This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

On locally concave functions on simplest non-convex domains thanks: Supported by the Russian Science Foundation grant 19-71-10023.

Dmitriy Stolyarov    Pavel Zatitskiy
Abstract

We prove that certain Bellman functions of several variables are the minimal locally concave functions. This generalizes earlier results about Bellman functions of two variables.

1 Introduction

The aim of the present paper is to extend the theory of [21]. The main result of that article said that certain Bellman functions of two variables coincide with minimal locally concave functions in the case when their domain is a set theoretic difference of two unbounded convex sets, the smaller lying strictly inside the larger one. The proof relied upon a special class of 2\mathbb{R}^{2}-valued martingales and the notion of monotonic rearrangement. We improve these results in two directions: we allow our Bellman functions to depend on more than two variables and also work with the case when the domain is bounded (and therefore, not simply connected in dimension 2). While the special martingales work perfectly in this setting, the notion of monotonic rearrangement is, seemingly, not applicable. We substitute it with the notion of homogenization of a function from [20].

The work is technical: we mostly combine the ideas and methods of two cited papers. Our main results are Theorems 4.4 and 5.3. Sections 2, 3, 4, and 5 contain definitions, examples, descriptions of previous development of the theory, and statements of the results. Sections 6, 7, and 8 contain the proofs. We also place two auxiliary results in Section 9.

We wish to thank Vasily Vasyunin for his attention to our work.

2 Classes of functions

Let Ω0\Omega_{0} be a non-empty proper open convex subset of d\mathbb{R}^{d}, here dd is a natural number. Usually d2d\geqslant 2. Let Ω1\Omega_{1} be another open convex set such that clΩ1Ω0\operatorname{cl}\Omega_{1}\subset\Omega_{0}. It will be convenient to use the notation

Ω=clΩ0Ω1.{\Omega=\operatorname{cl}\Omega_{0}\setminus\Omega_{1}.} (2.1)

We assume Ω1\Omega_{1} is non-empty for convenience (the case of empty Ω1\Omega_{1} may be considered by means of classical convex geometry). Let I\operatorname{I}\subset\mathbb{R} be an interval. Consider the class of d\mathbb{R}^{d} valued summable functions φ\varphi defined by the domains Ω0\Omega_{0} and Ω1\Omega_{1}:

𝑨={φ:IΩ0|Jsubinterval ofIφJΩ1}.{\boldsymbol{A}=\Big{\{}{\varphi\colon\operatorname{I}\to\partial\Omega_{0}}\;\Big{|}\,{\forall\ \operatorname{J}\ \text{subinterval of}\ \operatorname{I}\qquad\langle{\varphi}\rangle_{{}_{\operatorname{J}}}\notin\Omega_{1}}\Big{\}}.} (2.2)

Here and in what follows we use the notation

φE=1|E|Eφ(x)𝑑x{\langle{\varphi}\rangle_{{}_{E}}=\frac{1}{|E|}\int\limits_{E}\varphi(x)\,dx} (2.3)

for the average of a summable function φ\varphi over a measurable set EE whose Lebesgue measure satisfies the requirement 0<|E|<0<|E|<\infty. Sometimes we will call the points φJ\langle{\varphi}\rangle_{{}_{\operatorname{J}}}J\operatorname{J} being a subinterval of I\operatorname{I}, the Bellman points of φ\varphi. Now we will show how several useful classes of functions may be described using particular choices of Ω0\Omega_{0} and Ω1\Omega_{1}.

Muckenhoupt classes.

Let d=2d=2 and δ>1\delta>1. We pick particular domains

Ω0={(x,y)2|x0,y0,xy>1};\displaystyle\Omega_{0}=\{{(x,y)\in\mathbb{R}^{2}}\;|\,{x\geqslant 0,y\geqslant 0,\quad xy>1}\}; (2.4)
Ω1={(x,y)2|x0,y0,xy>δ}.\displaystyle\Omega_{1}=\{{(x,y)\in\mathbb{R}^{2}}\;|\,{x\geqslant 0,y\geqslant 0,\quad xy>\delta}\}.

See Fig. 1 for visualization. Consider the class 𝑨\boldsymbol{A} generated by these domains and a function φ𝑨\varphi\in\boldsymbol{A}. Let ww be the first coordinate of φ\varphi, i. e., φ(t)=(w(t),w1(t))\varphi(t)=(w(t),w^{-1}(t)) and w:I+w\colon\operatorname{I}\to\mathbb{R}_{+} is a scalar almost everywhere positive function. The condition φJΩ1\langle{\varphi}\rangle_{{}_{\operatorname{J}}}\notin\Omega_{1} in the definition (2.2) is rewritten in terms of ww as

wJw1Jδ.{\langle{w}\rangle_{{}_{\operatorname{J}}}\langle{w^{-1}}\rangle_{{}_{\operatorname{J}}}\leqslant\delta.} (2.5)

By our definition, this condition is fulfilled for any interval JI\operatorname{J}\subset\operatorname{I} by the requirement φ𝑨\varphi\in\boldsymbol{A}. Therefore, [w]A2δ[w]_{A_{2}}\leqslant\delta (see Chapter V in [18] for definition and basic properties of the Muckenhoupt classes ApA_{p}). More specifically, we have proved a simple lemma.

Lemma 2.1.

The condition [w]A2δ[w]_{A_{2}}\leqslant\delta is equivalent to φ𝐀\varphi\in\boldsymbol{A}, where the domains Ω0\Omega_{0} and Ω1\Omega_{1} are defined in (2.4) and φ=(w,w1)\varphi=(w,w^{-1}).

One may prove a similar statement for ApA_{p} classes when 1<p<1<p<\infty and for AA_{\infty} as well, provided the latter class is equipped with Hruschev’s norm. The only difference is that one should replace the expression xyxy in (2.4) with xyp1xy^{p-1} (and xeyxe^{-y} in the case p=p=\infty, see [24] for details). See Section 22 in [8] or Subsection 1.31.3 in [21] for more information.

Refer to caption
Refer to caption
Refer to caption
Figure 1: The domains corresponding to A2A_{2}, (2.4); scalar-valued BMOε\mathrm{BMO}_{\varepsilon}, (2.7); and BMOε(I,S1)\mathrm{BMO}_{\varepsilon}(\operatorname{I},S^{1}), (2.12).

Spaces BMO\mathrm{BMO} of vector-valued functions.

Let dd be an arbitrary natural number larger than one and let ε>0\varepsilon>0. The notation |z||z| means the Euclidean norm of zd1z\in\mathbb{R}^{d-1}:

|z|=(j=1d1zj2)12.{|z|=\Big{(}\sum\limits_{j=1}^{d-1}z_{j}^{2}\Big{)}^{\frac{1}{2}}.} (2.6)

Consider the case

Ω0\displaystyle\Omega_{0} ={(x,y)d1×+|y>|x|2};\displaystyle=\{{(x,y)\in\mathbb{R}^{d-1}\times\mathbb{R}_{+}}\;|\,{y>|x|^{2}}\}; (2.7)
Ω1\displaystyle\Omega_{1} ={(x,y)d1×+|y>|x|2+ε2}.\displaystyle=\{{(x,y)\in\mathbb{R}^{d-1}\times\mathbb{R}_{+}}\;|\,{y>|x|^{2}+\varepsilon^{2}}\}.

Let ψ(t)\psi(t) be the vector in d1\mathbb{R}^{d-1} formed by the first (d1)(d-1) coordinates of φ(t)\varphi(t), where φ𝑨\varphi\in\boldsymbol{A}, i. e., φ=(ψ,|ψ|2)\varphi=(\psi,|\psi|^{2}). Then, condition (2.2) turns into

|ψ|2J|ψJ|2+ε2,{\langle{|\psi|^{2}}\rangle_{{}_{\operatorname{J}}}\leqslant|\langle{\psi}\rangle_{{}_{\operatorname{J}}}|^{2}+\varepsilon^{2},} (2.8)

which may be rewritten as

|ψψJ|2Jε2.{\langle{|\psi-\langle{\psi}\rangle_{{}_{\operatorname{J}}}|^{2}}\rangle_{{}_{\operatorname{J}}}\leqslant\varepsilon^{2}.} (2.9)

Since the requirement φ𝑨\varphi\in\boldsymbol{A} means the above inequality holds true for any interval JI\operatorname{J}\subset\operatorname{I}, we have

ψBMO(I)ε,{\|\psi\|_{\mathrm{BMO}(\operatorname{I})}\leqslant\varepsilon,} (2.10)

provided we define the BMO(I)\mathrm{BMO}(\operatorname{I}) norm of a vectorial function by the rule

ψBMO(I)=(supJI1|J|J|ψ(t)1|J|Jψ(s)𝑑s|2𝑑t)12,{\|\psi\|_{\mathrm{BMO}(\operatorname{I})}=\Big{(}\sup\limits_{\operatorname{J}\subset\operatorname{I}}\frac{1}{|\operatorname{J}|}\int\limits_{\operatorname{J}}\Big{|}\psi(t)-\frac{1}{|\operatorname{J}|}\int\limits_{\operatorname{J}}\psi(s)\,ds\Big{|}^{2}\,dt\Big{)}^{\frac{1}{2}},} (2.11)

where the supremum is taken over all subintervals of I\operatorname{I}. We refer the reader to Chapter IV of [18] for the definition and basic properties of the BMO\mathrm{BMO} space of scalar functions; the quantitative properties of vectorial BMO\mathrm{BMO} functions are almost the same as that of scalar functions. With this definition at hand, we state yet another simple lemma.

Lemma 2.2.

The condition ψBMO(I)ε\|\psi\|_{\mathrm{BMO}(\operatorname{I})}\leqslant\varepsilon is equivalent to φ𝐀\varphi\in\boldsymbol{A}, where the domains are given in (2.7) and φ=(ψ,|ψ|2)\varphi=(\psi,|\psi|^{2}).

Note that we use the quadratic norm on BMO\mathrm{BMO} in (2.11). Usually, the definition of BMO\mathrm{BMO} is given with the more widespread L1L_{1}-based norm and after that it is proved via the John–Nirenberg inequality that the two norms are equivalent. Since we will be working with sharp constants, the choice of the particular norm is crucial.

Functions of bounded mean oscillation with values in the unit sphere.

Let d2d\geqslant 2 and let ε(0,1)\varepsilon\in(0,1). Consider the case

Ω0\displaystyle\Omega_{0} ={xd||x|<1};\displaystyle=\{{x\in\mathbb{R}^{d}}\;|\,{|x|<1}\}; (2.12)
Ω1\displaystyle\Omega_{1} ={xd||x|2<1ε2}.\displaystyle=\{{x\in\mathbb{R}^{d}}\;|\,{|x|^{2}<1-\varepsilon^{2}}\}.

Here and in what follows we use the Euclidean norms in d\mathbb{R}^{d}. We see that the functions φ𝑨\varphi\in\boldsymbol{A} attain values in the unit sphere Sd1S^{d-1}. Computations similar to those we did in the case of BMO\mathrm{BMO} functions lead to the following lemma.

Lemma 2.3.

Let φ:ISd1\varphi\colon\operatorname{I}\to S^{d-1} be a summable function. The condition φBMOε\|\varphi\|_{\mathrm{BMO}}\leqslant\varepsilon is equivalent to φ𝐀\varphi\in\boldsymbol{A}, where the domains are given in (2.12).

Following [3], we will call the class of spherically-valued functions whose BMO\mathrm{BMO} norm does not exceed ε\varepsilon the ε\varepsilon-ball of the space BMO(I,Sd1)\mathrm{BMO}(\operatorname{I},S^{d-1}) and denote it by BMOε(I,Sd1)\mathrm{BMO}_{\varepsilon}(\operatorname{I},S^{d-1}). Note that BMO(I,Sd1)\mathrm{BMO}(\operatorname{I},S^{d-1}) does not have linear structure.

Domain related to multiplicative inequalities.

Here we set d=3d=3. Pick some p(1,)p\in(1,\infty) and ε>0\varepsilon>0. Consider the domains

Ω0=intconv{(t,t2,|t|p)|t};\displaystyle\Omega_{0}=\operatorname{int}\operatorname{conv}\{{(t,t^{2},|t|^{p})}\;|\,{t\in\mathbb{R}}\}; (2.13)
Ω1={(x,y,z)3|y>x2+ε2,z>0}.\displaystyle\Omega_{1}=\{{(x,y,z)\in\mathbb{R}^{3}}\;|\,{y>x^{2}+\varepsilon^{2},z>0}\}.

The notation conv\operatorname{conv} designates the convex hull of a set. Note that these domains do not fulfill the requirement clΩ1Ω0\operatorname{cl}\Omega_{1}\subset\Omega_{0}. They appeared naturally in the study of multiplicative inequalities involving the BMO\mathrm{BMO} norm in [19] and [23]. In fact, the class 𝑨\boldsymbol{A} corresponds to the ε\varepsilon-ball of the BMO\mathrm{BMO} space. The additional third coordinate allows to keep track of the LpL_{p} norm. This example will be mostly used to show the limitation of our current tools.

Refer to caption
Refer to caption
Figure 2: The first domain satisfies the cone condition whereas the second domain does not.

We need to make further assumptions about the domains Ω0\Omega_{0} and Ω1\Omega_{1}. The following two conditions appear naturally in the theory, in particular, the reader may find them in [21].

(2.14)

Recall that a convex set is strictly convex provided its boundary does not contain linear segments. Equivalently, a set is strictly convex if and only if any point of its boundary is an exposed point, i. e., it is the unique point of intersection of the boundary with a supporting hyperplane. The second assumption somehow says Ω0\Omega_{0} and Ω1\Omega_{1} behave in a similar way at infinity. It may be restated: Ω0\Omega_{0} and Ω1\Omega_{1} have congruent maximal inscribed cones; this assumption is meaningless if Ω0\Omega_{0} is bounded. The domains on Fig. 1 satisfy conditions (LABEL:StrictConvexity) and (2) because the corresponding domains Ω\Omega defined in (2.1) do not contain infinite rays. The domain between two shifted hyperbolas (see Fig. 2),

Ω={(x,y)2|x,y>0,xy>1,andy<1x1+1whenx>1}{\Omega=\Big{\{}{(x,y)\in\mathbb{R}^{2}}\;\Big{|}\,{x,y>0,\quad xy>1,\ \text{and}\ y<\frac{1}{x-1}+1\ \text{when}\ x>1}\Big{\}}} (2.15)

contains infinite rays, e.g., the ones parallel to the coordinate axes. It still satisfies condition (2). The second domain on Fig. 2,

Ω={(x,y)2|x2+1yx2+2},{\Omega=\Big{\{}{(x,y)\in\mathbb{R}^{2}}\;\Big{|}\,{\sqrt{x^{2}+1}\leqslant y\leqslant x^{2}+2}\Big{\}},} (2.16)

does not satisfy the cone condition.

The domains of this type (i. e., a set theoretic difference of two strictly convex sets, the smaller one lying strictly inside the larger one, and such that (2) holds true) will be informally called lenses. In the proof of the following lemma we will use the notation [A,B][A,B] to denote the straight line segment that connects AA with BB.

Lemma 2.4.

Assume the domains Ω0\Omega_{0} and Ω1\Omega_{1} satisfy (LABEL:StrictConvexity). In the case d=2d=2 we additionally assume (2). Then, for any xΩx\in\Omega there exists a segment xΩ\ell_{x}\subset\Omega passing through xx and whose endpoints lie on Ω0\partial\Omega_{0}.

Refer to caption
Figure 3: Illustration to the proof of Lemma 2.4.
Proof.

Consider the case d=2d=2 first. Let x0x_{0} be the closest to xx point in clΩ1\operatorname{cl}\Omega_{1}, it is clear that x0Ω1x_{0}\notin\Omega_{1}. Let ll be a line passing through x0x_{0} that is perpendicular to [x,x0][x,x_{0}] (if x=x0Ω1x=x_{0}\in\partial\Omega_{1}, we take ll to be a supporting line to clΩ1\operatorname{cl}\Omega_{1} at x0x_{0}). Note that ll does not intersect Ω1\Omega_{1} and separates xx from Ω1\Omega_{1}. By (LABEL:StrictConvexity), the intersection of any translate of ll with Ω1\Omega_{1} is bounded. Thus, by (2), the intersection of the translate of ll passing through xx with Ω0\Omega_{0} is a finite segment, let it be x\ell_{x}. It remains to note that xΩ1=\ell_{x}\cap\Omega_{1}=\varnothing.

Now let us turn to the case d3d\geqslant 3. We will argue by induction over dimension. In fact, it suffices to show that for any xΩx\in\Omega there exists an affine hyperplane LL passing through xx such that LΩ0L\cap\Omega_{0} is bounded. If this assertion is proved, we may work inside the (d1)(d-1) dimensional plane LL (condition (2) holds true for bounded domains). The desired hyperplane LL is also easy to find: pick any supporting hyperplane to Ω0\Omega_{0} and translate it to xx (the intersection of this translated supported hyperplane and Ω0\Omega_{0} is compact by (LABEL:StrictConvexity)). ∎

Remark 2.5.

In the assumptions of Lemma 2.4, if xclΩ1x\notin\operatorname{cl}\Omega_{1}, then x\ell_{x} may be chosen in such a way that xclΩ1=\ell_{x}\cap\operatorname{cl}\Omega_{1}=\varnothing as well.

Corollary 2.6.

In the assumptions of Lemma 2.4, for any xΩx\in\Omega there exists φ𝐀\varphi\in\boldsymbol{A} such that φI=x\langle{\varphi}\rangle_{{}_{\operatorname{I}}}=x.

Proof.

Construct a segment x\ell_{x} with the help of Lemma 2.4: there exist points a,bΩ0a,b\in\partial\Omega_{0} and non-negative numbers α\alpha and β\beta with sum one such that

x=αa+βband[a,b]Ω1=.{x=\alpha a+\beta b\quad\text{and}\quad[a,b]\cap\Omega_{1}=\varnothing.} (2.17)

Then, the desired function φ\varphi may be constructed as

φ(t)={a,t[0,α];b,t(α,1],{\varphi(t)=\begin{cases}a,\quad t\in[0,\alpha];\\ b,\quad t\in(\alpha,1],\end{cases}} (2.18)

here we set I=[0,1]\operatorname{I}=[0,1] for convenience (all the considerations do not depend on the particular choice of I\operatorname{I}). By construction, for any J[0,1]\operatorname{J}\subset[0,1], the point φJ\langle{\varphi}\rangle_{{}_{\operatorname{J}}} lies inside x\ell_{x}. Therefore, φ𝑨\varphi\in\boldsymbol{A}. ∎


3 Functionals

Let f:Ω0f\colon\partial\Omega_{0}\to\mathbb{R} be a Borel measurable locally bounded function. We wish to find sharp estimates of the expression f(φ)I\langle{f(\varphi)}\rangle_{{}_{\operatorname{I}}} when φ𝑨\varphi\in\boldsymbol{A}. Note that a priori it is unclear whether the integral in question exists. The function

𝑩Ω,f(x)=sup{f(φ)I|φ𝑨,φI=x},xΩ,{\boldsymbol{B}_{\Omega,f}(x)=\sup\Big{\{}{\langle{f(\varphi)}\rangle_{{}_{\operatorname{I}}}}\;\Big{|}\,{\varphi\in\boldsymbol{A},\ \langle{\varphi}\rangle_{{}_{\operatorname{I}}}=x}\Big{\}},\quad x\in\Omega,} (3.1)

is well defined in the case where ff is bounded from below (though this function may attain the value ++\infty). The function

𝑩Ω,fb(x)=sup{f(φ)I|φ𝑨,φI=x,φL},xΩ,{\boldsymbol{B}^{\mathrm{b}}_{\Omega,f}(x)=\sup\Big{\{}{\langle{f(\varphi)}\rangle_{{}_{\operatorname{I}}}}\;\Big{|}\,{\varphi\in\boldsymbol{A},\ \langle{\varphi}\rangle_{{}_{\operatorname{I}}}=x,\quad\varphi\in L_{\infty}}\Big{\}},\qquad x\in\Omega,} (3.2)

is a meaningful object for any ff locally bounded from below. Note that all the Bellman functions in the paper do not depend on the choice of II, because the classes of functions 𝑨\boldsymbol{A} (see (2.2)) and the Bellman functions themselves are defined in terms of averages. We will often omit the symbols Ω\Omega and ff in the notation for our Bellman functions and simply write 𝑩\boldsymbol{B} and 𝑩b\boldsymbol{B}^{\mathrm{b}}. We also use the notation Ω=clΩ0clΩ1\Omega^{*}=\operatorname{cl}\Omega_{0}\setminus\operatorname{cl}\Omega_{1}. Of course, we expect that the functions 𝑩\boldsymbol{B} and 𝑩b\boldsymbol{B}^{\mathrm{b}} coincide in reasonable situations. However, the proof of this assertion might be unexpectedly difficult.

Remark 3.1.

Assume (LABEL:StrictConvexity) holds true. In the case d=2d=2 we also require (2). Corollary 2.6 says that in this case

<𝑩b(x)𝑩(x),xΩ.{-\infty<\boldsymbol{B}^{\mathrm{b}}(x)\leqslant\boldsymbol{B}(x),\qquad x\in\Omega.} (3.3)

Now we pass to the examples and show how the Bellman functions above help to find sharp constants in various inequalities. We do not provide many details for the first two examples since they are discussed in the cited papers.

Muckenhoupt classes.

Consider the domains (2.4) and the function f(x1,x11)=x1pf(x_{1},x_{1}^{-1})=x_{1}^{p}, where p>1p>1. The computation of the corresponding Bellman function 𝑩\boldsymbol{B} leads to sharp constants in various forms of the Reverse Hölder inequality for Muckenhoupt weights. See [25] and [26] for details. For weak-type bounds, one makes the choice f(x1,x11)=χ[1,)(x1)f(x_{1},x_{1}^{-1})=\chi_{[1,\infty)}(x_{1}), see [13].

Scalar BMO\mathrm{BMO} space.

Consider the domains (2.7) with d=2d=2. The choice f(x1,x12)=ex1f(x_{1},x_{1}^{2})=e^{x_{1}} leads to the sharp John–Nirenberg inequality in integral form, see [16]. The function f(x1,x12)=|x1|pf(x_{1},x_{1}^{2})=|x_{1}|^{p} was used to obtain sharp results on the constants in the inequalities that express the equivalence of different norms on BMO\mathrm{BMO}, see [17]. For weak-type estimates, we may choose the function f(x1,x12)=χ[0,)(x1)f(x_{1},x_{1}^{2})=\chi_{[0,\infty)}(x_{1}), see [22]. For quite general boundary values ff, the function (3.1) was computed in [10] (see a simpler version [9] and the short report [7]).

For larger dd similar Bellman functions will lead to sharp inequalities for vectorial functions.

Functions of bounded mean oscillation with values in the unit sphere.

A version of the John–Nirenberg inequality for BMO\mathrm{BMO} functions between manifolds may be found in Appendix B of [3]. In that paper the inequality is stated in its integral form, here we prefer to work with the classical ’tail estimate’ form as in [11]. For that consider the class 𝑨\boldsymbol{A} generated by the domains Ω0\Omega_{0} and Ω1\Omega_{1} given in (2.12). Pick some point x0Sd1x_{0}\in S^{d-1} and some δ(0,1)\delta\in(0,1). Consider the function

f(x)=χ[δ,2](|xx0|),xSd1,{f(x)=\chi_{[\delta,2]}(|x-x_{0}|),\quad x\in S^{d-1},} (3.4)

and the Bellman function (3.1) generated by this boundary value. The Bellman function delivers sharp estimate of the amount of points tIt\in\operatorname{I} such that

|φ(t)x0|δ,{|\varphi(t)-x_{0}|\geqslant\delta,} (3.5)

provided φ\varphi belongs to the ε\varepsilon-ball of BMO(I,Sd1)\mathrm{BMO}(\operatorname{I},S^{d-1}) and φI=x\langle{\varphi}\rangle_{{}_{\operatorname{I}}}=x. If we choose x=x0|x|x=x_{0}|x|, we obtain the sharp estimate for the quantity

1|I||{tI||φ(t)φI|δ~}|;δ~2=(1|x|)2+|x|δ2.{\frac{1}{|\operatorname{I}|}\Big{|}\Big{\{}{t\in\operatorname{I}}\;\Big{|}\,{|\varphi(t)-\langle{\varphi}\rangle_{{}_{\operatorname{I}}}|\geqslant\tilde{\delta}}\Big{\}}\Big{|};\qquad\tilde{\delta}^{2}=(1-|x|)^{2}+|x|\delta^{2}.} (3.6)

The John–Nirenberg inequality says this quantity decays exponentially as ε\varepsilon decreases down to zero. The Bellman function allows to find the sharp constants in the corresponding inequality, see the forthcoming paper [4].

Multiplicative inequalities.

Consider the domains given by (2.13) and the corresponding class 𝑨\boldsymbol{A}. Let f(t,t2,|t|p)=|t|rf(t,t^{2},|t|^{p})=|t|^{r}, where r(p,)r\in(p,\infty). The corresponding Bellman function delivers sharp upper estimates for the LrL_{r}-norm of a function φ\varphi provided its average, LpL_{p}-norm, and BMO\mathrm{BMO} norms are fixed. This, in particular leads to computation of sharp constants cp,rc_{p,r} in the multiplicative inequalities

φLrcp,rφLpp/rφBMO1p/r,{\|\varphi\|_{L_{r}}\leqslant c_{p,r}\|\varphi\|_{L_{p}}^{p/r}\|\varphi\|_{\mathrm{BMO}}^{1-p/r},} (3.7)

see [19] and [23].

4 Locally concave functions and martingales

Definition 4.1.

Let ωd\omega\subset\mathbb{R}^{d}. We say that a function G:ω{+}G\colon\omega\to\mathbb{R}\cup\{+\infty\} is locally concave provided its restriction G|G|_{\ell} to any segment ω\ell\subset\omega is concave.

Here and in what follows we will be using the convention that concave functions may attain infinite values, see [14]. It is important that we do not allow the value -\infty (see a pathological example at the end of this section). Locally concave functions play important role in the Bellman function theory; see [5] for applications to geometry. In the definition below, fixedω\partial_{\mathrm{fixed}}\omega is the set of all points xωx\in\omega such that there does not exist a segment ω\ell\subset\omega with xx lying in the interior of \ell; the latter set is called the fixed boundary (because we fix the boundary values of locally concave functions on this set). The remaining part of the boundary is called the free boundary and is denoted by freeω\partial_{\mathrm{free}}\omega. In our usual examples of a lens Ω=clΩ0Ω1\Omega=\operatorname{cl}\Omega_{0}\setminus\Omega_{1}, we have fixedΩ=Ω0\partial_{\mathrm{fixed}}\Omega=\partial\Omega_{0} and freeΩ=Ω1\partial_{\mathrm{free}}\Omega=\partial\Omega_{1}. From now on we will be using this notation.

Definition 4.2.

Let ωd\omega\subset\mathbb{R}^{d}, let f:fixedωf\colon\partial_{\mathrm{fixed}}\omega\to\mathbb{R} be a function. By Λω,f\Lambda_{{\omega},{f}} we denote the class of all locally concave on ω\omega functions GG that satisfy the boundary inequality G(x)f(x)G(x)\geqslant f(x) for any xfixedωx\in\partial_{\mathrm{fixed}}\omega.

Remark 4.3.

By Theorem 10.110.1 in [14], any function GΛω,fG\in\Lambda_{{\omega},{f}} is continuous on the interior of ω\omega as a mapping with values in {+}\mathbb{R}\cup\{+\infty\}.

We are ready to define the pointwise minimal locally concave function

𝔅ω,f(x)=inf{G(x)|GΛω,f},xω,{\mathfrak{B}_{\omega,f}(x)=\inf\{{G(x)}\;|\,{G\in\Lambda_{{\omega},{f}}}\},\quad x\in\omega,} (4.1)

(note that 𝔅ω,f(x)Λω,f\mathfrak{B}_{\omega,f}(x)\in\Lambda_{{\omega},{f}} if this function does not attain the value -\infty) and state our first main theorem. The notation freeΩC2\partial_{\mathrm{free}}\Omega\in C^{2} means that locally freeΩ\partial_{\mathrm{free}}\Omega coincides with a graph of a C2C^{2}-smooth function. Recall Ω=clΩ0clΩ1\Omega^{*}=\operatorname{cl}\Omega_{0}\setminus\operatorname{cl}\Omega_{1}.

Theorem 4.4.

Let the domains Ω0\Omega_{0} and Ω1\Omega_{1} satisfy the usual requirements clΩ1Ω0\operatorname{cl}\Omega_{1}\subset\Omega_{0}, (LABEL:StrictConvexity), (2), and let also freeΩC2\partial_{\mathrm{free}}\Omega\in C^{2}. Let the function ff be lower semi-continuous and bounded from below. Then,

𝑩Ω,f(x)=𝔅Ω,f(x),xΩ.{\boldsymbol{B}_{\Omega,f}(x)=\mathfrak{B}_{\Omega,f}(x),\qquad x\in\Omega^{*}.} (4.2)

This theorem generalizes, up to minor modifications, the main result of [21]; in that paper we had d=2d=2 and unbounded Ω1\Omega_{1} (and, therefore, unbounded Ω0\Omega_{0}). The methods of [21] relied upon the notion of monotonic (non-decreasing) rearrangement, which, seemingly, does not exist in the case where Ω\Omega is not simply connected. A good replacement for monotonic rearrangements was found in [20]. The disadvantage of this new method is that it does not allow to work with xfreeΩx\in\partial_{\mathrm{free}}\Omega. The main theorem in [21] states identity (4.2) for all xΩx\in\Omega. Seemingly, in the larger generality Theorem 4.4 is also true for xΩx\in\Omega, however, the methods from [20] do not allow to prove it (see Section 9 below for an explanation).

We need to survey the notions from [21] we will be using. We will be working with discrete time martingales adapted to a filtration ={n}n\mathcal{F}=\{\mathcal{F}_{n}\}_{n}. We refer the reader to [15] for the general martingale theory and present here the simplified definitions we will use.

By a filtration we mean a sequence of increasing set algebras (we consider finite algebras only), that is if AnA\in\mathcal{F}_{n}, then An+1A\in\mathcal{F}_{n+1} as well. A sequence {Mn}n\{M_{n}\}_{n} of random variables taking values in some linear space is called a martingale adapted to \mathcal{F} provided, first, each MnM_{n} is n\mathcal{F}_{n}-measurable, and second, for each nn we have Mn=𝔼(Mn+1n)M_{n}=\operatorname{\mathbb{E}}(M_{n+1}\mid\mathcal{F}_{n}). Note that since our algebras are simple, we may freely work in infinite dimensional spaces (in the case of general martingales, there are difficulties with the definition of conditional expectation, see Section 1.31.3 in [6]). All martingales we will be working with have the limit value ML1M_{\infty}\in L_{1}, a random variable that generates the martingale:

Mn=𝔼(Mn),n{0}.{M_{n}=\operatorname{\mathbb{E}}(M_{\infty}\mid\mathcal{F}_{n}),\qquad n\in\mathbb{N}\cup\{0\}.} (4.3)

Here we should take care, MM_{\infty} should attain values in a finite dimensional space since we wish to omit the theory of integration of functions taking values in infinite-dimensional spaces.

The main property of MM_{\infty} may be restated: for any atom ana\in\mathcal{F}_{n} (by an atom of a set algebra \mathcal{F} we mean a set aa\in\mathcal{F} of positive measure that is minimal by inclusion), one may restore the value of MnM_{n} on this atom by the formula

Mn(a)=1P(a)aM.{M_{n}(a)=\frac{1}{P(a)}\int\limits_{a}M_{\infty}.} (4.4)

We cite an important definition from [21]. See Fig. 4 for an example.

Definition 4.5.

Let ωd\omega\subset\mathbb{R}^{d}. An d\mathbb{R}^{d}-valued martingale M={Mn}nM=\{M_{n}\}_{n} adapted to ={n}\mathcal{F}=\{\mathcal{F}_{n}\} is called an ω\omega-martingale, provided

  1. 1)

    the algebra 0\mathcal{F}_{0} is trivial, i. e., consists of the whole probability space and the empty set;

  2. 2)

    there exists a random variable MM_{\infty} with the values in fixedω\partial_{\mathrm{fixed}}\omega such that MnMM_{n}\to M_{\infty} in L1L_{1} and almost everywhere (in particular, MM_{\infty} is summable itself);

  3. 3)

    for any atom ana\in\mathcal{F}_{n} there exists a convex set CaωC_{a}\subset\omega such that Mn+1|aM_{n+1}|_{a} lies in CaC_{a} almost surely.

Sometimes, we will need to use a slightly modified notion of an (ω,𝔇)(\omega,\mathfrak{D})-martingale introduced in [20].

Definition 4.6.

Let ωd\omega\subset\mathbb{R}^{d} and let 𝔇fixedω\mathfrak{D}\subset\partial_{\mathrm{fixed}}\omega. An ω\omega-martingale MM is called an (ω,𝔇)(\omega,\mathfrak{D})-martingale, provided M𝔇M_{\infty}\in\mathfrak{D} almost surely.

Remark 4.7.

We may also consider ω\omega or (ω,𝔇)(\omega,\mathfrak{D})-martingales for infinite-dimensional domains ω\omega. However, in this case we require that MM_{\infty} attains values in the intersection of fixedω\partial_{\mathrm{fixed}}\omega with a finite dimensional space. For example, this happens if MM_{\infty} attains finitely many values.

Refer to caption
Refer to caption
Refer to caption
Figure 4: An example of an ω\omega-martingale where ω\omega is given in (2.12); the numbers assigned to points denote the probability of the specific value.

Consider two other Bellman functions on ω\omega: the first one is defined for f:fixedωf\colon\partial_{\mathrm{fixed}}\omega\to\mathbb{R} measurable and bounded from below,

(x)=sup{𝔼f(M)|M0=x,Mis an ω-martingale},xω,{\mathcal{B}(x)=\sup\Big{\{}{\operatorname{\mathbb{E}}f(M_{\infty})}\;\Big{|}\,{M_{0}=x,\ M\ \text{is an\leavevmode\nobreak\ $\omega$-martingale}}\Big{\}},\quad x\in\omega,} (4.5)

and the second one is defined for arbitrary locally bounded from below measurable ff:

b(x)=sup{𝔼f(M)|M0=x,Mis an ω-martingale,ML},xω.{\mathcal{B}^{\mathrm{b}}(x)=\sup\Big{\{}{\operatorname{\mathbb{E}}f(M_{\infty})}\;\Big{|}\,{M_{0}=x,\ M\ \text{is an\leavevmode\nobreak\ $\omega$-martingale},\ M_{\infty}\in L_{\infty}}\Big{\}},\quad x\in\omega.} (4.6)

Recall that a martingale MM is called simple if there exists a natural number NN such that Mk=MNM_{k}=M_{N} for all kNk\geqslant N. We will use Theorem 2.212.21 from [21]. It uses the notion of a strongly martingale connected domain. A set ωd\omega\subset\mathbb{R}^{d} is called strongly martingale connected if for every xωx\in\omega there exists a simple ω\omega-martingale MM that starts at xx, that is M0=xM_{0}=x.

Lemma 4.8.

Assume (LABEL:StrictConvexity). The domains Ω\Omega and Ω\Omega^{*} are strongly martingale connected if either d3d\geqslant 3 or (2) holds true.

Proof.

Consider the case of the domain Ω\Omega first. Let xΩx\in\Omega, we wish to construct a simple Ω\Omega-martingale starting at xx. If xfixedΩx\in\partial_{\mathrm{fixed}}\Omega, then the desired martingale is constantly equal xx. So, we may assume xfixedΩx\notin\partial_{\mathrm{fixed}}\Omega. By Lemma 2.4, there exists a segment xΩ\ell_{x}\subset\Omega with the endpoints AA and BB lying on fixedΩ\partial_{\mathrm{fixed}}\Omega, passing through xx. In other words,

x=αA+βB,α+β=1,andα,β0.{x=\alpha A+\beta B,\qquad\alpha+\beta=1,\ \text{and}\ \alpha,\beta\geqslant 0.} (4.7)

Let us construct the martingale MM by the formula,

M0=x,M1={A,with probabilityα;B,with probabilityβ,Mn=M1,n1.{M_{0}=x,\quad M_{1}=\begin{cases}A,\quad\text{with probability}\ \alpha;\\ B,\quad\text{with probability}\ \beta,\end{cases}M_{n}=M_{1},\quad n\geqslant 1.} (4.8)

In other words, MM is a martingale that starts at xx, splits into AA and BB, and stops there. Since xΩ\ell_{x}\subset\OmegaMM is the desired simple Ω\Omega-martingale.

The reasoning for the domain Ω\Omega^{*} is completely similar. One relies upon Remark 2.5 instead of Lemma 2.4 in this case. ∎

Theorem 4.9 (Theorem 2.212.21 in [21]).

Let ωd\omega\subset\mathbb{R}^{d} be a strongly martingale connected domain. Let f:fixedωf\colon\partial_{\mathrm{fixed}}\omega\to\mathbb{R} be a bounded from below function such that 𝔅w,f\mathfrak{B}_{w,f} is continuous at every point of the fixed boundary. Then, 𝔅(x)=(x)\mathfrak{B}(x)=\mathcal{B}(x) for all xωx\in\omega.

Remark 2.222.22 of the same paper says that if ff is only locally bounded from below, 𝔅ω,f\mathfrak{B}_{\omega,f} is continuous at the points of the fixed boundary, then 𝔅=b\mathfrak{B}=\mathcal{B}^{\mathrm{b}}. We will need a slightly stronger statement. Let us introduce yet another Bellman function

s(x)=sup{𝔼f(M)|M0=x,Mis a simple Ω-martingale},xω.{\mathcal{B}^{\mathrm{s}}(x)=\sup\Big{\{}{\operatorname{\mathbb{E}}f(M_{\infty})}\;\Big{|}\,{M_{0}=x,\ M\ \text{is a simple\leavevmode\nobreak\ $\Omega$-martingale}}\Big{\}},\quad x\in\omega.} (4.9)

Clearly, sb\mathcal{B}^{\mathrm{s}}\leqslant\mathcal{B}^{\mathrm{b}}\leqslant\mathcal{B}. We claim that this chain of inequalities often turns into a chain of equalities.

Lemma 4.10.

Let ωd\omega\subset\mathbb{R}^{d} be a strongly martingale connected domain. Let f:fixedωf\colon\partial_{\mathrm{fixed}}\omega\to\mathbb{R} be any measurable function. Then, 𝔅w,f(x)=w,fs(x)\mathfrak{B}_{w,f}(x)=\mathcal{B}^{\mathrm{s}}_{w,f}(x) for any xωx\in\omega.

Proof.

The proof is a simplification of the proof of Theorem 2.222.22 in [21]; we provide a comment about simplifications. It suffices to prove the inequalities 𝔅(x)s(x)\mathfrak{B}(x)\leqslant\mathcal{B}^{\mathrm{s}}(x) and s(x)𝔅(x)\mathcal{B}^{\mathrm{s}}(x)\leqslant\mathfrak{B}(x) for any xωx\in\omega.

To show the first inequality, it suffices to check the inclusion sΛω,f\mathcal{B}^{\mathrm{s}}\in\Lambda_{{\omega},{f}} and use the definition of 𝔅\mathfrak{B}. It is clear that s\mathcal{B}^{\mathrm{s}} satisfies the boundary condition. By the requirement that ω\omega is strongly martingale connected, s\mathcal{B}^{\mathrm{s}} does not attain the value -\infty. The local concavity may be verified similar to the proof of Lemma 2.172.17 in [21].

To show the reverse inequality s(x)𝔅(x)\mathcal{B}^{\mathrm{s}}(x)\leqslant\mathfrak{B}(x), it suffices to prove that for any simple ω\omega-martingale MM with M0=xM_{0}=x and any GΛω,fG\in\Lambda_{{\omega},{f}}, the inequality

G(M0)𝔼G(M)𝔼f(M){G(M_{0})\geqslant\operatorname{\mathbb{E}}G(M_{\infty})\geqslant\operatorname{\mathbb{E}}f(M_{\infty})} (4.10)

holds true. This inequality follows from Lemma 2.102.10 in [21] that says that the quantity 𝔼G(Mn)\operatorname{\mathbb{E}}G(M_{n}) is non-increasing in this case; note that the assumption that MM is simple cancels the need for the limit argument (compare with the proof of Lemma 2.19 in [21]). ∎

A simple modification of the proof above allows to prove a version of Theorem 4.9.

Theorem 4.11.

Let ωd\omega\subset\mathbb{R}^{d} be a strongly martingale connected domain. Let f:fixedωf\colon\partial_{\mathrm{fixed}}\omega\to\mathbb{R} be a bounded from below function such that 𝔅w,f\mathfrak{B}_{w,f} is lower semi-continuous at every point of the fixed boundary. Then, 𝔅(x)=(x)\mathfrak{B}(x)=\mathcal{B}(x) for all xωx\in\omega.

We will also need a technical statement in the spirit of Proposition 2.72.7 in [21]. The proof is also very similar, so, we omit it.

Proposition 4.12.

Let ωd\omega\subset\mathbb{R}^{d}, let xfixedωx\in\partial_{\mathrm{fixed}}\omega. Suppose there exists a closed ball Br(x)B_{r}(x) such that Br(x)ωB_{r}(x)\cap\omega is a closed strictly convex set. Suppose 𝔅ω,f\mathfrak{B}_{\omega,f} nowhere equals ++\infty. Then, 𝔅\mathfrak{B} is lower semi-continuous at the point xx provided ff is.

Surprisingly, the assertion of Lemma 4.10 becomes false without the assumption ω\omega is strongly martingale connected. We use a little bit of complex variable notation to indicate points in the plane. Let ω\omega be given by the rule

ω={z2|z|1}j=0,1,2{z2||z12e2πij3|<341239}.\omega=\{z\in\mathbb{R}^{2}\mid|z|\leqslant 1\}\setminus\bigcup_{j=0,1,2}\Big{\{}{z\in\mathbb{R}^{2}}\;\Big{|}\,{\big{|}z-\frac{1}{2}e^{\frac{2\pi ij}{3}}\big{|}<\frac{\sqrt{3}}{4}-\frac{1}{239}}\Big{\}}. (4.11)
Refer to caption
Figure 5: An example of a domain that is not strongly martingale connected. The dotted lines mark the convex hull.

The main feature of the specific numbers in (4.11) is that the three small circles almost touch (see Fig. 5). By definition, fixedω={|z|=1}\partial_{\mathrm{fixed}}\omega=\{|z|=1\}. Let f0f\equiv 0 on the fixed boundary. In this case,

ω,fs(x)={,xωconv(j=0,1,2{z2||z12e2πij3|<341239});0,otherwise.\mathcal{B}^{\mathrm{s}}_{\omega,f}(x)=\begin{cases}-\infty,\quad&x\in\omega\cap\operatorname{conv}\bigg{(}\bigcup_{j=0,1,2}\Big{\{}{z\in\mathbb{R}^{2}}\;\Big{|}\,{\big{|}z-\frac{1}{2}e^{\frac{2\pi ij}{3}}\big{|}<\frac{\sqrt{3}}{4}-\frac{1}{239}}\Big{\}}\bigg{)};\\ 0,\quad&\hbox{otherwise}.\end{cases}

On the other hand, Lemma C.5\mathrm{C}.5 in [21] says 𝔅ω,f0\mathfrak{B}_{\omega,f}\geqslant 0 (the domain ω\omega is a cheese domain in the terminology of that paper). Therefore, 𝔅0\mathfrak{B}\equiv 0 and does not coincide with s\mathcal{B}^{\mathrm{s}}. The effect of similar nature arises when one defines rank-one or separate convex hulls, see [12].

5 Functions on the circle and the line

During the proof, we will need to work with functions defined on the circle 𝕋\mathbb{T} of unit length (in other words, its radius equals 1/(2π)1/(2\pi)). Let us equip 𝕋\mathbb{T} with the natural arc length measure. We will think of functions on 𝕋\mathbb{T} as of periodic functions on the line, i. e., identify the function φ:𝕋d\varphi\colon\mathbb{T}\to\mathbb{R}^{d} with its periodic version φper:d\varphi\hbox{\tiny{per}}\colon\mathbb{R}\to\mathbb{R}^{d}, here

φper(t)=φ(12πe2πit),t.{\varphi\hbox{\tiny{per}}(t)=\varphi\Big{(}\frac{1}{2\pi}e^{2\pi it}\Big{)},\qquad t\in\mathbb{R}.} (5.1)

Consider a version of the class (2.2) formed from summable d\mathbb{R}^{d}-valued functions on 𝕋\mathbb{T}:

𝑨={φ:𝕋Ω0|open setΩ^1such thatclΩ1Ω^1andclΩ^1Ω0and for any intervalJφperJΩ^1}.{\boldsymbol{A}^{\circ}=\Bigg{\{}{\varphi\colon\mathbb{T}\to\partial\Omega_{0}}\;\Bigg{|}\,{\begin{aligned} \exists\ \text{open set}\ \hat{\Omega}_{1}\ \text{such that}\ \operatorname{cl}\Omega_{1}\subset\hat{\Omega}_{1}\ \text{and}\ \operatorname{cl}\hat{\Omega}_{1}\subset\Omega_{0}\\ \text{and for any interval}\ \operatorname{J}\subset\mathbb{R}\quad\langle{\varphi\hbox{\tiny{per}}}\rangle_{{}_{\operatorname{J}}}\notin\hat{\Omega}_{1}\end{aligned}}\Bigg{\}}.} (5.2)

Note that in this definition we require the ’boundedness of oscillation’ over arcs that may wind around the circle several times. The additional domain Ω^1\hat{\Omega}_{1} is mostly needed for technical purposes, it seems to be unavoidable in Lemma 6.1 below.

Remark 5.1.

By Theorem 8.17 below, we may assume Ω^1\hat{\Omega}_{1} in (5.2) is strictly convex. In this case, the domain Ω^=clΩ0Ω^1\hat{\Omega}=\operatorname{cl}\Omega_{0}\setminus\hat{\Omega}_{1} satisfies (LABEL:StrictConvexity) and (2).

Recall Ω=clΩ0clΩ1\Omega^{*}=\operatorname{cl}\Omega_{0}\setminus\operatorname{cl}\Omega_{1}.

Lemma 5.2.

Assume the domains satisfy (LABEL:StrictConvexity). In the case d=2d=2 we additionally assume (2). For any xΩx\in\Omega^{*} there exists φ𝐀\varphi\in\boldsymbol{A}^{\circ} such that φ𝕋=x\langle{\varphi}\rangle_{{}_{\mathbb{T}}}=x.

Proof.

Construct a segment x\ell_{x} with the help of Remark 2.5: there exist points a,bΩ0a,b\in\partial\Omega_{0} and non-negative numbers α\alpha and β\beta with sum one such that x=αa+βbx=\alpha a+\beta b and x=[a,b]Ω\ell_{x}=[a,b]\subset\Omega^{*}. Take any open Ω^1\hat{\Omega}_{1} such that clΩ1Ω^1\operatorname{cl}\Omega_{1}\subset\hat{\Omega}_{1} and clΩ^1Ω0x\operatorname{cl}\hat{\Omega}_{1}\subset\Omega_{0}\setminus\ell_{x}. Construct the function φ\varphi as in (2.18) and extend it periodically to the entire line. Then, for any J\operatorname{J}\subset\mathbb{R} the point φperJ\langle{\varphi_{\hbox{\tiny{per}}}}\rangle_{{}_{\operatorname{J}}} lies on x\ell_{x} and therefore not in Ω^1\hat{\Omega}_{1}. ∎

Consider the Bellman functions

𝑩Ω,f(x)=sup{f(φ)I|φ𝑨,φ𝕋=x},xΩ,{\boldsymbol{B}^{\circ}_{\Omega,f}(x)=\sup\Big{\{}{\langle{f(\varphi)}\rangle_{{}_{\operatorname{I}}}}\;\Big{|}\,{\varphi\in\boldsymbol{A}^{\circ},\ \langle{\varphi}\rangle_{{}_{\mathbb{T}}}=x}\Big{\}},\quad x\in\Omega^{*},} (5.3)

and

𝑩Ω,f,b(x)=sup{f(φ)I|φ𝑨,φ𝕋=x,φL},xΩ.{\boldsymbol{B}^{\circ,\mathrm{b}}_{\Omega,f}(x)=\sup\Big{\{}{\langle{f(\varphi)}\rangle_{{}_{\operatorname{I}}}}\;\Big{|}\,{\varphi\in\boldsymbol{A}^{\circ},\ \langle{\varphi}\rangle_{{}_{\mathbb{T}}}=x,\quad\varphi\in L_{\infty}}\Big{\}},\quad x\in\Omega^{*}.} (5.4)

Similar to (3.1) and (3.2), we define the function (5.3) for ff that is bounded from below, whereas the function 𝑩,b\boldsymbol{B}^{\circ,\mathrm{b}} is a meaningful object for any ff locally bounded from below. Since φper|[0,1]𝑨\varphi_{\hbox{\tiny{per}}}|_{[0,1]}\in\boldsymbol{A} for any φ𝑨\varphi\in\boldsymbol{A}^{\circ}, we have

𝑩Ω,f(x)𝑩Ω,f(x)and𝑩Ω,f,b(x)𝑩Ω,fb(x),xΩ.{\boldsymbol{B}^{\circ}_{\Omega,f}(x)\leqslant\boldsymbol{B}_{\Omega,f}(x)\quad\text{and}\quad\boldsymbol{B}^{\circ,\mathrm{b}}_{\Omega,f}(x)\leqslant\boldsymbol{B}^{\mathrm{b}}_{\Omega,f}(x),\qquad x\in\Omega^{*}.} (5.5)

The theorem below is our second main result.

Theorem 5.3.

Let the domains Ω0\Omega_{0} and Ω1\Omega_{1} satisfy the usual requirements clΩ1Ω0\operatorname{cl}\Omega_{1}\subset\Omega_{0}, (LABEL:StrictConvexity), and (2). Let the function ff be lower semi-continuous and bounded from below. Then,

𝑩Ω,f(x)=𝔅Ω,f(x),xΩ.{\boldsymbol{B}^{\circ}_{\Omega,f}(x)=\mathfrak{B}_{\Omega^{*},f}(x),\qquad x\in\Omega^{*}.} (5.6)

The geometric functions on the right hand sides of Theorems 4.4 and 5.3 are closely related.

Proposition 5.4.

Let the domains Ω0\Omega_{0} and Ω1\Omega_{1} satisfy the requirements clΩ1Ω0\operatorname{cl}\Omega_{1}\subset\Omega_{0} and (LABEL:StrictConvexity). Let the function ff be locally bounded. Then,

𝔅Ω,f(x)=𝔅Ω,f(x),xΩ.{\mathfrak{B}_{\Omega,f}(x)=\mathfrak{B}_{\Omega^{*},f}(x),\qquad x\in\Omega^{*}.} (5.7)

Before we pass to the proofs, we need to survey the theory from [20].

Definition 5.5.

We say that two YY-valued random variables ζ1\zeta_{1} and ζ2\zeta_{2} are equidistributed provided they have the same distributions, which means

P(ζ1A)=P(ζ2A){P(\zeta_{1}\in A)=P(\zeta_{2}\in A)} (5.8)

for any measurable set AYA\subset Y.

Note that if ζ1\zeta_{1} and ζ2\zeta_{2} are equidistributed, then 𝔼f(ζ1)=𝔼f(ζ2)\operatorname{\mathbb{E}}f(\zeta_{1})=\operatorname{\mathbb{E}}f(\zeta_{2}) for any function ff such that one of these mathematical expectations exists. If φ𝑨\varphi\in\boldsymbol{A}^{\circ} and J\operatorname{J}\subset\mathbb{R} is an interval, then by μφ|J\mu_{\varphi|_{\operatorname{J}}} we denote the distribution of the random variable φper|J\varphi_{\hbox{\tiny{per}}}|_{\operatorname{J}}. In other words,

μφ|J(A)=1|J||{tJ|φper(t)A}|,Ais a Borel subset offixedΩ.{\mu_{\varphi|_{\operatorname{J}}}(A)=\frac{1}{|\operatorname{J}|}\big{|}\{{t\in\operatorname{J}}\;|\,{\varphi_{\hbox{\tiny{per}}}(t)\in A}\}\big{|},\qquad A\ \text{is a Borel subset of}\ \partial_{\mathrm{fixed}}\Omega.} (5.9)

Note that μφ|J\mu_{\varphi|_{\operatorname{J}}} is a probability measure on fixedΩ\partial_{\mathrm{fixed}}\Omega. Consider the set (fixedΩ)\mathcal{M}(\partial_{\mathrm{fixed}}\Omega) of all probability measures on fixedΩ\partial_{\mathrm{fixed}}\Omega with bounded first moment. The set (fixedΩ)\mathcal{M}(\partial_{\mathrm{fixed}}\Omega) is a convex subset of the space of all finite signed measures on fixedΩ\partial_{\mathrm{fixed}}\Omega with bounded first moment. Let 𝔚\mathfrak{W} be a subset of (fixedΩ)\mathcal{M}(\partial_{\mathrm{fixed}}\Omega). We will always impose two conditions on this set:

1) The set 𝔚 contains all delta measures, i. e., 𝔇={δx|xfixedΩ}𝔚;\displaystyle 1)\text{ The set }\mathfrak{W}\text{ contains all delta measures, i.\,e., }\mathfrak{D}=\{{\delta_{x}}\;|\,{x\in\partial_{\mathrm{fixed}}\Omega}\}\subset\mathfrak{W}; (5.10)
2) For any finite collection x1,x2,,xNfixedΩ the intersection of 𝔚\displaystyle 2)\text{ For any finite collection }x_{1},x_{2},\ldots,x_{N}\in\partial_{\mathrm{fixed}}\Omega\text{ the intersection of }\mathfrak{W}
 with the linear space generated by δx1,δx2,,δxN\displaystyle\quad\text{ with the linear space generated by }\delta_{x_{1}},\delta_{x_{2}},\ldots,\delta_{x_{N}}
    is open in the topology of the latter space.

Note that 𝔇fixed𝔚\mathfrak{D}\subset\partial_{\mathrm{fixed}}\mathfrak{W}. We will be working with simple martingales that attain their values in the space (fixedΩ)\mathcal{M}(\partial_{\mathrm{fixed}}\Omega); these martingales are easy to define since a simple martingale attains values in a finite dimensional linear space. We denote by μφ\mu_{\varphi} the distribution of the function φ\varphi itself, i. e., μφ=μφ|[0,1]\mu_{\varphi}=\mu_{\varphi|_{[0,1]}}.

Theorem 5.6 (Theorem 2.32.3 in [20] with slight modifications).

Let 𝔚\mathfrak{W} be a subset of (fixedΩ)\mathcal{M}(\partial_{\mathrm{fixed}}\Omega) that satisfies conditions (5.10). Let 𝕄\mathbb{M} be a simple (𝔚,𝔇)(\mathfrak{W},\mathfrak{D})-martingale in the sense of Remark 4.7. Then, there exists φ:𝕋fixedΩ\varphi\colon\mathbb{T}\to\partial_{\mathrm{fixed}}\Omega such that μφ=𝕄0\mu_{\varphi}=\mathbb{M}_{0} and for any interval J\operatorname{J}\subset\mathbb{R} we have μφ|J𝔚\mu_{\varphi|_{\operatorname{J}}}\in\mathfrak{W}.

The theorem above will serve as a tool to construct functions φ𝑨\varphi\in\boldsymbol{A}^{\circ} with prescribed distributions. We must say about the difference between the formulations above and in [20]. We require weaker openness condition on 𝔚\mathfrak{W} here, in [20] the set 𝔚\mathfrak{W} was open in weak-* topology. One may go through the proof in [20] and realize that everything happens in the finite dimensional space spanned by the values of 𝕄\mathbb{M}_{\infty}. We will apply the theorem to the sets

𝔚={μ(fixedΩ)|dx𝑑μ(x)clΩ1}.{\mathfrak{W}=\Big{\{}{\mu\in\mathcal{M}(\partial_{\mathrm{fixed}}\Omega)}\;\Big{|}\,{\int\limits_{\mathbb{R}^{d}}x\,d\mu(x)\notin\operatorname{cl}\Omega_{1}}\Big{\}}.} (5.11)

Note that the intersection of 𝔚\mathfrak{W} with any finite dimensional space VV spanned by delta measures is open. By definition, the condition φ𝑨\varphi\in\boldsymbol{A}^{\circ} is almost equivalent to the condition that μφ|J𝔚\mu_{\varphi|_{J}}\in\mathfrak{W} for any subinterval JJ\subset\mathbb{R} (the word ‘almost’ corresponds to the presence of the set Ω^1\hat{\Omega}_{1} in (5.2)). We will later show that any simple Ω\Omega-martingale MM generates a simple martingale 𝕄\mathbb{M} with values in (fixedΩ)\mathcal{M}(\partial_{\mathrm{fixed}}\Omega) that satisfies the conditions of Theorem 5.6 with 𝔚\mathfrak{W} given in (5.11). This observation will allow us to construct a function φ𝑨\varphi\in\boldsymbol{A}^{\circ} that has the same distribution as MM_{\infty} (see Lemma 6.2 below).

6 Proof of Theorem 5.3

The proof will be based on two lemmas below.

Lemma 6.1 (Splitting lemma).

Assume Ω\Omega satisfies (LABEL:StrictConvexity) and (2). Let φ𝐀\varphi\in\boldsymbol{A} and let Ω~1\tilde{\Omega}_{1} be an open set such that clΩ~1Ω1\operatorname{cl}\tilde{\Omega}_{1}\subset\Omega_{1}, let Ω~=Ω0Ω~1\tilde{\Omega}=\Omega_{0}\setminus\tilde{\Omega}_{1}. There exists an Ω~\tilde{\Omega}-martingale MM such that MM_{\infty} is equidistributed with φ\varphi.

We will call the domain Ω~\tilde{\Omega} as in lemma above an extension of Ω\Omega.

Lemma 6.2 (Gluing lemma).

Assume Ω\Omega satisfies (LABEL:StrictConvexity) and (2). Let MM be a simple Ω\Omega^{*}-martingale. There exists a function φ𝐀\varphi\in\boldsymbol{A}^{\circ} that is equidistributed with MM_{\infty}.

Proof of Theorem 5.3..

It suffices to prove the inequalities

(6.1)

Without loss of generality assume that 𝔅Ω,f\mathfrak{B}_{\Omega^{*},f} is finite. Let us first prove (LABEL:CircleThFirstIneq). Fix xΩx\in\Omega^{*}. By (5.3), for any ε>0\varepsilon>0 there exists ψ𝑨\psi\in\boldsymbol{A}^{\circ} such that

ψ𝕋=xandf(ψ)𝕋𝑩(x)ε.{\langle{\psi}\rangle_{{}_{\mathbb{T}}}=x\qquad\text{and}\qquad\langle{f(\psi)}\rangle_{{}_{\mathbb{T}}}\geqslant\boldsymbol{B}^{\circ}(x)-\varepsilon.} (6.2)

Let Ω^1\hat{\Omega}_{1} be the set that corresponds to ψ\psi in (5.2). By Remark 5.1 we assume Ω^1\hat{\Omega}_{1} is strictly convex and Ω^=clΩ0Ω^1\hat{\Omega}=\operatorname{cl}\Omega_{0}\setminus\hat{\Omega}_{1} satisfies (LABEL:StrictConvexity) and (2). Then, Ω\Omega is an extension of Ω^\hat{\Omega}. We apply Lemma 6.1 to the function ψper|[0,1]𝑨(Ω^)\psi_{\hbox{\tiny{per}}}|_{[0,1]}\in\boldsymbol{A}(\hat{\Omega}) with Ω\Omega in the role of extension of Ω^\hat{\Omega} and obtain an Ω\Omega-martingale MM such that MM_{\infty} is equidistributed with ψ\psi. Then,

𝑩Ω,f(x)f(ψ)𝕋+ε=𝔼f(M)+ε𝔅Ω,f(x)+ε,{\boldsymbol{B}^{\circ}_{\Omega,f}(x)\leqslant\langle{f(\psi)}\rangle_{{}_{\mathbb{T}}}+\varepsilon=\operatorname{\mathbb{E}}f(M_{\infty})+\varepsilon\leqslant\mathfrak{B}_{\Omega^{*},f}(x)+\varepsilon,} (6.3)

the last inequality follows from Theorem 4.11, Proposition 4.12, and Lemma 4.8. It remains to choose arbitrarily small ε\varepsilon.

Let us now prove (6). By Lemmas 4.8 and 4.10, for any xΩx\in\Omega^{*} there exists a simple Ω\Omega^{*}-martingale MM such that

𝔼f(M)𝔅Ω,f(x)ε,M0=x.{\operatorname{\mathbb{E}}f(M_{\infty})\geqslant\mathfrak{B}_{\Omega^{*},f}(x)-\varepsilon,\qquad M_{0}=x.} (6.4)

We apply Lemma 6.2 and obtain a function φ𝑨\varphi\in\boldsymbol{A}^{\circ} equidistributed with MM_{\infty}. Then,

φ𝕋=𝔼M=x,f(φ)𝕋=𝔼f(M)𝔅Ω,f(x)ε.{\langle{\varphi}\rangle_{{}_{\mathbb{T}}}=\operatorname{\mathbb{E}}M_{\infty}=x,\qquad\langle{f(\varphi)}\rangle_{{}_{\mathbb{T}}}=\operatorname{\mathbb{E}}f(M_{\infty})\geqslant\mathfrak{B}_{\Omega^{*},f}(x)-\varepsilon.} (6.5)

It remains to choose arbitrarily small ε\varepsilon to prove (6). ∎

The proof of Lemma 6.1 follows the lines of the proof of Theorem 3.73.7 in [21]. We introduce the function Δ:ΩfixedΩ\Delta\colon\Omega\setminus\partial_{\mathrm{fixed}}\Omega\to\mathbb{R}:

Δ(x)=sup{max(1,|xy||xz|)|x[y,z],yΩ,zclΩ~1}.{\Delta(x)=\sup\Big{\{}{\max\Big{(}1,\frac{|x-y|}{|x-z|}\Big{)}}\;\Big{|}\,{x\in[y,z],\ y\in\Omega,\ z\in\operatorname{cl}\tilde{\Omega}_{1}}\Big{\}}.} (6.6)

We will also often use the notion of a transversal segment. See Fig. 6 for visualization.

Refer to caption
Figure 6: The segment x\ell_{x} is transversal, the segment ee is not transversal in this case.
Definition 6.3.

Let xfreeΩx\in\partial_{\mathrm{free}}\Omega, let Ω\ell\subset\Omega be a segment with the endpoint xx. We say that xx is transversal provided the line containing it intersects Ω1\Omega_{1}.

Lemma 6.4.

Let Ω0\Omega_{0} and Ω1\Omega_{1} satisfy (LABEL:StrictConvexity). Let Ω~1Ω1\tilde{\Omega}_{1}\subset\Omega_{1} be such that clΩ~1Ω1\operatorname{cl}\tilde{\Omega}_{1}\subset\Omega_{1}. Then, the condition (2) is equivalent to the fact that for any choice of Ω~1\tilde{\Omega}_{1} the function Δ\Delta is uniformly bounded on any compact subset of Ω\Omega.

Proof.

Let us first verify the necessity of (2). Assume it does not hold and there exists a ray LΩ0L\subset\Omega_{0} such that it cannot be shifted inside Ω1\Omega_{1}. Without loss of generality, we may assume LL starts from xfreeΩx\in\partial_{\mathrm{free}}\Omega and does not intersect Ω1\Omega_{1}. By (LABEL:StrictConvexity), we may also assume LL is a transversal segment. Let us choose Ω~1\tilde{\Omega}_{1} in such a manner that it intersects the continuation of LL beyond xx, let zz belong to that intersection. Choosing yy as far as we wish on LL, we obtain that the ratio |xy|/|xz||x-y|/|x-z|, and thus, the value Δ(x)\Delta(x), is unbounded.

Now we turn to the sufficiency of (2). Let CΩC\subset\Omega be a compact set. First, we note that

|xz|dist(C,Ω~1)>0,xC,zΩ~1.{|x-z|\geqslant\operatorname{dist}(C,\tilde{\Omega}_{1})>0,\qquad x\in C,\ z\in\tilde{\Omega}_{1}.} (6.7)

Second, it suffices to prove that the quantity |yx||y-x| is uniformly bounded whenever xCx\in CyΩy\in\Omega, and there exists zclΩ~1z\in\operatorname{cl}\tilde{\Omega}_{1} such that x[y,z]x\in[y,z]. Assume the contrary. Let there exist sequences {xn}n\{x_{n}\}_{n},  {yn}n\{y_{n}\}_{n}, and {zn}n\{z_{n}\}_{n} such that

xnC,ynΩ,znclΩ~1,xn[yn,zn],and|ynxn|.{x_{n}\in C,\ y_{n}\in\Omega,\ z_{n}\in\operatorname{cl}\tilde{\Omega}_{1},\quad x_{n}\in[y_{n},z_{n}],\quad\text{and}\quad|y_{n}-x_{n}|\to\infty.} (6.8)

Without loss of generality, we may assume that xnxx_{n}\to x and (ynxn)/|ynxn|e(y_{n}-x_{n})/|y_{n}-x_{n}|\to e, where |e|=1|e|=1. By the closedness of Ω\Omega, the ray L=x+e+L=x+e\cdot\mathbb{R}_{+} lies inside Ω\Omega entirely. By (LABEL:StrictConvexity), Ω\Omega does not contain lines, so |zn||z_{n}| is uniformly bounded. We may assume znzclΩ~1z_{n}\to z\in\operatorname{cl}\tilde{\Omega}_{1}. This means LΩL\subset\Omega cannot be shifted to lie inside Ω1\Omega_{1}, which contradicts (2). ∎

Proof of Lemma 6.1..

Given a function φ𝑨\varphi\in\boldsymbol{A}, there exists a partition I=I1I2\operatorname{I}=\operatorname{I}_{1}\cup\operatorname{I}_{2}, with I1,I2\operatorname{I}_{1},\operatorname{I}_{2} being disjoint (up to a common point) intervals such that

[φI1,φI2]Ω~1=andmax(|I1||I2|,|I2||I1|)Δ(φI).{\big{[}\langle{\varphi}\rangle_{{}_{\operatorname{I}_{1}}},\langle{\varphi}\rangle_{{}_{\operatorname{I}_{2}}}\big{]}\cap\tilde{\Omega}_{1}=\varnothing\quad\text{and}\quad\max\Big{(}\frac{|\operatorname{I}_{1}|}{|\operatorname{I}_{2}|},\frac{|\operatorname{I}_{2}|}{|\operatorname{I}_{1}|}\Big{)}\leqslant\Delta(\langle{\varphi}\rangle_{{}_{\operatorname{I}}}).} (6.9)

The proof of this statement is completely similar to the proof of Lemma 3.93.9 in [21]. We apply it inductively to build a sequence {{Ikn}k=12n}n\{\{\operatorname{I}_{k}^{n}\}_{k=1}^{2^{n}}\}_{n} of partitions of I\operatorname{I} such that

  1. 1)

    for each nn the partition {Ikn+1}k\{\operatorname{I}_{k}^{n+1}\}_{k} is a subpartition of {Ikn}k\{\operatorname{I}_{k}^{n}\}_{k}, moreover, for each nn and kk1k2n1\leqslant k\leqslant 2^{n}, one has I2k1n+1I2kn+1=Ikn\operatorname{I}^{n+1}_{2k-1}\cup\operatorname{I}^{n+1}_{2k}=\operatorname{I}^{n}_{k};

  2. 2)

    for each nn and kk1k2n1\leqslant k\leqslant 2^{n}, the segment [φI2k1n+1,φI2kn+1]\Big{[}\langle{\varphi}\rangle_{{}_{\operatorname{I}_{2k-1}^{n+1}}},\langle{\varphi}\rangle_{{}_{\operatorname{I}^{n+1}_{2k}}}\Big{]} lies in Ω~\tilde{\Omega};

  3. 3)

    for each nn and kk1k2n1\leqslant k\leqslant 2^{n}max(|I2k1n+1||I2kn+1|,|I2kn+1||I2k1n+1|)Δ(φIkn)\max\Big{(}\frac{|\operatorname{I}^{n+1}_{2k-1}|}{|\operatorname{I}^{n+1}_{2k}|},\frac{|\operatorname{I}^{n+1}_{2k}|}{|\operatorname{I}^{n+1}_{2k-1}|}\Big{)}\leqslant\Delta(\langle{\varphi}\rangle_{{}_{\operatorname{I}^{n}_{k}}}).

Let n\mathcal{F}_{n} be generated by {Ikn}k=12n\{\operatorname{I}_{k}^{n}\}_{k=1}^{2^{n}}, let Mn=𝔼(φn)M_{n}=\operatorname{\mathbb{E}}(\varphi\mid\mathcal{F}_{n}). The martingale M={Mn}nM=\{M_{n}\}_{n} is the desired Ω~\tilde{\Omega}-martingale (the proof of this assertion is identical to the proof of Theorem 3.73.7 in [21]). ∎

Remark 6.5.

The assertion of Lemma 6.1 remains true if Ω\Omega does not necessarily satisfy (2), however, φL\varphi\in L_{\infty}. The proof should be modified as follows. We choose a compact convex set CdC\subset\mathbb{R}^{d} such that φC\varphi\in C almost surely. We consider the function

ΔC(x)=sup{max(1,|xy||xz|)|x[y,z],yΩC,zclΩ~1C},xCΩ.{\Delta_{C}(x)=\sup\Big{\{}{\max\Big{(}1,\frac{|x-y|}{|x-z|}\Big{)}}\;\Big{|}\,{x\in[y,z],\ y\in\Omega\cap C,\ z\in\operatorname{cl}\tilde{\Omega}_{1}\cap C}\Big{\}},\quad x\in C\cap\Omega.} (6.10)

This function is bounded since |xz||x-z| is separated away from zero and |xy||x-y| is bounded. Now we may repeat the proof of Lemma 6.1 with ΔC\Delta_{C} in the role of Δ\Delta.

Proof of Lemma 6.2..

Let MM be a simple Ω\Omega^{*} martingale. Note that we may choose the sets CaC_{a} in Definition 4.5 to be closed simplices. Let CC be the union of such simplices over all atoms of all algebras n\mathcal{F}_{n}. In fact, this is union of a finite number of simplices. Thus, CC is a compact subset of Ω\Omega^{*}. Therefore, CC is separated from Ω1\Omega_{1} and does not intersect with the sets Ωε\Omega_{\varepsilon} for ε\varepsilon sufficiently close to 11, here

Ωε=(1ε)Ω0+εΩ1;{\Omega_{\varepsilon}=(1-\varepsilon)\Omega_{0}+\varepsilon\Omega_{1};} (6.11)

we use the standard Minkowski addition. Fix some ε\varepsilon close to 11 such that CΩε=C\cap\Omega_{\varepsilon}=\varnothing and set Ω^1=Ωε\hat{\Omega}_{1}=\Omega_{\varepsilon}. Note that the corresponding lens Ω^=clΩ0Ω^1\hat{\Omega}=\operatorname{cl}\Omega_{0}\setminus\hat{\Omega}_{1} satisfies (LABEL:StrictConvexity) and (2)111Alternatively, the set Ω^1\hat{\Omega}_{1} may be constructed with the help of Theorem 8.17 below.. What is more, clΩ1Ω^1\operatorname{cl}\Omega_{1}\subset\hat{\Omega}_{1}. Thus, Ω^1\hat{\Omega}_{1} fits into formula (5.2) and MM is an Ω^\hat{\Omega}-martingale. Consider the set

𝔚={μ(fixedΩ)|dx𝑑μ(x)clΩ^1}.{\mathfrak{W}=\Big{\{}{\mu\in\mathcal{M}(\partial_{\mathrm{fixed}}\Omega)}\;\Big{|}\,{\int\limits_{\mathbb{R}^{d}}x\,d\mu(x)\notin\operatorname{cl}\hat{\Omega}_{1}}\Big{\}}.} (6.12)

This set satisfies the two requirements on the set 𝔚\mathfrak{W} listed in (5.10). It is high time to make our choice for the martingale 𝕄\mathbb{M}. Recall the martingale MM is simple. The desired martingale is defined by the formula

𝕄n(w)=μM|w,wis an atom ofn,{\mathbb{M}_{n}(w)=\mu_{M_{\infty}|_{w}},\qquad w\ \text{is an atom of}\ \mathcal{F}_{n},} (6.13)

here μζ\mu_{\zeta} denotes the distribution of the random variable ζ\zeta; we treat M|wM_{\infty}|_{w} as a random variable on the probability space ww equipped with the measure (P(w))1P|w(P(w))^{-1}P|_{w}. Let us prove that 𝕄\mathbb{M} is an (𝔚,𝔇)(\mathfrak{W},\mathfrak{D})-martingale with 𝔚\mathfrak{W} given in (6.12). We will firstly show that 𝕄\mathbb{M} is indeed a martingale. For that we choose an arbitrary nn and an atom wnw\in\mathcal{F}_{n}. Let w1,w2,,wjn+1w_{1},w_{2},\ldots,w_{j}\in\mathcal{F}_{n+1} be the kids of ww. The martingale property of 𝕄\mathbb{M} is

P(w)μM|w=jP(wj)μM|wj.{P(w)\mu_{M_{\infty}|w}=\sum\limits_{j}P(w_{j})\mu_{M_{\infty}|_{w_{j}}}.} (6.14)

To prove this identity in measures, we may test it against a Borel set AdA\subset\mathbb{R}^{d}:

P(w)P(M|wA)=jP(wj)P(M|wjA),{P(w)P(M_{\infty}|_{w}\in A)=\sum\limits_{j}P(w_{j})P(M_{\infty}|_{w_{j}}\in A),} (6.15)

which is true. It remains to verify the third property in Definition 4.5. We need to check that any convex combination jαjμM|wj\sum_{j}\alpha_{j}\mu_{M_{\infty}|_{w_{j}}} lies inside the set 𝔚\mathfrak{W}. This means

jΩ0αjx𝑑μM|wj(x)clΩ^1.{\sum\limits_{j}\int\limits_{\partial\Omega_{0}}\alpha_{j}x\,d\mu_{M_{\infty}|_{w_{j}}}(x)\notin\operatorname{cl}\hat{\Omega}_{1}.} (6.16)

This reduces to the fact that MM is an Ω^\hat{\Omega}-martingale since

Ω0x𝑑μM|wj(x)=Mn+1(wj).{\int\limits_{\partial\Omega_{0}}x\,d\mu_{M_{\infty}|_{w_{j}}}(x)=M_{n+1}(w_{j}).} (6.17)

We apply Theorem 5.6 to 𝕄\mathbb{M} and 𝔚\mathfrak{W} and obtain a function φ𝑨\varphi\in\boldsymbol{A}^{\circ} with μφ=𝕄0\mu_{\varphi}=\mathbb{M}_{0}. It remains to notice that 𝕄0\mathbb{M}_{0} is the distribution of MM_{\infty}. ∎

7 Proof of Proposition 5.4

First, the inequality

𝔅Ω,f(x)𝔅Ω,f(x),xΩ,{\mathfrak{B}_{\Omega,f}(x)\geqslant\mathfrak{B}_{\Omega^{*},f}(x),\qquad x\in\Omega^{*},} (7.1)

follows from (4.1) since whenever GΛΩ,fG\in\Lambda_{{\Omega},{f}}, its restriction G|ΩG|_{\Omega^{*}} to Ω\Omega^{*} belongs to ΛΩ,f\Lambda_{{\Omega^{*}},{f}}. Second, to prove the reverse inequality to (7.1), it suffices to construct a function GΛΩ,fG\in\Lambda_{{\Omega},{f}} such that

G(x)=𝔅Ω,f(x),xΩ.{G(x)=\mathfrak{B}_{\Omega^{*},f}(x),\qquad x\in\Omega^{*}.} (7.2)

The construction of the function GG is fairly straightforward, however, the verification of its local concavity will take some time. To construct GG, we will use special segments Ω\ell\subset\Omega. For any xfreeΩx\in\partial_{\mathrm{free}}\Omega let us choose some transversal (see Def. 6.3) segment x\ell_{x} with the endpoint xx. Define the function GG by the formula

G(x)={𝔅Ω,f(x),xΩ;limyxyx𝔅Ω,f(y),xfreeΩ,{G(x)=\begin{cases}\mathfrak{B}_{\Omega^{*},f}(x),\qquad&x\in\Omega^{*};\\ \lim\limits_{\genfrac{}{}{0.0pt}{-2}{y\to x}{y\in\ell_{x}}}\mathfrak{B}_{\Omega^{*},f}(y),\qquad&x\in\partial_{\mathrm{free}}\Omega,\end{cases}} (7.3)

where yy approaches xx along x\ell_{x}. Note that the limit in the formula always exists (though it might be equal to -\infty) since 𝔅Ω,f|x\mathfrak{B}_{\Omega^{*},f}|_{\ell_{x}} is a concave function.

Lemma 7.1.

Let Ω\Omega satisfy (LABEL:StrictConvexity), let xfreeΩx\in\partial_{\mathrm{free}}\Omega. Let \ell be a transversal segment with the endpoint xx. Let sΩs\subset\Omega be another segment with the endpoint xx. Then, the convex hull of \ell and ss also belongs to Ω\Omega entirely.

Proof.

It suffices to prove that the said convex hull is disjoint with Ω1\Omega_{1}. Assume the contrary, let yΩ1y\in\Omega_{1} lie in the convex hull of \ell and ss. Let zz be a point on the continuation of \ell over the point xx that is sufficiently close to xx. Since \ell is a transversal segment, zΩ1z\in\Omega_{1}. The segment [z,y][z,y] then lies inside Ω1\Omega_{1} since Ω1\Omega_{1} is convex. On the other hand, [z,y][z,y], clearly, intersects ss, which contradicts sΩs\subset\Omega. ∎

Remark 7.2.

In fact, the said convex hull lies inside Ω,\Omega^{*}, except for the point xx itself.

Proof of Proposition 5.4.

As we have said, it suffices to show GG given in (7.3) is locally concave on Ω\Omega (in particular, we need to verify that GG does not attain the value -\infty). The verification of local concavity consists of checking the inequalities

G(x)αG(a)+βG(b),x=αa+βb,α+β=1,α,β>0,[a,b]Ω.{G(x)\geqslant\alpha G(a)+\beta G(b),\qquad x=\alpha a+\beta b,\ \alpha+\beta=1,\ \alpha,\beta>0,\quad[a,b]\subset\Omega.} (7.4)

We are interested in the cases where one of the points x,ax,a, or bb lies on freeΩ\partial_{\mathrm{free}}\Omega. Let a,b,xa,b,x be distinct points.

Case xfreeΩx\in\partial_{\mathrm{free}}\Omega.

Note that in this case aa and bb do not lie on freeΩ\partial_{\mathrm{free}}\Omega by (LABEL:StrictConvexity). Thus, we may assume they are interior points of Ω\Omega. Consider the segment x\ell_{x}, some point yxy\in\ell_{x} (let yxy\neq x), and the points aγa_{\gamma}bγb_{\gamma}, and xγx_{\gamma} given by the rule

zγ=γy+(1γ)z,γ(0,1],{z_{\gamma}=\gamma y+(1-\gamma)z,\quad\gamma\in(0,1],} (7.5)

here zz stands either for aa, or for bb, or for xx. Note that [aγ,bγ]Ω[a_{\gamma},b_{\gamma}]\subset\Omega^{*} by Lemma 7.1 (with Remark 7.2). Thus,

𝔅Ω,f(xγ)α𝔅Ω,f(aγ)+β𝔅Ω,f(bγ).{\mathfrak{B}_{\Omega^{*},f}(x_{\gamma})\geqslant\alpha\mathfrak{B}_{\Omega^{*},f}(a_{\gamma})+\beta\mathfrak{B}_{\Omega^{*},f}(b_{\gamma}).} (7.6)

Note that zγzz_{\gamma}\to z as γ0\gamma\to 0. Thus, (7.4) is proved in this case since 𝔅Ω,f\mathfrak{B}_{\Omega^{*},f} is continuous at aa and bb.

The reasoning above also shows that G(x)>G(x)>-\infty. Indeed, we need to choose some points a,bΩa,b\in\Omega^{*} such that x[a,b]x\in[a,b] and use (7.4).

Case afreeΩa\in\partial_{\mathrm{free}}\Omega.

Consider the segment a\ell_{a}. Let yay\in\ell_{a}. We consider the points aγa_{\gamma}, bγb_{\gamma}, and xγx_{\gamma} defined by the same formula (7.5). By Lemma 7.1 (Remark 7.2), these points lie inside Ω\Omega^{*} together with the segment [aγ,bγ][a_{\gamma},b_{\gamma}]. Thus, (7.6) holds true and (7.4) follows by a limit argument. ∎

Remark 7.3.

One may prove that the definition of the function GG by (7.3) does not depend on the particular choice of transversal segments x,\ell_{x}, xfreeΩx\in\partial_{\mathrm{free}}\Omega.

8 Proof of Theorem 4.4

Theorem 4.4 immediately follows from (5.5), Theorem 5.3, and Proposition 5.4 once we prove the ’extension’ theorem below.

Theorem 8.1.

Let Ω\Omega be a lens that satisfies (LABEL:StrictConvexity) and (2). Assume freeΩ\partial_{\mathrm{free}}\Omega is C2C^{2}-smooth and ff is an arbitrary function. Then,

𝔅Ω,f(x)=inf{𝔅Ω~,f(x)|Ω~is an extension ofΩ},xΩ.{\mathfrak{B}_{\Omega,f}(x)=\inf\Big{\{}{\mathfrak{B}_{\tilde{\Omega},f}(x)}\;\Big{|}\,{\,\tilde{\Omega}\ \text{is an extension of}\ \Omega}\Big{\}},\quad x\in\Omega.} (8.1)

When proving Theorem 8.1, we may assume, without loss of generality, that 𝔅Ω,f\mathfrak{B}_{\Omega,f} is finite. Then, the identity (8.1) may be reformulated: for any xΩx\in\Omega and any ε>0\varepsilon>0 there exists an extension Ω~\tilde{\Omega} such that

𝔅Ω,f(x)𝔅Ω~,f(x)𝔅Ω,f(x)+ε.{\mathfrak{B}_{\Omega,f}(x)\leqslant\mathfrak{B}_{\tilde{\Omega},f}(x)\leqslant\mathfrak{B}_{\Omega,f}(x)+\varepsilon.} (8.2)

Note that Ω~\tilde{\Omega} may depend on ε\varepsilon and xx. In the case of smooth boundaries and sufficiently regular ff, we will prove a stronger statement, which is the main step toward the proof of Theorem 8.1.

Theorem 8.2.

Let Ω\Omega be a lens that satisfies (LABEL:StrictConvexity) and (2). Assume the boundaries fixedΩ,\partial_{\mathrm{fixed}}\Omega, freeΩ\partial_{\mathrm{free}}\Omega are C2C^{2}-smooth and the function ff is C2C^{2}-smooth as well. Then, for any ε>0\varepsilon>0 there exists an extension Ω~\tilde{\Omega} such that

𝔅Ω,f(x)𝔅Ω~,f(x)𝔅Ω,f(x)+ε{\mathfrak{B}_{\Omega,f}(x)\leqslant\mathfrak{B}_{\tilde{\Omega},f}(x)\leqslant\mathfrak{B}_{\Omega,f}(x)+\varepsilon} (8.3)

for any xΩx\in\Omega.

The proof of this theorem follows the proof of Theorem 4.14.1 in [21] with some modifications. These modifications do not require new ideas, however, some re-phrasing is needed to work in d\mathbb{R}^{d} with arbitrary dd instead of 2\mathbb{R}^{2}. The condition fC2f\in C^{2} may be immediately replaced with ff being merely continuous by approximation in the uniform norm. Here we have used the obvious inequalities

𝔅Ω,f𝔅Ω,g𝔅Ω,f+ε=𝔅Ω,f+ε,\mathfrak{B}_{\Omega,f}\leqslant\mathfrak{B}_{\Omega,g}\leqslant\mathfrak{B}_{\Omega,f+\varepsilon}=\mathfrak{B}_{\Omega,f}+\varepsilon,

provided fgf+εf\leqslant g\leqslant f+\varepsilon.

Corollary 8.3.

The statement of Theorem 8.2 holds true if fC(fixedΩ)f\in C(\partial_{\mathrm{fixed}}\Omega).

The method of [21] was to perturb the function 𝔅Ω,f\mathfrak{B}_{\Omega,f} a little bit to make it strongly concave and then to extend it through the free boundary. The reasoning naturally splits into two steps: first, we study the boundary behavior of minimal locally concave functions and, second, use this structure to construct the extension.

8.1 Boundary behavior of minimal locally concave functions

Definition 8.4.

Let ωd\omega\subset\mathbb{R}^{d}. We say that two points x,yωx,y\in\omega see each other if [x,y]ω[x,y]\subset\omega. The set

Visxω={yω|xandysee each other inω}{\operatorname{Vis}_{x}^{\omega}=\{{y\in\omega}\;|\,{x\ \text{and}\ y\ \text{see each other in}\ \omega}\}} (8.4)

is called the set of points visible from xx.

We will simply write Visx\operatorname{Vis}_{x} instead of Visxω\operatorname{Vis}_{x}^{\omega} when the ambient set ω\omega is clear from the context.

Proposition 8.5.

Let Ω\Omega be a lens that satisfies (LABEL:StrictConvexity) and (2). The set Visx\operatorname{Vis}_{x} is compact whenever xfreeΩx\in\partial_{\mathrm{free}}\Omega. The diameter of Visx\operatorname{Vis}_{x} is uniformly bounded when xx runs through a compact subset of freeΩ\partial_{\mathrm{free}}\Omega.

Proof.

It is clear that the set Visx\operatorname{Vis}_{x} is closed (since Ω\Omega is closed). So we only need to prove the second assertion. Assume the contrary, let there exist a sequence {xn}n\{x_{n}\}_{n} with values in a compact subset of freeΩ\partial_{\mathrm{free}}\Omega and a sequence {yn}\{y_{n}\} such that |yn||y_{n}|\to\infty and [xn,yn]Ω[x_{n},y_{n}]\subset\Omega. Without loss of generality, xnxfreeΩx_{n}\to x\in\partial_{\mathrm{free}}\Omega and yn/|yn|ySd1y_{n}/|y_{n}|\to y\in S^{d-1}. Then, by the closedness of Ω\Omega, the ray x++yx+\mathbb{R}_{+}y lies inside Ω\Omega. By (2), there exists zΩ1z\in\Omega_{1} such that z++yΩ1z+\mathbb{R}_{+}y\subset\Omega_{1}. This contradicts the strict convexity of Ω1\Omega_{1} since xclΩ1x\in\operatorname{cl}\Omega_{1}. ∎

Remark 8.6.

The set Visx\operatorname{Vis}_{x} is not necessarily bounded if xfreeΩx\notin\partial_{\mathrm{free}}\Omega as the left picture on Figure 2 shows.

We will often use the following simple principle (compare with Fact B.3B.3 in [21]).

Fact 8.7.

Let ω\omega be a strictly convex subset of d\mathbb{R}^{d}. Suppose that there are positive rr and RR and a point pωLp\in\omega\cap L such that Br(p)ωB_{r}(p)\subset\omega, BR(p)ωLB_{R}(p)\supset\omega\cap L. Then for C=r+RrC=\frac{r+R}{r} and any yωy\in\partial\omega one may find zωLz\in\partial\omega\cap L such that |yz|Cdist(y,L)|y-z|\leqslant C\operatorname{dist}(y,L).

For the hint to the proof see Figure 7.

Refer to caption
Figure 7: Hint to the proof of Fact 8.7.
Definition 8.8.

Let xωdx\in\omega\subset\mathbb{R}^{d} and let G:ωG\colon\omega\to\mathbb{R}. The set of linear functions LL such that G(y)G(x)+L(yx)G(y)\leqslant G(x)+L(y-x) holds for any yVisxωy\in\operatorname{Vis}_{x}^{\omega} is called the superdifferential of GG at xx. We will denote it by ðG|x\eth G|_{x}.

Fact 8.9.

Let ωd\omega\subset\mathbb{R}^{d} and let G:ωG\colon\omega\to\mathbb{R}. Assume for every xωx\in\omega the superdifferential ðG|x\eth G|_{x} of GG at xx is non-empty. Then, the function GG is locally concave on ω\omega.

Every concave function has a non-empty superdifferetial at interior points of its domain (see, e. g., Section 2323 in [14]). One may ask whether the superdifferential is non-empty for every xx in ω\omega provided GG is locally concave. The answer to this question is negative in general (consider the domain ω\omega formed by three lines passing through the origin in 2\mathbb{R}^{2}). However, for some good domains (lenses among them) and sufficiently good functions, the answer is positive.

Proposition 8.10.

Let Ω\Omega be a lens that satisfies (LABEL:StrictConvexity) and (2). Assume that the boundaries of Ω\Omega are C1C^{1}-smooth. Let f:fixedΩf\colon\partial_{\mathrm{fixed}}\Omega\to\mathbb{R} be a locally Lipschitz function such that 𝔅Ω,f\mathfrak{B}_{\Omega,f} is finite. For any xfreeΩx\in\partial_{\mathrm{free}}\Omega there exists a linear function L[𝔅,x]L[\mathfrak{B},x] such that

𝔅Ω,f(x)+L[𝔅,x](yx)𝔅Ω,f(y){\mathfrak{B}_{\Omega,f}(x)+L[\mathfrak{B},x](y-x)\geqslant\mathfrak{B}_{\Omega,f}(y)} (8.5)

for all yVisxΩy\in\operatorname{Vis}_{x}^{\Omega}. In other words, ð𝔅|x\eth\mathfrak{B}|_{x} is non-empty.

Remark 8.11.

One may replace the minimal locally concave function 𝔅\mathfrak{B} with an arbitrary locally concave function GG provided GG is locally Lipschitz.

Proof.

Without loss of generality, we may assume x=0x=0 and

T0Ω1={ydyd=0}.{T_{0}\Omega_{1}=\{y\in\mathbb{R}^{d}\mid y_{d}=0\}.} (8.6)

The symbol TzΩ1T_{z}\Omega_{1} denotes the tangent plane to Ω1\Omega_{1} at the point zz. We also assume yd>0y_{d}>0 on Ω1\Omega_{1}. By concavity of 𝔅|yd=0\mathfrak{B}|_{y_{d}=0}, there exists a linear function :d1\ell\colon\mathbb{R}^{d-1}\to\mathbb{R} such that

𝔅(z1,z2,,zd1,0)𝔅(0)+(z){\mathfrak{B}(z_{1},z_{2},\ldots,z_{d-1},0)\leqslant\mathfrak{B}(0)+\ell(z)} (8.7)

for any zd1z\in\mathbb{R}^{d-1} satisfying (z,0)Ω(z,0)\in\Omega. Let π\pi denote the orthogonal projection onto {ydyd=0}\{y\in\mathbb{R}^{d}\mid y_{d}=0\}. Let

a=sup{𝔅(y)𝔅(0)(π[y])yd|yVis0,yd<0}.{a=\sup\Big{\{}{\frac{\mathfrak{B}(y)-\mathfrak{B}(0)-\ell(\pi[y])}{-y_{d}}}\;\Big{|}\,{y\in\operatorname{Vis}_{0},y_{d}<0}\Big{\}}.} (8.8)

We claim that

a=sup{𝔅(y)𝔅(0)(π[y])yd|yVis0fixedΩ,yd<0}=sup{f(y)𝔅(0)(π[y])yd|yVis0fixedΩ,yd<0}.{a=\sup\Big{\{}{\frac{\mathfrak{B}(y)-\mathfrak{B}(0)-\ell(\pi[y])}{-y_{d}}}\;\Big{|}\,{y\in\operatorname{Vis}_{0}\cap\partial_{\mathrm{fixed}}\Omega,y_{d}<0}\Big{\}}\\ =\sup\Big{\{}{\frac{f(y)-\mathfrak{B}(0)-\ell(\pi[y])}{-y_{d}}}\;\Big{|}\,{y\in\operatorname{Vis}_{0}\cap\partial_{\mathrm{fixed}}\Omega,y_{d}<0}\Big{\}}.} (8.11)

Indeed, (8.11) clearly does not exceed (8.8). On the other hand, if

𝔅(y)byd+𝔅(0)+(π[y]){\mathfrak{B}(y)\leqslant-by_{d}+\mathfrak{B}(0)+\ell(\pi[y])} (8.12)

for all yVis0fixedΩy\in\operatorname{Vis}_{0}\cap\partial_{\mathrm{fixed}}\Omega and some bb\in\mathbb{R}, then the same inequality holds true for all yVis0y\in\operatorname{Vis}_{0} by minimality of 𝔅\mathfrak{B}. Indeed, if this is not the case, the function

y{𝔅(y),yΩVis0,min(𝔅(y),byd+𝔅(0)+(π[y])),yΩVis0{y\mapsto\begin{cases}\mathfrak{B}(y),&y\in\Omega\setminus\operatorname{Vis}_{0},\\ \min(\mathfrak{B}(y),-by_{d}+\mathfrak{B}(0)+\ell(\pi[y])),&y\in\Omega\cap\operatorname{Vis}_{0}\end{cases}} (8.13)

lies in ΛΩ,f\Lambda_{{\Omega},{f}} and is smaller than 𝔅\mathfrak{B}, which is a contradiction. Thus, the two supremums on the right hand sides of (8.8) and (8.11) coincide.

Now let us prove that the local Lipschitz property of ff implies aa is a finite quantity. Let {yn}n\{y^{n}\}_{n} be a sequence of points in Vis0fixedΩ\operatorname{Vis}_{0}\cap\partial_{\mathrm{fixed}}\Omega that realizes the supremum in (8.11). Assume ynyy^{n}\to y^{*}. If yd0y^{*}_{d}\neq 0, then aa is finite.

Consider the case yd=0y_{d}^{*}=0. By Fact 8.7, there exist points y~nΩ0\tilde{y}^{n}\in\partial\Omega_{0} such that y~dn=0\tilde{y}^{n}_{d}=0 and

|yny~n|ydn.{|y^{n}-\tilde{y}^{n}|\lesssim y_{d}^{n}.} (8.14)

Then,

f(yn)𝔅(0)(π[yn])=f(y~n)𝔅(0)(π[y~n])+O(|ydn|)O(|ydn|),{f(y^{n})-\mathfrak{B}(0)-\ell(\pi[y^{n}])=f(\tilde{y}^{n})-\mathfrak{B}(0)-\ell(\pi[\tilde{y}^{n}])+O(|y_{d}^{n}|)\leqslant O(|y_{d}^{n}|),} (8.15)

the constant in OO depends on the Lipschitz constant of ff and the constant CC in Fact 8.7. The last relation proves aa is finite. We may set

L[𝔅,0](y)=(π[y])ayd.{L[\mathfrak{B},0](y)=\ell(\pi[y])-ay_{d}.} (8.16)

Remark 8.12.

It follows from construction that the coefficients of the linear function L[𝔅,x]L[\mathfrak{B},x] are uniformly bounded when xx runs through a compact subset of freeΩ\partial_{\mathrm{free}}\Omega.

Proposition 8.13.

Let Ω\Omega be a lens that satisfies (LABEL:StrictConvexity) and (2). Assume that the boundaries of Ω\Omega are C2C^{2}-smooth. Let f:fixedΩf\colon\partial_{\mathrm{fixed}}\Omega\to\mathbb{R} be a C2C^{2}-smooth function such that 𝔅Ω,f\mathfrak{B}_{\Omega,f} is finite. Let L[𝔅,x]L[\mathfrak{B},x] be the linear functions constructed by formulas (8.11) and (8.16), here xfreeΩx\in\partial_{\mathrm{free}}\Omega. Then, there exists a point exfixedΩVisxe_{x}\in\partial_{\mathrm{fixed}}\Omega\cap\operatorname{Vis}_{x} such that

|𝔅(x)+L[𝔅,x](zx)f(z)|=O(|exz|2),zfixedΩVisx.{|\mathfrak{B}(x)+L[\mathfrak{B},x](z-x)-f(z)|=O(|e_{x}-z|^{2}),\qquad z\in\partial_{\mathrm{fixed}}\Omega\cap\operatorname{Vis}_{x}.} (8.17)

The constant in this inequality is uniform when xx runs through a compact set on freeΩ\partial_{\mathrm{free}}\Omega.

Proof.

Similar to the proof of Proposition 8.10, we assume x=0x=0,

T0Ω1={ydyd=0},{T_{0}\Omega_{1}=\{y\in\mathbb{R}^{d}\mid y_{d}=0\},} (8.18)

and yd>0y_{d}>0 on Ω1\Omega_{1}.

Let ee be a limit point of the sequence {yn}n\{y^{n}\}_{n} that realizes the supremum in (8.11). We will show that the choice ex:=ee_{x}:=e with x=0x=0 fulfills (8.17). Let us first prove 𝔅(0)+L[𝔅,0](e)=f(e)\mathfrak{B}(0)+L[\mathfrak{B},0](e)=f(e). If ed0e_{d}\neq 0, then this follows from the definition of L[𝔅,0]L[\mathfrak{B},0]. In the case ed=0e_{d}=0, we have 𝔅(e)=𝔅(0)+(e1,e2,,ed1)\mathfrak{B}(e)=\mathfrak{B}(0)+\ell(e_{1},e_{2},\ldots,e_{d-1}); if this identity does not hold, the supremum aa equals -\infty, which is definitely false. Therefore, 𝔅(0)+L[𝔅,0](e)=f(e)\mathfrak{B}(0)+L[\mathfrak{B},0](e)=f(e).

Let us consider the C2C^{2}-smoth function

A(z)=f(z)𝔅(0)L[𝔅,0](z),zfixedΩ.{A(z)=f(z)-\mathfrak{B}(0)-L[\mathfrak{B},0](z),\qquad z\in\partial_{\mathrm{fixed}}\Omega.} (8.19)

We have A(e)=0A(e)=0. If ed<0e_{d}<0, then AA is non-positive in a neighborhood of ee. Therefore, ee is a local maximum of AA, and (8.17) follows.

If ed=0e_{d}=0, then we know that A(z)0A(z)\leqslant 0 for zfixedΩz\in\partial_{\mathrm{fixed}}\Omega such that zd0z_{d}\leqslant 0. From (8.11) we know that for any δ>0\delta>0 we have A(yn)>δydnA(y^{n})>\delta y^{n}_{d} for sufficiently large nn. Let us consider the hyperplane that contains the intersection Te,Ω0T0,Ω1T_{e,\Omega_{0}}\cap T_{0,\Omega_{1}} and is parallel to the ydy_{d}-axis. Without loss of generality, we may assume this hyperplane is {yd1=0}\{y_{d-1}=0\}. Let A~\tilde{A} be the projection of AA onto this plane defined near ee: if π:fixedΩ{yd1=0}\pi\colon\partial_{\mathrm{fixed}}\Omega\to\{y_{d-1}=0\} is the orthogonal projection, then

A~(z~)=A(π1[z~]),z~d1=0,z~ is close to e.{\tilde{A}(\tilde{z})=A(\pi^{-1}[\tilde{z}]),\qquad\tilde{z}_{d-1}=0,\quad\tilde{z}\text{ is close to }e.} (8.20)

The function A~\tilde{A} is C2C^{2}-smooth in a neighborhood of ee, A~(e)=0\tilde{A}(e)=0, and A~(z)0\tilde{A}(z)\leqslant 0 when z~d0\tilde{z}_{d}\leqslant 0. What is more, there is a sequence y~n=(y~1n,,y~d2n,0,y~dn)\tilde{y}^{n}=(\tilde{y}^{n}_{1},\dots,\tilde{y}^{n}_{d-2},0,\tilde{y}^{n}_{d}) such that y~ne\tilde{y}^{n}\to e, y~dn<0\tilde{y}^{n}_{d}<0, and

0A~(y~n)>δy~dn{0\geqslant\tilde{A}(\tilde{y}^{n})>\delta\tilde{y}^{n}_{d}} (8.21)

for any fixed δ>0\delta>0 provided nn is sufficiently large.

Let us prove that A~(e)=0\nabla\tilde{A}(e)=0. The restriction of A~\tilde{A} to the section yd=0y_{d}=0 attains its maximum at ee, therefore, we only need to check that

A~yd(e)=0.{\frac{\partial\tilde{A}}{\partial y_{d}}(e)=0.} (8.22)

Note that the derivative on the left hand side cannot be negative since A~(z)0\tilde{A}(z)\leqslant 0 when z~d0\tilde{z}_{d}\leqslant 0. Let us prove it is non-positive. If this is not the case, then ydA~(e)>2δ\frac{\partial}{\partial y_{d}}\tilde{A}(e)>2\delta for some δ>0\delta>0, and, by C1C^{1}-continuity, it follows that ydA~>δ\frac{\partial}{\partial y_{d}}\tilde{A}>\delta in a neighborhood of ee. Then,

A~(y~n)A~(y~1n,,y~d2n,0,0)+δy~dnδy~dn,{\tilde{A}(\tilde{y}^{n})\leqslant\tilde{A}(\tilde{y}_{1}^{n},\dots,\tilde{y}_{d-2}^{n},0,0)+\delta\tilde{y}^{n}_{d}\leqslant\delta\tilde{y}^{n}_{d},} (8.23)

which contradicts (8.21). Therefore, A~(e)=0\nabla\tilde{A}(e)=0. Then, for any zfixedΩz\in\partial_{\mathrm{fixed}}\Omega sufficiently close to ee we have

A(z)=A~(z~)=O(|z~e|2)=O(|ze|2),z~=π[z],{A(z)=\tilde{A}(\tilde{z})=O(|\tilde{z}-e|^{2})=O(|z-e|^{2}),\qquad\tilde{z}=\pi[z],} (8.24)

which proves (8.17). ∎

Remark 8.14.

The condition fC2f\in C^{2} is superfluous. One may replace it with C1,1C^{1,1} as it can be seen from the proof.

Proposition 8.15.

Let Ω\Omega be a lens that satisfies (LABEL:StrictConvexity) and (2). Assume that the boundaries of Ω\Omega are C2C^{2}-smooth. Let f:fixedΩf\colon\partial_{\mathrm{fixed}}\Omega\to\mathbb{R} be a C2C^{2}-smooth function such that 𝔅Ω,f\mathfrak{B}_{\Omega,f} is finite. Let L[𝔅,x]L[\mathfrak{B},x] be the linear functions constructed by formulas (8.11) and (8.16), here xfreeΩx\in\partial_{\mathrm{free}}\Omega. Then,

𝔅(y)𝔅(x)+L[𝔅,x](yx)+O(|xy|3),x,yfreeΩ.{\mathfrak{B}(y)\leqslant\mathfrak{B}(x)+L[\mathfrak{B},x](y-x)+O(|x-y|^{3}),\qquad x,y\in\partial_{\mathrm{free}}\Omega.} (8.25)

The implicit constants hidden by the OO-notation are uniform when xx and yy run through a compact set on freeΩ\partial_{\mathrm{free}}\Omega.

Proof.

By compactness argument, it suffices to consider the case where yy lies in a neighborhood of xx. Similar to the proof of Proposition 8.10, we assume x=0x=0,

T0Ω1={ydyd=0},{T_{0}\Omega_{1}=\{y\in\mathbb{R}^{d}\mid y_{d}=0\},} (8.26)

and yd>0y_{d}>0 on Ω1\Omega_{1}. Let freeΩ\partial_{\mathrm{free}}\Omega be defined as the graph of the function h:Uh\colon U\to\mathbb{R}, where Ud1U\subset\mathbb{R}^{d-1} is a neighborhood of the origin. Consider the cone

Cy={zd|zdydC|yd¯||zd¯yd¯|},y=(yd¯,yd),z=(zd¯,zd),yd¯,zd¯d1,{C_{y}=\Big{\{}{z\in\mathbb{R}^{d}}\;\Big{|}\,{z_{d}-y_{d}\leqslant-C|y_{\bar{d}}||z_{\bar{d}}-y_{\bar{d}}|}\Big{\}},\quad y=(y_{\bar{d}},y_{d}),z=(z_{\bar{d}},z_{d}),\quad y_{\bar{d}},z_{\bar{d}}\in\mathbb{R}^{d-1},} (8.27)

where yfreeΩy\in\partial_{\mathrm{free}}\Omega is a point in a neighborhood of xx and CC is a sufficiently large constant. See Fig. 8 for visualization.

Refer to caption
Refer to caption
Figure 8: Illustration to the proof of Proposition 8.15. The first picture shows the cone CyC_{y} in the d=2d=2 case. The second picture shows the two-dimensional section that arises in the construction of ee^{\prime}.

The cone CyC_{y} does not intersect Ω1\Omega_{1}.

Let us prove this claim. Since the tangent plane to Ω1\Omega_{1} at the point (yd¯,h(yd¯))(y_{\bar{d}},h(y_{\bar{d}})) is described by the equation zdyd=h(yd¯),zd¯yd¯z_{d}-y_{d}=\langle{\nabla h(y_{\bar{d}})},{z_{\bar{d}}-y_{\bar{d}}}\rangle, it suffices to prove the inequality

zdydh(yd¯),zd¯yd¯,zCy.{z_{d}-y_{d}\leqslant\langle{\nabla h(y_{\bar{d}})},{z_{\bar{d}}-y_{\bar{d}}}\rangle,\qquad z\in C_{y}.} (8.28)

We estimate

zdydzCyC|yd¯||zd¯yd¯||h(yd¯)||zd¯yd¯|h(yd¯),zd¯yd¯.{z_{d}-y_{d}\stackrel{{\scriptstyle\scriptscriptstyle z\in C_{y}}}{{\leqslant}}-C|y_{\bar{d}}||z_{\bar{d}}-y_{\bar{d}}|\leqslant-|\nabla h(y_{\bar{d}})||z_{\bar{d}}-y_{\bar{d}}|\leqslant\langle{\nabla h(y_{\bar{d}})},{z_{\bar{d}}-y_{\bar{d}}}\rangle.} (8.29)

The inequality |h(yd¯)|C|yd¯||\nabla h(y_{\bar{d}})|\leqslant C|y_{\bar{d}}| for all xx in a compact set and all yy in a neighborhood of xx follows from the C2C^{2}-smoothness assumption about freeΩ\partial_{\mathrm{free}}\Omega. Thus, CyC_{y} indeed does not intersect Ω1\Omega_{1}.

A similar reasoning shows that

|zy||yd¯|,providedzCyandzd0.{|z-y|\lesssim|y_{\bar{d}}|,\quad\text{provided}\quad z\in C_{y}\ \text{and}\ z_{d}\geqslant 0.} (8.30)

In particular, in such a case zz cannot belong to fixedΩ\partial_{\mathrm{fixed}}\Omega.

Let exe_{x} be the point constructed in Proposition 8.13. We consider two cases: exCye_{x}\in C_{y} and exCye_{x}\notin C_{y}.

Case exCye_{x}\in C_{y}.

In this case, exVisye_{x}\in\operatorname{Vis}_{y} and the segment [ex,y][e_{x},y] intersects {zzd=0}\{z\mid z_{d}=0\} (since exe_{x} lies below the zd=0z_{d}=0 plane by (8.30)). Denote the point of intersection by PP. Then, the function GG defined by

G(z)=𝔅(z)𝔅(0)L[𝔅,0](z),zΩ,{G(z)=\mathfrak{B}(z)-\mathfrak{B}(0)-L[\mathfrak{B},0](z),\qquad z\in\Omega,} (8.31)

is concave on this segment, attains the value 0 at exe_{x} and is non-positive at PP. Thus, it is non-positive at yy as well, and we have proved 𝔅(y)𝔅(x)+L[𝔅,x](yx)\mathfrak{B}(y)\leqslant\mathfrak{B}(x)+L[\mathfrak{B},x](y-x), which is stronger than (8.25).

Consider the case exCye_{x}\notin C_{y}.

There exists a point efixedΩCye^{\prime}\in\partial_{\mathrm{fixed}}\Omega\cap C_{y} such that |exe||yd¯||e_{x}-e^{\prime}|\lesssim|y_{\bar{d}}|. For example, such a point can be found in the two-dimensional plane that passes through yyexe_{x}, and is orthogonal to the plane zd=0z_{d}=0 (here we use Proposition 8.5; see also the second drawing on Fig. 8). The segment [e,y][e^{\prime},y] intersects the plane {zzd=0}\{z\mid z_{d}=0\} at the point PP. By (8.30),

|yP||yd¯|and|ye|1.{|y-P|\lesssim|y_{\bar{d}}|\quad\text{and}\quad|y-e^{\prime}|\gtrsim 1.} (8.32)

By the concavity of GG (see (8.31) for the definition of GG),

G(P)|Pe||ey|G(y)+|Py||ey|G(e).{G(P)\geqslant\frac{|P-e^{\prime}|}{|e^{\prime}-y|}G(y)+\frac{|P-y|}{|e^{\prime}-y|}G(e^{\prime}).} (8.33)

Since G(P)0G(P)\leqslant 0,

G(y)|Py||Pe|G(e)(8.17)|Py||Pe||eex|2(8.32)|yd¯||yd¯|2=|yd¯|3.{G(y)\leqslant-\frac{|P-y|}{|P-e^{\prime}|}G(e^{\prime})\stackrel{{\scriptstyle\scriptscriptstyle{\eqref{C2Estimate}}}}{{\lesssim}}\frac{|P-y|}{|P-e^{\prime}|}|e^{\prime}-e_{x}|^{2}\stackrel{{\scriptstyle\scriptscriptstyle{\eqref{Eq829}}}}{{\lesssim}}|y_{\bar{d}}|\cdot|y_{\bar{d}}|^{2}=|y_{\bar{d}}|^{3}.} (8.34)

Corollary 8.16.

Let Ω\Omega be a lens that satisfies (LABEL:StrictConvexity) and (2). Assume that the boundaries of Ω\Omega are C2C^{2}-smooth. Let f:fixedΩf\colon\partial_{\mathrm{fixed}}\Omega\to\mathbb{R} be a C2C^{2}-smooth function such that 𝔅Ω,f\mathfrak{B}_{\Omega,f} is finite. For any ε>0\varepsilon>0 there exists a function G:ΩG\colon\Omega\to\mathbb{R} that is locally concave on Ω\Omega and satisfies the inequalities

𝔅Ω,f(x)G(x)𝔅Ω,f(x)+ε,xΩ;{\ \mathfrak{B}_{\Omega,f}(x)\leqslant G(x)\leqslant\mathfrak{B}_{\Omega,f}(x)+\varepsilon,\qquad x\in\Omega;} (8.35)

moreover, for any xfreeΩx\in\partial_{\mathrm{free}}\Omega there exists a linear function L[G,x]L[G,x] such that for any compact set KfreeΩK\subset\partial_{\mathrm{free}}\Omega the inequality

G(y)G(x)+L[G,x](yx)cK|xy|2,x,yK,|xy|<εK,{G(y)\leqslant G(x)+L[G,x](y-x)-c_{K}|x-y|^{2},\qquad x,y\in K,\quad|x-y|<\varepsilon_{K},} (8.36)

holds true with the positive constants cKc_{K} and εK\varepsilon_{K} depending on KK only. The coefficients of the linear functions L[G,x]L[G,x] are uniformly bounded when xx runs through a compact subset of freeΩ\partial_{\mathrm{free}}\Omega.

Proof.

Let g:Ω[0,1]g\colon\Omega\to[0,1] be a strongly convex function (i. e., it is C2C^{2}-smooth and with everywhere strictly positive definite Hesse matrix); such a function is easy to construct using (LABEL:StrictConvexity) and (2) (see Lemma 4.54.5 in [21]). We set

G(x)=𝔅Ω,f(x)+εg(x),{G(x)=\mathfrak{B}_{\Omega,f}(x)+\varepsilon g(x),} (8.37)

which leads to the natural choice of the linear functions L[G,x]L[G,x]:

L[G,x]=L[𝔅,x]+g(x),,xfreeΩ.{L[G,x]=L[\mathfrak{B},x]+\langle{\nabla g(x)},{\cdot}\rangle,\quad x\in\partial_{\mathrm{free}}\Omega.} (8.38)

The inequality (8.36) then follows from Proposition 8.15. ∎

8.2 Construction of the extension

Before we pass to the details, we will survey our method for construction of extensions. Assume GG is a locally concave function on some set ω\omega and assume its superdifferential is non-empty at each point yωy\in\omega. Let ω~\tilde{\omega} be a set such that ωω~\omega\subset\tilde{\omega}. Then, we can extend GG to ω~\tilde{\omega} by the formula

G~(x)=inf{G(y)+L(xy)|yω,LðG|y and yVisxω~},xω~.{\tilde{G}(x)=\inf\Big{\{}{G(y)+L(x-y)}\;\Big{|}\,{y\in\omega,\ L\in\eth G|_{y}\text{ and }y\in\operatorname{Vis}_{x}^{\tilde{\omega}}}\Big{\}},\quad x\in\tilde{\omega}.} (8.39)

The formula is not completely rigorous, because sometimes it is convenient to take all linear functions from the superdifferential of GG at yy, sometimes we may pick only one. There are two questions concerning formula (8.39): when do we obtain a finite function and when is it locally concave on ω~\tilde{\omega}? We will give two simple sufficient conditions.

The function G~\tilde{G} is finite when all LðG|yL\in\eth G|_{y} are uniformly bounded when yy runs through a compact set and for any xω~x\in\tilde{\omega}, the set Visxω~\operatorname{Vis}_{x}^{\tilde{\omega}} is compact.

The function G~\tilde{G} is locally concave provided for any xω~x\in\tilde{\omega} there exists a neighborhood UxU_{x} (in the relative topology of ω~\tilde{\omega}) and a point yωy\in\omega such that yViszω~y\in\operatorname{Vis}_{z}^{\tilde{\omega}} for any zUxz\in U_{x} and

G~(x)=G(y)+L(yx) for some LðG|y.{\tilde{G}(x)=G(y)+L(y-x)\quad\text{ for some }L\in\eth G|_{y}.} (8.40)

Then ðG|x\eth G|_{x}\neq\varnothing and the local concavity follows from Fact 8.9.

We will also need two theorems about approximation. For more details and proofs, see [2] (based on earlier work in [1]).

Theorem 8.17.

Let Ω\Omega be a non-empty strictly convex open proper subset of d\mathbb{R}^{d}. Let UU be an open set that contains clΩ\operatorname{cl}\Omega. There exists another open set Ω\Omega^{\prime} such that clΩΩ\operatorname{cl}\Omega\subset\Omega^{\prime}, clΩU\operatorname{cl}\Omega^{\prime}\subset U, and Ω\Omega^{\prime} is a strictly convex set with real-analytic boundary.

Theorem 8.18.

Let Ω\Omega be a non-empty strictly convex open proper subset of d\mathbb{R}^{d}. Let VV be a closed set that lies inside Ω\Omega. There exists another open set Ω\Omega^{\prime} such that VΩclΩΩV\subset\Omega^{\prime}\subset\operatorname{cl}\Omega^{\prime}\subset\Omega and Ω\Omega^{\prime} is a strictly convex set with real-analytic boundary.

Proposition 8.19.

Let Ω\Omega be a non-empty strictly convex open proper subset of d\mathbb{R}^{d}. Let ρ:Ω(0,1]\rho\colon\partial\Omega\to(0,1] be a continuous function. There exists a strictly convex open set Ω\Omega^{\prime} with real-analytic boundary such that clΩΩ\operatorname{cl}\Omega^{\prime}\subset\Omega and if x,yΩx,y\in\partial\Omega see each other in clΩΩ,\operatorname{cl}\Omega\setminus\Omega^{\prime}, then |xy|<ρ(x)|x-y|<\rho(x).

Proof.

Consider the set

V={zclΩ|z=x+y2,x,yΩ,|xy|ρ(x)}.{V=\Big{\{}{z\in\operatorname{cl}\Omega}\;\Big{|}\,{z=\frac{x+y}{2},\quad x,y\in\partial\Omega,\ |x-y|\geqslant\rho(x)}\Big{\}}.} (8.41)

This set is closed and lies inside Ω\Omega. If two points x,yΩx,y\in\partial\Omega see each other in clΩV\operatorname{cl}\Omega\setminus V, then |xy|<ρ(x)|x-y|<\rho(x). It remains to apply Theorem 8.18 to replace VV with a larger set Ω\Omega^{\prime}. ∎

The following corollary is obtained by combination of Corollary 8.16 and Proposition 8.19.

Corollary 8.20.

Let Ω\Omega be a lens that satisfies (LABEL:StrictConvexity) and (2). Assume that the boundaries of Ω\Omega are C2C^{2}-smooth. Let f:fixedΩf\colon\partial_{\mathrm{fixed}}\Omega\to\mathbb{R} be a C2C^{2}-smooth function such that 𝔅Ω,f\mathfrak{B}_{\Omega,f} is finite. For any ε>0\varepsilon>0 there exists an extension Ω\Omega^{\prime} whose boundaries are C2C^{2}-smooth and a function G:ΩG\colon\Omega\to\mathbb{R} that is locally concave on Ω\Omega and satisfies the inequalities

𝔅Ω,f(x)G(x)𝔅Ω,f(x)+ε,xΩ;{\ \mathfrak{B}_{\Omega,f}(x)\leqslant G(x)\leqslant\mathfrak{B}_{\Omega,f}(x)+\varepsilon,\qquad x\in\Omega;} (8.42)

moreover, for any xfreeΩx\in\partial_{\mathrm{free}}\Omega there exists a linear function L[G,x]L[G,x] such that for any compact set KfreeΩK\subset\partial_{\mathrm{free}}\Omega there exists cK>0c_{K}>0 such that (8.36) holds true whenever x,yKx,y\in K see each other in Ω\Omega^{\prime}. The coefficients of the linear functions L[G,x]L[G,x] are uniformly bounded when xx runs through a compact subset of freeΩ\partial_{\mathrm{free}}\Omega.

Now we are ready to define our extension G~\tilde{G} by a formula similar to (8.39):

G~(z)={G(z),zΩ;inf{G(y)+L[G,y](zy)|yViszΩfreeΩ},zΩΩ.{\tilde{G}(z)=\begin{cases}G(z),\quad&z\in\Omega;\\ \inf\{{G(y)+L[G,y](z-y)}\;|\,{y\in\operatorname{Vis}_{z}^{\Omega^{\prime}}\cap\partial_{\mathrm{free}}\Omega}\},\quad&z\in\Omega^{\prime}\setminus\Omega.\end{cases}} (8.43)
Lemma 8.21.

For any xfreeΩx\in\partial_{\mathrm{free}}\Omega there exists a relatively open set UxΩintΩU_{x}\subset\Omega^{\prime}\setminus\operatorname{int}\Omega that contains xx and such that for any zUxz\in U_{x} the value G~(z)\tilde{G}(z) is finite and the superdifferential of G~\tilde{G} at zz is non-empty.

Refer to caption
Figure 9: Illustration to the proof of Lemma 8.21.
Proof.

Let HH be a hyperplane that separates xx from freeΩ\partial_{\mathrm{free}}\Omega^{\prime} (say, HH is closer to xx than to the latter set). It suffices to prove that if zΩintΩz\in\Omega^{\prime}\setminus\operatorname{int}\Omega is sufficiently close to xx, then the infimum in (8.43) is attained at yy that lies on the same side of HH as xx. See Fig. 9 for visualization. Then yy sees a neighborhood of zz and L[G,y]L[G,y] belongs to the superdifferential of G~\tilde{G} at zz.

Let pfreeΩp\in\partial_{\mathrm{free}}\Omega be a point that lies on the other side of HH than xx and that sees zz in Ω\Omega^{\prime}. We wish to prove that there exists yy on the same side of HH as xx such that

G(y)+L[G,y](zy)G(p)+L[G,p](zp),{G(y)+L[G,y](z-y)\leqslant G(p)+L[G,p](z-p),} (8.44)

provided zz is sufficiently close to xx. Let yy be the intersection of the line passing through zz and pp with freeΩ\partial_{\mathrm{free}}\Omega lying on the same side of HH as xx. Let also KfreeΩK\subset\partial_{\mathrm{free}}\Omega be a compact set that contains xx with all the points it can see in Ω\Omega^{\prime} and all the points the latter points can see. Let MM be the supremum of the Lipschitz constants of the functions L[G,q]L[G,q] when qKq\in K. Then,

G(y)+L[G,y](zy)G(y)+M|zy|(8.36)G(p)+L[G,p](yp)cK|py|2+M|zy|G(p)+L[G,p](zp)cK|py|2+2M|zy|.{G(y)+L[G,y](z-y)\leqslant G(y)+M|z-y|\stackrel{{\scriptstyle\scriptscriptstyle{\eqref{Quadratische}}}}{{\leqslant}}G(p)+L[G,p](y-p)-c_{K}|p-y|^{2}+M|z-y|\leqslant\\ G(p)+L[G,p](z-p)-c_{K}|p-y|^{2}+2M|z-y|.} (8.47)

We see that (8.44) indeed holds true provided zz is sufficiently close to xx, because in this case yy is also sufficiently close to xx while |py||p-y| is bounded away from zero. ∎

Proof of Theorem 8.2..

We construct the set VΩ1V\subset\Omega_{1} as the complement to the union of the sets UxU_{x}xfreeΩx\in\partial_{\mathrm{free}}\Omega, provided by Lemma 8.21. We apply Theorem 8.18 and obtain a strictly convex set Ω~1\tilde{\Omega}_{1} such that VΩ~1clΩ~1Ω1V\subset\tilde{\Omega}_{1}\subset\operatorname{cl}\tilde{\Omega}_{1}\subset\Omega_{1}. Set Ω~=clΩ0Ω~1\tilde{\Omega}=\operatorname{cl}\Omega_{0}\setminus\tilde{\Omega}_{1}. Then, G~\tilde{G} constructed by (8.43) is locally concave on Ω~\tilde{\Omega} (it is locally concave by Fact 8.9). Therefore, 𝔅Ω~,f(z)G~(z)\mathfrak{B}_{\tilde{\Omega},f}(z)\leqslant\tilde{G}(z) for any zΩ~z\in\tilde{\Omega} and

𝔅Ω,f(x)𝔅Ω~,f(x)G(x)𝔅Ω,f(x)+ε,xΩ.{\mathfrak{B}_{\Omega,f}(x)\leqslant\mathfrak{B}_{\tilde{\Omega},f}(x)\leqslant G(x)\leqslant\mathfrak{B}_{\Omega,f}(x)+\varepsilon,\qquad x\in\Omega.} (8.48)

8.3 Proof of Theorem 8.1

In order to prove Theorem 8.1, we will need to extend a locally convex function via formula (8.39) over the fixed boundary.

Lemma 8.22.

Let Ω\Omega be a lens that satisfies the requirements (LABEL:StrictConvexity) and (2). Let ωΩ\omega\supset\Omega be a set whose interior contains ΩfreeΩ\Omega\setminus\partial_{\mathrm{free}}\Omega. Let G:ωG\colon\omega\to\mathbb{R} be a locally Lipschitz function that is locally concave on Ω\Omega and has non-empty superdifferential at each point. Assume Ω\Omega^{\prime} is an open convex set that contains Ω\Omega and such that each xΩx\in\Omega^{\prime} sees only a compact subset of fixedΩ\partial_{\mathrm{fixed}}\Omega in clΩΩ0\operatorname{cl}\Omega^{\prime}\setminus\Omega_{0}. Then, the function

G~(x)={G(x),xΩ;inf{G(y)+L(xy)|yΩVisxΩΩ1,LðG|y},xΩΩ0,{\tilde{G}(x)=\begin{cases}G(x),\quad&x\in\Omega;\\ \inf\Big{\{}{G(y)+L(x-y)}\;\Big{|}\,{y\in\Omega\cap\operatorname{Vis}_{x}^{\Omega^{\prime}\setminus\Omega_{1}},L\in\eth G|_{y}}\Big{\}},\quad&x\in\Omega^{\prime}\setminus\Omega_{0},\end{cases}} (8.49)

is finite and locally concave on ΩΩ1\Omega^{\prime}\setminus\Omega_{1}.

Proof.

By local concavity of GG, we may consider only yfixedΩVisxΩΩ0y\in\partial_{\mathrm{fixed}}\Omega\cap\operatorname{Vis}_{x}^{\Omega^{\prime}\setminus\Omega_{0}} when calculating the infimum in (8.49). Since the sets fixedΩVisxΩΩ0\partial_{\mathrm{fixed}}\Omega\cap\operatorname{Vis}_{x}^{\Omega^{\prime}\setminus\Omega_{0}} are compact and the function GG is locally Lipschitz on ω\omegaG~\tilde{G} does not attain the value -\infty. Thus, it remains to verify the local concavity of G~\tilde{G}. For that, we will prove G~\tilde{G} has a non-empty superdifferential at each point xΩΩ0x\in\Omega^{\prime}\setminus\Omega_{0}.

Let xΩΩ0x\in\Omega^{\prime}\setminus\Omega_{0} and assume G~(x)=G(y)+L(xy)\tilde{G}(x)=G(y)+L(x-y)yfixedΩVisxΩΩ0y\in\partial_{\mathrm{fixed}}\Omega\cap\operatorname{Vis}_{x}^{\Omega^{\prime}\setminus\Omega_{0}} and LðG|yL\in\eth G|_{y}. It suffices to prove G~(z)G(y)+L(zy)\tilde{G}(z)\leqslant G(y)+L(z-y) when zz lies in a sufficiently small neighborhood of xx. This is true since zz sees yy in ΩΩ1\Omega^{\prime}\setminus\Omega_{1} (now we are using the initial formula (8.49)). The reasoning in the case when the infimum that defines G~(x)\tilde{G}(x) is attained at a sequence of yny_{n} does not differ. ∎

Fact 8.23.

Let Ω0\Omega_{0} be a strictly convex open set, let xx be a point of its boundary. There exists a neighborhood UxU_{x} in dΩ0\mathbb{R}^{d}\setminus\Omega_{0} such that any point yUxy\in U_{x} can see only a compact subset of Ω0\partial\Omega_{0} in dΩ0\mathbb{R}^{d}\setminus\Omega_{0}.

Proof of Theorem 8.1.

Fix pΩfixedΩp\in\Omega\setminus\partial_{\mathrm{fixed}}\Omega and ε>0\varepsilon>0.

First, we construct an open strictly convex set Ω1\Omega_{-1} with real analytic boundary such that it contains the closure of Ω0\Omega_{0} and any point xclΩ1Ω0x\in\operatorname{cl}\Omega_{-1}\setminus\Omega_{0} can see only a compact subset of fixedΩ\partial_{\mathrm{fixed}}\Omega in Ω1Ω0\Omega_{-1}\setminus\Omega_{0}; this is done by a combination of Theorem 8.17 and Fact 8.23.

Second, we construct a strictly convex set Ω\Omega^{\prime} such that clΩΩ0\operatorname{cl}\Omega^{\prime}\subset\Omega_{0}clΩ1Ω\operatorname{cl}\Omega_{1}\subset\Omega^{\prime}pΩp\in\Omega^{\prime}, and if yy and zz from fixedΩ\partial_{\mathrm{fixed}}\Omega see each other in clΩ0Ω\operatorname{cl}\Omega_{0}\setminus\Omega^{\prime}, then |yz|1|y-z|\leqslant 1; this is done by application of Proposition 8.19 with ρδ\rho\equiv\delta, where δ\delta is sufficiently small.

Note that any point xclΩ1Ωx\in\operatorname{cl}\Omega_{-1}\setminus\Omega^{\prime} can see only a compact subset of Ω\partial\Omega^{\prime} in clΩ1Ω\operatorname{cl}\Omega_{-1}\setminus\Omega^{\prime}. We also note that the restriction of the function 𝔅Ω,f\mathfrak{B}_{\Omega,f} to the set Ω\Omega^{\prime} is locally Lipschitz and has a non-empty superdifferential at each point; the set {L|Lð𝔅Ω,f|x,xK}\{{\nabla L}\;|\,{L\in\eth\mathfrak{B}_{\Omega,f}|_{x},x\in K}\} is uniformly bounded for any compact set KΩK\subset\Omega^{\prime}. Thus, we may construct the function G:clΩ1Ω1G\colon\operatorname{cl}\Omega_{-1}\setminus\Omega_{1}\to\mathbb{R} by the formula

G(x)={𝔅Ω,f(x),xclΩΩ1;inf{𝔅Ω,f(y)+L(xy)|yΩVisxclΩ1Ω1,Lð𝔅Ω,f|y},xclΩ1Ω.{G(x)=\begin{cases}\mathfrak{B}_{\Omega,f}(x),\quad&x\in\operatorname{cl}\Omega^{\prime}\setminus\Omega_{1};\\ \inf\Big{\{}{\mathfrak{B}_{\Omega,f}(y)+L(x-y)}\;\Big{|}\,{y\in\Omega^{\prime}\cap\operatorname{Vis}_{x}^{\operatorname{cl}\Omega_{-1}\setminus\Omega_{1}},L\in\eth\mathfrak{B}_{\Omega,f}|_{y}}\Big{\}},\quad&x\in\operatorname{cl}\Omega_{-1}\setminus\Omega^{\prime}.\end{cases}} (8.50)

By Lemma 8.22GG is a locally concave function. What is more, GG is continuous on Ω1\partial\Omega_{-1}. Therefore, the function

𝔅clΩ1Ω1,G|Ω1{\mathfrak{B}_{\operatorname{cl}\Omega_{-1}\setminus\Omega_{1},G|_{\partial\Omega_{-1}}}} (8.51)

may be ’extended’ through the free boundary by Theorem 8.2. Let Ω~\tilde{\Omega} be the extension of clΩ1Ω1\operatorname{cl}\Omega_{-1}\setminus\Omega_{1}, let G~\tilde{G} be the ’extended’ function. Then,

G~(p)𝔅clΩ1Ω1,G|Ω1(p)+εG(p)+ε=𝔅Ω,f(p)+ε.{\tilde{G}(p)\leqslant\mathfrak{B}_{\operatorname{cl}\Omega_{-1}\setminus\Omega_{1},G|_{\partial\Omega_{-1}}}(p)+\varepsilon\leqslant G(p)+\varepsilon=\mathfrak{B}_{\Omega,f}(p)+\varepsilon.} (8.52)

9 Limitations of current methods

The lemma below justifies the appearance of the auxiliary domain Ω^1\hat{\Omega}_{1} in (5.2). It also shows why we are not able to prove (4.2) for the points xfreeΩx\in\partial_{\mathrm{free}}\Omega in Theorem 4.4. We recall the definition of a cheese domain from [21]. A set Ωd\Omega\subset\mathbb{R}^{d} is called a cheese domain if it may be represented as

Ω=clΩ0j=1NΩj,\Omega=\operatorname{cl}\Omega_{0}\setminus\bigcup_{j=1}^{N}\Omega_{j}, (9.1)

where the domains Ωj\Omega_{j}j=0,1,,N,j=0,1,\ldots,N, are strictly convex, open, and bounded; the ‘holes’ Ωj\Omega_{j}j=1,2,,Nj=1,2,\ldots,N, are mutually separated and also lie inside the interior of Ω0\Omega_{0}.

Lemma 9.1.

Let Ω2\Omega\subset\mathbb{R}^{2} be a cheese domain such that the domains Ωj,\Omega_{j}, j=1,2,,N,j=1,2,\ldots,N, in the representation (9.1) have C1C^{1}-smooth boundaries. Let φ:𝕋Ω0\varphi\colon\mathbb{T}\to\partial\Omega_{0} be such that for any arc J𝕋J\subset\mathbb{T} the point φJ\langle{\varphi}\rangle_{{}_{J}} lies in Ω\Omega. If φ𝕋Ωj\langle{\varphi}\rangle_{{}_{\mathbb{T}}}\in\partial\Omega_{j} for some j,j, then φ\varphi attains two values.

Proof.

Without loss of generality, φ𝕋Ω1\langle{\varphi}\rangle_{{}_{\mathbb{T}}}\in\partial\Omega_{1}. Let \ell be the tangent to Ω1\partial\Omega_{1} at the point x=φ𝕋x=\langle{\varphi}\rangle_{{}_{\mathbb{T}}}. We say that a point lies below \ell if it is strictly separated by \ell from Ω1\Omega_{1} and say that it lies above \ell in the case the point and Ω1\Omega_{1} belong to the same open half-plane generated by \ell.

We will first prove that φ\varphi does not attain values on the part of Ω0\partial\Omega_{0} lying below \ell. Assume the contrary. Let tt be a Lebesgue point of φ\varphi such that φ(t)\varphi(t) lies below \ell. Consider the point

φ𝕋[tε,t+ε]=112ε(x2εφ[tε,t+ε]).{\langle{\varphi}\rangle_{{}_{\mathbb{T}\setminus[t-\varepsilon,t+\varepsilon]}}=\frac{1}{1-2\varepsilon}\Big{(}x-2\varepsilon\langle{\varphi}\rangle_{{}_{[t-\varepsilon,t+\varepsilon]}}\Big{)}.} (9.2)

When ε\varepsilon is sufficiently close to 0, the point φ[tε,t+ε]\langle{\varphi}\rangle_{{}_{[t-\varepsilon,t+\varepsilon]}} lies close to φ(t)\varphi(t). Thus, the point (9.2) lies in a small neighborhood of xx and above \ell. In particular, it does not belong to Ω\Omega provided ε\varepsilon is sufficiently small. This is a contradiction.

Second, we will prove that φ\varphi does not attain values on the part of Ω0\partial\Omega_{0} lying above \ell. Consider an affine function LL that is equal to zero on \ell and is negative below \ell. Then, L(x)=0L(x)=0. On the other hand,

L(x)=𝕋L(φ(t))𝑑t>0L(x)=\int\limits_{\mathbb{T}}L\big{(}\varphi(t)\big{)}\,dt>0

provided L(φ)>0L(\varphi)>0 on a set of non-zero measure (since we have proved that L(φ)L(\varphi) is always non-negative). ∎

Finally we will show that the convexity of Ω1\Omega_{1} is necessary for (4.2) (note that the definition of the minimal locally concave function 𝔅Ω,f\mathfrak{B}_{\Omega,f} and the Bellman function 𝑩Ω,f\boldsymbol{B}_{\Omega,f} do not require convexity of Ω1\Omega_{1}).

Refer to caption
Figure 10: The domain Ω\Omega and the set of Bellman points of φ\varphi.

We will shortly construct an example of Ω\Omega and ff such that 𝑩Ω,f>𝔅Ω,f\boldsymbol{B}_{\Omega,f}>\mathfrak{B}_{\Omega,f}. Let Ω0\Omega_{0} be the unit circle. Consider three points

A=(1,0),B=(1,0),C=(0,1).A=(-1,0),\quad B=(1,0),\quad C=(0,-1).

Define the function φ:[0,1]Ω0\varphi\colon[0,1]\to\partial\Omega_{0} by the rule:

φ(t)={A,[t][0,13);C,[t][13,23);B,[t][23,1].\varphi(t)=\begin{cases}A,\quad&[t]\in[0,\frac{1}{3});\\ C,\quad&[t]\in[\frac{1}{3},\frac{2}{3});\\ B,\quad&[t]\in[\frac{2}{3},1].\end{cases}
Fact 9.2.

The set of Bellman points of φ\varphi, i. e., the points φJ\langle{\varphi}\rangle_{{}_{\operatorname{J}}}, is

ACBC{x=αA+βB+γCα+β+γ=1,γα0,γβ0}.AC\cup BC\cup\{x=\alpha A+\beta B+\gamma C\mid\alpha+\beta+\gamma=1,\gamma\geqslant\alpha\geqslant 0,\gamma\geqslant\beta\geqslant 0\}. (9.3)

Now we construct the domain Ω1\Omega_{1} and the function ff. Define Ω1\Omega_{1} by the rule

Ω1={x2|x(12,0)12.9}{x2|x(12,0)12.9}.\Omega_{1}=\Big{\{}x\in\mathbb{R}^{2}\,\Big{|}\;\Big{\|}x-\Big{(}-\frac{1}{2},0\Big{)}\Big{\|}\leqslant\frac{1}{2.9}\Big{\}}\cup\Big{\{}x\in\mathbb{R}^{2}\,\Big{|}\;\Big{\|}x-\Big{(}\frac{1}{2},0\Big{)}\Big{\|}\leqslant\frac{1}{2.9}\Big{\}}.

We draw Figure 10 for reader’s convenience (the picture has slightly different numeric parameters, what is important is that the two ‘erased’ circles do not intersect the set (9.3), however, the average of φ\varphi lies above the lower common tangent to the circles).

By Fact 9.2φ𝑨(Ω)\varphi\in\boldsymbol{A}(\Omega). Define f:fixedΩf\colon\partial_{\mathrm{fixed}}\Omega\to\mathbb{R} by the formula

f(y)={0,y20;y2,y2>0.f(y)=\begin{cases}0,\quad&y_{2}\leqslant 0;\\ -y_{2},\quad&y_{2}>0.\end{cases} (9.4)

Clearly, f(φ)[0,1]=0\langle{f(\varphi)}\rangle_{{}_{[0,1]}}=0, and, thus, 𝑩f(0,13)0\boldsymbol{B}_{f}(0,-\frac{1}{3})\geqslant 0.

Lemma 9.3.

Consider ff given by (9.4). Then, 𝔅Ω,f(0,13)<0\mathfrak{B}_{\Omega,f}(0,-\frac{1}{3})<0.

Proof.

It is clear that 𝔅Ω,f(y)f(y)\mathfrak{B}_{\Omega,f}(y)\leqslant f(y) provided we extend the function ff to Ω\Omega by the same formula (9.4). Consider the points EE and FF on the erased circles that have the smallest possible second coordinates. In other words, they are the points on the lower common tangent to these two circles. Similarly, let GG and HH be the points having the largest possible second coordinates.

Let LL be the linear function that coincides with ff on the segments EFEF and GHGH and let the part of Ω\Omega lying between EFEF and GHGH be called the channel. Define the new function Φ\Phi to be equal LL on the channel and to be equal ff everywhere else. It is easy to observe that Φ\Phi is locally concave. On the other hand, L<fL<f and thus, Φ(0,13)<0\Phi(0,-\frac{1}{3})<0. ∎

References

  • [1] D. Azagra, Global and fine approximation of convex functions, Proc. Lond. Math. Soc. (3) 107 (2013), no. 4, 799–824.
  • [2] D. Azagra and D. Stolyarov, Inner and outer smooth approximation of convex hypersurfaces. When is it possible?, https://arxiv.org/abs/2204.07498.
  • [3] H. Brezis and L. Nirenberg, Degree theory and BMO. I. Compact manifolds without boundaries, Selecta Math. 1 (1995), no. 2, 197–263.
  • [4] E. Dobronravov, On a minimal locally concave function on a ring, in preparation.
  • [5] B. Guan, The Dirichlet problem for Monge-Ampère equations in non-convex domains and spacelike hypersurfaces of constant Gauss curvature, Trans. Amer. Math. Soc. 350 (1998), no. 12, 4955–4971.
  • [6] T. Hytönen, J. van Neerven, M. Veraar, and L. Weis, Analysis in Banach spaces. Volume I: Martingales and Littlewood–Paley theory, Springer, 2016.
  • [7] P. Ivanishvili, N. N. Osipov, D. M. Stolyarov, V. I. Vasyunin, and P. B. Zatitskiy, Sharp estimates of integral functionals on classes of functions with small mean oscillation, C. R. Math. Acad. Sci. Paris 350 (2012), no. 11, 561–564.
  • [8]  , On Bellman function for extremal problems in BMO\mathrm{BMO}, C. R. Math. Acad. Sci. Paris 353 (2015), no. 12, 1081–1085.
  • [9] P. Ivanishvili, N. N. Osipov, D. M. Stolyarov, V. I. Vasyunin, and P. B. Zatitskiy, Bellman function for extremal problems in BMO\mathrm{BMO}, Trans. Amer. Math. Soc. 368 (2016), 3415–3468.
  • [10] P. Ivanishvili, D. M. Stolyarov, V. I. Vasyunin, and P. B. Zatitskiy, Bellman function for extremal problems on BMO\mathrm{BMO} II: evolution, Mem. Amer. Math. Soc. 255 (2018), no. 1220.
  • [11] F. John and L. Nirenberg, On functions of bounded mean oscillation, Comm. Pure Appl. Math. 14 (1961), 415–426.
  • [12] B. Kirchheim, S. Müller, and V. Šverák, Studying nonlinear pde by geometry in matrix space, Geometric analysis and nonlinear partial differential equations, Springer, Berlin, 2003, pp. 347–395.
  • [13] A. Reznikov, Sharp weak type estimates for weights in the class Ap1,p2A_{p_{1},p_{2}}, Rev. Mat. Iberoam. 29 (2013), no. 2, 433–478.
  • [14] R. T. Rockafellar, Convex analysis, Princeton Mathematical Series, No. 28, Princeton University Press, Princeton, N.J., 1970.
  • [15] A. N. Shiryaev, Probability. 2, Graduate Texts in Mathematics, vol. 95, Springer, New York, 2019, Third edition of [ MR0737192], Translated from the 2007 fourth Russian edition by R. P. Boas and D. M. Chibisov.
  • [16] L. Slavin and V. Vasyunin, Sharp results in the integral form John–Nirenberg inequality, Trans. Amer. Math. Soc. 363 (2011), no. 8, 4135–4169.
  • [17]  , Sharp LpL^{p} estimates on BMO\mathrm{BMO}, Indiana Univ. Math. J. 61 (2012), no. 3, 1051–1110.
  • [18] E. M. Stein, Harmonic analysis: real-variable methods, orthogonality, and oscillatory integrals, Princeton University Press, 1993.
  • [19] D. Stolyarov, V. Vasyunin, and P. Zatitskiy, Sharp multiplicative inequalities with BMO I, J. Math. Anal. Appl. 492 (2020), no. 2, 124479, 17.
  • [20] D. Stolyarov and P. Zatitskiy, Sharp transference principle for BMO and ApA_{p}, J. Funct. Anal. 281, no. 6, 109085.
  • [21] D. M. Stolyarov and P. B. Zatitskiy, Theory of locally concave functions and its applications to sharp estimates of integral functionals, Adv. Math. 291 (2016), 228–273.
  • [22] V. Vasyunin and A. Volberg, Sharp constants in the classical weak form of the John–Nirenberg inequality, Proc. Lond. Math. Soc. 108 (2014), no. 6, 1417–1434.
  • [23] V. Vasyunin, P. Zatitskiy, and I. Zlotnikov, Sharp multiplicative inequalities with BMO II, https://arxiv.org/abs/2111.05565.
  • [24] V. I. Vasyunin, The sharp constant in the John–Nirenberg inequality, PDMI preprints 20/2003, http://www.pdmi.ras.ru/preprint/2003/03-20.html.
  • [25]  , The exact constant in the inverse Hölder inequality for Muckenhoupt weights, Algebra i Analiz 15 (2003), no. 1, 73–117.
  • [26]  , Mutual estimates for LpL^{p}-norms and the Bellman function, Zap. Nauchn. Sem. S.-Peterburg. Otdel. Mat. Inst. Steklov. (POMI) 355 (2008), no. Issledovaniya po Lineĭnym Operatoram i Teorii Funktsiĭ. 36, 81–138, 237–238.

Dmitriy Stolyarov, [email protected],

St. Petersburg State University, Department of Mathematics and Computer Science;

Pavel Zatitskiy, [email protected],

St. Petersburg State University, Department of Mathematics and Computer Science.