1 Introduction and Preliminaries

The classical inequality of Bohr says that if f is an analytic function in the unit disk \(\mathbb {D}:=\{z\in \mathbb {C}:|z|<1\}\) with the following Taylor series expansion

$$\begin{aligned} f(z)=\sum _{n=0}^{\infty } a_nz^n \end{aligned}$$
(1.1)

such that \(|f(z)|<1\) in \(\mathbb {D}\). Then

$$\begin{aligned} \sum _{n=0}^{\infty }|a_n|r^n\le 1\;\text {for}\;|z|=r\le \frac{1}{3}. \end{aligned}$$
(1.2)

Here 1/3 is known as Bohr radius and it can’t be improved, while the inequality (1.2) is known as Bohr inequality. In 1914, H. Bohr [10] obtained the inequality (1.2) for \(r\le 1/6\) but later Weiner, Riesz and Schur [12] independently improved it to 1/3. An observation shows that the quantity \(1-|a_0|\) is equal to \(d\left( f(0), \partial \mathbb {D}\right) \), where d is the Euclidean distance and \(\partial \mathbb {D}\) is the boundary of \(\mathbb {D}\). Therefore, the inequality (1.2) also can be written as

$$\begin{aligned} \sum _{n=1}^{\infty }|a_nz^n|\le 1-|a_0|=d\left( f(0), \partial \mathbb {D}\right) \;\text {for}\;|z|=r\le \frac{1}{3}.\end{aligned}$$
(1.3)

It is important to note that the constant 1/3 is independent of the coefficients of the Taylor series (1.1). This fact can be elucidated by saying that Bohr inequality occurs in the class \(\mathcal {B}\) of analytic self maps of the unit disk \(\mathbb {D}\). Analytic functions \(f\in \mathcal {B}\) of the form (1.1) satisfying the inequality (1.2) for \(|z|=r\le 1/3\), are said to satisfy the classical Bohr phenomenon. The concept of Bohr phenomenon can be generalized to the class \(\mathcal {A}\) consisting of analytic functions of the form (1.1) which map from \(\mathbb {D}\) into a given domain \(\Theta \subseteq \mathbb {C}\) such that \(f(\mathbb {D})\subseteq \Theta \). The class \(\mathcal {A}\) is said to satisfy the Bohr phenomenon if \(\exists \) largest radius \(r_{\Theta }\in (0, 1)\) such that (1.3) holds for \(|z|=r \le r_{\Theta }\). Here \(r_{\Theta }\) is known as Bohr radius for the class \(\mathcal {A}\). The Bohr radius has been obtained for the class \(\mathcal {A}\) when \(\Theta \) is convex domain [4], simply connected domain [1], the exterior of the closed unit disk, the punctured unit disk, and concave wedge domain (see [5]). In 1997, Boas and Khavinson [9] generalized the Bohr inequality in several complex variables by finding multidimensional Bohr radius. In 2021, Liu and Ponnusamy [22] obtained multidimensional analogues of refined Bohr inequality.

There are many improved versions of Bohr’s inequality (1.2) in various forms obtained by several authors. In 2020, Kayumov and Ponnusamy [20] obtained several interesting improved versions of Bohr inequality. For more results on this, we refer the reader to glance through the articles (see [16,17,18,19,20,21, 23, 24]). In 2017, Kayumov and Ponnusamy [18] introduced Bohr–Rogosinski radius motivated by Rogosinski radius for bounded analytic functions in \(\mathbb {D}\). Rogosinski radius is defined as follows: Let \(f(z)=\sum _{n=0}^{\infty }a_nz^n\) be analytic in \(\mathbb {D}\) and its corresponding partial sum of f is defined by \(S_N(z):=\sum _{n=0}^{N-1}a_nz^n\). Then, for every \(N\ge 1\), we have \(|\sum _{n=0}^{N-1}a_nz^n|<1\) in the disk \(|z|<1/2\) and the radius 1/2 is sharp. Motivated by Rogosinski radius, Kayumov and Ponnusamy have considered the Bohr–Rogosinski sum \(R_N^f(z)\) which is defined by

$$\begin{aligned} R_N^f(z):=|f(z)|+\sum _{n=N}^{\infty }|a_n||z|^n. \end{aligned}$$
(1.4)

It is worth to point out that \(|S_N(z)|=\left| f(z)-\sum _{n=N}^{\infty }a_nz^n\right| \le R_N^f(z)\). Therefore, it is easy to see that the validity of Bohr-type radius for \(R_N^f(z)\), which is related to the classical Bohr sum (Majorant series) in which f(0) is replaced by f(z), gives Rogosinski radius in the case of bounded analytic functions in \(\mathbb {D}\). We have the following interesting results by Kayumov and Ponnusamy [18].

Theorem A

[18] Let \(f(z)=\sum _{n=0}^{\infty } a_nz^n\) be analytic in \(\mathbb {D}\) and \(|f(z)|\le 1\). Then

$$\begin{aligned} |f(z)|+\sum _{n=N}^{\infty }|a_n||z|^n\le 1\end{aligned}$$
(1.5)

for \(|z|=r\le R_N\), where \(R_N\) is the positive root of the equation \(\psi _N(r)=0\), \(\psi _n(r)=2(1+r)r^N-(1-r)^2\). The radius \(R_N\) is the best possible. Moreover,

$$\begin{aligned} |f(z)|^2+\sum _{n=N}^{\infty }|a_n||z|^n\le 1\end{aligned}$$
(1.6)

for \(R_N'\), where \(R_N'\) is the positive root of the equation \((1+r)r^N-(1-r)^2=0\). The radius \(R_N'\) is the best possible.

In 2020, Kayumov and Ponnusamy [20] have proved the following several improved versions of Bohr’s inequality for analytic functions.

Theorem B

[20] Let \(f(z)=\sum _{n=0}^{\infty } a_nz^n\) be analytic in \(\mathbb {D}\), \(|f(z)|\le 1\) and \(S_r\) denotes the area of the image of the subdisk \(|z|<r\) under mapping f. Then

$$\begin{aligned} B_1(r):=\sum _{n=0}^{\infty } |a_n|r^n+\frac{16}{9}\left( \frac{S_r}{\pi }\right) \le 1\;\text {for}\;r\le \frac{1}{3} \end{aligned}$$
(1.7)

and the numbers 1/3, 16/9 cannot be improved. Moreover,

$$\begin{aligned} B_2(r):=|a_0|^2+\sum _{n=1}^{\infty } a_nr^n+\frac{9}{8}\left( \frac{S_r}{\pi }\right) \le 1\;\text {for}\;r\le \frac{1}{2} \end{aligned}$$
(1.8)

and the numbers 1/2, 9/8 cannot be improved.

In 2020, Ponnusamy et al. [23] established the following refined Bohr inequality by applying a refined version of the coefficient inequalities.

Theorem C

[23] Let \(f(z)=\sum _{n=0}^{\infty } a_nz^n\) be analytic in \(\mathbb {D}\) and \(|f(z)|\le 1\). Then

$$\begin{aligned} \sum _{n=0}^\infty |a_n|r^n+\left( \frac{1}{1+|a_0|}+\frac{r}{1-r}\right) \sum _{n=1}^\infty |a_n|^2r^{2n}\le 1\end{aligned}$$
(1.9)

for \(r\le 1/(2+|a_0|)\) and the numbers \(1/(1+|a_0|)\) and \(1/(2+|a_0|)\) cannot be improved. Moreover,

$$\begin{aligned} |a_0|^2+\sum _{n=1}^\infty |a_n|r^n+\left( \frac{1}{1+|a_0|}+\frac{r}{1-r}\right) \sum _{n=1}^\infty |a_n|^2r^{2n}\le 1\end{aligned}$$

for \(r\le 1/2\) and the numbers \(1/(1+|a_0|)\) and 1/2 cannot be improved.

Bohr’s phenomenon for the complex-valued harmonic mappings have been studied extensively by many authors (see [1, 2, 6, 7]). Improved Bohr inequality for locally univalent harmonic mappings have been discussed by Evdoridis et al. [13].

A complex-valued function \(f=u+iv\) is harmonic if u and v are real-harmonic in \(\mathbb {D}\). Every harmonic function f has the canonical representation \(f=h+\overline{g}\), where h and g are analytic in \(\mathbb {D}\) known respectively as the analytic and co-analytic parts of f. A locally univalent harmonic function f is said to be sense-preserving if the Jacobian of f, defined by \(J_f(z):=|h'(z)|^2-|g'(z)|^2\), is positive in \(\mathbb {D}\) and sense-reversing if \(J_f(z)\) is negative in \(\mathbb {D}\). Let \(\mathcal {H}\) be the class of all complex- valued harmonic functions \(f=h+\overline{g}\) defined in \(\mathbb {D}\), where h and g are analytic in \(\mathbb {D}\) such that \(h(0)=h'(0)-1=0\) and \(g(0)=0\). If the co-analytic part \(g(z) \equiv 0\) in \(\mathbb {D}\), then the class \(\mathcal {H}\) reduces to the class \(\mathcal {A}\) of analytic functions in \(\mathbb {D}\) with \(f(0)=0\) and \(f'(0)=1\). A function \(f\in \mathcal {H}\) is said to be in \(\mathcal {H}_0\) if \(g'(0)=0\). Thus, every \(f=h+\overline{g}\in \mathcal {H}_0\) has the following form

$$\begin{aligned} f(z)=h(z)+\overline{g}(z)=z+\sum _{n=2}^\infty a_nz^n+\overline{\sum _{n=2}^\infty b_nz^n}. \end{aligned}$$
(1.10)

A domain \(\Omega \) is called starlike with respect to a point \(z_0\in \Omega \) if the line segment joining \(z_0\) to any point in \(\Omega \) lies in \(\Omega \). In particular, if \(z_0=0\), then \(\Omega \) is simply called starlike. A complex-valued harmonic mapping \(f\in \mathcal {H}\) is said to be starlike if \(f(\mathbb {D})\) is starlike. We denote the class of harmonic starlike functions in \(\mathbb {D}\) by \(\mathcal {S}_{\mathcal {H}}^*\). A domain \(\Omega \) is called convex if it is starlike with respect to every point in \(\Omega \). A function \(f\in \mathcal {H}\) is said to be convex if \(f(\mathbb {D})\) is convex. We denote \(\mathcal {K}_{\mathcal {H}}\) by the class of harmonic convex mappings in \(\mathbb {D}\).

Definition A

The polylogarithm \(Li_s(z)\), also known as Jonquière’s function, is a special function of order s and argument z

$$\begin{aligned} Li_s(z)=\sum _{n=1}^\infty \frac{z^n}{n^s}=z+\frac{z^2}{2^s}+\frac{z^3}{3^s}+\cdots \end{aligned}$$

defined in the complex plane over the unit disk. The special case \(s=1\) involves the ordinary natural logarithm, \(Li_1(z)=-\ln (1-z)\), while the special cases \(s=2\) and \(s=3\) are called the dilogarithm (also known as Spence’s function) and trilogarithm respectively.

In 2020, Ghosh and Allu [14] considered the following class for \(M>0\),

$$\begin{aligned} \mathcal {P}_{\mathcal {H}}^0(M)=\{f=h+\overline{g}\in \mathcal {H}: \text {Re}\left( zh''(z)\right) >-M+|zg''(z)|\;\text {with}\; g'(0)=0\;\text {for}\; z\in \mathbb {D}\}. \end{aligned}$$

The organization of this paper is: In Sect. 3, we discuss some geometric properties for functions in the class \(\mathcal {P}_{\mathcal {H}}^0(M)\). In Sect. 4, we obtain sharp Bohr–Rogosinski radius for the class \(\mathcal {P}_{\mathcal {H}}^0(M)\). In Sect. 5, we establish interesting sharp improved Bohr radius \(\mathcal {P}_{\mathcal {H}}^0(M)\). In Sect. 6, we prove sharp refined Bohr radius for the class \(\mathcal {P}_{\mathcal {H}}^0(M)\). The rest sections are for lemmas and proofs of the main results.

2 Some Lemmas

We have the following lemmas related to coefficient bounds and growth estimates for the class \(\mathcal {P}_{\mathcal {H}}^0(M)\).

Lemma 2.1

[14] The harmonic map \(f=h+\overline{g}\) belongs to \(\mathcal {P}_{\mathcal {H}}^0(M)\) if and only if the function \(F_\epsilon =h +\epsilon g\) belongs to \(\mathcal {P}(M\)) for \(|\epsilon |=1\), where \(\mathcal {P}(M)\) is defined by

$$\begin{aligned} \mathcal {P}(M):= \left\{ \phi \in \mathcal {A}: \text {Re}\left( z\phi ''(z)\right)>-M\;\;\text {for}\; M > 0\;\text {and}\;z\in \mathbb {D}\right\} . \end{aligned}$$
(2.1)

Lemma 2.2

[14] Let \(f \in \mathcal {P}_{\mathcal {H}}^0(M)\) for \(M>0\) be given by (1.10). Then for each \(n\ge 2\), we have \(|b_n|\le \frac{2M}{n(n-1)}\). The result is sharp with \(f(z)=z-\frac{M}{n(n-1)}\overline{z^n}\) being extremal.

Remark 2.1

We have found some typographical error in Lemma 2.2 of [14] and the correct statement is given below:

If \(f \in \mathcal {P}_{\mathcal {H}}^0(M)\) for \(M>0\) is given by (1.10). Then for each \(n\ge 2\), we have \(|b_n|\le \frac{M}{n(n-1)}\). The result is sharp with \(f(z)=z-\frac{M}{n(n-1)}\overline{z^n}\) being extremal.

Lemma 2.3

[14] Let \(f \in \mathcal {P}_{\mathcal {H}}^0(M)\) for \(M>0\) be given by (1.10). Then for any \(n\ge 2\), we have (i) \(|a_n|+|b_n|\le \frac{2M}{n(n-1)}\); (ii) \(\left| |a_n|-|b_n|\right| \le \frac{2M}{n(n-1)}\); (iii) \(|a_n|\le \frac{2M}{n(n-1)}\).

The result is sharp for the function f given by \(f'(z)=1-2M\ln (1-z)\).

Lemma 2.4

[7, 14] Let \(f=h+\overline{g}\in \mathcal {P}_{\mathcal {H}}^0(M)\) for \(M>0\) be given by (1.10). Then

$$\begin{aligned} |z|+2M\sum _{n=2}^\infty \frac{(-1)^{n-1}}{n(n-1)}|z|^n\le |f(z)|\le |z|+2M\sum _{n=2}^\infty \frac{1}{n(n-1)}|z|^n.\end{aligned}$$

Both the inequalities are sharp for the function \(f_M=z+2M\sum _{n=2}^\infty \frac{1}{n(n-1)}z^n\).

To prove our convolution results, we need the following definitions and lemmas.

Definition 2.1

[11, 25] Let \(\psi _1\) and \(\psi _2\) be two analytic functions in \(\mathbb {D}\) given by \(\psi _1(z)=\sum _{n=0}^{\infty }a_nz^n\) and \(\psi _2(z)=\sum _{n=0}^{\infty }b_nz^n\). The convolution (or, Hadamard product) is defined by

$$\begin{aligned} \left( \psi _1*\psi _2\right) (z)=\sum _{n=0}^{\infty }a_nb_nz^n\;\;\text {for}\; z\in \mathbb {D}.\end{aligned}$$

Definition 2.2

[15] For harmonic functions \(f_1=h_1+\overline{g_1}\) and \(f_2=h_2+\overline{g_2}\) in \(\mathcal {H}\), the convolution is defined as \(f_1*f_2=h_1*h_2+\overline{g_1*g_2}\).

Definition 2.3

[25] A sequence \(\{a_n\}\) of non-negative numbers is said to be a convex null sequence if \(a_n\rightarrow 0\) as \(n\rightarrow \infty \) and \(a_0-a_1\ge a_1-a_2\ge \cdots \ge a_{n-1}-a_n\ge \cdots \ge 0\).

Lemma 2.5

[25] Let \(\{a_n\}\) be a convex null sequence. Then the function p given by \(p(z)=\frac{a_0}{2}+\sum \limits _{n=1}^{\infty } a_nz^n\) is analytic in \(\mathbb {D}\) and \(\text {Re}(p(z))>0, z\in \mathbb {D}\).

Lemma 2.6

[25] Let the function p be analytic in \(\mathbb {D}\) with \(p(0)=1\) and \(\text {Re}\;p(z)>\frac{1}{2}\) in \(\mathbb {D}\). Then for any analytic function f in \(\mathbb {D}\), the function \(p*f\) takes values in the convex hull of the image of \(\mathbb {D}\) under f.

Lemma 2.7

Let \(\mathcal {P}(M)\) be the subclass of \(\mathcal {A}\) defined in (2.1). If \(F\in \mathcal {P}(M)\) with \(0<M\le \frac{3}{5}\). Then \(\text {Re}\left( \frac{F(z)}{z}\right) >\frac{1}{2}\).

Proof

Let \(F\in \mathcal {P}(M)\) be given by \(F(z)=z+\sum \limits _{n=2}^\infty A_nz^n\). Then, we have

$$\begin{aligned} \text {Re}\left( zF''(z)\right)>-M\Rightarrow & {} \text {Re}\left( M+\sum _{n=2}^\infty n(n-1)A_nz^{n-1}\right)>0\\\Rightarrow & {} \text {Re}\left( 1+\frac{1}{2M}\sum _{n=2}^\infty n(n-1)A_nz^{n-1}\right) >\frac{1}{2}\;\;\text {for}\;\;z\in \mathbb {D}. \end{aligned}$$

Let \(p(z)=1+\frac{1}{2M}\sum \nolimits _{n=2}^\infty n(n-1)A_nz^{n-1}\). Then \(p(0)=1\) and \(\text {Re}(p(z))>\frac{1}{2}\) in \(\mathbb {D}\). Now, we consider a sequence \(\{c_n\}\) defined by \(c_0=1\) and \(c_{n-1}=\frac{2M}{n(n-1)}\) for \(n\ge 2\). It is clear that \(c_n\rightarrow 0\) as \(n\rightarrow \infty \). Note that \(c_0-c_1=1-M\) and \(c_1-c_2=2\,M/3\). So \(c_0-c_1\ge c_1-c_2\ge \cdots \ge c_{n-1}-c_n\ge \cdots \ge 0\) is possible only when \(0<M\le 3/5\). Thus \(\{c_n\}\) is a convex null sequence. In view of Lemma 2.5, the function

$$\begin{aligned} q(z)=\frac{1}{2}+\sum _{n=2}^\infty \frac{2M}{n(n-1)}z^{n-1} \end{aligned}$$

is analytic in \(\mathbb {D}\) with \(\text {Re} (q(z))>0\). Now,

$$\begin{aligned} \frac{F(z)}{z}=1+\sum \limits _{n=2}^\infty A_nz^{n-1}= & {} p(z)*\left( 1+\sum _{n=2}^\infty \frac{2M}{n(n-1)}z^{n-1}\right) \nonumber \\= & {} p(z)*\left( q(z)+\frac{1}{2}\right) . \end{aligned}$$
(2.2)

In view of Lemma 2.6 and (2.2), we have \(\text {Re}\left( \frac{F(z)}{z}\right) >\frac{1}{2}\) for \(z\in \mathbb {D}\). This completes the proof. \(\square \)

Lemma 2.8

Let \(F_1,F_2\in \mathcal {P}(M)\) with \(0<M\le \frac{3}{5}\), where \(\mathcal {P}(M)\) is defined in (2.1). Then \(F_1*F_2\in \mathcal {P}(M)\).

Proof

Let \(F_1(z)=z+\sum \nolimits _{n=2}^\infty A_nz^n\) and \(F_2(z)=z+\sum \nolimits _{n=2}^\infty B_nz^n\). Then the convolution of \(F_1\) and \(F_2\) is given by

$$\begin{aligned} F(z)=F_1(z)*F_2(z)=z+\sum \limits _{n=2}^\infty A_nB_nz^n. \end{aligned}$$

Now,

$$\begin{aligned} zF''(z)=\sum \limits _{n=2}^\infty n(n-1)A_nB_nz^{n-1}=\left( \frac{F_2(z)}{z}\right) *\left( zF_1''(z)\right) .\end{aligned}$$
(2.3)

Since \(F_1, F_2\in \mathcal {P}(M)\), so \(\text {Re}(zF_1''(z))>-M\) and in view of Lemma 2.7, we have \(\text {Re}\left( \frac{F_2(z)}{z}\right) >\frac{1}{2}\). In view of Lemma 2.6 and (2.3), we have \(\text {Re}(zF''(z))>-M\) in \(\mathbb {D}\). Therefore \(F=F_1*F_2\in \mathcal {P}(M)\). This completes the proof. \(\square \)

Lemma 2.9

[8] Let \(f=h+\overline{g}\) be given by (1.10).

  1. (i)

    If \(\sum \nolimits _{n=2}^{\infty } n\left( |a_n|+|b_n|\right) \le 1\), then f is starlike in \(\mathbb {D}\);

  2. (ii)

    If \(\sum \nolimits _{n=2}^{\infty } n^2\left( |a_n|+|b_n|\right) \le 1\), then f is convex in \(\mathbb {D}\).

3 Convex Combinations and Convolutions

In this section, we will show that \(\mathcal {P}_{\mathcal {H}}^0(M)\) is closed under convex combinations and convolutions.

Theorem 3.1

The class \(\mathcal {P}_{\mathcal {H}}^0(M)\) is closed under convex combinations.

Proof

Let \(f_i=h_i+\overline{g_i}\in \mathcal {P}_{\mathcal {H}}^0(M)\) for \(1\le i\le n\) and \(\sum \nolimits _{i=1}^n t_i=1\), where \(0\le t_i\le 1\) for each i. Then, we have

$$\begin{aligned} \text {Re}\left( zh_i''(z)\right) >-M+|zg_i''(z)| \;\text {with}\; h_i(0)=g_i(0)=h_i'(0)-1=g_i'(0)=0\;\text {for}\;1\le i\le n. \end{aligned}$$

The convex combination of the \(f_i\)’s can be written as

$$\begin{aligned} f(z)=\sum _{i=1}^n t_if_i(z)=h(z)+\overline{g(z)},\end{aligned}$$

where \(h(z)=\sum \nolimits _{i=1}^n t_i h_i(z)\) and \(g(z)=\sum \nolimits _{i=1}^n t_ig_i(z)\). Then both h and g are analytic in \(\mathbb {D}\) with \(h(0)=g(0)=h'(0)-1=g'(0)=0\). Now,

$$\begin{aligned} \text {Re}\left( zh''(z)\right)= & {} \text {Re}\left( z\sum _{i=1}^nt_ih_i''\right) =\sum _{i=1}^n t_i\text {Re}\left( zh_i''\right) \\> & {} \sum _{i=1}^nt_i \left( -M+|zg_i''(z)|\right) =-M+\sum _{i=1}^nt_i\left| zg_i''(z)\right| \\\ge & {} -M+\left| z\left( \sum _{i=1}^nt_i g_i''(z)\right) \right| = -M+|zg''(z)|. \end{aligned}$$

This shows that \(f\in \mathcal {P}_{\mathcal {H}}^0(M)\). This completes the proof.\(\square \)

Theorem 3.2

Let \(F_1,F_2\in \mathcal {P}_{\mathcal {H}}^0(M)\) with \(0<M\le \frac{3}{5}\). Then \(F_1*F_2\in \mathcal {P}_{\mathcal {H}}^0(M)\).

Proof

Let \(F_1=h_1+\overline{g_1}\) and \(F_2=h_2+\overline{g_2}\) be two functions in \(\mathcal {P}_{\mathcal {H}}^0(M)\). Then the convolution of \(F_1\) and \(F_2\) is given by \(F_1*F_2=h_1*h_2+\overline{g_1*g_2}\). To show that \(F_1*F_2\in \mathcal {P}_{\mathcal {H}}^0(M)\), it is sufficient to show that \(F=h_1*h_2+\epsilon \left( g_1*g_2\right) \in \mathcal {P}(M)\) for each \(\epsilon \) with \(|\epsilon |=1\). By Lemma 2.1, we have \(h_1+\epsilon g_1, h_2+\epsilon g_2\in \mathcal {P}(M)\) for each \(\epsilon \) with \(|\epsilon |=1\). Thus, we deduce that

$$\begin{aligned} F=h_1*h_2+\epsilon \left( g_1*g_2\right) =\frac{1}{2}\left( (h_1-g_1)*(h_2-\epsilon g_2)\right) +\frac{1}{2}\left( (h_1+g_1)*(h_2+\epsilon g_2)\right) . \end{aligned}$$

In view of Lemma 2.8, we have \((h_1-g_1)*(h_2-\epsilon g_2), (h_1+g_1)*(h_2+\epsilon g_2)\in \mathcal {P}(M)\). Then in view of Theorem 3.1, we get \(F\in \mathcal {P}(M)\). Hence \(\mathcal {P}_{\mathcal {H}}^0(M)\) is closed under convolution. This completes the proof. \(\square \)

In 2002, Goodloe [15] considered the Hadamard product of a harmonic function with an analytic function as follows.

$$\begin{aligned} f\tilde{*}\phi =h*\phi +\overline{g*\phi },\end{aligned}$$

where \(f=h+\overline{g}\) is harmonic and \(\phi \) is an analytic function in \(\mathbb {D}\).

Theorem 3.3

Let \(f\in \mathcal {P}_{\mathcal {H}}^0(M)\) and \(\phi \in \mathcal {A}\) be such that \(\text {Re}\left( \frac{\phi (z)}{z}\right) >\frac{1}{2}\) for \(z\in \mathbb {D}\). Then \(f\tilde{*}\phi \in \mathcal {P}_{\mathcal {H}}^0(M)\).

Proof

Let \(f=h+\overline{g}\in \mathcal {P}_{\mathcal {H}}^0(M)\). In view of Lemma 2.1, we have \(f_1=h+\epsilon g\in \mathcal {P}(M)\) for each \(\epsilon \) with \(|\epsilon |=1\). To prove that \(f\tilde{*}\phi =h*\phi +\overline{g*\phi }\in \mathcal {P}_{\mathcal {H}}^0(M)\), it is sufficient to show that \(F(z)=h*\phi +\epsilon (g*\phi )\in \mathcal {P}(M)\) for each \(\epsilon (|\epsilon |=1)\). Since \(f_1\in \mathcal {P}(M)\) and \(\phi \in \mathcal {A}\), so we assume that \(f_1(z)=z+\sum _{n=2}^\infty A_nz^n\) and \(\phi (z)=z+\sum _{n=2}^\infty B_nz^n\). Then, we deduce that

$$\begin{aligned}{} & {} F=f_1*\phi =z+\sum \limits _{n=2}^\infty A_nB_nz^n\nonumber \\ \text {and}{} & {} zF''=\sum \limits _{n=2}^\infty n(n-1)A_nB_nz^{n-1}=\left( \frac{\phi (z)}{z}\right) *(zf_1''(z)). \end{aligned}$$
(3.1)

Since \(\text {Re}\left( \frac{\phi (z)}{z}\right) >\frac{1}{2}\) and \(f_1\in \mathcal {P}(M)\), \(\text {Re}(zf_1''(z))>-M\), so in view of Lemma 2.6 and (3.1), we have \(\text {Re}(zF''(z))>-M\) in \(\mathbb {D}\). Hence \(F\in \mathcal {P}(M)\). This completes the proof. \(\square \)

Corollary 3.1

Let \(f\in \mathcal {P}_{\mathcal {H}}^0(M)\) and \(\phi \in \mathcal {K}\), where \(\mathcal {K}\) denotes the family of all convex functions in \(\mathbb {D}\). Then \(f\tilde{*}\phi \in \mathcal {P}_{\mathcal {H}}^0(M)\).

Proof

Since \(\phi \in \mathcal {K}\), so \(\text {Re}\left( \frac{\phi (z)}{z}\right) >\frac{1}{2}\) for \(z\in \mathbb {D}\). The result immediately follows from Theorem 3.3.\(\square \)

Theorem 3.4

Let \(f\in \mathcal {P}_{\mathcal {H}}^0(M)\) and \(\phi \in \mathcal {A}\) be such that \(\text {Re}\left( \frac{\phi (z)}{z}\right) >\frac{1}{2}\) for \(z\in \mathbb {D}\). Then \(f*\left( \phi +\beta \overline{\phi }\right) \in \mathcal {P}_{\mathcal {H}}^0(M)\), where \(|\beta |=1\).

Proof

Let \(f=h+\overline{g}\in \mathcal {P}_{\mathcal {H}}^0(M)\). Then \(h(z)=z+\sum _{n=2}^\infty a_nz^n\), \(g(z)=\sum _{n=2}^\infty b_nz^n\). Now,

$$\begin{aligned} f*\left( \phi +\beta \overline{\phi }\right) =\left( h+\overline{g}\right) *\left( \phi +\overline{\overline{\beta }\phi }\right) =h*\phi +\overline{\overline{\beta }(g*\phi )}.\end{aligned}$$

To prove that \(f*\left( \phi +\beta \overline{\phi }\right) \in \mathcal {P}_{\mathcal {H}}^0(M)\), it is sufficient to show that \(f_{\epsilon }=h*\phi +\epsilon \overline{\beta }(g*\phi )\in \mathcal {P}(M)\) for each \(\epsilon (|\epsilon |=1)\). Let \(\phi (z)=z+\sum \nolimits _{n=2}^\infty C_nz^n\). For each \(|\epsilon |=1\), we have

$$\begin{aligned} zf_{\epsilon }''(z)=\left( \frac{\phi (z)}{z}\right) *\left( z\left( h(z)+\epsilon \overline{\beta }g(z)\right) ''\right) .\end{aligned}$$
(3.2)

Since \(f=h+\overline{g}\in \mathcal {P}_{\mathcal {H}}^0(M)\), so in view of Lemma 2.1, we have \(h+\epsilon \overline{\beta } g\in \mathcal {P}(M)\) for \(\epsilon \), \(\beta \) with \(|\epsilon \overline{\beta }|=1\), i.e., \(|\beta |=1\). Thus, we have

$$\begin{aligned} \text {Re}\left( z\left( h(z)+\epsilon \overline{\beta }g(z)\right) ''\right) >-M\;\text {for}\;z\in \mathbb {D}. \end{aligned}$$

Since \(\text {Re}\left( \frac{\phi (z)}{z}\right) >\frac{1}{2}\) for \(z\in \mathbb {D}\), so in view of Lemma 2.6 and (3.2), we have

$$\begin{aligned} \text {Re}\left( zf_{\epsilon }''(z)\right) >-M\;\text {for}\;z\in \mathbb {D}.\end{aligned}$$

Hence \(f_{\epsilon }\in \mathcal {P}(M)\). This completes the proof. \(\square \)

Corollary 3.2

Let \(f\in \mathcal {P}_{\mathcal {H}}^0(M)\) and \(\phi \in \mathcal {K}\), where \(\mathcal {K}\) denotes the family of all convex functions in \(\mathbb {D}\). Then \(f*\left( \phi +\beta \overline{\phi }\right) \in \mathcal {P}_{\mathcal {H}}^0(M)\), where \(|\beta |=1\).

Proof

Since \(\phi \in \mathcal {K}\), so \(\text {Re}\left( \frac{\phi (z)}{z}\right) >\frac{1}{2}\) for \(z\in \mathbb {D}\). The result immediately follows from Theorem 3.4.\(\square \)

By Lemmas 2.3 and 2.9, it is possible to show that each \(f\in \mathcal {P}_{\mathcal {H}}^0(M)\) is convex (resp. starlike) in some disk D, i.e., f(D) is a convex domain (resp. f(D) is a domain starlike with respect to the origin).

Theorem 3.5

Let \(f=h+\overline{g}\in \mathcal {P}_{\mathcal {H}}^0(M)\) be given by (1.10). Then f is starlike in \(|z|<1-e^{-\frac{1}{2M}}=r^*\) and convex in \(|z|<r_c\), where \(r_c\in (0,1)\) is the smallest root of the equation \(\frac{r}{1-r}-\ln (1-r)-\frac{1}{2M}=0\).

Proof

Let \(0<r<1\) and \(f_r(z)=\frac{1}{r}f(rz)=z+\sum _{n=2}^{\infty } a_nr^{n-1}z^n+\overline{\sum _{n=2}^{\infty } b_nr^{n-1}z^n}\) for \(z\in \mathbb {D}\). For convenience, we let \(S_1=\sum _{n=2}^{\infty } n\left( |a_n|+|b_n|\right) r^{n-1}\) and \(S_2=\sum _{n=2}^{\infty } n^2\left( |a_n|+|b_n|\right) r^{n-1}\). According to Lemma 2.9, it suffices to show that \(S_1\le 1\) for \(|z|=r<r^*\) and \(S_2\le 1\) for \(|z|=r<r_c\). In view of Lemma 2.3, we have

$$\begin{aligned} S_1\le 2M\sum _{n=1}^{\infty }\frac{r^n}{n}=-2M\ln (1-r).\end{aligned}$$

Thus, \(S_1\le 1\) if \(r< 1-e^{-\frac{1}{2M}}=r^*\). Again

$$\begin{aligned} S_2\le 2M\sum _{n=1}^{\infty }\left( r^n+\frac{r^n}{n}\right) =2M\left( \frac{r}{1-r}-\ln (1-r)\right) .\end{aligned}$$

Thus, \(S_2\le 1\) if \(r<r_c\), where \(r_c\in (0,1)\) is the smallest root of the equation \(\frac{r}{1-r}-\ln (1-r)-\frac{1}{2M}=0\). This completes the proof. \(\square \)

4 Bohr–Rogosinski Radius for the Class \(\mathcal {P}_{\mathcal {H}}^0(M)\)

In 2023, Ahamed et al. [3] obtained the following results regarding Bohr–Rogosinski radius for the class \(\mathcal {P}_{\mathcal {H}}^0(M)\).

Theorem D

[3] Let \(f\in \mathcal {P}_{\mathcal {H}}^0(M)\) be given by (1.10). Then

$$\begin{aligned} |f(z)|+\sum _{n=N}^\infty \left( |a_n|+|b_n|\right) |z|^n\le d\left( f(0), \partial f(\mathbb {D})\right) \end{aligned}$$
(4.1)

for \(|z|=r\le r_N(M)\) with \(N\ge 2\), where \(r_N(M)\in (0,1)\) is the smallest root of the equation

$$\begin{aligned} r-1+2M\left( 2r-1+2(1-r)\ln (1-r)-\sum _{n=2}^{N-1}\frac{r^n}{n(n-1)}+\ln 4\right) =0. \end{aligned}$$

The constant \(r_N(M)\) is the best possible.

Theorem E

[3] Let \(f\in \mathcal {P}_{\mathcal {H}}^0(M)\) be given by (1.10). Then

$$\begin{aligned} |f(z)|^2+\sum _{n=N}^\infty \left( |a_n|+|b_n|\right) |z|^n\le d\left( f(0), \partial f(\mathbb {D})\right) \end{aligned}$$
(4.2)

for \(|z|=r\le r_N(M)\) with \(N\ge 2\), where \(r_N(M)\in (0,1)\) is the smallest root of the equation

$$\begin{aligned}{} & {} \left( r+2M\left( r+(1-r)\ln (1-r)\right) \right) ^2\\{} & {} \quad +2M\left( r-1+(1-r)\ln (1-r) -\sum _{n=2}^{N-1}\frac{r^n}{n(n-1)}+\ln 4\right) =1.\end{aligned}$$

The constant \(r_N(M)\) is the best possible.

Theorem F

[3] Let \(f\in \mathcal {P}_{\mathcal {H}}^0(M)\) be given by (1.10). Then

$$\begin{aligned} |f(z^m)|+\sum _{n=N}^\infty |a_n||z|^n\le d\left( f(0), \partial f(\mathbb {D})\right) \end{aligned}$$
(4.3)

for \(|z|=r\le r_{m,N}(M)\) with \(N\ge 2\), where \(r_{m,N}(M)\in (0,1)\) is the smallest root of the equation

$$\begin{aligned}{} & {} r^m-2M\\{} & {} \quad \left( r^m+r-1+\left( 1-r^m\right) \ln \left( 1-r^m\right) +(1-r)\ln (1-r)+\sum _{n=2}^{N-1}\frac{r^n}{n(n-1)}+\ln 4\right) =1.\end{aligned}$$

The constant \(r_{m,N}(M)\) is the best possible.

Note that, \(0<M<\frac{1}{2(\ln 4-1)}\) in Theorems D-F. Now we focus on the following question.

Question 4.1

Can we further reduce the Bohr–Rogosinski radius for the class \(\mathcal {P}_{\mathcal {H}}^0(M)\) in Theorems D and E?

Corresponding to the question above, we first prove the following Bohr–Rogosinski radius for the class \(\mathcal {P}_{\mathcal {H}}^0(M)\).

Theorem 4.1

Let \(f\in \mathcal {P}_{\mathcal {H}}^0(M)\) for \(0<M<\frac{1}{2(\ln 4-1)}\) be given by (1.10). Then

$$\begin{aligned} |f(z)|+\sum _{n=N_1}^\infty |a_n||z|^n+\sum _{n=N_2}^\infty |b_n||z|^n\le d\left( f(0), \partial f(\mathbb {D})\right) \end{aligned}$$
(4.4)

for \(|z|=r\le r_{N_1,N_2}(M)\) with \(N_1\ge N_2\ge 2\), where \(r_{N_1,N_2}(M)\in (0,1)\) is the smallest root of the equation

$$\begin{aligned} r-1+2M\left( 2r-1+2\ln 2+2(1-r)\ln (1-r)-\sum _{n=2}^{N_1-1}\frac{r^n}{n(n-1)}+\sum _{n=N_2}^{N_1-1}\frac{r^n}{2n(n-1)}\right) =0. \end{aligned}$$

The constant \(r_{N_1,N_2}(M)\) is the best possible (Figs. 1 and 2).

Fig. 1
figure 1

The graphs of \(r_{3,2}(M)\) and \(r_{5,3}(M)\) in (0, 1)

Fig. 2
figure 2

The graphs of \(r_ {5, 4}(M)\) and \( r_ {7, 3}(M)\) in (0, 1)

Table 1 This table shows the values of the roots \(r_{N_1,N_2}(M)\) for different values of M in \(\left( 0,\frac{1}{2(\ln 4-1)}\right) \) with \(N_1=3,5,7\) and \(N_2=2,3,4\)

Remark 4.1

Clearly Theorem 4.1 holds for the small Bohr–Rogosinski radius than the radius in Theorem D. It can be checked from the Table 1, e.g., when \(N=3\), then \(r_3(0.4)=0.527\) [3] and \(r_{3,2}(0.4)=0.497629\).

Theorem 4.2

Let \(f\in \mathcal {P}_{\mathcal {H}}^0(M)\) for \(0<M<\frac{1}{2(\ln 4-1)}\) be given by (1.10). Then

$$\begin{aligned} |f(z^m)|^l+\sum _{n=N_1}^\infty |a_n||z^n+\sum _{n=N_2}^\infty |b_n||z|^n\le d\left( f(0), \partial f(\mathbb {D})\right) \end{aligned}$$
(4.5)

for \(|z|=r\le r_{l,m,N_1,N_2}(M)\) with \(N_1\ge N_2\ge 2\) and \(l,m\in \mathbb {N}\), where \(r_{l,m,N_1,N_2}(M)\in (0,1)\) is the smallest root of the equation

$$\begin{aligned}{} & {} \left[ r^m+2M\left\{ r^m+(1-r^m)\ln (1-r^m)\right\} \right] ^l+2M\left( (1-r)\ln (1-r)+(1-r)\sum _{n=1}^{N_1-2}\frac{r^n}{n}\right. \\ {}{} & {} \left. -1+2\ln 2+\frac{r^{N_1-1}}{N_1-1}+\sum _{n=N_2}^{N_1-1}\frac{r^n}{2n(n-1)}\right) -1=0.\end{aligned}$$

The constant \(r_{l,m,N_1,N_2}(M)\) is the best possible (Figs. 3 and 4).

Fig. 3
figure 3

The graphs of \(r_ {2, 1, 3, 2} (M)\) and \(r_ {2, 1, 5, 3} (M)\) in (0, 1)

Fig. 4
figure 4

The graphs of \(r_ {2, 2, 3, 2}(M)\) and \(r_ {3, 2, 5, 3} (M)\) in (0, 1)

Table 2 This table shows the values of the roots \(r_{l,m,N_1,N_2}(M)\) for different values of M in \(\left( 0,\frac{1}{2(\ln 4-1)}\right) \) with \(l=2,3\), \(m=1,2,3\), \(N_1=3,5,7\) and \(N_2=2,3\)

Remark 4.2

Clearly Theorem 4.2 holds for the small Bohr–Rogosinski radius than the radius in Theorem E. It can be checked from the Table 2, e.g., when \(N=3\), then \(r_3(1.29)=0.053\) [3], and \(r_{2,1,3,2}(1.29)=0.0434376\).

Theorem 4.3

Let \(f\in \mathcal {P}_{\mathcal {H}}^0(M)\) for \(0<M<\frac{1}{2(\ln 4-1)}\) be given by (1.10). Then

$$\begin{aligned} |f(z^m)|^l+\sum _{n=N}^\infty |a_nb_n||z|^n\le d\left( f(0), \partial f(\mathbb {D})\right) \end{aligned}$$
(4.6)

for \(|z|=r\le r_{l,m,N}(M)\) with \(N\ge 2\) and \(l,m\in \mathbb {N}\), where \(r_{l,m,N}(M)\in (0,1)\) is the smallest root of the equation

$$\begin{aligned}{} & {} \left[ r^m+2M\left\{ r^m+(1-r^m)\ln (1-r^m)\right\} \right] ^l-1+8M^2(r-1) \left( \ln (1-r)+\sum _{n=1}^{N-2}\frac{r^n}{n}\right) \\{} & {} \quad +4M^2(1+r)\left( Li_2(r)-\sum _{n=1}^{N-2}\frac{r^n}{n^2}\right) -8M^2\frac{r^{N-1}}{N-1}-4M^2\frac{r^{N-1}}{(N-1)^2}-2M(1-2\ln 2)=0,\end{aligned}$$

where \(Li_2(r)\) is the dilogarithm function. The constant \(r_{l,m,N}(M)\) is the best possible (Figs. 5 and 6; Table 3).

Fig. 5
figure 5

The graphs of \(r_ {1, 1, 3} (M) \) and \(r_ {1, 2, 3} (M) \) in (0, 1)

Fig. 6
figure 6

The graphs of \(r_ {1, 2, 4} (M)\) and \(r_ {2, 2, 4} (M)\) in (0, 1)

Table 3 This table shows the values of the roots \(r_{l,m,N}(M)\) for different values of M in \(\left( 0,\frac{1}{2(\ln 4-1)}\right) \) with \(N=3,4,5\), \(m=1,2\) and \(l=1,2\)

Theorem 4.4

Let \(f\in \mathcal {P}_{\mathcal {H}}^0(M)\) for \(0<M<\frac{1}{2(\ln 4-1)}\) be given by (1.10). Then

$$\begin{aligned} r+k_1|h(r)|^p+k_2|g(r)|^q+\sum _{n=N_1}^\infty |a_n|r^n+\sum _{n=N_2}^\infty |b_n|r^n\le d\left( f(0), \partial f(\mathbb {D})\right) \end{aligned}$$
(4.7)

for \(r\le r_{p,k_1,q,k_2,N_1,N_2}(M)\) with \(N_1\ge N_2\ge 2\), \(p,q\in \mathbb {N}\) and \(k_1,k_2\in \mathbb {R}\), where \(r_{p,k_1,q,k_2,N_1,N_2}(M)\in (0,1)\) is the smallest root of the equation

$$\begin{aligned}{} & {} r+k_1 r^p+k_2 r^q-1+2M\left( (1-r)\ln (1-r)-1+2\ln 2+(1-r)\sum _{n=1}^{N_1-2}\frac{r^n}{n}+\frac{r^{N_1-1}}{N_1-1}\right. \\ {}{} & {} \left. +\sum _{n=N_2}^{N_1-1}\frac{r^n}{2n(n-1)}\right) =0. \end{aligned}$$

The radius \(r_{p,k_1,q,k_2,N_1,N_2}(M)\) is the best possible (Figs. 7; Table 4).

Fig. 7
figure 7

The graphs of \(r_ {25, 100, 8, 1000, 3, 2}(M)\) and \(r_ {7, 200, 9, 200, 5, 2}(M)\) in (0, 1)

Table 4 This table shows the values of the roots \(r_{p,k_1,q,k_2,N_1,N_2}(M)\) for different values of M in \(\left( 0,\frac{1}{2(\ln 4-1)}\right) \) with \(p=7,25,70\), \(k_1=100,200,2000\), \(q=8,9,90\), \(k_2=200,1000,2000\), \(N_1=3,5\) and \(N_2=2,3\)

5 Proof of the Theorems 4.14.4

Proof of Theorem 4.1

Let \(f\in \mathcal {P}_{\mathcal {H}}^0(M)\) be as (1.10). Using Lemma 2.4, we get

$$\begin{aligned} |z| +2M\sum _{n=2}^{\infty }\frac{(-1)^n}{n(n-1)}|z|^n\le |f(z)|, \;\;\text {where}\;\;|z|<1.\end{aligned}$$
(5.1)

Since \(f(0)=0\), so the Euclidean distance between f(0) and the boundary of \(f(\mathbb {D})\) is \(d\left( f(0),\partial f(\mathbb {D})\right) :=\liminf \limits _{|z|\rightarrow 1}|f(z)-f(0)|=\liminf \limits _{|z|\rightarrow 1}|f(z)|\). Thus from (5.1), we get

$$\begin{aligned} 1+2M(1-2\ln 2)\le d\left( f(0),\partial f(\mathbb {D})\right) .\end{aligned}$$
(5.2)

In view of Lemmas 2.2, 2.3 and 2.4, we now deduce for \(N_1\ge N_2\ge 2\) that

$$\begin{aligned}{} & {} |f(z)|+\sum _{n=N_1}^{\infty }|a_n|r^n+\sum _{n=N_2}^{\infty }|b_n|r^n\nonumber \\{} & {} \quad =|f(z)|+\sum _{n=N_1}^{\infty }|a_n|r^n+\sum _{n=N_1}^{\infty }|b_n|r^n+\sum _{n=N_2}^{N_1-1}|b_n|r^n\nonumber \\{} & {} \quad \le r+2M\left( \sum _{n=2}^{\infty }\frac{r^n}{n(n-1)}+\sum _{n=N_1}^{\infty }\frac{r^n}{n(n-1)}+\sum _{n=N_2}^{N_1-1}\frac{r^n}{2n(n-1)}\right) \nonumber \\{} & {} \quad = r+2M\left( 2r+2(1-r)\ln (1-r)-\sum _{n=2}^{N_1-1}\frac{r^n}{n(n-1)}+\sum _{n=N_2}^{N_1-1}\frac{r^n}{2n(n-1)}\right) .\nonumber \\ \end{aligned}$$
(5.3)

Now, we deduce that

$$\begin{aligned}{} & {} r+2M\left( 2r+2(1-r)\ln (1-r)-\sum _{n=2}^{N_1-1}\frac{r^n}{n(n-1)}+\sum _{n=N_2}^{N_1-1}\frac{r^n}{2n(n-1)}\right) \nonumber \\{} & {} \quad \le 1+2M(1-2\ln 2) \end{aligned}$$
(5.4)

for \(0<r\le r_{N_1,N_2}(M)<1\), where \(r_{N_1,N_2}(M)\) is the smallest root of the equation

$$\begin{aligned} r-1+2M\left( 2r-1+2\ln 2+2(1-r)\ln (1-r)-\sum _{n=2}^{N_1-1}\frac{r^n}{n(n-1)}+\sum _{n=N_2}^{N_1-1}\frac{r^n}{2n(n-1)}\right) =0.\end{aligned}$$

To ensure about the existence of a root \(r_{N_1,N_2}(M)\), we construct the function \(\mathcal {G}_1:[0,1)\rightarrow \mathbb {R}\) such that

$$\begin{aligned}{} & {} \mathcal {G}_1(r):=r-1+2M\\{} & {} \quad \left( 2r-1+2\ln 2+2(1-r)\ln (1-r)-\sum _{n=2}^{N_1-1}\frac{r^n}{n(n-1)}+\sum _{n=N_2}^{N_1-1}\frac{r^n}{2n(n-1)}\right) .\end{aligned}$$

It is clear that (i) \(\mathcal {G}_1\) is continuous on [0, 1), (ii) \(\mathcal {G}_1(0)=-1+2M(2\ln 2-1)<0\) and (iii) \(\lim \limits _{r\rightarrow 1}\mathcal {G}_1(r)>0\), since \(\lim \limits _{r\rightarrow 1}(1-r)\ln (1-r)=0\). Thus the claim follows from Intermediate value theorem. Thus, we have

$$\begin{aligned}{} & {} r_{N_1,N_2}(M)-1+2M\left( 2r_{N_1,N_2}(M)-1+2\ln 2+2(1-r_{N_1,N_2}(M))\ln (1-r_{N_1,N_2}(M))\right. \nonumber \\{} & {} \left. -\sum _{n=2}^{N_1-1}\frac{r_{N_1,N_2}^n(M)}{n(n-1)}+\sum _{n=N_2}^{N_1-1}\frac{r_{N_1,N_2}^n(M)}{2n(n-1)}\right) =0. \end{aligned}$$
(5.5)

Combining (5.2), (5.3) and (5.4) for \(|z|=r\le r_{N_1,N_2}(M)\), we deduce that

$$\begin{aligned} |f(z)|+\sum _{n=N_1}^{\infty }|a_n|r^n+\sum _{n=N_2}^{\infty }|b_n|r^n\le d\left( f(0),\partial f(\mathbb {D})\right) .\end{aligned}$$

Now we show that the radius \(r_{N_1,N_2}(M)\) is the best possible. We set

$$\begin{aligned} f:=f_{M}(z)=z+2M\sum _{n=2}^{\infty }\frac{z^n}{n(n-1)}.\end{aligned}$$
(5.6)

Note that \(f_{M}(0)=0\), \(f_{M}\in \mathcal {P}_{\mathcal {H}}^0(M)\). For \(z=r\), we have

$$\begin{aligned}{} & {} |f_{M}(r)-f_{M}(0)|=|f_{M}(r)|=\left| r+2M\sum _{n=2}^{\infty }\frac{(r)^n}{n(n-1)}\right| =r+2M\sum _{n=2}^{\infty }\frac{r^n}{n(n-1)}\nonumber \\\text {and}{} & {} \liminf _{r\rightarrow 1^{-}}|f_{M}(r)|=1+2M. \end{aligned}$$
(5.7)

and for \(z=-r\), we have

$$\begin{aligned}{} & {} |f_{M}(-r)-f_{M}(0)|=\left| -r+2M\sum _{n=2}^{\infty }\frac{(-r)^n}{n(n-1)}\right| =r+2M\sum _{n=2}^{\infty }\frac{(-1)^{n-1}r^n}{n(n-1)}\nonumber \\\text {and}{} & {} \liminf _{r\rightarrow 1^{-}}|f_{M}(-r)|=1+2M(1-2\ln 2).\end{aligned}$$
(5.8)

From (5.7) and (5.8), we have

$$\begin{aligned}{} & {} d\left( f_{M}(0),\partial f_{M}(\mathbb {D})\right) =\liminf _{|z|=r\rightarrow 1^{-}}|f_{M}(z)-f_{M}(0)|=1+2M(1-2\ln 2). \end{aligned}$$
(5.9)

For \(|z|=r_{N_1,N_2}(M)\), it follows from (5.5) and (5.9) that

$$\begin{aligned}{} & {} |f(z)|+\sum _{n=N_1}^{\infty }|a_n|r_{N_{1,2}}^n(M)+\sum _{n=N_2}^{\infty }|b_n|r_{N_{1,2}}^n(M)\\{} & {} =r_{N_1,N_2}(M)+2M\left( 2r_{N_1,N_2}(M)+2\left( 1-r_{N_1,N_2}(M)\right) \ln \left( 1-r_{N_1,N_2}(M)\right) \right. \nonumber \\ {}{} & {} \left. -\sum _{n=2}^{N_1-1}\frac{r_{N_1,N_2}^n(M)}{n(n-1)}+\sum _{n=N_2}^{N_1-1}\frac{r_{N_1,N_2}^n(M)}{2n(n-1)}\right) =1+2M(1-2\ln 2)=d\left( f_{M}(0),\partial f_{M}(\mathbb {D})\right) . \end{aligned}$$

Thus the result follows. \(\square \)

Proof of Theorem 4.2

Let \(f\in \mathcal {P}_{\mathcal {H}}^0(M)\) be given by (1.10). Note that \(N_1\ge N_2\ge 2\) and \(l,m\in \mathbb {N}\). In view of Lemmas 2.2, 2.3 and 2.4, we get

$$\begin{aligned}{} & {} |f(z^m)|^l+\sum _{n=N_1}^\infty |a_n||z|^n+\sum _{n=N_2}^\infty |b_n||z|^n\nonumber \\{} & {} \le \left( r^m+2M\sum _{n=2}^\infty \frac{r^{mn}}{n(n-1)}\right) ^l+2M\left( \sum _{n=N_1}^{\infty }\frac{r^n}{n(n-1)}+\sum _{n=N_2}^{N_1-1}\frac{r^n}{2n(n-1)}\right) \nonumber \\{} & {} =\left[ r^m+2M\left\{ r^m+(1-r^m)\ln (1-r^m)\right\} \right] ^l+2M\left( (1-r)\ln (1-r)+(1-r)\sum _{n=1}^{N_1-2}\frac{r^n}{n}\right. \nonumber \\ {}{} & {} \left. +\frac{r^{N_1-1}}{N_1-1}+\sum _{n=N_2}^{N_1-1}\frac{r^n}{2n(n-1)}\right) \end{aligned}$$
(5.10)

Similarly as the proof of Theorem 4.1, we get (5.2) and

$$\begin{aligned}{} & {} \left[ r^m+2M\left\{ r^m+(1-r^m)\ln (1-r^m)\right\} \right] ^l+2M\left( (1-r)\ln (1-r)+(1-r)\sum _{n=1}^{N_1-2}\frac{r^n}{n}\right. \nonumber \\ {}{} & {} \left. +\frac{r^{N_1-1}}{N_1-1}+\sum _{n=N_2}^{N_1-1}\frac{r^n}{2n(n-1)}\right) \le 1+2M(1-2\ln 2)\end{aligned}$$
(5.11)

for \(0<r\le r_{l,m,N_1,N_2}(M)\), where \(r_{l,m,N_1,N_2}(M)\) is the smallest root of \(\mathcal {G}_2(r)=0\) in (0, 1), where \(\mathcal {G}_2:[0,1)\rightarrow \mathbb {R}\) is defined by

$$\begin{aligned}{} & {} \mathcal {G}_2(r)=\left[ r^m+2M\left\{ r^m+(1-r^m)\ln (1-r^m)\right\} \right] ^l\\{} & {} \quad +2M\left( (1-r)\ln (1-r)+(1-r)\sum _{n=1}^{N_1-2}\frac{r^n}{n}\right. \\{} & {} \quad \left. -1+2\ln 2+\frac{r^{N_1-1}}{N_1-1}+\sum _{n=N_2}^{N_1-1}\frac{r^n}{2n(n-1)}\right) -1.\end{aligned}$$

Similarly as the proof of Theorem 4.1, we have

\(\mathcal {G}_2(r_{l,m,N_1,N_2}(M))=0\), i.e.,

$$\begin{aligned}{} & {} \left[ r_{l,m,N_1,N_2}^m(M)\!+\!2M\left\{ r_{l,m,N_1,N_2}^m(M)+(1\!-\!r_{l,m,N_1,N_2}^m(M)) \ln (1\!-\!r_{l,m,\!\!N_1,\!N_2}^m(M))\right\} \right] ^l\nonumber \\{} & {} \quad +2M\left( (1-r_{l,m,N_1,N_2}(M))\ln (1-r_{l,m,N_1,N_2}(M))+(1-r_{l,m,N_1,N_2}(M))\right. \nonumber \\{} & {} \left. \sum _{n=1}^{N_1-2}\frac{r_{l,m,N_1,N_2}^n(M)}{n}\right. \nonumber \\{} & {} \quad \left. -1+2\ln 2+\frac{r_{l,m,N_1,N_2}^{N_1-1}(M)}{N_1-1}+\sum _{n=N_2}^{N_1-1} \frac{r_{l,m,N_1,N_2}^n(M)}{2n(n-1)}\right) -1=0.\end{aligned}$$
(5.12)

Combining (5.2), (5.10) and (5.11), we obtain for \(|z|=r\le r_{l,m,N_1,N_2}(M)\)

$$\begin{aligned} |f(z^m)|^l+\sum _{n=N_1}^\infty |a_n||z^n+\sum _{n=N_2}^\infty |b_n||z|^n\le d\left( f(0),\partial f(\mathbb {D})\right) .\end{aligned}$$

In order to show that \(r_{l,m,N_1,N_2}(M)\) is the best possible, we consider the function \(f=f_{M}\) defined by (5.6) and we again get (5.9). For \(|z|=r_{l,m,N_1,N_2}(M)\), it follows from (5.9) and (5.12) that

$$\begin{aligned}{} & {} |f(z^m)|^l+\sum _{n=N_1}^\infty |a_n||z^n+\sum _{n=N_2}^\infty |b_n||z|^n \\{} & {} \quad =\left[ r_{l,m,N_1,N_2}^m(M)+2M\left\{ r_{l,m,N_1,N_2}^m(M)+(1-r_{l,m,N_1,N_2}^m(M))\ln (1-r_{l,m,N_1,N_2}^m(M))\right\} \right] ^l\nonumber \\{} & {} \quad +2M(1-r_{l,m,N_1,N_2}(M))\ln (1-r_{l,m,N_1,N_2}(M))\nonumber \\ {}{} & {} \quad +2M\left( (1-r_{l,m,N_1,N_2}(M))\sum _{n=1}^{N_1-2}\frac{r_{l,m,N_1,N_2}^n(M)}{n}+\frac{r_{l,m,N_1,N_2}^{N_1-1}(M)}{N_1-1}+\sum _{n=N_2}^{N_1-1}\frac{r_{l,m,N_1,N_2}^n(M)}{2n(n-1)}\right) \\{} & {} \quad =1+2M(1-2\ln 2)=d\left( f_{M}(0),\partial f_{M}(\mathbb {D})\right) .\end{aligned}$$

Therefore, the radius \(r_{l,m,N_1,N_2}(M)\) is the best possible. Thus the result follows. \(\square \)

Proof of Theorem 4.3

Let \(f\in \mathcal {P}_{\mathcal {H}}^0(M)\) be given by (1.10). Note that \(N\ge 2\) and \(l,m\in \mathbb {N}\). In view of Lemmas 2.2, 2.3 and 2.4, we obtain

$$\begin{aligned}{} & {} |f(z^m)|^l+\sum _{n=N}^\infty |a_nb_n||z|^n\nonumber \\{} & {} \quad \le \left( r^m+2M\sum _{n=2}^\infty \frac{r^{mn}}{n(n-1)}\right) ^l+4M^2\sum _{n=N}^\infty \frac{r^{n}}{n^2(n-1)^2}.\end{aligned}$$
(5.13)

Similarly as the proof of Theorem 4.1, we get (5.2) and

$$\begin{aligned}{} & {} \left( r^m+2M\sum _{n=2}^\infty \frac{r^{mn}}{n(n-1)}\right) ^l+4M^2\sum _{n=N}^\infty \frac{r^{n}}{n^2(n-1)^2}\nonumber \\{} & {} \quad \le \left[ r^m\!+\!2M\left\{ r^m\!+\!(1\!-\!r^m)\ln (1\!-\!r^m)\right\} \right] ^l\!+\!8M^2(r\!-\!1) \left( \ln (1\!-\!r)\!+\!\sum _{n\!=\!1}^{N-2}\frac{r^n}{n}\right) \nonumber \\{} & {} \quad \quad +4M^2(1+r)\left( Li_2(r)-\sum _{n=1}^{N-2}\frac{r^n}{n^2}\right) -8M^2\frac{r^{N-1}}{N-1}-4M^2\frac{r^{N-1}}{(N-1)^2}\nonumber \\{} & {} \quad \le 1+2M(1-2\ln 2) \end{aligned}$$
(5.14)

for \(0< r\le r_{l,m,N}(M)\), where \(r_{l,m,N}(M)\) is the smallest root of \(\mathcal {G}_3(r)=0\) in (0, 1) and \(\mathcal {G}_3:[0,1)\rightarrow \mathbb {R}\) is defined by

$$\begin{aligned}{} & {} \mathcal {G}_3(r):= \left[ r^m+2M\left\{ r^m+(1-r^m)\ln (1-r^m)\right\} \right] ^l-1+8M^2(r-1)\\{} & {} \quad \left( \ln (1-r)+\sum _{n=1}^{N-2}\frac{r^n}{n}\right) +4M^2(1+r)\left( Li_2(r)-\sum _{n=1}^{N-2}\frac{r^n}{n^2}\right) -8M^2\frac{r^{N-1}}{N-1}-4M^2\frac{r^{N-1}}{(N-1)^2}-2M(1-2\ln 2).\end{aligned}$$

Similarly as the proof of Theorem 4.1, we have \(\mathcal {G}_3\left( r_{l,m,N}(M)\right) =0\), i.e.,

$$\begin{aligned}{} & {} \left[ r_{l,m,N}^m(M)+2M\left\{ r_{l,m,N}^m(M)+(1-r_{l,m,N}^m(M))\ln (1-r_{l,m,N}^m(M))\right\} \right] ^l\nonumber \\{} & {} \quad -1-2M(1-2\ln 2)\nonumber \\{} & {} \quad +8M^2(r_{l,m,N}(M)-1) \left( \ln (1-r_{l,m,N}(M))+\sum _{n=1}^{N-2}\frac{r_{l,m,N}^n(M)}{n}\right) -8M^2\frac{r_{l,m,N}^{N-1}(M)}{N-1}\nonumber \\{} & {} \quad +4M^2(1+r_{l,m,N}(M))\left( Li_2(r_{l,m,N}(M))-\sum _{n=1}^{N-2}\frac{r_{l,m,N}^n(M)}{n^2}\right) -4M^2\frac{r_{l,m,N}^{N-1}(M)}{(N-1)^2}=0.\nonumber \\ \end{aligned}$$
(5.15)

Combining (5.2), (5.13) and (5.14), we obtain for \(|z|=r\le r_{l,m,N}(M)\)

$$\begin{aligned} |f(z^m)|^l+\sum _{n=N}^\infty |a_nb_n||z|^n\le d\left( f(0),\partial f(\mathbb {D})\right) .\end{aligned}$$

In order to show that \(r_{l,m,N}(M)\) is the best possible, we consider the function \(f=f_ M\) defined by (5.6) and we again get (5.9). For \(|z|=r_{l,m,N}(M)\), it follows from (5.9) and (5.15) that

$$\begin{aligned}{} & {} |f(z^m)|^l+\sum _{n=N}^\infty |a_nb_n|r_{l,m,N}^n(M)\\{} & {} \quad =\left[ r_{l,m,N}^m(M)+2M\left\{ r_{l,m,N}^m(M)+(1-r_{l,m,N}^m(M))\ln (1-r_{l,m,N}^m(M))\right\} \right] ^l\\{} & {} \quad \quad +8M^2(r_{l,m,N}(M)-1) \left( \ln (1-r_{l,m,N}(M))+\sum _{n=1}^{N-2}\frac{r_{l,m,N}^n(M)}{n}\right) -8M^2\frac{r_{l,m,N}^{N-1}(M)}{N-1}\\{} & {} \quad \quad +4M^2(1+r_{l,m,N}(M))\left( Li_2(r_{l,m,N}(M))-\sum _{n=1}^{N-2}\frac{r_{l,m,N}^n(M)}{n^2}\right) -4M^2\frac{r_{l,m,N}^{N-1}(M)}{(N-1)^2}\\{} & {} \quad =1+2M(1-2\ln 2)=d\left( f_M(0),\partial f_M(\mathbb {D})\right) .\end{aligned}$$

Therefore, the radius \(r_{l,m,N}(M)\) is the best possible. Thus the results follows. \(\square \)

Proof of Theorem 4.4

Let \(f\in \mathcal {P}_{\mathcal {H}}^0(M)\) be given by (1.10). We know that, if \(\phi \) is analytic in \(\mathbb {D}\) with \(\phi (0)=0\) and \(|\phi (z)|<1\), \(\forall z\in \mathbb {D}\), then by Schwarz Lemma, we have \(|\phi (z)|\le |z|\). Note that \(N_1\ge N_2\ge 2\), \(p,q\in \mathbb {N}\) and \(k_1,k_2\in \mathbb {R}\). In view of Lemmas 2.2, 2.3 and 2.4, we obtain

$$\begin{aligned}{} & {} r+k_1|h(r)|^p+k_2|g(r)|^q+\sum _{n=N_1}^\infty |a_n|r^n+\sum _{n=N_2}^\infty |b_n|r^n\nonumber \\{} & {} \quad \le r+k_1 r^p+k_2 r^q+\sum _{n=N_1}^\infty \frac{2Mr^n}{n(n-1)}+\sum _{n=N_2}^{N_1-1}\frac{Mr^n}{n(n-1)}.\end{aligned}$$
(5.16)

Now, we deduce that

$$\begin{aligned}{} & {} r+k_1 r^p+k_2 r^q+\sum _{n=N_1}^\infty \frac{2Mr^n}{n(n-1)}+\sum _{n=N_2}^{N_1-1}\frac{Mr^n}{n(n-1)}\nonumber \\{} & {} \quad = r+k_1 r^p+k_2 r^q+2M\left( \sum _{n=N_1}^\infty \frac{r^n}{n(n-1)}+\sum _{n=N_2}^{N_1-1}\frac{r^n}{2n(n-1)}\right) \nonumber \\{} & {} \quad =r+k_1 r^p+k_2 r^q+2M\left( (1-r)\ln (1-r)+(1-r)\sum _{n=1}^{N_1-2}\frac{r^n}{n}+\frac{r^{N_1-1}}{N_1-1}\right. \nonumber \\{} & {} \left. \quad \quad +\sum _{n=N_2}^{N_1-1}\frac{r^n}{2n(n-1)}\right) \le 1+2M(1-2\ln 2)\end{aligned}$$
(5.17)

for \(0<r\le r_{p,k_1,q,k_2,N_1,N_2}(M)\), where \(r_{p,k_1,q,k_2,N_1,N_2}(M)\) is the smallest root of \(\mathcal {G}_4(r)=0\) in (0, 1) and \(\mathcal {G}_4:[0,1)\rightarrow \mathbb {R}\) is defined by

$$\begin{aligned}{} & {} \mathcal {G}_4(r):=r+k_1 r^p+k_2 r^q-1+2M\\{} & {} \left( (1-r)\ln (1-r)-1+2\ln 2+(1-r)\sum _{n=1}^{N_1-2}\frac{r^n}{n}\right. \\ {}{} & {} \left. +\frac{r^{N_1-1}}{N_1-1}+\sum _{n=N_2}^{N_1-1}\frac{r^n}{2n(n-1)}\right) .\end{aligned}$$

Similarly as the proof of Theorem 4.1, we have \(\mathcal {G}_4\left( r_{p,k_1,q,k_2,N_1,N_2}(M)\right) =0\), i.e.,

$$\begin{aligned}{} & {} r_{p,k_1,q,k_2,N_1,N_2}(M)+k_1 r_{p,k_1,q,k_2,N_1,N_2}^p(M)+k_2 r_{p,k_1,q,k_2,N_1,N_2}^q(M)-1\nonumber \\{} & {} \quad \quad +2M\left( (1-r_{p,k_1,q,k_2,N_1,N_2}(M))\ln (1-r_{p,k_1,q,k_2,N_1,N_2}(M)) -1+2\ln 2\right. \nonumber \\{} & {} \quad \quad +\left. \frac{r_{p,k_1,q,k_2,N_1,N_2}^{N_1-1}(M)}{N_1-1}\right. \nonumber \\{} & {} \quad \quad \left. +(1-r_{p,k_1,q,k_2,N_1,N_2}(M))\sum _{n=1}^{N_1-2}\frac{r_{p,k_1,q,k_2,N_1,N_2}^n(M)}{n}+\sum _{n=N_2}^{N_1-1}\frac{2r_{p,k_1,q,k_2,N_1,N_2}^n(M)}{2n(n-1)}\right) =0.\nonumber \\ \end{aligned}$$
(5.18)

Combining (5.2), (5.16) and (5.17), we obtain for \(|z|=r\le r_{p,k_1,q,k_2,N_1,N_2}(M)\)

$$\begin{aligned} r+k_1|h(r)|^p+k_2|g(r)|^q+\sum _{n=N_1}^\infty |a_n|r^n+\sum _{n=N_2}^\infty |b_n|r^n\le d\left( f(0),\partial f(\mathbb {D})\right) .\end{aligned}$$

In order to show that \(r_{p,k_1,q,k_2,N_1,N_2}(M)\) is the best possible, we consider the function \(f=f_M\) defined by (5.6) and we again get (5.9). For \(|z|=r_{p,k_1,q,k_2,N_1,N_2}(M)\), it follows from (5.9) and (5.18) that

$$\begin{aligned}{} & {} r_{p,k_1,q,k_2,N_1,N_2}(M)+k_1|h(r_{p,k_1,q,k_2,N_1,N_2}(M))|^p+k_2|g(r_{p,k_1,q,k_2,N_1,N_2}(M))|^q\\{} & {} \quad \quad +\sum _{n=N_1}^\infty |a_n|r_{p,k_1,q,k_2,N_1,N_2}^n(M)+\sum _{n=N_2}^\infty |b_n|r_{p,k_1,q,k_2,N_1,N_2}^n(M)\\{} & {} \quad =r_{p,k_1,q,k_2,N_1,N_2}(M)+k_1 r_{p,k_1,q,k_2,N_1,N_2}^p(M)+k_2 r_{p,k_1,q,k_2,N_1,N_2}^q(M)+2M\times \\{} & {} \left( (1-r_{p,k_1,q,k_2,N_1,N_2}(M))\ln (1-r_{p,k_1,q,k_2,N_1,N_2}(M))+\frac{r_{p,k_1,q,k_2,N_1,N_2}^{N_1-1}(M)}{N_1-1}\right. \nonumber \\{} & {} \quad \quad \left. +(1\!-\!r_{p,k_1,q,\!k_2,N_1,N_2}(M))\sum _{n\!=\!1}^{N_1\!-\!2}\frac{r_{p,k_1,q,k_2,N_1,N_2}^n(M)}{n}\!+\!\sum _{n\!=\!N_2}^{N_1-1}\frac{r_{p,k_1,q,k_2,N_1,N_2}^n(M)}{2n(n\!-\!1)}\right) \\{} & {} \quad =1+2M(1-2\ln 2)=d\left( f_M(0),\partial f_M(\mathbb {D})\right) .\end{aligned}$$

Therefore, the radius \(r_{p,k_1,q,k_2,N_1,N_2}(M)\) is the best possible. Thus the result follows. \(\square \)

6 Improved Bohr Radius for the Class \(\mathcal {P}_{\mathcal {H}}^0(M)\)

In 2023, Ahamed et al. [3] generalized the harmonic versions of Theorem B for the class \(\mathcal {P}_{\mathcal {H}}^0(M)\) and obtained the following result.

Theorem G

[3] Let \(f\in \mathcal {P}_{\mathcal {H}}^0(M)\) be given by (1.10). Then

$$\begin{aligned} r+\sum _{n=2}^\infty \left( |a_n|+|b_n|\right) r^n+P\left( \frac{S_r}{\pi }\right) \le d\left( f(0),\partial f(\mathbb {D})\right) \end{aligned}$$
(6.1)

for \(r\le r_N(M)\), where \(P\left( \omega \right) =\omega ^N+\omega ^{N-1}+\cdots +\omega \) and \(r_N(M)\in (0,1)\) is the smallest root of the equation

$$\begin{aligned} r-1-2M(r-1+(1-r)\ln (1-r)+2\ln 2)+P\left( r^2+4M^2G(r)\right) =0,\nonumber \\\end{aligned}$$
(6.2)

where G(r) is defined by \(G(r):=r^2\left( Li_2(r^2)-1\right) +(1-r^2)\ln (1-r^2)\). The constant \(r_N(M)\) is the best possible.

In order to generalize Theorem G, we consider a N-th degree polynomial of the form

$$\begin{aligned} P\left( \omega \right) =c_{N}\left( \omega \right) ^N+c_{N-1}\left( \omega \right) ^{N-1}+c_{N-2}\left( \omega \right) ^{N-2}+\cdots +c_{1}\omega , \end{aligned}$$
(6.3)

where \(c_i\in \mathbb {R}\) \((1\le i\le N)\) with \(c_N\not =0\). Concerning improved Bohr radius for the class \(\mathcal {P}_{\mathcal {H}}^0(M)\), we have obtain the following results.

Theorem 6.1

Let \(f\in \mathcal {P}_{\mathcal {H}}^0(M)\) for \(0<M<\frac{1}{2(\ln 4-1)}\) be given by (1.10). Then

$$\begin{aligned} r+\sum _{n=N_1}^\infty |a_n||z|^n+\sum _{n=N_2}^\infty |b_n||z|^n+P\left( \frac{S_r}{\pi }\right) \le d\left( f(0),\partial f(\mathbb {D})\right) \end{aligned}$$
(6.4)

for \(r\le r_{N_1,N_2}(M)\) with \(N_1\ge N_2\ge 2\), where \(P\left( \omega \right) \) is defined in (6.3) and \(r_{N_1,N_2}(M)\in (0,1)\) is the smallest root of the equation

$$\begin{aligned}{} & {} r-1+2M\left( (1-r)\ln (1-r)-1+2\ln 2+(1-r)\sum _{n=1}^{N_1-2}\frac{r^n}{n}+\frac{r^{N_1-1}}{N_1-1}\right. \\{} & {} \quad \left. +\sum _{n=N_2}^{N_1-1 }\frac{r^n}{2n(n-1)}\right) \\ {}{} & {} +P\left( r^2+4M^2\left( (r^2-1)\ln (1-r^2)+r^2(Li_2(r^2)-1)\right) \right) =0.\end{aligned}$$

The constant \(r_{N_1,N_2}(M)\) is the best possible (Fig. 8 and 9; Tables 5 and 6).

Fig. 8
figure 8

The graphs of \(r_{3,2}(M)\) and \(r_{10,5}(M)\) in (0, 1) when \(P(r)=r^3+6r^2+3r\)

Fig. 9
figure 9

The graphs of \(r_{3,2}(M)\) and \(r_{10,4}(M)\) in (0, 1) when \(P(r)=r^2+2r\)

Table 5 This table shows the values of the roots \(r_{N_1,N_2}(M)\) for different values of M in \(\left( 0,\frac{1}{2(\ln 4-1)}\right) \) with \(N_1=3,5,10\), \(N_2=2,3,5\) and \(P(r)=r^3+6r^2+3r\)
Table 6 This table shows the values of the roots \(r_{N_1,N_2}(M)\) for different values of M in \(\left( 0,\frac{1}{2(\ln 4-1)}\right) \) with \(N_1=3,10\), \(N_2=2,4\) and \(P(r)=r^2+2r\)

As a consequence of Theorem 6.1, we obtain the following corollary.

Corollary 6.1

Let \(f\in \mathcal {P}_{\mathcal {H}}^0(M)\) for \(0<M<\frac{1}{2(\ln 4-1)}\) be given by (1.10). Then,

$$\begin{aligned} r+\sum _{n=N_1}^\infty |a_n|r^n+\sum _{n=N_2}^\infty |b_n|r^n+\frac{S_r}{\pi }\le d\left( f(0),\partial f(\mathbb {D})\right) \end{aligned}$$
(6.5)

for \(r\le r_{N_1,N_2}(M)\), where \(r_{N_1,N_2}(M)\in (0,1)\) is the smallest root of the equation

$$\begin{aligned}{} & {} r-1+2M\left( (1-r)\ln (1-r)-1+2\ln 2+(1-r)\sum _{n=1}^{N_1-2}\frac{r^n}{n}+\frac{r^{N_1-1}}{N_1-1}\right. \\{} & {} \left. +\sum _{n=N_2}^{N_1-1} \frac{r^n}{2n(n-1)}\right) +r^2+4M^2\left( (r^2-1)\ln (1-r^2)+r^2(Li_2(r^2)-1)\right) =0.\end{aligned}$$

The constant \(r_{N_1,N_2}(M)\) is the best possible.

7 Proof of the Theorem 6.1

Proof of Theorem 6.1

Let \(f\in \mathcal {P}_{\mathcal {H}}^0(M)\) be given by (1.10). For the analytic functions h and g, the area \(S_r\) of the disk \(|z|<r\) under the harmonic map f is given by

$$\begin{aligned}{} & {} S_r=\iint \limits _{\begin{array}{c} |z|<r \end{array}}\left( |h'(z)|^2-|g'(z)|^2\right) dx dy, \end{aligned}$$
(7.1)
$$\begin{aligned} \text {where}{} & {} \frac{1}{\pi }\iint \limits _{\begin{array}{c} |z|<r \end{array}}|h'(z)|^2dx dy=\sum _{n=1}^\infty n|a_n|^2r^{2n} \end{aligned}$$
(7.2)
$$\begin{aligned} \text {and}{} & {} \frac{1}{\pi }\iint \limits _{\begin{array}{c} |z|<r \end{array}}|g'(z)|^2dx dy=\sum _{n=2}^\infty n|b_n|^2r^{2n}. \end{aligned}$$
(7.3)

Combining (7.1), (7.2), (7.3) and Lemma 2.3, we obtain

$$\begin{aligned} \frac{S_r}{\pi }= & {} \frac{1}{\pi }\iint \limits _{\begin{array}{c} |z|<r \end{array}}\left( |h'(z)|^2-|g'(z)|^2\right) dx dy \nonumber \\= & {} r^2+\sum _{n=2}^\infty n|a_n|^2r^{2n}-\sum _{n=2}^\infty n|b_n|^2r^{2n}\nonumber \\= & {} r^2+\sum _{n=2}^\infty n\left( |a_n|+|b_n|\right) \left( |a_n|-|b_n|\right) r^{2n}\nonumber \\\le & {} r^2+\sum _{n=2}^\infty \frac{4M^2r^{2n}}{n(n-1)^2}\nonumber \\= & {} r^2+4M^2\left( (r^2-1)\ln (1-r^2)+r^2(Li_2(r^2)-1)\right) .\end{aligned}$$
(7.4)

Note that, \(N_1\ge N_2\ge 2\). In view of Lemmas 2.2, 2.3 and 2.4, we obtain

$$\begin{aligned}{} & {} r+\sum _{n=N_1}^\infty |a_n|r^n+\sum _{n=N_2}^\infty |b_n|r^n+P\left( \frac{S_r}{\pi }\right) \nonumber \\{} & {} =r+\sum _{n=N_1}^\infty \left( |a_n|+|b_n|\right) r^n+\sum _{n=N_2}^{N_1-1}|b_n|r^n+P\left( \frac{S_r}{\pi }\right) \nonumber \\{} & {} \le r+\sum _{n=N_1}^\infty \frac{2Mr^n}{n(n-1)}+\sum _{n=N_2}^{N_1-1}\frac{Mr^n}{n(n-1)}+P\left( r^2+4M^2\left( (r^2-1)\ln (1-r^2)\right. \right. \nonumber \\ {}{} & {} \left. \left. +r^2(Li_2(r^2)-1)\right) \right) \nonumber \\{} & {} =r+2M\left( (1-r)\ln (1-r)+(1-r)\sum _{n=1}^{N_1-2}\frac{r^n}{n}+\frac{r^{N_1-1}}{N_1-1}+\sum _{n=N_2}^{N_1-1}\frac{r^n}{2n(n-1)}\right) \nonumber \\{} & {} +P\left( r^2+4M^2\left( (r^2-1)\ln (1-r^2)+r^2(Li_2(r^2)-1)\right) \right) \nonumber \\ {}{} & {} \le 1+2M(1-2\ln 2)\end{aligned}$$
(7.5)

for \(0<r\le r_{N_1,N_2}(M)\), where \(r_{N_1,N_2}(M)\) is the smallest root of \(\mathcal {G}_5(r)=0\) in (0, 1) and \(\mathcal {G}_5:[0,1)\rightarrow \mathbb {R}\) be defined by

$$\begin{aligned}{} & {} \mathcal {G}_5(r):=r-1+2M\left( (1-r)\ln (1-r)-1+2\ln 2+(1-r)\sum _{n=1}^{N_1-2}\frac{r^n}{n}+\frac{r^{N_1-1}}{N_1-1}\right. \\ {}{} & {} \left. +\sum _{n=N_2}^{N_1-1}\frac{r^n}{2n(n-1)}\right) +P\left( r^2+4M^2\left( (r^2-1)\ln (1-r^2)+r^2(Li_2(r^2)-1)\right) \right) .\end{aligned}$$

Similarly as the proof of Theorem 4.1, we have \(\mathcal {G}_5\left( r_{N_1,N_2}(M)\right) =0\), i.e.,

$$\begin{aligned}{} & {} r_{N_1,N_2}(M)-1+2M\left( (1-r_{N_1,N_2}(M))\ln (1-r_{N_1,N_2}(M))-1+2\ln 2\right. \nonumber \\{} & {} \left. +\sum _{n=N_2}^{N_1-1}\frac{r_{N_1,N_2}^n(M)}{2n(n-1)}\right. \nonumber \\{} & {} \left. +(1-r_{N_1,N_2}(M))\sum _{n=1}^{N_1-2}\frac{r_{N_1,N_2}^n(M)}{n} +\frac{r_{N_1,N_2}^{N_1-1}(M)}{N_1-1}\right) +P\left( r_{N_1,N_2}^2(M)+4M^2\times \right. \nonumber \\{} & {} \left. \left( \left( r_{N_1,N_2}^2(M)-1\right) \ln \left( 1-r_{N_1,N_2}^2(M)\right) +r_{N_1,N_2}^2(M)\left( Li_2(r_{N_1,N_2}^2(M))-1\right) \right) \right) =0.\nonumber \\ \end{aligned}$$
(7.6)

Combining (5.2), (7.5) and (7.6), we obtain for \(|z|=r\le r_{N_1,N_2}(M)\)

$$\begin{aligned} r+\sum _{n=N_1}^\infty |a_n|r^n+\sum _{n=N_2}^\infty |b_n|r^n+P\left( \frac{S_r}{\pi }\right) \le d\left( f(0),\partial f(\mathbb {D})\right) .\end{aligned}$$

In order to show that \(r_{N_1,N_2}(M)\) is the best possible, we consider the function \(f=f_M\) defined by (5.6) and we again get (5.9). For \(|z|=r_{N_1,N_2}(M)\), it follows from (5.9) and (7.6) that

$$\begin{aligned}{} & {} r_{N_1,N_2}(M)+\sum _{n=N_1}^\infty |a_n|r_{N_1,N_2}^n(M)+\sum _{n=N_2}^\infty |b_n|r_{N_1,N_2}^n(M)+P\left( \frac{S_r}{\pi }\right) \\{} & {} =r_{N_1,N_2}(M)+2M\left( (1-r_{N_1,N_2}(M))\ln (1-r_{N_1,N_2}(M))+\sum _{n=N_2}^{N_1-1}\frac{r_{N_1,N_2}^n(M)}{2n(n-1)}\right. \nonumber \\ {}{} & {} \left. +(1-r_{N_1,N_2}(M))\sum _{n=1}^{N_1-2}\frac{r_{N_1,N_2}^n(M)}{n}+\frac{r_{N_1,N_2}^{N_1-1}(M)}{N_1-1}\right) +P\left( r_{N_1,N_2}^2(M)+4M^2\times \right. \nonumber \\ {}{} & {} \left. \left( \left( r_{N_1,N_2}^2(M)-1\right) \ln \left( 1-r_{N_1,N_2}^2(M)\right) +r_{N_1,N_2}^2(M)\left( Li_2(r_{N_1,N_2}^2(M))-1\right) \right) \right) \\{} & {} =1+2M(1-2\ln 2)=d\left( f_M(0),\partial f_M(\mathbb {D})\right) .\end{aligned}$$

Hence the radius \(r_{N_1,N_2}(M)\) is the best possible. This completes the proof. \(\square \)

8 Refined Bohr Radius for the Class \(\mathcal {P}_{\mathcal {H}}^0(M)\)

In this section, we establish some sharp results on the harmonic analogue of Theorem C for the class \(\mathcal {P}_{\mathcal {H}}^0(M)\) as follows.

Theorem 8.1

Let \(f\in \mathcal {P}_{\mathcal {H}}^0(M)\) for \(0<M<\frac{1}{2(\ln 4-1)}\) be given by (1.10). Then

$$\begin{aligned}{} & {} |f(z)|^p+\sum _{n=N}^\infty \left( |a_n|+|b_n|\right) r^n+\frac{r^q}{1-r^q}\sum _{n=N}^\infty (n(n-1))^{q-1}\left( |a_n|+|b_n|\right) ^{q}r^{qn}\\{} & {} \le d\left( f(0),\partial f(\mathbb {D})\right) \end{aligned}$$

for \(r\le r_{p,q,N}(M)\) with \(p,q(\not =1)\in \mathbb {N}\) and \(N\ge 2\), where \(r_{p,q,N}(M)\in (0,1)\) is the smallest root of the equation

$$\begin{aligned}{} & {} \left[ r+2M\left\{ r+(1-r)\ln (1-r)\right\} \right] ^p-1+2M\left( (1-r)\ln (1-r)+(1-r)\sum _{n=1}^{N-2}\frac{r^n}{n}\right. \\{} & {} \left. -1+2\ln 2+\frac{r^{N-1}}{N-1}\right) +\frac{(2rM)^q}{1-r^q}\left( (1-r^q)\ln (1-r^q)+(1-r^q)\sum _{n=1}^{N-2}\frac{r^{nq}}{n}+\frac{r^{q(N-1)}}{N-1}\right) =0.\nonumber \end{aligned}$$
(8.1)

The constant \(r_{p,q,N}(M)\) is the best possible (Figs. 10 and 11; Table 7).

Fig. 10
figure 10

The graphs of \(r_ {1, 2, 3}(M) \) and \(r_ {1, 2, 10}(M)\) in (0, 1)

Fig. 11
figure 11

The graphs of \(r_ {2, 3, 5}(M)\) and \(r_ {3, 3, 5}(M)\) in (0, 1)

Table 7 This table shows the values of the roots \(r_{p,q,N}(M)\) for different values of M in \(\left( 0,\frac{1}{2(\ln 4-1)}\right) \) with \(p=1,2,3\), \(q=2,3\) and \(N=3,5,10\)

9 Proof of the Theorem 8.1

Proof of Theorem 8.1

Let \(f\in \mathcal {P}_{\mathcal {H}}^0(M)\) be given by (1.10). Note that \(p,q(\not =1)\in \mathbb {N}\) and \(N\ge 2\). In view of Lemmas 2.3 and 2.4, we obtain

$$\begin{aligned}{} & {} |f(z)|^p+\sum _{n=N}^\infty \left( |a_n|+|b_n|\right) r^n+\frac{r^q}{1-r^q}\sum _{n=N}^\infty (n(n-1))^{q-1}\left( |a_n|+|b_n|\right) ^{q}r^{qn}\nonumber \\{} & {} \le \left( r+2M\sum _{n=2}^\infty \frac{r^{n}}{n(n-1)}\right) ^p+\sum _{n=N}^\infty \frac{2Mr^{n}}{n(n-1)}+\frac{r^q}{1-r^q}\sum _{n=N}^\infty \frac{2^qM^qr^{qn}}{n(n-1)}.\end{aligned}$$
(9.1)

Now,

$$\begin{aligned}{} & {} \left( r+2M\sum _{n=2}^\infty \frac{r^{n}}{n(n-1)}\right) ^p+\sum _{n=N}^\infty \frac{2Mr^{n}}{n(n-1)}+\frac{r^q}{1-r^q}\sum _{n=N}^\infty \frac{2^qM^qr^{qn}}{n(n-1)}\\{} & {} =\left[ r+2M\left\{ r+(1-r)\ln (1-r)\right\} \right] ^p\\{} & {} +2M\left( (1-r)\ln (1-r)+(1-r) \sum _{n=1}^{N-2}\frac{r^n}{n}+\frac{r^{N-1}}{N-1}\right) \\ {}{} & {} +\frac{(2rM)^q}{1-r^q}\left( (1-r^q)\ln (1-r^q)+(1-r^q)\sum _{n=1}^{N-2}\frac{r^{nq}}{n}+\frac{r^{q(N-1)}}{N-1}\right) \\{} & {} \le 1+2M(1-2\ln 2) \end{aligned}$$

for \(0<r\le r_{p,q,N}(M)\), where \( r_{p,q,N}(M)\) is the smallest root of \(\mathcal {G}_7(r)=0\) in (0, 1) and \(\mathcal {G}_7:[0,1)\rightarrow \mathbb {R}\) be defined by

$$\begin{aligned}{} & {} \mathcal {G}_7(r):=\left[ r+2M\left\{ r+(1-r)\ln (1-r)\right\} \right] ^p-1+2M\left( (1-r)\ln (1-r)+(1-r)\sum _{n=1}^{N-2}\frac{r^n}{n}\right. \\{} & {} \left. -1+2\ln 2+\frac{r^{N-1}}{N-1}\right) +\frac{(2rM)^q}{1-r^q}\left( (1-r^q)\ln (1-r^q)+(1-r^q)\sum _{n=1}^{N-2}\frac{r^{nq}}{n}+\frac{r^{q(N-1)}}{N-1}\right) .\end{aligned}$$

Similarly as the proof of Theorem 4.1, we have \(\mathcal {G}_7\left( r_{p,q,N}(M)\right) =0\), i.e.,

$$\begin{aligned}{} & {} \left[ r_{p,q,N}(M)+2M\left\{ r_{p,q,N}(M)+(1-r_{p,q,N}(M))\ln (1-r_{p,q,N}(M))\right\} \right] ^p\nonumber \\{} & {} +2M\left( (1-r_{p,q,N}(M))\times \right. \nonumber \\{} & {} \left. \ln (1-r_{p,q,N}(M))+(1-r_{p,q,N}(M))\sum _{n=1}^{N-2}\frac{r_{p,q,N}^n(M)}{n}-1+2\ln 2+\frac{r_{p,q,N}^{N-1}(M)}{N-1}\right) -1\nonumber \\{} & {} +\frac{(2r_{p,q,N}(M)M)^q}{1-r_{p,q,N}^q(M)}\left( (1-r_{p,q,N}^q(M))\ln (1-r_{p,q,N}^q(M))+(1-r_{p,q,N}^q(M))\sum _{n=1}^{N-2}\frac{r_{p,q,N}^{nq}(M)}{n}\right. \nonumber \\ {}{} & {} \left. +\frac{r_{p,q,N}^{q(N-1)}(M)}{N-1}\right) =0.\end{aligned}$$
(9.2)

Combining (5.2), (9.1) and (9.2), we obtain for \(|z|=r\le r_{p,q,N}(M)\)

$$\begin{aligned}{} & {} |f(z)|^p+\sum _{n=N}^\infty \left( |a_n|+|b_n|\right) r^n+\frac{r^q}{1-r^q}\sum _{n=N}^\infty (n(n-1))^{q-1}\left( |a_n|+|b_n|\right) ^{q}r^{qn}\\{} & {} \le d\left( f(0),\partial f(\mathbb {D})\right) .\end{aligned}$$

In order to show that \(r_{p,q,N}(M)\) is the best possible, we consider the function \(f=f_M\) defined by (5.6) and we again get (5.9). For \(|z|=r_{p,q,N}(M)\), it follows from (5.9) and (9.2) that

$$\begin{aligned}{} & {} |f(z)|^p+\sum _{n=N}^\infty \left( |a_n|+|b_n|\right) r_{p,q,N}^n(M)+\frac{r_{p,q,N}^q(M)}{1-r_{p,q,N}^q(M)}\sum _{n=N}^\infty (n(n-1))^{q-1}\\{} & {} \left( |a_n|+|b_n|\right) ^{q}r_{p,q,N}^{qn}(M)\\{} & {} \quad =\left[ r_{p,q,N}(M)+2M\left\{ r_{p,q,N}(M)+(1-r_{p,q,N}(M))\ln (1-r_{p,q,N}(M))\right\} \right] ^p\\{} & {} \quad \quad +2M\left( (1-r_{p,q,N}(M))\times \right. \nonumber \\{} & {} \quad \left. \ln (1-r_{p,q,N}(M))+(1-r_{p,q,N}(M))\sum _{n=1}^{N-2}\frac{r_{p,q,N}^n(M)}{n} +\frac{r_{p,q,N}^{N-1}(M)}{N-1}\right) +\frac{(2r_{p,q,N}(M)M)^q}{1-r_{p,q,N}^q(M)}\times \nonumber \\{} & {} \left( (1-r_{p,q,N}^q(M))\ln (1-r_{p,q,N}^q(M))+(1-r_{p,q,N}^q(M))\sum _{n=1}^{N-2}\frac{r_{p,q,N}^{nq}(M)}{n}+\frac{r_{p,q,N}^{q(N-1)}(M)}{N-1}\right) \\{} & {} \quad =1+2M(1-2\ln 2)=d\left( f_M(0),\partial f_M(\mathbb {D})\right) .\end{aligned}$$

Hence the radius \(r_{p,q,N}(M)\) is the best possible. Thus the result follows. \(\square \)