Sbornik: Mathematics
RUS  ENG    JOURNALS   PEOPLE   ORGANISATIONS   CONFERENCES   SEMINARS   VIDEO LIBRARY   PACKAGE AMSBIB  
General information
Latest issue
Forthcoming papers
Archive
Impact factor
Guidelines for authors
License agreement
Submit a manuscript

Search papers
Search references

RSS
Latest issue
Current issues
Archive issues
What is RSS



Mat. Sb.:
Year:
Volume:
Issue:
Page:
Find






Personal entry:
Login:
Password:
Save password
Enter
Forgotten password?
Register


Sbornik: Mathematics, 2022, Volume 213, Issue 11, Pages 1488–1506
DOI: https://doi.org/10.4213/sm9611e
(Mi sm9611)
 

This article is cited in 2 scientific papers (total in 2 papers)

Necessary and sufficient conditions for extending a function to a Carathéodory function

V. I. Buslaev

Steklov Mathematical Institute of Russian Academy of Sciences, Moscow, Russia
References:
Abstract: A criterion deciding whether a function given by its values (with multiplicities) at a sequence of points in the disc $\mathbb D=\{|z|<1\}$ can be extended to a holomorphic function with nonnegative real part in $\mathbb D$ is stated and proved. In the case when this function is given by the values of its derivatives at $z=0$, this is the well-known Carathéodory criterion. It is also shown that Carathéodory's criterion is a consequence of Schur's criterion and, conversely, Schur's criterion follows from Carathéodory's.
Bibliography: 10 titles.
Keywords: continued fractions, Schur's algorithm, Carathéodory function, Hankel determinants.
Funding agency Grant number
Ministry of Science and Higher Education of the Russian Federation 075-15-2022-265
This work was performed at the Steklov International Mathematical Center and supported by the Ministry of Science and Higher Education of the Russian Federation (agreement no. 075-15-2022-265).
Received: 03.05.2021 and 23.08.2021
Russian version:
Matematicheskii Sbornik, 2022, Volume 213, Number 11, Pages 5–24
DOI: https://doi.org/10.4213/sm9611
Bibliographic databases:
Document Type: Article
MSC: Primary 30E05, 30H05; Secondary 30B70
Language: English
Original paper language: Russian

§ 1. The classical Carathéodory and Schur criteria

Recall that a holomorphic function $\mathscr F (z)$ in the disc $\mathbb D:=\{|z|<1\}$ is called a Carathéodory function if $\operatorname{Re} \mathscr F (z)\geqslant 0$ for $z\in\mathbb D$, and it is called a Schur function if $|\mathscr F (z)|\leqslant 1$ for $z\in\mathbb D$. We denote the sets of Carathéodory and Schur functions by $\mathfrak B ^{\mathtt{C}}$ and $\mathfrak B ^{\mathtt{S}}$, respectively.

In the sets $\mathfrak B ^{\mathtt{C}}$ and $\mathfrak B ^{\mathtt{S}}$ of Carathéodory and Schur functions authors usually distinguish the disjoint subsets $\mathfrak B_N^{\mathtt{C}}$ and $\mathfrak B_N^{\mathtt{S}}$, $N\in\mathbb Z_+^\infty :=\{\infty,0,1,2,\dots \}$, where

$$ \begin{equation*} \begin{aligned} \, \mathfrak B_0^{\mathtt{C}} &:=\bigl\{\mathscr F (z)\in\mathfrak B ^{\mathtt{C}}\colon \mathscr F (z)\equiv \lambda,\, \operatorname{Re} \lambda =0\bigr\}, \\ \mathfrak B_0^{\mathtt{S}} &:=\bigl\{\mathscr F (z)\in \mathfrak B ^{\mathtt{S}}\colon \mathscr F (z)\equiv \gamma,\, |\gamma |=1\bigr\}, \\ \mathfrak B_N^{\mathtt{C}} &:=\biggl\{\mathscr F (z)\in\mathfrak B ^{\mathtt{C}}\colon \mathscr F(z)=\lambda_0+\sum_{k=1}^N\lambda_k\dfrac{t_k-z}{t_k+z},\, \operatorname{Re} \lambda_0=0,\, \lambda_k>0,\, |t_k|=1, \\ &\qquad k=1,\dots,N,\, t_1,\dots,t_N\text{ are pairwise distinct}\biggr\}, \\ \mathfrak B_N^{\mathtt{S}} &:=\biggl\{\mathscr F (z)\in \mathfrak B ^{\mathtt{S}}\colon \mathscr F (z)=\gamma \prod_{k=1}^N\dfrac{z-e_k}{1-z\overline{e}_k},\,|\gamma |=1,\, e_k\in\mathbb D,\, k=1,\dots,N\biggr\} \end{aligned} \end{equation*} \notag $$
(a bar over a symbol denotes complex conjugation), $N=1,2,\dots$,
$$ \begin{equation*} \mathfrak B_\infty ^{\mathtt{C}}:=\mathfrak B ^{\mathtt{C}}\setminus \biggl(\bigcup_{N\in\mathbb Z_+} \mathfrak B_N^{\mathtt{C}}\biggr)\quad\text{and}\quad \mathfrak B_\infty^{\mathtt{S}}:=\mathfrak B ^{\mathtt{S}}\setminus \biggl(\bigcup_{N\in\mathbb Z_+} \mathfrak B_N^{\mathtt{S}}\biggr). \end{equation*} \notag $$

Let $T(z)$ denote the linear fractional map $T(z)=(1-z)/(1+z)$ taking the unit disc $\mathbb D$ to the right-hand half-plane $\{\operatorname{Re} z>0\}$ and taking this half-plane to the unit disc. It was shown in [1], pp. 229–230, that

$$ \begin{equation} \mathfrak B_N^{\mathtt{C}} =\bigl\{(T\circ \mathscr F )(z)\colon \mathscr F (z)\in\mathfrak B_N^{\mathtt{S}}\bigr\}, \quad\text{for } N\in\mathbb Z_+^\infty, \quad\text{where } (T\circ \mathscr F ) (z):=T(\mathscr F (z)) \end{equation} \tag{1} $$
(the function $\mathscr F (z)\equiv -1$ in $\mathfrak B_0^{\mathtt{S}}$, for which $(T\circ \mathscr F )(z)\equiv \infty$, is an exception here).

Set

$$ \begin{equation} t(\zeta ):= \begin{cases} {\mathtt{S}}&\text{if }\zeta = {\mathtt{C}}, \\ {\mathtt{C}}&\text{if }\zeta = {\mathtt{S}}. \end{cases} \end{equation} \tag{2} $$

Since $(T\circ T)(z)=z$, it follows from (1) (for $\mathscr F (z)\not\equiv -1$) that

$$ \begin{equation} \mathscr F (z)\in \mathfrak B_N^{\zeta} \quad\Longleftrightarrow\quad (T\circ \mathscr F )(z)\in \mathfrak B_N^{t(\zeta )} \quad\text{for } N\in\mathbb Z_+^\infty\quad\text{and} \quad \zeta =\mathtt{C},\mathtt{S}. \end{equation} \tag{3} $$

Let $f (z)=\sum_{k=0}^\infty a_kz^k$ be a formal power series and $I_n$ be the $n\times n $ identity matrix. Set

$$ \begin{equation} A_n^{f}: =\begin{pmatrix} a_0 & a_1 & \dots & a_{n-1}\\ 0 & a_0 & \dots & a_{n-2}\\ \dots & \dots & \dots & \dots \\ 0 & 0 & \dots & a_0 \end{pmatrix}, \qquad \widetilde{A}_n^{f}: =\begin{pmatrix} \overline{a}_0 & 0 & \dots & 0\\ \overline{a}_1 & \overline{a}_0 & \dots & 0\\ \dots & \dots & \dots & \dots \\ \overline{a}_{n-1} & \overline{a}_{n-2} & \dots & \overline{a}_0 \end{pmatrix} \end{equation} \tag{4} $$
and
$$ \begin{equation} M_n^{\zeta;f} := \begin{cases} \det (A_n^{f}+\widetilde{A}_n^{f})&\text{for }\zeta = {\mathtt{C}}, \\ \det(I_n-A_n^{f}\widetilde{A}_n^{f}) &\text{for }\zeta = {\mathtt{S}}, \end{cases} \qquad n=1,2,\dots\,. \end{equation} \tag{5} $$

Note that the matrices $A_n^{f}$ and $\widetilde{A}_n^{f}$ and the determinants $M_n^{\zeta;f}$ depend only on the first $n$ coefficients $a_0,\dots,a_{n-1}$ of the series $f (z)$. In particular,

$$ \begin{equation*} \begin{gathered} \, A_n^{f}=A_n^{f_m}, \quad\text{where } m\geqslant n\quad\text{and} \\ f_m(z)=\sum_{k=0}^{m-1} a_kz^k \text{ is the $(m-1)$st partial sum of the series } f(z). \end{gathered} \end{equation*} \notag $$

In the set of formal power series we distinguish the following subsets, where $N\in\mathbb Z_+^\infty$ and $\zeta =\mathtt{C}, \mathtt{S}$:

$$ \begin{equation} {\mathscr A} ^{\zeta}_N :=\bigl\{f(z)\colon M_p^{\zeta;f}>0, \, p=1,\dots,N,\, M_{N+p}^{\zeta;f}=0,\, p=1,2,\dots \bigr\} \end{equation} \tag{6} $$
(if $N=0$, then there are no inequalities of the form $M_p^{\zeta;f} > 0$ for $p=1,\dots,N$, and if $N=\infty$, then there are no equalities of the form $M_{N+p}^{\zeta;f} = 0$ for $p=1,2,\dots$).

If a formal power series $f(z)$ is the Taylor series of a Carathéodory or a Schur function, then we write $f(z)\lessdot \mathfrak B ^{\mathtt{C}}$ or $f(z)\lessdot \mathfrak B ^{\mathtt{S}}$, respectively. The expressions $f(z)\lessdot \mathfrak B_N^{\mathtt{C}}$ and $f(z)\lessdot \mathfrak B_N^{\mathtt{S}}$, $N\in\mathbb Z_+^\infty $, are interpreted similarly.

In [2] and [3] necessary and sufficient conditions were found which ensure that a given formal power series is the Taylor series of a Carathéodory function. A similar result for Schur functions was obtained in [1]. For brevity we formulate Carathéodory’s and Schur’s criteria as one statement, in which $\zeta =\mathtt{C}$ corresponds to Carathéodory’s criterion and $\zeta =\mathtt{S}$ to Schur’s criterion.

Carathéodory-Schur Criterion. Let $f (z)\!=\!\sum_{k=0}^\infty a_kz^k$ be a formal power series. Then using the notation (4) and (5),

$$ \begin{equation*} f(z)\lessdot \mathfrak B_N^{\zeta}\quad \Longleftrightarrow\quad f(z)\in{\mathscr A}_N^{\zeta}\quad\textit{for } N\in\mathbb Z_+^\infty \quad \textit{and}\quad \zeta =\mathtt{C}, \mathtt{S}. \end{equation*} \notag $$

In § 2 we establish the following result.

Theorem 1. Let $f (z)=\sum_{k=0}^\infty a_kz^k$ be a formal power series such that $a_0\neq -1$. Then using the notation (4) and (5), the following equalities hold:

$$ \begin{equation} M_n^{t(\zeta );T\circ f}=\frac{2^n}{|1+a_0|^{2n}}M_n^{\zeta;f}, \qquad n=1,2,\dotsc\,. \end{equation} \tag{7} $$

From (7) we obtain

$$ \begin{equation*} M_n^{t(\zeta );T\circ f}=0\ \Longleftrightarrow\ M_n^{\zeta;f}=0 \quad\text{and} \quad M_n^{t(\zeta );T\circ f}>0\ \Longleftrightarrow\ M_n^{\zeta;f}>0. \end{equation*} \notag $$
Hence for $a_0\neq -1$,
$$ \begin{equation} f(z)\in{\mathscr A}_N^{\zeta}\quad\Longleftrightarrow\quad(T\circ f)(z)\in{\mathscr A}_N^{t(\zeta )}. \end{equation} \tag{8} $$
Taking (3) and (8) into account Theorem 1 means that, with the exception of the case when $a_0=-1$, which can easily be examined separately, Carathéodory’s and Schur’s criteria are equivalent in the following sense: Schur’s criterion follows directly from Carathéodory’s criterion and Theorem 1 and, conversely, Carathéodory’s criterion is a direct consequence of Schur’s criterion and Theorem 1.

In § 3 we state and in § 4 we prove an analogue of Carathéodory’s and Schur’s criteria for functions defined (with multiplicities) at points $e_1,e_2,\dots$ in $\mathbb D$. Its particular case in the situation of functions defined by the values of their derivatives at zero (that is, for $e_1=e_2=\dots =0$) reads as follows.

Refinement of the Carathéodory-Schur Criterion for $N\in {\mathbb Z}_+$. Let $f (z)=\sum_{k=0}^\infty a_kz^k$ be a formal power series. Then

$$ \begin{equation*} f(z)\lessdot \mathfrak B_N^{\zeta} \quad\Longleftrightarrow\quad f(z)\in{\mathscr A}_N^{\zeta} \quad\Longleftrightarrow\quad f(z)\in\hat{\mathscr{A}}_N^{\zeta} \quad\Longleftrightarrow\quad f(z)\in\breve{{\mathscr A}}_N^{\zeta}, \end{equation*} \notag $$
where $N\in\mathbb Z_+$, $\zeta ={\mathtt{C}}, {\mathtt{S}}$,
$$ \begin{equation} \hat{\mathscr{A}}_N^{\zeta} :=\bigl\{f(z)\colon M_p^{\zeta;f}>0,\, p=1,\dots,N, \,M_{N+1}^{\zeta;f}= M_{N+2p}^{\zeta;f}=0,\, p=1,2,\dots \bigr\} \end{equation} \tag{9} $$
and
$$ \begin{equation} \breve{{\mathscr A}}^{\zeta}_N :=\bigl\{f(z)\colon M_p^{\zeta;f}>0,\, p=1,\dots,N, \, M_{N+1}^{\zeta;f}=M_{N+2p}^{\zeta;f_{N+p+1}}=0,\, p=1,2,\dots \bigr\} \end{equation} \tag{10} $$
(there are no inequalities of the form $M_p^{\zeta;f}>0$ for $N=0$), where $f_{n}(z)$ is the $(n-1)$st partial sum of the series $f(z)$, $n=1,2,\dotsc$ .

Note that the definition (9) of the sets $\hat{\mathscr{A}}_N^{\zeta}$ involves no equalities of the form

$$ \begin{equation*} M_{N+2p+1}^{\zeta;f} = 0, \qquad p=1,2,\dots, \quad N\in\mathbb Z_+, \end{equation*} \notag $$
which are present in the definition (6) of the sets ${\mathscr A} ^{\zeta}_N$. This means that for $N\in\mathbb Z_+$ the hypotheses in the implications
$$ \begin{equation*} f(z)\lessdot \hat{\mathscr{A}}_N^{\zeta} \quad\Longrightarrow\quad f(z)\in \mathfrak B_N^{\zeta} \end{equation*} \notag $$
form subsets of the hypotheses in the implications
$$ \begin{equation*} f(z)\lessdot {\mathscr A}_N^{\zeta} \quad\Longrightarrow\quad f(z)\in \mathfrak B_N^{\zeta}, \end{equation*} \notag $$
which are parts of Carathéodory’s and Schur’s criteria. Also note that from the standpoint of computations, the determinants $M_{{N+2p}}^{\zeta;f_{N+p+1}}$ in the definition (10) of the sets $\breve{{\mathscr A}}^{\zeta}_N$ are slightly simpler than the determinants $M_{{N+2p}}^{\zeta;f}$ in the definition (9) of the sets $\hat{\mathscr{A}}_N^{\zeta}$, because in the calculation of $M_{{N+2p}}^{\zeta;f_{N+p+1}}$ the coefficients $a_{N+p+1},\dots, a_{N+2p-1}$ involved in the determinants $M_{{N+2p}}^{\zeta;f}$, $p=2,3,\dots$, can be replaced by zeros (or any other numbers as shown below).

In [4] we proposed a proof of Schur’s criterion in terms of two-point Hankel determinants of the power series $f(z)=\sum_{k=0}^\infty a_kz^k$ ($a_0\neq 0$) with centre $z=0$ and the associated series

$$ \begin{equation*} f^* (z):=\bigl(\overline{f(\overline{z^{-1}})}\bigr)^{-1}=\sum_{k=0}^\infty a_k^* z^{-k}\, \end{equation*} \notag $$
with centre $z=\infty$; the latter comes into play because of the well-known relation between convergents with even and odd indices of the Schur continued fraction corresponding to $f(z)$. It turns out that two-point Hankel determinants of order $n$ for $f(z)$ and $f^*(z)$ coincide with the Schur determinants $M_n^{\mathtt{S};f}$ up to a factor of $(-1)^n\overline{a}_0^{\,n}$. Using this fact and a two-point analogue of Polya’s theorem (see [5] and [6]) on an upper estimate for the capacity of the singularities of a meromorphic function, in [7] we investigated the convergence and boundary behaviour of a limit periodic Schur continued fraction.

It is known (for instance, see [8]) that Schur’s classical algorithm described in [1] has a continuous analogue, which enables one to interpolate a function $\mathscr F (z)\in\mathfrak B ^{\mathtt{S}}$ by convergents of a multipoint Schur continued fraction at a prescribed sequence of points in the unit disc, rather than at the single point $z=0$ (with multiplicity).

In [9], for $E_n:=\{e_1,\dots,e_n\}\subset \mathbb D$ we discovered quantities $M_{E_n}^{\mathtt{S};f}$ (coinciding with the Schur determinants $M_n^{\mathtt{S};f}$ for $e_1=\dots =e_n=0$) in terms of which we could state a multipoint analogue of Schur’s criterion. The proof of this multipoint analogue of Schur’s criterion stated in [9] is based on a multipoint version of Schur’s algorithm; it was presented in [10].

In this paper we define certain quantities $M_{E_n}^{\mathtt{C};f}$ (coinciding with the Carathéodory determinants $M_n^{\mathtt{C};f}$ for $e_1=\dots =e_n=0$), and we state and prove a multipoint version of Carathéodory’s criterion in terms of these quantities.

As in the classical case, the multipoint Carathéodory criterion and the multipoint Schur criterion are equivalent in the following sense: each of them is a direct consequence of the other and relations between the quantities $M_{E_n}^{\mathtt{C};T\circ f}$ and $M_{E_n}^{\mathtt{S};f}$ (see Theorem 2 in § 2, of which Theorem 1 is a special case). Bearing in mind the proof of the multipoint Schur criterion in [10], the proof of the multipoint Carathéodory criterion in § 4 is based on the equivalence of the two criteria that we have revealed. Apart from the proof of Theorem 2, it reduces to verifying the conditions $f(z)\neq -1$ for $z\in E_n$, $n=1,2,\dots$, from the hypotheses of Theorem 2.

§ 2. Multipoint analogues of Carathéodory and Schur determinants

Let $F(z)$ be a function defined (with multiplicities taken into account) at the points in a set $E_n:=\{e_1,\dots,e_n\}$ so that if $\nu_j$ is the multiplicity of $e_j$ in the set $\{e_1,\dots,e_j\}$, $j=1,\dots,n$, then the $(\nu_j-1)$st derivative $F ^{(\nu_j-1)}(e_{j})$ of $F(z)$ at the point $e_{j}$ is defined.

Recall that if $F(z)$ and $G(z)$ are functions defined (with multiplicities) at the points in $E_n$, then the functions $(F\pm G)(z)$, $(FG)(z)$ and $(F/G)(z)$ are also defined there with multiplicities (provided that $G(z)\neq 0$ for $z\in E_n$ in the last case). In particular, if $F(z)\neq -1$ for $z\in E_n$, then the function $(T\circ F)(z)=(1-F(z))/ (1+F(z))$ is also defined with multiplicities.

It is easy to see that any function $\mathbf{F}(z)\in H(E_n)$ (that is, any function holomorphic in a neighbourhood of $E_n=\{e_1,\dots,e_n\}$) is defined with multiplicities at the points in any set $E^{j_1,\dots,j_p}:=\{e_{j_1},\dots,e_{j_p}\}$, where $e_{j_q}\in E_n$, $q=1,\dots,p$, $p\in\mathbb N$, and for any function $F(z)$ defined with multiplicities at the points in $E_n$ there exists a function $\mathbf{F}(z)\in H(E_n)$ that is equal to $F(z)$ on $E_n$ with multiplicities taken into account. In particular, there exists a polynomial $\mathbf{F}_n(z)$ of degree at most $n-1$ (the Lagrange interpolation polynomial) such that

$$ \begin{equation*} \mathbf{F}_n^{(\nu_j-1)}(e_j)=F^{(\nu_j-1)}(e_j), \qquad j=1,\dots,n, \end{equation*} \notag $$
where $\nu_j$ is the multiplicity of $e_j$ in $E_{n}$.

For the convenience of references we state the following obvious result.

Proposition 1. Let $F(z)$ be a function defined at the points in $E_n$ with multiplicities taken into account and such that $F(z)\neq -1$, $z\in E_n$. Let $\mathbf{F}(z)\in H(E_n)$ interpolate $F(z)$ with multiplicities at the points in $E_n$. Then the function $(T\circ \mathbf{F})(z)\in H(E_n)$ interpolates $(T\circ F)(z)$ with multiplicities at the points in $E_n$.

Let

$$ \begin{equation*} E_n=\{e_1,\dots,e_n\}\quad\text{and} \quad E^{j_1,\dots,j_p}:=\{e_{j_1},\dots,e_{j_p}\} \quad\text{for } 1\leqslant j_1<\dots <j_p\leqslant n. \end{equation*} \notag $$

By the restriction of a function $F(z)$, defined at the points in $E_n$ with multiplicities taken into account, to the set $E^{j_1,\dots,j_p}$ we mean the function defined withe multiplicities at the points in $E^{j_1,\dots,j_p}$ by means of the Lagrange interpolation polynomial $\mathbf{F}_n(z)$. For the Lagrange interpolation polynomial $\mathbf{F}^{j_1,\dots,j_p}(z)$ of this restriction of $F(z)$ to $E^{j_1,\dots,j_p}$ we have

$$ \begin{equation*} \frac{\mathbf{F}^{j_1,\dots,j_p}(z)-\mathbf{F}_n(z)}{(z- e_{j_1})\dotsb(z-e_{j_p})}\in H(E^{j_1,\dots,j_p}). \end{equation*} \notag $$

Note that we do not need to calculate the Lagrange polynomial $\mathbf{F}_n(z)$ to define the restriction; we can do this in terms of $F(z)$ itself without using this polynomial. We only mention the Lagrange polynomial to avoid explaining some subtleties related to the fact that the points $e_{j_q}$, $q=1,\dots,p$, can have different multiplicities in $E_n$ and $E^{j_1,\dots,j_p}$.

The multipoint analogue of the Carathéodory-Schur criterion that we establish here will be stated in terms of certain quantities $M_{E_n}^{\mathtt{C};F}$ and $M_{E_n}^{\mathtt{S};F}$. To define these we introduce some further notation.

Notation 1. Assume that an $n$-point set $E_n=\{e_1,\dots,e_n\}$ is rearranged in the form $E_n:=\bigl\{\{e_1\} ^{r_1},\dots,\{e_k\} ^{r_k}\bigr\}$, where $e_1,\dots,e_k$ are pairwise distinct, ${r_1+\dots +r_k}=n$ and $\{e\}^{r}:=\{\underbrace{e,\dots,e}_{r}\}$.

Let $F(z)$ be a function defined at the points in $E_n$ with multiplicities taken into account, and let $\varphi_p (z):=z^p$, $p=0,1,\dots$ (we set $z^0:=1$ even for $z=0$). Bearing in mind that each $(\varphi_pF)(z)$ is a function defined with multiplicities at the points in $E_n$, we let $A_{E_n}^F$ and $\widetilde{A}_{E_n}^F$ denote the matrices

$$ \begin{equation} { \begin{pmatrix} \dfrac{(\varphi_0F)(e_1)}{0!} &\dots & \dfrac{(\varphi_0F)^{(r_1-1)}(e_1)}{(r_1{-}\,1)!} &\dots & \dfrac{(\varphi_0F)(e_k)}{0!} &\dots & \dfrac{(\varphi_0F)^{(r_k-1)}(e_k)}{(r_k{-}\,1)!} \\ \dots &\dots &\dots &\dots &\dots &\dots &\dots \\ \dfrac{(\varphi_{n-1}F)(e_1)}{0!} &\dots & \dfrac{(\varphi_{n-1}F)^{(r_1-1)}(e_1)}{(r_1{-}\,1)!} &\dots & \dfrac{(\varphi_{n-1}F)(e_k)}{0!} &\dots & \dfrac{(\varphi_{n-1}F)^{(r_k-1)}(e_k)}{(r_k{-}\,1)!} \end{pmatrix}, } \end{equation} \tag{11} $$
and
$$ \begin{equation} { \begin{pmatrix} \dfrac{\overline{(\varphi_{n-1}F)^{(r_k-1)}(e_k)}}{(r_k{-}\,1)!} &\dots & \dfrac{\overline{(\varphi_{n-1}F)(e_k)}}{0!} &\dots & \dfrac{\overline{(\varphi_{n-1}F)^{(r_1-1)}(e_1)}}{(r_1{-}\,1)!} &\dots & \dfrac{\overline{(\varphi_{n-1}F)(e_1)}}{0!} \\ \dots &\dots &\dots &\dots &\dots &\dots &\dots \\ \dfrac{\overline{(\varphi_{0}F)^{(r_k-1)}(e_k)}}{(r_k{-}\,1)!} &\dots & \dfrac{\overline{(\varphi_{0}F)(e_k)}}{0!} &\dots & \dfrac{\overline{(\varphi_{0}F)^{(r_1-1)}(e_1)}}{(r_1{-}\,1)!} &\dots & \dfrac{\overline{(\varphi_{0}F)(e_1)}}{0!} \end{pmatrix}, } \end{equation} \tag{12} $$
respectively, and we set
$$ \begin{equation} W_{E_n}:=\det \begin{pmatrix} A_{E_n}^{\varphi_0} & \widetilde{A}_{E_n}^{\varphi_n} \\ A_{E_n}^{\varphi_n} & \widetilde{A}_{E_n}^{\varphi_0} \end{pmatrix}, \end{equation} \tag{13} $$
$$ \begin{equation} M_{E_n}^{\mathtt{C};F}: =\dfrac{\det \begin{pmatrix} A_{E_n}^{\varphi_0} & \widetilde{A}_{E_n}^{\varphi_0} \\ -A_{E_n}^{F} & \widetilde{A}_{E_n}^{F} \end{pmatrix}}{W_{E_n}} \quad\text{and}\quad M_{E_n}^{\mathtt{S};F}: =\dfrac{\det \begin{pmatrix} A_{E_n}^{\varphi_0} & \widetilde{A}_{E_n}^{F} \\ A_{E_n}^{F} & \widetilde{A}_{E_n}^{\varphi_0} \end{pmatrix}}{W_{E_n}}. \end{equation} \tag{14} $$

Note the following: we can obtain $\widetilde{A}_{E_n}^F$ from $A_{E_n}^F$ by performing complex conjugation and reverting the order of both columns and rows (if $A_{E_n}^F \,{=}\,(a_{k,j})_{k,j=1,\dots,n}$, then $\widetilde{A}_{E_n}^F =(\overline{a}_{n+1-k,n+1-j})_{k,j=1,\dots,n}$); $W_{E_n}\neq 0$ (provided that $E_n\subset\mathbb D$); each quantity $M_{E_{n}}^{\mathtt{C};F}$ and $M_{E_{n}}^{\mathtt{S};F}$ is real and invariant under rearrangements of the points in $E_n=\{e_1,\dots,e_n\}$.

Also note that we can replace $F(z)$ in (11) and (12) by any function equal to $F (z)$ on $E_n$ with multiplicities taken into account (for instance, by the Lagrange interpolation polynomial $\mathbf{F}_n(z)$), and the constants $0!, \dots,(r_j-1)!$ can be replaced by arbitrary nonzero constants (for instance, by ones as in [10], in the definition of $M_{E_{n}}^{\mathtt{S};F}$), because these constants are multiplied out of determinants and occur in a similar way in the numerators and denominators of the quantities $M_{E_n}^{\mathtt{C};F}$ and $M_{E_n}^{\mathtt{S};F}$ as defined in (14). Our choice of $0!, \dots,(r_j-1)!$ in the definitions (11) and (12) is explained by the convenience of comparison of $M_{E_n}^{\mathtt{C};F}$ and $M_{E_n}^{\mathtt{S};F}$, in the special case when $E_n=\{0\} ^n$ (so that all points in the $n$-point set $E_n$ coincide with zero) and $F (z)=\sum_{k=0}^{n-1} a_kz^k$, with the quantities $M_{n}^{\mathtt{C};F}$ and $M_{n}^{\mathtt{S};F}$ introduced before the Carathódory-Schur criterion (see (5)) in terms of the matrices $A_n^F$ and $\widetilde{A}_n^F$ (see (4)). Namely, the following holds.

Proposition 2. Let $n\in\mathbb N$, $E_n=\{0\} ^n$, $F (z)=\sum_{k=0}^{n-1} a_kz^k$ and $\zeta =\mathtt{C}, \mathtt{S} $. Then $M_{E_n}^{\zeta;F}=M_{n}^{\zeta;F}$, where the quantities $M_{E_n}^{\zeta;F}$ are defined by (11)(14) and the $M_{n}^{\zeta;F}$ are defined by (4) and (5).

In fact, if $E_n=\{0\} ^n$ and $F (z)=\sum_{k=0}^{n-1} a_kz^k$, then from (11)(14) and (4), (5) we obtain

$$ \begin{equation*} \begin{gathered} \, \frac{(\varphi_pF)^{(r)}(0)}{r!}= \begin{cases} 0 \quad &\text{for }0\leqslant r<p\leqslant n-1, \\ a_{r-p}\quad &\text{for } 0\leqslant p\leqslant r\leqslant n-1, \end{cases} \\ A_{E_n}^{F}=\begin{pmatrix} a_0 & a_1 & \dots & a_{n-1} \\ 0 & a_0 & \dots & a_{n-2} \\ \dots & \dots & \dots & \dots \\ 0 & 0 & \dots & a_0 \end{pmatrix} {=}\,A_n^F, \qquad \widetilde{A}_{E_n}^{F} = \begin{pmatrix} \overline{a}_0 & 0 & \dots & 0 \\ \overline{a}_1 & \overline{a}_0 & \dots & 0 \\ \dots & \dots & \dots & \dots \\ \overline{a}_{n-1} & \overline{a}_{n-2} & \dots & \overline{a}_0 \end{pmatrix} {=}\,\widetilde{A}_{n}^{F}, \\ A_{E_n}^{\varphi_0}=\begin{pmatrix} 1 & 0 & \dots & 0 \\ 0 & 1 & \dots & 0 \\ \dots & \dots & \dots & \dots \\ 0 & 0 & \dots & 1 \end{pmatrix} {=}\,\widetilde{A}_{E_n}^{\varphi_0}, \qquad A_{E_n}^{\varphi_n}=\begin{pmatrix} 0 & 0 & \dots & 0 \\ 0 & 0 & \dots & 0 \\ \dots & \dots & \dots & \dots \\ 0 & 0 & \dots & 0 \end{pmatrix} {=}\,\widetilde{A}_{E_n}^{\varphi_n}, \\ W_{E_n}=\det \begin{pmatrix} I_n & O_n \\ O_n & I_n \end{pmatrix} =1, \end{gathered} \end{equation*} \notag $$
where $I_n$ and $O_n$ are the $ n\times n $ identity and zero matrices,
$$ \begin{equation} M_{E_n}^{\mathtt{C};F}: =\det \begin{pmatrix} I_n & I_n \\ -A_n^F & \widetilde{A}_n^F \end{pmatrix}=\det(A_n^F +\widetilde{A}_n^F)=:M_{n}^{\mathtt{C};F} \end{equation} \tag{15} $$
and
$$ \begin{equation} M_{E_n}^{\mathtt{S};F}: =\det \begin{pmatrix} I_n & \widetilde{A}_n^F \\ A_n^F & I_n \end{pmatrix}=\det(I_n-A_n^F \widetilde{A}_n^F)=:M_{n}^{\mathtt{S};F}. \end{equation} \tag{16} $$
In (15) and (16) we have used the well-known equality (see, for instance, [1], § 5)
$$ \begin{equation*} \det\begin{pmatrix} P & Q \\ R & S \end{pmatrix} =\det(PS-RQ) \end{equation*} \notag $$
for $ n\times n $-matrices $P,Q,R$ and $S$ such that $PR=RP$.

A very useful auxiliary result is as follows.

Proposition 3. For $n\in\mathbb N$ let $E_n=\{e_1,\dots,e_n\}\subset\mathbb D$, let $F(z)$ be a function defined at the points in $E_n$ with multiplicities taken into account, $\mathbf{F}(z)\in H(E_n)$ be a function interpolating $F(z)$ with multiplicities at the points in $E_n$, and let $\zeta =\mathtt{C},\mathtt{S}$. Then

$$ \begin{equation*} M_{E_n}^{\zeta;F}=\lim_{\varepsilon_1\to 0}\dotsb \lim_{\varepsilon_n\to 0}M_{E_{n;\varepsilon_1,\dots,\varepsilon_n}}^{\zeta;\mathbf{F}}, \quad \textit{where } E_{n;\varepsilon_1,\dots,\varepsilon_n} :=\{e_1+\varepsilon_1,\dots,e_n+\varepsilon_n\}. \end{equation*} \notag $$

It was shown in [10], Proposition 5, that for $\zeta =\mathtt{S}$ Proposition 3 is a simple consequence of Taylor’s formula. Moreover, the arguments in [10] can also be used for $\zeta =\mathtt{C} $ almost word for word.

The next theorem extends Theorem 1 to the multipoint case.

Theorem 2. For $n\in\mathbb N$ let $E_n=\{e_1,\dots,e_n\}\subset\mathbb D$, and let $F(z)$ be a function defined at the points in $E_n$ with multiplicities taken into account and such that $F(z)\neq-1$ for $z\in E_n$. Then, using the notation (11)(14), the following equalities hold for the function $(T\circ F)(z)$ defined with multiplicities at the points in $E_n$:

$$ \begin{equation} M_{E_n}^{t(\zeta );T\circ F}=\frac{2^n}{\prod_{k=1}^n|1+F(e_k)|^{2}}M_{E_n}^{\zeta;F}, \qquad \zeta =\mathtt{C}, \mathtt{S}. \end{equation} \tag{17} $$
In particular,
$$ \begin{equation*} M_{E_n}^{t(\zeta );T\circ F}=0 \ \Longleftrightarrow\ M_{E_n}^{\zeta;F}=0\quad\textit{and} \quad M_{E_n}^{t(\zeta );T\circ F}>0 \ \Longleftrightarrow\ M_{E_n}^{\zeta;F}>0. \end{equation*} \notag $$

Proof. First assume that the points $e_1,\dots,e_n$ in $E_n$ are pairwise distinct, and set for brevity
$$ \begin{equation*} F_k:=F(e_k)\quad\text{and} \quad G_k:=(T\circ F)(e_k)=\frac{1-F_k}{1+F_k}, \qquad k=1,\dots,n. \end{equation*} \notag $$

Then the following chain of equalities holds:

$$ \begin{equation*} \begin{aligned} \, &W_{E_n}M_{E_{n}}^{\mathtt{C};T\circ F} =\det \begin{pmatrix} e_1^0 & \dots & e_n^0 & \overline{e_n^{n-1}} & \dots & \overline{e_1^{n-1}} \\ \dots & \dots & \dots & \dots & \dots & \dots \\ e_1^{n-1} & \dots & e_n^{n-1} & \overline{e_n^{0}} & \dots & \overline{e_1^{0}} \\ -e_1^0G_1 & \dots & -e_n^0G_n & \overline{e_n^{n-1}G_n} & \dots & \overline{e_1^{n-1}G_1} \\ \dots & \dots & \dots & \dots & \dots & \dots \\ -e_1^{n-1}G_1 & \dots & -e_n^{n-1}G_n & \overline{e_n^{0}G_n} & \dots & \overline{e_1^{0}G_1} \end{pmatrix} \\ &=\frac{\det { \begin{pmatrix} e_1^0(1+F_1) & \dots & e_n^0(1+F_n) & \overline{e_n^{n-1}(1+F_n)} & \dots & \overline{e_1^{n-1}(1+F_1)} \\ \dots & \dots & \dots & \dots & \dots & \dots \\ e_1^{n-1}(1+F_1) & \dots & e_n^{n-1}(1+F_n) & \overline{e_n^{0}(1+F_n)} & \dots & \overline{e_1^{0}(1+F_1)} \\ -e_1^0(1-F_1) & \dots & -e_n^0(1-F_n) & \overline{e_n^{n-1}(1-F_n)} & \dots & \overline{e_1^{n-1}(1-F_1)} \\ \dots & \dots & \dots & \dots & \dots & \dots \\ -e_1^{n-1}(1-F_1) & \dots & -e_n^{n-1}(1-F_n) & \overline{e_n^{0}(1-F_n)} & \dots & \overline{e_1^{0}(1-F_1)} \end{pmatrix}}}{\prod_{k=1}^n|1+F_k|^{2}} \\ &=\frac{\det { \begin{pmatrix} 2e_1^0 & \dots & 2e_n^0 & 2\overline{e_n^{n-1}F_n} & \dots & 2\overline{e_1^{n-1} F_1} \\ \dots & \dots & \dots & \dots & \dots & \dots \\ 2e_1^{n-1} & \dots & 2e_n^{n-1} & 2\overline{e_n^{0}F_n} & \dots & 2\overline{e_1^{0} F_1} \\ -e_1^0(1-F_1) & \dots & -e_n^0(1-F_n) & \overline{e_n^{n-1}(1-F_n)} & \dots & \overline{e_1^{n-1}(1-F_1)} \\ \dots & \dots & \dots & \dots & \dots & \dots \\ -e_1^{n-1}(1-F_1) & \dots & -e_n^{n-1}(1-F_n) & \overline{e_n^{0}(1-F_n)} & \dots & \overline{e_1^{0}(1-F_1)} \end{pmatrix}}}{\prod_{k=1}^n|1+F_k|^{2}} \\ &=C_{E_n}^F\det \begin{pmatrix} e_1^0 & \dots & e_n^0 & \overline{e_n^{n-1}F_n} & \dots & \overline{e_1^{n-1}F_1} \\ \dots & \dots & \dots & \dots & \dots & \dots \\ e_1^{n-1} & \dots & e_n^{n-1} & \overline{e_n^{0}F_n} & \dots & \overline{e_1^{0}F_1} \\ e_1^0F_1 & \dots & e_n^0F_n & \overline{e_n^{n-1}} & \dots & \overline{e_1^{n-1}} \\ \dots & \dots & \dots & \dots & \dots & \dots \\ e_1^{n-1}F_1 & \dots & e_n^{n-1}F_n & \overline{e_n^{0}} & \dots & \overline{e_1^{0}} \end{pmatrix} =C_{E_n}^FW_{E_n}M_{E_{n}}^{\mathtt{S};F}, \end{aligned} \end{equation*} \notag $$
where $C_{E_n}^F:=2^n/\prod_{k=1}^n|1+F(e_k)|^{2}$.

Some comments regarding this chain of equalities are in order here. The first equality follows from the definition (11)(14) of $M_{E_n}^{\mathtt{C};T\circ F}$; the second (taking the inequality $F_k\neq -1$ and the equalities $G_k=(1-F_k)/(1+F_k)$, $k=1,\dots,n$, into account) is deduced by multiplying and dividing by $\prod_{k=1}^n|1+F_k|^{2}$; the third is obtained by subtracting the $(n+k)$th row from the $k$th ($k=1,\dots,n$) in the determinant and the fourth by multiplying out the constant coefficient 2 from the first $n$ rows in the determinant, introducing the constant $C_{E_n}^F$ and adding the $k$th row of the transformed determinant to the ${(n+k)}$th; the fifth equality follows from the definition (11)(14) of $M_{E_n}^{\mathtt{S};F}$.

Since $W_{E_n}\neq 0$ for $E_n\subset\mathbb D$, we have thus completed the proof of (17) for $\zeta =\mathtt{S} $ and pairwise distinct points $e_1,\dots,e_n$.

In the general case let $\mathbf{F}(z)\in H(E_n)$ denote the function interpolating $F(z)$ with multiplicities at the points in $E_n$. Using Propositions 1 and 3, where the infinitesimals $\varepsilon_1,\dots,\varepsilon_n$ are selected so that the set $E_{n;\varepsilon_1,\dots,\varepsilon_n}$ consists of pairwise distinct points, from equality (17) for $\zeta =\mathtt{S} $ and pairwise distinct points, which we have established, we obtain

$$ \begin{equation*} \begin{aligned} \, M_{E_n}^{\mathtt{C};T\circ F} &=\lim_{\varepsilon_1\to 0}\dotsb \lim_{\varepsilon_n\to 0}M_{E_{n;\varepsilon_1,\dots,\varepsilon_n}}^{\mathtt{C};T\circ \mathbf{F}} \\ &=\lim_{\varepsilon_1\to 0}\dotsb \lim_{\varepsilon_n\to 0} C_{E_{n;\varepsilon_1,\dots,\varepsilon_n}}^{\mathbf{F}} M_{E_{n;\varepsilon_1,\dots,\varepsilon_n}}^{\mathtt{S};\mathbf{F}} =C_{E_n}^FM_{E_n}^{\mathtt{S};F}. \end{aligned} \end{equation*} \notag $$
This is the same as (17) for $\zeta =\mathtt{S} $ in the general case.

Replacing $F(z)$ by $(T\circ F)(z)$ in equality (17) for $\zeta =\mathtt{S} $, which we have already proved, and bearing in mind that

$$ \begin{equation*} (T\circ F)(z)\neq -1, \quad z\in E_n, \qquad (T\circ T\circ F)(z)=F(z)\quad\text{and} \quad C_{E_n}^{T\circ F}=(C_{E_n}^{F})^{-1}, \end{equation*} \notag $$

we obtain

$$ \begin{equation*} M_{E_n}^{\mathtt{C};F}=C_{E_n}^{T\circ F}M_{E_n}^{\mathtt{S};T\circ F}=(C_{E_n}^{F})^{-1}M_{E_n}^{\mathtt{S};T\circ F}, \end{equation*} \notag $$
which is (17) for $\zeta =\mathtt{C} $.

Theorem 2 is proved.

By Proposition 2, Theorem 1 is the special case of Theorem 2 for $E_n=\{0\} ^n$.

§ 3. Multipoint Carathéodory-Schur criterion

Let $e_1,e_2,\dots $ be an infinite sequence of points in the disc $\mathbb D$ and $F(z)$ be a function defined with multiplicities at these points. We can define the restrictions $F_n(z)$ of $F$ to the sets $E_{n}:=\{e_1,\dots,e_n\}$, $n=1,2,\dots $, in the natural way; in their turn these latter define for (sufficiently large) suitable $n$ the restrictions to any sets of the form $\{e_{j_1},\dots,e_{j_p}\}$, $1\leqslant j_1<\dots <j_p<\infty$ (see § 2).

Definition. We say that a function $F(z)$ defined at an infinite sequence of points $e_1,e_2,\dots $ in $\mathbb D$ with multiplicities taken into account admits an extension to a function in $\mathfrak B_N^\zeta$ ($N\in\mathbb Z_+^\infty$ and $\zeta =\mathtt{C}, \mathtt{S}$), and we write $F(z)\lessdot \mathfrak B_N^\zeta$ if there exists a function $\mathscr F_N^\zeta (z)\in \mathfrak B_N^\zeta$ such that

$$ \begin{equation} (\mathscr F_N^\zeta)^{(\nu_n-1)}(e_n)=F^{(\nu_n-1)}(e_n) \quad \text{for all } n=1,2,\dots, \end{equation} \tag{18} $$
where $\nu_n$ is the multiplicity of $e_n$ in the set $E_{n}=\{e_1,\dots,e_n\}$.

If $F(z)$ is such that $F(e_n)\neq -1$, $n=1,2,\dots $, then we can also define the function $(T\circ F)(z)$ at the points $e_1,e_2,\dots$ . Furthermore, by Proposition 1 equalities (18) are equivalent to

$$ \begin{equation} (T\circ \mathscr F_N^\zeta )^{(\nu_n-1)}(e_n)=(T\circ F)^{(\nu_n-1)}(e_n) \quad \text{for all } n=1,2,\dotsc\,. \end{equation} \tag{19} $$

Hence, as

$$ \begin{equation*} F(z)\lessdot \mathfrak B_N^{\zeta}\quad\text{for } N\in\mathbb Z_+^\infty\quad\text{and} \quad \zeta =\mathtt{C},\mathtt{S} \, \quad\Longrightarrow\quad F(e_n)\neq -1, \quad n=1,2,\dots \end{equation*} \notag $$
(except when $\zeta ={\mathtt{S}}$ and $F(z)\equiv -1$), taking (19) and (3) into account we obtain the following.

Proposition 4. Let $F(z)$ be a function defined with multiplicities at points $e_1, e_2,\dots$ in the disc $\mathbb D$. Then the following implications hold for $N\in\mathbb Z_+^\infty$:

$$ \begin{equation*} \begin{gathered} \, F(z)\lessdot \mathfrak B_N^{\mathtt{C}} \quad\Longrightarrow\quad (T\circ F)(z)\lessdot \mathfrak B_N^{\mathtt{S}}, \qquad (T\circ F)(e_n)\neq -1, \quad n=1,2,\dots, \\ F(z)\lessdot \mathfrak B_N^{\mathtt{S}}, \qquad F(z)\not\equiv -1 \quad\Longrightarrow\quad (T\circ F)(z)\lessdot \mathfrak B_N^{\mathtt{C}}. \end{gathered} \end{equation*} \notag $$

Apart from the quantities $M_{E_n}^{\mathtt{C};F}$ and $M_{E_n}^{\mathtt{S};F}$ introduced in § 2, to state the multipoint analogue of Schur’s criterion for functions defined with multiplicities at points $e_1,e_2,\dots $ in $\mathbb D$, which was established in [10] (and which we prove in our § 4), we need the following definitions.

Notation 2. Let $n\in\mathbb N$ and $N\in\mathbb Z_+$, let $n\geqslant N+2$, and let $\nu_n$ be the multiplicity of $e_n$ in $E_n :=\{e_1,\dots,e_n\}$. Set

$$ \begin{equation*} E_{n,N}:=\begin{cases} \{e_n\} ^{2\nu_n-N-2}&\text{for } \nu_n\geqslant N+2, \\ E^{j_1,\dots,j_{N+2}}&\text{for }\nu_n\leqslant N+2, \end{cases} \end{equation*} \notag $$
where the indices $j_1,\dots,j_{N+2}$ satisfy $1\leqslant j_1<\dots <j_{N+2}=n$ and the point $e_{j_{N+2}}=e_n$ has multiplicity $\nu_n$ in $E^{j_1,\dots,j_{N+2}}$ (so that for $ \nu_n\leqslant N+2\leqslant n$ the set $E_{n,N}$ is obtained from $E_n$ by removing arbitrary $n-N-2$ points distinct from $e_n$). It is easy to see that for $n=N+2$ we have $E_{N+2,N}=E_{N+2}$.

Let $F(z)$ be a function defined with multiplicities at points $e_1,e_2,\dots$ in $\mathbb D$, let ${m\in\mathbb N}$, and let $j_1,\dots,j_m$ be a set of indices such that $1\leqslant j_1<\dots <j_m<\infty$. Let $F_n(z)$ be the restriction of $F(z)$ to $E_n :=\{e_1,\dots,e_n\}$, $n=1,2,\dots$, $F^{j_1,\dots,j_m}(z)$ be the restriction of $F(z)$ to $E^{j_1,\dots,j_m}:=\{e_{j_1},\dots,e_{j_m}\}$, and

$$ \begin{equation*} {F}_{n,N}(z) \text{ be the restriction of }F_n(z)\text{ to } \begin{cases} E_{n,N}\subseteq E_n\quad &\text{if }\nu_n\leqslant N+2\leqslant n, \\ \{e_n\} ^{\nu_n}\subseteq E_n \quad &\text{if } N+2\leqslant \nu_n\leqslant n; \end{cases} \end{equation*} \notag $$
here for $ N+2\leqslant \nu_n\leqslant n$ the function ${F}_{n,N}(z)$ is arbitrarily extended to a function defined at the points in $E_{n,N}=\{e_n\} ^{2\nu_n-N-2}$ with multiplicities taken into account (that is, the restriction of $F_n(z)$ to the set $\{e_n\} ^{\nu_n}$ defined by the values of derivatives $F_n^{(0)}(e_n),\dots,F_n^{(\nu_n-1)}(e_n)$ is extended by selecting arbitrary values of derivatives $F_n^{(\nu_n)}(e_n),\dots,F_n^{(2\nu_n-N-3)}(e_n)$).

For the brevity of the statements that follow set

$$ \begin{equation*} E_{N+1,N}:=E_{N+1}\quad\text{and} \quad F_{N+1,N}(z):=F_{N+1}(z). \end{equation*} \notag $$

In the set of functions defined with multiplicities at $e_1,e_2,\dots $, for $N\in\mathbb Z_+^\infty$ and $\zeta =\mathtt{C},\mathtt{S}$ we distinguish the subsets

$$ \begin{equation} \Phi ^{\zeta}_N :=\bigl\{F(z)\colon M_{E_p}^{\zeta;F_p}>0, \, p=1,\dots,N, \,M_{E_{N+p,N}}^{\zeta;F_{N+p,N}}=0, \, p=1,2,\dots \bigr\} \end{equation} \tag{20} $$
(as above, if $N=0$, then there are no inequalities of the form $M_{E_p}^{\zeta;F_p}>0$ for $p=1,\dots,N$, while for $N=\infty$ there are no equalities of the form $M_{E_{N+p,N}}^{\zeta;F_{N+p,N}}=0$ for $p=1,2,\dots $).

For functions defined with multiplicities at some points in $\mathbb D$, in § 4 we prove a multipoint analogue of Carathéodory’s and Schur’s criteria.

Theorem 3. Let $F(z)$ be a function defined at points $e_1,e_2,\dots$ in $\mathbb D$ with multiplicities taken into account. Then in the notation introduced above

$$ \begin{equation*} F(z)\lessdot \mathfrak B_N^{\zeta} \quad\Longleftrightarrow\quad F(z)\in\Phi_N^{\zeta}, \qquad N\in\mathbb Z_+^\infty, \quad \zeta =\mathtt{C},\mathtt{S}. \end{equation*} \notag $$

We make some comments on Theorem 3.

Remark 1. The necessary conditions for the relation $F(z)\lessdot \mathfrak B_N^{\zeta}$ in Theorem 3 can also be expressed in a more elaborate form:

$$ \begin{equation} F(z)\lessdot \mathfrak B_N^{\zeta}\quad \Longrightarrow\quad \begin{cases} M_{E^{j_1,\dots,j_m}}^{\zeta;F^{j_1,\dots,j_m}}>0,&m\leqslant N, \\ M_{E^{j_1,\dots,j_m}}^{\zeta;F^{j_1,\dots,j_m}}=0,&m\geqslant N+1, \\ M_{E_{n,N}}^{\zeta;{F}_{n,N}} = 0,&n\geqslant N+1, \end{cases} \quad\Longrightarrow \quad F(z)\in\Phi_N^{\zeta}. \end{equation} \tag{21} $$

Remark 2. For $\zeta ={\mathtt{S}}$ Theorem 3 and the implication (21) were proved in [10] (Theorem 1 and an addendum to it).

Remark 3. Theorem 3 is independent of the choice of the values of the derivatives $F_n^{(\nu_n)}(e_n),\dots,F_n^{(2\nu_n-N-3)}(e_n)$ of ${F}_{n,N}(z)$ for $\nu_n\geqslant N+2$, which can be arbitrary.

Remark 4. Theorem 3 can also be stated in terms of formal Newton series. In fact, let $F(z)$ be the function defined at the points $e_1,e_2,\dots$ in $\mathbb D$ with multiplicities taken into account, and let $\mathbf{F}_{k}(z)$ be the Lagrange interpolation polynomial for the restriction of $F(z)$ to $E_k:=\{e_1,\dots,e_k\}$, $k=1,2,\dots$ . It is easy to see that $\mathbf{F}_{k+1}(z)-\mathbf{F}_k(z)$ is a polynomial of degree at most $k$ which vanishes with multiplicities at the points in $E_{k}$, so that

$$ \begin{equation*} \mathbf{F}_{k+1}(z)-\mathbf{F}_{k}(z)= a_k(z-e_1)\dotsb (z-e_k) \end{equation*} \notag $$
for some $a_k\in\mathbb C$. This means that we can assign to $F(z)$ the Newton series
$$ \begin{equation*} f(z)=\mathbf{F}_{1}(z)+\sum_{k=1}^{\infty}(\mathbf{F}_{k+1}(z)-\mathbf{F}_{k}(z)) =a_0+\sum_{k=1}^{\infty} a_k(z-e_1)\dotsb (z-e_k) \end{equation*} \notag $$
such that its $(n-1)$st partial sum coincides with the Lagrange interpolation polynomial for $F(z)$ on $E_n$.

Conversely, each Newton series $f(z)$ with nodes at points $e_1,e_2,\dots$ in $\mathbb D$ produces a function $F(z)$ defined at the points $e_1,e_2,\dots$ by the equalities

$$ \begin{equation*} F^{(\nu_n -1)}(e_n)=f_n^{(\nu_n -1)}(e_n), \qquad n=1,2,\dots, \end{equation*} \notag $$
with multiplicities taken into account. Here $\nu_n$ is the multiplicity of $e_n$ in the set $E_n=\{e_1,\dots,e_n\}$ and $f_n(z)$ is the $(n-1)$st partial sum of the Newton series $f(z)$.

If a Newton series $f(z)$ with nodes at points $e_1,e_2,\dots$ in $\mathbb D$ corresponds to a function $F(z)$ defined with multiplicities at $e_1,e_2,\dots$ and if $F(e_n)\neq -1$ for $n=1,2,\dots$, then we let $(T\circ f)(z)$ denote the Newton series corresponding to the function $(T\circ F)(z)$. In the case when

$$ \begin{equation*} e_n=e, \quad n=1,2,\dots,\quad\text{and} \quad f(z)=\sum_{k=0}^{\infty} a_k(z-e)^k, \quad a_0\neq -1, \end{equation*} \notag $$
the series $(T\circ f)(z)$ coincides with the series obtained by dividing the power series $1-f(z)$ (about $e$) formally by the series $1+f(z)$.

Given a Newton series $f(z)$, the expression $f(z)\lessdot \mathfrak B_N^\zeta$ means that there exists a function $\mathscr F_N^\zeta (z)\in \mathfrak B_N^\zeta$ such that the following inclusions hold:

$$ \begin{equation*} \frac{\mathscr F_N^\zeta (z)-f_n(z)}{(z-e_1)\dotsb (z-e_n)}\in H(\mathbb D) \quad \text{for all } n=1,2,\dots, \end{equation*} \notag $$
where $f_n(z)$ is the $(n-1)$st partial sum of the Newton series. Note that
$$ \begin{equation*} f(z)\lessdot \mathfrak B_N^\zeta \quad \Longleftrightarrow \quad F(z)\lessdot \mathfrak B_N^\zeta, \qquad N\in\mathbb Z_+^\infty, \quad \zeta =\mathtt{C},\mathtt{S}, \end{equation*} \notag $$
where $F(z)$ is the function corresponding to the series $f(z)$.

To formulate Theorem 3 in terms of formal Newton series we replace the notation $F$ by $f$ throughout the statement of Theorem 3 and the definition (20) of the sets $\Phi ^{\zeta}_N$ and treat $f(z)$ as a formal Newton series with nodes $e_1,e_2,\dots$ in $\mathbb D$, $f_n(z)$, $n=1,2,\dots$, as its $(n-1)$st partial sum, $f^{j_1,\dots,j_m}(z)$ as an arbitrary fixed polynomial such that

$$ \begin{equation*} \frac{f^{j_1,\dots,j_m}(z)-f_{n}(z)}{(z-e_{j_1})\dotsb (z-e_{j_m})}\in H(E^{j_1,\dots,j_m}), \qquad 1\leqslant j_1<\dots <j_m\leqslant n, \end{equation*} \notag $$
and ${f}_{n,N}(z)$ as an (arbitrary fixed) polynomial such that
$$ \begin{equation} \frac{f_{n,N}(z)-f_n(z)}{(z-e_{n})^{\nu_n}}\in H(e_{n}) \quad \text{for } n\geqslant \nu_n\geqslant N+2 \end{equation} \tag{22} $$
and
$$ \begin{equation*} \frac{f_{n,N}(z)-f_n(z)}{\prod_{e\in E_{n,N}}(z-e)}\in H(E_{n,N}) \quad \text{for } \nu_n\leqslant N+2\leqslant n \end{equation*} \notag $$
($\nu_n$ and $E_{n,N}$ were defined before Theorem 3).

Remark 5. The definitions of the sets $E_{n,N}$ for $N+2\leqslant n$ are considerably different in the cases when $e_1,e_2,\dots$ are pairwise distinct and when $e_1,e_2,\dots$ coincide with the same point $e\in\mathbb D$. In the first case, for $n\geqslant N+2$ we have $E_{n,N}=E^{j_1,\dots,j_{N+2}}$, where $1\leqslant j_1<\dots <j_{N+1}<n$ are arbitrary fixed indices and $j_{N+2}=n$. In particular, for $j_k=k$, $k=1,\dots,N+1$, we obtain

$$ \begin{equation*} E_{n,N}=E^{1,\dots,N+1,n}=\{e_1,\dots,e_{N+1}, e_n\}, \qquad n\geqslant N+2. \end{equation*} \notag $$
Hence the definition (20) has the following form:
$$ \begin{equation} \begin{aligned} \, \notag \Phi ^{\zeta}_N &:=\bigl\{F(z)\colon M_{E_p}^{\zeta;F_p}>0,\, p=1,\dots,N, \\ &\qquad M_{E_{N+1}}^{\zeta;{F}_{N+1}}= M_{\{e_1,\dots,e_{N+1}, e_{N+p}\}}^{\zeta;F^{1,\dots,N+1, N+p}} =0,\, p = 2,3,\dots \bigr\}, \end{aligned} \end{equation} \tag{23} $$
where $F^{1,\dots,N+1, n}(z)$ for $n\geqslant N+2$ is the restriction of $F_n(z)$ to the set ${E^{1,\dots,N+1,n} \subseteq E_n}$.

In the second case

$$ \begin{equation*} E_n=\{e\} ^n, \quad \nu_n=n, \qquad E_{n,N}=\{e\} ^{2n-N-2}, \end{equation*} \notag $$
and it is natural to take either $f_{2n-N-2}(z)$ or $f_{n}(z)$ as a polynomial $f_{n,N}(z)$ (${n\geqslant N+2}$) satisfying (22). In particular, for $e=0$, bearing in mind that ${M_{E_n}^{\zeta;f_n}=M_n^{\zeta;f}}$ by Proposition 2 we see that for $N\in\mathbb Z_+$ the sets $\Phi_N^\zeta$ (see (20) for $F$ replaced by $f$) either coincide with the $\hat{\mathscr{A}}_N^\zeta$ (see (9)) or with the $\breve{{\mathscr A}}_N^\zeta$ (see (10)). This means that Theorem 3, as reformulated in terms of the formal power series $f(z)=\sum_{k=0}^\infty a_kz^k$, yields the refinement of Carathéodory’s and Schur’s criteria for $N\in\mathbb Z_+$ that we stated in § 1.

Remark 6. It was shown in [10] that for a function $F(z)$ defined by the equalities $F(e_n)=\gamma$ for $n\in\mathbb N\setminus\{k\}$ at pairwise distinct points $e_1,e_2,\dots$ in $\mathbb D$, where $|\gamma |=1$ and $k\geqslant 3$, we have $M_{E_{n}}^{\mathtt{S};F_n}=0$ for all $n=1,2,\dots $ and any value of $F(e_k)$ (we also obtain a similar example for $\zeta =\mathtt{C}$, provided that we replace the condition $|\gamma |=1$ by $\operatorname{Re} \gamma =0$). This means that in contrast to the conditions in (23) for $N=0$, the conditions

$$ \begin{equation*} M_{E_p}^{\zeta;{F}_{p}} = 0, \qquad p=1,2,\dots, \end{equation*} \notag $$
are not sufficient to claim that $F(z)\lessdot \mathfrak B_0^{\mathtt{S}}$ in the case when $F(e_k)\neq\gamma$.

§ 4. Proof of the multipoint Carathéodory-Schur criterion

Theorem 3 supplemented by implication (21) covers two different cases, ${\zeta =\mathtt{C}}$ and ${\zeta =\mathtt{S}}$. Throughout this section, for brevity we call the case $\zeta = \mathtt{C} $ of Theorem 3, which corresponds to the multipoint version of Carathéodory’s theorem, Theorem 3$^{\mathtt{C}}$, and we call the case $\zeta =\mathtt{S}$, corresponding to the multipoint version of Schur’s theorem, Theorem 3$^{\mathtt{S}}$. In view of Remark 2 after the statement of Theorem 3, to establish Theorem 3 it is sufficient to prove Theorem 3$^{\mathtt{C}}$ alone. We obtain the proof of Theorem 3$^{\mathtt{C}}$ by reducing it to Theorem 3$^{\mathtt{S}}$ (established in [10]) with the help of Theorem 2, proved in § 2.

Turning to the proof of Theorem 3$^{\mathtt{C}}$ with implication (21) note that we have the chain of implications

$$ \begin{equation} \begin{aligned} \, \notag F(z)\lessdot \mathfrak B_N^{\mathtt{C}} &\quad\Longrightarrow\quad (T\circ F)(z)\lessdot \mathfrak B_N^{\mathtt{S}}, \qquad (T\circ F)(e_n)\neq -1, \quad n=1,2,\dots, \\ \notag &\quad\Longrightarrow \quad \begin{cases} M_{E^{j_1,\dots,j_m}}^{\mathtt{S};T\circ F^{j_1,\dots,j_m}}>0,& m\leqslant N, \\ M_{E^{j_1,\dots,j_m}}^{\mathtt{S};T\circ F^{j_1,\dots,j_m}}=0,& m\geqslant N+1, \\ M_{E_{n,N}}^{\mathtt{S};T\circ F_{n,N}}=0,& n\geqslant N+1, \end{cases} \\ & \quad\Longrightarrow \quad \begin{cases} M_{E^{j_1,\dots,j_m}}^{\mathtt{C};F^{j_1,\dots,j_m}}>0,&m\leqslant N, \\ M_{E^{j_1,\dots,j_m}}^{\mathtt{C};F^{j_1,\dots,j_m}}=0,&m\geqslant N+1, \\ M_{E_{n,N}}^{\mathtt{C};F_{n,N}}=0,& n\geqslant N+1, \end{cases} \quad\Longrightarrow \quad F(z)\in\Phi_N^{\mathtt{C}}, \end{aligned} \end{equation} \tag{24} $$
the first of which follows from Proposition 4, the second from Theorem 3$^{\mathtt{S}}$, proved in [10], as applied to the function $(T\circ F)(z)$, the third follows from Theorem 2 and the inequalities $(T\circ F)(e_n)\neq -1$, $n=1,2,\dots$, and the fourth from the definition (20) of the sets $\Phi_N^{\mathtt{C}}$.

Thus, to complete the proof of Theorem 3$^{\mathtt{C}}$ it suffices to show that

$$ \begin{equation} F(z)\in\Phi_N^{\mathtt{C}} \quad\Longrightarrow\quad F(z)\lessdot \mathfrak B_N^{\mathtt{C}}. \end{equation} \tag{25} $$

We prove (25) on the basis of Lemmas 1$^{\mathtt{C}}$ and 2$^{\mathtt{C}}$, which are of independent interest. They are analogues of Lemmas 1$^{\mathtt{S}}$ and 2$^{\mathtt{S}}$, which we proved in [10] (Lemmas 3–5). We combine their assertions into Lemmas 1 and 2, by adding some notation. Namely, for $E_n=\{e_1,\dots,e_n\}$ we set

$$ \begin{equation*} \omega_{E_{n}}(z):=(z-e_1)\dotsb (z-e_n). \end{equation*} \notag $$

Lemma 1. Let $F(z)$ be a polynomial, let $n\in\mathbb N$, $E_{n}=\{e_1,\dots,e_{n}\}\subset\mathbb D$, $\zeta =\mathtt{C},\mathtt{S}$, and let

$$ \begin{equation} M_{E_1}^{\zeta;F}>0,\quad \dots,\quad M_{E_{n}}^{\zeta;F}>0. \end{equation} \tag{26} $$

Then there exists a function $\mathscr F_{n}^\zeta (z)\in \mathfrak B ^\zeta \setminus (\bigcup_{k=0}^{n-1} \mathfrak B_k^\zeta )$ such that

$$ \begin{equation} \frac{\mathscr F_{n}^\zeta (z)-F(z)}{\omega_{E_n}(z)}\in H({E_{n}}). \end{equation} \tag{27} $$
In particular, $F(z)\neq -1$, $z\in E_n$.

Lemma 2. Let $F(z)$ be a polynomial, let $N\in \mathbb Z_+$, $p \in \mathbb N$, $E_{N+p} = \{e_1,\dots,e_{N+p}\} \subset \mathbb D$, $\zeta =\mathtt{C},\mathtt{S}$, and let

$$ \begin{equation} M_{E_1}^{\zeta;F}>0,\quad \dots,\quad M_{E_{N}}^{\zeta;F}>0\quad\textit{and} \quad M_{E_{N+1,N}}^{\zeta;F_{N+1,N}}=\dots = M_{E_{N+p,N}}^{\zeta;F_{N+p,N}}=0, \end{equation} \tag{28} $$
where the sets $E_{N+j,N}(z)$ and polynomials $F_{N+j,N}(z)$ were defined for $j=1,\dots,p$ before Theorem 3.

Then there exists a function $\mathscr F_{N}^\zeta (z)\in \mathfrak B_N^\zeta $ such that

$$ \begin{equation} \frac{\mathscr F_{N}^\zeta (z)-F(z)}{\omega_{E_{N+p}}(z)}\in H({E_{N+p}}). \end{equation} \tag{29} $$
In particular, $F(z)\neq -1$, $z\in E_{N+p}$ (except when $N=0$, $\zeta ={\mathtt{S}}$ and $\mathscr F_{N}^{\mathtt{S}} (z)\equiv -1$).

Proof of Lemma 1. Lemma 1$^{\mathtt{S}}$ was proved in [10] (Lemma 3). Now note that under the additional assumption that $F(z)\neq -1$ for $z\in E_{n}$, Lemma 1$^{\mathtt{C}}$ reduces to Lemma 1$^{\mathtt{S}}$ by use of Theorem 2. In fact, if conditions (26) are satisfied for $\zeta =\mathtt{C}$ and a function $F(z)\neq -1$ for $z\in E_{n}$, then by Theorem 2 the same conditions for $\zeta =\mathtt{S}$ hold for the function $(T\circ F)(z)$. Hence by Lemma 1$^{\mathtt{S}}$ there exists a function $\mathscr F_{n}^{\mathtt{S}} (z)\in \mathfrak B ^{\mathtt{S}} \setminus (\bigcup_{k=0}^{n-1} \mathfrak B_k^{\mathtt{S}} )$ such that
$$ \begin{equation*} \frac{\mathscr F_{n}^{\mathtt{S}} (z)-(T\circ F)(z)}{\omega_{E_n}(z)}\in H({E_{n}}). \end{equation*} \notag $$
By Proposition 1 this yields
$$ \begin{equation*} \frac{(T\circ \mathscr F_{n}^{\mathtt{S}}) (z)-F(z)}{\omega_{E_n}(z)}\in H({E_{n}}). \end{equation*} \notag $$
Therefore, setting $\mathscr F_{n}^{\mathtt{C}} (z)=(T\circ \mathscr F_{n}^{\mathtt{S}}) (z)$ and observing that $\mathscr F_{n}^{\mathtt{C}} (z)\in \mathfrak B ^{\mathtt{C}} \setminus (\bigcup_{k=0}^{n-1} \mathfrak B_k^{\mathtt{C}} )$ by Proposition 4 we obtain the required inclusion (27) for $\zeta =\mathtt{C}$.

Thus, to complete the proof of Lemma 1 it suffices to show that inequalities (26) for $\zeta =\mathtt{C}$ imply the inequalities $F(z)\neq -1$ for $z\in E_{n}$ in the hypotheses of Theorem 2.

Assume the converse: let $F(z)= -1$ for some $z\in E_{n}$. Taking the relation $F(e_1)\neq -1$, which follows from the first inequality in (26) (namely, $0<M_{E_1}^{\mathtt{C};F}=2\operatorname{Re} F(e_1)$), into account, this assumption means that $n\geqslant 2$ and there exists ${k\in\{2,\dots,n\}}$ such that

$$ \begin{equation} F(z)\neq -1, \quad z\in E_{k-1}:=\{e_1,\dots,e_{k-1}\}, \qquad F(e_{k})=-1. \end{equation} \tag{30} $$
In particular, it follows from (30) that $e_k$ is distinct from all points in $E_{k-1}$ (we write $e_{k}\notin E_{k-1}$ for brevity).

Consider the polynomial $G_{k,\varepsilon}(z)$ that is equal to $F(z)$ for $z\in E_{k-1}$ and to $F(e_k)+\varepsilon$ for $z=e_{k}$. We see directly from this definition and (30) that

$$ \begin{equation*} G_{k,\varepsilon}\neq -1 \quad \text{for } z\in E_{k}, \quad \varepsilon\neq 0. \end{equation*} \notag $$
Since by the definitions (11)(14) we have
$$ \begin{equation*} M_{E_{1}}^{\mathtt{C};G_{k,\varepsilon}}=M_{E_{1}}^{\mathtt{C};F}, \quad \dots,\quad M_{E_{k-1}}^{\mathtt{C};G_{k,\varepsilon}}=M_{E_{k-1}}^{\mathtt{C};F}\quad\text{and} \quad \lim_{\varepsilon\to 0}M_{E_{k}}^{\mathtt{C};G_{k,\varepsilon}}=M_{E_{k}}^{\mathtt{C};F}, \end{equation*} \notag $$
it follows from (26) for $\zeta =\mathtt{C}$ that for all sufficiently small $\varepsilon$ we have $M_{E_{j}}^{\mathtt{C};G_{k,\varepsilon}}>0$, $j=1,\dots,k$. Thus, $G_{k,\varepsilon}(z)$ for $\varepsilon\neq 0$ fulfils all assumptions under which Lemma 1$^{\mathtt{C}}$ (for $k$ in place of $n$) was proved. Hence for all sufficiently small $\varepsilon\neq 0$ there exist functions $\mathscr F_{k,\varepsilon}^{\mathtt{C}}(z)\in \mathfrak B ^{\mathtt{C}}$ such that
$$ \begin{equation} \frac{\mathscr F_{k,\varepsilon}^{\mathtt{C}}(z)-G_{k,\varepsilon}(z)}{\omega_{E_{k}}(z)}\in H(E_{k}). \end{equation} \tag{31} $$
It follows from (31) that
$$ \begin{equation*} \lim_{\varepsilon \to 0}\mathscr F_{k,\varepsilon}(e_{k})=\lim_{\varepsilon \to 0}G_{k,\varepsilon}(e_{k})=\lim_{\varepsilon \to 0}(F(e_{k})+\varepsilon )=F(e_{k})=-1, \end{equation*} \notag $$
which is impossible because $\mathscr F_{k,\varepsilon}^{\mathtt{C}}\in\mathfrak B ^{\mathtt{C}}$, and therefore $\operatorname{Re} \mathscr F_{k,\varepsilon}^{\mathtt{C}} (e_{k})\geqslant 0$. This means that the assumption $F(e_{k})=-1$ leads to a contradiction. Lemma 1$^{\mathtt{C}}$, and therefore also Lemma 1, are proved.

Proof of Lemma 2. Note that Lemma 2$^{\mathtt{S}}$ was proved in [10], explicitly for $p=1,2$ (see [10], Lemmas 3 and 4) and implicitly for $p=3,4,\dots$ (see Lemma 5 and the subsequent proof of Theorem 1 in [10]). Since Lemma 2$^{\mathtt{C}}$, under the additional assumption that $F(z)\neq -1$ for $z\in E_{N+p}$, reduces to Lemma 2$^{\mathtt{S}}$ by use of Theorem 2, to prove Lemma 2 it suffices to show that the hypotheses of Lemma 2$^{\mathtt{C}}$ imply the relations
$$ \begin{equation} F(z)\neq -1 \quad\text{for } z\in E_{N+p}. \end{equation} \tag{32} $$

For $N=0$ and $p=1$ the inequality $F(e_1)\neq -1$ clearly follows from the equality $0=M_{E_1}^{\mathtt{C};F}=2\operatorname{Re} F(e_1)$ in (28).

We prove (32) for $N\in\mathbb N$ and $p=1$. Then we must slightly modify the arguments used in the proof of Lemma 1 and rely on the valid equality $M_{E_{N+1}}^{\mathtt{C};F}=0$ in place of the strict inequality $M_{E_{N+1}}^{\mathtt{C};F}>0$ used in Lemma 1. Fix a function $\mathscr F ^{\mathtt{C}}(z)\in \mathfrak B ^{\mathtt{C}}$ such that

$$ \begin{equation} \frac{\mathscr F ^{\mathtt{C}} (z)-F(z)}{\omega_{E_N}(z)}\in H({E_{N}}). \end{equation} \tag{33} $$
It exists by Lemma 1 for $\zeta =\mathtt{C}$ (and $N$ in place of $n$), which we have already established. In particular, this inclusion yields the inequalities $F(z)\neq -1$, $z\in E_{N}$. Thus, to prove (32) for $p=1$ it suffices to show that conditions (28) (for $\zeta =\mathtt{C}$, $N\in\mathbb N$ and $p=1$) yield the inequality $F(e_{N+1})\neq -1$.

Assuming the converse we obtain

$$ \begin{equation*} e_{N+1}\notin E_N\quad\text{and} \quad G_{N+1,\varepsilon}\neq -1 \quad \text{for } z\in E_{N+1}\quad\text{and} \quad \varepsilon\neq 0, \end{equation*} \notag $$
where $G_{N+1,\varepsilon}(z)$ is the polynomial equal to $F(z)$ for $z\in E_N$ and to $F(e_{N+1})+\varepsilon$ for $z=e_{N+1}$.

We claim that there exists a null sequence of nonzero complex numbers $\{\varepsilon_l\}_{l=1}^\infty$ such that

$$ \begin{equation*} M_{E_{N+1}}^{\mathtt{C};G_{N+1,\varepsilon_l}}\geqslant 0 \quad \text{for } l=1,2,\dots\,. \end{equation*} \notag $$
Since $\mathscr F ^{\mathtt{C}}(z)\in \mathfrak B ^{\mathtt{C}}$, we have $M_{E_{N+1}}^{\mathtt{C};\mathscr F ^{\mathtt{C}}}\geqslant 0$ by (24). Hence from the definition of $G_{N+1,\varepsilon} (z)$ and(33) we obtain
$$ \begin{equation} M_{E_{N+1}}^{\mathtt{C};G_{N+1,\varepsilon_0}}=M_{E_{N+1}}^{\mathtt{C};\mathscr F ^{\mathtt{C}}}\geqslant 0, \quad \text{where } \varepsilon_0:=\mathscr F ^{\mathtt{C}}(e_{N+1})-F(e_{N+1}). \end{equation} \tag{34} $$
Note that $\varepsilon_0\neq 0$, because $\mathscr F ^{\mathtt{C}}(e_{N+1})\neq -1$ and $F(e_{N+1})=-1$ by assumption.

Since $e_{N+1}\notin E_N$, it follows from the definitions of $G_{N+1,\varepsilon} (z)$ and the quantities $M_{E_{N+1}}^{\mathtt{C};G_{N+1,\varepsilon}}$ (see (11)(14)) that

$$ \begin{equation} M_{E_{N+1}}^{\mathtt{C};G_{N+1,\varepsilon}}=A|\varepsilon |^2+B\varepsilon +\overline{B\varepsilon}+C, \end{equation} \tag{35} $$
where $A$, $B$ and $C $ are some coefficients depending on $N$ and $F$ and, furthermore,
$$ \begin{equation*} A \in\mathbb R\quad\text{and} \quad C =M_{E_{N+1}}^{\mathtt{C};G_{N+1,0}}=M_{E_{N+1}}^{\mathtt{C};F}=0. \end{equation*} \notag $$
If $B=0$, then $M_{E_{N+1}}^{\mathtt{C};G_{N+1,\varepsilon}}\geqslant 0$ for all $\varepsilon\in\mathbb C$ because
$$ \begin{equation*} M_{E_{N+1}}^{\mathtt{C};G_{N+1,\varepsilon}}=A|\varepsilon|^2 = \frac{M_{E_{N+1}}^{\mathtt{C};G_{N+1,\varepsilon_0}}|\varepsilon|^2}{|\varepsilon_0|^2}\geqslant 0 \end{equation*} \notag $$
by (35) and (34). If $B\neq 0$, then the equalities
$$ \begin{equation*} M_{E_{N+1}}^{\mathtt{C};G_{N+1,\varepsilon_l}}=A|\varepsilon_l|^2 +2\operatorname{Re} (B\varepsilon_l) = 0 \end{equation*} \notag $$
hold for each null sequence of nonzero complex numbers $\varepsilon_l=|\varepsilon_l|e^{i\theta_l}$, where
$$ \begin{equation*} 0<|\varepsilon_l|<2\biggl|\frac BA\biggr|\quad\text{and} \quad \cos (\arg B+\theta_l)=-\frac{A|\varepsilon_l|}{2|B|}. \end{equation*} \notag $$

Thus, the polynomials $G_{N+1,\varepsilon_l}$, $l=1,2,\dots$, fulfill the assumptions of either Lemma 1$^{\mathtt{C}}$ for $n=N+1$ or of Lemma 2$^{\mathtt{C}}$, which we proved above. Hence for all $l=1,2,\dots$ there exist functions $\mathscr F ^{\mathtt{C},l}(z)\in \mathfrak B ^{\mathtt{C}}$ such that

$$ \begin{equation} \frac{\mathscr F ^{\mathtt{C},l}(z)-G_{N+1,\varepsilon_l}(z)}{\omega_{E_{N+1}}(z)}\in H(E_{N+1}), \qquad l=1,2,\dotsc\,. \end{equation} \tag{36} $$
It follows from (36) that
$$ \begin{equation*} \begin{aligned} \, \lim_{l\to \infty}\mathscr F ^{\mathtt{C},l}(e_{N+1}) &=\lim_{l\to \infty}G_{N+1,\varepsilon_l}(e_{N+1}) \\ &=\lim_{l\to \infty}(F(e_{N+1})+\varepsilon_l )=F(e_{N+1})=-1, \end{aligned} \end{equation*} \notag $$
which is impossible because $\mathscr F ^{\mathtt{C},l}\in\mathfrak B ^{\mathtt{C}}$, $l=1,2,\dots$ . This means that the assumption $F(e_{N+1})=-1$ leads to a contradiction. Lemma 2$^{\mathtt{C}}$ is proved for $N\in \mathbb Z_+$ and $p=1$.

Let $p\geqslant 2$. We make the inductive assumption that (32) has been proved for all $N\in\mathbb Z_+$ and all positive indices up to $p-1$ inclusive. Now we prove these inequalities for the index $p$. By our inductive assumption there exists a function $\mathscr F ^{\mathtt{C}}(z)\in \mathfrak B ^{\mathtt{C}}$ such that (29) holds (for $\zeta =\mathtt{C}$ and $p$ replaced by $p-1$). In particular, this yields the inequalities $F(z)\neq -1$ for $z\in E_{N+p-1}$. Thus, to prove (32) is its sufficient to show that conditions (28) for $\zeta =\mathtt{C}$ and $N\in\mathbb N$ imply that $F(e_{N+p})\neq -1$. Assuming the converse we arrive at a contradiction by means of the arguments analogous to the ones used for $p=1$ (except that $N$ must now be replaced by $N+p-1$). This completes the proof of Lemma 2$^{\mathtt{C}}$, and therefore of Lemma 2.

To complete the proof of Theorem 3 we note that, given Lemmas 1 and 2 and Theorem 2, implication (25) follows from the chain of implications

$$ \begin{equation*} F(z)\in\Phi_N^{\mathtt{C}} \quad\Longrightarrow\quad (T\circ F)(z)\in\Phi_N^{\mathtt{S}} \quad\Longrightarrow\quad (T\circ F)(z)\lessdot \mathfrak B_N^{\mathtt{S}} \quad\Longrightarrow\quad F(z)\lessdot \mathfrak B_N^{\mathtt{C}}, \end{equation*} \notag $$
the first of which follows from the definition (20) of the sets $\Phi_N^{\zeta}$ for $\zeta = \mathtt{C},\mathtt{S}$ and Theorem 2, taking the inequalities
$$ \begin{equation*} F(z)\neq -1 \quad\text{for } z\in E_n, \quad n=1,2,\dots, \end{equation*} \notag $$
which follow from Lemmas 1 and 2, into account; the second implication follows from Theorem 3 for $\zeta =\mathtt{S}$, which was proved in [10], and the third follows from Proposition 4. The proof of (25) and therefore of Theorem 3 is complete.

In conclusion we note that the arguments in this section show that Theorem 3$^{\mathtt{C}}$ is a consequence of Theorems 2 and 3$^{\mathtt{S}}$. It is easy to see that, in a similar way, Theorem 3$^{\mathtt{S}}$ is a consequence of Theorems 2 and 3$^{\mathtt{C}}$. In other words, Theorem 2 reveals the equivalence of the multipoint analogues of Carathéodory’s and Schur’s criteria, just as Theorem 1 reveals the equivalence of the classical Carathéodory and Schur criteria, with the single reservation that the verification of the inequalities $F(e_n)\neq -1$, $n=1,2,\dots$, in the hypotheses of Theorem 2 is not as trivial as in the classical case.


Bibliography

1. J. Schur, “Über Potenzreihen, die im Innern des Einheitskreises beschränkt sind. I”, J. Reine Angew. Math., 1917:147 (1917), 205–232  crossref  mathscinet  zmath; II, 1918:148 (1918), 122–145  crossref  mathscinet  zmath
2. C. Carathéodory, “Über den Variabilitätsbereich der Koeffizienten von Potenzreihen, die gegebene Werte nicht annehmen”, Math. Ann., 64:1 (1907), 95–115  crossref  mathscinet  zmath
3. O. Toeplitz, “Über die Fourier'sche Entwickelung pozitiver Funktionen”, Rend. Circ. Mat. Palermo, 32 (1911), 191–192  crossref  zmath
4. V. I. Buslaev, “Schur's criterion for formal power series”, Mat. Sb., 210:11 (2019), 58–75  mathnet  crossref  mathscinet  zmath; English transl. in Sb. Math., 210:11 (2019), 1563–1580  crossref  adsnasa
5. G. Pólya, “Beitrag zur Verallgemeinerung des Verzerrungssatzes auf mehrfach zusammenhängende Gebiete. III”, Sitzungsber. Preuss. Akad. Wiss. Phys.-Math. Kl., 1929 (1929), 55–62  zmath
6. V. I. Buslaev, “An analogue of Polya's theorem for piecewise holomorphic functions”, Mat. Sb., 206:12 (2015), 55–69  mathnet  crossref  mathscinet  zmath; English transl. in Sb. Math., 206:12 (2015), 1707–1721  crossref  adsnasa
7. V. I. Buslaev, “Convergence of a limit periodic Schur continued fraction”, Mat. Zametki, 107:5 (2020), 643–656  mathnet  crossref  mathscinet  zmath; English transl. in Math. Notes, 107:5 (2020), 701–712  crossref
8. L. Baratchart, S. Kupin, V. Lunot and M. Olivi, “Multipoint Schur algorithm and orthogonal rational functions, I: Convergence properties”, J. Anal. Math., 114 (2011), 207–253  crossref  mathscinet  zmath
9. V. I. Buslaev, “Schur's criterion for formal Newton series”, Mat. Zametki, 108:6 (2020), 920–924  mathnet  crossref  mathscinet  zmath; English transl. in Math. Notes, 108:6 (2020), 884–888  crossref
10. V. I. Buslaev, “Necessary and sufficient conditions for extending a function to a Schur function”, Mat. Sb., 211:12 (2020), 3–48  mathnet  crossref  mathscinet  zmath; English transl. in Sb. Math., 211:12 (2020), 1660–1703  crossref  adsnasa

Citation: V. I. Buslaev, “Necessary and sufficient conditions for extending a function to a Carathéodory function”, Mat. Sb., 213:11 (2022), 5–24; Sb. Math., 213:11 (2022), 1488–1506
Citation in format AMSBIB
\Bibitem{Bus22}
\by V.~I.~Buslaev
\paper Necessary and sufficient conditions for extending a~function to a~Carath\'eodory function
\jour Mat. Sb.
\yr 2022
\vol 213
\issue 11
\pages 5--24
\mathnet{http://mi.mathnet.ru/sm9611}
\crossref{https://doi.org/10.4213/sm9611}
\mathscinet{http://mathscinet.ams.org/mathscinet-getitem?mr=4582602}
\adsnasa{https://adsabs.harvard.edu/cgi-bin/bib_query?2022SbMat.213.1488B}
\transl
\jour Sb. Math.
\yr 2022
\vol 213
\issue 11
\pages 1488--1506
\crossref{https://doi.org/10.4213/sm9611e}
\isi{https://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=Publons&SrcAuth=Publons_CEL&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=000992276000001}
\scopus{https://www.scopus.com/record/display.url?origin=inward&eid=2-s2.0-85165887441}
Linking options:
  • https://www.mathnet.ru/eng/sm9611
  • https://doi.org/10.4213/sm9611e
  • https://www.mathnet.ru/eng/sm/v213/i11/p5
  • This publication is cited in the following 2 articles:
    Citing articles in Google Scholar: Russian citations, English citations
    Related articles in Google Scholar: Russian articles, English articles
    Математический сборник Sbornik: Mathematics
    Statistics & downloads:
    Abstract page:301
    Russian version PDF:31
    English version PDF:64
    Russian version HTML:132
    English version HTML:74
    References:46
    First page:9
     
      Contact us:
     Terms of Use  Registration to the website  Logotypes © Steklov Mathematical Institute RAS, 2024