1 Introduction

In this paper, we use the standard notations of the Nevanlinna theory as presented in [11, 17, 50, 52]. By definition, two meromorphic functions \(F\) and \(G\) are said to share \(a\) IM if \(F-a\) and \(G-a\) assume the same zeros ignoring multiplicity. When \(a=\infty \), the zeros of \(F-a\) mean the poles of \(F\).

Let \(D\) be a domain in \(\mathbb {C}\) and let \({{\fancyscript{F}}}\) be meromorphic functions defined in the domain \(D\). Then \({{\fancyscript{F}}}\) is said to be normal in \(D\), in the sense of Montel, if for any sequence \(\{f_n\}\subset {{\fancyscript{F}}}\) there exists a subsequence \(\{f_{n_j}\}\) such that \(f_{n_j}\) converges spherically locally uniformly in \(D\), to a meromorphic function or \(\infty \) (cf. [15, 38]). For simplicity, we take \(\rightarrow \) to stand for convergence and \(\rightrightarrows \) for convergence spherically locally uniformly.

Let \({\fancyscript{M}}(D)\) (resp. \({\fancyscript{A}}(D)\)) be the set of meromorphic (resp. holomorphic) functions on \(D\). Let \(n\) be an integer and take a positive integer \(k\). We will study normality of the subset \({{\fancyscript{F}}}\) of \({\fancyscript{M}}(D)\) such that \(f^nf^{(k)}\) satisfies some conditions for each \(f\in {{\fancyscript{F}}}\).

First of all, we look at some background for the case \(n=0\). Hayman [17] proved that if \(F\in {\fancyscript{M}}(\mathbb {C})\) is transcendental, then either \(F^{(k)}\) assumes every finite non-zero complex number infinitely often for any positive integer \(k\) or \(F\) assumes every finite complex number infinitely often. A normality criterion corresponding to Hayman’s theorem is obtained by Gu [14] which is stated as follows: If \({{\fancyscript{F}}}\) is the family in \({\fancyscript{M}}(D)\) such that each \(f\in {{\fancyscript{F}}}\) satisfies \(f^{(k)}\not =a\) and \(f\not =b\), where \(a,b\) are two complex numbers with \(a\not =0\), then \({{\fancyscript{F}}}\) is normal in the sense of Montel. In particular, if \({{\fancyscript{F}}}\subset {\fancyscript{A}}(D)\), the normality criterion was conjectured by Montel (see [38], p. 125) for \(k=1\), and proved by Miranda [30]. Further, Yang [51] and Schwick [40] confirmed that the normality criterion due to Gu is true if \(a\) is replaced by a non-zero holomorphic function on \(D\). In 2001, Jiang and Gao [22] proved that if \({{\fancyscript{F}}}\) is the family in \({\fancyscript{A}}(D)\) such that the multiplicities of zeros of each \(f\in {{\fancyscript{F}}}\) are least \(k+m+2\) for another non-negative integer \(m\) and such that each pair \((f,g)\) of \({{\fancyscript{F}}}\) satisfies that \(f^{(k)}\) and \(g^{(k)}\) share \(a\) IM (ignoring multiplicity), where \(a\in {\fancyscript{A}}(D)\) and multiplicities of zeros of \(a\) are at most \(m\), then \(F\) is normal in \(D\), and obtained a similar result when \({{\fancyscript{F}}}\subset {\fancyscript{M}}(D)\). For other generations, see [35, 10, 23, 27, 28, 43, 44] and [46].

Next, we introduce some developments for the case \(n\ge 1\) and \(k=1\). In 1959, Hayman [16] proposed a conjecture: If \(F\in {\fancyscript{M}}(\mathbb {C})\) is transcendental, then \(F^nF'\) assumes every finite non-zero complex number infinitely often for any positive integer \(n\). Hayman himself [16, 18] showed that it is true for \(n\ge 3\), and for \(n=2, F\in {\fancyscript{A}}(\mathbb {C})\). Mues [31] confirmed the conjecture for \(n=2\) in 1979. Furthermore, the case of \(n=1\) was considered by Clunie [9] when \(F\in {\fancyscript{A}}(\mathbb {C})\), finally settled by Bergweiler and Eremenko [2], Chen and Fang [6]. Related to these results on value distribution, Hayman [18] conjectured that if \({{\fancyscript{F}}}\) is the family of \({\fancyscript{M}}(D)\) such that each \(f\in {{\fancyscript{F}}}\) satisfies \(f^nf'\not =a\) for a positive integer \(n\) and a non-zero complex number \(a\), then \({{\fancyscript{F}}}\) is normal. This conjecture has been confirmed by Yang and Zhang [54] (for \(n\ge 5\), and for \(n\ge 2\) with \({{\fancyscript{F}}}\subset {\fancyscript{A}}(D)\)), Gu [13] (for \(n=3, 4\)), Pang [34] (for \(n\ge 2\); cf. [12]) and Oshkin [32] (for \(n=1\) with \({{\fancyscript{F}}}\subset {\fancyscript{A}}(D)\); cf. [24]). Finally, Pang [34] (or see [6, 55, 56]) indicated that the conjecture for \(n=1\) is a consequence of his theorem and Chen-Fang’s theorem [6]. Recently, based on the ideas of sharing values, Zhang [58] proved that if \({{\fancyscript{F}}}\) is the family of \({\fancyscript{M}}(D)\) such that each pair \((f,g)\) of \({{\fancyscript{F}}}\) satisfies that \(f^nf'\) and \(g^ng'\) share a finite non-zero complex number \(a\) IM for \(n\ge 2\), then \({{\fancyscript{F}}}\) is normal. There are examples showing that this result is not true for the case \(n=1\). Further, Jiang [22] concluded that if \({{\fancyscript{F}}}\) is the family of \({\fancyscript{M}}(D)\) such that each pair \((f,g)\) of \({{\fancyscript{F}}}\) satisfies that \(f^nf'\) and \(g^ng'\) share \(a\) IM for \(n\ge 2m+2\), where \(a\in {\fancyscript{A}}(D)\) and multiplicities of zeros of \(a\) are at most \(m\), then \({{\fancyscript{F}}}\) is normal.

Similarly, we also have analogs related to some conditions of \(f\left( f^{(k)}\right) ^l\) for a positive integer \(l\). For example, Zhang and Song [60] announced that if \(F\in {\fancyscript{M}}(\mathbb {C})\) is transcendental; \(a(\not \equiv 0)\) a small function of \(F\); \(l\ge 2,\) then \(F\left( F^{(k)}\right) ^{l}-a\) has infinitely many zeros. A simple proof was given by Alotaibi [1]. The normality criterion corresponding to this result was obtained by Jiang and Gao [21] which is stated as follows: Let \(l, k\ge 2, m\ge 0\) be three integers such that \(m\) is divisible by \(l+1\) and suppose that \(a (\not \equiv 0)\) is a holomorphic function in \(D\) with zeros of multiplicity \(m\). If \({{\fancyscript{F}}}\) is the family of \({\fancyscript{A}}(D)\) (resp. \({\fancyscript{M}}(D)\)) such that multiplicities of zeros of each \(f\in {{\fancyscript{F}}}\) are at least \(k+m\) (resp. \(\max \{k+m,2m+2\}\)) and such that each pair \((f,g)\) of \({{\fancyscript{F}}}\) satisfies that \(f\left( f^{(k)}\right) ^{l}\) and \(g\left( g^{(k)}\right) ^{l}\) share \(a\) IM, then \({{\fancyscript{F}}}\) is normal. For more results related to this topic, see Hennekemper [19], Hu and Meng [20], Li [25, 26], Schwick [39], Wang and Fang [42], Yang et al. [49].

Finally, we consider general cases of \(n\ge 1\) and \(k\ge 1\). In 1994, Zhang and Li [61] proved that if \(F\in {\fancyscript{M}}(\mathbb {C})\) is transcendental, then \(F^nL[F]-a\) has infinitely many zeros for \(n\ge 2\) and \(a\ne 0, \infty \), where

$$\begin{aligned} L[F]=a_{k}F^{(k)}+a_{k-1}F^{(k-1)}+\cdots +a_0F \end{aligned}$$

in which \(a_{i}\) (\(i=0,1,2,\cdots ,k\)) are small functions of \(F\). In 1999, Pang and Zalcman [36] obtained a corresponding normality criterion as follows: If \({{\fancyscript{F}}}\) is the family of \({\fancyscript{A}}(D)\) such that zeros of each \(f\in {{\fancyscript{F}}}\) have multiplicities at least \(k\) and such that each \(f\in {{\fancyscript{F}}}\) satisfies \(f^nf^{(k)} \not =a\) for a non-zero complex number \(a\), then \({{\fancyscript{F}}}\) is normal. In 2005, Zhang [59] showed that when \(n\ge 2\), this result is also true if \(a\) is replaced by a non-vanishing holomorphic functions in \(D\). For other related results, see Meng and Hu [29], Qi [37], Wang [41], Xu [45], Yang and Hu [48], Yang and Yang [53].

Take three integers \(m\ge 0, k\ge 1\), and \(n\ge 2\). Let \(a\ (\not \equiv 0)\) be a holomorphic function in \(D\) such that multiplicities of zeros of \(a\) are at most \(m\) and divisible by \(n+1\). In this paper, we obtain the following normality criteria:

Theorem 1.1

Let \({{\fancyscript{F}}}\) be the family of \({\fancyscript{M}}(D)\) such that multiplicities of zeros of each \(f\in {{\fancyscript{F}}}\) are at least \(k+m\) and such that multiplicities of poles of \(f\) are at least \(m+1\) whenever \(f\) have zeros and poles. If each pair \((f,g)\) of \({{\fancyscript{F}}}\) satisfies that \(f^{n}f^{(k)}\) and \(g^{n}g^{(k)}\) share \(a\) IM, then \({{\fancyscript{F}}}\) is normal in \(D\).

In special, if \(a\) has no zeros, which means \(m=0\), then Theorem 1.1 has the following form:

Corollary 1.1

Let \({{\fancyscript{F}}}\) be the family of \({\fancyscript{M}}(D)\) such that multiplicities of zeros of each \(f\in {{\fancyscript{F}}}\) are at least \(k\). If each pair \((f,g)\) of \({{\fancyscript{F}}}\) satisfies that \(f^{n}f^{(k)}\) and \(g^{n}g^{(k)}\) share \(a\) IM, then \({{\fancyscript{F}}}\) is normal in \(D\).

It is easy to see that this result extends above normality criteria due to Pang and Zalcman [36], and Zhang [59]. Furthermore, we can improve partially the normality criterion due to Jiang [22] as follows:

Theorem 1.2

If \({{\fancyscript{F}}}\) is the family of \({\fancyscript{M}}(D)\) such that each \(f\in {{\fancyscript{F}}}\) satisfies that \(f^{n}f'\ne a\), then \({{\fancyscript{F}}}\) is normal in \(D\).

The condition \(a(z)\not \equiv 0\) in Theorem 1.1 and 1.2 is necessary. This fact can be illustrated by the following example:

Example 1.1

Let \(D=\{z\in \mathbb {C}\ |\ |z|<1\}\). Let \(a(z)\equiv 0\) and

$$\begin{aligned} {{\fancyscript{F}}}=\left\{ \left. f_j(z)=e^{j(z-\frac{1}{2})}\ \right| \ j=1,2,\ldots \right\} \end{aligned}$$

Obviously, \(f_i^{n}f_{i}^{(k)}\) and \(f_j^{n} f_{j}^{(k)}\) share \(a\) IM for distinct positive integers \(i\) and \(j\) \((\)resp. \(f_j^{n} f_{j}'\ne a)\), but the family \({{\fancyscript{F}}}\) is not normal at \(z=1/2\).

In Corollary 1.1, the condition that multiplicities of zeros of each \(f\in {{\fancyscript{F}}}\) are at least \(k\) is sharp. For example, we consider the following family:

Example 1.2

Denote \(D\) as in Example 1.1. Let \(a(z)=e^{z}\) and

$$\begin{aligned} {{\fancyscript{F}}}=\left\{ \left. f_j(z)=j\left( z-\frac{1}{2j}\right) ^{k-1}\ \right| \ j=1,2,\ldots \right\} . \end{aligned}$$

Any \(f_{j} \in {{\fancyscript{F}}}\) has only a zero of multiplicity \(k-1\) in \(D\) and for distinct positive integers \(i\) and \(j, f_i^{n} f_{i}^{(k)}\) and \(f_j^{n} f_{j}^{(k)}\) share \(a\) IM. However, the family \({{\fancyscript{F}}}\) is not normal at \(z=0\).

2 Preliminary Lemmas

In order to prove our results, we require the following Zalcman’s lemma (cf. [56]):

Lemma 2.1

Take a positive integer \(k\). Let \(\fancyscript{F}\) be a family of meromorphic functions in the unit disk \(\triangle \) with the property that zeros of each \(f\in \fancyscript{F}\) are of multiplicity at least \(k\). If \(\fancyscript{F}\) is not normal at a point \(z_0\in \Delta \), then for \(0\le \alpha < k\), there exist a sequence \(\{z_n\}\subset \Delta \) of complex numbers with \(z_n\rightarrow z_0\); a sequence \(\{f_n\}\) of \(\fancyscript{F}\); and a sequence \(\{\rho _n\}\) of positive numbers with \(\rho _n\rightarrow 0\) such that \(g_n(\xi )=\rho _n^{-\alpha }f_n(z_n+\rho _n\xi )\) locally uniformly \((\)with respect to the spherical metric\()\) to a non-constant meromorphic function \(g(\xi )\) on \(\mathbb {C}\). Moreover, the zeros of \(g(\xi )\) are of multiplicity at least \(k\), and the function \(g(\xi )\) may be taken to satisfy the normalization \(g^{\sharp }(\xi )\le g^{\sharp }(0)=1\) for any \(\xi \in \mathbb {C}\). In particular, \(g(\xi )\) has at most order 2.

This result is Pang’s generalization (cf. [33, 35, 47]) to the Main Lemma in [55] (where \(\alpha \) is taken to be \(0\)), with improvements due to Schwick [39], Chen and Gu [7]. In Lemma 2.1, the order of \(g\) is defined using the Nevanlinna’s characteristic function \(T(r,g)\):

$$\begin{aligned} \mathrm{ord}(g)=\limsup _{r\rightarrow \infty }\frac{\log T(r,g)}{\log r}. \end{aligned}$$

Here, as usual, \(g^{\sharp }\) denotes the spherical derivative

$$\begin{aligned} g^{\sharp }(\xi )=\frac{|g'(\xi )|}{1+|g(\xi )|^2}. \end{aligned}$$

Lemma 2.2

Let \(p\ge 0, k\ge 1\), and \(n\ge 2\) be three integers, and let \(a\) be a non-zero polynomial of degree \(p\). If \(f\) is a non-constant rational function which has only zeros of multiplicity at least \(k+p\) and has only poles of multiplicity at least \(p+1\), then \(f^nf^{(k)}-a\) has at least one zero.

Proof

If \(f\) is a polynomial, then \(f^{(k)}\not \equiv 0\) since \(f\) is non-constant and has only zeros of multiplicity at least \(k+p\) which further means \(\deg (f)\ge k+p\). Noting that \(n\ge 2\), we immediately obtain that

$$\begin{aligned} \deg \left( f^nf^{(k)}\right) \ge n\deg (f)\ge n(k+p)>p=\deg (a). \end{aligned}$$

Therefore, it follows that \(f^nf^{(k)}-a\) is also a non-constant polynomial, and hence \(f^nf^{(k)}-a\) has at least one zero. Next, we assume that \(f\) has poles. Set

$$\begin{aligned} f(z)=\frac{A(z-\alpha _1)^{m_1}(z-\alpha _2)^{m_2}\cdots (z-\alpha _{s})^{m_{s}}}{(z-\beta _1)^{n_1}(z-\beta _2)^{n_2}\cdots (z-\beta _{t})^{n_{t}}}, \end{aligned}$$
(2.1)

where \(A\) is a non-zero constant, \(\alpha _i\) distinct zeroes of \(f\) with \(s\ge 0\), and \(\beta _j\) distinct poles of \(f\) with \(t\ge 1\). For simplicity, we put

$$\begin{aligned}&m_1+m_2+\cdots +m_{s}=M\ge (k+p)s,\end{aligned}$$
(2.2)
$$\begin{aligned}&n_1+n_2+\cdots +n_{t}=N\ge (p+1)t. \end{aligned}$$
(2.3)

From Eq. (2.1), we obtain

$$\begin{aligned} f^{(k)}(z)=\frac{(z-\alpha _1)^{m_1-k}(z-\alpha _2)^{m_2-k}\cdots (z-\alpha _{s})^{m_{s}-k}g(z)}{(z-\beta _1)^{n_1+k}(z-\beta _2)^{n_2+k}\cdots (z-\beta _{t})^{n_{t}+k}}, \end{aligned}$$
(2.4)

where \(g\) is a polynomial of degree \(\le k(s+t-1)\). From Eqs. (2.1) and (2.7), we get

$$\begin{aligned} f^n(z)f^{(k)}(z)=\frac{A^n(z-\alpha _1)^{M_1}(z-\alpha _2)^{M_2}\cdots (z-\alpha _s)^{M_s}g(z)}{(z-\beta _1)^{N_1}(z-\beta _2)^{N_2}\cdots (z-\beta _{t})^{N_t}}, \end{aligned}$$
(2.5)

in which

$$\begin{aligned} M_{i}&= (n+1)m_{i}-k, \quad \ i=1,2,\cdots ,s,\\ N_{j}&= (n+1)n_{j}+k, \quad \ j=1,2,\cdots ,t. \end{aligned}$$

Differentiating Eq. (2.5) yields

$$\begin{aligned} \left\{ f^nf^{(k)}\right\} ^{({p}+1)}(z)=\frac{(z-\alpha _1)^{M_1-p-1} (z-\alpha _2)^{M_2-p-1}\cdots (z-\alpha _s)^{M_s-p-1}g_{0}(z)}{(z-\beta _1)^{N_1+p+1}\cdots (z-\beta _t)^{N_{t+p}+1}}, \end{aligned}$$
(2.6)

where \(g_{0}(z)\) is a polynomial of degree \(\le (p+k+1)(s+t-1)\). We assume, to the contrary, that \(f^{n}f^{(k)}-a\) has no zero, then

$$\begin{aligned} f^n(z)f^{(k)}(z)=a(z)+\frac{C}{(z-\beta _1)^{N_1}(z-\beta _2)^{N_2}\cdots (z-\beta _t)^{N_t}}, \end{aligned}$$
(2.7)

where \(C\) is a non-zero constant. Subsequently, Eq. (2.12) yields

$$\begin{aligned} \left\{ f^nf^{(k)}\right\} ^{(p+1)}(z)=\frac{g_{1}(z)}{(z-\beta _1)^{N_1+p+1}\cdots (z-\beta _t)^{N_t+p+1}}, \end{aligned}$$
(2.8)

where \(g_1(z)\) is a polynomial of degree \(\le (p+1)(t-1)\).

Comparing Eq. (2.6) with Eq. (2.8), we get

$$\begin{aligned} (p+1)(t-1)\ge \deg (g_1)\ge (n+1)M-ks-(p+1)s, \end{aligned}$$

and hence

$$\begin{aligned} M < \frac{p+k+1}{n+1}s+\frac{p+1}{n+1}t. \end{aligned}$$
(2.9)

From Eqs. (2.5) and (2.7), we have

$$\begin{aligned} (n+1)N+kt+p=(n+1)M-ks+\deg (g). \end{aligned}$$

Since \(\deg (g)\le k(s+t-1)\), we find

$$\begin{aligned} (n+1)N \le (n+1)M-ks+k(s+t-1)-kt-p, \end{aligned}$$

and thus

$$\begin{aligned} N < M. \end{aligned}$$
(2.10)

By Eqs. (2.9), (2.10) and noting that \(M\ge (k+p)s, N\ge (p+1)t\), we deduce that

$$\begin{aligned} M < \frac{p+k+1}{n+1}s+\frac{p+1}{n+1}t \le \left\{ \frac{p+k+1}{(n+1)(k+p)}+\frac{1}{n+1}\right\} M. \end{aligned}$$
(2.11)

Note that \(n \ge 2\) implies

$$\begin{aligned} \frac{p+k+1}{(n+1)(k+p)}+\frac{1}{n+1}=\frac{2(k+p)+1}{(n+1)(k+p)}\le 1. \end{aligned}$$

Hence it follows from Eq. (2.11) that \(M<M\), which is a contradiction. Lemma 2.2 is proved.\(\square \)

Lemma 2.3

Let \(p\ge 0, k\ge 1\), and \(n\ge 2\) be three integers, and let \(a\) be a non-zero polynomial of degree \(p\). If \(f\) is a non-constant rational function which has only zeros of multiplicity at least \(k+p\) and has only poles of multiplicity at least \(p+1\), then \(f^nf^{(k)}-a\) has at least two distinct zeros.

Proof

Lemma 2.2 implies that \(f^{n}f^{(k)}-a\) has at least one zero. Assume, to the contrary, that \(f^{n}f^{(k)}-a\) has only one zero \(z_{0}\). If \(f\) is a polynomial, then we can write

$$\begin{aligned} f^{n}(z)f^{(k)}(z)-a(z)=A'(z-z_{0})^{d}, \end{aligned}$$

where \(A'\) is a non-zero constant and \(d\) is a positive integer. Since \(f\) is a non-constant polynomial which has only zeros of multiplicity at least \(k+p\), we find \(f^{(k)}\not \equiv 0\), and hence

$$\begin{aligned} d= \deg (f^{n}f^{(k)}-a) \ge \deg (f^{n})\ge n(k+p) \ge 2p+2. \end{aligned}$$

By computing, we find

$$\begin{aligned} \left\{ f^nf^{(k)}\right\} ^{(p+1)}(z)=A'd(d-1)\ldots (d-p)(z-z_{0})^{{d}-p-1}, \end{aligned}$$

hence \(\left\{ f^nf^{(k)}\right\} ^{(p+1)}\) has a unique zero \(z_{0}\). Take a zero \(\xi _{0}\) of \(f\), then it is a zero of \(f^{n}\) with multiplicity at least \(2p+2\). It follows that \(\xi _{0}\) is a common zero of \(\left\{ f^nf^{(k)}\right\} ^{(p)}\) and \(\left\{ f^nf^{(k)}\right\} ^{(p+1)}\), which further implies that \(\xi _{0}=z_{0}\). Therefore, we obtain \(\left\{ f^nf^{(k)}\right\} ^{(p)}(z_{0})=0.\)

On the other hand, we get

$$\begin{aligned} \left\{ f^{n}f^{(k)}\right\} ^{(p)}(z)=a^{(p)}(z)+A'd(d-1)\ldots (d-p+1)(z-z_{0})^{{d-p}}, \end{aligned}$$

which means

$$\begin{aligned} \left\{ f^{n}f^{(k)}\right\} ^{(p)}(z_{0})=a^{(p)}(z_0)\ne 0 \end{aligned}$$

since \(\deg (a)=p\). This is contradictory to \(\left\{ f^{n}f^{(k)}\right\} ^{(p)}(z_{0})=0\).

If \(f\) has poles, we can express \(f\) by Eq. (2.1) again, and then find

$$\begin{aligned} f^n(z)f^{(k)}(z)=a(z)+\frac{C'(z-z_{0})^{l}}{(z-\beta _1)^{N_1}(z-\beta _2)^{N_2}\cdots (z-\beta _t)^{N_t}}, \end{aligned}$$
(2.12)

where \(C'\) is a non-zero constant and \(l\) is a positive integer. We distinguish two cases to deduce contradictions.

Case 1 \(p \ge l\). Since \(p\ge l\), the expression Eq. (2.5) together with Eq. (2.12) implies that

$$\begin{aligned} (n+1)N+kt+p=(n+1)M-ks+\deg (g). \end{aligned}$$

Therefore, we can also conclude Eq. (2.10), that is, \(N<M\). Differentiating Eq. (2.12), we obtain

$$\begin{aligned} \left\{ f^nf^{(k)}\right\} ^{(p+1)}(z)= \frac{g_{2}(z)}{(z-\beta _1)^{N_1+p+1}\cdots (z-\beta _t)^{N_t+p+1}}, \end{aligned}$$

where \(g_2(z)\) is a polynomial of degree at most \((p+1)t-(p-l+1)\), and hence

$$\begin{aligned} (p+1)t-(p-l+1)\ge \deg (g_2)\ge (n+1)M-ks-(p+1)s, \end{aligned}$$

where the last estimate follows from Eq. (2.6). Then we have

$$\begin{aligned} \frac{p-l}{n+1} < \frac{p+k+1}{n+1}s+\frac{p+1}{n+1}t -M\le \left\{ \frac{p+k+1}{(n+1)(k+p)}+\frac{1}{n+1}-1\right\} M\nonumber \\ \end{aligned}$$
(2.13)

since \(M\ge (k+p)s,N\ge (p+1)t,M>N\). It follows that

$$\begin{aligned} \frac{p+k+1}{(n+1)(k+p)}+\frac{1}{n+1} \le 1 \end{aligned}$$

since \(n \ge 2\). Therefore, from Eq. (2.13), we conclude that \(p-l < 0\), a contradiction with the assumption \(p\ge l\).

Case 2. \(l > p\). The expression Eq. (2.12) yields

$$\begin{aligned} \left\{ f^nf^{(k)}\right\} ^{(p+1)}(z)=\frac{(z-z_{0})^{l-p-1}g_{3}(z)}{(z-\beta _1)^{N_1+p+1}\cdots (z-\beta _t)^{N_t+p+1}}, \end{aligned}$$
(2.14)

where \(g_{3}(z)\) is a polynomial with \(\deg (g_{3}) \le (p+1)t\). We claim that \(z_0\not =\alpha _i\) for each \(i\). Otherwise, if \(z_0=\alpha _i\) for some \(i\), then Eq. (2.12) yields

$$\begin{aligned} a^{(p)}(z_0)=\left\{ f^nf^{(k)}\right\} ^{(p)}(\alpha _i)=0 \end{aligned}$$

because each \(\alpha _i\) is a zero of \(f^nf^{(k)}\) of multiplicity \(\ge n(k+p)\ge 2p+2\). This is impossible since \(\deg (a)=p\). Hence \((z-z_{0})^{l-p-1}\) is a factor of the polynomial \(g_0\) in Eq. (2.6). By Eqs. (2.6) and (2.14), we conclude that

$$\begin{aligned} (p+1)t \ge \deg (g_{3})\ge (n+1)M-ks-(p+1)s, \end{aligned}$$

which is equivalent to

$$\begin{aligned} M\le \frac{p+k+1}{n+1}s+\frac{p+1}{n+1}t. \end{aligned}$$
(2.15)

If  \(l\ne (n+1)N+kt+p\), then Eq. (2.5) together with Eq. (2.12) implies

$$\begin{aligned} (n+1)N+kt+p \le (n+1)M-ks+\deg (g), \end{aligned}$$

so we get \(N< M\) from \(\deg (g)\le k(s+t-1)\). Therefore, using the facts \(M\ge (k+p)s,N\ge (p+1)t\), Eq. (2.15) implies a contradiction

$$\begin{aligned} M<\left\{ \frac{p+k+1}{(n+1)(k+p)}+\frac{1}{n+1}\right\} M\le M. \end{aligned}$$

Hence \(l= (n+1)N+kt+p\).

Now we must have \(N\ge M\), otherwise, when \(N<M\), we can deduce the contradiction \(M<M\) from Eq. (2.15). Comparing Eq. (2.6) with Eq. (2.14), we find

$$\begin{aligned} (p+k+1)(s+t-1) \ge \deg (g_{0}) \ge l-p-1 \end{aligned}$$

since \((z-z_0)^{l-p-1}|g_0\), and hence

$$\begin{aligned} (n+1)N+kt+p = l \le (p+k+1)s+(p+k+1)t-k, \end{aligned}$$

which further yields

$$\begin{aligned} N<\frac{p+k+1}{n+1}s+\frac{p+1}{n+1}t. \end{aligned}$$

Since \(M \ge (k+p)s\) and \(N\ge (p+1)t\), it follows from Eq. (2.15) that

$$\begin{aligned} N< \frac{p+k+1}{(n+1)(k+p)}M+\frac{1}{n+1}N. \end{aligned}$$

Hence \(N\ge M\) yields

$$\begin{aligned} N < \left\{ \frac{p+k+1}{(n+1)(k+p)}+\frac{1}{n+1}\right\} N. \end{aligned}$$
(2.16)

Since \(n \ge 2\), we obtain consequently

$$\begin{aligned} \frac{p+k+1}{(n+1)(k+p)} + \frac{1}{n+1} \le 1. \end{aligned}$$

Hence Eq. (2.16) yields \(N < N\). This is a contradiction. Proof of Lemma 2.3 is completed.\(\square \)

Lemma 2.4

Let \(p\ge 0\) and \(n\ge 2\) be two integers such that \(p\) is divisible by \(n+1\), and let \(a\) be a non-zero polynomial of degree \(p\). If \(f\) is a non-constant rational function, then \(f^nf'-a\) has at least one zero.

Proof

If \(f\) is a non-constant polynomial, then \(f'\not \equiv 0\). We consequently conclude that

$$\begin{aligned} \deg \left( f^nf'\right) =(n+1)\deg (f)-1\ne p \end{aligned}$$

since \(p\) is divisible by \(n+1\). It follows that \(f^nf'-a\) is also a non-constant polynomial, so that \(f^nf'-a\) has at least one zero.

If \(f\) has poles, we can express \(f\) by Eq. (2.1) again, and then by differentiating Eq. (2.1), we deduce that

$$\begin{aligned} f'(z)=\frac{(z-\alpha _1)^{m_1-1}(z-\alpha _2)^{{m}_2-1}\cdots (z-\alpha _s)^{{m_s}-1}h(z)}{(z-\beta _1)^{n_1+1}(z-\beta _2)^{n_2+1}\cdots (z-\beta _t)^{n_t+1}}, \end{aligned}$$
(2.17)

where \(h(z)\) is a polynomial of form

$$\begin{aligned} h(z)=(M-N)z^{{s+t}-1}+\cdots . \end{aligned}$$

From Eqs. (2.1) and (2.17), we obtain

$$\begin{aligned} f^nf'=\frac{P}{Q}, \end{aligned}$$

in which

$$\begin{aligned}&\displaystyle P(z) = A^n(z-\alpha _1)^{(n+1){m}_1-1}(z-\alpha _2)^{(n+1){m}_2-1}\cdots (z-\alpha _s)^{(n+1){m}_s-1}h(z),\\&\displaystyle Q(z) = (z-\beta _1)^{(n+1)n_1+1}(z-\beta _2)^{(n+1)n_2+1}\cdots (z-\beta _t)^{(n+1)n_t+1}.\qquad \qquad \end{aligned}$$

We suppose, to the contrary, that \(f^{n}f'-a\) has no zero. When \(M\not =N\), we have

$$\begin{aligned} f^nf'=a+\frac{B}{Q}=\frac{P}{Q}, \end{aligned}$$

where \(B\) is a non-zero constant. Therefore, we obtain

$$\begin{aligned} \deg (P)=\deg (Qa+B)=\deg (Q)+p. \end{aligned}$$

This implies that

$$\begin{aligned} (n+1)M-s+(s+t-1)= (n+1)N+t+p, \end{aligned}$$

or equivalently

$$\begin{aligned} M-N=\frac{p+1}{n+1}, \end{aligned}$$

in which \(p\) is divisible by \(n+1\). This is impossible since \(M-N\) is an integer.

If \(M=N\), we can rewrite Eq. (2.1) as follows

$$\begin{aligned} f(z)=A+ \frac{B'(z-\gamma _1)^{l_1}(z-\gamma _2)^{l_2}\cdots (z-\gamma _r)^{l_{r}}}{(z-\beta _1)^{n_1}(z-\beta _2)^{n_2}\cdots (z-\beta _t)^{n_t}}, \end{aligned}$$

where \(B'\) is a non-zero constant, \(\gamma _i\) are distinct with \(l_i\ge 1, r\ge 0\), and

$$\begin{aligned} M'=l_1+\cdots +l_{r}<N. \end{aligned}$$

Thus we find

$$\begin{aligned} f'(z)= \frac{(z-\gamma _1)^{l_1-1}(z-\gamma _2)^{l_2-1}\cdots (z-\gamma _{r})^{l_r-1}\hbar (z)}{(z-\beta _1)^{n_1+1}(z-\beta _2)^{n_2+1}\cdots (z-\beta _t)^{n_t+1}} , \end{aligned}$$

where \(\hbar (z)\) is a polynomial of form

$$\begin{aligned} \hbar (z)=(M'-N)z^{r+t-1}+\cdots . \end{aligned}$$

Similarly, since \(\deg (P)=\deg (Q)+p\), we have

$$\begin{aligned} nM+M'-r+(r+t-1)=(n+1)N+t+p=(n+1)M+t+p, \end{aligned}$$

that is,

$$\begin{aligned} M'=M+p+1. \end{aligned}$$

This is impossible since \(M'<N=M\). Therefore, \(f^{n}f'-a\) has at least one zero.    \(\square \)

The following lemma is a direct consequence of a result from [61]:

Lemma 2.5

Let \(n, k\) be two positive integers with \(n\ge 2\), and let \(a\ (\not \equiv 0)\) be a polynomial. If \(f\) is a transcendental meromorphic function in \(\mathbb {C}\), then \(f^nf^{(k)}-a\) has infinitely zeros.

3 Proof of Theorem 1.1

Without loss of generality, we may assume that \(D=\{z\in \mathbb {C}\ |\ |z|<1\}\). For any point \(z_{0}\) in \(D\), either \(a(z_{0})=0\) or \(a(z_{0})\ne 0\) holds. For simplicity, we assume \(z_{0}=0\) and distinguish two cases.

Case 1 \(a(0) \ne 0\). To the contrary, we suppose that \({{\fancyscript{F}}}\) is not normal at \(z_{0}=0\). Then, by Lemma 2.1, there exist a sequence \(\{z_j\}\) of complex numbers with \(z_j\rightarrow 0\) \((j\rightarrow \infty )\); a sequence \(\{f_j\}\) of \({{\fancyscript{F}}}\); and a sequence \(\{\rho _j\}\) of positive numbers with \(\rho _j\rightarrow 0\) \((j\rightarrow \infty )\) such that

$$\begin{aligned} g_j(\xi )=\rho _j^{-\frac{k}{n+1}}f_j(z_j+\rho _j\xi ) \end{aligned}$$

converges uniformly to a non-constant meromorphic function \(g(\xi )\) in \(\mathbb {C}\) with respect to the spherical metric. Moreover, \(g(\xi )\) is of order at most 2. By Hurwitz’s theorem, the zeros of \(g(\xi )\) have at least multiplicity \(k+m\).

On every compact subset of \(\mathbb {C}\) which contains no poles of \(g\), we have uniformly

$$\begin{aligned}&f_j^{n}(z_j+\rho _j\xi ) f^{(k)}_j(z_j+\rho _j\xi ) -a(z_j+\rho _j\xi )\nonumber \\&\qquad = g_j^{n}(\xi )g^{(k)}_j(\xi )-a(z_j+\rho _j\xi )\rightrightarrows g^{n}(\xi ) g^{(k)}(\xi ) -a(0). \end{aligned}$$
(3.1)

If \(g^{n}g^{(k)}\equiv a(0)\), then \(g\) has no zeros and poles. Then there exist constants \(c_i\) such that \((c_1,c_2)\not =(0,0)\), and

$$\begin{aligned} g(\xi )=e^{c_0+c_1\xi +c_2\xi ^2} \end{aligned}$$

since \(g\) is a non-constant meromorphic function of order at most 2. Obviously, this is contrary to the case \(g^{n}g^{(k)}\equiv a(0)\). Hence we have \(g^{n}g^{(k)}\not \equiv a(0)\).

By Lemmas 2.3 and 2.5, the function \(g^{n}g^{(k)}-a(0)\) has two distinct zeros \(\xi _{0}\) and \(\xi _{0}^{*}\). We choose a positive number \(\delta \) small enough such that \(D_1\cap D_2=\emptyset \) and such that \(g^{n}g^{(k)}-a(0)\) has no other zeros in \(D_1 \cup D_2\) except for \(\xi _{0}\) and \(\xi _{0}^{*}\), where

$$\begin{aligned} D_1=\{\xi \in \mathbb {C}\ |\ |\xi -\xi _0|<\delta \},\quad D_2=\{\xi \in \mathbb {C}\ |\ |\xi -\xi _0^*|<\delta \}. \end{aligned}$$

By Eq. (3.1) and Hurwitz’s theorem, there exist points \(\xi _j\in D_1, \xi _j^*\in D_2\) such that

$$\begin{aligned} f^{n}_j(z_j+\rho _j\xi _j)f^{(k)}_j(z_j+\rho _j\xi _j)-a(z_j+\rho _j\xi _j)=0, \end{aligned}$$

and

$$\begin{aligned} f^{n}_j(z_j+\rho _j\xi _j^*)f^{(k)}_j(z_j+\rho _j\xi _j^*)-a(z_j+\rho _j\xi _j^{*})=0 \end{aligned}$$

for sufficiently large \(j\).

By the assumption in Theorem 1.1, \(f_1^{n} f_1^{(k)}\) and \(f_j^{n} f_j^{(k)}\) share \(a\) IM for each \(j\). It follows

$$\begin{aligned} f_1^{n}(z_j+\rho _j\xi _j) f_1^{(k)}(z_j+\rho _j\xi _j) -a(z_j+\rho _j\xi _j)=0, \end{aligned}$$

and

$$\begin{aligned} f_1^{n}(z_j+\rho _j\xi _j^*) f_1^{(k)}(z_j+\rho _j\xi _j^*) -a(z_j+\rho _j\xi _j^{*})=0. \end{aligned}$$

By letting \(j\rightarrow \infty \) and noting \(z_j+\rho _j\xi _j\rightarrow 0, z_j+\rho _j\xi _j^*\rightarrow 0\), we obtain

$$\begin{aligned} f^{n}_1(0) f^{(k)}_1(0) -a(0)=0. \end{aligned}$$

Since the zeros of \(f^{n}_1(\xi ) f^{(k)}_1(\xi ) -a(\xi )\) have no accumulation points, in fact we have

$$\begin{aligned} z_j+\rho _j\xi _j= 0, \quad \ z_j+\rho _j\xi _j^*= 0, \end{aligned}$$

or equivalently

$$\begin{aligned} \xi _j=-\frac{z_j}{\rho _j},\qquad \xi _j^*=-\frac{z_j}{\rho _j}. \end{aligned}$$

This contradicts with the facts that \(\xi _j\in D_1, \xi _j^*\in D_2, D_1\cap D_2=\emptyset \). Thus \({{\fancyscript{F}}}\) is normal at \(z_{0}=0\).

Case 2 \(a(0)=0\). We assume that \(z_{0}=0\) is a zero of \(a\) of multiplicity \(p\). Then we have \(p\le m\) by the assumption. Write \(a(z)=z^pb(z)\), in which \(b(0)=b_p \ne 0\). Since multiplicities of all zeros of \(a\) are divisible by \(n+1\), then \(d=p/(n+1)\) is just a positive integer. Thus we obtain a new family of \({\fancyscript{M}}(D)\) as follows

$$\begin{aligned} \fancyscript{H}= \left. \left\{ \frac{f(z)}{z^d}\ \right| \ f\in {{\fancyscript{F}}}\right\} . \end{aligned}$$

We claim that \( \fancyscript{H}\) is normal at \(0\).

Otherwise, if \(\fancyscript{H}\) is not normal at \(0\), then by lemma 2.1, there exist a sequence \(\{z_j\}\) of complex numbers with \(z_j\rightarrow 0\) \((j\rightarrow \infty )\); a sequence \(\{h_j\}\) of \(\fancyscript{H}\); and a sequence \(\{\rho _j\}\) of positive numbers with \(\rho _j\rightarrow 0\) \((j\rightarrow \infty )\) such that

$$\begin{aligned} g_j(\xi )=\rho _j^{-\frac{k}{n+1}}h_j(z_j+\rho _j\xi ) \end{aligned}$$
(3.2)

converges uniformly to a non-constant meromorphic function \(g(\xi )\) in \(\mathbb {C}\) with respect to the spherical metric, where \(g^{\sharp }(\xi ) \le 1, \mathrm{ord}(g)\le 2\), and \(h_j\) has the following form

$$\begin{aligned} h_j(z)=\frac{f_j(z)}{z^d}. \end{aligned}$$

We will deduce contradiction by distinguishing two cases.

Subcase 2.1 There exists a subsequence of \(z_{j} / \rho _{j}\), for simplicity we still denote it as \(z_{j} / \rho _{j}\), such that \(z_{j} / \rho _{j}\rightarrow c\) as \(j\rightarrow \infty \), where \(c\) is a finite number. Thus we have

$$\begin{aligned} F_{j}(\xi )=\frac{f_{j}(\rho _{j}\xi )}{\rho _{j}^{\frac{k}{n+1}+d}}= \frac{(\rho _{j}\xi )^{d}h_{j}(z_{j}+\rho _{j}(\xi -\frac{z_{j}}{\rho _{j}}))}{(\rho _{j})^{d}(\rho _{j})^{\frac{k}{n+1}}} \rightrightarrows \xi ^{d} g(\xi -c)=h(\xi ), \end{aligned}$$

and

$$\begin{aligned} F_{j}^{n}(\xi )F_{j}^{(k)}(\xi )-\frac{a(\rho _{j}\xi )}{\rho _{j}^p}= \frac{f_{j}^{n}(\rho _{j}\xi )f_{j}^{(k)}(\rho _{j}\xi )-a(\rho _{j}\xi )}{\rho _{j}^p}\rightrightarrows h^{n}(\xi )h^{(k)}(\xi )-b_p\xi ^p. \end{aligned}$$
(3.3)

Noting \(p\le m\), it follows from Lemmas 2.3 and 2.5 that \(h^{n}(\xi )h^{(k)}(\xi )-b_{p}\xi ^{p}\) has two distinct zeros at least. Additionally, with similar discussion to the proof of Case 1, we can conclude that \(h^{n}(\xi )h^{(k)}(\xi )-b_p\xi ^p \not \equiv 0\). Let \(\xi _{1}\) and \(\xi _{1}^{*}\) be two distinct zeros of \(h^{n}(\xi )h^{(k)}(\xi )-b_p\xi ^p\). We choose a positive number \(\gamma \) properly, such that \(D_3\cap D_4=\emptyset \) and such that \(h^{n}(\xi )h^{(k)}(\xi )-b_p\xi ^p\) has no other zeros in \(D_3 \cup D_4\) except for \(\xi _{1}\) and \(\xi _{1}^{*}\), where

$$\begin{aligned} D_3=\{\xi \in \mathbb {C}\ |\ |\xi -\xi _1|<\gamma \},\quad D_4=\{\xi \in \mathbb {C}\ |\ |\xi -\xi _1^*|<\gamma \}. \end{aligned}$$

By Eq. (3.3) and Hurwitz’s theorem, there exist points \(\zeta _j\in D_3, \zeta _j^*\in D_4\) such that

$$\begin{aligned} f^{n}_j(\rho _j\zeta _j)f^{(k)}_j(\rho _j\zeta _j)-a(\rho _j\zeta _j)=0, \end{aligned}$$

and

$$\begin{aligned} f^{n}_j(\rho _j\zeta _j^*)f^{(k)}_j(\rho _j\zeta _j^*)-a(\rho _j\zeta _j^{*})=0 \end{aligned}$$

for sufficiently large \(j\). By the similar arguments in Case 1, we obtain a contradiction.

Subcase 2.2 There exists a subsequence of \(z_{j} / \rho _{j}\), for simplicity we still denote it as \(z_{j} / \rho _{j}\), such that \(z_{j} / \rho _{j}\rightarrow \infty \) as \(j\rightarrow \infty \). Then

$$\begin{aligned}&f_{j}^{(k)}(z_{j}+\rho _{j}\xi ) =\left\{ (z_{j}+\rho _{j}\xi )^{d}h_{j}(z_{j}+\rho _{j}\xi )\right\} ^{(k)}\\&\quad =(z_{j}+\rho _{j}\xi )^{d}h_{j}^{(k)}(z_{j}+\rho _{j}\xi ) +\sum _{i=1}^{k}a_{i}(z_{j}+\rho _{j}\xi )^{{d}-i}h_{j}^{(k-i)}(z_{j}+\rho _{j}\xi )\\&\quad =(z_{j}+\rho _{j}\xi )^{d}\rho _{j}^{-\frac{nk}{n+1}}g_{j}^{(k)}(\xi ) +\sum _{i=1}^{k}a_{i}(z_{j}+\rho _{j}\xi )^{{d}-i}\rho _{j}^{-\frac{nk}{n+1}+i} g_{j}^{(k-i)}(\xi ), \end{aligned}$$

in which \(a_{i} (i=1,2,\cdots ,k)\) are all constants. Since \(z_{j} / \rho _{j}\rightarrow \infty , b(z_{j}+\rho _{j}\xi )\rightarrow b_p\) as \(j\rightarrow \infty \), it follows that

$$\begin{aligned}&b_p\frac{f_{j}^{n}(z_{j}+\rho _{j}\xi )f_{j}^{(k)}(z_{j} +\rho _{j}\xi )}{a(z_{j}+\rho _{j}\xi )}-b_p =b_p\frac{(z_{j}+\rho _{j}\xi )^{p}g_{j}^{n}(\xi )g_{j}^{(k)}(\xi )}{b(z_{j}+\rho _{j}\xi )(z_{j}+\rho _{j}\xi )^p}\nonumber \\&\quad +\sum _{i=1}^{k}b_p\frac{(z_{j}+\rho _{j}\xi )^pg_{j}^{n} (\xi )g_{j}^{(k-i)}(\xi )}{b(z_{j}+\rho _{j}\xi )(z_{j} +\rho _{j}\xi )^p}\left( \frac{\rho _{j}}{z_{j}+\rho _{j}\xi }\right) ^{i}\nonumber \\&\quad -b_p \rightrightarrows g^{n}(\xi )g^{(k)}(\xi )-b_p \end{aligned}$$
(3.4)

on every compact subset of \(\mathbb {C}\) which contains no poles of \(g\). Since all zeros of \(f_{j}\in {{\fancyscript{F}}}\) have at least multiplicity \(k+m\), then multiplicities of zeros of \(g\) are at least \(k\). Then from Lemmas 2.3 and 2.5, the function \(g^{n}(\xi )g^{(k)}(\xi )-b_p\) has at least two distinct zeros. With similar discussion to the proof of Case 1, we can get a contradiction.

Hence the claim is proved, that is, \(\fancyscript{H}\) is normal at \(z_{0}=0\). Therefore, for any sequence \(\{f_t\}\subset {{\fancyscript{F}}}\) there exist \(\Delta _{r}=\{z: |z|< r\}\) and a subsequence \(\{h_{t_{k}}\}\) of \(\{h_t(z)=f_{t(z)}/z^{d}\}\subset \fancyscript{H}\) such that \(h_{t_{k}} \rightrightarrows I\) or \(\infty \) in \(\Delta _{r}\), where \(I\) is a meromorphic function. Next, we distinguish two cases.

Case A Assume \(f_{t_{k}}(0)\ne 0\) when \(k\) is sufficiently large. Then \(I(0)=\infty \), and hence for arbitrary \(R>0\), there exists a positive number \(\delta \) with \(0<\delta <r\) such that \(|I(z)|> R\) when \(z\in \Delta _{\delta }\). Hence when \(k\) is sufficiently large, we have \(|h_{t_{k}}(z)|> R/2\), which means that \(1/ f_{t_{k}}\) is holomorphic in \(\Delta _{\delta }\). In fact, when \(|z|=\delta /2\),

$$\begin{aligned} \left| \frac{1}{f_{t_{k}}(z)}\right| =\left| \frac{1}{h_{t_{k}}(z)z^{d}}\right| \le M=\frac{2^{d+1}}{R\delta ^{d}}. \end{aligned}$$

By applying maximum principle, we have

$$\begin{aligned} \left| \frac{1}{f_{t_{k}}(z)}\right| \le M \end{aligned}$$

for \(z\in \Delta _{\delta /2}\). It follows from Motel’s normal criterion that there exists a convergent subsequence of \(\{f_{t_k}\}\), that is, \({{\fancyscript{F}}}\) is normal at \(0\).

Case B There exists a subsequence of \(f_{t_{k}}\), for simplicity we still denote it as \(f_{t_{k}}\), such that \(f_{t_{k}}(0)=0\). Then we get \(I(0)=0\) since \(h_{t_k}(z)=f_{t_{k}}(z)/ z^{d}\rightrightarrows I(z)\), and hence there exists a positive number \(\rho \) with \(0<\rho <r\) such that \(I(z)\) is holomorphic in \(\Delta _{\rho }\) and has a unique zero \(z=0\) in \(\Delta _{\rho }\). Therefore, we have \(f_{t_{k}}(z)\rightrightarrows z^{d}I(z)\) in \(\Delta _{\rho }\) since \(h_{t_{k}}\) converges spherically locally uniformly to a holomorphic function \(I\) in \(\Delta _{\rho }\). Thus \({{\fancyscript{F}}}\) is normal at \(0\).

Similarly, we can prove that \({{\fancyscript{F}}}\) is normal at arbitrary \(z_{0}\in D\), and hence \({{\fancyscript{F}}}\) is normal in \(D\).

4 Proof of Corollary 1.1

Using Lemmas 2.3 and 2.5, we find that if \(f\) is a non-constant meromorphic function which has only zeros of multiplicity at least \(k\), then \(f^nf^{(k)}-a\) has at least two distinct zeros for a non-zero complex number \(a\). Therefore, noting that \(a\) has no zeroes, we can verify that \({{\fancyscript{F}}}\) is normal in \(D\) by utilizing the same method in the proof of Theorem 1.1.

5 Proof of Theorem 1.2

Without loss of generality, we assume that \(D=\{z\in \mathbb {C}\ |\ |z|<1\}\) and \(z_{0}=0\). Now we distinguish two cases by either \(a(0)=0\) or \(a(0)\ne 0\).

Case 1 \(a(0) \ne 0\). To the contrary, we suppose that \({{\fancyscript{F}}}\) is not normal at \(0\). Using the notations in the proof of Theorem 1.1, we also obtain

$$\begin{aligned}&f_j^{n}(z_j+\rho _j\xi ) f_j'(z_j+\rho _j\xi ) -a(z_j+\rho _j\xi ) \nonumber \\&\quad = g_j^{n}(\xi )g_j'(\xi )-a(z_j+\rho _j\xi )\rightrightarrows g^{n}(\xi ) g'(\xi ) -a(0), \end{aligned}$$
(5.1)

where \(g^{n}g^{(k)}\not \equiv a(0)\).

By Lemmas 2.4 and 2.5, the function \(g^{n}g'-a(0)\) has a zero \(\xi _{2}\). By Eq. (5.1) and Hurwitz’s theorem, there exist points \(\eta _j \rightarrow \xi _{2}\) \((j\rightarrow \infty )\) such that for sufficiently large \(j, z_j+\rho _j\eta _j \in D\) and

$$\begin{aligned} f^{n}_j(z_j+\rho _j\eta _j)f_j'(z_j+\rho _j\eta _j)-a(z_j+\rho _j\eta _j)=0, \end{aligned}$$

which contradicts the assumption that \(f^{n}f'\ne a\).

Case 2 \(a(0)=0\). Using the notations in the proof of Theorem 1.1, we also get the formulas Eqs. (3.1)–(3.4). Therefore, with the similar method in Case 1, we can prove that \({{\fancyscript{F}}}\) is normal at \(z_{0}\), and hence \({{\fancyscript{F}}}\) is normal in \(D\).