Abstract:
Some mixed potential-theoretic equilibrium problems are treated. These problems are motivated by the asymptotic properties of rational approximants based on type II Hermite–Padé polynomials for Nikishin systems of functions.
Bibliography: 26 titles.
Gonchar and Rakhmanov applied in 1981 for the first time the potential-theoretic approach the asymptotic properties of Hermite–Padé polynomials. In [9] (see also [10]) they considered the case of type II Hermite–Padé polynomials for a system of $m$ functions, which also form an Angelesco system. Since $m\geqslant2$ could be an arbitrary natural number, the arising potential-theoretic equilibrium problem is a vector equilibrium problem described in terms of an $ m\times m $ matrix. In 1986 Nikishin [16] introduced a new system of functions which is now called a ‘Nikishin system’ after him. Following the ideas of Gonchar and Rakhmanov [9], Nikishin considered in [16] type I Hermite–Padé polynomials for $m$ functions forming a Nikishin system. The asymptotics of these Hermite–Padé polynomials were described in [16] in terms of an associated vector equilibrium problem, which was stated in terms of an $ m\times m $ matrix. For the further development of this powerful Gonchar–Rakhmanov method, see [1], [15] and the bibliography there.
In [18] (see also [5]) the case of two functions forming a Nikishin system was considered. Instead of the vector equilibrium problem for a logarithmic potential, in [18] a scalar equilibrium problem was proposed for a mixed Green-logarithmic potential. The asymptotics of the corresponding type I Hermite–Padé polynomials were described in [18] in terms of the unique solution of this scalar equilibrium problem. In [5] it was proved that for this equilibrium problem there is a conjugate equilibrium problem, which has a unique solution, equal to the balayage of the solution of the original equilibrium problem; for details see [5] and § 1.2 below. It has turned out [25] that the asymptotics of the type II Hermite–Padé polynomials (HP-polynomials) can be described in terms of the unique solution of the conjugate equilibrium problem.
It should be noted also that for some special cases the original equilibrium problem was stated and treated for the first time in [11], in connection with the problem of the convergence of the so-called ‘linear’ (or Frobenious-type) and ‘nonlinear’ (or Baker-type) Padé approximations to orthogonal expansions. This is quite natural since there is a well-known connection between type I HP-polynomials and linear Padé approximations to orthogonal expansions and, in particular, expansions in Chebyshev polynomials; see [12], [2], [19] and [26].
In both cases, of the original mixed equilibrium problem and of the conjugate one, the corresponding equilibrium measure $\lambda$ depends on a real nonnegative parameter $\theta$ (see (4) and (5) below), that is, $\lambda=\lambda(\theta)$. Since the rate of convergence of the corresponding constructive rational approximants depends on the Green’s potential of the equilibrium measure $\lambda(\theta)$, it would be natural to consider the property of monotonicity in $\theta$ of Green’s potential to compare these rates with one another. For the initial mixed equilibrium problem this was done in [11] for $\theta=0,1$ and $3$. In this work we consider the case of the conjugate mixed equilibrium problem. We note that the equilibrium problems that come from HP-theory are very closely connected with extremal problems of geometric function theory; see [17], [1] and [18] and cf. [3] and [8]. For the very recent results on the asymptotics of HP-polynomials see [20], [4], [24] and [21].
Note in conclusion that the asymptotic properties of Hermite–Padé polynomials found recently wide applications to the Rayleigh–Schrödinger perturbation theory; see, for example, [6] and [7] and the bibliography there.
Acknowledgement
The authors express their gratitude to the reviewers for the careful examination of the present work, and are thankful for their comments, which contributed to the corrections made in the final manuscript.
1.2.
For an arbitrary regular (in the sense of the solution of the Dirichlet problem) compact set $K\subset\mathbb{R}$ let us denote by $M_1(K)$ the set of all probability (Borel) measures with support on $K$. For a compact set $S\subset\mathbb{R}$ let $g_S(\zeta,z)$ be the Green’s function for the domain $\widehat{\mathbb{C}}\setminus{S}$ with a logarithmic singularity at the point $\zeta=z$. Let $V^\mu(z)$ be the logarithmic potential of the measure $\mu\in M_1(K)$ and $G^\mu_S(z)$ be the Green’s potential of $\mu$ corresponding to the Green’s function $g_S(\zeta,z)$, $S\cap K=\varnothing$:
It is well known (see [14]) that there exists a unique measure $\tau_K\in M_1(K)$ with the following property:
$$
\begin{equation}
V^{\tau_K}(x)\equiv\gamma_K=\mathrm{const}, \qquad x\in K.
\end{equation}
\tag{2}
$$
The measure $\tau_K$ is called the Robin measure (or the equilibrium measure) for the compact set $K$, and $\gamma_K$ is the Robin constant of $K$. Notice that $\operatorname{supp}{\tau_K}=K$. Let
be the Green’s function for the domain $D=D(K):=\widehat{\mathbb{C}}\setminus{K}$ with a logarithmic singularity at the point at infinity $z=\infty$. Then
Let $E\subset\mathbb{R}$ and $F\subset\mathbb{R}$ be two compact sets consisting of a finite number of disjoint closed intervals, $E=\bigsqcup_{j=1}^p E_j$ and $F=\bigsqcup_{k=1}^qF_k$, and such that $\operatorname{conv}(E)\cap \operatorname{conv}(F)=\varnothing$, where $\operatorname{conv}(\,{\cdot}\,)$ denotes the convex hull of the corresponding real set. For definiteness we suppose that $F$ is positioned to the right of $E$ (see Figure 1). We note that the ordered pair of compact sets $(E,F)$ forms a so-called Nuttall condenser (see [18], where this notion was first introduced). In the asymptotic theory of the Hermite–Padé polynomials for a pair of functions that form a Nikishin system, the Nuttall condenser plays the same role as the Stahl compact set plays in the theory of Padé polynomials. From the equilibrium equations (4) and (5) it follows that we have indeed an ordered pair of compact sets.
For a nonnegative real $\theta$ consider the mixed Green-logarithmic potential ${\theta V^\mu(z) +G^\mu_F(z)}$, where $\mu\in M_1(E)$. It is well known (see [11], [18] and [5]) that there exists a unique measure $\lambda_E=\lambda_E(\theta)$ with support on $E$, $\lambda_E(\theta)\in M_1(E)$, that has the following property:
Let us introduce another mixed Green-logarithmic potential, namely, ${\theta V^\nu(z) +G^\nu_E(z)}$, $\nu\in M_1(F)$, and consider for it the following equilibrium problem with the potential of an external field given by $g_E(z,\infty)$:
It is known (see [5]) that there exists a unique solution of (5), which is a measure $\lambda_F=\lambda_F(\theta)$ with support on $F$, $\lambda_F(\theta)\in M_1(F)$. Notice that $\operatorname{supp}{\lambda_F}=F$.
The following connection1[x]1In fact, it is a one-to-one correspondence. (see [5]) exists between $\lambda_E(\theta)\in M_1(E)$ and ${\lambda_F(\theta)\in M_1(F)}$:
where ${\mathfrak b}_F(\,{\cdot}\,)$ is the balayage of a given measure from the domain $G:=\widehat{\mathbb{C}}\setminus{F}$ onto its boundary $\partial G=F$.
Also, the following connection exists between the potentials:
(see [5], formula (24)). In view of relation (6) and identity (7), the potential-theoretic equilibrium problem (5) can be considered to be conjugate to the equilibrium problem (4).
§ 2. Statement of the problem
2.1.
It was proved in [11] (see also [12]) that the Green’s potentials of the equilibrium measures $\lambda_F(1)$ and $\lambda_F(3)$ determine the rates of convergence of the nonlinear ($F_n$) and linear ($\Phi_n$) Padé approximants, respectively, to the orthogonal expansion of a Markov-type function $f(z)=\widehat{\varkappa}(z)$ corresponding to $\operatorname{supp}{\varkappa}=F=[c,d]$, $b<c$. More precisely, let $N=2n+1=3m+1$ and let $c_0,c_1,\dots, c_{N-1}$ be the $N$ Fourier coefficients of the orthogonal expansion of the Markov-type function $f(z)$ with respect to a measure $\sigma$ on $E=[a,b]$ such that $\sigma'=d\sigma/dx>0$ almost everywhere on $[a,b]$. Then from these $N$ Fourier coefficients we can construct the nonlinear Padé–Fourier approximant $F_n$ of degree $(n,n)$ and the linear Padé–Fourier approximant $\Phi_m$ of degree $(m,m)$, where $3m=2n$. It was proved in [11] that the rates of convergence of the constructive (with respect to the given $c_0,c_1,\dots,c_{N-1}$; see [13], § 2) rational approximants $R_N=F_n$ and $R_N=\Phi_m$ are determined by the following relations, which take place locally uniformly2[x]2That is on compact subsets of $G$ . in the domain $G$:
Since $G^{\lambda_E(1)}_F(z) >\frac23G^{\lambda_E(3)}_F(z)$ for $z\in G$ (see [11], formula (9)), we have ${\delta_2(z)\,{<}\,\delta_1(z)}$ and thus, the nonlinear Padé–Fourier approximants $F_n$ are more efficient than the linear Padé–Fourier approximants $\Phi_m$ as $N\to\infty$.
Notice (see [11]) that the equilibrium measure $\lambda_E(0)$ and the corresponding function $G^{\lambda_E(0)}_F$ are associated with the best (Chebyshev) rational approximants of degree $(n,n)$ to the Markov type function $f(z)=\widehat{\varkappa}(z)$ on $E=[a,b]$. However, these approximants are not constructive in the sense of the paper [13] by Henrici (see [13], § 2).
2.2.
Now let $f(z)\in\mathscr{H}(\infty)$ be given by the following explicit representation:
where $1<A<B$ and the branch of the root function $(\,{\cdot}\,)^{1/2}$ is chosen so that $\varphi(z)=z+(z^2-1)^{1/2}\sim 2z$ and, as a result, $f(z)\sim1/\sqrt{AB}$ as $z\to\infty$. The function $f$ is an algebraic function of fourth degree with four branch points $\{\pm1,a,b\}$, where $a=(A+1/A)/2$ and $b=(B+1/B)/2$, $1<a<b$. The interval $E=[-1,1]$ is the Stahl compact set $S(f)$ for $f$ given by (9) and $D:=\widehat{\mathbb{C}}\setminus{E}$ is the corresponding Stahl domain. We denote the class of such analytic functions by $\mathscr Z([-1,1])$.
In [23] it was proved that $f\in\mathscr Z([-1,1])$ is a Markov-type function, and the pair $f$, $f^2$ and the triple $f$, $f^2$, $f^3$ form Nikishin systems [16], [17]:
where $\operatorname{supp}\sigma=\operatorname{supp}{s}_1=\operatorname{supp}{s}_2=E$, $s_1:=\langle \sigma,\sigma_2\rangle$, that is, $ds_1(z)=\widehat{\sigma}_2(z)\,d\sigma(z)$, $\operatorname{supp}{\sigma_2}=[a,b]$ and $s_2:=\langle \sigma,\langle \sigma_2,\sigma\rangle\rangle$. The measures $\sigma$ and $\sigma_2$ admit explicit representations; see [23], formulae (16) and (17).
Let $f\in\mathscr Z([-1,1])$, $f\in\mathscr{H}(\infty)$ and $N=2n+1=3m+1=4\ell+1$.
For the pair of functions $f$, $f^2$ let us define the type II HP-polynomials $P^{(2)}_{2m,0}\not\equiv0$, $P^{(2)}_{2m,1}$ and $P^{(2)}_{2m,2}$ of degree $\leqslant 2m$ by the relations
Analogously, for the triple of functions $f$, $f^2$, $f^3$ let us define the type II HP-polynomials $P^{(3)}_{3\ell,0}\not\equiv0$, $P^{(3)}_{3\ell,1}$, $P^{(3)}_{3\ell,2}$ and $P^{(3)}_{3\ell,3}$, $\operatorname{deg}P^{(3)}_{3\ell,j}\leqslant 3\ell$, by the relations:
where ‘$\xrightarrow{*}$’ denotes convergence in the space $M_1(E)$ in the weak-$*$ topology, ${j\!=\!0,1,2}$ and $k=0,1,2,3$.
2.3.
In [25] (see also [6]) the following result was announced.
Theorem 1 ([25]). Let $f\in\mathscr Z(E)$, $E=[-1,1]$. Then, as $N\to\infty$, the following relations are valid locally uniformly in the domain $D=\widehat{\mathbb{C}}\setminus{E}$:
$$
\begin{equation}
\delta_3(z)<\delta_2(z)<\delta_1(z)<1, \qquad z\in D.
\end{equation}
\tag{16}
$$
The main result of this paper is the following statement, which defines the monotonicity property of the potential $G_E^{\lambda_F(\theta)}(z)$ with equilibrium measure ${\lambda_F(\mkern-1mu\theta\mkern-1mu)\mkern-1mu\!\in\! M_1(\mkern-1mu F\mkern-1mu)}$ for $\theta\in[1,3]$.
Theorem 2. Let $1\leqslant\theta_1<\theta_2\leqslant3$. Then for $z\in{D}$
Let $\mathfrak{R}_3$ be the three-sheeted Riemann surface constructed in the following way (see Figure 1).
The zeroth sheet $\mathfrak{R}^{(0)}_3$ of the Riemann surface $\mathfrak{R}_3$ is a copy of the extended complex plane (Riemann sphere) $\widehat{\mathbb{C}}$, cut along the compact set $F=\bigsqcup_{k=1}^qF_k$. Thus we consider this new compact set as two-sided and denote it by $F^{(0)}_{+}\cup F^{(0)}_{-}$. The first sheet $\mathfrak{R}^{(1)}_3$ is constructed in a similar way: we consider a copy of the extended complex plane $\widehat{\mathbb{C}}$ cut along the two compact sets $E=\bigsqcup_{j=1}^pE_j$ and $F=\bigsqcup_{k=1}^qF_k$. Both these new compact sets are considered as two-sided and denoted by $E^{(1)}_{+}\cup E^{(1)}_{-}$ and $F^{(1)}_{+}\cup F^{(1)}_{-}$, respectively. Finally, to obtain the second sheet $\mathfrak{R}^{(2)}_3 $ we take a copy of the extended complex plane $\widehat{\mathbb{C}}$ cut along the compact set $E=\bigsqcup_{j=1}^pE_j$, consider this new compact set as two-sided and denote it by $E^{(2)}_{+}\cup E^{(2)}_{-}$. Finally, we construct the whole Riemann surface $\mathfrak{R}_3$ by gluing crosswise3[x]3In other words, we identify points on $F^{(0)}_{+}$ with the corresponding points on $F^{(1)}_{-}$ and points on $F^{(0)}_{-}$ with the corresponding points on $F^{(1)}_{+}$. Similarly, we identify points on the $E^{(2)}_{+}$ with the corresponding points on $E^{(1)}_{-}$ and points on $E^{(2)}_{-}$ with the corresponding points on $E^{(1)}_{+}$. the zeroth sheet $\mathfrak{R}^{(0)}_3$ with the first sheet $\mathfrak{R}^{(1)}_3$ along $F^{(0)}_{+}\cup F^{(0)}_{-}$ and $F^{(1)}_{+}\cup F^{(1)}_{-}$ and gluing crosswise the second sheet $\mathfrak{R}^{(2)}_3$ with the first sheet $\mathfrak{R}^{(1)}_3$ along $E^{(2)}_{+}\cup E^{(2)}_{-}$ and $E^{(1)}_{+}\cup E^{(1)}_{-}$. We denote the point on $\mathfrak{R}^{(0)}_3$ corresponding to a point $z$ in $G=\widehat{\mathbb{C}}\setminus{F}$ by $z^{(0)}$, the point on $\mathfrak{R}^{(1)}_3$ corresponds to $z$ in $\widehat{\mathbb{C}}\setminus({E\cup F})$ by $z^{(1)}$ and the point on $\mathfrak{R}^{(2)}_3$ corresponding to $z$ in $D=\widehat{\mathbb{C}}\setminus{E}$ by $z^{(2)}$. Set $F^{(2)}:=\{z^{(2)}\colon z\in F\}$ and $E^{(0)}:=\{z^{(0)}\colon z\in E\}$.
The boundary $\Gamma^{(0,1)}$ between the zeroth and first sheets consists of $q$ closed disjoint curves on the Riemann surface $\mathfrak{R}_3$: $\Gamma^{(0,1)}=\bigsqcup_{k=1}^q \Gamma^{(0,1)}_k$. Similarly, the boundary $\Gamma^{(1,2)}$ between the first and second sheets consists of $p$ closed disjoint curves on $\mathfrak{R}_3$: $\Gamma^{(1,2)}=\bigsqcup_{j=1}^p \Gamma^{(1,2)}_j$.
For a point on $\mathfrak{R}_3$ we use the notation $\mathbf{z}$. In particular, $\mathbf{z}=z^{(0)}\in\mathfrak{R}^{(0)}_3$, $\mathbf{z}=z^{(1)}\in\mathfrak{R}^{(1)}_3$ and $\mathbf{z}=z^{(2)}\in\mathfrak{R}^{(2)}_3$. Thus, there is a natural projection $\pi$ of $\mathfrak{R}_3$ onto $\widehat{\mathbb{C}}$, $\pi\colon\mathfrak{R}_3\to \widehat{\mathbb{C}}$, given by the equality $\pi(z^{(j)})=z$, $j=0,1,2$, for $z^{(j)}\in\mathfrak{R}^{(j)}_3$ and $z\notin(E\cup F)$. For the disjoint curves $\Gamma^{(0,1)}$ and $\Gamma^{(1,2)}$ on $\mathfrak{R}_3$ we have $\pi(\Gamma^{(0,1)})=F$ and $\pi(\Gamma^{(1,2)})=E$, respectively. Therefore, $\pi(\mathbf{z})\in F$ for $\mathbf{z}\in\Gamma^{(0,1)}$ and $\pi(\mathbf{z})\in E$ for $\mathbf{z}\in\Gamma^{(1,2)}$.
For points $\mathbf z\in\mathfrak R_3\setminus\bigl(\Gamma^{(0,1)}\cup\Gamma^{(1,2)}\bigr)$ we define a $*$-operation $\mathbf{z}\mapsto\mathbf{z}_{*}$ by the following rule:
where $\overline{z}$ is the complex conjugate of $z\in\mathbb{C}$.
Figure 1 shows the three-sheeted Riemann surface $\mathfrak{R}_3$ constructed, along with the ‘physical’ complex plane $\widehat{\mathbb{C}}$ which contains the compact sets $E$ and $F$. We denote by $E^{(0)}\subset\mathfrak{R}^{(0)}_3$ the compact set corresponding to $E=\bigsqcup_{j=1}^pE_j\subset\widehat{\mathbb{C}}$ and by $F^{(2)} \subset \mathfrak{R}^{(2)}_3$ the compact set corresponding to $F\subset\widehat{\mathbb{C}}$.
Note that all three sheets of Riemann surface $\mathfrak{R}_3$ are open subsets of $\mathfrak{R}_3$ and $\mathfrak{R}_3=\mathfrak{R}_3^{(0)} \sqcup\Gamma^{(0,1)}\sqcup\mathfrak{R}_3^{(1)} \sqcup\Gamma^{(1,2)}\sqcup\mathfrak{R}_3^{(2)}$.
3.2.
We define a function $v_\theta(\mathbf{z})$ in the following way:
Clearly, $v_\theta(\mathbf{z})\equiv0$ for $\mathbf{z}=z^{(0)}\in E^{(0)}$, $v_\theta(\mathbf{z})=4\theta\log|z|+O(1)$ as $\mathbf{z}\to\infty^{(2)}$, and $v_\theta(\mathbf{z})$ is a harmonic function on $\mathfrak{R}^{(0)}_3\setminus E^{(0)}$, $\mathfrak{R}^{(1)}$ and $\mathfrak{R}^{(2)}_3\setminus(F^{(2)}\cup\infty^{(2)})$. Thus, $v_\theta(\mathbf{z})$ is a piecewise harmonic function except at the point $\mathbf{z}=\infty^{(2)}$.
Lemma 1. The piecewise harmonic function $v_\theta(\mathbf{z})$ can be extended across the curves $\Gamma^{(0,1)}$ and $\Gamma^{(1,2)}$ to produce a harmonic function on $\mathfrak{R}_3\setminus(E^{(0)}\cup F^{(2)}\cup\infty^{(2)})$.
Proof. Let $ U^{(0,1)}\subset\mathfrak{R}_3$ be a neighbourhood of $\Gamma^{(0,1)}$ such that $ U^{(0,1)}\cap E^{(0)}=\varnothing$ and $U^{(0,1)}\cap \Gamma^{(1,2)}=\varnothing$; see Figure 1. From (5) it follows that
is a harmonic function in $ U^{(0)}:= U^{(0,1)}\cap \mathfrak{R}^{(0)}_3$ with the following property: it is continuous on $U^{(0)}\cup \Gamma^{(0,1)}$ and $u_1(z^{(0)})\equiv0$ for $z^{(0)}\in\Gamma^{(0,1)}$. Therefore, we can continue the function $u_1(\mathbf{z})$, $\mathbf{z}\in U^{(0)}$, across the curves $\Gamma^{(0,1)}$ to the open set $U^{(1)}:=U^{(0,1)}\cap \mathfrak{R}^{(1)}_3$ with opposite values at $*$-symmetric points:
The extended function $u_1(\mathbf{z})$ is a harmonic function in $U^{(0,1)}\supset\Gamma^{(0,1)}$.
Let $\widetilde{\lambda}={\mathfrak b}_E(\lambda_F(\theta))\in M_1(E)$ be the balayage of the measure $\lambda_F(\theta)\in M_1(F)$ from the domain $D$ onto its boundary $\partial{D}=E$. Then $V^{\widetilde{\lambda}}(z)\equiv V^{\lambda_F(\theta)}(z)-G_E^{\lambda_F(\theta)}(z)+ \mathrm{const}$, $z\in\widehat{\mathbb C}$. Since $\widetilde{\lambda}\in M_1(E)$, the potential $V^{\widetilde{\lambda}}(z)$ is a harmonic function in $\pi(U^{(0,1)})$. The function $V^{\widetilde{\lambda}}$ can be lifted to the set $U^{(0,1)}\subset\mathfrak{R}_3$, as well as to the sheets $\mathfrak{R}^{(0)}_3$, $\mathfrak{R}^{(1)}_3$ and $\mathfrak{R}^{(2)}_3$ in the following way:
From (20) and (21) it follows that we can define the new function $u_2(\mathbf z):=u_1(\mathbf z)-\theta V^{\widetilde{\lambda}}(\mathbf z)-\theta g_E(\mathbf z,\infty)+\mathrm{const}$, which is harmonic for $\mathbf{z}\in U^{(0,1)}$, and we obtain the following representation in $U^{(0,1)}$:
This function $u_2(\mathbf{z})$ is harmonic in $U^{(0,1)}$.
Let $U^{(1,2)}\subset\mathfrak{R}_3$ be a neighbourhood of $\Gamma^{(1,2)}$ such that $U^{(1,2)}\cap \Gamma^{(0,1)}=\varnothing$ and $U^{(1,2)}\cap F^{(2)}=\varnothing$. Since $\lambda_F(\theta)\in M_1(F)$, the function $V^{\lambda_F(\theta)}(z)$ is harmonic in $\pi(U^{(1,2)})$. Also, $G^{\lambda_F(\theta)}_E(z)\equiv0$ and $g_E(z,\infty)\equiv0$ for $z\in E$, so that the function
The function $v(\mathbf{z})$ is a continuous function on the Riemann surface $\mathfrak{R}_3$ and is a harmonic function in the domain $\mathfrak{R}_3\setminus(E^{(0)}\cup F^{(2)})$. From (25) and (26) it follows that
From here and (27), by the minimum principle for harmonic functions it follows that $v(\mathbf{z})>0$ in the domain $\mathfrak D:=\mathfrak{R}_3\setminus (E^{(0)}\cup F^{(2)})$. Therefore,
and Theorem 2 follows for $\theta_1=1$ and $\theta_2=3$. Clearly, the theorem is also true for ${1<\theta_1<\theta_2=3}$.
Now we prove the theorem in the general case of $1\leqslant\theta_1<\theta_2<3$.
If $v(z^{(2)})\geqslant0$ for $z^{(2)}\in F^{(2)}$, then $v(\mathbf{z})>0$ in the domain $\mathfrak D$, since $v(z^{(0)})\equiv0$ on $E^{(0)}$, and it is easy to check that $v(\mathbf{z})\not\equiv0$. Theorem 2 follows from the inequality $v(\mathbf{z})>0$ in the domain $\mathfrak D$.
Now assume that $m:=\min_{z\in F}v(z^{(2)})<0$. Then there exists a point $y_0\in F$ such that $m=v(y^{(2)}_0)<0$. Therefore,
From the minimum principle for harmonic functions it follows that $v(\mathbf{z})>m$ for ${\mathbf{z}\in\mathfrak D}$. Hence $v(y_0^{(0)})>m$, and from the definition (26) of the function $v(\mathbf{z})$ and (27) we obtain
The relation (29) implies the following inequality:
$$
\begin{equation}
\biggl(2-\frac2{\theta_1}\biggr)G^{\lambda_F(\theta_1)}_E(y_0)> \biggl(2-\frac2{\theta_2}\biggr)G^{\lambda_F(\theta_2)}_E(y_0), \qquad y_0\in F.
\end{equation}
\tag{30}
$$
In the case when $\theta_1=1$ relation (30) implies a contradiction since $G^{\lambda_F(\theta_2)}_E(y_0)>0$. Thus, we can now assume that $1<\theta_1<\theta_2<3$.
Therefore, $g(t)$ is a decreasing function on the interval $(1,3)$. This contradicts (32). Relation (32) holds under the assumption that $\min_{z\in F}v(z^{(2)})<0$. Therefore, it follows from this contradiction that $v(z^{(2)})\geqslant0$ on $F^{(2)}$. Theorem 2 is proved.
Bibliography
1.
A. I. Aptekarev and V. G. Lysov, “Systems of Markov functions generated by graphs and the asymptotics of their Hermite–Padé approximants”, Sb. Math., 201:2 (2010), 183–234
2.
A. I. Aptekarev, A. I. Bogolyubskii and M. L. Yattselev, “Convergence of ray sequences of Frobenius–Padé approximants”, Sb. Math., 208:3 (2017), 313–334
3.
F. G. Avkhadiev, I. R. Kayumov and S. R. Nasyrov, “Extremal problems in geometric function theory”, Russian Math. Surveys, 78:2 (2023), 211–271
4.
A. I. Aptekarev, S. Yu. Dobrokhotov, D. N. Tulyakov and A. V. Tsvetkova, “Plancherel–Rotach type asymptotic formulae for multiple orthogonal Hermite polynomials and recurrence relations”, Izv. Math., 86:1 (2022), 32–91
5.
V. I. Buslaev and S. P. Suetin, “On equilibrium problems related to the distribution of zeros of the Hermite–Padé polynomials”, Proc. Steklov Inst. Math., 290:1 (2015), 256–263
6.
E. O. Dobrolyubov, N. R. Ikonomov, L. A. Knizhnerman and S. P. Suetin, Rational Hermite–Padé approximants vs Padé approximants. Numerical results, arXiv: 2306.07063
7.
E. O. Dobrolyubov, I. V. Polyakov, D. V. Millionshchikov and S. V. Krasnoshchekov, “Vibrational resonance phenomena of the OCS isotopologues studied by resummation of high-order Rayleigh–Schrödinger perturbation theory”, J. Quant. Spectrosc. Radiat. Transf., 316 (2024), 108909, 13 pp.
8.
V. N. Dubinin, “Green energy of discrete signed measure on concentric circles”, Izv. Math., 87:2 (2023), 265–283
9.
A. A. Gonchar and E. A. Rakhmanov, “On the convergence of simultaneous Padé approximants for systems of functions of Markov type”, Proc. Steklov Inst. Math., 157 (1983), 31–50
10.
A. A. Gonchar and E. A. Rakhmanov, “Equilibrium measure and the distribution of zeros of extremal polynomials”, Math. USSR-Sb., 53:1 (1986), 119–130
11.
A. A. Gonchar, E. A. Rakhmanov and S. P. Suetin, “On the convergence of Padé approximations of orthogonal expansions”, Proc. Steklov Inst. Math., 200 (1993), 149–159
12.
A. A. Gonchar, E. A. Rakhmanov and S. P. Suetin, “Padé–Chebyshev approximants for multivalued analytic functions, variation of equilibrium energy, and the $S$-property of stationary compact sets”, Russian Math. Surveys, 66:6 (2011), 1015–1048
13.
P. Henrici, “An algorithm for analytic continuation”, SIAM J. Numer. Anal., 3:1 (1966), 67–78
14.
N. S. Landkof, Foundations of modern potential theory, Grundlehren Math. Wiss., 180, Springer-Verlag, New York–Heidelberg, 1972, x+424 pp.
15.
V. G. Lysov, “Mixed type Hermite–Padé approximants for a Nikishin system”, Proc. Steklov Inst. Math., 311 (2020), 199–213
16.
E. M. Nikishin, “The asymptotic behavior of linear forms for joint Padé approximations”, Soviet Math. (Iz. VUZ), 30:2 (1986), 43–52
17.
E. M. Nikishin and V. N. Sorokin, Rational approximations and orthogonality, Transl. Math. Monogr., 92, Amer. Math. Soc., Providence, RI, 1991, viii+221 pp.
18.
E. A. Rakhmanov and S. P. Suetin, “The distribution of the zeros of the Hermite–Padé polynomials for a pair of functions forming a Nikishin system”, Sb. Math., 204:9 (2013), 1347–1390
19.
E. A. Rakhmanov and S. P. Suetin, “Tschebyshev–Padé approximations for multivalued functions”, Tr. Mosk. Mat. Obs., 83, no. 2, MCCME, Moscow, 2022, 319–344; English transl., arXiv: 2106.01047
20.
V. N. Sorokin, “A generalization of the discrete Rodrigues formula for Meixner polynomials”, Sb. Math., 213:11 (2022), 1559–1581
21.
V. N. Sorokin, “On polynomials defined by the discrete Rodrigues formula”, Math. Notes, 113:3 (2023), 420–433
22.
H. Stahl, “The convergence of Padé approximants to functions with branch points”, J. Approx. Theory, 91:2 (1997), 139–204
23.
S. P. Suetin, “On an example of the Nikishin system”, Math. Notes, 104:6 (2018), 905–914
24.
S. P. Suetin, “Asymptotic properties of Hermite–Padé polynomials and Katz points”, Russian Math. Surveys, 77:6 (2022), 1149–1151
25.
S. P. Suetin, “Convergence of Hermite–Padé rational approximations”, Russian Math. Surveys, 78:5 (2023), 967–969
26.
L. N. Trefethen, “Numerical analytic continuation”, Jpn. J. Ind. Appl. Math., 40:3 (2023), 1587–1636
Citation:
N. R. Ikonomov, S. P. Suetin, “On some potential-theoretic problems related to the asymptotics of Hermite–Padé polynomials”, Mat. Sb., 215:8 (2024), 52–65; Sb. Math., 215:8 (2024), 1053–1064