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

Holomorphic stability for Carleman pairs of function spaces

Yong Han Yong HAN: College of Mathematics and Statistics, Shenzhen University, Shenzhen 518060, Guangdong, China [email protected] Yanqi Qiu Yanqi Qiu: School of Mathematics and Statistics, Wuhan University, Wuhan 430072, Hubei, China; Institute of Mathematics & Hua Loo-Keng Key Laboratory of Mathematics, AMSS, CAS, Beijing 100190, China [email protected]  and  Zipeng Wang Zipeng WANG: College of Mathematics and Statistics, Chongqing University, Chongqing 401331, China [email protected]
Abstract.

We introduce a notion of holomorphic stability for pairs of function spaces on a planar domain Ω\Omega. In the case of the open unit disk Ω=𝔻\Omega=\mathbb{D} equipped with a radial measure μ\mu, by establishing Bourgain-Brezis type inequalities, we show that the pair

(B2(𝔻,μ),h1(𝔻))(B^{2}(\mathbb{D},\mu),h^{1}(\mathbb{D}))

of weighted harmonic Bergman space and harmonic Hardy space is holomorphically stable if and only if μ\mu is a (1,2)(1,2)-Carleson measure. With some extra efforts, we also obtain an analogous result for the upper half plane equipped with horizontal-translation invariant measures.

Key words and phrases:
holomorphic stability, weigthed Bergman spaces, Hardy spaces, Zen-type spaces, Bourgain-Brezis type inequalities
1991 Mathematics Subject Classification:
Primary 30H10, 30H20; Secondary 32A10, 32A36
YH is supported by the grant NSFC 11688101. YQ is supported by grants NSFC Y7116335K1, NSFC 11801547 and NSFC 11688101. ZW is supported by the grant NSFC 11601296

1. Introduction

1.1. Holomorphic stability for Carleman pairs

Let Ω\Omega be a planar domain and let 𝒪(Ω)\mathcal{O}(\Omega) be the space of holomorphic functions on Ω\Omega. In what follows, we shall always consider a pair (X,Y)(X,Y) of vector spaces both consisting of functions on Ω\Omega.

Definition (Holomorphic stability).

The pair (X,Y)(X,Y) is called holomorphically stable if one of the following equivalent conditions is satisfied:

  • (i)

    (X+𝒪(Ω))YX(X+\mathcal{O}(\Omega))\cap Y\subset X;

  • (ii)

    (X+𝒪(Ω))(YX)=;(X+\mathcal{O}(\Omega))\cap(Y\setminus X)=\emptyset;

  • (iii)

    (X+Y)𝒪(Ω)=X𝒪(Ω).(X+Y)\cap\mathcal{O}(\Omega)=X\cap\mathcal{O}(\Omega).

Here X+𝒪(Ω)={f+g|fX,g𝒪(Ω)}X+\mathcal{O}(\Omega)=\{f+g|f\in X,g\in\mathcal{O}(\Omega)\} and X+Y={f+g|fX,gY}X+Y=\{f+g|f\in X,g\in Y\}.

The Venn diagram in Figure 1 illustrates a holomorphically stable pair (X,Y)(X,Y).

Refer to caption
Figure 1. (X+𝒪(Ω))YX.(X+\mathcal{O}(\Omega))\cap Y\subset X.\qquad\quad\quad

For any fXf\in X and g𝒪(Ω)g\in\mathcal{O}(\Omega), we interpret f+gf+g as a holomorphic perturbation of ff. Then the holomorphic stability of a pair (X,Y)(X,Y) means that a holomorphic perturbation of any element in XX remains in XX provided that it is contained in YY after the perturbation.

Note that a holomorphically stable pair (X,Y)(X,Y) always satisfies the condition

Y𝒪(Ω)X.Y\cap\mathcal{O}(\Omega)\subset X.

Such pairs will be referred as Carleman pairs. This terminology comes from the classical Carleman embeddings of some classical holomorphic-function spaces on the open unit disk 𝔻\mathbb{D} in the complex plane (see [Car21, Sai79, GK89, Vuk03]). For its generalization to complex domains of arbitrary dimension, see Hörmander [Hor67]. Note that if YXY\subset X, then the pair (X,Y)(X,Y) is trivially holomorphically stable. Therefore, the notion of the holomorphic stability is of interests only for the pairs (X,Y)(X,Y) satisfying

YX and Y𝒪(Ω)X.Y\not\subset X\text{\, and \,}Y\cap\mathcal{O}(\Omega)\subset X.

Our research is inspired by Da Lio, Rivière and Wettstein’s very recent work [DLRW21] on Bourgain-Brezis type inequalities, where they essentially proved that the pair

(B2(𝔻),h1(𝔻))(B^{2}(\mathbb{D}),h^{1}(\mathbb{D}))

of the harmonic Bergman space B2(𝔻)B^{2}(\mathbb{D}) and the classical harmonic Hardy space h1(𝔻)h^{1}(\mathbb{D}) is holomorphically stable (the precise definitions of B2(𝔻)B^{2}(\mathbb{D}) and h1(𝔻)h^{1}(\mathbb{D}) are given in §1.2). The notion of holomorphic stability leads to many natural questions. Generalization of our work in more general planar domains (including the non-simply connected ones) is more involved and will be given in the sequel to this paper.

1.2. Main results

The harmonic Hardy space h1(𝔻)h^{1}(\mathbb{D}) is defined by

h1(𝔻):={u𝒪h(𝔻)|uh1(𝔻)=sup0<r<112π02π|u(reiθ)|𝑑θ<},h^{1}(\mathbb{D}):=\Big{\{}u\in\mathcal{O}_{h}(\mathbb{D})\Big{|}\|u\|_{h^{1}(\mathbb{D})}=\sup_{0<r<1}\frac{1}{2\pi}\int_{0}^{2\pi}|u(re^{i\theta})|d\theta<\infty\Big{\}},

where 𝒪h(𝔻)\mathcal{O}_{h}(\mathbb{D}) denotes the space of all harmonic functions on 𝔻\mathbb{D}. The Hardy space H1(𝔻)H^{1}(\mathbb{D}) is then defined as H1(𝔻):=h1(𝔻)𝒪(𝔻)H^{1}(\mathbb{D}):=h^{1}(\mathbb{D})\cap\mathcal{O}(\mathbb{D}).

Throughout the paper, all measures are assumed to be positive measures. Given a measure μ\mu on 𝔻\mathbb{D}, the associated weighted harmonic Bergman space B2(𝔻,μ)B^{2}(\mathbb{D},\mu) and weighted Bergman space A2(𝔻,μ)A^{2}(\mathbb{D},\mu) are defined as

B2(𝔻,μ):=L2(𝔻,μ)𝒪h(𝔻) and A2(𝔻,μ):=L2(𝔻,μ)𝒪(𝔻),B^{2}(\mathbb{D},\mu):=L^{2}(\mathbb{D},\mu)\cap\mathcal{O}_{h}(\mathbb{D})\text{\, and \,}A^{2}(\mathbb{D},\mu):=L^{2}(\mathbb{D},\mu)\cap\mathcal{O}(\mathbb{D}),

both of which inherit the norm of L2(𝔻,μ)L^{2}(\mathbb{D},\mu). If μ\mu is the Lebesgue measure on 𝔻\mathbb{D}, then we use the simplified notation B2(𝔻)B^{2}(\mathbb{D}) and A2(𝔻)A^{2}(\mathbb{D}).

A measure μ\mu on 𝔻\mathbb{D} is called boundary-accessable if its support is not relatively compact in 𝔻\mathbb{D}. Note that if μ\mu is boundary-inaccessable, then it is trivial to verify that the pair (B2(𝔻,μ),h1(𝔻))(B^{2}(\mathbb{D},\mu),h^{1}(\mathbb{D})) is holomorphically stable. Therefore, in what follows, we always assume that μ\mu is boundary-accessable.

A measure μ\mu on 𝔻\mathbb{D} is called a (1,2)(1,2)-Carleson measure if there exists a constant C>0C>0 such that

(1.1) (𝔻|f(z)|2μ(dz))1/2CfH1(𝔻),fH1(𝔻).\displaystyle\Big{(}\int_{\mathbb{D}}|f(z)|^{2}\mu(dz)\Big{)}^{1/2}\leq C\|f\|_{H^{1}(\mathbb{D})},\quad\forall f\in H^{1}(\mathbb{D}).
Theorem 1.1.

Let μ\mu be a radial boundary-accessable measure on 𝔻\mathbb{D}. Then the pair

(B2(𝔻,μ),h1(𝔻))(B^{2}(\mathbb{D},\mu),h^{1}(\mathbb{D}))

is holomorphically stable if and only if μ\mu is a (1,2)(1,2)-Carleson measure.

A natural conjecture is

Conjecture.

For any boundary-accessable measure μ\mu on 𝔻\mathbb{D}, the pair (B2(𝔻,μ),h1(𝔻))(B^{2}(\mathbb{D},\mu),h^{1}(\mathbb{D})) is holomorphically stable if and only if μ\mu is a (1,2)(1,2)-Carleson measure.

For a finite measure μ\mu on 𝔻\mathbb{D} which is not necessarily radial, a simple situation (which in general is rather different from the situation in Theorem 1.1, see Theorem 1.2 below for more details) for the holomorphic stability of (B2(𝔻,μ),h1(𝔻))(B^{2}(\mathbb{D},\mu),h^{1}(\mathbb{D})) is provided as follows. Consider the linear map 𝒬+\mathcal{Q}_{+} defined on the space 𝒪h(𝔻)\mathcal{O}_{h}(\mathbb{D}) by

(1.2) 𝒬+(n0anzn+n1bnz¯n):=n0anzn.\displaystyle\mathcal{Q}_{+}\Big{(}\sum_{n\geq 0}a_{n}z^{n}+\sum_{n\geq 1}b_{n}\bar{z}^{n}\Big{)}:=\sum_{n\geq 0}a_{n}z^{n}.

Then the pair (B2(𝔻,μ),h1(𝔻))(B^{2}(\mathbb{D},\mu),h^{1}(\mathbb{D})) is holomorphically stable if both

(1.3) 𝒬+:h1(𝔻)A2(𝔻,μ)\displaystyle\mathcal{Q}_{+}:h^{1}(\mathbb{D})\longrightarrow A^{2}(\mathbb{D},\mu)

and

(1.4) 𝒬+:B2(𝔻,μ)A2(𝔻,μ)\displaystyle\mathcal{Q}_{+}:B^{2}(\mathbb{D},\mu)\longrightarrow A^{2}(\mathbb{D},\mu)

are bounded linear operators.

It is not hard to see that the operator (1.3) is bounded if and only if

(1.5) supθ[0,2π)𝔻μ(dz)|1eiθz|2<.\displaystyle\sup_{\theta\in[0,2\pi)}\int_{\mathbb{D}}\frac{\mu(dz)}{|1-e^{-i\theta}z|^{2}}<\infty.

Hence the boundedness of the operator (1.3) in general fails even for a radial (1,2)(1,2)-Carleson measure. On the other hand, the boundedness of (1.4) holds for all radial finite measure μ\mu on 𝔻\mathbb{D}. For general weights, a clear sufficient condition for the boundedness of the operator (1.4) is that the Bergman projection being bounded on L2(𝔻,μ)L^{2}(\mathbb{D},\mu) (which then is equivalent to the condition that μ\mu is a B2B_{2}-weight à la Békollé-Bonami, see [BB78] for more details on Bergman projections). Consequently, for any B2B_{2}-weight μ\mu on 𝔻\mathbb{D} satisfying (1.5), the pair (B2(𝔻,μ),h1(𝔻))(B^{2}(\mathbb{D},\mu),h^{1}(\mathbb{D})) is holomorphically stable.

For a radial boundary-accessable finite measure μ\mu on 𝔻\mathbb{D}, the space B2(𝔻,μ)B^{2}(\mathbb{D},\mu) is complete and B2(𝔻,μ)+h1(𝔻)B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}) is a Banach space equipped with the norm:

fB2(𝔻,μ)+h1(𝔻):=inf{gB2(𝔻,μ)+hh1(𝔻)|f=g+h,gB2(𝔻,μ) and hh1(𝔻)}.\|f\|_{B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})}:=\inf\Big{\{}\|g\|_{B^{2}(\mathbb{D},\mu)}+\|h\|_{h^{1}(\mathbb{D})}\Big{|}f=g+h,\,\,g\in B^{2}(\mathbb{D},\mu)\text{\, and \,}h\in h^{1}(\mathbb{D})\Big{\}}.

Recall that a closed subspace B1B_{1} of a Banach space BB is called complemented in BB if there exists a bounded linear projection from BB onto B1B_{1}.

Theorem 1.2.

Let μ\mu be a radial boundary-accessable (1,2)(1,2)-Carleson measure on 𝔻\mathbb{D}. Then

(B2(𝔻,μ)+h1(𝔻))𝒪(𝔻)(B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}))\cap\mathcal{O}(\mathbb{D})

is a closed subspace of B2(𝔻,μ)+h1(𝔻)B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}). Moreover, the above subspace is complemented in B2(𝔻,μ)+h1(𝔻)B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}) if and only if μ\mu satisfies the condition

(1.6) 𝔻μ(dz)1|z|2<.\displaystyle\int_{\mathbb{D}}\frac{\mu(dz)}{1-|z|^{2}}<\infty.
Remark.

For radial measures on 𝔻\mathbb{D}, the conditions (1.5) and (1.6) are clearly equivalent.

In general, the notion of holomorphic stability for the pair (B2(Ω,μ),h1(Ω))(B^{2}(\Omega,\mu),h^{1}(\Omega)) is not conformally invariant, where Ω\Omega is assumed to have a nice boundary and h1(Ω)h^{1}(\Omega) is then defined as the set of all the Poisson convolutions of functions in L1(Ω,ds)L^{1}(\partial\Omega,ds) (where dsds is the arc-length measure on Ω\partial\Omega). In particular, our following result for the upper half plane does not seem to be a direct consequence of the result on the unit disk and its proof requires more efforts.

Let ={z|(z)>0}\mathbb{H}=\{z\in\mathbb{C}|\Im(z)>0\} denote the upper half plane. In this case, the suitable spaces for studying the holomorphic stability are the harmonic Zen-type spaces (which reduces to the ordinary harmonic Bergman space when the weight is the Lebesgue measure on \mathbb{H}), see [Har09, JPP13].

A measure μ\mu on \mathbb{H} is called boundary-accessable if its support is not contained in

ε:={z|(z)>ε}\mathbb{H}_{\varepsilon}:=\{z\in\mathbb{C}|\Im(z)>\varepsilon\}

for any ε>0\varepsilon>0. Given a horizontal translation-invariant boundary-accessable measure μ\mu on \mathbb{H}, define the harmonic Zen-type space by

(1.7) 2(,μ):={g𝒪h()|g2(,μ)=supL>0(|g(z+iL)|2μ(dz))1/2<},\displaystyle\mathscr{B}^{2}(\mathbb{H},\mu):=\Big{\{}g\in\mathcal{O}_{h}(\mathbb{H})\Big{|}\|g\|_{\mathscr{B}^{2}(\mathbb{H},\mu)}=\sup_{L>0}\Big{(}\int_{\mathbb{H}}|g(z+iL)|^{2}\mu(dz)\Big{)}^{1/2}<\infty\Big{\}},

where 𝒪h()\mathcal{O}_{h}(\mathbb{H}) denotes the set of all harmonic functions on \mathbb{H}. It is easy to see that the above space 2(,μ)\mathscr{B}^{2}(\mathbb{H},\mu) is complete and thus is a Hilbert space.

The relation between 2(,μ)\mathscr{B}^{2}(\mathbb{H},\mu) and the ordinary weighted harmonic Bergman space

B2(,μ):=L2(,μ)𝒪h()B^{2}(\mathbb{H},\mu):=L^{2}(\mathbb{H},\mu)\cap\mathcal{O}_{h}(\mathbb{H})

is given as follows. For any g𝒪h()g\in\mathcal{O}_{h}(\mathbb{H}) and any y>0y>0, define gy:g_{y}:\mathbb{R}\rightarrow\mathbb{C} by

(1.8) gy(x):=g(x+iy),x.\displaystyle g_{y}(x):=g(x+iy),\,\forall\,x\in\mathbb{R}.

Recall the Poisson kernel for \mathbb{H} at the point z=x+iyz=x+iy\in\mathbb{H}:

(1.9) Pz(t)=1πy(xt)2+y2,t.\displaystyle P_{z}^{\mathbb{H}}(t)=\frac{1}{\pi}\frac{y}{(x-t)^{2}+y^{2}},\quad t\in\mathbb{R}.

Set

(1.10) Poi():={g𝒪h()|gyL2() and gy+y=Piygyy,y>0}.\displaystyle\mathrm{Poi}(\mathbb{H}):=\Big{\{}g\in\mathcal{O}_{h}(\mathbb{H})\Big{|}g_{y}\in L^{2}(\mathbb{R})\text{\, and \,}g_{y+y^{\prime}}=P_{iy^{\prime}}^{\mathbb{H}}*g_{y}\,\,\forall y,y^{\prime}>0\Big{\}}.

One can easily check that, for any horizontal translation-invariant boundary-accessable measure μ\mu on \mathbb{H},

(1.11) 2(,μ)=B2(,μ)Poi().\displaystyle\mathscr{B}^{2}(\mathbb{H},\mu)=B^{2}(\mathbb{H},\mu)\cap\mathrm{Poi}(\mathbb{H}).

The harmonic Hardy space h1()h^{1}(\mathbb{H}) is defined by

h1():={u𝒪h()|uh1()=supy>0|u(x+iy)|𝑑x<}h^{1}(\mathbb{H}):=\Big{\{}u\in\mathcal{O}_{h}(\mathbb{H})\Big{|}\|u\|_{h^{1}(\mathbb{H})}=\sup_{y>0}\int_{\mathbb{R}}|u(x+iy)|dx<\infty\Big{\}}

and the Hardy space H1()H^{1}(\mathbb{H}) is defined as H1():=h1()𝒪()H^{1}(\mathbb{H}):=h^{1}(\mathbb{H})\cap\mathcal{O}(\mathbb{H}). A measure μ\mu on \mathbb{H} is called a (1,2)(1,2)-Carleson measure if there exists a constant C>0C>0 such that

(1.12) (|f(z)|2μ(dz))1/2CfH1(),fH1().\displaystyle\Big{(}\int_{\mathbb{H}}|f(z)|^{2}\mu(dz)\Big{)}^{1/2}\leq C\|f\|_{H^{1}(\mathbb{H})},\quad\forall f\in H^{1}(\mathbb{H}).
Theorem 1.3.

Let μ\mu be a horizontal translation-invariant boundary-accessable measure on \mathbb{H}. Then the pair

(2(,μ),h1())(\mathscr{B}^{2}(\mathbb{H},\mu),h^{1}(\mathbb{H}))

is holomorphically stable if and only if μ\mu is a (1,2)(1,2)-Carleson measure.

1.3. Sketch of the proof

Here we give a sketch of the proof of Theorem 1.1. The proof is based on a generalization of the following one-dimensional Bourgain-Brezis-type inequality due to Da Lio-Rivière-Wettstein: there exists a universal constant C>0C>0 such that for any smooth function uC(𝕋)u\in C^{\infty}(\mathbb{T}) with u(eiθ)𝑑θ=0\int u(e^{i\theta})d\theta=0,

(1.13) uL2(𝕋)C((Δ)1/4uH1/2(𝕋)+L1(𝕋)+(Δ)1/4uH1/2(𝕋)+L1(𝕋)),\displaystyle\|u\|_{L^{2}(\mathbb{T})}\leq C\Big{(}\|(-\Delta)^{1/4}u\|_{H^{-1/2}(\mathbb{T})+L^{1}(\mathbb{T})}+\|\mathcal{H}(-\Delta)^{1/4}u\|_{H^{-1/2}(\mathbb{T})+L^{1}(\mathbb{T})}\Big{)},

where the Hilbert transform of uu is given by

u(eiθ)=nsgn(n)u^(n)einθ\mathcal{H}u(e^{i\theta})=\sum_{n\in\mathbb{Z}}\text{sgn}(n)\widehat{u}(n)e^{in\theta}

and the 1/41/4-fractional Laplace transform (Δ)1/4u(-\Delta)^{1/4}u is given by

(Δ)1/4u(eiθ)=n|n|1/2u^(n)einθ.(-\Delta)^{1/4}u(e^{i\theta})=\sum_{n\in\mathbb{Z}}|n|^{1/2}\widehat{u}(n)e^{in\theta}.

The inequality (1.13) implies

(1.14) fB2(𝔻)C(fB2(𝔻)+h1(𝔻)+fB2(𝔻)+h1(𝔻)),f𝒪h(𝔻)L(𝔻),\displaystyle\|f\|_{B^{2}(\mathbb{D})}\leq C(\|f\|_{B^{2}(\mathbb{D})+h^{1}(\mathbb{D})}+\|\mathcal{H}f\|_{B^{2}(\mathbb{D})+h^{1}(\mathbb{D})}),\quad\forall f\in\mathcal{O}_{h}(\mathbb{D})\cap L^{\infty}(\mathbb{D}),

where, slightly by abusing the notation, f\mathcal{H}f is defined by

(1.15) f(z)=n1anznn1bnz¯n,provided that f(z)=n0anzn+n1bnz¯n\displaystyle\mathcal{H}f(z)=\sum_{n\geq 1}a_{n}z^{n}-\sum_{n\geq 1}b_{n}\bar{z}^{n},\quad\text{provided that\,\,}f(z)=\sum_{n\geq 0}a_{n}z^{n}+\sum_{n\geq 1}b_{n}\bar{z}^{n}

In our situation, we are able to prove that, if μ\mu is a radial boundary-accessable measure, then μ\mu is a (1,2)(1,2)-Carleson measure on 𝔻\mathbb{D} if and only if there exists a constant Cμ>0C_{\mu}>0 such that for any bounded harmonic function f𝒪h(𝔻)L(𝔻)f\in\mathcal{O}_{h}(\mathbb{D})\cap L^{\infty}(\mathbb{D}),

(1.16) fB2(𝔻,μ)Cμ(fB2(𝔻,μ)+h1(𝔻)+fB2(𝔻,μ)+h1(𝔻)).\displaystyle\|f\|_{B^{2}(\mathbb{D},\mu)}\leq C_{\mu}(\|f\|_{B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})}+\|\mathcal{H}f\|_{B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})}).

The inequality (1.16) applied to holomorphic functions immediately gives the result stated in Theorem 1.1.

The proof of the inequality (1.16) relies on a weighted version of Bourgain-Brezis-type inequality obtained in Theorem 1.4 below. More precisely, define a Fourier multiplier operator by

(1.17) 𝒜μun(𝔻|z|2|n|μ(dz))1/2u^(n)einθ,uC(𝕋),\displaystyle\mathcal{A}_{\mu}u\sim\sum_{n\in\mathbb{Z}}\Big{(}\int_{\mathbb{D}}|z|^{2|n|}\mu(dz)\Big{)}^{-1/2}\widehat{u}(n)e^{in\theta},\quad u\in C^{\infty}(\mathbb{T}),

where

u^(n):=12π02πu(eiθ)𝑑θ,n.\widehat{u}(n):=\frac{1}{2\pi}\int_{0}^{2\pi}u(e^{i\theta})d\theta,\quad n\in\mathbb{Z}.

Define also a Sobolev-type space corresponding to the radial weight μ\mu by

(1.18) Hμ(𝕋):={vnv^(n)einθ|vHμ(𝕋)=(n|v^(n)|2𝔻|z|2|n|μ(dz))1/2<}.\displaystyle H_{\mu}(\mathbb{T}):=\Big{\{}v\sim\sum_{n\in\mathbb{Z}}\widehat{v}(n)e^{in\theta}\Big{|}\|v\|_{H_{\mu}(\mathbb{T})}=\Big{(}\sum_{n\in\mathbb{Z}}|\widehat{v}(n)|^{2}\int_{\mathbb{D}}|z|^{2|n|}\mu(dz)\Big{)}^{1/2}<\infty\Big{\}}.
Remark.

For a general μ\mu, the coefficients of a formal Fourier series vHμ(𝕋)v\in H_{\mu}(\mathbb{T}) may have non-polynomial growth and it may not represent a distribution in 𝒟(𝕋)\mathcal{D}^{\prime}(\mathbb{T}). However, if μ\mu is the Lebesgue measure on 𝔻\mathbb{D}, then Hμ(𝕋)H_{\mu}(\mathbb{T}) is the Sobolev space H1/2(𝕋)𝒟(𝕋)H^{-1/2}(\mathbb{T})\subset\mathcal{D}^{\prime}(\mathbb{T}).

Theorem 1.4.

Let μ\mu be a radial boundary-accessable finite measure on 𝔻\mathbb{D}. Then μ\mu is a (1,2)(1,2)-Carleson measure if and only if there exists a universal constant Cμ>0C_{\mu}>0 such that for any smooth function uC(𝕋)u\in C^{\infty}(\mathbb{T}),

(1.19) uL2(𝕋)Cμ(𝒜μuHμ(𝕋)+L1(𝕋)+𝒜μuHμ(𝕋)+L1(𝕋)).\displaystyle\|u\|_{L^{2}(\mathbb{T})}\leq C_{\mu}\Big{(}\|\mathcal{A}_{\mu}u\|_{H_{\mu}(\mathbb{T})+L^{1}(\mathbb{T})}+\|\mathcal{H}\mathcal{A}_{\mu}u\|_{H_{\mu}(\mathbb{T})+L^{1}(\mathbb{T})}\Big{)}.

2. Preliminaries on Carleson measures

Recall that throughout the paper, all measures are assumed to be positive. We shall use the famous geometric characterization of the (1,2)(1,2)-Carleson measures on 𝔻\mathbb{D} defined in (1.1). For any interval I𝕋I\subset\mathbb{T}, the Carleson box SIS_{I} is defined by

SI={z𝔻|z|z|I,1|I|2π|z|<1},S_{I}=\Big{\{}z\in\mathbb{D}\Big{|}\frac{z}{|z|}\in I,1-\frac{|I|}{2\pi}\leq|z|<1\Big{\}},

where |I||I| denotes the arc-length of II. Let |SI||S_{I}| denote the Lebesgue measure of SIS_{I}, then a measure μ\mu on 𝔻\mathbb{D} is a (1,2)(1,2)-Carleson measure if and only if (see [Car62] and [Dur69])

supI is an arc in 𝕋μ(SI)|SI|<.\sup_{\text{$I$ is an arc in $\mathbb{T}$}}\frac{\mu(S_{I})}{|S_{I}|}<\infty.

In particular, we have

Lemma 2.1.

Let μ(dz)=σ(dr)dθ\mu(dz)=\sigma(dr)d\theta be a radial measure on 𝔻\mathbb{D}. Then μ\mu is a (1,2)(1,2)-Carleson measure if and only if

(2.1) sup0<δ<1σ([1δ,1))δ<.\displaystyle\sup_{0<\delta<1}\frac{\sigma([1-\delta,1))}{\delta}<\infty.

The (1,2)(1,2)-Carleson measures on the upper half plane \mathbb{H} is defined in (1.12) and its geometric characterization (see, e.g., [Ryd20, Thm. 2.1]) is given as follows: a positive Radon measure μ\mu on \mathbb{H} is a (1,2)(1,2)-Carleson measure on \mathbb{H} if and only if

supI is an interval in μ(QI)|QI|<,\sup_{\text{$I$ is an interval in $\mathbb{R}$}}\frac{\mu(Q_{I})}{|Q_{I}|}<\infty,

where |QI||Q_{I}| is the Lebesgue measure of the Carleson box QIQ_{I} defined by

QI={z=x+iy|xI,0<y<|I|},Q_{I}=\Big{\{}z=x+iy\in\mathbb{H}\Big{|}x\in I,0<y<|I|\Big{\}},

here |I||I| denotes the Lebesgue measure of the interval II\subset\mathbb{R}. In particular, we have

Lemma 2.2.

Let μ(dz)=dxΠ(dy)\mu(dz)=dx\Pi(dy) be a horizontal translation-invariant measure on \mathbb{H}. Then μ\mu is a (1,2)(1,2)-Carleson measure if and only if

(2.2) supy>0Π((0,y])y<.\displaystyle\sup_{y>0}\frac{\Pi((0,y])}{y}<\infty.
Remark.

While all (1,2)(1,2)-Carleson measures on 𝔻\mathbb{D} are finite, a (1,2)(1,2)-Carleson measure on \mathbb{H} needs not be.

3. Holomorphic stability: the disk case

This section is mainly devoted to proving Theorems 1.1 and 1.2. We shall use the following elementary observation: if μ(dz)=σ(dr)dθ\mu(dz)=\sigma(dr)d\theta is a radial boundary-accessable finite measure on 𝔻\mathbb{D}, then

  • both A2(𝔻,μ)A^{2}(\mathbb{D},\mu) and B2(𝔻,μ)B^{2}(\mathbb{D},\mu) are closed in L2(𝔻,μ)L^{2}(\mathbb{D},\mu);

  • for any z𝔻z\in\mathbb{D}, the evaluation map evz:B2(𝔻,μ)+h1(𝔻)\mathrm{ev}_{z}:B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})\longrightarrow\mathbb{C} defined by

    (3.1) evz(f)=f(z)\displaystyle\mathrm{ev}_{z}(f)=f(z)

    is a continuous linear functional on B2(𝔻,μ)+h1(𝔻)B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D});

  • for any ρ(0,1)\rho\in(0,1) and any kk\in\mathbb{N},

    (3.2) σk=01r2kσ(dr)[ρ,1)r2kσ(dr)ρ2kσ([ρ,1)).\displaystyle\sigma_{k}=\int_{0}^{1}r^{2k}\sigma(dr)\geq\int_{[\rho,1)}r^{2k}\sigma(dr)\geq\rho^{2k}\sigma([\rho,1)).

Recall the definition (1.18) of the space Hμ(𝕋)H_{\mu}(\mathbb{T}): for any vHμ(𝕋)v\in H_{\mu}(\mathbb{T}), we set

(3.3) vHμ(𝕋)2=n|v^(n)|2𝔻|z|2|n|μ(dz)=2πn|v^(n)|2σn.\displaystyle\|v\|_{H_{\mu}(\mathbb{T})}^{2}=\sum_{n\in\mathbb{Z}}|\widehat{v}(n)|^{2}\int_{\mathbb{D}}|z|^{2|n|}\mu(dz)=2\pi\sum_{n\in\mathbb{Z}}|\widehat{v}(n)|^{2}\sigma_{n}.

By (3.2) and (3.3), for any r[0,1)r\in[0,1), the Poisson transformation Pr𝔻vP_{r}^{\mathbb{D}}*v of an element vHμ(𝕋)v\in H_{\mu}(\mathbb{T}) is a smooth function given by

Pr𝔻v(eiθ)=nr|n|v^(n)einθ.P_{r}^{\mathbb{D}}*v(e^{i\theta})=\sum_{n\in\mathbb{Z}}r^{|n|}\widehat{v}(n)e^{in\theta}.

Any vHμ(𝕋)v\in H_{\mu}(\mathbb{T}) has a natural harmonic extension (denoted again by vv) on 𝔻\mathbb{D}:

(3.6) v(z)=nv^(n)en(z)withen(z)={znif n0z¯|n|if n1.\displaystyle v(z)=\sum_{n\in\mathbb{Z}}\widehat{v}(n)e_{n}(z)\quad\text{with}\quad e_{n}(z)=\left\{\begin{array}[]{cc}z^{n}&\text{if $n\geq 0$}\vspace{2mm}\\ \bar{z}^{|n|}&\text{if $n\leq-1$}\end{array}\right..

3.1. The derivation of Theorem 1.1 from Theorem 1.4

If (B2(𝔻,μ),h1(𝔻))(B^{2}(\mathbb{D},\mu),h^{1}(\mathbb{D})) is a holomorphically stable pair, then we have set-theoretical inclusion

H1(𝔻)A2(𝔻,μ).H^{1}(\mathbb{D})\subset A^{2}(\mathbb{D},\mu).

It follows that μ\mu is a finite measure, which, when combined with the assumption of the theorem, implies that μ\mu is a radial boundary-accessable finite measure on 𝔻\mathbb{D}. Therefore, A2(𝔻,μ)A^{2}(\mathbb{D},\mu) is complete. Hence the embedding H1(𝔻)A2(𝔻,μ)H^{1}(\mathbb{D})\subset A^{2}(\mathbb{D},\mu) is continuous by the Closed Graph Theorem. In other words, μ\mu is a (1,2)(1,2)-Carleson measure on 𝔻\mathbb{D}.

Now assume that μ\mu is a radial boundary-accessable (1,2)(1,2)-Carleson measure μ\mu on 𝔻\mathbb{D}. To prove the holomorphic stability of the pair (B2(𝔻,μ),h1(𝔻))(B^{2}(\mathbb{D},\mu),h^{1}(\mathbb{D})), it suffices to show that there exists a constant C>0C>0 such that

(3.7) fA2(𝔻,μ)CfB2(𝔻,μ)+h1(𝔻),f𝒪(𝔻).\displaystyle\|f\|_{A^{2}(\mathbb{D},\mu)}\leq C\|f\|_{B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})},\quad\forall f\in\mathcal{O}(\mathbb{D}).

Indeed, assuming (3.7) and let uB2(𝔻,μ),f𝒪(𝔻)u\in B^{2}(\mathbb{D},\mu),f\in\mathcal{O}(\mathbb{D}) with u+fh1(𝔻)u+f\in h^{1}(\mathbb{D}), we obtain

fA2(𝔻,μ)Cu(u+f)B2(𝔻,μ)+h1(𝔻)C(uB2(𝔻,μ)+u+fh1(𝔻))<\|f\|_{A^{2}(\mathbb{D},\mu)}\leq C\|u-(u+f)\|_{B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})}\leq C(\|u\|_{B^{2}(\mathbb{D},\mu)}+\|u+f\|_{h^{1}(\mathbb{D})})<\infty

and hence fA2(𝔻,μ)B2(𝔻,μ)f\in A^{2}(\mathbb{D},\mu)\subset B^{2}(\mathbb{D},\mu). It follows that u+fB2(𝔻,μ)u+f\in B^{2}(\mathbb{D},\mu) and this gives the holomorphic stability of the pair (B2(𝔻,μ),h1(𝔻))(B^{2}(\mathbb{D},\mu),h^{1}(\mathbb{D})).

It remains to prove (3.7). For any f𝒪(𝔻)f\in\mathcal{O}(\mathbb{D}) and any 0<r<10<r<1, write fr(z):=f(rz)f_{r}(z):=f(rz). Then, since μ\mu is radial,

limr1frA2(𝔻,μ)=fA2(𝔻,μ)\lim_{r\to 1^{-}}\|f_{r}\|_{A^{2}(\mathbb{D},\mu)}=\|f\|_{A^{2}(\mathbb{D},\mu)}

and

lim supr1frB2(𝔻,μ)+h1(𝔻)fB2(𝔻,μ)+h1(𝔻).\limsup_{r\to 1^{-}}\|f_{r}\|_{B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})}\leq\|f\|_{B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})}.

Therefore, it suffices to show that (3.7) holds for all ff belonging to the following class:

𝒪(𝔻¯)={f|f is holomorphic in a neighborhood of 𝔻¯}.\mathcal{O}(\overline{\mathbb{D}})=\{f|\text{$f$ is holomorphic in a neighborhood of $\overline{\mathbb{D}}$}\}.

We now proceed to the derivation of the inequality (3.7) for any f𝒪(𝔻¯)f\in\mathcal{O}(\overline{\mathbb{D}}) from the inequality (1.19) obtained in Theorem 1.4. Observe that, any vB2(𝔻,μ)v\in B^{2}(\mathbb{D},\mu) has the form

v(z)=nanen(z),v(z)=\sum_{n\in\mathbb{Z}}a_{n}e_{n}(z),

where ene_{n} is defined as in (3.6). Then, by the radial assumption on μ\mu,

(3.8) vB2(𝔻,μ)2=n|an|2𝔻|z|2|n|μ(dz).\displaystyle\|v\|_{B^{2}(\mathbb{D},\mu)}^{2}=\sum_{n\in\mathbb{Z}}|a_{n}|^{2}\int_{\mathbb{D}}|z|^{2|n|}\mu(dz).

Comparing (3.8) and (1.18), we obtain a natural isometric isomorphism Hμ(𝕋)B2(𝔻,μ)H_{\mu}(\mathbb{T})\rightarrow B^{2}(\mathbb{D},\mu) that associates any vHμ(𝕋)v\in H_{\mu}(\mathbb{T}) to its harmonic extension in 𝔻\mathbb{D} defined by (3.6). Similarly, there is a natural isometric isomorphism L1(𝕋)h1(𝔻)L^{1}(\mathbb{T})\rightarrow h^{1}(\mathbb{D}). Therefore, we get a natural identification of the Banach spaces

Hμ(𝕋)+L1(𝕋)B2(𝔻,μ)+h1(𝔻).H_{\mu}(\mathbb{T})+L^{1}(\mathbb{T})\simeq B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}).

Hence, for any f𝒪(𝔻¯)f\in\mathcal{O}(\overline{\mathbb{D}}), by taking u=f|𝕋C(𝕋)u=f|_{\mathbb{T}}\in C^{\infty}(\mathbb{T}) in (1.19), we obtain

𝒜μ1(f|𝕋)L2(𝕋)Cμ(fB2(𝔻,μ)+h1(𝔻)+fB2(𝔻,μ)+h1(𝔻)),\|\mathcal{A}_{\mu}^{-1}(f|_{\mathbb{T}})\|_{L^{2}(\mathbb{T})}\leq C_{\mu}(\|f\|_{B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})}+\|\mathcal{H}f\|_{B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})}),

with 𝒜μ\mathcal{A}_{\mu} defined in (1.17), f\mathcal{H}f defined in (1.15). Since f=ff(0)\mathcal{H}f=f-f(0) for all f𝒪(𝔻¯)f\in\mathcal{O}(\overline{\mathbb{D}}) and |f(0)|fh1(𝔻)fB2(𝔻)+h1(𝔻)|f(0)|\leq\|f\|_{h^{1}(\mathbb{D})}\leq\|f\|_{B^{2}(\mathbb{D})+h^{1}(\mathbb{D})}, we get

𝒜μ1(f|𝕋)L2(𝕋)3CμfB2(𝔻,μ)+h1(𝔻),f𝒪(𝔻¯).\|\mathcal{A}_{\mu}^{-1}(f|_{\mathbb{T}})\|_{L^{2}(\mathbb{T})}\leq 3C_{\mu}\|f\|_{B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})},\quad\forall f\in\mathcal{O}(\overline{\mathbb{D}}).

Finally, notice that, for any f=n0cnzn𝒪(𝔻¯)f=\sum_{n\geq 0}c_{n}z^{n}\in\mathcal{O}(\overline{\mathbb{D}}),

𝒜μ1(f|𝕋)L2(𝕋)2=n0|cn|2𝔻|z|2|n|μ(dz)=fA2(𝔻,μ)2.\displaystyle\|\mathcal{A}_{\mu}^{-1}(f|_{\mathbb{T}})\|_{L^{2}(\mathbb{T})}^{2}=\sum_{n\geq 0}|c_{n}|^{2}\int_{\mathbb{D}}|z|^{2|n|}\mu(dz)=\|f\|_{A^{2}(\mathbb{D},\mu)}^{2}.

Thus we obtain the desired inequality (3.7) for all f𝒪(𝔻¯)f\in\mathcal{O}(\overline{\mathbb{D}}) and complete the derivation of Theorem 1.1 from Theorem 1.4.

3.2. The proof of Theorem 1.4

A pair (a,b)(a,b) of sequences a=(a(n))na=(a(n))_{n\in\mathbb{Z}} and b=(b(n))nb=(b(n))_{n\in\mathbb{Z}} is called μ\mu-adapted if the following conditions are satisfied:

  • (i)

    a(0)0a(0)\neq 0;

  • (ii)

    for any n={0}n\in\mathbb{Z}^{*}=\mathbb{Z}\setminus\{0\},

    (3.9) |a(n)|2b(n)sgn(n)=σn1,where σn:=01r2|n|σ(dr)>0;\displaystyle|a(n)|^{2}b(n)\mathrm{sgn}(n)=\sigma_{n}^{-1},\quad\text{where\,\,}\sigma_{n}:=\int_{0}^{1}r^{2|n|}\sigma(dr)>0;
  • (iii)

    there exists a constant CbC_{b} such that

    (3.10) 0<1Cb|b(n)|Cb,n.\displaystyle 0<\frac{1}{C_{b}}\leq|b(n)|\leq C_{b},\quad n\in\mathbb{Z}^{*}.

Note that the condition (3.9) implies in particular that b(n)b(n)\in\mathbb{R} for all nn\in\mathbb{Z}^{*}.

Proposition 3.1.

Suppose μ\mu is a radial boundary-accessable (1,2)(1,2)-Carleson measure on 𝔻\mathbb{D}. Let (a,b)(a,b) be a μ\mu-adapted pair of sequences. Then there exists a constant CC such that

(3.11) uL2(𝕋)C(𝒯auHμ(𝕋)+L1(𝕋)+𝒯b𝒯auHμ(𝕋)+L1(𝕋)),uC(𝕋),\displaystyle\|u\|_{L^{2}(\mathbb{T})}\leq C(\|\mathcal{T}_{a}u\|_{H_{\mu}(\mathbb{T})+L^{1}(\mathbb{T})}+\|\mathcal{T}_{b}\mathcal{T}_{a}u\|_{H_{\mu}(\mathbb{T})+L^{1}(\mathbb{T})}),\quad\forall u\in C^{\infty}(\mathbb{T}),

where 𝒯a\mathcal{T}_{a} and 𝒯b\mathcal{T}_{b} are the Fourier multipliers defined by

𝒯au^(n)=a(n)u^(n) and 𝒯bu^(n)=b(n)u^(n),n.\widehat{\mathcal{T}_{a}u}(n)=a(n)\widehat{u}(n)\text{\, and \,}\widehat{\mathcal{T}_{b}u}(n)=b(n)\widehat{u}(n),\quad n\in\mathbb{Z}.

The next criterion of radial (1,2)(1,2)-Carleson measures will be useful for us.

Lemma 3.2.

Let α(dr)\alpha(dr) be a finite measure on [0,1)[0,1). Then the inequality

(3.12) supθ[0,2π)|01sinθ(rcosθ)2+sin2θα(dr)|<\displaystyle\sup_{\theta\in[0,2\pi)}\Big{|}\int_{0}^{1}\frac{\sin\theta}{(r-\cos\theta)^{2}+\sin^{2}\theta}\alpha(dr)\Big{|}<\infty

holds if and only if

sup0<δ<1α([1δ,1))δ<.\sup_{0<\delta<1}\frac{\alpha([1-\delta,1))}{\delta}<\infty.

We postpone the proof of Lemma 3.2 for a while and proceed to the proof of Theorem 1.4.

Lemma 3.3.

Let μ(dz)=σ(dr)dθ\mu(dz)=\sigma(dr)d\theta be a radial boundary-accessable (1,2)(1,2)-Carleson measure on 𝔻\mathbb{D}, then there exists a function wσL(𝕋)w_{\sigma}\in L^{\infty}(\mathbb{T}) such that

(3.13) w^σ(n)=sgn(n)σn,n.\displaystyle\widehat{w}_{\sigma}(n)=\mathrm{sgn}(n)\sigma_{n},\quad n\in\mathbb{Z}.
Proof.

Under the assumption of the lemma, set

(3.14) wσ(eiθ)=2i01r2sinθ|r2eiθ|2σ(dr).\displaystyle w_{\sigma}(e^{i\theta})=2i\int_{0}^{1}\frac{r^{2}\sin\theta}{|r^{2}-e^{-i\theta}|^{2}}\sigma(dr).

We first show that wσL(𝕋)w_{\sigma}\in L^{\infty}(\mathbb{T}). Indeed, by change-of-variable r=sr=\sqrt{s},

wσ(eiθ)=2i01sinθ(cosθs)2+sin2θσ(ds),w_{\sigma}(e^{i\theta})=2i\int_{0}^{1}\frac{\sin\theta}{(\cos\theta-s)^{2}+\sin^{2}\theta}\sigma^{\prime}(ds),

where σ(ds)=sσ(ds)\sigma^{\prime}(ds)=s\sigma_{*}(ds) with σ(ds)\sigma_{*}(ds) being the push-forward of the measure σ\sigma under the map s=r2s=r^{2}. By Lemma 2.1, there exists a constant C>0C>0 such that

σ([1δ,1))Cδ,δ(0,1).\sigma([1-\delta,1))\leq C\delta,\quad\forall\delta\in(0,1).

Then, by the definition of σ\sigma^{\prime}, there exists a constant C>0C^{\prime}>0 such that

σ([1δ,1))Cδ,δ(0,1).\sigma^{\prime}([1-\delta,1))\leq C^{\prime}\delta,\quad\forall\delta\in(0,1).

Hence, wσL(𝕋)w_{\sigma}\in L^{\infty}(\mathbb{T}) by Lemma 3.2.

It remains to prove the equality (3.13). Since wσL(𝕋)w_{\sigma}\in L^{\infty}(\mathbb{T}), we have

supθ[0,2π)01r2|sinθ||r2eiθ|2σ(dr)=supθ[0,2π)|01r2sinθ|r2eiθ|2σ(dr)|<.\sup_{\theta\in[0,2\pi)}\int_{0}^{1}\frac{r^{2}|\sin\theta|}{|r^{2}-e^{-i\theta}|^{2}}\sigma(dr)=\sup_{\theta\in[0,2\pi)}\Big{|}\int_{0}^{1}\frac{r^{2}\sin\theta}{|r^{2}-e^{-i\theta}|^{2}}\sigma(dr)\Big{|}<\infty.

Therefore, for any nn\in\mathbb{Z}, by Fubini’s Theorem,

w^σ(n)\displaystyle\widehat{w}_{\sigma}(n) =12π02π(2i01r2sinθ|r2eiθ|2σ(dr))einθ𝑑θ\displaystyle=\frac{1}{2\pi}\int_{0}^{2\pi}\Big{(}2i\int_{0}^{1}\frac{r^{2}\sin\theta}{|r^{2}-e^{-i\theta}|^{2}}\sigma(dr)\Big{)}e^{-in\theta}d\theta
=01(12π02π2ir2sinθ|r2eiθ|2einθ𝑑θ)σ(dr).\displaystyle=\int_{0}^{1}\Big{(}\frac{1}{2\pi}\int_{0}^{2\pi}2i\frac{r^{2}\sin\theta}{|r^{2}-e^{-i\theta}|^{2}}e^{-in\theta}d\theta\Big{)}\sigma(dr).

Then by the elementary identity (which converges absolutely for any fixed 0r<10\leq r<1)

2ir2sinθ|r2eiθ|2=nsgn(n)einθr2|n|,r[0,1),2i\frac{r^{2}\sin\theta}{|r^{2}-e^{-i\theta}|^{2}}=\sum_{n\in\mathbb{Z}^{*}}\text{sgn}(n)e^{in\theta}r^{2|n|},\quad\forall r\in[0,1),

we have

12π02π2ir2sinθ|r2eiθ|2einθ𝑑θ=sgn(n)r2|n|\frac{1}{2\pi}\int_{0}^{2\pi}2i\frac{r^{2}\sin\theta}{|r^{2}-e^{-i\theta}|^{2}}e^{-in\theta}d\theta=\mathrm{sgn}(n)r^{2|n|}

and hence

w^σ(n)=sgn(n)01r2|n|σ(dr)=sgn(n)σn.\widehat{w}_{\sigma}(n)=\mathrm{sgn}(n)\int_{0}^{1}r^{2|n|}\sigma(dr)=\mathrm{sgn}(n)\sigma_{n}.

This is the desired equality (3.13). ∎

Proof of Proposition 3.1.

Let uC(𝕋)u\in C^{\infty}(\mathbb{T}). Note that uL2(𝕋)uu^(0)L2(𝕋)+|u^(0)|.\|u\|_{L^{2}(\mathbb{T})}\leq\|u-\widehat{u}(0)\|_{L^{2}(\mathbb{T})}+|\widehat{u}(0)|. Clearly, if 𝒯au=f+g\mathcal{T}_{a}u=f+g with fHμ(𝕋),gL1(𝕋)f\in H_{\mu}(\mathbb{T}),g\in L^{1}(\mathbb{T}), then a(0)u^(0)=f^(0)+g^(0)a(0)\widehat{u}(0)=\widehat{f}(0)+\widehat{g}(0) and

|a(0)u^(0)||f^(0)|+|g^(0)|fHμ(𝕋)+gL1(𝕋).|a(0)\widehat{u}(0)|\leq|\widehat{f}(0)|+|\widehat{g}(0)|\leq\|f\|_{H_{\mu}(\mathbb{T})}+\|g\|_{L^{1}(\mathbb{T})}.

It follows that |u^(0)||a(0)|1𝒯auHμ(𝕋)+L1(𝕋)|\widehat{u}(0)|\leq|a(0)|^{-1}\|\mathcal{T}_{a}u\|_{H_{\mu}(\mathbb{T})+L^{1}(\mathbb{T})}. Therefore, from now on, we may assume that u^(0)=0\widehat{u}(0)=0. Take any pairs of decompositions

(3.17) {𝒯au=f1+g1,𝒯b𝒯au=f2+g2,\displaystyle\left\{\begin{array}[]{ll}\mathcal{T}_{a}u&=f_{1}+g_{1},\\ \mathcal{T}_{b}\mathcal{T}_{a}u&=f_{2}+g_{2},\end{array}\right.

with f1,f2Hμ(𝕋)f_{1},f_{2}\in H_{\mu}(\mathbb{T}) and g1,g2L1(𝕋)g_{1},g_{2}\in L^{1}(\mathbb{T}). Then for any 0<r<10<r<1, we have (the following Poisson convolutions will be used in the proof of the equality (3.29) below)

(3.20) {Pr𝔻(𝒯au)=Pr𝔻f1+Pr𝔻g1,Pr𝔻(𝒯b𝒯au)=Pr𝔻f2+Pr𝔻g2.\displaystyle\left\{\begin{array}[]{ll}P_{r}^{\mathbb{D}}*(\mathcal{T}_{a}u)&=P_{r}^{\mathbb{D}}*f_{1}+P_{r}^{\mathbb{D}}*g_{1},\\ P_{r}^{\mathbb{D}}*(\mathcal{T}_{b}\mathcal{T}_{a}u)&=P_{r}^{\mathbb{D}}*f_{2}+P_{r}^{\mathbb{D}}*g_{2}.\end{array}\right.

That is,

(3.23) {r|n|a(n)u^(n)=r|n|f^1(n)+r|n|g^1(n),n,r|n|b(n)a(n)u^(n)=r|n|f^2(n)+r|n|g^2(n),n.\displaystyle\left\{\begin{array}[]{ll}r^{|n|}a(n)\widehat{u}(n)&=r^{|n|}\widehat{f}_{1}(n)+r^{|n|}\widehat{g}_{1}(n),\quad n\in\mathbb{Z},\\ r^{|n|}b(n)a(n)\widehat{u}(n)&=r^{|n|}\widehat{f}_{2}(n)+r^{|n|}\widehat{g}_{2}(n),\quad n\in\mathbb{Z}.\end{array}\right.

From (3.23), we have

(3.24) Pr𝔻uL2(𝕋)2=nr2|n||u^(n)|2=nr|n|a(n)f^1(n)r|n|u^(n)¯denoted by I+nr|n|a(n)g^1(n)r|n|u^(n)¯denoted by II.\displaystyle\begin{split}\|P_{r}^{\mathbb{D}}*u\|_{L^{2}(\mathbb{T})}^{2}&=\sum_{n\in\mathbb{Z}^{*}}r^{2|n|}|\widehat{u}(n)|^{2}\\ &=\underbrace{\sum_{n\in\mathbb{Z}^{*}}\frac{r^{|n|}}{a(n)}\hat{f}_{1}(n)r^{|n|}\overline{\hat{u}(n)}}_{\text{denoted by I}}+\underbrace{\sum_{n\in\mathbb{Z}^{*}}\frac{r^{|n|}}{a(n)}\hat{g}_{1}(n)r^{|n|}\overline{\hat{u}(n)}}_{\text{denoted by II}}.\end{split}

By Cauchy-Schwarz’s inequality,

(3.25) |I|(nr2|n||u^(n)¯|2)1/2(nr2|n||f^1(n)|2|a(n)|2)1/2Cb/2πPr𝔻uL2(𝕋)Pr𝔻f1Hμ(𝕋),\displaystyle\begin{split}|\text{I}|&\leq\Big{(}\sum_{n\in\mathbb{Z}^{*}}r^{2|n|}|\overline{\hat{u}(n)}|^{2}\Big{)}^{1/2}\Big{(}\sum_{n\in\mathbb{Z}^{*}}\frac{r^{2|n|}|\widehat{f}_{1}(n)|^{2}}{|a(n)|^{2}}\Big{)}^{1/2}\\ &\leq\sqrt{C_{b}/2\pi}\|P_{r}^{\mathbb{D}}*u\|_{L^{2}(\mathbb{T})}\|P_{r}^{\mathbb{D}}*f_{1}\|_{H_{\mu}(\mathbb{T})},\end{split}

where we used the fact that if (a,b)(a,b) is a μ\mu-adapted pair of sequences, then by (3.9) and (3.10), for any vHμ(𝕋)v\in H_{\mu}(\mathbb{T}),

(3.26) (n|v^(n)|2|a(n)|2)1/2=(n|v^(n)|2|b(n)|σn)1/2Cb/2πvHμ(𝕋).\displaystyle\Big{(}\sum_{n\in\mathbb{Z}^{*}}\frac{|\widehat{v}(n)|^{2}}{|a(n)|^{2}}\Big{)}^{1/2}=\Big{(}\sum_{n\in\mathbb{Z}^{*}}|\widehat{v}(n)|^{2}|b(n)|\sigma_{n}\Big{)}^{1/2}\leq\sqrt{C_{b}/2\pi}\|v\|_{H_{\mu}(\mathbb{T})}.

By (3.23),

(3.27) II =nr|n|(a(n)u^(n)f^1(n))|a(n)|2b(n)r|n|f^2(n)¯denoted by III+nr|n|g^1(n)r|n|g^2(n)¯|a(n)|2b(n)denoted by IV.\displaystyle=\underbrace{\sum_{n\in\mathbb{Z}^{*}}\frac{r^{|n|}(a(n)\widehat{u}(n)-\widehat{f}_{1}(n))}{|a(n)|^{2}b(n)}r^{|n|}\overline{\widehat{f}_{2}(n)}}_{\text{denoted by III}}+\underbrace{\sum_{n\in\mathbb{Z}^{*}}\frac{r^{|n|}\widehat{g}_{1}(n)r^{|n|}\overline{\widehat{g}_{2}(n)}}{|a(n)|^{2}b(n)}}_{\text{denoted by IV}}.

Then by Cauchy-Schwarz’s inequality,

|III||nr|n|u^(n)a(n)¯b(n)r|n|f^2(n)¯|+|nr|n|f^1(n)|a(n)|2b(n)r|n|f^2(n)¯|\displaystyle|\text{III}|\leq\Big{|}\sum_{n\in\mathbb{Z}^{*}}\frac{r^{|n|}\widehat{u}(n)}{\overline{a(n)}b(n)}r^{|n|}\overline{\widehat{f}_{2}(n)}\Big{|}+\Big{|}\sum_{n\in\mathbb{Z}^{*}}\frac{r^{|n|}\widehat{f}_{1}(n)}{|a(n)|^{2}b(n)}r^{|n|}\overline{\widehat{f}_{2}(n)}\Big{|}
Pr𝔻uL2(𝕋)(n|r|n|f^2(n)a(n)b(n)|2)1/2+(nr2|n||f^1(n)|2|a(n)|2|b(n)|)1/2(nr2|n||f^2(n)|2|a(n)|2|b(n)|)1/2.\displaystyle\leq\|P_{r}^{\mathbb{D}}*u\|_{L^{2}(\mathbb{T})}\Big{(}\sum_{n\in\mathbb{Z}^{*}}\Big{|}\frac{r^{|n|}\widehat{f}_{2}(n)}{a(n)b(n)}\Big{|}^{2}\Big{)}^{1/2}+\Big{(}\sum_{n\in\mathbb{Z}^{*}}\frac{r^{2|n|}|\widehat{f}_{1}(n)|^{2}}{|a(n)|^{2}|b(n)|}\Big{)}^{1/2}\Big{(}\sum_{n\in\mathbb{Z}^{*}}\frac{r^{2|n|}|\widehat{f}_{2}(n)|^{2}}{|a(n)|^{2}|b(n)|}\Big{)}^{1/2}.

Using similar inequality as (3.26), under the conditions (3.9) and (3.10), we have

(3.28) |III|\displaystyle|\text{III}| Cb/2πPr𝔻uL2(𝕋)Pr𝔻f2Hμ(𝕋)+12πPr𝔻f1Hμ(𝕋)Pr𝔻f2Hμ(𝕋).\displaystyle\leq\sqrt{C_{b}/2\pi}\|P_{r}^{\mathbb{D}}*u\|_{L^{2}(\mathbb{T})}\|P_{r}^{\mathbb{D}}*f_{2}\|_{H_{\mu}(\mathbb{T})}+\frac{1}{2\pi}\|P_{r}^{\mathbb{D}}*f_{1}\|_{H_{\mu}(\mathbb{T})}\|P_{r}^{\mathbb{D}}*f_{2}\|_{H_{\mu}(\mathbb{T})}.

We now proceed to the estimate of term IV\mathrm{IV} in the decomposition (3.27). Note that

g^¯2(n)=g~¯^2(n), where g~2(eiθ):=g2(eiθ).\overline{\widehat{g}}_{2}(n)=\widehat{\overline{\widetilde{g}}}_{2}(n),\text{\, where \,}\tilde{g}_{2}(e^{i\theta}):=g_{2}(e^{-i\theta}).

For any 0<r<10<r<1, set

hr=(Pr𝔻g1)(Pr𝔻g~¯2).h_{r}=(P_{r}^{\mathbb{D}}*g_{1})*(P_{r}^{\mathbb{D}}*\overline{\widetilde{g}}_{2}).

A priori, we only have g1g~¯2L1(𝕋)g_{1}*\overline{\widetilde{g}}_{2}\in L^{1}(\mathbb{T}), but hrL2(𝕋)h_{r}\in L^{2}(\mathbb{T}) for any 0<r<10<r<1. By (3.9),

IV=nsgn(n)σnr|n|g^1(n)r|n|g^2(n)¯=nh^r(n)sgn(n)σn.\displaystyle\text{IV}=\sum_{n\in\mathbb{Z}^{*}}\mathrm{sgn}(n)\sigma_{n}r^{|n|}\widehat{g}_{1}(n)r^{|n|}\overline{\widehat{g}_{2}(n)}=\sum_{n\in\mathbb{Z}^{*}}\widehat{h}_{r}(n)\text{sgn}(n)\sigma_{n}.

By Lemma 3.3, wσL(𝕋)L2(𝕋)w_{\sigma}\in L^{\infty}(\mathbb{T})\subset L^{2}(\mathbb{T}). Then the Plancherel’s identity implies

(3.29) IV=nh^r(n)w^σ(n)=nh^r(n)w^(n)¯=𝕋hrw¯σ𝕋hr𝕋w¯σ.\displaystyle\text{IV}=\sum_{n\in\mathbb{Z}^{*}}\widehat{h}_{r}(n)\widehat{w}_{\sigma}(n)=\sum_{n\in\mathbb{Z}^{*}}\widehat{h}_{r}(n)\overline{\widehat{w}(n)}=\int_{\mathbb{T}}h_{r}\bar{w}_{\sigma}-\int_{\mathbb{T}}h_{r}\int_{\mathbb{T}}\bar{w}_{\sigma}.

Hence

(3.30) |IV|2hrL1(𝕋)wσL(𝕋).\displaystyle|\text{IV}|\leq 2\|h_{r}\|_{L^{1}(\mathbb{T})}\|w_{\sigma}\|_{L^{\infty}(\mathbb{T})}.

By (3.24), (3.25), (3.27), (3.28) and (3.30), there is a constant C=C(a,b,μ)C=C(a,b,\mu), depending only on (a,b)(a,b) and the measure μ\mu but not on r(0,1)r\in(0,1), such that

Pr𝔻uL2(𝕋)2\displaystyle\|P_{r}^{\mathbb{D}}*u\|_{L^{2}(\mathbb{T})}^{2}\leq CPr𝔻uL2(𝕋)Pr𝔻f1Hμ(𝕋)+CPr𝔻uL2(𝕋)Pr𝔻f2Hμ(𝕋)+\displaystyle C\|P_{r}^{\mathbb{D}}*u\|_{L^{2}(\mathbb{T})}\|P_{r}^{\mathbb{D}}*f_{1}\|_{H_{\mu}(\mathbb{T})}+C\|P_{r}^{\mathbb{D}}*u\|_{L^{2}(\mathbb{T})}\|P_{r}^{\mathbb{D}}*f_{2}\|_{H_{\mu}(\mathbb{T})}+
+CPr𝔻f1Hμ(𝕋)Pr𝔻f2Hμ(𝕋)+CPr𝔻g1L1(𝕋)Pr𝔻g2L1(𝕋).\displaystyle+C\|P_{r}^{\mathbb{D}}*f_{1}\|_{H_{\mu}(\mathbb{T})}\|P_{r}^{\mathbb{D}}*f_{2}\|_{H_{\mu}(\mathbb{T})}+C\|P_{r}^{\mathbb{D}}*g_{1}\|_{L^{1}(\mathbb{T})}\|P_{r}^{\mathbb{D}}*g_{2}\|_{L^{1}(\mathbb{T})}.

Therefore, by a standard argument, there is a constant C=C(a,b,μ)C^{\prime}=C^{\prime}(a,b,\mu) such that

Pr𝔻uL2(𝕋)\displaystyle\|P_{r}^{\mathbb{D}}*u\|_{L^{2}(\mathbb{T})} C(Pr𝔻f1Hμ(𝕋)+Pr𝔻f2Hμ(𝕋)+Pr𝔻g1L1(𝕋)+Pr𝔻g2L1(𝕋))\displaystyle\leq C^{\prime}\Big{(}\|P_{r}^{\mathbb{D}}*f_{1}\|_{H_{\mu}(\mathbb{T})}+\|P_{r}^{\mathbb{D}}*f_{2}\|_{H_{\mu}(\mathbb{T})}+\|P_{r}^{\mathbb{D}}*g_{1}\|_{L^{1}(\mathbb{T})}+\|P_{r}^{\mathbb{D}}*g_{2}\|_{L^{1}(\mathbb{T})}\Big{)}
C(f1Hμ(𝕋)+g1L1(𝕋)+f2Hμ(𝕋)+g2L1(𝕋)),\displaystyle\leq C^{\prime}\Big{(}\|f_{1}\|_{H_{\mu}(\mathbb{T})}+\|g_{1}\|_{L^{1}(\mathbb{T})}+\|f_{2}\|_{H_{\mu}(\mathbb{T})}+\|g_{2}\|_{L^{1}(\mathbb{T})}\Big{)},

where the last inequality is due to the contractive property of the Poission convolution on both Hμ(𝕋)H_{\mu}(\mathbb{T}) and L1(𝕋)L^{1}(\mathbb{T}). Let rr approach to 11, then

uL2(𝕋)C(f1Hμ(𝕋)+g1L1(𝕋)+f2Hμ(𝕋)+g2L1(𝕋)).\|u\|_{L^{2}(\mathbb{T})}\leq C^{\prime}\Big{(}\|f_{1}\|_{H_{\mu}(\mathbb{T})}+\|g_{1}\|_{L^{1}(\mathbb{T})}+\|f_{2}\|_{H_{\mu}(\mathbb{T})}+\|g_{2}\|_{L^{1}(\mathbb{T})}\Big{)}.

Since the decompositions (3.17) are arbitrary, we obtain the desired inequality (3.11). ∎

Proof of Theorem 1.4.

If μ=σ(dr)dθ\mu=\sigma(dr)d\theta is a radial boundary-accessable (1,2)(1,2)-Carleson measure on 𝔻\mathbb{D}, then we obtain the inequality (1.19) from Proposition 3.1 by taking

a(n)=(𝔻|z|2|n|μ(dz))1/2=12πσn and b(n)=sgn(n).a(n)=\Big{(}\int_{\mathbb{D}}|z|^{2|n|}\mu(dz)\Big{)}^{-1/2}=\frac{1}{\sqrt{2\pi\sigma_{n}}}\text{\, and \,}b(n)=\mathrm{sgn}(n).

Conversely, if the inequality (1.19) holds, then by the argument in the first two paragraphs of §3.1, the measure μ\mu is a (1,2)(1,2)-Carleson measure on 𝔻\mathbb{D}. ∎

It remains to prove Lemma 3.2. We shall apply a result due to Garnett about the boundary behavior of Poisson integrals on the upper half plane \mathbb{H}.

Lemma 3.4 (Garnett, see, e.g., [RU88, pp. 210]).

Let ν\nu be a measure on \mathbb{R} with 11+t2ν(dt)<.\int_{\mathbb{R}}\frac{1}{1+t^{2}}\nu(dt)<\infty. Then the following two assertions are equivalent:

  • supy>0yt2+y2ν(dt)<\sup_{y>0}\int_{\mathbb{R}}\frac{y}{t^{2}+y^{2}}\nu(dt)<\infty

  • supL>0ν([L,L])2L<.\sup_{L>0}\frac{\nu([-L,L])}{2L}<\infty.

Proof of Lemma 3.2.

Note that

supθ[0,2π)|01sinθ(rcosθ)2+sin2θα(dr)|=supθ(0,π)01sinθ(rcosθ)2+sin2θα(dr).\displaystyle\sup_{\theta\in[0,2\pi)}\Big{|}\int_{0}^{1}\frac{\sin\theta}{(r-\cos\theta)^{2}+\sin^{2}\theta}\alpha(dr)\Big{|}=\sup_{\theta\in(0,\pi)}\int_{0}^{1}\frac{\sin\theta}{(r-\cos\theta)^{2}+\sin^{2}\theta}\alpha(dr).

For any θ(0,π)\theta\in(0,\pi), consider the point z=eiθ=cosθ+isinθz=e^{i\theta}=\cos\theta+i\sin\theta and recall the Poisson kernel PzP_{z}^{\mathbb{H}} at the point zz\in\mathbb{H} given in (1.9), then

01sinθ(rcosθ)2+sin2θσ(dr)=πPeiθ(t)1[0,1)(t)σ(dt).\int_{0}^{1}\frac{\sin\theta}{(r-\cos\theta)^{2}+\sin^{2}\theta}\sigma(dr)=\pi\int_{\mathbb{R}}P^{\mathbb{H}}_{e^{i\theta}}(t)1_{[0,1)}(t)\sigma(dt).

Consider the Möbius transformation ϕ\phi defined by ϕ(z)=z1z+1.\phi(z)=\frac{z-1}{z+1}. Then ϕ\phi is an automorphism of the upper half plane and

Pϕ(z)(ϕ(t))|ϕ(t)|=Pz(t),z,t{1}.P^{\mathbb{H}}_{\phi(z)}(\phi(t))|\phi^{\prime}(t)|=P^{\mathbb{H}}_{z}(t),\quad z\in\mathbb{H},\,t\in\mathbb{R}\setminus\{-1\}.

Note that when θ\theta ranges over (0,π)(0,\pi), the image ϕ(eiθ)\phi(e^{i\theta}) ranges over i+i\mathbb{R}_{+}. Therefore,

supθ(0,π)Peiθ(t)1[0,1)(t)σ(dt)\displaystyle\sup_{\theta\in(0,\pi)}\int_{\mathbb{R}}P^{\mathbb{H}}_{e^{i\theta}}(t)1_{[0,1)}(t)\sigma(dt) =supθ(0,π)Pϕ(eiθ)(ϕ(t))|ϕ(t)|1[0,1)(t)σ(dt)\displaystyle=\sup_{\theta\in(0,\pi)}\int_{\mathbb{R}}P^{\mathbb{H}}_{\phi(e^{i\theta})}(\phi(t))|\phi^{\prime}(t)|1_{[0,1)}(t)\sigma(dt)
=supy>001Piy(ϕ(t))|ϕ(t)|σ(dt).\displaystyle=\sup_{y>0}\int_{0}^{1}P_{iy}^{\mathbb{H}}(\phi(t))|\phi^{\prime}(t)|\sigma(dt).

By change-of-variable s=ϕ(t)s=\phi(t),

01Piy(ϕ(t))|ϕ(t)|σ(dt)=10Piy(s)(1s)22σϕ1(ds).\int_{0}^{1}P_{iy}^{\mathbb{H}}(\phi(t))|\phi^{\prime}(t)|\sigma(dt)=\int_{-1}^{0}P_{iy}^{\mathbb{H}}(s)\frac{(1-s)^{2}}{2}\sigma\circ\phi^{-1}(ds).

Then

supθ(0,π)01sinθ(rcosθ)2+sin2θσ(dr)\displaystyle\sup_{\theta\in(0,\pi)}\int_{0}^{1}\frac{\sin\theta}{(r-\cos\theta)^{2}+\sin^{2}\theta}\sigma(dr) =π2supy>0yy2+s2σ~(ds),\displaystyle=\frac{\pi}{2}\sup_{y>0}\int_{\mathbb{R}}\frac{y}{y^{2}+s^{2}}\widetilde{\sigma}(ds),

where σ~(ds)=(1s)2𝟙(1,0)(s)σϕ1(ds)\widetilde{\sigma}(ds)=(1-s)^{2}\mathds{1}_{(-1,0)}(s)\sigma\circ\phi^{-1}(ds). Clearly, σ~(ds)1+s2<\int_{\mathbb{R}}\frac{\widetilde{\sigma}(ds)}{1+s^{2}}<\infty. Therefore, by Lemma 3.4, the inequality (3.12) holds if and only if

(3.31) supL>0σ~([L,L])L<.\displaystyle\sup_{L>0}\frac{\widetilde{\sigma}([-L,L])}{L}<\infty.

By the definition of σ~\widetilde{\sigma}, it is easy to see that 12σ(IL)σ~([L,L])2σ(IL)\frac{1}{2}\sigma(I_{L})\leq\widetilde{\sigma}([-L,L])\leq 2\sigma(I_{L}), where ILI_{L} is the open interval

IL:=(1min(L,1)1+min(L,1),1)(0,1).I_{L}:=\Big{(}\frac{1-\min(L,1)}{1+\min(L,1)},1\Big{)}\subset(0,1).

It follows that, (3.31) holds if and only if supL>0σ(IL)L<\sup_{L>0}\frac{\sigma(I_{L})}{L}<\infty, which in turn is equivalent to

sup0<δ<1σ([1δ,1))δ<.\sup_{0<\delta<1}\frac{\sigma([1-\delta,1))}{\delta}<\infty.

By Lemma 2.1, the above inequality holds if and only if μ(dz)=σ(dr)dθ\mu(dz)=\sigma(dr)d\theta is a (1,2)(1,2)-Carleson measure on 𝔻\mathbb{D}. This completes the whole proof. ∎

3.3. Proof of Theorem 1.2

Fix a radial boundary-accessable (1,2)(1,2)-Carleson measure μ(dz)=σ(dr)dθ.\mu(dz)=\sigma(dr)d\theta. By Theorem 1.1,

(3.32) (B2(𝔻,μ)+h1(𝔻))𝒪(𝔻)=B2(𝔻,μ)𝒪(𝔻)=A2(𝔻,μ).\displaystyle(B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}))\cap\mathcal{O}(\mathbb{D})=B^{2}(\mathbb{D},\mu)\cap\mathcal{O}(\mathbb{D})=A^{2}(\mathbb{D},\mu).

By (3.7), there exists a constant C=Cμ>0C=C_{\mu}>0, such that for all f𝒪(𝔻)f\in\mathcal{O}(\mathbb{D}),

(3.33) 1CfA2(𝔻,μ)fB2(𝔻,μ)+h1(𝔻)fB2(𝔻,μ)=fA2(𝔻,μ).\displaystyle\frac{1}{C}\|f\|_{A^{2}(\mathbb{D},\mu)}\leq\|f\|_{B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})}\leq\|f\|_{B^{2}(\mathbb{D},\mu)}=\|f\|_{A^{2}(\mathbb{D},\mu)}.

That is, the identity map

id:A2(𝔻,μ)(B2(𝔻,μ)+h1(𝔻))𝒪(𝔻)B2(𝔻,μ)+h1(𝔻)id:A^{2}(\mathbb{D},\mu)\rightarrow(B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}))\cap\mathcal{O}(\mathbb{D})\subset B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})

is an isomorphic isomorphism. Thus (B2(𝔻,μ)+h1(𝔻))𝒪(𝔻)(B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}))\cap\mathcal{O}(\mathbb{D}) is a closed subspace of B2(𝔻,μ)+h1(𝔻)B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}).

Now suppose that the extra condition (1.6) is satisfied. We are going to show that the closed subspace (B2(𝔻,μ)+h1(𝔻))𝒪(𝔻)(B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}))\cap\mathcal{O}(\mathbb{D}) is complemented in B2(𝔻,μ)+h1(𝔻)B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}). Indeed, under the condition (1.6), for any θ[0,2π)\theta\in[0,2\pi), the following holomorphic function

kθ(z):=11eiθz=n0einθzn,z𝔻k_{\theta}(z):=\frac{1}{1-e^{-i\theta}z}=\sum_{n\geq 0}e^{-in\theta}z^{n},\quad z\in\mathbb{D}

belongs to B2(𝔻,μ)+h1(𝔻)B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}) and

(3.34) Mμ:=supθ[0,2π)kθB2(𝔻,μ)+h1(𝔻)supθ[0,2π)kθA2(𝔻,μ)=(𝔻μ(dz)1|z|2)1/2<.\displaystyle M_{\mu}:=\sup_{\theta\in[0,2\pi)}\|k_{\theta}\|_{B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})}\leq\sup_{\theta\in[0,2\pi)}\|k_{\theta}\|_{A^{2}(\mathbb{D},\mu)}=\Big{(}\int_{\mathbb{D}}\frac{\mu(dz)}{1-|z|^{2}}\Big{)}^{1/2}<\infty.

Recall the definition (1.2) of 𝒬+\mathcal{Q}_{+}. Clearly, since μ\mu is radial, 𝒬+\mathcal{Q}_{+} defines an orthgonal projection from B2(𝔻,μ)B^{2}(\mathbb{D},\mu) onto A2(𝔻,μ)A^{2}(\mathbb{D},\mu). Then

(3.35) 𝒬+(u)B2(𝔻,μ)+h1(𝔻)𝒬+(u)B2(𝔻,μ)uB2(𝔻,μ),uB2(𝔻,μ).\displaystyle\|\mathcal{Q}_{+}(u)\|_{B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})}\leq\|\mathcal{Q}_{+}(u)\|_{B^{2}(\mathbb{D},\mu)}\leq\|u\|_{B^{2}(\mathbb{D},\mu)},\quad\forall u\in B^{2}(\mathbb{D},\mu).

Note also that if v=nanenh1(𝔻)v=\sum_{n\in\mathbb{Z}}a_{n}e_{n}\in h^{1}(\mathbb{D}), that is,

v~:=naneinθL1(𝕋) and vh1(𝔻)=v~L1(𝕋),\widetilde{v}:=\sum_{n\in\mathbb{Z}}a_{n}e^{in\theta}\in L^{1}(\mathbb{T})\text{\, and \,}\|v\|_{h^{1}(\mathbb{D})}=\|\widetilde{v}\|_{L^{1}(\mathbb{T})},

then it is easy to see that

𝒬+(v)=𝒬+(nanen)=12π02πkθv~(eiθ)𝑑θ.\mathcal{Q}_{+}(v)=\mathcal{Q}_{+}\Big{(}\sum_{n\in\mathbb{Z}}a_{n}e_{n}\Big{)}=\frac{1}{2\pi}\int_{0}^{2\pi}k_{\theta}\widetilde{v}(e^{i\theta})d\theta.

Hence, by (3.34),

(3.36) 𝒬+(v)B2(𝔻,μ)+h1(𝔻)12π02πkθB2(𝔻,μ)+h1(𝔻)|v~(eiθ)|𝑑θMμv~L1(𝕋)=Mμvh1(𝔻).\displaystyle\begin{split}\|\mathcal{Q}_{+}(v)\|_{B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})}&\leq\frac{1}{2\pi}\int_{0}^{2\pi}\|k_{\theta}\|_{B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})}|\widetilde{v}(e^{i\theta})|d\theta\\ &\leq M_{\mu}\|\widetilde{v}\|_{L^{1}(\mathbb{T})}=M_{\mu}\|v\|_{h^{1}(\mathbb{D})}.\end{split}

By (3.35) and (3.36) and the definition of the norm on B2(𝔻,μ)+h1(𝔻)B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}),

𝒬+(f)B2(𝔻,μ)+h1(𝔻)MμfB2(𝔻,μ)+h1(𝔻),fB2(𝔻,μ)+h1(𝔻).\|\mathcal{Q}_{+}(f)\|_{B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})}\leq M_{\mu}\|f\|_{B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})},\quad\forall f\in B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}).

It follows that 𝒬+\mathcal{Q}_{+} defines a bounded linear projection from B2(𝔻,μ)+h1(𝔻)B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}) onto

(B2(𝔻,μ)+h1(𝔻))𝒪(𝔻).(B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}))\cap\mathcal{O}(\mathbb{D}).

Hence (B2(𝔻,μ)+h1(𝔻))𝒪(𝔻)(B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}))\cap\mathcal{O}(\mathbb{D}) is complemented in B2(𝔻,μ)+h1(𝔻)B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}).

Finally, assume that the condition (1.6) is not satisfied. Then

n0σn=n001r2nσ(dr)=01σ(dr)1r2=12π𝔻μ(dz)1|z|2=.\sum_{n\geq 0}\sigma_{n}=\sum_{n\geq 0}\int_{0}^{1}r^{2n}\sigma(dr)=\int_{0}^{1}\frac{\sigma(dr)}{1-r^{2}}=\frac{1}{2\pi}\int_{\mathbb{D}}\frac{\mu(dz)}{1-|z|^{2}}=\infty.

Let us show that (B2(𝔻,μ)+h1(𝔻))𝒪(𝔻)(B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}))\cap\mathcal{O}(\mathbb{D}) is not complemented in B2(𝔻,μ)+h1(𝔻)B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}). Otherwise, there exists a bounded linear projection operator

P:B2(𝔻,μ)+h1(𝔻)B2(𝔻,μ)+h1(𝔻)P:B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})\longrightarrow B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})

onto the closed subspace (B2(𝔻,μ)+h1(𝔻))𝒪(𝔻)(B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}))\cap\mathcal{O}(\mathbb{D}). That is,

  • PP=PP\circ P=P,

  • P(f)=fP(f)=f for all f(B2(𝔻,μ)+h1(𝔻))𝒪(𝔻)f\in(B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}))\cap\mathcal{O}(\mathbb{D}),

  • P(g)(B2(𝔻,μ)+h1(𝔻))𝒪(𝔻)P(g)\in(B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}))\cap\mathcal{O}(\mathbb{D}) for all gB2(𝔻,μ)+h1(𝔻)g\in B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}).

Since μ\mu is radial, for any θ[0,2π)\theta\in[0,2\pi), the rotation map τθ\tau_{\theta} defined by τθ(f)(z)=f(eiθz)\tau_{\theta}(f)(z)=f(e^{i\theta}z) preserves both the norms of functions in B2(𝔻,μ)B^{2}(\mathbb{D},\mu) and the norms of functions in h1(𝔻)h^{1}(\mathbb{D}). Therefore, τθ\tau_{\theta} preserves the norms of functions in B2(𝔻,μ)+h1(𝔻)B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}):

τθ(f)B2(𝔻,μ)+h1(𝔻)=fB2(𝔻,μ)+h1(𝔻),fB2(𝔻,μ)+h1(𝔻).\|\tau_{\theta}(f)\|_{B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})}=\|f\|_{B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})},\quad\forall f\in B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}).

Consequently, the operator-norm of the composition operator

Pθ=τθPτθ:B2(𝔻,μ)+h1(𝔻)B2(𝔻,μ)+h1(𝔻)P_{\theta}=\tau_{-\theta}\circ P\circ\tau_{\theta}:B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})\longrightarrow B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})

is bounded by that of PP:

PθP.\|P_{\theta}\|\leq\|P\|.

It can be easily checked that PθP_{\theta} is also a projection operator from B2(𝔻,μ)+h1(𝔻)B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}) onto (B2(𝔻,μ)+h1(𝔻))𝒪(𝔻)(B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}))\cap\mathcal{O}(\mathbb{D}).

Define a bounded linear operator 𝒫:B2(𝔻,μ)+h1(𝔻)B2(𝔻,μ)+h1(𝔻)\mathcal{P}:B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})\longrightarrow B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}) via the Bochner integral (see, e.g., [Yos95, Section V.5])

𝒫:=12π02πPθ𝑑θ.\mathcal{P}:=\frac{1}{2\pi}\int_{0}^{2\pi}P_{\theta}d\theta.

Then

(3.37) 𝒫supθ[0,2π)Pθ=P<\displaystyle\|\mathcal{P}\|\leq\sup_{\theta\in[0,2\pi)}\|P_{\theta}\|=\|P\|<\infty

and

(3.38) 𝒫(f)=12π02πPθ(f)𝑑θ,fB2(𝔻,μ)+h1(𝔻).\displaystyle\mathcal{P}(f)=\frac{1}{2\pi}\int_{0}^{2\pi}P_{\theta}(f)d\theta,\quad\forall f\in B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}).

Since the evalutation map evz\mathrm{ev}_{z} defined in (3.1) is a continuous linear functional on B2(𝔻,μ)+h1(𝔻)B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}) for any z𝔻z\in\mathbb{D},

(3.39) [𝒫(f)](z)=12π02π[Pθ(f)](z)𝑑θ,fB2(𝔻,μ)+h1(𝔻).\displaystyle[\mathcal{P}(f)](z)=\frac{1}{2\pi}\int_{0}^{2\pi}[P_{\theta}(f)](z)d\theta,\quad\forall f\in B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}).

Note that Pθ(f)=fP_{\theta}(f)=f for any f(B2(𝔻,μ)+h1(𝔻))𝒪(𝔻)f\in(B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}))\cap\mathcal{O}(\mathbb{D}) and any θ[0,2π)\theta\in[0,2\pi), thus

𝒫(f)=f,f(B2(𝔻,μ)+h1(𝔻))𝒪(𝔻).\mathcal{P}(f)=f,\quad\forall f\in(B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}))\cap\mathcal{O}(\mathbb{D}).

On the other hand, for any integer n1n\geq 1,

(3.40) 𝒫(en)=0,n1.\displaystyle\mathcal{P}(e_{-n})=0,\quad\forall n\geq 1.

Indeed, for any θ[0,2π)\theta\in[0,2\pi),

(τθ(en))(z)=(eiθz)¯n=einθz¯n=einθen(z).(\tau_{\theta}(e_{-n}))(z)=\overline{(e^{i\theta}z)}^{n}=e^{-in\theta}\bar{z}^{n}=e^{-in\theta}e_{-n}(z).

Thus Pτθ(en)=einθP(en)P\circ\tau_{\theta}(e_{-n})=e^{-in\theta}P(e_{-n}). By (3.32),

P(en)(B2(𝔻,μ)+h1(𝔻))𝒪(𝔻)=A2(𝔻,μ),P(e_{-n})\in(B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}))\cap\mathcal{O}(\mathbb{D})=A^{2}(\mathbb{D},\mu),

we can write

P(en)(z)=k=0ck(n)zkA2(𝔻,μ),P(e_{-n})(z)=\sum_{k=0}^{\infty}c^{(n)}_{k}z^{k}\in A^{2}(\mathbb{D},\mu),

with

(3.41) P(en)A2(𝔻,μ)2=2πk=0|ck(n)|2σk<.\displaystyle\|P(e_{-n})\|_{A^{2}(\mathbb{D},\mu)}^{2}=2\pi\sum_{k=0}^{\infty}|c_{k}^{(n)}|^{2}\sigma_{k}<\infty.

Thus, for all z𝔻z\in\mathbb{D},

Pθ(en)(z)\displaystyle P_{\theta}(e_{-n})(z) =[τθ(einθP(en))](z)=einθ[τθ(P(en))](z)\displaystyle=[\tau_{-\theta}(e^{-in\theta}P(e_{-n}))](z)=e^{-in\theta}[\tau_{-\theta}(P(e_{-n}))](z)
=einθk=0ck(n)(eiθz)k=k=0ck(n)ei(k+n)θzk,\displaystyle=e^{-in\theta}\sum_{k=0}^{\infty}c_{k}^{(n)}(e^{-i\theta}z)^{k}=\sum_{k=0}^{\infty}c_{k}^{(n)}e^{-i(k+n)\theta}z^{k},

where the last series converges absolutely by the inequalities (3.2), (3.41) and

(k=0|ck(n)zk|)2k=0|ck(n)|2σkk=0|z|2kσkk=0|ck(n)|2σkk=0|z|2kρ2kσ([ρ,1)),ρ(0,1).\Big{(}\sum_{k=0}^{\infty}|c_{k}^{(n)}z^{k}|\Big{)}^{2}\leq\sum_{k=0}^{\infty}|c_{k}^{(n)}|^{2}\sigma_{k}\sum_{k=0}^{\infty}\frac{|z|^{2k}}{\sigma_{k}}\leq\sum_{k=0}^{\infty}|c_{k}^{(n)}|^{2}\sigma_{k}\sum_{k=0}^{\infty}\frac{|z|^{2k}}{\rho^{2k}\sigma([\rho,1))},\quad\forall\rho\in(0,1).

Therefore, by (3.39), for all z𝔻z\in\mathbb{D},

[𝒫(en)](z)\displaystyle[\mathcal{P}(e_{-n})](z) =02πk=0ck(n)ei(k+n)θzkdθ2π=k=002πck(n)ei(k+n)θzkdθ2π=0.\displaystyle=\int_{0}^{2\pi}\sum_{k=0}^{\infty}c_{k}^{(n)}e^{-i(k+n)\theta}z^{k}\frac{d\theta}{2\pi}=\sum_{k=0}^{\infty}\int_{0}^{2\pi}c_{k}^{(n)}e^{-i(k+n)\theta}z^{k}\frac{d\theta}{2\pi}=0.

This is the desired equality (3.40).

However, if we take the harmonic extension of the Féjer kernel on 𝔻\mathbb{D}:

N(z)=j=NN(1|j|N)ej(z),N1,\mathcal{F}_{N}(z)=\sum_{j=-N}^{N}\Big{(}1-\frac{|j|}{N}\Big{)}e_{j}(z),\quad N\geq 1,

then

(3.42) NB2(𝔻,μ)+h1(𝔻)Nh1(𝔻)=N|𝕋L1(𝕋)=1.\displaystyle\|\mathcal{F}_{N}\|_{B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})}\leq\|\mathcal{F}_{N}\|_{h^{1}(\mathbb{D})}=\|\mathcal{F}_{N}|_{\mathbb{T}}\|_{L^{1}(\mathbb{T})}=1.

Since 𝒫\mathcal{P} is a projection onto (B2(𝔻,μ)+h1(𝔻))𝒪(𝔻)(B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D}))\cap\mathcal{O}(\mathbb{D}) and satisfies (3.40), we have

𝒫(N)=j=0N(1jN)ej\mathcal{P}(\mathcal{F}_{N})=\sum_{j=0}^{N}\Big{(}1-\frac{j}{N}\Big{)}e_{j}

and, by the radial assumption on μ\mu,

𝒫(N)A2(𝔻,μ)2=j=0N(1j/N)2ejA2(𝔻,μ)2=2πj=0N(1j/N)2σj.\|\mathcal{P}(\mathcal{F}_{N})\|_{A^{2}(\mathbb{D},\mu)}^{2}=\sum_{j=0}^{N}(1-j/N)^{2}\|e_{j}\|_{A^{2}(\mathbb{D},\mu)}^{2}=2\pi\sum_{j=0}^{N}(1-j/N)^{2}\sigma_{j}.

Then, by (3.33),

lim infN𝒫(N)B2(𝔻,μ)+h1(𝔻)2lim infN𝒫(N)A2(𝔻,μ)2C22πC2j=0σj=.\liminf_{N\to\infty}\|\mathcal{P}(\mathcal{F}_{N})\|_{B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})}^{2}\geq\liminf_{N\to\infty}\frac{\|\mathcal{P}(\mathcal{F}_{N})\|_{A^{2}(\mathbb{D},\mu)}^{2}}{C^{2}}\geq\frac{2\pi}{C^{2}}\sum_{j=0}^{\infty}\sigma_{j}=\infty.

This contradicts to the following inequality (which is a consequence of (3.37) and (3.42))

supN1𝒫(N)B2(𝔻,μ)+h1(𝔻)P<.\sup_{N\geq 1}\|\mathcal{P}(\mathcal{F}_{N})\|_{B^{2}(\mathbb{D},\mu)+h^{1}(\mathbb{D})}\leq\|P\|<\infty.

Hence we complete the whole proof of the theorem.

4. Holomorphic stability: the upper half plane case

In this section, we will prove Theorem 1.3. For any Radon measure Π\Pi on +=(0,)\mathbb{R}_{+}=(0,\infty) satisfying the condition (2.2), define

Π(ξ):={+e4πy|ξ|Π(dy)if ξ,0if ξ=0.\mathcal{L}_{\Pi}(\xi):=\left\{\begin{array}[]{cl}\int_{\mathbb{R}^{+}}e^{-4\pi y|\xi|}\Pi(dy)&\text{if $\xi\in\mathbb{R}^{*}$},\\ 0&\text{if $\xi=0$}.\end{array}\right.

Recall the following definition of the Fourier transform for fL1()f\in L^{1}(\mathbb{R}):

f^(ξ):=f(x)ei2πxξ𝑑x,ξ.\widehat{f}(\xi):=\int_{\mathbb{R}}f(x)e^{-i2\pi x\xi}dx,\quad\xi\in\mathbb{R}.

Recall the definition (1.7) of 2(,μ)\mathscr{B}^{2}(\mathbb{H},\mu). For any g2(,μ)g\in\mathscr{B}^{2}(\mathbb{H},\mu) and y>0y>0, recall the definition of the function gyg_{y} defined in (1.8). Note that the Fourier transform of the Poisson kernel PiyP_{iy}^{\mathbb{H}} given in (1.9) has the following form (see [Kat04, Chapter VI, p. 140]):

Piy^(ξ)=e2πy|ξ|,ξ.\widehat{P^{\mathbb{H}}_{iy}}(\xi)=e^{-2\pi y|\xi|},\quad\xi\in\mathbb{R}.

Then, by (1.10) and (1.11), we have g^y(ξ)=e2π(yy)|ξ|g^y(ξ)\widehat{g}_{y}(\xi)=e^{-2\pi(y-y^{\prime})|\xi|}\widehat{g}_{y^{\prime}}(\xi) for all 0<y<y0<y^{\prime}<y and hence

(4.1) e2πy|ξ|g^y(ξ)=e2πy|ξ|g^y(ξ), 0<y<y.\displaystyle e^{2\pi y|\xi|}\widehat{g}_{y}(\xi)=e^{2\pi y^{\prime}|\xi|}\widehat{g}_{y^{\prime}}(\xi),\quad\forall\,0<y^{\prime}<y.
Definition.

Let μ(dz)=dxΠ(dy)\mu(dz)=dx\Pi(dy) be a boundary-accessable (1,2)(1,2)-Carleson measure on \mathbb{H}. For any g2(,μ)g\in\mathscr{B}^{2}(\mathbb{H},\mu), define a function g^0\widehat{g}_{0} by

(4.2) g^0(ξ):=e2πy|ξ|g^y(ξ),y>0,\displaystyle\widehat{g}_{0}(\xi):=e^{2\pi y|\xi|}\widehat{g}_{y}(\xi),\,y>0,

where, by (4.1), the right hand side of the equality (4.2) is independent of y>0y>0.

By (4.2) and the Plancherel’s identity, the norm of any g2(,μ)g\in\mathscr{B}^{2}(\mathbb{H},\mu) has the form:

(4.3) g2(,μ)=(|g^0(ξ)|2Π(ξ)𝑑ξ)1/2.\displaystyle\|g\|_{\mathscr{B}^{2}(\mathbb{H},\mu)}=\Big{(}\int_{\mathbb{R}}|\widehat{g}_{0}(\xi)|^{2}\mathcal{L}_{\Pi}(\xi)d\xi\Big{)}^{1/2}.
Remark.

If the function g^0\widehat{g}_{0} defined in (4.2) belongs to L2()L^{2}(\mathbb{R}), then it is the Fourier transform of a function g0L2()g_{0}\in L^{2}(\mathbb{R}) and the equality (4.2) is equivalent to gy=Piyg0.g_{y}=P^{\mathbb{H}}_{iy}*g_{0}. However, the notation g^0\widehat{g}_{0} is only formal for a general g2(,μ)g\in\mathscr{B}^{2}(\mathbb{H},\mu), that is, it may not correspond to the Fourier transform of a generalized function g0g_{0} on \mathbb{R}.

Definition.

Suppose that μ(dz)=dxΠ(dy)\mu(dz)=dx\Pi(dy) is a boundary-accessable (1,2)(1,2)-Carleson measure on \mathbb{H}. Let Hμ()H_{\mu}(\mathbb{R}) be the Hilbert space defined by the norm completion as follows:

Hμ():={fL2()|fHμ()=(|f^(ξ)|2Π(ξ)𝑑ξ)1/2<}¯Hμ().H_{\mu}(\mathbb{R}):=\overline{\left\{f\in L^{2}(\mathbb{R})\Big{|}\|f\|_{H_{\mu}(\mathbb{R})}=\Big{(}\int_{\mathbb{R}}|\widehat{f}(\xi)|^{2}\mathcal{L}_{\Pi}(\xi)d\xi\Big{)}^{1/2}<\infty\right\}}^{\|\cdot\|_{H_{\mu}(\mathbb{R})}}.

For any fL2()f\in L^{2}(\mathbb{R}), set

𝒫(f)(z):=(Piyf)(x),z=x+iy.\mathcal{P}^{\mathbb{H}}(f)(z):=(P_{iy}^{\mathbb{H}}*f)(x),\quad z=x+iy\in\mathbb{H}.

Immediately from the definition of Hμ()H_{\mu}(\mathbb{R}), we see that the map

L2()f𝒫(f)2(,μ)L^{2}(\mathbb{R})\ni f\mapsto\mathcal{P}^{\mathbb{H}}(f)\in\mathscr{B}^{2}(\mathbb{H},\mu)

extends to a unitary map from Hμ()H_{\mu}(\mathbb{R}) to 2(,μ)\mathscr{B}^{2}(\mathbb{H},\mu).

Similar to the disk case, for a given measure μ(dz)=dxΠ(dy)\mu(dz)=dx\Pi(dy) on \mathbb{H}, a pair (a,b)(a,b) of two functions on \mathbb{R} is called μ\mu-adapted if the following conditions are satisfied:

  • (i)

    for any ξ\xi\in\mathbb{R}^{*},

    (4.4) |a(ξ)|2b(ξ)sgn(ξ)=Π(ξ)1;\displaystyle|a(\xi)|^{2}b(\xi)\mathrm{sgn}(\xi)=\mathcal{L}_{\Pi}(\xi)^{-1};
  • (ii)

    there exists a constant Cb>0C_{b}>0 such that

    (4.5) 1Cb|b(ξ)|Cb.\displaystyle\frac{1}{C_{b}}\leq|b(\xi)|\leq C_{b}.

Given any Radon measure Π\Pi on +\mathbb{R}_{+} satisfying (2.2), Garnett’s result stated in Lemma 3.4 implies that the following function WΠW^{\Pi} belongs to L()L^{\infty}(\mathbb{R}):

(4.6) WΠ(x):=i+πxy2+π2x2Π(dy),x.\displaystyle W^{\Pi}(x):=i\int_{\mathbb{R}_{+}}\frac{\pi x}{y^{2}+\pi^{2}x^{2}}\Pi(dy),\quad x\in\mathbb{R}.
Proposition 4.1.

Suppose that μ(dz)=dxΠ(dy)\mu(dz)=dx\Pi(dy) is a boundary-accessable (1,2)(1,2)-Carleson measure on \mathbb{H} and let (a,b)(a,b) be a μ\mu-adapted pair of functions defined on \mathbb{R}. Then for any uL2()u\in L^{2}(\mathbb{R}),

(4.7) uL2()C(𝒯auHμ()+L1()+𝒯a𝒯buHμ()+L1()),\displaystyle\|u\|_{L^{2}(\mathbb{\mathbb{R}})}\leq C(\|\mathcal{T}_{a}u\|_{H_{\mu}(\mathbb{R})+L^{1}(\mathbb{R})}+\|\mathcal{T}_{a}\mathcal{T}_{b}u\|_{H_{\mu}(\mathbb{R})+L^{1}(\mathbb{R})}),

where 𝒯a,𝒯b\mathcal{T}_{a},\mathcal{T}_{b} are the Fourier multipliers associated to a,ba,b given by

𝒯au^(ξ)=a(ξ)u^(ξ),𝒯b(u)^(ξ)=b(ξ)u^(ξ)\widehat{\mathcal{T}_{a}u}(\xi)=a(\xi)\widehat{u}(\xi),\quad\widehat{\mathcal{T}_{b}(u)}(\xi)=b(\xi)\widehat{u}(\xi)

and the constant C=C(b,Π)>0C=C(b,\Pi)>0 can be taken to be

(4.8) C(b,Π)=Cb+WΠL()+1<\displaystyle C(b,\Pi)=\sqrt{C_{b}+\|W^{\Pi}\|_{L^{\infty}(\mathbb{R})}+1}<\infty
Remark.

If either 𝒯au\mathcal{T}_{a}u or 𝒯a𝒯bu\mathcal{T}_{a}\mathcal{T}_{b}u does not belong to Hμ()+L1()H_{\mu}(\mathbb{R})+L^{1}(\mathbb{R}), then the right hand side of (4.7) is understood as \infty.

4.1. The derivation of Theorem 1.3 from Proposition 4.1

Let μ(dz)=dxΠ(dy)\mu(dz)=dx\Pi(dy) be a boundary-accessable Radon measure on \mathbb{H}. If the pair (2(,μ),h1())(\mathscr{B}^{2}(\mathbb{H},\mu),h^{1}(\mathbb{H})) is holomorphically stable, then H1()=h1()𝒪()2(,μ)H^{1}(\mathbb{H})=h^{1}(\mathbb{H})\cap\mathcal{O}(\mathbb{H})\subset\mathscr{B}^{2}(\mathbb{H},\mu) and by the Closed Graph Theorem, this embedding is continuous: there exists C>0C>0 such that

(4.9) f2(,μ)CfH1(),fH1().\displaystyle\|f\|_{\mathscr{B}^{2}(\mathbb{H},\mu)}\leq C\|f\|_{H^{1}(\mathbb{H})},\quad\forall f\in H^{1}(\mathbb{H}).

Recall the definition (1.10) of the space Poi()\mathrm{Poi}(\mathbb{H}). Since H1()Poi()H^{1}(\mathbb{H})\subset\mathrm{Poi}(\mathbb{H}),

(4.10) f2(,μ)2=|f(z)|2μ(dz),fH1().\displaystyle\|f\|_{\mathscr{B}^{2}(\mathbb{H},\mu)}^{2}=\int_{\mathbb{H}}|f(z)|^{2}\mu(dz),\quad\forall f\in H^{1}(\mathbb{H}).

The inequality (4.9) and the equality (4.10) together imply that the measure μ\mu is a (1,2)(1,2)-Carleson measure.

Suppose now that μ(dz)=dxΠ(dy)\mu(dz)=dx\Pi(dy) is a boundary-accessable (1,2)(1,2)-Carleson measure on \mathbb{H}. Assume that

(4.11) f=g+h with f𝒪(),g2(,μ),hh1().\displaystyle f=g+h\text{\, with \,}f\in\mathcal{O}(\mathbb{H}),g\in\mathscr{B}^{2}(\mathbb{H},\mu),h\in h^{1}(\mathbb{H}).

Then, the goal is to show that f2(,μ)f\in\mathscr{B}^{2}(\mathbb{H},\mu). It suffices to show

(4.12) f2(,μ)22+WΠL()(g2(,μ)+hh1()).\displaystyle\|f\|_{\mathscr{B}^{2}(\mathbb{H},\mu)}\leq 2\sqrt{2+\|W^{\Pi}\|_{L^{\infty}(\mathbb{R})}}(\|g\|_{\mathscr{B}^{2}(\mathbb{H},\mu)}+\|h\|_{h^{1}(\mathbb{H})}).

To avoid technical issues, we first consider the truncated measures of μ\mu. That is, for any R>0R>0, define

μR(dz)=dxΠR(dy), where ΠR(dy)=𝟙(y<R)Π(dy).\mu_{R}(dz)=dx\Pi_{R}(dy),\text{\, where \,}\Pi_{R}(dy)=\mathds{1}(y<R)\cdot\Pi(dy).

Define WΠRL()W^{\Pi_{R}}\in L^{\infty}(\mathbb{R}) in a similar way as in (4.6). Then WΠRL()WΠL()\|W^{\Pi_{R}}\|_{L^{\infty}(\mathbb{R})}\leq\|W^{\Pi}\|_{L^{\infty}(\mathbb{R})} for any R>0R>0. Therefore, the desired inequality (4.12) follows from

(4.13) f2(,μR)22+WΠRL()(g2(,μR)+hh1()).\displaystyle\|f\|_{\mathscr{B}^{2}(\mathbb{H},\mu_{R})}\leq 2\sqrt{2+\|W^{\Pi_{R}}\|_{L^{\infty}(\mathbb{R})}}\Big{(}\|g\|_{\mathscr{B}^{2}(\mathbb{H},\mu_{R})}+\|h\|_{h^{1}(\mathbb{H})}\Big{)}.

Now we are going to apply Proposition 4.1. For any R>0R>0, define a μR\mu_{R}-adapted pair (aR,b)(a_{R},b) of functions by

aR(ξ):=ΠR(ξ)1/2=(0Re4πy|ξ|Π(dy))1/2 and b(ξ)=sgn(ξ).a_{R}(\xi):=\mathcal{L}_{\Pi_{R}}(\xi)^{-1/2}=\Big{(}\int_{0}^{R}e^{-4\pi y|\xi|}\Pi(dy)\Big{)}^{-1/2}\text{\, and \,}b(\xi)=\mathrm{sgn}(\xi).

In particular, by (2.2),

(4.14) supξaR(ξ)1Π((0,R))<.\displaystyle\sup_{\xi\in\mathbb{R}}a_{R}(\xi)^{-1}\leq\sqrt{\Pi((0,R))}<\infty.

For any y>0y>0, define fy:f_{y}:\mathbb{R}\rightarrow\mathbb{C} and fy:f^{y}:\mathbb{H}\rightarrow\mathbb{C} by

fy(x)=f(x+iy),x and fy(z)=f(z+iy),z.f_{y}(x)=f(x+iy),\,x\in\mathbb{R}\text{\, and \,}f^{y}(z)=f(z+iy),\,z\in\mathbb{H}.

And gy,gy,hy,hyg_{y},g^{y},h_{y},h^{y} are defined similarly.

Claim I.

For any ε>0\varepsilon>0, the function fεf_{\varepsilon} belongs to the classical analytic Hardy space H2()H^{2}(\mathbb{R}) and hence

(4.15) supp(f^ε)[0,).\displaystyle\mathrm{supp}(\widehat{f}_{\varepsilon})\subset[0,\infty).
Remark.

The assertion (4.15) does not follow from the fact that fεf_{\varepsilon} is the restriction onto the real line of a holomorphic function defined on a neighborhood of the closed upper-half plane. For instance, the following function

K(x):=sin(πx)πx,xK(x):=\frac{\sin(\pi x)}{\pi x},\,x\in\mathbb{R}

belongs to L2()L^{2}(\mathbb{R}) and is the restriction of an entire function on the complex plane. However, supp(K^)=[1/2,1/2][0,)\mathrm{supp}(\widehat{K})=[-1/2,1/2]\not\subset[0,\infty).

Since hh1()h\in h^{1}(\mathbb{H}), there exists h0L1()h_{0}\in L^{1}(\mathbb{R}) with

(4.16) hy=Piyh0,y>0.\displaystyle h_{y}=P_{iy}^{\mathbb{H}}*h_{0},\quad\forall y>0.

Thus hyL1()L()L2().h_{y}\in L^{1}(\mathbb{R})\cap L^{\infty}(\mathbb{R})\subset L^{2}(\mathbb{R}). By (1.11), the assumption g2(,μ)g\in\mathscr{B}^{2}(\mathbb{H},\mu) implies that gyL2()L()g_{y}\in L^{2}(\mathbb{R})\cap L^{\infty}(\mathbb{R}). Consequently

(4.17) fy=gy+hyL2().\displaystyle f_{y}=g_{y}+h_{y}\in L^{2}(\mathbb{R}).

Again by (4.16) and (1.11), for any ε>0\varepsilon>0,

fy=Pi(yε)gε+Pi(yε)hε,yε.f_{y}=P_{i(y-\varepsilon)}^{\mathbb{H}}*g_{\varepsilon}+P_{i(y-\varepsilon)}^{\mathbb{H}}*h_{\varepsilon},\quad\forall y\geq\varepsilon.

Therefore, for any ε>0\varepsilon>0,

supy>ε(|f(x+iy)|2𝑑x)1/2gεL2()+hεL2()<.\displaystyle\sup_{y>\varepsilon}\Big{(}\int_{\mathbb{R}}|f(x+iy)|^{2}dx\Big{)}^{1/2}\leq\|g_{\varepsilon}\|_{L^{2}(\mathbb{R})}+\|h_{\varepsilon}\|_{L^{2}(\mathbb{R})}<\infty.

The above inequality combined with f𝒪()f\in\mathcal{O}(\mathbb{H}) implies that fεH2()f_{\varepsilon}\in H^{2}(\mathbb{R}). This completes the proof of Claim I.

Since g2(,μ)g\in\mathscr{B}^{2}(\mathbb{H},\mu), for any ε>0\varepsilon>0, the function gεg^{\varepsilon} belongs to 2(,μ)\mathscr{B}^{2}(\mathbb{H},\mu) and hence gε2(,μR)g^{\varepsilon}\in\mathscr{B}^{2}(\mathbb{H},\mu_{R}). Then by the natural unitary map between HμR()H_{\mu_{R}}(\mathbb{R}) and 2(,μR)\mathscr{B}^{2}(\mathbb{H},\mu_{R}),

gεHμR()=gε2(,μR).\|g_{\varepsilon}\|_{H_{\mu_{R}}(\mathbb{R})}=\|g^{\varepsilon}\|_{\mathscr{B}^{2}(\mathbb{H},\mu_{R})}.

Note that the equality (4.17) and the inequality (4.14) together imply that the function aR(ξ)1f^ε(ξ)a_{R}(\xi)^{-1}\widehat{f}_{\varepsilon}(\xi) belongs to L2()L^{2}(\mathbb{R}). Then there exists a unique function uεL2()u_{\varepsilon}\in L^{2}(\mathbb{R}) with

(4.18) u^ε(ξ)=aR(ξ)1f^ε(ξ).\displaystyle\widehat{u}_{\varepsilon}(\xi)=a_{R}(\xi)^{-1}\widehat{f}_{\varepsilon}(\xi).

Hence, by (4.15), supp(u^ε)[0,)\mathrm{supp}(\widehat{u}_{\varepsilon})\subset[0,\infty). It follows that,

f^ε(ξ)=aR(ξ)u^ε(ξ)=aR(ξ)sgn(ξ)u^ε(ξ)=aR(ξ)b(ξ)u^ε(ξ).\widehat{f}_{\varepsilon}(\xi)=a_{R}(\xi)\widehat{u}_{\varepsilon}(\xi)=a_{R}(\xi)\mathrm{sgn}(\xi)\widehat{u}_{\varepsilon}(\xi)=a_{R}(\xi)b(\xi)\widehat{u}_{\varepsilon}(\xi).

That is, fε=𝒯aRuε=𝒯aR𝒯buε.f_{\varepsilon}=\mathcal{T}_{a_{R}}u_{\varepsilon}=\mathcal{T}_{a_{R}}\mathcal{T}_{b}u_{\varepsilon}. Therefore, since uεL2()u_{\varepsilon}\in L^{2}(\mathbb{R}), we may apply (4.7) and get

uεL2()\displaystyle\|u_{\varepsilon}\|_{L^{2}(\mathbb{R})}\leq 22+WΠRL()fεHμR()+L1()\displaystyle 2\sqrt{2+\|W^{\Pi_{R}}\|_{L^{\infty}(\mathbb{R})}}\|f_{\varepsilon}\|_{H_{\mu_{R}}(\mathbb{R})+L^{1}(\mathbb{R})}
\displaystyle\leq 22+WΠRL()(gεHμR()+hεL1())\displaystyle 2\sqrt{2+\|W^{\Pi_{R}}\|_{L^{\infty}(\mathbb{R})}}\Big{(}\|g_{\varepsilon}\|_{H_{\mu_{R}}(\mathbb{R})}+\|h_{\varepsilon}\|_{L^{1}(\mathbb{R})}\Big{)}
=\displaystyle= 22+WΠRL()(gε2(,μR)+hεL1())\displaystyle 2\sqrt{2+\|W^{\Pi_{R}}\|_{L^{\infty}(\mathbb{R})}}\Big{(}\|g^{\varepsilon}\|_{\mathscr{B}^{2}(\mathbb{H},\mu_{R})}+\|h_{\varepsilon}\|_{L^{1}(\mathbb{R})}\Big{)}
\displaystyle\leq 22+WΠRL()(g2(,μR)+hH1()),\displaystyle 2\sqrt{2+\|W^{\Pi_{R}}\|_{L^{\infty}(\mathbb{R})}}\Big{(}\|g\|_{\mathscr{B}^{2}(\mathbb{H},\mu_{R})}+\|h\|_{H^{1}(\mathbb{H})}\Big{)},

where the last inequality is due to the simple observation: for any ε>0\varepsilon>0,

gε2(,μR)g2(,μR) and hεL1()h0L1()=hh1().\|g^{\varepsilon}\|_{\mathscr{B}^{2}(\mathbb{H},\mu_{R})}\leq\|g\|_{\mathscr{B}^{2}(\mathbb{H},\mu_{R})}\text{\, and \,}\|h_{\varepsilon}\|_{L^{1}(\mathbb{R})}\leq\|h_{0}\|_{L^{1}(\mathbb{R})}=\|h\|_{h^{1}(\mathbb{H})}.

Finally, by Plancherel’s identity, the equalities (4.18) and (4.3),

uεL2()2=|f^ε(ξ)|2aR(ξ)2𝑑ξ=|f^ε(ξ)|2ΠR(ξ)𝑑ξ=fε2(,μR)2.\|u_{\varepsilon}\|_{L^{2}(\mathbb{R})}^{2}=\int_{\mathbb{R}}\frac{|\widehat{f}_{\varepsilon}(\xi)|^{2}}{a_{R}(\xi)^{2}}d\xi=\int_{\mathbb{R}}|\widehat{f}_{\varepsilon}(\xi)|^{2}\mathcal{L}_{\Pi_{R}}(\xi)d\xi=\|f^{\varepsilon}\|_{\mathscr{B}^{2}(\mathbb{H},\mu_{R})}^{2}.

Thus,

fε2(,μR)22+WΠRL()(g2(,μR)+hh1()).\|f^{\varepsilon}\|_{\mathscr{B}^{2}(\mathbb{H},\mu_{R})}\leq 2\sqrt{2+\|W^{\Pi_{R}}\|_{L^{\infty}(\mathbb{R})}}\Big{(}\|g\|_{\mathscr{B}^{2}(\mathbb{H},\mu_{R})}+\|h\|_{h^{1}(\mathbb{H})}\Big{)}.

The inequality (4.13) now follows immediately since

limε0+fε2(,μR)=f2(,μR).\lim_{\varepsilon\to 0^{+}}\|f^{\varepsilon}\|_{\mathscr{B}^{2}(\mathbb{H},\mu_{R})}=\|f\|_{\mathscr{B}^{2}(\mathbb{H},\mu_{R})}.

4.2. The proof of Proposition 4.1

Lemma 4.2.

Let Π\Pi be a Radon measure on +\mathbb{R}_{+} satisfying (2.2). Then there is a function WΠL()W^{\Pi}\in L^{\infty}(\mathbb{R}) such that the following equality

(4.19) u(x)WΠ(x)¯𝑑x=u^(ξ)sgn(ξ)Π(ξ)𝑑ξ\displaystyle\int_{\mathbb{R}}u(x)\overline{W^{\Pi}(x)}dx=\int_{\mathbb{R}}\widehat{u}(\xi)\mathrm{sgn}(\xi)\mathcal{L}_{\Pi}(\xi)d\xi

holds for all uL1()L()u\in L^{1}(\mathbb{R})\cap L^{\infty}(\mathbb{R}) satisfying

(4.20) |u^(ξ)|Π(ξ)𝑑ξ<.\displaystyle\int_{\mathbb{R}}|\widehat{u}(\xi)|\mathcal{L}_{\Pi}(\xi)d\xi<\infty.
Remark.

The equality (4.19) means that the Fourier transform, in a certain distributional sense, of the function WΠW^{\Pi}, is given by WΠ^(ξ)=sgn(ξ)Π(ξ)\widehat{W^{\Pi}}(\xi)=\mathrm{sgn}(\xi)\mathcal{L}_{\Pi}(\xi). If Π(dy)=dy\Pi(dy)=dy is the Lebesgue measure on +\mathbb{R}_{+}, then

Π(ξ)=12|ξ| and WΠ(x)=iπ2sgn(x).\mathcal{L}_{\Pi}(\xi)=\frac{1}{2|\xi|}\text{\, and \,}W^{\Pi}(x)=\frac{i\pi}{2}\mathrm{sgn}(x).

In general, the Fourier transform of WΠW^{\Pi} can only be understood in a certain distributional sense and the condition (4.20) in Lemma 4.2 can not be removed.

The proof of Lemma 4.2 is postponed to the end of this section.

Proof of Proposition 4.1.

Take uL2()u\in L^{2}(\mathbb{R}). Suppose that we have decompositions

(4.21) 𝒯au(x)=f1(x)+g1(x) and 𝒯a𝒯bu(x)=f2(x)+g2(x)\displaystyle\mathcal{T}_{a}u(x)=f_{1}(x)+g_{1}(x)\text{\, and \,}\mathcal{T}_{a}\mathcal{T}_{b}u(x)=f_{2}(x)+g_{2}(x)

with f1,f2Hμ()f_{1},f_{2}\in H_{\mu}(\mathbb{R}) and g1,g2L1()g_{1},g_{2}\in L^{1}(\mathbb{R}). That is,

(4.22) u^(ξ)a(ξ)=f^1(ξ)+g^1(ξ),u^(ξ)a(ξ)b(ξ)=f^2(ξ)+g^2(ξ).\displaystyle\widehat{u}(\xi)a(\xi)=\widehat{f}_{1}(\xi)+\widehat{g}_{1}(\xi),\quad\widehat{u}(\xi)a(\xi)b(\xi)=\widehat{f}_{2}(\xi)+\widehat{g}_{2}(\xi).

For any fixed y>0y>0, applying the Poisson convolution to both sides of (4.21), we have

Piy𝒯au=Piyf1+Piyg1,Piy(𝒯a𝒯b)u=Piyf2+Piyg2.P_{iy}^{\mathbb{H}}*\mathcal{T}_{a}u=P_{iy}^{\mathbb{H}}*f_{1}+P_{iy}^{\mathbb{H}}*g_{1},\quad P_{iy}^{\mathbb{H}}*(\mathcal{T}_{a}\mathcal{T}_{b})u=P_{iy}^{\mathbb{H}}*f_{2}+P_{iy}^{\mathbb{H}}*g_{2}.

By Plancherel’s identity and (4.22),

(4.23) PiyuL2()2=|Piy^(ξ)|2|u^(ξ)|2𝑑ξ=|Piy^(ξ)|2u^(ξ)u^(ξ)¯𝑑ξ=|Piy^(ξ)|2f^1(ξ)+g^1(ξ)a(ξ)u^(ξ)¯𝑑ξ=|Piy^(ξ)|2f^1(ξ)a(ξ)u^(ξ)¯𝑑ξdenoted by I1+|Piy^(ξ)|2g^1(ξ)a(ξ)u^(ξ)¯𝑑ξdenoted by I2.\displaystyle\begin{split}\|P_{iy}^{\mathbb{H}}*u\|^{2}_{L^{2}(\mathbb{R})}=&\int_{\mathbb{R}}|\widehat{P_{iy}^{\mathbb{H}}}(\xi)|^{2}|\widehat{u}(\xi)|^{2}d\xi=\int_{\mathbb{R}}|\widehat{P_{iy}^{\mathbb{H}}}(\xi)|^{2}\widehat{u}(\xi)\overline{\widehat{u}(\xi)}d\xi\\ =&\int_{\mathbb{R}}|\widehat{P_{iy}^{\mathbb{H}}}(\xi)|^{2}\frac{\widehat{f}_{1}(\xi)+\widehat{g}_{1}(\xi)}{a(\xi)}\overline{\widehat{u}(\xi)}d\xi\\ =&\underbrace{\int_{\mathbb{R}}|\widehat{P_{iy}^{\mathbb{H}}}(\xi)|^{2}\frac{\widehat{f}_{1}(\xi)}{a(\xi)}\overline{\widehat{u}(\xi)}d\xi}_{\text{denoted by $\mathrm{I}_{1}$}}+\underbrace{\int_{\mathbb{R}}|\widehat{P_{iy}^{\mathbb{H}}}(\xi)|^{2}\frac{\widehat{g}_{1}(\xi)}{a(\xi)}\overline{\widehat{u}(\xi)}d\xi}_{\text{denoted by $\mathrm{I}_{2}$}}.\end{split}

Cauchy-Schwarz’s inequality and the conditions (4.4), (4.5) together imply

(4.24) |I1|(|Piy^(ξ)u^(ξ)|2)1/2(|Piy^(ξ)f^1(ξ)a(ξ)|2)1/2CbPiyuL2()Piyf1Hμ()CbPiyuL2()f1Hμ().\displaystyle\begin{split}|\mathrm{I}_{1}|&\leq\Big{(}\int_{\mathbb{R}}|\widehat{P_{iy}^{\mathbb{H}}}(\xi)\widehat{u}(\xi)|^{2}\Big{)}^{1/2}\Big{(}\int_{\mathbb{R}}\Big{|}\frac{\widehat{P_{iy}^{\mathbb{H}}}(\xi)\widehat{f}_{1}(\xi)}{a(\xi)}\Big{|}^{2}\Big{)}^{1/2}\\ &\leq\sqrt{C_{b}}\left\|P_{iy}^{\mathbb{H}}*u\right\|_{L^{2}(\mathbb{R})}\left\|P_{iy}^{\mathbb{H}}*f_{1}\right\|_{H_{\mu}(\mathbb{R})}\\ &\leq\sqrt{C_{b}}\left\|P_{iy}^{\mathbb{H}}*u\right\|_{L^{2}(\mathbb{R})}\|f_{1}\|_{H_{\mu}(\mathbb{R})}.\end{split}

And, by (4.22) and b(ξ)b(\xi)\in\mathbb{R}, the integral I2\mathrm{I}_{2} can be decomposed as

(4.25) I2=|Piy^(ξ)|2g^1(ξ)a(ξ)(f2^(ξ)+g^2(ξ)a(ξ)b(ξ))¯𝑑ξ=|Piy^(ξ)|2(a(ξ)u^(ξ)f^1(ξ))f^2(ξ)¯|a(ξ)|2b(ξ)𝑑ξdenoted by I3+|Piy^(ξ)|2g^1(ξ)g^2(ξ)¯|a(ξ)|2b(ξ)𝑑ξdenoted by I4.\displaystyle\begin{split}\mathrm{I}_{2}=&\int_{\mathbb{R}}|\widehat{P_{iy}^{\mathbb{H}}}(\xi)|^{2}\frac{\widehat{g}_{1}(\xi)}{a(\xi)}\overline{\Big{(}\frac{\widehat{f_{2}}(\xi)+\widehat{g}_{2}(\xi)}{a(\xi)b(\xi)}\Big{)}}d\xi\\ =&\underbrace{\int_{\mathbb{R}}|\widehat{P_{iy}^{\mathbb{H}}}(\xi)|^{2}\frac{(a(\xi)\widehat{u}(\xi)-\widehat{f}_{1}(\xi))\overline{\widehat{f}_{2}(\xi)}}{|a(\xi)|^{2}b(\xi)}d\xi}_{\text{denoted by $\mathrm{I}_{3}$}}+\underbrace{\int_{\mathbb{R}}|\widehat{P_{iy}^{\mathbb{H}}}(\xi)|^{2}\frac{\widehat{g}_{1}(\xi)\overline{\widehat{g}_{2}(\xi)}}{|a(\xi)|^{2}b(\xi)}d\xi}_{\text{denoted by $\mathrm{I}_{4}$}}.\end{split}

The integral I3\mathrm{I}_{3} can be easily controlled. Indeed, again by Cauchy-Schwarz’s inequality and (4.4), (4.5),

(4.26) |I3|(|Piy^(ξ)u^(ξ)|2𝑑ξ)1/2(|Piy^(ξ)f^2(ξ)a(ξ)b(ξ)|2𝑑ξ)1/2+(|Piy^(ξ)f^1(ξ)|2|a(ξ)|2|b(ξ)|𝑑ξ)1/2(|Piy^(ξ)f^2(ξ)|2|a(ξ)|2|b(ξ)|𝑑ξ)1/2CbPiyuL2()Piyf2Hμ()+Piyf1Hμ()Piyf2Hμ()CbPiyuL2()f2Hμ()+f1Hμ()f2Hμ().\displaystyle\begin{split}|\mathrm{I}_{3}|\leq&\Big{(}\int_{\mathbb{R}}|\widehat{P_{iy}^{\mathbb{H}}}(\xi)\widehat{u}(\xi)|^{2}d\xi\Big{)}^{1/2}\Big{(}\int_{\mathbb{R}}\Big{|}\frac{\widehat{P_{iy}^{\mathbb{H}}}(\xi)\widehat{f}_{2}(\xi)}{a(\xi)b(\xi)}\Big{|}^{2}d\xi\Big{)}^{1/2}\\ &+\Big{(}\int_{\mathbb{R}}\frac{|\widehat{P_{iy}^{\mathbb{H}}}(\xi)\widehat{f}_{1}(\xi)|^{2}}{|a(\xi)|^{2}|b(\xi)|}d\xi\Big{)}^{1/2}\Big{(}\int_{\mathbb{R}}\frac{|\widehat{P_{iy}^{\mathbb{H}}}(\xi)\widehat{f}_{2}(\xi)|^{2}}{|a(\xi)|^{2}|b(\xi)|}d\xi\Big{)}^{1/2}\\ \leq&\sqrt{C_{b}}\|P_{iy}^{\mathbb{H}}*u\|_{L^{2}(\mathbb{R})}\|P_{iy}^{\mathbb{H}}*f_{2}\|_{H_{\mu}(\mathbb{R})}+\|P_{iy}^{\mathbb{H}}*f_{1}\|_{H_{\mu}(\mathbb{R})}\|P_{iy}^{\mathbb{H}}*f_{2}\|_{H_{\mu}(\mathbb{R})}\\ \leq&\sqrt{C_{b}}\|P_{iy}^{\mathbb{H}}*u\|_{L^{2}(\mathbb{R})}\|f_{2}\|_{H_{\mu}(\mathbb{R})}+\|f_{1}\|_{H_{\mu}(\mathbb{R})}\|f_{2}\|_{H_{\mu}(\mathbb{R})}.\end{split}

It remains to estimate the integral I4\mathrm{I}_{4}. Since g1,g2L1()g_{1},g_{2}\in L^{1}(\mathbb{R}), for any y>0y>0, one can define

(4.27) Gy:=(Piyg1)(Piyg~2)=P2iy(g1g~2), where g~2(x):=g2(x).\displaystyle G_{y}:=(P_{iy}^{\mathbb{H}}*g_{1})*(P_{iy}^{\mathbb{H}}*\tilde{g}_{2})=P_{2iy}^{\mathbb{H}}*(g_{1}*\tilde{g}_{2}),\text{\, where \,}\tilde{g}_{2}(x):=g_{2}(-x).

In particular,

(4.28) Gy^(ξ)=|Piy^(ξ)|2g^1(ξ)g^2(ξ)¯=e4πy|ξ|g^1(ξ)g^2(ξ)¯.\displaystyle\widehat{G_{y}}(\xi)=|\widehat{P_{iy}^{\mathbb{H}}}(\xi)|^{2}\widehat{g}_{1}(\xi)\overline{\widehat{g}_{2}(\xi)}=e^{-4\pi y|\xi|}\widehat{g}_{1}(\xi)\overline{\widehat{g}_{2}(\xi)}.

Claim A.

For any y>0y>0, the function GyG_{y} defined in (4.27) satisfies

(4.29) GyL1()L()\displaystyle G_{y}\in L^{1}(\mathbb{R})\cap L^{\infty}(\mathbb{R})

and

(4.30) |Gy^(ξ)|Π(ξ)𝑑ξ<.\displaystyle\int_{\mathbb{R}}|\widehat{G_{y}}(\xi)|\mathcal{L}_{\Pi}(\xi)d\xi<\infty.

Indeed, g1g~2L1()g_{1}*\tilde{g}_{2}\in L^{1}(\mathbb{R}) since g1,g2L1()g_{1},g_{2}\in L^{1}(\mathbb{R}). Therefore, (4.29) follows from the definition (4.27) and the simple observation that P2iyL1()L()P_{2iy}^{\mathbb{H}}\in L^{1}(\mathbb{R})\cap L^{\infty}(\mathbb{R}). By (4.22),

g^1(ξ)a(ξ)=u^(ξ)f^1(ξ)a(ξ),g^2(ξ)a(ξ)b(ξ)=u^(ξ)f^2(ξ)a(ξ)b(ξ).\frac{\widehat{g}_{1}(\xi)}{a(\xi)}=\widehat{u}(\xi)-\frac{\widehat{f}_{1}(\xi)}{a(\xi)},\quad\frac{\widehat{g}_{2}(\xi)}{a(\xi)b(\xi)}=\widehat{u}(\xi)-\frac{\widehat{f}_{2}(\xi)}{a(\xi)b(\xi)}.

The assumptions f1,f2Hμ()f_{1},f_{2}\in H_{\mu}(\mathbb{R}) combined with the conditions (4.4), (4.5) on the pair (a,b)(a,b) imply that both functions f^1/a\widehat{f}_{1}/a and f^2/(ab)\widehat{f}_{2}/(ab) belong to L2()L^{2}(\mathbb{R}). Since uL2()u\in L^{2}(\mathbb{R}) and hence u^L2()\widehat{u}\in L^{2}(\mathbb{R}), we obtain, by using (4.4) again, that

|Gy^(ξ)|Π(ξ)𝑑ξ\displaystyle\int_{\mathbb{R}}|\widehat{G_{y}}(\xi)|\mathcal{L}_{\Pi}(\xi)d\xi =e4πy|ξ||g^1(ξ)||a(ξ)||g^2(ξ)||a(ξ)b(ξ)|𝑑ξ\displaystyle=\int_{\mathbb{R}}e^{-4\pi y|\xi|}\frac{|\widehat{g}_{1}(\xi)|}{|a(\xi)|}\cdot\frac{|\widehat{g}_{2}(\xi)|}{|a(\xi)b(\xi)|}d\xi
|u^(ξ)f^1(ξ)a(ξ)||u^(ξ)f^2(ξ)a(ξ)b(ξ)|𝑑ξ\displaystyle\leq\int_{\mathbb{R}}\Big{|}\widehat{u}(\xi)-\frac{\widehat{f}_{1}(\xi)}{a(\xi)}\Big{|}\cdot\Big{|}\widehat{u}(\xi)-\frac{\widehat{f}_{2}(\xi)}{a(\xi)b(\xi)}\Big{|}d\xi
u^f^1aL2()u^f^2abL2()<.\displaystyle\leq\Big{\|}\widehat{u}-\frac{\widehat{f}_{1}}{a}\Big{\|}_{L^{2}(\mathbb{R})}\Big{\|}\widehat{u}-\frac{\widehat{f}_{2}}{ab}\Big{\|}_{L^{2}(\mathbb{R})}<\infty.

By Claim A, the function GyG_{y} satisfies all the required conditions of Lemma 4.2. Hence, by (4.28), (4.4) and (4.19),

I4=Gy^(ξ)sgn(ξ)Π(ξ)𝑑ξ=Gy(x)WΠ(x)¯𝑑x.\mathrm{I}_{4}=\int_{\mathbb{R}}\widehat{G_{y}}(\xi)\mathrm{sgn}(\xi)\mathcal{L}_{\Pi}(\xi)d\xi=\int_{\mathbb{R}}G_{y}(x)\overline{W^{\Pi}(x)}dx.

It follows that

(4.31) |I4|WΠL()GyL1()=WΠL()P2iy(g1g~2)L1()WΠL()g1L1()g2L1().\displaystyle\begin{split}|\mathrm{I}_{4}|&\leq\|W^{\Pi}\|_{L^{\infty}(\mathbb{R})}\|G_{y}\|_{L^{1}(\mathbb{R})}=\|W^{\Pi}\|_{L^{\infty}(\mathbb{R})}\|P_{2iy}^{\mathbb{H}}*(g_{1}*\tilde{g}_{2})\|_{L^{1}(\mathbb{R})}\\ &\leq\|W^{\Pi}\|_{L^{\infty}(\mathbb{R})}\|g_{1}\|_{L^{1}(\mathbb{R})}\|g_{2}\|_{L^{1}(\mathbb{R})}.\end{split}

Combining (4.23), (4.24), (4.25), (4.26) and (4.31), we get

PiyuL2()2\displaystyle\|P_{iy}^{\mathbb{H}}*u\|^{2}_{L^{2}(\mathbb{R})}\leq CbPiyuL2()f1Hμ()+CbPiyuL2()f2Hμ()\displaystyle\sqrt{C_{b}}\left\|P_{iy}^{\mathbb{H}}*u\right\|_{L^{2}(\mathbb{R})}\|f_{1}\|_{H_{\mu}(\mathbb{R})}+\sqrt{C_{b}}\|P_{iy}^{\mathbb{H}}*u\|_{L^{2}(\mathbb{R})}\|f_{2}\|_{H_{\mu}(\mathbb{R})}
+f1Hμ()f2Hμ()+WΠL()g1L1()g2L1().\displaystyle+\|f_{1}\|_{H_{\mu}(\mathbb{R})}\|f_{2}\|_{H_{\mu}(\mathbb{R})}+\|W^{\Pi}\|_{L^{\infty}(\mathbb{R})}\|g_{1}\|_{L^{1}(\mathbb{R})}\|g_{2}\|_{L^{1}(\mathbb{R})}.

Therefore, by a standard argument, there exists a constant C>0C>0 depending only on the constants CbC_{b} and WΠL()\|W^{\Pi}\|_{L^{\infty}(\mathbb{R})} such that

PiyuL2()C(f1Hμ()+g1L1()+f2Hμ()+g2L1()).\|P_{iy}^{\mathbb{H}}*u\|_{L^{2}(\mathbb{R})}\leq C(\|f_{1}\|_{H_{\mu}(\mathbb{R})}+\|g_{1}\|_{L^{1}(\mathbb{R})}+\|f_{2}\|_{H_{\mu}(\mathbb{R})}+\|g_{2}\|_{L^{1}(\mathbb{R})}).

The constant CC in the above inequality can be taken to be

C=Cb+WΠL()+1.C=\sqrt{C_{b}+\|W^{\Pi}\|_{L^{\infty}(\mathbb{R})}+1}.

Since the decompositions (4.21) are arbitrary, we get

PiyuL2()C(𝒯auHμ()+L1()+𝒯b𝒯auHμ()+L1()).\|P_{iy}^{\mathbb{H}}*u\|_{L^{2}(\mathbb{R})}\leq C(\|\mathcal{T}_{a}u\|_{H_{\mu}(\mathbb{R})+L^{1}(\mathbb{R})}+\|\mathcal{T}_{b}\mathcal{T}_{a}u\|_{H_{\mu}(\mathbb{R})+L^{1}(\mathbb{R})}).

Finally, by taking the limit y0+y\to 0^{+} and using

limy0+PiyuL2()=uL2(),\lim_{y\to 0^{+}}\|P_{iy}^{\mathbb{H}}*u\|_{L^{2}(\mathbb{R})}=\|u\|_{L^{2}(\mathbb{R})},

we obtain the desired inequality (4.7) and complete the whole proof of the proposition. ∎

Proof of Lemma 4.2.

Fix a Radon measure Π\Pi on +\mathbb{R}_{+} satisfying (2.2). By Garnett’s result stated in Lemma 3.4, one can define a function WΠL()W^{\Pi}\in L^{\infty}(\mathbb{R}) by (4.6).

Now we show that WΠW^{\Pi} satisfies the equality (4.19). For any 0<ε<R<0<\varepsilon<R<\infty, set

(4.32) Wε,RΠ(x):=i+πxy2+π2x2Πε,R(dy), where Πε,R(dy)=𝟙(ε<y<R)Π(dy).\displaystyle W^{\Pi}_{\varepsilon,R}(x):=i\int_{\mathbb{R}_{+}}\frac{\pi x}{y^{2}+\pi^{2}x^{2}}\Pi_{\varepsilon,R}(dy),\text{\, where \,}\Pi_{\varepsilon,R}(dy)=\mathds{1}(\varepsilon<y<R)\cdot\Pi(dy).

Claim B.

For any 0<ε<R<0<\varepsilon<R<\infty, we have Wε,RΠL2()W^{\Pi}_{\varepsilon,R}\in L^{2}(\mathbb{R}) and the Fourier transform of Wε,RΠW^{\Pi}_{\varepsilon,R} is given by the Bochner integral for L2()L^{2}(\mathbb{R})-vector valued function:

(4.33) Wε,RΠ^=εRyΠ(dy),y(ξ):=sgn(ξ)e2y|ξ|.\displaystyle\widehat{W^{\Pi}_{\varepsilon,R}}=\int_{\varepsilon}^{R}\ell_{y}\Pi(dy),\quad\ell_{y}(\xi):=\mathrm{sgn}(\xi)e^{-2y|\xi|}.

In particular, Wε,RΠ^\widehat{W^{\Pi}_{\varepsilon,R}} can be identified with a C()C^{\infty}(\mathbb{R}^{*})-function by the formula

(4.34) Wε,RΠ^(ξ)=εRsgn(ξ)e2y|ξ|Π(dy),ξ={0}.\displaystyle\widehat{W^{\Pi}_{\varepsilon,R}}(\xi)=\int_{\varepsilon}^{R}\mathrm{sgn}(\xi)e^{-2y|\xi|}\Pi(dy),\quad\xi\in\mathbb{R}^{*}=\mathbb{R}\setminus\{0\}.

Indeed, for any y>0y>0, recall that the conjugate Poisson kernel (see, e.g., [Gra14, formula (4.1.16)]) of \mathbb{H} is given by

Qiy(x)=πxy2+π2x2,x.Q_{iy}^{\mathbb{H}}(x)=\frac{\pi x}{y^{2}+\pi^{2}x^{2}},\quad x\in\mathbb{R}.

Clearly, QiyL2()Q_{iy}^{\mathbb{H}}\in L^{2}(\mathbb{R}) for all y>0y>0 and the map yQiyy\mapsto Q_{iy}^{\mathbb{H}} is continuous from +\mathbb{R}_{+} to L2()L^{2}(\mathbb{R}), hence it is uniformly continuous from [ε,R][\varepsilon,R] to L2()L^{2}(\mathbb{R}). Consequently, using the definition (4.32) of Wε,RΠW_{\varepsilon,R}^{\Pi} and the fact that Πε,R\Pi_{\varepsilon,R} is a finite measure with support contained in [ε,R][\varepsilon,R], we obtain that Wε,RΠL2()W_{\varepsilon,R}^{\Pi}\in L^{2}(\mathbb{R}) and the following equality in the sense of the Bochner integral for L2()L^{2}(\mathbb{R})-vector valued functions:

Wε,RΠ^=iεRQiy^Π(dy).\widehat{W_{\varepsilon,R}^{\Pi}}=i\int_{\varepsilon}^{R}\widehat{Q_{iy}^{\mathbb{H}}}\Pi(dy).

Then the equality (4.33) follows immediately since (see, e.g., [Gra14, formula (4.1.33)])

Qiy^(ξ)=iy(ξ)=isgn(ξ)e2y|ξ|.\widehat{Q_{iy}^{\mathbb{H}}}(\xi)=-i\ell_{y}(\xi)=-i\mathrm{sgn}(\xi)e^{-2y|\xi|}.

Claim C.

For any φL1()\varphi\in L^{1}(\mathbb{R}),

(4.35) limε0+φ(x)[0επxy2+π2x2Π(dy)]𝑑x=0.\displaystyle\lim_{\varepsilon\rightarrow 0^{+}}\int_{\mathbb{R}}\varphi(x)\Big{[}\int_{0}^{\varepsilon}\frac{\pi x}{y^{2}+\pi^{2}x^{2}}\Pi(dy)\Big{]}dx=0.

Indeed, for any y>0y>0, set FΠ(y):=Π((0,y])F_{\Pi}(y):=\Pi((0,y]). Then Π(dy)=dFΠ(y)\Pi(dy)=dF_{\Pi}(y) and by the assumption (2.2), there exists a constant C>0C>0 such that

(4.36) FΠ(y)Cy,y>0.\displaystyle F_{\Pi}(y)\leq Cy,\quad\forall y>0.

By integration by parts for the absolutely continuous function FΠF_{\Pi},

(4.37) 0επxy2+π2x2Π(dy)=πxy2+π2x2FΠ(y)|y=0y=ε+0εFΠ(y)2πxy(y2+π2x2)2𝑑y.\displaystyle\int_{0}^{\varepsilon}\frac{\pi x}{y^{2}+\pi^{2}x^{2}}\Pi(dy)=\frac{\pi x}{y^{2}+\pi^{2}x^{2}}F_{\Pi}(y)\Big{|}_{y=0}^{y=\varepsilon}+\int_{0}^{\varepsilon}F_{\Pi}(y)\frac{2\pi xy}{(y^{2}+\pi^{2}x^{2})^{2}}dy.

In particular, if Π(dy)\Pi(dy) is the Lebesgue measure on +\mathbb{R}_{+}, then the equality (4.37) becomes

(4.38) arctan(επx)=0επxy2+π2x2𝑑y=πxy2+π2x2y|y=0y=ε+0εy2πxy(y2+π2x2)2𝑑y.\displaystyle\arctan\left(\frac{\varepsilon}{\pi x}\right)=\int_{0}^{\varepsilon}\frac{\pi x}{y^{2}+\pi^{2}x^{2}}dy=\frac{\pi x}{y^{2}+\pi^{2}x^{2}}y\Big{|}_{y=0}^{y=\varepsilon}+\int_{0}^{\varepsilon}y\frac{2\pi xy}{(y^{2}+\pi^{2}x^{2})^{2}}dy.

Comparing (4.37) and (4.38) and using (4.36), we obtain

0επ|x|y2+π2x2Π(dy)Carctan(επ|x|).\int_{0}^{\varepsilon}\frac{\pi|x|}{y^{2}+\pi^{2}x^{2}}\Pi(dy)\leq C\arctan\left(\frac{\varepsilon}{\pi|x|}\right).

Therefore, by dominated convergence theorem, for any φL1()\varphi\in L^{1}(\mathbb{R}),

lim supε0+|φ(x)0επxy2+π2x2Π(dy)𝑑x|\displaystyle\limsup_{\varepsilon\to 0^{+}}\Big{|}\int_{\mathbb{R}}\varphi(x)\int_{0}^{\varepsilon}\frac{\pi x}{y^{2}+\pi^{2}x^{2}}\Pi(dy)dx\Big{|} Clim supε0+|φ(x)|arctan(επ|x|)𝑑x=0.\displaystyle\leq C\limsup_{\varepsilon\to 0^{+}}\int_{\mathbb{R}}|\varphi(x)|\arctan\left(\frac{\varepsilon}{\pi|x|}\right)dx=0.

Claim D.

For any φL1()\varphi\in L^{1}(\mathbb{R}),

(4.39) limRφ(x)[Rπxy2+π2x2Π(dy)]𝑑x=0.\displaystyle\lim_{R\rightarrow\infty}\int_{\mathbb{R}}\varphi(x)\Big{[}\int_{R}^{\infty}\frac{\pi x}{y^{2}+\pi^{2}x^{2}}\Pi(dy)\Big{]}dx=0.

The proof of the equality (4.39) is similar to that of (4.35) and thus is omitted here.

Now fix any uL1()L()u\in L^{1}(\mathbb{R})\cap L^{\infty}(\mathbb{R}) satisfying (4.20). By (4.35) and (4.39),

(4.40) limε0+Ru(x)Wε,RΠ(x)¯𝑑x=u(x)WΠ(x)¯𝑑x.\displaystyle\lim_{\varepsilon\to 0^{+}\atop R\to\infty}\int_{\mathbb{R}}u(x)\overline{W^{\Pi}_{\varepsilon,R}(x)}dx=\int_{\mathbb{R}}u(x)\overline{W^{\Pi}(x)}dx.

Moreover, since both uu and WΠW^{\Pi} belong to L2()L^{2}(\mathbb{R}), the Plancherel’s identity implies

(4.41) u(x)Wε,RΠ(x)¯𝑑x=u^(ξ)Wε,RΠ^(ξ)¯𝑑ξ=u^(ξ)sgn(ξ)[εRe2y|ξ|Π(dy)]𝑑ξ.\displaystyle\int_{\mathbb{R}}u(x)\overline{W^{\Pi}_{\varepsilon,R}(x)}dx=\int_{\mathbb{R}}\widehat{u}(\xi)\overline{\widehat{W_{\varepsilon,R}^{\Pi}}(\xi)}d\xi=\int_{\mathbb{R}}\widehat{u}(\xi)\mathrm{sgn}(\xi)\Big{[}\int_{\varepsilon}^{R}e^{-2y|\xi|}\Pi(dy)\Big{]}d\xi.

Using the assumption (4.20), we obtain, by dominated convergence theorem,

(4.42) limε0+Ru^(ξ)sgn(ξ)[εRe2y|ξ|Π(dy)]𝑑ξ=u^(ξ)sgn(ξ)Π(ξ)𝑑ξ.\displaystyle\lim_{\varepsilon\to 0^{+}\atop R\to\infty}\int_{\mathbb{R}}\widehat{u}(\xi)\mathrm{sgn}(\xi)\Big{[}\int_{\varepsilon}^{R}e^{-2y|\xi|}\Pi(dy)\Big{]}d\xi=\int_{\mathbb{R}}\widehat{u}(\xi)\mathrm{sgn}(\xi)\mathcal{L}_{\Pi}(\xi)d\xi.

Combining (4.40), (4.41) and (4.42), we obtain the desired equality (4.19). ∎

References

  • [BB78] David Bekollé and Aline Bonami. Inégalités à poids pour le noyau de Bergman. C. R. Acad. Sci. Paris Sér. A-B, 286(18):A775–A778, 1978.
  • [Car21] Torsten Carleman. Zur Theorie der linearen Integralgleichungen. Math. Z., 9(3-4):196–217, 1921.
  • [Car62] Lennart Carleson. Interpolations by bounded analytic functions and the corona problem. Ann. of Math. (2), 76:547–559, 1962.
  • [DLRW21] Francesca Da Lio, Tristan Rivière, and Jerome Wettstein. Bergman-Bourgain-Brezis-type inequality. J. Funct. Anal., 281(9):Paper No. 109201, 33, 2021.
  • [Dur69] Peter L. Duren. Extension of a theorem of Carleson. Bull. Amer. Math. Soc., 75:143–146, 1969.
  • [GK89] Theodore W. Gamelin and Dmitry S. Khavinson. The isoperimetric inequality and rational approximation. Amer. Math. Monthly, 96(1):18–30, 1989.
  • [Gra14] Loukas Grafakos. Classical Fourier analysis, volume 249 of Graduate Texts in Mathematics. Springer, New York, third edition, 2014.
  • [Har09] Zen Harper. Boundedness of convolution operators and input-output maps between weighted spaces. Complex Anal. Oper. Theory, 3(1):113–146, 2009.
  • [Hor67] Lars Hormander. LpL^{p} estimates for (pluri-) subharmonic functions. Math. Scand., 20:65–78, 1967.
  • [JPP13] Birgit Jacob, Jonathan R. Partington, and Sandra Pott. On Laplace-Carleson embedding theorems. J. Funct. Anal., 264(3):783–814, 2013.
  • [Kat04] Yitzhak Katznelson. An introduction to harmonic analysis. Cambridge Mathematical Library. Cambridge University Press, Cambridge, third edition, 2004.
  • [RU88] Wade Ramey and David Ullrich. On the behavior of harmonic functions near a boundary point. Trans. Amer. Math. Soc., 305(1):207–220, 1988.
  • [Ryd20] Eskil Rydhe. On Laplace-Carleson embeddings, and LpL^{p}-mapping properties of the Fourier transform. Ark. Mat., 58(2):437–457, 2020.
  • [Sai79] Saburou Saitoh. The Bergman norm and the Szego norm. Trans. Amer. Math. Soc., 249(2):261–279, 1979.
  • [Vuk03] Dragan Vukotić. The isoperimetric inequality and a theorem of Hardy and Littlewood. Amer. Math. Monthly, 110(6):532–536, 2003.
  • [Yos95] Kōsaku Yosida. Functional analysis. Classics in Mathematics. Springer-Verlag, Berlin, 1995. Reprint of the sixth (1980) edition.