arxiv: v1 [math.ca] 31 Dec 2014

Similar documents
This article appeared in a journal published by Elsevier. The attached copy is furnished to the author for internal non-commercial research and

for all subintervals I J. If the same is true for the dyadic subintervals I D J only, we will write ϕ BMO d (J). In fact, the following is true

JUHA KINNUNEN. Harmonic Analysis

Bellman function approach to the sharp constants in uniform convexity

Lecture 4 Lebesgue spaces and inequalities

Introduction to Real Analysis Alternative Chapter 1

THE JOHN-NIRENBERG INEQUALITY WITH SHARP CONSTANTS

Examples of Dual Spaces from Measure Theory

Chapter 1. Measure Spaces. 1.1 Algebras and σ algebras of sets Notation and preliminaries

A NOTE ON WAVELET EXPANSIONS FOR DYADIC BMO FUNCTIONS IN SPACES OF HOMOGENEOUS TYPE

Part III. 10 Topological Space Basics. Topological Spaces

3 Integration and Expectation

Deposited on: 20 April 2010

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

Topological properties of Z p and Q p and Euclidean models

In this note we give a rather simple proof of the A 2 conjecture recently settled by T. Hytönen [7]. Theorem 1.1. For any w A 2,

3 (Due ). Let A X consist of points (x, y) such that either x or y is a rational number. Is A measurable? What is its Lebesgue measure?

Integration on Measure Spaces

Differentiation. Chapter 4. Consider the set E = [0, that E [0, b] = b 2

L p Spaces and Convexity

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

arxiv: v1 [math.ca] 25 Jul 2017

Measures. Chapter Some prerequisites. 1.2 Introduction

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

MATHS 730 FC Lecture Notes March 5, Introduction

Construction of a general measure structure

Measurable functions are approximately nice, even if look terrible.

Convexity in R n. The following lemma will be needed in a while. Lemma 1 Let x E, u R n. If τ I(x, u), τ 0, define. f(x + τu) f(x). τ.

2 (Bonus). Let A X consist of points (x, y) such that either x or y is a rational number. Is A measurable? What is its Lebesgue measure?

Geometric-arithmetic averaging of dyadic weights

Optimization and Optimal Control in Banach Spaces

VECTOR A 2 WEIGHTS AND A HARDY-LITTLEWOOD MAXIMAL FUNCTION

Homework 1 Real Analysis

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

The Hilbert Transform and Fine Continuity

Decomposition of Riesz frames and wavelets into a finite union of linearly independent sets

Mathematical Methods for Physics and Engineering

( f ^ M _ M 0 )dµ (5.1)

MATH 202B - Problem Set 5

Advanced Calculus: MATH 410 Real Numbers Professor David Levermore 5 December 2010

x i y i f(y) dy, x y m+1

Estimates for probabilities of independent events and infinite series

Geometric intuition: from Hölder spaces to the Calderón-Zygmund estimate

THEOREMS, ETC., FOR MATH 515

A RELATIONSHIP BETWEEN THE DIRICHLET AND REGULARITY PROBLEMS FOR ELLIPTIC EQUATIONS. Zhongwei Shen

Some Background Math Notes on Limsups, Sets, and Convexity

Chapter 2 Convex Analysis

GAUSSIAN MEASURE OF SECTIONS OF DILATES AND TRANSLATIONS OF CONVEX BODIES. 2π) n

MTH 404: Measure and Integration

MAT 570 REAL ANALYSIS LECTURE NOTES. Contents. 1. Sets Functions Countability Axiom of choice Equivalence relations 9

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

A WEIGHTED MAXIMAL INEQUALITY FOR DIFFERENTIALLY SUBORDINATE MARTINGALES

Herz (cf. [H], and also [BS]) proved that the reverse inequality is also true, that is,

Dyadic structure theorems for multiparameter function spaces

Probability and Measure

Invariant measures for iterated function systems

arxiv: v2 [math.fa] 27 Sep 2016

NECESSARY CONDITIONS FOR WEIGHTED POINTWISE HARDY INEQUALITIES

X. Tolsa: Analytic capacity, the Cauchy transform, and nonhomogeneous 396 pp

A VERY BRIEF REVIEW OF MEASURE THEORY

Metric Spaces and Topology

The small ball property in Banach spaces (quantitative results)

Continued fractions for complex numbers and values of binary quadratic forms

CHAPTER I THE RIESZ REPRESENTATION THEOREM

If Y and Y 0 satisfy (1-2), then Y = Y 0 a.s.

On the Set of Limit Points of Normed Sums of Geometrically Weighted I.I.D. Bounded Random Variables

1.1. MEASURES AND INTEGRALS

Regularizations of Singular Integral Operators (joint work with C. Liaw)

ON A MAXIMAL OPERATOR IN REARRANGEMENT INVARIANT BANACH FUNCTION SPACES ON METRIC SPACES

PARTIAL REGULARITY OF BRENIER SOLUTIONS OF THE MONGE-AMPÈRE EQUATION

ON THE REGULARITY OF SAMPLE PATHS OF SUB-ELLIPTIC DIFFUSIONS ON MANIFOLDS

1/12/05: sec 3.1 and my article: How good is the Lebesgue measure?, Math. Intelligencer 11(2) (1989),

HAUSDORFF DIMENSION AND ITS APPLICATIONS

Integral Jensen inequality

SCALE INVARIANT FOURIER RESTRICTION TO A HYPERBOLIC SURFACE

PREVALENCE OF SOME KNOWN TYPICAL PROPERTIES. 1. Introduction

Combinatorics in Banach space theory Lecture 12

Duality of multiparameter Hardy spaces H p on spaces of homogeneous type

ON THE UNIQUENESS PROPERTY FOR PRODUCTS OF SYMMETRIC INVARIANT PROBABILITY MEASURES

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

A NOTE ON WEAK CONVERGENCE OF SINGULAR INTEGRALS IN METRIC SPACES

Week 2: Sequences and Series

13. Examples of measure-preserving tranformations: rotations of a torus, the doubling map

Chapter 2 Metric Spaces

The Caratheodory Construction of Measures

Jordan Journal of Mathematics and Statistics (JJMS) 9(1), 2016, pp BOUNDEDNESS OF COMMUTATORS ON HERZ-TYPE HARDY SPACES WITH VARIABLE EXPONENT

The Skorokhod reflection problem for functions with discontinuities (contractive case)

Functional Analysis HW #3

Analysis II Lecture notes

Definitions of curvature bounded below

Signed Measures and Complex Measures

LECTURE 6. CONTINUOUS FUNCTIONS AND BASIC TOPOLOGICAL NOTIONS

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

VISCOSITY SOLUTIONS. We follow Han and Lin, Elliptic Partial Differential Equations, 5.

Final Exam Practice Problems Math 428, Spring 2017

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

It follows from the above inequalities that for c C 1

It follows from the above inequalities that for c C 1

On a Class of Multidimensional Optimal Transportation Problems

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

Transcription:

INEQUALITIES FOR BMO ON -TREES LEONID SLAVIN AND VASILY VASYUNIN arxiv:150100097v1 [mathca] 31 Dec 2014 Abstract We develop technical tools that enable the use of Bellman functions for BMO defined on -trees, which are structures that generalize dyadic lattices As applications, we prove the integral John Nirenberg inequality and an inequality relating L 1 - and L 2 - oscillations for BMO on -trees, with explicit constants When the tree in question is the collection of all dyadic cubes in R n, the inequalities proved are sharp We also reformulate the John Nirenberg inequality for the continuous BMO in terms of special martingales generated by BMO functions The tools presented can be used for any function class that corresponds to a non-convex Bellman domain 1 Preliminaries and main results Let D stand for the collection of all open dyadic cubes in R n This collection is uniquely defined by the choice of the root cube, say Q 0 = 0, 1 n If a cube Q is fixed, then DQ is the collection of all dyadic subcubes of Q By ϕ J we denote the average of a locally integrable function over a set J with respect to the Lebesgue measure; if a different measure, µ, is involved, we write ϕ J,µ Thus, ϕ J = 1 J For p > 0, and a function ϕ L p loc, let J ϕ, ϕ J,µ = 1 µj J ϕ dµ p,j ϕ = ϕ ϕ J p J and p,µ,j ϕ = ϕ ϕ J,µ p J,µ We will refer to both p,j and p,µ,j as the p-oscillation of ϕ over J; this will not cause confusion, as the measure is always fixed We will mainly need these definitions for p = 1 and p = 2 Observe that 2,µ,J ϕ = ϕ 2 J,µ ϕ 2 J,µ Let the dyadic BMO on R n be defined by 11 BMO d R n = { ϕ L 1 loc : ϕ 1/2 } BMOd := sup 2,J < J D We will also use BMO d Q when the supremum is taken over all J DQ for some cube Q For ε > 0, the symbols BMO d εr n and BMO d εq will stand for the set of all BMO d functions on the appropriate domain with norm not exceeding ε Elements of BMO are locally exponentially integrable The classical result that quantifies this property is the John Nirenberg inequality [3] Here we state it in the integral form 2010 Mathematics Subject Classification Primary 42A05, 42B35, 49K20 Key words and phrases BMO on trees, John Nirenberg inequality, explicit Bellman function L Slavin s research was supported in part by the National Science Foundation DMS-1041763 V Vasyunin s research was supported by the Russian Science Foundation grant 14-41-00010 1

2 LEONID SLAVIN AND VASILY VASYUNIN Theorem 11 John, Nirenberg; integral form There exists ε d 0 n > 0 such that for every 0 ε < ε d 0 n there is a function Cε, n > 0 such that for any ϕ BMOd εr n and any Q D 12 e ϕ Q Cε, ne ϕ Q Let us reserve the names ε d 0 n and Cε, n for the sharp values of these constants in 12, ie, the largest possible ε d 0 n and the smallest possible Cε, n The constant εd 0 n is of particular importance: it is easy to show that ε d 0 n is the supremum of ε > 0 such that for any ϕ BMO d R n, e εϕ/ ϕ BMO d is a dyadic A 2 weight In [11], we computed these constants for n = 1 : Theorem 12 [11] ε d 01 = 1 2 log 2, Cε, 1 = 2 e ε 2 e 2ε In that paper, a family of Bellman functions for the John Nirenberg inequality on the continuous as opposed to dyadic BMO was constructed, and an element of that family was identified as the Bellman function for the dyadic BMO d R Although it was intuitively clear at the time how to pick an element of the family that would work in dimensions greater than 1, the computation in [11] was unsuitable for higher dimensions In this paper, we develop several technical tools that allow us to prove the following theorem: Theorem 13 ε d 0n = 2n/2 2 n 1 n log 2, Cε, n = 2 n 1 2 n e 2 n/2ε e 2n/2 ε Theorem 13 itself is a partial corollary of the corresponding result for special structures that generalize dyadic lattices, which we now define Definition 14 Let X, µ be a measure space with 0 < µx < Let 0, 1/2] A collection T of measurable subsets of X is called an -tree, if the following conditions are satisfied: 1 X T 2 For every J T, there exists a subset CJ T such that a J = I CJ I, b the elements of CJ are pairwise disjoint up to sets of measure zero, c for any I CJ, µi µj 3 T = m T m, where T 0 = {X} and T m+1 = J T m CJ 4 The family T differentiates L 1 X, µ Given an -tree T on X, µ, a function ϕ on X is called T -simple, if there exists an N 0 such that ϕ is constant µ-ae on each element of T N Observe that each CJ is necessarily finite We will refer to the elements of CJ as children of J and to J as their parent Also note that T J := {I T : I J} is an -tree on J, µ J We write T k J for the collection of all descendants of J of the k-th generation relative to J; thus, T J = m T kj Remark 15 The definition just given is similar to the one used in Bellman-function contexts by Melas [6] and Melas, Nikolidakis, and Stavropoulos [7] In particular, like those authors, we restrict the size of children of each element of the tree The main distinction is that in those applications the trees were assumed homogeneous, meaning that each element of the

INEQUALITIES FOR BMO ON -TREES 3 tree was split into the same number of children, and all children had the same measure The prototypical such tree is the collection of all dyadic subcubes of a fixed cube in R n ; in our terminology it is a 2 n -tree However, homogeneous trees are too rigid for our purposes The elements of a general -tree have similar nesting properties, but this concept allows us to divide a parent into an arbitrary number of children of varying sizes, so long as none is too small In addition, we do not foreclose the possibility that µ has atoms and so a parent can have only one child Suppose a measure space X, µ supports an -tree T for some 0, 1/2] There is a natural associated BMO: ϕ BMOT ϕ BMOT := sup{ ϕ 2 J,µ ϕ 2 J,µ }1/2 < J T Let us make two formal definitions: 13 ε 0 = 1 log1/, K, ε = 1 e ε e ε In this notation we have the following theorem Theorem 16 If 0, 1/2], ε 0, ε 0, T is an -tree on a measure space X, µ, J T, and ϕ BMO ε T, then 14 e ϕ J,µ K, ε e ϕ J,µ Setting = 2 n gives the values of ε d 0 n and Cε, n in Theorem 13 The fact that those values are sharp does not follow from Theorem 16 and requires a separate construction of optimizers, ie, functions from BMO d for which equality is attained in 12 This construction is carried out in Section 3 Our technique works for other BMO inequalities as well We illustrate this point by establishing an inequality relating 1- and 2-oscillations of BMO functions, which implies equivalence of the corresponding BMO norms Specifically, for ϕ BMO d R n, let ϕ BMO d,1 R n = sup 1,J ϕ J D Since 1,J ϕ 2,J ϕ 1/2, we have ϕ BMO d,1 R n ϕ BMO d R n Importantly, both inequalities can be reversed Theorem 17 If ϕ BMO d R n, then for any J D, 15 2 n/2 2 n + 1 2,Jϕ ϕ BMO d R n 1,Jϕ This inequality us sharp for every value of ϕ BMO d R n Consequently, 16 2 n/2 2 n + 1 ϕ BMO d R n ϕ BMO d,1 R n As before, this theorem is a partial corollary of the corresponding result for -trees Theorem 18 If T is an -tree on a measure space X, µ, then for any ε > 0, any ϕ BMO ε T, and any J T, 17 1 + ε 2,µ,Jϕ 1,µ,J ϕ

4 LEONID SLAVIN AND VASILY VASYUNIN Consequently, 18 1 + ϕ BMOT ϕ BMO 1 T, where ϕ BMO 1 T = sup J T 1,µ,J ϕ Before proving all results just stated, let us put them in historical and methodological context The notion of a Bellman function in analysis goes back to Bukholder [1] and Nazarov, Treil, and Volberg [9], [8], [10] In the original, utilitarian meaning, a Bellman function was an inductive device with certain size and convexity properties that allowed one to estimate an integral functional by induction on dyadic or pseudo-dyadic scales in the underlying measure space In a more recent understanding, the Bellman function for an inequality is the solution of the corresponding extremal problem and, often, also is a solution of the homogeneous Monge Ampère equation on a Euclidean domain It turns out that such a solution, once in hand, not only allows one to perform the induction on scales, but also encodes information about optimizing functions or sequences The specific Bellman functions we use here have origins in our studies [11] and [12] and, like those earlier functions, they are defined on a parabolic domain in the plane However, those papers dealt with the one-dimensional BMO on an interval This meant that on each step of the induction one would split an interval into two subintervals, yielding three points in the Bellman domain that had to be controlled using the concavity/convexity of the Bellman function In dimension n there are 2 n + 1 such points, and since the domain is non-convex, it is not clear how to control all of them at the same time This has led to the introduction of - trees and the notion of -concave/-convex functions, defined the next section If one has such a tree and a function, one can run the induction three points at a time We do not provide a general recipe for constructing -concave/convex functions, but simply present natural -tree analogs of the Bellman functions from [11] and [12] With perhaps considerable effort, one can -ize any Bellman function for BMOR, and these are now plentiful: following [12], the papers [4] and [5] develop a rather general method for computing them; a somewhat different example is given in [13] Though motivated by inequalities for the dyadic BMOR n, Bellman analysis on -trees has wider applications First, it works in other geometric settings; one important example is supplied by spaces of homogeneous type, where the dyadic cubes of M Christ [2] give rise to -trees with 1/ comparable to the doubling constant Second, it naturally extends to other function classes that yield non-convex Bellman domains, such as A p and reverse Hölder classes, A p -weighted L p, etc Lastly, when properly modified to include the range 1/2, 1] it may allow one to obtain results for continuous BMO simply by taking a suitable supremum in In this modification, described in Section 5, each BMO function generates its own tree and thus yields what we call an -martingale The rest of the paper is organized as follows: in Section 2 we define -concave/convex functions and formalize Bellman induction on -trees using such functions We also present a set of easy-to-verify sufficient conditions for a function to be -concave/convex In Section 3 we prove Theorem 16, which allows us to compute the exact Bellman function for the John Nirenberg inequality for BMO d R n ; Theorem 13 then follows easily This sequence is repeated in Section 4: we first prove Theorem 18, then define and find the corresponding Bellman function, which immediately gives Theorem 17 Finally, in Section 5 we introduce the notion of an -martingale, state a general result about such martingales generated by BMO functions, and estimate the John Nirenberg constant for the continuous BMOR n in terms of a special martingale-related parameter of the space

INEQUALITIES FOR BMO ON -TREES 5 2 Main technical tools For ε > 0, let Ω ε = {x = x 1, x 2 : x 2 1 x 2 x 2 1 + ε 2 } Also, for any ξ 0, let Γ ξ = {x : x 2 = x 2 1 + ξ2 }; thus, Γ 0 and Γ ε are the lower and upper boundaries of Ω ε, respectively A line segment connecting points x and y in the plane will be denoted by [x, y] and the length of such a segment, by [x, y] Observe that if T is an -tree on X, µ and ϕ BMO ε T, then for any J T, ϕ 2 J,µ ϕ2 J,µ ϕ 2 + J,µ ε2 and so the point ϕ J,µ, ϕ 2 J,µ is in Ω ε This simple fact is the geometric foundation of the Bellman approach to the L 2 -based BMO, which consists of using an appropriate concave or convex function on Ω ε to perform induction on the generation of the tree and bound the desired integral in the limit However, since Ω ε is itself non-convex, we need to strengthen the usual notion of concavity/convexity Definition 21 A function on Ω ε is called locally concave respectively, locally convex, if it is concave respectively, convex on any convex subset of Ω ε Definition 22 If 0, 1 2], a function B on Ωε is called -concave if 21 Bβx + 1 βx + βbx + 1 βbx +, for any β [, 1 2] and any two points x ± Ω ε such that βx + 1 βx + Ω ε Similarly, b is called -convex on Ω ε if 22 bβx + 1 βx + βbx + 1 βbx +, for all β, x, and x + as above Armed with such a function, we can obtain the desired integral estimate using what is commonly referred to as Bellman induction The procedure depends on a simple geometric fact that replaces the average of N points in Ω ε with the average of just two points Lemma 23 If N 2, numbers 1,, N 0, 1 are such that N k=1 k = 1, and points P 1,, P N Ω ε are such that N k=1 kp k Ω ε, then there exists at least one j, 1 j N, for which the point R j := 1 1 j k j kp k is in Ω ε Proof Assume, to the contrary, that R j / Ω ε, j = 1,, N Since the set {x : x 2 x 2 1 } is convex, all R j are in this set Since none is in Ω ε, all are, in fact, in the set {x : x 2 > x 2 1 +ε2 }, which is also convex Therefore, their convex combination, 1 N 1 j R j = 1 N N N k P k j P j = k P k, N 1 N 1 j=1 j=1 k=1 is also in {x : x 2 > x 2 1 + ε2 }, which is a contradiction Lemma 24 Take 0, 1/2] and let T be an -tree on a measure space X, µ Let ϕ be a T -simple function; set ε = ϕ BMOT note that ϕ is bounded and thus in BMOT i If B is an -concave function on Ω ε, then B ϕ X,µ, ϕ 2 X,µ 1 µx ii If b is an -convex function on Ω ε, then b ϕ X,µ, ϕ 2 X,µ 1 µx X X Bϕ, ϕ 2 dµ bϕ, ϕ 2 dµ k=1

6 LEONID SLAVIN AND VASILY VASYUNIN Proof We will prove only statement i; the proof of ii is the same, except all inequality signs are reversed For all I T, let P I = ϕ I,µ, ϕ 2 I,µ and µi = µi Since ϕ BMO ε T, we have P I Ω ε Furthermore, P I = 1 µ I J T 1 I µ J P J We first claim that for any I T, 23 BP I 1 µ J BP J µ I J T 1 I If T 1 I has only one element, then 23 holds with equality Assume that T 1 I has two or more elements By Lemma 23, there exists an L T 1 I such that the point 1 R L := µ J P J µ I µ L J T 1 I J L is in Ω ε We have P I = µ L µi P L + 1 µ L µi RL, and, since T is an -tree, both µ L /µ I and 1 µ L /µ I By the -concavity of B, BP I µ L µ I BP L + 1 µ L µ I BR L If T 1 I has two elements, this is exactly statement 23 If T 1 I has more than two elements, we apply Lemma 23 to R L in place of P I : BR L µ K BP K + 1 µ K BR L,K, µ I µ L µ I µ L 1 for some K T 1 I, K L, and R L,K = µ I µ L µ K J L,K µ J P J This gives BP I µ L BP L + µ K BP K + 1 µ L + µ K BR L,K µ I µ I µ I Continuing in this fashion, we obtain 23 Now, let N be such that ϕ is constant on each J T N Note that this means that ϕ 2 J,µ = ϕ 2 for all such J A repeated application of 23 gives J,µ BP X 1 µ J BP J 1 µ R BP R = 1 µ J BP J µ X µ X µ X J T 1 J T 2 1 µ X J T 1 R T 1 J µ J BP J = 1 µ J B ϕ J,µ, ϕ 2 µ X J,µ J T N J T N The last expression is precisely 1 µ X X Bϕ, ϕ2 dµ, and the proof is complete Our next lemma gives sufficient conditions for a function on Ω ε to be -concave/convex Lemma 25 Let ε > 0 and 0, 1 2 ] Assume that functions B and b on Ω ε satisfy the following three conditions: 1 B is locally concave on Ω ε and b is locally convex on Ω ε, 2 B and b have non-tangential derivatives at every point Γ ε Furthermore, for any two distinct points on Γ ε, P = p, p 2 + ε 2 and Q = q, q 2 + ε 2 with p q 1 ε, where D P Q D P Q BP D P Q BQ, D P Q bp D P Q bq, denotes the derivative in the direction of the vector P Q

INEQUALITIES FOR BMO ON -TREES 7 3 For any P and Q as above, and S = 1 1 P Q, BP 1 BS + BQ, bp 1 bs + bq Then B is -concave on Ω ε and b is -convex on Ω ε Remark 26 Before proving this lemma, let us examine its conditions more closely First, Conditions 1 and 3 are clearly necessary for -concavity/convexity Second, Condition 2 is also quite natural: along any line segment fully contained in Ω ε the directional derivatives of B and b defined almost everywhere are monotone This condition can be seen as preserving this monotonicity of derivatives even for those segments that cross the boundary curve Γ ε at two points, as long as the part external to Ω ε [P, Q]in this case is not too large Note that B and b need not be defined along the segment [P, Q] itself Proof Again, we only prove the statement for B If the segment [x, x + ] is contained in Ω ε, inequality 21 holds because B is locally concave in Ω ε Thus, we need to consider only those segments that cross Γ ε at two points Take any β [, 1 2 ] and any x, x + Ω ε such that x 0 := β x + 1 β x + Ω ε Let l be the line containing the three points x, x 0, and x + and assume that l intersects Γ ε at two points, P and Q Then l also intersects Γ 0 at two points, say U and W, named so that the order of points on l is U, P, Q, W Without loss of generality, assume that x, x 0 [U, P ] and x + [Q, W ] Finally let S [U, W ] be such that [S, P ] = [S, Q], that is P = 1 S + Q Figure 1 shows the location of all these points for = 1 6 and β = 1 4 Figure 1 Picture for the proof of Lemma 25 with = 1 6 and β = 1 4 The essence of the lemma is that if Conditions 1 and 2 are fulfilled, then inequality 21 reduces to its special case assumed in Condition 3, ie, to the situation when β = and x 0, x + Γε In fact, it will be technically convenient to show a little more Specifically, we claim that if 24 BP 1 BS + BQ, then 25 Bx 0 [x 0, x + ] [x, x + ] Bx + [x, x 0 ] [x, x + ] Bx+,

8 LEONID SLAVIN AND VASILY VASYUNIN for any x [U, S], x 0 [x, P ], and x + [Q, W ] Parametrize the segment [U, W ] by xt = 1 tu + tw, 0 t 1 Let t, t S, t 0, t P, t Q, and t + be the values of the parameter corresponding to the points x, S, x 0, P, Q, and x +, respectively Let At = Bxt and define a new function F by on the domain F t, t 0, t + = t + t At 0 t + t 0 At t 0 t At + D = {t, t 0, t + : 0 t t S, t t 0 t P, t Q t + 1} In this notation, condition 24 is equivalent to 26 F t S, t P, t Q 0, and to prove that 24 implies 25 is the same as to prove that 26 implies 27 F t, t 0, t + 0, t, t 0, t + D Thus, assume that 26 holds Since B is concave in Ω ε, A is concave on [0, t P ], and, viewed as a function of t 0, F is just an affine transformation of A, with a positive coefficient Therefore, F is concave in t 0 on [t, t P ], and min D F = min{f t, t, t +, F t, t P, t + } Since F t, t, t + = 0, to prove 26 we need to show that F t, t P, t + 0 We have F t, t P, t + = t + t At P t + t P At t P t At + [ ] AtP At = t + t P t P t t P t At + At P t P t [ ] AtP At S t + t P t P t t P t At + At P t P t S = t P t F t S, t P, t + t P t S where we used the concavity of A on [0, t P ] To estimate F t S, t P, t +, we write: F t S, t P, t + = F t S, t P, t Q + t + t Q t P t S [ AtP At S t P t S At + At Q t + t Q The first term is non-negative by 26, while the expression in brackets can be estimated by A t P A t Q, which is non-negative by Condition 2 This completes the proof 3 The John Nirenberg inequality Here we prove Theorems 13 and 16 in the reverse order, by choosing the smallest - concave element from a family of locally concave functions developed in [11] We also obtain another significant result, not stated in the introduction: Theorem 34 gives the exact Bellman function for the John Nirenberg inequality on n-dimensional dyadic BMO Recall definitions 13 of ε 0 and K, ε Let g, δ, ε = 1 δ 2 ε 2 e δ+ δ 2 ε 2 K, ε 1 δ Lemma 31 For each 0, 1 2 ] and each ε 0, ε 0 the equation g, δ, ε = 0 has a unique solution δ, ε such that { 31 ε < δ, ε < min 1, 1 + } 2 ε ]

INEQUALITIES FOR BMO ON -TREES 9 Proof For any δ 0, 1, we have g δ = δδ δ 2 ε 2 1 δ 2 e δ+ δ 2 ε 2 > 0 We need to check the sign of g, δ, ε at the two endpoints, δ = ε and δ = min { 1, 1+ 2 ε} For δ = ε we compute: e ε [ ] g, ε, ε = 1 ε 1 e 1 ε e 1 ε e 1 1ε 1 1 ε = e ε 1 ε 1 e 1 ε [ k=3 ε k k! 1 ] k 1 k 1 k/2 < 0 To check the right endpoint, assume first that 0 < ε < 2 1+ Then, after a bit of algebra, we obtain g, 1 + 2 ε, ε = 2 1 ε 2 ε 1 + ε e K, ε = 2 e 1 3 2 ε [ 1 1 1+ 2 ε 1 e 1 ε 2 ε 1 1 ] cosh 2 ε sinh 2 ε > 0 Thus, the interval ε, 1+ 2 ε contains a unique root δ of the equation g, δ, ε = 0 Now, assume that 2 1+ ε < ε 0 Since g, δ, ε as δ 1, we conclude that there exists a unique root δ ε, 1 Let 32 B δ x = e δ 1 δ ex 1+rx 1 rx, where rx = δ 2 x 2 + x 2 1 This family of functions was obtained in [11] As shown there, for every δ 0 B δ is a solution of the Monge Ampère equation B x1 x 1 B x2 x 2 = Bx 2 1 x 2 on Ω ε satisfying B δ x 1, x 2 1 = ex 1 Furthermore, B δ is locally concave on Ω ε In fact, more can be said if δ is sufficiently large For ε > 0 and 0 < 1/2, let δ, ε be given by Lemma 31 Define 33 B,ε x = B δ,ε x, x Ω ε Since ε < δ, ε < 1, we have B,ε C Ω ε Lemma 32 The function B,ε is -concave on Ω ε Proof Let us write simply δ for δ, ε and B for B,ε We have to verify the three conditions of Lemma 25 for B Let µ = δ 2 ε 2 Observe that µ < 1 and that rx = µ for x Γ ε Condition 1: Direct differentiation or Lemma 3c of [11] shows that B is locally concave in Ω ε To verify Condition 2, take any p and q such that 0 < p q 1 ε and let P = p, p 2 + ε 2, Q = q 2, q 2 + ε 2, and ξ = q p/2 We have [ ] Bx = e δ 1 δ ex 1+rx 1 rx x1, Therefore, for some non-negative multiple C, [ D P Q BP D P Q BQ = Cq p 1 2 e p 1 µ + q p 2 P Q = q p [ ] 1 p + q e q 1 µ q p ] 2

10 LEONID SLAVIN AND VASILY VASYUNIN = 4Ce p+q/2 [ ξ 2 cosh ξ 1 µξ sinh ξ ] 4Ce p+q/2 [ ξ 2 cosh ξ ξ sinh ξ ] 0 Lastly, to verify Conidtion 3, take P and Q as above; let S = 1 q p 1 P Q and θ = 1 We have S = 1 1 p q, p2 q 2 +1 ε 2, so rs = δ 2 s 2 + s 2 1 = µ 2 + θ 2 Hence, BP 1 BS BQ = e δ 1 δ ] [e p+µ 1 µ 1 e p q 1 +rs 1 rs e q+µ 1 µ = e δ [ 1 δ eq e µ 1 µe 1 θ 1 e θ+ µ 2 +θ 2 1 ] µ 2 + θ 2 e δ =: 1 δ eq Gθ We now need to show that the function G is non-negative on the domain [ t 0, let ht = e t 1 t An easy computation [ gives G θ = e 1 θ h θ + ] µ 2 + θ 2 hµ ε ] ε, For Since h is a decreasing function, if θ 0, then G θ 0 Since G0 = 0, we conclude that Gθ 0 on [ ε, 0] ε The interval [0, ] requires a little more care To determine the sign of G we need to compare the quantities µ and θ + µ 2 + θ 2 note that the latter is non-negative: θ + µ 2 + θ 2 µ θ 2µ 1 In addition, 2µ 1 = 2 δ 2 ε 2 1 ε this inequality is equivalent to the right-hand inequality in 31 Therefore, G 2µ 0 on [0, 1 ] and G 0 on [ 2µ 1, ε ] Since G0 = 0, it remains to check that G ε 0 In fact, we have G ε = 0, since this is the slightly rewritten equation g, δ, ε = 0 This completes the proof Theorem 16 now follows easily Proof of Theorem 16 Take any ϕ BMO ε T and for N > 0 let ϕ N be the truncation of ϕ at the N -th generation of T, ϕ N = J T N ϕ J,µ χ J Note that ϕ N X,µ = ϕ X,µ and ϕ 2 N X,µ ϕ2 X,µ Since B,ε is -concave on Ω ε and B,ε x 1, x 2 1 = ex 1, we can apply Lemma 24: B,ε ϕn X,µ, ϕ 2 1 N X,µ B,ε ϕ µx N, ϕ 2 N dµ = 1 e ϕ N dµ µx X Since the family T differentiates L 1 X, µ, ϕ N ϕ µ-ae as N Thus, ϕ 2 N X,µ ϕ 2 X,µ, by Fatou s Lemma Because B is continuous on Ω ε, the left-hand side tends to B,ε ϕ X,µ, ϕ 2 X,µ Hence, again by Fatou s Lemma, e ϕ is integrable and 34 B,ε ϕ X,µ, ϕ 2 X,µ e ϕ X,µ For a fixed x 1, B attains its maximum on Γ ε, thus the left-hand side is bounded by e ϕ 1 δ 2 ε 2 X,µ e δ+ δ 2 ε 2 = e ϕ X,µ K, ε 1 δ This proves 14 with J = X The full result follows, since the set T J = {I T : I J} is itself an -tree and ϕ BMOT J ϕ BMOT X

INEQUALITIES FOR BMO ON -TREES 11 Observe that 34 gives more than the John Nirenberg inequality 14: with ϕ BMOT fixed, we get a continuum of more precise inequalities indexed by ϕ X,µ, ϕ 2 X,µ In the dyadic case, all these inequalities turn out to be sharp We can state this compactly using the concept of a Bellman function Take ε > 0 and a dyadic cube Q R n For every x Ω ε, let B n ε x = sup ϕ BMO d ε Q { e ϕ Q : ϕ Q = x 1, ϕ 2 Q = x 2 } The function B n ε is called the Bellman function for the integral John Nirenberg inequality for BMO d R n It is easy to see that the supremum above is taken over a non-empty set for every x Ω ε By rescaling, B n ε can be seen not to depend on Q One of the main results of [11] was the following theorem Theorem 33 [11] If ε < 2 log 2, then If ε 2 log 2, then B 1 εx = B 1 2,εx, x Ω ε B 1 εx = We now have the following theorem { e x 1, x Γ 0,, x Ω ε \ Γ 0 Theorem 34 Take n 1 If ε < ε d 0 n, then B n ε x = B 2 n,εx, x Ω ε If ε ε d 0 n, then B n ε x = { e x 1, x Γ 0,, x Ω ε \ Γ 0 A key role in the proof of this theorem is played by the following function from BMO d 0, 1 n Let Q = 0, 1 n and Q k = 0, 2 k n for all k 0 Now, let 35 ϕ t = 2 n/2 k2 n 1 1, t Q k \ Q k+1 Lemma 35 We have ϕ Q = 0, ϕ 2 Q = 1, and ϕ BMO d Q with ϕ BMO d Q = 1 Proof It is clear that to verify that ϕ BMO d Q, and to compute the norm, it suffices to check the average oscillations of ϕ only over the cubes Q k We have ϕ Qk = 2 nk 1/2 2 nj 2 nj+1 j2 n 1 1 and j=k = 2 n/2 2 n 1k ϕ 2 Qk = 2 nk 2 2 n 1 2 nj j2 n 1 1 2 j=k = 1 + 2 n 2 n 1 2 k 2 Setting k = 0, we obtain ϕ Q = 0 and ϕ 2 Q = 1 Furthermore, for any k, ϕ 2 Qk ϕ 2 = 1, which means that ϕ Q BMO d Q and ϕ k BMO d Q = 1, as claimed

12 LEONID SLAVIN AND VASILY VASYUNIN Proof of Theorem 34 Take any ε 0, ε d 0 n and let us write B for Bn ε, B for B 2 n,ε, and δ for δ2 n, ε Fix a point x Ω ε and pick any ϕ BMO d Q such that ϕ Q, ϕ 2 Q = x By 34, Bx e ϕ Q Taking the supremum over all such ϕ, we conclude that Bx Bx To show the converse, we make use of the function ϕ given by 35 Take any a R and consider the corresponding point on Γ ε, P a = a, a 2 + ε 2 Let ϕ a = εϕ + a Then ϕ a Q, ϕ 2 a Q = P a and ϕ a BMO d Q = ε Furthermore, e ϕa Q = e a ε2n/2 1 2 n = e ε2n/2 1 2 n 2 nj e ε2 n/2 2 n 1j j=0 j=0 2 n e ε2 n/2 2 n 1 j This sum converges if and only if ε < 2n/2 2 n 1 n log 2 = εd 0 n, in which case e ϕa Q = e a ε2n/2 1 2 n 1 1 2 n e = ε2 n/2 2 n 1 ea K2 n, ε = BP a Therefore, BP a BP a Thus, B = B on Γ ε Similarly, by considering the constant function ψ u = u corresponding to a point u, u 2 Γ 0, we conclude that B = B on Γ 0 Now, as shown in [11], for any number c, the function B is linear along the line l c with the equation x 2 = 2cx 1 + δ 2 c 2 ; such lines are tangent to Γ δ and they foliate Ω ε On the other hand, it is easy to show that B is locally concave in Ω ε this is done is [11] for n = 1 Take any point x Ω ε and let l be the unique tangent to Γ δ passing through x Then l intersects Γ 0 and Γ ε at some points U and V, respectively, such that the segment [U, V ] lies entirely in Ω ε We can write x as a convex combination of U and V : x = 1 γu + γv for some γ [0, 1] Then, Bx 1 γ BU + γ BV = 1 γ BU + γ BV = Bx Thus B B everywhere in Ω ε Now take ε ε d 0 n If x Γ 0, then Bx = e x 1 Indeed, if ϕ 2 Q = ϕ 2, then ϕ is ae Q constant on Q, and so e ϕ Q = e ϕ Q On the other hand, e ϕa Q =, which means that B = on Γ ε It is now an easy exercise to show that B = everywhere in Ω ε \ Γ 0 Theorem 13 now follows Proof of Theorem 13 Note that this theorem is stated for BMO d R n, while both Theorem 16 and Theorem 34 deal with BMO on sets of finite measure However, the difference is inconsequential Indeed, take any ϕ BMO d R n and let ε = ϕ BMO d R n Assume that ε < ε d 0 n Take any Q D Then ϕ BMO d Q ε, thus by Theorem 16, 12 holds To prove sharpness, consider again the function ϕ from 35 and extend it periodically to all of R n, by replicating it on every dyadic cube of measure 1 As before, for any a R, let ϕ a = εϕ + a Then ϕ a BMO d R n = ε ϕ BMO d 0,1 n = ε Moreover, if ε < εd 0 n, then inequality 12 becomes equality for Q = 0, 1 n and ϕ = ϕ a On the other hand, if ε ε d 0 n, we have eϕa Q =

INEQUALITIES FOR BMO ON -TREES 13 4 1-oscillations vs 2-oscillations and equivalence of BMO norms In this section, we illustrate how one can obtain lower bounds for integral functionals on BMO on trees using -convex functions A representative example is supplied by the task of sharply estimating the p-oscillation of a BMO function, p,µ,j ϕ, for p < 2, in terms of its 2-oscillation 2,µ,J ϕ and the BMO norm In the case of the continuous BMO on an interval, this problem was solved in [12], where we built a family of locally convex functions on Ω ε that yielded sharp estimates of L p -norms and, consequently, of p-oscillations Here, we consider the simplest non-trivial case, p = 1, and construct the largest -convex analog of the corresponding function from [12] This allows us to prove Theorems 17 and 18 As in Section 3, there is an additional result of interest: Theorem 42 gives the exact Bellman function for the L 1 -norm of a BMO d function in dimension n Take an ε > 0 and 0, 1/2] and write Ω ε = Ω 0 Ω 1, where Let { Ω 0 = x : Ω 1 = 1+ε x 1 x 2 x 2 1 + ε2 }, { x : x 1 ε, x 2 1 x 2 1+ε x 1 41 b,ε x = } {x : x 1 > } ε, x 2 1 x 2 x 2 1 + ε2 { 1+ε x 2, x Ω 0, x 1, x Ω 1 We will show that b,ε is -convex in Ω ε, which, in conjunction with Lemma 24, will yield the estimates of Theorem 18 We need one more definition: let { } Ω 2 = Ω 0 x Ω 1 : x 2 1+ε x 1 Observe that if x Ω 2, then b,ε x = max { x 1, and x 2 > 1 + ε 2 / Lemma 41 The function b,ε is -convex in Ω ε 1+ε x 2} ; if x Ωε \Ω 2, then x 1 > ε/ Proof Write b for b,ε Because b is piecewise linear, we can easily verify part 22 of Definition 22 directly, without using Lemma 25 Take any β [, 1/2] and any three points P, Q, S Ω ε such that P = 1 βs + βq If P Ω 1, then bp 1 β bs β bq p 1 1 β s 1 β q 1 0, because bx x 1 in Ω ε and x 1 is a convex function on the plane If P Ω 0, then p 1 ε and p 2 1 + ε 2 Thus, s 2 = p 2 βq 2 1 β 1 + 1 β ε2 1 + ε2, since 1 β 1/2 Therefore, S Ω 2 On the other hand, 1 βs 2 = p 2 βq 2 p 2 q 2 1 + ε 2 q 2 Thus, q 2 1+ε2, which means that Q Ω 2 Finally, since bx 1+ε x 2 in Ω 2, we have bp 1 β bs β bq 1+ε p 2 1 β s 2 β q 2 = 0 This completes the proof We can now prove Theorem 18

14 LEONID SLAVIN AND VASILY VASYUNIN Proof of Theorem 18 As in the proof of Theorem 16, take any ϕ BMO ε T and for N > 0 let ϕ N = J T N ϕ J,µ χ J Since b,ε is -convex in Ω ε, and b,ε x = x 1 on Γ 0, Lemma 24 gives b,ε ϕn X,µ, ϕ 2 1 N X,µ b,ε ϕ µx N, ϕ 2 N dµ = 1 ϕ X µx N dµ X Note that ϕ N ϕ N, thus the right-hand side does not exceed ϕ N X,µ = ϕ X,µ Moreover, as in Lemma 24, the left-hand side converges to b,ε ϕ X,µ, ϕ 2 X,µ as N Thus, 42 b,ε ϕ X,µ, ϕ 2 X,µ ϕ X,µ Recall definition 41 of b,ε Replacing ϕ with ϕ ϕ X,µ which does not affect the BMO norm of ϕ, we obtain b,ε 0, ϕ 2 X,µ ϕ 2 X,µ = 1 + ε 2,µ,Xϕ 1,µ,X ϕ, which proves 17 with J = X As before, by considering the tree T J in place of T we get the result for an arbitrary J T Thus, 43 1 + ε 2,µ,Jϕ 1,µ,J ϕ ϕ BMO 1 T To prove 18, take any ϕ BMOT and let ε = ϕ BMOT If ε = 0, 18 holds automatically; if ε > 0, we have a sequence {J k } of elements of T such that 2,µ,Jk ϕ ε 2 as k Replacing J with J k in 43 and taking the limit completes the proof In the dyadic case, we again state the corresponding result using a Bellman function In contrast with the John Nirenberg inequality, we are proving a sharp lower estimate, thus our Bellman function involves infimum instead of supremum Take a dyadic cube Q R n and for every x Ω ε, let b n ε x = Theorem 42 For any ε > 0, inf ϕ BMO d εq { ϕ Q : ϕ I = x 1, ϕ 2 I = x 2 } b n ε = b 2 n,ε Proof The proof is similar to that of Theorem 34 Write b for b n ε and b for b 2 n,ε Take any x Ω ε and any ϕ BMO d εq such that ϕ Q, ϕ 2 Q = x As shown in the proof of Theorem 18 equation 42 with X = Q and µ the Lebesgue measure, bx ϕ Q Taking the infimum over all such ϕ, we conclude that b b in Ω ε To prove the converse, we again let Q = 0, 1 n and use the function ϕ from 35 For all a such that a 2 n/2 ε, let ϕ a = εϕ + a Recall that ϕ a BMO d Q = ε and that ϕ a Q, ϕ 2 a Q = a, a 2 + ε 2 Furthermore, ϕ a is almost everywhere positive for a 2 n/2 ε and almost everywhere negative for a 2 n/2 ε, thus ba, a 2 + ε 2 ϕ a Q = ϕ a Q = a = ba, a 2 + ε 2 Hence, b = b on Γ ε Ω 1 On the other hand, by considering the constant function ψ u = u corresponding to a point u, u 2 Γ 0 we conclude that b = b on Γ 0 Take any x Ω 1 and let l x be any line that passes through x, a point U on Γ 0, and a point V on Γ ε, and such that the segment [U, V ] Ω 1 there are infinitely many such lines

INEQUALITIES FOR BMO ON -TREES 15 for each x Write x = 1 γu + γv for some γ [0, 1] Note that b is linear along [U, V ], while b is a locally convex function directly from its definition Therefore, 44 bx 1 γ bu + γ bv = 1 γ bu + γ bv = bx Thus, b = b in Ω 1 Now, take any x Ω 0 and write x a convex combination of the three corner points of Ω 0, 0, 0, 2 n/2 ε, 1 + 2 n ε 2, and 2 n/2 ε, 1 + 2 n ε 2 Since at these three points b = b, b is linear in Ω 0, and b is locally convex, arguing as in 44 proves that b b in Ω 0 Putting everything together, we conclude that b = b in Ω ε Proof of Theorem 17 Arguing as in the proof of Theorem 13, we obtain 15 and 16 from Theorem 18 To prove sharpness of 15, we do not present an explicit function as we did in Theorem 13 Instead, take any ε > 0 and any Q D Theorem 42 says that b n ε 0, ε 2 = 2n/2 2 n +1 ε Therefore, there exists a sequence of functions {ϕ k} from BMO d Q such that ϕ k Q = 0, ϕ k BMO d Q = ϕ2 k 1/2 = ε, and Q lim ϕ k ϕ k k Q Q = lim ϕ k k Q = 2n/2 2 n + 1 ε Extending each ϕ k periodically to all of R n we get ϕ k BMO d R n = ε, which completes the proof 5 -martingales and continuous BMO Any metric measure space has a canonical BMO defined on balls; we call this the continuous BMO If the space is equipped with an -tree T, one also has BMOT It is natural to ask whether our theory for BMO on trees can be used to obtain sharp inequalities for the continuous BMO If the tree in question is fixed from the outset, the answer is no; however, if one allows each BMO function to generate its own tree and, thus, a martingale, then the answer is yes, provided all such martingales are regular in a particular uniform sense To focus the discussion, consider the classical BMO on Euclidean cubes the continuous BMOR n in the L -metric Take a cube Q in R n and, as in 11, let BMOQ = { ϕ L 1 Q: ϕ BMOQ := sup cube J Q 2,J 1/2 < } We first give a geometric definition of an -concave function that works for the full range 0, 1]; -convex functions are defined symmetrically The reader can easily verify that this definition is equivalent to Definition 22 for 0, 1/2] Definition 51 Take ε > 0 and 0, 1] Let P and R be two distinct points from Ω ε The segment [P, R] is called -good, if the length of the portion of [P, R] that lies outside Ω ε does not exceed 1 [P, R] A function B on Ωε is called -concave, if for each -good segment [P, R] and each Q [P, R] Ω ε we have 51 BQ [Q, R] [P, R] BP + [P, Q] [P, R] BR Observe that the larger the, the weaker the condition of -concavity, and 1-concavity is the same as local concavity We now introduce special trees and martingales adapted to those trees Definition 52 Let Q be a cube in R n A disjoint collection T of measurable subsets of Q is called a binary tree on Q if

16 LEONID SLAVIN AND VASILY VASYUNIN 1 For each J T, J = J J +, with J, J + T ; 2 T = T k, where T 0 = {Q} and T k+1 = J Tk {J, J + }; 3 For all J T, J > 0, and lim k max J Tk J = 0 Definition 53 Let T be a binary tree on a cube Q, 0, 1], and ε > 0 Let {x J } J T be a collection of points from Ω ε such that for any J, J x J = J x J + J + x J + Let {f k } be the corresponding sequence of Ω ε -valued simple functions on Q given by f k = J T k x J χ J Assume that for each J T, either x J = x J + or the segment [x J, x J +] is -good Then f k is called an, ε-martingale adapted to the tree T Furthermore, if ϕ L 2 Q is such that for all k and all J T k, ϕ J, ϕ 2 J = f k J, then we say that ϕ generates f k Lastly, an -martingale is an, ϕ BMOQ -martingale generated by a function ϕ BMOQ Observe that by Lemma 23, as used in the proof of Lemma 24, any function ϕ BMO d Q generates a 2 n -martingale Combining the notions of -concave functions and -martingales gives the following general result, which can be viewed as a special form of Jensen s inequality The first part is contained in Lemma 24 and the second part follows from Fatou s Lemma, as in the proof of Theorem 16 Lemma 54 Let Q be a cube in R n, ε > 0, and 0, 1] Let B be an -concave function on Ω ε If f k is an, ε-martingale on Q, then for any k 0 Bf 0 1 Q Q Bf k Consequently, if ϕ L 2 Q generates an, ε-martingale and B is continuous on Ω ε, then B ϕ Q, ϕ 2 Q 1 Bϕ, ϕ 2 Q Let us illustrate this lemma in the case of the John Nirenberg inequality To state our result, we need two new parameters Take a cube Q Let { } ε 0 n = sup ε > 0 : ϕ BMOQ, sup e εϕ ϕ J / ϕ BMOQ J < cube J Q and 0 n = sup{ 0, 1] : every ϕ BMOQ generates an -martingale on Q} Of course, neither ε 0 n nor 0 n actually depends on Q Observe that both suprema are taken over non-empty sets: since every ϕ BMOQ is also in BMO d Q with at most the same norm, we have ε 0 n ε d 2n/2 0 n = 2 n 1 n log 2, and 0n 2 n Recall the function B,ε given by 33 Lemma 32 asserts that B,ε is -concave on Ω ε for 0, 1/2] In fact, the same proof also works for 1/2, 1] Using this function in Lemma 54 gives the following corollary Corollary 55 52 ε 0 n Q 0 n 1 1 0 n log 0 n

INEQUALITIES FOR BMO ON -TREES 17 A natural conjecture is that 52 holds with equality It does when n = 1, which is the only case where we currently know 0 n To explain: the main geometric result that underpins the now-well-developed Bellman theory for BMO in dimension 1 is Lemma 4c of [11] For an interval I, any δ 0, 1, and any ϕ BMOI, that lemma gives an explicit construction of a 1 δ-martingale generated by ϕ Therefore, 0 1 = 1 and so ε 0 1 1; a simple logarithmic example shows that ε 0 1 = 1 In higher dimensions we do not yet know either 0 n or ε 0 n However, in our view inequality 52 expresses the right idea: to find ε 0 n, first understand the nature of - martingales generated by BMO functions The trees that correspond to the best such martingales the ones with the largest may not be given by the procedure of Lemma 24, where on each step we split off a cube from a larger set Indeed, consider the function on the unit square Q = 0, 1 2 that equals zero on two of the four quarters of Q and ± 2 on the other two This function has BMO norm 1 If one builds a binary martingale by first splitting off one quarter-square of the four, then one of the remaining three, then splitting the last two, one obtains a 1/2-martingale If, however, one first splits Q into two halves, each containing one quarter on which the function is zero, one obtains a 1-martingale References [1] D L Burkholder Boundary value problems and sharp inequalities for martingale transforms Annals of Probability, Vol 12 1984, No 3, pp 647 702 [2] M Christ A T b theorem with remarks on analytic capacity and the Cauchy integral Colloq Math, Vol 60/61 1990, No 2, pp 601-628 [3] F John, L Nirenberg On functions of bounded mean oscillation Comm Pure Appl Math, Vol 14 1961, pp 415-426 [4] P Ivanishvili, N Osipov, D Stolyarov, V Vasyunin, P Zatitskiy, On Bellman function for extremal problems in BMO C R Math Acad Sci Paris 350 2012, No 11-12, pp 561-564 [5] P Ivanishvili, N Osipov, D Stolyarov, VVasyunin, P Zatitskiy Bellman function for extremal problems in BMO To appear in Transactions of the AMS, pp 1-91, http://arxivorg/pdf/12057018pdf [6] A Melas Dyadic-like maximal operators on L log L functions J Funct Anal, Vol 257 2009, No 6, pp 1631-1654 [7] A Melas, E Nikolidakis, T Stavropoulos Sharp local lower L p -bounds for dyadic-like maximal operators Proc Amer Math Soc, Vol 141 2013, No 9, pp 3171-3181 [8] F Nazarov, S Treil The hunt for Bellman function: applications to estimates of singular integral operators and to other classical problems in harmonic analysis RussianAlgebra i Analiz, Vol 8 1996, No 5, pp 32-162; English translation in St Petersburg Math J, Vol 8 1997, No 5, pp 721-824 [9] F Nazarov, S Treil, A Volberg The Bellman functions and two-weight inequalities for Haar multipliers 1995, Preprint, MSU, pp 1-25 [10] F Nazarov, S Treil, A Volberg The Bellman functions and two-weight inequalities for Haar multipliers Journal of the American Mathematical Society, Vol 12 1999, No 4, pp 909-928 [11] L Slavin, V Vasyunin Sharp results in the integral-form John Nirenberg inequality Trans Amer Math Soc, Vol 363 2011, No 8, pp 4135-4169 [12] L Slavin, V Vasyunin Sharp L p estimates on BMO norms Indiana Univ Math J, Vol 61 2012, No 3, pp 1051-1110 [13] V Vasyunin, A Volberg, Sharp constants in the classical weak form of the John Nirenberg inequality Proc Lond Math Soc 3, Vol 108 2014, No 6, pp 1417-1434 University of Cincinnati E-mail address: leonidslavin@ucedu St Petersburg Department of the V A Steklov Mathematical Institute, RAS E-mail address: vasyunin@pdmirasru