1 Introduction and main results

Let \(\mathbb {C}\) be the complex plane. For \(a\in \mathbb {C}\), let \(\mathbb {D}(a,r)=\{z:|z-a|<r\}\) \((r>0)\) and \(\mathbb {D}(0,r)=\mathbb {D}_{r}\). Also, we use the notations \(\mathbb {D}=\mathbb {D}_1\) and \(\mathbb {T}=\partial \mathbb {D}\), the boundary of \(\mathbb {D}\).

Let

$$\begin{aligned} A=\left( \begin{array}{cc} a &{} b \\ c &{} d \\ \end{array} \right) \in \mathbb {R}^{2\times 2}. \end{aligned}$$

We will consider the matrix norm

$$\begin{aligned} |A|=\sup \{|Az|:\; z\in \mathbb {C},\;|z|=1\} \end{aligned}$$

and the matrix function

$$\begin{aligned} l(A)=\inf \{|Az|:\;z\in \mathbb {C},\;|z|=1\}. \end{aligned}$$

Let D and \(\Omega \) be domains in \(\mathbb {C}\), and let \(f=u+iv\): \(D\rightarrow \Omega \) be a function that has both partial derivatives at \(z=x+iy\) in D, where u and v are real functions. The Jacobian matrix of f at z is denoted by

$$\begin{aligned} D f(z) = \left( \begin{array}{cc} u_x &{} u_y \\ v_x &{} v_y \\ \end{array} \right) . \end{aligned}$$

Set

$$\begin{aligned} \frac{\partial }{\partial {z}}=\frac{1}{2}\Big (\frac{\partial }{\partial {x}}-i\frac{\partial }{\partial {y}}\Big )\quad \hbox {and}\quad \frac{\partial }{\partial {\overline{z}}}=\frac{1}{2}\Big (\frac{\partial }{\partial {x}}+i\frac{\partial }{\partial {y}}\Big ). \end{aligned}$$

Then

$$\begin{aligned} |D f(z)|= & {} \sup \{|D f(z)\varsigma |:\; |\varsigma |=1\}=|f_z(z)|+|f_{\overline{z}}(z)|, \end{aligned}$$
(1.1)
$$\begin{aligned} l(D f(z))= & {} \inf \{|D f(z)\varsigma |:\; |\varsigma |=1\}=\big ||f_z(z)|-|f_{\overline{z}}(z)|\big | \end{aligned}$$
(1.2)

and

$$\begin{aligned} |J_{f}(z)|=|D f(z)|\cdot l(D f(z)), \end{aligned}$$

where \(J_{f}(z)\) stands for the Jacobian of f at z.

We denote by \(\Delta _{\alpha }\) the weighted Laplace operator corresponding to the so-called standard weight \(w_{\alpha }=(1-|z|^{2})^{\alpha }\), that is,

$$\begin{aligned} \Delta _{\alpha ,z}=\frac{\partial }{\partial z}(w_{\alpha })^{-1}\frac{\partial }{\partial \overline{z}}=\frac{\partial }{\partial z}(1-{|{z}|}^{2})^{-\alpha }\frac{\partial }{\partial \overline{z}} \end{aligned}$$

in \(\mathbb {D}\), where \(\alpha > -1\) (see [20, Proposition 1.5] for the reason for this constraint).

Olofsson and Wittsten [20] introduced this operator \(\Delta _{\alpha }\) and a counterpart of the classical Poisson integral formula was given.

We remark that in the study of Bergman spaces of \(\mathbb {D}\), one often considers the weights \(w_{\alpha }\) in \(\mathbb {D}\) \((\alpha >-1)\). For an account of recent developments in Bergman space theory, we mention the monograph by Hedenmalm et al. [13]. The case \(\alpha = 0\) is commonly referred to as the unweighted case, whereas the case \(\alpha = 1\) has attracted special attention recently with contributions by Hedenmalm, Shimorin and others (see for instance [14,15,16, 22] etc).

Of particular interest to us is the following \(\alpha \)-harmonic equation in \(\mathbb {D}\):

$$\begin{aligned} \Delta _{\alpha }(f)=0. \end{aligned}$$
(1.3)

Denote the associated Dirichlet boundary value problem of functions f satisfying the Eq. (1.3) by

$$\begin{aligned} \left\{ \begin{array}{ll} \Delta _{\alpha }(f)=0&{}\quad \text {in}\;\mathbb {D},\\ f=f^{*}&{}\quad \text {on}\;\mathbb {T}. \end{array} \right. \end{aligned}$$
(1.4)

Here the boundary data \(f^{*}\) is a distribution on \(\mathbb {T}\), i.e. \(f^{*}\in \mathcal {D}^{\prime }(\mathbb {T})\), and the boundary condition in the Eq. (1.4) is to be understood as \(f_{r}\rightarrow f^{*}\in \mathcal {D}^{\prime }(\mathbb {T})\) as \(r \rightarrow 1^{-}\), where

$$\begin{aligned} f_{r}(e^{i\theta })=f(re^{i\theta }) \end{aligned}$$
(1.5)

for \(\theta \in [0, 2\pi ]\) and \(r\in [0,1)\).

For simplicity, we introduce the following definition.

Definition 1.1

For \(\alpha >-1\), a complex-valued function f is said to be \(\alpha \)-harmonic if f is twice continuously differentiable in \(\mathbb {D}\) and satisfies the condition (1.3).

Olofsson and Wittsten [20] showed that if an \(\alpha \)-harmonic function f satisfies

$$\begin{aligned} \lim _{r\rightarrow 1^{-}}f_{r}=f^{*} \in \mathcal {D}^{\prime }(\mathbb {T})\;\; (\alpha > -1), \end{aligned}$$

then it has the form of a Poisson type integral

$$\begin{aligned} f(z)=\frac{1}{2\pi }\int ^{2\pi }_{0}\mathcal {P}_{\alpha } (ze^{-i\theta })f^{*}(e^{i\theta })d\theta \end{aligned}$$
(1.6)

in \(\mathbb {D}\), where

$$\begin{aligned} \mathcal {P}_{\alpha }(z)=\frac{(1-|z|^{2})^{\alpha +1}}{(1-z) (1-\overline{z})^{\alpha +1}}. \end{aligned}$$

In the following, we always assume that any \(\alpha \)-harmonic function has such a representation which plays a key role in the discussions of this paper.

Obviously, \(\alpha \)-harmonicity coincides with harmonicity when \(\alpha = 0\). See [12] and the references therein for the properties of harmonic mappings. Particularly, Colonna proved the following Schwarz–Pick type inequality.

Theorem A

[11, Theorems 3 and 4] Let f be a harmonic function of \(\mathbb {D}\) into \(\mathbb {D}\). Then for \(z\in \mathbb {D}\),

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

This estimate is sharp and all the extremal functions are

$$\begin{aligned} f(z)=\frac{2\delta }{\pi }\arg \Big (\frac{1+\psi (z)}{1-\psi (z)}\Big ), \end{aligned}$$

where \(\delta \in \mathbb {C}\), \(|\delta |=1\) and \(\psi \) is a conformal automorphism of \(\mathbb {D}\).

For the related discussions on this topic, see [2, 4, 7, 10, 17, 21] etc.

As the first aim of this paper, we shall generalize Theorem A to the case of \(\alpha \)-harmonic functions. Our first result is as follows.

Theorem 1.1

Suppose that f is an \(\alpha \)-harmonic function in \(\mathbb {D}\) with \(\alpha >-1\), that \(f^{*}\in C(\mathbb {T})\) and that \(\sup _{z\in \overline{\mathbb {D}}}|f(z)|\le M\), where M is a constant. Then for \(z\in \mathbb {D}\),

$$\begin{aligned} |D{f}(z)|\le \frac{M(\alpha +2)}{\mathrm {c}_{\alpha }}\cdot \frac{1}{1-|z|}\le \frac{2M(\alpha +2)}{\mathrm {c}_{\alpha }}\cdot \frac{1}{1-|z|^2}, \end{aligned}$$

where \(\mathrm {c}_{\alpha }=\frac{(\Gamma (\frac{\alpha }{2}+1))^{2}}{\Gamma (\alpha +1)}\) and \(\Gamma (s)=\int ^{\infty }_{0}t^{s-1}e^{-t}dt\) \((s>0)\) is the Gamma function.

In particular, if f maps \(\mathbb {D}\) into \(\mathbb {D}\), then

$$\begin{aligned} |D{f}(z)|\le \frac{2(\alpha +2)}{\mathrm {c}_{\alpha }}\cdot \frac{1}{1-|z|^2}. \end{aligned}$$

Let \(\lambda _{D}(z)|dz|\) be the hyperbolic metric of the domain D having constant Gaussian curvature \(-1\). The hyperbolic distance \(d_{h_D}(z_1, z_2)\) between two points \(z_1\) and \(z_2\) in D is defined by

$$\begin{aligned} \inf _{\gamma }\Big \{\int _{\gamma }\lambda _{D}(z)|dz|\Big \}, \end{aligned}$$

where the infimum is taken over all rectifiable curves \(\gamma \) in D connecting \(z_1\) and \(z_2\).

We have known that if \(D=\mathbb {D}\), then (cf. [1])

$$\begin{aligned} \lambda _{\mathbb {D}}(z)=\frac{2}{1-|z|^2}\;\;\hbox {and}\;\;d_{h_\mathbb {D}}(z_1, z_2)=\log \frac{|1-z_1\overline{z}_2|+|z_1-z_2|}{|1-z_1\overline{z}_2|-|z_1-z_2|}. \end{aligned}$$

As a consequence of Theorem 1.1, we have

Corollary 1.1

Under the assumptions of Theorem 1.1, if f maps \(\mathbb {D}\) into \(\mathbb {D}\), then for \(z_1\) and \(z_2\in \mathbb {D}\),

$$\begin{aligned} |f(z_1)-f(z_2)|\le \frac{\alpha +2}{\mathrm {c}_{\alpha }}d_{h_\mathbb {D}}(z_1, z_2). \end{aligned}$$

In [20], the authors got the following homogeneous expansion of \(\alpha \)-harmonic functions (see [20, Theorem 1.2]):

A function f in \(\mathbb {D}\) is \(\alpha \)-harmonic if and only if it has the following convergent power series expansion:

$$\begin{aligned} f(z)=\sum ^{\infty }_{k=0}c_{k}z^{k}+\sum ^{\infty }_{k=1}c_{-k}P_{\alpha ,k}(|z|^2)\overline{z}^{k}, \end{aligned}$$
(1.7)

where \(P_{\alpha ,k}(x)=\int ^{1}_{0}t^{k-1}(1-tx)^{\alpha }dt\) \((-1<x<1)\) and \(\{c_{k}\}^{\infty }_{k=-\infty }\) denotes a sequence of complex numbers with \( \lim _{|k|\rightarrow \infty }\sup |c_{k}|^{\frac{1}{|k|}}\le 1. \)

The second aim of this paper is to prove the following estimates on coefficients \(c_{k}\) and \(c_{-k}\).

Theorem 1.2

Suppose that f is an \(\alpha \)-harmonic function in \(\mathbb {D}\) with \(\alpha >-1\) and that \(\sup _{z\in \mathbb {D}}|f(z)|\le M\), where M is a constant. If f has the series expansion (1.7), then for \(k\in \{0,1,2,...\},\)

$$\begin{aligned} |c_{k}|\le M, \end{aligned}$$
(1.8)

and for \(k\in \{1,2,...\},\)

$$\begin{aligned} |c_{k}|+|c_{-k}|B(k, \alpha +1)\le \frac{4M}{\pi }, \end{aligned}$$
(1.9)

where B(pq) denotes the Beta function.

By [20, Definition 2.1], we find that

$$\begin{aligned} \mathcal {P}_{\alpha }(ze^{-i\theta })=\sum ^{\infty }_{k=0}e^{-ik\theta }z^{k} +\sum ^{\infty }_{k=1}\frac{\Gamma (k+\alpha +1)}{\Gamma (k)\Gamma (\alpha +1)}P_{\alpha ,k}(|z|^2)e^{ik\theta }\overline{z}^{k}. \end{aligned}$$

If \(|f^{*}(z)|\le M\), then by (1.6), we get

$$\begin{aligned} |c_{-k}|=\left| \frac{\Gamma (k+\alpha +1)}{\Gamma (k)\Gamma (\alpha +1)}\frac{1}{2\pi }\int ^{2\pi }_{0}e^{ik\theta }f^{*}(e^{i\theta })d\theta \right| \le M\frac{\Gamma (k+\alpha +1)}{\Gamma (k)\Gamma (\alpha +1)}\rightarrow \infty \end{aligned}$$

as \(k\rightarrow \infty \).

Moreover, from the proof of [20, Theorem 1.2], we see that

$$\begin{aligned} (1-{|{z}|}^{2})^{-\alpha }\frac{\partial }{\partial \overline{z}}f(z)=\overline{h(z)}, \end{aligned}$$

where \(h(z)=\sum ^{\infty }_{k=0}a_{k}z^k\), \(z\in \mathbb {D}\) and \(c_{-k}=\overline{a_{k-1}}\) for \(k\ge 1\). Note that if h(z) is a normalized (in the sense that \(h(0)=h'(0)-1=0\)) univalent analytic function in \(\mathbb {D}\), then by Louis de Branges’s theorem it is well-known that \(|a_{k}|\le k\) for all \(k\ge 2\) so that

$$\begin{aligned} c_{-1}=0,\quad c_{-2}=1\quad \hbox {and}\quad |c_{-k}|=|a_{k-1}|\le k-1\quad \hbox {for all}\;\;k\ge 3. \end{aligned}$$
(1.10)

The classical Landau theorem says that there is a \(\rho =\frac{1}{M+\sqrt{M^{2}-1}}\) such that every function f, analytic in \(\mathbb {D}\) with \(f(0)=f^{'}(0)-1=0\) and \(|f(z)|< M\), is univalent in the disk \(\mathbb {D}_{\rho }\). Moreover, the range \(f(\mathbb {D}_{\rho })\) contains a disk of radius \(M\rho ^{2}\), where \(M\ge 1\) is a constant (see [18]). Recently, many authors considered Landau type theorem for \(\alpha \)-harmonic functions f when \(\alpha =0\) (see [3, 5,6,7,8,9] etc).

As an application of Theorems 1.1 and 1.2, we get the following Landau type theorem for \(\alpha \)-harmonic functions.

Theorem 1.3

Suppose that f is an \(\alpha \)-harmonic function in \(\mathbb {D}\) with \(\alpha \ge 0\), that \(f^{*}\in C(\mathbb {T})\), that \(\sup _{z\in \overline{\mathbb {D}}}|f(z)|\le M\), where M is a constant, and that \(f(0)=|J_{f}(0)|-\beta =0\). If f satisfies (1.10), then we have the following:

  1. (1)

    f is univalent in \(\mathbb {D}_{\rho _{0}}\), where \(\rho _{0}\) satisfies the following equation

    $$\begin{aligned} \frac{\beta \mathrm {c}_{\alpha }}{M(\alpha +2)}-(M+5)\,\frac{\rho _{0}(2-\rho _{0})}{(1-\rho _{0})^{2}}=0; \end{aligned}$$
    (1.11)
  2. (2)

    \(f(\mathbb {D}_{\rho _{0}})\) contains a univalent disk \(\mathbb {D}_{R_{0}}\) with

    $$\begin{aligned} R_{0}\ge (M+5)\,\Big (\frac{\rho _{0}}{1-\rho _{0}}\Big )^2. \end{aligned}$$

The arrangement of the rest of this paper is as follows. In Sect. 2, we shall prove Theorem 1.1 and Corollary 1.1. Section 3 will be devoted to the proof of Theorem 1.2. In Sect. 4, Theorem 1.3 will be demonstrated.

2 Schwarz–Pick type inequality

The aim of this section is to prove Theorem 1.1 and Corollary 1.1. The proofs need a result from [19]. Before the statement of this result, we do some preparation.

In [19], the author considered the following integral means:

$$\begin{aligned} \mathcal {M}_{\alpha }(r)= \frac{1}{2\pi } \int _{0}^{2\pi }K_{\alpha }(re^{i\theta })d\theta , \end{aligned}$$
(2.1)

where \(r\in [0, 1)\) and

$$\begin{aligned} K_{\alpha }(z)=\mathrm {c}_{\alpha }|\mathcal {P}_{\alpha }(z)|=\mathrm {c}_{\alpha }\frac{(1-|z|^{2})^{\alpha +1}}{|1-z|^{\alpha +2}} \end{aligned}$$

in \(\mathbb {D}\).

Let us recall the following result from [19].

Theorem B

[19, Theorem 3.1] Let \(\alpha >-1\). The integral means function \(\mathcal {M}_{\alpha }(r)\) given by (2.1) satisfies the following assertions.

  1. (1)

    \(\lim _{r\rightarrow 1^{-}}\mathcal {M}_{\alpha }(r)=1;\)

  2. (2)

    \(\mathcal {M}_{\alpha }^{(n)}(r)\ge 0\) for \(r\in [0,1)\) and \(n\ge 0\).

The following result also plays a key role in the proof of Theorem 1.1.

Lemma 2.1

If \(\alpha >-1\) and \(f^{*}\in C(\mathbb {T})\), then

$$\begin{aligned} \frac{\partial }{\partial z}\int ^{2\pi }_{0}\mathcal {P}_{\alpha }(ze^{-i\theta })f^{*}(e^{i\theta })d\theta =\int ^{2\pi }_{0}\frac{\partial }{\partial z}\mathcal {P}_{\alpha }(ze^{-i\theta })f^{*}(e^{i\theta })d\theta \end{aligned}$$

and

$$\begin{aligned} \frac{\partial }{\partial \overline{z}}\int ^{2\pi }_{0}\mathcal {P}_{\alpha }(ze^{-i\theta })f^{*}(e^{i\theta })d\theta =\int ^{2\pi }_{0}\frac{\partial }{\partial \overline{z}}\mathcal {P}_{\alpha }(ze^{-i\theta })f^{*}(e^{i\theta })d\theta . \end{aligned}$$

Proof

By elementary calculations we see that the following equalities hold:

$$\begin{aligned} \frac{\partial }{\partial {z}}\mathcal {P}_{\alpha }(ze^{-i\theta }) =\frac{(1-|z|^{2})^{\alpha }\left[ e^{-i\theta }(1-|z|^{2})-(\alpha +1)\overline{z}(1-ze^{-i\theta })\right] }{(1-ze^{-i\theta })^{2}(1-\overline{z}e^{i\theta })^{\alpha +1}} \end{aligned}$$
(2.2)

and

$$\begin{aligned} \frac{\partial }{\partial {\overline{z}}}\mathcal {P}_{\alpha }(ze^{-i\theta })=\frac{(\alpha +1)(1-|z|^{2})^{\alpha }e^{i\theta }}{(1-\overline{z}e^{i\theta })^{\alpha +2}}. \end{aligned}$$
(2.3)

Then we know that functions

$$\begin{aligned} \frac{\partial }{\partial z}\mathcal {P}_{\alpha }(ze^{-i\theta })f^{*}(e^{i\theta })\quad \hbox {and}\quad \frac{\partial }{\partial \overline{z}}\mathcal {P}_{\alpha }(ze^{-i\theta })f^{*}(e^{i\theta }) \end{aligned}$$

are continuous on \(\overline{\mathbb {D}}_r\times [0, 2\pi ]\), where \(r\in [0, 1)\).

Let \(z=\rho e^{i\varphi }\in \overline{\mathbb {D}}_r\). It follows from

$$\begin{aligned} \frac{\partial }{\partial \rho } \mathcal {P}_{\alpha }(ze^{-i\theta })=\frac{\partial }{\partial z} \mathcal {P}_{\alpha }(ze^{-i\theta })e^{i\varphi } +\frac{\partial }{\partial \overline{z}}\mathcal {P}_{\alpha }(ze^{-i\theta })e^{-i\varphi } \end{aligned}$$

and

$$\begin{aligned} \frac{\partial }{\partial \varphi }\mathcal {P}_{\alpha }(ze^{-i\theta })=\frac{\partial }{\partial z} \mathcal {P}_{\alpha }(ze^{-i\theta })iz -\frac{\partial }{\partial \overline{z}}\mathcal {P}_{\alpha }(ze^{-i\theta })i\overline{z} \end{aligned}$$

that both

$$\begin{aligned} \frac{\partial }{\partial \rho } \mathcal {P}_{\alpha }(ze^{-i\theta })f(e^{i\theta })\;\;\hbox { and} \;\;\frac{\partial }{\partial \varphi }\mathcal {P}_{\alpha }(ze^{-i\theta })f(e^{i\theta }) \end{aligned}$$

are continuous in \(\overline{\mathbb {D}}_r\times [0, 2\pi ]\). Hence

$$\begin{aligned} \int ^{\rho }_{0}\int ^{2\pi }_{0}\frac{\partial }{\partial \rho } \mathcal {P}_{\alpha }(ze^{-i\theta })f^{*}(e^{i\theta })d\theta d\rho= & {} \int ^{2\pi }_{0}\int ^{\rho }_{0}\frac{\partial }{\partial \rho } \mathcal {P}_{\alpha }(ze^{-i\theta })f^{*}(e^{i\theta })d\rho d\theta \\= & {} \int ^{2\pi }_{0}\big (\mathcal {P}_{\alpha }(ze^{-i\theta })-\mathcal {P}_{\alpha }(0)\big )f^{*}(e^{i\theta })d\theta \end{aligned}$$

and

$$\begin{aligned} \int ^{\varphi }_{0}\int ^{2\pi }_{0}\frac{\partial }{\partial \varphi }\mathcal {P}_{\alpha }(ze^{-i\theta })f^{*}(e^{i\theta })d\theta d\varphi= & {} \int ^{2\pi }_{0}\int ^{\varphi }_{0}\frac{\partial }{\partial \varphi }\mathcal {P}_{\alpha }(ze^{-i\theta })f^{*}(e^{i\theta })d\varphi d\theta \\= & {} \int ^{2\pi }_{0}\big (\mathcal {P}_{\alpha }(ze^{-i\theta })-\mathcal {P}_{\alpha }(\rho e^{-i\theta })\big )f^{*}(e^{i\theta })d\theta . \end{aligned}$$

By differentiating with respect to \(\rho \) and \(\varphi \), respectively, we get

$$\begin{aligned} \int ^{2\pi }_{0}\frac{\partial }{\partial \rho } \mathcal {P}_{\alpha }(ze^{-i\theta })f^{*}(e^{i\theta })d\theta =\frac{\partial }{\partial \rho }\int ^{2\pi }_{0}\mathcal {P}_{\alpha }(ze^{-i\theta })f^{*}(e^{i\theta })d\theta \end{aligned}$$
(2.4)

and

$$\begin{aligned} \int ^{2\pi }_{0}\frac{\partial }{\partial \varphi }\mathcal {P}_{\alpha }(ze^{-i\theta })f^{*}(e^{i\theta })d\theta =\frac{\partial }{\partial \varphi }\int ^{2\pi }_{0}\mathcal {P}_{\alpha }(ze^{-i\theta })f^{*}(e^{i\theta })d\theta . \end{aligned}$$
(2.5)

Since

$$\begin{aligned} \frac{\partial }{\partial z} \mathcal {P}_{\alpha }(ze^{-i\theta })=\frac{e^{-i\varphi }}{2}\left( \frac{\partial }{\partial \rho } \mathcal {P}_{\alpha }(ze^{-i\theta })-\frac{i}{\rho }\frac{\partial }{\partial \varphi }\mathcal {P}_{\alpha }(ze^{-i\theta })\right) \end{aligned}$$

and

$$\begin{aligned} \frac{\partial }{\partial \overline{z}}\mathcal {P}_{\alpha }(ze^{-i\theta })=\frac{e^{i\varphi }}{2}\left( \frac{\partial }{\partial \rho } \mathcal {P}_{\alpha }(ze^{-i\theta })+\frac{i}{\rho }\frac{\partial }{\partial \varphi }\mathcal {P}_{\alpha }(ze^{-i\theta })\right) , \end{aligned}$$

it follows from (2.4) and (2.5) that the proof of the lemma is complete. \(\square \)

Now, we are ready to present the proofs of Theorem 1.1 and Corollary 1.1.

Proof of Theorem 1.1

From (2.2) and (2.3), we can easily get

$$\begin{aligned} \left| \frac{\partial }{\partial {z}}\mathcal {P}_{\alpha }(ze^{-i\theta })\right| \le \frac{1}{\mathrm {c}_{\alpha }}\cdot \frac{(\alpha +2)|z|+1}{1-|z|^{2}}K_{\alpha }(ze^{-i\theta }) \end{aligned}$$

and

$$\begin{aligned} \left| \frac{\partial }{\partial {\overline{z}}}\mathcal {P}_{\alpha }(ze^{-i\theta })\right| = \frac{\alpha +1}{\mathrm {c}_{\alpha }}\cdot \frac{1}{1-|z|^{2}}K_{\alpha }(ze^{-i\theta }). \end{aligned}$$

In the first inequality above, the fact “\(1-|z|\le |1-ze^{-i\theta }|\)” is applied. By (1.1), (1.6) and Lemma 2.1 yield

$$\begin{aligned} |D{f}(z)|=\left| \frac{1}{2\pi }\int _{0}^{2\pi }\frac{\partial }{\partial {z}}\mathcal {P}_{\alpha }(ze^{-i\theta })f^{*}(e^{i\theta })d\theta \right| +\left| \frac{1}{2\pi }\int _{0}^{2\pi }\frac{\partial }{\partial {\overline{z}}}\mathcal {P}_{\alpha }(ze^{-i\theta })f^{*}(e^{i\theta })d\theta \right| , \end{aligned}$$

we see from (2.1) and Theorem B that

$$\begin{aligned} |D{f}(z)| \le \frac{M(\alpha +2)}{\mathrm {c}_{\alpha }}\cdot \frac{1}{1-|z|}\mathcal {M}_{\alpha }(|z|) \le \frac{M(\alpha +2)}{\mathrm {c}_{\alpha }}\cdot \frac{1}{1-|z|}, \end{aligned}$$

and so the proof of Theorem 1.1 is complete. \(\square \)

Proof of Corollary 1.1

For any \(z_1\) and \(z_2\in \mathbb {D}\), let \(\gamma \) be the hyperbolic geodesic connecting \(z_1\) and \(z_2\). It follows from Theorem 1.1 that

$$\begin{aligned} |f(z_1)-f(z_2)|\le \int _{\gamma }|D{f}(z)|\cdot |dz| \le \frac{\alpha +2}{\mathrm {c}_{\alpha }}\int _{\gamma }\frac{2}{1-|z|^2}|dz| = \frac{\alpha +2}{\mathrm {c}_{\alpha }}d_{h_\mathbb {D}}(z_1, z_2), \end{aligned}$$

as required. \(\square \)

3 Estimates on coefficients

The aim of this paper is to prove Theorem 1.2. We start with a lemma.

Lemma 3.1

Under the assumptions of Theorem 1.2, if f has the series expansion (1.7), then

  1. (1)

    \(|c_k|\le M\) for \(k\ge 0\);

  2. (2)

    \((|c_{k}|+|c_{-k}|P_{\alpha ,k}(r^{2}))r^{k}\le \frac{4}{\pi }M\) for \(k>0\) and \(r\in (0, 1)\).

Proof

If \(k\not =0\), let \(z=re^{i\theta }\in {\mathbb {D}}\). Then by (1.7), we have

$$\begin{aligned} c_{k}r^{k}=\frac{1}{2\pi }\int _{0}^{2\pi }f(re^{i\theta })e^{-ik\theta }d{\theta }\quad \hbox {and}\quad c_{-k}P_{\alpha ,k}(r^{2})r^{k}=\frac{1}{2\pi }\int _{0}^{2\pi }f(re^{i\theta })e^{ik\theta }d{\theta }. \end{aligned}$$

Letting \(c_k=|c_k|e^{i\mu _k}\) and \(c_{-k}=|c_{-k}|e^{i\nu _k}\) leads to

$$\begin{aligned} (|c_{k}|+|c_{-k}|P_{\alpha ,k}(r^{2}))r^{k}= & {} \frac{1}{2\pi }\int ^{2\pi }_{0}f(re^{i\theta })\left( e^{-i(k\theta +\mu _k)}+e^{i(k\theta -\nu _k)}\right) d{\theta } \nonumber \\\le & {} \frac{1}{2\pi }\int ^{2\pi }_{0}|f(re^{i\theta })|\cdot \left| e^{-i(k\theta +\mu _k)}+e^{i(k\theta -\nu _k)}\right| d{\theta } \nonumber \\\le & {} \frac{M}{\pi }\int ^{2\pi }_{0}\left| \cos \left( k\theta +\frac{\mu _k-\nu _k}{2}\right) \right| d{\theta }, \end{aligned}$$

and so [4, Lemma 1] gives

$$\begin{aligned} (|c_{k}|+|c_{-k}|P_{\alpha ,k}(r^{2}))r^{k} \le \frac{4M}{\pi }. \end{aligned}$$

Thus the assertion (2) in the lemma is true.

To prove the assertions (1), we first recall from [20, Definition 2.1] that

$$\begin{aligned} \mathcal {P}_{\alpha }(ze^{-i\theta })=\sum ^{\infty }_{k=0}e^{-ik\theta }z^{k} +\sum ^{\infty }_{k=1}\frac{\Gamma (k+\alpha +1)}{\Gamma (k)\Gamma (\alpha +1)}P_{\alpha ,k}(|z|^2)e^{ik\theta }\overline{z}^{k}. \end{aligned}$$

Then by (1.6), we get

$$\begin{aligned} f(z)= & {} \sum ^{\infty }_{k=0}z^k\frac{1}{2\pi }\int ^{2\pi }_{0}e^{-ik\theta }f^{*}(e^{i\theta })d{\theta }\\&+\sum ^{\infty }_{k=1}\frac{\Gamma (k+\alpha +1)}{\Gamma (k)\Gamma (\alpha +1)}P_{\alpha ,k}(|z|^2)\overline{z}^{k} \frac{1}{2\pi }\int ^{2\pi }_{0}e^{ik\theta }f^{*}(e^{i\theta })d{\theta }, \end{aligned}$$

which implies

$$\begin{aligned} |c_k|=\left| \frac{1}{2\pi }\int ^{2\pi }_{0}e^{-ik\theta }f^{*}(e^{i\theta })d{\theta }\right| \le M \end{aligned}$$

as required. \(\square \)

Proof of Theorem 1.2

To prove this theorem, by Lemma 3.1, we only need to check (1.9) in the theorem. By letting \(r\rightarrow 1^{-}\) in Lemma 3.1(2), we see that the inequalities (1.9) easily follows. \(\square \)

4 Landau type theorem

This section consists of two subsections. In the first subsection, we shall prove an auxiliary result. In the second subsection, Theorem 1.3 will be checked.

4.1 A lemma

Lemma 4.1

For constants \(\alpha >-2\), \(\beta >0\) and \(M>0\), let

$$\begin{aligned} \varphi (x)=\frac{\beta \mathrm {c}_{\alpha }}{M(\alpha +2)}+(M+5)\,\frac{x(x-2)}{(1-x)^{2}} \end{aligned}$$

in [0, 1). Then

  1. (1)

    \(\varphi \) is continuous in [0, 1) and strictly decreasing in (0, 1);

  2. (2)

    there is a unique \(x_{0}\in (0,1)\) such that \(\varphi (x_{0})=0\).

Proof

For \(x\in [0,1)\), obviously,

$$\begin{aligned} \varphi '(x)=-\frac{2(M+5)}{(1-x)^{3}}<0. \end{aligned}$$

Hence \(\varphi (x)\) is continuous and strictly decreasing in [0, 1). It follows from

$$\begin{aligned} \varphi (0)=\frac{\beta \mathrm {c}_{\alpha }}{M(\alpha +2)}>0 \quad \text {and}\quad \lim _{x\rightarrow 1^{-}}\varphi (x)=-\infty <0 \end{aligned}$$

that there is a unique \(x_{0}\in (0,1)\) such that \(\varphi (x_{0})=0\). The proof of this lemma is complete. \(\square \)

4.2 Proof of Theorem 1.3

To prove this theorem, we need estimates on two quantities \(|f_{z}(z)-f_{z}(0)|+|f_{\overline{z}}(z)-f_{\overline{z}}(0)|\) and l(D f(0)). First, we estimate \(|f_{z}(z)-f_{z}(0)|+|f_{\overline{z}}(z)-f_{\overline{z}}(0)|\). Obviously, by (1.7), we see that

$$\begin{aligned} f_{z}(z)-f_{z}(0)=\sum _{k=2}^{\infty }kc_{k}z^{k-1}+\sum _{k=2}^{\infty }c_{-k}\frac{d}{dw}P_{\alpha , k}(w)\overline{z}^{k+1} \end{aligned}$$

and

$$\begin{aligned} f_{\overline{z}}(z)-f_{\overline{z}}(0)=\sum _{k=2}^{\infty }kc_{-k}P_{\alpha , k}(w)\overline{z}^{k-1}+\sum _{k=2}^{\infty }c_{-k}\frac{d}{dw}P_{\alpha , k}(w)z\overline{z}^{k}, \end{aligned}$$

where \(w=|z|^{2}\).

Since

$$\begin{aligned} \frac{d}{dw}P_{\alpha ,k}(w)=-\int ^{1}_{0}t^{k}\alpha (1-tw)^{\alpha -1}dt\le 0, \end{aligned}$$

we get that

$$\begin{aligned} P_{\alpha ,k}(w)\le P_{\alpha ,k}(0)=\frac{1}{k}. \end{aligned}$$
(4.1)

Moreover, since

$$\begin{aligned} P_{\alpha , k}(w)=\frac{1}{w^{k}}\int ^{w}_{0}x^{k-1}(1-x)^{\alpha }dx, \end{aligned}$$

we easily get

$$\begin{aligned} \frac{d}{dw}P_{\alpha ,k}(w)=-\frac{k}{w}P_{\alpha ,k}(w)+\frac{(1-w)^\alpha }{w}. \end{aligned}$$
(4.2)

Then (1.9), (1.10), (4.1) and (4.2) guarantee that

$$\begin{aligned} |f_{z}(z)-f_{z}(0)|+|f_{\overline{z}}(z)-f_{\overline{z}}(0)|\le & {} \sum _{k=2}^{\infty }k\big (|c_{k}|+|c_{-k}|P_{\alpha ,k}(w)\big )|z|^{k-1}\nonumber \\&+2\sum _{k=2}^{\infty }|c_{-k}|\big (kP_{\alpha ,k}(w)+1\big )|z|^{k-1} \nonumber \\\le & {} (M+5)\sum _{k=2}^{\infty }k|z|^{k-1} \nonumber \\= & {} (M+5)\,\frac{|z|(2-|z|)}{(1-|z|)^{2}}, \end{aligned}$$
(4.3)

which is what we want.

Next, we estimate l(D f(0)). Applying Theorem 1.1 leads to

$$\begin{aligned} \beta =|J_{f}(0)|=|D{f}(0)|\;l(D{f}(0))\le \frac{M(\alpha +2)}{\mathrm {c}_{\alpha }}\;l(D{f}(0)), \end{aligned}$$

which gives

$$\begin{aligned} l(D{f}(0))\ge \frac{\beta \mathrm {c}_{\alpha }}{M(\alpha +2)}. \end{aligned}$$
(4.4)

Now, we are ready to finish the proof of the theorem. First, we demonstrate the univalence of f in \(\mathbb {D}_{\rho _{0}}\), where \(\rho _{0}\) is determined by Eq. (1.11). For this, let \(z_{1}\), \(z_{2}\) be two points in \(\mathbb {D}_{\rho _{0}}\) with \(z_{1}\ne z_{2}\), and denote the segment from \(z_{1}\) to \(z_{2}\) with the endpoints \(z_{1}\) and \(z_{2}\) by \([z_{1},z_{2}]\). Since

$$\begin{aligned} |f(z_{2})-f(z_{1})|= & {} \left| \int _{[z_{1},z_{2}]}f_{z}(z)dz+f_{\overline{z}}(z)d\overline{z}\right| \nonumber \\\ge & {} \left| \int _{[z_{1},z_{2}]}f_{z}(0)dz+f_{\overline{z}}(0)d\overline{z}\right| \nonumber \\&-\left| \int _{[z_{1},z_{2}]}[f_{z}(z)-f_{z}(0)]dz+[f_{\overline{z}}(z)-f_{\overline{z}}(0)]d\overline{z}\right| , \end{aligned}$$

we see from (4.3), (4.4) and Lemma 4.1 that

$$\begin{aligned} |f(z_{2})-f(z_{1})|\ge & {} l(D{f}(0))\cdot |z_{2}-z_{1}|-(M+5)\int _{0}^{|z_{2}-z_{1}|}\frac{|z|(2-|z|)}{(1-|z|)^{2}}|dz| \nonumber \\ {}> & {} \left[ \frac{\beta \mathrm {c}_{\alpha }}{M(\alpha +2)}-(M+5)\,\frac{\rho _{0}(2-\rho _{0})}{(1-\rho _{0})^{2}}\right] |z_{2}-z_{1}| \nonumber \\= & {} 0. \end{aligned}$$

Thus, for arbitrary \(z_{1}\) and \(z_{2}\in \mathbb {D}_{\rho _{0}}\) with \(z_{1}\ne z_{2}\), we have

$$\begin{aligned} f(z_{1})\ne f(z_{2}), \end{aligned}$$

which implies the univalence of f in \(\mathbb {D}_{\rho _{0}}\).

Next, we prove Theorem 1.3(2). For any \(\zeta =\rho _{0}e^{i\theta }\in \partial \mathbb {D}_{\rho _{0}}\), we obtain that

$$\begin{aligned} |f(\zeta )-f(0)|= & {} \left| \int _{[0,\zeta ]}f_{z}(z)dz+f_{\overline{z}}(z)d\overline{z}\right| \nonumber \\\ge & {} \left| \int _{[0,\zeta ]}f_{z}(0)dz+f_{\overline{z}}(0)d\overline{z}\right| \nonumber \\&-\left| \int _{[0,\zeta ]}[f_{z}(z)-f_{z}(0)]dz+[f_{\overline{z}}(z)-f_{\overline{z}}(0)]d\overline{z}\right| \nonumber \\\ge & {} l(D{f}(0))\rho _{0}-(M+5)\int _{0}^{\rho _0}\frac{|z|(2-|z|)}{(1-|z|)^{2}}|dz| \quad (\text {by }4.3) \nonumber \\= & {} \frac{\beta \mathrm {c}_{\alpha }\rho _{0}}{M(\alpha +2)}-(M+5)\,\frac{\rho _{0}^2}{1-\rho _{0}}. \nonumber \\= & {} (M+5)\,\Big (\frac{\rho _{0}}{1-\rho _{0}}\Big )^2. \quad (\text {by }1.11) \end{aligned}$$

Hence \(f(\mathbb {D}_{\rho _{0}})\) contains a univalent disk \(\mathbb {D}_{R_{0}}\), where

$$\begin{aligned} R_{0}\ge (M+5)\,\Big (\frac{\rho _{0}}{1-\rho _{0}}\Big )^2. \end{aligned}$$

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