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

Zygmund’s theorem for harmonic quasiregular mappings

Suman Das Suman DasDepartment of Mathematics with Computer Science, Guangdong Technion - Israel Institute of Technology, Shantou, Guangdong 515063, P. R. China. suman.das@gtiit.edu.cn Jie Huang Jie Huang Department of Mathematics with Computer Science, Guangdong Technion - Israel Institute of Technology, Shantou, Guangdong 515063, P. R. China. jie.huang@gtiit.edu.cn  and  Antti Rasila Antti Rasila Department of Mathematics with Computer Science, Guangdong Technion - Israel Institute of Technology, Shantou, Guangdong 515063, P. R. China. Department of Mathematics, Technion - Israel Institute of Technology, Haifa 3200003, Israel. antti.rasila@gtiit.edu.cn; antti.rasila@iki.fi
Abstract.

Given an analytic function f=u+ivf=u+iv in the unit disk 𝔻{\mathbb{D}}, Zygmund’s theorem gives the minimal growth restriction on uu which ensures that vv is in the Hardy space h1h^{1}. This need not be true if ff is a complex-valued harmonic function. However, we prove that Zygmund’s theorem holds if ff is a harmonic KK-quasiregular mapping in 𝔻{\mathbb{D}}. Our work makes further progress on the recent Riesz-type theorem of Liu and Zhu (Adv. Math., 2023), and the Kolmogorov-type theorem of Kalaj (J. Math. Anal. Appl., 2025), for harmonic quasiregular mappings. We also obtain a partial converse, thus showing that the proposed growth condition is the best possible. Furthermore, as an application of the classical conjugate function theorems, we establish a harmonic analogue of a well-known result of Hardy and Littlewood.

Key words and phrases:
Hardy space; Harmonic functions; Quasiregular mappings; Zygmund theorem; Riesz theorem
2020 Mathematics Subject Classification:
primary: 31A05, 30H10; secondary: 30C62

1. Introduction

1.1. Notations and preliminaries

Suppose 𝔻{z:|z|<1}{\mathbb{D}}\coloneqq\{z\in{\mathbb{C}}:\,|z|<1\} is the open unit disk in the complex plane {\mathbb{C}}, and 𝕋{z:|z|=1}\mathbb{T}\coloneqq\{z\in{\mathbb{C}}:|z|=1\} is the unit circle. For a function ff analytic in 𝔻{\mathbb{D}}, the integral means of ff are defined as

Mp(r,f)(12π02π|f(reiθ)|p𝑑θ)1/pfor0<p<,M_{p}(r,f)\coloneqq\left(\frac{1}{2\pi}\int_{0}^{2\pi}|f(re^{i\theta})|^{p}d\theta\right)^{1/p}\quad\text{for}\quad 0<p<\infty,

and

M(r,f)sup|z|=r|f(z)|.M_{\infty}(r,f)\coloneqq\sup_{|z|=r}|f(z)|.

The function ff is said to be in the Hardy space HpH^{p} (0<p)(0<p\leq\infty) if

fplimr1Mp(r,f)<.\|f\|_{p}\coloneqq\lim_{r\to 1^{-}}M_{p}(r,f)<\infty.

Every function fHpf\in H^{p} has “nice” boundary behaviour, in the sense that the radial limit

f(eiθ)limr1f(reiθ)f(e^{i\theta})\coloneqq\lim_{r\to 1^{-}}f(r{e^{i\theta}})

exists in almost every direction, and is of class LpLp(𝕋)L^{p}\coloneqq L^{p}({\mathbb{T}}). Detailed surveys on Hardy spaces can be found, e.g., in the books of Duren [7] and Koosis [11]. Throughout this paper, we follow notations from [7].

A real-valued function u(x,y)u(x,y), twice continuously differentiable in 𝔻{\mathbb{D}}, is called harmonic if it satisfies the Laplace equation

Δu=2ux2+2uy2=0in𝔻.\Delta u=\frac{\partial^{2}u}{\partial x^{2}}+\frac{\partial^{2}u}{\partial y^{2}}=0\quad\text{in}\quad{\mathbb{D}}.

For a complex-valued function f=u+ivf=u+iv, the complex partial derivatives have the form

fz=12(fxify)andfz¯=12(fx+ify),f_{z}=\frac{1}{2}\left(f_{x}-if_{y}\right)\quad\text{and}\quad f_{\bar{z}}=\frac{1}{2}\left(f_{x}+if_{y}\right),

where z=x+iyz=x+iy\in{\mathbb{C}}. In view of this, it is easy to see that the Laplacian operator Δ\Delta can also be written as

Δ=42z¯z\Delta=4\,\frac{\partial^{2}}{\partial\bar{z}\partial z}\cdot

A function f=u+ivf=u+iv is harmonic in 𝔻{\mathbb{D}}, if uu and vv are real-valued harmonic functions in 𝔻{\mathbb{D}}. Every such function has a representation f=h+g¯f=h+\overline{g}, where hh and gg are analytic in 𝔻\mathbb{D}. This representation is unique up to an additive constant. In this paper we usually assume g(0)=0g(0)=0, unless mentioned otherwise. Similar to the HpH^{p} spaces, the harmonic Hardy spaces hph^{p} are defined as the class of functions ff, harmonic in 𝔻{\mathbb{D}}, such that

fp=limr1Mp(r,f)<.\|f\|_{p}=\lim_{r\to 1^{-}}M_{p}(r,f)<\infty.

We refer to the books of Duren [6] for the theory of planar harmonic mappings and Pavlović [15] for a concise survey on hph^{p} spaces.

Given K1K\geq 1, a sense-preserving harmonic function f=h+g¯f=h+\bar{g} is said to be KK-quasiregular if its complex dilatation ω=g/h\omega=g^{\prime}/h^{\prime} satisfies the inequality

|ω(z)|k<1(z𝔻),|\omega(z)|\leq k<1\quad(z\in{\mathbb{D}}),

where

kK1K+1.k\coloneqq\frac{K-1}{K+1}.

We say that ff is KK-quasiconformal if ff is KK-quasiregular and homeomorphic in 𝔻{\mathbb{D}}. One can find the HpH^{p}-theory for quasiregular and quasiconformal mappings, for example, in the survey of Astala and Koskela [2] and the paper of Adamowicz and González [1]. Some recent results on the Hardy space of harmonic quasiconformal mappings are obtained in [18].

1.2. Classical results of Riesz, Kolmogorov and Zygmund

Given a real-valued harmonic function uu in 𝔻{\mathbb{D}}, let vv be its harmonic conjugate, normalized by the condition v(0)=0v(0)=0. It is a natural question that if uu has a certain property, whether so does vv. In the context of boundary behaviour, M. Riesz established this principle in a remarkably precise form:

Theorem A.

[7, Theorem 4.1] If uhpu\in h^{p} for some pp, 1<p<1<p<\infty, then its harmonic conjugate vv is also of class hph^{p}. Furthermore, there is a constant ApA_{p}, depending only on pp, such that

Mp(r,v)ApMp(r,u),M_{p}(r,v)\leq A_{p}\,M_{p}(r,u),

for all uhpu\in h^{p}.

Curiously, the theorem fails for the cases p=1p=1 and p=p=\infty, examples are given in [7, p. 56]. Although the harmonic conjugate of an h1h^{1} function need not be in h1h^{1}, Kolmogorov crucially pointed out that it does belong to hph^{p} for all p<1p<1. The result, in a stronger form, is as follows.

Theorem B.

[7, Theorem 4.2] If uh1u\in h^{1}, then its harmonic conjugate vhpv\in h^{p} for all p<1p<1. Furthermore, there is a constant BpB_{p}, depending only on pp, such that

Mp(r,v)BpM1(r,u),M_{p}(r,v)\leq B_{p}\,M_{1}(r,u),

for all uh1u\in h^{1}.

It follows from the theorems of Riesz and Kolmogorov that the condition uh1u\in h^{1} is not enough to ensure vh1v\in h^{1}, but the slightly stronger hypothesis uhpu\in h^{p} (for some p>1p>1) is sufficient. Therefore, one naturally asks for the “minimal” growth restriction on uu which should imply vh1v\in h^{1}. As established by Zygmund, such a condition is the boundedness of the integrals

02π|u(reiθ)|log+|u(reiθ)|dθ,\int_{0}^{2\pi}|u(r{e^{i\theta}})|\log^{+}|u(r{e^{i\theta}})|\,d\theta,

where log+x=max{logx, 0}\log^{+}x=\max\{\log x,\,0\}. The class of such harmonic functions uu in 𝔻{\mathbb{D}} is denoted by 𝐡log+𝐡{{\bf h}\log^{+}{\bf h}}. The following is the precise statement of Zygmund’s theorem.

Theorem C.

[7, Theorem 4.3] If u𝐡log+𝐡u\in{{\bf h}\log^{+}{\bf h}}, then its conjugate vv is of class h1h^{1}, and

M1(r,v)12π02π|u(reiθ)|log+|u(reiθ)|dθ+3e.M_{1}(r,v)\leq\frac{1}{2\pi}\int_{0}^{2\pi}|u(r{e^{i\theta}})|\log^{+}|u(r{e^{i\theta}})|\,d\theta+3e.

It should be noted that hp𝐡log+𝐡h^{p}\subset{{\bf h}\log^{+}{\bf h}} for every p>1p>1, so that Zygmund’s theorem is indeed best possible. We refer the reader to the paper of Pichorides [17] for the sharp constants in the Riesz, Kolmogorov, and Zygmund theorems.

1.3. Recent developments for harmonic quasiregular mappings

Suppose f=u+ivf=u+iv is a harmonic function in 𝔻{\mathbb{D}}. If uhpu\in h^{p}, p>1p>1, then the imaginary part vv does not necessarily belong to the space hph^{p}, i.e., Riesz theorem is not true for harmonic functions. The question arises: Under what additional condition(s) does a harmonic analogue of the Riesz theorem hold? Recently, Liu and Zhu established in their breakthrough paper [12] that such a condition is the quasiregularity of ff.

Theorem D.

[12] Let f=u+ivf=u+iv be a harmonic KK-quasiregular mapping in 𝔻{\mathbb{D}} such that u0u\geq 0 and v(0)=0v(0)=0. If uhpu\in h^{p} for some p(1,2]p\in(1,2], then also vv is in hph^{p}. Furthermore, there is a constant C(K,p)C(K,p), depending only on KK and pp, such that

Mp(r,v)C(K,p)Mp(r,u).M_{p}(r,v)\leq C(K,p)M_{p}(r,u).

Moreover, if K=1K=1, i.e., ff is analytic, then C(1,p)C(1,p) coincides with the optimal constant in the Riesz theorem.

It is important to note that the assumption “ff is harmonic in 𝔻{\mathbb{D}}” cannot be removed. There exists a quasiconformal mapping for which the first coordinate function belongs to the Hardy space, but the whole function does not (see [2, p. 36]).

Later, Kalaj [10] refined Theorem D by removing the condition u0u\geq 0. Furthermore, he proved a Kolmogorov type theorem for harmonic quasiregular mappings in 𝔻{\mathbb{D}}, thereby making considerable progress in this direction.

Theorem E.

[10] Suppose f=u+ivf=u+iv is a harmonic KK-quasiregular mapping in 𝔻{\mathbb{D}} such that u>0u>0 and v(0)=0v(0)=0. Then uh1u\in h^{1} and vhpv\in h^{p} for every p<1p<1. Furthermore, they satisfy the inequalities

Mpp(r,v)sec(pπ2)(K2M1p(r,u)(K21)Mpp(r,u))M_{p}^{p}(r,v)\leq\sec\left(\frac{p\pi}{2}\right)\left(K^{2}M_{1}^{p}(r,u)-(K^{2}-1)M_{p}^{p}(r,u)\right)

and

(2K2)M1p(r,u)(2K2)Mpp(r,u)+cos(pπ2)Mpp(r,v).(2-K^{2})M_{1}^{p}(r,u)\leq(2-K^{2})M_{p}^{p}(r,u)+\cos\left(\frac{p\pi}{2}\right)M_{p}^{p}(r,v).

All the constants are asymptotically sharp as K1K\to 1.

Kalaj also points out that the second inequality is new even for analytic functions. Several other Riesz type theorems for harmonic functions can be found in the recent works of Kalaj [8, 9], Chen and Hamada [3], Melentijević [13], and Melentijević and Marković [14]. Furthermore, Chen and Huang [4] have proved Riesz theorem for pluriharmonic functions where the assumption u0u\geq 0 is dropped as well.

1.4. Main results

In this paper, we make further advance along this line and establish a Zygmund type theorem for harmonic quasiregular mappings. For a harmonic KK-quasiregular mapping f=u+ivf=u+iv in 𝔻{\mathbb{D}}, we find out the growth condition on uu that would imply fh1f\in h^{1} (and vh1v\in h^{1}). The following is the central result of this paper.

Theorem 1.

Let f=u+ivf=u+iv be a harmonic KK-quasiregular mapping in 𝔻{\mathbb{D}} such that u1u\geq 1 (or u1u\leq-1) and v(0)=0v(0)=0. If u𝐡log+𝐡u\in{{\bf h}\log^{+}{\bf h}}, then fh1f\in h^{1} and

M1(r,f)K22π02π|u(reiθ)|log+|u(reiθ)|dθ+|u(0)|[1K2log|u(0)|].M_{1}(r,f)\leq\frac{K^{2}}{2\pi}\int_{0}^{2\pi}|u(r{e^{i\theta}})|\log^{+}|u(r{e^{i\theta}})|\,d\theta+|u(0)|\left[1-K^{2}\log|u(0)|\right]. (1)

Consequently, vh1v\in h^{1} and M1(r,v)M_{1}(r,v) satisfies inequality (1).

Remark 1.

The conclusion fh1f\in h^{1} is true under the slightly weaker hypothesis that uCu\geq C for some constant CC, since adding a constant to uu does not affect vv or the quasiregularity of ff, although we do not know if the estimate (1) still holds in this case. However, we suspect that the assumption u1u\geq 1 (or uCu\geq C) may actually be redundant, but a proof remains elusive.

Theorem 1 is the best possible, in the sense that the growth restriction imposed on uu cannot be weakened. This can be seen from the following partial converse, which surprisingly does not require the quasiregularity hypothesis.

Theorem 2.

Let f=h+g¯f=h+\bar{g} be a harmonic function in 𝔻{\mathbb{D}} such that fC\real f\geq C for some constant CC, and Imh{\operatorname{Im}\,}h is non-vanishing. If fh1f\in h^{1}, then f𝐡log+𝐡\real f\in{{\bf h}\log^{+}{\bf h}}.

The proofs of the main theorems are presented in Section 2. In Section 3, we study the harmonic analogue of a well-known result of Hardy and Littlewood as an application of the classical Riesz and Kolmogorov theorems.

2. Proof of the main theorems

2.1. Proof of Theorem 1

Our technique is based on a method of Green’s theorem, which has also been used in [10] and [12].

First assume that u1u\geq 1, so that log+|u(reiθ)|=logu(reiθ)\log^{+}|u(r{e^{i\theta}})|=\log u(r{e^{i\theta}}). Let us write f=h+g¯f=h+\bar{g}, and let ω(z)=g(z)/h(z)\omega(z)=g^{\prime}(z)/h^{\prime}(z) be its complex dilatation. An elementary computation gives

Δ{ulogu}\displaystyle\Delta\{u\log u\} =(2x2+2y2){ulogu}=ux2+uy2u.\displaystyle=\left(\frac{\partial^{2}}{\partial x^{2}}+\frac{\partial^{2}}{\partial y^{2}}\right)\{u\log u\}=\frac{u_{x}^{2}+u_{y}^{2}}{u}.

Suppose F=h+gF=h+g. Then

F=h+g=f=u,\real F=\real h+\real g=\real f=u,

so that

F=h+g=uxiuy.F^{\prime}=h^{\prime}+g^{\prime}=u_{x}-i\,u_{y}.

This implies

ux2+uy2=|h+g|2.u_{x}^{2}+u_{y}^{2}=|h^{\prime}+g^{\prime}|^{2}.

Therefore, we have

Δ{ulogu}=|h|2|1+ω|2u(1k)2|h|2u,\displaystyle\Delta\{u\log u\}=\frac{|h^{\prime}|^{2}\,|1+\omega|^{2}}{u}\geq(1-k)^{2}\,\frac{|h^{\prime}|^{2}}{u},

as |ω(z)|k|\omega(z)|\leq k. On the other hand, one finds

Δ{|f|}\displaystyle\Delta\{|f|\} =42z¯z{|f|}\displaystyle=4\,\frac{\partial^{2}}{\partial\bar{z}\partial z}\{|f|\}
=2z¯(h(h¯+g)+(h+g¯)g|f|)\displaystyle=2\,\frac{\partial}{\partial\bar{z}}\left(\frac{h^{\prime}(\bar{h}+g)+(h+\bar{g})g^{\prime}}{|f|}\right)
=2|f|(|h|2+|g|2)(h(h¯+g)+(h+g¯)g)g¯(h¯+g)+(h+g¯)h¯2|f||f|2\displaystyle=2\,\frac{|f|(|h^{\prime}|^{2}+|g^{\prime}|^{2})-\left(h^{\prime}(\bar{h}+g)+(h+\bar{g})g^{\prime}\right)\frac{\overline{g^{\prime}}(\bar{h}+g)+(h+\bar{g})\overline{h^{\prime}}}{2|f|}}{|f|^{2}}
=2|f|2(|h|2+|g|2)(|f|2(|h|2+|g|2)+2(h¯gf2))|f|3\displaystyle=\frac{2\,|f|^{2}(|h^{\prime}|^{2}+|g^{\prime}|^{2})-\left(|f|^{2}(|h^{\prime}|^{2}+|g^{\prime}|^{2})+2\real(\overline{h^{\prime}}g^{\prime}f^{2})\right)}{|f|^{3}}
=|f|2(|h|2+|g|2)2(h¯gf2)|f|3.\displaystyle=\frac{|f|^{2}(|h^{\prime}|^{2}+|g^{\prime}|^{2})-2\real(\overline{h^{\prime}}g^{\prime}f^{2})}{|f|^{3}}.

Thus, it follows that

Δ{|f|}\displaystyle\Delta\{|f|\} |f|2(|h|2+|g|2)+2|h||g||f|2|f|3\displaystyle\leq\frac{|f|^{2}(|h^{\prime}|^{2}+|g^{\prime}|^{2})+2|h^{\prime}||g^{\prime}||f|^{2}}{|f|^{3}}
=(|h|+|g|)2|f|=(1+|ω|)2|h|2|f|\displaystyle=\frac{(|h^{\prime}|+|g^{\prime}|)^{2}}{|f|}=(1+|\omega|)^{2}\,\frac{|h^{\prime}|^{2}}{|f|}
(1+k)2|h|2|f|.\displaystyle\leq(1+k)^{2}\,\frac{|h^{\prime}|^{2}}{|f|}.

These inequalities imply

Δ{|f|}Δ{ulogu}(1+k1k)2u|f|(1+k1k)2=K2,\displaystyle\frac{\Delta\{|f|\}}{\Delta\{u\log u\}}\leq\left(\frac{1+k}{1-k}\right)^{2}\frac{u}{|f|}\leq\left(\frac{1+k}{1-k}\right)^{2}=K^{2},

and therefore,

Δ{|f|}K2Δ{ulogu}.\Delta\{|f|\}\leq K^{2}\Delta\{u\log u\}. (2)

The next step is to apply Green’s theorem in the form

r02πφr𝑑θ=|z|rΔφ𝑑x𝑑y,r\int_{0}^{2\pi}\frac{\partial\varphi}{\partial r}\,d\theta=\iint\limits_{|z|\leq r}\Delta\varphi\,dx\,dy,

where z=x+iyz=x+iy. Then, it follows from (2) that

ddr02π|f(reiθ)|𝑑θK2ddr02πu(reiθ)logu(reiθ)𝑑θ.\frac{d}{dr}\int_{0}^{2\pi}|f(r{e^{i\theta}})|\,d\theta\leq K^{2}\frac{d}{dr}\int_{0}^{2\pi}u(r{e^{i\theta}})\log u(r{e^{i\theta}})\,d\theta.

Now, integrating from 0 to rr and recalling that v(0)=0v(0)=0, we get

02π|f(reiθ)|𝑑θ2πu(0)K2[02πu(reiθ)logu(reiθ)𝑑θ2πu(0)logu(0)].\int_{0}^{2\pi}|f(r{e^{i\theta}})|\,d\theta-2\pi u(0)\leq K^{2}\left[\int_{0}^{2\pi}u(r{e^{i\theta}})\log u(r{e^{i\theta}})\,d\theta-2\pi u(0)\log u(0)\right].

Therefore

M1(r,f)K22π02πu(reiθ)logu(reiθ)𝑑θ+u(0)[1K2logu(0)],M_{1}(r,f)\leq\frac{K^{2}}{2\pi}\int_{0}^{2\pi}u(r{e^{i\theta}})\log u(r{e^{i\theta}})\,d\theta+u(0)\left[1-K^{2}\log u(0)\right],

which is the desired inequality.

For the case u1u\leq-1, let us write F=eiπfF=e^{i\pi}f. We see that FF is harmonic KK-quasiregular in 𝔻{\mathbb{D}} with uF=F=u1u_{F}=\real F=-u\geq 1, and vF=ImF=vv_{F}={\operatorname{Im}\,}F=-v satisfying vF(0)=0v_{F}(0)=0. Then, from what we have already proved, it follows that

M1(r,F)K22π02πuF(reiθ)loguF(reiθ)𝑑θ+uF(0)[1K2loguF(0)].M_{1}(r,F)\leq\frac{K^{2}}{2\pi}\int_{0}^{2\pi}u_{F}(r{e^{i\theta}})\log u_{F}(r{e^{i\theta}})\,d\theta+u_{F}(0)\left[1-K^{2}\log u_{F}(0)\right].

Since M1(r,f)=M1(r,F)M_{1}(r,f)=M_{1}(r,F) and uF=|u|u_{F}=|u|, the proof is complete.∎

The proof of the converse (Theorem 2) relies on the following lemma, which is of some independent interest.

Lemma 1.

Let f=h+g¯f=h+\bar{g} be a harmonic function in 𝔻{\mathbb{D}} such that Imh{\operatorname{Im}\,}h is non-vanishing. If fh1f\in h^{1}, then the analytic function F=h+gF=h+g is in H1H^{1}.

Proof.

Suppose fh1f\in h^{1}, and assume that h=a+ibh=a+ib and g=c+idg=c+id. Then

f=F=a+ch1,\real f=\real F=a+c\,\in h^{1},

and

Imf=bdh1,{\operatorname{Im}\,}f=b-d\,\in h^{1},

since f,Imf|f|\real f,\,{\operatorname{Im}\,}f\leq|f|. We recall [7, Theorem 1.1] that every real-valued harmonic function in h1h^{1} can be expressed as the difference of two positive harmonic functions. Let us write

bd=u1u2,b-d=u_{1}-u_{2},

where u1u_{1} and u2u_{2} are positive harmonic functions in 𝔻{\mathbb{D}}. Since Imh=b{\operatorname{Im}\,}h=b is continuous and non-vanishing, either b>0b>0 or b<0b<0 in the whole disk 𝔻{\mathbb{D}}. If b>0b>0, then

d=(b+u2)u1,d=(b+u_{2})-u_{1},

which is again the difference of two positive harmonic functions, and hence is in h1h^{1}. Similarly, if b<0b<0, we have

d=u2(u1b)h1.d=u_{2}-(u_{1}-b)\in h^{1}.

In either case,

ImF=b+d=2d+(u1u2)h1.{\operatorname{Im}\,}F=b+d=2d+(u_{1}-u_{2})\in h^{1}.

Therefore, FH1F\in H^{1}. This completes the proof. ∎

Remark 2.
  1. (i)

    The usual normalisation g(0)=0g(0)=0 implies Img(0)=0{\operatorname{Im}\,}g(0)=0. However, if we drop the condition g(0)=0g(0)=0, then Lemma 1 is true if either Imh{\operatorname{Im}\,}h or Img{\operatorname{Im}\,}g is non-vanishing, and the proof is essentially the same.

  2. (ii)

    If fh1f\in h^{1}, it seems that the corresponding FF need not be in H1H^{1}. In view of this, the hypothesis Imh0{\operatorname{Im}\,}h\neq 0 (or Img0{\operatorname{Im}\,}g\neq 0) seems to be a surprisingly simple assumption that ensures FH1F\in H^{1}. A similar condition appears in Kalaj’s refinement of the Liu-Zhu result (Theorem 2.2 of [10]), where it is assumed that f(𝔻)(,0)=f({\mathbb{D}})\cap(-\infty,0)=\emptyset.

2.2. Proof of Theorem 2

Suppose f=u+ivf=u+iv. Without any loss of generality, we may assume that u1u\geq 1, as adding a constant to uu does not affect vv or the condition fh1f\in h^{1}. Since Imh{\operatorname{Im}\,}h is non-vanishing, by Lemma 1 we have F=h+gH1F=h+g\in H^{1}. Let us write

F(z)=ReiΦ(|Φ|π2),F(z)=Re^{i\Phi}\quad\left(|\Phi|\leq\frac{\pi}{2}\right),

where z=reiθz=r{e^{i\theta}}, R=R(r,θ)R=R(r,\theta) and Φ=Φ(r,θ)\Phi=\Phi(r,\theta). Then

F=f=u=Rcos(Φ).\real F=\real f=u=R\cos(\Phi).

Now, the function F(z)logF(z)F(z)\log F(z) is analytic in 𝔻{\mathbb{D}}, with

[F(z)logF(z)]=Rcos(Φ)log(R)ΦRsin(Φ).\real[F(z)\log F(z)]=R\cos(\Phi)\log(R)-\Phi R\sin(\Phi).

We apply mean value theorem to this real part to obtain

02π[Rcos(Φ)log(R)ΦRsin(Φ)]𝑑θ\displaystyle\int_{0}^{2\pi}\left[R\cos(\Phi)\log(R)-\Phi R\sin(\Phi)\right]\,d\theta =2π[F(0)logF(0)]\displaystyle=2\pi\real\left[F(0)\log F(0)\right]
=2π[h(0)logh(0)],\displaystyle=2\pi\real\left[h(0)\log h(0)\right],

since g(0)=0g(0)=0. It is to be noted that Imh0{\operatorname{Im}\,}h\neq 0 implies h0h\neq 0, so that logh(0)\log h(0) is well-defined. Thus we have

02πRcos(Φ)log(R)𝑑θ=02πΦRsin(Φ)𝑑θ+2π[h(0)logh(0)].\int_{0}^{2\pi}R\cos(\Phi)\log(R)\,d\theta=\int_{0}^{2\pi}\Phi R\sin(\Phi)\,d\theta+2\pi\real\left[h(0)\log h(0)\right].

It follows that

02πu(reiθ)logu(reiθ)𝑑θ\displaystyle\int_{0}^{2\pi}u(r{e^{i\theta}})\log u(r{e^{i\theta}})\,d\theta 02πRcos(Φ)log(R)𝑑θ\displaystyle\leq\int_{0}^{2\pi}R\cos(\Phi)\log(R)\,d\theta
|Φ|02πR𝑑θ+2π|[h(0)logh(0)]|\displaystyle\leq|\Phi|\int_{0}^{2\pi}R\,d\theta+2\pi\left|\real\left[h(0)\log h(0)\right]\right|
π202π|F(reiθ)|𝑑θ+2π|h(0)logh(0)|,\displaystyle\leq\frac{\pi}{2}\int_{0}^{2\pi}|F(r{e^{i\theta}})|\,d\theta+2\pi|h(0)\log h(0)|,

which is finite since FH1F\in H^{1}. Therefore, u𝐡log+𝐡u\in{{\bf h}\log^{+}{\bf h}} and the proof is complete.∎

Remark 3.

In Theorem 1, we assume that v(0)=0v(0)=0, while in Theorem 2, the condition Imh0{\operatorname{Im}\,}h\neq 0 implies v(0)0v(0)\neq 0 (as g(0)=0g(0)=0). This is not really an inconsistency, since adding a constant to vv (therefore, to ff) does not affect uu or the condition fh1f\in h^{1}. So Theorems 1 and 2 are simultaneously valid.

3. An application of the theorems of Riesz and Kolmogorov

Here we discuss how the classical theorems of Riesz and Kolmogorov can be applied to obtain numerous results for planar harmonic Hardy spaces. We give an example of a famous theorem of Hardy and Littlewood, that seems to be unknown for complex-valued harmonic functions.

Let ff be an analytic function in 𝔻{\mathbb{D}} with f(z)=anznf(z)=\sum a_{n}z^{n}. If fHpf\in H^{p} for some pp, it is interesting to understand the behaviour of the Taylor coefficients ana_{n}. We should note that it makes sense only to talk about the eventual behaviour as nn\to\infty, since any finite number of coefficients can be changed arbitrarily without affecting the fact fHpf\in H^{p}. The following theorem of Hardy and Littlewood describes the Taylor coefficients of an HpH^{p}-function in a precise form.

Theorem F.

[7, Theorem 6.2] Let 0<p20<p\leq 2 and

f(z)=n=0anznHp.f(z)=\sum_{n=0}^{\infty}a_{n}\,z^{n}\in H^{p}.

Then we have

n=0(n+1)p2|an|p<.\sum_{n=0}^{\infty}(n+1)^{p-2}|a_{n}|^{p}<\infty.

The converse of this theorem is false for 0<p<20<p<2 but holds for the indices p2p\geq 2, and can be obtained from Theorem F via a duality argument. However, Pavlović [16] showed that if the sequence {an}\{a_{n}\} decreases to 0, then the converse is indeed true for all p1p\geq 1.

Theorem G.

[16] Suppose 1p<1\leq p<\infty. If an0a_{n}\downarrow 0 as nn\to\infty, then the function f(z)=anznf(z)=\sum a_{n}z^{n} is in HpH^{p} if and only if

n=0(n+1)p2anp<.\sum_{n=0}^{\infty}(n+1)^{p-2}a_{n}^{p}<\infty.

Similar problem for planar harmonic functions is not known, but can be easily deduced from the conjugate function theorems. It is worth mentioning that the case p1p\leq 1 is usually difficult to handle for harmonic functions. Nevertheless, we can produce a slightly different result that deals with this case (part (ii) of Theorem 3).

Let f=h+g¯f=h+\bar{g} be a harmonic function in 𝔻{\mathbb{D}} with

h(z)=n=0anzn,andg(z)=n=0bnzn.h(z)=\sum_{n=0}^{\infty}a_{n}\,z^{n},\quad\text{and}\quad g(z)=\sum_{n=0}^{\infty}b_{n}\,z^{n}. (3)

We should note that b0=0b_{0}=0 as g(0)=0g(0)=0, but for our forthcoming discussion, it is convenient to write the sum as n=0\sum_{n=0}^{\infty}.

Theorem 3.

If ff is a harmonic functions in 𝔻{\mathbb{D}} with the series representation (3), then the following statements hold:

  1. (i)

    If fhpf\in h^{p} for some p(1,2]p\in(1,2], then

    n=0(n+1)p2(|an|p+|bn|p)<.\sum_{n=0}^{\infty}(n+1)^{p-2}\left(|a_{n}|^{p}+|b_{n}|^{p}\right)<\infty.

    Furthermore, suppose an0a_{n}\downarrow 0 and bn0b_{n}\downarrow 0. Then fhpf\in h^{p} (p>1)(p>1) if and only if

    n=0(n+1)p2(anp+bnp)<.\sum_{n=0}^{\infty}(n+1)^{p-2}\left(a_{n}^{p}+b_{n}^{p}\right)<\infty.
  2. (ii)

    If fh1f\in h^{1}, then

    n=0(n+1)p2(|an|p+|bn|p)<\sum_{n=0}^{\infty}(n+1)^{p-2}\left(|a_{n}|^{p}+|b_{n}|^{p}\right)<\infty

    for every p<1p<1.

  3. (iii)

    Suppose Imh0{\operatorname{Im}\,}h\neq 0 and an0a_{n}\downarrow 0, bn0b_{n}\downarrow 0. If fh1f\in h^{1}, then

    n=0an+bnn+1<.\sum_{n=0}^{\infty}\frac{a_{n}+b_{n}}{n+1}<\infty.
Proof.

(i) The main step is to show that h,gHph,g\in H^{p}. This is contained in Lemma 2.2 of [12], but we give a slightly different proof based on ideas from the more general result [5, Lemma 1]. Let us write f=u+ivf=u+iv, where uu and vv are real-valued harmonic functions in 𝔻{\mathbb{D}}. Suppose u1u_{1} is the harmonic conjugate of uu normalised by u1(0)=v(0)u_{1}(0)=v(0), while v1v_{1} is the harmonic conjugate of vv with v1(0)=u(0)v_{1}(0)=-u(0). Assume that U=u+iu1U=u+iu_{1} and V=v+iv1V=v+iv_{1}. Then UU, VV are analytic in 𝔻{\mathbb{D}}, and

f=U+iV=12(U+U¯)+i2(V+V¯)=12(U+iV)+12(UiV)¯.f=\real U+i\,\real V=\frac{1}{2}\left(U+\overline{U}\right)+\frac{i}{2}\left(V+\overline{V}\right)=\frac{1}{2}\left(U+iV\right)+\frac{1}{2}\overline{\left(U-iV\right)}.

Moreover, the above normalisations suggest that

(UiV2)(0)=0,\left(\frac{U-iV}{2}\right)(0)=0,

so that the uniqueness of hh and gg implies

h=12(U+iV)andg=12(UiV).h=\frac{1}{2}(U+iV)\quad\text{and}\quad g=\frac{1}{2}(U-iV).

As |u|,|v||f||u|,|v|\leq|f|, clearly u,vhpu,v\in h^{p}. In view of Riesz theorem (Theorem A), we have u1,v1hpu_{1},v_{1}\in h^{p}. It follows that UU, VV are in HpH^{p}, therefore so are hh and gg.

Now, we apply Theorem F to the functions hh and gg to find

n=0(n+1)p2|an|p<andn=0(n+1)p2|bn|p<,\sum_{n=0}^{\infty}(n+1)^{p-2}|a_{n}|^{p}<\infty\quad\text{and}\quad\sum_{n=0}^{\infty}(n+1)^{p-2}|b_{n}|^{p}<\infty,

for 1<p21<p\leq 2. Combining these two inequalities, we get the desired result.

The proof for the second part of the theorem is identical, except appealing to Theorem G. The details are omitted.

(ii) The reasoning is similar to that in part (i), except for the fact that we need to use Kolmogorov theorem (Theorem B), instead of Riesz theorem. If fh1f\in h^{1}, using an analogous argument as above we find that h,gh,g are in HpH^{p} for every p<1p<1. Therefore, Theorem F gives

n=0(n+1)p2(|an|p+|bn|p)<,\sum_{n=0}^{\infty}(n+1)^{p-2}\left(|a_{n}|^{p}+|b_{n}|^{p}\right)<\infty,

for 0<p<10<p<1.

(iii) If fh1f\in h^{1} and Imh{\operatorname{Im}\,}h is non-vanishing, Lemma 1 shows that the analytic function FH1F\in H^{1}, where

F(z)=h(z)+g(z)=n=0(an+bn)zn.F(z)=h(z)+g(z)=\sum_{n=0}^{\infty}(a_{n}+b_{n})\,z^{n}.

If ana_{n} and bnb_{n} decrease to 0, so does an+bna_{n}+b_{n}. Therefore, it follows from Theorem G, for p=1p=1, that

n=0an+bnn+1<.\sum_{n=0}^{\infty}\frac{a_{n}+b_{n}}{n+1}<\infty.

The proof of the theorem is complete. ∎

Remark 4.

It is obvious that the converse of Theorem 3 (iii) holds even without the condition Imh0{\operatorname{Im}\,}h\neq 0. Indeed, if

n=0an+bnn+1<,\sum_{n=0}^{\infty}\frac{a_{n}+b_{n}}{n+1}<\infty,

one sees from Theorem G that h,gH1h,\,g\in H^{1}. Thus, fh1f\in h^{1}.

Acknowledgements

The first-named author thanks Petar Melentijević for helpful discussions on the space h1h^{1} that led to the classical theory of Zygmund. The research was partially supported by the Natural Science Foundation of Guangdong Province (Grant no. 2024A1515010467) and the Li Ka Shing Foundation STU-GTIIT Joint Research Grant (Grant no. 2024LKSFG06).

References

  • [1] T. Adamowicz and M. J. González. Hardy spaces for quasiregular mappings and composition operators. J. Geom. Anal., 31(11):11417–11427, 2021.
  • [2] K. Astala and P. Koskela. HpH^{p}-theory for quasiconformal mappings. Pure Appl. Math. Q., 7(1):19–50, 2011.
  • [3] S. Chen and H. Hamada. On (Fejér-)Riesz type inequalities, Hardy-Littlewood type theorems and smooth moduli. Math. Z., 305(4):Paper No. 64, 30, 2023.
  • [4] S. Chen and M. Huang. Riesz type theorems for κ\kappa-pluriharmonic mappings, invariant harmonic quasiregular mappings and harmonic quasiregular mappings. arXiv:2310.15452, 2023.
  • [5] S. Das and A. Sairam Kaliraj. Growth of harmonic mappings and Baernstein type inequalities. Potential Anal., 60(3):1121–1137, 2024.
  • [6] P. Duren. Harmonic mappings in the plane, volume 156 of Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge, 2004.
  • [7] P. L. Duren. Theory of HpH^{p} spaces. Pure and Applied Mathematics, Vol. 38. Academic Press, New York-London, 1970.
  • [8] D. Kalaj. On Riesz type inequalities for harmonic mappings on the unit disk. Trans. Amer. Math. Soc., 372(6):4031–4051, 2019.
  • [9] D. Kalaj. On M. Riesz conjugate function theorem for harmonic functions. Potential Anal., 2024.
  • [10] D. Kalaj. Riesz and Kolmogorov inequality for harmonic quasiregular mappings. J. Math. Anal. Appl., 542(1):Paper No. 128767, 15, 2025.
  • [11] P. Koosis. Introduction to HpH_{p} spaces, volume 115 of Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge, second edition, 1998.
  • [12] J. Liu and J.-F. Zhu. Riesz conjugate functions theorem for harmonic quasiconformal mappings. Adv. Math., 434:Paper No. 109321, 27, 2023.
  • [13] P. Melentijević. Hollenbeck-Verbitsky conjecture on best constant inequalities for analytic and co-analytic projections. Math. Ann., 388(4):4405–4448, 2024.
  • [14] P. Melentijević and M. Marković. Best constants in inequalities involving analytic and co-analytic projections and Riesz’s theorem in various function spaces. Potential Anal., 59(4):1599–1620, 2023.
  • [15] M. Pavlović. Introduction to function spaces on the disk, volume 20 of Posebna Izdanja [Special Editions]. Matematički Institut SANU, Belgrade, 2004.
  • [16] M. Pavlović. Analytic functions with decreasing coefficients and Hardy and Bloch spaces. Proc. Edinb. Math. Soc. (2), 56(2):623–635, 2013.
  • [17] S. K. Pichorides. On the best values of the constants in the theorems of M. Riesz, Zygmund and Kolmogorov. Studia Math., 44:165–179, 1972.
  • [18] Z.-G. Wang, X.-Y. Wang, A. Rasila, and J.-L. Qiu. On a problem of Pavlović involving harmonic quasiconformal mappings. arXiv:2405.19852, 2024.