1 Preliminaries and Main Results

Let \({\mathbb {C}} \cong {\mathbb {R}}^{2}\) be the complex plane. For \(a\in {\mathbb {C}}\) and \(r>0\), we let \({\mathbb D}(a,r)=\{z:\, |z-a|<r\}\) so that \({\mathbb {D}}_r:={\mathbb {D}}(0,r)\) and thus, \({\mathbb {D}}:={\mathbb D}_1\) denotes the open unit disk in the complex plane \({\mathbb {C}}\). Let \({\mathbb {T}}=\partial {\mathbb {D}}\) be the boundary of \({\mathbb {D}}\). We denote by \({\mathcal {C}}^{m}(\Omega )\) the set of all complex-valued m-times continuously differentiable functions from \(\Omega \) into \({\mathbb {C}}\), where \(\Omega \) is a subset of \({\mathbb {C}}\) and \(m\in {\mathbb {N}}_0:={\mathbb {N}}\cup \{0\}\). In particular, let \({\mathcal {C}}(\Omega ):={\mathcal {C}}^{0}(\Omega )\), the set of all continuous functions defined in \(\Omega \).

For a real \(2\times 2\) matrix A, we use the matrix norm \(\Vert A\Vert =\sup \{|Az|:\,|z|=1\}\) and the matrix function \(\lambda (A)=\inf \{|Az|:\,|z|=1\}\). For \(z=x+iy\in {\mathbb {C}}\), the formal derivative of the complex-valued functions \(f=u+iv\) is given by

$$\begin{aligned} D_{f}=\left( \begin{array}{cccc} \displaystyle u_{x}\;~~ u_{y}\\ \displaystyle v_{x}\;~~ v_{y} \end{array}\right) , \end{aligned}$$

so that

$$\begin{aligned} \Vert D_{f}\Vert =|f_{z}|+|f_{\overline{z}}| \quad \text{ and }\quad \lambda (D_{f})=\big | |f_{z}|-|f_{\overline{z}}|\big |, \end{aligned}$$

where

$$\begin{aligned} f_{z}=\frac{\partial f}{\partial z}=\frac{1}{2}\big ( f_x-if_y\big )\quad \text{ and }\quad f_{\overline{z}}=\frac{\partial f}{\partial \overline{z}}=\frac{1}{2}\big (f_x+if_y\big ). \end{aligned}$$

We use

$$\begin{aligned} J_{f}:=\det D_{f} =|f_{z}|^{2}-|f_{\overline{z}}|^{2} \end{aligned}$$

to denote the Jacobian of f and

$$\begin{aligned} \Delta f:=\frac{\partial ^{2}f}{\partial x^{2}}+\frac{\partial ^{2}f}{\partial y^{2}}=4f_{z \overline{z}} \end{aligned}$$

is the Laplacian of f.

For \(t\in {\mathbb {R}}\) and \(z, w\in {\mathbb {D}}\) with \(z\ne w\) and \(|z|+|w|\ne 0\), let

$$\begin{aligned} G(z,w)=\log \left| \frac{1-z\overline{w}}{z-w}\right| \quad \text{ and } \quad P(z,e^{it})=\frac{1-|z|^{2}}{|1-ze^{-it}|^{2}} \end{aligned}$$

be the Green function and Poisson kernel, respectively.

Let \(\psi :~{\mathbb {T}}\rightarrow {\mathbb {C}}\) be a bounded integrable function and let \(g\in {\mathcal {C}}({\mathbb {D}})\). For \(z\in {\mathbb {D}}\), the solution to the Poisson’s equation

$$\begin{aligned} \Delta f(z)=g(z) \end{aligned}$$

satisfying the boundary condition \(f|_{{\mathbb {T}}}=\psi \in L^{1}({\mathbb {T}})\) is given by

$$\begin{aligned} f(z)={\mathcal {P}}_{\psi }(z)-{\mathcal {G}}_{g}(z), \end{aligned}$$
(1.1)

where

$$\begin{aligned} {\mathcal {G}}_{g}(z)=\frac{1}{2\pi }\int _{{\mathbb {D}}}G(z,w)g(w)dA(w),~ ~{\mathcal {P}}_{\psi }(z)=\frac{1}{2\pi }\int _{0}^{2\pi }P(z,e^{it})\psi (e^{it})dt,\nonumber \\ \end{aligned}$$
(1.2)

and dA(w) denotes the Lebesgue measure in \({\mathbb {D}}\). It is well known that if \(\psi \) and g are continuous in \({\mathbb {T}}\) and in \(\overline{{\mathbb {D}}}\), respectively, then \(f={\mathcal {P}}_{\psi }-{\mathcal {G}}_{g}\) has a continuous extension \(\tilde{f}\) to the boundary, and \(\tilde{f}=\psi \) in \({\mathbb {T}}\) (see [18, pp. 118–120] and [2, 19, 20, 22]).

Heinz in his classical paper [17] proved the following result, which is called the Schwarz Lemma of complex-valued harmonic functions: If f is a complex-valued harmonic function from \({\mathbb {D}}\) into itself satisfying the condition \(f(0)=0\), then, for \(z\in {\mathbb {D}}\),

$$\begin{aligned} |f(z)|\le \frac{4}{\pi }\arctan |z|. \end{aligned}$$
(1.3)

Later, Pavlović [30, Theorem 3.6.1] removed the assumption \(f(0)=0\) and improved (1.3) into the following sharp form

$$\begin{aligned} \left| f(z)-\frac{1-|z|^{2}}{1+|z|^{2}}f(0)\right| \le \frac{4}{\pi }\arctan |z|, \end{aligned}$$
(1.4)

where f is a complex-valued harmonic function from \({\mathbb {D}}\) to itself.

The first aim of this paper is to extend (1.4) into mappings satisfying the Poisson’s equation as follows.

Theorem 1

For a given \(g\in {\mathcal {C}}(\overline{{\mathbb {D}}})\) and a given \(\psi \in {\mathcal {C}}({\mathbb {T}})\), if a complex-valued function f satisfies \(\Delta f=g\) in \({\mathbb {D}}\) and \(f=\psi \) in \({\mathbb {T}}\), then, for \(z\in \overline{{\mathbb {D}}}\),

$$\begin{aligned} \left| f(z)-\frac{1-|z|^{2}}{1+|z|^{2}}{\mathcal {P}}_{\psi }(0)\right| \le \frac{4\Vert {\mathcal {P}}_{\psi }\Vert _{\infty }}{\pi }\arctan |z|+\frac{\Vert g\Vert _{\infty }}{4}(1-|z|^{2}), \end{aligned}$$
(1.5)

where

$$\begin{aligned} {\mathcal {P}}_{\psi }(z)=\frac{1}{2\pi }\int _{0}^{2\pi }\!\! P(z,e^{it})\psi (e^{it})dt,~\Vert {\mathcal {P}}_{\psi }\Vert _{\infty }= \sup _{z\in {\mathbb {D}}}|{\mathcal {P}}_{\psi }(z)|~\text{ and }~\Vert g\Vert _{\infty }=\sup _{z\in {\mathbb {D}}}|g(z)|. \end{aligned}$$

If we take \(g(z)=-4M\) and \(f(z)=M(1-|z|^{2})\) for \(z\in \overline{{\mathbb {D}}}\), where M is a positive constant, then the inequality (1.5) is sharp in \(\overline{{\mathbb {D}}}\).

The following result is a classical Schwarz Lemma at the boundary.

Theorem A

(see [15]) Let f be a holomorphic function from \({\mathbb {D}}\) into itself. If f is holomorphic at \(z=1\) with \(f(0)=0\) and \(f(1)=1\), then \(f'(1)\ge 1\). Moreover, the inequality is sharp.

Theorem A has attracted much attention and has been generalized in various forms (See [6, 23, 26, 27] for holomorphic functions, and see [21] for harmonic functions). In the following result, applying Theorem 1, we establish a Schwarz Lemma at the boundary for mappings satisfying the Poisson’s equation, which is a generalization of Theorem A.

Theorem 2

For a given \(g\in {\mathcal {C}}(\overline{{\mathbb {D}}})\), let \(f\in {\mathcal {C}}^{2}({\mathbb {D}})\cap {\mathcal {C}}({\mathbb {T}})\) be a function of \({\mathbb {D}}\) into itself satisfying \(\Delta f=g,\) where \(\Vert g\Vert _{\infty }<\frac{8}{3\pi }.\) If \(f(0)=0\) and, for some \(\zeta \in {\mathbb {T}}\), \(\lim _{r\rightarrow 1^{-}}|f(r\zeta )|=1\), then

$$\begin{aligned} \liminf _{r\rightarrow 1^{-}}\frac{|f(\zeta )-f(r\zeta )|}{1-r}\ge \frac{2}{\pi }-\frac{3\Vert g\Vert _{\infty }}{4}. \end{aligned}$$
(1.6)

In particular, if \(\Vert g\Vert _{\infty }=0\), then the estimate of (1.6) is sharp.

In [14], Colonna proved a sharp Schwarz-Pick type Lemma of complex-valued harmonic functions, which is as follows: If f is a complex-valued harmonic function from \({\mathbb {D}}\) into itself, then, for \(z\in {\mathbb {D}}\),

$$\begin{aligned} \Vert D_{f}(z)\Vert \le \frac{4}{\pi }\frac{1}{1-|z|^{2}}. \end{aligned}$$
(1.7)

We extend (1.7) into the following form.

Theorem 3

For a given \(g\in {\mathcal {C}}(\overline{{\mathbb {D}}})\) and a given \(\psi \in {\mathcal {C}}({\mathbb {T}})\), if a complex-valued function f satisfies \(\Delta f=g\) in \({\mathbb {D}}\) and \(f=\psi \) in \({\mathbb {T}}\), then, for \(z\in {\mathbb {D}}\backslash \{0\},\)

$$\begin{aligned} \Vert D_{f}(z)\Vert \le \frac{4\Vert {\mathcal {P}}_{\psi }\Vert _{\infty }}{\pi }\frac{1}{1-|z|^{2}}+2\mu (|z|), \end{aligned}$$
(1.8)

where

$$\begin{aligned} \frac{\Vert g\Vert _{\infty }}{4}\le \mu (|z|)=\frac{\Vert g\Vert _{\infty }(1-|z|^{2})}{8|z|^{2}} \bigg [\frac{1+|z|^{2}}{1-|z|^{2}}-\frac{(1-|z|^{2})}{2|z|}\log \frac{1+|z|}{1-|z|}\bigg ]\le \frac{\Vert g\Vert _{\infty }}{3} \end{aligned}$$

and \(\mu (|z|)\) is decreasing on \(|z|\in (0,1)\). In particular, if \(z=0\), then

$$\begin{aligned} \Vert D_{f}(0)\Vert \le \lim _{|z|\rightarrow 0^{+}}\bigg (\frac{4\Vert {\mathcal {P}}_{\psi }\Vert _{\infty }}{\pi }\frac{1}{1-|z|^{2}}+2\mu (|z|)\bigg )= \frac{4}{\pi }\Vert {\mathcal {P}}_{\psi }\Vert _{\infty }+\frac{2}{3}\Vert g\Vert _{\infty }. \end{aligned}$$
(1.9)

Moreover, if \(\Vert g\Vert _{\infty }=0\), then the extremal functions

$$\begin{aligned} f(z)=\frac{2M\alpha }{\pi }\arg \left( \frac{1+\phi (z)}{1-\phi (z)}\right) \end{aligned}$$

show that the estimate of (1.8) and (1.9) are sharp, where \(|\alpha |=1\) and \(M>0\) are constants, and \(\phi \) is a conformal automorphism of \({\mathbb {D}}\).

We remark that if \(\Vert g\Vert _{\infty }=0\) and \(\Vert {\mathcal {P}}_{\psi }\Vert _{\infty }=1\) in Theorem 3, then (1.8) and (1.9) coincide with (1.7).

Let \({\mathcal {A}}\) denote the set of all analytic functions f defined in \({\mathbb {D}}\) satisfying the standard normalization: \(f(0)=f'(0)-1=0\). In the early 20th century, Landau [24] showed that there is a constant \(r>0\), independent of \(f\in {\mathcal {A}}\), such that \(f({\mathbb {D}})\) contains a disk of radius r. Let \(L_{f}\) be the supremum of the set of positive numbers r such that \(f({\mathbb {D}})\) contains a disk of radius r, where \(f\in {\mathcal {A}}\). Then we call \(\inf _{f\in {\mathcal {A}}}L_{f}\) the Landau–Bloch constant. One of the long standing open problems in geometric function theory is to determine the precise value of the Landau–Bloch constant. It has attracted much attention, see [4, 25, 28, 29, 32] and references therein. The Landau theorem is an important tool in geometric function theory of one complex variable (cf. [5, 33]). Unfortunately, for general class of functions, there is no Landau type theorem (see [7, 32]). In order to obtain some analogs of the Landau type theorem for more general classes of functions, it is necessary to restrict the class of functions considered (cf. [1, 3, 7,8,9,10,11, 13, 16, 32]). Let’s recall some known results as follows.

Theorem B

([7, Theorem 2]) Let f be a harmonic mapping in \({\mathbb {D}}\) such that \(f(0)=J_{f}(0)-1=0\) and \(|f(z)|<M\) for \(z\in {\mathbb {D}}\), where M is a positive constant. Then f is univalent in \({\mathbb {D}}_{\rho _{0}}\) with \(\rho _{0}=\pi ^{3}/(64mM^{2})\), and \(f({\mathbb {D}}_{\rho _{0}})\) contains a univalent disk \({\mathbb {D}}_{R_{0}}\) with

$$\begin{aligned} R_{0}=\frac{\pi }{8M}\rho _{0}=\frac{\pi ^{4}}{512mM^{3}}, \end{aligned}$$

where \(m\approx 6.85\) is the minimum of the function \((3-r^{2})/[r(1-r^{2})]\) for \(r\in (0,1)\).

Theorem C

([1, Theorem 1]) Let \(f(z)=|z|^{2}G(z)+K(z)\) be a biharmonic mapping, that is \(\Delta (\Delta f)=0\), in \({\mathbb {D}}\) such that \(f(0)= K(0)=J_{f}(0)-1=0\), where G and K are harmonic satisfying \(|G(z)|, ~|K(z)|<M\) for \(z\in {\mathbb {D}}\), where M is a positive constant. Then there is a constant \(\rho _{2}\in (0,1)\) such that f is univalent in \({\mathbb {D}}_{\rho _{2}}\). Specifically, \(\rho _{2}\) satisfies

$$\begin{aligned} \frac{\pi }{4M}-2\rho _{2}M-2M\left[ \frac{\rho _{2}^{2}}{(1-\rho _{2})^{2}}+\frac{1}{(1-\rho _{2})^{2}}-1\right] =0 \end{aligned}$$

and \(f({\mathbb {D}}_{\rho _{2}})\) contains a disk \({\mathbb {D}}_{R_{2}}\), where

$$\begin{aligned} R_{2}=\frac{\pi }{4M}\rho _{2}-2M\frac{\rho _{2}^{3}+\rho _{2}^{2}}{1-\rho _{2}}. \end{aligned}$$

For some \(g\in {\mathcal {C}}(\overline{{\mathbb {D}}})\), let \({\mathcal {F}}_{g}(\overline{{\mathbb {D}}})\) denote the class of all complex-valued functions \(f\in {\mathcal {C}}^{2}({\mathbb {D}})\cap {\mathcal {C}}({\mathbb {T}})\) satisfying \(\Delta f=g\) and \(f(0)=J_{f}(0)-1=0\). We extend Theorems B and C into the following from.

Theorem 4

For a given \(g\in {\mathcal {C}}(\overline{{\mathbb {D}}})\), let \(f\in {\mathcal {F}}_{g}(\overline{{\mathbb {D}}})\) satisfying \(\Vert g\Vert _{\infty }\le M_{1}\) and \(\Vert f\Vert _{\infty }\le M_{2}\), where \(M_{1}\ge 0\) and \(M_{2}>0\) are constants. Then f is univalent in \({\mathbb {D}}_{r_{0}}\), where \(r_{0}\) satisfies the following equation

$$\begin{aligned} \frac{1}{\frac{4}{\pi }M_{2}+\frac{2}{3}M_{1}}-\frac{4M_{2}}{\pi }\frac{r_{0}(2-r_{0})}{(1-r_{0})^{2}} -2M_{1}\big [\log 4(1+r_{0})- \log r_{0}\big ](2+r_{0})r_{0}=0. \end{aligned}$$

Moreover, \(f({\mathbb {D}}_{r_{0}})\) contains an univalent disk \({\mathbb {D}}_{R_{0}}\) with

$$\begin{aligned} R_{0}\ge \frac{2M_{2}}{\pi }\frac{r_{0}^{2}(2-r_{0})}{(1-r_{0})^{2}}. \end{aligned}$$

Remark 1.1

Theorem 4 gives an affirmative answer to the open problem of [13] for the u-gradient mapping\(f\in {\mathcal {C}}^{2}({\mathbb {D}})\). If g is harmonic, then all \(f\in {\mathcal {F}}_{g}(\overline{{\mathbb {D}}})\) are biharmonic. Furthermore, if \(\Vert g\Vert _{\infty }=0\), then all \(f\in {\mathcal {F}}_{g}(\overline{{\mathbb {D}}})\) are harmonic. Hence, Theorem 4 is also a generalization of a series of known results, such as [1, Theorem 2], [7, Theorems, 3, 4, 5 and 6], [8, Theorems 2 and 3].

In the following two Examples, we will show that there is no Landau type Theorem for \(f\in {\mathcal {F}}_{g}(\overline{{\mathbb {D}}})\) without the boundedness hypothesis of \(\Vert f\Vert _{\infty }\).

Example 1.10

For \(g\equiv 1\) and \(z=x+iy\in {\mathbb {D}},\) let \(f_{k}(z)=kx+|z|^{2}/4+i\frac{y}{k},\) where \(k\in \{1,2,\ldots \}\). Then, for all \(k\in \{1,2,\ldots \}\), \(f_{k}\) is univalent. For all \(k\in \{1,2,\ldots \}\), by simple calculations, we see that \(J_{f_{k}} (0)-1=f_{k}(0)=0\), and there is no an absolute constant \(\rho _{0}>0\) such that \({\mathbb {D}}_{\rho _{0}} \) is contained in \(f_{k}({\mathbb {D}})\).

Example 1.11

For \(\Vert g\Vert _{\infty }=0\) and \(z=x+iy\in {\mathbb {D}},\) let \(f_{k}(z)=kx+i\frac{y}{k},\) where \(k\in \{1,2,\ldots \}\). For all \(k\in \{1,2,\ldots \}\), it is not difficult to see that \(f_{k}\) is univalent and \(J_{f_{k}} (0)-1=f_{k}(0)=0\). Moreover, for all \(k\in \{1,2,\ldots \}\), \(f_{k}({\mathbb {D}})\) contains no disk with radius bigger than 1 / k. Hence, for all \(k\in \{1,2,\ldots \}\), there is no an absolute constant \(r_{0}>0\) such that \({\mathbb {D}}_{r_{0}} \) is contained in \(f_{k}({\mathbb {D}})\).

Corollary 1

Under the same hypothesis of Theorem 4, there is a \(r_{0}\in (0,1)\) such that f is bi-Lipschitz in \({\mathbb {D}}_{r_{0}}\).

The proofs of Theorems 1, 2, 3, 4 and Corollary 1 will be presented in Sect. 2.

2 Proofs of the Main Results

Proof of Theorem 1

For a given \(g\in {\mathcal {C}}({\mathbb {D}})\), by (1.1), we have

$$\begin{aligned} f(z)={\mathcal {P}}_{\psi }(z)-{\mathcal {G}}_{g}(z), \quad z\in {\mathbb {D}}, \end{aligned}$$
(2.1)

where \({\mathcal {P}}_{\psi }\) and \({\mathcal {G}}_{g}\) are defined in (1.2). Since \({\mathcal {P}}_{\psi }\) is harmonic in \({\mathbb {D}}\), by (1.4), we see that, for \(z\in {\mathbb {D}}\),

$$\begin{aligned} \left| {\mathcal {P}}_{\psi }(z)-\frac{1-|z|^{2}}{1+|z|^{2}}{\mathcal {P}}_{\psi }(0)\right| \le \frac{4\Vert {\mathcal {P}}_{\psi }\Vert _{\infty }}{\pi }\arctan |z|. \end{aligned}$$
(2.2)

On the other hand, for a fixed \(z\in {\mathbb {D}}\), let

$$\begin{aligned} \zeta =\frac{z-w}{1-\overline{z}w}, \end{aligned}$$

which is equivalent to

$$\begin{aligned} w=\frac{z-\zeta }{1-\overline{z}\zeta }. \end{aligned}$$

Then

$$\begin{aligned} \big |{\mathcal {G}}_{g}(z)\big |= & {} \bigg |\frac{1}{2\pi }\int _{{\mathbb {D}}}\left( \log \frac{1}{|\zeta |}\right) g\left( \frac{z-\zeta }{1-\overline{z}\zeta }\right) \frac{(1-|z|^{2})^{2}}{|1-\overline{z}\zeta |^{4}}dA(\zeta )\bigg |\nonumber \\\le & {} \frac{\Vert g\Vert _{\infty }}{2\pi }\bigg |\int _{{\mathbb {D}}}\left( \log \frac{1}{|\zeta |}\right) \frac{(1-|z|^{2})^{2}}{|1-\overline{z}\zeta |^{4}}dA(\zeta )\bigg |\nonumber \\= & {} (1-|z|^{2})^{2}\Vert g\Vert _{\infty }\int _{0}^{1}\left[ \bigg (\frac{1}{2\pi }\int _{0}^{2\pi }\frac{dt}{|1-\overline{z}re^{it}|^{4}}\bigg )r\log \frac{1}{r}\right] dr\nonumber \\= & {} (1-|z|^{2})^{2}\Vert g\Vert _{\infty }\int _{0}^{1}\left[ \bigg (\frac{1}{2\pi }\int _{0}^{2\pi }\frac{dt}{|(1-\overline{z}re^{it})^{2}|^{2}}\bigg )r\log \frac{1}{r}\right] dr\nonumber \\= & {} (1-|z|^{2})^{2}\Vert g\Vert _{\infty }\int _{0}^{1}\left[ \left( \frac{1}{2\pi }\int _{0}^{2\pi }\left| \sum _{n=0}^{\infty }(n+1)(r\overline{z})^{n}e^{int}\right| ^{2}dt\right) r\log \frac{1}{r}\right] dr \nonumber \\= & {} (1-|z|^{2})^{2}\Vert g\Vert _{\infty }\int _{0}^{1}\bigg (r\log \frac{1}{r}\bigg )\sum _{n=0}^{\infty }(n+1)^{2}|z|^{2n}r^{2n}dr\nonumber \\= & {} (1-|z|^{2})^{2}\Vert g\Vert _{\infty }\sum _{n=0}^{\infty }(n+1)^{2}|z|^{2n}\int _{0}^{1}r^{2n+1}\bigg (\log \frac{1}{r}\bigg )dr \nonumber \\= & {} \frac{(1-|z|^{2})^{2}\Vert g\Vert _{\infty }}{4}\sum _{n=0}^{\infty }|z|^{2n}\nonumber \\= & {} \frac{\Vert g\Vert _{\infty }}{4}(1-|z|^{2}). \end{aligned}$$
(2.3)

Hence, by (2.2) and (2.3), we conclude that

$$\begin{aligned} \left| f(z)-\frac{1-|z|^{2}}{1+|z|^{2}}{\mathcal {P}}_{\psi }(0)\right|\le & {} \left| {\mathcal {P}}_{\psi }(z)-\frac{1-|z|^{2}}{1+|z|^{2}}{\mathcal {P}}_{\psi }(0)\right| +\big |{\mathcal {G}}_{g}(z)\big |\\\le & {} \frac{4\Vert {\mathcal {P}}_{\psi }\Vert _{\infty }}{\pi }\arctan |z|+\frac{\Vert g\Vert _{\infty }}{4}(1-|z|^{2}). \end{aligned}$$

Now we prove the sharpness part. For \(z\in \overline{{\mathbb {D}}}\), let

$$\begin{aligned} g(z)=-4M\quad \text{ and }\quad f(z)=M(1-|z|^{2}), \end{aligned}$$

where M is a positive constant. Then \(\psi \equiv 0\) in \({\mathbb {T}}\) and

$$\begin{aligned} \left| f(z)-\frac{1-|z|^{2}}{1+|z|^{2}}{\mathcal {P}}_{\psi }(0)\right|= & {} |f(z)|=\left| \frac{1}{2\pi }\int _{{\mathbb {D}}}G(z,w)g(w)dA(w)\right| \\= & {} \frac{\Vert g\Vert _{\infty }}{4}(1-|z|^{2}), \end{aligned}$$

which shows (1.5) is sharp in \(\overline{{\mathbb {D}}}\). The proof of this theorem is complete. \(\square \)

Proof of Theorem 2

For a given \(g\in {\mathcal {C}}(\overline{{\mathbb {D}}})\), by (1.1) with f in place of \(\psi \), we have

$$\begin{aligned} f(z)={\mathcal {P}}_{f}(z)-{\mathcal {G}}_{g}(z), \quad z\in {\mathbb {D}}, \end{aligned}$$

where \({\mathcal {P}}_{f}\) and \({\mathcal {G}}_{g}\) are defined in (1.2). Since \(f(0)=0\), we see that

$$\begin{aligned} |{\mathcal {P}}_{f}(0)|= & {} |{\mathcal {G}}_{g}(0)|=\left| \frac{1}{2\pi }\int _{{\mathbb {D}}}\log \frac{1}{|w|}g(w)dA(w)\right| \nonumber \\\le & {} \frac{\Vert g\Vert _{\infty }}{2\pi }\int _{0}^{2\pi }dt\int _{0}^{1}r\log \frac{1}{r}dr\nonumber \\= & {} \frac{\Vert g\Vert _{\infty }}{4}. \end{aligned}$$
(2.4)

Let \(z=r\zeta \in {\mathbb {D}}\), where \(\zeta \in {\mathbb {T}}\) is as in the statement of the theorem. Then, by (2.4) and Theorem 1, we have

$$\begin{aligned} |f(\zeta )-f(r\zeta )|= & {} \left| f(\zeta )+{\mathcal {P}}_{f}(0)\frac{1-|z|^{2}}{1+|z|^{2}}-{\mathcal {G}}_{g}(0)\frac{1-|z|^{2}}{1+|z|^{2}}-f(r\zeta )\right| \\\ge & {} 1-\left| f(r\zeta )-{\mathcal {P}}_{f}(0)\frac{1-|z|^{2}}{1+|z|^{2}}\right| -|{\mathcal {G}}_{g}(0)|\frac{1-|z|^{2}}{1+|z|^{2}}\\\ge & {} 1-\frac{4}{\pi }\arctan |z|-\frac{\Vert g\Vert _{\infty }}{4}(1-|z|^{2})-|{\mathcal {G}}_{g}(0)|\frac{1-|z|^{2}}{1+|z|^{2}}\\\ge & {} 1-\frac{4}{\pi }\arctan |z|-\frac{\Vert g\Vert _{\infty }}{4}(1-|z|^{2})-\frac{\Vert g\Vert _{\infty }}{4}\frac{(1-|z|^{2})}{1+|z|^{2}}, \end{aligned}$$

which, together with L’Hospital’s rule, gives that

$$\begin{aligned} \liminf _{r\rightarrow 1^{-}}\frac{|f(e^{i\theta })-f(re^{i\theta })|}{1-r}\ge & {} \lim _{r\rightarrow 1^{-}}\frac{1-\frac{4}{\pi }\arctan r-\frac{\Vert g\Vert _{\infty }}{4}(1-r^{2})-\frac{\Vert g\Vert _{\infty }}{4}\frac{(1-r^{2})}{1+r^{2}}}{1-r}\\= & {} \lim _{r\rightarrow 1^{-}}\left[ \frac{4}{\pi }\frac{1}{1+r^{2}}-\Vert g\Vert _{\infty }\frac{r}{2}-\Vert g\Vert _{\infty }\frac{r}{(1+r^{2})^{2}}\right] \\= & {} \frac{2}{\pi }-\frac{3\Vert g\Vert _{\infty }}{4}. \end{aligned}$$

Now we prove the sharpness part. For \(z\in {\mathbb {D}}\), let

$$\begin{aligned} f(z)=\frac{2}{\pi }\arctan \frac{2\text{ Re }(z)}{1-|z|^{2}}. \end{aligned}$$

Then f is harmonic in \({\mathbb {D}}\) with \(f(0)=f(1)-1=0\), and

$$\begin{aligned} f(\rho )=\frac{4}{\pi }\arctan \rho , \end{aligned}$$

where \(\rho \in (-1,1)\). Elementary calculations show that

$$\begin{aligned} \liminf _{\rho \rightarrow 1^{-}}\frac{|f(1)-f(\rho )|}{1-\rho }=\frac{2}{\pi }, \end{aligned}$$

which implies that (1.6) is sharp for \(\Vert g\Vert _{\infty }=0\). The proof of this theorem is complete. \(\square \)

Theorem D

([31\(\text{ or }\) [19, Proposition 2.4]) Suppose that X is an open subset of \({\mathbb {R}}\), and \(\Omega \) a measure space. Suppose, further, that a function F :  \(X\times \Omega \rightarrow {\mathbb {R}}\) satisfies the following conditions:

  1. (1)

    F(xw) is a measurable function of x and w jointly, and is integrable with respect to w for almost every \(x\in X.\)

  2. (2)

    For almost every \(w\in \Omega \), F(xw) is an absolutely continuous function with respect to x. [This guarantees that \(\partial F(x,w)/\partial x\) exists almost everywhere.]

  3. (3)

    \(\partial F/\partial x\) is locally integrable, that is, for all compact intervals [ab] contained in X:

    $$\begin{aligned} \int _{a}^{b}\int _{\Omega }\left| \frac{\partial }{\partial x}F(x,w)\right| dwdx<\infty . \end{aligned}$$

Then, \(\int _{\Omega }F(x,w)dw\) is an absolutely continuous function with respect to x, and for almost every \(x\in X\), its derivative exists, which is given by

$$\begin{aligned} \frac{d}{dx}\int _{\Omega }F(x,w)dw=\int _{\Omega }\frac{\partial }{\partial x}F(x,w)dw. \end{aligned}$$

Proof of Theorem 3

For a given \(g\in {\mathcal {C}}(\overline{{\mathbb {D}}})\), by (2.1), we have

$$\begin{aligned} f(z)={\mathcal {P}}_{\psi }(z)-{\mathcal {G}}_{g}(z), \quad z\in {\mathbb {D}}, \end{aligned}$$

where \({\mathcal {P}}_{\psi }\) and \({\mathcal {G}}_{g}\) are the same as in (2.1). Applying [19, Lemma 2.3] and Theorem D, we have

$$\begin{aligned} \frac{\partial }{\partial z}{\mathcal {G}}_{g}(z)= & {} \frac{1}{2\pi }\int _{{\mathbb {D}}}\frac{\partial }{\partial z}G(z,w)g(w)dA(w)\nonumber \\= & {} \frac{1}{4\pi }\int _{{\mathbb {D}}}\frac{(1-|w|^{2})}{(z-w)(z\overline{w}-1)}g(w)dA(w)\in {\mathcal {C}}({\mathbb {D}}) \end{aligned}$$
(2.5)

and

$$\begin{aligned} \frac{\partial }{\partial \overline{z}}{\mathcal {G}}_{g}(z)= & {} \frac{1}{2\pi }\int _{{\mathbb {D}}}\frac{\partial }{\partial \overline{z}}G(z,w)g(w)dA(w)\\= & {} \frac{1}{4\pi }\int _{{\mathbb {D}}}\frac{(1-|w|^{2})}{(\overline{z}-\overline{w})(w\overline{z}-1)}g(w)dA(w)\in {\mathcal {C}}({\mathbb {D}}). \end{aligned}$$

For a fixed \(z\in {\mathbb {D}}\backslash \{0\}\), let

$$\begin{aligned} \zeta =\frac{z-w}{1-\overline{z}w} \end{aligned}$$
(2.6)

which implies that

$$\begin{aligned} w=\frac{z-\zeta }{1-\overline{z}\zeta },\quad 1-\overline{z}w=\frac{1-|z|^{2}}{1-\overline{z}\zeta }\quad \text{ and }\quad 1-|w|^{2}=\frac{(1-|\zeta |^{2})(1-|z|^{2})}{|1-\overline{z}\zeta |^{2}}. \nonumber \\ \end{aligned}$$
(2.7)

Then, by (2.5), (2.7) and the change of variables (2.6), we have

$$\begin{aligned} \left| \frac{\partial }{\partial z}{\mathcal {G}}_{g}(z)\right|\le & {} \frac{1}{4\pi }\int _{{\mathbb {D}}}\frac{(1-|w|^{2})}{|z-w||z\overline{w}-1|}|g(w)|dA(w)\nonumber \\\le & {} \frac{\Vert g\Vert _{\infty }}{4\pi }\int _{{\mathbb {D}}}\frac{(1-|w|^{2})}{|z-w||z\overline{w}-1|}dA(w)\nonumber \\= & {} \frac{\Vert g\Vert _{\infty }}{4\pi }\int _{{\mathbb {D}}}\frac{(1-|w|^{2})}{|\zeta ||1-\overline{z}w|^{2}}\frac{(1-|z|^{2})^{2}}{|1-\overline{z}\zeta |^{4}}dA(\zeta ) \nonumber \\= & {} \frac{\Vert g\Vert _{\infty }}{4\pi }\int _{{\mathbb {D}}}\frac{(1-|z|^{2})(1-|\zeta |^{2})}{|\zeta ||1-\overline{z}\zeta |^{4}}dA(\zeta )\nonumber \\= & {} \frac{\Vert g\Vert _{\infty }(1-|z|^{2})}{2}\int _{0}^{1}\left[ (1-r^{2})\bigg (\frac{1}{2\pi }\int _{0}^{2\pi }\frac{dt}{|1-\overline{z}re^{it}|^{4}}\bigg )\right] dr\nonumber \\= & {} \frac{\Vert g\Vert _{\infty }(1-|z|^{2})}{2}\int _{0}^{1} \left[ (1-r^{2})\left( \frac{1}{2\pi }\int _{0}^{2\pi }\left| \sum _{n=0}^{\infty }(n+1)(r\overline{z})^{n}e^{int}\right| ^{2}dt\right) \right] dr\nonumber \\= & {} \frac{\Vert g\Vert _{\infty }(1-|z|^{2})}{2}\int _{0}^{1} (1-r^{2})\left[ \sum _{n=0}^{\infty }(n+1)^{2}|z|^{2n}r^{2n}\right] dr\nonumber \\= & {} \frac{\Vert g\Vert _{\infty }(1-|z|^{2})}{2}\int _{0}^{1} \frac{(1-r^{2})(1+|z|^{2}r^{2})}{(1-|z|^{2}r^{2})^{3}}dr\nonumber \\= & {} \frac{\Vert g\Vert _{\infty }(1-|z|^{2})}{2}\bigg [-\frac{1}{|z|^{2}}I_{1}+ \left( \frac{3}{|z|^{2}}-1\right) I_{2} +2\left( 1-\frac{1}{|z|^{2}}\right) I_{3}\bigg ], \end{aligned}$$
(2.8)

where

$$\begin{aligned} I_{1}= & {} \int _{0}^{1}\frac{dr}{1-r^{2}|z|^{2}}=\frac{1}{|z|}\log \frac{1+|z|r}{\sqrt{1-|z|^{2}r^{2}}}\bigg |_{0}^{1}=\frac{1}{|z|}\log \frac{1+|z|}{\sqrt{1-|z|^{2}}}, \end{aligned}$$
(2.9)
$$\begin{aligned} I_{2}= & {} \int _{0}^{1}\frac{dr}{(1-r^{2}|z|^{2})^{2}}=\frac{1}{2|z|}\left( \log \frac{1+|z|r}{\sqrt{1-|z|^{2}r^{2}}}+ \frac{|z|r}{1-|z|^{2}r^{2}}\right) \bigg |_{0}^{1}\nonumber \\= & {} \frac{1}{2|z|}\log \frac{1+|z|}{\sqrt{1-|z|^{2}}}+\frac{1}{2(1-|z|^{2})} \end{aligned}$$
(2.10)

and

$$\begin{aligned} I_{3}= & {} \int _{0}^{1}\frac{dr}{(1-r^{2}|z|^{2})^{3}}=\frac{1}{4|z|}\left( \frac{|z|r}{(1-r^{2}|z|^{2})^{2}} +\frac{3}{2}\frac{|z|r}{1-r^{2}|z|^{2}}+\frac{3}{2}\log \frac{1+|z|r}{\sqrt{1-|z|^{2}r^{2}}}\right) \bigg |_{0}^{1}\nonumber \\= & {} \frac{1}{4(1-|z|^{2})^{2}}+\frac{3}{8(1-|z|^{2})}+\frac{3}{8|z|}\log \frac{1+|z|}{\sqrt{1-|z|^{2}}}. \end{aligned}$$
(2.11)

By (2.9), (2.10) and (2.11), we get

$$\begin{aligned}&-\frac{1}{|z|^{2}}I_{1}+ \left( \frac{3}{|z|^{2}}-1\right) I_{2} +2\left( 1-\frac{1}{|z|^{2}}\right) I_{3}\\&\quad =\frac{1}{4|z|^{2}}\bigg [\frac{1+|z|^{2}}{1-|z|^{2}}-\frac{(1-|z|^{2})}{2|z|}\log \frac{1+|z|}{1-|z|}\bigg ], \end{aligned}$$

which, together with (2.8), yields that

$$\begin{aligned} \left| \frac{\partial }{\partial z}{\mathcal {G}}_{g}(z)\right| \le \mu (|z|), \end{aligned}$$
(2.12)

where

$$\begin{aligned} \mu (|z|)=\frac{\Vert g\Vert _{\infty }(1-|z|^{2})}{8|z|^{2}} \bigg [\frac{1+|z|^{2}}{1-|z|^{2}}-\frac{(1-|z|^{2})}{2|z|}\log \frac{1+|z|}{1-|z|}\bigg ]. \end{aligned}$$

By a similar proof process of (2.12), we have

$$\begin{aligned} \left| \frac{\partial }{\partial \overline{z}}{\mathcal {G}}_{g}(z)\right| \le \mu (|z|). \end{aligned}$$
(2.13)

By direct calculation (or by [19, Lemma 2.3]), we obtain

$$\begin{aligned}&\lim _{|z|\rightarrow 0^{+}}\frac{\Vert g\Vert _{\infty }(1-|z|^{2})}{8|z|^{2}} \bigg [\frac{1+|z|^{2}}{1-|z|^{2}}-\frac{(1-|z|^{2})}{2|z|}\log \frac{1+|z|}{1-|z|}\bigg ]=\frac{\Vert g\Vert _{\infty }}{3},\\&\lim _{|z|\rightarrow 1^{-}}\frac{\Vert g\Vert _{\infty }(1-|z|^{2})}{8|z|^{2}} \bigg [\frac{1+|z|^{2}}{1-|z|^{2}}-\frac{(1-|z|^{2})}{2|z|}\log \frac{1+|z|}{1-|z|}\bigg ]=\frac{\Vert g\Vert _{\infty }}{4}\nonumber \end{aligned}$$
(2.14)

and \(\mu (|z|)\) is decreasing on \(|z|\in (0,1)\).

On the other hand, since \({\mathcal {P}}_{\psi }\) is harmonic in \({\mathbb {D}}\), by [14, Theorem 3] (see also [11, 12]), we see that, for \(z\in {\mathbb {D}}\),

$$\begin{aligned} \Vert D_{{\mathcal {P}}_{\psi }}(z)\Vert \le \frac{4\Vert {\mathcal {P}}_{\psi }\Vert _{\infty }}{\pi }\frac{1}{1-|z|^{2}}. \end{aligned}$$
(2.15)

Hence (1.8) follows from (2.12), (2.13) and (2.15). Furthermore, applying (1.8) and (2.14), we get (1.9). The proof of this theorem is complete. \(\square \)

Now we formulate the following well-known result.

Lemma 1

The improper integral

$$\begin{aligned} \int _{0}^{\frac{\pi }{2}}\log \sin x dx=\int _{0}^{\frac{\pi }{2}}\log \cos x dx=-\frac{\pi }{2}\log 2. \end{aligned}$$

Lemma 2

For \(z\in {\mathbb {D}}\backslash \{0\}\), the improper integral

$$\begin{aligned} \int _{{\mathbb {D}}}\frac{dA(w)}{|w||z-w|}= & {} \int _{0}^{2\pi }\log \big (1-r\cos t+\sqrt{1+r^{2}-2r\cos t}\big )dt\\&-\,2\pi \log r+2\pi \log 2\\\le & {} 2\pi \log 4(1+r)-2\pi \log r, \end{aligned}$$

where \(r=|z|\).

Proof

Let \(z=re^{i\alpha }\) and \(w=\rho e^{i\theta } \). Then

$$\begin{aligned} \int _{{\mathbb {D}}}\frac{dA(w)}{|w||z-w|}= & {} \int _{0}^{1}d\rho \int _{0}^{2\pi }\frac{d\theta }{\sqrt{r^{2}+\rho ^{2}-2\rho r\cos (\theta -\alpha )}}\nonumber \\= & {} \int _{0}^{1}d\rho \int _{0}^{2\pi }\frac{dt}{\sqrt{r^{2}+\rho ^{2}-2\rho r\cos t}}\nonumber \\= & {} \int _{0}^{2\pi }dt\int _{0}^{1}\frac{d\rho }{\sqrt{r^{2}+\rho ^{2}-2\rho r\cos t}}\nonumber \\= & {} \int _{0}^{2\pi }\bigg \{\frac{1}{2r \cos t}\bigg [\int _{0}^{1}\frac{2\rho d\rho }{\sqrt{r^{2}+\rho ^{2}-2\rho r\cos t}}\nonumber \\&-\int _{0}^{1}\frac{d(r^{2}+\rho ^{2}-2\rho r\cos t)}{\sqrt{r^{2}+\rho ^{2}-2\rho r\cos t}}\bigg ]\bigg \}dt\nonumber \\= & {} \int _{0}^{2\pi }\bigg [\frac{1}{r\cos t}\int _{0}^{1}\frac{\rho d\rho }{\sqrt{r^{2}+\rho ^{2}-2\rho r\cos t}} \nonumber \\&-\,\frac{1}{r\cos t}\big (\sqrt{1+r^{2}-2r\cos t}-r\big )\bigg ]dt\nonumber \\= & {} \int _{0}^{2\pi }\bigg [\frac{1}{r\cos t}\int _{0}^{1}\frac{\rho d\rho }{\sqrt{r^{2}+\rho ^{2}-2\rho r\cos t}}\nonumber \\&-\,\frac{\sqrt{1+r^{2}-2r\cos t}}{r\cos t}+\frac{1}{\cos t}\bigg ]dt. \end{aligned}$$
(2.16)

By calculations, we get

$$\begin{aligned} \int _{0}^{1}\frac{\rho d\rho }{\sqrt{r^{2}+\rho ^{2}-2\rho r\cos t}}= & {} H(\rho )|_{0}^{1}\nonumber \\= & {} \sqrt{1+r^{2}-2r\cos t}\nonumber \\&+\,r\cos t \log \left( 1-r\cos t+\sqrt{1+r^{2}-2r\cos t}\right) \nonumber \\&-\, r-r\cos t\log r(1-\cos t), \end{aligned}$$
(2.17)

where

$$\begin{aligned} H(\rho )=\sqrt{\rho ^{2}+r^{2}-2r\rho \cos t}+r \cos t\log \left( \rho -r\cos t+\sqrt{r^{2}+\rho ^{2}-2\rho r\cos t}\right) . \end{aligned}$$

By (2.16), (2.17) and Lemma 1, we see that

$$\begin{aligned} \int _{{\mathbb {D}}}\frac{dA(w)}{|w||z-w|}= & {} \int _{0}^{2\pi }\log \big (1-r\cos t+\sqrt{1+r^{2}-2r\cos t}\big )dt\nonumber \\&-\int _{0}^{2\pi }\log r(1-\cos t)dt\nonumber \\= & {} \int _{0}^{2\pi }\log \big (1-r\cos t+\sqrt{1+r^{2}-2r\cos t}\big )dt \nonumber \\&-\,2\pi \log r-\int _{0}^{2\pi }\log \left( 2\sin ^{2}\frac{t}{2}\right) dt \nonumber \\= & {} \int _{0}^{2\pi }\log \big (1-r\cos t+\sqrt{1+r^{2}-2r\cos t}\big )dt\nonumber \\&-\,2\pi \log 2r-8\int _{0}^{\frac{\pi }{2}}\log (\sin t) dt \nonumber \\= & {} \int _{0}^{2\pi }\log \big (1-r\cos t+\sqrt{1+r^{2}-2r\cos t}\big )dt \nonumber \\&-\,2\pi \log r+2\pi \log 2\nonumber \\\le & {} 2\pi \log 4(1+r)-2\pi \log r. \end{aligned}$$
(2.18)

The proof of this lemma is complete. \(\square \)

Lemma E

([10, Lemma 1]) Let f be a harmonic mapping of \({\mathbb {D}}\) into \({\mathbb {C}}\) such that \(|f(z)|\le M\) and \(f(z)=\sum _{n=0}^{\infty }a_{n}z^{n}+\sum _{n=1}^{\infty }\overline{b}_{n}\overline{z}^{n}\). Then \(|a_{0}|\le M\) and for all \(n\ge 1,\)

$$\begin{aligned} |a_{n}|+|b_{n}|\le \frac{4M}{\pi }. \end{aligned}$$

Lemma 3

For \(x\in (0,1)\), let

$$\begin{aligned} \phi (x)=\frac{1}{\frac{4}{\pi }M_{2}+\frac{2}{3}M_{1}}-\frac{4M_{2}}{\pi }\frac{x(2-x)}{(1-x)^{2}} -2M_{1}\big [\log 4(1+x)- \log x\big ](2+x)x, \end{aligned}$$

where \(M_{2}>0\) and \(M_{1}\ge 0\) are constant. Then \(\phi \) is strictly decreasing and there is an unique \(x_{0}\in (0,1)\) such that \(\phi (x_{0})=0.\)

Proof

For \(x\in (0,1)\), let

$$\begin{aligned} f_{1}(x)=\frac{4M_{2}}{\pi }\frac{x(2-x)}{(1-x)^{2}} \end{aligned}$$

and

$$\begin{aligned} f_{2}(x)=2M_{1}\big [\log 2(1+x)- \log x+\log 2\big ](2+x)x. \end{aligned}$$

Since, for \(x\in (0,1)\),

$$\begin{aligned} f_{1}'(x)=\frac{8M_{2}}{\pi }\frac{1}{(1-x)^{3}}>0 \end{aligned}$$

and

$$\begin{aligned} f_{2}'(x)= & {} 2M_{1}\left[ 2(x+1)\log \frac{4(1+x)}{x}-\frac{2+x}{1+x}\right] \\= & {} 2M_{1}\left\{ 2(x+1)\bigg [\log 4+\log \Big (1+\frac{1}{x}\Big )\bigg ]-\frac{2+x}{1+x}\right\} \\\ge & {} 2M_{1}\left\{ 2(x+1)\bigg [1+\frac{1}{1+x}\bigg ]-\frac{2+x}{1+x}\right\} \\= & {} 2M_{1}\frac{(2+x)(2x+1)}{1+x}\ge 0, \end{aligned}$$

we see that \(f_{1}+f_{2}\) is continuous and strictly increasing in (0, 1). Then \(\phi \) is continuous and strictly decreasing in (0, 1), which, together with

$$\begin{aligned} \lim _{x\rightarrow 0^{+}}\phi (x)=\frac{1}{\frac{4}{\pi }M_{2}+\frac{2}{3}M_{1}}\quad \text{ and }\quad \lim _{x\rightarrow 1^{-}}\phi (x)=-\infty , \end{aligned}$$

implies that there is an unique \(x_{0}\in (0,1)\) such that \(\phi (x_{0})=0.\)\(\square \)

Lemma 4

For \(x\in (0,1]\), let

$$\begin{aligned} \tau _{1}(x)=\frac{2-r_{0}x}{(1-r_{0}x)^{2}}\quad \text{ and }\quad \tau _{2}(x)=x\big [\log 4(1+r_{0}x)-\log (r_{0}x)\big ], \end{aligned}$$

where \(r_{0}\in (0,1)\) is a constant. Then \(\tau _{1}\) and \(\tau _{2}\) are increasing functions in (0, 1].

Proof of Theorem 4

As before, by (2.1) with f in place of \(\psi \), we have

$$\begin{aligned} f(z)={\mathcal {P}}_{f}(z)-{\mathcal {G}}_{g}(z), \quad z\in {\mathbb {D}}, \end{aligned}$$

where \({\mathcal {P}}_{f}\) and \({\mathcal {G}}_{g}\) are defined in (2.1). By [19, Lemma 2.3], Theorem D and Lemma 2, we have

$$\begin{aligned} \left| \frac{\partial {\mathcal {G}}_{g}(z)}{\partial z}-\frac{\partial {\mathcal {G}}_{g}(0)}{\partial z}\right|= & {} \bigg |\frac{1}{4\pi }\int _{{\mathbb {D}}}\frac{(1-|w|^{2})}{(z-w)(z\overline{w}-1)}g(w)dA(w)\nonumber \\&-\,\frac{1}{4\pi }\int _{{\mathbb {D}}}\frac{(1-|w|^{2})}{w}g(w)dA(w)\bigg |\nonumber \\= & {} \bigg |\frac{1}{4\pi }\int _{{\mathbb {D}}}\frac{z(1-|w|^{2})(1+|w|^{2}-z\overline{w})}{w(z-w)(z\overline{w}-1)}g(w)dA(w)\bigg |\nonumber \\\le & {} \frac{M_{1}|z|}{4\pi }\int _{{\mathbb {D}}}\frac{(1-|w|^{2})\big |1+|w|^{2}-z\overline{w}\big |}{|w||z-w||1-z\overline{w}|}dA(w)\nonumber \\\le & {} \frac{|z|(2+|z|)M_{1}}{4\pi }\int _{{\mathbb {D}}}\frac{(1+|w|)}{|w||z-w|}dA(w)\nonumber \\\le & {} \frac{|z|(2+|z|)M_{1}}{2\pi }\int _{{\mathbb {D}}}\frac{1}{|w||z-w|}dA(w)\nonumber \\\le & {} M_{1}\big [\log 4(1+|z|)- \log |z|\big ]|z|(2+|z|). \end{aligned}$$
(2.19)

By a similar argument, we get

$$\begin{aligned} \left| \frac{\partial {\mathcal {G}}_{g}(z)}{\partial \overline{z}}-\frac{\partial {\mathcal {G}}_{g}(0)}{\partial \overline{z}}\right|= & {} \bigg |\frac{1}{4\pi }\int _{{\mathbb {D}}}\frac{(1-|w|^{2})}{(\overline{z}-\overline{w})(w\overline{z}-1)}g(w)dA(w)\nonumber \\&-\,\frac{1}{4\pi }\int _{{\mathbb {D}}}\frac{(1-|w|^{2})}{\overline{w}}g(w)dA(w)\bigg |\nonumber \\\le & {} M_{1}\big [\log 4(1+|z|)- \log |z|\big ]|z|(2+|z|). \end{aligned}$$
(2.20)

On the other hand, \({\mathcal {P}}_{f}\) can be written by

$$\begin{aligned} {\mathcal {P}}_{f}(z)=\sum _{n=0}^{\infty }a_{n}z^{n}+\sum _{n=1}^{\infty }\overline{b}_{n}\overline{z}^{n} \end{aligned}$$

because \({\mathcal {P}}_{f}\) is harmonic in \({\mathbb {D}}\).

Since \(|{\mathcal {P}}_{f}(z)|\le M_{2}\) for \(z\in {\mathbb {D}}\), by Lemma E, we have

$$\begin{aligned} |a_{n}|+|b_{n}|\le \frac{4M_{2}}{\pi } \end{aligned}$$
(2.21)

for \(n\ge 1\).

By (2.21), we see that

$$\begin{aligned} \left| \frac{\partial {\mathcal {P}}_{f}(z)}{\partial z}- \frac{\partial {\mathcal {P}}_{f}(0)}{\partial z}\right| +\left| \frac{\partial {\mathcal {P}}_{f}(z)}{\partial \overline{z}}- \frac{\partial {\mathcal {P}}_{f}(0)}{\partial \overline{z}}\right|= & {} \left| \sum _{n=2}^{\infty }na_{n}z^{n-1}\right| +\left| \sum _{n=2}^{\infty }nb_{n}\overline{z}^{n-1}\right| \nonumber \\\le & {} \sum _{n=2}^{\infty }n\big (|a_{n}|+|b_{n}|\big )|z|^{n-1} \nonumber \\\le & {} \frac{4M_{2}}{\pi }\sum _{n=2}^{\infty }n|z|^{n-1}\nonumber \\= & {} \frac{4M_{2}}{\pi }\frac{|z|(2-|z|)}{(1-|z|)^{2}}. \end{aligned}$$
(2.22)

Applying Theorem 3, we obtain

$$\begin{aligned} 1=J_{f}(0)=\Vert D_{f}(0)\Vert \lambda (D_{f}(0))\le \lambda (D_{f}(0))\left( \frac{4}{\pi }M_{2}+\frac{2}{3}M_{1}\right) , \end{aligned}$$

which gives that

$$\begin{aligned} \lambda (D_{f}(0))\ge \frac{1}{\frac{4}{\pi }M_{2}+\frac{2}{3}M_{1}}. \end{aligned}$$
(2.23)

In order to prove the univalence of f in \({\mathbb {D}}_{r_{0}}\), we choose two distinct points \(z_{1}, z_{2}\in {\mathbb {D}}_{r_{0}}\) and let \([z_{1},z_{2}]\) denote the segment from \(z_{1}\) to \(z_{2}\) with the endpoints \(z_{1}\) and \(z_{2}\), where \(r_{0}\) satisfies the following equation

$$\begin{aligned} \frac{1}{\frac{4}{\pi }M_{2}+\frac{2}{3}M_{1}}-\frac{4M_{2}}{\pi }\frac{r_{0}(2-r_{0})}{(1-r_{0})^{2}}-2M_{1}\big [\log 4(1+r_{0})- \log r_{0}\big ](2+r_{0})r_{0}=0. \end{aligned}$$

By (2.19), (2.20), (2.22), (2.23), Lemmas 3 and 4, we have

$$\begin{aligned} f(z_{2})-f(z_{1})|= & {} \left| \int _{[z_{1},z_{2}]}f_{z}(z)dz+f_{\overline{z}}(z)d\overline{z}\right| \nonumber \\\ge & {} \left| \int _{[z_{1},z_{2}]}f_{z}(0)dz+f_{\overline{z}}(0)d\overline{z}\right| \nonumber \\&-\left| \int _{[z_{1},z_{2}]}\big (f_{z}(z)-f_{z}(0)\big )dz+\big (f_{\overline{z}}(z)-f_{\overline{z}}(0)\big )d\overline{z}\right| \nonumber \\\ge & {} \lambda (D_{f}(0))|z_{2}-z_{1}|\nonumber \\&-\int _{[z_{1},z_{2}]}\big (|f_{z}(z)-f_{z}(0)|+|f_{\overline{z}}(z)-f_{\overline{z}}(0)|\big )|dz|\nonumber \\\ge & {} \lambda (D_{f}(0))|z_{2}-z_{1}|\nonumber \\&-\int _{[z_{1},z_{2}]}\left( \bigg |\frac{\partial {\mathcal {G}}_{g}(z)}{\partial z}-\frac{\partial {\mathcal {G}}_{g}(0)}{\partial z}\bigg |+\bigg |\frac{\partial {\mathcal {G}}_{g}(z)}{\partial \overline{z}}-\frac{\partial {\mathcal {G}}_{g}(0)}{\partial \overline{z}}\bigg |\right) |dz|\nonumber \\&-\int _{[z_{1},z_{2}]}\left( \bigg |\frac{\partial {\mathcal {P}}_{f}(z)}{\partial z}- \frac{\partial {\mathcal {P}}_{f}(0)}{\partial z}\bigg | +\bigg |\frac{\partial {\mathcal {P}}_{f}(z)}{\partial \overline{z}}- \frac{\partial {\mathcal {P}}_{f}(0)}{\partial \overline{z}}\bigg |\right) |dz|\nonumber \\> & {} |z_{2}-z_{1}|\big \{\lambda (D_{f}(0))-\frac{4M_{2}}{\pi }\frac{r_{0}(2-r_{0})}{(1-r_{0})^{2}}\nonumber \\&-\,2M_{1}\big [\log 4(1+r_{0})- \log r_{0}\big ](2+r_{0})r_{0}\big \}\nonumber \\\ge & {} |z_{2}-z_{1}|\bigg \{\frac{1}{\frac{4}{\pi }M_{2}+\frac{2}{3}M_{1}}-\frac{4M_{2}}{\pi }\frac{r_{0}(2-r_{0})}{(1-r_{0})^{2}}\nonumber \\&-\,2M_{1}\big [\log 4(1+r_{0})- \log r_{0}\big ](2+r_{0})r_{0}\bigg \}\nonumber \\= & {} 0, \end{aligned}$$
(2.24)

which yields that \(f(z_{2})\ne f(z_{1})\). The univalence of f follows from the arbitrariness of \(z_{1}\) and \(z_{2}\).

Now, for all \(\zeta =r_{0}e^{i\theta }\in \partial {\mathbb {D}}_{r_{0}}\), by (2.19), (2.20), (2.22), (2.23), Lemmas 3 and 4, we obtain

$$\begin{aligned} |f(\zeta )-f(0)|= & {} \left| \int _{[0,\zeta ]}f_{z}(z)dz+f_{\overline{z}}(z)d\overline{z}\right| \\= & {} \left| \int _{[0,\zeta ]}f_{z}(0)dz+f_{\overline{z}}(0)d\overline{z}\right| \\&-\left| \int _{[0,\zeta ]}\big (f_{z}(z)-f_{z}(0)\big )dz+\big (f_{\overline{z}}(z)-f_{\overline{z}}(0)\big )d\overline{z}\right| \\\ge & {} \lambda (D_{f}(0))r_{0}\\&-\int _{[0,\zeta ]}\big (|f_{z}(z)-f_{z}(0)|+|f_{\overline{z}}(z)-f_{\overline{z}}(0)|\big )|dz|\\\ge & {} \lambda (D_{f}(0))r_{0}\\&-\int _{[0,\zeta ]}\left( \bigg |\frac{\partial {\mathcal {G}}_{g}(z)}{\partial z}-\frac{\partial {\mathcal {G}}_{g}(0)}{\partial z}\bigg |+\bigg |\frac{\partial {\mathcal {G}}_{g}(z)}{\partial \overline{z}}-\frac{\partial {\mathcal {G}}_{g}(0)}{\partial \overline{z}}\bigg |\right) |dz|\\&-\int _{[0,\zeta ]}\left( \bigg |\frac{\partial {\mathcal {P}}_{f}(z)}{\partial z}- \frac{\partial {\mathcal {P}}_{f}(0)}{\partial z}\bigg | +\bigg |\frac{\partial {\mathcal {P}}_{f}(z)}{\partial \overline{z}}- \frac{\partial {\mathcal {P}}_{f}(0)}{\partial \overline{z}}\bigg |\right) |dz| \\\ge & {} \frac{r_{0}}{\frac{4}{\pi }M_{2}+\frac{2}{3}M_{1}}-\frac{4M_{2}}{\pi }\int _{[0,\zeta ]}\frac{|z|(2-|z|)}{(1-|z|)^{2}}|dz|\\&-\,2M_{1}\int _{[0,\zeta ]}\big [\log 4(1+|z|)- \log |z|\big ]|z|(2+|z|)|dz|\nonumber \\= & {} \frac{r_{0}}{\frac{4}{\pi }M_{2}+\frac{2}{3}M_{1}}-\frac{4M_{2}r_{0}^{2}}{\pi }\int _{0}^{1}\frac{t(2-r_{0}t)}{(1-r_{0}t)^{2}}dt\\&-\,2M_{1}r_{0}^{2}\int _{0}^{1}\big [\log 4(1+r_{0}t)- \log (r_{0}t)\big ]t(2+r_{0}t)dt\\\ge & {} \frac{r_{0}}{\frac{4}{\pi }M_{2}+\frac{2}{3}M_{1}}-\frac{4M_{2}r_{0}^{2}}{\pi }\frac{(2-r_{0})}{(1-r_{0})^{2}}\int _{0}^{1}t dt\\&-\,2M_{1}r_{0}^{2}(2+r_{0})\int _{0}^{1}\big [\log 4(1+r_{0}t)- \log (r_{0}t)\big ]t dt\\\ge & {} r_{0}\bigg \{\frac{1}{\frac{4}{\pi }M_{2}+\frac{2}{3}M_{1}}-\frac{2M_{2}}{\pi }\frac{r_{0}(2-r_{0})}{(1-r_{0})^{2}}\\&-\,2M_{1}r_{0}(2+r_{0})\big [\log 4(1+r_{0})- \log r_{0}\big ]\bigg \}\\= & {} \frac{2M_{2}}{\pi }\frac{r_{0}^{2}(2-r_{0})}{(1-r_{0})^{2}}. \end{aligned}$$

Hence \(f({\mathbb {D}}_{r_{0}})\) contains an univalent disk \({\mathbb {D}}_{R_{0}}\) with

$$\begin{aligned} R_{0}\ge \frac{2M_{2}}{\pi }\frac{r_{0}^{2}(2-r_{0})}{(1-r_{0})^{2}}. \end{aligned}$$

The proof of this theorem is complete. \(\square \)

Proof of Corollary 1

For \(z_{1}, z_{2}\in {\mathbb {D}}_{r_{0}}\), by (2.24), we see that there is a positive constant \(L_{1}\) such that

$$\begin{aligned} L_{1}|z_{1}-z_{2}|\le |f(z_{1})-f(z_{2})|, \end{aligned}$$

where \(r_{0}\) satisfies the following equation

$$\begin{aligned} \frac{1}{\frac{4}{\pi }M_{2}+\frac{2}{3}M_{1}}-\frac{4M_{2}}{\pi }\frac{r_{0}(2-r_{0})}{(1-r_{0})^{2}}-2M_{1}\big [\log 4(1+r_{0})- \log r_{0}\big ](2+r_{0})r_{0}=0. \end{aligned}$$

On the other hand, for \(z_{1}, z_{2}\in {\mathbb {D}}_{r_{0}}\), we use Theorem 3 to get

$$\begin{aligned} |f(z_{2})-f(z_{1})|= & {} \left| \int _{[z_{1},z_{2}]}df(z)\right| \\\le & {} \int _{[z_{1},z_{2}]}\Vert D_{f}(z)\Vert |dz|\\\le & {} \int _{[z_{1},z_{2}]}\left( \frac{4M_{2}}{\pi }\frac{1}{1-r_{0}^{2}}+\frac{2}{3}M_{1}\right) |dz|\\= & {} \left( \frac{4M_{2}}{\pi }\frac{1}{1-r_{0}^{2}}+\frac{2}{3}M_{1}\right) |z_{1}-z_{2}|, \end{aligned}$$

where \([z_{1},z_{2}]\) is the segment from \(z_{1}\) to \(z_{2}\) with the endpoints \(z_{1}\) and \(z_{2}\). Therefore, f is bi-Lipschitz in \({\mathbb {D}}_{r_{0}}\). \(\square \)