THESE l 5- 7‘4 003843 53 This is to certify that the dissertation entitled BELLMAN FUNCTION AND BMO presented by LEONID SLAVIN has been accepted towards fulfillment of the requirements for the Doctoral degree in Mathematics Major Professor’s Signature flu; . / IL 2 c9 c9 4 Date MSU is an Affirmative Action/Equal Opportunity Institution LIBRARY Michigan State University PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 6/01 cJClRC/DateDue.p65~p.15 BELLMAN FUNCTION AND BMO By Leonid Slavin A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Mathematics 2004 ABSTRACT BELLMAN FUNCTION AND BMO By Leonid Slavin The Bellman function method is applied to three different problems in harmonic anal- ysis. The first, introductory chapter outlines the specifics of the method, addresses its stochastic control origins, and gives a harmonic analysis perspective. A brief descrip- tion of the main function space under consideration, BM 0, is also provided. In the second chapter, the integral form of the John-Nirenberg inequality for BM 0 func- tions is examined, the corresponding Bellman function explicitly found, and the sharp constants in the inequality as well as the exact bounds on the region of its validity established. Two cases, those of the continuous and dyadic BM 0, are treated and the results differ significantly between the cases. In the third chapter, the dyadic ver- sion of the Chang-Wilson—Wollf theorem for functions whose s-function is uniformly bounded is proved using a Bellman-type argument. Furthermore, a local version of the theorem is established, whereby the s-function is assumed to be bounded on a measurable subset E of [0,1]. Consequently, the exponential summability of the second order over E is derived. In the fourth chapter, the famous question of H 1 — BM 0 duality is considered. Two cases, the continuous and dyadic ones, are treated and the same key lemma, based on a Bellman-type argument, is used in both to establish the embedding of BM 0 in the corresponding dual space. Moreover, in the dyadic case, an explicit estimate for the norm of the embedding is found. To the memory of my father iii ACKNOWLEDGMENTS I would like to express my deepest and sincerest gratitude to my advisor, Professor Alexander Volberg. He has taught me all the skills of the trade I have had the ability to grasp and he is responsible for my reaching this stage in my career. I am very grateful for his patience and generosity with his time, ideas, and expertise. I am also very happy to be working in this field and credit him solely for the opportunity to be on the cutting edge of research. Whatever mathematical accomplishments I dare count to my name would not have existed, were it not for his help and encouragement. The time I have spent as his apprentice will forever be appreciated. I would also like to thank my mentor and coauthor, Professor Vasily Vasyunin. He has opened my eyes to many a subtle mathematical phenomenon in the process of our joint work and taken time to educate me on not so subtle matters. His friendship, truly infinite patience, and preparedness to give the youngling an Opportunity after opportunity to establish himself have helped me immensely. I am very thankful to Professor Michael Frazier for the many useful discussions I have had the pleasure and advantage of having with him. He has helped me prepare for my talks and his insight has given me a broader perspective on certain areas of research. I am also very thankful for his help in preparation for my job interviews. I would like to express my appreciation to Professors Alexander Volberg, Clifford Weil, Vladimir Peller, Joel Shapiro, and Kirill Vaninsky for agreeing to be on my committee, to have their patience tried with delays, and their kindness tested with this imperfect text. Last but not least, I would like to thank my best friends Leonid Freidovich and Zeynep Altinsel for being there for me all these years. iv Contents List of Figures vi 1 Preliminaries 1 1.1 Introduction ................................ 1 1.2 Stochastic control Bellman function ................... 4 1.3 Harmonic analysis Bellman function ................... 7 1.4 The space BMO ............................. 9 1.4.1 John-Nirenberg inequality .................... 9 1.4.2 Hl — BMO duality ....................... 11 1.5 The Bellman equation for the John-Nirenberg inequality ....... 12 1.6 Bellman-function-type proofs ...................... 13 2 Sharp constants and bounds in the J ohn-N irenberg inequality 16 2.1 Introduction ................................ 16 2.2 Main results ................................ 18 2.3 The continuous case ........................... 20 2.3.1 The key lemmas .......................... 20 2.3.2 How to find the Bellman function ................ 26 2.3.3 How to find the extremal function ................ 29 2.4 The dyadic case .............................. 32 2.4.1 Preliminary considerations .................... 32 2.4.2 Detailed considerations ...................... 2.4.3 Stage 1 ............................... 2.4.4 Stage 2 ............................... 3 Bellman-function-type proof of a local Chang-Wilson—Wolfl' theorem and related results 3. 1 Introduction ................................ 3.2 Main results ................................ 3.3 Bellman function considerations ..................... 4 Bellman function and the H1 — BM 0 duality 4.1 Introduction ................................ 4.2 The dyadic case .............................. 4.3 The continuous case ........................... 4.3.1 Multi-dimensional setting .................... 4.4 The proof of the key lemma ....................... 4.5 A sample Bellman function ........................ Research Prospects Bibliography vi 37 40 52 54 54 56 61 64 64 65 70 74 75 77 78 81 List of Figures 2.1 2.2 2.3 2.4 4.1 4.2 The initial splitting: a- = 0+ = = .r’“. [Olr— I The stopping time: [.L“,£] is tangent to the parabola I2 = If + 62. The counterexample to the conjecture 53 = 1. ............. The worst case scenario: the largest. portion of [1", .z'+] is outside 95. The region II, (6”). ........................... The clecomposition Q; = U TQ1. ................. IEDJCJ 25 33 35 71 Chapter 1 Preliminaries 1. 1 Introduction This work is a compilation of three results united by a common theme. In all three, a relatively new harmonic analysis technique is used to establish the crucial segments of the argument: the Bellman function method. With its origins in stochastic control, its timeline in analysis both short and rich, and its scope seemingly unlimited, the method is as novel as it is powerful. Every successful application, therefore, serves to further the method’s legacy as well as explore and develop its subtle and intricate aspects that only come to light in a particular problem. That is why applying the Bellman function technique to problems that have been solved by other methods merits a researcher’s time. Due to its nature, this technique often streamlines the proofs and / or reveals new properties of the function spaces under consideration. What is more, it has been applied to obtain sharp and dimensionless results in existing estimates and inequalities, establish their regions of validity, and, sometimes, disprove intriguing conjectures. The other theme linking the results presented is, as the title suggests, the function space for whose elements they are valid. It is BM 0, the space of functions of bounded mean oscillation (see below for detailed description). The first and the last results establish certain properties of B M 0 functions explicitly, whereas the second one deals with a space that is better (smaller) than BM 0, and whose elements thus possess properties that are similar in nature to those of BM 0 functions, but stronger; that is to say it proves exponential integrability but of higher order than that of BM 0. Chapter 2 is on the John-Nirenberg inequality for BM 0, perhaps the most fun- damental property of BM 0 functions. It establishes sharp constants and bounds previously unknown. This is a joint result with Professor Vasily Vasyunin of St. Petersburg Department of Steklov Mathematical Institute and Mathematics Depart- ment of St. Petersburg State University. This work was inspired by that of Professor Alexander Volberg of Michigan State University, the mathematician who is, perhaps, the most responsible for establishing and promoting the Bellman function method as a tool in harmonic analysis. It is his presentation of the stochastic control Bell- man set-up for the John-Nirenberg inequality in [15] and an earlier version of [24] that, at about the same time but at vastly distant venues, caught our attention and spurred further research. We then independently found the Bellman function for the problem, after which we joined the efforts to solve the problem completely. Many useful discussions with A. Volberg have allowed us to use his insight and expertise. Two formulations, the continuous and dyadic, are considered. Notably, the results are significantly different in the two cases. Chapter 3 is on a joint result with A. Volberg. It gives a Bellman-function—type proof of a famous result of Chang, Wilson, and Wolff for functions whose square function is bounded and uses that result to establish that the square of such a function is exponentially summable. More precisely, for any cp from the unit ball in the corresponding space (with the usual factorization over constant functions and the s- . . 2 . . . . function as the semi-norm), e0“? 18 integrable. Furthermore, we give a local versmn of the theorem, whereas the s-function is assumed to be from L°°(E), as opposed to L°°([0, 1]), for a measurable subset E of [0, 1]. The result is a weak-form estimate, similar to that in the Chang, Wilson, Wolff’s original paper [2] and a uniform bound on the integral f3 emf”! for some a > 0. The article [2] does not give the local version, although it seems that the authors’ reasoning can be modified to make it work for an arbitrary E. Our proof is short, straightforward, and constructive — the usual advantages of the method used. The formulation of an extremal problem is discussed, although the “Bellman function” used in the proof is a supersolution of the optimization problem, i.e. a majorate of the “true” Bellman function. The Bellman approach to this problem has very intriguing implications, as discussed in the section on research prospects. Chapter 4 is on a joint result with A.Volberg. It examines the old question of H1 — BM 0 duality, first addressed by Fefferman in [5]. We use a Bellman-type reasoning to establish a key lemma, which then works to prove two duality results: the continuous one and its dyadic analog. No optimization problem is set up; thus we use a hands-on, heuristic approach to find a “Bellman function,” whose properties are dictated by the differential estimates in the lemma. As is often the case with this technique, we are able to obtain an explicit embedding constant in the dyadic case, although we do not discuss whether it is sharp. Moreover, the same proof seems to work in a multidimensional setting and, with the types of H1 and BM 0 norms used, the constants of embedding are dimensionless. One common feature, the hallmark of a Bellman function or Bellman-function-type proof, that is present in all three results is the unwrapping of a certain integral sum, undoubtedly the main connection between stochastic control and harmonic analysis. We expand on this connection in the next two sections. 1.2 Stochastic control Bellman function We will formally derive the Bellman equation for a controlled stochastic process fol- lowing the exposition in [15, 24, 10]. Let :r‘ be an d-dimensional stochastic process, satisfying the stochastic differential equation t t 33‘ = a: +/ o(a",:r’) dw’ +/ b(a’,:r’) ds. (1.1) o 0 Here t is the time, w‘ is a d1 —dimensional Wiener process, o(a, y) is a d x d1 matrix, and b is a d-dimensional vector. Different choices of the control, a’, which is a d-dimensional stochastic process, give us different trajectories, i.e. different solutions of (1.1). The derivation we give will be entirely formal; thus we do not address questions of existence or uniqueness of solutions. Equation (1.1) is a part of an optimal control problem. Namely, given a profit function f“, on the trajectory x‘, for the interval [t,t + At], the profit is fat(a:‘)At + 0(At). Therefore, on the whole trajectory we earn / fa‘ (17‘) dt. 0 We want to choose the control a = {as} to maximize the average profit 210(33): IE / fa‘(xt)dt+;1i‘ffia(r(xt)), (1.2) 0 —ioo for the process starting at :5. Here F 2 0 is the bonus function — one gets it when one retires. The Bellman function for the process (1.1) is the optimal average gain, v(:r:) = sup v"(x), (1.3) 06.4 where A is the set of admissible controls; 1) satisfies the well-known Bellman (par- tial) differential equation, which is based on two ingredients: Bellman’s principle and Ito’s formula. Bellman’s principle states that v(-:c) —— suplE [A( f"‘3 x)ds + v(:r‘) . (1.4) 06A To explain it, we fix t > 0 and consider an individual trajectory. The profit for the [SM Suppose the trajectory has reached the point y at the moment t. The maximal interval [0, t] is given by average profit we can make starting at t and at the point y is precisely v(y). Indeed, since the increments of w‘ for s 2 t do not depend on wT,T < t, and equation (1.1) is time-invariant, there is no difference between starting at time 0 or at time t. Applying the full probability formula to take into account all possible endpoints y = x‘, we obtain (1.4). Let us now explain the version of Ito’s formula that we need. Fix a moment of time s and a small increment As. We want to estimate the difference v(:r’+A") — v(3:’). Let Aw’ = w’+A‘ -— w’. Assuming enough smoothness, we can use Taylor’s formula. Among others, we will have the term d1 d okj(a’, x”) )Awf-l- +2; a—(f’ )bk( (0 ,x’)As. _1 k: HQJIQD 623;. J: After taking the expectation, the first term will vanish, since each Aw}: is indepen- dent of :c’ and has zero mean. The second term can be rewritten as IE (£?s(:c’)v) (x’)As, with the first order differential operator .6? given by “ a 1 : gbk(a, S) a. The next term in the Taylor formula will be 2 '1': 6 v 2 UjkAwZ + M018, $3)AS Z UikAwi + b.-(a’, x")As . 2 k,j 611.1323]; 1: k Averaging over probability, taking into account the-fact that lEszAwfn = As if k = m and 0 otherwise, and omitting the terms with (As)2, we get IE (£38 (x’)v) ($’)As, where the second-order differential operator CS is given by d . a2 1 “1 £3 = Z av 33.3.; arm) = ,- ganammm). i,j=l =1 Gathering all the terms together and omitting the ones with powers of As greater than one, we obtain lE('U(iL‘t)) = v(:1:) + EA £08053) v(:c’) d3, (1.5) where L“ = Li" + £3. That is the application of Ito’s formula we need. Putting (1.5) into Bellman’s principle (1.4), we get 0 = sup [1: fa’w) + [at castes) v(:1:") d3] . 06A Dividing by t an taking the limit as t —» 0 (assuming it is justified), we get Bellman’s partial differential equation supplemented by the obstacle condition 1} 2 F :13 [£°‘(x)v(:r) + f°’($)] = 0, x E Q v(:1:) 2 F(m), a: E 9. 1.3 Harmonic analysis Bellman function The scope of applications of the Bellman function method has been exceptionally broad. While the method seems well suited for proving weighted norm inequalities (see the early 1995 version of paper [16], the ground-breaking proof of the matrix Hunt-Muckenhoupt-Wheeden theorem in [13], or a more recent result in [20]; alter- natively, see [12] for a two-weight negative result), it has found use in areas far from its origins. The (much needed) anthology of the existing Bellman function results is far beyond the scope of this chapter. An incomplete list of references, besides those already named, includes [7, 14, 17, 18, 19] and, perhaps the closest in spirit to the current work, [23]. This panoply of results seemed to necessitate the development of a uniform founda- tion. Such a foundation has been found in the very field from which the notion of a Bellman function first arose — that of stochastic control. The work on developing the stochastic control framework for harmonic analysis problems was begun in [15] and continued in [24]. It is in paper [15] that the famous result of Burkholder for martingale transforms [3] was interpreted from the stochastic perspective. We refer the reader to these articles for a thorough exposition and many interesting examples. Here, we would like to indicate the general principles of the stochastic control frame work. We will use them to develop the Bellman equation for the John-Nirenberg inequality after we introduce the space BM 0 in the next section. The problems that can be treated using the Bellman function from stochastic control and thus the machinery of the previous section are always dyadic. It is often possible to pass from a dyadic problem to the corresponding problem with analytic or harmonic function using some kind of Green’s formula. (That is exactly what is done in Chapter 4 in the continuous case.) It is, therefore, very interesting and unexpected that in the John-Nirenberg setting of Chapter 2, we, motivated by the stochastic control Bellman function formally derived for the inequality, first solve the continuous problem and only then use those results to return to the dyadic case. First of all, we make a choice of variables so that the function space under con- sideration is mapped onto a Euclidean domain. Namely, assume that we want to prove a certain inequality for all functions from the 6-ball, F5, of a space F. Then, with every pair (tp, J), where cp E F; and J is a dyadic interval, we associate a d-dimensional vector :1: = (1151,15,. . . ,xd) in a domain 05 whose geometry is de- termined by the space F. Very often, the coordinates x,- will have a martingale structure; we often see 1:10p, J) = (cp)J. (We make a choice like this when dealing with the John-Nirenberg inequality below.) The vector x then is the state vector of our system, i.e. a solution of (1.1). Thus, given a function (,0 6 F5, we know the state of the system on every dyadic level (for every generation of dyadic intervals), that is to say, at every moment t. Naturally, the time is now discrete and is equivalent to the order of the current generation. Instead of (1.1), we now have a discrete model 10”“ = :12” + 0(0", :13”)A"w + b(a”, :13"). (1.7) Most of the time, we have a = diag{a1,a2, . . . ,ad}. The exact form of the matrix o and the vector b is determined by the difference 23"“ — x”. The Wiener process is very often just a series of coin tosses: we either move to the left or the right half of the interval I. Thus the equation (1.7) often becomes xvi-+1: 1'" + anén + b(an,$n), where 6" is either 1 or —1. If we are maximizing a sum of the form that sum can be interpreted as the expectation of the cumulative gain f0°° f as (:r’) ds, allowing us to choose the profit function correctly. On the other hand, if we have a term of the form (g(1| > M) S cle—C2AIIMIBMO(I)- (1.10) Finding the sharp constants in the inequality (1.10) is a natural goal. In an important development, Korenovskii found the sharp constant 02 = 2 / e in [9]. A remarkable consequence of the John-Nirenberg inequality is that every BM 0 func- tion (,0 is in LP(I),l£p1|2dt) . (1.12) JcI IJI .1 We can — and this is the main reason for using p = 2 — rewrite (1.12) as 1/2 2 ”WIIBMO(1),2 = (93211) “902% - Will) . (1-13) 10 Phrthermore, we have the corresponding weak-form John-Nirenberg inequality m ({x e I : we) — «am > A» s ale-:2*/”r"BMOi2. (1.14) with constants c1 and C2 different from those in ( 1.10). One can rewrite (1.14) in the integral form, as follows. There exists so > 0 such that for every s < so and every 90 E BMO(I) 31101] that "‘pllBMO(I),2 _<. 5, we have or). s C(aeW (1.15) for some constant C (s) The inequality (1.15) is the reverse Jensen inequality. Find- ing the sharp value for so and the sharp expression for C(s) is highly desirable. Chapter 2 describes how it is done in both continuous and dyadic settings. 1.4.2 Hl—BMO duality A major reason BM 0 gained prominence is paper [5] in which Fefferman established it as dual to H1. This is the result we take up in Chapter 4. Both spaces are considered on the unit circle T. Again, we treat two cases, the continuous and dyadic. In the dyadic case, we use the following BM 0 norm ||1_)2lll. (1.16) JeDlJl 19 where D is the dyadic lattice rooted in T and 1., 1+ are the left and right halves of the dyadic arc I, correspondingly. Here, the advantage of the L2-formulation for BM 0 is evident, since fi 2 (M1, — <90)1_)2 III = 4 («02), — M3)- 11 (See Chapter 4 for a detailed explanation.) Thus we are using virtually the same BMO norm as in (1.12). In the continuous case, the equivalence of the norm (1.12) and 1 I 2 __ llcpll— sup m [01 moi (1 |€I)dA(€), (1.17) are ICT where 90(2) is the harmonic extension of (,0 into the disk and Q; is the Carleson square based on the arc I, can be found, for instance, in [21]. 1.5 The Bellman equation for the J ohn—N irenberg inequality To preserve history, we will develop the equation in the way we first encountered it, although the attentive reader will notice that we use a slightly different choice of variables in Chapter 2. According to the previous section, we have the follow- ing underlying dyadic problem: Given that ((90 — (90) J)2)J S 62, for every dyadic subinterval J of I , prove that (e‘P), S C(6)e("°>1. For every J, let $1 = (90)] , 3:2 = ((cp — (1= (e): = $1, (902); = $2}, ll‘PllBMO(I)SE does not depend on the interval I . This allows for effective modeling using the stochastic control technique. In the case of the local Chang-Wilson-Wolff theorem, given a measurable subset E of [0, l], the corresponding Bellman function Bf(x) = sup {/ ems) ds : (90)“),1] = 331,561] = 1:2} E ‘PEFQ is very E-specific. This is, of course, due to the fact that E does not scale as well as [0,1] itself does, i.e. dividing an interval in half, we will not necessarily divide that interval’s portion of E in half. Of equal importance is the fact that the variable S; defined in the chapter does not have a martingale character to it; thus it does not scale correctly either. A reasonable question then is whether one can obtain a more manageable Bellman function set-up in the case E = [0,1] (the original Chang-Wilson—Wolff case), if, instead of 5;, one uses the s-function itself as a variable. This remains subject of further investigation. The second category of Bellman-function-type arguments are the ones where the 14 extremal problem does not even enter the consideration. Chapter 4 deals with a result of this kind, the H1 — B M 0 duality. Although it is possible to think of an underlying optimization problem, the Bellman function involved appears simply as a means to carry out a clever segment of the proof. It is a very utilitarian approach — one attempts to unwrap the corresponding sum over dyadic intervals with a nonexistent (as of yet) function B and imposes such differential and quantitative properties on the function as are required so that the estimates work out the right way, the constants are not too big, etc. After that, one attempts to find the function satisfying the prOperties so determined and, perhaps, Optimize its parameters or tweak it otherwise. We daresay this is the way many explicit Bellman functions as well as their majorates are constructed. Admittedly, this is how the function in Chapter 3 was constructed (The function in Chapter 2, however, stands out as a pure product of stochastic control formalism.) What, then, distinguishes a Bellman-type proof from any other? Precisely the un- wrapping of an integral/dyadic sum, mentioned in the preceding paragraph. The three key lemmas of the three chapters that follow this introduction, the reader will notice, accomplish just that. We are now in a position to present the main results of this thesis. It is our hope that the reader will appreciate the unifying ideas behind the proofs and the flexibility of the method in dealing with subtleties of each particular case. Some prospects for future research are detailed in the last section. 15 Chapter 2 Sharp constants and bounds in the J ohn—N irenb erg inequality 2. 1 Introduction For any interval I C R and a function cp 6 DH), we denote by (<0) I the average of cp over I, (90), = l—i|f190(t)dt' We define the space BMO(I) as BMO(I) = {90 E L2(I) :/| 0 such that for every 0 S s < so there is C(s) > 0 such that for any function <0 6 BM 05(1 ), (er), s C(s)e<‘p)1. (2.4) We are interested in determining the sharp bound so and the exact expression for C (s) We will do that in the case of both, the conventional (continuous) BM 0 and the dyadic BM 0. In accordance with the ideology of the Bellman function method, with every ball BMO,-(I) (BM 03(1 )) and the set of all subintervals J C I we associate the domain 05 = {x = (2:1,:r2) : :31 E Rm? 3 :52 g at? + 82} as follows (so, J)*—* (<¢>,. ($51) (25) This map is well-defined because (cp)3 S (902)] (Cauchy inequality) and «p E BM 05(1 ) (BM 021(1 )) On 95, we define the following Bellman functions B.(x)= sup {,= (¢)1=w1.(902),=$2}, (2.6) ‘PEBMOsUI 17 Bg(.r) = sup {(e‘p), : (<0), = x1,((02)1 = 51:2}. (2.7) wEBMO‘gU) Observe that these functions do not depend on I . Finding them explicitly would provide us with the complete solution of the John-Nirenberg problem. That is exactly what we are able to accomplish. Remarkably and unlike many other Bellman function problems, the results for the continuous and dyadic cases differ significantly. Which is more, we first solve the continuous problem and only then use the corresponding Bellman function family to solve the dyadic case. 2.2 Main results Theorem 1. Let so = 1. For every 0 g s < so, let C(s) = . (2.8) Then, for any <0 6 BMOE(I), (e‘p), S C(s)e(‘p)1. (2.9) Moreover, so and C(s) are sharp. Theorem 2. Let s3 = x/2—log 2. For every 0 S s < s3, let 01(5) = 0(6), (2.10) where C(6) is defined by (2.8) and 6 = 6(5) is the unique solution of the equation (1 — «52 — €2)e .2-.2 [2 — 8W5] — (1 — (net-W5 = 0. (2.11) 18 Then, for any <0 6 BMO,?(I), (er), s C“(s)e("°>1. (2.12) Moreover, s3 and C“(s) are sharp. Theorems 1 and 2 are immediate consequences of the following results for the Bellman functions (2.6) and (2.7). Let _ 1/ 2 5- 85(13): 1 61 1.31:1 x2 exp (2:1 + V62 + :13? - x2 — 6). (2.13) Theorem 3. If 0 S s < 1, then BE(.v) = Bs(:r); (2.14) if s 2 1, then e‘cl if :02 = 113% +00 ifxg > in? Theorem 4. If 0 S s < x/210g 2, then I323(3) = Ba(.)(:r); (2-15) if s 2 filog 2, then eI1 if 2:2 = at? +00 ifxo > :10? Indeed, since the function t H (1 — t)e‘ is decreasing for positive t, B assumes its 19 maximum when :32 = a3? + 52, i.e. —€ e3E1 , e < B(x)__ 1-s giving (2.9) and (2.12) with the sharp constant (2.8). 2.3 The continuous case 2.3.1 The key lemmas We first consider the continuous case and prove Theorem 3. One can observe that the proof does not work in the dyadic case. We split the proof of the identity (2.14) into two parts. Lemma 1. For every 1: E 05, B5(x) 2 85(3). Proof. We prove this inequality by explicitly finding a function (0 for every point x E 9. such that (<¢)1.< 0. Take I = [0,1], a 6 (0,1], b 6 IR, 7 E IR\{O}. Let ylog%+b for0_<_t§a ‘pa.bn(t) “—— b foragtg 1. 20 Direct calculation shows that <00;w E BM 0|,|(I ), ( 32%. Then wecan set a: 1— fl2+x§—$2 and b=x1—7a, which yields (moon), = x1,(<0§,bfl)1 = 432. Now, if we put 7 = s 2 1, we get B£(:c) = 00. For '7 = s 6 (0,1), we get 1—\/s2+:c"1’—;r2 l—s BE(.r) _>_ (yam)! = exp (271+ s2 + :13? — 1:2 — s) = BE(:r). Cl Lemma 2. For every x 6 fig, BE(:c) S Bs(:r) (2.16) Proof. To establish (2.16), we first prove that B€(:r) _<_ 851(3), Vsl > s,Va: E 95, and take the limit as sl —+ s. (Observe that BE is continuous in s from above.) We need the following two results: Lemma 3. The function 8.; is concave in $25, i.e. Bs(oz_:r' + a+x+) 2 a_B€(:r’) + a+B€(;r+) (2.17) for any straight-line segment with the endpoints 23* that lies entirely in DE and an I! pair of nonnegative numbers at such that a- + 0+ = 1. Lemma 4. Fix s. Take any s1 > 5. Then for every interval I and every <0 6 BMOEU), there exists such a splitting I = I_ U 1+ that the whole straight—line 21 segment with the endpoints xi = ((<0)Ii,(<02)1i) is inside 95,. Moreover, the splitting parameter 0+ = [1+] / II [ can be chosen uniformly {with respect to <0 and I ) separated from 0 and 1. Assuming these lemmas for the moment, take <0 6 BM 05(1 ) Take any 51 > s. Observe that <0 6 BM 06(J ) for any subinterval J of I. Split I according to the rule from Lemma 4. Let 10'0 = I, 1110 = I_, 11'1 = 1+. Now split I. and 1+ according to the rule of Lemma 4 and continue this splitting. By I "I'" we denote the intervals of the n-th generation, as follows: I ”'2’“ = If-l’k and I "12““ = If”, so the second index runs from O to 2" - 1. The corresponding points given by (2.5) are indexed by the same pair of indices. Also, let awn = |I"I'"|/|I|. Since Lemma 4 provides for the value of (1+ uniformly separated from O and 1 on every step, we have max {|I”’k|} -—> 0 as n —> oo. k=0,1,...,2"—1 Then, using Lemma 3 repeatedly, we have II‘IOI II“! 0, , 1, 361(5): 0) Z EEO—[851(x10)+I-jO—,O-IBEI(SB 1) [Il’ol [[2,0] 2,0 l__Il’0l l___’_121lB 2,1 2 0,0 1,0 Bel(x )+— 0,0 1,0 Bel“ ) II III | II III I ”I’ll [122' 2,2 [Il’ll [12’3" 2,3 +I1‘09II1—IIIIB‘1“: H IIIIII I1—1»IIB€1“‘ ’ (2'18) [I2’Ol 2,0 [12,1] 2,1 [12'2l 2,2 _l___12’3lB 23 = l10,0|B€1($ )+ l10,0|B€1($ )+ [10,0lB€1($ )+— [[0,0] 851(55 ) 2"—1 2"-—1 1 2 II'”"IB — —- /B”""’ds, ,,.——., Z; III , where we have used the fact that 851(3) 2 erl and <0,, is the step function, 22 <0,,(s) = 2:?” for s E 1”“. Since <0” converges almost everywhere to 90, Fatou’s Lemma yields 1 1 I B51(:r) 2 —lim sup/e‘PMsM’ 2 — lim sup elm“) ds = — /e"’(3) ds = (6”)1. III I [I] I III 1 Taking supremum over all <0 with ((0), = x?” = x1 and (<02), = x310 = :02, we obtain the inequality Bil (x) Z B€($)7 thus proving the lemma. E] To finish the proof of Theorem 3, we need to prove Lemmas 3 and 4. In the section “How to find the Bellman function” below, the function BE is explicitly constructed to be concave in $25, which is what Lemma 3 states. Thus Theorem 3 is contingent on Lemma 4. Proof of Lemma 4. We fix an interval I and a function <0 6 BM 05(1 ) We now explicitly construct an algorithm to find the splitting I = I. U 1., i.e. choose the splitting parameters ai = |Ii|/|I|. As before, 501* = (<0) 2:; = (<02) 11' Also, 15:, put at? = (<0) I and $3 = (<02),. Lastly, by [s,t] we will denote the straight-line segment connecting two points 3 and t in the plane. 1 First, we take a- = (1+ = . If the whole segment [x‘,x+] is in 05,, we fix this [0| splitting. Assuming it is not the case, there exists a point a: on this segment with 2:2 — 2:? > sf. Observe that only one of the segments [:r",:1:°] and [23+, :00] contains such points. Call the corresponding endpoint ( :13“ or 23+ ) 6. Its position is completely defined by the choice of 0+. Define the function 0 by: p(a+) = maxxelmoflxo —- 10%}. By assumption, 0 G) > sf. We will now change 01+ 0 so that 6 approaches a: , i.e. we will increase a+ if .5 = x+ and deCrease it if E = :r‘. We stop when p(a+) = sf and fix that splitting. It remains to check that 23 + Figure 2.1: The initial splitting: a- = (1+ = , 5 = 2: . NIH such a moment occurs at all and that the corresponding 0+ is separated from 0 and 1. Without loss of generality, assume that 5 = 22". Let I = [a, b]. Since <0 6 L2(I), the functions 5101+) = 31: fbb—lllm <0(w)dw and {2(a+) = a—1+- fbb-llla+ <02(w)dw are continuous on the interval (0,1] and 5 (1) = 23° . Therefore, 0 is continuous on (0,1]. Since 0(%) > sf and 0(1) 3 52 < sf (recall, 22° E (25 ), we conclude that there is a point m, E [%, 1] with 0(a+) = sf. Having just proved that the desired point exists, we need to check that the cor- responding a+ is not too close to 0 or 1. If 5 = 2”“, we have 0+ > and l 2 {1 — 2:? = 2:? — 2:? = a.(2:1+ — 20f). Analogously, if 6 = 23‘, we have a- > % and £1— 2:? = 2:1- — 23(1): a+(:rl_ - 231+). Thus [£1 — $9] = min{ai}]2:1‘ — 25?]. For the stopping value of 0+, the straight line through the points 27,2:+ and 2:0 is tangent to the parabola 2:2 = 20? + sf at some point y. The equation of this line 24 Figure 2.2: The stopping time: [2:‘, 5] is tangent to the parabola 232 = 2:? + 62. is, therefore, 22 = 2231311 — y? + sf. The line intersects the graph of 222 = 22% + s2 at 2:(s)i =(y1 :1: Vs? — s2,y2 :l: 2y1(/s¥ — s2) and the graph of :02 = 2:? at the points the points 917(0)’c = (w i 61.112 i 21/161)- We then have lw(€)‘.x(€)+l C [300.6] C [$193+] C lx(0)‘,$(0)+l 25 and, therefore, 2 6i - 52 = |£I(€)f - $(€)f| S I23? - {II = min{ai}|$l - III S min{ai}|x(0)'f — 23(0)l"| = min{ai}2sl, which implies As promised, this estimate does not depend on <0 or I. E] 2.3.2 How to find the Bellman function We first observe that the Bellman function B must be of the form BE(2:) = exp {2:1 + w€(2:2 — 20%)} (2.19) for some positive function w on [0, 52] such that w€(0) = 0. Indeed, fix an interval I. Then <0 6 BMO,;(I) if and only if <0 + c E BMO,-(I), where c is an arbitrary constant. Let <0 = <0 + c. We have (all averages are over I ) (<0) = (<0) + c, (<02) = (902) + 2c (<0) + (:2, and (e‘f’) = e‘ (e‘P) . Then sup {(.¢).(,.)=.,,(,.2)=.,}-.c sup {=<¢>=x., 0. Let the function <0 be defined on I = (0,1] by 90l(2-(k+l)’2—k] = (k — 1)a, k = 0,1,. . . , (2.30) with the constant a to be determined later. We have the following picture for <0 5a“ 4a“ 3a“ 2a,_ ‘7 Q v—t '0 cal“ (Jolt--W .c-II-I NIH H r Figure 2.3: The counterexample to the conjecture s3 = 1. We now calculate the BM 0" norm of <0 and choose a so that ||<0||BMOd = s. The only dyadic intervals on which <0 is not constant and, hence, (<0?) - (<0)2 75 0 are 33 the ones with 0 as their left endpoint. Let 1,. = (0, 2‘"] . Then ”2“ 0° ha a 1 "-2 = 2" d = 2" _ = -2" _ = (80)!“ [0 ‘P(3) 3 k§12k+2 4 (2) n an and V2" 0° We2 a2 l "—2 ($02)“ = 2"] 902(3) ds = 2" 2: 5,23 = 42“ (2) (n2 + 2) = a2(n2 + 2), 0 k=n—-l where we have used the identities oo 1 k 1 N—2 oo 1 k 1 N—2 2 .(-) = t) I Z .2 <-) = (I (N2 + I k=N-1 2 2 k=N—1 2 2 Then 2 _ 2 _ 2 llfpllgMod _' styldgcl {J (99>J} = sup {(<02),n — (<0)?n} = sup {a2(n2 + 2) - a2n2} = 2a2. Setting ||<0||BMOd = s, we get a = s/\/2. Now, <0 —°° eke—m1 eak 1_ 2k+2—ZZ§ ' The latter sum converges if and only if e“ < 2, i.e. a < log 2. In terms of 53 from Theorem 2, we obtain the crucial estimate 53 S x/210g 2. (2.31) Informally, we have just shown that the space BM 0‘1 is substantially worse than BM 0. However, having developed the Bellman function family (2.13) for the con- tinuous BM 0, we would like to look for the dyadic Bellman function within that 34 family. From the example (2.30) above, it is clear that, should the dyadic Bellman function be in fact found in that family, we would have to have B: = 36(5) (2.32) for some 6(5) > 5. One straightforward approach would be to choose 6(5) large enough so that any straight-line segment [1", x+] with 23-, 17+ 6 05 would fit entirely inside 95(5). Then we would be able to use the chain of inequalities in Lemma 2 without the help of Lemma 4. Let us investigate how large the 6(5) so chosen would be with regard to 5. Consider the situation when the segment [x‘, 3*] “sticks out” the most. This hap- pens when the middle point x0 = 11,-(15’ + 3:“) as well as one of the endpoints, say 23+, are on the top boundary, m2 = 1:? + 52, and the other endpoint, 2:", is on the bottom boundary, x2 = mg. 2:2 2 x? + 62(5) $2=x§+52 Figure 2.4: The worst case scenario: the largest portion of [1:", 1+] is outside (25. We have :1:(2’=(:r(1))2 + 52, 2:; =(:171+)2 + 52, 11:3 = ($92 + 52, 35 and 0 xf+srf 0 LUZ-+33; :cl=—, x2=———. 2 2 Eliminating :r‘ and parameterizing everything by $9, we get 0 xf=$1+ 72'“ The equation of the line segment [2:0, x+] is then 332: (T;_—2-+2:r(1’) x1—::_-2—x1"—(a:(1))2+52, $93$13x?+% and the distance between this segment and the top boundary curve is E 5 5 dCL‘ —-$2+( +2x°):c——:r°— $02, x°<$ 1 = glaze )+ 513%), which completes the proof. Cl Observe that in the continuous case BM05(I) # BMO,;(L) U BMOE(I+) U {cp : (902), - (90)? S 52} , since there are other intervals to consider, those with the left endpoint in I- and the right one, in 1+. 37 We have just proved that B? is concave in $25. Phrthermore, the reasoning of (2.19) still works and we conclude that Bg(a:) = exp {1:1 + 10({132 — 1%)} (2.37) for a nonnegative function U) such that w(O) = 0. What is more, we expect the corresponding matrix —-d2Bg (assuming sufficient smoothness) to be degenerate, in order for the supremum to be attained for an extremal function. But we have already described all functions with these properties. They are the functions B; from (2.13) with 6 Z 5. This, somewhat heuristic, argument supports our believe that the dyadic Bellman function is a member of that family. In our desire to use Lemma 2, we have been trying to ensure that the segment [:r‘ , 33+] lies inside the domain of concavity of a certain function B, so that we can conclude that 1 1 and proceed with the unwrapping of the integral sum (2.18). Now, we try to enforce the condition (2.38) directly instead. Since we are searching for 6(6) such that B: = 85(6), we attempt to solve the extremal problem (5(5) = min {62 85(3)) 2 185017) + l35(174'), s<6/\) = 2(1- “Fifi-236 V 62—52 —(1 — a_)e9+°‘- — (1 — a+)e‘9+a+ — Ma": + a: — 2(62 — 52) — 292). From Vhl = O we get a_e6+°‘- — 2a_A = 0 a+e'9+°‘+ — 204A = O ——(1 — a-)e9+°- + (1 — a+)e-0+°‘+ + 40)\ = 0 (13+ a1 = 2(62 — 52) + 262. Therefore, e9+°- = e“9+"+ = 2). and (1+ = a_ + 26. Plugging this into the last 41 equation, we get (a- +6)2 = 62—52, a- = —6+\/62 —e2, a+'= 0+\/62 —e'~’, where the plus sign is chosen in the square root to ensure that each a is positive. Lastly, we obtain at any possible extremum point, h1 = 2(1— «62 — €2)eV62“€2 — 2(1— «62 — e2)e 52-52 = 0. 2. Face (1+ = 6 Let h2(a,a-,0,x\) = H(a,a_,6,0,)\) = 2(1— a)e°‘ — (1 — a_)e9+°‘- — (1 — 6)e“9+‘S — Mari + 62 — 2a2 — 262). From th = 0 we get 6“ = e9+°'- = 2A —(1 — a_)eg+°" +(1— 6)e"“HHS + 40A = 0 a2_ + 62 = 2012 + 262, from which we get a = 0 + a_, (1 -— (De—0+6 = (1 — a- — 26)e9+°‘-, and 62 = (a- +26)2. If a- = -—20+6, then we get (1 — 6)e" = (1 +6)e‘6, which only has the trivial solution 6 = 0. So a- = 20 + 6, a = —0 + 6 and, at any possible extremum point, h2 = 2(1+ 0 — 6)e_9+5 — (1+ 29 — 6)e‘9+6 — (1 — 6)e_9+6 = 0. 42 3. Face 0+ = V62 — 52 Let h3(a,a_,6, A) = H(a,a__, V62 _ 82,0,A) = 2(1_ (1)60 _(1_ a_)ea+a_ ‘ (1 - V52 - 826“” ‘52"2 - Mai + (62 — 52) — 2a2 — 292). We observe that this case is identical with the previous one if V62 — e2 is substituted for 6. Therefore, at any possible extremum point, h3 = O. Edges We have a total of 12 edges: (1)a=\/62———€7,a+—5 (2)0: 62—52, 01+: 62—52 (3)a=\/67-—52,a_=6 (4)0: 62-52,a_= 62—52 (5)a=6,a+—-6 (6)a=6,a+=\/67_—? (7)a=6,a_=6 (8)a=6,a_=m (9)a+=6, (1-26 (10)a+=6,a_= 62—52 (11)a+= 62—52,a_=6 ) Edges (5) through (8) have a = 6, which implies a- = (1+ = 6, t9 = O. The pairs (1)-(3), (2)-(4), and (10)—(11) are symmetric. On edge (12), using the fact that ag+afi =2a2+202, we get a2 =52—82—92, giving 0:0, 0:0- 201+ and f = 0. This leaves us with the following four (renumbered) edges to consider (out of 43 the symmetric pairs we chose the edges sharing a vertex): (1) a+=6, a_=6 (2) a+=\/'6—2-——e2, a_=6 (3) a=\/6"’——8_2, 01.26 (4) a=m, a+=m 1. Edge a+ = 6, a- = 6 Since a“: + 0?, = 20:2 + 202, we have a = V62 — 62, —e S 0 S e. The function to minimize is, therefore, 17(0) = 2(1 — J52 — gays/62*? — (1 — 6)e6(ee + e-O). We only need to show that F (00) 2 0 at any possible extremum point 90 E (-—e, e). We will, however, show more; namely, that F(6) Z 0, V0 6 {—6, 6], V6 E [0,1]. Let u = 6 + 0, v =6 — 0. Then, slightly abusing notation, we have F(u,v)=2(1—\/uv)e‘/fi— (1_u~2l-v) (eu-l-e”), u+v=26, 0Su,vS26. We will accomplish our goal if we show that F (u, v) 2 0, 0 S u S 2, OSvS2—u. 44 Every point (11,12) of minimum inside the region in the picture would have to satisfy the equations 1 Fuz—vem+§(e“+e”)— 1_u-;-v e“=0 1 Fv=—ue‘/fi+-2-(e“+e”)- 1_u-2l-v e”=0. Adding the two equations, we get e + e = em. (2.44) Subtracting the second equation from the first and using (2.44), we get e”(1 — v) = e“(1 — 11.), which implies (since u, v 2 O ) that u = v. The value of F on any possible extremum point in the interior is then F (u, u) = 0. On the boundary 1) = 0, 0 S u S 2, we have F(u,0) = 2 — (1 — g) (e“ +1). The equation [F(u,0)]' = 0 gives e"(1 — u) = 1. Since 11 Z 0, we conclude that u = 0. Then F(0, 0) = 0. The case u = O is identical. If u + v = 2, we have F(u,v) = F(u,2 — u) = 2(1 — \/u(2 — u))eV“(2'“). Since 0 S u(2 — u) = l — (u - 1)2 S 1, and the function 3 H e’(1 - s) is nonnegative for s 6 [0,1], we conclude that F 2 0 on this piece of the boundary. This completes the consideration of this edge. 2. Edge a+ = V62 — 52, a- = 6 Since oz?’_+oz_2+ = 2024-202, we have a: 62———l92,—E SOS %. The 45 function to minimize is, therefore, F(6)=2(1— \/62——2——02)e eV 62-27-92 ——(1 6).: “9— —1( —\/62——?)em-9. Assume that F has a local minimum for some 60 E (-—}v, 3‘5) . Then F’ (60) = O, 2 i.e. 62 5—2 —92 290a ’ 2 ° -— (1 — 6)e"+"0 + (1 — 62 — e2) eV62'52’90 = 0. Expressing (1 -— V62 — 52) 6" 62—52490 from (2.45), we get (2.45) F(60) = 2.990 [(1- (M62 — é — 03 — 60)) eV62—%_93_90 — (1 — (5)125] . (2.46) Observe that (1 - :1:)e3c - (1 — y)ey 2 0 if y 2 0 and —y S a: S y. Together with (2.46), this consideration implies that if 2 —6g\/62—52——ag—aoga, then F (60) Z 0. We now solve the inequality (2.47). First, since |60| S €/\/2 S 6, we have that ((52-5493 —6> —,6V06[—-£- 5—] 2 O 0 fi,\/§ ' Secondly, solving the (right-hand) inequality (2.47), we obtain that 52 2 62—3—90— —60S6 if and only if _ 2__2 __ 2_2 602 6+\/26 e or 003 6 6 e. 46 (2.47) _ _ 2- 2 . . . . . We must check whether —% S $42. This inequality is equivalent to V62—52S \/2€—6. Since €> >2—‘Q6 62 76, we can square the last inequality, getting as a result 2f 5>—6 again, which is true by the formulation of our extremal problem (2.42)—(2.43). Finally, we conclude that (2.47) holds if and only if _ _ 2_ 2 $003 6 5 5. (2.48) _ 2_ 2 6+\/6 e $903-$— 2 J2 We now try and fill the gap in the condition (2.48). Expressing (1 — 6)e‘5+90 from (2.45), we get ‘/ __2_ 2 F(90)=2€-9° [(1- (y/52-E23—03+00))e 62 90”“ (2.49) — (1 - V62 — 52) eV52‘52]. Similarly to (2.47), if 2 —x/62—52S (/62—52——63+90SV62—52, (2.50) then F (60) _>_ 0. Carefully solving this inequality, we obtain that (2.50) holds if and only if —\/62-€2 —6+\/62—€2 $903 2 - 2 (2.51) 47 To summarize, the fact that 60 E (—§5, %) implies either (2.48) or (2.51), which, in turn, imply (2.47 ) or (2.50), respectively. Either one of the latter inequalities implies that F(00) Z O. This completes the consideration of this edge. 3. Edge a=\/62—52, a_=6 Since of“: + 01 = 2012 + 262, we have 02+ = \/62 — 282 + 262, i S |0| S E. The function to minimize is, therefore, 15(0) = 2 (1 — «62 — e2) eV62-52—(1—5)e<‘+9—(1 — J52 — 252 + 292) eV52-2€2+292-9. We seek the absolute minimum of F. First, we observe that we only need to consider 6 < 0. Indeed, since 0 3 J62 - 262 + 202 g 6, we have (1 — 6):? g (1 — J62 — 252 + 262) #5245244"? So, if 6 Z 0, then (1 _ (5)66” + (1 - \/62 — 252 + 202) eV 62-2€2+202—9 g (1 _ (5)65—9 + (1 _ \/52 _ 252 + 292) e\/62—2£2+202+0. i.e. F(0) Z F(—6). Therefore, min F(6)= min F(6). e e 753955 —559$—75 We now look for the minimum of F over the interval —5 S 0 g —755. We have F’(6) = e9 [(1— (x/(s2 — 252 + 262 - 20)) eV52-2€2+292-29 — (1 — 6).? . 48 We claim that F’(6) 5 0, — 5 S 6 S —752-. Indeed, since 6 5 755 and 5 2 2496, we have 26 S —\/25 g —§6. Thus 6+26 S O and J62 — 252 + 262 - 26 2 6. Since the function .9 H (1 — s)e’ is decreasing for s 2 0, we conclude that F’(0) S 0. Finally, since there are no extremum points in the interior of the interval, and F’ (6) S 0, we conclude that 5: min F6: min F6=F --—, isms: ( ) asses-75; ( ) ( V2) i.e. the minimum of f over this edge is attained at a vertex. 4. Edge a 2 V62 — 52, 0+ = V62 — 52 Since 013 + (11: 2(12 + 262, we have a_ = J62 — e2 + 262, ——§§ S |6| 5 i. The function to minimize is, therefore, F(6)— - 2(1— V62 — 52)ev522-€ _(1_ V52 _ 52 + 292),r3 \/62-—e2+262+6 — (1— x/6—2-—52) 6 62‘52‘9. As in the previous case, we only need to consider 6 < 0. Indeed, assume that 6 Z 0. Since (1— \/62 — 52 + 262) e 62‘€2+292 S (1 _ 52 _ 52) em 2 we have (1 — V62 — e2 + 202) eV52-€2+292+9 + (1 _ m) 8 5242—9 S (1— V62 - 52 + 262) eV 62“2+292‘9 + (1 _ 52 _ 52) e\/62—-e2+6. Therefore, as above F(6) 2 F (-—6) and we have 49 Assume now that there exists a local extremum 60 in this interval. Then F’ (60) = 0. This gives (1— 007:2) 8(62-62—90 = (1.. (fl? _ 52 + 203 — 200) 3(62-62+298+90 and F(60) = 2800 [(1 __ W) e‘/¢§2_E2 __ (1— (\/62 _ 52 + 20(2) + 00)) BW+90] . We know that if W52 — 52 + 203 + 60 2 V62 - 52, then F(60) Z 0. Solving this inequality we obtain )0.) 2 2052 — 52. What to do if |6o| < 2 62 - E2? Assume that is the case. Here we have to consider the behavior of the derivative. We have F'(6) = 8-9 [(1 — M) 6/63:2 — (1 — («52 — e2 + 202 + 20)) eV62-E2+2"2+29] . Recall that (1— :1:)eJr — (1— y)ey Z 0 if y 2 0 and —y S :1: S y. Thus if —\/6"’—52 < \/62—52+262+26< V62—52, then F’ (6) < O. The left-hand inequality translates into 0 < |6| < 2V62 — 52, while the right-hand one, into |6| > 0. To summarize: we have F’(6) < 0 if —2\/62 —e2 < 6 < 0, thus there are no points of local extremum on this subinterval; if there is an extremum point 60 S —2\/62 — 52, then F(60) 2 O. Altogether, the two “interesting” points where the absolute minimum may be attained are 6 = 0 and 6 = f5, both corresponding to 50 some vertices of the cube, to be considered in the next subsection. This completes the consideration of the last edge. Vertices We formally have a total of eight vertices, but some are non-existent and only two give us nontrivial results. 1. As mentioned before, the set 35,5 intersects the face a = 6 at one point only, the vertex 0 = a- = 01+ = 6 where we have f = 0. 2. The vertex a = a- 2 0+ = V62 — 52 also gives 6 = O and f = 0. 3. The vertex (1 = V62 — 52, a- = 01... = 6 was considered above as a part of the edge a_ = (1+ = 6 on which we have f 2 0. 4. The vertex 0 = a- = \/62———€2, (2+ = 6 gives 6 = i%, Then fl9=755 = 2(1— m) e 62-52-(1— W) 6W+s/¢2_(1_5)es_em and f'6=— 2 = 2 (1 - W) e ”-62—(1 — W) .m—e/«a_(,_6,e.../a Since (1- V62 — 52) (2"‘52‘£2 2 (1 — 6)e‘5, we have 5. The vertex (1 = (1+ = V62 — 52, a- = 6 is symmetric to the previous one and gives the same result. We have thus solved the first part of the extremal problem (2.42)—(2.43) in the sense that the only possible nontrivial (meaning negative) minimum the function f can 51 have on 35,5 is given by 9(05) = (1 — W) «9/35752 (2 — (BE/‘5) — (1 — 5)e6-W~". (2.52) 2.4.4 Stage 2 We are now in a position to solve the minimization problem (2.43) 6(5) 2 min {6 E (e,min {7:28, 1}) : g(6,e) Z 0}. (2.53) Differentiating g with respect to 6, we get 96(5,€) : 6 [eé—s/fi _ e\/52_s2 (2 _ Bah/2)] . We observe that 95 > 0. Indeed, checking if 6 — i > V62 — 52, we obtain that this condition is equivalent to 6 < '273/"5 6, hence it is satisfied. Then 9505,13) Z 56 V 62—52 [—1 + (as/‘5] > 0. Therefore, if the equation g(6, a) = 0 has a solution, then it is unique and solves our extremal problem. We thus look for a solution of the equation g(6,€) = 0 (2.54) in the interval [5, min {237 5, 1}) . We have two cases 22 <_\£ 1. E 3 We seek the solution 6 E [5, 5—312 5) . We have g(€,€) .—.. 2 — 6W? — (1 - gee-6N5 < 0 52 and 3E 35 5 —,5 = 1__ 2.66% _ (1___)]em >0, g(zfi) l< 2fi)( ) 2f which implies that there is solution 6 of (2.54) inside the interval [5, 2.3/‘2 5) . Ni 2. —<5 <1 We seek the solution 6 6 [5,1). As above, g(5,5) < 0. On the other hand, 90,5) = (1— v1— 52) 5V 1‘52 [2 —eE/‘/§]. If 5 < s/210g 2, we have g(1,5) > O and, hence, there is a unique solution of (2.54) in the interval [249, \/210g 2) . Putting the two cases together, we see that, provided 5 E (0, x/2 log 2) there is a unique solution 6(5) of equation (2.54), which also solves our extremal problem (2.53). On the other hand, in light of example (2.30) this is the best we can hope for, since the example implies that 33(5) = 00 if 5 2 \/210g 2. Therefore, we have proved that 53 = \/2 log 2, as Theorem 2 asserts. Cl Some of the prospects and future directions of research on this topic are discussed at the end. We are now turning to the result that deals with a property not unlike the one this chapter has been devoted to. The space considered there, the Chang- Wilson-Wolff space, is better than BM 0. Thus the result is stronger: instead of summability of the exponent 5‘6 we get summability of 5°82. We deviate from the formalism of the Bellman function method, so the proofs are Bellman-function—type proofs. 53 Chapter 3 Bellman-function—type proof of a local Chang-Wilson—Wolff theorem and related results 3. 1 Introduction Let D be the dyadic lattice on [0,1]. For each I E D and every :1: 6 [0,1], define the dyadic cone I‘ [(513) to be F1(:1:)={JED: JQI,J3:I:}. (3.1) Let (p E L1([0, 1]) For every I E D, let (cp), = fif11+—<90)1_)2) . (3.2) IeI‘lO,” (1:) 54 be the s-function of 1p. Let E be a measurable subset of [0, 1]. We introduce the Chang-Wilson-Wolff space F(E) = {90 5 L1 = 11506111005) < oo}. (32») Then ||S¢p(a:)||Lw(E) defines a semi-norm in this space. (In the case E = [0, 1], it is actually a norm, with the usual factorization over constant functions). Fix (,0 E F(E). For every I E D such that m (If) E) 96 0, let 2 SI = Z (<‘PlJ+ " (WlL) (3-4) J6F1(:z:) L°°(E) and 51 = 0 if m(I DE) = 0. Obviously, 5; depends on 1p but we will not indicate this dependence when the context is unambiguous. We prove that if SIM] = ||S 0 and C > 0 such that 2 / e“(~°-10,11) s c. (3.5) E The integrability of 50"“2 over the disk D under the assumption f0 |Vu(:c)|2da: S 1 has been studied in [22] and [11]. In [1], the authors study a question like ours but for functions analytic in ID). Namely, they answer in the affirmative the ques- tion of Beurling and Moser as to whether the fact that In | f’ (z)|2 dz S 7r implies fr exp I f (15“9)|2 d6 S C for some absolute constant 0. Considering for an instant 11‘ instead of [0,1], we note that even in the case E = T (3.5) holds under weaker conditions, since if 1p 6 L2('ll‘) and 11%, such that 98% > 0 and there exists 6 6 (0,1) such that flab) 132*— < 4(1— 0) (3.6) 63 — , 811 (C7 d) for any (a,b), (c,d) E IR x IR+ such that b —- d S --662 and [a — cl S %6 for some 6 Z 0. Then, for every cp E F(E) and every I E D such that m(I n E) 79 0, 1 1 B «was» 2 53 (<

,_ .31-) + §B(<<0>1+ .81..) . (3.7) Proof. Assume the existence of such a function B. By definition (3.4), S, = maX{SI_1SI+}+ (0P);+ — <¢)1_)2- 56 Let L0 = maX{S[_,S]+} , L = S] = Led-62, L_ = S]__, L+ = 31+, CE = ((p)l, :1;- = ($0)1_ , 33+ =<‘P)1+, 5 = [33+ - x’l. Then B (<10, .51) - ,l-B (<0)._ .51-) — $30.01, .31..) = B(z,L) — é— (B(:r+,L+) + B(:c",L")) 1 2 301.52) — 5 (3501.). 30,5.» since %-L > 0. We continue = [B(:z:, Lo + 62) - 3(117, Lo + (952)] + [303,110 + 952) - 3(‘7’3 L0)] + B(:1:, L0) — % (B(:1:+, Lo) + B(:1:_, 110”] BB 1 823 >— —(:z:, fi)62(1 -6)— 4 83: ——2- 2 6L —(77 L0)6 for some 6 6 [Lo + 662, L0 + 62] and some 7 between 12‘ and :19“. 3(71L0) m0) [ 2 0’ since 7 S [as — x'l = la: — z+| = %6 and L0 — 6 S —662. 168 =10L“ 0)[4<1—6>— 58:13? Q) This completes the proof. E] The choice of B We introduce the family of functions B parametrized by t 2 0. Let B.(a:, L) = etI‘LAtzL (3.8) for some A 2 0. We find the best suitable A below. For now, we check that the 57 functions Bt satisfy the conditions of Lemma 1. Fix any 6 6 (0,1). Take a, b, c,d as in the formulation of Lemma 1. Then 828, (a b) 2 2 3x2 1 t exp(at + At b) 1 2 = = — - t At b—d 92% d, At2€xp(ct+At2d) Aexp< A» S 8—62/(2Asm'11). This suggests that we need to choose 6 so that A is the smallest possible. Let D = 4%,; Then the equation relating 6 and A becomes DeD/(49) = 1 — 0, (3.9) which defines D as a function of 6 in the interval (0,1). We are seeking the maximum of D. Differentiating (3.9) with respect to 6, we get D D2 D/(49) ’ + _ _ _ + — Setting D’ = 0 and solving for D gives 462 1—6' D = (3.10) 58 Solving (3.9) and (3.10) simultaneously yields 2+) D W (3...) 1 4 :1+2 i)’1+2w( where w is Lambert’s w-function, i.e. w(z) is the solution of the equation we‘” = 2:. Therefore, for the best A we obtain ————)- z 0.5291. (3.12) We are now in a position to prove the following lemma. Lemma 2. For every (p 6 F(E) and every t 2 O, / et(‘P(3)—(‘P)[0,1])ds S eAt2S[O,1]. (313) E Proof. Let B; be given by (3.8) with A given by (3.12). We will apply Lemma 1 repeatedly, at every step omitting the terms corresponding to dyadic intervals whose intersection with E has zero measure, i.e at every step we use the simple fact that Z IIJIB<<¢>J.SJ) 2 2 (JIB(((o>.,,sJ), ”'5’" 45:53:... which allows us to apply Lemma 1 to every term in the latter sum. Thus we have et<¢l[o,1]+At25[o,ll = B (((p)[0 1] ,S[o,1]) 1 1 -B (( 2 (1130.0 >,.SI)2 Z IIIB((<.o>,.S,) (1| =12- |1|=2--n m(IUE);£0 m(IUE)9éO 59 = 2 [Il et(1p)]+At2SI. |1|=2~n m(1uE)¢o Now let n -+ 00 and observe that, since SW] < 00, we have 5'; —) 0 as [I] —> 0. On the other hand, 2 |I|et1 —*/et""(’)ds asn-—>oo, E III=2-" m(IUE);éO which completes the proof. Cl Lemma 3. Let (,0 E F(E). Let E,\ = {x E E: w(m) — ((p)[0,1] > A}. Let S = 510,1]. Then m(E1) s erg/(“5), (3.14) where A is given by (3.12). Proof. We apply Chebyshev’s inequality to (3.13) with t = A/ (AS). m(E,\) 6A2/(AS) < / eA(SP(8)—(‘P)[0,1])/(A3)ds S €A2/(2AS), E concluding the proof. C] We are now in a position to prove the main result. Theorem. If ||(,0||p(3)S1, then 2 / $946100!) 3 C(01) (3.15) E for any a < —1-, with A given by (3.12. 2A 60 Proof. Lemma 3 implies that m (“9" ‘ (6)1011! - 2 1 F1x any 7' > 1 such that ar < 57. Let 2 2.)) g 2562““). r" r 1 Fk={z(o,1][37}1 k=0,2,--- (3-16) Take any a < fl. Then, using (3.16), 2 / e“(¢"(“°>[0,ll) S m(Fk)ear2k+2/A2 S 2e—1~2k/(2A3)ec1rr'2k'l'2/A2 = 2e—r2k(1/(2A)—orr2)/A2. Fl: Therefore, [60(‘p_(¢)[0,1])2 :/ (-

[01])2 E {Ir—(610.1151 } 1 °‘ 00 2k 3 m ({l“? ‘ (01ml S 2}) a” + 226’ (“WM-MW, k=0 where the last sum converges and depends only on 0 (provided the best choice of r has been made). This completes the proof. E] 3.3 Bellman function considerations Lemma 1 and Lemma 2 are where the Bellman function technique is used, the rest of the proof follows using standard arguments. One can observe that no extremal problem as such has been posed. However, in the spirit of Chapter 1, we can try and associate with every dyadic interval I such that m(I 0 E) aé O and every function (p E F(E) a point in a certain two-dimensional domain. First, we let Fa be the 61 “oz-ball” in F(E), Fa = {cp E F(E) : ||Scp(a:)||Lm(E) g a}. What is the domain in this case? This depends on the choice of the variables. Lemma 1 suggests the choice ($1,152) = ((90),,31). Then the domain associated with F0 is {20 = IR x [0,0]. We can define the Bellman function Bf(:l:) = sup {/ ems) ds : (99)“),1] = $1,561] = 1:2} . (3.17) E sOEFa If we let cfi = 90 + c, then (95) 2 (4p) + c and 397’ = .9". Furthermore, / em”) ds = etc / ems) ds. E E Taking the supremum in the last identity, we get B?(:c1 + c, 232) = 6“ Bf"(:r1,:rg). If we set c = —:c1, we obtain that B?($1,$2) = €trlft($2) (3-18) for some ft 2 0. This is one of the considerations that led us to choose the family (3.8) of Bellman function majorates. We observe that the inequality (3.7) New-)+éB<<¢>I+,s:+> NIH B(I’SI) 2. implies some sort of concave behavior about B, although the functions (3.8) are 828 decidedly not concave and (3.7) holds because of the subtle interaction between 5—13- and g—E. If one were to produce a concave function U of the form (3.18), which would also be increasing with respect to the second argument, one could conclude 62 %U(<

1 :51-) + -;-U ((99)1+ :SI+) S U((90)1,%(SI_ + SI+)) S U(<99>1’SI)’ where the last inequality is due to the definition of SI and the fact that % Z 0. However, a concave function of the form (3.18) does not exist, signaling the need for more delicate considerations. The approach used to obtain the results of this chapter may benefit from being put on a more formal Bellman function method footing. However, as the preceding discussion demonstrates, the questions of formulating the extremal problem exactly, the choice of the variables, and, of course, finding the Bellman function explicitly can be very complicated in this case. Some intriguing perspectives are discussed at the end. We now turn to a result, in which the Bellman function does not appear in con- nection with any extremal problem at all, although, doubtless, the corresponding formalism may be developed. This result brings us back to the space BM 0, the main space of interest in this work. The old and famous question of H1 — BM 0 du- ality is examined using the Bellman-function-type approach, which proves powerful; one Bellman-function lemma yields concise proofs in both, the dyadic and continuous settings. 63 Chapter 4 Bellman function and the H1 — BMO duality 4. 1 Introduction In this chapter, we aim to demonstrate technique rather than new results. Specif- ically, with the aid of the Bellman function method we prove one, the more tech- nically involved, direction of the famous Fefferman duality theorem by elementary means. Namely, we establish the fact that BM00('lI‘) C H1('Il‘)‘ (BM00('II‘) = {cp E BMO(T), (0,00) and M : D —) [0,1171], such that 1 S[_ = 31+ 2 S] and M1 2 §(M1_ + M1+),VI E D. (4.1) Lemma. Let S and M be as above. Assume there exists a 02-function B : (0,00) x [0, M] —+ R satisfying - 2 BB BB M 9—? < 0, @- 2 0. (4.2) o_<_B(x,y)s2M\/i 7537?? 8:132— 6y? Then, for any positive integer n, Z |J|\/(SJ+ - SJ) (MJ - g(MJ_ + MJ+)) S g 2'" Z \/S—J- (4-3) JED JED lJIZ2""+1 IJI=2-" We will prove the lemma and demonstrate our Bellman function later. For now, we will establish the main results. 4.2 The dyadic case d Consider the dyadic lattice D = DT on T. Let F? be the dyadic Triebel-Lizorkin space, 1/2 £92: feLl; /1~( Z ((f),+—(f),_)2) d61+-(f)1_)) do. 196:160 The definition of the dyadic BM 0 is the same (up to a constant multiple) as the one we used in Chapter 2: BMOdz cpELI: sup— JED'JI 12((cp)l+ )_)2 |1| < 00} (4.5) IcJ with the norm 1/2 ll‘PHBMOd: 5111;0le ((‘PM— ‘P>1_2) III) - ICJ To see the equivalence of the definitions (2.3) and (4.5), recall the Haar system: for every dyadic arc J, let 1 —— on J. (M h] = —— on J+ . (4.6) VI 0 elsewhere It is easy to check that {h J} K. D form an orthonormal system in L2 = {f E L2 (T) : f1. f (0) d0 = 0}, what is more (and well-known) is the fact that the Haar system actually is a basis for Lg. For any function f E L1 and every J E D one can compute the corresponding Haar coefficient, (M )= [fl (._ — m...) (4.7) For f 6 L3 we then have f = ZJeD(f,hJ)hJ and 66 my = Zuni)? = Z 1‘—,f1(._ — mg)? JED JED We state our main result. (I :- Theorem 1. BMOd 2 (F92) . Proof. The more difficult inclusion is handled using the Bellman-function lemma stated above. (I t Lemma 1. BM 0d C (F92) . More precisely, in terms of the Haar coefi'icients, for d every cpEBMOd and fEF‘fz, 2 law 1w.» = $2 HI I m.+ — ._ Hat. — (99)J_| (4.8) JED JeD 1 — MHWHBMOd “fIIFC‘i92 d Proof. Fix cp E BMOd, f 6 F3”. For every J E D define ICJ def Then 0 < M J < M— Define ”scum... and M.-§(MJ++MJ_) = ((99)., - (¢)J_)2- s. = 2 (ml, — ,_)2. 19.1 2 2 Then SJ.=SJ_=Z (m1, — ,_) and SJ.—SJ = 81:51 = (.;._) . 131 We thus see that the conditions (4.1) of the lemma are satisfied. 67 Assuming the existence of the function B in the lemma and using (4.3), we obtain 2: MI I m... — ._ ll «0)., — (m I s 92‘- 2-"2 Z (0)., — .,_)2. JED IJI22-n+1 |J|=2‘" Letting n —» 00, we get the statement (4.8) of the theorem. C] The converse inclusion follows along more conventional lines. 0' 2 "' d Lemma 2. (F?) C BMO . - d Proof. We want to show that for every continuous linear functional 1 on F92 there exists cp E BM 0" such that ”‘pllBMOd 5 CW“ (49) and 1m = [T 90(0)f(6) do, w e 5‘32. (4.10) d First, we observe that L3 6 F92. Indeed, for f 6 L3, 1 |If||:402= (fr ( Z ‘((f>1+-1+-z_)2d9 139;IED 139;IED =1; /r we) (1+ - ,_) «16 = 2m (at. — ,_) = 4||f|lig- 16D d * Let I 6 (F92) . We can apply the Riesz representation theorem to 1ng and conclude that there exists a function (,0 6 L3 such that W) = [T 90(9)f(9)d0. w 6 L3. (4.11) 68 We test I on appropriate elements of L3 to see that «,0 E BM 0". Let a; be an atom associated with a dyadic arc I , i.e. be supported on I with |a1| 3 fi’ ac. and f1 a(6) d6 = 0. We have 2 1/2 llazlllga,2 =/1~( Z ((cu).1+ - (a1)J_) ) d0 J39;JED 1/2 /<<>> . S (flJaaZJc1( 0 as F 1 n —+ 00. This concludes the proof of Lemma 2 and Theorem 1. Cl 4.3 The continuous case Following [21], we define H l = H1('ll‘) using the area integral; specifically = {f E L1 ; [T (Amie) lft(€)|2dA(§))l/2 d6 < 00} (4.12) with the corresponding norm Ilfllul = [r ( / .ae, award/1(a))” am. (4.13) Here f (z) is the harmonic extension of f into 1D. Pa(e‘9) is the cone-like region } . We will specify the angle lei” — z| l with vertex 6‘”: Fa(e‘9)= z E 11):, 1 - |z| sma a a little later. The corresponding definition of BM 00 = BM 00(T) is arc I CT BMOo = {w 6 L1: SUP fi/Q |¢'(€)l2(1-|€|)dA(€) < 00, w(O) = 0}, (4-14) I where 00. 1/2 . 1 I 2 _ I 2 3320 E: HI ('7, [TQJIMOI (1 KIM/1(6)) (AQJ|f(€)IdA(€)) ”lag-n+1 1/2 72 1/2 1/2 — Jl/2 ’ 21— dA ' 2dA —JEZD|| (fQ Imam |6|) (a) (fTQJIIIoI (a) > Z W” /QJ|

0: .2 MIN )ldA(€) 12%| I/QJIr I I I I .1— > C [D I

C" If acpgflog -—1—dA(§)| 1D |€| = C” (wf) 10g l—éldA €()‘ where we have used the fact that 6 Cl/T0 —a_B§_Bi>C21 for some positive constants 01,02. The proof of the key lemma using B suggests that we want to choose these constants in order to minimize the ratio Cl/\/Cg. To make the estimates “sharper,” we require that gay?- = 0. (We cannot require equality for £7?- 5 0 .) This means that B is a linear function of y. Furthermore, % must be negative. Because of the homogeneity in the way B is used in the lemma, we only need to choose one coefficient in that linear function. Thus we seek B in the form Therefore, __3£3_B_A—y>A—M Bar By_ 2 - 2 ’ since y S M. We have CI = A and for the ratio to be minimized _C_1____\/§A _ \/’C_2_\/A—M' The minimum of this ratio is attained at A = 2M, thus producing the function B(:v,y) = x/HQM - y), satisfying conditions (4.3). Some possible directions of future research on the Bellman function and H 1 - BM 0 duality are discussed in the next section. 77 Research prospects In this section, we discuss some possible directions for future research on the topics presented throughout the thesis. We do not attempt to embrace all possible research prospects but rather concentrate on those of immediate importance and the greatest promise. J ohn—Nirenberg inequality The most natural continuation of the research on this topic would be to find the Bellman function for the weak-form John-Nirenberg inequality (1.14) m ({a: E I : |gp(x) — (90)” > ,\}) S cle_C2A/”‘P”BMO(I), where the L2-based BM 0 norm is used. While it is entirely possible that the extremal functions are different for the weak and integral form of the inequality, it would be very interesting to find out how they are related. The corrasponding formulation would look something like this B(s,t) = sup{|{:1: 6 [0,1] : |[o,1] I > t}|, cp E BMO,” ”SOHBMOP S 3}. Finding B exactly would give sharp constants for the traditional John-Nirenberg inequality. This problem is open even in the case p = 2. 78 The next order of business would be to explore the possibility of finding the Bellman function for the integral form in the LP-based formulations with p aé 2. We observe that the ease with which we managed to associate a plane domain with the s-ball in BM 0 in the L2 setting will not be there. Somewhat similar formulations are possible in the case of even, positive p, however the number of variables involved would grow very fast, as would the number of constraints relating those variables. The essential feature of the formulation we have used is that, after considerations of homogeneity, the Bellman function can be sought as a function of one variable. That is unlikely to happen if the number of the variables in the set-up increases. Perhaps a more perspective direction of research (and, possibly, of more interest to the general mathematical audience) would be to try to use the results obtained to deal with BM 0(Q) with Q being a cube in IR". The John-Nirenberg inequality holds in higher-dimensional BM 0; hence the question of best constants. Chang-Wilson-Wolff theorem Having successfully treated the local Chang-Wilson-Wolff theorem in the dyadic case using the Bellman function technique, we would like to apply the method to the continuous version of the theorem. Namely, for f E LICK"), let P7(:c) = {(3}: t) 6 RnH- Ix - yl < 7t}. Let 1/2 A~f($) = (fr ( )lVyf(y,t)l2tl‘"dydt) - Assume A7 f E L°°(lR"). Then, as shown in [2], U.) ___|_2) su exp c < C , 1:..R..IT|/, (I'll/17f“; 2 79 where c1, C2 depend on 7 and the dimension. A successful application of the Bell- man function technique (how to do it is far from obvious in this case) may yield sharp constants. If the constants are independent 1of 7, one may replace A, by the g-function, g(f)(ac) = (foo |V3f(x,t)|2t cit) /, thus solving a famous open problem. 0 On the other hand, it would be instrumental to try and pose the extremal problem even in the dyadic case, so that the Bellman function could be found explicitly. The difficulties with the choice of variables and the right scaling (an impossibility in the case E 7:4 [0,1] ) need to be addressed. H1 - BM 0 duality The successful formulation of an optimization problem in this case would be a signif- icant accomplishment, since it would not only be the first of its kind, but also would shed some light on how to proceed to find the best function exactly. If the optimiza- tion problem is explicitly solved, the sharp constant of embedding would be produced, in the one- and multi-dimensional cases, which, to the best of our knowledge, is an open problem. 80 Bibliography [1] S.-Y. A. Chang, D. E. Marshall. On a sharp inequality concerning the Dirichlet integral. American Journal of Mathematics, Vol. 107 (1985), no. 5, pp. 1015- 1033. [2] S.-Y. A. Chang, J. M. Wilson, T. H. Wolff. Some weighted norm inequali- ties concerning the Schrodinger operators. Comment. Math. Helvetici, Vol. 60 (1985), no. 2, pp. 217-246. [3] D. L. Burkholder. Boundary value problems and sharp inequalities for martin- gale transforms. Annals of Probability, Vol. 12 (1984), no. 3, pp. 647-702. [4] D. Donoho, M. Frazier. Matrix boundedness on discrete Littlewood-Paley Spaces with applications to Haar and interpolating wavelets. Preprint, MSU, 2002. [5] C. Fefferman. Characterizations of bounded mean oscillations. Bull. Amer. Math. Soc., Vol. 77 (1971), pp. 587-588. [6] J. B. Garnett, Bounded Analytic functions. Pure and Applied Math. #96, Aca- demic Press, 1981. [7] S. Hucovic, S. Treil, A. Volberg. The Bellman functions and sharp weighted inequalitias for square functions. Operator Theory: Advances and Applications, the volume in memory of S. A. Vinogradov, Vol. 113, Birkhauser Verlag, 2000, pp. 97-113. [8] F. John, L. Nirenberg. On functions of bounded mean oscillation. Comm. Pure Appl. Math, Vol. 14 (1961), pp. 415-426. [9] A. Korenovskii. The connection between mean oscillations and exact exponents of summability of functions. Math. USSR-3b., Vol. 71 (1992), no. 2, pp. 561-567. [10] N. Krylov, Optimal control of diflusion processes. Springer-Verlag, New York, 1980, 308 p. [11] J. Moser. A sharp form of an inequality by N. Trudinger. Indiana University Mathematical Journal, Vol. 20 (1971), no. 11, pp. 1077-1092. [12] F. Nazarov. A counterexample to Sarason’s conjecture. Preprint, MSU, 1997. 81 [13] F. Nazarov, S. Treil. The hunt for Bellman function: applications to estimates of singular integral operators and to other classical problems in harmonic analysis. Algebm i Analis, Vol. 8 (1997), No. 5, pp. 32-162. [14] F. N azarov, S. Treil. The weighted norm inequalities for Hilbert transform are now trivial. C'. R. Acad. Sci. Paris Se’r. 1 Math, Vol. 323 (1996), no. 7, pp. 717— 722. [15] F. Nazarov, S. Treil, A. Volberg. Bellman function in stochastic control and harmonic analysis in Systems, Approximation, Singular integral operators, and related topics, ed. A Borichev, N. Nikolski, Operator Theory: Advances and Applications, Vol. 129, 2001, pp. 393-424, Birkhauser Verlag. [16] F. Nazarov, S. Treil, A. Volberg. The Bellman functions and two-weight in- equalities for Haar multipliers. Journal of the American Mathematical Society, Vol. 12 (1999), no. 4, pp. 909-928. [17] F. Nazarov, A. Volberg. Heating of the Ahlfors—Beurling operator and estimates of its norm. St. Petersburg Math. Journal, Vol. 14 (2003), no. 3. [18] F. Nazarov, A. Volberg. The Bellman function, the two—weight Hilbert trans- form, and embeddings of the model spaces K9. Dedicated to the memory of Thomas H. Wolff. J. Anal. Math, Vol. 87 (2002), pp. 385-414. [19] St. Petermichl, A. Volberg. Heating of the Ahlfors—Beurling Operator: weakly quasiregular maps on the plane are quasiregular. Duke Math. Journal, Vol. 112 (2002), no. 2, pp. 281-305. [20] St. Petermichl, J. Wittwer. A sharp estimate for the weighted Hilbert transform via Bellman functions. Michigan Math. Journal, Vol. 50 (2002), no. 1, pp. 71-87. [21] E. Stein, Harmonic Analysis: Real—variable Methods, Orthogonality, and Oscil- latory Integrals. Princeton University Press, Princeton, New Jersey, 1993. [22] N. Trudinger. On imbeddings into Orlicz spaces and some applications. Journal of Mathematics and Mechanics, Vol. 17 (1967), no. 5, pp. 473-483. [23] V. Vasyunin. The sharp constant in the reverse Holder inequality for Mucken- houpt weights. Algebra i Analis, Vol. 15 (2003), no. 1, pp. 73-117. [24] A. Volberg. Bellman approach to some problems in harmonic analysisEquations aux De’rive’es Par'tielles, Exposé n. XX, 2002. 82