Abstract:
Under the assumption that Stahl's S-compact set exists we give a short proof of the limiting distribution of the zeros of Padé polynomials and the convergence in capacity of diagonal Padé approximants for a generic class of algebraic functions. The proof is direct, rather than by contradiction as Stahl's original proof was. The ‘generic class’ means, in particular, that all the ramification points of the multisheeted Riemann surface of the algebraic function in question are of the second order (that is, all branch points of the function are of square root type). As a consequence, a conjecture of Gonchar relating to Padé approximations is proved for this class of algebraic functions. We do not use the relations of orthogonality for Padé polynomials. The proof is based on the maximum principle only.
Bibliography: 19 titles.
Keywords:
Padé approximant, convergence in capacity, Stahl's theorem, Riemann surface.
Stahl’s seminal theory on the convergence of the diagonal Padé approximants for multivalued analytic functions (see [17], [18], [1] and the bibliography there) consists of two parts, the geometric part and the analytic one. In the first, geometric, part it is proved that, given a multivalued function, there exists a unique admissible compact set S possessing the S-property or, briefly, an S-compact set. In Stahl’s general theory one assumes that the set of singularities of the multivalued analytic function has zero (logarithmic) capacity. Here we restrict our investigations to the case when there is a finite number of singular points. Note that a short proof of the existence of an S-compact set in this case was proposed by Rakhmanov (1994) in the unpublished paper [11] (also see [8]). Rakhmanov’s proof is based on the connections between the capacity and the potential of the equilibrium measure of a compact set.
In the second, analytic, part of Stahl’s theory it is proved that the Padé polynomials have a limiting distribution of zeros and the diagonal Padé approximants converge in capacity. Stahl’s original proof of the existence of a limiting distribution of the Padé polynomials is based on the properties of S-compact sets and goes by contradiction. In this paper we give a short and direct proof of the second part of Stahl’s theory for a generic class of algebraic functions satisfying certain conditions, under the assumption that the first geometric component of the theory is valid. We do not use the relations of orthogonality for Padé polynomials. The proof is based only on the maximum principle for subharmonic functions (cf. [10] and [15]).
Let introduce the class F of admissible multivalued functions under consideration in our paper.
For a multivalued function f we agree that f belongs to F if the following conditions are satisfied:
(I) f is an algebraic function all of whose branch points have the second order (that is, we assume that all branch points are of square root type);
(II) there exists a germ f∞∈H(∞) of f such that the S-compact set for it consists of a finite number of disjoint closed analytic arcs each of which contains exactly two branch points of f, the endpoints of the arc.
If conditions (I) and (II) are satisfied, then we write (f,f∞)∈F.
Some discussions relating to assumptions (I) and (II) are presented in § 3 (also see [12], [9], [10] and [6]). Note the following: condition (II) means that S contains no Chebotarev points (cf. [16] and [2]).
1.2
Let f∞∈H(∞), and let S=S(f∞) be Stahl’s S-compact set corresponding to the germ f∞. Let λS be the probability equilibrium measure with support on S and VλS(z) be the corresponding logarithmic potential of λS:
VλS(z)=γS−gS(z,∞),z∈D:=ˆC∖S,
where gS(z,∞) is the Green’s function for the Stahl domain D=ˆC∖S with logarithmic singularity at the point at infinity z=∞, and γS is the Robin constant for D (relative to the point z=∞).
Let Pn:=Cn[z] be the set of all algebraic polynomials with complex coefficients of degree ⩽n. For an arbitrary polynomial Q∈C[z], Q≢0, let χ(Q) denote the zero counting measure of Q: χ(Q):=∑ζ:Q(ζ)=0δζ, where each zero of Q is counted with multiplicity.
Given a germ f∞∈H(∞) and an arbitrary n, the Padé polynomials Pn,Qn∈Pn, Qn≢0, are (not uniquely) defined by the following relation:
Rn(z):=(Qnf∞−Pn)(z)=O(1zn+1),z→∞;
Rn(z) is the so-called error function, and [n/n]f∞:=Pn/Qn is the nth diagonal Padé approximant tof∞.
The following result holds (see [17], [18] and also [1]).
Theorem. Let (f,f∞)∈F. Then, as n→∞,
1nχ(Qn),1nχ(Pn)∗→λS
and
|f(z)−[n/n]f∞(z)|1/ncap→e−2gS(z,∞)in the interior of D.
Convergence ∗→ is understood as weak-∗ convergence in the space of measures. Convergence cap→ in the interior of a domain D means convergence in capacity on compact subsets of D (see § 2.5).
Note that in the case when (f,f∞)∈F, the existence of an S-compact set for a germ f∞ follows already from the paper [11], treating a more special case than Stahl’s theory does.
The proof of the above theorem has the following consequence (see inequality (18)).
Corollary. If (f,f∞)∈F then the sequences {n−degQn} and {n−degPn} are bounded.
This result means that one of Gonchar’s conjectures relating to Padé approximations (see [1], Ch. 1, § 6, Conjecture 7) holds in the class F.
Acknowledgement
The author is sincerely grateful to the referee, whose comments contributed to correcting some gaps in proofs and improving the presentation of results.
§ 2. Proof of the theorem
2.1
Given two positive sequences {αn} and {βn}, the relation αn≍βn means that 0<C1⩽αn/βn⩽C2<∞ for n=1,2,… and some constants C1 and C2 independent of n. Given two sequences {αn(z)} and {βn(z)} of holomorphic functions in a domain Ω, the relation αn≍βn means that for each compact set K⊂Ω and all n=1,2,… the inequality 0<C1⩽|αn(z)/βn(z)|⩽C2<∞ holds for z∈K, where the constants C1 and C2 depend on K, but not on n. For such pairs of sequences (of numbers or functions) we obviously have |αn/βn|1/n→1 as n→∞.
Since (f,f∞)∈F, we have S=S(f)=⨆pj=1Sj, where Sj=arc(a2j−1,a2j). Set w2=∏pj=1(z−a2j−1)(z−a2j). Then the two-sheeted hyperelliptic Riemann surface R2(w) of the function w is the Riemann surface associated with f∞∈H(∞) and f∈F in accordance with Stahl’s theory. A point z on R2(w) is given by z=(z,w). The Riemann surface R2(w) can be considered as a two-sheeted covering of the Riemann sphere ˆC with the canonical projection π:R2(w)→ˆC defined by π(z)=z. Let \boldsymbol\Gamma:=\pi^{-1}(S). Then \boldsymbol\Gamma partitions \mathfrak R_2(w) into two domains. We call them the (open) sheets of \mathfrak R_2(w).
The function w is single valued on this Riemann surface and takes opposite values on the sheets. We denote by \mathfrak R_2^{(0)} the sheet of \mathfrak R_2(w) where w(z)/z^{p}\to1 as z\to\infty and refer to it as the zeroth sheet. The other sheet is denoted by \mathfrak R^{(1)}_2 and referred to as the first sheet of \mathfrak R_2(w). Thus, \mathfrak R_2(w)=\mathfrak R_2^{(0)}\sqcup\boldsymbol\Gamma\sqcup\mathfrak R_2^{(1)}. Points on the two sheets of \mathfrak R_2(w) are denoted by z^{(j)}, j=0,1. Clearly, \pi(\mathfrak R_2^{(j)})=D.
Following tradition, we identify \mathfrak R_2^{(0)} with the Stahl domain D=\widehat{\mathbb C}\setminus{S} and \infty^{(0)} with \infty, and we consider f_\infty as a germ f_{\infty^{(0)}} defined on \mathfrak R_2(w). In general, \mathfrak R_2(w) is not the Riemann surface of f, so that f is not single valued on \mathfrak R_2(w). But since (f,f_\infty)\in\mathscr F, the germ f_{\infty^{(0)}} can be extended as a single-valued function from the point \infty^{(0)} at infinity to the whole of the zero sheet \mathfrak R^{(0)}_2 and even farther, to a neighbourhood V^{(0,1)} of the compact set \boldsymbol\Gamma such that V^{(0,1)}\cap\mathfrak R_2^{(1)}\neq\varnothing (in what follows we assume that this neighbourhood is sufficiently small). Since \pi(\boldsymbol\Gamma)=S is Stahl’s compact set, the Green’s function g_S(z,\infty) of D can be lifted to \mathfrak R_2(w) as a function g(\mathbf z) of \mathbf z with the following properties:
Moreover, R>1 is such that the number of connected components of the set \{z^{(1)}\colon g_S(z,\infty)= \log{R}\} is equal to that of S. Let \mathfrak D:=\mathfrak R_2^{(0)}\cup V^{(0,1)} be a domain on \mathfrak R_2(w), and assume that R is such that f_{\infty^{(0)}} extends to a (single-valued) meromorphic function f(\mathbf z) on \mathfrak D, f\in\mathscr M(\mathfrak D). Then R_n(z) is also extended to \mathfrak D as a meromorphic function R_n(\mathbf z). Let q_m(z)= z^m+\dotsb be the polynomial whose zeros coincide (with multiplicities) with the projections of poles of f(\mathbf z) that lie in \mathfrak D, and such that, moreover, the functions \widetilde{f}:=q_mf and q_mR_n are holomorphic in \mathfrak D. In what follows we assume that n>m. Finally, we assume without loss of generality that S\ni0. This is a purely technical condition: it ensures that the function g_S(z,\infty)\mkern-1mu-\mkern-1mu\log|z| is continuous in any domain \{z\colon\mkern-1mu g_S(z,\infty)\mkern-1mu>\mkern-1mu\log\rho\}, \rho\mkern-1mu>\mkern-1mu1 (see (35)).
Following tradition, below we identify the sheet \mathfrak R_2^{(0)}\,{=}\,\mathfrak R_2^{(0)}(w) of \mathfrak R_2(w) with the ‘physical’ extended complex plane \widehat{\mathbb C} cut along the arcs forming Stahl’s S-compact set.
Given an arbitrary \rho\in(1,R), where R is as above, we denote by \Gamma^{(1)}_\rho the set of points z^{(1)} such that g_S(z,\infty)=\log\rho for z^{(1)}\in\Gamma^{(1)}_\rho. Clearly, g(z^{(1)})=-\log\rho for z^{(1)}\in\Gamma^{(1)}_\rho. The set \Gamma^{(0)}_\rho is defined similarly: on it we have g(z^{(0)})=\log\rho. Set \Gamma_\rho:=\pi(\Gamma^{(0)}_\rho)=\pi(\Gamma^{(1)}_\rho). For \rho\in(1,R) let D^{(1)}_\rho be the subdomain of \mathfrak D with boundary \partial D^{(1)}_\rho=\Gamma^{(1)}_\rho such that \infty^{(0)}\in D^{(1)}_\rho. In a similar way D^{(0)}_\rho\subset\mathfrak D, \partial D^{(0)}_\rho=\Gamma^{(0)}_\rho and \infty^{(0)}\in D^{(0)}_\rho. Set
Below in this section we consider only \rho\in (1,R) such that |q_m(z)[f(z^{(0)})-f(z^{(1)})]|\geqslant C(\rho)>0 for z\in\Gamma_\rho (clearly, f(z^{(0)})-f(z^{(1)})\not\equiv0).
where g_{\Gamma_{\rho_1}}(z,\infty) is the Green’s function for the domain g_S(z,\infty)>\log\rho_1. Clearly, g_{\Gamma_{\rho_1}}(z,\infty)=g_S(z,\infty)-\log\rho_1. From (13) we obtain the estimate
Note that relations (15) are ‘comparisonal’, that is, they are preserved by renormalizations of the polynomial Q_n replacing Q_n(z)=z^{k_n}+\dotsb by c_nQ_n, c_n\neq0. In what follows we fix a certain analogue of the so-called spherical normalization for Q_n (see [5] and [3]).
2.3
Let D_\rho:=\{z\in\widehat{\mathbb C}\colon g_S(z,\infty)>\log\rho\}, \rho>1, and let g_{\Gamma_\rho}(z,\infty) be the Green’s function for D_\rho. Then
Since the function \widetilde{u}_n is subharmonic in D_\rho and \widetilde{u}_n\leqslant0 on \Gamma_\rho, it follows that \widetilde{u}_n\leqslant 0 in D_\rho and \widetilde{u}_n(\infty)\leqslant0. Thus we have \log\rho-\gamma_S\leqslant\log m_n(\rho)^{1/k_n}, and finally we obtain
For z\in K\Subset D_\rho and \zeta\in\Gamma_\rho we have g_{\Gamma_\rho}(\zeta,z)=0 by definition. Using the symmetry of the Green’s function in its arguments we obtain g_{\Gamma_\rho}(z,\zeta)=0 for z\in K\Subset D_\rho and \zeta\in\Gamma_\rho. Now let us extend g_{\Gamma_\rho}(z,\zeta) by identical zero inside \Gamma_\rho with respect to \zeta: g_{\Gamma_\rho}(z,\zeta)\equiv0 for z\in D_\rho and \zeta \in\operatorname{int} \Gamma_\rho. Throughout the rest of this section we consider only \rho>1 such that Q_n(z)\neq0 for z\in\Gamma_\rho and all n\in\mathbb N.
for z\in D_{\rho}. Then v_n is a harmonic function in D_\rho and v_n\leqslant0 on \Gamma_\rho. From this it follows that for z\in \Gamma_{\rho_2}, where \rho<\rho_2<R, we have
Now let the probability measure \mu be a limit point of the sequence \{\mu_n\}, where \mu_n=\dfrac1{k_n}\chi(Q_n), that is, let \mu_n\xrightarrow{\ast}\mu as n\to\infty, n\in\Lambda\subset\mathbb N. We also assume that z_n^*\to z^*\in \Gamma_{\rho_2} as n\to\infty, n\in\Lambda. Then by the descendence principle (see [7], Ch. I, § 3, Theorem 1.3; [14], Ch. I, Theorem 6.8; [4] and also the lemma below) we have
where \mu^{(\rho)}\!=\!\mu|_{\overline{D}_\rho}. It follows directly from (26) that \mu|_{\overline{D}_\rho}\!\!=\!0, so that {\operatorname{supp}{\mu}\!\subset\!\widehat{\mathbb C}\!\setminus\! D_\rho}. Since this is true for all \rho>1 except a countable set of values of \rho, we have \operatorname{supp}{\mu}\subset S.
2.4
Thus we have obtained that k_n/n\to1 and each limit point \mu of the sequence \biggl\{\dfrac1{k_n}\chi(Q_n)\biggr\} satisfies the condition \operatorname{supp}\mu\subset S. Now we show that \mu=\lambda_S.
We denote the error function corresponding to this normalization of the Padé polynomial by R^{*}_n. Let m^{*}_n(\rho') and M^{*}_{n,1}(\rho'), \rho'>1, be the analogues of the quantities m_n(\rho') for Q_n and M_{n,1}(\rho') for R_n, which are obtained by replacing Q_n by Q^{*}_n and R_n by R^{*}_n. Then (15) also holds for these quantities.
Set1[x]1Recall that we consider only \rho>1 such that Q_n(z)\neq0 for z\in\Gamma_\rho and all n\in\mathbb N.
Since k_n/n\to1 and \operatorname{supp}{\widetilde{\mu}_{n,2}}, \operatorname{supp}{\mu_{n,1}} \subset \widehat{\mathbb C}\setminus D_\rho, we see that, as n\to\infty, n\in\Lambda, we have
uniformly on compact subsets of D_\rho. In turn, from (15) and (31)–(35) we obtain (28). It follows from (28) that u(z)=\mathrm{const}, so that V^\mu(z)=V^{\lambda_S}(z) for z\in D. Since S contains no interior points, we finally obtain {\mu=\lambda_S}. Thus the equilibrium measure \lambda_S is the unique limit point of the sequence \dfrac1n\chi(Q_n).
Thus we have shown that the following limits exist for any \rho>1:
It is clear that the above relations do not depend on the choice of \rho>1 in the *-normalization (27) for Q_n. Thus, we can take the *-normalization with respect to R>1 which we fixed in § 2.2. We stick to this convention in what follows.
Relation (38) follows from (37) and (42) by the two-constants theorem (cf. [15], § 3.8, formulae (31)–(36), and [3]).
In fact, (38) means that for an arbitrary compact set K\subset D=\widehat{\mathbb C}\setminus{S} and each \varepsilon>0 we have (cf. [15], § 3.8, formulae (31)–(36), and [19], Theorem 1.1)
It follows from (42) that we must consider only the case when K_n(\varepsilon):=K_{2,n}(\varepsilon), that is, we must show that \operatorname{cap}(K_n(\varepsilon))\to0 as n\to\infty.
Let \rho_1 and \rho_2, 1<\rho<\rho_1<\rho_2, be numbers such that K\subset G:=D_{\rho_2}\setminus\overline{D}_{\rho_1}. Then K_n(\varepsilon)\subset G for all n\in\mathbb N. Throughout the rest of this subsection we consider only compact sets K lying in the open set G. Then the properties of the logarithmic capacity \operatorname{cap}(K) of K are equivalent to the properties of the Green’s capacity \operatorname{cap}_\rho(K) of K with respect to the compact set \Gamma_\rho=\partial D_\rho in the following sense: the relation \operatorname{cap}_\rho(K_n)\to0 as n\to\infty is equivalent to \operatorname{cap}(K_n)\to0 (see [7], [14], [4] and formula (49) below). Thus, we must prove that \operatorname{cap}_\rho(K_n(\varepsilon))\to0 as n\to\infty.
Assume the contrary: let \operatorname{cap}_\rho(K_n(\varepsilon))\geqslant\delta for some \delta>0 and n\in\Lambda, n\to\infty.
Since q_mR^{*}_n is a holomorphic function in D_\rho\supset K_n(\varepsilon), each point z\in K_n(\varepsilon) has a neighbourhood U(z) such that
Hence there exists a compact set F_n(\varepsilon)=\bigcup_{j=1}^N \overline{U}(z_j) such that F_n(z)\supset K_n(\varepsilon), F_n(\varepsilon)\subset G, F_n(\varepsilon) is a regular compact set, \operatorname{cap}_\rho(F_n(\varepsilon))\geqslant\delta>0 for n\in\Lambda and
Since q_m(z)R^{*}_n(z) is a holomorphic function in G\supset K, by the maximum principle inequality (43) also holds in the polynomial hull of F_n(\varepsilon). Thus we can assume that F_n(\varepsilon) does not separate the complex plane.
Set D_n(\varepsilon):=D_\rho\setminus F_n(\varepsilon). Then D_n(\varepsilon) is a domain with boundary \partial D_n(\varepsilon)=\Gamma_\rho\cup\partial F_n(\varepsilon). Let \omega_n(z) be the harmonic measure of \partial F_n(\varepsilon) with respect to \Gamma_\rho, that is, let \omega_n(z) be the harmonic function in D_n(\varepsilon) that is continuous in the closure of \overline{D}_\rho(\varepsilon) and satisfies \omega_n(z)\equiv 0 for z\in\Gamma_\rho and \omega_n(z)\equiv1 for z\in\partial F_n(\varepsilon). Set
For an arbitrary compact set K\subset G with positive capacity \operatorname{cap}(K) and for an arbitrary unit measure \mu with support in K, \operatorname{supp}(\mu)\subset K, we define the Green’s potential of \mu relative to the domain D_\rho by
is positive. Since F_n(\varepsilon) is a regular compact set and \operatorname{cap}_\rho(F_n(\varepsilon))\geqslant\delta>0, it follows that G_{\lambda_{F_n(\varepsilon)}}(z)\equiv\gamma_\rho(F_n(\varepsilon)) on \partial F_n(\varepsilon) and \gamma_\rho(F_n(\varepsilon))\leqslant\log(1/\delta) for n\in\Lambda. Hence the harmonic measure \omega_n(z) introduced above has the representation
In fact, the class of admissible multivalued analytic functions for which the above proof is valid is much wider than the class \mathscr F of algebraic functions satisfying conditions (I) and (II) in the above definition. It particular, our approach holds for a certain class of analytic functions produced by the inverse Joukowsky function. More precisely, let
where z\in\widehat{\mathbb C}\setminus\Delta, \Delta=[-1,1] and the branch of (\,\cdot\,)^{1/2} is chosen so that \varphi(z)/z\to2 as z\to\infty. Let 1<A<B<\infty and let a:=(A+1/A)/2 and b:=(B+1/B)/2. Then the function
is an algebraic function of the fourth order with square root singularities. Let \Sigma(\mathfrak f)=\{\pm1,a,b\} be the corresponding set of branch points of \mathfrak f(z;\Delta). Under the above condition on (\,\cdot\,)^{1/2} we have \mathfrak f_\infty\in\mathscr H(\widehat{\mathbb C}\setminus\Delta), and Stahl’s compact set for \mathfrak f_\infty is the interval [-1,1]: S(\mathfrak f_\infty)=[-1,1]. Now let \varphi_{\Delta_j}(z) be the inverse Joukowsky function corresponding to an interval \Delta_j:=[\alpha_j,\beta_j], j=1,\dots,m, where \Delta_j\cap \Delta_k=\varnothing, j\neq k. Set
where each \mathfrak f(z;\Delta_j) is defined by (55) for \varphi_{\Delta_j} in place of \varphi and some A_j and B_j in place of A and B.
If all intervals \Delta_j are real, \Delta_j\subset\mathbb R, j=1,\dots,m, then Stahl’s compact set has the following form: S(\mathfrak f)=\bigsqcup_{j=1}^m\Delta_j. Since all functions in \mathbb C(z,\mathfrak f) satisfy condition (II) in the definition of \mathscr F, our approach also holds for the germ f_\infty of an arbitrary function f in the class \mathbb C(z,\mathfrak f).
Another nontrivial admissible class of multivalued analytic functions arises when we assume that at least one of the branch points \alpha_j and \beta_j, j=1,\dots,m, does not belong to the real line.
We can also generalize (55) in the following way. Set
where \alpha,\beta,\dots,\gamma\in\mathbb C\setminus\mathbb Z and \alpha+\beta+\dots+\gamma\in\mathbb Z. Then we obtain a class of functions \mathbb C(z,\mathfrak f), which is not a subclass of \mathscr F, but for which the proof in § 2 is also valid.
3.2
Let D\subset\widehat{\mathbb C}, D\neq\widehat{\mathbb C}, be a regular domain and g_D(\zeta,z), z,\zeta\in D, be the Green’s function for D with logarithmic singularity at \zeta=z. For an arbitrary (positive Borel) measure \mu we define its Green potential G_D(z;\mu) with respect to D by
\begin{equation*}
G_D(z;\mu):=\int g_D(\zeta,z)\,d\mu(\zeta), \qquad z\in D.
\end{equation*}
\notag
Then the following result holds.
Lemma. Let K\subset D be a compact set, let \{z_n\}_{n\in\mathbb N}\subset K be a sequence of points such that z_n\to z^* as n\to\infty and \{\mu_n\} be a sequence of measures such that \operatorname{supp}\mu_n\subset\overline{D}, \mu_n\to\mu as n\to\infty, where \mu is a probability measure. Then
Proof. Since K\subset D, for all z\in K and each sufficiently small \varepsilon\in(0,\varepsilon_0), {\varepsilon_0\,{=}\,\varepsilon_0(K)}, the level curve L_\varepsilon:=\{\zeta\colon g_D(\zeta,z)= 1/\varepsilon\} is a closed analytic curve enclosing the point z and such that g_D(\zeta,z)> 1/\varepsilon inside L_\varepsilon and g_D(\zeta,z)<1/\varepsilon outside L_\varepsilon.
Let D_\varepsilon\ni z be the domain with boundary \partial D_\varepsilon=L_\varepsilon. For \varepsilon\in(0,\varepsilon_0) we set (cf. [7], Ch. I, § 3.7)
Then for any fixed \varepsilon\in(0,\varepsilon_0) the family of functions \{G^{(\varepsilon)}_D(z;\mu_n)\}_{n\in\mathbb N} is equicontinuous for z\in K. Moreover, for each fixed z\in K the functions g^{(\varepsilon)}_D(\zeta,z) and G^{(\varepsilon)}_D(z;\mu) are monotonically increasing as \varepsilon\to0. Thus,
A. I. Aptekarev, V. I. Buslaev, A. Martínez-Finkelshtein and S. P. Suetin, “Padé approximants, continued fractions, and orthogonal polynomials”, Uspekhi Mat. Nauk, 66:6(402) (2011), 37–122; English transl. in Russian Math. Surveys, 66:6 (2011), 1049–1131
2.
A. I. Aptekarev and M. L. Yattselev, “Padé approximants for functions with branch points – strong asymptotics of Nuttall-Stahl polynomials”, Acta Math., 215:2 (2015), 217–280
3.
V. I. Buslaev, “On a lower bound for the rate of convergence of multipoint Padé approximants of piecewise analytic functions”, Izv. Ross. Akad. Nauk Ser. Mat., 85:3 (2021), 13–29; English transl. in Izv. Math., 85:3 (2021), 351–366
4.
E. M. Chirka, “Capacities on a compact Riemann surface”, Tr. Mat. Inst. Steklova, 311 (2020), 41–83; English transl. in Proc. Steklov Inst. Math., 311 (2020), 36–77
5.
A. A. Gonchar and E. A. Rakhmanov, “Equilibrium distributions and degree of rational approximation of analytic functions”, Mat. Sb., 134(176):3(11) (1987), 306–352; English transl. in Sb. Math., 62:2 (1989), 305–348
6.
A. V. Komlov, “The polynomial Hermite-Padé m-system for meromorphic functions on a compact Riemann surface”, Mat. Sb., 212:12 (2021), 40–76; English transl. in Sb. Math., 212:12 (2021), 1694–1729
7.
N. S. Landkof, Foundations of modern potential theory, Nauka, Moscow, 1966, 515 pp. ; English transl., Grundlehren Math. Wiss., 180, Springer-Verlag, New York–Heidelberg, 1972, x+424 pp.
8.
A. Martínez-Finkelshtein, E. A. Rakhmanov and S. P. Suetin, “Variation of the equilibrium energy and the S-property of stationary compact sets”, Mt. Sb., 202:12 (2011), 113–136; English transl. in Sb. Math., 202:12 (2011), 1831–1852
9.
J. Nuttall and S. R. Singh, “Orthogonal polynomials and Padé approximants associated with a system of arcs”, J. Approx. Theory, 21:1 (1977), 1–42
10.
J. Nuttall, “Asymptotics of diagonal Hermite-Padé polynomials”, J. Approx. Theory, 42:4 (1984), 299–386
11.
E. A. Perevoznikova and E. A. Rakhmanov, Variation of equilibrium energy and the S-property of
compact sets with minimum capacity, Manuscript, 1994 (Russian)
12.
E. A. Rahmanov (Rakhmanov), “Convergence of diagonal Padé approximants”, Mat. Sb., 104(146):2(10) (1977), 271–291; English transl. in Sb. Math., 33:2 (1977), 243–260
13.
E. A. Rakhmanov, “Orthogonal polynomials and S-curves”, Recent advances in orthogonal polynomials, special functions, and their applications, Contemp. Math., 578, Amer. Math. Soc., Providence, RI, 2012, 195–239
14.
E. B. Saff and V. Totik, Logarithmic potentials with external fields, Appendix B by T. Bloom, Grundlehren Math. Wiss., 316, Springer-Verlag, Berlin, 1997, xvi+505 pp.
15.
H. Stahl, “Three different approaches to a proof of convergence for Padé approximants”, Rational approximation and applications in mathematics and physics (Łańcut 1985), Lecture Notes in Math., 1237, Springer, Berlin, 1987, 79–124
16.
H. Stahl, “Diagonal Padé approximants to hyperelliptic functions”, Ann. Fac. Sci. Toulouse Math. (6), 1996, special issue, 121–193
17.
H. Stahl, “The convergence of Padé approximants to functions with branch points”, J. Approx. Theory, 91:2 (1997), 139–204
18.
H. R. Stahl, Sets of minimal capacity and extremal domains, arXiv: 1205.3811
19.
M. L. Yattselev, “Convergence of two-point Padé approximants to piecewise holomorphic functions”, Mat. Sb., 212:11 (2021), 128–164; English transl. in Sb. Math., 212:11 (2021), 1626–1659
Citation:
S. P. Suetin, “A direct proof of Stahl's theorem for a generic class of algebraic functions”, Sb. Math., 213:11 (2022), 1582–1596