k+1 ×Z d .Considering w| [0,lk m k+1 −1) d as an element <strong>of</strong> Ω [0,l k m k+1 n k −n k ) d({0,1}), (5.3.1) implies thatS(w k ,W| [0,lk m k+1 n k −n k ) d,n kg,F k ) ≠ ∅, ∀(w k ,w,g) ∈ C k × ˜C k+1 ×Z d .As gcd(n k ,m k ) = 1, the set n k Z d runs over all residue classes modulo F k and so (the restrictionto [0,l k m k+1 n k −n k ) d <strong>of</strong>) every element <strong>of</strong> ˜C k+1 contains every element <strong>of</strong> C k centered at everyresidue class modulo F k .Obviously, C k+1 must not be equal to ˜C k+1 , because m k+1 divides l k m k+1 n k . Instead, wetake n k+1 = l k m k+1 n k + 1 and insert positions B i , i = 1,2,...,d, next to faces <strong>of</strong> the cube[0,l k m k+1 n k ) d . These are given byB i = {(r 1 ,...,r d ) ∈ A k+1 ; r i = l k m k+1 n k −n k }.There is a natural surjection Φ : Ω Ak+1 ({0,1}) → Ω [0,nk+1 −1) d({0,1}) obtained removing thepositions B 1 ,...,B d . More specifically, ifδ(r) = 0, if r < l k m k+1 n k −n k ,= 1, otherwiseand∆(r 1 ,...,r d ) = (δ(r 1 ),...,δ(r d )), (5.3.2)the map Φ is given byΦ(W)(g) = W (g +∆(g)), ∀(r 1 ,...,r d ) ∈ [0,n k+1 −1) d .We conclude the induction step taking C k+1 = Φ −1 (˜C k+1 ).


Chapter 5. Z d -actions with prescribed topological and ergodic properties 717 8 94 5 6Φ7 8 94 5 61 2 31 2 3Φ(w k+1 ) ∈ ˜C k+1 w k+1 ∈ C k+1Figure: example <strong>of</strong> Φ when d = 2.By definition, w k+1 and Φ(w k+1 ) coincide in [0,n k+1 −n k −1) d , for every w k+1 ∈ C k+1 . Thisimplies that every element <strong>of</strong> C k appears in every element <strong>of</strong> C k+1 centered at every residue classmodulo F k .Let us prove that (X,T) is totally minimal. Fix elements x,y ∈ X, a subgroup F ∈ F anda pattern x| A <strong>of</strong> x centered in 0 ∈ Z d . By the definition <strong>of</strong> X, x| A is a subconfiguration <strong>of</strong>some w k ∈ C k , for k large enough such that F k < F. As y is built concatenating elements <strong>of</strong>C k+1 , w k occurs in y centered in every residue class modulo F and the same happens to x| A . Inparticular, x| A occurs in y centered in some g ∈ F, which is exactly the required condition.5.3.3 Total strict ergodicityIn addition to the occurrence <strong>of</strong> configurations in every residue class <strong>of</strong> subgroups <strong>of</strong> Z d , wealso control their frequency. Consider a sequence (d k ) k≥1 <strong>of</strong> positive real numbers such that∑k≥1 d k < ∞. Assume that C 1 ,...,C k−1 ,C k are non-empty <strong>sets</strong> satisfying (i), (ii), (iii), (iv)and(vi) For every (w k−1 ,w k ,g) ∈ C k−1 ×C k ×Z d ,(1−dk−1fr(w k−1 ,w k ,g,F k−1 ) ∈m k−1d ·|C k−1 |,)1+d k−1·m k−1d ·|C k−1 |Before going to the inductive step, let us make an observation. Condition (vi) also controls thefrequency on subgroups F such that F k−1 < F. In fact, if ¯F k−1 is a complete residue set moduloF k−1 ,fr(w k−1 ,w k ,g,F) = ∑h∈ ¯F k−1h≡ F gfr(w k−1 ,w k ,h,F k−1 ) (5.3.3)and, as ∣ {h ∈ ¯Fk−1 ; h ≡ F g} ∣ = (F : Fk−1 ),(1−d k−1fr(w k−1 ,w k ,g,F) ∈(Z d : F)·|C k−1 |,)1+d k−1(Z d · (5.3.4): F)·|C k−1 |


Chapter 5. Z d -actions with prescribed topological and ergodic properties 72Weproceedthesamewayasintheprevioussubsection: takel k largeand ˜C k+1 ⊂ Ω [0,lk m k+1 ) d(C k)non-empty such thatfr(w k , ˜w k+1 ,g,F k ) ∈( 1−dkm kd ·|C k |,)1+d km kd ·|C k |(5.3.5)for every (w k , ˜w k+1 ,g) ∈ C k × ˜C k+1 ×Z d . Note that the non-emptyness <strong>of</strong> ˜C k+1 is guaranteed byCorollary 5.2.18. Also, let n k+1 = l k m k+1 n k +1 and C k+1 = Φ −1 (˜C k+1 ).Fix b A ∈ Ω ∗ (C). Using the big-O notation, we havefr(b A ,W k+1 ,g,F)−fr(b A ,W k+1 | [0,nk+1 −n k −1) d,g,F) = O(1/l k), (5.3.6)because these two frequencies differ by the frequency <strong>of</strong> b A in [n k+1 −n k −1,n k+1 ) d and(n k +1) d ( )n k +1 d= = O(1/ln d k ).k+1 l k m k+1 n k +1The same happens to fr(w k ,w k+1 ,g,F) and fr(w k ,Φ(w k+1 ),g,F), because ∆(g) = 0 for allg ∈ [0,n k+1 −n k −1) d . To simplify citation in the future, we write it down:fr(w k ,w k+1 ,g,F)−fr(w k ,Φ(w k+1 ),g,F) = O(1/l k ). (5.3.7)These estimates imply we can assume, taking l k large enough, that( ) 1−dk 1+d kfr(w k ,w k+1 ,g,F k ) ∈ , , ∀(w k ,w k+1 ,g) ∈ C k ×C k+1 ×Z d .m kd ·|C k | m kd ·|C k |We make a calculation to be used in the next proposition. Fix b A ∈ Ω ∗ (C) and F ∈ F.The main (and simple) observation is: if b A occurs in W k ∈ C k centered at g and w k occursin Φ(w k+1 ) ∈ ˜C k+1 centered at h ∈ [0,l k m k+1 −1) d , then b A occurs in W k+1 ∈ C k+1 centeredat g + n k h. This implies that, if ¯F is a complete residue set modulo F, the cardinality <strong>of</strong>S(b A ,W k+1 | [0,nk+1 −n k −1) d,g,F) is equal to∑ ∑|S(b A ,W k ,g −n k h,F)|+Th∈ ¯Fw k ∈C kw k occurring inw k+1 | [0,lk m k+1 −1) dat a vertex ≡ F h= ∑ h∈ ¯Fw k ∈C k∣ ∣∣S(wk,w k+1 | [0,lk m k+1 −1) d,h,F) ∣ ∣∣·|S(bA ,W k ,g −n k h,F)|+T,where T denotes the number <strong>of</strong> occurrences <strong>of</strong> b A in W k+1 | [0,nk+1 −n k −1) d not entirely containedin a concatenated element <strong>of</strong> C k . Observe that 40 ≤ T ≤ d·l k m k+1 ·n d−1 ·n k+1 < dn d−1 · nk+1 2n k,4 For each line parallel to a coordinate axis e iZ between two elements <strong>of</strong> C k in W k+1 or containing a line <strong>of</strong>ones, there is a rectangle <strong>of</strong> dimensions n×···×n×n k+1 ×n×···×n in which b A is not entirely contained ina concatenated element <strong>of</strong> C k .


Chapter 5. Z d -actions with prescribed topological and ergodic properties 73where n is the length <strong>of</strong> b A . Dividing(5.3.6), (5.3.7), we getfr(b A ,W k+1 ,g,F) =∣∣S(b A ,W k+1 | [0,nk+1 −n k −1) d,g,F) ∣ ∣∣ by nk+1 d and using( )nk+1 −1 d ∑fr(w k ,w k+1 ,h,F)fr(b A ,W k ,g −n k h,F)n k+1h∈ ¯Fw k ∈C k+O(1/l k−1 ). (5.3.8)We wish to show that fr(b A ,x,F) does not depend on x ∈ X. For this, defineα k (b A ,F) = min{fr(b A ,W k ,g,F); W k ∈ C k ,g ∈ Z d}{β k (b A ,F) = max fr(b A ,W k ,g,F); W k ∈ C k ,g ∈ Z d} .The required property is a direct consequence 5 <strong>of</strong> the next result.Proposition 5.3.1. If b A ∈ Ω ∗ (C) and F ∈ F, thenlim α k(b A ,F) = lim β k(b A ,F).k→∞ k→∞Pro<strong>of</strong>. By (5.3.3), if l is large such that F l < F, then(F : F l )·α k (b A ,F l ) ≤ α k (b A ,F) ≤ β k (b A ,F) ≤ (F : F l )·β k (b A ,F l ).This means that we can assume F = F l . We estimate α k+1 (b A ,F) and β k+1 (b A ,F) in terms <strong>of</strong>α k (b A ,F) and β k (b A ,F), for k ≥ l. As b A and F are fixed, denote the above quantities by α kand β k . Take W k+1 ∈ C k+1 . By (5.3.8),( )nk+1 −1 d ∑fr(b A ,W k+1 ,g,F) ≥ ·α k · fr(w k ,w k+1 ,h,F)+O(1/l k−1 )n k+1h∈ ¯Fw k ∈C k( )nk+1 −1 d= ·α k +O(1/l k−1 )n k+1and, as W k+1 and g are arbitrary, we get( )nk+1 −1 dα k+1 ≥ ·α k +O(1/l k−1 ). (5.3.9)n k+1Equality (5.3.8) also implies the upper bound( )nk+1 −1 d ∑fr(b A ,W k+1 ,g,F) ≤ ·β k ·n k+1==⇒ β k+1 ≤fr(w k ,w k+1 ,h,F)+O(1/l k−1 )h∈ ¯Fw k ∈C k( )nk+1 −1 d·β k +O(1/l k−1 )n k+1( )nk+1 −1 d·β k +O(1/l k−1 ). (5.3.10)n k+15 In fact, just take the limit in the inequality α k (b A,F) ≤ fr(b A,x| Ak ,0,F) ≤ β k (b A,F).


Chapter 5. Z d -actions with prescribed topological and ergodic properties 74Inequalities (5.3.9) and (5.3.10) show that α k+1 and β k+1 do not differ very much from α kand β k . The same happens to their difference. Consider w 1 ,w 2 ∈ C k+1 and g 1 ,g 2 ∈ Z d .Renamingg−n k hbyhin(5.3.8) andconsideringn −k theinverse<strong>of</strong> n k modulom l , thedifferencefr(b A ,W 1 ,g 1 ,F)−fr(b A ,W 2 ,g 2 ,F) is at most∑From (5.3.5),fr(b A ,W k ,h,F)|fr(w k ,w 1 ,n −k (g 1 −h),F)−fr(w k ,w 2 ,n −k (g 2 −h),F)|h∈ ¯Fw k ∈C k+O(1/l k−1 ).fr(b A ,W 1 ,g 1 ,F)−fr(b A ,W 2 ,g 2 ,F) ≤2d km ld ·|C k |+O(1/l k−1 )∑≤ 2d k +O(1/l k−1 ),fr(b A ,W k ,h,F)h∈ ¯Fw k ∈C kimplying that0 ≤ β k+1 −α k+1 ≤ 2d k +O(1/l k−1 ). (5.3.11)In particular, β k − α k converges to <strong>zero</strong> as k → +∞. The proposition will be proved if β kconverges. Let us estimate |β k+1 −β k |. On one side, (5.3.10) givesOn the other, by (5.3.9) and (5.3.11),β k+1 −β k ≤ O(1/l k−1 ). (5.3.12)β k+1 −β k ≥ α k+1 −β k( )nk+1 −1 d≥ ·α k −β k +O(1/l k−1 )n k+1( )nk+1 −1 d≥ ·[β k −2d k−1 −O(1/l k−2 )]−β k +O(1/l k−1 )n k+1which, together with (5.3.12), implies that|β k+1 −β k | ≤ 2d k−1 +β k ·[1−= 2d k−1 +O(1/l k−2 ).( ) ]nk+1 −1 d+O(1/l k−2 )n k+1As ∑ d k and ∑ 1/l k both converge, (β k ) k≥1 is a Cauchy sequence, which concludes the pro<strong>of</strong>.


Chapter 5. Z d -actions with prescribed topological and ergodic properties 75From now on, we consider (X,T) as the dynamical system constructed as above. Note thatwe have total freedom to choose C k with few or many elements. This is what controls theentropy <strong>of</strong> the system.5.3.4 Pro<strong>of</strong> <strong>of</strong> Theorem 5.1.1By (5.2.2), the topological entropy <strong>of</strong> the Z d -action (X,T) satisfieslog|C k |h(X,T) ≥ lim ·k→∞ n d kConsider a sequence (ν k ) k≥1 <strong>of</strong> positive real numbers. In the construction <strong>of</strong> C k+1 from C k , takel k large enough such that(vii) ν k ·n k+1 ≥ 1.(viii) |˜C k+1 | ≥ |C k | (l km k+1 ) d·(1−ν k ) .These inequalities implylog|C k+1 |≥ log|˜C k+1 |n d k+1 n d k+1≥ (l km k+1 ) d ·(1−ν k )·log|C k |n k+1d≥ (1−ν k ) d+1 · log|C k|n kdand thenlog|C k |n kdk−1∏≥ (1−ν i ) d+1 · log|C k−11| ∏= (1−νnd i ) d+1 ·log2.1i=1i=1If ν ∈ (0,1) is given and (ν k ) k≥1 are chosen also satisfyinglimk→∞i=1k∏(1−ν i ) d+1 = 1−ν,we obtain that h(X,T) ≥ (1−ν)log2 > 0. If, instead <strong>of</strong> {0,1}, we take C with more elementsand apply the construction verifying (i) to (viii), the topological entropy <strong>of</strong> the Z d -action is atleast (1−ν)log|C|. We have thus proved Theorem 5.1.1.


Chapter 5. Z d -actions with prescribed topological and ergodic properties 765.4 Pro<strong>of</strong> <strong>of</strong> Theorems 5.1.2 and 5.1.3Given a finite alphabet C, consider a configuration b A1 : A 1 → C and any A 2 ⊂ A 1 such that|A 2 | ≤ ε|A 1 |. If b A2 : A 2 → C, the element w ∈ Ω A1 (C) defined byw(g) = b A1 (g), if g ∈ A 1 \A 2 ,= b A2 (g), if g ∈ A 2has frequencies not too different from b A1 , depending on how small ε is. In fact, for any c ∈ C,|S(c,b A1 ,g,F)|−|A 2 | ≤ |S(c,w,g,F)| ≤ |S(c,b A1 ,g,F)|+|A 2 |and then |fr(c,b A1 ,g,F)−fr(c,w,g,F)| ≤ ε.Definition 5.4.1. The upper-Banach <strong>density</strong> <strong>of</strong> a set P ⊂ Z d is equal tod ∗ |P ∩[r 1 ,r 1 +n 1 )×···×[r d ,r d +n d )|(P) = limsup·n 1 ,...,n d →∞ n 1···n dConsider a set P ⊂ Z d <strong>of</strong> <strong>zero</strong> upper-Banach <strong>density</strong>. We will make l k grow quickly suchthat any pattern <strong>of</strong> P∩(A k +g) appears as a subconfiguration in an element <strong>of</strong> C k . Let’s explainthis better. Consider the d-dimensional cubes (A k ) k≥1 that define (X,T). For each k ≥ 1, letà k ⊂ A k be the region containing concatenated elements <strong>of</strong> C k−1 . Inductively, they are definedas Ã1 = {0} and⋃)à k+1 =(Ãk +n k g +∆(n k g) , ∀k ≥ 1,g∈[0,l k m k+1 ) dwhere ∆ is the function defined in (5.3.2).Lemma 5.4.2. If P ⊂ Z d has <strong>zero</strong> upper-Banach <strong>density</strong>, there exists a totally strictly ergodicZ d -action (X,T) with the following property: for any k ≥ 1, g ∈ Z d and b : P∩(A k +g) → {0,1},there exists w k ∈ C k such thatw k (h−g) = b(h), ∀h ∈ P ∩(A k +g).Pro<strong>of</strong>. We proceed by induction on k. The case k = 1 is obvious, since C 1∼ = {0,1}. Supposethe result is true for some k ≥ 1 and consider b : P ∩(A k+1 +g 0 ) → {0,1}. By definition, any0,1 configuration on A k+1 \Ãk+1 is admissible, so that we only have to worry about positionsbelonging to Ãk+1. For each g ∈ [0,l k m k+1 ) d , let)b g : P ∩(Ãk +n k g +∆(n k g)+g 0 → {0,1}


Chapter 5. Z d -actions with prescribed topological and ergodic properties 77)be the restriction <strong>of</strong> b to P∩(Ãk +n k g +∆(n k g)+g 0 . If ε > 0 is given and l k is large enough,=⇒∣∣P ∩(Ãk+1 +g 0)∣ ∣∣∣∣Ãk+1 +g 0∣ ∣∣ 0 is sufficiently small, ˜z ∈ C k+1 . By its own definition, ˜z satisfies the required conditions.The above lemma is the main property <strong>of</strong> our construction. It proves the following strongerstatement.Corollary 5.4.3. Let (X,T) be the Z d -action obtained by the previous lemma. For any b : P →{0,1}, there is x ∈ X such that x| P = b. Also, given x ∈ X, A ∈ R and b : P → {0,1}, thereare ˜x ∈ X and n ∈ N such that ˜x| A = x| A and ˜x(g) = b(g) for all g ∈ P\(−n,n) d .Pro<strong>of</strong>. The first statement is a direct consequence <strong>of</strong> Lemma 5.4.2 and a diagonal argument.For the second, remember that x is the concatenation <strong>of</strong> elements <strong>of</strong> C k and lines <strong>of</strong> <strong>zero</strong>es andones, for every k ≥ 1. Consider k ≥ 1 sufficiently large and z k ∈ C k such that x| A occurs inz k . For any z ∈ C k+1 , there is g ∈ Z d such that z| Ak +g = z k . Constructing ˜z from z makingall substitutions described in Lemma 5.4.2, except in the pattern z| Ak +g, we still have that˜z ∈ C k+1 .5.4.1 Pro<strong>of</strong> <strong>of</strong> Theorem 5.1.2Consider f : X → R given by f(x) = x(0). Then1 ∑ ( )|P ∩(−n,n) d f (T g x) = fr 1,x||P∩(−n,n) d .g∈P∩(−n,n) d


Chapter 5. Z d -actions with prescribed topological and ergodic properties 78For each n ≥ 1, consider the <strong>sets</strong>Λ n = ⋃ { ( ) }x ∈ X; fr 1,x| P∩(−k,k) d < 1/nk≥nΛ n = ⋃ { ( ) }x ∈ X; fr 1,x| P∩(−k,k) d > 1−1/n .k≥n{}Fixed k and n, the <strong>sets</strong> x ∈ X; fr(1,x| P∩(−k,k) d) < 1/n and is clearly open, so that the samehappens to Λ n . It is also dense in X, as we will now prove. Fix x ∈ X and ε > 0. Letk 0 ∈ N be large enough so that d(x,y) < ε whenever x| (−k0 ,k 0 ) d = y| (−k 0 ,k 0 ) d. Take y ∈ Xsuch that y| (−k0 ,k 0 ) d = x| (−k 0 ,k 0 ) d and y(g) = 0 for all g ∈ P\(−n,n)d as in Corollary 5.4.3. Asfr(1,y| (−k,k) d) approaches to <strong>zero</strong> as k approaches to infinity, y ∈ Λ n , proving that Λ n is densein X. The same argument show that Λ n is a dense open set. ThenX 0 = ⋂ n≥1(Λ n ∩Λ n )is a countable intersection <strong>of</strong> dense open <strong>sets</strong>, thus residual. For each x ∈ X 0 ,liminfn→∞limsupn→∞1|P ∩(−n,n) d |1|P ∩(−n,n) d |which concludes the pro<strong>of</strong> <strong>of</strong> Theorem 5.1.2.∑f (T g x) = 0g∈P∩(−n,n) d∑f (T g x) = 1,g∈P∩(−n,n) d5.4.2 Pro<strong>of</strong> <strong>of</strong> Theorem 5.1.3Choose an infinite set G = {g i } i≥1 in Z d disjoint from P such that P ′ = G∪P ∪{0} also has<strong>zero</strong> upper-Banach <strong>density</strong> and let (X,T) be the Z d -action given by Lemma 5.4.2 with respectto P ′ , that is: for every b : P ′ → {0,1}, there exists x b ∈ X such that x b | P ′ = b. Consider{}X 0 = x b ∈ X; b(0) = 0 and b(g) = 1, ∀g ∈ P .This is an uncountable set (it has the same cardinality <strong>of</strong> 2 G = 2 N ) and, for every x b ∈ X 0 andg ∈ P, the elements T g x b and x b differ at 0 ∈ Z d , implying that x b ∉ {T g x b ; g ∈ P}. Thisconcludes the pro<strong>of</strong>.


Bibliography[1] J. Aaronson, An overview <strong>of</strong> infinite ergodic theory, Descriptive set theory anddynamicalsystems, LMS Lecture Notes 277 (2000), 1–29.[2] T. Bedford and A. Fisher, Analogues <strong>of</strong> the Lebesgue <strong>density</strong> theorem for fractal <strong>sets</strong><strong>of</strong> reals and integers, Proceedings <strong>of</strong> the London Mathematical Society (3) 64 (1992), no.1, 95–124.[3] V. Bergelson, Ergodic Ramsey Theory - an Update, ErgodicTheory<strong>of</strong>Z d -actions, LondonMath. Soc. Lecture Note Series 228 (1996), 1–61.[4] V. Bergelson and A. Leibman, Polynomial extensions <strong>of</strong> van der Waerden’s and Szemerédi’stheorems, Journal <strong>of</strong> AMS 9 (1996), no. 3, 725–753.[5] A.S. Besicovitch, On existence <strong>of</strong> sub<strong>sets</strong> <strong>of</strong> finite measure <strong>of</strong> <strong>sets</strong> <strong>of</strong> infinite measure,Indagationes Mathematicae 14 (1952), 339–344.[6] J. Bourgain, Pointwise ergodic theorems for arithmetic <strong>sets</strong>, Publ.Math. IHES69(1989),5–45.[7] Z. Buczolich and D. Mauldin, Divergent square averagesx, Annals <strong>of</strong> Mathematics 171(2010), no. 3, 1479–1530.[8] K.L. Chung and P. Erdös, Probability limit theorems assuming only the first momentI, Mem. Amer. Math. Soc. 6 (1951).[9] T. Cusick and M. Flahive, The Markov and Lagrange spectra, Mathematical Surveysand Monographs 30, AMS.[10] R.O. Davies, Sub<strong>sets</strong> <strong>of</strong> finite measure in analytic <strong>sets</strong>, Indagationes Mathematicae 14(1952), 488–489.[11] K. Falconer, The geometry <strong>of</strong> fractal <strong>sets</strong>, Cambridge Tracts in Mathematics, Cambridge(1986).79


Chapter 5. Z d -actions with prescribed topological and ergodic properties 80[12] A. Fisher, Integer Cantor <strong>sets</strong> and an order-two ergodic theorem, Ergodic Theory andDynamical Systems 13 (1993), no. 1, 45–64.[13] H. Furstenberg, Intersections <strong>of</strong> Cantor <strong>sets</strong> and transversality <strong>of</strong> semigroups, Problemsin analysis (Sympos. Salomon Bochner, Princeton University, 1969), 41–59.[14] H. Furstenberg, Poincaré recurrence and number theory, Bull. Amer. Math. Soc. 5(1981), no. 3, 211–234.[15] H. Furstenberg, Strict ergodicity and transformations <strong>of</strong> the torus, Amer. J. <strong>of</strong> Math.83 (1961), 573–601.[16] B. Green and T. Tao, The primes contain arbitrarily long arithmetic progressions,Annals <strong>of</strong> Mathematics 167 (2008), 481–547.[17] F. Hahn and Y. Katznelson, On the entropy <strong>of</strong> uniquely ergodic transformations, Trans.Amer. Math. Soc. 126 (1967), 335–360.[18] M. Hall, On the sum and product <strong>of</strong> continued fractions, Annals <strong>of</strong> Mathematics 48(1947), 966–993.[19] M. Hochman and P. Shmerkin, Local entropy averages and projections <strong>of</strong> fractalmeasures, to appear in Annals <strong>of</strong> Mathematics.[20] A. Katok, Lyapunov exponents, entropy and periodic orbits for diffeomorphisms, Inst.Hautes Études Sci. Publ. Math. 51 (1980), 137–173.[21] R. Kaufman, On Hausdorff dimension <strong>of</strong> projections, Mathematika 15 (1968), 153–155.[22] B. Kitchens, Symbolic <strong>Dynamics</strong>: One-Sided, Two-Sided and Countable State MarkovShifts, Springer (1997).[23] Y. Lima, Z d -actions with prescribed topological and ergodic properties, to appear inErgodic Theory & Dynamical Systems.[24] Y. Lima and C.G. Moreira, A Marstrand theorem for sub<strong>sets</strong> <strong>of</strong> integers, available athttp://arxiv.org/abs/1011.0672.[25] Y. Lima and C.G. Moreira, A combinatorial pro<strong>of</strong> <strong>of</strong> Marstrand’s theorem for products<strong>of</strong> regular Cantor <strong>sets</strong>, to appear in Expositiones Mathematicae.[26] Y. Lima and C.G. Moreira, Yet another pro<strong>of</strong> <strong>of</strong> Marstrand’s theorem, to appear inBulletin <strong>of</strong> the Brazilian Mathematical Society.


Chapter 5. Z d -actions with prescribed topological and ergodic properties 81[27] J.M. Marstrand, Some fundamental geometrical properties <strong>of</strong> plane <strong>sets</strong> <strong>of</strong> fractionaldimensions, Proceedings <strong>of</strong> the London Mathematical Society 3 (1954), vol. 4, 257–302.[28] J.M. Marstrand, The dimension <strong>of</strong> cartesian product <strong>sets</strong>, Proceedings <strong>of</strong> the CambridgePhilosophical Society 50 (1954), 198–202.[29] P. Mattila, Hausdorff dimension, projections, and the Fourier transform, Publ. Mat. 48(2004), no. 1, 3–48.[30] P. Mattila and P. Saaranen, Ahlfors-David regular <strong>sets</strong> and bilipschitz maps, Ann.Acad. Sci. Fenn. Math. 34 (2009), no. 2, 487–502.[31] C.G. Moreira, A dimension formula for images <strong>of</strong> cartesian products <strong>of</strong> regular Cantor<strong>sets</strong> by differentiable real maps, in preparation.[32] C.G. Moreira, Geometric properties <strong>of</strong> Markov and Lagrange spectra, available athttp://w3.impa.br/∼gugu.[33] C.G. Moreira and J.C. Yoccoz, Stable Intersections <strong>of</strong> Cantor Sets with Large HausdorffDimension, Annals <strong>of</strong> Mathematics 154 (2001), 45–96.[34] C.G. Moreira and J.C. Yoccoz, Tangences homoclines stables pour des ensembleshyperboliques de grande dimension fractale, Annales scientifiques de l’ENS 43, fascicule 1(2010).[35] J. Naudts, Dimension <strong>of</strong> discrete fractal spaces, Journal <strong>of</strong> Physics A: Math. Gen. 21(1988), no. 2, 447–452.[36] J. Palis, Homoclinic orbits, hyperbolic dynamics and dimension <strong>of</strong> Cantor <strong>sets</strong>, ContemporaryMathematics 58 (1987), 203–216.[37] J. Palis and F. Takens, Hyperbolicity and sensitive chaotic dynamics at homoclinicbifurcations, Cambridge studies in advanced mathematics, Cambridge (1993).[38] J. Palis and F. Takens, Hyperbolicity and the creation <strong>of</strong> homoclinic orbits, Annals <strong>of</strong>Mathematics 125 (1987), no. 2, 337–374.[39] J. Palis and J.C. Yoccoz, On the Arithmetic Sum <strong>of</strong> Regular Cantor Sets, Annales del’Inst. Henri Poincaré, Analyse Non Lineaire 14 (1997), 439–456.[40] R. Pavlov, Some counterexamples in topological dynamics, Ergodic Theory & DynamicalSystems 28 (2008), 1291–1322.


Chapter 5. Z d -actions with prescribed topological and ergodic properties 82[41] Y. Peres and P. Shmerkin, Resonance between Cantor <strong>sets</strong>, Ergodic Theory & DynamicalSystems 29 (2009), 201–221.[42] K. Petersen, Ergodic Theory, Cambridge University Press (1983).[43] M. Rams, Exceptional parameters for iterated function systems with overlaps, Period.Math. Hungar. 37 (1998), no. 1-3, 111–119.[44] M. Rams, Packing dimension estimation for exceptional parameters, Israel J. Math. 130(2002), 125–144.[45] E. Szemerédi, On <strong>sets</strong> <strong>of</strong> integers containing no k elements in arithmetic progression,Acta Arith. 27 (1975), 199–245.[46] B. Weiss, Strictly ergodic models for dynamical systems, Bull. Amer. Math. Soc. 13(1985), no. 2, 143–146.[47] N. Wiener, The ergodic theorem, Duke Math. J. 5 (1939), no. 1, 1–18.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!