1 Introduction and Main Results

A family \(\mathcal{F }\) of meromorphic functions defined in a plane domain \(D\subset \mathbb{C }\) is said to be normal in \(D\), if each sequence \(\{f_n\}\subset \mathcal{F }\) contains a subsequence which converges spherically locally uniformly in \(D\) to a meromorphic function or \(\infty \). See [6, 11, 15].

The Gu’s normality criterion [5] which was conjectured by Hayman [6] says that a family \(\mathcal{F }\) of functions meromorphic on \(D\) is normal if \(f\not =0\) and \(f^{(k)}\not =1\) for each \(f\in \mathcal{F }\). The following generalization of Gu’s theorem was proved by Yang [14].

Theorem 1

Let \(\mathcal{F }\) be a family of meromorphic functions on \(D\), \(k\in \mathbb{N }\) and \(h(\not \equiv 0)\) be a holomorphic function on \(D\). If for every \(f\in \mathcal{F }\), \(f\not =0\) and \(f^{(k)}\ne h\) on \(D\), then \(\mathcal{F }\) is normal on \(D\).

In recent years, following Schwick [12], many normality criteria concerning shared values or functions have been proved. We say that two functions \(f\) and \(g\) share a value or a function \(\phi \) if the two equations \(f(z)=\phi (z)\) and \(g(z)=\phi (z)\) have the same solutions (ignoring multiplicity). Here, we want to generalize the following result of Fang and Zalcman [4] by replacing the constant 1 by a function.

Theorem 2

Let \(k\) be a positive integer and let \(\mathcal{F }\) a family of meromorphic functions on \(D\), all of whose zeros have multiplicity at least \(k+2\), such that for each pair of functions \(f\) and \(g\) in \(\mathcal{F }\), \(f\) and \(g\) share the value \(0\), and \(f^{(k)}\) and \(g^{(k)}\) share the value \(1\). Then, the family \(\mathcal{F }\) is normal.

In general, the constant 1 cannot be replaced by a function. For example, the family \(\{f_n\}\), where \(f_n(z)=nz^{k+2}\), is not normal at 0. However, each pair of functions \(f_n\) and \(f_m\) share the value \(0\), and \(f_n^{(k)}\) and \(f_m^{(k)}\) share the function \(z^2\).

So we need some additional conditions. We prove the following generalization of Theorem 2.

Theorem 3

Let \(k\in \mathbb{N }\) and \(h(\not \equiv 0)\) be a function holomorphic on \(D\). Let \(\mathcal{F }\) be a family of meromorphic functions in \(D\), all of whose zeros have multiplicity at least \(k+2\), such that for each pair of functions \(f\) and \(g\) in \(\mathcal{F }\), \(f\) and \(g\) share the value \(0\), and \(f^{(k)}\) and \(g^{(k)}\) share the function \(h\). Suppose additionally that at each common zero of \(f\) and \(h\) for every \(f\in \mathcal{F }\), the multiplicities \(m_f\) for \(f\) and \(m_h\) for \(h\) satisfy \(m_f\ge m_h+k+1\) for \(k>1\) and \(m_f\ge 2m_h+3\) for \(k=1\). Then, \(\mathcal{F }\) is normal in \(D\).

The above example shows that the assumption \(m_f\ge m_h+k+1\) is necessary. The condition \(m_f\ge 2m_h+3\) for \(k=1\) is also necessary and sharp as showed by the following example.

Example 1

Let \(\alpha \) be a positive integer and \(h(z)=z^\alpha \). Let for \(n\in \mathbb{N }\),

$$\begin{aligned} f_n(z)=\frac{z^{2\alpha +2}}{(\alpha +1)[z^{\alpha +1}-1/n]}. \end{aligned}$$

Then,

$$\begin{aligned} f_n^{\prime }(z)-h(z)=\frac{z^{3\alpha +2}-\frac{2}{n}z^{2\alpha +1}}{\left(z^{\alpha +1}-\frac{1}{n}\right)^2} -z^\alpha =-\frac{\frac{1}{n^2}z^{\alpha }}{\left(z^{\alpha +1}-\frac{1}{n}\right)^2}. \end{aligned}$$

Thus, each pair of functions \(f_n\) and \(f_m\) share the value \(0\), and \(f_n^{\prime }\) and \(f_m^{\prime }\) share the function \(h(z)\). However, the family \(\{f_n\}\) is not normal at \(0\).

Let us look at some more aspects. By fixing a function \(f_0\in \mathcal{F }\) and letting \(E=\{z\in D:f_0(z)=0\}\cup \{z\in D:f_0^{(k)}(z)=h(z)\}\), the shared condition of Theorem 3 is equivalent to saying that there exists a fixed set \(E\subset D\), which is independent of \(f\in \mathcal{F }\), such that \(f\not =0\) and \(f^{(k)}\not =h\) on \(D\setminus E\) for every \(f\in \mathcal{F }\). There are two cases for the set \(E\). One which is trivial is that \(E=D\). Then, every \(f\in \mathcal{F }\) satisfies \(f\equiv 0\) or \(f^{(k)}\equiv h\), and hence the normality of \(\mathcal{F }\) can be easily dealt with. The non-trivial case is that the exceptional set \(E\) is (locally) discrete in \(D\). We focus on the non-trivial case and consider here that the exceptional set \(E\) is dependent on \(f\). To state our result, we require the following definition.

Definition 1

The sets \(\{E_\lambda \}_{\lambda \in \Lambda }\) are said to be locally uniformly discrete in \(D\), if for each point \(z_0\in D\), there exists \(\delta >0\) such that every \(E_\lambda \) has at most one point lying in the disk \(\Delta (z_0,\delta )=\{z:\ |z-z_0|<\delta \}\).

For example, the sets \(\{E_n\}_{n\in \mathbb{N }}\), where \(E_n=\{\frac{m-1}{m}+\frac{1}{n}: m\in \mathbb{N }\}\), are locally uniformly discrete in the unit disk \(\Delta (0,1)\), but not in the domains that contain the point 1.

Theorem 4

Let \(k\in \mathbb{N }\) and \(h(\not \equiv 0)\) be a function holomorphic on \(D\). Let \(\mathcal{F }\) be a family of meromorphic functions in \(D\), all of whose zeros have multiplicity at least \(k+2\) or \(k+3\) and \(h\) has simple zeros. Suppose that the sets \(\{E_f\}_{f\in \mathcal{F }}\), where \(E_f=\{z\in D:f(z)=0\}\cup \{z\in D:f^{(k)}(z)=h(z)\}\), are locally uniformly discrete in \(D\). Suppose additionally that at the common zeros of \(f\in \mathcal{F }\) and \(h\), the multiplicities \(m_f\) for \(f\) and \(m_h\) for \(h\) satisfy \(m_f\ge m_h+k+1\) for \(k>1\) and \(m_f\ge 2m_h+3\) for \(k=1\). Then, \(\mathcal{F }\) is normal in \(D\).

The following example shows that it is necessary (and sharp) to assume that the multiplicity of zeros of \(f\in \mathcal{F }\) is at least \(k+3\) when \(h\) has simple zeros.

Example 2

Let \(h(z)=z\) and

$$\begin{aligned} f_n(z)=\frac{(z-1/n)^{k+2}}{(k+1)![z-(k+2)/n]},\ \ n\in \mathbb{N }. \end{aligned}$$

Then, \(E_{f_n}=\{1/n\}\), so that \(\{E_{f_n}\}\) are locally uniformly discrete in \(\mathbb{C }\). In fact, we have

$$\begin{aligned} f_n(z)&= \frac{\left[\left(z-\frac{k+2}{n}\right)+\frac{k+1}{n}\right]^{k+2}}{(k+1)!\left(z-\frac{k+2}{n}\right)} =\frac{1}{(k+1)!}\left[\left(z-\frac{k+2}{n}\right)^{k+1}\right.\\&\left.+\frac{(k+1)(k+2)}{n}\left(z-\frac{k+2}{n}\right)^{k}+P(z)+ \frac{\left(\frac{k+1}{n}\right)^{k+2}}{\left(z-\frac{k+2}{n}\right)}\right], \end{aligned}$$

where \(P\) is a polynomial of degree \(<k\), so that

$$\begin{aligned} f_n^{(k)}(z)=z+\frac{(-1)^k\left(\frac{k+1}{n}\right)^{k+2}}{(k+1)\left(z-\frac{k+2}{n}\right)^{k+1}}\not =h(z). \end{aligned}$$

However, \(\{f_n\}\) is not normal at \(0\), as \(f_n(1/n)=0\) while \(f_n((k+2)/n)=\infty \). Throughout in this paper, we denote by \(\mathbb{C }\) the complex plane, by \(\mathbb{C }^*\) the punctured complex plane \(\mathbb{C }\setminus \{0\}\), by \(\Delta (z_0,r)\) the open disk \(\{z: |z-z_0|<r\}\), and by \(\Delta ^\circ (z_0,r)\) the punctured disk \(\Delta (z_0,r)\setminus \{z_0\}=\{z: 0<|z-z_0|<r\}\), where \(z_0\in \mathbb{C }\) and \(r>0\).

2 Auxiliary Results

To prove our results, we require some preliminary results.

Lemma 5

[3] Let \(\mathcal{F }\) be a family of meromorphic functions in a domain \(D\), all of whose zeros have multiplicity at least \(k\). Then, if \(\mathcal{F }\) is not normal at \(z_0\), there exist, for each \(-1< \alpha < k\), points \(z_n\in D\) with \(z_n\rightarrow z_0\), functions \(f_n\in \mathcal{F }\) and positive numbers \(\rho _n\rightarrow 0\) such that \(g_n(\zeta ):=\rho _{n} ^{-\alpha }f_n(z_n+\rho _n\zeta )\) converges locally uniformly with respect to the spherical metric in \(\mathbb{C }\) to a non-constant meromorphic function \(g\) of finite order, all of whose zeros have multiplicity at least \(k\).

The original form (\(\alpha =0\)) of this rescaling lemma is due to Zalcman [16], while the case \(-1<\alpha <1\) was proved by Pang [8, 9]. The present form is due to Chen and Gu [3]. This lemma also holds for \(\alpha =k\) [10] under an additional condition.

Lemma 6

[6, 7] Let \(k\) be a positive integer, \(f\) be a transcendental meromorphic function and \(P(\not \equiv 0)\) be a polynomial. Then, either \(f\) or \(f^{(k)}-P\) has infinitely many zeros. If \(f\) is a non-constant rational function, then either \(f\) or \(f^{(k)}-1\) has at least one zero.

Lemma 7

[13] Let \(k\) be a positive integer and let \(f\) be a non-constant rational function such that \(f^{(k)}\not =1\) on \(\mathbb{C }\). Then, either \(f\) is a polynomial of degree at most \(k\) or \(f(z)=z^k/k!+P_{k}(z)+a(z-b)^{-n}\), where \(a(\not =0), b\in \mathbb{C }\) and \(n\in \mathbb{N }\) are constants, and \(P_{k}\) is a polynomial of degree less than \(k\). Furthermore, \(f\) has a zero whose multiplicity is at most \(k+1\).

Lemma 8

[2, Lemma 4] Let \(k\) be a positive integer and \(f\) be a non-constant rational function. If \(f(z)\not =0\) for \(z\in \mathbb{C }\), then \(f^{(k)}-1\) has at least \(k+1\) distinct zeros on \(\mathbb{C }\).

Lemma 9

Let \(k\) be a positive integer and \(f\) be a non-constant rational function. If \(f(z)[f^{(k)}(z)-1]\not =0\) for \(z\in \mathbb{C }\setminus \{z_0\}\), where \(z_0\in \mathbb{C }\), then \(z_0\) is a zero of \(f\) with multiplicity at most \(k+1\).

Proof

We claim that \(f(z_0)=0\). For otherwise, we would have \(f(z)\not =0\) for \(z\in \mathbb{C }\) by the condition, and hence, by Lemma 8, \(f^{(k)}-1\) has at least \(k+1\ge 2\) distinct zeros, which contradicts that \(f^{(k)}(z)\not =1\) for \(z\not =z_0\).

We now assume that the zero \(z_0\) of \(f\) has multiplicity at least \(k+2\). Then, \(f^{(k)}(z_0)=0\). Hence by the condition, we have \(f(z)\not =0\) for \(z\in \mathbb{C }\setminus \{z_0\}\), and \(f^{(k)}(z)\not =1\) for \(z\in \mathbb{C }\). Thus by Lemma 7, \(f\) has a zero whose multiplicity is at most \(k+1\). Since \(f(z)\not =0\) for \(z\in \mathbb{C }\setminus \{z_0\}\), this zero coincides with \(z_0\), which contradicts the assumption that the zero \(z_0\) of \(f\) has multiplicity at least \(k+2\).

Thus, \(z_0\) is a zero of \(f\) with multiplicity at most \(k+1\). \(\square \)

Lemma 10

Let \(k,m\) be positive integers and \(f\) be a non-constant rational function. If \(f^{(k)}(z)\not =z^m\) for \(z\in \mathbb{C }\) and if \(f(z)\not =0\) for \(z\not =z_0\), where \(z_0\in \mathbb{C }\), then \(m=1\), \(z_0\not =0\) and either \(f(z)=(z-z_0)^{k+1}/(k+1)!\) or

$$\begin{aligned} f(z)=\frac{1}{(k+1)!}\cdot \frac{(z-z_0)^{k+2}}{z-(k+2)z_0}. \end{aligned}$$

Proof

Consider first the case that \(f\) is a non-constant polynomial. Then by \(f(z)\not =0\) for \(z\not =z_0\), \(f(z)=C_1(z-z_0)^l\) for some constant \(C_1\not =0\) and \(l\in \mathbb{N }\); and by \(f^{(k)}(z)\not =z^m\), \(f^{(k)}(z)=z^m+C_2\) for some constant \(C_2\not =0\). Thus, \(C_1[(z-z_0)^l]^{(k)}=z^m+C_2\). It can be easily seen that \(z_0\not =0\), \(m=1\), \(l=k+1\) and \(C_1=1/(k+1)!\). Hence \(f(z)=(z-z_0)^{k+1}/(k+1)!\).

Now we assume that \(f\) is a non-polynomial rational function. By \(f^{(k)}(z)\not =z^m\) for \(z\in \mathbb{C }\), we have

$$\begin{aligned} \left[f(z)-\frac{m!}{(m+k)!}z^{m+k}+\frac{1}{k!}z^k\right]^{(k)}=f^{(k)}(z)-z^m+1\not =1. \end{aligned}$$

Thus by Lemma 7,

$$\begin{aligned} f(z)=\frac{m!}{(m+k)!}z^{m+k}+P_k(z)+\frac{a}{(z-b)^n}, \end{aligned}$$
(1)

where \(P_k,a,b,n\) are stated as in Lemma 7. Since \(f(z)\not =0\) for \(z\not =z_0\), we also get \(f(z)=C(z-z_0)^{l}(z-b)^{-n},\) for some constant \(C\not =0\) and integer \(l\ge 0\). This, combined with (1), yields that

$$\begin{aligned} \left[\frac{m!}{(m+k)!}z^{m+k}+P_{k}(z)\right](z-b)^n=C(z-z_0)^{l}-a. \end{aligned}$$
(2)

Comparing the degree and the coefficient of the leading term of (2) yields that \(l=m+k+n\) and \(C=\frac{m!}{(m+k)!}\). Further, since each zero of the right hand side of (2) is simple, we see that \(n=1\). Thus, \(l=m+k+1\) and we can deduce from (2) that

$$\begin{aligned} \left[\frac{m!}{(m+k)!}z^{m+k}+P_{k}(z)\right](z-b)=\frac{m!}{(m+k)!}(z-z_0)^{m+k+1}-a. \end{aligned}$$
(3)

Now by comparing the coefficients of the term \(z^{m+k}\), we get \(b=(m+k+1)z_0\).

We claim that \(z_0\not =0\). In fact, if \(z_0=0\), then \(b=0\), and hence \(a=0\) by taking \(z=0\) in (3). This is a contradiction.

Next by comparing the coefficients of the term \(z^{m+k-1}\), we see that \(m=1\). Thus, \(f\) has the second desired form. \(\square \)

Lemma 11

Let \(k,m\) be positive integers and \(f\) be a rational function. If \(f(z)\not =0\) for \(z\in \mathbb{C }\), and \(f^{(k)}(z)\not =z^m\) for \(z\not =z_0\), where \(z_0\in \mathbb{C }\), then \(f\) is a constant.

Proof

If \(f\) is a polynomial, then by \(f(z)\not =0\), \(f\) must be constant. Now suppose that \(f\) is a non-polynomial rational function. Then by Lemma 10, \(f^{(k)}(z)-z^m\) must have at least one zero. Hence by the condition, \(f^{(k)}(z_0)=z_0^m\). Thus by \(f(z)\not =0\) for \(z\in \mathbb{C }\) and \(f^{(k)}(z)\not =z^m\) for \(z\not =z_0\), we can write

$$\begin{aligned} f(z)=C_1\prod _{i=1}^n(z-z_i)^{-p_i},\ f^{(k)}(z)=z^m+C_2(z-z_0)^{l}\prod _{i=1}^n(z-z_i)^{-p_i-k}, \end{aligned}$$
(4)

where \(C_1, C_2\) are non-zero constants, \(l,n, p_i\) are positive integers, and \(z_i, 0\le i\le n\) are distinct complex numbers. By the expression of \(f\) in (4), we have

$$\begin{aligned} f^{(k)}(z)=P(z)\prod _{i=1}^n(z-z_i)^{-p_i-k}, \end{aligned}$$
(5)

where \(P\) is polynomial of degree \((n-1)k\). Thus by the two expressions of \(f^{(k)}\) in (4) and (5),

$$\begin{aligned} z^m\prod _{i=1}^n(z-z_i)^{p_i+k}+C_2(z-z_0)^{l}=P(z). \end{aligned}$$
(6)

Since

$$\begin{aligned} \deg (P)=(n-1)k<m+\sum _{i=1}^n(p_i+k)=\deg \left(z^m\prod _{i=1}^n(z-z_i)^{p_i+k}\right), \end{aligned}$$

by (6), we have

$$\begin{aligned} l=m+\sum _{i=1}^n(p_i+k)=m+kn+\sum _{i=1}^np_i \end{aligned}$$
(7)

and \(C_2=-1\). By letting \(z=1/t\) in (6), we get

$$\begin{aligned} \prod _{i=1}^n(1-z_it)^{p_i+k}=(1-z_0t)^l+t^{l}P(1/t)= (1-z_0t)^l\left[1+O(t^{m+k+\sum _{i=1}^np_i})\right]\nonumber \\ \end{aligned}$$
(8)

as \(t\) goes to \(0\). Thus by taking the logarithmic derivatives,

$$\begin{aligned} \sum _{i=1}^n\frac{(p_i+k)z_i}{1-z_it}-\frac{lz_0}{1-z_0t}=O(t^{m+k-1+\sum _{i=1}^np_i}). \end{aligned}$$
(9)

It follows that

$$\begin{aligned} \sum _{i=1}^n(p_i+k)z_i^j-lz_0^j=0\ \ \text{ for}\ \ 1\le j\le m+k-1+\sum _{i=1}^np_i. \end{aligned}$$
(10)

Since \(m+k-1+\sum _{i=1}^np_i\ge n+1\), it follows that the system of linear equations

$$\begin{aligned} \sum _{i=0}^nz_i^jx_i=0,\ \ \ 1\le j\le n+1 \end{aligned}$$
(11)

has a non-zero solution \((x_0,x_1,\ldots ,x_n)=(-l,p_1+k,\ldots ,p_n+k)\). This is impossible, as all \(z_i\) are distinct. \(\square \)

Lemma 12

Let \(k,m\) be positive integers, and let \(f\) be a non-constant rational function. If \(f(z)[f^{(k)}(z)-z^m]\not =0\) for \(z\not =0\), and the multiplicity is at least \(m+k+1\) when \(0\) is a zero of \(f\), then \(k=1\) and \(0\) is a zero of \(f\) with exact multiplicity \(2m+2\).

Proof

First, we show that \(f\) cannot be a polynomial. Suppose not, then by \(f(z)\not =0\) for \(z\not =0\), \(f(z)=Cz^s\) for some constant \(C\not =0\) and integer \(s\in \mathbb{N }\). Further, by the condition, \(s\ge m+k+1\). Thus, \(f^{(k)}(z)-z^m=Az^{s-k}-z^m=Az^m(z^{s-k-m}-1/A)\), where \(A\not =0\) is a constant. This contradicts that \(f^{(k)}(z)-z^m\not =0\) for \(z\not =0\).

Thus, \(f\) is a non-polynomial rational function. By Lemma 11, \(f\) has at least one zero, and hence by \(f(z)\not =0\) for \(z\not =0\), we must have \(f(0)=0\). Thus, we can write

$$\begin{aligned} f(z)=C_1z^l\prod _{i=1}^n(z-z_i)^{-p_i}, \end{aligned}$$
(12)

where \(C_1\not =0\) is constant, \(z_i\in \mathbb{C }, 1\le i\le n\) are distinct and non-zero, and \(n,l, p_i\in \mathbb{N }\) with \(l\ge m+k+1\) (by the condition). Thus, 0 is a zero of \(f^{(k)}\) with multiplicity \(l-k\ge m+1\), and hence 0 is a zero of \(f^{(k)}(z)-z^m\) with exact multiplicity \(m\). Hence, since \(f^{(k)}(z)-z^m\not =0\) for \(z\not =0\), we have

$$\begin{aligned} f^{(k)}(z)=z^m+C_2z^m\prod _{i=1}^n(z-z_i)^{-p_i-k}, \end{aligned}$$
(13)

for some constant \(C_2\not =0\). However, by (12), one can obtain by induction that

$$\begin{aligned} f^{(k)}(z)=C_1z^{l-k}P(z)\prod _{i=1}^n(z-z_i)^{-p_i-k}, \end{aligned}$$
(14)

where

$$\begin{aligned} P(z)=\prod _{j=0}^{k-1}\left(l-j-\sum _{i=1}^np_i\right)z^{nk}+\cdots \ (\not \equiv 0) \end{aligned}$$
(15)

is a polynomial of degree at most \(nk\). Thus by the two expressions of \(f^{(k)}\) in (13) and (14),

$$\begin{aligned} z^m\prod _{i=1}^n(z-z_i)^{p_i+k}+C_2z^m=C_1z^{l-k}P(z), \end{aligned}$$

and hence

$$\begin{aligned} \prod _{i=1}^n(z-z_i)^{p_i+k}=-C_2+C_1z^{l-k-m}P(z). \end{aligned}$$
(16)

Then, comparing the degrees of the both sides of (16) shows that

$$\begin{aligned} \sum _{i=1}^n(p_i+k)=l-k-m+\deg (P). \end{aligned}$$
(17)

Since \(\deg (P)\le nk\), it follows from (17) that \(l\ge m+k+\sum _{i=1}^np_i\). Thus by (15), \(\deg (P)=nk\) and then by (17),

$$\begin{aligned} l=m+k+\sum _{i=1}^np_i. \end{aligned}$$
(18)

By (16), we also have

$$\begin{aligned} C_2=-\prod _{i=1}^n(-z_i)^{p_i+k}. \end{aligned}$$
(19)

Now by taking the logarithmic derivatives of (16), we have

$$\begin{aligned} \sum _{i=1}^n\frac{p_i+k}{z-z_i}=O(z^{l-k-m-1})\ \ \text{ as}\ \ z\rightarrow 0. \end{aligned}$$
(20)

Thus,

$$\begin{aligned} \sum _{i=1}^n\frac{p_i+k}{(z_i)^j}=0\ \ \text{ for}\ \ 1\le j\le l-k-m-1. \end{aligned}$$
(21)

It follows that the system of linear equations

$$\begin{aligned} \sum _{i=1}^n\frac{x_i}{(z_i)^j}=0,\ \ \ 1\le j\le l-k-m-1 \end{aligned}$$
(22)

has a non-zero solution \((x_1,\ldots ,x_n)=(p_1+k,\ldots ,p_n+k)\). Since \(z_i\) are non-zero and distinct, we get \(l-k-m-1<n\), and hence by (18), \(\sum _{i=1}^np_i<n+1\). It follows that all \(p_i=1\). Thus, \(l=m+k+n\) and by (21),

$$\begin{aligned} \sum _{i=1}^n\frac{1}{(z_i)^j}=0,\ \ j=1,\ldots ,n-1. \end{aligned}$$
(23)

It follows from the well-known Newton’s formula that \(\prod _{i=1}^n(z-z_i)=z^n-r\), where \(r\not =0\) is a constant. Hence by (19)

$$\begin{aligned} C_2=-\prod _{i=1}^n(-z_i)^{p_i+k}=-(-r)^{k+1}. \end{aligned}$$
(24)

Thus by (12) and (13),

$$\begin{aligned} f(z)&= C_1z^{k+m+n}(z^n-r)^{-1}=C_1z^{k+m}(1-rz^{-n})^{-1},\\ \nonumber f^{(k)}(z)&= z^m-(-r)^{k+1}z^m(z^n-r)^{-k-1} \end{aligned}$$
(25)
$$\begin{aligned}&= z^m\left[1-(-rz^{-n})^{k+1}(1-rz^{-n})^{-k-1}\right]. \end{aligned}$$
(26)

By (25), we have

$$\begin{aligned} f(z)=C_1z^{k+m}\sum _{s=0}^\infty (rz^{-n})^s =C_1\sum _{s=0}^\infty r^sz^{k+m-sn} \end{aligned}$$

for \(z\) satisfying \(|rz^{-n}|<1\), and hence

$$\begin{aligned} f^{(k)}(z)&= C_1\sum _{s=0}^\infty r^s\prod _{j=1}^k(j+m-sn)z^{m-sn}\nonumber \\&= C_1z^m\sum _{s=0}^\infty \prod _{j=1}^k(j+m-sn)(rz^{-n})^s. \end{aligned}$$
(27)

Thus by (26) and (27) with writing \(w=rz^{-n}\),

$$\begin{aligned} 1-(-w)^{k+1}(1-w)^{-k-1}=C_1\sum _{s=0}^\infty \prod _{j=1}^k(j+m-sn)w^s\ \ \text{ for}\ \ |w|<1. \end{aligned}$$
(28)

Now comparing the coefficients of (28) yields that \(\prod _{j=1}^k(j+m-sn)=0\) for \( 1\le s\le k\). Thus, \(sn-m\in \{1,2,\ldots , k\}\) for each \(1\le s\le k\). In particular, \(1\le n-m\le k\) and \(1\le kn-m\le k\). It follows that \(k=1\) and \(n=m+1\), and hence \(0\) is a zero of \(f\) with exact multiplicity \(l=m+k+n=2m+2\). \(\square \)

3 Proof of Theorem 4

Let \(z_0\in D\) be a point and \(\{f_n\}\subset \mathcal{F }\) be a sequence. We have to prove that \(\{f_n\}\) has a subsequence which is normal at \(z_0\). We may assume that \(z_0=0\). Since \(\{E_{f_n}\}\) are locally uniformly discrete in \(D\), there exists \(\delta _0>0\) such that \(E_{f_n}\cap \Delta (0,\delta _0)\) contains at most one point \(z_{f_n}\). That is to say, we have \(f_n\not =0\) and \(f_n^{(k)}\not =h\) on \(\Delta (0,\delta _0)\setminus \{z_{f_n}\}\) for all \(f_n\). The following considerations for \(\{f_n\}\) are understood to always hold with respect to the disk \(\Delta (0,\delta _0)\).

Case 1

There exists \(0<\delta <\delta _0\) such that \(|z_{f_n}|\ge \delta \) for all \(f_n\) (with \(n\) sufficiently large). Then, we have \(f_n\not =0\) and \(f_n^{(k)}\not =h\) on \(\Delta (0,\delta )\) for all \(f_n\). Thus by Theorem 1, \(\{f_n\}\) is normal on \(\Delta (0,\delta )\), and hence at \(z_0=0\).

Case 2

There exists a subsequence of \(\{z_{f_n}\}\), which we continue to call \(\{z_{f_n}\}\), such that \(z_{f_n}\rightarrow z_0=0\). Then by Theorem 1, \(\{f_n\}\) is normal on \(\Delta ^\circ (0,\delta _0)\).

Suppose that \(\{f_n\}\) has no subsequence which is normal at \(0\). Next we consider two cases according to whether \(h(0)\) is \(0\) or not.

Case 3

For the case \(h(0)\not =0\), we may say that \(h(0)=1\). Since \(\{f_n\}\) is not normal at \(0\), by Lemma 5, there exist a subsequence of \(\{f_n\}\) which we continue to call \(\{f_n\}\), a sequence of points \(z_n\rightarrow 0\) and a sequence of positive numbers \(\rho _n\rightarrow 0\) such that

$$\begin{aligned} g_n(\zeta ):=\rho _n^{-k}f_n(z_n+\rho _n\zeta )\rightarrow g(\zeta ) \end{aligned}$$
(29)

spherically locally uniformly on \(\mathbb{C }\), where \(g\) is non-constant and meromorphic on \(\mathbb{C }\).

Claim 1

\(g^{(k)}(\zeta )\not \equiv 1\). In fact, if \(g^{(k)}(\zeta )\equiv 1\), then \(g\) is a polynomial with exact degree \(k\), so that \(g\) has a zero \(\zeta _0\in \mathbb{C }\) with multiplicity at most \(k\), and hence by applying Hurwitz’s theorem to (29), \(g_n\) (for sufficiently large \(n\)) has a zero \(\zeta _n\rightarrow \zeta _0\) with multiplicity at most \(k\). It follows that \(f_n\) has a zero \(z_n+\rho _n\zeta _n\rightarrow 0\) with multiplicity at most \(k\). This contradicts the assumption that all zeros of \(f_n\) have multiplicity at least \(k+2\).

Claim 2

\(g\) has at most one zero, and if it has, then the multiplicity is at least \(k+2\).

The latter assertion follows from an argument similar to that in Claim 1. We now prove the former. Suppose that \(g\) has at least two distinct zeros \(\zeta _1,\zeta _2\in \mathbb{C }\). Then by applying Hurwitz’s theorem to (29), \(g_n\) (for sufficiently large \(n\)) has two distinct zeros \(\zeta _n^{(1)}, \zeta _n^{(2)}\) tending to \(\zeta _1,\zeta _2\) respectively, and hence \(f_n\) has two distinct zeros \(z_n+\rho _n\zeta _n^{(1)}\) and \(z_n+\rho _n\zeta _n^{(2)}\), both tending to 0. This contradicts that \(f_n(z)\not =0\) for \(z\not =z_{f_n}\).

Claim 3

\(g^{(k)}-1\) has at most one zero.

By Claim 1 and the fact that

$$\begin{aligned} f_n^{(k)}(z_n+\rho _n\zeta )-h(z_n+\rho _n\zeta )=g_n^{(k)}(\zeta )-h(z_n+\rho _n\zeta )\rightarrow g^{(k)}(\zeta )-1(\not \equiv 0)\nonumber \\ \end{aligned}$$
(30)

locally uniformly on \(\mathbb{C }\setminus g^{-1}(\infty )\), an argument similar to that in Claim 2 yields this claim.

Claim 4

Either \(g\not =0\) or \(g^{(k)}\not =1\).

Suppose not, say \(g(\zeta _0^{(1)})=0\) and \(g^{(k)}(\zeta _0^{(2)})=1\). Since \(\zeta _0^{(1)}\) is a zero of \(g\) with multiplicity \(\ge k+2\), we have \(g^{(k)}(\zeta _0^{(1)})=0\), and hence \(\zeta _0^{(1)}\not =\zeta _0^{(2)}\). By applying Hurwitz’s theorem to (29) and (30), there exist points \(\zeta _n^{(i)}\rightarrow \zeta _0^{(i)}\) such that \(g_n(\zeta _n^{(1)})=0\) and \(g_n^{(k)}(\zeta _n^{(2)})-h(z_n+\rho _n\zeta _n^{(2)})=0\), and hence \(f_n(z_n+\rho _n\zeta _n^{(1)})=0\) and \(f_n^{(k)}(z_n+\rho _n\zeta _n^{(2)})-h(z_n+\rho _n\zeta _n^{(2)})=0\). Since \(f_n(z)\not =0\) and \(f^{(k)}_n(z)\not =h(z)\) for \(z\not =z_{f_n}\), it follows that \(z_n+\rho _n\zeta _n^{(1)}=z_n+\rho _n\zeta _n^{(2)}(=z_{f_n})\) so that \(\zeta _n^{(1)}=\zeta _n^{(2)}\), and hence \(\zeta _0^{(1)}=\zeta _0^{(2)}\). This is a contradiction.

Thus by Claims 2–4, \(g(g^{(k)}-1)\) has at most one zero, and hence by Lemma 6, \(g\) is a rational function. Further by Lemma 9, \(g\) has a zero with multiplicity at most \(k+1\). This contradicts Claim 2 which says that the zero of \(g\) has multiplicity at least \(k+2\).

Case 4

Next we consider the case that \(h(0)=0\). Then \( h(z)=z^m\phi (z)\), where \(m\in \mathbb{N }\) and \(\phi \) is holomorphic with \(\phi (0)\not =0\). We can say \(\phi \not =0\) in \(\Delta (0,\delta _0)\) with the normalization \(\phi (0)=1\). Set for each \(n\)

$$\begin{aligned} F_n(z)=z^{-m}f_n(z). \end{aligned}$$
(31)

Then \(F_n(z)\not =0\) for \(z\not =z_{f_n}\), since \(f_n(z)\not =0\) for \(z\not =z_{f_n}\).

We first show that \(\{F_n\}\) has no subsequence which is normal at 0. Suppose not, say \(\{F_n\}\) is normal at 0. Then \(\{F_n\}\) has a subsequence, which we continue to call \(\{F_n\}\), such that \(\{F_n\}\) converges spherically locally uniformly to \(\psi \) which may be \(\infty \) identically in some neighborhood \(\Delta (0,\eta _0)\).

If \(\psi (0)\not =\infty \), then for sufficiently large \(n\), \(f_n(z)\not =\infty \) and \(|F_n(z)|\le M\) in some closed domain \(\overline{\Delta }(0,\eta )\) with \(\eta <\eta _0\), where \(M>0\) is a constant. It follows that the functions \(f_n\) are holomorphic and satisfy \(|f_n(z)|\le M|z|^{m}\) on \(\overline{\Delta }(0,\eta )\). By the Montel’s theorem, \(\{f_n\}\) is normal at 0, which contradicts our assumption that \(\{f_n\}\) is not normal at 0.

If \(\psi (0)=\infty \), then for sufficiently large \(n\), \(|F_n(z)|>1\) in some closed domain \(\overline{\Delta }(0,\eta )\) with \(\eta <\eta _0\). We claim that \(f_n(z_{f_n})\not =0\) for all (sufficiently large) \(n\). In fact, if \(f_n(z_{f_n})=0\), then as \(F_n(z_{f_n})\not =0\), we see that \(z_{f_n}=0\), and so by the assumption on the multiplicities of common zeros of \(f_n\) and \(h\), \(z_{f_n}=0\) is a zero of \(f_n\) with multiplicity at least \(m+k+1>m\), and hence \(F_n(0)=0\), which contradicts that \(|F_n(z)|>1\). Thus, \(f_n(z_{f_n})\not =0\). This, combined with the fact that \(f_n(z)\not =0\) for \(z\not =z_{f_n}\), shows that the functions \(1/f_n\) are holomorphic. By \(|F_n(z)|>1\), we have \(|1/f_n(z)|<|z|^{-m}\) on \(\overline{\Delta }(0,\eta )\). Now the maximum modulus principle implies that \(|1/f_n(z)|\le \eta ^{-m}\) on \(\overline{\Delta }(0,\eta )\) and hence \(\{1/f_n\}\) is normal at 0 by Montel’s theorem. This again contradicts our assumption that \(\{f_n\}\) is not normal at 0.

Thus, \(\{F_n\}\) has no subsequence which is normal at 0.

Claim 5

If \(z_{f_n}\) is a zero of \(F_n\), then the multiplicity is at least \(k+1\), or \(k+2\) if \(z_{f_n}\not =0\). In fact, suppose \(F_n(z_{f_n})=0\), then \(f_n(z_{f_n})=0\). Thus for the case \(z_{f_n}\not =0\), the claim is true by the assumption on the multiplicities of the zeros of \(f_n\), while for the case \(z_{f_n}=0\), \(z_{f_n}=0\) is a common zero of \(f_n\) and \(h\), and hence by the assumption on the multiplicities of the common zeros of \(f_n\) and \(h\), 0 is a zero of \(f_n\) with multiplicity at least \(k+m+1\), and then the claim follows.

Thus by Claim 5 and the fact that \(F_n(z)\not =0\) for \(z\not =z_{f_n}\), Lemma 5 can be applied, and so there exist a subsequence of \(\{f_n\}\), which we continue to call \(\{f_n\}\), a sequence of points \(w_n\rightarrow 0\) and a sequence of positive numbers \(\eta _n\rightarrow 0\) such that

$$\begin{aligned} \widehat{g}_n(\zeta ):=\eta _n^{-k}F_n(w_n+\eta _n\zeta ) =\eta _n^{-k}(w_n+\eta _n\zeta )^{-m}f_n(w_n+\eta _n\zeta )\rightarrow \widehat{g}(\zeta ) \end{aligned}$$
(32)

spherically locally uniformly on \(\mathbb{C }\), where \(\widehat{g}\) is non-constant and meromorphic on \(\mathbb{C }\).

Claim 6

\(\widehat{g}\) has at most one zero.

In fact, if \(\widehat{g}\) has two distinct zeros \(\zeta _1,\zeta _2\in \mathbb{C }\), then applying Hurwitz’s theorem to (32), there exist points \(\zeta _{i,n}\rightarrow \zeta _i\) such that \(\widehat{g}_n(\zeta _{i,n})=0\) and hence \(F_n(w_n+\eta _n\zeta _{i,n})=0\). Since \(F_n(z)\not =0\) for \(z\not =z_{f_n}\), we get \(w_n+\eta _n\zeta _{1,n}=w_n+\eta _n\zeta _{2,n}(=z_{f_n})\). It follows that \(\zeta _{1,n}=\zeta _{2,n}\), and hence \(\zeta _1=\zeta _2\). This is a contradiction.

Next we consider two subcases according to whether the sequence \(\{w_n/\eta _n\}\) is bounded or unbounded.

Case 5

First assume that the sequence \(\{w_n/\eta _n\}\) is unbounded. Then, there exists a subsequence, which we continue to call \(\{w_n/\eta _n\}\), such that \(w_n/\eta _n\rightarrow \infty \).

Claim 7

If \(\widehat{g}\) has a zero, then the multiplicity is at least \(k+2\). And hence, \(\widehat{g}^{(k)}\not \equiv 1\).

In fact, if \(\widehat{g}(\zeta _{0})=0\), then as above, there exist points \(\zeta _{n}\rightarrow \zeta _0\) such that \(F_n(w_n+\eta _n\zeta _{n})=0\), and hence \(z_{f_n}=w_n+\eta _n\zeta _{n}\). As \(w_n/\eta _n\rightarrow \infty \), we see that \(z_{f_n}\not =0\), and hence by Claim 5, \(z_{f_n}\) is a zero of \(F_n\) with multiplicity at least \(k+2\). Thus, \(\zeta _n\) is a zero of \(\widehat{g}_n\) with multiplicity at least \(k+2\). The claim then follows.

Claim 8

\(\widehat{g}^{(k)}-1\) has at most one zero.

This follows from a similar argument as above with the following fact that

$$\begin{aligned} \frac{f_n^{(k)}(w_n+\eta _n\zeta )}{h(w_n+\eta _n\zeta )}&= \frac{[(\eta _n)^{-k}f_n(w_n+\eta _n\zeta )]^{(k)}}{h(w_n+\eta _n\zeta )}= \frac{[(w_n+\eta _n\zeta )^m\widehat{g}_n(\zeta )]^{(k)}}{h(w_n+\eta _n\zeta )}\nonumber \\&= \frac{\left[\left(\zeta \!+\!\frac{w_n}{\eta _n}\right)^m\widehat{g}_n(\zeta )\right]^{(k)}}{\left(\zeta \!+\!\frac{w_n}{\eta _n}\right)^m\phi (w_n+\eta _n\zeta )} =\frac{\sum _{i=0}^k\genfrac(){0.0pt}{}{k}{i}\left[\left(\zeta \!+\!\frac{w_n}{\eta _n}\right)^{m}\right]^{(k-i)} \widehat{g}_n^{(i)}(\zeta )}{\left(\zeta +\frac{w_n}{\eta _n}\right)^m\phi (w_n+\eta _n\zeta )}\nonumber \\&= \frac{1}{\phi (w_n+\eta _n\zeta )}\sum _{i=0}^{k} \frac{C_i\widehat{g}_n^{(i)}(\zeta )}{\left(\zeta +\frac{w_n}{\eta _n}\right)^{k-i}} \rightarrow \widehat{g}^{(k)}(\zeta ) \end{aligned}$$
(33)

locally uniformly on \(\mathbb{C }\setminus \widehat{g}^{-1}(\infty )\), where \(C_i=m(m-1)\cdots (m-(k-i)+1)\genfrac(){0.0pt}{}{k}{i}\) are constants, and in particular, \(C_k=1\).

Claim 9

Either \(\widehat{g}\not =0\) or \(\widehat{g}^{(k)}\not =1\).

Combined with the fact that \(F_n(z)\not =0\) and \(f_n^{(k)}(z)\not =h(z)\) for \(z\not =z_{f_n}\), this can be shown by an argument similar to the proof of Claim 4 by applying Hurwitz’s theorem to (32) and (33).

However, as in Case 2.1, the above Claims 6–9 lead to a contradiction.

Case 6

Now we consider the case that \(\{w_n/\eta _n\}\) is bounded. Then, there is a subsequence, which we continue to call \(\{w_n/\eta _n\}\), such that \(w_n/\eta _n\rightarrow \alpha \in \mathbb{C }\). It follows from (32) that

$$\begin{aligned} \frac{f_n(\eta _n\zeta )}{\eta _n^{k+m}\zeta ^m}=\widehat{g}_n\left(\zeta -\frac{w_n}{\eta _n}\right) \rightarrow \widehat{g}(\zeta -\alpha ) \end{aligned}$$
(34)

spherically locally uniformly on \(\mathbb{C }\), and hence

$$\begin{aligned} G_n(\zeta ):=\frac{f_n(\eta _n\zeta )}{\eta _n^{k+m}}=\zeta ^m\cdot \frac{f_n(\eta _n\zeta )}{\eta _n^{k+m}\zeta ^m}\rightarrow G(\zeta ):=\zeta ^m \widehat{g}(\zeta -\alpha ) \end{aligned}$$
(35)

spherically locally uniformly on \(\mathbb{C }^*=\mathbb{C }\setminus \{0\}\), or on \(\mathbb{C }\) if \(\widehat{g}(-\alpha )\not =\infty \). Obviously, \(G\) is meromorphic on \(\mathbb{C }\) and \(G\not \equiv 0\).

Claim 10

\(G\) is non-constant. Suppose not. Then, \( \widehat{g}(\zeta -\alpha )=A\zeta ^{-m}\) for some non-zero constant \(A\). Thus by (34), we see that

$$\begin{aligned} G_n(\zeta ):=\frac{f_n(\eta _n\zeta )}{\eta _n^{k+m}}\rightarrow A \end{aligned}$$
(36)

locally uniformly on \(\mathbb{C }\). Hence \(G^{(k)}_n(\zeta )\rightarrow 0\), so that

$$\begin{aligned} \eta _n^{-m}[f_n^{(k)}(\eta _n\zeta )-h(\eta _n\zeta )]=G_n^{(k)}(\zeta )-\zeta ^m\phi (\eta _n\zeta )\rightarrow -\zeta ^m \end{aligned}$$
(37)

locally uniformly on \(\mathbb{C }\). Thus by applying Hurwitz’s theorem to (37), there exist exactly \(m\) points \(\zeta ^{(j)}_n\rightarrow 0\), \(j=1,2,\ldots ,m\), such that \(f_n^{(k)}(\eta _n\zeta ^{(j)}_n)=h(\eta _n\zeta ^{(j)}_n)\). It follows from \(f_n^{(k)}(z)\not =h(z)\) for \(z\not =z_{f_n}\) that \(z_{f_n}=\eta _n\zeta ^{(j)}_n\) for all \(j\). This shows that the \(m\) points \(\zeta ^{(j)}_n\) are coincide with \(\zeta _n:=z_{f_n}/\eta _n\), and \(z_{f_n}\) is a zero of \(f_n^{(k)}(z)-h(z)\) with multiplicity \(m\), and by (36), \(f_n(z_{f_n})\not =0\). Thus, \(f_n(z)\not =0\) for \(z\in \Delta (0,\delta _0)\).

Since \(\{f_n\}\) is normal on \(\Delta ^\circ (0,\delta _0)\), but not normal at 0, it follows from \(f_n\not =0\) on \(\Delta (0,\delta _0)\) that there exists a subsequence of \(\{f_n\}\), which we continue to call \(\{f_n\}\), such that \(f_n\rightarrow 0\) and hence \(f_n^{(k)}\rightarrow 0\) locally uniformly on \(\Delta ^\circ (0,\delta _0)\).

Thus by the argument principle, we see that

$$\begin{aligned} n\left(\frac{\delta _0}{2},f_n^{(k)}-h\right)-n\left(\frac{\delta _0}{2},\frac{1}{f_n^{(k)}-h}\right) \rightarrow n\left(\frac{\delta _0}{2},h\right)-n\left(\frac{\delta _0}{2},\frac{1}{h}\right)=-m,\quad \end{aligned}$$
(38)

where \(n(r,f)\) is the number of poles of \(f\) in \(\Delta (0,r)\), and \(n(r,1/f)\) is the number of zeros of \(f\) in \(\Delta (0,r)\). Since both hand sides of (38) are integers, we see that for sufficiently large \(n\),

$$\begin{aligned} n\left(\frac{\delta _0}{2},f_n^{(k)}-h\right)=n\left(\frac{\delta _0}{2},\frac{1}{f_n^{(k)}-h}\right)-m=0. \end{aligned}$$
(39)

It follows from (39) that \(f_n\) are holomorphic on \(\Delta (0,\delta _0/2)\). Hence by \(f_n\rightarrow 0\) locally uniformly on \(\Delta ^\circ (0,\delta _0)\), we get \(f_n\rightarrow 0\) locally uniformly on \(\Delta (0,\delta _0)\). This contradicts that \(\{f_n\}\) is not normal at 0.

Claim 11

\(G\) has at most one zero on \(\mathbb{C }^*\), or on \(\mathbb{C }\) if \(\widehat{g}(-\alpha )\not =\infty \).

This can be proved by applying Hurwitz’s theorem to (35) with the fact that \(f_n(z)\not =0\) for \(z\not =z_{f_n}\).

Claim 12

If \(G\) has a zero on \(\mathbb{C }^*\), then \(z_{f_n}\not =0\) and the multiplicity of the zero of \(G\) on \(\mathbb{C }^*\) is at least \(k+2\), or \(k+3\) if \(m=1\).

Suppose that \(\zeta _0\in \mathbb{C }^*\) is a zero of \(G\). Then by applying Hurwitz’s theorem to (35), \(G_n\) has a zero \(\zeta _n\rightarrow \zeta _0\). Since \(\zeta _0\not =0\), \(\zeta _n\not =0\) (for \(n\) large enough). It follows that \(\eta _n\zeta _n\rightarrow 0\) is a non-zero zero of \(f_n\). Since \(f_n(z)\not =0\) for \(z\not =z_{f_n}\), we see that \(z_{f_n}=\eta _n\zeta _n\not =0\). Hence by the assumption, \(z_{f_n}\) is a zero of \(f_n\) with multiplicity at least \(k+2\), or \(k+3\) if \(m=1\). It follows that the multiplicity of the zero \(\zeta _0\) of \(G\) is at least \(k+2\), or \(k+3\) if \(m=1\).

Claim 13

If \(G(0)=0\), then \(f_n(0)=0\) and hence \(z_{f_n}=0\) (for sufficiently large \(n\)).

For otherwise, say \(f_n(0)\not =0\). Then by (34), 0 is a pole of \(\widehat{g}_n(\zeta -w_n/\eta _n)\) with multiplicity at least \(m\), and hence is a pole of \(\widehat{g}(\zeta -\alpha )\) with multiplicity at least \(m\). This shows that \(0\) is not a zero of \(G(\zeta )=\zeta ^m\widehat{g}(\zeta -\alpha )\), which is a contradiction.

By Claims 12 and 13, we see that if \(G(0)=0\), then \(G(\zeta )\not =0\) for \(\zeta \in \mathbb{C }^*\). Hence by Claim 11, \(G\) has at most one zero on \(\mathbb{C }\).

Claim 14

If \(0\) is a zero of \(G\), then the multiplicity is at least \(m+k+1\) for \(k>1\) and \(2m+3\) for \(k=1\).

In fact, if \(0\) is a zero of \(G\), then Claim 13 shows that \(0\) is a common zero of \(f_n\) and \(h\). Let \(m_{f_n}\) be the multiplicity of \(0\) as a zero of \(f_n\). Then by the assumption, \(m_{f_n}\ge m+k+1\) for \(k>1\) and \(m_{f_n}\ge 2m+3\) for \(k=1\). It follows from (34) that \(-w_n/\eta _n\) is a zero of \(\widehat{g}_n\) with multiplicity \(m_{f_n}-m\), so that \(-\alpha \) is a zero of \(\widehat{g}\) with multiplicity at least \(k+1\) for \(k>1\) and \(m+3\) for \(k=1\). The assertion then follows.

Now by (35), we have

$$\begin{aligned} \eta _n^{-m}[f_n^{(k)}(\eta _n\zeta )-h(\eta _n\zeta )]=G_n^{(k)}(\zeta )-\zeta ^m\phi (\eta _n\zeta )\rightarrow G^{(k)}(\zeta )-\zeta ^m \end{aligned}$$
(40)

locally uniformly on \(\mathbb{C }^*\setminus G^{-1}(\infty )\), or on \(\mathbb{C }\setminus G^{-1}(\infty )\) if \(\widehat{g}(-\alpha )\not =\infty \).

Claim 15

\(G^{(k)}(\zeta )\not \equiv \zeta ^m\).

For otherwise, \(G(\zeta )=c\zeta ^{m+k}+P(\zeta )\) for some constant \(c\not =0\) and polynomial \(P\) of degree at most \(k-1\). Thus, \(G\) has at least one zero in \(\mathbb{C }\). Let \(\zeta _0\) be a zero of \(G\), then by the fact that \(\zeta _0\) has multiplicity \(>k\), we get \(G^{(k)}(\zeta _0)=0\) and hence \(\zeta _0=0\) by \(G^{(k)}(\zeta )\equiv \zeta ^m\). It follows that \(G(\zeta )=c^{\prime }\zeta ^{l}\) for some \(c^{\prime }\in \mathbb{C }\) and \(l\in \mathbb{N }\). Thus, \(c\zeta ^{m+k}+P(\zeta )=c^{\prime }\zeta ^l\). By comparing the degrees and coefficients, we see that \(P\equiv 0\). Thus, \(G(\zeta )=c\zeta ^{m+k}\). This contradicts Claim 14 which says that the multiplicity is at least \(m+k+1\) if 0 is a zero of \(G\).

Claim 16

\(G^{(k)}(\zeta )-\zeta ^m\) has at most one zero on \(\mathbb{C }^*\), or on \(\mathbb{C }\) if \(\widehat{g}(-\alpha )\not =\infty \).

By Claim 15, this claim can be seen by applying Hurwitz’s theorem to (40) with the condition \(f_n^{(k)}(z)\not =h(z)\) for \(z\not =z_{f_n}\).

By Claims 10, 11 and 16, it follows from Lemma 6 that \(G\) is a non-constant rational function.

Claim 17

If \(G^{(k)}(0)=0\), then \(G\not =0\) on \(\mathbb{C }^*\).

Suppose that \(G(\zeta _0)=0\) for some \(\zeta _0\not =0\). Then \(G\) is holomorphic at 0. Since we have proved before Claim 14 that \(G\) has at most one zero on \(\mathbb{C }\), we get \(G(0)\not =0\). Thus by \(G(0)\not =0,\infty \), it follows from \(G(\zeta )=\zeta ^m \widehat{g}(\zeta -\alpha )\) that \(\widehat{g}(-\alpha )=\infty \) and that \(0\) is a pole of \(\widehat{g}(\zeta -\alpha )\) with exact multiplicity \(m\). Thus by (34), there exists a positive number \(\delta \) such that \(f_n(\eta _n\zeta )\not =\infty \) for \(|\zeta |\le \delta \) (and \(n\) large enough), and hence \(G_n(\zeta )\) is holomorphic on \(|\zeta |\le \delta \). Thus, on \(|\zeta |\le \delta \), the convergence of (40) is uniformly, and then Hurwitz’s theorem can be applied to (40). Since \(G^{(k)}(0)=0^m\) and \(G^{(k)}(\zeta )\not \equiv \zeta ^m\) by Claim 15, there exist points \(\zeta _n\rightarrow 0\) such that \(G_n^{(k)}(\zeta _n)=\zeta _n^m\phi (\eta _n\zeta _n)\) so that \(f_n^{(k)}(\eta _n\zeta _n)=h(\eta _n\zeta _n)\). Hence, by \(f_n^{(k)}(z)\not =h(z)\) for \(z\not =z_{f_n}\), we get \(z_{f_n}=\eta _n\zeta _n\). On the other hand, since \(G(\zeta _0)=0\), applying Hurwitz’s theorem to (35) yields that there exist points \(\zeta ^{\prime }_n\rightarrow \zeta _0\) such that \(G_n(\zeta ^{\prime }_n)=0\) so that \(f_n(\eta _n\zeta ^{\prime }_n)=0\). Thus by \(f_n(z)\not =0\) for \(z\not =z_{f_n}\), we get \(z_{f_n}=\eta _n\zeta ^{\prime }_n\). Hence \(\eta _n\zeta _n=\eta _n\zeta ^{\prime }_n(=z_{f_n})\) and then \(\zeta ^{\prime }_n=\zeta _n\). It follows that \(\zeta _0=0\), which is a contradiction.

Claim 18

If \(G^{(k)}(0)=0\), then \(G^{(k)}(\zeta )-\zeta ^m\not =0\) on \(\mathbb{C }^*\).

Suppose that \(G^{(k)}(\zeta _0)=\zeta _0^m\) for some \(\zeta _0\not =0\). Then \(G\) is holomorphic at 0. If \(G(0)=0\), then by Claim 14, \(\widehat{g}(-\alpha )=0\), and hence by Claim 16, \(G^{(k)}(\zeta )-\zeta ^m\) has at most one zero on \(\mathbb{C }\). This contradicts that \(G^{(k)}(0)=0^m\) and \(G^{(k)}(\zeta _0)=\zeta _0^m\). Thus, \(G(0)\not =0\). Now by an argument similar to that in the proof of Claim 17, there exists \(\delta >0\) such that on \(|\zeta |\le \delta \), the convergence of (40) is uniformly, and then Hurwitz’s theorem can be applied to (40). It follows from \(G^{(k)}(0)=0^m\) that there exist points \(\zeta _n\rightarrow 0\) such that \(G_n^{(k)}(\zeta _n)=\zeta _n^m\phi (\eta _n\zeta _n)\) so that \(f_n^{(k)}(\eta _n\zeta _n)=h(\eta _n\zeta _n)\). Thus by \(f_n^{(k)}(z)=h(z)\) for \(z\not =z_{f_n}\), we get \(z_{f_n}=\eta _n\zeta _n\). On the other hand, since \(G^{(k)}(\zeta _0)=\zeta _0^m\) with \(\zeta _0\not =0\), by applying Hurwitz’s theorem to (40), there exist points \(\zeta ^{\prime }_n\rightarrow \zeta _0\) such that \(G_n^{(k)}(\zeta ^{\prime }_n)=(\zeta ^{\prime }_n)^m\phi (\eta _n\zeta ^{\prime }_n)\) so that \(f_n^{(k)}(\eta _n\zeta ^{\prime }_n)=h(\eta _n\zeta ^{\prime }_n)\). Again by \(f_n^{(k)}(z)=h(z)\) for \(z\not =z_{f_n}\), we get \(z_{f_n}=\eta _n\zeta ^{\prime }_n\). Thus \(\eta _n\zeta _n=\eta _n\zeta ^{\prime }_n(=z_{f_n})\) and then \(\zeta ^{\prime }_n=\zeta _n\). It follows that \(\zeta _0=0\), which is a contradiction.

Thus by Claims 16 and 18, \(G^{(k)}(\zeta )-\zeta ^m\) has at most one zero on \(\mathbb{C }\).

Claim 19

At least one of \(G\) and \(G^{(k)}(\zeta )-\zeta ^m\) has no zero on \(\mathbb{C }^*\).

Suppose that \(G(\zeta _0^{(1)})=0\) and \(G^{(k)}(\zeta _0^{(2)})=(\zeta _0^{(2)})^m\) for some \(\zeta _0^{(1)},\zeta _0^{(2)}\in \mathbb{C }^*\). By Claim 12, \(\zeta _0^{(1)}\) is a zero of \(G\) with multiplicity \(\ge k+2\), and hence \(G^{(k)}(\zeta _0^{(1)})=0\). It follows that \(\zeta _0^{(1)}\not =\zeta _0^{(2)}\). Applying Hurwitz’s theorem to (35) shows that there exist points \(\zeta _n^{(1)}\rightarrow \zeta _0^{(1)}\) such that \(G_n(\zeta _n^{(1)})=0\). Thus \(f_n(\eta _n\zeta _n^{(1)})=0\), and hence by \(f_n(z)\not =0\) for \(z\not =z_{f_n}\), we get \(z_{f_n}=\eta _n\zeta _n^{(1)}\). Applying Hurwitz’s theorem to (40) shows that there exist points \(\zeta _n^{(2)}\rightarrow \zeta _0^{(2)}\) such that \(f_n^{(k)}(\eta _n\zeta _n^{(2)})=h(\eta _n\zeta _n^{(2)})\), and hence by \(f_n^{(k)}(z)\not =h(z)\) for \(z\not =z_{f_n}\), we get \(z_{f_n}=\eta _n\zeta _n^{(2)}\). Thus \(\zeta _n^{(1)}=\zeta _n^{(2)}\) and hence \(\zeta _0^{(1)}=\zeta _0^{(2)}\). This contradicts that \(\zeta _0^{(1)}\not =\zeta _0^{(2)}\).

Now, according to Claim 19, we are in one of the following three cases. Note that we have proved before Claim 17 that \(G\) is a non-constant rational function.

Case 7

Neither of \(G\) and \(G^{(k)}(\zeta )-\zeta ^m\) has a zero on \(\mathbb{C }^*\).

Then by Lemma 11, \(G\) has at least one zero on \(\mathbb{C }\), and hence \(G(0)=0\). It now follows from Claim 14 and Lemma 12 that Case A cannot occur.

Case 8

\(G(\zeta _0)=0\) for some \(\zeta _0\in \mathbb{C }^*\) and \(G^{(k)}(\zeta )-\zeta ^m\) has no zero on \(\mathbb{C }^*\).

Then \(G(\zeta )\not =0\) for \(\zeta \in \mathbb{C }\setminus \{\zeta _0\}\), since we have proved before Claim 14 that \(G\) has at most one zero on \(\mathbb{C }\). Also, by Claim 17, we have \(G^{(k)}(0)\not =0\), and hence \(G^{(k)}(\zeta )\not =\zeta ^m\) for \(\zeta \in \mathbb{C }\). It now follows from Claim 12 and Lemma 10 that Case B cannot occur.

Case 9

\(G^{(k)}(\zeta _0)=\zeta _0^m\) for some \(\zeta _0\in \mathbb{C }^*\) and \(G(\zeta )\) has no zero on \(\mathbb{C }^*\).

Then \(G^{(k)}(\zeta )\not =\zeta ^m\) for \(\zeta \in \mathbb{C }\setminus \{\zeta _0\}\), since we have proved before Claim 19 that \(G^{(k)}(\zeta )-\zeta ^m\) has at most one zero on \(\mathbb{C }\). Hence by Lemma 11, \(G\) has at least one zero on \(\mathbb{C }\), so that we must have \(G(0)=0\). It now follows from Claim 14 that \(G^{(k)}(0)=0\), which contradicts that \(G^{(k)}(\zeta )\not =\zeta ^m\) for \(\zeta \in \mathbb{C }\setminus \{\zeta _0\}\).

The proof of Theorem 4 is completed.

4 Proof of Theorem 3

From the proof of Theorem 4, we have the following result, which shows that the multiplicity \(k+3\) in Theorem 4 can be replaced by \(k+2\) when the set \(E_f\) is independent of \(f\). Note that the assumption that multiplicity \(\ge k+3\) in Theorem 4 is only used to rule out the Case B within the proof of Theorem 4.

Theorem 13

Let \(k\in \mathbb{N }\) and \(h(\not \equiv 0)\) be a function holomorphic on \(D\). Let \(\mathcal{F }\) be a family of meromorphic functions in \(D\), all of whose zeros have multiplicity at least \(k+2\). Suppose that there exists a point \(z_0\in D\) such that on \(D\setminus \{z_0\}\), \(f\not =0\) and \(f^{(k)}\not =h\) for all functions \(f\in \mathcal{F }\). Suppose additionally that if \(z_0\) is a common zero of \(f\in \mathcal{F }\) and \(h\), the multiplicities \(m_f\) for \(f\) and \(m_h\) for \(h\) satisfy \(m_f\ge m_h+k+1\) for \(k>1\) and \(m_f\ge 2m_h+3\) for \(k=1\). Then \(\mathcal{F }\) is normal in \(D\).

Proof

The proof is similar to, and more simple than, the proof of Theorem 4, since in this case, we only have to consider the case that \(z_{f_n}=0\) for all \(n\) in the proof of Theorem 4. In fact, since \(z_{f_n}=0\), we see from Claim 12 that \(G\) has no zeros on \(\mathbb{C }^*\), and a similar argument by applying Hurwitz’s theorem to (40) can be used to show that \(G^{(k)}(\zeta )-\zeta ^m\) has no zeros on \(\mathbb{C }^*\). So only the Case A appears and is required to be ruled out. \(\square \)

Proof of Theorem 3

Let \(z_0\in D\) and a sequence \(\{f_n\}\subset \mathcal{F }\). We have to prove that \(\{f_n\}\) is normal at \(z_0\).

Suppose first that there exists a neighborhood \(U_0\) of \(z_0\) such that on \(U_0\setminus \{z_0\}\), \(f_1\not =0\) and \(f_1^{(k)}\not =h\). Then by the condition, on \(U_0\setminus \{z_0\}\), \(f_n\not =0\) and \(f_n^{(k)}\not =h\) for all functions \(f_n\). Hence, by Theorem 13, \(\{f_n\}\) is normal in \(U_0\) and hence normal at \(z_0\).

The other case is that there exists a sequence \(z_n\rightarrow z_0,\ z_n\not =z_0\) such that either \(f_1(z_n)=0\) or \(f_1^{(k)}(z_n)=h(z_n)\). It follows that either \(f_1(z)\equiv 0\) or \(f_1^{(k)}(z)\equiv h(z)\). Thus by the condition, either \(f_n(z)\equiv 0\) or \(f_n^{(k)}(z)\equiv h(z)\) for all \(n\).

It is obviously that \(\{f_n\}\) is normal if \(f_n(z)\equiv 0\). Thus, we consider the case that \(f_n^{(k)}(z)\equiv h(z)\). If \(\{f_n\}\) is not normal at \(z_0\), then by Lemma 5, there exists points \(z_n\rightarrow z_0\), positive numbers \(\rho _n\rightarrow 0\) and functions \(f_n\in \{f_n\}\) such that \(g_n(\zeta )=f_n(z_n+\rho _n\zeta )\rightarrow g(\zeta )\) locally uniformly on \(\mathbb{C }\), where \(g\) is a non-constant meromorphic function, all of whose zeros have multiplicity at least \(k+2\). Since \(g_n^{(k)}(\zeta )=\rho _n^kf^{(k)}_n(z_n+\rho _n\zeta )=\rho _n^kh(z_n+\rho _n\zeta )\rightarrow 0\), we get \(g^{(k)}(\zeta )\equiv 0\). Since all zeros of \(g\) have multiplicity at least \(k+2\), it follows that \(g\) is a constant, which is a contradiction.

Thus, \(\{f_n\}\) is normal at \(z_0\). Theorem 3 is proved.