Hostname: page-component-745bb68f8f-s22k5 Total loading time: 0 Render date: 2025-02-11T13:59:33.033Z Has data issue: false hasContentIssue false

Non-normality, topological transitivity and expanding families

Published online by Cambridge University Press:  14 December 2021

THIERRY MEYRATH
Affiliation:
University of Luxembourg, Department of Mathematics, L-4364 Esch-sur-Alzette, Luxembourg. e-mail: thierry.meyrath@uni.lu
JÜRGEN MÜLLER
Affiliation:
Universität Trier, Fachbereich IV – Mathematik, D-54286 Trier, Germany. e-mail: jmueller@uni-trier.de
Rights & Permissions [Opens in a new window]

Abstract

We investigate the behaviour of families of meromorphic functions in the neighbourhood of points of non-normality and prove certain covering properties that complement Montel’s Theorem. In particular, we also obtain characterisations of non-normality in terms of such properties.

Type
Research Article
Copyright
© The Author(s), 2021. Published by Cambridge University Press on behalf of Cambridge Philosophical Society

1. Introduction

For an open set $\Omega \subset \mathbb C$ we denote by $M(\Omega)$ the set of meromorphic functions on $\Omega$ , by which we mean all functions whose restriction to a connected component of $\Omega$ is either meromorphic or constant infinity. Endowed with the topology of spherically uniform convergence (i.e. uniform convergence with respect to the chordal metric $\chi$ ) on compact subsets of $\Omega$ , the space $M(\Omega)$ becomes a complete metric space (e.g. [Reference Conway12, Chapter VII]). As usual, we say that a family $\mathcal F \subset M(\Omega)$ is normal in $\Omega$ , if every sequence $(f_n) \subset \mathcal F$ contains a subsequence $(f_{n_k})$ that converges spherically uniformly on compact subsets of $\Omega$ to a function $f \in M(\Omega)$ . The family $\mathcal F$ is called normal at a point $z_0 \in \Omega$ , if there exists an open neighbourhood U of $z_0$ , such that $\mathcal F$ is normal in U. By $J(\mathcal F)$ we denote the set of points in $\Omega$ , at which the family $\mathcal F$ is non-normal. If $z_0 \in J(\mathcal F)$ , the family $\mathcal F$ can still have infinite subfamilies $\tilde{\mathcal F} \subset \mathcal F$ that are normal at $z_0$ , in other words, $z_0 \in J(\mathcal F)$ does in general not imply $z_0 \in J(\tilde{\mathcal F})$ . We say that $\mathcal F$ is strongly non-normal at a point $z_0 \in \Omega$ , if we have $z_0 \in J(\tilde{\mathcal F})$ for every infinite subfamily $\tilde{\mathcal F} \subset \mathcal F$ . We further say that $\mathcal F$ is strongly non-normal on a relatively closed set $B \subset \Omega$ , if $\mathcal F$ is strongly non-normal at every $z_0 \in B$ , that is if $B \subset J(\tilde{\mathcal F})$ for every infinite subfamily $\tilde{\mathcal F} \subset \mathcal F$ . Moreover, we call $\mathcal F$ hereditarily non-normal on B, if some infinite subfamily $\tilde{\mathcal F} \subset \mathcal F$ is strongly non-normal on B. Note that on a single point set, hereditary non-normality is equivalent to non-normality, while this is in general not true for sets containing at least two points.

For a family $\mathcal F \subset M(\Omega)$ and an open set $U \subset \Omega$ , we write $\limsup \mathcal F(U)$ for the intersection of all $\bigcup_{f \in \tilde{\mathcal F}} f(U)$ , where $\tilde{\mathcal F}$ ranges over the cofinite subsets of $\mathcal F$ . Moreover, for $z_0 \in \Omega$ we denote by $\limsup_{z_0} \mathcal F$ the intersection of $\limsup \mathcal F(U)$ taken over all open neighbourhoods $U \subset \Omega$ of $z_0$ . Similarly, we write $\liminf \mathcal F(U)$ for the union of all $\bigcap_{f \in \tilde{\mathcal F}} f(U)$ , where $\tilde{\mathcal F}$ ranges over the cofinite subsets of $\mathcal F$ and $\liminf_{z_0} \mathcal F$ for the intersection of $\liminf \mathcal F(U)$ taken over all open neighbourhoods $U \subset \Omega$ of $z_0$ . Obviously, we have that $\liminf_{z_0}\mathcal F \subset \limsup_{z_0}\mathcal F$ , furthermore $\liminf_{z_0}\mathcal F=\bigcap_{\tilde{\mathcal F} \subset \mathcal F \text{ infinite}} \limsup_{z_0} \tilde{\mathcal F}$ . For instance, if $\mathcal F = \{f_n\;:\; n \in \mathbb N\}$ with $f_n(z)=nz$ for even integers n and $f_n(z) = z$ for odd n, then $\limsup_0 \mathcal F = \mathbb C$ and $\liminf_0 \mathcal F = \{0\}$ .

The classical Montel Theorem suggests that the behaviour of families $\mathcal F \subset M(\Omega)$ in neighbourhoods of points $z_0 \in J(\mathcal F)$ consists in some sense of spreading points, since it asserts that for every $z_0 \in J(\mathcal F)$ , the set $E_{z_0}(\mathcal F) \;:\!=\; \mathbb C_\infty \setminus \limsup_{z_0} \mathcal F$ contains at most two points, where $\mathbb C_\infty \;:\!=\; \mathbb C \cup \{\infty\}$ . Hence, for every neighbourhood U of $z_0$ , every point $a \in \mathbb C_\infty$ is covered by f(U) for infinitely many $f \in \mathcal F$ , with at most two exceptions. In case that $E_{z_0}(\mathcal F)$ contains two points and $\mathcal F$ is strongly non-normal at $z_0$ , a further consequence of Montel’s Theorem is that $\liminf_{z_0}\mathcal F = \limsup_{z_0}\mathcal F$ , so that for every neighbourhood U of $z_0$ , every point $a \in \mathbb C_\infty \setminus E_{z_0}(\mathcal F)$ is covered by f(U) for cofinitely many $f \in \mathcal F$ . Note, however, that Montel’s Theorem does not contain any information about the ‘size’ of the individual sets f(U), for instance, if U is any neighbourhood of a point $z_0 \in J(\mathcal F)$ , it is in general not clear if for a given set $A \subset \limsup_{z_0}\mathcal F$ we have $A \subset f(U)$ for infinitely many $f \in \mathcal F$ .

In this paper, we will further investigate the behaviour of (strongly) non-normal families near points of non-normality and show certain covering and ‘expanding’ properties that complement the statement of Montel’s Theorem. In particular, we will also derive different characterisations of (strong) non-normality in terms of these properties.

2. Non-normality and topological transitivity

In the sequel, for $\lambda > 0$ and $z_0 \in \mathbb C$ we set $D_{\lambda}(z_0) \;:\!=\; \{z \in \mathbb C\;:\; \left|{z - z_0}\right| < \lambda\}$ and denote by $\overline{D}_{\lambda}(z_0)$ the closure of $D_{\lambda}(z_0)$ in $\mathbb C$ . For $w_0 \in \mathbb C_\infty$ , we further set $D_{\lambda}^{\chi}(w_0) \;:\!=\; \{w \in \mathbb C_\infty\;:\; \chi(w, w_0) < \lambda\}$ . We say that a family $\mathcal F \subset M(\Omega)$ is (topologically) transitive with respect to a point $z_0 \in \Omega$ , if for every pair of non-empty open sets $U \subset \Omega$ and $V \subset \mathbb C_\infty$ with $z_0 \in U$ , there exists $f \in \mathcal F$ such that $f(U) \cap V \neq \emptyset$ . Note that in this case we have $f(U) \cap V \neq \emptyset$ for infinitely many $f \in \mathcal F$ . If $f(U) \cap V \neq \emptyset$ holds for cofinitely many $f \in \mathcal F$ , we say that $\mathcal F$ is (topologically) mixing with respect to $z_0$ . Furthermore, if for every non-empty open set $U \subset \Omega$ with $z_0 \in U$ and every pair of non-empty open sets $V_1, V_2 \subset \mathbb C_\infty$ , there exists $f \in \mathcal F$ such that $f(U) \cap V_i \neq \emptyset$ for $i = 1,2$ , we say that $\mathcal F$ is weakly mixing with respect to $z_0$ . Finally, we say that $\mathcal F$ is transitive (or (weakly) mixing) with respect to a relatively closed set $B \subset \Omega$ , if $\mathcal F$ is transitive (or (weakly) mixing) with respect to every $z_0 \in B$ .

With these notations, we obtain the following characterisation of (strong) non-normality.

Theorem 1. Let $\Omega \subset \mathbb C$ be open, $\mathcal F \subset M(\Omega)$ a family of meromorphic functions and $z_0 \in \Omega$ . Then we have:

  1. (a) $\mathcal F$ is strongly non-normal at $z_0$ if and only if $\mathcal F$ is mixing with respect to $z_0$ ;

  2. (b) the following are equivalent:

    1. (i) $\mathcal F$ is non-normal at $z_0$ ;

    2. (ii) there exists an infinite subfamily $\tilde{\mathcal F} \subset \mathcal F$ that is mixing with respect to $z_0$ ;

    3. (iii) $\mathcal F$ is weakly mixing with respect to $z_0$ .

Proof. (a) Let $\mathcal F$ be strongly non-normal at $z_0$ and suppose that $\mathcal F$ is not mixing with respect to $z_0$ . Then there exist non-empty open sets $U \subset \Omega$ and $V \subset \mathbb C_\infty$ with $z_0 \in U$ , and an infinite subfamily $\tilde{\mathcal F} \subset \mathcal F$ such that $f(U) \cap V = \emptyset$ for every $f \in \tilde{\mathcal F}$ . By Montel’s Theorem, we obtain that $\tilde{\mathcal F}$ is normal on U, hence also at $z_0$ , in contradiction to the strong non-normality of $\mathcal F$ at $z_0$ .

On the other hand, suppose that $\mathcal F$ is mixing with respect to $z_0 \in \Omega$ , but not strongly non-normal at $z_0$ . Then there exists an open neighbourhood U of $z_0$ and a sequence $(f_n) \subset \mathcal F$ , such that $(f_n)$ converges spherically uniformly on compact subsets of U to a function $f \in M(U)$ . Hence, for $\varepsilon > 0$ sufficiently small, we have that $\overline{D}_{\varepsilon}(z_0) \subset U$ and there exists $\delta > 0$ and $w_0 \in \mathbb C_\infty$ such that $f(\overline{D}_{\varepsilon}(z_0)) \cap D_{\delta}^{\chi}(w_0) = \emptyset$ . Since $(f_n)$ is mixing with respect to $z_0$ , we obtain that $f_n(D_{\varepsilon}(z_0)) \cap D_{\frac{\delta}{2}}^{\chi}(w_0) \neq \emptyset$ for all n sufficiently large, in contradiction to the spherically uniform convergence of $(f_n)$ to f on $\overline{D}_{\varepsilon}(z_0)$ .

(b) $\textrm{(i)} \Rightarrow \textrm{(ii)}:$ Since $\mathcal F$ is non-normal at $z_0$ , there exists an infinite subfamily $\tilde{\mathcal F} \subset \mathcal F$ that is strongly non-normal at $z_0$ . This subfamily is mixing with respect to $z_0$ according to the first statement of the theorem.

$\textrm{(ii)} \Rightarrow \textrm{(iii)}$ : This is clear, since a mixing family is also weakly mixing.

$\textrm{(iii)} \Rightarrow \textrm{(i)}$ : Suppose that $\mathcal F$ is weakly mixing with respect to $z_0$ . Further consider two non-empty open sets $V_1, V_2 \subset \mathbb C_\infty$ such that $\inf_{z \in V_1, w \in V_2} \chi(z ,w) > \varepsilon$ for some $\varepsilon > 0$ . For $k \in \mathbb N$ , we set $U_k \;:\!=\; D_{\frac{1}{k}}(z_0) \cap \Omega$ . By assumption, for every $k \in \mathbb N$ there is a function $f_k \in \mathcal F$ such that $f_k(U_k) \cap V_1 \neq \emptyset$ and $f_k(U_k) \cap V_2 \neq \emptyset$ , and hence points $z_k^{(1)}, z_k^{(2)} \in U_k$ such that $f_k(z_k^{(1)}) \in V_1$ and $f_k(z_k^{(2)}) \in V_2$ . Note that $z_k^{(1)}, z_k^{(2)} \in U_k$ implies that $z_k^{(1)} \to z_0$ and $z_k^{(2)} \to z_0$ for $k \to \infty$ , furthermore we have that $\chi(f_k(z_k^{(1)}), f_k(z_k^{(2)})) > \varepsilon$ for every $k \in \mathbb N$ , and hence

\begin{equation*}\chi(f_k(z_0), f_k(z_k^{(1)})) > \frac{\varepsilon}{2} \quad \text{or} \quad \chi(f_k(z_0), f_k(z_k^{(2)})) > \frac{\varepsilon}{2}.\end{equation*}

Hence, we can find a sequence $(z_k)$ with $z_k \to z_0$ for $k \to \infty$ and $\chi(f_k(z_0), f_k(z_k)) > \frac{\varepsilon}{2}$ for every $k \in \mathbb N$ , implying that the family $\mathcal F$ is not spherically equicontinuous at $z_0$ , and thus also not normal.

By Montel’s Theorem, it is clear that $z_0 \in J(\mathcal F)$ implies that $\mathcal F$ is transitive with respect to $z_0$ . On the other hand, it is easily seen that transitivity of a family with respect to some point $z_0 \in \Omega$ is in general not sufficient for non-normality at $z_0$ . For instance, if $(z_n)$ is a sequence that is dense in $\mathbb C_\infty$ , the family $(f_n)$ of constant functions $f_n \equiv z_n$ is transitive with respect to any $z_0 \in \Omega$ , while at the same time we have $J(f_n) = \emptyset$ . However, the following proposition shows that this example is in some sense typical:

Proposition 1. Let $\Omega \subset \mathbb C$ be open, $\mathcal F \subset M(\Omega)$ a family of meromorphic functions and $z_0 \in \Omega$ . Suppose that $\mathcal F$ is transitive with respect to $z_0$ and that $z_0 \notin J(\mathcal F)$ . Then $\{f(z_0)\;:\; f \in \mathcal F\}$ is dense in $\mathbb C_\infty$ .

Proof. Suppose that $\{f(z_0)\;:\; f \in \mathcal F\}$ is not dense in $\mathbb C_\infty$ . Then there is $w \in \mathbb C_\infty$ and $\varepsilon > 0$ , such that $\{f(z_0)\;:\;f \in \mathcal F\} \cap D_{\varepsilon}^{\chi}(w) = \emptyset$ . Consider now for $k \in \mathbb N$ the sets $U_k \;:\!=\; D_{\frac{1}{k}}(z_0) \cap \Omega$ . Since $\mathcal F$ is transitive with respect to $z_0$ , for every $k \in \mathbb N$ there is $f_k \in \mathcal F$ such that $f_k(U_k) \cap D_{\frac{\varepsilon}{2}}^{\chi}(w) \neq \emptyset$ . In particular, there is a sequence $(z_k)$ with $z_k \in U_k$ , and hence $z_k \to z_0$ for $k \to \infty$ , such that $f_k(z_k) \in D_{\frac{\varepsilon}{2}}^{\chi}(w)$ for $k \in \mathbb N$ . On the other hand, we have $f_k(z_0) \notin D_{\varepsilon}^{\chi}(w)$ for $k \in \mathbb N$ . Finally, we obtain that

\begin{equation*}\chi(f_k(z_0), f_k(z_k)) > \frac{\varepsilon}{2} \qquad \text{for every } k \in \mathbb N,\end{equation*}

so that $\mathcal F$ is not spherically equicontinuous at $z_0$ , and thus also not normal, that is $z_0 \in J(\mathcal F)$ .

Example 1.

  1. (i) Let f be an entire function that is neither constant nor a polynomial of degree 1, and let $\mathcal F\;:\!=\;\{f^{\circ n}\;:\;n \in \mathbb N\}$ be the family of iterates of f. Then $\mathcal F$ is strongly non-normal on the Julia set $J=J(\mathcal F)$ , as follows e.g. from the facts that the repelling periodic points are dense in J and that J is the boundary of the escaping set (e.g. [Reference Carleson and Gamelin6,Reference Fatou14,Reference Morosawa, Nishimura, Taniguchi and Ueda26,Reference Schleicher30]). Here we have $\liminf_{z_0}\mathcal F\supset \mathbb C \setminus E$ for each $z_0 \in J$ , where E is the (empty or one-point) set of Fatou exceptional values of f, that is the set of points $w \in \mathbb C$ whose backward orbit $O^{-}(w) \;:\!=\; \bigcup_{n \geq 1}\{z\;:\;f^{\circ n}(z) = w\}$ is finite. Indeed, consider $z_0 \in J$ and an infinite subfamily $\tilde{\mathcal F} = \{f^{\circ n_k}\;:\;k \in \mathbb N\}$ with $n_k > 2$ . It follows from Picard’s Theorem that if $a \in \mathbb C$ is not Fatou exceptional, there are points $a_1, a_2 \in \mathbb C$ with $a_1 \neq a_2$ and $f^{\circ 2}(a_1) = a = f^{\circ 2}(a_2)$ . Since $\mathcal F$ is strongly non-normal at $z_0$ , Montel’s Theorem implies that the set $\mathbb C \setminus \limsup_{z_0} \tilde{\mathcal F}^-$ contains at most one point, where $\tilde{\mathcal F}^- \;:\!=\; \{f^{\circ (n_k - 2)}\;:\;k \in \mathbb N\}$ . Hence, $\{a_1,a_2\} \cap \limsup_{z_0} \tilde{\mathcal F}^- \neq \emptyset$ , which implies $a \in \limsup_{z_0}{\tilde{\mathcal F}}$ .

  2. (ii) Let M denote the Mandelbrot set and let, with $p_0\;:\!=\;\mathrm{id}_\mathbb C$ , the family $(p_n)$ of polynomials of degree $2^n$ be recursively defined by $p_n\;:\!=\;p_{n-1}^2 + \mathrm{ id}_\mathbb C$ . Since $p_n \to \infty$ pointwise on $\mathbb C \setminus M$ for $n \to \infty$ and $|p_n|\le 2$ on M (e.g. [Reference Carleson and Gamelin6]), we have $\partial M \subset J(\mathcal F)$ , where $\mathcal F\;:\!=\;\{p_n\;:\;n \in \mathbb N_0\}$ , and no infinite subfamily of $\mathcal F$ can be normal at any point of $\partial M$ . Hence, $\mathcal F$ is strongly non-normal and thus mixing on $\partial M$ .

  3. (iii) A function $f \in M(\mathbb C)$ is called Yosida function, if it has bounded spherical derivative $${f^\# }$$ (e.g. [Reference Minda and Chuang24,Reference Yosida32]). Hence, if f is not a Yosida function, there exists a sequence $(z_n)$ in $\mathbb C$ with $z_n \to \infty$ and ${f^\# }(z_n) \to \infty$ for $n \to \infty$ . Marty’s Theorem (e.g. [Reference Schiff29, p.75]) implies that the family $(f_n)$ with $f_n(z) \;:\!=\; f(z + z_n)$ is strongly non-normal at 0, hence by Theorem 1, we obtain that $(f_n)$ is mixing with respect to 0. Note that it is easily seen that if $f \in M(\mathbb C)$ is a Yosida function, then its order of growth is at most 2, while entire Yosida functions are necessarily of exponential type (e.g. [Reference Clunie and Hayman11,Reference Minda and Chuang24]).

For a family of meromorphic functions $\mathcal F \subset M(\Omega)$ and $N \in \mathbb N$ , we consider the family $\mathcal F^{\times N} \;:\!=\; \{f^{\times N}\;:\;f \in \mathcal F\}$ , where $f^{\times N}\;:\;\Omega^N \to \mathbb C_\infty^N$ with $f^{\times N}(z_1,\dots,z_N) = (f(z_1), \dots, f(z_N))$ . We say that $\mathcal F^{\times N}$ is transitive with respect to $z \in \Omega^N$ , if for every pair of non-empty open sets $U \subset \Omega^N$ and $V \subset \mathbb C_\infty^N$ with $z \in U$ , there exists $f^{\times N} \in \mathcal F^{\times N}$ such that $f^{\times N}(U) \cap V \neq \emptyset$ . Furthermore, for a relatively closed set $B \subset \Omega$ , we say that $\mathcal F^{\times N}$ is transitive with respect to $B^N$ , if $\mathcal F^{\times N}$ is transitive with respect to every $z \in B^N$ . We then have the following characterisation of hereditary non-normality.

Proposition 2. Let $\Omega \subset \mathbb C$ be open, $\mathcal F \subset M(\Omega)$ a family of meromorphic functions and $B \subset \Omega$ closed in $\Omega$ . Then the following are equivalent:

  1. (i) $\mathcal F$ is hereditarily non-normal on B;

  2. (ii) there exists an infinite subfamily $\tilde{\mathcal F} \subset \mathcal F$ that is mixing with respect to B;

  3. (iii) for all $N \in \mathbb N$ the family $\mathcal F^{\times N}$ is transitive with respect to $B^N$ .

Proof. The equivalence of (i) and (ii) follows from Theorem 1.

$\textrm{(ii)} \Rightarrow \textrm{(iii)}$ : Without loss of generality consider $\tilde{\mathcal F}$ to be countable, $\tilde{\mathcal F} = \{f_n\;:\;n \in \mathbb N\}$ say. Let $N \in \mathbb N$ and consider non-empty open sets $U \subset \Omega^N$ and $V \subset \mathbb C_\infty^N$ with $B^N \cap U \neq \emptyset$ . Then there exist non-empty open sets $U_1, \ldots, U_N$ with $U_1\times \cdots \times U_N\subset U$ and $B \cap U_i \neq \emptyset$ for $i = 1,\dots,N$ , and non-empty open sets $V_1,\ldots,V_N\subset \mathbb C_\infty$ with $V_1\times \cdots \times V_N\subset V$ . According to the assumption, $\{f_n:n>m\}$ is transitive with respect to B, for all $m \in \mathbb N$ . Inductively, we can find a strictly increasing sequence $(n_k)$ in $\mathbb{N}$ with $f_{n_k}(U_1)\cap V_1\neq\emptyset$ for all $k\in\mathbb{N}$ . By assumption, the family $\{f_{n_k}:k\in\mathbb{N}\}$ is transitive with respect to B. Thus, the same argument as above yields the existence of a subsequence $(n^{(2)}_k)$ of $(n^{(1)}_k)\;:\!=\;(n_k)$ with $f_{n^{(2)}_k}(U_2)\cap V_2\neq\emptyset$ for all $k\in\mathbb{N}$ . Proceeding in the same way, for any j with $2 \le j \le N$ we find subsequences $(n_k^{(j)})$ of $(n_k^{(j-1)})$ with $f_{n^{(j)}_k}(U_j)\cap V_j\neq\emptyset$ for all $k\in\mathbb{N}$ . In particular, for $n\;:\!=\;n_1^{(N)}$ , we obtain that

\begin{equation*}(f_n(U_1)\times \cdots \times f_n(U_N)) \cap (V_1\times \cdots \times V_N) \neq \emptyset,\end{equation*}

hence also $f_n^{\times N}(U)\cap V \neq \emptyset$ , implying that $\mathcal F^{\times N}$ is transitive with respect to $B^N$ .

$\textrm{(iii)} \Rightarrow \textrm{(ii)}$ The proof follows along the same lines as the proof of the corresponding part of the Bès–Peris Theorem (e.g. [Reference Grosse-erdmann and Peris21, pp.76]).

Remark 1.

  1. (i) Let $\mathcal{K}(A)$ denote the hyperspace of $A \subset \mathbb C$ , that is the space of all non-empty compact subsets of A endowed with the Hausdorff metric, and suppose that B as in Proposition 2 has non-empty interior. Then [Reference Beise, Meyrath and Müller2, corollay 1·2] shows that, under the conditions of Proposition 2, for each $\mathbb C$ -closed set $A \subset B$ which coincides with the closure of its interior, the family $\mathcal F|_E$ is dense in $C(E,\mathbb C_\infty)$ for generically many sets $E \in \mathcal{K}(A)$ .

  2. (ii) We mention that Proposition 2 is an extension of Theorem 3·7 from the recent paper [Reference Bernal-gonzález, Jung and Müller4].

Example 2.

  1. (i) Consider a function $f(z) = \sum_{\nu = 0}^{\infty} a_{\nu} z^{\nu}$ that is holomorphic on the unit disk $\mathbb D$ . Suppose that f has at least one singularity on $\partial \mathbb D$ and denote by $D \subset \partial \mathbb D$ the set of all singularities. Then, denoting by $s_n(z) \;:\!=\; (s_n f)(z) \;:\!=\; \sum_{\nu = 0}^{n} a_{\nu} z^{\nu}$ the nth partial sum of f, the family $(s_n)$ is non-normal on $\partial \mathbb D$ and strongly non-normal on D. Moreover, in case $D \neq \partial\mathbb D$ , Vitali’s Theorem implies that a subsequence of $(s_n)$ forms a normal family at a point $z_0 \in \partial \mathbb D \setminus D$ if and only if it converges to an analytic continuation of f in some neighbourhood of $z_0$ . From refined versions of Ostrowski’s results on overconvergence ([Reference Gehlen16, Theorems 3 and 4]), it follows that a subsequence $(s_{n_k})$ is strongly non-normal at $z_0 \in \partial \mathbb D \setminus D$ if and only if $(s_{n})$ has no Hadamard–Ostrowski gaps relative to $(n_k)$ , that is, if and only if there is a sequence $(\delta_k)$ of positive numbers tending to 0 with

    \begin{equation*}\sup_{(1-\delta_k)n_k\le\nu \le {n_k}} |a_\nu|^{1/\nu} \to 1\end{equation*}
    as $k \to \infty$ . In this case, the sequence $(s_{n_k})$ is already strongly non-normal at all $z \in \partial \mathbb D$ . Since the non-normality of $(s_n)$ on $\partial \mathbb D$ implies that, given $z_0 \in \partial \mathbb D \setminus D$ , some subsequence of $(s_n)$ is strongly non-normal at $z_0$ , we finally obtain that the family $(s_n)$ is always hereditarily non-normal on $\partial \mathbb D$ . According to a result of Gardiner ([Reference Gardiner15, corollary 3]), for each f that is holomorphic on $\mathbb D$ and analytically continuable to some domain U such that $\mathbb C \setminus U$ is thin at some $z_0 \in \partial \mathbb D$ but not continuable to the point $z_0$ , the sequence $(s_n)$ has no Hadamard–Ostrowski gaps with respect to any $(n_k)$ , hence $(s_n)$ is strongly non-normal on $ \partial \mathbb D$ . In particular, this holds for each f that has an isolated singularity at some point $z_0 \in \partial \mathbb{D}$ .
  2. (ii) We write $H_0$ for the space of functions holomorphic on $\mathbb{C} \setminus \{1\}$ that vanish at $\infty$ . For $f(z)=1/(1-z)$ , the sequence $(s_n f)$ is the geometric series which tends to $\infty$ spherically uniformly on compact subsets of $\mathbb C \setminus \overline{\mathbb{D}}$ . From [Reference Beise, Meyrath and Müller3, theorem 1·1] it can be deduced that generically many functions $f \in H_0$ enjoy the property that some subsequence of the sequence $((f-s_{n} f)(z)/z^n)$ converges to $1/(1-z)$ spherically uniformly on compact subsets of $\mathbb C_\infty \setminus \{1\}$ . This implies that the corresponding subsequence of $(s_n f)$ converges to $\infty$ spherically uniformly on compact subsets of $\mathbb C \setminus \overline{\mathbb{D}}$ and thus forms a normal family on $\mathbb C \setminus \overline{\mathbb{D}}$ . In particular, $(s_n f)$ is not strongly non-normal at any point $z_0 \in \mathbb C \setminus \overline{\mathbb{D}}$ . On the other hand, if A is a countable and dense subset of $\mathbb C \setminus \mathbb{D}$ , from [Reference Melas23, theorem 2] it follows that for generically many functions $f \in H_0$ , a subsequence $(s_{n_k}f)$ of $(s_{n}f)$ converges to 0 pointwise on A. Since a result from [Reference Kalmes, Müller and Niess22] implies that for $f \in H_0$ , normality of a subsequence of $(s_nf)$ at a point $z_0 \in \mathbb C \setminus \overline{\mathbb D}$ forces the subsequence to tend to $\infty$ spherically uniformly on compact subsets of some neighbourhood of $z_0$ , it follows that no subsequence of $(s_{n_k} f)$ can form a normal family at any point of $\mathbb C \setminus \overline{\mathbb D}$ . By the previous example, $(s_n f)$ is strongly non-normal on $\partial \mathbb{D}$ for $f \in H_0$ , thus we obtain that for generically many $f \in H_0$ , the family $(s_n f)$ is hereditarily non-normal on $\mathbb C \setminus \mathbb D$ . By Remark 1, for generically many $f \in H_0$ , the sequence $(s_nf|_E)$ is dense in $C(E, \mathbb C_\infty)$ for generically many $E \in \mathcal{K}(\mathbb C\setminus \mathbb D)$ (see also [Reference Beise, Meyrath and Müller1, theorem 2]).

3. Non-normality and expanding families

We define the following ‘expanding’ property of families $\mathcal F \subset M(\Omega)$ .

Definition 1. Let $\Omega \subset \mathbb C$ be open, $\mathcal F \subset M(\Omega)$ a family of meromorphic functions and $z_0 \in \Omega$ . Consider further a set $A \subset \mathbb C_\infty$ . We say that $\mathcal F$ is expanding at $z_0$ with respect to A, if for every open neighbourhood $U \subset \Omega$ of $z_0$ and every compact set $K \subset A$ we have $K \subset f(U)$ for infinitely many $f \in \mathcal F$ . If $K \subset f(U)$ holds for cofinitely many $f \in \mathcal F$ , we say that $\mathcal F$ is strongly expanding at $z_0$ with respect to A. Finally, we say that $\mathcal F$ is (strongly) expanding on a set $B \subset \Omega$ with respect to A, if $\mathcal F$ is (strongly) expanding with respect to A at every $z_0 \in B$ .

Note that if $\mathcal F$ is expanding at $z_0$ with respect to A, there exists an infinite subfamily $\tilde{\mathcal F} \subset \mathcal F$ that is strongly expanding at $z_0$ with respect to A. Moreover, in this case we have that A is contained in $\limsup_{z_0}\mathcal F$ . Also note that $\mathcal F$ is strongly expanding at $z_0$ with respect to A if and only if every infinite subfamily $\tilde{\mathcal F} \subset \mathcal F$ is expanding at $z_0$ with respect to A, and in this case, A is contained in $\liminf_{z_0}\mathcal F$ . On the other hand, we remark that $A \subset \liminf_{z_0}\mathcal F$ does in general not imply that $\mathcal F$ is (strongly) expanding at $z_0$ with respect to A. This can for instance be seen by considering the family $\mathcal F \;:\!=\; \{e^{nz} + (1 - {1}/{n})\;:\;n \in \mathbb N\}$ , for which we have $\liminf_{0}\mathcal F = \mathbb C$ (note that for each neighbourhood U of 0 and each $w \in \mathbb C$ , there is some N such that $w - 1 + {1}/{n} \in \exp(nU)$ for all $n \geq N$ ), but $\mathcal F$ is not expanding at 0 with respect to any set $A \subset \mathbb C$ with $1 \in A^{\circ}$ .

Our next result establishes a relationship between strong non-normality and the expanding property. Here and in the following, we denote by $\left|{E}\right| \in \mathbb N_0\cup\{\infty\}$ the number of elements of a set $E \subset \mathbb C_\infty$ .

Theorem 2. Let $\Omega \subset \mathbb C$ be open, $\mathcal F \subset M(\Omega)$ a family of meromorphic functions and $z_0 \in \Omega$ . Then we have:

  1. (i) if $\mathcal F$ is strongly non-normal at $z_0$ , then for each infinite subfamily $\tilde{\mathcal F} \subset \mathcal F$ there exists $E \subset \mathbb C_\infty$ with $\left|{E}\right| \leq 2$ , such that $\tilde{\mathcal F}$ is expanding at $z_0$ with respect to $\mathbb C_\infty \setminus E$ . Moreover, $\mathcal F$ is strongly expanding at $z_0$ with respect to $\mathbb C_\infty \setminus \mathcal E$ , where $\mathcal E \;:\!=\; \bigcup_{\tilde{\mathcal F} \subset \mathcal F \text{ infinite}} E_{\tilde{\mathcal F}}$ with $E_{\tilde{\mathcal F}} \subset \mathbb C_\infty$ being some set such that $\tilde{\mathcal F}$ is expanding at $z_0$ with respect to $\mathbb C_\infty \setminus E_{\tilde{\mathcal F}}$ ;

  2. (ii) if $|\liminf_{z_0}\mathcal F|\ge 2$ , then $\mathcal F$ is strongly non-normal at $z_0$ . In particular, this holds if $\mathcal F$ is strongly expanding at $z_0$ with respect to some $A \subset \mathbb C_\infty$ with $\left|{A}\right| \geq 2$ .

Proof. (i) Suppose that $\mathcal F$ is strongly non-normal at $z_0$ and consider an infinite subfamily $\tilde{\mathcal F} \subset \mathcal F$ . Then $\tilde{\mathcal F}$ is strongly non-normal at $z_0$ and assuming that $\tilde{\mathcal F}$ is not expanding at $z_0$ with respect to $\mathbb C_\infty \setminus E$ for any $E \subset \mathbb C_\infty$ with $\left|{E}\right| \leq 2$ , we obtain that for every $E \subset \mathbb C_\infty$ with $\left|{E}\right| \leq 2$ , there is an open neighbourhood U of $z_0$ and a compact set $K \subset \mathbb C_\infty \setminus E$ , such that $K \setminus f(U) \neq \emptyset$ for cofinitely many $f \in \tilde{\mathcal F}$ . In particular, if $\tilde{\mathcal F}$ is not expanding at $z_0$ with respect to $\mathbb C_\infty$ , we can find an open neighbourhood $U_1$ of $z_0$ , a sequence $(f_n)$ in $\tilde{\mathcal F}$ , and a sequence $(a_n)$ in $\mathbb C_\infty$ with $a_n \to a \in \mathbb C_\infty$ for $n \to \infty$ , such that $a_n \notin f_n(U_1)$ for every $n \in \mathbb N$ . By assumption, $\tilde{\mathcal F}$ is not expanding at $z_0$ with respect to $\mathbb C_\infty \setminus \{a\}$ , hence, there is an open neighbourhood $U_2$ of $z_0$ and a compact set $K_2 \subset \mathbb C_\infty \setminus \{a\}$ , such that $K_2 \setminus f(U_2) \neq \emptyset$ for cofinitely many $f \in \tilde{\mathcal F}$ . In particular, there is a subsequence $(f_{n_k})$ of $(f_n)$ , and a sequence $(b_k)$ in $K_2$ with $b_k \to b \in K_2$ for $k \to \infty$ , such that $b_k \notin f_{n_k}(U_2)$ for every $k \in \mathbb N$ . Since $\tilde{\mathcal F}$ is not expanding at $z_0$ with respect to $\mathbb C_\infty \setminus \{a,b\}$ , a similar argumentation leads to an open neighbourhood $U_3$ of $z_0$ , a compact set $K_3 \subset \mathbb C_\infty \setminus \{a,b\}$ , a subsequence $(f_{n_{k_l}})$ of $(f_{n_k})$ and a sequence $(c_l)$ in $K_3$ with $c_l \to c \in K_3$ for $l \to \infty$ , such that $c_l \notin f_{n_{k_l}}(U_3)$ for every $l \in \mathbb N$ .

Finally, setting $U = U_1 \cap U_2 \cap U_3$ we obtain that

\begin{equation*}\{a_{n_{k_l}}, b_{k_l}, c_l\} \cap f_{n_{k_l}}(U) = \emptyset \quad \text{for every } l \in \mathbb N.\end{equation*}

Furthermore, since a,b,c are pairwise distinct, there exists $\varepsilon > 0$ such that

\begin{equation*}\chi(a_{n_{k_l}}, b_{k_l}) \, \chi(b_{k_l}, c_l) \, \chi(a_{n_{k_l}}, c_l) > \varepsilon,\end{equation*}

for $l \in \mathbb N$ sufficiently large, so that Carathéodory’s extension of Montel’s Theorem (e.g. [Reference Schiff29, p.104]) implies that $(f_{n_{k_l}}) \subset \tilde{\mathcal F}$ is normal in U, hence also at $z_0$ , in contradiction to the strong non-normality of $\tilde{\mathcal F}$ at $z_0$ .

To prove the second statement, suppose that $\mathcal F$ is not strongly expanding at $z_0$ with respect to $\mathbb C_\infty\setminus \mathcal E$ . Then there is an infinite subfamily $\tilde{\mathcal F} \subset \mathcal F$ that is not expanding at $z_0$ with respect to $\mathbb C_\infty \setminus \mathcal E$ , contradicting the fact that $\tilde{\mathcal F}$ is expanding at $z_0$ with respect to $\mathbb C_\infty \setminus E_{\tilde{\mathcal F}}$ for some set $E_{\tilde{\mathcal F}} \subset \mathbb C_\infty$ with $E_{\tilde{\mathcal F}} \subset \mathcal E$ .

(ii) Suppose that for some infinite subfamily $\tilde{\mathcal F} = \{f_n:n \in \mathbb N\}$ of $\mathcal F$ the sequence $(f_n)$ is spherically uniformly convergent on compact subsets of a neighbourhood of $z_0$ . Then $\limsup_{z_0} \tilde{\mathcal F}$ is a one-point set, and hence $|\liminf_{z_0}\mathcal F|\le 1$ . The second statement follows from the fact that in this case we have $A \subset \liminf_{z_0}\mathcal F$ .

Remark 2. Note that if $\mathcal F$ is strongly non-normal at $z_0$ , $\mathcal F$ does not need to be strongly expanding at $z_0$ with respect to any open set $A \subset \mathbb C_\infty$ . Indeed, let $(q_n)$ be an enumeration of the Gaussian rational numbers with $q_n^2/n \to 0$ as $n \to \infty$ and consider the family $(f_n)$ with $f_n(z)\;:\!=\;e^{nz}+q_n$ for $z \in \mathbb C$ . From Marty’s Theorem, it is easily seen that $(f_n)$ is strongly non-normal on the imaginary axis $i\mathbb R$ , but for a point $z_0 \in i\mathbb R$ and an open neighbourhood U of $z_0$ , there is no $N \in \mathbb N$ such that $K \subset f_n(U)$ holds for all $n \geq N$ for any compact set $K \subset \mathbb C$ with $K^{\circ} \neq \emptyset$ .

From Theorem 2 we easily obtain the following characterisation of non-normality in terms of the expanding property, which in some sense complements the statement of Montel’s Theorem:

Corollary 1. Let $\Omega \subset \mathbb C$ be open, $\mathcal F \subset M(\Omega)$ a family of meromorphic functions and $z_0 \in \Omega$ . Then the following are equivalent:

  1. (i) there exists $A \subset \mathbb C_\infty$ with $\left|{A}\right| \geq 2$ such that $\mathcal F$ is expanding at $z_0$ with respect to A;

  2. (ii) $\mathcal F$ is non-normal at $z_0$ ;

  3. (iii) there exists $E \subset \mathbb C_\infty$ with $\left|{E}\right| \leq 2$ such that $\mathcal F$ is expanding at $z_0$ with respect to $\mathbb C_\infty \setminus E$ .

Proof. $\textrm{(i)} \Rightarrow \textrm{(ii)}$ Suppose that $\mathcal F$ is expanding at $z_0$ with respect to some $A \subset \mathbb C_\infty$ with $\left|{A}\right| \geq 2$ . Then there exists an infinite subfamily $\tilde{\mathcal F} \subset \mathcal F$ that is strongly expanding at $z_0$ with respect to A. By Theorem 2, the family $\tilde{\mathcal F}$ is strongly non-normal at $z_0$ , hence $\mathcal F$ is non-normal at $z_0$ .

$\textrm{(ii)} \Rightarrow \textrm{(iii)}$ If $\mathcal F$ is non-normal at $z_0$ , there exists an infinite subfamily $\tilde{\mathcal F} \subset \mathcal F$ that is strongly non-normal at $z_0$ . By Theorem 2, there then exists $E \subset \mathbb C_\infty$ with $\left|{E}\right| \leq 2$ such that $\tilde{\mathcal F}$ is expanding at $z_0$ with respect to $\mathbb C_\infty \setminus E$ . The same then holds for the family $\mathcal F$ .

$\textrm{(iii)} \Rightarrow \textrm{(i)}$ is obvious.

Let $\mathcal F \subset M(\Omega)$ be a family that is non-normal at a point $z_0 \in \Omega$ and consider the set $E_{z_0}(\mathcal F) = \mathbb C_\infty \setminus \limsup_{z_0} \mathcal F$ . If $\mathcal F$ is expanding at $z_0$ with respect to $\mathbb C_\infty \setminus E$ for some set $E \subset \mathbb C_\infty$ , we obviously have $E_{z_0}(\mathcal F) \subset E$ . If $\mathcal F$ is a family of holomorphic functions on $\Omega$ that is (strongly) non-normal at $z_0$ , we have $\infty \in E_{z_0}(\mathcal F)$ , so that in this case we obtain that the expanding property of $\mathcal F$ at $z_0$ in Theorem 2 and Corollary 1 holds with respect to $\mathbb C \setminus E$ for some set $E \subset \mathbb C$ with $\left|{E}\right| \leq 1$ .

Example 3.

  1. (i) Consider a compact set $K \subset \mathbb C$ with connected complement and let f be a function that is continuous on K and holomorphic in $K^{\circ}$ . Further assume that f has at least one singularity on $\partial K$ and denote by $D \subset \partial K$ the set of all singularities. Let $(p_n)$ be a sequence of polynomials converging uniformly on K to f (such a sequence exists by Mergelian’s Theorem). Then, $(p_n)$ is strongly non-normal on D, hence also expanding at every point $z_0 \in D$ with respect to $\mathbb C \setminus E$ for some set $E \subset \mathbb C$ with $\left|{E}\right| \leq 1$ . Indeed, since otherwise there exists a point $z_0 \in D$ , an open neighbourhood U of $z_0$ , and a subsequence $(p_{n_k})$ of $(p_n)$ that converges uniformly on compact subsets of U to a function holomorphic in U, contradicting that f does not have an analytic continuation across $z_0 \in D$ .

  2. (ii) Consider the function $f(z) = \left|{z}\right|$ on the interval $[\!-\!1,1]$ and denote by $(p_n^{\star})$ the sequence of polynomials of best uniform approximation to f on $[\!-\!1,1]$ . Then, according to the previous example, $(p_n^{\star})$ is strongly non-normal at the point 0. However, since $p_n^{\star}(z) \to \infty$ for $n \to \infty$ spherically uniformly on compact subsets of $\mathbb C \setminus [\!-\!1,1]$ (e.g. [Reference Saff and Stahl28]), the family $(p_n^{\star})$ is strongly non-normal on $[\!-\!1,1]$ , hence expanding at every point $z_0 \in [\!-\!1,1]$ with respect to $\mathbb C \setminus E$ for some set $E \subset \mathbb C$ with $\left|{E}\right| \leq 1$ . (Note that the strong non-normality on $[\!-\!1,1]$ also holds for several specific ray sequences of best uniform rational approximants to f on $[\!-\!1,1]$ ([Reference Saff and Stahl28, corollary 1·3]).) In fact, [Reference Blatt, Blatt and Luh5, corollary 2] implies that $(p_n^{\star})$ is expanding on $[\!-\!1,1]$ with respect to $\mathbb C$ , as it shows the existence of a subsequence $(p_{n_k}^{\star})$ of $(p_n^{\star})$ that is strongly expanding on $[\!-\!1,1]$ with respect to $\mathbb C$ .

  3. (iii) Consider again a function $f(z) = \sum_{\nu = 0}^{\infty} a_{\nu} z^{\nu}$ that is holomorphic on $\mathbb D$ and has at least one singularity on $\partial \mathbb D$ . Then the family of partial sums $(s_n)$ is non-normal on $\partial \mathbb D$ , hence, $(s_n)$ is expanding at every $z_0 \in \partial \mathbb D$ with respect to $\mathbb C \setminus E$ for some set $E \subset \mathbb C$ with $\left|{E}\right| \leq 1$ . In fact, $(s_n)$ is expanding on $\partial \mathbb D$ with respect to $\mathbb C$ , as results in [Reference Blatt, Blatt and Luh5,Reference Dvoretzky13] show that if $(a_{n_k})$ is a sequence such that $\lim_{k \to \infty} \left|{a_{n_k}}\right|^{\frac{1}{n_k}} = 1$ , the subfamily $(s_{n_k})$ is strongly expanding on $\partial \mathbb D$ with respect to $\mathbb C$ .

A further consequence of Theorem 2 and the fact that we have $E_{z_0}(\mathcal F) \subset E$ if $\mathcal F \subset M(\Omega)$ is expanding at $z_0 \in \Omega$ with respect to $\mathbb C_\infty \setminus E$ is the following statement for the case $\left|{E_{z_0}(\mathcal F)}\right| = 2$ .

Corollary 2. Let $\Omega \subset \mathbb C$ be open and $\mathcal F \subset M(\Omega)$ be a family of meromorphic functions. Consider $z_0 \in \Omega$ and suppose that $\mathcal F$ is (strongly) non-normal at $z_0$ with $\left|{E_{z_0}(\mathcal F)}\right| = 2$ . Then $\mathcal F$ is (strongly) expanding at $z_0$ with respect to $\mathbb C_\infty \setminus E_{z_0}(\mathcal F)$ .

Proof. Suppose that $\mathcal F$ is non-normal at $z_0$ . By Corollary 1, there then exists $E \subset \mathbb C_\infty$ with $\left|{E}\right| \leq 2$ such that $\mathcal F$ is expanding at $z_0$ with respect to $\mathbb C_\infty \setminus E$ . Since $E_{z_0}(\mathcal F) \subset E$ , we obtain $E_{z_0}(\mathcal F) = E$ . If $\mathcal F$ is strongly non-normal at $z_0$ , every infinite subfamily $\tilde{\mathcal F} \subset \mathcal F$ is non-normal at $z_0$ with $E_{z_0}(\tilde{\mathcal F}) = E_{z_0}(\mathcal F)$ , hence expanding at $z_0$ with respect to $\mathbb C_\infty \setminus E_{z_0}(\mathcal F)$ .

Example 4.

  1. (i) Consider again the family $\mathcal F \;:\!=\; \{e^{nz} + (1 - {1}/{n})\;:\;n \in \mathbb N\}$ , which is strongly non-normal at the point 0. It is easily seen that $\mathcal F$ is strongly expanding at 0 with respect to $\mathbb C_\infty \setminus \{1,\infty\}$ , but since $E_{0}(\mathcal F) = \{\infty\}$ , this can not be derived from Corollary 2. On the other hand, the family $\mathcal F \;:\!=\; \{e^{nz} + (1 - {1}/{n!})\;:\;n \in \mathbb N\}$ is strongly non-normal at the point 0 with $E_{0}(\mathcal F) = \{1, \infty\}$ (note that for each neighbourhood U of 0 and each $1 \not = w \in \mathbb C$ , there is some N with $w - 1 + {1}/{n!} \in \exp(nU)$ for all $n\ge N$ , but ${1}/{n!} \not\in \exp(n\mathbb D)$ for sufficiently large n). So, in this case Corollary 2 can be applied.

  2. (ii) Consider again a power series $f(z) = \sum_{\nu = 0}^{\infty} a_{\nu} z^{\nu}$ with radius of convergence 1 and denote by $(s_n)$ its partial sums. As mentioned in Example 3, the family $\mathcal F = \{s_n\;:\;n \in \mathbb N\}$ is expanding on $\partial \mathbb D$ with respect to $\mathbb C$ , so that for every $z_0 \in \partial \mathbb D$ we have $E_{z_0}(\mathcal F) = \{\infty\}$ (note that this is also easily derived from the classical Jentzsch Theorem ([Reference Jentzsch19]) stating that for every $a \in \mathbb C$ , every $z_0 \in \partial \mathbb D$ is a limit point of a-points of the partial sums). However, a further result of Jentzsch ([Reference Jentzsch20]) states that there exist power series with radius of convergence 1, such that the zeros of some subsequence $(s_{n_k})$ of the partial sums do not have a finite limit point. Hence, in this case Corollary 2 shows that the family $\tilde{\mathcal F} = \{s_{n_k}\;:\;k \in \mathbb N\}$ is strongly expanding with respect to $\mathbb C \setminus \{0\}$ at every point $z_0 \in \partial \mathbb D$ at which the function does not admit an analytic continuation (there must be at least one such point), since $\tilde{\mathcal F}$ is strongly non-normal at such $z_0$ with $E_{z_0}(\tilde{\mathcal F}) = \{0, \infty\}$ . In a similar vein, it was shown in [Reference Ivanov, Saff and Totik18, theorem 1] that there exists a function f holomorphic on $\mathbb D$ and continuous on $\overline{\mathbb D}$ with at least one singularity on $\partial \mathbb D$ , for which the zeros of some subsequence $(p_{n_k}^{\star})$ of the sequence $(p_n^{\star})$ of polynomials of best uniform approximation do not have a finite limit point. Hence, as before, Corollary 2 can be applied to the family $\mathcal F = \{p_{n_k}^{\star}\;:\;k \in \mathbb N \}$ at every singular point $z_0 \in \partial \mathbb D$ of f, since $\mathcal F$ is strongly non-normal at $z_0$ (see Example 3 (i)) and we have $E_{z_0}(\mathcal F) = \{0, \infty\}$ . Moreover, [Reference Ivanov, Saff and Totik18, theorem 2] shows the existence of a function f that is holomorphic on $\mathbb D$ and continuous on $\overline{\mathbb D}$ with at least one singularity on $\partial \mathbb D$ , for which there is a sequence $(q_n)$ of polynomials of near-best uniform approximation that has no finite limit point of zeros. Hence, in this case Corollary 2 implies that the family $\mathcal F = \{q_n\;:\;n \in \mathbb N \}$ is strongly expanding with respect to $\mathbb C \setminus \{0\}$ at every singular point $z_0 \in \partial \mathbb D$ of f.

4. Expanding families of derivatives

In the following, we show that under certain conditions, (strong) non-normality of a family $\mathcal F \subset M(\Omega)$ at a point $z_0 \in \Omega$ implies that the family of derivatives is (strongly) expanding at $z_0$ with respect to $\mathbb C \setminus \{0\}$ , hence in particular (strongly) non-normal at $z_0$ . Throughout this section, we denote by $\mathcal F^{(k)}$ the family of kth derivatives of the functions in $\mathcal F$ , that is $\mathcal F^{(k)} = \{f^{(k)}\;:\;f \in \mathcal F\}$ , where k is some natural number.

Theorem 3. Let $\Omega \subset \mathbb C$ be open and $\mathcal F \subset M(\Omega)$ be a family of meromorphic functions. Consider $z_0 \in \Omega$ and suppose that $\mathcal F$ is (strongly) non-normal at $z_0$ . Further assume that $\mathcal F$ is not expanding at $z_0$ with respect to $\mathbb C$ . Then, for every $k \in \mathbb N$ , the family $\mathcal F^{(k)}$ is (strongly) expanding at $z_0$ with respect to $\mathbb C \setminus \{0\}$ .

Proof. We first assume that $\mathcal F$ is strongly non-normal at $z_0$ . By assumption, $\mathcal F$ is not expanding at $z_0$ with respect to $\mathbb C$ , hence there exists an open neighbourhood $U_1$ of $z_0$ and a compact set $K_1 \subset \mathbb C$ such that $K_1 \setminus f(U_1) \neq \emptyset$ holds for cofinitely many $f \in \mathcal F$ .

Now assume that there exists $k \in \mathbb N$ , such that $\mathcal F^{(k)}$ is not strongly expanding at $z_0$ with respect to $\mathbb C \setminus \{0\}$ . Then there exists an open neighbourhood $U_2$ of $z_0$ and a compact set $K_2 \subset \mathbb C \setminus \{0\}$ such that $K_2 \setminus f^{(k)}(U_2) \neq \emptyset$ holds for infinitely many $f \in \mathcal F$ .

In particular, we can find a sequence $(f_n)$ in $\mathcal F$ , and sequences $(c^{(1)}_n)$ in $K_1$ and $(c^{(2)}_n)$ in $K_2$ , such that the equations $f_n(z) = c_n^{(1)}$ and $f_n^{(k)}(z) = c_n^{(2)}$ have no roots in $U \;:\!=\; U_1 \cap U_2$ for every $n \in \mathbb N$ . From [Reference Chuang10, theorem 3·17], which is an extension of Gu’s famous normality criterion (e.g. [Reference Gu17,Reference Schiff29]), we obtain that $(f_n)$ is normal in U, hence also at $z_0$ , in contradiction to the strong non-normality of $\mathcal F$ at $z_0$ .

If $\mathcal F$ is non-normal at $z_0$ , there exists an infinite subfamily $\tilde{\mathcal F} \subset \mathcal F$ that is strongly non-normal at $z_0$ . By assumption, $\mathcal F$ is not expanding at $z_0$ with respect to $\mathbb C$ , hence the same holds for $\tilde{\mathcal F}$ , so that by the above argumentation, $\tilde{\mathcal F}^{(k)}$ is strongly expanding at $z_0$ with respect to $\mathbb C \setminus \{0\}$ for every $k \in \mathbb N$ . Hence, $\mathcal F^{(k)}$ is expanding at $z_0$ with respect to $\mathbb C \setminus \{0\}$ for every $k \in \mathbb N$ .

Remark 3. Note that the assumption that $\mathcal F$ is not expanding at $z_0$ with respect to $\mathbb C$ turns out to be necessary, as is seen e.g. by considering the sequence of polynomials $f_n(z) = nz$ and $z_0 = 0$ . Moreover, it is easily seen that a similar argumentation as in the proof of the theorem leads to the following result: Let $\Omega \subset \mathbb C$ be open and $\mathcal F \subset M(\Omega)$ be a family of meromorphic functions. Consider $z_0 \in \Omega$ and suppose that $\mathcal F$ is (strongly) non-normal at $z_0$ . Further assume that for some $k \in \mathbb N$ , the family $\mathcal F^{(k)}$ is not expanding at $z_0$ with respect to $\mathbb C \setminus \{0\}$ . Then, the family $\mathcal F$ is (strongly) expanding at $z_0$ with respect to $\mathbb C$ .

Corollary 3. Let $\Omega \subset \mathbb C$ be open and $\mathcal F \subset M(\Omega)$ be a family of meromorphic functions. Consider $z_0 \in \Omega$ and suppose that $\mathcal F$ is (strongly) non-normal at $z_0$ . Suppose further that there exists an open neighbourhood U of $z_0$ and a number $M > 0$ , such that for cofinitely many $f \in \mathcal F$ there is a point $a_f \in \mathbb C$ with $\left|{a_f}\right| < M$ and $a_f \notin f(U)$ . Then, for every $k \in \mathbb N$ , the family $\mathcal F^{(k)}$ is (strongly) expanding at $z_0$ with respect to $\mathbb C \setminus \{0\}$ .

Proof. Since it follows from the assumptions that $\mathcal F$ is not expanding at $z_0$ with respect to $\mathbb C$ , the statement follows from Theorem 3.

Note that the assumptions of Corollary 3 are fulfilled if $\mathcal F \subset M(\Omega)$ is (strongly) non-normal at $z_0 \in \Omega$ and for some $a \in \mathbb C$ we have $a \in E_{z_0}(\mathcal F)$ , hence in particular if $\left|{E_{z_0}(\mathcal F)}\right| = 2$ .

Example 5.

  1. (i) In Example 4 (ii) we considered strongly non-normal families $\mathcal F$ of polynomials for which $E_{z_0}(\mathcal F) = \{0, \infty\}$ , hence we obtain that the corresponding families of derivatives $\mathcal F^{(k)}$ are strongly expanding at $z_0$ with respect to $\mathbb C \setminus \{0\}$ for every $k \in \mathbb N$ .

  2. (ii) Consider the family $(f_n)$ with $f_n \;:\!=\; \exp^{\circ n}$ , the nth iterate of $e^z$ . Then $J(f_n)$ coincides with the Julia set of $e^z$ , which is known to equal $\mathbb C$ ([Reference Misiurewicz25]). According to Example 1 (i), we thus have that $(f_n)$ is strongly non-normal on $\mathbb C$ . Furthermore, we obviously have $0 \in E_{z_0}(f_n)$ for every $z_0 \in \mathbb C$ , so that Corollary 3 implies that for every $k \in \mathbb N$ , the family $(f_n^{(k)})$ is strongly expanding on $\mathbb C$ with respect to $\mathbb C \setminus \{0\}$ .

We mention that the statement of Corollary 3 remains valid to some extent, if instead of omitting a value $a_f$ in some neighbourhood of $z_0$ , cofinitely many functions $f \in \mathcal F$ have a value $a_f$ that they take with sufficiently high multiplicity in that neighbourhood.

Proposition 3. Let $\Omega \subset \mathbb C$ be open and $\mathcal F \subset M(\Omega)$ be a family of meromorphic functions. Consider $z_0 \in \Omega$ and suppose that $\mathcal F$ is (strongly) non-normal at $z_0$ . Suppose further that there exists an open neighbourhood U of $z_0$ , a number $M > 0$ and some $k \in \mathbb N$ , such that for cofinitely many $f \in \mathcal F$ there is a point $a_f \in \mathbb C$ with $\left|{a_f}\right| < M$ , such that the $a_f$ -points of f in U have multiplicity at least $k+2$ . Then the family $\mathcal F^{(k)}$ is (strongly) expanding at $z_0$ with respect to $\mathbb C \setminus \{0\}$ .

Proof. Again, we first consider the case that $\mathcal F$ is strongly non-normal at $z_0$ . Assuming that $\mathcal F^{(k)}$ is not strongly expanding at $z_0$ with respect to $\mathbb C \setminus \{0\}$ , there exists an open neighbourhood $U_1$ of $z_0$ and a compact set $K \subset \mathbb C \setminus \{0\}$ such that $K \setminus f^{(k)}(U_1) \neq \emptyset$ for infinitely many $f \in \mathcal F$ . In particular, we can find a sequence $(c_n)$ in K with $c_n \to c$ for some $c \neq 0$ , and a sequence $(f_n)$ in $\mathcal F$ such that $c_n \notin f_n^{(k)}(U_1)$ for every $n \in \mathbb N$ . For $n \in \mathbb N$ sufficiently large, say $n > N$ , there is a point $a_{f_n} \in \mathbb C$ with $\left|{a_{f_n}}\right| < M$ , such that the $a_{f_n}$ -points of $f_n$ in U have multiplicity at least $k+2$ . Setting $g_n(z) = f_n(z) - a_{f_n}$ for $n > N$ , we obtain that the functions $g_n$ only have zeros of multiplicity at least $k+2$ in $U' \;:\!=\; U \cap U_1$ . Furthermore, since $c_n \notin g_n^{(k)}(U')$ for every $n > N$ , it follows from [Reference Cheng and Xu9, lemma 2·7] that the family $\{g_n\;:\;n > N\}$ is normal in U , and as $\left|{a_{f_n}}\right| < M$ for every $n > N$ , the same holds for the family $\{f_n\;:\;n > N\}$ . This is in contradiction to the strong non-normality of $\mathcal F$ at $z_0$ .

If $\mathcal F$ is non-normal at $z_0$ , the statement follows as before from the fact that $\mathcal F$ contains a strongly non-normal subfamily.

In general, the number $k+2$ can not be replaced by $k+1$ in Proposition 3. Indeed, for fixed $k \in \mathbb N$ , the family $(f_n)$ with

\begin{equation*}f_n(z) = \frac{1}{k!} \, \frac{z^{k+1}}{(z - \frac{1}{n})}\end{equation*}

is strongly non-normal at the point 0 and has only zeros of multiplicity $k + 1$ (see also [Reference Wang and Fang31]). But as $f_n^{(k)}(z) \neq 1$ for every $n \in \mathbb N$ and every $z \in \mathbb C$ , the familiy $(f_n^{(k)})$ is obviously not expanding at 0 with respect to $\mathbb C \setminus \{0\}$ . Nevertheless, under certain additional conditions, $k+2$ can be replaced by $k+1$ :

Proposition 4. Under each of the following additional conditions, the statement of Proposition 3 remains valid if $k+2$ is replaced by $k+1$ :

  1. (i) the functions $f \in \mathcal F$ are holomorphic in $\Omega$ ;

  2. (ii) the functions $f \in \mathcal F$ only have multiple poles;

  3. (iii) there exists a sequence $(z_n)$ in $\Omega$ with $z_n \to z_0$ and $\mathcal F$ is strongly non-normal at $z_n$ for every $n \in \mathbb N$ .

Proof. Using [Reference Chen7, lemma 4] and [Reference Nevo, Pang and Zalcman27, lemma 6], respectively, the proofs of (i) and (ii) are similar to the proof of Proposition 3. In order to prove the third statement, we note that using [Reference Chen, Pang and Yang8, lemma 2·9], a similar argumentation as in the proof of Proposition 3 implies that the family $(g_n)$ with $g_n(z) = f_n(z) - a_{f_n}$ is quasinormal in some neighbourhood U of $z_0$ . Since $\left|{a_{f_n}}\right| < M$ for every $n \in \mathbb N$ , the same then holds for the family $(f_n)$ ([Reference Chuang10, lemma 5·2]). This contradicts the assumption that the set $\{z\;:\;\mathcal F \text{ is strongly non-normal at } z\}$ has an accumulation point in U.

Acknowledgements

The authors would like to thank the anonymous reviewer for his careful reading of the manuscript and his valuable comments and suggestions.

References

Beise, H.-P., Meyrath, T. and Müller, J.. Universality properties of Taylor series inside the domain of holomorphy. J. Math. Anal. Appl. 383 (2011), 234238.CrossRefGoogle Scholar
Beise, H.-P., Meyrath, T. and Müller, J.. Limit functions of discrete dynamical systems. Conform. Geom. Dyn. 18 (2014), 5664.CrossRefGoogle Scholar
Beise, H.-P., Meyrath, T. and Müller, J.. Mixing Taylor shifts and universal Taylor series. Bull. London Math. Soc. 47 (2015), 136142.CrossRefGoogle Scholar
Bernal-gonzález, L., Jung, A. and Müller, J.. Universality vs. non-normality of families of meromorphic functions. Proc. Amer. Math. Soc. 149 (2021), 761771.CrossRefGoogle Scholar
Blatt, H. P., Blatt, S. and Luh, W.. On a generalization of Jentzsch’s theorem. J. Approx. Theory 159 (2009), 2638.CrossRefGoogle Scholar
Carleson, L. and Gamelin, T. W.. Complex Dynamics (Springer, New York, 1993).Google Scholar
Chen, J.-F.. Exceptional functions and normal families of holomorphic functions with multiple zeros. Georgian Math. J. 18 (2011), 3138.CrossRefGoogle Scholar
Chen, Q., Pang, X. and Yang, P.. A new Picard type theorem concerning elliptic functions. Ann. Acad. Sci. Fenn. Math. 40 (2015), 1730.CrossRefGoogle Scholar
Cheng, C. and Xu, Y.. Normality concerning exceptional functions. Rocky Mt. J. Math. 45 (2015), 157168.CrossRefGoogle Scholar
Chuang, C.-T.. Normal families of meromorphic functions, (World Scientific, 1993).CrossRefGoogle Scholar
Clunie, J. and Hayman, W. K.. The spherical derivative of integral and meromorphic functions. Comment. Math. Helv. 40 (1966), 117148.CrossRefGoogle Scholar
Conway, J. B.. Functions of One Complex Variable, I 2nd ed., (Springer, New York, 1978).CrossRefGoogle Scholar
Dvoretzky, A.. On sections of power series. Ann. of Math. 51 (1950), 643696.CrossRefGoogle Scholar
Fatou, P.. Sur l’itération des fonctions transcendantes entières. Acta Math. 47 (1926), 337360.CrossRefGoogle Scholar
Gardiner, S.. Existence of universal Taylor series for nonsimply connected domains. Constr. Approx. 35 (2012), 245257.CrossRefGoogle Scholar
Gehlen, W.. Overconvergent power series and conformal maps. J. Math. Anal. Appl. 198 (1996), 490505.CrossRefGoogle Scholar
Gu, Y. X.. A normal criterion of meromorphic families. Sci. Sinica 1 (1979), 267274.Google Scholar
Ivanov, K. G., Saff, E. B. and Totik, V.. On the behavior of zeros of polynomials of best and near-best approximation. Canad. J. Math. 43 (1991), 10101021.CrossRefGoogle Scholar
Jentzsch, R.. Untersuchungen zur Theorie der Folgen analytischer Funktionen. Acta. Math. 41 (1918) 219251.CrossRefGoogle Scholar
Jentzsch, R.. Fortgesetzte Untersuchungen über die Abschnitte von Potenzreihen. Acta. Math. 41 (1918), 253270.CrossRefGoogle Scholar
Grosse-erdmann, K.-G. and Peris, A.. Linear Chaos, (Springer, London, 2011).CrossRefGoogle Scholar
Kalmes, T., Müller, J. and Niess, M.. On the behaviour of power series in the absence of Hadamard–Ostrowski gaps. C. R. Math. Acad. Sci. Paris 351 (2013), 255259.CrossRefGoogle Scholar
Melas, A.. Universal functions on nonsimply connected domains. Ann. Inst. Fourier (Grenoble) 51 (2001), 15391551.CrossRefGoogle Scholar
Minda, D.. Yosida functions, Lectures on Complex Analysis (Xian, 1987), Chuang, C.-T. (ed.) (World Scientific, Singapore, 1988), 197213.Google Scholar
Misiurewicz, M.. On iterates of $e^z$ . Ergod. Theory Dynam. Systems. 1 (1981), 103106.CrossRefGoogle Scholar
Morosawa, S., Nishimura, Y., Taniguchi, M. and Ueda, T.. Holomorphic Dynamics. (Cambridge University Press, Cambridge, 2000).Google Scholar
Nevo, S., Pang, X. and Zalcman, L.. Quasinormality and meromorphic functions with multiple zeros. J. Anal. Math. 101 (2007), 123.CrossRefGoogle Scholar
Saff, E. B. and Stahl, H.. Ray sequences of best rational approximants for $\left|{x}\right|^{\alpha}$ . Canad. J. Math. 49 (1997), 10341065.CrossRefGoogle Scholar
Schiff, J. L.. Normal Families, Springer, (New York, Berlin, Heidelberg, 1993).Google Scholar
Schleicher, D.. Dynamics of entire functions. In: Holomorphic Dynamical Systems. Lecture Notes in Math. vol. 1998 (Springer, Berlin, 2010), 295339.Google Scholar
Wang, Y. and Fang, M.. Picard values and normal families of meromorphic functions with multiple zeros. Acta Math. Sinica (N.S.) 14 (1998), 1726.Google Scholar
Yosida, K.. On a class of meromorphic functions. Proc. Phys.-Math. Soc. Japan 16 (1934), 227235.Google Scholar