1 Introduction and main results

Let D be a domain in the complex plane \(\mathbb {C}\), and \(\mathcal F\) be a family of meromorphic functions defined on D. \(\mathcal F\) is said to be normal on D, in the sense of Montel, if for any sequence \(\{f_n\} \subset \mathcal F\) there exists a subsequence \(\{f_{n_k}\}\), such that \(\{f_{n_k}\}\) converges spherically locally uniformly on D, to a meromorphic function or \(\infty \) (see [3],[8],[13]).

The following well-known normality criterion was conjectured by Hayman[3], and proved by Gu [2].

Theorem A

Let k be a positive integer. Let \(\mathcal {F}\) be a family of meromorphic functions defined in a domain D. If for each \(f\in \mathcal {F}\), \(f\ne 0\) and \(f^{(k)}\ne 1\), then \(\mathcal {F}\) is normal in D.

This result has undergone various extensions and improvements. In [5] (cf. [6, 11]), Pang-Yang-Zalcman obtained.

Theorem B

Let k be a positive integer. Let \(\mathcal {F}\) be a family of meromorphic functions defined in a domain D, all of whose zeros have multiplicity at least \(k+2\) and whose poles are multiple. Let \(h(z)(\not \equiv 0)\) be a holomorphic functions on D. If for each \(f\in \mathcal {F}\), \(f^{(k)}(z)\ne h(z)\), then \(\mathcal {F}\) is normal in D.

When \(k=1\), an example [19, Example 1] (cf. [6]) shows that the condition on the multiplicity of zeros of functions in \(\mathcal F\) cannot be weakened. Zhang-Pang-Zalcman[14] proved that when \(k\ge 2\) the multiplicity of zeros of functions in \(\mathcal F\) can be reduced from \(k+2\) to \(k+1\) in Theorem B.

Theorem E

Let \(k\ge 2\) be a positive integer. Let \(\mathcal {F}\) be a family of meromorphic functions defined in a domain D, all of whose zeros have multiplicity at least \(k+1\) and whose poles are multiple. Let \(h(z)(\not \equiv 0)\) be a holomorphic functions on D. If for each \(f\in \mathcal {F}\), \(f^{(k)}(z)\ne h(z)\), then \(\mathcal {F}\) is normal in D.

Also in [14], they indicated that one cannot further reduce the assumption on the multiplicity of the zeros from \(k+1\) to k, by considering the following example.

Example 1

(see [14])   Let \(\Delta =\{z: |z|<1\}\), \(h(z)=z\), and let

$$\begin{aligned} \mathcal{F}=\left\{ f_n(z)=nz^k\right\} . \end{aligned}$$

Clearly, all zeros of \(f_n\) are of multiplicity k, and \(f^{(k)}_n(z)=nk!\ne z\) on \(\Delta \). However, \(\mathcal F\) fails to be equicontinuous at 0, and then \(\mathcal F\) is not normal in \(\Delta \).

In this paper, we consider the case \(h(z)=z\), then \(f^{(k)}(z)\ne h(z)\) means that \(f^{(k)}\) has no fixed-points. We reduce the multiplicity of zeros of functions in \(\mathcal F\) to k, but restricting the values \(f^{(k)}\) can take at the zeros of f, as follows.

Theorem 1

Let \(k\ge 4\) be a positive integer, \(A>1\) be a constant. Let \(\mathcal F\) be a family of meromorphic functions in a domain D. If, for every function \(f\in \mathcal F\), f has only zeros of multiplicity at least k and satisfies the following conditions:

  1. (a)

    \(f(z)=0\Rightarrow |f^{(k)}(z)|\le A|z|\).

  2. (b)

    \( f^{(k)}(z)\ne z\).

  3. (c)

    All poles of f are multiple.

Then \(\mathcal F\) is normal in D.

For the case \(k=2\) or 3, the multiplicity of poles of \(f\in \mathcal F\) need be at least three.

Theorem 2

Let \(k=2\) or 3, \(A>1\) be a constant. Let \(\mathcal F\) be a family of meromorphic functions in a domain D. If, for every function \(f\in \mathcal F\), f has only zeros of multiplicity at least k and satisfies the following conditions:

  1. (a)

    \(f(z)=0\Rightarrow |f^{(k)}(z)|\le A|z|\).

  2. (b)

    \( f^{(k)}(z)\ne z\).

  3. (c)

    All poles of f have multiplicity at least 3.

Then \(\mathcal F\) is normal in D.

Example 1 shows that condition (a) in Theorems 1 and 2 cannot be removed. For the case \(k=1\), the above theorems are no longer true even if the multiplicities of poles of \(f\in \mathcal F\) are large enough, as is shown by the next example.

Example 2

Let j be a positive integer, \(\Delta =\{z: |z|<1\}\), and let

$$\begin{aligned} \mathcal{F}=\left\{ f_n(z)=\frac{z^{j+2}-1/n^{j+2}}{2z^{j}}\right\} . \end{aligned}$$

Clearly,

$$\begin{aligned} f'_n(z)=z+\frac{j}{2n^{j+2}z^{j+1}}\ne z. \end{aligned}$$

For each n, \(f_n\) has one pole \(z=0\) with multiplicity j, and \(j+2\) simple zeros \(z_m=\frac{1}{n}e^{i\frac{2m\pi }{j+2}}(m=0,1,\ldots ,j+1)\) in \(\Delta \). We have

$$\begin{aligned} f'_n(z_m)=z_m+\frac{j}{2n^{j+2}z_m^{j+1}}=\frac{j+2}{2n}e^{i\frac{2m\pi }{j+2}}, \end{aligned}$$

and then

$$\begin{aligned} |f'_n(z_m)|\le \frac{j+2}{2}|z_m|, \end{aligned}$$

that is, \(f_n(z)=0\Rightarrow |f'_n(z)|\le \frac{j+2}{2}|z|\). But, since \(f_n(1/n)=0\) and \(f_n(0)=\infty \), \(\mathcal F\) fails to be equicontinuous at \(z=0\), and then \(\mathcal F\) is not normal in \(\Delta \).

The following example shows that condition (c) in Theorem 2 is necessary, and the number 3 is best possible.

Example 3

Let \(\Delta =\{z: |z|<1\}\), and let

$$\begin{aligned} \mathcal{F}=\left\{ f_n(z)=\frac{(z-1/n)^3(z+1/n)^3}{24z^2}\right\} . \end{aligned}$$

Clearly,

$$\begin{aligned} f^{(3)}_n(z)=z+\frac{1}{n^{6}z^5}\ne z. \end{aligned}$$

For each n, \(f_n\) has two zeros \(z_1=1/n\) and \(z_2=-1/n\) of multiplicity 3. We have

$$\begin{aligned} f^{(3)}_n(\frac{1}{n})=\frac{2}{n},\quad f^{(3)}_n(-\frac{1}{n})=-\frac{2}{n}, \end{aligned}$$

and \(|f^{(3)}_n(z_i)|\le 2|z_i|(i=1,2),\) then \(f_n(z)=0\Rightarrow |f^{(3)}_n(z)|\le 2|z|\). However \(\mathcal F\) is not normal at 0 since \(f_n(1/n)=0\) and \(f_n(0)=\infty \).

The next example shows that condition (c) cannot be omitted in Theorem 1.

Example 4

Let k be a positive integer, \(\Delta =\{z: |z|<1\}\) and

$$\begin{aligned} \mathcal {F}=\left\{ f_{n}(z)=\frac{1}{(k+1)!}\frac{(z-1/n)^{k+2}}{z-(k+2)/n}\right\} . \end{aligned}$$

Clearly, the zero of \(f_{n}\) is of multiplicity \(k+2\), so that \(f(z)=0\Rightarrow |f^{(k)}(z)|\le |z|\); the pole of \(f_{n}\) is simple. On the other hand, since

$$\begin{aligned} f_{n}(z)=\frac{1}{(k+1)!}\left( z^{k+1}+P_{k-1}(z)+\frac{a}{z-(k+2)/n}\right) , \end{aligned}$$

where \(P_{k-1}(z)\) is a polynomial of degree \(k-1\) and a is a nonzero constant, we have \(f_{n}^{(k)}(z)\ne z\). But \(\mathcal {F}\) is not normal at 0 since \(f_n(1/n)=0\) and \(f_n((k+2)/n)=\infty \).

In this paper, we write \(\Delta =\{z: |z|<1\}\) and \(\Delta '=\{z: 0<|z|<1\}\). For \(z_0\in \mathbb {C}\) and \(r>0\), we write \(\Delta (z_0,r)=\{z: |z-z_0|<r\}\), and \(\Delta '(z_0,r)=\{z: 0<|z-z_0|<r\}\).

2 Preliminary results

To prove our results, we need the following lemmas.

Lemma 1

[4, Lemma 2] Let k be a positive integer and let \(\mathcal {F}\) be a family of meromorphic functions in a domain D, all of whose zeros have multiplicity at least k, and suppose that there exists \(A\ge 1\) such that \( |f^{(k)}(z) |\le A \) whenever \(f(z)=0, f\in \mathcal {F}\). If \(\mathcal {F}\) is not normal at \(z_{0} \in D\), then for each \(\alpha \), \(0\le \alpha \le k\), there exist a sequence of complex numbers \(z_{n} \in D\), \(z_{n}\rightarrow z_{0}\), a sequence of positive numbers \(\rho _{n} \rightarrow 0\), and a sequence of functions \(f_{n} \in \mathcal {F}\) such that

$$\begin{aligned}g_{n}(\zeta )=\frac{f_{n}(z_{n}+\rho _{n}\zeta )}{\rho _{n}^{\alpha }}\rightarrow g(\zeta )\end{aligned}$$

locally uniformly with respect to the spherical metric, where g is a nonconstant meromorphic function on \(\mathbb {C}\), all of whose zeros have multiplicity at least k , such that \( g^{\#}(\zeta ) \le g^{\#}(0)= kA+1 \). Moreover, \(g(\zeta )\) has order at most 2.

Here, as usual, \(g^{\#}(\zeta )=|g'(\zeta )|/(1+|g(\zeta )|^{2})\) is the spherical derivative.

Lemma 2

[11, Lemma 5] Let f be a transcendental meromorphic function, \(k(\ge 2), \ell \) be positive integers. If f has only zeros of order at least 3, then \(f^{(k)}-z^{\ell }\) has infinitely many zeros.

The next is a generalization of Hayman inequality, which is due to Yang [12].

Lemma 3

Let f be a transcendental meromorphic function, \(\varphi \) be a small meromorphic function of f, and \(k\in \mathbb {N}\). Then

$$\begin{aligned} T(r,f) \le 3N\left( r,\frac{1}{f}\right) +4N\left( r,\frac{1}{f^{(k)}-\varphi }\right) +S(r,f). \end{aligned}$$

Lemma 4

[1, Corollary 2] Let f be meromorphic in \(\mathbb {C}\) and of finite order \(\rho \) and E be the set of its critical values. If f has at most \(2\rho +card E'\) asymptotic values, where \(E'\) is the derived set of E.

Lemma 5

[7, Lemma 2.2] Let f be meromorphic in \(\mathbb {C}\) and suppose that the set of all finite critical and asymptotic values of f is bounded. Then there exists \(R>0\) such that if \(|z|>R\) and \(|f(z)|>R\), then

$$\begin{aligned} |f'(z)|\ge \frac{|f(z)|\log |f(z)|}{16 \pi |z|}. \end{aligned}$$

Lemma 6

Let f be a transcendental meromorphic function of finite order \(\rho \), and let \(k(\ge 2)\) be a positive integer. If f has only zeros of multiplicity at least k, and there exists \(A>1\) such that \(f(z)=0\Rightarrow |f^{(k)}(z)|\le A|z|\), then \(f^{(k)}\) has infinitely many fix-points.

Proof

Suppose that \(f^{(k)}\) has finitely many fix-points. Lemma 3 implies that f has infinitely many zeros, say \(z_n\)(\(n=1,2,\ldots \)). Clearly, \(z_n\rightarrow \infty \). Now set

$$\begin{aligned} g(z)=\frac{z^2}{2}-f^{(k-1)}(z). \end{aligned}$$

Then g is also of finite order \(\rho \), and \(g'(z)=z-f^{(k)}(z)\) has only finitely many zeros. By Lemma 4 or Denjoy-Carleman-Ahlfors’ theorem, g has at most \(2\rho \) asymptotic values, and then satisfies the hypotheses of Lemma 5 for some \(R>0\). It follows that

$$\begin{aligned} \frac{|z_ng'(z_n)|}{|g(z_n)|}\ge \frac{\log |g(z_n)|}{16\pi } \end{aligned}$$

for large n. Since \(g(z_n)=z_n^{2}/2\) and \(|g'(z_n)|=|z_n-f^{(k)}(z_n)|\le (A+1)|z_n|\), we have

$$\begin{aligned} 2(A+1)\ge \frac{1}{16\pi }[2\log |z_n|-\log 2]\rightarrow \infty \end{aligned}$$

as \(n\rightarrow \infty ,\) a contradiction. Lemma 6 is proved. \(\square \)

Lemma 7

[10, Lemma 5] Let f be meromorphic in \(\mathbb {C}\) and of finite order, and let \(k\ge 2\) be a positive integer and K be a positive number. Suppose that f has only zeros of multiplicity at least k, \(|f^{(k)}(z)|<K\) whenever \(f(z)=0\), and \(f^{(k)}(z)\ne 1\). Then one of the following two cases must occur:

  1. (1)
    $$\begin{aligned} f(z)=\alpha (z-\beta )^k, \end{aligned}$$
    (1)

    where \(\alpha , \beta \in \mathbb {C}\), and \( \alpha \cdot k!\ne 1.\)

  2. (2)

    If \(k=2\), then

    $$\begin{aligned} f(z)=\frac{(z-c_1)^2(z-c_2)^2}{2(z-c)^2}, \end{aligned}$$
    (2)

    or

    $$\begin{aligned} f(z)=\frac{(z-c_1)^3}{2(z-c)}. \end{aligned}$$
    (3)

    If \(k\ge 3\), then

    $$\begin{aligned} f(z)=\frac{1}{k!}\frac{(z-c_1)^{k+1}}{(z-c)}. \end{aligned}$$
    (4)

    Here \(c_1,c_2,c\) are distinct complex numbers.

Lemma 8

[9, Lemma 8] Let f be a non-polynomial rational function and k be a positive integer. If \(f^{(k)}(z)\ne 1\), then

$$\begin{aligned} f(z)=\frac{1}{k!}z^k+a_{k-1}z^{k-1}+\cdots +a_0+\frac{a}{(z-b)^m}, \end{aligned}$$

where \(a_{k-1},\ldots , a_0, a(\ne 0) ,b\) are constants and m is a positive integer.

Lemma 9

Let \(k(\ge 2)\) be a positive integer, and f be a rational function, all of whose zeros are of multiplicity at least k. If \(f^{(k)}(z)\ne z\), then one of the following three cases must occur:

  1. (1)
    $$\begin{aligned} f(z)=\frac{(z+c)^{k+1}}{(k+1)!}; \end{aligned}$$
    (5)
  2. (2)
    $$\begin{aligned} f(z)=\frac{(z-c_1)^{k+2}}{(k+1)!(z-b)}; \end{aligned}$$
    (6)
  3. (3)
    $$\begin{aligned}&f(z)=\frac{(z-c_1)^2(z-c_2)^3}{6(z-b)^2} \ (for\ k=2),\end{aligned}$$
    (7)
    $$\begin{aligned}&f(z)=\frac{(z-c_1)^3(z-c_2)^3}{24(z-b)^2} \ (for\ k=3), \end{aligned}$$
    (8)

where c is nonzero constant, and \(c_1, c_2\) and b are distinct constants.

Proof

Suppose first that f is a polynomial. Then \(f^{(k)}(z)=z+c\), where \(c(\ne 0)\) is a constant, so that

$$\begin{aligned} f^{(k-1)}(z)=\frac{z^2}{2}+cz+d \end{aligned}$$

where d is a constant. If f vanishes at \(z_0\), then \(f^{(k-1)}(z_0)=z_0^2/2+cz_0+d=0\) since f has only zeros of multiplicity at least k. It follows that f has at most two zeros. So f has either only one zero of multiplicity \(k+1\) or two distinct zeros of multiplicity exactly k. If f has two distinct zeros of multiplicity exactly k, then \(\deg f=2k\) and \(\deg f^{(k)}=k\), which contradicts the fact that \(f^{(k)}(z)=z+c\) and \(k\ge 2\). Thus, f has only one zero of multiplicity \(k+1\), and hence f has the form (5).

Suppose then that f is a nonpolynomial rational function. Set

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

Then \(g^{(k)}(z)\ne 1\), so by Lemma 8

$$\begin{aligned} g(z)=\frac{1}{k!}z^k+a_{k-1}z^{k-1}+\cdots +a_0+\frac{a}{(z-b)^m}, \end{aligned}$$

where \(a_{k-1},\ldots , a_0, a(\ne 0) ,b\) are constants and m is a positive integer. Thus

$$\begin{aligned} f(z)=p(z)+\frac{a}{(z-b)^m}=\frac{p(z)(z-b)^m+a}{(z-b)^m}, \end{aligned}$$
(9)

where

$$\begin{aligned} p(z)=\frac{1}{(k+1)!}z^{k+1}+a_{k-1}z^{k-1}+\cdots +a_0. \end{aligned}$$

Let \(c_1, c_2,\cdots , c_q \) be q distinct zeros of \(p(z)(z-b)^m+a\), with multiplicity \(n_1, n_2, \cdots , n_q\). Clearly, \(n_i\ge k\), \(c_i\ne b\), and \(c_i\) is a zero of \((p(z)(z-b)^m+a)'\) with multiplicity \(n_i-1\ge k-1(1\le i\le q)\). Since

$$\begin{aligned} \left( p(z)(z-b)^m+a\right) '=(z-b)^{m-1}\left( p'(z)(z-b)+mp(z)\right) , \end{aligned}$$
(10)

then \(c_i\) must be a zero of \(p'(z)(z-b)+mp(z)\) with multiplicity \(n_i-1(\ge k-1)\). Note that \(\deg [p'(z)(z-b)+mp(z)]=k+1\). Now we divide into three cases.

Case 1. \(k=2.\)

Then \(\deg [p'(z)(z-b)+mp(z)]=3\), and hence

  1. (a)

    \(p'(z)(z-b)+mp(z)\) has three simple zeros \(c_1, c_2\), and \(c_3\); or

  2. (b)

    \(p'(z)(z-b)+mp(z)\) has one simple zero \(c_1\) and one zero \(c_2\) with multiplicity 2; or

  3. (c)

    \(p'(z)(z-b)+mp(z)\) has only one zero \(c_1\) with multiplicity 3.

For case (a), we deduce that \(m=3\), and

$$\begin{aligned} p'(z)(z-b)+3p(z)=(z-c_1)(z-c_2)(z-c_3), \end{aligned}$$
$$\begin{aligned} p(z)(z-b)^3+a=\frac{1}{6}(z-c_1)^2(z-c_2)^2(z-c_3)^2. \end{aligned}$$

These, together with (10) give

$$\begin{aligned} (z-b)^2= \frac{1}{3}[(z-c_1)(z-c_2)+(z-c_1)(z-c_3)+(z-c_2)(z-c_3)]. \end{aligned}$$

Equating coefficients, we have \(b=(c_1+c_2+c_3)/3\) and \( b^2=(c_1c_2+c_1c_3+c_2c_3)/3\), so that

$$\begin{aligned} c_1^2+c_2^2+c_2^2=c_1c_2+c_1c_3+c_2c_3, \end{aligned}$$

that is,

$$\begin{aligned} (c_1-c_2)^2+(c_1-c_3)^2+(c_2-c_3)^2=0, \end{aligned}$$

and hence \(c_1=c_2=c_3\), a contradiction. Thus case (1) is ruled out.

For case (b), we deduce that \(m=2\) and

$$\begin{aligned} p(z)(z-b)^2+a=\frac{1}{6}(z-c_1)^2(z-c_2)^3. \end{aligned}$$

Then, by (9), f has the form (7).

For case (c), we can deduce that \(m=1\) and

$$\begin{aligned} p(z)(z-b)+a=\frac{1}{6}(z-c_1)^4, \end{aligned}$$

This, together with (9), gives that f has the form (6).

Case 2. \(k=3\).

Since \(\deg [p'(z)(z-b)+mp(z)]=4\), \(p'(z)(z-b)+mp(z)\) has two zeros \(c_1, c_2\) with multiplicity 2 or one zero \(c_1\) with multiplicity 4. It follows that \(m=2\) and

$$\begin{aligned} p(z)(z-b)^2+a=\frac{1}{24}(z-c_1)^3(z-c_2)^3 \end{aligned}$$

or \(m=1\) and

$$\begin{aligned} p(z)(z-b)+a=\frac{1}{24}(z-c_1)^5. \end{aligned}$$

Then, by (9), f has the form (6) or (8).

Case 3. \(k\ge 4\).

Noting that \(\deg [p'(z)(z-b)+mp(z)]=k+1\), we conclude that \(p'(z)(z-b)+mp(z)\) has only one zero \(c_1\) with multiplicity \(k+1\). In fact, if \(p'(z)(z-b)+mp(z)\) has at least two zeros \(c_1, c_2\) with multiplicity \(n_1, n_2,\ge k-1\), then \(2(k-1)\le k+1\), and thus \(k\le 3\), a contradiction. Thus \(m=1\) and \(p(z)(z-c)+b=\frac{1}{k!}(z-c_1)^{k+2}\), and hence f has the form (6). This completes the proof of Lemma 9. \(\square \)

Lemma 10

Let \(k\ge 3\) be a positive integer, \(A>1\) be a constant. Let \(\mathcal F\) be a family of meromorphic functions in a domain D. Suppose that, for every \(f\in \mathcal F\), f has only zeros of multiplicity at least k, and satisfies the following conditions:

  1. (a)

    \(f(z)=0\Rightarrow |f^{(k)}(z)|\le A|z|\).

  2. (b)

    \(f^{(k)}(z)\ne z\).

  3. (c)

    all poles of f are multiple.

Then \(\mathcal F\) is normal in \(D\backslash \{0\}\).

Proof

Suppose that \(\mathcal F\) is not normal at a point \(z_0\in D\backslash \{0\}\). Giving a small \(r>0\) such that \(\Delta (z_0,r)\subset D\backslash \{0\}\) and \(f(z)=0\Rightarrow |f^{(k)}(z)|\le A|z_0|+1\) for \(f\in \mathcal F\) and \(z\in \Delta (z_0,r)\). Then by Lemma 1, for \(\alpha =k\), there exist a sequence of functions \(f_n\in \mathcal F\), a sequence of complex numbers \(z_n\rightarrow z_0\) and a sequence of positive numbers \(\rho _n \rightarrow 0\), such that

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

converges sphericaly uniformly on compact subsets of \({\mathbb {C}}\), where g is a non-constant meromorphic functionon on \(\mathbb {C}\), all zeros of g have multiplicity at least k, and

$$\begin{aligned}g^{\#}(\zeta )\le g^{\#}(0)=k(A|z_0|+1)+1\end{aligned}$$

for all \(\zeta \in \mathbb {C}\). Moreover, , g is of finite order. By Hurwitz’s theorem, all poles of g are multiple.

We claim: (1) \(g=0\Rightarrow |g^{(k)}|\le A|z_0|\); (2) \(g^{(k)}(\zeta )\ne z_0\).

Let \(\zeta _0\) be a zero of \(g(\zeta )\). Then there exist \(\zeta _n, \zeta _n \rightarrow \zeta _0 \), such that \(g_n(\zeta _n)=\rho _n^{-k}f_n(z_n+\rho _n \zeta _n)=0\) for n sufficiently large. Thus \(f_n(z_n+\rho _n \zeta _n)=0\), so that \(|f^{(k)}_n(z_n+\rho _n \zeta _n)|\le A|z_n+\rho _n \zeta _n|\) for sufficiently large n. Since

$$\begin{aligned} g_n^{(k)}(\zeta _n)=f_n^{(k)}(z_n+\rho _n \zeta _n)\rightarrow g^{(k)}(\zeta _0), \end{aligned}$$

we have \(|g^{(k)}(\zeta _0)|\le A|z_0|\). We have proved (i).

Suppose that there exists \(\zeta _0\) such that \(g^{(k)}(\zeta _0)=z_0\). Since

$$\begin{aligned} 0\ne f^{(k)}_n(z_n+\rho _n \zeta )-(z_n+\rho _n \zeta )=g^{(k)}_n(\zeta )-(z_n+\rho _n \zeta )\rightarrow g^{(k)}(\zeta )-z_0, \end{aligned}$$

Hurwitz’s theorem implies that \(g^{(k)}(\zeta )\equiv z_0\). Note that g has only zeros of multiplicity at least k, we have

$$\begin{aligned} g(\zeta )=\frac{z_0}{k!}(z-\alpha )^k,\quad \alpha \in \mathbb {C}. \end{aligned}$$

A simple calculation shows that

$$\begin{aligned} g^{\#}(0)\le \left\{ \begin{array}{ll} k/2 \ &{} \quad \mathrm {if}\ |\alpha |\ge 1; \\ |z_0|\ &{} \quad \mathrm {if}\ |\alpha |<1. \end{array} \right. \end{aligned}$$

But this contradicts \(g^{\#}(0)=k(A|z_0|+1)+1\), and thus (2) is proved.

By Lemma 7, g has the form (1) or (4) in Lemma 7. Similarly as above, we exclude the case that g has the form (1), so that g has the form (4). But g has only multiple poles, a contradiction. This completes the proof of Lemma 10. \(\square \)

Lemma 11

Let \(\mathcal F\) be a family of meromorphic functions in a domain D, \(A>1\) be a constant. Suppose that, for every \(f\in \mathcal F\), f has only zeros of multiplicity at least k, and satisfies the following conditions:

  1. (a)

    \(f(z)=0\Rightarrow |f''(z)|\le A|z|\).

  2. (b)

    \(f''(z)\ne z\).

  3. (c)

    all poles of f are of multiplicity at least 3.

Then \(\mathcal F\) is normal in \(D\backslash \{0\}\).

This lemma can be proved almost the same as Lemma 10. We omit the details here.

3 Proof of Theorems 1 and 2

Proof of Theorem 1

Since normality is a local property, by Lemma 10, we only need to prove that \(\mathcal {F}\) is normal at \(z=0\). Without loss of generality, we may assume \(D=\Delta \). Suppose, on the contrary, \(\mathcal {F}\) is not normal at the origin. Our goal is to obtain a contradiction in the sequel.

Consider the family

$$\begin{aligned} \mathcal {G}=\left\{ g(z)=\frac{f(z)}{z}:f\in \mathcal {F}\right\} . \end{aligned}$$

We claim that \(f(0)\ne 0\) for every \(f\in \mathcal {F}\). Otherwise, if \(f(0)=0\), by the assumption of Theorem 1, \(|f^{(k)}(0)|\le 0\), and then \(f^{(k)}(0)=0\). But \(f^{(k)}(z)\ne z\), a contradiction. Thus, for each \(g\in \mathcal {G}\), \(g(0)=\infty \). Furthermore, all zeros of g(z) have multiplicity at least k. On the other hand, by simple calculation, we have

$$\begin{aligned} g^{(k)}(z)=\frac{f^{(k)}(z)}{z}-\frac{kg^{(k-1)}(z)}{z}. \end{aligned}$$
(11)

Since \(f(z)=0\Rightarrow |f^{(k)}(z)|\le A|z|\), we deduce that \(g(z)=0\Rightarrow |g^{(k)}(z)|\le A\).

We first prove that \(\mathcal {G}\) is normal at 0. Suppose not; by Lemma 1, there exist functions \(g_{n}\in \mathcal {G}\), points \(z_{n}\rightarrow 0\) and positive numbers \(\rho _{n}\rightarrow 0\) such that

$$\begin{aligned} G_{n}(\zeta )=\frac{g_{n}(z_{n}+\rho _{n}\zeta )}{\rho _{n}^k}\rightarrow G(\zeta ), \end{aligned}$$
(12)

converges spherically uniformly on compact subsets of \({\mathbb {C}}\), where G is a non-constant meromorphic functionon on \(\mathbb {C}\) and of finite order, all zeros of G have multiplicity at least k, and \(G^{\#}(\zeta )\le G^{\#}(0)=kA+1\) for all \(\zeta \in \mathbb {C}\).

We distinguish two cases.

Case 1.   \(z_{n}{/}\rho _{n}\rightarrow \infty \). Since \(G_{n}(-z_{n}{/}\rho _{n})=g_{n}(0){/}\rho _{n}^k\), the pole of \(G_n\) corresponding to that of \(g_n\) at 0 drifts to infity. Then, by Hurwitz’s theorem, G has only mutiple poles. By (11) and (12), we have

$$\begin{aligned} G_n^{(k)}(\zeta )= & {} g_n^{(k)}(z_n+\rho _n\zeta )\\= & {} \frac{f_n^{(k)}(z_n+\rho _n\zeta )}{z_n+\rho _n\zeta }- k\frac{g_n^{(k-1)}(z_n+\rho _n\zeta )}{\rho _n}\frac{\rho _n}{z_n+\rho _n\zeta }.\\ \end{aligned}$$

Noting that

$$\begin{aligned} \frac{\rho _n}{z_n+\rho _n\zeta }\rightarrow 0 \end{aligned}$$

uniformly on compact subsets of \(\mathbb {C}\), and \(g_n^{(k-1)}(z_n+\rho _n\zeta ){/}\rho _n\) is locally bounded on \(\mathbb {C}\backslash G^{-1}(\infty )\) since \(g_{n}(z_{n}+\rho _{n}\zeta ){/}\rho _{n}^k\rightarrow G(\zeta )\). Thus

$$\begin{aligned} \frac{f_n^{(k)}(z_n+\rho _n\zeta )}{z_n+\rho _n\zeta }\rightarrow G^{(k)}(\zeta ), \end{aligned}$$
(13)

uniformly on compact subsets of \(\mathbb {C}{\backslash } G^{-1}(\infty )\).

Claim: (I) \(G(\zeta )=0\Rightarrow |G^{(k)}(\zeta )|\le A\); (II) \( G^{(k)}(\zeta )\ne 1\).

Indeed, if \(G(\zeta _0)=0\), Hurwitz’s theorem and (12) imply that there exist \(\zeta _n, \zeta _n \rightarrow \zeta _0 \), such that \(g_n(z_n+\rho _n \zeta _n)=0\), and then \(f_n(z_n+\rho _n \zeta _n)=0\) for n sufficiently large. By assumption, \(|f_n^{(k)}(z_n+\rho _n\zeta _n)|\le A|z_n+\rho _n\zeta _n|\). It follows from (13) that \(|G^{(k)}(\zeta _0)|\le A\). Claim (I) is proved.

Since \(f_n^{(k)}(z)\ne z\), Hurwitz’s theorem and (13) yield that either \(G^{(k)}(\zeta )\ne 1\) or \(G^{(k)}(\zeta )\equiv 1\) for any \(\zeta \in \mathbb {C}\backslash G^{-1}(\infty )\). Clearly, these also hold for all \(\zeta \in \mathbb {C}\). If \(G^{(k)}(\zeta )\equiv 1\), noting that all zeros of G have multiplicity at least k, we have \(G(\zeta )=(\zeta -\alpha )^k{/}k!\)(\(\alpha \in \mathbb {C}\)). As in the proof of Lemma 10,

$$\begin{aligned} G^{\#}(0)\le \left\{ \begin{array}{ll} k/2 \ &{} \quad \mathrm {if}\ |\alpha |\ge 1; \\ 1\ &{} \quad \mathrm {if}\ |\alpha |<1. \end{array} \right. \end{aligned}$$

which contradicts \(G^{\#}(0)=kA+1\). Then Claim (II) is proved. Then by Lemma 7, G has the form (1) or (4) in Lemma 7. The form (1) can be ruled out similarly as above. Thus

$$\begin{aligned} G(\zeta )=\frac{1}{k!}\frac{(\zeta -c_1)^{k+1}}{(\zeta -c)}, \end{aligned}$$

where \(c_1, c\) are distinct complex numbers. But, this contradicts that G has only mutiple poles.

Case 2.   

figure a

. Taking subsequence, we can assume that \(z_{n}/\rho _{n}\rightarrow \alpha ,\) a finite complex number. Then

$$\begin{aligned} \frac{g_{n}(\rho _{n}\zeta )}{\rho _{n}^{k}}=G_{n}(\zeta -z_{n}/\rho _{n})\overset{\chi }{\rightarrow } G(\zeta -\alpha )=\widetilde{G}(\zeta ) \end{aligned}$$

on \(\mathbb {C}\). Clearly, all zeros of \(\widetilde{G}\) have multiplicity at least k, and all poles of \(\widetilde{G}\) are multiple, except possibly the pole at 0.

Set

$$\begin{aligned} H_{n}(\zeta )=\frac{f_{n}(\rho _{n}\zeta )}{\rho _{n}^{k+1}}. \end{aligned}$$
(14)

Then

$$\begin{aligned} H_{n}(\zeta )=\frac{f_{n}(\rho _{n}\zeta )}{\rho _{n}^{k+1}}=\zeta \frac{g_{n}(\rho _{n}\zeta )}{\rho _{n}^{k}}\rightarrow \zeta \widetilde{G}(\zeta )=H(\zeta ) \end{aligned}$$
(15)

spherically uniformly on compact subsets of \({\mathbb {C}}\), and

$$\begin{aligned} H_{n}^{(k)}(\zeta )=\frac{f_{n}^{(k)}(\rho _{n}\zeta )}{\rho _{n}} \rightarrow H^{(k)}(\zeta ) \end{aligned}$$
(16)

locally uniformly on \(\mathbb {C}\setminus H^{-1}(\infty )\). Obviously, all zeros of H have multiplicity at least k, and all poles of H are multiple. Since \(\widetilde{G}(0)=\infty \), \(H(0)\ne 0\).

Claim: (III) \(H(\zeta )=0\Rightarrow |H^{(k)}(\zeta )|\le A|\zeta |\); (IV) \( H^{(k)}(\zeta )\ne \zeta \).

If \(H(\zeta _0)=0\), by Hurwitz’s theorem and (15), there exist \(\zeta _n \rightarrow \zeta _0 \) such that \(f_n(\rho _n \zeta _n)=0\) for for n sufficiently large. By the assumption, \(|f_n^{(k)}(\rho _n\zeta _n)|\le A|\rho _n\zeta _n|\). Then, it follows from (16) that \(|H^{(k)}(\zeta _0)|\le A|\zeta _0|\). Claim (III) is proved.

Suppose that there exists \(\zeta _0\) such that \(H^{(k)}(\zeta _0)=\zeta _0\). By (16),

$$\begin{aligned} 0\ne \frac{f_{n}^{(k)}(\rho _{n}\zeta )-\rho _{n}\zeta }{\rho _{n}}=H_{n}^{(k)}(\zeta )-\zeta \rightarrow H^{(k)}(\zeta )-\zeta , \end{aligned}$$

uniformly on compact subsets of \(\mathbb {C}{\setminus } H^{-1}(\infty )\). Hurwitz’s theorem implies that \(H^{(k)}(\zeta )\equiv \zeta \) on \(\mathbb {C}{\setminus } H^{-1}(\infty )\), and then on \(\mathbb {C}\). It follows that H is a polynomial of degree \(k+1\). Since all zeros of H have multiplicity at least k, and noting that \(k\ge 4\), we know that H has a single zero \(\zeta _1\) with multiplicity \(k+1\), so that \(H^{(k)}(\zeta _1)=0\), and hence \(\zeta _1=0\) since \(H^{(k)}(\zeta )\equiv \zeta \). But \(H(0)\ne 0\), we arrive at a contradiction. This proves claim (IV).

Then, by Lemma 6, H must be a rational function, and thus Lemma 9 implies that H has the form (5) or (6) in Lemma 9. The form (6) can be excluded since all poles of H are multiple. Thus we have

$$\begin{aligned} H(\zeta )=\frac{(\zeta +c)^{k+1}}{(k+1)!} \end{aligned}$$
(17)

where \(c(\ne 0)\) is a constant.

Next we will show that (17) is impossible. Indeed, combining (15) and (17) gives

$$\begin{aligned} \frac{f_{n}(\rho _{n}\zeta )}{\rho _{n}^{k+1}}\rightarrow \frac{(\zeta +c)^{k+1}}{(k+1)!}. \end{aligned}$$
(18)

Note that all zeros of \(f_n\) have multiplicity at least k and \(k\ge 4\), there exist points \(\zeta _{n,0}\rightarrow -c\) such that \(z_{n,0}=\rho _{n}\zeta _{n,0}\) is a zero of \(f_{n}\) with multiplicity \(k+1\).

We now consider two subcases.

Case 2.1 There exists \(0<\delta \le 1\) such that the functions \(f_{n}(z)\) (for large n) are all holomorphic on \(\Delta (0,\delta )\).

Since \(\{f_{n}\}\) is normal on \(\Delta '(0,\delta )\), but not normal at 0, it follows from the maximum modulus principle that \(f_{n}\rightarrow \infty \) locally uniformly on \(\Delta '(0,\delta ).\)

Suppose that there exists \(0<\sigma <\delta \) such that each \(f_{n}\) has only one zero \(z_{n,0}\) in \(\Delta (0,\sigma ).\) Set

$$\begin{aligned} K_{n}(z)=\frac{f_{n}(z)}{(z-z_{n,0})^{k+1}}. \end{aligned}$$
(19)

Then \(\{K_{n}\}\) is a sequence of nonvanishing holomorphic functions on \(\Delta (0,\sigma )\), and \(K_{n}(z)\rightarrow \infty \) locally uniformly on \(\Delta '(0,\sigma )\). It follows that \(\{1{/}K_{n}\}\) is holomorphic on \(\Delta (0,\sigma )\), and \(1{/}K_{n}(z)\rightarrow 0\) locally uniformly on \(\Delta '(0,\sigma )\), and hence on \(\Delta (0,\sigma )\) by the maximum modulus principle. So \(K_{n}(z)\rightarrow \infty \) locally uniformly on \(\Delta (0,\sigma )\). In particular, \(K_{n}(2z_{n,0})\rightarrow \infty .\) But, by (18) and (19),

$$\begin{aligned} K_{n}(2z_{n,0})=\frac{f_{n}(2z_{n,0})}{z_{n,0}^{k+1}} =\frac{f_{n}(2\rho _{n}\zeta _{n,0})}{\rho _{n}^{k+1}\zeta _{n,0}^{k+1}}\rightarrow \frac{1}{(k+1)!}, \end{aligned}$$

a contradiction.

Hence, taking a subsequence if necessary, for any \(0<\sigma <\delta \), \(f_{n}\) has at least two distinct zeros in \(\Delta (0,\sigma )\) for sufficiently large n. We assume that \(z_{n,1}\) is a zero of \(f_{n}\) on \(\Delta (0,\sigma )\backslash \{z_{n,0}\}\). Clearly, \(z_{n,1}\rightarrow 0\). Let \(\zeta _{n,1}=z_{n,1}/\rho _{n}\), it follows froms (18) that \(\zeta _{n,1}\rightarrow \infty \). Hence \(z_{n,0}/z_{n,1}=\zeta _{n,0}/\zeta _{n,1}\rightarrow 0\). Set

$$\begin{aligned} L_{n}(z)=\frac{f_{n}(z_{n,1}z)}{z_{n,1}^{k+1}}. \end{aligned}$$

Then, for sufficiently large n, \(\{L_{n}\}\) is well-defined and holomorphic on each bounded set of \(\mathbb {C}\), and all of whose zeros have multiplicity at least k. By the assumption, we have \(L_{n}(z)=0\Rightarrow |L_{n}^{(k)}(z)|\le A|z|\), and \(L_{n}^{(k)}(z)\ne z\). By Lemma 10, \(\{L_{n}\}\) is normal on the punctured complex plane \(\mathbb {C}^{*}=\mathbb {C}\backslash \{0\}.\) We claim that \(\{L_{n}\}\) is also normal at 0. Otherwise, the maximum modulus principle implies that \(L_{n}\rightarrow \infty \) locally uniformly on \(\mathbb {C}^{*}\). But, this is impossible since \(L_{n}(1)=0\). Hence \(\{L_{n}\}\) is normal on the whole plane \(\mathbb {C}.\)

Taking a subsequence and renumbering, we assume that

$$\begin{aligned} L_{n}(z)\rightarrow L(z), \end{aligned}$$

and then

$$\begin{aligned} L_{n}^{(k)}(z)\rightarrow L^{(k)}(z) \end{aligned}$$
(20)

locally uniformly on \(\mathbb {C}\), where L is entire, all zeros of L have multiplicity at least k. Clearly, \(L(1)=0\). On the other hand, \(L_{n}(z_{n,0}/z_{n,1})=0\) and \(z_{n,0}/z_{n,1}\rightarrow 0\), we get that \(L(0)=0\). Since \(L_{n}(z)=0\Rightarrow |L_{n}^{(k)}(z)|\le A|z|\), an argument similar to that in Claim III yields that \(L(z)=0\Rightarrow |L^{(k)}(z)|\le |z|\). So it follows from \(L(0)=0\) that \(L^{(k)}(0)=0\). Since \(L_{n}^{(k)}(z)\ne z\), Hurwitz’s theorem and (20) imply that \(L^{(k)}(z)\equiv z\). Note that all zeros of L have multiplicity at least k and \(L(0)=0\), we deduce that \(L(z)=z^{k+1}/(k+1)!\). But, this in impossible since \(L(1)=0\).

Case 2.2 By taking a subsequence, if necessary, for any \(\delta >0\), \(f_{n}\) has at least one pole on \(\Delta (0,\delta )\) for all n.

Then there exist points \(z_{n,\infty }\rightarrow 0\) such that \(f_n(z_{n,\infty })=\infty \). We may assume that \(z_{n,\infty }\) is the pole of \(f_{n}\) of smallest modulus. Let \(\zeta _{n,\infty }=z_{n,\infty }{/}\rho _{n}\). It follows from (18) that \(\zeta _{n,\infty }\rightarrow \infty \), and then \(z_{n,0}/z_{n,\infty }=\zeta _{n,0}/\zeta _{n,\infty }\rightarrow 0\). Now set

$$\begin{aligned} M_{n}(z)=\frac{f_{n}(z_{n,\infty }z)}{z_{n,\infty }^{k+1}}. \end{aligned}$$

Then, for sufficiently large n, \(\{M_{n}\}\) is well-defined for each \(z\in \mathbb {C}\), all of whose zeros have multiplicity at least k and whose poles are are multiple. Moreover, \(\{M_{n}\}\) is holomorphic on \(\Delta \) for sufficiently large n. By the assumption, we have \(M_{n}(z)=0\Rightarrow |M_{n}^{(k)}(z)|\le A|z|\), and \(M_{n}^{(k)}(z)\ne z\). Lemma 10 implies that \(\{M_{n}\}\) is normal on \(\mathbb {C}^{*}.\) We claim that \(\{M_{n}\}\) is also normal at 0. Otherwise, \(\{M_{n}\}\) is normal on \(\Delta '\), but not normal at 0. Since \(\{M_{n}\}\) is holomorphic on \(\Delta \), he maximum modulus principle implies that \(M_n\rightarrow \infty \). But \(M_{n}(z_{n,0}/z_{n,\infty })=0\) and \(z_{n,0}/z_{n,\infty }\rightarrow 0\). This contradiction proves our claim. Hence, \(\{M_{n}\}\) is normal on \(\mathbb {C}.\)

Then, taking a subsequence and renumbering,

$$\begin{aligned} M_{n}(z)\rightarrow M(z) \end{aligned}$$

spherically uniformly on compact subsets of \({\mathbb {C}}\), where M is meromorphic, all of whose zeros have multiplicity at least k. Clearly, \(M(1)=\infty \). On the other hand, \(M_{n}(z_{n,0}/z_{n,\infty })=0\) and \(z_{n,0}/z_{n,\infty }\rightarrow 0\), we obtain \(M(0)=0\). Arguing as in Case 2.1 (for L(z)), we have \(M(z)=z^{k+1}/(k+1)!\). But, \(M(1)=\infty \), a contradiction. Then we have shown that (17) is impossible.

We thus have proved that \(\mathcal G\) is normal at 0.

We now turn to show that is normal at \(z=0\). Since \(\mathcal G\) is normal at 0, then the family \(\mathcal G\) is equicontinuous at 0 with respect to the spherical distance. On the other hand, \(g(0)=\infty \) for each \(g\in \mathcal G\), so there exists \(\delta >0\) such that \(|g(z)|\ge 1\) for all \(g\in \mathcal G\) and each \(z\in \Delta (0,\delta )\). It follows that \(f(z)\ne 0\) for all \(f\in \mathcal F\) and \(z\in \Delta (0,\delta )\). Since \(\mathcal F\) is normal on \(\Delta '\) but not normal at \(z=0\), the family \(1/\mathcal{F}={\{}1/f: f\in \mathcal{F}{\}}\) is holomorphic in \(D_{\delta }\) and normal on \(\Delta '(0,\delta )\), but not normal at \(z=0\). Thus there exists a sequence \(\{ 1/f_n \}\subset 1/\mathcal F\) which converges locally uniformly in \(\Delta '(0,\delta )\), but not on \(\Delta (0,\delta )\). The maximum modulus principle implies that \(1{/}f_n\rightarrow \infty \) in \(\Delta '(0,\delta )\). Thus \(f_n\rightarrow 0\) converges locally uniformly in \(\Delta '(0,\delta )\), and hence so does \(\{ g_n \}\subset \mathcal G\), where \(g_n(z)=f_n(z)/z\). But \(|g_n(z)|\ge 1\) for \(z\in \Delta (0,\delta )\), a contradiction. This completes the proof of Theorem 1. \(\square \)

Proof of Theorem 2

Using the same argument as in the proof of Theorem 1, we can prove Theorem 2. We here omit the details. \(\square \)