A FREE BOUNDARY PROBLEM, PROJECTIONS OF RANDOM CANTOR SETS, AND THE GEOMETRY OF CURVES WITH SMALL INTERSECTION WITH MANY LINES By Dimitrios Vardakis A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Mathematicsβ€”Doctor of Philosophy 2023 ABSTRACT Finding the geometric properties of a set is a very old problem. The present text consists of three chapters where we study such properties with techniques involving Complex and Harmonic Analy- sis, Probability, and Geometric Measure Theory. We specifically deal with a few considerations of free boundary problems, we calculate the decay rate of the projections of a certain random Cantor set, and we describe the shape of planar graphs which avoid having too many intersections with a positive cone of lines. To begin with, we introduce Schwarz functions; holomorphic functions on open domains Ξ© satisfying 𝑆(𝜁) = 𝜁 on Ξ“, part of Ω’s boundary. Sakai in 1991 gave a complete characterization of the boundary of a domain admitting a Schwarz function. In fact, if Ξ© is simply connected and Ξ“ = πœ•Ξ© ∩ 𝐷 (𝜁, π‘Ÿ), then Ξ“ has to be regular real analytic. Here, we attempt to describe Ξ“ when the boundary condition is slightly relaxed. In particular, three different scenarios over a simply connected domain Ξ© are treated: when 𝑓1 (𝜁) = 𝜁 𝑓2 (𝜁) on Ξ“ with 𝑓1 , 𝑓2 holomorphic and continuous up to the boundary, when U/V equals certain real analytic function on Ξ“ with U, V positive and harmonic on Ξ© and vanishing on Ξ“, and when 𝑆(𝜁) = Ξ¦(𝜁, 𝜁) on Ξ“ with Ξ¦ a holomorphic function of two variables. It turns out that the boundary piece Ξ“ can be, respectively, anything from real analytic to merely 𝐢 1 , regular except finitely many points, or regular except for a measure zero set. For the second chapter, we consider a model of randomness for self-similar Cantor sets of finite and positive 1-Hausdorff measure. We find the sharp rate of decay of the probability that a Buffon needle lands 𝛿-close to a Cantor set of this particular randomness. Two quite different models of randomness for Cantor sets, by Peres and Solomyak, and by Shiwen Zhang, appear to have the same 𝑐 order of decay for the Buffon needle probability: log 1 . Here, we prove the same rate of decay for 𝛿 a third model of randomness, which asserts a vague feeling that any β€œreasonable” random Cantor 𝑐 set of positive and finite length will have Favard length of order log 1 for its 𝛿-neighbourhood. The 𝛿 estimate from below was obtained long ago by Mattila. In the last chapter, we show the local Lipschitz property for a graph avoiding multiple-point intersection with lines directed in a given cone. The assumption is much stronger than those of the well-known Marstrand’s theorem, but the conclusion is much stronger too. Additionally, we find that a continuous curve with a similar property is 𝜎-finite with respect to Hausdorff length, and we give an estimate on the Hausdorff measure of each β€œpiece”. ACKNOWLEDGEMENTS Before anything else, I need to thank my parents for if it weren’t for their support, nothing from what I have achieved in my life would have been possible. I also want to thank the professors that have taught me and shaped me as a mathematician. I want to express my deepest gratitude and respect for my teacher Manolis G. Maragakis who taught me that a person needs to have ethos before anything else. Also, I am grateful to my professors Nikos Frantzikinakis and Themis Mitsis, who taught me how to think and showed me what Mathematics are. Last, I am thankful to my advisor Alexander Volberg who introduced me to the great vastness of the mathematical world of today. I want to thank all the friends I made here, who gave life a true flavour amidst the numbness of East Lansing and of this foreign country. I am greatly grateful to my friend Stratos Tsoukanis who bore with me throughout all of our conversations and our endless video calls. And I am especially glad I have met here EstefanΓ­a who has given me a fresh view of the world. Of course, I want to thank my committee who allowed me to continue with my plans for the future, as well as JosΓ© Conde Alonso, Irina Holmes, Ilya Kachkovskiy, Ignacio Uriarte-Tuero, and Alexander Volberg who wrote my reference letters and helped me make these plans a reality. Finally, I am grateful to the staff of the Michigan State University that have helped me all these years to guide through the different tasks and obstacles those being administrative or related to teaching. I am especially happy to have met Tsveta Sendova who really tried to take care of us all graduate students and struggled alongside us in our first teaching steps. iv TABLE OF CONTENTS CHAPTER 1 FREE BOUNDARY PROBLEMS VIA SAKAI’S THEOREM . . . . . . 1 1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1.2 Polynomials &analytic functions . . . . . . . . . . . . . . . . . . . . . . . . . 4 1.3 Nevanlinna domains and inner functions . . . . . . . . . . . . . . . . . . . . . 9 1.4 Boundary behaviour of conformal maps in πΎπœƒ . . . . . . . . . . . . . . . . . . 12 1.5 Holomorphic functions in C2 . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 1.6 The U-V problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19 1.7 Some open β€œfree boundary” problems in the spirit of Sakai . . . . . . . . . . . 26 BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28 CHAPTER 2 THE BUFFON’S NEEDLE PROBLEM FOR RANDOM PLANAR DISK-LIKE CANTOR SETS . . . . . . . . . . . . . . . . . . . . . . . 30 2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 2.2 Cantor Disks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32 2.3 Favard Length . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36 2.4 Statement and use of the main lemma . . . . . . . . . . . . . . . . . . . . . . 37 2.5 Proving the main lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38 2.6 Comparison with the other random models . . . . . . . . . . . . . . . . . . . . 44 BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46 CHAPTER 3 GEOMETRY OF PLANAR CURVES INTERSECTING MANY LINES IN A FEW POINTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48 3.1 The statement of the problem . . . . . . . . . . . . . . . . . . . . . . . . . . . 48 3.2 An example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54 3.3 Hausdorff measure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57 3.4 Higher dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66 3.5 Relationships with perturbation theory . . . . . . . . . . . . . . . . . . . . . . 67 BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70 v CHAPTER 1 FREE BOUNDARY PROBLEMS VIA SAKAI’S THEOREM 1.1 Introduction Let 𝐷 (𝜁0 , π‘Ÿ) βŠ‚ C denote the open disk centred at 𝜁0 ∈ C and of radius π‘Ÿ > 0. Let Ξ© be an open subset of 𝐷 (𝜁0 , π‘Ÿ) where 𝜁0 ∈ Ξ“ = πœ•Ξ© ∩ 𝐷 (𝜁0 , π‘Ÿ) is a non-isolated boundary point. A Schwarz function of Ξ© βˆͺ Ξ“ is a function 𝑆 : Ξ© βˆͺ Ξ“ β†’ C holomorphic on Ξ© and continuous on Ξ© βˆͺ Ξ“ that satisfies 𝑆(𝜁) = 𝜁 on Ξ“. (1.1.1) In his Acta Mathematica paper [13], Sakai proved that Schwarz functions completely charac- terize the shape of Ξ“. One of the technical tools used was the PhragmΓ©n–LindelΓΆf principle in the form below, but it is far from being the key to his proof; his paper is full of very subtle tricks. Theorem 1.1.1. Let Ξ© be an open set in C and let 𝜁0 be a non-isolated boundary point of Ξ©. Let 𝑓 be a holomorphic function on Ξ© and 𝐷 (𝜁0 , 𝛿) a ball satisfying the following: (i) lim sup | 𝑓 (𝑧)| ≀ 1 while Ξ© βˆ‹ 𝑧 β†’ 𝜁 for every 𝜁 ∈ πœ•Ξ© ∩ 𝐷 (𝜁0 , 𝛿) \ {𝜁0 } and (ii) | 𝑓 (𝑧)| ≀ 𝛼|𝑧 βˆ’ 𝜁0 | βˆ’π›½ in Ξ© ∩ 𝐷 (𝜁0 , 𝛿) for some positive constants 𝛼 and 𝛽. Then, lim sup | 𝑓 (𝑧)| ≀ 1 while Ξ© βˆ‹ 𝑧 β†’ 𝜁0 . In particular, Sakai proved the following, see [13, Theorem 5.2]. Theorem 1.1.2. Let Ξ© βŠ‚ 𝐷 (𝜁0 , π‘Ÿ) be a bounded open set in C and 𝜁0 an non-isolated point of its boundary, Ξ“ = πœ•Ξ© ∩ 𝐷 (𝜁0 , π‘Ÿ). Suppose 𝑆 is a Schwarz function on Ξ© βˆͺ Ξ“, that is, (i) 𝑆 is holomorphic on Ξ©, (ii) continuous on Ξ© βˆͺ Ξ“, and 1 (iii) 𝑆(𝜁) = 𝜁 on Ξ“. Then, for some small 0 < 𝛿 ≀ π‘Ÿ one of the following must occur (where we set 𝐷 = 𝐷 (𝜁0 , 𝛿)): (1) Ξ© ∩ 𝐷 is simply connected and Ξ“ ∩ 𝐷 is a regular real analytic simple arc through 𝜁0 ; (2a) Ξ“ ∩ 𝐷 determines uniquely a regular real analytic arc through 𝜁0 ; Ξ“ ∩ 𝐷 is either an infinite proper subset of this arc with 𝜁0 as an accumulation point or equal to it; also, Ξ© ∩ 𝐷 = 𝐷 \ Ξ“; (2b) Ξ© ∩ 𝐷 = Ξ©1 βˆͺ Ξ©2 where Ξ©1 and Ξ©2 are (open) simply connected and πœ•Ξ©1 ∩ 𝐷 and πœ•Ξ©2 ∩ 𝐷 are regular real analytic simple arcs through 𝜁0 and tangent at 𝜁0 ; (2c) Ξ© ∩ 𝐷 is simply connected and Ξ“ ∩ 𝐷 is a regular real analytic simple arc except for a cusp at 𝜁0 ; the cusp points into Ξ©. Recall that a regular arc means a differentiable arc whose derivative never vanishes and simple means that it is parametrized by an injective continuous function. Remarks 1.1.3. Here is an example of a cusp of (2c) at 𝜁0 = 0 with Schwarz function. There exist analytic functions 𝑇 on {|𝑧| ≀ πœ‚}, for some πœ‚ > 0, that have a zero of order 2 at 0, are  univalent on closed upper half-disk πΎπœ‚ ≑ |𝑧| ≀ πœ‚ : Im(𝑧) β‰₯ 0 , and satisfy Ξ“ ∩ 𝐷 βŠ‚ 𝑇 (βˆ’πœ‚, πœ‚) and 𝑇 (πΎπœ‚ ) βŠ‚ Ξ© βˆͺ Ξ“. In fact, it is easy to construct such functions. Every such 𝑇 leads to a Schwarz function on the domain Ξ© = 𝑇 ({|𝑧| < πœ‚, Im 𝑧 > 0}), which has two analytic arcs forming a cusp Ξ“ at 0. In order to have 𝑆(𝜁) = 𝜁¯ on Ξ“, it suffices to have a function analytic in {|𝑧| < πœ‚, Im 𝑧 > 0} and continuous up to (βˆ’πœ‚, πœ‚) such that 𝐴(π‘₯) = 𝑇 (π‘₯), π‘₯ ∈ (βˆ’πœ‚, πœ‚). Having such an 𝐴 we set 𝑆 = 𝐴 β—¦ 𝑇 βˆ’1 on Ξ©. On the other hand, using that 𝑇 is analytic in the whole ball {|𝑧| ≀ πœ‚}, we can choose 𝐴 as follows: 𝐴(𝑧) = 𝑇 ( 𝑧¯). Moreover, Sakai [13] showed that every Schwarz function on a cusp domain appears because of an analytic function 𝑇 as above. The converse of this theorem also holds, in the sense that if any of the conditions (1), (2a), (2b), or (2c) is satisfied, then Ξ© admits a Schwarz function. In order to distinguish between the cases, Sakai also showed an auxiliary result [13, Proposition 5.1], which we will also use here. 2 Theorem 1.1.4. Set 𝐷 β€² = 𝐷 (0, π‘Ÿ). Let Ξ©β€² βŠ‚ 𝐷 β€² be an open set and 0 an accumulation point of its boundary, Ξ“β€² = πœ•Ξ©β€² ∩ 𝐷 β€². Then, for some π‘Ÿ β€² ≀ π‘Ÿ, either (1) there exists a Schwarz function, 𝑆𝑑 , of (Ξ©β€² βˆͺ Ξ“β€²) ∩ 𝐷 (0, π‘Ÿ β€²) at 0 if and only if there exists a function Ξ¦1 defined on (Ξ©β€² βˆͺ Ξ“β€²) ∩ 𝐷 (0, 𝛿) for some 𝛿 > 0 such that (i) Ξ¦1 is holomorphic and univalent in Ξ©β€² ∩ 𝐷 (0, 𝛿), (ii) Ξ¦1 is continuous on (Ξ©β€² βˆͺ Ξ“β€²) ∩ 𝐷 (0, 𝛿), (iii) Ξ¦1 (𝜁) = |𝜁 | 2 on Ξ“β€² ∩ 𝐷 (0, 𝛿) or (2) there exists a Schwarz function, 𝑆𝑑 , of (Ξ©β€² βˆͺ Ξ“β€²) ∩ 𝐷 (0, π‘Ÿ β€²) at 0 if and only if there exists a function Ξ¦2 defined on (Ξ©β€² βˆͺ Ξ“β€²) ∩ 𝐷 (0, 𝛿) for some 𝛿 > 0 such that (i’) Ξ¦2 is holomorphic and univalent in Ξ©β€² ∩ 𝐷 (0, 𝛿), (ii’) Ξ¦22 is continuous on (Ξ©β€² βˆͺ Ξ“β€²) ∩ 𝐷 (0, 𝛿), (iii’) Ξ¦22 (𝜁) = |𝜁 | 2 on Ξ“β€² ∩ 𝐷 (0, 𝛿), (iv’) Ξ¦2 (Ξ©β€² ∩ 𝐷 (0, 𝛿)) βˆͺ (βˆ’πœ–, πœ–) contains a neighbourhood of 0 for πœ– > 0. √︁ In particular, the functions Ξ¦1 , Ξ¦2 are related to 𝑆𝑑 by Ξ¦1 (𝑧) = 𝑧𝑆𝑑 (𝑧) and Ξ¦2 (𝑧) = 𝑧𝑆𝑑 (𝑧). Unfortunately, Theorem 1.1.4 is only valid around 0 in this form. Nevertheless, we can β€œtranslate” the setup of Theorem 1.1.2 by setting Ξ©β€² = Ξ© βˆ’ 𝜁0 , Ξ“β€² = Ξ“ βˆ’ 𝜁0 and 𝑆𝑑 (𝑧) = 𝑆(𝑧 + 𝜁0 ) βˆ’ 𝜁 0 for 𝑧 ∈ Ξ©β€². Then, 𝑆𝑑 is a Schwarz function on Ξ©β€² βˆͺ Ξ“β€² at 0. Cases (1) of the two theorems correspond with one another as do (2a), (2b), and (2c) with (2). Sakai gave two applications of his results: the first one describes the local structure of the boundary of quadrature domains, while the second one deals with a free boundary problem of classical type, namely, what is the boundary of the set of positivity of a smooth non-negative function in the disk such that Δ𝑒 = 1 on the set {𝑒 > 0}. 3 It is natural to wonder how one can derive similar results for other forms of (1.1.1). In this text, we examine three different scenarios for a simply connected domain Ξ©. In Sections 1.2 to 1.4 equation (1.1.1) is replaced by 𝑓1 (𝜁) = 𝜁 𝑓2 (𝜁) for all 𝜁 ∈ πœ•Ξ© (1.1.2) where 𝑓1 , 𝑓2 are holomorphic functions continuous up to the boundary. This is closely related to the model subspaces πΎπœƒ and Nevanlinna domains, which will be important here. It is shown that there are domains so that (1.1.2) holds for which πœ•Ξ© is 𝐢 ∞ but not real analytic. Further, in Section 1.5 we replace the quantity 𝜁 𝑓2 (𝜁) with Ξ¦(𝜁, 𝜁), where Ξ¦ is a holomorphic function of two variables, to find that the boundary is locally composed of real analytic arcs. Finally, in Section 1.6 we consider two positive harmonic functions U and V that are zero on a Jordan arc, Ξ“, of the boundary. If their ratio on Ξ“ is equal to a real analytic function of the form | 𝐴| 2 , where 𝐴 is holomorphic, then Ξ“ is real analytic itself with the possible exception of some cusps. Our interests to the problems considered below also was spurred by an application, which originates from complex dynamics. A certain complex dynamics question naturally brought the second author to another free boundary problem described in Section 1.6. After that it was very natural to ask related questions, where the Sakai setup was generalized in yet two other ways. To our surprise the answers were quite different and required different techniques: from the use of Nevanlinna domains and pseudo-continuation to multivalued analytic functions. 1.2 Polynomials & analytic functions Let Ξ© be an open domain, 𝜁0 a non-isolated boundary point of Ξ©, and let Ξ“ = πœ•Ξ© ∩ 𝐷 (𝜁0 , π‘Ÿ) for some π‘Ÿ > 0. Suppose 𝑆 is a holomorphic function on Ξ© continuous on Ξ© βˆͺ Ξ“. We start with a simple yet important case. Instead of (1.1.1), we consider 𝑆(𝜁) = 𝜁 𝑝(𝜁) on Ξ“, (1.2.1) 𝑆(𝑧) where 𝑝 is a polynomial. We will shortly show that 𝑓 (𝑧) = 𝑝(𝑧) is, in fact, a Schwarz function on Ξ“. 4 Lemma 1.2.1. Assume that 𝑆 : Ξ© β†’ C is holomorphic on Ξ© βŠ‚ 𝐷 (𝜁0 , π‘Ÿ), continuous on Ξ© βˆͺ Ξ“, and that it satisfies 𝑆(𝜁) = 𝜁 (𝜁 βˆ’ 𝜁0 ) 𝑛 on Ξ“. Then, the function 𝑆𝑑 (𝑧) = 𝑆(𝑧 + 𝜁0 ) βˆ’ 𝜁0 𝑧 𝑛 is holomorphic on Ξ© βˆ’ 𝜁0 βŠ‚ 𝐷 (0, π‘Ÿ), continuous on (Ξ© βˆ’ 𝜁0 ) βˆͺ (Ξ“ βˆ’ 𝜁0 ) and it satisfies 𝑆𝑑 (𝜁) = 𝜁 𝜁 𝑛 on Ξ“ βˆ’ 𝜁0 . Proposition 1.2.2. Assume 0 ∈ Ξ“ is a non-isolated boundary point of Ξ© βŠ‚ 𝐷 (0, π‘Ÿ) and suppose 𝑆 is a holomorphic function on Ξ© continuous on Ξ© βˆͺ Ξ“ and satisfying 𝑆(𝜁) = 𝜁 𝜁 𝑛 on Ξ“. 𝑆(𝑧) Then, for any positive 𝛿 < π‘Ÿ the function 𝑧𝑛 is holomorphic on Ξ© ∩ 𝐷 (0, 𝛿) and continuous on (Ξ© βˆͺ Ξ“) ∩ 𝐷 (0, 𝛿) \ {0}. Moreover, the following holds while 𝑧 ∈ Ξ© βˆͺ Ξ“ \ {0}: 𝑆(𝑧) lim = 0. 𝑧→0 𝑧 𝑛 𝑆(𝑧) Proof. The function 𝑧𝑛 is clearly holomorphic on Ξ© ∩ 𝐷 (0, 𝛿) and continuous on (Ξ© βˆͺ Ξ“) ∩ 𝐷 (0, 𝛿) \ {0} for any 𝛿 ∈ (0, π‘Ÿ). It remains to see what happens at 0. Fix 𝛿 ∈ (0, π‘Ÿ). Since 𝑆 is bounded on Ξ© ∩ 𝐷 (0, π‘Ÿ), say by π‘š, we get 𝑆(𝑧) 𝑛 ≀ π‘š|𝑧| βˆ’π‘› on Ξ© ∩ 𝐷 (0, 𝛿) 𝑧 and additionally for any 𝜁 ∈ Ξ“ ∩ 𝐷 (0, 𝛿) \ {0} we have 𝑆(𝑧) lim = |𝜁 | ≀ 𝛿 while Ξ© βˆ‹ 𝑧 β†’ 𝜁 . 𝑧𝑛 Hence, by the PhragmΓ©n-LindelΓΆf principle 1.1.1 we obtain 𝑆(𝑧) lim sup ≀𝛿 while Ξ© βˆ‹ 𝑧 β†’ 0. 𝑧𝑛 This last inequality is true for any positive 𝛿 < π‘Ÿ and therefore lim 𝑆(𝑧) 𝑧 𝑛 = 0 as 𝑧 β†’ 0. β–‘ 5 Corollary 1.2.3. Let 𝑝 be a complex polynomial. Assume that 𝜁0 ∈ Ξ“ is a non-isolated boundary point of Ξ© other than zero and suppose 𝑆 is a holomorphic function of Ξ© βŠ‚ 𝐷 (𝜁0 , π‘Ÿ) continuous on Ξ© βˆͺ Ξ“ and satisfying 𝑆(𝜁) = 𝜁 𝑝(𝜁) on Ξ“. Set 𝑓 (𝑧) = 𝑆(𝑧)/𝑝(𝑧) on Ξ© βˆͺ Ξ“ \ {𝜁0 } and 𝑓 (𝜁0 ) = 𝜁0 . Then, 𝑓 is a Schwarz function of Ξ© βˆͺ Ξ“ on 𝐷 (𝜁0 , π‘Ÿ) for sufficiently small π‘Ÿ > 0. Proof. Take π‘Ÿ so small that 𝑝 has no zeros on 𝐷 (𝜁0 , π‘Ÿ) \ {𝜁0 }. If 𝑝(𝜁0 ) β‰  0, the result is immediate. If 𝑝(𝜁0 ) = 0, we only need to show that 𝑓 is continuous on (Ξ© βˆͺ Ξ“) ∩ 𝐷 (𝜁0 , π‘Ÿ). Denote by 𝑛 the order of 𝜁0 as a zero of 𝑝 and consider the function (𝑧 βˆ’ 𝜁0 ) 𝑛 𝑆𝑛 (𝑧) = 𝑆(𝑧) . 𝑝(𝑧) 𝑆𝑛 is holomorphic on Ξ©, is continuous on Ξ© βˆͺ Ξ“, and satisfies 𝑆𝑛 (𝜁) = 𝜁 (𝜁 βˆ’ 𝜁0 ) 𝑛 on Ξ“. From Lemma 1.2.1 we get (𝑆𝑛 )𝑑 (𝜁) = 𝜁 𝜁 𝑛 on Ξ“ βˆ’ 𝜁0 and from Proposition 1.2.2 we deduce that while (𝑆𝑛 )𝑑 (𝑧) 𝑆𝑛 (𝑧 + 𝜁0 ) βˆ’ 𝜁0 𝑧 𝑛 lim = 0 =β‡’ lim =0 𝑧→0 𝑧𝑛 𝑧→0 𝑧𝑛 𝑆𝑛 (𝑧) βˆ’ 𝜁0 (𝑧 βˆ’ 𝜁0 ) 𝑛 =β‡’ lim =0 π‘§β†’πœ0 (𝑧 βˆ’ 𝜁0 ) 𝑛 𝑆𝑛 (𝑧) =β‡’ lim 𝑓 (𝑧) = lim = 𝜁0 π‘§β†’πœ0 π‘§β†’πœ0 (𝑧 βˆ’ 𝜁0 ) 𝑛 𝑧 ∈ Ξ©, and the conclusion follows. β–‘ Notice that the same proof works with 𝑝 replaced by any function 𝐹 that is analytic in a neighbourhood of 𝜁0 . This along with Lemma 1.2.1 give us the following corollary. 6 Corollary 1.2.4. Assume 𝜁0 ∈ Ξ“ is a non-isolated boundary point of Ξ© βŠ‚ 𝐷 (𝜁0 , π‘Ÿ). Suppose 𝐹 is a function analytic around 𝜁0 and 𝑆 is a holomorphic function on Ξ© continuous on Ξ© βˆͺ Ξ“ and satisfying 𝑆(𝜁) = 𝜁 𝐹 (𝜁) on Ξ“. Set 𝑓 (𝑧) = 𝑆(𝑧)/𝐹 (𝑧) on (Ξ©βˆͺ Ξ“) ∩ 𝐷 (𝜁0 , 𝛿) \ {𝜁0 } for some sufficiently small 𝛿 > 0 and 𝑓 (𝜁0 ) = 𝜁0 . Then, 𝑓 is a Schwarz function of Ξ© βˆͺ Ξ“ on 𝐷 (𝜁0 , 𝛿). The converse of this corollary also holds true in the sense that if Ξ“ has certain shape, in particular, if it satisfies (1), (2a), (2b), or (2c) of 1.1.2, then there is a Schwarz function 𝑓 of Ξ© βˆͺ Ξ“ at 𝜁0 such that 𝑆(𝜁) = 𝜁 𝐹 (𝜁) on Ξ“ where 𝑆 = 𝐹 𝑓 . In fact, we can slightly modify the same proof to get a little more, again through the PhragmΓ©n- LindelΓΆf principle 1.1.1. Corollary 1.2.5. Let 𝑝 be a polynomial, 𝐹 a function analytic in a neighbourhood of Ξ©Μ„, and 𝑆 a function holomorphic on the (bounded) set Ξ© and continuous on Ξ© βˆͺ Ξ“. Suppose that for all 𝜁 ∈ Ξ“ we have 𝑆(𝜁) = 𝑝(𝜁)𝐹 (𝜁). Then, for every non-isolated point 𝜁0 of the boundary Ξ“ for which 𝑝′ (𝜁0 ) β‰  0, there is some 𝛿 > 0 such that the function 𝑝 βˆ’1 (𝑆/𝐹) is a Schwarz function of Ξ© βˆͺ Ξ“ on 𝐷 (𝜁0 , 𝛿). We wish to examine what happens in the more general case where 𝑝 in (1.2.1) is replaced with any analytic function of Ξ© continuous on its boundary, but not necessarily analytic on that boundary. More specifically, suppose that 𝑓1 and 𝑓2 are functions analytic on Ξ©, continuous on Ξ© βˆͺ Ξ“, and satisfying 𝑓1 (𝜁) = 𝜁 𝑓2 (𝜁) on Ξ“. (1.2.2) As above, if 𝑓2 (𝜁0 ) β‰  0, the function 𝑓 = 𝑓1 / 𝑓2 is a Schwarz function around 𝜁0 ∈ Ξ“ and no issues arise. However, if 𝑓2 (𝜁0 ) = 0, the situation is very complicated in general. We start with a lemma analogous to Lemma 1.2.1: 7 Lemma 1.2.6. Assume that 𝑓1 , 𝑓2 : Ξ© β†’ C are holomorphic on Ξ© βŠ‚ 𝐷 (𝜁0 , π‘Ÿ), continuous on Ξ© βˆͺ Ξ“, and that they satisfy 𝑓1 (𝜁) = 𝜁 𝑓2 (𝜁) on Ξ“. Then, there exist functions ( 𝑓1 )𝑑 and ( 𝑓2 )𝑑 holomorphic on Ξ© βˆ’ 𝜁0 , continuous on (Ξ© βˆ’ 𝜁0 ) βˆͺ (Ξ“ βˆ’ 𝜁0 ) and such that ( 𝑓1 )𝑑 (𝜁) = 𝜁 ( 𝑓2 )𝑑 (𝜁) on Ξ“ βˆ’ 𝜁0 . If additionally 𝑓2 (𝜁0 ) = 0, then ( 𝑓2 )𝑑 (0) = 0. Proof. Define ( 𝑓1 )𝑑 by ( 𝑓1 )𝑑 (𝑧) = 𝑓1 (𝑧 + 𝜁0 ) βˆ’ 𝜁0 𝑓2 (𝑧 + 𝜁0 ). Then for 𝜁 ∈ Ξ“ βˆ’ 𝜁0 we have ( 𝑓1 )𝑑 (𝜁) = 𝑓1 (𝜁 + 𝜁0 ) βˆ’ 𝜁0 𝑓2 (𝜁 + 𝜁0 ) = 𝜁 + 𝜁0 𝑓2 (𝜁 + 𝜁0 ) βˆ’ 𝜁0 𝑓2 (𝜁 + 𝜁0 ) = 𝜁 𝑓2 (𝜁 + 𝜁0 ) Setting ( 𝑓2 )𝑑 (𝑧) = 𝑓2 (𝑧 + 𝜁0 ), we have the desired identity. Clearly, ( 𝑓1 )𝑑 (0) = 0 and also if 𝑓2 (𝜁0 ) = 0, ( 𝑓2 )𝑑 (0) = 0. β–‘ Abusing the notation, we denote these new functions again by 𝑓1 and 𝑓2 . It remains to show a result analogous to Corollary 1.2.3 with 𝑝 replaced by 𝑓2 . In particular, we would like to show that the function 𝑓 = 𝑓1 / 𝑓2 is holomorphic on Ξ©, continuous on Ξ“, and that it satisfies 𝑓1 (𝜁) 𝑓 (𝜁) = =𝜁 for all 𝜁 ∈ Ξ“. 𝑓2 (𝜁) However, the limit of 𝑓1 (𝑧)/ 𝑓2 (𝑧) as Ξ© βˆ‹ 𝑧 β†’ 0 may even fail to exist when 𝑓2 (0) = 0, and we cannot apply the PhragmΓ©n-LindelΓΆf principle here. We will need to see this problem from a different scope. 8 1.3 Nevanlinna domains and inner functions We recall that a bounded simply connected domain Ξ© is called a Nevanlinna domain if there exist bounded holomorphic functions 𝑓1 , 𝑓2 in Ξ© such that 𝑓1 (πœ‘(𝑧)) πœ‘(𝑧) = 𝑓2 (πœ‘(𝑧)) for almost every 𝑧 ∈ T = {𝑧 : |𝑧| = 1}, where πœ‘ is a conformal mapping of the unit disk onto Ξ©. Note that this definition does not imply any additional regularity (for instance, continuity) of the functions 𝑓1 , 𝑓2 on πœ•Ξ©. We will restrict the above situation, and suppose there are holomorphic functions 𝑓1 , 𝑓2 : Ξ© β†’ C continuous up to the boundary that satisfy 𝑓1 (𝜁) = 𝜁 𝑓2 (𝜁) for 𝜁 ∈ Ξ“. (1.3.1) In order to better understand the situation, we rewrite (1.3.1) as 𝑓1 (𝜁) = 𝜁, (1.3.1’) 𝑓2 (𝜁) which is now fulfilled almost everywhere on Ξ“ except for the closed set Ξ“ ∩ 𝑓2βˆ’1 {0}, which has zero measure. Then, Ξ© is what we call a strong Nevanlinna domain and if such 𝑓1 and 𝑓2 exist, the ratio 𝑓1 / 𝑓2 is unique thanks to the Lusin-Privalov uniqueness theorem. Let πœ™ : D β†’ Ξ© be a conformal map and consider the functions 𝐹1 = 𝑓1 β—¦ πœ™ and 𝐹2 = 𝑓2 β—¦ πœ™. Formulas (1.3.1) and (1.3.1’) transform respectively to 𝐹1 (𝜁) = πœ™(𝜁)𝐹2 (𝜁) (1.3.2) and 𝐹1 (𝜁) = πœ™(𝜁) (1.3.2’) 𝐹2 (𝜁) both of which hold true in the sense of angular boundary values almost everywhere on T, because πœ™ may fail to extend β€œnicely” to DΜ„. By the factorization theorem, we can write 𝐹1 and 𝐹2 in D as 𝐹1 = πœƒ 1 F1 and 𝐹2 = πœƒ 2 F2 (1.3.3) 9 where the F𝑖 are the outer factors of 𝐹𝑖 and the πœƒ 𝑖 are their inner factors. Since 𝐹1 , 𝐹2 ∈ 𝐻 ∞ , also F1 , F2 ∈ 𝐻 ∞ , and from (1.3.2’) we get πœƒ 1 (𝜁) F1 (𝜁) = πœ™(𝜁), (1.3.4) πœƒ 2 (𝜁) F2 (𝜁) almost everywhere on T in the sense of angular boundary values. We distinguish between two cases: either πœƒ 2 divides πœƒ 1 , that is, πœƒ 1 /πœƒ 2 ∈ 𝐻 ∞ , or it does not. 1.3.1 πœƒ2 | πœƒ1. Let β„Ž = πœƒ 1 /πœƒ 2 ∈ 𝐻 ∞ . Then, the function (β„ŽF1 )/F2 belongs to the class 𝑁 + , defined as   + 𝑓 ∞ 𝑁 = : 𝑓 , 𝑔 ∈ 𝐻 , 𝑔 is an outer function , 𝑔 and its (angular) boundary values are equal almost everywhere on T to the (angular) boundary values of πœ™. However, since Ξ© is bounded, we see that πœ™ ∈ 𝐿 ∞ (T, π‘š) where π‘š is the normalized Lebesgue measure on T. Smirnov’s Theorem tells us that in fact (β„ŽF1 )/F2 ∈ 𝐻 ∞ . Therefore, we have a bounded holomorphic function on the disk that is equal to πœ™ almost everywhere on T. This is impossible whenever πœ™ is a bounded holomorphic function on D. We are necessarily left with the other case. 1.3.2 πœƒ2 ∀ πœƒ1. We begin with some notation and definition which will be important for the rest of this text. Let D𝑒 = Cb \ DΜ„. For any function β„Ž : D β†’ C we define e β„Ž as β„Ž(𝑧) = β„Ž(1/𝑧). e The notation 𝐻 e will stand for a function 𝐻 e : D𝑒 β†’ C and we will write 𝐻 instead of 𝐻 e for the e function 𝐻 e(1/𝑧). Observe that β„Ž ∈ 𝐻 ∞ if and only if e β„Ž ∈ 𝐻 ∞ (D𝑒 ), and β„Ž(0) = 0 if and only if β„Ž(∞) = 0. e We will also consider the backward shift operator, B : 𝐻 𝑝 β†’ 𝐻 𝑝 , for 𝑝 ∈ [1, ∞), that is 𝑓 (𝑧) βˆ’ 𝑓 (0) B : 𝑓 ↦→ . 𝑧 10 Definition 1.3.1. Let 𝑓 be a meromorphic function on D. We say that 𝑓 admits pseudo-continuation (across T) if there exists another meromorphic function 𝑔 on D𝑒 such that 𝑓 = 𝑔 almost everywhere (on T) in the sense of non-tangential limits. The pseudo-continuation of 𝑓 is called of bounded type or a Nevanlinna-type pseudo-continua- tion if 𝑔 is of the form 𝑔 = β„Ž1 /β„Ž2 for some β„Ž1 , β„Ž2 ∈ 𝐻 ∞ (D𝑒 ). Definition 1.3.2. A function 𝑓 ∈ 𝐻 𝑝 is called a cyclic vector for B, or simply cyclic for B if the set {B 𝑛 𝑓 }∞ 𝑛=0 spans the space 𝐻 . 𝑝 The following important result is due to Douglas, Shapiro, and Shields. Theorem 1.3.3. Consider 1 ≀ 𝑝 < ∞. A function 𝑓 ∈ 𝐻 𝑝 is not cyclic for B if and only if 𝑓 has a pseudo-continuation of bounded type. In the case when 𝑝 = 2, it is known that any non-cyclic function of B belongs to a proper B- invariant subspace. As a consequence of Beurling’s theorem, these spaces are of the form (πœƒπ» 2 ) βŠ₯ and are known as model spaces and denoted by πΎπœƒ . Here we will need the fact that πΎπœƒ = (πœƒπ» 2 ) βŠ₯ = 𝐻 2 (T) ∩ πœƒπ»02 (T), where in the last identity we mean the boundary values of the corresponding functions and where 𝐻02 = { 𝑓 ∈ 𝐻 2 : 𝑓 (0) = 0}. Now, we can proceed with the case when πœƒ 2 ∀ πœƒ 1 : After dividing both πœƒ 1 and πœƒ 2 by their greatest common divisor, we may assume that πœƒ 1 and πœƒ 2 have no common zeros and that the Borel supports of their singular measures are disjoint. Much as above, we see that the function 𝐹 = (πœƒ 1 F1 )/F2 = 𝐹1 /F2 belongs the class 𝑁 + and thus 𝐹 ∈ 𝐻 ∞ , because πœƒ 2 πœ™ ∈ 𝐿 ∞ (T, π‘š). Then the following is true in the sense of angular boundary values for 11 almost every 𝜁 ∈ T: πœƒ 1 (𝜁)F1 (𝜁) 𝐹 (𝜁) πœ™(𝜁) = = πœƒ 2 (𝜁)F2 (𝜁) πœƒ 2 (𝜁) ⇐⇒ πœ™(𝜁) = πœƒ 2 (𝜁)𝐹 (𝜁) (1.3.5) e(𝜁) 𝐹 ⇐⇒ πœ™(𝜁) = . (1.3.6) πœƒ 2 (𝜁) e Since 𝐹, ee πœƒ 2 ∈ 𝐻 ∞ (D𝑒 ), we see that πœ™ ∈ 𝐻 ∞ βŠ‚ 𝐻 2 admits pseudo-continuation across T of bounded type, and Theorem 1.3.3 shows that πœ™ is not cyclic for B. So, it has to belong to some model space πΎπœƒ . See [7, Theorem 1] for more details. In fact, from (1.3.5) and because we β€œneed” to have 𝐹 (0) = 0, it follows that either πœ™ ∈ πΎπœƒ 2 if πœƒ 1 (0) = 0, or πœ™ ∈ 𝐾 π‘§πœƒ 2 if πœƒ 1 (0) β‰  0. 1.4 Boundary behaviour of conformal maps in πΎπœƒ In this section we show that Theorem 1.1.2 fails when condition (iii) is replaced by (1.3.1). To this end, we will find a simply connected domain Ξ© and a conformal map πœ™ : D β†’ Ξ© continuous up the boundary that has a pseudo-continuation of bounded type and is smooth but not real analytic on T. The functions participating in this pseudo-continuation will also be continuous on the boundary. First, we go one step back and work with Nevanlinna domains. Thanks to [7, Theorem 1] by Fedorovskiy, this is equivalent to studying the model subspaces, πΎπœƒ , for different inner functions πœƒ. If πœƒ (𝑧0 ) = 0 for some 𝑧0 ∈ D, the function 1 πœ™(𝑧) = ∈ πΎπœƒ ∩ 𝐢 ∞ (T) 1 βˆ’ 𝑧0 𝑧 has bounded type pseudo-continuation across T and thus πœ™(D) is a Nevanlinna domain. In fact, πœ™ can be analytically extended on the whole closed disk, DΜ„, and πœ™(T) is real analytic. On the other hand, in a series of papers, [10, 4, 7, 11, 12, 3], it has been shown that the boundary of a Nevanlinna domain can be β€œarbitrarily bad”. In particular, it can be nowhere analytic [10], of class 𝐢 1 but not in any 𝐢 1,𝛼 for no 𝛼 > 0 [7], or even non-rectifiable [11]. We refer also to the Belov-Fedorovskiy paper [2], where the description is given of model spaces that contain bounded univalent functions. 12 We mention that the Hausdorff dimension of the accessible boundary of a Nevanlinna domain can be any number between 1 and 2 as shown in [3], another construction can be found in [12]. However, in all the above work the inner function πœƒ is a Blaschke product or has a Blaschke part. Moreover, in order to compare with Sakai’s theorem, we have to consider the case where the functions 𝐹 e2 ∈ 𝐻 ∞ (D𝑒 ) for which πœ™ = 𝐹 e1 , 𝐹 e1 /𝐹 e2 on T are continuous up to T. This is not always possible when πœƒ is not purely singular (see [4, Example 5.8]). Therefore, in this section πœƒ will be a singular inner function of the form  ∫  𝜁 +𝑧 πœƒ (𝑧) = exp βˆ’ π‘‘πœ‡πœƒ (𝜁) 𝜁 βˆ’π‘§ T with πœ‡πœƒ supported on a Carleson set, 𝐸 βŠ‚ T. We will show that there is a conformal map πœ™ ∈ πΎπœƒ continuous on DΜ„ which is in 𝐢 ∞ (T) but not real analytic on T. In view of [6, Theorem 2.1], since supp(πœ‡πœƒ ) is Carleson, the space πΎπœƒ then contains a non-trivial function from some smoothness class, for example a function 𝑔 ∈ 𝐻 ∞ ∩ 𝐢 ∞ (T) (or in a Bergman space, i.e., 𝑔 ∈ 𝐴 𝑝,1 for some 𝑝 > 1). Since 𝑔 ∈ πΎπœƒ , it admits a bounded type pseudo-continuation of the form 𝑔 = 𝐺/ e eπœƒ almost everywhere on T, where 𝐺 e ∈ 𝐻 ∞ (D𝑒 ) vanishes at infinity (see [5, Theorem 5.1.4]). Additionally, 𝑔 has an analytic continuation, say G, to C b \ supp(πœ‡). Of course, G = 𝐺/ e eπœƒ on D𝑒 and observe that G cannot be bounded in D𝑒 ; otherwise 𝑔 would be constant, as G|D = 𝑔 and G|D𝑒 coincide almost everywhere on T. Now, consider 𝛼 ∈ D𝑒 with πœƒ (1/𝛼) β‰  0 and the following aggregate: G(𝑧) βˆ’ G(𝛼) πœ™(𝑧) = . π‘§βˆ’π›Ό We will show that πœ™ ∈ πΎπœƒ ∩ 𝐢 ∞ (T) and πœ™ is conformal in DΜ„. Clearly, πœ™ is inside 𝐻 2 (D) and also πœƒ (𝜁)𝑔(𝜁) βˆ’ πœƒ (𝜁)G(𝛼) 𝐺 e(𝜁) βˆ’ e πœƒ (𝜁)G(𝛼) πœƒ (𝜁)πœ™(𝜁) = = . 𝜁 βˆ’π›Ό 𝜁 βˆ’π›Ό 13 For 𝑧 ∈ D𝑒 the function   e(𝑧) βˆ’ e 𝐺 πœƒ (𝑧)G(𝛼) 1 πœƒ e (𝑧) = e(𝑧) βˆ’ 𝐺 e(𝛼) 𝐺 π‘§βˆ’π›Ό π‘§βˆ’π›Ό πœƒ (𝛼) e is analytic around 𝛼 and vanishes at infinity. Hence, πœ™ ∈ πΎπœƒ . Furthermore, πœ™ is univalent in DΜ„. Indeed, suppose it is not. Then, there exist 𝑧, 𝑀 ∈ DΜ„ with 𝑧 β‰  𝑀 and πœ™(𝑧) = πœ™(𝑀) or equivalently 𝑔(𝑧) βˆ’ G(𝛼) 𝑔(𝑀) βˆ’ G(𝛼) = π‘§βˆ’π›Ό π‘€βˆ’π›Ό 𝑔(𝑧) 𝑔(𝑀) G(𝛼) G(𝛼) π‘§βˆ’π‘€ ⇐⇒ βˆ’ = βˆ’ = G(𝛼) π‘§βˆ’π›Ό π‘€βˆ’π›Ό π‘§βˆ’π›Ό π‘€βˆ’π›Ό (𝑧 βˆ’ 𝛼)(𝑀 βˆ’ 𝛼) 𝑔(𝑧) βˆ’ 𝑔(𝑀) 𝑔(𝑧) βˆ’ 𝑔(𝑀) ⇐⇒ βˆ’ 𝛼 +𝑀 βˆ’ 𝑔(𝑀) = G(𝛼). π‘§βˆ’π‘€ π‘§βˆ’π‘€ The left-hand side is bounded, because 𝑔 ∈ 𝐢 ∞ ( DΜ„), whereas we can pick 1 < 𝛼 < 2 so that |G(𝛼)| is arbitrarily large (recall G|D𝑒 is not bounded), a contradiction, and therefore πœ™ is univalent in DΜ„. Consequently, if 𝑔 ∈ πΎπœƒ ∩ 𝐢 ∞ (T) and πœƒ is a singular inner function, then πœ™ is univalent in DΜ„ and πœ™ ∈ πΎπœƒ ∩ 𝐢 ∞ (T). Also see [1, Section 4] for more details. At the same time, note that G cannot be analytically extended to the whole DΜ„, because it is unbounded near the unit circle, and thus neither can πœ™; this fails exactly on the Carleson set 𝐸. Now, since πœ™ ∈ πΎπœƒ , we can write πœ™ = πœƒπΉ ⇐⇒ πœƒπœ™ = 𝐹 (1.4.1) almost everywhere on T for some function 𝐹 ∈ 𝐻 2 with 𝐹 (0) = 0. In fact, 𝐹 ∈ 𝐻 ∞ because πœ™ ∈ 𝐢 ∞ ( DΜ„). It is known that there exists some analytic function, H , with H|𝐸 = 0 such that both H and H πœƒ are Lipschitz on DΜ„. In fact, we can further consider H to be an outer function in 𝐢 ∞ (T). Multiplying by H in (1.4.1), we get (H πœƒ)πœ™ = H 𝐹 (1.4.2) almost everywhere on T. In particular, the left-hand side is now smooth on the whole T and the same therefore holds true for the right-hand side. In a sense, H β€œannihilates” the singularities of πœƒ as (1.4.1) fails exactly on the support, 𝐸, of πœ‡πœƒ . 14 At this point, set 𝐹1 = H 𝐹, 𝐹2 = H πœƒ, and 𝑓 𝑗 = 𝐹 𝑗 β—¦ πœ™βˆ’1 for 𝑗 = 1, 2. Then, (1.4.2) becomes 𝐹1 = πœ™πΉ2 , which now is fulfilled on the whole boundary T, and in turn 𝑓1 (𝜁) = 𝜁 𝑓2 (𝜁) for all 𝜁 ∈ Ξ“. This is exactly the setup we were looking for, albeit it contrasts with Sakai’s result: Even though Ξ“ = πœ™(T) is 𝐢 ∞ -smooth, πœ™ cannot be analytic on the Carleson set 𝐸 and thus neither can Ξ“. It is worth mentioning that there are examples of Nevanlinna domains that come from singular inner functions with particularly irregular boundaries. Namely, in [3] one can find examples of univalent functions in a Paley-Wiener space such that they map the upper half-plane onto a Nevanlinna domain whose boundary can have any dimension between 1 and 2. 1.5 Holomorphic functions in C2 In this section we attempt to replace the function 𝜁 𝑓0 (𝜁) in (1.3.1) with a more general formula. For some positive π‘Ÿ > 0, let Ξ© βŠ‚ 𝐷 (𝜁0 , π‘Ÿ) be a simply connected open set, let Ξ“ = πœ•Ξ©βˆ©π· (𝜁0 , π‘Ÿ), and let 𝜁0 ∈ Ξ“. Here, we will also need the extra assumption that Ξ“ is a Jordan arc (or possibly a union of Jordan arcs). Let Ξ¦ be a holomorphic function of two variables, that is, a function of the form +∞ βˆ‘οΈ Ξ¦(𝑧, 𝑀) = 𝑏 π‘›π‘š 𝑧 𝑛 𝑀 π‘š 𝑛,π‘š=0 where each of the functions Ξ¦(𝑧, Β· ) and Ξ¦( Β· , 𝑀) is itself holomorphic. Suppose there exists a function 𝑅 which is (i) holomorphic on Ξ©, (ii) continuous on Ξ©Μ„, and (iii) satisfies 𝑅(𝜁) = Ξ¦(𝜁, 𝜁) on Ξ“. 15 In view of Lemma 1.2.6, we may assume that 𝜁0 = 0 and 𝑏 00 = 0 so that 𝑅(0) = Ξ¦(0, 0) = 0. Notice that 𝑅(𝑧) and Ξ¦(𝑧, 𝑧) are bounded on Ξ©Μ„ and thanks to the PhragmΓ©n-LindelΓΆf Principle 1.1.1, we may assume without loss of generality that there exists some non-negative integer π‘˜ for which πœ• πœ• π‘˜βˆ’1 πœ•π‘˜ Ξ¦(0, 0) = Ξ¦(0, 0) = Β· Β· Β· = Ξ¦(0, 0) = 0 and Ξ¦(0, 0) β‰  0 (1.5.1) πœ•π‘€ πœ•π‘€ π‘˜βˆ’1 πœ•π‘€ π‘˜ otherwise Ξ¦ would be identically zero. We would like to use the Weierstraß approximation theorem for the function Ξ¦(𝑧, 𝑀) βˆ’ 𝑅(𝑧) around 0, but 𝑅 is not holomorphic on the boundary. But since it is continuous by (ii) and Ξ“ is Jordan, we can use Mergelyan’s theorem to get a sequence of polynomials 𝑝 𝑛 that converge to 𝑅 uniformly on Ξ©Μ„. And we can pick this sequence so that 𝑝 𝑛 (0) = 0 for every 𝑛 = 0, 1, . . . . Next, we define the functions Ξ¨(𝑧, 𝑀) = Ξ¦(𝑧, 𝑀) βˆ’ 𝑅(𝑧) and Ψ𝑛 (𝑧, 𝑀) = Ξ¦(𝑧, 𝑀) βˆ’ 𝑝 𝑛 (𝑧). The Ψ𝑛 are holomorphic on C2 and converge uniformly to Ξ¨ on Ξ©Μ„ Γ— C. Observe that for all 𝑛 we have Ψ𝑛 (0, 0) = Ξ¦(0, 0) βˆ’ 𝑝 𝑛 (0) = 0 and also πœ•πœ… πœ•πœ… Ψ𝑛 = Ξ¦ for all integers πœ… β‰₯ 1 πœ•π‘€ πœ… πœ•π‘€ πœ… and all points (𝑧, 𝑀). Then, from (1.5.1) and from the Weierstraß approximation theorem, there exist unique holomorphic functions π‘Ž 0;𝑛 , . . . , π‘Ž π‘˜βˆ’1;𝑛 : C β†’ C and 𝑐 𝑛 : C2 β†’ C with π‘Ž 𝑗;𝑛 (0) = 0 and 𝑐 𝑛 (0, 0) β‰  0 such that   Ψ𝑛 (𝑧, 𝑀) = 𝑐 𝑛 (𝑧, 𝑀) 𝑀 π‘˜ + π‘Ž π‘˜βˆ’1;𝑛 (𝑧)𝑀 π‘˜βˆ’1 + Β· Β· Β· + π‘Ž 0;𝑛 (𝑧) . Following the proof of the Weierstraß theorem and since the convergence Ψ𝑛 β†’ Ξ¨ is uniform on Ξ©Μ„ Γ— C, we can find sufficiently small 𝛿 and 𝜌 with 𝜌 β‰₯ 𝛿 > 0 so that π‘Ž 0;𝑛 , . . . , π‘Ž π‘˜βˆ’1;𝑛 and the 𝑐 𝑛  converge uniformly on Ξ©Μ„ ∩ 𝐷 (0, 𝛿) and Ξ©Μ„ ∩ 𝐷 (0, 𝛿) Γ— 𝐷 (0, 𝜌), respectively, to some functions π‘Ž 0 , . . . , π‘Ž π‘˜βˆ’1 and 𝑐 with π‘Ž 𝑗 (0) = 0 and 𝑐(0, 0) β‰  0. Note that the functions π‘Ž 𝑗 are holomorphic on Ξ© ∩ 𝐷 (0, 𝛿) and continuous on Ξ©Μ„ ∩ 𝐷 (0, 𝛿). Subsequently, we get   Ξ¦(𝑧, 𝑀) βˆ’ 𝑅(𝑧) = 𝑐(𝑧, 𝑀) 𝑀 π‘˜ + π‘Ž π‘˜βˆ’1 (𝑧)𝑀 π‘˜βˆ’1 + Β· Β· Β· + π‘Ž 0 (𝑧) . (1.5.2) 16 Let us write 𝑃(𝑧, 𝑀) = 𝑀 π‘˜ + π‘Ž π‘˜βˆ’1 (𝑧)𝑀 π‘˜βˆ’1 + Β· Β· Β· + π‘Ž 0 (𝑧) for the polynomial factor. From (iii), (1.5.2) and since 𝑐(0, 0) β‰  0, we have π‘˜ π‘˜βˆ’1 𝑃(𝜁, 𝜁) = 𝜁 + π‘Ž π‘˜βˆ’1 (𝜁)𝜁 + Β· Β· Β· + π‘Ž 0 (𝜁) = 0 for all 𝜁 ∈ Ξ“ ∩ 𝐷 (0, 𝛿). (1.5.3) Remark. Functions of the form 𝑃(𝑧, 𝑧) = 𝑧 π‘˜ + π‘Ž π‘˜βˆ’1 (𝑧)𝑧 π‘˜βˆ’1 + Β· Β· Β· + π‘Ž 0 (𝑧), where π‘Ž 𝑗 are polynomials, are called polyanalytic polynomials. One can find more details on these in [8, 11] or [14]. We are interested in the roots of the polynomial 𝑃(𝑧, Β· ) when 𝑧 ∈ Ξ©Μ„ ∩ 𝐷 (0, 𝛿). In other words, we will study the equation (in 𝑀) 𝑃(𝑧, 𝑀) = 0 ⇐⇒ 𝑀 π‘˜ + π‘Ž π‘˜βˆ’1 (𝑧)𝑀 π‘˜βˆ’1 + Β· Β· Β· + π‘Ž 0 (𝑧) = 0 (1.5.4) when 𝑧 ∈ Ξ©Μ„ ∩ 𝐷 (0, 𝛿). Let D (𝑧) be the discriminant of 𝑃(𝑧, Β· ) (for any fixed 𝑧). Then, D (𝑧) is a polynomial of πœ• the coefficients π‘Ž 0 (𝑧), . . . , π‘Ž π‘˜βˆ’1 (𝑧) and is equal to 0 if, and only if, 𝑃(𝑧, 𝑀) and πœ•π‘€ 𝑃(𝑧, 𝑀) share a common factor. The roots of 𝑃(𝑧, Β· ) are given by a multivalued holomorphic function, W, depending on π‘Ž 1 , . . . , π‘Ž π‘˜βˆ’1 , and the points where W changes a branch inside Ξ© ∩ 𝐷 (0, 𝛿) are exactly the zeros of D (in Ξ© ∩ 𝐷 (0, 𝛿)). We distinguish between two cases: when D is identically 0 and when it is not. Before moving on, let us note that the set M (Ξ©, Ξ“, 𝛿) of all meromorphic functions on Ξ© ∩ 𝐷 (0, 𝛿) continuous up to (Ξ© βˆͺ Ξ“) ∩ 𝐷 (0, 𝛿) except possibly a (closed) measure zero subset of Ξ“ is a field with the usual operations of addition and multiplication. 1.5.1 Dβ‰ 0 Here 𝑃(𝑧, 𝑀) is irreducible over M (Ξ©, Ξ“, 𝛿). Since D is continuous on (Ξ© βˆͺ Ξ“) ∩ 𝐷 (0, 𝛿), the set D βˆ’1 {0} ∩ Ξ“ ∩ 𝐷 (0, 𝛿) is closed and of zero harmonic measure. Now, we decompose  Ξ“ \ D βˆ’1 {0} ∩ 𝐷 (0, 𝛿) into countably many open connected arcs.  17 Let 𝛾 be one of these arcs. Then, there exists a simply connected set 𝐷 βŠ‚ Ξ© ∩ 𝐷 (0, 𝛿) such that πœ•π· ∩ πœ•Ξ© = 𝛾. Since D has no zeros on 𝐷 βˆͺ 𝛾, by the monodromy theorem the multivalued function W β€œsplits” into π‘˜ distinct holomorphic functions, π‘Š 𝑗 ( 𝑗 = 1, . . . , π‘˜), and let 𝐢 𝑗 = {𝜁 ∈ 𝛾 : π‘Š 𝑗 (𝜁) = 𝜁 }. Notice that the 𝐢 𝑗 ’s are closed (in 𝛾), they cover 𝛾, and any two of them intersect at a (closed) set of zero harmonic measure. Unfortunately, 𝐢 𝑗 need not be connected, but we can further decompose each 𝐢˚ 𝑗 (whenever it is non-empty) into countably many open arcs as in 𝐢˚ 𝑗 = βˆͺ𝑖 𝛾 𝑖𝑗 , for 𝑗 = 1, . . . , π‘˜. Again, around each 𝛾 𝑖𝑗 we consider a neighbourhood 𝐷 𝑖𝑗 βŠ‚ 𝐷 with πœ•π· 𝑖𝑗 ∩ πœ•π· = 𝛾 𝑖𝑗 (these can, but need not be simply connected) and let π‘Š 𝑖𝑗 = π‘Š 𝑗 |𝐷 𝑖 βˆͺ𝛾 𝑖 . 𝑗 𝑗 Then, for each 𝑗 = 1, . . . , π‘˜ and 𝑖 = 1, 2 . . . the functions π‘Š 𝑖𝑗 are holomorphic on 𝐷 𝑖𝑗 , continuous on 𝐷 𝑖𝑗 βˆͺ 𝛾 𝑖𝑗 and satisfy π‘Š 𝑖𝑗 (𝜁) = 𝜁 for all 𝜁 ∈ 𝛾 𝑖𝑗 ; in other words, they are Schwarz functions on 𝐷 𝑖𝑗 βˆͺ 𝛾 𝑖𝑗 . Since Ξ“ is Jordan, all 𝛾 𝑖𝑗 are also Jordan and from Theorem 1.1.2 we conclude that each 𝛾 𝑖𝑗 is, in fact, a regular real analytic simple arc except possibly some cusps. 1.5.2 D=0 In this case, 𝑃(𝑧, 𝑀) has to be reducible over M (Ξ©, Ξ“, 𝛿). In particular, we can write 𝑃(𝑧, 𝑀) = 𝑃1 (𝑧, 𝑀) Β· Β· Β· 𝑃eπ‘˜ (𝑧, 𝑀) for some e π‘˜ ≀ π‘˜ where each π‘ƒπœ… (𝑧, 𝑀) has now coefficients in M (Ξ©, Ξ“, 𝛿) and is irreducible, i.e., Dπœ… β‰  0 where Dπœ… is the discriminant of π‘ƒπœ… (𝑧, Β· ). Since 𝑃(𝜁, 𝜁) = 0 for all 𝜁 ∈ Ξ“, we can split (Ξ“ \ 𝐸) ∩ 𝐷 (0, 𝛿), where 𝐸 is some closed zero-(harmonic)-measure set, into open sets 𝑂 πœ… for πœ… = 1, . . . , eπ‘˜ so that π‘ƒπœ… (𝜁, 𝜁) = 0 for all 𝜁 ∈ 𝑂 πœ… . Notice that 𝑂 πœ… ∩ 𝑂 πœ… β€² = βˆ… when π‘ƒπœ… and π‘ƒπœ… β€² are different. Observe that, since 𝑃(𝑧, 𝑀) factors into the polynomials π‘ƒπœ… (𝑧, 𝑀) (over M (Ξ©, Ξ“, 𝛿)) and the roots of 𝑃(𝑧, Β· ) are given by the multivalued holomorphic function W, the roots of each π‘ƒπœ… (𝑧, Β· ) are also given by a multivalued holomorphic function Wπœ… whose branches are comprised of branches of W. Working as above for each πœ… = 1, . . . , e π‘˜, we separate 𝑂 πœ… \ Dπœ…βˆ’1 {0} into countably many open arcs and for each such arc, 𝛾, we find some simply connected neighbourhood, 𝐷 βŠ‚ Ξ©, with πœ•π· ∩ πœ•Ξ© = 𝛾 so that Wπœ… β€œsplits” into its different branches. Again following the above arguments, 18 we can decompose 𝛾 β€” minus a zero-measure set β€” into countably many open arcs over which π‘Š 𝑗 (𝜁) = 𝜁 for some branch π‘Š 𝑗 of Wπœ… . Constructing appropriate neighbourhoods, we conclude that except a zero-measure set, 𝛾 is a countable union of regular real analytic simple arcs except possibly some cusps. In either case, the cusps (if they exist) point into Ξ© and may only accumulate on the endpoints of each open arc. Now we formulate the above results into a theorem. Theorem 1.5.1. Let Ξ© be a bounded simply connected domain such that Ξ“ = πœ•Ξ© ∩ 𝐷 (𝜁0 , π‘Ÿ) is a (union of) Jordan arc(s). Also, let Ξ¦ be a (non-trivial) holomorphic function of two variables defined in 𝐷 (𝜁0 , π‘Ÿ) Γ— 𝐷 ( 𝜁¯0 , π‘Ÿ), and suppose there exists a function 𝑅 (i) holomorphic on Ξ©, (ii) continuous on Ξ©Μ„, and such that (iii) 𝑅(𝜁) = Ξ¦(𝜁, 𝜁) for all 𝜁 ∈ Ξ“. Then, there exists a closed set, 𝐸 βŠ‚ Ξ“, of zero harmonic measure so that Ξ“ \ 𝐸 is a countable union of regular real analytic simple arcs except possibly for some cusps. The cusps (if they exist) point into Ξ© and may only accumulate on 𝐸. 1.6 The U-V problem In this section, we are interested in the following setup. Let Ξ© be a simply connected open set in C and let 𝜁0 ∈ πœ•Ξ© be a boundary point of Ξ©. Assume that for some 𝜌 > 0 the connected component, Ξ“, of πœ•Ξ© ∩ 𝐷 (𝜁0 , 𝜌) containing 𝜁0 is a Jordan curve. Note that 𝜌 β‰₯ dist(𝜁0 , πœ•Ξ©\ Ξ“) > 0. For convenience we will write simply Ξ© to denote Ω∩ 𝐷 (𝜁0 , 𝜌). Let 𝐴 be an analytic function in a neighbourhood, 𝐷 (𝜁0 , πœ–), of 𝜁0 and suppose we have two functions U and V defined on Ξ© that are not proportional and have the following properties: I) U and V are positive and harmonic on Ξ©, 19 II) they are continuous on Ξ© βˆͺ Ξ“, III) U = V = 0 on Ξ“, and U (𝜁) IV) V (𝜁) = | 𝐴(𝜁)| 2 β‰  const for 𝜁 ∈ Ξ“. Notice that since U β‰  𝑐V, the function | 𝐴| needs to be non-constant. Otherwise, we could have U = 𝑐V and all our conditions work trivially for any Ξ“. Also, we may assume that 𝜌 < πœ– without loss of generality (so that 𝐴 is defined over the whole Ξ©) to avoid unnecessary technical difficulties. Formula (IV) is to be understood in the sense of limits, i.e., the limit of U (𝑧)/V (𝑧) as Ξ© βˆ‹ 𝑧 β†’ 𝜁 ∈ Ξ“ exists and is equal to | 𝐴(𝜁)| 2 . In fact, this limit always exists when Ξ© is simply connected and Ξ“ is Jordan (see Remark 1.6.1), so the only assumption here is the values it takes. Consider a conformal map from the PoincarΓ© plane to Ξ©, πœ™ : H β†’ Ξ©. Since Ξ“ is connected and Jordan, CarathΓ©odory’s theorem implies that πœ™ extends conformally to a function (abusing the notation) πœ™ : H βˆͺ 𝛾 β†’ Ξ© βˆͺ Ξ“ which we can pick so that 𝛾 βŠ‚ R is some bounded open interval with πœ™(𝛾) = Ξ“ and πœ™(0) = 𝜁0 . Utilizing this πœ™, we can β€œtransfer” the information about U and V over Ξ© to information over H. Define 𝑒 ≑ U β—¦ πœ™, 𝑣 ≑Vβ—¦πœ™ and π‘Ž ≑ π΄β—¦πœ™ and note that π‘Ž is analytic on H and continuous on H βˆͺ 𝛾. As above, we have i) 𝑒 and 𝑣 are positive and harmonic on H, ii) they are continuous on H βˆͺ 𝛾, iii) 𝑒 = 𝑣 = 0 on 𝛾, and iv) 𝑒 𝑣 = |π‘Ž| 2 on 𝛾. Again, (iv) is to be understood in the sense of limits. 20 Now, harmonically extend 𝑒 and 𝑣 to H βˆͺ 𝛾 βˆͺ Hβˆ’ by ο£±  ο£±  ∈ π‘§βˆˆH      𝑒(𝑧), 𝑧 H    𝑣(𝑧),     ο£²  ο£²  𝑒 βˆ— (𝑧) = 0, 𝑧 ∈ 𝛾 and 𝑣 βˆ— (𝑧) = 0, π‘§βˆˆπ›Ύ         βˆ’  βˆ’π‘£(𝑧), 𝑧 ∈ Hβˆ’      βˆ’π‘’(𝑧), 𝑧 ∈ H   ο£³ ο£³ and let β„Ž be the function ο£± 𝑒 βˆ— (𝑧) ο£² 𝑣 βˆ— (𝑧) , 𝑧 ∈ H βˆͺ Hβˆ’ ,     β„Ž(𝑧) =  𝑒 βˆ—π‘¦ (𝑧)  𝑣 βˆ—π‘¦ (𝑧) , 𝑧 ∈ 𝛾.   ο£³ We claim that β„Ž is well defined and, in fact, real analytic on Hβˆͺ 𝛾 βˆͺHβˆ’ . Indeed, using Harnack’s inequality, for any (π‘₯, 0) ∈ 𝛾 there exists a constant 𝑐 > 0 (dependent on 𝑣 βˆ— ) such that 𝑦 2βˆ’π‘¦ 𝑐 ≀ 𝑣 βˆ— (π‘₯, 𝑦) ≀ 𝑐 for every 0 < 𝑦 < 1, or 2βˆ’π‘¦ 𝑦 1 βˆ— 𝑣 (π‘₯, 𝑦) 2βˆ’π‘¦ 𝑐 ≀ ≀𝑐 2 . (1.6.1) 2βˆ’π‘¦ 𝑦 𝑦 Recall that 𝑣 βˆ— (π‘₯, 0) = 0 and take limits as 𝑦 β†’ 0+ . Since 𝑣 βˆ— is harmonic on H βˆͺ 𝛾 βˆͺ Hβˆ’ , (1.6.1) guarantees that 𝑣 βˆ—π‘¦ > 0 on 𝛾 (the same holds true for 𝑒 βˆ— ) and therefore the limit βˆ— 𝑒 βˆ— (π‘₯, 𝑦) 𝑒 𝑦 (π‘₯, 0) lim = 𝑦→0 𝑣 βˆ— (π‘₯, 𝑦) 𝑣 βˆ—π‘¦ (π‘₯, 0) exists and is finite. Hence, β„Ž is a well-defined continuous function on H βˆͺ 𝛾 βˆͺ Hβˆ’ . In fact, since 𝑒 βˆ—π‘¦ and 𝑣 βˆ—π‘¦ are real analytic and non-zero around 𝛾, β„Ž is also real analytic on H βˆͺ 𝛾 βˆͺ Hβˆ’ . What is more is that 𝑒 βˆ—π‘¦ (πœ‰) 𝑒(𝑧) β„Ž(πœ‰) = = lim = |π‘Ž(πœ‰)| 2 for any πœ‰ ∈ 𝛾 (1.6.2) 𝑣 βˆ—π‘¦ (πœ‰) Hβˆ‹π‘§β†’πœ‰ 𝑣(𝑧) because of (iv) and therefore |π‘Ž| 2 is also real analytic on 𝛾. U Remark 1.6.1. The above is the reason why relation (IV) is meaningful. When we write V on Ξ“, it really means the limit of β„Ž β—¦ πœ™βˆ’1 as we approach Ξ“ from the inside of Ξ©. This limit always exist on a Jordan arc Ξ“ when Ξ© is simply connected thanks to Harnack’s inequality. It is worth mentioning the work of Jerison and Kenig who showed [9, Theorems 5.1 and 7.9] that equation (IV) makes sense whenever Ξ© is assumed to be a non-tangentially accessible (NTA) domain. 21 Next, consider β„Ž |𝛾 . Its power series around 0 ∈ 𝛾 is given by ∞ βˆ‘οΈ β„Ž |𝛾 (π‘₯) = 𝑏𝑛π‘₯𝑛 𝑛=0 for some real numbers 𝑏 0 , 𝑏 1 , . . . This readily extends to a complex analytic function, say π‘Ÿ, on some open neighbourhood, 𝐷 (0, πœ– β€²): ∞ βˆ‘οΈ π‘Ÿ (𝑧) = 𝑏𝑛 𝑧𝑛 , 𝑛=0 where we can choose πœ™ and πœ– β€² so that 𝛾 βŠ‚ 𝐷 (0, πœ– β€²). Of course, by construction and from (1.6.2) we get π‘Ÿ |𝛾 = β„Ž |𝛾 = |π‘Ž| 2 . At this point, we want to β€œshift” everything back at Ξ©. We set 𝑉 ≑ πœ™(H ∩ 𝐷 (0, πœ– β€²)) βŠ‚ Ξ© and observe that πœ•π‘‰ is a closed Jordan arc such that Ξ“ ⊊ πœ•π‘‰ ∩ 𝐷 (𝜁0 , 𝜌). Define a new function 𝑅 ≑ π‘Ÿ β—¦ (πœ™βˆ’1 ) |𝑉βˆͺΞ“ , (1.6.3) which is holomorphic on 𝑉, continuous on 𝑉 βˆͺ Ξ“, and on Ξ“ it satisfies 𝑅(𝜁) = | 𝐴(𝜁)| 2 . Now, consider the function Ξ¦(𝑧, 𝑀) = 𝐴(𝑧) 𝐴(𝑀). Ξ¦ is holomorphic on 𝐷 (𝜁0 , πœ–) Γ— 𝐷 (𝜁 0 , πœ–) and it satisfies Ξ¦(𝜁, 𝜁) = 𝐴(𝜁) 𝐴(𝜁) = | 𝐴(𝜁)| 2 when 𝑧 = 𝑀 = 𝜁 ∈ Ξ“. As a corollary to Theorem 1.5.1, the next theorem follows. Theorem 1.6.2. Let Ξ© be a bounded simply connected domain in C and let Ξ“ be an open Jordan arc of its boundary with 𝜁0 ∈ Ξ“. Suppose there are two positive non-proportional harmonic functions U and V on Ξ© continuous on Ξ© βˆͺ Ξ“ and such that U (𝜁) U (𝜁) = V (𝜁) = 0 and = | 𝐴(𝜁)| 2 for all 𝜁 ∈ Ξ“, V (𝜁) where 𝐴 is a non-trivial analytic function on a neighbourhood of Ξ©. Then, there exists some neighbourhood 𝐷 of 𝜁0 and a closed set 𝐸 βŠ‚ Ξ“ of zero harmonic measure so that (Ξ“ \ 𝐸) ∩ 𝐷 is a countable union of regular real analytic simple arcs except possibly for some cusps. The cusps (if they exist) point into Ξ© and may only accumulate on 𝐸 ∩ 𝐷. 22 Of course, Theorems 1.5.1 and 1.6.2 are somewhat far from Sakai’s result. Nevertheless, because of the special form of the function Ξ¦(𝑧, 𝑀) = 𝐴(𝑧) 𝐴(𝑀), we can actually say more in this case. Proposition 1.6.3. Let Ξ© be a bounded simply connected domain in C and let Ξ“ be an open Jordan arc of its boundary with 𝜁0 ∈ Ξ“. Suppose there are two positive non-proportional harmonic functions U and V on Ξ© continuous on Ξ© βˆͺ Ξ“ and such that U (𝜁) U (𝜁) = V (𝜁) = 0 and = | 𝐴(𝜁)| 2 for all 𝜁 ∈ Ξ“ V (𝜁) where 𝐴 is a non-trivial analytic function on a neighbourhood of Ξ“. Then, there exists a neighbourhood 𝐷 of 𝜁0 and a function 𝑅 satisfying the following: (i) 𝑅 is holomorphic on Ξ© ∩ 𝐷, (ii) 𝑅 is continuous on (Ξ© βˆͺ Ξ“) ∩ 𝐷 and (iii) 𝑅(𝜁) = | 𝐴(𝜁)| 2 for 𝜁 ∈ Ξ“ ∩ 𝐷. Additionally, for any 𝜁0 ∈ Ξ“ with 𝐴′ (𝜁0 ) β‰  0 either (1) there exist a function Ξ¨1 holomorphic and univalent on Ξ© ∩ 𝐷 such that Ξ¨1 is continuous on (Ξ© βˆͺ Ξ“) ∩ 𝐷, and Ξ¨1 (𝜁) = | 𝐴(𝜁) βˆ’ 𝐴(𝜁0 )| 2 for 𝜁 ∈ Ξ“ ∩ 𝐷, or (2) there exist a function Ξ¨2 holomorphic and univalent on Ξ© ∩ 𝐷 such that Ξ¨22 is continuous on (Ξ© βˆͺ Ξ“) ∩ 𝐷, and Ξ¨22 (𝜁) = | 𝐴(𝜁) βˆ’ 𝐴(𝜁0 )| 2 for 𝜁 ∈ Ξ“ ∩ 𝐷. Proof. We have already established the existence of such a function 𝑅 in (1.6.3). For the rest, 𝐴′ (𝜁0 ) β‰  0 and we may assume without loss of generality that 𝐴 is conformal on a neighbourhood of Ξ©Μ„. Recall that 𝑉 from the definition of 𝑅 in (1.6.3) is such that πœ•π‘‰ is Jordan and Ξ“ ⊊ πœ•π‘‰ ∩ 𝐷 (𝜁0 , 𝜌) when Ξ© βŠ‚ 𝐷 (𝜁0 , 𝜌). Since 𝐴 is continuous and injective on 𝑉, Β― there exists some small 𝛿, 0 < 𝛿 ≀ 𝜌, such that πœ• ( 𝐴(𝑉)) ∩ 𝐷 (𝜁0β€² , 𝛿) βŠ‚ 𝐴(Ξ“). 23 Now, let 𝐴(𝜁0 ) = 𝜁0β€² , Ξ©β€² = 𝐴(𝑉) ∩ 𝐷 (𝜁0β€² , 𝛿), and Ξ“β€² = πœ•Ξ©β€² ∩ 𝐷 (𝜁0β€² , 𝛿). The function 1 𝑆(𝑧) ≑ 𝑅 β—¦ π΄βˆ’1 (𝑧) (1.6.4) 𝑧 is a Schwarz function of Ξ©β€² βˆͺ Ξ“β€² in 𝐷 (𝜁0β€² , 𝛿): (i) 𝑆 is holomorphic on Ξ©β€², (ii) it is continuous on Ξ©β€² βˆͺ Ξ“β€², and (iii) 𝑆(𝜁) = 1𝜁 𝑅( π΄βˆ’1 (𝜁)) = 𝜁 on Ξ“β€². Notice that from (1.6.1) the functions π‘Ž = 𝐴 β—¦ πœ™ and 𝐴 are always non-zero and thus 𝑆 is a well-defined holomorphic function, because 0 cannot be a point of Ξ©Μ„β€². Finally, consider the function 𝑆𝑑 (𝑧) = 𝑆(𝑧 + 𝜁0β€² ) βˆ’ 𝜁0β€² , which is a Schwarz function on (Ξ©β€² βˆ’ 𝜁0β€² ) βˆͺ (Ξ“β€² βˆ’ 𝜁0β€² ) at 0. From Theorem 1.1.4, we know that one of the functions Ξ¦1 (𝑧) = 𝑧𝑆𝑑 (𝑧) and √︁ Ξ¦2 (𝑧) = 𝑧𝑆𝑑 (𝑧) is univalent on (Ξ©β€² βˆ’ 𝜁0β€² ) ∩ 𝐷 (0, 𝛿′) for some 𝛿′ ≀ 𝛿. Changing variables to get back to our initial domain Ξ©, we find that one of the following functions, Ξ¨1 or Ξ¨2 , has to be univalent on Ξ© ∩ 𝐷 β€²:   𝑅(𝑧) Ξ¨1 (𝑧) = ( 𝐴(𝑧) βˆ’ 𝐴(𝜁0 )) βˆ’ 𝐴(𝜁0 ) 𝐴(𝑧) and βˆšοΈ„   𝑅(𝑧) Ξ¨2 (𝑧) = ( 𝐴(𝑧) βˆ’ 𝐴(𝜁0 )) βˆ’ 𝐴(𝜁0 ) 𝐴(𝑧) for 𝑧 ∈ Ξ© ∩ 𝐷 β€², where 𝐷 β€² = π΄βˆ’1 (𝐷 (𝜁0β€² , 𝛿′)). The rest of the desired properties are obvious. β–‘ In the above proof, Ξ“β€² is the image of a Jordan arc under the (conformal) map 𝐴. Therefore, the existence of a Schwarz function, 𝑆, along with Theorem 1.1.2 imply that Ξ“β€², and in turn Ξ“, satisfy (1) or (2c) of Theorem 1.1.2. Case (1) corresponds to (1) of Proposition 1.6.3 and (2c) to (2), that is, Ξ“β€² (respectively, Ξ“) has a cusp if, and only if, the function βˆšοΈƒ 𝑧 𝑆(𝑧 + 𝜁0β€² ) βˆ’ 𝜁0β€²  24 is univalent on (Ξ©β€² βˆ’ 𝜁0β€² ) ∩ 𝐷 (0, 𝛿′) (respectively, Ξ¨2 on Ξ© ∩ 𝐷). As a consequence, we have the following theorem, which is the main result of this section. Theorem 1.6.4. Let Ξ© be a bounded simply connected domain in C and let Ξ“ be an open Jordan arc of its boundary with 𝜁0 ∈ Ξ“. Suppose there are two positive non-proportional harmonic functions U and V on Ξ© continuous on Ξ© βˆͺ Ξ“ and satisfying U (𝜁) U (𝜁) = V (𝜁) = 0 and = | 𝐴(𝜁)| 2 for all 𝜁 ∈ Ξ“, V (𝜁) where 𝐴 is a non-trivial analytic function on a neighbourhood of Ξ“. Then, for all but possibly finitely many points 𝜁0 ∈ Ξ“ there exists some small neighbourhood 𝐷 of 𝜁0 such that the following holds: Ξ“ ∩ 𝐷 is a regular real analytic simple arc through 𝜁0 except possibly a cusp at 𝜁0 . (1.6.5) The finitely many points around which (1.6.5) might fail are the points 𝜁 ∈ Ξ“ where 𝐴′ (𝜁) = 0, i.e., where 𝐴 might not be invertible. There is a cusp at 𝜁0 if and only if (2) of Proposition 1.6.3 holds true. Of course, one can ask at this point whether it is possible to actually have a cusp. The answer is yes as the next example shows. Example 1.6.5. Let Ξ© be open and Ξ“ = πœ•Ξ© ∩ 𝐷 (0, 𝜌) (with 𝜌 ≀ 1 sufficiently small) be such that Ξ“ has a cusp at 0 (i.e., 𝜁0 = 0). Then, from Remarks 1.1.3, for some πœ‚ > 0, there is a holomorphic function 𝑇 defined on {|𝑧| ≀ πœ‚} that maps conformally the closed upper half-disk πΎπœ‚ = {|𝑧| ≀ πœ‚ : Im(𝑧) β‰₯ 0} into Ξ© βˆͺ Ξ“ and Ξ“ ∩ 𝐷 βŠ‚ 𝑇 (βˆ’πœ‚, πœ‚) for some small neighbourhood 𝐷 of 0. Also, 𝑇 (0) = 0 with order 2. By dilating appropriately, we may assume that everything happens in the unit disk, that is, πœ‚ = 1, 𝑇 is defined on DΜ„ and is univalent on 𝐾1 = {|𝑧| ≀ 1 : Im(𝑧) β‰₯ 0}, 𝑇 (𝐾1 ) βŠ‚ Ξ© βˆͺ Ξ“, and Ξ“ ∩ 𝐷 (0, 𝜌) βŠ‚ 𝑇 (βˆ’1, 1). Next, consider two positive harmonic functions, 𝑒 and 𝑣, on the upper half-disk D βˆͺ H that are zero on (βˆ’1, 1). As we saw in the beginning of this section, 𝑒 and 𝑣 can be extended on the whole 25 disk and the ratio 𝑒/𝑣 is a positive analytic function on (βˆ’1, 1). Therefore, on (βˆ’1, 1) we can write 𝑒/𝑣 = |π‘Ž| 2 for some function π‘Ž holomorphic on D. Finally, construct the functions U = 𝑒 β—¦ 𝑇 βˆ’1 , V = 𝑣 β—¦ 𝑇 βˆ’1 , and 𝐴 = π‘Ž β—¦ 𝑇 βˆ’1 . Then, 𝐴 is holomorphic around the cusp at 0, and U, V are positive harmonic functions on Ξ© ∩ 𝐷 (0, 𝜌) and zero on the boundary Ξ“. Moreover, U and V satisfy U/V = | 𝐴| 2 on Ξ“. 1.7 Some open β€œfree boundary” problems in the spirit of Sakai All problems treated above are examples of the so-called free boundary problems (non- variational free boundary problems). We would like to call the attention of the reader to one open question: what can one say for the boundary of a domain Ξ© that is not simply connected but admits positive harmonic functions vanishing on its boundary and whose ratio is β€œnice” on that boundary? Finitely connected situations present no difficulties, but what if, for example, Ξ“ is a Cantor set and Ξ© = D \ Ξ“? Suppose we know that the ratio of two positive harmonic (non-proportional) functions U, V in Ξ© vanishing on the Cantor set Ξ“ has a well-defined ratio on Ξ“ (this happens for a wide class of Cantor sets Γ’s, for example for all regular Cantor sets of positive Hausdorff dimension). Suppose this ratio is equal to | 𝐴(𝜁)| 2 β‰  const for 𝜁 ∈ Ξ“, where 𝐴 is a holomorphic function on D. What we can say about the Cantor set Ξ“? The β€œdesired” answer is that this is impossible to happen on any Cantor set. This type of problems (we may call them β€œone-phase free boundary problems”) appear naturally in certain problems of complex dynamics, see, e.g., [15]. If we would know the aforementioned answer (we conjecture that no Cantor set would allow such a triple (U, V, 𝐴)), then a long-standing problem about the dimension of harmonic measure on Cantor repellers would be solved. Another similar one-phase boundary problem concerns functions in R𝑛 for 𝑛 > 2. Let Ξ© be a bounded domain in R𝑛 , 𝑛 > 2, and let Ξ“ = πœ•Ξ© ∩ 𝐷 (π‘₯, π‘Ÿ), where π‘₯ ∈ πœ•Ξ©. Again, let U, V be two positive (non-proportional) harmonic functions in Ξ© vanishing continuously on Ξ“. If Ξ© is assumed 26 to be a Lipschitz domain, then [9] claims that U/V makes sense on Ξ“ and is additionally a HΓΆlder function on Ξ“ (boundary Harnack principle). Here is a question. Let 𝑅 be a real analytic function on 𝐷 (π‘₯, π‘Ÿ), π‘₯ ∈ Ξ“, and let U/V = 𝑅 on Ξ“ ∩ 𝐷 (π‘₯, π‘Ÿ). Is it true that Ξ“ ∩ 𝐷 (π‘₯, π‘Ÿ) is real analytic, maybe with the exception of some lower dimensional singular set? 27 BIBLIOGRAPHY [1] A. Baranov and K. Fedorovskiy. β€œBoundary regularity of Nevanlinna domains and univalent functions in model subspaces”. In: Sbornik: Mathematics 202.12 (Dec. 2011), pp. 1723– 1740. DOI: 10.1070/sm2011v202n12abeh004205. [2] Y. S. Belov and K. Y. Fedorovskiy. β€œModel spaces containing univalent functions”. In: Russian Mathematical Surveys 73.1 (Feb. 2018), pp. 172–174. DOI: 10.1070/rm9807. [3] Y. Belov, A. Borichev, and K. Fedorovskiy. β€œNevanlinna domains with large boundaries”. In: Journal of Functional Analysis 277.8 (Oct. 2019), pp. 2617–2643. DOI: 10.1016/j.jfa. 2018.12.015. [4] J. J. Carmona, P. V. Paramonov, and K. Y. Fedorovskiy. β€œOn uniform approximation by polyanalytic polynomials and the Dirichlet problem for bianalytic functions”. In: Sbornik: Mathematics 193.10 (Oct. 2002), pp. 1469–1492. DOI: 10.1070/sm2002v193n10abeh000 690. [5] J. Cima and W. Ross. The Backward Shift on the Hardy Space. American Mathematical Society, June 2000. DOI: 10.1090/surv/079. [6] K. Dyakonov and D. Khavinson. Smooth functions in star-invariant subspaces. 2006. DOI: 10.1090/conm/393/07371. [7] K. Y. Fedorovskiy. β€œOn some properties and examples of Nevanlinna domains”. In: Proceedings of the Steklov Institute of Mathematics 253.1 (July 2006), pp. 186–194. DOI: 10.1134/s0081543806020155. [8] K. Y. Fedorovskiy. β€œUniform 𝑛-analytic polynomial approximations of functions on recti- fiable contours in C𝑛 ”. In: Mathematical Notes 59 (4 Apr. 1, 1996), pp. 435–439. DOI: 10.1007/BF02308692. [9] D. S. Jerison and C. E. Kenig. β€œBoundary behavior of harmonic functions in non-tangentially accessible domains”. In: Advances in Mathematics 46.1 (Oct. 1982), pp. 80–147. DOI: 10.1016/0001-8708(82)90055-x. [10] M. Y. Mazalov. β€œAn example of a nonconstant bianalytic function vanishing everywhere on a nowhere analytic boundary”. In: Mathematical Notes 62.4 (Oct. 1997), pp. 524–526. DOI: 10.1007/bf02358988. [11] M. Y. Mazalov. β€œExample of a non-rectifiable Nevanlinna contour”. In: St. Petersburg Mathematical Journal 27.4 (June 2016), pp. 625–630. DOI: 10.1090/spmj/1409. [12] M. Y. Mazalov. β€œOn Nevanlinna domains with fractal boundaries”. In: St. Petersburg Mathematical Journal 29.5 (July 2018), pp. 777–791. DOI: 10.1090/spmj/1516. 28 [13] M. Sakai. β€œRegularity of a boundary having a Schwarz function”. In: Acta Mathematica 166.0 (1991), pp. 263–297. DOI: 10.1007/bf02398888. [14] C. L. Siegel. Topics in complex function theory - Elliptic Functions and Uniformization Theory. Vol. 1. New York: Wiley, 1988. ISBN: 9780471608448. [15] A. Volberg. β€œOn the dimension of harmonic measure of Cantor repellers”. In: Michigan Mathematical Journal 40.2 (Jan. 1993). DOI: 10.1307/mmj/1029004751. 29 CHAPTER 2 THE BUFFON’S NEEDLE PROBLEM FOR RANDOM PLANAR DISK-LIKE CANTOR SETS 2.1 Introduction Let 𝐸 be a subset of the unit disk, D. The Buffon needle problem wants to determine the probability with which a random needle or line intersects 𝐸 provided that it already intersects the unit disk. At the same time, let 𝑙 πœƒ be the line passing through the origin and forming angle πœƒ with the horizontal axis. The Favard length of 𝐸 is the average length of the projection of 𝐸 onto 𝑙 πœƒ when averaging over all angles πœƒ. It turns out these two quantities are proportional. Now, consider the following picture: let us have 𝐿 many (𝐿 β‰₯ 3) disjoint closed disks (𝐷 1 , . . . , 𝐷 𝐿 ) of diameter 1/𝐿 and strictly inside D. These are disks of the first generation. Consider also a piecewise affine map 𝑓 = ( 𝑓1 , . . . , 𝑓 𝐿 ) from those disks onto D. Then, 𝑓 βˆ’1 (D) = 𝐷 1 βˆͺ Β· Β· Β· βˆͺ 𝐷 𝐿 . Furthermore, 𝑓 βˆ’1 (𝐷 1 βˆͺ Β· Β· Β· βˆͺ 𝐷 𝐿 ) is consists of 𝐿 2 disks (groups of 𝐿 many disks in each 𝐷 𝑖 ); we call those disks of the second generation. We can iterate this procedure: denoting by π‘ˆπ‘› the union of disks of the 𝑛-th generation, where π‘ˆ1 := 𝐷 1 βˆͺ Β· Β· Β· βˆͺ 𝐷 𝐿 , we form the self-similar Cantor set K = ∞ Γ‘ 𝑛=1 π‘ˆπ‘› . This has positive and finite 1-dimensional Hausdorff measure; thus it is completely unrectifiable in the sense of Besicovitch [8]; and thus its Favard length is zero [8]. Of course, the disks can be replaced by other shapes. For example, π‘ˆ1 can consist of 𝐿 disjoint squares with side-length 1/𝐿 inside the unit square [0, 1] 2 (where the word β€œstrictly” can be omitted but β€œdisjoint” cannot). One of such Cantor sets is a rather β€œfamous”, namely the 1/4-corner Cantor set, K1/4 (see [7]). The 𝐿 βˆ’π‘› -neighbourhood of such sets is roughly π‘ˆπ‘› , and therefore its Favard length Fav(π‘ˆπ‘› ) β†’ 0, as 𝑛 β†’ ∞. But what is Fav(π‘ˆπ‘› ), or what is the speed with which Fav(π‘ˆπ‘› ) decreases? Nobody knows exactly, but there has been considerable interest in recent years. It is now clear that the answer may depend on several factors; the magnitude of 𝐿; the geometry of π‘ˆ1 ; the subtle algebraic and number 30 theoretic properties of a certain trigonometric sum built by the centres of the disks of the first generation. See [2, 3, 4, 6, 10] and the survey paper [5]. For the 1/4-corner Cantor set K1/4 in particular, the best known estimate from above for its 4βˆ’π‘› -neighbourhood is πΆπœ– Fav(𝑁4βˆ’π‘› (K1/4 )) ≀ 1 , βˆ€πœ– > 0, 𝑛 6 βˆ’πœ– for all large 𝑛. We suspect that this estimate can be improved to πΆπœ– Fav(𝑁4βˆ’π‘› (K1/4 )) ≀ , βˆ€πœ– > 0, 𝑛1+πœ– but at this moment this is only a conjecture. On the other hand, there is a universal estimate from below obtained in [9] for every self-similar Cantor set constructed as above: 𝑐 Fav(𝑁4βˆ’π‘› (K)) β‰₯ . (2.1.1) 𝑛 For any concrete set, this bound from below could be improved. In fact, it is proven in [1] that for the same 1/4-corner Cantor set K1/4 𝑐 log 𝑛 Fav(𝑁4βˆ’π‘› (K1/4 )) β‰₯ . 𝑛 For random Cantor sets the situation should be simpler. With large probability, Mattila’s lower estimate (2.1.1) is met by the same estimate from above (with a different constant). The problem is that in general there can be many different models of randomness. In this note, we are interested in an analogue of the random Cantor set appearing in [11] and in [14]. In our case, this will come from the random Cantor disks constructed below at Section 2.2. The model of randomness presented here is somewhat different from the ones in the above two papers, but it amazingly exhibits the same behaviour, as we’ll see below in our main Theorem 2.3.1, which we contrast with [11, Theorem 2.2] and [14, Theorem 1]. In particular, we prove an analogue of [14, Theorem 1]. Unfortunately, the randomness of the disk model we study here is not equivalent to that of the random (square) Cantor set R = ∞ Γ‘ 𝑛=0 R 𝑛 31 from [11], but it is nonetheless closer compared to the one constructed in [14]. The essential π‘›βˆ’1 difference between [14] and our consideration are the angles πœ”1𝑛 , πœ”2𝑛 , . . . , πœ”4𝑛 , which are here allowed to be distinct and independent whereas in [14] are all equal. So, our model is a little β€œmore random” than the random Cantor sets of Zhang in [14]. We introduce our notations β€”some borrowed from [14]β€” in the next Section 2.2. The problem of interest, namely the Favard length of a random planar disk-like Cantor set, is explained in Section 2.3. Our results and their proofs are postponed to Sections 2.4 and 2.5. In Section 2.6, we compare the differences and difficulties between our work and that of Peres and Solomyak’s and Zhang’s. 2.2 Cantor Disks Our work will be heavy on notation; without any ado let us introduce our basic β€œvocabulary”. The letter 𝑛 will stand for a (large) positive integer. The letter πœ” will be used to denote angles with values inside the interval [0, πœ‹2 ]. Now, let us consider a word of length 𝑛 made of the alphabet of angles in [0, πœ‹2 ], i.e. a word of the form πœ”1 πœ”2 Β· Β· Β· πœ”π‘› . The subscript in πœ” π‘˜ denotes the position of the angle πœ” π‘˜ within such a word of length 𝑛. We refer to the position of an angle within a word as the depth of that angle. Our operators, which we will introduce below, are such that every choice of an angle, say, πœ”1 necessitates four different independent choices for the angle πœ”2 ; every choice of the angle πœ”2 necessitates four different independent choices for the angle πœ”3 ; and so on up until depth 𝑛 where we will have 4π‘›βˆ’1 different angles πœ”π‘› . In order to differentiate between all those, for each 𝑗 𝑗 π‘˜ = 1, 2, . . . , 4 π‘˜βˆ’1 we write πœ” π‘˜π‘˜ for the 𝑗 π‘˜ -th choice of an angle πœ” π‘˜ at depth π‘˜. Notice there are 4 π‘˜βˆ’1 such choices. Therefore, a typical word from our alphabet of angles looks as follows, where 𝑗 we note that πœ” π‘˜π‘˜ ∈ [0, πœ‹2 ]: 𝑗1 = 1 , 𝑗 π‘˜ = 1, 2, . . . , 4 π‘˜βˆ’1 , 𝑗 𝑗 𝑗 𝑗 πœ”11 πœ”22 Β· Β· Β· πœ” π‘˜π‘˜ Β· Β· Β· πœ”π‘›π‘› where 𝑗2 = 1, 2, 3, 4 , Β·Β·Β· Β·Β·Β· 𝑗 𝑛 = 1, 2, . . . , 4π‘›βˆ’1 . 32 At certain instances, we need to consider cumulatively all angles of a certain depth; given a collection of words of length 𝑛, for each π‘˜ = 1, 2, . . . , 𝑛 let πœ”β€²π‘˜ be the collection of all 4 π‘˜βˆ’1 many angles at depth π‘˜, that is πœ”β€²π‘˜ = (πœ”1π‘˜ , . . . , πœ”4π‘˜ π‘˜βˆ’1 ). With this notation, we may use the symbols πœ”1 , πœ”11 , πœ”11 , and πœ”β€²1 interchangeably as these all refer to the same single angle. 𝑗 All the above give to our angles the structure of a rooted tree of height 𝑛 with root πœ”1 and such that each parent has four children. The vertexes have values in [0, πœ‹2 ], and are independent from each other and from their predecessors and ancestors. This tree we denote by πœ”β€²1 Β· Β· Β· πœ”β€²π‘› ; the trimmed tree with root πœ”1 and height π‘˜ we denote as πœ”β€²1 Β· Β· Β· πœ”β€²π‘˜ (for any π‘˜ = 1, 2, . . . , 𝑛). For the 𝑗 subtree of height 𝑛 βˆ’ π‘˜ + 1 with root πœ” π‘˜π‘˜ , which reaches up to the leaves (that is, from depth π‘˜ till 𝑗 𝑗 depth 𝑛 with starting vertex πœ” π‘˜π‘˜ ) we write πœ”Β― π‘˜π‘˜ . Later on, we will be working with rooted subtrees 𝑗 𝑗 𝑗 of the form πœ”Β― π‘›βˆ’π‘˜+1 π‘›βˆ’π‘˜+1 . To reiterate, πœ”Β― π‘›βˆ’π‘˜+1 π‘›βˆ’π‘˜+1 consists of the angle πœ”π‘›βˆ’π‘˜+1 π‘›βˆ’π‘˜+1 (as its root) along with all 𝑗 the angles from depth 𝑛 βˆ’ π‘˜ + 1 till depth 𝑛 (which have πœ”π‘›βˆ’π‘˜+1 π‘›βˆ’π‘˜+1 as an ancestor). This has height 𝑗 π‘˜. Alternatively, πœ”Β― π‘›βˆ’π‘˜+1 π‘›βˆ’π‘˜+1 is the collection of all the words (from our alphabet of angles) which have 𝑗 depth π‘˜ and the first letter is πœ”π‘›βˆ’π‘˜+1 π‘›βˆ’π‘˜+1 . There are 4π‘›βˆ’π‘˜ such words. Next, we will need to introduce certain operators and sets. The main objects of interest will be the operators D π‘˜ (π‘˜ = 0, 1, . . . , 𝑛) which will act on trees of angles of depth π‘˜. To understand these we need some auxiliary constructions first. For any angle πœ” and for 𝛼 = 0, 1, 2, 3 consider the transformations 1 3 πœ‹ π‘‡π›Όπœ” (𝑧) = 𝑧 + 𝑒 (𝛼 2 βˆ’πœ”)𝑖 4 4 where 𝑧 is any number on the complex plane C. Observe that if D is the unit disk, 𝑇00 (D), 𝑇10 (D), 𝑇20 (D), and 𝑇30 (D) are disks of radius 1/4 centred respectively at (3/4, 0), (0, 3/4), (βˆ’3/4, 0), and (0, βˆ’3/4). Introducing an angle πœ” in π‘‡π›Όπœ” (D), rotates (about (0, 0)) the aforementioned disks by angle πœ” in the clockwise direction. 𝑗 𝑗 Moreover, given an angle πœ” π‘˜π‘˜ from depth π‘˜ let Ξ© π‘˜π‘˜ be the set 3 𝑗 Ø 1 πœ” π‘˜π‘—π‘˜ Ξ© π‘˜π‘˜ = 𝑇 (D). π‘˜βˆ’1 𝛼 𝛼=0 4 33 That is, Ξ© π‘˜π‘˜ is a collection of four disks of radius 4βˆ’π‘˜ with centres (0, Β±3/4 π‘˜ ) and (Β±3/4 π‘˜ , 0) rotated 𝑗 𝑗 clockwise by πœ” π‘˜π‘˜ . 𝑗 We also give an enumeration to all the disks for all depths. We number the disks of Ξ© π‘˜π‘˜ so π‘—π‘˜ 1 πœ” that 𝑇 π‘˜ (D) is the (4 𝑗 π‘˜ βˆ’ 3 + 𝛼)-th disk at depth π‘˜. We call this the π‘˜-depth enumeration (of 4 π‘˜βˆ’1 𝛼 π‘˜βˆ’1 1 πœ”1 1 πœ”1 1 πœ”4π‘˜ the disks lying at depth π‘˜). Illustratively, we note 𝑇 π‘˜ (D), 4 π‘˜βˆ’1 0 𝑇 π‘˜ (D), 4 π‘˜βˆ’1 1 𝑇 4 π‘˜βˆ’1 0 (D), and π‘˜βˆ’1 1 πœ”4π‘˜ 𝑇 4 π‘˜βˆ’1 3 (D) are respectively the 1st, 2nd, (4 π‘˜ βˆ’ 3)-th, and 4 π‘˜ -th disks of depth π‘˜. We retain this enumeration as we translate these disks at different positions on the plane. This will be useful to track down each disk at each step so that our subsequent constructions make better sense. Now, we are ready to introduce our main protagonists. The operator D π‘˜ acts on the collection of trees (of angles) of height π‘˜ and for each such tree outputs a certain collection of 4 π‘˜ disks of radius 4βˆ’π‘˜ . We define these inductively below. To begin with, set D0 = D to be the unit disk. Next, we define D1 by 3 πœ”1 Ø D1 (πœ”β€²1 ) = Ξ©11 = 𝑇𝛼 1 (D0 ), 𝛼=0 that is, D1 (πœ”β€²1 ) consists of four disks of radius 1/4 centred at (0, Β±3/4) and (Β±3/4, 0) rotated clockwise by πœ”1 . Recall these disks are enumerated as in Ξ©11 . For the operator D2 , consider a tree of height 2, πœ”β€²1 πœ”β€²2 , which consists of the angles πœ”11 , and πœ”12 , πœ”22 , πœ”32 , πœ”42 . Then, we define D2 (πœ”β€²1 πœ”β€²2 ) to be the collection of disks constructed as follows: Replace the 1st, 2nd, 3rd and 4th disk of D1 (πœ”β€²1 ) respectively by Ξ©12 , Ξ©22 , Ξ©32 and Ξ©42 . By replacing 𝑗 we mean the translation of Ξ©2 in such a way that (0, 0) is translated to the centre of the 𝑗-th disk of D1 (πœ”β€²1 ). Consequently, D2 (πœ”β€²1 πœ”β€²2 ) consists of 42 disks of radius 4βˆ’2 translated appropriately so that each Ξ©22 replaces one the disks from D1 (πœ”β€²1 ). The set, say, Ξ©12 is in fact a subset of the 1st disk of 𝑗 D1 (πœ”β€²); actually D2 (πœ”β€²1 πœ”β€²2 ) βŠ‚ D1 (πœ”β€²1 ). Again, the disks comprising D2 (πœ”β€²1 πœ”β€²2 ) are enumerated to match Ξ©12 , Ξ©22 , Ξ©32 and Ξ©42 as we described above. Also see the Figure 2.1 below. 34 D0 πœ”22 πœ”32 πœ”1 D2 (πœ”β€²1 πœ”β€²2 ) πœ”12 Ξ©12 πœ”12 D1 (πœ”β€²1 ) πœ”42 Figure 2.1 The collections D0 , D1 (πœ”β€²1 ), D2 (πœ”β€²1 πœ”β€²2 ) and Ξ©12 . Continuing inductively, the operator D π‘˜ acts on the tree πœ”β€²1 Β· Β· Β· πœ”β€²π‘˜ in this manner: Consider the collection D π‘˜βˆ’1 (πœ”β€²1 Β· Β· Β· πœ”β€²π‘˜βˆ’1 ). These are 4 π‘˜βˆ’1 many (enumerated) disks. Replace the 1st of them by Ξ©1π‘˜ , the 2nd of them by Ξ©2π‘˜ , etc., until every disk of D π‘˜βˆ’1 (πœ”β€²1 Β· Β· Β· πœ”β€²π‘˜βˆ’1 ) has been replaced by four 𝑗 smaller ones. This replacement is done so that (0, 0), as the β€œcentre” of Ξ© π‘˜ , is translated to the centre of the 𝑗-th disk of D π‘˜βˆ’1 (πœ”β€²1 Β· Β· Β· πœ”β€²π‘˜βˆ’1 ). That is, we substitute the 𝑗-th disk (from depth π‘˜ βˆ’ 1) with the (4 𝑗 βˆ’ 3)-, (4 𝑗 βˆ’ 2)-, (4 𝑗 βˆ’ 1)-, and 4 𝑗-th disks of depth π‘˜. The resulting collection, which has 4 π‘˜ many disks of radius 4βˆ’π‘˜ , is D π‘˜ (πœ”β€²1 . . . πœ”β€²π‘˜ ). It holds that D π‘˜ (πœ”β€²1 Β· Β· Β· πœ”β€²π‘˜ ) βŠ‚ Dπ‘˜βˆ’1 (πœ”β€²1 Β· Β· Β· πœ”β€²π‘˜βˆ’1 ). In the present work, we will study the collection of disks D𝑛 (πœ”β€²1 Β· Β· Β· πœ”β€²π‘› ) where the angles πœ” π‘˜π‘˜ 𝑗 (for 𝑗 π‘˜ = 1, . . . , 4 π‘˜βˆ’1 and all π‘˜ = 1, 2, . . . , 𝑛) of the tree πœ”β€²1 Β· Β· Β· πœ”β€²π‘› are chosen randomly with uniform and independent distributions on the interval [0, πœ‹2 ]. So, let us describe this picture once more before moving on further. The set D𝑛 (πœ”β€²1 Β· Β· Β· πœ”β€²π‘› ) consists of 4𝑛 disks of radius 4βˆ’π‘› . These can be separated into 4π‘›βˆ’1 groups of four, which are copies of 3 𝑗 Ø 1 πœ”π‘›π‘—π‘› Ω𝑛𝑛 = 𝑇 (D0 ) π‘›βˆ’1 𝛼 𝛼=0 4 35 (for 𝑗 𝑛 = 1, . . . , 4π‘›βˆ’1 ) translated appropriately within the unit disk. 2.3 Favard Length Recall the Favard length of a planar set 𝐸 βŠ‚ C is the integral ∫ πœ‹ 1 Fav(𝐸) = projπœƒ 𝐸 π‘‘πœƒ πœ‹ 0 where projπœƒ 𝐸 is the projection of 𝐸 onto the line with slope tan πœƒ passing through the origin, and | 𝐴| is the (1-dimensional) Lebesgue measure of 𝐴. Now, consider an infinite tree of angles from [0, πœ‹2 ] with root πœ”1 and four branches at each vertex, and let D be the limit set Γ™βˆž D= D𝑛 (πœ”β€²1 Β· Β· Β· πœ”β€²π‘› ). 𝑛=0 Notice that by construction, D a purely unrectifiable planar set. As such, Fav(D) = 0 and by dominated convergence Fav(D𝑛 (πœ”β€²1 Β· Β· Β· πœ”β€²π‘› )) β†’ 0 while 𝑛 β†’ ∞. In fact, if the angles are randomly chosen uniformly and independently over [0, πœ‹2 ], by dominated convergence and Fubini E[Fav(D)] = 0 and E[Fav(D𝑛 (πœ”β€²1 Β· Β· Β· πœ”β€²π‘› ))] β†’ 0 as 𝑛 β†’ ∞, where the expectation is taken over all such angles. The question arises as to the rate with which E[Fav(D𝑛 (πœ”β€²1 Β· Β· Β· πœ”β€²π‘› ))] goes to 0. This we answer in the following theorem: Theorem 2.3.1. Let 𝑛 ∈ N and consider a tree of angles of height 𝑛 with each vertex having 𝑗 four branches. Suppose that the angles πœ” π‘˜π‘˜ (for all 𝑗 π‘˜ = 1, 2, . . . , 4 π‘˜βˆ’1 and all π‘˜ = 1, 2, . . . , 𝑛) are chosen randomly with uniform and independent distributions on the interval [0, πœ‹2 ]. Also set πœ”β€²π‘˜ = (πœ”1π‘˜ , πœ”2π‘˜ , . . . , πœ”4π‘˜ π‘˜βˆ’1 ) for each π‘˜ = 1, 2, . . . , 𝑛. Then, there exists a constant 𝑐 > 0 such that for any πœƒ ∈ [0, πœ‹2 ] it holds that 𝑐 Eπœ”1β€² Β·Β·Β· πœ”β€²π‘› projπœƒ D𝑛 (πœ”β€²1 Β· Β· Β· πœ”β€²π‘› ) ≀ βˆ€π‘› ∈ N. (2.3.1) 𝑛 Consequently, 𝑐 Eπœ”1β€² Β·Β·Β· πœ”β€²π‘› [Fav(D𝑛 (πœ”β€²1 Β· Β· Β· πœ”β€²π‘› ))] ≀ βˆ€π‘› ∈ N (2.3.2) 𝑛 36 and also lim inf 𝑛 Fav(D𝑛 (πœ”β€²1 Β· Β· Β· πœ”β€²π‘› )) < ∞ βˆ€π‘› ∈ N almost surely. (2.3.3) π‘›β†’βˆž Clearly, (2.3.3) follows from (2.3.2) by an immediate application of Fatou’s lemma, whereas (2.3.2) follows from (2.3.1) through Fubini. 2.4 Statement and use of the main lemma The present and the following sections are dedicated to the proof of (2.3.1). Towards this goal, we need to introduce Lemma 2.4.1 below, which describes the decay of the average projection when transitioning from depth π‘˜ to depth π‘˜ + 1. The main difficulty will come from obtaining the square factor appearing in (2.4.1), which emanates from the naturally occurring overlap of the projections. From now on, suppose we are given a tree of angles of height 𝑛 with four branches at each vertex where the angles are uniformly and independently distributed random variables on the 𝑗 𝑗 interval [0, πœ‹2 ]. Recall that given such a tree πœ”Β― π‘›βˆ’π‘˜+1 π‘›βˆ’π‘˜+1 is the subtree of height π‘˜ with the vertex πœ”π‘›βˆ’π‘˜+1π‘›βˆ’π‘˜+1 as its root. Observe that πœ”Β― 11 = πœ”β€²1 Β· Β· Β· πœ”β€²π‘› is the full tree whilst πœ”Β― 𝑛𝑛 = πœ”π‘›π‘› ( 𝑗 𝑛 = 1, 2, . . . , 4π‘›βˆ’1 ) are 𝑗 𝑗 𝑗 the its leaves, i.e. trees of height 1. For any πœƒ ∈ [0, πœ‹2 ] and all π‘˜ = 1, 2, . . . , 𝑛, define the following quantities 𝑗 𝑗 𝐷 1𝑛 = Eπœ”Β― 𝑗𝑛 projπœƒ D1 ( πœ”Β― 𝑛𝑛 ) , 𝑗 𝑛 = 1, 2, . . . , 4π‘›βˆ’1 𝑛 𝑗 𝑗 𝐷 π‘˜π‘›βˆ’π‘˜+1 = Eπœ”Β― π‘—π‘›βˆ’π‘˜+1 projπœƒ D π‘˜ ( πœ”Β― π‘›βˆ’π‘˜+1π‘›βˆ’π‘˜+1 ) , 𝑗 π‘›βˆ’π‘˜+1 = 1, 2, . . . , 4π‘›βˆ’π‘˜ π‘›βˆ’π‘˜+1 𝐷 𝑛1 = 𝐷 1𝑛 = Eπœ”Β― 𝑗1 projπœƒ D𝑛 ( πœ”Β― 11 ) , 𝑗 𝑗 𝑗 1 = 1. 1 𝑗 Notice that, because we are averaging over the independent and identically distributed πœ” π‘˜π‘˜ , π‘›βˆ’π‘˜ 𝐷 1π‘˜ = 𝐷 2π‘˜ = Β· Β· Β· = 𝐷 4π‘˜ for any π‘˜ = 1, 2, . . . , 𝑛. Therefore, it suffices to work with 𝐷 1π‘˜ ; the rest should be identical. Also, note that 𝐷 1𝑛 = Eπœ”1β€² Β·Β·Β· πœ”β€²π‘› projπœƒ D𝑛 (πœ”β€²1 Β· Β· Β· πœ”β€²π‘› ) . Now, we are ready to state a simple but important lemma. Also, see [14, Lemma 2.1]. 37 𝑗 Lemma 2.4.1. With notation as above, if πœ” π‘˜π‘˜ are uniformly and independently distributed random variables on [0, πœ‹2 ], there exits a constant 𝑐 β‰₯ 4 such that for any 𝑛 ∈ N (and any πœƒ ∈ [0, πœ‹2 ]) 𝐷 1π‘˜+1 ≀ 𝐷 1π‘˜ βˆ’ π‘βˆ’1 (𝐷 1π‘˜ ) 2 for all π‘˜ = 1, . . . , 𝑛 βˆ’ 1. (2.4.1) In fact, we will see below that 𝑐 = 8. Provided this holds true we can give a very compact proof of Theorem 2.3.1 using induction: Proof of Theorem 2.3.1. Let 𝑐 be as in Lemma 2.4.1 and note that 𝐷 12 ≀ 𝐷 11 < 2 ≀ 𝑐 2. Also, 𝐷 11 < 𝑐. Next, assume 𝐷 1π‘˜ < 𝑐 π‘˜ for some 2 ≀ π‘˜ ≀ 𝑛 βˆ’ 1. From Lemma 2.4.1, and by the monotonicity of the function π‘₯ βˆ’ π‘₯ 2 /𝑐 in [0, 2𝑐 ], we see that 𝑐 𝑐 π‘˜ βˆ’1 𝑐 𝐷 1π‘˜+1 ≀ 𝐷 1π‘˜ βˆ’ π‘βˆ’1 (𝐷 1π‘˜ ) 2 < βˆ’ 2 =𝑐 2 < . π‘˜ π‘˜ π‘˜ π‘˜ +1 Therefore, 𝐷 1π‘˜ < 𝑐 π‘˜ holds for all for 1 ≀ π‘˜ ≀ 𝑛 βˆ’ 1 and thus for π‘˜ = 𝑛 we get 𝑐 Eπœ”1β€² Β·Β·Β· πœ”β€²π‘› projπœƒ D𝑛 (πœ”β€²1 Β· Β· Β· πœ”β€²π‘› ) = 𝐷 1𝑛 < . 𝑛 This is (2.3.1). Equation (2.3.2) follows after integrating with respect to πœƒ, and (2.3.3) after applying Fatou’s Lemma. β–‘ 2.5 Proving the main lemma Whatever follows is dedicated to the proof of (2.4.1). First, we rewrite the length of the projection of a set in more convenient way. Let 𝑙 πœƒ and 𝑙 πœƒβŠ₯ be two lines through the origin so that 𝑙 πœƒ forms an angle πœƒ with the horizontal axis and 𝑙 πœƒβŠ₯ is perpendicular to 𝑙 πœƒ . Also, let n be the unit normal vector of 𝑙 πœƒβŠ₯ . The length of the projection of a planar set 𝐸 βŠ‚ C onto the line 𝑙 πœƒ can be written as ∫ projπœƒ 𝐸 = {𝑑 ∈ R : (𝑙 πœƒβŠ₯ + 𝑑n) ∩ 𝐸 β‰  βˆ…} = 𝑑𝑑. (2.5.1) (𝑙 βŠ₯πœƒ +𝑑n)βˆ©πΈβ‰ βˆ… For brevity, we denote the line 𝑙 πœƒβŠ₯ + 𝑑n by 𝑙 πœƒβŠ₯ (𝑑) where 𝑑 ∈ R. Additionally, because of the symmetry of our considerations, we can assume without loss of generality that πœƒ = 0 β€”as we will average over all πœƒ at the end. So, we can simply omit writing πœƒ altogether from now on. 38 The idea behind Lemma 2.4.1 is to look at the collection D𝑛 (πœ”β€²1 Β· Β· Β· πœ”β€²π‘› ) at depth 𝑛 but β€œzoomed in” so that it looks like depth 1. Then, we go one level up and look at the disks of depth 𝑛 βˆ’ 1 and 𝑛 zooming in enough so that they to look like depth 2; and so forth. If we rewrite the projections in the form of (2.5.1), the average overlap at each level is of at least a square factor compared to the total average projection of the level above. This last comparison is paramount to the proof. It will follow from the fact that the disks in our constructions never get too close to one another. In fact, this observation is not true in the case of the random square Cantor sets, which is the reason why we cannot directly apply the arguments here to the setting of [11]. Let us proceed with the proof of (2.4.1). Fix some π‘˜ = 1, 2, . . . , 𝑛 and recall that by construction 3 πœ”1 Ø D1 (πœ”1π‘›βˆ’π‘˜+1 ) = 𝑇𝛼 π‘›βˆ’π‘˜+1 (D0 ). 𝛼=0 This means that each disk from the collection Dπ‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) lies inside one of the above four disks, and therefore we can separate D π‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) into four groups of disks depending on their positioning at depth 1. More precisely, for each 𝛼 = 0, 1, 2, 3 define Tπ›Όπ‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) as πœ”1 Γ™ Tπ›Όπ‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) = 𝑇𝛼 π‘›βˆ’π‘˜+1 (D0 ) Dπ‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ). That is, the set Tπ›Όπ‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) consists of those disks of D π‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) which lie inside the 41 -radius πœ”1 disk 𝑇𝛼 π‘›βˆ’π‘˜+1 (D0 ). We can think of Tπ›Όπ‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) as the East, North, West, and South parts of D π‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ), respectively for 𝛼 = 0, 1, 2, 3. From this definition, it is also clear that Ø3 D π‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) = Tπ›Όπ‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ). (2.5.2) 𝛼=0 In fact, Tπ›Όπ‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) depends only on the angle πœ”1π‘›βˆ’π‘˜+1 and the subtree πœ”Β― 4Β·1βˆ’3+𝛼 Β― 1+𝛼 π‘›βˆ’π‘˜+2 = πœ” π‘›βˆ’π‘˜+2 . (Recall our enumeration of the angles in Section 2.2.) Thus, we can write Tπ›Όπ‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) as Tπ›Όπ‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) = Tπ›Όπ‘˜ (πœ”1π‘›βˆ’π‘˜+1 , πœ”Β― 1+𝛼 π‘›βˆ’π‘˜+2 ). 39 2.5.1 Key observations There are two key observations regarding the sets Tπ›Όπ‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ). First, note that each point of πœ”1 the interval (βˆ’1, 1) can be covered by at most two of the projections proj 𝑇𝛼 π‘›βˆ’π‘˜+1 (D0 ) for different πœ”1 𝛼’s. Since Tπ›Όπ‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) βŠ‚ 𝑇𝛼 π‘›βˆ’π‘˜+1 (D0 ), the same holds true for proj Tπ›Όπ‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ); the intersection Γ‘ π‘˜ Β―1 𝛼 proj T𝛼 ( πœ” π‘›βˆ’π‘˜+1 ) is empty when the intersection is over more than two values of 𝛼. Second, we can compare the average projections of Tπ›Όπ‘˜ ( πœ”Β― π‘›βˆ’π‘˜+1 ) = Tπ›Όπ‘˜ (πœ”1π‘›βˆ’π‘˜+1 , πœ”Β― 1+𝛼 π‘›βˆ’π‘˜+2 ) and D π‘˜βˆ’1 ( πœ”Β― 1+𝛼 π‘›βˆ’π‘˜+2 ). Notice that both these collections consist of 4 π‘˜βˆ’1 many disks, which in fact have the same 𝑛-depth enumerations. This means that they correspond to same disks of the collection D𝑛 (πœ”β€²1 Β· Β· Β· πœ”β€²π‘› ). The difference is that the disks of the former are translated according to D1 (πœ”1π‘›βˆ’π‘˜+1 ) and have radius 4βˆ’π‘˜ , whereas the ones of the latter have radius 4βˆ’(π‘˜βˆ’1) . Consequently, Tπ›Όπ‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) is a shifted copy of D π‘˜βˆ’1 ( πœ”Β― 1+𝛼 π‘›βˆ’π‘˜+2 ) dilated by a factor of 1/4. As such, the (average of the) projections of Tπ›Όπ‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) and D π‘˜βˆ’1 ( πœ”Β― 1+𝛼 π‘›βˆ’π‘˜+2 ) should also differ by a factor of 1/4. In other words, for any 𝛼 = 0, 1, 2, 3 we have Eπœ”Β― 1 proj Tπ›Όπ‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) = Eπœ”1 Eπœ”Β― 1+𝛼 proj Tπ›Όπ‘˜ (πœ”1π‘›βˆ’π‘˜+1 , πœ”Β― 1π‘›βˆ’π‘˜+2 ) π‘›βˆ’π‘˜+1 π‘›βˆ’π‘˜+1 π‘›βˆ’π‘˜+2 (2.5.3) 1 = Eπœ”Β― 1+𝛼 proj D π‘˜βˆ’1 ( πœ”Β― 1+𝛼 π‘›βˆ’π‘˜+2 ) . 4 π‘›βˆ’π‘˜+2 2.5.2 The estimates Utilising the above, we can now estimate 𝐷 1π‘˜ in terms of 𝐷 1π‘˜βˆ’1 : 𝐷 1π‘˜ = Eπœ”Β― 1 proj D π‘˜1 ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) π‘›βˆ’π‘˜+1 βˆ‘οΈ3 ≀ Eπœ”Β― 1 proj Tπ›Όπ‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) βˆ’ Eπœ”Β― 1 proj T0π‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) ∩ proj T1π‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) π‘›βˆ’π‘˜+1 π‘›βˆ’π‘˜+1 𝛼=0 3 (2.5.3) 1 βˆ‘οΈ ======== Eπœ”Β― 1+𝛼 proj D π‘˜βˆ’1 ( πœ”Β― 1+𝛼 π‘›βˆ’π‘˜+2 ) 4 𝛼=0 π‘›βˆ’π‘˜+2 βˆ’ Eπœ”Β― 1 proj T0π‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) ∩ proj T1π‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) π‘›βˆ’π‘˜+1 40 where the inequality follows from the first observation above (see 2.5.1) and (2.5.2). This in turn gives 1 1 𝐷 1π‘˜ ≀ (𝐷 + 𝐷 2π‘˜βˆ’1 + 𝐷 3π‘˜βˆ’1 + 𝐷 4π‘˜βˆ’1 ) βˆ’ Eπœ”Β― 1 proj T0π‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) ∩ proj T1π‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) 4 π‘˜βˆ’1 π‘›βˆ’π‘˜+1 (2.5.4) = 𝐷 1π‘˜βˆ’1 βˆ’ Eπœ”Β― 1 proj T0 ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) π‘˜ ∩ proj T1 ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) π‘˜ , π‘›βˆ’π‘˜+1 since 𝐷 1+𝛼 1 π‘˜βˆ’1 = 𝐷 π‘˜βˆ’1 for any 𝛼 = 0, 1, 2, 3. The next step, is to estimate the overlap term proj T0π‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) ∩ proj T1π‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) from below. For this, recall T0π‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) and T1π‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) depend (aside from πœ”1π‘›βˆ’π‘˜+1 ) respectively on πœ”Β― 1π‘›βˆ’π‘˜+2 and πœ”Β― 2π‘›βˆ’π‘˜+2 . First, we average with respect to the subtrees πœ”Β― 1π‘›βˆ’π‘˜+2 and πœ”Β― 2π‘›βˆ’π‘˜+2 , and afterwards we integrate over their common ancestor πœ”1π‘›βˆ’π‘˜+1 . To simplify the notation, let us write πœ”Β― 1,2 π‘›βˆ’π‘˜+2 for both the subtrees πœ”Β― 1π‘›βˆ’π‘˜+2 and πœ”Β― 2π‘›βˆ’π‘˜+2 , and also πœ“ for the angle πœ”1π‘›βˆ’π‘˜+1 . Then, we have Eπœ”Β― 1,2 proj T0π‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) ∩ proj T1π‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) π‘›βˆ’π‘˜+2 = Eπœ”Β― 1,2 proj T0π‘˜ (πœ”1π‘›βˆ’π‘˜+1 , πœ”Β― 1π‘›βˆ’π‘˜+2 ) ∩ proj T1π‘˜ (πœ”1π‘›βˆ’π‘˜+1 , πœ”Β― 2π‘›βˆ’π‘˜+2 ) π‘›βˆ’π‘˜+2 ∫  (2.5.1) ======== Pπœ”Β― 1,2 𝑙 βŠ₯ (𝑑) ∩ T0π‘˜ (πœ”1π‘›βˆ’π‘˜+1 ,πœ”Β― 1π‘›βˆ’π‘˜+2 ) β‰  βˆ… π‘›βˆ’π‘˜+2  and 𝑙 βŠ₯ (𝑑) ∩ T1π‘˜ (πœ”1π‘›βˆ’π‘˜+1 , πœ”Β― 2π‘›βˆ’π‘˜+2 ) β‰  βˆ… 𝑑𝑑 ∫   = Pπœ”Β― 1,2 𝑙 βŠ₯ (𝑑) ∩ T0π‘˜ (πœ”1π‘›βˆ’π‘˜+1 ,πœ”Β― 1π‘›βˆ’π‘˜+2 ) β‰  βˆ… Β· π‘›βˆ’π‘˜+2   Β·Pπœ”Β― 1,2 𝑙 βŠ₯ (𝑑) ∩ T1π‘˜ (πœ”1π‘›βˆ’π‘˜+1 , πœ”Β― 2π‘›βˆ’π‘˜+2 ) β‰  βˆ… 𝑑𝑑 π‘›βˆ’π‘˜+2 ∫   = Pπœ”Β― 1 𝑙 βŠ₯ (𝑑) ∩ T0π‘˜ (πœ”1π‘›βˆ’π‘˜+1 ,πœ”Β― 1π‘›βˆ’π‘˜+2 ) β‰  βˆ… Β· π‘›βˆ’π‘˜+2   Β·Pπœ”Β― 2 𝑙 βŠ₯ (𝑑) ∩ T1π‘˜ (πœ”1π‘›βˆ’π‘˜+1 , πœ”Β― 2π‘›βˆ’π‘˜+2 ) β‰  βˆ… 𝑑𝑑 π‘›βˆ’π‘˜+2 =: E (πœ”1π‘›βˆ’π‘˜+1 ). The 3rd equality above holds because for a fixed angle πœ”1π‘›βˆ’π‘˜+1 the events {𝑙 βŠ₯ (𝑑) ∩ Tπ›Όπ‘˜ (πœ”1π‘›βˆ’π‘˜+1 , πœ”Β― 1+𝛼 π‘›βˆ’π‘˜+2 ) β‰  βˆ…} are independent. 41 It would be very nice if these two events would have the same probability. Then at the end, we would use HΓΆlder inequality to get ∫ h  i 2 E (πœ”1π‘›βˆ’π‘˜+1 ) = Pπœ”Β― 1 𝑙 βŠ₯ (𝑑) ∩ T0π‘˜ (πœ”1π‘›βˆ’π‘˜+1 , πœ”Β― 1π‘›βˆ’π‘˜+2 ) β‰  βˆ… 𝑑𝑑 π‘›βˆ’π‘˜+2 ∫   2 βŠ₯ π‘˜ 1 1 β‰₯𝐢 Pπœ”Β― 1 𝑙 (𝑑) ∩ T0 (πœ”π‘›βˆ’π‘˜+1 , πœ”Β― π‘›βˆ’π‘˜+2 ) β‰  βˆ… 𝑑𝑑 . π‘›βˆ’π‘˜+2 However, this is not the case. For brevity, let us temporarily denote 3 πœ“ := πœ”1π‘›βˆ’π‘˜+1 and 𝑠(πœ“) := (1 βˆ’ cos πœ“). (2.5.5) 4 Also, set 𝐹 (𝑑) := {𝑙 βŠ₯ (𝑑) ∩ T0π‘˜ (0, πœ”Β― 1π‘›βˆ’π‘˜+2 ) β‰  βˆ…}. (2.5.6) For fixed πœ“, the events {𝑙 βŠ₯ (𝑑) ∩ T0π‘˜ (πœ“, πœ”Β― 1π‘›βˆ’π‘˜+2 ) β‰  βˆ…} and {𝑙 βŠ₯ (𝑑) ∩ T1π‘˜ (πœ“, πœ”Β― 2π‘›βˆ’π‘˜+2 ) β‰  βˆ…} do not have the same probability; one should take into consideration that the probability of the non- empty intersection with 𝑙 βŠ₯ (𝑑) for the first T has the same probability as the non-empty intersection with 𝑙 βŠ₯ (𝑑 + 𝑠(πœ“))) with T0π‘˜ (0, πœ”Β― 1π‘›βˆ’π‘˜+2 ). (Notice what happens with πœ“!) And the probability of the non-empty intersection with 𝑙 βŠ₯ (𝑑) for the second T has the same probability as the non-empty intersection with 𝑙 βŠ₯ (𝑑 + 𝑠(πœ“ + πœ‹2 ))) for the event T0π‘˜ (0, πœ”Β― 1π‘›βˆ’π‘˜+2 ). In fact, a simple geometric consideration shows the following holds: Lemma 2.5.1. With notation as above we have that     Pπœ”Β― 1 𝑙 βŠ₯ (𝑑) ∩ T0π‘˜ (πœ“, πœ”Β― 1π‘›βˆ’π‘˜+2 ) β‰  βˆ… = Pπœ”Β― 1 𝑙 βŠ₯ (𝑑 + 𝑠(πœ“)) ∩ T0π‘˜ (0, πœ”Β― 1π‘›βˆ’π‘˜+2 ) β‰  βˆ… . (2.5.7) π‘›βˆ’π‘˜+2 π‘›βˆ’π‘˜+2 42 In other words, Lemma 2.5.1 shows that E (πœ“) = ∫   = Pπœ”Β― 1 𝑙 βŠ₯ (𝑑 + 𝑠(πœ“)) ∩ T0π‘˜ (0, πœ”Β― 1π‘›βˆ’π‘˜+2 ) β‰  βˆ… Β· π‘›βˆ’π‘˜+2  πœ‹  βŠ₯ π‘˜ 1 Β·Pπœ”Β― 1 𝑙 (𝑑 + 𝑠(πœ“ + )) ∩ T0 (0, πœ”Β― π‘›βˆ’π‘˜+2 ) β‰  βˆ… 𝑑𝑑 ∫ π‘›βˆ’π‘˜+2 2 πœ‹ = 𝐹 (𝑑 + 𝑠(πœ“)) Β· 𝐹 (𝑑 + 𝑠(πœ“ + ))𝑑𝑑. 2 Next, if we integrate over πœ“ ∈ [0, πœ‹2 ], we get that the ∫ ∫ ∫ πœ‹ Expectation of the overlap = E (πœ“)π‘‘πœ“ = 𝐹 (𝑑 + 𝑠(πœ“)) Β· 𝐹 (𝑑 + 𝑠(πœ“ + ))π‘‘πœ“π‘‘π‘‘. 2 Let’s make this change of variables: 𝑒 = 𝑑 + 34 (1 βˆ’ cos πœ“) and 𝑣 = 𝑑 + 43 (1 βˆ’ cos(πœ“ + πœ‹2 )). The Jacobian of this change is at most 43 , and thus ∫ ∫ ∫ 2 4 4 Expectation of the overlap β‰₯ 𝐹 (𝑒)𝐹 (𝑣)𝑑𝑒𝑑𝑣 = 𝐹 (𝑑)𝑑𝑑 . 3 3 Since there is no dependence on πœ”Β― 3π‘›βˆ’π‘˜+1 or πœ”Β― 4π‘›βˆ’π‘˜+1 , we get Eπœ”Β― 1 proj T0π‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) ∩ proj T1π‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) π‘›βˆ’π‘˜+1 = Expectation of the overlap 4 2 β‰₯ Eπœ”Β― 1 proj T0π‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) 3 π‘›βˆ’π‘˜+1 (2.5.3) 4 1  2 ======== Β· Eπœ”Β― 1 proj D π‘˜βˆ’1 ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) 3 16 π‘›βˆ’π‘˜+2 1 = (𝐷 1π‘˜βˆ’1 ) 2 . 12 Finally, combing the fact that 1 Eπœ”Β― 1 proj T0π‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) ∩ proj T1π‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) β‰₯ (𝐷 1π‘˜βˆ’1 ) 2 π‘›βˆ’π‘˜+1 12 with (2.5.4) and setting 𝑐 = 12 we get 𝐷 1π‘˜ ≀ 𝐷 1π‘˜βˆ’1 βˆ’ π‘βˆ’1 (𝐷 1π‘˜βˆ’1 ) 2 and Lemma 2.4.1 is proved. 43 2.6 Comparison with the other random models The random Cantor set in [14] is a very close relative of the random Cantor set in this note, the difference is that Zhang’s random construction of 𝑛 generations has 𝑛 independent rotations involved, whereas our construction has 1 + Β· Β· Β· + 4π‘›βˆ’1 independent rotations. There the disks of generation π‘˜ are rotated by the same angle πœ” π‘˜ , while in this note we have 4 π‘˜βˆ’1 independent rotations of disks of generation π‘˜. Naturally, it is more difficult to work in a more chaotic model such as ours, and the techniques here use independence in a more involved way than in [14]. It is just a little harder to make sense of the combinatorics involved in our model. On the other hand, there are many β€œcommon places”: the use of overlap as the way to see the rate of decays of successive approximations of the random Cantor set, the use of Lemma 2.4.1, as well as the technical Lemma 2.5.1. Concerning [11], there are two main differences which create difficulties. The first is the fact that at most two of the projections projπœƒ Tπ›Όπ‘˜ ( πœ”Β― 1π‘›βˆ’π‘˜+1 ) can intersect at each point on the line 𝑙 πœƒ . This is equivalent to line 𝑙 πœƒβŠ₯ (𝑑) intersecting at most two of the disks for any 𝑑, and is key to the square factor appearing in our calculations. However, this is simply not true in the case of squares. In fact, in the Peres and Solomyak case the corresponding line 𝑙 πœƒβŠ₯ (𝑑) can simultaneously intersect 3 squares of generation π‘˜ for any π‘˜ and any 𝑑. Because of this, the inequalities appearing here cannot be translated directly in the square setting. But even if this wasn’t an obstacle, the reader should pay attention to Lemma 2.5.1. Let’s pretend that we can repeat everything before this lemma for the model of Peres and Solomyak. The role of the angle πœ”1π‘›βˆ’π‘˜+1 will be played by the β€œFavard angle” πœƒ, the shift function 𝑠(πœ”1π‘›βˆ’π‘˜+1 ) will be replaced by 1 𝑆(πœƒ) = sin πœƒ, 2 and all seems to be following smoothly along the same lines. Also, the following equality ∫ ∫ E1{𝑙 βŠ₯ (𝑑)∩T π‘˜ (πœ”Β― 1 )β‰ βˆ… } π‘‘πœƒ = E1{𝑙 βŠ₯ (𝑑+𝑠(πœƒ))∩T π‘˜ (πœ”Β― 1 )β‰ βˆ… } π‘‘πœƒ, (2.6.1) πœƒ 0 π‘›βˆ’π‘˜+2 πœƒ 1 π‘›βˆ’π‘˜+2 44 which would be the analogue of (2.5.7), makes sense in principle if we understand πœ”β€™s as the random variables in the Peres–Solomyak model, which assume the values 0, 1, 2, 3 (instead of values in the interval [0, πœ‹2 ] as in our’s and Zhang’s models). But, there is a caveat. We reduced the function of two variables   𝐺 (πœ“, 𝑑) := π‘ƒπœ”Β― 1 𝑙 βŠ₯ (𝑑) ∩ T0π‘˜ (πœ“, πœ”Β― 1π‘›βˆ’π‘˜+2 ) β‰  βˆ… π‘›βˆ’π‘˜+2 to the composition with a function of one variable and the shift (see (2.5.6) for the definition of 𝐹): 𝐺 (πœ“, 𝑑) = 𝐺 (0, 𝑑 + 𝑠(πœ“)) = 𝐹 (𝑑 + 𝑠(πœ“)) (2.6.2) thanks to (2.5.7). But looking at (2.6.1), we can notice that the function G(πœƒ, 𝑑) := E1{𝑙 βŠ₯ (𝑑)∩T π‘˜ (πœ”Β― 1 )β‰ βˆ… } πœƒ 0 π‘›βˆ’π‘˜+2 cannot be written as some F (𝑑 + 𝑆(πœƒ)). As a result of this misfortune, we cannot write ∫ ∫ ∫ Expectation of the overlap = E (πœƒ)π‘‘πœƒ = F (𝑑 + 𝑆(πœƒ)) Β· F (𝑑)π‘‘πœƒπ‘‘π‘‘ ∫ as before. Working similarly, this would in turn bring about the term ( F 𝑑𝑑) 2 . Instead, we only have that ∫ ∫ ∫ Expectation of the overlap = E (πœƒ)π‘‘πœƒ = G(πœƒ, 𝑑) Β· G(πœƒ, 𝑑 + 𝑆(πœƒ)))π‘‘πœƒπ‘‘π‘‘, and it is not clear (at least to us) how to estimate this integral from below as no change of variables seems to be of help. 45 BIBLIOGRAPHY [1] M. Bateman and A. Volberg. β€œAn estimate from below for the Buffon needle probability of the four-corner Cantor set”. In: Mathematical Research Letters 17.5 (2010), pp. 959–967. DOI: 10.4310/mrl.2010.v17.n5.a12. [2] M. Bond, I. Łaba, and A. Volberg. β€œBuffon’s needle estimates for rational product Cantor sets”. In: American Journal of Mathematics 136.2 (2014), pp. 357–391. DOI: 10.1353/aj m.2014.0013. [3] M. Bond and A. Volberg. β€œBuffon needle lands in πœ–-neighborhood of a 1-dimensional Sierpinski Gasket with probability at most | log πœ– | βˆ’π‘ ”. In: Comptes Rendus Mathematique 348.11-12 (June 2010), pp. 653–656. DOI: 10.1016/j.crma.2010.04.006. [4] M. Bond and A. Volberg. β€œBuffon’s needle landing near Besicovitch irregular self-similar sets”. In: Indiana University Mathematics Journal 61.6 (2012), pp. 2085–2109. DOI: 10.1512/iumj.2012.61.4828. [5] I. Łaba. β€œRecent Progress on Favard Length Estimates for Planar Cantor Sets”. In: Operator- Related Function Theory and Time-Frequency Analysis. Springer International Publishing, Sept. 2014, pp. 117–145. DOI: 10.1007/978-3-319-08557-9_5. [6] I. Łaba and K. Zhai. β€œThe Favard length of product Cantor sets”. In: Bulletin of the London Mathematical Society 42.6 (Aug. 2010), pp. 997–1009. DOI: 10.1112/blms/bdq059. [7] P. Mattila. β€œHausdorff dimension, projections, and the Fourier transform”. In: Publicacions MatemΓ tiques 48 (Jan. 2004), pp. 3–48. DOI: 10.5565/publmat_48104_01. [8] P. Mattila. Fourier Analysis and Hausdorff Dimension. Cambridge University Press, 2015. DOI: 10.1017/cbo9781316227619. [9] P. Mattila. β€œHausdorff dimension, orthogonal projections and intersections with planes”. In: Annales Academiae Scientiarum Fennicae Series A I Mathematica 1 (1975), pp. 227–244. DOI: 10.5186/aasfm.1975.0110. [10] F. Nazarov, Y. Peres, and A. Volberg. β€œThe power law for the Buffon needle probability of the four-corner Cantor set”. In: St. Petersburg Mathematical Journal 22.1 (Feb. 2011), pp. 61–61. DOI: 10.1090/s1061-0022-2010-01133-6. [11] Y. Peres and B. Solomyak. β€œHow likely is Buffon’s needle to fall near a planar Cantor set?” In: Pacific Journal of Mathematics 204.2 (June 2002), pp. 473–496. DOI: 10.2140/pjm.20 02.204.473. [12] T. Tao. β€œA quantitative version of the Besicovitch projection theorem via multiscale analysis”. In: Proceedings of the London Mathematical Society 98.3 (Oct. 2008), pp. 559–584. DOI: 46 10.1112/plms/pdn037. [13] X. Tolsa. β€œAnalytic capacity, rectifiability, and the Cauchy integral”. In: Proceedings of the International Congress of Mathematicians Madrid, August 22–30. European Mathematical Society Publishing House, 2006, pp. 1505–1527. DOI: 10.4171/022-2/71. [14] S. Zhang. β€œThe exact power law for Buffon’s needle landing near some random Cantor sets”. In: Revista MatemΓ‘tica Iberoamericana 36.2 (Dec. 2019), pp. 537–548. DOI: 10.4171/rmi/1138. 47 CHAPTER 3 GEOMETRY OF PLANAR CURVES INTERSECTING MANY LINES IN A FEW POINTS 3.1 The statement of the problem The problem at hand is to better understand the structure of Borel sets in R2 that have a small intersection with parallel shifts of lines from a whole cone. Here, we work only with sets that are graphs and continuous curves. So we have strong assumptions. But the results claim some estimate on the Hausdorff measure (not merely the Hausdorff dimension). Initially, we show that a function’s graph intersecting all parallel shifts of lines from a nonde- generate cone in at most two points is locally Lipschitz and also present a counter-example showing this fails if more intersection points are allowed. Next, we prove that any curve that has finitely many intersections with a cone of lines is 𝜎-finite with respect to Hausdorff length and we find a bound on the Hausdorff measure of each β€œpiece.” On the other hand, in [1] it was shown that, given countably many graphs of functions, there is another function whose graph has only one intersection with all shifts of the given graphs but whose graph has dimension 2. This result shows that there is a β€œthick” graph having only one intersection with all shifts of countably many other graphs. In our turn, we show that the graph having finitely many intersection with shifts of the whole cone of linear functions must be in fact very β€œthin”. Proposition 3.1.1. Let πœ† > 0 be a fixed number and consider all the cones of lines with slopes between πœ† and βˆ’πœ† (containing the vertical line). If 𝑓 : (0, 1) β†’ R is a continuous function such that any line of these cones intersects its graph at at most two points, then 𝑓 is locally Lipschitz. Notice that our hypothesis implies that no three points of the graph of 𝑓 can lie on the same line that is a parallel shift of a line from a given cone. For the proof we will need the following lemmas. Lemma 3.1.2. Every convex (or concave) function on an open interval is locally Lipschitz. 48 Figure 3.1 Each line from any cone intersects the graph at at most two points. Lemma 3.1.3. If a function 𝑔 : (0, 1) β†’ R is continuous and has a unique local extremum, π‘₯, ˜ inside (0, 1), then it is strictly monotone in (0, π‘₯] ˜ and [ π‘₯, ˜ 1) with opposite monotonicity on each interval. Proof of Lemma 3.1.3. Suppose π‘₯˜ is a local minimum for 𝑔. We will show that 𝑔 is strictly monotone increasing in [ π‘₯, ˜ 1). Assume the contrary, i.e., consider two points π‘₯ 1 < π‘₯2 ∈ [π‘₯, ˜ 1) such that 𝑔(π‘₯1 ) β‰₯ 𝑔(π‘₯ 2 ). On the compact interval [π‘₯ 1 , π‘₯2 ], the function 𝑔 has to attain a minimum and a maximum, which respectively are at π‘₯ 2 and π‘₯1 otherwise the uniqueness of π‘₯˜ is contradicted. If π‘₯ 1 = π‘₯, ˜ the point π‘₯˜ is not a local minimum and so π‘₯˜ < π‘₯ 1 . Again, π‘₯˜ and π‘₯ 1 must be the minimum and maximum, respectively, of 𝑔 in [ π‘₯, ˜ π‘₯ 1 ], which in turn says π‘₯1 is a local maximum contradicting the uniqueness of π‘₯. ˜ Therefore, 𝑔(π‘₯ 1 ) < 𝑔(π‘₯ 2 ) and 𝑔 is strictly monotone increasing on [ π‘₯, ˜ 1). Similarly, on (0, π‘₯] ˜ 𝑔 is (strictly) monotone decreasing and the same arguments work for when π‘₯˜ is local maximum. β–‘ 𝑓 (π‘₯)βˆ’ 𝑓 (𝑦) Proof. Consider the slope function of 𝑓 , 𝑆(π‘₯, 𝑦) = π‘₯βˆ’π‘¦ , and note that 𝑓 (π‘₯) βˆ’ 𝑓 (𝑦) 𝑆(π‘₯, 𝑦) = = 𝜁 ⇐⇒ 𝑓 (π‘₯) βˆ’ 𝜁π‘₯ = 𝑓 (𝑦) βˆ’ 𝜁 𝑦. π‘₯βˆ’π‘¦ If for any two points π‘₯ < 𝑦 ∈ (0, 1) we have |𝑆(π‘₯, 𝑦)| < πœ†, then 𝑓 is Lipschitz (with Lipschitz constant at most πœ†). 49 Figure 3.2 𝑆(π‘₯0 , 𝑦 0 ) = πœ†β€² β‰₯ πœ†; The part of the graph of 𝑓 between π‘₯ 0 and 𝑦 0 cannot lie on different sides of πœ–πœ†β€² . Now suppose that there exist π‘₯0 , 𝑦 0 ∈ (0, 1) for which |𝑆(π‘₯ 0 , 𝑦 0 )| β‰₯ πœ† and consider the case where 𝑆(π‘₯ 0 , 𝑦 0 ) = πœ†β€² β‰₯ πœ†. Since 𝑆(π‘₯, 𝑦) = 𝑆(𝑦, π‘₯), we may assume that π‘₯ 0 < 𝑦 0 . We will denote the line passing through (π‘₯ 0 , 𝑓 (π‘₯ 0 )) and (𝑦 0 , 𝑓 (𝑦 0 )) by πœ–πœ†β€² . If there are numbers π‘₯0 < π‘Ž < 𝑏 < 𝑦 0 such that (𝑆(π‘₯ 0 , π‘Ž) βˆ’ πœ†β€²)(𝑆(π‘₯ 0 , 𝑏) βˆ’ πœ†β€²) ≀ 0, 𝑓 (π‘₯0 )βˆ’ 𝑓 (𝑐) then by the continuity of 𝑆(π‘₯, Β· ) there has to exist a number 𝑐 ∈ [π‘Ž, 𝑏] such that π‘₯ 0 βˆ’π‘ = 𝑓 (π‘₯ 0 )βˆ’ 𝑓 (𝑦 0 ) πœ†β€² = π‘₯0 βˆ’π‘¦ 0 . But this means that (π‘₯ 0 , 𝑓 (π‘₯ 0 )), (𝑐, 𝑓 (𝑐)) and (𝑦 0 , 𝑓 (𝑦 0 )) are colinear, which contradicts our hypothesis and therefore 𝑆(π‘₯ 0 , 𝑦) has to be constantly greater or constantly less than πœ†β€² for π‘₯0 < 𝑦 < 𝑦 0 (see Figure 3.2). For the same reasons 𝑆(π‘₯ 0 , 𝑦) has to be constantly greater or constantly less than πœ†β€² also for 𝑦 > 𝑦 0 and the same holds for 𝑆(π‘₯, 𝑦 0 ) for π‘₯ < π‘₯ 0 . Graphically, this means that πœ–πœ†β€² separates 𝑓 in three parts that do not intersect πœ–πœ†β€² ; one before π‘₯ 0 , one over (π‘₯0 , 𝑦 0 ), and one after 𝑦 0 . We proceed to show that the part over (π‘₯ 0 , 𝑦 0 ) lies on a different side of πœ–πœ†β€² from the other two. Let us consider the case when 𝑆(π‘₯ 0 , 𝑦) < πœ†β€² for π‘₯ 0 < 𝑦 < 𝑦 0 . Then, the function 𝑓 (π‘₯) βˆ’ πœ†β€²π‘₯ defined on [π‘₯ 0 , 𝑦 0 ] attains a maximum at π‘₯0 and at 𝑦 0 (which also implies that 𝑆(π‘₯, 𝑦 0 ) > πœ†β€² for π‘₯ 0 < π‘₯ < 𝑦 0 ) and let π‘¦Λœ ∈ (π‘₯0 , 𝑦 0 ) be the point where 𝑓 (π‘₯) βˆ’ πœ†β€²π‘₯ attains a minimum (see Figure 3.4). 50 𝑆(π‘₯ 0 , 𝑦) > πœ†β€² 𝑆(π‘₯ 0 , 𝑦) < πœ†β€² Figure 3.3 The two cases when π‘₯ 0 < 𝑦 < 𝑦 0 . Figure 3.4 If 𝑆(π‘₯ 0 , 𝑦) < πœ†β€² for every 𝑦 βˆ‰ (π‘₯ 0 , 1) \ {𝑦 0 }, by moving the line πœ–πœ†β€² slightly down, we get three points of intersection. Now, suppose additionally that 𝑆(π‘₯0 , 𝑦) < πœ†β€² also for 𝑦 > 𝑦 0 . Pick a number π‘˜ with 𝑓 (π‘₯0 ) βˆ’ πœ†β€²π‘₯ 0 > π‘˜ > max{ 𝑓 ( π‘¦Λœ ) βˆ’ πœ†β€² π‘¦Λœ , 𝑓 (𝑦) βˆ’ πœ†β€² 𝑦} for some 𝑦 > 𝑦 0 . Then, we have simultaneously 𝑓 ( π‘¦Λœ ) βˆ’ πœ†β€² π‘¦Λœ < π‘˜ < 𝑓 (π‘₯0 ) βˆ’ πœ†β€²π‘₯ 0 , 𝑓 ( π‘¦Λœ ) βˆ’ πœ†β€² π‘¦Λœ < π‘˜ < 𝑓 (𝑦 0 ) βˆ’ πœ†β€² 𝑦 0 , 𝑓 (𝑦) βˆ’ πœ†β€² 𝑦 < π‘˜ < 𝑓 (𝑦 0 ) βˆ’ πœ†β€² 𝑦 0 . The continuity of 𝑓 and the above inequalities imply that there must exist numbers π‘Ž, 𝑏, and 𝑐 in (π‘₯ 0 , π‘¦Λœ ), ( π‘¦Λœ , 𝑦 0 ), and (𝑦 0 , 𝑦) respectively such that 𝑓 (π‘Ž) βˆ’ πœ†β€² π‘Ž = 𝑓 (𝑏) βˆ’ πœ†β€² 𝑏 = 𝑓 (𝑐) βˆ’ πœ†β€² 𝑐 = π‘˜ 51 Figure 3.5 If 𝑓 attains a local minimum at another point π‘¦Λœ β€² > π‘¦Λœ , we can find a line of slope greater than πœ†β€² intersecting 𝑓 at three points. which implies that (π‘Ž, 𝑓 (π‘Ž)), (𝑏, 𝑓 (𝑏)), and (𝑐, 𝑓 (𝑐)) are colinear, a contradiction, and therefore 𝑆(π‘₯ 0 , 𝑦) has to be greater than πœ†β€² for 𝑦 > 𝑦 0 . Working similarly, we see that 𝑆(π‘₯, 𝑦 0 ) < πœ†β€² for π‘₯ < π‘₯0 . An identical argument gives us that π‘¦Λœ is the only point in [π‘₯ 0 , 𝑦 0 ], and eventually in [π‘₯0 , 1), where 𝑓 (π‘₯) βˆ’ πœ†β€²π‘₯ attains a local minimum (see Figure 3.5) and from Lemma 3.1.3 we deduce that 𝑓 (π‘₯) βˆ’ πœ†β€²π‘₯ has to be monotone increasing in [ π‘¦Λœ , 1). Hence, for any π‘₯, 𝑦 β‰₯ π‘¦Λœ we have: π‘₯<𝑦 π‘₯ < 𝑦 ⇐⇒ 𝑓 (π‘₯) βˆ’ πœ†β€²π‘₯ < 𝑓 (𝑦) βˆ’ πœ†β€² 𝑦 ⇐=β‡’ 𝑆(π‘₯, 𝑦) > πœ†β€² . However, observe that for any π‘₯ and 𝑦 for which 𝑆(π‘₯, 𝑦) > πœ†β€², the function 𝑆(π‘₯, Β· ) has to be 1-1 otherwise our hypothesis fails in a similar way as above and, since it is continuous, it has to be monotone in (π‘₯, 1) for every π‘₯ ∈ [ π‘¦Λœ , 1). Therefore, 𝑓 is either convex or concave in [ π‘¦Λœ , 1) and thus locally Lipschitz in ( π‘¦Λœ , 1) thanks to Lemma 3.1.2. In particular, 𝑓 has to be convex in [ π‘¦Λœ , 1). Indeed, assume 𝑓 is concave and let π‘₯ be any number in ( π‘¦Λœ , 𝑦 0 ), see Figure 3.6. By concavity, the point ( π‘¦Λœ , 𝑓 ( π‘¦Λœ )) has to lie below the line passing through (𝑦 0 , 𝑓 (𝑦 0 )) with slope 𝜁 = 𝑆(π‘₯, 𝑦 0 ) and, since 𝜁 = 𝑆(π‘₯, 𝑦 0 ) > 𝑆(π‘₯ 0 , 𝑦 0 ) = πœ†β€² β‰₯ πœ†, the point (π‘₯ 0 , 𝑓 (π‘₯ 0 )) lies above. Hence, this line will intersect the graph of 𝑓 at some point (𝑐, 𝑓 (𝑐)) with 52 Figure 3.6 𝑆(π‘₯ 0 , 𝑦) has to be strictly monotone increasing in (𝑦 0 , 1). 𝑐 ∈ (π‘₯ 0 , π‘¦Λœ ) and the points (𝑐, 𝑓 (𝑐)), (π‘₯, 𝑓 (π‘₯)), and (𝑦 0 , 𝑓 (𝑦 0 )) are colinear, a contradiction. If we instead assume that 𝑆(π‘₯ 0 , 𝑦) > πœ†β€² for π‘₯ 0 < 𝑦 < 𝑦 0 , working similarly we conclude that there must exist π‘¦Λœ ∈ [π‘₯ 0 , 𝑦 0 ] such that 𝑓 is concave in (0, π‘¦Λœ ]. The case when there exist π‘₯0 , 𝑦 0 ∈ (0, 1) for which 𝑆(π‘₯0 , 𝑦 0 ) = πœ†β€² ≀ βˆ’πœ† is identical and gives us the reverse implications. To sum up, we conclude that there are points π‘₯, ˜ π‘¦Λœ ∈ (0, 1) such that 𝑓 has some particular convexity on (0, π‘₯] ˜ and on [ π‘¦Λœ , 1). These intervals cannot overlap, because otherwise 𝑓 would be a line segment of slope at least πœ† (or at most βˆ’πœ†) on [ π‘¦Λœ , π‘₯], ˜ which contradicts our hypothesis and so π‘₯˜ ≀ π‘¦Λœ . Let π‘₯˜ be the maximal point so that 𝑓 is, for instance, convex on (0, π‘₯], ˜ and π‘¦Λœ the minimal so that 𝑓 is convex on [ π‘¦Λœ , 1). When π‘₯˜ β‰  π‘¦Λœ , for every points π‘₯, 𝑦 ∈ [ π‘₯, ˜ π‘¦Λœ ] we have |𝑆(π‘₯, 𝑦)| ≀ πœ† and 𝑓 is Lipschitz in [π‘₯, ˜ π‘¦Λœ ] with Lipschitz constant πœ†. This concludes the proof. β–‘ Of course, any continuous function that satisfies the condition of the proposition and has different convexity on (π‘Ž, π‘₯] ˜ and on [ π‘¦Λœ , 𝑏) has to additionally satisfy limπ‘₯β†’π‘Ž+ ,𝑦→𝑏 βˆ’ |𝑆(π‘₯, 𝑦)| < πœ†. Furthermore, notice that the fact that the cone is vertical (or at least that it contains the vertical line) is essential to get the locally Lipschitz property. Indeed, if 𝐢 is a cone avoiding the vertical √ line, we can restrict the function 3 π‘₯ to a sufficiently small interval around 0 so that it intersects all √ the lines of the cone at at most two points. But 3 π‘₯ is clearly not Lipschitz around 0. However, we 53 Figure 3.7 All the possible ways the graph of 𝑓 can look like. do have the following corollary. Corollary. Let πœ† 1 > 0 > πœ† 2 be some fixed numbers and consider all the cones of lines with slopes between πœ† 1 and πœ† 2 (containing the vertical line). If 𝑓 : (0, 1) β†’ R is a continuous function satisfying the same condition as above, then it is locally Lipschitz. Proof. The inequalities |𝑆(π‘₯, 𝑦)| < πœ† and |𝑆(π‘₯, 𝑦)| β‰₯ πœ† in this case correspond to πœ†2 < 𝑆(π‘₯, 𝑦) < πœ†1 and 𝑆(π‘₯, 𝑦) β‰₯ πœ† 1 π‘œπ‘Ÿ 𝑆(π‘₯, 𝑦) ≀ πœ† 2 , respectively. The proof is the same as before and on the regions where 𝑓 is not convex or concave it is Lipschitz with Lipschitz constant the maximum of πœ† 1 and βˆ’πœ† 2 . β–‘ Remark. All the above remains true for any interval (π‘Ž, 𝑏). It is not hard to see that the same proof also works in the case where 𝑓 is defined on a closed interval, but Lemma 3.1.2 cannot be used in this setting. However, if 𝑓 : [0, 1] β†’ R, its restriction 𝑓 |(0,1) is locally Lipschitz. 3.2 An example It is natural then to ask whether our assumption still gives us the locally Lipschitz property when we allow more points of intersection. It turns out this fails even for at most 3 points of intersection in the sense that there can be infinitely many points around where the function cannot be locally Lipschitz. Here, we construct such a function whose graph intersects a certain cone of lines at at most three points. 54 1 Consider the sequence π‘Ž π‘˜ = 2 βˆ’ 21π‘˜ for π‘˜ β‰₯ 1, and on the each of the intervals [π‘Ž π‘˜ , π‘Ž π‘˜+1 ] define a continuous function 𝑓 π‘˜ with the following properties: i) 𝑓1 (0) = 0, 𝑓1 ( 41 ) = 𝑓2 ( 41 ) = πœ†4 ; ii) 𝑓 π‘˜+1 (π‘Ž π‘˜+1 ) = 𝑓 π‘˜ (π‘Ž π‘˜+1 ); 1 iii) 𝑓 π‘˜ (π‘Ž π‘˜+1 ) = 2 ( 𝑓 π‘˜ (π‘Ž π‘˜ ) + 𝑓 π‘˜βˆ’1 (π‘Ž π‘˜βˆ’1 )); iv) 𝑓2π‘˜ is monotone decreasing and convex on [π‘Ž 2π‘˜ , π‘Ž 2π‘˜+1 ] and 𝑓2π‘˜βˆ’1 is monotone increasing and concave on [π‘Ž 2π‘˜βˆ’1 , π‘Ž 2π‘˜ ]; v) the tangent line to 𝑓 π‘˜ at (π‘Ž π‘˜ , 𝑓 π‘˜ (π‘Ž π‘˜ )) is vertical. Let 𝑓 : [0, 1] β†’ R be the function given by ο£±  𝑓 π‘˜ (π‘₯) if π‘₯ ∈ [π‘Ž π‘˜ , π‘Ž π‘˜+1 ),       ο£²  𝑓 (π‘₯) = 𝑓 π‘˜ (1 βˆ’ π‘₯) if π‘₯ ∈ (1 βˆ’ π‘Ž π‘˜+1 , 1 βˆ’ π‘Ž π‘˜ ],     1  πœ†   6 if π‘₯ = 2 ο£³ for all π‘˜ β‰₯ 1 (Figure 3.8), which is clearly continuous in (0, 1) \ { 12 } because of (ii). Observe that 𝑏 π‘˜ +𝑏 π‘˜βˆ’1 the sequence (𝑏 π‘˜ ) = ( 𝑓 π‘˜ (π‘Ž π‘˜ )) is recursively defined by 𝑏 π‘˜+1 = 2 (through property (iii)) and 𝑏 π‘˜+1 βˆ’π‘ π‘˜ it converges. In particular, we have 𝑏 π‘˜ βˆ’π‘ π‘˜βˆ’1 = βˆ’ 12 and therefore  βˆ’1  π‘˜βˆ’1   βˆ’1  π‘˜βˆ’1  1 𝑏 π‘˜+1 = 𝑏 π‘˜ + (𝑏 2 βˆ’ 𝑏 1 ) =β‡’ 𝑏 π‘˜+1 = 𝑏2 βˆ’ 1 βˆ’ (𝑏 2 βˆ’ 𝑏 1 ). (3.2.1) 2 3 2 In our case, we have 𝑏 1 = 𝑓1 (0) = 0, 𝑏 2 = 𝑓2 ( 14 ) = πœ†4 , and also   βˆ’1  π‘˜βˆ’1  πœ† 𝑓 π‘˜ (π‘Ž π‘˜ ) = 1 βˆ’ , 6 2 hence lim π‘˜β†’+∞ 𝑓 π‘˜ (π‘Ž π‘˜ ) = πœ† 6. But note that for every π‘₯ ∈ (0, 12 ) there is an 𝑛 β‰₯ 1 for which π‘₯ ∈ [π‘Ž 𝑛 , π‘Ž 𝑛+1 ) and, since each 𝑓 π‘˜ is monotone in [π‘Ž π‘˜ , π‘Ž π‘˜+1 ) for every π‘˜, we get   min 𝑓𝑛 (π‘Ž 𝑛 ), 𝑓𝑛+1 (π‘Ž 𝑛+1 ) ≀ 𝑓 (π‘₯) ≀ max 𝑓𝑛 (π‘Ž 𝑛 ), 𝑓𝑛+1 (π‘Ž 𝑛+1 ) . 55 Figure 3.8 At most 3 points of intersection with any line inside the cones. Therefore, we have limπ‘₯β†’ 1 βˆ’ 𝑓 (π‘₯) = πœ† 6 = 𝑓 ( 21 ), and similarly for π‘₯ ∈ ( 12 , 1), which means that 𝑓 is 2 also continuous at 12 . However, by construction 𝑓 is locally Lipschitz on (0, 1) \ { 12 } except at around π‘Ž π‘˜ and 1 βˆ’ π‘Ž π‘˜ , 1 1 π‘˜ β‰₯ 1, and therefore it is not locally Lipschitz around 2 either, because π‘Ž π‘˜ β†’ 2 as π‘˜ β†’ +∞. Now we proceed to show the graph of 𝑓 has at most 3 intersection points with any line inside a vertical cone with slopes between πœ† and βˆ’πœ†. Each 𝑓 π‘˜ is monotone and has certain concavity on [π‘Ž π‘˜ , π‘Ž π‘˜+1 ], hence its graph is contained inside the triangle π‘‡π‘˜ with vertices (π‘Ž π‘˜ , 𝑓 (π‘Ž π‘˜ )), (π‘Ž π‘˜+1 , 𝑓 π‘˜+1 (π‘Ž π‘˜+1 )), and (π‘Ž π‘˜ , 𝑓 (π‘Ž π‘˜+1 )) (see Figure 3.9) and therefore any line intersecting the graph of 𝑓 (at at least two points) has to pass through some of these triangles. Notice, however, that if a line passes through two nonconsecutive triangles, say π‘‡π‘˜ and π‘‡π‘˜+ 𝑗 ( 𝑗 > 1), then it falls outside the admissible cone of lines. In particular, (because of properties (ii) through (iv)) each π‘‡π‘˜+1 is half the size of π‘‡π‘˜ and they are placed is such a way that the maximum and minimum slope a line through them can have are respectively the maximum and the minimum of the quantities 𝑓 π‘˜+ 𝑗 (π‘Ž π‘˜+ 𝑗 ) βˆ’ 𝑓 π‘˜ (π‘Ž π‘˜ ) 𝑓 π‘˜+ 𝑗 (π‘Ž π‘˜+ 𝑗+1 ) βˆ’ 𝑓 π‘˜ (π‘Ž π‘˜+1 ) and , π‘Ž π‘˜+ 𝑗 βˆ’ π‘Ž π‘˜ π‘Ž π‘˜+ 𝑗 βˆ’ π‘Ž π‘˜+1 when one of the numbers π‘˜ and π‘˜ + 𝑗 is even and the other is odd, and the maximum and minimum 56 of the quantities 𝑓 π‘˜+ 𝑗 (π‘Ž π‘˜+ 𝑗+1 ) βˆ’ 𝑓 π‘˜ (π‘Ž π‘˜ ) 𝑓 π‘˜+ 𝑗 (π‘Ž π‘˜+ 𝑗 ) βˆ’ 𝑓 π‘˜ (π‘Ž π‘˜+1 ) and , π‘Ž π‘˜+ 𝑗 βˆ’ π‘Ž π‘˜ π‘Ž π‘˜+ 𝑗 βˆ’ π‘Ž π‘˜+1 when π‘˜ and π‘˜ + 𝑗 are both even or both odd. Using (3.2.1) we can see that each of the above is bounded in absolute value by πœ† whenever 𝑗 > 1. For the same reasons any admissible line passing through ( 12 , πœ†6 ) intersects the graph only at that point, because πœ† 𝑓 π‘˜ (π‘Ž π‘˜ ) βˆ’ 6 πœ† 1 = < πœ†. π‘Žπ‘˜ βˆ’ 2 3 Therefore, the admissible lines intersecting the graph necessarily pass through two (or maybe only one) consecutive triangles and each such line intersects the graph of 𝑓 π‘˜ at at most two points because of (iv). Furthermore, due to the difference in concavity of 𝑓 π‘˜ and 𝑓 π‘˜+1 , a line cannot intersect both of their graphs at two points, because then it would need to have both negative and positive slope, which is absurd. An example of a sequence ( 𝑓 π‘˜ ) of functions with the above properties is the following:   βˆ’1  π‘˜βˆ’1  (βˆ’1) π‘˜+1πœ† √ πœ† 𝑓 π‘˜ (π‘₯) = 1 βˆ’ + π‘₯ βˆ’ π‘Žπ‘˜ . 6 2 2 2 π‘˜+1 3.3 Hausdorff measure Marstrand in [5, Theorem 6.5.III] proved that if a Borel set on the plane has the property that if the lines in a positive measure of directions intersect this Borel set at a set of Hausdorff (3.3.1) dimension zero, then the Hausdorff dimension of this Borel set is at most 1. In particular, this happens if the intersections are at most countable. The Borel assumption is essential. That said, Marstrand’s theorem does not in general guarantee the Hausdorff measure of the Borel set is finite. Our next goal will be to deal with the Hausdorff measure of a continuous curve and also generalise to arbitrarily many points of intersection with our cones (still finitely many, though). It turns out that the curve has to always be 𝜎-finite with respect to the H 1 measure. In order to proceed we need set up things more rigorously: 57 Figure 3.9 The case when π‘˜ and π‘˜ + 𝑗 are both odd. Notations. Let 𝐢 (πœ™, 0) = {(π‘₯, 𝑦) ∈ R2 : |𝑦| β‰₯ tan(πœ™) |π‘₯|} denote the vertical closed cone in between the lines through the point (0, 0) with slopes tan(πœ™) and βˆ’ tan(πœ™) (where 0 < πœ™ < πœ‹2 ). By 𝐢+ (πœ™, 0) we will denote the upper half of the cone 𝐢 (πœ™, 0), that is 𝐢+ (πœ™, 0) = {(π‘₯, 𝑦) ∈ R2 : |𝑦| β‰₯ tan(πœ™) |π‘₯|, 𝑦 β‰₯ 0}, and by πΆβˆ’ (πœ™, 0) its lower half. Let 𝐢 (πœ™, 𝜌) be the cone’s counter-clockwise rotation by angle 𝜌, 𝐢 (πœ™, 0, β„Ž) = 𝐡0 (β„Ž) ∩ 𝐢 (πœ™, 0), where 𝐡π‘₯ (π‘Ÿ) = 𝐡(π‘₯, π‘Ÿ) is the closed ball centred at π‘₯ with radius π‘Ÿ, and 𝐢𝑃 (πœ™, 0) the translation of 𝐢 (πœ™, 0) so that its vertex is the point 𝑃. Finally, 𝐢 βˆ— will denote the dual cone of 𝐢, that is 𝐢 βˆ— (πœ™, 0) = 𝐢 (πœ™, 0) 𝐢 . We will be combining different notation in a natural way, for example 𝐢+ (πœ™, 𝜌, β„Ž) is the upper half of the truncated and rotated cone with vertex at 0. 𝛾 : [0, 1] β†’ R2 will be a continuous curve. 58 3.3.1 The main hypothesis Fix an integer π‘˜ β‰₯ 2. Fix an angle πœ™ ∈ (0, πœ‹2 ) and a rotation 𝜌 ∈ [0, 2πœ‹). A line contained (3.3.2) inside the cone 𝐢𝑃 (πœ™, 𝜌) for any point 𝑃 ∈ R2 intersects the curve 𝛾 at at most π‘˜ points. Any such line will be called admissible. A cone consisting of only admissible lines will also be called admissible. 3.3.2 𝛾 is 𝜎-finite For simplicity and without loss of generality we will assume the the curve 𝛾 : [0, 1] β†’ R2 is bounded inside the unit square and that (0, 0), (1, 1) ∈ 𝛾. We additionally assume that the cones of our hypothesis are vertical, i.e., that 𝜌 = 0. Theorem 3.3.1. 𝛾 can be split into countably many sets 𝛾𝑛 with finite H 1 measure. In particular, 𝛾 is 1-rectifiable. The following lemma plays a key role in the proof of this theorem, but we will postpone its proof until later. Lemma 3.3.2. For every point 𝑃 ∈ 𝛾 there exists an admissible cone 𝐢𝑃 (πœƒ, 𝜌, β„Ž) that avoids the curve 𝛾 except at 𝑃, that is 𝐢𝑃 (πœƒ, 𝜌, β„Ž) ∩ 𝛾 = {𝑃}. In view of Lemma 3.3.2 β€” by slightly tilting 𝜌, enlarging πœƒ and monotone decreasing β„Ž β€” we may assume the triplet (πœƒ, 𝜌, β„Ž) consists of rational numbers. If {(πœƒ 𝑛 , πœŒπ‘› , β„Žπ‘› )} is an enumeration of all rational triples that still lie within our admissible set, then we can decomposed 𝛾 into the countably many sets  𝛾𝑛 = 𝑃 ∈ 𝛾 : 𝐢𝑃 (πœƒ 𝑛 , πœŒπ‘› , β„Žπ‘› ) ∩ 𝛾 = {𝑃} (see Figure 3.10). Note that 𝛾𝑛 are not necessarily disjoint for different values of 𝑛. We proceed to prove each one of them has finite H 1 measure. Note that this is not new knowledge and it can be found, for example, in [2, Lemma 3.3.5] or [6, Lemma 15.13] in a more general setup. Nevertheless, we present it here for completeness. For the rest of this section 𝑛 will be fixed. 59 Figure 3.10 The curve 𝛾 and its part 𝛾𝑛 for πœƒ 𝑛 , πœŒπ‘› = 0, and β„Žπ‘› . 2π‘˜ Lemma 3.3.3. H 1 (𝛾𝑛 ) < cos(πœƒ 𝑛 ) . Proof. Without loss of generality we may assume the cone 𝐢𝑃 (πœƒ 𝑛 , πœŒπ‘› , β„Žπ‘› ) is vertical, i.e., that πœŒπ‘› = 0. Let us now split the unit square into 𝑁 vertical strips, 𝑆 𝑗 ( 𝑗 = 1, 2, . . . , 𝑁), of base length 1 1 𝑁 with 𝑁 sufficiently large so that 𝑁 < cos(πœƒ 𝑛 ) β„Žπ‘› . Let 𝐽 be the set of indices 𝑗 for which 𝑆 𝑗 ∩ 𝛾𝑛 β‰  βˆ… and for any point 𝑃 ∈ 𝛾 denote the connected component of 𝛾 inside 𝑆 𝑗 through 𝑃 ∈ 𝑆 𝑗 ∩ 𝛾 by Ξ“π‘ƒβˆ— ( 𝑗). 1 Fix a 𝑗 ∈ 𝐽 and consider a point 𝑃 ∈ 𝑆 𝑗 ∩ 𝛾𝑛 . Since 𝑁 < cos(πœƒ 𝑛 ) β„Žπ‘› , the sides of 𝑆 𝑗 necessarily intersect both sides of the cone 𝐢𝑃 (πœƒ 𝑛 , 0, β„Žπ‘› ) creating thus two triangles both contained inside 1 (see Figure 3.11). For any point 𝑃′ ∈ 𝑆 𝑗 ∩ 𝛾𝑛 other than 𝑃 there are two  the ball 𝐡 𝑃 𝑁 cos(πœƒ 𝑛) 60 1 Figure 3.11 Each cone intersects a strip of length 𝑁 < cos(πœƒ 𝑛 ) β„Žπ‘› . cases: either |𝑃 βˆ’ 𝑃′ | ≀ β„Žπ‘› or |𝑃 βˆ’ 𝑃′ | > β„Žπ‘› . In the first case, the sets Ξ“π‘ƒβˆ— ( 𝑗) and Ξ“π‘ƒβˆ— β€² ( 𝑗) are both contained inside the two triangles πΆπ‘ƒβˆ— (πœƒ 𝑛 , 0) ∩ 𝑆 𝑗 . In the second, they are necessarily disjoint, because 𝐢𝑃 (πœƒ 𝑛 , 0, β„Žπ‘› ) is free from points of 𝛾 (other than 𝑃). These additionally imply that there 1 can be no more than sin(πœƒ 𝑛 ) β„Ž 𝑛 such distinct paths inside 𝑆 𝑗 . In particular,   𝑃 ∈ Ξ“π‘ƒβˆ— ( 𝑗) βŠ‚ 𝑆 𝑗 ∩ 𝛾 ∩ 𝐡 𝑃 (β„Žπ‘› ) βŠ‚ πΆπ‘ƒβˆ— (πœƒ 𝑛 , 0, β„Žπ‘› ) ∩ 𝑆 𝑗 βŠ‚ 𝐡 𝑃 1 𝑁 cos(πœƒ 𝑛 ) . Now, let P 𝑗 be a maximal set of points in 𝑆 𝑗 ∩ 𝛾𝑛 such that the sets Ξ“π‘ƒβˆ— ( 𝑗) for 𝑃 ∈ P 𝑗 are 1  all disjoint and observe that 𝑆 𝑗 ∩ 𝛾𝑛 is covered by the balls 𝐡 𝑃 𝑁 cos(πœƒ 𝑛 ) with 𝑃 ∈ P 𝑗 . Indeed, 1 Ð  if 𝑃0 ∈ 𝑆 𝑗 ∩ 𝛾𝑛 is not inside the set π‘ƒβˆˆP 𝑗 𝐡 𝑃 𝑁 cos(πœƒ 𝑛 ) , then by construction it is also outside 61 𝐡 𝑃 (β„Žπ‘› ) and therefore Ξ“π‘ƒβˆ— 0 ( 𝑗) and Ξ“π‘ƒβˆ— ( 𝑗) are disjoint for all 𝑃 ∈ P 𝑗 , which contradicts the Ð π‘ƒβˆˆP 𝑗 maximality of P 𝑗 . Moreover, due to the connectedness of 𝛾, the set {𝑃} has to be path-connected with (0, 0) and (1, 1) and therefore each Ξ“π‘ƒβˆ— ( 𝑗) has to intersect at least one side of the strip 𝑆 𝑗 . Hence, because of (3.3.2), there can be at most 2π‘˜ of these paths, i.e., #(P 𝑗 ) ≀ min{2π‘˜, sin(πœƒ1𝑛 ) β„Žπ‘› } ≀ 2π‘˜ for every 𝑗 ∈ 𝐽. Therefore, Ø   Ø Ø   1 1 𝛾𝑛 ∩ 𝑆 𝑗 βŠ‚ 𝐡𝑃 𝑁 cos(πœƒ 𝑛 ) =β‡’ 𝛾𝑛 βŠ‚ 𝐡𝑃 𝑁 cos(πœƒ 𝑛 ) π‘ƒβˆˆP 𝑗 𝑗 ∈𝐽 π‘ƒβˆˆP 𝑗 and the total sum of the radii of these balls is at most 1 2π‘˜ 2π‘˜ #(𝐽) ≀ . 𝑁 cos(πœƒ 𝑛 ) cos(πœƒ 𝑛 ) Finally, if π›ΎΛœ 𝑛 = {𝑃 ∈ 𝛾 : 𝐢𝑃 (πœƒ 𝑛 , 0, β„Žπ‘› /2) ∩ 𝛾 = {𝑃}}, then 𝛾𝑛 βŠ‚ π›ΎΛœ 𝑛 . Repeating the above 1 β„Žπ‘› construction with 𝑁 < cos(πœƒ 𝑛 ) 2 , we get a cover of π›ΎΛœ 𝑛 β€” and thus of 𝛾𝑛 β€” consisting of balls 2π‘˜ with a total sum of radii at most cos(πœƒ 𝑛 ) . The result follows. β–‘ Ð Remark. In the above construction we are in fact able to cover the whole part of 𝛾 inside 𝑗 ∈𝐽 𝑆𝑗 with the same balls, and not merely 𝛾𝑛 . Eventually, the curve 𝛾 has to be 𝜎-finite. 3.3.3 Cones free of 𝛾 Here we prove Lemma 3.3.2. Fix 𝑃 ∈ 𝛾. Since 𝛾 is bounded, there must exist an β„ŽΛœ > 0 such that 𝐢𝑃 (πœ™, 0)βˆ©π›Ύ = 𝐢𝑃 (πœ™, 0, β„Ž)βˆ©π›Ύ. ˜ If 𝐢𝑃 (πœ™β€², 0) ∩ 𝛾 = {𝑃} or 𝐢𝑃 (πœ™β€², 0, β„Ž) ∩ 𝛾 = {𝑃} for some πœ™β€² ∈ [πœ™, πœ‹2 ) and some β„Ž > 0, then we are done. Suppose this does not happen. Then, for all πœ™β€² ∈ [πœ™, πœ‹2 ) and for all sufficiently small β„Ž > 0 we have 𝐢𝑃 (πœ™β€², 0, β„Ž) ∩ 𝛾 \ {𝑃} β‰  βˆ…. (3.3.3) Lemma 3.3.4. For any 𝑃 ∈ 𝛾 the set 𝐢𝑃 (πœ™, 0) βˆ©π›Ύ has finitely many (closed) connected components. 62 Proof. Since 𝛾 is connected, every point of 𝐢𝑃 (πœ™, 0) ∩ 𝛾 has to be path-connected with the point 𝑃 through some part of the curve 𝛾. There are two possibilities: either that path is entirely contained inside 𝐢𝑃 (πœ™, 0) or it has to pass through its sides. If a path does not intersect the sides, then it necessarily has to pass through 𝑃 otherwise 𝛾 would not be connected. This yields precisely one connected component β€” the one containing 𝑃 β€” and all the rest (if any) have to intersect the sides of the cone. If these components are infinitely many, there have to exist also infinitely many points of intersection on the sides of the cone; at least one for each connected component. But this contradicts (3.3.2). β–‘ Remark. The connected components of Lemma 3.3.4 total at most 2π‘˜ and 𝑃 need not be a point of the curve. This lemma is still valid regardless of the cone we are working with as soon as it is in our admissible family of cones. Let Γ𝑃 (πœ™, 0) be the connected component of 𝐢𝑃 (πœ™, 0) ∩ 𝛾 that contains the point 𝑃, which because of (3.3.3) cannot be precisely the point set {𝑃}. Because of Lemma 3.3.4, the set 𝐢𝑃 (πœ™, 0) βˆ©π›Ύ\Γ𝑃 (πœ™, 0) is compact and thus there exists β„Ž0 > 0 such that 𝐢𝑃 (πœ™, 0, β„Ž0 )βˆ©π›Ύ βŠ‚ Γ𝑃 (πœ™, 0). Observe that 𝐢𝑃 (πœ™, 0) ∩ 𝛾 \ Γ𝑃 (πœ™, 0) could be empty in general in which case β„Ž0 = ∞, however, we can always assume that β„Ž0 ≀ β„Ž. ˜ Next, we bisect our cone into two new identical cones sharing one common side 𝐢𝑃 (πœ™, 0) = 𝐢𝑃 (πœ™1 , 𝜌1 ) βˆͺ 𝐢𝑃 (πœ™1 , βˆ’πœŒ1 ), πœ™ where πœ™1 = πœ‹ 4 + 2 and 𝜌1 = πœ‹ 4 βˆ’ πœ™2 , and repeat the above arguments for each new cone: If 𝐢𝑃 (πœ™β€², 𝜌1 ) ∩ 𝛾 = {𝑃} or 𝐢𝑃 (πœ™β€², 𝜌1 , β„Ž) ∩ 𝛾 = {𝑃} for some πœ™β€² ∈ [πœ™1 , πœ‹2 ) and some β„Ž > 0, then we are done. Similarly for βˆ’πœŒ1 in place of 𝜌1 . Suppose none of these happen. Then, for all πœ™β€² ∈ [πœ™1 , πœ‹2 ) and for all sufficiently small β„Ž and β„Žβ€² we have 𝐢𝑃 (πœ™β€², 𝜌1 , β„Ž) ∩ 𝛾 \ {𝑃} β‰  βˆ… and 𝐢𝑃 (πœ™β€², βˆ’πœŒ1 , β„Žβ€²) ∩ 𝛾 \ {𝑃} β‰  βˆ…. (3.3.4) 63 Figure 3.12 Finding a cone free from points of 𝛾. The parameters π‘Ÿ, 𝑑, and β„Ž determine the radius. We denote by Γ𝑃 (πœ™1 , 𝜌1 ) and Γ𝑃 (πœ™1 , βˆ’πœŒ1 ) the connected component of 𝐢𝑃 (πœ™1 , 𝜌1 ) ∩ 𝛾 and 𝐢𝑃 (πœ™1 , βˆ’πœŒ1 ) ∩ 𝛾 containing 𝑃, respectively. Then, the sets 𝐢𝑃 (πœ™1 , 𝜌1 ) ∩ 𝛾 \ Γ𝑃 (πœ™1 , 𝜌1 ) and 𝐢𝑃 (πœ™1 , βˆ’πœŒ1 ) ∩ 𝛾 \ ˜ such that Γ𝑃 (πœ™1 , βˆ’πœŒ1 ) are compact (thanks to Lemma 3.3.4) and thus there exist β„Ž1,0 , β„Ž1,1 ∈ (0, β„Ž] 𝐢𝑃 (πœ™1 , 𝜌1 , β„Ž1,0 ) ∩ 𝛾 βŠ‚ Γ𝑃 (πœ™1 , 𝜌1 ) and 𝐢𝑃 (πœ™1 , βˆ’πœŒ1 , β„Ž1,1 ) ∩ 𝛾 βŠ‚ Γ𝑃 (πœ™1 , βˆ’πœŒ1 ). We iterate this construction indefinitely (Figure 3.12). If at any step we get 𝐢𝑃 (πœ™β€², 𝜌, β„Ž) ∩ 𝛾 = {𝑃} (3.3.5) for some πœ™β€², 𝜌, and β„Ž, then we have found our desired cone and we stop. Otherwise, we get an 64 infinite sequence of smaller and smaller cones satisfying the following: {𝑃} ⊊ 𝐢𝑃 (πœ™π‘› , πœŒπ‘›,𝑖 , β„Žπ‘›,𝑖 ) ∩ 𝛾 βŠ‚ Γ𝑃 (πœ™π‘› , πœŒπ‘›,𝑖 ) βŠ‚ 𝐢𝑃 (πœ™π‘› , πœŒπ‘›,𝑖 ) for all 𝑖 = 0, 1, . . . , 2𝑛 βˆ’ 1 for all 𝑛 β‰₯ 0 where πœ‹ πœ™ πœ‹ πœ™π‘›βˆ’1 πœ™0 = πœ™ πœ™1 = + πœ™π‘› = + 4 2 4 2 πœ‹ πœ™ 2(πœ™π‘› βˆ’ πœ™) 𝜌0,0 = 0 𝜌1,0 = 𝜌1 = βˆ’ 𝜌1,1 = βˆ’πœŒ1 πœŒπ‘›,𝑖 = (πœ™π‘› βˆ’ πœ™) βˆ’ 𝑖 4 2 2𝑛 βˆ’ 1 β„Ž0,0 = β„Ž0 0 < β„Žπ‘›,𝑖 ≀ β„Ž. ˜ Note that at the 𝑛th iteration we have exactly 2𝑛 truncated closed cones separated by the lines  𝑙 𝑛,𝑖 = 𝑃 + (π‘₯, 𝑦) : 𝑦 = tan(πœ‹ βˆ’ πœ™π‘› + πœŒπ‘›,𝑖 ) π‘₯ through 𝑃. The sets Γ𝑃 (πœ™π‘› , πœŒπ‘›,𝑖 ) might intersect these lines, but this can happen at at most π‘˜ may points due to (3.3.2). Let π‘Ÿ 𝑛,𝑖 be the smallest distance between these points of intersection (if any) and 𝑃, that is  π‘Ÿ 𝑛,𝑖 = dist 𝑃, 𝑙 𝑛,𝑖 ∩ Γ𝑃 (πœ™π‘› , πœŒπ‘›,𝑖 ) \ {𝑃} (again we can arbitrarily set some 0 < π‘Ÿ 𝑛,𝑖 ≀ β„ŽΛœ if 𝑙 𝑛,𝑖 ∩ Γ𝑃 (πœ™π‘› , πœŒπ‘›,𝑖 ) \ {𝑃} = βˆ…) and let n o 𝑑𝑛,𝑖 = min sup{𝑑 (𝑃, Γ𝑃+ (𝑑) \ 𝑃) : 𝑑 ∈ (0, 1]}, sup{𝑑 (𝑃, Ξ“π‘ƒβˆ’ (𝑑) \ 𝑃) : 𝑑 ∈ (0, 1]} where Γ𝑃+ (𝑑) and Ξ“π‘ƒβˆ’ (𝑑) are parametrisations of the sets Γ𝑃 (πœ™π‘› , πœŒπ‘›,𝑖 )βˆ©πΆπ‘ƒ+ (πœ™π‘› , πœŒπ‘›,𝑖 ) and Γ𝑃 (πœ™π‘› , πœŒπ‘›,𝑖 )∩ πΆπ‘ƒβˆ’ (πœ™π‘› , πœŒπ‘›,𝑖 ) respectively (which in general could be precisely the point set {𝑃}) with Γ𝑃+ (0) = Ξ“π‘ƒβˆ’ (0) = 𝑃. Finally, we set β„Žπ‘› = min{π‘Ÿ 𝑛,𝑖 , 𝑑𝑛,𝑖 , β„Žπ‘›,𝑖 : 𝑖 = 0, 1, . . . , 2𝑛 βˆ’ 1}. Since the above set is finite, β„Žπ‘› > 0. From this construction for every 𝑛 β‰₯ 0 we get a collection of truncated cones 𝐢𝑃 (πœ™π‘› , πœŒπ‘›,𝑖 , β„Žπ‘› ), for 𝑖 = 0, 1, . . . , 2𝑛 βˆ’ 1, (see Figure 3.12) that have the following 65 property. There is a path (part of 𝛾) lying inside the cone that connects the point 𝑃 with at least one of the two arcs of length (πœ‹ βˆ’ 2πœ™π‘› )β„Žπ‘› which bound the cone 𝐢𝑃 (πœ™π‘› , πœŒπ‘›,𝑖 , β„Žπ‘› ). Moreover, (3.3.6) these paths avoid any other intersections with that cone’s boundary aside 𝑃 and the (closed) arc(s). Now, fix 𝑛 sufficiently large so that 2𝑛 β‰₯ 2π‘˜ + 3. Then, we can find at least π‘˜ + 2 of the cones 𝐢𝑃 (πœ™π‘› , πœŒπ‘›,𝑖 , β„Žπ‘› ) that contain some path of those mentioned at (3.3.6) all lying on the same half-cone, say on 𝐢𝑃+ (πœ™, 0, β„Žπ‘› ). Consider one of the sides of our initial cone 𝐢𝑃 (πœ™, 0), say 𝑙 = 𝑃 + {(π‘₯, 𝑦) : 𝑦 = tan(πœ™) π‘₯}, fix 0 < πœ– < β„Žπ‘› sin(πœ‹ βˆ’ 2πœ™π‘› ) and translate 𝑙 vertically by πœ–: 𝑙 πœ– = 𝑙 + (0, πœ–). Then, 𝑙 πœ– necessarily intersects all the 2𝑛 different sectors of the ball 𝐡 𝑃 (β„Žπ‘› ) inside 𝐢𝑃+ (πœ™, 0, β„Žπ‘› ), but only the right-most one, 𝐢𝑃+ (πœ™π‘› , πœŒπ‘›,2𝑛 βˆ’1 , β„Žπ‘› ), at its arc-like part of the boundary. In particular, 𝑙 πœ– has to intersect the sides of at least π‘˜ + 1 sectors that contain the paths described in (3.3.6) and therefore also intersects these paths. Hence, 𝑙 πœ– is one of our admissible lines that has at least π‘˜ + 1 intersections with 𝛾, a contradiction. Lemma 3.3.2 is proved. β–‘ Remarks. i) In the definition of β„Žπ‘› , three different parameters occur, π‘Ÿ 𝑛,𝑖 , 𝑑𝑛,𝑖 , and β„Žπ‘›,𝑖 . Without β„Žπ‘›,𝑖 , (3.3.5) automatically fails; 𝑑𝑛,𝑖 is to ensure Γ𝑃 (πœ™π‘› , πœŒπ‘›,𝑖 ) will always intersect the boundary of the corresponding cone and π‘Ÿ 𝑛,𝑖 forces this intersection to avoid the sides. ii) In the above construction we bisected the initial cone into 2, 4, 8 etc. smaller cones every time. However, any possible way to cut the cones would still work as soon as it eventually yields an infinite sequence. iii) The same proof can be applied to any cone within our admissible set of directions. 3.4 Higher dimensions Mattila in [7, Lemma 6.4] generalised Marstrand’s results from [5] and showed the following. Lemma 3.4.1 (Mattila). Let 𝐸 be an H 𝑠 measurable subset of R𝑛 with 0 < H 𝑠 (𝐸) < ∞. Then, dim(𝐸 ∩ (𝑉 + π‘₯)) β‰₯ 𝑠 + π‘š βˆ’ 𝑛 66 for almost all (π‘₯, 𝑉) ∈ 𝐸 Γ— 𝐺 (𝑛, π‘š). In particular, for a Borel set in, say, R2 we have: if any 2-dimensional plane in a positive measure of directions intersects this Borel set at a set of Hausdorff dimension at most 1, then the Hausdorff dimension of this Borel set is at most 2. Furthermore, if every line in the direction of some 2-dimensional cone intersects a Borel set (not merely the graph of some continuous function) at at most countably many points, then any 2- dimensional plane in a positive measure of directions intersects this Borel set by a set of Hausdorff dimension at most 1 (Marstrand) and then the Hausdorff dimension of this Borel set is at most 2 (Mattila). Of course, the same is also true in R𝑛 , that is, if a Borel set has countable intersection with a certain cone of lines, then its dimension does not exceed 𝑛 βˆ’ 1. Now, we restrict our attention to what happens with only 2 points of intersection in higher dimensions and we would like to generalize Proposition 3.1.1 to R𝑛 . Suppose we have a continuous function 𝑧 = 𝑓 (π‘₯, 𝑦), say, on a square in R2 , satisfying the property that any line in the direction of a certain open cone with axis along a vector v ∈ R3 intersects (3.4.1) the graph at at most two points. Then, we would want 𝑓 to obey the same rule. Namely we ask the following: Question. Is a continuous function on (βˆ’1, 1) 2 having property (3.4.1) locally Lipschitz? 3.5 Relationships with perturbation theory The problem we consider in this note grew from a question in perturbation theory of self-adjoint operators (see [4]). The question was to better understand the structure of Borel sets in R𝑛 that have a small intersection with a whole cone of lines. Marstrand’s and Mattila’s theorems in [5] and [7], respectively, give a lot of information about the exceptional set of finite-rank perturbations of a given self-adjoint operator. The exception happens when singular parts of unperturbed and 67 perturbed operators are not mutually singular. It is known that this is a rare event in the sense that its measure is zero among all finite-rank perturbations. The paper [4] proves a stronger claim: the dimension of a bad set of perturbations actually drops. Let us explain what was the thrust from [4] and why that paper naturally gives rise to the questions considered above: what is the structure of Borel sets in R𝑛 that have a small intersection with all the lines filling a whole cone and their parallel shifts? In [4], a family of finite rank (self-adjoint) perturbations, 𝐴𝛼 , of a self-adjoint (suppose bounded for simplicity) operator 𝐴 in a Hilbert space H is considered: 𝐴𝛼 := 𝐴 + π΅π›Όπ΅βˆ— parametrized by self-adjoint operators 𝛼 : C𝑑 β†’ C𝑑 (i.e., Hermitian matrices). The operator 𝐡 : C𝑑 β†’ H is a fixed injective and bounded operator. It is also assumed that range of 𝐡 is cyclic with respect to 𝐴. In the case when 𝑑 = 1 (rank-one perturbations), the Aronszajn-Donoghue theorem states that the singular parts of the spectral measures of 𝐴 and 𝐴𝛼 are always mutually singular. However, it is known that for 𝑑 > 1 the singular parts of the spectral measures of unperturbed and perturbed operators are not always mutually singular. Notice that the space of perturbations, that is the space 𝐻 (𝑑) of Hermitian (𝑑 Γ— 𝑑) matrices, has dimension 𝑑 2 . In [3], it was proved that, given a singular measure 𝜈, the scalar spectral measure πœ‡π›Ό of the perturbation 𝐴𝛼 is not singular with respect to 𝜈 for the set of 𝛼’s having zero Lebesgue measure in 𝐻 (𝑑). Such 𝛼’s are called exceptional, and this result shows that even though the set of exceptional 𝛼’s can be non-empty (for 𝑑 > 1), it is a thin set. But is it maybe thinner? In fact, the following result was proved in [3]. Fix 𝛼0 , 𝛼1 ∈ 𝐻 (𝑑) where 𝛼1 is in the cone of positive Hermitian matrices and consider 𝛼(𝑑) = 𝛼0 + 𝑑𝛼1 . Then, for any such 𝛼0 , 𝛼1 there are at most countably many 𝑑 ∈ R such that the 𝛼(𝑑) is exceptional. This extra information allowed the authors in [4] to prove that the Hausdorff dimension of exceptional perturbations is actually at most 𝑑 2 βˆ’ 1. The reader might have noticed an underlying geometric measure theory fact: a Borel set in R𝑛 (here 𝑛 = 𝑑 2 ) that has an at most countable intersection with a whole cone of lines and their parallel 68 shifts is, in fact, of dimension 𝑛 βˆ’ 1. Thus the dimension drop detected in Marstrand’s and Mattila’s theorems was instrumental for the drop in dimension for exceptional perturbations. It seems enticing to understand the structure of the sets that have even less than countable intersection with all parallel shifts of all lines from a fixed cone. Suppose the Borel set under investigation intersects only at at most two, or at most π‘˜ < ∞, points with these lines. What additional knowledge one can obtain about this set? This question motivated the work presented in the previous sections. 69 BIBLIOGRAPHY [1] V. Eiderman and M. Larsen. β€œA β€œrare” plane set with Hausdorff dimension 2”. In: (Apr. 18, 2019). arXiv: 1904.09034v2 [math.CA]. [2] H. Federer. Geometric measure theory. Springer, 1996, p. 676. ISBN: 3540606564. [3] C. Liaw and S. Treil. β€œMatrix measures and finite rank perturbations of self-adjoint operators”. In: Journal of Spectral Theory 10.4 (Oct. 2020), pp. 1173–1210. DOI: 10.4171/jst/324. [4] C. Liaw, S. Treil, and A. Volberg. β€œDimension of the Exceptional Set in the Aronszajn- Donoghue Theorem for Finite Rank Perturbations”. In: International Mathematics Research Notices 2022.5 (Nov. 2020), pp. 3297–3307. DOI: 10.1093/imrn/rnaa281. [5] J. M. Marstrand. β€œSome Fundamental Geometrical Properties of Plane Sets of Fractional Dimensions”. In: Proceedings of the London Mathematical Society s3-4.1 (1954), pp. 257– 302. DOI: 10.1112/plms/s3-4.1.257. [6] P. Mattila. Geometry of Sets and Measures in Euclidean Spaces. Cambridge University Press, Apr. 1995. DOI: 10.1017/cbo9780511623813. [7] P. Mattila. β€œHausdorff dimension, orthogonal projections and intersections with planes”. In: Annales Academiae Scientiarum Fennicae Series A I Mathematica 1 (1975), pp. 227–244. DOI: 10.5186/aasfm.1975.0110. 70