Carleson operators on doubling metric measure spaces

10 Two-sided Metric Space Carleson

We prove a variant of Theorem 1.0.2 for a two-sided Calderón–Zygmund kernel on the doubling metric measure space \((X,\rho ,\mu ,a)\), i.e. a one-sided Calderón–Zygmund kernel \(K\) which additionally satisfies for all \(x,x',y\in X\) with \(x\neq y\) and \(2\rho (x,x') \leq \rho (x,y)\),

\begin{equation} \label{eqkernel-x-smooth} |K(x,y) - K(x',y)| \leq \left(\frac{\rho (x,x')}{\rho (x,y)}\right)^{\frac{1}{a}}\frac{2^{a^3}}{V(x,y)}. \end{equation}
10.0.1

By the additional regularity, we can weaken the assumption 1.0.18 to a family of operators that is easier to work with in applications. Namely, for \(r {\gt} 0\), \(x\in X\), and a bounded, measurable function \(f:X\to {\mathbb {C}}\) supported on a set of finite measure, we define

\begin{equation} \label{def-T-r} T_r f(x):= \int _{r\le \rho (x,y)} K(x,y) f(y) \, d\mu (y) = \int _{X\setminus B(x,r)} K(x,y) f(y) \, d\mu (y). \end{equation}
10.0.2

Theorem 10.0.1 two-sided metric space Carleson
#

For all integers \(a \ge 4\) and real numbers \(1{\lt}q\le 2\) the following holds. Let \((X,\rho ,\mu ,a)\) be a doubling metric measure space. Let \({\Theta }\) be a cancellative compatible collection of functions and let \(K\) be a two-sided Calderón–Zygmund kernel on \((X,\rho ,\mu ,a)\). Assume that for every bounded measurable function \(g\) on \(X\) supported on a set of finite measure and all \(r{\gt}0\) we have

\begin{equation} \label{two-sided-Hr-bound-assumption} \| T_r g\| _{2} \leq 2^{a^3} \| g\| _2\, . \end{equation}
10.0.3

Then for all Borel sets \(F\) and \(G\) in \(X\) and all Borel functions \(f:X\to {\mathbb {C}}\) with \(|f|\le \mathbf{1}_F\), we have, with \(T\) defined in 1.0.17,

\begin{equation} \label{two-sided-resweak} \left|\int _{G} T f \, \mathrm{d}\mu \right| \leq \frac{2^{452a^3}}{(q-1)^6} \mu (G)^{1-\frac{1}{q}} \mu (F)^{\frac{1}{q}}\, . \end{equation}
10.0.4

For the remainder of this chapter, fix an integer \(a\ge 4\), a doubling metric measure space \((X,\rho ,\mu ,a)\) and a two-sided Calderón–Zygmund kernel \(K\) as in Theorem 10.0.1.

The following lemma is proved in Section 10.1.

Lemma 10.0.2 nontangential-from-simple

Assume 10.0.3 holds. Then, for every bounded measurable function \(g : X \to {\mathbb {C}}\) supported on a set of finite measure we have

\begin{equation} \label{concretetstarbound} \| T_*g\| _2\le 2^{3a^3}\| g\| _2. \end{equation}
10.0.5

Proof of Theorem 10.0.1

Let \(1{\lt}q\le 2\) be a real number. Let \(\Theta \) be a cancellative compatible collection of functions. By the assumption 10.0.3, we can apply Lemma 10.0.2 to obtain for every bounded measurable \(g:X\to {\mathbb {C}}\) supported on a set of finite measure,

\begin{equation} \label{original-operator-assumption} \| T_*g\| _2\le 2^{3a^3}\| g\| _2. \end{equation}
10.0.6

Define

\begin{equation*} K’(x,y):= 2^{-2a^3} K(x,y) \, . \end{equation*}

Then \(K'\) is a two-sided Calderón–Zygmund kernel on \((X,\rho ,\mu ,a)\). Denote the corresponding maximally truncated non-tangential singular operator by \(T_*'\) and the corresponding generalized Carleson operator by \(T'\). With 10.0.6, we obtain for \(g\) as above,

\begin{equation} \label{modified-operator-assumption} \| T_*'g\| _2\le 2^{a^3}\| g\| _2. \end{equation}
10.0.7

Applying Theorem 1.0.2 for \(K'\) yields that for all Borel sets \(F\) and \(G\) in \(X\) and all Borel functions \(f:X\to {\mathbb {C}}\) with \(|f|\le \mathbf{1}_F\), we have

\begin{equation*} \left|\int _{G} T’ f \, \mathrm{d}\mu \right| \leq \frac{2^{450a^3}}{(q-1)^6} \mu (G)^{1-\frac{1}{q}} \mu (F)^{\frac{1}{q}}\, . \end{equation*}

This finishes the proof since for all \(x\in X\),

\begin{equation*} T’f(x) = 2^{-2a^3} Tf(x) \, . \end{equation*}

The proof of Lemma 10.0.2 relies on the following auxiliary lemma which is proved in Section 10.2.

Lemma 10.0.3 Weak 1 1

Let \(f:X\to {\mathbb {C}}\) be a bounded measurable function supported on a set of finite measure and assume for some \(r{\gt}0\) that for every bounded measurable function \(g:X\to {\mathbb {C}}\) supported on a set of finite measure,

\begin{equation} \label{eq-strong-2-2-assumption} \| T_rg\| _{2}\le 2^{a^3} \| g\| _2. \end{equation}
10.0.8

Then for all \(\alpha {\gt}0\), we have

\begin{equation} \label{eq-weak-1-1} \mu \left(\{ x\in X: |T_r f(x)|{\gt}\alpha \} \right)\le \frac{2^{a^3 + 19a}}{\alpha } \int |f(y)|\, d\mu (y). \end{equation}
10.0.9

Throughout ??, for any measurable bounded function \(w: X \to {\mathbb {C}}\), let \(Mw: X \to [0, \infty )\) denote the corresponding Hardy–Littlewood maximal function defined in Proposition 2.0.6. Apart from Proposition 2.0.6, ?? have no dependencies in the previous chapters.

10.1 Proof of Cotlar’s Inequality

Lemma 10.1.1

For all real numbers \(x\ge 4\),

\begin{equation*} \sum _{n=0}^\infty 2^{-\frac{n}{x}} \le 2^x. \end{equation*}
Proof

By convexity, for all \(0\le \lambda \le 1\)

\begin{equation*} 2^{\lambda (-\frac{1}{4})} \le \lambda 2^{-\frac{1}{4}} + (1-\lambda )2^0. \end{equation*}

For \(\lambda :=\frac{4}{x}\), we obtain

\begin{equation*} 2^{-\frac1x} \le 1 - (1-2^{-\frac14}) \frac{4}{x}. \end{equation*}

We conclude

\begin{equation*} \sum _{n=0}^\infty 2^{-\frac{n}{x}} = \frac{1}{1-2^{-\frac1x}} \le \frac{1}{4(1-2^{-\frac14})} x \le 2^x. \end{equation*}
Lemma 10.1.2 estimate x shift

Let \(0{\lt}r\) and \(x\in X\). Let \(g:X\to {\mathbb {C}}\) be a bounded measurable function supported on a set of finite measure. Then for all \(x'\) with \(\rho (x,x')\le r\).

\begin{equation*} \left| T_r g(x) - T_r g(x’) \right| \le 2^{a^3 + 2a + 2} Mg(x)\, . \end{equation*}
Proof

By definition,

\begin{equation} \label{xx'difference} \left| T_r g(x) - T_r g(x') \right| =\left|\int _{r\le \rho (x,y)} K(x,y) g(y) \, d\mu (y) -\int _{r\le \rho (x',y)} K(x',y) g(y)\, d\mu (y) \right| \, . \end{equation}
10.1.1

We split the first integral in 10.1.1 into the domains \(r\le \rho (x,y){\lt}2r\) and \(2r\le \rho (x,y)\). The integral over the first domain we estimate by \(\eqref{firstxx'}\) below. For the second domain, we observe with \(\rho (x,x')\le r\) and the triangle inequality that \(r\le \rho (x',y)\). We therefore combine on this domain with the corresponding part of the second integral in 10.1.1 and estimate that by \(\eqref{secondxx'}\) below. The remaining part of the second integral in 10.1.1 we estimate by \(\eqref{thirdxx'}\). Overall, we have estimated 10.1.1 by

\begin{equation} \label{firstxx'} \int _{r\le \rho (x,y){\lt} 2r} |K(x,y)| |g(y)| \, d\mu (y) \end{equation}
10.1.2

\begin{equation} \label{secondxx'} + \left|\int _{2r\le \rho (x,y)} (K(x,y)-K(x',y)) g(y) \, d\mu (y) \right| \end{equation}
10.1.3

\begin{equation} \label{thirdxx'} +\int _{r\le \rho (x',y), r\le \rho (x,y){\lt}2r} |K(x',y)| |g(y)|\, d\mu (y)\, . \end{equation}
10.1.4

Using the bound on \(K\) in 1.0.14 and the doubling condition 1.0.5, we estimate 10.1.2 by

\begin{align} \int _{r\le \rho (x,y){\lt}2r} \frac{2^{a^3}}{V(x,y)} |g(y)| \, d\mu (y)\, & \le \frac{2^{a^3}}{\mu (B(x,r))} \int _{r\le \rho (x,y){\lt}2r} |g(y)| \, d\mu (y)\, \\ \label{firstxx'b} & \le \frac{2^{a^3} \cdot 2^a}{\mu (B(x,2r))} \int _{\rho (x,y){\lt}2r} |g(y)| \, d\mu (y)\, . \end{align}

Using the definition of \(Mg\), we estimate 10.1.6 by

\begin{equation} \label{firstxx'c} \le 2^{a^3 + a} {Mg(x)}\, . \end{equation}
10.1.7

Similarly, in the domain of 10.1.4 we note by the triangle inequality and assumption on \(x'\) that \(\rho (x',y){\lt}3r\) and thus we estimate 10.1.4 by

\begin{equation} \label{thirdxx'b} \frac{2^{a^3}}{\mu (B(x',r))} \int _{\rho (x',y){\lt}4r} |g(y)| \, d\mu (y)\le 2^{a^3 + 2a} Mg(x) \end{equation}
10.1.8

We turn to the remaining term. Using 10.0.1, we estimate 10.1.3 by

\begin{equation} \label{secondxx'b} \int _{2r\le \rho (x,y)} \left(\frac{\rho (x,x')}{\rho (x,y)}\right)^{\frac{1}{a}}\frac{2^{a^3}}{V(x,y)} |g(y)| \, d\mu (y) \end{equation}
10.1.9

We decompose and estimate 10.1.3 with the triangle inequality by

\begin{align} & \sum _{j=1}^\infty \int _{2^jr\le \rho (x,y){\lt} 2^{j+1}r} \left(\frac{\rho (x,x')}{\rho (x,y)}\right)^{\frac{1}{a}}\frac{2^{a^3}}{V(x,y)} |g(y)| \, d\mu (y)\, \\ \le & \sum _{j=1}^\infty \left( 2^{-j} \right)^{\frac{1}{a}} \int _{2^jr\le \rho (x,y){\lt} 2^{j+1}r} \frac{2^{a^3}}{\mu (B(x,2^j r))} |g(y)| \, d\mu (y) \\ \le & \sum _{j=1}^\infty 2^{-\frac{j}{a}} \frac{2^{a^3 + a}}{\mu (B(x,2^{j+1} r))} \int _{\rho (x,y){\lt}2^{j+1}r} |g(y)| \, d\mu (y) \\ \label{secondxx'c} \le & 2^{a^3 + a} \sum _{j=1}^\infty 2^{-\frac{j}{a}} Mg(x)\, . \end{align}

Using Lemma 10.1.1, we estimate 10.1.13 by

\begin{equation} \label{secondxx'd} \le 2^{a^3 + 2a} Mg(x) \, . \end{equation}
10.1.14

Summing the estimates for 10.1.2, 10.1.3, and 10.1.4 proves the lemma.

Lemma 10.1.3 Cotlar control

Let \(0{\lt}r\le R\) and \(x\in X\). Let \(g:X\to {\mathbb {C}}\) be a bounded measurable function supported on a set of finite measure. Then for all \(x'\in X\) with \(\rho (x,x')\le \frac{R}{4}\) we have

\begin{equation} \label{eq-cotlar-control} \left|T_R g(x) \right|\le |T_r(g-g\mathbf{1}_{B(x,\frac{R}2)})(x')| + 2^{a^3 + 4a + 1} Mg(x)\, . \end{equation}
10.1.15

Proof

Let \(x\) and \(x'\) be given with \(\rho (x,x')\le \frac{R}{4}\). By an application of Lemma 10.1.2, we estimate the left-hand-side of 10.1.15 by

\begin{equation} \label{eqcotlar0} |T_R(g)(x')|+ 2^{a^3 + 2a + 2} Mg(x)\, . \end{equation}
10.1.16

We have

\begin{equation} \label{eqcotlar-1} T_R(g)(x')= \int _{R\le \rho (x',y)} K(x',y) g(y) \, d\mu (y)\, . \end{equation}
10.1.17

On the domain \(R\le \rho (x',y)\), we have \(\frac{R}2\le \rho (x,y)\). Hence we may write for 10.1.17

\begin{equation*} T_R(g)(x’)=\int _{R\le \rho (x',y)} K(x’,y) (g-g\mathbf{1}_{B(x,\frac{R}2)})(y) \, d\mu (y) \end{equation*}

\begin{equation} \label{eqcotlar1} =T_R(g-g\mathbf{1}_{B(x,\frac{R}2)})(x')\, . \end{equation}
10.1.18

Combining the estimate 10.1.16 with the identification 10.1.18, we obtain

\begin{equation} \label{eqcotlar5} \left|T_R g(x) \right|\le |T_R(g-g\mathbf{1}_{B(x,\frac{R}2)})(x')|+ 2^{a^3 + 2a + 2} Mg(x)\, . \end{equation}
10.1.19

We have

\begin{equation*} (T_r-T_R)(g-g\mathbf{1}_{B(x,\frac{R}2)})(x’) \end{equation*}
\begin{equation*} = \int _{B(x',R)\setminus B(x',r)} K(x’,y) (g-g\mathbf{1}_{B(x,\frac{R}2)})(y) \, d\mu (y) \end{equation*}

\begin{equation} \label{eqcotlar2} = \int _{B(x',R)\setminus (B(x',r) \cup B(x,\frac{R}{2}))} K(x',y) g(y) \, d\mu (y) \end{equation}
10.1.20

As \(\frac{R}{2}\le \rho (x,y)\) together with \(\rho (x,x')\le \frac{R}{4}\) implies \(\frac{R}{4}\le \rho (x',y)\), we can estimate the absolute value of 10.1.20 with 1.0.14 by

\begin{align*} \le & \frac{2^{a^3}}{\mu (B(x',\frac{R}{4}))} \int _{B(x,2R)\setminus B(x',\frac{R}{4})} |g(y)|\, d\mu (y) \\ \le & \frac{2^{a^3+a}}{\mu (B(x',\frac{R}{2}))} \int _{B(x,2R)} |g(y)|\, d\mu (y) \\ \le & 2^{a^3+a} \frac{\mu (B(x,2R))}{\mu (B(x,\frac{R}{4}))} Mg(x) \le 2^{a^3 + 4a} Mg(x)\, . \end{align*}

By the triangle inequality, 10.1.15 follows now from 10.1.19 and the estimate for 10.1.20.

Lemma 10.1.4 Cotlar sets

Assume that 10.0.3 holds. Let \(0{\lt}r\le R\) and \(x\in X\). Let \(g:X\to {\mathbb {C}}\) be a bounded measurable function supported on a set of finite measure. Then the measure \(|F_1|\) of the set \(F_1\) of all \(x'\in B(x,\frac{R}4)\) such that

\begin{equation} \label{first-cotlar-exception} |T_rg(x')|{\gt} 4 M(T_rg)(x) \end{equation}
10.1.21

is less than or equal to \(\mu (B(x,\frac{R}{4}))/4\). Moreover, the measure \(|F_2|\) of the set \(F_2\) of all \(x'\in B(x,\frac{R}4)\) such that

\begin{equation} \label{second-cotlar-exception} |T_r(g\mathbf{1}_{B(x,\frac{R}2)})(x')| {\gt} 2^{a^3 + 20a + 2} Mg(x) \end{equation}
10.1.22

is less than or equal to \(\mu (B(x,\frac{R}{4}))/4\).

Proof

Let \(r\), \(R\), \(x\) and \(g\) be given. If \(M(T_rg)(x)=0\), then \(T_rg\) is zero almost everywhere and the estimate on \(|F_1|\) is trivial. Assume \(M(T_rg)(x){\gt}0\). We have with 10.1.21

\begin{equation} M(T_rg)(x)\ge \frac1{\mu (B(x,\frac{R}{4}))}\int _{B(x,\frac{R}{4})}|T_rg(x')|\, dx' \end{equation}
10.1.23

\begin{equation} \ge \frac1{\mu (B(x,\frac{R}{4}))}\int _{F_1} 4 M(T_rg)(x)\, dx' \, . \end{equation}
10.1.24

Dividing by \(M(T_rg)(x)\) gives

\begin{equation} 1\ge \frac{4}{\mu (B(x,\frac{R}{4}))} |F_1|\, . \end{equation}
10.1.25

This gives the desired bound for the measure of \(F_1\). We turn to the set \(F_2\). Similarly as above we may assume \(Mg(x){\gt}0\). The set \(F_2\) is then estimated with Lemma 10.0.3 by

\begin{equation} \frac{2^{a^3+19a}}{2^{a^3 + 20a + 2}Mg(x)}\int |g\mathbf{1}_{B(x,\frac{R}2)}|(y)\, d\mu (y) \end{equation}
10.1.26

\begin{equation} \le \frac{1}{2^{a+2} Mg(x)}\mu (B(x,\frac{R}{2})) Mg(x) \le \frac{\mu (B(x,\frac{R}{4}))}{4} \, . \end{equation}
10.1.27

This gives the desired bound for the measure of \(F_2\).

Lemma 10.1.5 Cotlar estimate

Assume that 10.0.3 holds. Let \(0{\lt}r\le R\) and \(x\in X\). Let \(g:X\to {\mathbb {C}}\) be a bounded measurable function supported on a set of finite measure. Then

\begin{equation} \label{eq-cotlar-estimate} |T_Rg(x)| \le 2^{2}M(T_rg)(x)+ 2^{a^3+20a+3} Mg(x) \, . \end{equation}
10.1.28

Proof

By Lemma 10.1.4, the set of all \(x'\in B(x,\frac{R}4)\) such that at least one of the conditions 10.1.21 and 10.1.22 is satisfied has measure less than or equal to \(\mu (B(x,\frac{R}{4}))/2\) and hence is not all of \(B(x,\frac{R}4)\). Pick an \(x'\in B(x,\frac{R}4)\) such that both conditions are not satisfied. Applying Lemma 10.1.3 for this \(x'\) and using the triangle inequality estimates the left-hand side of 10.1.28 by

\begin{equation} 4 M(T_rg)(x) + 2^{a^3 + 20a + 2} Mg(x) + 2^{a^3 + 4a + 1} Mg(x)\, . \end{equation}
10.1.29

This proves the lemma.

Lemma 10.1.6 simple nontangential operator

Assume that 10.0.3 holds. For every \(r{\gt}0\) and every bounded measurable function \(g\) supported on a set of finite measure we have

\begin{equation} \label{trzerobound} \| T_{*}^rg\| _2\le 2^{a^3+24a+6}\| g\| _2, \end{equation}
10.1.30

where

\begin{equation} \label{eq-simple--nontangential} T_{*}^r g(x):=\sup _{r{\lt}R}\sup _{x'\in B(x,R)} |T_R(g)(x')| \, . \end{equation}
10.1.31

Proof

With Lemma 10.1.2 and the triangle inequality, we estimate for every \(x\in X\)

\begin{equation} T_{*}^r g(x) \le 2^{a^3 + 2a + 2} Mg(x)+\sup _{r{\lt}R} |T_R(g)(x)|\, . \end{equation}
10.1.32

Using further Lemma 10.1.5, we estimate

\begin{equation} T_{*}^r g(x) \le 2^{a^3+2a+2}Mg(x) + 2^{a^3+20a+3}Mg(x) + 2^{2}M(T_rg)(x)\, . \end{equation}
10.1.33

Taking the \(L^2\) norm and using Proposition 2.0.6 with \(a=4\) and \(p_2=2\) and \(p_1=1\) , we obtain

\begin{equation} \| T_{*}^r g\| _2 \le 2^{a^3+20a+4} \| Mg\| _2 + 2^{2}\| M(T_rg)\| _2 \end{equation}
10.1.34

\begin{equation} \le 2^{a^3+24a+5} \| g\| _2 + 2^{4a+3}\| T_r g\| _2\, . \end{equation}
10.1.35

Applying 10.0.3 gives

\begin{equation} \| T_{*}^r g\| _2\le 2^{a^3+24a+5}\| g\| _2 + 2^{a^3+4a+3}\| g\| _2\, . \end{equation}
10.1.36

This shows 10.1.30 and completes the proof of the lemma.

In order to pass from the one-sided truncation in \(T_r\) and \(T_{*}^r\) to the two-sided truncation in \(T_*\), we show in the following two lemmas that the integral in 1.0.16 can be exchanged for an integral over the difference of two balls.

Lemma 10.1.7

Let \(f:X\to {\mathbb {C}}\) be a bounded measurable function supported on a set of finite measure. Let \(x\in X\) and \(R{\gt}0\). Then, for all \(\epsilon {\gt}0\), there exists some \(\delta {\gt}0\) such that

\begin{equation} \left| \int _{R{\lt}\rho (x,y){\lt}R+\delta } K(x,y) f(y) \, d\mu (y) \right| \le \epsilon \end{equation}
10.1.37

and

\begin{equation} \left| \int _{R-\delta {\lt}\rho (x,y){\lt}R} K(x,y) f(y) \, d\mu (y) \right| \le \epsilon \, . \end{equation}
10.1.38

Proof

We only prove the second inquality, the first one is analogous. Note that the integrand is bounded in \(X\setminus B(x,\frac{R}{2})\). So for \(0{\lt}\delta \le \frac{R}{2}\),

\begin{align*} & \left| \int _{R-\delta {\lt}\rho (x,y){\lt}R} K(x,y) f(y) \, d\mu (y) \right| \\ \le & \frac{2^{a^3}}{\mu (B(x,\frac{R}{2}))} \, \sup _{y\in X}|f(x)| \cdot \mu (\{ y\in X: R-\delta {\lt}\rho (x,y){\lt}R\} ). \end{align*}

By continuity from above of \(\mu \), the right factor becomes arbitrarily small as \(\delta \rightarrow 0\). Thus, for small enough \(\delta \), the whole expression is \(\le \epsilon \).

Lemma 10.1.8

Let \(f:X\to {\mathbb {C}}\) be a bounded measurable function supported on a set of finite measure. For all \(x\in X\),

\begin{equation} \label{tang-unm-op-eq} T_*f(x) = \sup _{R_1 {\lt} R_2} \sup _{x'\in B(x,R_1)} \left|\int _{B(x',R_2)\setminus B(x',R_1)} K(x',y) f(y) \, \mathrm{d}\mu (y) \right| \end{equation}
10.1.39

Proof

We show two inequalities. Let \(\epsilon {\gt}0\). Let \(R_1{\lt}R_2\) and \(x'\in B(x,R_1)\). Then for small enough \(\delta {\gt}0\),

\begin{alignat}{3} \label{eq-without-suprema-1} & & & \left|\int _{R_1{\lt}\rho (x',y){\lt}R_2} K(x’,y) f(y) \, \mathrm{d}\mu (y) \right| \\ \label{eq-diff-small-1} \le & & & \left|\int _{R_1{\lt}\rho (x',y){\lt}R_1+\delta } K(x’,y) f(y) \, \mathrm{d}\mu (y) \right| \\ \label{eq-other-1} & +& & \left|\int _{R_1+\delta \le \rho (x',y){\lt}R_2} K(x’,y) f(y) \, \mathrm{d}\mu (y) \right| . \end{alignat}

By Lemma 10.1.7, we can choose \(\delta \) such that 10.1.41 is bounded by \(\epsilon \). Without loss of generality, we can assume \(R_1+\delta {\lt}R_2\). Then 10.1.42 is bounded by the right hand side of 10.1.39 and we obtain

\begin{equation*} \le \epsilon + \sup _{R_1 {\lt} R_2} \sup _{x'\in B(x,R_1)} \left|\int _{B(x',R_2)\setminus B(x',R_1)} K(x’,y) f(y) \, \mathrm{d}\mu (y) \right| . \end{equation*}

The inequality still holds when taking the suprema over \(R_1{\lt}R_2\) and \(\rho (x,x'){\lt}R_1\) in 10.1.40. Since \(\epsilon {\gt}0\) was arbitrary, this proves the first inequality.

The other direction is similar. Let \(\epsilon {\gt}0\). Let \(R_1{\lt}R_2\) and \(x'\in B(x,R_1)\). Then for \(\delta {\gt}0\),

\begin{alignat}{3} \label{eq-without-suprema-2} & & & \left|\int _{B(x',R_2)\setminus B(x',R_1)} K(x’,y) f(y) \, \mathrm{d}\mu (y) \right| \\ \label{eq-diff-small-2} \le & & & \left|\int _{R_1-\delta {\lt}\rho (x',y){\lt} R_1} K(x’,y) f(y) \, \mathrm{d}\mu (y) \right| \\ \label{eq-other-2} & +& & \left|\int _{R_1-\delta {\lt}\rho (x',y){\lt}R_2} K(x’,y) f(y) \, \mathrm{d}\mu (y) \right| . \end{alignat}

By Lemma 10.1.7, we can choose \(\delta \) such that 10.1.44 is bounded by \(\epsilon \). Without loss of generality, we can assume \(\rho (x,x'){\lt}R_1-\delta \). Then 10.1.45 is bounded by the left hand side of 10.1.39 and we obtain

\begin{equation*} \le \epsilon + \sup _{R_1 {\lt} R_2} \sup _{x'\in B(x,R_1)} \left|\int _{R_1{\lt}\rho (x',y){\lt}R_2} K(x’,y) f(y) \, \mathrm{d}\mu (y) \right| . \end{equation*}

The inequality still holds when taking the suprema over \(R_1{\lt}R_2\) and \(\rho (x,x'){\lt}R_1\) in 10.1.40. Since \(\epsilon {\gt}0\) was arbitrary, this proves the second inequality.

Proof of Lemma 10.0.2

Fix \(g\) as in the Lemma. Applying Lemma 10.1.6 with a sequence of \(r\) tending to \(0\) and using Lebesgue monotone convergence shows

\begin{equation} \label{tzerobound} \| T_*^{0}g\| _2\le 2^{a^3+24a+6}\| g\| _2, \end{equation}
10.1.46

where

\begin{equation} \label{eq-simpler--nontangential} T_*^{0} g(x):=\sup _{0{\lt}R}\sup _{x'\in B(x,R)} \left|\int _{X\setminus B(x',R)} K(x',y) g(y)\, d\mu (y)\right|\, . \end{equation}
10.1.47

We now write using Lemma 10.1.8 and the triangle inequality,

\begin{alignat*}{3}\label{concretetstartriangle} T_* g(x)\le & & & \sup _{0{\lt}R_1{\lt}R_2}\sup _{x'\in B(x,R_1)} \left|\int _{X\setminus B(x',R_1)}K(x’,y) g(y)\, d\mu (y)\right| \\ & +& & \sup _{0{\lt}R_1{\lt}R_2}\sup _{x'\in B(x,R_1)} \left|\int _{X\setminus B(x',R_2)} K(x’,y) g(y)\, d\mu (y)\right|\, . \end{alignat*}

Noting that the first integral does not depend on \(R_2\) and estimating the second summand by the larger supremum over all \(x'\in B(x,R_2)\), at which time the integral does not depend on \(R_1\), we estimate further

\begin{alignat*}{3} \le & & & \sup _{0{\lt}R_1}\sup _{x'\in B(x,R_1)} \left|\int _{X\setminus B(x',R_1)}K(x’,y) g(y)\, d\mu (y)\right| \\ & + & & \sup _{0{\lt}R_2}\sup _{x'\in B(x,R_2)} \left|\int _{X\setminus B(x',R_2)} K(x’,y) g(y)\, d\mu (y)\right|\, . \end{alignat*}

Applying the triangle inequality on the left-hand side of 10.0.5 and applying 10.1.46 twice proves 10.0.5. This completes the proof of Lemma 10.0.2.

10.2 Calderón-Zygmund Decomposition

Calderón-Zygmund decomposition is a tool to extend \(L^2\) bounds to \(L^p\) bounds with \(p{\lt}2\) or to the so-called weak \((1, 1)\) type endpoint bound. It is classical and can be found in [ .

The following lemma is Theorem 3.1(b) in [ . The proof uses Proposition 2.0.6.

Lemma 10.2.1 Maximal theorem

Let \(f: X \to {\mathbb {C}}\) be bounded, measurable, supported on a set of finite measure, and let \(\alpha {\gt} 0\). Then

\begin{equation} \label{maximal-theorem-equation} \mu (\{ x\in X : Mf(x) {\gt} \alpha \} ) \le \frac{2^{2a}}{\alpha } \int |f(y)|\, d\mu (y). \end{equation}
10.2.1

Proof

By definition, for each \(x\in X\) with \(Mf(x) {\gt} \alpha \), there exists a ball \(B_x\) such that \(x\in B_x\) and

\begin{equation} \label{maximal-theorem-a} \alpha \mu (B_x) {\lt} \int _{B_x} |f(y)|\, d\mu (y). \end{equation}
10.2.2

Since \(\{ x\in X : Mf(x) {\gt} \alpha \} \) is open and \(\mu \) is inner regular on open sets, it suffices to show that

\begin{equation*} \mu (E) \le \frac{2^{2a}}{\alpha } \int |f(y)|\, d\mu (y) \end{equation*}

for every compact \(E\subset \{ x\in X : Mf(x) {\gt} \alpha \} \). For such an \(E\), by compactness, we can select a finite subcollection \(\mathcal{B} \subset \{ B_x: x\in E\} \) that covers \(E\). By 2.0.43 applied to 10.2.2,

\begin{equation} \alpha \mu (\bigcup \mathcal{B}) \le 2^{2a} \int |f(y)|\, d\mu (y) \end{equation}
10.2.3

and hence

\begin{equation*} \mu (E) \le \mu (\bigcup \mathcal{B}) \le \frac{2^{2a}}{\alpha } \int |f(y)|\, d\mu (y). \end{equation*}
Lemma 10.2.2 Lebesgue differentiation
#

Let \(f\) be a bounded measurable function supported on a set of finite measure. Then for \(\mu \) almost every \(x\), we have

\[ \lim _{n\to \infty } \frac{1}{\mu (B_n)}\int _{B_n} f(y)\, dy= f(x), \]

where \(\{ B_n\} _{n\geq 1}\) is a sequence of balls with radii \(r_n{\gt}0\) such that \(x\in B_n\) for each \(n\geq 1\) and

\begin{equation*} \lim _{n\to \infty } r_n=0 \, . \end{equation*}
Proof

This follows from the Lebesgue differentiation theorem, which is already formalized in Lean.

Lemma 10.2.3 Disjoint family countable

In a doubling metric measure space \((X,\rho ,\mu , a)\), every disjoint family of balls \(B_j = B(x_j, r_j)\), \(j\in J\), is countable.

Proof

Choose an arbitrary \(x\in X\) as reference point. For \(q, Q\in {\mathbb {Q}}_+\), let \(J_{q,Q}\) denote the set of all \(j\in J\) such that \(B_j\subset B(x, Q)\) and \(r_j \ge q\). It suffices to show that all the \(J_{q,Q}\) are finite. Indeed, for all \(j\in J_{q,Q}\),

\begin{equation*} \mu (B(x, Q)) \le \mu (B(x_j, 2Q)) = \mu (B(x_j, \frac{2 Q}{r_j} r_j)) \le 2^{a\log _2{\lceil \frac{2 Q}{r_j}\rceil }} \mu (B_j). \end{equation*}

Since the \(B_j\) are disjoint,

\begin{equation} |J_{q,Q}| \mu (B(x, Q)) \le 2^{a\log _2{\lceil \frac{2Q}{q}\rceil }} \sum _{j\in J_{q,Q}} \mu (B_j) \le 2^{a\log _2{\lceil \frac{2Q}{q}\rceil }} \mu (B(x,Q)) \end{equation}
10.2.4

and hence \(|J_{q,Q}| \le 2^{a\log _2{\lceil \frac{2Q}{q}\rceil }}\).

The following lemma corresponds to Lemma 3.2 in [ with additional proof of the bounded intersection property taken from the proof of Proposition 7.1 .

Lemma 10.2.4 Ball covering

Given an open set \(O\ne X\), there exists a countable family of balls \(B_j = B(x_j, r_j)\) such that

\begin{equation} \label{balls-disjoint} B_j \cap B_{j'} = \emptyset \quad \text{ for } j \ne j' \end{equation}
10.2.5

and for \(B_j^* := B(x_j, 3r_j)\),

\begin{equation} \label{balls-covering} \bigcup _j B_j^* = O \end{equation}
10.2.6

and for \(B_j^{**} := B(x_j, 7r_j)\),

\begin{equation} \label{enlarged-balls-intersect-complement} B_j^{**} \cap (X \setminus O) \ne \emptyset \quad \text{ for all } j \end{equation}
10.2.7

and we have the bounded intersection property that each \(x\in O\) is contained in at most \(2^{6a}\) of the \(B_j^*\).

Proof

Define for \(x\in O\),

\begin{equation} \delta (x):= \sup \{ \delta \in {\mathbb {R}}: B(x,\delta )\subset O\} . \end{equation}
10.2.8

Since \(O\) is open, and \(O\ne X\), we have

\begin{equation} 0 {\lt} \delta (x) {\lt} \infty \, . \end{equation}
10.2.9

Using Zorn’s Lemma, we select a maximal disjoint subfamily of \(\{ B(x,\frac{\delta (x)}{6}) : x \in O\} \). We obtain a (by Lemma 10.2.3 countable) family of balls \(B_j = B(x_j, \frac{\delta (x_j)}{6}), j \in J\) such that 10.2.5. 10.2.7 and \(\bigcup _j B_j^* \subset O\) are also immediate. For the other inclusion, first observe that for \(x,y\in X\), if \(B(x,\frac{\delta (x)}{6}) \cap B(y,\frac{\delta (y)}{6}) \ne \emptyset \), then

\begin{equation*} \delta (x) \le \rho (x,y) + \delta (y) \le (\frac{\delta (x)}{6} + \frac{\delta (y)}{6}) + \delta (y) = \frac{\delta (x)}{6} + \frac{7\delta (y)}{6}, \end{equation*}

so

\begin{equation} \label{control-distance-growth} \delta (x) \le 2 \delta (y). \end{equation}
10.2.10

Now let \(z\in X\). By maximality, there exists some \(j\in J\) with \(B(z,\frac{\delta (z)}{6}) \cap B_j \ne \emptyset \). By 10.2.10,

\begin{equation*} \rho (z,x_j){\lt} \frac{\delta (z)}{6} + \frac{\delta (x_j)}{6} \le \frac{3\delta (x_j)}{6} = 3r_j \end{equation*}

and thus \(z\in B_j^*\).

We now turn to the bounded intersection property. Assume that for some \(j_1,\dots ,j_N\),

\begin{equation} z\in \bigcap _{k=1}^N B_{j_k}^*. \end{equation}
10.2.11

Similarly as above, observe for \(1\le k \le N\),

\begin{equation} \label{control-distance-growth-b} \delta (z) \le \rho (z,x_{j_k}) + \delta (x_{j_k}) \le \frac{\delta (x_{j_k})}{2} + \delta (x_{j_k}) = \frac{3\delta (x_{j_k})}{2} \end{equation}
10.2.12

and

\begin{equation*} \delta (x_{j_k}) \le \rho (x_{j_k},z) + \delta (z) \le \frac{\delta (x_{j_k})}{2} + \delta (z), \end{equation*}

so

\begin{equation} \label{control-distance-growth-c} \delta (x_{j_k}) \le 2 \delta (z). \end{equation}
10.2.13

By 10.2.12 and 10.2.13, for all \(1\le k \le N\), \(B(z,\frac{\delta (z)}{6}) \subset B(x_{j_k}, 5r_{j_k})\) and \(B_{j_k} \subset B(z,\frac{8\delta (z)}{6})\). Using this and 10.2.5, we obtain

\begin{align} N \mu (B(z,\frac{\delta (z)}{6})) & \le \sum _{k=1}^N \mu (B(x_{j_k}, 5r_j)) \le 2^{3a} \sum _{k=1}^N \mu (B_{j_k}) \\ & = 2^{3a} \mu (\bigcup _{k=1}^N B_{j_k}) \le 2^{3a} \mu (B(z,\frac{8\delta (z)}{6})) \le 2^{6a} \mu (B(z,\frac{\delta (z)}{6})) \end{align}

and conclude \(N\le 2^{6a}\).

Most of the next lemma and its proof is taken from Theorem 4.2 in [ .

Lemma 10.2.5 Calderon Zygmund decomposition

Let \(f\) be a bounded, measurable function supported on a set of finite measure and let \(\alpha {\gt}\frac{1}{\mu (X)}\int |f|\, d\mu \). Then there exists a measurable function \(g\), a countable family of balls \(B_j^*\) (where we allow \(B_1^* = X\) in the special case that \(\mu (X){\lt}\infty \)) such that each \(x\in X\) is contained in at most \(2^{6a}\) of the \(B_j^*\), and a countable family of measurable functions \(\{ b_j\} _{j\in J}\) such that for all \(x \in X\)

\begin{equation} \label{eq-gb-dec} f(x)= g(x) + \sum _{j} b_j(x) \end{equation}
10.2.16

and such that the following holds. For almost every \(x\in X\),

\begin{equation} \label{eq-g-max} |g(x)|\leq 2^{3a} \alpha \, . \end{equation}
10.2.17

We have

\begin{equation} \label{eq-g-L1-norm} \int |g(y)|\, d\mu (y)\leq \int |f(y)|\, d\mu (y). \end{equation}
10.2.18

For every \(j\)

\begin{equation} \label{eq-supp-bj} \operatorname {supp} b_j \subset B_j^*\, . \end{equation}
10.2.19

For every \(j\)

\begin{equation} \label{eq-bad-mean-zero} \int _{B_j^*} b_j(x)\, d\mu (x)=0, \end{equation}
10.2.20

and

\begin{equation} \label{eq-bj-L1} \int _{B_j^*} |b_j(x)|\, d\mu (x) \leq 2^{2a+1} \alpha \mu (B_j^*). \end{equation}
10.2.21

We have

\begin{equation} \label{eq-bset-length-sum} \sum _j \mu (B_j^*)\leq \frac{2^{4a}}{\alpha }\int |f(y)|\, d\mu (y) \end{equation}
10.2.22

and

\begin{equation} \label{eq-b-L1} \sum _{j}\int _{B_j^*} |b_j(y)|\, d\mu (y)\leq 2 \int |f(y)|\, d\mu (y)\, . \end{equation}
10.2.23

Proof

Let \(E_\alpha :=\{ x\in X: Mf(x){\gt}\alpha \} \). Then \(E_\alpha \) is open. Assume first that \(E_\alpha \ne X\). We apply Lemma 10.2.4 with \(O=E_\alpha \) to obtain the family \(B_j, j\in J,\). Without loss of generality, we can assume \(J={\mathbb {N}}\). Define inductively

\begin{equation} Q_j := B_j^* \setminus \left(\bigcup _{i{\lt}j} Q_i \cup \bigcup _{i{\gt}j} B_i \right). \end{equation}
10.2.24

Then \(B_j\subset Q_j\subset B_j^*\), the \(Q_j\) are pairwise disjoint and \(\bigcup _j Q_j = E_\alpha \). Define

\begin{equation} \label{eq-g-def} g(x):=\begin{cases} f(x), & x\in X\setminus E_\alpha ,\\ \frac{1}{\mu (Q_j)}\int _{Q_j} f(y)\, d\mu (y), & x\in Q_j, \end{cases} \end{equation}
10.2.25

and, for each \(j\),

\begin{equation} b_j(x):= \mathbf{1}_{Q_j}(x) \left(f(x)-\frac{1}{\mu (Q_j)}\int _{Q_j} f(y)\, d\mu (y) \right). \end{equation}
10.2.28

Then 10.2.16, 10.2.19 and 10.2.20 are true by construction. For 10.2.17, we first do the case \(x\in X\setminus E_\alpha \). By definition of \(Mf\),

\begin{equation} \frac{1}{\mu (B)}\int _B |f(y)|\, d\mu (y) \le \alpha \end{equation}
10.2.29

for every ball \(B\subset X\) with \(x\in B\). It follows by Lemma 10.2.2 that for almost every \(x\in X\setminus E_\alpha \), \(|f(x)|\le \alpha \). In the case \(x\in E_\alpha \), there exists some \(j\in J\) with \(x\in Q_j\) and we have that

\begin{equation} \label{large-ball-estimate} \frac{1}{\mu (B_j^{**})} \int _{B_j^{**}} |f(y)| \, d\mu (y) \le \alpha \end{equation}
10.2.30

because \(B_j^{**}\cap (X\setminus E_\alpha ) \ne \emptyset \). We get

\begin{equation} |g(x)| \le \frac{1}{\mu (Q_j)}\int _{Q_j} |f(y)|\, d\mu (y) \le \frac{1}{\mu (B_j)}\int _{B_j^{**}} |f(y)| \, d\mu (y) \le 2^{3a}\alpha . \end{equation}
10.2.31

To prove 10.2.18, we estimate

\begin{align*} \int |g(z)|\, d\mu (z) & \le \int _{X\setminus E_\alpha } |f(z)|\, d\mu (z) + \sum _{j} \int _{Q_j}\frac{1}{\mu (Q_j)}\int _{Q_j}|f(y)|\, d\mu (y)\, d\mu (z) \\ & = \int |f(z)|\, d\mu (z). \end{align*}

Using the triangle inequality, we have that

\begin{align} \label{eq-bj-int} \int _{B_j^*} |b_j(y)|\, dy & \le \int _{Q_j} |f(y)|\, d\mu (y) + \int _{Q_j} \frac{1}{\mu (Q_j)}\int _{Q_j} |f(x)|\, d\mu (x)\, d\mu (y) \\ & = 2 \int _{Q_j} |f(y)|\, dy. \end{align}

With 10.2.30, we estimate further

\begin{equation} \le 2 \int _{B_j^{**}} |f(y)|\, dy \le 2\mu (B_j^{**})\alpha \le 2^{2a+1} \alpha \mu (B_j^*) \end{equation}
10.2.34

to obtain 10.2.21. Further, summing up 10.2.32 in \(j\) yields 10.2.23. At last, we estimate with Lemma 10.2.1

\begin{equation} \sum _j \mu (B_j^*) \le 2^{2a} \sum _j \mu (B_j) \le 2^{2a} \mu (E_\alpha ) \le \frac{2^{4a}}{\alpha }\int |f(y)|\, d\mu (y), \end{equation}
10.2.35

proving 10.2.22.

Assume now that \(E_\alpha = X\). It follows from Lemma 10.2.1 that then \(\mu (X){\lt}\infty \). Define

\begin{equation*} g := \frac{1}{\mu (X)} \int |f(y)|\, d\mu (y) \end{equation*}

and

\begin{equation*} b_1 := f - g. \end{equation*}

Then \(f = g + b_1\) and \(\operatorname{\operatorname {supp}}b_1 \subset B_1^*:=X\) and 10.2.16, 10.2.18, 10.2.19, 10.2.20 all hold immediately. By assumption, \(\alpha {\gt}\frac{1}{\mu (X)}\int |f|\, d\mu = g\), so 10.2.17 holds. We also have, using the definitions and the same assumption,

\begin{equation} \int |b_1(y)|\, d\mu (y) \le 2 \int |f(y)|\, d\mu (y) \le 2\alpha \mu (X), \end{equation}
10.2.36

which verifies both 10.2.23 and 10.2.21. Finally, by Lemma 10.2.1,

\begin{equation*} \mu (X) \le \frac{2^{2a}}{\alpha } \int |f(y)|\, d\mu (y), \end{equation*}

which shows 10.2.22.

We use Lemma 10.2.5 to prove Lemma 10.0.3. For the remainder of this section, let \(f:X\to {\mathbb {C}}\), \(r{\gt}0\) and \(\alpha {\gt}0\) as in the lemma. We define the constant

\begin{equation} \label{weak-1-1-proof-cz-const} c:= 2^{-a^3-12a-4} \end{equation}
10.2.37

and \(\alpha ' := c\alpha \). If \(\alpha '\le \frac{1}{\mu (X)}\int |f|\, d\mu \), then we directly have

\begin{equation*} \mu \left(\{ x\in X: |T_r f(x)|{\gt}\alpha \} \right)\le \mu (X) \le \frac{1}{\alpha '} \int |f(y)|\, d\mu (y) \le \frac{2^{a^3 + 19a}}{\alpha } \int |f(y)|\, d\mu (y), \end{equation*}

which proves 10.0.9. So assume from now on that \(\alpha '{\gt}\frac{1}{\mu (X)}\int |f|\, d\mu \). Using Lemma 10.2.5 for \(f\) and \(\alpha '\), we obtain the decomposition

\begin{equation*} f=g+b=g+\sum _j b_j \end{equation*}

such that the properties 10.2.16-10.2.23 are satisfied (with \(\alpha '\) replacing \(\alpha \)). We rename \(B_j^*\) to \(B_j\) and let

\begin{equation} \label{eq-Ij-cj} B_j=B(x_j, r_j). \end{equation}
10.2.38

Define

\begin{equation} \label{eq-Ij*} B_j':=B(x_j, 2r_j). \end{equation}
10.2.39

(In the special case \(B_j=X\), we define \(B_j':=X\).) Then \(B_j'\) is a ball with the same center as \(B_j\) but with

\begin{equation} \label{eq-Ij*-dim} \mu (B_j')\le 2^{a} \mu (B_j). \end{equation}
10.2.40

Let

\begin{equation} \label{eq-omega} \Omega :=\bigcup _j B_j'. \end{equation}
10.2.41

We deal with \(T_rg\) and \(T_rb\) separately in the following lemmas.

Lemma 10.2.6 Estimate good
\begin{equation*} \mu \left(\{ x\in X: |T_r g(x)|{\gt}{\alpha }/2\} \right) \le \frac{2^{2a^3+3a+2}c}{\alpha } \int |f(y)|\, d\mu (y). \end{equation*}
Proof

We estimate using monotonicity of the integral

\begin{equation*} \mu \left(\{ x\in X: |T_r g(x)|{\gt}{\alpha }/2\} \right)\leq \frac{4}{\alpha ^2} \int |T_r g(y)|^2\, d\mu (y). \end{equation*}

Using 10.0.8 followed by 10.2.17 and 10.2.18, we estimate the right hand side above by

\begin{equation} \label{eq-Hr-g} \leq \frac{4\cdot 2^{2a^3}}{\alpha ^2} \int |g(y)|^2\, d\mu (y)\leq \frac{2^{2a^3+3a+2}c}{\alpha } \int |g(y)|\, dy \le \frac{2^{2a^3+3a+2}c}{\alpha } \int |f(y)|\, d\mu (y). \end{equation}
10.2.42

Lemma 10.2.7

Let \(x\in X\setminus \Omega \). Then

\begin{equation*} |T_rb(x)| \le 3F(x)+\alpha /8, \end{equation*}

where

\begin{equation*} F(x) := 2^{a^3+2a+1} c\alpha \sum _{j\in J} \left(\frac{r_j}{\rho (x,x_j)}\right)^{\frac{1}{a}}\frac{\mu (B_j)}{V(x,x_j)}. \end{equation*}
Proof

We decompose the index set \(J\) into the following disjoint sets:

\begin{align*} \mathcal{J}_1(x)& :=\{ j\, : r+r_j \le \rho (x,x_j) \} ,\\ \mathcal{J}_2(x)& :=\{ j\, : r-r_j \le \rho (x,x_j) {\lt} r+r_j\} ,\\ \mathcal{J}_3(x)& :=\{ j\, : \rho (x,x_j) {\lt} r-r_j\} . \end{align*}

Then

\begin{alignat}{3} \label{eq-b-dec-1} |T_r b(x)|\le & & & \sum _{j\in \mathcal{J}_1(x)} |T_rb_j(x)| \\ \label{eq-b-dec-2} & +& & \sum _{j\in \mathcal{J}_2(x)} |T_rb_j(x)| \\ \label{eq-b-dec-3} & +& & \sum _{j\in \mathcal{J}_3(x)} |T_rb_j(x)|. \end{alignat}

For all \(j\in \mathcal{J}_3(x)\), \(\operatorname{\operatorname {supp}}b_j\subset B_j\subset B(x,r)\), and thus \(T_rb_j(x)=0\), so \(\eqref{eq-b-dec-3} = 0\).

Next, for \(j\in \mathcal{J}_1(x)\), \(\operatorname{\operatorname {supp}}b_j\subset B_j\subset X \setminus B(x,r)\), and we have

\begin{equation*} T_rb_j(x)=\int _{X\setminus B(x,r)} K(x,y) b_j(y)\, d\mu (y)=\int _{B_j} K(x,y) b_j(y)\, d\mu (y)\, . \end{equation*}

Using 10.2.20, the above is equal to

\begin{equation*} \int _{B_j} (K(x,y)-K(x,x_j)) b_j(y)\, d\mu (y)\, . \end{equation*}

Since \(x\in X\setminus \Omega \), we have for each \(y\in B_j\) that

\begin{equation} \label{eq-Om-cj} \rho (x,x_j)\ge 2r_j {\gt} 2\rho (x_j,y), \end{equation}
10.2.46

so we can apply 1.0.15 to estimate

\begin{equation*} \eqref{eq-b-dec-1} \le \sum _{j\in \mathcal{J}_1(x)}\int _{B_j}\left(\frac{\rho (x_j,y)}{\rho (x,x_j)}\right)^{\frac{1}{a}}\frac{2^{a^3}}{V(x,x_j)} |b_j(y)|\, d\mu (y) \end{equation*}

\begin{equation} \le 2^{a^3} \sum _{j} \left(\frac{r_j}{\rho (x,x_j)}\right)^{\frac{1}{a}}\frac{1}{V(x,x_j)}\int _{B_j} |b_j(y)|\, dy \end{equation}
10.2.47

and by 10.2.21,

\begin{equation} \label{eq-J1-diff-est} \le 2^{a^3+2a+1} c\alpha \sum _{j} \left(\frac{r_j}{\rho (x,x_j)}\right)^{\frac{1}{a}}\frac{\mu (B_j)}{V(x,x_j)} = F(x). \end{equation}
10.2.48

Next, we estimate 10.2.44. For each \(j\in \mathcal{J}_2(x)\), set

\begin{equation*} d_j:=\frac{1}{\mu (B_j)}\int _{B_j} \mathbf{1}_{X\setminus B(x,r)}(y) b_j(y)\, dy. \end{equation*}

Then by 10.2.21

\begin{equation} \label{eq-dj-est} |d_j|\le 2^{2a+1} c\alpha . \end{equation}
10.2.49

For each \(j\in \mathcal{J}_2(x)\), we have

\begin{align*} T_r b_j(x)& =\int _{B_j} K(x,y) (\mathbf{1}_{X\setminus B(x,r)}(y)b_j(y)-d_j)\, dy + \int _{B_j} d_j K(x,y) \, dy \\ & = \int _{B_j} (K(x,y)-K(x,x_j)) (\mathbf{1}_{X\setminus B(x,r)}(y)b_j(y)-d_j)\, dy + \int _{B_j} d_j K(x,y) \, dy. \end{align*}

Thus, using the triangle inequality, the equation above and 10.2.49, we obtain

\begin{equation*} |T_r b_j(x)|\le \end{equation*}

\begin{equation} \label{eq-J2-diff-est} \int _{B_j} |K(x,y)-K(x,x_j)| \left(|b_j(y)|+2^{2a+1} c\alpha \right)\, dy +2^{2a+1} c\alpha \int _{B_j} |K(x,y)| \, dy. \end{equation}
10.2.50

By 10.2.46, we can apply 1.0.15 and arguing as in 10.2.48, we get that

\begin{equation} \label{eq-J2-diff-est-2} \eqref{eq-b-dec-2} \le 2F(x) + 2^{2a+1}c\alpha \sum _{j\in \mathcal{J}_2(x)} \int _{B_j} |K(x,y)|\, d\mu (y), \end{equation}
10.2.51

with \(F\) as in 10.2.48. Define

\begin{equation*} A := \bigcup _{j \in \mathcal{J}_2(x)} B_j. \end{equation*}

We claim that

\begin{equation} \label{eq-J2-union-subset} A\subset B(x,3r) \setminus B(x,\frac{r}{3}). \end{equation}
10.2.52

Indeed, for each \(j\in \mathcal{J}_2(x)\) and \(y\in B_j\), using again 10.2.46,

\begin{equation*} \rho (x,x_j) {\lt} r+r_j \le r + \frac{1}{2} \rho (x,x_j) \implies \rho (x,x_j) {\lt} 2r \end{equation*}

and hence

\begin{equation*} \rho (x,y) \le \rho (x,x_j) + \rho (x_j,y) {\lt} 2r + r_j \le 2r + \frac{1}{2}\rho (x,x_j) {\lt} 3r. \end{equation*}

For the lower bound, we observe

\begin{equation*} \rho (x,x_j) \ge r-r_j \ge r - \frac{1}{2} \rho (x,x_j) \implies \rho (x,x_j) \ge \frac{2}{3}r, \end{equation*}

and conclude

\begin{equation*} \rho (x,y) \ge \rho (x,x_j) - \rho (y,x_j) \ge \rho (x,x_j) - r_j \ge \rho (x,x_j) - \frac{1}{2}\rho (x,x_j) \ge \frac{1}{3} r. \end{equation*}

Using the bounded intersection property of the \(B_j\), 10.2.52 and 1.0.14, we get

\begin{align} \sum _{j\in \mathcal{J}_2(x)} \int _{B_j} |K(x,y)|\, d\mu (y) & \le 2^{6a} \int _{A} |K(x,y)|\, d\mu (y) \\ & \le 2^{6a} \int _{B(x,3r) \setminus B(x,\frac{r}{3})} |K(x,y)|\, d\mu (y) \\ & \le 2^{6a} \int _{B(x,3r) \setminus B(x,\frac{r}{3})} \frac{2^{a^3}}{V(x,y)} \, d\mu (y) \\ & \le 2^{a^3+6a} \int _{B(x,3r) \setminus B(x,\frac{r}{3})} \frac{1}{\mu (B(x,\frac{r}{3}))} \, d\mu (y) \\ & \le 2^{a^3+6a} \frac{\mu (B(x,3r))}{\mu (B(x,\frac{r}{3}))} \\ \label{eq-J2-diff-est-4} & \le 2^{a^3+10a}. \end{align}

Combining the estimates 10.2.48 for 10.2.43, 10.2.51 for 10.2.44, and 10.2.58, we get

\begin{equation*} |T_rb(x)|\leq 3F(x)+2^{a^3+12a+1}c\alpha . \end{equation*}

By the definition 10.2.37 of \(c\), this equals

\begin{equation*} 3F(x)+\alpha /8. \end{equation*}
Lemma 10.2.8

For \(F\) as defined in Lemma 10.2.7, we have

\begin{equation} \label{eq-F-X-minus-Omega} \mu (\{ x\in X\setminus \Omega : F(x){\gt}\alpha /8\} ) \le \frac{2^{a^3+9a+4}}{\alpha } \int |f(y)|\, d\mu (y)\, . \end{equation}
10.2.59

Proof

We estimate

\begin{align} \mu (\{ x\in X\setminus \Omega : F(x){\gt} \alpha /8\} ) & \le \frac{8}{\alpha } \int _{X\setminus \Omega } F(x)\, d\mu (x) \\ & \le \frac{8}{\alpha } \int _{X\setminus \Omega } 2^{a^3+2a+1} c\alpha \sum _{j} \left(\frac{r_j}{\rho (x,x_j)}\right)^{\frac{1}{a}}\frac{\mu (B_j)}{V(x,x_j)}\, d\mu (x) \\ \label{eq-F-est-1} & \le 2^{a^3+2a+4} c \sum _{j} \mu (B_j) \int _{X\setminus B_j'} \left(\frac{r_j}{\rho (x,x_j)}\right)^{\frac{1}{a}}\frac{1}{V(x,x_j)}\, d\mu (x) \end{align}

Using

\begin{equation*} V(x,x_j) = \mu (B(x,\rho (x,x_j))) \ge 2^{-a}\mu (B(x,2\rho (x,x_j))) \ge 2^{-a} \mu (B(x_j,\rho (x_j,x))), \end{equation*}

we have for all \(j\in J\),

\begin{align*} & \int _{X\setminus B_j'} \left(\frac{r_j}{\rho (x,x_j)}\right)^{\frac{1}{a}}\frac{1}{V(x,x_j)}\, d\mu (x) \\ \le & 2^a \int _{X\setminus B_j'} \left(\frac{r_j}{\rho (x,x_j)}\right)^{\frac{1}{a}}\frac{1}{\mu (B(x_j,\rho (x_j,x)))}\, d\mu (x) \\ \le & 2^a \sum _{n=1}^\infty \int _{B(x_j,2^{n+1}r_j)\setminus B(x_j,2^n r_j)} \left(\frac{r_j}{2^n r_j}\right)^{\frac{1}{a}}\frac{1}{\mu (B(x_j,2^n r_j))}\, d\mu (x) \\ \le & 2^a \sum _{n=1}^\infty 2^{-\frac{n}{a}} \frac{\mu (B(x_j,2^{n+1}r_j))}{\mu (B(x_j,2^n r_j))} \\ \le & 2^{3a}, \end{align*}

where we used Lemma 10.1.1 in the last step. Plugging this into 10.2.62 and using 10.2.22, we conclude that

\begin{equation*} \mu (\{ x\in X\setminus \Omega : F(x){\gt} \alpha /8\} ) \le \frac{2^{a^3+9a+4}}{\alpha } \int |f(y)|\, d\mu (y)\, . \end{equation*}
Lemma 10.2.9 Estimate bad

We have

\begin{equation*} \mu \left({\{ x\in X: |T_r b(x)|{\gt}\alpha /2\} }\right) \le \frac{\frac{2^{5a}}{c} + 2^{a^3+9a+4}}{\alpha } \int |f(y)|\, d\mu (y) \, . \end{equation*}
Proof

We estimate

\begin{equation*} \mu \left(\{ x\in X: |T_r b(x)|{\gt}\alpha /2\} \right) \end{equation*}

\begin{equation} \label{eq-set-dec-2} \le \mu (\Omega ) + \mu \left(\{ x\in X\setminus \Omega : |T_r b(x)|{\gt}{\alpha }/2\} \right)\, . \end{equation}
10.2.63

Using 10.2.40 and 10.2.22, we conclude that

\begin{equation} \label{eq-omega-bd} \mu (\Omega ) \le \sum _{j} \mu (B_j') \le 2^{a} \sum _{j} \mu (B_j) \le \frac{2^{5a}}{c\alpha } \int |f(y)|\, d\mu (y)\, . \end{equation}
10.2.64

It follows from Lemma 10.2.7 and the triangle inequality that

\begin{equation} \label{eq-set-dec-3} \mu ({\{ x\in X\setminus \Omega : |T_r b(x)|{\gt}\alpha /2\} }) \le \mu (\{ x\in X\setminus \Omega : F(x){\gt} \alpha /8\} )\, . \end{equation}
10.2.65

The claim now follows from 10.2.63, 10.2.65 and 10.2.59.

Proof of Lemma 10.0.3

It follows by the triangle inequality and subadditivity of \(\mu \) that

\begin{equation*} \mu \left(\{ x\in X: |T_r f(x)|{\gt}\alpha \} \right) \end{equation*}
\begin{equation*} \label{eq-set-dec-1} \le \mu \left(\{ x\in X: |T_r g(x)|{\gt}{\alpha }/2\} \right) + \mu \left(\{ x\in X: |T_r b(x)|{\gt}{\alpha }/2\} \right). \end{equation*}

Using Lemma 10.2.6, Lemma 10.2.9 and the definition 10.2.37 of \(c\), we get

\begin{align*} \le & \frac{2^{2a^3+3a+2}c + \frac{2^{5a}}{c} + 2^{a^3+9a+4}}{\alpha } \int |f(y)|\, d\mu (y) \\ =& \frac{2^{a^3-9a-2} + 2^{a^3+17a+4} + 2^{a^3+9a+4}}{\alpha } \int |f(y)|\, d\mu (y) \\ \le & \frac{2^{a^3 + 19a}}{\alpha } \int |f(y)|\, d\mu (y). \end{align*}

This proves 10.0.9.