1 Introduction

In this paper, we assume the reader is familiar with Nevanlinna theory of meromorphic functions. Let D be a domain in \(\mathbb {C}\) and let \(\mathcal {F}\) be a family of meromorphic functions in D. We say that \(\mathcal {F}\) is normal in D (in the sense of Montel) if each sequence \(\{f_n\}\) in \(\mathcal {F}\) has a subsequence \(\{f_{n_j}\}\) that converges locally uniformly in D, with respect to the spherical metric, to a meromorphic function or \(\infty \) (see [7, 15, 17]).

For simplicity, we take \(\rightarrow \) to stand for convergence and \(\Rightarrow \) for convergence spherically locally uniformly.

Let f(z) and g(z) be two meromorphic functions in a domain D, and let h(z) be a holomorphic function in D. If \(f(z)-h(z)\) and \(g(z)-h(z)\) have the same zeros ignoring multiplicity (counting multiplicity), then we say that f(z) and g(z) share h(z) IM (CM) in D.

The following normality criterion was conjectured by Hayman [8] and proved by several authors (see [1, 4, 6, 10, 16]).

Theorem 1

Let n be a positive integer, and let \(\mathcal {F}\) be a family of meromorphic functions in D. If, for each \(f\in \mathcal {F}\), \(f^nf' \ne 1\), then \(\mathcal {F}\) is normal in D.

For other related results, see Bergweiler and Langley [2], Pang and Zalcman [11], Wu and Xu [14] and Tan et al. [13].

In 2008, Zhang [18] considered the case of shared value and obtained.

Theorem 2

Let \(\mathcal {F}\) be a family of meromorphic functions in D, and let \(n({\ge }2)\) be a positive integer. If, for any two functions \(f, g\in \mathcal {F}\), \(f^nf'\) and \(g^ng'\) share a nonzero value a IM in D, then \(\mathcal {F}\) is normal in D.

In 2015, Meng and Hu [9] studied the case of \(f^nf^{(k)}(n\ge 2)\) sharing a holomorphic function and obtained

Theorem 3

Let \(k({\ge }1),n({\ge }2), m({\ge }0)\) be three integers, let \(h(z)({\not \equiv }0)\) be a holomorphic function in a domain D with all zeros that have multiplicity at most m and divisible by \(n + 1\), and let \(\mathcal {F}\) be a family of meromorphic functions in domain D such that each \(f \in \mathcal {F}\) has zeros of multiplicity at least \(k + m\) and poles of multiplicity at least \(m + 1\). If, for any two functions \(f, g\in \mathcal {F}\), \(f^n(z)f^{(k)}(z)\) and \(g^n(z)g^{(k)}(z)\) share h(z) IM in D, then \(\mathcal {F}\) is normal in D.

By Theorems 2 and 3, it is nature to ask that: can we get rid of the condition “all zeros of h(z) have multiplicity divisible by \(n+1\)” and “all poles of f have multiplicity at least \(m+1\) in Theorem 3”?

In this paper, we studied the question and gave an affirmative answer to the question.

Theorem 4

Let \(k({\ge }1),n({\ge }2), m({\ge }0)\) be three integers, let \(h(z)({\not \equiv }0)\) be a holomorphic function in a domain D with all zeros that have multiplicity at most m, and let \(\mathcal {F}\) be a family of meromorphic functions in domain D such that each \(f \in \mathcal {F}\) has zeros of multiplicity at least \(k + m\). If, for any two functions \(f, g\in \mathcal {F}\), \(f^n(z)f^{(k)}(z)\) and \(g^n(z)g^{(k)}(z)\) share h(z) IM in D, then \(\mathcal {F}\) is normal in D.

In fact, we proved the following more general result:

Theorem 5

Let \(k({\ge }1),n({\ge }2), m({\ge }0)\) be three integers, let \(h(z)({\not \equiv }0)\) be a holomorphic function in a domain D with all zeros that have multiplicity at most m, and let \(\mathcal {F}\) be a family of meromorphic functions in a domain D such that each \(f \in \mathcal {F}\) has zeros of multiplicity at least \(k + m\). If, for any two functions \(f, g\in \mathcal {F}\), \(f^n(z)f^{(k)}(z)-h(z)\) has at most one distinct zero in D, then \(\mathcal {F}\) is normal in D.

The following examples show that the conditions in Theorem 5 are necessary.

Example 1

[9] Let \(D =\{z\in \mathbb {C}|\ |z|< 1\}\), let \(h(z)\equiv 0\) and let

$$\begin{aligned} \mathcal {F}=\left\{ f_j(z)=e^{jz} \mid j=1,2,\ldots \right\} . \end{aligned}$$

Obviously, \(f^n_j(z)f_j^{(k)}(z)-h(z)\) does not have zero in D for each positive integer j. But the family \(\mathcal {F}\) is not normal at \(z =0\). This shows that \(h(z)\not \equiv 0\) is necessary Theorem 5.

Example 2

Let \(D =\{z\in \mathbb {C}|\ |z|< 1\}\), let \(h(z)=\frac{1}{z^{n+k+1}}\) and let

$$\begin{aligned} \mathcal {F}=\left\{ f_j(z)=\frac{1}{jz} \mid j=1,2,\ldots , \text {and} \ j^{n+1}\ne (-1)^k k!\right\} . \end{aligned}$$

Obviously, \(f^n_j(z)f_j^{(k)}(z)-h(z)\) does not have zero in D for each positive integer j. But the family \(\mathcal {F}\) is not normal at \(z =0\). This shows that Theorem 5 is not valid if h(z) is a meromorphic function in D.

Example 3

Let \(D =\{z\in \mathbb {C}|\ |z|< 1\}\), let \(h(z)=1\) and let

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

Then \(f^n_j(z)f_j^{(k)}(z)-h(z)\) does not have zero in D for each positive integer j. But the family \(\mathcal {F}\) is not normal at \(z =0\). This shows that the condition “all zeros of f have multiplicity at least \(k + m\) ” in Theorem 5 is best.

Example 4

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

$$\begin{aligned} \mathcal {F}=\left\{ f_j(z)=jz \mid j=1,2,\ldots \right\} . \end{aligned}$$

Obviously, \(f^2_j(z)f'_j(z)-h(z)=j^3z^2-1\) have exactly two distinct zeros in D for each positive integer j. But the family \(\mathcal {F}\) is not normal at \(z =0\). This shows that the condition “\(f^n(z)f^{(k)}(z)-h(z)\) has at most one distinct zero” in Theorem 5 is necessary.

2 Some Lemmas

For the proofs of our theorems, we require the following results.

Lemma 1

[12, 17] Let \(\mathcal {F}\) be a family of meromorphic functions in the unit disk \(\varDelta \) such that all zeros of functions in \(\mathcal {F}\) have multiplicity \(\ge l\). Let \(\alpha \) be a real number satisfying \(-l<\alpha <1\). Then \(\mathcal {F}\) is not normal in any neighborhood of \(z_0\in \varDelta \) if and only if there exist

  1. (a)

    points \(z_j\in \varDelta ,\)\(z_j \rightarrow z_0;\)

  2. (b)

    functions \(f_j\in {\mathcal {F}};\) and

  3. (c)

    positive numbers \(\rho _j\rightarrow 0\)

such that \(g_j(\xi )=\rho _{j} ^{\alpha }f_j(z_j+\rho _j\xi )\Rightarrow g(\xi )\) spherically uniformly on compact subsets of \(\mathbb {C} \), where \(g(\xi )\) is a non-constant meromorphic function in \(\mathbb {C}\) satisfying that all zeros of g have multiplicity at least l.

Lemma 2

[15] Let \(f_1\) and \(f_2\) be two non-constant meromorphic functions in \(\mathbb {C}\), then

$$\begin{aligned} N\left( r,f_1f_2\right) -N\left( r,\frac{1}{f_1f_2}\right) =N\left( r,f_1\right) +N\left( r,f_2\right) -N\left( r,\frac{1}{f_1}\right) -N\left( r,\frac{1}{f_2}\right) . \end{aligned}$$

The following Lemma was proved by Zhang and Li [19] when f is a transcendental meromorphic function, and by Meng and Hu [9] when f is a rational function.

Lemma 3

Let \(n({\ge }2), k({\ge }1)\) be three integers, let \(a\ne 0\) be a finite complex number, and let f(z) be a non-constant meromorphic in \(\mathbb {C}\) with all zeros that have multiplicity at least k. Then \(f^n(z)f^{(k)}(z)-a\) have at least two distinct zeros.

Lemma 4

Let \(n({\ge }1), k({\ge }1), M({\ge }1)\) be three integers, let p(z) be a polynomial with \(\deg p=M\), and let f(z) be a non-constant rational function in \(\mathbb {C}\) with \(f(z)\ne 0\). Then \(f^n(z)f^{(k)}(z)-p(z)\) has at least \(n+k+1\) distinct zeros.

The proof of Lemma 4 is almost the same with Chang [3] and Lemma 11 in Deng etc. [5], we omit the detail.

Lemma 5

Let \(n({\ge }2), k({\ge }1), m({\ge }1)\) be three integers, let p(z) be a polynomial with \(\deg p=m\), and let f(z) be a non-constant meromorphic in \(\mathbb {C}\) with all zeros that have multiplicity at least \(k+m\). Then \(f^n(z)f^{(k)}(z)-p(z)\) has at least two distinct zeros.

Proof

Set

$$\begin{aligned} \frac{1}{f^{n+1}}=\frac{f^nf^{(k)}}{pf^{n+1}}-\frac{p[f^nf^{(k)}]'-p'f^nf^{(k)}}{pf^{n+1}}\frac{f^nf^{(k)}-p}{p[f^nf^{(k)}]'-p'f^nf^{(k)}}. \end{aligned}$$

Then by \(m(r,\frac{f^{(i)}}{f})=S(r,f)(i\ge 1)\), \(m(r,p)=m\log r+O(1)\), \(m\big (r,\frac{1}{p}\big )=O(1)\), Lemma 2 and Nevanlinna’s elementary theory, we get

$$\begin{aligned} \begin{aligned} \left( n+1\right) m\left( r,\frac{1}{f}\right)&\le m\left( r,\frac{f^nf^{\left( k\right) }}{pf^{n+1}}\right) +m\left( r,\frac{p\left( f^nf^{\left( k\right) }\right) '-p'f^nf^{\left( k\right) }}{pf^{n+1}}\right) \\&\quad +\,m\left( r,\frac{f^nf^{\left( k\right) }-p}{p\left[ f^nf^{\left( k\right) }\right] '-p'f^nf^{\left( k\right) }}\right) +S\left( r,f\right) \\&\le T\left( r,\frac{f^nf^{\left( k\right) }-p}{p\left[ f^nf^{\left( k\right) }\right] '-p'f^nf^{\left( k\right) }}\right) \\&\quad -\,N\left( r,\frac{f^nf^{\left( k\right) }-p}{p\left[ f^nf^{\left( k\right) }\right] '-p'f^nf^{\left( k\right) }}\right) +S\left( r,f\right) \\&=m\left( r,\frac{p\left[ f^nf^{\left( k\right) }\right] '-p'f^nf^{\left( k\right) }}{f^nf^{\left( k\right) }-p}\right) \\&\quad +\,N\left( r,\frac{p\left[ f^nf^{\left( k\right) }\right] '-p'f^nf^{\left( k\right) }}{f^nf^{\left( k\right) }-p}\right) \\&\quad -\,N\left( r,\frac{f^nf^{\left( k\right) }-p}{p\left[ f^nf^{\left( k\right) }\right] '-p'f^nf^{\left( k\right) }}\right) +S\left( r,f\right) \\&=m\left( r,\frac{p\left[ \frac{f^nf^{\left( k\right) }}{p}-1\right] '}{\frac{f^nf^{\left( k\right) }}{p}-1}\right) \\&\quad +\,N\left( r,p\left[ f^nf^{\left( k\right) }\right] '-p'f^nf^{\left( k\right) }\right) +N\left( r,\frac{1}{f^nf^{\left( k\right) }-p}\right) \\&\quad -\,N\left( r,\frac{1}{p\left[ f^nf^{\left( k\right) }\right] '-p'f^nf^{\left( k\right) }}\right) \\&\quad -\,N\left( r,f^nf^{\left( k\right) }-p\right) +S\left( r,f\right) \\&\le \overline{N}\left( r,f\right) +N\left( r,\frac{1}{f^nf^{\left( k\right) }-p}\right) \\&\quad -\,N\left( r,\frac{1}{p\left[ f^nf^{\left( k\right) }\right] '-p'f^nf^{\left( k\right) }}\right) +m\log r+S\left( r,f\right) . \end{aligned} \end{aligned}$$

Let \(z_1\) is a zero of f with multiplicity \(l_1\ge k+m\). Then \(z_1\) is a zero of \(p[f^nf^{(k)}]'-p'f^nf^{(k)}\) with multiplicity at least \((n+1)l_1-k-1\).

Let \(z_2\) is a zero of \(f^nf^{(k)}-p\) with multiplicity \(l_2\). Obviously, we have

$$\begin{aligned} p[f^nf^{(k)}]'-p'f^nf^{(k)}=p[f^nf^{(k)}-p]'-p'[f^nf^{(k)}-p]. \end{aligned}$$

Then \(z_2\) is a zero of \(p[f^nf^{(k)}]'-p'f^nf^{(k)}\) with multiplicity at least \(l_2-1\).

Hence, we have

$$\begin{aligned} \left( n+1\right) T\left( r,f\right)\le & {} \overline{N}\left( r,f\right) + \left( n+1\right) N\left( r,\frac{1}{f}\right) +N\left( r,\frac{1}{f^nf^{\left( k\right) }-p}\right) \nonumber \\&\quad +\,m\log r-N\left( r,\frac{1}{p\left[ f^nf^{\left( k\right) }\right] '-p'f^nf^{\left( k\right) }}\right) +S\left( r,f\right) \nonumber \\\le & {} \overline{N}\left( r,f\right) + \left( k+1\right) \overline{N}\left( r,\frac{1}{f}\right) +\overline{N}\left( r,\frac{1}{f^nf^{\left( k\right) }-p}\right) \nonumber \\&\quad +\,m\log r+S\left( r,f\right) \nonumber \\\le & {} \overline{N}\left( r,f\right) + \frac{k+1}{k+m}N\left( r,\frac{1}{f}\right) +\overline{N}\left( r,\frac{1}{f^nf^{\left( k\right) }-p}\right) \nonumber \\&\quad +\,m\log r+S\left( r,f\right) . \end{aligned}$$
(2.1)

Suppose that \(f^n(z)f^{(k)}(z)-p(z)\) has at most one distinct zero.

Next we consider two cases.

Case 1\(m\ge 2\). Then by (2.1), we have

$$\begin{aligned} T(r,f)<\left( n-\frac{k+1}{k+m}\right) T(r,f)\le (m+1)\log r+S(r,f). \end{aligned}$$

Thus, f is a rational function with \(\deg f<m+1\). Since all zeros of f have multiplicity at least \(k+m\ge 1+m\), we deduce that \(f(z)\ne 0\). Then by Lemma 4, we obtain that \(f^n(z)f^{(k)}(z)-p(z)\) has at least \(n+k+1>2\) distinct zeros, a contradiction.

Case 2\(m=1\).

If \(f^n(z)f^{(k)}(z)-p(z)\ne 0\), then by (2.1), we get \(T(r,f)\le \log r+S(r,f)\). It follows that f is a rational function with \(\deg f\le 1\). We deduce that \(f(z)\ne 0\), since all zeros of f have multiplicity at least \(k+m\ge 2\), Then by Lemma 4, we get \(f^n(z)f^{(k)}(z)-p(z)\) has at least \(n+k+1>2\) distinct zeros, a contradiction.

Thus \(f^n(z)f^{(k)}(z)-p(z)\) has exactly one distinct zero. By (2.1), we have

$$\begin{aligned} nT(r,f)\le 2\log r+\overline{N}(r,f)+S(r,f). \end{aligned}$$
(2.2)

If \(n\ge 3\), by (2.2), we obtain \(T(r,f)\le \log r+S(r,f)\). It follows that f is a rational function with \(\deg f \le 1\). Since all zeros of f have multiplicity at least \(k+m\ge 2\), we obtain \(f(z)\ne 0\), then by Lemma 4, we get \(f^n(z)f^{(k)}(z)-p(z)\) has at least \(n+k+1>2\) distinct zeros, a contradiction.

Thus \(n=2\). By (2.2) again, we get \(T(r,f)\le 2\log r+S(r,f)\). It follows that f is a rational function with \(\deg f \le 2\). If \(k\ge 2\), since all zeros of f have multiplicity at least \(k+m\ge 3\), we get \(f(z)\ne 0\), then by Lemma 4, we get a contradiction. Hence \(k=1\), then f has one zero with multiplicity 2 at most. If f has no zero, then by Lemma 4, a contradiction. Thus, f(z) has exactly one distinct zero with multiplicity 2, and of the following forms:

$$\begin{aligned} A_1{:}\,f(z)= & {} a(z-\alpha )^2;\quad A_2{:}\,f(z)=a\frac{(z-\alpha )^2}{z-\beta };\\ A_3{:}\,f(z)= & {} \frac{a(z-\alpha )^2}{(z-\beta _1)(z-\beta _2)};\quad A_4{:}\, f(z)=a\frac{(z-\alpha )^2}{(z-\beta )^2}. \end{aligned}$$

If f(z) has the form \(A_1\) or \(A_2\) or \(A_4\), we have \(\overline{N}(r,f) \le \log r =1/2T(r,f)+O(1)\). Then by (2.2), we get \(T(r,f)\le 4/3\log r+S(r,f)\), this contradicts with \(T(r,f)=2\log r +O(1)\).

Then

$$\begin{aligned} f(z)=\frac{a(z-\alpha )^2}{(z-\beta _1)(z-\beta _2)}. \end{aligned}$$
(2.3)

It follows from (2.3) that

$$\begin{aligned} f'\left( z\right) =\frac{a\left( z-\alpha \right) \left[ \left( 2\alpha -\beta _1-\beta _2\right) z+2\beta _1\beta _2-\alpha \left( \beta _1+\beta _2\right) \right] }{\left( z-\beta _1\right) ^2\left( z-\beta _2\right) ^2}. \end{aligned}$$
(2.4)

By (2.3) and (2.4), we get

$$\begin{aligned} f^2(z)f'(z)=\frac{a^3(z-\alpha )^5\left[ \left( 2\alpha -\beta _1-\beta _2\right) z+2\beta _1\beta _2-\alpha (\beta _1+\beta _2)\right] }{(z-\beta _1)^4(z-\beta _2)^4}. \end{aligned}$$
(2.5)

Since \(\deg p=m=1\), we may set \(p(z)=b(z-z_0)\), where \(b\ne 0\) is a constant. Since \(f^n(z)f^{(k)}(z)-p(z)\) has exactly one distinct zero, by (2.5), we may set

$$\begin{aligned} f^2(z)f'(z)=b(z-z_0)-\frac{b(z-w)^9}{(z-\beta _1)^4(z-\beta _2)^4}, \end{aligned}$$
(2.6)

where \(w\ne \alpha \). Otherwise, if \(w= \alpha \), then by (2.5), we get \(\alpha \) is a zero of \((f^2(z)f'(z))''\) with multiplicity 3. But from (2.6), we get \(\alpha \) is a zero of \((f^2(z)f'(z))''\) with multiplicity 7, a contradiction.

Differentiating (2.5) two times, we obtain,

$$\begin{aligned}{}[f^2(z)f'(z)]''=\frac{(z-\alpha )^3g_1(z)}{(z-\beta _1)^6(z-\beta _2)^6}, \end{aligned}$$
(2.7)

where \(g_1(z)\) is a polynomial with \(\deg g_1\le 5\).

On the other hand, differentiating (2.6) two times, we obtain,

$$\begin{aligned}{}[f^2(z)f'(z)]''=\frac{(z-w)^7g_2(z)}{(z-\beta _1)^6(z-\beta _2)^6}, \end{aligned}$$
(2.8)

where \(g_2(z)\) is a polynomial with \(\deg g_2\le 4\).

From (2.7)–(2.8), and \(w\ne \alpha \), we get \(7\le \deg g_1\le 5\), a contradiction.

This completes the proof of Lemma 5.

Lemma 6

Let \(n({\ge }2), k({\ge }1)\) be three integers, and let \(\{f_j\}\) be a sequence of meromorphic functions in domain D, \(\{h_j(z)\}\) be a sequence of holomorphic functions in D such that \(h_j(z)\Rightarrow h(z)\), where \(h(z)\not = 0\) be a holomorphic function. If, for each \(j\in N^+\), all zeros of function \(f_j(z)\) have multiplicity at least k, and \(f^n_j(z)f_j^{(k)}(z)-h_j(z)\) has at most one distinct zero in D, then \(\{f_j\}\) is normal in D.

Proof

Suppose that \(\{f_j\}\) is not normal at \(z_0\in D\). By Lemma 1, there exists a sequence \(z_j\) of complex numbers \(z_j\rightarrow z_0\), a sequence \(\rho _j\) of positive numbers \(\rho _j\rightarrow 0\) , and a subsequence of \(\{f_j\}\) (we may still denote by \(\{f_j\}\)) such that

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

locally uniformly on compact subsets of \(\mathbb {C}\), where \(g(\xi )\) is a non-constant meromorphic function in \(\mathbb {C}\). By Hurwitz’s theorem, all zeros of \(g(\xi )\)have multiplicity at least k. Then, we have

$$\begin{aligned} g^n_j(\xi )g_j^{(k)}(\xi )-h_j(z_j+\rho _j\xi )= & {} f^n_j(z_j+\rho _j\xi )f_j^{(k)}(z_j+\rho _j\xi )-h_j(z_j+\rho _j\xi )\\\Rightarrow & {} g^n(\xi )g^{(k)}(\xi )-h(z_0). \end{aligned}$$

for all \(\xi \in \mathbb {C}/\{g^{-1}(\infty )\}\).

Obviously, \(g^n(\xi )g^{(k)}(\xi )-h(z_0)\not \equiv 0\). Otherwise, suppose that

$$\begin{aligned} g^n(\xi )g^{(k)}(\xi )-h(z_0) \equiv 0, \end{aligned}$$
(2.9)

then we have \(g(\xi )\ne 0\) since \(h(z_0)\ne 0\).

It follows from (2.9) that

$$\begin{aligned} \frac{1}{g^{n+1}(\xi )}\equiv \frac{g^{(k)}(\xi )}{h(z_0)g(\xi )}. \end{aligned}$$

Then, we get

$$\begin{aligned} (n+1)m\left( r,\frac{1}{g}\right) =m\left( r, \frac{g^{(k)}}{h(z_0)g}\right) =S(r,g). \end{aligned}$$

It follows that \(T(r,g)=S(r,g)\) since \(g\ne 0\). Hence g is a constant, a contradiction.

We claim that \(g^n(\xi )g^{(k)}(\xi )-h(z_0)\) has at most one distinct zero. Otherwise, suppose that \(\xi _1, \xi _2\) are two distinct zeros of \(g^n(\xi )g^{(k)}(\xi )-h(z_0)\). We choose a positive number \(\sigma \) small enough such that \(D_1\cap D_2=\emptyset \) and \(g^n(\xi )g^{(k)}(\xi )-h(z_0)\) has no other zeros in \(D_1\bigcup D_2\) except for \(\xi _1\) and \(\xi _2\), where \(D_1=\{\xi {:}\,\mid \xi -\xi _1\mid < \sigma \}\) and \(D_2=\{\xi {:}\mid \xi -\xi _2\mid < \sigma \}\).

By Hurwitz’s theorem, for sufficiently large j there exist points \(\xi _{1,j}\rightarrow \xi _1\) and \(\xi _{2,j}\rightarrow \xi _2\) such that

$$\begin{aligned}&f^n_j(z_j+\rho _j\xi _{1,j})f_j^{(k)}(z_j+\rho _j\xi _{1,j})-h_j(z_j+\rho _j\xi _{1,j})=0;\\&f^n_j(z_j+\rho _j\xi _{2,j})f_j^{(k)}(z_j+\rho _j\xi _{2,j})-h_j(z_j+\rho _j\xi _{2,j})=0. \end{aligned}$$

By the assumption in Lemma 6, \(f^n_jf_j^{(k)}(z)-h_j(z)\) has at most one zero in D, it follows that \(z_j+\rho _j\xi _{1,j}=z_j+\rho _j\xi _{2,j}\), that is \(\xi _{1,j}=\xi _{2,j}=(z_0-z_j)/\rho _j\), which contradicts with the facts \(D_1\cap D_2=\emptyset \).

The claim is proved. On the other hand, it follows from Lemma 3 that \(g^n(\xi )g^{(k)}(\xi )-h(z_0)\) has at least two distinct zeros, a contradiction. Thus \(\{f_j\}\) is normal in D.

3 Proof of Theorems

Proof of Theorem 5

By Lemma 6, it is enough to prove that \(\mathcal {F}\) is normal at the point \(z_0\), where \(h(z_0)=0\). By making standard normalization, we may assume that \(z_0=0\), and \(h(z)=z^tb(z)\) where \(1\le t\le m\) is a positive integer, and \(b(0)= 1\).

Suppose that \(\mathcal {F}\) is not normal at \(z_0=0\). By Lemma 1, there exists a sequence \(z_j\) of complex numbers \(z_j\rightarrow 0\), a sequence \(\rho _j\) of positive numbers \(\rho _j\rightarrow 0\), and a sequence of functions \(\{f_j\}\subseteq \mathcal {F}\) such that

$$\begin{aligned} g_j(\xi )=\frac{f_j(z_j+\rho _j\xi )}{\rho _j^{\frac{k+t}{n+1}}}{\Rightarrow }g(\xi ) \end{aligned}$$
(3.1)

locally uniformly on compact subsets of \(\mathbb {C}\), where \(g(\xi )\) is a non-constant meromorphic function in \(\mathbb {C}\). By Hurwitz’s theorem, all zeros of \(g(\xi )\) have multiplicity at least \(k+m\). Next we consider two cases.

Case 1\(z_j/ \rho _j \rightarrow \infty \). Set

$$\begin{aligned} F_j(\xi )=\frac{f_j(z_j+z_j\xi )}{z_j^{\frac{k+t}{n+1}}}. \end{aligned}$$

Then, we have

$$\begin{aligned}&F^n_j(\xi )F_j^{(k)}(\xi )-(1+\xi )^tb(z_j+z_j\xi )\\&\quad =\frac{f^n_j(z_j+z_j\xi )f_j^{(k)}(z_j+z_j\xi )-h(z_j+z_j\xi )}{z_j^t}. \end{aligned}$$

As the same argument as in Lemma 6, we deduce that \(F_j^n(\xi )F_j^{(k)}(\xi )-(1+\xi )^tb(z_j+z_j\xi )\) has at most one distinct zero in \(\varDelta =\{\xi : \mid \xi \mid <1\}\).

Since all zeros of \(F_j\) have multiplicity at least \(k+m\), and \((1+\xi )^tb(z_j+z_j\xi )\rightarrow (1+\xi )^t \ne 0\) when \(\xi \in \varDelta \). Then by Lemma 6, \(\{F_j\}\) is normal in \(\varDelta \).

So, there exists a subsequence of functions [we still denote as \({F_j (\xi )}\)] and a function \(F(\xi )\) (a meromorphic function or \(\infty \)), such that \(F_j(\xi ){\Rightarrow }F(\xi ).\)

If \(F(0)\ne \infty \), then it follows from \(k+m-1-\frac{k+t}{n+1}>0\) that

$$\begin{aligned} g^{\left( k+m-1\right) }\left( \xi \right)= & {} \underset{j\rightarrow \infty }{\lim }g_j^{\left( k+m-1\right) }\left( \xi \right) =\underset{j\rightarrow \infty }{\lim }\frac{f_j^{\left( k+m-1\right) }\left( z_j+\rho _j\xi \right) }{\rho _j^{\frac{k+t}{n+1}-\left( k+m-1\right) }}\\= & {} \underset{j\rightarrow \infty }{\lim }\left( \frac{\rho _j}{z_j}\right) ^{k+m-1-\frac{k+t}{n+1}}F_j^{\left( k+m-1\right) }\left( \frac{\rho _j}{z_j}\xi \right) =0, \end{aligned}$$

for all \(\xi \in \mathbb {C}/\{g^{-1}(\infty )\}\).

Thus we deduce that \(g^{(k+m-1)}\equiv 0\). Hence g is a polynomial of degree at most \(k+m-1\). Since all zeros of g have multiplicity at least \(k+m\), it follows that \(g(\xi )\) is a constant, a contradiction.

If \(F(0)= \infty \), then by

$$\begin{aligned} \frac{1}{F_j\left( \frac{\rho _j}{z_j}\xi \right) }=\frac{z_j^{\frac{k+t}{n+1}}}{f_j\left( z_j+\rho _j\xi \right) }\rightarrow \frac{1}{F(0)}= 0, \end{aligned}$$

when \(\xi \in \mathbb {C}/\{g^{-1}(0)\}\), we obtain that,

$$\begin{aligned} \frac{1}{g\left( \xi \right) }=\underset{j\rightarrow \infty }{\lim }\frac{\rho _j^{\frac{k+t}{n+1}}}{f_j\left( z_j+\rho _j\xi \right) }=\underset{j\rightarrow \infty }{\lim }\left( \frac{\rho _j}{z_j}\right) ^{\frac{k+t}{n+1}}\frac{z_j^{\frac{k+t}{n+1}}}{f_j\left( z_j+\rho _j\xi \right) }=0. \end{aligned}$$

Thus \(g(\xi ) \equiv \infty \), which contradicts that \(g(\xi )\) is a non-constant meromorphic function.

Case 2\(z_j/ \rho _j \rightarrow \alpha \), where \(\alpha \) is a finite complex number. Then by (3.1), we have

$$\begin{aligned}&g^n_j\left( \xi \right) g_j^{\left( k\right) }\left( \xi \right) -\left( \xi +\frac{z_j}{\rho _j}\right) ^tb\left( z_j+\rho _j\xi \right) \\&\quad =\frac{f^n_j\left( z_j+\rho _j\xi \right) f_j^{\left( k\right) }\left( z_j+\rho _j\xi \right) -h\left( z_j+\rho _j\xi \right) }{\rho _j^t} \\&\quad \Rightarrow g^n\left( \xi \right) g^{\left( k\right) }\left( \xi \right) -\left( \xi +\alpha \right) ^t \end{aligned}$$

for all \(\xi \in \mathbb {C}/ \{g^{-1}(\infty )\} \).

Since for sufficiently large j, \(f^n_j(z_j+\rho _j\xi )f_j^{(k)}(z_j+\rho _j\xi )-h(z_j+\rho _j\xi )\) has one distinct zero, it follows from the proof of Lemma 6 that \(g^n(\xi )g^{(k)}(\xi )-(\xi +\alpha )^t\) has at most one distinct zero.

But from Lemma 5, \(g^n(\xi )g^{(k)}(\xi )-(\xi +\alpha )^t\) have at least two distinct zeros. Hence \(g(\xi )\) is a constant, a contradiction.

This completes the proof of Theorem 5.

Proof of Theorem 4

Let \(z_0\in D\). We show that \(\mathcal {F}\) is normal at \(z_0\). Let \(f\in \mathcal {F}\).

We consider two cases.

Case 1\(f^n(z_0)f^{(k)}(z_0)\ne h(z_0)\). Then there exists a disk \(D_\delta (z_0)=\{z{:}\,\mid z- z_0\mid < \delta \}\) such that \(f^n(z)f^{(k)}(z)\ne h(z)\) in \(D_\delta (z_0)\). Since for each pair of functions \((f,g)\in \mathcal {F}\), \(f^n(z)f^{(k)}(z)\) and \(g^n(z)g^{(k)}(z)\) share h(z) in D. Thus, for every \(g\in \mathcal {F}\), \(g^n(z)g^{(k)}\ne h(z)\) in \(D_\delta (z_0)\). By Theorem 5, \(\mathcal {F}\) is normal in \(D_\delta (z_0)\). Hence \(\mathcal {F}\) is normal at \(z_0\).

Case 2\(f^n(z_0)f^{(k)}(z_0)= h(z_0)\). Then there exists a disk \(D_\delta (z_0)=\{z{:}\,\mid z- z_0\mid < \delta \}\) such that \(f^n(z)f^{(k)}(z)\ne h(z)\) in \(D^{0}_\delta (z_0)\). Since for each pair of functions \((f,g)\in \mathcal {F}\), \(f^n(z)f^{(k)}(z)\) and \(g^n(z)g^{(k)}\) share h(z) in D. Thus, for every \(g\in \mathcal {F}\), \(g^n(z)g^{(k)}\ne h(z)\) in \(D^{0}_\delta (z_0)\) and \(g^n(z_0)g^{(k)}(z_0)= h(z_0)\). So, \(g^n(z)g^{(k)}- h(z)\) have only distinct zero in \(D_\delta (z_0)\). By Theorem 5, \(\mathcal {F}\) is normal in \(D_\delta (z_0)\). Hence \(\mathcal {F}\) is normal at \(z_0\).

This completes the proof of Theorem 4.